text
stringlengths
4
2.78M
meta
dict
--- author: - 'Ivan Bliznets [^1]' - 'Fedor V. Fomin [^2]' - 'Marcin Pilipczuk [^3]' - 'Michał Pilipczuk [^4]' bibliography: - '../completion.bib' title: | A subexponential parameterized algorithm for\ <span style="font-variant:small-caps;">Interval Completion</span>[^5] --- Introduction {#sec:intro} ============ Preliminaries {#sec:prelims} ============= Overview of the algorithm {#sec:overview} ========================= Modules and neighborhood classes {#sec:neighbors} ================================ Listing potential maximal cliques and sections {#sec:pmc} ============================================== Guessing fill-in edges with fixed endpoint {#sec:fill-in} ========================================== Small-separation lemma {#sec:left-right} ====================== Dynamic programming {#sec:dp} =================== Conclusions {#sec:conc} =========== Appendix {#sec:boring .unnumbered} ======== [^1]: St. Petersburg Academic University of the Russian Academy of Sciences, Russia, `[email protected]`. [^2]: Department of Informatics, University of Bergen, Norway, `[email protected]`. [^3]: Department of Computer Science, University of Warwick, United Kingdom, `[email protected]`. [^4]: Faculty of Mathematics, Computer Science, and Mechanics, University of Warsaw, Poland, `[email protected]`. [^5]: The research leading to these results has received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP/2007-2013) / ERC Grant Agreement n. 267959
{ "pile_set_name": "ArXiv" }
--- author: - 'D. Drechsel, S. S. Kamalov, L. Tiator' date: - 'Received: date / Revised version: date' - 'October 1, 2007' title: 'Unitary Isobar Model - MAID2007' --- Introduction ============ Our knowledge about the excitation spectrum of the nucleon was originally provided by elastic pion-nucleon scattering [@Hohler79]. All the resonances listed in the Particle Data Tables [@PDG06] have been identified by partial-wave analyses of this process with both Breit-Wigner and pole extraction techniques. From such analyses we know the resonance masses, widths, and branching ratios into the $\pi N$ and $\pi\pi N$ channels. These are reliable parameters for the resonances in the 3- and 4-star tiers, with only few exceptions. In particular, there remains some doubt about the structure of two prominent resonances, the Roper $P_{11}(1440)$, which appears unusually broad, and the $S_{11}(1535)$, where the pole can not be uniquely determined, because it lies close to the $\eta N$ threshold.\ On the basis of these relatively firm grounds, additional information can be obtained for the electromagnetic (e.m.) $\gamma N N^*$ couplings through pion photo- and electroproduction. These couplings are described by electric, magnetic, and charge transition form factors, $G_E^*(Q^2)$, $G_M^*(Q^2)$, and $G_C^*(Q^2)$, or by linear combinations thereof as helicity amplitudes $A_{1/2}(Q^2)$, $A_{3/2}(Q^2)$, and $S_{1/2}(Q^2)$. So far we have some reasonable knowledge of the transverse amplitudes $A_{1/2}$ and $A_{3/2}$ at the real photon point, which are tabulated in the Particle Data Tables. For finite $Q^2$ the information found in the literature is scarce and until recently practically nonexistent for the longitudinal amplitudes $S_{1/2}$. But even for the transverse amplitudes only few results have remained firm over the recent years, such as the $G_M^*$ form factor of the $P_{33}(1232)$ or $\Delta(1232)$ resonance up to $Q^2\approx 10$ GeV$^2$, the $A_{1/2}(Q^2)$ for the $S_{11}(1535)$ up to $Q^2\approx 5$ GeV$^2$, and the helicity asymmetry ${\mathcal {A}}(Q^2)$ for the resonances $D_{13}(1520)$ and $F_{15}(1680)$ up to $Q^2\approx 3$ GeV$^2$ [@Boffi96]. Frequently also data points for other resonances, e.g., the Roper resonance, are shown together with quark model calculations. However, the statistical errors are often quite large and the model dependence of the analysis may be even larger. In this context it is worth mentioning that also the notion of a ‘data point’ is somewhat misleading because the photon couplings and amplitudes can only be derived indirectly by a partial-wave analysis. It is in fact prerequisite to analyze a particular experiment within a framework based on the “world data”. The only exception from this caveat is the $\Delta(1232)$ resonance. For this lowest-lying and strongest resonance of the nucleon, the analysis is facilitated by two important constraints: the validity of (I) the Watson theorem at the 1 % level and (II) the truncation of the multipole series to $S$ and $P$ waves as a good first-order approximation. With these assumptions the e.m. couplings have been directly determined in the real photon limit by a complete experiment with polarized photons and detecting both neutral and charged pions in the final state, thus allowing also for an isospin separation [@Beck97]. Moreover, a nearly complete separation of the possible polarization observables has recently provided the basis to extend such a “model-independent” analysis also to electroproduction [@Kelly]. However, we are still far from such a situation for all the higher resonances. Neither are the mentioned constraints valid nor are we close to a complete experiment. Until recently the data base was rather limited, the error bars were large, and no data were available from target or recoil polarization experiments. Even now there exist only very few data points from double-polarization experiments at energies above the $\Delta(1232)$. However, the situation for unpolarized $e+p\rightarrow e'+p+\pi^0$ reaction has considerably improved, mainly by new JLab experiments in all three halls A, B, and C. These data cover a large energy range from the $\Delta(1232)$ up to the third resonance region with a wide angular range in $\theta_\pi$. Furthermore, electron beam polarization has been used in several experiments at JLab, MIT/Bates, and MAMI/Mainz. Because of the large coverage in the azimuthal angle by the modern large-acceptance detectors, a separation of all 4 partial cross sections in the unpolarized experiment becomes possible. But even without a Rosenbluth separation of the transverse ($\sigma_T$) and longitudinal ($\sigma_L$) cross sections, there is an enhanced sensitivity to the longitudinal amplitudes due to the interference terms $\sigma_{LT}$ and $\sigma_{LT'}$. Such data are the basis of our new partial-wave analysis with an improved version of the Mainz unitary isobar model MAID.\ We proceed by presenting a brief history of the unitary isobar model in Sect. 2. The formalism of pion photo- and electroproduction is summarized in Sect. 3. In the following Sect. 4 we present our results for photoproduction as obtained from the latest version MAID2007, and in Sect. 5 this analysis is extended to electroproduction. We conclude with a short summary in Sect. 6. History of MAID =============== - [**MAID98**]{}\ In 1998 the first version of the Unitary Isobar Model was developed and implemented on the web to give an easy access for the community. MAID98 was constructed with a limited set of nucleon resonances described by Breit-Wigner forms and a non-resonant background constructed from Born terms and t-channel vector-meson contributions [@Maid]. In order to have the right threshold behavior and a reasonable description at the higher energies, the Born terms were introduced with an energy-dependent mixing of pseudovector and pseudoscalar $\pi NN$ coupling. Each partial wave was unitarized up to the two-pion threshold by use of Watson’s theorem. Specifically, the unitarization was achieved by introducing additional phases $\phi_R$ in the resonance amplitudes in order to adjust the phase of the total amplitude. Only the following 4-star resonances were included: $P_{33}(1232)$, $P_{11}(1440)$, $D_{13}(1520)$, $S_{11}(1535)$, $S_{11}(1650)$, $F_{15}(1680)$, and $D_{33}(1700)$. The e.m. vertices of these resonances were extracted from a best fit to the VPI/GWU partial-wave analysis [@VPI97]. For the $P_{33}(1232)$ resonance, we determined the following ratios of transition amplitudes: (I) electric quadrupole to magnetic dipole transition, $R_{EM}=E2/M1=-2.2\%$, and (II) electric Coulomb to magnetic dipole transition, $R_{CM}=C2/M1=-3.6\%$, independent of the 4-momentum transfer $Q^2$. The $Q^2$-dependence of the resonance amplitudes in the second and third resonance regions was expressed in terms of the quark electric and magnetic multipoles [@Burk]. The non-unitarized background contributions were determined using standard Born terms and vector-meson exchange. In order to preserve gauge invariance, the Born terms were expressed by the usual dipole form for the Sachs form factors, and both the pion and the axial form factor were set equal to the isovector Dirac form factor, $ F_{\pi}(Q^2)=G_A(Q^2)= F_1^p(Q^2)-F_1^n(Q^2)$. - [**MAID2000**]{}\ In this version of MAID, the background contribution was unitarized for the multipoles up to $F$ waves according to the prescription of K-matrix theory. The $S$-wave multipoles $E_{0+}$ and $L_{0+}$ were modified in order to improve their energy dependence in the threshold region. With the new unitarization procedure, the pion photoproduction multipoles of SAID and some selected data for pion photo- and electroproduction in the energy range up to $W=1.6$ GeV were fitted [@NSTAR2001]. The ratios of the $\Delta(1232)$ multipoles were found to be $R_{EM}=-2.2\%$ and $R_{CM}=-6.5\%$, still independent of $Q^2$. - [**MAID2003**]{}\ In accordance with results of Ref. [@DMT], the $Q^2$ dependence of the electric and Coulomb excitations of the $\Delta(1232)$ resonance was modified. The ratio $R_{EM}$ was found to change sign at $Q^2 \approx 3.3$ GeV$^2$ from negative to positive values, and $R_{CM}$ decreased from $-6.5\%$ at $Q^2=0$ to $-13.5\%$ at $Q^2=4$ GeV$^2$. Moreover, the following 4-star resonances were included in MAID2003: $S_{31}(1620)$, $D_{15}(1675)$, $P_{13}(1720)$, $F_{35}(1905)$, $P_{31}(1910)$, and $F_{37}(1950)$. In contrast to previous versions, the helicity amplitudes of all 13 resonances were input parameters and their $Q^2$ dependence was parameterized by polynomials. With this new version of MAID we directly analyzed all the pion photo- and electroproduction data available since 1960, and for the first time we made local (single energy) and global (energy dependent) fits, independent of the GWU/SAID group. - [**MAID2005**]{}\ The $Q^2$ dependence of the Sachs form factors in the Born terms was replaced by the more recent parameterization of Ref. [@Kelly04], and at the e.m. vertices of the pion-pole and seagull terms, realistic pion and axial form factors were introduced. As a result the description of charged pion electroproduction was much improved. On the basis of a large amount of new data from MIT/Bates, ELSA/Bonn, Grenoble, Mainz, and Jefferson Lab, we performed new local and global as well as single-$Q^2$ fits and obtained a better description of the data in the energy range 1.6 GeV$<W<$2 GeV [@MAID05]. - [**MAID2007**]{}\ The present version of MAID is presented in some detail in the following sections. As far as pion photoproduction is concerned, this version is identical to MAID2005. The main changes are related to the $Q^2$ evolution of e.m. form factors. In particular, the $Q^2$-dependence of the $\Delta(1232)$ transition form factors are remodeled to be consistent with the Siegert theorem. As a result the ratio $R_{CM}$ decreases sharply if $Q^2$ approaches zero. Furthermore, our analysis of the recent high-$Q^2$ data [@Ungaro06] led to the conclusion that $R_{EM}$ remains negative in the range of the existing experiments. Formalism for pion photo- and electroproduction =============================================== Let us first define the kinematics of pion photo- and electroproduction on a nucleon, $$\gamma^{\ast}(k) + N(p) \to \pi(q) + N'(p')\,,\label{eq:3.1}$$ where the variables in brackets denote the 4-momenta of the participating particles. In the pion-nucleon center-of mass (c.m.) system, we define $$k^{\mu} = (\omega_{\gamma}, {\vec k})\,,\; q^{\mu} = (\omega_{\pi}, {\vec q})\,,\;p^{\mu} = (E_N, -{\vec k}) \,,\label{eq:3.2}$$ where $$\begin{aligned} k(W,Q^2)& = & |\vec k|= \sqrt{\left( \frac{W^2-m^2-Q^2}{2W}\right)^2+Q^2}\,,\label{eq:3.3}\\ q(W)& = & |\vec q|= \sqrt{\left(\frac{W^2-m^2+m_{\pi}^2}{2W}\right)^2-m_{\pi}^2}\,, \label{eq:3.4}\end{aligned}$$ with $W=\omega_{\gamma}+E_N$ the total c.m. energy and $Q^2=k^2-\omega_{\gamma}^2>0$ the square of 4-momentum deposited by the photon at the nucleon vertex, also referred to as the virtuality of the photon. In order to simplify the notation, we use the following abbreviations: $$k_W =k (W,0)= \frac{W^2-m^2}{2W}\,,\label{eq:3.5}\,$$ describing the momentum of a real photon, and $$k_R =k (M_R,0)\,,\; q_R =q (M_R)\,,\label{eq:3.6}\,$$ for the real photon and pion momenta at the resonance position, $W=M_R$.\ The basic equations used for MAID2007 are taken from the dynamical Dubna-Mainz-Taipei (DMT) model [@DMT; @Yang85; @KY99]. In this approach the t-matrix for pion photo- and electroproduction takes the form $$t_{\gamma\pi}(W)=v_{\gamma\pi}(W)+v_{\gamma\pi}(W)\,g_0(W)\,t_{\pi N}(W)\,, \label{eq:3.7}$$ with $v_{\gamma\pi}$ the transition potential for the reaction $\gamma^{\ast}N \rightarrow \pi N$, $t_{\pi N}$ the $\pi N$ scattering matrix, and $g_0$ the free $\pi N$ propagator. In a resonant channel the transition potential $v_{\gamma\pi}$ consists of two terms, $$v_{\gamma\pi}(W)=v_{\gamma\pi}^B(W) +v_{\gamma\pi}^R(W)\,, \label{eq:3.8}$$ with $v_{\gamma\pi}^B$ the background transition potential and $v_{\gamma\pi}^R$ the contribution of the “bare” resonance excitation. The resulting t-matrix can be decomposed into two terms [@KY99] $$t_{\gamma\pi}(W)=t_{\gamma\pi}^B(W) + t_{\gamma\pi}^{R}(W)\,,\label{eq:3.9}$$ where $$\begin{aligned} t_{\gamma\pi}^B(W) &=& v_{\gamma\pi}^B(W)+v_{\gamma\pi}^B(W)\,g_0(W)\,t_{\pi N}(W)\,, \label{eq:3.10}\\ t_{\gamma\pi}^R(W) &=& v_{\gamma\pi}^R(W)+v_{\gamma\pi}^R(W)\,g_0(W)\,t_{\pi N}(W)\,, \label{eq:3.11}\end{aligned}$$ with $t_{\gamma\pi}^B$ including the contributions from both the non-resonant background and the $\gamma^*NR$ vertex renormalization. The decomposition in resonance and background contributions is not unique, however, our definition has the advantage that all the processes starting with the e.m. excitation of a bare resonance are summed up in $t_{\gamma\pi}^R$. Unitarized background --------------------- The multipole decomposition of Eq. (\[eq:3.10\]) yields the background contribution to the physical amplitudes in the channels $\alpha=(\xi,\ell,j,I)$ [@Yang85], where $\ell$, $j$ and $I$ denote the orbital momentum, the total angular momentum, and the isospin of the pion-nucleon final state, and $\xi$ stands for the magnetic ($\xi=M$), electric ($\xi=E$), and Coulomb or “scalar” ($\xi=S$) transitions, $$\begin{aligned} t^{B,\alpha}_{\gamma\pi}(q,k,Q^2)& = & v_{\gamma\pi}^{B,\alpha}(q,k,Q^2)\label{eq:3.12}\\ &+& \int_0^{\infty} dq' \frac{q'^2 t_{\pi N}^{\alpha}(q,q';W)\,v_{\gamma\pi}^{B,\alpha}(q',k,Q^2)} {W-W_{\pi N}(q')+i\epsilon}\,,\nonumber\end{aligned}$$ where $W_{\pi N}(q')$ is the hadronic c.m. energy in the intermediate state. The pion electroproduction potential $v^{B,\alpha}_{\gamma\pi}$ is constructed as in Ref. [@Maid] and contains contributions from the Born terms described by an energy-dependent mixing of pseudovector (PV) and pseudoscalar (PS) $\pi NN$ coupling as well as t-channel vector meson exchange. The quasi-potential $v^{B,\alpha}_{\gamma\pi}$ depends on 5 parameters: the PV-PS mixing parameter $\Lambda_m$ as defined in Eq. (12) of Ref. [@Maid] and 4 coupling constants for the vector-meson exchange. The on-shell parts of $v^{B,\alpha}_{\gamma\pi}$ and $t^{B,\alpha}_{\gamma\pi}$ depend on two variables only, i.e., $$\begin{aligned} v^{B,\alpha}_{\gamma\pi}(q,k,Q^2) &=& v^{B,\alpha}_{\gamma\pi}(W,Q^2)\label{eq:3.13}\\ t^{B,\alpha}_{\gamma\pi}(q,k,Q^2) &=& t^{B,\alpha}_{\gamma\pi}(W,Q^2)\,.\label{eq:3.14}\end{aligned}$$ The $Q^2$ evolution of the s- and u-channel nucleon pole terms of the background is described by the form factors of Ref. [@Kelly]. At the e.m. vertices of the pion-pole and seagull terms we apply a monopole form for the pion form factor and a dipole form for the axial form factor, while the standard dipole form factor is used for the vector-meson exchange.\ We note that the background contribution of MAID98 was defined by $t_{\gamma\pi}^{B,\alpha}({\rm MAID}98)=v_{\gamma\pi}^{B,\alpha}(W,Q^2)$ and assumed to be a real and smooth function. The unitarization of the total amplitude was then provided by an additional phase $\phi_{\alpha}$ in the resonance contribution such that the phase of the total amplitude had the phase $\delta_{\alpha}$ of the respective $\pi N$ scattering state. In MAID2007, however, the background contributions are complex functions defined according to K-matrix theory, $$t^{B,\alpha}_{\gamma\pi}(W,Q^2)=v^{B,\alpha}_{\gamma\pi}(W,Q^2)\,[1+it_{\pi N}^{\alpha}(W) ]\, , \label{eq:3.15}$$ where the pion-nucleon elastic scattering amplitudes, $t^{\alpha}_{\pi N}=[\eta_{\alpha} \exp(2i\delta_{\alpha})-1]/2i$, are described by the phase shifts $\delta_{\alpha}$ and the inelasticity parameters $\eta_{\alpha}$ taken from the GWU/SAID analysis [@VPI]. The assumed structure of the background corresponds to neglecting the principal value integral in the pion-rescattering term of Eq. (\[eq:3.12\]). Our previous studies of the $P$-wave multipoles in the (3,3) channel [@DMT; @KY99] showed that the “pion cloud” contributions of the principal value integral are effectively included by the dressing of the $\gamma N N^{\ast}$ vertex.\ Furthermore, the threshold behavior of the $S$ waves was improved. The results of the dynamical approaches [@DMTthr] show that the pion cloud contributions are very important to obtain a good description of the $E_{0+}$ multipole in the $\pi^0 p$ channel. For this purpose we have introduced the following phenomenological term: $$E^{\rm {corr}}_{0+}(W,Q^2)= \frac{A}{(1 + B^2 q^2)^2}\,G_D(Q^2)\,, \label{eq:3.16}$$ with $A$ and $B$ free parameters fixed by fitting the low-energy $\pi^0$ photoproduction data, and $G_D$ the standard nucleon dipole form factor. The threshold correction for the $L_{0+}$ multipole we will consider later in Sect. 5.3. As a result the background contribution of MAID now depends on 8 parameters. We furthermore account for the cusp effect in the $\pi^0 p$ channel appearing at the $\pi^+n$ threshold by the term [@Laget; @Bernard] $$E^{\rm {cusp}}_{0+}= - a_{\pi N} \,\omega_c\,Re E_{0+}^{\gamma\pi^+}\,\sqrt{1-\frac{\omega_{\pi}^2}{\omega_c^2}}\,, \label{eq:3.17}$$ where $\omega_c=140$ MeV is the $\pi^0$ c.m. energy at the cusp and $a_{\pi N}=0.124/m_{\pi^+}$ the pion charge-exchange amplitude. Resonance contributions ----------------------- For the resonance contributions we follow Ref. [@Maid] and assume Breit-Wigner forms for the resonance shape, $$t_{\gamma\pi}^{R,\alpha}(W,Q^2)\,=\,{\bar{\cal A}}_{\alpha}^R(W,Q^2)\, \frac{f_{\gamma N}(W)\Gamma_{tot}\,M_R\,f_{\pi N}(W)}{M_R^2-W^2-iM_R\, \Gamma_{tot}} \,e^{i\phi_R}\,, \label{eq:3.18}$$ where $f_{\pi N}(W)$ is the usual Breit-Wigner factor describing the decay of a resonance with total width $\Gamma_{tot}(W)$, partial $\pi N$ width $\Gamma_{\pi N}$, and spin $j$, $$f_{\pi N}(W)=C_{\pi N}\left[\frac{1}{(2j+1)\pi}\frac{k_W}{q} \frac{m}{M_R}\frac{\Gamma_{\pi N}}{\Gamma_{tot}^2}\right]^{1/2}\,, \label{eq:3.19}$$ with $C_{\pi N}=\sqrt{3/2}$ and $-1/\sqrt{3}$ for isospin $\frac {3}{2}$ and $\frac {1}{2}$, respectively. The energy dependence of the partial width is given by $$\Gamma_{\pi N}(W)=\beta_{\pi}\,\Gamma_R\,\left(\frac{q}{q_R} \right)^{2l+1}\,\left(\frac{X^2_R+q_R^2}{X^2_R+q^2}\right)^{\ell} \, \frac{M_R}{W}\,, \label{eq:3.20}$$ with $\Gamma_R=\Gamma_{tot}(M_R)$ and $X_R$ a damping parameter and $\beta_{\pi}$ the single-pion branching ratio. The expression for the total width $\Gamma_{tot}$ is given in Ref. [@Maid]. The $\gamma NN^*$ vertex is assumed to have the following dependence on $W$: $$f_{\gamma N}(W)=\left(\frac{k_W}{k_R}\right)^n\, \left(\frac{X^2_R+k_R^2}{X^2_R+k_W^2}\right)\,, \label{eq:3.21}$$ where $n$ is obtained from a best fit to the real photon data, and with the normalization condition $f_{\gamma N}(M_R)=1$. The phase $\phi_R(W)$ in Eq. (\[eq:3.18\]) is introduced to adjust the total phase such that the Fermi-Watson theorem is fulfilled below two-pion threshold. For the $S$- and $P$-wave multipoles we extend this unitarization procedure up to $W=1400$ MeV. Because of a lack of further information, we assume that the phases $\phi_R$ are constant at the higher energies. In particular we note that the phase $\phi_R$ for the $P_{33}(1232)$ excitation vanishes at $W=M_R=1232$ MeV for all values of $Q^2$. For this multipole we may even apply the Fermi-Watson theorem up to $W \approx 1600$ MeV because the inelasticity parameter $\eta_{\alpha}$ remains close to 1. For the $D$- and $F$-wave resonances, the phases $\phi_R$ are assumed to be constant and determined from the best fit.\ Whereas MAID98 [@Maid] included only the 7 most important nucleon resonances, essentially with only transverse e.m. couplings, our present version contains all 13 resonances of the 4-star tier below 2 GeV with transverse electric (${\bar{\cal A}}_{\alpha}^R=\bar{E}_{l\pm}$), transverse magnetic (${\bar{\cal A}}_{\alpha}^R=\bar{M}_{l\pm}$), and Coulomb (${\bar{\cal A}}_{\alpha}^R=\bar{S}_{l\pm}$) couplings: $P_{33}(1232)$, $P_{11}(1440)$, $D_{13}(1520)$, $S_{11}(1535)$, $S_{31}(1620)$, $S_{11}(1650)$, $D_{15}(1675)$, $F_{15}(1680)$, $D_{33}(1700)$, $P_{13}(1720)$, $F_{35}(1905)$, $P_{31}(1910)$, and $F_{37}(1950)$. Because we determine the isovector amplitudes from the proton channels, the number of the e.m. couplings is 34 for the proton and 18 for the neutron channels, that is 52 parameters altogether. These are taken to be constant in a single-Q$^2$ analysis, e.g., in photoproduction but also at any fixed $Q^2$ if sufficient data are available in the chosen energy and angular range. Alternatively, the couplings have also been parameterized as functions of $Q^2$, as is discussed in Sec. 5.\ $N^{\ast}$ $\bar{E}$ $\bar{M}$ $\bar{S}$ ------------------- -- -------------------------------------------------- -- -- --------------------------------------------------- -- --------------------------------- $S_{11}$/$S_{31}$ $-A_{1/2}$ — $-\sqrt{2}S_{1/2}$ $P_{13}$/$P_{33}$ $\frac{1}{2}(\frac{1}{\sqrt{3}}A_{3/2}-A_{1/2})$ $-\frac{1}{2}(\sqrt{3}A_{3/2}+A_{1/2})$ $-\frac{1}{\sqrt{2}}S_{1/2}$ $P_{11}$/$P_{31}$ — $A_{1/2}$ $-\sqrt{2}S_{1/2}$ $D_{13}$/$D_{33}$ $-\frac{1}{2}(\sqrt{3}A_{3/2}+A_{1/2})$ $-\frac{1}{2}(\frac{1}{\sqrt{3}}A_{3/2}-A_{1/2})$ $-\frac{1}{\sqrt{2}}S_{1/2}$ $D_{15}$/$D_{35}$ $\frac{1}{3}(\frac{1}{\sqrt{2}}A_{3/2}-A_{1/2})$ $-\frac{1}{3}(\sqrt{2}A_{3/2}+A_{1/2})$ $-\frac{\sqrt{2}}{3}S_{1/2}$ $F_{15}$/$F_{35}$ $-\frac{1}{3}(\sqrt{2}A_{3/2}+A_{1/2})$ $-\frac{1}{3}(\frac{1}{\sqrt{2}}A_{3/2}-A_{1/2})$ $-\frac{\sqrt{2}}{3}S_{1/2}$ $F_{17}$/$F_{37}$ $\frac{1}{4}(\sqrt{\frac{3}{5}}A_{3/2}-A_{1/2})$ $-\frac{1}{4}(\sqrt{\frac{5}{3}}A_{3/2}+A_{1/2})$ $-\frac{1}{2\,\sqrt{2}}S_{1/2}$ The more commonly used helicity amplitudes $A_{1/2}$, $A_{3/2}$, and $S_{1/2}$ are given by linear combinations of the e.m. couplings $\bar{\cal A}_{\alpha}^R$. These relations take the form $$\begin{aligned} A^{\ell +}_{1/2} &=& -\frac{1}{2} [(\ell +2) \bar{E}_{\ell+} + \ell \bar{M}_{\ell+} ] \,, \nonumber\\ A^{\ell +}_{3/2} &=& \frac{1}{2}\sqrt{\ell(\ell+2)} (\bar{E}_{\ell+} - \bar{M}_{\ell+}) \,, \label{3.22}\\ S^{\ell +}_{1/2} &=& -\frac{\ell+1}{\sqrt{2}} \bar{S}_{\ell+} \nonumber\end{aligned}$$ for resonances with total spin $j=\ell +\frac {1}{2}$, and $$\begin{aligned} A^{\ell -}_{1/2} &=& \frac{1}{2} [(\ell +1) \bar{M}_{\ell -} - (\ell -1) \bar{E}_{\ell -}]\,, \nonumber\\ A^{\ell -}_{3/2} &=& -\frac{1}{2}\sqrt{(\ell -1)(\ell+1)}(\bar{E}_{\ell -} + \bar{M}_{\ell -})\,, \label{3.23}\\ S^{\ell -}_{1/2} &=& -\frac{\ell}{\sqrt{2}} \bar{S}_{\ell -} \nonumber\end{aligned}$$ for total spin $j=\ell - \frac {1}{2}$. The inverse relations for the partial waves are listed in Table \[tab:amplitudes\]. The helicity amplitudes are related to matrix elements of the e.m. current $J_{\mu}$ between the nucleon and the resonance states, e.g., as obtained in the framework of quark models, $$\begin{aligned} A_{1/2} &=& -\sqrt{\frac{2\pi\alpha_{\rm {em}}}{k_W}} <R,\frac{1}{2}\,|\,J_{+}\,|\,N,-\frac{1}{2}>\, \zeta\,, \nonumber\\ A_{3/2} &=& -\sqrt{\frac{2\pi\alpha_{\rm {em}}}{k_W}} <R,\frac{3}{2}\,|\,J_{+}\,|\,N,\frac{1}{2}>\,\zeta\,, \label{3.24}\\ S_{1/2} &=& -\sqrt{\frac{2\pi\alpha_{\rm {em}}}{k_W}} <R,\frac{1}{2}\,|\,\rho\,|\,N,\frac{1}{2}>\, \zeta \,, \nonumber\end{aligned}$$ where $J_{+}=-\frac{1}{\sqrt{2}}(J_x+iJ_y)\,$ and $\alpha_{\rm {em}}=1/137$. However, these equations define the couplings only up to a phase $\zeta$, which in principle can be obtained from the pionic decay of the resonance calculated within the same model. Because this phase is ignored in most of the literature, the comparison of the sign is not always meaningful, especially in critical cases such as the Roper resonance whose correct sign is not obvious from the data. In contrast with MAID98 and MAID2000, our present version uses the helicity amplitudes $A_{1/2}$, $A_{1/2}$, and $S_{1/2}$ for photoproduction as input parameters, except for the $P_{33}(1232)$ resonance which is directly described by the 3 e.m. amplitudes $\bar{\cal A}_{\alpha}$. Partial-wave analysis of pion photoproduction data ================================================== The unitary isobar model MAID2007 has been developed to analyze the world data of pion photo- and electroproduction. In this section we fix (I) the background parameters and the helicity amplitudes for pion photoproduction ($Q^2=0$) and (II) the dependence of the resonance contributions on the c.m. energy $W$. These results are then generalized to pion electroproduction in the next section. Data base for pion photoproduction and fit procedure ---------------------------------------------------- The main part of the photoproduction data was taken from the GWU/SAID compilation of SAID2000, which includes the data published between 1960 and 2000, a total of 14700 data points. A separation of these data in different physical channels and observables is given in Table \[tab:data\_SAID\]. In the following years the data base was extended by including recent results from MAMI (Mainz) [@Leukel; @GDH01; @Ilia; @GDH06], GRAAL (Grenoble) [@GRAAL02; @GRAAL05], LEGS (Brookhaven) [@LEGS04], and ELSA (Bonn) [@Bonn05] as listed in Table \[tab:data\_new\]. Altogether 4976 more data points were added. As a result our full data base contains 19676 points within the energy range 140 MeV$<E_{\gamma}<$1610 MeV. channel $d\sigma$ $\Sigma$ $T$ $P$ total ---------- ----------- ---------- ----- ----- ------- $n\pi^+$ 4646 760 645 205 6256 $p\pi^0$ 4936 673 353 540 6502 $p\pi^-$ 1554 206 94 88 1942 : Number of data points from the SAID2000 data base for differential cross sections ($d\sigma$), photon asymmetries ($\Sigma$), target asymmetries ($T$), and recoil asymmetries ($P$).[]{data-label="tab:data_SAID"} channel range $E_{\gamma}$(MeV) data points (observable) Ref. ---------- ------------------------- ---------------------------------------- ------------ $p\pi^0$ 202-790 1129 ($d\sigma$) + 357 ($\Sigma$) [@Leukel] $p\pi^0$ 310-780 174 ($d\sigma$) + 138 ($\Delta\sigma$) [@GDH01] $p\pi^+$ 180-450 205 ($d\sigma$) + 129 ($\Delta\sigma$) [@Ilia] $p\pi^+$ 463-783 204 ($d\sigma$) + 102 ($\Delta\sigma$) [@GDH06] $p\pi^+$ 800-1454 237 ($\Sigma$) [@GRAAL02] $p\pi^0$ 555-1541 861 ($d\sigma$) + 469 ($\Sigma$) [@GRAAL05] $n\pi^-$ 285-769 300 ($d\sigma$) [@LEGS04] $p\pi^0$ 513-1575 671 ($d\sigma$) [@Bonn05] total 4976 : Number of data points collected after 2000 for differential cross sections ($d\sigma$), photon asymmetries ($\Sigma$), and helicity asymmetries $\Delta\sigma=d\sigma_{1/2}-d\sigma_{3/2}$.[]{data-label="tab:data_new"} Our strategy for the data analysis is as follows. First, we try to find a global (energy dependent) solution by fitting all the data in the range 140 MeV$\le E_{\gamma} \le$1610 MeV. This allows us to determine the phase of the multipoles, i.e., the ratio $ {\rm {Im}}\, t_{\gamma\pi}^{\alpha}/{\rm {Re}}\, t_{\gamma\pi}^{\alpha}$ above the two-pion threshold. At the lower energies this phase is constrained by the $\pi N$ scattering phase. In a second step we perform local (single energy) fits to the data, in energy bins of 10 MeV in the range 140 MeV$\le E_{\gamma} \le$460 MeV and of 20 MeV for the higher energies, by varying the absolute values of the multipoles but keeping the phase as previously determined. Similar to the prescription of the SAID group we minimize the modified $\chi^2$ function $$\chi^2 = \sum_i^{N_{\rm {data}}} \left(\frac{\Theta_i-\Theta_i^{\rm {exp}}}{\delta\Theta_i}\right)^2 + \sum_j^{N_{\rm {mult}}} \left(\frac{X_j-1}{\Delta}\right)^2\,. \label{eq:4.1}$$ The first term on the r.h.s. of this equation is the standard $\chi^2$ function with $\Theta_i$ the calculated and $\Theta_i^{\rm {exp}}$ the measured observables, $\delta\Theta_i$ the statistical errors, and $N_{\rm {data}}$ the number of data points. In the second term, $N_{\rm {mult}}$ is the number of the varied multipoles and $X_j$ describes the deviation from the global fit. The fitting procedure starts with the initial value $X_j=1$ corresponding to the global solution, and the quantity $\Delta$ enforces a smooth energy dependence of the single-energy solution. In the limit of $\Delta\rightarrow\infty$ we obtain the standard $\chi^2_{\rm {std}}$, and for $\Delta\rightarrow 0$ the single-energy and the global solutions become identical. The optimum value for $\Delta$ is chosen from the condition $1<\chi^2/\chi^2_{\rm {std}}<1.05$. The described two-step fitting procedure can be repeated several times by adjusting the energy dependence of the global solution, for example by changing the parameters $X_R$ and $n$ in Eqs. (\[eq:3.20\]-\[eq:3.21\]) in order to improve the agreement between the global and local solutions. Results for pion photoproduction -------------------------------- Our results for $\chi^2$ are summarized in Table \[tab:chisq\] by comparing the local and global solutions for different energy ranges and channels. We recall that the number of varied multipoles $N_{\rm {mult}}$ in the proton and neutron channels is different. Since the number of data points in the proton channels ($\gamma\pi^0$ and $\gamma\pi^+$) is about one order of magnitude larger than for the neutron channel ($\gamma\pi^-$), we proceed as follows. First, we analyze the proton channel and extract the multipoles $_pE_{l\pm}^{1/2}$, $_pM_{l\pm}^{1/2}$, $E_{l\pm}^{3/2}$, and $M_{l\pm}^{3/2}$ as defined in Ref. [@Maid]. Second, with the thus obtained values for the isospin 3/2 multipoles, we extract the multipoles $_nE_{l\pm}^{1/2}$ and $_nM_{l\pm}^{1/2}$ from the neutron channel. In this way we minimize the pressure from the large number of proton data on the results in the neutron channel. The number of varied multipoles also depends on the energy. For $E_{\gamma}<$450 MeV we vary all the $S$- and $P$-wave multipoles plus $_{p,n}E_{2-}^{1/2}$ and $E_{2-}^{3/2}$. At the higher energies we include all the multipoles up to the $F$ waves.\ proton --------------------- ------------------ ------------------ --------------------- --------------------- $E_{\gamma}$\[MeV\] $N_{\rm {mult}}$ $N_{\rm {data}}$ $\chi^2_{\rm {se}}$ $\chi^2_{\rm {gl}}$ 140—200 10 990 787 2346 200—450 10 5622 5454 14236 450—850 24 6403 6996 22700 850—1210 24 2965 4133 24990 1210—1610 24 1454 4737 12594 total 17434 22107 76866 neutron $E_{\gamma}$\[MeV\] $N_{\rm {mult}}$ $N_{\rm {data}}$ $\chi^2_{\rm {se}}$ $\chi^2_{\rm {gl}}$ 140—200 5 51 113 151 200—450 5 872 1613 2203 450—850 12 902 1748 3414 850—1210 12 334 651 2311 1210—1610 12 83 107 583 total 2242 4222 9262 : Results for $\chi^2$ from single-energy (se) and global (gl) solutions.[]{data-label="tab:chisq"} The best fit for the background and the resonance parameters yields the results listed in Tables \[tab:par\_backgr\] and \[tab:par\_res\], respectively. The PS-PV mixing parameter and the vector-meson coupling constants are defined as in Ref. [@Maid]. However we note that in the present version we do not use form factors at the hadronic vertices involving vector-meson exchange. $m_{V}$\[MeV\] $\lambda_V$ $\tilde g_{V1}$ $\tilde g_{V2}/\tilde g_{V1}$ -------------- ---------------- ------------- ----------------- ------------------------------- $\omega$ 783 0.314 16.3 -0.94 $\rho$ 770 0.103 1.8 12.7 $B= 0.71 fm$ : Masses and coupling constants for vector mesons, PS-PV mixing parameter $\Lambda_m$, and parameter $A$ for the low-energy correction of Eq. (\[eq:3.16\]).[]{data-label="tab:par_backgr"} ---------------- -------------- ------------------- --------------- ----------------- -------------- -------- ------- --------- ------- proton neutron $N^{\ast}$ $M_R$\[MeV\] $\Gamma_R$\[MeV\] $\beta_{\pi}$ $\phi_R$\[deg\] $X_R$\[MeV\] $n_E$ $n_M$ $n_E$ $n_M$ $P_{33}(1232)$ 1232 130 1.0 0.0 570 -1 2 -1 2 $P_{11}(1440)$ 1440 350 0.70 -15 470 — 0 — -1 $D_{13}(1520)$ 1530 130 0.60 32 500 3 4 7 2 $S_{11}(1535)$ 1535 100 0.40 8.2 500 2 — 2 — $S_{31}(1620)$ 1620 150 0.25 23 470 5 — 5 — $S_{11}(1650)$ 1690 100 0.85 7.0 500 4 — 4 — $D_{15}(1675)$ 1675 150 0.45 20 500 3 5 3 4 $F_{15}(1680)$ 1680 135 0.70 10 500 3 3 2 2 $D_{33}(1700)$ 1740 450 0.15 61 700 4 5 4 5 $P_{13}(1720)$ 1740 250 0.20 0.0 500 3 3 3 3 $F_{35}(1905)$ 1905 350 0.10 40 500 4 5 4 5 $P_{31}(1910)$ 1910 250 0.25 35 500 — 1 — 1 $F_{37}(1950)$ 1945 280 0.40 30 500 6 6 6 6 ---------------- -------------- ------------------- --------------- ----------------- -------------- -------- ------- --------- ------- In Tables \[tab:helicity\_p\] and \[tab:helicity\_n\] we compare the helicity amplitudes obtained from MAID2003 and MAID2007 with the results of the PDG [@PDG06] and GWU/SAID [@SAID02; @SAID06] analysis. As is very typical for a global analysis with about 20,000 data points fitted to a small set of 20-30 parameters, the fit errors appear unrealistically small. However, one should realize that these errors only reflect the statistical uncertainty of the experimental error, whereas the model uncertainty can be larger by an order of magnitude. We therefore do not list our fit errors, which in fact are very similar in the GW02 or GW06 fits of the SAID group [@SAID02; @SAID06]. The only realistic error estimate is obtained by comparing different analysis, such as SAID, MAID, and coupled-channels approaches.\ PDG GW06 2003 2007 ---------------- ----------- -------------- ---------------- ------ ------ $P_{33}(1232)$ $A_{1/2}$ -135$\pm$6 -139.1$\pm$3.6 -140 -140 $A_{3/2}$ -250$\pm$8 -257.6$\pm$4.6 -265 -265 $E2/M1$ (%) -2.5$\pm$0.5 -2.2 -2.2 $P_{11}(1440)$ $A_{1/2}$ -65 $\pm$4 -50.6 $\pm$1.9 -77 -61 $D_{13}(1520)$ $A_{1/2}$ -24 $\pm$9 -28.0 $\pm$1.9 -30 -27 $A_{3/2}$ 166 $\pm$5 143.1 $\pm$2.0 166 161 $S_{11}(1535)$ $A_{1/2}$ 90 $\pm$30 91.0 $\pm$2.2 73 66 $S_{31}(1620)$ $A_{1/2}$ 27 $\pm$11 49.6$\pm$2.2 71 66 $S_{11}(1650)$ $A_{1/2}$ 53$\pm $16 22.2 $\pm$7.2 32 33 $D_{15}(1675)$ $A_{1/2}$ 19 $\pm$8 18.0 $\pm$2.3 23 15 $A_{3/2}$ 15 $\pm$9 21.2 $\pm$1.4 24 22 $F_{15}(1680)$ $A_{1/2}$ -15 $\pm$6 -17.3 $\pm$1.4 -25 -25 $A_{3/2}$ 133 $\pm$12 133.6 $\pm$1.6 134 134 $D_{33}(1700)$ $A_{1/2}$ 104 $\pm$15 125.4 $\pm$3.0 135 226 $A_{3/2}$ 85 $\pm$22 105.0 $\pm$3.2 213 210 $P_{13}(1720)$ $A_{1/2}$ 18 $\pm$30 96.6 $\pm$3.4 55 73 $A_{3/2}$ -19 $\pm$20 -39.0 $\pm$3.2 -32 -11 $F_{35}(1905)$ $A_{1/2}$ 26 $\pm$11 21.3 $\pm$3.6 14 18 $A_{3/2}$ -45 $\pm$20 -45.6$\pm$4.7 -22 -28 $F_{37}(1950)$ $A_{1/2}$ -76 $\pm$12 -78 -94 $A_{3/2}$ -97 $\pm$10 -101 -121 : Proton helicity amplitudes at $Q^2=0$ for the major nucleon resonances, in units $10^{-3}$ GeV$^{-1/2}$. The results with MAID2003 and MAID2007 are compared to the PDG [@PDG06] and GWU/SAID [@SAID06] analysis.[]{data-label="tab:helicity_p"} PDG GW02 2003 2007 ---------------- ----------- -------------- ------------- ------ ------ $P_{11}(1440)$ $A_{1/2}$ 40$\pm$10 47$\pm$5 52 54 $D_{13}(1520)$ $A_{1/2}$ -59 $\pm$9 -67 $\pm$4 -85 -77 $A_{3/2}$ -139 $\pm$11 -112 $\pm$3 -148 -154 $S_{11}(1535)$ $A_{1/2}$ -46$\pm$27 -16$\pm$5 -42 -51 $S_{11}(1650)$ $A_{1/2}$ -15$\pm$21 -28$\pm$4 27 9 $D_{15}(1675)$ $A_{1/2}$ -43 $\pm$12 -50 $\pm$4 -61 -62 $A_{3/2}$ -58 $\pm$13 -71 $\pm$5 -74 -84 $F_{15}(1680)$ $A_{1/2}$ 29 $\pm$10 29 $\pm$6 25 28 $A_{3/2}$ -33 $\pm$9 -58 $\pm$9 -35 -38 $P_{13}(1720)$ $A_{1/2}$ 1 $\pm$15 17 -3 $A_{3/2}$ -29 $\pm$61 -75 -31 : Neutron helicity amplitudes at $Q^2=0$ for the major nucleon resonances. GW02 are the results GWU/SAID analysis [@SAID02]. Further notation as in Tab. \[tab:helicity\_p\].[]{data-label="tab:helicity_n"} Next we present our results for the multipoles starting with the threshold region. In Fig. \[fig:E0thr\] we demonstrate the effects of the low-energy correction and the cusp effect for $\pi^0$ photoproduction, as described by Eqs. (\[eq:3.16\]) and (\[eq:3.17\]), respectively. The prediction of MAID98 for $\pi^0$ photoproduction at threshold (dotted lines) lies substantially below the data. In accordance with Ref. [@DMTthr], the phenomenological term $E^{corr}_{0+}$ simulates the pion off-shell rescattering or pion-loop contributions of ChPT. The cusp term of Eq. (\[eq:3.17\]) describes the strong energy dependence near $\pi^+$ threshold, which has its origin in the pion mass difference and the strong coupling with the $\pi^+n$ channel. The figure shows that the off-shell pion rescattering substantially improves the agreement with the data. However, one problem still remains in the threshold region. The experimental photon asymmetry $\Sigma$ in $\pi^0$ photoproduction at $E_{\gamma}\approx 160$ MeV takes positive values, whereas the MAID results are negative in this region. As has been demonstrated in Refs. [@DMTthr; @HDT], this observable is very sensitive to the $M_{1-}$ multipole which strongly depends on the details of the low-energy behavior of Roper resonance, vector meson and off-shell pion rescattering contributions. Therefore, a slight modification of one or all of these mechanisms can drastically change the photon asymmetry.\ Figures \[fig:2\_photo\]-\[fig:4\_photo\] display the results for the most important $S$ and $P$ waves in the $\Delta$(1232) region. However, a look at Fig. \[fig:5\_photo\] shows that also the $D$-wave amplitudes $_pE_{2-}^{1/2}$, $E_{2-}^{3/2}$, and $_nE_{2-}^{1/2}$, give sizable contributions in this region, in particular through their real parts. In these figures, we present the MAID and SAID global (energy dependent) solutions, together with our local (single energy) fit obtained for energy bins of 10 MeV. In general the MAID and SAID results are close, which is not too surprising because the phases are constrained by the Fermi-Watson theorem. However, there are much larger differences in the $_pE_{0+}^{1/2}$ and $_pE_{2-}^{1/2}$ amplitudes, which indicates that the present data base is still too limited to determine these background amplitudes in a reliable way.\ More substantial discrepancies between the MAID and SAID analyses are found in the second and third resonance regions. A detailed comparison of the two models is shown in Figs. \[fig:6\_photo\]-\[fig:13\_photo\]. As pointed out in Sect. 3.1, it is prerequisite to know the phases of the multipoles in order to get correct single-energy solutions above the two-pion threshold. In MAID2007 these phases are determined by Eqs. (\[eq:3.9\]), (\[eq:3.15\]), and (\[eq:3.18\]). The SAID analysis is based on the following parametrization of the partial wave amplitudes: $$\begin{aligned} t_{\gamma\pi}\,&=&\,({\rm {Born}}+A)\,(1+i t_{\pi N}) + B\,t_{\pi N} \nonumber\\ & & + (C+iD)\,({\rm {Im}}\,t_{\pi N}\,-\mid t_{\pi N} \mid^2)\,, \label{eq:4.2}\end{aligned}$$ where $A$, $B$ $C$ and $D$ are polynomials in the energy with real coefficients, and $t_{\pi N}$ is the pion-nucleon elastic scattering amplitude of Eq. (\[eq:3.15\]). As seen in the first resonance region, the most serious differences between MAID and SAID are again found for the real parts of the multipoles $_pE_{0+}^{1/2}$ and $_pE_{2-}^{1/2}$. We have checked the phases of these multipoles by independent calculations on the basis of dispersion relations [@DR02; @aznauryan]. The result confirmed our phase relations. Concerning the small amplitudes, the most sizable differences between SAID and MAID are in the $M_{1-}^{3/2}$ $_pM_{1+}^{1/2}$ and $_pE_{1+}^{1/2}$ multipoles. In the neutron channel, the largest differences are in the multipoles $_nE_{0+}^{1/2}$, $_nE_{3-}^{1/2}$, $_nE_{1+}^{1/2}$, and $_nM_{1+}^{1/2}$. In the last two cases this is due to the large contribution from the $P_{13}(1720)$ resonance which is not found in the SAID analysis (see Table \[tab:helicity\_p\]).\ Let us finally discuss the possible contributions of the weaker resonances. As discussed in Ref. [@S11], the two additional $S_{11}$ resonances found with masses of about 1800 and 2000 MeV might also show up in pion photoproduction. This conclusion was mainly based on the single-energy solution of the SAID group. As illustrated by Fig. \[fig:14\_photo\], our present analysis requires only one additional $S_{11}$ resonance with mass $M_R\approx$ 1950 MeV, and our single-energy solution shows no resonance at $M_R\approx$ 1800 MeV. Of course, the solution of the problem is certainly correlated with the way how the resonance and background contributions are separated. As demonstrated both in Ref. [@S11] and by Fig. \[fig:6\_photo\], the background is very important for this particular channel. In this context we recall that we use the same form of the unitarized background contribution (Born, $\omega$ and $\rho$ exchange) for all the partial waves. Another interesting topic deserving further experimental and theoretical studies, concerns the Roper or $P_{11}$ channel. As clearly seen in Fig. \[fig:15\_photo\], both our and the SAID analysis yield a second resonance structure of the $_pM_{1-}^{1/2}$ multipole at $E_{\gamma}\approx$ 1070 MeV or $W\approx$ 1700 MeV. However, our analysis yields a very small width of $\Gamma_{\rm {tot}}\approx$ 30-60 MeV, whereas the PDG lists $\Gamma_{\rm {tot}}\approx$ 50-250 MeV for this resonance. Moreover, also the helicity amplitude differs, whereas our result is $A_{1/2}\approx -0.024$ GeV$^{-1/2}$, the PDG lists $0.009\pm 0.022$ GeV$^{-1/2}$. Of course, these numbers do strongly depend on the values for the single-pion branching ratio. On the other hand, we do not anticipate large effects from different definitions of the background in this channel, because the background contribution is very small in the resonance region (see Fig. \[fig:7\_photo\]). Partial-wave analysis of pion electroproduction =============================================== In most of the pion electroproduction experiments the five-fold differential cross section was measured. However, different conventions exist for the partial cross sections, and therefore we recall the definitions used in MAID. For an unpolarized target the cross sections written as the product of the virtual-photon flux factor $\Gamma_v$ and the virtual photon cross section $d\sigma_v/d\Omega_{\pi}$ [@DT92], $$\frac{d\sigma}{d\Omega_f^L\,dE_f^L\,d\Omega_{\pi}} = \Gamma_v\, \frac{d\sigma_v}{d\Omega_{\pi}}\,, \label{eq:5.1}$$ $$\begin{aligned} \frac{d\sigma_v}{d\Omega_{\pi}} & = & \frac{d\sigma_T}{d\Omega_{\pi}} + \epsilon \,\frac{d\sigma_L}{d\Omega_{\pi}} + \sqrt{2\epsilon (1+\epsilon)}\,\,\frac{d\sigma_{LT}}{d\Omega_{\pi}}\, \cos{\Phi_{\pi}}\label{eq:5.2} \\ &+& \epsilon\,\frac{d\sigma_{TT}}{d\Omega_{\pi}}\, \cos{2\Phi_{\pi}} + h\sqrt{2\epsilon (1-\epsilon)}\,\,\frac{d\sigma_{LT'}} {d\Omega_{\pi}}\,\sin{\Phi_{\pi}}\,,\nonumber\end{aligned}$$ where $\epsilon$ and $h$ describe the polarizations of the virtual photon and the electron, respectively. We further note that the hadronic kinematics is expressed in the c.m. system, whereas the electron and virtual photon kinematics is written in the lab frame, as indicated by $L$ in the following variables: the initial and final electron energies $E_i^L$ and $E_f^L$, respectively, the electron scattering angle $\theta_L$, the photon energy $\omega_L=E_i^L-E_f^L$, and the photon three-momentum ${\bf k}_L$. With these definitions the virtual photon flux and the transverse photon polarization take the form $$\epsilon=\frac {1} {1 + 2\frac{{\bf k}^2_L}{Q^2}\tan^2 \frac{\theta_L}{2}},\; \Gamma_v = \frac{\alpha_{\rm {em}}}{2\pi^2}\ \frac{E_f^L}{E_i^L}\,\frac{K}{Q^2}\ \frac{1}{1-\epsilon}\,. \label{eq:5.3}$$ As in our previous notation [@DT92], the flux is denoted by the photon “equivalent energy” in the lab frame, $K=K_H=(W^2-m^2)/2m$ as originally introduced by Hand [@Hand]. Another definition was given by Gilman [@Gilman] who used $K=K_G=\mid{{\bf k}_L}\mid$.\ The first two terms on the r.h.s. of Eq. (\[eq:5.2\]) are the transverse ($T$) and longitudinal ($L$) cross sections. They do not depend on the pion azimuthal angle $\Phi_{\pi}$. The third and fifth terms describe longitudinal-transverse interferences ($LT$, $LT'$). They contain an explicit factor $\sin\theta_{\pi}$ and therefore are vanishing along the axis of momentum transfer. The same is true for the fourth term, a transverse-transverse interference ($TT$) proportional to $\sin^2\theta_{\pi}$. It is useful to express these 5 cross sections in terms of hadronic response functions depending only on 3 independent variables, i.e., $R_i=R_i(Q^2,W,\theta_{\pi})$. The corresponding relations take the form $$\begin{aligned} \nonumber \frac{d\sigma_T}{d\Omega_{\pi}}=\frac{q}{k_W}\,R_T\,,\, \frac{d\sigma_{TT}}{d\Omega_{\pi}}=\frac{q}{k_W}\,R_{TT}\,,\, \frac{d\sigma_L}{d\Omega_{\pi}}=\frac{q}{k_W} \frac{Q^2}{\omega_{\gamma}^2}\,R_L\,,\end{aligned}$$ $$\frac{d\sigma_{LT}}{d\Omega_{\pi}}=\frac{q}{k_W}\, \frac{Q}{\omega_{\gamma}}\,R_{LT}\,,\, \frac{d\sigma_{LT'}}{d\Omega_{\pi}}=\frac{q}{k_W}\, \frac{Q}{\omega_{\gamma}}\,R_{LT'}\,. \label{eq:5.4}$$ As a result of this equation, the longitudinal ($L$) and longitudinal-transverse ($LT$ and $LT'$) response functions must be proportional to $\omega^2_{\gamma}$ and $\omega_{\gamma}$, respectively, in order to avoid non-physical singularities at the energy for which the c.m. virtual photon energy passes through zero. The 5 response functions may be expressed in terms of 6 independent CGLN amplitudes $F_1,...,F_6$ [@CGLN], or in terms of the helicity amplitudes $H_1,...,H_6$, which are linear combinations of the CGLN amplitudes. The relevant expressions can be found in Refs. [@DT92; @KDT].\ Data base for pion electroproduction and fit procedure ------------------------------------------------------ The main part of our data base for pion electroproduction includes the compilation of the GWU/SAID group [@SAID] in 2000 and recent data from Bonn and JLab (see Table \[database\]). Altogether this base contains about 70000 data points within the energy range 1.074 GeV$< W < 2~$GeV and photon virtuality range 0.1 GeV$^2 \leq Q^2 \leq $6 GeV$^2$. In addition we have analyzed high precision data from Bates [@Mer01; @Stave06], Mainz [@Pos01; @Elsner06], and JLab [@Kelly; @Lav04]. Our fitting procedure was as follows. In a first step we fitted the data sets at constant values of $Q^2$ (single-$Q^2$ fit). This procedure is similar to the partial-wave analysis for pion photoproduction except for the additional longitudinal couplings of the resonances. Second, we introduced a smooth $Q^2$ evolution of the e.m. transition form factors and parameterized the 3 helicity amplitudes accordingly. In a combined fit with the complete electroproduction data base and information from the single-$Q^2$ fits we finally constructed the $Q^2$-dependent solution (super-global fit). This new solution (MAID2007) was then compared with the previous solution (MAID2003) in terms of $\chi^2$ as presented in Table \[database\]. In most cases the new fit improves the description of the data, in particular for the $n\pi^+$ channel. -------------------- ----------------- ------------------ -------------------------------- Ref. $W$ (MeV) $N_{\rm {data}}$ $\chi^2/N_{\rm {data}}$ (2003) channel $Q^2$ (GeV$^2$) observables $\chi^2/N_{\rm {data}}$ (2007) SAID00 1074-1895 13152 3.238 $p\pi^0$ 0.1-4.3 d$\sigma$, ... 3.172 SAID00 1125-1975 5464 3.297 $n\pi^+$ 0.117-4.4 d$\sigma$, ... 4.188 Bonn02 [@Ban02] 1153-1312 4914 1.378 $p\pi^0$ 0.63 d$\sigma$ 1.400 CLAS02 [@Joo02] 1110-1680 31810 1.907 $p\pi^0$ 0.4-1.80 d$\sigma$ 1.952 CLAS03 [@Joo03] 1100-1660 223 4.881 $p\pi^0$ 0.4-0.65 d$\sigma_{LT'}$ 3.490 CLAS04 [@Joo04] 1100-1660 224 4.879 $n\pi^+$ 0.4-0.65 d$\sigma_{LT'}$ 2.196 CLAS06 [@Egi06] 1110-1570 4179 10.04 $n\pi^+$ 0.3-0.60 d$\sigma$ 4.954 CLAS06 [@Ungaro06] 1110-1390 8491 1.691 $p\pi^0$ 3.0-6.0 d$\sigma$ 1.335 total 1074-1975 68457 2.724 $p\pi^0$, $n\pi^+$ 0.1-6.0 d$\sigma$, ... 2.437 SAID00 1253-1976 799 2.100 $p\pi^-$ 0.54-1.36 d$\sigma$ 2.264 -------------------- ----------------- ------------------ -------------------------------- : \[database\] The number of data points, $N_{\rm {data}}$, and the $\chi^2$ value per data point obtained with MAID2003 and MAID2007. Results for the $\Delta(1232)$ form factors -------------------------------------------- In the literature the e.m. properties of the $N \Delta(1232)$ transition are described by either the magnetic ($G_M^*$), electric ($G_E^*$), and Coulomb ($G_C^*$) form factors or the helicity amplitudes $A_{1/2}$, $A_{3/2}$, and $S_{1/2}$, which can be derived from the reduced e.m. amplitudes $\bar{\mathcal A}_\alpha$ as defined by Eq. (\[eq:3.18\]). It is worthwhile pointing out that these amplitudes are related to the multipoles over the full energy region, that is, they are the primary target of the fitting procedure. The form factors and helicity amplitudes are then obtained by evaluating the reduced e.m. amplitudes at the resonance position $W=M_\Delta$=1232 MeV. The respective relations take the following form: $$\begin{aligned} G_M^*(Q^2) &=& -c_\Delta (A_{1/2}+\sqrt{3} A_{3/2})=2 c_\Delta \, \bar{\mathcal A}_M^\Delta(M_\Delta,Q^2)\,,\nonumber\\ G_E^*(Q^2) &=& \,\,\, c_\Delta (A_{1/2}-\frac{1}{\sqrt{3}} A_{3/2})=-2 c_\Delta \,\bar{\mathcal A}_E^\Delta(M_\Delta,Q^2)\,,\nonumber\\ G_C^*(Q^2) &=& \sqrt{2} c_\Delta \frac{2M_{\Delta}}{k_\Delta} S_{1/2}=-2 c_\Delta \, \frac{2M_{\Delta}}{k_\Delta}\bar{\mathcal A}_S^\Delta(M_\Delta,Q^2) \,,\nonumber\\ \mbox{with} \;\; c_\Delta &=& \left( \frac{m^3 k_W^\Delta}{4\pi\alpha_{\rm {em}} M_\Delta k_\Delta^2} \right)^{1/2}\,, \label{eq:5.5}\end{aligned}$$ and where $k_\Delta=k_\Delta (Q^2)=k(M_{\Delta},Q^2)$ and $k_W^\Delta=k(M_\Delta,0)$ are the virtual photon momentum and the photon equivalent energy at resonance. Because the $\Delta(1232)$ is very close to an ideal resonance, the real part of the amplitudes vanishes for $W=M_{\Delta}$ and the form factors can be directly expressed by the imaginary parts of the corresponding multipoles at the resonance position, $$\begin{aligned} G^{\ast}_M(Q^2) &=& b_\Delta\, {\rm {Im}}\{M_{1+}^{(3/2)}(M_\Delta,Q^2)\} \,,\nonumber\\ G^{\ast}_E(Q^2) &=& -b_\Delta\, {\rm {Im}}\{E_{1+}^{(3/2)}(M_\Delta,Q^2)\} \,,\label{eq:5.6}\\ G^{\ast}_C(Q^2) &=& -b_\Delta \frac{2M_{\Delta}}{k_{\Delta}}\, {\rm {Im}}\{S_{1+}^{(3/2)}(M_\Delta, Q^2)\}\,,\nonumber\\ \mbox{where}\;\; b_\Delta &=& \left( \frac{8\, m^2\, q_\Delta\, \Gamma_\Delta}{3\,\alpha_{\rm {em}}\, k_\Delta^2} \right)^{1/2} \,, \nonumber\end{aligned}$$ and with $\Gamma_{\Delta}=115$ MeV and $q_{\Delta}=q(M_{\Delta})$ the pion momentum at resonance. The above definition of the form factors is due to Ash [@Ash67]. The form factors of Jones and Scadron  [@Jon73] are obtained by multiplying our form factors with $\sqrt{1+Q^2/(M_N +M_\Delta)^2}$. We note that the form factor $G_C^{\ast}$ differs from our previous work [@Tiat03] by the factor $2M_\Delta/k_{\Delta}$ in Eq. (\[eq:5.6\]). With these definitions all 3 transition form factors remain finite at pseudo-threshold (Siegert limit). In the literature, the following ratios of multipoles have been defined: $$\begin{aligned} R_{EM} &=& -\frac{G_E^{\ast}}{G_M^{\ast}}=\frac{A_{1/2}-\frac{1}{\sqrt{3}} A_{3/2}}{A_{1/2}+\sqrt{3} A_{3/2}}\,,\label{eq:5.7}\\ R_{SM} &=& -\frac{k_\Delta}{2M_{\Delta}}\frac{G_C^{\ast}}{G_M^{\ast}}=\frac{\sqrt{2} S_{1/2}}{A_{1/2}+\sqrt{3} A_{3/2}}\,. \label{eq:5.8}\end{aligned}$$ In MAID2003 the $Q^2$ dependence of the e.m. amplitudes $\bar{\mathcal A}_\alpha^\Delta$ was parameterized as follows: $$\bar{\mathcal A}_\alpha^\Delta(W,Q^2)=A_\alpha^0 (1 + \beta_\alpha Q^{2n_\alpha}) \frac{ k}{k_W} e^{-\gamma_\alpha Q^2}G_D(Q^2), \label{eq:5.9}$$ where $G_D(Q^2)=1/(1+Q^2/0.71\,{\rm {GeV}}^2)^2$ is the dipole form factor. MAID2007 follows this parametrization for the magnetic and electric amplitudes, although with somewhat different values of the parameters (see Table \[tab:2\]). In order to fulfill the Siegert theorem, we have however changed the description of the Coulomb amplitude as specified below. M E S model ----------------- ------ -------- -------- ------- $A_\alpha^0$ 300 -6.50 -19.50 2003 300 -6.37 -12.40 2007 $\beta_\alpha$ 0 -0.306 0.017 2003 0.01 -0.021 0.12 2007 $\gamma_\alpha$ 0.21 0.21 0.21 2003 0.23 0.16 0.23 2007 $n_\alpha$ 1 1 3 2003 1 1 — 2007 : Parameters for the $N \Delta$ amplitudes given by Eqs. (\[eq:5.9\]) and (\[eq:5.16\]). The amplitudes $A_\alpha^0$ are in units 10$^{-3}$ GeV$^{-1/2}$, the parameters $\beta$ and $\gamma$ in GeV$^{-2}$. For the Coulomb amplitude in MAID2007 we use Eq. (\[eq:5.16\]) with $d$=4.9.[]{data-label="tab:2"} The results of MAID2003 and MAID2007 for $G_M^{\ast}(Q^2)$ are compared in Fig. \[fig:gmstar\]. Because our single-$Q^2$ analysis follows the global fit closely, it is not shown in the figure. We find an excellent agreement with the data, which also include the new high-$Q^2$ data of the JLab/CLAS Collaboration [@Ungaro06]. At this point a word of caution is in order. Because the form factors are extracted from the multipoles by Eq. (\[eq:5.6\]), they are proportional to $\sqrt {\Gamma_{\Delta}}$. The MAID fit to the experimental data yields $\Gamma_{\Delta}$=130 MeV, which is different from the usually assumed value of about 115 MeV. Therefore, in order to compare with form factor values of other analyses, we scale our predicted form factor with $\sqrt{115/130}$. As shown in Fig. \[fig:gmstar\], $G_M^{\ast}(0)/3$ will then take the usual value of 1 to an accuracy of 1%. From this number we can determine the $N\rightarrow\Delta$ magnetic transition moment, $\mu_{N\Delta}=3.46\pm 0.03$, in units of the nuclear magneton. Siegert theorem and ratios $R_{EM}$ and $R_{SM}$ ------------------------------------------------ Let us next discuss our results for the $R_{EM}$ and $R_{SM}$ ratios. In all previous solutions these ratios were nearly constant for $Q^2<1$ GeV$^2$. However, calculations in effective field theories [@Gel99; @Pas05] and dynamical models [@DMT; @KY99; @SL01] indicated a rapid rise of $R_{SM}$ for $Q^2 \rightarrow +0$. This dependence is rather model-independent, because it reflects the behavior of the multipoles at physical threshold (pion momentum ${\bf {q}} \rightarrow 0)$ and pseudothreshold (Siegert limit, photon momentum ${\bf {k}}\rightarrow 0)$ [@Dre06]. The longitudinal and Coulomb multipoles are related by gauge invariance, ${\bf {k}} \cdot \vec J = \omega_\gamma \rho$, which leads to $$|{\bf {k}}| \, L_{\ell \pm}^I (W, Q^2)= \omega_\gamma \, S_{\ell \pm}^I (W, Q^2)\,. \label{eq:5.10}$$ Since the photon c.m. energy $\omega_\gamma$ vanishes for $Q^2=Q^2_0=W^2-m^2$, the longitudinal multipole must have a zero at that momentum transfer, $L_{\ell \pm}^I(W,Q^2_0)=0$. Furthermore, gauge invariance implies that the longitudinal and Coulomb multipoles take the same value in the real photon limit, $L_{\ell \pm}^I(W,Q^2=0)=S_{\ell \pm}^I(W,Q^2=0)$. Finally, the multipoles obey the following model-independent relations at physical threshold (${\bf {q}}\rightarrow 0$) and pseudothreshold (${\bf {k}}\rightarrow 0$): $$\begin{aligned} (E_{\ell+}^I, L_{\ell+}^I) &\rightarrow& k^\ell q^\ell \;\; (\ell \ge 0)\, \nonumber\\ (M_{\ell+}^I, M_{\ell-}^I) &\rightarrow& k^\ell q^\ell\;\; (\ell \ge 1)\, \label{eq:5.11}\\ (L_{\ell-}^I) &\rightarrow& kq \;\;\;\;\; (\ell = 1) \nonumber\\ (E_{\ell-}^I, L_{\ell-}^I) &\rightarrow& k^{\ell-2} q^\ell\;\; (\ell \ge 2)\ \,. \nonumber\end{aligned}$$ According to Eq. (\[eq:5.10\]) the Coulomb amplitudes acquire an additional factor $k$ at pseudothreshold, i.e., $S_{\ell \pm}^I \sim k L_{\ell \pm}^I$. This limit is reached at $Q^2=Q_{\rm{pt}}^2=-(W-m)^2$ (pseudo-threshold), and because no direction is defined for ${\bf {k}} = 0$, the electric and longitudinal multipoles are no longer independent at this point, $$E_{\ell+}^I/L_{\ell+}^I \rightarrow 1 \;\; \mbox{and} \;\; E_{\ell-}^I/ L_{\ell-}^I \rightarrow -\ell/(\ell-1)\;\; \mbox{if} \;\; k \rightarrow 0. \label{eq:5.12}$$ In the case of the $N \Delta$ multipoles, Eq. (\[eq:5.12\]) yields the following relation in the limit ${\bf {k}}\rightarrow 0$: $L_{1+}^{3/2} \rightarrow E_{1+}^{3/2} \rightarrow {\mathcal O}(k)$ and consequently $S_{1+}^{3/2}= kE_{1+}^{3/2}/\omega_\gamma\rightarrow {\mathcal O}(k^2)$. Although the pseudo-threshold is reached at the unphysical point $Q_{\rm{pt}}^2=-(M_{\Delta}-m)^2 \approx -0.084$ GeV$^2$, it still influences the multipoles near $Q^2=0$ because of the relatively small excitation energy of the $\Delta(1232)$. In particular we get the following relation for $Q^2\rightarrow Q_{\rm{pt}}^2$: $$R_{SM}=\frac{S_{1+}^{(3/2)}}{M_{1+}^{(3/2)}} =\frac{k}{\omega_\gamma}\frac{E_{1+}^{(3/2)}}{M_{1+}^{(3/2)}} \rightarrow \frac{k}{M_\Delta-m}R_{EM}\,. \label{eq:5.13}$$ With increasing value of $Q^2$, the Siegert relation fails to describe the experimental data. Moreover, it contains a singularity at $\omega_\gamma$=0, which occurs in $\Delta(1232)$ electroproduction already at $Q^2=0.64$ GeV$^2$. However, we obtain a good overall description by using the idea of Ref. [@Buchmann] that the ratio $R_{SM}$ is related to the (elastic) form factors of the neutron, $$R_{SM}(Q^2)=\frac{m \, k_\Delta(Q^2)\,G_E^n(Q^2)} {2\, Q^2 \, G_M^n(Q^2)}\,. \label{eq:5.14}$$ This relation gives the necessary proportionality to the photon momentum at small $Q^2$, describes the experimental value of the ratio over a wide range of $Q^2$, and yields an asymptotic behavior consistent with the prediction of perturbative QCD that $R_{SM}$ should approach a constant for $Q^2\rightarrow\infty$. This leads to the following simple parametrization: $$R_{SM}(Q^2)=-\frac{k_\Delta(Q^2)}{8m}\,\frac{a}{1+d\tau}\,, \label{eq:5.15}$$ with $\tau=Q^2/(4m^2)$, and the parameters $a$ and $d$ to be determined by a fit to the data. On the basis of this ansatz, the Coulomb coupling has been modified as follows: $$\begin{aligned} \bar{\mathcal A}_S^\Delta(W,Q^2) =A_S^0 \frac{1 + \beta_S Q^2}{1+d\tau} \frac{k^2}{k_W\,k_W^\Delta} e^{-\gamma_S Q^2}G_D(Q^2), \label{eq:5.16}\end{aligned}$$ with parameters given in Table \[tab:2\]. This leads to the multipole ratio $$R_{SM}(Q^2)=\frac{A_S^0}{A_M^0}\,\frac{1}{1+d\tau}\left(\frac{1+\beta_S Q^2} {1+\beta_M Q^2}\right)\,\frac{k_\Delta}{k_W^\Delta}\,. \label{eq:5.17}$$ By construction this ratio vanishes in the Siegert limit, $Q^2\rightarrow Q_{\rm{pt}}^2$, and approaches a (negative) constant for $Q^2\rightarrow \infty$ in agreement with perturbative QCD. However, a word of caution has to be added at this point. The polynomials and gaussians used to fit the data in the range of low and intermediate virtualities, $Q^2<10$ GeV$^2$, should not be expected to yield realistic extrapolations to the higher values of $Q^2$.\ The correct Siegert limit is even more important for pion $S$-wave production in the threshold region, in which case the pseudo-threshold comes as close as $Q^2_{\rm{pt}}=-m_{\pi}^2\approx -0.02$ GeV$^2$. The term describing the pion cloud contribution has therefore been parameterized as follows: $$L_{0+}^{\rm {corr}}(W,Q^2)=\frac{\omega_\gamma}{\omega_{\rm{pt}}}\, e^{-\beta(Q^2-Q_{pt}^2)} \,E_{0+}^{\rm {corr}}(W,Q^2)\,, \label{eq:5.18}$$ where $\omega_{\rm{pt}}^2=-Q_{\rm{pt}}^2=(W-m)^2$. From a fit to $\pi^0$ electroproduction data near threshold [@Wei07], we obtain $\beta=10$ GeV$^{-2}$. In the future we intend to study pion electroproduction near threshold in more detail.\ Figures \[fig:emratios\] and \[fig:smratios\] display the super-global solutions of MAID2003 (dashed lines) and MAID2007 (solid lines) for the ratios $R_{EM}$ and $R_{SM}$ in comparison with other analyses. Different from our previous solution, the ratio $R_{EM}$ of MAID2007 stays always below the zero line, in agreement with the original analysis of the data [@Ungaro06; @Fro99] and also with the dynamical model of Sato and Lee [@SL01] who concluded that $R_{EM}$ remains negative and tends towards more negative values with increasing $Q^2$. This indicates that the predicted helicity conservation at the quark level is irrelevant for the present experiments. We also analyzed the new data of Ref. [@Ungaro06] in the range of 3 GeV$^2\leq Q^2 \leq 6$ GeV$^2$ and found slightly decreasing values of $R_{EM}$ from our single-$Q^2$ analysis. In this analysis we varied both the $\Delta$ and the Roper multipoles. For the ratio $R_{SM}$ both the super-global and the single-$Q^2$ solutions yield ratios that asymptotically tend to a negative constant. This result is in good agreement with the prediction of Ref. [@Buchmann] (dash-dotted curve in Fig. \[fig:smratios\]) but disagrees with our previous solution and with the analysis of Ref. [@Ungaro06]. As discussed before, the new solution has a large slope at small $Q^2$ as a consequence of the Siegert theorem. The following Fig. \[helicity\_P33\] displays the $Q^2$ dependence of the helicity amplitudes for the $N \Delta(1232)$ transition. Our single-$Q^2$ fit is in excellent agreement with the super-global solution, except for the values of $S_{1/2}$ at $Q^2=0.4$ and 0.525 GeV$^2$. Results for the higher resonances --------------------------------- Above the two-pion threshold we can no longer apply the two-channel unitarity and consequently the Watson theorem does not hold. Therefore, the background amplitude of the partial waves does not vanish at resonance as was the case for the $\Delta(1232)$ resonance. As an immediate consequence the resonance-background separation becomes more model-dependent. In MAID2007 we choose to separate the background and resonance contributions according to the K-matrix approximation. Furthermore, we recall that the absolute values of the helicity amplitudes are correlated with the values used for the total resonance width $\Gamma_R$ and the single-pion branching ratio $\beta_\pi$. On the experimental side, the data at the higher energies are no longer as abundant as in the $\Delta$ region. However, the large data set recently obtained by the CLAS collaboration (see Table \[database\]) enabled us to determine the transverse and longitudinal helicity couplings as functions of $Q^2$ for all the 4-star resonances below 1700 MeV. These data are available in the kinematical region of $1100~\mbox{MeV}<W<1680~\mbox{MeV}$ and $0.4~\mbox{GeV}^2<Q^2<1.8~\mbox{GeV}^2$.\ The helicity amplitudes for the Roper resonance $P_{11}(1440)$ are shown in Fig. \[helicity\_P11\]. Our latest super-global solution (solid lines) is in reasonable agreement with the single-$Q^2$ analysis. The figure shows a zero crossing of the transverse helicity amplitude at $Q^2\approx 0.7$ GeV$^2$ and a maximum at the relatively large momentum transfer $Q^2\approx 2.5$ GeV$^2$. The longitudinal Roper excitation rises to large values around $Q^2\approx 0.5$ GeV$^2$ and in fact produces the strongest longitudinal amplitude we can find in our analysis. This answers the question raised by Li and Burkert [@Burk92] whether the Roper resonance is a radially excited 3-quark state or a quark-gluon hybrid, because in the latter case the longitudinal coupling should vanish completely. From the global fit we find the following parametrization for the $Q^2$ dependence of the Roper amplitudes for the proton and neutron channels: $$\begin{aligned} A_{1/2}^p(Q^2) &=& A_{1/2}^{0,p}(1 -1.22\, Q^2-0.55 \,Q^8)\, e^{-1.51 Q^2}\,,\nonumber\\ S_{1/2}^p(Q^2) &=& S_{1/2}^{0,p}\,(1 + 40\, Q^2 +1.5\,Q^8)\,e^{-1.75Q^2}\,, \label{eq:5.19}\\ A_{1/2}^n(Q^2) &=& A_{1/2}^{0,n}(1 +0.95\, Q^2)\,e^{-1.77 Q^2}\,,\nonumber\\ S_{1/2}^n(Q^2) &=& S_{1/2}^{0,n} \,(1 + 2.98\, Q^2)\,e^{-1.55 Q^2}\,, \label{eq:5.20}\end{aligned}$$ where $Q^2$ should be inserted in units of GeV$^2$. The numerical values of the helicity amplitudes for real photons are given in Table \[tab:3\]. At $Q^2$=0 the fit yields a large neutron value for the Coulomb amplitude $S_{1/2}^0$, but with increasing $Q^2$ the proton and neutron amplitudes become comparable.\ ---------------- -------- ------------- -- -------- ------------- -- $A_{1/2}^0$ $S_{1/2}^0$ proton neutron proton neutron $P_{11}(1440)$ -61.4 54.1 4.2 -41.5 ---------------- -------- ------------- -- -------- ------------- -- : Helicity amplitudes for the $P_{11}(1440)$ resonance at $Q^2$=0 in units 10$^{-3}$ GeV$^{-1/2}$.[]{data-label="tab:3"} For all the higher resonances the transverse and longitudinal helicity amplitudes are simply parameterized by the form $$A_{\lambda}(Q^2) = A_\lambda^0\,(1 +\alpha\, Q^2) e^{- \beta Q^2}\,. \label{eq:5.21}$$ The values of the fit parameters $A_\lambda^0$, $\alpha$ and $\beta$ are listed in Tables \[tab:4\] and \[tab:5\]. In the following Fig. \[helicity\_S11\] we present the results for the $S_{11}(1535)$. As is also known from $\eta$ electroproduction, the transverse form factor falls off very slowly. At a virtuality of $Q^2\approx 3$ GeV$^2$ this resonance is much stronger than the $\Delta (1232)$ or the $D_{13} (1520)$ and only comparable to the Roper. However, due to its much smaller width as compared to the Roper, the $S_{11}$ dominates over the Roper at large $Q^2$. This result is in agreement with the inclusive electroproduction cross section on the proton, which clearly shows the dominance of the $\Delta(1232)$ at small momentum transfer whereas at the larger momentum transfers the second resonance region takes over.\ In Fig. \[aznar\] we compare our results to those of Aznauryan *et al.* [@Azna05] who used a similar set of the CLAS data in the second resonance region. Our super-global solutions (solid lines) agree generally quite well with the JLab-Yerevan analysis, which was performed with both an isobar model and dispersion analysis. ---------------- ----------------- ----------------- ----------------- ------------- $A_{1/2}$ $A_{3/2}$ $S_{1/2}$ $S_{1/2}^0$ proton $\alpha$$\beta$ $\alpha$$\beta$ $\alpha$$\beta$ $D_{13}(1520)$ 7.771.09 0.692.10 4.193.40 -63.6 $S_{11}(1535)$ 1.610.70 —— 23.90.81 -2.0 $S_{31}(1620)$ 1.862.50 —— 2.832.00 16.2 $S_{11}(1650)$ 1.450.62 —— 2.880.76 -3.5 $D_{15}(1675)$ 0.102.00 0.102.00 0.000.00 0.00 $F_{15}(1680)$ 3.981.20 1.002.22 3.141.68 -44.0 $D_{33}(1700)$ 1.911.77 1.972.20 0.000.00 0.00 $P_{13}(1720)$ 1.891.55 16.01.55 2.461.55 -53.0 ---------------- ----------------- ----------------- ----------------- ------------- : The proton parameters for the higher resonances: $\alpha$ and $\beta$ as defined by Eq. (\[eq:5.21\]), in units GeV$^{-2}$, and $S_{1/2}^0$, the longitudinal amplitude at $Q^2=0$, in units 10$^{-3}$ GeV$^{-1/2}$. The values for the transverse amplitudes $A_{1/2,3/2}^0$ are determined by the real photon physics and listed in Table \[tab:helicity\_p\].[]{data-label="tab:4"} ---------------- ----------------- ----------------- ----------------- ------------- $A_{1/2}$ $A_{3/2}$ $S_{1/2}$ $S_{1/2}^0$ neutron $\alpha$$\beta$ $\alpha$$\beta$ $\alpha$$\beta$ $D_{13}(1520)$ -0.531.55 0.581.75 15.71.57 13.6 $S_{11}(1535)$ 4.751.69 —— 0.361.55 28.5 $S_{11}(1650)$ 0.131.55 —— -0.501.55 10.1 $D_{15}(1675)$ 0.012.00 0.012.00 0.000.00 0.00 $F_{15}(1680)$ 0.001.20 4.091.75 0.000.00 0.00 $P_{13}(1720)$ 12.71.55 4.991.55 0.000.00 0.00 ---------------- ----------------- ----------------- ----------------- ------------- : The neutron parameters for the higher resonances. The values for the transverse amplitudes $A_{1/2,3/2}^0$ are given in Table \[tab:helicity\_n\]. Further notation as in Table \[tab:4\].[]{data-label="tab:5"} The following Fig. \[helicity\_D13\_F15\] displays our super-global and single-$Q^2$ fits for the $D_{13}(1520)$ and $F_{15}(1680)$ resonances. The figure demonstrates that (I) the helicity non-conserving amplitude $A_{3/2}$ dominates for real photons and (II) with increasing values of $Q^2$, $A_{3/2}$ drops faster than the helicity conserving amplitude $A_{1/2}$. As a consequence the asymmetry $${\mathcal {A}}(Q^2)=\frac{\mid A_{1/2} \mid^2 - \mid A_{3/2} \mid^2} {\mid A_{1/2} \mid^2 + \mid A_{3/2} \mid^2} \label{eq:5.22}$$ changes rapidly from values close to $-1$ to values near $+1$ over a small $Q^2$ range. As is seen in Fig. \[fig:asym\], the asymmetry crosses the zero line at $Q^2\approx 0.5$ GeV$^2$ for the $D_{13}(1520)$ resonance and at $Q^2\approx 0.8$ GeV$^2$ for the $F_{15}(1680)$. As a comparison, the asymmetry $A$ for the $\Delta(1232)$ resonance is practically constant over this $Q^2$ range with a value $\approx -0.5$. This again shows the special role of the $\Delta$ resonance, where the helicity conservation is not observed. Conclusion ========== Using the world data base of pion photo- and electroproduction and recent data from Bates/MIT, ELSA/Bonn, MAMI/Mainz, and Jefferson Lab, we have extracted the longitudinal and transverse helicity amplitudes of nucleon resonance excitation for all the 4-star resonances below $W=2$ GeV. For this purpose we have extended our unitary isobar model MAID and parameterized the $Q^2$ dependence of the transition amplitudes. The comparison between such super-global solutions with the corresponding single-$Q^2$ fits gives us confidence in the obtained helicity couplings for the $P_{33}(1232)$, $P_{11}(1440)$, $S_{11}(1535)$, $D_{13}(1520)$, and $F_{15}(1680)$ resonances, even though the model uncertainty is still quite large, particularly for the longitudinal amplitudes.\ For the higher 4-star and all 3-star resonances the situation is less clear. This deplorable situation reflects the fact that a model-independent analysis requires precision data over a large kinematical range. In some cases double-polarization experiments will be helpful, as has already been shown for pion photoproduction. Furthermore, charged pion electroproduction data are needed with the same quantity and quality as for neutral pions, in order to resolve the ambiguities in the isospin structure, in particular for the $S_{11}$ and $S_{31}$ resonances. While we have mostly discussed the electroproduction from proton targets, also the existing neutron data have been analyzed. The latter are of course less abundant, and moreover no new neutron data have been reported over the recent years. Because the isospin symmetry is most likely on safe grounds in the resonance region, only the electromagnetic neutron couplings with isospin $1/2$ are still lacking. In spite of the discussed problems, we have implemented a super-global solution also for the neutron amplitudes in MAID07.\ Pion photo- and electroproduction are invaluable tools to study the resonance structure of the nucleon. With the advent of the new c.w. electron accelerators, new precision experiments have afforded a host of new data to unravel this structure in the first and second resonance regions. In particular, electroproduction has provided new insights in the spatial distribution of the nucleon-resonance transition densities. However, in order to get the whole picture, several challenges remain: - [Dedicated experiments to investigate the higher energy region, which have to include an intense study of the polarization degrees of freedom. Experience has shown that even the physics of the $\Delta~(1232)$ requires a full-fledged program to measure the spin observables in order to understand the background of the non-resonating multipoles.]{} - [A fresh approach to also determine the excitation spectrum of the neutron. As an example, the comparison of the Roper or $P_{11}~(1440)$ helicity amplitudes for proton and neutron will shed light on the structure of this enigmatic resonance.]{} - [The open question of the excitation spectrum in the third resonance region and above deserves further studies in both theory and experiment. This includes “missing” and “exotic”, e.g., 5-quark resonances as well as more mundane second and third resonances in a multipole, which show up in a particular analysis and not in another one.]{} In conclusion we hope that MAID2007, just as other approaches based on partial-wave analysis, dynamic models, coupled-channels calculations, and dispersion theory, will contribute to settle the mentioned issues and thus to improve our still somewhat vestigial knowledge of the nucleon’s resonance structure.\ **Acknowledgment** This work was supported by the Deutsche Forschungsgemeinschaft through the SFB 443, by the joint project NSC/DFG 446 TAI113/10/0-3 and by the joint Russian-German Heisenberg-Landau program.\ G. Höhler [*et al.*]{}, *Handbook of Pion-Nucleon Scattering*, Physics Data 12-1 (Karlsruhe, 1979). W.-M. Yao [*et al.*]{} \[Particle Data Group\], J. Phys. G [**33**]{}, 1 (2006). For an overview and further references see S. Boffi, C. Giusti, F.D. Pacati, and M. Radici, *Electromagnetic Response of Atomic Nuclei*, Clarendon Press, Oxford (1996) p. 114 pp. R. Beck *et al.*, Phys. Rev. Lett. [**78**]{}, 606 (1997). J.J. Kelly *et al.*, Phys. Rev. Lett. [**95**]{}, 102001 (2005) and Phys. Rev. C [**75**]{}, 025201 (2007). D. Drechsel, O. Hanstein, S.S. Kamalov, and L. Tiator, Nucl. Phys. A [**645**]{}, 145 (1999); http://www.kph.uni-mainz.de/MAID/. R.A. Arndt, I.I. Strakovsky, and R.L.  Workman, Phys. Rev. C [**53**]{}, 430 (1996) (SP97 solution of the VPI analysis). V. Burkert, Proc. of Topical Workshop on “Excited Baryons 1988”, Troy, New York (1988), eds. G. Adams, N. Mukhopadhyay, and P. Stoler, World Scientific, Singapore (1989) p. 122. S.S. Kamalov, D. Drechsel, L. Tiator, and S.N. Yang, Proc. of Workshop on “The Physics of Excited Nucleons”, Mainz, Germany (2001), eds. D. Drechsel and L. Tiator, P. World Scientific, Singapore (2001), p. 197. S.S. Kamalov, S.N. Yang, D. Drechsel, O. Hanstein, L. Tiator, Phys. Rev. C [**64**]{}, 032201 (2001); http://www.kph.uni-mainz.de/MAID/DMT/. J.J. Kelly, Phys. Rev. C [**70**]{}, 068202 (2004). L. Tiator and S.S. Kamalov, Proc. of Workshop on “The Physics of Excited Nucleons”, Tallahassee, FL, USA (2005), eds. S. Capstick, V. Crede and P. Eugenio, World Scientific, Singapore (2006), p. 16. M. Ungaro *et al.*, Phys. Rev. Lett. C [**90**]{}, 112003 (2006). S.N. Yang, J. Phys. G [**11**]{}, L205 (1985). S.S. Kamalov, and S.N. Yang, Phys. Rev. Lett. [**83**]{}, 4494 (1999). R.A. Arndt, I.I. Strakovsky, and R.L. Workman, Phys. Rev. C [**53**]{}, 430 (1996) (SP99 solution of the GWI/SAID analysis). S.S. Kamalov, G.Y. Chen, S.N. Yang, D. Drechsel, and L. Tiator, Phys. Lett. B [**522**]{}, 27 (2001). J.M. Laget, Phys. Rep. [**69**]{}, 1 (1981). V. Bernard, N. Kaiser, and Ulf-G. Mei[ß]{}ner, Phys. Lett. B [**268**]{}, 291 (1991); Z. Phys. C [**70**]{}, 483 (1996). R. Leukel, PhD thesis, Mainz (2001), http://wwwa2.kph.uni-mainz.de/A2/. J. Ahrens *et al.*, Phys. Rev. Lett. [**87**]{}, 022003 (2001) and [**88**]{}, 232002 (2002). I. Preobrajenski, PhD thesis, Mainz (2001), http://wwwa2.kph.uni-mainz.de/A2/. J. Ahrens, *et al.*, Phys. Rev. C [**87**]{}, 045204 (2006). O. Bartalini *et al.*, Phys. Lett. B [**554**]{}, 113 (2002). O. Bartalini *et al.*, Eur. Phys. J, A [**26**]{}, 399 (2005). A. Shafi *et al.* Phys. Rev. C [**70**]{}, 035204 (2004). O. Bartholomy *et al.*, Phys. Rev. Lett. [**94**]{}, 012003 (2005). R.A. Arndt, W.J. Briscoe, I.I. Strakovsky, and R.L. Workman, Phys. Rev. C [**66**]{}, 055213 (2002); http://gwdac.phys.gwu.edu/. M. Dugger *et al.* Phys. Rev. C [**76**]{}, 025211 (2007) J.C. Bergstrom *et al.*, Phys. Rev. C [**53**]{}, R1052 (1996) and [**55**]{}, 2016 (1997). A. Schmidt *et al.*, Phys. Rev. Lett. [**87**]{}, 232501 (2001). O. Hanstein, D. Drechsel, and L. Tiator, Phys. Lett. B [**385**]{}, 45 (1996) and Nucl. Phys. A [**632**]{}, 561 (1998). S.S. Kamalov, L. Tiator, D. Drechsel, R.A. Arndt, C. Bennhold, I.I. Strakovsky, and R.L. Workman, Phys. Rev. C [**66**]{}, 065206 (2002). I.G. Aznauryan, Phys. Rev. C [**67**]{}, 015209 (2003). G.Y. Chen, S. Kamalov, S.N. Yang, D. Drechsel, and L. Tiator, Nucl. Phys. A [**723**]{}, 447 (2003). D. Drechsel and L. Tiator, J. Phys. G: Nucl. Phys. [**18**]{}, 449 (1992). L.H. Hand, Phys. Rev. [**129**]{}, 1834 (1963). F.J. Gilman, Phys. Rev. [**167**]{}, 1365 (1968). G.F. Chew, M.L. Goldberger, F.E. Low, and Y. Nambu, Phys. Rev, [**106**]{}, 1345 (1957). G. Knöchlein, D. Drechsel, and L. Tiator, Z. Phys. A [**352**]{}, 327 (1995). R.A. Arndt, W.J. Briscoe, I.I. Strakovsky, and R.L. Workman, http://gwdac.phys.gwu.edu/. T. Bantes and R. Gothe, private communication. K. Joo *et al.*, Phys. Rev. Lett. [**88**]{}, 122001-1 (2002). K. Joo *et al.*, Phys. Rev. C [**68**]{}, 032201 (2003). K. Joo *et al.*, Phys. Rev. C [**70**]{}, 042201 (2004). H. Egiyan *et al.*, Phys. Rev. C [**73**]{}, 025204 (2006). C. Mertz *et al.*, Phys. Rev. Lett. [**86**]{}, 2963 (2001). S. Stave *et al.*, Eur. Phys. J. A [**30**]{}, 471 (2006). Th. Pospischil *et al.*, Phys. Rev. Lett. [**86**]{}, 2959 (2001). D. Elsner *et al.*, Eur. Phys. J. A [**27**]{}, 91 (2006). G. Laveissiere *et al.*, Phys. Rev. C [**69**]{}, 045202 (2004). W.W. Ash et al., Phys. Lett. B [**24**]{}, 165 (1967). H.F. Jones and M.D. Scadron, Annals Phys. [**81**]{}, 1 (1973). L. Tiator, D. Drechsel, S.S. Kamalov, and S.N. Yang, Eur. Phys. J. A [**17**]{}, 357 (2003). G.C. Gellas, T.R. Hemmert, C.N. Ktorides, and G.I. Poulis, Phys. Rev. D **60**, 054022 (1999); T.A. Gail and T.R. Hemmert, Eur. Phys. J. A [**28**]{}, 91 (2006). V. Pascalutsa and M. Vanderhaeghen, Phys. Rev. Lett. **95**, 232001 (2005) and Phys. Rev. D **73**, 034003 (2005). T. Sato and T.S.H. Lee, Phys. Rev. C [**63**]{}, 055201 (2001). D. Drechsel and L. Tiator, AIP Conf. Proc. [**904**]{} 129 (2007). A.J. Buchmann, Phys. Rev. Lett. [**93**]{}, 212301-1 (2004). R.A. Arndt, R.L. Workman, Z. Li and L.D. Roper, Phys. Rev. C [**42**]{}, 1864 (1990). M. Weis *et al.* (A1 Collaboration), arXiv:0705.3816 \[nucl-ex\]. V.V. Frolov *et al.*, Phys. Rev. Lett. [**82**]{}, 45 (1999). Z. Li, V. Burkert, and Z. Li, Phys. Rev. D [**46**]{}, 70 (1992). I. Aznauryan, V. Burkert, H. Egiyan, K. Joo, R. Minehart and L.C. Smith, Phys. Rev. C [**71**]{}, 015201 (2005).
{ "pile_set_name": "ArXiv" }
--- abstract: 'Quantum cosmology has traditionally been studied at the level of symmetry-reduced minisuperspace models, analyzing the behavior of wave functions. However, in the absence of a complete full setting of quantum gravity and detailed knowledge of specific properties of quantum states, it remained difficult to make testable predictions. For quantum cosmology to be part of empirical science, it must allow for a systematic framework in which corrections to well-tested classical equations can be derived, with any ambiguities and ignorance sufficiently parameterized. As in particle and condensed-matter physics, a successful viewpoint is one of effective theories, adapted to specific issues one encounters in quantum cosmology. This review presents such an effective framework of quantum cosmology, taking into account, among other things, space-time structures, covariance, the problem of time and the anomaly issue.[^1]' --- [Quantum Cosmology: Effective Theory]{}\ Martin Bojowald[^2]\ Institute for Gravitation and the Cosmos, The Pennsylvania State University,\ 104 Davey Lab, University Park, PA 16802, USA\ Introduction ============ If quantum cosmology is ever to be part of empirical science, it must be described by a good effective theory. There is no hope of exactly solving its equations in realistic models or to tame conceptual quantum issues made even more severe in the context of cosmology. The derivation of testable predictions requires systematic approximations at semiclassical order and beyond, and by experience with other areas of physics, effective actions or equations are the best available tools. While the speculative part of quantum cosmology, addressing for instance the Planck regime or the status of multiverses, requires all subtleties of quantum physics to be considered — such as choices of Hilbert spaces, self-adjointness properties of Hamiltonians or unitarity of evolution, an understanding of deep conceptual issues and the measurement problem — physics can and must proceed without all these problems being solved.[^3] Compare this situation for instance with quantum field theory, for which no rigorous interacting and non-integrable version is known. And yet, at the effective level it is the key tool behind the success of elementary particle physics. Given the immensity of the Planck scale, potentially observable effects in quantum cosmology are realized at low energies where semiclassical quantum gravity, with the first few orders in $\hbar$ taken into account, suffices. This feature makes effective theory in quantum gravity and cosmology even more powerful than in other settings [@EffectiveGR; @BurgessLivRev]. As we will see in the course of this review, somewhat surprisingly, even conceptual problems of quantum gravity can advantageously be addressed with effective methods, especially with an extension to effective constraints. Given the amount of research on effective theories and their applications, one may think that deriving an effective theory of quantum cosmology is a simple and well-understood problem. However, this is not at all the case. Quite general and powerful techniques of effective actions or potentials do exist, employed with great success in particle physics and condensed-matter physics alike. Quantum cosmology, on the other hand, is unique by virtue of several features. It requires aspects of effective equations not encountered in other fields, related for instance to the prevalence of canonical methods, the generally covariant setting lacking evolution by a unique time parameter, or the absence or inapplicability of non-perturbative ground states or other distinguished classes of states. These technical problems will be discussed in due course. For now, as a motivation of our detailed look into effective theory, we state the following two general problems by which quantum cosmology differs from other fields. First, quantum cosmology is much like condensed-matter physics, with microscopic quantum degrees of freedom manifesting themselves on length scales far larger than their own. While the precise nature of microscopic degrees of freedom (strings, loops, …) remains unclear and disputed among the different approaches, their presence in some form is widely agreed upon. By considering large structures or space-time regions made from many building blocks, quantum cosmology differs significantly from most situations encountered in elementary-particle physics, where events with a comparatively small number of particles in excited states close to the vacuum are studied. Quantum cosmology is a many-body problem, a situation in which it is difficult to derive and justify valid effective descriptions. The effective view has been put to good use in condensed-matter physics, but only thanks to rich and merciless experimental input to weed out wrong ideas and stimulate new successful ones. Quantum cosmology is not (yet?) subject to experimental pressure, and many (good and bad) ideas are sprouting. Effective cosmological theory must be able to stand on its own, requiring a systematic and rigorous formulation taking into account all features and consistency conditions to be imposed in quantum gravity. As the second problem, we observe that the theoretical foundation of quantum cosmology is much weaker than that of condensed-matter physics to which it is otherwise quite close. We know well which Hamiltonian we should use to find all states and energies of excitations in a crystal, but mathematically the problem is challenging and calls for the approximations of effective theory. In quantum cosmology, we don’t even know which precise Hamiltonian or other underlying object to use for the dynamics of a universe. Even if we choose one particular approach to quantum gravity, its mathematical objects or its specializations to cosmology are incompletely known or understood, opening wide the door for ambiguities and spurious constructions. We need a well-understood theory to pinpoint places where best to look for observational effects, and we need observations to guide our theoretical constructions. Quantum cosmology, with an incompletely understood theory and no current observations, is a slippery subject, depriving us of a good handle to grasp its implications. In the absence of experiments, we can only rely on conceptual arguments and internal mathematical consistency conditions which, however, come along with their own problems. In this situation, effective techniques have proven to be one of the few reliable approaches for physical evaluations of the theory, allowing one to include all crucial quantum effects in equations with clear physical meaning, and to take into account ambiguities by sufficiently general parameterizations. This general framework and its current status in the context of quantum cosmology are the topics of this review.[^4] Cosmological consistency ======================== One source of technical and conceptual problems in quantum cosmology and quantum gravity in general is that the theory deals with relativistic space-time, in the absence of a unique Hamiltonian to generate evolution in time. Instead, many choices for time and corresponding Hamiltonians or evolution equations are possible, and they must all lead to the same physics. While this invariance is guaranteed classically, it implies a complicated problem after quantization, presenting the strongest set of consistency conditions to restrict possible choices of quantum cosmologies. Unfortunately, these issues are often set aside in research on quantum cosmology and even quantum gravity, owing to their complicated nature. This intentional oversight implies a large number of ambiguities, fixed in those contexts only by ad-hoc constructions. To keep this review focused, we will not discuss the rather large body of works in such directions, for instance those crucially using the distinction of a time variable by gauge-fixing or deparameterization in canonical settings, and only mention shortcomings in contexts in which they become apparent. Generally speaking, results derived with a distinguished choice of time (or gauge) cannot be considered physically reliable unless one can make sure that they do not depend on one’s choice of time. In this section, we will discuss the main features that an effective theory of quantum cosmology must deal with, which includes covariance and state properties. The former ensures independence of choices of time, the latter deals with additional freedom in quantum theories. By these considerations, we will be guided toward suitable ingredients for the mathematical formulation of effective theory. Covariance {#s:Cova} ---------- Effective equations of quantum cosmology are supposed to modify some cosmological version of Einstein’s equation by quantum corrections. In most cases, such a cosmological version is a reduction to isotropic Friedmann–Lemaître–Robertson–Walker space-times or homogeneous Bianchi models (minisuperspaces), a restriction to some specific form of inhomogeneous degrees of freedom such as Lemaître–Tolman–Bondi or Gowdy geometries (midisuperspaces), or an inclusion of unrestricted but perturbative inhomogeneity around a background in the former classes of models. ### Homogeneous models and automatic consistency In homogeneous models, the dynamics is completely determined by one equation for gravitational degrees of freedom (such as the Friedmann equation) and one for matter (such as the continuity equation). The Friedmann equation of isotropic models, $$\label{Friedmann} \left(\frac{\dot{a}}{a}\right)^2 +\frac{k}{a^2}= \frac{8\pi G}{3}\rho\,,$$ depends only on first-order derivatives and is therefore a constraint, to be satisfied by initial values of second-order equations of motion. When interpreted as a constraint (the Hamiltonian constraint), it is usually written in the form of an energy-balance law: $$\label{constraint} H:= -\frac{3}{8\pi G}(\dot{a}^2a+ka)+ E_{\rm matter}=0$$ with the matter energy $E_{\rm matter}=\rho a^3$ contained in some region of unit coordinate volume. (See [@Springer] for a detailed discussion of coordinate factors when the volume is not fixed, especially in the context of quantization.) In this form, the Hamiltonian constraint of gravity is obtained by varying the action by the lapse function $N=\sqrt{-g_{00}}$. (For more on constraints and canonical gravity, see [@CUP].) Given the Friedmann equation and the continuity equation $$\dot{\rho}+3\frac{\dot{a}}{a}(\rho+P)=0\,,$$ with pressure $P$, one can derive a second-order equation of motion by taking a time derivative of (\[Friedmann\]) and eliminating $\dot{\rho}$: the Raychaudhuri equation $$\frac{\ddot{a}}{a} = -\frac{4\pi G}{3} (\rho+3P)\,.$$ At this stage, we have all equations expected from the components of the isotropic Einstein tensor: the time-time component providing the Friedmann equation, the (identical) diagonal components of the spatial part amounting to the Raychaudhuri equation, and all off-diagonal components vanishing identically. The equations obtained are automatically consistent with each other: By construction, the time derivative of the Friedmann equation vanishes if the Raychaudhuri equation holds. Therefore, if the constraint imposed by the Friedmann equation holds for initial values at some time, it holds at all times. This latter property is realized for a large class of systems describing versions of isotropic (or homogeneous) cosmology, not just for the classical one resulting from Einstein’s equation. As a phase-space function, $H(a,p_a)$ in (\[constraint\]), with the momentum $p_a= -3(4\pi G)^{-1} a\dot{a}$ as it follows from the variation $\partial L_{\rm grav}^{\rm iso}/\partial \dot a $ of the Einstein–Hilbert action reduced to isotropy, plays the role of the Hamiltonian generating all equations of motion in proper time. The Raychaudhuri equation indeed follows from the Hamiltonian equations of motion $\dot{a}=\partial H/\partial p_a=\{a,H\}$ and $\dot{p}_a=-\partial H/\partial a=\{p_a,H\}$ (the first of which is identical to the definition of the momentum $p_a$). The matter terms $\rho$ and $P$ are realized by $$\rho=\frac{E_{\rm matter}}{a^3}\quad\mbox{ and }\quad P=-\frac{1}{3a^2}\frac{\partial E_{\rm matter}}{\partial a}\,,$$ the negative change of energy by volume change. Matter equations of motion follow once $E_{\rm matter}$ in (\[constraint\]) is expressed in terms of canonical degrees of freedom. The phase-space function $H(a,p_a)$ itself evolves according to the same Hamiltonian law, $\dot{H}(a,p_a)=\{H(a,p_a),H\}=0$ and is automatically constant in time, no matter what form $H(a,p_a)$ has. In homogeneous situations, the single Hamiltonian constraint that determines evolution is automatically preserved and consistent with evolution equations. We can easily modify $H$ by any form of quantum corrections without encountering consistency problems, issuing a powerful license to cosmological model builders. Consistency remains valid when we consider different choices of time. So far, the equations were in proper time $\tau$. All other choices $t$ in homogeneous models are related to $\tau$ by $\tau(t)=\int^{t} N(t'){\rm d}t'$, with a lapse function $N$ that enters Hamiltonian equations as well: If ${\rm d}f(a,p_a)/{\rm d}\tau=\{f,H\}$, we have $$\frac{{\rm d}f(a,p_a)}{{\rm d} t} = \frac{{\rm d}\tau}{{\rm d}t}\frac{{\rm d}f(a,p_a)}{{\rm d}\tau} =N\{f,H\}\approx \{f,N H\}$$ for any other time with ${\rm d}\tau/{\rm d}t=N$. In the last step, we are allowed to pull the lapse function $N$ inside the Poisson bracket, keeping the equality satisfied as a “weak” one, one that is valid provided the constraint $H=0$ holds. Applying the general law to the constraint itself, we again observe consistency: ${\rm d}H/{\rm d}t= \{H,NH\}\approx 0$. The Hamiltonian constraint generates not only evolution with respect to a given time choice, but also the transition between different choices as a gauge transformation. Infinitesimally, with $N$ close to one, we have $\tau= t+\epsilon$ with $\epsilon=\int (N-1){\rm d}t$, and for any function $f$, $$\delta_{\epsilon}f= f(\tau)-f(t)=\epsilon \frac{{\rm d}f}{{\rm d}t} =\epsilon \{f,H\}\approx \{f,\epsilon H\}\,.$$ The Friedmann equation, amounting to the Hamiltonian constraint, is automatically invariant under changes of time, and so are its solutions. Also the evolution equations are invariant if we use the Jacobi identity for Poisson brackets: $$\begin{aligned} \left\{\frac{{\rm d}f}{{\rm d}t},\epsilon H\right\}&=& \{\{f,NH\},\epsilon H\}= \{\{f,\epsilon H\},NH\}+ \{f,\{NH,\epsilon H\}\}\\ &=& \frac{{\rm d}\{f,\epsilon H\}}{{\rm d}t}- \{f,({\rm d}\epsilon/{\rm d}t) H\}+ \{f,(\delta_{\epsilon}N)H\}\,.\end{aligned}$$ Rearranging, we see that the gauge-transformed $f$ evolves in agreement with the gauge-transformed ${\rm d}f/{\rm d}t$, with a correction taking into account a possible gauge transformation of $N$ and time dependence of $\epsilon$. The presence of a single Hamiltonian constraint therefore ensures dynamical consistency and invariance, even if quantum modifications occur. For this reason, homogeneous minisuperspace models are a simple and popular tool to investigate possible consequences of quantum gravity and cosmology. But for the very same reason, extreme care must be exercised when such models are used for physical predictions: The trivialization of consistency conditions does not hold in a more general context, and therefore spurious results can easily be produced in their absence. ### Inhomogeneity and covariance Compared with minisuperspace models, inhomogeneous cosmology presents a very different situation regarding consistency, even if inhomogeneity is small and treated perturbatively. For gravity and matter to fit together in Einstein’s equation or a quantum modification thereof, a version of the contracted Bianchi identity must hold. But this identity, relating different types of dynamical equations, is easily destroyed if quantum corrections, for instance those found in minisuperspace models, are inserted blindly. When amending inhomogeneous equations by quantum corrections, one must face the problem of anomaly freedom or covariance. Certain relations between different dynamical equations and gauge generators, classically implemented by the contracted Bianchi identity, must be preserved in the presence of quantum corrections. This well-known classical fact is often overlooked in quantum treatments, especially those making use of gauge fixing or deparameterization. Gauge generators disappear when the gauge is fixed. A simple but illegitimate way out of difficult consistency problems is therefore to fix the gauge before quantum corrections are inserted. However, one then dispenses with ways to check consistency and cannot be sure that results obtained are physically viable. Most cases in which cosmological perturbations have been computed by gauge fixing in loop quantum cosmology, for instance, have by now been shown to be incorrect; important effects such as signature change have been overlooked. See also the instructive discussion of [@GaugeInvTransPlanck] in this context. We will come back to this issue later and for now continue with a classical discussion to provide more details. The first implication of the Bianchi identity is the presence of constraints. If we write $\nabla_{\nu}G^{\nu}_{\mu}=0$ in the form $$\label{ContractedSplit} \partial_0G^0_{\mu}= -\partial_aG^a_{\mu}-\Gamma^{\nu}_{\nu\kappa}G^{\kappa}_{\mu}+ \Gamma^{\kappa}_{\nu\mu}G^{\nu}_{\kappa}\,,$$ with spatial indices “$a$,” it becomes evident that the components $G^0_{\mu}$ of the Einstein tensor cannot contain second-order time derivatives: On the right-hand side, all factors in the three terms are at most second order in time, and there is one explicit time derivative on the left-hand side, leaving only the option of first time derivatives in $G^0_{\mu}$. These components of the Einstein tensor (minus $8\pi G$ times the corresponding stress-energy components $T_{\mu}^0$ if there is matter) are constraints on initial values, while the remaining components provide evolution equations. In contrast to minisuperspace models, we are dealing with a larger constrained system of four independent and functional constraints, the Hamiltonian constraint $H=G^0_0-N^aG_a^0$ and the diffeomorphism constraint $D=NG^0_a$, which are to be imposed pointwise or for all possible multiplier functions $N$ and $N^a$ in $$\label{Constraints} H[N]=-\int{\rm d}^3x N(G^0_0-N^aG^0_a)\quad\mbox{ and }\quad D[N^a]= -\int{\rm d}^3x N^a N G^0_a\,.$$ We are dealing with an infinite number of constraints. (To define these integrations, one introduces a foliation of space-time into spatial surfaces $t={\rm const}$, also used to set up canonical variables. The time direction $t^a$ at each point, used to define time derivatives in evolution equations, may be different from the normal direction $n^a$ to spatial slices in space-time, a freedom parameterized as $t^a=Nn^a+N^a$ with the lapse function $N$ and the shift vector field $N^a$. The linear combinations of $G^0_0$ and $G^0_a$ in (\[Constraints\]), depending on the shift $N^a$ and lapse $N$, take into account that constraints refer to directions normal and tangential to spatial slices, not to coordinate directions such as the zero-index of the Einstein tensor for a component along the time-evolution vector field $t^a$.) As before, for consistency the constraints must always hold provided they are imposed for initial values, a feature that is guaranteed by the Bianchi identity as well. If we combine the Einstein tensor and the stress-energy tensor $T_{\mu\nu}$ of matter in (\[ContractedSplit\]), we see that $\partial_0(G^0_{\mu}- 8\pi GT^0_{\mu})$ vanishes at any time provided the constraints $G_{\mu}^0-8\pi G T_{\mu}^0$ themselves (and therefore their spatial derivatives) vanish at that time and the evolution equations hold. By virtue of the contracted Bianchi identity, the constraints are consistent with evolution. Finally, as a third consequence, we see that all equations, Einstein’s equation and the contracted Bianchi identity as a consistency condition, are covariant and independent of coordinates used. Therefore, solutions will be covariant. All equations hold irrespective of the choice of coordinates, a well-known feature in perturbative cosmology which allows one to express all equations explicitly in terms of gauge-invariant variables [@Bardeen; @CosmoPert]. As in our discussion of minisuperspace models, a canonical view is useful to analyze which consistency conditions are satisfied automatically and which ones are non-trivial. There is a Hamiltonian constraint, and therefore Hamiltonian equations $\dot{f}=\{f,C\}$ are generated by a constraint $C$. However, in the inhomogeneous context, the constraint is not unique (up to a pre-factor $N$), nor is time evolution. We have a much larger choice of possible time variables to generate evolution. The most general version of equations of motion is obtained if we use all our constraints in a linear combination, defining $H[N,N^a]:= H[N]+D[N^a]$. For fixed $N$ and $N^a$, the Hamiltonian flow $\dot{f}=\{f,H[N,N^a]\}$ is then equivalent to Lie derivatives $\dot{f}={\cal L}_tf$ along the time-evolution vector field $t^a=Nn^a+N^a$ in space-time, foliated by spatial slices with unit normals $n^a$. For all constraints to be preserved by the evolution equations they generate, we need $\{H[M,M^a],H[N,N^a]\}\approx 0$ for all $M$, $N$, $M^a$ and $N^a$. If this condition is satisfied, the constraints are said to form a first-class system. Unlike in homogeneous models, where $H[M]$ always commutes weakly with $H[N]$, the general condition is highly non-trivial and provides strong restrictions on consistent modifications of Einstein’s equation. Quantum corrections can no longer be inserted at will. The same constraints that generate evolution provide gauge transformations, classically equivalent to coordinate changes. We use the same general combination as before, $H[\epsilon,\epsilon^a]$, but interpret the multipliers $\epsilon$ and $\epsilon^a$ differently, not related to a time-evolution vector field. Instead, the gauge transformation $\delta_{\epsilon^{\mu}}f= \{f,H[\epsilon]+D[\epsilon^a]\}$ with the classical constraints is equivalent to a coordinate transformation or the Lie derivative ${\cal L}_{\xi}f$ along the space-time vector field $\xi^{\mu}$ with components such that $\epsilon=N\xi^0$ and $\epsilon^a=\xi^a+N^a\xi^0$ [@LapseGauge]. The factors of $N$ and $N^a$ again result because space-time coordinate changes and the components of $\xi^{\mu}$ refer to coordinate directions, while constraints refer to directions normal and tangential to spatial slices with normal $n^a=N^{-1}(t^a-N^a)$. Also regarding gauge invariance, the condition of a first-class constraint algebra is then sufficient for consistency: In this case, all constraints are gauge invariant, $\delta_{\epsilon}H[N,N^a]=\{H[N,N^a],H[\epsilon,\epsilon^a]\}\approx 0$, and so are the evolution equations they generate. With this full set of gauge transformations, one can freely change the constant-time spatial surfaces used to define canonical variables and to integrate the constraints (\[Constraints\]). We are then dealing with a covariant theory of space-time, not just with a theory on a fixed spatial foliation. ### Hypersurface-deformation algebra The crucial consistency condition for any classical constrained system, its quantization, or an effective theory thereof, is therefore that it be first class: all constraints $H[N,N^a]$ must have Poisson brackets that vanish when the constraints are imposed, $$\{H[M,M^a],H[N,N^a]\}\approx 0\,.$$ For gravity, or any generally covariant space-time theory, the specific form is the hypersurface-deformation algebra [@DiracHamGR] $$\begin{aligned} \{D[M^a],D[N^a]\}&=& D[{\cal L}_{N^b}M^a] \label{DD} \\ \{H[M],D[N^a]\}&=& H[{\cal L}_{N^b}M] \label{HD}\\ \{H[M],H[N]\}&=& D[q^{ab}(M\nabla_bN-N\nabla_bM)] \label{HH}\end{aligned}$$ with the spatial metric $q^{ab}$. For a consistent quantization, an algebra of this form must be realized with commutators for constraint operators instead of Poisson brackets, and an effective constrained system must have quantum-corrected constraints such that a first-class algebra holds with Poisson brackets. If this is realized, no gauge transformations are broken by quantization and the quantum or effective theory is called anomaly-free. If this condition is satisfied, all consistency conditions that are classically implied by the contracted Bianchi identity hold in the presence of quantum corrections, and the (quantum) theory describes space-time rather than just a family of spatial slices. This property may be achieved with exactly the same form of the algebra, or with one that shows quantum corrections not just in the constraints but also in the structure functions of the algebra, as long as two constraints still commute up to another constraint. One universal example found in loop quantum gravity, as the most prominent result regarding hypersurface deformations with quantum corrections, has (\[DD\]) and (\[HD\]) unchanged, but (\[HH\]) modified to [@ConstraintAlgebra] $$\{H[M],H[N]\}= D[\beta q^{ab}(M\nabla_bN-N\nabla_bM)] \label{HHbeta}$$ with some phase-space function $\beta$. If the classical hypersurface-deformation algebra is modified, gauge transformations no longer correspond to Lie derivatives by space-time vector fields. Not just the dynamics but even the structure of space-time may be modified by quantum effects. We will discuss specific examples and results in later parts of this review. By analyzing quantum or effective constraints and their algebra, one can draw conclusions about quantum space-time structures. For instance, once the full hypersurface-deformation algebra is known, one can specialize it to Poincaré transformations by using linear $N$ and $N^a$. With $N(x)=\Delta t+v x$, for instance, we have a combination of a time translation by $\Delta t$ and a boost by $v$. With $\beta\not=1$ in (\[HHbeta\]), the usual Poincaré relations are modified. Although this may look like a version of deformed special relativity [@DSR; @DSR2; @Rainbow], there is no direct relation: In deformed special relativity, one has non-linear realizations of the Poincaré algebra, with structure constants depending on the algebra generators. In (\[HHbeta\]), we have corrections of structure functions depending on phase-space degrees of freedom, not directly on the algebra generators $H[N]$ and $D[N^a]$. A deformed version of special relativity would require a relation between phase-space variables, such as extrinsic curvature, and some space-time generators, such as energy. Relations of this form do exist in some regimes, for instance in asymptotically flat ones using the ADM energy, but not in general. ### Consistent and inconsistent quantum cosmology In general terms, the problem of canonical quantum cosmology can be formulated as follows: Find quantum corrections in $H[N,N^a]$, perhaps motivated by some full theory of quantum gravity, such that these constraints remain first class. This statement presents a well-defined mathematical problem of classifying deformed algebras. For every consistent version that may exist, we can compute and analyze equations of motion by standard means, as explicitly written out above in the general classical case. At the present stage, several examples of consistent deformations of constraints remaining first class are known, mainly from loop quantum gravity, but there is no general classification of these infinite-dimensional algebras. As already mentioned, there are attempts to shortcut through the difficult calculation of these algebraic structures by fixing the gauge before quantum corrections or other modifications are put in. If the gauge is fixed, one would no longer consider $H[\epsilon,\epsilon^a]$ as gauge generators, but only use $H[N,N^a]$ as constraints and to generate evolution. Some consistency conditions still need to be satisfied because the constraints are to be preserved by evolution, but this can usually be achieved more easily than in the non-gauge fixed case in which more fields are present. However, even if formally consistent versions of preserved constraints can be found in this way, such as those in [@ScalarHolEv], they are not guaranteed to be consistent because only a subclass of the constraint algebra can be tested when some modes are eliminated beforehand. Even if these are classical gauge modes, some consistency conditions and physical effects are overlooked. Moreover, the procedure is intrinsically inconsistent because one would first fix the gauge as it is determined by the classical constraints, and then proceed to modify the constraints that generate the gauge. A consistent scheme could be obtained only when the modified gauge structure is taken into account from the very beginning, but for that one would have to know a consistent version of non-gauge fixed constraints, not just of the gauge-fixed ones. Finally, having fixed the gauge, there is no way of calculating general gauge-invariant observables. Also here, one could only refer to the classical invariant variables, whose form however must be modified when quantum corrections are put into the constraints. Note that quantum space-time structures and modified constraints in most cases imply departures from classical manifold pictures, as we will see explicitly in the examples provided later. One can no longer refer to the usual form of coordinate transformations to compute gauge-invariant variables without using the constraints explicitly. In modified space-times, the constraints are the only means to compute gauge flows and invariants, but this can be done only if the gauge has not been fixed. Algebraic conditions can be simplified even more when gauge fixing is combined with deparameterization. With the latter procedure, one chooses a phase-space degree of freedom, for instance from matter, to rewrite constraint equations as relational evolution equations with respect to this variable. Not just gauge transformations but even constraints then disappear from the system, and no strong consistency conditions remain. A popular example is the coupling of a free, massless scalar field $\phi$, whose homogeneous mode $\bar{\phi}$ in an expansion $\phi=\bar{\phi}+\delta\phi$ around some background can be treated as a global time function. The canonical scalar Hamiltonian $$\begin{aligned} H_{\rm scalar}&=& \frac{1}{2}\int{\rm d}^3x \left(\frac{p_{\phi}^2}{\sqrt{\det q}}+ \sqrt{\det q} q^{ab}(\partial_a\phi)(\partial_b\phi)\right)\\ &=& \frac{1}{2} \left(\frac{\bar{p}_{\phi}^2}{\sqrt{\det \bar{q}}}+ \int{\rm d}^3x \delta(\cdots)\right)\end{aligned}$$ with the momentum $p_{\phi}=\bar{p}_{\phi}+\delta p_{\phi}$ then has a purely kinetic background term and is completely independent of $\bar{\phi}$. The momentum $\bar{p}_{\phi}$ is therefore a constant of motion and never becomes zero unless it vanishes identically. The background scalar $\bar{\phi}$ has no turning points and is monotonic, serving as a global internal time along classical trajectories. The Hamiltonian generating evolution with respect to this variable is obtained by solving the Hamiltonian constraint $H_{\rm gravity}+H_{\rm scalar}=0$ (with fixed $N$ as part of the gauge choice) for $\bar{p}_{\phi}=H_{\bar{\phi}}(q_{ab},p^{ab},\delta \phi,\delta p_{\phi})$: we have $\dot{\bar{\phi}}=\{\bar{\phi},H_{\bar{\phi}}\}=1$ and $\dot{\bar{p}_{\phi}}=\{\bar{p}_{\phi},H_{\bar{\phi}}\}=0$, consistent with evolution with respect to $\bar{\phi}$ where the dot stands for ${\rm d}/{\rm d}\bar{\phi}$, as well as Hamiltonian equations for the remaining variables, such as $\dot{q}_{ab}=\{q_{ab},H_{\bar{\phi}}\}$ and $\dot{\delta}\phi=\{\delta\phi,H_{\bar{\phi}}\}$. There is just one Hamiltonian generating all evolution equations, instead of a set of infinitely many constraints. Almost as in minisuperspace models one can then implement in $H_{\bar{\phi}}$ any quantum corrections one may desire. (In the context of cosmology, this approach has been suggested for instance in [@Hybrid] in an analysis of Gowdy models and possible quantizations.) Deparameterization is a powerful mathematical tool to derive properties of physical Hilbert spaces, for which no general method exists in the absence of deparameterizability. (See also Section \[s:EffCons\].) In quantum cosmology, this method has been applied in [@Blyth], and used recently to derive several aspects of self-adjointness and unitarity of evolution [@SelfAdFlat; @NonSelfAd; @DensityOp]. But it cannot serve as a valid procedure to evade the anomaly problem and do physical evaluations. To start with, most realistic systems are not globally deparameterizable, with one variable having no turning points at all. Moreover, also this procedure, on its own, cannot weed out physical inconsistencies in spite of its formal consistency. One has the same drawbacks as in the gauge-fixed approach, and on top of that one has distinguished (or even introduced) one degree of freedom as time. For physical consistency, one should then show that results for observables do not depend on one’s choice of time after quantization, but no systematic procedures exist to this end, constituting part of the problem of time in quantum gravity [@KucharTime; @Isham:Time; @AndersonTime]. (See [@ReducedKasner] for a discussion in the context of deparameterization or reduced phase-space quantization.) And even if one thinks that the distinguished time $\bar{\phi}$ should be sufficient for all physical purposes and does not worry about relating results in different times, the fact that there is no analog of the Bianchi identity to constrain modifications should arouse suspicion. Gauge fixing and deparameterization are not necessarily bad, but they provide reliable results only when they are used [*after*]{} a consistent version of quantum-corrected constraints has been found. When this is the case, it is clear that all equations are consistent and gauge covariant, and instead of computing complete gauge-invariant variables for the corrected constraints, one may well pick a gauge or choose an internal time and work out physical implications. There is also a possibility of providing consistent results with gauge fixing or deparameterization before quantization, but then one would have to show that all possible gauge fixings or deparameterizations would lead to the same physical observables. This requirement can be achieved in cases of simple constraints, such as the Gauss constraint of gravity in triad or connection variables as described at the end of Section \[s:SphSymm\], but presents a much more involved problem for the complicated Hamiltonian constraint. Problems and physical inconsistencies arise always when the gauge or time is chosen according to the classical system, and modifications of the constraints and thereby gauge transformations are inserted later. States {#s:States} ------ Covariance or analogs of the contracted Bianchi identity severely constrain consistent versions of quantum corrections, raising the manifold and metric structures underlying space-time in general relativity to the effective or quantum level. Although space-time structures and their covariance principles may then differ significantly from well-known classical ones, there is still a consistent dynamical theory independent of coordinates and gauge choices, and the set of all equations has meaningful solutions available for well-defined predictions. In addition to these restrictions on quantum corrections based on gauge aspects, the quantum theory of gravitational degrees of freedom itself should tell us what form of modifications we can have. ### Moments Effective equations, in general terms, describe quantum evolution by a smaller, more manageable number of degrees of freedom compared with the full quantum theory considered. When we go from classical physics to quantum physics, every degree of freedom we have at first is replaced by infinitely many parameters, for instance values the whole wave function takes at all points in configuration space. It is difficult to deal with values of wave functions in physical terms. Another parameterization, more convenient for effective equations as it turns out, is the set of expectation values of basic operators, $\langle\hat{a}\rangle$ and $\langle\hat{p}_a\rangle$ in Wheeler–DeWitt quantum cosmology, together with fluctuations and higher moments $$\Delta(a^bp_a^c):= \langle (\hat{a}-\langle\hat{a}\rangle)^b (\hat{p}_a-\langle\hat{p}_a\rangle)^c\rangle_{\rm Weyl}$$ with operators in totally symmetric, or Weyl, ordering. For generic states, all these moments with integer $b$ and $c$ such that $b+c\geq 2$, are independent of one another and of the expectation values, and therefore provide infinitely many degrees of freedom. (For $b+c=2$, for instance, we have the two fluctuations $\Delta(a^2)=(\Delta a)^2$ and $\Delta(p_a^2)=(\Delta p_a)^2$, and the covariance $\Delta(ap_a)= \frac{1}{2}\langle\hat{a}\hat{p}_a+ \hat{p}_a\hat{a}\rangle- \langle\hat{a}\rangle \langle\hat{p}_a\rangle$.) Any set of effective equations provides a description between the classical limit and the full quantum theory of the system considered, making use of finitely many degrees of freedom for each classical one. There may be additional degrees of freedom compared with the classical theory, but not infinitely many more. By the new degrees of freedom included and their coupling to expectation values — the variables with a direct classical analog — quantum corrections result. This is our general definition of effective theories, which we distinguish from coarse-grained theories. The latter may remove some quantum degrees of freedom by integrating them out, but typically leave whole towers of moments intact and therefore present infinitely many variables for some classical degrees of freedom. The behavior of the moments computed in simple unsqueezed Gaussian states with fluctuation parameter $\sigma$ [@HigherMoments], $$\label{Gaussian} \Delta(a^bp_a^c)= 2^{-(b+c)} \hbar^c\sigma^{b-c}\frac{b!\,c!}{(b/2)! (c/2)!} \quad{\rm if}\, b \,{\rm and}\, c\, {\rm are}\,{\rm even}$$ and $\Delta(a^bp_a^c)=0$ otherwise, suggest a natural organization of effective theories of different degrees. A Gaussian state saturates the uncertainty relation $$\label{Uncert} (\Delta a)^2(\Delta p_a)^2- \Delta(ap_a)^2\geq \frac{\hbar^2}{4}$$ and, for position and momentum fluctuations of the same order, has $\sigma=\Delta a=O(\hbar^{1/2})$. The product $\hbar^c\sigma^{b-c}\sim O(\hbar^{(b+c)/2})$ then shows that moments of order $n=b+c$ are of the order $n/2$ in $\hbar$. We use this observation to generalize the notion of semiclassical states from simple Gaussians to a much larger class: If the moments of a state behave as $\Delta(a^bp_a^c)\sim O(\hbar^{(b+c)/2})$ (said to obey a $\hbar$ierarchy), the state is called semiclassical. Moments of higher orders then affect only terms of high orders in $\hbar$ and can be neglected in an approximation. Ignoring all moments of order higher than some $n$, only finitely many quantum degrees of freedom are left, and we obtain an effective theory. With $n+1$ moments of order $n$, an effective theory with moments up to order $n$ has $\sum_{k=2}^{n}(k+1)=\frac{1}{2}n^2+\frac{3}{2}n-2$ state parameters in addition to the two basic expectation values. ### Dynamics and quantum back-reaction Effective solutions, describing the evolution of a quantum state, depend not only on classical variables but also on the quantum state used, specified for instance by initial values of its moments. In quantum cosmology, one first promotes the constraint (\[constraint\]) expressed in canonical variables $a$ and $p_a$ as $$H(a,p_a)= -\frac{2\pi G}{3} \frac{p_a^2}{a}+E_{\rm matter}$$ to an operator $\hat{H}$, choosing some ordering of $\hat{a}$ and $\hat{p}_a$. Dirac quantization then implements the classical constraint equation $H(a,p_a)=0$ by the condition $\hat{H}|\psi\rangle=0$ on physical states. In particular, the expectation value $\langle\hat{H}\rangle$ must vanish in all physical states. If we express this equation as a functional equation on the space of expectation values and moments parameterizing states, we obtain an expression such as $$\label{Qconstraint} \langle\hat{H}\rangle= -\frac{2\pi G}{3} \left(\frac{\langle\hat{p}_a\rangle^2}{\langle\hat{a}\rangle}+ \frac{(\Delta p_a)^2}{\langle\hat{a}\rangle}+\cdots \right)+ \langle\hat{E}_{\rm matter}\rangle$$ where we have used $\langle\hat{p}_a^2\rangle= \langle\hat{p}_a\rangle^2+ (\Delta p_a)^2$ and the dots indicate additional terms that contain the covariance of $a$ and $p_a$ as well as higher moments, and depend on the specific ordering chosen. If we take into account only expectation values in $\langle\hat{H}\rangle=0$, the classical Hamiltonian and the classical constraint surface are obtained. With fluctuations and higher moments, quantum corrections result that couple moments to expectation values and change the classical constraint surface. A systematic derivation and analysis of such moment terms, starting with expectation values of Hamiltonians and constraints, is the key ingredient of effective theories. The moments couple to expectation values, thereby providing quantum corrections to the classical motion, and their values therefore enter effective equations. A complete effective theory cannot leave the moments in equations as unknowns and instead provides evolution equations or other conditions for them as well. But some freedom always remains, for instance in the initial values chosen for evolution equations of moments. If effective theories are formulated for expectation values without including additional degrees of freedom such as moments to describe the evolution of classes of states, the specific states must either be restricted to provide unique effective equations, or be parameterized for a more general set. Often, such a state dependence enters effective theories only implicitly, for instance in the unique-looking low-energy effective action [@EffAcQM] free of any state parameters. This effective action describes low-energy effects of states near the vacuum of the interacting theory. In this way, the class of states is specified, certainly a rather small set compared to all possible states. In quantum cosmology, we are often interested in high-energy, Planckian phenomena and cannot restrict attention to the low-energy effective action. Even in low-energy regimes, which would be all we need to make contact with potential observations, it is not clear what low-energy state should be used. Quantum cosmology in its non-perturbative form, or quantum gravity in general, does not have a vacuum or other distinguished low-energy state. Effective theories of quantum cosmology must therefore be more general than the low-energy effective action, leaving more freedom for states parameterized in some suitable way. Even if we restrict attention to semiclassical regimes, the class of states to be considered may be large: simple Gaussians provide at most a 2-parameter family within a large set of semiclassical states when they are fully squeezed. For uncorrelated Gaussians, wave functions $$\psi_{\sigma}(a)= N\exp\left(-\frac{(a-\langle\hat{a}\rangle)^2}{4\sigma^2}\right) \exp(-ia\langle\hat{p}_a\rangle/\hbar)$$ with a normalization constant $N$, we have, besides the two expectation values $\langle\hat{a}\rangle$ and $\langle\hat{p}_a\rangle$ on which the state is peaked, only one quantum parameter, the variance $\sigma$. The most general Gaussian state is of the form $\psi_z(a)=\exp(-z_1a^2+z_2a+z_3)$ with three complex numbers $z_i$ such that ${\rm Re}z_1>0$ for normalizability. Out of these six real parameters, the two contained in $z_3$ do not matter for moments because the real part is fixed by normalization and the imaginary part contributes only a phase factor. For the remaining parameters, writing $z_1=\alpha_1+i\beta_1$ and $z_2=\alpha_2+i\beta_2$ with real $\alpha_i$ and $\beta_i$, we compute expectation values $$\langle\hat{a}\rangle= \frac{\alpha_2}{2\alpha_1}\quad,\quad \langle\hat{p}_a\rangle= \hbar\frac{\alpha_1\beta_2-\alpha_2\beta_1}{\alpha_1}$$ and second-order moments $$(\Delta a)^2 = \frac{1}{4\alpha_1}\quad,\quad (\Delta p_a)^2 = \hbar^2\alpha_1+\hbar^2\frac{\beta_1^2}{\alpha_1}\quad,\quad \Delta(ap_a) = -\hbar\frac{\beta_1}{2\alpha_1}$$ One easily confirms that the uncertainty relation (\[Uncert\]) is always saturated. Deviations from saturation are much more difficult to parameterize, but easily occur for evolved semiclassical states even if they start out as a Gaussian. And even if one stays close to the saturation condition and does not vary second-order moments much beyond the values they can take for Gaussians, a Gaussian determines all higher moments in terms of the real and imaginary parts of $z_1$. Our general semiclassicality condition $\Delta(a^bp_a^c)\sim O(\hbar^{(b+c)/2})$ provides an infinite-parameter family instead of the special 2-parameter one realized for Gaussians. Unless one can motivate Gaussians by other means, for instance proximity to the Gaussian harmonic-oscillator ground state or the vacuum of a free field theory, using them exclusively may easily be too restrictive. (Given the large parameter space of states, one may be tempted to refer to probabilistic arguments to pick “likely” states, following ideas that go back to an analysis of inflation in quantum cosmology [@MeasureUniverses; @MeasureInflation]. However, such probability considerations, though popular, are difficult, if not impossible, to make sense of in quantum cosmology [@MeasureCosmo].) Not just the absence of a vacuum state but several other special properties of quantum cosmology are important when we consider possible states: - Quantum cosmology considers long-term evolution. Even if we may be able to choose a specific form of semiclassical states at large volume and small curvature, it may change much when states are evolved back to high densities to infer possible implications at the big bang. - As already mentioned, even the form of semiclassical states is unclear. It is customary to explore semiclassical features using simple and nicely peaked Gaussian states. In quantum mechanics, such states provide interesting information, and they are realized in exactly this form as coherent states or the ground state of the harmonic oscillator. In quantum field theory, Gaussian states are then close to the perturbative vacuum even for interacting theories. The dynamics of quantum cosmology, however, is not near that of the harmonic oscillator (except for some special models), and the unquestioned use of Gaussians is more difficult to justify. But going beyond Gaussians is complicated in terms of wave functions, whose parameters then become much less controlled. - It is not clear how precisely quantum cosmology can be derived from some full theory of quantum gravity, but the number of degrees of freedom is certainly reduced either by exact symmetries or by perturbing around some background. In such situations, when degrees of freedom are eliminated, pure states easily become mixed. For general evolution equations able to model a full state, we should therefore allow for the possibility of mixed states, again going beyond pure Gaussians or other specific wave functions. Moments provide a parameterization of mixed states as well since they are based only on the notion of expectation values. - The question of covariance affects also the choice of classes of states. A state chosen in a minisuperspace model must have a chance of being the reduction of a full state that does not break covariance. This question may be difficult to analyze at the level of wave functions, but it also shows that a sufficiently large freedom in the choice of states must be included in considerations that aim to provide a consistent formulation of quantum cosmology. Since state properties are important for effective actions and quantum back-reaction, we should be as general as possible with the choice of states we consider. With wave functions or density matrices for mixed states, such a generality is difficult to achieve, but it is possible with parameterizations such as the one by moments. Moments provide a general form of semiclassical states, as already introduced. In effective theories, they are subject to their own evolution equations which show how they may change as high-density regimes are approached. They go well beyond the 2-parameter family of squeezed Gaussians, and describe pure and mixed states alike. ### Quantum phase space and covariance Covariance at the level of effective equations with quantum back-reaction can be addressed by combining the moment parameterization with the methods of the previous section, deriving consistent constrained systems amended by quantum corrections that include the moments. For the last question, we must be able to fit moments into a phase-space structure, so as to be able to compute Poisson brackets such as $\{H[M,M^a],H[N,N^a]\}$ with constraints that may include moment terms, such as an inhomogeneous version of (\[Qconstraint\]). This construction is indeed possible: Together with expectation values, the moments form a quantum phase space with Poisson brackets defined by the commutator [@EffAc], first for expectation values of arbitrary operators: $$\label{Poisson} \{\langle\hat{A}\rangle,\langle\hat{B}\rangle\}= \frac{\langle[\hat{A},\hat{B}]\rangle}{i\hbar}\,.$$ This bracket satisfies the Jacobi identity and is linear. If we extend it to polynomials of expectation values by imposing the Leibniz rule, all laws for a Poisson bracket are satisfied, and we can apply the definition to moments. For instance, we obtain $\{\langle\hat{a}\rangle,\langle\hat{p}_a\rangle\}=1$ and $\{\langle\hat{a}\rangle,\Delta(a^bp_a^c)\}=0= \{\langle\hat{p}_a\rangle,\Delta(a^bp_a^c)\}$ for all $b$ and $c$. The moments, as defined here, are symplectically orthogonal to expectation values, a convenient feature for calculations. Poisson brackets between different moments have been computed explicitly but are lengthy; see [@EffAc] and the correction of a typo in [@HigherMoments]. For low orders, it is usually more convenient to compute Poisson brackets directly from the definition (\[Poisson\]). For instance, we calculate $$\begin{aligned} \label{FluctBracket} \{(\Delta q)^2,(\Delta p)^2\}&=& \{\langle\hat{q}^2\rangle-\langle\hat{q}\rangle^2, \langle\hat{p}^2\rangle-\langle\hat{p}\rangle^2\}\\ &=& \frac{\langle[\hat{q}^2,\hat{p}^2]\rangle}{i\hbar} - 2\langle\hat{p}\rangle \frac{\langle[\hat{q}^2,\hat{p}]\rangle}{i\hbar} - 2\langle\hat{q}\rangle \frac{\langle[\hat{q},\hat{p}^2]\rangle}{i\hbar} + 4\langle\hat{q}\rangle\langle\hat{p}\rangle \frac{\langle[\hat{q},\hat{p}]\rangle}{i\hbar}\nonumber \\ & =& 2\langle\hat{q}\hat{p}+\hat{p}\hat{q}\rangle- 4\langle\hat{q}\rangle\langle\hat{p}\rangle= 4\Delta(qp)\nonumber\end{aligned}$$ using the Leibniz identity and (\[Poisson\]). If we know how a state enters effective constraints via its moments, a question which we will address in due course, we can compute Poisson brackets of effective constraints and see whether they provide a consistent deformation of the classical constraint algebra. If a consistent deformation is realized, the system can be analyzed further by standard canonical means to arrive at observables and dynamical equations. Note, however, that the Poisson tensor for moments truncated to some order is in general not invertible, for instance on the three second-order moments which form an odd-dimensional Poisson manifold: With (\[FluctBracket\]) and similar calculations for the other second-order moments, we obtain $$\begin{aligned} \label{FluctAlg} \{(\Delta q)^2,(\Delta p)^2\}&=& 4\Delta(qp)\,,\\ \{(\Delta q)^2,\Delta(qp)\}&=& 2(\Delta q)^2\,,\\ \{(\Delta p)^2,\Delta(qp)\}&=&-2(\Delta p)^2\,,\end{aligned}$$ the 3-dimensional Poisson manifold of second-degree polynomials. Symplectic geometry therefore cannot be used for effective theories, while Poisson geometry is available from the definition (\[Poisson\]). All relevant properties of constrained systems, such as the distinction between first and second class or properties of gauge transformations, can be formulated at the level of Poisson geometry [@brackets]. ### Quantum-gravity states In summary of this section, we note that there are several specific issues in the derivation of effective theories for quantum cosmology, compared to other fields in which such methods are in use. The central theme is covariance, an issue which can be addressed only if one goes beyond the traditional quantum-cosmological realm of minisuperspace models. The notion of covariance and the question whether it is realized consistently may be affected by new quantum space-time structures introduced by the specific form of quantum geometry in one’s approach to quantum gravity. The second, less obvious source of potential modifications of covariance is the form of dynamical quantum states used. Effective actions or equations depend on the classes of quantum states whose evolution they approximate, which should be specific solutions of some underlying quantum theory of gravity. Even if a theory such as quantum gravity is covariant, the selection of specific solutions may always break this symmetry. Covariance may then be realized in a deformed way, or only partially within one effective theory. For instance, even if the states used are peaked on covariant classical fields, giving rise to covariant effective terms depending on expectation values, their fluctuations or higher moments may not be fully covariant. Changing the gauge in quantum gravity may then require the transition to a different effective theory, in which different state parameters have been used (reminiscent of the application of background-field methods in standard quantum field theory). On the other hand, if the class of states is sufficiently large, all gauge-related parameters could be encompassed within one effective theory. These considerations highlight the importance of the selection of states in the derivation of effective theories. Both quantum geometry and quantum dynamics must be part of one consistent quantum theory of gravity; in a complete treatment, their effects therefore cannot be separated from each other. However, they are derived by different means, making use of different expansions of the expectation-value function $\langle\hat{H}\rangle$ of the Hamiltonian or Hamiltonian constraint in the class of states used. In what follows, as in most derivations in the literature, we will split the treatment into one of quantum-geometry corrections first, as they are easier to see, followed by a discussion of quantum-dynamics corrections. When their expressions are known, we can compare implications of different effects and see if some are more relevant than others in specific regimes. In the canonical setting, quantum-geometry corrections are specific to loop quantum gravity which has given rise to many results regarding background-independent quantization [@Rov; @ThomasRev; @ALRev]. The effects in $\langle\hat{H}\rangle$ are not unique, but characteristic enough to show implications for quantum space-time structure. After an overview of these terms in the next section, we will discuss dynamical quantum back-reaction, obtained from a further expansion of $\langle\hat{H}\rangle$ by the moments parameterizing states. We will see the relation of canonical formulations to the low-energy effective action used in particle physics and the role of higher time derivatives in effective equations. Finally, we will put together our results to find properties of general effective actions taking into account all possible effects of canonical quantum gravity, and thereby shed light on quantum space-time structure. Quantum geometry of space {#QGS} ========================= Canonical gravity implements space-time structure by imposing the constraints $H[N]+D[N^a]$ and requiring invariance under the gauge flow they generate. (There may be additional constraints, such as the Gauss constraint if triad variables are used. Such constraints, however, restrict auxiliary degrees of freedom not related to transformations of space-time.) The diffeomorphism (and Gauss) constraint generates a simple flow by Lie derivatives, and has a direct action on quantum states. If we use states $\psi[q_{ab}]$ as in Wheeler–DeWitt quantum gravity (leaving aside the complicated question of how to define an inner product) we have a formal action $\hat{D}[N^a]\psi[q_{bc}]= \psi[{\cal L}_{N^a}q_{bc}]$. States are annihilated by the diffeomorphism constraint if they depend only on spatial invariants, of the same form as classically, and expectation values in non-invariant states transform according to the classical gauge transformations. In loop quantum gravity, the action on states is different because discrete spatial structures do not allow infinitesimal diffeomorphisms to be represented. But for finite diffeomorphisms, the situation is exactly as just described. Therefore, we do not expect significant quantum modifications of the spatial diffeomorphism constraint (but see Section \[s:Diffeo\]). Similarly, the Gauss constraint, if there is one, has a simple and direct action that does not suggest modifications; see also the end of Section \[s:SphSymm\]. Hamiltonian constraint ---------------------- The Hamiltonian constraint $\hat{H}[N]$ presents a different story. It cannot be quantized directly by promoting its (complicated) classical gauge flow to an action on states. The only procedure to arrive at a constraint operator is tedious: inserting basic operators quantizing the classical kinematical phase space into the expression for the Hamiltonian constraints. As a phase-space function, the constraint contains rather involved combinations of the basic fields, which require regularization and sometimes even other modifications for a well-defined operator to result [@RS:Ham; @QSDI]. Also on general grounds, we do expect quantum corrections in the Hamiltonian constraint because it includes all about the dynamics of the theory. The theory being interacting, quantum corrections must result. The Hamiltonian constraint is therefore the place where we should look for characteristic quantum corrections of different theories, as well as possible restrictions by consistency requirements. ### Gauge flow {#s:Gauge} A constraint operator $\hat{H}[N]$ restricts states by $\hat{H}[N]|\psi\rangle_{\rm phys}=0$ and generates a gauge flow $|\psi\rangle_{\epsilon}= \exp(-i\hat{H}[\epsilon]/\hbar)|\psi\rangle$. Physical states, on which $\exp(-i\hat{H}[\epsilon]/\hbar)$ acts as the identity, are gauge-invariant, and the gauge flow need not be considered separately if first-class constraint operators are solved. But exact solutions are complicated to find, and when quantum corrections are computed by systematic approximation schemes, the situation is quite different. Expectation values are constrained by $\langle\hat{H}[N]\rangle$, a weaker condition than $\hat{H}[N]|\psi\rangle=0$. We have additional constraints $\langle(\hat{f}-\langle\hat{f}\rangle)\hat{H}[N]\rangle=0$, an enlarged set of constraints which all vanish in physical states, for arbitrary $\hat{f}$. These constraints in general are all independent, constraining not just expectation values but also moments. In an effective theory, as spelled out in detail later, one solves this infinite set of quantum phase-space constraints order by order in the moments. To any given order, the sharp condition $\hat{H}[N]|\psi\rangle_{\rm phys}=0$ for physical states is not fully implemented, and on the corresponding solution space there is a non-trivial gauge flow by quantum constraints. Moreover, the effective treatment at this stage is more general because one does not assume a (kinematical) Hilbert-space structure when solving the constraints for moments; therefore the standard argument of a trivial gauge flow does not apply. At the effective level, there are non-trivial gauge transformations by quantum constraints, bringing solution procedures closer to classical ones: one solves phase-space constraints and factors out their gauge. For an effective theory, the key ingredient is therefore the expectation value $\langle\hat{H}[N]\rangle$ in a sufficiently large class of states, or the general expression parameterized by the moments of states. As in (\[Qconstraint\]), moments then appear in quantum corrections that change the constraint surface. With a modified constraint surface, the gauge flows must receive quantum corrections as well for them to be tangential to the constraint surface mapping solutions to constraints into other solutions, as required for a first-class system. Indeed, the expectation value of the constraint, an expression that includes quantum corrections, determines the gauge flow on the quantum phase space with Poisson brackets (\[Poisson\]) by $$\label{dOdt} \delta_{\epsilon}\langle\hat{O}\rangle(\epsilon)= \epsilon \frac{\delta}{\delta\epsilon} \langle\psi |\exp(i\hat{H}[\epsilon]/\hbar) \hat{O} \exp(-i\hat{H}[\epsilon]/\hbar)|\psi\rangle= \frac{\langle[\hat{O},\hat{H}[\epsilon]]\rangle}{i\hbar}= \{\langle\hat{O}\rangle,\langle\hat{H}[\epsilon]\rangle\}$$ for expectation values $\langle\hat{O}\rangle(\epsilon)= {}_{\epsilon}\langle\psi|\hat{O}|\psi\rangle_{\epsilon}$, and therefore for any quantum phase-space function such as the moments by using the Leibniz rule. If the constraint surface changes by corrections in $\hat{H}$, so do the gauge transformations. For the quantum corrected constraint surface and the gauge flow we are therefore required to compute $\langle\hat{H}[\epsilon]\rangle$ in general kinematical states. If quantum constraint operators are represented in an anomaly-free way, such that any commutator $[\hat{C}_1,\hat{C}_2]$ is an operator of the form $f(\hat{q},\hat{p})\hat{C}_3$ with another constraint operator $\hat{C}_3$ and structure functions $f(q,p)$, the quantum constraints $\langle\hat{C}_i\rangle$ form a consistent first-class system: $\{\langle\hat{C}_1\rangle,\langle\hat{C}_2\rangle\}= \langle[\hat{C}_1,\hat{C}_2]\rangle/i\hbar= \langle f(\hat{q},\hat{p})\hat{C}_3\rangle$. Since all expressions such as $\langle f(\hat{q},\hat{p})\hat{C}_3\rangle$ vanish when computed in physical states, they provide effective constraints (in general independent of $\langle\hat{C}_3\rangle$ as phase-space functions). If all these constraints are imposed, the effective constrained system is first class and has a consistent gauge flow generated by all these constraints [@EffCons; @EffConsRel; @EffConsComp]. Such systems of infinitely many constraints for each local classical degree of freedom can be difficult to analyze. However, just as the moments obey a $\hbar$ierarchy in semiclassical regimes, the effective constraints can be truncated to finite sets to any given order in $\hbar$ or in the moments. The leading corrections can be found in direct expectation values $\langle\hat{C}_i\rangle$, restricting expectation values with quantum corrections that depend on the moments. For instance, (\[Qconstraint\]), when imposed as a constraint, shows how the classical constraint surface and the gauge flow of $\langle\hat{a}\rangle$ and $\langle\hat{p}_a\rangle$ are changed by fluctuations $(\Delta p_a)^2$. These corrections depend on the value of $(\Delta p_a)^2$, which is constrained and subject to gauge flows by higher-order constraints such as $\langle\hat{p}_a\hat{H}\rangle$. When all constraints have been solved and gauge flows factored out to a certain order, the condition $\hat{H}|\psi\rangle=0$ has been implemented for states used in expectation values and moments. One has then computed observables in physical states, sidestepping the complicated problem of computing an integral form of the physical inner product. This is one example for the use of effective methods to tame complicated technical and conceptual issues in quantum gravity. Even the problem of time can be solved, at least at the semiclassical level: Different choices of time are related by mere gauge transformations in the quantum phase space [@EffTime; @EffTimeLong; @EffTimeCosmo]. We will come back to these conceptual question in Section \[s:EffCons\], and for now note the lesson that expectation values of constraints supply the key ingredient for effective gauge theories. ### Quantum-geometry effects We should then apply effective techniques to the Hamiltonian constraint of gravity, computing $\langle\hat{H}[N]\rangle$. Expressions for the Hamiltonian constraint are rather complicated even classically, given by $$\label{HADM} H_{\rm grav}(q_{ab},\pi^{cd})= -\frac{16\pi G}{\sqrt{\det q}} \left(\pi_{ab}\pi^{ab}- \frac{1}{2}(\pi^a_a)^2\right)+ \frac{\sqrt{\det q}}{16\pi G} {}^{(3)}R$$ in ADM variables, the spatial metric $q_{ab}$ and its momentum $$\pi^{cd}= \frac{\sqrt{\det q}}{16\pi G} (K^{cd}-K^a_aq^{cd})$$ in terms of extrinsic curvature (with the spatial Ricci scalar ${}^{(3)}R$). Loop quantum gravity uses a different set of variables, the densitized triad $E^a_i$ instead of the spatial metric $q^{ab}=E^a_iE^b_i/\det (q_{cd})$ and the Ashtekar–Barbero connection $A_a^i=\Gamma_a^i+\gamma K_a^i$ [@AshVar; @AshVarReell], defined by combining the spin connection $\Gamma_a^i$ compatible with the densitized triad and extrinsic curvature $K_a^i=K_{ab}E^{bi}/\sqrt{|\det (E^c_i)|}$. The basic Poisson brackets are $$\label{AE} \{A_a^i(x),E^b_j(y)\}= 8\pi\gamma G \delta_a^b \delta_j^i \delta(x,y) \,.$$ The Barbero–Immirzi parameter $\gamma>0$ [@AshVarReell; @Immirzi] does not play a role classically but appears in quantum spectra and corrections. In these variables, the Hamiltonian constraint is $$\label{HAsh} H_{\rm grav}(A_a^i,E^b_j) = -\frac{E^a_iE^b_j\epsilon^{ij}{}_k}{16\pi G \sqrt{|\det (E^c_l)|}} \left(F_{ab}^k+(1+\gamma^{-2}) \epsilon^k{}_{mn} (A_a^m-\Gamma_a^m)(A_b^n-\Gamma_b^n)\right)$$ with the curvature $F_{ab}^k$ of the Ashtekar–Barbero connection. Both expressions are rather complicated to quantize, owing for instance to the expressions of $^{(3)}R$ or $\Gamma_a^i$, to be written in terms of the metric or densitized-triad operators. Fortunately, there are several characteristic features, common to both versions: The Hamiltonian constraint is quadratic in the connection or extrinsic curvature, and it requires an inverse of the spatial metric or the densitized triad. These two features, when combined with quantum-representation properties, imply characteristic structures of quantum geometry and associated corrections to classical equations. These corrections, in turn, can be analyzed for potential implications even if their precise form cannot be determined from a complete calculation of $\langle\hat{H}\rangle$ in semiclassical states. And even if expectation values could be computed for a specific $\hat{H}$ and some states, ambiguities in the construction of $\hat{H}$ or the choice of states would be so severe that only general features and characteristic effects could be trusted. In a Wheeler–DeWitt quantization, we have formal quantum constraint equations with $\pi_{ab}$ in (\[HADM\]) replaced by functional derivatives with respect to $q_{ab}$, acting on wave functions $\psi[q_{ab}]$. The formal nature leaves precise representation properties unclear, and therefore does not show specific quantum effects; one simply takes the classical expression and performs the usual substitution of momenta by derivative operators. One does not expect strong quantum-geometry effects, simply because quantum geometry has not been completed in this setting. Loop quantum gravity, on the other hand, has celebrated its greatest success so far at this level of quantum representations, and indeed sheds considerable light on questions of quantum corrections resulting from quantum geometry. ### Loop representation and background independence Loop quantum gravity looks closely at representation properties of basic operators. To eliminate the need for formal functional derivatives and the associated delta functions in quantizations of (\[AE\]), the classical fields $A_a^i$ and $E^b_j$ are smeared or integrated over suitable sets in space: holonomies $h_e(A_a^i)= {\cal P}\exp(\int_e {\rm d}\lambda t_e^a A^i_a \tau_i)$ and fluxes $F^{(f)}_S(E^b_j)= \int_S{\rm d}^2y n^S_a E^a_if^i$ with the tangent vector $t_e^a$ to a curve $e$ and the co-normal $n^S_a$ to a surface $S$ (on which an su(2)-valued smearing function $f^i$ is chosen). Allowing for all curves and surfaces, all information about the fields can be recovered. By the integrations, the delta function in $\{A_a^i(x),E^b_j(y)\}=8\pi\gamma G\delta^b_a\delta^i_j\delta(x,y)$ is eliminated, and a well-defined holonomy-flux algebra results: $$\label{HolFlux} \{h_e(A_a^i),F_S(E^b_j)\}= 8\pi\gamma G h_{e\rightarrow x}\tau_if^i(x) h_{x\rightarrow e}$$ if there is only one intersection point $\{x\}=e\cap S$, denoting by $h_{e\rightarrow x}$ and $h_{x\rightarrow e}$ the holonomy along $e$ up to $x$ and starting at $x$, respectively. With strong uniqueness properties of possible quantum representations [@FluxAlg; @Meas; @LOST; @WeylRep], spatial quantum geometry in this setting is under excellent control. Representations of the holonomy-flux algebra then provide operators to be inserted in (\[HAsh\]) to obtain a quantized Hamiltonian constraint. The kinematical Hilbert space is spanned by cylindrical states $\Psi(A_a^i)= \psi(h_{e_1}(A_a^i),\ldots, h_{e_n}(A_a^i))$, each of which depends on the connection via a finite number of holonomies. The full state space, has no restriction on the number of holonomies that may appear, thereby representing the continuum theory rather than some lattice model. The inner product is obtained by integrating the product of two cylindrical functions over as many copies of SU(2) as there are non-trivial dependencies on holonomies in both states, using the normalized Haar measure. By completion of the space of cylindrical functions, the kinematical Hilbert space is obtained [@ALMMT]. Holonomies $\hat{h}_e$ then act as multiplication operators, changing the dependence of a cylindrical state on $h_e$, or creating a new dependence if the state was independent of $h_e$ before acting. Flux operators, representing the densitized triad, become derivative operators in the connection representation used, and can be expressed in terms of invariant derivative operators on SU(2) (or angular-momentum operators). They have discrete spectra, indicating modifications to the classical spatial structure [@AreaVol; @Area; @Vol2]: distance, areas or volumes computed from $E^a_i$ can, after quantization, no longer increase continuously even if the underlying curves, surfaces, or regions are deformed by homotopies. As a further consequence of discreteness, it turns out that holonomy operators do not continuously depend on the curves used. A cylindrical state depending only on $h_e$ and one depending only on $h_{e\circ e'}$ with a piece $e'$ appended to the curve refer to two different holonomies, and are orthogonal according to the inner product just described, even if $e$ is a 1-point set. While we classically have $t_{e'}^a A_a^i(x)\tau_i= \lim_{e'\to\{x\}} (h_{e'}(A_a^i)-1)/|e'|$, where $t_{e'}^a$ is the tangent vector of $e'$ at $x$ and $|e'|$ its coordinate length, the sequence $\hat{h}_{e'}|\psi\rangle$ of states after quantization does not converge for any $|\psi\rangle$ if $e'$ is changed. The edge dependence disappears when one implements the diffeomorphism constraint, which has not been assumed in the previous constructions. States $\hat{h}_{e'}|\psi\rangle$ for different $e'$ are no longer orthogonal, but they all give rise to the same state when diffeomorphism are factored out. The limit can then be taken, but always equals zero. Again, there is no way of deriving a connection operator. Unlike with classical expressions, it is not possible to take a derivative of $h_e(A_a^i)$, for instance by the endpoint of $e$, to obtain an expression for a quantized $A_a^i$. Loop quantum gravity does not offer connection operators; all connection dependence in the Hamiltonian constraint must be expressed in terms of holonomies. No quadratic function as it appears in the constraint can exactly agree with a linear combination of exponentials, and modifications arise, motivated by background-independent quantum geometry. ### Holonomy corrections When the Hamiltonian constraint is quantized in loop quantum gravity, holonomies are used instead of connection components [@RS:Ham], providing a “regularization” necessary to render the constraint expressible by basic operators. However, the limit in which the “regulator” — the specific curves used for holonomies — is removed does not exist; we are not dealing with a proper regularization. (In [@QSDI], the limit is argued to exist and be trivial if spatially diffeomorphism-invariant states are used. But with this assumption there is no handle on the full off-shell algebra and its anomaly problem.) Instead, one usually interprets the difference of a holonomy-modified constraint or its effective versions with the classical expression as a series of higher-order corrections, amending the classical Hamiltonian by higher powers of the connection, or intrinsic and extrinsic curvature. In this viewpoint, the modification is similar to expected higher-curvature corrections — except for the issue of covariance that we will have to address. (In some models of loop quantum cosmology, it is possible to represent an exponentiated version of the constraint in terms of holonomy operators without introducing modifications to the classical expression [@NonExpLQC]. However, the procedure seems to depend sensitively on specific properties of the model used and is not available in general.) Holonomy corrections therefore appear as higher-order terms such as $$\label{DerExp} -2 {\rm tr}(\tau_ih_e(A_b^j))-t_e^aA_a^i(x)= -\frac{1}{24}(t_e^aA_a^i(x)(t_e^bA_b^j(x))(t_e^cA_{cj}(x))+ t_e^at_e^b \partial_bA_a^i(x)+\cdots$$ obtained from a Taylor expansion of the exponential and a derivative expansion of the integration. These terms depend on the routing of the curve $e$, to be chosen suitably for a quantization of the Hamiltonian constraint. The condition of anomaly-freedom puts strong restrictions on the possible routings [@TwoPlusOneDef] which, however, are difficult to evaluate. Moreover, the expansion is done in an effective derivation, requiring the calculation of expectation values that lead to additional moment terms. All these calculations are difficult to perform explicitly, but the form of corrections is clear and quite characteristic: higher orders as well as higher spatial derivatives in the connection. With these types of corrections, suitably parameterized, one can look for consistent deformations of the constraint algebra to find versions for a physical evaluation of these corrections, or to derive further restrictions on the quantization choices made. In effective equations, the use of holonomies as basic operators of the quantum theory has another consequence: The holonomy-flux algebra (\[HolFlux\]) is not canonical, with non-constant Poisson brackets. Moments of states used for effective descriptions are based on expectation values of basic operators, holonomies and fluxes in loop-quantized models. The general constructions of effective theories still go through: Poisson brackets on the quantum phase space, the quantum Hamiltonian or constraints, and so on. However, the explicit Poisson relations (\[Poisson\]) evaluated between individual moments of the form $\Delta(h^bF^c)$ are different from the canonical case. In particular, moments no longer Poisson-commute with expectation values. This feature requires care and complicates some calculations. These problems, however, are only technical; see the examples provided later in Section \[s:HarmLQC\] and in [@BouncePert; @BounceCohStates]. ### Inverse-triad corrections Fluxes are linear in the densitized triad and do not suggest the same kind of modification as holonomies. Nevertheless, there is a characteristic effect associated also with them. Flux operators have discrete spectra, containing zero as an eigenvalue. Such operators do not have densely-defined inverses, and yet we need an inverse of the densitized triad for the classical Hamiltonian constraint. To obtain such quantizations, a more indirect route is taken in loop quantum gravity, which does result in well-defined operators but introduces another kind of correction, called inverse-triad correction. To see the form of these corrections, we show a lattice calculation of inverse-triad operators. As proposed in [@QSDI; @QSDV], the combination of triad components required for the Hamiltonian constraint (\[HAsh\]) is first rewritten as $$\label{Inverse} 2\pi\gamma G \epsilon^{ijk}\epsilon_{abc} \frac{E^b_jE^c_k}{{\sqrt{|\det E|}}}=\left\{A_a^i,\int{\sqrt{|\det E|}}\mathrm{d}^3x\right\}\,.$$ In the new form, no inverse is needed, $A_a^i$ can be expressed by holonomies, the volume operator can be used for $\int{\sqrt{|\det E|}}\mathrm{d}^3x$, and the Poisson bracket be turned into a commutator divided by $i\hbar$. The resulting operators are rather contrived, especially with SU(2) holonomies and derivatives involved. But the presence of corrections and their qualitative form can be illustrated by a U(1)-calculation, for which we also assume regular cubic lattices. On a cubic lattice, we can assign a unique plaquette to each link $e$, by which we then label fluxes. Our basic operators are $\hat{h}_e$, which as a multiplication operator by $\exp(i\int_eA_at_e^a{\rm d}\lambda)$ takes values in U(1), and $\hat{F}_e$ for a fixed set of edges $e$ in a regular cubic lattice. We are therefore computing inverse-triad operators and their expectation values for a fixed subset of cylindrical states, but the lattice can be as fine as we want, allowing us to capture all continuum degrees of freedom. For this U(1)-simplification, the holonomy-flux algebra, quantizing an Abelian version of (\[HolFlux\]), reads $[\hat{h}_e,\hat{F}_e]= -8\pi\gamma\ell_{\rm P}^2\hat{h}_e$ while all operators commute if they belong to different edges. Moreover, the U(1)-valuedness of holonomies implies the reality condition $\hat{h}_e\hat{h}_e^{\dagger}=1$, which we will make use of below. We first rewrite (\[Inverse\]) in terms of holonomies instead of connection components, and express the volume $V=\int\sqrt{|\det E|}{\rm d}^3x$ by lattice fluxes $\sqrt{|F_1F_2F_3|}$ per vertex, with the three fluxes through plaquettes in all three directions around the vertex: $t_e^a\{A_a,V\}=ih_e\{h_e^{-1},\sqrt{|F_1F_2F_3|}\}$ or, more symmetrically, $\frac{1}{2}i(h_e\{h_e^{-1},\sqrt{|F_1F_2F_3|}\}- h_e^{-1}\{h_e,\sqrt{|F_1F_2F_3|}\})$, computes the Poisson bracket at the vertex. Since $h_e$ commutes with all but one of the $F_I$, we can focus on one of them, $\sqrt{|\hat{F}_e|}$. Keeping the power more general, the quantization of some inverse power of flux takes the form $$\widehat{(|F|^{r-1} {\rm sgn}F)_e}= \frac{\hat{h}_e^{\dagger}|\hat{F}_e|^r\hat{h}_e- \hat{h}_e|\hat{F}_e|^r\hat{h}_e^{\dagger}}{16\pi Gr\gamma\ell_{\rm P}^2} =: \hat{I}_e\,.$$ For any $0<r<1$ we quantize an inverse power of $F$ but need not use any inverse in the commutator. Following [@InflTest], we can now easily simplify these operators, if we observe the relations of the U(1)-holonomy-flux algebra, together with the reality condition. These relations imply $$\hat{h}_e^{\dagger}|\hat{F}_e|^r\hat{h}_e= |\hat{F}_e+8\pi\gamma\ell_{\rm P}^2|^r \quad,\quad \hat{h}_e|\hat{F}_e|^r\hat{h}_e^{\dagger}= |\hat{F}_e-8\pi\gamma\ell_{\rm P}^2|^r\,,$$ such that $$\hat{I}_e= {\frac{|\hat{F}_e+8\pi\gamma\ell_{\rm P}^2|^r- |\hat{F}_e-8\pi\gamma\ell_{\rm P}^2|^r}{16\pi Gr\gamma\ell_{\rm P}^2}}\,.$$ Eigenvalues of this operator can easily be computed, with eigenstates equal to flux eigenstates [@InvScale; @Ambig]. All eigenvalues are finite, as required for a densely-defined operator, and show how the classical divergence of $|F|^{r-1}$ at $F=0$ is cut off. For inverse-triad corrections in effective Hamiltonians, however, we need expectation values of $\hat{I}_e$ in semiclassical states. Explicit calculations would require good knowledge of semiclassical wave functions or coherent states. For general effective equations it is sufficient, and even more useful, to perform a moment expansion, keeping the specific state free and parameterized by moments. Staying at the expectation-value order of effective expressions, we have $$\label{IExp} \langle\hat{I}_e\rangle= {\frac{|\langle\hat{F}_e\rangle+8\pi\gamma\ell_{\rm P}^2|^r- |\langle\hat{F}_e\rangle-8\pi\gamma\ell_{\rm P}^2|^r}{16\pi Gr\gamma\ell_{\rm P}^2}}+ \mbox{moment terms}\,.$$ Already to this order we see characteristic corrections (depending on $\hbar$ via the Planck length). Inverse-triad corrections therefore have a contribution independent of quantum back-reaction. Interpreting $\langle\hat{F}\rangle=:L^2$ as the discrete quantum-gravity scale (the lattice spacing as measured by flux operators), we find the correction function $$\label{alpha} \alpha_r(L):= \frac{\langle\hat{I}\rangle}{I_{\rm class}}= \frac{|L^2+8\pi\gamma\ell_{\rm P}^2|^r- |L^2-8\pi\gamma\ell_{\rm P}^2|^r}{16\pi\gamma r\ell_{\rm P}^2}L^{2(1-r)}$$ that will appear in an effective Hamiltonian constraint. To leading order in an expansion by $\hbar$ (or $\ell_{\rm P}^2/L^2$), the correction function equals one. But even if no moment terms are included, there are quantum corrections in the full form of $\alpha(L)$. Corrections are strong for $L^2\sim 8\pi\gamma\ell_{\rm P}^2$ or smaller, typically in the deep quantum regime, where $\alpha(L)$ drops to zero at $L=0$. However, even for larger $L$, $\alpha_r(L)$ is not identical to one and implies interesting corrections. In addition to corrections contained in (\[alpha\]) and quantum back-reaction from moment terms, the flux dependence implies corrections from a derivative expansion of the integrations involved, as already seen for holonomies. Moreover, non-Abelian holonomies do not lead to exact cancellations in the substitution of $h_e\{h_e^{-1},V\}$ for $t_e^a\{A_a^i,V\}$ and rather imply additional higher-order corrections by powers of $A_a^i$ [@DegFull]. As noted in the context of holonomy corrections, such extra terms mix with higher-curvature corrections. The leading term in (\[alpha\]), on the other hand, shows a different dependence on parameters that distinguish a given cosmological regime and are more characteristic. Their effects can thus be studied in isolation. Diffeomorphism constraint {#s:Diffeo} ------------------------- We have already stated that the diffeomorphism constraint can be quantized by its direct action on spatial functions or other objects such as curves and surfaces. In loop quantum gravity, for instance, a diffeomorphism $\Phi$ acts by shifting all arguments of a cylindrical function by $h_e\mapsto h_{\Phi(e)}$, the usual pull-back of functions. The representation of the holonomy-flux algebra is diffeomorphism covariant under this action, showing that no quantum corrections to classical diffeomorphisms result. It is not possible to compute or represent an infinitesimal action or the diffeomorphism constraint because two states that differ by a non-trivial diffeomorphism are either identical (if the diffeomorphism does not change the underlying graph) or orthogonal. But finite diffeomorphisms suffice to remove the related gauge, which is done without quantum corrections. Nevertheless, the situation is not completely satisfactory because the diffeomorphism constraint is a crucial ingredient of the hypersurface-deformation algebra. If diffeomorphisms are represented without quantum corrections, there should be no deformations of the relations (\[DD\]) and (\[HD\]) of the hypersurface-deformation algebra for commutators involving at least one spatial deformation. However, the diffeomorphism constraint also appears on the right-hand side of (\[HH\]), the crucial part for space-time structure. On the left-hand side, we have two Hamiltonian constraints, which we do quantize in loop quantum gravity and whose commutators we can, in principle, compute. The result should be a well-defined operator, which must vanish on physical states for the quantization to be anomaly-free. However, classically it corresponds to a diffeomorphism constraint, which cannot be represented directly. To check for anomaly freedom, one must then find an operator version of the right-hand side of (\[HH\]), taking into account the structure function $q^{ab}$, to be turned into an operator as well. This is one of the most important but still outstanding issues in loop quantum gravity, which was evaded by the arguments of [@QSDI] and only partially addressed by the advanced constructions of [@LM:Vertsm; @Consist]. More recently, the issue has been revisited in several models [@TwoPlusOneDef], with encouraging results. At least in U(1)-versions of $2+1$-dimensional gravity, one can indeed make sense of the right-hand side of (\[HH\]) as an operator, in such a way that the quantum constraint algebra is anomaly-free. As a side product, the same deformation (\[HHbeta\]) with inverse-triad corrections as seen by effective methods [@ConstraintAlgebra] appears. (Holonomy corrections and their deformation of the constraint algebra could not be seen by the methods of [@TwoPlusOneDef], going back to [@LM:Vertsm; @Consist], because the consistency conditions of anomaly freedom are tested only at vertices.) In [@TwoPlusOneDef], the diffeomorphism constraint itself did not have to be amended by quantum corrections. However, other considerations in the same context have been put forward that may suggest such terms [@DiffeoOp]. At present, the status regarding quantum corrections in the diffeomorphism constraint is incomplete, but a consistent implementation does not appear to be easy. From the point of view of effective theory, corrections to diffeomorphisms do not seem required because, in any canonical space-time theory, one is dealing with fields as functions on space. These functions are represented using some coordinates, but physics as always must be independent of the choice. There must therefore be a part of the gauge content of the theory that requires independence under arbitrary changes of spatial coordinates or, infinitesimally, invariance under spatial Lie derivatives. But then, a gauge transformation that amounts to a Lie derivative of all fields must have a generator identical to the diffeomorphism constraint uniquely associated with the fields [@KucharHypI]. The spatial structure assumed in canonical formulations leaves no room for corrections in the diffeomorphism constraint. The space-time structure is not presupposed in canonical quantum gravity and may well change, as indicated by some quantum corrections in the Hamiltonian constraint. Space-time, unlike space, has dynamical content and can easily receive quantum corrections, as borne out in loop quantum gravity. Having the classical structure of space but modified space-time is therefore consistent. Nevertheless, in an effort to relax some of the general assumptions of canonical formulations, one could expect changes to the spatial manifold structure as well, as perhaps indicated by potential corrections in the diffeomorphism constraint such as those in [@DiffeoOp]. Quantum-geometry effects ------------------------ Comparing holonomy and inverse-triad corrections, we have several important properties: - Holonomy corrections crucially add higher powers of the connection to the classical quadratic form of the Hamiltonian constraint. In flat isotropic cosmological models, the connection is proportional to the Hubble parameter ${\cal H}$, which in turn is proportional to the square root of the energy density. Holonomy corrections in cosmological models therefore depend on the dimensionless parameters $\ell_{\rm P}{\cal H}$ or $\sqrt{\rho/\rho_{\rm P}}$, both of which are tiny in observationally accessible regimes. Inverse-triad corrections, on the other hand, depend on the ratio $\ell_{\rm P}^2/L^2$ with the discrete quantum-gravity scale $L$ in (\[alpha\]). It is not easy to estimate $L$, but the dimensionless ratio associated with it certainly need not be small. Inverse-triad corrections can be more significant than holonomy corrections in observationally accessible regimes. (The scale $L$ may change in time, depending on the form of lattice refinement realized [@InhomLattice; @CosConst].) - Holonomy corrections and inverse-triad corrections are both obtained from properties of integrated objects, holonomies and fluxes. One should therefore expect not just higher-order terms as in the expansions already discussed, but also higher spatial derivatives in a derivative expansion of inhomogeneous models. For holonomy corrections, higher-derivative terms are crucial because they should be part of higher-curvature corrections together with higher powers of the connection that immediately arise from expanded holonomies. Only a suitable combination of higher powers and derivatives can result in consistent covariant versions. - Following up on the last item, we also need higher time derivatives to complete higher-order corrections to covariant objects related to curvature. Such corrections should be present even in homogeneous models, but are not easy to see directly from the form of holonomies. However, such terms cannot be ignored, because high-curvature regimes have significant contributions from higher-order and higher-derivative terms. In isotropic models, ${\cal H}^2$ and $\dot{\cal H}$ are of similar orders, both related to linear combinations of stress-energy components by the Friedmann and Raychaudhuri equations. An expansion of holonomies only by ${\cal H}$ (related to the isotropic connection; see Section \[s:ModFried\]) but ignoring higher time derivatives would be inconsistent. To see how higher time derivatives arise in canonical quantum theories, we have to pause our description of loop quantum gravity and return to more details of quantum back-reaction. Quantum back-reaction {#s:QBR} ===================== For a canonical effective theory, quantum Hamiltonians and quantum constraints $\langle\hat{H}\rangle$, generating evolution or gauge flows by (\[dOdt\]), must be expanded systematically by moments of states to see all quantum effects. This is also the case for individual non-linear correction functions such as $\langle\hat{I}\rangle$ of inverse triads (\[IExp\]) or $\langle\hat{h}\rangle$ of holonomies as they may be implied by quantum-geometry effects of loop quantum gravity. Additional terms, products of expectation values and moments, are then added to the constraints. Effective quantum mechanics --------------------------- The correctness of the quantum dynamics resulting from a moment-expanded $\langle\hat{H}\rangle$ can be illustrated with a quantum-mechanical example. We start with the well-known Ehrenfest equations $$\label{Ehrenfest} \frac{{\rm d}\langle\hat{q}\rangle}{{\rm d}t}=\langle\hat{p}\rangle/m \quad,\quad \frac{{\rm d}\langle\hat{p}\rangle}{{\rm d}t}= -\langle V'(\hat{q})\rangle$$ for basic expectation values, computed using (\[dOdt\]). These equations have been analyzed by [@Hepp] in the limit $\hbar\to 0$ to prove that quantum mechanics has the correct classical limit. Effective equations go beyond this limit by performing a systematic expansion in $\hbar$. The first Ehrenfest equation takes exactly the classical form, while the momentum expectation value is subject to quantum corrections: ${\rm d}\langle\hat{p}\rangle/{\rm d}t=-\langle V'(\hat{q})\rangle$ does not equal the classical force $F(\langle\hat{q}\rangle)= -V'(\langle\hat{q}\rangle)$ at position $\langle\hat{q}\rangle$ (unless the potential is at most quadratic). Moments as quantifiers of corrections arise when we expand the quantum force $F_Q={\rm d}\langle\hat{p}\rangle/{\rm d}t$ as $$\begin{aligned} -\langle V'(\hat{q})\rangle= -\langle V'(\langle\hat{q}\rangle+ (\hat{q}- \langle\hat{q}\rangle))\rangle= -V'(\langle\hat{q}\rangle)- \sum_{n=2}^{\infty} \frac{1}{n!}\frac{\partial^{n+1} V(\langle\hat{q}\rangle)}{\partial\langle\hat{q}\rangle^{n+1}} \Delta(q^n)\nonumber\\ = F_Q(\langle\hat{q}\rangle, \Delta(q^n))\,, \label{FQ}\end{aligned}$$ or the quantum potential as $$\label{QPot} V_Q(\langle\hat{q}\rangle, \Delta(q^n))= \langle V(\langle\hat{q}\rangle+ (\hat{q}- \langle\hat{q}\rangle))\rangle= V(\langle\hat{q}\rangle)+ \sum_{n=2}^{\infty} \frac{1}{n!}\frac{\partial^{n} V(\langle\hat{q}\rangle)}{\partial\langle\hat{q}\rangle^n} \Delta(q^n)$$ such that $-\langle V'(\hat{q})\rangle= -\partial V_Q/\partial\langle\hat{q}\rangle$. ### Quantum Hamiltonian The quantum potential is defined as a function on the infinite-dimensional quantum phase space of expectation values and moments, whose Poisson structure is given by (\[Poisson\]). In a quantum Hamiltonian $$\label{HQ} H_Q=\langle\hat{H}\rangle = \frac{1}{2m}(\langle\hat{p}\rangle^2+ \Delta(p^2))+ V_Q(\langle\hat{q}\rangle, \Delta(q^n))\,,$$ we therefore have terms generating a dynamical flow of the moments by $$\label{Deltadot} \dot{\Delta}(q^bp^c)= \{\Delta(q^bp^c),H_Q\}\,.$$ The coupled set of equations for expectation values and moments, (\[Ehrenfest\]) and (\[Deltadot\]), is equivalent to the Schrödinger flow of quantum mechanics, but its solutions do not provide wave functions but rather variables directly related to observations. It can be solved with different approximations, most importantly a semiclassical one by the order of moments, sometimes combined with an adiabatic one. In the latter case, applied to anharmonic oscillators, effective equations are equivalent to those of the low-energy effective action [@EffAc]. The validity and usefulness of the canonical effective scheme is thereby established. In the context of quantum gravity, the feature of higher time derivatives in effective equations, a crucial ingredient of higher-curvature corrections, is of particular interest. The moments are related to such terms, although not in a direct way. Equation (\[Ehrenfest\]) combined with (\[FQ\]) already shows that a specific linear combination of the moments, with coefficients depending on expectation values, amounts to the time derivative of the momentum, or the second derivative of the position expectation value. Higher than second time derivatives of $\langle\hat{q}\rangle$ can be computed by taking further derivatives of (\[Ehrenfest\]) and inserting $\dot{\Delta}(q^n)$ and, for higher than third order, $\dot{\Delta}(q^bp^c)$ according to equations of motion (\[Deltadot\]) generated by the quantum Hamiltonian. Different combinations of the moments therefore provide all higher time derivatives of $\langle\hat{q}\rangle$. With the scheme just sketched, it is difficult to invert the equations to find expressions for moments in terms of higher time derivatives, or to eliminate all moments and end up with a higher-derivative equation just for $\langle\hat{q}\rangle$ instead of the moment-coupled (\[Ehrenfest\]), (\[FQ\]) and (\[Deltadot\]). But with more-refined methods, as well as an adiabatic expansion, this task can be performed. For quantum cosmology, we learn that we must study quantum back-reaction to see all terms relevant for higher-curvature corrections. In semiclassical regimes, the moments by definition obey the $\hbar$ierarchy $\Delta(q^bp^c)\sim O(\hbar^{(b+c)/2})$, as can easily be verified in Gaussians; see (\[Gaussian\]) and [@HigherMoments]. We can therefore consider the first term $\frac{1}{2} V''(\langle\hat{q}\rangle) (\Delta q)^2$ for $n=1$ in (\[QPot\]) as the leading semiclassical correction, providing a quantum Hamiltonian $$\label{HQsecond} H_Q=\langle\hat{H}\rangle = \frac{1}{2m}\langle\hat{p}\rangle^2+ V(\langle\hat{q}\rangle)+\frac{1}{2m} (\Delta p)^2+ \frac{1}{2} V''(\langle\hat{q}\rangle) (\Delta q)^2\,.$$ (The kinetic term contributes $(\Delta p)^2/2m$, potentially of the same order as $\frac{1}{2}V''(\langle\hat{q}\rangle)(\Delta q)^2$. But it does not appear in a product with expectation values and therefore does not cause quantum back-reaction.) For equations of motion of expectation values and second-order moments, relevant to this order, we use the Poisson brackets (\[FluctAlg\]). Applied to our second-order quantum Hamiltonian, we find $$\begin{aligned} \frac{{\rm d} \langle\hat{q}\rangle}{{\rm d}t} &=& \frac{\langle\hat{p}\rangle}{m}\label{dqdt}\\ \frac{{\rm d} \langle\hat{p}\rangle}{{\rm d}t} &=& -V'(\langle\hat{q}\rangle)- \frac{1}{2}V'''(\langle\hat{q}\rangle) (\Delta q)^2\label{dpdt}\\ \frac{{\rm d}(\Delta q)^2}{{\rm d}t} &=& \frac{2}{m}\Delta(qp) \label{dqqdt} \\ \frac{{\rm d}\Delta(qp)}{{\rm d}t} &=& \frac{1}{m} (\Delta p)^2- V''(\langle\hat{q}\rangle) (\Delta q)^2\label{dqpdt}\\ \frac{{\rm d}(\Delta p)^2}{{\rm d}t} &=& -2V''(\langle\hat{q}\rangle)\Delta(qp)\,. \label{dppdt} \end{aligned}$$ For a given potential, one may solve these equations numerically. However, it would be more instructive to compute $(\Delta q)^2$ and insert it in (\[dpdt\]) to see what quantum corrections result. So far, all equations are coupled to one another and one cannot solve independently for $(\Delta q)^2$ (unless $V''$ is constant, the case of the harmonic oscillator, a constant force or a free particle). But with an additional adiabatic approximation for the moments, decoupling can be achieved. ### Adiabatic approximation {#s:Adia} To zeroth order in an adiabatic approximation, we assume the moments (but not expectation values) to be time independent. We will denote the adiabatic order by an integer subscript. Eqs. (\[dqqdt\]) and (\[dppdt\]) then imply $\Delta_0(qp)=0$ at zeroth adiabatic order, and (\[dqpdt\]) shows that $(\Delta_0 p)^2=mV''(\langle\hat{q}\rangle) (\Delta_0 q)^2$. With the last equation, we see that the zeroth-order adiabatic approximation cannot be valid unless $\langle\hat{q}\rangle$ is constant in time as well. To avoid such a restrictive condition, we proceed to higher adiabatic orders, from order $i$ to order $i+1$ by inserting time derivatives of $\Delta_i(q^bp^c)$ on the left-hand sides of (\[dqqdt\])–(\[dppdt\]) to compute $\Delta_{i+1}(q^bp^c)$ on the right-hand sides. (For a systematic implementation of the adiabatic approximation, see [@EffAc; @HigherTime].) With time derivatives known from preceding orders, the equations to solve for the moments are initially algebraic, but additional consistency conditions relating different orders sometimes imply differential equations for coefficients, as we will see in this example. To first adiabatic order, $$\begin{aligned} \Delta_1(qp)&=&\frac{1}{2}m \frac{{\rm d}(\Delta_0 q)^2}{{\rm d}t} = -\frac{1}{2V''(\langle\hat{q}\rangle)} \frac{{\rm d}(\Delta_0p)^2}{{\rm d}t} \label{Delta1qp}\\ &=& -\frac{1}{2}m\left( \frac{V'''(\langle\hat{q}\rangle)}{V''(\langle\hat{q}\rangle)} \frac{{\rm d}\langle\hat{q}\rangle}{{\rm d}t} (\Delta_0q)^2+ \frac{{\rm d} (\Delta_0q)^2}{{\rm d}t}\right) \end{aligned}$$ using (\[dqqdt\]), (\[dppdt\]) and our zeroth-order condition relating $(\Delta_0p)^2$ to $(\Delta_0q)^2$. The two lines can both hold only if $$\frac{{\rm d}(\Delta_0q)^2}{{\rm d}t} =- \frac{1}{2} \frac{V'''(\langle\hat{q}\rangle)}{V''(\langle\hat{q}\rangle)} \frac{{\rm d}\langle\hat{q}\rangle}{{\rm d}t} (\Delta_0q)^2\,,$$ solved by $$\label{Delta0q} (\Delta_0q)^2= \frac{C}{\sqrt{V''(\langle\hat{q}\rangle)}}$$ with a constant $C$. Our zeroth-order adiabatic relation between the moments then shows that $(\Delta_0p)^2= mV''(\langle\hat{q}\rangle) (\Delta_0q)^2= mC \sqrt{V''(\langle\hat{q}\rangle)}$. Inserting these solutions in the quantum Hamiltonian (\[HQsecond\]), we obtain a correction $$\label{EffPotsecond} \frac{1}{2m}(\Delta_0p)^2+ \frac{1}{2}V''(\langle\hat{q}\rangle) (\Delta_0q)^2= C\sqrt{V''(\langle\hat{q}\rangle)}$$ to the classical Hamiltonian. As one goes to higher orders in the adiabatic approximation, one takes more and more time derivatives of $\Delta_0(q^bp^c)$. We can see this feature already with the low-order equations found here. So far, we have used the first adiabatic order only to restrict the zeroth-order solutions. But with the solution (\[Delta0q\]) found for $(\Delta_0q)^2$, we obtain from (\[Delta1qp\]) the moment $$\Delta_1(qp)= \frac{1}{2}m\frac{{\rm d}(\Delta_0q)^2}{{\rm d}t}= -\frac{1}{4} Cm \frac{V'''(\langle\hat{q}\rangle)}{V''(\langle\hat{q}\rangle)^{3/2}} \frac{{\rm d}\langle\hat{q}\rangle}{{\rm d}t}\,,$$ depending on a first-order derivative of $\langle\hat{q}\rangle$. We do not need $\Delta(qp)$ in the quantum Hamiltonian, but this pattern continues for all moments at higher adiabatic orders, including $\Delta_i(q^n)$. When we go beyond second adiabatic order and insert solutions into expectation-value equations, higher-derivative effective equations will be obtained; see [@HigherTime] for explicit derivations. Quantum back-reaction by moments is responsible for these higher-derivative corrections, but there is no direct correspondence between the moments as independent quantum degrees of freedom and new degrees of freedom that appear in higher-derivative equations because more initial values need to be specified. It is not the moment expansion itself which gives rise to higher derivatives, but rather the adiabatic expansion of individual moments. The order of moments corresponds to a semiclassical expansion, according to $\Delta(q^bp^c)\sim O(\hbar^{(b+c)/2})$ in semiclassical states, not to a derivative expansion. Any fixed order in $\hbar$ can produce arbitrarily high orders of time derivatives if the adiabatic expansion is pushed further. ### State dependence The parameter $C$ in (\[Delta0q\]), related to second-order moments, is of the order $\hbar$ in semiclassical states; the correction (\[EffPotsecond\]) is therefore the first-order semiclassical correction under the assumption of zeroth adiabatic order for the moments. We cannot choose arbitrary values for $C$ because the uncertainty relation (\[Uncert\]) must be obeyed, such that $C=m^{-1/2} \Delta_0q\Delta_0p\geq \frac{1}{2}\hbar/\sqrt{m}$. Requiring the uncertainty relation to be saturated determines $C$. In general, this condition may be too strong because we would assume saturation at all times, amounting to the existence of a dynamical coherent state which is not guaranteed for general potentials. But to zeroth adiabatic order, with the solutions found here, such an assumption is consistent: all dependence on $\langle\hat{q}\rangle$ drops out in the product of $(\Delta_0q)^2(\Delta_0p)^2$ (and we have $\Delta_0(qp)=0$). Without additional assumptions on the states solved for, or initial conditions for the moment equations (\[dqqdt\])–(\[dppdt\]), the constant $C$ remains undetermined. One possibility to fix $C$, in the class of models of this example, is to assume that solutions are close to the harmonic-oscillator vacuum or some other specific state. If the potential $V(q)=\frac{1}{2}m\omega^2q^2$ is harmonic, $(\Delta_0q)^2= C/\sqrt{m\omega^2}$ is constant — in this case there are states for which the adiabatic approximation is exact — and equals the Gaussian spread $\sigma^2$ in a coherent state: we may write $C=\sigma^2\sqrt{m\omega^2}$. For the harmonic oscillator, dynamical coherent states do exist and the uncertainty relation may be satisfied at all times. In this case, $C=\frac{1}{2}\hbar/\sqrt{m}$, or $(\Delta_0q)^2= \frac{1}{2}\hbar/m\omega$, the correct relation for position fluctuations in the ground state. With $(\Delta_0p)^2=\frac{1}{2}m\hbar\omega$, the non-classical terms in the quantum Hamiltonian amount to the zero-point energy $\frac{1}{2}\hbar\omega$. For a general potential $V$, we do not have the frequency parameter $\omega$ to refer to, but we can define it as the square root of $2/m$ times the coefficient of the quadratic term in a Taylor expansion $V(q)=V_0+ V_1q+\frac{1}{2}m\omega^2q^2+\cdots$, assuming that the coefficient is not zero. In this way, we treat higher than second-order terms in the potential as an anharmonicity. Specifying the class of states solved for as those that are close to a harmonic-oscillator ground state, we can therefore write $(\Delta_0q)^2= \frac{1}{2}\hbar/\sqrt{mV''(\langle\hat{q}\rangle)}$. The correction $\frac{1}{2}\hbar\sqrt{V''(\langle\hat{q}\rangle)/m}$ in the effective Hamiltonian (\[EffPotsecond\]) then agrees with that found for the low-energy effective action [@EffAcQM], a relation that holds to higher adiabatic orders as well [@EffAc]. The canonical picture of quantum back-reaction provides an interpretation of moment-coupling terms as an analog of loop diagrams in quantum field theory, with moments taking the place of $n$-point functions. A formulation of the canonical effective scheme for quantum field theory is not fully worked out yet, but its implications for quantum gravity and cosmology can nevertheless be seen. Already in minisuperspace models there are characteristic effects which show cosmological implications of quantum corrections. ### Notes on the WKB approximation The WKB approximation is often seen as implementing a semiclassical regime, in the sense that leading terms in powers of $\hbar$ are considered in the quantum evolution equation for states, expanded as $\psi(q)= \exp\left(i\hbar^{-1}\sum_{n=0}^{\infty} \hbar^n S_n(q)\right)$ with an asymptotic series. With this ansatz in the Schrödinger equation, one can solve order by order in $\hbar$ to find expressions for the $S_n$: in quantum mechanics, $$\frac{1}{2m} \left(\frac{{\rm d}S_0}{{\rm d}q}\right)^2+ V(q)=E \quad,\quad i\frac{{\rm d}^2S_0}{{\rm d}q^2}+ 2\frac{{\rm d}S_0}{{\rm d} q} \frac{{\rm d}S_1}{{\rm d}q}=0$$ for zeroth and first order in $\hbar$ implies $S_1=-\frac{1}{4}i \log(2m(E-V(q))+{\rm const}$, while $S_0$ satisfies the classical Hamilton–Jacobi equation. Solutions obtained by the WKB approximation do not directly provide observables such as expectation values, for which additional integrations would be necessary. Such integrations are usually complicated to perform not just analytically but also numerically, given the strongly oscillating nature of WKB solutions in semiclassical regimes. Moreover, WKB solutions do not show how quantum corrections can be included in classical equations as the dominant quantum effects. In particular, although quantum back-reaction is implicitly contained in solutions to the WKB equations, it does not appear in the form of effective potentials or quantum forces useful for intuitive explanations of quantum effects. In the WKB approximation, $S_0$ satisfies exactly the classical Hamilton–Jacobi equation, without any quantum corrections. Corrections to the dynamics arise by higher orders of $S_n$ in the wave function, but they do not appear in a form added to the Hamilton–Jacobi (or another classical) equation. While the WKB approximation, as an expansion in $\hbar$, does have a semiclassical flavor, it can more generally be viewed as a formal expansion to produce solutions for wave functions. The WKB equations are obtained by solving the Schrödinger equation exactly at every order of $\hbar$: An equation $\sum_{n=0}^{\infty}E_n\hbar^n=0$ is interpreted as implying $E_n=0$ for all $n$. From a semiclassical perspective, on the other hand, one would interpret an equation $\sum_{n=0}^{\infty}E_n\hbar^n=0$ as providing a tower of quantum corrections $\sum_{n=1}^{\infty}E_n\hbar^n$ to the classical expression $E_0$, and then be interested in solutions to the equations $\sum_{n=0}^{N}E_n\hbar^n=0$ cut off at finite orders of $\hbar$. Additional consistency conditions are needed to determine the $E_n$ showing up in quantum corrections. Usually, the $E_n$ for $n>0$ depend on state parameters such as fluctuations, while $E_0$ depends only on expectation values and equals the classical expression. A dynamical equation $\sum_{n=0}^{N}E_n\hbar^n=0$ then encodes the quantum back-reaction of state parameters on the expectation values, implying deviations from classical behavior, as derived systematically by effective equations. In principle, one could derive such quantum corrections from WKB solutions by computing expectation values of the $\hbar$-expanded wave functions. But the WKB approximation does not automatically arrange the terms in its equation by semiclassical relevance. While canonical effective equations have a direct correspondence to the low-energy effective action, as already seen, the WKB approximation does not produce all terms [@EffAcWKB]. Another question, important in the context of quantum gravity and quantum cosmology, is the treatment of quantum constraints (or the physical Hilbert space), which remains open in the context of WKB solutions. (For instance, one may solve $\hat{H}|\psi\rangle=0$ with WKB techniques, but for approximate solutions, the gauge flow $\exp(-i\hat{H}[\epsilon]/\hbar)|\psi\rangle_{\rm WKB}$ does not automatically vanish.) Canonical effective techniques, on the other hand, apply to constrained systems as well and even help to solve some long-standing conceptual problems of quantum gravity related to constraints and gauge. ### Effective constraints and the problem of time {#s:EffCons} As already indicated in Section \[s:Gauge\], a quantum constrained system with constraint operators $\hat{C}$ produces quantum constraints $C_Q:=\langle\hat{C}\rangle$, defined just like a quantum Hamiltonian (\[HQ\]), but also independent quantum phase-space functions $C_f=\langle (f(\hat{q},\hat{p})-\langle f(\hat{q},\hat{p})\rangle)\hat{C}\rangle$ (in this ordering) constrained to vanish in physical states [@EffCons; @EffConsRel]. In semiclassical expansions, calculating order by order in the moments, polynomial $f(q,p)$ are sufficient. To fixed order in the moments, only finitely many constraints are then present. Their number is larger than the number of classical constraints because they remove not only expectation values of constrained degrees of freedom but also the corresponding moments. With the ordering of constraint operators to the right of $f(q,p)$ chosen in effective constraints, they are automatically first class if the constraint operators are first class. There are then constraint equations to be solved, and gauge flows to be factored out. The gauge flow is computed using the Poisson brackets (\[Poisson\]), affecting also the moments. Standard techniques of constrained systems can then be used, except that moments truncated to a fixed order usually define a non-symplectic Poisson manifold. This feature requires some care and may have several consequences, for instance that the number of independent gauge flows does not equal the number of first-class constraints. Nevertheless, the usual classification of constraints and gauge flows applies [@brackets]. To see the treatment of effective constraints we consider a Hamiltonian constraint operator $\hat{C}=\hat{p}_{\phi}^2- \hat{p}^2+W(\hat{\phi})$ for a free, massless relativistic particle $(q,p)$ coupled to a second degree of freedom $(\phi,p_{\phi})$ with an arbitrary $\phi$-dependent potential $W(\phi)$. Depending on the form of $W(\phi)$, $p_{\phi}$ may become zero along trajectories generated by the Hamiltonian constraint, in which case $\phi$ does not serve as global internal time. On the other hand, with a $q$-independent Hamiltonian constraint, we could deparameterize by $q$, obtaining evolution by the classical Hamiltonian $p= \pm \sqrt{ p_{\phi}^2+W(\phi)}$. We have equations of motion ${\rm d}\phi/{\rm d}q= \pm p_{\phi}/\sqrt{ p_{\phi}^2+W(\phi)}$ and ${\rm d}p_{\phi}/{\rm d}q= \mp \frac{1}{2}W'(\phi)/\sqrt{ p_{\phi}^2+W(\phi)}$. The momentum $p_{\phi}$ evolves, and could indeed become zero. The constraint operator gives rise to the effective constraints [@EffTime; @EffTimeLong] $$\begin{aligned} C_Q&=&\langle\hat{p}_{\phi}\rangle^2-\langle\hat{p}\rangle^2+ (\Delta p_{\phi})^2-(\Delta p)^2+W(\langle\hat{\phi}\rangle)+ {\textstyle\frac{1}{2}}W''(\langle\hat{\phi}\rangle)(\Delta \phi)^2 \label{Cnon}\\ C_{\phi}&=& 2\langle\hat{p}_{\phi}\rangle\Delta(\phi p_{\phi})+ i \hbar \langle\hat{p}_{\phi}\rangle -2p\Delta(\phi p)+W'(\langle\hat{\phi}\rangle) (\Delta \phi)^2\\ C_{p_{\phi}}&=& 2\langle\hat{p}_{\phi}\rangle(\Delta p_{\phi})^2- 2\langle\hat{p}\rangle\Delta(p_{\phi}p)+ W'(\langle\hat{\phi}\rangle)(\Delta(\phi p_{\phi})- {\textstyle\frac{1}{2}}i\hbar)\end{aligned}$$ expanded to second order in the moments, together with additional constraints $C_q$, $C_p$, $C_{qp}$ and so on, which we will not make use of. These constraints can be solved to find the quantum-corrected constraint surface, and their gauge flows can be computed to find moments of observables in physical states. Once the non-symplectic nature of the Poisson manifold of second-order moments is taken into account, these calculations are not very different from standard procedures. The effective constraints shown here illustrate another important feature: the complexity of constraints and their solutions. It comes about because effective constraints, to be first class, are defined in a non-symmetric ordering, while moments are by definition Weyl ordered. Reorderings required to express effective constraints as functions of the moments then introduce imaginary contributions by the commutator $[\hat{q},\hat{p}]=i\hbar$. For $C_{\phi}$ and $C_{p_{\phi}}$ to vanish, some moments must be complex. While moments before the imposition of constraints, belonging to a kinematical Hilbert space, should be real as the expectation values of Weyl-ordered operators, after solving the constraints one moves to the physical Hilbert space, in general not related to a subspace of the kinematical one. After solving the constraints, the original kinematical moments may therefore take complex values, as long as physical observables of the quantum constrained system are subject to reality conditions. For further consequences, we study the problem of time in this system, using the variable $\phi$ as internal time even though it does not deparameterize the system globally. At the full quantum level, local internal times, free of turning points only for finite ranges of evolution, cannot easily be made sense of: if internal time exists only for a finite range, evolution cannot be unitary even in this range. (See for instance the discussion in [@RovelliTimeModel; @RovelliTime; @RovelliTimeReply]. If states are evolved by local internal times past their turning points, evolution freezes: expectation values are stuck at constant values [@WaldTime; @WaldTimeModels].) This consequence is the reason why the problem of time is much more severe at the quantum level, compared to the classical one. At the effective level, as we will see, the problem of time can be overcome, allowing consistent derivations of observables without using artificial deparameterizations [@EffTime; @EffTimeLong; @EffTimeCosmo]. If $\phi$ is used as (local) internal time, it is not represented as an operator on the resulting physical Hilbert space, whatever it may be. No generally manageable techniques are known to derive physical Hilbert spaces and evolution in non-deparameterizable systems (for some possibilities of Hilbert-space derivations, see e.g.[@WaldTime; @WaldTimeModels; @Master; @MasterTesting]). At the effective level, it is sufficient to distinguish $\phi$ as non-operator time by requiring that its moments in effective constraints vanish, $$\label{Zeitgeist} (\Delta \phi)^2=\Delta(\phi q)=\Delta(\phi p)=0$$ while its expectation value $\langle\phi\rangle$ (denoted without the hat to indicate that $\hat{\phi}$ no longer acts as an operator) will become the time parameter. In fact, the conditions (\[Zeitgeist\]) implement a good gauge fixing of the second-order constraints $C_{\phi}$, $C_{p_{\phi}}$, $C_q$ and $C_p$ after quantization. (With constraints on a non-symplectic Poisson manifold, only three gauge-fixing conditions are required for four constraints. The remaining second-order moment involving $\phi$, $\Delta(\phi p_{\phi})$, is fixed by the constraints, as we will see shortly.) In the terminology of [@EffTime], these conditions implement the Zeitgeist during which $\langle\phi\rangle$ as local internal time is current. Imposing (\[Zeitgeist\]) initiates the transition to physical moments — moments computed for states in the physical Hilbert space on which $\phi$ does not act as an operator. Solving the effective constraints in the given Zeitgeist, we have $\Delta(\phi p_{\phi})=-\frac{1}{2}i\hbar$ from $C_{\phi}=0$, which then implies $$(\Delta p_{\phi})^2= \frac{\langle\hat{p}\rangle^2}{\langle\hat{p}_{\phi}\rangle^2} (\Delta p)^2+ \frac{1}{2}i\frac{W'(\langle\phi\rangle)\hbar}{\langle\hat{p}_{\phi}\rangle}$$ from $C_{p_{\phi}}=0$. Inserted in (\[Cnon\]), this implies the reduced constraint $$\label{Cimag} C=\langle\hat{p}_{\phi}\rangle^2-\langle\hat{p}\rangle^2+ \frac{\langle\hat{p}\rangle^2- \langle\hat{p}_{\phi}\rangle^2}{\langle\hat{p}_{\phi}\rangle^2}(\Delta p)^2+ \frac{1}{2}i\frac{W'(\langle\phi\rangle)\hbar}{\langle\hat{p}_{\phi}\rangle} + W(\langle\phi\rangle)$$ amounting to the quantum constraint $C_Q=\langle\hat{C}\rangle$ on the space on which $C_{\phi}$ and $C_{p_{\phi}}$ are solved in the given Zeitgeist. Solving $C=0$ for $\langle\hat{p}_{\phi}\rangle$, we obtain the time-dependent Hamiltonian for $\langle\phi\rangle$-evolution, including quantum back-reaction. However, it still contains complex terms. In (\[Cimag\]), all terms except the last two are expected to be real-valued because $\langle\hat{p}\rangle$ and $\Delta p$ are physical observables, and $\langle\hat{p}_{\phi}\rangle$ can be interpreted physically as the local energy value. The constraint can then be satisfied, only if we allow for an imaginary part of $\langle\phi\rangle$, calculated from $$\label{Vtime} \frac{1}{2}i\frac{W'(\langle\phi\rangle)\hbar}{\langle\hat{p}_{\phi}\rangle}+ W(\langle\phi\rangle)=0\,.$$ For semiclassical states, to which this approximation of effective constraints refers, we can Taylor expand the potential $$W(\langle\phi\rangle)=W({\rm Re}\langle\phi\rangle+i\,{\rm Im}\langle\phi\rangle)= W({\rm Re}\langle\phi\rangle)+i\,{\rm Im}\langle\phi\rangle\, W'({\rm Re}\langle\phi\rangle)+ O(({\rm Im}\langle\phi\rangle)^2)$$ by the imaginary term, expected to be at least of the order $\hbar$ because it vanishes classically. To this order, the imaginary contribution ${\rm Im}C=0$ to $C$ in (\[Cimag\]) implies that $$\label{imt} {\rm Im}\langle\phi\rangle= -\frac{\hbar}{2\langle\hat{p}_{\phi}\rangle}\,.$$ The remaining terms, $${\rm Re}C= \langle\hat{p}_{\phi}\rangle^2-\langle\hat{p}\rangle^2+ \frac{\langle\hat{p}\rangle^2- \langle\hat{p}_{\phi}\rangle^2}{\langle\hat{p}_{\phi}\rangle^2}(\Delta p)^2 + W({\rm Re}\langle\phi\rangle)=0$$ provide the physical ${\rm Re}\langle\phi\rangle$-Hamiltonian upon solving the constraint equation for $\langle\hat{p}_{\phi}\rangle$. At this stage, the Hamiltonian and its solutions, corresponding to evolving observables with respect to $\phi$, are all real: physical reality conditions are imposed and we have solutions corresponding to states in the physical Hilbert space. Although imaginary parts may be unexpected, a detailed analysis of this and other models shows that they are fully consistent [@EffTimeLong]. In models in which one can compute a physical Hilbert space, results equivalent with those shown here are obtained. Within the effective treatment of constraints, if we transform to a different internal time such as $q$, which is done by a gauge transformation in the effective constrained system so that a new Zeitgeist — the gauge-fixing (\[Zeitgeist\]) — is realized, the imaginary parts are automatically transferred from $\langle\phi\rangle$ to $\langle q\rangle$, in such a way that observables remain real. By successive gauge transformations, one can evolve through turning points of local internal times, without freezing the evolution of physical observables; see in particular the cosmological example analyzed in [@EffTimeCosmo]. The imaginary part of time can be seen as a remnant of non-unitarity problems of evolution in local-time quantum systems, but unlike in Hilbert-space treatments, it does not pose any problems at the effective level. Gauge transformations in effective constrained systems show that physical results are independent of the choice of (local) internal time. One may deparameterize the effective system in different ways to solve the resulting equations, without affecting observables. This conclusion, one example for effective solutions to the traditional problems of canonical quantum gravity, indicates that deparameterization can be used consistently. However, in complicated systems subject to ambiguities such as factor-ordering choices, each deparameterization must be formulated in a specific way so that they all can result from one non-deparameterized system, effective or not. At the effective level, all quantum constraints and Zeitgeists must be computed and implemented with the same operator $\hat{C}$ for different local time choices to produce mutually consistent results. In many constructions of physical Hilbert spaces, however, one quantizes a system with a specific deparameterization in mind, choosing factor orderings and using possible simplifications. In such a case, there is no guarantee that results can agree with those obtained from other parameterizations, and the independence of physical results of the choice of time is put at risk. Modified Friedmann equations, or: the sins of sines {#s:ModFried} --------------------------------------------------- In quantum cosmology, a systematic derivation of quantum back-reaction is required especially for reliable evaluations of holonomy corrections in the Hamiltonian constraint, as they both are relevant in high-curvature regimes and contribute to higher-curvature terms. Constraints appear in such systems, but for simplicity we will refer to deparameterized toy models. Holonomy corrections have provided a popular class of models within loop quantum cosmology, in which the Friedmann equation is modified in a simple way [@AmbigConstr]: The classical constraint equivalent to the Friedmann equation is first modified to $$\label{Hmod} H_{\rm mod}= -\frac{3}{8\pi G} \frac{\sin^2(\ell c)}{\gamma^2\ell^2} \sqrt{|p|}+ \rho |p|^{3/2}=0$$ with a holonomy parameter $\ell$ that could possibly depend on $p$. According to loop quantum cosmology [@LivRev; @Springer], the Hamiltonian is written in canonical triad and connection variables, with $A_a^i=c\delta_a^i$ and $E^b_j=p\delta^b_j$, $\{c,p\}=8\pi\gamma G/3$, under the assumption of isotropy. The densitized-triad component $p$ can be positive and negative, according to the orientation of the triad, and is related to the scale factor by $|p|=a^2$. (Without loss of generality regarding effective equations, we take $p$ to be positive in what follows.) For spatially flat models, as assumed here, $c=\gamma\dot{a}$ is proportional to the proper-time derivative of the scale factor. By the modification in (\[Hmod\]), the periodic form of holonomies is implemented, replacing the quadratic connection dependence of the classical expression. Computing Hamiltonian equations of motion for $p$ allows us to eliminate $c$ in favor of $\dot{p}$, upon which the constraint equation takes the form of some kind of Friedmann equation. We have $\dot{p}=\{p,H_{\rm mod}\}= (\gamma\ell)^{-1} \sin(2\ell c)\sqrt{p}$. Using trigonometric identities, we find $\sin^2(\ell c)= \frac{1}{2}(1-\sqrt{1-4\gamma^2\ell^2\dot{a}^2})$ with $\dot{a}=\dot{p}/(2\sqrt{p})$. Inserting this in the modified constraint and solving for $\dot{a}^2$, the modified Friedmann equation becomes $$\label{ModFried} \left(\frac{\dot{a}}{a}\right)^2= \frac{8\pi G}{3} \rho\left(1-\frac{8\pi G}{3} \gamma^2\ell^2 a^2\rho\right)\,.$$ This simple and interesting equation, with just a quadratic correction to the energy density, has served as the basis of many ad-hoc investigations of potential effects of loop quantum cosmology. Equation (\[ModFried\]) is clearly not an effective equation in the generality written here. Quantum back-reaction is ignored, while all terms in the complete series expansion of holonomy corrections $$\label{sinexp} \frac{\sin^2(\ell c)}{(\ell c)^2}-1=-\frac{1}{3}\ell^2c^2+ \frac{4}{45} \ell^4c^4+\cdots$$ are taken into account for the calculation. A consistent treatment would include only those higher-order terms in an expansion of holonomy modifications that are larger than any quantum back-reaction or other term that has been ignored. The relation of quantum back-reaction to higher-curvature corrections indicates that $c^2$-corrections in (\[sinexp\]) should be of comparable size to $\dot{c}$-corrections from quantum back-reaction, the latter of which are not included in (\[ModFried\]). Including holonomy corrections but ignoring quantum back-reaction is therefore inconsistent, even at leading order in the $c$-expansion, unless one considers only models in which quantum back-reaction is weak. (Such models do indeed exist, as we will show later, but they are very special.) Keeping all terms in the $c$-expansion to arbitrary orders then leads to a questionable equation. One could think that keeping small higher-order terms is harmless, but it turns out that our cautionary considerations do matter for the form of modified Friedmann equations. To see this concretely, let us look at a few examples in which the holonomy modification in the Hamiltonian constraint is expanded first, followed by a calculation of Hamiltonian equations of motion and a modified Friedmann equation. The first order of $c$-corrections provides a constraint $$H_1=-\frac{3}{8\pi G}\frac{c^2}{\gamma^2} \left(1-\frac{1}{3} \ell^2c^2\right) \sqrt{p}+\rho p^{3/2}=0\,.$$ Proceeding as before, we compute $\dot{p}=\{p,H_1\}= 2\gamma^{-1}c(1-\frac{2}{3}\ell^2c^2)\sqrt{p}$, solve for $c$ in terms of $\dot{p}$, insert the result in $H_1$, and rewrite as $$\left(\frac{\dot{a}}{a}\right)^2= \frac{8\pi G}{3} \rho\left(1-\frac{8\pi G}{3} \gamma^2\ell^2 a^2\rho\right)\,.$$ Rather surprisingly, the result agrees exactly with the one obtained with the full holonomy modification, (\[ModFried\]). However, this outcome does not mean that higher orders in the $c$-expansion do not matter. It rather shows that the specific form of the full modification by $\sin^2(\ell c)$ is arranged so delicately that all higher-order contributions beyond the $c^2$-correction precisely cancel one another. To confirm this, we go one order beyond the quadratic correction, modifiying the Hamiltonian constraint by $$H_2= -\frac{3}{8\pi G}\frac{c^2}{\gamma^2} \left(1-\frac{1}{3} \ell^2c^2 +\frac{4}{45}\ell^4c^4\right) \sqrt{p}+\rho p^{3/2}=0\,.$$ Again we proceed as before. (The higher higher-order polynomial equations to be solved for a relation of $c$ to $\dot{p}$ can be handled easily within the perturbative scheme of the $c$-expansion.) The result, $$\left(\frac{\dot{a}}{a}\right)^2= \frac{8\pi G}{3} \rho\left(1-\frac{8\pi G}{3} \gamma^2\ell^2 a^2\rho+ \frac{157}{45} \left(\frac{8\pi G}{3}\right)^2 \gamma^4\ell^4 a^4\rho^2\right)\,,$$ now has a higher-than-quadratic correction in the energy density. Going to higher orders in $c$ and following this scheme shows that also the energy-order increases to include all possible powers. The leading corrections in $\rho$, such as $(8\pi G/3) \gamma^2\ell^2 a^2\rho$ with the same coefficient in all modified Friedmann equations, do not change if one goes to higher orders in the $c$-expansion and can therefore be used consistently — provided one stays in energy ranges in which it is the dominant term. When the energy density approaches Planckian levels and holonomy corrections are strong, however, $8\pi G\gamma^2\ell^2a^2 \rho$ is close to one and all terms in the energy expansion are relevant. Bounce scenarios, for instance, cannot be formulated with a consistent version of the equation. (In this context, notice that the next term beyond $\rho^2$ enters with a positive sign. At high density, it may well be larger than the correction in (\[ModFried\]), in which case no zero of $\dot{a}$ and no bounce would be reached.) As anticipated, the sine-modification has its infinitely many higher-order terms arranged such that all but the quadratic energy correction disappear. A consistent perturbative treatment keeping only the relevant orders instead produces a whole series expansion by the energy density. If one has reasons to trust the whole sine function and to exclude all other corrections, (\[ModFried\]) is correct. But if there are additional corrections, however weak, it is not consistent to keep all terms in a $c$-expansion of the sine function; instead, one has a modified Friedmann equation with a perturbative expansion in $\rho$. Those additional corrections then unhinge the fine balance in the sine terms that eliminated all $\rho$-corrections beyond second order, and their form must be known for a reliable derivation of correct effective Friedmann equations. The main source of such extra terms is, of course, quantum back-reaction, producing higher time derivatives that compete with higher powers of $c$. (Higher orders are also sensitive to quantization ambiguities, as analyzed for instance in [@AmbigBounce].) Harmonic cosmology ------------------ To understand the interrelation between different corrections, we should have a more detailed look at quantum back-reaction in quantum cosmology. In an effective description of Wheeler–DeWitt minisuperspace models, one considers the dynamics of expectation values $\langle\hat{a}\rangle$ and $\langle\hat{p}_a\rangle$ coupled to fluctuations and higher moments $\Delta(a^bp_a^c)$. The coupled dynamics, including quantum back-reaction, is usually complicated and unruly, but it simplifies considerably if perturbations around a simple model such as the harmonic oscillator in quantum mechanics can be used. As an analog of the harmonic oscillator in quantum mechanics with simple effective equations, quantum cosmology has a harmonic model given by a free, massless scalar in a spatially flat isotropic geometry [@BouncePert]. To realize the model as one with a Hamiltonian generating evolution, we must pick a time variable which we do by parameterizing, using the scalar $\phi$ as time. Since it is free and massless, the Hamiltonian constraint $$H(a,p_a,p_{\phi})= -\frac{2\pi G}{3} \frac{p_a^2}{a}+ \frac{1}{2}\frac{p_{\phi}^2}{a^3}$$ implies that $p_{\phi}$ is a constant of motion and $\phi$ has no turning points where $p_{\phi}$ would move through zero. The scalar therefore provides a global internal time. Deparameterized models, as discussed before, cannot produce reliable physical predictions unless one can show that results do not depend on the choice of time. In the present context, we use the model merely to illustrate properties of quantum cosmological dynamics. For realistic effects, one can avoid deparameterization before quantization and the dependence on time choices by using effective constraints instead of effective deparameterized Hamiltonians [@EffCons; @EffConsRel; @EffTime; @EffTimeLong]. It is an interesting coincidence that the same model is easily deparameterizable and at the same time, as we will see, harmonic, without quantum back-reaction. Both features imply that the model is extremely special even among symmetry-reduced isotropic systems; its implications must therefore be interpreted with a great amount of care. We perform deparameterization by solving the Hamiltonian constraint $H(a,p_a,p_{\phi})=0$ for $$p_{\phi}(a,p_a)= \pm \sqrt{\frac{4\pi G}{3}} |ap_a|\,,$$ the Hamiltonian generating evolution with respect to $\phi$. Equations of motion for $a(\phi)$ and $p_a(\phi)$ are then obtained via Poisson brackets with $p_{\phi}(a,p_a)$. If solutions are to be transferred back to coordinate time, such as proper time, we solve ${\rm d}\phi/{\rm d}\tau=\{\phi,H(a,p_a,p_{\phi})\}= p_{\phi}/a(\phi)^3$ for $\phi(\tau)$ (with a constant $p_{\phi}$) and insert this function in our solutions for $a(\phi)$ and $p_a(\phi)$. ### Effective Wheeler–DeWitt equations Effective deparameterized equations are generated by the quantum Hamiltonian $\langle p_{\phi}(\hat{a},\hat{p}_a)\rangle$, $$\label{dOdphi} \frac{{\rm d}\langle\hat{O}\rangle}{{\rm d}\phi}= \frac{\langle[\hat{O},p_{\phi}(\hat{a},\hat{p}_a)]\rangle}{i\hbar}= \{\langle\hat{O}\rangle,\langle p_{\phi}(\hat{a},\hat{p}_a)\rangle\}$$ using the Poisson brackets (\[Poisson\]). The absolute value in $p_{\phi}(a,p_a)$ makes a completely general expansion in moments complicated, but for $|p_{\phi}|$ not close to zero, there is a simple effective Hamiltonian. If we can ensure positivity of $\widehat{ap_a}$ in evolved states, the absolute value can be dropped. This is possible in particular for an initial state supported solely on the positive part of the spectrum of $\widehat{ap_a}$ (an operator for which we will assume Weyl ordering). Since $ap_a$ is preserved by the motion it generates, also after quantization, the evolved state will remain supported on the positive part of the spectrum of $\widehat{ap_a}$. Unless $|p_{\phi}|$ is close to zero, it is easy to find initial states supported only on the positive part of the spectrum of $\widehat{ap_a}$ and with specified initial expectation values for $\hat{a}$ and $\hat{p}_a$: Projecting out negative contributions will not change the basic expectation values much. We are then allowed to write (\[dOdphi\]) as $$\begin{aligned} \frac{{\rm d}_+\langle\hat{O}\rangle_+}{{\rm d}\phi} = \pm \sqrt{\frac{4\pi G}{3}} \frac{_+\langle [\hat{O},\widehat{|ap_a|}]\rangle_+}{i\hbar}= \pm \sqrt{\frac{4\pi G}{3}} \frac{_+\langle [\hat{O},\widehat{ap_a}]\rangle_+}{i\hbar}\nonumber\\ = \pm \sqrt{\frac{4\pi G}{3}} \{_+\langle\hat{O}\rangle_+, _+\langle \widehat{ap_a}\rangle_+\} \end{aligned}$$ using $\widehat{|ap_a|}|\psi\rangle_+= \widehat{ap_a}|\psi\rangle_+$ (and $\widehat{|ap_a|}^{\dagger}|\psi\rangle_+= \widehat{ap_a}^{\dagger}|\psi\rangle_+$) on states $|\psi\rangle_+$ with support only on the positive part of the spectrum of $\widehat{ap_a}$. On such positively supported states, the $\phi$-Hamiltonian is quadratic and can easily be expanded in moments. We have the quantum Hamiltonian $$H_Q= \pm \sqrt{\frac{4\pi G}{3}} (\langle\hat{a}\rangle \langle\hat{p}_a\rangle +\Delta(ap_a))\,,$$ free of coupling terms of expectation values and moments: there is no quantum back-reaction. We compute and solve equations of motion for expectation values, resulting in $$\langle\hat{a}\rangle(\phi)= \exp(\pm \sqrt{4\pi G/3}\: \phi) \quad \mbox{and} \quad \langle\hat{p}_a\rangle(\phi)= \exp(\mp \sqrt{4\pi G/3}\: \phi)\,.$$ To transform to proper time, we solve ${\rm d}\phi/{\rm d}\tau= p_{\phi}\exp(\mp \sqrt{12\pi G}\phi)$ for $$\phi(\tau)= \pm\frac{\log(\pm \sqrt{12\pi G}p_{\phi}\tau)}{\sqrt{12\pi G}}$$ and obtain $\langle\hat{a}\rangle(\tau)= (\pm \sqrt{12\pi G} p_{\phi}\tau)^{1/3}$, the classical dependence on proper time with a stiff matter source. (In particular, even after Wheeler–DeWitt quantization the system remains singular: infinite density $p_{\phi}^2/(2\langle\hat{a}\rangle^3)$ is reached at finite proper time.) In addition to these expectation-value solutions, the state evolves such that its second-order moments change by $$\frac{{\rm d}(\Delta a)^2}{{\rm d}\phi}= \pm 2\sqrt{\frac{4\pi G}{3}} (\Delta a)^2 \quad,\quad \frac{{\rm d}(\Delta p_a)^2}{{\rm d}\phi}= \mp 2\sqrt{\frac{4\pi G}{3}} (\Delta p_a)^2$$ and ${\rm d}\Delta(ap_a)/{\rm d}\phi=0$, with solutions such that $(\Delta a)/\langle\hat{a}\rangle$ and $(\Delta p_a)/\langle\hat{p}_a\rangle$ are constant. Semiclassicality is preserved exactly throughout evolution in this harmonic model, even at high density. Note that $\Delta a$ and $\Delta p_a$ change nonetheless, but have often been assumed constant when state evolution was modeled by Gaussians. Wrong quantum corrections then result, which is especially significant in the presence of quantum back-reaction when the harmonic model is generalized. Especially curvature fluctuations are important because they grow when one evolves to high density, and they do show up in effective constraints such as (\[Qconstraint\]) as a simple example. This issue is another illustration of the importance of complete effective equations including the moment dynamics. With additional ingredients such as spatial curvature, a cosmological constant, a scalar mass or self-interaction, anisotropy or inhomogeneity, the system is no longer harmonic and becomes subject to quantum back-reaction. Deviations from the classical trajectory will then occur, to be captured by effective equations. (Some of these ingredients also remove deparameterizability, but at the effective level we can still use local internal times.) With a cosmological constant, for instance, the $\phi$-Hamiltonian becomes $p_{\phi}(a,p_a)= \pm\sqrt{4\pi G/3}\: a \sqrt{p_a^2- 4\Lambda a^4}$, a non-quadratic expression that entails coupling terms between expectation values and moments in a quantum Hamiltonian. ### Harmonic loop quantum cosmology {#s:HarmLQC} At first sight, it seems that quantum-geometry corrections in loop quantum cosmology imply quantum back-reaction by deviations from the quadratic nature if (\[Hmod\]) is used. This expectation is correct for inverse-triad corrections, but holonomy corrections, although they change the quadratic nature by higher-order terms, still lead to a harmonic model free of quantum back-reaction [@BouncePert]. To see this, we first change to connection variables $c=\gamma \dot{a}= -(4\pi\gamma G/3) p_a/a$ and $p=a^2$, with $\{c,p\}=8\pi\gamma G/3$. The $\phi$-Hamiltonian is still quadratic in these variables, proportional to $|cp|$, but the holonomy modification leads us to replace $c$ by $\sin(\ell c)/\ell$ with some $\ell$ that may depend on $p$. After this, the Hamiltonian $p_{\phi}(c,p)$ is no longer quadratic in $c$ and $p$. However, if we introduce a new variable $J:= 3p\exp(i\ell c)/8\pi\gamma G$, we have a linear $\phi$-Hamiltonian $$\label{pphi} p_{\phi}= \pm 2\sqrt{\frac{4\pi G}{3}} \frac{|{\rm Im}J|}{\ell}$$ if $\ell$ is constant. More generally, we can assume a power law for the $p$-dependence of $\ell(p)=\ell_0 p^x$, and define new basic variables $$\label{VJ} V:= \frac{3p^{1-x}}{8\pi \gamma G(1-x)} \quad,\quad J:=V\exp(i\ell_0 p^xc)$$ so that the $\phi$-Hamiltonian remains linear (and is just multiplied with $1-x$ compared to (\[pphi\])). Moreover, and importantly, we have a (non-canonical) closed algebra of basic variables, $$\{V,J\}= -i\ell_0J\quad,\quad \{V,\bar{J}\}=i\ell_0\bar{J}\quad,\quad \{J,\bar{J}\}= 2i\ell_0V\,,$$ with the Hamiltonian a linear combination of the generators. Given these properties, upon quantization the Ehrenfest equations still provide closed equations for expectation values, without coupling to moments and quantum back-reaction. The Hamiltonian operator $$\label{pphiJ} \hat{p}_{\phi}= \pm \sqrt{\frac{4\pi G}{3}} (1-x) \left|\frac{\hat{J}-\hat{J}^{\dagger}}{i\ell_0}\right|$$ is linear in $\hat{J}$ and its adjoint, and the quantum Hamiltonian is linear in $\langle\hat{J}\rangle$ and its complex conjugate. The only additional condition to impose is a reality condition because we have used partially complex variables. If we initially keep $J$ and $\bar{J}$ as independent variables, valid solutions must satisfy $J\bar{J}=V^2$. We quantize by turning $V$ and $J$ into operators, choosing an ordering of $J$ with the exponential to the right. The classical Poisson algebra is then replaced by the closed commutator algebra $$\label{VJAlg} [\hat{V},\hat{J}]= \ell_0\hbar\hat{J}\quad,\quad [\hat{V},\hat{J}^{\dagger}]= -\ell_0\hbar\hat{J}^{\dagger}\quad,\quad [\hat{J},\hat{J}^{\dagger}]= -\ell_0\hbar(2\hat{V}+\ell_0\hbar)\,,$$ where the last $\hbar$ comes from reordering exponentials. (Note that this non-canonical algebra implies $(V,J)$-moments not commuting with expectation values on the quantum phase space.) The reality condition, with the same ordering, reads $\hat{J}\hat{J}^{\dagger}-\hat{V}^2=0$, which implies conditions on moments upon taking an expectation value, possibly preceded by multiplication with basic operators. For second-order moments, we have $$\label{Reality} |\langle\hat{J}\rangle|^2-(\langle\hat{V}\rangle+\ell_0\hbar/2)^2= (\Delta V)^2- \Delta(J\bar{J})+\frac{1}{4}\ell_0^2\hbar^2\,.$$ (For conditions on higher moments, see [@HighDens].) The reality condition is a Casimir of the commutator algebra of type ${\rm sl}(2,{\mathbb R})$ and therefore commutes with the quantum Hamiltonian proportional to $i(\hat{J}-\hat{J}^{\dagger})$. If reality holds for initial expectation values and moments, it holds at all times. This statement is true not only for the harmonic model but for any Hamiltonian because the Casimir commutes with all $\hat{V}$, $\hat{J}$ and $\hat{J}^{\dagger}$ individually, and therefore with any function of these variables. Solutions for expectation values obtained from the linear quantum Hamiltonian are $$\begin{aligned} \langle\hat{V}\rangle(\phi)&=& A\exp(C\phi)+B\exp(-C\phi)\quad\mbox{ and}\\ \langle\hat{J}\rangle(\phi)&=& A\exp(C\phi)-B\exp(-C\phi)+ \frac{i\ell_0}{C} p_{\phi}\,,\end{aligned}$$ with two integration constants $A$ and $B$ as well as the constant $C=\pm2\sqrt{4\pi G/3}(1-x)$. The reality condition then requires that $$|\langle\hat{J}\rangle|^2-\langle\hat{V}\rangle^2= -4AB+ \frac{\ell_0^2p_{\phi}^2}{C^2}$$ is of the order $\langle\hat{V}\rangle\ell_0\hbar$ (the size of semiclassical fluctuations $(\Delta V)^2$ and $\Delta(J\bar{J})$ in (\[Reality\])), much smaller than $p_{\phi}^2$ for a universe which has a large amount of matter and is semiclassical at least once. (Recall that it is sufficient to impose the reality condition at just one time, for instance when semiclassicality is realized.) The product $AB$ must therefore be positive, close to $\ell_0^2p_{\phi}^2/C^2$, and the function $\langle\hat{V}\rangle(\phi)\propto \cosh(C\phi-C\phi_0)$ never becomes zero. Holonomy modifications in the harmonic model replace the classical singularity by a bounce. As already seen in the beginning of this section, the bounce property of holonomy-modified equations is very sensitive to quantum back-reaction (or other possible corrections). With quantum back-reaction, the equations become more complicated but can still be analyzed numerically as long as moments do not become large. The approach to high densities can therefore be studied, but the Planck regime remains poorly controlled. In general, it is not known whether loop models always exhibit a bounce. ### Quantum Friedmann equation As in the case of holonomy-modified constraints, one can put effective equations into the form of quantum Friedmann equations. In the case of harmonic loop quantum cosmology, we eliminate $\langle\hat{J}\rangle$ in favor of $\dot{\langle\hat{V}\rangle}$ by using the effective equations and, importantly, the reality condition. We have ${\rm d}\langle\hat{V}\rangle/{\rm d}\phi=\{\langle\hat{V}\rangle, \langle\hat{p}_{\phi}\rangle\} \propto {\rm Re} \langle\hat{J}\rangle$ using (\[pphiJ\]) and (\[VJAlg\]), which we turn into a proper-time derivative, as needed for a Friedmann-type equation, using ${\rm d}\phi/{\rm d}\tau= p_{\phi}/a(\phi)^3$ where $a(\phi)$ is taken as some power of $\langle\hat{V}\rangle(\phi)$, depending on $x$ according to (\[VJ\]). (More precisely, we write the Friedmann equation in terms of $V$ by $\dot{a}/a= \dot{V}/(2V(1-x))$ and then use $\langle\hat{V}\rangle(\phi)$. There is no direct relation between $\langle\hat{a}\rangle$ and $\langle\hat{V}\rangle$ because these are expectation values of different powers of the basic $\hat{V}$ unless $x=1/2$.) The real part of $\langle\hat{J}\rangle$, which is proportional to ${\rm d}\langle\hat{V}\rangle/{\rm d}\phi$, is related to the imaginary part by the reality condition (\[Reality\]). The imaginary part, finally, is proportional to the deparameterized quantum Hamiltonian $p_{\phi}$. When all these relations are used and $p_{\phi}$ is expressed via the energy density of the free massless scalar assumed here, we have $$\label{ReJ} {\rm Re}\langle\hat{J}\rangle= \pm\langle\hat{V}\rangle\sqrt{1-(8\pi G/3)\gamma^2(\ell a)^2\rho_Q}$$ with $$\rho_Q=\rho_{\rm free}+ (3/8\pi G)(\gamma\ell a)^{-2} (\Delta(J\bar{J})-(\Delta V)^2)/\langle\hat{V}\rangle^2$$ the free energy density corrected by moment terms. If (\[ReJ\]) is squared and suitable factors are inserted to express all terms by $a$, an equation for $(\dot{a}/a)^2$ follows: $$\label{HarmFried} \left(\frac{\dot{a}}{a}\right)^2= \frac{8\pi G}{3}\rho_{\rm free} \left(1-\frac{8\pi G}{3}\gamma^2(\ell a)^2\rho_Q\right)\,.$$ Except for the fluctuation terms in $\rho_Q$, this is Eq. (\[ModFried\]). There is another, more important difference: The derivation of (\[HarmFried\]) holds only in the harmonic model, in which moments enter just by the reality condition, not by quantum back-reaction. The quantum Friedmann equation is valid in this form only if the sole matter source is a free, massless scalar, and there is no spatial curvature, a cosmological constant, or deviations from isotropy. When any one of these conditions is violated, quantum back-reaction results and there are additional corrections, not just in $\rho_Q$ but also in the general form of (\[HarmFried\]): As already anticipated by considering holonomy expansions in Section \[s:ModFried\], a whole series of corrections in a density expansion appears. The quantum Friedmann equation reads $$\label{EffFriedW} \left(\frac{\dot{a}}{a}\right)^2 = \frac{8\pi G}{3}\left(\rho \left(1- \frac{8\pi G}{3}\gamma^2(\ell a)^2 \rho_Q\right) \pm\frac{1}{2}\sqrt{1-\frac{8\pi G}{3}\gamma^2(\ell a)^2\rho_Q} \:\eta W+ \frac{a^6W^2}{2p_{\phi}^2}\eta^2 \right)$$ where $W(\phi)$ is a possible scalar potential, and we have a general quantum parameter $\eta=\sum_k\eta_{k+1}(a^6W/p_{\phi}^2)^k$ with coefficients $\eta_k$ that depend on the moments, especially correlation parameters [@QuantumBounce; @BounceSqueezed]. The expansion by $a^6W/p_{\phi}^2$ can be interpreted as one by $(\rho-P)/(\rho+P)$ with pressure $P$. (For a free, massless scalar, $\rho=P$.) A more-specific evaluation requires the detailed computation of quantum back-reaction to analyze how the moments evolve and what values they take especially at high density. Techniques for numerical studies have been provided in [@HighDens]. While it is difficult to find general information about the values of moments, it is clear that they contribute, among other effects, the canonical analog of higher-time derivatives as they appear in higher-curvature corrections. Moment terms and quantum back-reaction should therefore be large in high-density regimes, near the big bang. Reliable results in loop quantum cosmology can only state that the singularity is avoided by a bounce when matter is kinetic dominated, in which case $W\ll \rho_{\rm kin}$ and the $\eta$-dependent terms in (\[EffFriedW\]) can be ignored unless $\eta$ is extremely large. This conclusion coincides with numerical investigations [@APS; @APSCurved; @NegCosNum] of the underlying difference equation for wave functions. However, such numerical studies suffer from the choices required for wave functions, for instance by an initial state. With such methods, it is difficult to capture general-enough effects, which in quantum cosmology with its lack of distinguished states make robust conclusions difficult. As shown by the generality of the quantum Friedmann equations displayed here, effective techniques allow one to draw conclusions and confirm the regime-dependent validity of some effects even when no specific states are chosen. ### Cosmic forgetfulness and signature change There are additional and more-surprising properties of the high-density regime that invalidate a traditional bounce interpretation even in the harmonic model. First, staying in the isotropic context, there is cosmic forgetfulness [@BeforeBB; @Harmonic]: When crossing the bounce in $\phi$-evolution, some moments change in ways so sensitive to initial values that the pre-bounce state cannot be recovered precisely from what may be known post bounce. Using solutions of moment equations [@Harmonic] or considerations of semiclassical wave functions [@LoopScattering], one can derive an inequality $$\label{Forget} \left|1-\frac{(\Delta V)_{\phi\to\infty}}{(\Delta V)_{\phi\to-\infty}}\right|\leq \frac{\Delta p_{\phi}/\langle\hat{p}_{\phi}\rangle}{(\Delta V/\langle\hat{V}\rangle)_{\phi\to\infty}}$$ bounding the ratio of volume fluctuations at early and late times. For a state with fluctuations symmetric around the high-density regime near the minimum of $\langle\hat{V}\rangle(\phi)$, the left-hand side would be near zero. The estimate can therefore be used to shed light on the question of how much a quantum state may change while evolving through high density, and how much possible changes can be controlled. With matter fluctuations $\Delta p_{\phi}/\langle\hat{p}_{\phi}\rangle$ usually much larger than geometry fluctuations $(\Delta V/\langle\hat{V}\rangle)_{\phi\to\infty}$ at large volume, where quantum field theory on curved space-time should be a good approximation, the right-hand side of the inequality is much larger than one, and $(\Delta V)_{\phi\to-\infty}$ can differ significantly from $(\Delta V)_{\phi\to\infty}$. The inequality can be saturated by highly squeezed dynamical coherent states [@Harmonic], showing that control on pre-bounce fluctuations cannot be improved unless states are restricted further, more strongly than by semiclassicality. Several classes of specific states, especially ones with weak correlations of the canonical variables — the volume and the Hubble parameter — show more-symmetric behavior of volume fluctuations. Most wave functions that can be constructed explicitly, for instance using ${\rm sl}(2,{\mathbb R})$-coherent states based on the algebra (\[VJAlg\]) as used in [@GroupLQC], are only weakly correlated and do not show all possible asymmetries. Again, the effective viewpoint using moments instead of wave functions provides larger generality. And even though wave functions are not provided in explicit terms, one can show that wave functions for the moment solutions even at saturation of (\[Forget\]) do indeed exist: examples for such states are dynamical coherent states [@BounceCohStates] saturating the uncertainty relation, whose existence can be shown by general methods well-known from quantum mechanics. The second feature preventing a bounce interpretation brings us back to quantum space-time structure. For a reliable cosmological model, we must embed a holonomy-modified isotropic version within a consistent deformation of the constraint algebra. Only then can we be sure that the model describes consistent evolution of quantum space-time. No complete extension of holonomy corrections to inhomogeneity is known, but as we will see in the next section, existing versions of holonomy modifications at high density imply drastic modifications with signature change, turning space-time into a quantum version of 4-dimensional Euclidean space. This happens right where the bounce would be, but without time and evolution in Euclidean space, a bounce interpretation is not valid even though the model remains non-singular. Quantum geometry and dynamics of space-time {#s:QGST} =========================================== So far, in Section \[QGS\], we have seen the quantum geometry of [ *space*]{}, with its characteristic features of discrete structures in loop quantum gravity. To fit these modifications into a covariant quantum [ *space-time*]{} structure, we must find a consistent deformation of the hypersurface-deformation algebra of which the modified Hamiltonian constraint is a part. As has by now become clear from many examples, derived at different levels of effective and operator calculations, the classical constraint algebra is then indeed deformed: quantum-geometry corrections imply modified quantum space-time structures [@ConstraintAlgebra]. The possibility of such deformations has also been suggested based on Wheeler–DeWitt quantization [@BohmEuclidean]. Instead of the classical commutator of two time deformations, we have $$\{H[N_1],H[N_2]\}= D[\beta q^{ab}(N_1\nabla_bN_2-N_2\nabla_bN_1)]$$ with a phase-space function $\beta\not=1$. The typical and rather universal form of these deformations is as follows. Holonomy corrections result in a curvature-dependent $\beta(K)=\cos(2\ell K)$, where $\ell$, related to the quantum-gravity scale $L$, is a holonomy parameter depending on the curves used to integrate the connection, and $K$ is a curvature component such as the Hubble parameter for perturbations around isotropic models [@ScalarHol] or the rate of change of orbit areas in spherical symmetry [@JR; @LTBII]. Such a deformation has also been found by operator calculations in $2+1$-models [@ThreeDeform]. For inverse-triad corrections, $\beta=\alpha^2$ depends on the inverse-triad correction function $\alpha$ as in (\[alpha\]), which in turn depends on the quantum-gravity scale $L$ [@ConstraintAlgebra; @JR; @LTBII]. Also here, operator calculations have provided supporting evidence [@TwoPlusOneDef]. Example: spherical symmetry {#s:SphSymm} --------------------------- Spherically symmetric models with their reduced number of free fields provide an interesting testing ground for different quantum space-time structures, and at the same time allow physical applications for instance to black-hole physics. In Ashtekar–Barbero variables, used to compute deformations with corrections from loop quantum gravity, we express the canonical structure by four fields, two scalars $A_{\varphi}$ and $E^x$ and two densitized scalars $A_x$ and $E^{\varphi}$. They appear as components of spherically symmetric SU(2)-connections $$\label{A} A=A_x(x)\tau_3{\rm d} x+ A_{\varphi}(x)\bar{\Lambda}_A{\rm d} \vartheta+ A_{\varphi}(x)\Lambda_A\sin\vartheta{\rm d}\varphi+ \tau_3\cos\vartheta{\rm d}\varphi$$ and densitized triads $$\label{E} E=E^x(x)\tau_3\sin\vartheta\frac{\partial}{\partial x}+ E^{\varphi}(x)\bar{\Lambda}^E\sin\vartheta\frac{\partial}{\partial\vartheta}+ E^{\varphi}(x)\Lambda^E\frac{\partial}{\partial\varphi}\,.$$ For the general derivation see [@SymmRed; @SphSymm; @Springer]. The su(2)-valued fields $\Lambda^E=\tau_1\cos(\zeta(x))+\tau_2 \sin(\zeta(x))$ and $\bar{\Lambda}^E=\tau_3^{-1}\Lambda\tau_3$ describe a U(1)-gauge freedom with gauge rotations by $\exp(\lambda(x)\tau_3)$, remnant from the initial SU(2)-freedom. Similarly, the densitized triad has independent su(2)-matrices $\Lambda_A$ and $\bar{\Lambda}_A$. Also $A_x$ is affected by these gauge transformations, under which it changes to $A_x+{\rm d}\lambda/{\rm d}x$ like a U(1)-connection, but the combination $A_x+{\rm d}\zeta/{\rm d}x$ is invariant, and happens to agree with an extrinsic-curvature component $K_x$ (up to a factor of $\gamma$). With different matrices $\Lambda_A$ and $\Lambda^E$, the components $E^{\varphi}$ and $A_{\varphi}$, unlike $E^x$ and $A_x$ or $K_x$, are not canonically conjugate. However, if we switch to extrinsic curvature also for $\varphi$-components, we obtain canonical pairs $\{K_x(x_1),E^x(x_2)\}= 2\gamma G \delta(x_1,x_2)$ and $\{K_{\varphi}(x_1),E^{\varphi}(x_2)\}= \gamma G \delta(x_1,x_2)$, as shown in [@SphSymmHam]. In these canonical variables, we have the diffeomorphism constraint $$\label{DSph} D_{\rm grav}[N^x]=\int{\rm d} x N^x(2K_{\varphi}'E^{\varphi}-K_xE^x{}')$$ and the Hamiltonian constraint $$\label{HL} H[N] = -\frac{1}{2G}\int {\rm d} x N |E^x|^{-1/2}\left( (1-\Gamma_{\varphi}^2+K_{\varphi}^2)E^{\varphi}+2 |E^x|K_{\varphi}K_x+ 2|E^x|\Gamma_{\varphi}' \right) \,.$$ with $\Gamma_{\varphi}= -(E^x)'/2E^{\varphi}$ a spin-connection component. (Primes denote derivatives by $x$.) The appearance of inverses of $E^x$ and quadratic expressions in $K_x$ and $K_{\varphi}$, the latter as canonical and gauge-invariant versions of the connection, suggests inverse-triad and holonomy corrections from loop quantum gravity. Hamiltonian constraint operators have been constructed in [@SphSymmHam], and more generally for Gowdy models in [@EinsteinRosenAsh; @EinsteinRosenQuant], in which inverse-triad operators and holonomy operators indeed appear. An effective Hamiltonian would be obtained from expectation values of these operators, but the calculations are complicated. Moreover, so far these Hamiltonians could not be ensured to be anomaly-free at the operator level. Instead, we can parameterize effective Hamiltonians by correction functions originating from inverse-triad and holonomy operators, and compute Poisson brackets of the modified constraints to see under which conditions they can be anomaly-free [@JR]. We write a general modified Hamiltonian constraint as $$\begin{aligned} \label{HQSph} H^Q_{\rm grav}[N]&=&-\frac{1}{2G}\int {{\mathrm{d}}}x\, N\bigg(\alpha|E^x|^{-1/2} E^{\varphi}f_1(K_{\varphi},K_x)+ 2\bar{\alpha} |E^x|^{1/2}f_2(K_{\varphi},K_x) \nonumber\\ && \qquad\qquad+\alpha_{\Gamma}|E^x|^{-1/2}(1-\Gamma_{\varphi}^2)E^{\varphi}+ 2\bar{\alpha}_{\Gamma}\Gamma_{\varphi}'|E^x|^{1/2} \bigg)\,,\end{aligned}$$ with inverse-triad correction functions $\alpha$, $\bar{\alpha}$, $\alpha_{\Gamma}$ and $\bar{\alpha}_{\Gamma}$ initially left independent of one another, and holonomy correction functions $f_1$ and $f_2$. All these functions may in principle depend on all canonical variables, although the triad dependence of inverse-triad correction functions and the curvature dependence of holonomy correction functions should be primary. Moreover, an anomaly-free commutator with the diffeomorphism constraint shows that inverse-triad correction functions can only depend on $E^x$, not on the density-weighted $E^{\varphi}$. Computing Poisson brackets of two modified Hamiltonian constraints with different lapse functions, it turns out that anomaly freedom can be realized if $f_1=F_1^2$ and $f_2=K_x F_2$ provided that $F_2= F_1(\partial F_1/\partial K_{\varphi}) \alpha/\alpha_{\Gamma}$ [@JR]. Choosing a function $F_1$ periodic in $K_{\varphi}$, holonomy modifications for this component are realized. The second correction function $F_2$ is then fixed, showing how anomaly-freedom can put restrictions on possible modifications and quantum corrections. In fact, the corrections seem even stronger for the $K_x$-dependence, left unmodified in the function $f_2$ shown here. An extension to a holonomy-corrected $K_x$-dependence appears more difficult than one of the $K_{\varphi}$-dependence. Moreover, while $K_{\varphi}$ would give rise only to pointwise exponentials $\exp(i\gamma\ell K_{\varphi})$ as holonomies, the curve integration along angular directions, in which $K_{\varphi}$ points, being trivial in spherically symmetric models, $K_x$ would be replaced by a holonomy $\exp(i\gamma\int_I{\rm d}x K_x)$ integrated along some interval $I$. Derivative corrections should therefore result as well, or the constraint would become non-local if integrations are left unexpanded. If we take $F_1(K_{\varphi})=(\gamma\ell)^{-1} \sin(\gamma\ell K_{\varphi})$, suitable for holonomy corrections as in the cosmological example (\[Hmod\]), we have $F_2(K_{\varphi},E^x)=(2\gamma\ell)^{-1}\sin(2\gamma\ell K_{\varphi})\alpha/\alpha_{\Gamma}$. The algebraic deformation is then given by $\beta(E^x,K_{\varphi})=\bar{\alpha}\bar{\alpha}_{\Gamma} \partial F_2/\partial K_{\varphi}$ [@JR]. For the example provided, this means $\beta(E^x,K_{\varphi})= \alpha(E^x)\bar{\alpha}(E^x)(\bar{\alpha}_{\Gamma}(E^x)/\alpha_{\Gamma}(E^x)) \cos(2\gamma\ell K_{\varphi})$, a function that is negative for $\gamma\ell K_{\varphi}\sim \pi/2$, at curvatures where the correction function $f_1$ is near its maximum and a strong modification of the classical linear function. This property is realized generically: Combining the previous equations, we can write $$\beta(E^x,K_{\varphi})= \frac{1}{2} \alpha\bar{\alpha}\frac{\bar{\alpha}_{\Gamma}}{\alpha_{\Gamma}} \frac{\partial^2 f_1}{\partial K_{\varphi}^2}$$ which is negative around maxima of $f_1$, irrespective of its functional form. Consequences of negative $\beta$ will be discussed in more detail soon. Note also that holonomy corrections and inverse-triad corrections are rather independent of each other in their effect on the deformation, affecting $\beta$ multiplicatively. The inverse-triad correction functions must satisfy $$\label{alphaGamma} (\bar{\alpha}\alpha_{\Gamma}-\alpha\bar{\alpha}_{\Gamma}) (E^x)' + 2(\bar{\alpha}'\bar{\alpha}_{\Gamma}- \bar{\alpha}\bar{\alpha}_{\Gamma}') E^x=0$$ for a closed constraint algebra [@JR]. If $F_1$ is independent of $E^x$, or at least depends on this triad variable in a way different from inverse-triad corrections, one can show that both terms in (\[alphaGamma\]) must vanish individually, and we have $\alpha_{\Gamma}=\alpha$ and $\bar{\alpha}_{\Gamma}=\bar{\alpha}$. See also [@Action]. For consistent deformations in the presence of cosmological perturbations, anomaly-freedom is implemented in the same spirit, but with an extra ingredient. Without any symmetry assumptions, requiring irregular lattices and non-Abelian SU(2)-features, it is complicated to derive inverse-triad operators or to parameterize holonomy corrections. One therefore starts using all information about such correction functions that can be obtained in tractable models, such as a homogeneous background, and inserts those background functions just like $\alpha$, $\bar{\alpha}$, $\alpha_{\Gamma}$, $\bar{\alpha}_{\Gamma}$, $f_1$ and $f_2$ in spherically symmetric models. These functions refer only to the background variables, but depending on the order of cosmological perturbations, also the dependence on inhomogeneity is required. The corresponding terms, in many cases, cannot be computed directly from operators; instead, one inserts “counterterms” in the Hamiltonian constraint expanded by inhomogeneity, taking into account all possible terms to the given order that could be generated by correction functions depending on homogeneous fields [@ConstraintAlgebra]. Terms that cannot be computed from operators are left unspecified as free functions. In many cases, counterterms contribute derivative corrections, adding for instance terms containing $\partial_aE^b_i$ for inverse-triad corrections. A dependence of correction functions on integrated variables such as fluxes is therefore realized even if $\alpha$ initially depends only on local triad values. The condition of anomaly-freedom is often so restrictive that the counterterms can be derived uniquely from known inverse-triad or holonomy correction functions of the background. Consistent constraints can therefore be computed even if not all quantum corrections are known in detail. Another small difference between spherical symmetry and cosmological perturbations is the treatment of the SU(2)-gauge. In spherical symmetry, (\[DSph\]) and (\[HQSph\]) are manifestly invariant under these transformations, even after deformation by quantum corrections. In perturbative treatments, on the other hand, one usually fixes a background triad, including its SU(2)-gauge. But then, one can see that physical results and the deformation of the space-time algebra are independent of the specific SU(2)-fixing chosen. The constructions are therefore consistent even if some gauge has been fixed: All gauge fixings produce the same results. (The same arguments can be applied to the time gauge used to descend from space-time tetrads to spatial triads.) Such consistent derivations in the presence of gauge fixing are possible in simple cases such as the Gauss constraint, which moreover survives unmodified after quantization. Making gauge-fixings consistent is much less trivial if attempted for complicated constraints such as $H[N]$. First, it is difficult to find any good gauge fixing in general terms; having to study even all possible gauge fixings and to make sure that physical results do not depend on the choice is then nearly impossible. In such cases, the only manageable approach is to forgo gauge fixing before quantization or deformation, even if it makes derivations more complicated than in one given gauge. The meaning of deformed hypersurface deformations {#s:Deform} ------------------------------------------------- The hypersurface-deformation algebra encodes the space-time structure of generally covariant theories just as the Poincaré algebra encodes special relativity’s structure. One can recover the Poincaré relations by using functions $N$ and $N^a$ linear in some coordinates amounting to Minkowski space-time or a local Minkowski patch. For instance, two linear lapse functions of the form $N=\Delta t+ \vec{v}\cdot \Delta \vec{x}$, inserted in (\[HH\]), provide the commutator of Lorentz boosts by velocity $v$ and time translations $\Delta t$. Inserting two such functions, $N_1=\vec{v}\cdot\Delta \vec{x}$ and $N_2=\Delta t-\vec{v}\cdot\Delta\vec{x}$, on the right-hand side of (\[HH\]) shows a commutator that amounts to the displacement $\Delta \vec{x}=\vec{v}\Delta t$. In terms of linear hypersurface deformations, the relation follows from elementary geometry; see Fig. \[HypDefLinMink\]. ![Two linear deformations of spatial slices, first by $N_1=v\Delta x$ along the normals and then by $N_2=\Delta t-v\Delta x$ (top) and in the opposite ordering (bottom), commute up to a spatial displacement by $\Delta x=v\Delta t$. Normals are drawn according to Minkowski geometry, corresponding to Lorentzian signature. \[HypDefLinMink\]](HypDefLinMink){height="5cm"} If the algebra is deformed, as in (\[HHbeta\]), the same choice of linear deformations along the normals gives rise to a rescaled relation $\Delta x=\beta v\Delta t$. Quantum space-time, with its discrete structure that is responsible for the algebraic deformation via holonomy and inverse-triad corrections, changes the relation between velocity and displacement; discrete space-time speeds up or slows down motion. Such a phenomenon is well-known from condensed-matter physics and should not come as a surprise, although the form in which it is realized here is rather different owing to the more-basic notions of space and time involved. When $\beta$ becomes negative, as happens with holonomy corrections at high density, the relation $\Delta x=\beta v\Delta t$ is rendered counter-intuitive. However, the change of sign can be interpreted easily if one redraws Fig. \[HypDefLinMink\] in Euclidean space, especially regarding the directions of normals to spatial slices. As shown in Fig. \[HypDefLin\] compared with Fig. \[HypDefLinMink\], the displacement then indeed points in the opposite direction (and does not change magnitude). The relation $\Delta x=-v\Delta t$ does not describe motion but rather, despite the notation of variables, a rotation. When $\beta$ turns negative, we have Euclidean signature rather than Lorentzian. Even if $\beta$ is not exactly $-1$ but negative, the structure is best described as Euclidean even though we do not have classical Euclidean space (just as we do not have classical Minkowski space if $\beta$ is positive but not exactly $+1$). ![With normals drawn according to Euclidean geometry, the displacement seen in Fig. \[HypDefLinMink\] points in the opposite direction: The sign in the commutator of two normal deformations is reversed. \[HypDefLin\]](HypDefLin1){height="5cm"} Signature change is not only a drastic reminder that we cannot take much of our usual concepts for granted when space-time is quantized. It also shows the limitations of approaches in which gauge-fixing or deparameterization is used before quantization or modifications of the constraint. If one fixes the gauge before quantization, one can only assume the classical space-time structure, which does not allow signature change. After gauge fixing and quantization, the full space-time structure can no longer be accessed, leaving one with the conclusion that space-time is unmodified. However, since the constraints are modified in quantization after gauge fixing and determine the gauge and space-time structure, the procedure becomes intrinsically inconsistent. Deparameterization cannot capture all quantum space-time effects either. When one distinguishes a phase-space degree of freedom to measure change and evolution, there is no guarantee that this degree of freedom actually behaves like time in a space-time sense. Also Euclidean theories can formally be deparameterized (internal time simply parameterizes gauge orbits of the Hamiltonian constraint), showing that deparameterized “evolution” does not necessarily imply evolution in a temporal sense. Again, only a complete analysis of the off-shell constraint algebra, without eliminating some of its more complicated ingredients by gauge-fixing or deparameterization, can show what space-time structure is realized. In our diagrams so far, we have assumed that normals are drawn using either Minkowski or Euclidean geometry. We did not use quantum corrections of angles even though distances and displacements did receive corrections as a consequence of the algebra (\[HHbeta\]). One may expect angles to change too, in particular angles of normals to spatial slices drawn to visualize the commutator of two time deformations. Such corrections could indeed happen, in general deformations of space-time structures, but since the angle between space and time directions would be involved, they would amount to deformations of the commutator (\[HD\]) of a time and a space deformation. A temporal and a spatial deformation commute up to a temporal deformation, as illustrated in Fig. \[HDFig\]. If there are quantum corrections to this relation, one might interpret them as modified spatial displacements, rescaled compared to the classical relations, or a modification in space-time angles used to define temporal deformations along the normals. In the former case, also the purely spatial commutator (\[DD\]) should be modified, which is not the case. A modified (\[HD\]) therefore indicates quantum corrections to space-time angles, not just to distances as indicated by a modified (\[HH\]). For instance, modifications which add a term of $D[\beta' NN^a]$ to the classical result of $\{H[N],D[N^a]\}$, as possible for vector modes [@VectorHol] but not with scalar modes [@ScalarHol], would have an additional spatial shift of the open circles in (\[HDFig\]). (Note that Fig. \[HDFig\] also illustrates the fact that (\[HD\]) is not subject to a change in sign in Euclidean signature. If the normals are drawn according to Euclidean geometry, just the rescaling of $\Delta x$ under boosts changes, but not the direction of the temporal displacement.) ![The commutator of linear time and space deformations: A Lorentz boost produces the solid inclined line, the same boost performed after a spatial displacement the dashed line. A spatial point (full circle) is mapped to two different positions (open circles) depending on whether the spatial displacement is performed before or after the boost. The final positions differ by a time translation (dotted), according to $\{H[v x],D[\Delta x]\}= H[v\Delta x]$ using (\[HD\]). \[HDFig\]](HD){width="11cm"} Based on algebraic calculations in effective loop quantum gravity, modifications of (\[HD\]) that seemed possible for vector modes subject to holonomy corrections [@VectorHol] turned out not to be consistent with scalar modes [@ScalarHol]. It therefore appears that space-time angles are not affected by the corrections of loop quantum gravity, but as discussed in Section \[s:Diffeo\], stronger corrections to the diffeomorphism constraint than used so far may be realized. For now, all consistent deformations point to quantum corrections only in the commutator of two time deformations. Consistency and space-time structures {#s:SpaceTimeCons} ------------------------------------- With a deformed hypersurface-deformation algebra the theory is fully consistent. The full amount of gauge transformations exists to remove spurious degrees of freedom, and constraints are guaranteed to be preserved by evolution. Cosmological observables, for instance, can be computed by standard Hamiltonian means, as developed for deformed algebras in [@ConstraintAlgebra; @ScalarGaugeInv]. (The classical Hamiltonian formalism for cosmological perturbations, of [@HamGaugePert], assumed certain features of observables that are modified by deformations.) However, other familiar notions used often in general relativity no longer apply. Even the line element, one of the most basic mathematical objects of differential geometry, must be treated with care — or altogether avoided. Constraints obeying a modified algebra generate gauge transformations for metric components $q_{ab}$ that cannot agree with Lie derivatives or space-time coordinate transformations — otherwise, the form of gauge transformations would imply the classical hypersurface-deformation algebra. But if $q_{ab}$, or the space-time metric $g_{\mu\nu}$ completed in the usual way using lapse and shift in $${\rm d}s^2= g_{\mu\nu}{\rm d}x^{\mu}{\rm d}x^{\nu}= -N^2{\rm d}t^2+ q_{ab}({\rm d}x^a+N^a{\rm d}t)({\rm d}x^b+N^b{\rm d}t)\,,$$ does not transform by classical coordinate changes, the contraction $g_{\mu\nu}{\rm d}x^{\mu}{\rm d}x^{\nu}$ with standard coordinate differentials ${\rm d}x^{\mu}$ is not invariant and cannot be used as a line element. One would have to modify transformations of coordinate differentials as well to make the contraction invariant, in which way one could possibly make contact with non-commutative [@Connes] or fractional geometry [@Fractional]. (Modified hypersurface-deformation algebras have also been found with higher-derivative dispersion relations for matter [@TensorialDeform], but the form does not appear related to what is suggested by loop quantum gravity.) Lacking a line element and related notions such as geodesics or trapped surfaces, black-hole properties and even their definitions must be rethought. One can only rely on canonical formulations of horizon conditions that capture the well-known notions of standard space-time. Classically, there are several different definitions of horizons which, at least in simple cases, all provide the same results. However, when they are reformulated canonically and adapted to quantum space-time, results may differ. As analyzed in [@ModifiedHorizon], the consistency of quantum space-time structures also in this case helps to arrive at unambiguous answers. For calculations in quantum space-time, one must rely on Hamiltonian methods, using directly the modified constraints. First applied systematically in [@ScalarGaugeInv], one obtains unambiguous expressions for gauge-invariant variables in cosmology as well as their evolution equations. Without using an action, it may not always be obvious how to combine all equations obtained to just one Mukhanov-type equation for the analog of the curvature perturbation, but examples have been found for inverse-triad corrections [@LoopMuk] and holonomy corrections [@ScalarHolMuk]. Quantum corrected wave equations are then obtained, which directly show the modified speeds of modes as expected from a deformed hypersurface-deformation algebra: electromagnetic and gravitational waves obey the equation $$\label{Wave} -\frac{\partial^2w}{\partial t^2}+ \beta \Delta w+ f(a,\dot{a})w=0$$ with the correction function $\beta$ from (\[HHbeta\]) and a function $f$ that shows how the evolution of modes depends on an expanding background. Density perturbations obey a similar equation, but with a differently modified speed for inverse-triad corrections [@LoopMuk]. Interesting phenomenological effects are therefore suggested. (For holonomy corrections in currently existing versions, the modified speeds of density perturbations and gravitational waves are identical [@ScalarTensorHol].) Holonomy modifications at high density are especially drastic: they imply signature change [@Action; @SigChange]. We are no longer dealing with quantum space-time but with a quantum version of 4-dimensional Euclidean space, shown by the sign change of $\beta(K)$ for large $\ell K$. The deformed hypersurface-deformation algebra (\[HHbeta\]) then belongs to Euclidean-type space, and (\[Wave\]) shows the elliptic nature of linear mode equations. For positive as well as negative $\beta$, we do not have standard space or space-time unless the value is exactly $\pm1$; quantum space-time effects always occur. But the distinction between positive and negative values of $\beta$, as opposed to two different positive values, is much more important because it changes the type of initial-value or boundary problems in mode equations. Signature change defined by the sign of $\beta$ is therefore physically relevant, even in the absence of a standard classical space-time structure. In these high-density regimes, however, it is no longer consistent to treat holonomy corrections in isolation because they mix with quantum back-reaction, together forming higher-curvature corrections. So far, no consistent deformation of the constraint algebra has been found for quantum corrections caused by quantum back-reaction, adding moment-dependent terms to the classical constraints. The problem is well-defined because we know the Poisson algebra of moments, and even though canonical effective field theory techniques remain incomplete, one could use those of quantum mechanical models in loop quantum gravity restricted to fixed graphs. In any finite region, there would then be a finite, though large, number of degrees of freedom given by link holonomies and plaquette fluxes. (Most calculations in the full theory are done with fixed graphs, anyway.) Since Poisson brackets of moments always produce other moments, quantum corrections must be arranged so that surplus terms delicately cancel in the constraint algebra. No such version has been found yet, which is not altogether surprising because a successful implementation would imply a consistent version of quantum space-time, including moments of a state preserving covariance. Anomaly problem --------------- The problem of finding consistent deformations of the hypersurface-deformation algebra is related to the long-standing anomaly issue of canonical quantum gravity. If one can find an anomaly-free representation of quantum constraint operators, turning the classical constraint algebra into some operator version, effective constraints will automatically be first class and consistent: If two constraint operators $\hat{C}_1$ and $\hat{C}_2$ commute up to another constraint operator, the quantum constraints $\langle\hat{C}_1\rangle$ and $\langle\hat{C}_2\rangle$ Poisson commute up to another quantum constraint, thanks to $\{\langle\hat{C}_1\rangle,\langle\hat{C}_2\rangle\}= \langle[\hat{C}_1,\hat{C}_2]\rangle/i\hbar$. This statement also extends to higher-order quantum constraints $\langle f(\hat{q},\hat{p})\hat{C}\rangle$ [@EffCons]. Structure functions in the constraint algebra cause several problems in trying to find anomaly-free operator versions. But if such a version has been found, there are no additional problems in the transition to effective constraints. The product $[\hat{C}_1,\hat{C}_2]=\hat{f}\hat{C}$ with a quantized structure function $\hat{f}$ will, after taking an expectation value, be one of the higher-order quantum constraints. However, there are obstructions to simple attempts at solving the anomaly problem for constraint operators, related for instance to the interrelation of ordering and self-adjointness issues especially in the presence of structure functions [@NonHerm]. If Hamiltonian constraint operators are self-adjoint, the commutator $[\hat{H}[M],\hat{H}[N]]/i\hbar$ quantizing $\{H[M],H[N]\}$ in (\[HH\]) is self-adjoint too. But classically the bracket equals a product of the local diffeomorphism constraint $D_a$ with the structure function $q^{ab}$, two expressions that have a non-vanishing Poisson bracket because $q^{ab}$ is not diffeomorphism invariant. Any quantization of the product $q^{ab}D_a$ can then be self-adjoint and have a chance of agreeing with $[\hat{H}[M],\hat{H}[N]]/i\hbar$ only if a symmetric ordering is used. But if one simply reorders “$\hat{q}^{ab}\hat{D}_a$” symmetrically, a metric factor would come to lie to the right of the diffeomorphism constraint, and the product would no longer annihilate physical states. An anomalous version of the constraints would be obtained. In an effective description, the situation is more manageable. First, one can use calculations of Poisson brackets instead of commutators of operators, even in the presence of quantum corrections and ordering choices. Moreover, quantum corrections may be suitably parameterized, to take into account different ordering choices, quantization ambiguities in the representation of inverse-triad and holonomy operators, and general classes of states in terms of their moments. One can compute Poisson brackets with ambiguity functions, such as (\[alpha\]), or moment terms unspecified, and see what conditions anomaly-freedom imposes on them. An operator calculation with free functions or states, by comparison, would be much more involved. The self-adjointness question can be left open at first by allowing for complex-valued constraints — kinematical moments appearing in effective constraints are complex-valued anyway, even in the absence of structure-function issues, as seen in Section \[s:EffCons\]. With this strategy the first consistent deformations of the form (\[HHbeta\]) have been found [@ConstraintAlgebra], by now confirmed also by operator calculations [@TwoPlusOneDef]. The effective view is therefore reliable and powerful also in the context of the anomaly problem. Quantum back-reaction and higher time derivatives ------------------------------------------------- Moments and their quantum back-reaction, if they appear in a consistent deformation of the constraint algebra, are the canonical analog of higher-curvature terms with their higher time derivatives. However, their form is not directly one of higher time derivatives, and additional steps and expansions, primarily an adiabatic one, are required to put canonical effective equations in the form of higher-derivative ones; see Section \[s:Adia\]. The adiabatic approximation is not always applicable. It serves well for anharmonic oscillators expanded around the harmonic vacuum with its constant moments. Moments of the anharmonic vacuum change, but only slowly, and can be treated adiabatically. Solving for moments order by order in the adiabatic expansion then shows how they are related to higher time derivatives of expectation values. The same statements hold true in quantum field theory, where one expands around the free vacuum in order to describe excitations around the interacting vacuum using the low-energy effective action. These features have given rise to the expectation that quantum gravity should be subject only to higher-curvature corrections. However, two hidden assumptions are required for this conclusion. First, one assumes that quantum gravity implies corrections only in the dynamics of gravitons, say, not in the underlying space-time structure. Since gravity theories are fundamentally about space-time structure, this assumption, valid in a perturbative context of excitations on a fixed background, need not be true in general. The second assumption is that quantum gravity can be realized perturbatively around some free theory of the usual form. Also this statement may be true for perturbations around a fixed background and perhaps some other situations, but it is not always valid. Quantum gravity is not expected to have a non-perturbative ground state, and other distinguished states may be very different from Gaussians as they appear in the harmonic-oscillator ground state or free vacuum states. Such states may not allow adiabaticity of the moments, and therefore cannot have a dynamics fully expressed by higher time derivatives. Indeed, examples in quantum cosmology are known in which the adiabaticity assumption is difficult to realize consistently [@BouncePot]. Even for anharmonic oscillators, the adiabatic approximation is not valid when one perturbs around correlated coherent states, such as fully squeezed Gaussians, of the harmonic oscillator instead of the ground state [@EffAc; @HigherTime]. Instead of a higher-curvature or other higher-derivative effective theory, canonical quantum gravity has higher-dimensional effective systems in which the moments play the role of independent degrees of freedom in addition to expectation values. They are subject to their own equations of motion, and by quantum back-reaction couple to them, as e.g. in (\[dqdt\])–(\[dppdt\]). Their effect can be formulated by higher-derivative terms only in certain regimes, but not in general. Effective actions ----------------- So far, we have stayed in the canonical framework and dealt with effective equations and constraints. Complementary information can be obtained if a corresponding effective action is found. In principle, there is a one-to-one correspondence between canonical and Lagrangian formulations, but in perturbative settings, especially those that imply higher derivatives, the transformation is far from obvious. A derivative expansion of a Lagrangian may then correspond to a complicated resummation of a derivative-expanded Hamiltonian. It is therefore of interest to look for and study effective actions even if effective canonical equations are already known. There are several examples in which effective actions have proven their usefulness in quantum cosmology. Euclidean path-integral techniques have been developed and applied to questions such as tunneling probabilities and some semiclassical issues [@EffAcOneLoop; @EffAcTunneling; @EffAcInflation; @FermionDecoherence]. As expected, in semiclassical regimes, corrections to Einstein’s equation are of higher-curvature form. Also causal dynamical triangulations have led to effective actions of cosmological systems by comparing detailed numerical studies of volume fluctuations with the dynamics of minisuperspace models [@CDTEffAc]. By matching volume correlation functions with results expected from a higher-curvature effective action, quantum corrections can be derived. A general comparison involving also possible modifications to quantum space-time structure on top of higher-curvature corrections has not yet been completed. Effective actions of such forms often show more directly than other types of effective equations when quantum effects become significant. Effective actions based on path integrals are subject to modified quantum space-time structures just as canonical effective equations and constraints. Path integration for gravitational theories requires an integration over all metrics, with a suitable measure that preserves covariance and does not introduce anomalies. Such a measure has not been found in complete generality, indicating that one encounters in this approach the same difficulties that appear when one tries to represent the canonical constraint algebra consistently. The path-integral measure is indeed the place to look at when one is interested in possible space-time modifications, but its incomplete nature does not give many clues. Canonical theories, with consistent deformations of the constraint algebra found in recent years, have been able to make more progress in this direction. (The spin-foam approach [@Rov:Loops; @SpinFoamRev; @PerezLivRev] attempts to take the results of loop quantum gravity regarding background-independent representations to a level comparable to path-integrals [@SurfaceSum]. However, while some aspects of integrations over the space of metrics can be clarified, the issue of the correct measure, or spin-foam face amplitudes, remains open also here [@Anomaly]. Notwithstanding these problems, traditional effective-action techniques have been applied to spin foams in [@EffAcSpinFoam].) Another question in path-integral based approaches is how the state dependence enters effective actions. The low-energy effective action can be computed quite conveniently with path integrals, but it hides the fact that one is expanding around the ground state. If one tries to go beyond this limitation, which in quantum cosmology with its lack of ground states is a severe one, the required calculations become much more involved. Suitable states would have to be implemented by additional wave-function factors in the path integral, and integrations would no longer be Gaussian for general states. Moreover, since path integrals in their most common form provide transition amplitudes between two states separated by some finite time, one would have to put in the initial and final state. Compared to effective canonical equations, which only require initial moments to be specified, more a-priori information about the physical system is therefore required. In quantum cosmology, this information is not easily found by independent means, but is important for an analysis especially of the Planckian regime. Regained dynamics: from canonical to Lagrangian ----------------------------------------------- In this situation, a combination of canonical and Lagrangian effective methods is of interest. As already mentioned, it may be difficult to perform a transformation for constrained systems, especially those with higher-derivative terms. If the relation between the momenta and time derivatives of configuration variables involves higher derivatives, as easily happens with quantum-gravity corrections such as holonomy modifications, the Lagrangian would appear as a complicated resummation of the higher-derivative expansion of the Hamiltonian. Moreover, even if one can Legendre transform the Hamiltonian constraint to arrive at a Lagrangian, one would, for a comparison with path integrals, still have to look for a measure under which the modified Lagrangian is covariant. Also integrating the Lagrangian to an action requires new constructions even though just a classical measure on space-time is needed. However, with space-time structures modified and effective metrics non-existent — see Section \[s:SpaceTimeCons\] — defining space-time integrations is non-trivial. Measure issues, therefore, cannot be resolved easily, but at least an effective Lagrangian for further classical-type analysis may still be found. Instead of attempting a Legendre transformation, the canonical methods of [@Regained; @LagrangianRegained] can be used to derive a Lagrangian (or the constraints themselves) directly from a modified constraint algebra, to any given order in derivatives. If only second-order derivatives are assumed and the constraint algebra is classical, the Einstein–Hilbert action as a two-parameter family with Newton’s and the cosmological constant is obtained as the unique solution. If higher derivative orders are allowed, one expects higher-curvature corrections as well. And if even the constraint algebra is quantum corrected, as suggested by loop quantum gravity, stronger corrections, also to second derivative order, are obtained. With this procedure of “regaining” a Lagrangian from the constraint algebra one can sidestep the complicated resummations that a Legendre transformation from modified constraints to the higher-derivative Lagrangian would imply. Not all coefficients in the Lagrangian may follow uniquely, especially if higher derivatives are included leaving several options of higher-curvature invariants of the same order. But the general form of modifications and implications for quantum space-time structure can still be found. ### Functional equations To spell out the general procedure, let us assume that we have a Hamiltonian or Hamiltonian constraint $H(q,p)$ depending on canonical fields $q(x)$ and $p(x)$ in space. We introduce $\delta H/\delta p(x)=:v(x)$ as a new independent variable in place of $p$, and then expand equations by this newly defined $v$. If $H$ is a Hamiltonian generating evolution in some fixed time parameter, we have $v(x)=\dot{q}(x)$ by Hamilton’s equations. If $H$ is a Hamiltonian constraint in the absence of an absolute time, $v= (\delta N)^{-1}\{q,H[\delta N]\}$ is the derivative of $q$ in a direction normal to spatial slices. For gravitational variables with $q$ the metric or triad, $v$ would therefore be related to extrinsic curvature. This change of variables amounts to what is needed for a Legendre transformation from $(q,p)$ with Hamiltonian $H$ to $(q,v)$ with Lagrangian $L=pv-H$, whose form will result as a solution of the $v$-expanded equations. Note that we cannot always assume the Hamiltonian to be local and free of derivatives of $p$, which would imply that partial derivatives could be used to compute $v$. Holonomy corrections in loop quantum gravity, for instance, introduce higher spatial derivatives of the momentum conjugate to the densitized triad. In such situations, there is no local relation between the Hamiltonian and the Lagrangian, and explicitly performing a Legendre transformation is complicated. Instead of computing the transformation, we intend to calculate the Lagrangian directly from the constraint algebra, assuming from now on the relations of the (deformed) hypersurface-deformation algebra. Using the definition of $v$ and $$\left.\frac{\delta H}{\delta q(x')}\right|_{p(x)} = -\left.\frac{\delta L}{\delta q(x')}\right|_{v(x)}$$ for the unsmeared Hamiltonian constraint and the Lagrangian density, we write the Poisson bracket (\[HHbeta\]) of two Hamiltonian constraints as $$\begin{aligned} \{H[N],H[M]\} &=& -\int{\rm d}^3x\int{\rm d}^3y \frac{\delta L(y)}{\delta q(x)} v(x) N(y)M(x)- (N\leftrightarrow M)\\ &=& \int{\rm d}^3x \beta D^a(x) (N\nabla_aM-M\nabla_aN)\end{aligned}$$ with the local diffeomorphism constraint $D^a$. Taking functional derivatives by $N$ and $M$, we arrive at the functional equation $$\label{FunctionalL} \frac{\delta L(x)}{\delta q(x')}v(x') + \beta(x)D^a(x)\nabla_a\delta(x,x')-(x\leftrightarrow x')= 0$$ for $L(x)$, which can be solved once an expression for the diffeomorphism constraint $D^a$ is inserted, depending on whether $q$ refers to gravity or some matter field. In all cases, $D^a$ is linear in the momenta. A linear equation for $L$ is thus obtained [@LagrangianRegained]. If (\[DD\]) and (\[HD\]) are unmodified, standard expressions for $D^a$ can be used. (See Section \[s:Diffeo\] for a discussion of possible further modifications.) ### Matter To illustrate the regaining procedure, we look at a scalar matter field $\phi$ without derivative couplings, whose Hamiltonian obeys the (deformed) hypersurface-deformation algebra on its own, without adding the gravitational piece. The Lagrangian density must be of the form ${\cal L}= \sqrt{\det g} L(\phi,v,\psi)$ where $v=(\delta N)^{-1}\{\phi,H[\delta N]\}$, as before, is the normal scalar velocity and $\psi= q^{ab} \nabla_a\phi\nabla_b\phi$ is the only remaining scalar that can be formed from $\phi$ and its derivatives, to a total derivative order of at most two. Higher derivatives may easily result in interacting matter or quantum-gravity theories, with higher time derivatives from quantum back-reaction and higher spatial ones, additionally, from possible discretizations. (See e.g.[@ConsistDisc; @UniformDisc; @PerfectAction; @CanSimp] for information about discretized space-time theories.) For now, however, we look for modifications implied by corrections that leave the classical derivative order unchanged. With the canonical variables of a scalar field and its diffeomorphism constraint $D^a=p_{\varphi}\nabla^a\phi$, Eq. (\[FunctionalL\]) assumes the form $$\frac{\delta L(x)}{\delta\phi(x')} v(x')+ \beta \frac{\partial L(x)}{\partial v(x)} (\nabla^a\phi(x)) \nabla_a\delta(x,x') - (x\leftrightarrow x')=0\,.$$ As in [@LagrangianRegained], we write $$\frac{\delta L(x)}{\delta \phi(x')}= \frac{\partial L(x)}{\partial \phi(x)} \frac{\delta\phi(x)}{\delta\phi(x')}+ 2\frac{\partial L(x)}{\partial\psi(x)} (\nabla^a\phi(x)) \nabla_a\delta(x,x')\,.$$ It follows that $$A^a:= (\nabla^a\phi) \left(\beta\frac{\partial L}{\partial v}+ 2v\frac{\partial L}{\partial\psi}\right)$$ satisfies the equation $A^a(x)\nabla_a\delta(x,x')-(x\leftrightarrow x')=0$, shown in [@LagrangianRegained] to imply $A^a=0$. Thus, $$\beta\frac{\partial L}{\partial v}+ 2v\frac{\partial L}{\partial\psi}=0$$ and $L$ must be of the form $L(\phi,\psi-v^2/\beta)$. With non-trivial deformation, $\beta\not=1$, the scalar field therefore obeys a modified dispersion relation. The kinetic term of the Lagrangian does not depend on $\psi-v^2= g^{\mu\nu}(\nabla_{\mu}\phi)(\nabla_{\nu}\phi)$ in space-time terms, but has its time derivatives in $\psi-v^2/\beta$ rescaled by the correction function $\beta$. The resulting modified dispersion relation is in agreement with the wave equation (\[Wave\]). At the canonical level, for comparison, we begin with a matter Hamiltonian density of the form $$H= \nu \frac{p_{\phi}^2}{2\sqrt{\det q}}+ \frac{1}{2}\sigma \sqrt{\det q}\psi+ \sqrt{\det q} \: W(\phi)$$ with general inverse-triad correction functions $\nu$ and $\sigma$, and some potential $W(\phi)$. The corresponding Lagrangian density, with $v=\nu p_{\phi}/\sqrt{\det q}$, takes the form $$L= \sqrt{\det q}\left(\frac{v^2}{2\nu}-\frac{\sigma\psi}{2}- W(\phi)\right)= -\sqrt{\det q}\:\frac{\sigma}{2}\left(\psi- \frac{v^2}{\beta}\right)- \sqrt{\det q} W(\phi)\,.$$ This function has the same kinetic dependence as derived above, provided that $\beta=\nu\sigma$, exactly the requirement for an anomaly-free constraint algebra in the presence of inverse-triad corrections, where $\beta=\alpha^2$ [@ConstraintAlgebra]. Purely canonically, this consistency condition can be seen to ensure causality in the sense that gravitational waves on quantum space-time travel at the speed of light, modified by a factor of $\sqrt{|\beta|}$ compared to the classical speed [@tensor]. No super-luminal propagation happens when anomaly-freedom is taken into account, even if propagation speeds may be larger than the classical speed of light for $\alpha>1$. ### Effective action for inverse-triad corrections {#s:EffInv} For gravity, one example has been worked out in quite some detail with these methods: inverse-triad corrections of loop quantum gravity. These corrections have a characteristic component independent of higher derivatives, and therefore can be analyzed already at the level of second-order equations. The resulting second-order effective Lagrangian, regained from a modified constraint algebra with a correction function $\beta$ independent of curvature components, is $$\label{EffAc} L_{\beta}=\frac{1}{16\pi G} \sqrt{\det q} \left(\frac{{\rm sgn} \beta}{\sqrt{|\beta|}} \frac{v_{ab}v^{ab}- v^a_a v^b_b}{4} + \sqrt{|\beta|} \,{}^{(3)}\!R-2\lambda\right)$$ with “velocities” $v_{ab}$, defined again as normal derivatives [@Action]. For classical gravity, $v_{ab}=2K_{ab}$ would be proportional to extrinsic curvature. Compared with the classical action obtained for $\beta=1$, the notion of covariance has changed: Space and time derivatives are corrected by different coefficients $\sqrt{|\beta|}$ of $^{(3)}R$ and $|\beta|^{-1/2}{\rm sgn}(\beta)$ of $\frac{1}{4}(v_{ab}v^{ab}- v^a_av^b_b)$, respectively. This result is consistent with the fact that the underlying constraint algebra is modified, taking the form (\[HHbeta\]). For this reason, as already mentioned in the context of line elements, the manifold structure required to integrate $L_{\beta}$ to an effective action is not clear. Nevertheless, the effective Lagrangian shows several characteristic effects. First, inverse-triad corrections, having implications even without higher-derivative or higher-order terms in $v_{ab}$, can easily be separated from holonomy effects and higher-curvature corrections. They are especially significant for small fluxes, where $\beta$ becomes small. With these constructions, it is possible to use inverse-triad corrections even when they imply strong modifications, regimes in which one would otherwise have to include full holonomy and higher-curvature corrections as well. These latter types of corrections are indeed present, but affect only higher orders in the $v$-expansion. Inverse-triad effects up to quadratic order in $v$ are reliable even if the other corrections are not known precisely. Time derivatives are then dominant compared to the spatial Ricci scalar when $\beta$ approaches zero at small fluxes, indicating that near-singular geometries are controlled by homogeneous dynamics, strengthening the classical BKL scenario [@BKL] by a no-singularity scenario in loop quantum gravity [@NoSing]. While holonomy effects cannot easily be seen at this level since they would manifest themselves only at higher orders in $v_{ab}$ and mix with higher-derivative terms, they have one drastic implication at high density if they are dominant. Then, the holonomy correction function $\beta$ becomes negative, a feature taken into account by the sign factors in (\[EffAc\]). When the sign changes, the signature does too: at the Lagrangian level, one can interpret the effect by turning $t$ into $it$ in (\[EffAc\]) with $v_{ab}$ interpreted as first-order time derivatives. (The constraint algebra again provides a rigorous interpretation of this signature change; see Section \[s:Deform\].) At Planckian densities, holonomy effects are so strong that they turn space-time into a quantum version of 4-dimensional Euclidean space, lacking time and evolution. Temporal interpretations of high-density holonomy implications, such as bounces, are incorrect. Only large higher-derivative terms could prevent $\beta$ from turning negative, but then other holonomy effects such as bounces would go away too. Implications for phenomenology and potential tests ================================================== It is difficult to test quantum cosmology by any observational means, and given the substantial lack of control over deep quantum regimes, devising high-density scenarios of the universe remains a highly speculative exercise. Higher-curvature corrections, as they always appear in quantum gravity and cosmology except in the most simple harmonic models, are not relevant at currently accessible scales. They are certainly important in the Planckian regime, but then the present theories are so uncontrolled that it is impossible to derive clear effects, and even if some could be suggested, they would most likely be washed away by the immense amount of subsequent cosmic expansion. If one goes beyond higher-curvature corrections as computed for Wheeler–DeWitt quantum cosmology in [@WdWCMB], using for instance quantum-geometry effects from loop quantum gravity, the situation has a chance of being more optimistic. Several investigations have been performed [@SuperInflTensor; @CCPowerLoop; @TensorBackground; @BounceTensor; @SuperInflPowerSpec; @TensorHalfII; @TensorSlowRollInv; @TensorSlowRoll; @GravWaveLQC], but not all are based on consistent implementations of inhomogeneity. The general picture is therefore still incomplete. One type of corrections, holonomy modifications, provides contributions of higher powers of the connection or extrinsic curvature, very similar to some parts of higher-curvature corrections. Holonomy corrections therefore cannot be separated from higher-curvature effects, and cannot provide more-sizeable consequences regarding observations. Moreover, the quantum space-time structure corresponding to holonomy corrections remains incompletely understood. Consistent deformations of the classical constraint algebra with some holonomy-like effects are known in spherically symmetric models [@JR; @LTBII], $2+1$-dimensional gravity [@ThreeDeform] and for cosmological perturbations [@ScalarHol]. However, in spherically symmetric and cosmological models, only “pointwise” holonomy modifications have been implemented, replacing connection components $c$ by $\exp(i\ell c)$ but not integrating over curves. Curve integrations, on the other hand, provide additional terms which are non-local or, when a derivative expansion (\[DerExp\]) is used, introduce higher spatial derivatives. Keeping only higher powers of $c$ as in an expansion of the exponential, but ignoring spatial derivatives is not a consistent approximation: In strong curvature regimes, where holonomy effects should be significant, higher powers and higher derivatives of connection components both contribute to the same order of curvature. Inverse-triad corrections, fortunately, are much better-behaved. First, the derivation of consistent deformations based on them is more complete, achieved in the same kind of models — cosmological perturbations [@ConstraintAlgebra], spherical symmetry [@JR; @LTBII], and $2+1$-dimensional models [@TwoPlusOneDef] in which operator calculations can be performed — but with all crucial terms included. Also inverse-triad corrections should come along with higher-derivative terms because they depend on fluxes, or integrated densitized triads. However, the non-derivative contribution is significant, as seen in Section \[s:EffInv\], and, unlike derivative terms, depends on parameters unrelated to curvature components. Moreover, additional derivative terms are included by the counterterms of [@ConstraintAlgebra], while no connection-derivative counterterms have been used for holonomy corrections in [@ScalarHol]. Inverse-triad corrections are therefore more reliable than holonomy corrections at the present stage of developments in loop quantum cosmology. Inverse-triad corrections are also more interesting from an observational perspective. Because they do not directly refer to the curvature scale but rather to the discrete quantum-gravity scale related to the Planck length, there is no a-priori reason why they should be small at low curvature. They can play a role in standard cosmological scenarios, for instance during inflation. Indeed, in such a combined scenario, the window allowed for inverse-triad effects is much smaller than the one for curvature or holonomy modifications. For inverse-triad corrections, a parameter range of about four orders of magnitude is consistent with observations [@InflTest; @InflConsist], while curvature corrections have an allowed range of about ten orders of magnitude, one compared to the ratio of densities in observationally accessible regimes to the Planck density. There are also indications of interesting and characteristic effects in non-Gaussianity [@NonGaussInvVol], although the required equations of motion second order in inhomogeneity still have to be made consistent. By inverse-triad effects, loop quantum gravity becomes falsifiable. For a more detailed review of phenomenological implications, see [@GCPheno]. Outlook ======= Any quantum system can be evaluated consistently and reliably only when all possible quantum effects and the relevant degrees of freedom are taken into account. For quantum cosmology, this means that one must go beyond the traditional minisuperspace models and find consistent extensions to inhomogeneity. Quantum-representation issues then become much more involved, but can be handled for instance with methods of loop quantum gravity. In this canonical setting, a large set of effective techniques, described in the main part of this review, is now available. These methods allow one to forgo ad-hoc assumptions, to implement full (but possibly deformed) space-time covariance, and to derive a complete phenomenological setting in which all relevant quantum effects are included. Unlike in traditional canonical quantizations and derivations of wave functions, there do not appear to be major obstacles on the way toward systematic comparisons with observations. Quantum cosmology is therefore empirically testable. Concretely working out all terms and studying the necessary parameterizations of quantization and state ambiguities still remains to be completed. Even in isotropic models beyond the harmonic one, not much is known about the evolution of generic quantum states and the robustness of singularity avoidance (see Sections \[s:States\] and \[s:QBR\]). Control on inhomogeneous modes, necessary for most physical questions in cosmology and an understanding of quantum space-time, remains poor in strong quantum regimes, suffering from quantization ambiguities and the difficult anomaly problem (see Sections \[s:Cova\] and \[s:QGST\]). Effective techniques, especially effective constraints, have relieved some of the pressure caused by the failure of traditional methods to address these problems, but they have not been evaluated in sufficient detail to provide a reliable view on Planckian stages in cosmology. Further in-depth investigations are required to change this situation and to provide a complete and reliable phenomenology of quantum cosmology. Acknowledgements {#acknowledgements .unnumbered} ================ This work was supported in part by NSF grant PHY0748336. References {#references .unnumbered} ========== [100]{} J. F. Donoghue, General relativity as an effective field theory: The leading quantum corrections, 50 (1994) 3874–3888, \[gr-qc/9405057\] C. P. Burgess, Quantum Gravity in Everyday Life: General Relativity as an Effective Field Theory, 7 (2004), \[gr-qc/0311082\], http://www.livingreviews.org/lrr-2004-5 M. Bojowald and C. Kiefer, Quantum Cosmology: Fundamentals, \[to appear\] M. Bojowald, , Springer, New York, 2011 M. Bojowald, , Cambridge University Press, Cambridge, 2010 S. Shankaranarayanan and M. Lubo, Gauge-invariant perturbation theory for trans-Planckian inflation, 72 (2005) 123513, \[hep-th/0507086\] J. M. Bardeen, Gauge-invariant cosmological perturbations, 22 (1980) 1882–1905 V. F. Mukhanov, H. A. Feldman, and R. H. Brandenberger, Theory of cosmological perturbations, 215 (1992) 203–333 J. M. Pons, D. C. Salisbury, and L. C. Shepley, Gauge transformations in the Lagrangian and Hamiltonian formalisms of generally covariant theories, 55 (1997) 658–668, \[gr-qc/9612037\] P. A. M. Dirac, The theory of gravitation in Hamiltonian form, 246 (1958) 333–343 M. Bojowald, G. Hossain, M. Kagan, and S. Shankaranarayanan, Anomaly freedom in perturbative loop quantum gravity, 78 (2008) 063547, \[arXiv:0806.3929\] J. Kowalski-Glikman, Introduction to Doubly Special Relativity, 669 (2005) 131–159, \[hep-th/0405273\] J. Magueijo and L. Smolin, Lorentz Invariance with an Invariant Energy Scale, 88 (2002) 190403 J. Magueijo and L. Smolin, Gravity’s Rainbow, 21 (2004) 1725–1736, \[gr-qc/0305055\] E. Wilson-Ewing, Holonomy Corrections in the Effective Equations for Scalar Mode Perturbations in Loop Quantum Cosmology, 29 (2012) 085005, \[arXiv:1108.6265\] M. Mart[í]{}n-Benito, L. J. Garay, and G. A. Mena Marugán, Hybrid Quantum Gowdy Cosmology: Combining Loop and Fock Quantizations, 78 (2008) 083516, \[arXiv:0804.1098\] W. F. Blyth and C. J. Isham, Quantization of a Friedmann universe filled with a scalar field, 11 (1975) 768–778 W. Kaminski and J. Lewandowski, The flat FRW model in LQC: the self-adjointness, 25 (2008) 035001, \[arXiv:0709.3120\] W. Kaminski and T. Pawlowski, The LQC evolution operator of FRW universe with positive cosmological constant, 81 (2010) 024014, \[arXiv:0912.0162\] W. Kaminski, J. Lewandowski, and T. Pawlowski, Physical time and other conceptual issues of QG on the example of LQC, 26 (2009) 035012 K. V. Kuchař, Time and interpretations of quantum gravity, In G. Kunstatter, D. E. Vincent, and J. G. Williams, editors, [ *Proceedings of the 4th Canadian Conference on General Relativity and Relativistic Astrophysics*]{}, Singapore, 1992. World Scientific C. J. Isham, Canonical Quantum Gravity and the Question of Time, In J. Ehlers and H. Friedrich, editors, [*Canonical Gravity: From Classical to Quantum*]{}, pages 150–169. Springer-Verlag, Berlin, Heidelberg, 1994 E. Anderson, The Problem of Time in Quantum Gravity, \[arXiv:1009.2157\] P. Malkiewicz, Reduced phase space approach to Kasner universe and the problem of time in quantum theory, 29 (2012) 075008, \[arXiv:1105.6030\] M. Bojowald, D. Brizuela, H. H. Hernandez, M. J. Koop, and H. A. Morales-Técotl, High-order quantum back-reaction and quantum cosmology with a positive cosmological constant, 84 (2011) 043514, \[arXiv:1011.3022\] F. Cametti, G. Jona-Lasinio, C. Presilla, and F. Toninelli, Comparison between quantum and classical dynamics in the effective action formalism, In [*Proceedings of the International School of Physics “Enrico Fermi”, Course CXLIII*]{}, pages 431–448, Amsterdam, 2000. IOS Press, \[quant-ph/9910065\] G. W. Gibbons, S. W. Hawking, and J. M. Stewart, A Natural Measure On The Set Of All Universes, 281 (1987) 736 S. W. Hawking and D. N. Page, How probable is inflation?, 298 (1988) 789–809 J. S. Schiffrin and R. M. Wald, Measure and Probability in Cosmology, \[arXiv:1202.1818\] M. Bojowald and A. Skirzewski, Effective Equations of Motion for Quantum Systems, 18 (2006) 713–745, \[math-ph/0511043\] M. Bojowald and T. Strobl, Poisson Geometry in Constrained Systems, 15 (2003) 663–703, \[hep-th/0112074\] C. Rovelli, , Cambridge University Press, Cambridge, UK, 2004 T. Thiemann, , Cambridge University Press, Cambridge, UK, 2007, \[gr-qc/0110034\] A. Ashtekar and J. Lewandowski, Background independent quantum gravity: A status report, 21 (2004) R53–R152, \[gr-qc/0404018\] C. Rovelli and L. Smolin, The physical Hamiltonian in nonperturbative quantum gravity, 72 (1994) 446–449, \[gr-qc/9308002\] T. Thiemann, Quantum Spin Dynamics [(QSD)]{}, 15 (1998) 839–873, \[gr-qc/9606089\] M. Bojowald, B. Sandhöfer, A. Skirzewski, and A. Tsobanjan, Effective constraints for quantum systems, 21 (2009) 111–154, \[arXiv:0804.3365\] M. Bojowald and A. Tsobanjan, Effective constraints for relativistic quantum systems, 80 (2009) 125008, \[arXiv:0906.1772\] M. Bojowald and A. Tsobanjan, Effective constraints and physical coherent states in quantum cosmology: A numerical comparison, 27 (2010) 145004, \[arXiv:0911.4950\] M. Bojowald, P. A. Höhn, and A. Tsobanjan, An effective approach to the problem of time, 28 (2011) 035006, \[arXiv:1009.5953\] M. Bojowald, P. A. Höhn, and A. Tsobanjan, An effective approach to the problem of time: general features and examples, 83 (2011) 125023, \[arXiv:1011.3040\] P. A. Höhn, E. Kubalova, and A. Tsobanjan, Effective relational dynamics of the closed FRW model universe minimally coupled to a massive scalar field, \[arXiv:1111.5193\] A. Ashtekar, New Hamiltonian Formulation of General Relativity, 36 (1987) 1587–1602 J. F. Barbero G., Real Ashtekar Variables for Lorentzian Signature Space-Times, 51 (1995) 5507–5510, \[gr-qc/9410014\] G. Immirzi, Real and Complex Connections for Canonical Gravity, 14 (1997) L177–L181 H. Sahlmann, Some Comments on the Representation Theory of the Algebra Underlying Loop Quantum Gravity, \[gr-qc/0207111\] H. Sahlmann, When Do Measures on the Space of Connections Support the Triad Operators of Loop Quantum Gravity?, \[gr-qc/0207112\] J. Lewandowski, A. Okołów, H. Sahlmann, and T. Thiemann, Uniqueness of diffeomorphism invariant states on holonomy-flux algebras, 267 (2006) 703–733, \[gr-qc/0504147\] C. Fleischhack, Representations of the Weyl Algebra in Quantum Geometry, 285 (2009) 67–140, \[math-ph/0407006\] A. Ashtekar, J. Lewandowski, D. Marolf, J. Mourão, and T. Thiemann, Quantization of Diffeomorphism Invariant Theories of Connections with Local Degrees of Freedom, 36 (1995) 6456–6493, \[gr-qc/9504018\] C. Rovelli and L. Smolin, Discreteness of Area and Volume in Quantum Gravity, 442 (1995) 593–619, \[gr-qc/9411005\], Erratum: [*Nucl. Phys. B*]{} 456 (1995) 753 A. Ashtekar and J. Lewandowski, Quantum Theory of Geometry I: Area Operators, 14 (1997) A55–A82, \[gr-qc/9602046\] A. Ashtekar and J. Lewandowski, Quantum Theory of Geometry II: Volume Operators, 1 (1998) 388–429, \[gr-qc/9711031\] M. Varadarajan, On the resolution of the big bang singularity in isotropic Loop Quantum Cosmology, 26 (2009) 085006, \[arXiv:0812.0272\] A. Henderson, A. Laddha, and C. Tomlin, Constraint algebra in LQG reloaded : Toy model of a ${\rm U}(1)^{3}$ Gauge Theory I, \[arXiv:1204.0211\] M. Bojowald, Large scale effective theory for cosmological bounces, 75 (2007) 081301(R), \[gr-qc/0608100\] M. Bojowald, Dynamical coherent states and physical solutions of quantum cosmological bounces, 75 (2007) 123512, \[gr-qc/0703144\] T. Thiemann, : Quantum Gravity as the Natural Regulator of Matter Quantum Field Theories, 15 (1998) 1281–1314, \[gr-qc/9705019\] M. Bojowald, G. Calcagni, and S. Tsujikawa, Observational test of inflation in loop quantum cosmology, 11 (2001) 046, \[arXiv:1107.1540\] M. Bojowald, Inverse Scale Factor in Isotropic Quantum Geometry, 64 (2001) 084018, \[gr-qc/0105067\] M. Bojowald, Quantization ambiguities in isotropic quantum geometry, 19 (2002) 5113–5130, \[gr-qc/0206053\] M. Bojowald, Degenerate Configurations, Singularities and the Non-Abelian Nature of Loop Quantum Gravity, 23 (2006) 987–1008, \[gr-qc/0508118\] J. Lewandowski and D. Marolf, Loop Constraints: A Habitat and their Algebra, 7 (1998) 299–330, \[gr-qc/9710016\] R. Gambini, J. Lewandowski, D. Marolf, and J. Pullin, On the Consistency of the Constraint Algebra in Spin Network Quantum Gravity, 7 (1998) 97–109, \[gr-qc/9710018\] A. Laddha and M. Varadarajan, The Diffeomorphism Constraint Operator in Loop Quantum Gravity, \[arXiv:1105.0636\] K. V. Kuchař, Geometry of hypersurfaces. I, 17 (1976) 777–791 M. Bojowald, Loop quantum cosmology and inhomogeneities, 38 (2006) 1771–1795, \[gr-qc/0609034\] M. Bojowald, The dark side of a patchwork universe, 40 (2008) 639–660, \[arXiv:0705.4398\] K. Hepp, The Classical Limit for Quantum Mechanical Correlation Functions, 35 (1974) 265–277 M. Bojowald, S. Brahma, and E. Nelson, Higher time derivatives in effective equations of canonical quantum systems, \[arXiv:1208.1242\] N. C. Dias, A. Mikovic, and J. N. Prata, Coherent States Expectation Values as Semiclassical Trajectories, 47 (2006) 082101, \[hep-th/0507255\] C. Rovelli, Quantum Mechanics Without Time: A Model, 42 (1990) 2638–2646 C. Rovelli, Time in Quantum Gravity: An Hypothesis, 43 (1991) 442–456 C. Rovelli, Quantum evolving constants: Reply to comment on ‘Time in quantum gravity: An Hypothesis’, 44 (1991) 1339–1341 R. M. Wald, A Proposal for solving the ‘problem of time’ in canonical quantum gravity, 48 (1993) 2377–2381, \[gr-qc/9305024\] A. Higuchi and R. M. Wald, Applications of a new proposal for solving the ‘problem of time’ to some simple quantum cosmological models, 51 (1995) 544–561, \[gr-qc/9407038\] T. Thiemann, The Phoenix Project: Master Constraint Programme for Loop Quantum Gravity, 23 (2006) 2211–2248, \[gr-qc/0305080\] B. Dittrich and T. Thiemann, Testing the Master Constraint Programme for Loop Quantum Gravity I. General Framework, 23 (2006) 1025–1066, \[gr-qc/0411138\] K. Vandersloot, On the Hamiltonian Constraint of Loop Quantum Cosmology, 71 (2005) 103506, \[gr-qc/0502082\] M. Bojowald, Loop Quantum Cosmology, 11 (2008) 4, \[gr-qc/0601085\], O. Hrycyna, J. Mielczarek, and M. Szyd[ł]{}owski, Effects of the quantisation ambiguities on the Big Bounce dynamics, 41 (2009) 1025–1049, \[arXiv:0804.2778\] M. Bojowald, D. Mulryne, W. Nelson, and R. Tavakol, The high-density regime of kinetic-dominated loop quantum cosmology, 82 (2010) 124055, \[arXiv:1004.3979\] M. Bojowald, Quantum nature of cosmological bounces, 40 (2008) 2659–2683, \[arXiv:0801.4001\] M. Bojowald, How quantum is the big bang?, 100 (2008) 221301, \[arXiv:0805.1192\] A. Ashtekar, T. Pawlowski, and P. Singh, Quantum Nature of the Big Bang: An Analytical and Numerical Investigation, 73 (2006) 124038, \[gr-qc/0604013\] A. Ashtekar, T. Pawlowski, P. Singh, and K. Vandersloot, Loop quantum cosmology of k=1 FRW models, 75 (2007) 024035, \[gr-qc/0612104\] E. Bentivegna and T. Pawlowski, Anti-deSitter universe dynamics in LQC, 77 (2008) 124025, \[arXiv:0803.4446\] M. Bojowald, What happened before the big bang?, 3 (2007) 523–525 M. Bojowald, Harmonic cosmology: How much can we know about a universe before the big bang?, 464 (2008) 2135–2150, \[arXiv:0710.4919\] W. Kaminski and T. Pawlowski, Cosmic recall and the scattering picture of Loop Quantum Cosmology, 81 (2010) 084027, \[arXiv:1001.2663\] E. R. Livine and M. Martín-Benito, Group theoretical Quantization of Isotropic Loop Cosmology, \[arXiv:1204.0539\] N. Pinto-Neto and E. S. Santini, Must Quantum Spacetimes Be Euclidean?, 59 (1999) 123517, \[arXiv:gr-qc/9811067\] T. Cailleteau, J. Mielczarek, A. Barrau, and J. Grain, Anomaly-free scalar perturbations with holonomy corrections in loop quantum cosmology, 29 (2012) 095010, \[arXiv:1111.3535\] J. D. Reyes, , PhD thesis, The Pennsylvania State University, 2009 M. Bojowald, J. D. Reyes, and R. Tibrewala, Non-marginal LTB-like models with inverse triad corrections from loop quantum gravity, 80 (2009) 084002, \[arXiv:0906.4767\] A. Perez and D. Pranzetti, On the regularization of the constraints algebra of Quantum Gravity in $2+1$ dimensions with non-vanishing cosmological constant, 27 (2010) 145009, \[arXiv:1001.3292\] M. Bojowald and H. A. Kastrup, Symmetry Reduction for Quantized Diffeomorphism Invariant Theories of Connections, 17 (2000) 3009–3043, \[hep-th/9907042\] M. Bojowald, Spherically Symmetric Quantum Geometry: States and Basic Operators, 21 (2004) 3733–3753, \[gr-qc/0407017\] M. Bojowald and R. Swiderski, Spherically Symmetric Quantum Geometry: Hamiltonian Constraint, 23 (2006) 2129–2154, \[gr-qc/0511108\] K. Banerjee and G. Date, Loop quantization of polarized Gowdy model on $T^3$: Classical theory, 25 (2008) 105014, \[arXiv:0712.0683\] K. Banerjee and G. Date, Loop Quantization of Polarized Gowdy Model on $T^3$: Quantum Theory, 25 (2008) 145004, \[arXiv:0712.0687\] M. Bojowald and G. M. Paily, Deformed General Relativity and Effective Actions from Loop Quantum Gravity, \[arXiv:1112.1899\] J. Mielczarek, A. Cailleteau, Barrau, T. and J. Grain, Anomaly-free vector perturbations with holonomy corrections in loop quantum cosmology, 29 (2012) 085009, \[arXiv:1106.3744\] M. Bojowald, G. Hossain, M. Kagan, and S. Shankaranarayanan, Gauge invariant cosmological perturbation equations with corrections from loop quantum gravity, 79 (2009) 043505, \[arXiv:0811.1572\] D. Langlois, Hamiltonian formalism and gauge invariance for linear perturbations in inflation, 11 (1994) 389–407 A. Connes, Formule de trace en geometrie non commutative et hypothese de Riemann, 323 (1996) 1231–1235 G. Calcagni, Fractal universe and quantum gravity, 104 (2010) 251301, \[arXiv:0912.3142\] K. Giesel, F. P. Schuller, C. Witte, and M. N. R. Wohlfahrt, Gravitational dynamics for all tensorial spacetimes carrying predictive, interpretable and quantizable matter, 85 (2012) 104042, \[arXiv:1202.2991\] M. Bojowald, G. M. Paily, J. D. Reyes, and R. Tibrewala, Black-hole horizons in modified space-time structures arising from canonical quantum gravity, 28 (2011) 185006, \[arXiv:1105.1340\] M. Bojowald and G. Calcagni, Inflationary observables in loop quantum cosmology, 1103 (2011) 032, \[arXiv:1011.2779\] T. Cailleteau and A. Barrau, Gauge invariance in Loop Quantum Cosmology : Hamilton-Jacobi and Mukhanov-Sasaki equations for scalar perturbations, \[arXiv:1111.7192\] T. Cailleteau, A. Barrau, J. Grain, and F. Vidotto, Consistency of holonomy-corrected scalar, vector and tensor perturbations in Loop Quantum Cosmology, \[arXiv:1206.6736\] J. Mielczarek, Signature change in loop quantum cosmology, \[arXiv:1207.4657\] A. Komar, Consistent Factor Ordering Of General Relativistic Constraints, 20 (1979) 830–833 M. Bojowald, H. Hernández, and A. Skirzewski, Effective equations for isotropic quantum cosmology including matter, 76 (2007) 063511, \[arXiv:0706.1057\] A. O. Barvinsky and A. Yu. Kamenshchik, One loop quantum cosmology: The Normalizability of the Hartle-Hawking wave function and the probability of inflation, 7 (1990) L181–L186 A. O. Barvinsky and A. Yu. Kamenshchik, Tunneling geometries. 1. Analyticity, unitarity and instantons in quantum cosmology, 50 (1994) 5093–5114, \[gr-qc/9311022\] A. O. Barvinsky and A. Yu. Kamenshchik, Quantum origin of the early universe and the energy scale of inflation, 5 (1996) 825–844, \[gr-qc/9510032\] A. O. Barvinsky, A. Yu. Kamenshchik, and C. Kiefer, Effective action and decoherence by fermions in quantum cosmology, 552 (1999) 420–444, \[gr-qc/9901055\] J. Ambj[ø]{}rn, A. Gorlich, J. Jurkiewicz, R. Loll, J. Gizbert-Studnicki, and T. Trzesniewski, The Semiclassical Limit of Causal Dynamical Triangulations, 849 (2011) 144–165, \[arXiv:1102.3929\] C. Rovelli, Loop Quantum Gravity, 1 (1998) 1, \[gr-qc/9710008\], A. Perez, Spin Foam Models for Quantum Gravity, 20 (2003) R43, \[gr-qc/0301113\] A. Perez, The Spin Foam Approach to Quantum Gravity, \[arXiv:1205.2019\] M. Reisenberger and C. Rovelli, “Sum over Surfaces” form of Loop Quantum Gravity, 56 (1997) 3490–3508, \[gr-qc/9612035\] M. Bojowald and A. Perez, Spin Foam Quantization and Anomalies, 42 (2010) 877, \[gr-qc/0303026\] A. Mikovic and M. Vojinovic, Effective action and semiclassical limit of spin foam models, 28 (2011) 225004, \[arXiv:1104.1384\] S. A. Hojman, K. Kuchař, and C. Teitelboim, Geometrodynamics Regained, 96 (1976) 88–135 K. V. Kuchař, Geometrodynamics regained: A Lagrangian approach, 15 (1974) 708–715 R. Gambini and J. Pullin, Canonical quantization of general relativity in discrete space-times, 90 (2003) 021301, \[gr-qc/0206055\] M. Campiglia, C. Di Bartolo, R. Gambini, and J. Pullin, Uniform discretizations: a new approach for the quantization of totally constrained systems, 74 (2006) 124012, \[gr-qc/0610023\] B. Bahr and B. Dittrich, Improved and Perfect Actions in Discrete Gravity, 80 (2009) 124030, \[arXiv:0907.4323\] B. Dittrich and P. A. Höhn, Canonical simplicial gravity, \[arXiv:1108.1974\] M. Bojowald and G. Hossain, Quantum gravity corrections to gravitational wave dispersion, 77 (2008) 023508, \[arXiv:0709.2365\] V. A. Belinskii, I. M. Khalatnikov, and E. M. Lifschitz, A general solution of the Einstein equations with a time singularity, 13 (1982) 639–667 M. Bojowald and G. M. Paily, A no-singularity scenario in loop quantum gravity, \[arXiv:1206.5765\] C. Kiefer and M. Kraemer, Quantum Gravitational Contributions to the CMB Anisotropy Spectrum, 108 (2012) 021301, \[arXiv:1103.4967\] E. J. Copeland, D. J. Mulryne, N. J. Nunes, and M. Shaeri, The gravitational wave background from super-inflation in Loop Quantum Cosmology, 79 (2009) 023508, \[arXiv:0810.0104\] G. Calcagni and M. V. Cortês, Inflationary scalar spectrum in loop quantum cosmology, 24 (2007) 829–853, \[gr-qc/0607059\] J. Mielczarek and M. Szydłowski, Relic gravitons as the observable for Loop Quantum Cosmology, 657 (2007) 20–26, \[arXiv:0705.4449\] J. Mielczarek, Gravitational waves from the Big Bounce, 0811 (2008) 011, \[arXiv:0807.0712\] M. Shimano and T. Harada, Observational constraints of a power spectrum from super-inflation in Loop Quantum Cosmology, 80 (2009) 063538, \[arXiv:0909.0334\] A. Barrau and J. Grain, Cosmological footprint of loop quantum gravity, 102 (2009) 081301, \[arXiv:0902.0145\] J. Grain, A. Barrau, and A. Gorecki, Inverse volume corrections from loop quantum gravity and the primordial tensor power spectrum in slow-roll inflation, 79 (2009) 084015, \[arXiv:0902.3605\] J. Grain, T. Cailleteau, A. Barrau, and A. Gorecki, Fully LQC-corrected propagation of gravitational waves during slow-roll inflation, 81 (2010) 024040, \[arXiv:0910.2892\] J. Mielczarek, T. Cailleteau, J. Grain, and A. Barrau, Inflation in loop quantum cosmology: Dynamics and spectrum of gravitational waves, 81 (2010) 104049, \[arXiv:1003.4660\] M. Bojowald, G. Calcagni, and S. Tsujikawa, Observational constraints on loop quantum cosmology, 107 (2011) 211302, \[arXiv:1101.5391\] L.-F. Li, R.-G. Cai, Z.-K. Guo, and B. Hu, Non-Gaussian features from the inverse volume corrections in loop quantum cosmology, \[arXiv:1112.2785\] G. Calcagni, Observational Effects from Quantum Cosmology, (2012) to appear, \[arXiv:1209.0473\] [^1]: Topical Review [*Class. Quantum Grav.*]{} 29 (2012) 213001 [^2]: e-mail address: [[email protected]]{} [^3]: Especially in the loop-quantum-gravity community with its long and proud history of mathematical-physics primacy, it is sometimes said that one must become “less rigorous” in order to find useful and interesting physical results. This statement is, of course, incorrect; one does not become a physicist by being a mathematician first and then turning a little less rigorous. Physics requires as much rigor as mathematics, but a different kind of rigor. [^4]: Fundamental issues of quantum cosmology are reviewed in the companion article [@QCFund].
{ "pile_set_name": "ArXiv" }
Bhabani Prasad Mandal [^1] and Saurabh Gupta [^2] Department of Physics, Banaras Hindu University, Varanasi-221005, INDIA. [**Abstract**]{} We consider a couple of examples to study the pseudo-Hermitian interaction in relativistic quantum mechanics. Rasbha interaction, commonly used to study the spin Hall effect, is considered with imaginary coupling. The corresponding Dirac Hamiltonian is shown to be parity pseudo-Hermitian. In the other example we consider parity pseudo-Hermitian scalar interaction with arbitrary parameter in Dirac theory. In both the cases we show that the energy spectrum is real and all the other features of non-relativistic pseudo-Hermitian formulation are present. Using the spectral method the positive definite metric operator ($\eta$) has been calculated explicitly for both the models to ensure positive definite norms for the state vectors. Introduction ============ Non-Hermitian quantum mechanics has recently created a lot of interest. This is due to the observation by Bender and Boettcher [@ben; @ben11] that with properly defined boundary conditions, the spectrum of the system described by the Hamiltonian $ H= p^2 + x^2 (ix)^\nu, \ \ \nu \geq 0$ is real positive and discrete. The reality of the spectrum is a consequence of unbroken PT (combined Parity \[P\] & Time reversal \[T\]) invariance of the Hamiltonian, $ [H,PT]\psi =0$ and the spectrum becomes partially complex when the PT symmetry is broken spontaneously [@pt3; @pt31]. This new result has given rise to growing interest in the literature, see for examples, [@pt3]-[@co]. Past few years many non-Hermitian but PT symmetric systems have been investigated including (i) quantum mechanics of single particle in one space dimension [@ben]-[@kha], (ii) exactly solvable many particle quantum systems in one space dimension [@bm]-[@bn], (iii) field theoretic models [@ft]-[@conf]. To develop a consistent quantum theory for these systems, one encounters severe difficulties [@wei; @jap]. Particularly, the eigenstates of PT symmetric non-Hermitian Hamiltonians, with real eigenvalues only, do not satisfy standard completeness relations. More importantly the eigenstates have negative norms if one takes the natural inner product associated with such systems defined as $$<f\mid g>_{PT} = \int d^3 x [PT f(x)] g(x).$$ These problems are overcome by introducing a new symmetry \[C\], analogous to charge conjugation symmetry, associated with all such systems with equal number of negative and positive norm states [@pt1; @co]. This allows to introduce an inner product structure associated with CPT conjugate as $$<f\mid g>_{CPT} = \int d^3 x [CPT f(x)] g(x), \label{cpt}$$ for which the norms of the quantum states are positive definite and one gets usual completeness relation. As a result, the Hamiltonian and its eigenstates can be extended to complex domain so that the associated eigenvalues are real and underlying dynamics is unitary. Thus we have a fully consistent quantum theory for non-Hermitian but PT invariant systems. In an alternative approach [@alir]-[@gey], it has been shown that the reality of the spectrum of a non-Hermitian system is due to so called pseudo-Hermiticity properties of the Hamiltonian. A Hamiltonian is called $\eta $ pseudo-Hermitian if it satisfies the relation $$\eta H \eta ^{-1} = H^\dagger\label{ps},$$ where $\eta $ is some linear Hermitian and invertible operator. All PT symmetric non-Hermitian systems are pseudo-Hermitian where parity operator plays the role of $\eta $. However there are many pseudo-Hermitian systems with completely real spectrum which are not PT symmetric [@bpm]-[@yb]. In the formulation of pseudo-Hermitian quantum mechanics, the scalar product is defined as $$<f\mid g>_\eta = <f\mid\eta g>=\int d^3 x f^*\eta g \label{eta},$$ to get rid of the conceptual difficulties arise in such theories. Equation (\[cpt\]) which defines the scalar product in case of non-Hermitian PT invariant theory is a special case of Eq. (\[eta\]) [@alin]. If the Hamiltonian is diagonalizable and has a real and discrete spectrum, it is always possible to find a positive definite $\eta$ [@alir; @rev0] such that the norms of all the eigenstates become positive definite and they satisfy usual completeness relation. Thus one can have completely consistent quantum theory for the pseudo-Hermitian systems provided atleast one positive definite $\eta $ is contructed. However, most of the previous works in the pseudo-Hermitian quantum mechanics have been carried out in the non-relativistic framework [^3]. The purpose of this paper is to extend the results of pseudo-Hermitian quantum mechanics for relativistic systems. Here we consider a couple of examples of pseudo-Hermitian Hamiltonian and show that the energy spectrum obtained by solving corresponding Dirac equation is also real. In the first example, we study the famous Rashba interaction [^4] [@ras] with imaginary coupling which is pseudo-Hermitian with respect to parity operator. We consider a Dirac particle in 1+1 dimension, moving under arbitrary scalar pseudo-Hermitian interaction in other example. The norms of the state vectors, defined according to the modified rule of scalar product \[ i.e. Eq. (\[eta\])\] will be positive definite if we find a positive $\eta $ for the above two models. We have constructed positive definite $\eta $ for both the models using spectral method [@spec; @spec1]. This paper is arranged as follows. In section II, we consider the motion of a Dirac particle in the x-y plane experiencing Rashba interaction with imaginary coupling. The pseudo-Hermitian scalar interaction of Dirac particle in 1+1 dimension is discussed in section III. Positive definite $\eta $ for both the models have been constructed explicitly in section IV. Last section reveals for conclusion and summary. Dirac equation with Rashba interaction ======================================= We consider the motion of a particle in the x-y plane described by the Hamiltonian $$H= c\mbox{\boldmath $\alpha $} \cdot {\mathbf p } +\beta m_0 c^2 +\lambda_1 (\mbox{\boldmath $\sigma $}\times{\mathbf p})\cdot {\mathbf\hat{z}} \label{2.1} ,$$ where the last term (with real $\lambda_1$) is known as the Rashba interaction term. This type of interaction is widely used to discuss spin Hall effect [@she; @she11; @she12; @she13] in non-relativistic formulation. However, the Hamiltonian is no longer Hermitian if we consider the $\lambda_1 $ is complex. Further it can be shown that this Hamiltonian is a parity pseudo-Hermitian when $\lambda_1 (\equiv i\lambda)$ is purely imaginary. $$\begin{aligned} H^\dag &=& c\mbox{\boldmath $ \alpha $}\cdot {\mathbf p} +\beta m_0 c^2 -i\lambda (\mbox{\boldmath $\sigma $}\times{\mathbf p})\cdot{\mathbf \hat{z}} \nonumber \\ &\neq & H \nonumber \\ &=& P H P^{-1},\end{aligned}$$ where $P (=\beta e^{i\delta})$ is parity operator in Dirac theory. ([$\sigma $]{} $\times {\mathbf p} )\cdot {\mathbf \hat{z}}$ changes sign under parity transformation $$P(\mbox{\boldmath $\sigma $}\times {\mathbf p} )\cdot {\mathbf \hat{z}}P^{-1} = -(\mbox{\boldmath $\sigma $}\times {\mathbf p} )\cdot {\mathbf \hat{z}},$$ as under parity: [$\sigma $]{}$\rightarrow $[$\sigma $]{} ,$ \ {\mathbf p}\rightarrow {\mathbf -p}, \ {\mathbf \hat{z}}\rightarrow -{\mathbf \hat{z}} $ and $\beta$ anti-commutes with $\sigma_x , \sigma_y$. Now we solve the Dirac equation $H\psi =E\psi $ for this system to find the energy eigenvalues. To do that we consider the wave function in terms of components, $ \psi =\left (\begin{array}{c} \phi \\ \chi \end{array} \right ) $ and restrict ourselves to 2-dimension only. Further, we consider the particular representation of Dirac matrices in 2-dimension as, $ {\mathbf\alpha}={\mathbf\sigma} , \beta =\left (\begin{array}{cc} 1 & 0 \\ 0 &-1 \end{array}\right ),$ where $ {\mathbf\sigma_i} [\mbox{ for } i=x,y,z]$ are Pauli spin matrices. Then the Hamiltonian in equation (\[2.1\]) can be written as $$\begin{aligned} H&=& c\sigma _x p_x + c\sigma_y p_y +\beta m_0c^2 + i \lambda(\sigma_xp_y -\sigma_y p_x) \nonumber \\ &=& \left [ \begin{array}{cc} m_0 c^2 & (c-\lambda)p_- \\ (c+\lambda)p_+ & -m_0c^2 \end{array} \right ], \label{2.2}\end{aligned}$$ where $ p_{\pm} = p_x\pm ip_y. $ The Dirac equation $H\psi = E\psi $ can be written as pair of coupled equations of the components $\psi $ and $\chi$ as $$\begin{aligned} (c-\lambda)p_- \chi &=& (E-m_0c^2)\phi , \nonumber\\ (c+\lambda)p_+ \phi &=& (E+m_0c^2)\chi \label{2.3}.\end{aligned}$$ On eliminating one component $\chi$ in terms of others, we obtain the equation $$p_-p_+\phi = \epsilon\phi \label{2.4},$$ where $\epsilon =\frac{E^2-m_0^2c^4}{c^2-\lambda^2}$. Equation (\[2.4\]) can be interpreted as Schrodinger equation for a free particle moving in the x-y plane. The corresponding energy eigenvalues $\epsilon$ are given as $\hbar^2(k_x^2 +k_y^2)$, which are real and positive. Therefore the energy eigenvalues E for the Dirac Hamiltonian for the pseudo-Hermitian system, described in Eq.(\[2.1\]), are given by $$E=\pm \sqrt{m_0^2 c^4 +(c^2-\lambda^2)\hbar^2(k_x^2+ k_y^2)} \label{2.5}.$$ Thus the energy spectrum for a pseudo-Hermitian \[ Rashba interaction with imaginary coupling \] relativistic system is completely real for $\lambda^2 < c^2 $. Same conclusion can be drawn by eliminating other component $\phi $ from the Eqs. in (\[2.3\]). Dirac equation with scalar pseudo-Hermitian interaction ======================================================= Let us consider a Dirac particle of rest mass $m_0$ subjected to a scalar pseudo-Hermitian potential $V_s$. The dynamics of such a system can be described by the Hamiltonian $$H= c\mbox{\boldmath $\alpha $}\cdot{\mathbf p} +\beta m_0 c^2 +V_s .\label{3.1}$$ For simplicity, we restrict ourselves to one space dimension and we choose the scalar pseudo-Hermitian potential as $$V_s = V(x) \left ( \begin{tabular}{c c} 0&1 \\ -1 &0 \end{tabular} \right ), \label{3.2}$$ where $V(x)$ is an arbitrary real function and assumed to have even parity, $V(-x) = V(x)$, for later convenience. We note that the Hamiltonian given in Eq. (\[3.1\]) is not Hermitian as $V_s^\dag = -V_s \neq V_s$, but it is parity pseudo-Hermitian, i.e. $$H^\dag = PH P^{-1},$$ where $P$ is the parity operator and in relativistic formulation is given by $P= \beta e^{i\delta}$. This is so because, $ c\alpha\cdot {\mathbf p} +\beta m_0 c^2$ is a parity invariant term and $$PV_s P^{-1} = \beta e^{i\delta} V(x)\left ( \begin{tabular}{c c} 0&1 \\ -1 & 0 \end{tabular} \right )\beta^{-1}e^{-i\delta}= -V_s = V_s^\dag ,$$ where we have chosen a particular representation of Dirac matrices $\alpha$ and $\beta$ in 1+1 dimension as $$\alpha_x = \sigma_x \mbox{ and }\beta= \left ( \begin{tabular}{c c} 1&0 \\ 0 &-1 \end{tabular} \right ),$$ $\sigma_x$ is the first of Pauli’s $ 2\times 2$ spin matrices. 1+1 dimensional Dirac equation for the system can be written as $$\left [ c\sigma_x p_x +\beta m_0 c^2 +V_s\right ]\psi= E\psi . \label{3.3}$$ By considering $\psi=\left ( \begin{array}{c} \phi \\ \chi \end{array} \right ),$ the above equation can be written in component form $$\begin{aligned} \left [ V(x)- ic\hbar\frac{d}{dx} \right ] \chi &=& (E-m_0c^2)\phi , \label{3.35} \\ \left [- V(x)- ic\hbar\frac{d}{dx} \right ] \phi &=& (E+m_0c^2)\chi \label{3.4}.\end{aligned}$$ By eliminating $\chi$ from the above coupled differential equations, we obtain $$\left [-\hbar^2\frac{d^2}{dx^2}+U \right ]\phi =\epsilon \phi \label{sch},$$ where $$U= -i\hbar\frac{dV}{dx}-V^2 \mbox{ and } \epsilon= E^2-m_0^2c^4 .$$ Eq. (\[sch\]) is nothing but non-relativistic Schrodinger equation for a particle of mass $m=\frac{1}{2}$ subjected to a complex potential $U$ and $\epsilon $ is the energy eigenvalues for the particle. Even though the potential U is complex, remarkably it is parity pseudo-Hermitian as $$U^* = \tilde{P}U\tilde{P}^{-1}.$$ where $\tilde{P}$ is parity operator in non-relativistic quantum theory. Note we have assumed $V(-x) = V(x) $. Hence the energy eigenvalues,$\epsilon$ for this effective non-relativistic theory is real or occure in complex conjugate pairs[@ali]depending on whether the symmetry is broken or not. . The energy eigenvalues of the relativistic particle is given in terms of $\epsilon$ by $$E= \pm\sqrt{\epsilon+m_0^2c^4} \label{e}.$$ Now if $\epsilon$ is real and positive, the whole spectrum of the relativistic particle is real without any restriction. On the other hand, if $\epsilon$ is real but negative i.e. $\epsilon= -|\epsilon|$, then also the spectrum is real provided $|\epsilon|<m_0^2c^4 $. However for arbitrary $V(x)$ energy eigenvalues are not real always as expected. Exactly similar conclusion can be drawn by eliminating $\phi$ from the equations (\[3.35\]) and (\[3.4\]). Now we consider a special case when, $V(x)= V_0$, independent of x $$V_s = V_0 \left ( \begin{tabular}{c c} 0&1 \\ -1 &0 \end{tabular} \right ), \label{3.2}$$ where $V_0$ is an arbitrary real constant. The Dirac equation in Eq. \[(\]3.1) can be written as $$\begin{aligned} \left [ \begin{array}{cc} m_0c^2 & cp_x+V_0 \\ cp_x-V_0 &-m_0c^2 \end{array} \right ] \ \left ( \begin{array}{c} \phi \\ \chi \end{array} \right ) = E\left ( \begin{array}{c} \phi \\ \chi \end{array} \right ), \label{3.4}\end{aligned}$$ and in terms of components, $$\begin{aligned} (cp_x+V_0)\chi &=& (E-m_0c^2)\phi , \nonumber \\ (cp_x-V_0)\phi &=& (E+m_0c^2)\chi .\end{aligned}$$ By solving the above coupled differential equations, we obtain the energy eigenvalues as $$E= \pm\sqrt{\hbar^2 c^2 k_x^2 +m^2_0c^4-V_0^2} \label{3.7}$$ with $$p_x^2\phi = \hbar^2k_x^2\phi$$ The energy eigenvalues are always real for sufficiently weak pseudo-Hermitian interaction. Construction of positive definite $\eta $ ========================================== We have shown in both the examples that even in relativistic quantum mechanics we can have real eigenvalues if the relativistic Hamiltonian is parity pseudo-Hermitian. In order to have a consistent formulation for these systems we need to construct a positive definite $\eta $ such that inner product is well defined. In this section we intend to construct the positive definite $\eta $ explicitly by using the spectral method as defined in Refs. [@spec; @spec1]. The positive definite $\eta $ for a $\eta $-pseudo Hermitian theory described by the Hamiltonian $H$ is defined as [@spec; @spec1], $${\bf \eta }= \sum_{i=1}^2|u_i><u_i|, \label{5.1}$$ where $|u_i>;\ \ i=1,2$ are the Dirac spinors associated with $H^\dagger ( = \eta H \eta ^{-1})$. Following this method we construct the positive definite $\eta $ for both the models discussed above. [**Case I: Pseudo-Hermitian Rasbha Interaction:**]{} In this case, $H^\dagger$ can be written from Eq. (\[2.2\]) as $$H^\dagger = \left [ \begin{array}{cc} m_0 c^2 & (c+\lambda){\bf p_-} \\ (c-\lambda){\bf p_+} & -m_0c^2 \end{array} \right ], \label{5.2}$$ and let $\left ( \begin{array}{c} v_1 \\ v_2 \end{array}\right ) $ be the two component spinor for $H^\dagger $ corresponding to the energy E, and satisfy, $$H^\dagger \left ( \begin{array}{c} v_1 \\ v_2 \end{array}\right ) =E\left ( \begin{array}{c} v_1 \\ v_2 \end{array}\right )$$ The spinors are calculated as $$|u_1>= \left ( \begin{array}{c} 1 \\ \frac{(c-\lambda)p_+}{E+m_0c^2} \end{array}\right ); \mbox{and } |u_2>= \left ( \begin{array}{c} \frac{-(c+\lambda)p_-}{E+m_0c^2 }\\ 1 \end{array}\right ) \label{5.4}$$ when $E$ is given in Eq. (\[2.5\]) and $p_+, p_-$ are eigenvalues of the operator [**$p_+$**]{} and [**$p_-$**]{} respectively. Substituting Eq. (\[5.4\]) in Eq. (\[5.1\]) we obtain the positive definite $\eta $, as $$\eta_I = \left [ \begin{array}{cc} 1+\frac{(c+\lambda)^2}{(E-m_0c^2)^2}p_+p_- & \frac{2p_-(cE+ \lambda m_0c^2)}{E^2-m_0^2c^4} \\ \frac{2p_+(cE+ \lambda m_0c^2)}{E^2-m_0^2c^4} & 1+\frac{(c-\lambda)^2}{(E+m_0c^2)^2}p_+p_- \end{array} \right ]$$ [**Case II: Pseudo-Hermitian scalar interaction**]{}\ It is difficult to obtain the spinors associated with $H^\dagger $ in this case for an arbitary $V(x)$ because $H^\dagger $ and the operators $p_{\pm}$ do not have simaltaneous eigen functions as $[ H^\dagger, p_{\pm}]\neq 0$. However for the special case when $V(x)$ is independent of x we can substitutes the operators $p_{\pm}$ by their eigenvalues as $[ H^\dagger, p_{\pm}]= 0$. We therefore construct the positive definite $\eta $ for the special case For this model, $$H^\dagger = \left [ \begin{array}{cc} m_0 c^2 & c{\bf p_x}-V_0 \\ c{\bf p_x}+V_0 & -m_0c^2 \end{array} \right ]. \label{5.6}$$ Let $\left ( \begin{array}{c} v_1 \\ v_2 \end{array}\right ) $ be the two components spinor for $H^\dagger $ corresponding to the energy E, and satisfy, $$H^\dagger \left ( \begin{array}{c} v_1 \\ v_2 \end{array}\right ) =E\left ( \begin{array}{c} v_1 \\ v_2 \end{array}\right ),$$ $|u_1>$ and $|u_2>$ are calculated as $$|u_1>= \left ( \begin{array}{c} 1 \\ \frac{cp_x+V_0}{E+m_0c^2} \end{array}\right ); \mbox{and } |u_2>= \left ( \begin{array}{c} \frac{-(cp_x-V_0)}{E+m_0c^2 }\\ 1 \end{array}\right ), \label{5.7}$$ when $E$ is given by Eq. (\[3.7\]). Putting $|u_1>$ and $|u_2>$ from Eq. (\[5.7\]) in Eq. (\[5.1\]) we obtain $$\eta_{II} = \left [ \begin{array}{cc} 1+(\frac{cp_x+V_0}{E-m_0c^2})^2- & \frac{2Ecp_x-2V_0 m_0c^2}{E^2-m_0^2c^4} \\ \frac{2Ecp_x-2V_0 m_0c^2}{E^2-m_0^2c^4} & 1+(\frac{cp_x-V_0}{(E+m_0c^2)})^2\end{array} \right ]$$ We have constructed a positive definite $\eta $ for both the examples discussed in this paper. These positive definite $\eta $’s ensure the positive definite norms for all the state vectors associated with this theory and hence lead to a fullly consistent relativistic pseudo-Hermitian theory. Conclusion ========== We have considered two completely different pseudo-Hermitian interactions in relativistic quantum mechanics. In the first example, we have dealt with Rashba interaction with imaginary coupling which plays an important role in studying spin Hall effect [@she; @she11; @she12; @she13]. This interaction is shown to be parity pseudo-Hermitian and the complete spectrum which we obtain by solving Dirac equation on the plane is real. It will be interesting to see whether the solutions of imaginary Rashba interaction lead to any new consequence in the study of spin Hall effect. Scalar pseudo-Hermitian interaction has been constructed with an arbitrary real parameter in 1+1 dimension and is shown that the spectrum is real by solving corresponding Dirac equation. Using spectral method we have further constructed a positive definite metric operator, $\eta $ for both the examples. Such positive definite $\eta $ ensures the positive definite norm for the state vectors. Thus we have a fully consistent quantum theory for the pseudo-Hermitian interactions in relativistic theory. [ **Acknowledgment:**]{} We are grateful to Prof. Ali Mostafazadeh for making valuable comments/suggestions on the earlier version of the manuscript which have helped improve the paper substantially. [99]{} C.M. Bender and S. Boettcher, [*Phys. Rev. Lett.*]{} [**80**]{} (1998) 5243. C.M. Bender, S. Boettcher and P.N. Meisinger, [*J. Math. Phys.*]{} [**40**]{} (1999) 2210. A. Khare and B. P. Mandal, [*Phys. Lett.* ]{} [**A 272**]{} (2000) 53. M Znojil and G Levai, [*Mod. Phys. Lett.*]{} [**A 16**]{}, (2001) 2273. A. Mostafazadeh , [*J. Phys* ]{} [**A 38**]{} (2005) 6657, Erratum-ibid.[**A 38**]{} (2005) 8185. M Znojil, [*Phys. Lett.*]{} [**A 259**]{}, (1999) 220. M Znojil, [*J. Phys.*]{} [**A 35**]{} (2002) 8793, ibid [**A 35**]{} (2002) 2341. C. M. Bender, D. C.Brody, J. Chen, H. F. Jones, K. A. Milton and M. C. Ogilvie, Phy. Rev. [**D 74**]{}(2006) 025016 and see refs therein. C M Bender, K. Besseghir, H F Jones and X. Yin, [**arXiv: 0906:1291**]{} (2009). G Levai, P. Siegl and M Znojil, [**arXiv: 0906:2333**]{}, (2009). A. Khare and B. P. Mandal, [*Spl issue of Pramana J of Physics*]{} [**73**]{} (2009), 387. Z. Ahmed, [*Phys. Lett.*]{} [**A 324**]{} (2004) 152; ibid. [**A 294** ]{} (2002), 287. P. Dorey, C. Dunning and R. Tateo, [*J. Phys A: Math. Theor..*]{} [**34**]{} (2001) 5679. B.Bagchi, C. Quesne, [*Phys. Lett.*]{} [**A273**]{} (2000) 256. S. Rajvjani, A.K. Kapoor and P. K. Panigrahi, [**quant-ph/0403054**]{} (2004). A. Khare and U. Sukhatme, [**quant-ph/0402106**]{} (2004). B. Basu-Mallick and B.P. Mandal, [*Phys. Lett.* ]{} [**A 284**]{} (2001) 231. B. Basu-Mallick, [*Int. J. of Mod. Phys. B*]{}, [**16**]{} (2002) 1875. B. Basu-Mallick, T. Bhattacharyya A. Kundu, and B. P. Mandal [*Czech. J. Phys* ]{} [**54**]{} (2004) 5. B. Basu-Mallick, T. Bhattacharyya and B. P. Mandal, [ *Mod. Phys.Lett.*]{} [**A 20** ]{}, (2004), 543. P. K. Ghosh and K. S. Gupta, [*Phys. Lett.*]{} [**A 323**]{} (2004), 29. Y. Brihaye and A. Ninimahazwe, [*Int.J. Mod. Phys.*]{} [**A 19**]{} (2004), 517. C.M. Bender, D.C. Brody and H. F. Jones, [*Phys. Rev.*]{} [**D 70**]{} (2004), 025001; Erratum-ibid. [**D 71**]{} (2005) 049901. C. M. Bender, S.F. Brandt, J.Chen and Q. Wang, [*Phys. Rev.*]{} [**D 71**]{} (2005) 065010. C. M. Bender, H.F. Jones and R. J. Rivers, [*Phys. Lett.*]{} [**B 625**]{} (2005) 333. K. A. Milton, [*Czech J. Phys.*]{} [**54**]{} (2004) 1069. P. Dorey, C Dunning and R. Tateo, [**arXiv:0906.1130**]{} (2009). S. Weigert, [*Phys. Rev.* ]{} [**A 68**]{} (2003) 062111, [*J. Opt.*]{} [**B 5**]{} (2003), S416. G. Japaridze, [*J. Phys. A*]{} [**35**]{} (2002), 1709. C.M. Bender and S. Boettcher, [*Phys. Rev. Lett.*]{} [**89**]{} (2002) 270401-1; Erratum-ibid. [**92**]{} (2004) 119902 . C.M. Bender, J. Brod, A. Refig and M. Reuter, [**quant-ph/0402026.**]{} A. Mostafazadeh, [**arXiv: 0810.5643**]{}, (2008). A. Mostafazadeh, [*J. Math Phys.*]{} [**43** ]{} (2002) 205; [**43**]{} (2002) 2814; [**43**]{} (2002) 3944. A. Mostafazadeh, [**arXiv: 0901:4472**]{} (To appear in PRL) (2009). A. Mostafazadeh and A. Batal, [*J. Phys A: Math. and theor.*]{} , [**37**]{},(2004) 11645. A. Mostafazadeh, [*J. Phys A: Math. and theor.*]{} , [**36**]{},(2003) 7081. A. Mostafazadeh, [*Nucl. Phys. B*]{} [**640**]{} (2002) 419; [*J. Math. Phys.*]{} [**44**]{} (2003) 974; [*IJMPA*]{}, [**21**]{}, 2553 (2006) ; [*J. Phys.*]{} [**A 37**]{} (2004) 10193; F. G. Scholtz, H. B. Geyer, F. J. W. Hahne, [*Ann. Phys.*]{} [**213** ]{} (1992) 74. B. P. Mandal, [*Mod. Phys. Lett.*]{} [**A 20**]{} (2005) 655. P.K. Ghosh, [*J. Phys.*]{} [**A 38**]{} (2005), 7313. Y. Brihaye and A. Nininahazwe, [**quant-ph/0506249**]{}. A. Mostafazadeh, [*J. of Phys.* ]{} [**A 38**]{} (2005), 3213. A. Mostafazadeh and F. Zamani, [*Ann Phy*]{}, [**321**]{},(2006) 2183 ; ibid (2006) 2210. F. Kleefeld, [*Czech J. Phys*]{}, [**56**]{},(2006) 999 . A. Mostafazadeh,[*Ann. Phys*]{} [**309**]{} (2004), 1. F. Cannata and A. Ventura, [*J. Phys A: Math. Theor.*]{} [**41**]{}(2008) 505305. A. Sinha, P.Roy, [*Mod. Phys. Lett.*]{} [**A 20**]{} (2005), 2377. E. I. Rashba, [*Phy. Rev.*]{} [**B 68**]{} (2003), 241315; ibid. [**B 70**]{}, (2004), 201309. H. Engle, E.I. Rashba, B. I. Halperin, [**cond-mat/0603306**]{}, and refs. therein. J. E. Hirsch, [*Phys. Rev. Lett.*]{} [**83**]{} (1999) 1834. C. L. Kane & E. J. Male [*Phys. Rev. Lett.*]{} [**95**]{} (2005) 226801. J. Shi, P. Zhang, D. Xiao and Q. Niu, [*Phys. Rev. Lett*]{} [**96**]{} (2006) 076604. [^1]: [email protected] ; [email protected] [^2]: [email protected] [^3]: Some attempts to study relativistic PT invariant non-Hermitian system can be found in Refs. [@rev0]-[@rev]. [^4]: This type of spin orbit coupling is widely used to study spin Hall effect [@she; @she11; @she12; @she13]. In the simplest version of spin Hall effect, an electric current passes through a sample with spin orbit interaction and induces a spin polarization near the lateral edges, with opposite polarization at opposite edges. This effect does not require an external magnetic field or magnetic order in the equilibrium state before the current is applied.
{ "pile_set_name": "ArXiv" }
--- author: - | Hadi Salman[^1]\ `[email protected]`\ Microsoft Research - | Andrew Ilyas\ `[email protected]`\ MIT - | Logan Engstrom\ `[email protected]`\ MIT - | Ashish Kapoor\ `[email protected]`\ Microsoft Research - | Aleksander Mdry\ `[email protected]`\ MIT bibliography: - 'bibliography/bib.bib' title: 'Do Adversarially Robust ImageNet Models Transfer Better?' --- Introduction {#sec:intro} ============ Motivation: Fixed-Feature Transfer Learning {#sec:motivation} =========================================== Adversarial Robustness and Full-Network Fine Tuning =================================================== Analysis and Discussion {#sec:analysis} ======================= Related Work ============ Conclusion ========== Experimental Setup {#app:experimental-setup} ================== Detailed Numerical Results {#app:numerical-results} ========================== [^1]: Equal contribution.
{ "pile_set_name": "ArXiv" }
--- abstract: 'It is argued that the Unruh effect may be interpreted as a duality of a theory on different backgrounds. This issue is revisited in String Theory in the path integral formalism. By using T-duality and the Unruh effect, the T-dual transformation for acceleration is investigated and a maximum effective physical acceleration for observers in String Theory is found.' author: - 'Heliudson Bernardo[^1]' bibliography: - 'References.bib' title: Unruh Duality and Maximum Acceleration in String Theory --- Introduction {#sec:intro} ============ String theory is a promising candidate for a physical theory of quantum gravity. It is based on the assumption that one-dimensional extended fundamental strings play a major role in the theory in many regimes. There is only one dimension-full parameter in the theory, the string length $\alpha'^{1/2}$, and all other parameters are vacuum expectation values of dynamical fields of the theory [@Polchinski:1998rq; @Polchinski:1998rr]. In some sense, this fundamental length is the minimum possible distance probed in String Theory. In scattering of strings at weak coupling, the hard limit of the amplitudes indicates objects of size $\alpha'^{1/2}$ [@Veneziano:1974dr]. Also, on toroidal compactifications of String Theory, there is an equivalence of large and small radii, when compared to the string length. This equivalence and its generalizations are referred to as T-duality between backgrounds (see [@Giveon:1994fu] and references therein). The self-dual radius, given exactly by the string length, yields a minimum effective physical length for the radii of the tori. Considering the low-energy limit of String Theory, a Supergravity solution is generally not supposed to hold in regions where the curvature is bigger than the string scale or at distances smaller than its characteristic length. In this case, corrections due to string-size effects are necessary. Usually, in any attempt to construct a field theory or to describe physical phenomena at energy scales closer to the inverse string length, new stringy effects appear [@Green:1987sp; @Green:1987mn]. Heuristically, relativistic theories with a minimum fundamental length $\lambda$ have an upper bound on the possible maximum acceleration $a_c$ given by $$a_c \sim \frac{1}{\lambda}.$$ Extended objects have also a critical acceleration due to causality arguments. In fact, for an object with proper length $\lambda$, if some point on it has an acceleration bigger than $1/\lambda $ then it will be causally disconnected from other points on the object, since it will perceive a Rindler horizon at distance smaller than $\lambda$ [@Misner:1974qy]. So, consistent acceleration of fundamental extended objects should satisfy $a < 1/\lambda$. From this, one sees that accelerated fundamental strings should be treated carefully. One should also recall that in String Theory, the fundamental objects are $\textit{quantum}$ objects, such that simple classical heuristic arguments may not apply directly for fundamental strings. In fact, quantum mechanically, we also find possible limits on acceleration [@Caianiello1984]. For strings, in [@Sanchez:1989cw; @Gasperini:1990pn] it was shown that arbitrary large acceleration leads to instabilities in the string proper’s length. A universal maximum critical acceleration was first recognized as an intrinsic property of String Theory in [@Frolov:1990ct] (see also [@Gasperini:1992jz; @Feoli:1992vmd; @Martellini:1987ug; @Sanchez:1991vv]). Also, strings in Rindler spacetime were considered in [@deVega:1987um], motivated by the connection between accelerated frames and black hole backgrounds. A divergence in string partition functions was interpreted as maximal acceleration close to Rindler horizon in [@Mertens:2015adr], though the authors concluded that such divergence is due to an extrapolation of field theory to strings. In any analysis of quantum relativistic systems on accelerated frames one should consider the Fulling-Davies-Unruh, Davies-Unruh, or simply Unruh effect [@Fulling:1972md; @Davies:1974th; @Unruh:1976db; @Unruh:1983ac] (for a review, see [@Crispino:2007eb]). An observer accelerating in the inertial vacuum will be in a thermal bath with temperature that depends on its acceleration. So, quantization in its frame should be done taking this thermal bath into account. The Unruh effect in String Theory was considered in several papers from different perspectives [@deVega:1987um; @Parentani:1989gq; @Feoli:1993ew; @Witten:2018xfj] (for connections with holography, see [@Paredes:2008cr]). A more systematic approach in terms of path integrals, analogous to the analysis of [@Unruh:1983ac] is presented in section \[sec3\]. Strings at finite temperature have interesting new properties when compared with particles [@Atick:1988si] (see also references therein). More striking is the existence of a maximum temperature for a gas of strings, the Hagedorn temperature[^2] $T_H$ [@Hagedorn:1965st]. Also, due to the existence of winding modes in the string case, T-duality implies that there is a duality between temperatures greater and smaller than a self-dual temperature of the order of $T_H$. Thus, effectively, there is a maximum possible temperature for a thermal bath of strings [@Atick:1988si]. From this observation and the connection between temperature and acceleration through the Unruh effect, we may suspect that there should possibly be a maximum acceleration for observers in String Theory. A relation between T-duality of temperature and acceleration seems very plausible at least. The goal of this work is to study and elucidate the connection between the Unruh effect, acceleration, temperature and T-duality in String Theory. The paper is organized as follows: In section \[sec2\] the Unruh effect as formulated in the path integral formalism is reviewed and it is argued that it may be seen as a duality (technical details are given in appendix \[appendix\]); in section \[sec3\], a path integral proof of the Unruh effect first for relativistic point particle and then for strings is given; based on T-duality of temperature and the Unruh effect, a T-duality transformation of acceleration is proposed in section \[sec4\]; section \[sec5\] is a discussion of the results and a possible relation with the Hagedorn phase transition; in section \[sec6\] the conclusions are presented. #### On notation. {#on-notation. .unnumbered} Natural units are adopted throughout the paper, $c=1,\; \hbar=1,\; k_B =1$, and the subscript $E$ in some time variables is used to emphasize Euclidean signature. The Unruh effect as a duality {#sec2} ============================= Quantum Field Theory is constructed in a Lorentz invariant way, such that the vacuum of a theory is the same for all inertial observers. The fact that two observers non-related by a Lorentz transformation assign different states for their respective vacua is the main reason for the Unruh Effect, which states that a uniform accelerated observer will experience a thermal bath in the vacuum state of inertial observers. The temperature $T$ of the thermal state as seen by the accelerated observer is related with its proper acceleration $a$ by the Unruh temperature formula, $$\label{unruhtemperature} T= \frac{a}{2\pi}.$$ In practice, one can show this by noticing that creation and annihilation operators in the mode expansion of fields in the accelerated frame are related to the corresponding operators in the inertial frame by a Bogoliubov transformation [@Crispino:2007eb]. It was also shown in [@Unruh:1983ac] that the Unruh effect can be derived directly from the partition function of the theory. In fact, the Euclidean path integral for a field theory defined on a Rindler wedge, with Rindler metric $$ds^2 = G_{\mu\nu}^{(a)}(x) dx^\mu dx^\nu= a^2r^2d\eta_E^2 + dr^2 +\delta_{ij}dx_{\perp}^idx_{\perp}^j,$$ and with time compactified on a circle of radius $\beta/2\pi$ is identical to the path integral in Euclidean space provided $a\beta = 2\pi$, that is, $$\label{unruhduality} Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu} = G_{\mu\nu}^{(2\pi/\beta)}\right]= Z_{\mathbb{R}^D}\left[G_{\mu\nu}= \delta_{\mu\nu}\right].$$ The subscript in $Z$ is a reminder of where the coordinates appearing in the metric take values. For the Rindler wedge with Euclidean periodic time, we have $\left(\eta_E, r, x^i_\perp\right)\in S^1_{\beta/2\pi} \times\mathbb{R}_{+}^{*} \times\mathbb{R}$, as $r>0$, and the form of the Rindler metric implies that the theory in the left hand side of (\[unruhduality\]) is defined on a cone (with deficit angle $2\pi - a \beta $) times flat directions. Thus, for $a\beta = 2\pi$ it is topologically equivalent to a flat space. Since the Euclidean path integral with time taking values in $S^1_{\beta/2\pi}$ defines a field theory at finite temperature with temperature $T=1/\beta$, the condition on the acceleration for the Rindler metric in the left side of equation (\[unruhduality\]) is equivalent to equation (\[unruhtemperature\]). Therefore, equation (\[unruhduality\]) is a formal way to state the Unruh effect: the Euclidean generator for correlators for the accelerated observer is the same as the inertial one provided the thermal bath is taken into account *and* its temperature is related with the proper acceleration via (\[unruhtemperature\]). These results were proved for any interacting field theory containing scalars and spin-$1/2$ fields, and in the presence of sources in [@Unruh:1983ac]. For a review of the calculation, see appendix \[appendix\]. So, *if* $a \beta =2\pi$, any correlation function to be measured by the accelerated observer at finite temperature should be equal to the correlation functions that inertial observers would measure[^3]. Generally, the equality of partition functions implies that the theories are the same or that at least they are dual to each other. Having theories in different backgrounds giving the same partition function is the definition of a duality between the backgrounds. So, from equation (\[unruhduality\]) we conclude that the Unruh effect may be seen as a duality between different backgrounds for the same relativistic quantum theory. Finite temperature String Theory on the Rindler wedge {#sec3} ===================================================== Similar to what was done for scalar and spin-$1/2$ fields in [@Unruh:1983ac] (see also appendix \[appendix\]), we can consider the string partition function on a Rindler wedge with compactified time and prove the Unruh effect in String Theory in the path integral formalism. Before doing that, let us consider the relativistic point particle case, $$\begin{aligned} Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(a)}\right] = \int [dXde]_R &\exp\left\{-\frac{1}{2}\int d\tau \left[e^{-1}\left(a^2r^2{\Dot{\eta}_E}^2 +\right.\right.\right.\nonumber\\ &+\left.\left.\left.\Dot{r}^2+ \delta_{ij}\Dot{X}^i_{\perp}\Dot{X}^j_{\perp}\right)+em^2\right]\vphantom{\frac{1}{2}}\right\},\end{aligned}$$ where $e(\tau)$ is the einbein of the worldline $X^\mu(\tau)=(\eta_E(\tau), r(\tau), X^i_{\perp}(\tau))$ parameterized by $\tau$ and overdots denotes derivation with respect to this parameter. The subscript $R$ indicates that the path integral is over worldlines in the compactfied Rindler wedge, i.e., $\left(\eta_E(\tau), r(\tau)\right) \in S^1_{\beta/2\pi}\times \mathbb{R}_{+}^{*}$. Making the following change of variables in the path integral, $$T(\tau) = r(\tau)\sin(a\eta_E(\tau)), \quad X(\tau) = r(\tau)\cos(a\eta(\tau)),$$ gives $$\begin{aligned} Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(a)}\right] = \int [dXde]_R&\exp\left\{-\frac{1}{2}\int d\tau \left[e^{-1}\left(\Dot{T}^2+ \Dot{X}^2 + \delta_{ij}\Dot{X}^i_{\perp}\Dot{X}^j_{\perp}\right) em^2 \vphantom{\frac{1}{2}}\right]\right\} . \end{aligned}$$ Due to the periodicity of $\eta_E$, not all the $(T,X)$ plane is accessible for the integrated worldlines. But for the particular value $\beta = 2\pi/a$, the path integral in the new variables is over worldlines in all coordinate space, that is, for this specific value of $\beta$ we have $-\infty <T,X,X^i_{\perp}< \infty$. In this case we can drop the subscript $R$ and write $$\begin{aligned} Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(a= 2\pi/\beta)}\right] &= \int [dXde]\exp\left\{-\frac{1}{2}\int d\tau \left[e^{-1}\delta_{\mu\nu}\Dot{X}^\mu\Dot{X}^\nu+em^2\right]\right\}\nonumber\\ &= Z_{\mathbb{R}^D}\left[\delta_{\mu\nu}\right],\end{aligned}$$ which proves the Unruh duality for a relativistic point particle. Similarly, we can consider the string partition function on a Rindler wedge with compactfied time, $$Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(a)}\right] = \int [dX d\gamma]_R \exp\left\{-\frac{1}{4\pi\alpha'}\int d^2\sigma \sqrt{-\gamma}\gamma^{ab}\partial_a X^\mu \partial_b X^\nu G_{\mu\nu}^{(a)} \right\},$$ where the subscript $R$ in the functional measure is a reminder that we are summing over embeddings of the worldsheet onto the compactified Rindler wedge, i.e., that the target space function’s components take values in $S^1_{\beta/2\pi}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}$. We have $$\begin{aligned} Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(a)}\right] = \int [dX d\gamma]_R \exp&\left\{-\frac{1}{4\pi\alpha'}\int d^2\sigma \sqrt{-\gamma}\gamma^{ab}\left[a^2r^2\partial_a \eta_E \partial_b \eta_E + \right.\right.\nonumber\\ &+\left.\left.\partial_a r\partial_b r + \delta_{ij}\partial_a X_{\perp}^i \partial_b X^j_{\perp} \right]\vphantom{\frac{1}{2}} \right\},\end{aligned}$$ with $$\quad 0< \eta< \beta, \quad 0 < r< \infty, \quad -\infty< X^i_{\perp}<\infty.$$ Making the following change of variables $$\begin{aligned} X(\sigma) &= r(\sigma)\cos{(a\eta(\sigma))},\\ T(\sigma) &= r(\sigma)\sin{(a\eta(\sigma))},\end{aligned}$$ we get $$\begin{aligned} Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(a)}\right] = \int [dX d\gamma]_R \exp&\left\{-\frac{1}{4\pi\alpha'}\int d^2\sigma \sqrt{-\gamma}\gamma^{ab}\left[\partial_a T \partial_b T +\right.\right.\\ &+\left.\left.\partial_a X \partial_b X + \delta_{ij}\partial_a X_{\perp}^i \partial_b X^j_{\perp} \right] \right\}\nonumber\\ = \int [dX d\gamma]_R \exp&\left\{-\frac{1}{4\pi\alpha'}\int d^2\sigma \sqrt{-\gamma}\gamma^{ab}\partial_a X^\mu \partial_b X^\nu \delta_{\mu\nu} \right\}.\end{aligned}$$ If we now consider $a\beta =2\pi$, as in the point particle case, the new variables cover the entire target space and we have[^4] $$\label{thermalstringduality} Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(a = \frac{2\pi}{\beta})}\right] =Z_{\mathbb{R}^D}\left[\delta_{\mu\nu}\right].$$ So, the Polyakov path integral for an accelerating observer in a stringy thermal bath is equal to the Polyakov path integral for an inertial observer *if* the acceleration and temperature are related by the Unruh temperature formula (\[unruhtemperature\]). Notice that this case is physically different of having an accelerating string as seen by an inertial observer. T-duality transformation for acceleration {#sec4} ========================================= Let us step back and consider strings at finite temperature as seen by inertial observers. This case is different from having finite temperature strings for accelerated observers with a Rindler metric, that was considered in the previous section; as the vacuum state for inertial observers appears as a thermal state for the accelerated observers, a thermal state for inertial observers *is not* a thermal state for the accelerated ones. For the finite temperature strings in Minskowski spacetime, since the thermal partition function has a compactified Euclidean direction, we can do a T-duality transformation on this direction and this should leave the string partition function invariant. Indeed, in the inertial background, this may be used to find the T-duality transformation of temperature, the radius $R = \beta/2\pi$ transforming as $$\frac{\beta}{2\pi} \rightarrow \frac{\Tilde{\beta}}{2\pi} = \frac{2\pi}{\beta}\alpha' \implies \Tilde{\beta} = \frac{4\pi^2\alpha'}{\beta}.$$ This implies that the T-dual temperature is (see chapter 9 in [@Polchinski:1998rq]) $$\label{tdualitytemperature} \Tilde{T} = \frac{4 T^2_H}{T},$$ where $T_H \equiv 1/(4\pi\alpha'^{1/2})$ is the Hagedorn temperature [@Atick:1988si]. From this T-duality transformation we can calculate the self-dual temperature $T_*$, given by $$T_* = 2T_H = \frac{1}{2\pi \alpha'^{1/2}}.$$ Now, consider two accelerated observers with different constant proper accelerations, but such that the corresponding thermal baths have T-dual temperatures. If String Theory describe their physics, the observers would never differ between these two temperatures, as both are physically equivalent. In a sense, they would be “dual" to each other, as their measurements of physical quantities associated with the thermal baths would be the same. Thus, by the relation between temperature and acceleration, it is possible to define a “dual" acceleration. To find how an observer’s acceleration “transforms" under T-duality, we use the Unruh duality, equation (\[thermalstringduality\]) twice: for the observer that measures a thermal bath at temperature $T$ and then for the observer at dual temperature $\Tilde{T}$, $$\begin{aligned} \label{unruhstring} Z_{S^1_{\frac{\beta}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(a= 2\pi/\beta)}\right] &= Z_{\mathbb{R}^D}\left[\eta_{\mu\nu}\right]= Z_{S^1_{\frac{\Tilde{\beta}}{2\pi}}\times\mathbb{R}_{+}^{*} \times\mathbb{R}^{D-2}}\left[G_{\mu\nu}^{(\Tilde{a})}\right],\end{aligned}$$ so we should have $$\Tilde{a} = \frac{2\pi}{\Tilde{\beta}} = \frac{\beta}{2\pi\alpha'} = \frac{1}{a\alpha'}.$$ Therefore, we found the T-duality transformation for observer’s acceleration, $$a \rightarrow \Tilde{a} = \frac{1}{a\alpha'}.$$ The self-dual acceleration is given by $$a_*= \frac{1}{\alpha'^{1/2}},$$ and it corresponds to the self-dual temperature by (\[unruhtemperature\]). So, from Unruh duality and T-duality, we conclude that the thermal state for the observer with acceleration $a$ is a thermal state with T-dual temperature for the observer with acceleration $\Tilde{a}$. Since their respective partition functions are equal, both observers would measure the same correlators and so amplitudes, upon considering their respective thermal baths with temperatures related by T-duality. As far as string amplitudes are concerned, they could not distinguish between being accelerated at uniform acceleration $a$ and in thermal bath at temperature $T$ or being accelerated at uniform acceleration $1/(a\alpha')$ and in thermal bath at temperature $4T^2_H/T$, by probing *only* their thermal baths. This is in the same level as physical equivalence of compactifications on tori with radius $R$ or $\alpha'/R$. Therefore, we conclude that effectively, there is a maximal possible acceleration in String Theory, given by the self-dual value $a_*$, on the same grounds as there is an effective minimal length given by the self-dual radius $R_*= \alpha'^{1/2}$. Discussion {#sec5} ========== T-duality is a duality between different but equivalent backgrounds in String Theory. The results in this paper indicates that T-duality may also be used to relate measurements of different observers in String Theory, a property that seems not to have been explored before. Notice that this “application” of T-duality depends more on the state of motion of the T-dual-related observers than on the geometry of the background. In the case presented here, we were able to find the “T-dual motion” of proper acceleration. Generalizations of this case are under investigation. It is worth emphasizing that the equalities in equation (\[unruhstring\]) are *not* due to T-duality transformation of the metric. In fact, using Buscher’s rules [@Buscher:1987sk; @Buscher:1987qj] to find out the T-dual Rindler metric in the $S^1_{\beta/2\pi}$ direction, we would get $$d\Tilde{s}^2 = \frac{1}{a^2r^2}d\Tilde{\eta}_E^2 + dr^2 +\delta_{ij}dx_{\perp}^idx_{\perp}^j,$$ with $\Tilde{\eta}_E \in S^1_{\Tilde{\beta}/2\pi}$, which is not a Rindler metric for acceleration $1/a$. But the proper times of the observers with accelerations connected by T-duality are the same, and this indicates that they are related[^5]. Regardless of such observations, T-duality was only used to construct the thermal bath with T-dual temperature and to argue that the physics in the presence of both thermal baths is the same. Since thermal baths in String Theory were considered, one possible question is the relation of the present work with the Hagedorn phase transition [@Atick:1988si]. It is known that the thermal partition function for a string gas diverges at temperatures equal or greater than Hagedorn temperature, physically due to the exponential energy growth in the density of string states. This divergence is believed to signal a first order phase transition, not totally understood (see also [@Antoniadis:1991kh]). But the temperature due to the Unruh effect cannot give rise to phase transitions, since the thermal state as seen by the accelerated observer is just the vacuum state of the inertial observers. Accelerated fields or strings were not considered in [@Unruh:1983ac] or in the current work. What was done in this paper is different than calculating the thermal partition function for a gas of strings in Minskowski spacetime. Only the usual Polyakov path integral for an inertial string was used, but as seen by different observers. So, this work does not offer any new approach for the Hagedorn phase transition. A potential problem in the arguments of the previous section is the fact that the self-dual temperature $T_*$ is larger than the Hagedorn temperature and so its usage is subtle. But note that the value of the self-dual temperature comes from the T-duality transformation formula, equation (\[tdualitytemperature\]), that holds as long as the target space has an compactified Euclidean time direction. Also, there are thermal string constructions free of Hagedorn instabilities (see for example [@Florakis:2010is; @Kounnas:2011fk; @PandoZayas:2002hh; @Chaudhuri:2014hoa]) and the arguments of the previous section can be applied for the bosonic sector of them. Also, there could be potential applications to Euclidean type $\text{II}^*$ theories[^6] (see [@Hull:1998vg] and references therein), as they can be seen as T-dual to type II strings (though in the time-like direction). Notice that the arguments to find the maximum acceleration presented in the previous section are generally valid, due to the universality of T-duality and the Unruh effect. In particular, the relation between the maximum acceleration $a_{\text{max}}$ and the self-dual temperature, $$a_{\text{max}} = 2\pi T_*,$$ is expected to hold true in general and it is the main result of the present work. Conclusion {#sec6} ========== In this paper, a novel proof of the Unruh effect in String Theory was given, in the path integral formalism. Using this result and T-duality, a T-dual transformation of an observer’s proper acceleration was found and it was argued that there is an effective maximum acceleration in String Theory, as would be heuristically expected from a theory with minimal length. All of these results were obtained in bosonic String Theory. Generalizations of this particular case are under investigation, even though only mild modifications are physically expected. Acknowledgments {#acknowledgments .unnumbered} =============== The author would like to thank Jéssica Martins, George Matsas, Gabriel Cozzella, Yigit Yargic and Guilherme Franzmann for interesting discussions and Robert Brandenberger, Horatiu Nastase and Keshav Dasgupta for useful comments on previous versions of the manuscript. This work is fully supported by CAPES-Brazil. The author is also thankful to McGill University for hospitality during the execution of this work. Unruh duality in field theory {#appendix} ============================= In [@Unruh:1983ac] it was shown that all vacuum Green’s functions between spacetime points within the same Rindler wedge for inertial observers are the same as the Green’s functions for accelerated observers in thermal equilibrium at a temperature $T = a/2\pi$. This was done starting from the phase space path integral for scalar and spin-$1/2$ fields. In this appendix, the core of such results is reviewed, by showing equation (\[thermalstringduality\]) for spin-0 and spin-$1/2$ theories. Let us consider scalar fields first. The starting point is the Euclidean partition function for the accelerated observers: $$Z^R(\beta) = \int [d\phi]_P\exp\left\{-S_E^R(\beta)\right\},$$ where $$\begin{aligned} S^R_E(\beta) = \int_0^{\beta}d\eta_E\int_{r>0}dr d^{D-2}x_{\perp}\sqrt{G^E}\left[\frac{1}{2}(G^E)^{\alpha \beta}\partial_{\alpha}\phi\partial_{\beta}\phi + V(\phi)\right],\end{aligned}$$ with $$(ds^E)^2 = G^E_{\alpha \beta}dx^{\alpha}dx^{\beta}= a^2r^2d\eta_E^2 + dr^2+ \delta_{ij}dx_{\perp}^idx_{\perp}^j.$$ The superscript $R$ indicates that the region of integration in the action is over the Rindler wedge and the subscript $P$ denotes that the path integral is over field configurations with periodic boundary condition in time with period $\beta$, ${\phi(\eta_E=0)=\phi(\eta_E=\beta)}$. Changing variables to $$\label{changeofvar} t_E = r\sin(a\eta_E), \qquad x = r\cos(a\eta_E),$$ with $\beta a \leq 2\pi$ (in order to this transformation be single valued), we have $$(ds^E)^2 = \delta_{\mu\nu}dx^{\mu}dx^{\nu}=dt_E^2 + dx^2 + \delta_{ij}dx_{\perp}^idx_{\perp}^j,$$ and the Euclidean path integral is $$\begin{aligned} Z^R(\beta) = \int[d\phi]_P\exp\left\{- \int_R dt_Edx d^{D-2}x_{\perp}\left[\frac{1}{2}\delta^{\mu\nu}\partial_{\mu}\phi \partial_{\nu}\phi + V(\phi)\right]\right\}. \end{aligned}$$ The periodic boundary condition becomes $$\phi(0) = \phi(r\sin(a\beta)),$$ and then, if $\beta = 2\pi/a$, it turns into a consistency condition. Also, for this value of $\beta$, the region $R$ is the full $(t_E,x,x_{\perp}^i)$ space and so we get $$Z^R(\beta = 2\pi/a) = \int [d\phi] \exp \left\{-S_E\right\},$$ which is the Euclidean generating functional for the theory in inertial coordinates, i.e., for inertial observers. This proves Unruh duality for scalar theories. For spin-$1/2$ case, we need to consider the Dirac action in Rindler coordinates. Let us write the action for a Dirac fermion in covariant form, $$S = -\int d^Dx e(x)\Bar{\psi}(x)(\gamma^a\nabla_a+m)\psi(x),$$ where $e = \det e_\mu^{\;\;a}$ is the determinant of the vielbein fields $e_{\mu}^{\;\;a}$, $\gamma^a$ are the Dirac matrices and $$\gamma^a\nabla_a = \gamma^ce_c^{\;\;\mu}\left(\partial_\mu - \frac{i}{2}(\omega_\mu)^{ab}\Sigma_{ab}\right),$$ with $(\omega_\mu)^{ab}$ the spin connection and $\Sigma^{ab} = -i/4[\gamma^a,\gamma^b]$ the Lorentz generators in spin-$1/2$ representation. For the Rindler metric, the only non-vanishing independent component of the spin connection is $$\omega^0_{\;\;1} = ad\eta,$$ such that the Dirac action on the Rindler wedge $r> 0$ is $$\begin{aligned} S^R= -\int_{r>0}d\eta dr d^{D-2}x_{\perp}ar\Bar{\psi}\left(\frac{1}{ar}\gamma^0 \partial_{\eta} + \gamma^1\partial_r + \gamma^i\partial_i+\frac{1}{2r}\gamma^1 + m\right)\psi.\end{aligned}$$ Now, consider the thermal partition function for the accelerated observers, $$Z^R(\beta) = \int[d\Bar{\psi}d\psi]_{\psi(0) = -\psi(\eta_E = \beta)}\exp\left\{-S_E^R(\beta)\right\},$$ where the fermion fields have antiperiodic boundary condition in time with period $\beta$ and $$\begin{aligned} S_E^R(\beta) = \int^{\beta}_0d\eta_E\int_{r>0}dr d^{D-2}x_{\perp}ar \Bar{\psi}\left(\frac{i}{ar}\gamma^0 \partial_{\eta_E} +\gamma^1\partial_r + \gamma^i\partial_i+\frac{1}{2r}\gamma^1 + m\right)\psi,\end{aligned}$$ is the Euclidean Dirac action in Rindler coordinates. Performing a change of variables in the functional path integration by $$\begin{aligned} \psi(x) \rightarrow M(\eta_E)\psi(x) &\equiv \left(\cos(a\eta_E/2)+ i\gamma^0\gamma^1\sin(a\eta_E/2)\right)\psi(x),\\ \Bar{\psi}(x) \rightarrow \Bar{\psi}(x)M^{\dagger}(\eta_E) &\equiv \Bar{\psi}(x)\left(\cos(a\eta_E/2)- i\gamma^0\gamma^1\sin(a\eta_E/2)\right),\end{aligned}$$ leads to $$\begin{aligned} Z^R(\beta) = \int[d\Bar{\psi}d\psi]_A &\exp\left\{-\int_0^\beta d\eta_E \int_{r>0} dr d^{D-2}x_{\perp} ar \left[\frac{i}{ar}\Bar{\psi}\Bar{\Sigma}^0 \partial_{\eta_E}\psi + \Bar{\psi}\Bar{\Sigma}^1\partial_r\psi +\right.\right.\nonumber\\ &+\left.\left.\Bar{\psi}\gamma^i\partial_i\psi+ m\Bar{\psi}\psi\right]\vphantom{\frac{1}{2}}\right\}.\end{aligned}$$ The subscript $A$ in the functional measure indicates the antiperiodic condition that is now written as $$\psi(\eta_E= 0) = - \left(\cos(a\beta/2)+ i\gamma^0\gamma^1\sin(a\beta/2)\right)\psi(\eta_E = \beta).$$ Using the same change of variable as in the scalar case equation (\[changeofvar\]), after some algebra we get $$Z^R(\beta) = \int[d\Bar{\psi}d\psi]_A\exp\left\{-\int_Rdt_Edxd^{D-2}x_{\perp}\Bar{\psi}\left[i\gamma^0\partial_{t_E}+ \gamma^1\partial_x + \gamma^i \partial_i +m\right]\psi \right\}.$$ If now we take $\beta =2\pi/a$, the antiperiodic condition becomes a consistency condition and the region $R$ turns into all space in $(t_E,x,x^i_{\perp})$ coordinates. Thus, $$Z^R(\beta = 2\pi/a) = \int[d\Bar{\psi}d\psi]\exp\left\{-S_E \right\},$$ where $S_E$ is the Dirac action in inertial coordinates. This is the Unruh duality for spin-$1/2$ fields. In [@Unruh:1983ac] it is argued that we can use the results presented in this appendix to prove the Unruh effect for any interacting field theory involving scalars fields and Dirac spinors. [^1]: Email: <[email protected]> [^2]: For investigations on string thermodynamics near the Hagedorn temperature and in Rindler spacetime, see [@Mertens:2013pza; @Mertens:2013zya] and references therein. [^3]: The local operators in the correlators of both theories should be evaluated at the corresponding spacetime points for both observers. Also, one should consider the relative transformation of the operators for each observer. [^4]: This result was also explored in [@Mertens:2016tqv] by explicitly computing the partition functions. [^5]: Also, even considering $a\beta\neq2\pi$, it is not clear if the fact that the Rindler coordinates do not cover the entire space plays a big role in “T-dualizing” observers. [^6]: The author is grateful to Keshav Dasgupta for bringing that to his attention.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present quantitative investigations of the weak lensing effect on the two-point correlation functions of local maxima (hotspots), ${\xi_{\rm pk-pk}}(\theta)$, in the cosmic microwave background (CMB) maps. The lensing effect depends on the projected mass fluctuations between today and the redshift $z_{\rm rec}\approx1100$. If adopting the Gaussian assumption for the primordial temperature fluctuations field, the peak statistics can provide an additional information about the intrinsic distribution of hotspots that those pairs have some characteristic separation angles. The weak lensing then redistributes hotspots in the observed CMB maps from the intrinsic distribution and consequently imprints non-Gaussian signatures onto ${\xi_{\rm pk-pk}}(\theta)$. Especially, since the intrinsic ${\xi_{\rm pk-pk}}(\theta)$ has a pronounced depression feature around the angular scale of $\theta\approx 70''$ for a flat universe, the weak lensing induces a large smoothing at the scale. We show that the lensing signature therefore has an advantage to effectively probe mass fluctuations with large wavelength modes around $\lambda\approx 50 h^{-1}{\rm Mpc}$. To reveal the detectability, we performed numerical experiments with specifications of [*MAP*]{} and [*Planck Surveyor*]{} including the instrumental effects of beam smoothing and detector noise. It is then found that our method can successfully provide constraints on amplitude of the mass fluctuations and cosmological parameters in a flat universe with and without cosmological constant, provided that we use maps with $65\%$ sky coverage expected from Planck.' author: - Masahiro Takada and Toshifumi Futamase title: 'Detectability of Gravitational Lensing Effect on the Two-point Correlation Function of Hotspots in the CMB maps' --- Introduction ============ The temperature anisotropies in the comic microwave background (CMB) contain detailed information about the underlying cosmological model ([@HSS]). The recent high precision balloon-borne experiments, [Boomerang]{} ([@boom]; [@lange]) and MAXIMA ([@Maxima]; [@balbi]), revealed that the measured angular power spectrum $C_l$ are in good agreement with that predicted by standard inflation paradigm. On the other hand, the large-scale structure of the universe imprints secondary effects on the primordial temperature fluctuations. One of them is the weak lensing effect; the CMB photons are randomly deflected by the gravitational field due to the intervening large-scale structure during the propagations from the last scattering surface to a telescope. The weak lensing can be a powerful probe for mapping inhomogeneous distribution of dark matter in the universe ([@Gunn]; see [@BSreview] for a review), which is not directly attainable by any other means. In fact, several groups ([@Ludovic]; [@Bacon]; [@wtk]; [@kaiser]) recently have reported significant detections of the coherent distortion of faint galaxies images arising from the gravitational lensing by the foreground large-scale structure, and showed that those results can provide some constraints on the cosmological parameters. However, it would be still extremely interesting to be able to measure the lensing effects on the CMB and the detection would be very precise for constraining the cosmological parameters because there is no ambiguity in theoretical understanding of the unlensed CMB physics and about the distance of the source plane. The inflationary scenarios also predict that the primordial temperature fluctuations are Gaussian. In this case, the statistical properties of any unlensed CMB field can be exactly predicted based on the Gaussian random theory developed by Bardeen et al. (1986; hereafter BBKS) and Bond & Efstathiou (1987; hereafter BE) for three- and two-dimensional cases, respectively. However, the weak lensing then induces the non-Gaussian signatures in the observed CMB maps. Based on these considerations, some specific features on the lensed temperature fluctuations field have been revealed. Bernardeau (1998) investigated how the weak lensing alters the probability density function (PDF) of the ellipticities defined from the local curvature matrix of the temperature fluctuations field. The unlensed PDF indeed has specific statistical properties for the two dimensional Gaussian field, and then the gravitational distortion induces an excess of elongated structures in the CMB maps in the similar way as the distortions of distant galaxies. Although the method could be a powerful probe to measure the lensing signatures around the characteristic curvature scale ($\sim 5'$) of the unlensed temperature field, the instrumental effect of a finite beam size is crucial for the detection because the beam smearing effect again tends to circularize the deformed local structures. Hence, Van Wearbeke, Bernardeau & Benabed (1999) investigated how the weak lensing causes a coherent distribution of the relative orientation between the CMB and distant galactic ellipticities, and proposed that it can be a efficient tool because the orientation of the CMB ellipticities is robust against the beam smearing. We ([@TKF]; hereafter TKF) recently investigated the weak lensing effect on the two-point correlation function of local maxima ([*hotspots*]{}), say $\xi_{\rm pk-pk}(\theta)$, in the two dimensional CMB maps. Since the distribution of hotspots is a point process, the analysis focused on the secondary effect how the weak lensing redistributes hotspots in the observed CMB maps from the intrinsic distribution. The unlensed ${\xi_{\rm pk-pk}}(\theta)$ can be then accurately predicted by the Gaussian random theory once $C_l$ is given (BE; [@HS] hereafter HS). According to the acoustic peaks in the $C_l$, the pairs of hotspots have some characteristic angular separations on the last scattering surface. We then found that the weak lensing fairly smooths out the oscillatory shape of $\xi_{\rm pk-pk}(\theta)$. In particular, the most interesting result is that the lensing contribution to ${\xi_{\rm pk-pk}}(\theta)$ at angular scales ($\approx 70'$) corresponding to the first Doppler peak of $C_l$ is relatively large, and we thus expect that ${\xi_{\rm pk-pk}}(\theta)$ can be a sensitive statistical tool to measure the projected mass fluctuations at such larger scales than the other methods do. The crucial quantity of our method is the lensing fluctuations of relative angular separation between two CMB photons, and the lensing signatures to ${\xi_{\rm pk-pk}}(\theta)$ at such large scales is the consequence of large scale modes of lensing deflection angles. The simulated maps indeed illustrate that each displacement of peak positions in the lensed maps from the unlensed maps is relatively large even though both global features of pattern of temperature fluctuations nearly trace each other ([@Zald] and also see Figure \[fig:cmbmap\]). Furthermore, these considerations lead to the expectation that our method is not particularly affected by the beam smearing effect. The purpose of this paper, therefore, is to quantitatively investigate the weak lensing effect on ${\xi_{\rm pk-pk}}(\theta)$ and reveal in detail the physical interpretations of the effect. For the practical purpose we perform quantitative investigations of the problem whether the future satellite missions, [*MAP*]{}[^1] and [*Planck Surveyor*]{}[^2], can measure the lensing signatures to ${\xi_{\rm pk-pk}}(\theta)$ for constraining the cosmological parameters. This can be done by using numerical experiments of both unlensed and lensed CMB maps including the instrumental effects of finite beam size and detector noise. This paper is organized as follows. In the next section, we introduce the formalism to investigate the lensing effect on ${\xi_{\rm pk-pk}}(\theta)$. In Section 4, the formalism is applied to some specific cosmological models, and we then compute the signal-to-noise ratios of the lensing signatures to ${\xi_{\rm pk-pk}}(\theta)$ using the numerical experiments in Section 5. Finally, in Section 6 we present discussions and conclusions. Weak Lensing Effect on Two-point Correlations of Hotspots in the CMB: Formalism =============================================================================== A bundle of CMB photons is randomly deflected by the inhomogeneous matter distributions of the intervening large-scale structure as it propagates from the last scattering surface to a telescope. The two CMB bundles observed with a certain angular separation $\theta$ thus have a different angular separation when emitted from the last scattering surface. The ensemble averages of the second moment of the relative separation fluctuations and the following characteristic function can be then easily calculated by using the power spectrum approach developed by Seljak (1996); $$\begin{aligned} &&\sigma_{\rm GL}^2(\theta)\equiv2^{-1} {\left\langle{(\delta{\mbox{\boldmath$\theta$}}_1-\delta{\mbox{\boldmath$\theta$}}_2)^2}\right\rangle} _{|{\mbox{\boldmath$\theta$}}_1 -{\mbox{\boldmath$\theta$}}_2|=\theta} =\sigma_{\rm GL,0}^2(\theta)+\sigma^2_{\rm GL,2}(\theta), \nonumber \\ &&{\left\langle{\exp[i{\mbox{\boldmath$l$}}\cdot(\delta{\mbox{\boldmath$\theta$}}_1-\delta{\mbox{\boldmath$\theta$}}_2)]}\right\rangle} _{|{\mbox{\boldmath$\theta$}}_1 -{\mbox{\boldmath$\theta$}}_2|=\theta}\simeq 1-\frac{l^2}{2}\left[\sigma^2_{\rm GL,0}(\theta) +\cos(2\varphi_l)\sigma^2_{\rm GL,2}(\theta)\right], \label{eqn:avergl}\end{aligned}$$ where $\delta {\mbox{\boldmath$\theta$}}_1(\equiv \delta{\mbox{\boldmath$\theta$}}({\mbox{\boldmath$\theta$}}_1))$ and $\delta {\mbox{\boldmath$\theta$}}_2(\equiv \delta{\mbox{\boldmath$\theta$}}({\mbox{\boldmath$\theta$}}_2))$ are the angular excursions of the two bundles, and ${\left\langle{\ \ }\right\rangle}_\theta$ observationally means the average performed over all pairs with a fixed observed angular separation $\theta$. $\sigma_{\rm GL,0}(\theta)$ and $\sigma_{\rm GL,2}(\theta)$ represent isotropic and anisotropic contributions to the lensing dispersion, respectively. Although the anisotropic one is ignored in TKF for simplicity, we also take into account the contribution in this paper. It is convenient to express those dispersions in terms of the logarithmic angular power spectrum of the deflection angle, $P_{\rm GL}(l)$, as $$\begin{aligned} &&\sigma_{\rm GL,0}^2(\theta)=\frac{1}{2\pi}\int\!\!\frac{dl}{l} P_{\rm GL}(l)[1-J_0(l\theta)], \nonumber \\ &&\sigma_{\rm GL,2}^2(\theta)=\frac{1}{2\pi}\int\!\!\frac{dl}{l} P_{\rm GL}(l)J_2(l\theta), \label{eqn:gldisp}\end{aligned}$$ with $$P_{\rm GL}(l)=9H^4_0\Omega_{\rm m0}^2\int^{\chi_{\rm rec}}_0\!d\chi a^{-2}(\tau)W^2(\chi,\chi_{\rm rec}) P_\delta\left(k=\frac{l}{r(\chi)},\chi\right). \label{eqn:lensps}$$ The statistical properties of the lensing effects are thus entirely determined by $P_{\rm GL}(l)$, because the lensing field is expected to be also Gaussian at relevant angular scales ($\theta{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}}10'$). $\tau$ is a conformal time, $\chi\equiv \tau_0-\tau$, $J_n(x)$ is the Bessel function of order $n$, and the subscript $0$ and “rec” denote values at present and a recombination time, respectively. $P_\delta(k,\tau)$ is the power spectrum of matter fluctuations field, $H_0=100h$ km s$^{-1}$ Mpc$^{-1}$ and $\Omega_{\rm m0}$ denote the present Hubble parameter and the present energy density of matter, respectively. $r(\chi)$ is the corresponding comoving angular diameter distance, defined as $K^{-1/2}\sin K^{1/2}\chi$, $\chi$, $(-K)^{-1/2}\sinh (-K)^{1/2}\chi$ for $K>0, K=0, K<0$, respectively, where the curvature parameter $K$ is represented as $K=(\Omega_{\rm m0}+\Omega_{\lambda 0}-1)H_0^2$ and $\Omega_{\lambda 0}$ is the present vacuum energy density. The projection operator $W(\chi,\chi_{\rm rec})$ on the celestial sphere is given by $W(\chi,\chi_{\rm rec})=r(\chi_{\rm rec}-\chi)/r(\chi_{\rm rec}$). In the derivation of equation (\[eqn:avergl\]), we have employed two approximations. First is the flat-sky approximation where the two dimension Fourier transformation is used neglecting the curvature of the celestial sphere. This is based on the consideration that the lensing are important only on small angular scales. Second is the Born approximation that the integral can be evaluated along the unperturbed null-geodesics of CMB photon. Hu (2000) recently investigated the correction to the flat sky approach by directly evaluating the lensing effect in harmonic space. This correction to our method, however, is small as will be discussed later. The Born approximation is valid as long as the $\sigma_{{\rm GL}}(\theta)/\theta\ll 1$ is satisfied, and we numerically confirmed this on the relevant angular scales for all cosmological models considered in this paper. Importantly, magnitude of the lensing dispersion (\[eqn:avergl\]) is particularly sensitive to $\Omega_{\rm m0}$ and the normalization of matter power spectrum, which is conventionally expressed in terms of the rms mass fluctuations of a sphere of $8h^{-1}$Mpc, i.e., $\sigma_8$. The hotspots are local maxima in the two dimensional CMB sky map, and hence the distribution obeys a point process. Once the angular power spectrum of the temperature fluctuations field $C_l$ is given, the Gaussian assumption allows us to exactly predict statistical properties of the intrinsic distribution of hotspots following the methods developed by BBKS and BE. Since $C_l$ has oscillatory features such as series of Doppler peaks, the pairs of hotspots are distributed with some [*characteristic*]{} separation angles on the last scattering surface as shown by HS and TKF. This result leads to the following expectation. Let us consider all pairs of hotspots separated with the certain characteristic angular scale. Although all those pairs should be observed with the characteristic scale in the absence of the lensing, they are actually observed with various different separations in random lines of sight because of the weak lensing. The probability distribution of observed separation angles then has the lensing dispersion (\[eqn:avergl\]) at the characteristic scale. Since the effect can be measured in only a statistical sense, in this paper we focus on investigations of the lensing effect on the [*two-point*]{} correlation function of hotspots. Note that our method do not consider spurious hotspots created by the lensing, but it is a good approximation because it has been shown that an additional power of the anisotropies generated by the weak lensing is very small and important only at small scales ($\theta{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}1'$) ([@Metcalf97]; [@SZ99]; [@Zald]). These features are indeed illustrated by the numerically simulated CMB maps (see Figure \[fig:cmbmap\]) To calculate the weak lensing effect on the two-point correlation function of hotspots in the CMB maps, we first define the number density fluctuations field of hotspots as $$\delta n_{\rm pk}({\mbox{\boldmath$\theta$}})=\frac{n_{\rm pk}({\mbox{\boldmath$\theta$}}) -\bar{n}_{\rm pk}}{\bar{n}_{\rm pk}},$$ where $n_{\rm pk}({\mbox{\boldmath$\theta$}})$ and $\bar{n}_{\rm pk}$ are the number density field and the mean number density of hotspots above a certain threshold $\nu$, respectively. The threshold is conventionally expressed in units of the rms temperature fluctuations as $\nu=\Delta_{\rm pk}/\sigma_0$, where $\Delta_{\rm pk}$ is the value of temperature fluctuation at the peak position defined by $\Delta({\mbox{\boldmath$\theta$}}_{\rm pk})\equiv\delta T({\mbox{\boldmath$\theta$}}_{\rm pk})/T_{\rm CMB}$ and the dispersion is defined by $\sigma_0^2\equiv{\langle{\Delta^2({\mbox{\boldmath$\theta$}})}\rangle}$. Similarly, the other spectral parameters are defined by $\sigma_1^2\equiv{\langle{(\nabla\Delta)^2}\rangle}$ and $\sigma_2^2\equiv {\langle{(\nabla^2\Delta)^2}\rangle}$. These parameters can be expressed in terms of $C_l$ in the context of the small angular approximation (BE) as $\sigma_n^2\equiv\int(ldl/(2\pi))C_l l^{2n}$. The analytical expression of $\bar{n}_{\rm pk}(>\nu)$ has been derived by BE and is in detail presented in appendix \[app:num\]. Because of the mapping effect due to weak lensing, the lensed (observed) fluctuations field, $\delta n^{\rm GL}_{\rm pk}({\mbox{\boldmath$\theta$}})$, at a certain angular position ${\mbox{\boldmath$\theta$}}$ is the intrinsic field at another position ${\mbox{\boldmath$\theta$}}+\delta{\mbox{\boldmath$\theta$}}$ on the last scattering surface, where $\delta {\mbox{\boldmath$\theta$}}$ is the deflection angle. Thus $\delta n^{\rm GL}_{\rm pk}({\mbox{\boldmath$\theta$}})$ can be expressed as $$\delta n^{\rm GL}_{\rm pk}({\mbox{\boldmath$\theta$}})=\delta n_{\rm pk}({\mbox{\boldmath$\theta$}} +\delta{\mbox{\boldmath$\theta$}})=\int\!\!\frac{d^2{\mbox{\boldmath$l$}}}{(2\pi)^2}\delta n_{{\mbox{\boldmath$l$}}}e^{i{\mbox{\boldmath$l$}} \cdot({\mbox{\boldmath$\theta$}}+\delta{\mbox{\boldmath$\theta$}})},$$ where $\delta n_{{\mbox{\boldmath$l$}}}$ is the Fourier component of unlensed field $\delta n({\mbox{\boldmath$\theta$}})$. Therefore, since the lensing deflection angle induced by the large-scale structure and the CMB field on the last scattering surface are statistically independent, the lensed (observed) two-point correlation function of hotspots, $\xi^{\rm GL}_{\rm pk-pk}(\theta)$, can be calculated with help of equation (\[eqn:avergl\]) as $$\begin{aligned} {\xi_{\rm pk-pk}}^{\rm GL}(\theta)&=&{\left\langle{\delta n^{\rm GL}({\mbox{\boldmath$\theta$}}_1) \delta n^{\rm GL}({\mbox{\boldmath$\theta$}}_2)}\right\rangle}_{|{\mbox{\boldmath$\theta$}}_1-{\mbox{\boldmath$\theta$}}_2|=\theta} \nonumber \\ &=& \int\!\!\frac{d^2{\mbox{\boldmath$l$}}}{(2\pi)^2}\int\!\!\frac{d^2{\mbox{\boldmath$l$}}'}{(2\pi)^2} e^{i({\mbox{\boldmath$l$}}\cdot{\mbox{\boldmath$\theta$}}_1-{\mbox{\boldmath$l$}}'\cdot{\mbox{\boldmath$\theta$}}_2)} {\left\langle{\delta n_{{\mbox{\boldmath$l$}}}\delta n_{{\mbox{\boldmath$l$}}'}}\right\rangle}{\left\langle{e^{i({\mbox{\boldmath$l$}}\cdot \delta{\mbox{\boldmath$\theta$}}_1 -{\mbox{\boldmath$l$}}'\cdot\delta{\mbox{\boldmath$\theta$}}_2)}}\right\rangle}\nonumber \\ &=& \int^\infty_0 \!\frac{ldl}{2\pi}P_{\rm pk-pk}(l)\left[ \left(1-\frac{l^2}{2}\sigma^2_{\rm GL,0}(\theta)\right)J_0(l\theta) +\frac{l^2}{2}\sigma^2_{\rm GL,2}(\theta)J_2(l\theta)\right], \label{eqn:xigl0}\end{aligned}$$ where we have used the following Gaussian random property of $\delta n_{{\mbox{\boldmath$l$}}}$; $${\left\langle{\delta n_{{\mbox{\boldmath$l$}}}\delta n_{{\mbox{\boldmath$l$}}'}}\right\rangle}=(2\pi)^2P_{\rm pk-pk}(l) \delta^2({\mbox{\boldmath$l$}}-{\mbox{\boldmath$l$}}').$$ $P_{\rm pk-pk}(l)$ is the angular power spectrum of unlensed two-point correlation function of hotspots, ${\xi_{\rm pk-pk}}(\theta)$. The derivation of ${\xi_{\rm pk-pk}}(\theta)$ is presented in detail in the appendix \[app:unlensxi\]. The relation between ${\xi_{\rm pk-pk}}(\theta)$ and $P_{\rm pk-pk}(l)$ can be expressed as $$P_{\rm pk-pk}(l)=2\pi\int_0^\pi\! \theta d\theta {\xi_{\rm pk-pk}}(\theta)J_0(l\theta). \label{eqn:powerxi}$$ By using this equation, equation (\[eqn:xigl0\]) can be therefore rewritten as $${\xi_{\rm pk-pk}}^{\rm GL}(\theta)=\int\!\!d\theta'\theta'\int\!\!ldl\xi_{\rm pk-pk}(\theta') J_{0}(l\theta')\left[ \left(1-\frac{l^2}{2}\sigma^2_{\rm GL,0}(\theta)\right)J_0(l\theta) +\frac{l^2}{2}\sigma^2_{\rm GL,2}(\theta)J_2(l\theta)\right]. \label{eqn:lensxi}$$ This is the equation which we use for theoretical predictions of the lensing effect on ${\xi_{\rm pk-pk}}(\theta)$. If we ignore the anisotropic lensing dispersion $\sigma_{\rm GL,2}$, the expression (\[eqn:lensxi\]) can be further simplified (see appendix \[app:xianaly\]). Importantly, this equation indicates that the lensing contribution to ${\xi_{\rm pk-pk}}(\theta)$ at a certain scale $\theta$ arises only from the lensing dispersion $\sigma_{\rm GL}(\theta)$ at the same scale. Hence, detections of scale dependences of the lensing signatures to ${\xi_{\rm pk-pk}}(\theta)$ allow us to reconstruct the lensing dispersion at the respective scales, more interestingly, to reconstruct the projected matter power spectrum. Cosmological models =================== To make some quantitative predictions, one needs to specify cosmological models. For this reason, we adopt following adiabatic cold dark matter models with $\Omega_{\rm m0}=1$, $\Omega_{\rm \lambda 0}=0$, $h=0.5$ (hereafter SCDM) and $\Omega_{\rm m0}=0.3$, $\Omega_{\rm \lambda 0}=0.7$, $h=0.7$ (hereafter LCDM), respectively. These models are motivated by the fact that the recent high precision measurements of $C_l$ supported a flat universe under the adiabatic condition as suggested by standard inflationary scenarios ([@boom]; [@Maxima]). The baryon density is chosen to satisfy $\Omega_{\rm b0}h^2=0.019$, which is consistent with values obtained from the measurements of the primeval deuterium abundance ([@tytler]). As for the matter power spectrum, we employed the Harrison-Zel’dovich spectrum and the BBKS transfer function with the shape parameter from Sugiyama (1995). The free parameter in each model is only the normalization of the present-day matter power spectrum, i.e., $\sigma_8$. The nonlinear evolution of the power spectrum can be modeled using the fitting formula given by Peacock & Dodds (1996). To compute the angular power spectrum $C_l$, we used helpful CMBFAST code developed by Seljak & Zaldarriaga (1996). Theoretical Results =================== Dependence of lensing dispersion on cosmological parameters ----------------------------------------------------------- In Figure \[fig:dispgl\] we plot the relative lensing dispersion $\sigma_{\rm GL}(\theta)/\theta$ as a function of the separation angle $\theta$ for different sets of $\sigma_8$ in SCDM and LCDM models, which are computed using equations (\[eqn:avergl\]). The dependence of $\sigma_8$ in each model is demonstrated by choices of $\sigma_8=0.5, 1$, and $1.5$ (solid lines) in SCDM, and $\sigma_8=1, 1.5$ and $2.0$ (dashed lines) in LCDM from bottom to top, respectively. The normalizations from the [*COBE*]{} 4-year measurements (e.g. [@Bunn]) and the [*X*]{}-ray cluster abundance ([@Eke]; [@KS]) roughly correspond to $\sigma_8=1.2$ and $0.5$ for SCDM, and $\sigma_8\simeq 1-1.5$ for LCDM, respectively. Furthermore, the recent several measurements of cosmic shear (e.g. [@Ludovic]) have suggested $\sigma_8=1.5\pm 0.5$ for the current favored LCDM models, while uncertainties involved in redshift distributions of distant galaxies, the cosmic variance of the variance of the shear and the systematic errors of the signals still remain unresolved. Figure \[fig:dispgl\] clearly shows that the magnitude of $\sigma_{\rm GL}(\theta)$ at a certain angle $\theta$ has a strong dependence on the amplitude of $\sigma_8$ in each cosmological model. Since the logarithmic angular power spectrum (\[eqn:lensps\]) of $\sigma_{\rm GL}$ has contributions from the proportional factor $\Omega^2_{\rm m0}$ and the distance and the growth factors, the combination yields a dependence of $\Omega_{\rm m0}$ on $\sigma_{\rm GL}(\theta)$. Although the Hubble parameter $h$ affects $\sigma_{\rm GL}$ mainly through the shape parameter $\Gamma \approx \Omega_{\rm m0}h$ in the matter power spectrum, the dependence is weaker. The thin dot-dashed line shows a result of using the linear matter power spectrum for $\sigma_8=1.5$ in SCDM, and it reveals that the effect of the nonlinear evolution on $\sigma_{\rm GL}$ is not important at $\theta{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}}10'$, where the two-point correlations function of hotspots has most power of correlations. It is therefore expected that the measurements of $\sigma_{\rm GL}$ can generally provide constraints on $\sigma_8-\Omega_{\rm m0}$ plane. To break the degeneracy between $\Omega_{\rm m0}$ and $\sigma_8$, we have to measure the scale dependence of $\sigma_{\rm GL}(\theta)$ with respect to $\theta$. As shown by equations (\[eqn:gldisp\]) and (\[eqn:lensps\]), the lensing dispersion arises from the projected matter power spectrum $P_{\delta}(k)$ weighted with $a^{-2}W^2(\chi,\chi_{\rm rec})$. Figure \[fig:2dps\] shows the logarithmic angular power spectrum $P_{\rm GL}(l)$ defined by equation (\[eqn:lensps\]) as a function of $l$. The thin and thick curves are results of using the linear and nonlinear matter power spectra, respectively, and the figure demonstrates that the nonlinear effect is important on $l{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}}1000$, which corresponds to angular scales of $\theta{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}20'$ from the relation of $l\approx 2\pi/\theta$. The shape of $P_{\rm GL}$ peaks around $l\approx 100$ and thus the contributions to the deflection angle of [*each*]{} CMB photon come mainly from modes with such large $l$ (see Figure \[fig:cmbmap\]). The essential quantity of lensing effect on ${\xi_{\rm pk-pk}}$ is lensing fluctuations of relative separation angle, and the term including Bessel function in $\sigma_{\rm GL,0}(\theta)$ of equation (\[eqn:gldisp\]) indicates that the lensing fluctuations for two CMB bundles separated with a certain angle $\theta$ arises dominantly from the integrations of $P_{\rm GL}(l)$ over $l>2\pi/\theta$ in $l-$space. This physically means that the gravitational lensing effect on the relative angular separation is caused mainly by the projected matter fluctuations lied [*between*]{} the two bundles. Accordingly the two bundles with smaller separation are more strongly affected by the smaller scale structures of the universe. Furthermore, from these interpretations we can conclude that corrections to the flat sky approximation proposed by Hu (2000) do not affect our results at $l{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}}100$ because the corrections are important only at $l<10$. Since the more fundamental quantity is the three dimensional density fluctuations characterized by the power spectrum $P_\delta(k)$, we have to see the relation between the two power spectra, namely $P_\delta(k)$ and $P_{\rm GL}(l)$. Figure \[fig:psdkdz\](a) shows the logarithmic contribution to $P_{\rm GL}(l)$ for a given $l$-mode as a function of three dimensional wavenumber $k$ in the same models as in Figure \[fig:2dps\], where normalizations of those curves are arbitrarily scaled. These functions have relatively broad shape and peaks at $\lambda=k/2\pi\approx 630 h^{-1}{\rm Mpc}$ for $l=10$, at $k\approx 125h^{-1}{\rm Mpc}$ for $l=100$, and $k\approx 21h^{-1}{\rm Mpc}$ for $l=1000$ in the considered SCDM model, respectively. These relations between $l$ and $k$ depend on the shape of matter power spectrum. On the other hand, since in the LCDM model the comoving angular diameter distance at a certain redshift is larger than the corresponding distance in SCDM, the peak-wavelengths are smaller than those in the SCDM cases for each $l$ mode. The figure also demonstrates that the higher $l$ mode is affected more strongly by the matter fluctuations with smaller wavelengths. The next question is the redshift distribution of the contribution to a given $P_{\rm GL}(l)$. Figure \[fig:psdkdz\](b) shows the logarithmic contribution to $P_{\rm GL}(l)$ for each $l$ mode as a function of $(1+z)$. Since the window function $W/a$ in $P_{\rm GL}(l)$ is a monotonous decreasing function with respect to $z$ and has no characteristic redshift, the question which redshift structures give the dominant contribution depends on the shape of matter power spectrum. The figure clearly demonstrates that each curve peaks at a certain redshift; for low $l$ mode the contribution is dominated by the low-$z$ structures and, on the other hand, high $l$ mode has wide range contributions in the redshift space. These results can be explained as follows. As shown in Figure \[fig:psdkdz\](a), there is the peak scale $k_{\rm max}$ of three dimensional mass fluctuations that provides most dominant contribution to $P_{\rm GL}(l)$ for each $l$ mode. Because of the projection effect on the celestial sphere, the contribution from the $k_{\rm max}$ mode comes from structures at a specific redshift $z_{\rm max}$ which satisfies the relation of $k_{\rm max} \approx l/r(z_{\rm max})$. As a result, for lower $l$ mode the contribution will be dominated by lower $z$ structures. On the other hand, for $l=1000$ mode the contribution peaks at $z\approx 0.5$ with a long tail expanding to higher $z$, where the tail is caused by the fact that the window function $W/a$ has larger power at lower $z$. In fact, Figure \[fig:psdkdz\](b) confirms these interpretations. We could therefore directly probe dark matter clustering at low and high redshifts for low and high $l$ modes, respectively, in principle by using the weak lensing signatures. Unlensed two-point correlation function of hotspots with the instrumental effects --------------------------------------------------------------------------------- The unlensed two-point correlation functions of local maxima ([*hotspots*]{}), say $\xi_{\rm pk-pk}(\theta)$, in the CMB maps can be accurately predicted based on the Gaussian random theory. The derivation is presented in the appendix \[app:unlensxi\] in detail. To show the shape of ${\xi_{\rm pk-pk}}(\theta)$, we performed a six dimensional numerical integration using equation (\[eqn:unlensxi\]) because the two of the eight dimensional integration in equation (\[eqn:unlensxi\]) can be analytically done. For the practical purpose, we also take into account instrumental effects of finite beam size and detector noise. The beam smearing effect on the temperature fluctuations field can be modeled by the Gaussian beam approximation characterized by a filter function $F_l=\exp[-l^2 \theta_s^2/2]$ in $l$-space, where the smoothing angle $\theta_s$ is expressed in terms of the full-width at half-maximum angle $\theta_{\rm fwhm}$ of a telescope as $\theta_s=\theta_{\rm fwhm}/\sqrt{8\ln 2}$. The noise level of detectors is conventionally expressed in terms of the temperature fluctuations per a pixel on a side of FWHM extent as $\Delta_{\rm noise}=\sigma_{\rm sens}$. If we assume that the primordial temperature and the noise fields are uncorrelated, by modifying the angular power spectrum $C_l$ to $\tilde{C}_l=(C_l+\sigma_{\rm sens}^2\theta^2_{\rm fwhm})\exp[-l^2\theta_s^2]$, these instrumental effects can be approximately included into the theoretical predictions (HS), because in the Gaussian random theory $C_l$ contains complete information about statistical properties of any intrinsic CMB field. The numerical experiments indeed show that this treatment works well. Figure \[fig:xipk\] shows the unlensed ${\xi_{\rm pk-pk}}(\theta)$ with and without the experimental effects as a function of separation angle $\theta$ in the SCDM and LCDM models, where we considered the two hotspots of height above the threshold $\nu=1(\Delta_{\rm pk}=\sigma_0)$. As for the instrumental effects, we have employed the specifications of Planck $217{\rm GHz}$ channel and MAP $90{\rm GHz}$ channel; the beam size and noise level are assumed to be $\theta_{\rm fwhm}=5.5'$ and $\sigma_{\rm sens}=4.3\times 10^{-6}$ for Planck while $\theta_{\rm fwhm}=12.6'$ and $\sigma_{\rm sens}=2.5\times 10^{-5}$ for MAP. The solid lines in each panel show ${\xi_{\rm pk-pk}}(\theta)$ including both effects of the beam size and noise, and the dotted lines in the upper left and bottom left panels show an ideal case without those effects. The ideal cases demonstrate that the intrinsic ${\xi_{\rm pk-pk}}(\theta)$ has a prominent peak at $\theta\approx 10'$ and the damping tail at $\theta<10'$. These features physically mean that the primary temperature field has the characteristic curvature scale of the order of $10'$, which can be estimated by $\theta_\ast\equiv \sqrt{2}\sigma_1/\sigma_2 \sim 5'$ in Appendix \[app:num\], and has smooth structures at scales of $\theta<\theta_\ast$ as actually shown by the simulated maps. The beam smearing on the intrinsic ${\xi_{\rm pk-pk}}(\theta)$ then appears as a cutoff at scales below the beam size although the effect moderately changes the global shape of ${\xi_{\rm pk-pk}}(\theta)$ in the MAP case (top right and bottom right panels). This is because the beam smearing causes an incorporation of intrinsic hotspots contained within one beam. To clarify the noise effect explicitly, we also show ${\xi_{\rm pk-pk}}(\theta)$ only with beam smearing effect (dot-dashed lines). The curves explain that spurious hotspots due to the detector noises generate an extra power of correlations on ${\xi_{\rm pk-pk}}(\theta)$. In particular, the predicted shape of ${\xi_{\rm pk-pk}}(\theta)$ for MAP is largely affected by the noise effect up to $\theta=80'$ while the Planck cases have slight changes only at $\theta{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}20'$. By using equation (\[eqn:meannum\]), we can predict values of mean number density of hotspots above $\nu=1$ for those cases. The values with both the beam smearing and noise effects and only with beam effect in SCDM model are $\bar{n}_{\rm pk}=8.17\times 10^3$ and $7.74\times 10^3[{\rm rad}^{-2}]$ for Planck (upper left), respectively, while $\bar{n}_{\rm pk}=5.31\times10^3$ and $3.01\times10^3[{\rm rad}^{-2}]$ for MAP (upper right), respectively. Similarly, in LCDM model $\bar{n}_{\rm pk}=7.73\times 10^3$ and $7.33\times 10^3[{\rm rad}^{-2}]$ for Planck (bottom left) and $\bar{n}_{\rm pk}=5.06\times 10^2$ and $2.88\times 10^3 [{\rm rad}^{-2}]$ for MAP (bottom right), respectively. These values actually coincide with results of the number counts of hotspots in the simulated CMB maps within the Poisson error. Moreover, to reveal the differences of shapes between ${\xi_{\rm pk-pk}}(\theta)$ and the conventionally used two-point correlation function of the temperature fluctuations field itself defined by $C(\theta)\equiv {\left\langle{\Delta({\mbox{\boldmath$\theta$}}_1) \Delta({\mbox{\boldmath$\theta$}}_2)}\right\rangle}_{|{\mbox{\boldmath$\theta$}}_1-{\mbox{\boldmath$\theta$}}_2|=\theta}$, the dashed line in upper left panel shows $C(\theta)$ normalized to agree with the value of solid line at $\theta=200'$. It is clear that ${\xi_{\rm pk-pk}}(\theta)$ has much more oscillatory features than $C(\theta)$ does (HS; TKF). This reason is as follows. In the peak statistics we need statistical properties of the gradient and second derivative fields of the temperature fluctuations field in order to identify the local maxima (or minima) in the CMB maps. The power spectra of those fields per logarithmic interval in $l$ are then $l^4 C_l/(2\pi)$ and $l^6 C_l/(2\pi)$, respectively, while the power spectrum of the temperature fluctuations filed is $l^2C_l/(2\pi)$ whose integral over $l$ space produces $C(\theta)$. Figure \[fig:Clln\] shows the power spectra of $\Delta_{,i}$ and $\Delta_{,ij}$ as a function of $1/l$ $(\propto \theta)$, and clearly explains that they strongly enhance the oscillations of $l^2C_l$. If recalling the relation of $\theta\approx 2\pi/l$, there are indeed correspondences between both oscillations of ${\xi_{\rm pk-pk}}(\theta)$ and $C_l l^2/(2\pi)$; the depression at $\theta\approx 70'$ corresponding to the scales around the first Doppler peak, a prominent peak at $\theta \approx 13'$, and a damping tail at $\theta<10'$ associated with the Silk damping in $C_l$. Most importantly, these oscillatory features of ${\xi_{\rm pk-pk}}(\theta)$ physically mean that pairs of hotspots are discretely distributed with some characteristic separation angles on the last scattering surface. The peak statistics thus produces an additional information that could be convenient for the study of the weak lensing although it relys on only a subset of available information (the location of peaks, peak height and their profile). Furthermore, we expect that the distribution of hotspots are more robust against the systematic observational errors. The previous works already concluded that the lensing effect on $C_l$ is small at $l<3000$ (e.g. see Seljak 1996 and references therein), and we therefore expect that the observed $C_l$ will provide the accurate prediction of unlensed ${\xi_{\rm pk-pk}}(\theta)$. The lensing signatures to ${\xi_{\rm pk-pk}}(\theta)$ are then extracted as non-Gaussian signatures, which are differences between the observed (lensed) and the predicted two-point correlation functions. Lensed two-point correlation function of hotspots ------------------------------------------------- We present the theoretical predictions of the lensed two-point correlation function of hotspots. By using equation (\[eqn:lensxi\]) we then compute the lensing effect on the intrinsic ${\xi_{\rm pk-pk}}(\theta)$ into which the experimental effects are already included by the method presented in the previous subsection. Although the order of computing those secondary effects is not correct, the treatment makes the calculations much simpler and is a good approximation because of the following reasons. As explained, the instrumental effects on ${\xi_{\rm pk-pk}}(\theta)$ can be approximately included only by modifying the angular power spectrum $C_l$ based on the Gaussian theory and actually the predictions are in good agreement with the numerical experiments of ${\xi_{\rm pk-pk}}(\theta)$ (HS and also see Figure \[fig:obsxi\]). This means that the distribution of hotspots as a point process can be accurately predicted even in the CMB maps altered by the beam smearing and the detector noise. Then, since we are here interested in a problem how the weak lensing redistributes the ‘key’ hotspots, which can survive after the beam smearing effect, in an actual observed map, we can approximately consider the lensing effect on the ${\xi_{\rm pk-pk}}(\theta)$ after the instrumental effects as long as the weak lensing does not create a lot of spurious hotspots on the map. Moreover, we can at least say that the important lensing signature to ${\xi_{\rm pk-pk}}$ on large angular scales such as $\theta\approx70'$ is not directly coupled to the beam smearing effect of $\theta_{\rm fwhm}{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}10'$. The validity of our treatment has been confirmed by the numerical experiments on the relevant angular scales. Figure \[fig:lensxipk\] shows both unlensed and lensed two-point correlation functions of hotspots, say ${\xi_{\rm pk-pk}}^{\rm GL}(\theta)$ and ${\xi_{\rm pk-pk}}(\theta)$, where the threshold $\nu=1$ is similarly assumed and we employ the same models of matter power spectrum as in Figure \[fig:dispgl\]. Although $\nu=1$ is assumed throughout this paper for simplicity, the two-point correlation function of hotspots above height of an arbitrary threshold can be predicted under the Gaussian theory and we could use this freedom for measuring the lensing effect as will be discussed later. The upper left and bottom left panels for Planck case clearly demonstrate that the weak lensing effect fairly smooths out the shape of unlensed ${\xi_{\rm pk-pk}}(\theta)$ (solid line). The magnitude of the lensing effect is strongly sensitive to the amplitude of $\sigma_8$ in each cosmological models and larger at angular scales where ${\xi_{\rm pk-pk}}(\theta)$ has more oscillatory features (TKF). In particular, we stress that the weak lensing causes a distinct smoothing effect on the depression feature of ${\xi_{\rm pk-pk}}(\theta)$ at scales around $\theta\approx 70'$. We therefore expect that the observed depth of depression relative to the plateau shape at larger or smaller scales than the scale can be a robust indicator of the weak lensing signatures and depends only on the magnitude of $\sigma_{\rm GL}$ if it is large sufficiently to detect. However, since in the MAP cases (upper right and bottom right panels) the detector noise effect decreases largely the depth of the intrinsic depression as shown in Figure \[fig:xipk\], the lensing effect is hidden at the scale for all models. If using the relation between $k$ and $l$ shown in Figure \[fig:psdkdz\](a), the measure of $\sigma_{\rm GL}(\theta)$ from the lensing signature at $\theta\approx 70'$ can provide a constraint on the amplitude of mass fluctuations with large wavelength modes such as $\lambda\approx 50h^{-1}{\rm Mpc}$. On the other hand, the measurement of $\sigma_{\rm GL}$ at the prominent peak scale such as $\theta\approx 20'$ is sensitive to the smaller scale structures such as $\lambda\approx 10h^{-1}{\rm Mpc}$, while the shape of ${\xi_{\rm pk-pk}}$ is also sensitive to the experimental effects at such scales. Detectability of the lensing signatures ======================================= In this section, by using numerical experiments we quantitatively investigate how accurately non-Gaussian signatures on ${\xi_{\rm pk-pk}}(\theta)$ induced by the weak lensing can be detected with the expected future MAP and Planck surveys. Numerical experiments --------------------- We perform simulations of the CMB maps with and without the lensing effect by using the following procedures. A realistic unlensed temperature maps on a fixed square grid can be generated from a given power spectrum, $C_l$, based on the Gaussian assumption. Each simulated map is initially on $40\times40$ square degree area with $4096\times4096$ pixels. After regridding to take into account the beam smearing effect, the actual pixel number is reduced to $1200\times1200$ and $480\times480$ for the Planck and MAP cases, respectively. To compute the lensing effect on the CMB maps, we first need the convergence field on each grid. We then employ the following angular power spectrum of convergence field ([@Blandford]; [@miralda]; [@Kaiser92]); $$P_\kappa(l)=\frac{9}{4}H_0^4\Omega_{\rm m0}^2 \int^{\chi_{\rm rec}}_0\!\!d\chi a^{-2}W^2(\chi,\chi_{\rm rec})P_{\delta} \left(k=\frac{l}{r(\chi)},\chi\right), \label{eqn:pskappa}$$ where the convergence field $\kappa$ is expressed in terms of the radial integral of three dimensional density fluctuations field $\delta$ as $\kappa =(3/2)H_0^2\Omega_{\rm m0}\int\!d\chi W(\chi,\chi_{\rm rec}) r(\chi)a^{-1}\delta$. Note that the second moment of convergence field is then ${\left\langle{\kappa^2({\mbox{\boldmath$\theta$}})}\right\rangle}=\int(ldl/2\pi)P_\kappa(l)$. Using the power spectrum $P_\kappa(l)$, the convergence field can be generated as a realization of a Gaussian process. To compute the displacement vector $\delta{\mbox{\boldmath$\theta$}}$, we transform the convergence field in the Fourier space, and compute the Fourier component of the displacement vector, $\delta\tilde{{\mbox{\boldmath$\theta$}}}({\mbox{\boldmath$l$}})$, by using the relation of $$\delta\tilde{{\mbox{\boldmath$\theta$}}}({\mbox{\boldmath$l$}})=2i \frac{{\mbox{\boldmath$l$}}}{l^2}\tilde{\kappa}({\mbox{\boldmath$l$}}),$$ where we have used the relation of $2\kappa({\mbox{\boldmath$\theta$}})=\partial(\delta \theta_1)/\partial \theta_1+\partial(\delta\theta_2)/\partial \theta_2$ derived by the cosmological lens equation. If transforming it back to real space, for each point on the observed temperature map we can obtain the corresponding displacement vector to map the point on a irregular grid of the primary temperature map on the last scattering surface. The lensed temperature map can be then obtained by using cloud-in-cell interpolation to compute the value on the original regular grid of observed map. In the case of taking into account the instrumental effects of beam smearing and noise, we furthermore smooth out the temperature map by convolving the Gaussian window function ${\cal F}({\mbox{\boldmath$\theta$}},{\mbox{\boldmath$\theta$}}') =1/(2\pi\theta_s^2)\exp[-|{\mbox{\boldmath$\theta$}}-{\mbox{\boldmath$\theta$}}'|^2/(2\theta_s^2)]$ and then add randomly the noise field into each pixel. Figure \[fig:cmbmap\] illustrates an example of realizations of the simulated unlensed and lensed maps for the SCDM model with $\sigma_8=1.5$, which provides the largest lensing signatures of our considered models. Interestingly, the figure shows that the displacement of each position of hotspots in the lensed map from the unlensed map is relatively large even though the global patterns of temperature fluctuations in both maps are not considerably changed. This is the consequence of the large scale modes of the deflection angle, which play an important role to the lensing effect on ${\xi_{\rm pk-pk}}(\theta)$. The signal to noise ratio of the lensing signatures --------------------------------------------------- The observational errors associated with measurements of ${\xi_{\rm pk-pk}}(\theta)$ arise from the cosmic variance and the instrumental resolutions. To accurately compute the cosmic variance, we have used $100$ independent realizations of both unlensed and lensed temperature maps, respectively, with $40\times 40$ square degree area. In this paper we assume $65\%$ sky coverage for MAP and Planck surveys, and this corresponds to the assumption that we can use about $17$ independent simulated maps for those surveys in order to obtain the averaged ${\xi_{\rm pk-pk}}(\theta)$. In Figure \[fig:obsxi\] we show an example of the averaged [*lensed*]{} two-point correlation functions of hotspots computed from one set of 17 realizations for the expected Planck runs, where the error-bars in each bin can be estimated by rescaling the variance of the estimates obtained from the $100$ realizations by a factor of $\sqrt{17}$ and the resolution of bins in $\theta$ is limited by the pixel size of simulated CMB maps. The unlensed cases have been already presented by HS. The bottom panels also show the results around the depression scale $\theta\approx70'$ of ${\xi_{\rm pk-pk}}(\theta)$ in each cosmological model. Figure \[fig:obsxi\] clearly shows that the theoretical predictions are in remarkable good agreement with the numerical results. In particular, the lensing signatures at the depression scale definitely deviate from the unlensed cases, and we therefore expect that the non-Gaussian signatures should be detected by Planck with high significance for some adequate values of $\sigma_8$. On the other hand, for the lensing signatures at smaller scales such as the prominent scale $\theta\approx20'$, it seems to be slightly difficult to distinguish them because the sampling variance are larger at smaller scales and the shapes of ${\xi_{\rm pk-pk}}(\theta)$ is also sensitive to the instrumental effects of beam smearing and noise as shown in Figure \[fig:xipk\]. On the other hand, we concluded that it is difficult for MAP survey to distinguish the lensing signatures on ${\xi_{\rm pk-pk}}(\theta)$ for all cosmological models considered in this paper. For this reason, we present only results from Planck in the following. We so far have adopted the specific shape of matter power spectrum in each considered cosmological model, and therefore a free parameter of our model is only $\sigma_8$. While the actual observable quantity of our method is the lensing dispersion $\sigma_{\rm GL}(\theta)$, we here investigate the dependence of the lensing signatures to ${\xi_{\rm pk-pk}}(\theta)$ on $\sigma_8$ for simplicity and the problem will be discussed in the final section. In Table \[tab:1\] we summarize the results obtained for the determination of $\sigma_8$ with a best fit and the error associated with this determination. The error then has been determined as the variance of the best fit values of $\sigma_8$ among a lot of sets of the numerical experiments performed by fitting the simulated results to the theoretical templates of lensed ${\xi_{\rm pk-pk}}(\theta)$ so that the $\chi$-square is minimum. Each best fit of theoretical curves to the numerical experiments is restricted mainly from the data at $\theta>20'$, in particular the data around the depression scale of ${\xi_{\rm pk-pk}}(\theta)$. One can see that the signal to noise ratio indeed grows with $\sigma_8$ in each cosmological model. Furthermore, if comparing results obtained from SCDM and LCDM models with the same value of $\sigma_8$ ($\sigma_8=1.5$), it is clear that more robust constraints can be provided in SCDM, more generally for background models with larger $\Omega_{\rm m0}$, as expected. The noise level of Planck is independent on the $\sigma_8$ estimations, but we have confirmed that the beam size is rather important for the detections. Importantly, the accuracies of $\sigma_8$ determinations expected from Planck reach about $8\%$ and $11\%$ for models with $\sigma_8=1.5$ in SCDM and LCDM, respectively. Furthermore, in the Gaussian random theory the two-point correlation function of local minima ([*coldspots*]{}) should have the same shape as that of hotspots, and therefore combing the measurements of coldspots correlation function improves the lensing signals by about factor $\sqrt{2}$. We could also improve the significance by combining the measurements of ${\xi_{\rm pk-pk}}$ with another different thresholds, but the independence of data then has to be carefully investigated because the angular positions of same hotspots are used many times in the fitting. input values of $\sigma_8$ SCDM $(\Omega_{\rm m0}=1,h=0.5)$ LCDM $(\Omega_{\rm m0}=0.3, \Omega_{\rm \lambda 0}=0.7, h=0.7)$ ---------------------------- ---------------------------------- ----------------------------------------------------------------- 0 $0.30\pm0.31$ $0.23\pm 0.35$ 0.5 $0.47\pm0.24$ - 1.0 $1.03\pm0.13$ $0.93\pm0.23$ 1.5 $1.52\pm0.12$ $1.49\pm0.17$ 1.5 (without noise) $1.47\pm0.12$ $1.53\pm0.16$ 2.0 - $1.97\pm0.16$ : Summary of best fit of $\sigma_8$ determinations from the lensing signatures to ${\xi_{\rm pk-pk}}(\theta)$ using the numerical experiments for the expected Planck survey in SCDM and LCDM models. We have assumed $65\%$ sky coverage, and the $1\sigma$ error in each determination represents uncertanities caused by the cosmic variance and the instrumental effects due to the beam smearing effect and the detector noise (see text). \[tab:1\] Our arguments presented in this section rely on the expectation that we can accurately predict the shape of unlensed ${\xi_{\rm pk-pk}}(\theta)$ as a function of sets of cosmological parameters constrained from the measured angular power spectrum $C_l$ within the limit of the observational errors. However, we should bear in mind the fact that only the lensed $C_l$ is measurable. Then, one may imagine an approach to compare the measured ${\xi_{\rm pk-pk}}$ to the [*fake*]{} prediction, say $\tilde{\xi}_{\rm pk-pk}$, computed directly from the lensed (measured) $C_l$ based on the Gaussian theory. Since the lensed temperature fluctuations field weakly deviates from the Gaussian ([@Bern97]), the distribution of hotspots in the lensed maps can be no longer characterized only by the lensed $C_l$. However, it will be still interesting to see differences between the exact and fake predictions of the lensing effect on ${\xi_{\rm pk-pk}}(\theta)$. Figure \[fig:fakexi\] thus shows the shape of $\tilde{\xi}_{\rm pk-pk}$ and reveals that $\tilde{\xi}_{\rm pk-pk}$ overestimates the power of correlations on scales of $\theta<40'$, while $\tilde{\xi}_{\rm pk-pk}$ mimics the smoothing effect on the depression feature of ${\xi_{\rm pk-pk}}$ resulting from the fact that the weak lnsing already causes the smoothing of $C_l$. Consequently, for example, the value of reduced $\chi$-square for the fitting between $\tilde{\xi}_{\rm pk-pk}$ and the simulated ${\xi_{\rm pk-pk}}$ becomes worse to be $\chi^2\approx 162/95$ against the fact that the fitting of using the exact theoretical predictions (\[eqn:lensxi\]) of lensed ${\xi_{\rm pk-pk}}$ produces values around almost unity. Hence, the important problem how we extract the lensing information only from the measurable CMB quantities has to be carefully investigated and this problem will be again discussed in the next section. Discussion and Conclusions {#summary} ========================== In this paper, we have quantitatively shown that the peak-peak correlation function for the CMB maps can be an efficient measure to probe the mass fluctuations at relatively large scales. The non-Gaussian signatures on ${\xi_{\rm pk-pk}}(\theta)$ caused by the weak lensing appear at several scales as smoothing effects on the oscillatory features of ${\xi_{\rm pk-pk}}(\theta)$ (TKF). In particular, by using numerical experiments of the CMB maps including effects of beam smoothing and detector noise as well as the lensing displacements, we found that the lensing signatures at about [*one degree*]{} scales, where the intrinsic ${\xi_{\rm pk-pk}}(\theta)$ has the pronounced depression feature, should be detected most significantly from the expected Planck survey if the signatures are adequately large for the detections. On the other hand, unfortunately we concluded that it is difficult for MAP to detect the lensing signatures for the cosmological models considered in this paper. The direct observable quantity of our method is the dispersion of lensing deflection angle as shown by equation (\[eqn:lensxi\]). We have then revealed that the one-degree scale dispersions are sensitive to amplitudes of three dimensional mass fluctuations around wavelength $\lambda\approx 50h^{-1}{\rm Mpc}$ because of the projection effect. Furthermore, such lensing dispersions have contributions from structures of the universe with a wide redshift distribution in the range of $1{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}z{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}3$ (see Figure \[fig:psdkdz\]). This projection effect would be thus a serious issue for extracting the cosmological information from weak lensing contributions in a general case. However, since the mass fluctuations around $\lambda\sim 50h^{-1}{\rm Mpc}$ are now still in the linear regime and the evolution history in the redshift space is theoretically well understood in the context of gravitational instability in an expanding universe, the lensing signatures can accurately determine the cosmological parameters associated with amplitude and evolution of mass fluctuations at the scales. It is therefore expected that our method can provide robust constraints on cosmological parameters $\sigma_8$ and $\Omega_{\rm m0}$ without much specifying the shape of matter power spectrum. Our numerical experiments indeed revealed that significant signal to noise rations for determinations of $\sigma_8$ from the lensing signatures to ${\xi_{\rm pk-pk}}(\theta)$ are obtained for some input values of $\sigma_8$ in SCDM and LCDM models, respectively. It was also shown that for the same value of $\sigma_8$ more significant signal can be obtained in SCDM than in LCDM. To break the degeneracy in $\sigma_8-\Omega_{\rm m0}$ determinations only by using our method, one must measure the scale dependence of lensing signatures. This seems to be difficult even for Planck survey, because the shape of ${\xi_{\rm pk-pk}}(\theta)$ at small scales of $\theta\approx20'$ where ${\xi_{\rm pk-pk}}$ has the secondly significant lensing signatures is also sensitive to the instrumental effects of beam smoothing and noise. Anyway, it is very interesting that our method could provide constraints on the mass fluctuations in the linear regime independently of those provided by the survey of galaxies clustering and the primary CMB anisotropies alone, which cannot directly probe the power spectrum of dark matter. In the results presented in this paper, we have discussed how accurately the non-Gaussian signatures of the lensed ${\xi_{\rm pk-pk}}^{\rm GL}$ can be detected as a deviation from the unlensed ${\xi_{\rm pk-pk}}$ as shown in Figure \[fig:obsxi\]. This strategy is based on the assumption that we can accurately predict the unlensed ${\xi_{\rm pk-pk}}(\theta)$ from the measured $C_l$ because the lensing effect on $C_l$ is small ([@Seljak96]). However, since an actual measurable quantity is only the lensed $C_l$, there remains an important problem that we have to carefully investigate. Usually, the measured angular power spectrum $C_l$ can be used to constrain the sets of cosmological parameters under a specific scenario, for example, within the framework of inflationary-motivated models (e.g. [@lange]). The detailed analysis of using higher $l$ modes such as $l>3000$ will need to take into account the lensing contributions to $C_l$ for the accurate determinations ([@Stompor]). Therefore, one approach toward detecting the lensing effect on ${\xi_{\rm pk-pk}}$ is that we first construct a lot of templates of the unlensed ${\xi_{\rm pk-pk}}$ as a function of sets of cosmological parameters constrained by the measured $C_l$ within the observational errors, and then we compare the templates with the directly measured ${\xi_{\rm pk-pk}}$ in the observed (lensed) CMB maps by taking into account the lensing contributions. Based on the considerations, we are now investigating the detailed dependence of cosmological parameters on the shape of ${\xi_{\rm pk-pk}}$. As a result, we confirm to a extent that, if we focus on the depth of depression feature of ${\xi_{\rm pk-pk}}(\theta)$ at one degree scales relative to the plateau shapes at larger or smaller scales than that scale, the measured depth could be a robust indicator relatively independently of the cosmological parameters because the weak lensing with same magnitude of $\sigma_{\rm GL}$ can shallow the measured depth to a same amount. This should be further investigated carefully and will be presented elsewhere. The important thing that we stress in this paper is that we quantitatively proposed a new statistical method for measuring the lensing effect based on the peak statistics. It is not evident that our method and the other methods can measure the small lensing signatures from the observed CMB data with exactly [*same*]{} statistical significance. Therefore, several independent methods should be performed to measure the lensing signatures complementarily. Undoubtedly, we have to carefully consider how secondary anisotropies and the foreground sources affect our conclusions. The most important source of secondary anisotropies is the (thermal) Sunyaev-Zel’dovich (SZ) effect, which is essentially caused by hot electrons inside the clusters of galaxies. Since SZ induces redistributions of the photon energy from low frequency to high frequency part of the black body spectrum, the effect could mimic peaks in the observed temperature map. However, the SZ anisotropies are dominated by the Poisson contribution from the individual clusters at relevant angular scales ([@KK]), and therefore the contribution of the cross correlation between spurious peaks due to SZ and the intrinsic peaks will be smaller than the amplitude of the intrinsic peak-peak correlation because the peaks on the last scattering surface and the SZ clusters are statistically uncorrelated. Furthermore, we should emphasize that the SZ effect can be always removed by either observing at $217$ GHz or by taking advantage of its specific spectral properties. The other secondary anisotropies such as Rees-Sciama effect do not affect our method because the amplitudes of those effects have a very small contribution at $l<3000$ ([@SeljakRees]). Finally, with regard to the effect of the extragalactic sources, since it is expected that the amplitude of anisotropies due to the discrete sources in the $100-200$ GHz range are well bellow the amplitude of primordial fluctuations ([@Toffolatti]) and the sources will be also eliminated by using the multi frequency observations in principle, it can be safely concluded that this is not a serious for our method. Recent measurements of cosmic shear (e.g. see [@Ludovic]) have provided constraints on the combination of $\sigma_8$ and $\Omega_{\rm m0}$. The cosmic shear can probe the mass fluctuations at angular scales of $1'{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}\theta {\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}30'$ because of the limited survey volume, where the nonlinear clustering effect of dark matter could play an important role. It has been also shown that the deformation effects on the CMB maps could be measured by Planck ([@Bern98]; [@WBB]), and it is sensitive to the projected mass fluctuations at scales around $5'$ which is the characteristic curvature scale of the primary temperature field. Therefore, although $\Omega_{\rm m0}$ parameter cannot be determined accurately form the primary CMB anisotropies alone even with Planck because of the so-called cosmic (geometrical) degeneracy ([@BET]), it is expected that our method and those other independent methods of using the weak lensing will break the cosmic degeneracy with high precision because the amplitude of lensing contributions is also sensitive to $\Omega_{m0}$ as explained. Another challenging possibility is that those independent methods could allow us to observationally reconstruct the shape of power spectrum of dark matter including the evolution history in the redshift space by combining those measurements of amplitudes of mass fluctuations at respective angular scales. Acknowledgments {#acknowledgments .unnumbered} =============== We thank E. Komatsu for careful reading of the manuscript, frequent discussions and critical comments. We also thank an anonymous referee for useful comments which have considerably improved this manuscript. We are also grateful to U. Seljak and M. Zaldarriaga for their available CMBFAST code. M.T. thanks to the Japan Society for Promotion of Science (JSPS) Research Fellowships for Young Scientist. Mean density of hotspots in the 2D CMB map {#app:num} ========================================== In this appendix, we briefly review the derivation of the mean number density of peaks in the two dimensional Gaussian field ([@BBKS; @BE]). Following BBKS and BE, we introduce the spectral parameters defined by $$\gamma\equiv\frac{\sigma_1^2}{\sigma_0\sigma_2},\hspace{4em} \theta_\ast=\sqrt{2}\frac{\sigma_1}{\sigma_2},$$ where $$\sigma_n^2\equiv\int\!\!\frac{ldl}{2\pi}C_l l^{2n}.$$ The characteristic curvature scale of primary temperature field can be estimated as $\theta_\ast\sim5'$ for the cosmological models considered in this paper. The problem of the two dimensional peak statistics is to consider the statistical properties of the point process. We can then express the point process entirely in terms of the field and its derivatives. At hotspots (coldspots) the gradient $\Delta_i(\equiv\partial \Delta/\partial \theta^i)$ vanishes, and the eigenvalues of curvature matrix $\Delta_{ij}(\equiv \partial^2\Delta/(\partial \theta^i\partial\theta^j))$ are all negative (positive). We therefore need to consider six independent variables ${\mbox{\boldmath$v$}}=(\Delta,\Delta_x,\Delta_y,\Delta_{xx},\Delta_{yy} ,\Delta_{xy})$ to specify [*one*]{} local maximum. For the Gaussian field, the probability density function (PDF) for ${\mbox{\boldmath$v$}}$ is $$p_1({\mbox{\boldmath$v$}})=\frac{1}{(2\pi)^3|{\rm det}(M_{ij})|^{1/2}}\exp\left( -\frac{1}{2}v_iM^{-1}_{ij}v_j\right),$$ where the covariance matrix is $M_{ij}={\left\langle{v_iv_j}\right\rangle}$ because of ${\left\langle{v_i}\right\rangle}=0$ in the present case, and $M^{-1}_{ij}$ is the inverse matrix of $M_{ij}$. Following HS, we introduce the notations for convenience as $$\begin{aligned} &&\nu\equiv \frac{\Delta}{\sigma_0},\hspace{2em}\eta_i\equiv\frac{\Delta_i} {\sigma_1} \ (i=x,y) \nonumber \\ && X\equiv-\frac{\Delta_{xx} +\Delta_{yy}}{\sigma_2},\hspace{2em}Y\equiv\frac{\Delta_{xx}-\Delta_{yy}} {\sigma_2},\hspace{2em}Z\equiv\frac{2\Delta_{xy}}{\sigma_2}. \label{eqn:gaussvari}\end{aligned}$$ The non-zero second moments of these variables are $$\begin{aligned} &&{\left\langle{\nu^2}\right\rangle}={\langle{X^2}\rangle}=2{\langle{Y^2}\rangle}=2{\langle{Z^2}\rangle} =2{\left\langle{\eta_i^2}\right\rangle}=1,\hspace{1em} {\langle{\nu X}\rangle}=\gamma,\end{aligned}$$ where $\gamma\equiv \sigma_1^2/(\sigma_0\sigma_2)$. By using these equations, we can obtain the following PDF for the variables ${\mbox{\boldmath$v$}}=(\nu,X,Y,Z,\eta_i)$ with the simple form, and it gives a probability that the field point has values in the ranges of $\nu$ to $\nu+d\nu$, $X$ to $X$ to $X+dX$ and so on: $$\begin{aligned} p_1(\nu,X,Y,Z,{\mbox{\boldmath$\eta$}}_i)d\nu dXdYdZd^2\eta_i &=&\frac{2^2}{(2\pi)^3\sqrt{1-\gamma^2}}\exp\left[-Q\right] d\nu dXdYdZd^2\eta_i \label{eqn:1ppdf}\end{aligned}$$ with $$2Q=\frac{(\nu-\gamma X)^2}{1-\gamma^2}+X^2+2Y^2+2Z^2+2{\mbox{\boldmath$\eta$}}^2.$$ The point process of hotspots (or coldspots) can be described by the number density ‘operator’ $$n_{\rm pk}({\mbox{\boldmath$\theta$}})=\sum_p\delta^{D}({\mbox{\boldmath$\theta$}}-{\mbox{\boldmath$\theta$}}_{{\rm pk},p}), \label{eqn:sumpk}$$ where $\delta^{D}(x)$ is the Dirac delta function. In the neighborhood of a hotspot point ${\mbox{\boldmath$\theta$}}_{\rm pk}$ we can expand the field $\Delta({\mbox{\boldmath$\theta$}})$ in a Taylor series: $$\Delta({\mbox{\boldmath$\theta$}})\approx\Delta({\mbox{\boldmath$\theta$}}_{\rm pk})+\frac{1}{2}\Delta_{ij} ({\mbox{\boldmath$\theta$}}_{\rm pk}) (\theta-\theta_{\rm pk})_i(\theta-\theta_{\rm pk})_j.$$ Using this equation, therefore, the number density field can be expressed in terms of the Gaussian variables: $$n_{\rm pk}({\mbox{\boldmath$\theta$}})=|{\rm det}(\Delta_{ij}({\mbox{\boldmath$\theta$}}))| \delta^D(\Delta_i({\mbox{\boldmath$\theta$}})). \label{eqn:pkgauss}$$ The summation symbol in equation (\[eqn:sumpk\]) can be eliminated because the delta function of the above equation picks out all of the extremal points which are zero of $\Delta_i({\mbox{\boldmath$\theta$}})$ are maximum in the two dimensional map. Hence the ensemble average of equation (\[eqn:pkgauss\]) produces the differential mean number density of hotspots of height in the range of $\nu$ and $\nu+d\nu$: $$\begin{aligned} \bar{n}_{\rm pk}(\nu)&=&{\left\langle{|{\rm det}(\Delta_{ij})|\delta^D(\Delta_i)}\right\rangle} =\frac{1}{2\theta_\ast^2}{\left\langle{|X^2-Y^2-Z^2|\delta^D(\eta_i)}\right\rangle} \nonumber \\ &=&\frac{1}{2\theta_\ast^2}\int^\infty_0\!\!dX\int^X_{-X}\!\!dY \int^{\sqrt{X^2-Y^2}}_{-\sqrt{X^2-Y^2}}dZ(X^2-Y^2-Z^2) p_1(\nu,X,Y,Z,\eta_i=0),\end{aligned}$$ where we have adopted the conditions for a hotspot of $X>0$ and $Y^2+Z^2<X^2$. BE derived the analytical expression for $\bar{n}_{\rm pk}(\nu)$; $$\bar{n}_{\rm pk}(\nu)=\frac{1}{(2\pi)^{3/2}\theta^2_\ast} \exp(-\nu^2/2)G(\gamma,\gamma\nu), \label{eqn:meannum}$$ where $$\begin{aligned} G(\gamma,x_\ast)&\equiv&(x_\ast^2-\gamma^2)\left\{1-\frac{1}{2} {\rm erfc}\left[ \frac{x_\ast}{\sqrt{2(1-\gamma^2)}}\right]\right\}+x_\ast(1-\gamma^2) \frac{\exp\{-x_\ast^2/ [2(1-\gamma^2)]\}}{\sqrt{2\pi(1-\gamma^2)}}\nonumber \\ &&+\frac{\exp[-x_\ast^2/(3-2\gamma^2)]}{(3-2\gamma^2)^{1/2}} \left\{1-\frac{1}{2}{\rm erfc}\left[\frac{x_\ast}{\sqrt{2 (1-\gamma^2)(3-2\gamma^2)}}\right] \right\}. \label{eqn:diffnum}\end{aligned}$$ In this paper, we have often used the mean number density of peak of height above a certain threshold $\nu_{\rm thresh}$ obtained by integrating equation (\[eqn:meannum\]) over $\nu>\nu_{\rm thresh}$. Naturally, the mean number density of coldspots of height below $-\nu$ ($\Delta_{\rm pk}<-\nu\sigma_0$) is symmetrically given by $\bar{n}_{\rm cold}(<-\nu)=\bar{n}_{\rm pk}(>\nu)$. The unlensed two-point correlation function of hotspots {#app:unlensxi} ======================================================= Based on the peak statistics for the two dimensional Gaussian field, we briefly review the derivation of two-point correlation function of hotspots following HS. For this purpose, let us consider two hotspots separated by the angular scale $\theta$. In the same way as in appendix \[app:num\], we then need the joint probability density function for the 12 independent variables ${\mbox{\boldmath$v$}}=({\mbox{\boldmath$v$}}_1,{\mbox{\boldmath$v$}}_2) =({\mbox{\boldmath$v$}}({\mbox{\boldmath$\theta$}}_1),{\mbox{\boldmath$v$}}({\mbox{\boldmath$\theta$}}_2))$ with $\theta=|{\mbox{\boldmath$\theta$}}_1-{\mbox{\boldmath$\theta$}}_2|$. If using the variables (\[eqn:gaussvari\]) for each hotspot, we can block the covariance matrix $M_{ij}$ for the 12 variables in the order $(\nu_1,X_1,\eta_{1x},Y_1,\nu_2,X_2, \eta_{2x},Y_2,\eta_{1y} ,\eta_{2y},Z_1, Z_2 )$ into the $8\times 8$-matrix $M_{(8)}$ (upper left) and the $4\times4$-matrix $M_{(4)}$ (bottom right); $$M_{(8)}=\left( \begin{array}{cccccccc} 1&\tilde{\gamma}&0&0 &\lambda_{000}& \lambda_{020}&-\lambda_{011}&\lambda_{022}\\ \ddots &1&0&0&\lambda_{020}&\lambda_{220}&-\lambda_{121} & \lambda_{222}\\ \ddots &\ddots &1/2&0&\lambda_{011}&\lambda_{121} &\frac{1}{2}(\lambda_{110} - \lambda_{112})&\frac{1}{2}(\lambda_{123} - \lambda_{121})\\ \ddots&\ddots&\ddots&1/2& \lambda_{022}&\lambda_{222}& -\frac{1}{2}(\lambda_{123} - \lambda_{121})&\frac{1}{2}(\lambda_{220} + \lambda_{224})\\ \ddots &\ddots&\ddots&\ddots&1&\gamma&0&0\\ \ddots &\ddots&\ddots&\ddots&\ddots&1&0&0\\ \ddots &\ddots&\ddots&\ddots&\ddots&\ddots&1/2&0\\ \ddots &\ddots&\ddots&\ddots&\ddots&\ddots&\ddots&1/2 \end{array} \right),$$ $$M_{(4)}=\left( \begin{array}{cccc} 1/2 &\frac{1}{2}(\lambda_{110} + \lambda_{112})& 0 & -\frac{1}{2}(\lambda_{121} + \lambda_{123})\\ \ddots&1/2&-\frac{1}{2}(\lambda_{121} + \lambda_{123})&0\\ \ddots &\ddots&1/2&\frac{1}{2}(\lambda_{220} - \lambda_{224})\\ \ddots &\ddots&\ddots&1/2 \end{array} \right),$$ where we have introduced the notations (HS): $$\begin{aligned} &&\lambda_{ijn}\equiv\frac{1}{\sigma_i\sigma_j}\int \frac{ldl}{2\pi}l^{i+j}C_lJ_n(l\theta). \label{eqn:cov2}\end{aligned}$$ Note that both $M_8$ and $M_4$ are symmetric. The joint PDF for the two hotspots can thus be expressed in terms of the above covariant matrix; $$p_2({\mbox{\boldmath$v$}}_1,{\mbox{\boldmath$v$}}_2)=\frac{1}{(2\pi)^6|{\rm det}(M_{ij})|^{1/2}} \exp\left(-\frac{1}{2}v_iM^{-1}_{ij}v_j\right).$$ Therefore, as in the similar way of the derivation of mean number density, we can calculate the unlensed two-point correlation function of hotspots which are separated by $\theta$ and have heights above $\nu_1$ and $\nu_2$, respectively; $$\begin{aligned} 1+{\xi_{\rm pk-pk}}(\theta)&=&\frac{1}{\bar{n}_{\rm pk}(>\nu_1)\bar{n}_{\rm pk}(>\nu_2)} {\left\langle{n_{\rm pk}({\mbox{\boldmath$\theta$}}_1)n_{\rm pk}({\mbox{\boldmath$\theta$}}_2)}\right\rangle}_{|{\mbox{\boldmath$\theta$}}_1 -{\mbox{\boldmath$\theta$}}_2|=\theta} \nonumber \\ &=&\frac{1}{4\theta_\ast^4 \bar{n}_{\rm pk}(>\nu_1)\bar{n}_{\rm pk}(>n_2)} {\left\langle{(X^{\prime 2}_1-Y^{\prime 2}_1-Z^{\prime 2}_1) (X^{\prime 2}_2-Y^{\prime 2}_2-Z^{\prime 2}_2)\delta^D(\eta'_{1i}) \delta^D(\eta'_{2i})}\right\rangle}_{|{\mbox{\boldmath$\theta$}}_1-{\mbox{\boldmath$\theta$}}_2|=\theta} \nonumber \\ &=&\frac{1}{4\theta_\ast^4 \bar{n}_{\rm pk}(>\nu_1)\bar{n}_{\rm pk}(>n_2)} \int_{\nu_1}^\infty\!\!d\nu_1'\int_{\nu_2}^\infty\!\!d\nu_2' \int_0^\infty\!\!dX_1'\int_0^\infty\!\!dX_2'\int^{X_1'}_{-X_1'}\!\!dY_1' \int^{X_2'}_{-X_2'}\!\!dY_2' \nonumber \\ &&\times\int^{\sqrt{X_1^{\prime 2}-Y_1^{\prime 2}}} _{-\sqrt{X_1^{\prime 2}-Y_1^{\prime 2}}}\!\!dZ_1' \int^{\sqrt{X_2^{\prime 2}-Y_2^{\prime 2}}} _{-\sqrt{X_2^{\prime 2}-Y_2^{\prime 2}}}\!\!dZ_2' (X^{\prime 2}_1-Y^{\prime 2}_1-Z^{\prime 2}_1) (X^{\prime 2}_2-Y^{\prime 2}_2-Z^{\prime 2}_2)\nonumber \\ && \times p_2(\nu_1',X_1',Y_1',Z_1',\eta'_{1i}=0,\nu_2',X_2',Y_2',Z_2', \eta'_{2i}=0). \label{eqn:unlensxi}\end{aligned}$$ To obtain ${\xi_{\rm pk-pk}}(\theta)$, we performed a six dimensional numerical integration of the equation (\[eqn:unlensxi\]) because the two integrals can be done analytically. Alternative representation of the lensed hotspots correlation function {#app:xianaly} ====================================================================== The lensing dispersion $\sigma_{\rm GL}(\theta)$ generally has isotropic and anisotropic contributions, which are expressed by $\sigma_{\rm GL,0}(\theta)$ and $\sigma_{\rm GL,2}(\theta)$, respectively, as shown in equation (\[eqn:gldisp\]). For the standard power spectra of mass fluctuation as adopted in this paper, the contribution of $\sigma_{{\rm GL},2}(\theta)$ is smaller than that of $\sigma_{{\rm GL},0}(\theta)$. If the neglecting the contribution, the key equation (\[eqn:lensxi\]) of this paper can be further simplified. In this case, the equation (\[eqn:lensxi\]) becomes $$\begin{aligned} \tilde{\xi}_{\rm pk-pk}(\theta)\approx . \label{eqn:xianaly}\int\!\!d\theta'\theta'\int\!\!ldl \xi_{\rm pk-pk}(\theta') J_{0}(l\theta')\exp\left[-\frac{l^2}{2}\sigma^2_{\rm GL,0}(\theta)\right] J_0(l\theta), \end{aligned}$$ and this can be then rewritten as $$\tilde{\xi}_{\rm pk-pk}(\theta)=\int\!\!\theta'd\theta'\xi_{\rm pk-pk}(\theta') K(\theta,\theta'),$$ where the kernel $K(\theta,\theta')$ is given by $$K(\theta,\theta')\equiv \int^{\infty}_0\!\!ldl J_0(l\theta)J_0(l\theta') \exp\left[-\frac{l^2}{2}\sigma^2_{{\rm GL},0}(\theta)\right] =\frac{1}{\sigma^2_{{\rm GL},0}(\theta)} \exp\left[-\frac{\theta^2+\theta'^{2}}{2\sigma^2_{{\rm GL},0}(\theta)}\right]I_0\left[\frac{\theta\theta'}{\sigma^2_{{\rm GL},0}(\theta)}\right].$$ $I_0(x)$ is the modified zeroth-order Bessel function, and we have used equation (6.633.2) of Gradshteyn & Ryzhik (1994). As shown in Figure \[fig:dispgl\], $\sigma_{\rm GL,0}(\theta)/\theta<1$ and the argument of $I_0$ is generally a large number. Therefore, if using the approximation $I_0(x)\approx (2\pi x)^{-1/2}\exp(x)$ for $x\rightarrow \infty$, we can rewrite equation (\[eqn:xianaly\]) in the following form; $$\xi^{\rm GL}_{\rm pk-pk}(\theta)\approx \frac{1}{(2\pi \theta)^{1/2}\sigma_{\rm GL,0}(\theta)} \int\!\!\theta'd\theta'\xi_{\rm pk-pk}(\theta')\exp \left[-\frac{(\theta-\theta')^2} {2\sigma_{\rm GL,0}^2(\theta)}\right]$$ This expression explicitly explains that the lensing effect on ${\xi_{\rm pk-pk}}(\theta)$ appears as a Gaussian smoothing with relative width $\sigma_{\rm GL,0}$, and then the asymptotic behavior at $\sigma_{\rm GL,0}\rightarrow 0$ is naturally ${\xi_{\rm pk-pk}}^{\rm GL} (\theta)\rightarrow {\xi_{\rm pk-pk}}(\theta)$. Bacon, D., Refregier, A., & Ellis, R. 2000,  submitted, astro-ph/0003008 Balbi, A. et al. 2000, astro-ph/0005124 Bardeen, J. M., Bond., J. R., Kaiser, N., & Szalay, A. S. 1986, , 304, 15 (BBKS) Bartelmann, M., & Schneider, P. 1999, astro-ph/9912508 Bernardeau, F. 1997, , 324, 15 Bernardeau, F. 1998, , 338, 767 de Bernardis, P., et al. 2000, , 404, 955 Blandford, R. D., Saust, A. B., Brainerd, T. G., & Villumsen, J. V. 1991, , 251, 600 Bond, J. R., & Efstathiou, G. P. 1987, , 226, 655 (BE) Bond, J. R., Efstathiou, G. P., & Tegmark, M. 1997, , 291, L33 Bunn, E. F., & White, M. 1997, , 480, 6 Burles, S. & Tytler, D. 1998 , 507, 732 Eke, V., Cole, S., & Frenk, C. S. 1996, , 282, 263 Gradshteyn, I. S., & Ryzhik, I. M. 1994, Tabel of Integrals, Series, and Products, Academic Press, 5th edition Gunn, J., 1967, , 147, 61 Heavens, A. F., & Sheth, R. K. 1999, , 310, 1062 (HS) Hanany, S. et al. 2000, astro-ph/0005123 Hu, W. 2000, , 62, 043007 Hu, W., Sugiyama, N., & Silk, J. 1997, , 386, 37 Kaiser, N. 1992, , 388, 272 Kaiser, N., Wilson, G., & Luppino, G. 2000,  Letters submitted, astro-ph/0003338 Kitayama, T., & Suto, Y. 1997, , 490, 557 Komatsu, E., & Kitayama, T. 1999, , 526, L1 Lange, A. E. [*et al.*]{} 2000, astro-ph/0005004 Metcalf, R. B., & Silk, J. 1997, , 489, 1 Miralda-Escude, J. 1991, , 380, 1 Peacock, J. A., & Dodds, S. J. 1996, , 280, L19 Seljak, U. 1996a, , 460, 549 Seljak, U. 1996b, , 463, 1 Seljak, U., & Zaldarriaga, M. 1996, , 469, 437 Seljak, U., & Zaldarriaga, M. 2000, , 538, 57 Stompor, R., & Efstathiou, G. 1999, , 302, 735 Sugiyama, N. 1995 , 100, 281 Takada, M., Komatsu, E., & Futamase, T. 2000, , 533, L83 (TKF) Toffolatti, M. et al. 1998, , 297, 117 Van Waerbeke, L., et al. 2000, 358, 30 Van Waerbeke, L., Bernardeau, F., & Benabed, K. 1999, astro-ph/9910366 Wittman, D., Tyson, J., Kirkman, D., Dell’Antonio, I., & Bernstein, G. 2000, , 405, 143 Zaldarriaga, M., & Seljak, U. 1999, , 59, 123507 Zaldarriaga, M. 2000, , 62, 063510 [^1]: [http://map.gsfc.nasa.gov]{} [^2]: [http://astro.estec.esa.nl/SA-general/Projects/Planck/]{}
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study the topological pump for a lattice fermion model mainly in three spatial dimensions. We first calculate the U(1) current density for the Dirac model defined in continuous space-time to review the known results as well as to introduce some technical details convenient for the calculations of the lattice model. We next investigate the U(1) current density for a lattice fermion model, a variant of the Wilson-Dirac model. The model we introduce is defined on a lattice in space but in continuous time, which is suited for the study of the topological pump. For such a model, we derive the conserved U(1) current density and calculate it directly for the $1+1$ dimensional system as well as $3+1$ dimensional system in the limit of the small lattice constant. We find that the current includes a nontrivial lattice effect characterized by the Chern number, and therefore, the pumped particle number is quantized by the topological reason. Finally we study the topological temporal pump in $3+1$ dimensions by numerical calculations. We discuss the relationship between the second Chern number and the first Chern number, the bulk-edge correspondence, and the generalized Streda formula which enables us to compute the second Chern number using the spectral asymmetry.' author: - Takahiro Fukui - Takanori Fujiwara title: Topological magnetoelectric pump in three dimensions --- Introduction ============ In a topological background such as a soliton or a vortex, the vacuum state of the Dirac fermion shows a nontrivial topological structure. [@Goldstone:1981aa; @Wilczek:1987ab] One of famous examples in condensed matter physics is the mid-gap states of the SSH soliton,[@Su:1979aa; @Su:1980aa] which can be effectively described by a Dirac fermion model with a nontrivial mass term.[@Takayama:1980aa] Recent discovery of topological insulators [@Kane:2005aa; @Qi:2008aa] tells us that topological states of matter are richer than we expected, [@Hasan:2010fk; @Qi:2011kx] and the Dirac fermion model is very convenient for the classification of symmetry classes. [@Schnyder:2008aa; @Ryu:2010uq] The topological pump in one dimensional (1D) systems has been proposed by Thouless,[@Thouless:1983fk] and experimentally observed quite recently. [@Nakajima:2016aa; @Lohse:2016aa] This can also be described very simply by the Dirac fermion, as already seen in Ref. [@Goldstone:1981aa]. The topological pump has been generalized to three dimensional (3D) systems.[@Qi:2008aa] The 3D pump is unique, since it is a part of topological magneto-electric effect, [@Qi:2008aa] which has close relationship with the chiral anomaly of the Dirac fermion.[@Wilczek:1987ab] Here, the pumping parameter plays a role of the axion field.[@Wilczek:1987ab] The magneto-electric response in generic systems has also been studied by developing the theory of the orbital magneto-electric polarization. [@Essin:2009aa; @Essin:2010aa; @Malashevich:2010aa] The chiral magnetic effect (CME), originally proposed for the quark-gluon plasma, [@Fukushima:2008aa] has also been attracting much interest in condensed matter physics. In particular, the discovery of the Weyl semimetal [@Murakami:2007aa; @Burkov:2011ab; @Wan:2011aa; @Xiong:2015aa; @1503.08179; @Zhang:2017aa] has led our interest to the observation of the chiral anomaly in a crystal through the magneto-electric response, [@Nielsen:1983aa; @Chen:2013aa; @Parameswaran:2014aa] including the CME, [@Basar:2014aa; @Zyuzin:2012aa; @Jian-Hui:2013aa; @Chang:2015aa; @Ma:2015aa; @Sumiyoshi:2016ab; @Alavirad:2016aa; @OBrien:2017aa; @Sekine:2016aa; @Sekine:2016ab] the anomalous Hall effect (AHE), [@Yang:2011aa; @Zyuzin:2012aa; @Grushin:2012aa; @Vazifeh:2013aa; @Goswami:2013aa; @Sekine:2016aa; @Sekine:2016ab] axial-magneto-electric effect, [@Grushin:2016aa; @Pikulin:2016aa] and Z$_2$ anomaly in Dirac semimetals, [@Burkov:2016aa] etc. Thus, the chiral anomaly and its related phenomena have been one of hot topics in condensed matter physics. In this paper, we examine mainly the topological pump in a 3D system using a lattice fermion model. In the next section \[s:D\], we present the U(1) current of the Dirac fermion in continuous space-time. Here, some notations are fixed and some techniques convenient to the lattice model are given. In Sec. \[s:WD\_la\], we introduce a Wilson-Dirac model defined on the spatial 1D or 3D lattice but in continuous time to study the topological temporal pump in Sec. \[s:MEP\]. Throughout the paper, this model is simply referred to as Wilson-Dirac model. In Sec. \[s:WD\_cc\], we derive the conserved U(1) current density for the Wilson-Dirac model. Because of continuous time, the charge density is the same as that of the Dirac fermion, whereas the current density includes some lattice effects. Therefore, we calculate the charge density and the current density separately in Sec. \[s:ChaDen\] and Appendix \[s:CurDen\]. In Sec. \[s:E\], we calculate the Chern numbers exactly. In particular, we show that the 3D model has nontrivial second Chern number $1$ or $-2$. It should be noted that such second Chern numbers are due to the Berry curvature of the wave function in the zero field limit. Derivation of the exact conserved current and the second Chern number for the lattice Wilson-Dirac model are one of main results of the present paper. In Sec. \[s:MEP\], we restrict our discussions to the temporal 3D pump, and present some numerical results in detail. Various numerical analyses of the 3D pump which give physical interpretations of the exact results in Sec. \[s:WD\] are another main results of the present paper. In Sec. \[s:3dpump\], the number of pumped particles is explicitly derived in the case of a static and uniform magnetic field. This number is proportional to the second Chern number as well as the magnetic field and the system size.[@Qi:2008aa] In 3D systems, particles are pumped toward the direction of the applied magnetic field. [@Qi:2008aa] Therefore, it can also be viewed as a 1D pump. Based on the Thouless formula for the 1D pump, [@Thouless:1983fk] the number of the pumped particle is rederived, which is given by the first Chern number. Here, the first Chern number is computed using the wave function under the magnetic field. The equivalence of the two formula gives an interesting relation between two kinds of Chern numbers. We present numerical calculation of the first Chern number, and show that this relation is valid as far as the mass gap is open. The relationship between the second Chern number associated with the Berry curvature under zero magnetic field and the first Chern number associated with the Berry curvature under a finite magnetic field is one of the main results. In Sec. \[s:BEC\], we consider the system with boundaries perpendicular to the magnetic field. Then it is possible to define the center of mass of the occupied particles along the pumped direction, which played a central role in the experimental observation of the 1D pump. [@Wang:2013fk_pump; @Nakajima:2016aa; @Lohse:2016aa] We show that the bulk-edge correspondence recently established for the 1D pump [@Hatsugai:2016aa] is also valid for the 3D pump. Thus, from the behavior of the center of mass as the function of time, we can compute the number of the pumped particle, and hence the second Chern number. Finally in Sec. \[s:Streda\], we give an alternative method of computing the second Chern number (and hence, the number of the pumped particles) using the generalized Streda formula.[@Fukui:2016aa] This method is based on the four dimensional Hamiltonian in discrete time whose spectral asymmetry gives the chiral anomaly for the lattice fermion. Since the generalized Streda formula is the only one method of numerically computing the second Chern number for generic models, the success in reproducing the exact Chern number for the present model is of significance. In Sec. \[s:sd\], we give summary and discussion including outlook. Dirac model {#s:D} =========== The main topic of this paper is to calculate the U(1) current for the Dirac fermion on the lattice, including external scalar and pseudscalar fields which serve as nonuniform mass terms. Calculations on the lattice, however, is quite complicated, so that we first examine the Dirac fermion defined in continuous space-time, which may be of help in Sec. \[s:WD\]. Let us directly calculate the vector current for the model described by the Lagrangian density $$\begin{aligned} 1 {\cal L}=\bar\psi\left(i\gamma^\mu D_\mu-\sigma-i\gamma_5\pi\right)\psi, \label{LagCon}$$ where $\mu=0,1$ with $\gamma^0=\sigma^1$, $\gamma^1=-i\sigma^2$, and $\gamma_5=\sigma^3$ for a $d=1+1$ model, whereas $\mu=0,\cdots,3$ for a $d=3+1$ model with $\gamma$ matrices defined by $$\begin{aligned} 1 \gamma^0=\begin{pmatrix}0&\bm1\\\bm1&0\end{pmatrix}, \quad \bm\gamma=\begin{pmatrix}0&-\bm\sigma\\\bm\sigma&0\end{pmatrix}, \quad \gamma_5=\begin{pmatrix}\bm1&0\\0&-\bm1\end{pmatrix}.\end{aligned}$$ These satisfy $\{\gamma^\mu,\gamma^\nu\}=2g^{\mu\nu}$, where $g^{\mu\nu}=\mbox{diag}(1,-1,-1,-1)$, and $\gamma_5=i\gamma^0\gamma^1\gamma^2\gamma^3$. The covariant derivative under a background electro-magnetic field is defined by $D_\mu=\partial_\mu-ieA_\mu(x)$. Note $e<0$ for electrons. We have introduced external scalar and pseudscalar fields denoted by $\sigma$ and $\pi$. This model has been studied in Ref.[@Goldstone:1981aa] for the fractional fermion numbers on solitons and in Ref. [@Sekine:2016ab] for the CME and AHE for an antiferromagnetic topological insulator with broken time reversal and parity symmetries. In this paper, we assume that $\sigma$ and $\pi$ depend on the coordinates $x=(t,\bm x)$ through a single parameter $\theta(x)$ such that $\sigma(x)=m\cos\theta(x)$ and $\pi(x)=m\sin\theta(x)$. Then, we readily see that $\sigma+i\gamma_5\pi=me^{i\gamma_5\theta}$ in Eq. (\[LagCon\]). This model is invariant not only U(1) but also axial U(1) transformations, $\psi\rightarrow e^{i\gamma_5 \alpha}\psi$, $\bar\psi\rightarrow \bar\psi e^{i\gamma_5 \alpha}$, and $\theta\rightarrow \theta-2\alpha$. In what follows, we calculate the U(1) vector current in the $m\rightarrow\infty$ limit. The U(1) vector current is defined by $$\begin{aligned} 1 \langle j^{\mu}(x)\rangle&=\langle0|\bar\psi(x)\gamma^\mu\psi(x)|0\rangle \nonumber\\ &=-\lim_{x'\rightarrow x}{{\rm tr}\,}\gamma^\mu\langle0|T\psi(x)\bar\psi(x')|0\rangle, \label{CurCon}$$ where the propagator is given by $$\begin{aligned} 1 \langle0|T\psi(x)\bar\psi(x')|0\rangle=\frac{i}{i\slashed D-me^{i\gamma_5\theta}+i\epsilon}\delta(x-x').\end{aligned}$$ The positive infinitesimal constant $\epsilon$ will be sometimes suppressed for simplicity below. Inserting $$\begin{aligned} 1 \delta(x-x')=\int\frac{d^dk}{(2\pi)^d}e^{ik(x-x')},\end{aligned}$$ with $kx=\omega t-\bm k\cdot\bm x$ and $d=2$ $(4)$ for a $1+1$ ($1+3$) dimensional system, we can write Eq. (\[CurCon\]) as $$\begin{aligned} 1 \langle j^\mu(x)\rangle=\int\frac{d^dk}{i(2\pi)^d}e^{-ikx}{{\rm tr}\,}\gamma^\mu \frac{1}{i\slashed D-me^{i\gamma_5\theta}+i\epsilon}e^{ikx}. \label{CurCon2}$$ To calculate the topological sector of the current, it is convenient to use the identity $$\begin{aligned} 1 &\frac{1}{i\slashed D-me^{i\gamma_5\theta}}=(-i\slashed D-me^{-i\gamma_5\theta}) \frac{1}{\slashed D^2+m^2+me^{i\gamma_5\theta}\gamma_5\slashed\partial\theta}. \label{ConIde}$$ In the denominator, we further have $\slashed D^2 =D^\mu D_\mu-\frac{i\gamma^\mu\gamma^\nu}{2}eF_{\mu\nu}$, where $F_{\mu\nu}=\partial_\mu A_\nu-\partial_\nu A_\mu$ is the field strength of the background electro-magnetic field. Note also $e^{-ikx}\slashed D e^{ikx}=\slashed D+i\slashed k$. Finally, let us make a scale transformation for the momentum, $k^\mu\rightarrow mk^\mu$, and expand the propagator with respect to $1/m$. Then, Eq. (\[CurCon2\]) becomes $$\begin{aligned} 1 \langle j^\mu(x)\rangle &=\int\frac{d^dk}{i(2\pi)^d}m^{d-2}{{\rm tr}\,}\gamma^\mu (-i\slashed D+m\slashed k-me^{-i\gamma_5\theta}) \frac{1}{1-k^2+e^{i\gamma_5\theta}\gamma_5\slashed\partial\theta/m- \frac{i\gamma^\rho\gamma^\sigma}{2} eF_{\rho\sigma}/m^2+\mathcal{O}/m^2} \nonumber\\ &=\int\frac{d^dk}{i(2\pi)^d}\sum_{n=0}^\infty\frac{m^{d-1-n}}{(1-k^2-i\epsilon)^{n+1}} {{\rm tr}\,}\gamma^\mu \left(-i\frac{\slashed{D}}{m}+\slashed k-e^{-i\gamma_5\theta}\right) \left(-e^{i\gamma_5\theta}\gamma_5\slashed\partial\theta+ \frac{i\gamma^\rho\gamma^\sigma}{2m}{eF}_{\rho\sigma}+\frac{\mathcal{O}}{m}\right)^n, \label{CurCon3}$$ where $k^2=\omega^2-\bm k^2$ and $\mathcal{O}=2ik^\mu D_\mu+D^\mu D_\mu$. The topological sector of the current is associated with the terms including $\gamma_5$. For the $d=1+1$ system, using (\[TraGam1D\]), it turns out that the $n=1$ term in Eq. (\[CurCon3\]) survives in the $m\rightarrow\infty$ limit. Thus, we have $$\begin{aligned} 1 \langle j^\mu(x)\rangle &=\int\frac{d^2k}{i(2\pi)^2}\frac{{{\rm tr}\,}\gamma^\mu\gamma_5\gamma^\nu\partial_\nu\theta}{(1-k^2-i\epsilon)^2} +O(m^{-1}) =\frac{1}{2\pi}\epsilon^{\mu\nu}\partial_\nu\theta, \label{DCur1}$$ where we have used Eq. (\[Int\_1\]). For the $d=3+1$ system, using (\[TraGam3D\]), we see that among terms with $\gamma_5$, only those included in the $n=2$ term in Eq. (\[CurCon3\]) survive, $$\begin{aligned} 1 \langle j^\mu(x)\rangle &=\int\frac{d^4k}{i(2\pi)^4}\frac{m}{(1-k^2-i\epsilon)^{3}} {{\rm tr}\,}\gamma^\mu\cdot 2 \gamma_5\gamma^\nu\partial_\nu\theta \frac{i\gamma^\rho\gamma^\sigma}{2m}{eF}_{\rho\sigma}+O(m^{-1}) =-\frac{e}{8\pi^2}\epsilon^{\mu\nu\rho\sigma}(\partial_\nu\theta) F_{\rho\sigma}. \label{DCur3_1}$$ This result can be written more explicitly as $$\begin{aligned} 1 &\rho(x)=\frac{-e}{4\pi^2}\nabla\theta(x)\cdot\bm B(x), \nonumber\\ &\bm j(x)=\frac{e}{4\pi^2}\left[\dot\theta(x)\bm B(x)+\nabla\theta(x)\times\bm E(x)\right]. \label{DCur3_2}\end{aligned}$$ Thus, we have reached the known result.[@Wilczek:1987ab; @Qi:2008aa] In the next section, we use a similar method to calculate the current for the Wilson-Dirac model. Wilson-Dirac model {#s:WD} ================== In this section, we calculate the U(1) current density for the Wilson-Dirac model based on a similar technique demonstrated in Sec. \[s:D\]. Lattice action {#s:WD_la} -------------- We consider the Wilson-Dirac Hamiltonian[@Qi:2008aa] defined on the 1D or 3D regular lattices, $$\begin{aligned} 1 &H=\sum_{\bm x}a^{d-1}\psi^\dagger(t,\bm x) {\cal H}(t,\bm x) \psi(t,\bm x), \nonumber\\ &{\cal H}(t,\bm x)\equiv -i\bm\alpha\cdot \bm D^{\rm L} -\beta\left(\frac{m}{a} e^{i\gamma_5\theta}+\frac{ba}{2}\Delta^{\rm L}\right), \label{WDHam}$$ where $a$ is the lattice constant, and $\bm\alpha$ and $\beta$ are given by the $\gamma$-matrices in Sec. \[s:D\], $\bm\alpha=\beta\bm\gamma$ and $\beta=\gamma^0$. The mass is written by $m/a$, where $m$ is a dimensionless parameter. We will keep $m$ finite as $a\rightarrow0$. This corresponds to taking $m\rightarrow\infty$ limit in the continuum theory in Sec. \[s:D\]. The lattice operators are introduced by $$\begin{aligned} 1 D_j^{\rm L}&=\frac{1}{2}(\nabla_j+\nabla_j^*),\quad \Delta^{\rm L}=\frac{1}{a}\sum_{j}(\nabla_j-\nabla_j^*), \label{LatDirOpe}\end{aligned}$$ where the forward and backward covariant differences are defined by $$\begin{aligned} 1 &\nabla_j\psi(t,\bm x)=\frac{1}{a}\left[U_j(t,\bm x)\psi(t,\bm x+a\hat j)-\psi(t,\bm x)\right], \nonumber\\ &\nabla_j^*\psi(t,\bm x)=\frac{1}{a}\left[\psi(t,\bm x)-U^\dagger_j( t,\bm x-a\hat j)\psi(t,\bm x-a\hat j)\right], \label{CovDifOpe}$$ with the gauge field $U_j(t,\bm x)=e^{-ieaA_j(t,\bm x)}$, and $\hat j$ stands for the unit vector in the $j$ direction. The term associated with the Laplacian on the lattice, $\Delta^{\rm L}$, is the Wilson term originally introduced to avoid the doubling of fermions. It violates the axial U(1) symmetry and is the origin of the axial anomaly. In the condensed matter physics literature, it is known to control the topological property of the ground state. In this paper, we consider the case $b>0$. We derive the current of the lattice model in the limit $a\rightarrow0$ in a way similar to the continuum model in Sec. \[s:D\]. We treat time $t$ as continuous variable, so that the lattice action reads $$\begin{aligned} 1 S=\int_{-\infty}^\infty dt\sum_{\bm x}a^{d-1}\bar\psi(x) \left( i\slashed D^{\rm L}-\frac{m}{a}e^{i\gamma_5\theta}-\frac{ba}{2}\Delta^L \right)\psi(x), \label{LatAct}$$ where $x=(t,\bm x)$ and the covariant derivative with respect to $t$ is the same as that in the continuum model, $D_0\equiv D_0^{\rm L}=\partial_0-ieA_0$. The fermion propagator on the lattice is $$\begin{aligned} 1 \langle0&|T\psi(x)\bar\psi(x')|0\rangle \nonumber\\ &=\frac{1}{i\slashed D^{\rm L}-\frac{m}{a}e^{i\gamma_5\theta}-\frac{ba}{2}\Delta^{\rm L}+i\epsilon}\frac{i}{a^{d-1}} \delta(t-t')\delta_{\bm x,\bm x'} , \label{WilDirPro}$$ where $\epsilon$ is a positive infinitesimal constant to implement the time ordering, which will be sometimes suppressed for simplicity below. This follows from the identity $$\begin{aligned} 1 0&=\frac{1}{Z}\int{\cal D}\psi{\cal D\bar\psi}\frac{\delta}{\delta\bar\psi(x)}\left[e^{iS}\bar\psi(x')\right] \nonumber\\ &=ia^{d-1}\left(i\slashed D^{\rm L}-\frac{m}{a}e^{i\gamma_5\theta}-\frac{ba}{2}\Delta^{\rm L}\right) \langle\psi(x)\bar\psi(x')\rangle \nonumber\\ &\qquad+\delta(t-t')\delta_{\bm x,\bm x'} . $$ Conserved U(1) current {#s:WD_cc} ---------------------- The action (\[LatAct\]) is invariant under the gauge transformation $$\begin{aligned} 1 &\psi'(x)=\Lambda(x)\psi(x), \nonumber\\ &\bar\psi'(x)=\bar\psi(x)\Lambda^\dagger(x), \nonumber\\ &eA_0'(x)=eA_0(x)-i\Lambda^\dagger(x)\partial_0\Lambda(x), \nonumber\\ &U'_j(x)=\Lambda(x)U_j(x)\Lambda^\dagger(x+a\hat j).\end{aligned}$$ For the U(1) gauge theory, the conserved current density can be obtained by considering an infinitesimal gauge transformation, $\Lambda(x)=1+i\lambda(x)$, which induces $$\begin{aligned} 1 &\delta A_0(x)=\partial_0\lambda(x), \nonumber\\ &\delta U_j(x)=-ia\left(\partial_j\lambda(x)\right)U_j(x), \nonumber\\ &\delta U_j^\dagger(x-a\hat j)=ia\left(\partial_j^*\lambda(x)\right)U_j^\dagger(x-a\hat j), \label{InfGauTra}$$ where we have introduced the forward and backward differences by $$\begin{aligned} 1 &\partial_j\lambda(x)=\frac{1}{a}\left[\lambda(x+a\hat j)-\lambda(x)\right], \nonumber\\ &\partial^*_j\lambda(x)=\frac{1}{a}\left[\lambda(x)-\lambda(x-a\hat j)\right]. \label{DifOpe}$$ It is only in this subsection \[s:WD\_cc\] to use the symbols $\partial_j$ and $\partial_j^*$ as differences. The change of the action (\[LatAct\]) under the infinitesimal gauge transformation reads $$\begin{aligned} 1 \delta S&=\int dt\sum_{\bm x}a^{d-1}\Big\{ \bar\psi(x)i\gamma^0\left[-ie\delta A_0(x)\right]\psi(x) +\frac{i}{2a}\bar\psi(x)\sum_j\gamma^j \left[\delta U_j(x)\psi(x+a\hat j)-\delta U_j^\dagger(x-a\hat j)\psi(x-a\hat j)\right] \nonumber\\ &\qquad\qquad -\frac{b}{2a}\bar\psi(x)\sum_j \left[\delta U_j(x)\psi(x+a\hat j)+\delta U_j^\dagger(x-a\hat j)\psi(x-a\hat j)\right] \Big\} .\end{aligned}$$ Substituting the infinitesimal gauge transformation (\[InfGauTra\]) into the above equation and using the relations $\sum_x\partial_jf(x)g(x)=-\sum_xf(x)\partial_j^* g(x)$, and $\partial_jf(x)=\partial_j^*f(x+a\hat j)$, we have $$\begin{aligned} 1 \delta S&=-\int dt\sum_xa^{d-1}\lambda(x)\Big\{ \partial_0\left[\bar\psi(x)\gamma^0\psi(x)\right] +\frac{1}{2}\sum_j\partial_j^*\left[\bar\psi(x)\gamma^jU_j(x)\psi(x+a\hat j) +\bar\psi(x+a\hat j)\gamma^jU_j^\dagger(x)\psi(x)\right] \nonumber\\ &\qquad\qquad+\frac{ib}{2}\sum_j\partial_j^*\left[ \bar\psi(x) U_j(x)\psi(x+a\hat j)-\bar\psi(x+a\hat j)U_j^\dagger(x)\psi(x)\right] \Big\} \nonumber\\ &\equiv -\int dt\sum_{\bm x}a^{d-1}\lambda(x)\left[\partial_0j^0(x)+\sum_l\partial_l^*j^l(x)\right] ,\end{aligned}$$ where $\partial_0=\partial_t$ is the derivative with respect to $t$ and $\partial_j^*$ is the backward difference operator in Eq. (\[DifOpe\]). Thus, we reach the conserved U(1) current density $$\begin{aligned} 1 &j^0(x)=\bar\psi(x)\gamma^0\psi(x), \nonumber\\ &j^l(x)=\bar\psi(x)\gamma^l\psi(x) +\frac{a}{2} \left[ \bar\psi(x)\gamma^l\nabla_l\psi(x)+\bar\psi(x)\overleftarrow{\nabla}_l\gamma^l\psi(x) \right] +\frac{iba}{2} \left[ \bar\psi(x)\nabla_l\psi(x)-\bar\psi(x)\overleftarrow{\nabla}_l\psi(x) \right], \label{WilDirCur}$$ where repeated $l$ in the middle term on the rhs of the second equation is not summed. We have also defined $$\begin{aligned} 1 a\bar\psi(x)\overleftarrow{\nabla}_l\equiv\bar\psi(x+a\hat l)U^\dagger(x)-\bar\psi(x).\end{aligned}$$ While the charge density is the same as that of the continuum theory, the current density includes lattice effects described by the differences. In what follows, we calculate the charge density and current density, separately. Charge density in the continuum limit {#s:ChaDen} ------------------------------------- The computation of the charge density $j^0(x)$ is much simpler than that of the current density, since we have regarded $t$ as continuous variable. Therefore, let us first start with the charge density, $$\begin{aligned} 1 \langle j^0(x)\rangle&= \langle0|\bar\psi(x)\gamma^0\psi(x)|0\rangle \nonumber\\ &= \lim_{x'\rightarrow x}(-){{\rm tr}\,}\langle0|T\gamma^0\psi(x)\bar\psi(x')|0\rangle .\end{aligned}$$ This can be calculated in a way similar to the continuum model. Namely, substituting the propagator (\[WilDirPro\]) and inserting the plane-wave representation of the $\delta$ function, we have $$\begin{aligned} 1 \langle j^0(x)\rangle &= \lim_{x'\rightarrow x}{{\rm tr}\,}\gamma^0 \frac{1}{i\slashed D^{\rm L}-\frac{m}{a}e^{i\gamma_5\theta}-\frac{b}{2}a\Delta^{\rm L} +i\epsilon}\frac{-i}{a^{d-1}}\delta(t-t')\delta_{\bm x,\bm x'} \nonumber\\ &=\frac{1}{a^{d-1}}\int_{-\infty}^\infty\frac{d\omega}{2\pi i}\int_{-\pi}^\pi\frac{d^{d-1}k}{(2\pi)^{d-1}} e^{-i(\omega t-\frac{\bm k\cdot \bm x}{a})} {{\rm tr}\,}\gamma^0 \frac{1}{i\slashed D^{\rm L}-\frac{m}{a}e^{i\gamma_5\theta}-\frac{b}{2}a\Delta^{\rm L}+i\epsilon} e^{i(\omega t-\frac{\bm k\cdot \bm x}{a})} \nonumber\\ &=\frac{1}{a^{d-1}} \int_{-\infty}^\infty\frac{d\omega}{2\pi i} \int\frac{d^{d-1}k}{(2\pi)^{d-1}} e^{-i\frac{kx}{a}} {{\rm tr}\,}\gamma^0 \frac{1}{ia\slashed D^{\rm L}-me^{i\gamma_5\theta}-\frac{b}{2}a^2\Delta^{\rm L}+i\epsilon} e^{i\frac{kx}{a}}, \label{J0_1}$$ where in the last line, we have rescaled $\omega\rightarrow \omega/a$, and $kx$ stands for the abbreviation of $\omega t-\bm k\cdot \bm x$. We will carry out the above integral in the limit $a\rightarrow 0$, implying the large mass limit $m/a\rightarrow \infty$ in the continuum model. Notice that $$\begin{aligned} 1 &e^{-i\frac{kx}{a}} a\nabla_j e^{i\frac{kx}{a}}=e^{-ik_j}a\nabla_j+e^{-ik_j}-1, \nonumber\\ &e^{-i\frac{kx}{a}} a\nabla_j^* e^{i\frac{kx}{a}}=e^{ik_j}a\nabla_j^*-e^{ik_j}+1,\end{aligned}$$ where $j$ denotes the spatial direction. Therefore, in the limit $a\rightarrow 0$, the difference becomes $$\begin{aligned} 1 e^{-i\frac{kx}{a}} aD^{\rm L}_j e^{i\frac{kx}{a}} &=-i\sin k_j+\cos k_j aD_j +O(a^2) \nonumber\\ &\equiv -is_j+a\widetilde D_j +O(a^2), \label{DiraOpeCon_1}\end{aligned}$$ where $D_j$ in the first line is the covariant derivative in the continuum limit in Sec. \[s:D\], and in the second line, the abbreviations $s_j$ (and $c_j$) mean $s_j\equiv \sin k_j$ (and $c_j=\cos k_j$), and repeated $j$ in the definition of $\widetilde D_j=c_jD_j$ is not summed. As for the time component, we simply have $$\begin{aligned} 1 e^{-i\frac{kx}{a}} aD^{\rm L}_0 e^{i\frac{kx}{a}}&=i\omega+aD_0 \equiv is_0+a\widetilde D_0. \label{DiraOpeCon_2}\end{aligned}$$ Thus, the differences $D_\mu^{\rm L}$ in Eqs. (\[DiraOpeCon\_1\]) and (\[DiraOpeCon\_2\]) are summarized as $$\begin{aligned} 1 e^{-i\frac{kx}{a}} aD^{\rm L}_\mu e^{i\frac{kx}{a}}&=is_\mu+a\widetilde D_\mu +O(a^2), \label{aLDt}\end{aligned}$$ where $s_\mu=(\omega,-s_j)$, and $\widetilde D_\mu=c_\mu D_\mu$ (no summation over $\mu$) with $c_\mu=(1,c_j)$. Likewise, the Laplacian on the lattice becomes $$\begin{aligned} 1 e^{-i\frac{k\cdot x}{a}} a^2\Delta^{\rm L} e^{i\frac{k\cdot x}{a}} &=2\sum_{j=1}^{d-1}\left(\cos k_j-1-i\sin k_j aD_j\right) +O(a^2) \nonumber\\ &=2\left(c_{\rm s}-ia\widetilde D_{\rm s}\right)+O(a^2), \label{LLapt}\end{aligned}$$ where $c_{\rm s}=\sum_{j=1}^{d-1} (c_j-1)$ and $\widetilde D_{\rm s}\equiv\sum_{j=1}^{d-1} s_jD_j$. Using these, the propagator in Eq. (\[J0\_1\]) can be written as $$\begin{aligned} 1 e^{-i\frac{kx}{a}}& \frac{1} {ia\slashed D^L-me^{i\gamma_5\theta}-\frac{b}{2}a^2\Delta^L} e^{i\frac{kx}{a}} =\frac{1}{i(i\slashed{s}+a\slashed{\widetilde D}) -me^{i\gamma_5\theta}-b(c_{\rm s}-ia\widetilde D_{\rm s})} \nonumber\\ &=\left\{-i(i\slashed{s}+a\widetilde{\slashed D})-me^{-i\gamma_5\theta} -b(c_{\rm s}-ia\widetilde D_{\rm s})\right\} \nonumber\\ &\qquad\qquad\times\frac{1} { \mu^2-s^2+me^{i\gamma_5\theta}\gamma_5a \tilde{\slashed{\partial}}\theta -mbe^{-i\gamma_5\theta}\gamma_5a\tilde\partial_{\rm s}\theta -\frac{i\gamma^\rho\gamma^\sigma}{2}ea^2\widetilde F_{\rho\sigma} +ib\gamma^\rho ea^2\widetilde F_{\rho,{\rm s}}+\tilde{\mathcal{O}} }, \label{LatProMom}$$ where $s^2=\omega^2-\bm s^2$, $\tilde{\partial}_\mu\equiv c_\mu\partial_\mu$ (no sum over $\mu$), $\tilde\partial_{\rm s}=\sum_{j=1}^{d-1} s_j\partial_j$, and $$\begin{aligned} 1 \mu^2=m^2\sin^2\theta+(m\cos\theta+bc_{\rm s})^2 .\end{aligned}$$ We have also introduced two kinds of the field strength: First, $\widetilde F_{\mu\nu}\equiv c_\mu c_\nu F_{\mu\nu}$ (no sum over $\mu,\nu$) which follows from $$\begin{aligned} 1 [\widetilde D_\mu,\widetilde D_\nu]=-ie\widetilde F_{\mu\nu},\end{aligned}$$ and second, $\widetilde F_{\mu,{\rm s}}\equiv \sum_{j=1}^3c_\mu s_j F_{\mu j}$ (no sum over $\mu$) coming from $$\begin{aligned} 1 [\widetilde D_\mu,\widetilde D_{\rm s}]=-ie\widetilde F_{\mu,{\rm s}},\end{aligned}$$ where $F_{\mu\nu}$ is the field strength of the electro-magnetic field in the continuum model defined in Sec. \[s:D\]. In Eq. (\[LatProMom\]) the other operators without $\gamma$-matrices are simply denoted as $\tilde{\mathcal{O}}\equiv a^2\widetilde D^\mu\widetilde D_\mu+2is^\mu a\widetilde{D}_\mu -b^2(a^2\widetilde{D}_{\rm s}^2+2ic_{\rm s}a\widetilde{D}_{\rm s})-2imb\cos\theta a\widetilde{D}_{\rm s}$. Furthermore, we have ignored the $O(a^2)$ terms in Eqs. (\[aLDt\]) and (\[LLapt\]) since they do not contribute to the charge density in the continuum limit. For sufficiently small $a$, we can expand the denominator on the rhs of Eq. (\[LatProMom\]). The charge density (\[J0\_1\]) then can be written as $$\begin{aligned} 1 \langle j^0(x)\rangle &=\frac{1}{a^{d-1}} \int_{-\infty}^\infty\frac{d\omega}{2\pi i} \int\frac{d^{d-1}k}{(2\pi)^{d-1}}\sum_{n=0}^\infty\frac{1}{(\mu^2-s^2-i\epsilon)^{n+1}} {{\rm tr}\,}\gamma^0 \left(\slashed{s}-me^{-i\gamma_5\theta}-bc_{\rm s}-ia\widetilde{\slashed D}+ia\widetilde D_{\rm s}\right) \nonumber\\ &\times \left( -me^{i\gamma_5\theta}\gamma_5a \tilde{\slashed{\partial}}\theta +mbe^{-i\gamma_5\theta}\gamma_5a\tilde\partial_{\rm s}\theta +\frac{i\gamma^\rho\gamma^\sigma}{2}ea^2\widetilde F_{\rho\sigma} -ib\gamma^\rho ea^2\widetilde F_{\rho,{\rm s}}+\tilde{\mathcal{O}} \right)^{n}. \label{J0_2}$$ This equation for the lattice Wilson-Dirac fermion corresponds to Eq. (\[CurCon3\]) for the continuum Dirac fermion. ### $d=1+1$ system We are interested in the terms with $\gamma_5$ in Eq. (\[J0\_2\]) which survive in the limit $a\rightarrow0$. Due to Eq. (\[TraGam1D\]), it is enough to consider the $n=1$ term in Eq. (\[J0\_2\]). $$\begin{aligned} 1 \langle j^0(x)\rangle &=\frac{1}{a} \int_{-\infty}^\infty\frac{d\omega}{2\pi i}\int_{-\pi}^\pi\frac{dk_1}{2\pi} \frac{ {{\rm tr}\,}\gamma^0\left(\slashed{s}-me^{-i\gamma_5\theta}-bc_{\rm s}\right)(-ma\gamma_5) \left(e^{i\gamma_5\theta} \tilde{\slashed{\partial}}\theta -be^{-i\gamma_5\theta}\tilde\partial_{\rm s}\theta \right) } {(\mu^2-s^2-i\epsilon)^2} . \label{ChaDen1Ori}$$ It turns out that the trace in the numerator of the above equation yields $$\begin{aligned} 1 (-ma)\left[ {{\rm tr}\,}\gamma_5\gamma^0\left(me^{-i\gamma_5\theta}+bc_{\rm s}\right) e^{i\gamma_5\theta}\tilde{\slashed{\partial}}\theta -b{{\rm tr}\,}\gamma_5\gamma^0\slashed se^{-i\gamma_5\theta}\tilde\partial_{\rm s}\theta) \right] =2am\left[(m+bc_{\rm s}\cos\theta)c_1+b\cos\theta s_1^2\right]\epsilon^{01}\partial_1\theta .\end{aligned}$$ The denominator becomes $\mu^2-s^2=\mu^2+\bm s^2-\omega^2\equiv \Omega^2(k_1,\theta)-\omega^2$, where we have introduced a generic expression for later convenience, $$\begin{aligned} 1 \Omega^2(\bm k,\theta)&\equiv\bm s^2+\mu^2 \nonumber\\ & =\sum_{j=1}^{d-1}s_j^2+m^2\sin^2\theta+\left[m\cos\theta+b\sum_{j=1}^{d-1}(c_j-1)\right]^2. \label{OmeD}$$ Note $d=2$ in the present 1D system. Then, using the integration over $\omega$ in Eq. (\[IntOme1\]), we finally obtain $$\begin{aligned} 1 \langle j^\mu(x)\rangle=G_1(\theta)\epsilon^{\mu\nu}\partial_\nu\theta, \label{ChaDen1}$$ where we have derived the $\mu=0$ charge density in this subsection, although Eq. (\[ChaDen1\]) is valid for the $\mu=1$ current density, as we will show in Appendix \[s:CurDen\], and we have introduced $$\begin{aligned} 1 &G_{d-1}(\theta)=N_{d-1}\int_{-\pi}^\pi d^{d-1}k \frac{\Theta(\bm k,\theta)} {\Omega^{d+1}(\bm k,\theta)}, \nonumber\\ &\Theta(\bm k,\theta)\equiv m\left[m+b\sum_{j=1}^{d-1}(\sec k_j-1)\cos\theta\right]\prod_{j=1}^{d-1}\cos k_j, \label{Gd}$$ with $N_1=1/(4\pi)$. It should be noted that $G_{d-1}(\theta)$ is independent of the electro-magnetic field, and furthermore, it depends on $x$ only through $\theta$. Thus, $$\begin{aligned} 1 \int_0^{2\pi}G_1(\theta)d\theta=c_1 \label{FirChe}\end{aligned}$$ does not depend on $x$ any longer. This is nothing but the first Chern number, which can also be computed using the wave functions of the Wilson-Dirac Hamiltonian (\[WDHam\]). See Sec. \[s:E\]. In the case of $\theta=\theta(x)$, Eq. (\[ChaDen1\]) as well as Eq. (\[DCur1\]) have attracted much interest as the topological number on solitons in quantum field theory.[@Goldstone:1981aa; @Qi:2008aa] On the other hand, when $\theta=\theta(t)$, we nowadays know that they describe the topological pump, which is of current interest. Both phenomena mentioned above are related with the anomaly in two dimensional Dirac fermions, as will be discussed in Sec. \[s:sd\]. ### $d=3+1$ system We are also interested in the terms with $\gamma_5$ in Eq. (\[J0\_2\]) which survive in the limit $a\rightarrow0$. Considering Eq. (\[TraGam3D\]), it is enough to take the $n=2$ term in Eq. (\[J0\_2\]) into account. $$\begin{aligned} 1 \langle j^0(x)\rangle &=\frac{1}{a^3} \int_{-\infty}^\infty\frac{d\omega}{2\pi i} \int\frac{d^3k}{(2\pi)^3} \frac{1}{(\mu^2-s^2-i\epsilon)^3} \nonumber\\ &\times {{\rm tr}\,}\gamma^0\left(\slashed{s}-me^{-i\gamma_5\theta}-bc_{\rm s}\right) \left( -me^{i\gamma_5\theta}\gamma_5a \tilde{\slashed{\partial}}\theta +mbe^{-i\gamma_5\theta}\gamma_5a\tilde\partial_{\rm s}\theta +\frac{i}{2}\gamma^\rho\gamma^\sigma ea^2\widetilde F_{\rho\sigma} -ib\gamma^\rho ea^2\widetilde F_{\rho,{\rm s}} \right)^2 . \label{J0_3}\end{aligned}$$ Among various terms in the above equation associated with the trace of the $\gamma$ matrices, the product terms between $\partial\theta$ and $\widetilde F$ give finite contributions, which have indeed $a^3$, $$\begin{aligned} 1 ea^3\Big\{& {{\rm tr}\,}\gamma^0(-me^{-i\gamma_5\theta}-bc_{\rm s})(-me^{i\gamma_5\theta}\gamma_5a\tilde{\slashed\partial}\theta) (i\gamma^\rho\gamma^\sigma \widetilde F_{\rho\sigma}) +{{\rm tr}\,}\gamma^0\slashed smbe^{-i\gamma_5\theta}\gamma_5\tilde\partial_{\rm s}\theta (i\gamma^\rho\gamma^\sigma \widetilde F_{\rho\sigma}) \nonumber\\ &+2{{\rm tr}\,}\gamma^0\slashed s(-me^{i\gamma_5\theta}\gamma_5\tilde{\slashed\partial}\theta) (-ib\gamma^\rho\widetilde F_{\rho,{\rm s}})\Big\} \nonumber\\ &=-mea^3\Big\{ i(m+bc_{\rm s}\cos\theta){{\rm tr}\,}\gamma_5\gamma^0 (\tilde{\slashed\partial}\theta) \gamma^\rho\gamma^\sigma \widetilde F_{\rho\sigma} -ib\cos\theta{{\rm tr}\,}\gamma_5\gamma^0\slashed s\tilde\partial_{\rm s}\theta \gamma^\rho\gamma^\sigma \widetilde F_{\rho\sigma} -2ib\cos\theta{{\rm tr}\,}\gamma_5\gamma^0\slashed s\tilde{\slashed\partial}\theta \gamma^\rho\widetilde F_{\rho,{\rm s}}\Big\} \nonumber\\ &=-4ea^3m\left[m+b\sum_{j=1}^3(\sec k_j-1)\cos\theta\right]\left(\prod_{j=1}^3\cos k_j\right) \epsilon^{0\nu\rho\sigma}(\partial_\nu\theta) F_{\rho\sigma}.\end{aligned}$$ Thus, we finally reach $$\begin{aligned} 1 \langle j^\mu(x)\rangle=-e\frac{G_3(\theta)}{4\pi}\epsilon^{\mu\nu\rho\sigma}\partial_\nu\theta(x) F_{\rho\sigma}(x), \label{ChaDen3}$$ where we have derived the $\mu=0$ charge density in this subsection, although Eq. (\[ChaDen3\]) is valid for the $\mu=1,2,3$ current density, as we will show in Appendix \[s:CurDen\]. $G_3(\theta)$ is defined by Eq. (\[Gd\]) with $N_3=3/(8\pi^2)$, which follows from Eq. (\[IntOme2\]). Note that integration of $G_3(\theta)$ over $\theta$ gives the second Chern number $$\begin{aligned} 1 \int_0^{2\pi}G_3(\theta)d\theta=c_2. \label{SecChe}$$ It should be stressed here that $G_3(\theta)$ does not depends on the electro-magnetic field. The limit $a\rightarrow 0$ implies, therefore, small field limit as well. The Chern number $c_2$ is alternatively calculated using the Berry curvature with respect to the eigenfunctions of the Wilson-Dirac Hamiltonian with zero fields. See Ref. [@Qi:2008aa] or [@Fukui:2016aa], and also Sec. \[s:E\]. Equation (\[ChaDen3\]) or its continuum version (\[DCur3\_1\]) show the topological magneto-electric effects, including CME and AHE. For Weyl semimetals, $\theta$ is induced by the Zeeman term violating time reversal symmetry and/or energy imbalance between the Weyl nodes breaking the inversion symmetry. In contrast, for the present Wilson-Dirac fermion, $\theta$ describes a rotation between two kinds of mass terms which keeps a finite mass gap in the spectrum. Regardless of whether the system is massless or massive, the magneto-electric effects are associated with the chiral anomaly, as will be discussed in Sec. \[s:sd\]. $k_1$ $k_2$ $\qquad \xi_3\Omega\qquad$ $\Theta(k_1,\theta)$ ------- ------- ---------------------------- ---------------------- $0$ $0$ $m$ $m^2$ $\pi$ $0$ $m-2b$ $-m(m-2b)$ $0$ $\pi$ $-m$ $m^2$ $\pi$ $\pi$ $-m-2b$ $-m(m+2b)$ : List of $\xi_3\Omega=m\cos k_2+b(\cos k_1-1)$ and $\Theta(k_1,\theta)=m[m+b(\sec k_1-1)\cos k_2]\cos k_1$ at $k_\mu=0$ or $\pi$. []{data-label="t:1"} Chern numbers {#s:E} ------------- For the study of the temporal pump in the next section \[s:MEP\], we need an explicit value of the Chern number. Therefore, we calculate the first and second Chern numbers in Eqs. (\[FirChe\]) and (\[SecChe\]). Fortunately, the Wilson-Dirac model is so simple that one can calculate the second Chern number exactly. ### First Chern number We can regard $k_1$ and $k_2\equiv\theta$ as the coordinates of a two-dimensional torus T$^2$. A mapping $f$ from T$^2$ to S$^2$ can be defined by $$\begin{aligned} 1 &\xi_1=\frac{\sin k_1}{\Omega(k_1,\theta)},\quad \xi_2=\frac{m\sin\theta}{\Omega(k_1,\theta)}, \nonumber\\ &\xi_3=\frac{m\cos\theta+b(\cos k_1-1)}{\Omega(k_1,\theta)},\end{aligned}$$ where $\Omega(k_1,\theta)$ is given by Eq. (\[OmeD\]) for $d=2$. It is known that the Chern number is the degree of the mapping which can be computed by $$\begin{aligned} 1 \mbox{deg }f=\frac{1}{\mbox{Vol(S}^2)}\int\frac{1}{2!}\epsilon^{\alpha\beta\gamma}\xi_\alpha d\xi_\beta d\xi_\gamma, \label{Degf1}\end{aligned}$$ where $\mbox{Vol(S}^2)=4\pi$ and $d\xi_\alpha$ is the coordinate differential 1-form. Indeed, it is not difficult to rewrite Eq. (\[Degf1\]) by $k_\mu$ ($\mu=1,2$) to show $\mbox{deg\:}f=c_1$, where $c_1$ is given by the lhs of Eq. (\[FirChe\]) with $G_1(\theta)$ in Eq. (\[Gd\]). It is readily seen that only the points $k_\mu=0$ or $\pi$ on T$^2$ are mapped to $\xi_\pm=(0,0,\pm1)$ on S$^2$. The degree of the mapping $f$ is then given by $$\begin{aligned} 1 \mbox{deg }f=&\sum_{k\in f^{-1}(\xi_+)} \mbox{sgn}\left[\Theta(k_1,\theta)\right], $$ where $\Theta(k_1,\theta)$ is defined in Eq. (\[Gd\]) for $d=2$. From Table \[t:1\], we can read the coordinates $k_\mu$ mapped to $\xi_+$ with its degree of mapping (or winding number) $\pm1$. For example, when $0<m<2b$, only $(0,0)$ is mapped to $\xi_+$ with a positive winding number $m^2>0$, whereas others $(0,\pi)$, $(\pi,0)$, and $(\pi,\pi)$ are mapped to $\xi_-$. Thus, we find $c_1=1$ in this case. The other cases are likewise. Thus, Table \[t:1\] leads to $$\begin{aligned} 1 \mbox{deg }f=c_1= \left\{ \begin{array}{rl} +1&(0<|m|<2b)\\ 0&(2b<|m|) \end{array} \right. .\end{aligned}$$ ($k_1$, $k_2$, $k_3$) $k_4$ $\qquad \xi_5\Omega\qquad$ $\Theta(\bm k,\theta)$ ----------------------- ------- ---------------------------- ------------------------ no $\pi$ $0$ $m$ $m^2$ one $\pi$ $0$ $m-2b$ $-m(m-2b)$ two $\pi$ $0$ $m-4b$ $m(m-4b)$ three $\pi$ $0$ $m-6b$ $-m(m-6b)$ no $\pi$ $\pi$ $-m$ $m^2$ one $\pi$ $\pi$ $-m-2b$ $-m(m+2b)$ two $\pi$ $\pi$ $-m-4b$ $m(m+4b)$ three $\pi$ $\pi$ $-m-6b$ $-m(m+6b)$ : List of $\xi_5\Omega=m\cos k_4+b\sum_j^3(\cos k_j-1)$ and $\Theta(\bm k,\theta)=m[m+b\sum_j^3(\sec k_j-1)\cos k_4]\prod_j^3\cos k_j$ at $k_\mu=0$ or $\pi$. “no $\pi$” in the first column means all $k_\mu=0$, whereas “one $\pi$" means that one of $k_\mu=\pi$ and others are $0$, and so on. []{data-label="t:2"} ### Second Chern number The second Chern number can be obtained by extending the computation of the first Chern number for the S$^2$ to S$^4$. Let us introduce $\xi_\mu$ ($\mu=1,\cdots,5$) by $$\begin{aligned} 1 &\xi_j=\frac{\sin k_j}{\Omega(\bm k,\theta)},\quad (j=1,2,3),\quad \xi_4=\frac{m\sin\theta}{\Omega(\bm k,\theta)}, \nonumber\\ &\xi_5=\frac{m\cos\theta+b\sum_{l=1}^{3}(\cos k_l-1)}{\Omega(\bm k,\theta)},\end{aligned}$$ where $\Omega(\bm k,\theta)$ is defined in Eq. (\[OmeD\]) for $d=4$. These define a mapping $f$ from T$^4$ spanned by $(k_1,k_2,k_3,\theta\equiv k_4)$ to S$^4$ spanned by $\xi_\mu$ satisfying $\xi_\mu^2=1$. The Chern number is the degree of the mapping given by $$\begin{aligned} 1 \mbox{deg }f=\frac{1}{\mbox{Vol(S}^4)}\int\frac{1}{4!}\epsilon^{\alpha\beta\gamma\delta\epsilon}\xi_\alpha d\xi_\beta d\xi_\gamma d\xi_\delta d\xi_\epsilon, \label{Degf}\end{aligned}$$ where Vol(S$^4$)$=8\pi^2/3$. It is straightforward to rewrite Eq. (\[Degf\]) by $k_j$ and $\theta$ to show $\mbox{deg }f=c_2 $, where $c_2$ is given by the lhs of Eq. (\[SecChe\]) with $G_3(\theta)$ in Eq. (\[Gd\]). It is readily seen that the points $k_\mu=0$ or $\pi$ on T$^4$ are mapped to $\xi_\pm=(0,0,0,0,\pm1)$ on S$^4$. The degree of the mapping $f$ is then given by $$\begin{aligned} 1 \mbox{deg }f=&\sum_{k\in f^{-1}(\xi_+)} \mbox{sgn}\left[\Theta(\bm k,\theta)\right] ,\end{aligned}$$ where $\Theta(\bm k,\theta)$ is defined in Eq. (\[Gd\]) for $d=4$. Thus, the second Chern number can be obtained from Table \[t:2\] in the similar way to the first Chern number: $$\begin{aligned} 1 \mbox{deg }f=c_2= \left\{ \begin{array}{rl} +1&(0<|m|<2b)\\ -2\quad&(2b<|m|<4b)\\ +1&(4b<|m|<6b)\\ 0&(6b<|m|)\\ \end{array} \right. . \label{ExpCheNum2}$$ Magneto-electric pump {#s:MEP} ===================== In Secs. \[s:ChaDen\] and \[s:CurDen\], we have established the U(1) current (\[ChaDen3\]) in the 3D Wilson-Dirac model. This result is obtained in the limit $a\rightarrow0$. It also implies that the result is valid only in a weak field limit. On the other hand, the pump is topological so that the result may be valid as long as the mass gap is open. To check this point, we study the pump by numerical calculations. In this section, we restrict our discussions to the 3D temporal particle pump. 3D pump {#s:3dpump} ------- Assume that the electro-magnetic field is static, and consider the case in which $\theta$ depends only on $t$, $\theta=\theta(t)$ with a period $T$, $\theta(t+T)=\theta(t)$. When $1/T\ll m$, we can regard the process of changing $t$ as an adiabatic process. Then, integration of Eq. (\[ChaDen3\]) over $t$ in one period yields the [*pumped*]{} particle density $$\begin{aligned} 1 \bm q(x)=e\frac{c_2}{2\pi}\bm B(x),\end{aligned}$$ where $c_2$ is the second Chern number defined in Eq. (\[SecChe\]) and explicitly given by (\[ExpCheNum2\]). We have stressed there that the second Chern number (\[SecChe\]) is that of the Wilson-Dirac model in zero field limit. Namely, the Chern number can be computed directly using the eigenfunctions of the Hamiltonian without the magnetic field. Without loss of generality, we assume the magnetic field in the $z$ direction, $(0,0,B)$. The total pumped particle number $Q^z$ can be defined by integrating over the $xy$ surface, $$\begin{aligned} 1 Q^z=ec_2\frac{\Phi}{2\pi}, \label{Qz1}$$ where $\Phi$ is the total flux penetrating the surface of an area under consideration. This integration can be simply carried out, since $c_2$ is independent of $B$, as stressed already. For numerical computations, we further consider a simpler case where the magnetic field $B$ is uniform, and system is periodic in $xy$ surface whose size in the $x$ ($y$) direction is $N_x$ ($N_y$). Then, $\Phi=B a^2N_xN_y$, and the flux per plaquette should be rational $$\begin{aligned} 1 |e|\phi=|e|Ba^2=\frac{2\pi p}{q}\equiv \phi_0, \label{FluPla}$$ where $p$ and $q (>0)$ are integers describing a rational magnetic flux per plaquette. [@Hofstadter:1976aa; @Thouless:1982uq; @kohmoto:85] Thus, it turns out that the total pumped particle number $Q^z$ is given by $$\begin{aligned} 1 Q^z=\mbox{sgn(}e) c_2\frac{Ba^2N_xN_y}{2\pi}=\mbox{sgn(}e)c_2\frac{p}{q} N_xN_y. \label{TotPumCha2}$$ The elementary pump is due to a nonzero second Chern number, and $pN_xN_y/q$ can be considered as the geometrical multiplicity of pumped particles associated with the magnetic flux. 3D pump as a set of 1D pump {#s:1dpump} --------------------------- The magneto-electric pump discussed so far has deep relationship with the chiral anomaly, so that the $3+1$ dimensionality plays a crucial role. However, the pumping itself is toward one direction of the applied magnetic field, and hence, it can also be viewed as a 1D pump. In this section, we derive the pumped particle number by the Thouless formula for the 1D pump.[@Thouless:1983fk; @Xiao:2010fk] Consider the snapshot (for a fixed $t$, i.e., fixed $\theta$) many-body ground state eigenfunction obeying $H{|\Psi_n\rangle}=E_n{|\Psi_n\rangle}$. We assume that ${|\Psi_0\rangle}$ is the ground state and have a spectral gap $E_n(t)-E_0(t)>0$ at any time during the pump. Then, the ground state ${|G\rangle}$ which satisfies the time-dependent Schrödinger equation at the first order of $\hbar$ is given by $$\begin{aligned} 1 {|G\rangle}=e^{-\frac{i}{\hbar}\int^tE_0 dt}\left({|\Psi_0\rangle} +i\hbar\sum_{n\ne0}\frac{{|\Psi_n\rangle}{\langle\Psi_n|\dot\Psi_0\rangle}}{E_n-E_0}\right). \label{PerG}$$ For non-interacting case, it is enough to consider the single particle states. Let ${\cal H}(\theta,\bm k;\bm B)$ be the Fourier-transformed Hamiltonian given by Eq. (\[WDHam\]), and let $\psi(\theta,\bm k;\bm B)$ be the ground state (i.e., negative energy) multiplet wave functions of single-particle states satisfying $$\begin{aligned} 1 {\cal H}(\theta,\bm k;\bm B)\psi(\theta,\bm k;\bm B)=\psi(\theta,\bm k;\bm B){\cal E}(\theta,\bm k;\bm B),\end{aligned}$$ where ${\cal E}(\theta,\bm k,\bm B)$ is the diagonal matrix of the energy eigenvalues. Let us now consider the case where uniform magnetic field is applied in the $z$ direction, as studied in Sec. \[s:3dpump\]. Regarding $k_x$ and $k_y$ as parameters, the current toward the $z$ direction, $J^z$, with respect to the state Eq. (\[PerG\]) is given by [@Thouless:1983fk] $$\begin{aligned} 1 \langle J^z\rangle= \int_0^{2\pi}\frac{dk_z}{2\pi i}f(\theta,k_z;k_x,k_y,B), $$ where $f(\theta,k_z;k_x,k_y,B)\equiv \partial_{k_z}\psi^\dagger\partial_{\theta}\psi-\partial_\theta\psi^\dagger\partial_{k_z}\psi$ is the Berry curvature with respect to $\theta$ and $k_z$ with fixed $k_x$ and $k_y$. Integrating $\langle J^z\rangle$ with respect to $t$ in one period as well as with respect to $k_x$ and $k_y$, we obtain the total number of pumped particle, $$\begin{aligned} 1 Q^z=\sum_{k_x,k_y}c_1(B), \label{1ChaNum}$$ where $$\begin{aligned} 1 c_1(B)=\frac{1}{2\pi i}\int_0^{2\pi}d\theta \int_0^{2\pi}dk_zf(\theta,k_z;k_x,k_y,B), \label{SecFirCheNum}$$ is the first Chern number on the section specified by fixed $k_x,k_y$. However, it should be noted that the Chern number $c_1(B)$ does depend on the applied magnetic field, but does not depend on $k_x$ ($\tilde k_x$) and $k_y$, provided that the mass gap always opens. When we compute the eigenfunctions in the Landau gauge $U_j(t,\bm x)=e^{-ieaA_j(t,\bm x)}$ with $$\begin{aligned} 1 e\bm A(\bm x)=(0,\mbox{sgn(}e) \phi\frac{x}{a},0), \label{LanGau}\end{aligned}$$ we can take the $q$ sites in the $x$-direction as a unit cell, and therefore, we set $N_x=q\tilde N_x$ for the periodic boundary condition. In this case, $k_y=2\pi n_y/N_y$ with $n_y=1,\cdots,N_y$ , whereas $\tilde k_x=2\pi n_x/\widetilde N_x$ with $n_x=1,\cdots,\widetilde N_x$. It follows from Eq. (\[1ChaNum\]) and from the fact that $c_1(B)$ is independent of $k_x$ and $k_y$ that $$\begin{aligned} 1 Q^z=c_1(B,\mbox{sgn}(e))\tilde N_xN_y, \label{Qz2}$$ where in Eq. (\[1ChaNum\]) $k_x$ is replaced by $\tilde k_x$, and the dependence of $c_1$ on the sign of the charge $e$ through Eq. (\[LanGau\]) has been explicitly denoted. Comparing Eq. (\[TotPumCha2\]), we have a simple relationship between the second Chern number $c_2$ in the zero magnetic field and $c_1$ in the magnetic field $B$, $$\begin{aligned} 1 c_1(B,\mbox{sgn}(e))=\mbox{sgn}(e) c_2 p, \label{RelC1C2}$$ where $p$ is given by Eq. (\[FluPla\]). This relation may be useful for computing the second Chern number, since the numerical method of computing the first Chern number has already been established.[@FHS05] -------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ![ The Hofstadter butterfly diagrams for fixed $\theta=0$ and for $m=b$ (upper) and $m=3b$ (lower). We set $q=50$, and the system size is $\widetilde N_x=1$ and $N_y=N_z=50(=q)$. []{data-label="f:but"}](butterfly_m_1.eps "fig:") ![ The Hofstadter butterfly diagrams for fixed $\theta=0$ and for $m=b$ (upper) and $m=3b$ (lower). We set $q=50$, and the system size is $\widetilde N_x=1$ and $N_y=N_z=50(=q)$. []{data-label="f:but"}](butterfly_m_3.eps "fig:") -------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ---------------- ------- ---------------- ------- --------------- ------- ---------------- ------- ---------------- ------- --------------- ------- $\frac{p}{q}$ $c_1$ $\frac{p}{q}$ $c_1$ $\frac{p}{q}$ $c_1$ $\frac{p}{q}$ $c_1$ $\frac{p}{q}$ $c_1$ $\frac{p}{q}$ $c_1$ $\frac{1}{20}$ 1 $\frac{1}{20}$ $-2$ $\frac{2}{20}$ 2 $\frac{1}{10}$ 1 $\frac{2}{20}$ $-4$ $\frac{1}{10}$ $-2$ $\frac{3}{20}$ 3 $\frac{3}{20}$ $-6$ $\frac{4}{20}$ 4 $\frac{2}{10}$ 2 $\frac{1}{5}$ 1 $\frac{4}{20}$ $-8$ $\frac{2}{10}$ $-4$ $\frac{1}{5}$ $-2$ ---------------- ------- ---------------- ------- --------------- ------- ---------------- ------- ---------------- ------- --------------- ------- : The (section) first Chern number (\[SecFirCheNum\]) computed on the discretized $(\theta,k_z)$ Brillouin zone [@FHS05] in the case $e<0$ for several $\frac{p}{q}=\frac{\phi_0}{2\pi}$. []{data-label="t:3"} We show in Table \[t:3\] the list of the section Chern number $c_1(B)$ in (\[SecFirCheNum\]) for various $\phi_0$, and in Fig. \[f:but\] the corresponding Hofstadter butterfly diagrams.[@Hofstadter:1976aa] The Hofstadter diagrams tell that the mass gap at $\phi_0=0$ becomes smaller as a function of $\phi_0$, and eventually closed around $\phi_0\sim \pi/2$ for both cases $m=b$ and $m=3b$. Thus, it turns out from the Table \[t:3\] that the relationship (\[RelC1C2\]) is indeed valid as long as the mass gap is finite. This implies the absence of the higher order corrections for the chiral anomaly. [@Adler:1969ab] ------------------------- ------------------------- ![image](spct_m0_1.eps) ![image](spct_m0_3.eps) ![image](ncom_m0_1.eps) ![image](ncom_m0_3.eps) ------------------------- ------------------------- Bulk-edge correspondence {#s:BEC} ------------------------ So far we have discussed the topological property of the bulk system. In this subsection, we discussed the surface states of the 3D topological pump, considering a system with boundaries. In Ref.[@Hatsugai:2016aa], the bulk-edge correspondence in the 1D Thouless pump has been discussed. Consider the system with boundaries. Then one can define the center of mass of the occupied particles. [@Wang:2013fk_pump] It has been shown that the change of the center of mass is just the number of pumped particle. [@Wang:2013fk_pump] Consider here a system coupled with a particle reservoir. Then, the center of mass as a function of time shows discontinuity due to sudden change of the ground state when the chemical potential crosses the edge states.[@Hatsugai:2016aa] After one period, the center of mass returns to the initial value. This implies that the amount of pumped particles in the bulk are compensated by these discontinuities. Thus, from the discontinuities, one can know the number of the pumped particle.[@Hatsugai:2016aa] Consider the Wilson-Dirac Hamiltonian with bottom and top surfaces, labeled by $j_z=0$ and $j_z=N_z$, respectively, perpendicular to the $z$ axis. Let $\psi^{n}(\theta,\tilde k_x,k_y;B)$ be the $n$th normalized eigenfunction of the Hamiltonian with the boundaries ${\cal H}(\theta,\tilde k_x,k_y;B)$, where we have assumed the Landau gauge in Sec. \[s:1dpump\]. Then, we can define the normalized center of mass of the ground state along the $z$ axis as $$\begin{aligned} 1 &p_z(\theta)=\frac{1}{p\widetilde N_x N_y} \sum_{n~{\rm occ.}}\sum_{\tilde k_x,k_y}\sum_{j_z=0}^{N_z}\left(\frac{j_z}{N_z}-\frac{1}{2}\right)|\psi_{j_z}^{n}(\theta,\tilde k_x.k_y)|^2,\end{aligned}$$ where the normalization factor is due to the multiplicity associated with the total flux in Eq. (\[TotPumCha2\]). It follows from Eq. (\[TotPumCha2\]) that the normalized center of mass gives the second Chern number $$\begin{aligned} 1 c_2=-\mbox{sgn}(e) \left[\mbox{ sum of discontinuities of }p_z(\theta)\right].\end{aligned}$$ In Fig. \[f:com\], we show the spectral flow as a function of $\theta$ and the normalized center of mass $p_z(\theta)$. In both cases in Fig. \[f:com\], the structure of the vacuum (negative energy states) changes at $\theta=0$, and shows the discontinuity in the center of mass, from which the second Chern number can be obtained. The result is consistent with Eq. (\[ExpCheNum2\]). ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ![ Upper figure shows the Hofstadter diagram of the Hamiltonian (\[ModWDHam\]) as a function of the magnetic flux per plaquette (\[FluPla\]). The parameters are $q=50$, $N_y=N_z=6$, $\tilde q=6$, and $\tilde p=1$ is fixed, whereas $p$ is changed by $\Delta p=1$. The model parameter is $m=3b$. Lower figure shows the difference of the number of the negative energy states defined in the rhs of Eq. (\[StrFor\]). []{data-label="f:asy"}](butterfly_4d_m0_3.eps "fig:") ![ Upper figure shows the Hofstadter diagram of the Hamiltonian (\[ModWDHam\]) as a function of the magnetic flux per plaquette (\[FluPla\]). The parameters are $q=50$, $N_y=N_z=6$, $\tilde q=6$, and $\tilde p=1$ is fixed, whereas $p$ is changed by $\Delta p=1$. The model parameter is $m=3b$. Lower figure shows the difference of the number of the negative energy states defined in the rhs of Eq. (\[StrFor\]). []{data-label="f:asy"}](spct_asy_m0_3.eps "fig:") ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- Generalized Streda formula {#s:Streda} -------------------------- For reference, we here present a method of computing the second Chern number based on the generalized Streda formula.[@Fukui:2016aa] Without an electro-magnetic field, the Wilson-Dirac Hamiltonian (\[WDHam\]) in the momentum space reads $$\begin{aligned} 1 {\cal H}(k)=&\Gamma^j\sin k_j+m\Gamma^4\sin k_4 \nonumber\\ &-\Gamma_5\left[m\cos k_4+b\sum_{j=1}^3(\cos k_j-1)\right],\end{aligned}$$ where we have introduced $k_4=\theta$, and new hermitian $\Gamma$ matrices $\Gamma^j=\alpha^j$ ($j=1,2,3$), $\Gamma^4=-i\beta\gamma_5$, and $\Gamma^5=\beta$, with $\{\Gamma^\mu,\Gamma^\nu\}=2\delta^{\mu\nu}$. Now let us regard $k_4$ as the frequency of discrete imaginary time. Then, we can reconstruct an equivalent lattice fermion such that $$\begin{aligned} 1 {\cal H}=&\Gamma^j D_j^{\rm L}+m\Gamma^4 D_4^{\rm L} -\Gamma^5\left[\frac{m}{2}(\Delta_4^{\rm L}+2)+\frac{ba}{2}\Delta^{\rm L}\right], \label{ModWDHam}\end{aligned}$$ where $D_4^{\rm L}\equiv (\nabla_4+\nabla_4^*)/2$ and $\Delta_4^{\rm L}=(\nabla_4-\nabla_4^*)/a$ are defined for the new coordinate $x^4$ in the same way as Eqs. (\[LatDirOpe\]) and (\[CovDifOpe\]). Using this four dimensional Hamiltonian, one can define the overlap Dirac operator, [@Neuberger:1998ab; @Neuberger:1998aa] which obeys the Ginsparg-Wilson relation. [@Ginsparg:1982aa] This enables us to define the chiral anomaly on the lattice. [@Luscher:1998aa; @Kikukawa:1999aa; @Suzuki01071999] Taking the continuum limit, it indeed reproduces the chiral anomaly in arbitrary dimensions with a nontrivial Chern number as a coefficient.[@1126-6708-2002-09-025] It has been shown that the chiral anomaly thus obtained is given by the spectral asymmetry of the above Hamiltonian, and hence one can compute the second Chern number from the spectral flow of the Hamiltonian. [@Fukui:2016aa] This is referred to as the generalized Streda formula. In what follows, we use $N_-$ rather than the spectral asymmetry $\eta=(N_+-N_-)/2$, where $N_\pm$ is the number of positive and negative energy states. Consider the system which includes a static and uniform magnetic field in the $x^3=z$ direction, as studied in Secs. \[s:3dpump\], \[s:1dpump\], and \[s:BEC\], and introduce a fictitious electric field associated with the imaginary time $x^4$ direction. As shown in Ref. [@Fukui:2016aa], the density of occupied (negative energy) states $$\begin{aligned} 1 n_-=\frac{1}{2V}{\rm Tr }\left(1-\frac{{\cal H}}{\sqrt{\cal H}^2}\right)=\frac{N_-}{V}\end{aligned}$$ of the Hamiltonian (\[ModWDHam\]) as a function of the magnetic field yields the second Chern number, $$\begin{aligned} 1 \frac{\partial n_-}{\partial (\bm B\cdot\bm E)}=-\frac{c_2}{(2\pi)^2},\end{aligned}$$ where $\bm E$ is the fictitious electric field. To be concrete, assume that the magnetic field is included in the Landau gauge in Sec. \[s:1dpump\]. The fictitious electric field is also included in the $z$ direction such that $$\begin{aligned} 1 e\bm A(\bm x)=(0,\mbox{sgn(}e) \phi\frac{x}{a},\mbox{sgn(}e) \tilde\phi\frac{x^4}{a}), \label{LanGau_2}\end{aligned}$$ where $\tilde \phi=2\pi \tilde p/\tilde q$, gives the fictitious electric field $(0,0,E)$ such that $E=-\partial A_z/\partial x^4=-\mbox{sgn}(e)\tilde\phi/a^2$. For numerical calculations, we set the system size as $q$, $N_y$, $N_z$, and $\tilde q$ sites for $x$, $y$, $z$, and $x^4$ directions, respectively, with periodic boundary conditions imposed in all directions. Then, the volume of the system is $V=qN_yN_z\tilde qa^4$, and therefore, $$\begin{aligned} 1 c_2=-(2\pi)^2\frac{\Delta N_-}{qN_yN_z\tilde qa^4\Delta(BE)}=\frac{\Delta N_-}{N_yN_z\tilde p\Delta p}, \label{StrFor}\end{aligned}$$ where $\Delta p=1$, and we have assumed that the electric field is fixed. In Fig. \[f:asy\], we show the spectrum as a function $\phi_0$, and corresponding difference of the density of the occupied states. This figure tells that the the second Chern number is $c_2=-2$, consistent with the previous results. In passing, we mention that in the case $m=b$, we can also reproduce $c_2=1$ in the same manner. The Hofstadter diagram in Fig. \[f:asy\] is different from that in Fig. \[f:but\], since the former includes a finite fictitious electric field. In other words, such a difference enables us to compute the second Chern number. Without the fictitious electric field, the anomaly is trivially vanishing. Summary and discussion {#s:sd} ====================== In summary, we studied mainly the 3D topological pump analytically and numerically in detail in this paper. We introduced a variant of the Wilson-Dirac model defined on the spatial lattice but in continuous time, including two kinds of mass terms depending generically on $\bm x$ as well as $t$. We derived the conserved current density on the lattice and calculate it in the continuum limit, or in other words in the small field limit. For the temporal pump, the result was checked by numerical calculations from various methods as follows. Firstly, the 3D pump governed by the second Chern number (\[Qz1\]) or (\[TotPumCha2\]) can be viewed as a set of 1D pump described by the first Chern number (\[1ChaNum\]) or (\[Qz2\]). It should be noted that the latter description is valid even in a strong magnetic field as long as the mass gap is open. Both results lead to the relationship between two Chern numbers (\[RelC1C2\]). We showed by the numerical calculation of the first Chern number using the Berry curvature of the wave functions that Eq. (\[RelC1C2\]) is valid in a strong magnetic field regime up to the gap closing point. It would be an interesting problem to ask whether the relationship (\[RelC1C2\]) is restricted only to the present system or more applicable to other cases. Since the second Chern number is generically due to non-Abelian Berry curvature, its numerical calculation is very hard, and therefore, a simple relationship like (\[RelC1C2\]) is quite helpful. Secondly, as the Chern number description of the number of the pumped particles is for the bulk system, the bulk-edge correspondence enables us to observe the 3D pump as the flow of the surface states. We showed that the bulk-edge correspondence established in a 1D pump [@Hatsugai:2016aa] can be applied to the present 3D system, and discontinuities of the center of mass of the occupied particles, which are the contribution from the surface states, reproduces the correct number of the pumped particles. The center of mass is one of important observables for the topological pump: [@Wang:2013fk_pump] Indeed, in the recent experiments of the 1D pump, [@Nakajima:2016aa; @Lohse:2016aa] the center of mass played a central role. Thus, it would be expected that 3D pump can be detected experimentally using the center of mass of the occupied particles. This would also imply the experimental observation of the chiral anomaly. In passing, we would like to add a comment on the effect of interactions. So far we have discussed the bulk-edge correspondence for a particle pump of a noninteracting system by observing the discontinuities of the center of mass of the ground state [*in contact with a particle reservoir*]{}. Since the present pumping is of topological origin, small interactions cannot change the quantized discontinuities of the center of mass. On the other hand, for systems with strong interactions, fractional pumping have been proposed for 1D systems with degenerate ground states. [@Zeng:2016aa; @Li:2017aa; @Taddia:2017aa] In such cases, the discontinuities of the center of mass is not obvious, but if we introduce a particle reservoir also for such systems and consider the ground states with different number of particles, we could discuss the discontinuities for the degenerate ground states. This is because of the universality of the bulk-edge correspondence: The bulk topological properties should be closely related with the edge states also for interacting systems. Thirdly, we applied the generalized Streda formula, which is based on the chiral anomaly of the Dirac fermion, to compute the second Chern number. Let us here mention the anomaly of the present system. The current (\[ChaDen3\]) may be derived from the effective action, as has been done in Ref. [@Qi:2008aa]. Although we did not directly calculate the effective action, we can derive it from the expressions of the current (\[ChaDen1\]) and (\[ChaDen3\]) as follows: Let $\Gamma_{\rm eff}[\theta,A]$ be the effective action defined by $$\begin{aligned} 1 i\Gamma_{\rm eff}[\theta,A]=\ln{\rm Det}\left[(i\slashed D-me^{i\gamma_5\theta})/(i\slashed D-m)\right].\end{aligned}$$ Then, the current can be obtained by $$\begin{aligned} 1 \langle j^\mu(x)\rangle=\frac{1}{e}\frac{\delta \Gamma_{\rm eff}[\theta,A]}{\delta A_\mu(x)},\end{aligned}$$ from which we have for $d=1+1$ system $$\begin{aligned} 1 \Gamma_{\rm eff}[\theta,A]=\frac{e}{2}\int d^2x\epsilon^{\mu\nu}P_1(\theta)F_{\mu\nu},\end{aligned}$$ and for $d=3+1$ system $$\begin{aligned} 1 \Gamma_{\rm eff}[\theta,A]=-\frac{e^2}{16\pi}\int d^4x\epsilon^{\mu\nu\rho\sigma}P_3(\theta) F_{\mu\nu}F_{\rho\sigma},\end{aligned}$$ where the charge polarization ($d=1+1$) or the magneto-electric polarization ($d=3+1$), $P_{d-1}(\theta)$, is defined by [@Qi:2008aa] $$\begin{aligned} 1 P_{d-1}(\theta)=\int_0^\theta d\theta G_{d-1}(\theta).\end{aligned}$$ For $\theta=2\pi$, the effective action gives the chiral anomaly. On the other hand, in Sec. \[s:Streda\] in this paper, we demonstrated the manifestation of the anomaly based on a related method developed in Ref. [@Fukui:2016aa]. Namely, we calculated the second Chern number from spectral asymmetry of the four dimensional Hamiltonian with an electric field as well as a magnetic field. As shown in Ref. [@Fukui:2016aa], the spectral asymmetry gives the chiral anomaly of the overlap Dirac operator [@Neuberger:1998ab; @Neuberger:1998aa] obeying the Ginsparg-Wilson relation. It may be usually natural to use the Wilson-Dirac Hamiltonian to construct the overlap operator $D=\frac{1}{a}\left(1-\gamma_5\frac{\cal H}{\sqrt{{\cal H}^2}}\right)$. However, for any gapped Hamiltonian, the overlap operator obeys the Ginsparg-Wilson relation.[@Ginsparg:1982aa] Thus, it would be an interesting future problem to seek the possible Hamiltonian for the overlap operator. Finally, we would like to add a comment on recent observations concerning the four dimensional (4D) topological pump. In Refs. [@1705.08361; @1705.08371], the well-known two dimensional (2D) (or $d=1+1$) topological pump models such as the Harper pump model [@Hatsugai:2016aa] or Rice-Mele model [@Rice:1982qf; @Xiao:2010fk] $H(k,t)$ are extended to a 4D model considering the direct sum, $H(k_1,t)+H'(k_2,s)$. This allows a simple relationship between the second Chern number that governs the topological properties of a 4D system and the first Chern number of each 2D (or $d=1+1$) subsystem. In spite of some weak couplings between two subsystems in the experimental setup, the expected second Chern number has been observed indeed. However, if the couplings become larger, the topological change may be expected, which is an interesting future issue to explore. To this end, we note that for the lowest non-degenerate band studied in Refs. [@1705.08361; @1705.08371], the second Chern number associated with the U(1) Berry curvature can be computed directly on the lattice. [@Luescher:1999uq; @Luescher:1999fk; @1126-6708-2002-09-025] Also it may be interesting to develop several numerical techniques studied in Sec. \[s:MEP\] for these non-Dirac systems, or to apply the techniques of the entanglement Chern number, which can separate the Chern number into those of subsystems.[@Fukui:2014qv; @Araki:2017aa]. Acknowledgments {#acknowledgments .unnumbered} =============== We would like to thank Y. Hatsugai for fruitful discussions. This work was supported in part by Grants-in-Aid for Scientific Research Numbers 17K05563 and 17H06138 from the Japan Society for the Promotion of Science. Current density in the continuum limit {#s:CurDen} ====================================== In this Appendix, we calculate the current density given by Eq. (\[WilDirCur\]). It is similar to the charge density, but it includes the lattice corrections. From Eq. (\[WilDirCur\]), we need to calculate $$\begin{aligned} 1 \langle j^l(x)\rangle&= -\lim_{x'\rightarrow x} {{\rm tr}\,}\langle0| T\Big\{\gamma^l\psi(x)\bar\psi(x') +\frac{a}{2}\gamma^l\left[\nabla_l\psi(x)\bar\psi(x') +\psi(x)\bar\psi(x')\overleftarrow{\nabla}_l'\right] +\frac{iba}{2}\left[\nabla_l\psi(x)\bar\psi(x') -\psi(x)\bar\psi(x')\overleftarrow{\nabla}_l'\right]\Big\}|0\rangle,\end{aligned}$$ where the repeated $l$ in the middle term is not summed, as have been noticed. Using the propagator (\[WilDirPro\]), this can be written as $$\begin{aligned} 1 \langle j^l(x)\rangle &=\frac{1}{a^{d-1}} \int_{-\infty}^\infty\frac{d\omega}{2\pi i} \int\frac{d^{d-1}k}{(2\pi)^{d-1}} e^{-i\frac{kx}{a}} {{\rm tr}\,}\Bigg\{\left(\gamma^l+\frac{a}{2}\gamma^l\nabla_l+\frac{iba}{2}\nabla_l\right) \frac{1}{ia\slashed D^{\rm L}-me^{i\gamma_5\theta}-\frac{b}{2}a^2\Delta^{\rm L}+i\epsilon} \nonumber\\ &\qquad+\frac{1}{ia\slashed D^{\rm L}-me^{i\gamma_5\theta}-\frac{b}{2}a^2\Delta^{\rm L}+i\epsilon} \left(\frac{a}{2}\gamma^l\overleftarrow{\nabla}_l-\frac{iba}{2}\overleftarrow{\nabla}_l\right) \Bigg\}e^{i\frac{kx}{a}} . \label{Jl_1}$$ Note that in the limit $a\rightarrow 0$, the difference $\overleftarrow{\nabla}_l$ becomes $$\begin{aligned} 1 e^{-i\frac{kx}{a}} a\overleftarrow{\nabla}_j e^{i\frac{kx}{a}}=e^{ik_j}a\overleftarrow{\nabla}_j+e^{ik_j}-1 =-e^{ik_j}aD_j+e^{ik_j}-1+O(a^2).\end{aligned}$$ Using this, together with (\[DiraOpeCon\_1\]), we have $$\begin{aligned} 1 \langle j^l(x)\rangle =&\frac{1}{a^{d-1}} \int_{-\infty}^\infty\frac{d\omega}{2\pi i} \int\frac{d^{d-1}k}{(2\pi)^{d-1}} {{\rm tr}\,}(\gamma^lc_l+bs_l) \left(\slashed{s}-me^{-i\gamma_5\theta}-bc_{\rm s}\right) \nonumber\\ &\times\frac{1} { \mu^2-s^2+me^{i\gamma_5\theta}\gamma_5a \tilde{\slashed{\partial}}\theta -mbe^{-i\gamma_5\theta}\gamma_5a\tilde\partial_{\rm s}\theta -\frac{i\gamma^\rho\gamma^\sigma}{2}ea^2\widetilde F_{\rho\sigma} +ib\gamma^\rho ea^2\widetilde F_{\rho,{\rm s}} }, \label{Jl_2}$$ where the repeated $l$ in the rhs is not summed, and we have already omitted irrelevant operators after the limit $a\rightarrow0$ as well as after the trace over the $\gamma$ matrices. Expanding the propagator in (\[Jl\_2\]) with respect to $a$ as we did to compute the charge density in Sec. \[s:ChaDen\], we can calculate the current density. In what follows, we briefly show several steps of the calculations separately in $d=1+1$ and $d=3+1$. $d=1+1$ system -------------- Corresponding to Eq. (\[ChaDen1Ori\]), the following expression can be obtained for the current density, $$\begin{aligned} 1 \langle j^1(x)\rangle =&\frac{1}{a} \int_{-\infty}^\infty\frac{d\omega}{2\pi i} \int\frac{dk_1}{2\pi} \frac{ {{\rm tr}\,}(\gamma^1c_1+bs_1) \left(\slashed{s}-me^{-i\gamma_5\theta}-bc_{\rm s}\right)(-ma\gamma_5) (e^{i\gamma_5\theta}\tilde{\slashed{\partial}}\theta -be^{-i\gamma_5\theta}\tilde\partial_{\rm s}\theta) } {(\mu^2-s^2-i\epsilon)^2}. \label{Jl_d1}$$ In the numerator, the terms which survive after the trace and integration over $\omega$ are $$\begin{aligned} 1 (-ma)\left[ {{\rm tr}\,}\gamma_5\gamma^1c_1(me^{-i\gamma_5\theta}+bc_{\rm s})e^{i\gamma_5\theta}\tilde{\slashed\partial}\theta - bs_1{{\rm tr}\,}\gamma_5\slashed se^{i\gamma_5\theta}\tilde{\slashed\partial}\theta \right] =2am\left[(m+bc_{\rm s}\cos\theta)c_1+bs_1^2\cos\theta\right]\epsilon^{10}\partial_0\theta .\end{aligned}$$ This is nothing but Eq. (\[ChaDen1\]) for $\mu=1$. Thus, we have established that the result in Eq. (\[ChaDen1\]) is valid for any $\mu$. $d=3+1$ system -------------- From Eq. (\[Jl\_2\]), we obtain the following equation similar to Eq. (\[J0\_3\]), $$\begin{aligned} 1 \langle j^l(x)\rangle &=\frac{1}{a^3} \int_{-\infty}^\infty\frac{d\omega}{2\pi i} \int\frac{d^3k}{(2\pi)^3} \frac{1}{(\mu^2-s^2-i\epsilon)^3} \nonumber\\ &\times {{\rm tr}\,}(\gamma^lc_l+bs_l)\left(\slashed{s}-me^{-i\gamma_5\theta}-bc_{\rm s}\right) \left( -me^{i\gamma_5\theta}\gamma_5a \tilde{\slashed{\partial}}\theta +mbe^{-i\gamma_5\theta}\gamma_5a\tilde\partial_{\rm s}\theta +\frac{i}{2}\gamma^\rho\gamma^\sigma ea^2\widetilde F_{\rho\sigma} -ib\gamma^\rho ea^2\widetilde F_{\rho,{\rm s}} \right)^2 .\end{aligned}$$ As in the case of the charge density, the product terms between $\partial\theta$ and $\widetilde F$ survive in the limit $a\rightarrow0$ and after the trace over the $\gamma$ matrices. To be concrete, the trace for the $\gamma$ matrices yields $$\begin{aligned} 1 ea^3\Big\{& {{\rm tr}\,}\gamma^lc_l(-me^{-i\gamma_5\theta}-bc_{\rm s})(-me^{i\gamma_5\theta}\gamma_5a\tilde{\slashed\partial}\theta) (i\gamma^\rho\gamma^\sigma \widetilde F_{\rho\sigma}) +{{\rm tr}\,}\gamma^lc_l\slashed smbe^{-i\gamma_5\theta}\gamma_5\tilde\partial_{\rm s}\theta (i\gamma^\rho\gamma^\sigma \widetilde F_{\rho\sigma}) \nonumber\\ &+2{{\rm tr}\,}\gamma^lc_l\slashed s(-me^{i\gamma_5\theta}\gamma_5\tilde{\slashed\partial}\theta) (-ib\gamma^\rho\widetilde F_{\rho,{\rm s}}) +bs_l{{\rm tr}\,}\slashed s(-me^{i\gamma_5\theta}\gamma_5\tilde{\slashed\partial}\theta) (i\gamma^\rho\gamma^\sigma\widetilde F_{\rho\sigma}) \Big\} \nonumber\\ =&-mea^3i\Big\{ (m+bc_{\rm s}\cos\theta)c_l{{\rm tr}\,}\gamma_5\gamma^l (\tilde{\slashed\partial}\theta) \gamma^\rho\gamma^\sigma \widetilde F_{\rho\sigma} -b\cos\theta c_l{{\rm tr}\,}\gamma_5\gamma^l\slashed s\tilde\partial_{\rm s}\theta \gamma^\rho\gamma^\sigma \widetilde F_{\rho\sigma} \nonumber\\ & -2b\cos\theta c_l{{\rm tr}\,}\gamma_5\gamma^l\slashed s\tilde{\slashed\partial}\theta \gamma^\rho\widetilde F_{\rho,{\rm s}} -bs_l\cos\theta{{\rm tr}\,}\gamma_5\slashed s\tilde{\slashed\partial}\theta\gamma^\rho\gamma^\sigma\widetilde F_{\rho\sigma} \Big\} \nonumber\\ =&-4ea^3m\left[m+b\sum_{j=1}^3(\sec k_j-1)\cos\theta\right]\left(\prod_{j=1}^3\cos k_j\right) \epsilon^{l\nu\rho\sigma}(\partial_\nu\theta) F_{\rho\sigma} .\end{aligned}$$ Thus, we finally reach Eq. (\[ChaDen3\]) for $\mu=l$, and the formula (\[ChaDen3\]) has been established for any $\mu$. Mathematical formulas ===================== In this Appendix, we show some mathematical formulas used in the text. Trace for $\gamma$ matrices --------------------------- The trace of the $\gamma$ matrices including $\gamma_5$ is summarized as follows: In $d=1+1$ dimensions, $$\begin{aligned} 1 {{\rm tr}\,}\gamma_5={{\rm tr}\,}\gamma_5\gamma^\mu=0,\quad {{\rm tr}\,}\gamma_5\gamma^\mu\gamma^\nu=-2\epsilon^{\mu\nu} . \label{TraGam1D}$$ In $d=3+1$ dimensions, $$\begin{aligned} 1 &{{\rm tr}\,}\gamma_5={{\rm tr}\,}\gamma_5\gamma^\mu={{\rm tr}\,}\gamma_5\gamma^\mu\gamma^\nu={{\rm tr}\,}\gamma_5\gamma^\mu\gamma^\nu\gamma^\rho=0, \nonumber\\ &{{\rm tr}\,}\gamma_5\gamma^\mu\gamma^\nu\gamma^\rho\gamma^\sigma=-4i\epsilon^{\mu\nu\rho\sigma}. \label{TraGam3D}$$ Integral -------- For the $d$ dimensional momentum integration in the continuum model, we use $$\begin{aligned} 1 \int\frac{d^dk}{i(2\pi)^d}\frac{1}{(1-k^2-i\epsilon)^n}=\frac{\Gamma(n-d/2)}{(4\pi)^{d/2}\Gamma(n)}. \label{Int_1}$$ As to the integration over $\omega$ in the lattice model, $$\begin{aligned} 1 I_n\equiv\int_{-\infty}^\infty\frac{d\omega}{2\pi i}\frac{1}{(\Omega^2-\omega^2-i\epsilon)^n}.\end{aligned}$$ we simply have in the $n=1$ case, $$\begin{aligned} 1 I_1=\frac{1}{2\Omega}.\end{aligned}$$ Then, we have $$\begin{aligned} 1 &I_2=-\frac{d}{d\Omega^2}I_1=\frac{1}{4\Omega^3}, \label{IntOme1} \\ &I_3=-\frac{1}{2}\frac{d}{d\Omega^2}I_2=\frac{3}{16\Omega^3}. \label{IntOme2}\end{aligned}$$ [70]{} natexlab\#1[\#1]{}bibnamefont \#1[\#1]{}bibfnamefont \#1[\#1]{}citenamefont \#1[\#1]{}url \#1[`#1`]{}urlprefix\[2\][\#2]{} \[2\]\[\][[\#2](#2)]{} , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.47.986>. , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.58.1799>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.42.1698>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.22.2099>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.21.2388>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevLett.95.146802>. , , , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.78.195424>. , ****, (), <http://link.aps.org/doi/10.1103/RevModPhys.82.3045>. , ****, (), <http://link.aps.org/doi/10.1103/RevModPhys.83.1057>. , , , , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.78.195125>. , , , , ****, (), , <http://arXiv.org/abs/0912.2157>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.27.6083>. , , , , , , , , ****, (), <http://dx.doi.org/10.1038/nphys3622>. , , , , , ****, (), <http://dx.doi.org/10.1038/nphys3584>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.102.146805>. , , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.81.205104>. , , , , ****, (), <http://stacks.iop.org/1367-2630/12/i=5/a=053032>. , , , ****, (), <http://link.aps.org/doi/10.1103/PhysRevD.78.074033>. , ****, (), <http://stacks.iop.org/1367-2630/9/i=9/a=356>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevLett.107.127205>. , , , , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.83.205101>. , , , , , , , , ****, (). , , , , , , (), . , , , , , , , , , , , ****, (), <http://dx.doi.org/10.1038/ncomms13741>. , ****, (), <http://www.sciencedirect.com/science/article/pii/0370269383915290>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.88.125105>. , , , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevX.4.031035>. , , , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.89.035142>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.86.115133>. , , , , ****, (), <http://stacks.iop.org/0256-307X/30/i=2/a=027101>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.91.115203>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.92.235205>. , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.116.166601>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.94.115160>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.118.207701>. , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.116.096401>. , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.93.094510>. , , , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.84.075129>. , ****, (), <https://link.aps.org/doi/10.1103/PhysRevD.86.045001>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevLett.111.027201>. , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.88.245107>. , , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevX.6.041046>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevX.6.041021>. , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.117.136602>. , , , ****, (), <http://link.aps.org/doi/10.1103/PhysRevLett.111.026802>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.94.041102>. , ****, (), <http://dx.doi.org/10.7566/JPSJ.85.124709>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevB.14.2239>. , , , , ****, (), <http://link.aps.org/doi/10.1103/PhysRevLett.49.405>. , ****, (). , , , ****, (), <http://link.aps.org/doi/10.1103/RevModPhys.82.1959>. , , , ****, (), <http://dx.doi.org/10.1143/JPSJ.74.1674>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRev.182.1517>. , ****, (), <http://www.sciencedirect.com/science/article/pii/S0370269397013683>. , ****, (), <http://www.sciencedirect.com/science/article/pii/S0370269398003554>. , ****, (), <http://link.aps.org/doi/10.1103/PhysRevD.25.2649>. , ****, (), <http://www.sciencedirect.com/science/article/pii/S0370269398004237>. , ****, (), <http://www.sciencedirect.com/science/article/pii/S0370269399000210>. , ****, (), , <http://ptp.oxfordjournals.org/content/102/1/141.abstract>. , , , ****, (), <http://stacks.iop.org/1126-6708/2002/i=09/a=025>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.94.235139>. , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.96.085444>. , , , , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevLett.118.230402>. , , , , , , (), . , , , , (), . , ****, (), <http://link.aps.org/doi/10.1103/PhysRevLett.49.1455>. , ****, (). , ****, (). , ****, (), <http://dx.doi.org/10.7566/JPSJ.83.113705>. , , , ****, (), <https://link.aps.org/doi/10.1103/PhysRevB.96.165139>.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We consider a bilevel optimisation approach for parameter learning in higher-order total variation image reconstruction models. Apart from the least squares cost functional, naturally used in bilevel learning, we propose and analyse an alternative cost, based on a Huber regularised TV-seminorm. Differentiability properties of the solution operator are verified and a first-order optimality system is derived. Based on the adjoint information, a quasi-Newton algorithm is proposed for the numerical solution of the bilevel problems. Numerical experiments are carried out to show the suitability of our approach and the improved performance of the new cost functional. Thanks to the bilevel optimisation framework, also a detailed comparison between $\TGV^2$ and $\ICTV$ is carried out, showing the advantages and shortcomings of both regularisers, depending on the structure of the processed images and their noise level.' address: - '$^1$Research Center on Mathematical Modelling (MODEMAT), Escuela Politécnica Nacional, Quito, Ecuador.' - '$^2$Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Cambridge, United Kingdom.' author: - 'J.C. De los Reyes$^1$, C.-B. Schönlieb$^2$ and T. Valkonen$^{2}$' bibliography: - 'abbrevs.bib' - 'bib.bib' - 'bib-own.bib' title: 'Bilevel parameter learning for higher-order total variation regularisation models$^*$' --- [^1] Introduction ============ In this paper we propose a bilevel optimisation approach for parameter learning in higher-order total variation regularisation models for image restoration. The reconstruction of an image from imperfect measurements is essential for all research which relies on the analysis and interpretation of image content. Mathematical image reconstruction approaches aim to maximise the information gain from acquired image data by intelligent modelling and mathematical analysis. A variational image reconstruction model can be formalised as follows. Given data $f$ which is related to an image (or to certain image information, e.g. a segmented or edge detected image) $u$ through a generic forward operator (or function) $K$ the task is to retrieve $u$ from $f$. In most realistic situations this retrieval is complicated by the ill-posedness of $K$ as well as random noise in $f$. A widely accepted method that approximates this ill-posed problem by a well-posed one and counteracts the noise is the method of Tikhonov regularisation. That is, an approximation to the true image is computed as a minimiser of $$\label{reg1:eq} \alpha~ R(u) + d(K(u),f),$$ where $R$ is a regularising energy that models a-priori knowledge about the image $u$, $d(\cdot , \cdot)$ is a suitable distance function that models the relation of the data $f$ to the unknown $u$, and $\alpha>0$ is a parameter that balances our trust in the forward model against the need of regularisation. The parameter $\alpha$ in particular, depends on the amount of ill-posedness in the operator $K$ and the amount (amplitude) of the noise present in $f$. A key issue in imaging inverse problems is the correct choice of $\alpha$, image priors (regularisation functionals $R$), fidelity terms $d$ and (if applicable) the choice of what to measure (the linear or nonlinear operator $K$). Depending on this choice, different reconstruction results are obtained. While functional modelling constitutes a mathematically rigorous and physical way of setting up the reconstruction of an image – providing reconstruction guarantees in terms of error and stability estimates – it is limited with respect to its adaptivity for real data. On the other hand, data-based modelling of reconstruction approaches is set up to produce results which are optimal with respect to the given data. However, in general it neither offers insights into the structural properties of the model nor provides comprehensible reconstruction guarantees. Indeed, we believe that for the development of reliable, comprehensible and at the same time effective models it is essential to aim for a unified approach that seeks tailor-made regularisation and data models by combining model- and data-based approaches. To do so we focus on a bilevel optimisation strategy for finding an optimal setup of variational regularisation models . That is, for a given training pair of noisy and original clean images $(f,f_0)$, respectively, we consider a learning problem of the form $$\label{eq:generprob} \min F(u^*) = cost(u^*,f_0) \quad \textrm{subject to} \quad u^*\in \argmin_u \left\{\alpha~ R(u) + d(K(u),f)\right\},$$ where $F$ is a generic cost functional that measures the fitness of $u^*$ to the original image $f_0$. The argument of the minimisation problem will depend on the specific setup (i.e. the degrees of freedom) in the constraint problem . In particular, we propose a bilevel optimisation approach for learning optimal parameters in higher-order total variation regularisation models for image reconstruction in which the arguments of the optimisation constitute parameters in front of the first- and higher-order regularisation terms. Rather than working on the discrete problem, as is done in standard parameter learning and model optimisation methods, we optimise the regularisation models in infinite dimensional function space. We will explain this approach in more detail in the next section. Before, let us give an account to the state of the art of bilevel optimisation for model learning. In machine learning bilevel optimisation is well established. It is a semi-supervised learning method that optimally adapts itself to a given dataset of measurements and desirable solutions. In [@tappen2007utilizing; @domke2012generic; @chen2014], for instance the authors consider bilevel optimization for finite dimensional Markov random field models. In inverse problems the optimal inversion and experimental acquisition setup is discussed in the context of optimal model design in works by Haber, Horesh and Tenorio [@haber2003learning; @haber2010numerical], as well as Ghattas et al. [@bui2008model; @biegler2011large]. Recently parameter learning in the context of functional variational regularisation models also entered the image processing community with works by the authors [@de2013image; @calatronidynamic], Kunisch, Pock and co-workers [@kunisch2013bilevel; @Chen2012], Chung et al. [@chung2014optimal] and Hintermüller et al. [@hintermuller2014bilevel]. Apart from the work of the authors [@de2013image; @calatronidynamic], all approaches so far are formulated and optimised in the discrete setting. Our subsequent modelling, analysis and optimisation will be carried out in function space rather than on a discretisation of . While digitally acquired image data is of course discrete, the aim of high resolution image reconstruction and processing is always to compute an image that is close to the real (analogue, infinite dimensional) world. Hence, it makes sense to seek images which have certain properties in an infinite dimensional function space. That is, we aim for a processing method that accentuates and preserves qualitative properties in images independent of the resolution of the image itself [@viola2012unifying]. Moreover, optimisation methods conceived in function space potentially result in numerical iterative schemes which are resolution and mesh-independent upon discretisation [@hintermuller2006infeasible]. Higher-order total variation regularisation has been introduced as an extension of the standard total variation regulariser in image processing. As the [**T**]{}otal [**V**]{}ariation (TV) [@Rud1992] and many more contributions in the image processing community have proven, a non-smooth first-order regularisation procedure results in a nonlinear smoothing of the image, smoothing more in homogeneous areas of the image domain and preserving characteristic structures such as edges. In particular, the TV regulariser is tuned towards the preservation of edges and performs very well if the reconstructed image is piecewise constant. The drawback of such a regularisation procedure becomes apparent as soon as images or signals (in 1D) are considered which do not only consist of constant regions and jumps, but also possess more complicated, higher-order structures, e.g. piecewise linear parts. The artefact introduced by TV regularisation in this case is called staircasing [@ring2000structural]. One possibility to counteract such artefacts is the introduction of higher-order derivatives in the image regularisation. Chambolle and Lions [@chambolle97image], for instance, propose a higher order method by means of an infimal convolution of the TV and the TV of the image gradient called [**I**]{}nfimal-[**C**]{}onvolution [**T**]{}otal [**V**]{}ariation (ICTV) model. Other approaches to combine first and second order regularisation originate, for instance, from Chan, Marquina, and Mulet [@chan2000highorder] who consider total variation minimisation together with weighted versions of the Laplacian, the Euler-elastica functional [@masnou1998level; @chan2002euler] which combines total variation regularization with curvature penalisation, and many more [@lysaker2006iterative; @papafitsoros2012combined] just to name a few. Recently Bredies et al. have proposed [**T**]{}otal [**G**]{}eneralized [**V**]{}ariation (TGV) [@bredies2011tgv] as a higher-order variant of TV regularisation. In this work we mainly concentrate on two second-order total variation models: the recently proposed TGV [@bredies2011tgv] and the ICTV model of Chambolle and Lions [@chambolle97image]. We focus on second-order TV regularisation only since this is the one which seems to be most relevant in imaging applications [@knoll2010second; @bredies2012total]. For $\Omega\subset\mathbb R^2$ open and bounded and $u\in BV(\Omega)$, the ICTV regulariser reads $$\label{eq:ictv} \ICTV_{\alpha,\beta}(u)\defeq \min_{v\in W^{1,1}(\Omega),~ \nabla v\in BV(\Omega)} \alpha \norm{Du-\nabla v}_{\Meas(\Omega; \R^2)} + \beta \norm{D\nabla v}_{\Meas(\Omega; \R^{2\times 2})}.$$ On the other hand, second-order TGV [@sampta2011tgv; @l1tgv] for $u\in BV(\Omega)$ reads $$\label{eq:tgv} \TGV^2_{\alpha,\beta}(u) \defeq \min_{w \in BD(\Omega)} \alpha \norm{Du-w}_{\Meas(\Omega; \R^2)} + \beta \norm{Ew}_{\Meas(\Omega; \Sym^2(\R^2))}.$$ Here $ \BDspace(\Omega) \defeq \{ w \in L^1(\Omega; \R^\DIMdomain) \mid \norm{Ew}_{\Meas(\Omega; \R^{\DIMdomain \times \DIMdomain})} < \infty \} $ is the space of vector fields of bounded deformation on $\Omega$, $E$ denotes the *symmetrised gradient* and $\mathrm{Sym}^2(\mathbb{R}^2)$ the space of symmetric tensors of order $2$ with arguments in $\mathbb{R}^2$. The parameters $\alpha,\beta$ are fixed positive parameters and will constitute the arguments in the special learning problem á la we consider in this paper. The main difference between and is that we do not generally have that $w=\nabla v$ for any function $v$. That results in some qualitative differences of ICTV and TGV regularisation, compare for instance [@benning2011higher]. Substituting $\alpha R(u)$ in by $\TGV^2_{\alpha,\beta}(u)$ or $\ICTV_{\alpha,\beta}(u)$ gives the TGV image reconstruction model and the ICTV image reconstruction model, respectively. In this paper we only consider the case $K=Id$ identity and $d(u,f)=\|u-f\|_{L^2(\Omega)}^2$ in which corresponds to an image de-noising model for removing Gaussian noise. With our choice of regulariser the former scalar $\alpha$ in has been replaced by a vector $(\alpha,\beta)$ of two parameters in and . The choice of the entries in this vector now do not only determine the overall strength of the regularisation (depending on the properties of $K$ and the noise level) but those parameters also balance between the different orders of regularity of the function $u$, and their choice is indeed crucial for the image reconstruction result. Large $\beta$ will give regularised solutions that are close to TV regularised reconstructions, compare Figure \[fig:intro\]. Large $\alpha$ will result in TV$^2$ type solutions, that is solutions that are regularised with TV of the gradient [@hinterberger2006variational; @papafitsoros2012combined], compare Figure \[fig:intro2\]. With our approach described in the next section we propose a learning approach for choosing those parameters optimally, in particular optimally for particular types of images. [0.3]{} [0.3]{} [0.3]{} [0.3]{} [0.3]{} [0.3]{} For the existence analysis of an optimal solution as well as for the derivation of an optimality system for the corresponding learning problem we will consider a smoothed version of the constraint problem – which is the one in fact used in the numerics. That is, we replace $R(u)$ – being TV, TGV or ICTV in this paper – by a Huber regularised version and add an $H^1$ regularisation with a small weight to . In this setting and under the special assumption of box constraints on $\alpha$ and $\beta$ we provide a simple existence proof for an optimal solution. A more general existence result that holds also for the original non-smooth problem and does not require box constraints is derived in [@tuomov-interior] and we refer the reader to this paper for a more sophisticated analysis on the structure of solutions. A main challenge in the setup of such a learning approach is to decide what is the best way to measure fitness (optimality) of the model. In our setting this amounts to choosing an appropriate distance $F$ in that measures the fitness of reconstructed images to the ‘perfect’, noise-free images in an appropriate training set. We have to formalise what we mean by an optimal reconstruction model. Classically, the difference between the original, noise-free image $f_0$ and its regularised version $u_{\alpha,\beta}$ is computed with an $\costltwo$ cost functional $$\label{eq:costL2} \costf_{\costltwo}(u_{\alpha,\beta}) = \norm{u_{\alpha,\beta} - f_0}_{L^2(\Omega)}^2,$$ which is closely related to the PSNR quality measure. Apart from this, we propose in this paper an alternative cost functional based on a Huberised total variation cost $$\label{eq:costTVhub} \costf_{\costhubertv}(u_{\alpha,\beta}) \defeq \int_\Omega \huber[\gamma]{D(u_{\alpha,\beta}-f_0)}~ dx,$$ where the Huber regularisation $\huber[\gamma]{\cdot}$ will be defined later on in Definition \[def:huber\]. We will see that the choice of this cost functional is indeed crucial for the qualitative properties of the reconstructed image. The proposed bilevel approach has an important indirect consequence: It establishes a basis for the comparison of the different total variation regularisers employed in image denoising tasks. In the last part of the paper we exhaustively compare the performance of $\TV$, $\TGV^2$ and $\ICTV$ for various image datasets. The parameters are chosen optimally, according to the proposed bilevel approach, and different quality measures (like PSNR and SSIM) are considered for the comparison. The obtained results are enlightening about when to use each one of the considered regularisers. In particular, $\ICTV$ appears to behave better for images with arbitrary structure and moderate noise levels, whereas $\TGV^2$ behaves better for images with large smooth areas.\ [**Outline of the paper**]{} In Section \[sec:probstatex\] we state the bilevel learning problem for the two higher-order total variation regularisation models, TGV and ICTV, and prove existence of an optimal parameter pair $\alpha,\beta$. The bilevel optimization problem is analysed in Section \[sec:Lagrange\], where existence of Lagrange multipliers is proved and an optimality system, as well as a gradient formula, are derived. Based on the optimality condition, a BFGS algorithm for the bilevel learning problem is devised in Section \[sec:denoising\]. For the numerical solution of each denoising problem an infeasible semi-smooth Newton method is considered. Finally, we discuss the performance of the parameter learning method by means of several examples for the denoising of natural photographs in Section \[sec:experiments\]. Therein, we also present a statistical analysis on how TV, ICTV and TGV regularisation compare in terms of returned image quality, carried out on 200 images from the Berkeley segmentation dataset BSDS300. Problem statement and existence analysis {#sec:probstatex} ======================================== We strive to develop a parameter learning method for higher-order total variation regularisation models that maximises the fit of the reconstructed images to training images simulated for an application at hand. For a given noisy image $f\in L^2(\Omega)$, $\Omega\subset\mathbb R^2$ open and bounded, we consider $$\label{tgvdenoise} \min_u \left\{R_{\alpha, \beta}(u) + \frac{1}{2} \|u-f\|_{L^2(\Omega)}^2\right\}.$$ where, $\alpha,\beta\in\mathbb R$. We focus on TGV$^2$ and ICTV image denoising: $$\begin{gathered} R_{\alpha,\beta}(u)=\TGV^2_{\alpha,\beta}(u)\defeq \min_{w \in BD(\Omega)} \norm{\alpha~(Du-w)}_{\Meas(\Omega; \R^2)} \\+ \norm{\beta~Ew}_{\Meas(\Omega; \Sym^2(\R^2))}.\end{gathered}$$ and with spatial dependence $$\begin{gathered} R_{\alpha,\beta}(u)=\ICTV_{\alpha, \beta}(u)\defeq\min_{\substack{v\in W^{1,1}(\Omega)\\ \nabla v\in BV(\Omega)}} \norm{\alpha~(Du-\nabla v)}_{\Meas(\Omega; \R^2)} \\ + \norm{\beta~D\nabla v}_{\Meas(\Omega; \R^{2\times 2})},\end{gathered}$$ for $u\in BV(\Omega)$. For this model, we want to determine the optimal choice of $\alpha,\beta$, given a particular type of images and a fixed noise level. More precisely, we consider a training pair $(f,f_0)$, where $f$ is a noisy image corrupted by normally distributed noise with a fixed variation, and the image $f_0$ represents the ground truth or an image that approximates the ground truth within a desirable tolerance. Then, we determine the optimal choice of $\alpha,\beta$ by solving the following problem $$\label{optimalcontrolbasic} \min_{(\alpha,\beta)\in\mathbb R^{2}} ~ \costf(u_{\alpha,\beta}) \quad \textrm{ s.t. } \alpha, \beta \geq 0,$$ where $\costf$ equals the $\costltwo$ cost or the Huberised TV cost and $u_{\alpha,\beta}$ for a given $f$ solves a regularised version of the minimization problem that will be specified in the next section, compare problem . This regularisation of the problem is a technical requirement for solving the bilevel problem that will be discussed in the sequel. In contrast to learning $\alpha,\beta$ in in finite dimensional parameter spaces (as is the case in machine learning) we aim for novel optimisation techniques in infinite dimensional function spaces. Formal statement ---------------- Let $\Omega \subset \R^\DIMdomain$ be an open bounded domain with Lipschitz boundary. This will be our image domain. Usually $\Omega=(0, w) \times (0, h)$ for $w$ and $h$ the width and height of a two-dimensional image, although no such assumptions are made in this work. Our data $f$ and $f_0$ are assumed to lie in $L^2(\Omega)$. In our learning problem, we look for parameters $(\alpha,\beta)$ that for some cost functional $F: H^1(\Omega) \to \R$ solve the problem \[eq:learn\] $$\min_{(\alpha,\beta)\in\mathbb R^{2}} ~ \costf(u_{\alpha,\beta})$$ subject to $$\begin{aligned} & u_{\alpha,\beta} \in \argmin_{u\in H^1(\Omega)} \Jmu[\alpha,\beta]{u} \label{eq:denoise-huber1} \\ & \alpha, \beta \geq 0,\end{aligned}$$ where $$\Jmu[\alpha,\beta]{u} \defeq \frac12 \norm{u-f}_{L^2(\Omega)}^2 + R_{\alpha,\beta}^{\gamma,\mu}(u).$$ Here $\Jmu[\alpha,\beta]{\cdot}$ is the regularised denoising functional that amends the regularisation term in by a Huber regularised version of it with parameter $\gamma>0$, and an elliptic regularisation term with parameter $\mu>0$. In the case of TGV$^2$ the modified regularisation term $R_{\alpha,\beta}^{\gamma,\mu}(u)$ then reads for $u\in H^1(\Omega)$ $$\begin{gathered} \TGV^{2,\gamma,\mu}_{\alpha,\beta}(u)\defeq \min_{w \in H^1(\Omega)} \int_\Omega \alpha~|Du-w|_\gamma~ dx \\ + \int_\Omega \beta~|Ew|_\gamma~ dx + \frac{\mu}{2} \left(\norm{u}_{H^1(\Omega)}^2 + \norm{w}_{\mathbb H^1(\Omega)}^2\right)\end{gathered}$$ and in the case of ICTV we have $$\begin{gathered} \ICTV_{\alpha, \beta}^{\gamma,\mu}(u)\defeq\min_{\substack{v\in W^{1,1}(\Omega)\\\nabla v\in BV(\Omega,\mathbb R^n)\cap \mathbb H^1(\Omega)}} \int_\Omega \alpha ~|Du-\nabla v|_\gamma~dx \\ + \int_\Omega \beta~|D\nabla v|_\gamma~ dx + \frac{\mu}{2} \left(\norm{u}_{H^1(\Omega)}^2 + \norm{\nabla v}_{\mathbb H^1(\Omega)}^2\right).\end{gathered}$$ Here, $\mathbb H^1(\Omega)=H^1(\Omega;\mathbb R^n)$ and the Huber regularisation $|\cdot|_\gamma$ is defined as follows. \[def:huber\] Given $\gamma \in (0, \infty]$, we define for the norm $\norm{\freevar}_2$ on $\R^m$, the Huber regularisation $$\huber[\gamma]{g} = \begin{cases} \norm{g}_2 - \frac{1}{2\gamma}, & \norm{g}_2 \ge 1/\gamma, \\ \frac{\gamma}{2}\norm{g}_2^2, & \norm{g}_2 < 1/\gamma. \end{cases}$$ For the cost functional $\costf$, given noise-free data $f_0 \in L^2(\Omega)$ and a regularised solution $u\in H^1(\Omega)$, we consider in particular the $L^2$ cost $$\costf_{\costltwo}(u) \defeq \frac{1}{2}\norm{f_0 - u}_{\SPACEkuLtwo}^2,$$ as well as the Huberised total variation cost $$\costf_{\costhubertv}(u) \defeq \int_\Omega\huber[\gamma]{D(f_0-u)}~ dx$$ with noise-free data $f_0 \in \BVspace(\Omega)$. Existence of an optimal solution -------------------------------- The existence of an optimal solution for the learning problem is a special case of the class of bilevel problems considered in [@tuomov-interior], where existence of optimal parameters in $(0,+\infty]^{2N}$ is proven. For convenience, we provide a simplified proof for the case where box constraints on the parameters are imposed. We start with an auxiliary lower semicontinuity result for the Huber regularised functionals. \[lem:huberlsc\] Let $u,v\in L^p(\Omega)$, $1\leq p<\infty$. Then, the functional $u \mapsto \int_\Omega |u-v|_\gamma~ dx$, where $|\cdot|_\gamma$ is the Huber regularisation in Definition \[def:huber\], is lower semicontinuous with respect to weak\* convergence in $\Meas(\Omega; \R^\DIMkurange)$ Recall that for $g \in \R^m$, the Huber-regularised norm may be written in dual form as $$ \notag \huber[\gamma]{g}= \sup \Bigl\{ \iprod{q}{g} - \frac{\gamma}{2} \norm{q}_2^2 : \norm{q}_2 \le 1 \Bigr\}.$$ Therefore, we find that $$\begin{gathered} G(u)\defeq\int_\Omega |u-v|_\gamma~ dx=\sup\Big\{ \int_\Omega u(x)\cdot\varphi(x)~ dx - \int_\Omega \frac{\gamma}{2} \norm{\varphi(x)}_2^2 \d x : \\ \varphi \in C_c^\infty(\Omega),\ \norm{\varphi(x)}_2 \le \alpha \text{ for every } x \in \Omega \Big\}.\end{gathered}$$ The functional $G$ is of the form $G(u) = \sup\{\iprod{u}{\varphi}-G^*(\varphi)\}$, where $G^*$ is the convex conjugate of $G$. Now, let $\{u^i\}_{i=1}^\infty$ converge to $u$ weakly\* in $\Meas(\Omega; \R^\DIMkurange)$. Taking a supremising sequence $\{\varphi^j\}_{j=1}^\infty$ for this functional at any point $u$, we easily see lower semicontinuity by considering the sequences $\{\iprod{u^i}{\varphi^j}-G^*(\varphi^j)\}_{i=1}^\infty$ for each $j$. Our main existence result is the following. We consider the learning problem for TGV$^2$ and ICTV regularisation, optimising over parameters $(\alpha,\beta)$ such that $0 \leq \alpha \leq \bar \alpha, 0 \leq \beta \leq \bar \beta$. Here $(\bar\alpha,\bar\beta)<\infty$ is an arbitrary but fixed vector in $\mathbb R^{2}$ that defines a box constraint on the parameter space. Then, there exists an optimal solution $(\hat\alpha,\hat\beta)\in\mathbb R^{2}$ for this problem for both choices of cost functionals, $\costf=\costltwo$ and $\costf=\costf_{\costhubertv}$. Let $(\alpha_n,\beta_n)\subset \mathbb R^{2}$ be a minimising sequence. Due to the box constraints we have that the sequence $(\alpha_n,\beta_n)$ is bounded in $ \mathbb R^{2}$. Moreover, we get for the corresponding sequences of states $u_n:= u_{(\alpha_n,\beta_n)}$ that $$\Jmu[\alpha_n,\beta_n]{u_n} \leq \Jmu[\alpha_n,\beta_n]{u}, \quad \forall u\in H^1(\Omega),$$ in particular this holds for $u=0$. Hence, $$\label{eq:compbound} \frac12 \norm{u_n-f}_{L^2(\Omega)}^2 + R_{\alpha_n,\beta_n}^{\gamma,\mu}(u_n) \leq \frac12 \norm{f}_{L^2(\Omega)}^2.$$ Exemplarily, we consider here the case for the TGV regulariser, that is $R_{\alpha_n,\beta_n}^{\gamma,\mu} = \TGV^{2,\gamma,\mu}_{\alpha,\beta}$. The proof for the ICTV regulariser can be done in a similar fashion. Inequality in particular gives $$\norm{u_n}_{H^1(\Omega)}^2 + \norm{w_n}_{\mathbb H^1(\Omega)}^2 \leq \frac{1}{\mu} \norm{f}_{L^2(\Omega)},$$ where $w_n$ is the optimal $w$ for $u_n$. This gives that $(u_n,w_n)$ is uniformly bounded in $H^1(\Omega)\times \mathbb H^1(\Omega)$ and that there exists a subsequence $\{(\alpha_n,\beta_n,u_n,w_n)\}$ which converges weakly in $\mathbb R^{2}\times H^1(\Omega)\times \mathbb H^1(\Omega)$ to a limit point $(\hat\alpha,\hat\beta,\hat u,\hat w)$. Moreover, $u_n\rightarrow \hat u$ strongly in $L^p(\Omega)$ and $w_n\rightarrow \hat w$ in $L^p(\Omega;\mathbb R^n)$. Using the continuity of the $L^2$ fidelity term with respect to strong convergence in $L^2$, and the weak lower semicontinuity of the $H^1$ term with respect to weak convergence in $H^1$ and of the Huber regularised functional even with respect to weak$*$ convergence in $\mathcal M$ (cf. Lemma \[lem:huberlsc\]) we get $$\begin{aligned} & & & \frac12 \norm{\hat u-f}_{L^2(\Omega)}^2 + \int_\Omega \hat \alpha~|D\hat u-\hat w|_\gamma~ dx + \int_\Omega \hat\beta~|Ew|_\gamma~ dx \\ & & & + \frac{\mu}{2} \left(\norm{\hat u}_{H^1(\Omega)}^2 + \norm{\hat w}_{\mathbb H^1(\Omega)}^2\right) \\ & \leq & & \liminf_n \frac12 \norm{u_n-f}_{L^2(\Omega)}^2 + \int_\Omega \hat\alpha~|Du_n-w_n|_\gamma~ dx + \int_\Omega \hat\beta~|Ew_n|_\gamma~ dx \\ & & & + \frac{\mu}{2} \left(\norm{u_n}_{H^1(\Omega)}^2 + \norm{w_n}_{\mathbb H^1(\Omega)}^2\right) \\ & \leq & &\liminf_n \frac12 \norm{u_n-f}_{L^2(\Omega)}^2 + \int_\Omega \alpha_n~|Du_n-w_n|_\gamma~ dx + \int_\Omega \beta_n~|Ew_n|_\gamma~ dx \\ & & & + \frac{\mu}{2} \left(\norm{u_n}_{H^1(\Omega)}^2 + \norm{w_n}_{\mathbb H^1(\Omega)}^2\right), \end{aligned}$$ where in the last step we have used the boundedness of the sequence $R_{\alpha_n,\beta_n}^{\gamma,\mu}(u_n)$ from and the convergence of $(\alpha_n,\beta_n)$ in $\mathbb R^{2}$. This shows that the limit point $\hat u$ is an optimal solution for $(\hat\alpha,\hat\beta)$. Moreover, due to the weak lower semicontinuity of the cost functional $F$ and the fact that the set $\{(\alpha,\beta):~ 0 \leq \alpha \leq \bar\alpha,0 \leq \beta \leq \bar\beta \}$ is closed, we have that $(\hat\alpha,\hat\beta,\hat u)$ is optimal for . \[rem: existence\] - Using the existence result in [@tuomov-interior], in principle we could allow infinite values for $\alpha$ and $\beta$. This would include both $\TV^2$ and $\TV$ as possible optimal regularisers in our learning problem. - In [@tuomov-interior], in the case of the $L^2$ cost and assuming that $$R_{\alpha,\beta}^{\gamma}(f)>R_{\alpha,\beta}^{\gamma}(f_0),$$ we moreover show that the parameters $(\alpha,\beta)$ are strictly larger than $0$. In the case of the Huberised TV cost this can only be proven in a discretised setting. Please see [@tuomov-interior] for details. - The existence of solutions with $\mu=0$, that is without elliptic regularisation, is also proven in [@tuomov-interior]. Note that here, we focus on the $\mu>0$ case since the elliptic regularity is required for proving the existence of Lagrange multipliers in the next section. Lagrange multipliers {#sec:Lagrange} ==================== In this section we prove the existence of Lagrange multipliers for the learning problem and derive an optimality system that characterizes its solution. Moreover, a gradient formula for the reduced cost functional is obtained, which plays an important role in the development of fast solution algorithms for the learning problems (see Section \[sec:denoising\]). In what follows all proofs are presented for the $\TGV^2$ regularisation case, that is $R_{\alpha,\beta}^{\gamma}=\TGV^{2,\gamma}_{\alpha,\beta}$. However, possible modifications to cope with the ICTV model will also be commented. We start by investigating the differentiability of the solution operator. Differentiability of the solution operator {#sec:diff} ------------------------------------------ We recall that the $\TGV^2$ denoising problem is given by $$u=(v,w)= \argmin_{BV \times BD} \left\{ \frac{1}{2}\int_\Omega |v-f|^2 + \int_\Omega \alpha |Dv-w|_\gamma + \int_\Omega \beta |E w|_\gamma \right\}.$$ Using an elliptic regularization we then get $$\label{eq: denoising problem in diff proof} u= \argmin_{H^1(\Omega) \times \mathbb H^1(\Omega)} \left\{ a(u,u)+ \frac{1}{2} \int_\Omega |v-f|^2 + \int_\Omega \alpha |Dv-w|_\gamma + \int_\Omega \beta |E w|_\gamma \right\},$$ where $a(u,u)= \mu \left( \|v\|_{H^1}^2 + \|w\|_{\mathbb H^1}^2 \right)$. A necessary and sufficient optimality condition for the latter is then given by the following variational equation $$\begin{gathered} \label{eq: denoising problem in diff proof var form} a(u, \Psi)+ \int_\Omega \alpha h_\gamma(Dv-w)(D \phi- \varphi) \,dx\\ + \int_\Omega \beta h_\gamma(E w) E \varphi \,dx +\int_\Omega(v-f)\phi \,dx=0, \text{ for all } \Psi \in U,\end{gathered}$$ where $\Psi=(\phi,\varphi)$ and $U=H^1(\Omega) \times \mathbb H^1(\Omega)$. The solution operator $S: \mathbb R^2 \mapsto U$, which assigns to each pair $(\alpha, \beta) \in \mathbb R^{2}$ the corresponding solution to the denoising problem , is Fréchet differentiable and its derivative is characterized by the unique solution $z=S'(\alpha, \beta)[\theta_1, \theta_2] \in U$ of the following linearized equation: $$\begin{gathered} \label{eq: linearized equation} a(z, \Psi)+ \int_\Omega \theta_1 \ h_\gamma(Dv-w)(D \phi- \varphi) \,dx\\ + \int_\Omega \alpha h'_\gamma(Dv-w)(Dz_1-z_2)(D \phi- \varphi) \,dx + \int_\Omega \theta_2 \ h_\gamma(E w) E \varphi \,dx\\+ \int_\Omega \beta h'_\gamma(E w) E z_2 E \varphi \,dx + \int_\Omega z_1 \phi \,dx=0, \text{ for all } \Psi \in U.\end{gathered}$$ Thanks to the ellipticity of $a(\cdot, \cdot)$ and the monotonicity of $h_\gamma$, existence of a unique solution to the linearized equation follows from the Lax-Milgram theorem. Let $\xi:=u^+- u -z$, where $u=S(\alpha, \beta)$ and $u^+=S(\alpha+\theta_1, \beta+ \theta_2)$. Our aim is to prove that $\| \xi \|_U= o(|\theta|).$ Combining the equations for $u^+$, $u$ and $z$ we get that $$\begin{gathered} a(\xi, \Psi)+ \int_\Omega (\alpha +\theta_1) \ h_\gamma(Dv^+-w^+)(D \phi- \varphi) \,dx- \int_\Omega \alpha \ h_\gamma(Dv-w)(D \phi- \varphi) \,dx\\ - \int_\Omega \theta_1 \ h_\gamma(Dv-w)(D \phi- \varphi) \,dx - \int_\Omega \alpha h'_\gamma(Dv-w)(Dz_1-z_2)(D \phi- \varphi) \,dx\\ + \int_\Omega (\beta+\theta_2) h_\gamma(E w^+) E \varphi \,dx- \int_\Omega \beta h_\gamma(E w) E \varphi \,dx\\ - \int_\Omega \theta_2 \ h_\gamma(E w) E \varphi \,dx - \int_\Omega \beta \ h'_\gamma(E w) E z_2 E \varphi \,dx +2 \int_\Omega \xi_1 \phi \,dx=0, \text{ for all } \Psi \in U,\end{gathered}$$ where $\xi:=(\xi_1,\xi_2) \in H^1(\Omega) \times \mathbb H^1(\Omega)$. Adding and subtracting the terms $$\int_\Omega \alpha h'_\gamma(Dv-w)(D \delta_v - \delta_w)(D \phi- \varphi) \,dx$$ and $$\int_\Omega \beta h'_\gamma(E w)E \delta_w: E \varphi \,dx,$$ where $\delta_v:=v_{\alpha+\theta}-v$ and $\delta_w:=w_{\alpha+\theta}-w$, we obtain that $$\begin{gathered} a(\xi, \Psi)+ \int_\Omega \alpha h'_\gamma(Dv-w)(D \xi_1- \xi_2)(D \phi- \varphi)\\+ \int_\Omega \beta h'_\gamma(E w) E \xi_2 :E \varphi \,dx +2 \int_\Omega \xi_1 \phi \,dx\\ =- \int_\Omega \alpha \left[ h_\gamma(Dv^+-w^+) -h_\gamma(Dv-w)- h'_\gamma(Dv-w)(D \delta_v - \delta_w)\right] (D \phi- \varphi)\\ - \int_\Omega \theta_1 \ \left[ h_\gamma(Dv^+-w^+)-h_\gamma(Dv-w) \right] (D \phi- \varphi) \,dx\\ - \int_\Omega \beta \left[h_\gamma(E w^+)-h_\gamma(E w)- h'_\gamma(E w) E \delta_w \right] :E \varphi \,dx\\ - \int_\Omega \theta_2 \ \left[ h_\gamma(E w_{\alpha+\theta})-h_\gamma(E w) \right]: E \varphi \,dx, \text{ for all } \Psi \in U.\end{gathered}$$ Testing with $\Psi=\xi$ and using the monotonicity of $h_\gamma'(\cdot)$ we get that $$\begin{gathered} \|\xi\|_U \leq C \left\{ |\alpha| \left\| h_\gamma(Dv^+-w^+)-h_\gamma(Dv-w)- h'_\gamma(Dv-w)(D \delta_v - \delta_w) \right\|_{L^2} \right.\\ +|\theta_1| \left\| h_\gamma(Dv^+-w^+)-h_\gamma(Dv-w) \right\|_{L^2}\\ \left. +|\beta| \left\| h_\gamma(E w^+)-h_\gamma(E w)- h'_\gamma(E w) E \delta_w \right \|_{L^2} \right.\\ \left. +|\theta_2| \left\| h_\gamma(E w_{\alpha+\theta})-h_\gamma(E w) \right\|_{L^2} \right\},\end{gathered}$$ for some generic constant $C >0$. Considering the differentiability and Lipschitz continuity of $h_\gamma'(\cdot)$, it then follows that $$\begin{gathered} \label{eq: differentiability estimate} \|\xi\|_U \leq C \left( |\alpha|~ o (\left\| u^+-u \right\|_{1,p}) +|\theta_1| \left\| u_{\alpha+\theta}-u \right \|_{U} \right.\\ +\left. |\beta|~ o (\left\| w^+-w \right\|_{1,p}) +|\theta_2| \left\| w_{\alpha+\theta}-w \right \|_{\mathbb H^1(\Omega)} \right),\end{gathered}$$ where $\| \cdot \|_{1,p}$ stands for the norm in the space $\mathbb W^{1,p}(\Omega)$. From regularity results for second order systems (see [@Groeger1989 Thm. 1, Rem. 14]), it follows that $$\begin{aligned} \left\| u^+-u \right\|_{1,p} & \leq L |\theta| \left( \|\mathrm{Div}~ h_\gamma(Dv -w)\|_{-1,p}+ \|h_\gamma(Dv -w)\|_{-1,p}+\|\mathrm{Div}~ h_\gamma(E w)\|_{-1,p} \right)\\ & \leq L |\theta| \left(2 \|h_\gamma(Dv -w)\|_{L^\infty}+\|h_\gamma(E w)\|_{L^\infty} \right)\\ & \leq \tilde L |\theta|,\end{aligned}$$ since $|h_\gamma (\cdot)| \leq 1$. Inserting the latter in estimate , we finally get that $$\| \xi \|_U= o(|\theta|).$$ The Fréchet differentiability proof makes use of the quasilinear structure of the $\TGV^2$ variational form, making it difficult to extend to the ICTV model without further regularisation terms. For the latter, however, a Gateaux differentiability result may be obtained using the same proof technique as in [@de2013image]. The adjoint equation {#sec:optsys} -------------------- Next, we use the Lagrangian formalism for deriving the adjoint equations for both the $\TGV^2$ and ICTV learning problems. Existence of a solution to the adjoint equation then follows from the well-posedness of the linearized equation. Defining the Lagrangian associated to $\TGV^2$ learning problem by: $$\begin{gathered} \mathcal L(v,w,\alpha,\beta,p_1,p_2) = F(u) +\mu (v, p_1)_{H^1}+ \mu (w, p_2)_{\mathbb H^1}\\ + \int_\Omega \alpha h_\gamma (Dv - w)(D p_1- p_2)+ \int_\Omega \beta h_\gamma(E w) E p_2 + \int_\Omega (v-f) p_1,\end{gathered}$$ and taking the derivative with respect to the state variable $(v,w)$, we get the necessary optimality condition $$\begin{gathered} \mathcal L_{(u,v)}'(u,v,\alpha,\beta,p_1,p_2)[(\delta_v, \delta_w)]= F'(u)\delta_u +\mu (p_1, \delta_v)_{H^1}+ \mu (p_2, \delta_w)_{\mathbb H^1}\\ + \int_\Omega \alpha h_\gamma' (Dv - w)(D \delta_v- \delta_w)(D p_1- p_2)\\+ \int_\Omega \beta h_\gamma' (E w) E \delta_w E p_2 + \int_\Omega p_1 \delta_v=0.\end{gathered}$$ If $\delta_w=0$, then $$\mu (p_1, \delta_v)_{H^1}+ \int_\Omega \alpha h_\gamma' (Dv - w)(D p_1- p_2) D \delta_v+ \int_\Omega p_1 \delta_v=-\nabla_vF(u)\delta_v,$$ whereas if $\delta_v=0$, then $$\begin{gathered} \mu (p_2, \delta_w)_{\mathbb H^1} - \int_\Omega \alpha h_\gamma' (Dv - w)(D p_1- p_2) \delta_w\\+ \int_\Omega \beta h_\gamma' (E w) \ E p_2 \ E \delta_w =-\nabla_w F(u)\delta_w.\end{gathered}$$ Existence of a unique solution then follows from the transposition method, since the linearised equation is well-posed. For the ICTV model it is possible to proceed formally with the Lagrangian approach. We recall that a necessary and sufficient optimality condition for the ICTV functional is given by $$\begin{gathered} \mu (u, \phi)_{H^1}+ \mu (\nabla v, \nabla \varphi)_{\mathbb H^1} + \int_\Omega \alpha h_\gamma (Du - \nabla v)(D \phi- \nabla \varphi)\\+ \int_\Omega \beta h_\gamma(D \nabla v) D \nabla \varphi + \int_\Omega (u-f)\phi=0, \text{ for all }(\phi, \varphi) \in H^1(\Omega) \times \mathbb H^1(\Omega)\end{gathered}$$ and the correspondent Lagrangian functional $\mathcal L$ is given by $$\begin{gathered} \mathcal L(u,v,\alpha,\beta,p_1,p_2) = F(u) +\mu (u, p_1)_{H^1}+ \mu (\nabla v, \nabla p_2)_{\mathbb H^1}\\ + \int_\Omega \alpha h_\gamma (Du - \nabla v)(D p_1- \nabla p_2)+ \int_\Omega \beta h_\gamma(D \nabla v) D \nabla p_2 + \int_\Omega (u-f) p_1.\end{gathered}$$ Deriving the Lagrangian with respect to the state variable $(u,v)$ and setting it equal to zero yields $$\begin{gathered} \mathcal L_{(u,v)}'(u,v,\alpha,\beta,p_1,p_2)[(\delta_u, \delta_v)]= F'(u)\delta_u +\mu (p_1, \delta_u)_{H^1}+ \mu (\nabla p_2, \nabla \delta_v)_{\mathbb H^1}\\ + \int_\Omega \alpha h_\gamma' (Du - \nabla v)(D \delta_u- \nabla \delta_v)(D p_1- \nabla p_2)\\+ \int_\Omega \beta h_\gamma' (D \nabla v) D \nabla \delta_v D \nabla p_2 + \int_\Omega p_1 \delta_u=0.\end{gathered}$$ By taking succesively $\delta_v=0$ and $\delta_u=0$, the following system is obtained $$\begin{gathered} \mu (p_1, \delta_u)_{H^1}+ \int_\Omega \alpha h_\gamma' (Du - \nabla v)(D p_1- \nabla p_2) D \delta_u+ \int_\Omega p_1 \delta_u=-F'(u)\delta_u.\end{gathered}$$ $$\begin{gathered} \mu (\nabla p_2, \nabla \delta_v)_{\mathbb H^1} + \int_\Omega \alpha h_\gamma' (Du - \nabla v)(D p_1- \nabla p_2) \nabla \delta_v\\+ \int_\Omega \beta h_\gamma' (D \nabla v) D \nabla p_2 D \nabla \delta_v =0.\end{gathered}$$ Optimality condition -------------------- Using the differentiability of the solution operator and the well-posedness of the adjoint equation, we derive next an optimality system for the characterization of local minima of the bilevel learning problem. Besides the optimality condition itself, a gradient formula arises as byproduct, which is of importance in the design of solution algorithms for the learning problems. \[thm: optimality system\] Let $(\bar \alpha, \bar \beta) \in \mathbb R^2_+$ be a local optimal solution for problem . Then there exist Lagrange multipliers $\Pi \in U$ and $\lambda_1, \lambda_2 \in L^2(\Omega)$ such that the following system holds: $$\begin{gathered} a(u, \Psi)+\alpha \int_\Omega h_\gamma(Dv-w)(D \phi- \varphi) \,dx\\ + \beta \int_\Omega h_\gamma(E w) E \varphi \,dx +2 \int_\Omega(v-f)\phi \,dx=0, \text{ for all } \Psi \in H^1(\Omega) \times \mathbb H^1(\Omega),\end{gathered}$$ $$\begin{gathered} \label{eq: adjoint equation TGV} a(\Pi, \Psi)+\alpha \int_\Omega h_\gamma' (Dv-w)(D p_1-p_2)(D \phi- \varphi) \,dx\\ + \beta \int_\Omega h_\gamma' (E w) \ E p_2 \ E \varphi \,dx +2 \int_\Omega p_1 \phi \,dx=-F_u(u)[\Psi], \text{ for all } \Psi \in H^1(\Omega) \times \mathbb H^1(\Omega),\end{gathered}$$ $$\lambda_1= \int_\Omega h_\gamma (Dv-w)(D p_1 -p_2)$$ $$\lambda_2= \int_\Omega h_\gamma (Ew) \ E p_2$$ $$\lambda_1 \geq 0, \qquad \lambda_2 \geq 0$$ $$\lambda_1 \cdot \bar \alpha = \lambda_2 \cdot \bar \beta=0.$$ Consider the reduced cost functional $\mathcal F(\alpha, \beta)=F(u(\alpha, \beta)).$ The bilevel optimization problem can then be formulated as $$\begin{aligned} & \min_{(\alpha, \beta) \in C} \mathcal F(\alpha, \beta),\end{aligned}$$ where $\mathcal F: \mathbb R^{2} \to \mathbb R$ and $C$ corresponds to the positive orthant in $\mathbb R^2$. From [@ZoweK1979 Thm. 3.1], there exist multipliers $\lambda_1, \lambda_2$ such that $$\begin{aligned} \lambda_1= \nabla_\alpha \mathcal F(\bar \alpha, \bar \beta)\\ \lambda_2= \nabla_\beta \mathcal F(\bar \alpha, \bar \beta)\\ \lambda_1 \geq 0, \quad \lambda_2 \geq 0\\ \lambda_1 \cdot \bar \alpha = \lambda_2 \cdot \bar \beta=0,\end{aligned}$$ By taking the derivative with respect to $(\alpha, \beta)$ and denoting by $u'$ the solution to the linearized equation we get, together with the adjoint equation , that $$\begin{aligned} \mathcal F'(\alpha, \beta)[\phi]& =F_u(u)u'(\alpha,\beta)[\phi]\\ &= -a(\Pi,u')- \alpha \int_\Omega h_\gamma'(Dv-w)(D p_1 -p_2 )(D v'-w')\\& \qquad -\beta \int_\Omega h_\gamma'(E w) E p_2 \ E w'-2 \int_\Omega p_1 v'\\ &= -a(u',\Pi) - \alpha \int_\Omega h_\gamma'(Dv-w)(D v'-w') (D p_1 -p_2 )\\& \qquad -\beta \int_\Omega h_\gamma'(E w) E w' \ E p_2-2 \int_\Omega v' p_1\end{aligned}$$ which, taking into account the linearized equation, yields $$\label{eq: derivative of reduced cost} \mathcal F'(\alpha, \beta)[\phi]=\phi_1 \int_\Omega h_\gamma(Dv-w)(D p_1 -p_2 )+ \phi_2 \int_\Omega h_\gamma(E w)E p_2.$$ Altogether we proved the result. From the existence result (see Remark \[rem: existence\]), we actually know that, under some assumptions, $\bar \alpha$ and $\bar \beta$ are strictly greater than zero. This implies that the multipliers $\lambda_1=\lambda_2=0$ and the problem is of unconstrained nature. This plays an important role in the design of solution algorithms, since only a mild treatment of the constraints has to be taken into account, as will be showed in Section 6. Numerical algorithms {#sec:denoising} ==================== In this section we propose a second order quasi-Newton method for the solution of the learning problem with scalar regularisation parameters. The algorithm is based on a BFGS update, preserving the positivity of the iterates through the line search strategy and updating the matrix cyclically depending on the satisfaction of the curvature condition. For the solution of the lower level problem, a semismooth Newton method with a properly modified Jacobi matrix is considered. Moreover, warm initialisation strategies have to be taken into account in order to get convergence for the $\TGV^2$ problem. The developed algorithm is also extended to a simple linear polynomial case. BFGS algorithm {#sec:bfgs} -------------- Thanks to the gradient characterization obtained in Theorem \[thm: optimality system\], we next devise a BFGS algorithm to solve the bilevel learning problems. We employ a few technical tricks to ensure convergence of the classical method. In particular, for numerical stability we need to avoid the boundary of the constraint set on the parameters, so we pick $0 < \theta < \Theta$, considered numerically almost zero or infinity, respectively, and require the box constraints $$\label{eq:bfgs-constr} \theta \le \alpha, \beta \le \Theta. $$ We also limit the step length to get at most a fraction closer to the boundary. As we show in [@tuomov-interior] the solution is in the interior for the regularisation and cost functionals we are interested in. Below this limit, we use Armijo line search. Moreover, the good behaviour of the BFGS method depends upon the BFGS matrix staying positive definite. This would be ensured by the Wolfe conditions, but because of our step length limitation, the curvature condition is not necessarily satisfied. (The Wolfe conditions are guaranteed to be satisfied for some step length $\sigma$, if our domain is unbounded, but the range where the step satisfies the criterion, may be beyond our maximum step length, and is not necessarily satisfied closer to the current point.) Instead we skip the BFGS update if the curvature is negative. Overall our learning algorithm may be written as follows. \[algorithm:bfgs-learn\] Pick Armijo line search constant $c$, and target residual $\rho$. Pick initial iterate $(\alpha^0,\beta^0)$. Solve the denoising problem for $(\alpha,\beta)=(\alpha^0,\beta^0)$, yielding $u^0$. Initialise $B^1=I$. Set $i \defeq 0$, and iterate the following steps: 1. \[step:adjoint\] Solve the adjoint equation for $\Pi^i$, and calculate $\grad \mathcal F (\alpha^i,\beta^i)$ from . 2. If $i \ge 2$ do the following: 1. Set $s \defeq (\alpha^i,\beta^i)-(\alpha^{i-1}, \beta^{i-1})$, and $r \defeq \grad \mathcal F(\alpha^i,\beta^i)-\grad \mathcal F(\alpha^{i-1},\beta^{i-1})$. 2. Perform the BFGS update $$B^i \defeq \begin{cases} B^{i-1}, & s^Tr < 0, \\ B^{i-1} - \frac{B^{i-1} s \otimes B^{i-1}s}{t^T B^{i-1} s} + \frac{r \otimes r}{s^Tr} & s^Tr \ge 0. \end{cases}$$ 3. Compute $\delta_{\alpha, \beta}$ from $$B^i \delta_{\alpha, \beta} = g^i.$$ 4. Initialise $\sigma \defeq \min\{1, \sigma_{\max}/2\}$, where $$\sigma_{\max} \defeq \max \{ \sigma \ge 0 \mid (\alpha^i, \beta^i)+\sigma \delta_{\alpha, \beta} \text{ satisfies } \eqref{eq:bfgs-constr}\}.$$ Repeat the following: 1. \[item:linesearch-1\] Let $(\alpha_\sigma, \beta_\sigma) \defeq (\alpha^i, \beta^i)+\sigma \delta_{\alpha, \beta}$, and solve the denoising problem for $(\alpha, \beta)=(\alpha_\sigma, \beta_\sigma)$, yielding $u_\sigma$. 2. If the residual $\norm{(\alpha_\sigma, \beta_\sigma) - (\alpha^i, \beta^i)}/\norm{(\alpha_\sigma, \beta_\sigma)} < \rho$ do the following: 1. If $\min_\sigma \mathcal F(\alpha_\sigma, \beta_\sigma) < \mathcal F(\alpha^i, \beta^i)$ over all $\sigma$ tried, choose $\sigma^*$ the minimiser, set $(\alpha^{i+1}, \beta^{i+1}) \defeq (\alpha_{\sigma^*}, \beta_{\sigma^*})$, $u^{i+1} \defeq u_{\sigma^*}$, and continue from Step \[step:res-check\] 2. Otherwise end the algorithm with solution $(\alpha^*, \beta^*) \defeq (\alpha^i, \beta^i)$. 3. Otherwise, if Armijo condition $\mathcal F(\alpha_\sigma, \beta_\sigma) \le \mathcal F(\alpha^i, \beta^i) + \sigma c \grad \mathcal F(\alpha^i,\beta^i)^T \delta_{\alpha, \beta}$ holds, set $(\alpha^{i+1}, \beta^{i+1}) \defeq (\alpha_{\sigma}, \beta_{\sigma})$, $u^{i+1} \defeq u_{\sigma}$, and continue from Step \[step:res-check\]. 4. In all other cases, set $\sigma \defeq \sigma/2$ and continue from Step \[item:linesearch-1\]. 5. \[step:res-check\] If the residual $\norm{(\alpha^{i+1}, \beta^{i+1}) - (\alpha^i, \beta^i)}/\norm{(\alpha^{i+1}, \beta^{i+1})} < \rho$, end the algorithm with $(\alpha^* , \beta^*) \defeq (\alpha^{i+1}, \beta^{i+1})$. Otherwise continue from Step \[step:adjoint\] with $i \defeq i+1$. Step (4) ensures that the iterates remain feasible, without making use of a projection step. This is justified since it’s been analytically proved that the optimal parameters are greater than zero (see [@tuomov-interior]). An infeasible semi-smooth Newton method {#sec:seminewton} --------------------------------------- In variational form, the $\TGV^2$ denoising problem can be written as $$\mu \int_\Omega (Dv \cdot D \phi+ v \phi)+ \int_\Omega \alpha h_\gamma (Dv - w) D \phi + \int_\Omega (v-f) \phi =0, \quad \forall \phi \in H^1(\Omega)$$ $$\begin{gathered} \mu \int_\Omega (Ew : E \varphi+ w \varphi) - \int_\Omega \alpha h_\gamma (Dv - w) D \varphi \\+ \int_\Omega \beta h_\gamma (E w) \ E \varphi =0, \quad \forall \varphi \in \mathbb H^1(\Omega)\end{gathered}$$ or, in general abstract primal-dual form, as \[eq:ssn-sys\] $$\begin{gathered} \label{eq:ssn-sys1} L u + \sum_{i=1}^{\NA} A_j^* q_j = f \quad \text{ in } \Omega \\ \label{eq:ssn-sys2} \max\{1/\gamma, \abs{[A_j u](x)}_2\} q_j(x) - \alpha_j [A_j u](x) = 0 \text{ a.e. in }\Omega, \quad j=1,\ldots,\NA.\end{gathered}$$ where $L \in \mathcal L (\SPACEuHdual,\SPACEuHdual')$ is a second order linear elliptic operator, $A_j, ~j=1, \dots, N$, are linear operators acting on $u$ and $q_j(x), ~j=1, \dots, N$, correspond to the dual multipliers. Let us set $$\maxO_j(u) \defeq \max\{1/\gamma, \abs{[A_j u](x)}_2\}.$$ Let us also define the diagonal application $\DiagO(u): \SPACEuLtwo \to \SPACEuLtwo$ by $$[\DiagO(u) q](x) = u(x) q(x), \quad (x \in \Omega)$$ We may derive $\grad_u [\DiagO(\maxO_j(u)) q_j]$ being defined by $$\grad_u [\DiagO(\maxO_j(u)) p_j] = A_j^* \DiagO(q_j) \nO(A_j u) \quad \text{where} \quad \nO(z) \defeq \begin{cases} 0, & \abs{z(x)}_2 < 1/\gamma \\ \frac{z(x)} {\abs{z(x)}_2}, & \abs{z(x)}_2 \geq 1/\gamma. \\ \end{cases}$$ Then , may be written as $$\begin{gathered} \notag L u + \sum_{i=1}^{\NA} A_j^* q_j = f \quad \text{ in } \Omega \\ \notag \DiagO(\maxO_j(u)) q_j - \alpha_j A_j u = 0, \quad \text{ a.e. in } \Omega, \quad (j=1,\ldots,\NA).\end{gathered}$$ Linearising, we obtain the system $$\label{eq:ssn} \tag{SSN-1} \begin{pmatrix} L & A_1^* & \ldots & A_\NA^* \\ - \alpha_1 A_1 + \nO(A_1 u) \DiagO(q_1) A_1 & \DiagO(\maxO_j(u)) & 0 & 0 \\ \vdots & 0 & \ddots & 0 \\ - \alpha_\NA A_\NA + \nO(A_\NA u) \DiagO(q_\NA) A_\NA & 0 & 0 & \DiagO(\maxO_\NA(u)) \\ \end{pmatrix} \begin{pmatrix} \Step u \\ \Step q_1 \\ \vdots \\ \Step q_\NA \\ \end{pmatrix} = R$$ where $$R \defeq \begin{pmatrix} -L u - \sum_{i=1}^{\NA} A_j^* q_j + f \\ \alpha_1 A_1 u - \DiagO(\maxO_1(u)) q_1\\ \vdots \\ \alpha_\NA A_\NA u - \DiagO(\maxO_\NA(u)) q_\NA \end{pmatrix}.$$ The semi-smooth Newton method solves at a current iterate $(u^i, q_1^i, \ldots q_\NA^i)$. It then updates $$\tag{SSN-2} \label{eq:ssn-step} (u^{i+1}, \tilde q_1^{i+1}, \ldots \tilde q^{i+1}_\NA) := (u^i + \tau \Step u, q_1^i + \tau \Step q_1, q_\NA^i + \tau \Step q_\NA),$$ for a suitable step length $\tau$, allowing $\tilde q^{i+1}$ to become infeasible in the process. That is, it may hold that $\abs{\tilde q_j^{i+1}(x)}_2 > \alpha_j$, which may lead to non-descent directions. In order to globalize the method, one projects $$\label{eq:ssn-proj} \tag{SSN-3} q_j^{i+1} \defeq \projO(\tilde q_j^{i+1}; \alpha_j), \quad \text{where} \quad \projO(q, \alpha)(x) \defeq \sign(q(x)) \min\{\alpha, \abs{q(x)}\},$$ in the building of the Jacobi matrix. Following [@hintermuller2006infeasible; @sun1997newton], it can be shown that a discrete version of the method – converges globally and locally superlinearly near a point where the subdifferentials of the operator on $(u, q_1, \ldots q_\NA)$ corresponding are non-singular. Further dampening as in [@hintermuller2006infeasible] guarantees local superlinear convergence at any point. We do not represent the proof, as going into the discretisation and dampening details would expand this work considerably. \[rem:ssn-simplify\] The system can be simplified, which is crucial to obtain acceptable performance with $\TGV^2$. Indeed observe that $B$ is invertible, so we may solve $\Step u$ from $$\label{eq:delta-u-upd} B \Step u = R_1 - \sum_{j=1}^N A_j^* \Step q_j.$$ Thus we may simplify $\Step u$ out of , and only solve for $\Step q_1, \ldots, \Step q_N$ using a reduced system matrix. Finally we calculate $\Step u$ from . For the denoising sub-problem we use the method – with the reduced system matrix of Remark \[rem:ssn-simplify\]. Here, we denote by $z$ in the case of TGV$^2$ the parameters $$z=(v,w),$$ and in the case of ICTV $$z=(u,v).$$ For the calculation of the step length $\tau$, we use Armijo line search with parameter $c=1\ee^{-4}$. We end the SSN iterations when $$\tau\frac{\norm{\Step y^i}}{\max\{1,\norm{y^i}\}} \le 1\ee^{-5},$$ where $\Step y^i=(\Step z^i,\Step q_1^i, \ldots, \Step q_N^i)$, and $y^i=(z^i, q_1^i, \ldots, q_N^i)$. Warm initialisation {#sec:numdisc} ------------------- In our numerical experimentation we generally found Algorithm \[algorithm:bfgs-learn\] to perform well for learning the regularisation parameter for $\TV$ denoising as was done in [@de2013image]. For learning the two (or even more) regularisation parameters for $\TGV^2$ denoising, we found that a warm initialisation is needed to obtain convergence. More specifically, we use $\TV$ as an aid for discovering both the initial iterate $(\alpha^0,\beta^0)$ as well as the initial BFGS matrix $B^1$. This is outlined in the following algorithm. \[algorithm:bfgs-learn-tgv2\] Pick a heuristic factor $\delta_0 > 0$. Then do the following: 1. Solve the corresponding problem for $\TV$ using Algorithm \[algorithm:bfgs-learn\]. This yields optimal $\TV$ denoising parameter $\alpha_\TV^*$, as well as the BFGS estimate $B_\TV$ for $\grad^2 \mathcal F (\alpha_\TV^*)$. 2. Run Algorithm \[algorithm:bfgs-learn\] for $\TGV^2$ with initialisation $(\alpha^0,\beta^0) \defeq (\alpha_\TV^* \delta_0, \alpha_\TV^*)$, and initial BFGS matrix $B^1 \defeq \mathrm{diag}(B_\TV \delta_0, B_\TV)$. With $\Omega=(0, 1)^2$, we pick $\delta_0=1/\pixels$, where the original discrete image has $\pixels \times \pixels$ pixels. This corresponds to the heuristic [@tuomov-dtireg; @tuomov-phaserec] that if $\pixels \approx 128$ or $256$ and the discrete image is mapped into the corresponding domain $\Omega=(0, \pixels)^2$ directly (corresponding to spatial step size of one in the discrete gradient operator), then $\beta \in (\alpha, 1.5 \alpha)$ tends to be a good choice. We will later verify this through the use of our algorithms. Now, if $f \in \BVspace((0, \pixels)^2)$ is rescaled to $\BVspace((0, 1)^2)$, i.e. $\tilde f(x) \defeq f(x/\pixels)$, then with $\tilde u(x) \defeq u(x/\pixels)$ and $\tilde w(x) \defeq w(x/\pixels)/\pixels$, we have $$\begin{gathered} \frac{1}{2}\norm{f-u}_{L^2((0, \pixels)^2)}^2 +\alpha\norm{Du-w}_{\Meas((0, \pixels)^2; \R^2)} +\beta\norm{Ew}_{\Meas((0, \pixels)^2; \R^{2 \times 2})} \\ = n^2\left( \frac{1}{2}\norm{\tilde f-\tilde u}_{L^2((0,1)^2)}^2 +n\alpha\norm{D\tilde u-\tilde w}_{\Meas((0,1)^2; \R^2)} +n^2\beta\norm{E\tilde w}_{\Meas((0,1)^2; \R^{2 \times 2})} \right). \end{gathered}$$ This introduces the factor $1/\pixels=\abs{\Omega}^{-1/2}$ between rescaled $\alpha$, $\beta$. Experiments {#sec:experiments} =========== In this section we present some numerical experiments to verify the theoretical properties of the bilevel learning problems and the efficiency of the proposed solution algorithms. In particular, we exhaustively compare the performance of the new proposed cost functional with respect to well-known quality measures, showing a better behaviour of the new cost for the chosen tested images. The performance of the proposed BFGS algorithm, combined with the semismooth Newton method for the lower level problem, is also examined. Gaussian denoising {#sec:denoising} ------------------ We tested Algorithm \[algorithm:bfgs-learn\] for $\TV$ and Algorithm \[algorithm:bfgs-learn-tgv2\] for $\TGV^2$ Gaussian denoising parameter learning on various images. Here we report the results for two images, the parrot image in Figure \[fig:res-dataset2:kodim23gray-crop\], and the geometric image in Figure \[fig:res-dataset11\]. We applied synthetic noise to the original images, such that the PSNR of the parrot image is $24.7$, and the PSNR of the geometric image is $24.8$. In order to learn the regularisation parameter $\alpha$ for $\TV$, we picked initial $\alpha^0=0.1/\pixels$. For $\TGV^2$ initialisation by $\TV$ was used as in Algorithm \[algorithm:bfgs-learn\]. We chose the other parameters of Algorithm \[algorithm:bfgs-learn\] as $c=1\ee^{-4}$, $\rho=1\ee^{-5}$, $\theta=1\ee{-8}$, and $\Theta=10$. For the SSN denoising method the parameters $\gamma=100$ and $\mu=1\ee^{-10}$ were chosen. We have included results for both the $L^2$-squared cost functional $\costltwo$ and the Huberised total variation cost functional $\costhubertv$. The learning results are reported in Table \[table:res-dataset2\] for the parrot images, and Table \[table:res-dataset11\] for the geometric image. The denoising results with the discovered parameters can be found in the aforementioned Figure \[fig:res-dataset2\] and Figure \[fig:res-dataset11\]. We report the resulting optimal parameter values, the cost functional value, PSNR, SSIM [@wang2004ssim], as well as the number of iterations taken by the outer BFGS method. Our first observation is that all approaches successfully learn a denoising parameter that gives a good-quality denoised image. Secondly, we observe that the gradient cost functional $\costhubertv$ performs visually and in terms of SSIM significantly better for $\TGV^2$ parameter learning than the cost functional $\costltwo$. In terms of PSNR the roles are reversed, as should be, since the $\costltwo$ is equivalent to PSNR. This again confirms that PSNR is a poor quality measure for images. For $\TV$ there is no significant difference between different cost functionals in terms of visual quality, although the PSNR and SSIM differ. We also observe that the optimal $\TGV^2$ parameters $(\alpha^*, \beta^*)$ generally satisfy $\beta^*/\alpha^* \in (0.75, 1.5)/\pixels$. This confirms the earlier observed heuristic that if $\pixels \approx 128,\, 256$ then $\beta \in (1, 1.5) \alpha$ tends to be a good choice. As we can observe from Figure \[fig:res-dataset2\] and Figure \[fig:res-dataset11\], this optimal $\TGV^2$ parameter choice also avoids the stair-casing effect that can be observed with $\TV$ in the results. [0.47]{} ![Cost functional value versus $(\alpha, \beta)$ for $\TGV^2$ denoising, for the parrot test images, for both and cost functionals. The illustrations are contour plots of function value versus $(\alpha, \beta)$. []{data-label="fig:landscape-tgv2"}](resimg/SCAN-tgv2-dataset2-huberonly "fig:"){width="\textwidth"} [0.47]{} ![Cost functional value versus $(\alpha, \beta)$ for $\TGV^2$ denoising, for the parrot test images, for both and cost functionals. The illustrations are contour plots of function value versus $(\alpha, \beta)$. []{data-label="fig:landscape-tgv2"}](resimg/SCAN-tgv2-dataset2-l2 "fig:"){width="\textwidth"} In Figure \[fig:landscape-tgv2\], we have plotted by the red star the discovered regularisation parameter $(\alpha^*, \beta^*)$ reported in Figure \[fig:res-dataset2\]. Studying the location of the red star, we may conclude that Algorithm \[algorithm:bfgs-learn\] and Algorithm \[algorithm:bfgs-learn-tgv2\] manage to find a nearly optimal parameter in very few BFGS iterations. \ [lll|lllll|l]{} Denoise & Cost & Initial $(\alpha, \beta)$ & Result $(\alpha^*, \beta^*)$ & Cost & SSIM & PSNR & Its. & Fig.\ [synth]{} [Original image]{}\ [lll|lllll|l]{} Denoise & Cost & Initial $\alphavec$ & Result $\alphavec^*$ & Value & SSIM & PSNR & Its. & Fig.\ Statistical testing ------------------- To obtain a statistically significant outlook to the performance of different regularisers and cost functionals, we made use of the Berkeley segmentation dataset BSDS300 [@MartinFTM01], displayed in Figure \[fig:bsds300\]. We resized each image to 128 pixels on its shortest edge, and take the $128\times 128$ top-left square of the image. To this data set, we applied pixelwise Gaussian noise of variance $\sigma^2=2,10$, and $20$. We tested the performance of both cost functionals, $\costhubertv$ and $\costltwo$, as well as the $\TGV^2$, $\ICTV$, and $\TV$ regularisers on this dataset, for all noise levels. In the first instance, reported in Figures \[fig:bsds300-individual-huber-2\]–\[fig:bsds300-individual-ltwo-20\] (noise levels $\sigma^2=2,20$ only), and Tables \[table:bsds300-individual-2\]–\[table:bsds300-individual-20\], we applied the proposed bi-level learning model on each image individually, to learn the optimal parameters specifically for that image, and a correponding noisy image for all of the noise levels separately. For the algorithm, we use the same parametrisation as in Section \[sec:denoising\]. The figures display the noisy images, and indicate by colour coding the best result as judged by the structural similarity measure SSIM [@wang2004ssim], PSNR, and the objective function value ($\costhubertv$ or $\costltwo$ cost). These criteria are, respectively, the top, middle, and bottom rows of colour-coding squares. Red square indicates that $\TV$ performed the best, green square indicates that $\ICTV$ performed the best, and blue square indicates that $\TGV^2$ performed the best—this is naturally for the optimal parameters for the corresponding regulariser and cost functional discovered by our algorithms. In the tables, we report the information in a more concise numerical fashion, indicating the mean, standard deviation, and median for all the different criteria (SSIM, PSNR, and cost functional value), as well as the number of images for which each regulariser performed the best. We recall that SSIM is normalised to $[0, 1]$, with higher value better. Moreover, we perform a statistical 95% one-tailed paired t-test on each of the criteria, and a pair of regularisers, to see whether any pair of regularisers can be ordered. If so, this is indicated in the last row of each of the tables. Overall, studying the t-test and other data, the ordering of the regularisers appears to be $$\ICTV > \TGV^2 > \TV.$$ This is rather surprising, as in many specific examples, $\TGV^2$ has been observed to perform better than $\ICTV$, see our Figures \[fig:res-dataset2\] and \[fig:res-dataset11\], as well as [@bredies2011tgv; @benning2011higher]. Only when the noise is high, appears $\TGV^2$ to come on par with $\ICTV$ with the $\costhubertv$ cost functional in Figure \[fig:bsds300-individual-huber-20\] and Table \[table:bsds300-individual-20\]. A more detailed study of the results in Figures \[fig:bsds300-individual-huber-2\]–\[fig:bsds300-individual-ltwo-20\] seems to indicate that $\TGV^2$ performs better than $\ICTV$ when the image contains large smooth areas, but $\ICTV$ generally performs better for more chaotic images. This observation agrees with the results in Figures \[fig:res-dataset2\] and \[fig:res-dataset11\], as well as [@bredies2011tgv; @benning2011higher], where the images are of the former type. \[fig:bsds300-individual-huber-2\] \[fig:bsds300-individual-ltwo-2\] \[fig:bsds300-individual-huber-20\] \[fig:bsds300-individual-ltwo-20\] \[table:bsds300-individual-2\] \[table:bsds300-individual-10\] \[table:bsds300-individual-20\] One possible reason for the better performance of $\ICTV$ could be that $\TGV^2$ has more degrees of freedom—in $\ICTV$ we essentially constrain $w=\grad v$ for some function $v$—and therefore overfits to the noisy data, until the noise level becomes so high that overfitting would become too high for any parameter. To see whether this is true, we also performed batch learning, learning a single set of parameters for all images with the same noise level. That is, we studied the model $$\min_{\alphavec} \sum_{i=1}^N F_i(u_{i,\alphavec}) \quad \text{s.t.}\quad u_{i,\alphavec} \in \argmin_{u\in H^1(\Omega)} \frac{1}{2}\norm{f_i-u}_{L^2(\Omega)}^2 + R_{\alphavec}^{\gamma,\mu}(u),$$ with $$F_i(u)=\frac{1}{2}\norm{f_{0,i}-u}^2_{L^2(\Omega)}, \quad\text{or}\quad F_i(u)=\int_\Omega \abs{\grad(f_{0,i}-u)}_\gamma \d x,$$ where $\vec \alpha=(\alpha, \beta)$, $f_1,\ldots,f_N$ are the $N=200$ noisy images with the same noise level, and $f_{0,1},\ldots,f_{0,N}$ the original noise free images. The results are in Figures \[fig:bsds300-batch-huber-2\]–\[fig:bsds300-batch-ltwo-20\] (noise levels $\sigma^2=2,20$ only), and Tables \[table:bsds300-batch-2\]–\[table:bsds300-batch-20\]. The results are still roughly the same as with individual learning. Again, only with high noise in Table \[table:bsds300-batch-20\], does $\TGV^2$ not lose to $\ICTV$. Another interesting observation is that $\TV$ starts to be frequently the best regulariser for individual images, although still statistically does worse than either $\ICTV$ or $\TGV^2$. For the first image of the data set, $\ICTV$ does in all of the Figures \[fig:bsds300-individual-huber-2\]–\[fig:bsds300-batch-ltwo-20\] better than $\TGV^2$, while for the second image, the situation is reversed. We have highlighted these two images for the $\costhubertv$ cost in Figures \[fig:ictv-better-2\]–\[fig:tgv-better-20\], for both noise levels $\sigma=2$ and $\sigma=20$. In the case where $\ICTV$ does better, hardly any difference can be observed by the eye, while for second image $\TGV^2$ clearly has less stair-casing in the smooth areas of the image, especially with the noise level $\sigma=20$. Based on this study, it therefore seems that $\ICTV$ is the most reliable regulariser of the ones tested, when the type of image being processed is unknown, and low SSIM, PSNR or $\costhubertv$ cost functional value is desired. But as can be observed for individual images, it can within large smooth areas exhibit artefacts that are avoided by the use of $\TGV^2$. \[fig:bsds300-batch-huber-2\] \[fig:bsds300-batch-ltwo-2\] \[fig:bsds300-batch-huber-20\] \[fig:bsds300-batch-ltwo-20\] \[table:bsds300-batch-2\] \[table:bsds300-batch-10\] \[table:bsds300-batch-20\] [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=2$[]{data-label="fig:ictv-better-2"}]({img/bsds300/orig_\EXictv}.jpg "fig:"){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=2$[]{data-label="fig:ictv-better-2"}]({resimg/tgv2-huberonly-dataset\EXictvnoise2-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=2$[]{data-label="fig:ictv-better-2"}]({resimg/tgv2-huberonly-bsds300-ALL\EXictvnoise2-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=2$[]{data-label="fig:ictv-better-2"}]({img/bsds300/noisy2_\EXictv}.jpg "fig:"){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=2$[]{data-label="fig:ictv-better-2"}]({resimg/ictv-huberonly-dataset\EXictvnoise2-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=2$[]{data-label="fig:ictv-better-2"}]({resimg/ictv-huberonly-bsds300-ALL\EXictvnoise2-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=20$[]{data-label="fig:ictv-better-20"}]({img/bsds300/orig_\EXictv}.jpg "fig:"){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=20$[]{data-label="fig:ictv-better-20"}]({resimg/tgv2-huberonly-dataset\EXictvnoise20-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=20$[]{data-label="fig:ictv-better-20"}]({resimg/tgv2-huberonly-bsds300-ALL\EXictvnoise20-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=20$[]{data-label="fig:ictv-better-20"}]({img/bsds300/noisy20_\EXictv}.jpg "fig:"){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=20$[]{data-label="fig:ictv-better-20"}]({resimg/ictv-huberonly-dataset\EXictvnoise20-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\ICTV$ performs better than $\TGV^2$, $\sigma=20$[]{data-label="fig:ictv-better-20"}]({resimg/ictv-huberonly-bsds300-ALL\EXictvnoise20-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=2$[]{data-label="fig:tgv-better-2"}]({img/bsds300/orig_\EXtgv}.jpg "fig:"){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=2$[]{data-label="fig:tgv-better-2"}]({resimg/tgv2-huberonly-dataset\EXtgvnoise2-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=2$[]{data-label="fig:tgv-better-2"}]({resimg/tgv2-huberonly-bsds300-ALL\EXtgvnoise2-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=2$[]{data-label="fig:tgv-better-2"}]({img/bsds300/noisy2_\EXtgv}.jpg "fig:"){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=2$[]{data-label="fig:tgv-better-2"}]({resimg/ictv-huberonly-dataset\EXtgvnoise2-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=2$[]{data-label="fig:tgv-better-2"}]({resimg/ictv-huberonly-bsds300-ALL\EXtgvnoise2-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=20$[]{data-label="fig:tgv-better-20"}]({img/bsds300/orig_\EXtgv}.jpg "fig:"){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=20$[]{data-label="fig:tgv-better-20"}]({resimg/tgv2-huberonly-dataset\EXtgvnoise20-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=20$[]{data-label="fig:tgv-better-20"}]({resimg/tgv2-huberonly-bsds300-ALL\EXtgvnoise20-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=20$[]{data-label="fig:tgv-better-20"}]({img/bsds300/noisy20_\EXtgv}.jpg "fig:"){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=20$[]{data-label="fig:tgv-better-20"}]({resimg/ictv-huberonly-dataset\EXtgvnoise20-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} [0.32]{} ![Image for which $\TGV^2$ performs better than $\ICTV$, $\sigma=20$[]{data-label="fig:tgv-better-20"}]({resimg/ictv-huberonly-bsds300-ALL\EXtgvnoise20-linesearch1-bfgs1-beta0-uinitreg_0.200000_}.png){width="\textwidth"} The choice of cost functional ----------------------------- The $\costltwo$ cost functional naturally obtains better PSNR than $\costhubertv$, as the two former are equivalent. Comparing the results for the two cost funtionals in Tables \[table:bsds300-individual-2\]–\[table:bsds300-individual-20\], we may however observe that for low noise levels $\sigma^2=2,10$, and generally for batch learning, $\costhubertv$ attains better (higher) SSIM. Since SSIM better captures [@wang2004ssim] the visual quality of images than PSNR, this recommends the use of our novel total variation cost functional $\costhubertv$. Of course, one might attempt to optimise the SSIM. This is however a non-convex functional, which will pose additional numerical challenges avoided by the convex total variation cost. Conclusion and Outlook {#conclusion-and-outlook .unnumbered} ====================== In this paper we propose a bilevel optimisation method in function space for learning the optimal choice of parameters in higher-order total variation regularisation. We present a rigorous analysis of this optimisation problem as well as a numerical discussion in the context of image denoising. In particular, we make use of the bilevel learning approach to compare the performance – in terms of returned image quality – of TV, ICTV and TGV regularisation. A statistical analysis, carried out on a dataset of 200 images, suggest that ICTV performs slightly better than TGV, and both perform better than TV, in average. For denoising of images with a high noise level ICTV and TGV score comparably well. For images with large smooth areas TGV performs better than ICTV. Moreover, we propose a new cost functional for the bilevel learning problem, which exhibits interesting theoretical properties and has a better behaviour with respect to the PSNR related L$^2$ cost used previously in the literature. This study raises the question of other, alternative cost functionals. For instance, one could be tempted to used the SSIM as cost, but its non-convexity might present several analytical and numerical difficulties. The new cost functional, proposed in this paper, turns out to be a good compromise between image quality measure and analytically tractable cost term. Acknowledgements {#acknowledgements .unnumbered} ================ This project has been supported by King Abdullah University of Science and Technology (KAUST) Award No. KUK-I1-007-43, EPSRC grants Nr. EP/J009539/1 and Nr. EP/M00483X/1, the Escuela Politécnica Nacional de Quito under award PIS 12-14 and the MATHAmSud project SOCDE ‘Sparse Optimal Control of Differential Equations’. While in Quito, T. Valkonen has moreover been supported by a Prometeo scholarship of the Senescyt (Ecuadorian Ministry of Science, Technology, Education, and Innovation). [^1]: $^*$This research has been supported by King Abdullah University of Science and Technology (KAUST) Award No. KUK-I1-007-43, EPSRC grants Nr. EP/J009539/1 “Sparse & Higher-order Image Restoration” and Nr. EP/M00483X/1 “Efficient computational tools for inverse imaging problems”, Escuela Politécnica Nacional de Quito Award No. PIS 12-14 and MATHAmSud project SOCDE “Sparse Optimal Control of Differential Equations”. While in Quito, T. Valkonen has moreover been supported by SENESCYT (Ecuadorian Ministry of Higher Education, Science, Technology and Innovation) under a Prometeo Fellowship.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Flapan, [*et al*]{} [@FHLM] showed that every spatial embedding of $K_{10}$, the complete graph on ten vertices, contains a non-split three-component link ($K_{10}$ is [*intrinsically triple-linked*]{}). The papers [@BF] and [@FFNP] extended the list of known intrinsically triple-linked graphs in $\mathbb{R}^3$ to include several other families of graphs. In this paper, we will show that while some of these graphs can be embedded 3-linklessly in $\mathbb{R}P^3$, $K_{10}$ is intrinsically triple-linked in $\mathbb{R}P^3$.' author: - 'J. Federman [^1]' - 'J. Foisy [^2]' - 'K. McNamara [^3]' - 'E.Stark [^4]' title: 'Intrinsically triple-linked graphs in $\mathbb{R}P^3$ [^5]' --- Introduction ============ [*Real projective 3-space*]{}, $\mathbb{R}P^3$, is defined to be the quotient $S^3 / \sim$, where $\sim$ is the antipodal relation $x \sim -x$ and can be thought of as the disk, $D^3$, with antipodal boundary points identified. Projective space has a non-trivial first homology group, $H_1 \cong \mathbb{Z}/2\mathbb{Z}$. The generator for the group, $g$, is the cycle originating from the line in $D^3$ that runs between the north and south poles. Mroczkowski [@M] has shown that every knot in $\mathbb{R}P^3$ can be transformed into either the trivial cycle or $g$ by crossing changes and generalized Reidemeister moves on an $\mathbb{R}P^2$ projection of the knot. Thus, there are two non-equivalent unknots in $\mathbb{R}P^3$. Cycles that can be “unknotted" into a cycle homologous to $g$ will be referred to as [*1-homologous cycles*]{}. Cycles that can be “unknotted" into a trivial cycle will be referred to as [*0-homologous cycles*]{}. A [*link*]{} in $\mathbb{R}P^3$ is [*splittable*]{} if one of the components can be contained within a sphere, embedded in the space, while the other component remains in the complement of the sphere. Otherwise, the link in $\mathbb{R}P^3$ is [*non-split*]{}. A non-split link can be formed one of three ways in $\mathbb{R}P^3$: two 0-homologous cycles, a 0-homologous cycle with a 1-homologous cycle, and two 1-homologous cycles. Note: Two disjoint 1-homologous cycles will always form a non-split link. Similarly, a [*non-split triple-link*]{} is a non-split link of three components. In this paper we will refer to non-split linked cycles as [*linked cycles*]{} and an embedding of a graph as [*linked*]{} if it contains a non-split link. We will refer to a non-split triple-link as a [*triple-link*]{} and an embedding of a graph as [*triple-linked*]{} if it contains a non-split triple-link. A graph $H$ is a [*minor*]{} of a graph $G$ if $H$ can be obtained from $G$ through a series of vertex removals, edge removals, or edge contractions. A graph $G$ is [*minor-minimal*]{} with respect to a property $P$ if $G$ has property $P$ but no minor of $G$ has property $P$. If $G$ is a graph, define an [*induced subgraph*]{}, $G[v_1, v_2, ... , v_n]$, of $G$ to be the subgraph of $G$ on vertices $\{v_1,v_2, ... , v_n\}$ and the set of edges in $G$ with both endpoints in the set $\{v_1, v_2, ... , v_n\}$. A graph $G$ is [*intrinsically linked in*]{} ${\mathbb R}^3$ if and only if $G$ contains a non-split link in every spatial embedding. We define [*intrinsically linked in*]{} ${\mathbb R}P^3$ analogously. It has been shown that the complete set of minor-minimal intrinsically linked graphs in ${\mathbb R}^3$ is the set of Petersen Family graphs [@RST] (including $K_6$ and graphs obtained from $K_6$ by $\Delta - Y$ and $Y - \Delta$ exchanges). However, all Petersen Family graphs except for $K_{4,4}-e$ embed linklessly in ${\mathbb R}P^3$ [@REU07]. While [@REU07] characterizes several families of graphs that are minor-minimally intrinsically linked in ${\mathbb R}P^3$, the complete set of minor-minally intrinsically linked graphs in ${\mathbb R}P^3$, which is finite due to the result in [@RS], remains to be found. A graph $G$ is [*intrinsically triple-linked in*]{} $\mathbb{R}^3$ if and only if $G$ contains a non-split link of three components in every spatial embedding. We define [*intrinsically triple-linked in*]{} $\mathbb{R}P^3$ analogously. An embedding is said to be 3-[*linkless*]{} if and only if it does not contain a triple-link. While Conway, Gordon [@CG], and Sachs [@S:83; @S:84] showed that $K_6$ is intrinsically linked in $\mathbb{R}^3$, $K_6$ can be linklessly embedded in $\mathbb{R}P^3$; it has been shown that 7 is the smallest $n$ for which $K_n$ is intrinsically linked in $\mathbb{R}P^3$ [@REU07]. In contrast, while 10 was shown to be the smallest $n$ for which $K_n$ is intrinsically triple-linked in $\mathbb{R}^3$ [@FNP], we have shown that 10 is also the smallest $n$ for which $K_n$ is intrinsically triple-linked in $\mathbb{R}P^3$. It remains to be shown whether $K_{10}$ is minor-minimal with respect to triple-linking in $\mathbb{R}P^3$. Additionally, we have shown two other intrinsically triple-linked graphs in $\mathbb{R}^3$ can be embedded without a triple-link in $\mathbb{R}P^3$. A complete set of minor-minimal intrinsically triple-linked graphs remains to be found, in both $\mathbb{R}^3$ and $\mathbb{R}P^3$. Such sets are finite due to the result in [@RS]. Intrinsically triple-linked complete graphs on $n$ vertices =========================================================== We will need the following lemmas: \[K7\] [@REU07] [*The graphs obtained by removing two edges from $K_7$ and removing one edge from $K_{4,4}$ are intrinsically linked in $\mathbb{R}P^3$.*]{} \[OHK4\] [@REU07] [*Given a linkless embedding of $K_6$ in $\mathbb{R}P^3$, no $K_4$ subgraph can have all 0-homologous cycles.*]{} We also use the following elementary observation. \[evenK4\] [*For every embedding into $\mathbb{R}P^3$, $K_4$ has an even number of 1-homologous cycles.*]{} The following lemma was shown true in $\mathbb{R}^3$ by Flapan, Naimi, and Pommersheim [@FNP] and the proof holds true analogously in $\mathbb{R}P^3$. \[LL\] [*Let $G$ be a graph embedded in $\mathbb{R}P^3$ that contains cycles $C_1$, $C_2$, $C_3$ and $C_4$. Suppose $C_1$ and $C_4$ are disjoint from each other and from $C_2$ and $C_3$ and suppose $C_2 \cap C_3$ is a simple path. If $lk(C_1, C_2) \not = 0$ and $lk(C_3, C_4) \not = 0$, then $G$ contains a non-split three-component link.*]{} The following proposition is not the main result of this paper. However, the proof is included because it is concise and since its method does not hold for proving $K_{10}$ is also triple-linked. [*The graph $K_{11}$ is intrinsically triple-linked in $\mathbb{R}P^3$.*]{} Let $G$ be a complete graph on the vertex set $\{1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11\}$. Embed $G$ in $\mathbb{R}P^3$. Consider $G[1, 2, 3, 4, 5, 6, 7]$. Since $K_7$ is intrinsically linked in $\mathbb{R}P^3$, this subgraph contains a pair of linked cycles that can be reduced to two linked $3$-cycles. Without loss of generality, let $C_1 = (1, 2, 3)$ and $C_2 = (4, 5, 6)$ be the pair of linked cycles in $G[1, 2, 3, 4, 5, 6, 7]$. Now consider $G[ 5, 6, 7, 8, 9, 10, 11]$. Since $K_7$ is intrinsically linked in $\mathbb{R}P^3$, this subgraph contains a pair of linked cycles that can be reduced to two linked $3$-cycles. In $G[ 5, 6, 7, 8, 9, 10, 11]$, one cycle must use $\{v_5\}$ and the other cycle must use $\{v_6\}$, or Lemma \[LL\] would apply immediately. Without loss of generality, let $C_3 = (5, 7, 9)$ and $C_4 = (6, 8, 10)$ be the pair of linked cycles in $G[ 5, 6, 7, 8, 9, 10, 11]$. Consider $G[1, 2, 3, 4, 6, 11]$. By Lemma \[OHK4\], $G[1, 2, 3, 11]$ must contain a $1$-homologous cycle or $G[1, 2, 3, 4, 6, 11]$ contains a pair of linked cycles and Lemma \[LL\] applies with $C_3$ and $C_4$. Thus by Lemma \[evenK4\], two cycles in $A = \{ (1, 2, 3), (1, 2, 11), (1, 3, 11), (2, 3, 11)\}$ must be $1$-homologous $3$-cycles. Now consider $G[6, 7, 8, 9, 10, 11]$. By Lemma \[OHK4\], $G[7, 8, 9, 10]$ must contain a $1$-homologous cycle or $G[6, 7, 8, 9, 10, 11]$ contains a pair of linked cycles and Lemma \[LL\] applies with $C_1$ and $C_2$. Thus by Lemma \[evenK4\], two cycles in $B = \{ (7, 8, 9), (7, 8, 10), (7, 9, 10), (8, 9, 10)\}$ must be $1$-homologous $3$-cycles. Since every cycle in $A$ is disjoint from every cycle in $B$, and at least two cycles in each set are $1$-homologous, there exists a link using one cycle from $A$ and one cycle from $B$. Lemma \[LL\] then applies since every cycle in $A$ shares at least a simple path with $C_1$, and $C_2$ and the cycle from $B$ are disjoint from each other, $C_1$, and the cycle from $A$. Thus, $G$ contains a triple-link. -1.5in ![A projection of a linkless embedding of $K_6$ in $\mathbb{R}P^3$.[]{data-label="linkless"}](LinklessK6.eps "fig:") -1.5in [*If $G$ is $K_6$ embedded in $\mathbb{R}P^3$ and $G$ has two disjoint $0$-homologous cycles, then $G$ contains a non-split link.*]{} Assume $G$ can be embedded so that it has two disjoint 0-homologous cycles and so that it does not have a non-split link. Without loss of generality, let $(1, 2, 3)$ and $(4, 5, 6)$ be $0$-homologous cycles in $G$. Consider $G[ 1, 2, 3, 4]$. Since $G$ is not linked, by Lemma \[OHK4\] and Lemma \[evenK4\], $G[ 1, 2, 3, 4]$ must have two $1$-homologous cycles. Without loss of generality, let $(1, 2, 4)$ and $(1, 3, 4)$ be $1$-homologous cycles. Similarly, $G[ 2, 4, 5, 6]$ must also have two $1$-homologous cycles. Since $(4, 5, 6)$ is $0$-homologous by assumption and $(2, 5, 6)$ is disjoint from $(1, 3, 4)$, $(2, 4, 5)$ and $(2, 4, 6)$ are $1$-homologous cycles. Similarly, $G[ 1, 2, 3, 6]$ has two $1$-homologous cycles. Since $(1, 2, 3)$ is $0$-homologous by assumption and $(1, 3, 6)$ is disjoint from $(2, 4, 5)$, $(1, 2, 6)$ and $(2, 3, 6)$ are $1$-homologous cycles or $G$ would contain a pair of linked cycles. Now consider $G[ 1, 3, 5, 6]$, which must also have two $1$-homologous cycles by Lemma 2 and Lemma 3. Since $(1, 3, 5)$ is disjoint from $(2, 4, 6)$, $(1, 3, 6)$ is disjoint from $(2, 4, 5)$, and $(3, 5, 6)$ is disjoint from $(1, 2, 4)$, $(1, 3, 5)$, $(1, 3, 6)$, and $(3, 5, 6)$ must be $0$-homologous. This forces $G[1,3,5,6]$ to contain only 0-homologous cycles, and thus $G$ is linked by \[OHK4\]. Thus, $G$ cannot have two disjoint $0$-homologous cycles and not be linked. [*Up to ambient isotopy and crossing changes, Figure \[linkless\] is the only way to linklessly embed $K_6$ in $\mathbb{R}P^3$.*]{} Let $G$ be a complete graph on the vertex set {1, 2, 3, 4, 5, 6}. Embed $G$ in $\mathbb{R}P^3$. The graph $G$ has a $0$-homologous 3-cycle, else $G$ has disjoint $1$-homologous cycles and is thus linked by Proposition 6. Without loss of generality, let $(4, 5, 6)$ be a $0$-homologous 3-cycle. Now consider vertices {1, 2, 3}. If $(1, 2, 3)$ is $0$-homologous, $G$ is linked; thus, we assume $(1, 2, 3)$ is a $1$-homologous cycle. Mroczkowski [@M] showed that any cycle can be made into an unknotted $0$- or $1$-homologous cycle by crossing changes, so we can assume after crossing changes and ambient isotopy the embedding has a projection as drawn in Figure \[linkless\] (except the edges between the vertices $\{1, 2, 3\}$ and $\{4, 5, 6\}$ may be more complicated than in the Figure) with vertices $\{1, 2, 3\}$ on the boundary and the edges between them on the boundary. We may use ambient isotopy and crossing changes so that edges from $\{1, 2, 3\}$ to $\{4, 5, 6\}$ connect in the projection without crossing the boundary of $D^2$. We now show that we may connect them, without loss of generality, as depicted in Figure 1. If vertex $v \in \{1, 2, 3\}$, $v$ must connect to at least one of {4, 5, 6} from $v_A$ and to at least one of {4, 5, 6} from $v_B$, else there would be a $0$-homologous $K_4$ and $G$ would be linked by Lemma 2. Without loss of generality, assume $v_2$ connects to $v_4$ and $v_6$ from $v_{2_B}$ and to $v_5$ from $v_{2_A}$. If $v_1$ connects to $v_4$ and $v_6$ from $v_{1_B}$, then $G[1, 2, 4, 6]$ is a $0$-homologous $K_4$ and $G$ is linked by Lemma 2. Thus, $v_1$ connects to either $v_4$ or $v_6$ from $v_{1_B}$ and connects to the other from $v_{1_A}$. Without loss of generality, let $v_{1_B}$ connect to $v_4$; so, $v_{1_A}$ connects to $v_6$. If $v_{1_B}$ connects to $v_5$, then either $v_{3_A}$ or $v_{3_B}$ must connect to both $v_5$ and $v_6$ and the other to $v_4$, else $G$ has a $0$-homologous $K_4$. Without loss of generality, let $v_{3_A}$ connect to $v_5$ and $v_6$ and $v_{3_B}$ connect to $v_4$. Then, $(1, 2, 5)$ and $(3, 4, 6)$ are disjoint $1$-homologous cycles so $G$ is linked. Thus, $v_{1_A}$ connects to $v_5$. Now, if $v_{3_A}$ connects to either $v_4$ and $v_6$ or $v_5$ and $v_6$, then $G$ has a $0$-homologous $K_4$ and is linked by Lemma 2. So, $v_{3_A}$ must connect to $v_6$ and $v_{3_B}$ must connect to $v_4$ and $v_5$. -1.5in ![A signed linkless embedding of $K_6$ in $\mathbb{R}P^3$.[]{data-label="signed"}](K6signed.eps "fig:") -1.5in Signed graphs, that is, graphs with each edge assigned a $+$ or a $-$ sign, have been studied extensively and were first introduced by Harary [@H], see also [@ZA]. An embedding of a graph $G$ into $\mathbb{R}P^3$ induces a signed graph of $G$ as follows: deform the embedding so that no vertices touch the line at infinity and all intersections of edges with the line at infinity are transverse. Assign $+$ edges to be edges that hit the boundary an even number of times and $-$ edges to be edges that hit the boundary an odd number of times. If a cycle has an odd number of $-$ edges, then the cycle is $1$-homologous. Two embeddings, $G_1$ and $G_2$, of a graph $G$ are [*crossing-change equivalent*]{} if and only if $G_1$ can be obtained from $G_2$ by crossing changes and ambient isotopy. Thus, by Proposition 7, a linkless $K_6$ is crossing-change equivalent to the embedding in Figure \[signed\]. That is, $(1, 2)$, $(1, 3)$, $(2, 3)$, $(1, 4)$, $(2, 5)$, and $(3, 6)$ are $-$ edges, and the other nine edges are $+$ edges. [*The graph $K_{10}$ is intrinsically triple-linked in $\mathbb{R}P^3$.*]{} Let $G$ be a graph isomorphic to $K_{10}$ on the vertex set $\{1, 2, 3, 4, 5, 6, 7, 8, 9, 10\}$. Embed $G$ in $\mathbb{R}P^3$. Assume the embedding is 3-linkless. If every subgraph of $G$ isomorphic to $K_6$ is linked, then Flapan, Naimi, and Pommersheim’s proof [@FNP] that $K_{10}$ is intrinsically linked in $\mathbb{R}^3$ nearly works, except they do use the fact that $K_{3,3,1}$ is intrinsically linked at the very end and $K_{3,3,1}$ is not intrinsically linked in $\mathbb{R}P^3$. Bowlin-Foisy [@BF], however, modify [@FNP] slightly so that only the fact that $K_6$ is intrinsically linked is needed. Thus, in the case that every subgraph of $G$ isomorphic to $K_6$ is linked, then $G$ is triple-linked. So we may assume there exists a linkless $K_6$ subgraph in $G$. Without loss of generality, assume this linkless $K_6$ is on vertices $\{1, 2, 3, 4, 5, 6\}$. By Proposition 7, this $K_6$ has an embedding crossing-change equivalent to that in drawn in Figure \[signed\]. [**Claim:**]{} The embedded induced subgraph $G\left[7, 8, 9, 10\right]$ is $0$-homologous. Assume $G[7, 8, 9, 10]$ has a $1$-homologous cycle. Without loss of generality, let $(7, 8, 9)$ be a $1$-homologous cycle. Now consider $G[4, 5, 6, 10]$. If $G[4, 5, 6, 10]$ is not $0$-homologous, then two of $(4, 5, 10)$, $(4, 6, 10)$, and $(5, 6, 10)$ are $1$-homologous by Lemma \[evenK4\]. Then $(1, 2, 3)$, $(7, 8, 9)$, and a cycle from $G[4, 5, 6, 10]$ comprise three disjoint $1$-homologous cycles, so $G$ is triple-linked. Thus, $G[4, 5, 6, 10]$ is $0$-homologous and so $G[1, 2, 4, 5, 6, 10]$ has a pair of linked cycles by Lemma \[OHK4\]. Since $(7, 8, 9)$ is $1$-homologous, and $(7, 8, 9)$ is disjoint from all the $1$-homologous cycles in the second column of Table 1, Lemma \[LL\] applies and $G$ has a triple-link. Thus, $G[7, 8, 9, 10]$ is $0$-homologous. ---------------------------------- ------------------------------------ Possible Linked $1$-Homologous Cycle that Cycles in $G[1, 2, 4, 5, 6, 10]$ shares an edge with a linked cycle $(1, 2, 4)$, $(5, 6, 10)$ $(1, 2, 3)$ $(1, 2, 5)$, $(4, 6, 10)$ $(1, 2, 3)$ $(1, 2, 6)$, $(4, 5, 10)$ $(1, 2, 3)$ $(1, 2, 10)$, $(4, 5, 6)$ $(1, 2, 3)$ $(1, 4, 5)$, $(2, 6, 10)$ $(1, 3, 5)$ $(1, 4, 6)$, $(2, 5, 10)$ $(1, 4, 6)$ $(1, 4, 10)$, $(2, 5, 6)$ $(2, 5, 6)$ $(1, 5, 6)$, $(2, 4, 10)$ $(1, 3, 5)$ $(1, 5, 10)$, $(2, 4, 6)$ $(1, 3, 5)$ $(1, 6, 10)$, $(2, 4, 5)$ $(2, 4, 5)$ ---------------------------------- ------------------------------------ Table $1$. Since $G[7, 8, 9, 10]$ is $0$-homologous, we may assume all edges in $G[7, 8, 9, 10]$ are $+$ edges. The edges in $G[1, 2, 3, 4, 5, 6]$ are $+$ and $-$ edges as defined in Figure \[signed\]. The following arguments will use this modified embedding of $G$, however, since ambient isotopy and crossing changes do not change the homology of the cycles, the linking arguments will still hold for the original embedding. Similar to the argument highlighted in Table 1, many of the following arguments rely on $K_6$ subgraphs of $G$ that must have a pair of linked cycles. The modified embedding may have a different pair of linked cycles in the subgraph than in the original embedding, however a pair of linked cycles still exists and the argument does not rely on which cycles are linked. We now consider the signs of the edges connecting $G[1, 2, 3, 4, 5, 6]$ to $G[7, 8, 9, 10]$. [**Claim:**]{} If $v \in \{1, 2, 3\}$, then edges from $v$ to $G[7, 8, 9, 10]$ must all be $+$ edges or all $-$ edges. Assume vertex $v_1$ does not connect by all $+$ edges or all $-$ edges to $G[7, 8, 9, 10]$. Without loss of generality, let $(1, 7)$ be a $+$ edge and $(1,8)$ be a $-$ edge. Then, $(1, 7, 8)$ is a $1$-homologous cycle. Consider $G[3, 4, 6, 9]$. Since $(3, 4, 6)$ is a $1$-homologous cycle, $G[3, 4, 6, 9]$ must have another $1$-homologous cycle by Lemma \[evenK4\]. If $(3, 4, 9)$ is $1$-homologous, then $(1, 7, 8)$, $(2, 5, 6)$, and $(3, 4, 9)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. If $(3, 6, 9)$ is $1$-homologous, then $(1, 7, 8)$, $(2, 4, 5)$, and $(3, 6, 9)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. Thus, $(4, 6, 9)$ is a $1$-homologous cycle. Now consider $G[2, 3, 4, 9]$. Since $(2, 3, 4)$ is a $1$-homologous cycle, $G[2, 3, 4, 9]$ must have another $1$-homologous cycle by Lemma \[evenK4\]. If $(3, 4, 9)$ is $1$-homologous, then $(1, 7, 8)$, $(2, 5, 6)$, and $(3, 4, 9)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. If $(2, 4, 9)$ is $1$-homologous, then $(1, 7, 8)$, $(2, 4, 9)$, and $(3, 5, 6)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. Thus, $(2, 3, 9)$ is a $1$-homologous cycle. Similarly, consider $G[3, 5, 6, 9] $. Since $(3, 5, 6)$ is a $1$-homologous cycle, $G[3, 5, 6, 9]$ must have another $1$-homologous cycle by Lemma \[evenK4\]. If $(3, 6, 9)$ is $1$-homologous, then $(1, 7, 8)$, $(2, 4, 5)$, and $(3, 6, 9)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. If $(5, 6, 9)$ is $1$-homologous, then $(1, 7, 8)$, $(2, 3, 4)$, and $(5, 6, 9)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. Thus, $(3, 5, 9)$ is a $1$-homologous cycle. Since $(1, 7, 8)$ and $(4, 6, 9)$ are $1$-homologous, $G[2, 3, 5, 10]$ is $0$-homologous or else there are three disjoint $1$-homologous cycles. Thus, $G[2, 3, 4, 5, 6, 10]$ has a pair of linked cycles by Lemma \[OHK4\]. Since $(1, 7, 8)$ is $1$-homologous, and $(1, 7, 8)$ is disjoint from all the $1$-homologous cycles in the second column of Table 2, Lemma \[LL\] applies and $G$ has a pair of links. Thus, vertex $v_1$ must have all $+$ edges or all $-$ edges to $G[7, 8, 9, 10]$. Similar reasoning applies to vertices $v_2$ and $v_3$. ---------------------------------- ------------------------------------ Possible Linked $1$-Homologous Cycle that Cycles in $G[2, 3, 4, 5, 6, 10]$ shares an edge with a linked cycle $(2, 3, 4)$, $(5, 6, 10)$ $(2, 3, 4)$ $(2, 3, 5)$, $(4, 6, 10)$ $(4, 6, 9)$ $(2, 3, 6)$, $(4, 5, 10)$ $(2, 3, 9)$ $(2, 3, 10)$, $(4, 5, 6)$ $(2, 3, 9)$ $(2, 4, 5)$, $(3, 6, 10)$ $(2, 4, 5)$ $(2, 4, 6)$, $(3, 5, 10)$ $(4, 6, 9)$ $(2, 4, 10)$, $(3, 5, 6)$ $(3, 5, 9)$ $(2, 5, 6)$, $(3, 4, 10)$ $(2, 5, 6)$ $(2, 5, 10)$, $(3, 4, 6)$ $(4, 6, 9)$ $(2, 6, 10)$, $(3, 4, 5)$ $(3, 5, 9)$ ---------------------------------- ------------------------------------ Table $2$. [**Claim:**]{} If $v \in \{4, 5, 6\}$, then edges from $v$ to $G[7, 8, 9, 10]$ must all be $+$ edges or all $-$ edges. Assume vertex $v_4$ does not have all $+$ edges or all $-$ edges to $G[7, 8, 9, 10]$. Without loss of generality, let $(4, 7)$ be a $+$ edge and $(4,8)$ be a $-$ edge. Then, $(4, 7, 8)$ is a $1$-homologous cycle. Consider $G[1, 2, 3, 9]$. Since $(1, 2, 3)$ is a $1$-homologous cycle, $G[1, 2, 3, 9]$ must have another $1$-homologous cycle by Lemma \[evenK4\]. If $(1, 3, 9)$ is $1$-homologous, then $(1, 3, 9)$, $(2, 5, 6)$, and $(4, 7, 8)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. If $(1, 2, 9)$ is $1$-homologous, then $(1, 2 ,9)$, $(3, 5, 6)$, and $(4, 7, 8)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. Thus, $(2, 3, 9)$ is a $1$-homologous cycle. Now consider $G[2, 5, 6, 9]$. Since $(2, 5, 6)$ is a $1$-homologous cycle, $G[2, 5, 6, 9]$ must have another $1$-homologous cycle by Lemma \[evenK4\]. If $(2, 6, 9)$ is $1$-homologous, then $(1, 3, 5)$, $(2, 6, 9)$, and $(4, 7, 8)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. If $(5, 6, 9)$ is $1$-homologous, then $(1, 2, 3)$, $(4, 7, 8)$, and $(5, 6, 9)$ form three disjoint $1$-homologous cycles, so $G$ is triple-linked. Thus, $(5, 6, 9)$ is a $1$-homologous cycle. Since $(2, 3, 9)$ and $(4, 7, 8)$ are $1$-homologous, $G[1, 5, 6, 10]$ is $0$-homologous or else there are three disjoint $1$-homologous cycles. Thus, by Lemma \[OHK4\], $G[1, 2, 3, 5, 6, 10]$ has a pair of linked cycles. Since $(4, 7, 8)$ is $1$-homologous, and $(4, 7, 8)$ is disjoint from all the $1$-homologous cycles in the second column of Table 3, Lemma \[LL\] applies and $G$ has a pair of links. Thus, vertex $v_4$ must have all $+$ edges or all $-$ edges to $G[7, 8, 9, 10]$. Similarly, vertices $v_5$ and $v_6$ must have all $+$ edges or all $-$ edges to $G[7, 8, 9, 10]$. ---------------------------------- ------------------------------------ Possible Linked $1$-Homologous Cycle that Cycles in $G[1, 2, 3, 5, 6, 10]$ shares an edge with a linked cycle $(1, 2, 3)$, $(5, 6, 10)$ $(1, 2, 3)$ $(1, 2, 5)$, $(3, 6, 10)$ $(2, 5, 9)$ $(1, 2, 6)$, $(3, 5, 10)$ $(1, 2, 6)$ $(1, 2, 10)$, $(3, 5, 6)$ $(3, 5, 6)$ $(1, 3, 5)$, $(2, 6, 10)$ $(1, 3, 5)$ $(1, 3, 6)$, $(2, 5, 10)$ $(2, 5, 9)$ $(1, 3, 10)$, $(2, 5, 6)$ $(2, 5, 9)$ $(1, 5, 6)$, $(2, 3, 10)$ $(2, 3, 9)$ $(1, 5, 10)$, $(2, 3, 6)$ $(2, 3, 9)$ $(1, 6, 10)$, $(2, 3, 5)$ $(2, 3, 9)$ ---------------------------------- ------------------------------------ Table $3$. Since each vertex in $G[1, 2, 3, 4, 5, 6]$ has either all $+$ edges or all $-$ edges to $G[7, 8, 9, 10]$, there are $2^6$ possible embedding classes, given our restrictions on how $G[1, 2, 3, 4, 5, 6]$ and $G[7, 8, 9, 10]$ are embedded. We consider all the cases. Note: If vertex $v_1$ connects to $G[7, 8, 9, 10]$ with all $+$ edges, we write $v_{1+}$, else we write $v_{1-}$. Consider the embedding of $G$ with $v_{1+}$ and $v_{4+}$. If we have one of the following embeddings: $v_{2+}$, $v_{3+}$, $v_{5+}$, and $v_{6+}$; $v_{2+}$, $v_{3-}$, $v_{5+}$, and $v_{6-}$; $v_{2-}$, $v_{3+}$, $v_{5-}$, and $v_{6+}$; $v_{2-}$, $v_{3-}$, $v_{5-}$, and $v_{6-}$, then $(1, 4, 7)$, $(2, 5, 8)$, and $(3, 6, 9)$ form three disjoint $1$-homologous cycles, so $G$ has a triple-link. If we have one of the following embeddings: $v_{2+}$, $v_{3+}$, $v_{5+}$, and $v_{6-}$; $v_{2+}$, $v_{3+}$, $v_{5-}$, and $v_{6+}$; $v_{2-}$, $v_{3-}$, $v_{5-}$, and $v_{6+}$; $v_{2-}$, $v_{3-}$, $v_{5+}$, and $v_{6-}$, then $(1, 4, 7)$, $(2, 3, 8)$, and $(5, 6, 9)$ form three disjoint $1$-homologous cycles, so $G$ has a triple-link. If we have one of the following embeddings: $v_{2+}$, $v_{3+}$, $v_{5-}$, and $v_{6-}$; $v_{2-}$, $v_{3-}$, $v_{5+}$, and $v_{6+}$, then $(1, 4, 7)$, $(2, 6, 8)$, and $(3, 5, 9)$ form three disjoint $1$-homologous cycles, so $G$ has a triple-link. If the embedding is $v_{2-}$, $v_{3+}$, $v_{5+}$, $v_{6-}$, since $(1, 4, 7)$ and $(5, 6, 8)$ are $1$-homologous cycles, $G[2, 3, 9, 10]$ must be $0$-homologous or $G$ is triple-linked, so, by Lemma \[OHK4\], $G[1, 2, 3, 4, 9, 10]$ has a pair of links in this embedding class. Since $(5, 6, 8)$ is $1$-homologous, and $(5, 6, 8)$ is disjoint from all the $1$-homologous cycles in the second column of Table 4, Lemma \[LL\] applies and $G$ has a pair of links. If the embedding is $v_{2+}$, $v_{3-}$, $v_{5-}$, $v_{6+}$, $G$ is linked by a similar argument. ---------------------------------- ------------------------------------ Possible Linked $1$-Homologous Cycle that Cycles in $G[1, 2, 3, 4, 9, 10]$ shares an edge with a linked cycle $(1, 2, 3)$, $(4, 9, 10)$ $(1, 2, 3)$ $(1, 2, 4)$, $(3, 9, 10)$ $(1, 4, 7)$ $(1, 2, 9)$, $(3, 4, 10)$ $(1, 2, 7)$ $(1, 2, 10)$, $(3, 4, 9)$ $(1, 2, 7)$ $(1, 3, 4)$, $(2, 9, 10)$ $(1, 4, 7)$ $(1, 3, 9)$, $(2, 4, 10)$ $(1, 3, 7)$ $(1, 3, 10)$, $(2, 4, 9)$ $(1, 3, 7)$ $(1, 4, 9)$, $(2, 3, 10)$ $(1, 4, 7)$ $(1, 4, 10)$, $(2, 3, 9)$ $(1, 4, 7)$ $(1, 9, 10)$, $(2, 3, 4)$ $(2, 3, 4)$ ---------------------------------- ------------------------------------ Table $4$. If the embedding is $v_{2+}$, $v_{3-}$, $v_{5+}$, and $v_{6+}$, since $G[7, 8, 9, 10]$ is $0$-homologous, by Lemma \[OHK4\], $G[1, 4, 7, 8, 9, 10]$ has a pair of links. Since $(3, 5, 6)$ is $1$-homologous, and $(3, 5, 6)$ is disjoint from all the $1$-homologous cycles in the second column of Table 5, Lemma \[LL\] applies and $G$ has a pair of links. If we have one of the following embeddings: $v_{2-}$, $v_{3+}$, $v_{5-}$, and $v_{6-}$; $v_{2+}$, $v_{3-}$, $v_{5-}$, and $v_{6-}$; $v_{2-}$, $v_{3+}$, $v_{5+}$, and $v_{6+}$, then $G$ is linked by a similar argument. ---------------------------------- ------------------------------------ Possible Linked $1$-Homologous Cycle that Cycles in $G[1, 4, 7, 8, 9, 10]$ shares an edge with a linked cycle $(1, 4, 7)$, $(8, 9, 10)$ $(1, 4, 7)$ $(1, 4, 8)$, $(7, 9, 10)$ $(1, 4, 8)$ $(1, 4, 9)$, $(7, 8, 10)$ $(1, 4, 9)$ $(1, 4, 10)$, $(7, 8, 9)$ $(1, 4, 10)$ $(1, 7, 8)$, $(4, 9, 10)$ $(1, 2, 7)$ $(1, 7, 9)$, $(4, 8, 10)$ $(1, 2, 7)$ $(1, 7, 10)$, $(4, 8, 9)$ $(1, 2, 7)$ $(1, 8, 9)$, $(4, 7, 10)$ $(1, 2, 8)$ $(1, 8, 10)$, $(4, 7, 9)$ $(1, 2, 8)$ $(1, 9, 10)$, $(4, 7, 8)$ $(1, 2, 9)$ ---------------------------------- ------------------------------------ Table $5$. This list exhausts the possible embeddings if both we have both $v_{1+}$ and $v_{4+}$. The same arguement holds if the embedding is $v_{1-}$ and $v_{4-}$. Thus, we can now assume the edges from $v_1$ and $v_4$ to $G[7, 8, 9, 10]$ have different signs. Consider the embedding of $G$ with $v_{1+}$ and $v_{4-}$. We can assume that the pairs $\{3, 6\}$, and $\{2, 5\}$ have different signs or the same arguments for $v_1$ and $v_4$ with the same sign holds from above. If the embedding is $v_{2+}$, $v_{3+}$, $v_{5-}$, and $v_{6-}$, then $(1, 6, 7)$, $(3, 5, 8)$, and $(2, 4, 9)$ form three $1$-homologous cycles, so $G$ has a triple-link. If the embedding is $v_{2+}$, $v_{3-}$, $v_{5-}$, and $v_{6+}$, since $G[7, 8, 9, 10]$ is $0$-homologous, by Lemma \[OHK4\], $G[4, 6, 7, 8, 9, 10]$ has a pair of links. Since $(1, 2, 3)$ is $1$-homologous, and $(1, 2, 3)$ is disjoint from all the $1$-homologous cycles in the second column of Table 6, Lemma \[LL\] applies and $G$ has a pair of links. A similar argument holds if the embedding is $v_{2-}$, $v_{3+}$, $v_{5+}$, and $v_{6-}$. ---------------------------------- ------------------------------------ Possible Linked $1$-Homologous Cycle that Cycles in $G[4, 6, 7, 8, 9, 10]$ shares an edge with a linked cycle $(4, 6, 7)$, $(8, 9, 10)$ $(4, 6, 7)$ $(4, 6, 8)$, $(7, 9, 10)$ $(4, 6, 8)$ $(4, 6, 9)$, $(7, 8, 10)$ $(4, 6, 9)$ $(4, 6, 10)$, $(7, 8, 9)$ $(4, 6, 10)$ $(4, 7, 8)$, $(6, 9, 10)$ $(5, 6, 9)$ $(4, 7, 9)$, $(6, 8, 10)$ $(5, 6, 8)$ $(4, 7, 10)$, $(6, 8, 9)$ $(5, 6, 8)$ $(4, 8, 9)$, $(6, 7, 10)$ $(5, 6, 7)$ $(4, 8, 10)$, $(6, 7, 9)$ $(5, 6, 7)$ $(4, 9, 10)$, $(6, 7, 8)$ $(5, 6, 7)$ ---------------------------------- ------------------------------------ Table $6$. If the embedding is $v_{2-}$, $v_{3-}$, $v_{5+}$, and $v_{6+}$, since $G[7, 8, 9, 10]$ is $0$-homologous, by Lemma \[OHK4\], $G[4, 6, 7, 8, 9, 10]$ has a pair of links. Since $(1, 2, 3)$ is $1$-homologous, and $(1, 2, 3)$ is disjoint from all the $1$-homologous cycles in the second column of Table 7, Lemma \[LL\] applies and $G$ has a pair of links. ---------------------------------- ------------------------------------ Possible Linked $1$-Homologous Cycle that Cycles in $G[4, 6, 7, 8, 9, 10]$ shares an edge with a linked cycle $(4, 6, 7)$, $(8, 9, 10)$ $(4, 6, 7)$ $(4, 6, 8)$, $(7, 9, 10)$ $(4, 6, 8)$ $(4, 6, 9)$, $(7, 8, 10)$ $(4, 6, 9)$ $(4, 6, 10)$, $(7, 8, 9)$ $(4, 6, 10)$ $(4, 7, 8)$, $(6, 9, 10)$ $(4, 5, 7)$ $(4, 7, 9)$, $(6, 8, 10)$ $(4, 5, 7)$ $(4, 7, 10)$, $(6, 8, 9)$ $(4, 5, 7)$ $(4, 8, 9)$, $(6, 7, 10)$ $(4, 5, 8)$ $(4, 8, 10)$, $(6, 7, 9)$ $(4, 5, 8)$ $(4, 9, 10)$, $(6, 7, 8)$ $(4, 5, 9)$ ---------------------------------- ------------------------------------ Table $7$. This list exhausts the possible embeddings with $v_{1+}$ and $v_{4-}$. The same argument holds for the embedding has $v_{1-}$ and $v_{4+}$. Thus, in every embedding of $G$ in $\mathbb{R}P^3$, $G$ has a triple-link. Flapan, Naimi, and Pommersheim [@FNP] showed that $K_9$ can be embedded 3-linklessly in $\mathbb{R}^3$, and so $K_9$ can be embedded 3-linklessly in $\mathbb{R}P^3$. Thus, 10 is the smallest $n$ for which $K_n$ is intrinsically triple-linked in $\mathbb{R}P^3$. Other intrinsically triple-linked graphs in $\mathbb{R}P^3$ =========================================================== [*A graph composed of $n$ disjoint copies of an intrinsically $n$-linked graph in $\mathbb{R}^3$ is intrinsically $n$-linked in $\mathbb{R}P^3$. In particular, three disjoint copies of intrinsically triple-linked graphs in $\mathbb{R}^3$ are intrinsically triple-linked in $\mathbb{R}P^3$*]{} If any of the three copies of the graph has all 0-homologous cycles, then it is crossing-change equivalent to a spatial embedding, and thus triple-linked, as its disjoint cycle pairs would have the same linking numbers as a spatial embedding. Else, all three copies have at least one 1-homologous cycle. Then we have three disjoint 1-homologous cycles, and thus have a triple-link. As shown above, $K_{10}$ is an example of a one-component graph that is intrinsically triple-linked in $\mathbb{R}^3$. In the following section, we will exhibit two examples of minor-minimal intrinsically triple-linked graphs, each comprised of two components, that are intrinsically triple-linked in $\mathbb{R}^3$. The question remains whether there exists a minor-minimal intrinsically triple-linked graph of three components in $\mathbb{R}P^3$. We will use the following theorem: [[@BF] *Let $G$ be a graph containing two disjoint graphs from the Petersen family, $G_1$ and $G_2$ as subgraphs. If there are edges between the two subgraphs $G_1$ and $G_2$ such that the edges form a 6-cycle with vertices that alternate between $G_1$ and $G_2$, then $G$ is minor-minimal intrinsically triple-linked in $\mathbb{R}^3$.*]{} -2.1in ![A 3-linkless embedding of $K_6$ connected to $K_6$ along a 6-cycle in $\mathbb{R}P^3$.[]{data-label="K6K6"}](K6K6.eps "fig:") -1.7in If $G_1$ and $G_2$ are isomorphic to $K_6$, this result does not hold in $\mathbb{R}P^3$, as seen in Figure \[K6K6\]. [*If $G_1$ and $G_2$ are disjoint copies of $K_6$ connected to $K_6$ along a $6$-cycle with vertices that alternate between the copies of $K_6$, then $G = G_1 \sqcup G_2$ is minor-minimal intrinsically triple-linked in $\mathbb {R}P^3$.*]{} Embed $G$ in $\mathbb{R}P^3$. If $G_1$ or $G_2$ has all $0$-homologous cycles, $G$ will have a triple-link since $K_6$ connected to $K_6$ along a $6$-cycle with vertices that alternate between the copies of $K_6$ is triple-linked in $\mathbb{R}^3$. Thus, $G_1$ and $G_2$ each have a $1$-homologous cycle. Let $G_1$ be a graph on the vertex set $\{1, 2, 3, 4, 5, 6, A, B, C, D, E, F\}$ where $G[1, 2, 3, 4, 5, 6]$ and $G[A, B, C, D, E, F]$ are the copies of $K_6$ and the connecting edges are $(4, A)$, $(4, C)$, $(5, A)$, $(5, B)$, $(6, B)$, and $(6, C)$. Up to isomorphsim, there are five $3$-cycle equivalence classes in $G_1$. Consider $S = \{(1, 2, 3), (1, 2, 4), (1, 4, 5), (4, 5, 6), (4, 5, A)\}$, which contains one representative from each $3$-cycle class. We assume, without loss of generality, one cycle in $S$ is $1$-homologous. Consider $G[A, B, C, D, E, F]$. If there is a one homologous cycle in $G[B, C, E, F]$ then this cycle will link with the cycle in $S$ that is $1$-homologous. Since the cycle from $S$ links with the $1$-homologous cycle in $G_2$, we have a triple-link in $G$. Thus, we assume every cycle in $G[B, C, E, F]$ is $0$-homologous and so $G[A, B, C, D, E, F]$ has a pair of linked cycles. By the pigeon-hole principle, at least two edges connecting vertices from the set $\{A, B, C\}$ are in a linked cycle in $G[A, B, C, D, E, F]$, so, without loss of generality, we may assume $v_A$ and $v_B$ are in one cycle. If the $1$-homologous cycle is in the subset $S_1 = \{ (1, 2, 3), (1, 2, 4), (1, 4, 5), (4, 5, 6)\}$, then there are disjoint edges from the $6$-cycle that connect the cycle from $S_1$ to the cycle containing $v_A$ and $v_B$. So, by Lemma \[LL\], $G$ has a triple-link. If $(4, 5, A)$ is the $1$-homologous cycle, consider $G[1, 2, 3, 4, 5, 6]$. If there is a one homologous cycle in $G[1, 2, 3, 6]$ then this cycle will link with $(4, 5, A)$ and the $1$-homologous cycle in $G_2$, so $G$ will have a triple-link. Else, $G[1, 2, 3, 4, 5, 6]$ has a pair of linked cycles. By the pigeon-hole principle, at least two vertices in the set $\{4, 5, 6\}$ are in a linked cycle within the embedding of one copy of $K_6$. Similarly, at least two vertices of $\{A, B, C\}$ are in a linked cycle in the other copy of $K_6$. As a result of the $6$-cycle, there are two disjoint edges between the cycles and Lemma \[LL\] then applies and $G$ is triple-linked. To see $G$ is minor-minimal with respect to intrinsic triple-linking in $\mathbb {R}P^3$, embed $G$ so that $G_1$ is embedded as in the drawing in Figure \[K6K6\] and $G_2$ is contained in a sphere that lies in the complement of $G_1$. Therefore, $G_1$ does not have any triple-links and no cycle in $G_1$ is linked with a cycle in $G_2$. Without loss of generality, if we delete an edge, contract an edge or delete any vertex on $G_2$, it will have an affine linkless embedding. Thus, we can re-embed $G_2$ within the sphere in each case. Thus, $G$ is minor-minimal for intrinsic triple-linking. -1.5in ![A 3-linkless embedding of $K_7$ connected to $K_7$ along an edge in $\mathbb{R}P^3$.[]{data-label="K7K7"}](K7K7.eps "fig:") -1in [@BF] [*Let $G$ be a graph formed by identifying an edge of $K_7$ with an edge from another copy of $K_7$. Then $G$ is intrinsically triple-linked in $\mathbb{R}^3$.*]{} If $G$ is isomorphic to $K_7$ connected to $K_7$ along an edge, this result does not hold in $\mathbb{R}P^3$, as seen in Figure \[K7K7\]. We will need the following lemma: \[5V\] [@REU07] [*Let $P$ be a Petersen-family graph and $v$ be a vertex of P. If every cycle of $P\setminus\{v\}$ is $0$-homologous in an embedding $f:P \rightarrow \mathbb{R}P^3$, then $f(P)$ contains a non-trivial link.*]{} [*[If $G_1$ and $G_2$ are disjoint copies of $K_7$ connected to $K_7$ along an edge, then $G = G_1 \sqcup G_2$ is intrinsically triple-linked in $\mathbb {R}P^3$.]{}*]{} If $G_1$ or $G_2$ have all $0$-homologous cycles, $G$ will have a triple-link since $K_7$ connected to $K_7$ along an edge is triple-linked in $\mathbb{R}^3$. Thus, $G_1$ and $G_2$ each have a $1$-homologous cycle. Let $G_1$ be a graph on the vertex set $\{1, 2, 3, 4, 5, 6, 7, A, B, C, D, E\}$ where $G[1, 2, 3, 4, 5, 6, 7]$ and $G[6, 7, A, B, C, D, E]$ are the copies of $K_7$ and the connecting edge is $(6, 7)$. Up to isomorphsim, there are three $3$-cycle equivalence classes in $G_1$. We consider $S = \{ (1, 2, 3), (1, 2, 7), (1, 6, 7)\}$, which contains one representative from each $3$-cycle class. We can assume, without loss of generality, at least one cycle of $S$ is $1$-homologous. [**Case 1:**]{} Let $(1, 2, 3)$ be a $1$-homologous cycle in $G_1$. Then $(1, 2, 3)$ links with the $1$-homologous cycle in $G_2$. Consider $G[A, B, C, D, E, 6]$. If $G[ A, B, C, D, E, 6]$ has a $1$-homologous cycle, then there are three disjoint $1$-homologous cycles, so we assume $G[A, B, C, D, E, 6]$ must be $0$-homologous and so $G[ A, B, C, D, E, 6]$ has a pair of linked cycles. Lemma \[LL\] applies with $v_7$ connecting to the cycle that uses $v_6$, following the proof in [@BF]. [**Case 2:**]{} Let $(1, 2, 7)$ be a $1$-homologous cycle in $G_1$. $(1, 2, 7)$ links with the $1$-homologous cycle in $G_2$. Consider $G[A, B, C, D, E, 6]$. If $G[ A, B, C, D, E, 6]$ has a $1$-homologous cycle, then there are three disjoint $1$-homologous cycles, so we assume $G[ A, B, C, D, E, 6]$ must be $0$-homologous and so $G[A, B, C, D, E, 6]$ has a pair of linked cycles. Lemma \[LL\] applies with $v_7$ connecting to the cycle that uses $v_6$, following the proof in [@BF]. [**Case 3:**]{} Let $(1, 7, 6)$ be a $1$-homologous cycle in $G_1$. $(1, 7, 6)$ links with the $1$-homologous cycle in $G_2$. Consider $G[A, B, C, D, E, 6]$. If $G[ A, B, C, D, E]$ has a $1$-homologous cycle, then there are three disjoint $1$-homologous cycles, so we assume $G[ A, B, C, D, E]$ must be $0$-homologous. Then, by Lemma \[5V\], $G[A, B, C, D, E, 6]$ has a pair of linked cycles. Lemma \[LL\] applies with $v_7$ connecting to the cycle with $v_6$, following the proof in [@BF]. We note that if $K_7$ connected to $K_7$ along an edge is minor-minimal with respect to triple-linking in $\mathbb{R}^3$, then we would also have that two disjoint copies of $K_7$ connected to $K_7$ along an edge is minor-minimal intrinsically triple-linked in $\mathbb{R}P^3$. However, the minor-minimality of this graph is still unknown in $\mathbb{R}^3$. We also note that $G(n)$, as defined in [@FFNP], is a one-component minor-minimal intrinsically $(n+1)$-linked graph in $\mathbb{R}P^3$, by the same arguement given in [@FFNP], since $K_{4,4}-e$ is intrinsically linked in both $\mathbb{R}^3$ and $\mathbb{R}P^3$. Graphs with linking number $\geq$ 1 in $\mathbb{R}P^3$ ====================================================== In $\mathbb{R}P^3$, there are intrinsically linked graphs for which there exists an embedding in which every pair of disjoint cycles has linking number less than 1. Work has been done in $\mathbb{R}^3$ to find graphs containing disjoint cycles with large linking number in every spatial embedding. Using the fact that $K_{10}$ is triple-linked in $\mathbb{R}^3$, Flapan [@F] showed that every spatial embedding of $K_{10}$ contains a two-component link $L\cup J$ such that, for some orientation, $lk(L, J) \geq 2.$ A similar argument using Theorem 8 yields the following proposition. [*Every projective embedding of $K_{10}$ contains a two-component link $L\cup J$ such that, for some orientation, $lk(L, J) \geq 1.$*]{} It remains an open question to determine if 10 is that smallest number for which this property holds. At this point, we know the smallest $n$ is such that $7 < n \leq 10$. [999]{} C. Adams, [*The Knot Book*]{}, American Mathematical Society, Providence, RI, 2004. G. Bowlin and J. Foisy, [*Some new intrinsically $3$-linked graphs*]{}, J. Knot Theory Ramifications, [**13**]{}, no. 8 (2004), 1021–1027. J. Bustamante, J. Federman, J. Foisy, K. Kozai, K. Matthews, K. McNamara, E. Stark, and K. Trickey, [*Intrinsically linked graphs in projective space*]{}, arXiv:0809.0454v1. J. H. Conway and C. McA. Gordon, [*Knots and links in spatial graphs*]{}, J. Graph Theory, [**7**]{}, no. 4 (1983), 445–453. R. Diestel, [*Graph Theory*]{}, Springer-Verlag Inc., New York, NY, 1997. E. Flapan, [*Intrinsic knotting and linking of complete graphs*]{}, Algebraic & Geometric Topology, [**2**]{} (2002), 371-380. E. Flapan, J. Foisy, R. Naimi, J. Pommersheim, [*Intrinsically $n$-linked graphs*]{}, J. Knot Theory Ramifications [**10**]{}, no. 8 (2001), 1143-1154. E. Flapan, H. Howards, D. Lawrence, B. Mellor, [*Intrinsic linking and knotting of graphs in arbitrary 3-manifolds*]{}, Algebraic and Geometric Number Topology [**6**]{} (2006), 1025-1035. E, Flapan, R.Naimi and J. Pommersheim, J. [ *Intrinsically triple-linked complete graphs*]{}, Top. App., [**115**]{} (2001), 239–246. F. Harary, [*On the notion of balance of a signed graph*]{}, Michigan Math J., [**2**]{} (1953-54), 143-146 (1955). R. Motwani, A. Raghunathan, and H. Saran, [ *Constructive results from graph minors: Linkless embeddings*]{}, 29th Annual Symposium on Foundations of Computer Science, IEEE, 1988, pp. 398-409. M. Mroczkowski, [*Diagrammatic unknotting of knots and links in the projective space*]{}, J. Knot Theory and Its Ramifications, [**12**]{} (2003), 637-651. N. Robertson and P. Seymour, [*Graph minors XX. Wagner’s Conjecture*]{}, J. Combin Theory Ser B, [**92**]{}, no. 2 (2004), 325-357. N. Robertson, P. Seymour,and R. Thomas, [ *Sachs’ linkless embedding conjecture*]{}, J. Combin. Theory Ser. B, [ **64**]{}, no. 2 (1995), 185–227. H. Sachs, [*On spatial representation of finite graphs*]{}, (Proceedings of a conference held in Lagow, February 10-13, 1981, Poland), Lecture Notes in Math., 1018, Springer-Verlag, Berlin, Heidelberg, New York, and Tokyo (1983). H. Sachs, [*On spatial representations of finite graphs, finite and infinite sets*]{}, (A. Hajnal, L. Lovasz, and V. T. Sós, eds), colloq. Math. Soc. János Bolyai, vol. 37, North-Holland, Budapest, (1984), 649-662. T. Zaslavsky, [*The projective-planar signed graphs*]{}, Discrete Math. 113 (1993), 233-247. [^1]: Department of Mathematics, SUNY Potsdam, Potsdam, NY 13676 [^2]: Department of Mathematics, SUNY Potsdam, Potsdam, NY 13676 [^3]: Department of Mathematics & Statistics, James Madison University, Harrisonburg, VA 22807 [^4]: Department of Mathematics, Pomona College, Claremont, CA 91711 [^5]: This material is based upon work supported by the National Science Foundation under Grant No. 0646847, and the National Security Administration under Grant No. 42652.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study the eigenvalues of Schrödinger operators $-\Delta_{{\mathbb{R}}^2} + V_{\varepsilon}$ on ${\mathbb{R}}^2$ with rapidly oscillatory potential $V_{\varepsilon}(x) = W(x,x/{\varepsilon})$, where $W(x,y) \in C^\infty_0({\mathbb{R}}^2 \times {{\mathbb{T}}}^2)$ satisfies $\int_{{{\mathbb{T}}}^2} W(x,y) dy = 0$. We show that for ${\varepsilon}$ small enough, such operators have a unique negative eigenvalue, that is exponentially close to $0$.' author: - Alexis Drouot title: Bound states for rapidly oscillatory Schrödinger operators in dimension $2$ --- Introduction. ============= We study the $L^2$-eigenvalues of the Schrödinger operator $-\Delta_{{\mathbb{R}}^2}+V_{\varepsilon}$ on ${\mathbb{R}}^2$, where ${\varepsilon}$ is a small parameter and $V_{\varepsilon}$ is a real-valued, compactly supported, and rapidly oscillatory potential on ${\mathbb{R}}^2$: $$V_{\varepsilon}(x) = W(x,x/{\varepsilon}), \ \ W(x,y) = \sum_{k \in {\mathbb{Z}}^2 \setminus 0} W_k(x) e^{iky}.$$ The functions $W_k$ are smooth, with support in ${\mathbb{B}}(0,L) = \{ x \in {\mathbb{R}}^2, |x| < L \}$, and satisfy $\overline{W_k}=W_{-k}$. The potential $V_{\varepsilon}$ is a first approximation to model disordered medias with scale of heterogeneity $\sim {\varepsilon}$ – oscillations play here the role of randomness. This note provides a simple proof of a conjecture of Duchêne–Vukićević–Weinstein [@DVW]: \[thm:1\] For ${\varepsilon}$ small enough, the operator $-\Delta_{{\mathbb{R}}^2}+V_{\varepsilon}$ has a unique bound state, with energy $E_{\varepsilon}$ given by $$\label{eq:1t} E_{\varepsilon}= - \exp\left( - \dfrac{4\pi}{{\varepsilon}^2 \int_{{\mathbb{R}}^2} \Lambda_0(x) dx + o({\varepsilon}^2) } \right), \ \ \ \Lambda_0(x) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}\sum_{k \neq 0} \dfrac{|W_k(x)|^2}{|k|^2}.$$ In one dimension, the study of the spectral properties of $-{{\partial}}_x^2+V_{\varepsilon}$ with $V_{\varepsilon}$ rapidly oscillatory originated with Borisov–Gadyl’shin [@BG], who gave a sufficient condition for the existence of a bound state, and derived the asymptotic of the corresponding energy: $$\label{eq:1o} E_{\varepsilon}= -\dfrac{{\varepsilon}^4}{4} \int_{{\mathbb{R}}} \Lambda_0(x) dx, \ \ \Lambda_0(x) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}\sum_{k \neq 0} \dfrac{|W_k(x)|^2}{|k|^2}.$$ This study was continued later in Borisov [@B] then Duchêne–Weinstein [@DW], in particular to include less regular potentials. Duchêne–Vukićević–Weinstein [@DVW] studied the behavior of scattering quantities of Schrödinger operators with potentials which are the sum of a slowly varying term $W_0$ and a rapidly oscillatory term $V_{\varepsilon}$. They showed that the transmission coefficient of the *effective potential* $W_0-{\varepsilon}^2 \Lambda_0$, which is a $O({\varepsilon}^2)$-perturbation of $W_0$, differs from the one of $W_0+V_{\varepsilon}$ by $O({\varepsilon}^3)$. When $W_0 = 0$, they recovered the result of [@BG]: $-{{\partial}}_x^2+V_{\varepsilon}$ has a unique negative eigenvalue satisfying for small ${\varepsilon}$. They proved a uniform weighted dispersive estimate for the propagator $e^{it(-{{\partial}}_x^2+V_{\varepsilon})}$ as ${\varepsilon}\rightarrow 0$, despite the presence of an eigenvalue near the edge of the continuous spectrum. Recently, Duchêne–Raymond [@DR] obtained homogenization results for potentials of the form ${\varepsilon}^{-\beta} V_{\varepsilon}$, $\beta \in (0,2)$, in dimension $1$, using a normal form approach. Dimassi [@Di] applied ${\varepsilon}$-semiclassical calculus to show a trace formula and a Weyl law for potentials of the form ${\varepsilon}^{-2} V_{\varepsilon}$, in any dimension $d$. Motivated by [@DVW] and by Christiansen [@Ch06], we used a different approach to study in [@D] the resonances and eigenvalues of $-\Delta_{{\mathbb{R}}^d}+W_0+V_{\varepsilon}$, in any odd dimension $d$. When $W_0 = 0$, we proved that the resonances and eigenvalues of $V_{\varepsilon}$ escape all bounded regions as ${\varepsilon}\rightarrow 0$ (except the one converging to $0$ when $d=1$). When $W_0 \neq 0$, we showed that the resonances and eigenvalues of $W_0+V_{\varepsilon}$ converge to the one of $W_0$, with a complete expansion in powers of ${\varepsilon}$. We improved upon the homogeneization result of [@DVW], refining the effective potential $W_0-{\varepsilon}^2 \Lambda_0$ to $W_0-{\varepsilon}^2 \Lambda_0-{\varepsilon}^3 \Lambda_1$, and deriving it for any odd $d$. We refer to [@D Figure 2] for numerical results and to [@D §1] for additional references. A famous result of Simon [@Si Theorem 3.4] in dimension $2$ predicts that under suitable conditions, a Schrödinger operator with small negative potential $-\Delta_{{\mathbb{R}}^2} + \epsilon\Lambda$ has a unique bound state with energy $$\label{eq:1f} E_{\varepsilon}= -\exp\left( -\dfrac{4\pi}{\epsilon \int_{{\mathbb{R}}^2} \Lambda(x) dx+ o(\epsilon)} \right).$$ This identity put together with supports the main idea of [@DVW]: the scattering quantities of rapidly oscillatory potentials are similar to the one of suitable small potentials. To prove Theorem \[thm:1\], we first follow [@Si §3]: we use a modified Fredholm determinant to reduce the study of eigenvalues of $-\Delta_{{\mathbb{R}}^2}+V_{\varepsilon}$ to an equation involving a certain trace. Then, we provide estimates on this trace following some ideas of [@D], ending the proof. We mention that the result of Theorem \[thm:1\] still applies when $W$ is not smooth but satisfies instead the weaker bound $$\sum_{k \neq 0} |W_k|_\infty + \dfrac{|W_k|_{C^1}}{|k|}+ \dfrac{|W_k|_{C^2}}{|k|^2}+ \dfrac{|W_k|_{C^3}}{|k|^3} + \sum_{0 \neq k \neq \ell} \dfrac{|W_k|_{C^3}|W_\ell|_{C^3}}{|k-\ell|^{5/2}} < \infty,$$ with no change in the proof. **Aknowledgement.** We would like to thanks Maciej Zworski for valuable discussions. This research was supported by the NSF grant DMS-1500852 and the Fondation CFM pour la recherche. Notations {#notations .unnumbered} --------- We will use the following: - For $x\in {\mathbb{R}}^2$, ${\langle x \rangle}$ denotes the Japanese bracket of $x$: ${\langle x \rangle} {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}(1+|x|^2)^{1/2}$. - The function $z \mapsto \ln(z)$ denotes the holomorphic logarithm on ${\mathbb{C}}\setminus (-\infty,0]$. - $C_0^\infty({\mathbb{R}}^2,{\mathbb{R}})$ is the set of real smooth compactly supported functions on ${\mathbb{R}}^2$. - $L^2$ denotes the Hilbert space of functions that are squared integrable, with $L^2$-norm denoted by $|f|_2$ and scalar product given by ${\langle f, g \rangle}_2 = \int_{{\mathbb{R}}^2} \overline{f} g$. - $C^k$ is the space of functions on ${\mathbb{R}}^2$ with $k$ continuous and bounded derivatives, with norm $|f|_{C^k} = |f|_\infty+|f'|_\infty+...+|f^{(k)}|_\infty$. - We write $\Delta$ for the Laplacian $\Delta_{{\mathbb{R}}^2}$, and $\hat{f}$ for the Fourier transform of $f$: $\hat{f}(\xi) = \frac{1}{2\pi} \int_{{\mathbb{R}}^2} e^{-ix\xi} f(x) dx$. $H^s$ denotes the standard Sobolev space on ${\mathbb{R}}^2$, with norm $|f|_{H^s} = |{\langle \xi \rangle}^s \hat{f}|_2 = |({{\operatorname{Id}}}- \Delta)^{s/2}f|_2$. - ${\mathcal{B}}$ is the Banach space of bounded linear operators from $L^2$ to itself. - ${{\mathcal{L}}}^2$ is the Banach space of Hilbert–Schmidt operators on $L^2$. General properties. =================== Let $R_0(\lambda,x,y)$ be the kernel of the free resolvent $(-\Delta + \lambda^2)^{-1}$. It its a Hankel function of $\lambda |x-y|$ and one can write $$R_0(\lambda,x,y) = - \dfrac{1}{2\pi} \ln(\lambda) + H_0(\lambda,x,y),$$ where for every $x \neq y$ the function $\lambda \mapsto H_0(\lambda,x,y)$ is holomorphic in $\{{\operatorname{Re}}\lambda > 0\}$ and admits a continuous extension to $\{ {\operatorname{Re}}\lambda \geq 0 \}$ – see [@Si (12)]. For $\lambda$ with ${\operatorname{Re}}\lambda \geq 0$, let $H_0(\lambda)$ be the operator with kernel $H_0(\lambda,x,y)$. When $\lambda \in (0,\infty)$, $R_0(\lambda)$ is selfadjoint and $\ln(\lambda)$ is real, hence $H_0(\lambda)$ is selfadjoint. Let ${{\mathcal{V}}}\in C_0^\infty({\mathbb{R}}^2,{\mathbb{R}})$, with support in ${\mathbb{B}}(0,L) = \{ x \in {\mathbb{R}}^2, |x| \leq L \}$, and $\rho \in C_0^\infty({\mathbb{R}}^2, {\mathbb{R}})$, equal to $1$ on the support of ${{\mathcal{V}}}$ and $0$ outside ${\mathbb{B}}(0,L)$. We will be interested in the behavior of $K_{{\mathcal{V}}}(\lambda) = \rho R_0(\lambda) {{\mathcal{V}}}$ for $ \lambda$ close to $0$. Write $K_{{\mathcal{V}}}(\lambda) = \Pi_{{\mathcal{V}}}\ln(\lambda) + L_{{\mathcal{V}}}(\lambda)$, $$\Pi_{{\mathcal{V}}}{\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}-\dfrac{1}{2\pi} \rho \otimes {{\mathcal{V}}}, \ \ \ L_{{\mathcal{V}}}(\lambda) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}\rho H_0(\lambda) {{\mathcal{V}}}, \ \ \ {\operatorname{Re}}\lambda \geq 0.$$ An operator $A : C^\infty_0({\mathbb{R}}^2) \rightarrow {{\mathcal{D}}}'({\mathbb{R}}^2)$ belongs to ${{\mathcal{L}}}^2$ (the Hilbert–Schmidt class) if and only if its kernel $A(x,y)$ is in $L^2({\mathbb{R}}^4,dxdy)$. In this case the ${{\mathcal{L}}}^2$-norm is $$|A|_{{{\mathcal{L}}}^2} {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}\left(\int_{{\mathbb{R}}^2 \times {\mathbb{R}}^2} |A(x,y)|^2 dx dy\right)^{1/2}.$$ We prove the following result, similar to [@Si Proposition 3.2]: \[lem:1b\] The operator $L_\rho(\lambda)$ is in ${{\mathcal{L}}}^2$. Moreover, uniformly locally in $\lambda, \mu$ with nonnegative real parts, $$\label{eq:1b} |L_\rho(\lambda)-L_\rho(\mu)|_{\mathcal{B}}= O(|\lambda \ln(\lambda)-\mu \ln(\mu)|), \ \ \ \ |L_\rho(\lambda)|_{L^2 \rightarrow H^2} = O(1).$$ To prove the first part of , we write the kernel of $H_0({{\lambda}})$ as $$H_0(\lambda,x,y) = -\dfrac{1}{2\pi} \ln |x-y| + F(\lambda|x-y|).$$ The function $F : {\mathbb{C}}\rightarrow {\mathbb{C}}$ takes the form $F(\zeta) = \zeta h(\zeta) \ln(\zeta) + g(\zeta)$ for some entire functions $g,h$ – see [@Si (13)]. For $x \in {\mathrm{supp}}(\rho), y \in {\mathrm{supp}}(\rho)$ and $\lambda$ in a compact set, $\lambda |x-y|$ is uniformly bounded. Hence, there exists a constant $C$ such that for such values of $x,y,\mu,\lambda$, $$|F(\lambda|x-y|) - F(\mu |x-y|)| \leq C |\lambda \ln(\lambda) - \mu \ln(\mu)| \cdot |\ln |x-y||.$$ The bound $|L_\rho(\lambda)-L_\rho(\mu)|_{\mathcal{B}}= O(|\lambda \ln(\lambda)-\mu \ln(\mu)|)$ follows now from Schur’s test and from the local integrability of $\ln|x-y|$ on ${\mathbb{R}}^2$. The operator $L_\rho(\lambda)$ belongs to $ {{\mathcal{L}}}^2$ by [@Si Proposition 3.2]. The second estimate of amounts to prove that $\Delta L_\rho(\lambda)$ belongs to ${\mathcal{B}}$, uniformly locally for $\lambda \in K$. By the same argument as in [@DZ Theorem 2.1], it suffices to show that $\rho \Delta H_0(\lambda) \rho$ belongs to ${\mathcal{B}}$. Recall that $\Delta \ln|x| = -2\pi \delta$, so that $$\label{eq:1g} \Delta_x H_0(\lambda,x,y) = \delta(x-y) + \Delta_x (F(\lambda|x-y|))$$ We now compute the Laplacian of $F(\lambda|x-y|)$ with respect to $x$. Note that $$\label{eq:1w} \Delta (g(\lambda |z|)) = 4{\dfrac{\partial ^2 g(\lambda |z|)}{\partial z {{\partial}}{\overline{z}}}} = \dfrac{\lambda}{|z|} g'(\lambda |z|) + \lambda^2 g''(\lambda |z|).$$ Define $f(\lambda |z|) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}\lambda |z| h(\lambda |z|)$. By applying to $f \cdot \ln$ instead of $g$, we get \[eq:1x\] (f(|z|) (|z|)) = -2f(|z|) + + ( + \^2 f”(|z|)) (|z|)\ = + ( + \^2 f”(|z|)) (|z|). In the second line we used that the product of a smooth function vanishing at $0$ with $\delta$ vanishes. The right hand sides of and both define locally integrable functions of $z \in {\mathbb{C}}$, uniformly locally in $\lambda$. Since $\rho$ is compactly supported, Schur’s test combined with shows that $\rho \Delta H_0(\lambda) \rho$ belongs to ${\mathcal{B}}$ uniformly in $\lambda$. This concludes the proof. Since $L_{{\mathcal{V}}}(\lambda)$ is in ${{\mathcal{L}}}^2$, we can define the modified Fredholm determinant $d_{{\mathcal{V}}}(\lambda)$ by $$d_{{\mathcal{V}}}(\lambda) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}{{\operatorname{Det}}}({{\operatorname{Id}}}+ \Psi(L_{{\mathcal{V}}}(\lambda))), \ \ \Psi(z) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}(1+z)e^{-z}-1, \ \ {\operatorname{Re}}(\lambda) \geq 0,$$ see [@DZ Appendix B]. As $K_{{\mathcal{V}}}(\lambda)$ is the sum of the rank one operator $\Pi_{{\mathcal{V}}}\ln(\lambda)$ with $L_{{\mathcal{V}}}(\lambda)$, $K_{{\mathcal{V}}}(\lambda)$ belongs to ${{\mathcal{L}}}^2$. We define $$D_{{\mathcal{V}}}(\lambda) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}{{\operatorname{Det}}}({{\operatorname{Id}}}+ \Psi(K_{{\mathcal{V}}}(\lambda))), \ \ {\operatorname{Re}}(\lambda) > 0.$$ The negative eigenvalues of $-\Delta + {{\mathcal{V}}}$ are exactly the numbers of the form $-\lambda^2$, where $\lambda \in (0,\infty)$ is a zero of $D_{{\mathcal{V}}}(\lambda)$, see [@GLMZ05 Theorem 5.4]. The next result is similar to [@Si Theorem 3.3]. \[lem:1c\] Let $\lambda$ such that ${\operatorname{Re}}\lambda > 0$ and ${{\operatorname{Id}}}+ L_V(\lambda)$ is invertible on $L^2$. Then, $$D_{{\mathcal{V}}}(\lambda) = 0 \ \Leftrightarrow \ 1+ \ln(\lambda) \varphi_{{\mathcal{V}}}(\lambda) = 0, \ \ \ \ \varphi_{{\mathcal{V}}}(\lambda) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}{{\operatorname{Tr}}}\left(({{\operatorname{Id}}}+ L_{{\mathcal{V}}}(\lambda))^{-1} \Pi_{{\mathcal{V}}}\right).$$ If moreover $\lambda \in (0,\infty)$ then $\varphi_{{\mathcal{V}}}(\lambda) \in {\mathbb{R}}$. To simplify the notations of this proof, we simply write $L_{{\mathcal{V}}}, K_{{\mathcal{V}}}$ for the operators $L_{{\mathcal{V}}}(\lambda), K_{{\mathcal{V}}}(\lambda)$. Let $\lambda$ with ${\operatorname{Re}}\lambda > 0$ and ${{\operatorname{Id}}}+ L_{{\mathcal{V}}}$ invertible on $L^2$. Since $K_{{\mathcal{V}}}= \Pi_{{\mathcal{V}}}\ln(\lambda) + L_{{\mathcal{V}}}$ we have $$\label{eq:1h} {{\operatorname{Id}}}+ K_{{\mathcal{V}}}=({{\operatorname{Id}}}+ L_{{\mathcal{V}}}) \left( {{\operatorname{Id}}}+ ({{\operatorname{Id}}}+ L_{{\mathcal{V}}})^{-1} \Pi_{{\mathcal{V}}}\ln(\lambda) \right).$$ Recall that $1+z = (1+\Psi(z)) e^z$ so that for every bounded operator $A$, ${{\operatorname{Id}}}+ A = ({{\operatorname{Id}}}+ \Psi(A)) e^A$. This identity combined with shows that $$\label{eq:1i} e^{K_{{\mathcal{V}}}}\left({{\operatorname{Id}}}+ \Psi(K_{{\mathcal{V}}}) \right) =e^{L_{{\mathcal{V}}}}({{\operatorname{Id}}}+ \Psi(L_{{\mathcal{V}}})) \left( {{\operatorname{Id}}}+ ({{\operatorname{Id}}}+ L_{{\mathcal{V}}})^{-1} \Pi_{{\mathcal{V}}}\ln(\lambda) \right).$$ We observe that $K_{{\mathcal{V}}}-L_{{\mathcal{V}}}= \ln(\lambda) \Pi_{{\mathcal{V}}}$. Since $\Pi_{{\mathcal{V}}}$ has kernel $-\rho \otimes {{\mathcal{V}}}/(2\pi)$, its trace is equal to ${\alpha}= -\int_{{\mathbb{R}}^2} {{\mathcal{V}}}/(2\pi)$. Taking the determinant on both sides of we obtain $$\lambda^{\alpha}D_{{\mathcal{V}}}(\lambda) = d_{{\mathcal{V}}}(\lambda) \cdot {{\operatorname{Det}}}\left( {{\operatorname{Id}}}+ ({{\operatorname{Id}}}+ L_{{\mathcal{V}}})^{-1} \Pi_{{\mathcal{V}}}\ln(\lambda) \right).$$ Since the operator $({{\operatorname{Id}}}+ L_{{\mathcal{V}}})^{-1} \Pi_{{\mathcal{V}}}\ln(\lambda)$ is of rank one, the determinant of ${{\operatorname{Id}}}+ ({{\operatorname{Id}}}+ L_{{\mathcal{V}}})^{-1} \Pi_{{\mathcal{V}}}\ln(\lambda)$ is equal to $1+{{\operatorname{Tr}}}\left(({{\operatorname{Id}}}+ L_{{\mathcal{V}}})^{-1} \Pi_{{\mathcal{V}}}\ln(\lambda) \right)$. Defining $\varphi_{{\mathcal{V}}}(\lambda) = {{\operatorname{Tr}}}\left(({{\operatorname{Id}}}+ L_{{\mathcal{V}}})^{-1} \Pi_{{\mathcal{V}}}\right)$, we obtain $$\label{eq:1u} \lambda^{\alpha}D_{{\mathcal{V}}}(\lambda) = d_{{\mathcal{V}}}(\lambda) \cdot \left( 1+\ln(\lambda) \varphi_{{\mathcal{V}}}(\lambda) \right).$$ Since ${\operatorname{Re}}\lambda > 0$ we have $\lambda \neq 0$ and the first part of the lemma follows. For the second part of the lemma, it suffices to prove that $D_{{\mathcal{V}}}(\lambda)$ and $d_{{\mathcal{V}}}(\lambda)$ are both real when $\lambda > 0$, thanks to . For such $\lambda$, $R_0(\lambda)$ is selfadjoint. Hence, $$\overline{D_{{\mathcal{V}}}(\lambda)} = {{\operatorname{Det}}}({{\operatorname{Id}}}+ \Psi(K_{{\mathcal{V}}}^*)) = {{\operatorname{Det}}}({{\operatorname{Id}}}+ \Psi({{\mathcal{V}}}R_0(\lambda) \rho)).$$ Since $\Psi$ vanishes at $0$, there exists $\psi$ entire with $\Psi(z) = z\psi(z)$. Therefore, $$\label{eq:1v} \overline{D_{{\mathcal{V}}}(\lambda)} = {{\operatorname{Det}}}({{\operatorname{Id}}}+ {{\mathcal{V}}}R_0(\lambda) \rho \psi({{\mathcal{V}}}R_0(\lambda) \rho)) = {{\operatorname{Det}}}({{\operatorname{Id}}}+ \rho R_0(\lambda) \rho \cdot \psi({{\mathcal{V}}}R_0(\lambda) \rho) {{\mathcal{V}}}).$$ In the last equality, we used that if $B \in {\mathcal{B}}$ and $A$ is trace-class then ${{\operatorname{Det}}}({{\operatorname{Id}}}+ BA) = {{\operatorname{Det}}}({{\operatorname{Id}}}+ AB)$, see [@DZ (B.5.13)]. The power series expansion of $\psi(z) = \sum_{m=0}^\infty a_m z^m$ implies that $\rho R_0(\lambda) \rho \cdot \psi({{\mathcal{V}}}R_0(\lambda) \rho) {{\mathcal{V}}}$ $$= \rho R_0(\lambda) \rho \sum_{m=0}^\infty a_m ({{\mathcal{V}}}R_0(\lambda) \rho)^m {{\mathcal{V}}}= \sum_{m=0}^\infty a_m (\rho R_0(\lambda) {{\mathcal{V}}})^{m+1} = \Psi(L_{{\mathcal{V}}}).$$ Equation now shows that $D_{{\mathcal{V}}}(\lambda) = \overline{D_{{\mathcal{V}}}(\lambda)}$ when $\lambda \in (0,\infty)$. The same arguments (using that $H_0(\lambda)$ is selfadjoint for $\lambda \in (0,\infty)$) shows that $d_{{\mathcal{V}}}(\lambda) \in {\mathbb{R}}$ if $\lambda \in (0,\infty)$. Hence, shows that $\varphi_{{\mathcal{V}}}(\lambda) \in {\mathbb{R}}$ when ${{\lambda}}\in (0,\infty)$. Bound state exponentially close to zero. ======================================== We now focus on the case of a potential $V = V_{\varepsilon}$ given by $$V_{\varepsilon}(x) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}W\left(x,\dfrac{x}{{\varepsilon}} \right), \ \ \ \ W(x,y) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}\sum_{k \in {\mathbb{Z}}^2 \setminus 0} W_k(x)e^{iky}.$$ Here the functions $W_k \in C^\infty({\mathbb{R}}^2)$ have supports in ${\mathbb{B}}(0,L) = \{ x \in {\mathbb{R}}^2, |x| < L \}$, and $\overline{W_k} = W_{-k}$. To simplify notations, we will drop the index ${\varepsilon}$ in $V_{\varepsilon}$ and write $V = V_{\varepsilon}$. We first investigate the invertibility of ${{\operatorname{Id}}}+ L_V(\lambda)$ for ${\varepsilon}$ small enough. We will need the following lemma regarding the behavior as ${\varepsilon}\rightarrow 0$ of certain oscillatory integrals: \[lem:1a\] The following estimates hold: $$|V|_{H^{-2}} = O({\varepsilon}^2), \ \ \ \ \ \left| {\langle D \rangle}^{-2} V {\langle D \rangle}^{-2} \right|_{\mathcal{B}}= O({\varepsilon}^2), \ \ {\langle D \rangle} {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}({{\operatorname{Id}}}- \Delta)^{1/2}.$$ We estimate $|V|_{H^{-2}}$ in a similar way as in the proof of [@D Theorem 1]: |V|\_[H\^[-2]{}]{} = | \^[-2]{} \_[k 0]{} (-k/) |\_2 \_[k 0]{} | \^[-2]{}[-k/]{}\^[-2]{}[-k/]{}\^2(-k/) |\_2\ \_[k 0]{} |\^[-2]{}[-k/]{}\^[-2]{}|\_|W\_k|\_[H\^2]{} C\_[k 0]{} [k/]{}\^[-2]{} |W\_k|\_[H\^2]{} = O(\^2). In the last line we used Peetre’s inequality: for $t > 0$, $x,y \in {\mathbb{R}}^2$, $$\label{eq:1m} {\langle x \rangle}^{-t} {\langle y \rangle}^{-t} \leq 2^t {\langle x-y \rangle}^{-t}.$$ This shows the first estimate. The boundedness of ${\langle D \rangle}^{-2} V {\langle D \rangle}^{-2}$ on $L^2$ is equivalent to the boundedness of the multiplication operator by $V$ from $H^2$ to $H^{-2}$. We recall that $H^2$ is an algebra in dimension $2$, and that there exists $C > 0$ such that for any $u,f \in H^2$, $|fu|_{H^2} \leq C|f|_{H^2} |u|_{H^2}$. The multiplication operator by $u \in H^2$ is bounded from $H^2$ to $H^2$, hence from $H^{-2}$ to $H^{-2}$ with same norm and we deduce the inequality $|fu|_{H^{-2}} \leq C|f|_{H^{-2}} |u|_{H^2}$. Applying this with $V=f$ shows that $|V|_{H^2 \rightarrow H^{-2}} \leq C |V|_{H^{-2}} = O({\varepsilon}^2)$. This ends the proof \[lem:1d\] For every compact subset $K$ of $\{\lambda : {\operatorname{Re}}(\lambda) \geq 0\}$, there exist $C,{\varepsilon}_0 > 0$ such that for all $0 < {\varepsilon}\leq {\varepsilon}_0$, the inverse of ${{\operatorname{Id}}}+ L_V(\lambda)$ exist, is given by a convergent Neumann series, and satisfies $$\left|({{\operatorname{Id}}}+ L_V(\lambda))^{-1}\right|_{\mathcal{B}}\leq C.$$ To prove the lemma it suffices to show that $L_V(\lambda)^2$ is bounded with $|L_V(\lambda)^2|_{\mathcal{B}}< \frac{1}{2}$, for ${\varepsilon}$ small enough. Recall that $V \in L^\infty$; and (Lemma \[lem:1b\]) that for $\lambda \in K$, $L_\rho(\lambda)$ is uniformly bounded from $L^2$ to $H^2$ (hence by the adjoint bound, from $H^{-2}$ to $L^2$). These facts imply $$|L_V(\lambda)^2|_{\mathcal{B}}\leq C|L_\rho(\lambda) V L_\rho(\lambda)|_{\mathcal{B}}\leq C |{\langle D \rangle}^{-2}V{\langle D \rangle}^{-2}|_{\mathcal{B}}.$$ The RHS is $O({\varepsilon}^2)$ by Lemma \[lem:1a\]. This ends the proof. The negative eigenvalues $-\lambda^2$ of the operator $-\Delta + V$ on $L^2$ all belong to a fixed compact set: if $-\lambda^2$ is a negative eigenvalue, then there exists $0 \neq u \in L^2$ such that $$-\Delta u + V u + \lambda^2 u = 0.$$ Multiplying by $\overline{u}$ on both sides and integrating by parts, we obtain $$|\nabla u|_2^2 + {\langle Vu, u \rangle}_2 + \lambda^2 |u|_2^2 = 0.$$ Thus, $\lambda^2 |u|_2^2 \leq |V|_\infty |u|_2^2$, which shows that $-\lambda^2 \in [-|V|_\infty, 0] \subset [-M^2,0]$ where $M = 2+\sum_{k \neq 0} |W_k|_\infty$ is independent of ${\varepsilon}$. This fact, together with Lemma \[lem:1c\] and \[lem:1d\] (applied with $K = K_0 {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}[0, M]$), implies that for ${\varepsilon}$ small enough the set of negative eigenvalues of $-\Delta + V$ is exactly $\{-\lambda^2\}$, where $\lambda$ solves the equation $$1+ \ln(\lambda) \varphi_V(\lambda) = 0, \ \ \lambda \in K_0, \ \ \varphi_V(\lambda) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}{{\operatorname{Tr}}}\left(({{\operatorname{Id}}}+ L_V(\lambda))^{-1} \Pi_V\right).$$ \[lem:1f\] Uniformly for $\lambda, \mu$ in $K_0 = [0,M]$, \[eq:1e\] |\_V()| = O(\^2),     |\_V()-\_V()| = O(|() - ()|). We work exclusively with $\lambda, \mu \in K_0$ and ${\varepsilon}$ small enough so that Lemma \[lem:1d\] applies. All the estimates below are uniform for such $\lambda, {\varepsilon}$. Write $({{\operatorname{Id}}}+ L_V(\lambda))^{-1} = {{\operatorname{Id}}}- L_V(\lambda) ({{\operatorname{Id}}}+ L_V(\lambda))^{-1}$ to get \[eq:1r\] \_V() = (\_V) - (L\_V() (+ L\_V())\^[-1]{} \_V) = ,\ f\_(+ L\_V())\^[-1]{}. The term ${\langle \rho,V \rangle}_2$ is clearly independent of $\lambda$ and equal to $O({\varepsilon}^\infty)$. Let $P_k = (k_1 D_{x_1} + k_2 D_{x_2})/|k|^2$, so that $P_k e^{ikx/{\varepsilon}} = {\varepsilon}e^{ikx/{\varepsilon}}$. Since $P_k$ is selfadjoint, ${\langle L_V(\lambda) f_\lambda, V \rangle}_2 =$ $${\varepsilon}^2 \sum_{k \neq 0} {\langle W_k \cdot L_V(\lambda) f_\lambda, \rho P_k^2 e^{-ik\bullet/{\varepsilon}} \rangle}_2 = {\varepsilon}^2 \sum_{k \neq 0} {\langle P_k^2 \left(W_k \cdot L_V(\lambda) f_\lambda\right), \rho e^{-ik\bullet/{\varepsilon}} \rangle}_2.$$ The Cauchy–Schwarz inequality yields |\_V()| O(\^) + C\^2 \_[k 0]{} |P\_k\^2 (W\_k L\_V() f\_)|\_2. By Lemma \[lem:1d\], the operator $({{\operatorname{Id}}}+ L_V(\lambda))^{-1}$ is uniformly bounded in ${\mathcal{B}}$, hence $|f_\lambda|_2 = O(1)$. Since $L_V(\lambda)$ maps $L^2$ to $H^2$ the operator $P_k^2 W_k L_V(\lambda)$ maps $L^2$ to $L^2$ with $$\left|P_k^2 W_k L_V(\lambda)\right|_{\mathcal{B}}\leq C\dfrac{|W_k|_{C^2}}{|k|^2} |V|_\infty.$$ Thus $|\varphi_V(\lambda)| = O({\varepsilon}^2)$ as claimed. Regarding the second estimate in , we use and the Cauchy–Schwarz inequality to see that |\_V()-\_V()| |L\_V() f\_- L\_V() f\_|\_2 |V|\_2\ C |L\_()-L\_()|\_+ C |f\_-f\_|\_2 C |L\_()-L\_()|\_. In the above we used $f_\lambda = ({{\operatorname{Id}}}+ L_V(\lambda))^{-1} \rho$, the resolvent identity, and the uniform boundedness of $({{\operatorname{Id}}}+ L_V(\lambda))^{-1}$. The conclusion follows now from Lemma \[lem:1b\]. We now identify $\lim_{{\varepsilon}\rightarrow 0} {\varepsilon}^{-2}\varphi_V(\lambda)$: \[lem:1e\] Uniformly for $\lambda \in K_0$, $$\label{eq:1c} \varphi_V(\lambda) = \dfrac{{\varepsilon}^2}{2\pi} \int_{{\mathbb{R}}^2} \Lambda_0(x) dx + o({\varepsilon}^2), \ \ \Lambda_0(x) {\mathrel{\stackrel{\makebox[0pt]{\mbox{\normalfont\tiny def}}}{=}}}\sum_{k \neq 0} \dfrac{|W_k(x)|^2}{|k|^2}.$$ 1\. We claim that it is enough to show for $\lambda \in [1,2]$. Indeed, Lemma \[lem:1f\] shows that ${\varepsilon}^{-2} \varphi_V$ is uniformly bounded. In addition, it is a holomorphic function of $\lambda \in K_0$. Hence, after possibly passing to a subsequence, it converges uniformly locally to a holomorphic function. If holds for $\lambda \in [1,2]$, ${\varepsilon}^{-2} \varphi_V$ has only one accumulation point on $[1,2]$, hence – by the unique continuation principle – only one accumulation point on $K_0$. This would prove that holds uniformly on $K_0$. Below we show for $\lambda \in [1,2]$. We will always assume that ${\varepsilon}$ is small enough so that Lemma \[lem:1d\] holds. 2\. By expanding $({{\operatorname{Id}}}+ L_V(\lambda))^{-1}$ into a finite Born series, \_V() = (\_V) - (L\_V() \_V) + (L\_V() (+ L\_V())\^[-1]{} L\_V() \_V)\ = - \_[\^2]{} V(x) dx + [L\_V() , V ]{}\_2 - [(+ L\_V())\^[-1]{} L\_V() , L\_V()\^\* V ]{}\_2. The first term is $O({\varepsilon}^\infty)$. We note that since $\lambda \in [1,2]$, $L_V(\lambda)^* = L_V(\lambda)$ and the third term can be bounded using the Cauchy–Schwarz inequality: $$\label{eq:1d} \left|{\langle ({{\operatorname{Id}}}+ L_V(\lambda))^{-1} L_V(\lambda) \rho, L_V(\lambda) V \rangle}_2\right| \leq \left|({{\operatorname{Id}}}+ L_V(\lambda))^{-1}\right|_{\mathcal{B}}|L_V(\lambda) \rho|_2 |L_V(\lambda) V|_2.$$ We note that $L_V(\lambda) \rho = L_\rho(\lambda) V$ and we recall that $L_V(\lambda)$ is bounded from $H^{-2}$ to $L^2$. Hence Lemma \[lem:1a\] implies that $|L_V(\lambda) \rho|_2 \leq C |V|_{H^{-2}} = O({\varepsilon}^2)$. Similarly, $$|L_V(\lambda) V|_2 \leq C |L_\rho(\lambda) V| = O({\varepsilon}^2)$$ This proves that the RHS of is $O({\varepsilon}^4)$. The steps below are devoted to estimating the leading term in the expansion of $\varphi_V(\lambda)$: ${\langle L_V(\lambda) \rho, V \rangle}_2$. 3\. Using $L_V(\lambda) = K_V(\lambda)-\Pi_V \ln(\lambda)$, we have uniformly on $[1,2]$ \[eq:1z\] [L\_V() , V ]{}\_2 = \_[\^2 \^2]{} R\_0(,x,y) V(y) V(x) dx dy + (\_[\^2]{}V(x) dx)\^2\ = \_[\^2 \^2]{} \_(x-y) V(y) V(x) dx dy + O(\^). Here the function $\phi_\lambda$ has Fourier transform equal to $\widehat{\phi_\lambda}(\xi) = (|\xi|^2+\lambda^2)^{-1}$. The identity and the Plancherel formula show that ${\langle L_V(\lambda) \rho, V \rangle}_2$ is equal modulo $O({\varepsilon}^\infty)$ to [ , ]{}\_2 = \_[\^2]{} d. As the Fourier transform of $V$ is given by $\sum_{k \neq 0} \widehat{W_k}(\xi-k/{\varepsilon})$ we obtain [L\_V() , V ]{}\_2 = \_[k,]{} I\[W\_k,W\_\] + O(\^),      I\[W\_k,W\_\] \_[\^2]{} d. We will study $I[W_k,W_\ell]$ depending whether $k+\ell \neq 0$ or $k+\ell = 0$. 4\. Assume that $k+\ell \neq 0$. We have |I\[W\_k,W\_\]| C |W\_k|\_[C\^3]{} |W\_|\_[C\^3]{} \_[\^2]{} \^[-2]{} [-k/]{}\^[-3]{} [-+ /]{}\^[-3]{} d\ C\^[5/2]{} \_[\^2]{} \^[-2]{} [-k/]{}\^[-1/2]{} [-+ /]{}\^[-1/2]{} d To get the second line, we controlled ${\langle \xi-k/{\varepsilon}\rangle}^{-5/2} {\langle -\xi + \ell/{\varepsilon}\rangle}^{-5/2}$ by ${\langle (k-\ell)/{\varepsilon}\rangle}^{-5/2}$ with Peetre’s inequality . Hölder’s inequality yields $$|I[W_k,W_\ell]| \leq C{\varepsilon}^{5/2} \dfrac{|W_k|_{C^3} |W_\ell|_{C^3}}{|k-\ell|^{5/2}}.$$ 5\. We now focus on the case $k+\ell=0$. A substitution yields $$I[W_k, W_{-k}] = {\varepsilon}^{2} \int_{{\mathbb{R}}^2} \dfrac{\widehat{W_k}(\xi) \widehat{W_{-k}}(-\xi)}{|{\varepsilon}\xi+k|^2+({\varepsilon}\lambda)^2} d\xi.$$ We split the domain of integration ${\mathbb{R}}^2$ into two parts: ${\mathbb{B}}(0,{\varepsilon}^{-{3/4}})$ and ${\mathbb{R}}^2 \setminus {\mathbb{B}}(0,{\varepsilon}^{-{3/4}})$. If $\xi \in {\mathbb{R}}^2 \setminus {\mathbb{B}}(0,{\varepsilon}^{-{3/4}})$ then $|{\varepsilon}\xi + k| + ({\varepsilon}\lambda)^2 \geq ({\varepsilon}\lambda)^2 \geq {\varepsilon}^2$ since $\lambda \in [1,2]$. Therefore |\_[\^2 (0,\^[-1/2]{})]{} d| \_[|| \^[-[3/4]{}]{}]{} |() (-)| d\ C \_[|| \^[-[3/4]{}]{}]{} \^[-6]{} |W\_k|\_[C\^[3]{}]{} |W\_[-k]{}|\_[C\^[3]{}]{} dC |W\_k|\_[C\^[3]{}]{} |W\_[-k]{}|\_[C\^[3]{}]{}. We now assume that $\xi \in {\mathbb{B}}(0,{\varepsilon}^{-{3/4}})$, hence $|{\varepsilon}\xi + k| \geq |k|-{\varepsilon}^{1/4} \geq |k|/2$. This implies - = = O(\^[1/4]{}|k|\^[-3]{}). This implies \_[(0,\^[-1/2]{})]{} d= \_[(0,\^[-1/2]{})]{} d+ O(\^[[1/4]{}]{})\ = \_[\^2]{} d+ O(\^[1/4]{}) . 6\. Combine the results of steps 1-5 to get \^[-2]{} \_V() = \_[k 0]{} I\[W\_k,W\_[-k]{}\] + o(1) = \_[k 0]{} \_[\^2]{} () (-) d+ o(1)\ = \_[k 0]{} [W\_k, ]{}\_2 + o(1) = \_[\^2]{} \_0(x) dx+ o(1),    \_0(x) = \_[k0]{} . In the above we used that as $V$ is real-valued, $W_k=\overline{W_{-k}}$. This estimate completes the proof of the lemma. We finally turn to the proof of Theorem \[thm:1\]. By the discussion following the proof of Lemma \[lem:1d\], for ${\varepsilon}$ small enough the negative eigenvalues of $-\Delta+V$ are all in a fixed compact set and are the negative squares of the resonances of $V$ in $K_0$. We now work for $\lambda \in K_0$. For ${\varepsilon}$ small enough, ${{\operatorname{Id}}}+ L_V(\lambda)$ is invertible by Lemma \[lem:1d\]. Lemma \[lem:1c\] shows that the resonances of $V$ are the zeros of $1+\ln(\lambda) \varphi_V(\lambda)$. Since $\varphi_V(\lambda) = \frac{{\varepsilon}^2}{2\pi} \int_{{\mathbb{R}}^2} \Lambda_0 + o({\varepsilon}^2)$ uniformly on $K_0 = [0,M]$, if ${\varepsilon}$ is small enough then $$\lim_{\lambda \rightarrow 0} 1+ \ln(\lambda) \varphi_V(\lambda) = -\infty, \ \ \ \ 1+\ln(M) \varphi_V(M) \geq \dfrac{1}{2}.$$ We recall that $\varphi_V$ is real-valued on $K_0$. The intermediate value theorem shows that $1 + \ln(\lambda) \varphi_V(\lambda)$ has at least one zero $\lambda$ on $[0,M]$. It must satisfy $$\label{eq:1p} \ln(\lambda) = -\dfrac{1}{\varphi_V(\lambda)} = - \dfrac{2\pi}{{\varepsilon}^2 \int_{{\mathbb{R}}^2} \Lambda_0(x) dx + o({\varepsilon}^2)},$$ or equivalently, $$\lambda = \exp\left( - \dfrac{2\pi}{{\varepsilon}^2 \int_{{\mathbb{R}}^2} \Lambda_0(x) dx + o({\varepsilon}^2) } \right).$$ It remains to show that $\lambda$ is the unique resonance of $V$ in $K_0$ for ${\varepsilon}$ small enough. We argue as in [@Si Theorem 2.3]. If $\mu \neq \lambda$ is another resonance in $K_0$ then $1+\ln(\mu) \varphi_V(\mu) = 0$. Hence, $$\label{eq:1q} \left| \dfrac{1}{\ln(\mu)} - \dfrac{1}{\ln(\lambda)} \right| = |\varphi_V(\mu)-\varphi_V(\lambda)|.$$ In addition, for ${\varepsilon}$ small enough, $|\ln(\lambda)|^{-1} \geq c {\varepsilon}^2$ and $|\ln(\mu)|^{-1} \geq c {\varepsilon}^2$. Hence, a lower bound for the right hand side of is $$\left| \dfrac{1}{\ln(\mu)} - \dfrac{1}{\ln(\lambda)} \right| \geq \dfrac{1}{|\ln(\lambda)| |\ln(\mu)|} \left|\int_\lambda^\mu \dfrac{dt}{t}\right| \geq \dfrac{|\lambda-\mu|}{|\ln(\lambda)| |\ln(\mu)||\lambda + \mu|} \geq \dfrac{c^2{\varepsilon}^4 |\lambda-\mu|}{|\lambda + \mu|}.$$ An upper bound for the left hand side of is provided by Lemma \[lem:1f\]: $|\varphi_V(\lambda)-\varphi_V(\mu)|\leq C|\lambda\ln(\lambda)-\mu \ln(\mu)|$. Therefore, implies $$\dfrac{c^2 {\varepsilon}^4}{|\lambda + \mu|} \leq C \dfrac{|\lambda \ln(\lambda)-\mu \ln(\mu)|}{|\lambda-\mu|}.$$ Since the derivative of $t\ln(t)$ is $1 + \ln(t)$, and since $\lambda, \mu$ remain in a bounded set with $\ln(\lambda) = O({\varepsilon}^{-2})$ (and a similar bound for $\mu$), we deduce that ${\varepsilon}^4 \leq C{\varepsilon}^{-2}|\lambda + \mu|$. In addition, $|\ln(\lambda)|^{-1} \geq c {\varepsilon}^2$ implies $|\lambda| \leq e^{-c^{-1}{\varepsilon}^{-2}}$ (and a similar bound for $\mu$). Hence, the existence of $\mu$ implies $c^2{\varepsilon}^4 \leq 2C {\varepsilon}^{-2} e^{-c^{-1}{\varepsilon}^{-2}} = O({\varepsilon}^\infty)$. This is a contradiction. [2]{} D. I. Borisov, *Some singular perturbations of periodic operators.* Theoret. and Math. Phys. 151(2) (2007), 614-624. D. I. Borisov and R. R. Gadyl’shin, *On the spectrum of the Schrödinger operator with a rapidly oscillating compactly supported potential.* Theoret. and Math. Phys. 147 (2006), no. 1, 496-500. T. Christiansen, *Schrödinger operators with complex-valued potentials and no resonances.* Duke Math Jour. 133 (2006), 313-323. M. Dimassi, *Semi-classical asymptotics for Schrödinger operator with oscillating decaying potential.* To appear in Canad. Math. Bull. A. Drouot, *Scattering resonances for highly oscillatory potentials.* Preprint, [arXiv:1509.04198](http://arxiv.org/abs/1509.04198). V. Duchêne and N. Raymond, *Spectral asymptotics for the Schrödinger operator on the line with spreading and oscillating potentials.* Preprint, [arXiv:1609.01990](https://arxiv.org/abs/1609.01990). V. Duchêne, I. Vukićević and M. I. Weinstein, *Scattering and localization properties of highly oscillatory potentials.* Comm. Pure Appl. Math. 67 (2014), no. 1, 83-128. V. Duchêne and M. I. Weinstein, *Scattering, homogenization, and interface effects for oscillatory potentials with strong singularities.* Multiscale Model. Simul. 9 (2011), no. 3, 1017-1063. S. Dyatlov and M. Zworski, *Mathematical theory of scattering resonances.* Lecture notes available [online](http://math.mit.edu/~dyatlov/res/res_20160319.pdf). F. Gesztesy, Y. Latushkin, M. Mitrea and M. Zinchenko, *Nonselfadjoint operators, infinite determinants, and some applications.* Russ. J. Math. Phys. 12 (2005), no. 4, 443-471. B. Simon, *The bound state of weakly coupled Schrödinger operators in one and two dimensions.* Ann. Physics 97 (1976), no. 2, 279-288.
{ "pile_set_name": "ArXiv" }
--- author: - 'Natalia da Silva, Dianne Cook & Eun-Kyung Lee' bibliography: - 'ppfbiblio.bib' title: A Projection Pursuit Forest Algorithm for Supervised Classification --- Abstract {#abstract .unnumbered} ======== This paper presents a new ensemble learning method for classification problems called projection pursuit random forest (PPF). PPF uses the [[PPtree]{}]{} algorithm introduced in [@lee2013pptree]. In PPF, trees are constructed by splitting on linear combinations of randomly chosen variables. Projection pursuit is used to choose a projection of the variables that best separates the classes. Utilizing linear combinations of variables to separate classes takes the correlation between variables into account which allows PPF to outperform a traditional random forest when separations between groups occurs in combinations of variables. The method presented here can be used in multi-class problems and is implemented into an R [@RCore] package, [[PPforest]{}]{}, which is available on CRAN. Introduction ============ There are two main aspects of a random forest [@breiman2001random], bootstrap aggregation and [@breiman1996bagging; @breiman1996heuristics] random predictor selection [@amit1997shape; @ho1998random], that are broadly applicable to build ensemble classifiers from any basic method. Bagging stabilizes the variance and random predictor selection reduces correlation between trees in the forest. This paper presents the projection pursuit random forest (PPF), a new ensemble learning method for classification problems, built on combinations of predictors in the tree construction. PPF builds on the projection pursuit tree (PPtree) algorithm [@lee2013pptree], available in the R package [[PPtreeViz]{}]{} [@PPtreeViz] which fits a single multi-class tree to the data. Projection pursuit is used to find the linear combination of variables that best separates groups, and many different rules to make the actual split are provided. Trees that use linear combinations of predictors in a split are known in the literature as oblique trees [@kim2001; @brodley; @tan2005mml; @truong2009fast; @lee2013pptree]. All these algorithms use different approaches for finding linear combinations of predictors upon which to make a split. Some of the methods used for selecting the linear combination include random coefficient generation, linear discriminant analysis, and linear support vector machines. Theoretically, these could also be used as a base underlying PPF. For each split, a random sample of predictors is selected, then an optimal linear combination for separating the classes is computed by using a projection pursuit index. The algorithm is targeted for problems where classes can be separated by linear combinations of predictors, which define separating hyperplanes that are oblique to the axes rather than orthogonal to them. Additionally PPF accommodates class imbalance by using stratified bootstrap samples and variable importance measures are computed using the coefficients of the projections. PPF can be used for multi-class problems and is implemented into an R package, called [[PPforest]{}]{}. Only the LDA and PDA projection pursuit indexes are available in PPF. In the machine learning literature numerous work has been conducted on algorithms for building forests from oblique trees [@tan2006decision], [@menze2011oblique] and [@do2010classifying]. The performance is reported to be better than random forests, which is what we have determined with our algorithm also. A limitation of building on these approaches is the lack of readily available software. This paper is organized as follows. Section \[PPT\] explains the projection pursuit tree underlying PPF. Section \[PPFsec\] describes the PPF algorithm; diagnostics, including how to compute variable importance and the implementation details. Section \[perfsec\] evaluates the algorithm using a simulation study and performance on benchmark machine learning data in comparison with other methods. Section \[options\] discusses the choice of parameters, and compares the diagnostics relative to random forests. Section \[discpp1\] discusses possible extensions and future directions. Background on the projection pursuit tree =========================================  \[PPT\] The projection pursuit algorithm searches for a low dimensional projection that optimizes a continuous function which measures some aspect of interest; for PPF, this is class separation. [@friedman1973projection] coined the term “projection pursuit”, but the ideas existed earlier than this [@kruskal1969toward]. [@lee2005projection] developed an index, derived from the linear discriminant analysis, for finding projections that separate classes. Let $\mathbf{x_{gi}}$ be a $p$-dimensional data vector, $i$-th observation of the $g$-th class, $g = \{1,\ldots, G\}$, $G$ is the number of classes, $i = \{1,\ldots , n_g\}$, and $n_g$ is the number of observations in class $g$. The LDA index is defined as follows: $$I_{LDA}(A) = \left\{ \begin{array}{l l} 1-\frac{|A^T WA|}{|A^T(W+B)A|} & \text{for} |A^T(W+B)A|\neq 0\\ 0 & \text{for} |A^T(W+B)A|= 0 \end{array} \right.$$ where $B=\sum_{g=1}^G n_g(\bar{\mathbf{x}}_\mathbf{{g.}}-\bar{\mathbf{x}}_\mathbf{{..}})(\bar{\mathbf{x}}_\mathbf{{g.}}-\bar{\mathbf{x}}_\mathbf{{..}})^{T}$ is the between-group sums of squares, and $W=\sum_{g=1}^{G}\sum_{i=1}^{n_g}(\mathbf{x}_\mathbf{{ig}}-\bar{\mathbf{x}}_\mathbf{{g.}})(\mathbf{x}_\mathbf{{ig}}-\bar{\mathbf{x}}_\mathbf{{.g}})^T$ is the within-group sums of squares. If the LDA index value is high, there is a large difference between classes. A second index, PDA, was developed to address large $p$, small $n$ data [@lee2010projection]. The main idea used in construction of the index is that when $n\leq p$ or the variables are highly correlated, the maximum likelihood variance-covariance matrix estimator will be close to being singular, and this will affect the inverse calculation. The PDA index adjusts the variance-covariance matrix calculation, and is defined as follows: $$I_{PDA}(A,\lambda)=1-\frac{|A^T W_{PDA}A|}{|A^T (W_{PDA}+B) A|}$$ where A is an orthonormal projection onto a $k$-dimensional space and $\lambda \in [0,1)$ is a pre-determined parameter. $B$ is the between-class sums of squares and $W_{PDA}=\mbox{diag}(W)+(1-\lambda)\mbox{offdiag}(W)$. ![Comparison of decision boundaries for the [[rpart]{}]{} (left) and PPtree (right) algorithms on 2D simulated data. The partitions generated by PPtree algorithm are oblique to the axis, incorporating the association between the two variables.\[bounds\]](boundss-1){width="\maxwidth"} The PPtree algorithm uses a multi-step approach to fit a multi-class model by finding linear combinations to split on. Figure \[bounds\] compares the boundaries that would result from a classification tree fitted using the rpart algorithm [@therneau2010rpart] and the PPtree algorithm. Figure \[diagpp1\] illustrates the PPtree algorithm for three classes, and the algorithm steps are detailed below. Let $d_n =\{(\mathbf{x_i},y_i)\}_{i=1}^n$ be the data set where $\mathbf{x_i}$ is a p-dimensional vector of explanatory variables and $y_i\in \mathscr{G}$ ($\mathscr{G} =\{1,2,\ldots G\}$) represents class information with $i=1,\ldots n$. 1. Optimize a projection pursuit index to find an optimal one-dimensional projection, $\alpha^*$, for separating all classes in the current data yielding projected data $z = \alpha^*x$. 2. On the projected data, $z$, redefine the problem into a two class problem by comparing means, and assign a new label, either $g_1^*$ or $g_2^*$ to each observation, generating a new class variable $y_i^*$. The new groups $g_1^*$ and $g_2^*$ can contain more than one original class. 3. Find an optimal one-dimensional projection $\alpha^{**}$, using $\{(\mathbf{x_i},y_i^*)\}_{i=1}^n$ to separate the two class problem $g_1^*$ and $g_2^*$. The best separation of $g_1^*$ and $g_2^*$ is determined in this step providing the decision rule for the node, > if $\alpha^{**T}M_1< c$ then assign $g_1^*$ to the left node else assign $g_2^*$ to the right node, where $M_1$ is the mean of $g_1^*$. 4. For each group, all the previous steps are repeated until $g_1^*$ and $g_2^*$ have only one class from the original classes. The depth of PPtree is at most the number of classes. ![Illustration of the PPtree algorithm for $g=3$ classes. It is a dual pass algorithm for multiclass problems, for each split. It first finds the best separation and combines classes into two super-groups. It then searches again for the best separation between these two super-groups and splits on this. It proceeds sequentially on the subsets in the nodes, but only $g-1$ splits are allowed. \[diagpp1\]](diag2.pdf){width="1\linewidth"} Projection pursuit random forest {#PPFsec} ================================ This section provides the definition of PPF for classification and the algorithm. Diagnostics for the classifier are also defined. Definition ---------- Let the random vector of predictor variables $\mathbf{X}\in R^p$ and the output random variable $Y \in \mathscr{G}$, where $\mathscr{G}$ is a finite set such that $\mathscr{G}=\{1,2, \ldots, G\}$. The training sample is defined as $D_n=\{(\mathbf{X_1}, Y_1), \ldots (\mathbf{X}_n, Y_n)\}$ of i.i.d $\Re^p \times \mathscr{G}$ random variables $(p\geq 2)$. The objective is to build a classifier which predicts $y$ from $\mathbf{x}$ using $D_n$ given an ensemble of classifiers $h$. A projection pursuit classification random forest can be defined as a collection of randomized classification trees $\{h_n(\mathbf{x}, \Theta_m, D_n), m\geq 1\}$ where $\{\Theta_m\}$ are i.i.d. random vectors. $\Theta_m$ includes the two sources of randomness in the tree (random variable selection and random bootstrap sample), then $\Theta_m$ has information about which variables were selected in each partition and which cases were selected in the bootstrap sample. For each tree, $h_n$, a unique vote is collected based on the most popular class for the selected predictor variables. Equation \[rfesti\] defines the PPF estimator based on combining the trees. $$\begin{aligned} \label{rfesti} f_n(\mathbf{X}, D_n )&=& \operatorname*{arg\,max}_{g\in \mathscr{G}} \{E_{\Theta}(I[h_n(\mathbf{X}, \Theta, D_n)=g])\}\\ \nonumber &=& \operatorname*{arg\,max}_{g\in \mathscr{G}} P_{\Theta}(h_n(\mathbf{X}, \Theta, D_n)=g)\end{aligned}$$ $E_{\Theta}$ is the expectation wrt $\Theta$, conditionally on $\mathbf{X}$ and $D_n$. In practice, the PPF estimator is evaluated by generating $B$ random trees and take the average of the individual outcomes. This procedure is justified in a similar way to the original random forest defined by [@breiman2001random], and is based on the Law of the Large Numbers [@athreya2006measure]. Equation \[predfor\] describes the prediction of a new observation $\mathbf{x_0}$. $$\hat f_n(\mathbf{x_0})= \operatorname*{arg\,max}_{g\in \mathscr{G}} \sum_{k=1}^B I [ h_n (\mathbf{x_0}, \Theta_{bk} )= g] \label{predfor}$$ Algorithm ---------- 1. Let $n=\sum_{i=1}^G n_i$ the total number of cases in the training set $d_n=\{\mathbf{x_i}, y_i\}_{i=1}^n$. $B$ stratified bootstrap samples from $d_n$ are taken. Then for each class, independently and uniformly re-sample cases from $d_{ng}$ (training data set for group $g$) with size $n_g$ to create a stratified bootstrap data set $\{bk= b_{k1}, b_{k2}, \ldots b_{kg}\}$. 2. Use a bootstrap sample $bk$ to grow a PPtree $(h_n(\mathbf{x}, \Theta_{bk}))$ to the largest extent possible without pruning. (Note that the depth of the PPtree is at most $G-1$, where $G$ is the number of classes). 1. Start with all the cases in $b_k$ in the root node. 2. A simple random sample of $m$ predictor variables from the set of all the predictor variables $M$ is drawn, where $m<<M$. 3. Find the optimal one-dimensional projection $\alpha^*$ to separate all the classes in $b_k$. 4. If more than two class, then reduce the number of classes to two by comparing means, and assign new labels, $g_1^*$ and $g_2^*$ to each case (called the new response $y_i^*$ in $b_k$). 5. Find the optimal one-dimensional projection, $\alpha^{**}$, using the bootstrap data set with the relabeled response, $y^*$, to separate $g_1^*$ and $g_2^*$. The linear combination is computed by optimizing a projection pursuit index to get a projection of the variables that best separates the classes using the $m$ random selected variables. Two index options are available LDA or PDA. 6. Compute the decision boundary $c$. Eight different rules to define the cutoff value of each node can be used. All the rules are defined in [@lee2018pptreeviz]. 7. Keep $\alpha^{**}$ and $c$. 8. Separate the data into two groups using the new labels $g_1^*$ and $g_2^*$. 9. Repeat from (b) to (h) if $g_1^*$ or $g_2^*$ have more than two original classes. 3. Repeat 2 for $k = 1,\ldots B$. 4. The output is the ensemble of PPtrees, $\{h_n^{bk}\}_{k=1}^B$. Split values on the projected data can be computed by one of eight methods, which use the group means, or medians, sample size and variance or IQR weighting Figure \[diagppf\] has a diagram illustrating the PPforest algorithm. ![Illustration of the PPforest algorithm. It is effectively the same as a random forest algorithm except that the PPtree classifier is used on each bootstrap sample. \[diagppf\]](diagram.pdf){width="1\linewidth"} Implementation -------------- The initial code for PPforest was developed entirely in R. It was subsequently profiled using [[profvis]{}]{} [@profvis], and two code optimization strategies were employed: translate main functions into [[Rcpp]{}]{} [@eddelbuettel2011rcpp] and parallelization using [[plyr]{}]{}. The [[microbenchmark]{}]{} package was used to compare the speed before and after optimization. Figure \[ratiotim\] shows the performance before and after optimization. The decrease in speed is linear as the number of groups increases. The improvement is between 3- and 9-fold for this range of parameters. The machine used for this comparison was a MacBook Pro with a processor of 2.4 GHz Intel Core i7 with a memory of 8GB and 1867MHz LPDDR3. Parameters Values -------------------------- ------------------- $g=$ number of classes $(3, 3^2, 3^3)$ $n=$ obs. by class $(10^1, 10^2)$ $p=$ number of variables $(10^1, 10^2)$ $m=$ number of trees (50, 500) $cr=$ numbers of cores $(1, 2, 4)$ $v$ PPforest version (only R,  C code) : Optimization assessment simulation design \[respp1\] ![Computational performance for different sample sizes and number variables, with purely R code (green) and with C++ code (purple).\[ratiotim\]](resus-1){width="15cm" height="15cm"} PPF diagnostics --------------- The process of bagging and combining results from multiple trees produces numerous diagnostics which can provide a lot of insight into the class structure in high dimensions. Because ensemble methods are composed of many models fitted to subsets of the data, many statistics can be calculated to be analyzed as a separate data set. This provides the ability to understand how the model is working. The diagnostics of interest are the error rate, variable importance measure, vote matrix, and proximity matrix. ### Error rate Using the out-of-bag (oob) cases from bagged trees in the forest construction allows ongoing estimates of the generalization error for an ensemble of trees, described in [@breiman2001random]. Given a training data set $d_n$, $B$ bootstrap samples from $d_n$ are taken. For each bootstrap sample ($b= 1, 2, \ldots B$), a ``tree classifier $h_n(\mathbf{x}, \Theta_b)$ is constructed, and a majority vote is used to get the PPF predictor. The oob cases are used to get the error rate estimates. For each $\{\mathbf{x_i}, y_i\}$ in $d_n$, the votes are aggregated only for the classifiers $h_n(\mathbf{x}, \Theta_b)$ that do not contain $\{\mathbf{x_i}, y_i\}$. Hence, PPF is called the out-of-bag classifier, and the error rate for this classifier (out-of-bag error rate) is the estimate of the generalized error. The out-of-bag error rate is a measure for each model that is combined in the ensemble and is used to provide the overall error of the ensemble. ### Variable importance PPF calculates variable importance in two ways: (1) permuted importance using accuracy, and (2) importance based on projection coefficients on standardized variables. The permuted variable importance is comparable to the measure defined in the classical random forest algorithm. It is computed using the oob cases for the tree $k\;\;(B^{(k)})$ for each $X_j$ predictor variable. Then the permuted importance of the variable $X_j$ in the tree $k$ can be defined as: $$IMP^{(k)}(X_j) = \frac{\sum_{i \in B^{(k)} } I(y_i=\hat y_i^{(k)})-I(y_i=\hat y_{i,P_j}^{(k)})}{|B^{(k)}|}$$ where $\hat y_i^{(k)}$ is the predicted class for the observation $i$ in the tree $k$, and $y_{i,P_j}^{(k)}$ is the predicted class for the observation $i$ in the tree $k$ after permuting the values for variable $X_j$. The global permuted importance measure is the average importance over all the trees in the forest.This measure is based on comparing the accuracy of classifying oob observations using the true class with permuted (nonsense) class. For the second importance measure, the coefficients of each projection are examined. The magnitude of these values indicates importance if the variables have been standardized. The variable importance for a single tree is computed by a weighted sum of the absolute values of the coefficients across node, then the weights take the number of classes in each node into account($cl_{nd}$) [@lee2013pptree] . The importance of the variable $X_j$ in the PPtree $k$ can be defined as: $$IMP_{pptree}^{(k)}(X_j)=\sum_{nd = 1}^{nn}\frac{|\alpha_{nd}^{(k)}|}{cl_{nd} }$$ where $\alpha_{nd}^{(k)}$ is the projected coefficient for node $ns$ and variable $k$ and $nn$ the total number of node partitions in the tree $k$. The global variable importance in a PPforest then can be defined in different ways. The most intuitive are the average variable importance from each PPtree across all the trees in the forest. $$IMP_{ppforest1}(X_j)=\frac{\sum_{k=1}^K IMP_{pptree}^{(k)}(X_j)}{K}$$ Alternatively, a global importance measure is defined for the forest as a weighted mean of the absolute value of the projection coefficients across all nodes in every tree. The weights are based on the projection pursuit indexes in each node ($Ix_{nd}$), and 1-(OOB-error of each tree)($acc_k$). $$IMP_{ppforest2}(X_j)=\frac{\sum_{k=1}^K acc_k \sum_{nd = 1}^{nn}\frac{Ix_{nd}|\alpha_{nd}^{(k)}|}{nn }}{K}$$ ### Vote matrix An uncertainty measure for each observation, across models, is the proportion of times that a case is predicted to be in each class. If a case is always predicted to be the one class, there is no uncertainty about its group, and if this matches the true class then it is correctly labeled. Cases that are proportionately predicted to be multiple classes indicate difficult-to-classify observations. These cases may be important in that they might indicate special attention is needed in some neighborhoods of the data space, or more simply, could be errors in measurements in the data. ### Proximity matrix In a tree, each pair of observations can be in the same terminal node or not. Tallying this up across all trees in a forest gives the proximity matrix, an $n\times n$ matrix of the proportion of trees that the pair shares a terminal node. A proximity matrix can be considered to be a similarity matrix. This is typically used to do a follow-up cluster analysis to assess the strength of the class structure, and whether there are additional unlabeled clusters. ### Summary These diagnostics are used to assess model complexity; individual model contributions; variable importance and dimension reduction; and uncertainty in prediction associated with individual observations. Performance comparison {#perfsec} ====================== This section presents simulation results and a benchmark data study to examine the predictive performance of PPF in comparison to other methods. In the benchmark data study, PPF is compared with PPtree, CART and RF. The simulation results are designed to compare PPF with RF on data with linear projections defining class differences. Benchmark data study -------------------- The performance of PPF is compared with the classification methods, PPtree, CART and RF using 10 benchmark data sets taken from the UCI Machine Learning archive [@Lichman]. Table \[bench.tab\] presents summary information about the benchmark data, number of groups, cases, and predictors for each data set. The *imbalance* between groups is measured by the range of group size proportions and *correlation* is the average of all pairwise correlation coefficients among predictor variables. DF Cases Predictors Groups Imbalance Correlation ----------- ------- ------------ -------- ----------- ------------- crab 200 5 4 0.00 0.95 lymphoma 80 50 3 0.41 0.75 NCI60 61 30 8 0.07 0.56 parkinson 195 22 2 0.51 0.50 fishcatch 159 6 7 0.31 0.46 leukemia 72 40 3 0.40 0.44 olive 572 8 9 0.32 0.35 wine 178 13 3 0.13 0.30 image 2310 18 7 0.00 0.28 glass 214 9 6 0.31 0.23 : Summary of benchmark data. Imbalance and correlation indicating relative class sizes, and separations in combinations of variables.[]{data-label="bench.tab"} For each benchmark data set, $2/3$ of the observations are randomly chosen and used for training while the remaining $1/3$ are used as test data for computing predictive error. This procedure is repeated 200 times and the mean error rate is reported in Table \[respp1\]. In PPF, the number of variables selected in each node partition is a tuning parameter, the proportion of variables selected at each partition. Three different values were used (0.6, 0.9 and the RF default). The test error reported for PPF is the best from these. The results show that PPF has a better performance in the test data set than the other methods for the crab, fishcatch, leukemia, lymphoma, olive and wine data, while the RF test error is smaller for glass, image, NCI60 and parkinson data. ----------- ------- ---------- -------- ------- ------- ---------- -------- ------- Data CART PPforest PPtree RF CART PPforest PPtree RF crab 0.277 0.046 0.044 0.244 0.453 0.057 0.057 0.238 fishcatch 0.118 0.000 0.000 0.193 0.184 0.011 0.012 0.191 glass 0.237 0.306 0.331 0.240 0.330 0.390 0.403 0.224 image 0.069 0.079 0.067 0.024 0.082 0.083 0.073 0.024 leukemia 0.037 0.000 0.000 0.033 0.146 0.030 0.049 0.032 lymphoma 0.052 0.000 0.000 0.100 0.155 0.053 0.069 0.081 NCI60 0.503 0.019 0.000 0.458 0.676 0.388 0.423 0.376 olive 0.072 0.037 0.048 0.053 0.119 0.048 0.068 0.052 parkinson 0.081 0.112 0.175 0.107 0.159 0.171 0.229 0.101 wine 0.050 0.001 0.001 0.019 0.127 0.018 0.021 0.021 ----------- ------- ---------- -------- ------- ------- ---------- -------- ------- : Comparison of PPtree, CART, RF and PPF results with various data sets. The mean of training and test error rates from 200 re-samples is shown. (Order of rows is same as in Table ef[bench.tab]{}.) PPF performs favorably compared to the other methods. \[res\] Figures \[parallel\] displays the performance comparison graphically. Each line connects the errors for one data set. Even though RF outperforms PPF on almost half the data (Table \[respp1\]) PPF tends to have consistently low error. ![Benchmark data results shown graphically. PPF performs consistently well across most of the data sets. \[parallel\]](parbench-1){width="\maxwidth"} Boundary comparison with random forest -------------------------------------- To illustrate why and where PPF outperforms RF, results from a small simulation are shown. We expect PPF to outperform RF when the separation between classes is in linear combinations of variables. The simulated data is similar to the crab data. ![Boundaries in rotated trangle simulation, for RF and PPF, for different rotations (a-d), and (e) the average error over 20 repetitions for different angle, solid line is RF, and dashed is PPF. PPF beats RF uniformly in this type of data, and RF produces inferior boundaries. A little surprising that RF does worse with a small rotation.[]{data-label="trianglesim"}](triangle-sim-1){width="\maxwidth"} Each 2D simulated data set was rotated from 0 through 90$^o$, and 20 replications were conducted. Average (and standard deviation) of error was computed. Figure \[trianglesim\] shows the boundaries for two of the rotations generated by the RF and PPF models, and shows the summary of the errors by rotation angle. PPF uniformly outperforms RF in this scenario and produces better boundaries. Diagnostics comparison ======================  \[options\] The diagnostics computed by PPF (Section \[PPFsec\]) and RF are compared for the lymphona data, which helps to understand why and how PPF outperforms RF with this data. Variable importance ------------------- Figure \[globalimp\] illustrates how the variable importance differs, using the lymphoma data. PPF outperformed RF for this data. There are three groups, and it is a high-dimension, low sample size data set. With PPF, the PDA index is used, and the 60% of variables are available at each node. The number of trees used is the same as the RF default. Only the top ten most important variables are shown. There are some common on both lists and a some differences. Showing just the first two variables from each list is sufficient to illustrate the different type of boundaries induced by the classifiers. The two ways of computing importance in PPF do produce a different hierarchy of variables. With the global average importance, Gene35 and Gene50 are the top two, and these distinguish the small group FL best. With the global importance, Gene35 and Gene44 are featured, and together these find a big gap between DLBCL and the other two groups. PPF is utilizing the association between variables to classify groups, as would be expected. ![Comparison of importance measures for the lymphoma data, where PPF outperformed RF. Top row shows the top 10 variables by each method, with two ways of calculating with PPF. Bottom row shows the top two variables from each, which illustrates the difference between methods. PPF is detecting differences between groups when there is association between variables. Using the global average importance (left), Gene35 and Gene50 better distinguish group FL. Using the global importance, Gene35 and Gene44 find a big gap between group DLBCL and the other two. \[globalimp\]](globalimpoe-1){width="\maxwidth"} ![Comparison of importance measures for the lymphoma data, where PPF outperformed RF. Top row shows the top 10 variables by each method, with two ways of calculating with PPF. Bottom row shows the top two variables from each, which illustrates the difference between methods. PPF is detecting differences between groups when there is association between variables. Using the global average importance (left), Gene35 and Gene50 better distinguish group FL. Using the global importance, Gene35 and Gene44 find a big gap between group DLBCL and the other two. \[globalimp\]](globalimpoevars-1){width="\maxwidth"} ![Comparison of the out-of-bag vote matrix for the three groups of the lymphoma data, returned by PPF and RF: (top) ternary plot, (bottom) side-by-side jittered dotplots. This illustrates the difference between methods. PPF votes more decidedly for most cases, than RF. Especially this is true for the DLCBL class, where all but one are almost always predicted to the true class.[]{data-label="voteplots"}](side-1){width="\maxwidth"} Vote matrix ----------- Figure \[voteplots\] shows the vote matrices returned by PPF and RF for three classes of the lymphona data. It is represented in two ways: as a ternary plot and as a side-by-side jittered dotplot. The vote matrix has three columns corresponding to the proportion of times the case was predicted to be class B-CLL, DLBCL or FL, and thus is constrained to lie in a 2D triangle in 3D space. A ternary diagram is created using a helmert transformation of the vote matrix to capture the 2D subspace. The way to read it is: points near the vertex are clearly predicted to be one class, points along an edge are confused between two classes, and points in the middle are confused between the three classes. PPF provides more distinct classification of observations than RF, because the points are more concentrated in the vertices, and along one edge. The side-by-side jittered dotplot is an alternative representation that readily can be used for any number of classes. The proportion each case is classified to a group is diplayed vertically along a horizontal axis representing the categorical class variable. Points are jittered a little horizontally to better see the distribution of proportions, and colour represents the true class. Points concentrated at the top part indicate cases that are clearly grouped into a class, and if the colour matches the true class then these are correct classifications. The message is similar to the ternary diagram: DLBCL is much more clearly distinguished by PPF, and FL is actually distinguishable from B-CLL by PPF but confused by RF. Proximity --------- ![Examining similarity between cases, using pairwise plots of multidimensional scaling on the proximity matrix from PPF and RF fits of the lymphoma data. It can be seen that most cases are grouped closely with their class in PPF while in RF FL and B-CLL are mixed. \[prox1\]](mds-1){width="\maxwidth"} Figure \[prox1\] shows multidimensional scaling plots of the proximity matrix produced by PPF and RF classification of the lymphoma data. PPF provides the cleaner proximities. This means that more frequently observations from the same class reside in the same terminal node of the trees making up the PPF, than those of RF. Parameter selection =================== The primary parameters for PPF are mostly the same as those for RF: number of trees, and number of variables used in each node partition, with the addition of $\lambda$ when PDA is used as the index. Figure \[parameters\] (left) shows the effect of proportion of variables for the benchmark data comparison. The average error over 200 training/test splits is shown. For all data sets error is lower when the more variables are used. Most converge to low error rate when half the variables are included. The right plot compares the number of trees needed to optimise the OOB error for both PPF and RF on the lymphona data. Both need around 100 trees to produce best performance. ![Illustrating model tuning using error rate reduction. The average error rate plotted against proportion of variables in all the benchmark data is shown at left. The error rate tends to be better with more variables, but it does vary substantially by data set. OOB error is plotted against number of trees (right) on the lymphoma data for both PPF and RF. PPF has the consistently lower error, but both would indicate about 100 trees is sufficient to get the best results. \[parameters\] ](errorrate-1){width="\maxwidth"} Discussion {#discpp1} ========== This article has presented a new ensemble method (PPF) for classification problems, that is built on an oblique tree classifier (PPtree). PPF takes the correlation between variables into account. The forest algorithm enhances the single tree performance, adding diagnostics to assess variable importance, confusion of observations between groups and proximity of observations. It is best for medium sized data sets, both in number of observations and variables. The benchmark data study showed that PPF predictive performance is always at least as good, or better, than CART and PPtree, and often better than RF. Simulation results show that PPF performs better than RF when the classes are separated by a linear combination of variables and when the correlation between variables increases. The variable importance diagnostic shows that different variables are combined to create the classification using a PPF than RF. There are several directions where the work could be extended. The two projection pursuit indexes, LDA and PDA, can be readily supplemented by other indices. An example would be to add a regression index for a continuous response. Another direction is to adapt the PPtree algorithm to allow more than $g-1$ splits. This constraint protects the single tree model from overfitting. There is some protection against this with the bagging, and we expect it would enable deeper non-linear boundaries to be constructed by PPF. Lastly, because the accuracy of each tree is collected, automatic pruning of poor performing trees is a possibility. Acknowledgements ================ The code for PPF are implemented in an R [@RCore] package, [[PPforest]{}]{}, which is available on CRAN, with development versions at <https://github.com/natydasilva/PPforest>. This paper was written with the R packages knitr ([@xie:2015]), ggplot2 ([@hadley:2009]) and dplyr ([@dplyr]).
{ "pile_set_name": "ArXiv" }
--- abstract: '[According to the shock jump conditions, the total fluid’s mass, momentum, and energy should be conserved in the entire simulation box. We perform the dynamical Monte Carlo simulations with the multiple scattering law for energy analysis. The various energy functions of time are obtained by monitoring the total particles’ mass, momentum, and energy in the simulation box. In conclusion, the energy analysis indicates that the smaller energy losses in the prescribed scattering law are, the harder the energy spectrum produced is.]{}' author: - Xin Wang - Yihua Yan title: The energy analysis for the monte carlo simulations of a diffusive shock --- Introduction {#Introduction} ============ The gradual solar energetic particles with a power-law energy spectrum are generally thought to be accelerated by the first-Fermi acceleration mechanism at the interplanetary shocks (IPs) [@axford77; @krymsky77; @bell78; @bo78]. It is well known that the diffusive shock accelerated the particles efficiently by the accelerated particles scattering off the instability of Alfven waves which are generated by the accelerated particles themselves [@gosling81; @cvm90; @Lee86; @bkv03; @plm06]. The diffusive shock acceleration (DSA) is so efficient that the back-reaction of the accelerated particles on the shock dynamics cannot be neglected. So the theoretical challenge is how to efficiently model the full shock dynamics [@bv06; @ckvj10; @zank00; @li03; @Lee05]. To efficiently model the shock dynamics and the particles’ acceleration, there are largely three basic approaches: stationary Monte Carlo simulations, fully numerical simulations, and semi-analytic solutions. In the stationary Monte Carlo simulations, the full particle population with a prescribed scattering law is calculated based on the particle-in-cell (PIC) techniques [@ebj96; @veb06]. In the fully numerical simulations, a time-dependent diffusion-convection equation for the CR transport is solved with coupled gas dynamics conservation laws [@kj07; @ZA10]. In the semi-analytic approach, the stationary or quasi-stationary diffusion-convection equations coupled to the gas dynamical equations are solved [@bac07; @mdv00]. Since the velocity distribution of suprathermal particles in the Maxwellian tail is not isotropic in the shock frame, the diffusion-convection equation cannot directly follow the injection from the non-diffusive thermal pool into the diffusive CR population. So considering both the quasi-stationary analytic models and the time-dependent numerical models, the injection of particles into the acceleration mechanism is based on an assumption of the transparency function for thermal leakage [@bgv05; @kj07; @vainio07]. Thus, the dynamical Monte Carlo simulations based on the PIC techniques are expected to model the shock dynamics time-dependently and also can eliminate the suspicion arising from the assumption of the injection [@knerr96; @wang11]. In plasma simulation (PIC and hybrid), there is no distinction between thermal and non-thermal particles, hence particle injection is intrinsically defined by the prescribed scattering properties, and so it is not controlled with a free parameter [@ckvj10]. Actually, Wang [[*et al. *]{}]{}[@wang11] have extended the dynamical Monte Carlo models with an anisotropic scattering law. Unlike the previous isotropic prescribed scattering law, the Gaussian scattering angular distribution is used as the complete prescribed scattering law. According to the extended prescribed scattering law, we obtained a series of similar energy spectrums with little difference in terms of the power-law tail. However, it is not clear how such a prescribed scattering law can affect the particles’ diffusion and the shock dynamics evolution. To probe these problems, we expect to diagnose the energy losses in the simulations by monitoring all of the behaviors of the simulated particles. In the time-dependent Monte Carlo models coupled with a Gaussian angular scattering law, the results show that the total energy spectral index and the compression ratio are both effected by the prescribed scattering law. Specifically, the total energy spectral index is an increasing function of the dispersion of the scattering angular distribution, but the subshock’s energy spectral index is a digressive function of the dispersion of the scattering angular distribution [@wang11]. In the dynamical Monte Carlo simulations, one find that it is the only way for the particles to escape upstream via free escaped boundary (FEB). With the same size of the FEB which limited the maximum energy of accelerated particles, we find that different Gaussian scattering angular distribution generate different maximum energy particles through the scattering process at the same simulation time. In an effort to verify the efficiency of the energy transfer from the thermal to superthermal and the effect of the shock dynamics evolution on the shock structures, we perform a dynamical Monte Carlo code on Matlab with Gaussian scattering angular distribution by monitoring the particles’ mass, momentum and energy as functions of time. Our Gaussian scattering angular distribution algorithm consists of four cases involving four specific standard deviation values. This aim is to know if the various particle’s loss functions are dependent on the prescribed scattering law and the various kinds of losses can directly determine the total compression ratio and final energy spectral index with the same timescale of the shock evolutions and the same size of FEB. In Section \[sec-model\], the basic simulation method is introduced with respect to the Gaussian scattering angular distributions for monitoring the particles’ mass, momentum and energy as function of time in each case. In Section \[sec-results\], we present the shock simulation results and the energy analysis for all cases with four assumptions of scattering angle distributions. Section \[sec-summary\] includes a summary and the conclusions. Method {#sec-model} ====== The Monte Carlo model is a general model, although it is considerably expensive computationally, and it is important in many applications to include the dynamical effects of nonlinear DSA in simulations. Since the prescribed scattering law in Monte Carlo model instead of the field calculation in hybrid simulations [@Giacalone04; @wo96], we assume that particle scatters elastically according to a Gaussian distribution in the local plasma frame and that the mean free path (mfp) is proportional to the gyroradius (i.e., $\lambda \propto r_{g}$), where $r_{g}=pc/(qB)$, and its value is proportional to its momentum. Under the prescribed scattering law, the injection is correlated with those “thermal" particles which manage to diffuse the shock front for obtaining additional energy gains and become superthermal particles [@ebg05]. However, the basic theoretical limit to the accelerated particle’s energy arises from the accelerated particle’s Larmor radius, which must be smaller than the dimensions of the acceleration region by at least $v_{s}/c$ [@hillas84], where $v_{s}$ is the shock velocity. The limitation to the maximum energy of accelerated particles due to the large Larmor radius would be ameliorated if the scattering angular distribution is varied. In these simulations, the size of the FEB is set as a finite length scale matched with the maximum diffusive length scale. Here, we further investigate the possibility that the accelerated particles scatter off the background magnetic field in the acceleration region with not only an isotropic scattering angular distribution, but also with an anisotropic scattering angular distribution. Actually, this anisotropic distribution would probably produce an important effect on simulation results. With the same size of FEB, the scattering law applied isotropic distribution or anisotropic distribution would produce different maximum energy particles. So an anisotropic scattering law in the theory of the CR-diffusion is also needed [@bell04]. ![The entire evolutional velocity profiles in four cases. The dashed line denotes the FEB position in each plot. The precursor is located in the area between the downstream region and the upstream region in each case.[]{data-label="fig:shock"}](fig1a "fig:"){width="2.5in"} ![The entire evolutional velocity profiles in four cases. The dashed line denotes the FEB position in each plot. The precursor is located in the area between the downstream region and the upstream region in each case.[]{data-label="fig:shock"}](fig1b "fig:"){width="2.5in"}\ ![The entire evolutional velocity profiles in four cases. The dashed line denotes the FEB position in each plot. The precursor is located in the area between the downstream region and the upstream region in each case.[]{data-label="fig:shock"}](fig1c "fig:"){width="2.5in"} ![The entire evolutional velocity profiles in four cases. The dashed line denotes the FEB position in each plot. The precursor is located in the area between the downstream region and the upstream region in each case.[]{data-label="fig:shock"}](fig1d "fig:"){width="2.5in"} The particle-in-cell techniques are applied in these dynamical Monte Carlo simulations. The simulation box is divided up into some number of cells and the field momentum is calculated at the center of each cell [@forslund85; @Spitkovsky08; @nps08]. The total size of a one-dimensional simulation box is set as $X_{max}$ and it is divided into the number of grids $N_{max}$. Upstream bulk speed $U_{0}$ with an initial Maxwellian thermal velocity $V_{L}$ in their local frame and the inflow in a “pre-inflow box" (PIB) are both moving along one-dimensional simulation box. The parallel magnetic field $B_{0}$ is along the $\hat{x}$ axis direction in the simulation box. A free escaped boundary (FEB) with a finite size in front of the shock position is used to decouple the escaped particles from the system as long as the accelerated particles beyond the position of the FEB. The simulation box is a dynamical mixture of three regions: upstream, precursor and downstream. The bulk fluid speed in upstream is $U=U_{0}$, the bulk fluid speed in downstream is $U=0$, and the bulk fluid speed with a gradient of velocity in the precursor region is $U_{0}>U>0$. Because of the prescribed scattering law instead of the particle’s movement in the fields, the injected particles from the thermalized downstream into the precursor for diffusive processes are controlled by the free elastic scatter mechanism. To obtain the information of the total particles in different regions at any time, we build a database for recording the velocities, positions, and time of the all particles, as well as the index and the bulk speeds of the grids. The scattering angle distributions are presented by Gaussian distribution function with a standard deviation $\sigma$, and an average value $\mu$ involving four cases: (1) Case A: $\sigma=\pi$/4, $\mu=0$. (2) Case B: $\sigma=\pi$/2, $\mu=0$. (3) Case C: $\sigma=\pi$, $\mu=0$. (4) Case D: $\sigma=\infty$, $\mu=0$. These presented simulations are all based on a one-dimensional simulation box and the specific parameters are based on [@wang11]. Results & analysis {#sec-results} ================== We present the entire shock evolution with the velocity profiles of the time sequences in each case as shown in Figure \[fig:shock\]. The total velocity profiles are divided into three parts with respect to the shock front and the FEB locations. The upstream bulk speed $U_{0}$ dynamically slows down by passing though the precursor region, and its value decreases to zero in the downstream region (i.e. $U_{d}=0$). The precursor explicitly shows a different slope of the bulk velocity and different final FEB locations in different cases. The present velocity profiles are similar to the density profiles in the previous simulations by @wang11, and the different prescribed scattering law leads to the different shock structures. Items Case A Case B Case C Case D ----------------- ----------- ----------- ----------- ----------- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- $M_{loss}$ 1037 338 182 127 $P_{loss}$ 0.0352 0.0189 0.0123 0.0084 $E_{loss}$ 0.7468 0.5861 0.4397 0.3014 $E_{feb}$ 0.8393 0.5881 0.5310 0.5022 $E_{in}$ 1.5861 1.1742 0.9707 0.8036 $E_{tot}$ 3.3534 3.4056 3.3574 3.4026 $R_{in}$ 38.25% 25.67% 19.98% 14.80% $R_{loss}$ 22.27% 17.21% 13.10% 8.86% $r_{tot} $ 8.1642 6.3532 5.6753 5.0909 $r_{sub}$ 2.0975 3.0234 3.1998 3.9444 $\Gamma_{tot} $ 0.7094 0.7802 0.8208 0.8667 $\Gamma_{sub} $ 1.8668 1.2413 1.1819 1.0094 $v_{sh}$ -0.0419 -0.0560 -0.0642 -0.0733 $v_{sub}$ 0.0805 0.1484 0.1613 0.2159 $V_{Lmax}$ 11.4115 14.2978 17.2347 20.5286 $E_{peak}$ 0.1650keV 0.1723keV 0.1986keV 0.2870keV $E_{max}$ 1.23MeV 1.93MeV 2.80MeV 4.01MeV [ The units of mass, momentum, and energy are normal to the proton mass $m_{p}$, initial momentum $P_{0}$, and initial energy $E_{0}$, respectively. The last two rows are shown as scaled values.]{} The various losses of the particles and the calculated results of the shocks at the end of the simulation for the four cases are listed in Table \[tab:res\]. The initial box energy is $E_{0}$. The subshock’s compression ratio $r_{sub}$ and the total compression ratio $r_{tot}$ are calculated from the fine velocity structures in the shock frame in each case, respectively. The total energy spectral index $\Gamma_{tot} $ and the subshock’s energy spectral index $\Gamma_{sub} $ are deduced from the corresponding total compression ratio $r_{tot}$ and the subshock’s compression ratio $r_{sub}$ in each case. The $M_{loss}$, $P_{loss}$ and $E_{loss}$ are the mass, momentum, and energy losses which are produced by the escaped particles via the FEB, respectively. We have monitored the mass, momentum and energy of the total particles in each time step in each case. Figure \[fig:eng\] shows all the types of energy functions with respect to time. The total energy $E_{tot}$ is the energy summation over the time in the entire simulation box at any instant in time. The box energy $E_{box}$ presents the actual energy in the simulation box at any instant in time. The supplement energy $E_{PIB}$ is the summation of the amount of energy from the pre-inflow box (at the left boundary of the simulation box) which enters into the simulation box with a constant flux. The $E_{feb}$ presents a summation of the amount of energy held in the precursor region. The $E_{out}$ indicates a summation of the amount of energy which escapes via the FEB. Clearly, the total energy in the simulation at any instant in time is not equal to the actual energy in the box at any instant in time in each plot. ![Various energy values vs. time (all normalized to the initial total energy $E_{0}$ in the simulation box) in each case. All quantities are calculated in the box frame.[]{data-label="fig:eng"}](fig2a "fig:"){width="2.5in"} ![Various energy values vs. time (all normalized to the initial total energy $E_{0}$ in the simulation box) in each case. All quantities are calculated in the box frame.[]{data-label="fig:eng"}](fig2b "fig:"){width="2.5in"}\ ![Various energy values vs. time (all normalized to the initial total energy $E_{0}$ in the simulation box) in each case. All quantities are calculated in the box frame.[]{data-label="fig:eng"}](fig2c "fig:"){width="2.5in"} ![Various energy values vs. time (all normalized to the initial total energy $E_{0}$ in the simulation box) in each case. All quantities are calculated in the box frame.[]{data-label="fig:eng"}](fig2d "fig:"){width="2.5in"} It can be shown that the incoming particles (upstream) decrease their energy (as viewed in the box frame) as they scatter in the shock precursor region. If each incoming particle loses a small amount of energy as it first encounters the shock, this would produce a constant linear divergence between the curves for $E_{box}$ and $E_{tot}$. Actually, the cases in these simulations produce the non-linear divergence between the curves for $E_{box}$ and $E_{tot}$, consistent with Figure \[fig:eng\]. Such behavior is evident from individual particle’s trajectories. Physically, all the kinds of the losses occur in the precursor region owing to the “back reaction" of the accelerated ions and the escaped particles via FEB. Various energy functions obviously show that the case applying the anisotropic scattering law produces a higher energy loss, while the case applying the isotropic scattering law produces a lower energy loss. Consequently, we consider that the prescribed scattering law dominate the energy losses. Energy losses ------------- With monitoring each particle in the grid of the simulation box in any increment of the time, the escaped particles’ mass, momentum, and energy loss functions with the time are obtained and shown in Figure \[fig:loss\]. Among these energy functions, the inverse flow function is obtained from the injected particles from the thermalized downstream to the precursor. ![ The four plots denote the mass losses, momentum losses, energy losses and the inverse energy, respectively. The solid line, dashed line, dash-dotted line and the dotted line represent the cases A, B, C and D in the first three plots. In the last plot, the dashed lines marked with the signal of dot, plus, and cross, and the solid line represent the cases A, B, C and D, respectively. The units are normal to the proton mass $M_{p}$, initial total momentum $P_{0}$ and initial total energy $E_{0}$, respectively.[]{data-label="fig:loss"}](fig3a "fig:"){width="2.5in"} ![ The four plots denote the mass losses, momentum losses, energy losses and the inverse energy, respectively. The solid line, dashed line, dash-dotted line and the dotted line represent the cases A, B, C and D in the first three plots. In the last plot, the dashed lines marked with the signal of dot, plus, and cross, and the solid line represent the cases A, B, C and D, respectively. The units are normal to the proton mass $M_{p}$, initial total momentum $P_{0}$ and initial total energy $E_{0}$, respectively.[]{data-label="fig:loss"}](fig3b "fig:"){width="2.5in"}\ ![ The four plots denote the mass losses, momentum losses, energy losses and the inverse energy, respectively. The solid line, dashed line, dash-dotted line and the dotted line represent the cases A, B, C and D in the first three plots. In the last plot, the dashed lines marked with the signal of dot, plus, and cross, and the solid line represent the cases A, B, C and D, respectively. The units are normal to the proton mass $M_{p}$, initial total momentum $P_{0}$ and initial total energy $E_{0}$, respectively.[]{data-label="fig:loss"}](fig3c "fig:"){width="2.5in"} ![ The four plots denote the mass losses, momentum losses, energy losses and the inverse energy, respectively. The solid line, dashed line, dash-dotted line and the dotted line represent the cases A, B, C and D in the first three plots. In the last plot, the dashed lines marked with the signal of dot, plus, and cross, and the solid line represent the cases A, B, C and D, respectively. The units are normal to the proton mass $M_{p}$, initial total momentum $P_{0}$ and initial total energy $E_{0}$, respectively.[]{data-label="fig:loss"}](fig3d "fig:"){width="2.5in"} Since the FEB is limited to be in front of the shock position with the same size in each case, once a single accelerated particle moves backward to the shock beyond the position of the FEB, we exclude this particle from the total system as the loss term in the mass, momentum, and the energy conservation equations. According to the Rankine-Hugoniot (RH) relationships, the compression ratio of the nonrelativistic shock with a large Mach number is not allowed to be larger than the standard value of four [@Pelletier01]. Owing to the energy losses which inevitably exist in the simulations, the calculated results show a decreasing values of the loss of the mass, momentum, and the energy from the cases A, B, and C to D, respectively. Simultaneously, the inverse energy injected from downstream to precursor is also a decreasing values of $(E_{in})_{A}=1.5861$, $(E_{in})_{B}=1.1742$, $(E_{in})_{C}=0.9707$, and $(E_{in})_{D}=0.8036$ from the cases A, B, and C to D, respectively. The accurate energy losses are the values of $(E_{loss})_{A}=0.7468$, $(E_{loss})_{B}=0.5861$, $(E_{loss})_{C}=0.4397$, and $(E_{loss})_{D}=0.3041$ in each case. The inverse energy is the summation of the energy loss $E_{loss}$ and the net energy $E_{feb}$ in the precursor (i.e. $E_{in}=E_{loss}+E_{feb}$). So it is no wonder that the total shock ratios are all larger than four because of the existence of energy losses in all cases. Therefore, the difference of the energy losses and the inverse energy can directly affect all aspects of the simulation shocks. Subshock structure {#subsec:SS} ------------------ Figure \[fig:rsub\] shows the subshock structure in each case at the end of the simulation time. The specific structure in each plot consists of three main parts: precursor, subshock and downstream. The smooth precursor with a larger scale is between the FEB and the subshock’s position $X_{sub}$, where the bulk velocity gradually decreases from the upstream bulk speed $U_{0}$ to $v_{sub1}$. And the size of the precursor is almost equal to the diffusive length of the maximum energy particle accelerated by the diffusion process. The sharp subshock with a shorter scale only spans three-grid-lengthsb involving a deep deflection of the bulk speed abruptly decreasing from $v_{sub1}$ to $v_{sub2}$, where the scale of the three-grid-length is largely equal to the mean free path of the averaged thermal particles in the thermalized downstream. The value of the subshock’s velocity can be defined by the difference value of the two boundaries of the subshock. $$\label{eq:vsub} v_{sub}= |v_{sub1}-v_{sub2}|.$$ With the upstream bulk speed slowing down from $U_{0}$ to zero, the size of the downstream region is increasing with its constant shock velocity $v_{sh}$ in each case, and the bulk speed is $U=0$ owing to the dissipation processes which characterize the downstream. The gas subshock is just an ordinary discontinuous classical shock embedded in the total shock with a comparably larger scale [@bere99]. ![Final subshock fine structures in the four cases. The vertical solid and dashed lines indicate the positions of the shock front and subshock in each plot, respectively. The horizontal solid, dashed, dash-dotted and dotted lines show the values of the shock velocity $v_{sh}$, subshock velocity $v_{sub2}$, subshock velocity $v_{sub1}$ and initial bulk velocity $U_{0}$, respectively. Three vertical blocks in each plot represent the three deflections of velocity: precursor region, subshock region and downstream region. All values of the velocity are based on the box frame.[]{data-label="fig:rsub"}](fig4a "fig:"){width="2.5in"} ![Final subshock fine structures in the four cases. The vertical solid and dashed lines indicate the positions of the shock front and subshock in each plot, respectively. The horizontal solid, dashed, dash-dotted and dotted lines show the values of the shock velocity $v_{sh}$, subshock velocity $v_{sub2}$, subshock velocity $v_{sub1}$ and initial bulk velocity $U_{0}$, respectively. Three vertical blocks in each plot represent the three deflections of velocity: precursor region, subshock region and downstream region. All values of the velocity are based on the box frame.[]{data-label="fig:rsub"}](fig4b "fig:"){width="2.5in"}\ ![Final subshock fine structures in the four cases. The vertical solid and dashed lines indicate the positions of the shock front and subshock in each plot, respectively. The horizontal solid, dashed, dash-dotted and dotted lines show the values of the shock velocity $v_{sh}$, subshock velocity $v_{sub2}$, subshock velocity $v_{sub1}$ and initial bulk velocity $U_{0}$, respectively. Three vertical blocks in each plot represent the three deflections of velocity: precursor region, subshock region and downstream region. All values of the velocity are based on the box frame.[]{data-label="fig:rsub"}](fig4c "fig:"){width="2.5in"} ![Final subshock fine structures in the four cases. The vertical solid and dashed lines indicate the positions of the shock front and subshock in each plot, respectively. The horizontal solid, dashed, dash-dotted and dotted lines show the values of the shock velocity $v_{sh}$, subshock velocity $v_{sub2}$, subshock velocity $v_{sub1}$ and initial bulk velocity $U_{0}$, respectively. Three vertical blocks in each plot represent the three deflections of velocity: precursor region, subshock region and downstream region. All values of the velocity are based on the box frame.[]{data-label="fig:rsub"}](fig4d "fig:"){width="2.5in"} According to the subtle shock structures, the shock compression ratio can be cataloged into two classes: one class presents the entire shock named the total compression ratio $r_{tot}$ and the other class characterizes the subshock named the subshock’s compression ratio $r_{sub}$. The values of the two kinds of compression ratios can be reduced by the following formulas, respectively. $$\label{eq:rtot} r_{tot}=u_{1}/u_{2},\\ $$ $$\label{eq:rsub} r_{sub}= (v_{sub} +|v_{sh}|) /|v_{sh}|,$$ where $u_{1}=u_{0}+|v_{sh}|$ , $ u_{2}=|v_{sh}|$, and $u_{1}(u_{2})$ is the upstream (downstream) velocity in the shock frame, $v_{sub}$ is the subshock’s velocity determined by Equation \[eq:vsub\], and the shock velocity $v_{sh}$ at the end of the simulation is decided by the following: $$v_{sh}=(X_{max}-X_{sh})/T_{max},\label{eq:vsh}$$ where $X_{max}$ is the total length of the simulation box, $T_{max}$ is the total simulation time, and $X_{sh}$ is the position of the shock at the end of the simulation. The specific calculated results are shown in Table \[tab:res\]. The values of the subshock’s compression ratios are $(r_{sub})_{A}$=2.0975 , $(r_{sub})_{B}$=3.0234, $(r_{sub})_{C}$=3.1998 and $(r_{sub})_{D}$=3.9444 corresponding to the cases A, B, C and D, respectively. The total shock compression ratios with the values of $(r_{tot})_{A}$=8.1642, $(r_{tot})_{B}$=6.3532, $(r_{tot})_{C}$=5.6753, and $(r_{tot})_{D}$=5.0909 also correspond to the cases A, B, C and D, respectively. In comparison, the total shock compression ratios are all larger than the standard value four and the subshock’s compression ratios are all lower than the standard value four. Additionally, the value of the total shock compression ratio decreases from Cases A, B, and C to D, while the subshock’s compression ratio increases from Cases A, B, and C to D. These differences are naturally attributed to the different fine subshock structures. Maximum energy -------------- ![The individual particles with their thermal velocity in the local frame vs their positions with respect to time in each plot. The shaded area indicates the shock front, the solid line in the bottom plane denotes the position of the FEB in each case, respectively. Some irregular curves trace the individual particle’s trajectories near the shock front with time. The maximum energy of accelerated particles in each case is marked with the value of the velocity, respectively.[]{data-label="fig:acc"}](fig5a "fig:"){width="2.5in"} ![The individual particles with their thermal velocity in the local frame vs their positions with respect to time in each plot. The shaded area indicates the shock front, the solid line in the bottom plane denotes the position of the FEB in each case, respectively. Some irregular curves trace the individual particle’s trajectories near the shock front with time. The maximum energy of accelerated particles in each case is marked with the value of the velocity, respectively.[]{data-label="fig:acc"}](fig5b "fig:"){width="2.5in"}\ ![The individual particles with their thermal velocity in the local frame vs their positions with respect to time in each plot. The shaded area indicates the shock front, the solid line in the bottom plane denotes the position of the FEB in each case, respectively. Some irregular curves trace the individual particle’s trajectories near the shock front with time. The maximum energy of accelerated particles in each case is marked with the value of the velocity, respectively.[]{data-label="fig:acc"}](fig5c "fig:"){width="2.5in"} ![The individual particles with their thermal velocity in the local frame vs their positions with respect to time in each plot. The shaded area indicates the shock front, the solid line in the bottom plane denotes the position of the FEB in each case, respectively. Some irregular curves trace the individual particle’s trajectories near the shock front with time. The maximum energy of accelerated particles in each case is marked with the value of the velocity, respectively.[]{data-label="fig:acc"}](fig5d "fig:"){width="2.5in"} We select some individual particles from the phase-space-time database recording the all particles’ information. The trajectories of the selected particles are shown in Figure \[fig:acc\]. One of these trajectories clearly shows the route of the maximum energy of accelerated particle which undergoes the multiple collisions with the shock front in each case. The maximum value of the energy is apparently different in each case with increasing values of $(VL_{max})_{A}=11.4115$, $(VL_{max})_{B}=14.2978$, $(VL_{max})_{C}=17.2347$, and $(VL_{max})_{D}=20.5286$ from the Cases A, B, and C to D, respectively. And those particles with a value higher than the cutoff energy are unavailable owing to their escaping from the FEB. The statistical data show the number of escaped particles in each case decreases with the number of particles $(n_{esc})_{A}=1037$, $(n_{esc})_{B}=338$, $(n_{esc})_{C}=182$, and $(n_{esc})_{D}=127$ from Cases A, B, and C to D, correspondingly. Except for the maximum energy of the particle in each case, the other particles show that some of them obtained finite energy accelerations from the multiple crossings with the shock and some of them do not have additional energy gains owing to their lack of probability for crossing back into the precursor due to their small diffusive length scale. The statistical data also exhibit that the inverse energy injected from the downstream back to upstream is characterized by a decreasing reflux rate of $(R_{in})_{A}=38.25\%$, $(R_{in})_{B}=25.67\%$, $(R_{in})_{C}=19.98\%$, and $(R_{in})_{D}=14.80\%$ in corresponding Cases A, B, C, and D. With the decrease of the inverse energy from Cases A, B, and C to D, the corresponding energy losses are also reduced at the rate of $(R_{loss})_{A}=22.27\%$, $(R_{loss})_{B}=17.21\%$, $(R_{loss})_{C}=13.10\%$, and $(R_{loss})_{D}=8.86\%$ in each case, respectively. Although the maximum energy of the accelerated particles should be identical because of the limitation of the same size of the FEB in four cases, the cutoff energy values are still modified by the existence of the energy losses in the different cases applied with the different prescribed Gaussian scattering laws. Energy spectrum {#subsec:spectrum} --------------- ![The two plots present the final energy spectrums on the downstream and the precursor region, respectively. The thick solid line with a narrow peak at $E = $1.3105keV in each plot represents the same initial Maxwell energy distributions in each case. The solid, dashed, dash-dotted and dotted extended curves with the “power-law" tail present the energy spectral distributions corresponding to Cases A, B, C and D, respectively. All these energy spectrum distributions are plotted in the same shock frame. []{data-label="fig:spec"}](fig6a "fig:"){width="2.5in"} ![The two plots present the final energy spectrums on the downstream and the precursor region, respectively. The thick solid line with a narrow peak at $E = $1.3105keV in each plot represents the same initial Maxwell energy distributions in each case. The solid, dashed, dash-dotted and dotted extended curves with the “power-law" tail present the energy spectral distributions corresponding to Cases A, B, C and D, respectively. All these energy spectrum distributions are plotted in the same shock frame. []{data-label="fig:spec"}](fig6b "fig:"){width="2.5in"} The energy spectrums with the “power-law“ tail are calculated in the shock frame from the downstream region and the precursor region at the end of simulation, respectively. The same initial Maxwellian distribution in each case is shown in each plot in Figure \[fig:spec\]. As shown in Figure \[fig:spec\], the calculated energy spectrums indicate that the four extended curves in the downstream region with an increasing value of the central energy peak $(E_{peak})_{A}=0.1650keV$, $(E_{peak})_{B}=0.1723keV$, $(E_{peak})_{C}=0.1986keV$, and $(E_{peak})_{D}=0.2870keV$ characterize the Maxwellian distributions in the ”heated-downstream“ from Cases A ,B, and C to D, respectively. The value of the total energy spectral index $(\Gamma_{tot})_{A}=0.7094$, $(\Gamma_{tot})_{B}=0.7802$, $(\Gamma_{tot})_{C}=0.8208$, and $(\Gamma_{tot})_{D}=0.8667$ in each case indicates the Maxwellian distribution in the ”heated-downstream" with a decreasing deviation to the “power-law" distribution from Cases A, B, and C to D, correspondingly. But the value of the subshock’s energy spectral index $(\Gamma_{sub})_{A}=1.8668$, $(\Gamma_{sub})_{B}=1.2413$, $(\Gamma_{sub})_{C}=1.1819$, and $(\Gamma_{sub})_{D}=1.0094$ present the energy spectrum distribution with a “power-law" tail in each case implying there is an increasing rigidity of the spectrum from the Cases A, B, and C to D, respectively. The cutoff energy at the “power-law" tail in the energy spectrum is given with an increasing value of $(E_{max})_{A} $=1.23 MeV, $(E_{max})_{B}$=1.93 MeV, $(E_{max})_{C}$=2.80 MeV and $(E_{max})_{D}$=4.01 MeV from the Cases A, B, C and D, respectively. In the precursor region, the final energy spectrum is divided into two very different parts in each case. The part in the range from the low energy to the central peak shows an irregular fluctuation in each case. The irregular fluctuation indicates that the cold upstream fluid slows down and becomes the “thermal fluid" by the nonlinear “back reaction" processes. And the other part in the range beyond the central peak energy shows a smooth “power-law" tail in each case. ![Two plots show the correlation of the compression ratio vs the energy losses and the correlation of the energy spectral index vs the inverse energy, respectively. The triangles represent the total compression ratios and the total energy spectral index of the all cases in each plot. The circles indicate the subshock’s compression ratios and the subshock’s energy spectral index of all cases in each plot. []{data-label="fig:ratio-index"}](fig7a "fig:"){width="2.5in"} ![Two plots show the correlation of the compression ratio vs the energy losses and the correlation of the energy spectral index vs the inverse energy, respectively. The triangles represent the total compression ratios and the total energy spectral index of the all cases in each plot. The circles indicate the subshock’s compression ratios and the subshock’s energy spectral index of all cases in each plot. []{data-label="fig:ratio-index"}](fig7b "fig:"){width="2.5in"} As shown in Figure \[fig:ratio-index\], the two kinds of shock compression ratios are both apparently dependent on the energy losses with respect to these four presented simulations. As viewed from Cases A, B, and C to D, the total compression ratio is a decreasing function of the energy losses and each value is larger than the standard value four, the subshock’s compression ratio is an increasing function of the energy losses and each value is lower than four. However, both the total compression ratio and the subshock’s compression ratio approximate the standard value of four as the energy loss decreases. According to the DSA theory, if the energy loss is limited to be the minimum, the simulation models based on the computer will more closely fit the realistic physical situation. Additionally, the energy spectral index is also fairly dependent on the inverse energy from the thermalized downstream region into the precursor region. Summary and conclusions {#sec-summary} ======================= In summary, we performed the dynamical Monte Carlo simulations using the Gaussian scattering angular distributions based on the Matlab platform by monitoring the particle’s mass, momentum and energy at any instant in time. The specific mass, momentum and energy loss functions with respect to time are presented. A series of analyses of the particle losses are obtained in the four cases. We successfully examine the relationship between the shock compression ratio and the energy losses, as well as verify the consistency of the energy spectral index with the inverse energy injected from the downstream to precursor region in the simulation cases which are applied with the prescribed Gaussian scattering angular distributions. In conclusion, the relationship of the shock compression ratio with the energy losses via FEB verify that the energy spectral index is determined by the inverse energy function with time. In fact, these energy losses simultaneously depend on the assumption of the prescribed scattering law. As expected, the maximum energy of accelerated particles is limited by the size of the FEB according to the maximum mean free path in each case. However, there is still a fairly large difference between the maximum energy of the particle from the different cases with the same size of the FEB. We find that the total energy spectral index increases as the standard deviation value of the scattering angular distribution increases, but the subshock’s energy spectral index decreases as the standard deviation value of the scattering angular distribution increases. In these multiple scattering angular distribution simulations, the prescribed scattering law dominates the energy losses and the inverse energy. Consequently, the case of applying a prescribed law which leads to the minimum energy losses will produce a harder subshock’s energy spectrum than those in the cases with larger energy losses. These relationships will drive us to find a newly prescribed scattering law which leads to the minimum energy losses, making the shock compression ratio more closely approximate the standard value of four for a nonrelativistic shock with high Mach number in astrophysics. [99]{} Axford, W.I., Leer, E., & Skadron, G.,  1977  in Proc. 15th Int. Comsmic Ray Conf. (Plovdiv), 132 Bell, A. R.,  1978,  MNRAS, 182, 147. Bell, A. R.,  2004,  MNRAS, 353, 550. Berezhko, E. G. & Ellison, D. C. 1999, , 526, 385 , E. G., [Ksenofontov]{}, L. T. & [V[" o]{}lk]{} H. J.  2003, , 412, L11. , E. G., & [V[" o]{}lk]{} H. J.  2006, , 451, 981. Blandford, R. D., & Ostriker, J. ,P.  1978, , 221, L29. Blasi, P., Amato, E., & Caprioli, D.,  2007, , 375,1471 Blasi, P., Gabici, S., & Vannoni, G.,  2005, , 361,907 Cane, H. V., von Rosenvinge, T. T., & McGuire, R. E., 1990, , 95, 6575. Caprioli, D., Kang, H., Vladimirov, A. E. & Jones, T. W., 2010, , 407,1773 Ellison, D. C., Baring, M. G., & Jones, F. C. ,  1996, 473,1029 Ellison, D. C., Möbius, E. & Paschmann, G. Ellison, D. C., Blasi & Gabici ,  2005,  in Proc. 29th Int. Comsmic Ray Conf. (India). Forslund, D. W. , J. ,  2004, , 609, 452. Gosling, J.T., Asbridge, J.R., Bame, S.J., Feldman, W.C., Zwickl,R. D.,Paschmann, G.,Sckopke, N., & Hynds, R. J. 1981, , 866, 547 Hillas,  1984,  ARA&A, 22, 425. Jones, F. C., & Ellison, D. C.,  1991,  Space Science Reviews, 58, 259. Kang H. & Jones T.W. 2007,  A. Ph., 28, 232 Knerr, J. M., Jokipii, J. R. & Ellison, D. C.  1996, , 458, 641 Krymsky, G. F.,  1977,  Akad. Nauk SSSR Dokl., 243, 1306 Lee, M. A., & Ryan, J. M.,  1986,  , 303, 829 Lee, M. A.,  2005,  APJS, 158, 38 Malkov, M. A., Diamond, P. H., & V$\ddot{o}$lk, H. J.,  2000, , 533,171 , J. , [Pohl]{}, M. , [Stroman]{}, T. , & [Nishikawa]{}, K.-I., 2008, , 684, 1174. Li, G., Zank, G. P., & Rice, W. K. M.,  2003,  JGR 108,1082 Ostrowski, M. , 1988, , 233, 257 Pelletier, G.  2001, [[ Lecture Notes in Physics, ]{}]{}, 576, 58 , G. , [Lemoine]{}, M.  & [Marcowith]{}, A. ,  2006, , 453,181. , A. ,  2008, , 673, L39. Vainio, R., & Laitinen, T.,  2007, , 658, 622. Vladimirov, A., Ellison, D. C., & Bykov, A.,  2006, , 652,1246 Wang, X., & Yan, Y.,  2011, , 530, A92. , D. , & [Omidi]{}, N. ,  1996, , 101,17287–17304. Zank, G., Rice, W.K.M., & Wu, C. C.,  2000,  JGR, 105, 25079 Zirakashvili, V. N. & Aharonian, F. A.,   2010, , 708, 965
{ "pile_set_name": "ArXiv" }
--- author: - 'Verena Zuber [^1] and Korbinian Strimmer' bibliography: - 'preamble.bib' - 'econ.bib' - 'genome.bib' - 'stats.bib' - 'array.bib' - 'sysbio.bib' - 'misc.bib' - 'molevol.bib' - 'med.bib' - 'entropy.bib' date: '4 February 2009; last revision 16 July 2009' title: Gene ranking and biomarker discovery under correlation --- [^1]: Institute for Medical Informatics, Statistics and Epidemiology (IMISE), University of Leipzig, Härtelstr. 16–18, D-04107 Leipzig, Germany
{ "pile_set_name": "ArXiv" }
--- author: - | V. Bocci$^a$, D. Chao$^c$, G. Chiodi$^a$, R. Faccini$^{a,b}$, F. Ferroni$^{a,b}$, R. Lunadei$^a$, G. Martellotti$^{a}$, G. Penso$^{a,b}$, D. Pinci$^{a}$, L. Recchia$^{a}$\ INFN Sezione di Roma, Roma, Italy\ Dipartimento di Fisica, Sapienza Università di Roma, Roma, Italy\ California Institute of Technology, Pasadena, California, USA\ E-mail: title: 'Dependence of the energy resolution of a scintillating crystal on the readout integration time.' --- Introduction ============ Electromagnetic calorimeters are often composed of inorganic scintillating crystals viewed by photodetectors. The energy resolution attainable depends primarily on the number of optical photons emitted by a scintillating crystal for a given energy deposit. Usually only a fraction of these photons are collected by a photodetector. In the following we consider a photomultiplier (PM), where the collected optical photons are converted in photoelectrons with an efficiency characteristic of the photocathode and amplified up to the anode by the dynode system. Besides a possible non-linear crystal response [@dorenbos], the anode charge pulse is therefore proportional to the energy deposited in the crystal by the primary particle and the fluctuations of this charge determine the energy resolution of the detector. The main contribution to this resolution comes from the statistics of the photoelectrons. In addition a small but not negligible contribution comes from the fluctuations in the PM gain and in particular from the gain of the first dynode. This resolution is worsened if the readout electronics integrates only part of the total charge delivered by the anode of the PM. This eventuality can occur when scintillating crystals with a long decay time are used in high-rate experiments where short integration times are needed. In the present paper we will consider the possibility of using BGO crystals and PM’s for high rate experiments. In that case to limit the pile-up and the dead time, the PM pulses must be integrated over a time interval shorter than the scintillation time of the BGO ($\tau_{scint} = 300$ ns). The deposited energy is then deduced from the maximum amplitude of the integrated pulse, measured by a peak-sensitive circuit. With this procedure only a fraction of the total charge is measured and a faster response is obtained at the cost of a worse resolution. This effect was experimentally studied for different integration times and the results were compared with a Monte Carlo (MC) simulation. Experimental setup {#sec:exp} ================== The experimental setup is sketched in Fig. \[figsetup\]. Photons with an energy of 662 keV from a $^{137}$Cs radioactive source are detected by a $2 \times 2\times 18$ cm$^3$ BGO crystal, read at both ends by two EMI-9814B PMs. 3truemm One of the two photomutipliers (PM-1) was used to trigger the acquisition of the pulses from the other photomultiplier (PM-2). In order to get rid of the noise, the trigger-signal from PM-1 was amplified and shaped with a gated biased amplifier (ORTEC-444) having an integration time of 250 ns. The signal from PM-2 was processed by a filter amplifier (ORTEC-474) that has a variable gain and an integration[^1] time ($\tau_{int}$) that can be set to 20, 50, 100, 200 and 500 ns. The output signal of that module was then acquired by a Lecroy WavePro 7300A digital oscilloscope, having a bandwidth of 300 MHz and a sampling rate of 250 MS/s. Equivalent circuit {#equicircuit} ================== The ORTEC 474 integrating amplifier and its connection to the PM-2 anode can be represented by the equivalent circuit reported in Fig. \[amplifier\]. When a given energy is released in the crystal at the time $t=0$, the anode current[^2] delivered by PM-2 is: 2truemm $$I(t) = I_0 \; e^{-t/\tau_{scint}} \times u(t) \label{eqexp}$$ 3truemm where $\tau_{scint}$ is the decay time of the scintillator and $u(t)$ is the unit step function. The total charge per pulse released by the anode of PM-2 is . Using the standard Laplace transform method, the output signal $V(t)$ of the circuit on Fig. \[amplifier\] turns out to be: 3truemm $$V(t) = \frac{G \; I_0 \; R_{in}}{1-\alpha} \; (e^{-t/\tau_{scint}}-e^{-t/\tau_{int}}) \label{eqvt}$$ 3truemm where $R_{in}$ is the input resistance of the first buffer amplifier, $G$ is the overall gain, $\tau_{int} = RC$ is the integration time and $\alpha=\tau_{int} /\tau_{scint}$. For $N_{pe} \to \infty$, the maximum amplitude of the output signal is: 3truemm $$A_{N_{pe} \to \infty} = I_0 \; R_{in} \; \alpha^{\;\; \alpha/(1- \; \alpha)} \label{eqA}$$ 3 truemm which occurs at the time: -1 truemm $$T_{N_{pe} \to \infty} =\tau_{int} \; \frac{ ln \, \alpha}{\alpha-1} \label{eqT}$$ 4 truemm From Eq. \[eqvt\] the total charge ($Q_{out}$) of an output pulse is proportional to $Q$: 3truemm $$Q_{out}= \int_0^\infty \! \frac{V(t)}{R_{out}} \, dt = G \; I_0 \; R_{in}\;\tau_{scint}= G\;R_{in}\;Q \label{eqQout}$$ Data taking =========== About 15000 pulses were recorded for each of the five possible values $\tau_{int}=$ 20, 50, 100, 200 and 500 ns. Each pulse was sampled every 4 ns during 10 $\mu$s and the resulting 2500 istantaneous amplitudes were acquired. A thousand pulses are shown in Fig. \[PULSES\]a, while in Fig. \[PULSES\]b the average waveform is reported for the five $\tau_{int}$ values. To check the reproducibility of the results, four data-sets ([**A, B, C**]{} and [**D**]{}) were acquired under different conditions for all values of $\tau_{int}$. - [**A**]{} and [**B**]{} were taken under the same conditions (HV of PM-2 equal to 1800 V and gain of the ORTEC-474 set to 10) but in different days in order to test the reproducibility of the results; - [**C**]{} : HV of PM-2 equal to 1800 V and the ORTEC-474 gain set to 2; - [**D**]{} : HV of PM-2 equal to 1750 V and the ORTEC-474 gain set to 10. During the acquisition of the four data-sets the configuration on the trigger side of the set-up ( and ORTEC 444) was kept fixed. The trigger level of the oscilloscope was set sufficiently low to accept all the pulses from the 662 keV photons. The electronics noise was measured by acquiring data with a random triggers. The width of these noise spectra being 10 times smaller than that of the source signal, the contribution of the electronics noise to the energy resolution was neglected. Data analysis ============= 2truemm The acquired waveforms were analysed off-line. For each value of $\tau_{int}$ and for each pulse, the maximum amplitude ($A$), the peaking time ($T$) and the total charge ($Q_{out}$) were evaluated. Total charge fluctuations {#sectotch} ------------------------- 2truemm When the total charge of each pulse is measured, the energy resolution of the BGO crystal is mainly determined by the statistical fluctuations of the total number ($N_{pe}$) of the photoelectrons and by the fluctuations of the gain of the 12 PM dynodes ($g_1,...,g_{12}$). Assuming a Poisson distribution for $N_{pe}$ and considering that in general $g_1\;>g_2\;=\;g_3\;=\;...\;=\;g_{12}\;\equiv\;g$ the energy resolution[^3] is given by [@RTC; @PM1; @PM2; @PM3; @PM4]: 5truemm $$\frac{\sigma_E}{\overline{E}} = \frac{\sigma_Q}{\overline{Q}} = \frac{\sigma_{Q_{out}}}{\overline{Q}_{out}} = \frac{1}{\sqrt{\overline{N}_{pe}}} \; \sqrt{1+\frac{1}{\overline{g}_1} \left(\sum\limits_{i=0}^{k-1}\frac{1}{\overline{g}^{\;i}} \right)} \simeq \frac{1}{\sqrt{\overline{N}_{pe}}} \; \sqrt{1+\frac{1}{\overline{g}_1} \left(\frac{1}{1-1/g}\right)} \label{eqsigmaEsuEOrig}$$ 6truemm where $E$ is the measured energy and $k\;=\;12$ is the number of dynodes. Since $g$ is rather larger than 1, Eq.\[eqsigmaEsuEOrig\] was approximated as: $$\frac{\sigma_E}{\overline{E}} = \frac{\sigma_Q}{\overline{Q}} = \frac{\sigma_{Q_{out}}}{\overline{Q}_{out}} \simeq \frac{1}{\sqrt{\overline{N}_{pe}}} \; \sqrt{1+\frac{1}{\overline{g}_1}} \label{eqsigmaEsuE}$$ that is equivalent to take into account only the contribution of the fluctuations of the first dynode. In Fig. \[figsigmaQsuQ\]a a typical experimental spectrum of $Q_{out}$ is shown. Taking into account the HV of PM-2, the characteristics of the 9814B tube [@ET] and of the BeCu dynodes [@RTC] the average gain of the first dynode was assumed to be $\overline{g}_1\simeq 6$.\ -6truemm The energy resolution becomes: 1 truemm $$\frac{\sigma_E}{\overline{E}} = \frac{\sigma_{Q_{out}}}{\overline{Q}_{out}} = \frac{1.08}{\sqrt{\overline{N}_{pe}}} \label{eqNpe}$$ 1 truemm In Fig. \[figsigmaQsuQ\]b the experimantal values of $\sigma_{Q_{out}}/\overline{Q}_{out}$ are reported for the five values of $\tau_{int}$ and for the four sets of data taking. They show, as expected, a flat behaviour with respect to $\tau_{int}$. A constant fit to these data allows to determine from Eq. \[eqNpe\] the average number of photoelectrons $\overline{N}_{pe} = 130^{+13}_{-11}$. 5 truemm Maximum amplitude fluctuations {#secmaxamp} ------------------------------ 3truemm When the measurement of the total charge $Q$ takes too long, the energy deposited in the crystal can be inferred from the maximum amplitude $A$ of the integrated signal. In Fig. \[figsigmaAsuA\]a a spectrum of $A$ obtained in the present test with $\tau_{int} = 100$ ns is shown. The fluctuations ($\sigma_A$) on $A$ are obtained by a gaussian fit to the data. In Fig. \[figsigmaAsuA\]b the experimental resolution $\sigma_A/\overline{A}$ is reported[^4] as a function of $\tau_{int}$. At large values of $\tau_{int}$ and in particular for $\tau_{int} \gg \tau_{scint}$, $\sigma_A/\overline{A}$ tends to the value of $\sigma_{Q_{out}}/\overline{Q}_{out}$, while at smaller values of $\tau_{int}$ the resolution worsen. To clarify the dependence of $\sigma A / A$ on $\tau_{int}$, a naive Poissonian model based on an extension of Eq. \[eqNpe\] was adopted. According to this model the resolution is given by: $$\frac{\sigma_E}{\overline{E}}= \frac{\sigma_A}{\overline{A}} = \frac{1.08}{\sqrt{\overline{n}_{pe}}} \label{eqnpe}$$ where $n_{pe}$ is the number of photoelectrons which contribute, for each event, to its maximum amplitude $A$, i.e. those emitted before the peaking time $T$ of that event. The average value $\overline{n}_{pe}$ was approximated to the fraction $F$ of the average total number of photoelectron emitted before the experimentally measured values of $\overline{T}$: $$\begin{aligned} {\overline{n}_{pe}}& = F \overline{N}_{pe} \\ F& = (1 - e^{-\overline{\vphantom{|^.}T}/\tau_{scint}}) \label{approxnpe}\end{aligned}$$ In Fig. \[T\] the measured dependence of $\overline{\vphantom{|^.}T}$ on $\tau_{int}$ is reported. In Fig. \[figsigmaAsuA\]b the predictions of Eq.’s \[eqnpe\] and \[approxnpe\] are compared with the experimental results. While for $\tau_{int} = 500$ ns the agreement is good, at lower values of $\tau_{int}$ the experimental resolution is better than predicted with the naive model. To understand this discrepancy, a detailed Monte Carlo simulation of the experimental situation was performed. The Monte Carlo simulation {#secMC} ========================== The formulae reported in Sec. \[equicircuit\] represent the response of an RC integrator excited by an exponentially decreasing current composed of a very large number of electrons, so that the charge quantization is washed out. In the situation we are considering, the average number of photoelectrons per pulse is relatively small so that the response of the device must be simulated with a MC. The input current $I(t)$ is described as a sum of delta functions, each one corresponding to an incoming photoelectron. Then the amplitude of the integrated pulse at a time $t$ turns out to be the sum of the contributions from all the photoelectrons emitted before that time[^5]: $$V(t) = R_{in} \; (\frac{q}{\tau_{int}}) \sum\limits_{i=1}^{n(\,t)}G_i \; e^{-(t-t_i)/\tau_{int}} \label{eqpulseshape}$$ where $q$ is the electron charge, $n(t)$ is the number of photoelectrons emitted before the time $t$, $t_i$ ($i=1,n$) is the emission time of the $i$-th electron ($0 < t_i < t$) and $G_i$ is the PM gain for the $i$-th electron. The probability distribution function of the $t_i$ is a decreasing exponential with a decay time equal to $\tau_{scint}$. For a fixed $\; t$, $n(t)$ follows a Poisson distribution with a mean $$\overline{n}(t) = \overline{N}_{pe} ( 1 - e^{-t/\tau_{scint}}) \label{mean}$$ It is worthwile to note that Eq. \[eqpulseshape\] represents a single pulse that can be used to measure the energy, only if there is an effective pile-up of the contributions of many photoelectrons belonging to the same detected particle. This occurs if the integration time $\tau_{int}$ is much larger than the average time interval between two consecutive photoelectrons ($\thickapprox \tau_{scint}/\overline{N}_{pe}$): $$K \equiv \overline{N}_{pe} \frac{\tau_{int}}{\tau_{scint}} \gg 1 \label{k}$$ while if $K \lesssim 1$ the energy released in the scintillator gives rise only to a series of single photoelectron pulses. In the present experiment $K$ ranges from 8.7 (at $\tau_{int}=20$ ns) to 217 (at $\tau_{int}=500$ ns). In the MC simulation all the aforementioned effects were taken into account. The MC was run with $10 \leq \overline{N}_{pe} \leq 10^4$ and for $\tau_{int}$ ranging from 10 ns to 1 $\mu$s. For each of the $\sim 100$ pairs of values $(\overline{N}_{pe},\tau_{int})$ about 10000 pulses were generated. For each simulated pulse the MC calculates its amplitude $V(t)$ every ns during an interval of 2 $\mu$s. The maximum amplitude $A$ and the time $T$ at which this maximum occurs were recorded for each pulse and the relative fluctuations $\sigma_A/\overline{A}$ were determined. Comparison with experimental data --------------------------------- For a comparison with the experimental data, the MC was run with $\overline{N}_{pe} = 130 $. The Poisson fluctuations of the gain $g$ of the first dynode, with a mean value $\overline{g} = 6$, were also taken into account. In Fig. \[MCexp\] a typical pulse generated by MC with $\tau_{int}=500$ ns is compared with an experimental pulse recorded with the same integration time. The agreement between the two shapes is quite good. In Fig. \[figsigmaAsuA\]b the dependence of $\sigma_A/\overline{A}$ and in Fig. \[T\] the mean value of the peaking time distribution ($\overline{T}$) on $\tau_{int}$, calculated with the MC, are compared with the experimental points. In both cases the agreement is quite good. This confirms that in the present experimental conditions the experimental resolution is better than predicted by the naive model based on a Poissonian statistcs. These checks give confidence in the MC simulation and allow to use it to predict, in the most general experimental situation, which is the energy resolution attainable with an integrator followed by a peak-sensitive electronics. 5 truemm Energy resolution in the general case ------------------------------------- The resolutions $\sigma_A/\overline{A}$, calculated as a function of $\alpha\;=\;\tau_{int}/\tau_{scint}$ and for different values of $\overline{N}_{pe}$, are reported in Fig. \[lookuptable\] which is therefore a general utility to evaluate the resolution attainable with a scintillator having a decay time $\tau_{scint}$, read by a photodetector followed by an integrator and a peak-sensitive electronics. To present these results in a general form the fluctuations on the photodetector gain have not been included because they depend on the particular type used. For a PM the effect of these fluctuations can be taken into account by multiplying the values of $\sigma_A/\overline{A}$ read on Fig. \[lookuptable\] by the corrective factor of Eq. \[eqsigmaEsuEOrig\]. From Fig. \[lookuptable\] it appears that to perform an integration with $\tau_{int} < \tau_{scint}$, at least 10 photoelectrons are needed. As already pointed out in the naive model, only the $n_{pe}$ photoelectrons emitted before the peaking time $T$ contribute, for each event, to the the maximum amplitude $A$. Assuming a Poisson distribution for $n(T)$ the relative fluctuations on $A$ is given by: $$\frac{\sigma_A}{\overline{A}} = \frac{1}{\sqrt{\overline{n}_{pe}}} \label{naive}$$ -2truemm Contrarily to the experimental situation, in the MC simulation $n_{pe}$ is a known quantity for each event, so that the simple model represented by Eq. \[naive\] can be tested. In Fig. \[figab\]a $\sigma_A / \overline{A}$, calculated with the MC, is reported as a function of $1/\sqrt{\overline{n}_{pe}}$, for different values of $\overline{N}_{pe}$ and $\alpha $. It appears that for $\overline{N}_{pe}\gtrsim 500$ Eq. \[naive\] is satisfied for any of the considered $\alpha $ values so that the statistics of $n_{pe}$ is Poissonian and the naive model is valid. For smaller values of $\overline{N}_{pe}$ and in the range where Eq. \[k\] is satisfied, the resolution is better than predicted by Eq. \[naive\] so that in that region the statistics is . The reason of this behaviour is clear from the curves reported in Fig. \[figab\]b: for a fixed integration time a positive (negative) variation of $N_{pe}$ with respect of its average value $\overline{N}_{pe}$ results in a negative (positive) variation of the corresponding peaking time, which partially compensates the variation on $N_{pe}$. This anticorrelation between $N_{pe}$ and $T$ is responsible for the sub-Poissonian fluctuations at the lower values of $\overline{N}_{pe}$. The results reported in Fig. \[lookuptable\] and Fig. \[figab\] are valid for any integration time and for any decay time of the scintillating light, when the readout electronics measures the maximum amplitude of the integrated pulse. Conclusions =========== The possibility of using a scintillating crystal with a slow decay time (like BGO) for an electromagnetic calorimeter in a high-rate experiment was investigated. In these experimental conditions a fast measurement of the energy deposited in the crystal is needed. This can be obtained, at the cost of a lower energy resolution, by integrating the output signal of the photodetector over a short time and by acquiring the maximum amplitude of the integrated signal. An experimental test and a Monte Carlo simulation show that the energy resolution comes from the statistics of the number of photoelectrons emitted before the peaking time of the integrated pulse. While for a large number of photoelectrons the statistics follows a Poisson distribution, at a lower number of photoelectrons the statistics becomes due to an anticorrelation between the fluctuations of the number of photoelectrons per pulse and the peaking time of that pulse. The results are reported in a general form which allows to evaluate the contribution of the photoelectron statistics to the resolution of a calorimeter equipped with a scintillating crystal read by a photomultiplier, followed by an integrator and a peak-sensitive electronics. [99]{} P. Dorenbos et al., *Non-proportionality in the scintillation response and the energy resolution obtainable with scintillation crystals*, *IEEE Trans. Nucl. Sci.* [**42**]{} (1995) 2190. *Les Photomultiplicateurs*, *RTC Radiotechnique-Compelec ed.*, Paris 1981. F.J. Lombard and F. Martin, *Statistics of Electron Multiplication*, *Rev. Sci. Instr.* [**32**]{} (1961) 200. http://dx.doi.org/10.1063/1.1717310 M. Brault and C. Gazier, *Etude des fluctuations d’amplitude des photomultiplicateurs*, *J. de Physique*, [**24**]{} (1963) 345. E. Gatti and V. Svelto, *Review of theories and experiments of resolving time with scintillation counters*, *Nucl. Instr. Meth.* [**43**]{} (1966) 248. S. Donati et al., *An equivalent circuit for the statistical behaviour of the scintillation counter*, *Nucl. Instr. Meth.* [**46**]{} (1967) 165. ET Enterprises, electron tubes. http://www.et-enterprises.com/photomultipliers. [^1]: The differentiation control of the 474 module was set in the $out$ position. This corresponds to a differentiation time of 0.2 ms which has a negligible effect on the output signals. [^2]: The formulas reported in this section hold for a very large number of photoelectrons per pulse, so that the statistical fluctuations are negligible. [^3]: Throughout this paper $\overline{x}$ and $\sigma_x$ indicate respectively the mean value and the r.m.s. of a Gaussian fit to the $x$ distribution. [^4]: The data from the four data sets A, B, C and D have been averaged. [^5]: It can be shown that when
{ "pile_set_name": "ArXiv" }
--- abstract: 'If $G$ is a finite Abelian group, define $s_{k}(G)$ to be the minimal $m$ such that a sequence of $m$ elements in $G$ always contains a $k$-element subsequence which sums to zero. Recently Bitz et al. proved that if $n = exp(G)$, then $s_{2n}(C_{n}^{r}) > \frac{n}{2}[\frac{5}{4}-O(n^{-\frac{3}{2}})]^{r}$ and $s_{k n}(C_{n}^{r}) > \frac{k n}{4} [1+\frac{1}{e k}-O(\frac{1}{n})]^{r}$ for $k > 2$. In this note, we sharpen their general bound by showing that $s_{k n}(C_{n}^{r}) > \frac{k n}{4} [1+\frac{(k-1)^{(k-1)}}{k^k}-O(\frac{1}{n})]^{r}$ for $k > 2$.' author: - 'Jesse Geneson (PSU)' title: 'Improved lower bound on generalized Erdos-Ginzburg-Ziv constants' --- Lower bound =========== The function $s_{k n}(G)$ is known as the $k^{th}$ generalized Erdos-Ginzburg-Ziv constant of $G$. The first result about these constants was proved in [@egz], and there has been a long history of results since then [@cl; @eg; @ga; @ha; @ku; @re; @ro], which were detailed in [@bgh]. Our improvement is the theorem below. \[main\] For $k > 2$, we have $s_{k n}(C_{n}^{r}) > \frac{k n}{4} [1+\frac{(k-1)^{(k-1)}}{k^k}-O(\frac{1}{n})]^{r}$. To prove this result, we use the bound of Sondow et al. [@so; @sz] for binomial coefficients of the form $\binom{k n}{n}$. Specifically they proved the bounds $\frac{1}{4(k-1)n}[\frac{k^k}{(k-1)^{(k-1)}}]^{n} < \binom{k n}{n} < [\frac{k^k}{(k-1)^{(k-1)}}]^{n}$ for $n$ a positive integer and $k \geq 2$ a real number. The proof of Theorem \[main\] is nearly identical to the proof in [@bgh], but just with this sharper bound on the binomial coefficients. As in [@bgh], define $N = \frac{k n A^{r}}{4}$ for $A$ to be chosen at the end of the proof. Randomly pick a sequence $X$ of $N$ vectors in $\left\{0,1\right\}^{r}$ by letting each coordinate be $1$ with probability $q$, and let $Z$ be the number of subsequences of length $k n$ in $X$ that sum to $0$. We show that $E[Z] < 1$ with $A = 1+\frac{(k-1)^{(k-1)}}{k^k}-O(\frac{1}{n})$. First we calculate the probability $Q$ that a given coordinate sums to $0$, which is the sum of the probabilities $P_{i n}$ that the coordinate sums to $i n$ for $0 \leq i \leq k$, which are equal to $P_{i n} = \binom{k n}{i n} q^{i n} (1-q)^{(k-i)n}$. We want the terms $P_{0}$ and $P_{n}$ to dominate, so we set $(1-q)^{k n} = \binom{k n}{n}(1-q)^{(k-1)n} q^{n}$. Combining the bounds of Sondow et al. with this equality, we obtain $\frac{1}{4(k-1)n}[\frac{k^k}{(k-1)^{(k-1)}}]^{n} (1-q)^{(k-1)n} q^{n} < (1-q)^{k n} < [\frac{k^k}{(k-1)^{(k-1)}}]^{n} (1-q)^{(k-1)n} q^{n}$. This implies that $(1-o(1))[\frac{k^k}{(k-1)^{(k-1)}}] q < 1-q < [\frac{k^k}{(k-1)^{(k-1)}}] q$, and thus that $\frac{1}{1+\frac{k^k}{(k-1)^{(k-1)}}} < q < \frac{1}{1+(1-o(1))\frac{k^k}{(k-1)^{(k-1)}}}$. $P_{0}$ and $P_{n}$ are approximately equal in this range, so that $Q < (k+1)(1-q)^{k n}$ and $E[Z] = \binom{N}{k n} Q^{r} < (\frac{4N}{k n})^{k n} (k+1)^{r} (1-q)^{k n r}$. We want $E[Z] < 1$, so $A < \frac{1}{(k+1)^{1/kn}(1-q)}$ will suffice, and thus we may let $A = 1+\frac{(k-1)^{(k-1)}}{k^k}-O(\frac{1}{n})$. [7]{} J. Bitz, C. Griffith, X. He. Exponential lower bounds on the generalized Erdos-Ginzburg-Ziv constant. https://arxiv.org/abs/1712.00861 E. Croot, V. Lev, and P. Pach, Progression-free sets in Zn4 are exponentially small, Annals of Mathematics 185 (2017), 331-337. J. Ellenberg and D. Gijswijt, On large subsets of Fnq with no three-term arithmetic progressions, Annals of Mathematics 185 (2016), 1-4. P. Erdos, A. Ginzburg, and A. Ziv, Theorem in the additive number theory, Bull. Research Council Israel 10 (1961), 41-43. W. Gao and A. Geroldinger, Zero-sum problems in finite abelian groups: A survey, Expositiones Mathematicae 24 (2006), 337-369. H. Harborth, Ein Extremalproblem fur Gitterpunkte, J. Reine Angew. Math. 262 (1973), 356-360. S. Kubertin, Zero-sums of length kq in Zdq, Acta Arithmetica 116 (2005), 145-152. C. Reiher, On Kemnitz’ conjecture concerning lattice-points in the plane, Ramanujan Journal 13 (2007), 333-337. L. Ronyai, On a Conjecture of Kemnitz, Combinatorica 20 (2000), 569-573. J. Sondow, Problem 11132, Amer. Math. Monthly 112, 180, 2005. J. Sondow and W. Zudilin, Euler’s Constant, q-Logarithms, and Formulas of Ramanujan and Gosper, Ramanujan J. 12, 225-244, 2006.
{ "pile_set_name": "ArXiv" }
--- abstract: 'It was recently suggested that the discrepancy between two methods of measuring the lifetime of the neutron may be a result of an unseen decay mode into a dark matter particle which is almost degenerate with the neutron. We explore the consequences of this for the properties of neutron stars, finding that their known properties are in conflict with the existence of such a particle.' address: - 'CSSM and ARC Centre of Excellence for Particle Physics at the Terascale, Department of Physics, University of Adelaide SA 5005 Australia' - 'IRFU-CEA, Université Paris-Saclay, F91191 Gif sur Yvette, France' - 'CSSM and ARC Centre of Excellence for Particle Physics at the Terascale, Department of Physics, University of Adelaide SA 5005 Australia' author: - 'T. F. Motta' - 'P. A. M. Guichon' - 'A. W. Thomas' title: Implications of Neutron Star Properties for the Existence of Light Dark Matter --- Dark matter, neutron star, equation of state of dense matter Introduction ============ For some time there has been a discrepancy of about 8 seconds ( $3.5 \sigma$)between two techniques used to determine the lifetime of a free neutron. Following earlier suggestions that this discrepancy might result from an oscillation to a mirror neutron [@Serebrov:2007gw], it was recently proposed by Fornal and Grinstein [@Fornal:2018eol] that it could rather be caused by a new decay mode of the neutron to an almost degenerate dark matter particle, which would not be visible in one of the measurments. Several authors have already placed further constraints on this mechanism. Czarnecki and collaborators [@Czarnecki:2018okw] noted a degree of tension between the current best value of the neutron axial charge and this explanation but could not rule it out. On the basis of a new experimental measurement at the Los Alamos ultra-cold neutron facility, Tang [*et al.*]{} [@Tang:2018eln] ruled out the decay mode $n \rightarrow DM + \gamma$, where $DM$ represents the hypothesised dark matter particle, however this leaves the decay to two dark matter particles as a viable possibility. More recently it has been suggested by Serebrov [*et al.*]{} [@Serebrov:2018mva] that a so-called reactor anti-neutrino anomaly might also be explained by the same mechanism. Here we examine the consequences of the existence of such a dark matter particle in a very different environment, namely a neutron star. If there were indeed a dark matter particle almost degenerate with the neutron, then as the density of nuclear matter increases and the chemical potential of the neutrons rises above the mass of the dark matter particle the neutrons must decay to restore chemical equilibrium. Since the composition of the nuclear matter in the core of a neutron star is dominated by neutrons with a chemical potential well above the neutron mass, in this new scenario one must expect that almost half of the energy density of matter in the core would now be dark. As we shall see this dramatically reduces the pressure for a given energy density, which in turn reduces the maximum mass of the neutron stars which can be formed. Indeed, our calculations suggest that this scenario is incompatible with the existence of neutron stars whose masses are already well established. Equations of state and solution of the TOV equations ==================================================== In order to calculate the equation of state (EoS) of dense matter, which is required as input to the Tolman-Oppenheimer-Volkov (TOV) equations [@Tolman:1939jz; @Oppenheimer:1939ne] used to compute the mass and radius of a given neutron star, we use the quark-meson coupling (QMC) model [@RikovskaStone:2006ta; @Guichon:1987jp; @Guichon:1995ue]. This model yields a relativistic EoS for hadrons starting from the self-consistent solution for the structure of hadrons moving in relativistic mean-fields corresponding to the $\sigma , \omega$ and $\rho$ mesons. It naturally yields three-body forces between hadrons with no additional parameters [@Guichon:2004xg] and as a consequence can support neutron stars with masses in the region of two M$_\odot$, [*even when*]{} hyperons are included [@RikovskaStone:2006ta], while producing realistic hypernuclear binding energies [@Guichon:2008zz; @Saito:2005rv]. ![Comparison of the equation of state (EoS) for nuclear matter in $\beta$-equilibrium, with and without the inclusion of a dark matter particle degenerate with the neutron. Note the dramatic softening of the EoS. []{data-label="fig:EoS"}](eos2A) On the other hand, for the present application many of these features are not needed, as so much of the energy density is carried by dark matter that the maximum stellar mass is already reached before hyperons can enter the EoS, so it enough to consider matter consisting of neutrons, protons, electrons and muons in $\beta$-equilibrium, in chemical equilibrium with dark matter. In this case, treating the DM as degenerate with the neutron (as a difference of order 1 MeV will make no meaningful difference), the chemical equilibrium equations are simply $$\mu_n = \mu_p + \mu_e \,\,\, , \,\,\, \mu_n=\mu_DM \, , \label{eq:chemical}$$ where $\mu_{n(p)}$ are the proton and neutron chemical potentials, $\mu_e$ is the electron chemical potential and we also include the muon if it is energetically allowed. ![Comparison of the mass versus radius curves resulting from the solution of the TOV equations for the case where only nucleons are allowed and where dark matter is included. The reduction in the maximum possible mass is dramatic, from around 2.2 M$_\odot$ to just 0.7 M$_\odot$. []{data-label="fig:massvsR"}](massaraioA) ![Illustration of the species fractions appearing in matter in $\beta$ and chemical equilibrium when dark matter is included, as a function of baryon density. Clearly dark matter begins to dominate beyond 2 $\rho_0$. []{data-label="fig:species"}](speciesfraction2) To calculate the EoS we use the model QMC700, which is fully explained in Ref. [@RikovskaStone:2006ta]. In the case where only nucleons are included, like most other relativistic mean field treatments this yields a maximum neutron star mass around 2.2 M$_\odot$. The expression for the pressure is $$P = \sum_i \mu_i n_i - \epsilon \, , \label{eq:pressure}$$ where $n_i$ are number densities of the species present, $\mu_i$ their chemical potentials and $\epsilon$ the total energy density. From Eq. (\[eq:pressure\]) it is obvious that at the same energy density the lower neutron chemical potential resulting when DM is included will result in a lower pressure. While the sign of the effect is obvious, the size of the reduction in practice is remarkable, as we see in Fig. \[fig:EoS\]. Such a reduction in pressure for a given energy density might be expected to lead to stars with a lower maximum mass but the extent of the reduction is very large. In particular, as illustrated in Fig. \[fig:massvsR\], the maximum mass of a neutron star decreases from around 2.2 M$_\odot$ to as low as 0.7 M$_\odot$ once DM is allowed. In part this can be understood by the argument presented earlier, below Eq. (\[eq:pressure\]). However, the effect is significantly enhanced by the change in composition shown in Fig. \[fig:species\], where we see that the composition of the matter in the core of the star is dominated by DM at relatively low baryon density. Concluding remarks ================== In concluding, it seems worthwhile to put our results in some perspective. Many authors have investigated the consequences for neutron stars if they capture dark matter of various kinds [@PerezGarcia:2011hh; @Perez-Garcia:2014dra; @Li:2012ii; @PerezGarcia:2010ap]. The situation here is very different. Under the scenario proposed by Fornal and Grinstein the lifetime of a neutron which is forbidden by energy conservation and the Pauli principle to $\beta$-decay will be of the order of days. Thus in a supernova explosion a proto-neutron star would form and then decay over a period of days and weeks, as the neutrons high in the Fermi sea are converted to dark matter. Most of the neutron stars which have been observed have masses around 1.4-.15 M$_\odot$ [@Kiziltan:2013oja], while the maximum mass neutron stars that have been discovered have masses around 2 M$_\odot$ [@Demorest:2010bx; @Antoniadis:2013pzd], corresponding to a central density of around 5-6 times nuclear matter density [@RikovskaStone:2006ta; @Whittenbury:2013wma]. The latter corresponds to a total baryon number of order $3.2 \times 10^{57}$. It seems reasonable that the almost degenerate dark matter particles formed when the neutrons decay would remain trapped by gravity while their lighter decay partners escape, in which case the total number of dark matter particles plus nucleons after the decay process would be approximately equal to the total number of nucleons before decay. Such a star is well beyond the maximum mass consistent with stability against collapse and would become a black hole. Indeed, the number of dark matter particles plus nucleons in the dark matter star with the maximum mass is just $9.2 \times 10^{56}$. This corresponds to a star, before decay, of mass around 0.7 M$_\odot$. We have explored the consequences for the maximum mass of a neutron star of the proposed explanation for the discrepancy between modern measurements of the lifetime of the neutron involving a decay mode to a dark matter particle which is almost degenerate with the neutron. The dark matter is assumed to be non-interacting. As the central density of the star increases it is unavoidable that the fraction of particles that are dark begins to dominate, with the consequence that for a given energy density the pressure is dramatically lowered. As a consequence, when the TOV equations are solved the pressure is unable to sustain stars as massive as those found without dark matter. Indeed the maximum mass allowed is reduced to just 0.7 M$_\odot$. We have verified that alternate parameter sets with or without the inclusion of hyperons make no significant difference. The effect of this hypothesised dark matter particle is so dramatic that the conclusion that one cannot generate stable stars with the masses observed in Nature, typically around 1.4 M$_\odot$ but as large as 2 M$_\odot$, is model independent. As a result we are led to conclude that this explanation of the discrepancy in neutron lifetimes cannot be correct and the proposed dark matter particle does not exist. Acknowledgements {#acknowledgements .unnumbered} ================ This work was supported by the University of Adelaide and by the Australian Research Council through the ARC Centre of Excellence for Particle Physics at the Terascale (CE110001104) and Discovery Project DP150103164. References {#references .unnumbered} ========== [55]{} B. Fornal and B. Grinstein, arXiv:1801.01124 \[hep-ph\]. A. P. Serebrov [*et al.*]{}, Phys. Lett. B [**663**]{} (2008) 181 doi:10.1016/j.physletb.2008.04.014 \[arXiv:0706.3600 \[nucl-ex\]\]. A. Czarnecki, W. J. Marciano and A. Sirlin, arXiv:1802.01804 \[hep-ph\]. Z. Tang [*et al.*]{}, arXiv:1802.01595 \[nucl-ex\]. A. P. Serebrov, R. M. Samoilov, I. A. Mitropolsky and A. M. Gagarsky, arXiv:1802.06277 \[nucl-ex\]. R. C. Tolman, Phys. Rev.  [**55**]{} (1939) 364. doi:10.1103/PhysRev.55.364 J. R. Oppenheimer and G. M. Volkoff, Phys. Rev.  [**55**]{} (1939) 374. doi:10.1103/PhysRev.55.374 J. Rikovska-Stone, P. A. M. Guichon, H. H. Matevosyan and A. W. Thomas, Nucl. Phys. A [**792**]{} (2007) 341 doi:10.1016/j.nuclphysa.2007.05.011 \[nucl-th/0611030\]. P. A. M. Guichon, Phys. Lett. B [**200**]{} (1988) 235. doi:10.1016/0370-2693(88)90762-9 P. A. M. Guichon, K. Saito, E. N. Rodionov and A. W. Thomas, Nucl. Phys. A [**601**]{} (1996) 349 doi:10.1016/0375-9474(96)00033-4 \[nucl-th/9509034\]. P. A. M. Guichon and A. W. Thomas, Phys. Rev. Lett.  [**93**]{} (2004) 132502 doi:10.1103/PhysRevLett.93.132502 \[nucl-th/0402064\]. P. A. M. Guichon, A. W. Thomas and K. Tsushima, Nucl. Phys. A [**814**]{} (2008) 66 doi:10.1016/j.nuclphysa.2008.10.001 \[arXiv:0712.1925 \[nucl-th\]\]. K. Saito, K. Tsushima and A. W. Thomas, Prog. Part. Nucl. Phys.  [**58**]{} (2007) 1 doi:10.1016/j.ppnp.2005.07.003 \[hep-ph/0506314\]. M. A. Perez-Garcia and J. Silk, Phys. Lett. B [**711**]{} (2012) 6 doi:10.1016/j.physletb.2012.03.065 \[arXiv:1111.2275 \[astro-ph.CO\]\]. M. Ángeles Pérez-García and J. Silk, Phys. Lett. B [**744**]{} (2015) 13 doi:10.1016/j.physletb.2015.03.026 \[arXiv:1403.6111 \[astro-ph.SR\]\]. A. Li, F. Huang and R. X. Xu, Astropart. Phys.  [**37**]{} (2012) 70 doi:10.1016/j.astropartphys.2012.07.006 \[arXiv:1208.3722 \[astro-ph.SR\]\]. M. A. Perez-Garcia, J. Silk and J. R. Stone, Phys. Rev. Lett.  [**105**]{} (2010) 141101 doi:10.1103/PhysRevLett.105.141101 \[arXiv:1007.1421 \[astro-ph.CO\]\]. B. Kiziltan, A. Kottas, M. De Yoreo and S. E. Thorsett, Astrophys. J.  [**778**]{} (2013) 66 doi:10.1088/0004-637X/778/1/66 \[arXiv:1309.6635 \[astro-ph.SR\]\]. P. Demorest, T. Pennucci, S. Ransom, M. Roberts and J. Hessels, Nature [**467**]{} (2010) 1081 doi:10.1038/nature09466 \[arXiv:1010.5788 \[astro-ph.HE\]\]. J. Antoniadis [*et al.*]{}, Science [**340**]{} (2013) 6131 doi:10.1126/science.1233232 \[arXiv:1304.6875 \[astro-ph.HE\]\]. D. L. Whittenbury, J. D. Carroll, A. W. Thomas, K. Tsushima and J. R. Stone, Phys. Rev. C [**89**]{} (2014) 065801 doi:10.1103/PhysRevC.89.065801 \[arXiv:1307.4166 \[nucl-th\]\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'When performing asymptotic expansions using the strategy of expansion by regions, it is a non-trivial task to find the relevant regions. The recently published Mathematica code `asy.m` automates this task, but it has not been able to detect potential regions in threshold expansions or Glauber regions. In this work we present an algorithm and its implementation in the update `asy2.m` which also reveals potential and Glauber regions automatically.' author: - Bernd Jantzen - 'Alexander V. Smirnov' - 'Vladimir A. Smirnov' date: 'Received: 9 June 2012 / Revised: 23 July 2012' title: | Expansion by regions:\ revealing potential and Glauber regions automatically --- Introduction ============ If a given Feynman integral depends on kinematic invariants and masses which essentially differ in scale, a natural idea is to expand it in ratios of small and large parameters. As a result, the integral is written as a series of simpler quantities than the original integral itself and it can be substituted by a sufficiently large number of terms of such an expansion. For limits typical of Euclidean space (for example, the off-shell large-momentum limit or the large-mass limit), one can write down the corresponding asymptotic expansion in terms of a sum over certain subgraphs of a given graph [@Chetyrkin:1988zz; @Chetyrkin:1988cu; @Gorishnii:1989dd; @Smirnov:1990rz; @Smirnov:1994tg; @books1a]. This prescription of expansion by subgraphs has been mathematically proven (see [@Smirnov:1990rz] and Appendix B.2 of [@books1a]). Moreover, there is an automated tool [@Seidensticker:1999bb; @Harlander:1997zb] where such an expansion by subgraphs is implemented. For limits typical of Minkowski space (i.e. which cannot be formulated in Euclidean space) the universal strategy of expansion by regions [@Beneke:1997zp; @Smirnov:1998vk; @Smirnov:1999bza; @books1a] is available. It consists of the following prescriptions: - Divide the space of the loop momenta into various regions and, in every region, expand the integrand in a Taylor series with respect to the parameters that are considered small there. - Integrate the integrand, expanded in the appropriate way in every region, over the *whole integration domain* of the loop momenta. - Set to zero any scaleless integral. As shown in [@Smirnov:1999bza], this prescription can also be applied to parametric representations of Feynman integrals, i.e. alpha parameters (or generalized Feynman parameters) integrated from 0 to $\infty$, eventually restricted by a delta function. Then the regions are specified by scaling relations between the parameters. There is no mathematical proof that this prescription is correct in all situations. But also no examples are known where a proper application of the expansion by regions leads to wrong results. An indirect proof exists for limits typical of Euclidean space because here the strategy of regions is equivalent to the mathematically proven expansion by subgraphs. A systematic study of the expansion by regions was presented recently in [@Jantzen:2011nz]. There it was shown explicitly and illustrated using various one-loop examples that one can start from a decomposition of a given integral into non-intersecting domains and arrive at an expansion by regions in the above sense. This requires certain conditions on the choice and completeness of the considered regions which were derived in [@Jantzen:2011nz]. As pointed out there, the appearance of additional overlap contributions can be avoided by adequate choices of the regions and regularization parameters. While these findings provide some hints on the proper choice of the regions, it remains a non-trivial task to actually reveal the typical regions for a given limit. Usually, one starts from considering one-loop examples, checks the results against known analytical results, then proceeds in two loops etc. One can also use the second version [@FIESTA2] of the code [FIESTA]{} [@FIESTA] to obtain numerically several first terms of a given asymptotic expansion. Recently an algorithm for an automatic search of regions was suggested and implemented on a computer as the open source [Mathematica]{} code `asy.m` [@asy]. The algorithm uses a geometric approach based on finding the convex hull of a set of points determined from a parametric representation of the Feynman integral. In this way all possible sets of scalings for the (Feynman) parameters are found which lead to non-vanishing (because non-scaleless) integrals. These regions may then be used to expand the parametric integral, or they can be translated into regions for expanding the integral in loop-momentum space. This code works successfully for a large variety of limits, at least in cases where the function ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ in the corresponding parametric representation which depends on kinematic invariants and masses is only composed of terms with the same sign. Moreover, it was shown in [@asy] that in this case there are no regions except for the ones produced by the code (all others result in scaleless integrals). In particular, the code [asy.m]{} works for Sudakov-type limits which are typical of Minkowski space. As it was pointed out in [@asy] the code does not reveal potential regions in threshold expansions, and similarly it fails to detect the so-called Glauber regions. Our goal in the present paper is to provide an algorithm and the corresponding update [asy2.m]{} of the code `asy.m` which automatically identifies all regions relevant for a given integral, including potential and Glauber regions. We start in Section \[sec:preliminaries\] by introducing the parametric representations of loop integrals which we use later and by explaining how the contribution of a given region is obtained in the language of such parametric integrals. Then we elaborate our algorithm for revealing potential and Glauber regions and explain how the code `asy2.m` is applied in such cases. This is done in Section \[sec:potential\] for an example with a potential region, and Section \[sec:Glauber\] deals with the more complicated problem of revealing Glauber regions. In both Sections \[sec:potential\] and \[sec:Glauber\], we first formulate simple changes of variables and decompositions of a given Feynman integral, using instructive one-loop examples, which lead to integrals where [asy2.m]{} is able to detect the relevant regions and print their list in terms of the scalings of the parameters. Then, for both cases, we explain how to use the new features of [asy2.m]{} to perform these algorithmic steps automatically. In the case of Section \[sec:Glauber\] with Glauber regions, the structure of the regions differs depending on whether the expansion is performed in loop-momentum space or at the level of the parametric integral. We show in Section \[sec:Glauber\_l\] how to disentangle and match the various regions arising in this problem by using generic propagator powers, and how [asy2.m]{} can be employed to automate such an analysis. A summary of the new features and the syntax of `asy2.m` (together with a download link) is provided in Section \[sec:syntax\]. In Section \[sec:conclusion\] we conclude by discussing the mathematical problem of proving the expansion by regions for a simple example which is not related to Feynman integrals, but where `asy2.m` works successfully. Expansion by regions in parametric representations {#sec:preliminaries} ================================================== We are dealing with dimensionally regularized Feynman integrals $$\begin{aligned} F(q_1,\ldots,q_n;a_1,\ldots,a_N;d) &= \idotsint \prod_{i=1}^h {\mbox{d}}^d k_i \, \frac{1}{\prod_{l=1}^N E_l^{a_l}}\,, \label{eqbn-d1}\end{aligned}$$ where $h$ is the number of loops, the indices $a_l$ are general powers of the propagators, the dimension is $d=4-2{\varepsilon}$ and the denominators $E_l$ are given by $$\begin{aligned} E_{l}&=&\sum_{i\geq j \geq 1}^h A^{i j}_{l} \, k_i \cdot k_j + \sum_{i=1}^h B^{i}_{l}\cdot k_i + D_l + i0 \;, \label{denom-d1}\end{aligned}$$ i.e. they are quadratic or linear functions of the external momenta $q_i$ and the loop momenta $k_i$ with the usual infinitesimal imaginary part $+i0$. Monomials in the numerator are taken into account as denominators raised to negative powers. The alpha representation of (\[eqbn-d1\]) takes the form $$\begin{aligned} &F(q_1,\ldots,q_n;a_1,\ldots,a_N;d) = (i \pi^{d/2})^h \, \frac{e^{-i \pi (a + h d/2)/2}}{\prod_{l=1}^N {\Gamma}(a_l)} {\nonumber}\\ &\qquad \times \int_0^\infty \cdots \int_0^\infty \prod_{l=1}^N \left({\mbox{d}}{\alpha}_l \, {\alpha}_l^{a_l-1}\right) {{\text{\usefont{OMS}{cmsy}{m}{n}U}}}^{\;-d/2} \, e^{-i \,{{\text{\usefont{OMS}{cmsy}{m}{n}F}}}/{{\text{\usefont{OMS}{cmsy}{m}{n}U}}}} \; , \label{alpha-d}\end{aligned}$$ where $a=\sum_l a_l$. The functions ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ depend polynomially on the alpha parameters ${\alpha}_l$. Furthermore, ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ are homogeneous functions of the alpha parameters with the homogeneity degrees $h$ and $h+1$, respectively. The function ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ is linear in the kinematic invariants and/or squared masses which we denote by $s_i,\;i=1,2,\ldots$, while the function ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ is independent of the $s_i$. If (\[eqbn-d1\]) is an integral with standard propagators $1/(p_l^2-m_l^2+i0)$ associated with the lines of a graph, then the functions ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ are called Symanzik polynomials and are given by the well-known formulae in terms of trees and $2$-trees. For a general Feynman integral of the form (\[eqbn-d1\]) one can obtain these functions using the simple public code `UF.m`[^1] [@UF.m] which is also part of the codes `asy.m` and `asy2.m`. If some of the indices $a_l$ are negative integers, i.e. they correspond to numerators instead of denominators of the integral (\[eqbn-d1\]), the alpha representation (\[alpha-d\]) is to be understood in the limit where these indices tend to their negative integer values. Effectively, the integration over the corresponding parameters ${\alpha}_l$ is replaced by differentiating with respect to these parameters and setting them to zero. Closely related to (\[alpha-d\]) is the (generalized) Feynman parametric representation $$\begin{aligned} &F(q_1,\ldots,q_n;a_1,\ldots,a_N;d) = (i \pi^{d/2})^h \, \frac{e^{-i \pi a} \, {\Gamma}(a - h d/2)}{\prod_{l=1}^N {\Gamma}(a_l)} {\nonumber}\\ &\times \int_0^\infty {\mbox{d}}x_1 \cdots \int_0^\infty{\mbox{d}}x_N \, \delta\!\left( \sum_{l\in \nu} x_l-1\right) I(x_1,\ldots,x_N;s_1,s_2,\ldots) \; , \label{alpha-d-mod}\end{aligned}$$ where $\nu$ in the argument of the delta function is an arbitrary non-empty subset of $\{1,\ldots,N\}$, $$I(x_1,\ldots,x_N;s_1,s_2,\ldots)= \prod_{l=1}^N x_l^{a_l-1} \, {{\text{\usefont{OMS}{cmsy}{m}{n}U}}}^{\;a-(h+1) \frac{d}{2}} \, ({{\text{\usefont{OMS}{cmsy}{m}{n}F}}}- i0)^{h \frac{d}{2}-a} \label{integrand}$$ and the functions ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ are the same as those in (\[alpha-d\]) with the parameters ${\alpha}_l$ replaced by $x_l$. It is well known that the formula (\[alpha-d-mod\]) holds for any choice of the subset $\nu$ in the argument of the delta function.[^2] This feature is related to the above-mentioned homogeneity properties of the functions ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$.[^3] If one chooses $\nu = \{1,\ldots,N\}$, the standard Feynman parametrization is recovered. Let us suppose that we have to study the asymptotic behaviour in a one-scale limit, i.e. every mass and kinematic invariant has a certain scaling $s_i\to s'_i= \rho^{\kappa_i} s_i$, $i=1,2,\ldots$, expressed in powers of the small parameter of the problem, $\rho$. The strategy of expansion by regions formulated in terms of parametric integrals (\[alpha-d\]) or (\[alpha-d-mod\]) [@Smirnov:1999bza; @books1a] states that the asymptotic expansion in such a limit is given by a sum over regions which are specified by the scalings of the parameters ${\alpha}_l$ or $x_l$ expressed in powers $r_l$ of the expansion parameter $\rho$. Each region $r$ is labelled by the list $r=\{r_1,\ldots,r_N\}$ of its scaling powers. The contribution of the region $r$ is obtained by scaling the masses and kinematic invariants according to the given limit as specified above, by substituting ${\alpha}_l \to {\alpha}'_l= \rho^{r_l} {\alpha}_l,\;l=1\ldots,N$, in the integrand of (\[alpha-d\]) or $x_l \to x'_l= \rho^{r_l} x_l$ in the integrand of (\[alpha-d-mod\]) and by expanding the integrand in powers of $\rho$. Here the product of the differentials ${\mbox{d}}{\alpha}_l$ or ${\mbox{d}}x_l$ provides another factor $\rho^{\sum_l r_l}$ to the power counting. Explicitly, the contribution of the region $r$ is given by the prefactor in (\[alpha-d-mod\]) times $\rho^{\sum_l r_l}$ times the integral $$\int_0^\infty {\mbox{d}}x_1 \cdots\int_0^\infty{\mbox{d}}x_N \, \delta\!\left( \sum_{l\in \nu} x'_l-1\right) I(x'_1,\ldots,x'_N;s'_1,s'_2,\ldots) \label{alpha-d-mod-scaled}$$ with the integrand expanded in powers of $\rho$. This expansion also involves the argument of the delta function in (\[alpha-d-mod-scaled\]), such that, under the expansion, certain parameters drop out of the argument of the delta function and are integrated from 0 to $\infty$. For this reason the upper integration limit of all Feynman parameters should be kept at infinity and not switched to $1$ even if, before the expansion, their integration is restricted by the delta function. One may avoid expanding the delta function by choosing the original subset $\nu$ in (\[alpha-d-mod\]) sufficiently small. Let us write down the leading-order (LO) contribution of a given region in a more explicit way. For the two basic functions in (\[integrand\]) we have $$\begin{aligned} {{\text{\usefont{OMS}{cmsy}{m}{n}U}}}(x'_1,\ldots,x'_N) &= \sum_{j=u_{\text{min}}}^{u_{\text{max}}} \rho^j \, {{\text{\usefont{OMS}{cmsy}{m}{n}U}}}_j(x_1,\ldots,x_N)\;, {\nonumber}\\ {{\text{\usefont{OMS}{cmsy}{m}{n}F}}}(x'_1,\ldots,x'_N;s'_1,s'_2,\ldots) &= \sum_{j=f_{\text{min}}}^{f_{\text{max}}} \rho^j \, {{\text{\usefont{OMS}{cmsy}{m}{n}F}}}_j(x_1,\ldots,x_N;s_1,s_2,\ldots)\;, \label{lpolyn}\end{aligned}$$ where the arguments of the polynomials ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ on the left-hand side indicate that they are expressed in terms of the scaled parameters $x'_l$ and $s'_i$, while their expansion coefficients on the right-hand side are expressed in terms of $x_l$ and $s_i$. According to the prescription formulated above, the LO contribution of the region $r$ is represented as $$\begin{aligned} &\rho^{\sum_l r_l a_l + u_{\text{min}}\left(a-(h+1)\frac{d}{2}\right) + f_{\text{min}}\left(h \frac{d}{2} - a\right)} {\nonumber}\\ &\times (i \pi^{d/2})^h \, \frac{e^{-i \pi a} \, {\Gamma}(a - h d/2)}{\prod_{l=1}^N {\Gamma}(a_l)} \int_0^\infty \cdots \int_0^\infty \prod_{l=1}^N \left({\mbox{d}}x_l \, x_l^{a_l-1}\right) {\nonumber}\\ &\times \delta\!\left( \sum_{l\in \nu_0} x_l-1\right) {{\text{\usefont{OMS}{cmsy}{m}{n}U}}}_{u_{\text{min}}}^{\;a-(h+1)\frac{d}{2}} \, ({{\text{\usefont{OMS}{cmsy}{m}{n}F}}}_{f_{\text{min}}}-i0)^{h \frac{d}{2} - a} \;. \label{alpha-d-mod0}\end{aligned}$$ In principle, the argument of the delta function in (\[alpha-d-mod0\]) contains a sum over only those scaled parameters $x'_l = \rho^{r_l} x_l$ with the minimal scaling power $r_l = r_{\text{min}}= \min\{r_1,\ldots,r_N\}$. But by rescaling $x_l \to \rho^{-r_{\text{min}}} \, x_l \, \forall l$, the delta function is transformed into its standard form without powers of $\rho$, while the rest of the integral remains invariant due to the homogeneity of the polynomials ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}_{u_{\text{min}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}_{f_{\text{min}}}$. Finally, as for the original Feynman parametric representation (\[alpha-d-mod\]), one can choose again an arbitrary non-empty subset $\nu_0$ for the sum in the argument of the delta function in (\[alpha-d-mod0\]). The list of scalings $r=\{r_1,\ldots,r_N\}$ of a region is only determined up to adding the same arbitrary real number $c$ to each entry, because the corresponding contribution stays the same under $r_l \to r_l + c \, \forall l$. In particular, the LO behaviour presented in (\[alpha-d-mod0\]) is independent of such a shift $c$, because $u_{\text{min}}\to u_{\text{min}}+ hc$ and $f_{\text{min}}\to f_{\text{min}}+ (h+1)c$, due to the homogeneity properties of ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$. The terms of the expansion come from various regions and can be ordered according to accompanying powers of $\rho$. After keeping some first terms of the expansion one can set $\rho=1$ and write down the given Feynman integral as these selected first terms plus a remainder which vanishes sufficiently fast in the given limit. It turns out that only a limited number of regions contribute to the expansion because for the rest of the regions one obtains integrals without scale which are set to zero. It is the subject of this paper and the task of the code [asy.m]{} and its updated version [asy2.m]{} to find all relevant regions for a given integral. Revealing potential contributions {#sec:potential} ================================= Let us consider the one-loop propagator diagram with two massive lines in the threshold limit, i.e. when $y=m^2-q^2/4\to 0$ with $q$ being the external momentum: $$F(q^2,m^2)=\int \frac{{\mbox{d}}^d k}{(k^2-m^2)\; \bigl((k-q)^2-m^2\bigr)}\;, \label{prop1_integral}$$ where the usual $+i 0$ is implied in all the propagators. Within the strategy of expansion by regions, the hard and the potential regions give contributions to the expansion [@Beneke:1997zp; @books1a]. The previous version of the code [asy.m]{} [@asy] reported only about the hard region. The reason for this can be seen in the corresponding parametric representation, $$\begin{aligned} F(q^2,y)&=i\pi^{d/2} \, {\Gamma}({\varepsilon}) {\nonumber}\\* &\times \iint \frac{(x_1+x_2)^{2{\varepsilon}-2}\;\delta\left(x_1+x_2-1\right)\; {\mbox{d}}x_1{\mbox{d}}x_2 }{\left[ \frac{q^2}{4}(x_1-x_2)^2 + y(x_1+x_2)^2 -i 0\right]^{{\varepsilon}}}\;, \label{prop1_alpha}\end{aligned}$$ where the parameters $x_i$ are integrated from 0 to $\infty$ (restricted by the delta function). As it was pointed out in [@asy], it is the region where $x_1\approx x_2$ (more precisely ) which causes problems. In other words, the polynomial in the square brackets in (\[prop1\_alpha\]) (considered at positive $q^2$ and $y$) has terms of different sign, such that cancellations occur because of the presence of the negative term $-q^2 x_1x_2/2$. To reveal the missing potential contribution, let us perform a simple trick. We decompose the integration domain into two subdomains, $x_1\leq x_2$ and $x_2\leq x_1$. The two resulting integrals are equal to each other, but such an equality will not generally take place for any integral. In the first domain we turn to new variables by $x_1=x_1'/2,\;x_2=x'_2+x'_1/2$, remove the primes at $x_i$ and obtain the integral (again from 0 to $\infty$ with the usual restrictions via the delta function) $$i\pi^{d/2} \, \frac{{\Gamma}({\varepsilon})}{2} \iint \frac{(x_1+x_2)^{2{\varepsilon}-2}\;\delta\left(x_1+x_2-1\right)\; {\mbox{d}}x_1{\mbox{d}}x_2} {\left[ \frac{q^2}{4}x_2^2 + y(x_1+x_2)^2 -i 0\right]^{{\varepsilon}}}\;. \label{prop1_alpha_1}$$ The goal of this trick was to make the line $x_1=x_2$ (in the old variables) the border of an integration domain which turned out to be (in the new variables) $x_2=0$. Now we can run the code [asy2.m]{}. Since this is a parametrical integral rather than a Feynman integral we use the newly introduced command[^4] [WilsonExpand\[\]]{} for integrals where all parameters are integrated from 0 to $\infty$: WilsonExpand[q^2/4*x2^2 + y*(x1 + x2)^2, x1 + x2, {x1, x2}, {q -> 1, y -> x}, Delta -> True] The first two arguments of `WilsonExpand[]` are the polynomials ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$, respectively, as defined in Section \[sec:preliminaries\]. They can easily be determined from the square brackets in the denominator of the parametric integral (\[prop1\_alpha\_1\]) and from the round brackets in the numerator. The third argument is the list of integration parameters, and the fourth argument specifies the scaling of the kinematic quantities with respect to the small parameter which is labelled by the global symbol `x`. Here by `y -> x` we tell the code that $y$ is the small expansion parameter, and by `q -> 1` we specify that the momentum $q$ scales as $y^0=1$. The option `Delta -> True` tells `WilsonExpand[]` that, under the integration, the sum over an arbitrary non-empty subset of the integration parameters is restricted to 1 by a delta function. Note that `WilsonExpand[]` can only take into account such a delta function if the specific choice of the sum over parameters in the argument of the delta function is irrelevant. This is the case for the generalized Feynman parametric integral (\[alpha-d-mod\]) introduced in Section \[sec:preliminaries\]. The integrals (\[prop1\_alpha\]) and (\[prop1\_alpha\_1\]) are special cases of (\[alpha-d-mod\]) such that we could e.g. replace $\delta(x_1+x_2-1)$ by $\delta(x_1-1)$ without changing the value of the integrals. If, however, a specific form of the delta function is assumed, e.g. by replacing $x_1+x_2 \to 1$ under the integral, then the option `Delta` of `WilsonExpand[]` does not apply (see Section \[sec:syntax\] for details). Alternatively, `WilsonExpand[]` can be used without the option `Delta` after eliminating one of the integrations, e.g. via $\delta(x_1-1)$. The output of the above-mentioned call of `WilsonExpand[]` is a list of regions specified by the scaling of the parameters $x_i$ in powers of the small parameter $y$: {{0, 0}, {0, 1/2}} This specification of a region corresponds to the list of scalings $\{r_1,\ldots,r_N\}$ introduced in Section \[sec:preliminaries\]. The first entry in the output, $\{0,0\}$, refers to the hard region with the scaling $x_1\sim y^0$, $x_2\sim y^0$ which has already been found by `asy.m`. But now also the region $\{0, 1/2\}$ is found with the scaling $x_1\sim y^0$, $x_2\sim y^{1/2}$ which provides the potential contribution. The contribution of the hard region starts with the order $y^0$. Every term of the expansion can be evaluated in terms of gamma functions for general ${\varepsilon}$. According to the prescriptions for writing down the contribution of a region formulated in Section \[sec:preliminaries\], the contribution of the $k$-th order expansion of (\[prop1\_alpha\_1\]) in the potential region reads $$i\pi^{d/2} \, \frac{{\Gamma}({\varepsilon})}{2 \, k!} \iint {\mbox{d}}x_1{\mbox{d}}x_2 \, x_2^k \left(\frac{\partial}{\partial x_1}\right)^k \, \frac{x_1^{2{\varepsilon}-2}\;\delta\left(x_1-1\right)} {\left(\frac{q^2}{4}x_2^2 + y \, x_1^2\right)^{{\varepsilon}}}\;. \label{prop1_alpha_p}$$ Only the leading order ($k=0$), $$i\pi^{d/2} \, \frac{{\Gamma}({\varepsilon})}{2} \int_0^\infty \frac{{\mbox{d}}x_2}{\left(\frac{q^2}{4}x_2^2 + y\right)^{{\varepsilon}}}\;, \label{prop1_alpha_p0}$$ yields a non-vanishing contribution which is evaluated in terms of gamma functions at general ${\varepsilon}$. Taking into account that we have two identical integrals after our decomposition, we arrive at the following result for the potential contribution which is of order $y^{1/2-{\varepsilon}}$: $$i \pi^{d/2} \, {\Gamma}({\varepsilon}-1/2) \sqrt{\frac{\pi y}{q^2}} \, y^{-{\varepsilon}} \label{prop1_pot}$$ in agreement with [@books1a]. In fact, such a trick of making manifest squares of some linear combination of the integration parameters was already used in the code [FIESTA]{} [@FIESTA] in order to evaluate numerically Feynman integrals at a threshold. Using the implementation of this procedure in [FIESTA]{} it turned out to be possible to automate the above trick for a general Feynman integral. In the present version [asy2.m]{}, the user may call the command `AlphaRepExpand[]` with the additional option `PreResolve` enabled which automatically looks for the change of variables described above: AlphaRepExpand[{k}, {k^2 - m^2, (k - q)^2 - m^2}, {q^2 -> qq, m^2 -> qq/4 + y}, {qq -> 1, y -> x}, PreResolve -> True] As in the previous version of `asy.m`, the arguments of `AlphaRepExpand[]` are the list of loop momenta, the list of denominators of the loop integral, a list of replacements for the kinematic quantities, and the list of scalings with respect to the small parameter `x`. The output is a list of entries which indicate for each region the changes of variables, the Jacobian of the integral transformation, and the scalings of the new variables: {{{x[1] -> y[1] + y[2]/2, x[2] -> y[2]/2}, 2, {0, -1/2}}, {{x[1] -> y[1] + y[2]/2, x[2] -> y[2]/2}, 2, {0, 0}}, {{x[1] -> y[1]/2, x[2] -> y[1]/2 + y[2]}, 2, {0, 0}}, {{x[1] -> y[1]/2, x[2] -> y[1]/2 + y[2]}, 2, {0, 1/2}}} The original Feynman parameters are labelled `x[1]`, `x[2]`, $\ldots$, while the new parameters are labelled `y[1]`, `y[2]`, $\ldots$. The scaling relations which determine the regions are specified for the new parameters. As explained in Section \[sec:preliminaries\], a region remains invariant if all of its scalings are shifted by the same amount, i.e. if all parameters are rescaled simultaneously. So the scaling $\{0,-1/2\}$ shown above is equivalent to $\{1/2,0\}$, i.e. the first new parameter is suppressed by $y^{1/2}$ with respect to the second one. For both parts of the integral decomposition the hard and potential regions are found. The preresolution algorithm implemented in [asy2.m]{} (and switched on with the option `PreResolve`) tries to eliminate factorized combinations of terms in the function ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ which potentially cancel each other, like $(x_1-x_2)^2$ in the example above. It checks all pairs of variables (say, $x$ and $y$) which are part of monomials with opposite sign. For all those pairs the code tries to build a linear combination $z$ of $x$ and $y$ such that in the variables $x$ and $z$ or $y$ and $z$ this monomial disappears. The code checks whether in the new variables the number of monomials with opposite sign decreases. For all such pairs the code recursively repeats the initial procedure in the new variables. As a result it creates a tree of possible bisections and corresponding replacements of variables. A leaf of this tree is a set of sectors and functions such that one cannot decrease the number of monomials with opposite sign any longer. Ideally it means that all monomials now have the same sign. The code analyzes all leafs and chooses one of those with the minimal number of opposite-sign monomials (or the minimal number of sectors if the numbers of monomials with opposite sign coincide). After finishing with the preresolution, the code performs the replacements and looks for regions in all those sectors, using the algorithm of the original code `asy.m` described in [@asy]. Note that the algorithm can only find the necessary variable transformations if it is able to determine the relative signs of all terms in the polynomial ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$. As the signs of symbols are unknown to `asy2.m`, the substitution rules of the third and fourth arguments of `AlphaRepExpand[]` must replace all kinematic quantities by numbers (integers or fractions of integers) or powers of the small parameter `x`. We have checked that the updated version [asy2.m]{} works in various examples of the threshold expansion (considered in [@Beneke:1997zp; @books1a]): a triangle, a box, the two-loop propagator diagram (with the masses $m,m,m,m,0$), a two-loop vertex diagram. Because of the decomposition of a given integration domain into subdomains, the number of resulting integrals for various regions increases a little bit. For example, the (hard-hard) region for the two-loop propagator diagram is described by six integrals, the (potential-ultrasoft) region is also described by six integrals, etc. However, the (potential-hard) region is described by four integrals with some regions (with scalings composed of powers 1, 1/2 and 0), and four more integrals with a set of regions of a different type (composed of 1 and 0). Let us finally mention that the preresolution algorithm also works for threshold expansions with unequal masses. Returning to the one-loop example (\[prop1\_integral\]), but now with different masses $m_1$ and $m_2$ in the propagators and the expansion parameter defined as $y = (m_1+m_2)^2/4 - q^2/4 \to 0$, we call: AlphaRepExpand[{k}, {k^2 - m1^2, (k - q)^2 - m2^2}, {q^2 -> (m1 + m2)^2 - 4*y}, {y -> x, m1 -> 2, m2 -> 3/2}, PreResolve -> True] We have set the values of the masses to different rational numbers, $m_1 \to 2$ and $m_2 \to 3/2$ (the actual values are irrelevant), thus permitting the preresolution algorithm to know about their sign without assuming any equality or other relation between them. The polynomial ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ in the parametric representation of the Feynman integral reads ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}= (m_1 x_1 - m_2 x_2)^2 + 4y x_1 x_2$, and cancellations occur where $m_1 x_1 = m_2 x_2$. These cancellations are automatically made explicit by adequate changes of variables. The output {{{x[1] -> y[1] + 3*y[2]/7, x[2] -> 4*y[2]/7}, 7/4, {0, -1/2}}, {{x[1] -> y[1] + 3*y[2]/7, x[2] -> 4*y[2]/7}, 7/4, {0, 0}}, {{x[1] -> 3*y[1]/7, x[2] -> 4*y[1]/7 + y[2]}, 7/3, {0, 0}}, {{x[1] -> 3*y[1]/7, x[2] -> 4*y[1]/7 + y[2]}, 7/3, {0, 1/2}}} shows that `asy2.m` detects both regions. The variable transformations are always normalized such that sums of the parameters (here $x_1+x_2$) remain invariant. So, we now have a manifestly Lorentz-invariant treatment of threshold expansion and a code that automatically provides the set of relevant regions. Revealing Glauber contributions {#sec:Glauber} =============================== ![One-loop five-point integral exhibiting a Glauber contribution[]{data-label="fig:Glauber"}](Glauber-5point){width="\columnwidth"} Let us consider the one-loop five-point integral in Fig. \[fig:Glauber\], where two initial-state partons both perform a collinear splitting into two partons each with momenta $p_1,p_2$ and $q_1,q_2$, respectively. While two partons, one of each pair, collide with a large centre-of-mass energy $Q=\sqrt{(p_2+q_2)^2}$, the two remaining partons exchange a particle with the small mass $m$. We will use the simplified kinematics $p_1 = p_2 = p$ and $q_1 = q_2 = q$ with $p^2 = q^2 = 0$ and $(p+q)^2 = 2p\cdot q = Q^2$ in the limit $m^2/Q^2 \to 0$: $$\begin{aligned} F(Q^2,m^2)&=\int \frac{{\mbox{d}}^d k}{(k^2-m^2)(k^2-2p\cdot k) (k^2+2p\cdot k)} {\nonumber}\\* &\times \frac{1}{(k^2-2q\cdot k) (k^2+2q \cdot k)} \;. \label{5pt_integral}\end{aligned}$$ Before we search for regions using `asy.m`, we notice that this five-point integral is similar to the Sudakov form factor example treated in Section 6 of [@Jantzen:2011nz]. From the viewpoint of the convergence of the expansions, the second and third propagators of (\[5pt\_integral\]) are equivalent, and so are the fourth and fifth propagators. Effectively, the five-point integral has only three different types of propagators which are equivalent to the ones of the three-point integral in [@Jantzen:2011nz]. So the integral (\[5pt\_integral\]) can be expanded in loop-momentum space employing the regions and convergence domains known from [@Jantzen:2011nz] (and using generic propagator powers as analytic regulators where necessary): - a hard region where $k \sim Q$, - a 1-collinear region where $k^2 \sim p\cdot k \sim m^2$ and $q\cdot k \sim Q^2$, - a 2-collinear region where $k^2 \sim q\cdot k \sim m^2$ and $p\cdot k \sim Q^2$, - a Glauber region where $p\cdot k \sim q\cdot k \sim m^2$, and the components of $k$ perpendicular to the plane spanned by $p,q$ scale as $k_\perp \sim m$. The collinear-plane region mentioned in [@Jantzen:2011nz] yields only scaleless contributions. But, in contrast to the three-point integral, the five-point integral has a non-vanishing Glauber contribution. The Glauber region even provides the leading contribution scaling as $(m^2)^{-2-{\varepsilon}}$, whereas the collinear contributions start with $(m^2)^{-1-{\varepsilon}}$ and the hard contribution starts with $(m^2)^0$. The five-point integral (\[5pt\_integral\]) can be represented in terms of an integral over Feynman parameters, $$\begin{aligned} &F(Q^2,m^2)=-i\pi^{d/2} \, {\Gamma}(3+{\varepsilon}) \idotsint {\mbox{d}}x_1\cdots{\mbox{d}}x_5 {\nonumber}\\* &\;\;\times \frac{\delta\left(\sum_i x_i-1\right) \; (x_1+\ldots+x_5)^{1+2{\varepsilon}} }{\bigl[x_1(x_1+\ldots+x_5) m^2 +(x_2-x_3)(x_4-x_5)Q^2-i 0\bigr]^{3+{\varepsilon}} }\;, \label{Fp_integral}\end{aligned}$$ where one can choose the sum in the argument of the delta function in an appropriate way, i.e. restrict only the sum over a subset of the parameters to 1 and extend the integration over the rest of the parameters to the whole domain $[0,\infty)$. Applying the strategy of expansion by regions in Feynman-parameter space and trying to reveal regions relevant to the given limit $m^2/Q^2 \to 0$ with the help of the code [asy.m]{} [@asy], we call: AlphaRepExpand[{k}, {k^2 - m^2, k^2 - 2*p*k, k^2 + 2*p*k, k^2 - 2*q*k, k^2 + 2*q*k}, {p^2 -> 0, q^2 -> 0, p*q -> Q^2/2}, {Q -> 1, m^2 -> x}] The output states the following set of three regions: {{0, 0, 0, 0, 0}, {0, 0, 0, 1, 1}, {0, 1, 1, 0, 0}} As before, the regions are specified by the scaling of the Feynman parameters in terms of powers of the small parameter $m^2$. For example, for the second region we have $x_1 \sim x_2 \sim x_3 \sim(m^2)^0$, $x_4 \sim x_5 \sim(m^2)^1$. The first region is hard; its contribution starts with $(m^2)^0$. The second and third regions start with order $(m^2)^{-1-{\varepsilon}}$. They correspond to the two collinear regions stated for the momentum-space expansion above. But `asy.m` does not find anything corresponding to the Glauber region; in particular, none of the regions found by `asy.m` provides the leading $(m^2)^{-2-{\varepsilon}}$ contribution. We notice that, as in the previous section about potential contributions, the polynomial in the square brackets of (\[Fp\_integral\]) has terms of different sign. The missing Glauber contribution stems from the parameter region where either $(x_2-x_3) \sim (m^2)^1$ or $(x_4-x_5) \sim (m^2)^1$. So let us decompose the parametric integral into four parts corresponding to the domains where the two factors $(x_2-x_3)$ and $(x_4-x_5)$ are either positive or negative and then introduce new variables in such a way that this product takes the form $\pm x'_2 x'_4$. For example, in the domain $x_2\leq x_3$, $x_5\leq x_4$ we change the variables by $x_2=x'_3/2,\;x_3=x'_2+x'_3/2$ and by $x_4=x'_4+x'_5/2,\;x_5=x'_5/2$, similarly to our example in the previous section. However, in the threshold expansion the cancelling terms appeared in squared form such that a transformation between one pair of variables was sufficient. Here two separate factors involve cancellations, which requires a twofold change of variables. Removing the primes from the variables $x_i$, the parametric integral reads $$F(Q^2,m^2)=2(I_+ + I_-)$$ with $$\begin{aligned} I_{\pm} &=-i\pi^{d/2} \, \frac{{\Gamma}(3+{\varepsilon})}{4} \idotsint {\mbox{d}}x_1\cdots{\mbox{d}}x_5 {\nonumber}\\* &\times \frac{\delta\left(x_1-1\right) \; (x_1+x_2+x_3+x_4+x_5)^{1+2{\varepsilon}} }{\left[x_1(x_1+x_2+x_3+x_4+x_5) m^2 \pm x_2 x_4Q^2 -i 0\right]^{3+{\varepsilon}} }\;, \label{Fp_integral_A}\end{aligned}$$ where we have chosen the argument of the delta function as $x_1-1$, so that we may also write $$\begin{aligned} I_{\pm} &=-i\pi^{d/2} \, \frac{{\Gamma}(3+{\varepsilon})}{4} \int_0^\infty \cdots \int_0^\infty {\mbox{d}}x_2\cdots{\mbox{d}}x_5 {\nonumber}\\* &\times \frac{(1+x_2+x_3+x_4+x_5)^{1+2{\varepsilon}} }{\left[(1+x_2+x_3+x_4+x_5) m^2 \pm x_2 x_4Q^2 -i 0\right]^{3+{\varepsilon}} }\;. \label{Fp_integral_B}\end{aligned}$$ It is sufficient to consider the expansion of $I_+$ and obtain a result for $I_-$ by analytically continuing $Q^2 \to -Q^2-i0$, taking into account that the dependence on $Q^2$ is power-like. Now we can apply [asy2.m]{} to the integral $I_+$ using either the command WilsonExpand[x1*(x1 + x2 + x3 + x4 + x5)*m^2 + x2*x4*Q^2, x1 + x2 + x3 + x4 + x5, {x1, x2, x3, x4, x5}, {Q^2 -> 1, m^2 -> x}, Delta -> True] for integrals (\[Fp\_integral\_A\]) restricted by a delta function (see Section \[sec:syntax\]), yielding the output: {{0, 0, 0, 0, 0}, {0, 1, 0, 0, 0}, {0, 0, 0, 1, 0}} Or we use the command WilsonExpand[(1 + x2 + x3 + x4 + x5)*m^2 + x2*x4*Q^2, 1 + x2 + x3 + x4 + x5, {x2, x3, x4, x5}, {Q^2 -> 1, m^2 -> x}] for integrals (\[Fp\_integral\_B\]) over variables from $0$ to $\infty$ without any restriction, and obtain the output: {{0, 0, 0, 0}, {1, 0, 0, 0}, {0, 0, 1, 0}} The two results are equivalent, as $x_1 \sim (m^2)^0$ is implied in the second case. So we obtain again three regions. We will see in a moment that this list of regions is indeed correct and complete. But first, let us emphasize that the new preresolution algorithm in `asy2.m` is capable of performing the transformation of the integral from (\[Fp\_integral\]) to (\[Fp\_integral\_A\]) automatically: AlphaRepExpand[{k}, {k^2 - m^2, k^2 - 2*p*k, k^2 + 2*p*k, k^2 - 2*q*k, k^2 + 2*q*k}, {p^2 -> 0, q^2 -> 0, p*q -> Q^2/2}, {Q -> 1, m^2 -> x}, PreResolve -> True] The output of this command lists four different variable transformations according to the twofold decomposition described above. For each of the integrals over new parameters, the regions $\{0,0,0,0,0\}$, $\{0,1,0,0,0\}$ and $\{0,0,0,1,0\}$ are found, up to permutations in the order of the parameters from different changes of variables. The evaluation of the contributions to each region, as found by `WilsonExpand[]` or `AlphaRepExpand[]` (including the `PreResolve` option), is straightforward. The first region is the hard one. The contributions of the second and third regions are not individually regularized by dimensional regularization, as it often happens for Sudakov-type limits. We use an auxiliary analytic regularization by introducing additional powers $x_2^{{\delta}_2} x_3^{{\delta}_3} x_4^{{\delta}_4} x_5^{{\delta}_5}$ of the new variables into the integrand of (\[Fp\_integral\_B\]), taking the limit ${\delta}_2,{\delta}_3,{\delta}_4,{\delta}_5 \to 0$ in the end. The leading-order (LO) contribution of the second and third regions to the integral $F(Q^2,m^2)$ reads $$\begin{aligned} -i\pi^{d/2} \, \frac{i\pi \Gamma({\varepsilon})}{2Q^2(m^2)^{2+{\varepsilon}}}\;. \label{Fp_LO}\end{aligned}$$ This agrees with the leading contribution of the Glauber region in the momentum-space expansion. We have found the leading Glauber contribution of order $(m^2)^{-2-{\varepsilon}}$. But we seem to have lost the two collinear regions with the scalings $\{0,0,0,1,1\}$ and $\{0,1,1,0,0\}$ found before the change of variables. In fact, we can evaluate the contributions from these two regions by expanding the integral (\[Fp\_integral\_B\]). The resulting integrals are scaleless and regularized by the parameters ${\delta}_3, {\delta}_5$, so they vanish, and `asy2.m` is right in omitting these two regions. We are also able to solve the integral (\[Fp\_integral\_B\]) including the auxiliary analytic regularization factor $x_2^{{\delta}_2} x_3^{{\delta}_3} x_4^{{\delta}_4} x_5^{{\delta}_5}$ in terms of a onefold Mellin–Barnes representation: $$\begin{aligned} I_{\pm} &=-i\pi^{d/2} \, \frac{{\Gamma}(1+{\delta}_3) {\Gamma}(1+{\delta}_5)}{4} {\nonumber}\\ &\times \frac{1}{2\pi i} \int {\mbox{d}}z \, (m^2)^z (\pm Q^2-i0)^{-3-{\varepsilon}-z} {\nonumber}\\ &\times {\Gamma}(-z) {\Gamma}(-2-{\varepsilon}+{\delta}_2-z) {\Gamma}(-2-{\varepsilon}+{\delta}_4-z) {\nonumber}\\ &\times \frac{{\Gamma}(1-{\delta}_2-{\delta}_3-{\delta}_4-{\delta}_5+z) {\Gamma}(3+{\varepsilon}+z)}{{\Gamma}(-1-2{\varepsilon}-z)} \;. \label{Fp_integral_MB}\end{aligned}$$ The asymptotic expansion of $I_\pm$ in the limit $m^2/Q^2 \to 0$ is obtained by taking the residues of the poles of the functions ${\Gamma}(\ldots-z)$. The poles of ${\Gamma}(-z)$ correspond to the hard region, while the poles of the two functions ${\Gamma}(-2-{\varepsilon}+{\delta}_{2,4}-z)$ provide the contributions of the second and third regions. So `asy2.m` has found all contributing regions. In the Mellin–Barnes integral (\[Fp\_integral\_MB\]) we can safely take the limit ${\delta}_2,{\delta}_3,{\delta}_4,{\delta}_5 \to 0$, add up $I_+$ and $I_-$, and arrive at the Mellin–Barnes representation $$\begin{aligned} &F(Q^2,m^2) = i\pi^{d/2} \, \frac{i}{2} \, \frac{1}{2\pi i} \int {\mbox{d}}z \, (m^2)^z (Q^2)^{-3-{\varepsilon}-z} e^{i\pi({\varepsilon}+z)/2} {\nonumber}\\* &\;\;\times {\Gamma}(-z) {\Gamma}(-2-{\varepsilon}-z) {\Gamma}\left(\tfrac{-1-{\varepsilon}-z}{2}\right) \, \frac{{\Gamma}(1+z) {\Gamma}\left(\frac{3+{\varepsilon}+z}{2}\right)}{{\Gamma}(-1-2{\varepsilon}-z)} \;. \label{Fp_integral_MB2}\end{aligned}$$ The LO contribution to $F(Q^2,m^2)$ is obtained from the residue of the single pole at $z=-2-{\varepsilon}$, in agreement with (\[Fp\_LO\]). The next-to-leading-order (NLO) contribution stems from the residue of the double pole at $z=-1-{\varepsilon}$ and reads $$\begin{aligned} i\pi^{d/2} \, \frac{ \Gamma (1+{\varepsilon})}{(Q^2)^2(m^2)^{1+{\varepsilon}}} \biggl( & i \frac{\pi}{2} +2 \psi(-{\varepsilon})- \psi(1+{\varepsilon})+ {\gamma}_{\rm E} {\nonumber}\\* & -\ln\frac{Q^2}{m^2} -1 \biggr)\;. \label{Fp_NLO}\end{aligned}$$ This agrees with the NLO contributions of the second and third regions. At next-to-next-to-leading order (NNLO) there is a contribution from the residue of the single pole at $z=-{\varepsilon}$ which reads $$-i\pi^{d/2} \, \frac{i\pi \Gamma (2+{\varepsilon})}{4(Q^2)^3(m^2)^{{\varepsilon}}} \label{Fp_NNLO_g}$$ and agrees with the NNLO contributions of the second and third regions. The second NNLO contribution comes from the residue of the single pole at $z=0$. It is given by $$-i\pi^{d/2} \, \frac{i \, e^{i\pi{\varepsilon}/2} \, {\Gamma}(-2-{\varepsilon}) {\Gamma}\left(\frac{1+{\varepsilon}}{2}\right) {\Gamma}\left(\frac{1-{\varepsilon}}{2}\right)}{ 2 (Q^2)^{3+{\varepsilon}} \, {\Gamma}(-1-2{\varepsilon})} \label{Fp_NNLO_h}$$ and agrees with the LO contribution of the hard region. So indeed all contributions to the five-point integral up to NNLO are correctly reproduced by the contributions of the three regions found by `asy2.m` after the decomposition of the integral and the change of variables. When revealing Glauber regions for a general diagram, the preresolution algorithm of `asy2.m` tries to eliminate monomials with opposite sign in the polynomial ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ by automatically separating the integration into domains and performing changes of variables. If the option `PreResolve` is enabled for `AlphaRepExpand[]`, the code warns the user if the elimination of monomials with opposite sign has not been successful, such that possibly not all regions are revealed. This is the case if some monomials of opposite sign remain in the polynomial ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ after tries to eliminate them, or if symbols with unknown signs are present in the polynomial. We are therefore convinced that `AlphaRepExpand[]`, with the preresolution enabled, either reveals all relevant regions or issues a warning. As for the threshold expansion of the previous section, `asy2.m` can also treat more complicated kinematical situations, e.g. the five-point integral depicted in Fig. \[fig:Glauber\] with $p_1 \ne p_2$ and $q_1 \ne q_2$ (retaining $p_1 \parallel p_2$ and $q_1 \parallel q_2$ such that $p_i \cdot p_j = q_i \cdot q_j = 0$ and $2p_i \cdot q_j = P_i Q_j$ with $P_i,Q_j>0$): AlphaRepExpand[{k}, {k^2 - m^2, k^2 - 2*p1*k, k^2 + 2*p2*k, k^2 - 2*q2*k, k^2 + 2*q1*k}, {p1^2 -> 0, p2^2 -> 0, p1*p2 -> 0, q1^2 -> 0, q2^2 -> 0, q1*q2 -> 0, p1*q1 -> P1*Q1/2, p1*q2 -> P1*Q2/2, p2*q1 -> P2*Q1/2, p2*q2 -> P2*Q2/2}, {m^2 -> x, P1 -> 1, P2 -> 3, Q1 -> 2, Q2 -> 3/2}, PreResolve -> True] All kinematic invariants are replaced by rational numbers in order to enable the preresolution algorithm to work. The code correctly decomposes the integral into four pieces and finds the three regions for each of them. Let us finally discuss the reason why, besides the hard region which is always present, the expansion in loop-momentum space requires two collinear regions and one Glauber region, whereas the expansion of the parametric integrals (\[Fp\_integral\_B\]) has two regions providing the leading Glauber contribution and no further collinear regions. In fact, the momentum-space expansion is also valid for loop integrals (\[5pt\_integral\]) where each propagator is raised to an arbitrary, even non-integer power. For the decomposition of the parametric integral and the change of variables, however, we have assumed the specific form (\[5pt\_integral\]) with each propagator present exactly once. In this case, the loop integrand can be expanded into partial fractions as follows: $$\begin{aligned} \lefteqn{\frac{1}{(k^2-m^2)(k^2-2p\cdot k) (k^2+2p\cdot k) (k^2-2q\cdot k) (k^2+2q \cdot k)}} \; {\nonumber}\\ & = \frac{1}{4 (m^2)^2} \left( \frac{1}{k^2-m^2} - \frac{1}{k^2} - \frac{m^2}{(k^2)^2} \right) {\nonumber}\\* &\times \left( \frac{1}{k^2-2p\cdot k} + \frac{1}{k^2+2p\cdot k} \right) \left( \frac{1}{k^2-2q\cdot k} + \frac{1}{k^2+2q\cdot k} \right) .\end{aligned}$$ Expanding this product of terms, one obtains twelve three-point integrals, which are well known. Because they only depend on $Q^2 = 2p\cdot q$, we recognize from the last two factors the structure $F(Q^2,m^2) = 2(I_+ + I_-)$, where $I_+$ and $I_-$ are related by $Q^2 \to -Q^2$ as before. The three-point integrals with the massless propagators $1/k^2$ or $1/(k^2)^2$ only have a hard region. The massive three-point integral with propagator $1/(k^2-m^2)$ is known to possess a hard and two collinear regions. Its LO and NLO hard contributions are cancelled by the massless three-point integrals, such that the uncancelled hard contributions start with $(m^2)^2/(m^2)^2 = (m^2)^0$, as for the five-point integral. The LO collinear contributions of the three-point integrals, enhanced by the $1/(m^2)^2$ prefactor, scale as $(m^2)^{-2-{\varepsilon}}$. So, in the special case when all propagator powers are equal to 1, the five-point integral reduces to a linear combination of three-point integrals revealing the same structure of regions as found from the expansion of the parametric integral (\[Fp\_integral\_B\]). This picture changes when generic propagator powers are introduced as analytic regulators, which is done in the next section. Disentangling regions via propagator powers {#sec:Glauber_l} =========================================== In the previous section we have seen different patterns of regions arising when expanding either in loop-momentum space or in parametric space. The individual contributions can be disentangled more easily when the dependence on the propagator powers is retained. Instead of (\[5pt\_integral\]), let us consider the integral $$\begin{aligned} &F(Q^2,m^2)=\int \frac{{\mbox{d}}^d k}{(k^2-m^2)^{1+{\lambda}_1} (k^2-2p\cdot k)^{1+{\lambda}_2}} {\nonumber}\\* &\quad\times \frac{1}{(k^2+2p\cdot k)^{1+{\lambda}_3} (k^2-2q\cdot k)^{1+{\lambda}_4} (k^2+2q\cdot k)^{1+{\lambda}_5}}\;, \label{5pt_integral_l}\end{aligned}$$ where the analytic regularization parameters ${\lambda}_i$ make the propagator powers different from the previous case. The asymptotic expansion in loop-momentum space yields contributions from the four regions listed in the beginning of Section \[sec:Glauber\]. The LO hard contribution still scales as $(m^2)^0$, but now the LO 1-collinear contribution scales as $(m^2)^{-1-{\varepsilon}-{\lambda}_1-{\lambda}_2-{\lambda}_3}$, the LO 2-collinear contribution as $(m^2)^{-1-{\varepsilon}-{\lambda}_1-{\lambda}_4-{\lambda}_5}$ and the LO Glauber contribution as $(m^2)^{-2-{\varepsilon}-{\lambda}_1-{\lambda}_2-{\lambda}_3-{\lambda}_4-{\lambda}_5}$. So all regions are characterized by a distinct scaling. When performing the expansion in parametric space for generic ${\lambda}_i$, we are able to disentangle the contributions from each region and match them with the regions obtained in loop-momentum space. For generic ${\lambda}_1,\ldots,{\lambda}_5$, the parametric integral corresponding to (\[Fp\_integral\]) reads $$\begin{aligned} \lefteqn{ F(Q^2,m^2)=-i\pi^{d/2} \, e^{-i\pi({\lambda}_1+\ldots+{\lambda}_5)} } \quad {\nonumber}\\ &\times \frac{{\Gamma}(3+{\varepsilon}+{\lambda}_1+\ldots+{\lambda}_5)}{{\Gamma}(1+{\lambda}_1) \cdots {\Gamma}(1+{\lambda}_5)} \idotsint {\mbox{d}}x_1\cdots{\mbox{d}}x_5 {\nonumber}\\ &\times \delta\biggl(\sum_i x_i-1\biggr) \; x_1^{{\lambda}_1} \cdots x_5^{{\lambda}_5} \; (x_1+\ldots+x_5)^{1+2{\varepsilon}+{\lambda}_1+\ldots+{\lambda}_5} {\nonumber}\\ &\times \bigl[x_1(x_1+\ldots+x_5) m^2 {\nonumber}\\* &\qquad +(x_2-x_3)(x_4-x_5)Q^2-i 0\bigr]^{-(3+{\varepsilon}+{\lambda}_1+\ldots+{\lambda}_5)} \;. \label{Fp_integral_l}\end{aligned}$$ The change of variables performed in Section \[sec:Glauber\], e.g. $x_2 = x'_2 + x'_3/2$, $x_3 = x'_3/2$ for $x_2 \ge x_3$, is complicated by the presence of the factors $x_2^{{\lambda}_2} x_3^{{\lambda}_3} x_4^{{\lambda}_4} x_5^{{\lambda}_5}$, where parts of the monomials will change into polynomials. If we want to keep the simple form $F(Q^2,m^2) = 2(I_+ + I_-)$ from the previous section, then we have to require ${\lambda}_3 = {\lambda}_2$ and ${\lambda}_5={\lambda}_4$. Under this restriction of the parameters ${\lambda}_i$, the parametric integrals can be written as $$\begin{aligned} I_\pm&=-i\pi^{d/2} \, \frac{e^{-i\pi({\lambda}_1+2{\lambda}_2+2{\lambda}_4)}\, {\Gamma}(3+{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4)}{ 4^{1+{\lambda}_2+{\lambda}_4} \, {\Gamma}(1+{\lambda}_1) {\Gamma}^2(1+{\lambda}_2) {\Gamma}^2(1+{\lambda}_4)} {\nonumber}\\ &\times \idotsint{\mbox{d}}x_1\cdots{\mbox{d}}x_5 \; \delta\left(x_1-1\right) \; x_1^{{\lambda}_1} (2x_2+x_3)^{{\lambda}_2} x_3^{{\lambda}_2} {\nonumber}\\ &\times \frac{(2x_4+x_5)^{{\lambda}_4} x_5^{{\lambda}_4} \; (x_1+\ldots+x_5)^{1+2{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4} }{\bigl[x_1(x_1+\ldots+x_5) m^2 \pm x_2 x_4 Q^2-i 0\bigr]^{3+{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4} } \;. \label{Fp_integral_l_A}\end{aligned}$$ In order to find the regions for the asymptotic expansion of (\[Fp\_integral\_l\_A\]), we have to provide the additional polynomial factors $(2x_2+x_3)$ and $(2x_4+x_5)$ to `WilsonExpand[]`. We can do this by multiplying the new polynomials to the second argument of the command: WilsonExpand[x1*(x1 + x2 + x3 + x4 + x5)*m^2 + x2*x4*Q^2, (x1 + x2 + x3 + x4 + x5) * (2*x2 + x3) * (2*x4 + x5), {x1, x2, x3, x4, x5}, {Q^2 -> 1, m^2 -> x}, Delta -> True] The output of this command reads: {{0, 0, 0, 1, 0}, {0, 0, 0, 1, 1}, {0, 1, 0, 0, 0}, {0, 1, 1, 0, 0}, {0, 0, 0, 0, 0}} In addition to the regions present in the analysis of Section \[sec:Glauber\], we retrieve the two collinear regions with scalings $\{0,0,0,1,1\}$ and $\{0,1,1,0,0\}$. The updated code `asy2.m` is capable of taking into account generic propagator powers automatically when told through the additional option `GenericPowers`: AlphaRepExpand[{k}, {k^2 - m^2, k^2 - 2*p*k, k^2 + 2*p*k, k^2 - 2*q*k, k^2 + 2*q*k}, {p^2 -> 0, q^2 -> 0, p*q -> Q^2/2}, {Q -> 1, m^2 -> x}, PreResolve -> True, GenericPowers -> True] When the option `GenericPowers` is enabled, the polynomial ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ obtained from the loop integral is multiplied by the product of all Feynman parameters, $x_1 x_2 \cdots$, before the changes of variables are performed. Some of these additional factors then turn into polynomials through replacements in the preresolution algorithm, while others remain monomials and are therefore irrelevant for the analysis of `asy2.m`. The call of `AlphaRepExpand[]` stated above correctly yields all five regions for each of the variable transformations. For the evaluation of the Glauber contributions with scalings $\{0,0,0,1,0\}$ and $\{0,1,0,0,0\}$, an additional analytic regularization is needed, and we choose to multiply the integrand of (\[Fp\_integral\_l\_A\]) by $x_2^{{\delta}_2} x_4^{{\delta}_4}$. (The parameters ${\delta}_3,{\delta}_5$ from Section \[sec:Glauber\] are not needed here due to the presence of ${\lambda}_2,{\lambda}_4$.) The two Glauber contributions are individually singular in the limit ${\delta}_2,{\delta}_4 \to 0$, but this singularity cancels in the sum of the two contributions. The leading contribution to the integral $F(Q^2,m^2)$ originates from the sum of the LO Glauber contributions. It reads $$\begin{aligned} &-i\pi^{d/2} \, \frac{i \, e^{-i\pi({\lambda}_1+2{\lambda}_2+2{\lambda}_4)} \, {\Gamma}\left(\tfrac{1}{2}+{\lambda}_2\right) {\Gamma}\left(\tfrac{1}{2}+{\lambda}_4\right)}{ 2 Q^2 \, (m^2)^{2+{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4}} {\nonumber}\\* &\quad\times \frac{{\Gamma}(2+{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4) {\Gamma}(-1-{\varepsilon}-2{\lambda}_2-2{\lambda}_4)}{ {\Gamma}(1+{\lambda}_1) {\Gamma}(1+{\lambda}_2) {\Gamma}(1+{\lambda}_4) {\Gamma}(1-{\varepsilon})} \;. \label{Fp_l_LO}\end{aligned}$$ This agrees with the LO contribution from the one Glauber region in the momentum-space expansion of (\[5pt\_integral\_l\]) for ${\lambda}_3 = {\lambda}_2$ and ${\lambda}_5={\lambda}_4$. Here, in the expansion of the parametric integrals, we have two regions producing the Glauber contribution. This is possible because the contributions from both regions have the same scaling (for ${\delta}_2={\delta}_4=0$), starting with $(m^2)^{-2-{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4}$. For ${\lambda}_i=0$, the result (\[Fp\_l\_LO\]) reproduces (\[Fp\_LO\]). Among the NLO contributions to $F(Q^2,m^2)$, we expect NLO Glauber contributions and LO collinear contributions. However, the NLO Glauber contributions vanish exactly for general ${\lambda}_1$, ${\lambda}_2={\lambda}_3$ and ${\lambda}_4={\lambda}_5$, due to non-trivial cancellations between the pieces which contribute to the NLO expansion of (\[Fp\_integral\_l\_A\]) for either of the scalings $\{0,0,0,1,0\}$ and $\{0,1,0,0,0\}$. The same happens in loop-momentum space, where the NLO Glauber contribution is proportional to $({\lambda}_3-{\lambda}_2)({\lambda}_5-{\lambda}_4)$, thus vanishing in the case considered here. So the NLO contribution to $F(Q^2,m^2)$ is made up entirely from the LO collinear contributions. The 1-collinear region provides $$\begin{aligned} &-i\pi^{d/2} \, \frac{e^{-i\pi({\lambda}_1+2{\lambda}_2+2{\lambda}_4)} \, e^{i\pi{\lambda}_4} \, {\Gamma}({\lambda}_2-{\lambda}_4) {\Gamma}(1-2{\lambda}_4)}{ 2 (Q^2)^{2+2{\lambda}_4} \, (m^2)^{1+{\varepsilon}+{\lambda}_1+2{\lambda}_2} \, {\Gamma}(1+{\lambda}_1) {\Gamma}(1+{\lambda}_2)} {\nonumber}\\* &\quad\times \frac{{\Gamma}(1+{\varepsilon}+{\lambda}_1+2{\lambda}_2) {\Gamma}(-{\varepsilon}-2{\lambda}_2)}{{\Gamma}(1-{\lambda}_4) {\Gamma}(-{\varepsilon}-2{\lambda}_4)} \; \frac{1}{1+2{\lambda}_4} \;, \label{Fp_l_NLO_1c}\end{aligned}$$ in agreement with the momentum-space expansion. The 2-collinear contribution is obtained from this by exchanging ${\lambda}_2 \leftrightarrow {\lambda}_4$. Adding the two collinear contributions together and performing the limit ${\lambda}_1,{\lambda}_2,{\lambda}_4 \to 0$, the result (\[Fp\_NLO\]) is reproduced. Considering finally the NNLO contributions to $F(Q^2,m^2)$, we expect NNLO Glauber contributions, NLO collinear contributions and a LO hard contribution. But here the NLO collinear contributions vanish exactly for ${\lambda}_3 = {\lambda}_2$ and ${\lambda}_5={\lambda}_4$, both in loop-momentum space and when expanding the parametric integrals. So we are left with the NNLO Glauber contributions yielding $$\begin{aligned} &-i\pi^{d/2} \, \frac{i \, e^{-i\pi({\lambda}_1+2{\lambda}_2+2{\lambda}_4)} \, {\Gamma}\left({\lambda}_2-\tfrac{1}{2}\right) {\Gamma}\left({\lambda}_4-\tfrac{1}{2}\right)}{ 16 (Q^2)^3 \, (m^2)^{{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4}} {\nonumber}\\* &\quad\times \frac{{\Gamma}({\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4) {\Gamma}(1-{\varepsilon}-2{\lambda}_2-2{\lambda}_4)}{ {\Gamma}(1+{\lambda}_1) {\Gamma}(1+{\lambda}_2) {\Gamma}(1+{\lambda}_4) {\Gamma}(-1-{\varepsilon})} \label{Fp_l_NNLO_g}\end{aligned}$$ and the LO hard contribution, $$\begin{aligned} &i\pi^{d/2} \, \frac{i \, e^{-i\pi({\lambda}_1+2{\lambda}_2+2{\lambda}_4)} \, e^{i\pi({\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4)/2}}{ 2\sqrt{\pi} \, (2Q^2)^{3+{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4} \, {\Gamma}(1+{\lambda}_2) {\Gamma}(1+{\lambda}_4)} {\nonumber}\\ &\quad\times \frac{{\Gamma}\left(\tfrac{-1-{\varepsilon}-{\lambda}_1-2{\lambda}_2}{2}\right) {\Gamma}\left(\tfrac{-1-{\varepsilon}-{\lambda}_1-2{\lambda}_4}{2}\right)}{ {\Gamma}(-1-2{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4)} {\nonumber}\\ &\quad\times {\Gamma}\left(\tfrac{-2-{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4}{2}\right) {\Gamma}\left(\tfrac{3+{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4}{2}\right) \,, \label{Fp_l_NNLO_h}\end{aligned}$$ both consistent between the expansions of the loop integral and of the parametric integrals. Setting all ${\lambda}_i = 0$ in (\[Fp\_l\_NNLO\_g\]) and (\[Fp\_l\_NNLO\_h\]), we recover the results from (\[Fp\_NNLO\_g\]) and (\[Fp\_NNLO\_h\]), respectively. We may also evaluate the parametric integral (\[Fp\_integral\_l\_A\]) in terms of a onefold Mellin–Barnes representation: $$\begin{aligned} I_{\pm} &= -i\pi^{d/2} \, \frac{e^{-i\pi({\lambda}_1+2{\lambda}_2+2{\lambda}_4)}}{ 4\pi \, {\Gamma}(1+{\lambda}_1) {\Gamma}(1+{\lambda}_2) {\Gamma}(1+{\lambda}_4)} {\nonumber}\\ &\times \frac{1}{2\pi i} \int {\mbox{d}}z \, (m^2)^z (\pm 4 Q^2-i0)^{-3-{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4-z} {\nonumber}\\ &\times \frac{{\Gamma}(1+{\lambda}_1+z) {\Gamma}(3+{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4+z)}{ {\Gamma}(-1-2{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4-z)} {\nonumber}\\ &\times {\Gamma}(-z) {\Gamma}\left(\tfrac{-1-{\varepsilon}-{\lambda}_1-2{\lambda}_2-z}{2}\right) {\Gamma}\left(\tfrac{-1-{\varepsilon}-{\lambda}_1-2{\lambda}_4-z}{2}\right) {\nonumber}\\ &\times {\Gamma}^2\left(\tfrac{-2-{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4-z}{2}\right) \;. \label{Fp_integral_l_MB}\end{aligned}$$ The relevant regions can easily be determined from the gamma functions in the last two lines. In particular, the squared gamma function indicates that the expansion of the parametric integrals $I_\pm$ requires *two* regions for the Glauber contribution, both scaling as $(m^2)^{-2-{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4}$ at leading order. When combining $F(Q^2,m^2) = 2(I_+ + I_-)$, one of these gamma functions is cancelled, and we obtain $$\begin{aligned} \lefteqn{ F(Q^2,m^2) = i\pi^{d/2} \, \frac{i \, e^{-i\pi({\lambda}_1+2{\lambda}_2+2{\lambda}_4)} e^{i\pi({\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4)/2}}{ 2\sqrt{\pi} \, {\Gamma}(1+{\lambda}_1) {\Gamma}(1+{\lambda}_2) {\Gamma}(1+{\lambda}_4)} } \quad {\nonumber}\\ &\times \frac{1}{2\pi i} \int {\mbox{d}}z \, (m^2)^z (2 Q^2)^{-3-{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4-z} \, e^{i\pi z/2} {\nonumber}\\ &\times \frac{{\Gamma}(1+{\lambda}_1+z) {\Gamma}\left(\tfrac{3+{\varepsilon}+{\lambda}_1+2{\lambda}_2+2{\lambda}_4+z}{2}\right)}{ {\Gamma}(-1-2{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4-z)} {\nonumber}\\ &\times {\Gamma}(-z) {\Gamma}\left(\tfrac{-1-{\varepsilon}-{\lambda}_1-2{\lambda}_2-z}{2}\right) {\Gamma}\left(\tfrac{-1-{\varepsilon}-{\lambda}_1-2{\lambda}_4-z}{2}\right) {\nonumber}\\ &\times {\Gamma}\left(\tfrac{-2-{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4-z}{2}\right) \;. \label{Fp_integral_l_MB2}\end{aligned}$$ From this representation the contributions to the asymptotic expansion in the limit $m^2/Q^2 \to 0$ can be extracted: The hard contributions stem from the residues of the poles at $z=n$, the 1- and 2-collinear contributions from $z=-1+2n-{\varepsilon}-{\lambda}_1-2{\lambda}_2$ and $z=-1+2n-{\varepsilon}-{\lambda}_1-2{\lambda}_4$, respectively, the Glauber contributions from $z=-2+2n-{\varepsilon}-{\lambda}_1-2{\lambda}_2-2{\lambda}_4$ (with $n=0,1,2,\ldots$). All LO, NLO and NNLO contributions reported in (\[Fp\_l\_LO\])–(\[Fp\_l\_NNLO\_h\]) are confirmed by the corresponding residue contributions from (\[Fp\_integral\_l\_MB2\]). In particular, the structure of the poles in (\[Fp\_integral\_l\_MB2\]) clearly shows that the Glauber region does not contribute to the NLO result, and that the collinear contributions are absent at NNLO, as obtained before. The results reported in this section show that the use of generic propagator powers helps disentangling the Glauber and collinear regions from each other, making all regions contribute in the same way to the asymptotic expansion in loop-momentum space and to the expansion of the parametric integrals. Keeping the dependence of the contributions and their scalings on the propagator powers also facilitates the identification of regions found by `asy2.m` in parametric space for a subsequent expansion at the level of the loop integration. Summary of `asy2.m` {#sec:syntax} =================== The updated version of the code, `asy2.m`, can be downloaded from the known web site [@asy_web], where further installation instructions are found. The `Mathematica` code is loaded using `<<asy2.m`. The main function `AlphaRepExpand[]` identifies all regions which contribute to the asymptotic expansion of a given loop integral: AlphaRepExpand[{k1, k2, ...}, {(k1 + p1)^2 - m1^2, (k2 + p2)^2 - m2^2, ...}, {p1^2 -> Q1, p2^2 -> Q2, p1*p2 -> Q3, ...}, {m1^2 -> x, m2^2 -> x^2, Q1 -> 1, Q2 -> 3/2, ...}, options] The first argument is the list of loop momenta. The second argument lists the denominators of the propagators. The third argument contains replacement rules for all kinematic invariants. In particular, all external momenta appearing in the denominators must be replaced here, otherwise they are not identified correctly as vectors. The fourth argument sets the scaling of the parameters by replacing all symbols with powers of the expansion parameter, labelled by the global symbol `x`, and rational numbers (integers or explicit fractions of integers). The output is a list of regions, specified by the scaling (in powers of the expansion parameter) of the Feynman parameters $x_1,x_2,\ldots$ corresponding to the propagators in the order stated in the second argument. E.g. the output {..., {0, 2, 1, ...}, ...} indicates that there is a region $\{0,2,1,\ldots\}$ specified by the scaling $x_1 \sim x^0$, $x_2 \sim x^2$, $x_3 \sim x^1$, $\ldots$ of the Feynman parameters, where $x$ is the small parameter of the problem. Possible options of `AlphaRepExpand[]` are: - `PreResolve -> True`: Try to eliminate cancellations between terms in the parametric representation by decomposing the integral and performing changes of variables. The output contains for each region a list of three entries, i.e. a region is e.g. specified by: {{x[1] -> y[1]/2, x[2] -> y[1]/2 + y[2]}, 2, {0, 1/2}} The first entry in this list is the transformation between the original Feynman parameters `x[1]`, `x[2]`, $\ldots$ and the new variables `y[1]`, `y[2]`, $\ldots$. The second entry is the Jacobian of the integral transformation (here “2”, i.e. ${\mbox{d}}y_1 {\mbox{d}}y_2 = 2 \, {\mbox{d}}x_1 {\mbox{d}}x_2$). The last entry specifies the region in the usual form with the scalings of the new variables in powers of the small parameter, here $y_1 \sim x^0$, $y_2 \sim x^{1/2}$. When the option `PreResolve` is enabled, the code warns if it fails to eliminate all possible cancellations in the parametric representation. When no warning is issued, all regions are found. Without this option, however, regions occurring at such cancellations will not be revealed. - `GenericPowers -> True`: Take into account generic (in particular non-integer) powers of the propagators, e.g. when these powers are used as analytic regulators. Without this option, the preresolution algorithm triggered by the option `PreResolve` only finds all regions for integrals with propagators raised to positive integer powers. - `Verbose -> True`: Print verbose internal information. - `Scalar -> True`: Permit more complex structures of the denominators by specifying scalar products of momenta via the function `Scalar[k,p]` instead of simple products `k*p` or `Scalar2[k]` instead of `k^2`, e.g.: AlphaRepExpand[{k}, {Scalar[k, k] - m^2, Scalar2[k - q] - m^2}, {Scalar2[q] -> qq, m^2 -> qq/4 + y}, {qq -> 1, y -> x}, PreResolve -> True, Scalar -> True] For expanding more general integrals, which need not originate from Feynman diagrams, the command `WilsonExpand[]` may be used: WilsonExpand[F, U, {x1, x2, ...}, {... -> x, ...}, options] Traditionally, the first two arguments are the polynomials ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$, respectively, from the alpha parametric representation (\[alpha-d\]) of the Feynman integral or from the Feynman parametric representation (\[alpha-d-mod\]), (\[integrand\]). But, more generally, `WilsonExpand[]` reveals regions for integrals over parameters $x_1,x_2,\ldots$, integrated from 0 to $\infty$ each, where all non-trivial polynomials of the parameters $x_i$ occurring in the integrand are specified either in the first or the second argument. The third argument of `WilsonExpand[]` is the list of integration parameters. The fourth argument specifies the scaling of all quantities with the expansion parameter `x`, in the same way as in the fourth argument of `AlphaRepExpand[]`. The output is a list of regions, specified by the scaling of the integration parameters in powers of the expansion parameter, exactly as described for the output of `AlphaRepExpand[]` (without the `PreResolve` option). Possible options of `WilsonExpand[]` are: - `Delta -> True`: Under the integral, the sum over an arbitrary non-empty subset of the integration parameters is restricted to 1 via a delta function. The specific choice of the sum in this delta function must be irrelevant for the integral, which is the case for the generalized Feynman parametric representation (\[alpha-d-mod\]). - `Verbose -> True`: Print verbose internal information. More generally, the option `Delta` of `WilsonExpand[]` works correctly for all integrals of the form $$\begin{aligned} \int_0^\infty \cdots \int_0^\infty {\mbox{d}}x_1\cdots{\mbox{d}}x_N \, \delta\left(\sum_{i=1}^N a_i x_i - 1\right) f(x_1,\ldots,x_N) \;, \label{delta_int}\end{aligned}$$ where the linear combination in the argument of the delta function has no negative and at least one positive coefficient ($a_i \ge 0 \, \forall i$ and $\exists \, a_i>0$), and where the function $f$ scales homogeneously with the parameters $x_i$ as $$\begin{aligned} f({\lambda}x_1,\ldots,{\lambda}x_N) = {\lambda}^{-N} f(x_1,\ldots,x_N) \; \forall {\lambda}>0 \;, \label{delta_int_hom}\end{aligned}$$ the degree of homogeneity being equal to minus the number of integration parameters. It can be shown that for such integrals (\[delta\_int\]) the specific choice of the coefficients $a_i$ is irrelevant.[^5] The integrand (\[integrand\]) of the Feynman parametric representation (\[alpha-d-mod\]) and the integrands of all parametric representations of Feynman integrals used in this paper (without the additional analytic regularization factors $x_i^{{\delta}_i}$) fulfill the homogeneity condition (\[delta\_int\_hom\]), so they do not depend on the specific choice for the arguments of their delta functions. Conclusion {#sec:conclusion} ========== We have presented an algorithm for identifying all regions which are relevant for the asymptotic expansion of a given loop integral at the level of its parametric representation. In contrast to the previous version `asy.m` of the code, also potential regions and Glauber regions are found now. The necessary decompositions and variable transformations of the integral are automated by the updated `Mathematica` code `asy2.m` [@asy_web]. When the command `AlphaRepExpand[]` is used with its option `PreResolve` enabled, we are convinced that it either reveals all relevant regions or issues a warning. In particular, regions corresponding to cancellations between large positive and negative terms in the parametric representation of the loop integrals (such as potential and Glauber regions) will now be found. Let us emphasize that to prove expansion by regions at least for some specific limit typical of Minkowski space is a natural mathematical problem. Perhaps, this problem is not specifically related to Feynman integrals. Let us present an example of a one-dimensional parametric integral, without any relevance to Feynman integrals, and show that expansion by regions works successfully. To do this, we will use [asy2.m]{}. Let us consider the integral $$F(t)=\int_0^\infty(t + u + u^2)^{\lambda}{\mbox{d}}u\;, \label{toy_example}$$ with ${\lambda}$ a complex parameter, in the limit $t\to 0$. We assume that ${\lambda}$ is in the domain Re${\lambda}<-1/2$ in order to have an absolute convergence of the integral which then can be continued analytically to the whole complex plane as an analytic function of ${\lambda}$. Running WilsonExpand[t + u + u^2, 1, {u}, {t -> x}] we obtain the two regions `{{1}, {0}}`. The leading-order terms from each region can be evaluated analytically in terms of gamma functions at general ${\lambda}$, with the results $$\frac{t^{\lambda +1} \Gamma (-\lambda -1)}{\Gamma (-\lambda )}$$ and $$\frac{\Gamma (-2 \lambda -1) \Gamma (\lambda +1)}{\Gamma (-\lambda )} \;.$$ They can be checked easily by deriving the onefold Mellin–Barnes representation $$\begin{aligned} F(t)&=\frac{1}{2\pi i}\frac{1}{\Gamma (-\lambda )} {\nonumber}\\* &\times \int \Gamma (-z) \Gamma (\lambda -z+1) \Gamma (-2 \lambda +2z-1) \, t^z \, {\mbox{d}}z\end{aligned}$$ and evaluating the first terms of the asymptotic expansion in the limit $t\to 0$ by shifting the contour to the right and taking residues at the poles of the two gamma functions in the integrand. This work is supported by the Deutsche Forschungsgemeinschaft Sonderforschungsbereich/Transregio 9 “Computergestützte Theoretische Teilchenphysik”. The work of A.S. and V.S. is also supported by the Russian Foundation for Basic Research through grant 11-02-01196. The authors thank M. Beneke and A. Pak for helpful discussions. [99]{} K.G. Chetyrkin, Theor. Math. Phys.  [**75**]{}, 346 (1988) \[Teor. Mat. Fiz.  [**75**]{}, 26 (1988)\] K.G. Chetyrkin, Theor. Math. Phys.  [**76**]{}, 809 (1988) \[Teor. Mat. Fiz.  [**76**]{}, 207 (1988)\] S.G. Gorishny, Nucl. Phys. B [**319**]{}, 633 (1989) V.A. Smirnov, Commun. Math. Phys.  [**134**]{}, 109 (1990) V.A. Smirnov, Mod. Phys. Lett. A [**10**]{}, 1485 (1995) \[hep-th/9412063\] V.A. Smirnov, [*Applied Asymptotic Expansions in Momenta and Masses*]{}, Springer Tracts in Modern Physics 177 (Springer, Berlin, 2002) T. Seidensticker, hep-ph/9905298 R. Harlander, T. Seidensticker, M. Steinhauser, Phys. Lett. B [**426**]{}, 125 (1998) \[hep-ph/9712228\] M. Beneke, V.A. Smirnov, Nucl. Phys. B [**522**]{}, 321 (1998) \[hep-ph/9711391\] V.A. Smirnov, E.R. Rakhmetov, Theor. Math. Phys.  [**120**]{}, 870 (1999) \[Teor. Mat. Fiz.  [**120**]{}, 64 (1999)\] \[hep-ph/9812529\] V.A. Smirnov, Phys. Lett. B [**465**]{}, 226 (1999) \[hep-ph/9907471\] B. Jantzen, JHEP [**12**]{}, 076 (2011) \[arXiv:1111.2589 \[hep-ph\]\] A.V. Smirnov, V.A. Smirnov, M. Tentyukov, Comput. Phys. Commun.  [**182**]{}, 790 (2011) \[arXiv:0912.0158 \[hep-ph\]\] A.V. Smirnov, M.N. Tentyukov, Comput. Phys. Commun.  [**180**]{}, 735 (2009) \[arXiv:0807.4129 \[hep-ph\]\] A. Pak, A. Smirnov, Eur. Phys. J. C [**71**]{}, 1626 (2011) \[arXiv:1011.4863 \[hep-ph\]\] A.V. Smirnov, <http://science.sander.su/Tools-UF.htm> V.A. Smirnov, [*Feynman Integral Calculus*]{} (Springer, Berlin, 2006) [](http://www-ttp.particle.uni-karlsruhe.de/~asmirnov/Tools-Regions.htm) [^1]: The function `UF[]` from `UF.m` [@UF.m] is called with three arguments: The list of loop momenta, the list of denominators of the propagators and a list of replacement rules for all kinematic invariants. The output is a list with the following entries: the function ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$, the function ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ and the number of loops. In order to obtain ${{\text{\usefont{OMS}{cmsy}{m}{n}U}}}$ and ${{\text{\usefont{OMS}{cmsy}{m}{n}F}}}$ with the correct sign, denominators have to be specified with the opposite sign as in (\[denom-d1\]), i.e. corresponding to a negative imaginary part $-i0$: `UF[{k1,k2,...}, {-E1,-E2,...}, {}]`. [^2]: See e.g. the discussion in Section 3.4 of [@Smirnov:2006ry]. [^3]: See also (\[delta\_int\]), (\[delta\_int\_hom\]) and Footnote \[fn:delta\_scaling\] (p. ) for a general proof. [^4]: The name of the command refers to its application to parametric integrals contributing to Wilson loops. [^5]: \[fn:delta\_scaling\] To see this, multiply the integrand of (\[delta\_int\]) by 1 in the form $x_j \int_0^\infty {\mbox{d}}t \, e^{-t x_j}$, where $x_j$ is any of the integration parameters. Then, inside the $t$-integration, transform the integration variables as $x_i \to x_i/t$, $i = 1,\ldots,N$, and use the homogeneity relation (\[delta\_int\_hom\]) with ${\lambda}=1/t$. Finally evaluate the $t$-integration first, yielding $\int_0^\infty {\mbox{d}}t \, \delta(\sum_{i=1}^N a_i x_i - t) = 1$, independent of the coefficients $a_i$, as long as the linear combination is positive. The integral (\[delta\_int\]) is given by $\int_0^\infty \cdots \int_0^\infty {\mbox{d}}x_1\cdots{\mbox{d}}x_N \, f(x_1,\ldots,x_N) \, x_j \, e^{-x_j}$.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Sidon sequences and their generalizations have found during the years and especially recently various applications in coding theory. One of the most important applications of these sequences is in the connection of synchronization patterns. A few constructions of two-dimensional synchronization patterns are based on these sequences. In this paper we present sufficient conditions that a two-dimensional synchronization pattern can be transformed into a Sidon sequence. We also present a new construction for Sidon sequences over an alphabet of size $q(q-1)$, where $q$ is a power of a prime.' author: - title: | Sidon Sequences and Doubly Periodic\ Two-Dimensional Synchronization Patterns --- Introduction {#sec:introduction} ============ Let ${{\cal A}}$ be an abelian group and let ${{\cal D}}= \{ a_1 , a_2 , \ldots , a_m \} \subseteq {{\cal A}}$ be a subset of $m$ distinct elements of ${{\cal A}}$. ${{\cal D}}$ is a *Sidon sequence* (or a $B_2$-sequence) over ${{\cal A}}$ if all the sums $a_{i_1} + a_{i_2}$ with $1 \leq i_1 \leq i_2 \leq m$ are distinct (if $i_1 < i_2$ in the definition the sequence is called a *weak Sidon sequences*). Sidon sequences have found many applications in coding and communication. For example, weak Sidon sequences are used for construction of constant weight codes with minimum Hamming distance 6 [@BSSS], and constructions of location-correcting codes [@RoSe96]. Sidon sequences were used in constructions of two-dimensional synchronization patterns [@BEMP1; @Etz11]. There is a generalization to $B_h$ sequences (all sums of $h$ elements are distinct) and they applied for example in multihop paths related to wireless sensor networds [@BEMP2] and error-correcting codes for rank modulation [@BaMa10]. A comprehensive survey on $B_2$-sequences and their generalizations was given by O’Bryant [@Bry04]. Even so in a Sidon sequence all sums of pairs of elements from ${{\cal D}}$ (not necessarily distinct elements) are distinct there is a trivial connection to a set in which all differences of ordered pairs of elements are distinct. \[thm:diff\] A subset ${{\cal D}}= \{ a_1 , a_2 , \ldots , a_m \} \subseteq {{\cal A}}$ is a Sidon sequence over ${{\cal A}}$ if and only if all the differences $a_{i_1} - a_{i_2}$ with $1 \leq i_1 \neq i_2 \leq m$ are distinct in ${{\cal A}}$. A Sidon sequence with $m$ elements over an abelian group with $n$ elements is called *optimal* if all Sidon sequences over an abelian group with $n$ elements have at most $m$ elements. In view of Theorem \[thm:diff\] bounds on the size of a Sidon sequence (on the number of elements $m$) can be derived by considering difference and not sums. This is important since the number of distinct sums is $\binom{m}{2} +m = \frac{m^2 +m}{2}$ while the number of distinct differences is considerably higher, $m(m-1) = m^2 -m$. This yields a better upper bound on $m$. A Sidon sequence ${{\cal D}}$ is a set of $m$ elements. If the abelian group is ${{\mathbb{Z}}}_n$ then ${{\cal D}}$ can be represented as a binary cyclic sequence $s=[s_0 s_1 , \cdots , s_{n-1} ]$, where $s_i =1$ if $i \in {{\cal D}}$. One-dimensional synchronization patterns were first introduced by Babcock in connection with radio interference [@Bab]. Other applications are discussed in details in [@BlGo77] and some more are given in [@ASU; @LaSa88]. The two-dimensional applications and related structures were first introduced in [@GoTa82] and discussed in many papers, e.g. [@GoTa84; @Rob85; @Games87; @BlTi88; @Rob97]. Recent new application in keys predistribution for wireless sensor networks [@BEMP] led to new related two-dimensional problems concerning these patterns [@BEMP1; @BEMP2]. Difference pattern and Sidon sequences have an important role in the construction of synchronization patterns. Some of the applications of Sidon sequence is due to the difference properties implied by Theorem \[thm:diff\]. This property is also the basis of the applications to two-dimensional synchronization patterns. There are various papers, e.g. [@BEMP1; @Etz11; @Rob97] in which an one-dimensional sequence (as a Sidon sequence or a ruler) is transformed into a two-dimensional synchronization pattern. The main goal of this paper is to establish the inverse transformation, in which a two-dimensional synchronization pattern is transformed into a Sidon sequence, which is a one-dimensional sequence. The rest of this paper is organized as follows. In Section \[sec:period\] we define what is a period in a two-dimensional array and as a result we obtain a definition for a cyclic two-dimensional array. In Section \[sec:DDCs\] we discuss various types of two-dimensional synchronization patterns. In particular we discuss periodic two-dimensional synchronization patterns. In Section \[sec:folding\] we present two operations, namely, folding and unfolding. Folding generates a two-dimensional array from an one-dimensional sequence. Unfolding is the inverse operation and it generates an one-dimensional sequence from a two-dimensional array. In particular we will prove that these operations relate periodic sequences to periodic two-dimensional arrays and vice-versa. Moreover, they relate an one-dimensional synchronization sequence to a two-dimensional synchronization array and vice-versa, if the two-dimensional array is periodic. As a consequence we obtain the main result of the paper that two-dimensional periodic synchronization arrays which can be unfolded are equivalent to Sidon sequences over ${{\mathbb{Z}}}_n$, where $n$ is the size of one period in the array. In Section \[sec:new\] we present a construction of optimal Sidon sequences with $q-1$ elements over a group with $q(q-1)$ elements, where $q$ is a power of a prime. This generalizes a similar result where $q$ is a prime. Section \[sec:conclude\] contains conclusions and problems for further research. Periodicity of Two-Dimensional Arrays {#sec:period} ===================================== periodic sequences and arrays ----------------------------- It is very simple to define the periodicity for one-dimensional sequences. An infinite sequence $S= \ldots s_{-1} , s_0 , s_1 , s_2 , \ldots$ is periodic if there exists an integer $\pi$ such that $s_{i+\pi} = s_i$ for each $i \in {{\mathbb{Z}}}$. If $\pi$ is the smallest integer for which the sequence has this property then we say that $\pi$ is the *period* of the sequence and write the sequence as $[s_0 , s_1 ,\ldots , s_{\pi -1}]$, and say that the sequences is a *cyclic sequence* or a *cycle*. It is well known that \[thm:period\_one\] If $\pi$ is the period of a sequence $S$, and there exists an integer $\rho$ such that $s_{i+\rho} = s_i$ for each $i \in {{\mathbb{Z}}}$, then $\pi$ divides $\rho$. Usually, an infinite two-dimensional array ${{\cal A}}$ is said to be doubly periodic if there exists two integers $\kappa$ and $\eta$ such that for each $i,j \in {{\mathbb{Z}}}$ we have ${{\cal A}}(i+\kappa ,j)={{\cal A}}(i,j+\eta)={{\cal A}}(i,j)$. But, it appears that this definition is too restricted. A generalized definition, which give more information, is as follows. An infinite two-dimensional array $A$ is *doubly periodic* if there exists two linearly independent integer vectors $(\pi_1 , \pi_2 )$ and $(\xi_1 , \xi_2 )$ such that each $i,j \in {{\mathbb{Z}}}$ satisfy ${{\cal A}}(i+\pi_1 ,j+\pi_2 )={{\cal A}}(i+\xi_1 ,j+\xi_2 )={{\cal A}}(i,j)$. How we can define the smallest vectors with this property? What is the period of the array? and what is a cyclic two-dimensional array? These questions will be answered after two necessary definitions, of tiling and lattices, will be presented. Tiling ------ Tiling is one of the most basic concepts in combinatorics. We say that a two-dimensional shape ${{\cal S}}$ tiles the two-dimensional square grid ${{\mathbb{Z}}}^2$ if disjoint copies of ${{\cal S}}$ cover ${{\mathbb{Z}}}^2$. This cover of ${{\mathbb{Z}}}^2$ with disjoint copies of ${{\cal S}}$ is called a [*tiling*]{} of ${{\mathbb{Z}}}^2$ with ${{\cal S}}$. For each shape ${{\cal S}}$, in the tiling, we distinguish one of the points of ${{\cal S}}$ to be the [*center*]{} of ${{\cal S}}$. Each copy of ${{\cal S}}$ in a tiling has the center in the same related point. The set ${{\cal T}}$ of centers in a tiling defines the tiling, and hence the tiling is denoted by the pair $({{\cal T}},{{\cal S}})$. Given a tiling $({{\cal T}},{{\cal S}})$ and a grid point $(i_1,i_2)$ we denote by $c(i_1,i_2)$ the center of the copy of ${{\cal S}}$, ${{\cal S}}'$, for which $(i_1,i_2) \in {{\cal S}}'$. We will also assume that the origin is a center of a copy of ${{\cal S}}$. The first lemma given in [@Etz11] can be easily verified. \[lem:center\] For a given tiling $({{\cal T}},{{\cal S}})$ and a point $(i_1,i_2)$ the point $(i_1,i_2)-c(i_1,i_2)$ belongs to the shape ${{\cal S}}$ whose center is in the origin. Lattices and Lattice Tiling --------------------------- One of the most common types of tiling is a [*lattice tiling*]{}. A two-dimensional [*lattice*]{} $\Lambda$ is a discrete, additive subgroup of the real two-dimensional space ${{\mathbb{R}}}^2$. W.l.o.g., we can assume that $$\label{eq:lattice_def} \Lambda = \{ u_1 v_1 + u_2v_2 ~:~ u_1, u_2 \in {{\mathbb{Z}}}\}$$ where $v_1, ~v_2$ are two linearly independent vectors in ${{\mathbb{R}}}^2$. A lattice $\Lambda$ defined by (\[eq:lattice\_def\]) is a sublattice of ${{\mathbb{Z}}}^2$ if and only if $\{ v_1,v_2\} \subset {{\mathbb{Z}}}^2$. We will be interested solely in sublattices of ${{\mathbb{Z}}}^2$. The vectors $v_1,v_2$ are called [*basis*]{} for $\Lambda \subseteq {{\mathbb{Z}}}^2$, and the $2 \times 2$ matrix $${\bf G}=\left[\begin{array}{cc} v_{11} & v_{12} \\ v_{21} & v_{22} \end{array}\right]$$ having these vectors as its rows is said to be the [*generator matrix*]{} for $\Lambda$. Note, that it is always possible to use a generator matrix ${\bf G}$ in which all the four entries are nonzeroes. It is also always possible to have in ${\bf G}$ exactly one [*zero*]{} entry. The [*volume*]{} of a lattice $\Lambda$, denoted $V( \Lambda )$, is inversely proportional to the number of lattice points per unit volume. More precisely, $V( \Lambda )$ may be defined as the volume of the [*fundamental parallelogram*]{} $\Pi(\Lambda)$ in ${{\mathbb{R}}}^2$, which is given by $$\Pi(\Lambda) {\mbox{$\stackrel{\rm def}{=}$}}\ \{ \xi_1 v_1 + \xi_2 v_2 ~:~ 0 \leq \xi_i < 1, ~ ,i=1,2 \}$$ There is a simple expression for the volume of $\Lambda$, namely, $V(\Lambda)=| \det {\bf G} |$. We say that $\Lambda$ induces a [*lattice tiling*]{} of ${{\cal S}}$ if the lattice points can be taken as the set ${{\cal T}}$ to form a tiling $({{\cal T}},{{\cal S}})$. Cyclic Arrays and Periods ------------------------- We are now in a position to define the period of a doubly periodic array and to define cyclic two-dimensional arrays. Let ${{\cal A}}$ be a doubly periodic two-dimensional array. Let $(\pi_1 , \pi_2 )$ and $(\xi_1 , \xi_2 )$ two linearly independent vectors such that each $i,j \in {{\mathbb{Z}}}$ satisfy ${{\cal A}}(i+\pi_1 ,j+\pi_2 )={{\cal A}}(i+\xi_1 ,j+\xi_2 )={{\cal A}}(i,j)$. Let $s$ be the volume of the lattice formed from $(\pi_1 , \pi_2 )$ and $(\xi_1 , \xi_2 )$. Let $(\pi_3 , \pi_4 )$ and $(\xi_3 , \xi_4 )$ be two linearly independent vectors for which, each $i,j \in {{\mathbb{Z}}}$ satisfy ${{\cal A}}(i+\pi_3 ,j+\pi_4 )={{\cal A}}(i+\xi_3 ,j+\xi_4 )={{\cal A}}(i,j)$. Let $s'$ be the volume of the lattice formed from $(\pi_3 , \pi_4 )$ and $(\xi_3 , \xi_4 )$. If $s' \geq s$ for each such pair of linearly independent vectors then we say that $\{ (\pi_1 , \pi_2 ) , (\xi_1 , \xi_2 ) \}$ is the *period* of ${{\cal A}}$ and $s$ is the *volume* of ${{\cal A}}$. The period in the one-dimensional case has the role of the period and the volume in the two-dimensional case. Clearly, the period of a two-dimensional array is not unique. The volume of the array is unique and can be calculated from the given period. We have a theorem in the two-dimensional case which is akin to Theorem \[thm:period\_one\]. Let ${{\cal A}}$ be a doubly periodic array with period $\{ (\pi_1 , \pi_2 ) , (\xi_1 , \xi_2 ) \}$. Let $(\pi_3 , \pi_4 )$ and $(\xi_3 , \xi_4 )$ be two linearly independent vectors for which, each $i,j \in {{\mathbb{Z}}}$ satisfy ${{\cal A}}(i+\pi_3 ,j+\pi_4 )={{\cal A}}(i+\xi_3 ,j+\xi_4 )={{\cal A}}(i,j)$. Let $s'$ be the volume of the lattice formed from $(\pi_3 , \pi_4 )$ and $(\xi_3 , \xi_4 )$. If $s$ is the volume of the lattice formed from $(\pi_1 , \pi_2 )$ and $(\xi_1 , \xi_2 )$ then $s$ divides $s'$. A shape ${{\cal S}}$ will be called *cyclic* if there is a lattice tiling $\Lambda$ for ${{\cal S}}$. In a cyclic sequence the order of the elements, in the sequence, is obvious. It is less obvious for a two-dimensional shape. We will discuss this order in Section \[sec:folding\]. Two-Dimensional Synchronization Arrays {#sec:DDCs} ====================================== Several types of two-dimensional synchronization arrays are defined in the literature. We start we a general definition which was given in [@BEMP1; @BEMP2]. Let ${{\cal S}}$ be a given shape, on the square grid, with $m$ dots on grid points. ${{\cal S}}$ is called a *distinct difference configuration* (DDC) if the $\binom{m}{2}$ lines connecting dots are distinct either in their length or in their slope. Several types of DDCs were defined in the literature. The main focus of research which was done on this topic is related to Costas arrays. A [*Costas array*]{} is an $m \times m$ permutation array having exactly one dot in each row and each column. Some results on Costas arrays are given in [@GoTa82; @GoTa84; @Gol84; @Dra06; @GoGo07]. We now present a definition for a doubly periodic DDC. A *doubly periodic ${{\cal S}}$-DDC* is a doubly periodic two-dimensional array ${{\cal A}}$ with period $\{ (\pi_1 , \pi_2 ) , (\xi_1 , \xi_2 ) \}$ such that the following three properties are satisfied. - The lattice formed by $(\pi_1 , \pi_2 )$ and $(\xi_1 , \xi_2 )$ is a lattice tiling for ${{\cal S}}$. - Each copy of ${{\cal S}}$ on the two-dimensional arrays ${{\cal A}}$ is a DDC. - In each two copies of ${{\cal S}}$ in the tiling, the positions of the dots are the same. Doubly periodic DDCs and in particular Costas array were considered in the past, e. g. [@Etz11; @MGC97; @MoGo06]. There are two essential constructions for Costas arrays, both of them form doubly periodic DDCs. The first construction is due to Welch and the second Construction is due to Golomb (with a variant of Lempel) [@GoTa82; @GoTa84; @Gol84]. We will present both of them in their doubly periodic version. [**The periodic Welch Construction:**]{} Let $\alpha$ be a primitive root modulo a prime $p$ and let ${{\cal A}}$ be the square grid. For any integers $i$ and $j$, there is a dot in ${{\cal A}}(i,j)$ if and only if $\alpha^i \equiv j \bmod p$. Let ${{\cal A}}$ be the array of dots from the Periodic Welch Construction. Then ${{\cal A}}$ is a doubly periodic ${{\cal S}}$-DDC with period $\{ (0,p),(p-1,0) \}$ and ${{\cal S}}$ is a $p \times (p-1)$ rectangle. [**The periodic Golomb Construction:**]{} Let $\alpha$ and $\beta$ be two primitive elements in GF($q$), where $q$ is a prime power. For any integers $i$ and $j$, there is a dot in ${{\cal A}}(i,j)$ if and only if $\alpha^i + \beta^j =1$. \[thm:periodicGolomb\] Let ${{\cal A}}$ be the array of dots from the Periodic Golomb Construction. Then ${{\cal A}}$ is a doubly periodic ${{\cal S}}$-DDC with period $\{ (0,q-1),(q-1,0) \}$ and ${{\cal S}}$ is a $(q-1) \times (q-1)$ square. There are many important questions concerning Costas arrays. A few of them are related to the periodicity of the arrays. In particular we have the following two question: 1. Is the Welch construction generates all singly periodic Costas arrays? where a singly periodic Costas array of order $n$ is an $n \times \infty$ array in which each $n \times n$ sub-array is a Costas array. Welch Construction has this property for $n=p-1$. 2. Are there more constructions for Costas arrays with periodicity property? The conjecture is NO for both questions. Some evidence that this conjecture is true is given in [@EGT89]. In fact it should be said that is very likely that most if not all Costas arrays are known since they derived from the known constructions [@DIR10]. In what follows we will throw more evidence for the difficulty to produce new doubly periodic ${{\cal S}}$-DDCs with many dots, different from those constructed by folding [@Etz11]. Costas arrays are only one family of DDCs, and doubly periodic DDCs. Two other families which were considered in the literature, are the sonar sequences [@GoTa82; @Games87; @MGC97; @EGRT92], and the Golomb rectangles [@Rob85; @Rob97]. The Folding and Unfolding Methods {#sec:folding} ================================= This section is devoted to with a transformation of a periodic sequence into a doubly periodic array and a transformation of a doubly periodic array into a periodic sequence. The two transformations will be called folding and unfolding, respectively, and as one might expect these two transformations are inverse of each other. As a consequence of the definition of folding we will be able to define the order of the elements in a cyclic shape. We will give the known theorems on the necessary and sufficient conditions that a folding exists. Based on these results we will define the inverse operation of unfolding. This will lead to the main theorem which will state when a doubly periodic two-dimensional DDC (or a cyclic DDC) is unfolded into a Sidon sequences. The definition of folding involves a lattice tiling $({{\cal T}},{{\cal S}})$, where ${{\cal S}}$ is the shape on which the folding is performed. A [*direction*]{} is a nonzero integer vector $(d_1,d_2)$, where $d_1 , d_2 \in {{\mathbb{Z}}}$. Let ${{\cal S}}$ be a two-dimensional shape and let $\delta=(d_1,d_2)$ be a direction. Let $\Lambda$ be a lattice tiling for a shape ${{\cal S}}$, and let ${{\cal S}}_1$ be the copy of ${{\cal S}}$, in the related tiling, which includes the origin. We define recursively a *folded-row* starting in the origin. If the point $(i_1,i_2)$ is the current point of ${{\cal S}}_1$ in the folded-row, then the next point on its folded-row is defined as follows: - If the point $(i_1+d_1,i_2+d_2)$ is in ${{\cal S}}_1$ then it is the next point on the folded-row. - If the point $(i_1+d_1,i_2+d_2)$ is in ${{\cal S}}_2 \neq {{\cal S}}_1$ whose center is in the point $(c_1,c_2)$ then $(i_1+d_1-c_1,i_2+d_2-c_2)$ is the next point on the folded-row (by Lemma \[lem:center\] this point is on ${{\cal S}}_1$). The definition of folding is based on a lattice $\Lambda$, a shape ${{\cal S}}$, and a direction $\delta$. The triple $(\Lambda,{{\cal S}},\delta)$ defines a folding if the definition yields a folded-row which includes all the elements of ${{\cal S}}$. It appears that only $\Lambda$ and $\delta$ determines whether the triple $(\Lambda,{{\cal S}},\delta)$ defines a folding. The role of ${{\cal S}}$ is only in the order of the elements in the folded-row; and of course $\Lambda$ must define a lattice tiling for ${{\cal S}}$. The first two lemmas proved in [@Etz11] are an immediate consequence of the definitions and provide us concise conditions whether the triple $(\Lambda,{{\cal S}},\delta)$ defines a folding. Let $({{\cal T}},{{\cal S}})$ be a lattice tiling defined by the two-dimensional lattice $\Lambda$ and let $\delta = (d_1,d_2)$ be a direction. $(\Lambda,{{\cal S}},\delta)$ defines a folding if and only if the set $\{ (i \cdot d_1,i \cdot d_2)-c(i \cdot d_1,i \cdot d_2) ~:~ 0 \leq i < | {{\cal S}}| \}$ contains $| {{\cal S}}|$ distinct elements. Let $({{\cal T}},{{\cal S}})$ be a lattice tiling defined by the two-dimensional lattice $\Lambda$ and let $\delta = (d_1,d_2)$ be a direction. $(\Lambda,{{\cal S}},\delta)$ defines a folding if and only if $(|{{\cal S}}| \cdot d_1,{{\cal S}}| \cdot d_2)-c(|{{\cal S}}| \cdot d_1,|{{\cal S}}| \cdot d_2)=(0,0)$ and for each $i$, $0 < i < |{{\cal S}}|$ we have $(i \cdot d_1,i \cdot d_2)-c(i \cdot d_1,i \cdot d_2) \neq (0,0)$. The next theorem determine precisely when the triple $(\Lambda,{{\cal S}},\delta)$ defines a folding. \[thm:new\_cond\_fold2D\] Let $\Lambda$ be a lattice whose generator matrix is given by $${\bf G}=\left[\begin{array}{cc} v_{11} & v_{12} \\ v_{21} & v_{22} \end{array}\right]~,$$ where all the entries of ${\bf G}$ are nonzeroes. Let $d_1$ and $d_2$ be two positive integers and $\tau = \text{g.c.d.}(d_1 , d_2)$. If $\Lambda$ defines a lattice tiling for the shape ${{\cal S}}$ then the triple $(\Lambda,{{\cal S}},\delta)$ defines a folding - with the direction $\delta =(+d_1,+d_2)$ if and only if $\text{g.c.d.}(\frac{d_1 v_{22}-d_2 v_{21}}{\tau},\frac{d_2 v_{11}-d_1 v_{12}}{\tau})=1$ and $\text{g.c.d.}(\tau , | {{\cal S}}|)=1$; - with the direction $\delta =(+d_1,-d_2)$ if and only if $\text{g.c.d.}(\frac{d_1 v_{22}+d_2 v_{21}}{\tau},\frac{d_2 v_{11}+d_1 v_{12}}{\tau})=1$ and $\text{g.c.d.}(\tau , | {{\cal S}}|)=1$; - with the direction $\delta =(+d_1,0)$ if and only if $\text{g.c.d.}(v_{12},v_{22})=1$ and $\text{g.c.d.}(d_1 , | {{\cal S}}|)=1$; - with the direction $\delta =(0,+d_2)$ if and only if $\text{g.c.d.}(v_{11},v_{21})=1$ and $\text{g.c.d.}(d_2 , | {{\cal S}}|)=1$. A direction $\delta$ for which $(\Lambda,{{\cal S}},\delta)$ defines a folding also defines the order of the elements in a cyclic shape ${{\cal S}}$. This order is exactly the order of the elements in the folded-row. It is easy to verify that only $|{{\cal S}}|-1$ directions should be considered for the existence of a folding. The order of elements on ${{\cal S}}$ is clearly not unique as it was proved in [@Etz11] that if one direction defines a folding then $\phi (| {{\cal S}}|)$ directions define a folding (and they come in pairs of reverse order), where $\phi$ is the Euler totient function. The unfolding operation is defined directly from the folding operation. Let $\Lambda$ be a lattice tiling for a two-dimensional shape ${{\cal S}}$ and a let $\delta$ be a direction, for which $(\Lambda,{{\cal S}},\delta)$ defines a folding. Then the folded-row is the *unfolded sequence* generated from the shape ${{\cal S}}$. In the folding, the folded-row indicates to which position of the array, each element of a given one-dimensional sequence will be assigned. In the unfolding, the folded-row is actually an unfolded-row and it indicates to which position of the sequence, each element of the array is assigned. These definitions are completely natural and there is no surprise. What is more interesting is the following theorem which connects two-dimensional doubly periodic ${{\cal S}}$-DDCs with Sidon sequences. \[thm:2DtoSidon\] Let ${{\cal A}}$ be a two-dimensional doubly periodic ${{\cal S}}$-DDC with period $\{ (\pi_1 , \pi_2 ) , (\xi_1 , \xi_2 ) \}$. Let $\Lambda$ be the lattice tiling of ${{\cal S}}$ formed from $(\pi_1 , \pi_2 )$ and $(\xi_1 , \xi_2 )$. If $\delta$ is a direction for which $(\Lambda,{{\cal S}},\delta)$ defines a folding then the folded-row generated by the unfolding of ${{\cal S}}$ is a Sidon Sequence. We will give a sketch of the proof. A proof with all the details will appear in the full version of this work. We assign colors to the points of the square grid as follows. The points of the shape ${{\cal S}}$ whose center is in the origin (say ${{\cal S}}_0$) are assigned colors by the order of the folded-row, where the origin is assigned with a *zero*. Each other copy of ${{\cal S}}$ in the tiling is assigned with the same colors as ${{\cal S}}_0$ in the same related positions. Assume the contrary, that the folded-row is not a Sidon sequence. It follows that there exists two distinct pairs of integers $(i_1 , i_2 )$ and $(i_3,i_4)$, where $i_1$, $i_2$, $i_3$, and $i_4$, are positions with dots on the folded-row, such that $i_4 - i_3 \equiv i_2 - i_1 ~ (mod~n)$, where $n=|{{\cal S}}|$. Let ${{\cal S}}_j$, $1 \leq j \leq 4$, a copy of ${{\cal S}}$ on the grid in which a point colored with $i_j$ is the center. It can be shown that in one of these four copies the two pair of lines which connects the points colored with $i_1$ and $i_2$ ad those colored with $i_3$ and $i_4$ are equal in length and slope. A contradiction to the assumption that ${{\cal A}}$ is a two-dimensional doubly periodic ${{\cal S}}$-DDC. In [@Etz11] the following theorem was proved. \[thm:Sidonto2D\] Let $\Lambda$ be a lattice tiling for a two-dimensional shape ${{\cal S}}$, $n=|{{\cal S}}|$, and let $\delta$ be a direction. Let ${{\cal B}}$ be a Sidon sequence with $m$ elements over ${{\mathbb{Z}}}_n$. If $(\Lambda,{{\cal S}},\delta)$ defines a folding then there exists a two-dimensional doubly periodic ${{\cal S}}$-DDC ${{\cal A}}$ with $m$ dots in each copy of ${{\cal S}}$ of ${{\cal A}}$. In view of Theorems \[thm:2DtoSidon\] and \[thm:Sidonto2D\] it is tempting to prove that “a Sidon sequence over ${{\mathbb{Z}}}_n$ with $m$ elements exists if and only if a two-dimensional doubly periodic ${{\cal S}}$-DDC with $m$ dots in each copy of ${{\cal S}}$ exists”. But, this claim is not correct. As we will conclude from the following discussion, which applies Theorem \[thm:2DtoSidon\] on the known doubly periodic constructions for Costas Arrays. The following $7 \times 6$ array was obtained by the periodic Welch Construction for $p=7$ and the primitive root 3 modulo 7. $$\begin{array}{|c|c|c|c|c|c|} \hline &&&&&\\ \hline &&&\bullet&&\\ \hline &&&&&\bullet\\ \hline &&&&\bullet&\\ \hline &\bullet&&&&\\ \hline &&\bullet&&&\\ \hline \bullet&&&&&\\ \hline \end{array}$$ By using unfolding with direction $(1,1)$, where the lower left dot is taken on the origin we obtain the Sidon sequence $\{ 0,8,10,11,33,37\}$ modulo 42. A construction such as the periodic Golomb Construction cannot produce any Sidon sequence since the shape ${{\cal S}}$ is a square and the is no direction which defines a folding when ${{\cal S}}$ is a square. New Optimal Sidon Sequences {#sec:new} =========================== The two celebrating constructions of optimal Sidon sequences are the ones of Singer [@Sin38] and Bose [@Bose42]. Let $q$ be a power of a prime number. Singer’s construction, which is based on projective planes, produces a Sidon sequence with $q+1$ elements over ${{\mathbb{Z}}}_{q^2 +q+1}$. Bose’s construction, which is based on affine planes, produces a Sidon set with $q$ elements over ${{\mathbb{Z}}}_{q^2-1}$. The construction of Ruzsa [@Ruz93] generates optimal Sidon sequences with $p-1$ elements taken modulo $p^2 -p$, where $p$ is a prime number. In this section we generalize this construction to obtain a Sidon sequence with $q-1$ elements taken over $(q-1) \times$GF($q$), where $q$ is any power of a prime. Given a power of a prime $q$ and a primitive element $\alpha$ in GF($q$), we construct the set $A_{q,\alpha}$ defined by $$A_{q,\alpha} = \{ (i, \alpha^i ) ~:~ 0 \leq i \leq q-2 ~ \}$$ The set $A_{q,\alpha}$ is an optimal Sidon sequence. We have to prove that given four integers $i_1$, $i_2$, $i_3$, and $i_4$, $0 \leq i_1 ,i_2 ,i_3 ,i_4 \leq q-2$, such that $i_1 \neq i_3$ and $i_2 \neq i_3$ then the two pairs $(i_1 +i_2 , \alpha^{i_1} + \alpha^{i_2})$ and $(i_3 +i_4 , \alpha^{i_3} + \alpha^{i_4})$ are not equal. Assume the contrary, that for four such integers we have $$\label{eq:base} i_1 + i_2 \equiv i_3 + i_4 ~(\text{mod}~q-1),~~~~ \alpha^{i_1} + \alpha^{i_2} = \alpha^{i_3} +\alpha^{i_4} ~. \vspace{-0.1cm}$$ Let $t \equiv i_1 - i_3 \equiv i_4 - i_2 ~(\text{mod}~q-1)$, where clearly we can assume that $0 < t < q-1$. Hence, we replace (\[eq:base\]) with the equations $$\label{eq:newbase} i_1 - i_3 \equiv i_4 - i_2 ~(\text{mod}~q-1),~~~~ \alpha^{i_1} - \alpha^{i_3} = \alpha^{i_4} -\alpha^{i_2} \vspace{-0.1cm}$$ We can substitute $i_1 \equiv i_3 +t ~(\text{mod}~q-1)$ and $i_4 \equiv i_2 +t ~(\text{mod}~q-1)$ in (\[eq:newbase\]) to obtain the equation $$\alpha^{t+i_3} - \alpha^{i_3} = \alpha^{t+i_2} - \alpha^{i_2}, \vspace{-0.1cm}$$ which is equivalent to the equation $$\label{eq:last} \alpha^{i_3} (\alpha^t -1) = \alpha^{i_2} (\alpha^t -1) ~. \vspace{-0.1cm}$$ Since $0 < t < q-1$, it follows that $\alpha^t -1 \neq 0$ and hence from (\[eq:last\]) we have that $\alpha^{i_3} = \alpha^{i_2}$, i.e., $i_3 = i_2$ which contradicts the original choice of $i_2$ and $i_3$. Therefore, $A_{q,\alpha}$ is a Sidon sequence. The optimality of the sequence is a straight forward enumeration. The new construction of Sidon sequences does not help in constructing new doubly periodic ${{\cal S}}$-DDC since the abelian group is not ${{\mathbb{Z}}}_n$. conclusions and Future Research {#sec:conclude} =============================== Sidon sets have many applications in coding theory and in communication problems. Constructions of optimal Sidon sequences are rare. We defined periodicity and cyclic arrays in the two-dimensional case. We proved that unfolded optimal doubly periodic two-dimensional synchronization patterns are optimal Sidon sequences. Thus, forming some equivalence between the two structures. All the results concerning two-dimensional arrays are generalized readily to higher dimensions. We presented a new construction of optimal Sidon sequences, with $q-1$ elements, over an alphabet with $q(q-1)$ elements, where $q$ is a power of a prime. The main problem for future research in this direction is to find new constructions for optimal doubly periodic DDCs and new constructions for optimal Sidon sequences. Acknowledgment {#acknowledgment .unnumbered} ============== This work was supported in part by the United States-Israel Binational Science Foundation (BSF), Jerusalem, Israel, under Grant No. 2006097. [10]{} \[1\][\#1]{} url@rmstyle \[2\][\#2]{} A. E. Brouwer, J. B. Shearer, N. J. A. Sloane, and W. D. Smith, “A new table of constant weight codes”, *IEEE Trans.Inform. Theory*, vol.36, pp.1334–1380, November 1990. R. M. Roth and G. Seroussi, “Location-Correcting Codes”, *IEEE Trans. Inform. Theory*, vol.42, pp.554–565, March 1996. S. R. Blackburn, T. Etzion, K. M. Martin, and M. B. Paterson, “Two-Dimensional Patterns with Distinct Differences – Constructions, Bounds, and Maximal Anticodes”, *IEEE Trans. on Inform. Theory*, vol. IT-56, pp. 1216-1229, March 2010. T. Etzion, “Sequence folding, lattice tiling, and multidimensional coding”, *IEEE Trans. on Inform. Theory*, to appear. S. R. Blackburn, T. Etzion, K. M. Martin, and M. B. Paterson, “Distinct difference configurations: multihop paths and key predistribution in sensor networks”, *IEEE Trans. on Inform. Theory*, vol. IT-56, pp. 3961-3972, August 2010. A. Barg and A. Mazumdar, “Codes in permutations and error correction for rank modulation”, *IEEE Trans. on Inform. Theory*, vol.IT-56, pp.3158–3165, July 2010. K. O’Bryant, “A complete annotated bibliography of work related to Sidon sequences”, *The Elec. J. of Combin.*, DS11, pp.1–39, July 2004. W. C. Babcock, “Intermodulation interference in radio systems,” *Bull. Sys. Tech. Journal*, pp.63–73, June 1953. G. S. Bloom and S. W. Golomb, “Applications of numbered undirected graphs”, *Proceedings of the IEEE*, vol.65, pp.562–570, April 1977. M. D. Atkinson, N. Santoro, and J. Urrutia, “Integer sets with distinct sums and differences and carrier frequency assignments for nonlinear repeaters”, *IEEE Transactions on Communications*, vol.COM-34, pp.614–617, 1986. A. W. Lam and D. V. Sarwate, “On optimum time-hopping patterns”, *IEEE Transactions on Communications*, vol.COM-36, pp.380–382, 1988. S. W. Golomb and H. Taylor, “Two-dimensional synchronization patterns for minimum ambiguity”, *IEEE Trans. Inform.Theory*, vol.IT-28, pp.600–604, 1982. S. W. Golomb and H. Taylor, “Constructions and properties of Costas arrays”, *Proceedings of the IEEE*, vol.72, pp.1143–1163, 1984. J. P. Robinson, “Golomb rectangles”, *IEEE Trans. Inform.Theory*, vol.IT-31, pp.781–787, 1985. R. A. Games, “An algebraic construction of sonar sequences using M-sequences”, *SIAM Journal on Algebraic and Discrete Methods*, vol.8, pp.753–761, October 1987. A. Blokhuis and H. J. Tiersma, “Bounds for the size of radar arrays”, *IEEE Trans. Inform. Theory*, vol.IT-34, pp.164–167, January 1988. J. P. Robinson, “Golomb rectangles as folded ruler”, *IEEE Trans. Inform. Theory*, vol.IT-43, pp.290–293, 1997. S. R. Blackburn, T. Etzion, K. M. Martin, and M. B. Paterson, “Efficient key predistribution for grid-based wireless sensor networks,” *Lecture Notes in Computer Science*, vol. 5155, pp.54–69, August 2008. R. A. Games, “Algebraic constructions for Costas arrays”, *Journal Combinatorial Theory, Ser. A*, vol.37, pp.13–21, 1984. K. Drakakis, “A review of Costas arrays”, *J. Appl. Math.*, vol.???, pp.1–32, 2006. S. W. Golomb and G. Gong, “The status of Costas arrays”, *IEEE Trans.Inform. Theory*, vol.53, pp.4260–4265, November 2007. O. Moreno, S. W. Golomb, and C. Corrada, “Extended sonar sequences”, *IEEE Trans. Inform. Theory*, vol.43, pp.1999–2005, November 1997. O. Moreno and S. Golomb, “A new optimal double periodical construction of one target two-dimensional arrays”, *40th Annual Conference on Information Sciences and Systems*, pp. 518–522, March 2006. T. Etzion, S. W. Golomb, and H. Taylor, “Tuscan-$k$ squares”, *Advances  in  Applied  Mathematics*, vol.10, pp.164–174, 1989. K. Drakakis, F. Iorio, and S. Rickard “The enumeration of Costas arrays of order 28”, *Information Theory Workshop*, Dublin, September 2010. P. Erdős, R. Graham, I. Z. Ruzsa, and H. Taylor, “Bounds for arrays of dots with distinct slopes or lengths”, *Combinatorica*, vol.12, pp.39–44, 1992. J. Singer, “A theorem in finite projective geometry and some applications to number theory”, *Trans. Amer. Math. Soc.*, vol. 43, pp. 377-385, 1938. R. C. Bose, “An affine analogue of Singer’s theorem”, *J. Indian Math. Soc. (N.S.)*, vol. 6, pp. 1-15, 1942. I. Z. Ruzsa “Solving a linear equation in a set of integers”, *ActaArith.*, vol.65, pp.259–282, 1993.
{ "pile_set_name": "ArXiv" }
--- abstract: | We study the magnetic response of a superconducting double strip, *i.e.*, two parallel coplanar thin strips of width $2w$, thickness $d \ll w$ and of infinite length, separated by a gap of width $2s$ and subject to a perpendicular magnetic field $H$. The magnetic properties of this system are governed by the presence of a geometric energy barrier for vortex penetration which we investigate as a function of applied field $H$ and gap parameter $s$. The new results deal with the case of a narrow gap $s \ll w$, where the field penetration from the inner edges is facilitated by large flux focusing. Upon reducing the gap width $2s$, we observe a considerable rearrangement of the screening currents, leading to a strong reduction of the penetration field and the overall magnetization loop, with a suppression factor reaching $\sim (d/w)^{1/2}$ as the gap drops below the sample thickness, $2s < d$. We compare our results with similar systems of different shapes (elliptic, rectangular platelet) and include effects of surface barriers as well. Furthermore, we verify that corrections arising from the magnetic response of the Shubnikov phase in the penetrated state are small and can be omitted. Extending the analysis to multiple strips, we determine the specific sequence of flux penetrations into the different strips. Our studies are relevant for the understanding of platelet shaped samples with cracks or the penetration into layered superconductors at oblique magnetic fields. author: - 'R. Willa' - 'V.B. Geshkenbein' - 'G. Blatter' title: Suppression of Geometric Barrier in Type II Superconducting Strips --- Introduction {#sec:introduction} ============ The characteristic properties of a superconductor are its diamagnetic response [@Meissner_33] $M$ to an external magnetic field $H$ and its ability to transport electric current without dissipation[@Onnes_11]. In the Meissner phase the magnetic induction $B = H + 4 \pi M$ vanishes inside the superconductor and the linear response $M = -H/4\pi$ is that of a perfect (bulk) diamagnet. In type II superconductors, a sufficiently large magnetic field $H>{H_{p}}$ penetrates the material via quantized flux lines (with flux $\Phi_{0}=hc/2e$); we denote with ${H_{p}}$ the field of first penetration. Within the mixed (or Shubnikov[@Shubnikov_37]) phase the presence of vortices reduces the bulk diamagnetic signal and the magnetization $M(H)$ decreases in magnitude. The magnetic properties of the material then depend on the behavior of the vortex state. In this paper, we determine the magnetic response of superconducting samples of more complex shape, in particular a double strip, two parallel coplanar thin strips of infinite length and subject to a perpendicular magnetic field $H$, see [Fig. ]{}\[fig:sketch-double-strip\]. The response of such a system is hysteretic and dominated by the so-called geometrical barrier[@Zeldov_94; @Benkraouda_96], i.e., an energy barrier retarding the magnetic field penetration. Our main result is an apparent suppression of the geometrical barrier for the situation where the two strips are closeby, i.e., separated by a narrow gap or crack. Such a suppression of geometrical barriers may be of practical interest in experiments, as has been the case in disentangling the vortex lattice melting- and irreversibility lines in layered BiSCCO superconductors [@Majer_95] or in separating apart the phenomenon of bulk vortex pinning by defects. So far, the geometrical barrier has been deliberately suppressed by polishing the sample into the shape of a prism [@Majer_95]; the suppression of the geometrical barrier observed when tilting the magnetic field applied to the sample [@Segev_11] and attributed to the appearance of Josephson vortex stacks resembles the mechanism reported in the present paper. ![Side-view representation ($xz$-plane) of two flat superconducting strips (parallel to $y$) subject to a perpendicular magnetic field $H$ (directed along $z$). The cross-sections of the strips have a width $2w$ and a thickness $d \ll w$, while the separation $2s$ between their inner edges measures the width of the gap. The outer edges of the strips at $\pm (2w + s)$ are denoted by $\pm W$. Any position in the $xz$-plane is described by the complex coordinate $\xi = x + i z$.[]{data-label="fig:sketch-double-strip"}](plot_09.eps){width="7cm"} The precise shape of the magnetization curve depends on the specific configuration assumed by the vortices after penetration, which is determined by the sample shape and its surface properties (we assume a sample free of defects). The sample surface is relevant in the determination of the penetration field ${H_{p}}$ as defined in the asymptotic region far away from the sample. A flat surface parallel to the field generates an image vortex which results in a surface barrier hindering vortices from entering the sample [@Bean_64; @Clem_74]. The metastable Meissner state survives until the local field at the surface is increased beyond the critical value ${H_{s}}$ which is of the order of the thermodynamic critical field $H_c$, ${H_{s}}\sim H_c > H_{c1}$, with $H_{c1}$ the lower critical field. For a non-ideal surface the effective surface barrier is reduced and assumes a value ${H_{s}}$ between $H_{c1}$ and $H_c$. The sample shape is relevant, too, in the determination of the penetration field ${H_{p}}$. This is well known for elliptic-shaped samples, cf. Fig.\[fig:geometries\], where the magnetic field is enhanced near the sample edge: for a cylindrical shaped diamagnetic (i.e., $\mu = 0$) sample with an elliptic cross section of height $d$ and width $2w$, the demagnetization factor[@Osborn_45] $n= 2w / (2w+d)$ generates a field enhancement ${H_{\mathrm{edge}}}= (1-n)^{-1} H$. Correspondingly, the penetration field is given by ${H_{p}}= (1-n){H_{s}}= d/(2w+d) {H_{s}}$. Once the penetration field is reached, vortices enter the sample, reversibly in the absence of a surface barrier (i.e., if ${H_{s}}= H_{c1}$) and irreversibly else. Without surface barrier, the vortices distribute homogeneously inside the sample, a result that is consistent with the constant induction inside a magnetic ellipsoid[@Landau_60]. On a microscopic level, this corresponds to an exact matching of the energy gain of vortex motion in the field of the screening current and the energy cost $\varepsilon_l$ associated with the increasing vortex length upon penetration, see Fig. \[fig:geometries\]. ![Top: sketch of field enhancement near the edges of an elliptic- (left) and a rectangular- (right) shaped sample. Below penetration $H < {H_{p}}$, for both geometries, the field is enhanced by the factor $\sim \sqrt{w/d}$ a distance $d$ away from the edges. The field remains unchanged on approaching the rectangular edge but increases by a further factor $\sim \sqrt{w/d}$ for the elliptic geometry. Upon increasing $H$ beyond ${H_{p}}$, the field penetrates homogeneously into the elliptic shaped sample and concentrates in a central dome for the rectangular sample. This is due to the different potential landscapes $U_{\mathrm{geo}}(x)$ (see bottom sketch) felt by the vortices penetrating the sample at $H \sim {H_{p}}$, flat for the ellipse (dotted line) and attractive for the rectangle (solid line). Note that the penetration fields differ by the factor $\sqrt{d/w}$ for the elliptic and the rectangular sample. The sketch illustrates the situation without additional surface barrier.[]{data-label="fig:geometries"}](plot_00.eps){width="45.00000%"} For a platelet shaped sample (of width $2w$ and thickness $d$) with a rectangular edge, the field at the boundary is enhanced as well, although (effectively) less than for the elliptic sample. A distance $d$ away from the edge[@footnote:enhancement-ellipse], the applied field $H$ is enhanced by a factor $\sim (w/d)^{1/2}$, resulting in a penetration field ${H_{p}}\sim (d/w)^{1/2} {H_{s}}$. At this field strength, the barrier for vortex entry into the sample has vanished and vortices move to the center of the sample where they accumulate in a dome-shaped form, cf. Fig. \[fig:geometries\]. Under further increase of the external field $H$, the vortex dome grows both in height and width until the sample is fully penetrated. In this geometry, the cost $\varepsilon_l d$ to create the vortex is payed right upon vortex entry at the sample edge; beyond the edge region the energy gain in the current field $I(x)$ is no longer balanced by the energy cost and the vortex is driven to the sample center. Hence the field penetration into the platelet shaped sample is irreversible even in the absence of a surface barrier, what is due to the presence of a geometric barrier defined through the energy cost for flux entry. It is this type of geometric barrier effects [@Zeldov_94; @Benkraouda_96] which is at the focus of the present paper. Another situation arises in dirty samples where vortices are pinned onto defects. Once the surface and geometrical barriers are overcome, the vortex arrangement may be dominated by bulk pinning and the magnetic induction (or magnetization) is given by a Bean profile[@Bean_62]. What is common to all three cases, surface-, geometric-, and bulk pinning is the irreversible, hysteretic behavior of the magnetization $M(H)$ with changing external field $H$. In this paper, we concentrate on the defect-free case and thus ignore possible modifications due to bulk pinning. The motivation to study geometrical barriers in samples of complex shape is manifold: Originally, the understanding of the flux penetration and vortex lattice melting in layered high-$T_c$ superconductors necessitated a proper analysis of the vortex state in platelet-shaped samples[@Zeldov_94]. On the technological side, the structuring of current-carrying strips [@Benkraouda_98; @Mawatari_01] enhances their critical current as the incorporation of slits generates geometrical barriers hindering vortex motion. Recently, Segev *et al.*[@Segev_11] observed a structured vortex dome in layered $\mathrm{Bi_{2}Sr_{2}CaCu_{2}O_{8+\delta}}$ samples subject to a tilted magnetic field. This finding can be interpreted as arising from stacks of in-plane (Josephson) vortices reducing the superconducting order parameter[@Koshelev_99] and acting as weak links for the perpendicular field (pancake vortices[@Feigelman_90; @Clem_91]). Our analysis of vortex penetration into a double-strip with a narrow gap, see Fig.\[fig:sketch-double-strip\], may serve as a first step towards the understanding of flux penetration in this geometry. From a general perspective, the magnetic response associated with superconducting samples can be calculated numerically. Effects of complex sample shapes, inhomogeneous material equations, and time-dependent perturbations can then be studied quantitatively[@Brandt_99]. On the other hand, analytic approaches give more qualitative insights into the system’s behavior. Earlier work on geometrical barriers in samples with more complex shapes considered the case of two coplanar thin strips in the Meissner phase[@Brojeny_02] and the full magnetization curve for a strip-shaped sample with a slit[@Mawatari_03], i.e., two strips shunted at their ends; this ring-type topology with circulating currents exhibits a markedly different magnetization $M(H)$ as compared to our unshunted situation. The situation of an unshunted double stripline in the critical state was investigated in Ref.[\[\]]{}. In our work, we go beyond these results in various ways, including the situation where the sample thickness $d$ plays an important role. ![Sketch of the geometric energy barrier $U_{b}$ for vortex penetration as a function of the applied field $H$ and the gap parameter $s$, see also [Fig. ]{}\[fig:sketch-double-strip\]. In this Figure we neglect an additional surface barrier, i.e., ${H_{s}}= H_{c1}$. The thick black curve marks the geometric barrier height $U_{b}^{\mathrm{eq}}(s)$ at the equilibrium field ${H_{\mathrm{eq}}}$ as defined in [Eq. ]{} and provides a measure for the irreversibility of the sample. Note the rapid decrease of the geometric barrier $U_{b}^{\mathrm{eq}}(s\ll d,H)$ with increasing field $H$ at small separation $s$ between the two strips; the small geometric barrier $U_{b}^{\mathrm{eq}}(s)$ tells that irreversibility is reduced when $s~\ll~d$. Still, a finite irreversibility remains with the geometrical barrier rapidly reinstalled when reversing the applied field.[]{data-label="fig:geometric-barrier"}](plot_21_mod.eps){width="45.00000%"} The most pertinent new result is the dramatic suppression of the geometrical barrier which we illustrate in [Fig. ]{}\[fig:geometric-barrier\]. This suppression is driven by a large flux-focussing into the gap between the strips, forcing the flux penetration into the sample to start from the inner edges. In tracing the evolution of the penetration field ${H_{p}}$ as a function of separation $s$ between the strips, we find it decay from ${H_{p}}\sim \sqrt{d/w} \,{H_{s}}$ at large $s$ to ${H_{p}}\sim \sqrt{sd/w^2} \log(w/s)\, {H_{s}}$ at intermediate separation $d < s < w$ to ${H_{p}}\sim (d/w) \, {H_{s}}$ at small $s \ll d$; the latter coincides with the result for the elliptic sample where the geometrical barrier is absent alltogether. We emphasize, however, that the narrow-gap double-strip still differs from the ellipse as the geometrical barrier remains present but rapidly collapses from $\varepsilon_l d$ to zero with increasing field, hence maintaining the hysteretic magnetization. The latter strongly decreases with the separation $s$ between strips as well: Within the individual strips, the penetrated field assumes a dome-like shape which is increasingly skewed towards the gap when $s$ becomes small. Following the change in shape of the magnetization curve through the various regimes, we find it to shrink by a factor $\propto (s/w)^{1/2} \log(w/s)$ when $s < w$ and by a factor $\propto (d/w)^{1/2}$ for narrow gaps $s \ll d$ when compared to the single platelet sample; this decay of the magnetization with decreasing $s$ ends up in a flat and nearly constant value $M = -({H_{s}}/4\pi) (4wd)$ at small $s \log(W/s) \ll d $. In the following, we briefly recall the key features of the magnetic response for elliptically shaped strips in Sec. \[sec:sec:ellipse\] and proceed with the description of coplanar parallel rectangular strips for the case where the thickness $d$ is the smallest geometric length in the problem (Sec. \[sec:sec:formalism\]). We review the appearance and consequences of a geometric barrier in a single strip (Sec. \[sec:sec:single-strip\]) and continue with the analysis of two adjacent strips (Sec.\[sec:sec:double-strip\]) discussing the behavior of the Meissner- and penetrated states. In Sec. \[sec:finite-thickness\], we analyze the double strip for the situation where the separation $2s$ between the strips is smaller than the strip thickness $d$, $s \ll d$. Section \[sec:several-strips\] is devoted to multi-strips and a summary and conclusions are given in Sec. \[sec:conclusions\]. Thin Strips {#sec:thin-strips} =========== Introduction - Elliptical strip {#sec:sec:ellipse} ------------------------------- Before considering samples with rectangular geometries, it is instructive to revisit the magnetic properties of a flat superconducting strip with an *elliptic* cross-section. The strip extends infinitely in the $y$-direction and the semi-axes along $x$ and $z$ are $w$ and $d/2$ ($d \ll w$) respectively, with the upper/lower sample surface parametrized by $z_{\pm}(x) = \pm(d/2w)\sqrt{w^2-x^2}$. The magnetic field $H$ is applied parallel to the $z$-axis; outside the sample, $\boldsymbol{B} = \boldsymbol{H}$, while $B_{\mathrm{el}} = \mu(B_{\mathrm{el}}) H_{\mathrm{el}}$ is constant and parallel to the $z$-axis inside the elliptic sample[@Landau_60], a consequence of the special elliptic shape. Here, $$\begin{aligned} \label{eq:mu-from-free-energy} \mu(B) = \frac{B}{4\pi}\Big(\frac{dF}{dB}\Big)^{-1}\end{aligned}$$ is the magnetic permeability of the material as obtained from the free energy density $F(B)$. The magnetic field at the sample edges $(\pm w,0)$ is continuous[@Landau_60], $H_{\mathrm{el}} = {H_{\mathrm{edge}}}$, where ${H_{\mathrm{edge}}}$ denotes the magnetic field strength at the sample edge. The latter is modified due to demagnetization effects of the sample which are described by the geometric demagnetizing factor[@Osborn_45] $n = 2w / (2w + d) \approx 1 - d/2w$. Exploiting the fact that the magnetic induction $B_\mathrm{el}$ is constant within the ellipse, we decompose the total field $\boldsymbol{B}(x,z)$ into two components, a constant one $\boldsymbol{B}_\mathrm{el} = (0,0,B_\mathrm{el})$, and the remaining field $\boldsymbol{B}_0(x,z)$ which does not penetrate the sample. Far away from the sample, all fields point along $z$, $\boldsymbol{B}_0 \equiv (0,0,B_0^\infty)$ and we have $B_\mathrm{el} + B_0^\infty = H$. The component $\boldsymbol{B}_0(x,z)$ then describes the field of a perfectly diamagnetic ellipse in the reduced external field $B_0^\infty = H-B_\mathrm{el}$. The magnetic field at the sample edge ($x = \pm w$) points along $z$, involves the two components $B_\mathrm{el}$ and $B_0 = B_0^\infty/(1-n)$, the latter enhanced by demagnetization effects, and reads $$\begin{aligned} \label{eq:edge-field-general} {H_{\mathrm{edge}}}&= B_\mathrm{el} + \frac{B_0^\infty}{1-n}.\end{aligned}$$ Using $B_0^\infty = H-B_{\mathrm{el}}$ as well as $B_{\mathrm{el}} = \mu(B_{\mathrm{el}}) H_{\mathrm{el}} = \mu(B_{\mathrm{el}}) {H_{\mathrm{edge}}}$, we obtain the standard formula for the field strength inside the sample[@Landau_60] $$\begin{aligned} \label{eq:elliptic-induction} B_{\mathrm{el}} = \frac{\mu(B_{\mathrm{el}})}{1-n[1-\mu(B_{\mathrm{el}})]}H,\end{aligned}$$ where the value for $B_{\mathrm{el}}$ has to be determined self-consistently. For notational simplicity we denote by $\mu$ the value for $\mu(B_{\mathrm{el}})$ after solving the above equation. The $\boldsymbol{B}$-field at the surface outside of the ellipse has both a normal ($\perp$) and a tangential ($\|$) component. Their magnitudes can be determined from the boundary conditions[@Landau_60], telling that $B_{\perp}$ and $B_{\|}/\mu$ are continuous across the surface. For the upper surface $z=z_+(x)$ of the ellipse we find $$\begin{aligned} \label{eq:elliptic-field-parallel-perp} \boldsymbol{(}B_{\|}(x),B_{\perp}(x) \boldsymbol{)} &= \frac{H}{1-n(1-\mu)}\boldsymbol{(}\sin(\alpha), \mu \cos(\alpha)\boldsymbol{)},\end{aligned}$$ where $$\begin{aligned} \alpha(x) &= \arctan\bigg(\frac{d}{2w}\frac{-x}{\sqrt{w^2-x^2}}\bigg)\end{aligned}$$ measures the angle between the external field orientation ($z$-axis) and the direction normal to the elliptic surface at the position $\boldsymbol{(}x, z_{+}(x)\boldsymbol{)}$. In most of the strip region (when $w-|x| \gg d^2/w$) the surface of the ellipse is almost parallel to the $x$-axis and the above field expression simplifies to $$\begin{aligned} \label{eq:elliptic-field-parallel-perp-approx} \boldsymbol{(}B_{\|}(x),B_{\perp}(x) \boldsymbol{)} &\approx \frac{H}{1-n(1-\mu)}\Big(\frac{-x(1-n)}{\sqrt{w^2-x^2}},\mu \Big).\end{aligned}$$ The discontinuity of the field parallel to the boundary determines the surface current that generates the magnetization of the sample. Using Ampère’s law and defining the sheet current density $I(x) = \int_{z_-}^{z_+} dz\, j(x,z)$ across the sample, we find $$\begin{aligned} I(x)&\approx \frac{2 c}{4\pi} \big[ B_{\|}(x) - B_{\mathrm{el}} \sin(\alpha)\big]\\ \label{eq:approx-field-at-the-surface-of-ellipse} &\approx -\frac{Hc}{2\pi} \frac{(1-n)(1-\mu)}{1-n(1-\mu)} \frac{x}{\sqrt{w^2-x^2}}\\ \label{eq:current-density-ellipse} &= -\frac{(H-B_{\mathrm{el}})c}{2\pi} \frac{x}{\sqrt{w^2-x^2}}.\end{aligned}$$ The factor 2 originates from the two current contributions at the upper and lower sample surface. The last expression shows that only the expelled component $B_0^\infty = H-B_{\mathrm{el}}$ contributes to the shielding currents. The magnetization $M$ (per unit length) is obtained from the relation $4\pi M / A = B_{\mathrm{el}} - H_{\mathrm{el}}$, where $A = \pi w d/2$ is the area of the strip’s cross-section. Using $H_{\mathrm{el}} = {H_{\mathrm{edge}}}$ and [Eq. ]{} gives for the magnetization $$\begin{aligned} \label{eq:magnetization-ellipse-general} M&= -\frac{B_{0}^{\infty}}{4} w^2 = -\frac{H}{4} \frac{(1-n)(1-\mu)}{1-n(1-\mu)} w^2.\end{aligned}$$ In the last equality we used $B_0^\infty = H-B_{\mathrm{el}}$ and [Eq. ]{}. ### Meissner state {#sec:sec:sec:meissner-ellipse} At low fields, the superconducting elliptic strip remains in the Meissner state ($\mu = 0$), resulting in a vanishing induction, i.e., $B_\mathrm{el} = B=0$. The field strength at the edge, see [Eq. ]{}, is enhanced by the geometric factor $1/(1-n) \approx 2w/d$ as compared to the applied field $H$. At the sample surface, the field is everywhere tangential and its strength is given by $H \sin(\alpha)/(1-n)$ $(\approx -H x/ \sqrt{w^{2}- x^{2}})$ as obtained from [Eqs. ]{} and . The resulting sheet current density inside the sample is obtained from [Eq. ]{}, $$\begin{aligned} \label{eq:meissner-current-ellipse} I(x) &\approx -\frac{H c}{2\pi} \frac{x}{\sqrt{w^2-x^2}}.\end{aligned}$$ The perfectly diamagnetic response \[[Eq. ]{} with $\mu = 0$\] $$\begin{aligned} \label{eq:magnetization-meissner-ellipse} M &= -\frac{H}{4}w^{2}.$$ lasts until the magnetic flux starts penetrating the superconducting sample in the form of vortices. To bring a vortex to the position $x$ inside the sample costs an energy $U_{\scriptscriptstyle{L}}(x) = \varepsilon_{l} \ell(x)$, gradually rising with the vortex length $\ell(x) = z_{+}(x) - z_{-}(x)$ from zero at the sample edges to $d$ in the sample center; here, the line-energy $\varepsilon_{l} = \varepsilon_{0} \log(\lambda/\xi) = \Phi_{0}(dF/dB)|_{B=0}$ is the cost per unit length associated with the nucleation of a single vortex in the bulk superconductor. On the other hand, the work gained from the Lorentz force \[due to the current $I(x)$ in [Eq. ]{}\] drives the vortex entrance. The two energy contributions can be combined to an effective potential landscape[@Likharev_71] for a single vortex $$\begin{aligned} \label{eq:energy-profile} U_{\mathrm{geo}}(x) &= U_{\scriptscriptstyle{L}}(x) - \frac{\Phi_{0}}{c} \int\limits_{w}^{x} du\,I(u).\end{aligned}$$ In the elliptical geometry, the functional form of the driving energy due to the current [Eq. ]{} coincides with the geometrical thickness $\ell(x) = d\sqrt{1-x^{2}/w^{2}}$ of the sample and the energy profile reduces to $$\begin{aligned} U_{\mathrm{geo}}(x) &= \varepsilon_{l}\ell(x) \bigg(1- \frac{H \Phi_{0}}{4\pi \varepsilon_{l}} \frac{2w}{d}\bigg).\end{aligned}$$ The barrier then vanishes throughout the sample at the penetration field $$\begin{aligned} \label{eq:single-penetration-field-ellipse} {H_{p}}&= \frac{4\pi \varepsilon_{l}}{\Phi_{0}}\frac{d}{2w} = H_{c1} \frac{d}{2w}\end{aligned}$$ where the *local* field strength at the edge reaches $H_{c1}$ and the magnetization (per unit length) as obtained from [Eq. ]{} amounts to $$\begin{aligned} \label{eq:single-maximal-magnetization-ellipse} M_{p} = -\frac{H_{c1}}{8} wd = -\frac{H_{c1}}{4\pi} \frac{\pi w d}{2},\end{aligned}$$ with $\pi w d/2$ the cross-section of the strip. ### Penetrated state {#sec:sec:sec:penetrated-ellipse} Beyond the field of first penetration ${H_{p}}$, vortices homogeneously flood the sample, and the potential landscape takes the form (we replace $B_\mathrm{el} \to B$) $$\begin{aligned} \label{eq:energy-profile-penetrated-state} U_{\mathrm{geo}}(x) &= \varepsilon_{l}(B)\frac{d}{w}\sqrt{w^{2}-x^{2}} - \frac{\Phi_{0}}{c} \int\limits_{w}^{x} du\,I(u).\end{aligned}$$ The line energy $\varepsilon_{l}(B)$ describes the energy difference (per unit length) between the vortex state and the homogeneous field configuration, i.e., $$\begin{aligned} \label{eq:effective-line-energy} \varepsilon_{l}(B) &= \Phi_{0}\frac{d}{d B}\bigg[F(B) -\frac{B^{2}}{8\pi}\bigg] = \frac{\Phi_{0} B}{4\pi} \frac{1 -\mu(B)}{\mu(B)}\end{aligned}$$ with $F(B)$ the free energy density of the superconducting state. The second term on the right-hand side of [Eq. ]{} is modified as well, since only the non-penetrating (diamagnetic) part $H-B$ of the field drives the diamagnetic currents in [Eq. ]{}. The resulting state remains in equilibrium for all $H > {H_{p}}$, i.e., $U_{\mathrm{geo}}(x) \equiv 0$, and the reversible magnetic response follows the form in Eq. $$\begin{aligned} \label{eq:magnetization-superconducting-ellipse} M&= -\frac{H-B}{4} w^2\end{aligned}$$ with $B$ determined by the self-consistency equation . A finite surface barrier as discussed further below will retard the vortex penetration and generate a hysteretic response. In order to illustrate the above results, we consider a superconductor with the Abrikosov (bulk) induction[@Tinkham_96] $$\begin{aligned} \label{eq:toy-bulk-magn} B &= C_{1} H_{c1} \Big[\log\Big(\frac{C_{2}H_{c1}}{H-H_{c1}}\Big)\Big]^{-2}\end{aligned}$$ near the penetration field, with $C_{1,2}$ constants of order unity. In this equation, $H = H_\mathrm{edge}$ is the local field strength at the surface of the bulk sample. The magnetic permeability $\mu(B)$, can be extracted from the above expression via the relation $\mu(B) = B/H(B)$ and we find $$\begin{aligned} \label{eq:toy-permeability} \mu(B) = \frac{B}{H_{c1}} \bigg[1 + C_{2}\exp\bigg(\!-\sqrt{\frac{C_{1}H_{c1}}{B}}\ \bigg)\bigg]^{-1}.\end{aligned}$$ The linear slope $1/H_{c1}$ of the permeability near $B=0$ follows from the vertical onset of the induction (see [Eq. ]{}) beyond $H_{c1}$. Dropping the exponential term in [Eq. ]{} close to the penetration ($B \ll H_{c1}$) and substituting $\mu$ to the self-consistency equation we obtain the induction $$\begin{aligned} \label{eq:ellipcit-induction} B(H) = (H-{H_{p}})/n,\end{aligned}$$ resulting in a linear decrease of the diamagnetic response, $$\begin{aligned} \label{eq:elliptic-magnetization} M(H) = -\frac{{H_{p}}}{4n} w^{2}\,\Big(1-\frac{H}{H_{c1}}\Big).\end{aligned}$$ Note that for small inductions $B \ll H_{c1}$, the diamagnetic response is very different from the usual bulk Abrikosov magnetization (see, e.g., Ref. [\[\]]{}). The linear decrease in [Eq. ]{} extrapolates to $M=0$ at $H=H_{c1}$. The full solution of [Eq. ]{} for the permeability leads to the magnetic response illustrated in [Fig. ]{}\[fig:elliptic-magnetization\]. ![Magnetic response of a superconducting elliptic strip with demagnetizing factor $n \approx d/2w$ (here $n=0.9$) as obtained from [Eq. ]{} and with material properties described by (solid line). For comparison, we show the bulk (Abrikosov) magnetization with the same permeability $\mu(B)$ (thin solid line). The vertical onset in the bulk magnetization goes over into the linear reduction of $M$ in the ellipse, extrapolating to $M=0$ at $H = H_{c1}$ (thin dashed line).[]{data-label="fig:elliptic-magnetization"}](plot_15.eps){width=".45\textwidth"} Rectangular strips - Formalism {#sec:sec:formalism} ------------------------------ Having familiarized ourselves with the results for the elliptic strip, we turn our attention to strips with rectangular shape, i.e., samples with constant thickness $d$ as opposed to the ellipse where the height is changing over the entire sample width. Specifically, we will consider (smooth) sample edges with a typical radius of curvature $\gtrsim d$ in contrast to the much sharper edge of the ellipse where the radius of curvature is $d^{2}/4w \ll d$. We consider a set of coplanar (in the $xy$-plane) and parallel superconducting strips of infinite length (along $y$), each with a rectangular shape of width $2w$ (along $x$) and thickness $d \ll 2w$ (along $z$), subject to a perpendicular magnetic field $H$ along $z$. The strip thickness $d$ is assumed to be the smallest geometric length and is set to zero in the following mathematical analysis; its finite value is properly reinstalled through appropriate boundary conditions. Because the system is effectively two-dimensional, we express the magnetic field $\boldsymbol{B}(x,z)$ in the $xz$-plane through the complex function[@Zeldov_94] $\mathcal{B}(\xi) = B_{z}(x,z) + iB_{x}(x,z)$, with the two-dimensional coordinate $(x,z)$ replaced by the complex variable $\xi = x + i z$. The magnetostatic problem of solving the Laplace equation ($\Delta \boldsymbol{B} = 0$) for $\boldsymbol{B}$ is translated to a problem in complex analysis, where the holomorphic function $\mathcal{B}(\xi)$ satisfies the Cauchy-Riemann equations (correspnding to the magnetostatic equations $\nabla\cdot \boldsymbol{B}=0$ and $\nabla \wedge \boldsymbol{B}=0$) in the superconductor-free region; the presence of the superconductor is accounted for through appropriate boundary conditions. The latter derive from two physical conditions: on the one hand, no vortices are present in regions where current is flowing, i.e., $$\begin{aligned} \label{eq:bcBzI} B_z(x) &= 0\quad \mathrm{ when } \quad I(x) \neq 0.\intertext{Here and below we simply call `current' the sheet current density $I(x)$ flowing between $z_{\pm}=\pm d/2$. On the other hand, no currents flow in the vortex-filled regions, } \label{eq:bcIBz} I(x) &= 0 \quad \mathrm{ when } \quad B_{z}(x) \neq 0.\end{aligned}$$ This last condition neglects the microscopic structure of the vortex state by treating the penetrated region as magnetically inactive, $\mu = 1$; the accuracy of this simplification will be discussed later in this section. Using Ampère’s law $$\begin{aligned} \label{eq:ampere-law} I(x) = \frac{c}{2\pi} B_{x}(x,0^{+}) = \frac{c}{2\pi} \mathrm{Im}[\mathcal{B}(x + i 0^{+})],\end{aligned}$$ the boundary conditions and transform to $$\begin{aligned} \label{eq:bcBzBx} B_z(x) &= 0 \quad \mathrm{ when }\quad B_{x}(x) \neq 0,\ \mathrm{and}\\ \label{eq:bcBxBz} B_x(x) &= 0 \quad \mathrm{ when }\quad B_{z}(x) \neq 0. \end{aligned}$$ For a single strip centered a the origin ($\xi = 0$) the holomorphic field $$\begin{aligned} \label{eq:ant-single-strip-function} \mathcal{B}(\xi) &= H \sqrt{\frac{\xi^{2}-{b_{0}}^{2}(H)}{\xi^{2}-w^{2}}}\end{aligned}$$ is known to satisfy all the above requirements[@Zeldov_94]; the parameter ${b_{0}}$ then determines the field configuration of the entire system. For the double strip studied below, the corresponding expression reads $$\begin{aligned} \label{eq:ant-double-strip-function} \mathcal{B}(\xi) &= H \sqrt{\frac{[\xi^{2}-{b_{1}}^{2}(H)][\xi^{2}-{b_{2}}^{2}(H)]} {(\xi^{2}-s^{2})(\xi^{2}-W^{2})}}.\end{aligned}$$ Here, the strips are arranged symmetrically, extending between $\pm s$ and $\pm W$ (with $W=s+2w$) on the $x$-axis. In order to specify the field and current distributions for these geometries, the parameters ${b_{0}}$, ${b_{1}}$, and ${b_{2}}$ (with $0\leq{b_{0}}<w$ and $s < {b_{1}}\leq {b_{2}}< W$) describing the boundaries of the field-filled region have to be determined from two physical conditions: First, the net current along each strip vanishes, i.e., $$\begin{aligned} \label{eq:current-neutrality-condition-general} \int_{\mathrm{strip}}\!\!\!\!\!dx\,I(x) &= 0.\end{aligned}$$ This (first) condition is independent of the magnetic state of the strips, Meissner or Shubnikov. The second condition regulates the penetration process of vortices into the superconducting sample. In the Meissner phase, no field penetrates the superconductor and the width of the vortex dome vanishes, imposing the (second) condition $$\begin{aligned} \label{eq:con2} \left.\begin{aligned} {b_{0}}&= 0 &&\textrm{for the single, or}\\ {b_{1}}&= {b_{2}}&&\textrm{for the double} \end{aligned}\right.\end{aligned}$$ strip geometry. The second condition for the penetrated state derives from the analysis of vortex penetration at the sample edge. We consider a smooth edge of shape $z_{\pm}(r) = \pm \ell(r)/2$ with $r$ measured from the sample edge, rising to $\ell = d$ within a distance $r \approx d/2$ (e.g., $\ell(r<d/2) = \sqrt{2rd}$). The (tangential) field ${H_{\mathrm{edge}}}$ at the surface is assumed constant and generates a current density $j =c {H_{\mathrm{edge}}}/4 \pi\lambda$ at the sample boundary, with $\lambda$ denoting the London penetration depth, $\lambda \ll d$. A simple geometrical consideration provides us with the sheet current $I(r) = 2 (c{H_{\mathrm{edge}}}/4\pi) \sqrt{1+[\ell'(r)/2]^2}$ and using [Eq. ]{}, we obtain the rise of the vortex energy near the edge $$\begin{aligned} \label{eq:Ugeo-edge-general} U_{\mathrm{geo}}(r) &= \varepsilon_l \ell(r) - \frac{\Phi_{0}{H_{\mathrm{edge}}}}{2\pi} \int\limits_{0}^{r} du\, \sqrt{1+[\ell'(u)/2]^2}.\end{aligned}$$ For a smooth edge with radius of curvature $\gtrsim d$ we have $\ell' \gg 1$ for $r \ll d$ (consistent with a roughly constant field ${H_{\mathrm{edge}}}$) and we can simplify the above expression to read $$\begin{aligned} U_{\mathrm{geo}}(r) &= \varepsilon_{l} \ell(r) \Big[1 - \frac{\Phi_{0}{H_{\mathrm{edge}}}}{4\pi \varepsilon_{l}}\Big].\end{aligned}$$ Hence, we find that the energy barrier for vortex entry is eliminated when the local field strength reaches the first critical field ${H_{\mathrm{edge}}}= H_{c1} = 4\pi \varepsilon_{l}/\Phi_{0}$. Once the edge region of width $d$ has been overcome, the vortices are driven to the sample center where they arrange within the vortex dome. The vortices deep inside the sample reduce the field at the edge and the penetration of flux is stopped when ${H_{\mathrm{edge}}}$ drops below $H_{c1}$. With a further increase of the external field, vortices continue to penetrate the sample when the condition ${H_{\mathrm{edge}}}= H_{c1}$ is satisfied again. This stop and go criterion for vortex penetration then is the second condition imposed on the fields in [Eqs. ]{} and and determines, together with [Eq. ]{}, the parameters ${b_{0}}$, ${b_{1}}$, and ${b_{2}}$. The above discussion ignores the possible presence of a surface barrier[@Clem_74] appearing on small length scales below $\lambda$. In the most effective case, this barrier further retards the penetration of vortices until the local field reaches the critical strength ${H_{\mathrm{edge}}}\sim H_c$. In order to deal with the general situation accounting for effects due to a surface barrier we denote the local critical field for vortex penetration by ${H_{s}}$ ($H_{c1} < {H_{s}}< H_{c}$). The second condition determining the fields [Eqs. ]{} and in the penetrated ($H > {H_{p}}$) state then can be cast in the form $$\begin{aligned} \label{eq:condition-penetration-field} {H_{\mathrm{edge}}}&= {H_{s}}.\end{aligned}$$ The above equation replaces the condition [Eq. ]{} valid for the Meissner phase. In the regime of very high fields, $H > {H_{s}}$, diamagnetic screening becomes small and the field strength at the sample edge lines up with the applied field, ${H_{\mathrm{edge}}}\approx H$; however, this large-field limit will not be considered below. Finally, we comment on the precision of this second condition: The field strengths in [Eqs. ]{} and show square-root singularities near the sample edges. The description of the spacial dependence of the field when approaching the edges to distances smaller than $d$ then requires a detailed analysis of the edge region. On the other hand, the typical scale for the field strength needed for overcoming the edge region can be obtained by the considerations presented above, once we have a proper definition for the edge field ${H_{\mathrm{edge}}}$ at our disposal. Below, we identify this field strength with the field evaluated a distance $d/2$ away from the edge, ${H_{\mathrm{edge}}}= B_{z}(r=-d/2)$. The surface barrier retarding the penetration of flux appears on the small length scale between $\lambda$ (at low fields of order $H_{c1}$) and $\xi$ (near $H_c$). On the contrary, the *geometric energy barrier* is a macroscopic object appearing on the scale $d$. We define the *geometric barrier* $U_{b}$ of a platelet sample as the maximum of [Eq. ]{} that is reached near $d$. The second term in [Eq. ]{} then reduces the geometric barrier linearly to zero at ${H_{\mathrm{edge}}}= H_{c1}$ and the barrier takes the functional form $$\begin{aligned} \label{eq:geometric-barrier} U_{b} = \varepsilon_{l} d \Big(1 - \frac{{H_{\mathrm{edge}}}}{H_{c1}}\Big) = \varepsilon_{l} d \Big(1 - \frac{H}{{H_{p}}}\frac{{H_{s}}}{H_{c1}}\Big)\end{aligned}$$ where the first (second) equality expresses the barrier in terms of the local (asymptotic) field (note that field penetration only starts when ${H_{\mathrm{edge}}}= {H_{s}}$, where the additional surface barrier has disappeared). While the geometric barrier only vanishes when the local field reaches $H_{c1}$, the vortex state may become thermodynamically stable at a lower *equilibrium* field ${H_{\mathrm{eq}}}$, defined as the applied field where a global minimum of the energy profile [Eq. ]{} develops inside the sample. For the single (double) strip, this minimum appears at $x_{0} = 0$ ($x_{0} = {b}$) and ${H_{\mathrm{eq}}}$ is determined from the condition $$\begin{aligned} \label{eq:def-heq} \varepsilon_{l} d - \frac{\Phi_{0}}{c} \int\limits_{e}^{x_{0}} du\, I(u)\Bigg|_{H = {H_{\mathrm{eq}}}} \!\!\!\!= 0,\end{aligned}$$ where $e$ denotes the position of the sample edge penetrated first, $e = w$ for the single strip and $e=s$ for the double strip, see Sec.\[sec:sec:double-strip\]. The geometrical barrier at the thermodynamic field ${H_{\mathrm{eq}}}$ $$\begin{aligned} \label{eq:eq-geometric-barrier} U_{b}^{\mathrm{eq}} = \varepsilon_{l} d \Big(1 - \frac{{H_{\mathrm{eq}}}}{{H_{p}}}\frac{{H_{s}}}{H_{c1}}\Big)\end{aligned}$$ then provides us with a measure for the irreversibility of the system, see Fig.\[fig:geometric-barrier\]. Having analyzed and determined the conditions determining the parameters ${b_{0}}$, ${b_{1}}$, and ${b_{2}}$ in the expressions and for the magnetic field, we now are in a position to evaluate the magnetic response (magnetization) of the sample. For this purpose, we make use of Ampère’s law and write the holomorphic field in the form (Biot-Savart, see also Ref. [\[\]]{}) $$\begin{aligned} \label{eq:biot-savart} \mathcal{B}(\xi) &= H - \frac{2}{c}\int_{\text{strips}}\!\!\!\! du \, \frac{I(u)}{\xi - u}.\end{aligned}$$ This field assumes the asymptotic form (we expand for $|\xi|\gg w$) $$\begin{aligned} \label{eq:multipole-expansion} \mathcal{B}(\xi) &= H - \frac{2}{c\, \xi^{2}}\int_{\text{strips}}\!\!\!\!\!\! du\, u\, I(u) + \mathcal{O}(\xi^{-4}),\end{aligned}$$ where we have used that the total current in each strip vanishes. The second term in [Eq. ]{} describes the field of a line of magnetic dipoles distributed along the $y$-axis ($\xi = 0$). We thus identify the magnetization $M$ per unit length (from here on called magnetization) with the expression $$\begin{aligned} \label{eq:definition-magnetization} M &= \frac{1}{c} \int_{\mathrm{strips}}\!\!\!\!du\,u\, I(u).\end{aligned}$$ This result differs from the usual textbook formula[@Jackson_62; @Landau_60] $$\begin{aligned} \label{eq:magnetization-landau-lifshitz} \mathcal{M} = \frac{1}{2c} \int d^{3}r\;\boldsymbol{r} \times \boldsymbol{j}(\boldsymbol{r})\end{aligned}$$ relating the total magnetic moment $\mathcal{M}$ to its generating current density $\boldsymbol{j}(\boldsymbol{r})$ flowing in a loop. The translation invariant 2D result can easily be shown to be consistent with the 3D textbook formula for a finite size ($2L$ along $y$) strip taking also into account the currents $j_{x}(y)$ flowing near the ends $y = \pm L$ of the strips and closing the loop. Formally expanding the left-hand side of [Eq. ]{} in $\xi^{-2}$ and comparing terms, the magnetization can be rewritten as $$\begin{aligned} \label{eq:magnetization-through-B} M(H) &= -\frac{1}{2} \left.\frac{\partial \mathcal{B}(\xi)}{\partial (1/\xi^{2})}\right|_{\xi^{-2} \to 0}\!.\end{aligned}$$ The magnetic responses of the single and double strip geometries \[as obtained from [Eqs. ]{}, , and \] take the particularly simple form $$\begin{aligned} \label{eq:magnetization-via-B-single} M(H) &= -\frac{H}{4} (w^{2}-{b_{0}}^{2}),\\ \label{eq:magnetization-via-B-double} M(H) &= -\frac{H}{4} (W^{2} + s^{2} - {b_{1}}^{2} - {b_{2}}^{2}).\end{aligned}$$ Single strip {#sec:sec:single-strip} ------------ We briefly review the physics of geometrical barriers for a single strip derived by Zeldov and co-workers[@Zeldov_94]. The function $\mathcal{B}(\xi)$, holomorphic in the superconductor-free region and satisfying the required boundary conditions, is given by [Eq. ]{}. On the $x$-axis ($z = 0$), the magnetic field component along $z$ is given by $$\begin{aligned} \label{eq:single-field-general} B_{z}(x) &= \left\{\begin{aligned} &H\sqrt{\frac{{b_{0}}^{2}-x^{2}}{w^{2}-x^{2}}} && \mathrm{for\ } |x| \leq {b_{0}},\\ &H\sqrt{\frac{x^{2}-{b_{0}}^{2}}{x^{2}-w^{2}}} && \mathrm{for\ }w \leq |x|,\\ &0 && \mathrm{for\ }{b_{0}}\leq |x| \leq w. \end{aligned}\right.\end{aligned}$$ The region $|x| \leq {b_{0}}$ describes the field-penetrated part of the sample where $B_z$ is finite. The current $I(x)$ flows in the complementary regions ${b_{0}}\leq |x| \leq w$ inside the strip; making use of [Eq. ]{} and Ampere’s law in the form of [Eq. ]{} we obtain the current $$\begin{aligned} \label{eq:single-current-general} I(x) &= -\frac{H c}{2\pi} \frac{x}{|x|} \sqrt{\frac{x^{2}-{b_{0}}^{2}}{w^{2}-x^{2}}}.\end{aligned}$$ The anti-symmetry of $I(x)$ guarantees the vanishing of the total current as required by [Eq. ]{}. The diamagnetic response resulting from these currents can be obtained with the formula given in [Eq. ]{} or directly via [Eq. ]{}. ### Meissner state {#sec:sec:sec:Meissner-1} In the Meissner state the field is fully expelled from the strip, ${b_{0}}=0$, and [Eqs. ]{} and simplify to $$\begin{aligned} \label{eq:single-field-meissner} B_{z}(x) &= \left\{\begin{aligned} &H\frac{x}{\sqrt{x^{2}-w^{2}}} && \mathrm{for\ }w \leq |x|,\\ &0 && \mathrm{for\ } |x| \leq w., \end{aligned}\right.\end{aligned}$$ and $$\begin{aligned} \label{eq:single-current-meissner} I(x) &= -\frac{H c}{2\pi} \frac{x}{\sqrt{w^{2}-x^{2}}},\end{aligned}$$ respectively. This anti-symmetric current density preserves the Meissner state and is identical to the one for the elliptic strip discussed before, see [Eq. ]{}. The divergencies in [Eq. ]{} at $x =\pm w$ have to be cut at the distance $\sim d$ away from the edges and we choose the specific value $d/2$. The local field strength at the edge (we drop corrections of higher order in $d/w$) $$\begin{aligned} \label{eq:single-field-enhancement} {H_{\mathrm{edge}}}&\equiv B_{z}(w + d/2) \simeq H\sqrt{\frac{w}{d}}\end{aligned}$$ then is enhanced by the factor $\sqrt{w/d}$. This enhancement is parametrically smaller as compared to the flat ellipsoid with corresponding dimensions where the enhancement factor is $2w/d$. The response of the superconducting strip in the Meissner state produces the magnetization \[see [Eq. ]{} with ${b_{0}}= 0$\] $$\begin{aligned} \label{eq:single-magnetization-meissner} M(H) &= - \frac{H}{4} w^{2},\end{aligned}$$ corresponding to the expulsion of the field $H$ from a region of size $\sim w^{2}$. Similar to the currents, the diamagnetic response is identical with that of an elliptic sample, see [Eq. ]{}. The Meissner state becomes unstable at $H = {H_{p}}$ as determined by the condition [Eq. ]{}; with the field enhancement given in [Eq. ]{}, we find $$\begin{aligned} \label{eq:single-penetration-field} {H_{p}}& \simeq {H_{s}}\sqrt{\frac{d}{w}}\end{aligned}$$ and the (maximum) magnetization at penetration reads $$\begin{aligned} \label{eq:single-m-at-penetration-field} M_{p} &= - \frac{{H_{s}}}{4} w^2 \sqrt{\frac{d}{w}} = - \frac{{H_{p}}}{4} w^2.\end{aligned}$$ As discussed above, the precise value for ${H_{p}}$ depends on the details of the edge geometry; the latter will modify the result by a numerical factor of order unity and affect all further results in this section in a straightforward way. For an elliptic strip, the larger field enhancement near the edges causes the penetration field [Eq. ]{} to be parametrically ($\sim\!\! \sqrt{d/w}$) smaller than that of the platelet sample. Although penetration is delayed to ${H_{p}}$, a field-filled state is thermodynamically stable (yet inaccessible due to the geometric barrier) beyond the equilibrium field $$\begin{aligned} \label{eq:heq-single} {H_{\mathrm{eq}}}= H_{c1} \frac{d}{2w}\end{aligned}$$ as obtained from evaluating [Eq. ]{}. The geometric barrier height \[from [Eq. ]{} with ${H_{s}}= H_{c1}$\] at that specific field amounts to $$\begin{aligned} \label{eq:eq-barrier-single} U_{b}^{\mathrm{eq}} = \varepsilon_{l} d \, \big(1 - \sqrt{d/4w}\big).\end{aligned}$$ ### Penetrated state {#sec:sec:sec:Shubnikov-1} Increasing the external field $H$ beyond ${H_{p}}$, vortices accumulate inside the strip in a dome-like density distribution of width $2{b_{0}}$. The field (current) profile along the $x$-axis ($z=0$) is given by the general form \[\]. The absence of a net current inside the strip is satisfied by symmetry, $I(-x) = -I(x)$. The evolution $$\begin{aligned} \label{eq:single-dome-width} {b_{0}}^{2}(H) &\simeq w^{2} [1-({H_{p}}/H)^{2}]\end{aligned}$$ of the dome width as a function of the applied field $H$ is determined by imposing a critical field strength at the edges, i.e., by solving [Eq. ]{} for ${H_{\mathrm{edge}}}= B_{z}(w + d/2)$. The induction in the vortex dome takes the maximal value $({b_{0}}/w)H$ at the gap center. For a largely penetrated strip, $w-{b_{0}}\ll w$, the induction is almost uniform and equal to the external field, $B(x) \approx H$. The presence of vortices inside the superconductor reduces the diamagnetic response, see [Eq. ]{} $$\begin{aligned} \label{eq:single-penetrated-magnetization} M(H) &\simeq - \frac{{H_{p}}^{2}}{4H} w^{2} = - \frac{{H_{s}}^{2}}{4H} w d .\end{aligned}$$ The applicability of the expressions and is limited to the regime where the screening currents flow in regions much wider than the sample thickness ($w-{b_{0}}\gg d$), a limit reached when the external field $H$ is very large, of order ${H_{s}}$. At this point, the strip is almost uniformly penetrated by the field with $B_{z} \approx {H_{s}}$, while the remaining screening currents flow in a narrow region of width $\sim d$ near the edges, maintaining a diamagnetic response $$\begin{aligned} \label{eq:magnetization-at-full-penetration} M({H_{s}}) &\approx - \frac{{H_{p}}}{4} w^{2} \sqrt{\frac{d}{w}}.\end{aligned}$$ Predictions on the system’s behavior for very large applied fields $H > {H_{s}}$ require a precise knowledge of the field distribution near the sample edge, a topic which is beyond our present analysis. The penetration process of vortices across a geometric energy barrier in a platelet strip features a hysteretic behavior[@Zeldov_94; @Zeldov_94_2]; upon reduction of the external field from a maximal value $H^{\star}$, the flux $\phi_{d}^{\star}=\phi_{d}(H^{\star})$ of vortices through the sample, where $$\begin{aligned} \label{eq:definition-flux-through-dome} \phi_{d} = \int\limits_{-{b_{0}}}^{{b_{0}}} dx\, B_{z}(x),\end{aligned}$$ is trapped unless the vortex dome boundaries reach the sample edges. Evaluating the above flux with the field , we find $$\begin{aligned} \label{eq:flux-through-dome} \phi_{d} = 2wH\Big[\operatorname{E}({b_{0}}/w) - \frac{w^{2}-{b_{0}}^{2}}{w^{2}}\operatorname{K}({b_{0}}/w)\Big],\end{aligned}$$ with $\operatorname{K}$ ($\operatorname{E}$) the complete elliptic integral of the first (second) kind defined according to standard textbooks on mathematical functions; e.g., see Eqs. (17.2.18)-(17.3.3) of Ref. [\[\]]{}, $$\begin{aligned} \label{eq:K} \operatorname{K}(\kappa) &= \int_{0}^{\pi/2} \frac{d\theta}{\sqrt{1-\kappa^{2}\sin\!{}^{2}(\theta)}},\\ \label{eq:E} \operatorname{E}(\kappa) &= \int_{0}^{\pi/2} d\theta \sqrt{1-\kappa^{2} \sin\!{}^{2} (\theta)}.\end{aligned}$$ For $\kappa \ll 1$, the elliptic functions show the limiting behavior $$\begin{aligned} \label{eq:K-small} \operatorname{K}(\kappa) &= \frac{\pi}{2}\Big[ 1 + \frac{\kappa^{2}}{4} + \frac{9\kappa^{4}}{64}+ \mathcal{O}(\kappa^{6}) \Big],\\ \label{eq:E-small} \operatorname{E}(\kappa) &= \frac{\pi}{2}\Big[ 1 - \frac{\kappa^{2}}{4} - \frac{3\kappa^{4}}{64}+ \mathcal{O}(\kappa^{6}) \Big],\end{aligned}$$ while for the opposite limit, $\kappa = \sqrt{1-\nu}$ with $\nu \ll 1$, we find $$\begin{aligned} \label{eq:K-large} \operatorname{K}(\sqrt{1-\nu}) &= \frac{1}{2}\log\Big(\frac{16}{\nu}\Big) - \frac{\nu}{8}\Big[2 - \log\Big(\frac{16}{\nu}\Big)\Big] + \mathcal{O}(\nu^{2}),\\ \label{eq:E-large} \operatorname{E}(\sqrt{1-\nu}) &= 1 - \frac{\nu}{4}\Big[1 - \log\Big(\frac{16}{\nu}\Big)\Big] + \mathcal{O}(\nu^{2}).\end{aligned}$$ The constraint $\phi_{d}(H<H^{\star}) = \phi_{d}^{\star}$ reduces to a condition for the dome width ${b_{0}}(H)$ of the form $$\begin{aligned} \label{eq:descending-branch-condition-single} \operatorname{E}({b_{0}}/w) - \frac{w^{2}-{b_{0}}^{2}}{w^{2}}\operatorname{K}({b_{0}}/w) &= \frac{\phi_{d}^{\star}}{2w\, H},\end{aligned}$$ in agreement with Ref. [\[\]]{}. The left-hand side is limited by unity from above (for ${b_{0}}= w$). Upon decreasing $H$, the vortex dome expands over the sample until reaching the edge. Since for a large dome, $w - {b_{0}}\ll w$, the induction is uniform and equal to $H$, we find that vortices leave the sample at $H = H_{\mathrm{ex}}= \phi_{d}^{\star}/2w$ where formally ${b_{0}}= w$ and $M=0$. ![The magnetization for the descending field branches are shown for different values of the turning field $H^{\star}$ ($H^{\star}/{H_{p}}$ = 1.25, 1.5, 2.5, 3). The numerical solution of [Eq. ]{} (thick solid lines) is compared to the magnetic response obtained from small domes (thin solid lines) featuring a constant Meissner slope, see [Eq. ]{}. For $H^{\star}$ = 1.25${H_{p}}$ and 1.5${H_{p}}$, the dotted curves show the magnetic response as obtained from a next-to-leading order expansion of [Eq. ]{} in ${b_{0}}/w$, see [Eq. ]{}. For $H^{\star}$ = 2.5${H_{p}}$ and 3${H_{p}}$, where the dome is sufficiently large at $H^{\star}$, i.e., $\nu^{\star} = ({H_{p}}/H^{\star}) \ll 1$, the magnetization is well described by the expression shown as dashed lines.[]{data-label="fig:descending-branch-single"}](plot_19.eps){width="48.00000%"} For a narrow dome ${b_{0}}/w \ll 1$, the above condition can be simplified using the asymptotic expressions and for the elliptic functions; to lowest (quadratic) order in $\kappa = {b_{0}}/w$, we find $$\begin{aligned} \label{eq:descending-branch-single-small-dome-1} \frac{{b_{0}}^{2}}{w^{2}} &= \frac{4}{\pi}\frac{H_{\mathrm{ex}}}{H}.\end{aligned}$$ The growth of the dome width ${b_{0}}^{2} = {b_{0}}^{\star}{}^{2} H^{\star}/H$ \[with ${b_{0}}^{\star} = {b_{0}}(H^{\star})$\] results in a magnetic response of the form $$\begin{aligned} \label{eq:magnetization-descending-single-small-dome-1} M(H) &= -\frac{H}{4} w^{2} + \frac{H^{\star}}{4} {{b_{0}}^{\star} }^2\end{aligned}$$ with a slope identical to the Meissner state. Higher order corrections (quartic in $\kappa = {b_{0}}/w$) are straight forwardly obtained from [Eqs. ]{} and : the condition then yields $$\begin{aligned} \label{eq:descending-branch-single-small-dome-4} \frac{{b_{0}}^{2}}{w^{2}} &= 4 \bigg\{ \sqrt{1 + \frac{H^{\star}}{H} \Big[ \Big(1 + \frac{1}{4}\frac{{b_{0}}^{\star}{}^{2}}{w^{2}} \Big)^{2} - 1 \Big] } -1 \bigg\}.\end{aligned}$$ Inserting this solution into the expression for the magnetization, the Meissner slope is corrected according to $$\begin{aligned} \nonumber \frac{dM}{dH} &= - \frac{w^{2}}{4} \bigg\{ 1 - \frac{1}{2}\Big(\frac{H^{\star}}{H}\Big)^{2} \Big[ \Big(1 + \frac{1}{4}\frac{{b_{0}}^{\star}{}^{2}}{w^{2}} \Big)^{2} - 1 \Big]^{2} \bigg\}\\ &\approx - \frac{w^{2}}{4} \Big[1- \frac{1}{8} \Big( \frac{H^{\star}}{H} \frac{{b_{0}}^{\star}{}^{2}}{w^{2}} \Big)^{2} \Big].\end{aligned}$$ The rapid growth of the dome-width on both the field-increasing (filling the dome with additional flux) and decreasing (expanding the dome at fixed flux) branches leads to a fast violation of the condition ${b_{0}}\ll w$ assumed above and hence these results have a rather limited range of validity. Another limit is reached when ${b_{0}}$ is large, $w - {b_{0}}\ll w$. Defining $\nu = 1 - {b_{0}}^{2}/w^{2}$, the asymptotic expressions and can be used to simplify (up to linear order in $\nu$) the condition to $$\begin{aligned} \label{eq:simplified-returning-branch} 1 - \frac{\nu}{4} \Big[\log\Big(\frac{16}{\nu}\Big) + 1\Big] &= \frac{H_{\mathrm{ex}}}{H}.\end{aligned}$$ In most of the $M$-$H$-diagram, the system’s magnetic response on the descending branch then is given by $M(H) = -H w^{2} \nu(H)/4$. Taking the derivative of $M$ with respect to $H$, the slope of the descending branch can be evaluated and, after some reordering, we find that $$\begin{aligned} \label{eq:slope} \frac{dM}{dH} &= -\frac{w^{2}}{4} \frac{4 - \nu}{\log(16/\nu)}.\end{aligned}$$ The derivative deviates from the Meissner slope $-w^{2}/4$ by a numerical factor which assumes the value $\approx 1.01$ for $\nu \sim 1/2$, when the previous approach of a narrow dome predicts a perfect Meissner slope \[see [Eq. ]{}\]. In the regime of applicability, where $\nu$ may change by several orders of magnitude, the factor $(4-\nu)/\log(16/\nu)$ changes noticeably but not parametrically. Typically, the slope of the descending branch is numerically close to the Meissner slope within the parameter range under consideration, see Fig. \[fig:descending-branch-single\]. For large reversal fields $H^{\star} \gg {H_{p}}$, we replace the parameter $\nu(H)$ by its value at the field reversal $\nu(H^\star) = \nu^{\star} = 1 - {b_{0}}^{\star}{}^{2}/w^{2} = ({H_{p}}/H^\star)^{2}$, where we have used [Eq. ]{}. The magnetization $$\begin{aligned} \label{eq:analytic-returning-single-strip} M(H) = M(H^{\star}) - \frac{H-H^{\star}}{4} w^{2} \frac{4 - \nu^{\star}}{\log(16/\nu^{\star})} \end{aligned}$$ as obtained from [Eq. ]{} and integration from $H^{\star}$ to $H$ provides a good description of the descending branch in this regime, see [Fig. ]{}\[fig:descending-branch-single\]. As the boundaries of the dome approach the edges of the strip to a distance $\sim d$ (which is the case when $H \approx [1+ \mathcal{O}(d/w)] H_{\mathrm{ex}}$) the precise geometric shape of the sample edge needs to be taken into account, requiring a more accurate analysis going beyond the present description. An attempt to cope with this situation has been undertaken by Zeldov and co-workers in Refs.[\[\]]{}. ### Magnetization of the vortex dome {#sec:sec:sec:magnetic-medium-1} The physical properties of quantized flux lines appeared in the above analysis merely as a criterion for vortex entry at the sample edges. The vortex dome in the penetrated state has been described by a smooth field $B_z(x) \neq 0$ residing in a magnetically inactive medium with $\mu = 1$ whose extend $[-b_0,b_0]$ derives from the solution ${\cal B}(\xi)$ of the boundary value problem. In reality, the vortex state in the dome is described by a field $h(x)$ modulated on the scale of the inter-vortex distance due to vortex currents. In the following, we show that the currents associated with the vortex state in the dome generate a magnetization which remains small as compared to the magnetization produced by the screening currents flowing in the field-free regions. An analogous problem appears in the context of surface barriers as discussed by Clem[@Clem_74] and by Koshelev[@Koshelev_94]: quite similar to our analysis, in Ref. [\[\]]{} the vortex-penetrated bulk, separated from the boundary by a layer of screening (Meissner) currents, has been described by an induction $B_z$ averaged over the inter-vortex spacing. This approximation neglects all field and current modulations due to the vortex state and the resulting magnetization density is given by[@Clem_74] $$\begin{aligned} \label{eq:magnetization-surf-barrier-Clem} m(H) &= -\frac{H}{4\pi} \big[1 - \sqrt{1 - ({H_{s}}/H)^{2}}\,\big].\end{aligned}$$ A way to account for the local currents in the vortex state has been proposed by Koshelev[@Koshelev_94], who found that these contribute a paramagnetic correction $\delta m = (\sqrt{3}/48) (\Phi_{0}/4\pi \lambda^{2})$ to the magnetization density $m(H)$ in the limit $B \gg \Phi_{0}/\lambda^{2}$. Following a similar ideology as in Ref. [\[\]]{}, we describe the flux-filled region in terms of a vortex lattice along $z$ with vortex rows aligned along $y$ and separated by ${b_{\scriptscriptstyle{\triangle}}}$ in the $x$-direction with ${b_{\scriptscriptstyle{\triangle}}}^{2} = (3/4)^{1/2}\Phi_{0}/B_{z}$. While in Ref.[\[\]]{} $B_z(x)$ was determined self-consistently, here, we estimate the corrections to the magnetization by adopting the averaged field $B_z(x)$ obtained from the above analytic solution. In our strip geometry, the spacing ${b_{\scriptscriptstyle{\triangle}}}$ between vortex-rows slowly varies along $x$, as the induction $B_{z}$ changes on macroscopic length scales. The connection between the local field $h(x)$ and the induction $B_{z}(x)$ is given by the average $$\begin{aligned} \label{eq:average-induction} B_{z}(x) = \frac{1}{{b_{\scriptscriptstyle{\triangle}}}}\int\limits_{x-{b_{\scriptscriptstyle{\triangle}}}/2}^{x+{b_{\scriptscriptstyle{\triangle}}}/2} dx' h(x').\end{aligned}$$ The local field $h(x)$ satisfies the one-dimensional London equation $\lambda^{2} h''(x) + h(x) = 0$ between the vortex rows with the boundary conditions replaced by the constraint . For a slowly varying dome profile, i.e., ${b_{\scriptscriptstyle{\triangle}}}\partial_{x} B_{z}(x) \ll B_{z}(x)$, we obtain the field modulation between vortex rows $$\begin{aligned} h(x) &\approx B_{z}(x_{c}) \frac{{b_{\scriptscriptstyle{\triangle}}}}{2\lambda} \frac{\cosh[(x-x_{c})/\lambda]}{\sinh({b_{\scriptscriptstyle{\triangle}}}/2\lambda)},\end{aligned}$$ with $x_{c}$ the center between the two adjacent rows and $|x-x_{c}| < {b_{\scriptscriptstyle{\triangle}}}/2$. Ampère’s law then provides us with the current profile $$\begin{aligned} j(x) &\approx - \frac{B_{z}(x_{c})c}{4\pi} \frac{{b_{\scriptscriptstyle{\triangle}}}}{2\lambda^{2}} \frac{\sinh[(x-x_{c})/\lambda]}{\sinh({b_{\scriptscriptstyle{\triangle}}}/2\lambda)}\end{aligned}$$ and we can evaluate the associated average magnetization density at the vortex location $x_v$ $$\begin{aligned} m(x_{v}) &\approx \frac{1}{{b_{\scriptscriptstyle{\triangle}}}c} \int\limits_{x_{v}-{b_{\scriptscriptstyle{\triangle}}}/2}^{x_{v}+{b_{\scriptscriptstyle{\triangle}}}/2} dx' x' j(x')\\ \label{eq:material-magnetization} &\approx \frac{B_{z}(x_{v})}{4\pi} \Big[1 - \frac{{b_{\scriptscriptstyle{\triangle}}}}{2\lambda} \frac{1}{\sinh({b_{\scriptscriptstyle{\triangle}}}/2\lambda)}\Big].\end{aligned}$$ For small fields $B_{z} \ll H_{c1}$, we find that $m(x) \approx B_{z}(x)/4\pi$, while the magnetization density saturates at $(\Phi_{0}/4\pi\lambda^{2}) \sqrt{3}/48$ for large fields $B_{z} \gg H_{c1}$, consistent with the results presented in Ref. [\[\]]{}. In order to estimate the correction to the strips’ magnetic response, we introduce the upper bound $$\begin{aligned} \label{eq:ub-magnetization} m(x) \leq \frac{B_{z}(x)}{4\pi} \frac{H_{c1}}{H_{c1}+B_{z}(x)}\end{aligned}$$ with the correct asymptotic behavior for $B \ll H_{c1}$ and logarithmically \[$\propto \log(\lambda/\xi)$\] overestimating the magnetization when $B \gg H_{c1}$. Integrating $m(x)$ over the dome and replacing the dome profile $B_z(x)$ by its maximum $H {b_{0}}/w$ at the center, see [Eq. ]{}, we obtain the bound $$\begin{aligned} \label{eq:magnetic-correction-estimated} \delta M < \frac{H}{4\pi} \frac{H_{c1}}{H_{c1} + H {b_{0}}/w} \frac{{b_{0}}^{2}}{w^{2}}2w d.\end{aligned}$$ For a small dome, ${b_{0}}\ll w$, this expression simplifies to $$\begin{aligned} \label{eq:magnetic-correction-estimated-low-field} \delta M < \frac{H}{4\pi} \Big(1 - \frac{{H_{p}}^{2}}{H^{2}} \Big)2w d,\end{aligned}$$ whereas for a large part of the penetrated region $d \ll w-{b_{0}}(H) \ll w$, we find that $$\begin{aligned} \label{eq:magnetic-correction-estimated-high-field} \delta M < \frac{H}{4\pi} \frac{H_{c1}}{H_{c1} + H} 2w d.\end{aligned}$$ As a result, the correction $\delta M$ due to the reversible magnetization measured on the magnetization $M$ of the screening currents [Eq. ]{} is bounded from above by $$\begin{aligned} \label{eq:relative-corrections-single} \frac{\delta M}{M} &< \frac{2}{\pi} \frac{H^{2}}{{H_{s}}^{2}} \frac{H_{c1}}{H_{c1}+H}.\end{aligned}$$ In the absence of a surface barrier (${H_{s}}= H_{c1}$) these corrections are small and become of order unity at the largest fields $H \sim H_{c1}$ where our analysis applies. In the presence of a large surface barrier where ${H_{s}}\gg H_{c1}$, the corrections are even smaller and reach a maximum $\sim H_{c1}/{H_{s}}\ll 1$ when $H \sim {H_{s}}$. We conclude that the corrections arising from the vortex currents can be omitted in the single strip geometry. Double strip {#sec:sec:double-strip} ------------ We now investigate the double-strip configuration defined in [Fig. ]{}\[fig:sketch-double-strip\], a system of two coplanar, parallel strips of width $2w$ each and separated by a gap $2s$. Assuming a gap that is large as compared to the strip thickness, $s \gg d$, the system can be treated within the framework introduced in [Sec. ]{}\[sec:sec:formalism\]. The holomorphic function has been presented in [Eq. ]{}, from which the \[symmetric, $B_z(-x) = B_z(x)$\] field and \[anti-symmetric, $I(-x) = -I(x)$\] current distribution on the $x$-axis can be readily deduced $$\begin{aligned} \label{eq:double-field-general} \frac{B_{z}(x)}{H} &= \left\{ \begin{aligned} &\sqrt{\frac{({b_{1}}^{2}-x^{2})({b_{2}}^{2}-x^{2})} {(s^{2}-x^{2})(W^{2}-x^{2})}} && \mathrm{for\ } 0 \leq x \leq s,\\ &\sqrt{\frac{(x^{2}-{b_{1}}^{2})({b_{2}}^{2}-x^{2})} {(x^{2}-s^{2})(W^{2}-x^{2})}} && \mathrm{for\ } {b_{1}}\leq x \leq {b_{2}},\\ &\sqrt{\frac{(x^{2}-{b_{1}}^{2})(x^{2}-{b_{2}}^{2})} {(x^{2}-s^{2})(x^{2}-W^{2})}} && \mathrm{for\ } W \leq x,\\ &0 && \mathrm{otherwise}, \end{aligned}\right.\end{aligned}$$ and $$\begin{aligned} \label{eq:double-current-general} \frac{2\pi I(x)}{cH} &= \left\{ \begin{aligned} &\sqrt{\frac{({b_{1}}^{2}-x^{2})({b_{2}}^{2}-x^{2})} {(x^{2}-s^{2})(W^{2}-x^{2})}} && \mathrm{for\ } s \leq x \leq {b_{1}},\\ &-\sqrt{\frac{(x^{2}-{b_{1}}^{2})(x^{2}-{b_{2}}^{2})} {(x^{2}-s^{2})(W^{2}-x^{2})}} && \mathrm{for\ } {b_{2}}\leq x \leq W,\\ &0 && \mathrm{otherwise}. \end{aligned}\right.\end{aligned}$$ The resulting magnetization is given by [Eq. ]{}. ### Meissner state {#sec:sec:sec:Meissner-2} In the (low-field) Meissner state the parameters ${b_{1}}$, ${b_{2}}$ in [Eq. ]{} coincide, ${b_{1}}= {b_{2}}= {b}$, with $\pm {b}$ marking the the positions inside the strips where the current density changes sign (see [Fig. ]{}\[fig:double-strip-meissner-d-ll-s\]). The magnetic field component $B_{z}$ \[from [Eq. ]{}\] is non-vanishing whenever $|x|\leq s$ or $W \leq |x|$ and reads $$\begin{aligned} \label{eq:double-field-meissner} B_{z}(x) &= H \frac{|x^{2}-{b}^{2}|} {\sqrt{(x^{2}-s^{2})(x^{2}-W^{2})}}.\end{aligned}$$ In the complementary region $s \leq |x| \leq W$, the screening current $$\begin{aligned} \label{eq:double-current-meissner} I(x) &= - \frac{H c}{2\pi} \frac{x}{|x|} \frac{x^{2}-{b}^{2}} {\sqrt{(x^{2}-s^{2})(W^{2}-x^{2})}}\end{aligned}$$ guarantees a perfect diamagnetic (Meissner) response $$\begin{aligned} \label{eq:double-strip-magnetization-meissner} M(H) &= -\frac{H}{4} (W^{2} + s^{2} - 2{b}^{2}),\end{aligned}$$ with ${b}$ independent of $H$. The condition that no net current flows along each strip requires that $$\begin{aligned} \label{eq:double-zero-current-cond} \int\limits_{s}^{W} \frac{dx\,x^{2}}{\sqrt{(x^{2}-s^{2})(W^{2}-x^{2})}} &= \int\limits_{s}^{W} \frac{dx\,{b}^{2}}{\sqrt{(x^{2}-s^{2})(W^{2}-x^{2})}},\end{aligned}$$ from which we find the value of ${b}$, $$\begin{aligned} \label{eq:double-zero-current-result} {b}^{2} &= W^{2} \frac{\operatorname{E}(\kappa')}{\operatorname{K}(\kappa')},\end{aligned}$$ in agreement with Ref. [\[\]]{}. Here, $\operatorname{K}$ ($\operatorname{E}$) is the complete elliptic integral of the first (second) kind, as defined in [Eq. ]{} \[\], and $\kappa' = \sqrt{1-\kappa^{2}}$ is the complementary modulus of $\kappa = s/W$. For large gaps, the double strip behaves as two independent strips: indeed, for $s/w \to \infty$, the parameter ${b}$ approaches the sample center $w+s$ and the magnetization assumes the asymptotic value $M(H) \to -H w^{2}/2$, twice that of an isolated strip, see [Eq. ]{}. ![Normalized current density $2 \pi I(x)/H c$ (solid line) flowing along the $y$-direction and dimensionless magnetic field $B_{z}(x)/H$ (dashed line) of a double-strip in the Meissner state. The two strips with width $2w$ and thickness $d$ ($d/w \to 0$) are separated by a gap $2s$ (here $w/s=100$). According to [Eqs. ]{} and , the local current reverts its sign at $\pm {b}$, with ${b}\approx 0.38 \, W$. The magnetic field inside the gap between the strips (see inset) is far above the range of this graph, $B_{z}(|x|<s)/H \geq {b}^{2}/Ws \approx 30$.[]{data-label="fig:double-strip-meissner-d-ll-s"}](plot_01_mod.eps){width="48.00000%"} Let us then focus on the opposite limit $s \ll W=2w+s$, where the right hand side of [Eq. ]{} shows a logarithmic divergence $\propto \log(W/s)$, while the left hand side is regular; in this limit, the parameter ${b}$ takes the asymptotic form $$\begin{aligned} \label{eq:a-over-W-ratio} {b}^{2} &= \frac{W^{2}}{\log{(4W/s)}},\end{aligned}$$ and the position ${b}$ where the current $I(x)$ changes sign is no longer at the sample center but has shifted towards the inner edge, see Figs.\[fig:double-strip-meissner-d-ll-s\], \[fig:b1-n-b2\] and \[fig:a-of-s\]. The magnetization (per unit length) to leading order in $s/W$ reads, $$\begin{aligned} \label{eq:meiss-mag-double-strip} M(H) &= -\frac{H}{4}W^{2}\Big[1 - \frac{2}{\log(4W/s)}\Big].\end{aligned}$$ In the limit $s/W \to 0$, the Meissner slope approaches that of a single strip with double width, see [Eq. ]{}. We conclude that over the full range of gap widths $s$ (from $s \gg W$ down to $s/W \to 0$) the slope in the magnetization of the Meissner state increases only by a factor 2. For the double strip geometry, the flux (per unit length) $\phi_{g}$ passing through the gap $|x| < s$ is defined as the $z$-component of the magnetic field integrated over the gap width, $$\begin{aligned} \label{eq:flux} \phi_{g} &=\!\!\int\limits_{-s}^{s}\! dx\,B_{z}(x) = 2 W \Big[\operatorname{E}(\kappa) - \Big(1-\frac{{b}^{2}}{W^{2}}\Big)\operatorname{K}(\kappa)\Big] H,\end{aligned}$$ where the elliptic functions are evaluated at $\kappa = s/W$. In the regime of almost independent strips, $s \gg W$, the flux $2s H$ of the homogeneous field in the empty gap region is enhanced by half of the flux $\phi_{b} = 4w H$ blocked by the two strips, thus adding up to $\phi_{g} \approx (2s + 2w)H$. In the opposite limit $s \ll W$, the expression for the flux in the gap simplifies to $$\begin{aligned} \label{eq:flux-t-s-l} \phi_{g} &\simeq \frac{\pi {b}^{2}}{W} H \simeq \frac{\pi W}{\log(4W/s)} H.\end{aligned}$$ An essential part (up to a logarithmic factor) of the blocked flux $\phi_{b} = 2W H$ is pushed through the gap. This slow reduction of $\phi_{g}$ upon reducing $s$ goes hand in hand with an enhancement of the field strength at the gap center $$\begin{aligned} \label{eq:double-field-enhancement-gap-center} B_{z}(0) &= H \frac{{b}^{2}}{s W} = \frac{2}{\pi}\frac{\phi_{g}}{2s} = H \frac{W/s}{\log(4W/s)}\end{aligned}$$ and near the inner edges $$\begin{aligned} \label{eq:double-field-enhancement} B_{z}(s - d/2) &\simeq H\frac{{b}^{2}}{\sqrt{sd}\,W} = H\frac{W/\sqrt{sd}}{\log(4W/s)}.\end{aligned}$$ This last expression is parametrically larger than the enhancement observed at the edge of an isolated strip, see [Eq. ]{}. Note that the field inside the gap is far from constant, but increases by a factor $\sqrt{s/d}$ from the gap center to one strip edge, see inset in [Fig. ]{}\[fig:double-strip-meissner-d-ll-s\]). On the other hand, the field strength near the outer edges $$\begin{aligned} \label{eq:double-field-enhancement-outer} B_{z}(W + d/2) &\simeq H \frac{W^{2}-{b}^{2}}{W\sqrt{Wd}} = H\sqrt{\frac{W}{d}}\Big[1 - \frac{1}{\log(4W/s)}\Big] \end{aligned}$$ is comparable to that of an isolated strip, see [Eq. ]{}. From this analysis we conclude that the local critical field ${H_{s}}$ is first reached near the inner edges, such that the penetration of vortices occurs from *inside*. The field of first penetration ${H_{p}}$ then is determined by the condition $$\begin{aligned} \label{eq:criticality-condition-double-strip} {H_{\mathrm{edge}}}&= B_{z}(s - d/2) = {H_{s}}\end{aligned}$$ and making use of [Eq. ]{} we find that the penetration field for small gaps $s \ll W$ $$\begin{aligned} \label{eq:double-penetration-field} {H_{p}}&\simeq {H_{s}}\sqrt{\frac{s d} {W^{2}}} \frac{W^{2}}{{b}^{2}} = {H_{s}}\sqrt{\frac{s d}{W^{2}}} \log{(4W/s)}\end{aligned}$$ is substantially reduced as compared to the one for isolated strips ${H_{p}}\simeq {H_{s}}\sqrt{d/w}$. As discussed for the single strip, see [Eq. ]{} and thereafter, the precise edge geometry will alter the above expression for ${H_{p}}$ by a numerical factor of order unity (the same factor as for the single strip), a correction that will be neglected in the following. At penetration $H={H_{p}}$, the Meissner state reaches the maximal diamagnetic response \[see [Eq. ]{}\] $$\begin{aligned} \label{eq:max-magnetization-double-thin-strip} M_{p} &= -\frac{{H_{s}}}{4}W \sqrt{s d}\, \big[\log{(4W/s)} - 2\big].\end{aligned}$$ Upon reducing the gap width $s$, the penetration field diminishes and the geometrical barrier is more strongly suppressed, see [Eq. ]{}. Vortices become energetically favorable (deep) inside the sample beyond the equilibrium field (we use [Eq. ]{} in the regime $s \ll W$) $$\begin{aligned} \label{eq:heq-double-thin} {H_{\mathrm{eq}}}= H_{c1} \frac{d}{2W} \Big\{1 - \frac{\log[4\log(4W/s)]+1}{2 \log(4W/s)}\Big\}^{-1},\end{aligned}$$ resulting in a geometric barrier at ${H_{\mathrm{eq}}}$ which decreases with $s$, $$\begin{aligned} \label{eq:eq-barrier-double-thin} \frac{U_{b}^{\mathrm{eq}}(s)}{\varepsilon_{l} d}= 1 - \frac{\sqrt{d/s}} {2 \log(4W/s) - \log[4\log(4W/s)]-1}.\end{aligned}$$ ### Penetrated state {#sec:sec:sec:Shubnikov-2-tsl} Increasing the external field beyond its critical value, ${H_{p}}$, vortices penetrate the superconductor from the inner edges at $x = \pm s$ and accumulate near the position ${b}$ inside the strips where the potential $U_{\mathrm{geo}}(x)$ is minimal. The field and currents take the general form given in [Eqs. ]{} and , with the non-trivial vortex state determined by the two boundaries of the vortex dome ${b_{1}}$ and ${b_{2}}$. ![Dimensionless field $B_{z}(x)/H$ (dashed line) and current $2\pi I(x)/H c$ (solid line) as a function of $x$ (for $z=0$) for the same geometry ($w/s = 100$ and $s/d = 100$) as in [Fig. ]{}\[fig:double-strip-meissner-d-ll-s\] and an external field $H = 1.2 {H_{p}}$ above first penetration. In the penetrated state above ${H_{p}}$, vortices accumulate in a finite region inside each strip (the vortex dome), with boundaries given by $\pm{b_{1}}$ and $\pm{b_{2}}$.[]{data-label="fig:double-strip-penetrated-d-ll-s"}](plot_05.eps){width="48.00000%"} These two parameters satisfy the constraint of vanishing net current in each strip $$\begin{aligned} \label{eq:double-no-net-current} \int_{s}^{{b_{1}}} dx\, I(x) + \int_{{b_{2}}}^{W} dx\, I(x) &= 0,\end{aligned}$$ together with the condition \[from [Eq. ]{}\] $$\begin{aligned} \label{eq:critical-condition-penetrated-double} {H_{\mathrm{edge}}}= B_{z}(s-d/2) &= {H_{s}}.\end{aligned}$$ While this constraint locks the field strength at the inner edge to ${H_{s}}$, the field strength near the outer edge continuously grows, but remains below ${H_{s}}$. In [Fig. ]{}\[fig:double-strip-penetrated-d-ll-s\] we show the field and current profiles in the penetrated state for $H = 1.2{H_{p}}$ as obtained from solving [Eqs. ]{} and numerically. The evolution of the dome’s boundaries and its width ${b_{2}}- {b_{1}}$ with increasing field is shown in [Fig. ]{}\[fig:b1-n-b2\]. The maximal field value in the dome can be estimated with the interpolation formula $B_\mathrm{dome} \sim H (b_2-b_1)/W$. In order to find analytic results describing the penetrated state, we have to simplify the problem of determining the parameters ${b_{1}}$, ${b_{2}}$. Evaluating the condition for the field and expressing the result through the penetration field ${H_{p}}$, the dome boundaries ${b_{1}}(H)$ and ${b_{2}}(H)$ are related via $$\begin{aligned} \label{eq:critical-current-condition} \frac{\sqrt{{b_{1}}^{2}-s^{2}}\ {b_{2}}}{{b}^{2}} = \frac{{H_{p}}}{H},\end{aligned}$$ where ${b}$ is the (field-independent) zero-current location in the Meissner state, [Eq. ]{}. It turns out that a perturbative calculation around the penetration field with the small parameter $h = (H-{H_{p}})/{H_{p}}\ll 1$ produces results with a very limited range of validity. This is due to the rapid growth of the dome width ${b_{2}}- {b_{1}}$ with increasing $h$, leading to a fast break-down of the approximation. Approaching the problem from the high field limit $H \gg {H_{p}}$ is more successful: starting from the regime where the dome extends over a large fraction of the strip ${b_{1}}\ll {b_{2}}$, we can adopt another perturbative approach which provides accurate results all the way down to ${H_{p}}$. We use the Ansatz ${b_{2}}= W (1-\nu)^{1/2}$ with $\nu(H) < 1$. For $s\ll {b_{1}}$, where [Eq. ]{} simplifies to $$\begin{aligned} \label{eq:appr-critical-current-condition} {b_{1}}= W \frac{{b}^{2}}{W^{2}} \frac{{H_{p}}}{H} \frac{1}{\sqrt{1-\nu}},\end{aligned}$$ the constraint of vanishing net current in the strips can be written as $$\begin{aligned} \label{eq:appr-neutral-current-condition} \frac{{H_{p}}}{H}\frac{{b}^{2}}{W^{2}} \left[\log{\Big(\frac{4W}{s} \frac{{b}^{2}}{W^{2}} \frac{{H_{p}}}{H} \frac{1}{\sqrt{1-\nu}}\Big)}-1\right]&\\[.5em] = \operatorname{E}(\sqrt{\nu})-&(1-\nu)\operatorname{K}(\sqrt{\nu}).\nonumber\end{aligned}$$ Solving this equation to leading order in ${H_{p}}/H$ where $\nu \ll 1$, we find that $$\begin{aligned} \label{eq:mu-t-s-l} \nu(H) &= \frac{4}{\pi}\frac{{H_{p}}}{H}\frac{{b}^{2}}{W^{2}} \left[\log{\Big(\frac{4W}{s} \frac{{b}^{2}}{W^{2}} \frac{{H_{p}}}{H}\Big)}-1\right].\end{aligned}$$ ![Evolution of the dome edges ${b_{1}}(H)$ and ${b_{2}}(H)$ with increasing field $H$ for parameters $w/s = 100$ and $s/d = 100$. For small fields $H < {H_{p}}$, the double strip is in the Meissner phase and ${b_{1}}= {b_{2}}= {b}$, where ${b}$ is shifted away from the sample center $w+s$. In the penetrated state $H>{H_{p}}$, vortices accumulate inside the strip and the dome width ${b_{2}}- {b_{1}}$ widens. The field and current profiles for the field $H = 1.2 {H_{p}}$ are shown in [Fig. ]{}\[fig:double-strip-penetrated-d-ll-s\].[]{data-label="fig:b1-n-b2"}](plot_13_mod.eps){width=".48\textwidth"} To leading order in $\nu(H)$, the magnetic response in [Eq. ]{} takes the form $M \simeq -H \nu(H) W^{2}/4$, resulting in a logarithmic field-dependence $$\begin{aligned} \label{eq:appr-mag-t-s-l-original-parameters} M(H) &\simeq -\frac{{H_{s}}}{4\pi} (2W d) \bigg\{2\sqrt{\frac{s}{d}} \Big[\log{\Big(\frac{4 {H_{s}}}{H}\sqrt{\frac{d}{s}\,}\Big)}-1\Big] \bigg\}.\end{aligned}$$ ![Magnetization of a double strip system obtained from numerical evaluation (solid line) and from the analytic solutions (dashed lines) for parameters $w/s = 100$ and $s/d = 100$. The expression in [Eq. ]{} is applicable in the field range ${H_{p}}\ll H \ll H|_{{b_{1}}\sim2s}$. It turns out, that the analytic approximation is accurate almost down to ${H_{p}}$, where the magnetization is $M_{p} = M({H_{p}})$, see [Eq. ]{}. For very large fields, $H > H|_{{b_{1}}\sim2s}$, where the distance between the dome boundary ${b_{1}}$ and the sample edge $s$ falls below $s$, the magnetic response is well described by the asymptotic result in [Eq. ]{}. The dome reaches the edges at a distance $d$ only when $H \sim {H_{s}}\gg H|_{{b_{1}}\sim 2s}$. Both approximations and are shown in their domain of applicability.[]{data-label="fig:numeric-vs-analytic"}](plot_07.eps){width=".43\textwidth"} Because of the simplification in [Eq. ]{}, the validity of the result is limited to fields $H \ll H|_{{b_{1}}\sim2s}\sim ({b}^{2}/sW){H_{p}}\approx \sqrt{d/s\,} {H_{s}}$, where the restriction $s \ll {b_{1}}$ is satisfied. As shown in [Fig. ]{}\[fig:numeric-vs-analytic\], the expression is in good agreement with the numerical solution and describes the evolution of the magnetic response over a large range of fields ${H_{p}}\lesssim H \ll H|_{{b_{1}}\sim2s}$. For ${b_{1}}- s \ll s$, the same Ansatz ${b_{2}}= W (1-\nu)^{1/2}$ allows to simplify the constraint [Eq. ]{} to $$\begin{aligned} \label{eq:no-net-current-for-b1minuss-ll-s} {b_{1}}&= s + W\nu/2,\end{aligned}$$ while [Eq. ]{} takes the form $$\begin{aligned} \label{eq:appr-critical-current-condition-2} s + W\nu/2 &= \sqrt{s^{2} + W^{2}\Big(\frac{{H_{p}}}{H}\frac{{b}^{2}}{W^{2}}\Big)^{2}}.\end{aligned}$$ To leading order in ${H_{p}}/H$ we find $$\begin{aligned} \nu(H) &= \frac{W}{s}\Big(\frac{{H_{p}}}{H} \frac{{b}^{2}}{W^{2}}\Big)^{2}\end{aligned}$$ and the magnetization reads $$\begin{aligned} \label{eq:magnetization-in-large-field-assymptotic} M(H) = -\frac{{H_{s}}^{2}}{4H} Wd.\end{aligned}$$ For the strongly penetrated double strip, when the current-carrying regions are smaller than $s$ but still wider than $d$, the mutual influence of the two strips becomes negligible. The magnetization thus approaches that of two independent single strips of width $W$ each \[see [Eq. ]{}\]. Pushing the above ‘thin strip’ solution obtained for $s \gg d$ to the limit $s = d$, we find for the penetration field in [Eq. ]{} $$\begin{aligned} \label{eq:penetration-field-s-is-d-thin-strips} {H_{p}}&\approx {H_{s}}\frac{d}{W} \log{(4W/d)},\end{aligned}$$ which is substantially smaller than that of an isolated strip as given in [Eq. ]{}. Similarly, in this limit the magnetization as approximated by [Eq. ]{} becomes $$\begin{aligned} \label{eq:appr-mag-s-eq-d} M(H) &= -\frac{{H_{s}}}{4\pi} 4Wd \,\big[\log{(4 {H_{s}}/H)}-1\big].\end{aligned}$$ This expression is valid for fields up to $H|_{{b_{1}}\sim 2d} \sim {H_{s}}$ where the dome reaches the edges and is consistent with the limit $s {\scriptstyle{\ \nearrow\ }} d$ approaching the thickness $d$ from below as discussed in Sec. \[sec:finite-thickness\] below. In order to understand the penetration mechanism in the double strip for the full range of strip separations $2s$, below we extend our analysis to a system where the gap width $2s$ is much smaller than the thickness $d$ of the strips, $s \ll d$, see [Sec. ]{}\[sec:finite-thickness\]. Before doing that, we briefly elaborate on the corrections due to the vortex structure in the dome. ![The magnetic response of the double strip is shown for different separation $s$ (solid lines) and a fixed ratio $w/d=10^{3}$. For $s$ larger or equal to $d$, we adopt the thin-strip approach of [Sec. ]{}\[sec:sec:double-strip\], while for $s < d$ (see top expansion), the determination of the magnetization curves has to account for the finite thickness of the strips as described in [Sec. ]{}\[sec:finite-thickness\]. The magnetization $M_{p}$ at the penetration field ${H_{p}}$ is largest for isolated strips ($s/w \to \infty$) and reduces upon decreasing $s$. The parametric curve $({H_{p}}, M_{p})$ as a function of $s$ is indicated by the dotted line. In the limit $s/w \to 0$ the slope of the magnetization curve in the Meissner state doubles as compared to that for isolated strips ($s/w \to \infty$). The dashed line indicates the magnetization of a single strip of width $2W$, corresponding to $s = 0$.[]{data-label="fig:magnetization-s=d-compared-to-single-and-isolated-strip"}](plot_04_mod_test.eps){width=".48\textwidth"} ### Magnetization of the vortex dome {#sec:sec:sec:magnetic-medium-2} To estimate the quantitative effects arising from the currents around the flux lines in the vortex dome, we give an upper bound to the corrections of the magnetic response in [Eqs. ]{} and . Following the analysis presented in [Sec. ]{}\[sec:sec:sec:magnetic-medium-1\], we find an upper bound $$\begin{aligned} \label{eq:magn-correction-estimated-2} \delta M < \frac{H}{4\pi} \frac{H_{c1}}{H_{c1}+H} 2W d\end{aligned}$$ for the magnetization corrections. In the regime $s \ll W$ the relative correction to the magnetic response is bounded by $$\begin{aligned} \label{eq:relative-corrections-double-1} \frac{\delta M}{M} &< \frac{H}{{H_{s}}} \frac{H_{c1}}{H_{c1}+H} \sqrt{\frac{d}{4s}}\frac{1}{\log(4{H_{s}}/H\sqrt{d/s})-1}\end{aligned}$$ in the low field range $H < {H_{s}}\sqrt{d/s}$ \[see [Eq. ]{}\] and by $$\begin{aligned} \label{eq:relative-corrections-double-2} \frac{\delta M}{M} &< \frac{H^{2}}{{H_{s}}^{2}} \frac{H_{c1}}{H_{c1}+H} \frac{2}{\pi}\end{aligned}$$ for higher fields field, $H > {H_{s}}\sqrt{d/s}$ \[see [Eq. ]{}\]. The first expression is always small by the order $d/s$, while the second expression predicts small corrections $\propto (H/{H_{s}})^{2}$ in the field range $H \ll {H_{s}}$, spanning the range of validity for the results presented in this section. We conclude, that the corrections arising due to the vortex state inside the superconducting strips are small, justifying the simplified model for the penetrated state ($\mu = 1$) used in our analysis. Strips with Finite thickness $\boldsymbol{d}$ {#sec:finite-thickness} ============================================= Introduction {#sec:f-t-formalism} ------------ We now explore the double strip geometry for narrow gaps $2s \ll d$. In order to simplify our discussion, the penetration depth $\lambda$ is assumed to be negligible[@footnote:finite-lambda], $\lambda \ll s$. With the gap-width $s$ the smallest geometric length and using $d \ll w$, the results are presented to leading order in $s/d$ and $d/w$, respectively; in particular, the half-width $W$ of the system is approximated by the width $2w$ of one strip. The solutions for infinitely thin strips derived in the previous sections have been regularized near the sample edges with a cut-off $\delta$ of the order of the thickness, $\delta~\sim~d$. This approach is not appropriate anymore when the spacial solution near (inside) the gap is determined by the length scale $s$ rather than $d$. The appropriate boundary conditions then have to be taken into account on the entire rectangular cross-section and the strips cannot be treated as infinitely thin anymore. ![Left panel: Field lines for the estuary problem (solid lines in the ${\tilde{\xi}}$-plane) as calculated numerically from [Eq. ]{}. Right panel: the field lines of a point source in the upper half $\zeta$-plane from which the estuary flow is derived via the inverse Schwarz-Christoffel transformation $\zeta({\tilde{\xi}})$. The field lines of the estuary problem approach that of a point source (dashed lines in the left panel) within a distance $s$ away from the opening.[]{data-label="fig:estuary-flow"}](plot_02_mod.eps){width="45.00000%"} The detailed derivation of the field distribution in the vicinity of the narrow ($2s$) and elongated ($d$) gap (see [Fig. ]{}\[fig:estuary-flow\]) presented in [Sec. ]{}\[sec:sec:sec:near-field\] below will provide us with a uniform field inside the gap of strength $$\begin{aligned} \label{eq:anticipate-field-in-the-gap} B_{g} &= \frac{\phi_{g}}{2s},\end{aligned}$$ where the flux $\phi_{g}$ through the gap has to be determined consistently with the field distribution far away from the gap. For distances $s < |\boldsymbol{r}| \ll w$ away from the upper ($+$) and lower ($-$) gap opening, the field assumes the form of a monopole with radial decay $$\begin{aligned} \label{eq:anticipate-field-away-from-the-opening} \boldsymbol{B}(\boldsymbol{r}) &= \pm \frac{\phi_{g}}{\pi} \frac{\boldsymbol{r}}{|\boldsymbol{r}|^{2}}.\end{aligned}$$ The corresponding result expressed through the holomorphic field reads $$\begin{aligned} \label{eq:ant-field-away-from-the-opening} \mathcal{B}(\xi) &= i \,\frac{\phi_{g}}{\pi\xi}.\end{aligned}$$ In [Sec. ]{}\[sec:sec:sec:far-field\] we find the field distribution far away from the gap, match the far-field solution with the solution in the gap, and thereby find the flux $\phi_{g}$ through the gap. Along with this derivation, we will discuss the consequences on the double strip solution originating from the current and field distribution in and around the gap. Estuary Problem {#sec:sec:sec:near-field} --------------- The field distribution inside the gap and near the opening at $\xi_{{\mathrm{out}}} = 0 + i d/2$ is described by a so-called estuary flow, i.e., the flow into open space of an incompressible fluid leaving a canal of width $2s$ and large (infinite) length $d$, see Fig.\[fig:estuary-flow\]. We define the shifted coordinate system ${\tilde{\xi}}= \xi - \xi_{{\mathrm{out}}}$ centered at the gap opening two-dimensional estuary geometry and determine the holomorphic function $\mathcal{B}({\tilde{\xi}})$. For a diamagnetic superconductor, the field component perpendicular to the surface vanishes everywhere such that $\mathcal{B}({\tilde{\xi}})$ is purely real ($B_{x} = 0$ and $B_{z} \neq 0$) at the surfaces inside the gap ($\operatorname{Re}{[{\tilde{\xi}}]} = \pm s$, $\operatorname{Im}{[{\tilde{\xi}}]} <0$) and imaginary ($B_{x} \neq 0$ and $B_{z} = 0$) on the surfaces along $x$, i.e., for $\operatorname{Im}{[{\tilde{\xi}}]} = 0$ and $|\!\operatorname{Re}{[{\tilde{\xi}}]}| \geq s$. ![Illustration of the stereographic projection of a (conventional) triangle from the Euclidean plane (top left) to the Riemann sphere (top right). Replacing one edge of this triangle by its complement (passing through infinity) generates an unbound triangle. This situation is illustrated both in the Euclidean plane (bottom left) and on the Riemann sphere (bottom right) after a stereographic projection. []{data-label="fig:stereographic-projection-example"}](plot_17_4in1.eps){width="48.00000%"} This boundary value problem can be solved with the help of a Schwarz-Christoffel transformation[@Henrici_74] describing a biholomorphic mapping of the upper complex half-plane $\zeta$, $\operatorname{Im}{[\zeta]} \geq 0$, onto the inner of a polygon. Indeed, the field-allowed region in the estuary geometry is a special case of an unbounded triangle (visualized in Figs.\[fig:stereographic-projection-example\] and \[fig:stereographic-projection\]), with vertices ${\tilde{\xi}}_{v}$ at $-s$, $-i\infty$, and $s$ and internal angles $3\pi/2$, $0$, and $3\pi/2$. The corresponding Schwarz-Christoffel transformation takes the form $$\begin{aligned} \label{eq:sc-transformation} {\tilde{\xi}}(\zeta) &= s + \frac{2}{\pi} \bigg[\sqrt{\zeta^{2}-s^{2}} - 2s \arctan{\sqrt{\frac{\zeta - s}{\zeta + s}}}\bigg]\end{aligned}$$ and maps the upper half-plane $\zeta$ ([Fig. ]{}\[fig:estuary-flow\], right) to the estuary plane ${\tilde{\xi}}$ ([Fig. ]{}\[fig:estuary-flow\], left). The flux $\phi_{g}$ emanating from the vertex at ${\tilde{\xi}}_{v} = -i \infty$ in the estuary is conserved in the transformation [Eq. ]{} and maps to a point source of strength $\phi_{g}$ at $\zeta = 0$, with field lines dispersing into the upper half plane $\operatorname{Im}[\zeta] \geq 0$. The complex potential[@Thomson_60] $$\begin{aligned} \label{eq:complex-potenital} \bar\Omega(\zeta) = \frac{i\phi_g}{\pi} \log\zeta\end{aligned}$$ is generating the field $\mathcal{\bar B}(\zeta)= d\bar\Omega/d\zeta = i\phi_{g}/\pi \zeta$ of this point source in the upper half-plane. Transforming back to the estuary geometry, the potential $\Omega({\tilde{\xi}}) = \bar \Omega[\zeta({\tilde{\xi}})]$ generates the field $$\begin{aligned} \label{eq:estuary-solution} \mathcal{B}({\tilde{\xi}})= \frac{d\Omega}{d{\tilde{\xi}}} = \frac{i\phi_g}{2} \frac{1}{\sqrt{\zeta({\tilde{\xi}})\big.^2-s^2}}.\end{aligned}$$ The last equality was obtained by using the Schwarz-Christoffel transformation . Alternatively, the analysis on the level of fields involves the solution $\mathcal{\bar B}(\zeta) = i \phi_{g}/\pi \zeta$ for a point source and the transformation back involves an additional derivative, $\mathcal{B}({\tilde{\xi}}) = (d\zeta/d{\tilde{\xi}}) \mathcal{\bar B}[\zeta({\tilde{\xi}})]$. In [Fig. ]{}\[fig:estuary-flow\], we show the resulting field lines of [Eq. ]{} as obtained from inverting [Eq. ]{} numerically. In our further discussion it is sufficient to determine the field distribution in the asymptotic regimes where analytic results are available. Deep inside the gap ($-\operatorname{Im}{[{\tilde{\xi}}]} \gg s$) the inverse of [Eq. ]{} takes the form $$\begin{aligned} \label{eq:gap-inversion} \zeta({\tilde{\xi}}) &= 2s \, e^{- [i \pi ({\tilde{\xi}}- 1)/ 2s] -1}\end{aligned}$$ and using [Eq. ]{}, we find \[up to corrections $\propto \exp({\pi \tilde{z}/2s})$\] a uniform field directed along $z$ of strength $$\begin{aligned} \label{eq:field-in-the-gap} B_{g} &= \frac{\phi_{g}}{2s}.\end{aligned}$$ Near the corner of the estuary, $|{\tilde{\xi}}- s| \ll s$, the transformation [Eq. ]{} reads $$\begin{aligned} \label{eq:sc-near-corner-s} \frac{{\tilde{\xi}}-s}{2s} &\sim \frac{2}{3\pi} \left(\frac{\zeta-s}{2s}\right)^{3/2}\end{aligned}$$ and a similar expression is found near ${\tilde{\xi}}= -s$. For both corners, the holomorphic field shows a power law singularity $\propto |{\tilde{\xi}}\pm s|^{-1/3}$, which will be regularized in a real sample by the partial penetration (at a depth $\sim s$) of vortices into the sample corners. ![Visualization of the field-allowed region (light gray) of the estuary geometry on the Riemann sphere (filled region) via a stereographic projection. The triangular shape of the boundary of the estuary, with vertices ${\tilde{\xi}}_{v}$ at $\pm s$ and $-i\infty$ is clearly visible on the Riemann sphere representation, see also [Fig. ]{}\[fig:stereographic-projection-example\]. Here, the north pole corresponds to the origin of the complex plane ${\tilde{\xi}}$, while the complex infinity is projected onto the south pole. A line in the original plane ${\tilde{\xi}}$ is mapped to a circle (passing through the south pole) on the Riemann sphere.[]{data-label="fig:stereographic-projection"}](plot_10.eps){width="48.00000%"} Far away from the opening, $|{\tilde{\xi}}| \gg s$ and $\operatorname{Im}{[{\tilde{\xi}}]} \geq 0$, the inverse transform becomes $\zeta({\tilde{\xi}}) = \pi {\tilde{\xi}}/ 2$ and the holomorphic function assumes the limiting form $$\begin{aligned} \label{eq:field-away-from-the-opening} \mathcal{B}({\tilde{\xi}}) &= i \,\frac{\phi_{g}}{\pi{\tilde{\xi}}}\end{aligned}$$ describing a point source of strength $\phi_{g}$ located at ${\tilde{\xi}}= 0$. Narrow gap double strip {#sec:sec:ngds} ----------------------- \[sec:sec:sec:far-field\] Away from the gap and from the outer strip edges, a thin-strip description similar to the one discussed in [Sec. ]{}\[sec:thin-strips\] is applicable, with the holomorphic field taking the form $$\begin{aligned} \label{eq:double-field-general-narrow-gap} \mathcal{B}(\xi) &= H \sqrt{\frac{(\xi^{2}-{b_{1}}^{2})(\xi^{2}-{b_{2}}^{2})} {\xi^{2}(\xi^{2}-W^{2})}}.\end{aligned}$$ The factor $\xi^{2}$ in the denominator \[replacing $(\xi^{2}-s^{2})$ in [Eq. ]{}\] captures the flux emanating from the point-like source as derived in [Eq. ]{}. From the above expression, the field distribution along the $x$-axis is given by $$\begin{aligned} \label{eq:double-field-general-narrow-gap_xz} \frac{B_{z}(x)}{H} &= \left\{ \begin{aligned} &\sqrt{\frac{(x^{2}-{b_{1}}^{2})({b_{2}}^{2}-x^{2})} {x^{2}(W^{2}-x^{2})}}, && \mathrm{for\ } {b_{1}}\leq x \leq {b_{2}},\\ &\sqrt{\frac{(x^{2}-{b_{1}}^{2})(x^{2}-{b_{2}}^{2})} {x^{2}(x^{2}-W^{2})}}, && \mathrm{for\ } W \leq x. \end{aligned}\right.\end{aligned}$$ Comparing [Eqs. ]{} and in the regime $|\xi| \ll {b_{1}}$, we find the flux $$\begin{aligned} \label{eq:flux-general-narrow-gap} \phi_{g}= H W \pi {b_{1}}{b_{2}}/W^{2}\end{aligned}$$ and the uniform field strength inside the gap $|x| < s$ takes the form $$\begin{aligned} \label{eq:field-general-in-the-gap} B_{g} = H \frac{\pi W}{2 s} \frac{{b_{1}}{b_{2}}}{W^{2}}.\end{aligned}$$ Note, that for $s \ll d \ll w$, the difference between the shifted coordinate ${\tilde{\xi}}$ and $\xi$ is beyond our resolution, such that ${\tilde{\xi}}= \xi$. The current contribution from the region away from the gap is obtained from the holomorphic field in [Eq. ]{} via Ampères law and reads $$\begin{aligned} \label{eq:double-current-general-narrow-gap} I(x) &= \left\{\begin{aligned} &\frac{H c}{2\pi} \sqrt{\frac{({b_{1}}^{2}-x^{2})({b_{2}}^{2}-x^{2})} {x^{2}(W^{2}-x^{2})}}, && s \leq |x| \leq {b_{1}},\\ &\frac{-H c}{2\pi} \sqrt{\frac{(x^{2}-{b_{1}}^{2})(x^{2}-{b_{2}}^{2})} {x^{2}(W^{2}-x^{2})}}, && {b_{2}}\leq |x| \leq W,\\ &0, && \mathrm{otherwise}. \end{aligned}\right.\end{aligned}$$ The $1/x$ dependence of the current is applicable only for $|x| \gg s$. However, it turns out that the deviation of $I(x)$ (as obtained from solving [Eq. ]{} numerically) from $1/x$ is not relevant for the further analysis, and the expression given above for the sheet current density $I(x)$ can be used down to $|x| = s$. The homogeneous field inside the gap is generated by a screening current density $$\begin{aligned} \label{eq:current-density-in-the-gap} j_{g}(x,|z|<d/2) &= \frac{x}{|x|} \frac{B_{g}\, c}{4\pi} \,\delta(|x|-s)\end{aligned}$$ flowing along $y$ at the gap surfaces ($x=\pm s$, $|z| \leq d/2$). Here $\delta$ is the Dirac delta function, which accounts for the assumption $\lambda \to 0$. The two current channels at $x = \pm s$ provide a significant contribution to the total current in the strips. Note that these channels exist for $s \gg d$ as well; for large gaps their contribution to the total current is negligible, though. To treat these currents on equal footing with the sheet current flowing in the strips ($s \leq x \leq W$) we define the sheet current density for the gap currents $$\begin{aligned} \label{eq:current-in-the-gap} I_{g}(x) &= \frac{x}{|x|} \frac{B_{g} c}{4\pi} \,d\,\delta(|x|-s)\end{aligned}$$ by integrating [Eq. ]{} over the strip thickness $d$. The currents flowing along the vertical surfaces at the outer edges ($|x| = W$) are parametrically smaller as compared to the contributions near the gap ($|x| = s$) and are neglected here. The two dominant current contributions then add up to the total current, $I_{\mathrm{tot}}(x) = I(x) + I_{g}(x)$. This current distribution, when compared to the thin strip case, corresponds to a rearrangement of the current densities towards the inner edges of the strips, see [Fig. ]{}\[fig:double-strip-meissner-s-ll-d\]. ![Dimensionless current $2\pi I(x)/H c$ (thick solid line) and reduced magnetic field $B_{z}(x)/H$ (dashed line) as a function of $x$ and for $z=0$ in a double strip in the Meissner state. The profiles are calculated for the case $s \ll d \ll 2w$ with the parameters $w/d = d/s = 100$. The current profile inside the strips changes sign at $\pm {b}$, with $|{b}| \approx 0.21 w$. The additional current $I_{g}(x)$ from [Eq. ]{} flowing near the inner edges of the strips changes the condition of zero net current dramatically. The thin lines show the current and field profiles of the double strip in the thin strip limit, $d/s \to 0$, for fixed $s = 10^{-4}w$.[]{data-label="fig:double-strip-meissner-s-ll-d"}](plot_03_mod.eps){width="45.00000%"} The diamagnetic contribution from the current $I(x)$ in the strip, $$\begin{aligned} \label{eq:double-magnetization-general-narrow-gap} M(H) &= -\frac{H}{4}(W^{2}-{b_{1}}^{2}-{b_{2}}^{2}),\end{aligned}$$ as obtained from evaluating [Eq. ]{} with the field , is parametrically larger ($\propto W/d$) than the paramagnetic contribution $$\begin{aligned} \label{eq:magnetization-in-the-gap} M_{g}(H) &= \frac{H}{4} W^{2} \frac{{b_{1}}{b_{2}}d}{W^{3}}\end{aligned}$$ from the current $I_{g}(x)$ along the gap surface and we neglect the latter in the following. The problem is then, once again, reduced to finding the parameters ${b_{1}}$ and ${b_{2}}$ within the Meissner- and penetrated states. ### Meissner state {#sec:sec:sec:Meissner-3} In the Meissner state were ${b_{1}}= {b_{2}}= {b}$, the field $B_{z}$ \[[Eq. ]{}\] along the $x$-axis simplifies to $$\begin{aligned} \label{eq:double-field-meissner-narrow-gap} B_{z}(x) &= H \frac{x^{2}-{b}^{2}} {|x| \sqrt{x^{2}-W^{2}}}.\end{aligned}$$ for $|x|>W$ and is constant \[[Eq. ]{}\], $$\begin{aligned} \label{eq:field-meissner-in-the-gap} B_{g} = H \pi \frac{{b}^{2}}{2 s W},\end{aligned}$$ inside the gap ($|x|<s$). The total sheet current density reads $$\begin{aligned} \label{eq:double-current-meissner-narrow-gap} I_{\mathrm{tot}}(x) &= - \frac{H c}{2\pi}\bigg[ \frac{x^{2}-{b}^{2}}{x \sqrt{W^{2}-x^{2}}} - \frac{x}{|x|} \frac{\pi {b}^{2}}{4sW} \,d\,\delta(|x|-s) \bigg]\end{aligned}$$ and the general expression for the magnetization takes the form $$\begin{aligned} \label{eq:double-magnetization-meissner-narrow-gap} M(H) &= -\frac{H}{4}(W^{2}-2{b}^{2}).\end{aligned}$$ The value of the parameter ${b}$ is fixed by the constraint of vanishing net current given as $$\begin{aligned} \label{eq:no-net-current-explicit} \int\limits_{s}^{W} \frac{dx \, x}{\sqrt{W^{2}-x^{2}}} = {b}^{2} \bigg[ \frac{\pi d}{4sW} + \int\limits_{s}^{W} \frac{dx}{x \sqrt{W^{2}-x^{2}}}\bigg].\end{aligned}$$ The above integrals simplify in the limit $s \ll W$ and the parameter ${b}$ takes the asymptotic form $$\begin{aligned} \label{eq:a-over-W-ratio-2} {b}^{2} &= \frac{W^{2}}{\pi d/4 s + \log\big(2W/s\big)}.\end{aligned}$$ In contrast to the result for thin strips \[see [Eq. ]{}\], where ${b}^{2}$ changes logarithmically with $s$, in the present case the dependence on $s$ is dominated by the linear term $d/s$ in the denominator. As a result, the parameter ${b}$ is substantially reduced when $s \ll d$, see [Fig. ]{}\[fig:a-of-s\], which is due to the additional currents $I_g$ flowing at the (vertical) gap surface and producing a substantial rearrangement of the overall current density as shown in [Fig. ]{}\[fig:double-strip-meissner-s-ll-d\]. We note that the numerical factor $\pi/4$ of the term $d/s$ in the above expression is precisely known since it derives from the current $I_{g}(x)$ originating from screening the uniform field inside the gap, [Eq. ]{}. The prefactor under the logarithm, however, will be modified if the field distribution at the opening of the estuary is accurately taken into account. Indeed, approaching the corner $(s,d/2)$ from both surfaces $(x,d/2)$ and $(s,z)$ the field deviates from the assumed behavior $B_{x}(x)\propto 1/x$ and $B_{z}(z) = \mathrm{const}$ \[following from [Eqs. ]{} and respectively\]. The precise field distribution (and its related current profile) can be derived by solving [Eq. ]{} numerically and inserting the result into [Eq. ]{}. Neglecting partial penetration of the edge corners, [Eq. ]{} will be modified to $$\begin{aligned} {b}^{2} &= \frac{W^{2}}{\pi d/4 s + \log\big(2.38 W/s\big)}.\end{aligned}$$ Since the precision of this expression also suffers from corrections (e.g. from partial penetration of the edge corners), we will use the relation in the following. The diamagnetic response in the Meissner phase follows from and reduces to $$\begin{aligned} \label{eq:mag-meissner-double-strip-s-ll-d} M(H) &\approx -\frac{H}{4}W^{2} \bigg[1-\frac{8s/\pi d}{1 + (4s/\pi d)\log\big(2W/s\big)}\bigg].\end{aligned}$$ This result approaches that of a single strip of double width \[[*cf. *]{}[Eq. ]{} with $w \to W$\] upon reducing $s$ far below $d$. ![The parameter ${b}$ characterizing the Meissner phase of the double strip is plotted against the half-width $s$ of the gap between the strips. All lengths are normalized to the half width $w$ of the strips. The fixed strip thickness $d = 10^{-3}w$ separates two regimes; in the thin strip regime $d \ll s$, ${b}$ depends on $s$ via [Eq. ]{}. For $s \ll d$, ${b}(s)$ is given through [Eq. ]{}. In between, i.e., for $s \sim d$, a smooth cross-over (dashed line) connects the two limits. In both far asymptotic limits $s \lll d$ and $s \ggg d$, the position ${b}(s)$ follows a simple behavior ${b}(s) = 2w\sqrt{4s/\pi d}$ and ${b}(s) = w + s$, respectively (thin lines).[]{data-label="fig:a-of-s"}](plot_08.eps){width="45.00000%"} Using [Eq. ]{} in the expression for the flux through the gap, we find that $$\begin{aligned} \label{eq:flux-meissner-finite-thickness} \phi_{g} &= H 2W \frac{2s/d}{1 + (4s/\pi d)\log\big(2W/s\big)}\end{aligned}$$ shrinks (up to logarithmic corrections) linearly with decreasing $s$ (note that ${b_{1}}{b_{2}}= {b}^{2}$). Compared to the blocked flux $\phi_{b} \approx H\, 2W$, only a small fraction $\sim 2s/d$ passes through the narrow gap of width $2s$ and length $d$. Consequently, the field strength inside the gap, $$\begin{aligned} \label{eq:double-field-meissner-in-the-gap} B_{g} = H \frac{2W}{d} \frac{1}{1 + (4s/\pi d)\log\big(2W/s\big)},\end{aligned}$$ does not diverge for $s/d \to 0$ but saturates at $H\, 2W/d$. As the field profile inside the gap from where vortices start penetrating the sample is precisely known, the present penetration process is more accurately described than the one for the thin-strip limit where the strips are separated by a distance larger than $d$. Corrections originating from the precise field distribution near the opening of the estuary affect the results only to the next-to-leading order. As before, the penetration starts when the field inside the gap reaches the strength ${H_{s}}$, i.e., at the penetration field $$\begin{aligned} \label{eq:double-penetration-field-narrow-gap} {H_{p}}&= {H_{s}}\frac{2s}{\pi W} \frac{W^{2}}{{b}^{2}} = {H_{s}}\frac{d}{2W} \Big[1 + \frac{4s}{\pi d} \log \Big(\frac{2W}{s}\Big)\Big].\end{aligned}$$ In the limit $s/d \to 0$, the penetration field asymptotically reaches the value ${H_{s}}d/2W$, that is the penetration field of the elliptic strip, [*cf. *]{}[Eq. ]{}. The retardation of field penetration originating from the geometrical barrier has completely disappeared in this limit. At penetration, $H = {H_{p}}$, the diamagnetic response $$\begin{aligned} \label{eq:magnetization-at-penetration-point} M_{p} &= -\frac{{H_{s}}}{16}(2Wd) \Big\{1 + \frac{2s}{\pi d} \Big[2\log \Big(\frac{2W}{s}\Big)-1\Big]\Big\}\\ &= -\frac{{H_{p}}}{4}W^{2} \Big[1-\frac{8s/\pi d}{1 + (4s/\pi d)\log\big(2W/s\big)}\Big],\end{aligned}$$ has collapsed by a factor $\sim (d/W)^{1/2}$ as compared to that of a single strip, [Eq. ]{}, for which the geometrical barrier is fully active. This so-called ‘suppression of the geometrical barrier’, the collapse of ${H_{p}}$ and of $M(H)$, is a central result of this work. Although in the limiting case $s/d \to 0$, the Meissner response and the field of first penetration coincide with that of an elliptically shaped strip, beyond ${H_{p}}$, the magnetic signatures of the double strip still differs substantially from those of the elliptic sample, see Figs.\[fig:numeric-vs-analytic-f-t-l\] and \[fig:descending-double-strip\] as well as the discussion below. Upon decreasing the gap width $s$, the penetration field ${H_{p}}$ is reduced, what leads to a stronger suppression of the geometrical barrier as follows from [Eq. ]{}. The calculation of the equilibrium field defined through [Eq. ]{} provides the result $$\begin{aligned} \label{eq:double-eq-field-narrow-gap} {H_{\mathrm{eq}}}= H_{c1} \frac{d}{2W} \Big[1 - \frac{\log(W^{2}/b^{2})-1} {\pi d / 2s + \log(4W^{2}/s^{2})}\Big]^{-1},\end{aligned}$$ approaching $H_{c1} d/2W$ and the corresponding geometrical barrier height vanishes as $s \log(s)$, $$\begin{aligned} U_{b}^{\mathrm{eq}} &= \varepsilon_{l} d \Big(1 - \Big\{1 + \frac{2s}{\pi d} \big[\log(4{b}^{2}/s^{2})-1\big]\Big\}^{-1}\Big)\\ &\approx \varepsilon_{l} d\ \frac{2s}{\pi d} \big[\log(16W^{2}/\pi s d)-1\big],\end{aligned}$$ where we have assumed that $s \log(W/s) \ll d$ for the last equality. ### Penetrated state {#sec:sec:sec:Shubnikov-2-ft} The field and current distributions [Eqs. ]{}, and [Eqs. ]{}, describe the penetrated state once the parameters ${b_{1}}$ and ${b_{2}}$ have been found; the latter have to respect the limits ${b_{1}}-s \gg s$ and $W-{b_{2}}\gg d$ and are determined by the usual conditions governing the evolution of the vortex dome, the vanishing of the total currents in the strips, $$\begin{aligned} \label{eq:no-net-current-2} \int\limits_{s_{-}}^{{b_{1}}} dx\, I_{\mathrm{tot}}(x) + \int\limits_{{b_{2}}}^{W} dx\, I_{\mathrm{tot}}(x) &= 0\end{aligned}$$ and the condition of criticality at the edge regulating the vortex entrance, here $B_{g} = {H_{s}}$. The latter condition is equivalent to the requirement that the flux $\phi_{g}$ in saturates at ${H_{s}}2s$, or $$\begin{aligned} \label{eq:critical-field-condition} \frac{{b_{1}}{b_{2}}}{{b}^{2}} &= \frac{{H_{p}}}{H},\end{aligned}$$ as expressed through the penetration field ${H_{p}}$ and the zero-current position $b$ of the Meissner state. As before, the perturbative calculation around the penetration field ${H_{p}}$ is very limited due to the rapid growth of the dome width ${b_{2}}-{b_{1}}$ beyond ${H_{p}}$ and we concentrate on the high-field expansion where the vortex domes occupy a large fraction of the strips ${b_{1}}\ll {b_{2}}$, providing results over a large field-range. The two conditions regulating the dome evolution then can be simplified and an analytic solution can be given. With the Ansatz ${b_{2}}= W (1-\nu)^{1/2}$ with $\nu < 1$, the inner dome edge $$\begin{aligned} \label{eq:bone} {b_{1}}&= \frac{{b}^{2}}{W} \frac{{H_{p}}}{H} \frac{1}{\sqrt{1-\nu}}\end{aligned}$$ is expressed through $\nu$ with the help of [Eq. ]{}. Assuming $s \ll {b_{1}}$ and ${b_{1}}\ll {b_{2}}$, the requirement of vanishing net current in [Eq. ]{} simplifies to $$\begin{aligned} \label{eq:appr-neutral-current-condition-f-t-l} \frac{{H_{p}}}{H}\frac{{b}^{2}}{W^{2}} \left[\frac{W^{2}}{{b}^{2}} + \log{\Big(\frac{{b}^{2}}{W^{2}} \frac{{H_{p}}}{H} \frac{1}{\sqrt{1-\nu}} \Big)}-1\right] &\\[.5em] = \operatorname{E}(\sqrt{\nu})-(1&-\nu)\operatorname{K}(\sqrt{\nu}).\nonumber\end{aligned}$$ For large fields $H \gg {H_{p}}$, where $\nu$ is small, the above equation can be expanded in $\nu$. Solving for $\nu(H)$ to second order in ${H_{p}}/H$, we obtain $$\begin{aligned} \label{eq:mu-second-order-f-t-l-0} \nu(H) &\approx \frac{4}{\pi}\frac{{H_{p}}}{H} \bigg\{ 1 + \frac{{b}^2}{W^2} \bigg[\log{\left(\frac{{b}^2}{W^2}\frac{{H_{p}}}{H}\right)}-1\bigg] \bigg\}\\ &\ - \frac{2}{\pi^2}\frac{{H_{p}}^2}{H^2} \bigg\{1+\frac{{b}^2}{W^2} \bigg[\log{\bigg(\frac{{b}^2}{W^2}\frac{{H_{p}}}{H}\bigg)}-1\bigg] \bigg\}^2\nonumber\\ &\ + \frac{8}{\pi^2}\frac{{H_{p}}^2}{H^2} \frac{{b}^2}{W^2} \bigg\{1+\frac{{b}^2}{W^2} \bigg[\log{\bigg(\frac{{b}^2}{W^2}\frac{{H_{p}}}{H}\bigg)}-1\bigg] \bigg\}.\nonumber $$ The magnetic response given in [Eq. ]{} simplifies to $M(H) = - H \nu(H)W^{2}/4$ and [Fig. ]{}\[fig:numeric-vs-analytic-f-t-l\] shows the result of combining this expression with $\nu(H)$ from [Eq. ]{}. Although the range of applicability ${H_{p}}\ll H$ of the above expression does not a-priori cover the regime near penetration, the results are still in good agreement with the numerical solution down to $H \approx {H_{p}}$. ![The diamagnetic response $M(H)$ (solid line) for the double strip in the limit $s \ll d \ll w$ (here $w/d = d/s = 100$) as obtained from solving [Eqs. ]{} and numerically. In addition the dotted (dashed) line shows the analytic result for the magnetization $M(H) = -H \nu(H) W^{2}/4$, where $\nu(H)$ is obtained from solving [Eq. ]{} to linear (quadratic) order in ${H_{p}}/H$, see [Eq. ]{}. It is necessary to express the solution to second order in ${H_{p}}/H$ as the first order solution gives only poor results close to ${H_{p}}$. The magnetization of a single elliptic strip of width $2W$ and thickness $d$ (long thin dashes) is reversible and with linear slope beyond ${H_{p}}$ (as also shown in [Fig. ]{}\[fig:elliptic-magnetization\]).[]{data-label="fig:numeric-vs-analytic-f-t-l"}](plot_11_merged.eps){width="45.00000%"} Neglecting irrelevant terms of order $({b}/W)^{2}({H_{p}}/H)^{2}$ and higher in [Eq. ]{}, the magnetization reads $$\begin{aligned} \label{eq:mag-second-order-f-t-l} M(H) &\approx \bar{M}\bigg\{ 1 + \frac{{b}^2}{W^2} \bigg[ \log{\bigg(\frac{{b}^2}{W^2}\frac{{H_{p}}}{H} \bigg)}-1 \bigg] - \frac{1}{2\pi}\frac{{H_{p}}}{H} \bigg\},\end{aligned}$$ with $\bar{M} = -{H_{p}}W^{2}/\pi$. In contrast to the thin strip case \[see [Eq. ]{}\], where the magnetization depends logarithmically $\propto \log{({H_{s}}/H)}$ on the applied field, the magnetic response in the present limit is dominated by a field-independent contribution, $$\begin{aligned} \label{eq:appr-mag-f-t-l} \bar{M} &\approx -\frac{{H_{p}}}{\pi} W^{2} \approx -\frac{{H_{s}}}{4\pi} (2Wd),\end{aligned}$$ producing an almost flat magnetization. This flatness is the result of the particular current distribution inside the strips: The current flowing close to the inner edge is dominated by the contribution $I_{g}(x)$ from the gap, i.e., $$\begin{aligned} \label{eq:current-estimation-near-the-gap} \int\limits_{s_{-}}^{{b_{1}}} dx\, I_{\mathrm{tot}}(x) &\approx \int\limits_{s_{-}}^{s_{+}} dx\, I_{g}(x) = \frac{{H_{s}}c}{4\pi} \,d.\end{aligned}$$ To satisfy the condition of vansihing net current, the current density $I(x)$ between ${b_{2}}$ and $W$ has to compensate the gap contribution, leading to $$\begin{aligned} \label{eq:current-estimation-outer-edge} \int\limits_{{b_{2}}}^{W} dx\, I_{\mathrm{tot}}(x) &\approx -\frac{{H_{s}}c}{4\pi} \,d.\end{aligned}$$ Once the dome occupies a large fraction of the sample, these currents flow at the outer edge, i.e., a distance $\sim W$ away from the origin and produce the dominant (field-independent) contribution $-{H_{s}}(2Wd)/4\pi$ \[[*cf. *]{}[Eq. ]{}\] to the magnetization at large fields (the factor 2 originates from the integration over both strips). Note that in the limit $s/d \to 0$, the leveling out of the magnetization at the value given in [Eq. ]{} is by a factor $4/\pi$ larger than its value at penetration ${H_{p}}$ \[see [Eq. ]{}\]. Although almost constant, the magnetization assumes a maximal diamagnetic signal $$\begin{aligned} \label{eq:max-magnetization} M(H_{m}) &= \bar{M} \bigg\{1 + \frac{4s}{\pi d} \bigg[\log\bigg(\frac{32s^{2}}{\pi d^{2}}\bigg) -2\bigg]\bigg\}\end{aligned}$$ at the applied field $$\begin{aligned} \label{eq:H-max} H_{m} &\approx {H_{p}}\frac{d}{8s} \approx \frac{{H_{s}}}{16}\frac{d^{2}}{s W}.\end{aligned}$$ For $s/d \lesssim d/16W$, the diamagnetic response monotonically increases up to $H \sim {H_{s}}$. On the other hand, we may extrapolate the expression to $s \lesssim d$ and predict a value of the gap parameter $s \sim d/8$ where $H_{m}$ merges with the penetration field ${H_{p}}$ upon increasing $s$. The “flatness” of the magnetization curve in the penetrated state is quantified by relating the slope $M'(H)$ \[as obtained from [Eq. ]{}\] to the Meissner slope $- W^{2}/4$, yielding $$\begin{aligned} \label{eq:relative-magnetization-slope} -\frac{4 M'(H)}{W^{2}} = \frac{2}{\pi^{2}} \bigg(\frac{{H_{p}}^{2}}{H^{2}} - \frac{8s}{d}\frac{{H_{p}}}{H}\bigg) \ll 1.\end{aligned}$$ As for thin strips and wide gaps (see [Sec. ]{}\[sec:sec:sec:Shubnikov-2-tsl\]), we can push the results obtained in the limit $s \ll d$ to the extreme case $s \to d$. The penetration field $$\begin{aligned} \label{eq:penetration-field-s-is-d-thick-strips} {H_{p}}&\approx {H_{s}}\frac{d}{2W} \Big[1 + \frac{4}{\pi} \log \Big(\frac{2W}{d}\Big)\Big]\end{aligned}$$ as obtained from [Eq. ]{} with $s = d$, agrees up to numbers of order unity with the result obtained from the opposite limit $s \gg d$, see [Eq. ]{}. Taking the limit $s \to d$ from the regime of small gaps $s \ll d$, the magnetization is dominated by the first term in [Eq. ]{} and simplifies to $$\begin{aligned} \label{eq:magnetization-s-is-d-thick-strips} M(H) &\approx -\frac{{H_{s}}}{4\pi} (4Wd) \Big[\frac{2}{\pi} \log{\Big(\frac{4{H_{s}}}{\pi H}\Big)} + \frac{4 - \pi}{2\pi} \Big].\end{aligned}$$ This expression agrees well with the corresponding expression obtained in the limit $s {\scriptstyle{\ \searrow\ }}d$ approaching the thickness $d$ from above. Although the penetration field asymptotically ($s/d \to 0$) approaches the equilibrium field , a finite irreversibility persists and the geometric barrier rapidly reappears upon reducing the applied field. Upon decreasing the magnetic field from a maximal value $H^{\star}$, the vortex dome expands while keeping the trapped flux constant, $$\begin{aligned} \label{eq:double-flux-through-dome} \phi_{d}(H) &\equiv \int\limits_{{b_{1}}}^{{b_{2}}} dx\, H \sqrt{\frac{(x^{2}-{b_{1}}^{2})({b_{2}}^{2}-x^{2})} {x^{2}(W^{2}-x^{2})}} = \phi_{d}^{\star},\end{aligned}$$ where $\phi_{d}^{\star} = \phi_{d}(H^{\star})$. This constraint for the decreasing field replaces the constraint $B_{g}={H_{s}}$ for the increasing field. Excluding a narrow field range $H^{\star} \simeq {H_{p}}$, the dome extends over a large fraction of the sample and the constraint of conserved trapped flux can be simplified under the assumptions ${b_{1}}\ll {b_{2}}$ and $W-{b_{2}}\ll W$ to read $$\begin{aligned} H W \Big\{1 - \frac{\nu}{4} \Big[\log\Big(\frac{16}{\nu}\Big) + 1\Big] \Big\} &= \phi_{d}^{\star},\end{aligned}$$ where we used $\nu(H) = 1-[{b_{2}}(H)/W]^{2} \ll 1$ as before. Similar to the single strip calculations \[see [Eq. ]{}\], the slope of the magnetic response $M(H) = - H \nu(H)W^{2}/4$ is given by $$\begin{aligned} \label{eq:slope-2} \frac{dM}{dH} &= -\frac{W^{2}}{4} \frac{4 - \nu}{\log(16/\nu)},\end{aligned}$$ which is numerically close to that of the Meissner phase ($-W^{2}/4$). At the onset of the descending branch, i.e., $H^{\star}-H \ll H^{\star}$, we find an analytic expression for the magnetization of the form $$\begin{aligned} \label{eq:analytic-descending-branch-2} M(H) = M(H^{\star}) - \frac{H-H^{\star}}{4} W^{2} \frac{4 - \nu^{\star}}{\log(16/\nu^{\star})},\end{aligned}$$ where $\nu^{\star} = \nu(H^{\star})$ is obtained from [Eq. ]{}. The above expression and the result of an exact numerical calculation of the magnetization are shown in [Fig. ]{}\[fig:descending-double-strip\]. ![Numerical solution for the magnetization of the descending branch for narrow gaps $s \ll d$. The two conditions of vanishing net current and conserved trapped flux is solved for $H^{\star} = n {H_{p}}$ (with integer $2\leq n \leq 5$). Parameters are $w/d = d/s = 100$. The analytic result (dashed lines) as obtained from an expansion close to $H = H^{\star}$ gives a reasonable description of the numerical solution over a wide field range.[]{data-label="fig:descending-double-strip"}](plot_18.eps){width=".45\textwidth"} ### Magnetization of the vortex dome {#sec:sec:sec:magnetic-medium-3} In the limit $s \ll d$, the diamagnetic response of the double strip is flat and small by the factor $\sim \sqrt{d/W}$ as compared to the single strip at $H_p$; hence, we should verify that the magnetic response of the vortex state in the flux-filled region does not substantially alter the above results. Following again the analysis discussed earlier in [Sec. ]{}\[sec:sec:sec:magnetic-medium-1\], the corrections to the magnetic response are bounded from above by the function $$\begin{aligned} \label{eq:magn-correction-estimated-3} \delta M < \frac{H}{4\pi} \frac{H_{c1}}{H_{c1}+H} 2W d\end{aligned}$$ leading to relative corrections $$\begin{aligned} \label{eq:relative-corrections-double-3} \frac{\delta M}{M} &< \frac{H}{{H_{s}}} \frac{H_{c1}}{H_{c1}+H} \frac{4}{\pi}\end{aligned}$$ that are small as long as $H \ll {H_{s}}$. Without surface barrier, ${H_{s}}= H_{c1}$, the corrections become of order unity only when $H \sim H_{c1}$. On the other hand, for a large surface barrier ${H_{s}}\gg H_{c1}$, the corrections remain small when $H \sim {H_{s}}$. We conclude, that the contribution of the equilibrium magnetization of the Shubnikov state to the overall magnetization of the double strip geometry (with $s \ll d \ll w$) is small and can, in most cases, be neglected in the entire field range $H < H_{c1}$. Several strips {#sec:several-strips} ============== So far, we have given a detailed description of the single and double strip geometries. A discussion of three coplanar strips will reveal additional features as compared to the previous systems, and allows for a qualitative understanding of the response of a system of a *finite* number $n \geq 3$ of coplanar strips in a parallel arrangement. The general holomorphic field for $n$ ($n \geq 1$) parallel strips arranged symmetrically around the origin $\xi=0$ assumes the form $$\begin{aligned} \label{eq:ant-n-strip-function} \mathcal{B}(\xi) &= H \sqrt{\prod\limits_{i}\frac{\xi^{2}-b_{i}^{2}(H)}{\xi^{2}-e_{i}^{2}}},\end{aligned}$$ where $\pm e_{i}$ denote the strip edges and the parameters $b_{i}(H)$ define the boundaries of the vortex states. For an even number $n = 2m$ of strips, the $k$th strip ($0 < k \leq m$) as counted along the positive $x$-axis ranges from $e_{2k-1}$ to $e_{2k}$ and vortices fill the region $b_{2k-1}$ to $b_{2k}$. Every strip has a symmetric counterpart on the negative $x$-axis. For an odd number $n=2m+1$ of strips, the above remains unchanged except for an additional innermost strip ranging from $-e_{0}$ to $e_{0}$ with a dome between $-{b_{0}}$ and ${b_{0}}$. The product in [Eq. ]{} runs from 1 to $n$ (from 0 to $n-1$) for the even (odd) numbered configurations. The expressions and are special cases for the single and double strip geometries. The magnetization of the $n$-strip system as obtained from [Eq. ]{} reads $$\begin{aligned} \label{eq:n-strip-magnetization} M(H) &= -\frac{H}{4} \sum\limits_{i}(e_{i}^{2}-b_{i}^{2}).\end{aligned}$$ In this section, we consider strips of equal width $2w$ and separated by a gap $2s$. We also limit the analysis to the thin strip case, i.e., the thickness $d$ of the strips is the smallest of all geometric lengths. Three strips {#sec:sec:three-strips} ------------ The holomorphic field for three parallel strips reads $$\begin{aligned} \label{eq:3-strip-function} \mathcal{B}(\xi) &= H \sqrt{ \frac{[\xi^{2}-{b_{0}}^{2}(H)] [\xi^{2}-{b_{1}}^{2}(H)] [\xi^{2}-{b_{2}}^{2}(H)]} {(\xi^{2}-e_{0}^{2}) (\xi^{2}-e_{1}^{2}) (\xi^{2}-e_{2}^{2})}},\end{aligned}$$ with ${e_{0}}= w$, ${e_{1}}= w + 2s$, and ${e_{2}}= 3w + 2s$. ### Meissner state {#sec:sec:sec:3-strip-Meissner} For small fields, where the entire system is in the Meissner state, i.e., no vortices have penetrated in either of the strips, the holomorphic field reduces to $$\begin{aligned} \label{eq:field-function-three-strips} \mathcal{B}(\xi) &= H\sqrt{\frac{\xi^{2} (\xi^{2}-{b}^{2})^{2}} {(\xi^{2}-{e_{0}}^{2})(\xi^{2}-{e_{1}}^{2})(\xi^{2}-{e_{2}}^{2})}},\end{aligned}$$ where the remaining parameter ${b}$ ($={b_{1}}= {b_{2}}$, note that ${b_{0}}= 0$) is determined from requiring a vanishing total current in the outer strip pair, $$\begin{aligned} \label{eq:neutral-current-3-strips} \int\limits_{{e_{1}}}^{{e_{2}}} &\frac{dx\, x\, b^{2}} {\sqrt{(x^{2}-{e_{0}}^{2})(x^{2}-{e_{1}}^{2})({e_{2}}^{2}-x^{2})}} \\ &\quad\qquad\qquad\qquad= \int\limits_{{e_{1}}}^{{e_{2}}} \frac{dx\, x^{3}}{\sqrt{(x^{2}-{e_{0}}^{2})(x^{2}-{e_{1}}^{2})({e_{2}}^{2}-x^{2})}}. \nonumber\end{aligned}$$ With the substitution $x^{2} \to {e_{1}}^{2} + ({e_{2}}^{2}-{e_{1}}^{2}) t^{2}$ the solution can formally be expressed through $$\begin{aligned} \label{eq:a-for-3-strips} {b}^{2} &= {e_{0}}^{2} + ({e_{1}}^{2} - {e_{0}}^{2})\frac{\operatorname{E}(\kappa)}{\operatorname{K}(\kappa)},\end{aligned}$$ where the elliptic integrals, defined in [Eqs. ]{} and , are evaluated at the imaginary argument $\kappa = \sqrt{({e_{2}}^{2}-{e_{1}}^{2}) /({e_{0}}^{2}-{e_{1}}^{2})}$, $\kappa^{2}<0$. In two asymptotic regimes the above result simplifies to $$\begin{aligned} \label{eq:a-for-3-strips-limits} {b}^{2} &\simeq \left\{\begin{aligned} &4(w+s)^{2} &\mathrm{for\ } s \gg w,\\ &w^{2} \Big[1 + \frac{16}{\log{(32w/s)}}\Big] &\mathrm{for\ } s \ll w.\\ \end{aligned}\right.\end{aligned}$$ The first limit ($s \gg w$) describes three almost isolated strips, while in the latter case of nearby strips with $s \ll w$ a logarithmic dependence of ${b}$ on $s$ shows up, analogous to the expression for two strips. Focusing on the regime of nearby strips $s \ll w$, we find that the flux $$\begin{aligned} \label{eq:flux-3-strips} \phi_{g} &=\int\limits_{e_{0}}^{e_{1}}\! dx\,B_{z}(x) \approx H 6w \frac{\pi \sqrt{2}}{3} \frac{1}{\log{(32w/s)}}\end{aligned}$$ passing through each of the two gaps carries a substantial fraction of the flux $\phi_{\mathrm{b}} = H 6w$ that is blocked by the strips. The field enhancement $\sim H\sqrt{w^{2}/sd}$ at the strip edges ${e_{0}}$ and ${e_{1}}$ \[see [Eq. ]{}\] is found to be parametrically $\sim\sqrt{w/s}$ larger than at the outermost edge ${e_{2}}$. A more detailed calculation reveals, that the field strength is largest near ${e_{0}}$, followed by a slightly lower field near ${e_{1}}$, $$\begin{aligned} \frac{B_{z}({e_{1}}-d/2)}{B_{z}({e_{0}}+d/2)} = 1 - \frac{21 w^{2} - 5 {b}^{2}}{{b}^{2}-w^{2}}\frac{s}{4w}.\end{aligned}$$ We conclude that the critical field ${H_{s}}$ \[as discussed in [Eq. ]{} \] is first reached at the edges $\pm {e_{0}}$ (strip index $k=0$) where the field enhancement is most pronounced. Thus, the geometrical barrier is first suppressed in the central strip and vortices start to penetrate the innermost strip beyond $$\begin{aligned} \label{eq:penetration-field-3-strips} {H_{p}}^{k=0} &\approx {H_{s}}\sqrt{\frac{s d}{8 w^{2}}} \log(32w/s).\end{aligned}$$ This critical field is parametrically similar to the field of first penetration of the double strip geometry, see [Eq. ]{}. ### Penetrated state(s) {#sec:sec:sec:3-strip-Penetrated} In general, for a multiple strip geometry, the strips are not equivalent and the penetration of vortices starts at a different field value for each strip. The penetration sequence may depend on the geometrical setup as well as on the boundary condition at $y \to \pm \infty$ (shunted vs. unshunted ends). In particular, we shall compare our results to the findings by Mawatari *et al.*, who considered a system of three *shunted* strips in Ref. [\[\]]{}. As the external field $H$ increases beyond ${H_{p}}^{k=0}$, vortices populate the innermost strip (${b_{0}}\neq 0$), while the two other strips remain free of flux (${b_{1}}= {b_{2}}= {b}$). The field distribution then is given by $$\begin{aligned} \label{eq:field-function-three-strips-penetrated} \mathcal{B}(\xi) &= H\sqrt{\frac{(\xi^{2}-{b_{0}}^{2}) (\xi^{2}-{b}^{2})^{2}} {(\xi^{2}-{e_{0}}^{2})(\xi^{2}-{e_{1}}^{2})(\xi^{2}-{e_{2}}^{2})}},\end{aligned}$$ where the two parameters ${b_{0}}$ and ${b}$ (now both depending on $H$) are fixed by the constraints of critical field strength ${H_{s}}$ near the edge ${e_{0}}$ and vanishing net current $$\begin{aligned} \label{eq:vanishing-current-three-strips} \int\limits_{{e_{1}}}^{{e_{2}}} dx\,I(x) = 0\end{aligned}$$ in the outer strips. The outer strip pair will first be penetrated by vortices only at a higher field ${H_{p}}^{k=1}$, where a critical field strength ${H_{s}}$ is reached at the edge ${e_{1}}$. At this particular field, the requirement that the field strength is critical at both edges ${e_{0}}$ and ${e_{1}}$ while the outer dome has not yet developed (${b_{1}}= {b_{2}}= {b}$), gives a relation between ${b_{0}}$ and ${b}$ of the form $$\begin{aligned} \label{eq:condition-for-hp2} {b_{0}}^{2} &= w^{2}\Big[1 - \frac{8({b}^{2}-w^{2})} {16w^{2} + 5({b}^{2}-w^{2})}\Big].\end{aligned}$$ Inserting this relation ${b_{0}}({b})$ into the constraint of vanishing net current in the outer strip fixes the last degree of freedom ${b}$ and permits to express the second penetration field through $$\begin{aligned} \label{eq:second-penetration-field} {H_{p}}^{k=1} &= {H_{s}}\sqrt{\frac{32 w^{4}sd} {(w^{2}-{b_{0}}({b})^{2})({b}^{2}-w^{2})^{2}}}.\end{aligned}$$ ![Penetration fields ${H_{p}}^{k=0}$ (dotted) and ${H_{p}}^{k=1}$ (dashed) as a function of $s/w$ for a system of three coplanar superconducting strips. The vertical scale is fixed by the specific choice of the ratio $d/w$. The position of the parameters ${b_{0}}$ and ${b}={b_{1}}={b_{2}}$ corresponding to these two penetration fields is shown in figure \[fig:dome-evolution-at-hp2\].[]{data-label="fig:hp2"}](plot_16_bis.eps){width=".482\textwidth"} Solving [Eqs. ]{} and numerically, we show the results for the two penetration fields ${H_{p}}^{k=0}$ and ${H_{p}}^{k=1}$ in [Fig. ]{}\[fig:hp2\]. Using the same numerical solution, we visualize in [Fig. ]{}\[fig:dome-evolution-at-hp2\] the dome boundaries ${b_{0}}$ and ${b}$ within the strips at the first (second) penetration field ${H_{p}}^{k=0}$ (${H_{p}}^{k=1}$) for different values of the gap width $2s$. We observe that ${b}(H)$ changes only little between ${H_{p}}^{k=0}$ and ${H_{p}}^{k=1}$. This finding allows us to give an estimate for ${H_{p}}^{k=1}$; indeed, inserting ${b}({H_{p}}^{k=1}) \approx {b}({H_{p}}^{k=0})$ in [Eq. ]{} where ${b}({H_{p}}^{k=0})$ is taken from [Eq. ]{}, we find $$\begin{aligned} \frac{{H_{p}}^{k=1}}{{H_{p}}^{k=0}} & \approx \sqrt{\frac{w^{2}}{w^{2}-{b_{0}}({b})^{2}}} \approx \sqrt{\frac{5+\log(32w/s)}{8}}.\end{aligned}$$ ![Position of the vortex dome boundaries ${b_{0}}$ and ${b}= {b_{1}}= {b_{2}}$ of the three-strip system at the penetration fields ${H_{p}}^{k=0}$ (dotted lines) and ${H_{p}}^{k=1}$ (dashed lines) (see also [Fig. ]{}\[fig:hp2\]) as a function of $s/w$. The strip areas along the positive $x$-axis are indicated in gray. For $H \leq {H_{p}}^{k=0}$, the system is described by one non-vanishing parameter ${b}$ (${b_{0}}= 0$); its dependence on $s$ is shown as a dotted line. Increasing $H$ beyond ${H_{p}}^{k=0}$, a vortex dome forms in the innermost strip ($0< {b_{0}}< w$) reaching a finite width at the second penetration field ${H_{p}}^{k=1}$ (dashed line). For that field, the position ${b}$ has shifted towards the center of the outer strip $2w+2s$ (dashed line). Beyond ${H_{p}}^{k=1}$, a pair of domes forms in the two outer strips (${b_{1}}\neq {b_{2}}$, not shown here).[]{data-label="fig:dome-evolution-at-hp2"}](plot_16.eps){width=".45\textwidth"} Beyond ${H_{p}}^{k=1}$, all three strips are penetrated, and the dome widths are determined by the restriction of no net current in the outer strips and the two critical field conditions at the edges ${e_{0}}$ and ${e_{1}}$. Note, that the central strip is penetrated from both edges while the strips of the pair $k=1$ are penetrated from the inner edges $\pm{e_{1}}$ only. The order in which the strips are populated with vortices depends on the specification of the problem. Indeed, if the same geometrical configuration was *shunted* at both ends ($y \to \pm \infty$), as considered in Ref. [\[\]]{}, the outer strip pair would be populated by vortices from $\pm e_2$, while the innermost strip remains free of flux until much higher fields. Many strips - General picture {#sec:sec:many-strips} ----------------------------- For any finite number of *unshunted* strips, the field enhancement in the Meissner state is strongest at the innermost edges, i.e., the inner edges at $\pm{e_{1}}$ for an even number of strips and the two edges at $\pm{e_{0}}$ of the central strip for an odd number of strips (the only case where a strip is penetrated from both sides). Subsequently, the strips are always penetrated asymmetrically from the inner edges. Specifically, when the applied field is raised beyond the first penetration field, vortices start to populate the innermost strip(s) through the respective edges, while all other strips are still free of flux. Under further increase of $H$, the critical field strength ${H_{s}}$ is successively reached at the inner strip edges $e_{2k-1}$ and vortices penetrate the strip pair $k$, when $H > {H_{p}}^{k}$, with ${H_{p}}^{k} > {H_{p}}^{k-1}$, with $k$ starting from 1 (2) in the case of an odd (even) number of strips. In the limit of a large number of strips, the field strengths in the different gaps are almost the same, such that vortex penetration starts within a narrow field range in all the strips. Summary and Conclusions {#sec:conclusions} ======================= In summary, we have investigated the magnetic response of (two) aligned superconducting strips subject to a perpendicular magnetic field $H$. The penetration of vortices in these systems is dominated by a macroscopic energy barrier, the so-called geometrical barrier. We have found that a narrow slit between the strips completely suppresses this geometrical barrier as manifested in an early field penetration and the collapse of the hysteretic magnetization loop. We have compared the results for a pair of rectangular (platelet shaped) strips to those of various other shapes and geometries. Of particular interest is the comparison with a single elliptic strip, i.e., the generic shape defining demagnetization effects, and a single platelet strip, the simplest system exhibiting a geometric barrier. In the elliptic case, vortices penetrate the sample above a field ${H_{p}}= {H_{s}}d/2w$ ($H_{c1} \leq {H_{s}}< H_{c}$), distribute uniformly inside the sample, and produce a reversible response (in the absence of a surface barrier). The penetration of vortices in a platelet sample is impeded by a geometrical energy barrier (and potentially an additional surface barrier) at the sample edges, ${H_{p}}= {H_{s}}\sqrt{d/w}$. Once this barrier is overcome, vortices occupy a finite region inside the sample (vortex dome), while the rest carries the diamagnetic shielding currents. Upon decreasing the applied field, the strip shows a irreversible response, where the penetrated flux is trapped inside the sample until the vortex dome expands to the sample edges. These qualitative features remain valid for an array of rectangular strips. The attention of the present work has mainly focused on the double strip. In the regime of a small gap parameter $s \ll w$, where the strip system is equivalent to a single strip of width $2W = 4w + 2s$ cut in half by a narrow gap, the geometrical barrier is overcome at much lower applied fields $H$. When $d$ is the smallest geometric length, the situation still resembles the one of the single strip; modifications concern the field of first penetration, ${H_{p}}= {H_{s}}\sqrt{d/W} \boldsymbol{[}\sqrt{s/W} \log(4W/s)\boldsymbol{]}$, the exclusive penetration from the inner edges, and the asymmetric shape of the vortex domes leaning towards the gap. When the gap width $2s$ drops below the thickness $d$, the currents rearrange strongly, piling up at the inner surfaces and channeling a larger field through the central opening. In the limit $s/d \to 0$, the geometrical barrier is maximally suppressed and the penetration field ${H_{p}}= {H_{s}}d/2W$ of the double strip coincides with that of a single ellipse with aspect ratio $d/2W$. In contrast to the previously discussed cases where the magnetization decreases beyond penetration, here the magnetic response levels off at the magnitude $M = {H_{s}}W d/2\pi$, a factor $4/\pi$ above the magnetization at the penetration field. In order to study the irreversibility due to the geometric barrier, we have examined the descending branch in the magnetization upon reduction of the field. Remarkably, our analytic results show, that the initial slope of the descending branch is close to the Meissner response, with a correction factor approaching unity when the reversing field approaches ${H_{p}}$ from above; reversing the field at larger $H$, the Meissner slope is changed by a factor $(4-\nu)/\log(16/\nu)$, where $\nu$ depends only on the point $(H,M)$ where the slope is evaluated, $\nu = -4M/Hw^2$. Surprisingly, the correction factor remains close to unity over a wide range of $\nu < 1$. The latter result is equally valid for both the single and narrow-gap ($s \ll d$) double strip. In addition, we have examined the influence of the vortex currents in the Shubnikov phase; while our complex-analysis approach describes the vortex-phase in terms of a non-magnetic medium, it intrinsically exhibits a finite magnetic response. We have shown that the magnetization due to the structure of the vortex state in the dome remains small within the region $H < {H_{s}}$ where our analysis is valid. Finally, we have extended our analysis to $n$ ($\geq 3$) strips and have given a qualitative discussion of the field penetration for this more complex geometry. The suppression of the geometrical barrier can be beneficial in many circumstances, as the hysteretic behavior due to geometrical effects often obscures other interesting physical phenomena. E.g., this has been the case in the identification of the vortex-lattice melting-transition in platelet-shaped layered BiSCCO samples, where the irreversibility line potentially interferes with the first-order melting line: polishing the sample into a prism shape, the geometrical barrier could be suppressed, what allowed to demonstrate experimentally that melting and irreversibility are uncorrelated phenomena [@Majer_95]. Another example is the competition between bulk pinning of vortices and pinning due to surface- and shape effects as analyzed in the present work: again, the suppression of the geometrical barrier provides access to an unambiguous study of bulk pinning phenomena. In evaluating different means to suppress the geometrical barrier, the generation of a simple gap or crack in the sample appears as a rather simple alternative. Recently, the suppression of geometrical barriers in platelet BiSCCO samples has also been observed when tilting the magnetic field [@Segev_11]. This finding has been related to the appearance of Josephson vortex stacks due to the parallel field component weakening the superconductor and channeling the perpendicular component of the magnetic field into the sample. Relating our present study to this experiment, we have modeled a stack of Josephson vortices by a sample crack (of width $2s$) and observe a similar suppression of the geometrical barrier. Another topic where the suppression of geometrical barriers is advantageous is the generation of low-density vortex states, which are difficult to realize in bulk samples due to the rapid accumulation of vortices when increasing $H$ beyond ${H_{s}}$. In elliptic samples, low vortex densities of the order of ${H_{p}}/\Phi_{0} \sim {H_{s}}(d/w)/\Phi_0$ could be achieved; however, it appears difficult to fabricate samples with this shape. In a realistic platelet-shape sample, typical vortex densities are larger, of order ${H_{s}}\sqrt{d/w} /\Phi_{0}$. Introducing a narrow gap in the sample suppresses the geometrical barrier and low vortex densities ${H_{s}}(d/w)\Phi_{0}$ can be reached. Further possible applications of the narrow-gap double strip include the lensing of magnetic fields near the gap, what may be useful for focusing weak magnetic signals. Finally, the analysis and results discussed in this paper may be of relevance in the design of superconducting atom chips for the manipulation of ultra-cold atoms [@Bernon_13]. We acknowledge illuminating discussions with Alexei Koshelev and Eli Zeldov and the financial support of the Swiss National Fonds under the program NCCR MaNEP. [34]{} natexlab\#1[\#1]{}bibnamefont \#1[\#1]{}bibfnamefont \#1[\#1]{}citenamefont \#1[\#1]{}url \#1[`#1`]{}urlprefix\[2\][\#2]{} \[2\]\[\][[\#2](#2)]{} , ****, (). , (). , , , , ****, (). , , , , , , , , ****, (). , ****, (). , , , ****, (). , , , , , , , , , ****, (). , ****, (). , in **, edited by , , (, ), vol. , p. . , **** (). , **, vol.  (, ). , ****, (). , ****, (). , ****, (). , ****, (). , , , ****, (). , ****, (). , ****, (). , , , , ****, (). , ****, (). , ****, (). , ****, () \[Izv. Vuzov Radiofiz. **14**, 819 (1971)\]. , ** (, ). , ** (, ). , , , , , , , ****, (). , ** (, ), ed. , , , , ****, (). , , , , ****, (). , ****, (). , **, vol.  (, ). , ** (, ), ed. , , , , , , , , , , , ****, ().
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present multi-wavelength observations of the brightest galaxies in four X-ray luminous groups at $z\sim0.37$ that will merge to form a cluster comparable in mass to Coma. Ordered by increasing stellar mass, the four brightest group galaxies (BGGs) present a time sequence where BGG-1, 2, & 3 are in merging systems and BGG-4 is a massive remnant ($M_{\ast}=6.7\times10^{11}M_{\odot}$). BGG-1 & 2 have bright, gravitationally bound companions and BGG-3 has two nuclei separated by only 2.5 kpc, thus merging at $z<0.5$ increases the BGG mass by $\gtrsim40$% ($t_{MGR}<2$ Gyr) and $V$-band luminosity by $\sim0.4$ mag. The BGGs’ rest-frame $(B-V)$ colors correspond to stellar ages of $>3$ Gyr, and their tight scatter in $(B-V)$ color ($\sigma_{BV}=0.032$) confirms they formed the bulk of their stars at $z>0.9$. Optical spectroscopy shows no signs of recent ($<1.5$ Gyr) or ongoing star formation. Only two BGGs are weakly detected at [$24\mu$m]{}, and X-ray and optical data indicate the emission in BGG-2 is due to an AGN. All four BGGs and their companions are early-type (bulge-dominated) galaxies, and they are embedded in diffuse stellar envelopes up to $\sim140$ kpc across. The four BGG systems must evolve into the massive, red, early-type galaxies dominating local clusters. Our results show that: 1) massive galaxies in groups and clusters form via dissipationless merging; and 2) the group environment is critical for this process.' author: - 'Kim-Vy H. Tran$^1$, John Moustakas$^2$, Anthony H. Gonzalez$^3$, Lei Bai$^4$, Dennis Zaritsky$^4$, & Stefan J. Kautsch$^3$' bibliography: - '/Users/vy/aastex/tran.bib' title: The Late Stellar Assembly of Massive Cluster Galaxies Via Major Merging --- Introduction ============ Whether massive cluster galaxies at $z\sim0$ build up a significant fraction of their stellar masses at $z<1$ via merging continues to be an intensely debated question. In the paradigm of hierarchical formation, galaxies continue to grow via accretion of smaller satellites [@peebles:70], and recent models predict that brightest cluster galaxies assemble half their stellar masses at $z<0.5$ [@delucia:07b]. However, direct merging is not favored in virialized clusters due to their high velocity dispersions ($\sigma_{1D}\sim1000$ [ km s$^{-1}$]{}). Brightest cluster galaxies (BCG) are particularly compelling tracers of mass growth because they are the most massive galaxies in the universe, can be observed to high redshifts [@gunn:75], and should have a rich merger history. A number of BCGs in the local universe have multiple nuclei [@hoessel:85; @lauer:86; @laine:03] which suggests recent merging. However, studies of BCGs in the more distant universe disagree as to whether BCGs have grown significantly in stellar mass since redshift $z\sim1$ [@brough:02; @whiley:08]. While merging of red, luminous members is observed in groups and dynamically young clusters [@vandokkum:99; @tran:05b; @mcintosh:07], the merging frequency and thus its importance remains uncertain. The key may be environment: merging is rare in established clusters but more frequent in groups that are accreted onto clusters. Because groups have a higher fraction of passive early-type galaxies relative to the field [@wilman:05; @jeltema:07], merging between such systems will form massive, passive galaxies. Our study focuses on the brightest galaxies in four X-ray luminous galaxy groups at $z\sim0.37$ discovered in the [*Las Campanas Distant Cluster Survey*]{}[^1] [@gonzalez:01]. These four galaxy groups are gravitationally bound to each other and form an extended structure that will assemble into a galaxy cluster comparable in mass to Coma by the present epoch [@gonzalez:05 hereafter G05]. By studying galaxies in the group environment prior to cluster assembly, we directly test whether galaxy-galaxy merging on group scales can drive stellar growth in the massive galaxies observed in local clusters, and whether a large fraction of new stars is formed in the process. Throughout the paper, we use $H_0=70$ km s$^{-1}$ Mpc$^{-1}$, $\Omega_M=0.3$, and $\Omega_{\Lambda}=0.7$; at $z=0.37$, this corresponds to a scale of 5.12 kpc per arcsec and a lookback time of 4 Gyr. Observations ============ Follow-up $Chandra$ ACIS-I imaging and optical spectroscopy of the unique super-group 1120 field confirmed the existence of four X-ray luminous groups at $z\sim0.37$ (hereafter SG1120). These groups lie within a projected region of $\sim3$ Mpc and have X-ray temperatures of $T_X=1.7-3.0$ keV (G05). Our study combines imaging and spectroscopy from six different observatories[^2]. Optical & Near-Infrared Imaging {#kcorr} ------------------------------- The optical photometry is measured from VLT/VIMOS imaging in $BVR$ for a nearly contiguous $18'\times20'$ region, and from a contiguous mosaic ($11'\times18'$) composed of 10 pointings taken with the HST/ACS in F814W. The high resolution ACS imaging is critical for identifying close galaxy pairs and merger signatures. Two pointings with KPNO/FLAMINGOS also provide a contiguous $K_s$ mosaic ($16'\times19'$). Using the HST/ACS mosaic as the master detection image, line-matched catalogs were generated using SExtractor v2.5.0 [@bertin:96]. Rest-frame absolute magnitudes, K-corrections, and stellar masses were determined with [*k-correct*]{} v4.1 [@blanton:07] which uses templates based on models from @bruzual:03 [hereafter BC03]. We used MAG$_{-}$AUTO photometry from the $BVRK_s$ imaging and assumed minimum photometric uncertainties in each bandpass of 0.05 mag. The photometry have also been corrected for foreground Galactic extinction using the @schlegel:98 dust maps and the @odonnell:94 Milky Way extinction curve, assuming $R_V=3.1$. Optical Spectroscopy -------------------- The initial survey using VLT/VIMOS confirmed that four of the five X-ray regions are galaxy groups at $z\sim0.37$ while the fifth is a cluster at $z=0.48$ (G05). Targets were selected from a magnitude-limited catalog ($R\leq22.5$ mag), and the spectra reduced using a combination of IRAF[^3] routines and custom software; see @tran:05a for details on the reduction and redshift determination. Further spectroscopy was obtained using Magellan/LDSS3 and VLT/FORS2. In both cases, targets were selected from $K_s$ catalogs ($K_s\leq20$ mag); for reference, the brightest supergroup galaxy has $K_s=14.62$ mag. Guided by the redshift distribution, we define the supergroup’s redshift range to be $0.32\leq z\leq0.39$ where the brightest group galaxies lie at $0.354\leq z\leq0.372$ (Table \[tab:bgg\]). These spectra provided 97 new members for a total of 198 confirmed supergroup galaxies across the approximately $20'\times20'$ region. The medium resolution spectroscopy corresponds to 0.7Å pix$^{-1}$ (LDSS3), 1.65Å pix$^{-1}$ (FORS2), and 2.5Å pix$^{-1}$ (VIMOS). The spectral range for all three instruments covers \[OII\]$\lambda3727$ to \[OIII\]$\lambda5007$ for most supergroup members. [lrrrrrr]{} 1a & 0.3540 & $-22.46$ & 0.86 & $2.9\times10^{11}$ & & $329\pm66$\ 1b & 0.3532 & $-21.49$ & 0.85 & $1.2\times10^{11}$ & & $329\pm66$\ 2a & 0.3706 & $-22.97$ & 0.94 & $5.0\times10^{11}$ & $1.5\times10^{44}$ & $421\pm83$\ 2b & 0.3704 & $-22.04$ & 0.92 & $2.1\times10^{11}$ & & $421\pm83$\ 3 & 0.3713 & $-23.20$ & 0.94 & $5.4\times10^{11}$ & $8.0\times10^{43}$ & $588\pm90$\ 4 & 0.3720 & $-23.38$ & 0.91 & $6.7\times10^{11}$ & & $565\pm130$\ \[tab:bgg\] [$24\mu$m]{} Imaging -------------------- Wide-field MIPS imaging ($0.36^{\circ}\times1^{\circ}$) was taken in slow scan mode, and the data processed with the MIPS Data Analysis Tool [DAT version 3.02; @gordon:05] where array-averaged background subtraction was applied to improve the flat-fielding. At [$24\mu$m]{}, the exposure time is $\sim1000$ s pix$^{-1}$ and the 3$\sigma$ point source detection limit $\sim100\mu$Jy. We use DAOPHOT II [@stetson:87] to detect sources and measure their fluxes, and we follow the same strategy as described in @papovich:04. The spectroscopically confirmed members are correlated with the [$24\mu$m]{} catalog using a $2''$ search radius. Total IR luminosities are determined using star-forming galaxy templates developed by @dale:02, and $L_{IR}$ converted to star formation rates using Eq. 3 in @kennicutt:98b. Results ======= Brightest Group Galaxies ------------------------ Images of the brightest group galaxy (BGG) within each of the four X-ray luminous groups at $z\sim0.37$ are shown in Fig. \[fig:tnails\]. The BGGs are ordered in increasing stellar mass (see §\[kcorr\]) and range from $2.9-6.7\times10^{11}M_{\odot}$ (Table \[tab:bgg\]). All of the BGGs are early-type galaxies with no visual signs of ongoing star formation. These BGGs are in groups with masses less than that of galaxy clusters ($\sim10^{14}M_{\odot}$ vs. $\sim10^{15}M_{\odot}$) at $\sim1/3$ the lookback time, yet they already morphologically resemble the massive early-type galaxies that are common in $z\sim0$ clusters, $e.g.$ Coma [@davis:76]. The BGGs are even comparable in stellar mass to the most massive brightest [*cluster*]{} galaxies in the local universe [$M_{\ast}\sim5\times10^{11}M_{\odot}$; @vonderlinden:07] Spectroscopy uncovers the second striking result: the two lowest mass BGGs are part of merging systems (Fig. \[fig:tnails\]). BGG-1 has a companion galaxy ($D_{proj}=34$ kpc, $\delta$v=170[ km s$^{-1}$]{}) where both are embedded in a common, extended envelope that has a semi-major axis of $\sim140$ kpc; this stellar envelope is especially prominent in the $R$-band imaging and is reminiscent of the stellar plume associated with a major merger forming a brightest cluster galaxy at $z=0.39$ [@rines:07]. BGG-2 also has a companion galaxy ($D_{proj}=48$ kpc, $\delta$v=53[ km s$^{-1}$]{}; note $\delta$v$\approx\sigma_{cz}$) within a faint, extended stellar halo. In the ground-based imaging (PSF$\sim0.7''$), BGG-3 looks like a normal early-type galaxy but the high resolution ACS imaging ($0.05''$ pix$^{-1}$) reveals its spectacular double nucleus (Fig. \[fig:tnails\]). As part of our spectroscopic survey, we aligned a slit along the two nuclei and confirm that both are part of the same galaxy. The nuclei are separated by only 2.5 kpc. BGG-3 also has a diffuse stellar halo. BGG-4 is the most massive galaxy in SG1120 and lies in the group with the highest X-ray temperature (G05). Other than a diffuse stellar halo, BGG-4 has no obvious signs of recent ($t<1$ Gyr) merging, $e.g.$ multiple nuclei. However, recent simulations indicate that massive ellipticals ($M_{\ast}>6\times10^{11}M_{\odot}$) like BGG-4 form via dissipationless merging of two spheroids [@naab:06]. Taken as a whole, these four BGGs present a time sequence where luminous ($L>L^{\ast}$) galaxies continue to merge at late times, as predicted by hierarchical formation [@peebles:70]. The merging galaxies are spheroids that have no visible signs of current or recent star formation, but do have extended stellar envelopes. During the merging process, stars from these spheroids are gravitationally disrupted and likely to contribute to the intracluster light [@mihos:03]. The resulting merger remnant will be a massive, early-type galaxy embedded in an extended, diffuse stellar envelope, $e.g.$ BGG-4. Color-Magnitude Diagram & Stellar Mass -------------------------------------- Figure \[fig:cmd\] shows the rest-frame $(B-V)$ color versus absolute $V$-band magnitude for the 146 supergroup members that fall on the HST/ACS mosaic. Passive members (N=88; $\sim60$%), are well-fit using the color-magnitude (CM) relation measured in CL 1358+62, a massive galaxy cluster at $z=0.33$ [@vandokkum:98a hereafter vD98]. The CM relation is normalized to the passive members where outliers have been iteratively removed using the median absolute deviation. We note that $\sim60$% of the members are E/S0 galaxies. All of the BGGs lie within $\Delta (B-V)>-0.1$ of the CM relation and are thus red. While they all lie blue-ward of the CM relation, assuming the BGGs evolve passively their $(B-V)$ colors will increase such that the BGGs will be tightly clustered around the CM relation by $z\sim0$. Their $(B-V)$ colors already correspond to single stellar populations that are at least 3 Gyr old (solar metallicity; BC03). Included in Fig. \[fig:cmd\] and Table \[tab:bgg\] are the companion galaxies to BGG-1 and BGG-2. Once these companions merge with their BGGs via, $e.g.$ dynamical friction, the resulting BGGs will be $\sim0.4$ magnitudes brighter. The stellar masses determined from the photometry provide further evidence of BGG build-up. The companion galaxies to BGG-1 and BGG-2 are each $\sim40$% of their respective BGG’s mass, and the two nuclei in BGG-3 have a luminosity ratio of about 1:2. Our data indicate that the BGGs grow in mass by $\gtrsim40-50$% during the most recent merger. We estimate the merging timescale by setting it equal to the dynamical friction timescale [@binney:87] and find that $t_{MGR}\lesssim2$ Gyr, consistent with merging timescales measured for luminous red galaxies [@conroy:07] and for close pairs in cosmological simulations [@kitzbichler:08]. A powerful feature of a color-magnitude analysis is using the scatter in color to determine the relative stellar ages of different galaxy populations [@bower:98]. The $(B-V)$ scatter (RMS) for the six members in the BGG systems is $0.032$; this is significantly smaller than the color scatter of the passive members ($\sigma_{BV}=0.116$; Fig. \[fig:cmd\]), $i.e.$ the BGGs are more homogeneous in stellar age than the passive population. Because $(B-V)$ scatter decreases as stellar populations age, we also can constrain the formation redshift of the BGGs. For simplicity, we model the BGGs as single stellar populations (SSPs) and allow SSPs to form any time up to $t$ where $t_0$ is the epoch at $z=0.37$. Using Eq. A3 in vD98, we estimate the $(B-V)$ scatter from 5000 SSPs generated at random times and find that to have $\sigma_{BV}=0.032$, all of the SSPs must form by $t=0.65t_0$. In the current cosmology, this corresponds to all of the stars in the BGGs forming at $z>0.9$. Assuming the BGGs continue to evolve passively, their $(B-V)$ scatter decreases to $\sigma_{BV}\sim0.020$ by $z=0$. This is even smaller than the intrinsic scatter of $\sigma_{BV}=0.028\pm0.004$ measured using the 104 brightest $H$-band early-types in Coma at $z=0.02$ [@eisenhardt:07]. Note that $\sim60-70$% of the supergroup galaxies are already passive and/or red (Fig. \[fig:cmd\]), thus later merging events are unlikely to significantly change the massive BGGs’ mean stellar ages. Composite Optical Spectrum -------------------------- Figure \[fig:bgg1d\] shows the composite rest-frame optical spectrum of the six galaxies in the BGG systems. The composite BGG spectrum has the strong H&K absorption features associated with an old ($>10^9$ Gyr) stellar population and has no emission features nor Balmer absorption. @jeltema:07 also find that intermediate-redshift ($0.2<z<0.6$) BGGs lack \[OII\]$\lambda3727$ emission. Optical emission lines are physically driven by ongoing star formation while strong Balmer absorption, H$\delta$ in particular, provides a fossil record of any star formation that ended within the last $\sim1.5$ Gyr [$e.g.$ @couch:87]. The absence of these features in the composite BGG spectrum indicates that these members have not formed any stars recently, and the spectrum’s $D_N(4000)$ index of 1.72 corresponds to a mean stellar age of $\sim1.7$ Gyr (solar metallicity; BC03). [$24\mu$m]{} Detections ----------------------- Only BGG-2 & 3 are detected at [$24\mu$m]{}, and assuming the mid-infrared emission is due to dusty star formation leads to rates of 6.7 and 3.6 $M_{\odot}$ yr$^{-1}$, respectively. However, dusty star-forming red-sequence galaxies detected at [$24\mu$m]{} tend to show weak \[OII\]$\lambda3727$ emission in their optical spectra [@wolf:05], and neither BGG-2 nor 3 do (Fig. \[fig:bgg1d\]). While [$24\mu$m]{} emission is often attributed to ongoing star formation, it can also be due to an active galactic nucleus (AGN). BGG-2 is an X-ray point source, thus both its X-ray and [$24\mu$m]{} emission are likely due to an AGN. BGG-3’s weaker [$24\mu$m]{} detection makes it difficult to discriminate between a faint AGN and star formation. BGG-2’s multi-wavelength properties are consistent with an old stellar population ($t_{\ast}\gtrsim2$ Gyr) and AGN. Current models favor the formation of luminous red galaxies via merging of progenitors with supermassive black holes; the gravitational perturbations bring fresh gas to the small central region and trigger an AGN phase [$e.g.$ @springel:05a]. BGG-2, and possibly BGG-3, are compatible with such a model. Discussion & Conclusions ======================== These four BGG snapshots provide unprecedented observational evidence for the late ($z<0.5$) stellar assembly of massive spheroidal galaxies in clusters via dissipationless merging of massive progenitors. Two of the four BGGs are part of merging pairs and another has a double nucleus with a separation of only 2.5 kpc. When ordered by increasing stellar mass ($2.9-6.7\times10^{11}M_{\odot}$), the BGGs present a time sequence where galaxy-galaxy merging increases the BGG mass by $\gtrsim40$% ($t_{MGR}\lesssim2$ Gyr) and $V$-band luminosity by $\sim0.4$ mag. Our results are remarkably consistent with recent claims from semi-analytic models that brightest cluster galaxies tend to assemble at $z<0.5$ [@delucia:07b]; such a model also naturally explains the observed number of brightest cluster galaxies in the local universe with multiple nuclei [$\sim20-40$%; @hoessel:85; @laine:03]. Despite the merger activity, the BGGs are dominated by old ($>10^9$ Gyr) stellar populations and show little sign of ongoing or recent star formation. The BGGs have rest-frame $(B-V)$ colors corresponding to mean stellar ages of $>3$ Gyr, and their tight scatter in color ($\sigma_{BV}=0.032$) confirms that the BGGs formed their stars at $z>0.9$. If the BGGs continue to age passively, their color scatter by $z\sim0$ will be even smaller than that of the $\sim100$ brightest early-type galaxies in Coma ($\sigma_{BV}=0.020$ vs. 0.028). Their composite optical spectrum shows no sign of star formation in the last $\sim1.5$ Gyr. While BGG-2 & 3 are weakly detected at [$24\mu$m]{}, the X-ray and optical data indicate that BGG-2’s [$24\mu$m]{} emission is due to AGN rather than star formation, and possibly for BGG-3 as well. SG1120 thus provides unique and powerful confirmation of hierarchical formation on both galaxy and cluster scales because 1) SG1120 has massive galaxies assembling from galaxy-galaxy merging, and 2) it is a proto-cluster assembling from the merging of galaxy groups. Our findings show that the most massive galaxies in groups and clusters form via dissipationless merging, and that the galaxy group environment is critical to forming the massive galaxies dominating local clusters. K.T. acknowledges support from the Swiss National Science Foundation (grant PP002-110576) and thanks J. Blakeslee for help during the initial ACS reduction. J.M. acknowledges support from NASA-06-GALEX06-0030 and Spitzer G05-AR-50443, and L.B. from NASA Spitzer programs through JPL subcontracts \#1255094 and \#1256318. Support was also provided by NASA HST G0-10499, JPL/Caltech SST GO-20683, and Chandra GO2-3183X3. [^1]: The LCDCS was designed to find galaxy clusters at $z\sim0.35-0.9$ and was sensitive to groups only at the lowest redshifts. [^2]: This work is based on observations made with 1) the NASA/ESA [*Hubble Space Telescope*]{}, obtained at the Space Telescope Science Institute, which is operated by AURA, Inc., under NASA contract NAS 5-26555; 2) the [*Spitzer Space Telescope*]{}, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA; 3) the [*Chandra X-ray Telescope*]{}, obtained at the Chandra X-ray Observatory Center, which is operated by the Smithsonian Astrophysics Observatory for and on behalf of NASA under contract NAS8-03060; 4) the 4 meter Mayall telescope at Kitt Peak National Observatory; 5) the 6.5 meter Magellan Telescopes at Las Campanas Observatory, Chile; and 6) the ESO Telescopes at the Paranal Observatories (072.A-0367, 076.B-0362, and 078.B-0409). [^3]: IRAF is distributed by the National Optical Astronomy Observatories.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We propose an information-theoretic alternative to the popular Cronbach alpha coefficient of reliability. Particularly suitable for contexts in which instruments are scored on a strictly nonnumeric scale, our proposed index is based on functions of the entropy of the distributions of defined on the sample space of responses. Our reliability index tracks the Cronbach alpha coefficient uniformly while offering several other advantages discussed in great details in this paper.' address: - | School of Mathematical Sciences\ Rochester Institute of Technology, Rochester, New York, USA\ - | Department of Statistics, Faculty of Science\ Gazi University, Ankara, Turkey\ author: - - bibliography: - 'fg-entropy-ref.bib' title: 'An Information-Theoretic Alternative to the Cronbach’s Alpha Coefficient of Item Reliability' --- , Introduction ============ Suppose that we are given a dataset represented by an $n \times p$ matrix $\bfX$ whose $i$th row ${\mathbf{x}}_i^\top \equiv ({\mathtt{x}}_{i1}, {\mathtt{x}}_{i2},\cdots,{\mathtt{x}}_{ip})$ denotes the $p$-tuple of characteristics, with each ${\mathtt{x}}_{ij} \in \{1,2,3,4,5\}$ representing the Likert-type level (order) of preference of respondent $i$ on item $j$. This Likert-type score is obtained by translating/mapping the response levels $\{{\tt Strong \, Disagree}, {\tt Disagree}, {\tt Neutral}, {\tt Agree}, {\tt Strongly \, Agree}\}$ into pseudo-numbers $\{1,2,3,4,5\}$. [Strong Disagree]{} [Disagree]{} [Neutral]{} [Agree]{} [Strongly Agree]{} --------------------- -------------- ------------- ------------ -------------------- $\bigcirc$ $\bigcirc$ $\bigcirc$ $\bigcirc$ $\bigcirc$ $1$ $2$ $3$ $4$ $5$ A usually crucial part in the analysis of questionnaire data is the calculation of Cronbach’s alpha coefficient which measures the internal consistency or reliability/quality of the data. Let $X= (X_1,X_2,\cdots,X_p)^\top$ be a $p$-tuple representing the $p$ items of a questionnaire. Initially proposed by [@Cronbach:1951:1] and later used and re-explained extensively by thousands of researchers and practitioners like [@Bland:1997:1] Cronbach’s alpha coefficient is a function of the ratio of the sum of the idiosyncratic item variances over the variance of the sum of the items, and is given by $$\alpha = \left(\frac{p}{p-1}\right)\left[1-\frac{\sum_{j=1}^p{\mathbb{V}(X_j)}}{\mathbb{V}\left(\sum_{\ell=1}^p{X_\ell}\right)}\right].$$ The coefficient of Cronbach $\alpha$ will be $1$ if the items are all the same and $0$ if none is related to another. Because it is depend on the variance of the sum of a group of independent variables and the sum of their variances. If the variables are positively correlated, the variance of the sum will be increased. If the items making up the score are all identical and so perfectly correlated, all the ${\mathbb{V}(X_j)}$ will be equal and ${\mathbb{V}\left(\sum_{\ell=1}^p{X_\ell}\right)}=p^2 {\mathbb{V}(X_j)}$, so that $\frac{\sum_{j=1}^p{\mathbb{V}(X_j)}}{\mathbb{V}\left(\sum_{\ell=1}^p{X_\ell}\right)} = \frac{1}{p}$ and $\alpha = 1$.\ The empirical version of Cronbach’s alpha coefficient of internal consistency is given by $$\widehat{{\alpha}} = \left(\frac{p}{p-1}\right)\left[1-\frac{\displaystyle \sum_{j=1}^p{\sum_{i=1}^n{\left({\mathtt{x}}_{ij}-\frac{1}{n}\sum_{i=1}^n{{\mathtt{x}}_{ij}}\right)^2}}} {\displaystyle \sum_{i=1}^n{\left(\sum_{j=1}^p{{\mathtt{x}}_{ij}}-\frac{1}{n}\sum_{i=1}^n{\sum_{j=1}^p{{\mathtt{x}}_{ij}}}\right)^2}}\right].$$ Let $\mathcal{D}=\{{\mathbf{x}}_1,{\mathbf{x}}_2,\cdots,{\mathbf{x}}_n\}$ be a dataset with ${\mathbf{x}}_i^\top = ({\mathtt{x}}_{i1}, {\mathtt{x}}_{i2},\cdots,{\mathtt{x}}_{ip})$. An observation vector ${\mathbf{x}}_i$ will be called a [*zero variation*]{} vector if ${\mathtt{x}}_{ij}= {\tt constant},\,\, j=1,\cdots,p$. Respondents with [*zero variation*]{} response vectors will be referred to as [single minded]{} respondents/evaluators. In fact, [*zero variation*]{} responses essentially reduce a $p$ items survey to a single item survey. Let $X= (X_1,X_2,\cdots,X_p)^\top$ be a $p$-tuple representing the $p$ items of a questionnaire. If $X$ is [*zero variation*]{}, then the Cronbach’s alpha coefficient will be equal to $1$. If $X= (X_1,X_2,\cdots,X_p)^\top$ is [*zero variation*]{}, then $X_j=W$ for $j=1,\cdots,p$, and $\sum_{j=1}^p{X_j}=pW$. As a result, $\sum_{j=1}^p{\mathbb{V}(X_j)}=p\mathbb{V}(W)$ and $\mathbb{V}\left(\sum_{j=1}^p{X_j}\right)=\mathbb{V}(pW)=p^2\mathbb{V}(W)$. Therefore, $$\alpha = \left(\frac{p}{p-1}\right)\left[1-\frac{p\mathbb{V}(W)}{p^2\mathbb{V}(W)}\right] = \left(\frac{p}{p-1}\right)\left[1-\frac{1}{p}\right] = 1$$ We use a straightforward adaptation of the Cronbach’s alpha coefficient to measure [*respondent reliability*]{}. Let $\mathcal{D}=\{{\mathbf{x}}_1,{\mathbf{x}}_2,\cdots,{\mathbf{x}}_n\}$ be a dataset with ${\mathbf{x}}_i^\top = ({\mathtt{x}}_{i1}, {\mathtt{x}}_{i2},\cdots,{\mathtt{x}}_{ip})$. Let the estimated variance of the $i$th respondent be ${\tilde{S}_i^2} = \sum_{j=1}^p{({\mathtt{x}}_{ij}-\bar{{\mathtt{x}}}_i)^2/(p-1)}$. Let $Z_j = \sum_{i=1}^n{{\mathtt{x}}_{ij}}$ represent the sum of the scores given by all the $n$ respondents to item $j$. Our respondent reliability is estimated by $$\widehat{\tilde{\alpha}} = \left(\frac{n}{n-1}\right)\left[1-\frac{\displaystyle \sum_{i=1}^n{\sum_{j=1}^p{\left({\mathtt{x}}_{ij}-\frac{1}{p}\sum_{j=1}^p{{\mathtt{x}}_{ij}}\right)^2}}} {\displaystyle \sum_{j=1}^p{\left(\sum_{i=1}^n{{\mathtt{x}}_{ij}}-\frac{1}{p}\sum_{j=1}^p{\sum_{i=1}^n{{\mathtt{x}}_{ij}}}\right)^2}}\right]$$ Given a data matrix $\bfX$, respondent reliability can be computed in practice by simply taking the Cronbach’s alpha coefficient of $\bfX^\top$, the transpose of the data matrix $\bfX$. Let $m$ be the number of [*nonzero variation*]{}. If $m \ll p$ and $m/n$ is very small, then respondent reliability will be very poor. Despite its widespread use of Likert-type data since it creation, Cronbach’s alpha coefficient is rigorously speaking not suitable for categorical data for the simple reason that averages on ordinal measurements are often difficult to interpret at best and misleading at worst. For many years researchers working on the clustering of Likert-type inappropriately resorted to average-driven methods like kMeans clustering. Fortunately, there has been a surge of contributions to the clustering of categorical data whereby appropriate methods have been used. At the heart of the clustering of categorical data is the need to define appropriate measure of similarity. Recognizing the possibility to preprocess Likert-type questionnaire data into a collection of estimate probability distributions over the sample spaces of responses, many authors have developed powerful, scalable and highly techniques for clustering categorical data, most of them based on information-theoretic [@Cover:1991:1] concepts like entropy [@Huang:1998:1], [@Guha:2000:1], [@Barbara:2002:1], [@San:2004:1], [@Li:2004:1], [@Chen:2005:1], [@Li:2006:1], [@Meila:2007:1], [@Cai:2007:1], mutual information, variation of information [@Meila:2003:1], along with many other distances and measures on probability distributions like the Bhattacharya distance [@Bhattacharya:1943:1], [@Mak:1996:1], [@Choi:2003:1], [@Goudail:2004:1], [@You:2009:1], [@Reyes:2006:1], the Kullback-Leibler divergence and the Hellinger distance just to name a few. In this paper, we use information-theoretic tools and concepts to create several measures of internal consistency of questionnaire data. Information-Theoretic Measures of Internal Data Consistency =========================================================== Let $X_j$ represent one of the questions on the questionnaire, and consider the $n$ responses, $\{{\mathtt{x}}_{1j}, \cdots, {\mathtt{x}}_{ij}, \cdots,{\mathtt{x}}_{nj}\}$ provided by the $n$ evaluators. Let ${\mathbf{v}}_j = ({\mathrm{v}}_{j1}, \cdots,{\mathrm{v}}_{jk},\cdots,{\mathrm{v}}_{jK})^\top$ denote the vector containing the relative frequencies of each Likert level for question $j$. With a total of $n$ questionnaires collected, we have $$\begin{aligned} \widehat{{\mathrm{v}}}_{jk} = \frac{1}{n}\sum_{i=1}^n{I({\rm x}_{ij}=k)}, \qquad k=1,2,\cdots,K \quad \text{and} \quad j=1,2,\cdots,p. \label{eq:entr:item:0}\end{aligned}$$ Using , one can then form probabilistic vectors $\widehat{{\mathbf{v}}}_j^\top = (\widehat{{\mathrm{v}}}_{j1}, \cdots, \widehat{{\mathrm{v}}}_{jk}, \cdots,\widehat{{\mathrm{v}}}_{jK})$, for $j=1,2,\cdots,p$. Each vector $\widehat{{\mathbf{v}}}_j$ essentially represents an approximate probability distribution on the sample space made up of the $K$ response levels. Using this probabilistic representation of each question $j$, we can compare the variability of each item of the questionnaire using the entropy, specifically $$\begin{aligned} {H}(\widehat{{\mathbf{v}}}_j) = -\sum_{k=1}^K{\widehat{{\mathrm{v}}}_{jk}\log_2(\widehat{{\mathrm{v}}}_{jk})} \label{eq:entr:item:1}\end{aligned}$$ We can imagine a transformation of the $n \times p$ data matrix ${\mathbf{X}}$ into a probabilistic $p \times K$ counterpart ${\mathbf{V}}$ where each row represent the approximate probability distribution of the corresponding question (item). The entropy of each question indicates the variability of the answers given by students on that question. For a given course and a given instructor, a small value of this entropy would indicate a greater degree of agreement of his/her student on that item, and therefore suggest a more careful examination of the scores on that item. As far as the relationship between items is concerned, information theory also provides a wealth of measures. The symmetrized Kullback-Leibler divergence given by $${\tt KL}_2({\mathbf{v}}_i, {\mathbf{v}}_j) = \frac{1}{{2}}\Big\{{\tt KL}({\mathbf{v}}_i, {\mathbf{v}}_j) +{\tt KL}({\mathbf{v}}_j, {\mathbf{v}}_i)\Big\} = \frac{1}{{2}}{\sum_{k=1}^K\left\{{\mathrm{v}}_{ik}\log\left(\frac{{\mathrm{v}}_{ik}}{{\mathrm{v}}_{jk}}\right)+ {\mathrm{v}}_{jk}\log\left(\frac{{\mathrm{v}}_{jk}}{{\mathrm{v}}_{ik}}\right)\right\}},$$ where $${\tt KL}({\mathbf{v}}_i, {\mathbf{v}}_j) = \sum_{k=1}^K{{\mathrm{v}}_{ik}\log\left(\frac{{\mathrm{v}}_{ik}}{{\mathrm{v}}_{jk}}\right)} \quad \textrm{and} \quad {\tt KL}({\mathbf{v}}_j, {\mathbf{v}}_i) = \sum_{k=1}^K{{\mathrm{v}}_{jk}\log\left(\frac{{\mathrm{v}}_{jk}}{{\mathrm{v}}_{ik}}\right)},$$ is usually the default measure used by most authors. The [*Kullback-Leibler*]{} divergence is closely related the [*mutual information*]{} $$I({\mathbf{v}}_i, {\mathbf{v}}_j) = \sum_{k=1}^K\left\{\sum_{l=1}^K\left\{{\mathrm{v}}_{ik,jl}\log_2\left(\frac{{\mathrm{v}}_{ik,jl}}{{\mathrm{v}}_{ik}{\mathrm{v}}_{jl}}\right)\right\}\right\},$$ which has been used extensively in machine learning to define a distance known as the [*Variation of Information*]{}, and defined by $${\tt VI}({\mathbf{v}}_i, {\mathbf{v}}_j) = H({\mathbf{v}}_i) + H({\mathbf{v}}_j)- 2 I({\mathbf{v}}_i, {\mathbf{v}}_j).$$ Many other non-information-theoretic similarity and variation measures operating on probabilistic vectors can be used to further investigate several aspects of the categorical data at hand. One that have been extensively used in the machine learning and data mining community is the Bhattacharya distance [@Bhattacharya:1943:1] is given by $${\tt BC}({\mathbf{v}}_i, {\mathbf{v}}_j) = -\log {\tt F}({\mathbf{v}}_i, {\mathbf{v}}_j),$$ where $${\tt F}({\mathbf{v}}_i, {\mathbf{v}}_j) = \sum_{k=1}^K{\sqrt{{\mathrm{v}}_{ik}{\mathrm{v}}_{jk}}},$$ is known as the Bhattacharya coefficient or Fidelity coefficient. The Bhattacharya distance ${\tt BC}({\mathbf{v}}_i, {\mathbf{v}}_j)$ measures the overlap between ${\mathbf{v}}_i$ and ${\mathbf{v}}_j$. The Bhattacharya distance has been immensely used in various data mining and machine learning applications [@Mak:1996:1], [@Choi:2003:1], [@Goudail:2004:1], [@You:2009:1]. It is interesting to note that the Bhattachrya distance is related to [*total variation*]{} measure defined by $$\Delta({\mathbf{v}}_i, {\mathbf{v}}_j) = \frac{1}{{2}}\sum_{k=1}^K{|{\mathrm{v}}_{ik}-{\mathrm{v}}_{jk}|} = \frac{1}{{2}}\|{\mathbf{v}}_i-{\mathbf{v}}_j\|_1$$ where $\|\cdot\|_1$ is the $\ell_1$ norm. Another very commonly used distance is the Hellinger distance between ${\mathbf{v}}_i$ and ${\mathbf{v}}_j$ is given by $${\tt Hellinger}({\mathbf{v}}_i, {\mathbf{v}}_j) = \frac{1}{\sqrt{2}}\sqrt{\sum_{k=1}^K{(\sqrt{{\mathrm{v}}_{ik}}-\sqrt{{\mathrm{v}}_{jk}})^2}} =\frac{1}{\sqrt{2}}\|\sqrt{{\mathbf{v}}_i}-\sqrt{{\mathbf{v}}_j}\|_2,$$ where $\|\cdot\|_2$ is the Euclidean norm or $\ell_2$ norm, $\sqrt{{\mathbf{v}}_i} = (\sqrt{{\mathrm{v}}_{i1}},\cdots,\sqrt{{\mathrm{v}}_{iK}})$ and $\sqrt{{\mathbf{v}}_j} = (\sqrt{{\mathrm{v}}_{j1}},\cdots,\sqrt{{\mathrm{v}}_{jK}})$. Let $Q$ denote an instrument (questionnaire) for which the realized matrix of obtained responses is given by $\bfX$ with entries ${\mathtt{x}}_{ij} \in \{1,2, \cdots, K\}$. We propose an information-theoretic measure of the reliability of $Q$, referred to as the information consistency ratio of $Q$ and given by $$\begin{aligned} \varphi = 1 - \frac{\underset{i=1,\cdots,n}{\mathtt{min}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\}} {\underset{{\mathbf{z}}}{\mathtt{max}}\Big\{H\left({{\mathbf{z}}}\right)\Big\}} = 1 - \frac{\underset{i=1,\cdots,n}{\mathtt{min}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\}}{H\left(\frac{1}{K},\cdots,\frac{1}{K}\right)}= 1 - \frac{\underset{i=1,\cdots,n}{\mathtt{min}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\}}{\log_2(K)}, \label{eq:entr:index:1}\end{aligned}$$ where each $\widehat{{\mathbf{z}}}_i = \{\widehat{{\mathrm{z}}}_{ik}, \, k=1,2,\cdots,K\}$ defines an approximate probability distribution on the sample space of possible responses, and $H(\cdot)$ is the entropy function, with $$\begin{aligned} \widehat{{\mathrm{z}}}_{ik} = \frac{1}{p}\sum_{j=1}^p{I({\rm x}_{ij}=k)} \quad \text{and} \quad {H}(\widehat{{\mathbf{z}}}_i) = -\sum_{k=1}^K{\widehat{{\mathrm{z}}}_{ik}\log_2(\widehat{{\mathrm{z}}}_{ik})}. \label{eq:entr:stud:1}\end{aligned}$$ Let ${\mathbf{z}}$ denote any probability measure defined on some $K$-dimensional sample space, with each ${\mathrm{z}}_k = \Pr\{E_k\}, \, k=1.2,\cdots,K$. Let $H(\cdot)$ denote the entropy function, such that for every ${\mathbf{z}}$, we have $H({\mathbf{z}}) = -\sum_{k=1}^K{{{\mathrm{z}}}_{k}\log_2({{\mathrm{z}}}_{k})}$. Then $$\underset{{\mathbf{z}}}{\mathtt{max}}\Big\{H\left({{\mathbf{z}}}\right)\Big\} = \log_2(K).$$ Since entropy essentially measures uncertainty (disturbance), the probability measure for which the uncertainty is the largest is the probability measure ${\mathbf{z}}^*$ in which all the events are equally likely, i.e., ${\mathrm{z}}_k^* = \Pr\{E_k\}=\frac{1}{K}, \, k=1.2,\cdots,K$. $$\begin{aligned} \underset{{\mathbf{z}}}{\mathtt{max}}\Big\{H\left({{\mathbf{z}}}\right)\Big\} = H({\mathbf{z}}^*) = H\left(\frac{1}{K},\cdots,\frac{1}{K}\right) = -\sum_{k=1}^K{\frac{1}{K}\log_2\left(\frac{1}{K}\right)}=\log_2(K).\end{aligned}$$ Let $Q_0$ denote a special questionnaire whose items are all mutually independent (unrelated). Then the corresponding information consistency ratio $\varphi_0$ of $Q_0$, is such that $$\underset{p \rightarrow \infty}{\lim}{\varphi_0} = 0.$$ With $Q_0$ denoting a questionnaire whose items that are all mutually independent (unrelated), the matrix of realized responses has entries ${\mathtt{x}}_{ij}$ that a realization of the discrete uniform distribution on $\{1,2,\cdots,K\}$, or specifically, ${\mathtt{x}}_{ij} \sim {\tt uniform}(1,2,\cdots,K)$. It follows that for each $i=1,2,\cdots,n$, we must have $$\underset{p \rightarrow \infty}{\lim}{\widehat{{\mathrm{z}}}_{ik}} = \underset{p \rightarrow \infty}{\lim}\left\{\frac{1}{p}\sum_{j=1}^p{I({\rm x}_{ij}=k)}\right\} = \frac{1}{K}, \quad k=1,2,\cdots,K.$$ In other words, given enough questions (items), the empirical proportion of answers will converge to its theoretical counterpart by the law of large number. We therefore have the uniform generation of answers, the limiting distribution $$\underset{p \rightarrow \infty}{\lim}{\widehat{{\mathbf{z}}}_{i}} = {\mathbf{z}}^* = \left(\frac{1}{K},\cdots,\frac{1}{K}\right).$$ Finally, since all the response distributions will tend to converge to the same maximal measure ${\mathbf{z}}^*$, i.e. $\widehat{{\mathbf{z}}}_{i} \overset{\mathscr{D}}{\rightarrow} {\mathbf{z}}^*$, for $i=1,2,\cdots,n$, we must have $$\underset{i=1,\cdots,n}{\mathtt{min}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\} \overset{P}{\rightarrow} H({\mathbf{z}}^*) = \underset{{\mathbf{z}}}{\mathtt{max}}\Big\{H\left({{\mathbf{z}}}\right)\Big\},$$ and therefore $$\underset{p \rightarrow \infty}{\lim}{\varphi_0} = 1 - \frac{\underset{i=1,\cdots,n}{\mathtt{min}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\}} {\underset{{\mathbf{z}}}{\mathtt{max}}\Big\{H\left({{\mathbf{z}}}\right)\Big\}} = 1 - \frac{H({\mathbf{z}}^*)}{H({\mathbf{z}}^*)} = 1-1 = 0.$$ Let $Q_+$ denote a special questionnaire whose items are all identical. Then the corresponding information consistency ratio $\varphi_+$ of $Q_+$, is such that ${\lim}{\varphi_+} = 1$. With $Q_+$ denoting a questionnaire whose items that are all identical, the matrix of realized responses has entries ${\mathtt{x}}_{ij} = c$, for some constant $c \in \{1,2,\cdots,K\}$. Then for each $i=1,2,\cdots,n$, there exists $k_+ \in \{1,2,\cdots,K\}$ such that $$\widehat{{\mathrm{z}}}_{ik} = \left\{ \begin{array}{ll} 1 & \quad k = k_+ \\ 0 & \quad k \neq k_+ \end{array}\right.$$ In other words, with $Q_+$, the approximate distributions $\widehat{{\mathbf{z}}}_i$ of the answers of each respondent are of the form $(1,0,\cdots,0)$, or $(0,1,\cdots,0)$ or $(0,0,\cdots,1)$. Therefore, for $Q_+$, we must have $H(\widehat{{\mathbf{z}}}_i) = 0, \quad i=1,\cdots,n$, with the result being $\underset{i=1,\cdots,n}{\mathtt{min}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\}=0$, and therefore $$\varphi_+ = 1 - \frac{\underset{i=1,\cdots,n}{\mathtt{min}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\}} {\underset{{\mathbf{z}}}{\mathtt{max}}\Big\{H\left({{\mathbf{z}}}\right)\Big\}} = 1 - \frac{0}{H({\mathbf{z}}^*)} = 1-0 = 1.$$ \[def:phi:2\] Let $Y_i$ represent the most frequently occurring answer in respondent $i$’s vector of $p$ answers. It is easy to see that $Y_i$ has the same sample space as each question/item, namely the same Likert scale in our case. Using the random variables $Y_i$, we can then define $\widehat{{\mathbf{w}}} = (\widehat{{\mathtt{w}}}_{1}, \cdots,\widehat{{\mathtt{w}}}_{k},\cdots,\widehat{{\mathtt{w}}}_{K})^\top$ in the same manner that we define ${\mathbf{v}}_j$ earlier. More specifically, we have $$\begin{aligned} Y_i = \underset{k=1,\cdots,K}{\mathtt{argmax}}\left\{\frac{1}{p}\sum_{j=1}^p{I(X_{ij}=k)}\right\} \quad \text{and} \quad \widehat{{\mathtt{w}}}_{k} = \frac{1}{n}\sum_{i=1}^n{I(Y_{i}=k)}. \label{eq:entr:max:1}\end{aligned}$$ The entropy of $\widehat{{\mathbf{w}}}$ is given by $$\begin{aligned} H(\widehat{{\mathbf{w}}}) = -\sum_{k=1}^K{\widehat{{\mathtt{w}}}_{k}\log_2(\widehat{{\mathtt{w}}}_{k})}. \label{eq:entr:max:2}\end{aligned}$$ [*The random variable $Y_i$ is maximal in a set-theoretic sense, and and can be thought of as the categorical analogue of the sum of numeric $X_j$’s.*]{} Using $\widehat{{\mathbf{w}}}$, an alternative definition of the [*information consistency ratio*]{} $\varphi$ is $$\begin{aligned} \varphi= 1 - \frac{\underset{i=1,\cdots,n}{\mathtt{min}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\}} {H\left({\widehat{{\mathbf{w}}}}\right)}. \label{eq:entr:index:2}\end{aligned}$$ An even more stringent measure of the information consistency ratio is given by $$\begin{aligned} \varphi= 1 - \frac{\underset{i=1,\cdots,n}{\mathtt{max}}\Big\{H\left(\widehat{{\mathbf{z}}}_i\right)\Big\}} {H\left({\widehat{{\mathbf{w}}}}\right)}. \label{eq:entr:index:3}\end{aligned}$$ Demonstration of Properties of $\varphi$ ======================================== We use a simple simulation setup to empirically compare the different measures presented in this paper. We set $p=50$ and $n=1000$ and we vary the ratio of perfectly reliable components from $10\%$ to $100\%$ by $10\%$. For $i=1,\cdots,n$ and $j=1,\cdots,n$, draw the ${\mathtt{x}}_{ij}$’s uniformly with replacement from $\{1,2,\cdots,K\}$, that is, $${\tt Draw} \quad {\mathtt{x}}_{ij} \sim {\tt uniform}(1,2,\cdots,K).$$ Randomly replace $100{\tt c}\%$ of the columns of ${\mathbf{X}}$ with the same column of constant values, where ${\tt c} \in \{0.1,0.2,\cdots,0.9,1\}$. Table shows the simulated values of the information consistency ratio and Cronbach’s alpha coefficient for different fractions of of reliable components in the instrument. Figure is a direct pictorial representation of the numbers from Table , and we can see that the Cronbach alpha coefficient is less strick than the information consistency ratio. [@ lrrrrrrrrrr]{}\ \ Fraction & $10$ & $20$ & $30$ & $40$ & $50$ & $60$ & $70$ & $80$ & $90$ & $100$\ $\varphi_1$ & $0.230$ & $0.270$ & $0.330$ & $0.440$ & $0.520$ & $0.630$ & $0.740$ & $0.900$ & $1.000$ & $1.000$\ $\varphi_2$ & $0.000$ & $0.000$ & $0.020$ & $0.080$ & $0.140$ & $0.240$ & $0.360$ & $0.520$ & $0.720$ & $1.000$\ $\varphi_3$ & $0.230$ & $0.270$ & $0.330$ & $0.440$ & $0.520$ & $0.630$ & $0.740$ & $0.900$ & $1.000$ & $1.000$\ $\varphi_4$ & $0.000$ & $0.000$ & $0.020$ & $0.080$ & $0.140$ & $0.240$ & $0.360$ & $0.520$ & $0.720$ & $1.000$\ ${\tt Cronbach}$ & $0.380$ & $0.700$ & $0.820$ & $0.910$ & $0.940$ & $0.960$ & $0.980$ & $0.990$ & $1.000$ & $1.000$\ \ ![Comparative curves of $\varphi$ and Cronbach alpha as measures of internal consistency.[]{data-label="fig:entropy:vs:cronbach:1"}](fg-index-1 "fig:"){width="10cm" height="8cm"}\ Conclusion and Discussion ========================= We have proposed and developed an information-theoretic measure of internal data consistency et demonstrated via straightforward simulation that it does indeed capture the amount of information potentially contained in the data for,the purposes of performing all kinds of pattern for the data. We have also provided several many other measures of similarity over probabilistic vectors that we intend to use for further refined our proposed information consistency ratio $\varphi$. We intend to conduct a larger simulation study to establish our proposed measure on a stronger footing. We also plan to compare the predictive power of ICR to Cronbach’s alpha coefficient on various real and simulated data. Acknowledgements {#acknowledgements .unnumbered} ================ Ernest Fokoué wishes to express his heartfelt gratitude and infinite thanks to Our Lady of Perpetual Help for Her ever-present support and guidance, especially for the uninterrupted flow of inspiration received through Her most powerful intercession.
{ "pile_set_name": "ArXiv" }
--- abstract: 'This article provides a pedagogical introduction to the basic concepts of chiral perturbation theory and is designed as a text for a two-semester course on that topic. Chapter 1 serves as a general introduction to the empirical and theoretical foundations which led to the development of chiral perturbation theory. Chapter 2 deals with QCD and its global symmetries in the chiral limit; the concept of Green functions and Ward identities reflecting the underlying chiral symmetry is elaborated. In Chap. 3 the idea of a spontaneous breakdown of a global symmetry is discussed and its consequences in terms of the Goldstone theorem are demonstrated. Chapter 4 deals with mesonic chiral perturbation theory and the principles entering the construction of the chiral Lagrangian are outlined. Various examples with increasing chiral orders and complexity are given. Finally, in Chap. 5 the methods are extended to include the interaction between Goldstone bosons and baryons in the single-baryon sector, with the main emphasis put on the heavy-baryon formulation. At the end, the method of infrared regularization in the relativistic framework is discussed.' author: - | Stefan Scherer[^1]\ Institut für Kernphysik\ Johannes Gutenberg-Universität Mainz\ J. J. Becher Weg 45\ D-55099 Mainz\ Germany date: | MKPH-T-02-09\ July 23, 2002 title: Introduction to Chiral Perturbation Theory --- Introduction {#chap_i} ============ Scope and Aim of the Review {#sec_sar} --------------------------- The present review has evolved from two courses I have taught several times at the Johannes Gutenberg-Universität, Mainz. The first course was an introduction to chiral perturbation theory (ChPT) which only covered the purely mesonic sector of the theory. In the second course the methods were extended to also include baryons. I have tried to preserve the spirit of those lectures in this article in the sense that it is meant to be a [*pedagogical introduction*]{} to the basic concepts of chiral perturbation theory. By this I do not mean that the material covered is trivial, but that rather I have deliberately also worked out those pieces which by the “experts” are considered as well known. In particular, I have often included intermediate steps in derivations in order to facilitate the understanding of the origin of the final results. My intention was to keep a balance between mathematical rigor and illustrations by means of (numerous) simple examples. This article addresses both experimentalists and theorists! Ideally, it would help a graduate student interested in theoretical physics getting started in the field of chiral perturbation theory. However, it is also written for an experimental graduate student with the purpose of conveying some ideas why the experiment she/he is performing is important for our theoretical understanding of the strong interactions. My precedent in this context is the review by A. W. Thomas [@Thomas] which appeared in this series many years ago and served for me as an introduction to the cloudy bag model. Finally, this article clearly is not intended to be a comprehensive overview of the numerous results which have been obtained over the past two decades. For obvious reasons, I would, right at the beginning, like to apologize to all the researchers who have made important contributions to the field that have not been mentioned in this work. Readers interested in the present status of applications are referred to lecture notes and review articles as well as conference proceedings [@Bernstein:zq; @Bernstein:pm; @Bernstein:2002]. The present article is organized as follows. Chapter 1 contains a general introduction to the empirical and theoretical foundations which led to the development of chiral perturbation theory. Many of the technical aspects mentioned in the introduction will be treated in great detail later on. Chapter 2 deals with QCD and its global symmetries in the chiral limit and the concept of Green functions and Ward identities reflecting the underlying chiral symmetry is elaborated. In Chap. 3 the idea of a spontaneous breakdown of a global symmetry is discussed and its consequences in terms of the Goldstone theorem are demonstrated. Chapter 4 deals with mesonic chiral perturbation theory and the principles entering the construction of the chiral Lagrangian are outlined. Various examples with increasing chiral orders and complexity are given. Finally, in Chap. 5 the methods are extended to include the interaction between Goldstone bosons and baryons in the single-baryon sector with the main emphasis put on the heavy-baryon formulation. At the end, the method of infrared regularization in the relativistic framework is discussed. Some technical details and simple illustrations are relegated to the Appendices. Introduction to Chiral Symmetry and Its Application to Mesons and Single Baryons {#sec_icsamsb} -------------------------------------------------------------------------------- In the 1950’s a description of the strong interactions in the framework of quantum field theory seemed to fail due to an ever increasing number of observed hadrons as well as a coupling constant which was too large to allow for a sensible application of perturbation theory [@Gross:1998jx]. The rich spectrum of hadrons together with their finite sizes (i.e., non-point-like behavior showing up, e.g., in elastic electron-proton scattering through the existence of form factors) were the first hints pointing to a substructure in terms of more fundamental constituents. A calculation of the anomalous magnetic moments of protons and neutrons in the framework of a pseudoscalar pion-nucleon interaction gave rise to values which were far off the empirical ones (see, e.g., [@Bethe:1956]). On the other hand, a simple quark model analysis [@Beg:1964nm; @Morpurgo:1965] gave a prediction $-3/2$ for the ratio $\mu_p/\mu_n$ which is very close to the empirical value of $-1.46$. Nevertheless, the existence of quarks was hotly debated for a long time, since these elementary building blocks, in contrast to the constituents of atomic or nuclear physics, could not be isolated as free particles, no matter what amount of energy was supplied to, say, the proton. Until the early 1970’s it was common to talk about “fictitious” constituents allowing for a simplified group-theoretical classification of the hadron spectrum [@Gell-Mann:nj; @Zweig:1964jf], which, however, could not be interpreted as dynamical degrees of freedom in the context of quantum field theory. In our present understanding, hadrons are highly complex objects built from more fundamental degrees of freedom. These are on the one hand matter fields with spin 1/2 (quarks) and on the other hand massless spin-1 fields (gluons) mediating the strong interactions. Many empirical results of medium- and high-energy physics [@Altarelli:1989ue] such as, e.g., deep-inelastic lepton-hadron scattering, hadron production in electron-positron annihilation, and lepton-pair production in Drell-Yan processes may successfully be described using perturbative methods in the framework of an SU(3) gauge theory, which is referred to as quantum chromodynamics (QCD) [@Gross:1973id; @Weinberg:un; @Fritzsch:pi]. Of particular importance in this context is the concept of asymptotic freedom [@Gross:1973id; @Gross:ju; @Politzer:fx], referring to the fact that the coupling strength decreases for increasing momentum transfer $Q^2$, providing an explanation of (approximate) Bjorken scaling in deep inelastic scattering and allowing more generally for a perturbative approach at high energies. Sometimes perturbative QCD is used as a synonym for asymptotic freedom. In Refs. [@Coleman:1973sx; @Zee:gn] it was shown that Yang-Mills theories, i.e., gauge theories based on non-Abelian Lie groups, provide the only possibility for asymptotically free theories in four dimensions. At present, QCD is compatible with all empirical phenomena of the strong interactions in the asymptotic region. However, one should also keep in mind that many phenomena cannot be treated by perturbation theory. For example, simple (static) properties of hadrons cannot yet be described by [*ab initio*]{} calculations from QCD and this remains one of the largest challenges in theoretical particle physics [@Gross:1998jx]. In this context it is interesting to note that, of the three open problems of QCD at the quantum level, namely, the “gap problem,” “quark confinement,” and (spontaneous) “chiral symmetry breaking” [@Jaffe:2000], the Yang-Mills existence of a mass gap has been chosen as one of the Millennium Prize Problems [@prizeproblems] of the Clay Mathematics Institute. From a physical point of view this problem relates to the fact that nuclear forces are strong and short-ranged. One distinguishes among six quark flavors $u$ (up), $d$ (down), $s$ (strange), $c$ (charm), $b$ (bottom), and $t$ (top), each of which coming in three different color degrees of freedom and transforming as a triplet under the fundamental representation of color SU(3). The interaction between the quarks and the eight gauge bosons does not depend on flavor, i.e., gluons themselves are flavor neutral. On the other hand, due to the non-Abelian character of the group SU(3), also gluons carry “color charges” such that the QCD Lagrangian contains gluon self interactions involving vertices with three and four gluon fields. As a result, the structure of QCD is much more complicated than that of Quantum Electrodynamics (QED) which is based on a local, Abelian U(1) invariance. However, it is exactly the non-Abelian nature of the theory which provides an anti-screening due to gluons that prevails over the screening due to $q\bar{q}$ pairs, leading to an asymptotically free theory [@Nielsen:1981]. Since neither quarks nor gluons have been observed as free, asymptotic states, one assumes that any observable hadron must be in a so-called color singlet state, i.e., a physically observable state is invariant under SU(3) color transformations. The strong increase of the running coupling for large distances possibly provides a mechanism for color confinement [@Gross:ju]. In the framework of lattice QCD this can be shown in the so-called strong coupling limit [@Wilson:1974sk]. However, one has to keep in mind that the continuum limit of lattice gauge theory is approached for a weak coupling and a mathematical proof for color confinement is still missing [@Jaffe:2000]. There still exists no [*analytical*]{} method for the description of QCD at large distances, i.e., at low energies. For example, how the asymptotically observed hadrons, including their rich resonance spectrum, are created from QCD dynamics is still insufficiently understood.[^2] This is one of the reasons why, for many practical purposes, one makes use of phenomenological, more-or-less QCD-inspired, models of hadrons (see, e.g., ). Besides the [*local*]{} SU(3) color symmetry, QCD exhibits further [*global*]{} symmetries. For example, in a strong interaction process, a given quark cannot change its flavor, and if quark-antiquark pairs are created or annihilated during the interaction, these pairs must be flavor neutral. In other words, for each flavor the difference in the number of quarks and antiquarks (flavor number) is a constant of the motion. This symmetry originates in a global invariance under a direct product of U(1) transformations for each quark flavor and is an exact symmetry of QCD independent of the value of the quark masses. Other symmetries are more or less satisfied. It is well known that the hadron spectrum may be organized in terms of (approximately) degenerate basis states carrying irreducible representations of isospin SU(2). Neglecting electromagnetic effects, such a symmetry in QCD results from equal $u$- and $d$-quark (current) masses. The extension including the $s$ quark leads to the famous flavor SU(3) symmetry [@EightfoldWay] which, however, is already significantly broken due to the larger $s$-quark mass. The masses of the three light quarks $u$, $d$, and $s$ are small in comparison with the masses of “typical” light hadrons such as, e.g., the $\rho$ meson (770 MeV) or the proton (938 MeV). On the other hand, the eight lightest pseudoscalar mesons are distinguished by their comparatively small masses.[^3] Within the pseudoscalar octet, the isospin triplet of pions has a significantly smaller mass (135 MeV) than the mesons containing strange quarks. One finds a relatively large mass gap of about 630 MeV between the isospin triplets of the pseudoscalar and the vector mesons, with the gap between the corresponding multiplets involving strange mesons being somewhat smaller. In the limit in which the masses of the light quarks go to zero, the left-handed and right-handed quark fields are decoupled from each other in the QCD Lagrangian. At the “classical” level QCD then exhibits a global $\mbox{U(3)}_L\times\mbox{U(3)}_R$ symmetry. However, at the quantum level (including loops) the singlet axial-vector current develops an anomaly [@Adler:1969gk; @Adler:1969er; @Bardeen:1969md; @Bell:1969ts; @Adler:1970] such that the difference in left-handed and right-handed quark numbers is not a constant of the motion. In other words, in the so-called chiral limit, the QCD-Hamiltonian has a $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$ symmetry. Naturally the question arises, whether the hadron spectrum is, at least approximately, in accordance with such a symmetry of the Hamiltonian. The $\mbox{U(1)}_V$ symmetry is connected to baryon number conservation, where quarks and antiquarks are assigned the baryon numbers $B=1/3$ and $B=-1/3$, respectively. Mesons and baryons differ by their respective baryon numbers $B=0$ and $B=1$. Since baryon number is additive, a nucleus containing $A$ nucleons has baryon number $B=A$. On the other hand, the $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ symmetry is not even approximately realized by the low-energy spectrum. If one constructs from the 16 generators of the group $G=\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ the linear combinations $Q^a_V=Q^a_R+Q^a_L$ and $Q^a_A=Q^a_R-Q^a_L$, $a=1,\cdots,8$, the generators $Q^a_V$ form a Lie algebra corresponding to a SU(3)$_V$ subgroup $H$ of $G$. It was shown in Ref. [@Vafa:tf] that, in the chiral limit, the ground state is necessarily invariant under the group $H$, i.e., the eight generators $Q^a_V$ annihilate the ground state. The symmetry with respect to $H$ is said to be realized in the so-called Wigner-Weyl mode. As a consequence of Coleman’s theorem [@Coleman:1966], the symmetry pattern of the spectrum follows the symmetry of the ground state. Applying one of the axial generators $Q^a_A$ to an arbitrary state of a given multiplet of well-defined parity, one would obtain a degenerate state of opposite parity, since $Q^a_A$ has negative parity and, by definition, commutes with the Hamiltonian in the chiral limit. However, due to Coleman’s theorem such a conclusion tacitly assumes that the ground state is annihilated by the $Q^a_A$. Since such a parity doubling is not observed in the spectrum one reaches the conclusion that the $Q^a_A$ do [*not*]{} annihilate the ground state. In other words, the ground state is not invariant under the full symmetry group of the Hamiltonian, a situation which is referred to as spontaneous symmetry breaking or the Nambu-Goldstone realization of a symmetry [@Nambu:xd; @Nambu:tp; @Nambu:fr]. As a consequence of the Goldstone theorem [@Goldstone:eq; @Goldstone:es], each generator which commutes with the Hamiltonian but does not annihilate the ground state is associated with a massless Goldstone boson, whose properties are tightly connected with the generator in question. The eight generators $Q_A^a$ have negative parity, baryon number zero, and transform as an octet under the subgroup SU(3)$_V$ leaving the vacuum invariant. Thus one expects the same properties of the Goldstone bosons, and the light pseudoscalar octet ($\pi,K,\eta$) qualifies as candidates for these Goldstone bosons. The finite masses of the physical multiplet are interpreted as a consequence of the explicit symmetry breaking due to the finite $u$-, $d$-, and $s$-quark masses in the QCD Lagrangian [@Gasser:1982ap]. Of course, the above (global) symmetry considerations were long known before the formulation of QCD. In the 1960’s they were the cornerstones of the description of low-energy interactions of hadrons in the framework of various techniques, such as the current-algebra approach in combination with the hypothesis of a partially conserved axial-vector current (PCAC) [@Gell-Mann:1964tf; @Adler:1968; @Treiman:1972; @DeAlfaro:1973], the application of phenomenological Lagrangians , and perturbation theory about a symmetry realized in the Nambu-Goldstone mode [@Dashen:eg; @Dashen:ez; @Li:1971vr; @Pagels:se]. All these methods were equivalent in the sense that they produced the same results for “soft” pions [@Dashen:ez] . Although QCD is widely accepted as the fundamental gauge theory underlying the strong interactions, we still lack the analytical tools for [*ab initio*]{} descriptions of low-energy properties and processes. However, new techniques have been developed to extend the results of the current-algebra days and [*systematically*]{} explore corrections to the soft-pion predictions based on symmetry properties of QCD Green functions. The method is called chiral perturbation theory (ChPT) [@Weinberg:1978kz; @Gasser:1983yg; @Gasser:1984gg] and describes the dynamics of Goldstone bosons in the framework of an effective field theory. Although one returns to a field theory in terms of non-elementary hadrons, there is an important distinction between the early quantum field theories of the strong interactions and the new approach in the sense that, now, one is dealing with a so-called [*effective*]{} field theory. Such a theory allows for a [*perturbative*]{} treatment in terms of a momentum—as opposed to a coupling-constant—expansion. The starting point is a theorem of Weinberg stating that a perturbative description in terms of the most general effective Lagrangian containing all possible terms compatible with assumed symmetry principles yields the most general $S$ matrix consistent with the fundamental principles of quantum field theory and the assumed symmetry principles [@Weinberg:1978kz]. The proof of the theorem relies on Lorentz invariance and the absence of anomalies [@Leutwyler:1993iq; @D'Hoker:1994ti] and starts from the observation that the Ward identities satisfied by the Green functions of the symmetry currents are equivalent to an invariance of the generating functional under [*local*]{} transformations [@Leutwyler:1993iq]. For that reason, one considers a [*locally*]{} invariant, effective Lagrangian although the symmetries of the underlying theory may originate in a global symmetry. If the Ward identities contain anomalies, they show up as a modification of the generating functional, which can explicitly be incorporated through the Wess-Zumino-Witten construction [@Wess:yu; @Witten:tw]. All other terms of the effective Lagrangian remain locally invariant. In the present case, the assumed symmetry is the $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$ symmetry of the QCD Hamilton operator in the chiral limit, in combination with a restricted $\mbox{SU(3)}_V\times\mbox{U(1)}_V$ symmetry of the ground state. For center-of-mass energies below the $\rho$-meson mass, the only asymptotic states which can explicitly be produced are the Goldstone bosons. For the description of processes in this energy regime one organizes the most general, chirally invariant Lagrangian for the pseudoscalar meson octet in an expansion in terms of momenta and quark masses. Such an ansatz is naturally suggested by the fact that the interactions of Goldstone bosons are known to vanish in the zero-energy limit. Since the effective Lagrangian by construction contains an infinite number of interaction terms, one needs for any practical purpose an organization scheme allowing one to compare the importance of, say, two given diagrams. To that end, for a given diagram, one analyzes its behavior under a linear rescaling of the [*external*]{} momenta, $p_i\to t p_i$, and a quadratic rescaling of the light quark masses, $m_i\to t^2 m_i$. Applying Weinberg’s power counting scheme [@Weinberg:1978kz], one finds that any given diagram behaves as $t^D$, where $D\geq 2$ is determined by the structure of the vertices and the topology of the diagram in question. For a given $D$, Weinberg’s formula unambiguously determines to which order in the momentum and quark mass expansion the Lagrangian needs to be known. Furthermore, the number of loops is restricted to be smaller than or equal to $D/2\!-\!1$, i.e., Weinberg’s power counting establishes a relation between the momentum expansion and the loop expansion.[^4] Effective field theories are not renormalizable in the usual sense. However, this is no longer regarded as a serious problem, since by means of Weinberg’s counting scheme the infinities arising from loops may be identified order by order in the momentum expansion and then absorbed in a renormalization of the coefficients of the most general Lagrangian. Thus, in any arbitrary order the results are finite. Of course, there is a price to pay: the rapid increase in the number of possible terms as the order increases. Practical applications will hence be restricted to low orders. The lowest-order mesonic Lagrangian, ${\cal L}_2$, is given by the nonlinear $\sigma$ model coupled to external fields [@Gasser:1983yg; @Gasser:1984gg]. It contains two free parameters: the pion-decay constant and the scalar quark condensate, both in the chiral limit. The specific values are not determined by chiral symmetry and must, ultimately, be explained from QCD dynamics. When calculating processes in the phenomenological approximation to ${\cal L}_2$, i.e., considering only tree-level diagrams, one reproduces the results of current algebra [@Weinberg:1978kz]. Since tree-level diagrams involving vertices derived from a Hermitian Lagrangian are always real, one has to go beyond the tree level in order not to violate the unitarity of the $S$ matrix. A calculation of one-loop diagrams with ${\cal L}_2$, on the one hand, leads to infinities which are not of the original type, but also contributes to a perturbative restoration of unitarity. Due to Weinberg’s power counting, the divergent terms are of order ${\cal O}(p^4)$ and can thus be compensated by means of a renormalization of the most general Lagrangian at ${\cal O}(p^4)$. The most general, effective Lagrangian at ${\cal O}(p^4)$ was first constructed by Gasser and Leutwyler [@Gasser:1984gg] and contains 10 physical low-energy constants as well as two additional terms containing only external fields. Out of the ten physically relevant structures, eight are required for the renormalization of the infinities due to the one-loop diagrams involving ${\cal L}_2$. The finite parts of the constants represent free parameters, reflecting our ignorance regarding the underlying theory, namely QCD, in this order of the momentum expansion. These parameters may be fixed phenomenologically by comparison with experimental data [@Gasser:1983yg; @Gasser:1984gg; @Bijnens:1994qh]. There are also theoretical approaches for estimating the low-energy constants in the framework of QCD-inspired models [@Ebert:1985kz; @Espriu:1989ff; @Ebert:1991xd; @Bijnens:1992uz], meson-resonance saturation [@Ecker:yg; @Ecker:1988te; @Donoghue:ed; @Knecht:2001xc; @Leupold:2001vs] and lattice QCD [@Myint:yw; @Golterman:2000mg]. Without external fields (i.e., pure QCD) or including electromagnetic processes only, the effective Lagrangians ${\cal L}_2$ and ${\cal L}_4$ have an additional symmetry: they contain interaction terms involving exclusively an even number of Goldstone bosons. This property is often referred to as normal or even [*intrinsic*]{} parity, but is obviously not a symmetry of QCD, because it would exclude reactions of the type $\pi^0\to\gamma\gamma$ or $K^+K^-\to\pi^+\pi^-\pi^0$. In Ref. [@Witten:tw], Witten discussed how to remove this symmetry from the effective Lagrangian and essentially re-derived the Wess-Zumino anomalous effective action which describes the chiral anomaly [@Wess:yu]. The corresponding Lagrangian, which is of ${\cal O}(p^4)$, cannot be written as a standard local effective Lagrangian in terms of the usual chiral matrix $U$ but can be expressed directly in terms of the Goldstone boson fields. In particular, for the above case, by construction it contains interaction terms with an odd number of Goldstone bosons (odd intrinsic parity). In contrast to the Lagrangian of Gasser and Leutwyler, the Wess-Zumino-Witten (WZW) effective action does not contain any free parameter apart from the number of colors. The excellent description of the neutral pion decay $\pi^0\to\gamma\gamma$ for $N_c=3$ is regarded as a key evidence for the existence of [*three*]{} color degrees of freedom. Chiral perturbation theory to ${\cal O}(p^4)$ has become a well-established method for describing the low-energy interactions of the pseudoscalar octet. For an overview of its many successful applications the interested reader is referred to Refs.  . In general, due to the relatively large mass of the $s$ quark, the convergence in the SU(3) sector is somewhat slower as compared with the SU(2) version. Nevertheless, ChPT in the SU(3) sector has significantly contributed to our understanding of previously open questions. A prime example is the decay rate of $\eta\to\pi\pi\pi$ which current algebra predicts to be much too small. In Ref. [@Gasser:1984pr] it was shown that one-loop corrections substantially increase the theoretical value and remove the previous discrepancy between theory and experiment. For obvious reasons, the question of convergence of the method is of utmost importance. The so-called chiral symmetry breaking scale $\Lambda_{\rm CSB}$ is the dimensional parameter which characterizes the convergence of the momentum expansion [@Manohar:1983md; @Georgi]. A “naive” dimensional analysis of loop diagrams suggests that this scale is given by $\Lambda_{\rm CSB}\approx 4\pi F_0$, where $F_0\approx 93\,\mbox{MeV}$ denotes the pion-decay constant in the chiral limit and the factor $4\pi$ originates from a geometric factor in the calculation of loop diagrams in four dimensions. A second dimensional scale is provided by the masses of the lightest excitations which have been “integrated out” as explicit dynamical degrees of freedom of the theory—in the present case, typically the lightest vector mesons. In a phenomenological approach the exchange of such particles leads to a propagator of the type $(q^2-M^2)^{-1}\approx-M^{-2}(1+q^2/M^2+\cdots)$, where the expansion only converges for $|q^2|< M^2$. The corresponding scale is approximately of the same size as $4\pi F_0$. Assuming a reasonable behavior of the coefficients of the momentum expansion leads to the expectation that ChPT converges for center-of-mass energies sufficiently below the $\rho$-meson mass. Of course, the validity of such a statement depends on the specific process under consideration and the quantum numbers of the intermediate states. Clearly, for a given process, it would be desirable to have an idea about the size of the next-to-leading-order corrections. In the odd-intrinsic-parity sector such a calculation is at least of order ${\cal O}(p^6)$, because the WZW action itself is already of order ${\cal O}(p^4)$. Thus, according to Weinberg’s power counting, one-loop diagrams involving exactly one WZW vertex and an arbitrary number of ${\cal L}_2$ vertices result in corrections of ${\cal O}(p^6)$. Several authors have shown that quantum corrections to the Wess-Zumino-Witten classical action do not renormalize the coefficient of the Wess-Zumino-Witten term . Furthermore, the one-loop counter terms lead to conventional chirally invariant structures at ${\cal O}(p^6)$ . There have been several attempts to construct the most general odd-intrinsic-parity Lagrangian at ${\cal O}(p^6)$ and only recently two independent calculations have found the same number of 23 independent structures in the SU(3) sector [@Ebertshauser:2001nj; @Bijnens:2001bb]. For an overview of the application of ChPT to anomalous processes, the interested reader is referred to Ref. [@Bijnens:xi]. In general, next-to-leading-order corrections to processes in the even-intrinsic-parity sector are of ${\cal O}(p^4)$. However, there are also processes which receive their leading-order contributions at ${\cal O}(p^4)$. In particular, the reactions $\gamma\gamma\to\pi^0\pi^0$ and $\eta\to\pi^0\gamma\gamma$ [@Ametller:1991dp; @Bellucci:1995ay; @Ko:rg; @Bel'kov:1995fj; @Jetter:1995js] have received considerable attention, because the predictions at ${\cal O}(p^4)$ [@Bijnens:1987dc; @Donoghue:ee] and [@Ametller:1991dp], respectively, were in disagreement with experimental results ([@Marsiske:1990hx] and [@Groom:in], respectively). In the case of $\gamma\gamma\to\pi^0\pi^0$ loop corrections at ${\cal O}(p^6)$ lead to a considerably improved description, with the result only little sensitive to the tree-level diagrams at ${\cal O}(p^6)$ [@Bellucci:1994eb]. The opposite picture emerges for the decay $\eta\to\pi^0\gamma\gamma$, where the tree-level diagrams at ${\cal O}(p^6)$ play an important role. A second class of ${\cal O}(p^6)$ calculations includes processes which already receive contributions at ${\cal O}(p^2)$ such as $\pi\pi$ scattering [@Bijnens:1995yn] or $\gamma\gamma\to\pi^+\pi^-$ [@Burgi:1996mm]. Here, ${\cal O}(p^6)$ calculations may be viewed as precision tests of ChPT. The first process is of fundamental importance because it provides information on the mechanism of spontaneous symmetry breaking in QCD [@Bijnens:1995yn]. The second reaction is of particular interest because an old current-algebra low-energy theorem [@Terentev:ix] relates the electromagnetic polarizabilities $\bar{\alpha}$ and $\bar{\beta}$ of the charged pion at ${\cal O}(p^4)$ to radiative pion decay $\pi^+\to e^+\nu_e\gamma$. Corrections at ${\cal O}(p^6)$ were shown to be 12% and 24% of the ${\cal O}(p^4)$ values for $\bar{\alpha}$ and $\bar{\beta}$, respectively [@Burgi:1996mm]. On the other hand, experimental results for the polarizabilities scatter substantially and still have large uncertainties (see, e.g., Ref. [@Unkmeir:2001gw]) and new experimental data are clearly needed to test the accuracy of the chiral predictions. In the SU(3) sector, the first construction of the most general even-intrinsic parity Lagrangian at ${\cal O}(p^6)$ was performed in Ref. [@Fearing:1994ga]. Although it was later shown that the original list of terms contained redundant structures [@Bijnens:1999sh], even the final number of 90 + 4 free parameters is very large, such that, in contrast to the Lagrangian ${\cal L}_4$ of Gasser and Leutwyler, it seems unlikely that all parameters can be fixed through comparison with experimental data. However, chiral symmetry relates different processes to each other, such that groups of interaction terms may be connected with each other and through comparison with experiment the consistency conditions of chiral symmetry may be tested. Furthermore, the same theoretical methods which have been applied to predict the coefficients of ${\cal O}(p^4)$ may be extended to the next order [@Belkov:1994qg] which, however, involves much more work. Chiral perturbation theory has proven to be highly successful in the mesonic sector and, for obvious reasons, one would like to have a generalization including the interaction of Goldstone bosons with baryons. The group-theoretical foundations for a nonlinear realization of chiral symmetry were developed in Refs. [@Weinberg:de; @Coleman:sm; @Callan:sn], which also included the coupling of Goldstone bosons to other isospin or, for the more general case, SU($N$)-flavor multiplets. Numerous low-energy theorems involving the pion-nucleon interaction and its SU(3) extension were derived in the 1960’s by use of current-algebra methods and PCAC. However, a [*systematic*]{} study of chiral corrections to the low-energy theorems has only become possible when the methods of mesonic ChPT were extended to processes with one external nucleon line [@Gasser:1987rb]. The situation turned out to be more involved than in the pure mesonic sector because the loops have a more complicated structure due to the nucleon mass which, in contrast to the Goldstone boson masses, does not vanish in the chiral limit. This introduces a third scale into the problem beyond the pion decay constant and the scalar quark condensate. In particular, it was shown that the relativistic formulation, at first sight, does not provide such a simple connection between the chiral expansion and the loop expansion as in the mesonic sector [@Gasser:1987rb], i.e., higher-order loop diagrams also contribute to lower orders in the chiral expansion of a physical quantity. This observation was taken as evidence for a breakdown of power counting in the relativistic formulation. Subsequently, techniques borrowed from heavy-quark physics were applied to the baryon sector [@Jenkins:1990jv; @Bernard:1992qa], providing a heavy-baryon formulation of ChPT (HBChPT), where the Lagrangian is not only expanded in the number of derivatives and quark masses but also in powers of inverse nucleon masses. The technique is very similar to the Foldy-Wouthuysen method [@Foldy:1949wa]. There have been many successful applications of HBChPT to “traditional” current-algebra processes such as pion photoproduction [@Bernard:1992nc; @Bernard:1996ti] and radiative pion capture [@Fearing:2000uy], pion electroproduction [@Bernard:1992ys; @Bernard:1993bq; @Bernard:1994dt; @Bernard:2000qz], pion-nucleon scattering [@Bernard:1995pa; @Mojzis:1997tu; @Fettes:1998ud; @Fettes:2001cr], to name just a few (for an extensive overview, see Ref. [@Bernard:1995dp]). In all these cases, ChPT has allowed one to either systematically calculate corrections to the old current-algebra results or to obtain new predictions which are beyond the scope of the old techniques. Other applications include the calculation of static properties such as masses and various form factors of baryons [@Bernard:1992ys; @Bernard:1996cc; @Fearing:1997dp; @Kubis:1999xb]. The role of the pionic degrees of freedom has been extensively discussed for real Compton scattering off the nucleon in terms of the electromagnetic polarizabilities . The new frontier of virtual scattering off the nucleon [@Guichon:1995pu; @Drechsel:1996ag; @Roche:2000ng] has also been addressed in the framework of ChPT [@Hemmert:1996gr; @Hemmert:1997at; @Hemmert:1999pz; @L'vov:2001fz]. As in the mesonic sector, the most general chiral Lagrangian in the single-baryon sector is needed which, due to the spin degree of freedom, is more complicated [@Gasser:1987rb; @Krause:xc; @Ecker:1995rk; @Fettes:2000gb]. In the baryonic sector, the $\Delta(1232)$ resonance plays a prominent role because its excitation energy is only about two times the pion mass and its (almost) 100 % branching ratio to the decay mode $N \pi$. In Ref. [@Hemmert:1996xg], the formalism of the so-called small scale expansion was developed, which also treats the nucleon-delta mass splitting as a “small” quantity like the pion mass. Subsequently, the formalism was applied to Compton scattering [@Hemmert:1997tj], baryon form factors [@Bernard:1998gv], the $N\Delta$ transition [@Gellas:1998wx] and virtual Compton scattering [@Hemmert:1999pz]. While the heavy-baryon formulation provided a useful low-energy expansion scheme, it was realized in the context of the isovector spectral function entering the calculation of the nucleon electromagnetic form factor that the corresponding perturbation series fails to converge in part of the low-energy region [@Bernard:1996cc]. Various methods have been suggested to generate a power counting which is also valid for the relativistic approach and which respects the singularity structure of Green functions . The so-called “infrared regularization” of Ref. [@Becher:1999he] decomposes one-loop diagrams into singular and regular parts. The singular parts satisfy power counting, whereas the regular parts can be absorbed into local counter terms of the Lagrangian. This technique solves the power counting problem of relativistic baryon chiral perturbation theory at the one-loop level and has already been applied to the calculation of baryon masses in SU(3) ChPT [@Ellis:1999jt], of form factors [@Kubis:2000zd; @Zhu:2000zf; @Kubis:2000aa], pion-nucleon scattering [@Becher:2001hv] as well as the generalized Gerasimov-Drell-Hearn sum rule [@Bernard:2002bs]. At present, the procedure has not yet been generalized to higher-order loop diagrams. In Ref. [@Gegelia:1999qt] another approach, based on choosing appropriate renormalization conditions, was proposed, leading to the correct analyticity structure and a consistent power counting, which can also be extended to higher loops. Finally, the techniques of effective field theory have also been applied to the nucleon-nucleon interaction (see, e.g., Refs.  ). Clearly this is a very important topic in its own right but is beyond the scope of the present work. QCD and Chiral Symmetry {#chap_qcdcs} ======================= Chiral perturbation theory (ChPT) provides a systematic framework for investigating strong-interaction processes at [*low*]{} energies, as opposed to a perturbative treatment of quantum chromodynamics (QCD) at high momentum transfers in terms of the “running coupling constant.” The basis of ChPT is the global $\mbox{SU(3)}_L\times \mbox{SU(3)}_R\times\mbox{U}(1)_V$ symmetry of the QCD Lagrangian in the limit of massless $u$, $d$, and $s$ quarks. This symmetry is assumed to be spontaneously broken down to $\mbox{SU(3)}_V\times\mbox{U(1)}_V$ giving rise to eight massless Goldstone bosons. In this chapter we will describe in detail one of the foundations of ChPT, namely the symmetries of QCD and their consequences in terms of QCD Green functions. Some Remarks on SU(3) {#sec_srsu3} --------------------- The group SU(3) plays an important role in the context of strong interactions, because on the one hand it is the gauge group of QCD and, on the other hand, flavor SU(3) is approximately realized as a global symmetry of the hadron spectrum (Eightfold Way [@Ne'eman:1961cd; @Gell-Mann:xb; @EightfoldWay]), so that the observed (low-mass) hadrons can be organized in approximately degenerate multiplets fitting the dimensionalities of irreducible representations of SU(3). Finally, as will be discussed later in this chapter, the direct product $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ is the chiral-symmetry group of QCD for vanishing $u$-, $d$-, and $s$-quark masses. Thus, it is appropriate to first recall a few basic properties of SU(3) and its Lie algebra su(3) [@Balachandran:ab; @O'Raifeartaigh:vq; @Jones:ti]. The group SU(3) is defined as the set of all unitary, unimodular, $3 \times 3$ matrices $U$, i.e.$U^\dagger U=1$,[^5] and $\mbox{det}(U)=1$. In mathematical terms, SU(3) is an eight-parameter, simply connected, compact Lie group. This implies that any group element can be parameterized by a set of eight independent real parameters $\Theta=(\Theta_1,\cdots, \Theta_8)$ varying over a continuous range. The Lie-group property refers to the fact that the group multiplication of two elements $U(\Theta)$ and $U(\Psi)$ is expressed in terms of eight [*analytic*]{} functions $\Phi_i(\Theta;\Psi)$, i.e. $U(\Theta)U(\Psi)=U(\Phi)$, where $\Phi=\Phi(\Theta;\Psi)$. It is simply connected because every element can be connected to the identity by a continuous path in the parameter space and compactness requires the parameters to be confined in a finite volume. Finally, for compact Lie groups, every finite-dimensional representation is equivalent to a unitary one and can be decomposed into a direct sum of irreducible representations (Clebsch-Gordan series). Elements of SU(3) are conveniently written in terms of the exponential representation[^6] $$\label{2:1:uexp} U(\Theta)=\exp\left(-i\sum_{a=1}^8 \Theta_a \frac{\lambda_a}{2}\right),$$ with $\Theta_a$ real numbers, and where the eight linearly independent matrices $\lambda_a$ are the so-called Gell-Mann matrices, satisfying $$\begin{aligned} \label{2:1:gmme1} \frac{\lambda_a}{2}&=&i\frac{\partial U}{\partial \Theta_a}(0,\cdots,0),\\ \label{2:1:gmme2} \lambda_a&=&\lambda_a^\dagger,\\ \label{2:1:gmme3} \mbox{Tr}(\lambda_a \lambda_b)&=&2\delta_{ab},\\ \label{2:1:gmme4} \mbox{Tr}(\lambda_a)&=&0.\end{aligned}$$ An explicit representation of the Gell-Mann matrices is given by $$\begin{aligned} \label{2:1:gmm} &&\lambda_1=\left(\begin{array}{rrr} 0&1&0\\1&0&0\\0&0&0 \end{array} \right),\quad \lambda_2=\left(\begin{array}{rrr} 0&-i&0\\i&0&0\\0&0&0 \end{array} \right),\quad \lambda_3=\left(\begin{array}{rrr} 1&0&0\\0&-1&0\\0&0&0 \end{array} \right),\nonumber\\ &&\lambda_4=\left(\begin{array}{rrr} 0&0&1\\0&0&0\\1&0&0 \end{array} \right),\quad \lambda_5=\left(\begin{array}{rrr} 0&0&-i\\0&0&0\\i&0&0 \end{array} \right),\quad \lambda_6=\left(\begin{array}{rrr} 0&0&0\\0&0&1\\0&1&0 \end{array} \right),\nonumber\\ && \lambda_7=\left(\begin{array}{rrr} 0&0&0\\0&0&-i\\0&i&0 \end{array} \right),\quad \lambda_8=\sqrt{\frac{1}{3}}\left(\begin{array}{rrr} 1&0&0\\0&1&0\\0&0&-2 \end{array} \right).\end{aligned}$$ The set $\{i\lambda_a\}$ constitutes a basis of the Lie algebra su(3) of SU(3), i.e., the set of all complex traceless skew Hermitian $3\times 3$ matrices. The Lie product is then defined in terms of ordinary matrix multiplication as the commutator of two elements of su(3). Such a definition naturally satisfies the Lie properties of anti-commutativity $$\label{2:1:anticom} [A,B]=-[B,A]$$ as well as the Jacobi identity $$\label{2:1:jacobi} [A,[B,C]]+[B,[C,A]]+[C,[A,B]]=0.$$ In accordance with Eqs. (\[2:1:uexp\]) and (\[2:1:gmme1\]), elements of su(3) can be interpreted as tangent vectors in the identity of SU(3). The structure of the Lie group is encoded in the commutation relations of the Gell-Mann matrices, $$\label{2:1:crgmm} \left[\frac{\lambda_a}{2},\frac{\lambda_b}{2}\right] =i f_{abc}\frac{\lambda_c}{2},$$ where the totally antisymmetric real structure constants $f_{abc}$ are obtained from Eq. (\[2:1:gmme3\]) as $$\label{2:1:fabc} f_{abc}=\frac{1}{4i}\mbox{Tr}([\lambda_a,\lambda_b]\lambda_c).$$ The independent non-vanishing values are explicitly summarized in the scheme of Table \[table:2:1:su3structurconstants\]. Roughly speaking, these structure constants are a measure of the non-commutativity of the group SU(3). $abc$ 123 147 156 246 257 345 367 458 678 ----------- ----- --------------- ---------------- --------------- --------------- --------------- ---------------- ----------------------- ----------------------- $f_{abc}$ 1 $\frac{1}{2}$ $-\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{2}$ $-\frac{1}{2}$ $\frac{1}{2}\sqrt{3}$ $\frac{1}{2}\sqrt{3}$ : \[table:2:1:su3structurconstants\] Totally antisymmetric non-vanishing structure constants of SU(3). The anticommutation relations read $$\label{2:1:acrgmm} \{\lambda_a,\lambda_b\} = \frac{4}{3}\delta_{ab} +2 d_{abc} \lambda_c,$$ where the totally symmetric $d_{abc}$ are given by $$\label{2:1:dabc} d_{abc}=\frac{1}{4}\mbox{Tr}(\{\lambda_a,\lambda_b\}\lambda_c),$$ and are summarized in Table \[table:2:1:su3dsymbols\]. Clearly, the anticommutator of two Gell-Mann matrices is not necessarily a Gell-Mann matrix. For example, the square of a (nontrivial) skew-Hermitian matrix is not skew Hermitian. $abc$ 118 146 157 228 247 256 338 344 ----------- ---------------------- ---------------- ---------------- ------------------------ ------------------------ ------------------------ ------------------------ ----------------------- $d_{abc}$ $\frac{1}{\sqrt{3}}$ $\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{\sqrt{3}}$ $-\frac{1}{2}$ $\frac{1}{2}$ $\frac{1}{\sqrt{3}}$ $\frac{1}{2}$ $abc$ 355 366 377 448 558 668 778 888 $d_{abc}$ $\frac{1}{2}$ $-\frac{1}{2}$ $-\frac{1}{2}$ $-\frac{1}{2\sqrt{3}}$ $-\frac{1}{2\sqrt{3}}$ $-\frac{1}{2\sqrt{3}}$ $-\frac{1}{2\sqrt{3}}$ $-\frac{1}{\sqrt{3}}$ : \[table:2:1:su3dsymbols\] Totally symmetric non-vanishing $d$ symbols of SU(3). Moreover, it is convenient to introduce as a ninth matrix $$\lambda_0 =\sqrt{2/3}\,\mbox{diag}(1,1,1),$$ such that Eqs. (\[2:1:gmme2\]) and (\[2:1:gmme3\]) are still satisfied by the nine matrices $\lambda_a$. In particular, the set $\{i\lambda_a|a=0,\cdots, 8\}$ constitutes a basis of the Lie algebra u(3) of U(3), i.e., the set of all complex skew Hermitian $3\times 3$ matrices. Finally, an [*arbitrary*]{} $3\times 3$ matrix $M$ can then be written as $$\label{2:1:matrixa} M=\sum_{a=0}^8 \lambda_a M_a,$$ where $M_a$ are complex numbers given by $$M_a=\frac{1}{2}\mbox{Tr}(\lambda_a M).$$ The QCD Lagrangian {#sec_qcdl} ------------------ The gauge principle has proven to be a tremendously successful method in elementary particle physics to generate interactions between matter fields through the exchange of massless gauge bosons (for a detailed account see, e.g., [@Abers:qs; @O'Raifeartaigh:vq]). The best-known example is, of course, quantum electrodynamics (QED) which is obtained from promoting the global U(1) symmetry of the Lagrangian describing a free electron,[^7] $$\label{2:2:freel} \Psi\mapsto \mbox{exp}(-i\Theta)\Psi: {\cal L}_{\rm free}=\bar{\Psi}(i\gamma^\mu\partial_\mu -m)\Psi \mapsto {\cal L}_{\rm free},$$ to a local symmetry. In this process the parameter $0\leq \Theta\leq 2\pi$ describing an element of U(1) is allowed to vary smoothly in space-time, $\Theta\to\Theta(x)$, which is referred to as gauging the U(1) group. To keep the invariance of the Lagrangian under local transformations one introduces a four-potential ${\cal A}_\mu$ into the theory which transforms under the gauge transformation ${\cal A}_\mu\mapsto {\cal A}_\mu-\partial_\mu\Theta/e$. The method is referred to as gauging the Lagrangian with respect to U(1): $$\label{2:2:lqed} {\cal L}_{\rm QED}=\bar{\Psi}[i\gamma^\mu(\partial_\mu-i e {\cal A}_\mu)-m]\Psi -\frac{1}{4} {\cal F}_{\mu\nu} {\cal F}^{\mu\nu},$$ where ${\cal F}_{\mu\nu}=\partial_\mu {\cal A}_\nu -\partial_\nu {\cal A}_\mu$.[^8] The covariant derivative of $\Psi$, $$D_\mu \Psi\equiv(\partial_\mu -ie {\cal A}_\mu)\Psi,$$ is defined such that under a so-called gauge transformation of the second kind $$\label{2:2:gtsk} \Psi(x)\mapsto\exp[-i\Theta(x)]\Psi(x),\quad {\cal A}_\mu(x)\mapsto {\cal A}_\mu(x)-\partial_\mu\Theta(x)/e,$$ it transforms in the same way as $\Psi$ itself: $$\label{2:2:cdt} D_\mu\Psi(x)\mapsto\exp[-i\Theta(x)] D_\mu\Psi(x).$$ In Eq. (\[2:2:lqed\]), the term containing the squared field strength makes the gauge potential a dynamical degree of freedom as opposed to a pure external field. A mass term $M^2 {\cal A}^2/2$ is not included since it would violate gauge invariance and thus the gauge principle requires massless gauge bosons.[^9] In the present case we identify the ${\cal A}_\mu$ with the electromagnetic four-potential and ${\cal F}_{\mu\nu}$ with the field strength tensor containing the electric and magnetic fields. The gauge principle has (naturally) generated the interaction of the electromagnetic field with matter. If the underlying gauge group is non-Abelian, the gauge principle associates an independent gauge field with each independent continuous parameter of the gauge group. QCD is the gauge theory of the strong interactions [@Gross:1973id; @Weinberg:un; @Fritzsch:pi] with color SU(3) as the underlying gauge group.[^10] The matter fields of QCD are the so-called quarks which are spin-1/2 fermions, with six different flavors in addition to their three possible colors (see Table \[2:2:table:quarks\]). Since quarks have not been observed as asymptotically free states, the meaning of quark masses and their numerical values are tightly connected with the method by which they are extracted from hadronic properties (see Ref. [@Manohar_PDG] for a thorough discussion). Regarding the so-called current-quark-mass values of the light quarks, one should view the quark mass terms merely as symmetry breaking parameters with their magnitude providing a measure for the extent to which chiral symmetry is broken [@Scheck:ur]. For example, [*ratios*]{} of the light quark masses can be inferred from the masses of the light pseudoscalar octet (see Ref.  [@Leutwyler:1996qg]). Comparing the proton mass, $m_p$ = 938 MeV, with the sum of two up and one down current-quark masses (see Table \[2:2:table:quarks\]), $$\label{2:2:mp} m_p\gg 2m_u+m_d,$$ shows that an interpretation of the proton mass in terms of current-quark mass parameters must be very different from, say, the situation in the hydrogen atom, where the mass is essentially given by the sum of the electron and proton masses, corrected by a small amount of binding energy. flavor u d s -------------- --------------------- --------------------- ----------------------- charge \[e\] $2/3$ $-1/3$ $-1/3$ mass \[MeV\] $5.1\pm 0.9$ $9.3\pm 1.4$ $175\pm 25$ [@Leutwyler:1996qg] [@Leutwyler:1996qg] [@Bijnens:1994ci] flavor c b t charge \[e\] $2/3$ $-1/3$ $2/3$ mass \[GeV\] $1.15-1.35$ $4.0 - 4.4$ $174.3\pm 3.2\pm 4.0$ [@Manohar_PDG] [@Manohar_PDG] [@Manohar_PDG] : \[2:2:table:quarks\] Quark flavors and their charges and masses. The absolute magnitude of $m_s$ is determined using QCD sum rules. The result is given for the $\overline{\mbox{MS}}$ running mass at scale $\mu = 1$ GeV. The light quark masses are obtained from the mass ratios found using chiral perturbation theory, using the strange quark mass as input. The heavy-quark masses $m_c$ and $m_b$ are determined by the charmonium and D masses, and the bottomium and B masses, respectively. The top quark mass $m_t$ results from the measurement of lepton + jets and dilepton + jets channels in the D$\emptyset$ and CDF experiments at Fermilab. The QCD Lagrangian obtained from the gauge principle reads [@Marciano:su; @Altarelli:1981ax] $$\label{2:2:lqcd} {\cal L}_{\rm QCD}=\sum_{f={u,d,s, \atop c,b,t}} \bar{q}_f(i D\hspace{-.6em}/ -m_f)q_f -\frac{1}{4}{\cal G}_{\mu\nu,a}{\cal G}^{\mu\nu}_a.$$ For each quark flavor $f$ the quark field $q_f$ consists of a color triplet (subscripts $r$, $g$, and $b$ standing for “red,” “green,” and “blue”), $$\label{2:2:qf} q_f=\left(\begin{array}{c}q_{f,r} \\q_{f,g}\\q_{f,b}\end{array}\right),$$ which transforms under a gauge transformation $g(x)$ described by the set of parameters $\Theta(x)=[\Theta_1(x),\cdots,\Theta_8(x)]$ according to[^11] $$\label{2:2:qft} q_f\mapsto q_f'=\exp\left[-i\sum_{a=1}^8 \Theta_a(x) \frac{\lambda_a^C}{2}\right]q_f=U[g(x)]q_f.$$ Technically speaking, each quark field $q_f$ transforms according to the fundamental representation of color SU(3). Because SU(3) is an eight-parameter group, the covariant derivative of Eq. (\[2:2:lqcd\]) contains eight independent gauge potentials ${\cal A}_{\mu,a}$, $$\label{2:2:ka} D_\mu\left(\begin{array}{l} q_{f,r}\\q_{f,g}\\q_{f,b}\end{array} \right) =\partial_\mu \left(\begin{array}{l} q_{f,r}\\q_{f,g}\\q_{f,b}\end{array} \right) -ig\sum_{a=1}^8 \frac{\lambda_a^C}{2}{\cal A}_{\mu,a} \left(\begin{array}{l} q_{f,r}\\q_{f,g}\\q_{f,b}\end{array} \right).$$ We note that the interaction between quarks and gluons is independent of the quark flavors. Demanding gauge invariance of ${\cal L}_{\rm QCD}$ imposes the following transformation property of the gauge fields $$\label{2:2:atraf} \frac{\lambda_a^C}{2} {\cal A}_{\mu,a}(x)\mapsto U[g(x)]\frac{\lambda_a^C}{2} {\cal A}_{\mu,a}(x)U^\dagger[g(x)] -\frac{i}{g}\partial_\mu U[g(x)]U^\dagger[g(x)].$$ Again, with this requirement the covariant derivative $D_\mu q_f$ transforms as $q_f$, i.e. $D_\mu q\mapsto D'_\mu q'=U(g)D_\mu q$. Under a gauge transformation of the first kind, i.e., a global SU(3) transformation, the second term on the right-hand side of Eq. (\[2:2:atraf\]) would vanish and the gauge fields would transform according to the adjoint representation. So far we have only considered the matter-field part of ${\cal L}_{\rm QCD}$ including its interaction with the gauge fields. Equation (\[2:2:lqcd\]) also contains the generalization of the field strength tensor to the non-Abelian case, $$\label{2:2:gmunu} {\cal G}_{\mu\nu,a}=\partial_\mu {\cal A}_{\nu,a}-\partial_\nu {\cal A}_{\mu,a} +g f_{abc}{\cal A}_{\mu,b} {\cal A}_{\nu,c},$$ with the SU(3) structure constants given in Table \[table:2:1:su3structurconstants\] and a summation over repeated indices implied. Given Eq. (\[2:2:atraf\]) the field strength tensor transforms under SU(3) as $$\label{2:2:gtrafo} {\cal G}_{\mu\nu}\equiv \frac{\lambda_a^C}{2} {\cal G}_{\mu\nu,a} \mapsto U[g(x)]{\cal G}_{\mu\nu} U^\dagger[g(x)].$$ Using Eq. (\[2:1:gmme3\]) the purely gluonic part of ${\cal L}_{\rm QCD}$ can be written as $$-\frac{1}{2}\mbox{Tr}_C({\cal G}_{\mu\nu} {\cal G}^{\mu\nu}),$$ which, using the cyclic property of traces, $\mbox{Tr}(AB)=\mbox{Tr}(BA)$, together with $UU^\dagger=1$, is easily seen to be invariant under the transformation of Eq. (\[2:2:gtrafo\]). In contradistinction to the Abelian case of QED, the squared field strength tensor gives rise to gauge-field self interactions involving vertices with three and four gauge fields of strength $g$ and $g^2$, respectively. Such interaction terms are characteristic of non-Abelian gauge theories and make them much more complicated than Abelian theories. From the point of view of gauge invariance the strong-interaction Lagrangian could also involve a term of the type $$\label{2:2:ltheta} {\cal L}_\theta=\frac{g^2\bar{\theta}}{64\pi^2}\epsilon^{\mu\nu\rho\sigma} \sum_{a=1}^8{\cal G}^a_{\mu\nu}{\cal G}^a_{\rho\sigma},$$ where $\epsilon_{\mu\nu\rho\sigma}$ denotes the totally antisymmetric Levi-Civita tensor.[^12] The so-called $\theta$ term of Eq. (\[2:2:ltheta\]) implies an explicit $P$ and $CP$ violation of the strong interactions which, for example, would give rise to an electric dipole moment of the neutron (for an upper limit, see Ref. [@Harris:jx]). The present empirical information indicates that the $\theta$ term is small and, in the following, we will omit Eq. (\[2:2:ltheta\]) from our discussion and refer the interested reader to Refs.  [@Pich:1991fq; @Borasoy:2000pq; @Kaiser:2000gs]. Accidental, Global Symmetries of ${\cal L}_{\rm QCD}$ {#sec_agsl} ----------------------------------------------------- ### Light and Heavy Quarks {#subsec_lhq} The six quark flavors are commonly divided into the three light quarks $u$, $d$, and $s$ and the three heavy flavors $c$, $b$, and $t$, $$\label{2:3:mq} \left(\begin{array}{r}m_u=0.005\,\mbox{GeV}\\ m_d=0.009\,\mbox{GeV}\\ m_s=0.175\,\mbox{GeV}\end{array}\right) \ll 1\, \mbox{GeV}\le \left(\begin{array}{r} m_c= (1.15 - 1.35)\, \mbox{GeV}\\ m_b= (4.0 - 4.4)\, \mbox{GeV}\\ m_t=174\,\mbox{GeV}\end{array}\right),$$ where the scale of 1 GeV is associated with the masses of the lightest hadrons containing light quarks, e.g.,  $m_\rho$= 770 MeV, which are not Goldstone bosons resulting from spontaneous symmetry breaking. The scale associated with spontaneous symmetry breaking, $4\pi F_\pi\approx$ 1170 MeV, is of the same order of magnitude [@Pagels:se; @Manohar:1983md; @Georgi]. The masses of the lightest meson and baryon containing a charmed quark, $D^+=c\bar{d}$ and $\Lambda^+_c=udc$, are $(1869.4\pm 0.5)\, \mbox{MeV}$ and $(2284.9\pm 0.6)\,\mbox{MeV}$, respectively [@Groom:in]. The threshold center-of-mass energy to produce, say, a $D^+ D^-$ pair in $e^+ e^-$ collisions is approximately 3.74 GeV, and thus way beyond the low-energy regime which we are interested in. In the following, we will approximate the full QCD Lagrangian by its light-flavor version, i.e., we will ignore effects due to (virtual) heavy quark-antiquark pairs $h\bar{h}$. In particular, Eq. (\[2:2:mp\]) suggests that the Lagrangian ${\cal L}_{\rm QCD}^0$, containing only the light-flavor quarks in the so-called chiral limit $m_u,m_d,m_s\to 0$, might be a good starting point in the discussion of low-energy QCD: $$\label{2:3:lqcd0} {\cal L}^0_{\rm QCD}= \sum_{l=u,d,s}\bar{q}_l i D\hspace{-.6em}/\hspace{.3em} q_l %+\sum_{h=c,b,t}\bar{q}_h(i D\hspace{-.6em}/\hspace{.3em} -m_h) q_h -\frac{1}{4}{\cal G}_{\mu\nu,a} {\cal G}^{\mu\nu}_a.$$ We repeat that the covariant derivative $D\hspace{-.6em}/\hspace{.3em} q_{l}$ acts on color and Dirac indices only, but is independent of flavor. ### Left-Handed and Right-Handed Quark Fields {#subsec_lhrhqf} In order to fully exhibit the global symmetries of Eq. (\[2:3:lqcd0\]), we consider the chirality matrix $\gamma_5=\gamma^5=i\gamma^0\gamma^1 \gamma^2\gamma^3$, $\{\gamma^\mu,\gamma_5\}=0$, $\gamma_5^2=1$,[^13] and introduce projection operators $$\label{2:3:prpl} P_R=\frac{1}{2}(1+\gamma_5)=P_R^\dagger,\quad P_L=\frac{1}{2}(1-\gamma_5)=P_L^\dagger,$$ where the indices $R$ and $L$ refer to right-handed and left-handed, respectively, as will become more clear below. Obviously, the $4\times 4$ matrices $P_R$ and $P_L$ satisfy a completeness relation, $$\label{2:3:prplcompleteness} P_R+P_L=1,$$ are idempotent, i.e., $$\label{2:3:prplidempotent} P_R^2=P_R,\quad P_L^2=P_L,$$ and respect the orthogonality relations $$\label{2:3:prplorthogonality} P_R P_L=P_L P_R=0.$$ The combined properties of Eqs. (\[2:3:prplcompleteness\]) – (\[2:3:prplorthogonality\]) guarantee that $P_R$ and $P_L$ are indeed projection operators which project from the Dirac field variable $q$ to its chiral components $q_R$ and $q_L$, $$\label{2:3:qlr} q_R=P_R q,\quad q_L=P_L q.$$ We recall in this context that a chiral (field) variable is one which under parity is transformed into neither the original variable nor its negative [@Doughty_1990].[^14] Under parity, the quark field is transformed into its parity conjugate, $$P:q(\vec{x},t)\mapsto \gamma_0 q(-\vec{x},t),$$ and hence $$q_R(\vec{x},t)=P_R \,q(\vec{x},t) \mapsto P_R \gamma_0 q(-\vec{x},t) =\gamma_0 P_L q(-\vec{x},t) \neq \pm q_R(-\vec{x},t),$$ and similarly for $q_L$.[^15] The terminology right-handed and left-handed fields can easily be visualized in terms of the solution to the free Dirac equation. For that purpose, let us consider an extreme relativistic positive-energy solution with three-momentum $\vec{p}$,[^16] $$u(\vec{p},\pm)=\sqrt{E+M}\left(\begin{array}{c} \chi_\pm\\ \frac{\vec{\sigma}\cdot\vec{p}}{E+M}\chi_\pm\end{array}\right) \stackrel{\mbox{$E\gg M$}}{\to} \sqrt{E} \left(\begin{array}{r}\chi_\pm\\ \pm\chi_\pm\end{array} \right)=u_\pm(\vec{p}\,),$$ where we assume that the spin in the rest frame is either parallel or antiparallel to the direction of momentum $$\vec{\sigma}\cdot \hat{p} \chi_{\pm}=\pm \chi_\pm.$$ In the standard representation of Dirac matrices we find $$P_R=\frac{1}{2}\left(\begin{array}{rr}1_{2\times2}& 1_{2\times 2}\\1_{2\times 2}&1_{2\times 2}\end{array}\right),\quad P_L=\frac{1}{2}\left(\begin{array}{rr}1_{2\times 2}&-1_{2\times 2}\\ -1_{2\times 2}&1_{2\times 2}\end{array}\right),$$ such that $$P_R u_+=\sqrt{E}\frac{1}{2} \left(\begin{array}{rr}1_{2\times 2}&1_{2\times 2}\\ 1_{2\times 2}&1_{2\times 2}\end{array}\right) \left(\begin{array}{r}\chi_+\\ \chi_+\end{array}\right) =\sqrt{E}\left(\begin{array}{r}\chi_+\\ \chi_+\end{array}\right)=u_+,$$ and similarly $$P_L u_+=0,\quad P_R u_-=0,\quad P_L u_-=u_-.$$ In the extreme relativistic limit (or better, in the zero-mass limit), the operators $P_R$ and $P_R$ project to the positive and negative helicity eigenstates, i.e., in this limit chirality equals helicity. Our goal is to analyze the symmetry of the QCD Lagrangian with respect to independent global transformations of the left- and right-handed fields. In order to decompose the 16 quadratic forms into their respective projections to right- and left-handed fields, we make use of [@Gasser_1989] $$\label{2:3:qgq} \bar{q}\Gamma_i q=\left \{\begin{array}{lcl} \bar{q}_R\Gamma_1 q_R+\bar{q}_L\Gamma_1 q_L&\mbox{for}& \Gamma_1\in\{\gamma^\mu,\gamma^\mu\gamma_5\}\\ \bar{q}_R\Gamma_2 q_L +\bar{q}_L\Gamma_2 q_R&\mbox{for}& \Gamma_2 \in\{1,\gamma_5,\sigma^{\mu\nu}\} \end{array} \right.,$$ where $\bar{q}_R=\bar{q}P_L$ and $\bar{q}_L=\bar{q}P_R.$ Equation (\[2:3:qgq\]) is easily proven by inserting the completeness relation of Eq. (\[2:3:prplcompleteness\]) both to the left and the right of $\Gamma_i$, $$\bar{q}\Gamma_i q=\bar{q}(P_R+P_L)\Gamma_i(P_R+P_L)q,$$ and by noting $\{\Gamma_1,\gamma_5\}=0$ and $[\Gamma_2,\gamma_5]=0$. Together with the orthogonality relations of Eq.  (\[2:3:prplorthogonality\]) we then obtain $$P_R\Gamma_1 P_R=\Gamma_1P_L P_R=0,$$ and similarly $$P_L \Gamma_1 P_L=0,\quad P_R \Gamma_2 P_L=0,\quad P_L \Gamma_2 P_R=0.$$ We stress that the validity of Eq. (\[2:3:qgq\]) is general and does not refer to “massless” quark fields. We now apply Eq. (\[2:3:qgq\]) to the term containing the contraction of the covariant derivative with $\gamma^\mu$. This quadratic quark form decouples into the sum of two terms which connect only left-handed with left-handed and right-handed with right-handed quark fields. The QCD Lagrangian in the chiral limit can then be written as $$\label{2:3:lqcd0lr} {\cal L}^0_{\rm QCD}=\sum_{l=u,d,s} (\bar{q}_{R,l}iD\hspace{-.6em}/\hspace{.3em}q_{R,l}+\bar{q}_{L,l}iD \hspace{-.6em}/\hspace{.3em} q_{L,l})-\frac{1}{4}{\cal G}_{\mu\nu,a} {\cal G}^{\mu\nu}_a.$$ Due to the flavor independence of the covariant derivative ${\cal L}^0_{\rm QCD}$ is invariant under $$\begin{aligned} \label{2:3:u3lu3r} \left(\begin{array}{c}u_L\\d_L\\s_L\end{array}\right) \mapsto U_L\left(\begin{array}{c}u_L\\d_L\\s_L\end{array}\right) =\exp\left(-i\sum_{a=1}^8 \Theta^L_a \frac{\lambda_a}{2}\right) e^{-i\Theta^L}\left(\begin{array}{c}u_L\\d_L\\s_L\end{array}\right), \nonumber\\ \left(\begin{array}{c}u_R\\d_R\\s_R\end{array}\right) \mapsto U_R\left(\begin{array}{c}u_R\\d_R\\s_R\end{array}\right) =\exp\left(-i\sum_{a=1}^8 \Theta^R_a \frac{\lambda_a}{2}\right) e^{-i\Theta^R}\left(\begin{array}{c}u_R\\d_R\\s_R\end{array}\right),\end{aligned}$$ where $U_L$ and $U_R$ are independent unitary $3\times 3$ matrices. Note that the Gell-Mann matrices act in flavor space. ${\cal L}^0_{\rm QCD}$ is said to have a classical [*global*]{} $\mbox{U(3)}_L\times\mbox{U(3)}_R$ symmetry. Applying Noether’s theorem (see, for example, [@Hill_1951; @DeAlfaro:1973]) from such an invariance one would expect a total of $2\times(8+1)=18$ conserved currents. ### Noether’s Theorem {#subsec_nt} In order to identify the conserved currents associated with this invariance, we briefly recall the method of Ref. [@Gell-Mann:np] and consider the variation of Eq. (\[2:3:lqcd0lr\]) under a [*local*]{} infinitesimal transformation.[^17] For simplicity we consider only internal symmetries. To that end we start with a Lagrangian ${\cal L}$ depending on $n$ independent fields $\Phi_i$ and their first partial derivatives, $$\label{2:3:l} {\cal L}={\cal L}(\Phi_i,\partial_\mu\Phi_i),$$ from which one obtains $n$ equations of motion: $$\label{2:3:eom} \frac{\partial \cal L}{\partial\Phi_i}-\partial_\mu \frac{\partial\cal L}{\partial\partial_\mu\Phi_i}=0,\quad i=1,\cdots,n.$$ For each of the $r$ generators of infinitesimal transformations representing the underlying symmetry group, we consider a [*local*]{} infinitesimal transformation of the fields [@Gell-Mann:np],[^18] $$\label{2:3:ltraf} \Phi_i(x)\mapsto\Phi'_i(x)=\Phi_i(x)+\delta\Phi_i(x) =\Phi_i(x)-i\epsilon_a(x) F^a_i[\Phi_j(x)],$$ and obtain, neglecting terms of order $\epsilon^2$, as the variation of the Lagrangian, $$\begin{aligned} \label{2:3:dl} \delta{\cal L}&=& {\cal L}(\Phi'_i,\partial_\mu\Phi'_i) -{\cal L}(\Phi_i,\partial_\mu\Phi_i)\nonumber\\ &=&\frac{\partial\cal L}{\partial\Phi_i}\delta\Phi_i +\frac{\partial\cal L}{\partial\partial_\mu\Phi_i} \partial_\mu\delta\Phi_i\nonumber\\ &=&\epsilon_a(x)\left(-i\frac{\partial\cal L}{\partial\Phi_i} F^a_i-i\frac{\partial\cal L}{\partial\partial_\mu\Phi_i} \partial_\mu F^a_i\right) +\partial_\mu\epsilon_a(x)\left(-i\frac{\partial\cal L}{ \partial\partial_\mu\Phi_i}F^a_i\right)\nonumber\\ &\equiv&\epsilon_a(x)\partial_\mu J^{\mu,a} +\partial_\mu\epsilon_a(x)J^{\mu,a}.\end{aligned}$$ According to this equation we define for each infinitesimal transformation a four-current density as $$\label{2:3:strom} J^{\mu,a}=-i\frac{\partial\cal L}{\partial\partial_\mu\Phi_i} F^a_i.$$ By calculating the divergence $\partial_\mu J^{\mu,a}$ of Eq. (\[2:3:strom\]) $$\begin{aligned} \partial_\mu J^{\mu,a}&=&-i\left(\partial_\mu\frac{\partial\cal L}{\partial \partial_\mu\Phi_i}\right)F^a_i-i\frac{\partial\cal L}{\partial \partial_\mu\Phi_i}\partial_\mu F^a_i\\ &=& -i\frac{\partial\cal L}{\partial \Phi_i} F^a_i-i\frac{\partial\cal L}{\partial\partial_\mu\Phi_i} \partial_\mu F^a_i,\end{aligned}$$ where we made use of the equations of motion, Eq. (\[2:3:eom\]), we explicitly verify the consistency with the definition of $\partial_\mu J^{\mu,a}$ according to Eq. (\[2:3:dl\]). From Eq. (\[2:3:dl\]) it is straightforward to obtain the four-currents as well as their divergences as $$\begin{aligned} \label{2:3:strom2} J^{\mu,a}&=&\frac{\partial \delta\cal L}{\partial \partial_\mu \epsilon_a},\\ \label{2:3:divergenz} \partial_\mu J^{\mu,a}&=&\frac{\partial \delta\cal L}{\partial \epsilon_a}.\end{aligned}$$ For a conserved current, $\partial_\mu J^{\mu,a}=0$, the charge $$\label{2:3:charge} Q^a(t)=\int d^3 x J^a_0(\vec{x},t)$$ is time independent, i.e., a constant of the motion, which is shown in the standard fashion by applying the divergence theorem for an infinite volume with appropriate boundary conditions for $R\to \infty$. So far we have discussed Noether’s theorem on the classical level, implying that the charges $Q^a(t)$ can have any continuous real value. However, we also need to discuss the implications of a transition to a quantum theory. After canonical quantization, the fields $\Phi_i$ and their conjugate momenta $\Pi_i=\partial{\cal L}/\partial (\partial_0 \Phi_i)$ are considered as linear operators acting on a Hilbert space which, in the Heisenberg picture, are subject to the equal-time commutation relations $$\begin{aligned} \label{2:3:gzvr} [\Phi_i(\vec{x},t),\Pi_j(\vec{y},t)]&=&i\delta^3(\vec{x}-\vec{y}) \delta_{ij},\nonumber\\ {[}\Phi_i(\vec{x},t),\Phi_j(\vec{y},t)]&=&0,\nonumber\\ {[}\Pi_i(\vec{x},t),\Pi_j(\vec{y},t)]&=&0.\end{aligned}$$ As a special case of Eq. (\[2:3:ltraf\]) let us consider infinitesimal transformations which are [*linear*]{} in the fields, $$\label{2:3:lt} \Phi_i(x)\mapsto \Phi'_i(x)=\Phi_i(x)-i\epsilon_a(x)t^a_{ij}\Phi_j(x),$$ where the $t^a_{ij}$ are constants generating a mixing of the fields. From Eq. (\[2:3:strom\]) we then obtain[^19] $$\begin{aligned} \label{2:3:j} J^{\mu,a}(x)&=&-it^a_{ij}\frac{\partial {\cal L}}{\partial \partial_\mu\Phi_i}\Phi_j,\\ \label{2:3:q} Q^{a}(t)&=&-i\int d^3x\, \Pi_i(x) t^a_{ij}\Phi_j(x),\end{aligned}$$ where $J^{\mu,a}(x)$ and $Q^{a}(t)$ are now operators. In order to interpret the charge operators $Q^a(t)$, let us make use of the equal-time commutation relations, Eqs. (\[2:3:gzvr\]), and calculate their commutators with the field operators, $$\begin{aligned} \label{2:3:qphi} [Q^a(t),\Phi_k(\vec{y},t)]&=&-it^a_{ij}\int d^3x\, [\Pi_i(\vec{x},t)\Phi_j(\vec{x},t),\Phi_k(\vec{y},t)]\nonumber\\ &=&-t^a_{kj}\Phi_j(\vec{y},t).\end{aligned}$$ Note that we did not require the charge operators to be time independent. On the other hand, for the transformation behavior of the Hilbert space associated with a global infinitesimal transformation, we make an ansatz in terms of an infinitesimal unitary transformation[^20] $$\label{1:5:tz} |\alpha'\rangle=[1+i\epsilon_a G^a(t)]|\alpha\rangle,$$ with Hermitian operators $G^a$. Demanding $$\label{2:3:at} \langle\beta|A|\alpha\rangle=\langle\beta'|A'|\alpha'\rangle\quad \forall\, |\alpha\rangle, |\beta\rangle, \epsilon_a,$$ in combination with Eq. (\[2:3:lt\]) yields the condition $$\begin{aligned} \langle\beta|\Phi_i(x)|\alpha\rangle &=&\langle\beta'|\Phi_i'(x)|\alpha'\rangle\nonumber\\ &=&\langle\beta|[1-i\epsilon_a G^a(t)][\Phi_i(x)-i\epsilon_b t^b_{ij}\Phi_j(x)][1+i\epsilon_c G^c(t)]|\alpha\rangle.\end{aligned}$$ By comparing the terms linear in $\epsilon_a$ on both sides, $$0=-i\epsilon_a[G^a(t),\Phi_i(x)] \underbrace{-i\epsilon_a t^a_{ij}\Phi_j(x)}_{\mbox{$ i\epsilon_a[Q^a(t),\Phi_i(x)]$}},$$ we see that the infinitesimal generators acting on the states of Hilbert space which are associated with the transformation of the fields are identical with the charge operators $Q^a(t)$ of Eq. (\[2:3:q\]). Finally, evaluating the commutation relations for the case of several generators, $$\begin{aligned} \label{2:3:qaqbkom} [Q^a(t),Q^b(t)] &=&-i(t^a_{ij}t^b_{jk}-t^b_{ij}t^a_{jk}) \int d^3x\,\Pi_i(\vec{x},t)\Phi_k(\vec{x},t),\end{aligned}$$ we find the right-hand side of Eq. (\[2:3:qaqbkom\]) to be again proportional to a charge operator, if $$\label{2:3:lrel} t^a_{ij}t^b_{jk}-t^b_{ij}t^a_{jk}=iC_{abc}t^c_{ik},$$ i.e., in that case the charge operators $Q^a(t)$ form a Lie algebra $$\label{2:3:liealgebra} [Q^a(t),Q^b(t)]=iC_{abc}Q^c(t)$$ with structure constants $C_{abc}$. The quantization of the charges (as opposed to continuous values in the classical case) can be understood in analogy to the algebraic construction of the angular momentum eigenvalues in quantum mechanics starting from the su(2) algebra. Of course, for conserved currents, the charge operators are time independent, i.e., they commute with the Hamilton operator of the system. From now on we assume the validity of Eq. (\[2:3:lrel\]) and interpret the constants $t^a_{ij}$ as the entries in the $i$th row and $j$th column of an $n\times n$ matrix $T^a$, $$T^a=\left(\begin{array}{ccc} t^a_{11}&\cdots&t^a_{1n}\\ \vdots & &\vdots\\ t^a_{n1}&\cdots &t^a_{nn} \end{array} \right).$$ Because of Eq. (\[2:3:lrel\]), these matrices form an $n$-dimensional representation of a Lie algebra, $$[T^a,T^b]=iC_{abc}T^c.$$ The infinitesimal, linear transformations of the fields $\Phi_i$ may then be written in a compact form, $$\label{2:3:ltrafo} \left(\begin{array}{c}\Phi_1(x)\\\vdots\\ \Phi_n(x)\end{array} \right)=\Phi(x)\mapsto\Phi'(x)=(1-i\epsilon_a T^a)\Phi(x).$$ In general, through an appropriate unitary transformation, the matrices $T_a$ may be decomposed into their irreducible components, i.e., brought into block-diagonal form, such that only fields belonging to the same multiplet transform into each other under the symmetry group. ### Global Symmetry Currents of the Light Quark Sector {#subsec_gsclqs} The method of Ref. [@Gell-Mann:np] can now easily be applied to the QCD Lagrangian by calculating the variation under the infinitesimal, local form of Eqs. (\[2:3:u3lu3r\]), $$\label{2:3:dlqcd0} \delta{\cal L}^0_{\rm QCD}=\bar{q}_R\left(\sum_{a=1}^8 \partial_\mu \Theta^R_a \frac{\lambda_a}{2}+\partial_\mu \Theta^R\right)\gamma^\mu q_R +\bar{q}_L\left(\sum_{a=1}^8 \partial_\mu \Theta^L_a \frac{\lambda_a}{2}+\partial_\mu \Theta^L\right)\gamma^\mu q_L,$$ from which, by virtue of Eqs. (\[2:3:strom2\]) and (\[2:3:divergenz\]), one obtains the currents associated with the transformations of the left-handed or right-handed quarks $$\begin{aligned} \label{2:3:str} L^{\mu,a}&=&\bar{q}_L\gamma^\mu \frac{\lambda^a}{2}q_L,\quad \partial_\mu L^{\mu,a}=0,\nonumber\\ R^{\mu,a}&=&\bar{q}_R\gamma^\mu \frac{\lambda^a}{2}q_R,\quad \partial_\mu R^{\mu,a}=0.\end{aligned}$$ The eight currents $L^{\mu,a}$ transform under $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ as an $(8,1)$ multiplet, i.e., as octet and singlet under transformations of the left- and right-handed fields, respectively. Similarly, the right-handed currents transform as a $(1,8)$ multiplet under $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$. Instead of these chiral currents one often uses linear combinations, $$\begin{aligned} \label{2:3:v} V^{\mu,a}&=& R^{\mu,a}+L^{\mu,a}=\bar{q}\gamma^\mu\frac{\lambda^a}{2}q,\\ \label{2:3:a} A^{\mu,a}&=&R^{\mu,a}-L^{\mu,a}=\bar{q}\gamma^\mu\gamma_5 \frac{\lambda^a}{2}q,\end{aligned}$$ transforming under parity as vector and axial-vector current densities, respectively, $$\begin{aligned} \label{2:3:pv} P:V^{\mu,a}(\vec{x},t)\mapsto V^a_\mu(-\vec{x},t),\\ \label{2:3pa} P: A^{\mu,a}(\vec{x},t)\mapsto -A_\mu^a(-\vec{x},t).\end{aligned}$$ From Eqs. (\[2:3:strom2\]) and (\[2:3:divergenz\]) one also obtains a conserved singlet vector current resulting from a transformation of all left-handed and right-handed quark fields by the [*same*]{} phase, $$\label{2:3:sv} V^\mu=\bar{q}_R\gamma^\mu q_R+\bar{q}_L\gamma^\mu q_L= \bar{q}\gamma^\mu q, \quad \partial_\mu V^\mu=0.$$ The singlet axial-vector current, $$\begin{aligned} \label{2:3:sa} A^\mu&=&\bar{q}_R \gamma^\mu q_R -\bar{q}_L\gamma^\mu q_L =\bar{q}P_L\gamma^\mu P_R q -\bar{q}P_R \gamma^\mu P_Lq \nonumber\\ &=& \bar{q}\gamma^\mu P_R q -\bar{q}\gamma^\mu P_L q =\bar{q}\gamma^\mu\gamma_5q,\end{aligned}$$ originates from a transformation of all left-handed quark fields with one phase and all right-handed with the [*opposite*]{} phase. However, such a singlet axial-vector current is only conserved on the [*classical*]{} level. This symmetry is not preserved by quantization and there will be extra terms, referred to as anomalies [@Adler:1969gk; @Adler:1969er; @Bardeen:1969md; @Bell:1969ts; @Adler:1970], resulting in[^21] $$\label{2:3:divsa} \partial_\mu A^\mu=\frac{3 g^2}{32\pi^2}\epsilon_{\mu\nu\rho\sigma} {\cal G}^{\mu\nu}_a {\cal G}^{\rho\sigma}_a,\quad \epsilon_{0123}=1,$$ where the factor of 3 originates from the number of flavors. ### The Chiral Algebra {#subsec_ca} The invariance of ${\cal L}^0_{\rm QCD}$ under global $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$ transformations implies that also the QCD Hamilton operator in the chiral limit, $H^0_{\rm QCD}$, exhibits a global $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$ symmetry. As usual, the “charge operators” are defined as the space integrals of the charge densities, $$\begin{aligned} \label{2:3:ql} Q^a_L(t)&=&\int d^3x\, q^\dagger_L(\vec{x},t)\frac{\lambda^a}{2} q_L(\vec{x},t),\quad a=1,\cdots, 8,\\ \label{2:3:qr} Q^a_R(t)&=&\int d^3x\, q^\dagger_R(\vec{x},t)\frac{\lambda^a}{2} q_R(\vec{x},t),\quad a=1,\cdots, 8,\\ \label{2:3:qv} Q_V(t)&=&\int d^3x\, \left[q^\dagger_L(\vec{x},t)q_L(\vec{x},t)+ q^\dagger_R(\vec{x},t)q_R(\vec{x},t)\right].\end{aligned}$$ For conserved symmetry currents, these operators are time independent, i.e., they commute with the Hamiltonian, $$\label{2:3:vrhq} [Q_L^a,H^0_{\rm QCD}]=[Q_R^a,H^0_{\rm QCD}]=[Q_V,H^0_{\rm QCD}]=0.$$ The commutation relations of the charge operators with each other are obtained by using the equal-time commutation relations of the quark fields in the Heisenberg picture, $$\begin{aligned} \label{2:3:comrelf1} \{q_{\alpha,r}(\vec{x},t),q^\dagger_{\beta,s}(\vec{y},t)\}&=& \delta^3(\vec{x}-\vec{y})\delta_{\alpha\beta}\delta_{rs},\\ \label{2:3:comrelf2} \{q_{\alpha,r}(\vec{x},t),q_{\beta,s}(\vec{y},t)\}&=&0,\\ \label{2:3:comrelf3} \{q_{\alpha,r}^\dagger(\vec{x},t),q^\dagger_{\beta,s}(\vec{y},t)\}&=&0,\end{aligned}$$ where $\alpha$ and $\beta$ are Dirac indices and $r$ and $s$ flavor indices, respectively.[^22] The equal-time commutator of two quadratic quark forms is of the type $$\begin{aligned} \label{2:3:fk} \lefteqn{[q^\dagger(\vec{x},t) \Gamma_1 F_1 q(\vec{x},t), q^\dagger(\vec{y},t)\Gamma_2 F_2 q(\vec{y},t)]=}\nonumber\\ &&\Gamma_{1,\alpha\beta}\Gamma_{2,\gamma\delta} F_{1,rs}F_{2,tu} [q^\dagger_{\alpha,r}(\vec{x},t) q_{\beta,s}(\vec{x},t),q^\dagger_{\gamma,t}(\vec{y},t) q_{\delta,u}(\vec{y},t)],\end{aligned}$$ where $\Gamma_i$ and $F_i$ are $4\times 4$ Dirac matrices and $3\times 3$ flavor matrices, respectively. Using $$\label{2:3:abcdfk} [ab,cd]=a\{b,c\}d-ac\{b,d\}+\{a,c\}db-c\{a,d\}b,$$ we express the commutator of Fermi fields in terms of anticommutators and make use of the equal-time commutation relations of Eqs. (\[2:3:comrelf1\]) – (\[2:3:comrelf3\]) to obtain $$\begin{aligned} \lefteqn{[q^\dagger_{\alpha,r}(\vec{x},t) q_{\beta,s}(\vec{x},t),q^\dagger_{\gamma,t}(\vec{y},t) q_{\delta,u}(\vec{y},t)]=}\\ && q^\dagger_{\alpha,r}(\vec{x},t)q_{\delta,u}(\vec{y},t)\delta^3( \vec{x}-\vec{y})\delta_{\beta\gamma}\delta_{st} -q^\dagger_{\gamma,t}(\vec{y},t)q_{\beta,s}(\vec{x},t) \delta^3(\vec{x}-\vec{y})\delta_{\alpha\delta}\delta_{ru}.\end{aligned}$$ With this result Eq. (\[2:3:fk\]) reads $$\begin{aligned} \label{2:3:fkf} \lefteqn{[q^\dagger(\vec{x},t) \Gamma_1 F_1 q(\vec{x},t), q^\dagger(\vec{y},t)\Gamma_2 F_2 q(\vec{y},t)]=}\nonumber\\ &&\delta^3(\vec{x}-\vec{y})\left[ q^\dagger(\vec{x},t)\Gamma_1\Gamma_2 F_1 F_2 q(\vec{y},t) -q^\dagger(\vec{y},t)\Gamma_2 \Gamma_1 F_2 F_1 q(\vec{x},t)\right].\end{aligned}$$ After inserting appropriate projectors $P_{L/R}$, Eq. (\[2:3:fkf\]) is easily applied to the charge operators of Eqs. (\[2:3:ql\]), (\[2:3:qr\]), and (\[2:3:qv\]), showing that these operators indeed satisfy the commutation relations corresponding to the Lie algebra of $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$, $$\begin{aligned} \label{2:2:crqll} [Q_L^a,Q_L^b]&=&if_{abc}Q_L^c,\\ \label{2:3:crqrr} {[Q_R^a,Q_R^b]}&=&if_{abc}Q_R^c,\\ \label{2:3:crqlr} {[Q_L^a,Q_R^b]}&=&0,\\ \label{2:3:crqlvrv} {[Q_L^a,Q_V]}&=&[Q_R^a,Q_V]=0.\end{aligned}$$ It should be stressed that, even without being able to explicitly solve the equation of motion of the quark fields entering the charge operators of Eqs. (\[2:3:ql\]) - (\[2:3:qv\]), we know from the equal-time commutation relations and the symmetry of the Lagrangian that these charge operators are the generators of infinitesimal transformations of the Hilbert space associated with $H^0_{\rm QCD}$. Furthermore, their commutation relations with a given operator specify the transformation behavior of the operator in question under the group $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$. ### Chiral Symmetry Breaking Due to Quark Masses {#subsec_csbdqm} The finite $u$-, $d$-, and $s$-quark masses in the QCD Lagrangian result in explicit divergences of the symmetry currents. As a consequence, the charge operators are, in general, no longer time independent. However, as first pointed out by Gell-Mann, the equal-time-commutation relations still play an important role even if the symmetry is explicitly broken [@Gell-Mann:xb]. As will be discussed later on in more detail, the symmetry currents will give rise to chiral Ward identities relating various QCD Green functions to each other. Equation (\[2:3:divergenz\]) allows one to discuss the divergences in the presence of quark masses. To that end, let us consider the quark-mass matrix of the three light quarks and project it on the nine $\lambda$ matrices of Eq. (\[2:1:matrixa\]), $$\begin{aligned} \label{2:3:qmm} M&=&\left(\begin{array}{ccc} m_u&0&0\\ 0&m_d&0\\ 0&0&m_s \end{array} \right)\nonumber\\ &=&\frac{m_u+m_d+m_s}{\sqrt{6}}\lambda_0 +\frac{(m_u+m_d)/2-m_s}{\sqrt{3}}\lambda_8 +\frac{m_u-m_d}{2}\lambda_3.\end{aligned}$$ In particular, applying Eq. (\[2:3:qgq\]) we see that the quark mass term mixes left- and right-handed fields, $$\label{2:3:lm} {\cal L}_M= -\bar{q}Mq= -(\bar{q}_R M q_L +\bar{q}_L M q_R).$$ The symmetry-breaking term transforms under $\mbox{SU(3)}_L\times \mbox{SU(3)}_R$ as a member of a $(3,3^\ast)+(3^\ast,3)$ representation, i.e., $$\bar{q}_{R,i} M_{ij} q_{L,j} +\bar{q}_{L,i} M_{ij} q_{R,j} \mapsto U_{L,jk}U^\ast_{R,il} \bar{q}_{R,l}M_{ij} q_{L,k}+ (L\leftrightarrow R),$$ where $(U_L,U_R)\in\mbox{SU(3)}_L\times \mbox{SU(3)}_R$. Such symmetry-breaking [*patterns*]{} were already discussed in the pre-QCD era in Refs.[@Glashow:1967rx; @Gell-Mann:rz]. From ${\cal L}_M$ one obtains as the variation $\delta{\cal L}_M$ under the transformations of Eqs. (\[2:3:u3lu3r\]), $$\begin{aligned} \label{2:3:dlm} \delta {\cal L}_M&=& -i\left[ \sum_{a=1}^8 \Theta_a^R \left( \bar{q}_R \frac{\lambda_a}{2}M q_L -\bar{q}_L M \frac{\lambda_a}{2} q_R \right) +\Theta^R \left(\bar{q}_RMq_L-\bar{q}_LM q_R\right)\right.\nonumber\\ &&+\left.\sum_{a=1}^8 \Theta_a^L \left( \bar{q}_L \frac{\lambda_a}{2}M q_R -\bar{q}_R M \frac{\lambda_a}{2} q_L \right) +\Theta^L \left(\bar{q}_LMq_R-\bar{q}_RM q_L\right)\right],\nonumber\\\end{aligned}$$ which results in the following divergences, $$\begin{aligned} \label{2:3:dslr} \partial_\mu L^{\mu,a}&=&\frac{\partial \delta {\cal L}_M}{\partial \Theta^L_a} =-i\left(\bar{q}_L\frac{\lambda_a}{2}M q_R -\bar{q}_R M \frac{\lambda_a}{2} q_L\right),\nonumber\\ \partial_\mu R^{\mu,a}&=&\frac{\partial \delta {\cal L}_M}{\partial \Theta^R_a} =-i\left(\bar{q}_R\frac{\lambda_a}{2}M q_L -\bar{q}_L M \frac{\lambda_a}{2} q_R\right),\nonumber\\ \partial_\mu L^{\mu}&=&\frac{\partial \delta {\cal L}_M}{\partial \Theta^L} =-i\left(\bar{q}_L M q_R -\bar{q}_R M q_L\right),\nonumber\\ \partial_\mu R^{\mu}&=&\frac{\partial \delta {\cal L}_M}{\partial \Theta^R} =-i\left(\bar{q}_R M q_L -\bar{q}_L M q_R\right).\end{aligned}$$ The anomaly has not yet been considered. Applying Eq. (\[2:3:qgq\]) to the case of the vector currents and inserting projection operators as in the derivation of Eq. (\[2:3:sa\]) for the axial-vector current, the corresponding divergences read $$\begin{aligned} \label{2:3:dsva} \partial_\mu V^{\mu,a}&=& i\bar{q}[M,\frac{\lambda_a}{2}]q,\nonumber\\ \partial_\mu A^{\mu,a}&=& i\left(\bar{q}_L\{\frac{\lambda_a}{2},M\}q_R -\bar{q}_R\{\frac{\lambda_a}{2},M\}q_L\right) =i\bar{q}\{\frac{\lambda_a}{2},M\}\gamma_5q,\nonumber\\ \partial_\mu V^\mu&=&0,\nonumber\\ \partial_\mu A^\mu&=&2i\bar{q}M\gamma_5 q+ \frac{3 g^2}{32\pi^2}\epsilon_{\mu\nu\rho\sigma} {\cal G}^{\mu\nu}_a {\cal G}^{\rho\sigma}_a,\quad \epsilon_{0123}=1, \end{aligned}$$ where the axial anomaly has also been taken into account. We are now in the position to summarize the various (approximate) symmetries of the strong interactions in combination with the corresponding currents and their divergences. - In the limit of massless quarks, the sixteen currents $L^{\mu,a}$ and $R^{\mu,a}$ or, alternatively, $V^{\mu,a}$ and $A^{\mu,a}$ are conserved. The same is true for the singlet vector current $V^\mu$, whereas the singlet axial-vector current $A^\mu$ has an anomaly. - For any value of quark masses, the individual flavor currents $\bar{u}\gamma^\mu u$, $\bar{d}\gamma^\mu d$, and $\bar{s}\gamma^\mu s$ are always conserved in strong interactions reflecting the flavor independence of the strong coupling and the diagonality of the quark mass matrix. Of course, the singlet vector current $V^\mu$, being the sum of the three flavor currents, is always conserved. - In addition to the anomaly, the singlet axial-vector current has an explicit divergence due to the quark masses. - For equal quark masses, $m_u=m_d=m_s$, the eight vector currents $V^{\mu,a}$ are conserved, because $[\lambda_a,1]=0$. Such a scenario is the origin of the SU(3) symmetry originally proposed by Gell-Mann and Ne’eman [@EightfoldWay]. The eight axial currents $A^{\mu,a}$ are not conserved. The divergences of the octet axial-vector currents of Eq. (\[2:3:dsva\]) are proportional to pseudoscalar quadratic forms. This can be interpreted as the microscopic origin of the PCAC relation (partially conserved axial-vector current) which states that the divergences of the axial-vector currents are proportional to renormalized field operators representing the lowest lying pseudoscalar octet (for a comprehensive discussion of the meaning of PCAC see Refs. [@Gell-Mann:1964tf; @Adler:1968; @Treiman:1972; @DeAlfaro:1973]). - Various symmetry-breaking patterns are discussed in great detail in Ref. [@Pagels:se]. Green Functions and Chiral Ward Identities {#sec_gfcwi} ------------------------------------------ ### Chiral Green Functions {#subsec_cgf} For conserved currents, the spatial integrals of the charge densities are time independent, i.e., in a quantized theory the corresponding charge operators commute with the Hamilton operator. These operators are generators of infinitesimal transformations on the Hilbert space of the theory. The mass eigenstates should organize themselves in degenerate multiplets with dimensionalities corresponding to irreducible representations of the Lie group in question.[^23] Which irreducible representations ultimately appear, and what the actual energy eigenvalues are, is determined by the dynamics of the Hamiltonian. For example, SU(2) isospin symmetry of the strong interactions reflects itself in degenerate SU(2) multiplets such as the nucleon doublet, the pion triplet and so on. Ultimately, the actual masses of the nucleon and the pion should follow from QCD (for a prediction of hadron masses in lattice QCD see, e.g., Refs. [@Butler:1994em; @AliKhan:2001tx]). It is also well-known that symmetries imply relations between $S$-matrix elements. For example, applying the Wigner-Eckart theorem to pion-nucleon scattering, assuming the strong-interaction Hamiltonian to be an isoscalar, it is sufficient to consider two isospin amplitudes describing transitions between states of total isospin $I=1/2$ or $I=3/2$ (see, for example, [@Ericson:gk]). All the dynamical information is contained in these isospin amplitudes and the results for physical processes can be expressed in terms of these amplitudes together with geometrical coefficients, namely, the Clebsch-Gordan coefficients. In quantum field theory, the objects of interest are the Green functions which are vacuum expectation values of time-ordered products.[^24] Pictorially, these Green functions can be understood as vertices and are related to physical scattering amplitudes through the Lehmann-Symanzik-Zimmermann (LSZ) reduction formalism [@Lehmann:1955rq]. Symmetries provide strong constraints not only for scattering amplitudes, i.e. their transformation behavior, but, more generally speaking, also for Green functions and, in particular, [*among*]{} Green functions. The famous example in this context is, of course, the Ward identity of QED associated with U(1) gauge invariance [@Ward:1950xp], $$\label{2:4:qedwardidentity} \Gamma^\mu(p,p)=-\frac{\partial}{\partial p_\mu}\Sigma(p),$$ which relates the electromagnetic vertex of an electron at zero momentum transfer, $\gamma^\mu+\Gamma^\mu(p,p)$, to the electron self energy, $\Sigma(p)$. Such symmetry relations can be extended to non-vanishing momentum transfer and also to more complicated groups and are referred to as Ward-Fradkin-Takahashi identities [@Ward:1950xp; @Fradkin:1955jr; @Takahashi:xn] (or Ward identities for short). Furthermore, even if a symmetry is broken, i.e., the infinitesimal generators are time dependent, conditions related to the symmetry breaking terms can still be obtained using equal-time commutation relations [@Gell-Mann:xb]. At first, we are interested in time-ordered products of color-neutral, Hermitian quadratic forms involving the light quark fields evaluated between the vacuum of QCD. Using the LSZ reduction formalism [@Lehmann:1955rq; @Itzykson:rh] such Green functions can be related to physical processes involving mesons as well as their interactions with the electroweak gauge fields of the Standard Model. The interpretation depends on the transformation properties and quantum numbers of the quadratic forms, determining for which mesons they may serve as an interpolating field. In addition to the vector and axial-vector currents of Eqs. (\[2:3:v\]), (\[2:3:a\]), and (\[2:3:sv\]) we want to investigate scalar and pseudoscalar densities,[^25] $$\label{2:4:quadraticforms} S_a(x)=\bar{q}(x)\lambda_a q(x),\quad P_a(x)=i\bar{q}(x)\gamma_5 \lambda_a q(x),\quad a=0,\cdots, 8,$$ which enter, for example, in Eqs. (\[2:3:dsva\]) as the divergences of the vector and axial-vector currents for nonzero quark masses. Whenever it is more convenient, we will also use $$\label{2:4:SP} S(x)=\bar{q}(x) q(x),\quad P(x)=i\bar{q}(x)\gamma_5 q(x),$$ instead of $S_0$ and $P_0$. Later on, we will also consider similar time-ordered products evaluated between a single nucleon in the initial and final states in addition to the vacuum Green functions. This will allow us to discuss properties of the nucleon as well as dynamical processes involving a single nucleon. Generally speaking, a chiral Ward identity relates the divergence of a Green function containing at least one factor of $V^{\mu,a}$ or $A^{\mu,a}$ \[see Eqs. (\[2:3:v\]) and (\[2:3:a\])\] to some linear combination of other Green functions. The terminology [*chiral*]{} refers to the underlying $\mbox{SU(3)}_L\times \mbox{SU(3)}_R$ group. To make this statement more precise, let us consider as a simple example the two-point Green function involving an axial-vector current and a pseudoscalar density,[^26] $$\begin{aligned} \label{2:4:gfaav} G^{\mu,ab}_{AP}(x,y)&=&\langle 0| T[A^\mu_a(x) P_b(y)]|0\rangle\nonumber\\ &=&\Theta(x_0-y_0)\langle 0|A^\mu_a(x) P_b(y)|0\rangle +\Theta(y_0-x_0)\langle 0|P_b(y) A^\mu_a(x)|0\rangle,\nonumber\\\end{aligned}$$ and evaluate the divergence $$\begin{aligned} \lefteqn{\partial_\mu^x G^{\mu,ab}_{AP}(x,y)}\\ &=& \delta(x_0-y_0)\langle 0| A_0^a(x) P_b(y)|0\rangle -\delta(x_0-y_0)\langle 0| P_b(y)A_0^a (x)|0\rangle\\ && +\Theta(x_0-y_0)\langle 0|\partial_\mu^x A^\mu_a(x) P_b(y)|0\rangle +\Theta(y_0-x_0)\langle 0| P_b(y)\partial_\mu^x A^\mu_a(x)|0\rangle\\ &=&\delta(x_0-y_0)\langle 0|[A^a_0(x),P_b(y)]|0\rangle +\langle 0|T[\partial_\mu^x A^\mu_a(x) P_b(y)]|0\rangle,\end{aligned}$$ where we made use of $\partial_\mu^x \Theta(x_0-y_0)=\delta(x_0-y_0) g_{0\mu}=-\partial_\mu^x \Theta(y_0-x_0)$. This simple example already shows the main features of (chiral) Ward identities. From the differentiation of the theta functions one obtains equal-time commutators between a charge density and the remaining quadratic forms. The results of such commutators are a reflection of the underlying symmetry, as will be shown below. As a second term, one obtains the divergence of the current operator in question. If the symmetry is perfect, such terms vanish identically. For example, this is always true for the electromagnetic case with its U(1) symmetry. If the symmetry is only approximate, an additional term involving the symmetry breaking appears. For a soft breaking such a divergence can be treated as a perturbation. Via induction, the generalization of the above simple example to an -point Green function is symbolically of the form $$\begin{aligned} \label{2:4:gendmug} \lefteqn{\partial_\mu^x \langle 0|T\{J^\mu(x) A_1(x_1)\cdots A_n(x_n) \}|0\rangle=}\nonumber\\ &&\langle 0|T\{[\partial_\mu^x J^\mu(x)] A_1(x_1)\cdots A_n(x_n)\}|0\rangle\nonumber\\ &&+\delta(x^0-x_1^0)\langle 0|T\{[J_0(x),A_1(x_1)] A_2(x_2)\cdots A_n(x_n)\}| 0\rangle\nonumber\\ &&+\delta(x^0-x_2^0)\langle 0|T\{A_1(x_1)[J_0(x),A_2(x_2)] \cdots A_n(x_n)\}|0\rangle\nonumber\\ &&+\cdots+\delta(x^0-x_n^0) \langle 0|T\{A_1(x_1)\cdots [J_0(x),A_n(x_n)]\}|0\rangle,\end{aligned}$$ where $J^\mu$ stands generically for any of the Noether currents. ### The Algebra of Currents {#subsec_ac} In the above example, we have seen that chiral Ward identities depend on the equal-time commutation relations of the [*charge densities*]{} of the symmetry currents with the relevant quadratic quark forms. Unfortunately, a naive application of Eq. (\[2:3:fkf\]) may lead to erroneous results. Let us illustrate this by means of a simplified example, the equal-time commutator of the time and space components of the ordinary electromagnetic current in QED. A naive use of the canonical commutation relations leads to $$\begin{aligned} \label{2:4:schwinger} [J_0(\vec{x},t), J_i(\vec{y},t)]&=& [\Psi^\dagger(\vec{x},t)\Psi(\vec{x},t),\Psi^\dagger(\vec{y},t)\gamma_0\gamma_i \Psi(\vec{y},t)]\nonumber\\ &=&\delta^3(\vec{x}-\vec{y})\Psi^\dagger(\vec{x},t)[1,\gamma_0\gamma_i] \Psi(\vec{x},t)=0,\end{aligned}$$ where we made use of the delta function to evaluate the fields at $\vec{x}=\vec{y}$. It was noticed a long time ago by Schwinger that this result cannot be true [@Schwinger:xd]. In order to see this, consider the commutator $$[J_0(\vec{x},t),\vec{\nabla}_y\cdot \vec{J}(\vec{y},t)]= -[J_0(\vec{x},t),\partial_t J_0(\vec{y},t)],$$ where we made use of current conservation, $\partial_\mu J^\mu=0$. If Eq. (\[2:4:schwinger\]) were true, one would necessarily also have $$0=[J_0(\vec{x},t),\partial_t J_0(\vec{y},t)],$$ which we evaluate for $\vec{x}=\vec{y}$ between the ground state, $$\begin{aligned} 0&=&\langle 0|[J_0(\vec{x},t),\partial_t J_0(\vec{x},t)]|0\rangle\\ &=&\sum_n \Big(\langle 0|J_0(\vec{x},t)|n\rangle\langle n| \partial_t J_0(\vec{x},t)|0\rangle -\langle 0|\partial_t J_0(\vec{x},t)|n\rangle\langle n | J_0(\vec{x},t)|0\rangle\Big)\\ &=&2i\sum_n(E_n-E_0)|\langle 0|J_0(\vec{x},t)|n\rangle|^2.\end{aligned}$$ Here, we inserted a complete set of states and made use of $$\partial_t J_0(\vec{x},t)=i[H,J_0(\vec{x},t)].$$ Since every individual term in the sum is non-negative, one would need $\langle 0|J_0(\vec{x},t)|n\rangle=0$ for any intermediate state which is obviously unphysical. The solution is that the starting point, Eq. (\[2:4:schwinger\]), is not true. The corrected version of Eq. (\[2:4:schwinger\]) picks up an additional, so-called Schwinger term containing a derivative of the delta function. Quite generally, by evaluating commutation relations with the component $\Theta^{00}$ of the energy-momentum tensor one can show that the equal-time commutation relation between a charge density and a current density can be determined up to one derivative of the $\delta$ function [@Jackiw:1972], $$\label{2:4:j0jigen} [J_0^a(\vec{x},0),J_i^b(\vec{y},0)]=iC_{abc} J_i^c(\vec{x},0)\delta^3 (\vec{x}-\vec{y})+S_{ij}^{ab}(\vec{y},0)\partial^j\delta^3(\vec{x}-\vec{y}),$$ where the Schwinger term possesses the symmetry $$S_{ij}^{ab}(\vec{y},0)=S_{ji}^{ba}(\vec{y},0),$$ and $C_{abc}$ denote the structure constants of the group in question. However, in our above derivation of the chiral Ward identity, we also made use of the [*naive*]{} time-ordered product ($T$) as opposed to the [*covariant*]{} one ($T^\ast$) which, typically, differ by another non-covariant term which is called a seagull. Feynman’s conjecture [@Jackiw:1972] states that there is a cancelation between Schwinger terms and seagull terms such that a Ward identity obtained by using the naive T product and by simultaneously omitting Schwinger terms ultimately yields the correct result to be satisfied by the Green function (involving the covariant $T^\ast$ product). Although this will not be true in general, a sufficient condition for it to happen is that the time component algebra of the full theory remains the same as the one derived canonically and does not possess a Schwinger term. For a detailed discussion, the interested reader is referred to Ref. [@Jackiw:1972]. Keeping the above discussion in mind, the complete list of equal-time commutation relations, omitting Schwinger terms, reads $$\begin{aligned} \label{2:4:letcr} [V^a_0(\vec{x},t),V^\mu_b(\vec{y},t)] &=&\delta^3(\vec{x}-\vec{y})if_{abc} V^\mu_c(\vec{x},t),\nonumber\\ {[}V^a_0(\vec{x},t),V^\mu(\vec{y},t)]&=&0,\nonumber\\ {[}V^a_0(\vec{x},t),A^\mu_b(\vec{y},t)] &=&\delta^3(\vec{x}-\vec{y})if_{abc} A^\mu_c(\vec{x},t),\nonumber\\ {[}V^a_0(\vec{x},t),S_b(\vec{y},t)] &=&\delta^3(\vec{x}-\vec{y})if_{abc} S_c(\vec{x},t),\quad b=1,\cdots,8, \nonumber\\ {[}V^a_0(\vec{x},t),S_0(\vec{y},t)]&=&0,\nonumber\\ {[}V^a_0(\vec{x},t),P_b(\vec{y},t)] &=&\delta^3(\vec{x}-\vec{y})if_{abc} P_c(\vec{x},t),\quad b=1,\cdots,8, \nonumber\\ {[}V^a_0(\vec{x},t),P_0(\vec{y},t)]&=&0,\nonumber\\ {[}A^a_0(\vec{x},t),V^\mu_b(\vec{y},t)] &=&\delta^3(\vec{x}-\vec{y})if_{abc} A^\mu_c(\vec{x},t),\nonumber\\ {[}A^a_0(\vec{x},t),V^\mu(\vec{y},t)]&=&0,\nonumber\\ {[}A^a_0(\vec{x},t),A^\mu_b(\vec{y},t)] &=&\delta^3(\vec{x}-\vec{y})if_{abc} V^\mu_c(\vec{x},t),\nonumber\\ {[}A^a_0(\vec{x},t),S_b(\vec{y},t)] &=&\delta^3(\vec{x}-\vec{y})if_{abc} P_c(\vec{x},t),\quad b=1,\cdots,8, \nonumber\\ {[}A^a_0(\vec{x},t),S_0(\vec{y},t)]&=&0,\nonumber\\ {[}A^a_0(\vec{x},t),P_b(\vec{y},t)] &=&\delta^3(\vec{x}-\vec{y})if_{abc} S_c(\vec{x},t),\quad b=1,\cdots,8, \nonumber\\ {[}A^a_0(\vec{x},t),P_0(\vec{y},t)]&=&0.\end{aligned}$$ ### Two Simple Examples {#subsec_tse} We now return to our specific example, namely, the divergence of Eq. (\[2:4:gfaav\]). Inserting the results of Eqs. (\[2:3:dsva\]) and (\[2:4:letcr\]) one obtains $$\begin{aligned} \label{2:4:dgfaav} \partial_\mu^x G^{\mu,ab}_{AP}(x,y) &=&\delta^4(x-y) i f_{abc}\langle 0|S_c(x)|0\rangle\nonumber\\ &&+i\langle 0|T[\bar{q}(x)\{\frac{\lambda_a}{2},M\}\gamma_5 q(x) P_b(y)]|0\rangle.\end{aligned}$$ The second term on the right-hand side of Eq. (\[2:4:dgfaav\]) can be re-expressed using Eq. (\[2:3:qmm\]) and the anti-commutation relations of Eq. (\[2:1:acrgmm\]) in combination with the $d$ coefficients of Table \[table:2:1:su3dsymbols\] (no summation over $a$ implied), $$\begin{aligned} \lefteqn{i\bar{q}(x)\{\frac{\lambda_a}{2},M\}\gamma_5q(x)=}\\ &&\left[\frac{1}{3}(m_u+m_d+m_s) +\frac{1}{\sqrt{3}}\left(\frac{m_u+m_d}{2}- m_s\right)d_{aa8}\right]P_a(x)\\ &&+\left[\sqrt{\frac{1}{6}}(m_u-m_d)\delta_{a3} +\frac{\sqrt{2}}{3}\left(\frac{m_u+m_d}{2}-m_s\right)\delta_{a8}\right]P_0(x)\\ &&+\frac{m_u-m_d}{2} \sum_{c=1}^8d_{a3c} P_c(x).\end{aligned}$$ Equation (\[2:4:dgfaav\]) serves to illustrate two distinct features of chiral Ward identities. The first term of Eq. (\[2:4:dgfaav\]) originates in the algebra of currents and thus represents a consequence of the [*transformation properties*]{} of the quadratic quark forms entering the Green function. In general, depending on whether the appropriate equal-time commutation relation of Eq. (\[2:4:letcr\]) vanishes or not, the resulting term in the divergence of an $n$-point Green function vanishes or is proportional to an $(n-1)$-point Green function. In our specific example, the divergence of the Green function involving the axial-vector current and the pseudoscalar density is related to the so-called scalar quark condensate which will be discussed in more detail in Sec. \[subsec\_sqc\]. The second term of Eq. (\[2:4:dgfaav\]) is due to an explicit symmetry breaking resulting from the quark masses. This shows the second property of chiral Ward identities, namely, symmetry breaking terms give rise to another $n$-point Green function. To summarize, chiral Ward identities incorporate both transformation properties of quadratic quark forms as well as symmetry breaking patterns. As another well-known and simple example, let us briefly consider, for the two-flavor case, the nucleon matrix element of the axial-vector current operator[^27] $$\label{2:4:avcn} \langle N(p_f)|A^i_\mu(x)|N(p_i)\rangle= \langle N(p_f)|\bar{q}(x)\gamma_\mu\gamma_5 \frac{\tau_i}{2}q(x)|N(p_i)\rangle.$$ This matrix element serves as an illustration of chiral Ward identities which are taken between one-nucleon states instead of the vacuum. According to Eq. (\[2:3:dsva\]), the divergence of Eq. (\[2:4:avcn\]) is related to the pseudoscalar density evaluated between one-nucleon states. Of course, in the chiral limit $M=0$ and the axial-vector current is conserved. ### QCD in the Presence of External Fields and the Generating Functional {#subsec_qcdpefgf} Here, we want to consider the consequences of Eqs. (\[2:4:letcr\]) for the Green functions of QCD (in particular, at low energies). In principle, using the techniques of the last section, for each Green function one can [*explicitly*]{} work out the chiral Ward identity which, however, becomes more and more tedious as the number $n$ of quark quadratic forms increases. However, there exists an elegant way of formally combining all Green functions in a generating functional. The (infinite) set of [*all*]{} chiral Ward identities is encoded as an invariance property of that functional. To see this, one has to consider a coupling to external c-number fields such that through functional methods one can, in principle, obtain all Green functions from a generating functional. The rationale behind this approach is that, in the absence of anomalies, the Ward identities obeyed by the Green functions are equivalent to an invariance of the generating functional under a [*local*]{} transformation of the external fields [@Leutwyler:1993iq]. The use of local transformations allows one to also consider divergences of Green functions. For an illustration of this statement, the reader is referred to Appendix A. Following the procedure of Gasser and Leutwyler [@Gasser:1983yg; @Gasser:1984gg], we introduce into the Lagrangian of QCD the couplings of the nine vector currents and the eight axial-vector currents as well as the scalar and pseudoscalar quark densities to external c-number fields $v^\mu (x)$, $v^\mu_{(s)}$, $a^\mu (x)$, $s(x)$, and $p(x)$, $$\label{2:4:lqcds} {\cal L}={\cal L}^0_{\rm QCD}+{\cal L}_{\rm ext} ={\cal L}^0_{\rm QCD}+\bar{q}\gamma_\mu (v^\mu +\frac{1}{3}v^\mu_{(s)} +\gamma_5 a^\mu )q -\bar{q}(s-i\gamma_5 p)q.$$ The external fields are color-neutral, Hermitian $3\times 3$ matrices, where the matrix character, with respect to the (suppressed) flavor indices $u$, $d$, and $s$ of the quark fields, is[^28] $$\label{2:4:mch} v^\mu=\sum_{a=1}^8\frac{\lambda_a}{2}v_a^\mu,\quad a^\mu=\sum_{a=1}^8\frac{\lambda_a}{2}a_a^\mu,\quad s=\sum_{a=0}^8 \lambda_a s_a,\quad p=\sum_{a=0}^8\lambda_a p_a.$$ The ordinary three flavor QCD Lagrangian is recovered by setting $v^\mu=v^\mu_{(s)}=a^\mu=p=0$ and $s=\mbox{diag}(m_u,m_d,m_s)$ in Eq. (\[2:4:lqcds\]). If one defines the generating functional[^29] $$\label{2:4:genfun} \exp(i Z[v,a,s,p])=\langle 0|T\exp\left[i\int d^4 x {\cal L}_{\rm ext}(x)\right]|0\rangle,$$ then any Green function consisting of the time-ordered product of color-neutral, Hermitian quadratic forms can be obtained from Eq. (\[2:4:genfun\]) through a functional derivative with respect to the external fields. The quark fields are operators in the Heisenberg picture and have to satisfy the equation of motion and the canonical anti-commutation relations. The actual value of the generating functional for a given configuration of external fields $v$, $a$, $s$, and $p$ reflects the dynamics generated by the QCD Lagrangian. The generating functional is related to the vacuum-to-vacuum transition amplitude in the presence of external fields [@Gasser:1983yg; @Gasser:1984gg], $$\label{2:4:genfunvv} \exp[i Z(v,a,s,p)]= \langle 0_{\rm out}|0_{\rm in}\rangle_{v,a,s,p},$$ where the dynamics is determined by the Lagrangian of Eq. (\[2:4:lqcds\]). For example,[^30] the $\bar{u} u$ component of the scalar quark condensate in the chiral limit, $\langle 0| \bar{u}u|0\rangle_0$, is given by $$\begin{aligned} \label{2:4:sqc} \lefteqn{\langle 0|\bar{u}(x) u(x)|0\rangle_0 =}\nonumber\\ &&\left.\frac{i}{2}\left[\sqrt{\frac{2}{3}}\frac{\delta}{\delta s_0(x)} +\frac{\delta}{\delta s_3(x)} +\frac{1}{\sqrt{3}} \frac{\delta}{\delta s_8(x)}\right] \exp(iZ [v,a,s,p])\right|_{v=a=s=p=0},\nonumber\\\end{aligned}$$ where we made use of Eq. (\[2:1:matrixa\]). Note that both the quark field operators and the ground state are considered in the chiral limit, which is denoted by the subscript 0. As another example, let us consider the two-point function of the axial-vector currents of Eq. (\[2:3:a\]) of the “real world,” i.e., for $s=\mbox{diag}(m_u,m_d,m_s)$, and the “true vacuum” $|0\rangle$, $$\begin{aligned} \label{2:4:tpfavc} \lefteqn{\langle 0|T[A_\mu^a(x) A_\nu^b(0)]|0\rangle =}\nonumber\\ &&\left. (-i)^2 \frac{\delta^2}{\delta a^\mu_a(x)\delta a^\nu_b(0)} \exp(iZ[v,a,s,p])\right|_{v=a=p=0,s=\mbox{diag}(m_u,m_d,m_s)}.\nonumber\\\end{aligned}$$ Requiring the total Lagrangian of Eq. (\[2:4:lqcds\]) to be Hermitian and invariant under $P$, $C$, and $T$ leads to constraints on the transformation behavior of the external fields. In fact, it is sufficient to consider $P$ and $C$, only, because $T$ is then automatically incorporated owing to the $CPT$ theorem. Under parity, the quark fields transform as $$\label{2:4:qtrafop} q_f(\vec{x},t)\stackrel{\mbox{$P$}}{\mapsto}\gamma^0 q_f(-\vec{x},t),$$ and the requirement of parity conservation, $$\label{2:4:parinv} {\cal L}(\vec{x},t) \stackrel{\mbox{$P$}}{\mapsto} {\cal L}(-\vec{x},t),$$ leads, using the results of Table \[2:4:parity\], to the following constraints for the external fields, $$\label{2:4:eftrafop} v^\mu\stackrel{\mbox{$P$}}{\mapsto}v_\mu,\quad v^\mu_{(s)}\stackrel{\mbox{$P$}}{\mapsto}v_\mu^{(s)},\quad a^\mu\stackrel{\mbox{$P$}}{\mapsto}-a_\mu,\quad s\stackrel{\mbox{$P$}}{\mapsto}s,\quad p\stackrel{\mbox{$P$}}{\mapsto}-p.$$ In Eq. (\[2:4:eftrafop\]) it is understood that the arguments change from $(\vec{x},t)$ to $(-\vec{x},t)$. $\Gamma$ $1$ $\gamma^\mu$ $\sigma^{\mu\nu}$ $\gamma_5$ $\gamma^\mu\gamma_5$ ---------------------------- ----- -------------- ------------------- ------------- ----------------------- $\gamma_0 \Gamma \gamma_0$ $1$ $\gamma_\mu$ $\sigma_{\mu\nu}$ $-\gamma_5$ $-\gamma_\mu\gamma_5$ : \[2:4:parity\] Transformation properties of the Dirac matrices $\Gamma$ under parity. Similarly, under charge conjugation the quark fields transform as $$\label{2:4:qtrafc} q_{\alpha,f}\stackrel{\mbox{$C$}}{\mapsto}C_{\alpha\beta}\bar{q}_{\beta,f}, \quad \bar{q}_{\alpha,f}\stackrel{\mbox{$C$}}{\mapsto} -q_{\beta,f}C^{-1}_{\beta\alpha},$$ where the subscripts $\alpha$ and $\beta$ are Dirac spinor indices, $C=i\gamma^2\gamma^0=-C^{-1}=-C^\dagger=-C^T$ is the usual charge conjugation matrix in the convention of Ref. [@Bjorken_1964] and $f$ refers to flavor. Using Eq. (\[2:4:qtrafc\]) in combination with Table \[2:4:chargeconjugation\] it is straightforward to show that invariance of ${\cal L}_{\rm ext}$ under charge conjugation requires the transformation properties[^31] $$\label{2:4:eftrafoc} v_\mu\stackrel{C}{\rightarrow}-v_\mu^T,\quad v_\mu^{(s)}\stackrel{C}{\rightarrow}-v_\mu^{(s)T},\quad a_\mu\stackrel{C}{\rightarrow}a_\mu^T,\quad s,p\stackrel{C}{\rightarrow}s^T,p^T,$$ where the transposition refers to the flavor space. --------------------------------------------------------------------------------------------- $\Gamma$ $1$ $\gamma^\mu$ $\sigma^{\mu\nu}$ $\gamma_5$ $\gamma^\mu\gamma_5$ -------------- ----- --------------- -------------------- ------------ ---------------------- $-C\Gamma^TC $1$ $-\gamma^\mu$ $-\sigma^{\mu\nu}$ $\gamma_5$ $\gamma^\mu\gamma_5$ $ --------------------------------------------------------------------------------------------- : \[2:4:chargeconjugation\] Transformation properties of the Dirac matrices $\Gamma$ under charge conjugation. Finally, we need to discuss the requirements to be met by the external fields under local $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U}(1)_V$ transformations. In a first step, we write Eq. (\[2:4:lqcds\]) in terms of the left- and right-handed quark fields. Besides the properties of Eqs. (\[2:3:prplcompleteness\]) - (\[2:3:prplorthogonality\]) we make use of the auxiliary formulae $$\gamma_5 P_R=P_R\gamma_5=P_R, \quad \gamma_5 P_L=P_L\gamma_5=-P_L,$$ and $$\gamma^\mu P_R=P_L\gamma^\mu,\quad \gamma^\mu P_L=P_R\gamma^\mu,$$ to obtain $$\begin{aligned} \bar{q}\gamma^\mu(v_\mu+\frac{1}{3}v_\mu^{(s)} +\gamma_5 a_\mu)q&=&\frac{1}{2}\bar{q} \gamma^\mu[r_\mu+l_\mu+\frac{2}{3}v_\mu^{(s)}+\gamma_5(r_\mu-l_\mu)] q\\ &=&\bar{q}_R\gamma^\mu \left(r_\mu +\frac{1}{3}v_\mu^{(s)}\right)q_R +\bar{q}_L\gamma^\mu \left(l_\mu+\frac{1}{3}v_\mu^{(s)}\right) q_L,\end{aligned}$$ where $$\label{2:4:vrlarl} v_\mu=\frac{1}{2}(r_\mu+l_\mu),\quad a_\mu=\frac{1}{2}(r_\mu-l_\mu).$$ Similarly, we rewrite the second part containing the external scalar and pseudoscalar fields, $$\begin{aligned} \bar{q}(s-i\gamma_5 p)q&=& \bar{q}_L(s-ip)q_R+\bar{q}_R(s+ip)q_L,\end{aligned}$$ yielding for the Lagrangian of Eq. (\[2:4:lqcds\]) $$\begin{aligned} \label{2:4:lqcdsn} {\cal L}&=&{\cal L}_{\rm QCD}^0 +\bar{q}_L\gamma^\mu\left(l_\mu+\frac{1}{3}v^{(s)}_\mu\right)q_L +\bar{q}_R\gamma^\mu\left(r_\mu+\frac{1}{3}v^{(s)}_\mu\right)q_R\nonumber\\ &&-\bar{q}_R(s+ip)q_L-\bar{q}_L(s-ip)q_R.\end{aligned}$$ Equation (\[2:4:lqcdsn\]) remains invariant under [*local*]{} transformations $$\begin{aligned} \label{2:4:qrl} q_R&\mapsto&\exp\left(-i\frac{\Theta(x)}{3}\right) V_R(x) q_R,\nonumber\\ q_L&\mapsto&\exp\left(-i\frac{\Theta(x)}{3}\right) V_L(x) q_L,\end{aligned}$$ where $V_R(x)$ and $V_L(x)$ are independent space-time-dependent SU(3) matrices, provided the external fields are subject to the transformations $$\begin{aligned} \label{2:4:sg} r_\mu&\mapsto& V_R r_\mu V_R^{\dagger} +iV_R\partial_\mu V_R^{\dagger},\nonumber\\ l_\mu&\mapsto& V_L l_\mu V_L^{\dagger} +iV_L\partial_\mu V_L^{\dagger}, \nonumber\\ v_\mu^{(s)}&\mapsto&v_\mu^{(s)}-\partial_\mu\Theta,\nonumber\\ s+ip&\mapsto& V_R(s+ip)V_L^{\dagger},\nonumber\\ s-ip&\mapsto& V_L(s-ip)V_R^{\dagger}.\end{aligned}$$ The derivative terms in Eq. (\[2:4:sg\]) serve the same purpose as in the construction of gauge theories, i.e., they cancel analogous terms originating from the kinetic part of the quark Lagrangian. There is another, yet, more practical aspect of the local invariance, namely: such a procedure allows one to also discuss a coupling to external gauge fields in the transition to the effective theory to be discussed later. For example, we have seen in Sec. 2.2 that a coupling of the electromagnetic field to point-like fundamental particles results from gauging a U(1) symmetry. Here, the corresponding U(1) group is to be understood as a subgroup of a local $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$. Another example deals with the interaction of the light quarks with the charged and neutral gauge bosons of the weak interactions. Let us consider both examples explicitly. The coupling of quarks to an external electromagnetic field ${\cal A}_\mu$ is given by $$\label{2:4:rla} r_\mu=l_\mu=-e Q {\cal A}_\mu,$$ where $Q=\mbox{diag}(2/3,-1/3,-1/3)$ is the quark charge matrix: $$\begin{aligned} {\cal L}_{\rm ext}&=&-e {\cal A}_\mu(\bar{q}_L Q\gamma^\mu q_L +\bar{q}_R Q \gamma^\mu q_R)\\ &=&-e {\cal A}_\mu \bar{q}Q\gamma^\mu q\\ &=&-e {\cal A}_\mu\left(\frac{2}{3}\bar{u}\gamma^\mu u -\frac{1}{3} \bar{d}\gamma^\mu d -\frac{1}{3}\bar{s}\gamma^\mu s\right)\\ &=&-e {\cal A}_\mu J^\mu.\end{aligned}$$ On the other hand, if one considers only the two-flavor version of ChPT one has to insert for the external fields $$\label{2:4:rlasu2} r_\mu=l_\mu=-e\frac{\tau_3}{2}{\cal A}_\mu,\quad v_\mu^{(s)}=-\frac{e}{2}{\cal A}_\mu.$$ In the description of semileptonic interactions such as $\pi^-\to \mu^-\bar{\nu}_\mu$, $\pi^-\to\pi^0e^-\bar{\nu}_e$, or neutron decay $n\to p e^-\bar{\nu}_e$ one needs the interaction of quarks with the massive charged weak bosons ${\cal W}^\pm_\mu=({\cal W}_{1\mu}\mp i {\cal W}_{2\mu})/\sqrt{2}$, $$\label{2:4:rlw} r_\mu=0,\quad l_\mu=-\frac{g}{\sqrt{2}} ({\cal W}^+_\mu T_+ + h.c.),$$ where $h.c.$ refers to the Hermitian conjugate and $$T_+=\left(\begin{array}{rrr}0&V_{ud}&V_{us}\\0&0&0\\0&0&0\end{array}\right).$$ Here, $V_{ij}$ denote the elements of the Cabibbo-Kobayashi-Maskawa quark-mixing matrix describing the transformation between the mass eigenstates of QCD and the weak eigenstates [@Groom:in], $$|V_{ud}|=0.9735\pm 0.0008,\quad |V_{us}|=0.2196\pm 0.0023.$$ At lowest order in perturbation theory, the Fermi constant is related to the gauge coupling $g$ and the $W$ mass as $$G_F=\sqrt{2} \frac{g^2}{8 M^2_W}=1.16639(1)\times 10^{-5}\,\mbox{GeV}^{-2}.$$ Making use of $$\begin{aligned} \bar{q}_L\gamma^\mu {\cal W}_\mu^+ T_+ q_L&=& {\cal W}_\mu^+ (\bar{u}\,\,\bar{d}\,\, \bar{s}) P_R\gamma^\mu \left(\begin{array}{rrr}0&V_{ud}&V_{us}\\0&0&0\\0&0&0\end{array} \right) P_L \left(\begin{array}{c}u\\ d\\ s\end{array}\right)\\ &=&{\cal W}_\mu^ +(\bar{u}\,\,\bar{d}\,\,\bar{s})\gamma^\mu \frac{1}{2}(1-\gamma_5) \left(\begin{array}{c}V_{ud} d+ V_{us} s\\0\\0\end{array} \right)\\ &=&\frac{1}{2}{\cal W}_\mu^+[V_{ud}\bar{u}\gamma^\mu(1-\gamma_5)d +V_{us}\bar{u}\gamma^\mu(1-\gamma_5)s],\end{aligned}$$ we see that inserting Eq. (\[2:4:rlw\]) into Eq. (\[2:4:lqcdsn\]) leads to the standard charged-current weak interaction in the light quark sector, $$\begin{aligned} {\cal L}_{\rm ext}&=&-\frac{g}{2\sqrt{2}}\left\{{\cal W}^+_\mu[ V_{ud}\bar{u}\gamma^\mu(1-\gamma_5)d+V_{us}\bar{u}\gamma^\mu(1-\gamma_5)s] +h.c.\right\}.\end{aligned}$$ The situation is slightly different for the neutral weak interaction. Here, the SU(3) version requires a coupling to the singlet axial-vector current which, because of the anomaly of Eq. (\[2:3:divsa\]), we have dropped from our discussion. On the other hand, in the SU(2) version the axial-vector current part is traceless and we have $$\begin{aligned} \label{2:4:rlz} r_\mu&=&e \tan(\theta_W) \frac{\tau_3}{2} {\cal Z}_\mu,\nonumber\\ l_\mu&=&-\frac{g}{\cos(\theta_W)}\frac{\tau_3}{2} {\cal Z}_\mu+ e \tan(\theta_W) \frac{\tau_3}{2} {\cal Z}_\mu, \nonumber\\ v_\mu^{(s)}&=&\frac{e\tan(\theta_W)}{2}{\cal Z}_\mu,\end{aligned}$$ where $\theta_W$ is the weak angle. With these external fields, we obtain the standard weak neutral-current interaction [@Groom:in] $$\begin{aligned} {\cal L}_{\rm ext}&=&-\frac{g}{2\cos(\theta_W)}{\cal Z}_\mu\left( \bar{u}\gamma^\mu\left\{\left[\frac{1}{2}-\frac{4}{3}\sin^2(\theta_W)\right] -\frac{1}{2}\gamma_5\right\}u\right.\nonumber\\ &&\left.+\bar{d}\gamma^\mu\left\{\left[-\frac{1}{2} +\frac{2}{3}\sin^2(\theta_W)\right] +\frac{1}{2}\gamma_5\right\}d\right),\end{aligned}$$ where we made use of $e=g\sin(\theta_W)$. ### PCAC in the Presence of an External Electromagnetic Field {#subsec_pcacpeef} Finally, the technique of coupling the QCD Lagrangian to external fields also allows us to determine the current divergences for rigid external fields, i.e., which are [*not*]{} simultaneously transformed. For the sake of simplicity we restrict ourselves to the SU(2) sector. (The generalization to the SU(3) case is straightforward.) If the external fields are not simultaneously transformed and one considers a [*global*]{} chiral transformation only, the divergences of the currents read \[see Eq. (\[2:3:divergenz\])\] $$\begin{aligned} \label{2:4:divv} \partial_\mu V^\mu_i&=&i\bar{q}\gamma^\mu[\frac{\tau_i}{2},v_\mu]q +i\bar{q}\gamma^\mu\gamma_5[\frac{\tau_i}{2},a_\mu]q -i\bar{q}[\frac{\tau_i}{2},s]q-\bar{q}\gamma_5[\frac{\tau_i}{2},p]q, \nonumber\\ \label{2:4:diva}\\ \partial_\mu A^\mu_i&=&i\bar{q}\gamma^\mu\gamma_5[\frac{\tau_i}{2},v_\mu]q +i\bar{q}\gamma^\mu[\frac{\tau_i}{2},a_\mu]q +i\bar{q}\gamma_5\{\frac{\tau_i}{2},s\}q +\bar{q}\{\frac{\tau_i}{2},p\}q.\nonumber\\\end{aligned}$$ As an example, let us consider the QCD Lagrangian for a finite light quark mass $m_q$ in combination with a coupling to an external electromagnetic field ${\cal A}_\mu$ \[see Eq. (\[2:4:rlasu2\]), $a_\mu=0=p$\]. In this case the expressions for the divergence of the vector and axial-vector currents, respectively, read $$\begin{aligned} \label{2:4:divvsc} \partial_\mu V^\mu_i&=&-\epsilon_{3ij}e{\cal A}_\mu \bar{q}\gamma^\mu \frac{\tau_j}{2}q=-\epsilon_{3ij}e{\cal A}_\mu V^\mu_j,\\ \label{2:4:divasc} \partial_\mu A^\mu_i &=&m_q P_i-e {\cal A}_\mu \epsilon_{3ij} A^\mu_j +\delta_{i3} \frac{e^2 N_c}{96\pi^2}\epsilon_{\mu\nu\rho\sigma}{\cal F}^{\mu\nu} {\cal F}^{\rho\sigma},\end{aligned}$$ where we have introduced the isovector pseudoscalar density $$\label{2:4:psd} P_i=i\bar{q}\gamma_5 \tau_i q,$$ and ${\cal F}_{\mu\nu}=\partial_\mu{\cal A}_\nu-\partial_\nu{\cal A}_\mu$ is the electromagnetic field strength tensor. The third component of the axial-vector current, $A^\mu_3$, has an anomaly [@Adler:1969gk; @Adler:1969er; @Bardeen:1969md; @Bell:1969ts; @Adler:1970] which is related to the decay $\pi^0\to\gamma\gamma$. We emphasize the formal similarity of Eq. (\[2:4:divasc\]) to the (pre-QCD) PCAC relation obtained by Adler through the inclusion of the electromagnetic interactions with minimal electromagnetic coupling (see the Appendix of Ref. [@Adler:1965]).[^32] Since in QCD the quarks are taken as truly elementary, their interaction with an (external) electromagnetic field is of such a minimal type. Spontaneous Symmetry Breaking and the Goldstone Theorem {#chap_ssbgt} ======================================================= So far we have concentrated on the chiral symmetry of the QCD Hamiltonian and the [*explicit*]{} symmetry breaking through the quark masses. We have discussed the importance of chiral symmetry for the properties of Green functions with particular emphasis on the relations [*among*]{} different Green functions as expressed through the chiral Ward identities. Now it is time to address a second aspect which, for the low-energy structure of QCD, is equally important, namely, the concept of [*spontaneous*]{} symmetry breaking. A (continuous) symmetry is said to be spontaneously broken or hidden, if the ground state of the system is no longer invariant under the full symmetry group of the Hamiltonian. In this chapter we will first illustrate this by means of a discrete symmetry and then turn to the case of a spontaneously broken continuous global symmetry. Degenerate Ground States {#sec_dgs} ------------------------ Before discussing the case of a [*continuous*]{} symmetry, we will first have a look at a field theory with a [*discrete*]{} internal symmetry. This will allow us to distinguish between two possibilities: a dynamical system with a unique ground state or a system with a finite number of distinct degenerate ground states. In particular, we will see how, for the second case, an infinitesimal perturbation selects a particular vacuum state. To that end we consider the Lagrangian of a real scalar field $\Phi(x)$ [@Georgi] $$\label{3:1:lphi} {\cal L}(\Phi,\partial_\mu\Phi)=\frac{1}{2}\partial_\mu \Phi \partial^\mu \Phi -\frac{m^2}{2}\Phi^2-\frac{\lambda}{4}\Phi^4,$$ which is invariant under the discrete transformation $R: \Phi\to -\Phi$. The corresponding classical energy density reads $$\label{3:1:ked} {\cal H}=\Pi\dot{\Phi}-{\cal L}=\frac{1}{2}\dot{\Phi}^2 +\frac{1}{2}(\vec{\nabla}\Phi)^2+ \underbrace{\frac{m^2}{2}\Phi^2+\frac{\lambda}{4}\Phi^4}_{ \mbox{${\cal V}(\Phi)$}},$$ where one chooses $\lambda > 0$ so that $\cal H$ is bounded from below. The field $\Phi_0$ which minimizes the Hamilton density ${\cal H}$ must be constant and uniform since in that case the first two terms take everywhere their minimum values of zero. It must also minimize the potential since ${\cal V}(\Phi(x))\geq {\cal V}(\Phi_0)$ (see Fig. \[3:1:potphi02\]), from which we obtain the condition $${\cal V}'(\Phi)=\Phi(m^2+\lambda \Phi^2)=0.$$ We now distinguish two different cases: - $m^2>0$ (see Fig. \[3:1:potww\]): In this case the potential $\cal V$ has its minimum for $\Phi=0$. In the quantized theory we associate a unique ground state $|0\rangle$ with this minimum. Later on, in the case of a continuous symmetry, this situation will be referred to as the Wigner-Weyl realization of the symmetry. - $m^2<0$ (see Fig. \[3:1:potng\]): Now the potential exhibits two distinct minima. (In the continuous symmetry case this will be referred to as the Nambu-Goldstone realization of the symmetry.) We will concentrate on the second situation, because this is the one which we would like to generalize to a continuous symmetry and which ultimately leads to the appearance of Goldstone bosons. In the present case, ${\cal V}(\Phi)$ has a maximum for $\Phi=0$ and [*two*]{} minima for $$\label{3:1:vmin} \Phi_\pm=\pm \sqrt{\frac{-m^2}{\lambda}}\equiv \pm \Phi_0.$$ As will be explained below, the quantized theory develops two degenerate vacua $|0,+\rangle$ and $|0,-\rangle$ which are distinguished through their vacuum expectation values of the field $\Phi(x)$:[^33] $$\begin{aligned} \label{3:1:vewphi} \langle0,+|\Phi(x)|0,+\rangle&=& \langle0,+|e^{iP\cdot x}\Phi(0)e^{-iP\cdot x}|0,+\rangle =\langle0,+|\Phi(0)|0,+\rangle\equiv \Phi_0,\nonumber\\ \langle0,-|\Phi(x)|0,-\rangle&=&-\Phi_0.\end{aligned}$$ We made use of translational invariance, $\Phi(x)=e^{iP\cdot x}\Phi(0)e^{-iP\cdot x}$, and the fact that the ground state is an eigenstate of energy and momentum. We associate with the transformation $R:\Phi\mapsto\Phi'=-\Phi$ a unitary operator $\cal R$ acting on the Hilbert space of our model, with the properties $${\cal R}^2=1, \quad {\cal R}={\cal R}^{-1}={\cal R}^\dagger.$$ In accord with Eq. (\[3:1:vewphi\]) the action of the operator $\cal R$ on the ground states is given by $$\label{3:1:rvpm} {\cal R}|0,\pm\rangle=|0,\mp\rangle.$$ For the moment we select one of the two expectation values and expand the field with respect to $\pm \Phi_0$:[^34] $$\begin{aligned} \label{3:1:entw} \Phi&=&\pm \Phi_0+\Phi',\nonumber\\ \partial_\mu \Phi&=&\partial_\mu \Phi'.\end{aligned}$$ A short calculation yields $$\begin{aligned} {\cal V}(\Phi)&=& \tilde{\cal V}(\Phi')= -\frac{\lambda}{4}\Phi_0^4+\frac{1}{2}(-2m^2)\Phi'^2 \pm\lambda\Phi_0\Phi'^3+\frac{\lambda}{4}\Phi'^4,\end{aligned}$$ such that the Lagrangian in terms of the shifted dynamical variable reads $$\label{3:1:lphip} {\cal L}'(\Phi',\partial_\mu \Phi') =\frac{1}{2}\partial_\mu\Phi'\partial^\mu \Phi' -\frac{1}{2}(-2m^2)\Phi'^2\mp\lambda\Phi_0\Phi'^3-\frac{\lambda}{4}\Phi'^4 +\frac{\lambda}{4}\Phi_0^4.$$ In terms of the new dynamical variable $\Phi'$, the symmetry $R$ is no longer manifest, i.e., it is hidden. Selecting one of the ground states has led to a spontaneous symmetry breaking which is always related to the existence of several degenerate vacua. At this stage it is not clear why the quantum mechanical ground state should be one or the other of $|0,\pm\rangle$ and not a superposition of both. For example, the linear combination $$\frac{1}{\sqrt{2}}\left(|0,+\rangle+|0,-\rangle\right)$$ is invariant under ${\cal R}$ as is the original Lagrangian of Eq. (\[3:1:lphi\]). However, this superposition is not stable against any infinitesimal external perturbation which is odd in $\Phi$ (see Fig. \[3:1:potngsb\]), $${\cal R}(\epsilon H'){\cal R}^\dagger =-\epsilon H'.$$ Any such perturbation will drive the ground state into the vicinity of either $|0,+\rangle$ or $|0,-\rangle$ rather than $\frac{1}{\sqrt{2}}(|0,+\rangle\pm|0,-\rangle)$. This can easily be seen in the framework of perturbation theory for degenerate states. Consider $$|1\rangle=\frac{1}{\sqrt{2}}(|0,+\rangle+|0,-\rangle),\quad |2\rangle=\frac{1}{\sqrt{2}}(|0,+\rangle-|0,-\rangle),$$ such that $${\cal R}|1\rangle=|1\rangle\quad {\cal R}|2\rangle=-|2\rangle.$$ The condition for the energy eigenvalues of the ground state, $E=E^{(0)}+\epsilon E^{(1)}+\cdots$, to first order in $\epsilon$ results from $$\mbox{det}\left(\begin{array}{cc} \langle1|H'|1\rangle-E^{(1)}&\langle1|H'|2\rangle\\\ \langle2|H'|1\rangle&\langle2|H'|2\rangle-E^{(1)}\end{array}\right) =0.$$ Due to the symmetry properties of Eq. (\[3:1:rvpm\]), we obtain $$\langle 1|H'|1\rangle =\langle 1|{\cal R}^{-1}{\cal R}H'{\cal R}^{-1}{\cal R}|1\rangle= \langle 1|-H'|1\rangle=0$$ and similarly $\langle2|H'|2\rangle=0$. Setting $\langle1|H'|2\rangle=a>0$, which can always be achieved by multiplication of one of the two states by an appropriate phase, one finds $$\langle2|H'|1\rangle \stackrel{H'=H'^\dagger}{=} \langle1|H'|2\rangle^\ast=a^\ast=a=\langle1|H'|2\rangle,$$ resulting in $$\mbox{det}\left(\begin{array}{cc}-E^{(1)}&a\\a&-E^{(1)}\end{array}\right) ={E^{(1)}}^2-a^2\stackrel{!}{=}0,\quad \Rightarrow\quad E^{(1)}_{1/2}=\pm a.$$ In other words, the degeneracy has been lifted and we get for the energy eigenvalues $$\label{3:1:eew} E_{1/2}=E^{(0)}\pm\epsilon a+\cdots.$$ The corresponding eigenstates of zeroth order in $\epsilon$ are $|0,+\rangle$ and $|0,-\rangle$, respectively. We thus conclude that an arbitrarily small external perturbation which is odd with respect to $R$ will push the ground state to either $|0,+\rangle$ or $|0,-\rangle$. In the above discussion, we have tacitly assumed that the Hamiltonian and the field $\Phi(x)$ can simultaneously be diagonalized in the vacuum sector, i.e. $\langle 0,+|0,-\rangle =0$. Following Ref. [@Weinberg:kr], we will justify this assumption which will also be crucial for the continuous case to be discussed later. Dispersion relation $E=\sqrt{1+\vec{p}\,^2}$ and asymptote $E=|\vec{p}\,|$. For an infinite volume, a general vacuum state $|v\rangle$ is defined as a state with momentum eigenvalue $\vec{0}$, $$\vec{P}|v\rangle=\vec{0},$$ where $\vec{0}$ is a [*discrete*]{} eigenvalue as opposed to an eigenvalue of single- or many-particle states for which $\vec{p}=0$ is an element of a continuous spectrum (see Fig. \[3:1:energie\]). We deal with the situation of several degenerate ground states which will be denoted by $|u\rangle$, $|v\rangle$, [*etc*]{}[^35] and start from the identity $$\label{3:1:0ucv} 0=\langle u|[H,\Phi(x)]|v\rangle\,\,\forall\,\,x,$$ from which we obtain for $t=0$ $$\begin{aligned} \label{3:1:0uvh} \int d^3 y \langle u|{\cal H}(\vec{y},0) \Phi(\vec{x},0)|v\rangle&=& \int d^3 y \langle u|\Phi(\vec{x},0) {\cal H}(\vec{y},0)|v\rangle.\end{aligned}$$ Let us consider the left-hand side, $$\begin{aligned} \lefteqn{\int d^3 y \langle u|{\cal H}(\vec{y},0) \Phi(\vec{x},0)|v\rangle =\sum_w \langle u|H|w\rangle\langle w|\Phi(0)|v\rangle}\\ && +\int d^3 y\int d^3p\sum_n\langle u|{\cal H}(\vec{y},0)|n,\vec{p}\,\rangle \langle n,\vec{p}\,|\Phi(0)|v\rangle e^{-i\vec{p}\cdot \vec{x}},\end{aligned}$$ where we inserted a complete set of states which we split into the vacuum contribution and the rest, and made use of translational invariance. We now define $$f_n(\vec{y},\vec{p}\,)=\langle u|{\cal H}(\vec{y},0)|n,\vec{p}\,\rangle \langle n,\vec{p}\,|\Phi(0)|v\rangle$$ and assume $f_n$ to be reasonably behaved such that one can apply the lemma of Riemann and Lebesgue, $$\lim_{|\vec{x}|\to\infty}\int d^3p f(\vec{p}\,)e^{-i\vec{p}\cdot\vec{x}}=0.$$ At this point the assumption of an infinite volume, $|\vec{x}|\to\infty$, is crucial. Repeating the argument for the right-hand side and taking the limit $|\vec{x}|\to\infty$, only the vacuum contributions survive in Eq.(\[3:1:0uvh\]) and we obtain $$\sum_w\langle u|H|w\rangle\langle w|\Phi(0)|v\rangle =\sum_w\langle u|\Phi(0)|w\rangle\langle w|H|v\rangle$$ for arbitrary ground states $|u\rangle$ and $|v\rangle$. In other words, the matrices $(H_{uv})\equiv (\langle u|H|v\rangle)$ and $(\Phi_{uv})\equiv(\langle u|\Phi(0)|v \rangle)$ commute and can be diagonalized simultaneously. Choosing an appropriate basis, one can write $$\langle u|\Phi(0)|v\rangle=\delta_{uv}v,\quad v\in R,$$ where $v$ denotes the expectation value of $\Phi$ in the state $|v\rangle$. In the above example, the ground states $|0,+\rangle$ and $|0,-\rangle$ with vacuum expectation values $\pm \Phi_0$ are thus indeed orthogonal and satisfy $$\langle 0,+|H|0,-\rangle=\langle 0,-|H|0,+\rangle=0.$$ Spontaneous Breakdown of a Global, Continuous, Non-Abelian Symmetry {#sec_sbgcnas} ------------------------------------------------------------------- We now extend the discussion to a system with a continuous, non-Abelian symmetry such as SO(3). To that end, we consider the Lagrangian $$\begin{aligned} \label{3:2:lphi} {\cal L}(\vec{\Phi},\partial_\mu\vec{\Phi}) &=&{\cal L}(\Phi_1,\Phi_2,\Phi_3,\partial_\mu\Phi_1, \partial_\mu\Phi_2,\partial_\mu\Phi_3)\nonumber\\ &=&\frac{1}{2}\partial_\mu \Phi_i\partial^\mu \Phi_i-\frac{m^2}{2}\Phi_i\Phi_i -\frac{\lambda}{4}(\Phi_i\Phi_i)^2,\end{aligned}$$ where $m^2<0$, $\lambda>0$, with Hermitian fields $\Phi_i$. The Lagrangian of Eq. (\[3:2:lphi\]) is invariant under a global “isospin” rotation,[^36] $$\label{3:2:phitrafo} g\in \mbox{SO(3)}:\,\,\Phi_i\to\Phi_i'=D_{ij}(g)\Phi_j= (e^{-i\alpha_k T_k})_{ij}\Phi_j.$$ For the $\Phi_i'$ to also be Hermitian, the Hermitian $T_k$ must be purely imaginary and thus antisymmetric. The $iT_k$ provide the basis of a representation of the so(3) Lie algebra and satisfy the commutation relations $[T_i,T_j]=i\epsilon_{ijk} T_k$. We will use the representation with the matrix elements given by $t_{jk}^i=-i\epsilon_{ijk}$. As in Sec. 3.1, we now look for a minimum of the potential which does not depend on $x$ and find $$\label{3:2:phimin} |\vec{\Phi}_{\rm min}|=\sqrt{\frac{-m^2}{\lambda}}\equiv v, \quad |\vec{\Phi}|=\sqrt{\Phi_1^2+\Phi_2^2+\Phi_3^2}.$$ Since $\vec{\Phi}_{\rm min}$ can point in any direction in isospin space we now have a non-countably infinite number of degenerate vacua. In analogy to the discussion of the last section, any infinitesimal external perturbation which is not invariant under SO(3) will select a particular direction which, by an appropriate orientation of the internal coordinate frame, we denote as the 3 direction, $$\label{3:2:phimin3} \vec{\Phi}_{\rm min}=v \hat{e}_3.$$ Clearly, $\vec{\Phi}_{\rm min}$ of Eq. (\[3:2:phimin3\]) is [*not*]{} invariant under the full group $G=\mbox{SO(3)}$ since rotations about the 1 and 2 axis change $\vec{\Phi}_{\rm min}$.[^37] To be specific, if $$\vec{\Phi}_{\rm min}=v\left(\begin{array}{r}0\\0\\1\end{array}\right),$$ we obtain $$\label{3:2:t12phimin} T_1 \vec{\Phi}_{\rm min}= v\left(\begin{array}{r}0\\-i\\0\end{array}\right),\quad T_2 \vec{\Phi}_{\rm min}= v\left(\begin{array}{r}i\\0\\0\end{array}\right), \quad T_3 \vec{\Phi}_{\rm min}=0.$$ Note that the set of transformations which do not leave $\vec{\Phi}_{\rm min}$ invariant does [*not*]{} form a group, because it does not contain the identity. On the other hand, $\vec{\Phi}_{\rm min}$ is invariant under a subgroup $H$ of $G$, namely, the rotations about the 3 axis: $$\label{3:2:phimintrafoh} h\in H:\quad \vec{\Phi}'=D(h)\vec{\Phi}=e^{-i\alpha_3 T_3}\vec{\Phi}, \quad D(h)\vec{\Phi}_{\rm min}=\vec{\Phi}_{\rm min}.$$ In analogy to Eq. (\[3:1:entw\]), we expand $\Phi_3$ with respect to $v$, $$\label{3:2:entw} \Phi_3=v+\eta,$$ where $\eta(x)$ is a new field replacing $\Phi_3(x)$, and obtain the new expression for the potential $$\begin{aligned} \label{3:2:ventw} \tilde{\cal V} &=&\frac{1}{2}(-2m^2)\eta^2 +\lambda v\eta (\Phi_1^2+\Phi_2^2+\eta^2) +\frac{\lambda}{4}(\Phi_1^2+\Phi_2^2+\eta^2)^2-\frac{\lambda}{4}v^4. \nonumber\\\end{aligned}$$ Upon inspection of the terms quadratic in the fields, one finds after spontaneous symmetry breaking two massless Goldstone bosons and one massive boson: $$\begin{aligned} \label{3:2:masses} m_{\Phi_1}^2=m_{\Phi_2}^2&=&0,\nonumber\\ m_\eta^2&=&-2m^2.\end{aligned}$$ The model-independent feature of the above example is given by the fact that for each of the two generators $T_1$ and $T_2$ which do not annihilate the ground state one obtains a [*massless*]{} Goldstone boson. By means of a two-dimensional simplification (see the “Mexican hat” potential shown in Fig. \[3:2:pot2dim\]) the mechanism at hand can easily be visualized. Infinitesimal variations orthogonal to the circle of the minimum of the potential generate quadratic terms, i.e., “restoring forces linear in the displacement,” whereas tangential variations experience restoring forces only of higher orders. Two-dimensional rotationally invariant potential: ${\cal V}(x,y)=-(x^2+y^2)+\frac{(x^2+y^2)^2}{4}$. Now let us generalize the model to the case of an arbitrary compact Lie group $G$ of order $n_G$ resulting in $n_G$ infinitesimal generators.[^38] Once again, we start from a Lagrangian of the form [@Goldstone:es] $$\label{3:2:lallg} {\cal L}(\vec{\Phi},\partial_\mu\vec{\Phi}) =\frac{1}{2}\partial_\mu \vec{\Phi}\cdot \partial^\mu \vec{\Phi}- {\cal V}(\vec{\Phi}),$$ where $\vec{\Phi}$ is a multiplet of scalar (or pseudoscalar) Hermitian fields. The Lagrangian ${\cal L}$ and thus also ${\cal V}(\vec{\Phi})$ are supposed to be globally invariant under $G$, where the infinitesimal transformations of the fields are given by $$\label{3:2:symmtrans} g\in G:\quad \Phi_i\to\Phi_i+\delta\Phi_i,\quad \delta \Phi_i=-i\epsilon_a t^a_{ij}\Phi_j.$$ The Hermitian representation matrices $T^a=(t^a_{ij})$ are again antisymmetric and purely imaginary. We now assume that, by choosing an appropriate form of ${\cal V}$, the Lagrangian generates a spontaneous symmetry breaking resulting in a ground state with a vacuum expectation value $\vec{\Phi}_{\rm min}=\langle\vec{\Phi}\rangle$ which is invariant under a continuous subgroup $H$ of $G$. We expand ${\cal V}(\vec{\Phi})$ with respect to $\vec{\Phi}_{\rm min}$, $|\vec{\Phi}_{\rm min}|=v$, i.e., $\vec{\Phi}=\vec{\Phi}_{\rm min}+\vec{\chi}$, $$\label{3:2:venta} {\cal V}(\vec{\Phi})={\cal V}(\vec{\Phi}_{\rm min}) +\underbrace{\frac{\partial {\cal V}(\vec{\Phi}_{\rm min})}{\partial \Phi_i}}_{ \mbox{$0$}}\chi_i +\frac{1}{2}\underbrace{\frac{\partial^2 {\cal V}(\vec{\Phi}_{\rm min})}{\partial \Phi_i\partial \Phi_j}}_{\mbox{$m^2_{ij}$}}\chi_i\chi_j+\cdots.$$ The matrix $M^2=(m^2_{ij})$ must be symmetric and, since one is expanding around a minimum, positive semidefinite, i.e., $$\label{3:2:m2} \sum_{i,j}m^2_{ij}x_i x_j\ge 0\quad \forall \quad \vec{x}.$$ In that case, all eigenvalues of $M^2$ are nonnegative. Making use of the invariance of ${\cal V}$ under the symmetry group $G$, $$\begin{aligned} \label{3:2:vinv} {\cal V}(\vec{\Phi}_{\rm min})&=&{\cal V}(D(g)\vec{\Phi}_{\rm min}) ={\cal V}(\vec{\Phi}_{\rm min}+\delta\vec{\Phi}_{\rm min})\nonumber\\ &\stackrel{\mbox{(\ref{3:2:venta})}}{=}& {\cal V}(\vec{\Phi}_{\rm min})+\frac{1}{2}m^2_{ij}\delta\Phi_{\rm min,i} \delta{\Phi}_{\rm min,j}+\cdots,\end{aligned}$$ one obtains, by comparing coefficients, $$\label{3:2:kv} m^2_{ij}\delta\Phi_{\rm min,i}\delta\Phi_{\rm min,j}=0.$$ Differentiating Eq. (\[3:2:kv\]) with respect to $\delta\Phi_{\rm min,k}$ and using $m^2_{ij}=m^2_{ji}$ results in the matrix equation $$\label{3:2:rel1} M^2\delta\vec{\Phi}_{\rm min}=\vec{0}.$$ Inserting the variations of Eq. (\[3:2:symmtrans\]) for arbitrary $\epsilon_a$, $\delta\vec{\Phi}_{\rm min}=-i\epsilon_a T^a\vec{\Phi}_{\rm min}$, we conclude $$\label{3:2:result} M^2T^a \vec{\Phi}_{\rm min}=\vec{0}.$$ The solutions of Eq. (\[3:2:result\]) can be classified into two categories: 1. $T^a$, $a=1,\cdots, n_H$, is a representation of an element of the Lie algebra belonging to the subgroup $H$ of $G$, leaving the selected ground state invariant. In that case one has $$T^a \vec{\Phi}_{\rm min}=\vec{0}, \quad a=1,\cdots,n_H,$$ such that Eq. (\[3:2:result\]) is automatically satisfied without any knowledge of $M^2$. 2. $T^a$, $a=n_H+1,\cdots, n_G,$ is [*not*]{} a representation of an element of the Lie algebra belonging to the subgroup $H$. In that case $T^a\vec{\Phi}_{\rm min}\neq\vec{0}$, and $T^a\vec{\Phi}_{\rm min}$ is an eigenvector of $M^2$ with eigenvalue 0. To each such eigenvector corresponds a massless Goldstone boson. In particular, the different $T^a\vec{\Phi}_{\rm min}\neq \vec{0}$ are linearly independent, resulting in $n_G-n_H$ independent Goldstone bosons. (If they were not linearly independent, there would exist a nontrivial linear combination $$\vec{0}=\sum_{a=n_H+1}^{n_G}c_a (T^a\vec{\Phi}_{\rm min})= \underbrace{\left(\sum_{a=n_H+1}^{n_G}c_a T^a\right)}_{\mbox{$:=T$}} \vec{\Phi}_{\rm min},$$ such that $T$ is an element of the Lie algebra of $H$ in contradiction to our assumption.) Let us check these results by reconsidering the example of Eq. (\[3:2:lphi\]). In that case $n_G=3$ and $n_H=1$, generating 2 Goldstone bosons \[see Eq. (\[3:2:masses\])\]. We conclude this section with two remarks. First, the number of Goldstone bosons is determined by the structure of the symmetry groups. Let $G$ denote the symmetry group of the Lagrangian, with $n_G$ generators and $H$ the subgroup with $n_H$ generators which leaves the ground state after spontaneous symmetry breaking invariant. For each generator which does not annihilate the vacuum one obtains a massless Goldstone boson, i.e., the total number of Goldstone bosons equals $n_G-n_H$. Second, the Lagrangians used in [*motivating*]{} the phenomenon of a spontaneous symmetry breakdown are typically constructed in such a fashion that the degeneracy of the ground states is built into the potential at the classical level (the prototype being the “Mexican hat” potential of Fig. \[3:2:pot2dim\]). As in the above case, it is then argued that an [*elementary*]{} Hermitian field of a multiplet transforming non-trivially under the symmetry group $G$ acquires a vacuum expectation value signaling a spontaneous symmetry breakdown. However, there also exist theories such as QCD where one cannot infer from inspection of the Lagrangian whether the theory exhibits spontaneous symmetry breaking. Rather, the criterion for spontaneous symmetry breaking is a non-vanishing vacuum expectation value of some Hermitian operator, not an elementary field, which is generated through the dynamics of the underlying theory. In particular, we will see that the quantities developing a vacuum expectation value may also be local Hermitian operators composed of more fundamental degrees of freedom of the theory. Such a possibility was already emphasized in the derivation of Goldstone’s theorem in Ref. [@Goldstone:es]. Goldstone’s Theorem {#sec_gt} ------------------- By means of the above example, we motivate another approach to Goldstone’s theorem without delving into all the subtleties of a quantum field-theoretical approach [@Bernstein]. Given a Hamilton operator with a global symmetry group $G=\mbox{SO(3)}$, let $\vec{\Phi}(x)=(\Phi_1(x),\Phi_2(x),\Phi_3(x))$ denote a triplet of local Hermitian operators transforming as a vector under $G$, $$\begin{aligned} \label{3:3:phitrafo} g\in G:&&\vec{\Phi}(x)\mapsto \vec{\Phi}'(x)= e^{i\sum_{k=1}^3\alpha_k Q_k}\vec{\Phi}(x)e^{-i\sum_{l=1}^3\alpha_l Q_l} \nonumber\\ &&=e^{-i\sum_{k=1}^3 \alpha_k T_k} \vec{\Phi}(x) \neq \vec{\Phi}(x),\end{aligned}$$ where the $Q_i$ are the generators of the SO(3) transformations on the Hilbert space satisfying $[Q_i,Q_j]=i\epsilon_{ijk}Q_k$ and the $T_i=(t^i_{jk})$ are the matrices of the three dimensional representation satisfying $t^i_{jk}=-i\epsilon_{ijk}$. We assume that one component of the multiplet acquires a non-vanishing vacuum expectation value: $$\label{3:3:phi3vac} \langle0|\Phi_1(x)|0\rangle =\langle0|\Phi_2(x)|0\rangle=0,\quad \langle0|\Phi_3(x)|0\rangle=v\neq 0.$$ Then the two generators $Q_1$ and $Q_2$ do not annihilate the ground state, and to each such generator corresponds a massless Goldstone boson. In order to prove these two statements let us expand Eq. (\[3:3:phitrafo\]) to first order in the $\alpha_k$: $$\vec{\Phi}'=\vec{\Phi}+i\sum_{k=1}^3\alpha_k[Q_k,\vec{\Phi}] =(1-i\sum_{k=1}^3\alpha_k T_k)\vec{\Phi} =\vec{\Phi}+\vec{\alpha}\times\vec{\Phi}.$$ Comparing the terms linear in the $\alpha_k$ $$i[\alpha_k Q_k,\Phi_l]=\epsilon_{lkm}\alpha_k\Phi_m$$ and noting that all three $\alpha_k$ can be chosen independently, we obtain $$i[Q_k,\Phi_l]=-\epsilon_{klm}\Phi_m,$$ which, of course, simply expresses the fact that the field operators $\Phi_i$ transform as a vector. Using $\epsilon_{klm}\epsilon_{kln}=2\delta_{mn}$, we find $$-\frac{i}{2}\epsilon_{kln}[Q_k,\Phi_l]=\delta_{mn}\Phi_m=\Phi_n.$$ In particular, $$\label{3:3:phi3zykl} \Phi_3=-\frac{i}{2}([Q_1,\Phi_2]-[Q_2,\Phi_1]),$$ with cyclic permutations for the other two cases. In order to prove that $Q_1$ and $Q_2$ do not annihilate the ground state, let us consider Eq. (\[3:3:phitrafo\]) for $\vec{\alpha}=(0,\pi/2,0)$, $$\begin{aligned} e^{-i\frac{\pi}{2} T_2}\vec{\Phi}&=& \left(\begin{array}{ccc} \cos(\pi/2)&0&\sin(\pi/2)\\ 0&1&0\\ -\sin(\pi/2)&0&\cos(\pi/2)\end{array}\right) \left(\begin{array}{c} \Phi_1\\ \Phi_2 \\ \Phi_3\end{array}\right) =\left(\begin{array}{c}\Phi_3\\ \Phi_2 \\ -\Phi_1\end{array}\right)\\ &=&e^{i\frac{\pi}{2}Q_2}\left(\begin{array}{c} \Phi_1\\ \Phi_2 \\ \Phi_3\end{array}\right) e^{-i\frac{\pi}{2}Q_2}.\end{aligned}$$ From the first row we obtain $$\Phi_3=e^{i\frac{\pi}{2}Q_2}\Phi_1 e^{-i\frac{\pi}{2}Q_2}.$$ Taking the vacuum expectation value $$v=\langle 0| e^{i\frac{\pi}{2}Q_2}\Phi_1 e^{-i\frac{\pi}{2}Q_2}|0\rangle$$ and using Eq. (\[3:3:phi3vac\]) clearly $Q_2|0\rangle\neq 0$, since otherwise the exponential operator could be replaced by unity and the right-hand side would vanish. A similar argument shows $Q_1|0\rangle\neq 0$. At this point let us make two remarks. The “states” $Q_{1(2)}|0\rangle$ cannot be normalized. In a more rigorous derivation one makes use of integrals of the form $$\int d^3 x \langle0|[J^{0,b}(\vec{x},t),\Phi_c(0)]|0\rangle,$$ and first determines the commutator before evaluating the integral [@Bernstein]. Some derivations of Goldstone’s theorem right away start by assuming $Q_{1(2)}|0\rangle$ $\neq 0$. However, for the discussion of spontaneous symmetry breaking in the framework of QCD it is advantageous to establish the connection between the existence of Goldstone bosons and a non-vanishing expectation value. Let us now turn to the existence of Goldstone bosons, taking the vacuum expectation value of Eq. (\[3:3:phi3zykl\]) $$0\neq v=\langle0|\Phi_3(0)|0\rangle = -\frac{i}{2}\langle0|\left([Q_1,\Phi_2(0)]-[Q_2,\Phi_1(0)]\right)|0\rangle \equiv -\frac{i}{2}(A-B).$$ We will first show $A=-B$. To that end we perform a rotation of the fields as well as the generators by $\pi/2$ about the 3 axis \[see Eq. (\[3:3:phitrafo\]) with $\vec{\alpha}=(0,0,\pi/2)$\]: $$\begin{aligned} e^{-i\frac{\pi}{2}T_3}\vec{\Phi} &=&\left(\begin{array}{r}-\Phi_2\\ \Phi_1\\ \Phi_3\end{array}\right)= e^{i\frac{\pi}{2}Q_3} \left(\begin{array}{c}\Phi_1\\ \Phi_2\\ \Phi_3\end{array}\right) e^{-i\frac{\pi}{2}Q_3},\end{aligned}$$ and analogously for the charge operators $$\left(\begin{array}{r} -Q_2\\Q_1\\Q_3\end{array}\right) =e^{i\frac{\pi}{2}Q_3}\left(\begin{array}{r}Q_1\\Q_2\\Q_3\end{array} \right)e^{-i\frac{\pi}{2}Q_3}.$$ We thus obtain $$\begin{aligned} B=\langle0|[Q_2, \Phi_1(0)]|0\rangle&=& \langle0|\Big(e^{i\frac{\pi}{2}Q_3}(- Q_1) \underbrace{e^{-i\frac{\pi}{2}Q_3}e^{i\frac{\pi}{2}Q_3}}_{\mbox{1}} \Phi_2(0) e^{-i\frac{\pi}{2}Q_3}\\ &&- e^{i\frac{\pi}{2}Q_3}\Phi_2(0) e^{-i\frac{\pi}{2}Q_3}e^{i\frac{\pi}{2}Q_3} (-Q_1) e^{-i\frac{\pi}{2}Q_3}\Big)|0\rangle\\ &=&-\langle0|[Q_1,\Phi_2(0)]|0\rangle=-A,\end{aligned}$$ where we made use of $Q_3|0\rangle=0$, i.e., the vacuum is invariant under rotations about the 3 axis. In other words, the non-vanishing vacuum expectation value $v$ can also be written as $$\begin{aligned} \label{3:3:vq1phi2} 0\neq v&=&\langle0|\Phi_3(0)|0\rangle=-i\langle0|[Q_1,\Phi_2(0)]|0\rangle \nonumber\\ &=&-i\int d^3 x\langle0|[J^1_0(\vec{x},t),\Phi_2(0)]|0\rangle.\end{aligned}$$ We insert a complete set of states $1=\sum_n\hspace{-1.4em}\int \hspace{0.5em} |n\rangle \langle n|$ into the commutator[^39] $$\begin{aligned} v&=&-i\sum_n\hspace{-1.1em}\int \int d^3x \left(\langle0|J^{1}_0(\vec{x},t)|n\rangle \langle n|\Phi_2(0)|0\rangle-\langle0|\Phi_2(0)|n\rangle \langle n|J^{1}_0(\vec{x},t)|0\rangle\right),\end{aligned}$$ and make use of translational invariance $$\begin{aligned} &=&-i\sum_n\hspace{-1.1em}\int \int d^3x\left(e^{-iP_n\cdot x} \langle0|J^{1}_0(0)|n\rangle\langle n|\Phi_2(0)|0\rangle -\cdots\right)\\ &=&-i\sum_n\hspace{-1.1em}\int (2\pi)^3\delta^3(\vec{P}_n) \left(e^{-iE_n t} \langle0|J^{1}_0(0)|n\rangle\langle n|\Phi_2(0)|0\rangle\right.\nonumber\\ &&\left.-e^{iE_n t}\langle 0|\Phi_2(0)|n\rangle \langle n|J^{1}_0(0)|0\rangle\right).\end{aligned}$$ Integration with respect to the momentum of the inserted intermediate states yields an expression of the form $$=-i(2\pi)^3 \sum_n' \left(e^{-iE_n t}\cdots -e^{iE_n t}\cdots\right),$$ where the prime indicates that only states with $\vec{P}=0$ need to be considered. Due to the Hermiticity of the symmetry current operators $J^{\mu,a}$ as well as the $\Phi_l$, we have $$c_n:=\langle 0|J^{1}_0(0)|n\rangle\langle n|\Phi_2(0)|0\rangle =\langle n|J^{1}_0(0)|0\rangle^\ast \langle0|\Phi_2(0)|n\rangle^\ast,$$ such that $$\label{3:3:vresult} v=-i(2\pi)^3\sum_n' \left(c_n e^{-iE_n t}-c_n^\ast e^{iE_n t}\right).$$ From Eq. (\[3:3:vresult\]) we draw the following conclusions. 1. Due to our assumption of a non-vanishing vacuum expectation value $v$, there must exist states $|n\rangle$ for which both $\langle0|J^{0}_{1(2)}(0)|n\rangle$ and $\langle n|\Phi_{1(2)}(0)|0\rangle$ do not vanish. The vacuum itself cannot contribute to Eq. (\[3:3:vresult\]) because $\langle0|\Phi_{1(2)}(0)|0\rangle=0$. 2. States with $E_n>0$ contribute ($\varphi_n$ is the phase of $c_n$) $$\begin{aligned} \frac{1}{i}\left(c_n e^{-iE_n t}-c_n^\ast e^{iE_n t}\right) &=&\frac{1}{i}|c_n|\left(e^{i\varphi_n}e^{-iE_n t} -e^{-i\varphi_n}e^{iE_n t}\right)\\ &=&2|c_n|\sin(\varphi_n-E_n t)\end{aligned}$$ to the sum. However, $v$ is time-independent and therefore the sum over states with $(E_n>0,\vec{0})$ must vanish. 3. The right-hand side of Eq. (\[3:3:vresult\]) must therefore contain the contribution from states with zero energy as well as zero momentum thus zero mass. These zero-mass states are the Goldstone bosons. Explicit Symmetry Breaking: A First Look {#sec_esbfl} ---------------------------------------- Finally, let us illustrate the consequences of adding to our Lagrangian of Eq. (\[3:2:lphi\]) a small perturbation which [*explicitly*]{} breaks the symmetry. To that end, we modify the potential of Eq. (\[3:2:lphi\]) by adding a term $a\Phi_3$, $$\label{3:4:pot} {\cal V}(\Phi_1,\Phi_2,\Phi_3)= \frac{m^2}{2}\Phi_i\Phi_i +\frac{\lambda}{4}(\Phi_i\Phi_i)^2 + a\Phi_3,$$ where $m^2<0$, $\lambda>0$, and $a>0$, with Hermitian fields $\Phi_i$. Clearly, the potential no longer has the original O(3) symmetry but is only invariant under O(2). The conditions for the new minimum, obtained from $\vec{\nabla}_\Phi {\cal V}=0$, read $$\Phi_1=\Phi_2=0,\quad \lambda \Phi_3^3+m^2\Phi_3+a=0.$$ Let us solve the cubic equation for $\Phi_3$ using the perturbative ansatz $$\label{3:4:phi3ansatz} \langle \Phi_3 \rangle=\Phi^{(0)}_3+a\Phi^{(1)}_3+{\cal O}(a^2),$$ from which we obtain $$\Phi^{(0)}_3=\pm \sqrt{-\frac{m^2}{\lambda}},\quad \Phi^{(1)}_3=\frac{1}{2m^2}.$$ Of course, $\Phi^{(0)}_3$ corresponds to our result without explicit perturbation. The condition for a [*minimum*]{} \[see Eq. (\[3:2:m2\])\] excludes $\Phi^{(0)}_3=+\sqrt{-\frac{m^2}{\lambda}}$. Expanding the potential with $\Phi_3=\langle\Phi_3\rangle +\eta$ we obtain, after a short calculation, for the masses $$\begin{aligned} \label{3:4:masses} m_{\Phi_1}^2=m_{\Phi_2}^2&=&a \sqrt{\frac{\lambda}{-m^2}},\nonumber\\ m_\eta^2&=&-2m^2+3 a\sqrt{\frac{\lambda}{-m^2}}.\end{aligned}$$ The important feature here is that the original Goldstone bosons of Eq. (\[3:2:masses\]) are now massive. The squared masses are proportional to the symmetry breaking parameter $a$. Calculating [*quantum*]{} corrections to observables in terms of Goldstone-boson loop diagrams will generate corrections which are non-analytic in the symmetry breaking parameter such as $a\ln(a)$ [@Li:1971vr]. Such so-called chiral logarithms originate from the mass terms in the Goldstone boson propagators entering the calculation of loop integrals. We will come back to this point in Chapter 4 when we discuss the masses of the pseudoscalar octet in terms of the quark masses which, in QCD, represent the analogue to the parameter $a$ in the above example. Chiral Perturbation Theory for Mesons {#chap_cptm} ===================================== Chiral perturbation theory provides a systematic method for discussing the consequences of the global flavor symmetries of QCD at low energies by means of an [*effective field theory*]{}. The effective Lagrangian is expressed in terms of those hadronic degrees of freedom which, at low energies, show up as observable asymptotic states. At very low energies these are just the members of the pseudoscalar octet ($\pi,K,\eta$) which are regarded as the Goldstone bosons of the [*spontaneous*]{} breaking of the chiral $\mbox{SU(3)}_L\times\mbox{SU}(3)_R$ symmetry down to $\mbox{SU}(3)_V$. The non-vanishing masses of the light pseudoscalars in the “real” world are related to the explicit symmetry breaking in QCD due to the light quark masses. We will first consider the indications for a spontaneous breakdown of chiral symmetry in QCD and then, in quite general terms, discuss the transformation properties of Goldstone bosons under the symmetry groups of the Lagrangian and the ground state, respectively. This will lead us to the concept of a nonlinear realization of a symmetry. After introducing the lowest-order effective Lagrangian relevant to the spontaneous breakdown from $\mbox{SU(3)}_L\times\mbox{SU}(3)_R$ to $\mbox{SU}(3)_V$, we will illustrate how Weinberg’s power counting scheme allows for a systematic classification of Feynman diagrams in the so-called momentum expansion. We will then outline the principles entering the construction of the effective Lagrangian and discuss how, at lowest order, the results of current algebra are reproduced. After presenting the Lagrangian of Gasser and Leutwyler and the Wess-Zumino-Witten action we will discuss some applications at chiral order ${\cal O}(p^4)$. We will conclude the presentation of the mesonic sector with referring to some selected examples at ${\cal O}(p^6)$. Spontaneous Symmetry Breaking in QCD {#sec_ssbqcd} ------------------------------------ While the toy model of Sec. \[sec\_sbgcnas\] by construction led to a spontaneous symmetry breaking, it is not fully understood theoretically why QCD should exhibit this phenomenon [@Jaffe:2000]. We will first motivate why experimental input, the hadron spectrum of the “real” world, indicates that spontaneous symmetry breaking happens in QCD. Secondly, we will show that a non-vanishing singlet scalar quark condensate is a sufficient condition for a spontaneous symmetry breaking in QCD. ### The Hadron Spectrum {#subsec_hs} We saw in Sec. \[sec\_agsl\] that the QCD Lagrangian possesses a $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times \mbox{U(1)}_V$ symmetry in the chiral limit in which the light quark masses vanish. From symmetry considerations involving the Hamiltonian $H^0_{\rm QCD}$ only, one would naively expect that hadrons organize themselves into approximately degenerate multiplets fitting the dimensionalities of irreducible representations of the group $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$. The $\mbox{U(1)}_V$ symmetry results in baryon number conservation[^40] and leads to a classification of hadrons into mesons ($B=0$) and baryons ($B=1$). The linear combinations $Q^a_V=Q^a_R+Q^a_L$ and $Q^a_A=Q^a_R-Q^a_L$ of the left- and right-handed charge operators commute with $H^0_{\rm QCD}$, have opposite parity, and thus for any state of positive parity one would expect the existence of a degenerate state of negative parity (parity doubling) which can be seen as follows. Let $|i,+\rangle$ denote an eigenstate of $H^0_{\rm QCD}$ with eigenvalue $E_i$, $$H^0_{\rm QCD}|i,+\rangle=E_i|i,+\rangle,$$ having positive parity, $$P|i,+\rangle=+ |i,+\rangle,$$ such as, e.g., a member of the ground state baryon octet (in the chiral limit). Defining $|\phi\rangle= Q_A^a|i,+\rangle$, because of $[H^0_{\rm QCD},Q_A^a]=0$, we have $$H^0_{\rm QCD}|\phi\rangle =H^0_{\rm QCD} Q_A^a|i,+\rangle = Q_A^a H^0_{\rm QCD}|i,+\rangle = E_i Q_A^a|i,+\rangle = E_i |\phi\rangle,$$ i.e, the new state $|\phi\rangle$ is also an eigenstate of $H^0_{\rm QCD}$ with the same eigenvalue $E_i$ but of opposite parity: $$P|\phi\rangle= PQ_A^a P^{-1} P|i,+\rangle=-Q_A^a(+|i,+\rangle) =-|\phi\rangle.$$ The state $|\phi\rangle$ can be expanded in terms of the members of the multiplet with negative parity, $$|\phi\rangle=Q^a_A|i,+\rangle=-t^a_{ij}|j,-\rangle.$$ However, the low-energy spectrum of baryons does not contain a degenerate baryon octet of negative parity. Naturally the question arises whether the above chain of arguments is incomplete. Indeed, we have tacitly assumed that the ground state of QCD is annihilated by $Q^a_A$. Let $a^\dagger_i$ symbolically denote an operator which creates quanta with the quantum numbers of the state $|i,+\rangle$, whereas $b_j^\dagger$ creates degenerate quanta of opposite parity. Let us assume the states $|i,+\rangle$ and $|j,-\rangle$ to be members of a basis of an irreducible representation of $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$. In analogy to Eq. (\[2:3:qphi\]), we assume that under $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ the creation operators are related by $$[Q^a_A,a^\dagger_i]= -t^a_{ij} b_j^\dagger.$$ The usual chain of arguments then works as $$\begin{aligned} \label{4:1:pardoub} Q^a_A|i,+\rangle&=&Q^a_A a^\dagger_i|0\rangle =\Big([Q^a_A,a^\dagger_i]+a_i^\dagger \underbrace{Q_A^a}_{ \mbox{$\hookrightarrow 0$}}\Big)|0\rangle = -t^a_{ij} b_j^\dagger |0\rangle.\end{aligned}$$ However, if the ground state is [*not*]{} annihilated by $Q_A^a$, the reasoning of Eq. (\[4:1:pardoub\]) does no longer apply. Two empirical facts about the hadron spectrum suggest that a spontaneous symmetry breaking happens in the chiral limit of QCD. First, $\mbox{SU(3)}$ instead of $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ is approximately realized as a symmetry of the hadrons. Second, the octet of the pseudoscalar mesons is special in the sense that the masses of its members are small in comparison with the corresponding $1^-$ vector mesons. They are candidates for the Goldstone bosons of a spontaneous symmetry breaking. In order to understand the origin of the SU(3) symmetry let us consider the vector charges $Q^a_V=Q_R^a+Q^a_L$ \[see Eq. (\[2:3:v\])\]. They satisfy the commutation relations of an SU(3) Lie algebra \[see Eqs. (\[2:2:crqll\]) - (\[2:3:crqlr\])\], $$\label{4:1:su3v} [Q_R^a+Q_L^a,Q_R^b+Q_L^b]=[Q_R^a,Q_R^b]+[Q_L^a,Q_L^b] =if_{abc} Q_R^c+if_{abc}Q_L^c=if_{abc}Q^c_V.$$ In Ref. [@Vafa:tf] it was shown that, in the chiral limit, the ground state is necessarily invariant under $\mbox{SU(3)}_V\times\mbox{U(1)}_V$, i.e., the eight vector charges $Q^a_V$ as well as the baryon number operator[^41] $Q_V/3$ annihilate the ground state, $$Q^a_V|0\rangle =Q_V|0\rangle =0.$$ If the vacuum is invariant under $\mbox{SU(3)}_V\times\mbox{U(1)}_V$, then so is the Hamiltonian [@Coleman:1966] (but not vice versa). Moreover, the invariance of the ground state [*and*]{} the Hamiltonian implies that the physical states of the spectrum of $H^0_{\rm QCD}$ can be organized according to irreducible representations of $\mbox{SU(3)}_V\times\mbox{U(1)}_V$. The index $V$ (for vector) indicates that the generators result from integrals of the zeroth component of vector current operators and thus transform with a positive sign under parity. Let us now turn to the linear combinations $Q^a_A=Q^a_R-Q^a_L$ satisfying the commutation relations \[see Eqs. (\[2:2:crqll\]) - (\[2:3:crqlr\])\] $$\begin{aligned} \label{4:1:crqaa} [Q^a_A,Q^b_A]&=&[Q^a_R-Q^a_L,Q^b_R-Q^b_L] =[Q^a_R,Q^b_R]+[Q^a_L,Q^b_L]\nonumber\\ &=&if_{abc}Q^c_R+if_{abc}Q^c_L= if_{abc}Q^c_V,\nonumber\\ \label{4:1:crqva} {[Q_V^a,Q^b_A]}&=&[Q_R^a+Q_L^a,Q_R^b-Q_L^b]= [Q_R^a,Q^b_R]-[Q^a_L,Q^b_L]\nonumber\\ &=&if_{abc}Q^c_R-if_{abc}Q^c_L= if_{abc}Q^c_A.\end{aligned}$$ Note that these charge operators do [*not*]{} form a closed algebra, i.e., the commutator of two axial charge operators is not again an axial charge operator. Since the parity doubling is not observed for the low-lying states, one assumes that the $Q_A^a$ do [*not*]{} annihilate the ground state, $$\label{4:1:qav} Q^a_A|0\rangle\neq 0,$$ i.e., the ground state of QCD is not invariant under “axial” transformations. According to Goldstone’s theorem , to each axial generator $Q^a_A$, which does not annihilate the ground state, corresponds a massless Goldstone boson field $\phi^a(x)$ with spin 0, whose symmetry properties are tightly connected to the generator in question. The Goldstone bosons have the same transformation behavior under parity, $$\label{4:1:parityphi} \phi^a(\vec{x},t)\stackrel{P}{\mapsto}-\phi^a(-\vec{x},t),$$ i.e., they are pseudoscalars, and transform under the subgroup $H=\mbox{SU(3)}_V$, which leaves the vacuum invariant, as an octet \[see Eq. (\[4:1:crqva\])\]: $$\label{4:1:transformationphiqv} [Q^a_V,\phi^b(x)]=if_{abc}\phi^c(x).$$ In the present case, $G=\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ with $n_G=16$ and $H=\mbox{SU(3)}_V$ with $n_H=8$ and we expect eight Goldstone bosons. ### The Scalar Quark Condensate $\langle \bar{q}q\rangle$ {#subsec_sqc} In the following, we will show that a non-vanishing scalar quark condensate in the chiral limit is a sufficient (but not a necessary) condition for a spontaneous symmetry breaking in QCD.[^42] The subsequent discussion will parallel that of the toy model in Sec. \[sec\_gt\] after replacement of the elementary fields $\Phi_i$ by appropriate composite Hermitian operators of QCD. Let us first recall the definition of the nine scalar and pseudoscalar quark densities: $$\begin{aligned} \label{4:1:sqd} S_a(y)&=&\bar{q}(y)\lambda_a q(y), \quad a=0,\cdots,8,\\ \label{4:1:psqd} P_a(y)&=&i\bar{q}(y)\gamma_5\lambda_a q(y), \quad a=0,\cdots,8.\end{aligned}$$ The equal-time commutation relation of two quark operators of the form $A_i(x)=q^\dagger(x)\hat{A}_i q(x)$, where $\hat{A}_i$ symbolically denotes Dirac- and flavor matrices and a summation over color indices is implied, can compactly be written as \[see Eq. (\[2:3:fkf\])\] $$\label{4:1:comrel} [A_1(\vec{x},t),A_2(\vec{y},t)]=\delta^3(\vec{x}-\vec{y}) q^\dagger(x)[\hat{A}_1,\hat{A}_2]q(x).$$ With the definition $$Q_V^a(t)=\int d^3 x q^\dagger(\vec{x},t) \frac{\lambda^a}{2} q(\vec{x},t),$$ and using $$\begin{aligned} [\frac{\lambda_a}{2},\gamma_0\lambda_0]&=&0,\\ {[}\frac{\lambda_a}{2},\gamma_0\lambda_b]&=& \gamma_0 i f_{abc} \lambda_c,\end{aligned}$$ we see, after integration of Eq. (\[4:1:comrel\]) over $\vec{x}$, that the scalar quark densities of Eq. (\[4:1:sqd\]) transform under $\mbox{SU(3)}_V$ as a singlet and as an octet, respectively, $$\begin{aligned} \label{4:1:sitr} [Q^a_V(t),S_0(y)]&=&0,\quad a=1,\cdots,8,\\ \label{4:1:octr} {[Q^a_V(t),S_b(y)]}&=&i\sum_{c=1}^8f_{abc}S_c(y), \quad a,b=1,\cdots,8,\end{aligned}$$ with analogous results for the pseudoscalar quark densities. In the $\mbox{SU(3)}_V$ limit and, of course, also in the even more restrictive chiral limit, the charge operators in Eqs. (\[4:1:sitr\]) and (\[4:1:octr\]) are actually time independent.[^43] Using the relation $$\sum_{a,b=1}^8 f_{abc}f_{abd}=3\delta_{cd}$$ for the structure constants of SU(3), we re-express the octet components of the scalar quark densities as $$\label{4:1:soktett} S_a(y)=-\frac{i}{3}\sum_{b,c=1}^8f_{abc}[Q_V^b(t),S_c(y)],$$ which represents the analogue of Eq. (\[3:3:phi3zykl\]) in the discussion of Goldstone’s theorem. In the chiral limit the ground state is necessarily invariant under $\mbox{SU(3)}_V$ [@Vafa:tf], i.e., $Q_V^a|0\rangle=0$, and we obtain from Eq. (\[4:1:soktett\]) $$\label{4:1:saun} \langle 0|S_a(y)|0\rangle =\langle 0|S_a(0)|0\rangle \equiv\langle S_a\rangle =0,\quad a=1,\cdots,8,$$ where we made use of translational invariance of the ground state. In other words, the octet components of the scalar quark condensate [*must*]{} vanish in the chiral limit. From Eq. (\[4:1:saun\]), we obtain for $a=3$ $$\langle\bar{u}u\rangle-\langle\bar{d}d\rangle=0,$$ i.e. $\langle\bar{u}u\rangle=\langle\bar{d}d\rangle$ and for $a=8$ $$\langle\bar{u}u\rangle+\langle\bar{d}d\rangle -2\langle\bar{s}s\rangle=0,$$ i.e. $\langle\bar{u}u\rangle=\langle\bar{d}d\rangle= \langle\bar{s}s\rangle. $ Because of Eq. (\[4:1:sitr\]) a similar argument cannot be used for the singlet condensate, and if we assume a non-vanishing singlet scalar quark condensate in the chiral limit, we thus find using Eq. (\[4:1:saun\]) $$\label{4:1:cqc} 0\neq \langle \bar{q}q\rangle =\langle\bar{u}u+\bar{d}d+\bar{s}s\rangle =3\langle\bar{u}u\rangle = 3\langle\bar{d}d\rangle =3\langle \bar{s}s\rangle.$$ Finally, we make use of (no summation implied!) $$(i)^2 [\gamma_5 \frac{\lambda_a}{2},\gamma_0\gamma_5\lambda_a] =\lambda^2_a\gamma_0$$ in combination with $$\begin{aligned} \lambda_1^2=\lambda_2^2=\lambda_3^2&=& \left( \begin{array}{rrr} 1&0&0\\ 0&1&0\\ 0&0&0 \end{array} \right),\\ \lambda_4^2=\lambda_5^2&=& \left( \begin{array}{rrr} 1&0&0\\ 0&0&0\\ 0&0&1 \end{array} \right),\\ \lambda_6^2=\lambda_7^2&=& \left( \begin{array}{rrr} 0&0&0\\ 0&1&0\\ 0&0&1 \end{array} \right),\\ \lambda_8^2&=& \frac{1}{3} \left( \begin{array}{rrr} 1&0&0\\ 0&1&0\\ 0&0&4 \end{array} \right)\end{aligned}$$ to obtain $$\label{4:1:crqapsqd} i[Q_a^A(t), P_a(y)] = \left \{\begin{array}{cl} \bar{u}u+\bar{d}d, & a=1,2,3\\ \bar{u}u+\bar{s}s, & a=4,5\\ \bar{d}d+\bar{s}s, & a=6,7\\ \frac{1}{3}(\bar{u}u+\bar{d}d+4\bar{s}s), & a=8 \end{array} \right.$$ where we have suppressed the $y$ dependence on the right-hand side. We evaluate Eq. (\[4:1:crqapsqd\]) for a ground state which is invariant under $\mbox{SU(3)}_V$, assuming a non-vanishing singlet scalar quark condensate, $$\label{4:1:crqc} \langle 0|i[Q_a^A(t),P_a(y)]|0\rangle =\frac{2}{3}\langle\bar{q}q\rangle,\quad a=1,\cdots,8,$$ where, because of translational invariance, the right-hand side is independent of $y$. Inserting a complete set of states into the commutator of Eq. (\[4:1:crqc\]) yields, in complete analogy to Sec. \[sec\_gt\] \[see the discussion following Eq. (\[3:3:vq1phi2\])\] that both the pseudoscalar density $P_a(y)$ as well as the axial charge operators $Q^a_A$ must have a non-vanishing matrix element between the vacuum and massless one particle states $|\phi^b\rangle$. In particular, because of Lorentz covariance, the matrix element of the axial-vector current operator between the vacuum and these massless states, appropriately normalized, can be written as $$\label{4:1:acc} \langle 0|A^a_\mu(0)|\phi^b(p)\rangle=ip_\mu F_0 \delta^{ab},$$ where $F_0\approx 93$ MeV denotes the “decay” constant of the Goldstone bosons in the chiral limit. Assuming $Q_A^a|0\rangle\neq 0$, a non-zero value of $F_0$ is a necessary and sufficient criterion for spontaneous chiral symmetry breaking. On the other hand, because of Eq. (\[4:1:crqc\]) a non-vanishing scalar quark condensate $\langle \bar{q} q\rangle$ is a sufficient (but not a necessary) condition for a spontaneous symmetry breakdown in QCD. Table \[table:4:1:comparison\] contains a summary of the patterns of spontaneous symmetry breaking as discussed in Sec. \[sec\_gt\], the generalization of Sec. \[sec\_sbgcnas\] to the so-called O($N$) linear sigma model, and QCD. ------------------------ -------------------------- ------------------------------------ -------------------------------------- Sec. 3.3 O($N$) linear QCD sigma model Symmetry group $G$ of O(3) O($N$) $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ the Lagrangian density Number of 3 $N(N-1)/2$ 16 generators $n_G$ Symmetry group $H$ O(2) O($N-1$) SU(3)$_V$ of the ground state Number of 1 $(N-1)(N-2)/2$ 8 generators $n_H$ Number of 2 $N-1$ 8 Goldstone bosons $n_G-n_H$ Multiplet of $(\Phi_1(x),\Phi_2(x))$ $(\Phi_1(x),\cdots,\Phi_{N-1}(x))$ $i\bar{q}(x)\gamma_5\lambda_a q(x)$ Goldstone boson fields Vacuum expectation $v=\langle\Phi_3\rangle$ $v=\langle\Phi_N\rangle$ $v=\langle\bar{q}q\rangle$ value ------------------------ -------------------------- ------------------------------------ -------------------------------------- : \[table:4:1:comparison\]Comparison of spontaneous symmetry breaking. Transformation Properties of the Goldstone Bosons {#sec_tpgb} ------------------------------------------------- The purpose of this section is to discuss the transformation properties of the field variables describing the Goldstone bosons [@Weinberg:de; @Coleman:sm; @Callan:sn; @Balachandran:zj; @Leutwyler:1991mz]. We will need the concept of a [*nonlinear realization*]{} of a group in addition to a [*representation*]{} of a group which one usually encounters in Physics. We will first discuss a few general group-theoretical properties before specializing to QCD. ### General Considerations {#subsec_gc} Let us consider a physical system with a Hamilton operator $\hat{H}$ which is invariant under a compact Lie group $G$. Furthermore we assume the ground state of the system to be invariant under only a subgroup $H$ of $G$, giving rise to $n=n_G-n_H$ Goldstone bosons. Each of these Goldstone bosons will be described by an independent field $\phi_i$ which is a continuous real function on Minkowski space $M^4$.[^44] We collect these fields in an $n$-component vector $\Phi$ and define the vector space $$\label{4:2:m1} M_1\equiv\{\Phi:M^4\to R^n|\phi_i:M^4\to R\,\,\mbox{continuous}\}.$$ Our aim is to find a mapping $\varphi$ which uniquely associates with each pair $(g,\Phi)\in G\times M_1$ an element $\varphi(g,\Phi)\in M_1$ with the following properties: $$\begin{aligned} \label{4:2:condmap1} &&\varphi(e,\Phi)=\Phi\,\,\forall\,\,\Phi\in M_1,\, e\,\, \mbox{identity of}\,\, G,\\ \label{4:2:condmap2} &&\varphi(g_1,\varphi(g_2,\Phi))=\varphi(g_1 g_2,\Phi)\,\,\forall\,\, g_1,g_2\in G,\,\forall\,\Phi\in M_1.\end{aligned}$$ Such a mapping defines an [*operation*]{} of the group $G$ on $M_1$. The second condition is the so-called group-homomorphism property [@Balachandran:ab; @O'Raifeartaigh:vq; @Jones:ti]. The mapping will, in general, [*not*]{} define a [*representation*]{} of the group $G$, because we do not require the mapping to be linear, i.e., $\varphi(g,\lambda \Phi)\neq \lambda\varphi(g,\Phi)$. Let $\Phi=0$ denote the “origin” of $M_1$ [@Leutwyler:1991mz] which, in a theory containing Goldstone bosons only, loosely speaking corresponds to the ground state configuration. Since the ground state is supposed to be invariant under the subgroup $H$ we require the mapping $\varphi$ to be such that all elements $h\in H$ map the origin onto itself. In this context the subgroup $H$ is also known as the little group of $\Phi=0$. Given that such a mapping indeed exists, we need to verify for infinite groups that (see Chap. 2.4 of [@Jones:ti]): 1. $H$ is not empty, because the identity $e$ maps the origin onto itself. 2. If $h_1$ and $h_2$ are elements satisfying $\varphi(h_1,0)= \varphi(h_2,0)=0$, so does $\varphi(h_1 h_2,0)= \varphi(h_1,\varphi(h_2,0))=\varphi(h_1,0)=0$, i.e., because of the homomorphism property also the product $h_1 h_2\in H$. 3. For $h\in H$ we have $$\varphi(h^{-1},0)=\varphi(h^{-1},\varphi(h,0)) =\varphi(h^{-1}h,0)=\varphi(e,0).$$ i.e., $h^{-1}\in H$. Following Ref. [@Leutwyler:1991mz] we will establish a connection between the Goldstone boson fields and the set of all left cosets $\{gH|g\in G\}$ which is also referred to as the quotient $G/H$. For a subgroup $H$ of $G$ the set $gH=\{gh|h\in H\}$ defines the left coset of $g$ (with an analogous definition for the right coset) which is one element of $G/H$.[^45] For our purposes we need the property that cosets either completely overlap or are completely disjoint (see, e.g., [@Jones:ti]), i.e, the quotient is a set whose elements themselves are sets of group elements, and these sets are completely disjoint. Let us first show that for all elements of a given coset, $\varphi$ maps the origin onto the same vector in $R^n$: $$\varphi(gh,0)=\varphi(g,\varphi(h,0)) =\varphi(g,0)\,\,\forall\,\, g\in G\, \mbox{and}\, h\in H.$$ Secondly, the mapping is injective with respect to the cosets, which can be proven as follows. Consider two elements $g$ and $g'$ of $G$ where $g'\not\in g H$. We need to show $\varphi(g,0)\neq \varphi(g',0)$. Let us assume $\varphi(g,0)=\varphi(g',0)$: $$0=\varphi(e,0)=\varphi(g^{-1}g,0) =\varphi(g^{-1},\varphi(g,0)) =\varphi(g^{-1},\varphi(g',0))=\varphi(g^{-1}g',0).$$ However, this implies $g^{-1}g'\in H$ or $g'\in gH $ in contradiction to the assumption. Thus $\varphi(g,0)=\varphi(g',0)$ cannot be true. In other words, the mapping can be inverted on the image of $\varphi(g,0)$. The conclusion is that there exists an [*isomorphic mapping*]{} between the quotient $G/H$ and the Goldstone boson fields.[^46] Now let us discuss the transformation behavior of the Goldstone boson fields under an arbitrary $g\in G$ in terms of the isomorphism established above. To each $\Phi$ corresponds a coset $\tilde{g}H$ with appropriate $\tilde{g}$. Let $f=\tilde{g}h\in \tilde{g}H$ denote a representative of this coset such that $$\Phi=\varphi(f,0)=\varphi(\tilde{g}h,0).$$ Now apply the mapping $\varphi(g)$ to $\Phi$: $$\varphi(g,\Phi)=\varphi(g,\varphi(\tilde{g}h,0)) =\varphi(g\tilde{g}h,0)=\varphi(f',0)=\Phi',\quad f'\in g(\tilde{g}H).$$ In other words, in order to obtain the transformed $\Phi'$ from a given $\Phi$ we simply need to multiply the left coset $\tilde{g}H$ representing $\Phi$ by $g$ in order to obtain the new left coset representing $\Phi'$. This procedure uniquely determines the transformation behavior of the Goldstone bosons up to an appropriate choice of variables parameterizing the elements of the quotient $G/H$. ### Application to QCD {#subsec_aqcd} Now let us apply the above general considerations to the specific case relevant to QCD and consider the group $G=\mbox{SU($N$)}\times\mbox{SU($N$)}=\{(L,R)| L\in \mbox{SU($N$)}, R\in \mbox{SU($N$)}\}$ and $H=\{(V,V)|V\in \mbox{SU($N$)}\}$ which is isomorphic to $\mbox{SU($N$)}$. Let $\tilde{g}=(\tilde{L},\tilde{R})\in G$. We may uniquely characterize the left coset of $\tilde{g}$, $\tilde{g}H=\{(\tilde{L}V,\tilde{R}V)|V\in \mbox{SU($N$)}\}$, through the SU($N$) matrix $U=\tilde{R}\tilde{L}^\dagger$ [@Balachandran:zj], $$(\tilde{L}V,\tilde{R}V)=(\tilde{L}V,\tilde{R}\tilde{L}^\dagger\tilde{L}V) =(1,\tilde{R}\tilde{L}^\dagger)\underbrace{(\tilde{L}V,\tilde{L}V)}_{ \mbox{$\in H$}},\quad \mbox{i.e.} \quad \tilde{g}H=(1,\tilde{R}\tilde{L}^\dagger)H,$$ if we follow the convention that we choose the representative of the coset such that the unit matrix stands in its first argument. According to the above derivation, $U$ is isomorphic to a $\Phi$. The transformation behavior of $U$ under $g=(L,R)\in G$ is obtained by multiplication in the left coset: $$g \tilde{g}H=(L, R\tilde{R}\tilde{L}^\dagger)H= (1,R\tilde{R}\tilde{L}^\dagger L^\dagger)(L,L)H= (1,R(\tilde{R} \tilde{L}^\dagger)L^\dagger)H,$$ i.e.$$\label{4:2:utrafo} U=\tilde{R}\tilde{L}^\dagger \mapsto U'=R(\tilde{R}\tilde{L}^\dagger)L^\dagger =RUL^\dagger.$$ As mentioned above, we finally need to introduce an $x$ dependence so that $$\label{4:2:utrfafo} U(x)\mapsto R U(x) L^\dagger.$$ Let us now restrict ourselves to the physically relevant cases of $N=2$ and $N=3$ and define $$M_1\equiv\left\{ \begin{array}{l} \{\Phi: M^4\to R^3|\phi_i: M^4\to R\,\,\mbox{continuous}\}\,\,\mbox{for $N=2$,}\\ \{\Phi: M^4\to R^8|\phi_i: M^4 \to R\,\,\mbox{continuous}\}\,\,\mbox{for $N=3$}. \end{array} \right.$$ Furthermore let $\tilde{\cal H}(N)$ denote the set of all Hermitian and traceless $N\times N$ matrices, $$\tilde{\cal H}(N)\equiv\{A\in \mbox{gl}(N, C)|A^\dagger=A\wedge \mbox{Tr}(A)=0\},$$ which under addition of matrices defines a real vector space. We define a second set $M_2\equiv\{\phi:M^4\to\tilde{\cal H}(N)|\phi\,\, \mbox{continuous}\}$, where the entries are continuous functions. For $N=2$ the elements of $M_1$ and $M_2$ are related to each other according to $$\begin{aligned} \phi(x)&=&\sum_{i=1}^3\tau_i\phi_i(x) =\left(\begin{array}{cc} \phi_3 & \phi_1-i\phi_2\\ \phi_1+i\phi_2&-\phi_3 \end{array}\right) \equiv \left(\begin{array}{cc} \pi^0&\sqrt{2}\pi^+\\ \sqrt{2}\pi^-&-\pi^0 \end{array}\right),\\\end{aligned}$$ where the $\tau_i$ are the usual Pauli matrices and $\phi_i(x)=\frac{1}{2} \mbox{Tr}[\tau_i \phi(x)]$. Analogously for $N=3$, $$\begin{aligned} \phi(x)=\sum_{a=1}^8 \lambda_a \phi_a(x) &=&\left(\begin{array}{ccc} \phi_3+ \frac{1}{\sqrt{3}}\phi_8&\phi_1-i\phi_2&\phi_4-i\phi_5\\ \phi_1+i\phi_2& -\phi_3+ \frac{1}{\sqrt{3}}\phi_8&\phi_6-i\phi_7\\ \phi_4+i\phi_5&\phi_6+i\phi_7&-\frac{2}{\sqrt{3}}\phi_8 \end{array}\right)\\ &\equiv& \left(\begin{array}{ccc} \pi^0+\frac{1}{\sqrt{3}}\eta &\sqrt{2}\pi^+&\sqrt{2}K^+\\ \sqrt{2}\pi^-&-\pi^0+\frac{1}{\sqrt{3}}\eta&\sqrt{2}K^0\\ \sqrt{2}K^- &\sqrt{2}\bar{K}^0&-\frac{2}{\sqrt{3}}\eta \end{array}\right),\\\end{aligned}$$ with the Gell-Mann matrices $\lambda_a$ and $\phi_a(x)=\frac{1}{2}\mbox{Tr}[\lambda_a \phi(x)]$. Again, $M_2$ forms a real vector space. Let us finally define $$M_3\equiv\left\{U:M^4\to \mbox{SU}(N)|U(x) =\exp\left(i\frac{\phi(x)}{F_0}\right), \phi\in M_2\right\}.$$ At this point it is important to note that $M_3$ does not define a vector space because the sum of two SU($N$) matrices is not an SU($N$) matrix. We are now in the position to discuss the so-called nonlinear realization of $\mbox{SU($N$)}\times\mbox{SU($N$)}$ on $M_3$. The homomorphism $$\varphi: G\times M_3 \to M_3\quad\mbox{with}\quad \varphi[(L,R),U](x)\equiv R U (x)L^\dagger,$$ defines an operation of $G$ on $M_3$, because 1. $RUL^\dagger\in M_3$, since $U\in M_3$ and $R, L^\dagger\in \mbox{SU}(N)$. 2. $\varphi[(1_{N\times N},1_{N\times N}),U](x)=1_{N\times N}U(x) 1_{N\times N}=U(x).$ 3. Let $g_i=(L_i,R_i)\in G$ and thus $g_1 g_2=(L_1 L_2,R_1 R_2)\in G$. $$\begin{aligned} \varphi[g_1,\varphi[g_2,U]](x) &=&\varphi[g_1,(R_2 U L_2^\dagger)](x)=R_1 R_2 U(x) L_2^\dagger L_1^\dagger,\\ \varphi[g_1g_2,U](x)&=&R_1 R_2 U(x)(L_1 L_2)^\dagger= R_1 R_2 U(x) L_2^\dagger L_1^\dagger.\end{aligned}$$ The mapping $\varphi$ is called a nonlinear realization, because $M_3$ is [*not*]{} a vector space. The origin $\phi(x)=0$, i.e. $U_0=1$, denotes the ground state of the system. Under transformations of the subgroup $H=\{(V,V)|V\in \mbox{SU($N$)}\}$ corresponding to rotating both left- and right-handed quark fields in QCD by the same $V$, the ground state remains invariant, $$\varphi[g=(V,V),U_0]=VU_0 V^\dagger=V V^\dagger= 1=U_0.$$ On the other hand, under “axial transformations,” i.e. rotating the left-handed quarks by $A$ and the right-handed quarks by $A^\dagger$, the ground state does [*not*]{} remain invariant, $$\varphi[g=(A,A^\dagger),U_0]=A^\dagger U_0 A^\dagger=A^\dagger A^\dagger \neq U_0,$$ which, of course, is consistent with the assumed spontaneous symmetry breakdown. Let us finally discuss the transformation behavior of $\phi(x)$ under the subgroup $H=\{(V,V)|V\in \mbox{SU($N$)}\}$. Expanding $$U=1+i\frac{\phi}{F_0}-\frac{\phi^2}{2F_0^2}+\cdots,$$ we immediately see that the realization restricted to the subgroup $H$, $$\label{4:2:uhtrafo} 1+i\frac{\phi}{F_0}-\frac{\phi^2}{2F_0^2}+\cdots\mapsto V(1+i\frac{\phi}{F_0}-\frac{\phi^2}{2F_0^2}+\cdots)V^\dagger =1+i\frac{V \phi V^\dagger}{F_0}-\frac{V\phi V^\dagger V\phi V^\dagger}{2F_0^2} +\cdots,$$ defines a linear representation on $M_2 \ni \phi\mapsto V\phi V^\dagger \in M_2$, because $$\begin{aligned} && (V\phi V^\dagger)^\dagger= V\phi V^\dagger,\quad \mbox{Tr}(V\phi V^\dagger)=\mbox{Tr}(\phi)=0,\\ &&V_1 (V_2 \phi V_2^\dagger) V_1^\dagger=(V_1 V_2) \phi (V_1 V_2)^\dagger.\end{aligned}$$ Let us consider the SU(3) case and parameterize $$V=\exp\left(-i\Theta^V_a\frac{\lambda_a}{2}\right),$$ from which we obtain, by comparing both sides of Eq. (\[4:2:uhtrafo\]), $$\label{4:2:phihtrafo} \phi=\lambda_b\phi_b\stackrel{\mbox{$h\in$ SU(3)$_V$}}{\mapsto} V\phi V^\dagger=\phi-i\Theta^V_a \underbrace{[\frac{\lambda_a}{2},\phi_b\lambda_b]}_{ \mbox{$\phi_b if_{abc}\lambda_c$}} +\cdots =\phi+f_{abc}\Theta^V_a\phi_b\lambda_c +\cdots.$$ However, this corresponds exactly to the adjoint representation, i.e., in SU(3) the fields $\phi_a$ transforms as an octet which is also consistent with the transformation behavior we discussed in Eq.(\[4:1:transformationphiqv\]): $$\begin{aligned} \label{4:2:phivergf} e^{i\Theta^V_a Q^a_V}\lambda_b\phi_b e^{-i\Theta^V_a Q^a_V} &=&\lambda_b\phi_b +i\Theta^V_a\lambda_b\underbrace{[Q^a_V,\phi_b]}_{ \mbox{$if_{abc}\phi_c$}}+\cdots\nonumber\\ &=&\phi+f_{abc}\Theta^V_a\phi_b\lambda_c +\cdots.\end{aligned}$$ For group elements of $G$ of the form $(A,A^\dagger)$ one may proceed in a completely analogous fashion. However, one finds that the fields $\phi_a$ do [*not*]{} have a simple transformation behavior under these group elements. In other words, the commutation relations of the fields with the [*axial*]{} charges are complicated nonlinear functions of the fields [@Weinberg:de]. The Lowest-Order Effective Lagrangian {#sec_loel} ------------------------------------- Our goal is the construction of the most general theory describing the dynamics of the Goldstone bosons associated with the spontaneous symmetry breakdown in QCD. In the chiral limit, we want the effective Lagrangian to be invariant under $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$. It should contain exactly eight pseudoscalar degrees of freedom transforming as an octet under the subgroup $H=\mbox{SU(3)}_V$. Moreover, taking account of spontaneous symmetry breaking, the ground state should only be invariant under $\mbox{SU(3)}_V\times\mbox{U(1)}_V$. Following the discussion of Sec. \[subsec\_aqcd\] we collect the dynamical variables in the SU(3) matrix $U(x)$, $$\begin{aligned} \label{4:3:upar} U(x)&=&\exp\left(i\frac{\phi(x)}{F_0}\right),\nonumber\\ \phi(x)&=&\sum_{a=1}^8 \lambda_a \phi_a(x)\equiv \left(\begin{array}{ccc} \pi^0+\frac{1}{\sqrt{3}}\eta &\sqrt{2}\pi^+&\sqrt{2}K^+\\ \sqrt{2}\pi^-&-\pi^0+\frac{1}{\sqrt{3}}\eta&\sqrt{2}K^0\\ \sqrt{2}K^- &\sqrt{2}\bar{K}^0&-\frac{2}{\sqrt{3}}\eta \end{array}\right).\end{aligned}$$ The most general, chirally invariant, effective Lagrangian density with the minimal number of derivatives reads $$\label{4:3:l2} {\cal L}_{\rm eff} =\frac{F^2_0}{4}\mbox{Tr}\left(\partial_\mu U \partial^\mu U^\dagger \right),$$ where $F_0\approx 93$ MeV is a free parameter which later on will be related to the pion decay $\pi^+\to\mu^+\nu_\mu$ (see Sec.\[subsec\_pdpmn\]). First of all, the Lagrangian is invariant under the [*global*]{} $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ transformations of Eq. (\[4:2:utrafo\]): $$\begin{aligned} U&\mapsto& R U L^\dagger,\\ \partial_\mu U&\mapsto&\partial_\mu(R U L^\dagger)= \underbrace{\partial_\mu R}_{\mbox{0}}UL^\dagger+R\partial_\mu U L^\dagger +RU\underbrace{\partial_\mu L^\dagger}_{\mbox{0}}=R\partial_\mu U L^\dagger,\\ U^\dagger&\mapsto& L U^\dagger R^\dagger,\\ \partial_\mu U^\dagger&\mapsto&L\partial_\mu U^\dagger R^\dagger,\end{aligned}$$ because $${\cal L}_{\rm eff} \mapsto\frac{F^2_0}{4}\mbox{Tr}\Big(R\partial_\mu U \underbrace{L^\dagger L}_{\mbox{1}}\partial^\mu U^\dagger R^\dagger\Big) =\frac{F^2_0}{4}\mbox{Tr}\Big(\underbrace{R^\dagger R}_{\mbox{1}} \partial_\mu U \partial^\mu U^\dagger\Big) ={\cal L}_{\rm eff},$$ where we made use of the trace property $\mbox{Tr}(AB)=\mbox{Tr}(BA)$. The global $\mbox{U(1)}_V$ invariance is trivially satisfied, because the Goldstone bosons have baryon number zero, thus transforming as $\phi\mapsto\phi$ under $\mbox{U(1)}_V$ which also implies $U \mapsto U$. The substitution $\phi_a(\vec{x},t)\mapsto -\phi_a(\vec{x},t)$ or, equivalently, $U(\vec{x},t)\mapsto U^\dagger(\vec{x},t)$ provides a simple method of testing, whether an expression is of so-called even or odd [*intrinsic*]{} parity,[^47] i.e., even or odd in the number of Goldstone boson fields. For example, it is easy to show, using the trace property, that the Lagrangian of Eq. (\[4:3:l2\]) is even. The purpose of the multiplicative constant $F^2_0/4$ in Eq. (\[4:3:l2\]) is to generate the standard form of the kinetic term $\frac{1}{2}\partial_\mu \phi_a\partial^\mu \phi_a$, which can be seen by expanding the exponential $U=1+i\phi/F_0+\cdots$, $\partial_\mu U=i\partial_\mu\phi/F_0 +\cdots$, resulting in $$\begin{aligned} {\cal L}_{\rm eff}&=& \frac{F^2_0}{4}\mbox{Tr}\left[\frac{i\partial_\mu\phi}{F_0} \left(-\frac{i\partial^\mu\phi}{F_0}\right)\right]+\cdots =\frac{1}{4}\mbox{Tr}(\lambda_a\partial_\mu\phi_a\lambda_b\partial^\mu \phi_b)+\cdots\\ &=&\frac{1}{4}\partial_\mu\phi_a\partial^\mu \phi_b\mbox{Tr}(\lambda_a\lambda_b)+\cdots =\frac{1}{2}\partial_\mu\phi_a\partial^\mu\phi_a +{\cal L}_{\rm int},\end{aligned}$$ where we made use of $\mbox{Tr}(\lambda_a \lambda_b)=2\delta_{ab}$. In particular, since there are no other terms containing only two fields (${\cal L}_{\rm int}$ starts with interaction terms containing at least four Goldstone bosons) the eight fields $\phi_a$ describe eight independent [*massless*]{} particles.[^48] A term of the type $\mbox{Tr}[(\partial_\mu\partial^\mu U) U^\dagger]$ may be re-expressed as[^49] $$\mbox{Tr}[(\partial_\mu\partial^\mu U) U^\dagger] =\partial_\mu[\mbox{Tr}(\partial^\mu U U^\dagger)] -\mbox{Tr}(\partial^\mu U \partial_\mu U^\dagger),$$ i.e., up to a total derivative it is proportional to the Lagrangian of Eq. (\[4:3:l2\]). However, in the present context, total derivatives do not have a dynamical significance, i.e. they leave the equations of motion unchanged and can thus be dropped. The product of two invariant traces is excluded at lowest order, because $\mbox{Tr}(\partial_\mu U U^\dagger)=0$. Let us prove the general SU($N$) case by considering an SU($N$)-valued field $$U=\exp\left(i \frac{\Lambda_a\phi_a(x)}{F_0}\right),$$ with $N^2-1$ Hermitian, traceless matrices $\Lambda_a$ and real fields $\phi_a(x)$. Defining $\Phi=\Lambda_a\phi_a/F_0$, we expand the exponential $$U=1+i\Phi+\frac{1}{2}(i\Phi)^2+\frac{1}{3!}(i\Phi)^3+\cdots$$ and consider the derivative[^50] $$\partial_\mu U=i\partial_\mu \Phi+\frac{1}{2}(i\partial_\mu \Phi i\Phi +i\Phi i\partial_\mu \Phi)+\frac{1}{3!}[i\partial_\mu\Phi (i\Phi)^2+i\Phi i\partial_\mu\Phi i\Phi+(i\Phi)^2i\partial_\mu\Phi]+ \cdots.$$ We then find $$\begin{aligned} \label{4:3:trpmuuud} \mbox{Tr}(\partial_\mu U U^\dagger) &=&\mbox{Tr}[i\partial_\mu\Phi U^\dagger +\frac{1}{2}(i\partial_\mu \Phi i\Phi+i\Phi i\partial_\mu\Phi)U^\dagger +\cdots]\nonumber\\ &=& \mbox{Tr}[i\partial_\mu \Phi U^\dagger +i\partial_\mu \Phi i\Phi U^\dagger +\frac{1}{2}i\partial_\mu\Phi (i\Phi)^2 U^\dagger +\cdots]\nonumber\\ &=&\mbox{Tr}(i\partial_\mu \Phi \underbrace{U U^\dagger}_{\mbox{1}})=\mbox{Tr}(i\partial_\mu\Phi)= i\partial_\mu \phi_a\underbrace{\mbox{Tr}(\Lambda_a)}_{\mbox{0}}=0,\end{aligned}$$ where we made use of $[\Phi, U^\dagger]=0$. Let us turn to the vector and axial-vector currents associated with the global $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ symmetry of the effective Lagrangian of Eq. (\[4:3:l2\]). To that end, we parameterize $$\begin{aligned} \label{4:3:l} L&=&\exp\left(-i\Theta^L_a\frac{\lambda_a}{2}\right),\\ \label{4:3:r} R&=&\exp\left(-i\Theta^R_a\frac{\lambda_a}{2}\right).\end{aligned}$$ In order to construct $J^{\mu,a}_L$, set $\Theta^R_a=0$ and choose $\Theta^L_a=\Theta^L_a(x)$ (see Sec. \[subsec\_nt\]). Then, to first order in $\Theta^L_a$, $$\begin{aligned} \label{dul} U&\mapsto& U'=R U L^\dagger=U\left(1+i\Theta^L_a\frac{\lambda_a}{2}\right), \nonumber\\ U^\dagger&\mapsto&U'^\dagger= \left(1-i\Theta^L_a\frac{\lambda_a}{2}\right)U^\dagger, \nonumber\\ \partial_\mu U&\mapsto&\partial_\mu U' =\partial_\mu U \left(1+i\Theta^L_a\frac{\lambda_a}{2}\right) +U i\partial_\mu\Theta_a^L\frac{\lambda_a}{2}, \nonumber\\ \partial_\mu U^\dagger&\mapsto&\partial_\mu U'^\dagger =\left(1-i\Theta_a^L\frac{\lambda_a}{2}\right)\partial_\mu U^\dagger -i\partial_\mu \Theta_a^L\frac{\lambda_a}{2} U^\dagger,\end{aligned}$$ from which we obtain for $\delta \cal L_{\rm eff}$: $$\begin{aligned} \label{4:3:dll} \delta{\cal L}_{\rm eff}&=&\frac{F^2_0}{4}\mbox{Tr}\left[ U i\partial_\mu \Theta_a^L\frac{\lambda_a}{2}\partial^\mu U^\dagger +\partial_\mu U \left(-i\partial^\mu \Theta_a^L\frac{\lambda_a}{2}U^\dagger \right)\right] \nonumber\\ &=& \frac{F^2_0}{4}i\partial_\mu \Theta^L_a\mbox{Tr}\left[ \frac{\lambda_a}{2}(\partial^\mu U^\dagger U-U^\dagger\partial^\mu U) \right]\nonumber\\ &=&\frac{F^2_0}{4}i\partial_\mu \Theta^L_a\mbox{Tr}\left(\lambda_a \partial^\mu U^\dagger U\right).\end{aligned}$$ (In the last step we made use of $$\partial^\mu U^\dagger U=-U^\dagger \partial^\mu U,$$ which follows from differentiating $U^\dagger U=1$.) We thus obtain for the left currents $$\label{4:3:jl} J^{\mu,a}_L=\frac{\partial \delta {\cal L}_{\rm eff}}{ \partial \partial_\mu \Theta_a^L} =i\frac{F^2_0}{4}\mbox{Tr}\left(\lambda_a \partial^\mu U^\dagger U\right),$$ and, completely analogously, choosing $\Theta_a^L=0$ and $\Theta_a^R=\Theta_a^R(x)$, $$\label{4:3:jr} J^{\mu,a}_R=\frac{\partial \delta {\cal L}_{\rm eff} }{\partial \partial_\mu \Theta^R_a} =-i\frac{F^2_0}{4}\mbox{Tr}\left(\lambda_a U \partial^\mu U^\dagger\right)$$ for the right currents. Combining Eqs. (\[4:3:jl\]) and (\[4:3:jr\]) the vector and axial-vector currents read $$\begin{aligned} \label{4:3:jv} J^{\mu,a}_V&=&J^{\mu,a}_R+J^{\mu,a}_L=-i\frac{F^2_0}{4} \mbox{Tr}\left(\lambda_a[U,\partial^\mu U^\dagger]\right),\\ \label{4:3:ja} J^{\mu,a}_A&=&J^{\mu,a}_R-J^{\mu,a}_L=-i\frac{F^2_0}{4} \mbox{Tr}\left(\lambda_a\{U,\partial^\mu U^\dagger\}\right).\end{aligned}$$ Furthermore, because of the symmetry of ${\cal L}_{\rm eff}$ under $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$, both vector and axial-vector currents are conserved. The vector current densities $J^{\mu,a}_V$ of Eq. (\[4:3:jv\]) contain only terms with an even number of Goldstone bosons, $$\begin{aligned} J^{\mu,a}_V&\stackrel{\mbox{$\phi\mapsto-\phi$}}{\mapsto}& -i\frac{F^2_0}{4}\mbox{Tr}[\lambda_a(U^\dagger\partial^\mu U-\partial^\mu U U^\dagger)]\\ &=&-i\frac{F^2_0}{4}\mbox{Tr}[\lambda_a(-\partial^\mu U^\dagger U +U \partial^\mu U^\dagger)] =J^{\mu,a}_V.\end{aligned}$$ On the other hand, the expression for the axial-vector currents is [*odd*]{} in the number of Goldstone bosons, $$\begin{aligned} J^{\mu,a}_A&\stackrel{\mbox{$\phi\mapsto-\phi$}}{\mapsto}& -i\frac{F^2_0}{4}\mbox{Tr}[\lambda_a(U^\dagger\partial^\mu U+\partial^\mu U U^\dagger)]\\ &=&i\frac{F^2_0}{4}\mbox{Tr}[\lambda_a(\partial^\mu U^\dagger U +U \partial^\mu U^\dagger)] =-J^{\mu,a}_A.\end{aligned}$$ To find the leading term let us expand Eq. (\[4:3:ja\]) in the fields, $$J^{\mu,a}_A=-i\frac{F^2_0}{4}\mbox{Tr}\left(\lambda_a\left\{1+\cdots, -i\frac{\lambda_b\partial^\mu \phi_b}{F_0}+\cdots\right\}\right)= -F_0\partial^\mu\phi_a+\cdots$$ from which we conclude that the axial-vector current has a non-vanishing matrix element when evaluated between the vacuum and a one-Goldstone boson state \[see Eq. (\[4:1:acc\])\]: $$\begin{aligned} \langle 0|J^{\mu,a}_A(x)|\phi^b(p)\rangle &=&\langle 0|-F_0\partial^\mu\phi_a(x)|\phi^b(p)\rangle\\ &=&-F_0\partial^\mu \exp(-ip\cdot x)\delta^{ab} =ip^\mu F_0\exp(-ip\cdot x)\delta^{ab}.\end{aligned}$$ In Sec. \[subsec\_pdpmn\] $F_0$ will be related to the pion-decay constant entering $\pi^+\to\mu^+\nu_\mu$. So far we have assumed a perfect $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ symmetry. However, in Sec. \[sec\_esbfl\] we saw, by means of a simple example, how an explicit symmetry breaking may lead to finite masses of the Goldstone bosons. As has been discussed in Sec. \[subsec\_csbdqm\], the quark mass term of QCD results in such an explicit symmetry breaking, $$\label{4:3:qmt} {\cal L}_M=-\bar{q}_RM q_L-\bar{q}_L M^\dagger q_R,\quad M=\left(\begin{array}{ccc}m_u&0&0\\0&m_d&0\\0&0&m_s\end{array}\right).$$ In order to incorporate the consequences of Eq. (\[4:3:qmt\]) into the effective-Lagrangian framework, one makes use of the following argument [@Georgi]: Although $M$ is in reality just a constant matrix and does not transform along with the quark fields, ${\cal L}_M$ of Eq. (\[4:3:qmt\]) [*would be*]{} invariant [*if*]{} $M$ transformed as $$\label{4:3:mgtrafo} M\mapsto R M L^\dagger.$$ One then constructs the most general Lagrangian ${\cal L}(U,M)$ which is invariant under Eqs. (\[4:2:utrfafo\]) and (\[4:3:mgtrafo\]) and expands this function in powers of $M$. At lowest order in $M$ one obtains $$\label{4:3:lqm} {\cal L}_{\rm s.b.}=\frac{F^2_0 B_0}{2}\mbox{Tr}(MU^\dagger+UM^\dagger),$$ where the subscript s.b. refers to symmetry breaking. In order to interpret the new parameter $B_0$ let us consider the energy density of the ground state ($U=U_0=1$), $$\label{4:3:Heff} \langle{\cal H}_{\rm eff}\rangle=-F_0^2 B_0(m_u+m_d+m_s),$$ and compare its derivative with respect to (any of) the light quark masses $m_q$ with the corresponding quantity in QCD, $$\left.\frac{\partial \langle 0|{\cal H}_{\rm QCD}|0\rangle}{\partial m_q} \right|_{m_u=m_d=m_s=0}=\frac{1}{3}\langle 0|\bar{q}{q}|0\rangle_0 =\frac{1}{3}\langle \bar{q}q\rangle,$$ where $\langle\bar{q}{q}\rangle$ is the chiral quark condensate of Eq. (\[4:1:cqc\]). Within the framework of the lowest-order effective Lagrangian, the constant $B_0$ is thus related to the chiral quark condensate as $$\label{4:3:b0} 3 F^2_0B_0=-\langle\bar{q}q\rangle.$$ Let us add a few remarks. 1. A term $\mbox{Tr}(M)$ by itself is not invariant. 2. The combination $\mbox{Tr}(MU^\dagger-U M^\dagger)$ has the wrong behavior under parity $\phi(\vec{x},t)\mapsto-\phi(-\vec{x},t)$, because $$\begin{aligned} \mbox{Tr}[M U^\dagger(\vec{x},t)-U(\vec{x},t)M^\dagger] &\stackrel{P}{\mapsto} &\mbox{Tr}[M U(-\vec{x},t)-U^\dagger(-\vec{x},t)M^\dagger] \\ &\stackrel{M=M^\dagger}{=}&-\mbox{Tr}[M U^\dagger(-\vec{x},t)- U(-\vec{x},t) M^\dagger].\end{aligned}$$ 3. Because $M=M^\dagger$, ${\cal L}_{\rm s.b.}$ contains only terms even in $\phi$. In order to determine the masses of the Goldstone bosons, we identify the terms of second order in the fields in ${\cal L}_{\rm s.b.}$, $$\label{4:3:lmzo} {\cal L}_{\rm s.b}=-\frac{B_0}{2}\mbox{Tr}(\phi^2M) +\cdots.$$ Using Eq. (\[4:3:upar\]) we find $$\begin{aligned} \mbox{Tr}(\phi^2M) &=&2(m_u+m_d)\pi^+\pi^- +2(m_u+m_s)K^+ K^- +2(m_d+m_s)K^0\bar{K}^0\\ &&+(m_u+m_d)\pi^0\pi^0 +\frac{2}{\sqrt{3}}(m_u-m_d)\pi^0\eta +\frac{m_u+m_d+4m_s}{3}\eta^2.\end{aligned}$$ For the sake of simplicity we consider the isospin-symmetric limit $m_u=m_d=m$ so that the $\pi^0\eta$ term vanishes and there is no $\pi^0$-$\eta$ mixing. We then obtain for the masses of the Goldstone bosons, to lowest order in the quark masses, $$\begin{aligned} \label{4:3:mpi2} M^2_\pi&=&2 B_0 m,\\ \label{4:3:mk2} M^2_K&=&B_0(m+m_s),\\ \label{4:3:meta2} M^2_\eta&=&\frac{2}{3} B_0\left(m+2m_s\right).\end{aligned}$$ These results, in combination with Eq. (\[4:3:b0\]), $B_0=-\langle\bar{q}q\rangle/(3 F_0^2)$, correspond relations obtained in Ref. [@Gell-Mann:rz] and are referred to as the Gell-Mann, Oakes, and Renner relations. Furthermore, the masses of Eqs. (\[4:3:mpi2\]) - (\[4:3:meta2\]) satisfy the Gell-Mann-Okubo relation $$\label{4:3:gmof} 4M^2_K=4B_0(m+m_s)=2B_0(m+2m_s)+2B_0m=3M^2_\eta+M^2_\pi$$ independent of the value of $B_0$. Without additional input regarding the numerical value of $B_0$, Eqs. (\[4:3:mpi2\]) - (\[4:3:meta2\]) do not allow for an extraction of the absolute values of the quark masses $m$ and $m_s$, because rescaling $B_0\to \lambda B_0$ in combination with $m_q\to m_q/\lambda$ leaves the relations invariant. For the ratio of the quark masses one obtains, using the empirical values of the pseudoscalar octet, $$\begin{aligned} \frac{M^2_K}{M^2_\pi}=\frac{m+m_s}{2m}&\Rightarrow&\frac{m_s}{m}=25.9, \nonumber\\ \frac{M^2_\eta}{M^2_\pi}=\frac{2m_s+m}{3m}&\Rightarrow&\frac{m_s}{m}=24.3.\end{aligned}$$ Let us conclude this section with the following remark. We saw in Sec. \[subsec\_sqc\] that a non-vanishing quark condensate in the chiral limit is a sufficient but not a necessary condition for a spontaneous chiral symmetry breaking. The effective Lagrangian term of Eq. (\[4:3:lqm\]) not only results in a shift of the vacuum energy but also in finite Goldstone boson masses.[^51] These are related via the parameter $B_0$ and we recall that it was a symmetry argument which excluded a term $\mbox{Tr}(M)$ which, at leading order in $M$, would decouple the vacuum energy shift from the Goldstone boson masses. The scenario underlying ${\cal L}_{\rm s.b.}$ of Eq. (\[4:3:lqm\]) is similar to that of a Heisenberg ferromagnet [@Ashcroft; @Leutwyler:1991mz] which exhibits a spontaneous magnetization $\langle \vec{M}\rangle$, breaking the O(3) symmetry of the Heisenberg Hamiltonian down to O(2). In the present case the analogue of the order parameter $\langle \vec{M}\rangle$ is the quark condensate $\langle \bar{q} q\rangle$. In the case of the ferromagnet, the interaction with an external magnetic field is given by $-\langle \vec{M}\rangle\cdot \vec{H}$, which corresponds to Eq. (\[4:3:Heff\]), with the quark masses playing the role of the external field $\vec{H}$. However, in principle, it is also possible that $B_0$ vanishes or is rather small. In such a case the quadratic masses of the Goldstone bosons might be dominated by terms which are nonlinear in the quark masses, i.e., by higher-order terms in the expansion of ${\cal L}(U,M)$. Such a scenario is the origin of the so-called generalized chiral perturbation theory [@Knecht:1995tr; @Knecht:1995ai; @Stern:1997]. The analogue would be an antiferromagnet which shows a spontaneous symmetry breaking but with $\langle \vec{M}\rangle=0$. The analysis of recent data on $K^+\to \pi^+\pi^-e^+\nu_e$ [@Pislak:2001bf] in terms of the isoscalar $s$-wave scattering length $a_0^0$ [@Colangelo:2001sp] supports the conjecture that the quark condensate is indeed the leading order parameter of the spontaneously broken chiral symmetry. For a recent discussion on the relation between the quark condensate and $s$-wave $\pi\pi$ scattering the interested reader is referred to Ref. [@Leutwyler:2001zc]. Effective Lagrangians and Weinberg’s Power Counting Scheme {#sec_elwpcs} ---------------------------------------------------------- An essential prerequisite for the construction of effective field theories is a “theorem” of Weinberg stating that a perturbative description in terms of the most general effective Lagrangian containing all possible terms compatible with assumed symmetry principles yields the most general $S$ matrix consistent with the fundamental principles of quantum field theory and the assumed symmetry principles [@Weinberg:1978kz]. The corresponding effective Lagrangian will contain an infinite number of terms with an infinite number of free parameters. Turning Weinberg’s theorem into a practical tool requires two steps: one needs some scheme to organize the effective Lagrangian and a systematic method of assessing the importance of diagrams generated by the interaction terms of this Lagrangian when calculating a physical matrix element. In the framework of mesonic chiral perturbation theory, the most general chiral Lagrangian describing the dynamics of the Goldstone bosons is organized as a string of terms with an increasing number of derivatives and quark mass terms, $$\label{4:4:ll2l4} {\cal L}_{\rm eff}={\cal L}_2 + {\cal L}_4 + {\cal L}_6 +\cdots,$$ where the subscripts refer to the order in the momentum and quark mass expansion. The index 2, for example, denotes either two derivatives or one quark mass term. In the context of Feynman rules, derivatives generate four-momenta, whereas the convention of counting quark mass terms as being of the same order as two derivatives originates from Eqs. (\[4:3:mpi2\]) - (\[4:3:meta2\]) in conjunction with the on-shell condition $p^2=M^2$. In an analogous fashion, ${\cal L}_4$ and ${\cal L}_6$ denote more complicated terms of so-called chiral orders ${\cal O}(p^4)$ and ${\cal O}(p^6)$ with corresponding numbers of derivatives and quark mass terms. With such a counting scheme, the chiral orders in the mesonic sector are always even \[${\cal O}(p^{2n})$\] because Lorentz indices of derivatives always have to be contracted with either the metric tensor $g^{\mu\nu}$ or the Levi-Civita tensor $\epsilon^{\mu\nu\rho\sigma}$ to generate scalars, and the quark mass terms are counted as ${\cal O}(p^2)$. Weinberg’s power counting scheme [@Weinberg:1978kz] analyzes the behavior of a given diagram under a linear rescaling of all the [*external*]{} momenta, $p_i\mapsto t p_i$, and a quadratic rescaling of the light quark masses, $m_q\mapsto t^2 m_q$, which, in terms of the Goldstone boson masses, corresponds to $M^2\mapsto t^2 M^2$. The chiral dimension $D$ of a given diagram with amplitude ${\cal M}(p_i,m_q)$ is defined by $$\label{4:4:mr1} {\cal M}(tp_i, t^2 m_q)=t^D {\cal M}(p_i,m_q),$$ and thus $$\label{4:4:mr2} D=2+\sum_{n=1}^\infty2(n-1)N_{2n} +2N_L,$$ where $N_{2n}$ denotes the number of vertices originating from ${\cal L}_{2n}$, and $N_L$ is the number of independent loops. Clearly, for small enough momenta and masses diagrams with small $D$, such as $D=2$ or $D=4$, should dominate. Of course, the rescaling of Eq. (\[4:4:mr1\]) must be viewed as a mathematical tool. While external three-momenta can, to a certain extent, be made arbitrarily small, the rescaling of the quark masses is a theoretical instrument only. Note that loop diagrams are always suppressed due to the term $2N_L$ in Eq. (\[4:4:mr2\]). It may happen, though, that the leading-order tree diagrams vanish and therefore that the lowest-order contribution to a certain process is a one-loop diagram. An example is the reaction $\gamma\gamma\to\pi^0\pi^0$ [@Bijnens:1987dc]. In order to prove Eq. (\[4:4:mr2\]) we start from the usual Feynman rules for evaluating an $S$-matrix element (see, e.g., Appendix A-4 of Ref. [@Itzykson:rh]). Each internal meson line contributes a factor $$\begin{aligned} \label{4:4:intlines} \int\frac{d^4k}{(2\pi)^4} \frac{i}{k^2-M^2+i\epsilon} &\stackrel{\mbox{$(M^2\mapsto t^2 M^2)$}}{\mapsto}& t^{-2}\int\frac{d^4k}{(2\pi)^4} \frac{i}{k^2/t^2-M^2+i\epsilon}\nonumber\\ &\stackrel{\mbox{$(k=tl)$}}{=}& t^2 \int\frac{d^4l}{(2\pi)^4} \frac{i}{l^2-M^2+i\epsilon}.\end{aligned}$$ For each vertex, originating from ${\cal L}_{2n}$, we obtain symbolically a factor $p^{2n}$ together with a four-momentum conserving delta function resulting in $t^{2n}$ for the vertex factor and $t^{-4}$ for the delta function. At this point one has to take into account the fact that, although Eq. (\[4:4:mr1\]) refers to a rescaling of [*external*]{} momenta, a substitution $k=tl$ for internal momenta as in Eq. (\[4:4:intlines\]) acts in exactly the same way as a rescaling of external momenta: $$\begin{aligned} \delta^4(p+k)&\stackrel{\mbox{$p\mapsto tp, k=tl$}}{\mapsto}t^{-4} \delta^4(p+l),\\ p^{2n-m}k^m&\stackrel{\mbox{$p\mapsto tp, k=tl$}}{\mapsto}t^{2n}p^{2n-m}l^m,\end{aligned}$$ where $p$ and $k$ denote external and internal momenta, respectively. So far we have discussed the rules for determining the power $D_S$ referring to the $S$-matrix element which is related to the invariant amplitude through a four-momentum conserving delta function, $$S\sim \delta^4(P_f-P_i){\cal M}.$$ The delta function contains external momenta only, and thus re-scales under $p_i\mapsto tp_i$ as $t^{-4}$, so $$t^{D_S}=t^{-4}t^D.$$ We thus find as an intermediate result $$\label{4:4:d} D=4+2N_I+\sum_{n=1}^\infty N_{2n}(2n-4),$$ where $N_I$ denotes the number of internal lines. The number of independent loops, total number of vertices, and number of internal lines are related by[^52] $$N_L=N_I-(N_V-1),$$ because each of the $N_V$ vertices generates a delta function. After extracting one overall delta function this yields $N_V-1$ conditions for the internal momenta. Using $N_V=\sum_{n} N_{2n}$ we finally obtain from Eq. (\[4:4:d\]) $$D=4+2(N_L+N_V-1)+\sum_{n=1}^\infty N_{2n}(2n-4) =2+2 N_L +\sum_{n=1}^\infty N_{2n}(2n-2).\quad$$ By means of a simple example we will illustrate how the mechanism of rescaling actually works. To that end we consider as a toy model of an effective field theory the self interaction of a scalar field, $$\label{4:4:l2bsp} {\cal L}_2=g \Phi^2\partial_\mu\Phi\partial^\mu\Phi,$$ where the coupling constant $g$ has the dimension of $\mbox{energy}^{-2}$.[^53] The Feynman rules give the amplitude corresponding to the simple tree diagram of Fig. \[4:4:figexamplerescaling1\] for the scattering of two particles, $$\begin{aligned} \label{4:4:frl2bsp} {\cal M}(p_1,p_2,;p_3,p_4) &=&4ig \left[(p_1+p_2)\cdot(p_3+p_4)-p_1\cdot p_2-p_3\cdot p_4\right] \nonumber\\ &\stackrel{\mbox{$p_i\mapsto t p_i$}}{\mapsto}& t^2 {\cal M}(p_1,p_2;p_3,p_4).\end{aligned}$$ As expected, the behavior under rescaling is in agreement with Eq. (\[4:4:mr2\]) for $N_L=0$, $N_2=1$, and $N_{2n}=0$ for all remaining $n$. Now let us consider a typical loop diagram of Fig.  \[4:4:figexamplerescaling\] contributing to the same process, where the 2 in the interaction blob indicates the ${\cal L}_2$ term in the Lagrangian containing two derivatives. Applying the usual Feynman rules, with the vertex of Eq. (\[4:4:frl2bsp\]), we obtain $$\begin{aligned} \label{4:4:mbsp2} {\cal M}&=&\frac{1}{2} \int \frac{d^4k}{(2\pi)^4}\nonumber\\ &&\times4ig\left[ (p_1+p_2-k+k)\cdot(p_3+p_4)-(p_1+p_2-k)\cdot k-p_3\cdot p_4\right]\nonumber\\ &&\times\frac{i}{k^2-M^2+i\epsilon}\nonumber\\ &&\times\frac{i}{(p_1+p_2-k)^2-M^2+i\epsilon}\nonumber\\ &&\times 4ig\left[(p_1+p_2)\cdot(p_3+p_4-k+k)-p_1\cdot p_2-(p_1+p_2-k)\cdot k\right] \nonumber\\ &=& 8g^2\int \frac{d^4k}{(2\pi)^4} \left[(p_1+p_2)\cdot(p_3+p_4)-(p_1+p_2-k)\cdot k-p_3\cdot p_4\right] \nonumber\\ &&\times\frac{1}{k^2-M^2+i\epsilon} \frac{1}{(p_1+p_2-k)^2-M^2+i\epsilon}\nonumber\\ &&\times \left[(p_1+p_2)\cdot(p_3+p_4)-p_1\cdot p_2-(p_1+p_2-k)\cdot k\right] \nonumber\\ &\stackrel{\stackrel{p_i\mapsto tp_i}{M^2\mapsto t^2 M^2}}{\mapsto}& 8 g^2 \int \frac{d^4k}{(2\pi)^4}\left[ (p_1+p_2)\cdot(p_3+p_4)-(p_1+p_2-\frac{k}{t})\cdot \frac{k}{t}- p_3\cdot p_4\right] \nonumber\\ &&\times \frac{1}{\frac{k^2}{t^2}-M^2+i\epsilon} \frac{1}{(p_1+p_2-\frac{k}{t})^2-M^2+i\epsilon}\nonumber\\ &&\times \left[(p_1+p_2)\cdot(p_3+p_4)-p_1\cdot p_2-(p_1+p_2-\frac{k}{t}) \cdot \frac{k}{t}\right]\nonumber\\ &\stackrel{\mbox{$tl=k$}}{=} &8 g^2 \int \frac{t^4d^4l}{(2\pi)^4}\left[ (p_1+p_2)\cdot(p_3+p_4)-(p_1+p_2-l)\cdot l-p_3\cdot p_4\right] \nonumber\\ &&\times \frac{1}{l^2-M^2+i\epsilon} \frac{1}{(p_1+p_2-l)^2-M^2+i\epsilon}\nonumber\\ &&\times\left[(p_1+p_2)\cdot(p_3+p_4)-p_1\cdot p_2-(p_1+p_2-l)\cdot l\right] \nonumber\\ &=&t^4 {\cal M},\end{aligned}$$ This agrees with the value $D=4$ given by Eq. (\[4:4:mr2\]) for $N_L=1$ and $N_{2}=2$. For the sake of completeness, let us comment on the symmetry factor 1/2 in Eq. (\[4:4:mbsp2\]). When deriving the Feynman rule of Eq. (\[4:4:frl2bsp\]), we took account of $4!=24$ distinct combinations of contracting four field operators with four external lines. The “product” of two such vertices thus contains $24\times 24$ combinations. However, from each vertex two lines have to be selected as internal lines and there exist 6 possibilities to choose one pair out of 4 field operators to form internal lines. For the two remaining operators one has two possibilities of contracting them with external lines. Finally, the respective pairs of internal lines of the first and second vertices may be contracted in two ways with each other, leaving us with $12\times 12\times 2= (24\times 24)/2$ combinations. In the discussion of the loop integral we did not address the question of convergence. This needs to be addressed since applying the substitution $tl=k$ in Eq. (\[4:4:mbsp2\]) is well-defined only for convergent integrals. Later on we will regularize the integrals by use of the method of dimensional regularization, introducing a renormalization scale $\mu$ which also has to be rescaled linearly. However, at a given chiral order, the sum of all diagrams will, by construction, not depend on the renormalization scale. Finally, the proof of Weinberg’s theorem [@Leutwyler:1993iq; @D'Hoker:1994ti] for chiral perturbation theory is rather technical and lengthy and beyond the scope of this review. In Ref. [@Leutwyler:1993iq] it was shown that global symmetry constraints alone do not suffice to fully determine the low-energy structure of the effective Lagrangian. In fact, a determination of the (low-energy) Green functions of QCD off the mass shell, i.e., for momenta which do not correspond to the mass-shell conditions for Goldstone bosons, one needs to study the Ward identities, and therefore the symmetries have to be extended to the local level. One thus considers a [*locally*]{} invariant, effective Lagrangian although the symmetries of the underlying theory originate in a global symmetry. If the Ward identities contain anomalies, they show up as a modification of the generating functional, which can explicitly be incorporated through the Wess-Zumino-Witten construction [@Wess:yu; @Witten:tw]. Construction of the Effective Lagrangian {#sec_cel} ---------------------------------------- In Sec. \[sec\_loel\] we have derived the lowest-order effective Lagrangian for a [*global*]{} $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ symmetry. On the other hand, the Ward identities originating in the global $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ symmetry of QCD are obtained from a [*locally*]{} invariant generating functional involving a coupling to external fields (see Sec. \[subsec\_qcdpefgf\] and App. \[app\_gfwi\]). Our goal is to approximate the “true” generating functional $Z_{\rm QCD}[v,a,s,p]$ of Eq. (\[2:4:genfun\]) by a sequence $Z^{(2)}_{\rm eff}[v,a,s,p] + Z^{(4)}_{\rm eff}[v,a,s,p] +\cdots$, where the effective generating functionals are obtained using the effective field theory. Therefore, we need to promote the global symmetry of the effective Lagrangian to a local one and introduce a coupling to the [*same*]{} external fields $v$, $a$, $s$, and $p$ as in QCD. In the following we will outline the principles entering the construction of the effective Lagrangian for a local $G=\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ symmetry (see Refs. [@Fearing:1994ga; @Bijnens:1999sh; @Ebertshauser:2001nj] for details).[^54] The matrix $U$ transforms as $U \mapsto U'=V_R U V_L^{\dagger}$, where $V_L(x)$ and $V_R(x)$ are independent space-time-dependent SU(3) matrices. As in the case of gauge theories, we need external fields $l_\mu^a(x)$ and $r_\mu^a(x)$ \[see Eqs. (\[2:4:mch\]), (\[2:4:vrlarl\]), and (\[2:4:sg\]) and Table \[4:5:table\_trafprop\]\] corresponding to the parameters $\Theta^L_a(x)$ and $\Theta^R_a(x)$ of $V_L(x)$ and $V_R(x)$, respectively. For any object $A$ transforming as $V_R A V_L^\dagger$ such as, e.g., $U$ we define the covariant derivative $D_\mu A$ as $$\begin{aligned} \label{4:5:kaa} D_\mu A&\equiv&\partial_\mu A -i r_\mu A+iA l_\mu\nonumber\\ &\mapsto&\partial_\mu(V_R A V_L^\dagger)-i(V_R r_\mu V_R^\dagger+i V_R\partial_\mu V_R^\dagger)V_R A V_L^\dagger\nonumber\\ &&+iV_R A V_L^\dagger (V_L l_\mu V_L^\dagger +i V_L \partial_\mu V_L^\dagger) \nonumber\\ &=&\partial_\mu V_R A V_L^\dagger+ V_R\partial_\mu A V_L^\dagger +V_R A \partial_\mu V_L^\dagger-i V_R r_\mu A V_L^\dagger -\partial_\mu V_R A V_L^\dagger\nonumber\\ && +i V_R A l_\mu V_L^\dagger -V_R A \partial_\mu V_L^\dagger\nonumber\\ &=& V_R(\partial_\mu A -i r_\mu A+iA l_\mu)V_L^\dagger= V_R (D_\mu A) V_L^\dagger, \end{aligned}$$ where we made use of $V_R\partial_\mu V_R^\dagger=-\partial_\mu V_R V_R^\dagger$. Again, the defining property for the covariant derivative is that it should transform in the same way as the object it acts on.[^55] Since the effective Lagrangian will ultimately contain arbitrarily high powers of derivatives we also need the field strength tensors $f^L_{\mu\nu}$ and $f^R_{\mu\nu}$ corresponding to the gauge fields, $$\begin{aligned} \label{4:5:fr} f_{\mu\nu}^R&\equiv&\partial_\mu r_\nu-\partial_\nu r_\mu-i{[r_\mu,r_\nu]},\\ \label{4:5:fl} f_{\mu\nu}^L&\equiv&\partial_\mu l_\nu-\partial_\nu l_\mu-i{[l_\mu,l_\nu]}.\end{aligned}$$ The field strength tensors are traceless, $$\label{4:5:trflfr} \mbox{Tr}(f^L_{\mu\nu})=\mbox{Tr}(f^R_{\mu\nu})=0,$$ because $\mbox{Tr}(l_\mu)=\mbox{Tr}(r_\mu)=0$ and the trace of any commutator vanishes. Finally, following the convention of Gasser and Leutwyler we introduce the linear combination $\chi\equiv 2B_0(s+ip)$ with the scalar and pseudoscalar external fields of Eq. (\[2:4:mch\]), where $B_0$ is defined in Eq. (\[4:3:b0\]). Table \[4:5:table\_trafprop\] contains the transformation properties of all building blocks under the group ($G$), charge conjugation ($C$), and parity ($P$). ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- element $G$ $C$ $P$ ----------------------------------------- --------------------------------------------------------- --------------------------------------------- --------------------------------------------------- $U$ $V_R U V_L^\dagger$ $U^T$ $U^\dagger$ $D_{\lambda_1}\cdots D_{\lambda_n}U$ $V_R D_{\lambda_1}\cdots D_{\lambda_n}U V_L^\dagger$ $(D_{\lambda_1}\cdots D_{\lambda_n}U)^T$ $(D^{\lambda_1}\cdots D^{\lambda_n}U)^\dagger$ $\chi$ $V_R \chi V_L^\dagger$ $\chi^T$ $\chi^\dagger$ $D_{\lambda_1}\cdots D_{\lambda_n}\chi$ $V_R D_{\lambda_1}\cdots D_{\lambda_n}\chi V_L^\dagger$ $(D_{\lambda_1}\cdots D_{\lambda_n}\chi)^T$ $(D^{\lambda_1}\cdots D^{\lambda_n}\chi)^\dagger$ $r_\mu$ $V_R r_\mu V_R^\dagger+iV_R\partial_\mu $-l_\mu^T$ $l^\mu$ V^\dagger_R$ $l_\mu$ $V_L l_\mu V_L^\dagger+iV_L\partial_\mu $-r_\mu^T$ $r^\mu$ V^\dagger_L$ $f^R_{\mu\nu}$ $V_R f^R_{\mu\nu}V_R^\dagger$ $-(f_{\mu\nu}^L)^T$ $f_L^{\mu\nu}$ $f^L_{\mu\nu}$ $V_L f^L_{\mu\nu}V_L^\dagger$ $-(f_{\mu\nu}^R)^T$ $f_R^{\mu\nu}$ ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- : \[4:5:table\_trafprop\] Transformation properties under the group ($G$), charge conjugation ($C$), and parity ($P$). The expressions for adjoint matrices are trivially obtained by taking the Hermitian conjugate of each entry. In the parity transformed expression it is understood that the argument is $(-\vec{x},t)$ and that partial derivatives $\partial_{\mu}$ act with respect to $x$ and not with respect to the argument of the corresponding function. In the chiral counting scheme of chiral perturbation theory the elements are counted as: $$\label{4:5:powercounting} U = {\cal O}(p^0),\, D_{\mu} U = {\cal O}(p),\, r_{\mu},l_{\mu} = {\cal O}(p),\, f^{L/R}_{\mu\nu} = {\cal O}(p^2),\, \chi = {\cal O}(p^2).$$ The external fields $r_{\mu}$ and $l_{\mu}$ count as ${\cal O}(p)$ to match $\partial_\mu A$, and $\chi$ is of ${\cal O}(p^2)$ because of Eqs. (\[4:3:mpi2\]) - (\[4:3:meta2\]). Any additional covariant derivative counts as ${\cal O}(p)$. The construction of the effective Lagrangian in terms of the building blocks of Eq. (\[4:5:powercounting\]) proceeds as follows.[^56] Given objects $A,B,\dots$, all of which transform as one can form invariants by taking the trace of products of the type $A B^{\dagger}$: $$\begin{aligned} \mbox{Tr}(A B^\dagger)&\mapsto& \mbox{Tr}[V_R A V_L^\dagger (V_R B V_L^\dagger)^\dagger]=\mbox{Tr}(V_R A V_L^\dagger V_L B^\dagger V_R^\dagger)=\mbox{Tr}(A B^\dagger V_R^\dagger V_R)\\ &=&\mbox{Tr}(A B^\dagger).\end{aligned}$$ The generalization to more terms is obvious and, of course, the product of invariant traces is invariant: $$\label{4:5:ketten} \mbox{Tr}(AB^\dagger C D^\dagger),\quad \mbox{Tr}(A B^\dagger)\mbox{Tr}(C D^\dagger),\quad \cdots.$$ The complete list of elements up to and including order ${\cal O}(p^2)$ transforming as $V_R\cdots V_L^\dagger$ reads $$\label{4:5:lis} U, D_\mu U, D_\mu D_\nu U,\chi, U f_{\mu\nu}^L, f^R_{\mu\nu}U.$$ For the invariants up to ${\cal O}(p^2)$ we then obtain $$\begin{aligned} \label{4:5:invariants} {\cal O}(p^0)&:& \mbox{Tr}(UU^\dagger)=3,\nonumber\\ {\cal O}(p)&:& \mbox{Tr}(D_\mu U U^\dagger) \stackrel{\ast}{=}-\mbox{Tr}[U (D_\mu U)^\dagger] \stackrel{\ast}{=}0,\nonumber \\ {\cal O}(p^2)&:& \mbox{Tr}(D_\mu D_\nu U U^\dagger) \stackrel{\ast\ast}{=}-\mbox{Tr}[D_\nu U (D_\mu U)^\dagger] \stackrel{\ast\ast}{=}\mbox{Tr}[U(D_\nu D_\mu U)^\dagger],\nonumber\\ &&\mbox{Tr}(\chi U^\dagger),\nonumber\\ &&\mbox{Tr}(U \chi^\dagger),\nonumber\\ &&\mbox{Tr}(U f^L_{\mu\nu}U^\dagger)=\mbox{Tr}(f^L_{\mu\nu})=0,\nonumber\\ &&\mbox{Tr}(f^R_{\mu\nu})=0.\end{aligned}$$ In $\ast$ we made use of two important properties of the covariant derivative $D_\mu U$: $$\begin{aligned} \label{4:5:kauprop1} D_\mu U U^\dagger &=& -U (D_\mu U)^\dagger,\\ \label{4:5:kauprop2} \mbox{Tr}(D_\mu U U^\dagger)&=&0.\end{aligned}$$ The first relation results from the unitarity of $U$ in combination with the definition of the covariant derivative, Eq. (\[4:5:kaa\]). Equation (\[4:5:kauprop2\]) is shown using $\mbox{Tr}(r_\mu)=\mbox{Tr}(l_\mu)=0$ together with Eq. (\[4:3:trpmuuud\]), $\mbox{Tr}(\partial_\mu U U^\dagger)=0$: $$\begin{aligned} \mbox{Tr}(D_\mu U U^\dagger)&=&\mbox{Tr}(\partial_\mu U U^\dagger -ir_\mu U U^\dagger +iUl_\mu U^\dagger)=0.\end{aligned}$$ The relations $\ast\ast$ can either be verified by explicit calculation or, more elegantly, using the product rule of Ref. [@Fearing:1994ga] for the covariant derivatives. Finally, we impose Lorentz invariance, i.e., Lorentz indices have to be contracted, resulting in three candidate terms: $$\begin{aligned} \label{4:5:lis1} &&\mbox{Tr}[D_\mu U (D^\mu U)^\dagger],\\ \label{4:5:lis2} &&\mbox{Tr}(\chi U^\dagger\pm U \chi^\dagger).\end{aligned}$$ The term in Eq. (\[4:5:lis2\]) with the minus sign is excluded because it has the wrong sign under parity (see Table \[4:5:table\_trafprop\]), and we end up with the most general, [*locally*]{} invariant, effective Lagrangian at lowest chiral order,[^57] $$\label{4:5:l2} {\cal L}_2=\frac{F_0^2}{4}\mbox{Tr}[D_\mu U (D^\mu U)^\dagger] +\frac{F^2_0}{4}\mbox{Tr}(\chi U^\dagger + U\chi^\dagger).$$ Note that ${\cal L}_2$ contains two free parameters: the pion-decay constant $F_0$ and $B_0$ of Eq. (\[4:3:b0\]) (hidden in the definition of $\chi$). Let us finally derive the equations of motion associated with the lowest-order Lagrangian. These are important because they can be used to eliminate so-called equation-of-motion terms in the construction of the higher-order Lagrangians by applying field transformations [@Chisholm; @Kamefuchi:sb]. To that end we need to consider an infinitesimal change of the SU(3) matrix $U(x)$. Since the set of SU(3) matrices forms a group, to each pair of elements $U$ and $U'$ corresponds a unique element $\tilde{U}$, connecting the two via $U'=\tilde{U}U$. Let us parameterize $\tilde{U}$ by means of the Gell-Mann matrices, $$\label{4:5:utilde} \tilde{U}=\exp(i\Delta),\quad \Delta=\sum_{a=1}^8 \lambda_a \Delta_a, \quad \Delta_a\in R,$$ and consider small variations of the SU(3) matrix as $$\label{4:5:deltau} U'(x)=U(x)+\delta U(x)=\left(1+i\sum_{a=1}^8 \Delta_a(x)\lambda_a\right) U(x),$$ where the $\Delta_a(x)$ are now real functions. With such an ansatz, the matrix $U'$ satisfies both conditions $$\label{4:5:upcond} U' U'^\dagger= 1, \quad \mbox{det}(U')=1,$$ up to and including the terms linear in $\Delta_a(x)$.[^58] Given the fields at $t_1$ and $t_2$, the dynamics is determined by the principle of stationary action. We obtain for the variation of the action $$\begin{aligned} \label{4:5:deltas} \delta S&=&\frac{F^2_0}{4}\int_{t_1}^{t_2}dt\int d^3 x\, \mbox{Tr}\left[ D_\mu \delta U(D^\mu U)^\dagger +D_\mu U (D^\mu \delta U)^\dagger +\chi \delta U^\dagger +\delta U \chi^\dagger\right]\nonumber\\ &=&\frac{F^2_0}{4}\int_{t_1}^{t_2}dt\int d^3 x\,\mbox{Tr}\left[ -\delta U (D_\mu D^\mu U)^\dagger -D_\mu D^\mu U \delta U^\dagger +\chi \delta U^\dagger + \delta U \chi^\dagger \right]\nonumber\\ &=&i\frac{F^2_0}{4}\int_{t_1}^{t_2}dt \int d^3 x \sum_{a=1}^8 \Delta_a(x)\nonumber\\ && \times \mbox{Tr}\left\{\lambda_a[ D_\mu D^\mu U U^\dagger- U(D_\mu D^\mu U)^\dagger -\chi U^\dagger + U\chi^\dagger]\right\}.\end{aligned}$$ In the second equation we made use of the standard boundary conditions $\Delta_a(t_1,\vec{x})=\Delta_a(t_2,\vec{x})=0$, the divergence theorem, and the definition of the covariant derivative of Eq. (\[4:5:kaa\]). The third equality results from $\delta U^\dagger=-U^\dagger\delta U U^\dagger$ and the invariance of the trace with respect to cyclic permutations. The functions $\Delta_a(x)$ may be chosen arbitrarily, and we obtain eight Euler-Lagrange equations $$\label{4:5:eom8} \mbox{Tr}\left\{\lambda_a[D^2U U^\dagger - U (D^2 U)^\dagger -\chi U^\dagger + U\chi^\dagger]\right\}=0, \quad a=1,\cdots, 8.$$ Since any $3\times 3$ matrix $A$ can be written as $$\label{4:5:apar} A=a_0 1_{3\times 3}+\sum_{i=1}^8 a_i \lambda_i,\quad a_0=\frac{1}{3}\mbox{Tr}(A), \quad a_i=\frac{1}{2}\mbox{Tr}(\lambda_i A),$$ the eight equations of motion of Eq. (\[4:5:eom8\]) may compactly be written in matrix form[^59] $$\label{4:5:eom} {\cal O}_{\rm EOM}^{(2)}(U)\equiv D^2 U U^\dagger - U (D^2 U)^\dagger -\chi U^\dagger + U \chi^\dagger +\frac{1}{3}\mbox{Tr}(\chi U^\dagger- U \chi^\dagger)=0.$$ The additional term involving the trace is included to guarantee that the component proportional to the identity matrix vanishes identically and thus one does not erroneously generate a ninth equation of motion. Applications at Lowest Order {#sec_alo} ---------------------------- Let us consider two simple examples at lowest order $D=2$. According to Eq. (\[4:4:mr2\]) we only need to consider tree-level diagrams with vertices of ${\cal L}_2$. ### Pion Decay $\pi^+\to \mu^+\nu_\mu$ {#subsec_pdpmn} Our first example deals with the weak decay $\pi^+\to \mu^+\nu_\mu$ which will allow us to relate the free parameter $F_0$ of ${\cal L}_2$ to the pion-decay constant. At the level of the degrees of freedom of the Standard Model, pion decay is described by the annihilation of a $u$ quark and a $\bar{d}$ antiquark, forming the $\pi^+$, into a $W^+$ boson, propagation of the intermediate $W^+$, and creation of the leptons $\mu^+$ and $\nu_\mu$ in the final state (see Fig. \[4:6:piondecay\]). The coupling of the $W$ bosons to the leptons is given by $$\label{4:6:lwl} {\cal L}=-\frac{g}{2\sqrt{2}}\left[{\cal W}^+_\mu\bar{\nu}_\mu \gamma^\mu(1-\gamma_5)\mu+{\cal W}^-_\mu\bar{\mu}\gamma^\mu (1-\gamma_5)\nu_\mu\right],$$ whereas their interaction with the quarks forming the Goldstone bosons is effectively taken into account by inserting Eq. (\[2:4:rlw\]) into the Lagrangian of Eq. (\[4:5:l2\]). Let us consider the first term of Eq. (\[4:5:l2\]) and set $r_\mu=0$ with, at this point, still arbitrary $l_\mu$. Using $D_\mu U=\partial_\mu U+iUl_\mu$ we find $$\begin{aligned} \frac{F^2_0}{4}\mbox{Tr}[D_\mu U (D^\mu U)^\dagger] &=&\frac{F^2_0}{4}\mbox{Tr}[(\partial_\mu U +iU l_\mu) (\partial^\mu U^\dagger -i l^\mu U^\dagger)]\\ &=&\cdots +i\frac{F^2_0}{4}\mbox{Tr}(Ul_\mu \partial^\mu U^\dagger -l^\mu\underbrace{U^\dagger\partial_\mu U}_{\mbox{$-\partial_\mu U^\dagger U$}} )+\cdots\\ &=&i\frac{F^2_0}{2}\mbox{Tr}(l_\mu \partial^\mu U^\dagger U)+\cdots,\end{aligned}$$ where only the term linear in $l_\mu$ is shown. If we parameterize $$l_\mu =\sum_{a=1}^8\frac{\lambda_a}{2}l^a_\mu,$$ the interaction term linear in $l_\mu$ reads $$\label{4:6:lwli} {\cal L}_{\rm int}=\sum_{a=1}^8l_\mu^a\left[i \frac{F_0^2}{4}\mbox{Tr}( \lambda_a \partial^\mu U^\dagger U)\right]= \sum_{a=1}^8l^a_\mu J^{\mu,a}_L,$$ where we made use of Eq. (\[4:3:jl\]) defining $J^{\mu,a}_L$. Again, we expand $J^{\mu,a}_L$ by using Eq. (\[4:3:upar\]) to first order in $\phi$, $$\begin{aligned} \label{4:6:jlent} J^{\mu,a}_L&=& \frac{F_0}{2}\partial^\mu \phi^a+O(\phi^2),\end{aligned}$$ from which we obtain the matrix element $$\label{4:6:lpi} \langle 0|J^{\mu,a}_L(0)|\phi^b(p)\rangle =\frac{F_0}{2}\langle 0|\partial^\mu \phi^a(0)|\phi^b(p)\rangle =-ip^\mu \frac{F_0}{2}\delta^{ab}.$$ Inserting $l_\mu$ of Eq. (\[2:4:rlw\]), we find for the interaction term of a single Goldstone boson with a $W$ $${\cal L}_{W\phi}=\frac{F_0}{2}\mbox{Tr}(l_\mu \partial^\mu \phi) = -\frac{g}{\sqrt{2}}\frac{F_0}{2}\mbox{Tr} [({\cal W}_\mu^+T_+ + {\cal W}^-_\mu T_-)\partial^\mu\phi].$$ Thus, we need to calculate[^60] $$\begin{aligned} \lefteqn{\mbox{Tr}(T_+\partial^\mu\phi)}\\ &=& \mbox{Tr}\left[ \left(\begin{array}{ccc}0&V_{ud}&V_{us}\\0&0&0\\0&0&0\end{array} \right) \partial^\mu \left(\begin{array}{ccc} \pi^0+\frac{1}{\sqrt{3}}\eta &\sqrt{2}\pi^+&\sqrt{2}K^+\\ \sqrt{2}\pi^-&-\pi^0+\frac{1}{\sqrt{3}}\eta&\sqrt{2}K^0\\ \sqrt{2}K^- &\sqrt{2}\bar{K}^0&-\frac{2}{\sqrt{3}}\eta \end{array}\right)\right]\\ &=& V_{ud}\sqrt{2}\partial^\mu\pi^-+V_{us}\sqrt{2}\partial^\mu K^-,\\ \lefteqn{\mbox{Tr}(T_-\partial^\mu\phi)}\\ &=&\mbox{Tr}\left[ \left(\begin{array}{ccc}0&0&0\\ V_{ud}&0&0\\ V_{us}&0&0 \end{array} \right) \partial^\mu \left(\begin{array}{ccc} \pi^0+\frac{1}{\sqrt{3}}\eta &\sqrt{2}\pi^+&\sqrt{2}K^+\\ \sqrt{2}\pi^-&-\pi^0+\frac{1}{\sqrt{3}}\eta&\sqrt{2}K^0\\ \sqrt{2}K^- &\sqrt{2}\bar{K}^0&-\frac{2}{\sqrt{3}}\eta \end{array}\right)\right]\\ &=&V_{ud} \sqrt{2}\partial^\mu\pi^++V_{us}\sqrt{2}\partial^\mu K^+.\end{aligned}$$ We then obtain for the interaction term $$\label{4:6:lwphi} {\cal L}_{W\phi}= -g \frac{F_0}{2} [{\cal W}_\mu^+(V_{ud}\partial^\mu\pi^-+V_{us}\partial^\mu K^-) +{\cal W}_\mu^-(V_{ud}\partial^\mu \pi^++V_{us}\partial^\mu K^+)].$$ In combination with the Feynman propagator for W bosons, $$\label{4:6:wprop} \frac{-g_{\mu\nu}+\frac{k_\mu k_\nu}{M^2_W}}{k^2-M^2_W} =\frac{g_{\mu\nu}}{M^2_W}+O(\frac{kk}{M^4_W}),$$ the Feynman rule for the invariant amplitude for the weak pion decay reads $$\begin{aligned} \label{4:6:mpionzerfall} {\cal M}&=&i\left[-\frac{g}{2\sqrt{2}}\bar{u}_{\nu_\mu} \gamma^\nu (1-\gamma_5)v_{\mu^+}\right] \frac{ig_{\nu\mu}}{M^2_W} i\left[-g\frac{F_0}{2}V_{ud} (-ip^\mu)\right]\nonumber\\ &=&-G_F V_{ud} F_0 \bar{u}_{\nu_\mu} p\hspace{-.4em}/(1-\gamma_5) v_{\mu^+},\end{aligned}$$ where $p$ denotes the four-momentum of the pion and $$G_F=\frac{g^2}{4\sqrt{2}M^2_W}= 1.16639(1)\times 10^{-5}\,\mbox{GeV}^{-2}$$ is the Fermi constant. The evaluation of the decay rate is a standard textbook exercise and we only quote the final result[^61] $$\label{4:6:zr} \frac{1}{\tau}=\frac{G^2_F |V_{ud}|^2}{4\pi} F^2_0 M_\pi m_\mu^2 \left(1-\frac{m_\mu^2}{M_\pi^2}\right)^2.$$ The constant $F_0$ is referred to as the pion-decay constant in the chiral limit.[^62] It measures the strength of the matrix element of the axial-vector current operator between a one-Goldstone-boson state and the vacuum \[see Eq. (\[4:1:acc\])\]. Since the interaction of the $W$ boson with the quarks is of the type $l_\mu^a L^{\mu,a}=l_\mu^a(V^{\mu,a}-A^{\mu,a})/2$ \[see Eq. (\[2:4:rlw\])\] and the vector current operator does not contribute to the matrix element between a single pion and the vacuum, pion decay is completely determined by the axial-vector current. The degeneracy of a single constant $F_0$ in Eq. (\[4:1:acc\]) is lifted at ${\cal O}(p^4)$ [@Gasser:1984gg] once SU(3) symmetry breaking is taken into account. The empirical numbers for $F_\pi$ and $F_K$ are $92.4$ MeV and $113$ MeV, respectively.[^63] ### Pion-Pion Scattering {#subsec_pps} Our second example deals with the prototype of a Goldstone boson reaction: $\pi\pi$ scattering. For the sake of simplicity we will restrict ourselves to the SU(2)$\times$SU(2) version of Eq. (\[4:5:l2\]). We will contrast two different methods of calculating the scattering amplitude: the “direct” calculation in terms of the Goldstone boson fields of the effective Lagrangian versus the calculation of the QCD Green function in combination with the LSZ reduction formalism. Loosely speaking, the “direct” calculation is somewhat more along the spirit of Weinberg’s original paper [@Weinberg:1978kz]: one considers the most general Lagrangian satisfying the general symmetry constraints and calculates $S$-matrix elements with that Lagrangian. The second method will allow one to also consider QCD Green functions “off shell,” i.e., for arbitrary squared invariant momenta. We will discuss under which circumstances the two methods are equivalent and also work out the more general scope of the Green function approach. For the “direct” calculation we set to zero all external fields except for the quark mass term, $\chi=2 B_0 \mbox{diag}(m_q,m_q)=M_\pi^2 1_{2\times 2}$ \[see Eq. (\[4:3:mpi2\])\], $$\label{4:6:l21} {\cal L}_2=\frac{F^2_0}{4}\mbox{Tr}(\partial_\mu U\partial^\mu U^\dagger) +\frac{F_0^2 M_\pi^2}{4}\mbox{Tr}(U^\dagger+U).$$ In our general discussion of the transformation behavior of Goldstone bosons at the end of Sec. \[subsec\_gc\] we argued that we still have a choice how to represent the variables parameterizing the elements of the set of cosets $G/H$. In the present case these are elements of SU(2) and we will illustrate this freedom by making use of two different parameterizations of the matrix $U$ [@Fearing:1999fw],[^64] $$\begin{aligned} \label{4:6:u1} U(x)&=&\frac{1}{F_0}\left[\sigma(x)+i\vec{\tau}\cdot\vec{\pi}(x) \right],\quad \sigma(x)=\sqrt{F^2_0-\vec{\pi}\,^2(x)}, \\ \label{4:6:u2} U(x)&=&\exp\left[i\frac{\vec{\tau}\cdot\vec{\phi}(x)}{F_0}\right],\end{aligned}$$ where in both cases the three Hermitian fields $\pi_i$ and $\phi_i$ describe pion fields transforming as isovectors under SU(2)$_V$. The fields in the two parameterizations are non-linearly related, $$\label{4:6:ft} \frac{\vec{\pi}}{F_0}=\hat{\phi}\sin\left(\frac{|\vec\phi|}{F_0}\right) =\frac{\vec{\phi}}{F_0}\left(1-\frac{1}{6}\frac{\vec{\phi}\,^2}{F^2_0} +\cdots\right).$$ This can be interpreted in terms of a change of variables which leaves the free-field part of the Lagrangian unchanged [@Chisholm; @Kamefuchi:sb]. As a consequence of the equivalence theorem of field theory [@Chisholm; @Kamefuchi:sb] the result for a physical observable should not depend on the choice of variables. The substitution $U\leftrightarrow U^\dagger$ corresponding, respectively, to $\vec{\pi}\mapsto -\vec{\pi}$ and $\vec{\phi}\mapsto -\vec{\phi}$ tells us that ${\cal L}_2$ generates only interaction terms containing an even number of pion fields. Since there exists no vertex involving 3 Goldstone bosons, $\pi\pi$ scattering must be described by a contact interaction at ${\cal O}(p^2)$. By inserting the expressions for $U$ of Eqs. (\[4:6:u1\]) and (\[4:6:u2\]) into Eq. (\[4:6:l21\]) and collecting those terms containing four pion fields we obtain the interaction Lagrangians $$\begin{aligned} \label{4:6:l24pi} {\cal L}_2^{4\pi}&=&\frac{1}{2F^2_0}\partial_\mu \vec{\pi}\cdot\vec{\pi} \partial^\mu \vec{\pi}\cdot\vec{\pi} -\frac{M_\pi^2}{8 F^2_0}(\vec{\pi}\,^2)^2,\\ \label{4:6:l24phi} {\cal L}_2^{4\phi}&=&\frac{1}{6F^2_0}(\partial_\mu \vec{\phi}\cdot\vec{\phi} \partial^\mu \vec{\phi}\cdot\vec{\phi}-\vec{\phi}\,^2 \partial_\mu\vec{\phi} \cdot \partial^\mu\vec{\phi}) +\frac{M_\pi^2}{24 F^2_0}(\vec{\phi}\,^2)^2.\end{aligned}$$ Observe that the two interaction Lagrangians depend differently on the respective pion fields. The corresponding Feynman rules are obtained in the usual fashion by considering all possible ways of contracting pion fields of $i{\cal L}_{\rm int}$ with initial and final pion lines, with the derivatives $\partial_\mu$ generating $-i p_\mu$ ($i p_\mu$) for an initial (final) line. For Cartesian isospin indices $a,b,c,d$ the Feynman rules for the scattering process $\pi^a(p_a)+\pi^b(p_b)\to\pi^c(p_c)+\pi^d(p_d)$ as obtained from Eqs. (\[4:6:l24pi\]) and (\[4:6:l24phi\]) read, respectively, $$\begin{aligned} \label{4:6:mpipi1} {\cal M}_2^{4\pi}&=&i\left[\delta^{ab}\delta^{cd}\frac{s-M^2_\pi}{F^2_0} +\delta^{ac}\delta^{bd}\frac{t-M^2_\pi}{F^2_0} +\delta^{ad}\delta^{bc}\frac{u-M^2_\pi}{F^2_0}\right],\\ \label{4:6:mpipi2} {\cal M}_2^{4\phi}&=&i\left[\delta^{ab}\delta^{cd}\frac{s-M^2_\pi}{F^2_0} +\delta^{ac}\delta^{bd}\frac{t-M^2_\pi}{F^2_0} +\delta^{ad}\delta^{bc}\frac{u-M^2_\pi}{F^2_0}\right]\nonumber\\ &&-\frac{i}{3F_0^2} \left(\delta^{ab}\delta^{cd}+\delta^{ac}\delta^{bd}+\delta^{ad}\delta^{bc} \right) \left(\Lambda_a+\Lambda_b+\Lambda_c+\Lambda_d\right),\end{aligned}$$ where we introduced $\Lambda_k=p_k^2-M^2_\pi$ and the usual Mandelstam variables $$s=(p_a+p_b)^2=(p_c+p_d)^2,$$ $$t=(p_a-p_c)^2=(p_d-p_b)^2,$$ $$u=(p_a-p_d)^2=(p_c-p_b)^2,$$ which are related by $s+t+u=p_a^2+p_b^2+p_c^2+p_d^2$. If the initial and final pions are all on the mass shell, i.e., $\Lambda_k=0$, the scattering amplitudes are the same, in agreement with the equivalence theorem [@Chisholm; @Kamefuchi:sb].[^65] The on-shell result also agrees with the current-algebra prediction for low-energy $\pi\pi$ scattering [@Weinberg:1966kf]. We will come back to $\pi\pi$ scattering in Sec.  \[subsec\_eppsop6\] when we also discuss corrections of higher order [@Bijnens:1995yn]. On the other hand, if one of the momenta of the external lines is off mass shell, the amplitudes of Eqs. (\[4:6:mpipi1\]) and (\[4:6:mpipi2\]) differ. In other words, a “direct” calculation gives a unique result independent of the parameterization of $U$ only for the on-shell matrix element. The second method, developed by Gasser and Leutwyler [@Gasser:1983yg], deals with the Green functions of QCD and their interrelations as expressed in the Ward identities. In particular, these Green functions can, in principle, be calculated for any value of squared momenta even though ChPT is set up only for a low-energy description. For the discussion of $\pi\pi$ scattering one considers the four-point function [@Gasser:1983yg] $$\label{4:6:fpfpppp} G_{PPPP}^{abcd}(x_a,x_b,x_c,x_d)\equiv \langle 0|T[P_a(x_a)P_b(x_b)P_c(x_c)P_d(x_d)]| 0\rangle$$ with the pseudoscalar quark densities of Eq. (\[4:1:psqd\]). In order so see that Eq. (\[4:6:fpfpppp\]) can indeed be related to $\pi\pi$ scattering, let us first investigate the matrix element of the pseudoscalar density evaluated between a single-pion state and the vacuum, which is defined in terms of the coupling constant $G_\pi$ [@Gasser:1983yg]: $$\label{4:6:pipiv} \langle 0|P_i(0)|\pi_j(q)\rangle =\delta_{ij}G_\pi.$$ At ${\cal O}(p^2)$ we determine the coupling of an external pseudoscalar source $p$ to the Goldstone bosons by inserting $\chi=2 B_0 i p$ into the Lagrangian of Eq. (\[4:5:l2\]) (see Fig. \[4:6:figppion\]), $$\label{4:6:l2ext2} {\cal L}_{\rm ext} =i\frac{F_0^2B_0}{2}\mbox{Tr}(pU^\dagger-Up)= \left\{\begin{array}{l} 2B_0F_0p_i \pi_i,\\ 2B_0F_0p_i\phi_i[1-\vec{\phi}\,^2/(6F_0^2)+\cdots], \end{array}\right.$$ where the first and second lines refer to the parameterizations of Eqs. (\[4:6:u1\]) and (\[4:6:u2\]), respectively. From Eq. (\[4:6:l2ext2\]) we obtain $G_\pi = 2B_0 F_0$ independent of the parameterization used which, since the pion is on-shell, is a consequence of the equivalence theorem [@Chisholm; @Kamefuchi:sb]. As a consistency check, let us verify the PCAC relation of Eq. (\[2:4:divasc\]) (without an external electromagnetic field) evaluated between a single-pion state and the vacuum. For the axial-vector current matrix element, we found at ${\cal O}(p^2)$ $$\label{4:6:axialcurrentpion} \langle 0|A^\mu_i(x)|\pi_j(q)\rangle = i q^\mu F_0 e^{-iq\cdot x}\delta_{ij}.$$ Taking the divergence we obtain $$\begin{aligned} \langle 0|\partial_\mu A^\mu_i(x)|\pi_j(q)\rangle &=& i q^\mu F_0 \partial_\mu e^{-iq\cdot x}\delta_{ij} =M^2_\pi F_0 e^{-iq\cdot x}\delta_{ij} = 2m_q B_0F_0 e^{-iq\cdot x}\delta_{ij},\end{aligned}$$ where we made use of Eq. (\[4:3:mpi2\]) for the pion mass. Multiplying Eq. (\[4:6:pipiv\]) by $m_q$ and using $G_\pi = 2B_0 F_0$ we explicitly verify the PCAC relation. Every field $\Phi_i(x)$, which satisfies the relation $$\label{4:6:intpolfield} \langle 0| \Phi_i(x)|\pi_j(q)\rangle = \delta_{ij} e^{-iq \cdot x},$$ can serve as a so-called [*interpolating*]{} pion field [@Borchers:1960] in the LSZ reduction formulas [@Lehmann:1955rq; @Itzykson:rh]. For the case of $\pi^a(p_a)+\pi^b(p_b)\to\pi^c(p_c)+\pi^d(p_d)$ the reduction formula relates the $S$-matrix element to the Green function of the interpolating field as $$\begin{aligned} S_{fi}&=& i^4 \int d^4 x_a \cdots d^4 x_d \,e^{-i p_a \cdot x_a} \cdots e^{i p_d \cdot x_d}\\ &&\times (\Box_a+M_\pi^2)\cdots (\Box_d+M_\pi^2) \langle 0 |T[\Phi_a(x_a)\Phi_b(x_b)\Phi_c(x_c)\Phi_d(x_d)]|0\rangle.\end{aligned}$$ After partial integrations, the Klein-Gordon operators convert into inverse free propagators $$\begin{aligned} S_{fi}&=& (-i)^4(p_a^2-M_\pi^2)\cdots(p_d^2-M_\pi^2)\\ &&\times \int d^4 x_a \cdots d^4 x_d \,e^{-i p_a \cdot x_a} \cdots e^{i p_d \cdot x_d} \langle 0 |T[\Phi_a(x_a)\Phi_b(x_b)\Phi_c(x_c)\Phi_d(x_d)]|0\rangle.\end{aligned}$$ In the present context, we will use $$\label{4:6:pionfield} \Phi_i(x) = \frac{P_i(x)}{G_\pi}=\frac{P_i(x)}{2B_0F_0}= \frac{m_q P_i(x)}{M_\pi^2F_0},$$ which then relates the $S$-matrix element of $\pi\pi$ scattering to the QCD Green function involving four pseudoscalar densities $$\begin{aligned} S_{fi}&=& \left(\frac{-i}{G_\pi}\right)^4(p_a^2-M_\pi^2)\cdots(p_d^2-M_\pi^2)\\ &&\times\int d^4 x_a \cdots d^4 x_d \,e^{-i p_a \cdot x_a} \cdots e^{i p_d \cdot x_d} G_{PPPP}^{abcd}(x_a,x_b,x_c,x_d).\end{aligned}$$ Using translational invariance, let us define the momentum space Green function as $$\begin{aligned} \label{4:6:msgf} \lefteqn{ (2\pi)^4 \delta^4(p_a+p_b+p_c+p_d)F^{abcd}_{PPPP}(p_a,p_b,p_c,p_d)=}\nonumber \\ && \int d^4 x_a d^4 x_b d^4 x_c d^4 x_d\, e^{-i p_a \cdot x_a} e^{-ip_b\cdot x_b} e^{-i p_c\cdot x_c} e^{-i p_d x_d} G_{PPPP}^{abcd}(x_a,x_b,x_c,x_d),\nonumber\\\end{aligned}$$ where we define all momenta as incoming. The usual relation between the $S$ matrix and the $T$ matrix, $S=I+iT$, implies for the $T$-matrix element $\langle f|T|i\rangle= (2\pi)^4\delta^4(P_f-P_I){\cal T}_{fi}$ and, finally, for ${\cal M}=i{\cal T}_{fi}$: $$\label{4:6:calMlsz} {\cal M}= \frac{1}{G_\pi^4} \left[\prod_{k=a,b,c,d}\lim_{p_k^2\to M_\pi^2} (p_k^2-M_\pi^2)\right] F^{abcd}_{PPPP}(p_a,p_b,-p_c,-p_d).$$ We will now determine the Green function $F^{abcd}_{PPPP}(p_a,p_b,-p_c,-p_d)$ using the parameterizations of Eqs.(\[4:6:u1\]) and (\[4:6:u2\]) for $U$. In the first parameterization we only obtain a linear coupling between the external pseudoscalar field and the pion field \[see Eq. (\[4:6:l2ext2\])\] so that only the Feynman diagram of Fig. \[4:6:figpppp1\] contributes $$\begin{aligned} \label{4:6:fabcd1} F^{abcd}_{PPPP}(p_a,p_b,-p_c,-p_d)&=&(2B_0 F_0)^4 \frac{i}{p_a^2-M_\pi^2}\cdots\frac{i}{p_d^2-M_\pi^2} {\cal M}^{4\pi}_2, \nonumber\\\end{aligned}$$ where ${\cal M}^{4\pi}_2$ is given in Eq. (\[4:6:mpipi1\]). The Green function depends on six independent Lorentz scalars which can be chosen as the squared invariant momenta $p^2_k$ and the three Mandelstam variables $s$, $t$, and $u$ satisfying the constraint $s+t+u=\sum_k p_k^2$. Using the second parameterization we will obtain a contribution which is of the same form as Fig. \[4:6:figpppp1\] but with ${\cal M}^{4\pi}_2$ replaced by ${\cal M}^{4\phi}_2$ of Eq. (\[4:6:mpipi2\]). Clearly, this is not yet the same result as Eq. (\[4:6:fabcd1\]) because of the terms proportional to $\Lambda_k$ in Eq. (\[4:6:mpipi2\]). However, in this parameterization the external pseudoscalar field also couples to three pion fields \[see Eq. (\[4:6:l2ext2\])\], resulting in four additional diagrams of the type shown in Fig. \[4:6:addfigpppp2\]. For example, the contribution shown in Fig. \[4:6:addfigpppp2\] reads $$\begin{aligned} \label{4:6:deltafabcd2} \lefteqn{\Delta_a F^{abcd}_{PPPP}(p_a,p_b,-p_c,-p_d)}\nonumber\\ &=& (2B_0 F_0)^3 \frac{i}{p_b^2-M_\pi^2} \frac{i}{p_c^2-M_\pi^2} \frac{i}{p_d^2-M_\pi^2} \left(-\frac{2B_0}{3F_0}\right) (\delta^{ab}\delta^{cd} +\delta^{ac}\delta^{bd}+\delta^{ad}\delta^{bc})\nonumber\\ &=&(2B_0 F_0)^4 \frac{i}{p_a^2-M_\pi^2}\cdots\frac{i}{p_d^2-M_\pi^2} \frac{i\Lambda_a}{3 F_0^2}(\delta^{ab}\delta^{cd} +\delta^{ac}\delta^{bd}+\delta^{ad}\delta^{bc}),\end{aligned}$$ where $\Lambda_a=(p_a^2-M_\pi^2)$. In combination with the contribution of the remaining three diagrams, we find a complete cancelation with those terms proportional to $\Lambda_k$ of Fig. \[4:6:figpppp1\] (in the second parameterization) and the end result is identical with Eq. (\[4:6:fabcd1\]). Finally, using $G_\pi=2 B_0F_0$ and inserting the result of Eq. (\[4:6:fabcd1\]) into Eq. (\[4:6:calMlsz\]) we obtain the same scattering amplitude as in the “direct” calculation of Eqs.  (\[4:6:mpipi1\]) and (\[4:6:mpipi2\]) evaluated for on-shell pions. This example serves as an illustration that the method of Gasser and Leutwyler generates unique results for the Green functions of QCD for arbitrary four-momenta. There is no ambiguity resulting from the choice of variables used to parameterize the matrix $U$ in the effective Lagrangian. These Green functions can be evaluated for arbitrary (but small) four-momenta. Using the reduction formalism, on-shell matrix elements such as the $\pi\pi$ scattering amplitude can be calculated from the QCD Green functions. The result for the $\pi\pi$ scattering amplitude as derived from Eq. (\[4:6:calMlsz\]) agrees with the “direct” calculation of the on-shell matrix elements of Eqs.  (\[4:6:mpipi1\]) and (\[4:6:mpipi2\]). On the other hand, the Feynman rules of Eqs. (\[4:6:mpipi1\]) and (\[4:6:mpipi2\]), when taken [*off shell*]{}, have to be considered as intermediate building blocks only and thus need not be unique. The Chiral Lagrangian at Order ${\cal O}(p^4)$ {#sec_clop4} ---------------------------------------------- Applying the ideas outlined in Sec. \[sec\_cel\] it is possible to construct the most general Lagrangian at ${\cal O}(p^4)$. Here we only quote the result of Ref. [@Gasser:1984gg]: $$\begin{aligned} \label{4:7:l4gl} \lefteqn{{\cal L}_4= L_1 \left\{\mbox{Tr}[D_{\mu}U (D^{\mu}U)^{\dagger}] \right\}^2 + L_2 \mbox{Tr} \left [D_{\mu}U (D_{\nu}U)^{\dagger}\right] \mbox{Tr} \left [D^{\mu}U (D^{\nu}U)^{\dagger}\right]}\nonumber\\ & & + L_3 \mbox{Tr}\left[ D_{\mu}U (D^{\mu}U)^{\dagger}D_{\nu}U (D^{\nu}U)^{\dagger} \right ] + L_4 \mbox{Tr} \left [ D_{\mu}U (D^{\mu}U)^{\dagger} \right ] \mbox{Tr} \left( \chi U^{\dagger}+ U \chi^{\dagger} \right ) \nonumber \\ & & +L_5 \mbox{Tr} \left[ D_{\mu}U (D^{\mu}U)^{\dagger} (\chi U^{\dagger}+ U \chi^{\dagger})\right] + L_6 \left[ \mbox{Tr} \left ( \chi U^{\dagger}+ U \chi^{\dagger} \right ) \right]^2 \nonumber \\ & & + L_7 \left[ \mbox{Tr} \left ( \chi U^{\dagger} - U \chi^{\dagger} \right ) \right]^2 + L_8 \mbox{Tr} \left ( U \chi^{\dagger} U \chi^{\dagger} + \chi U^{\dagger} \chi U^{\dagger} \right ) \nonumber \\ & & -i L_9 \mbox{Tr} \left [ f^R_{\mu\nu} D^{\mu} U (D^{\nu} U)^{\dagger} + f^L_{\mu\nu} (D^{\mu} U)^{\dagger} D^{\nu} U \right ] + L_{10} \mbox{Tr} \left ( U f^L_{\mu\nu} U^{\dagger} f_R^{\mu\nu} \right ) \nonumber \\ & & + H_1 \mbox{Tr} \left ( f^R_{\mu\nu} f^{\mu\nu}_R + f^L_{\mu\nu} f^{\mu\nu}_L \right ) + H_2 \mbox{Tr} \left ( \chi \chi^{\dagger} \right ).\end{aligned}$$ The numerical values of the low-energy coupling constants $L_i$ are not determined by chiral symmetry. In analogy to $F_0$ and $B_0$ of ${\cal L}_2$ they are parameters containing information on the underlying dynamics and should, in principle, be calculable in terms of the (remaining) parameters of QCD, namely, the heavy-quark masses and the QCD scale $\Lambda_{\rm QCD}$. In practice, they parameterize our inability to solve the dynamics of QCD in the non-perturbative regime. So far they have either been fixed using empirical input [@Gasser:1983yg; @Gasser:1984gg; @Bijnens:1994qh] or theoretically using QCD-inspired models [@Ebert:1985kz; @Espriu:1989ff; @Ebert:1991xd; @Bijnens:1992uz], meson-resonance saturation [@Ecker:yg; @Ecker:1988te; @Donoghue:ed; @Knecht:2001xc; @Leupold:2001vs], and lattice QCD [@Myint:yw; @Golterman:2000mg]. From a practical point of view the coefficients are also required for another purpose. When calculating one-loop graphs, using vertices from ${\cal L}_2$ of Eq. (\[4:5:l2\]), one generates infinities which, according to Weinberg’s power counting of Eq. (\[4:4:mr2\]), are of ${\cal O}(p^4)$, i.e., which cannot be absorbed by a renormalization of the coefficients $F_0$ and $B_0$. In the framework of dimensional regularization (see App. \[app\_drb\]) these divergences appear as poles at space-time dimension $n=4$. In Refs. [@Gasser:1983yg; @Gasser:1984gg] the poles, together with the relevant counter terms, were given in closed form. To that end, Gasser and Leutwyler made use of the so-called saddle-point method which, in the path-integral approach, allows one to identify the one-loop contribution to the generating functional. The action is expanded around the classical solution and the path integral is performed with respect to the terms quadratic in the fluctuations about the classical solution. The resulting one-loop piece of the generating functional is treated within the dimensional-regularization procedure and the poles are isolated by applying the so-called heat-kernel technique.[^66] Except for $L_3$ and $L_7$ the low-energy coupling constants $L_i$ and the “contact terms”—i.e., pure external field terms—$H_1$ and $H_2$ are required in the renormalization of the one-loop graphs [@Gasser:1984gg]. Since $H_1$ and $H_2$ contain only external fields, they are of no physical relevance [@Gasser:1984gg]. By construction Eq. (\[4:7:l4gl\]) represents the most general Lagrangian at ${\cal O}(p^4)$, and it is thus possible to absorb the one-loop divergences by an appropriate renormalization of the coefficients $L_i$ and $H_i$ [@Gasser:1984gg]: $$\begin{aligned} \label{4:7:li} L_i&=&L_i^r+\frac{\Gamma_i}{32\pi^2}R, \quad i=1,\cdots,10,\\ \label{4:7:hi} H_i&=&H^r_i+\frac{\Delta_i}{32\pi^2}R,\quad i=1,2,\end{aligned}$$ where $R$ is defined as (see App. \[app\_drb\]) $$\label{4:7:R} R=\frac{2}{n-4}-[\mbox{ln}(4\pi)-\gamma_E+1],$$ with $n$ denoting the number of space-time dimensions and $\gamma_E=-\Gamma'(1)$ being Euler’s constant. The constants $\Gamma_i$ and $\Delta_i$ are given in Table \[4:7:tableli\]. The renormalized coefficients $L_i^r$ depend on the scale $\mu$ introduced by dimensional regularization \[see Eq. (\[app:drb:im22\])\] and their values at two different scales $\mu_1$ and $\mu_2$ are related by $$\label{4:7:limu1mu2} L^r_i(\mu_2)=L^r_i(\mu_1) +\frac{\Gamma_i}{16\pi^2}\ln\left(\frac{\mu_1}{\mu_2}\right).$$ We will see that the scale dependence of the coefficients and the finite part of the loop-diagrams compensate each other in such a way that physical observables are scale independent. Coefficient Empirical Value $\Gamma_i$ ------------- ----------------- ------------------ $L_1^r$ $ 0.4\pm 0.3$ $\frac{3}{32}$ $L_2^r$ $ 1.35\pm 0.3$ $\frac{3}{16}$ $L_3^r$ $-3.5\pm 1.1$ $0$ $L_4^r$ $-0.3\pm 0.5$ $\frac{1}{8}$ $L_5^r$ $ 1.4\pm 0.5$ $\frac{3}{8}$ $L_6^r$ $-0.2\pm 0.3$ $\frac{11}{144}$ $L_7^r$ $-0.4\pm 0.2$ $0$ $L_8^r$ $ 0.9\pm 0.3$ $\frac{5}{48}$ $L_9^r$ $ 6.9\pm 0.7$ $\frac{1}{4}$ $L_{10}^r$ $-5.5\pm 0.7$ $-\frac{1}{4}$ : \[4:7:tableli\] Renormalized low-energy coupling constants $L_i^r$ in units of $10^{-3}$ at the scale $\mu=M_\rho$ [@Bijnens:1994qh]. $\Delta_1=-1/8$, $\Delta_2=5/24$. We finally discuss the method of using field transformations to eliminate redundant terms in the most general effective Lagrangian . From a “naive” point of view the two structures $$\label{4:7:addstruc} \mbox{Tr}[D^2 U (D^2 U)^\dagger],\quad \mbox{Tr}[D^2 U \chi^\dagger +\chi (D^2 U)^\dagger]$$ would qualify as independent terms of order ${\cal O}(p^4)$. Loosely speaking, by using the classical equation of motion of Eq. (\[4:5:eom\]) these terms can be shown to be redundant. We will justify this statement in terms of field transformations. To that end let us consider another SU(3) matrix $U'(x)$ which is related to $U(x)$ by a field transformation of the form $$\label{4:7:up} U(x)=\exp[iS(x)] U'(x).$$ Since both $U$ and $U'$ are SU(3) matrices, $S(x)$ must be a Hermitian traceless $3\times 3$ matrix. We demand that $U'(x)$ satisfies the same symmetry properties as $U(x)$ (see Table \[4:5:table\_trafprop\]), $$\label{4:7:uptrafo} U'\stackrel{G}{\mapsto} V_R U' V_L^\dagger,\quad U'(\vec{x},t)\stackrel{P}{\mapsto} U'^\dagger (-\vec{x},t),\quad U'\stackrel{C}{\mapsto}U'^T,$$ from which we obtain the following conditions for $S$: $$\label{4:7:strafo} S\stackrel{G}{\mapsto} V_R S V_R^\dagger,\quad S(\vec{x},t)\stackrel{P}{\mapsto} -U'^\dagger(-\vec{x},t)S(-\vec{x},t) U'(-\vec{x},t),\quad S\stackrel{C}{\mapsto}(U'^\dagger SU')^T.$$ The most general transformation is constructed iteratively in the momentum and quark-mass expansion, $$\label{4:7:trafoit} U=\exp[iS_2(x)]U^{(1)}(x),\quad U^{(1)}(x)=\exp[iS_4(x)]U^{(2)}(x),\quad \cdots,$$ where the matrices $S_{2n}$ are of ${\cal O}(p^{2n})$, satisfy the properties of Eq. (\[4:7:strafo\]), and have to be constructed from the same building blocks as the effective Lagrangian. To be explicit, let us derive the most general matrix $S_2(x)$. At ${\cal O}(p^2)$, the field strength tensors cannot contribute as building blocks because of their antisymmetry under interchange of the Lorentz indices. Imposing the transformation behavior under the group $G=\mbox{SU(3)}_L\times\mbox{SU(3)}_R$, we obtain a list of five terms $$\label{4:7:bausteines2} D^2 U' U'^\dagger,\quad U' (D^2 U')^\dagger,\quad D_\mu U' (D^\mu U')^\dagger,\quad \chi U'^\dagger,\quad U'\chi^\dagger.$$ Parity eliminates three combinations and we are left with $$\label{4:7:bausteines22} D^2 U' U'^\dagger-U'(D^2 U')^\dagger,\quad \chi U'^\dagger - U' \chi^\dagger.$$ Demanding Hermiticity and a vanishing trace, we end up with two terms at ${\cal O}(p^2)$: $$\label{4:7:s2} S_2=i\alpha_1[D^2 U' U'^\dagger- U'(D^2 U')^\dagger] +i\alpha_2[\chi U'^\dagger - U' \chi^\dagger -\frac{1}{3}\mbox{Tr}(\chi U'^\dagger- U'\chi^\dagger)],$$ where $\alpha_1$ and $\alpha_2$ are real numbers. At ${\cal O}(p^2)$, charge conjugation does not provide an additional constraint. What are the consequences of working with $U'(x)$ instead of $U(x)$? In Sec. \[subsec\_pps\] we have already argued, by means of a simple example, that the results for the Green functions are independent of the parameterizations of $U(x)$ of Eqs. (\[4:6:u1\]) and (\[4:6:u2\]). Expressing $U(x)$ of Eq. (\[4:7:up\]) by using Eq. (\[4:7:s2\]) and inserting the result into ${\cal L}_2$ of Eq. (\[4:5:l2\]), we obtain $${\cal L}_2(U)={\cal L}_2(U')+\Delta {\cal L}_2(U'),$$ where $\Delta {\cal L}_2$, to leading order in $S_2$, is given by $$\label{4:7:dl2} \Delta {\cal L}_2(U')=\frac{F^2_0}{4}\mbox{Tr}[iS_2 {\cal O}_{\rm EOM}^{(2)}(U')]+ O(S^2_2).$$ The functional form of ${\cal O}_{\rm EOM}^{(2)}$ has been defined in Eq. (\[4:5:eom\]). Note, however, that we do [*not*]{} assume ${\cal O}_{\rm EOM}^{(2)}=0$. We have dropped a total derivative, since it does not modify the dynamics. Both $S_2$ and ${\cal O}_{\rm EOM}^{(2)}$ are of order ${\cal O}(p^2)$ so that $\Delta{\cal L}_2$ is of order ${\cal O}(p^4)$. Of course, higher powers of $S_2$ in Eq. (\[4:7:dl2\]) induce additional terms of higher orders in the momentum expansion which we will discuss in a moment. Through a suitable choice of the parameters $\alpha_1$ and $\alpha_2$ it is possible to eliminate two structures at order ${\cal O}(p^4)$, i.e., one generates a new Lagrangian with a different functional form which, however, according to the equivalence theorem leads to the same observables [@Chisholm; @Kamefuchi:sb]. Such a procedure is commonly referred to as using the classical equation of motion to eliminate terms. For example, it is straightforward but tedious to re-express the two structures of Eq. (\[4:7:addstruc\]) through the terms of Gasser and Leutwyler, Eq. (\[4:7:l4gl\]), and the following two terms $$\label{4:7:addstruc2} c_1 \mbox{Tr}\left([D^2 U U^\dagger- U (D^2 U)^\dagger] {\cal O}_{\rm EOM}^{(2)}\right) + c_2 \mbox{Tr} \left((\chi U^\dagger - U\chi^\dagger){\cal O}_{\rm EOM}^{(2)}\right).$$ Choosing $\alpha_1=4 c_1/F_0^2$ and $\alpha_2=4 c_2/F_0^2$ in Eq. (\[4:7:s2\]), the two terms of Eq. (\[4:7:addstruc2\]) and the modification $\Delta {\cal L}_2$ of Eq. (\[4:7:dl2\]) precisely cancel and one is left with the canonical form of Gasser and Leutwyler. A field redefinition, of course, also leads to modifications of the functional form of the effective Lagrangians of higher orders. However, for $S_2$ such terms are at least of order ${\cal O}(p^6)$ as are the higher-order terms in Eq. (\[4:7:dl2\]). Thus one proceeds iteratively [@Scherer:1994wi]. Using $S_2$ one generates the simplest form of ${\cal L}_4$. Next one constructs $S_4$, inserts it again into ${\cal L}_2$ to simplify ${\cal L}_6$, etc. From a point of view of [*constructing*]{} the simplest Lagrangian at a given order it is sufficient to identify those terms proportional to the classical, i.e. lowest-order, equation of motion and drop them right from the beginning using the argument that, by choosing appropriate generators, they can be transformed away. A completely different situation arises if one tries to express the effective Lagrangian obtained within the framework of a specific model in the canonical form. In such a case it is necessary to explicitly perform the iteration process consistently to a given order and, in particular, take into account the modification of the higher-order coefficients due to the transformation. An explicit example is given in Appendix AII of Ref. [@Belkov:1994qg]. The Effective Wess-Zumino-Witten Action {#sec_ewzwa} --------------------------------------- The Lagrangians ${\cal L}_2$ and ${\cal L}_4$ discussed so far exhibit a larger symmetry than the “real world.” For example, if we consider the case of “pure” QCD, i.e., no external fields except for $\chi=2B_0 M$ with the quark mass matrix $M$ of Eq. (\[4:3:qmt\]), the two Lagrangians are invariant under the substitution $\phi(x)\mapsto -\phi(x)$. As discussed in Sec. \[sec\_loel\] they contain interaction terms with an even number of Goldstone bosons only, i.e., they are of even intrinsic parity, and it would not be possible to describe the reaction $K^+K^-\to\pi^+\pi^-\pi^0$.[^67] Analogously, the process $\pi^0\to\gamma\gamma$ cannot be described by ${\cal L}_2$ and ${\cal L}_4$ in the presence of external electromagnetic fields. These observations lead us to a discussion of the effective Wess-Zumino-Witten action [@Wess:yu; @Witten:tw]. Whereas normal Ward identities are related to the [*invariance*]{} of the generating functional under local transformations of the external fields, the anomalous Ward identities [@Adler:1969gk; @Adler:1969er; @Bardeen:1969md; @Bell:1969ts; @Adler:1970], which were first obtained in the framework of renormalized perturbation theory, give a particular form to the [*variation*]{} of the generating functional [@Wess:yu; @Gasser:1983yg]. Wess and Zumino derived consistency or integrability relations which are satisfied by the anomalous Ward identities and then explicitly constructed a functional involving the pseudoscalar octet which satisfies the anomalous Ward identities [@Wess:yu]. In particular, Wess and Zumino emphasized that their interaction Lagrangians cannot be obtained as part of a chiral invariant Lagrangian. In the construction of Witten [@Witten:tw] the simplest term possible which breaks the symmetry of having only an even number of Goldstone bosons at the Lagrangian level is added to the equation of motion of Eq. (\[4:5:eom\]) for the case of massless Goldstone bosons without any external fields,[^68] $$\label{4:8:eomadd} \partial_\mu\left(\frac{F_0^2}{2}U\partial^\mu U^\dagger\right) +\lambda \epsilon^{\mu\nu\rho\sigma} U\partial_\mu U^\dagger U\partial_\nu U^\dagger U\partial_\rho U^\dagger U\partial_\sigma U^\dagger=0,$$ where $\lambda$ is a (purely imaginary) constant. Substituting $U\leftrightarrow U^\dagger$ in Eq. (\[4:8:eomadd\]) and subsequently multiplying from the left by $U$ and from the right by $U^\dagger$, we verify that the two terms transform with opposite relative signs. Recall that a term which is even (odd) in the Lagrangian leads to a term which, in the equation of motion, is odd (even). However, the action functional corresponding to the new term cannot be written as the four-dimensional integral of a Lagrangian expressed in terms of $U$ and its derivatives. Rather, one has to extend the range of definition of the fields to a hypothetical fifth dimension, $$\label{4:8:ualpha} U(y)=\exp\left(i\alpha\frac{\phi(x)}{F_0}\right), \quad y^i=(x^\mu,\alpha),\,\,i=0,\cdots, 4,\,\, 0\leq\alpha\leq 1,$$ where Minkowski space is defined as the surface of the five-dimensional space for $\alpha =1$. Let us first quote the result of the effective Wess-Zumino-Witten action in the absence of external fields (denoted by a superscript 0): $$\begin{aligned} \label{4:8:swzw1} S_{\rm ano}^0&=&n S_{\rm WZW}^0,\\ \label{4:8:swzw2} S_{\rm WZW}^0&=&-\frac{i}{240\pi^2}\int_0^1 d \alpha \int d^4 x \epsilon^{ijklm} \mbox{Tr}\left( {\cal U}^L_i {\cal U}^L_j {\cal U}^L_k {\cal U}^L_l {\cal U}^L_m \right),\end{aligned}$$ where the indices $i,\cdots,m$ run from 0 to 4, $y_4=y^4=\alpha$, $\epsilon_{ijklm}$ is the completely antisymmetric tensor with $\epsilon_{01234}=-\epsilon^{01234}=1$, and ${\cal U}^L_i=U^\dagger \partial U/\partial y^i$. By calculating the variation of the action functional as in Eq.(\[4:5:deltas\]) we find that the constant $\lambda$ of Eq. (\[4:8:eomadd\]) and $n$ of Eq. (\[4:8:swzw1\]) are related by $\lambda=in/(48\pi^2)$. Using topological arguments Witten showed that the constant $n$ appearing in Eq. (\[4:8:swzw1\]) must be an integer. Below, $n$ will be identified with the number of colors $N_c$. Expanding the SU(3) matrix $U(y)$ in terms of the Goldstone boson fields, $U(y)=1+i\alpha \phi(x)/F_0+O(\phi^2)$, one obtains an infinite series of terms, each involving an odd number of Goldstone bosons, i.e., the WZW action $S_{\rm WZW}^0$ is of odd intrinsic parity. For each individual term the $\alpha$ integration can be performed explicitly resulting in an ordinary action in terms of a four-dimensional integral of a local Lagrangian. For example, the term with the smallest number of Goldstone bosons reads $$\begin{aligned} \label{4:8:swzw5phi} S_{\rm WZW}^{5\phi}&=&\frac{1}{240\pi^2F^5_0}\int_0^1 d\alpha\int d^4 x \epsilon^{ijklm}\mbox{Tr}[\partial_i(\alpha\phi) \partial_j(\alpha\phi) \partial_k(\alpha\phi) \partial_l(\alpha\phi) \partial_m(\alpha\phi)]\nonumber\\ &=&\frac{1}{240\pi^2F^5_0}\int_0^1 d\alpha \int d^4 x \epsilon^{ijklm}\partial_i\mbox{Tr}[\alpha\phi \partial_j(\alpha\phi) \partial_k(\alpha\phi) \partial_l(\alpha\phi) \partial_m(\alpha\phi)]\nonumber\\ &=&\frac{1}{240\pi^2 F^5_0}\int d^4 x \epsilon^{\mu\nu\rho\sigma}\mbox{Tr}(\phi\partial_\mu\phi\partial_\nu\phi \partial_\rho\phi\partial_\sigma\phi).\end{aligned}$$ In the last step we made use of the fact that exactly one index can take the value 4. The term involving $i=4$ has been integrated with respect to $\alpha$ whereas the other four possibilities cancel each other because the $\epsilon$ tensor in four dimensions is antisymmetric under a cyclic permutation of the indices whereas the trace is symmetric under a cyclic permutation. In particular, the WZW action without external fields involves at least five Goldstone bosons [@Wess:yu]. The connection to the number of colors $N_c$ is established by introducing a coupling to electromagnetism [@Wess:yu; @Witten:tw]. In the presence of external fields there will be an additional term in the anomalous action, $$\label{4:8:sanofull} S_{\rm ano}= S_{\rm ano}^0+ S_{\rm ano}^{\rm ext}=n(S_{\rm WZW}^0+S_{\rm WZW}^{\rm ext}),$$ given by [@Chou:1983qy; @Pak:1984bn; @Manes:1984gk; @Bijnens:xi] $$\label{4:8:deltaswzw} S_{\rm WZW}^{\rm ext}= -\frac{i}{48\pi^2}\int d^4 x\, \epsilon^{\mu\nu\rho\sigma}\mbox{Tr}(Z_{\mu\nu\rho\sigma})$$ with $$\begin{aligned} \label{4:8:zmunialphabeta} Z_{\mu\nu\rho\sigma}&=& \frac{1}{2}\ U l_\mu U^\dagger r_\nu U l_\rho U^\dagger r_\sigma\nonumber\\ &&+U l_\mu l_\nu l_\rho U^\dagger r_\sigma - U^\dagger r_\mu r_\nu r_\rho U l_\sigma\nonumber\\ &&+i U\partial_\mu l_\nu l_\rho U^\dagger r_\sigma - i U^\dagger\partial_\mu r_\nu r_\rho U l_\sigma\nonumber\\ &&+i \partial_\mu r_\nu U l_\rho U^\dagger r_\sigma - i \partial_\mu l_\nu U^\dagger r_\rho U l_\sigma\nonumber\\ &&-i {\cal U}_\mu^L l_\nu U^\dagger r_\rho U l_\sigma + i {\cal U}_\mu^R r_\nu U l_\rho U^\dagger r_\sigma\nonumber\\ &&-i {\cal U}_\mu^L l_\nu l_\rho l_\sigma + i {\cal U}_\mu^R r_\nu r_\rho r_\sigma\nonumber\\ &&+\frac{1}{2}{\cal U}_\mu^L U^\dagger\partial_\nu r_\rho U l_\sigma -\frac{1}{2}{\cal U}_\mu^R U\partial_\nu l_\rho U^\dagger r_\sigma\nonumber\\ &&+\frac{1}{2}{\cal U}_\mu^L U^\dagger r_\nu U\partial_\rho l_\sigma -\frac{1}{2}{\cal U}_\mu^R U l_\nu U^\dagger\partial_\rho r_\sigma\nonumber\\ &&-{\cal U}_\mu^L{\cal U}_\nu^L U^\dagger r_\rho U l_\sigma + {\cal U}_\mu^R{\cal U}_\nu^R U l_\rho U^\dagger r_\sigma\nonumber\\ &&+{\cal U}_\mu^L l_\nu\partial_\rho l_\sigma - {\cal U}_\mu^R r_\nu\partial_\rho r_\sigma\nonumber\\ &&+{\cal U}_\mu^L\partial_\nu l_\rho l_\sigma - {\cal U}_\mu^R\partial_\nu r_\rho r_\sigma\nonumber\\ &&+\frac{1}{2}\ {\cal U}_\mu^L l_\nu {\cal U}_\rho^L l_\sigma - \frac{1}{2}\ {\cal U}_\mu^R r_\nu {\cal U}_\rho^R r_\sigma\nonumber\\ &&-i {\cal U}_\mu^L{\cal U}_\nu^L{\cal U}_\rho^L l_\sigma + i {\cal U}_\mu^R{\cal U}_\nu^R{\cal U}_\rho^R r_\sigma,\end{aligned}$$ where we defined the abbreviations ${\cal U}_\mu^L=U^\dagger\partial_\mu U$ and ${\cal U}_\mu^R=U\partial_\mu U^\dagger$. In the commonly used expression [@Bijnens:xi], we performed the replacement $$\begin{aligned} \lefteqn{{\cal U}_\mu^L U^\dagger\partial_\nu r_\rho U l_\sigma- {\cal U}_\mu^R U\partial_\nu l_\rho U^\dagger r_\sigma\to}\\ && \frac{1}{2}{\cal U}_\mu^L U^\dagger\partial_\nu r_\rho U l_\sigma -\frac{1}{2}{\cal U}_\mu^R U\partial_\nu l_\rho U^\dagger r_\sigma +\frac{1}{2}{\cal U}_\mu^L U^\dagger r_\nu U\partial_\rho l_\sigma -\frac{1}{2}{\cal U}_\mu^R U l_\nu U^\dagger\partial_\rho r_\sigma,\end{aligned}$$ in order to generate a manifestly $C$ invariant and Hermitian action. Without this replacement charge-conjugation invariance and Hermiticity are satisfied up to a total derivative only. As a special case, let us consider a coupling to external electromagnetic fields by inserting $$r_\mu=l_\mu=-e Q {\cal A}_\mu,$$ where $Q$ is the quark charge matrix \[see Eq. (\[2:4:rla\])\]. The terms involving three and four electromagnetic four-potentials vanish upon contraction with the totally antisymmetric tensor $\epsilon^{\mu\nu\rho\sigma}$, because their contributions to $Z_{\mu\nu\rho\sigma}$ are symmetric in at least two indices, and we obtain $$\begin{aligned} \label{4:8:lanoelm} n{\cal L}^{\rm ext}_{\rm WZW}&=&-e n {\cal A}_\mu J^\mu +i \frac{n e^2 }{48\pi^2}\epsilon^{\mu\nu\rho\sigma} \partial_\nu {\cal A}_\rho{\cal A}_\sigma \mbox{Tr}[2Q^2(U\partial_\mu U^\dagger - U^\dagger \partial_\mu U) \nonumber\\ && - Q U^\dagger Q \partial_\mu U +Q U Q \partial_\mu U^\dagger].\end{aligned}$$ We note that the current $$\label{4:8:jmu} J^\mu=\frac{\epsilon^{\mu\nu\rho\sigma}}{48\pi^2} \mbox{Tr}(Q\partial_\nu U U^\dagger \partial_\rho U U^\dagger \partial_\sigma U U^\dagger +Q U^\dagger \partial_\nu U U^\dagger \partial_\rho U U^\dagger \partial_\sigma U),\quad \epsilon_{0123}=1,$$ by itself is not gauge invariant and the additional terms of Eq. (\[4:8:lanoelm\]) are required to obtain a gauge-invariant action. The identification of the constant $n$ with the number of colors $N_c$ [@Witten:tw] results from finding in Eq. (\[4:8:lanoelm\]) the interaction Lagrangian which is relevant to the decay $\pi^0\to\gamma\gamma$. Since $U=1+i\mbox{diag}(\pi^0,-\pi^0,0)/F_0+\cdots$, Eq. (\[4:8:lanoelm\]) contains a piece $$\label{4:8:lpi0gg} {\cal L}_{\pi^0\gamma\gamma}=-\frac{n e^2}{96\pi^2}\epsilon^{\mu\nu \rho\sigma}{\cal F}_{\mu\nu}{\cal F}_{\rho\sigma}\frac{\pi^0}{F_0},$$ where we made use of a partial integration to shift the derivative from the pion field onto the electromagnetic four-potential. The corresponding invariant amplitude reads $$\label{4:8:mpi0gg} {\cal M}=i\frac{n e^2}{12\pi^2 F_0}\epsilon^{\mu\nu\rho\sigma} q_{1\mu}\epsilon_{1\nu}^\ast q_{2\rho}\epsilon_{2\sigma}^\ast,$$ which agrees with a direct calculation of the anomaly term in terms of $u$ and $d$ quarks with $n=N_c$ colors (see, e.g., Ref. [@Veltman:wz]). After summation over the final photon polarizations and integration over phase space, the decay rate reads $$\label{} \Gamma_{\pi^0\to\gamma\gamma}= \frac{\alpha^2 M^3_{\pi^0} n^2}{576 \pi^3 F^2_0} =7.6\,\mbox{eV}\times\left(\frac{n}{3}\right)^2,$$ which is in good agreement with the experimental value $(7.7\pm 0.6)$ eV for $n=N_c=3$ [@Groom:in]. Applications at Order ${\cal O}(p^4)$ {#sec_aop4} ------------------------------------- ### Masses of the Goldstone Bosons {#subsec_mgb} A discussion of the masses at ${\cal O}(p^4)$ will allow us to illustrate various properties typical of chiral perturbation theory: 1. The relation between the bare low-energy coupling constants $L_i$ and the renormalized coefficients $L_i^r$ in Eq. (\[4:7:li\]) is such that the divergences of one-loop diagrams are canceled. 2. Similarly, the scale dependence of the coefficients $L^r_i(\mu)$ on the one hand and of the finite contributions of the one-loop diagrams on the other hand leads to scale-independent predictions for physical observables. 3. A perturbation expansion in the explicit symmetry breaking with respect to a symmetry that is realized in the Nambu-Goldstone mode generates corrections which are non-analytic in the symmetry breaking parameter [@Li:1971vr], here the quark masses. Let us consider ${\cal L}_2 + {\cal L}_4$ for QCD with finite quark masses but in the absence of external fields. We restrict ourselves to the limit of isospin symmetry, i.e., $m_u=m_d=m$. In order to determine the masses we calculate the self energies $\Sigma(p^2)$ of the Goldstone bosons. The propagator of a (pseudo-) scalar field is defined as the Fourier transform of the two-point Green function: $$\label{4:8:propdef} i\Delta(p)=\int d^4 x e^{-ip\cdot x}\langle 0|T\left[\Phi_0(x)\Phi_0(0)\right]|0\rangle,$$ where the index 0 refers to the fact that we still deal with the bare unrenormalized field—not to be confused with a free field without interaction. At lowest order ($D=2$) the propagator simply reads $$\label{4:8:prop} i\Delta(p)=\frac{i}{p^2-M^2_0+i0^+},$$ where the lowest-order masses $M_0$ are given in Eqs. (\[4:3:mpi2\]) - (\[4:3:meta2\]): $$\begin{aligned} M^2_{\pi,0}&=&2 B_0 m,\\ M^2_{K,0}&=&B_0(m+m_s),\\ M^2_{\eta,0}&=&\frac{2}{3} B_0\left(m+2m_s\right).\end{aligned}$$ The loop diagrams with ${\cal L}_2$ and the contact diagrams with ${\cal L}_4$ result in so-called proper self-energy insertions $-i\Sigma(p^2)$, which may be summed using a geometric series (see Fig. \[4:8:fullprop\]): $$\begin{aligned} \label{4:8:prop1} i\Delta(p)&=&\frac{i}{p^2-M^2_0+i0^+}+\frac{i}{p^2-M^2_0+i0^+} [-i\Sigma(p^2)]\frac{i}{p^2-M^2_0+i0^+}+\cdots\nonumber\\ &=&\frac{i}{p^2-M^2_0-\Sigma(p^2)+i0^+}.\end{aligned}$$ Note that $-i\Sigma(p^2)$ consists of one-particle-irreducible diagrams only, i.e., diagrams which do not fall apart into two separate pieces when cutting an arbitrary internal line. The physical mass, including the interaction, is defined as the position of the pole of Eq. (\[4:8:prop1\]), $$\label{4:8:mdef} M^2-M^2_0-\Sigma(M^2)\stackrel{!}{=}0.$$ Let us assume that $\Sigma(p^2)$ can be expanded in a series around $p^2=\lambda^2$, $$\label{4:8:sigmaexp} \Sigma(p^2)=\Sigma(\lambda^2) +(p^2-\lambda^2)\Sigma'(\lambda^2)+\tilde{\Sigma}(p^2),$$ where the remainder $\tilde{\Sigma}(p^2)$ depends on the choice of $\lambda^2$ and satisfies $\tilde{\Sigma}(\lambda^2)=\tilde{\Sigma}'(\lambda^2)=0$. We then obtain for the propagator $$\label{4:8:prop2} i\Delta(p)=\frac{i}{p^2-M^2_0 -\Sigma(\lambda^2)-(p^2-\lambda^2)\Sigma'(\lambda^2) -\tilde{\Sigma}(p^2)+i0^+}.$$ Taking $\lambda^2=M^2$ in Eq. (\[4:8:prop2\]) and applying the condition of Eq.  (\[4:8:mdef\]), the propagator may be written as $$i\Delta(p)=\frac{i}{(p^2-M^2)[1-\Sigma'(M^2)]-\tilde{\Sigma}(p^2)+i0^+} =\frac{iZ_\Phi}{p^2-M^2-Z_\Phi\tilde{\Sigma}(p^2)+i0^+},$$ where we have introduced the wave function renormalization constant $$Z_\Phi=\frac{1}{1-\Sigma'(M^2)}.$$ Introducing renormalized fields as $\Phi_R=\Phi_0/\sqrt{Z_\Phi}$, the renormalized propagator is given by $$\begin{aligned} i\Delta_R(p)&=&\int d^4 x e^{-ip\cdot x} \langle 0|T[\Phi_R(x)\Phi_R(0)]|0\rangle\nonumber\\ &=&\frac{i}{p^2-M^2-Z_\Phi\tilde{\Sigma}(p^2)+i0^+}.\end{aligned}$$ In particular, since $\tilde{\Sigma}(M^2) =\tilde{\Sigma}'(M^2)=0$, in the vicinity of the pole, the renormalized propagator behaves as a free propagator with physical mass $M^2$. Let us now turn to the calculation within the framework of ChPT (see, e.g., Ref. [@Rudy:1994qb]). Since ${\cal L}_2$ and ${\cal L}_4$ without external fields generate vertices with an even number of Goldstone bosons only, the candidate terms at $D=4$ contributing to the self energy are those shown in Fig. \[4:8:selfenergy\]. For our particular application with exactly two external meson lines, the relevant interaction Lagrangians can be written as [@Rudy:1994qb] $${\cal L}_{\rm int}={\cal L}_2^{4\phi}+{\cal L}_4^{2\phi},$$ where ${\cal L}_2^{4\phi}$ is given by $$\label{4:8:l24phi} {\cal L}^{4\phi}_2=\frac{1}{24 F^2_0}\left\{\mbox{Tr}( [\phi,\partial_\mu \phi]\phi \partial^\mu \phi) +B_0\mbox{Tr}(M\phi^4)\right\}.$$ The terms of ${\cal L}_4$ proportional to $L_9$, $L_{10}$, $H_1$, and $H_2$ do not contribute, because they either contain field-strength tensors or external fields only. Since $\partial_\mu U=O(\phi)$, the $L_1$, $L_2$, and $L_3$ terms of Eq. (\[4:7:l4gl\]) are $O(\phi^4)$ and need not be considered. The only candidates are the $L_4$ - $L_8$ terms, of which we consider the $L_4$ term as an explicit example,[^69] $$\begin{aligned} \lefteqn{L_4\mbox{Tr}(\partial_\mu U \partial^\mu U^\dagger) \mbox{Tr}(\chi U^\dagger +U \chi^\dagger)=}\\ &&L_4 \frac{2}{F_0^2}[\partial_\mu \eta \partial^\mu \eta +\partial_\mu \pi^0 \partial^\mu \pi^0 +2\partial_\mu \pi^+\partial^\mu\pi^-+2\partial_\mu K^+\partial^\mu K^-\\ &&+2\partial_\mu K^0\partial^\mu \bar{K}^0+O(\phi^4)] [4B_0(2m+m_s)+O(\phi^2)].\end{aligned}$$ The remaining terms are treated analogously and we obtain for ${\cal L}_4^{2\phi}$ $$\begin{aligned} \label{4:8:l42phi} {\cal L}_4^{2\phi}&=& \frac{1}{2}\left(a_\eta\partial_\mu\eta\partial^\mu\eta-b_\eta \eta^2\right) \nonumber\\ &&+\frac{1}{2}\left(a_\pi\partial_\mu\pi^0\partial^\mu\pi^0-b_\pi\pi^0\pi^0 \right)\nonumber\\ &&+a_\pi\partial_\mu\pi^+\partial^\mu\pi^--b_\pi\pi^+\pi^-\nonumber\\ &&+a_K\partial_\mu K^+\partial^\mu K^- - b_K K^+K^-\nonumber\\ &&+a_K\partial_\mu K^0\partial^\mu\bar{K}^0-b_K K^0\bar{K}^0,\end{aligned}$$ where the constants $a_\phi$ and $ b_\phi$ are given by $$\begin{aligned} \label{4:8:ab} a_\eta&=&\frac{16B_0}{F^2_0}\left[(2m+m_s)L_4+\frac{1}{3}(m+2m_s)L_5\right], \nonumber\\ b_\eta&=&\frac{64 B^2_0}{3F^2_0}\left[(2m+m_s)(m+2m_s)L_6+2(m-m_s)^2L_7 +(m^2+2m_s^2)L_8\right],\nonumber\\ a_\pi&=&\frac{16 B_0}{F^2_0}\left[(2m+m_s)L_4+mL_5\right],\nonumber\\ b_\pi&=&\frac{64 B^2_0}{F^2_0}\left[(2m+m_s)mL_6+m^2L_8\right],\nonumber\\ a_K&=&\frac{16 B_0}{F^2_0}\left[(2m+m_s)L_4+\frac{1}{2}(m+m_s)L_5\right], \nonumber\\ b_K&=&\frac{32B^2_0}{F^2_0}\left[(2m+m_s)(m+m_s)L_6+\frac{1}{2}(m+m_s)^2L_8 \right].\end{aligned}$$ At ${\cal O}(p^4)$ the self energies are of the form $$\label{4:8:sigmaphi} \Sigma_\phi(p^2)=A_\phi+B_\phi p^2,$$ where the constants $A_\phi$ and $B_\phi$ receive a tree-level contribution from ${\cal L}_4$ and a one-loop contribution with a vertex from ${\cal L}_2$ (see Fig. \[4:8:selfenergy\]). For the tree-level contribution of ${\cal L}_4$ this is easily seen, because the Lagrangians of Eq. (\[4:8:l42phi\]) contain either exactly two derivatives of the fields or no derivatives at all. For example, the contact contribution for the $\eta$ reads $$-i\Sigma_\eta^{\rm contact}(p^2)=i 2\left[\frac{1}{2}a_\eta (ip_\mu)(-ip^\mu) -\frac{1}{2}b_\eta\right]=i(a_\eta p^2-b_\eta),$$ where, as usual, $\partial_\mu \phi$ generates $-ip_\mu$ and $ip_\mu$ for initial and final lines, respectively, and the factor two takes account of two combinations of contracting the fields with external lines. For the one-loop contribution the argument is as follows. The Lagrangian ${\cal L}_2^{4\phi}$ contains either two derivatives or no derivatives at all which, symbolically, can be written as $\phi\phi\partial\phi\partial\phi$ and $\phi^4$, respectively. The first term results in $M^2$ or $p^2$, depending on whether the $\phi$ or the $\partial \phi$ are contracted with the external fields. The “mixed” situation vanishes upon integration. The second term, $\phi^4$, does not generate a momentum dependence. As a specific example, we evaluate the pion-loop contribution to the $\pi^0$ self energy (see Fig. \[4:8:pi0seloop\]) by applying the Feynman rule of Eq. (\[4:6:mpipi2\]) for $a=c=3$, $p_a=p_c=p$, $b=d=j$, and $p_b=p_d=k$:[^70] $$\begin{aligned} \label{4:8:diag} && \frac{1}{2}\int \frac{d^4k}{(2\pi)^4}i \left[ \underbrace{\delta^{3j}\delta^{3j}}_{\mbox{$1$}} \frac{(p+k)^2-M_{\pi,0}^2}{F_0^2} +\underbrace{\delta^{33}\delta^{jj}}_{\mbox{$3$}} \frac{-M_{\pi,0}^2}{F_0^2} +\underbrace{\delta^{3j}\delta^{j3}}_{\mbox{$1$}} \frac{(p-k)^2-M_{\pi,0}^2}{F_0^2}\right.\nonumber\\ &&\left.-\frac{1}{3 F_0^2} \underbrace{(\delta^{3j}\delta^{3j}+\delta^{33}\delta^{jj} +\delta^{3j}\delta^{j3})}_{\mbox{5}}(2p^2+2k^2-4M_{\pi,0}^2)\right] \frac{i}{k^2-M_{\pi,0}^2+i0^+}\nonumber\\ &&=\frac{1}{2}\int \frac{d^4k}{(2\pi)^4}\frac{i}{3F_0^2} [-4p^2-4k^2+5 M^2_{\pi,0}] \frac{i}{k^2-M_{\pi,0}^2+i0^+},\end{aligned}$$ where $1/2$ is a symmetry factor, as explained in Sec. \[sec\_elwpcs\]. The integral of Eq. (\[4:8:diag\]) diverges and we thus consider its extension to $n$ dimensions in order to make use of the dimensional-regularization technique described in App. \[app\_drb\]. In addition to the loop-integral of Eq. (\[app:drb:im22\]), $$\begin{aligned} \label{4:8:im22} I(M^2,\mu^2,n)&=&\mu^{4-n}\int\frac{d^nk}{(2\pi)^n}\frac{i}{k^2-M^2+i0^+} \nonumber\\ &=&\frac{M^2}{16\pi^2}\left[ R+\ln\left(\frac{M^2}{\mu^2}\right)\right]+O(n-4),\end{aligned}$$ where $R$ is given in Eq. (\[4:7:R\]), we need $$\mu^{4-n}i\int \frac{d^n k}{(2\pi)^n}\frac{k^2}{k^2-M^2+i0^+}= \mu^{4-n}i\int \frac{d^n k}{(2\pi)^n}\frac{k^2-M^2+M^2}{k^2-M^2+i0^+},$$ where we have added $0=-M^2+M^2$ in the numerator. We make use of $$\mu^{4-n}i\int \frac{d^n k}{(2\pi)^n}=0$$ in dimensional regularization (see the discussion at the end of Appendix \[subsec\_jpin\]) and obtain $$\mu^{4-n}i\int \frac{d^n k}{(2\pi)^n}\frac{k^2}{k^2-M^2+i0^+}= M^2 I(M^2,\mu^2,n),$$ with $I(M^2,\mu^2,n)$ of Eq. (\[4:8:im22\]). The pion-loop contribution to the $\pi^0$ self energy is thus $$\frac{i}{6 F^2_0}(-4p^2+M^2_{\pi,0})I(M^2_{\pi,0},\mu^2,n),$$ which is indeed of the type discussed in Eq. (\[4:8:sigmaphi\]) and diverges as $n\to 4$. After analyzing all loop contributions and combining them with the contact contributions of Eqs. (\[4:8:ab\]), the constants $A_\phi$ and $B_\phi$ of Eq. (\[4:8:sigmaphi\]) are given by $$\begin{aligned} \label{4:8:AB} A_\pi&=&\frac{M^2_\pi}{F^2_0}\Bigg\{ \underbrace{-\frac{1}{6}I(M^2_\pi) -\frac{1}{6}I(M^2_\eta)-\frac{1}{3}I(M^2_K)}_{\mbox{one-loop contribution}} \nonumber\\ &&\underbrace{+32[(2m+m_s)B_0L_6+mB_0L_8]}_{\mbox{contact contribution}} \Bigg\},\nonumber\\ B_\pi&=&\frac{2}{3}\frac{I(M^2_\pi)}{F^2_0}+\frac{1}{3} \frac{I(M^2_K)}{F^2_0}-\frac{16B_0}{F^2_0}\left[ (2m+m_s)L_4+mL_5\right],\nonumber\\ A_K&=&\frac{M^2_K}{F^2_0}\Bigg\{\frac{1}{12}I(M^2_\eta) -\frac{1}{4}I(M^2_\pi)-\frac{1}{2}I(M^2_K)\nonumber\\ &&+32\left[(2m+m_s)B_0L_6+\frac{1}{2}(m+m_s)B_0L_8\right]\Bigg\},\nonumber\\ B_K&=&\frac{1}{4}\frac{I(M^2_\eta)}{F^2_0} +\frac{1}{4}\frac{I(M^2_\pi)}{F^2_0} +\frac{1}{2}\frac{I(M^2_K)}{F^2_0}\nonumber\\ &&-16 \frac{B_0}{F^2_0}\left[(2m+m_s)L_4+\frac{1}{2}(m+m_s)L_5\right], \nonumber\\ A_\eta&=&\frac{M^2_\eta}{F^2_0}\left[-\frac{2}{3}I(M^2_\eta)\right] +\frac{M^2_\pi}{F^2_0}\left[\frac{1}{6}I(M^2_\eta)-\frac{1}{2}I(M^2_\pi) +\frac{1}{3}I(M^2_K)\right]\nonumber\\ &&+\frac{M^2_\eta}{F^2_0}[16M^2_\eta L_8+32(2m+m_s)B_0L_6]\nonumber\\ &&+\frac{128}{9}\frac{B^2_0(m-m_s)^2}{F^2_0}(3L_7+L_8),\nonumber\\ B_\eta&=&\frac{I(M^2_K)}{F^2_0}-\frac{16}{F^2_0}(2m+m_s)B_0L_4 -8\frac{M^2_\eta}{F^2_0}L_5,\end{aligned}$$ where, for simplicity, we have suppressed the dependence on the scale $\mu$ and the number of dimensions $n$ in the integrals $I(M^2,\mu^2,n)$ \[see Eq. (\[4:8:im22\])\]. Furthermore, the squared masses appearing in the loop integrals of Eq. (\[4:8:AB\]) are given by the predictions of lowest order, Eqs. (\[4:3:mpi2\]) - (\[4:3:meta2\]). Finally, the integrals $I$ as well as the bare coefficients $L_i$ (with the exception of $L_7$) have $1/(n-4)$ poles and finite pieces. In particular, the coefficients $A_\phi$ and $B_\phi$ are [*not*]{} finite as $n\to 4$. The masses at ${\cal O}(p^4)$ are determined by solving the general equation $$\label{4:8:mse} M^2=M_0^2+\Sigma(M^2)$$ with the predictions of Eq. (\[4:8:sigmaphi\]) for the self energies, $$M^2=M_0^2+A+BM^2,$$ where the lowest-order terms, $M^2_0$, are given in Eqs.  (\[4:3:mpi2\]) - (\[4:3:meta2\]). We then obtain $$M^2=\frac{M_0^2+A}{1-B}=M_0^2(1+B)+A+ {\cal O}(p^6),$$ because $A={\cal O}(p^4)$ and $\{B, M_0^2\}={\cal O}(p^2)$. Expressing the bare coefficients $L_i$ in Eq. (\[4:8:AB\]) in terms of the renormalized coefficients by using Eq. (\[4:7:li\]), the results for the masses of the Goldstone bosons at ${\cal O}(p^4)$ read $$\begin{aligned} \label{4:8:mpi24} M^2_{\pi,4}&=&M^2_{\pi,2}\Bigg\{1+\frac{M^2_{\pi,2}}{32\pi^2F^2_0} \ln\left(\frac{M^2_{\pi,2}}{\mu^2}\right)-\frac{M^2_{\eta,2}}{96\pi^2F^2_0} \ln\left(\frac{M^2_{\eta,2}}{\mu^2}\right)\nonumber\\ &&+\frac{16}{F^2_0}\left[(2m+m_s)B_0(2L^r_6-L^r_4) +mB_0(2L^r_8-L^r_5)\right]\Bigg\},\\ \label{4:8:mk24} M^2_{K,4}&=&M^2_{K,2}\Bigg\{1+\frac{M^2_{\eta,2}}{48\pi^2F^2_0} \ln\left(\frac{M^2_{\eta,2}}{\mu^2}\right)\nonumber\\ &&+\frac{16}{F^2_0}\left[(2m+m_s)B_0(2L^r_6-L^r_4) +\frac{1}{2}(m+m_s)B_0(2L^r_8-L^r_5)\right]\Bigg\},\nonumber\\ &&\\ \label{4:8:meta24} M^2_{\eta,4}&=&M^2_{\eta,2}\left[1+\frac{M^2_{K,2}}{16\pi^2F^2_0} \ln\left(\frac{M^2_{K,2}}{\mu^2}\right) -\frac{M^2_{\eta,2}}{24\pi^2F^2_0}\ln\left(\frac{M^2_{\eta,2}}{\mu^2}\right) \right.\nonumber\\ &&\left.+\frac{16}{F^2_0}(2m+m_s)B_0(2L^r_6-L^r_4) +8\frac{M^2_{\eta,2}}{F^2_0}(2L^r_8-L^r_5)\right]\nonumber\\ &&+M^2_{\pi,2}\left[\frac{M^2_{\eta,2}}{96\pi^2F^2_0} \ln\left(\frac{M^2_{\eta,2}}{\mu^2}\right) -\frac{M^2_{\pi,2}}{32\pi^2F^2_0} \ln\left(\frac{M^2_{\pi,2}}{\mu^2}\right)\right.\nonumber\\ &&\left. +\frac{M^2_{K,2}}{48\pi^2F^2_0} \ln\left(\frac{M^2_{K,2}}{\mu^2}\right)\right]\nonumber\\ &&+\frac{128}{9}\frac{B^2_0(m-m_s)^2}{F^2_0} (3L^r_7+L^r_8).\end{aligned}$$ In Eqs. (\[4:8:mpi24\]) - (\[4:8:meta24\]) we have included the subscripts 2 and 4 in order to indicate from which chiral order the predictions result. First of all, we note that the expressions for the masses are finite. The bare coefficients $L_i$ of the Lagrangian of Gasser and Leutwyler must be infinite in order to cancel the infinities resulting from the divergent loop integrals. Furthermore, at ${\cal O}(p^4)$ the masses of the Goldstone bosons vanish, if the quark masses are sent to zero. This is, of course, what we had expected from QCD in the chiral limit but it is comforting to see that the self interaction in ${\cal L}_2$ (in the absence of quark masses) does not generate Goldstone boson masses at higher order. At ${\cal O}(p^4)$, the squared Goldstone boson masses contain terms which are analytic in the quark masses, namely, of the form $m^2_q$ multiplied by the renormalized low-energy coupling constants $L_i^r$. However, there are also non-analytic terms of the type $m^2_q \ln(m_q)$—so-called chiral logarithms—which do not involve new parameters. Such a behavior is an illustration of the mechanism found by Li and Pagels [@Li:1971vr], who noticed that a perturbation theory around a symmetry which is realized in the Nambu-Goldstone mode results in both analytic as well as non-analytic expressions in the perturbation. Finally, the scale dependence of the renormalized coefficients $L_i^r$ of Eq. (\[4:7:limu1mu2\]) is by construction such that it cancels the scale dependence of the chiral logarithms. Thus, physical observables do not depend on the scale $\mu$. Let us verify this statement by differentiating Eqs. (\[4:8:mpi24\]) - (\[4:8:meta24\]) with respect to $\mu$. Using Eq. (\[4:7:limu1mu2\]), $$L_i^r(\mu)=L_i^r(\mu')+\frac{\Gamma_i}{16\pi^2}\ln\left(\frac{\mu'}{\mu} \right),$$ we obtain $$\frac{d L_i^r(\mu)}{d\mu}=-\frac{\Gamma_i}{16\pi^2\mu}$$ and, analogously, for the chiral logarithms $$\frac{d}{d\mu}\ln\left(\frac{M^2}{\mu^2}\right)= 2\frac{d}{d\mu}\left[\ln(M)-\ln(\mu)\right]=-\frac{2}{\mu}.$$ As a specific example, let us differentiate the expression for the pion mass $$\begin{aligned} \frac{d M^2_{\pi,4}}{d\mu}&=&\frac{M^2_{\pi,2}}{16\pi^2\mu F_0^2}\Bigg\{ \frac{M^2_{\pi,2}}{2}(-2)-\frac{M^2_{\eta,2}}{6}(-2)\\ &&+16[(2m+m_s)B_0(-2\Gamma_6+\Gamma_4)+mB_0(-2\Gamma_8+\Gamma_5)]\Bigg\}\\ &=&\frac{M^2_{\pi,2}}{16\pi^2\mu F_0^2}\Bigg\{-2B_0m +\frac{2}{9}(m+2m_s)B_0\\ &&+16\Bigg[(2m+m_s)B_0 \underbrace{\left(-2\frac{11}{144}+\frac{1}{8}\right)}_{\mbox{$-\frac{1}{36}$}} +mB_0\underbrace{\left(-2\frac{5}{48}+\frac{3}{8}\right) }_{\mbox{$\frac{1}{6}$}} \Bigg]\Bigg\}\\ &=&\frac{M^2_{\pi,2}}{16\pi^2\mu F_0^2}\Bigg\{ B_0m \left(-2+\frac{2}{9}-\frac{8}{9}+\frac{8}{3}\right) +B_0m_s\left(\frac{4}{9}-\frac{16}{36}\right)\Bigg\}\\ &=&0,\end{aligned}$$ where we made use of the $\Gamma_i$ of Table \[4:7:tableli\]. ### The Electromagnetic Form Factor of the Pion {#subsec_emffp} As a second application at ${\cal O}(p^4)$, we discuss the electromagnetic (or vector) form factor of the pion in SU(2)$\times$SU(2) chiral perturbation theory. We will work with two commonly used versions of the ${\cal O}(p^4)$ $\mbox{SU(2)}\times\mbox{SU(2)}$ mesonic Lagrangian [@Gasser:1983yg; @Gasser:1987rb] which are related by a field transformation (see App. \[app\_sec\_glvgss\]). Furthermore, we will perform the calculation with the two parameterizations for $U$ of Eqs. (\[4:6:u1\]) and (\[4:6:u2\]). We will thus be able to extend the observations of Sec. \[subsec\_pps\] regarding the invariance of physical results under a change of variables to the one-loop level. According to Eq. (\[2:4:rlasu2\]), in the two-flavor sector the coupling to the electromagnetic field ${\cal A}_\mu$ contains both isoscalar and isovector terms: $${\cal L}_{\rm ext}=\bar{q}\gamma^\mu\left(\frac{1}{3}v_\mu^{(s)}+v_\mu \right)q =-e{\cal A}_\mu\bar{q}\gamma^\mu\left(\frac{1}{6}+\frac{\tau_3}{2}\right)q =-e {\cal A}_\mu J^\mu,$$ i.e. $$\begin{aligned} \label{4:9rlasu2} v_\mu&=&r_\mu=l_\mu=-e\frac{\tau_3}{2}{\cal A}_\mu,\\ \label{4:9vsasu2} v_\mu^{(s)}&=&-\frac{e}{2}{\cal A}_\mu.\end{aligned}$$ When evaluating the electromagnetic current operator $$J^\mu=\frac{1}{6}\bar{q}\gamma^\mu q+\bar{q}\gamma^\mu\frac{\tau_3}{2} q$$ between $|\pi^i(p)\rangle$ and $\langle \pi^j(p')|$, the isoscalar first term does not contribute,[^71] and the matrix element of the electromagnetic current operator must be of the form[^72] $$\langle \pi^j(p')|J^\mu(0)|\pi^i(p)\rangle=i\epsilon_{3ij}(p'+p)^\mu F(q^2), \quad q=p'-p.$$ In other words, we only need to consider Eq. (\[4:9rlasu2\]) which corresponds to a coupling to the third component of the isovector current operator. In the calculations that follow we make use of the parameterization of Eq. (\[4:6:u1\]) for the SU(2) matrix $U(x)$: $$\label{4:9:sqrt} U(x)=\frac{1}{F_0}\left[\sigma(x)+i\vec{\tau}\cdot\vec{\pi}(x) \right],\quad \sigma(x)=\sqrt{F^2_0-\vec{\pi}\,^2(x)},$$ but we will comment along the way on the features, which would differ when using the parameterization of Eq. (\[4:6:u2\]). (The equivalence theorem guarantees that physical observables do not depend on the specific choice of parameterization of $U$ [@Chisholm; @Kamefuchi:sb].) The covariant derivative of $U$ with the external fields of Eq. (\[4:9rlasu2\]) reads \[see Eq. (\[4:5:kaa\])\] $$D_\mu U=\partial_\mu U+\frac{i}{2}e{\cal A}_\mu[\tau_3,U]$$ and generates, when inserted into the lowest-order Lagrangian of Eq. (\[4:5:l2\]), the interaction term $$\label{4:9:l2gpp} {\cal L}_2^{\gamma \pi\pi}=-e\epsilon_{3ij}\pi_i\partial^\mu \pi_j {\cal A}_\mu.$$ At ${\cal O}(p^2)$, therefore, the interaction is that of point-like pions with form factor $F(q^2)=1$, resulting in the Feynman amplitude (see Fig.\[4:9:pionffl2\]) $$\label{4:9:m2gpp} e\epsilon_{3ij}\epsilon\cdot(p'+p),$$ where $\epsilon$ denotes the polarization vector of the external real or virtual photon.[^73] In particular, using the parameterization of Eq. (\[4:9:sqrt\]), all interaction terms containing one electromagnetic field and $2n$ pions vanish for $n\geq 2$. This is not the case for the exponential parameterization of Eq. (\[4:6:u2\]) which generates the more complicated interaction Lagrangian $$\label{4:9:l2exp} -e\epsilon_{3ij}\phi_i\partial^\mu \phi_j{\cal A}_\mu \underbrace{\frac{F_0^2}{|\vec\phi|^2} \sin^2\left(\frac{|\vec{\phi}|}{F_0}\right)}_{ \mbox{$1-\frac{1}{3}\frac{|\vec{\phi}|^2}{F_0^2}+\cdots$}}$$ which is the same as Eq. (\[4:9:l2gpp\]) only at lowest order in the fields. At ${\cal O}(p^4)$ we need to consider a contact term of ${\cal L}_4$ (Fig. \[4:9:pionffl4\]) and one-loop diagrams with vertices from ${\cal L}_2$ (Figs. \[4:9:pionffloop1\] and \[4:9:pionffloop2\]). We will first work with the Lagrangian of Gasser and Leutwyler [@Gasser:1983yg], Eq. (\[app:glvgss:l4gl\]) of Appendix \[app\_sec\_glvgss\],[^74] $$\label{4:9:l4gl} {\cal L}^{\rm GL}_4=\cdots +i\frac{l_6}{2}\mbox{Tr}[ f^R_{\mu\nu} D^{\mu} U (D^{\nu} U)^{\dagger} + f^L_{\mu\nu} (D^{\mu} U)^{\dagger} D^{\nu} U]+\cdots,$$ which produces the contact interaction $$\label{4:9:l4gpp} {\cal L}_4^{\gamma \pi\pi}=e \frac{l_6}{F_0^2} \epsilon_{3ij}\partial^\mu\pi_i\partial^\nu \pi_j {\cal F}_{\mu\nu},$$ resulting in the Feynman amplitude $$\label{4:9:m4gpp} \frac{e l_6 \epsilon_{3ij}}{F_0^2} \left[-q^2 \epsilon\cdot(p'+p)+\epsilon\cdot(p'-p)(p'^2-p^2)\right]$$ which vanishes for real photons, $q^2=q\cdot\epsilon=0$. The second term vanishes if both pions are on the mass shell, $p^2=p'^2=M_\pi^2$, but can be of relevance if the vertex is used as an intermediate building block in a calculation such as virtual Compton scattering off the pion [@Unkmeir:1999md]. The Feynman amplitude resulting from Eq. (\[4:9:l4gl\]) is the same for both parameterizations. When using the ${\cal L}_4$ Lagrangian of Ref. [@Gasser:1987rb], Eq. (\[app:glvgss:l4gss\]), one obtains an additional contact interaction, $$\label{4:9:addl4gpp} -2 e \frac{l_4 M^2_{\pi,2}}{F_0^2} \epsilon_{3ij}\pi_i\partial^\mu \pi_j {\cal A}_\mu,$$ where $M^2_{\pi,2}=2 B_0 m$. For both parameterizations of $U$ we obtain the additional term $$\label{4:9:addlmgpp} \frac{2 l_4 M^2_{\pi,2}}{F_0^2} e \epsilon_{3ij}\epsilon\cdot(p'+p).$$ Let us now turn to the one-loop diagram of Fig. \[4:9:pionffloop1\]. The corresponding Feynman amplitude in the parameterization of Eq. (\[4:9:sqrt\]) using the pion-pion vertex of Eq. (\[4:6:mpipi1\]) reads $$\begin{aligned} \label{4:9:loopc1} &&\frac{1}{2} \frac{ie\epsilon_{3ij}}{F_0^2} \int \frac{d^4 k}{(2\pi)^4} \frac{(2\epsilon\cdot k+\epsilon\cdot q)[(2p+q)\cdot(2k+q)]}{ [k^2-M^2_{\pi,2}+i0^+][(k+q)^2-M^2_{\pi,2}+i0^+]},\end{aligned}$$ where the $1/2$ is a symmetry factor. The integral diverges and its extension to $n$ dimensions is given by $$\begin{aligned} \label{4:9:loopc1a} &&\frac{1}{2} \frac{e\epsilon_{3ij}}{F_0^2} \left\{\epsilon\cdot(p'+p)4 q^2 B_{21}(q^2,M_{\pi,2}^2)\right.\nonumber\\ &&\left. +\epsilon\cdot q (p'+p)\cdot q [4 B_{20}(q^2,M_{\pi,2}^2) +4 B_1(q^2,M_{\pi,2}^2) +B_0(q^2,M_{\pi,2}^2)]\right\},\nonumber\end{aligned}$$ where the functions $B_0$, $B_1$, $B_{20}$, and $B_{21}$ are defined in Eqs. (\[app:ipipib\]), (\[app:B1\]), and (\[app:B20B21\]) of Appendix \[app\_sec\_olims\]. Inserting the results of Eqs. (\[app:q2B21\]) and (\[app:B20\]) the one-loop contribution of Fig. \[4:9:pionffloop1\] finally reads $$\begin{aligned} \label{4:9:loopc1b} &&e\epsilon_{3ij}\left\{\epsilon\cdot(p'+p) \frac{M_{\pi,2}^2}{16\pi^2F_0^2} \left[R+\ln\left(\frac{M_{\pi,2}^2}{\mu^2}\right)\right]\right.\nonumber\\ && -\frac{1}{96\pi^2 F_0^2}[q^2 \epsilon\cdot(p'+p)-\epsilon\cdot q (p'^2-p^2)] \nonumber\\ &&\left.\times \left[R+\ln\left(\frac{M_{\pi,2}^2}{\mu^2}\right) +\frac{1}{3}+\left(1-\frac{M^2_{\pi,2}}{q^2}\right) J^{(0)}\left(\frac{q^2}{M_{\pi,2}^2}\right)\right]\right\}+O(n-4).\nonumber\\\end{aligned}$$ The (infinite) contribution of the first term will be precisely canceled by taking the pion wave function renormalization into account. The second structure is separately gauge invariant and also contains an infinite piece which will be canceled by a corresponding infinite piece of the bare coefficient $l_6$ \[see Eq. (\[4:9:m4gpp\])\]. Finally, a calculation of the one-loop diagram of Fig. \[4:9:pionffloop1\] with the exponential parameterization of Eq. (\[4:6:u2\]) and the pion-pion vertex of Eq. (\[4:6:mpipi2\]) yields exactly the same result as Eq. (\[4:9:loopc1b\]). Using the exponential parameterization there is a $4\pi\gamma$ vertex at ${\cal O}(p^4)$ \[see Eq. (\[4:9:l2exp\])\] resulting in the additional loop diagram of Fig. \[4:9:pionffloop2\]. The corresponding contribution in dimensional regularization, $$\label{4:9:loopc2a} -\frac{5}{3} e\epsilon_{3ij} \epsilon\cdot(p'+p) \frac{M^2_{\pi,2}}{16\pi^2 F_0^2}\left[R+\ln\left(\frac{M^2_{\pi,2}}{\mu^2} \right) \right]+O(n-4),$$ also generates an infinite contribution to the vertex at zero four-momentum transfer. The renormalized vertex is obtained by adding the bare contributions and multiplying the result by a factor $\sqrt{Z_\pi}$ for each external pion line. The wave function renormalization constant $Z_\pi$ is not an observable and depends on both the parameterization for $U$ and the ${\cal O}(p^4)$ Lagrangian. The corresponding results are summarized in Table \[app:dp:tab:abz\] of Appendix \[app\_sec\_dp\]. We add the bare contributions which were obtained using the parameterization of Eq. (\[4:9:sqrt\]) and the ${\cal O}(p^4)$ Lagrangian of Eq. (\[4:9:l4gpp\]), Eqs. (\[4:9:m2gpp\]), (\[4:9:m4gpp\]), and (\[4:9:loopc1b\]), and multiply the result by the appropriate wave function renormalization constant \[see entry “GL, Eq. (\[app:dp:sqrt\])” of Table \[app:dp:tab:abz\]\], $$\begin{aligned} \label{4:9:mr} \lefteqn{{\cal M}_R=e\epsilon_{3ij}\left( \epsilon\cdot(p'+p)\left\{1+\frac{M_{\pi,2}^2}{16\pi^2F_0^2} \left[R+\ln\left(\frac{M_{\pi,2}^2}{\mu^2}\right)\right]\right\}\right. }\nonumber\\ &&-\frac{q^2 \epsilon\cdot(p'+p)-\epsilon\cdot q (p'^2-p^2)}{F_0^2} \nonumber\\ &&\left.\times \left\{l_6 +\frac{1}{96\pi^2} \left[R+\ln\left(\frac{M_{\pi,2}^2}{\mu^2}\right) +\frac{1}{3}+\left(1-\frac{M^2_{\pi,2}}{q^2}\right) J^{(0)}\left(\frac{q^2}{M_{\pi,2}^2}\right)\right]\right\}\right) \nonumber\\ &&\times \left\{1-\frac{M_{\pi,2}^2}{16\pi^2F_0^2} \left[R+\ln\left(\frac{M_{\pi,2}^2}{\mu^2}\right)\right]\right\} +O(n-4).\end{aligned}$$ The factor $\epsilon\cdot(p'+p)$ counts as ${\cal O}(p^2)$, because the external electromagnetic field, represented by the polarization vector $\epsilon$, counts as ${\cal O}(p)$ \[see Eq. (\[4:5:powercounting\])\]. It is multiplied by $$\begin{aligned} 1+I(M^2_{\pi,2})&=&1+\frac{M^2_{\pi,2}}{16\pi^2}\left[ R+\ln\left(\frac{M^2_{\pi,2}}{\mu^2}\right)\right]+O(n-4)\\ &=&1+{\cal O}(p^2).\end{aligned}$$ Since the wave function renormalization constant is $Z_\pi=1-I(M^2_{\pi,2})=1+{\cal O}(p^2)$ (see Appendix \[app\_sec\_dp\]), it is only the tree-level contribution derived from ${\cal L}_2$ which gets modified. The product $[1+I(M^2_{\pi,2})][1-I(M^2_{\pi,2})]=1+{\cal O}(p^4)$ is such that the renormalized vertex is properly normalized to the charge at ${\cal O}(p^4)$. The factor $[q^2 \epsilon\cdot(p'+p)-\epsilon\cdot q (p'^2-p^2)]$ is ${\cal O}(p^4)$ and thus the apparent infinity $R$ cannot be canceled through the wave function renormalization. Here it is the connection between the bare parameter $l_6$ and the renormalized parameter $l_6^r(\mu)$, $l_6^r=l_6+R/(96\pi^2)$, which cancels the divergence \[see Eq. (9.6) of Ref. [@Gasser:1983yg]\]. Moreover, the explicit dependence on the renormalization scale $\mu$ cancels with a corresponding scale dependence of the parameter $l_6^r$. Finally, using the exponential parameterization or the Lagrangian of Ref. [@Gasser:1987rb] results in the same expression as Eq. (\[4:9:mr\]). The additional contributions from Eqs. (\[4:9:addl4gpp\]) and/or (\[4:9:loopc2a\]) to the unrenormalized vertex are precisely canceled by modified wave function renormalization constants, resulting in the same renormalized vertex. On the mass shell $p^2=p'^2=M_\pi^2$, and we obtain for the electromagnetic form factor [@Gasser:1983yg] $$\label{4:9:emffpi} F(q^2)=1-l_6^r\frac{q^2}{F^2_\pi}-\frac{1}{6}\frac{q^2}{(4\pi F_\pi)^2} \left[\ln\left(\frac{M^2_\pi}{\mu^2}\right) +\frac{1}{3} +\left(1-4\frac{M^2_\pi}{q^2}\right)J^{(0)}\left(\frac{q^2}{M^2_\pi} \right)\right],$$ where we replaced the ${\cal O}(p^2)$ quantities $F_0$ and $M^2_{\pi,2}$ by their physical values, the error introduced being of order ${\cal O}(p^6)$. Given a spherically symmetric charge distribution $e Z \rho(r)$ normalized so that $\int d^3 x \rho(r)=1$, the form factor $F(|\vec{q}\,|)$ in a nonrelativistic framework is given by $$F(|\vec{q}\,|)=\int d^3 x e^{i\vec{q}\cdot\vec{x}}\rho(r) =4\pi \int_0^\infty dr r^2 j_0(|\vec{q}\,|r) \rho(r) =1-\frac{1}{6}|\vec{q}\,|^2\langle r^2\rangle+\cdots,$$ where $\langle r^2\rangle$ denotes the mean square radius.[^75] In analogy, the Lorentz-invariant form factor of Eq. (\[4:9:emffpi\]) is expanded for small $q^2$ as[^76] $$F(q^2)=1+\frac{q^2}{6}\langle r^2\rangle +\cdots,$$ and the charge radius of the pion is [*defined*]{} as $$\langle r^2\rangle_\pi=6 \left.\frac{d F(q^2)}{dq^2}\right|_{q^2=0} = -\frac{6}{F_\pi^2}\left\{l_6^r(\mu)+\frac{1}{96\pi^2} \left[1+\ln\left(\frac{M^2_\pi}{\mu^2}\right)\right]\right\},$$ where we made use of $J^{(0)}(x)=-x/6+O(x^2)$. Following Ref. [@Gasser:1983yg], we introduce a scale-independent quantity (see Appendix \[app\_sec\_glvgss\]) $$\bar{l}_6=-96\pi^2 l_6^r(\mu)-\ln\left(\frac{M^2_\pi}{\mu^2}\right)$$ which can be determined using the empirical information on the charge radius of the pion: $\bar{l}_6=16\pi^2 F_\pi^2\langle r^2\rangle_\pi+1$. In a two-loop calculation of the vector form factor [@Bijnens:1998fm], higher-order terms in the chiral expansion terms were also taken into account and a fit to several experimental data sets was performed with the result $\bar{l}_6=16.0\pm 0.5\pm 0.7$, where the last error is of theoretical origin. Once the value of the parameter $\bar{l}_6$ has been determined it can be used to predict other processes such as, e.g., virtual Compton scattering off the pion [@Unkmeir:1999md; @Unkmeir:2001gw]. The results for the electromagnetic form factors of the charged pion, and the charged and neutral kaons in SU(3)$\times$SU(3) chiral perturbation theory at ${\cal O}(p^4)$ can be found in Refs. [@Gasser:1984ux; @Rudy:1994qb]. The calculation is very similar to the SU(2)$\times$SU(2) case and the mean square radii of the charged pions and kaons are dominated by the low-energy parameter $L^r_9$, whereas the one-loop diagrams generate a small contribution only:[^77] $$\begin{aligned} \label{4:9:r2pipkp} \langle r^2\rangle_{\pi^+}&=&12 \frac{L^r_9}{F^2_0} -\frac{1}{32 \pi^2 F^2_0}\left[3 + 2 \ln\left(\frac{M^2_{\pi,2}}{\mu^2}\right) +\ln\left(\frac{M^2_{K,2}}{\mu^2}\right)\right]\nonumber\\ &=&(0.37\pm 0.04+0.07)\,\mbox{fm}^2=(0.44\pm 0.04)\,\mbox{fm}^2,\nonumber\\ \langle r^2\rangle_{K^+}&=&12 \frac{L^r_9}{F^2_0} -\frac{1}{32 \pi^2 F^2_0}\left[3 + 2 \ln\left(\frac{M^2_{K,2}}{\mu^2}\right) +\ln\left(\frac{M^2_{\pi,2}}{\mu^2}\right)\right]\nonumber\\ &=&(0.37 \pm 0.04 + 0.03)\,\mbox{fm}^2 = (0.40 \pm 0.04)\,\mbox{fm}^2.\end{aligned}$$ In Ref. [@Gasser:1984ux] the empirical value $\langle r^2\rangle_\pi=(0.439\pm 0.030)\,\mbox{fm}^2$ of [@Dally:1982zk] was used to fix $L^r_9$.[^78] The result for the mean square radius of the charged kaon is then a prediction which has to be compared with the empirical values $\langle r^2 \rangle_{K^-}=(0.28\pm 0.05)\,\mbox{fm}^2$ of [@Dally:1980dj] and $\langle r^2 \rangle_{K^-}=(0.34\pm 0.05)\,\mbox{fm}^2$ of [@Amendolia:1986ui]. In Ref. [@Bijnens:2002hp] the empirical data on the charged pion and kaon form factors were analyzed at two-loop order and the low-energy constant including the $p^6$ terms was determined as $L_9^r(\mu=770\,\mbox{MeV}) =(5.93\pm 0.43)\times 10^{-3}$. At ${\cal O}(p^4)$ the form factor of the $K^0$ receives one-loop contributions only and thus is predicted in terms of the pion-decay constant and the Goldstone boson masses. The mean square radius is given by $$\langle r^2\rangle_{K^0}=-\frac{1}{32\pi^2F^2_0}\ln\left(\frac{M^2_K}{M^2_\pi}\right) =-0.037\, \mbox{fm}^2$$ which has to be compared with the empirical value $\langle r^2\rangle_{K^0}= (-0.054\pm 0.026)\, \mbox{fm}^2$ [@Molzon:1978py]. For a two-loop analysis of the neutral-kaon form factor, see Refs. [@Post:1997dk; @Post:2000gk]. Since the neutral pion and the eta are their own antiparticles, their electromagnetic vertices vanish because of charge conjugation symmetry as noted in footnote \[fnic\] for the case of the $\pi^0$. Chiral Perturbation Theory at ${\cal O}(p^6)$ {#sec_cptop6} --------------------------------------------- Mesonic chiral perturbation theory at order ${\cal O}(p^4)$ has led to a host of successful applications and may be considered as a full-grown and mature area of low-energy particle physics. In this section we will briefly touch upon its extension to ${\cal O}(p^6)$ , which naturally divides into the even- and odd-intrinsic-parity sectors. Calculations in the even-intrinsic-parity sector start at ${\cal O}(p^2)$ and two-loop calculations at ${\cal O}(p^6)$ are thus of next-to-next-to-leading order (NNLO). NNLO calculations at ${\cal O}(p^6)$ have been performed for $\gamma\gamma\to\pi^0\pi^0$ [@Bellucci:1994eb], vector two-point functions [@Golowich:1995kd; @Maltman:1995jg; @Durr:1999dp; @Amoros:1999dp] and axial-vector two-point functions [@Amoros:1999dp], $\pi\pi$ scattering [@Bijnens:1995yn], $\gamma\gamma\to\pi^+\pi^-$ [@Burgi:1996mm], $\tau\to \pi\pi \nu_\tau$ [@Colangelo:1996hs], $\pi\to l \nu \gamma$ [@Bijnens:1996wm], Sirlin’s combination of SU(3) form factors [@Post:1997dk], scalar and electromagnetic form factors of the pion [@Bijnens:1998fm], the $K\to\pi\pi l\nu$ ($K_{l4}$) form factors [@Amoros:2000qq], the electromagnetic form factor of the $K^0$ [@Post:2000gk], the $K\to\pi l\nu$ ($K_{l3}$) form factors [@Post:2001si], and the electromagnetic form factors of pions and kaons in SU(3)$\times$SU(3) ChPT [@Bijnens:2002hp]. Further applications deal with more technical aspects such as the evaluation of specific two-loop integrals [@Post:1996gg; @Gasser:1998qt] and the renormalization of the even-intrinsic-parity Lagrangian at ${\cal O}(p^6)$ [@Bijnens:1999hw]. The odd-intrinsic-parity sector starts at ${\cal O}(p^4)$ with the anomalous WZW action, as discussed in Sec. \[sec\_ewzwa\]. In this sector next-to-leading-order (NLO), i.e. one-loop calculations, are of ${\cal O}(p^6)$. It has been known for some time that quantum corrections to the WZW classical action do not renormalize the coefficient of the WZW term . The counter terms needed to renormalize the one-loop singularities at ${\cal O}(p^6)$ are of a conventional chirally invariant structure. The inclusion of the photon as a dynamical degree of freedom in the odd-intrinsic-parity sector has been discussed in Ref.  [@Ananthanarayan:2002kj]. For an overview of applications in the odd-intrinsic-parity sector, we refer to Ref. [@Bijnens:xi]. A two-loop calculation at ${\cal O}(p^8)$ for $\gamma\pi\to\pi\pi$ was performed in Ref. [@Hannah:2001ee]. Here, we will mainly be concerned with some aspects of the construction of the most general mesonic chiral Lagrangian at ${\cal O}(p^6)$ and discuss as an application a two-loop calculation for the $s$-wave $\pi\pi$ scattering lengths [@Bijnens:1995yn]. ### The Mesonic Chiral Lagrangian at Order ${\cal O}(p^6)$ {#subsec_mclop6} The rapid increase in the number of free parameters when going from ${\cal L}_2$ to ${\cal L}_4$ naturally leads to the expectation of a very large number of chirally invariant structures at ${\cal O}(p^6)$. One of the problems with the construction of effective Lagrangians at higher orders is that it is far too easy to think of terms satisfying the necessary criteria of Lorentz invariance and invariance under the discrete symmetries as well as chiral transformations. To our knowledge there is neither a general formula, even at ${\cal O}(p^4)$, for determining the number of independent structures to expect nor an algorithm to decide whether a set of given structures is independent or not. Experience has shown that for almost any sector of higher-order effective chiral Lagrangians the number of terms found to be independent has gone down with time (see Refs.  for the odd-intrinsic-parity sector, [@Fearing:1994ga; @Bijnens:1999sh] for the even-intrinsic-parity sector, and [@Ecker:1995rk; @Fettes:2000gb] for the heavy-baryon $\pi N$ Lagrangian, respectively). For that reason, it is important to define a strategy for obtaining all of the independent terms without generating a lot of extraneous terms which have to be eliminated by hand. In the following, we will outline the main principles entering the construction of the ${\cal O}(p^6)$ Lagrangian and refer the reader to Refs. [@Fearing:1994ga; @Bijnens:1999sh; @Ebertshauser:2001nj; @Bijnens:2001bb] for more details. The effective Lagrangian is constructed from the elements $U$, $U^\dagger$, $\chi$, $\chi^\dagger$, and the field strength tensors $f^L_{\mu\nu}$ and $f^R_{\mu\nu}$ (see Sec. \[sec\_cel\], in particular, Table \[4:5:table\_trafprop\]). The external fields $l_\mu$ and $r_\mu$ only appear in the field strength tensors or the covariant derivatives which we define as $$\begin{aligned} \label{4:10:covder1} A\stackrel{G}{\mapsto}V_RAV_L^\dagger: &\quad&D_\mu A\equiv\partial_\mu A-ir_\mu A+iA l_\mu,\quad\mbox{e.g.,} \,\, U,\chi,\nonumber\\ B\stackrel{G}{\mapsto}V_LBV_R^\dagger: &\quad&D_\mu B\equiv\partial_\mu B+iBr_\mu -il_\mu B, \quad\mbox{e.g.,}\,\, U^\dagger,\chi^\dagger,\nonumber\\ C\stackrel{G}{\mapsto}V_RCV_R^\dagger: &\quad&D_\mu C\equiv\partial_\mu C-ir_\mu C+iCr_\mu, \quad\mbox{e.g.,}\,\, f^R_{\mu\nu},\nonumber\\ D\stackrel{G}{\mapsto}V_LDV_L^\dagger: &\quad&D_\mu D\equiv\partial_\mu D-il_\mu D+iDl_\mu, \quad\mbox{e.g.,}\,\, f^L_{\mu\nu},\nonumber\\ E\stackrel{G}{\mapsto}E: &\quad&D_\mu E\equiv\partial_\mu E,\quad \mbox{e.g.,}\,\,\mbox{Tr}(\chi\chi^\dagger).\end{aligned}$$ In other words, the covariant derivative knows about the transformation property under $G=\mbox{SU(3)}_L\times \mbox{SU(3)}_R$ of the object it acts on and adjusts itself accordingly. With such a convention a product rule analogous to that for ordinary derivatives holds. Given the product $Z=XY$ where $X,Y,Z$ have, according to Eq. (\[4:10:covder1\]), well-defined but not necessarily the same transformation behavior, the product rule $$\label{4:10:pr} D_\mu Z = D_\mu (XY)=(D_\mu X) Y+X( D_\mu Y)$$ applies, which can be easily verified using the definitions of Eq. (\[4:10:covder1\]). This product rule is valuable as an intermediate step in a number of the derivations of various relations. In order to avoid unnecessary and tedious repetitions during the process of construction, one would like to perform as many manipulations as possible on a formal level without explicit reference to the specific building blocks. It is thus convenient to handle the external field terms $\chi$, $f^R_{\mu\nu}$, and $f^L_{\mu\nu}$ in the same way. To that end we define $$\label{4:10:ghdefinition} G^{\mu\nu}\equiv f_R^{\mu\nu}U+Uf_L^{\mu\nu},\quad H^{\mu\nu}\equiv f_R^{\mu\nu}U-Uf_L^{\mu\nu},$$ and introduce $\chi^{\mu\nu}$ as a common abbreviation for any of the building blocks $\chi$ ($\equiv \chi^{\mu\mu}$), $G^{\mu\nu}$ and $H^{\mu\nu}$ ($\mu\neq\nu$). With these definitions, we have only two basic building blocks $U$, $\chi^{\mu\nu}$, covariant derivatives acting on them and the respective adjoints. Due to the product rule of Eq. (\[4:10:pr\]) it is not necessary to consider derivatives acting on products of these basic terms. All building blocks then transform as $U$ (or $U^\dagger$). In terms of the momentum expansion, $U$ is of order 1, $\chi^{\mu\nu}$ of order $p^2$ and each covariant derivative $D_\mu$ of order $p$. Up to this point we have treated a building block and its adjoint on a different footing which we will now remedy by defining the Hermitian and anti-Hermitian combinations $$\label{4:10:apm} (A)_\pm=u^\dagger A u^\dagger \pm u A^\dagger u,$$ where $A$ is taken as $\chi^{\mu\nu}$ or $D_\mu U$, or as some number of covariant derivatives acting on $\chi^{\mu\nu}$ or $D_\mu U$.[^79] Here $u$ is defined as the square root of $U$, i.e., $u^2\equiv U$. In order to discuss the transformation behavior of Eq.  (\[4:10:apm\]), we define the SU(3)-valued function $K(V_L,V_R,U)$, referred to as the compensator field [@Ecker:1995gg], through [@Gasser:1987rb] $$\label{4:10:kdef} u(x)\mapsto u'(x)=\sqrt{V_R U V_L^\dagger}\equiv V_R u K^{-1}(V_L,V_R,U),$$ from which one obtains $$\label{4:10:k} K(V_L,V_R,U)=u'^{-1} V_R u=\sqrt{V_R U V_L^\dagger}^{-1}V_R\sqrt{U}.$$ From a group-theoretical point of view, $K$ defines a nonlinear realization of SU(3)$\times$SU(3) [@Gasser:1987rb], because[^80] $$\begin{aligned} \label{4:10:kreal} \lefteqn{K(V_{L1}, V_{R1}, V_{R2} U V_{L2}^\dagger) K (V_{L2},V_{R2},U)}\nonumber\\ &=& \sqrt{V_{R1}(V_{R2}U V_{L2}^\dagger) V_{L1}^\dagger}^{-1} V_{R1} \sqrt{V_{R2}U V_{L2}^\dagger} \sqrt{V_{R2}U V_{L2}^\dagger}^{-1} V_{R2}\sqrt{U}\nonumber\\ &=&\sqrt{V_{R1} V_{R2} U (V_{L1} V_{L2})^\dagger}^{-1} V_{R1} V_{R2} \sqrt{U} \nonumber\\ &=&K((V_{L1} V_{L2}), (V_{R1}V_{R2}),U).\end{aligned}$$ It is important to note that the first $K$ has the transformed $U'=V_{R2} U V_{L2}^\dagger$ as its argument. With these definitions, the building blocks $(A)_\pm$ transform as $$\label{4:10:apmtrans} (A)_\pm\mapsto K (A)_\pm K^\dagger.$$ The corresponding covariant derivative is defined as[^81] $$\label{4:10:covder} \nabla_\mu(A)_\pm\equiv\partial_\mu(A)_\pm+[\Gamma_\mu,(A)_\pm],$$ where $\Gamma_\mu$ is the so-called connection [@Ecker:1995gg], and is given by $$\label{4:10:chiralconnection} \Gamma_\mu=\frac{1}{2}\left[u^\dagger,\partial_\mu u\right] -\frac{i}{2}u^\dagger r_\mu u-\frac{i}{2}u l_\mu u^\dagger.$$ As usual, the covariant derivative transforms in the same way as the object it acts on, $$\nabla_\mu(A)_\pm\stackrel{G}{\mapsto} K \nabla_\mu(A)_\pm K^\dagger.$$ Invariants under $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ are constructed by forming products of objects, each transforming as $K\cdots K^\dagger$, and then taking the trace. For example, consider the trace $$\label{4:10:twoobjects} \mbox{Tr}[(A_1)_\pm \cdots (A_n)_\pm] \stackrel{G}{\mapsto} \mbox{Tr}[K (A_1)_\pm K^\dagger \cdots K (A_n)_\pm K^\dagger] =\mbox{Tr}[(A_1)_\pm \cdots (A_n)_\pm],$$ where we made use of $K^\dagger K=1$ and the invariance of the trace under cyclic permutations. Obviously, products of such traces are also invariant. Basically, the construction of the most general Lagrangian then proceeds by forming products of elements $(A)_\pm$, where $A$ is either $\chi^{\mu\nu}$ or $D^\mu U$ or covariant derivatives of these objects, taking appropriate traces, and forming Lorentz scalars by contracting the Lorentz indices with the metric tensor $g^{\mu\nu}$ or the totally antisymmetric tensor $\epsilon^{\mu\nu\rho\sigma}$. After having chosen the set of building blocks and their transformation behavior, we define a strategy concerning the order of constructing invariants at ${\cal O}(p^6)$. As shown in Refs. [@Fearing:1994ga; @Ebertshauser:2001nj], by applying the product rule, it is sufficient to restrict oneself to $(D^m U)_-$, $(D^n G)_+$, $(D^n H)_+$, and $(D^n \chi)_\pm$ ($m, n$ integer with $m>0, n\ge 0$), because the other combinations $(D^m U)_+$, $(D^n G)_-$, and $(D^n H)_-$ can either be expressed in terms of these or vanish. One then immediately finds that all possible terms at ${\cal{O}}{(p^6)}$ can either include no, one, two, or three $(D^n \chi_{\mu\nu})_\pm$ blocks which naturally defines four distinct levels to be considered: 1. terms with six $D_\mu$’s; 2. terms with four $D_\mu$’s and one $\chi^{\mu\nu}$; 3. terms with two $D_\mu$’s and two $\chi^{\mu\nu}$’s; 4. terms with three $\chi^{\mu\nu}$’s. We always try to get rid of terms as high in the hierarchy (with the most $D_\mu$’s) as possible. In particular, with this strategy one ensures that the number of terms is minimal also for the special case in which all external fields are set equal to zero.[^82] The motivation for such an approach is that at each level there exist relations which allow one to eliminate structures, as long as one keeps [*all*]{} terms at the lower levels of the hierarchy. To be specific, when considering multiple covariant derivatives, just one general (i.e., non-contracted) index combination is actually independent in the sense of the hierarchy defined above. For double derivatives this statement reads $$\label{4:10:dda} (D_\mu D_\nu A)_\pm=(D_\nu D_\mu A)_\pm+\frac{i}{4} [ (A)_\pm , (G_{\mu\nu})_+ ] - \frac{i}{4} \{ (A)_\mp , (H_{\mu\nu})_+ \},$$ i.e., it is sufficient to only keep the block $(D_\mu D_\nu A)_\pm$ as long as one considers all terms lower in the hierarchy. The construction then proceeds as follows. First of all, write down all conceivable Lorentz-invariant structures satisfying $P$ and $C$ invariance, Hermiticity, and chiral order $p^6$ in terms of the basic building blocks defined above. Then collect as many relations as possible among these structures and use these to eliminate structures. The relations can follow from any of the following mechanisms: 1. partial integration; 2. equation-of-motion argument; 3. epsilon relations; 4. Bianchi identities; 5. trace relations. Let us illustrate the meaning of each of the above items by selected examples. The partial-integration or total-derivative argument refers to the fact that a total derivative in the Lagrangian density does not change the equation of motion. One thus generates relations of the following type $$\begin{aligned} \lefteqn{ \underbrace{\partial_\mu\mbox{Tr}[(A_1)_\pm \cdots (A_m)_\pm]}_{\mbox{tot.~der.}} +\underbrace{\mbox{Tr}\{[\Gamma_\mu,(A_1)_\pm\cdots (A_m)_\pm ]\}}_{\mbox{0}}} \nonumber\\ &=&\mbox{Tr}\{\nabla_\mu [ (A_1)_\pm\cdots (A_m)_\pm ]\}\nonumber\\ &=&\mbox{Tr}[\nabla_\mu (A_1)_\pm\cdots (A_m)_\pm]+\cdots +\mbox{Tr}[(A_1)_\pm\cdots \nabla_\mu (A_m)_\pm],\end{aligned}$$ where we made use of Eq. (\[4:10:covder\]) and the product rule. This derivative shifting procedure is also valid for multiple traces. At this stage we note the advantage of working with the basic building blocks of Eq. (\[4:10:apm\]) in comparison with those of Ref.  [@Fearing:1994ga] due to the relatively simple connection between the covariant derivative $\nabla_\mu$ outside the block brackets and the covariant derivative $D_\mu$ inside when $(A)_\pm$ are used: $$\begin{aligned} \label{4:10:nabla} \nabla_\mu (A)_\pm&=&(D_\mu A)_\pm - \frac{1}{4}\{(D_\mu U)_-,(A)_\mp\}.\end{aligned}$$ From a technical point of view Eq. (\[4:10:nabla\]) is important because it helps avoid extremely tedious algebraic manipulations one had to perform in the old framework of Ref. [@Fearing:1994ga]. The combination of shifting derivatives back and forth and interchanging indices of multiple derivatives is referred to as index exchange. In Ref. [@Fearing:1994ga] not all total-derivative terms were properly identified which has led to subsequent reductions in the number of terms in both the even-intrinsic-parity sector [@Bijnens:1999sh] and the odd-intrinsic-parity sector [@Ebertshauser:2001nj; @Bijnens:2001bb]. The equation-of-motion argument makes use of the invariance of physical observables under field transformations, as discussed in Sec. \[sec\_clop4\]. The aim is to collect as many terms as possible containing a factor $(D^2 U)_-$. Such terms can be supplemented by corresponding $(\chi)_-$ terms lower in the hierarchy to generate an equation-of-motion term which can be eliminated by a field redefinition. The epsilon relations refer to the odd-intrinsic-parity sector with the basic idea being as follows. Consider a structure with six Lorentz indices transforming under parity as a Lorentz pseudotensor, i.e., $Q_{\lambda\mu\nu\rho\sigma\tau}(\vec{x},t) \mapsto -Q^{\lambda\mu\nu\rho\sigma\tau}(-\vec{x},t)$. In order to form a Lorentz scalar, one needs to contract two indices pairwise and the remaining four with the totally antisymmetric tensor $\epsilon^{\alpha\beta\gamma\delta}$ in four dimensions. Suppose $Q_{\lambda\mu\nu\rho\sigma\tau}(\vec{x},t)$ is neither symmetric nor antisymmetric under the exchange of any pair of indices. Naively one would then expect $5+4+\cdots+1=15$ independent contractions. However, such a counting does not take the totally antisymmetric nature of the epsilon tensor into account [@Akhoury:1990px], from which one obtains, for the above case of no symmetry in the indices, five additional conditions [@Akhoury:1990px; @Fearing:1994ga]. These additional identities have not been considered in the pioneering construction of Ref. [@Issler:1990nj]. In general, $Q_{\lambda\mu\nu\rho\sigma\tau}(\vec{x},t)$ has some symmetry in its indices, and not all five epsilon relations are independent. For example, using the transformation behavior of Table \[4:5:table\_trafprop\] it is easy to verify that $$\mbox{Tr}\{(G_{\lambda\mu})_+[(G_{\nu\rho})_+(H_{\sigma\tau})_+ +(H_{\sigma\tau})_+(G_{\nu\rho})_+]\}$$ is an example for a pseudotensor which is invariant under $G$ (and Hermitian). Its symmetries are given by $$Q_{\lambda\mu\nu\rho\sigma\tau}=-Q_{\mu\lambda\nu\rho\sigma\tau} =-Q_{\lambda\mu\rho\nu\sigma\tau}=-Q_{\lambda\mu\nu\rho\tau\sigma} =Q_{\nu\rho\lambda\mu\sigma\tau},$$ from which one would naively end up with a single combination $$\mbox{Tr}\{(G_{\mu\nu})_+[(G_{\lambda\alpha})_+({H^\lambda}_\beta)_+ +({H^\lambda}_\beta)_+(G_{\lambda\alpha})_+]\}\epsilon^{\mu\nu\alpha\beta}$$ which, however, vanishes due to the epsilon relation. The Bianchi identities refer to certain relations among covariant derivatives of field-strength tensors. Starting from the Jacobi identity $$\label{4:10:jacobi} [A,[B,C]]+[B,[C,A]]+[C,[A,B]]=0,$$ we consider the linear combination $$\begin{aligned} D_\mu f_{\nu\rho}^R+D_\nu f^R_{\rho\mu}+ D_\rho f^R_{\mu\nu} &\equiv&\sum_{\mbox{\scriptsize c.p.}\{\mu,\nu,\rho\}}D_\mu f^R_{\nu\rho} =\sum_{\mbox{\scriptsize c.p.}\{\mu,\nu,\rho\}} (\partial_\mu f^R_{\nu\rho}-i[r_\mu,f^R_{\nu\rho}])\nonumber\\ &=&\sum_{\mbox{\scriptsize c.p.}\{\mu,\nu,\rho\}} \bigg(\partial_\mu\partial_\nu r_\rho-\partial_\mu\partial_\rho r_\nu -i[\partial_\mu r_\nu, r_\rho]\nonumber\\ &-&i[r_\nu,\partial_\mu r_\rho] -i[r_\mu,\partial_\nu r_\rho-\partial_\rho r_\nu] -[r_\mu,[r_\nu, r_\rho]]\bigg)\nonumber\\ &=&0, \label{sumcp}\end{aligned}$$ where use of the Schwarz theorem, $\partial_\mu\partial_\nu \cdots= \partial_\nu\partial_\mu\cdots$, relabeling of indices, and the Jacobi identity, Eq. (\[4:10:jacobi\]), has been made. Observe that the cyclic permutations of the indices $\mu,\nu,$ and $\rho$ has been denoted by $\mbox{c.p.}\{\mu,\nu,\rho\}$. The same arguments hold for the independent field strength tensor $f^L_{\mu\nu}$, and we can summarize the constraints as $$\label{4:10:bi} \sum_{\mbox{\scriptsize c.p.}\{\mu,\nu,\rho\}}D_\mu f^{L/R}_{\nu\rho}=0,$$ which, because of their similarity to an analogous equation for the Riemann-Christoffel curvature tensor in general relativity, are referred to as the Bianchi identities (see, e.g., Refs. [@Ryder:wq; @Weinberg:kr]). Equation (\[4:10:bi\]) does not require that $f^{R/L}_{\mu\nu}$ satisfy any equations of motion. In terms of the building blocks $(D_\mu U)_-$, $(G_{\mu\nu})_+$, and $(H_{\mu\nu})_+$ the Bianchi identities read $$\begin{aligned} \label{4:10:bianchig} \sum_{\mbox{\scriptsize c.p.}\{\mu,\nu,\rho\}}(D_\mu G_{\nu\rho})_+&=& -\frac{1}{4}\sum_{\mbox{\scriptsize c.p.}\{\mu,\nu,\rho\}} [(D_\mu U)_-,(H_{\nu\rho})_+],\\ \label{4:10:bianchih} \sum_{\mbox{\scriptsize c.p.}\{\mu,\nu,\rho\}}(D_\mu H_{\nu\rho})_+&=& -\frac{1}{4}\sum_{\mbox{\scriptsize c.p.}\{\mu,\nu,\rho\}} [(D_\mu U)_-,(G_{\nu\rho})_+].\end{aligned}$$ Given the definition of Eq. (\[4:10:ghdefinition\]), the Bianchi identities can be used to generate relations among the building blocks in terms of structures kept in lower orders of the hierarchy defined above. In Ref. [@Fearing:1994ga] each term of the sum on the left-hand side of Eqs. (\[4:10:bianchig\]) and (\[4:10:bianchih\]) was treated as an independent element so that the final list of supposedly independent structures contained redundant elements. Finally, the trace relations refer to the fact that the construction of invariants uses traces and products of traces. One is thus particularly interested in finding any relations among those traces. We know from the Cayley-Hamilton theorem that any $n\times n$ matrix $A$ is a solution of its associated characteristical polynomial $\chi_A$. For $n=2$ this statement reads $$\begin{aligned} \label{4:10:CalHam} 0=\chi_A (A) & = & A^2 - \mbox{Tr}(A) A + \det (A) 1_{2\times 2}\nonumber\\ &=& A^2 - \mbox{Tr}(A) A + \frac{1}{2} \{ [\mbox{Tr}(A)]^2 - \mbox{Tr}(A^2)\} 1_{2\times 2}.\end{aligned}$$ Setting $A=A_1+A_2$ in (\[4:10:CalHam\]) and making use of $\chi_{A_1} (A_1) = 0 =\chi_{A_2} (A_2) $ one ends up with the matrix equation $$\begin{aligned} \label{4:10:F2} 0=F_2(A_1,A_2)&\equiv&\{A_1,A_2\} - \mbox{Tr}(A_1) A_2 - \mbox{Tr}(A_2) A_1 \nonumber\\ &&+ \mbox{Tr}(A_1)\mbox{Tr}(A_2) 1_{2\times 2} - \mbox{Tr}(A_1 A_2) 1_{2\times 2}\end{aligned}$$ which is the central piece of information needed to derive the trace relations in the SU(2)$\times$SU(2) sector. The analogous $n=3$ equation is slightly more complex $$\begin{aligned} \label{4:10:F3} 0&=&F_3(A_1,A_2,A_3)\nonumber\\ &\equiv& A_1\{A_2,A_3\} + A_2\{A_3,A_1\} + A_3\{A_1,A_2\}\nonumber\\ &&-\mbox{Tr}(A_1)\{A_2,A_3\} - \mbox{Tr}(A_2)\{A_3,A_1\} - \mbox{Tr}(A_3)\{A_1,A_2\}\nonumber\\ &&+\mbox{Tr}(A_1)\mbox{Tr}(A_2)A_3 +\mbox{Tr}(A_2)\mbox{Tr}(A_3)A_1 +\mbox{Tr}(A_3)\mbox{Tr}(A_1)A_2\nonumber\\ &&-\mbox{Tr}(A_1 A_2)A_3 -\mbox{Tr}(A_3 A_1)A_2 -\mbox{Tr}(A_2 A_3)A_1\nonumber\\ &&-\mbox{Tr}(A_1 A_2 A_3) 1_{3\times 3} -\mbox{Tr}(A_1 A_3 A_2) 1_{3\times 3}\nonumber\\ &&+\mbox{Tr}(A_1 A_2)\mbox{Tr}(A_3)1_{3\times 3} +\mbox{Tr}(A_3 A_1)\mbox{Tr}(A_2)1_{3\times 3} +\mbox{Tr}(A_2 A_3)\mbox{Tr}(A_1)1_{3\times 3}\nonumber\\ &&-\mbox{Tr}(A_1)\mbox{Tr}(A_2)\mbox{Tr}(A_3) 1_{3\times 3}.\end{aligned}$$ We can now derive trace relations by simply multiplying Eq.  (\[4:10:F2\]) or Eq. (\[4:10:F3\]) with another arbitrary matrix of the same dimension and finally taking the trace of the whole construction, i.e., $$\begin{aligned} \label{4:10:Tracesu2} 0&=&\mbox{Tr}[F_2(A_1,A_2)A_3],\\ \label{4:10:tracesu3} 0&=&\mbox{Tr}[F_3(A_1,A_2,A_3)A_4].\end{aligned}$$ Note that $A_i$ may be any $n\times n$ matrix, even a string of our basic building blocks. For example, Eq. (\[app:glvgss:tr2by2\]) of Appendix \[app\_sec\_glvgss\], is identical to Eq. (\[4:10:Tracesu2\]). In principle, the ideas developed above apply to the general $\mbox{SU}(N_f)_L\times\mbox{SU}(N_f)_R$ case and only at the end it is necessary to specify the number of flavors $N_f$. The reduction to the cases $N_f=2$ and $N_f=3$ is achieved in terms of the trace relations summarized in Eqs.  (\[4:10:Tracesu2\]) and (\[4:10:tracesu3\]). Although we have never come across a trace relation that could not be obtained in the manner explained above, we are not aware of a general proof showing that any kind of trace relation must be related to the Caley-Hamilton theorem. In the even-intrinsic-parity sector the Lagrangian has 112 in principle measurable + 3 contact terms for the general $\mbox{SU}(N_f)_L\times\mbox{SU}(N_f)_R$ case, 90 + 4 for the $\mbox{SU}(3)_L\times\mbox{SU}(3)_R$ case, and 53 + 4 for the $\mbox{SU}(2)_L\times\mbox{SU}(2)_R$ case [@Bijnens:1999sh]. The contact terms refer to structures which can be expressed in terms of only external fields such as the $H_1$ and $H_2$ terms of the ${\cal L}_4$ Lagrangian of Eq. (\[4:7:l4gl\]). The reduction in the number of terms in comparison with the 111 $\mbox{SU}(3)_L\times\mbox{SU}(3)_R$ structures of Ref. [@Fearing:1994ga] is due to a more complete application of the partial-integration relations, the use of additional trace relations, and the use of four relations due to the Bianchi identities which were not taken into account in Ref. [@Fearing:1994ga]. The odd-intrinsic-parity sector was reconsidered in Refs.  [@Ebertshauser:2001nj; @Bijnens:2001bb]. Both analyses found 24 $\mbox{SU}(N_f)_L\times\mbox{SU}(N_f)_R$, 23 $\mbox{SU}(3)_L\times\mbox{SU}(3)_R$, and 5 $\mbox{SU}(2)_L\times\mbox{SU}(2)_R$ terms. Moreover, 8 additional terms due to the extension of the chiral group to $\mbox{SU}(N_f)_L\times\mbox{SU}(N_f)_R\times U(1)_V$ were found, which are of some relevance when considering the electromagnetic interaction for the two-flavor case. In comparison to Ref. [@Fearing:1994ga], the new analysis of Ref. [@Ebertshauser:2001nj] could eliminate two structures via partial integration, 6 via Bianchi identities and one by a trace relation. It is unlikely that the coefficients of all the terms of ${\cal L}_6$ will be determined from experiment. However, usually a much smaller subset actually contributes to most simple processes, and it is possible to get information on some of the corresponding coefficients. ### Elastic Pion-Pion Scattering at ${\cal O}(p^6)$ {#subsec_eppsop6} Elastic pion-pion scattering represents a nice example of the success of mesonic chiral perturbation theory. A complete analytical calculation at two-loop order was performed in Ref. [@Bijnens:1995yn]. Let us consider the $T$-matrix element of the scattering process $\pi^a(p_a)+\pi^b(p_b)\to\pi^c(p_c)+\pi^d(p_d)$, $$\label{4:10:tpipi} T^{ab;cd}(p_a,p_b;p_c,p_d)=\delta^{ab}\delta^{cd}A(s,t,u) +\delta^{ac}\delta^{bd}A(t,s,u) +\delta^{ad}\delta^{bc}A(u,t,s),$$ where $s=(p_a+p_b)^2$, $t=(p_a-p_c)^2$, and $u=(p_a-p_d)^2$ denote the usual Mandelstam variables, the indices $a,\cdots,d$ refer to the Cartesian isospin components, and the function $A$ satisfies $A(s,t,u)=A(s,u,t)$ [@Weinberg:1966kf]. Since the pions form an isospin triplet, the two isovectors of both the initial and final states may be coupled to $I=0,1,2$. For $m_u=m_d=m$ the strong interactions are invariant under isospin transformations, implying that scattering matrix elements can be decomposed as $$\label{4:10:tdec} \langle I',I_3'|T|I,I_3\rangle=T^I \delta_{II'}\delta_{I_3 I_3'}.$$ For the case of $\pi\pi$ scattering the three isospin amplitudes are given in terms of the invariant amplitude $A$ of Eq. (\[4:10:tpipi\]) by [@Gasser:1983yg] $$\begin{aligned} \label{4:10:isk} T^{I=0}&=&3A(s,t,u)+A(t,u,s)+A(u,s,t),\nonumber\\ T^{I=1}&=&A(t,u,s)-A(u,s,t),\nonumber\\ T^{I=2}&=&A(t,u,s)+A(u,s,t).\end{aligned}$$ For example, the physical $\pi^+\pi^+$ scattering process is described by $T^{I=2}$. Other physical processes are obtained using the appropriate Clebsch-Gordan coefficients. Evaluating the $T$ matrices at threshold, one obtains the $s$-wave $\pi\pi$-scattering lengths[^83] $$\label{$:10:swsl} T^{I=0}|_{\rm thr}=32\pi a^0_0,\quad T^{I=2}|_{\rm thr}=32\pi a^2_0,$$ where the subscript $0$ refers to $s$ wave and the superscript to the isospin. ($T^{I=1}|_{\rm thr}$ vanishes because of Bose symmetry). The current-algebra prediction of Ref. [@Weinberg:1966kf] is identical with the lowest-order result obtained from Eqs. (\[4:6:mpipi1\]) or (\[4:6:mpipi2\]), $$\label{4:10:a00a02lo} a_0^0=\frac{7 M_\pi^2}{32 \pi F_\pi^2}=0.156,\quad a_0^2=-\frac{M_\pi^2}{16 \pi F_\pi^2}=-0.045,$$ where we replaced $F_0$ by $F_\pi$ and made use of the numerical values $F_\pi=93.2$ MeV and $M_\pi=139.57$ MeV of Ref. [@Bijnens:1995yn]. In order to obtain the results of Eq. (\[4:10:a00a02lo\]), use has been made of $s_{\rm thr}=4 M_\pi^2$ and $t_{\rm thr}=u_{\rm thr}=0$. The predictions for the $s$-wave scattering lengths at ${\cal O}(p^6)$ read [@Bijnens:1995yn] $$\begin{aligned} a_0^0&=& \overbrace{0.156}^{\mbox{${\cal O}(p^2)$}} +\overbrace{\underbrace{0.039}_{\mbox{L}} +\underbrace{0.005}_{\mbox{anal.}}}^{\mbox{${\cal O}(p^4)$: +28\%}} +\overbrace{\underbrace{0.013}_{\mbox{$k_i$}} +\underbrace{0.003}_{\mbox{L}} +\underbrace{0.001}_{\mbox{anal.}}}^{\mbox{${\cal O}(p^6)$: +8.5\%}}= \overbrace{0.217}^{\mbox{total}},\\ a_0^0-a_0^2&=& \overbrace{0.201}^{\mbox{${\cal O}(p^2)$}} +\overbrace{\underbrace{0.036}_{\mbox{L}} +\underbrace{0.006}_{\mbox{anal.}}}^{\mbox{${\cal O}(p^4)$: +21\%}} +\overbrace{\underbrace{0.012}_{\mbox{$k_i$}} +\underbrace{0.003}_{\mbox{L}} +\underbrace{0.001}_{\mbox{anal.}}}^{\mbox{${\cal O}(p^6)$: +6.6\%}}= \overbrace{0.258}^{\mbox{total}}.\end{aligned}$$ The corrections at ${\cal O}(p^4)$ consist of a dominant part from the chiral logarithms (L) of the one-loop diagrams and a less important analytical contribution (anal.) resulting from the one-loop diagrams as well as the tree graphs of ${\cal L}_4$. The total corrections at ${\cal O}(p^4)$ amount to 28% and 21% of the ${\cal O}(p^2)$ predictions, respectively. At ${\cal O}(p^6)$ one obtains two-loop corrections, one-loop corrections, and ${\cal L}_6$ tree-level contributions. Once again, the loop corrections ($k_i$, involving double chiral logarithms, and L) are more important than the analytical contributions. The influence of ${\cal L}_6$ was estimated via scalar- and vector-meson exchange and found to be very small. The empirical results for the $\pi\pi$ $s$-wave scattering lengths are, so far, obtained from the $K_{e4}$ decay $K^+\to\pi^+\pi^-e^+\nu_e$ and the $\pi^\pm p\to \pi^\pm \pi^+ n$ reactions. In the former case, the connection with low-energy $\pi\pi$ scattering stems from a partial-wave analysis of the form factors relevant for the $K_{e4}$ decay in terms of $\pi\pi$ angular momentum eigenstates. In the low-energy regime the phases of these form factors are related by (a generalization of) Watson’s theorem [@Watson:1954uc] to the corresponding phases of $I=0$ $s$-wave and $I=1$ $p$-wave elastic scattering [@Colangelo:2001sp]. Using a dispersion-theory approach in terms of the Roy equations [@Roy:1971tc; @Ananthanarayan:2000ht], the most recent analysis of $K^+\to\pi^+\pi^-e^+\nu_e$ [@Pislak:2001bf] has obtained $$\label{4:10:swslexpold} a_0^0=0.228\pm 0.012 \pm 0.003.$$ This result has to be compared with older determinations [@Rosselet:1976pu; @Froggatt:hu; @Nagels:xh] $$\label{swslexp} a_0^0=0.26\pm 0.05,\quad a_0^2=-0.028 \pm 0.012,$$ and the more recent one from $\pi^\pm p\to \pi^\pm \pi^+ n$ [@Kermani:gq] $$\label{4:10:swslexptriumf} a_0^0=0.204\pm 0.014\, (stat.) \pm 0.008\, (syst.),$$ which makes use of an extrapolation to the pion pole to extract the $\pi\pi$ amplitude. In particular, when analyzing the data of Ref. [@Pislak:2001bf] in combination with the Roy equations, an upper limit $|\bar{l}_3|\leq 16$ was obtained in Ref. [@Colangelo:2001sp] for the scale-independent low-energy coupling constant which is related to $l_3$ of the SU(2)$\times$SU(2) Lagrangian of Gasser and Leutwyler (see Appendix \[app\_sec\_glvgss\]). The great interest generated by this result is to be understood in the context of the pion mass at ${\cal O}(p^4)$ \[see Eq. (\[app:dp:Mpi2\]) of App. \[app\_sec\_dp\]\], $$\label{4:10:Mpi2} M_\pi^2=M^2-\frac{\bar{l}_3}{32\pi^2 F^2_0} M^4+ {\cal O}(M^6),$$ where $M^2=(m_u+m_d) B_0$. Recall that the constant $B_0$ is related to the scalar quark condensate in the chiral limit \[see Eq. (\[4:3:b0\])\] and that a nonvanishing quark condensate is a sufficient criterion for spontaneous chiral symmetry breakdown in QCD (see Sec. \[subsec\_sqc\]). If the expansion of $M_\pi^2$ in powers of the quark masses is dominated by the linear term in Eq. (\[4:10:Mpi2\]), the result is often referred to as the Gell-Mann-Oakes-Renner relation [@Gell-Mann:rz]. If the terms of order $m^2$ were comparable or even larger than the linear terms, a different power counting or bookkeeping in ChPT would be required [@Knecht:1995tr; @Knecht:1995ai; @Stern:1997]. The estimate $|\bar{l}_3|\leq 16$ implies that the Gell-Mann-Oakes-Renner relation [@Gell-Mann:rz] is indeed a decent starting point, because the contribution of the second term of Eq. (\[4:10:Mpi2\]) to the pion mass is approximately given by $$-\frac{\bar{l}_3 M_\pi^2}{64\pi^2 F_\pi^2} M_\pi =-0.054 M_\pi\,\,\mbox{for}\,\, \bar{l}_3=16,$$ i.e., more than 94 % of the pion mass must stem from the quark condensate [@Colangelo:2001sp]. Chiral Perturbation Theory for Baryons {#chap_cptb} ====================================== So far we have considered the purely mesonic sector involving the interaction of Goldstone bosons with each other and with the external fields. However, ChPT can be extended to also describe the dynamics of baryons at low energies. Here we will concentrate on matrix elements with a single baryon in the initial and final states. With such matrix elements we can, e.g, describe static properties such as masses or magnetic moments, form factors, or, finally, more complicated processes, such as pion-nucleon scattering, Compton scattering, pion photoproduction etc. Technically speaking, we are interested in the baryon-to-baryon transition amplitude in the presence of external fields (as opposed to the vacuum-to-vacuum transition amplitude of Sec. \[subsec\_qcdpefgf\]) [@Gasser:1987rb; @Krause:xc], $$\label{4:btbta} {\cal F}(\vec{p}\,',\vec{p};v,a,s,p)=\langle {\vec{p}\,'};{\rm out}| {\vec{p}\,};{\rm in}\rangle^{\rm c}_{v,a,s,p}, \quad \vec{p}\neq\vec{p}\,',$$ determined by the Lagrangian of Eq. (\[2:4:lqcds\]), $$\label{5:lqcds} {\cal L}={\cal L}^0_{\rm QCD}+{\cal L}_{\rm ext} ={\cal L}^0_{\rm QCD}+\bar{q}\gamma_\mu (v^\mu +\frac{1}{3}v^\mu_{(s)} +\gamma_5 a^\mu )q -\bar{q}(s-i\gamma_5 p)q.$$ In Eq. (\[4:btbta\]) $|\vec{p};{\rm in}\rangle$ and $|\vec{p}\,';{\rm out}\rangle$ denote asymptotic one-baryon in- and out-states, i.e., states which in the remote past and distant future behave as free one-particle states of momentum $\vec{p}$ and $\vec{p}\,'$, respectively. The functional ${\cal F}$ consists of connected diagrams only (superscript c). For example, the matrix elements of the vector and axial-vector currents between one-baryon states are given by [@Krause:xc] $$\begin{aligned} \label{5:vc} \langle \vec{p}\,'|V^{\mu,a}(x)|\vec{p}\,\rangle &=&\left. \frac{\delta}{i\delta v^a_\mu(x)} {\cal F}(\vec{p}\,',\vec{p};v,a,s,p)\right|_{v=0,a=0,s=M,p=0},\\ \label{5:avc} \langle \vec{p}\,'|A^{\mu,a}(x)|\vec{p}\,\rangle &=&\left. \frac{\delta}{i\delta a^a_\mu(x)} {\cal F}(\vec{p}\,',\vec{p};v,a,s,p)\right|_{v=0,a=0,s=M,p=0},\end{aligned}$$ where $M=\mbox{diag}(m_u,m_d,m_s)$ denotes the quark-mass matrix and $$V^{\mu,a}(x)=\bar{q}(x)\gamma^\mu\frac{\lambda^a}{2} q(x),\quad A^{\mu,a}(x)=\bar{q}(x)\gamma^\mu \gamma_5 \frac{\lambda^a}{2} q(x).$$ As in the mesonic sector the method of calculating the Green functions associated with the functional of Eq. (\[4:btbta\]) consists of an effective Lagrangian-approach in combination with an appropriate power counting. Specific matrix elements will be calculated applying the Gell-Mann and Low formula of perturbation theory [@Gell-Mann:1951rw]. The group-theoretical foundations of constructing phenomenological Lagrangians in the presence of spontaneous symmetry breaking have been developed in Refs. [@Weinberg:de; @Coleman:sm; @Callan:sn]. The fields entering the Lagrangian are assumed to transform under irreducible representations of the subgroup $H$ which leaves the vacuum invariant whereas the symmetry group $G$ of the Hamiltonian or Lagrangian is nonlinearly realized (for the transformation behavior of the Goldstone bosons, see Sec. \[sec\_tpgb\]). Transformation Properties of the Fields {#sec_tpf} --------------------------------------- Our aim is a description of the interaction of baryons with the Goldstone bosons as well as the external fields at low energies. To that end we need to specify the transformation properties of the dynamical fields entering the Lagrangian. Our discussion follows Refs. [@Georgi; @Gasser:1987rb]. To be specific, we consider the octet of the $\frac{1}{2}^+$ baryons (see Fig. \[sec:tf:tab:octetbaryons\]). With each member of the octet we associate a complex, four-component Dirac field which we arrange in a traceless $3\times 3$ matrix $B$, (10,6) (0,0)[(1,0)[10]{}]{} (10,-1)[$I_3$]{} (0,0)[(0,1)[6]{}]{} (-1,6)[$S$]{} (3,1) (2.0,0.5)[$n(940)(udd)$]{} (7,1) (6.0,0.5)[$p(938)(uud)$]{} (1,3) (0,3.5)[$\Sigma^-(1197)(dds)$]{} (5,3) (4.0,3.5)[$\Sigma^0(1193)(uds)$]{} (4.0,2.5)[$\Lambda(1116)(uds)$]{} (9,3) (8.0,3.5)[$\Sigma^+(1189)(uus)$]{} (3,5) (2.0,4.5)[$\Xi^-(1321)(dss)$]{} (7,5) (6.0,4.5)[$\Xi^0(1315)(uss)$]{} (0,1) (-0.5,1)[0]{} (0,3) (-0.5,3)[-1]{} (0,5) (-0.5,5)[-2]{} (1,0) (1,-0.5)[-1]{} (3,0) (3,-0.5)[-1/2]{} (5,0) (5,-0.5)[0]{} (7,0) (7,-0.5)[1/2]{} (9,0) (9,-0.5)[1]{} $$\label{5:1:su3oktett} B=\sum_{a=1}^8 \lambda_a B_a= \left(\begin{array}{ccc} \frac{1}{\sqrt{2}}\Sigma^0+\frac{1}{\sqrt{6}}\Lambda&\Sigma^+&p\\ \Sigma^-&-\frac{1}{\sqrt{2}}\Sigma^0+\frac{1}{\sqrt{6}}\Lambda&n\\ \Xi^-&\Xi^0&-\frac{2}{\sqrt{6}}\Lambda \end{array}\right),$$ where we have suppressed the dependence on $x$. For later use, we have to keep in mind that each entry of Eq. (\[5:1:su3oktett\]) is a Dirac field, but for the purpose of discussing the transformation properties under global flavor SU(3) this can be ignored, because these transformations act on each of the four components in the same way. In contrast to the mesonic case of Eq. (\[4:3:upar\]), where we collected the fields of the Goldstone octet in a Hermitian traceless matrix $\phi$, the $B_a$ of the spin $1/2$-case are not real (Hermitian), i.e., $B\neq B^\dagger$. Now let us define the set $$\label{5:1:setm} M\equiv\{B(x)|B(x)\, \mbox{complex, traceless $3\times 3$ matrix}\}$$ which under the addition of matrices is a complex vector space. The following homomorphism is a representation of the abstract group $H=\mbox{SU}(3)_V$ on the vector space $M$ \[see also Eq. (\[4:2:uhtrafo\])\]: $$\begin{aligned} \label{5:1:su3hom} &&\varphi: H\to \varphi(H),\quad V\mapsto \varphi(V)\quad \mbox{where}\quad \varphi(V): M\to M,\nonumber\\ &&B(x)\mapsto B'(x)=\varphi(V)B(x)\equiv V B(x) V^\dagger.\end{aligned}$$ First of all, $B'(x)$ is again an element of $M$, because $\mbox{Tr}[B'(x)] =\mbox{Tr}[VB(x)V^\dagger]=\mbox{Tr}[B(x)]=0$. Equation (\[5:1:su3hom\]) satisfies the homomorphism property $$\begin{aligned} \varphi(V_1)\varphi(V_2) B(x)&=&\varphi(V_1)V_2B(x) V_2^\dagger =V_1 V_2 B(x) V_2^\dagger V_1^\dagger=(V_1 V_2)B(x) (V_1 V_2)^\dagger\\ &=&\varphi(V_1V_2) B(x)\end{aligned}$$ and is indeed a [*representation*]{} of SU(3), because $$\begin{aligned} \varphi(V)[\lambda_1B_1(x)+\lambda_2 B_2(x)]&=& V[\lambda_1B_1(x)+\lambda_2 B_2(x)]V^\dagger\\ &=&\lambda_1 VB_1(x)V^\dagger+\lambda_2VB_2(x)V^\dagger\\ &=&\lambda_1\varphi(V)B_1(x) +\lambda_2\varphi(V)B_2(x).\end{aligned}$$ Equation (\[5:1:su3hom\]) is just the familiar statement that $B$ transforms as an octet under (the adjoint representation of) SU(3)$_V$.[^84] Let us now turn to various representations and realizations of the group $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$. We consider two explicit examples and refer the interested reader to Ref. [@Georgi] for more details. In analogy to the discussion of the quark fields in QCD, we may introduce left- and right-handed components of the baryon fields \[see Eq. (\[2:3:qlr\])\]: $$\label{5:1:blr} B_1=P_L B_1+P_R B_1=B_L + B_R.$$ We define the set $M_1\equiv\{(B_L(x),B_R(x))\}$ which under the addition of matrices is a complex vector space. The following homomorphism is a representation of the abstract group $G=\mbox{SU}(3)_L\times \mbox{SU(3)}_R$ on $M_1$: $$\begin{aligned} \label{5:1:su3lrhom} (B_L,B_R)\mapsto (B'_L,B'_R)%=\varphi(L,R)(B_L,B_R) \equiv (L B_L L^\dagger,R B_R R^\dagger),\end{aligned}$$ where we have suppressed the $x$ dependence. The proof proceeds in complete analogy to that of Eq. (\[5:1:su3hom\]). As a second example, consider the set $M_2\equiv\{B_2(x)\}$ with the homomorphism $$\label{5:1:su3lrhomb2} B_2\mapsto B_2'\equiv L B_2 L^\dagger,$$ i.e. the transformation behavior is independent of $R$. The mapping defines a representation of the group $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$, although the transformation behavior is drastically different from the first example. However, the important feature which both mappings have in common is that under the subgroup $H=\{(V,V)|V\in \mbox{SU($3$)}\}$ of $G$ both fields $B_i$ transform as an octet: $$\begin{aligned} B_1=B_L+B_R&\stackrel{H}{\mapsto}&VB_LV^\dagger+VB_R V^\dagger=VB_1V^\dagger,\\ B_2&\stackrel{H}{\mapsto}&VB_2V^\dagger.\end{aligned}$$ We will now show how in a theory also containing Goldstone bosons the various realizations may be connected to each other using field redefinitions. The procedure is actually very similar to Sec. \[subsec\_mclop6\], where we discussed how, by an appropriate multiplication with $U$ or $U^\dagger$, all building blocks of the mesonic effective Lagrangian could be made to transform in the same way. Here we consider the second example, with the fields $B_2$ of Eq. (\[5:1:su3lrhomb2\]) and $U$ of Eq. (\[4:3:upar\]) transforming as $$B_2\mapsto LB_2L^\dagger,\quad U\mapsto RUL^\dagger,$$ and define new baryon fields by $$\tilde{B}\equiv UB_2,$$ so that the new pair $(\tilde B,U)$ transforms as $$\tilde{B}\mapsto RUL^\dagger L B L^\dagger=R\tilde{B}L^\dagger,\quad U\mapsto RU L^\dagger.$$ Note in particular that $\tilde{B}$ still transforms as an octet under the subgroup $H=\mbox{SU(3)}_V$. Given that physical observable are invariant under field transformations we may choose a description of baryons that is maximally convenient for the construction of the effective Lagrangian [@Georgi] and which is commonly used in chiral perturbation theory. We start with $G=\mbox{SU(2)}_L\times\mbox{SU(2)}_R$ and consider the case of $G=\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ later. Let $$\label{5:1:Psi} \Psi=\left(\begin{array}{c}p\\n\end{array}\right)$$ denote the nucleon field with two four-component Dirac fields for the proton and the neutron and $U$ the SU(2) matrix containing the pion fields. We have already seen in Sec. \[subsec\_aqcd\] that the mapping $U\mapsto RUL^\dagger$ defines a nonlinear realization of $G$. We denote the square root of $U$ by $u$, $u^2(x)=U(x)$, and define the SU(2)-valued function $K(L,R,U)$ by \[see Eqs. (\[4:10:kdef\]) and (\[4:10:k\])\] $$\label{5:1:kdef} u(x)\mapsto u'(x)=\sqrt{RUL^\dagger}\equiv RuK^{-1}(L,R,U),$$ i.e. $$K(L,R,U)=u'^{-1}Ru=\sqrt{RUL^\dagger}^{-1}R \sqrt{U}.$$ The following homomorphism defines an operation of $G$ on the set $\{(U,\Psi)\}$ in terms of a nonlinear realization: $$\label{5:1:su2real} \varphi(g):\left(\begin{array}{c}U\\ \Psi\end{array}\right)\mapsto \left(\begin{array}{c}U'\\ \Psi'\end{array}\right) =\left(\begin{array}{c}RUL^\dagger\\K(L,R,U)\Psi\end{array}\right),$$ because the identity leaves $(U,\Psi)$ invariant and \[see Sec. \[subsec\_aqcd\] and Eq. (\[4:10:kreal\])\] $$\begin{aligned} \varphi(g_1)\varphi(g_2) \left(\begin{array}{c}U\\ \Psi\end{array}\right)&=&\varphi(g_1) \left(\begin{array}{c}R_2UL_2^\dagger\\K(L_2,R_2,U)\Psi\end{array}\right)\\ &=&\left(\begin{array}{c}R_1R_2UL_2^\dagger L_1^\dagger\\ K(L_1,R_1,R_2UL_2^\dagger)K(L_2,R_2,U)\Psi \end{array}\right)\\ &=&\left(\begin{array}{c}R_1 R_2U(L_1L_2)^\dagger\\ K(L_1L_2,R_1R_2,U)\Psi\end{array}\right)\\ &=&\varphi(g_1g_2)\left(\begin{array}{c}U\\ \Psi\end{array}\right).\end{aligned}$$ Note that for a general group element $g=(L,R)$ the transformation behavior of $\Psi$ depends on $U$. For the special case of an isospin transformation, $R=L=V$, one obtains $u'=VuV^\dagger$, because $$U'=u'^2=VuV^\dagger VuV^\dagger=Vu^2V^\dagger=VUV^\dagger.$$ Comparing with Eq. (\[5:1:kdef\]) yields $K^{-1}(V,V,U)=V^\dagger$ or $K(V,V,U)=V$, i.e., $\Psi$ transforms linearly as an isospin doublet under the isospin subgroup $\mbox{SU(2)}_V$ of $\mbox{SU(2)}_L\times\mbox{SU(2)}_R$. A general feature here is that the transformation behavior under the subgroup which leaves the ground state invariant is independent of $U$. Moreover, as already discussed in Sec. \[subsec\_aqcd\], the Goldstone bosons $\phi$ transform according to the adjoint representation of SU(2)$_V$, i.e., as an isospin triplet. For the case $G=\mbox{SU(3)}_L\times \mbox{SU(3)}_R$ one uses the nonlinear realization $$\label{5:1:su3real} \varphi(g):\left(\begin{array}{c}U\\ B\end{array}\right)\mapsto \left(\begin{array}{c}U'\\ B'\end{array}\right) =\left(\begin{array}{c}RUL^\dagger\\K(L,R,U)B K^\dagger(L,R,U) \end{array}\right),$$ where $K$ is defined completely analogously to Eq. (\[5:1:kdef\]) after inserting the corresponding SU(3) matrices. Lowest-Order Effective Baryonic Lagrangian {#sec_loebl} ------------------------------------------ Given the dynamical fields of Eqs. (\[5:1:su2real\]) and (\[5:1:su3real\]) and their transformation properties, we will now discuss the most general effective baryonic Lagrangian at lowest order. As in the vacuum sector, chiral symmetry provides constraints among the single-baryon Green functions contained in the functional of Eq. (\[4:btbta\]). Analogous to the mesonic sector, these Ward identities will be satisfied if the Green functions are calculated from the most general effective Lagrangian coupled to external fields with a [*local*]{} invariance under the chiral group (see Appendix \[app\_gfwi\]). Let us start with the construction of the $\pi N$ effective Lagrangian ${\cal L}^{(1)}_{\pi N}$ which we demand to have a [*local*]{} $\mbox{SU}(2)_L\times\mbox{SU(2)}_R\times\mbox{U(1)}_V$ symmetry. The transformation behavior of the external fields is given in Eq. (\[2:4:sg\]), whereas the nucleon doublet and $U$ transform as $$\label{5:2:psitrans} \left(\begin{array}{c}U(x)\\ \Psi(x)\end{array}\right)\mapsto \left(\begin{array}{c}V_R(x)U(x)V_L^\dagger(x)\\ \exp[-i\Theta(x)]K[V_L(x),V_R(x),U(x)]\Psi(x)\end{array}\right).$$ The local character of the transformation implies that we need to introduce a covariant derivative $D_\mu \Psi$ with the usual property that it transforms in the same way as $\Psi$ \[compare with Eq. (\[2:2:cdt\]) for the case of QED\]: $$\label{5:2:kovder} D_\mu \Psi(x)\mapsto [D_\mu \Psi(x)]'\stackrel{!}{=} \exp[-i\Theta(x)]K[V_L(x),V_R(x),U(x)]D_\mu\Psi(x).$$ Since $K$ not only depends on $V_L$ and $V_R$ but also on $U$, we may expect the covariant derivative to contain $u$ and $u^\dagger$ and their derivatives. In fact, the connection of Eq. (\[4:10:chiralconnection\]) (recall $\partial_\mu u u^\dagger=-u\partial_\mu u^\dagger$), $$\label{5:2:gamma} \Gamma_\mu=\frac{1}{2}\left[u^\dagger(\partial_\mu-ir_\mu)u +u(\partial_\mu-il_\mu)u^\dagger\right],$$ is also an integral part of the covariant derivative of the nucleon doublet: $$\label{5:2:kovderpsi} D_\mu\Psi=(\partial_\mu+\Gamma_\mu-iv_\mu^{(s)})\Psi.$$ What needs to be shown is $$\label{5:2tbs} D'_\mu\Psi'=[\partial_\mu+\Gamma_\mu'-i(v_\mu^{(s)}-\partial_\mu\Theta)] \exp(-i\Theta)K\Psi =\exp(-i\Theta)K(\partial_\mu+\Gamma_\mu-iv_\mu^{(s)})\Psi.$$ To that end, we make use of the product rule, $$\partial_\mu[\exp(-i\Theta)K\Psi]=-i\partial_\mu\Theta \exp(-i\Theta)K\Psi +\exp(-i\Theta)\partial_\mu K\Psi +\exp(-i\Theta) K \partial_\mu \Psi,$$ in Eq. (\[5:2tbs\]) and multiply by $\exp(i\Theta)$, reducing it to $$\partial_\mu K=K\Gamma_\mu-\Gamma_\mu'K.$$ Using Eq. (\[5:1:kdef\]), $$\begin{aligned} K&=&u'^\dagger V_R u=\underbrace{u'u'^\dagger}_{\mbox{1}}u'^\dagger V_R u =u' U'^\dagger V_R u= u' V_L \underbrace{U^\dagger}_{\mbox{$u^\dagger u^\dagger$}} \underbrace{V_R^\dagger V_R}_{\mbox{1}} u =u'V_Lu^\dagger,\end{aligned}$$ we find $$\begin{aligned} 2(K\Gamma_\mu-\Gamma_\mu'K)&=& K\left[u^\dagger(\partial_\mu-ir_\mu)u\right] -\left[u'^\dagger(\partial_\mu-iV_R r_\mu V_R^\dagger +V_R\partial_\mu V_R^\dagger)u'\right]K\\ &&+(R\to L,u\leftrightarrow u^\dagger,u'\leftrightarrow u'^\dagger)\\ &=&u'^\dagger V_R(\partial_\mu u-ir_\mu u) -u'^\dagger\partial_\mu u' \underbrace{K}_{\mbox{$u'^\dagger V_R u$}}\\ && +iu'^\dagger V_R r_\mu\underbrace{V_R^\dagger u'K}_{\mbox{$u$}} -u'^\dagger V_R\partial_\mu V^\dagger_R\underbrace{u'K}_{\mbox{$V_R u$}}\\ &&+(R\to L,u\leftrightarrow u^\dagger,u'\leftrightarrow u'^\dagger)\\ &=&u'^\dagger V_R\partial_\mu u -iu'^\dagger V_R r_\mu u -\underbrace{u'^\dagger\partial_\mu u'u'^\dagger}_{ \mbox{$-\partial_\mu u'^\dagger$}} V_R u\\ &&+iu'^\dagger V_R r_\mu u -u'^\dagger\underbrace{V_R\partial_\mu V_R^\dagger V_R}_{\mbox{$-\partial_\mu V_R$}}u\\ &&+(R\to L,u\leftrightarrow u^\dagger,u'\leftrightarrow u'^\dagger)\\ &=&u'^\dagger V_R\partial_\mu u +\partial_\mu u'^\dagger V_R u + u'^\dagger \partial_\mu V_R u\\ &&+(R\to L,u\leftrightarrow u^\dagger,u'\leftrightarrow u'^\dagger)\\ &=&\partial_\mu(u'^\dagger V_Ru+u'V_Lu^\dagger)=2\partial_\mu K,\end{aligned}$$ i.e., the covariant derivative defined in Eq. (\[5:2:kovderpsi\]) indeed satisfies the condition of Eq. (\[5:2:kovder\]). At ${\cal O}(p)$ there exists another Hermitian building block, the so-called vielbein [@Ecker:1995gg],[^85] $$\label{5:2:chvi} u_\mu\equiv i\left[u^\dagger(\partial_\mu-i r_\mu)u-u(\partial_\mu-i l_\mu)u^\dagger\right],$$ which under parity transforms as an axial vector: $$u_\mu\stackrel{P}{\mapsto} i\left[u(\partial^\mu-il^\mu)u^\dagger -u^\dagger(\partial^\mu-ir^\mu)u\right]=-u^\mu.$$ The transformation behavior under $\mbox{SU(2)}_L\times\mbox{SU(2)}_R\times \mbox{U}(1)_V$ is given by $$u_\mu \mapsto Ku_\mu K^\dagger,$$ which is shown using \[see Eq. (\[5:1:kdef\])\] $$u'=V_R u K^\dagger=K u V_L^\dagger$$ and the corresponding adjoints. We obtain $$\begin{aligned} u_\mu&\mapsto&i[u'^\dagger(\partial_\mu-iV_R r_\mu V_R^\dagger +V_R\partial_\mu V_R^\dagger)u'\\ &&-u'(\partial_\mu-iV_L l_\mu V_L^\dagger +V_L\partial_\mu V_L^\dagger) u'^\dagger]\\ &=&i[Ku^\dagger V_R^\dagger(\partial_\mu-iV_R r_\mu V_R^\dagger +V_R\partial_\mu V_R^\dagger)V_R u K^\dagger\\ &&-Ku V_L^\dagger(\partial_\mu-iV_L l_\mu V_L^\dagger +V_L\partial_\mu V_L^\dagger) V_L u^\dagger K^\dagger]\\ &=&i[Ku^\dagger V_R^\dagger\partial_\mu V_Ru K^\dagger +Ku^\dagger\partial_\mu u K^\dagger +K\partial_\mu K^\dagger\\ &&-iK u^\dagger r_\mu u K^\dagger +K u^\dagger\underbrace{\partial_\mu V_R^\dagger V_R}_{\mbox{$-V_R^\dagger \partial_\mu V_R$}} u K^\dagger\\ &&-Ku V_L^\dagger\partial_\mu V_L u^\dagger K^\dagger -Ku\partial_\mu u^\dagger K^\dagger -K\partial_\mu K^\dagger\\ &&+iKul_\mu u^\dagger K^\dagger -Ku\underbrace{\partial_\mu V_L^\dagger V_L}_{\mbox{$ -V_L^\dagger \partial_\mu V_L$}}u^\dagger K^\dagger]\\ &=&iK[u^\dagger(\partial_\mu -ir_\mu)u -u(\partial_\mu -i l_\mu)u^\dagger]K^\dagger\\ &=&Ku_\mu K^\dagger.\end{aligned}$$ The most general effective $\pi N$ Lagrangian describing processes with a single nucleon in the initial and final states is then of the type $\bar{\Psi} \widehat{O} \Psi$, where $\widehat{O}$ is an operator acting in Dirac and flavor space, transforming under $\mbox{SU(2)}_L\times\mbox{SU(2)}_R\times \mbox{U}(1)_V$ as $K\widehat{O}K^\dagger$. As in the mesonic sector, the Lagrangian must be a Hermitian Lorentz scalar which is even under the discrete symmetries $C$, $P$, and $T$. The most general such Lagrangian with the smallest number of derivatives is given by [@Gasser:1987rb][^86] $$\label{5:2:l1pin} {\cal L}^{(1)}_{\pi N}= \bar{\Psi}\left(iD\hspace{-.6em}/ -\stackrel{\circ}{m}_N +\frac{\stackrel{\circ}{g}_A}{2}\gamma^\mu \gamma_5 u_\mu\right)\Psi.$$ It contains two parameters not determined by chiral symmetry: the nucleon mass $\stackrel{\circ}{m}_N$ and the axial-vector coupling constant $\stackrel{\circ}{g}_A$, both taken in the chiral limit (denoted by $\circ$). The overall normalization of the Lagrangian is chosen such that in the case of no external fields and no pion fields it reduces to that of a free nucleon of mass $\stackrel{\circ}{m}_N$. Since the nucleon mass $m_N$ does not vanish in the chiral limit, the zeroth component $\partial^0$ of the partial derivative acting on the nucleon field does not produce a “small” quantity. We thus have to address the new features of chiral power counting in the baryonic sector. The counting of the external fields as well as of covariant derivatives acting on the mesonic fields remains the same as in mesonic chiral perturbation theory \[see Eq. (\[4:5:powercounting\]) of Sec.  \[sec\_cel\]\]. On the other hand, the counting of bilinears $\bar{\Psi}\Gamma\Psi$ is probably easiest understood by investigating the matrix elements of positive-energy plane-wave solutions to the free Dirac equation in the Dirac representation: $$\psi^{(+)}(\vec{x},t)=\exp(-ip_N\cdot x) \sqrt{E_N+m_N}\left( \begin{array}{c} \chi\\ \frac{\vec{\sigma}\cdot\vec{p}_N}{E_N+m_N}\chi \end{array} \right),$$ where $\chi$ denotes a two-component Pauli spinor and $p_N^\mu=(E_N,\vec{p}_N)$ with $E_N=\sqrt{\vec{p}\,_N^2+m_N^2}$. In the low-energy limit, i.e. for nonrelativistic kinematics, the lower (small) component is suppressed as $|\vec{p}_N|/m_N$ in comparison with the upper (large) component. For the analysis of the bilinears it is convenient to divide the 16 Dirac matrices into even and odd ones, ${\cal E}= \{1, \gamma_0,\gamma_5 \gamma_i,\sigma_{ij}\}$ and ${\cal O}=\{\gamma_5,\gamma_5 \gamma_0,\gamma_i,\sigma_{i0}\}$ [@Foldy:1949wa; @Fearing:ii], respectively, where odd matrices couple large and small components but not large with large, whereas even matrices do not. Finally, $i\partial^\mu$ acting on the nucleon solution produces $p_N^\mu$ which we write symbolically as $p_N^\mu=(m_N,\vec{0})+(E_N-m_N,\vec{p}_N\,)$ where we count the second term as ${\cal O}(p)$, i.e., as a small quantity. We are now in the position to summarize the chiral counting scheme for the (new) elements of baryon chiral perturbation theory [@Krause:xc]: $$\begin{aligned} \label{5:2:powercounting} &&\Psi,\bar{\Psi} = {\cal O}(p^0),\, D_{\mu} \Psi = {\cal O}(p^0),\, (iD\hspace{-.6em}/ -\stackrel{\circ}{m}_N)\Psi={\cal O}(p),\nonumber\\ &&1,\gamma_\mu,\gamma_5\gamma_\mu,\sigma_{\mu\nu}={\cal O}(p^0),\, \gamma_5 ={\cal O}(p),\end{aligned}$$ where the order given is the minimal one. For example, $\gamma_\mu$ has both an ${\cal O}(p^0)$ piece, $\gamma_0$, as well as an ${\cal O}(p)$ piece, $\gamma_i$. A rigorous nonrelativistic reduction may be achieved in the framework of the Foldy-Wouthuysen method [@Foldy:1949wa] or the heavy-baryon approach [@Jenkins:1990jv; @Bernard:1992qa] which will be discussed later (for a pedagogical introduction see Ref. [@Holstein:1997]). The construction of the $\mbox{SU(3)}_L\times\mbox{SU(3)}_R$ Lagrangian proceeds similarly except for the fact that the baryon fields are contained in the $3\times 3$ matrix of Eq. (\[5:1:su3oktett\]) transforming as $K B K^\dagger$. As in the mesonic sector, the building blocks are written as products transforming as $K\cdots K^\dagger$ with a trace taken at the end. The lowest-order Lagrangian reads [@Georgi; @Krause:xc] $$\label{5:2:l1su3} {\cal L}^{(1)}_{MB}=\mbox{Tr}\left[\bar{B}\left(iD\hspace{-.7em}/\hspace{.2em} -M_0\right)B\right] -\frac{D}{2}\mbox{Tr}\left(\bar{B}\gamma^\mu\gamma_5\{u_\mu,B\}\right) -\frac{F}{2}\mbox{Tr}\left(\bar{B}\gamma^\mu\gamma_5[u_\mu,B]\right),$$ where $M_0$ denotes the mass of the baryon octet in the chiral limit. The covariant derivative of $B$ is defined as $$\label{5:2:kovderb} D_\mu B=\partial_\mu B +[\Gamma_\mu,B],$$ with $\Gamma_\mu$ of Eq. (\[5:2:gamma\]) \[for $\mbox{SU(3)}_L\times\mbox{ SU(3)}_R$\]. The constants $D$ and $F$ may be determined by fitting the semi-leptonic decays $B\to B'+e^-+\bar{\nu}_e$ at tree level [@Borasoy:1998pe]: $$\label{5:2:df} D=0.80,\quad F=0.50.$$ Applications at Tree Level {#sec_aatl} -------------------------- ### Goldberger-Treiman Relation and the Axial-Vector Current Matrix Element {#subsec_gtravcme} We have seen in Sec. \[subsec\_csbdqm\] that the quark masses in QCD give rise to a non-vanishing divergence of the axial-vector current operator \[see Eq. (\[2:3:dsva\])\]. Here we will discuss the implications for the matrix elements of the pseudoscalar density and of the axial-vector current evaluated between single-nucleon states in terms of the lowest-order Lagrangians of Eqs. (\[4:5:l2\]) and (\[5:2:l1pin\]). In particular, we will see that the Ward identity $$\label{5:3:axwi} \langle N(p')|\partial_\mu A^\mu_i(0)|N(p)\rangle= \langle N(p')|m_q P_i(0)|N(p)\rangle,$$ where $m_q=m_u=m_d$, is satisfied. The nucleon matrix element of the pseudoscalar density can be parameterized as $$\label{5:3:def_gt} m_q\langle N(p')| P_i (0) | N(p) \rangle = \frac{M_\pi^2 F_\pi}{M_\pi^2 - t} G_{\pi N}(t)i\bar{u}(p') \gamma_5 \tau_i u(p),$$ where $t=(p'-p)^2$. Equation (\[5:3:def\_gt\]) [*defines*]{} the form factor $G_{\pi N}(t)$ in terms of the QCD operator $m_q P_i(x)$. As we have seen in the discussion of $\pi\pi$ scattering of Sec. \[subsec\_pps\], the operator $m_q P_i(x)/(M_\pi^2 F_\pi)$ serves as an interpolating pion field \[see Eq. (\[4:6:pionfield\])\], and thus $G_{\pi N}(t)$ is also referred to as the pion-nucleon form factor (for this specific choice of the interpolating pion field). The pion-nucleon coupling constant $g_{\pi N}$ is defined through $G_{\pi N}(t)$ evaluated at $t=M_\pi^2$. The Lagrangian ${\cal L}_{\pi N}^{(1)}$ of Eq. (\[5:2:l1pin\]) does not generate a direct coupling of an external pseudoscalar field $p_i(x)$ to the nucleon, i.e., it does not contain any terms involving $\chi$ or $\chi^\dagger$. At lowest order in the chiral expansion, the matrix element of the pseudoscalar density is therefore given in terms of the diagram of Fig. \[5:3:fig:pionnucleonformfactor\], i.e., the pseudoscalar source produces a pion which propagates and is then absorbed by the nucleon. The coupling of a pseudoscalar field to the pion in the framework of ${\cal L}_2$ has already been discussed in Eq. (\[4:6:l2ext2\]), $$\label{5:3:l2ext2} {\cal L}_{\rm ext} =i\frac{F_0^2B_0}{2}\mbox{Tr}(pU^\dagger-Up)=2B_0F_0p_i\phi_i+\cdots.$$ When working with the nonlinear realization of Eq. (\[5:1:su2real\]) it is convenient to use the exponential parameterization of Eq. (\[4:6:u2\]), $$U(x)=\exp\left[i\frac{\vec{\tau}\cdot\vec{\phi}(x)}{F_0}\right],$$ because in that case the square root is simply given by $$u(x)=\exp\left[i\frac{\vec{\tau}\cdot\vec{\phi}(x)}{2F_0}\right].$$ According to Fig. \[5:3:fig:pionnucleonformfactor\], we need to identify the interaction term of a nucleon with a single pion. In the absence of external fields the vielbein of Eq. (\[5:2:chvi\]) is odd in the pion fields, $$\label{5:3:umupin} u_\mu=i\left[u^\dagger\partial_\mu u-u\partial_\mu u^\dagger \right]\stackrel{\phi^a\mapsto -\phi^a}{\mapsto} i\left[u\partial_\mu u^\dagger -u^\dagger \partial_\mu u\right] =-u_\mu.$$ Expanding $u$ and $u^\dagger$ as $$\label{5:3:uentwi} u=1+i\frac{\vec{\tau}\cdot\vec{\phi}}{2 F_0}+O(\phi^2),\quad u^\dagger=1-i\frac{\vec{\tau}\cdot\vec{\phi}}{2 F_0}+O(\phi^2),$$ we obtain $$\begin{aligned} \label{5:3:umuentwi} u_\mu &=&-\frac{\vec{\tau}\cdot\partial_\mu\vec{\phi}}{F_0}+O(\phi^3),\end{aligned}$$ which, when inserted into ${\cal L}^{(1)}_{\pi N}$ of Eq. (\[5:2:l1pin\]), generates the following interaction Lagrangian: $$\label{5:3:lpin1} {\cal L}_{\rm int}=-\frac{1}{2}\frac{\stackrel{\circ}{g}_A}{F_0} \bar{\Psi}\gamma^\mu\gamma_5 \underbrace{\vec{\tau}\cdot\partial_\mu\vec{\phi}}_{\mbox{$ \tau^b\partial_\mu\phi^b$}}\Psi.$$ (Note that the sign is opposite to the conventionally used pseudovector pion-nucleon coupling.[^87]) The Feynman rule for the vertex of an incoming pion with four-momentum $q$ and Cartesian isospin index $a$ is given by $$\label{5:3:pionnucleonvertex} i\left(-\frac{1}{2}\frac{\stackrel{\circ}{g}_A}{F_0}\right) \gamma^\mu\gamma_5\tau^b \delta^{ba}(-i q_\mu)= -\frac{1}{2}\frac{\stackrel{\circ}{g}_A}{F_0} q\hspace{-.45em}/ \gamma_5 \tau^a.$$ On the other hand, the connection of Eq. (\[5:2:gamma\]) with the external fields set to zero is even in the pion fields, $$\label{5:3:gammapin} \Gamma_\mu=\frac{1}{2}\left[u^\dagger\partial_\mu u+u\partial_\mu u^\dagger \right]\stackrel{\phi^a\mapsto -\phi^a}{\mapsto} \frac{1}{2}\left[u\partial_\mu u^\dagger +u^\dagger \partial_\mu u\right] =\Gamma_\mu,$$ i.e., it does not contribute to the single-pion vertex. We now put the individual pieces together and obtain for the diagram of Fig. \[5:3:fig:pionnucleonformfactor\] $$\begin{aligned} \lefteqn{m_q 2 B_0 F_0 \frac{i}{t-M_\pi^2}\bar{u}(p')\left( -\frac{1}{2}\frac{\stackrel{\circ}{g}_A}{F_0} q\hspace{-.45em}/ \gamma_5 \tau_i\right)u(p)}\\ &=&M_\pi^2 F_0 \frac{\stackrel{\circ}{m}_N \stackrel{\circ}{g}_A}{F_0} \frac{1}{M_\pi^2-t}\bar{u}(p')\gamma_5 i \tau_i u(p),\end{aligned}$$ where we used $M_\pi^2=2B_0 m_q$, and the Dirac equation to show $\bar{u}q\hspace{-.5em}/ \gamma_5 u=2 \stackrel{\circ}{m}_N\bar{u}\gamma_5 u$. At ${\cal O}(p^2)$ $F_\pi=F_0$ so that, by comparison with Eq. (\[5:3:def\_gt\]), we can read off the lowest-order result $$\label{5:3:gt1} G_{\pi N}(t)= \frac{\stackrel{\circ}{m}_N}{F_0} {\stackrel{\circ}{g}}_A,$$ i.e., at this order the form factor does not depend on $t$. In general, the pion-nucleon coupling constant is defined at $t=M_\pi^2$ which, in the present case, simply yields $$\label{5:3:gpinn} g_{\pi N}=G_{\pi N}(M_\pi^2)= \frac{\stackrel{\circ}{m}_N}{F_0} {\stackrel{\circ}{g}}_A.$$ Equation (\[5:3:gpinn\]) represents the famous Goldberger-Treiman relation [@Goldberger:1958tr; @Goldberger:1958vp; @Nambu:xd] which establishes a connection between quantities entering weak processes, $F_\pi$ and $g_A$ (to be discussed below), and a typical strong-interaction quantity, namely the pion-nucleon coupling constant $g_{\pi N}$. The numerical violation of the Goldberger-Treiman relation, as expressed in the so-called Goldberger-Treiman discrepancy $$\label{5:3:gtd} \Delta_{\pi N}\equiv1-\frac{g_A m_N}{g_{\pi N}F_\pi},$$ is at the percent level,[^88] although one has to keep in mind that [*all four*]{} physical quantities move from their chiral-limit values $\stackrel{\circ}{g}_A$ etc. to the empirical ones $g_A$ etc. Using Lorentz covariance and isospin symmetry, the matrix element of the axial-vector current between initial and final nucleon states—excluding second-class currents [@Weinberg:1958ut]—can be parameterized as[^89] $$\label{5:3:axial_current} \left<N(p')| A_i^\mu (0) | N(p) \right> = \bar{u}(p') \left[ \gamma^\mu G_A(t) + \frac{(p'-p)^\mu}{2m_N} G_P(t) \right] \gamma_5 \frac{\tau_i}{2} u(p),$$ where $t = (p'-p)^2$, and $G_A(t)$ and $G_P(t)$ are the axial and induced pseudoscalar form factors, respectively. At lowest order, an external axial-vector field $a_\mu^i$ couples directly to the nucleon as $$\label{5:3:lnucleonavfc} {\cal L}_{\rm ext}={\stackrel{\circ}{g}}_A\bar{\Psi}\gamma^\mu \gamma_5 \frac{\tau_i}{2}\Psi a_\mu^i+\cdots,$$ which is obtained from ${\cal L}^{(1)}_{\pi N}$ through $u_\mu=(r_\mu-l_\mu)+\cdots = 2 a_\mu+\cdots$. The coupling to the pions is obtained from ${\cal L}_2$ with $r_\mu=-l_\mu=a_\mu$, $$\label{5:3:lpionavfc} {\cal L}_{\rm ext}=-F_0 \partial^\mu \phi_i a_\mu^i+\cdots,$$ which gives rise to a diagram similar to Fig.  \[5:3:fig:pionnucleonformfactor\], with $m_q p_i$ replaced by $a^\mu_i$. The matrix element is thus given by $$\begin{aligned} \bar{u}(p')\left\{{\stackrel{\circ}{g}}_A \gamma^\mu\gamma_5 \frac{\tau_i}{2} +\left[-\frac{1}{2}\frac{{\stackrel{\circ}{g}}_A}{F_0} (p'\hspace{-.70em}/\hspace{.45em}-p\hspace{-.45em}/)\gamma_5 \tau_i\right] \frac{i}{q^2-M^2_\pi} (-iF_0 q^\mu)\right\}u(p),\end{aligned}$$ from which we obtain, by applying the Dirac equation, $$\begin{aligned} \label{5:3:ga} G_A(t) &=& {\stackrel{\circ}{g}}_A,\\ \label{5:3:gp} G_P(t) &=& - \frac{ 4 \stackrel{\circ}{m}^2_N {\stackrel{\circ}{g}}_A}{ t-M_\pi^2}.\end{aligned}$$ At this order the axial form factor does not yet show a $t$ dependence. The axial-vector coupling constant is defined as $G_A(0)$ which is simply given by ${\stackrel{\circ}{g}}_A$. We have thus identified the second new parameter of ${\cal L}^{(1)}_{\pi N}$ besides the nucleon mass $\stackrel{\circ}{m}_N$. The induced pseudoscalar form factor is determined by the pion exchange which is the simplest version of the so-called pion-pole dominance. The $1/(t-M_\pi^2)$ behavior of $G_P$ is not in conflict with the book-keeping of a calculation at chiral order ${\cal O}(p)$, because, according to Eq.  (\[4:5:powercounting\]), the external axial-vector field $a_\mu$ counts as ${\cal O}(p)$, and the definition of the matrix element contains a momentum $(p'-p)^\mu$ and the Dirac matrix $\gamma_5$ \[see Eq.  (\[5:2:powercounting\])\] so that the combined order of all elements is indeed ${\cal O}(p)$. It is straightforward to verify that the form factors of Eqs. (\[5:3:gt1\]), (\[5:3:ga\]), and (\[5:3:gp\]) satisfy the relation $$\label{ff_relation} 2m_N G_A(t) + \frac{t}{2m_N} G_P(t) = 2\frac{M_\pi^2 F_\pi}{M_\pi^2 - t} G_{\pi N}(t),$$ which is required by the Ward identity of Eq. (\[5:3:axwi\]) with the parameterizations of Eqs. (\[5:3:def\_gt\]) and (\[5:3:axial\_current\]) for the matrix elements. In other words, only two of the three form factors $G_A$, $G_P$, and $G_{\pi N}$ are independent. Note that this relation is not restricted to small values of $t$ but holds for any $t$. In the chiral limit, Eq. (\[5:3:axwi\]) implies $$\label{5:3:gagpchirallimit} 2 \stackrel{\circ}{m}_N\, \stackrel{\circ}{G}_A(t)+\frac{t}{2 \stackrel{\circ}{m}_N} \stackrel{\circ}{G}_P(t)=0,$$ which also follows from Eq. (\[5:3:ga\]) and Eq. (\[5:3:gp\]) for $M_\pi^2=0$. Equation (\[5:3:gagpchirallimit\]) for $\stackrel{\circ}{G}_A(0)\neq 0$ requires that in the chiral limit the induced pseudoscalar form factor has a pole in the limit $t\to 0$. The interpretation of this pole is, of course, given in terms of the exchange of a massless Goldstone pion. To understand this in more detail consider the most general contribution of the pion exchange to the axial-vector current matrix element: $$\begin{aligned} \langle N(p')|A^\mu_i(0)|N(p)\rangle_\pi &=& -\frac{2 F_{\pi}(t) g_{\pi N}(t)}{t-M_\pi^2-\tilde{\Sigma}(t)} \bar{u}(p')q^\mu\gamma_5 \frac{\tau_i}{2}u(p),\end{aligned}$$ where $\tilde{\Sigma}(M^2_\pi)=\tilde{\Sigma}'(M^2_\pi)=0$ for the renormalized propagator \[see Eq. (\[4:8:sigmaexp\])\]. The functions $F_\pi(t)$ and $g_{\pi N}(t)$ denote the most general parameterizations for the pion-decay vertex and the pion-nucleon vertex (note that we have [*not*]{} specified the interpolating pion field). For general $t$ their values depend on the interpolating field, but for $t=M_\pi^2$ they are identical with the pion-decay constant $F_\pi$ and the pion-nucleon coupling constant $g_{\pi N}$, respectively. In the chiral limit, $M_\pi^2\to 0$, we obtain $$- \frac{2 F_0(t) \stackrel{\circ}{g}_{\pi N}(t)}{ t-\stackrel{\circ}{\tilde{\Sigma}}(t)}2\stackrel{\circ}{m}_N \bar{u}(p')\frac{q^\mu}{2\stackrel{\circ}{m}_N}\gamma_5 \frac{\tau_i}{2}u(p),$$ where $\stackrel{\circ}{\tilde{\Sigma}}(0)= \stackrel{\circ}{\tilde{\Sigma}'}(0)=0$. In other words, the most general contribution of a massless pion to the induced pseudoscalar form factor in the chiral limit is given by $$\stackrel{\circ}{G}_{P,\pi}(t)= -\frac{4\stackrel{\circ}{m}_N F_0(t)\stackrel{\circ}{g}_{\pi N}(t)}{ t-\stackrel{\circ}{\tilde{\Sigma}}(t)}.$$ We divide the pseudoscalar form factor into the pion contribution and the rest. Making use of Eq. (\[5:3:gagpchirallimit\]), we consider the limit $$\begin{aligned} \lim_{t\to 0}\, t[\stackrel{\circ}{G}_{P,\pi}(t) +\stackrel{\circ}{G}_{P,R}(t)] &=& -4 \stackrel{\circ}{m}_N F_0 \stackrel{\circ}{g}_{\pi N}\\ &\stackrel{!}{=}&-4 \stackrel{\circ}{m}_N^2 \stackrel{\circ}{G}_A(0)\end{aligned}$$ from which we obtain the Goldberger-Treiman relation $$\frac{\stackrel{\circ}{g}_A}{F_0}= \frac{\stackrel{\circ}{g}_{\pi N}}{\stackrel{\circ}{m}_N}.$$ Of course, we have assumed that there is no other massless particle in the theory which could produce a pole in the residual part $\stackrel{\circ}{G}_{P,R}(t)$ as $t\to 0$. ### Pion-Nucleon Scattering at Tree Level {#subsec_apnstl} As another example, we will consider pion-nucleon scattering and show how the effective Lagrangian of Eq. (\[5:2:l1pin\]) reproduces the Weinberg-Tomozawa predictions for the $s$-wave scattering lengths [@Weinberg:1966kf; @Tomozawa]. We will contrast the results with those of a tree-level calculation within pseudoscalar (PS) and pseudovector (PV) pion-nucleon couplings. Before calculating the $\pi N$ scattering amplitude within ChPT we introduce a general parameterization of the invariant amplitude ${\cal M}=iT$ for the process $\pi^a(q)+N(p)\to\pi^b(q')+N(p')$ [@Chew:1957; @Brown:1971pn]:[^90] $$\begin{aligned} \label{5:3:mpinpar} T^{ab}(p,q;p',q')&=&\bar{u}(p')\Bigg\{ \underbrace{\frac{1}{2}\{\tau^b,\tau^a\}}_{ \mbox{$\delta^{ab}$}}A^+(\nu,\nu_B) +\underbrace{\frac{1}{2}[\tau^b,\tau^a]}_{ \mbox{$-i\epsilon_{abc}\tau^c$}} A^-(\nu,\nu_B)\nonumber\\ &&+\frac{1}{2}(q\hspace{-.45em}/+q'\hspace{-.7em}/\hspace{.2em})\left[ \delta^{ab}B^+(\nu,\nu_B) -i\epsilon_{abc}\tau^c B^-(\nu,\nu_B)\right]\Bigg\} u(p),\nonumber\\\end{aligned}$$ with the two independent scalar kinematical variables $$\begin{aligned} \label{5:3:nu} \nu&=&\frac{s-u}{4m_N}=\frac{(p+p')\cdot q}{2m_N} =\frac{(p+p')\cdot q'}{2m_N},\\ \label{5:3:nub} \nu_B&=&-\frac{q\cdot q'}{2 m_N}=\frac{t-2M_\pi^2}{4 m_N},\end{aligned}$$ where $s=(p+q)^2$, $t=(p'-p)^2$, and $u=(p'-q)^2$ are the usual Mandelstam variables satisfying $s+t+u=2 m_N^2+ 2M_\pi^2$. From pion-crossing symmetry $T^{ab}(p,q;p',q')= T^{ba}(p,-q';p',-q)$ we obtain for the crossing behavior of the amplitudes $$\begin{aligned} &&A^+(-\nu,\nu_B)=A^+(\nu,\nu_B),\quad A^-(-\nu,\nu_B)=-A^-(\nu,\nu_B),\\ &&B^+(-\nu,\nu_B)=-B^+(\nu,\nu_B),\quad B^-(-\nu,\nu_B)=B^-(\nu,\nu_B).\end{aligned}$$ As in $\pi\pi$ scattering one often also finds the isospin decomposition as in Eq. (\[4:10:tdec\]), $$\langle I',{I'}\hspace{-.3em}_3|T|I,I_3\rangle=T^I \delta_{II'} \delta_{I_3 {I'}\hspace{-.1em}_3},$$ where the relation between the two sets is given by [@Ericson:gk] $$\begin{aligned} \label{5:3:trel} T^{\frac{1}{2}}&=&T^++2 T^-,\nonumber\\ T^{\frac{3}{2}}&=&T^+- T^-.\end{aligned}$$ Let us turn to the tree-level approximation to the $\pi N$ scattering amplitude as obtained from ${\cal L}^{(1)}_{\pi N}$ of Eq. (\[5:2:l1pin\]). In order to derive the relevant interaction Lagrangians from Eq. (\[5:2:l1pin\]), we reconsider the connection of Eq.  (\[5:2:gamma\]) with the external fields set to zero and obtain $$\begin{aligned} \label{5:3:gammaentwi} \Gamma_\mu&=& \frac{i}{4F^2_0}\vec{\tau}\cdot\vec{\phi}\times\partial_\mu\vec{\phi} +O(\phi^4).\end{aligned}$$ The linear pion-nucleon interaction term was already derived in Eq.(\[5:3:lpin1\]) so that we end up with the following interaction Lagrangian: $$\label{5:3:lpin} {\cal L}_{\rm int}=-\frac{1}{2}\frac{\stackrel{\circ}{g}_A}{F_0} \bar{\Psi}\gamma^\mu\gamma_5 \tau^b\partial_\mu\phi^b\Psi -\frac{1}{4F^2_0}\bar{\Psi}\gamma^\mu\underbrace{\vec{\tau}\cdot\vec{\phi} \times\partial_\mu\vec{\phi}}_{\mbox{$\epsilon_{cde}\tau^c \phi^d\partial_\mu\phi^e$}}\Psi.$$ The first term is the pseudovector pion-nucleon coupling and the second the contact interaction with two factors of the pion field interacting with the nucleon at a single point. The Feynman rules for the vertices derived from Eq. (\[5:3:lpin\]) read - for an incoming pion with four-momentum $q$ and Cartesian isospin index $a$: $$\label{5:3fr1} -\frac{1}{2}\frac{\stackrel{\circ}{g}_A}{F_0} q\hspace{-.45em}/ \gamma_5 \tau^a,$$ - for an incoming pion with $q,a$ and an outgoing pion with $q',b$: $$\label{5:3:fr2} i\left(-\frac{1}{4F_0^2}\right)\gamma^\mu\epsilon_{cde}\tau^c\left( \delta^{da}\delta^{eb}iq'_\mu+\delta^{db}\delta^{ea}(-iq)_\mu\right) =\frac{q\hspace{-.45em}/ +q'\hspace{-.7em}/}{4 F^2_0}\epsilon_{abc}\tau^c.$$ The latter gives the contact contribution to ${\cal M}$, $$\label{5:3:cont} {\cal M}_{\rm cont}=\bar{u}(p') \frac{q\hspace{-.45em}/+q'\hspace{-.7em}/\hspace{.2em}}{4 F^2_0} \underbrace{\epsilon_{abc}\tau^c}_{\mbox{$i\frac{1}{2}[\tau^b,\tau^a]$}} u(p) =i \frac{1}{2 F^2_0}\bar{u}(p')\frac{1}{2}[\tau^b,\tau^a] \frac{1}{2}(q\hspace{-.45em}/+q'\hspace{-.7em}/\hspace{.2em})u(p).$$ We emphasize that such a term is not present in a conventional calculation with either a pseudoscalar or a pseudovector pion-nucleon interaction. For the $s$- and $u$-channel nucleon-pole diagrams the pseudovector vertex appears twice and we obtain $$\begin{aligned} \label{5:3:sukanal} {\cal M}_{s+u}&=&i\frac{\stackrel{\circ}{g}_A^2}{4 F_0^2}\bar{u}(p') \tau^b\tau^a(-q'\hspace{-.7em}/\hspace{.2em})\gamma_5 \frac{1}{ p'\hspace{-.7em}/\hspace{.2em}+q'\hspace{-.7em}/\hspace{.2em} -\stackrel{\circ}{m}_N} q\hspace{-.45em}/\gamma_5u(p)\nonumber\\ &&+i\frac{\stackrel{\circ}{g}_A^2}{4 F_0^2}\bar{u}(p') \tau^a\tau^b q\hspace{-.45em}/\gamma_5 \frac{1}{p'\hspace{-.7em}/\hspace{.2em}-q\hspace{-.45em}/\hspace{.2em} -\stackrel{\circ}{m}_N} (-q'\hspace{-.7em}/\hspace{.2em})\gamma_5 u(p).\end{aligned}$$ The $s$- and $u$-channel contributions are related to each other through pion crossing $a\leftrightarrow b$ and $q\leftrightarrow -q'$. In what follows we explicitly calculate only the $s$ channel and make use of pion-crossing symmetry at the end to obtain the $u$-channel result. Moreover, we perform the manipulations such that the result of pseudoscalar coupling may also be read off. Using the Dirac equation, we rewrite $$q\hspace{-.45em}/\gamma_5 u(p)=(p'\hspace{-.7em}/\hspace{.2em} +q'\hspace{-.7em}/\hspace{.2em} -\stackrel{\circ}{m}_N +\stackrel{\circ}{m}_N\!-p\hspace{-.45em}/)\gamma_5 u(p)= (p'\hspace{-.7em}/\hspace{.2em}+q'\hspace{-.7em}/\hspace{.2em} -\stackrel{\circ}{m}_N)\gamma_5 u(p) +2\stackrel{\circ}{m}_N\gamma_5 u(p)$$ and obtain $$\begin{aligned} {\cal M}_s&=&i\frac{\stackrel{\circ}{g}_A^2}{4 F_0^2}\bar{u}(p')\tau^b\tau^a (-q'\hspace{-.7em}/\hspace{.2em})\gamma_5 \frac{1}{p'\hspace{-.7em}/\hspace{.2em} +q'\hspace{-.7em}/\hspace{.2em}-\stackrel{\circ}{m}_N} \left[(p'\hspace{-.7em}/\hspace{.2em}+q'\hspace{-.7em}/\hspace{.2em} -\stackrel{\circ}{m}_N) +2\stackrel{\circ}{m}_N\right]\gamma_5 u(p)\nonumber\\ &\stackrel{\gamma_5^2=1}{=}& i\frac{\stackrel{\circ}{g}_A^2}{4 F_0^2}\bar{u}(p')\tau^b\tau^a \left[(-q'\hspace{-.7em}/\hspace{.2em}) +(-q'\hspace{-.7em}/\hspace{.2em})\gamma_5 \frac{1}{p'\hspace{-.7em}/\hspace{.2em}+q'\hspace{-.7em}/\hspace{.2em} -\stackrel{\circ}{m}_N} 2\stackrel{\circ}{m}_N\gamma_5 \right]u(p).\end{aligned}$$ We repeat the above procedure $$\bar{u}(p')q'\hspace{-.7em}/\hspace{.2em}\gamma_5 =\bar{u}(p')[-2 \stackrel{\circ}{m}_N\!\gamma_5 -\gamma_5(p\hspace{-.45em}/+q\hspace{-.45em}/ -\stackrel{\circ}{m}_N)],$$ yielding $$\label{5:2ms1} {\cal M}_s=i\frac{\stackrel{\circ}{g}_A^2}{4 F_0^2}\bar{u}(p')\tau^b\tau^a [(-q'\hspace{-.7em}/\hspace{.2em}) +\underbrace{4m_N^2\gamma_5 \frac{1}{p'\hspace{-.7em}/\hspace{.2em}+q'\hspace{-.7em}/\hspace{.2em} -\stackrel{\circ}{m}_N} \gamma_5}_{\mbox{PS coupling}} +2\stackrel{\circ}{m}_N]u(p),$$ where, for the identification of the PS-coupling result, one has to make use of the Goldberger-Treiman relation [@Goldberger:1958tr; @Goldberger:1958vp; @Nambu:xd] (see Sec. \[subsec\_gtravcme\]) $$\frac{\stackrel{\circ}{g}_A}{F_0}=\frac{\stackrel{\circ}{g}_{\pi N} }{\stackrel{\circ}{m}_N},$$ where $\stackrel{\circ}{g}_{\pi N}$ denotes the pion-nucleon coupling constant in the chiral limit. Using $$s-m_N^2=2m_N(\nu-\nu_B),$$ we find $$\begin{aligned} \bar{u}(p')\gamma_5 \frac{1}{p'\hspace{-.7em}/\hspace{.2em}+q'\hspace{-.7em}/\hspace{.2em}- \stackrel{\circ}{m}_N} \gamma_5 u(p)&=& \bar{u}(p')\gamma_5 \frac{p'\hspace{-.7em}/\hspace{.2em}+q'\hspace{-.7em}/\hspace{.2em}+ \stackrel{\circ}{m}_N}{ (p'+q')^2-\stackrel{\circ}{m}_N^2}\gamma_5 u(p)\nonumber\\ &=&\frac{1}{2\stackrel{\circ}{m}_N(\nu-\nu_B)}\left[-\frac{1}{2}\bar{u}(p') (q\hspace{-.45em}/+q'\hspace{-.7em}/\hspace{.2em})u(p)\right],\end{aligned}$$ where we again made use of the Dirac equation. We finally obtain for the $s$-channel contribution $$\label{5:3:msf} {\cal M}_{s}=i\frac{\stackrel{\circ}{g}_A^2}{4 F_0^2} \bar{u}(p')\tau^b\tau^a\left[ 2\stackrel{\circ}{m}_N +\frac{1}{2}(q\hspace{-.45em}/+q'\hspace{-.7em}/\hspace{.2em})\left( -1-\frac{2\stackrel{\circ}{m}_N}{\nu-\nu_B}\right)\right]u(p).$$ As noted above, the expression for the $u$ channel results from the substitution $a\leftrightarrow b$ and $q\leftrightarrow -q'$ $$\label{5:3:ukanal} {\cal M}_{u}=i\frac{\stackrel{\circ}{g}_A^2}{4 F_0^2} \bar{u}(p')\tau^a\tau^b\left[ 2\stackrel{\circ}{m}_N +\frac{1}{2}(q\hspace{-.45em}/+q'\hspace{-.7em}/\hspace{.2em})\left( 1-\frac{2\stackrel{\circ}{m}_N}{\nu+\nu_B}\right)\right]u(p).$$ We combine the $s$- and $u$-channel contributions using $$\tau^b\tau^a=\frac{1}{2}\{\tau^b,\tau^a\}+\frac{1}{2}[\tau^b,\tau^a], \quad \tau^a\tau^b=\frac{1}{2}\{\tau^b,\tau^a\}-\frac{1}{2}[\tau^b,\tau^a],$$ and $$\frac{1}{\nu-\nu_B}\pm\frac{1}{\nu+\nu_B}= \frac{\left\{\begin{array}{c}2\nu\\ 2\nu_B\end{array}\right\}}{ \nu^2-\nu_B^2}$$ and summarize the contributions to the functions $A^\pm$ and $B^\pm$ of Eq. (\[5:3:mpinpar\]) in Table \[5:3:tableresults\] \[see also Eq. (A.26) of Ref. [@Gasser:1987rb]\]. -------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- amplitude$\backslash$origin PS $\Delta$PV contact sum ----------------------------- --------------------------------------------------- --------------------------------------------------------------- -------------------- --------------------------------------------------------------- $A^+$ 0 $\frac{\stackrel{\circ}{g}_A^2 \stackrel{\circ}{m}_N}{F^2_0}$ 0 $\frac{\stackrel{\circ}{g}_A^2 \stackrel{\circ}{m}_N}{F^2_0}$ $A^-$ 0 0 0 0 $B^+$ $-\frac{\stackrel{\circ}{g}_A^2}{F^2_0} 0 0 $-\frac{\stackrel{\circ}{g}_A^2}{F^2_0} \frac{\stackrel{\circ}{m}_N\nu}{\nu^2-\nu_B^2}$ \frac{\stackrel{\circ}{m}_N\nu}{\nu^2-\nu_B^2}$ $B^-$ $-\frac{\stackrel{\circ}{g}_A^2}{F^2_0} $-\frac{\stackrel{\circ}{g}_A^2}{2F^2_0}$ $\frac{1}{2F^2_0}$ $\frac{1-\stackrel{\circ}{g}_A^2}{2F^2_0} \frac{\stackrel{\circ}{m}_N\nu_B}{\nu^2-\nu_B^2}$ -\frac{\stackrel{\circ}{g}_A^2}{F^2_0} \frac{\stackrel{\circ}{m}_N\nu_B}{\nu^2-\nu_B^2}$ -------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- : \[5:3:tableresults\] Tree-level contributions to the functions $A^\pm$ and $B^\pm$ of Eq. (\[5:3:mpinpar\]). The second column (PS) denotes the result using pseudoscalar pion-nucleon coupling (using the Goldberger-Treiman relation). The sum of the second and third column (PS+$\Delta$PV) represents the result of pseudovector pion-nucleon coupling. The contact term is specific to the chiral approach. The last column, the sum of the second, third, and fourth columns, is the lowest-order ChPT result. In order to extract the scattering lengths, let us consider threshold kinematics $$\label{5:3:schkin} p^\mu=p'^\mu=(m_N,0),\quad q^\mu=q'^\mu=(M_\pi,0),\quad \nu|_{\rm thr}=M_\pi,\quad \nu_B|_{\rm thr}=-\frac{M_\pi^2}{2m_N}.$$ Since we only work at lowest-order tree level, we replace $\stackrel{\circ}{m}_N\to m_N$, etc. Together with[^91] $$u(p)\to \sqrt{2 m_N}\left(\begin{array}{c}\chi\\ 0\end{array}\right),\quad \bar{u}(p')\to \sqrt{2 m_N}\left(\chi'^\dagger\,\, 0\right)$$ we find for the threshold matrix element $$\label{5:3:mthr} T|_{\rm thr}=2 m_N \chi'^\dagger\left[\delta^{ab}\left(A^++M_\pi B^+ \right)-i\epsilon_{abc}\tau^c\left(A^-+M_\pi B^-\right)\right]_{\rm thr} \chi.$$ Using $$\left[\nu^2-\nu_B^2\right]_{\rm thr} =M^2_\pi\left(1-\frac{\mu^2}{4}\right),\quad \mu=\frac{M_\pi}{m_N}\approx \frac{1}{7}$$ we obtain $$\begin{aligned} \label{5:3:mthrres} T|_{\rm thr}&=& 2m_N\chi'^\dagger\Bigg[\delta^{ab} \underbrace{\Bigg( \frac{g^2_A m_N}{F_\pi^2}+\underbrace{ M_\pi\left(-\frac{g^2_A}{F_\pi^2}\right) \frac{m_N}{M_\pi}\frac{1}{1-\frac{\mu^2}{4}}}_{\mbox{PS}}\Bigg) }_{\mbox{ChPT = PV}} \nonumber\\ &&-i\epsilon_{abc}\tau^c M_\pi\underbrace{\Bigg( \frac{1}{2F_\pi^2} \underbrace{-\frac{g^2_A}{2F_\pi^2} \underbrace{-\frac{g^2_A}{F_\pi^2}\left(-\frac{1}{2}\right) \frac{1}{1-\frac{\mu^2}{4}}}_{\mbox{PS}}}_{\mbox{PV}} \Bigg)}_{\mbox{ChPT}}\Bigg]\chi,\end{aligned}$$ where we have indicated the results for the various coupling schemes. Let us discuss the $s$-wave scattering lengths resulting from Eq. (\[5:3:mthrres\]). Using the above normalization for the Dirac spinors, the differential cross section in the center-of-mass frame is given by [@Ericson:gk] $$\label{5:3:dscm} \frac{d\sigma}{d\Omega}=\frac{|\vec{q}\,'|}{|\vec{q}\,|} \left(\frac{1}{8\pi \sqrt{s}}\right)^2 |T|^2,$$ which, at threshold, reduces to $$\label{5:3:dsthr} \left.\frac{d\sigma}{d\Omega}\right|_{\rm thr} =\left(\frac{1}{8\pi(m_N+M_\pi)}\right)^2 |T|^2\stackrel{!}{=}|a|^2.$$ The $s$-wave scattering lengths are defined as[^92] $$\label{5:3:apm} a^\pm_{0+}= \frac{1}{8\pi(m_N+M_\pi)}T^{\pm}|_{\rm thr} =\frac{1}{4\pi(1+\mu)}\left[A^\pm+M_\pi B^\pm\right]_{\rm thr}.$$ The subscript $0+$ refers to the fact that the $\pi N$ system is in an orbital $s$ wave ($l=0$) with total angular momentum $1/2=0+1/2$. Inserting the results of Table \[5:3:tableresults\] we obtain[^93] $$\begin{aligned} \label{5:3:aminus} a^-_{0+}&=&\frac{M_\pi}{8\pi(1+\mu)F_\pi^2}\left(1+\frac{g_A^2\mu^2}{4} \frac{1}{1-\frac{\mu^2}{4}}\right) =\frac{M_\pi}{8\pi(1+\mu)F_\pi^2}[1+{\cal O}(p^2)],\nonumber\\ &&\\ \label{5:3:aplus} a^+_{0+}&=& -\frac{g_A^2 M_\pi}{16\pi(1+\mu)F_\pi^2}\frac{\mu}{1-\frac{\mu^2}{4}} = {\cal O}(p^2),\end{aligned}$$ where we have also indicated the chiral order. Taking the linear combinations $a^\frac{1}{2}=a^+_{0+}+2 a^-_{0+}$ and $a^\frac{3}{2}=a^+_{0+}-a^-_{0+}$ \[see Eq. (\[5:3:trel\])\], we see that the results of Eqs.(\[5:3:aminus\]) and (\[5:3:aplus\]) indeed satisfy the Weinberg-Tomozawa relation [@Weinberg:1966kf; @Tomozawa]:[^94] $$\label{5:3:weinbergtomozawa} a^I=-\frac{M_\pi}{8\pi (1+\mu)F_\pi^2}[I(I+1)-\frac{3}{4}-2].$$ As in $\pi\pi$ scattering, the scattering lengths vanish in the chiral limit reflecting the fact that the interaction of Goldstone bosons vanishes in the zero-energy limit. The pseudoscalar pion-nucleon interaction produces a scattering length $a^+_{0+}$ proportional to $m_N$ instead of $\mu M_\pi$ and is clearly in conflict with the requirements of chiral symmetry. Moreover, the scattering length $a^-_{0+}$ of the pseudoscalar coupling is too large by a factor $g_A^2$ in comparison with the two-pion contact term of Eq. (\[5:3:cont\]) (sometimes also referred to as the Weinberg-Tomozawa term) induced by the nonlinear realization of chiral symmetry. On the other hand, the pseudovector pion-nucleon interaction gives a totally wrong result for $a^-_{0+}$, because it misses the two-pion contact term of Eq. (\[5:3:cont\]). Using the values $$\begin{aligned} \label{5:3:par} &&g_A=1.267,\quad F_\pi=92.4\,\mbox{MeV},\nonumber\\ && m_N=m_p=938.3\,\mbox{MeV},\quad M_\pi=M_{\pi^+}=139.6\,\mbox{MeV},\end{aligned}$$ the numerical results for the scattering lengths are given in Table \[5:3:tecomp\]. We have included the full results of Eqs. (\[5:3:aminus\]) and (\[5:3:aplus\]) and the consistent corresponding prediction at ${\cal O}(p)$. The results of heavy-baryon chiral perturbation theory (HBChPT) (see Sec. \[sec\_hbf\]) are taken from Ref. [@Mojzis:1997tu]. At ${\cal O}(p^3)$ the calculation involves nine low-energy constants of the chiral Lagrangian which have been fit to the extrapolated threshold parameters of the partial wave analysis of Ref. [@Koch:ay], the pion-nucleon $\sigma$ term and the Goldberger-Treiman discrepancy. Up to and including ${\cal O}(p^4)$ the HBChPT calculation contains 14 free parameters [@Fettes:2000xg]. In Ref. [@Fettes:2000xg] the complete one-loop amplitude at ${\cal O}(p^4)$ was fit to the phase shifts provided by three different partial wave analyses [@Koch:bn] \[I\], [@Matsinos:1997pb] \[II\], and SP98 of [@SAID] \[III\]. Table \[5:3:tecomp\] includes the results for the $s$-wave scattering lengths obtained from those fits in combination with the empirical values of the three analyses. Finally, the results of the recently proposed manifestly Lorentz-invariant form of baryon ChPT \[R(elativistic)BChPT\] [@Becher:1999he] (see Sec. \[sec\_mir\]) are included up to ${\cal O}(p^4)$ [@Becher:2001hv]. The first entries (a) refer to a dispersive representation of the function $D=A+\nu B$ entering the threshold matrix element \[see Eq. (\[5:3:apm\]) and recall $\nu_{\rm thr}=M_\pi$\] whereas the second entries (b) involve only the one-loop approximation. Whereas for $a_{0+}^-$ there is no difference, the value for $a_{0+}^+$ differs substantially which has been interpreted as the result of an insufficient approximation of the one-loop representation to allow for an extrapolation from the Cheng-Dashen point \[$(\nu=0,\nu_B=0)$\] to the physical region [@Becher:2001hv]. The empirical results quoted have been taken from low-energy partial-wave analyses [@Koch:bn; @Matsinos:1997pb] and recent precision X-ray experiments on pionic hydrogen and deuterium [@Schroder:rc]. ---------------------------------------------------------------------------------------------------------------------- Scattering length $a^+_{0+}$ \[MeV$^{-1}$\] $a^-_{0+}$ \[MeV$^{-1}$\] ------------------------------------------------- --------------------------------- ---------------------------------- Tree-level result $-6.80\times 10^{-5}$ $+5.71\times 10^{-4}$ ChPT ${\cal O}(p)$ $0$ $+5.66\times 10^{-4}$ HBChPT ${\cal O}(p^2)$ [@Mojzis:1997tu] $-1.3\times 10^{-4}$ $+5.5\times 10^{-4}$ HBChPT ${\cal O}(p^3)$ [@Mojzis:1997tu] $(-7\pm 9)\times 10^{-5}$ $(+6.7\pm1.0)\times 10^{-4}$ HBChPT ${\cal O}(p^4)$ \[I\] [@Fettes:2000xg] $-6.9\times 10^{-5}$ $+6.47\times 10^{-4}$ HBChPT ${\cal O}(p^4)$ \[II\] [@Fettes:2000xg] $+3.2\times 10^{-5}$ $+5.52\times 10^{-4}$ HBChPT ${\cal O}(p^4)$ \[III\] [@Fettes:2000xg] $+1.9\times 10^{-5}$ $+6.21\times 10^{-4}$ RChPT ${\cal O}(p^4)$ (a) [@Becher:2001hv] $-6.0\times 10^{-5}$ $+6.55 \times 10^{-4}$ RChPT ${\cal O}(p^4)$ (b) [@Becher:2001hv] $-9.4\times 10^{-5}$ $+6.55 \times 10^{-4}$ PS $-1.23 \times 10^{-2}$ $+9.14 \times 10^{-4}$ PV $-6.80\times 10^{-5}$ $+5.06 \times 10^{-6}$ Empirical values [@Koch:bn] $ (-7\pm 1)\times 10^{-5}$ $(6.6 \pm 0.1) \times 10^{-4}$ Empirical values [@Matsinos:1997pb] $(2.04\pm 1.17) \times 10^{-5}$ $(5.71\pm 0.12) \times 10^{-4}$ $(5.92\pm 0.11) \times 10^{-4}$ Experiment [@Schroder:rc] $(-2.7\pm 3.6)\times 10^{-5}$ $(+6.59\pm0.30)\times 10^{-4}$ ---------------------------------------------------------------------------------------------------------------------- : \[5:3:tecomp\] $s$-wave scattering lengths $a_{0+}^\pm$. Examples of Loop Diagrams {#sec_eld} ------------------------- In Sec. \[sec\_elwpcs\] we saw that, in the purely mesonic sector, contributions of $n$-loop diagrams are at least of order ${\cal O}(p^{2n+2})$, i.e., they are suppressed by $p^{2n}$ in comparison with tree-level diagrams. An important ingredient in deriving this result was the fact that we treated the squared pion mass as a small quantity of order $p^2$. Such an approach is motivated by the observation that the masses of the Goldstone bosons must vanish in the chiral limit. In the framework of ordinary chiral perturbation theory $M_\pi^2\sim m_q$ \[see Eq. (\[4:3:mpi2\]) and the discussion at the end of Sec.  \[subsec\_eppsop6\]\] which translates into a momentum expansion of observables at fixed ratio $m_q/p^2$. On the other hand, there is no reason to believe that the masses of hadrons other than the Goldstone bosons should vanish or become small in the chiral limit. In other words, the nucleon mass entering the pion-nucleon Lagrangian of Eq. (\[5:2:l1pin\]) should—as already anticipated in the discussion following Eq. (\[5:2:l1pin\])—not be treated as a small quantity of, say, order ${\cal O}(p)$. Naturally the question arises how all this affects the calculation of loop diagrams and the setup of a consistent power counting scheme. We will follow Ref. [@Gasser:1987rb] and consider, for illustrative purposes, two examples: a one-loop contribution to the nucleon mass and a loop diagram contributing to $\pi N$ scattering. ### First Example: One-Loop Correction to the Nucleon Mass {#subsec_feolcnm} The discussion of the modification of the nucleon mass due to pion loops is very similar to that of Sec. \[subsec\_mgb\] for the masses of the Goldstone bosons. The lowest-order Feynman propagator of the nucleon, corresponding to the free-field part of ${\cal L}_{\pi N}^{(1)}$ of Eq. (\[5:2:l1pin\]), $$\label{5:4:sf} iS_F(p)=\frac{i}{p\hspace{-.45em}/ -\stackrel{\circ}{m}_N+i0^+},$$ is modified by the self energy $\Sigma(p)$ (see for example the one-loop contribution of Fig. \[5:4:fig:nsepl\]) in a way analogous to the modification of the meson propagator in Eq. (\[4:8:prop1\]), $$\frac{i}{p\hspace{-.45em}/ -\stackrel{\circ}{m}_N+i0^+} + \frac{i}{p\hspace{-.45em}/ -\stackrel{\circ}{m}_N+i0^+} [-i\Sigma(p)] \frac{i}{p\hspace{-.45em}/ -\stackrel{\circ}{m}_N+i0^+} +\cdots,$$ resulting in the full (but still unrenormalized) propagator $$\label{5:4:fullurprop} iS(p)=\frac{i}{p\hspace{-.45em}/ -\stackrel{\circ}{m}_N-\Sigma(p)+i0^+}.$$ In the absence of external fields (but including the quark mass term), the most general expression for the self energy can be written as $$\label{5:4:seansatz} \Sigma(p)=-f(p^2)p\hspace{-.45em}/+g(p^2) \stackrel{\circ}{m}_N,$$ where $f$ and $g$ are as yet undetermined functions of the invariant $p^2$. We assume that $f$ and $g$ may be determined in a perturbative (momentum or loop) expansion which, symbolically, we denote by some indicator $\lambda$, $$\begin{aligned} \label{5:4:fgexp} f(p^2,\lambda)=f_0(p^2)+\lambda f_1(p^2)+\lambda^2 f_2(p^2)+\cdots,\nonumber\\ g(p^2,\lambda)=g_0(p^2)+\lambda g_1(p^2)+\lambda^2 g_2(p^2)+\cdots.\end{aligned}$$ When switching off the interaction, we would like to recover the lowest-order result of Eq. (\[5:4:sf\]), i.e. $\Sigma\to 0$, implying $f_0=g_0=0$. The mass of the nucleon is defined through the position of the pole of the full propagator, i.e., for $p\hspace{-.45em}/=m_N$ we require $$m_N-\stackrel{\circ}{m}_N+f(m_N^2)m_N -g(m_N^2)\stackrel{\circ}{m}_N=0,$$ from which we obtain $$\label{5:4:mn} m_N=\stackrel{\circ}{m}_N\frac{1+g(m_N^2)}{1+f(m_N^2)}.$$ The perturbative result to first order in $\lambda$ reads $$\label{5:4:mnpert} m_N=\stackrel{\circ}{m}_N\frac{1+\lambda g_1(\stackrel{\circ}{m}_N^2)+\cdots}{ 1+\lambda f_1(\stackrel{\circ}{m}_N^2)+\cdots} = \stackrel{\circ}{m}_N\left\{1+\lambda\left[g_1(\stackrel{\circ}{m}_N^2) -f_1(\stackrel{\circ}{m}_N^2)\right]+\cdots\right\}.$$ \[Note that the argument $m_N^2$ of the functions $f$ and $g$ also has to be expanded in powers of $\lambda$, $m_N^2=\,\,\stackrel{\circ}{m}^2_N +\,\, O(\lambda)$.\] The wave function renormalization constant is defined through the residue at $p\hspace{-.45em}/=m_N$, $$\label{5:4:zndef} S(p)\to \frac{Z_N}{p\hspace{-.45em}/ -m_N+i0^+}\,\, \mbox{for}\,\, p\hspace{-.45em}/\to m_N,$$ i.e., the renormalized propagator, defined through $S(p)=Z_N S_R(p)$, has a pole at $p\hspace{-.45em}/=m_N$ with residue 1. Using $(p^2-m_N^2)^n=(p\hspace{-.45em}/-m_N)^n(p\hspace{-.45em}/+m_N)^n$ and Eq. (\[5:4:mn\]) we find that for $p\hspace{-.45em}/\to m_N$ $$\begin{aligned} S(p)&=&\frac{1}{p\hspace{-.45em}/[1+f(p^2)]-\stackrel{\circ}{m}_N[1+g(p^2)]}\\ &=&\left\{\vphantom{\stackrel{\circ}{m}_N} p\hspace{-.45em}/[1+f(m_N^2)+(p\hspace{-.45em}/-m_N) (p\hspace{-.45em}/+m_N)f'(m_N^2)+\cdots]\right.\\ &&\left. -\stackrel{\circ}{m}_N [1+g(m_N^2)+(p\hspace{-.45em}/-m_N) (p\hspace{-.45em}/+m_N)g'(m_N^2)+\cdots]\right\}^{-1} \\ &\to&\frac{1}{(p\hspace{-.45em}/-m_N)[1+f(m_N^2)+2 m_N^2 f'(m_N^2) -2\stackrel{\circ}{m}_N m_N g'(m_N^2)]},\end{aligned}$$ yielding for the wave function renormalization constant $$\begin{aligned} \label{5:4:wdrc} Z_N&=& \frac{1}{1+f(m_N^2)+2m_N^2 f'(m_N^2)-2\stackrel{\circ}{m}_N m_N g'(m_N^2)} \nonumber\\ &=&1-\lambda\left\{f_1(\stackrel{\circ}{m}_N^2) +2\stackrel{\circ}{m}_N^2[f'_1(\stackrel{\circ}{m}_N^2)- g'_1(\stackrel{\circ}{m}_N^2)]\right\} +\cdots.\end{aligned}$$ With these definitions let us consider the contribution of Fig.  \[5:4:fig:nsepl\] to the self energy, where, for the sake of simplicity, we perform the calculation in the chiral limit $M_\pi^2=0$. Using the vertex of Eq. (\[5:3:pionnucleonvertex\]) we obtain the contribution of the self energy $$\begin{aligned} \label{5:4:sel1} -i\stackrel{\circ}{\Sigma}_{\rm loop}(p)&=& \int \frac{d^4 k}{(2\pi)^4}\left[-\frac{1}{2} \frac{\stackrel{\circ}{g}_A}{F_0} (-k\hspace{-.45em}/)\gamma_5 \tau_i\right]\frac{i}{k^2+i0^+}\nonumber\\ &&\times \frac{i}{p\hspace{-.45em}/-k\hspace{-.45em}/- \stackrel{\circ}{m}_N+i0^+} \left[-\frac{1}{2} \frac{\stackrel{\circ}{g}_A}{F_0} k\hspace{-.45em}/\gamma_5 \tau_i\right].\end{aligned}$$ Counting powers we see that the integral has a cubic divergence. We make use of (normal) dimensional regularization [@Jegerlehner:2000dz], where the integrand is first simplified using[^95] $$\label{5:4:gammarelations} \{\gamma_\mu,\gamma_\nu\}=2 g_{\mu\nu},\quad \quad {g_\mu}^\mu=n,\quad \{\gamma_\mu,\gamma_5\}=0, \quad \gamma_5^2=1.$$ In the standard fashion, we first insert $$1=\frac{p\hspace{-.45em}/-k\hspace{-.45em}/+ \stackrel{\circ}{m}_N-i0^+}{ p\hspace{-.45em}/-k\hspace{-.45em}/+ \stackrel{\circ}{m}_N-i0^+},$$ simplify the numerator using Eq. (\[5:4:gammarelations\]), $$k\hspace{-.45em}/\gamma_5 (p\hspace{-.45em}/-k\hspace{-.45em}/+ \stackrel{\circ}{m}_N) k\hspace{-.45em}/\gamma_5 =-(p\hspace{-.45em}/+\stackrel{\circ}{m}_N)k^2 +(p^2-\stackrel{\circ}{m}_N^2)k\hspace{-.45em}/ -[(k-p)^2-\stackrel{\circ}{m}_N^2]k\hspace{-.45em}/,$$ and obtain, with $\tau_i\tau_i=3$ $$\begin{aligned} \label{5:4:sigmal} \stackrel{\circ}{\Sigma}_{\rm loop}(p)&=& \frac{3 \stackrel{\circ}{g}_A^2}{4 F_0^2} \left\{-(p\hspace{-.45em}/+\stackrel{\circ}{m}_N) \mu^{4-n}\int \frac{d^n k}{(2\pi)^n}\frac{i}{(k-p)^2-\stackrel{\circ}{m}_N^2 +i0^+}\right.\nonumber\\ &&+(p^2-\stackrel{\circ}{m}_N^2)\mu^{4-n}\int \frac{d^n k}{(2\pi)^n} \frac{i k\hspace{-.45em}/}{(k^2+i0^+)[(k-p)^2-\stackrel{\circ}{m}_N^2 +i0^+]}\nonumber\\ &&\left.-\mu^{4-n}\int \frac{d^n k}{(2\pi)^n} \frac{i k\hspace{-.45em}/}{k^2+i0^+}\right\}.\end{aligned}$$ Indeed, when discussing the contribution to the nucleon mass \[see Eq. (\[5:4:mnpert\])\] we only need to consider the first integral of Eq. (\[5:4:sigmal\]), because the second term does not contribute at $p^2=\stackrel{\circ}{m}_N^2$ and the third term vanishes in dimensional regularization because the integrand is odd in $k$. Using Eqs. (\[app:ipi\]) and (\[app:ipib\]) of Appendix \[app\_subsec\_ipi\] with the replacement $M_\pi\to\, \stackrel{\circ}{m}_N$ we obtain, in the language of Eq. (\[5:4:fgexp\]), $$\lambda f_1(\stackrel{\circ}{m}_N^2)= \frac{3 \stackrel{\circ}{g}_A^2}{4 F_0^2} I_N(0),\quad \lambda g_1(\stackrel{\circ}{m}_N^2)= -\frac{3 \stackrel{\circ}{g}_A^2}{4 F_0^2} I_N(0).$$ Applying Eq. (\[5:4:mnpert\]) we find for the nucleon mass including the one-loop contribution of Fig. \[5:4:fig:nsepl\] \[see Eq.(4.1) of Ref. [@Gasser:1987rb]\] $$\label{5:4:mnil} m_N=\stackrel{\circ}{m}_N\left[1-\frac{3 \stackrel{\circ}{g}_A^2}{2 F_0^2} I_N(0)\right],$$ where $$\begin{aligned} I_N(0)&=&\frac{\stackrel{\circ}{m}_N^2}{16\pi^2}\left[R+ \ln\left(\frac{\stackrel{\circ}{m}_N^2}{\mu^2}\right) \right]+O(n-4),\\ R&=&\frac{2}{n-4}-[\ln(4\pi)+\Gamma'(1)+1].\end{aligned}$$ The pion loop of Fig. \[5:4:fig:nsepl\] generates an (infinite) contribution to the nucleon mass, even in the chiral limit, i.e., the parameter $\stackrel{\circ}{m}_N$ of ${\cal L}_{\pi N}^{(1)}$ needs to be renormalized. The same is true for the second parameter $\stackrel{\circ}{g}_A$ [@Gasser:1987rb]. This situation is completely different from the mesonic sector, where the two parameters $F_0$ and $B_0$ of the lowest-order Lagrangian do not change due to higher-order corrections in the chiral limit. For example, in the SU(2)$\times$SU(2) sector, the pion-decay constant at ${\cal O}(p^4)$ is given by \[see Eq. (12.2) of Ref. [@Gasser:1983yg]\] $$\label{5:4:fpi} F_\pi=F_0\left[1+\frac{M^2}{16 \pi^2 F_0^2}\bar{l}_4+{\cal O}(M^4)\right],$$ where $M^2=2 B_0 m_q$, and the scale-independent low-energy parameter $\bar{l}_4$ is defined in Eq. (\[app:barli\]). Since $F_\pi\to F_0$ in the chiral limit $M^2\to 0$, the pion-decay constant in the chiral limit is still given by $F_0$ of ${\cal L}_2$. Similarly, in the chiral limit the Goldstone boson masses vanish, not only at ${\cal O}(p^2)$ but also at higher orders, as we have seen in Eqs. (\[4:8:mpi24\]) - (\[4:8:meta24\]). ### Second Example: One-Loop Correction to $\pi N$ Scattering In the previous section we have seen that the parameters of the lowest-order Lagrangian must be renormalized in the chiral limit. As a second example, we will discuss the $\pi N$-scattering loop diagram of Fig. \[5:4:schleife\], which will allow us to draw some further conclusions regarding the differences between the mesonic and baryonic sectors of ChPT. Given the Feynman rule of Eq. (\[5:3:fr2\]), the contribution of Fig. \[5:4:schleife\] to the invariant amplitude reads $$\begin{aligned} {\cal M}_{\rm loop}&=& \int \frac{d^4 k}{(2\pi)^4} \bar{u}(p') \frac{k\hspace{-.45em}/+q'\hspace{-.7em}/}{4 F^2_0} \epsilon_{cbd}\tau^d \frac{i}{p\hspace{-.45em}/+q\hspace{-.45em}/-k\hspace{-.45em}/- \stackrel{\circ}{m}_N+i0^+}\nonumber\\ && \times \frac{i}{k^2-M_\pi^2+i0^+} \frac{q\hspace{-.45em}/+k\hspace{-.45em}/}{4 F^2_0} \epsilon_{ace}\tau^e u(p),\nonumber\end{aligned}$$ where, counting powers, we expect the integral to have a cubic divergence. The isospin structure is given by $$\epsilon_{cbd}\epsilon_{ace} \tau^d\tau^e= (\delta_{be}\delta_{da}-\delta_{ba}\delta_{de})\tau^d\tau^e =\tau^a\tau^b-\delta^{ba}\underbrace{\tau^d\tau^d}_{\mbox{3}} =-\underbrace{(2\delta^{ab}+\frac{1}{2}[\tau^b,\tau^a])}_{\mbox{ isospin}},$$ i.e., the diagram contributes to both $\pm$ isospin amplitudes. We obtain $$\begin{aligned} {\cal M}_{\rm loop}&=&\mbox{isospin}\,\frac{1}{16F^4_0} \int\frac{d^4 k}{(2\pi)^4} \underbrace{\bar{u}(p') (k\hspace{-.45em}/+q'\hspace{-.7em}/) (p\hspace{-.45em}/+q\hspace{-.45em}/-k\hspace{-.45em}/+\stackrel{\circ}{m}_N) (q\hspace{-.45em}/+k\hspace{-.45em}/) u(p)}_{\mbox{$ \bar{u}(p')\widehat{O}(k)u(p)$}}\\ &&\times\frac{1}{(p+q-k)^2-\stackrel{\circ}{m}_N^2+i0^+} \frac{1}{k^2-M^2_\pi+i0^+}.\end{aligned}$$ We will outline the evaluation of the integral using dimensional regularization. To do this, we first combine the denominators using Feynman’s trick, Eq. (\[app:ipipiftrick\]) of Appendix \[app\_subsec\_ipipi\], yielding $$\int_0^1 dz \frac{1}{[k^2-2z (p+q)\cdot k +z(s-\stackrel{\circ}{m}_N^2)+(z-1)M^2_\pi+i0^+]^2},$$ where $s=(p+q)^2$. Shifting the integration variables as $k\to k+z(p+q)$, the amplitude reads $$\begin{aligned} {\cal M}_{\rm loop}&=&\mbox{isospin}\,\frac{1}{16F^4_0} \mu^{4-n}\int_0^1 dz\int\frac{d^n k}{(2\pi)^n} \bar{u}(p')\widehat{O}[k+z(p+q)]u(p)\\ &&\times \frac{1}{[k^2+z(s-\stackrel{\circ}{m}_N^2)-z^2s+(z-1)M^2_\pi+i0^+]^2}.\end{aligned}$$ For the final conclusions, it is actually sufficient to consider the chiral limit, $M_\pi^2=0$, which simplifies the discussion of the loop integral. We define $$A(z)\equiv s z^2+(\stackrel{\circ}{m}^2_N-s)z =z(sz+\stackrel{\circ}{m}^2_N-s)$$ and will discuss the properties of the function $A$ in more detail below. Note that $A$ is a real, but not necessarily positive, number. The numerator of our integral is of the form $$\widehat{O}[k+z(p+q)]=\widehat{O}_0 +\widehat{O}_1^\mu k_\mu +\widehat{O}_2^{\mu\nu} k_\mu k_\nu +\widehat{O}_3^{\mu\nu\rho} k_\mu k_\nu k_\rho,$$ generating integrals of the type $$\label{5:4:integrals} \mu^{4-n}\int\frac{d^n k}{(2\pi)^n} \frac{\{1,k_\mu,k_\mu k_\nu, k_\mu k_\nu k_\rho\}}{(k^2-A+i0^+)^2},$$ where the integrals with an odd power of integration momenta in the numerator vanish in dimensional regularization, because of an integration over a symmetric interval. (The denominator is even). Let us discuss the scalar integral (numerator 1) of Eq.  (\[5:4:integrals\]).[^96] After a Wick rotation \[see Eq. (\[app:drb:wickrotation\])\], one chooses $n$-dimensional spherical coordinates for the Euclidean integral, and the angular integration is carried out as in Eq. (\[app:drb:winkelintegration\]). The remaining one-dimensional integration can be done using Eq.(\[app:drb:allgint\]), and the result is expanded for small $\epsilon\equiv4-n$, $$\begin{aligned} \label{5:4:integralbsp} \lefteqn{\mu^{4-n}\int\frac{d^n k}{(2\pi)^n} \frac{1}{(k^2-A+i0^+)^2}=}\nonumber\\ && \frac{i}{(4\pi)^2} \Bigg[-\frac{2}{n-4}+\ln(4\pi)+\Gamma'(1) -\ln\left(\frac{A-i0^+}{\mu^2}\right)+O(n-4)\Bigg], \nonumber\\\end{aligned}$$ where $-\Gamma'(1)=\gamma_E=0.5772\cdots$ is Euler’s constant. The infinity as $n\to 4$ must be canceled by some counter term of the effective $\pi N$ Lagrangian. In order to perform the remaining integration over the Feynman parameter $z$, we make use of Eq. (\[app:drb:lna\]), $$\label{5:4:ln} \ln(A-i0^+)=\ln(|A|)-i\pi\Theta(-A)\quad\mbox{for}\,A\in R,$$ i.e., we need to discuss $A$ as a function of $z\in [0,1]$ (for, in principle, arbitrary $s$). It is easy to show that $A$ can take negative values in the interval $0\leq z\leq 1$ only if $s>\,\,\stackrel{\circ}{m}_N^2$ in which case $A\leq 0$ for $0\leq z\leq 1-\stackrel{\circ}{m}_N^2/s$. In combination with Eq. (\[5:4:ln\]) we obtain $$\begin{aligned} \int_0^1 dz\ln\left(\frac{A(z)-i0^+}{\mu^2}\right)&=& -i\pi\frac{s-\stackrel{\circ}{m}^2_N}{s}\Theta \left(s-\stackrel{\circ}{m}_N^2\right)\\ &&+ \int_0^1 dz\ln\left(\frac{|sz^2+(\stackrel{\circ}{m}_N^2-s)z|}{\mu^2}\right).\end{aligned}$$ The remaining integral can be evaluated using elementary methods, and the final expression is $$\begin{aligned} \label{5:4:integral} \int_0^1 dz\ln\left(\frac{A(z)-i0^+}{\mu^2}\right)&=& -i\pi\frac{s-\stackrel{\circ}{m}_N^2}{s} \Theta\left(s-\stackrel{\circ}{m}_N^2\right)\nonumber\\ && + \ln\left(\frac{\stackrel{\circ}{m}_N^2}{\mu^2}\right)-2 +\frac{s-\stackrel{\circ}{m}_N^2}{s}\ln\left( \frac{|s-\stackrel{\circ}{m}_N^2|}{\stackrel{\circ}{m}_N^2}\right).\nonumber\\\end{aligned}$$ At this point, we refrain from presenting the final expression of ${\cal M}_{\rm loop}$ in detail, because Eq. (\[5:4:integral\]) suffices to point out the difference between one-loop diagrams in the mesonic and the baryonic sectors. To do this, we expand $s$ for small pion four-momenta in the chiral limit about $s_0=\stackrel{\circ}{m}^2_N$: $$\begin{aligned} \frac{s-\stackrel{\circ}{m}^2_N}{\stackrel{\circ}{m}^2_N} &=&\frac{(p+q)^2-\stackrel{\circ}{m}^2_N}{\stackrel{\circ}{m}^2_N} =\frac{2p\cdot q}{\stackrel{\circ}{m}^2_N}\equiv\alpha,\nonumber\\ \label{5:4:smmbs} \frac{s-\stackrel{\circ}{m}_N^2}{s} &=&\frac{s-\stackrel{\circ}{m}_N^2}{\stackrel{\circ}{m}_N^2 +s-\stackrel{\circ}{m}_N^2} =\frac{\alpha}{1+\alpha}=\alpha-\alpha^2+\alpha^3+\cdots.\end{aligned}$$ \[Note that $\alpha$ is a small quantity of chiral order ${\cal O}(p)$.\] Taking into account that the two extracted Dirac structures (which we have not displayed) are (at least) of order ${\cal O}(p^2)$ \[see Eq. (4.3) of Ref. [@Gasser:1987rb]\], one can draw the following conclusions [@Gasser:1987rb]: - The counter term needed to renormalize the contribution of Fig.\[5:4:schleife\] must contain terms which are of order ${\cal O}(p^2)$ and ${\cal O}(p^3)$. - The finite part of the loop diagram has a logarithmic singularity of the form $p^3\ln(p)$. - Expanding the finite part of the diagram in terms of small external momenta one obtains an infinite series with arbitrary powers of (small) momenta $p$ \[see Eq. (\[5:4:smmbs\])\]. In combination with the result of the previous section we see that a loop calculation with the relativistic Lagrangians ${\cal L}^{(1)}_{\pi N}$ and ${\cal L}_2$ using dimensional regularization leads to rather different properties in the mesonic and baryonic sectors. The example of the nucleon mass shows that loop diagrams may contribute at the same order as the tree diagrams which has to be contrasted with the mesonic sector where, according to the power counting of Eq. (\[4:4:mr2\]), loops are always suppressed by a factor $p^{2N_L}$, with $N_L$ denoting the number of independent loops. In particular, with each new order of the loop expansion one has to expect that the low-energy coefficients including those of the lowest-order Lagrangian ${\cal L}^{(1)}_{\pi N}$ have to be renormalized. On the other hand, in the mesonic sector a one-loop calculation in the even-intrinsic parity sector leads to a renormalization of the ${\cal O}(p^4)$ coefficients (and possibly higher-order coefficients if vertices of higher order are used), a two-loop calculation to a renormalization of the ${\cal O}(p^6)$ and so on. A second difference refers to the orders produced by a loop contribution. In the mesonic sector, a one-loop calculation involving vertices of ${\cal L}_2$ produces exclusively an ${\cal O}(p^4)$ contribution. We have seen in the $\pi N$-scattering example above that in the baryonic sector [*all*]{} higher orders are generated, even though, in principle, there is nothing wrong with such a result as long as one can organize and predict the leading order of the corresponding contribution beforehand. In the next section we will discuss the so-called heavy-baryon formulation of ChPT [@Jenkins:1990jv; @Bernard:1992qa], which provides a framework allowing for a power counting scheme which is very similar to the mesonic sector. One trades the manifestly covariant formulation for the systematic power counting. Moreover, under certain circumstances, the results obtained in HBChPT do not converge in all of the low-energy region. This problem has recently been solved in the framework of the so-called infrared regularization [@Becher:1999he] which will be discussed in Sec. \[sec\_mir\]. The Heavy-Baryon Formulation {#sec_hbf} ---------------------------- We have already seen in Sec. \[sec\_loebl\] that the baryonic sector introduces another energy scale—the nucleon mass—which does not vanish in the chiral limit. Furthermore, the mass of the nucleon has about the same size as the scale $4\pi F_0$ which appears in the calculation of pion-loop contributions \[see, for example, the discussion of $\pi N$ scattering, where the tree-level contributions of Table \[5:3:tableresults\] are $\sim$ $1/F_0^2$, whereas the one-loop diagram of Fig. \[5:4:schleife\] is $\sim$ $1/(F_0^2 (4\pi F_0)^2)$\]. The heavy-baryon formulation of ChPT [@Jenkins:1990jv; @Bernard:1992qa] consists in an expansion (of matrix elements) in terms of $$\frac{p}{4 \pi F_0}\quad\mbox{and}\quad \frac{p}{\stackrel{\circ}{m}_N},$$ where $p$ represents a small external momentum. Clearly $p$ cannot simply be the four-momentum of the initial and final nucleons of Eq. (\[4:btbta\]), because the energy components $E_i$ and $E_f$ are not small. Instead, a method has been devised which separates an external nucleon four-momentum into a large piece of the order of the nucleon mass and a small residual component. The approach is similar to the nonrelativistic reduction of Foldy and Wouthuysen [@Foldy:1949wa] which provides a systematic procedure to block-diagonalize a relativistic Hamiltonian in $1/m$ and produce a decoupling of positive- and negative energy states to any desired order in $1/m$. A criterion for the Foldy-Wouthuysen method to work is that the potentials in the Dirac Hamiltonian (corresponding to the interaction with external fields) are small in comparison with the nucleon mass. This may be considered as the analogue of treating external fields as small quantities of order ${\cal O}(p)$ ($r_\mu$ and $l_\mu$) or ${\cal O}(p^2)$ ($f^R_{\mu\nu}$, $f^L_{\mu\nu}$, $\chi$, and $\chi^\dagger$) in ChPT. As in the previous cases we will discuss the lowest-order Lagrangian in quite some detail. For a discussion of the higher-order Lagrangians, the reader is referred to Refs. [@Ecker:1995rk; @Bernard:1997gq; @Fettes:2000gb]. ### Nonrelativistic Reduction {#subsec_nr} Before discussing the heavy-baryon framework let us start with the more familiar nonrelativistic limit of the Dirac equation for a charged particle interacting with an external electromagnetic field. Using this example, we will later be able to develop a better understanding of a peculiarity inherent in the heavy-baryon formulation regarding wave function (re)normalization. Our presentation will closely follow Refs. [@Okubo:54; @Das:jx]. Consider the Dirac equation of a point-particle of charge $q$ and mass $m$ interacting with an electromagnetic four-potential[^97] $$\label{5:5:dirac1} i\partial_0 \Psi=[\vec{\alpha}\cdot(\vec{p}-q\vec{A})+\beta m +q A_0]\Psi \equiv H\Psi,$$ where $\alpha_i$ and $\beta$ are the usual Dirac matrices $$\alpha_i=\left( \begin{array}{cc} 0_{2\times 2}&\sigma_i\\ \sigma_i&0_{2\times 2} \end{array} \right),\quad \beta=\left( \begin{array}{rr} 1_{2\times 2}& 0_{2\times 2}\\ 0_{2\times 2}&-1_{2\times 2} \end{array} \right),$$ and $\vec{p}=\vec{\nabla}/i$ is the momentum operator. For simplicity, we consider the interaction with a static external electric field, $$\vec{E}=-\vec{\nabla} A_0,\quad \vec{A}=0.$$ Since we want to describe a nonrelativistic particle-like solution, it is convenient to separate a factor $\exp(-imt)$ from the wave function,[^98] $$\Psi(\vec{x},t)=e^{-imt}\Psi'(\vec{x},t),$$ so that the Dirac equation (after multiplication with $e^{+imt}$) results in $$\label{5:5:dirac2} i\partial_0 \Psi'=[\vec{\alpha}\cdot\vec{p}+(\beta-1)m+q A_0]\Psi'\equiv H'\Psi'.$$ Note that both $H$ and $H'$ are Hermitian operators. In the spirit of the nonrelativistic reduction, we write $\Psi'$ in terms of a pair of two-component spinors $\Psi_L$ and $\Psi_S$ ($L$ for large and $S$ for small) $$\label{5:5:psiptwocomponent} \Psi'= \frac{1}{2}(1+\beta)\Psi' +\frac{1}{2}(1-\beta)\Psi' = \left( \begin{array}{c} \Psi_L\\ \Psi_S \end{array} \right),$$ and obtain, after insertion into Eq. (\[5:5:dirac2\]), a set of two coupled partial differential equations $$\begin{aligned} \label{5:5:dirac3a} (i\partial_0-q A_0)\Psi_L&=&\vec{\sigma}\cdot\vec{p}\,\Psi_S,\\ \label{5:5:dirac3b} (i\partial_0-q A_0+2m)\Psi_S&=&\vec{\sigma}\cdot\vec{p}\,\Psi_L.\end{aligned}$$ The second equation can formally be solved for $\Psi_S$, $$\label{5:5:psilsolution} \Psi_S=(2m+i\partial_0-q A_0)^{-1}\vec{\sigma}\cdot\vec{p}\,\Psi_L \equiv A \Psi_L,$$ where, for later use, we have introduced the abbreviation $A$ for the operator $(2m+i\partial_0-q A_0)^{-1}\vec{\sigma}\cdot\vec{p}$. We expand Eq. (\[5:5:psilsolution\]) in terms of $1/m$ up to and including order $1/m^2$, $$\begin{aligned} \label{5:5:apsilexp} A\Psi_L&=&\frac{\vec{\sigma}\cdot\vec{p}}{2m}\,\Psi_L -\frac{i\partial_0-q A_0}{2m} \frac{\vec{\sigma}\cdot\vec{p}}{2m}\,\Psi_L +O\left(\frac{1}{m^3}\right)\nonumber\\ &=&\frac{\vec{\sigma}\cdot\vec{p}}{2m} \left(1-\frac{i\partial_0-q A_0}{2m}\right) \Psi_L-iq\frac{\vec{\sigma}\cdot\vec{E}}{4m^2}\,\Psi_L +O\left(\frac{1}{m^3}\right)\nonumber\\ &=&\left(\frac{\vec{\sigma}\cdot\vec{p}}{2m}-iq \frac{\vec{\sigma}\cdot\vec{E}}{4m^2}\right)\Psi_L +O\left(\frac{1}{m^3}\right),\end{aligned}$$ where we made use of the commutation relation $[A_0,\vec{p}\,]=i(\vec{\nabla}A_0) =-i\vec{E}$ and of $(i\partial_0-q A_0)\Psi_L=O(1/m)\Psi_L$ \[see Eqs. (\[5:5:dirac3a\]) and (\[5:5:psilsolution\])\]. Inserting this result into the right-hand side of Eq. (\[5:5:dirac3a\]) and using $\vec{\sigma}\cdot\vec{A}\,\vec{\sigma}\cdot\vec{B}=\vec{A}\cdot\vec{B} +i\vec{\sigma}\cdot\vec{A}\times\vec{B}$, we obtain the Schrödinger-type equation $$\label{5:5:schroedinger} (i\partial_0-q A_0)\Psi_L= \left\{\frac{\vec{p}\,^2}{2m}-\frac{q}{4m^2}\left[ (\vec{\nabla}\cdot\vec{E}) +\vec{\sigma}\cdot\vec{E}\times \vec{p}+i\vec{E}\cdot\vec{p}\,\right] \right\}\Psi_L.$$ As already noted by Okubo [@Okubo:54], the last term on the right-hand side of Eq. (\[5:5:schroedinger\]) is not Hermitian and, when written as $V=-\vec{d}\cdot\vec{E}$, represents the interaction of an electric field with a (momentum-dependent) imaginary electric dipole moment $\vec{d}=iq\vec{p}/(4m^2)$ [@Das:jx].[^99] As pointed out in Ref. [@Das:jx], the non-Hermiticity of the Hamilton operator of Eq. (\[5:5:schroedinger\]) is a consequence of the procedure for eliminating the small-component spinors. The method can be thought of as applying the transformation $$\label{5:5:strans} S=\left(\begin{array}{cc} 1_{2\times 2} & 0_{2\times 2}\\ -A & 1_{2\times 2} \end{array} \right)$$ to the four-component spinor $\Psi'$ to generate a four-component spinor consisting exclusively of the upper component $\Psi_L$, and then solving the corresponding transformed Dirac equation. Since $S$ is [*not*]{} a unitary operator, i.e., $$\left(\begin{array}{cc} 1_{2\times 2} &-A^\dagger\\ 0_{2\times 2} & 1_{2\times 2} \end{array}\right) =S^\dagger\neq S^{-1} =\left(\begin{array}{cc} 1_{2\times 2}&0_{2\times 2}\\ A&1_{2\times 2} \end{array} \right),$$ the norm of the original spinor $\Psi'$ and the transformed spinor $\Psi_L$, in general, will not be the same $$\label{5:5:norm} \int d^3 x \Psi'^\dagger\Psi'=\int d^3 x (\Psi^\dagger_L\Psi_L+ \Psi^\dagger_S\Psi_S)= \int d^3x \Psi^\dagger_L(1+A^\dagger A)\Psi_L\neq \int d^3 x \Psi_L^\dagger\Psi_L.$$ Equation (\[5:5:norm\]) suggests considering a field redefinition of the form [@Okubo:54; @Das:jx] $$\label{5:5:psird} \widetilde{\Psi}_L=(1+A^\dagger A)^{\frac{1}{2}}\Psi_L,$$ so that the new spinor $\widetilde{\Psi}_L$ has the same norm as $\Psi'$. For the specific Hamiltonian of Eq. (\[5:5:dirac2\]) we have $$A=\frac{\vec{\sigma}\cdot\vec{p}}{2m}+O\left(\frac{1}{m^2}\right),$$ so that we find[^100] $$\label{5:5:psilpsiltilde} \Psi_L=(1+A^\dagger A)^{-\frac{1}{2}}\widetilde{\Psi}_L =\left[1-\frac{\vec{p}\,^2}{8m^2}+O\left(\frac{1}{m^3}\right)\right] \widetilde{\Psi}_L.$$ When inserting Eq. (\[5:5:psilpsiltilde\]) into Eq. (\[5:5:schroedinger\]), we make use of $$A_0 \vec{p}\,^2 \widetilde{\Psi}_L= \vec{p}\,^2 (A_0 \widetilde{\Psi}_L)- (\vec{\nabla}\cdot\vec{E}) \widetilde{\Psi}_L -2i \vec{E}\cdot\vec{p}\,\widetilde{\Psi}_L$$ and, as above, $(i\partial_0-q A_0)\widetilde{\Psi}_L= O(1/m)\widetilde{\Psi}_L$, yielding the Schrödinger equation for the two-component spinor $\widetilde{\Psi}_L$, including relativistic corrections up to order $1/m^2$, $$\label{5:5:schroedinger2} (i\partial_0-q A_0)\widetilde{\Psi}_L= \left[\frac{\vec{p}\,^2}{2m} -\frac{q}{4m^2}\vec{\sigma}\cdot\vec{E}\times \vec{p} -\frac{q}{8m^2}(\vec{\nabla}\cdot\vec{E}) \right]\widetilde{\Psi}_L,$$ where the second term, for a central potential, corresponds to the usual spin-orbit interaction and the last term is the so-called Darwin term [@Bjorken_1964; @Itzykson:rh]. Note that the Hamiltonian here [*is*]{} Hermitian, i.e., the imaginary dipole moment has disappeared. Moreover, because of Eqs. (\[5:5:norm\]) and (\[5:5:psird\]), the spinors are normalized conventionally. The result of Eq. (\[5:5:schroedinger2\]) is identical with a nonrelativistic reduction using the Foldy-Wouthuysen method [@Foldy:1949wa] which uses a sequence of unitary transformations to block-diagonalize a relativistic Hamiltonian of the form $$\label{5:5:hoe} H=\beta m+{\cal O}+{\cal E}$$ to any desired order in $1/m$. In Eq. (\[5:5:hoe\]) ${\cal O}$ and ${\cal E}$ denote the so-called odd and even operators of $H$, respectively, where odd operators couple large and small components whereas even operators do not. In the present case we have $${\cal O}=\vec{\alpha}\cdot\vec{p},\quad {\cal E}=q A_0,$$ and after three successive transformations one obtains the block-diagonal Hamiltonian (see, e.g., Refs. [@Bjorken_1964; @Itzykson:rh; @Fearing:ii]) $$\begin{aligned} \label{5:5:hfw3} H_{\rm FW}^{(3)}&=&\beta\left(m+\frac{\vec{p}\,^2}{2m}\right) +q A_0-\frac{1}{8m^2} [\vec{\alpha}\cdot\vec{p},[\vec{\alpha}\cdot\vec{p},q A_0]] +O\left(\frac{1}{m^3}\right)\nonumber\\ &=&\beta\left(m+\frac{\vec{p}\,^2}{2m}\right)+q A_0 -\frac{q}{8m^2}(\vec{\nabla}\cdot\vec{E}) -\frac{q}{4m^2}\vec{\Sigma}\cdot\vec{E}\times \vec{p}\, +O\left(\frac{1}{m^3}\right),\nonumber\\\end{aligned}$$ where $$\label{5:5:Sigma} \vec{\Sigma}= \left( \begin{array}{cc} \vec{\sigma}&0_{2\times 2}\\ 0_{2\times 2}&\vec{\sigma} \end{array} \right).$$ Restricting ourselves to the upper left block of Eq. (\[5:5:hfw3\]) and noting that in Eq. (\[5:5:dirac2\]) we have already separated the time dependence $\exp(-imt)$ from $\Psi$, we find that Eqs.  (\[5:5:schroedinger2\]) and (\[5:5:hfw3\]) are indeed identical. In Ref. [@Das:jx], the equivalence of the two approaches was explicitly shown to order $1/m^5$. We will see that the heavy-baryon approach proceeds along lines very similar to the nonrelativistic reduction leading from Eq. (\[5:5:dirac1\]) to (\[5:5:schroedinger\]). In analogy to Eqs. (\[5:5:norm\]) and (\[5:5:psird\]) we thus have to be alert to surprises related to the normalization of the relevant wave functions. ### Light and Heavy Components {#subsec_lhc} As mentioned above, the idea of the heavy-baryon approach consists of separating the large nucleon mass from the external four-momenta of the nucleons in the initial and final states and, in a sense to be discussed in Sec. \[subsec\_lol\] below, eliminating it from the Lagrangian. The starting point is the relativistic Lagrangian of Eq.  (\[5:2:l1pin\]), $$\label{5:4:lpin1} {\cal L}^{(1)}_{\pi N}=\bar{\Psi}\left(iD\hspace{-.7em}/ -m +\frac{{g}_A}{2}u\hspace{-.5em}/\gamma_5\right)\Psi,$$ where the covariant derivative $D_\mu \Psi$ and $u_\mu$ are defined in Eqs. (\[5:2:kovderpsi\]) and (\[5:2:chvi\]), respectively. The corresponding Euler-Lagrange equation for the nucleon field reads $$\label{5:5:eom1} -\partial_\mu \frac{\partial{\cal L}^{(1)}_{\pi N}}{\partial \partial_\mu \bar{\Psi}} +\frac{\partial{\cal L}^{(1)}_{\pi N}}{\partial \bar{\Psi}}= \left(iD\hspace{-.7em}/-m +\frac{{g}_A}{2}u\hspace{-.5em}/\gamma_5\right) \Psi=0.$$ (For notational convenience we replace $\stackrel{\circ}{m}_N\to m$ and $\stackrel{\circ}{g}_A\to g_A$ in Secs. \[subsec\_lhc\] and \[subsec\_lol\]). For a general four-vector $v^\mu$ with the properties $v^2=1$ and $v^0\geq 1$, we define the projection operators[^101] $$\label{5:4:ppmdef} P_{v\pm}\equiv\frac{1\pm v\hspace{-.5em}/}{2},\quad P_{v+}+P_{v-}=1,\quad P_{v\pm}^2=P_{v\pm},\quad P_{v\pm}P_{v\mp}=0,$$ and introduce the so-called velocity-dependent fields ${\cal N}_v$ and ${\cal H}_v$ as $$\label{5:5:nh} {\cal N}_v\equiv e^{imv\cdot x}P_{v+}\Psi, \quad {\cal H}_v\equiv e^{imv\cdot x}P_{v-}\Psi,$$ so that $\Psi$ can be written as $$\label{5:5:psinh} \Psi(x)=e^{-imv\cdot x}\left[{\cal N}_v(x)+{\cal H}_v(x)\right].$$ The fields ${\cal N}_v$ and ${\cal H}_v$ satisfy the properties $$\label{5:5:nhprop} v\hspace{-.5em}/ {\cal N}_v={\cal N}_v,\quad v\hspace{-.5em}/ {\cal H}_v=-{\cal H}_v.$$ For a particle with four-momentum $p^\mu=(E,\vec{p}\,)$ the particular choice $v^\mu=p^\mu/m$ corresponds to its world velocity which is why $v$ is also referred to as a four-velocity. The fields ${\cal N}_v$ and ${\cal H}_v$ are often called the light and heavy components of the field $\Psi$, which will become clearer below. In order to motivate the ansatz of Eq. (\[5:5:psinh\]) let us consider a positive-energy plane wave solution to the free Dirac equation with three-momentum $\vec{p}$: $$\begin{aligned} \psi^{(+)(\alpha)}_{\vec{p}}(\vec{x},t)&=& u^{(\alpha)}(\vec{p}\,) e^{-ip\cdot x},\\ u^{(\alpha)}(\vec{p}\,)&=& \sqrt{E(\vec{p}\,)+m} \left(\begin{array}{c}\chi^{(\alpha)} \\ \frac{\vec{\sigma}\cdot \vec{p}}{E(\vec{p}\,)+m}\chi^{(\alpha)}\end{array}\right),\end{aligned}$$ where $$\chi^{(1)}=\left(\begin{array}{c}1\\0\end{array}\right),\quad \chi^{(2)}=\left(\begin{array}{c}0\\1\end{array}\right),$$ are ordinary two-component Pauli spinors, and $E(\vec{p}\,)=\sqrt{m^2+\vec{p}\,^2}$. We can think of $\psi^{(+)(\alpha)}_{\vec{p}}(\vec{x},t)$ entering the calculation of, say, an $S$-matrix element through covariant perturbation theory in terms of the matrix element of an in-field $\Psi_{\rm in}(x)$ between the vacuum and a single-nucleon state: $$\langle 0| \Psi_{\rm in}(x)|N(\vec{p},\alpha),{\rm in}\rangle= u^{(\alpha)}(\vec{p}\,) e^{-ip\cdot x}\chi_N,$$ where $\chi_N$ denotes the nucleon isospinor. For the special case $v^\mu=(1,0,0,0)\equiv v^\mu_1$, i.e. $$P_{v_1+}=\left(\begin{array}{cc}1_{2\times 2}&0_{2\times 2}\\ 0_{2\times 2}&0_{2\times 2}\end{array}\right),\quad P_{v_1-}=\left(\begin{array}{cc}0_{2\times 2}&0_{2\times 2}\\ 0_{2\times 2}&1_{2\times 2}\end{array}\right),$$ the components $N_{v_1}$ and $H_{v_1}$ are, up to the modified time dependence, equivalent to the large and small components of the “one-particle wave function” $$\begin{aligned} \label{5:5:nhspwf} N_{v_1}^{(\alpha)}(x)&=&\sqrt{E(\vec{p}\,)+m} \left(\begin{array}{c}\chi^{(\alpha)} \\ 0_{2\times 1}\end{array}\right) e^{-i[E(\vec{p}\,)-m]t+i\vec{p}\cdot\vec{x}},\nonumber\\ H_{v_1}^{(\alpha)}(x)&=&\sqrt{E(\vec{p}\,)+m} \left(\begin{array}{c}0_{2\times 1}\\ \frac{\vec{\sigma}\cdot \vec{p}}{E(\vec{p}\,)+m}\chi^{(\alpha)}\end{array}\right) e^{-i[E(\vec{p}\,)-m]t+i\vec{p}\cdot\vec{x}}.\end{aligned}$$ In other words, for this choice of $v$ the light and heavy components of the positive-energy solutions are closely related to the large and small components of the nonrelativistic reduction discussed in Sec. \[subsec\_nr\]. Moreover, assuming $|\vec{p}\,|\ll m$, $\exp[-i(E-m)t]$ varies slowly with time in comparison with $\exp(-iEt)$ of $\psi^{(+)(\alpha)}_{\vec{p}}(\vec{x},t)$, with the result that a time derivative $i\partial/\partial t$ generates a factor $(E-m)$ which is small in comparison with $m$. Another choice is $v^\mu=p^\mu/m\equiv v^\mu_2$, in which case $P_{v_2+}$ and $P_{v_2-}$ correspond to the usual projection operators for positive- and negative-energy states $$P_{v_2\pm}=\Lambda_{\pm}(p) =\frac{\pm p\hspace{-.45em}/\hspace{.2em} + m}{2m}.$$ For this case we find $$\begin{aligned} \label{5:5:nhspwf2} N_{v_2}^{(\alpha)}(x)&=&u^{(\alpha)}(\vec{p}\,),\nonumber\\ H_{v_2}^{(\alpha)}(x)&=&0,\end{aligned}$$ i.e., the $x$ dependence has completely disappeared in $N_{v_2}$ and, due to the projection property $\Lambda_{-}(p) u^{(\alpha)}(\vec{p}\,)=0$, $H_{v_2}$ vanishes identically. In general, one decomposes the four-momentum $p^\mu$ of a low-energy nucleon into $m v^\mu$ and a residual momentum $k^\mu$,[^102] $$\label{5:5:pmvk} p^\mu=m v^\mu+k^\mu,$$ so that $$\label{5:5:vkcond} v\cdot k=-\frac{k^2}{2m}\stackrel{v^\mu=(1,0,0,0)}{=}k_0=E-m\ll m.$$ For $v^\mu$ in the vicinity of $(1,0,0,0)$, a partial derivative $i\partial^\mu$ acting on $e^{-ip\cdot x+imv\cdot x}$ produces a small residual momentum $k^\mu$ and, in particular, $$iv\cdot\partial\mapsto v\cdot k \ll m.$$ The actual choice of $v^\mu$ is, to some extent, a matter of convenience. For low-energy processes involving a single nucleon in the initial and final states, the four-momentum $q^\mu$ transferred in the reaction is defined as $q=p_f-p_i$, and is considered as a small quantity of chiral order ${\cal O}(p)$. For $p_i=m v+k_i$ and $p_f=mv+k_f$, where, say, $k_i$ is a small residual momentum in the sense of Eq. (\[5:5:vkcond\]), also $k_f=k_i+q$ is a small four-momentum. The implications on a chiral power-counting scheme will be discussed in Sec. \[subsec\_pcs\] below. ### Lowest-Order Lagrangian {#subsec_lol} In order to proceed with the construction of the lowest-order heavy-baryon Lagrangian we insert Eq. (\[5:5:psinh\]) into the EOM of Eq. (\[5:5:eom1\]),[^103] $$\begin{aligned} \lefteqn{\left(iD\hspace{-.7em}/-m +\frac{{g}_A}{2}u\hspace{-.5em}/\gamma_5\right) e^{-imv\cdot x}\left({\cal N}_v+{\cal H}_v\right)=}\\ &&e^{-imv\cdot x}\left(mv\hspace{-.5em}/ +iD\hspace{-.7em}/-m +\frac{{g}_A}{2}u\hspace{-.5em}/\gamma_5\right) \left({\cal N}_v+{\cal H}_v\right)=0,\end{aligned}$$ make use of Eq. (\[5:5:nhprop\]), multiply by $e^{im v\cdot x}$, and obtain $$\label{5:5:eom2} \left(iD\hspace{-.7em}/+\frac{g_A}{2}u\hspace{-.5em}/\gamma_5\right){\cal N}_v +\left(iD\hspace{-.7em}/-2m +\frac{g_A}{2}u\hspace{-.5em}/\gamma_5\right) {\cal H}_v=0.$$ In the next step we would like to separate the $P_{v+}$ and the $P_{v-}$ part of the EOM of Eq. (\[5:5:eom2\]). To that end we make use of the algebra of the gamma matrices to derive $$\begin{aligned} \label{5:5:ps} P_{v+}A\hspace{-.6em}/\hspace{.2em}P_{v+}&=&v\cdot A\, P_{v+},\nonumber\\ P_{v+}A\hspace{-.6em}/P_{v-}&=&A\hspace{-.6em}/_\bot P_{v-} =P_{v+}A\hspace{-.6em}/_\bot, \nonumber\\ P_{v-}A\hspace{-.6em}/\hspace{.2em}P_{v-}&=&-v\cdot A\, P_{v-},\nonumber\\ P_{v-}A\hspace{-.6em}/\hspace{.2em}P_{v+}&=&A\hspace{-.6em}/_\bot P_{v+}= P_{v-}A\hspace{-.6em}/_\bot, \nonumber\\ P_{v+}B\hspace{-.6em}/\hspace{.2em} \gamma_5 P_{v+}&=&B\hspace{-.6em}/_\bot\gamma_5 P_{v+}, \nonumber\\ P_{v+}B\hspace{-.6em}/\hspace{.2em}\gamma_5 P_{v-}&=&v\cdot B\,\gamma_5 P_{v-} =v\cdot B \,P_{v+}\gamma_5,\nonumber\\ P_{v-}B\hspace{-.6em}/\hspace{.2em} \gamma_5 P_{v-}&=&B\hspace{-.6em}/_\bot\gamma_5 P_{v-}, \nonumber\\ P_{v-}B\hspace{-.6em}/\hspace{.2em} \gamma_5 P_{v+}&=&-v\cdot B\gamma_5 P_{v+} =-v\cdot B P_{v-}\gamma_5,\end{aligned}$$ where $$P_{v\pm}=\frac{1\pm v\hspace{-.5em}/}{2}, \quad v^2=1,\quad A_\bot^\mu= A^\mu- v\cdot A\, v^\mu,\quad v\cdot A_\bot=0,\quad A\hspace{-.6em}/_\bot=A_\bot^\mu \gamma_\mu.$$ As an example, let us explicitly show the first relation of Eq.(\[5:5:ps\]) $$\begin{aligned} P_{v+}A\hspace{-.6em}/\hspace{.2em}P_{v+} &=&\frac{1}{2}(1+v\hspace{-.5em}/)A\hspace{-.6em}/\hspace{.2em}P_{v+} =\frac{1}{2}(A\hspace{-.6em}/+v\hspace{-.5em}/ \hspace{.2em}A\hspace{-.6em}/\hspace{.2em})P_{v+} =\frac{1}{2}(A\hspace{-.6em}/-A\hspace{-.6em}/ \hspace{.2em}v\hspace{-.5em}/+2v\cdot A)P_{v+} \\ &=&(A\hspace{-.6em}/\hspace{.2em}P_{v-}+v\cdot A)P_{v+} =v\cdot A\, P_{v+}.\end{aligned}$$ The remaining results of Eq. (\[5:5:ps\]) follow analogously. Using Eqs. (\[5:4:ppmdef\]) and (\[5:5:ps\]) we are now in the position to project onto the $P_{v+}$ and $P_{v-}$ parts of the EOM of Eq. (\[5:5:eom2\]), $$\begin{aligned} \label{5:5:eom3a} \left(iv\cdot D+\frac{g_A}{2}u\hspace{-.5em}/_\bot\gamma_5\right){\cal N}_v +\left(iD\hspace{-.7em}/_\bot+\frac{g_A}{2}v\cdot u \gamma_5\right) {\cal H}_v&=&0,\\ \label{5:5:eom3b} \left(iD\hspace{-.7em}/_\bot-\frac{g_A}{2}v\cdot u \gamma_5\right){\cal N}_v +\left(-iv\cdot D-2m+\frac{g_A}{2}u\hspace{-.5em}/_\bot\gamma_5\right) {\cal H}_v&=&0,\end{aligned}$$ which corresponds to Eqs. (\[5:5:dirac3a\]) and (\[5:5:dirac3b\]) of the nonrelativistic reduction of Sec. \[subsec\_nr\]. We formally solve Eq. (\[5:5:eom3b\]) for ${\cal H}_v$, $$\label{5:5:solutionhv} {\cal H}_v=\left(2m+iv\cdot D-\frac{g_A}{2}u\hspace{-.5em}/_\bot \gamma_5\right)^{-1} \left(iD\hspace{-.7em}/_\bot-\frac{g_A}{2}v\cdot u \gamma_5\right) {\cal N}_v,$$ which, similar to the discussion of Sec. \[subsec\_nr\], shows that the heavy component ${\cal H}_v$ is formally suppressed by powers of $1/m$ relative to the light component ${\cal N}_v$.[^104] Inserting Eq. (\[5:5:solutionhv\]) into Eq. (\[5:5:eom3a\]), the EOM for the light component reads $$\begin{aligned} \label{5:5:eomnv} \lefteqn{\left(iv\cdot D+\frac{g_A}{2}u\hspace{-.5em}/_\bot \gamma_5\right){\cal N}_v +\left(i D\hspace{-.7em}/_\bot+\frac{g_A}{2}v\cdot u \gamma_5\right)} \nonumber\\ &&\times \left(2m+iv\cdot D-\frac{g_A}{2}u\hspace{-.5em}/_\bot \gamma_5\right)^{-1} \left(iD\hspace{-.7em}/_\bot-\frac{g_A}{2}v\cdot u \gamma_5\right) {\cal N}_v=0,\end{aligned}$$ which represents the analogue of Eq. (\[5:5:schroedinger\]). This EOM may be obtained from applying the variational principle to the effective Lagrangian[^105] $$\begin{aligned} \label{5:5:lhneff} {\cal L}_{\rm eff}&=&\bar{\cal N}_v\left(iv\cdot D +\frac{g_A}{2}u\hspace{-.5em}/_\bot\gamma_5\right){\cal N}_v + \bar{\cal N}_v \left(i D\hspace{-.7em}/_\bot+\frac{g_A}{2}v\cdot u \gamma_5\right)\nonumber\\ &&\times \left(2m+iv\cdot D -\frac{g_A}{2}u\hspace{-.5em}/_\bot \gamma_5\right)^{-1} \left(i D\hspace{-.7em}/_\bot-\frac{g_A}{2}v\cdot u \gamma_5\right) {\cal N}_v.\end{aligned}$$ Note that the second term is suppressed by $1/m$ relative to the first term. Equation (\[5:5:lhneff\]) corresponds to the leading-order result for Eq. (A.10) of Ref. [@Bernard:1992qa] which was obtained in the framework of the path-integral approach, but does not yet represent the final form commonly used in HBChPT.[^106] Having the discussion following Eq. (\[5:5:strans\]) in mind, in order for the two Lagrangians of Eqs. (\[5:4:lpin1\]) and (\[5:5:lhneff\]) to describe the same observables, we cannot expect both fields $\Psi$ and ${\cal N}_v$ to be normalized in the same way. We will come back to this question in Sec. \[subsec\_nfs\]. To obtain the heavy-baryon Lagrangian we define the spin matrix $S^\mu_v$ as[^107] $$\label{5:5:spinop} S^\mu_v=\frac{i}{2}\gamma_5\sigma^{\mu\nu}v_\nu=-\frac{1}{2}\gamma_5 (\gamma^\mu v\hspace{-.5em}/-v^\mu),\quad S_v^{\mu\dagger}=\gamma_0S_v^\mu \gamma_0,$$ which, in four dimensions, has the properties $$\label{5:5:seig} v\cdot S_v=0,\quad \{S^\mu_v, S^\nu_v\}=\frac{1}{2}(v^\mu v^\nu-g^{\mu\nu}),\quad [S^\mu_v,S^\nu_v]=i\epsilon^{\mu\nu\rho\sigma} v_\rho S_\sigma^v.$$ Using the properties of Eq. (\[5:5:nhprop\]) together with Eq. (\[5:5:seig\]), the 16 combinations $\bar{\cal N}_v \Gamma {\cal N}_v$ may be written as \[see Eqs. (9) - (12) of Ref. [@Jenkins:1990jv]\] $$\begin{aligned} \label{5:5:nvbilineare} (\bar{\cal N}_v1_{4\times 4}{\cal N}_v&=&\bar{\cal N}_v1_{4\times 4} {\cal N}_v,) \nonumber\\ \bar{\cal N}_v\gamma_5{\cal N}_v&=&0,\nonumber\\ \bar{\cal N}_v\gamma^\mu{\cal N}_v&=&v^\mu \bar{\cal N}_v{\cal N}_v,\nonumber\\ \bar{\cal N}_v\gamma^\mu\gamma_5{\cal N}_v &=&2 \bar{\cal N}_v S^\mu_v{\cal N}_v,\nonumber\\ \bar{\cal N}_v\sigma^{\mu\nu}{\cal N}_v &=&2\epsilon^{\mu\nu\rho\sigma}v_\rho \bar{\cal N}_v S_\sigma^v {\cal N}_v,\nonumber\\ \bar{\cal N}_v\sigma^{\mu\nu}\gamma_5{\cal N}_v &=&2i (v^\mu \bar{\cal N}_v S^\nu_v{\cal N}_v -v^\nu\bar{\cal N}_v S^\mu_v{\cal N}_v).\end{aligned}$$ For example, $$\bar{\cal N}_v\gamma_5 {\cal N}_v=\bar{\cal N}_v\gamma_5 v\hspace{-.5em}/ \hspace{.2em} {\cal N}_v=-\bar{\cal N}_v v\hspace{-.5em}/\hspace{.2em}\gamma_5{\cal N}_v =-\bar{\cal N}\gamma_5{\cal N}_v=0,$$ where we made use of Eq. (\[5:5:nhprop\]). The remaining relations are shown analogously. Eqs. (\[5:5:nvbilineare\]) result in a nice simplification of the Dirac structures in the heavy-baryon approach, because one ultimately only deals with two groups of $4\times 4$ matrices, the unit matrix and $S^\mu_v$, instead of the original six groups on the left-hand side of Eq. (\[5:5:nvbilineare\]). Expanding Eq. (\[5:5:lhneff\]) formally into a series in $1/m$, $$\label{5:5:lhneffexp} {\cal L}_{\rm eff}=\bar{\cal N}_v\left(iv\cdot D +\frac{g_A}{2}u\hspace{-.5em}/_\bot\gamma_5\right){\cal N}_v +\sum_{n=1}^\infty \frac{1}{(2m)^n}{\cal L}_{{\rm eff},n},$$ and applying Eq. (\[5:5:nvbilineare\]), the lowest-order term reads $$\label{5:5:l1pinhb} \widehat{\cal L}^{(1)}_{\pi N}= \bar{\cal N}_v\left[iv\cdot D +g_A S_v\cdot u \right]{\cal N}_v,$$ where we made use of the fourth relation of Eq. (\[5:5:nvbilineare\]) and the first relation of Eq. (\[5:5:seig\]). (Recall that the second term of Eq. (\[5:5:lhneff\]) is of order $1/m$.) Equation (\[5:5:l1pinhb\]) represents the lowest-order Lagrangian of heavy-baryon chiral perturbation theory (HBChPT), indicated by the symbol $\widehat{}$. When comparing with the relativistic Lagrangian of Eq. (\[5:4:lpin1\]), we see that the nucleon mass has disappeared from the leading-order Lagrangian. It only shows up in higher orders as powers of $1/m$, as will be discussed in Sec. \[subsec\_cfo1om\]. In the power counting scheme $\widehat{\cal L}^{(1)}_{\pi N}$ counts as ${\cal O}(p)$, because the covariant derivative $D_\mu$ and the chiral vielbein $u_\mu$ both count as ${\cal O}(p)$. When calculating loop diagrams with the Lagrangian of Eq.  (\[5:5:l1pinhb\]) one will encounter divergences which are treated in the framework of (normal) dimensional regularization \[see Eq. (\[5:4:gammarelations\])\]. Since the definition of the spin matrix $S^\mu_v$ contains $\gamma_5$ and the commutator of two such spin matrices, in four dimensions, involves the epsilon tensor, one needs some convention for dealing with products of spin matrices when evaluating integrals in $n$ dimensions. To be on the safe side, we always reduce the gamma matrices using only the rules of Eq. (\[5:4:gammarelations\]). Let us consider the following example, which appears in the calculation of the pion-nucleon form factor at the one-loop level:[^108] $$\begin{aligned} \label{5:5:sss} S_\mu^v S_\nu^v S_v^\mu&=&-\frac{1}{8}\gamma_5(\gamma_\mu v\hspace{-.5em}/ -v_\mu)\gamma_5(\gamma_\nu v\hspace{-.5em}/-v_\nu) \gamma_5(\gamma^\mu v\hspace{-.5em}/ -v^\mu)\nonumber\\ &=&-\frac{1}{8}(n-3)\gamma_5(\gamma_\nu v\hspace{-.5em}/-v_\nu) =\frac{n-3}{4} S_\nu^v, \end{aligned}$$ where we consistently made use of the anticommutation relations of Eq. (\[5:4:gammarelations\]). In contrast, using the anticommutator and commutator of Eq.  (\[5:5:seig\]), one ends up with $S_\nu^v/4$ which only coincides with Eq. (\[5:5:sss\]) for $n=4$. However, the factor $(n-3)$ needs to be written as $1+(n-4)$ when it is multiplied with a singularity of the form $C/(n-4)$ in order to produce the constant non-divergent term $C$ in the product. Similarly, using the conventions of Eq. (\[5:4:gammarelations\]), the squared spin operator in $n$ dimensions reads $$\label{5:5:s2} S^2_v=\frac{1-n}{4}.$$ ### Normalization of Fields and States {#subsec_nfs} So far we have calculated matrix elements of the relativistic Lagrangian ${\cal L}^{(1)}_{\pi N}$ of Eq. (\[5:2:l1pin\]) in the framework of covariant perturbation theory based on the formula of Gell-Mann and Low [@Gell-Mann:1951rw] in combination with Wick’s theorem [@Wick:ee]. Let us recall that, for a generic field $\Phi(x)$ described by the Lagrangian $${\cal L}={\cal L}_0+{\cal L}_{\rm int},$$ the “magic formula of covariant perturbation theory” [@Haag:hx] allows one to calculate the Green functions $$\label{5:5:tau} \tau_n(x_1,\cdots,x_n)=\langle \Omega|T[\Phi(x_1)\cdots\Phi(x_n)]|\Omega\rangle$$ in terms of the generating functional $$\begin{aligned} \label{5:5:tgf} {\cal T}[f]&=&\frac{N[f]}{N[0]},\\ \label{5:5:nf} N[f]&=&\langle\Omega_0|T\exp\left\{i \int d^4 x \left[\Phi^0(x) f(x) + {\cal L}^0_{\rm int}(x)\right]\right\}|\Omega_0\rangle.\end{aligned}$$ While the Green functions of Eq. (\[5:5:tau\]) involve the interacting field $\Phi(x)$ and the vacuum $\Omega$ of the corresponding interacting theory, the formula of Gell-Mann and Low, in principle, provides an explicit expression for the generating functional in terms of the quantities $\Phi^0(x)$ and $\Omega_0$ defined in the free theory. Note that the Green functions of Eq. (\[5:5:tau\]) are expressed in terms of the bare fields and, in the end, still have to be renormalized (see, e.g, Sec. \[subsec\_mgb\]). Here we want to address the question of how to establish contact between matrix elements evaluated perturbatively using the relativistic Lagrangian of Eq. (\[5:2:l1pin\]), on the one hand, and the heavy-baryon Lagrangian of Eq. (\[5:5:lhneff\]) on the other hand. The presentation will make use of the ideas developed in Refs.  [@Dugan:1991ak; @Balk:1993ev], where this issue was discussed for the case of a heavy-quark effective theory. A different route was taken in Refs.  [@Ecker:1997dn; @Steininger:1998ya; @Kambor:1998pi], where the path-integral approach to the generating functional was used to define the wave function renormalization constant (including interaction). Later on we will explicitly see that the two approaches yield identical results at ${\cal O}(p^3)$. For later comparison with the heavy-baryon approach, we first need to collect a few properties of the free Dirac field operator $\Psi^0(x)$ which we decompose in the standard fashion in terms of the solutions of the free Dirac equation[^109] $$\label{5:5:psifeld} \Psi^0(x)=\sum_{\alpha=1}^2 \int \frac{d^3 p}{(2\pi)^3 2 E(\vec{p}\,)} \left[b_\alpha(\vec{p}\,) u^{(\alpha)}(\vec{p}\,)e^{-ip\cdot x} +d^\dagger_{\alpha}(\vec{p}\,) v^{(\alpha)}(\vec{p}\,) e^{ip\cdot x}\right],$$ where $p_0=E(\vec{p}\,)=\sqrt{m^2+\vec{p}\,^2}$. The Dirac spinors have the following properties $$\begin{aligned} \label{5:5:spinorprops} (p\hspace{-.5em}/-m)u^{(\alpha)}(\vec{p}\,)&=&0,\nonumber\\ (p\hspace{-.5em}/+m)v^{(\alpha)}(\vec{p}\,)&=&0,\nonumber\\ \bar{u}^{(\alpha)}(\vec{p}\,) u^{(\beta)}(\vec{p}\,)&=& -\bar{v}^{(\alpha)}(\vec{p}\,)v^{(\beta)}(\vec{p}\,) =2m\delta_{\alpha\beta},\nonumber\\ u^{(\alpha)\dagger}(\vec{p}\,)u^{(\beta)}(\vec{p}\,) &=&v^{(\alpha)\dagger}(\vec{p}\,)v^{(\beta)}(\vec{p}\,) =2 E(\vec{p}\,)\delta_{\alpha\beta},\nonumber\\ u^{(\alpha)\dagger}(\vec{p}\,)v^{(\beta)}(-\vec{p}\,)&=& \bar{u}^{(\alpha)}(\vec{p}\,)v^{(\beta)}(\vec{p}\,) =0.\end{aligned}$$ The creation operators $b_\alpha^\dagger$ and $d_\alpha^\dagger$ (annihilation operators $b_\alpha$ and $d_\alpha$) of particles and antiparticles, respectively, satisfy the anticommutation relations $$\begin{aligned} \{b_\alpha(\vec{p}\,),b_\beta^\dagger(\vec{p}\,')\}&=&(2\pi)^3 2 E(\vec{p}\,) \delta^3(\vec{p}-\vec{p}\,')\delta_{\alpha\beta},\\ {\{}d_\alpha(\vec{p}\,),d^\dagger_\beta(\vec{p}\,')\}&=&(2\pi)^3 2 E(\vec{p}\,) \delta^3(\vec{p}-\vec{p}\,')\delta_{\alpha\beta},\end{aligned}$$ where all remaining anticommutators such as, e.g., ${\{}b_\alpha(\vec{p}\,),b_\beta(\vec{p}\,')\}$ vanish. With this convention, single-particle states $|\vec{p},\alpha,+\rangle =b_\alpha^\dagger(\vec{p}\,)|0\rangle$ are normalized as $$\label{5:5:statenormalization} \langle \vec{p}\,',\beta,+|\vec{p},\alpha,+\rangle =(2\pi)^3 2 E(\vec{p}\,) \delta^3(\vec{p}\,'-\vec{p}\,)\delta_{\alpha\beta}.$$ Let us now turn to the heavy-baryon formulation and consider the leading-order term of Eq. (\[5:5:l1pinhb\]) which we write as $$\label{5:5:widehatl} \widehat{\cal L}^{(1)}_{\pi N}= \bar{\cal N}_v iv\cdot\partial {\cal N}_v +\widehat{\cal L}_{\rm int}\equiv \widehat{\cal L}_{0}+\widehat{\cal L}_{\rm int}$$ for later use in the formula of Gell-Mann and Low. We decompose the solution to the free equation of motion $$\label{5:5:loeom} iv\cdot\partial {\cal N}^0_v(x)=0,\quad v\hspace{-.5em}/\hspace{.2em}{\cal N}^0_v={\cal N}^0_v,$$ as $$\label{5:5:n0vdec} {\cal N}^0_v(x)=\sum_{\alpha=1}^2\int \frac{d^3 k}{(2\pi)^3 2m v_0} e^{-ik\cdot x}b_{\alpha,v}(\vec{k}) u^{(\alpha)}_v,$$ where $k_0$, at leading order, is defined through $v\cdot k=0$ in order to satisfy Eq. (\[5:5:loeom\]), i.e., $$k_0=\frac{\vec{v}\cdot\vec{k}}{v_0},\quad k\cdot x=-\vec{k}\cdot\left(\vec{x}-\vec{v}\,\frac{x_0}{v_0}\right).$$ The spinors are given by $$\label{5:5:ualphav0} u^{(\alpha)}_v=\sqrt{\frac{2m}{v_0}}P_{v+}\left( \begin{array}{c} \chi^{(\alpha)}\\ 0_{2\times 1} \end{array} \right),\quad \bar{u}^{(\alpha)}_v u^{(\beta)}_v=2m\delta_{\alpha\beta},$$ with $\chi^{(\alpha)}$ ordinary two-component Pauli spinors. Note that, at lowest order in $1/m$, the spinors do not depend on the residual momentum $\vec{k}$. Moreover, for an arbitrary choice of $v$, the $u^{(\alpha)}_v$ are four-component objects which only for the special case $v^\mu=(1,0,0,0)$ effectively reduce to two-component spinors. As will be shown below, the operators $b_{\alpha,v}(\vec{k})$ and $b_{\alpha,v}^\dagger(\vec{k})$ destroy and create a nucleon (isospin index suppressed) with residual three-momentum $\vec{k}$. They satisfy the anticommutation relations $$\label{5:5:banticomm} \{b_{\alpha,v}(\vec{k}), b_{\beta,v}^\dagger(\vec{k}')\}= 2m v_0 (2\pi)^3\delta^3(\vec{k}-\vec{k}\,')\delta_{\alpha\beta},$$ where, as usual, the anticommutator of two annihilation or two creation operators vanishes. Accordingly, the single-particle states are normalized as $$\label{5:5:statenormalizationhb} \langle v,\vec{k}\,',\beta|v,\vec{k},\alpha\rangle =2 m v_0 (2\pi)^3 \delta^3(\vec{k}\,'-\vec{k}\,)\delta_{\alpha\beta}.$$ Note that the normalization of the states of Eqs. (\[5:5:statenormalization\]) and (\[5:5:statenormalizationhb\]) coincide only at leading order in $1/m$ (or as $m\to \infty$). Using Eq. (\[5:5:banticomm\]) it is straightforward to verify that the (free) theory has been quantized “canonically” [@Dugan:1991ak], i.e. $$\{\Pi_v^0(\vec{x},t),{\cal N}_v^0(\vec{y},t)\} =iv_0\{\bar{\cal N}_v^0(\vec{x},t),{\cal N}_v^0(\vec{y},t)\} =i\delta^3(\vec{x}-\vec{y})P_{v+},$$ where $\Pi_v^0=\partial\widehat{\cal L}_0/\partial(\partial_0 {\cal N}_v) =iv_0\bar{\cal N}^0_v$ is the momentum conjugate to ${\cal N}^0_v$ and we made use of the completeness relation $$\label{5:5:completeness} \sum_{\alpha=1}^2 u^{(\alpha)}_v\bar{u}^{(\alpha)}_v=2m P_{v+}.$$ Constructing the energy-momentum tensor corresponding to $\widehat{\cal L}_{0}$, $$\Theta^{\mu\nu}_v=\partial^\nu\bar{\cal N}_v \frac{\partial\widehat{\cal L}_0}{\partial(\partial_\mu\bar{\cal N}_v)} +\frac{\partial\widehat{\cal L}_0}{\partial(\partial_\mu{\cal N}_v)} \partial^\nu{\cal N}_v-g^{\mu\nu}\widehat{\cal L}_0= \bar{\cal N}^0_v iv^\mu\partial^\nu {\cal N}_v,$$ (we made use of the equation of motion) we obtain for the four-momentum operator $$\label{5:5:kmuv} K^\mu_v=\int d^3 x \bar{\cal N}^0_v iv_0\partial^\mu {\cal N}_v =\sum_{\alpha=1}^2\int \frac{d^3 k}{(2\pi)^3 2m v_0} k^\mu b_{\alpha,v}^\dagger(\vec{k}) b_{\alpha,v}(\vec{k}).$$ Using $[ab,c]=a\{b,c\}-\{a,c\}b$ it is then straightforward to verify the commutation relations $$\label{5:5:kmucomrel} [K^\mu_v,b_{\alpha,v}(\vec{k})]=-k^\mu b_{\alpha,v}(\vec{k}),\quad [K^\mu_v,b_{\alpha,v}^\dagger(\vec{k})]=k^\mu b_{\alpha,v}^\dagger(\vec{k}).$$ Eq. (\[5:5:kmucomrel\]) implies that $b_{\alpha,v}(\vec{k})$ and $b_{\alpha,v}^\dagger(\vec{k})$ destroy and create quanta with (residual) four-momentum $k^\mu$ and total four-momentum $p^\mu=mv^\mu+k^\mu$. This can be seen by comparing with the four-momentum operator of the free relativistic theory $$P^\mu=\int d^3 x \bar{\Psi}^0\gamma^0 i\partial^\mu {\Psi}^0,$$ which, to leading order in $1/m$, is related to Eq. (\[5:5:kmuv\]) by $P^\mu=K^\mu_v+m v^\mu N$, where $N=\int d^3x \Psi^{0\dagger}\Psi^0$ is the number operator [@Dugan:1991ak]. Using the orthogonality relations of Eq. (\[5:5:ualphav0\]) we may express the creation and annihilation operators in terms of the fields in the standard way $$\begin{aligned} \label{5:5:bbd} b_{\alpha,v}^\dagger(\vec{k})&=&v_0\int d^3 x\bar{\cal N}_v(x)u_v^{(\alpha)} e^{-ik\cdot x},\nonumber\\ b_{\alpha,v}(\vec{k})&=&v_0\int d^3 x e^{ik\cdot x}\bar{u}^{(\alpha)}_v {\cal N}_v(x).\end{aligned}$$ Eqs. (\[5:5:bbd\]) are the starting point for the LSZ reduction [@Lehmann:1955rq; @Bjorken_1964qf; @Itzykson:rh] in the framework of the heavy-baryon approach. We consider the matrix element of Eq. (\[4:btbta\]) for the transition in the presence of external fields $v$, $a$, $s$, and $p$ (we omit spin and isospin labels) $$\begin{aligned} \label{5:5:ffunc} \lefteqn{{\cal F}(\vec{p}\,',\vec{p};v,a,s,p)=\langle {\vec{p}\,'};{\rm out}| {\vec{p}\,};{\rm in}\rangle^{\rm c}_{v,a,s,p}}\nonumber\\ &=&\sqrt{\frac{E}{mv_0}} \langle {\vec{p}\,'};{\rm out}|b_{v,\rm in}^\dagger(\vec{k})|\Omega \rangle^{\rm c}_{v,a,s,p}\nonumber\\ &=&\sqrt{\frac{E}{mv_0}} v_0\int d^3 x\langle {\vec{p}\,'};{\rm out}|\bar{\cal N}_{v,\rm in}(x)| \Omega\rangle^{\rm c}_{v,a,s,p}u_v e^{-ik\cdot x}\nonumber\\ &=&\sqrt{\frac{E}{mv_0}} \lim_{t\to-\infty}v_0\int d^3 x\langle {\vec{p}\,'};{\rm out}| \frac{\bar{\cal N}_v(x)}{\sqrt{Z}}| \Omega\rangle^{\rm c}_{v,a,s,p}u_v e^{-ik\cdot x}\nonumber\\ &=&\cdots\nonumber\\ &=&\left(\frac{-i}{\sqrt{Z}}\right)^2 N N' \int d^4x d^4 y \nonumber\\ &&\times e^{ik'\cdot y}\bar{u}_v iv\cdot \stackrel{\rightarrow}{ \partial_y}\langle\Omega|T[{\cal N}_v(y)\bar{\cal N}_v(x)]|\Omega\rangle^{ \rm c}_{v,a,s,p}(-iv\cdot\stackrel{\leftarrow}{\partial_x}) u_v e^{-ik\cdot x}.\nonumber\\\end{aligned}$$ The intermediate steps indicated by $\cdots$ proceed in complete analogy to the usual reduction formula as described in, e.g., Refs.  [@Bjorken_1964qf; @Itzykson:rh]. In Eq. (\[5:5:ffunc\]), the factors of the type $N=\sqrt{E/(mv_0)}$ are related to the relative normalization of the states \[see Eq.  (\[5:5:statenormalization\]) vs. (\[5:5:statenormalizationhb\])\], whereas $\sqrt{Z}$ refers to the wave function renormalization in the framework of the heavy-baryon Lagrangian. The Green function entering Eq. (\[5:5:ffunc\]) will be calculated perturbatively using the formula of Gell-Mann and Low [@Gell-Mann:1951rw], $$\label{5:5:gmlhb} \langle\Omega|T[{\cal N}_v(y)\bar{\cal N}_v(x)]|\Omega\rangle^{ \rm c}_{v,a,s,p} =\langle\Omega_0|T[{\cal N}_v^0(y)\bar{\cal N}_v^0(x) \exp\left(i\int d^4 z \widehat{\cal L}\,^0_{\rm int}(z)\right) ]|\Omega_0\rangle^{ \rm c},$$ where, on the right-hand side, $|\Omega_0\rangle$ denotes the vacuum of the free theory, and the external fields are part of the Lagrangian $\widehat{\cal L}\,^0_{\rm int}(z)$.[^110] ### Propagator at Lowest Order {#subsec_plo} We will now discuss the propagator of the lowest-order Lagrangian both on the “classical level” as well as in the quantized theory of the last section. The lowest-order equation of motion corresponding to Eq.  (\[5:5:l1pinhb\]) reads $$\label{5:5:l1pinhbeom} (iv\cdot D+g_A S_v\cdot u){\cal N}_v=0, \quad P_{v+} {\cal N}_v={\cal N}_v,$$ where the second relation implies $P_{v-}{\cal N}_v=0$ \[see Eq. (\[5:5:nhprop\])\]. We define the propagator corresponding to Eq. (\[5:5:l1pinhbeom\]) through $$\label{5:5:prop} (iv\cdot D+g_A S_v\cdot u)G_v(x,x')=P_{v+}\delta^4(x-x'),\quad P_{v-}G_v(x,x')=0.$$ In order to solve Eq. (\[5:5:l1pinhbeom\]) perturbatively, we re-write the equation of motion in the standard form as $$iv\cdot \partial {\cal N}_v(x)=V(x){\cal N}_v(x),$$ where $V$ denotes the interaction term, and search for the unperturbed Green function $G_v^0(x,x')$ satisfying the properties $$\begin{aligned} \label{5:5:greenprop1} i v\cdot \partial G_v^0(x,x')&=&\delta^4(x-x')P_{v+},\\ \label{5:5:greenprop2} P_{v-} G_v^0(x,x')&=&0,\\ \label{5:5:greenprop3} G_v^0(x,x')&=&0\,\,\,\mbox{for}\,\,\,x_0'>x_0.\end{aligned}$$ In terms of $G^0_v$ the propagator $G_v$ is then given by $$G_v(x,x')=G_v^0(x,x')+\int d^4 y G^0_v(x,y)V(y) G_v(y,x').$$ Inserting the standard ansatz in terms of a Fourier decomposition $$\label{5:5:g0f} G_v^0(x,x')=\int \frac{d^4 k}{(2\pi)^4}e^{-ik\cdot(x-x')}G_v^0(k)$$ into Eq. (\[5:5:greenprop1\]), $$\int \frac{d^4 k}{(2\pi)^4}e^{-ik\cdot(x-x')} v\cdot k G_v^0(k) =\delta^4(x-x')P_{v+}= \int \frac{d^4 k}{(2\pi)^4}e^{-ik\cdot(x-x')}P_{v+},$$ we obtain by comparing both sides $$G_v^0(k)=\frac{P_{v+}}{v\cdot k}\,\, \mbox{for}\,\, v\cdot k\neq 0.$$ The boundary condition of Eq. (\[5:5:greenprop3\]) may be incorporated by introducing an infinitesimally small imaginary part into the denominator: $$\label{5:5:gv0k} G_{v}^0(k)=\frac{P_{v+}}{v\cdot k +i0^+}.$$ That this is indeed the correct choice is easily seen by evaluating the integral $$\int_{-\infty}^\infty\frac{dk_0}{2\pi}e^{-ik_0(x_0-x'_0)}\frac{1}{ k_0-\frac{\vec{v}\cdot\vec{k}}{v_0}+i0^+} =-i\Theta(x_0-x_0') \exp\left[-i\frac{(x_0-x_0')\vec{v}\cdot\vec{k}}{v_0}\right]$$ as a contour integral in the complex $k_0$ plane (see Fig. \[5:5:fig:integration\]). For $x_0>x_0'$ the contour is closed in the lower half plane and one makes use of the residue theorem. On the other hand, for $x_0<x_0'$ the contour is closed in the upper half plane and, since the contour does not contain a pole, the integral vanishes. We then obtain $$\begin{aligned} \label{5:5:gv0f} G_v^0(x,x')&=&-i\frac{\Theta(x_0-x_0')}{v_0} \int \frac{d^3 k}{(2\pi)^3} \exp\left[i\vec{k}\cdot\left(\vec{x}-\vec{x}\,'-\vec{v}\,\frac{x_0-x_0'}{v_0} \right)\right] P_{v+}\nonumber\\ &=&-i\frac{\Theta(x_0-x_0')}{v_0}\delta^3\left(\vec{x}-\vec{x}\,'-\vec{v}\, \frac{x_0-x_0'}{v_0}\right)P_{v+}.\end{aligned}$$ For the special choice $v^\mu=(1,0,0,0)\equiv{\tilde v}^\mu$ the propagator reduces to that of a static source $$G_{\tilde{v}}^0(x,x')=-i\Theta(x_0-x_0') \delta^3(\vec{x}-\vec{x}\,') \left( \begin{array}{cc} 1_{2\times 2}&0_{2\times 2}\\ 0_{2\times 2}&0_{2\times 2} \end{array} \right).$$ (10,10) (0,5)[(1,0)[10]{}]{} (10,4)[$Re(k_0)$]{} (0.5,4.9)[(1,0)[9]{}]{} (0.5,5.1)[(1,0)[9]{}]{} (5.5,4.9)[(1,0)[1]{}]{} (7.5,5.1)[(1,0)[1]{}]{} (5,0)[(0,1)[10]{}]{} (3,10)[$Im(k_0)$]{} (7,4)[$\times$]{} (7,3) (5,5.1)[(9,8.9)\[tl\]]{} (5,5.1)[(9,8.9)\[tr\]]{} (8,8)[$x_0<x_0'$]{} (5,4.9)[(9,8.9)\[br\]]{} (5,4.9)[(9,8.9)\[bl\]]{} (1,1)[$x_0>x_0'$]{} (7,9.55)[(-1,0)[0.1]{}]{} (7,0.45)[(-1,0)[0.1]{}]{} Finally, it is easy to show that a definition of the propagator in terms of the field operators ${\cal N}^0_v$ and $\bar{\cal N}^0_v$ [@Dugan:1991ak], $$G^0_v(x,x')=-i\Theta(x_0-x_0')\langle\Omega_0|{\cal N}_v^0(x) \bar{\cal N}^0_v(x')|\Omega_0\rangle,$$ yields the same result as Eq. (\[5:5:gv0f\]). To that end, one inserts for each of the two operators a sum according to Eq. (\[5:5:n0vdec\]), commutes the creation and annihilation operators using Eq. (\[5:5:banticomm\]), applies the completeness relation of Eq. (\[5:5:completeness\]), and makes use of $v\cdot k=0$ for the individual Fourier components. Performing the remaining integration over $\vec{k}$ one ends up with Eq. (\[5:5:gv0f\]), i.e., as expected the two methods yield the same result. ### Example: $\pi N$ Scattering at Lowest Order As a simple example, let us return to $\pi N$ scattering, but now in the framework of the heavy-baryon Lagrangian of Eq. (\[5:5:l1pinhb\]). The four-momenta of the initial and final nucleons are written as $p=\,\stackrel{\circ}{m}_N v+k$ and $p'=\,\stackrel{\circ}{m}_N v+k'$, respectively, with $v\cdot k=0=v\cdot k'$ to leading order in $1/\stackrel{\circ}{m}_N$. The relevant interaction Lagrangian is obtained in complete analogy to Eq. (\[5:3:lpin\]), $$\label{5:5:lpinhb} \widehat{\cal L}^{(1)}_{\rm int}=-\frac{\stackrel{\circ}{g}_A}{F_0} \bar{\cal N}_v S^\mu_v \vec{\tau}\cdot\partial_\mu\vec{\phi}{\cal N}_v -\frac{1}{4 F_0^2} v^\mu \bar{\cal N}_v\vec{\tau}\cdot\vec{\phi}\times \partial_\mu\vec{\phi}{\cal N}_v,$$ and the corresponding Feynman rules for the vertices derived from Eq. (\[5:5:lpinhb\]) read - for a single incoming pion with four-momentum $q$ and Cartesian isospin index $a$: $$\label{5:5fr1} -\frac{\stackrel{\circ}{g}_A}{F_0} S_v\cdot q \tau^a,$$ - for an incoming pion with $q,a$ and an outgoing pion with $q',b$: $$\label{5:6:fr2} \frac{v\cdot(q+q')}{4 F^2_0}\epsilon_{abc}\tau^c.$$ As in the case of the relativistic calculation of Sec.  \[subsec\_apnstl\] the latter gives rise to a contact contribution to ${\cal M}^v$ $$\begin{aligned} \label{5:5:mconthb} {\cal M}_{\rm cont}^v&=& N' N \bar{u}_v'\frac{v\cdot(q+q')}{4 F^2_0}\epsilon_{abc}\tau^c u_v,\end{aligned}$$ where the spinors are given in Eq. (\[5:5:ualphav0\]) and $N$ and $N'$ are the normalization factors appearing in the reduction formula of Eq. (\[5:5:ffunc\]). The result for the direct-channel nucleon pole term reads $$\begin{aligned} \label{5:5:mdcconthb} {\cal M}_{\rm d}^v=-i\frac{\stackrel{\circ}{g}_A^2}{F_0^2} N'N \tau^b\tau^a \bar{u}_v'S_v\cdot q' \frac{P_{v+}}{v\cdot(k+q)} S_v\cdot{q} u_v,\end{aligned}$$ where, at leading order, we can make use of $v\cdot k=0$. The crossed channel is obtained from Eq. (\[5:5:mdcconthb\]) by the replacement $a\leftrightarrow b$ and $q\leftrightarrow -q'$ (pion crossing). The evaluation of the total matrix element ${\cal M}^v= {\cal M}^v_{\rm cont}+{\cal M}^v_{\rm d}+{\cal M}^v_{\rm c}$ is particularly simple for the special choice $v^\mu=(1,0,0,0)\equiv\tilde{v}^\mu$, for which we have $$P_{\tilde{v}+}=\left(\begin{array}{cc}1_{2\times 2}&0_{2\times 2}\\ 0_{2\times 2}&0_{2\times 2}\end{array}\right).$$ In that case, the calculation effectively reduces to that of a two-component theory as in the Foldy-Wouthuysen transformation, because the $4\times 4$ matrices of the vertices are multiplied both from the left and the right by $P_{\tilde{v}+}$ originating from either the propagator of Eq. (\[5:5:gv0k\]) or the spinors of Eq. (\[5:5:ualphav0\]). To be specific, for a $4\times 4$ matrix $\Gamma$ of the type $$\Gamma= \left(\begin{array}{cc} A&B\\ C&D \end{array} \right),$$ where each block $A$, $B$, $C$, and $D$ is a $2\times 2$ matrix, one has $$P_{\tilde{v}+}\Gamma P_{\tilde{v}+}= \left(\begin{array}{cc} A&0_{2\times 2}\\ 0_{2\times 2}&0_{2\times 2} \end{array} \right)$$ and $$P_{\tilde{v}+}\Gamma_1P_{\tilde{v}+}\Gamma_2P_{\tilde{v}+}= \left(\begin{array}{cc} A_1 A_2&0_{2\times 2}\\ 0_{2\times 2}&0_{2\times 2} \end{array} \right).$$ Moreover, the spin matrix of Eq. (\[5:5:spinop\]) is very simple for $\tilde{v}$, $$\label{5:5:spinopvtilde} S^0_{\tilde{v}}=0,\quad \vec{S}_{\tilde{v}}=\frac{1}{2}\vec{\Sigma},$$ where $\vec{\Sigma}$ has been defined in Eq. (\[5:5:Sigma\]). With this special choice of $v$ the $T$ matrix in the center-of-mass frame reads $$\label{5:5:tmatrix} T=2\stackrel{\circ}{m}_N \chi'^\dagger\left[-i\epsilon_{abc}\tau^c\left(\frac{E_\pi}{2F_0^2} -\frac{\stackrel{\circ}{g}_A^2}{F_0^2}\frac{\vec{q}\cdot\vec{q}\,'}{2E_\pi} \right) +\delta^{ab}\frac{\stackrel{\circ}{g}_A^2}{F_0^2} \left(-\frac{i\vec{\sigma}\cdot\vec{q}\,'\times\vec{q}}{2E_\pi}\right)\right] \chi.$$ Performing a nonrelativistic reduction of Eq. (\[5:3:mpinpar\]) in the center-of-mass frame, $$\label{5:5:tnonrelred} T=2\stackrel{\circ}{m}_N \chi'^\dagger\left[A+ \left(E_\pi +\frac{\vec{q}\,^2+\vec{q}\,'\cdot\vec{q}}{2\stackrel{\circ}{m}_N}\right)B +i\frac{\vec{\sigma}\cdot \vec{q}\,'\times\vec{q}}{2 \stackrel{\circ}{m}_N} B +\cdots\right] \chi,$$ and using $$\begin{aligned} \frac{1}{\nu-\nu_B}+\frac{1}{\nu+\nu_B}&=&\frac{1}{E_\pi}\left[2- \frac{\vec{q}\,^2+\vec{q}\cdot\vec{q}\,'}{E_\pi \stackrel{\circ}{m}_N }+O\left(\frac{1}{\stackrel{\circ}{m}_N^2}\right) \right],\\ \frac{1}{\nu-\nu_B}-\frac{1}{\nu+\nu_B}&=&\frac{1}{E_\pi}\left[ -\frac{E_\pi}{\stackrel{\circ}{m}_N}+ \frac{\vec{q}\cdot\vec{q}\,'}{E_\pi\stackrel{\circ}{m}_N} +O\left(\frac{1}{\stackrel{\circ}{m}_N^2}\right)\right]\end{aligned}$$ in the expansion of $A$ and $B$ of Table \[5:3:tableresults\], one verifies that, at leading order in $1/\!\!\stackrel{\circ}{m}_N$, the relativistic Lagrangian of Eq. (\[5:2:l1pin\]) and the heavy-baryon Lagrangian of Eq.  (\[5:5:l1pinhb\]) indeed generate the same $\pi N$ scattering amplitude. We emphasize that in order to obtain this equivalence of the two approaches an expansion of Eq. (\[5:5:tnonrelred\]) to $1/\!\!\stackrel{\circ}{m}_N$ is mandatory, because the functions $A^{(+)}$ and $B^{(+)}$ contain terms of leading order $\stackrel{\circ}{m}_N$. These terms disappear through a cancellation in the final result.[^111] ### Corrections at First Order in $1/m$ {#subsec_cfo1om} So far we have concentrated on the leading-order, $m$-independent, heavy-baryon Lagrangian of Eq. (\[5:5:l1pinhb\]). In comparison with Eq. (\[5:2:powercounting\]), the chiral counting scheme of HBChPT is different, because an ordinary partial derivative acting on a heavy-baryon field ${\cal N}_v$ produces a small residual [*four*]{}-momentum \[see also Eq. (\[4:5:powercounting\]) for the mesonic sector\]: $$\label{5:5:hbchppowercounting} {\cal N}_v,\bar{\cal N}_v = {\cal O}(p^0),\, D_{\mu} {\cal N}_v = {\cal O}(p),\, v_\mu, S^v_\mu, 1_{4\times 4}={\cal O}(p^0).$$ In the heavy-baryon approach four-momenta are considered small if their components are small in comparison with either the nucleon mass $m_N$ or the chiral symmetry breaking scale $4\pi F_\pi$, both of which we denote by a common scale $\Lambda\simeq 1$ GeV.[^112] It is clear that the Lagrangian of Eq. (\[5:5:lhneffexp\]) also generates terms of higher order in $1/m$ and, in analogy to the mesonic sector, we also expect additional new chiral structures from the most general chiral Lagrangian at higher orders. Recall that in the baryonic sector the chiral orders increase in units of one, because of the additional possibility of forming Lorentz invariants by contracting (covariant) derivatives with gamma matrices (see Sec. \[sec\_loebl\]). (The relativistic $\pi N$ Lagrangian at ${\cal O}(p^2)$ has (partially) been given in Ref. [@Gasser:1987rb].) Let us first consider the $1/m$ correction resulting from Eq. (\[5:5:lhneffexp\]) $$\begin{aligned} &&\frac{1}{2m}\bar{\cal N}_v \left(iD\hspace{-.7em}/_\bot+\frac{g_A}{2}v\cdot u \gamma_5\right) \left(iD\hspace{-.7em}/_\bot-\frac{g_A}{2}v\cdot u \gamma_5\right){\cal N}_v \nonumber\\ &=& \frac{1}{2m}\bar{\cal N}_v\Bigg[-D\hspace{-.7em}/_\bot D\hspace{-.7em}/_\bot -i\frac{g_A}{2}D\hspace{-.7em}/_\bot v\cdot u \gamma_5 +i\frac{g_A}{2}v\cdot u \gamma_5 D\hspace{-.7em}/_\bot -\frac{g^2_A}{4}(v\cdot u)^2\Bigg]{\cal N}_v.\end{aligned}$$ We make use of Eqs. (\[5:5:nvbilineare\]) to identify the relevant replacements in the heavy-baryon bilinears: $$\begin{aligned} D\hspace{-.7em}/_\bot\gamma_5&=& D\hspace{-.7em}/\hspace{.2em}\gamma_5-v\cdot D v\hspace{-.5em}/\hspace{.2em} \gamma_5 \mapsto 2 D\cdot S_v - 2v\cdot D \underbrace{v\cdot S_v}_{\mbox{0}} = 2 D\cdot S_v,\\ \gamma_5 D\hspace{-.7em}/_\bot&\mapsto&-2 D\cdot S_v,\\ D\hspace{-.7em}/_\bot D\hspace{-.7em}/_\bot&=& (D_\mu-v\cdot D v_\mu)(D_\nu-v\cdot D v_\nu)\hspace{-1em} \underbrace{\gamma^\mu\gamma^\nu}_{\mbox{$g^{\mu\nu}-i\sigma^{\mu\nu}$}}\\ &=&(D^2-v\cdot D v\cdot D)-i\hspace{-1.8em}\underbrace{\sigma^{\mu\nu}}_{ \mbox{$\mapsto 2 \epsilon^{\mu\nu\rho\sigma}v_\rho S_\sigma^v$}} \hspace{-1.8em} (D_\mu-v\cdot D v_\mu)(D_\nu-v\cdot D v_\nu)\\ &\mapsto&D^2-(v\cdot D)^2-i\epsilon^{\mu\nu\rho\sigma}[D_\mu,D_\nu]v_\rho S_\sigma^v\\ &=& D^2-(v\cdot D)^2 -\frac{i}{4}\epsilon^{\mu\nu\rho\sigma} [u_\mu,u_\nu]v_\rho S_\sigma^v\nonumber\\ && -\frac{1}{2}\epsilon^{\mu\nu\rho\sigma} f^+_{\mu\nu}v_\rho S_\sigma^v -\epsilon^{\mu\nu\rho\sigma}v^{(s)}_{\mu\nu} v_\rho S_\sigma^v,\end{aligned}$$ where the expression for the commutator $[D_\mu,D_\nu]$ of the covariant derivative of Eq. (\[5:2:kovderpsi\]) is obtained after straightforward algebra. The field-strength tensors are defined as $$\label{fpm} f^{\pm}_{\mu\nu} = u f_{\mu\nu}^L u^\dagger \pm u^\dagger f_{\mu\nu}^R u , \quad v^{(s)}_{\mu\nu}=\partial_\mu v^{(s)}_{\nu}-\partial_\nu v^{(s)}_{\mu},$$ where $f_{\mu\nu}^R$ and $f_{\mu\nu}^L$ are given in Eqs.  (\[4:5:fr\]) and (\[4:5:fl\]), respectively. Collecting all terms, we finally obtain as the contribution of Eq.  (\[5:5:lhneffexp\]) of order $1/m$ (returning to the notation in terms of expressions in the chiral limit) $$\begin{aligned} \label{5:5:l2pinhba} \widehat{\cal L}^{(2)}_{\pi N,1/m} &=&\frac{1}{2\stackrel{\circ}{m}_N} \bar{\cal N}_v\Bigg[(v\cdot D)^2-D^2-i\stackrel{\circ}{g}_A \{S_v\cdot D, v\cdot u\}-\frac{\stackrel{\circ}{g}_A^2}{4}(v\cdot u)^2\nonumber\\ &&+\frac{1}{2}\epsilon^{\mu\nu\rho\sigma} v_\rho S_\sigma^v \left(i u_\mu u_\nu+ f^+_{\mu\nu}+2 v^{(s)}_{\mu\nu}\right) \Bigg]{\cal N}_v.\end{aligned}$$ Applying the counting rules of Eqs. (\[4:5:powercounting\]) and (\[5:5:hbchppowercounting\]), we see that Eq. (\[5:5:l2pinhba\]) is indeed of ${\cal O}(p^2)$, where the suppression relative to (\[5:5:l1pinhb\]) is of the form $p/\stackrel{\circ}{m}_N$. At ${\cal O}(p^2)$ the heavy-baryon Lagrangian $\widehat{\cal L}^{(2)}_{\pi N}$ contains another contribution which, in analogy to $\widehat{\cal L}^{(1)}_{\pi N}$ in Sec. \[subsec\_lol\], may be obtained as the projection of the relativistic Lagrangian ${\cal L}^{(2)}_{\pi N}$ of [@Gasser:1987rb] onto the light components. Here we quote the result in the convention of Ref.  [@Bernard:1997gq](except for the $c_6$ and $c_7$ terms, where, following Ref. [@Ecker:1995rk], we explicitly separate the traceless and isoscalar terms)[^113] $$\begin{aligned} \label{5:5:l2pinhbb} \widehat{\cal L}^{(2)}_{\pi N, c_i}&=& \bar{\cal N}_v\Bigg[c_1 \mbox{Tr}(\chi_+) +c_2 (v\cdot u)^2 +c_3 u\cdot u +c_4 [S^\mu_v,S^\nu_v]u_\mu u_\nu\nonumber\\ && +c_5\left[\chi_+-\frac{1}{2}\mbox{Tr}(\chi_+)\right] -ic_6[S^\mu_v,S^\nu_v]f^+_{\mu\nu} -ic_7[S^\mu_v,S^\nu_v]v^{(s)}_{\mu\nu} \Bigg]{\cal N}_v,\nonumber\\\end{aligned}$$ where $$\label{5:5:chipm} \chi_{\pm} = u^\dagger \chi u^\dagger \pm u\chi^\dagger u.$$ In the parameterization of Eq. (\[5:5:l2pinhbb\]) the constants $c_i$ carry the dimension of an inverse mass and should be of the order of $1/\Lambda$ in order to produce a reasonable convergence of the chiral expansion. (The details of convergence generally depend on the observable in question). The complete heavy-baryon Lagrangian at ${\cal O}(p^2)$ is then given by the sum of Eqs. (\[5:5:l2pinhba\]) and (\[5:5:l2pinhbb\]), $$\label{5:5:l2pinhb} \widehat{\cal L}^{(2)}_{\pi N}= \widehat{\cal L}^{(2)}_{\pi N,1/m}+ \widehat{\cal L}^{(2)}_{\pi N, c_i}.$$ It is worthwhile mentioning that the contribution of Eq. (\[5:5:l2pinhba\]) to $\widehat{\cal L}^{(2)}_{\pi N}$ contains chirally invariant structures that are [*not*]{} part of Eq. (\[5:5:l2pinhbb\]). Unless such terms can be transformed away by a field transformation (see below) their coefficients are fixed in terms of the parameters of the lowest-order Lagrangian. As stressed by Ecker [@Ecker:1995gg], these fixed coefficients are a consequence of the Lorentz covariance of the whole approach. A related issue is the so-called reparameterization invariance, i.e., if a heavy particle of physical four-momentum $p$ is described by, say, $p=mv+k$ with $p^2=m^2$ and $v^2=1$ implying $2mv\cdot k+k^2=0$, physical observables should not change under the replacement $(v,k)\to (v+q/m,k-q)$ giving rise to an equivalent parameterization $p=mv'+k'$, if $q$ satisfies $(v+q/m)^2=1$ [@Luke:1992cs]. As a result, some coefficients of terms in the effective Lagrangian which are of different order in the $1/m$ expansion are related. For a detailed discussion, the reader is referred to Refs.  [@Luke:1992cs; @Chen:te; @Finkemeier:1997re]. The seven low-energy constants $c_i$ are determined by comparison with experimental information. For example, if we consider the interaction with an external electromagnetic field \[see Eq. (\[2:4:rlasu2\])\], $$r_\mu=l_\mu=-e\frac{\tau_3}{2}{\cal A}_\mu,\quad v_\mu^{(s)}=-\frac{e}{2}{\cal A}_\mu,$$ we obtain $$f_{\mu\nu}^+=-e\tau_3 {\cal F}_{\mu\nu}+\cdots,\quad v_{\mu\nu}^{(s)}=-\frac{e}{2}{\cal F}_{\mu\nu},\quad {\cal F}_{\mu\nu}=\partial_\mu{\cal A}_\nu-\partial_\nu{\cal A}_\mu,$$ so that the interaction with the field-strength tensor is given by $$\label{5:5:l2int} \widehat{\cal L}^{(2)}_{\rm int}= -e\epsilon_{\mu\nu\rho\sigma} {\cal F}^{\mu\nu} v^{\rho} \bar{\cal N}_v S^\sigma_v\left[\frac{1}{4\stackrel{\circ}{m}_N}+\frac{c_7}{2} +\tau_3\left(\frac{1}{4\stackrel{\circ}{m}_N}+c_6\right)\right]{\cal N}_v.$$ \[We made use of Eq. (\[5:5:seig\]).\] For the special choice $v^\mu=(1,0,0,0)=\tilde{v}^\mu$ we find \[see Eq. (\[5:5:spinopvtilde\])\] $$\epsilon_{\mu\nu\rho\sigma}{\cal F}^{\mu\nu} \tilde{v}^{\rho}S^\sigma_{\tilde{v}} =\epsilon_{ijk}{\cal F}^{ij} \frac{1}{2}\Sigma^k =-\vec{\Sigma}\cdot\vec{B},$$ and the interaction term reduces to[^114] $$\label{5:5:lintmm} \frac{e}{2\stackrel{\circ}{m}_N} \bar{\cal N}_{\tilde{v}}\, \vec{\sigma}\cdot\vec{B}\, {\cal N}_{\tilde{v}}\left[\frac{1}{2} \left(1+2\stackrel{\circ}{m}_N c_7\right) +\frac{\tau_3}{2}\left(1+4\stackrel{\circ}{m}_N c_6\right)\right],$$ which describes the interaction Lagrangian of a magnetic field with the magnetic moment of the nucleon. We define the isospin decomposition of the magnetic moment (in units of the nuclear magneton $e/2 m_p$) as $$\mu=\frac{1}{2}\mu^{(s)}+\frac{\tau_3}{2}\mu^{(v)} =\frac{1}{2}(1+\kappa^{(s)})+\frac{\tau_3}{2}(1+\kappa^{(v)}),$$ where $\kappa^{(s)}$ and $\kappa^{(v)}$ denote the isoscalar and isovector [*anomalous*]{} magnetic moments of the nucleon, respectively, with empirical values $\kappa^{(s)}=-0.120$ and $\kappa^{(v)}=3.706$. A comparison with Eq. (\[5:5:lintmm\]) shows that the constants $c_6$ and $c_7$ are related to the [*anomalous*]{} magnetic moments of the nucleon in the chiral limit $$\stackrel{\circ}{\kappa}^{(s)}=2\stackrel{\circ}{m}_N c_7,\quad \stackrel{\circ}{\kappa}^{(v)}=4\stackrel{\circ}{m}_N c_6.$$ The results for $\kappa^{(s)}$ and $\kappa^{(v)}$ up to and including ${\cal O}(p^3)$ [@Bernard:1995dp; @Fearing:1997dp] $$\begin{aligned} \kappa^{(s)}&=&\stackrel{\circ}{\kappa}^{(s)}+\,\,{\cal O}(p^4),\\ \kappa^{(v)}&=&\stackrel{\circ}{\kappa}^{(v)} -\frac{M_\pi m_N g_A^2}{4\pi F_\pi^2}+{\cal O}(p^4),\end{aligned}$$ are used to express the parameters $c_6$ and $c_7$ in terms of physical quantities. Note that the numerical correction of $-1.96$ \[parameters of Eq. (\[5:3:par\])\] to the isovector anomalous magnetic moment is substantial. Differences by factors of about 1.5 were generally observed for the determination of the $c_i$ at ${\cal O}(p^2)$ and to one-loop accuracy ${\cal O}(p^3)$ [@Bernard:1995dp; @Bernard:1997gq]. The numerical values of the low-energy constants $c_1,\cdots,c_4$ have been determined in Ref. [@Bernard:1997gq] by performing a best fit to a set of nine pion-nucleon scattering observables at ${\cal O}(p^3)$ which do not contain any new low-energy constants from the ${\cal O}(p^3)$ Lagrangian. Finally, $c_5$ was determined in terms of the strong contribution to the neutron-proton mass difference. The results in units of GeV$^{-1}$ are given by (see also Ref.[@Buttiker:1999ap]) $$\begin{aligned} \label{5:5:ciresults} &&c_1=-0.93\pm 0.10,\quad c_2=3.34\pm 0.20,\quad c_3=-5.29\pm0.25,\nonumber\\ &&c_4=3.63\pm 0.10,\quad c_5=-0.09\pm 0.01.\end{aligned}$$ For a phenomenological interpretation of the low-energy constants in terms of (meson and $\Delta$) resonance exchanges see Ref. [@Bernard:1997gq]. We will see in the next section that the constants $c_i$ are [*not*]{} required to compensate divergences of one-loop integrals. Such infinities first appear at ${\cal O}(p^3)$. The Lagrangian of Eq. (\[5:5:l2pinhba\]) still contains terms of the type $v\cdot D$ which appears in the lowest-order equation of motion of Eq. (\[5:5:l1pinhbeom\]). As discussed in detail for the mesonic sector in Sec. \[sec\_clop4\] and Appendix \[app\_sec\_glvgss\], such terms can be eliminated by appropriate field redefinitions. For example, the field transformation eliminating in Eq. (\[5:5:l2pinhba\]) the term $$\frac{1}{2\stackrel{\circ}{m}_N} \bar{\cal N}_v(v\cdot D)^2{\cal N}_v$$ is given by [@Ecker:1995rk] $$\label{5:5:nvft} {\cal N}_v=\left[1+\frac{iv\cdot D}{4\stackrel{\circ}{m}_N} -\frac{\stackrel{\circ}{g}_A S_v\cdot u}{4\stackrel{\circ}{m}_N} \right] \tilde{\cal N}_v.$$ Inserting Eq. (\[5:5:nvft\]) into the lowest-order Lagrangian of Eq. (\[5:5:l1pinhb\]) yields $$\begin{aligned} \label{5:5:dell1} &&\bar{\tilde{\cal N}}_v(iv\cdot D+ \stackrel{\circ}{g}_AS_v\cdot u)\tilde{\cal N}_v -\frac{1}{2\stackrel{\circ}{m}_N} \bar{\tilde{\cal N}}_v(v\cdot D)^2\tilde{\cal N}_v -\frac{\stackrel{\circ}{g}_A^2}{2\stackrel{\circ}{m}_N} \bar{\tilde{\cal N}}S_v\cdot u S_v\cdot u \tilde{\cal N}_v\nonumber\\ &&+\,\,\mbox{total derivative} +O\left(\frac{1}{\stackrel{\circ}{m}_N^2}\right).\end{aligned}$$ The second term cancels the equation-of-motion term, whereas rewriting the last term by using Eq. (\[5:5:seig\]), $$S_v\cdot u S_v\cdot u=\frac{1}{4}[(v\cdot u)^2-u\cdot u]+ \frac{1}{2}i\epsilon^{\mu\nu\rho\sigma}v_\rho S^v_\sigma u_\mu u_\nu,$$ we find that some of the coefficients at ${\cal O}(p^2)$ (and at higher orders) are modified. As in the case of the SU(2)$\times$SU(2) mesonic Lagrangian at ${\cal O}(p^4)$ (see Appendix \[app\_sec\_glvgss\]) one finds [*equivalent*]{} parameterizations of $\widehat{\cal L}^{(2)}_{\pi N}$ (and also of the higher-order Lagrangians) in the baryonic sector. For the sake of completeness we quote the result of Ecker and Mojžiš [@Ecker:1995rk], $$\begin{aligned} \label{5:5:l2eckermojzis} \widehat{\cal L}^{(2)}_{\pi N}&=&\bar{\cal N}_v\Bigg\{ -\frac{1}{2\stackrel{\circ}{m}_N}\left(D^2+i\stackrel{\circ}{g}_A \{S_v\cdot D,v\cdot u\}\right) +\frac{a_1}{\stackrel{\circ}{m}_N}\mbox{Tr}(u\cdot u)\nonumber\\ &&+\frac{a_2}{\stackrel{\circ}{m}_N}\mbox{Tr}\left[(v\cdot u)^2\right] +\frac{a_3}{\stackrel{\circ}{m}_N}\mbox{Tr}(\chi_+) +\frac{a_4}{\stackrel{\circ}{m}_N} \left[\chi_+-\frac{1}{2}\mbox{Tr}(\chi_+)\right]\nonumber\\ &&+\frac{1}{\stackrel{\circ}{m}_N}\epsilon^{\mu\nu\rho\sigma}v_\rho S^v_\sigma\left[ia_5 u_\mu u_\nu+a_6 f^+_{\mu\nu}+a_7 v^{(s)}_{\mu\nu}\right] \Bigg\}{\cal N}_v,\end{aligned}$$ where the relation to the coefficients $c_i$ of Eq.  (\[5:5:l2pinhbb\]) is given by $$\begin{aligned} \label{5:5:parrel} &&a_1=\frac{\stackrel{\circ}{m}_N\! c_3}{2}+\frac{\stackrel{\circ}{g}_A^2}{16}, \quad a_2=\frac{\stackrel{\circ}{m}_N\! c_2}{2}-\frac{\stackrel{\circ}{g}_A^2}{8}, \quad a_3=\,\stackrel{\circ}{m}_N\! c_1,\quad a_4=\,\stackrel{\circ}{m}_N\! c_5,\nonumber\\ && a_5=\,\stackrel{\circ}{m}_N\! c_4+\frac{1-\stackrel{\circ}{g}_A^2}{4}, \quad a_6=\,\stackrel{\circ}{m}_N\! c_6+\frac{1}{4},\quad a_7=\,\stackrel{\circ}{m}_N\! c_7+\frac{1}{2}. \quad\end{aligned}$$ Of course, the Lagrangians of Eq. (\[5:5:l2pinhb\]) and (\[5:5:l2eckermojzis\]) yield the same results for physical observables, provided their parameters are related by Eq. (\[5:5:parrel\]). However, they will differ for intermediate mathematical quantities such as vertices or wave function renormalization constants as observed in Ref. [@Fearing:1997dp] for the case of the nucleon wave function renormalization constant. We repeat that the coefficients of the first two terms of Eq. (\[5:5:l2eckermojzis\]) are fixed in terms of $\stackrel{\circ}{m}_N$ and $\stackrel{\circ}{g}_A$, whereas the constants $a_i$ are free parameters which have to be determined by comparison with experimental information. ### The Power Counting Scheme {#subsec_pcs} The power counting scheme of HBChPT may be formulated in close analogy to the mesonic sector (see Sec. \[sec\_elwpcs\]). On the scale of either the nucleon mass $m_N$ or $4\pi F_\pi$ we consider as small external momenta the four-momenta of pions, the four-momenta transferred by external sources, and the residual momenta $k^\mu$ of the nucleon appearing in the separation $p^\mu=\,\stackrel{\circ}{m}_N\! v^\mu + k^\mu$. For a given Feynman diagram we introduce - the number of independent loop momenta $N_L$, - the number of internal pion lines $I_M$, - the number of pion vertices $N^M_{2n}$ originating from ${\cal L}_{2n}$, - the total number of pion vertices $N_M=\sum_{n=1}^\infty N^M_{2n}$, - the number of internal nucleon lines $I_B$, - the number of baryonic vertices $N^B_n$ originating from $\widehat{\cal L}^{(n)}_{\pi N}$, - and the total number of baryonic vertices $N_B=\sum_{n=1}^\infty N^B_n$. As in the mesonic sector, the internal momenta appearing in the loop integration are not necessarily small. However, via the four-momentum conserving delta functions at the vertices and a substitution of integration variables, the rescaling of the external momenta is transferred to the internal momenta (see Sec. \[sec\_elwpcs\]). The chiral dimension $D$ of a given diagram is then given by [@Weinberg:1991um; @Ecker:1995gg] $$\label{5:5:d1} D=4 N_L - 2 I_M - I_B +\sum_{n=1}^\infty 2n N^M_{2n} +\sum_{n=1}^\infty n N_n^B.$$ We make use of the topological relation \[see, e.g., Eq. (2.130) of Ref. [@Cheng:bj]\] $$\label{5:5:NLint} N_L=I_M +I_B-N_M-N_B+1$$ to eliminate $I_M$ from Eq. (\[5:5:d1\]) $$\label{5:5:d2} D=2 N_L +I_B +2 +\sum_{n=1}^\infty 2(n-1)N^M_{2n} +\sum_{n=1}^\infty (n-2)N_n^B.$$ For processes containing exactly one nucleon in the initial and final states we have[^115] $N_B=I_B+1$ and we thus obtain $$\label{5:5:d3} D=2 N_L+1 +\sum_{n=1}^\infty 2(n-1)N^M_{2n} +\sum_{n=1}^\infty (n-1)N_n^B.$$ The power counting is very similar to the mesonic sector. We first observe that $D\ge 1$. Moreover, as already mentioned in Sec. \[subsec\_cfo1om\], loops start contributing at $D=3$. In other words, the low-energy coefficients $c_i$ of $\widehat{\cal L}^{(2)}_{\pi N}$ are not needed to renormalize infinities from one-loop calculations. Again, we have a connection between the number of loops and the chiral dimension $D$: $N_L\le (D-1)/2$. Each additional loop adds two units to the chiral dimension. As an example, let us consider the two-loop contribution to the nucleon self energy of Fig. \[5:5:fig:twoloopselfenergy\]. First of all, the number of independent loops is $N_L=2$ in agreement with Eq. (\[5:5:NLint\]) for $I_M=2$, $I_B=3$, $N_M=0$, and $N_B=4$. The counting of the chiral dimension is most intuitively performed in the framework of Eq. (\[5:5:d1\]), because it associates with each building block a unique term which is easy to remember ($+4$ for each independent loop, $-2$ for each internal meson propagator, etc). For $N^M_{2n}=0$, $N^B_1=2$, and $N^B_2=2$ we obtain $D=8-4-3+0+2+4=7$. ### Application at ${\cal O}(p^3)$: One-Loop Correction to the Nucleon Mass {#subsec_aop3olcnm} As a simple example, we will return to the modification of the nucleon mass through higher-order terms in the heavy-baryon approach. The calculation will proceed along the lines of Ref. [@Fearing:1997dp], where use was made of the Lagrangian of Ecker and Mojžiš [@Ecker:1995rk] \[see Eq. (\[5:5:l2eckermojzis\])\]. The determination of the physical nucleon mass and the discussion of the wave function renormalization factor will be very similar to Secs.\[subsec\_mgb\] for the masses of the Goldstone bosons and \[subsec\_feolcnm\] for the nucleon mass in the relativistic approach. Let us denote the four-momentum of the nucleon by $p = \, \stackrel{\circ}{m}_N\! v +r$, where, since we are interested in the propagator, we must allow the four-momentum to be off the mass shell. The on-shell case is, of course, given by $p^2=m_N^2$ with $m_N$ denoting the [*physical*]{} nucleon mass. Let us stress that, due to the interaction, we must expect the physical mass to be different from the mass $\stackrel{\circ}{m}_N$ in the chiral limit. We start from the lowest-order propagator of Eq. (\[5:5:gv0k\]), $$\label{5:5:prop0} \frac{P_{v+}}{v\cdot r +i0^+} =\frac{P_{v+}}{v\cdot p\,-\stackrel{\circ}{m}_N +\,i0^+},$$ and first determine its modification in terms of the tree-level contribution of Eq. (\[5:5:l2eckermojzis\]) to the self energy.[^116] Neglecting isospin-symmetry breaking effects proportional to $m_u-m_d$, we obtain at ${\cal O}(p^2)$ $$\label{5:5:sigma2} \Sigma^{(2)}(p)=-\frac{r^2}{2\stackrel{\circ}{m}_N} -\frac{4a_3 M^2}{\stackrel{\circ}{m}_N},$$ where $M^2=2B_0 m_q$ denotes the squared pion mass at ${\cal O}(p^2)$. The $r^2$ term comes from the term in $\widehat{\cal L}^{(2)}_{\pi N}$ proportional to $-\partial^2/2\stackrel{\circ}{m}_N$ which involves no pions. In the spirit of the reduction formula of Eq. (\[5:5:ffunc\]), in combination with the formula of Gell-Mann and Low of Eq. (\[5:5:gmlhb\]), we choose to include this as part of the interaction rather than part of the free Lagrangian and reserve for the free Lagrangian the $i v \cdot \partial$ term from $\widehat{\cal L}^{(1)}_{\pi N}$. The term involving $a_3$ is a contact term coming ultimately from Eq.  (\[5:5:l2pinhbb\]), where $a_3=\,\stackrel{\circ}{m}_N\! c_1$. The heavy-baryon Lagrangian at ${\cal O}(p^3)$ [@Ecker:1995rk] does not produce a contact contribution to the self energy, because all the structures contain at least one pion or an external field. Moreover, given the Feynman rule of Eq. (\[5:6:fr2\]), the second one-loop diagram of Fig. \[5:5:fig:hbnseloop\] (b) vanishes, because $\epsilon_{iij}\tau_j=0$. In other words, the only contribution at ${\cal O}(p^3)$ results from the one-loop diagram of Fig. \[5:5:fig:hbnseloop\] (a). Using the vertex of Eq. (\[5:5fr1\]) and the propagator of Eq.(\[5:5:gv0k\]) we obtain for the one-loop contribution at ${\cal O}(p^3)$ $$\begin{aligned} \label{5:5:sigmaloop1} -i\Sigma^{(3)}_{\rm loop}(p)&=&\int\frac{d^4 k}{(2\pi)^4}\left[ \frac{\stackrel{\circ}{g}_A}{F_0}(-S_v\cdot k)\tau_i\right] \frac{i}{v\cdot(r-k)+i0^+}\nonumber\\ &&\times \frac{i}{k^2-M^2+i0^+} \left[\frac{\stackrel{\circ}{g}_A}{F_0}S_v\cdot k\tau_i\right].\end{aligned}$$ As in the relativistic case of Eq. (\[5:4:sel1\]), counting powers, we expect the integral to have a cubic divergence. Extending the integral to $n$ dimensions, using $\tau_i\tau_i=3$, performing the substitution $k\to -k$, and applying Eq. (\[app:C20C21\]) of Appendix \[subsec\_jpin\], we obtain the intermediate result $$\Sigma^{(3)}_{\rm loop}(p)=3 \frac{\stackrel{\circ}{g}_A^2}{F_0^2} S^\mu_v S^\nu_v[v_\mu v_\nu C_{20}(v\cdot r, M^2)+g_{\mu\nu} C_{21} (v\cdot r, M^2)].$$ Since $S_v\cdot v=0$ and $S^2_v=(1-n)/4$ \[see Eq. (\[5:5:s2\])\], we obtain, applying the first equality of Eq. (\[app:C21\]), $$\Sigma^{(3)}_{\rm loop}(p)=-\frac{3}{4}\frac{\stackrel{\circ}{g}_A^2}{F_0^2} \left\{[M^2-(v\cdot r)^2]J_{\pi N}(0; v\cdot r)+v\cdot r I_\pi(0)\right\},$$ where the integrals $J_{\pi N}$ and $I_\pi$ are given in Eqs.(\[app:jpinb\]) and (\[app:ipi\]), respectively. Combining with Eq. (\[5:5:sigma2\]) and using Eq. (\[app:jpinb\]) we thus obtain for the (unrenormalized) nucleon self energy at ${\cal O}(p^3)$ [@Fearing:1997dp] $$\begin{aligned} \label{5:6:sigmaN} \lefteqn{\Sigma(p) = -\frac{r^2}{2\stackrel{\circ}{m}_N} -\frac{4a_3 M^2}{\stackrel{\circ}{m}_N}}\nonumber\\ &&-\frac{3\stackrel{\circ}{g}_A^2}{(4\pi F_0)^2}\left( \frac{v\cdot r}{4}\left\{[3M^2-2(v\cdot r)^2] \left[R+\ln\left(\frac{M^2}{\mu^2}\right)\right] -\frac{1}{2}\left[M^2-(v\cdot r)^2\right]\right\}\right.\nonumber\\ &&\left.+[M^2-(v\cdot r)^2]^\frac{3}{2}\arccos\left(-\frac{v\cdot r}{M}\right) \right),\end{aligned}$$ for $(v\cdot r)^2<M^2$. Clearly, the self-energy contribution generated by the loop diagram of Fig. \[5:5:fig:hbnseloop\] (a) contains a divergent piece proportional to $R$ of Eq. (\[app:constantR\]). We have chosen to express the self energy as a function of the four-momentum $p$. In the relativistic case of Eq. (\[5:4:seansatz\]) we needed two scalar functions depending on $p^2$ to parameterize the self energy. In contrast to the relativistic case, the heavy-baryon self energy of Eq.(\[5:6:sigmaN\]) is given by one function depending on two scalar variables for which one can take, say, $r^2$ and $v\cdot r$ or $$\label{5:5:etaxi} \eta\equiv v\cdot p-m_N,\quad \xi\equiv (p-m_N v)^2.$$ Making use of $r=(m_N-\stackrel{\circ}{m}_N)v+(p-m_N v)$, the two sets are related by $$\begin{aligned} \label{5:5:rvretaxi1} r^2&=&(m_N-\stackrel{\circ}{m}_N)^2 +2(m_N-\stackrel{\circ}{m}_N)\eta +\xi,\\ \label{5:5:rvretaxi2} v\cdot r&=&m_N-\stackrel{\circ}{m}_N+\,\eta.\end{aligned}$$ The choice of Eq. (\[5:5:etaxi\]) is convenient for the determination of the physical nucleon mass $m_N$ and renormalization constant $Z_N$, because, in view of Eq. (\[5:5:prop0\]), we want the full (but yet unrenormalized) propagator to have a pole at $p=m_N v$ which includes both the mass-shell condition $p^2=m_N^2$ and $v \cdot p =m_N$. In the vicinity of the pole at $p=m_N v$ the second choice of variables corresponds to terms which are, respectively, first and second order in the (small) distance from the pole. Thus in the following discussion we will use both notations $\Sigma(p)$ and $\Sigma(\eta,\xi)$ for the self energy, where it should be clear from the context which expression applies. In analogy to the mesonic case discussed in Sec. \[subsec\_mgb\] the full heavy-baryon propagator is written as \[see Eq. (\[4:8:prop2\])\] $$\begin{aligned} \label{5:5:fullhbp} iG_v(p) &=& \frac{i}{v \cdot p\,\, -\stackrel{\circ}{m}_N-\, \Sigma(p)}\nonumber\\ &=&\frac{i}{v \cdot p\,\, -\stackrel{\circ}{m}_N -\,\Sigma(0,0)-\eta \Sigma^\prime(0,0)-\tilde{\Sigma}(\eta,\xi)}\nonumber \\ &=& \frac{i}{[1-\Sigma^\prime (0,0)] \left\{\eta - \frac{\tilde{\Sigma}(\eta,\xi)}{[1-\Sigma^\prime (0,0)]} \right\}}\nonumber\\ &=& \frac{i Z_N}{\eta - Z_N \tilde{\Sigma}(\eta,\xi)},\end{aligned}$$ where $$\begin{aligned} \label{5:5:mndef} m_N&=&\stackrel{\circ}{m}_N+\,\Sigma(0,0),\\ \label{5:5:zndef} Z_N&=&\frac{1}{1-\Sigma^\prime (0,0)}.\end{aligned}$$ In these equations $\Sigma^\prime (0,0)$ denotes the first partial derivative of $\Sigma(\eta,\xi)$ with respect to $\eta$ evaluated at $(\eta,\xi)=(0,0)$, $$\Sigma^\prime (0,0)=\left.\frac{\partial\Sigma(\eta,\xi)}{\partial\eta} \right|_{(\eta,\xi)=(0,0)},$$ and $\tilde{\Sigma}(\eta,\xi)$ is at least of second order in the distance from the pole. For the evaluation of $\Sigma(0,0)$, $\Sigma^\prime (0,0)$, and $\tilde{\Sigma}(\eta,\xi)$ we need to expand Eq. (\[5:6:sigmaN\]). To the order we are working the $a_3$ term contributes only to $\Sigma(0,0)$ whereas the loop piece contributes to all three. In contrast to the mesonic sector at ${\cal O}(p^4)$, $\tilde{\Sigma}$ is not zero in this case. Using Eq. (\[5:5:rvretaxi1\]), we obtain for the $r^2$ term $$\label{5:5:rsq} \frac{r^2}{2\stackrel{\circ}{m}_N} =\frac{(m_N-\stackrel{\circ}{m}_N)^2}{2\stackrel{\circ}{m}_N} +\frac{(m_N-\stackrel{\circ}{m}_N)}{\stackrel{\circ}{m}_N} \eta + \frac{\xi}{2\stackrel{\circ}{m}_N}.$$ The first term on the right-hand side contributes to $\Sigma(0,0)$ but is ${\cal O}(1/m_N^3)$, since, as we will see, the difference $(m_N-\stackrel{\circ}{m}_N)$ is ${\cal O}(1/m_N)$. The second term is ${\cal O}(1/m_N^2)$ and will contribute to $\Sigma^\prime (0,0)$. Finally the third term contributes only to $\tilde{\Sigma}$. Applying Eq. (\[5:5:mndef\]) we obtain for the physical mass $$\label{5:5:nucmass1} m_N =\, \stackrel{\circ}{m}_N - \frac{(m_N-\stackrel{\circ}{m}_N)^2}{2\stackrel{\circ}{m}_N} -\frac{4a_3 M^2}{\stackrel{\circ}{m}_N} +\Sigma^{(3)}_{\rm loop}(0,0),$$ which implies $m_N-\stackrel{\circ}{m}_N={\cal O}(1/m_N)$.[^117] We can thus neglect the second term on the right-hand side of Eq.(\[5:5:nucmass1\]). The loop contribution is only a function of $v\cdot r$ and thus a function only of $\eta$, and, neglecting terms of higher order in $1/m_N$, we may replace $v\cdot r$ by 0, yielding $$\Sigma^{(3)}_{\rm loop}(0,0)=-\frac{3{\stackrel{\circ}{g}_A^2}M^3} {(4\pi F_0)^2}\arccos(0).$$ We finally obtain for the physical nucleon mass $$\label{5:5:nucmass2} m_N \simeq\, \stackrel{\circ}{m}_N \left[1 -\frac{4a_3 M_\pi^2}{m_N^2}-\frac{3\pi{g_A^2}M_\pi^3} {2m_N(4\pi F_\pi)^2}\right],$$ where, in the expression between the brackets, we have replaced all quantities in terms of the physical quantities, because the difference is of higher order in the chiral expansion. In the chiral limit, both the counter-term contribution $\sim M^2\sim m_q$ and the pion-loop correction $\sim M^3\sim m_q^{3/2}$ disappear. In other words, in the heavy-baryon framework the situation is again as in the mesonic sector, where the parameters of the lowest-order Lagrangian do not get modified due to higher-order corrections in the chiral limit. The same is actually true for the second parameter $\stackrel{\circ}{g}_A$ of Eq. (\[5:5:l1pinhb\]) \[see, e.g., Eq. (50) of Ref.  [@Fearing:1997dp]\]. Using the parameters of Eqs. (\[5:3:par\]) and (\[5:5:ciresults\]) one finds that the counter term and the pion loop generate contributions to the physical nucleon mass of $0.0733$ and $-0.0163$ in units of $\stackrel{\circ}{m}_N$, respectively. The wave function renormalization constant $Z_N$ is obtained from Eq. (\[5:5:zndef\]) as $$\label{5:5:znresulttemp} Z_N = \frac{1}{1-\Sigma'(0,0)} \approx 1+\Sigma'(0,0) =1-\frac{m_N-\stackrel{\circ}{m}_N}{\stackrel{\circ}{m}_N} +\Sigma'_{\rm loop}(0,0).$$ To the order we are considering, we have from Eqs. (\[5:5:nucmass2\]) and (\[5:6:sigmaN\]), respectively, $$\begin{aligned} \frac{m_N-\stackrel{\circ}{m}_N}{\stackrel{\circ}{m}_N}&=& -\frac{4 a_3 M^2}{\stackrel{\circ}{m}_N^2},\\ \Sigma'_{\rm loop}(0,0)&=&-\frac{9\stackrel{\circ}{g}_A^2M^2}{4(4\pi F_0)^2} \left[R+{\rm ln}\left(\frac{M^2}{\mu^2}\right)+\frac{2}{3}\right].\end{aligned}$$ Finally, expressing all quantities in terms of physical quantities, the wave function renormalization constant $Z_N$ reads $$\label{5:5:znresult} Z_N = 1+\frac{4a_3M_\pi^2}{m_N^2}-\frac{9g_A^2M_\pi^2}{4(4\pi{F_\pi})^2} \left[R+{\rm ln}\left(\frac{M_\pi^2}{\mu^2}\right)+\frac{2}{3}\right].$$ As in the pion case (see Table \[app:dp:tab:abz\] of Appendix \[app\_sec\_dp\]) $Z_N$ contains the infinite constant $R$ entering through dimensional regularization, i.e., $Z_N$ is not a finite quantity. However, this is not a problem, because the wave function renormalization constant is not a physical observable. Moreover, as we have seen explicitly for the pion case, and as discussed in Ref. [@Fearing:1997dp] for the heavy-baryon Lagrangian, $Z_N$ will also depend on the specific parameterization of the Lagrangian. In Ref. [@Ecker:1997dn] it was shown that the wave function renormalization “constant” $Z_N$ is in fact a non-trivial differential operator and should, in momentum space, depend on the momentum of the initial or final nucleon. Here we argue that the findings of Ref. [@Ecker:1997dn] and the method used above do not seem to be in conflict with each other. To that end, we first note that Ref. [@Ecker:1997dn] made use of the external spinor $u_+(\vec{p}\,)=P_{v+}u(\vec{p}\,)$. Using relativistic spinors normalized as in Eq. (\[5:5:spinorprops\]) this corresponds to a normalization of the heavy baryon spinors to $\bar{u}_+(\vec{p}\,)u_+(\vec{p}\,)=(p\cdot v+ m_N)$. To facilitate the comparison, let us consider the special choice $v^\mu=(1,0,0,0)$. In the framework of the reduction formula of Eq. (\[5:5:ffunc\]), we work with a factor $N u^{\alpha}_v/\sqrt{Z_N}$ for an external nucleon in the initial state, where Ecker and Mojžiš would have $u_+(\vec{p}\,)/\sqrt{Z_N^{\rm EM}}$. It is now straightforward to show that the normalization factor $N$ exactly produces the additional term which, in the approach of Ref. [@Ecker:1997dn], results from the additional term in the wave function renormalization. This explains why the two approaches, at least up to order ${\cal O}(p^3)$, generate the same result. For further discussion on this topic, the reader is referred to [@Fearing:1997dp; @Ecker:1997dn; @Steininger:1998ya; @Kambor:1998pi]. The Method of Infrared Regularization {#sec_mir} ------------------------------------- In the discussion of the one-loop corrections to the nucleon self energy and pion-nucleon scattering of Sec. \[sec\_eld\], we saw that the relativistic framework for baryons did not naturally provide a simple power counting scheme as for mesons. One major difference in comparison with the mesonic sector is related to the fact that the nucleon remains massive in the chiral limit which also introduces another mass scale into the problem. Thus, because of the zeroth component one can no longer argue that a derivative acting on the baryon field results in a small four-momentum. This problem is avoided in the heavy-baryon approach discussed in Sec. \[sec\_hbf\], where, through a field redefinition, the mass dependence has been shifted into an (infinite) string of vertices which are suppressed by powers of $1/m$. Since the derivatives in the heavy-baryon Lagrangian produce small residual four-momenta in the low-energy regime, a power counting scheme analogous to the mesonic sector can be formulated \[see Eqs. (\[5:5:d1\]) and (\[5:5:d3\])\]. A vast majority of applications of chiral perturbation theory in the baryonic sector were performed in the framework of the heavy-baryon approach. However, it was realized some time ago that the heavy-baryon approach, under certain circumstances, may generate Green functions which do not satisfy the analytic properties resulting from a (fully) relativistic field theory [@Bernard:1996cc]. Clearly, it would be desirable to have a method which combines the advantages of the relativistic and the heavy-baryon approaches and, at the same time, avoids their shortcomings—absence of a power counting scheme on the one hand and failure of convergence on the other hand. Such approaches have been proposed and developed by various authors and here we will briefly outline the ideas of the so-called infrared regularization [@Becher:1999he]. Our presentation will closely follow Refs.  [@Becher:1999he; @Becher:2000mb] to which we refer the reader for technical details. Some recent applications of the new approach deal with the electromagnetic form factors of the nucleon [@Kubis:2000zd] and the baryon octet [@Kubis:2000aa], $\pi N$ scattering [@Becher:2001hv], axial-vector current matrix elements [@Zhu:2000zf], and the generalized Gerasimov-Drell-Hearn sum rule [@Bernard:2002bs]. In order to understand the problems of the heavy-baryon approach regarding the analytic behavior of invariant functions let us start with a simple example [@Becher:2000mb]. To that end we consider the $s$ channel of pion-nucleon scattering (see Sec. \[subsec\_apnstl\]). The invariant amplitudes $B^\pm$ of Table \[5:3:tableresults\] develop poles for $\nu=\pm \nu_B$ (the upper and lower signs correspond to $s=m_N^2$ and $u=m_N^2$, respectively). For example, the singularity due to the nucleon pole in the $s$ channel is understood in terms of the relativistic propagator $$\label{5:6:relprop} \frac{1}{(p+q)^2-m_N^2}=\frac{1}{2p\cdot q+M_\pi^2},$$ which, of course, has a pole at $2p\cdot q=-M_\pi^2$ or, equivalently, $s=m_N^2$. (Analogously, a second pole results from the $u$ channel at $u=m_N^2$.) We also note that the propagator of Eq. (\[5:6:relprop\]) counts as ${\cal O}(p^{-1})$, because it is part of a tree-level diagram so that the four-momentum $q$ is assumed to be small, i.e., of ${\cal O}(p)$. Although both poles are not in the physical region of pion-nucleon scattering, analyticity of the invariant amplitudes requires these poles to be present in the amplitudes. Let us compare the situation with a heavy-baryon type of expansion, where, for simplicity, we choose as the four-velocity $p^\mu=m_N v^\mu$, $$\label{5:6:relpropv} \frac{1}{2p\cdot q+M_\pi^2}= \frac{1}{2m_N}\frac{1}{v\cdot q+ \frac{M_\pi^2}{2m_N}}= \frac{1}{2 m_N}\frac{1}{v\cdot q}\left(1-\frac{M_\pi^2}{2 m_N v\cdot q} +\cdots\right).$$ Clearly, to any finite order the heavy-baryon expansion produces poles at $v\cdot q=0$ instead of a simple pole at $v\cdot q=-M_\pi^2/(2 m_N)$ and will thus not generate the (nucleon) pole structures of the functions $B^\pm$. As a second example, we consider the so-called triangle diagram of Fig. \[5:6:fig:triangle\] which will serve to illustrate the different analytic properties of invariant functions obtained from loop diagrams in the relativistic and heavy-baryon approaches. A diagram of this type appears in many calculations such as the scalar or electromagnetic form factors of the nucleon, where $\bullet$ represents an external scalar or electromagnetic field, or $\pi N$ or Compton scattering, where $\bullet$ stands for two pion or electromagnetic fields. In all of these cases a four-momentum $q$ is transferred to the nucleon and the analytic properties of the Feynman diagram as a function of $t\equiv q^2$ are determined by the pole structure of the propagators. Thus we need to discuss some properties of the integral $$\label{5:6:gammat} \gamma(t)\equiv i\int \frac{d^4 k}{(2\pi)^4} \frac{1}{k^2-M^2_\pi+i0^+} \frac{1}{(k+q)^2-M^2_\pi+i0^+}\frac{1}{(p-k)^2-m^2_N+i0^+},$$ where $t=q^2$. We assume the initial and final nucleons to be on the mass shell, $p^2=m^2_N=(p+q)^2$ which implies $2p\cdot q=-t$. Counting powers we see that the integral converges. The function $\gamma(t)$ is analytic in $t$ except for a cut along the positive real axis starting at $t= 4M^2_\pi$ which expresses the fact that two on-shell pions can be produced for $t\geq 4M^2_\pi$. In the following discussion of the analytic properties of Eq.  (\[5:6:gammat\]) we will concentrate on the imaginary part of $\gamma(t)$ which we will derive applying the Cutkosky (or cutting) rules [@Cutkosky:1960sp; @LeBellac:cq; @Peskin:ev]. The rules, as summarized in [@Peskin:ev], read: In order to obtain $2i {\rm Im}\{\gamma(t)\}$ first cut through the diagram in all possible ways such that the cut propagators can simultaneously be put on shell. Next, for each cut one need to replace each cut propagator $1/(p^2-m^2+i0^+)$ by $-2\pi i\delta(p^2-m^2)$. Finally, sum the contributions of all possible cuts. In the present case, the two terms where the nucleon propagator is simultaneously cut with either the first or the second pion propagator do not contribute. The result from cutting the two pion propagators reads $$\label{5:6:img1t} 2i{\rm Im}\{\gamma(t)\}=(-2\pi i)^2 i \int \frac{d^4 k}{(2\pi)^4} \frac{\delta(k^2-M_\pi^2)\delta((k+q)^2-M_\pi^2)}{(p-k)^2-m_N^2+i0^+}.$$ In order to evaluate Eq. (\[5:6:img1t\]), we choose a frame where $q^\mu=(q_0,\vec{0})$ with $q_0=\sqrt{t}>0$, and $p^\mu=(-q_0/2,\vec{p}\,)$. Using $$\delta(k^2-M_\pi^2)\delta((k+q)^2-M_\pi^2)=\delta(k^2-M_\pi^2) \frac{1}{2\sqrt{t}}\delta\left(k_0+\frac{\sqrt{t}}{2}\right)$$ we find, as an intermediate result, $$\label{5:6:img2t} {\rm Im}\{\gamma(t)\}=-\frac{1}{16\pi^2\sqrt{t}}\int d^3 k \delta\left(\vec{k}^2 +M_\pi^2-\frac{t}{4}\right)\frac{1}{ -\frac{t}{2}+2\vec{p}\cdot\vec{k}+M_\pi^2+i0^+}.$$ For $t<4 M_\pi^2$, the delta function in Eq. (\[5:6:img2t\]) always vanishes, showing that the cut starts, as anticipated, at $t=4M_\pi^2$. Applying the mass-shell condition $p^2=m_N^2$, we write $$\begin{aligned} \vec{p}=i\frac{\sqrt{4m_N^2-t}}{2} \hat{e}_z&\mbox{for}&4M_\pi^2\leq t\leq 4m_N^2,\\ \vec{p}=\frac{\sqrt{t-4m_N^2}}{2} \hat{e}_z&\mbox{for}&4m_N^2\leq t.\end{aligned}$$ Performing the integration using spherical coordinates, the result for the first case reads $$\begin{aligned} \label{5:6:imgammat} {\rm Im}\{\gamma(t)\}&=&\frac{\sqrt{t-4M_\pi^2}}{16\pi \sqrt{t}}\int_{-1}^1 dz \frac{1}{t-2M_\pi^2-i\sqrt{4m_N^2-t}\sqrt{t-4M_\pi^2}\, z-i0^+}\nonumber\\ &=&\frac{i}{16\pi \sqrt{t}\sqrt{4m_N^2-t}}\ln\left(\frac{1-iy}{1+iy}\right) \nonumber\\ &=&\frac{1}{8\pi \sqrt{t(4m_N^2-t)}}\arctan(y),\end{aligned}$$ where $$y=\frac{\sqrt{(t-4M^2_\pi)(4m^2_N-t)}}{t-2M^2_\pi},\quad 4M_\pi^2\leq t\leq 4m_N^2.$$ The second case, $t>4m_N^2$, is obtained analogously by the replacement $i\sqrt{4m_N^2-t} \to \sqrt{t-4m_N^2}$: $$\label{5:6:img3ta} {\rm Im}\{\gamma(t)\}=\frac{1}{16\pi\sqrt{t(t-4m_N^2)}} \ln\left(\frac{t-2M_\pi^2+\sqrt{t-4m_N^2}\sqrt{t-4M_\pi^2}}{ t-2M_\pi^2-\sqrt{t-4m_N^2}\sqrt{t-4M_\pi^2}}\right).$$ Equations (\[5:6:imgammat\]) and (\[5:6:img3ta\]) agree with the results given in Eq. (B.43) of Ref. [@Gasser:1987rb]. In the low-energy region $t\ll m^2_N$, and Eq. (\[5:6:imgammat\]) becomes $$\label{5:6:imgammatexp} \mbox{Im}\{\gamma(t)\}\approx\frac{1}{16\pi m_N\sqrt{t}}\arctan(x),\quad x=\frac{2m_N\sqrt{t-4M^2_\pi}}{t-2M^2_\pi}.$$ Taking the factors resulting from the vertices and the relevant tensor structures of the loop integral into account, the contribution of Fig. \[5:6:fig:triangle\] to the imaginary part of the scalar form factor of the nucleon reads [@Gasser:1987rb; @Becher:1999he] $${\rm Im}\{\sigma(t)\}=\frac{3 g^2_A M_\pi^2 m_N}{4 F_\pi^2}(t-2M_\pi^2) {\rm Im}\{\gamma(t)\},$$ where $\sigma(t)$ is defined in terms of the $u$- and $d$-quark scalar densities $\bar{u}u$ and $\bar{d}d$ as $$\label{5:6:sffdef} \langle N(p')| m[\bar{u}(0)u(0)+\bar{d}(0)d(0)]|N(p)\rangle= \bar{u}(p')u(p)\sigma(t),$$ where $m=m_u=m_d$ and $t=q^2=(p'-p)^2$. We will now investigate two limiting procedures. First, we consider a fixed value $t>4 M_\pi^2$ and let $m_N\to \infty$. In that case $x\gg 1$, and one would use the expansion $$\arctan(x)=\frac{\pi}{2}-\frac{1}{x}+\frac{1}{3 x^3}-\cdots, \quad x>1.$$ Keeping only the leading order term, we find $$\label{5:6:imsfhbchpt} {\rm Im}\{\sigma(t)\}= \frac{3 g_A^2 M_\pi^2}{128 F_\pi^2}\frac{t-2 M_\pi^2}{\sqrt{t}},$$ which corresponds exactly to the result of HBChPT at ${\cal O}(p^3)$ [@Bernard:1992qa]. This result corresponds to the standard chiral expansion which treats the quantity $x$ of Eq. (\[5:6:imgammatexp\]) as ${\cal O}(p^{-1})$, because $m_N={\cal O}(p^0)$ and $t,M_\pi^2={\cal O}(p^2)$. However, for a fixed $m_N$ we may also consider a small enough $t$ close to the threshold value $t_{\rm thr}=4M^2_\pi$ so that $x<1$. In that case the expansion of the arctan reads $$\arctan(x)=x-\frac{x^3}{3}+\cdots$$ yielding $$\label{5:6:imsfrchpt} {\rm Im}\{\sigma(t)\}\approx \frac{3 g_A^2 M_\pi^2 m_N}{32 \pi F_\pi^2}\frac{\sqrt{t-4 M_\pi^2}}{\sqrt{t}},$$ where we have neglected higher powers of $x$. The critical value of $t$ corresponding to $x=1$ is given by $$t_{\rm cr}=4M_\pi^2\left[1+\frac{\mu^2}{4}+O(\mu^4)\right],$$ where $\mu=M_\pi/m_N$. Clearly the behavior of Eq. (\[5:6:imsfrchpt\]) is very different from the chiral expansion of Eq. (\[5:6:imsfhbchpt\]) and, similar to the discussion of Eq. (\[5:6:relpropv\]), a finite sum of terms in HBChPT cannot reproduce such a threshold behavior [@Becher:1999he]. The rapid variation of the imaginary part can be understood in terms of the analytic properties of the arctan which, as a function of the complex variable $z$, is analytic in the entire complex plane save for cuts along the positive and negative imaginary axis starting at $\pm i$. These branch points corresponding to $x=\pm i$ are obtained for $t=4M_\pi^2(1-\mu^2/4)$ which is just below the physical threshold $t_{\rm thr}=4M_\pi^2$. For that reason an expansion around $x=0$ corresponding to $t=4M_\pi^2$ has a small radius of convergence. Clearly, the heavy-baryon approach does not produce the correct analytic structure as generated by the relativistic loop diagram. Moreover the low-energy behavior of Eq. (\[5:6:imgammatexp\]) cannot be accounted for in the standard chiral analysis because the argument $x$ is of order ${\cal O}(p^{-1})$. What is needed is a method which produces both the relevant analytic structure and a consistent power counting. Here we will illustrate the method of Ref. [@Becher:1999he] by means of the nucleon self energy diagram of Fig. \[5:4:fig:nsepl\]. For a—at this stage—qualitative discussion of its properties we focus on the scalar loop integral $$\label{5:6:defh} H(p^2,n)=-i\int \frac{d^n k}{(2\pi)^n}\frac{1}{k^2-M_\pi^2+i0^+} \frac{1}{k^2-2p\cdot k+(p^2-m_N^2)+i0^+},$$ where, as usual, the right-hand side is thought of as a Feynman integral which has to be analytically continued as a function of the space-time dimension $n$. Counting powers, we see that, for $n=4$, the integrand behaves for large values of the integration variable $k$ as $k^3/k^4$, producing a logarithmic ultraviolet divergence while, on the other hand, the integral converges for $n<4$. Let us now consider the limit $M_\pi^2\to 0$. In this case, for both $p^2=m_N^2$ and $p^2\neq m_N^2$, the integral is infrared regular for $n=4$ because, for small momenta, the integrand behaves as $k^3/k^3$ and $k^3/k^2$, respectively. For $n=3$ the integral is infrared regular for $p^2\neq m_N^2$ but singular for $p^2=m_N^2$. For any smaller value of $n$ it is infrared singular for arbitrary $p^2$. The infrared singularity as $M_\pi^2\to 0$ originates in the region, where the integration variable $k$ is small, i.e., of the order ${\cal O}(p)$. Counting powers of momenta, we (naively) expect this part to be of order ${\cal O}(p^{n-3})$. On the other hand, for loop momenta of the order of and larger than the nucleon mass we expect power counting to fail, because the momentum of the nucleon propagating in loop integral is not constrained to be small in contrast to the case of tree-level diagrams \[see Eq. (\[5:6:relprop\])\]. In order to explain these qualitative statements let us discuss the integral in more detail. We first introduce the Feynman parameterization[^118] $$\label{5:6:feynmanparh} H(p^2,n)=-i\int \frac{d^n k}{(2\pi)^n}\int_0^1 dz \frac{1}{[az+b(1-z)]^2},$$ with $a=k^2-2k\cdot p+p^2-m_N^2+i0^+$ and $b=k^2-M_\pi^2+i0^+$, perform the shift $k\to k+pz$, and obtain $$H(p^2,n)=-i \int_0^1 dz\int\frac{d^n k}{(2\pi)^n}\frac{1}{[k^2-A(z) +i0^+]^2},$$ where $$A(z)=z^2 p^2-z(p^2-m_N^2+M_\pi^2)+M_\pi^2.$$ We then apply Eq. (\[app:drb:moregenint\]) of Appendix \[app\_drb\], $$\label{5:6:defh1} H(p^2,n)= \frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^{\frac{n}{2}}} \int_0^1 dz [A(z)-i0^+]^{\frac{n}{2}-2}.$$ The relevant properties can nicely be displayed at the threshold $p^2_{\rm thr}=(m_N+M_\pi)^2$, where $A(z)=[z(m_N+M_\pi)-M_\pi]^2$ is particularly simple. The small imaginary part can be dropped in this case, because $A(z)$ is never negative. Splitting the integration interval into $[0,z_0]$ and $[z_0,1]$ with $z_0=M_\pi/(m_N+M_\pi)$, we have, for $n>3$, $$\begin{aligned} \int_0^1 dz [A(z)]^{\frac{n}{2}-2}&=&\int_0^{z_0}dz [M_\pi-z(m_N+M_\pi)]^{n-4} \\ &&+\int_{z_0}^1dz [z(m_N+M_\pi)-M_\pi]^{n-4}\\ &=&\frac{1}{(n-3)(m_N+M_\pi)}(M_\pi^{n-3}+m_N^{n-3}),\end{aligned}$$ yielding, through analytic continuation, for arbitrary $n$ $$\label{5:6:defhthr} H((m_N+M_\pi)^2,n)= \frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^{\frac{n}{2}}(n-3)} \left(\frac{M_\pi^{n-3}}{m_N+M_\pi}+\frac{m_N^{n-3}}{m_N+M_\pi}\right).$$ The first term, proportional to $M_\pi^{n-3}$, is defined as the so-called infrared singular part $I$ which, as $M_\pi\to 0$, behaves as in the qualitative discussion above. Since $M_\pi\to 0$ implies $p^2_{\rm thr} \to m_N^2$ this term is singular for $n\leq 3$. The second term, proportional to $m_N^{n-3}$, is defined as the infrared regular part $R$ and can be thought of as originating from an integration region where $k$ is of order $m_N$ so that the tree-level counting rules no longer apply \[see Eq. (\[5:6:relprop\])\]. Note that for non-integer $n$ the infrared singular part contains non-integer powers of $M_\pi$, while an expansion of the regular part always contains non-negative integer powers of $M_\pi$ only. Let us now turn to a [*formal*]{} definition of the infrared singular and regular parts [@Becher:1999he] which makes use of the Feynman parameterization of Eq. (\[5:6:defh1\]). Introducing the dimensionless variables $$\label{5:6:alphaomegadef} \alpha=\frac{M_\pi}{m_N},\quad \Omega=\frac{p^2-m_N^2-M_\pi^2}{2m_N M_\pi},$$ counting as ${\cal O}(p)$ and ${\cal O}(p^0)$ \[$p^2-m_N^2={\cal O}(p)$\], respectively, we rewrite $A(z)$ as $$A(z)=m_N^2[z^2-2\alpha\Omega z(1-z)+\alpha^2(1-z)^2]\equiv m_N^2 C(z),$$ so that $H$ is now given by $$\label{5:6:defh2} H(p^2,n)=\kappa(n) \int_0^1 dz [C(z)-i0^+]^{\frac{n}{2}-2},$$ where $$\label{5:6:kappan} \kappa(n) =\frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^{\frac{n}{2}}}m_N^{n-4}.$$ The infrared singularity originates from small values of $z$, where the function $C(z)$ goes to zero as $M_\pi\to 0$. In order to isolate the divergent part one scales the integration variable $z\equiv \alpha x$ so that the upper limit $z=1$ in Eq. (\[5:6:defh2\]) corresponds to $x=1/\alpha\to \infty$ as $M_\pi\to 0$. An integral $I$ having the same infrared singularity as $H$ is then defined which is identical to $H$ except that the upper limit is replaced by $\infty$: $$\label{5:6:Idef} I\equiv\kappa(n)\int_0^\infty dz[C(z)-i0^+]^{\frac{n}{2}-2} =\kappa(n) \alpha^{n-3}\int_0^\infty [D(x)-i0^+]^{\frac{n}{2}-2},$$ where $$D(x)=1-2\Omega x+x^2+2\alpha x(\Omega x-1)+\alpha^2 x^2.$$ (The pion mass $M_\pi$ is not sent to zero.) Accordingly, the regular part of $H$ is defined as $$\label{5:6:Rdef} R\equiv-\kappa (n)\int_1^\infty dz [C(z)-i0^+]^{\frac{n}{2}-2},$$ so that $$\label{5:6:hir} H=I+R.$$ Let us verify that the definitions of Eqs. (\[5:6:Idef\]) and (\[5:6:Rdef\]) indeed reproduce the behavior of Eq. (\[5:6:defhthr\]). To that end we make use of $\Omega_{\rm thr}=1$, yielding $$\label{5:6:ithr1} I_{\rm thr}=\kappa(n)\alpha^{n-3}\int_0^\infty dx \left\{[(1+\alpha)x-1]^2-i0^+ \right\}^{\frac{n}{2}-2},$$ which converges for $n<3$. In order to continue the integral to $n>3$, we write [@Becher:1999he] $$\begin{aligned} \lefteqn{\left\{[(1+\alpha)x-1]^2-i0^+\right\}^{\frac{n}{2}-2}=}\\ &&=\frac{(1+\alpha)x-1}{(1+\alpha)(n-4)}\frac{d}{dx} \left\{[(1+\alpha)x-1]^2-i0^+\right\}^{\frac{n}{2}-2},\end{aligned}$$ and make use of a partial integration $$\begin{aligned} \lefteqn{\int_0^\infty dx \left\{[(1+\alpha)x-1]^2-i0^+\right\}^{\frac{n}{2}-2} =}\\ &&\left[\frac{(1+\alpha)x-1}{(1+\alpha)(n-4)} \left\{[(1+\alpha)x-1]^2-i0^+\right\}^{\frac{n}{2}-2} \right]_0^\infty\\ &&-\frac{1}{n-4} \int_0^\infty dx \left\{[(1+\alpha)x-1]^2-i0^+\right\}^{\frac{n}{2}-2}.\end{aligned}$$ For $n<3$, the first expression vanishes at the upper limit and, at the lower limit, yields $1/[(1+\alpha)(n-4)]$. Bringing the second expression to the left-hand side, we may then continue the integral analytically as $$\label{5:6:intanc} \int_0^\infty dx\left\{[(1+\alpha)x-1]^2-i0^+ \right\}^{\frac{n}{2}-2}=\frac{1}{(n-3)(1+\alpha)},$$ so that we obtain for $I_{\rm thr}$ $$\label{5:6:ithr} I_{\rm thr}=\kappa(n)\alpha^{n-3}\frac{1}{(n-3)(1+\alpha)} = \frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^{\frac{n}{2}}(n-3)} \frac{M_\pi^{n-3}}{m_N+M_\pi},$$ which agrees with the infrared singular part $I$ of Eq. (\[5:6:defhthr\]). The threshold value of the regular part of Eq. (\[5:6:Rdef\]) is obtained by analytic continuation from $n<3$ to $n>3$: $$\begin{aligned} \label{5:6:Rthr} R_{\rm thr}&=&-\frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^{\frac{n}{2}}} \int_1^\infty [z(m_N+M_\pi)-M_\pi]^{n-4}\nonumber\\ &=& -\frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^{\frac{n}{2}}} \frac{1}{(n-3)(m_N+M_\pi)}(\infty^{n-3}-m_N^{n-3})\nonumber\\ &\stackrel{\mbox{$n<3$}}{=}&\frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^{\frac{n}{2}}(n-3)} \frac{m_N^{n-3}}{m_N+M_\pi},\end{aligned}$$ which is indeed the regular part $R$ of Eq. (\[5:6:defhthr\]). What distinguishes $I$ from $R$ is that, for non-integer values of $n$, the chiral expansion of $I$ gives rise to non-integer powers of ${\cal O}(p)$, whereas the regular part $R$ may be expanded in an ordinary Taylor series. For the threshold integral, this can nicely be seen by expanding $I_{\rm thr}$ and $R_{\rm thr}$ in the pion mass counting as ${\cal O}(p)$. On the other hand, it is the regular part which does not satisfy the counting rules valid at tree level. The basic idea of the infrared regularization consists of replacing the general integral $H$ of Eq. (\[5:6:defh\]) by its infrared singular part $I$, defined in Eq. (\[5:6:Idef\]), and dropping the regular part $R$, defined in Eq. (\[5:6:Rdef\]). In the low-energy region $H$ and $I$ have the same analytic properties whereas the contribution of $R$, which is of the type of an infinite series in the momenta, can be included by adjusting the coefficients of the most general effective Lagrangian. As discussed in detail in Ref. [@Becher:1999he], the method can be generalized to an arbitrary one-loop graph. Using techniques similar to those of Appendices \[app\_subsec\_ipipi\] and \[subsec\_jpin\], it is first argued that tensor integrals involving an expression of the type $k^{\mu_1}\cdots k^{\mu_2}$ in the numerator may always be reduced to scalar loop integrals of the form $$-i\int\frac{d^n k}{(2\pi)^n}\frac{1}{a_1\cdots a_m}\frac{1}{b_1\cdots b_n},$$ where $a_i=(q_i+k)^2-M_\pi^2+i0^+$ and $b_i=(p_i-k)^2-m_N^2+i0^+$ are inverse meson and nucleon propagators, respectively. Here, the $q_i$ refer to four-momenta of ${\cal O}(p)$ and the $p_i$ are four-momenta which are not far off the nucleon mass shell, i.e., $p_i^2=m_N^2+{\cal O}(p)$. Using the Feynman parameterization, all pion propagators and all nucleon propagators are separately combined, and the result is written in such a way that it is obtained by applying $(m-1)$ and $(n-1)$ partial derivatives with respect to $M_\pi^2$ and $m_N^2$, respectively, to a master formula. A simple illustration is given by $$\frac{1}{a_1 a_2}=\int_0^1 dz \frac{1}{[a_1 z+a_2(1-z)]^2} =\frac{\partial}{\partial M_\pi^2}\int_0^1 dz \frac{1}{a_1 z+a_2 (1-z)},$$ where $a_i=(q_i+k)^2-M_\pi^2+i0^+$. Of course, the expressions become more complicated for larger numbers of propagators. The relevant property of the above procedure is that the result of combining the meson propagators is of the type $1/A$ with $A=(k+q)^2-M_\pi^2+i0^+$, where $q$ is a linear combination of the $m$ momenta $q_i$, with an analogous expression $1/B$ for the nucleon propagators. Finally, the expression $$-i\int\frac{d^n k}{(2\pi)^n}\frac{1}{AB}$$ may then be treated in complete analogy to $H$ of Eq. (\[5:6:defh\]), i.e., the denominators are combined as in Eq. (\[5:6:feynmanparh\]), and the infrared singular and regular pieces are identified by writing $\int_0^1 dz\cdots = \int_0^\infty dz \cdots-\int_1^\infty dz \cdots$. A crucial question is whether the infrared regularization respects the constraints of chiral symmetry as expressed through the chiral Ward identities. The argument given in Ref. [@Becher:1999he] that this is indeed the case is as follows. The total nucleon-to-nucleon transition amplitude of Eq. (\[4:btbta\]) is chirally symmetric, i.e., invariant under a simultaneous local transformation of the quark fields and the external fields (see Appendix \[app\_gfwi\] for an illustration). In terms of the effective theory, the contribution from all the tree-level diagrams is chirally symmetric so that the loop contribution must also be chirally symmetric. Since we work in dimensional regularization this statement holds for an arbitrary $n$. However, as we have seen in the example of Eq. (\[5:6:defhthr\]), the separation into infrared singular and regular parts amounts to distinguishing between contributions of non-integer and non-negative integer powers in the momentum expansion. Since these powers do not mix for arbitrary $n$, the infrared singular and regular parts must be separately chirally symmetric. Finally, the regular part can be expanded in powers of either momenta or quark masses, and thus may as well be absorbed in the (modified) tree-level contribution. Let us finally establish the connection between the infrared singular part $I$ and the corresponding result in HBChPT. To that end, we first consider the relativistic propagator by expressing the (off-shell) four-momentum as $p=m_N v+r$, $$\begin{aligned} \label{5:6:relprophb} \lefteqn{\frac{i}{p\hspace{-.5em}/\hspace{.2em}-m_N+i0^+}= i\frac{p\hspace{-.5em}/\hspace{.2em}+m_N}{p^2-m_N^2+i0^+} =i\frac{p\hspace{-.5em}/\hspace{.2em}+m_N}{2m_N v\cdot r+ r^2+i0^+}} \nonumber\\ &=&i\frac{p\hspace{-.5em}/\hspace{.2em}+m_N}{2m_N v\cdot r+i0^+} \,\frac{1}{1+\frac{r^2}{2m_N v\cdot r+i0^+}}\nonumber\\ &\mapsto&\frac{p\hspace{-.5em}/\hspace{.2em}+m_N}{2m_N} \frac{i}{v\cdot r+i0^+}\left[1+\frac{ir^2}{2m_N}\frac{i}{v\cdot r+i0^+}+ \left(\frac{ir^2}{2m_N}\frac{i}{v\cdot r+i0^+}\right)^2+\cdots\right].\nonumber \\\end{aligned}$$ In the last step, we have assumed that $r$ is small enough to allow for an expansion in terms of a geometric series. The result of Eq. (\[5:6:relprophb\]) is displayed in Fig. \[5:6:fig:relprophb\] and may be interpreted as an infinite series in terms of the heavy-baryon propagator $i/(v\cdot r+i0^+)$ and the self-energy insertion $-i\Sigma=i r^2/2m_N$ which has the form of a non-relativistic kinetic energy. (Note that the expression still involves the operator $(p\hspace{-.5em}/\hspace{.2em}+m_N)/2m_N$.) Let us apply Eq. (\[5:6:relprophb\]) to the loop integral $H$ of Eq. (\[5:6:defh\]) by first expanding the integrand and then performing the summation. This corresponds to the prescription proposed in Refs.  [@Tang:1996ca; @Ellis:1997kc] for identifying the low-energy or, in the nomenclature of Ref. [@Tang:1996ca], soft contribution to a Feynman graph. In the case at hand we obtain $$H(p^2,n)\to\sum_{j=0}^\infty I_j,$$ where $$\label{5:6:ijdef} I_j=\frac{-i}{2m_N}\int\frac{d^n k}{(2\pi)^n}\frac{1}{k^2-M_\pi^2+i0^+} \,\frac{1}{v\cdot (r-k)+i0^+} \left[\frac{-(r-k)^2}{2m_N v\cdot(r-k)+i0^+}\right]^j,$$ which is somewhat easier to handle if we perform the shift $k\to k+r$ and then the substitution $k\to -k$, $$\label{5:6:ijdef2} I_j=i\frac{(-)^{j+1}}{(2m_N)^{j+1}}\int\frac{d^n k}{(2\pi)^n} \frac{1}{(k-r)^2-M_\pi^2+i0^+} \,\frac{(k^2)^j}{(v\cdot k+i0^+)^{j+1}}.$$ As above, we explicitly discuss the threshold $p^2_{\rm thr}=(m_N+M_\pi)^2$ by inserting $r=M_\pi v$ into Eq. (\[5:6:ijdef2\]). Defining $X=k^2-2M_\pi v\cdot k+i0^+$ and $Y=v\cdot k+i0^+$, we have $$\label{5:6:ijthr} I_{j,\rm thr}=i \frac{(-)^{j+1}}{(2m_N)^{j+1}}\int\frac{d^n k}{(2\pi)^n} \frac{(X+2M_\pi Y)^j}{X Y^{j+1}}.$$ The different $I_{j,\rm thr}$ are related by a simple recursion relation $$\label{5:6:ijp1jthr} I_{j+1,\rm thr}=-\frac{M_\pi}{m_N} I_{j,\rm thr},\quad j\geq 0,$$ implying $$I_{j,\rm thr}=(-)^j \left(\frac{M_\pi}{m_N}\right)^j I_{0,\rm thr}, \quad j\geq 0.$$ Equation (\[5:6:ijp1jthr\]) is easily verified: $$\begin{aligned} I_{j+1,\rm thr}&=&i\frac{(-)^{j+2}}{(2m_N)^{j+2}} \int\frac{d^n k}{(2\pi)^n} \frac{(X+2M_\pi Y)^j}{X Y^{j+1}}\, \frac{X+2M_\pi Y}{Y}\\ &=&i\frac{(-)^{j+2}}{(2m_N)^{j+2}}\int \frac{d^n k}{(2\pi)^n} \frac{(k^2)^j}{(v\cdot k+i0^+)^{j+2}}\\ &&-\frac{M_\pi}{m_N}i\frac{(-1)^{j+1}}{(2m_N)^{j+1}}\int \frac{d^n k}{(2\pi)^n}\frac{(X+2M_\pi Y)^j}{X Y^{j+1}}\\ &=&-\frac{M_\pi}{m_N}I_{j,\rm thr},\end{aligned}$$ where we made use of the fact that the first term in the second line vanishes in dimensional regularization \[see Eq. (\[app:dnkk2pvkmq\])\]. We then obtain for the series, evaluated at threshold, $$\sum_{j=0}^\infty I_{j,\rm thr}=I_{0,\rm thr}\sum_{j=0}^{\infty} (-)^j \left(\frac{M_\pi}{m_N}\right)^j =\frac{m_N}{m_N+M_\pi} I_{0,\rm thr}.$$ What remains to be determined is the threshold integral $$I_{0,{\rm thr}}=\frac{-i}{2m_N} \int\frac{d^n k}{(2\pi)^n}\frac{1}{k^2-2M_\pi v\cdot k+i0^+} \frac{1}{v\cdot k+i0^+}.$$ Performing a shift $k\to k+M_\pi v$, combining the denominators as in Eq. (\[app:feynmantricksphb\]), performing another shift $k\to k-yv$, and making use of Eq. (\[app:drb:moregenint\]), one finds $$I_{0,{\rm thr}}=\frac{1}{m_N} \frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^\frac{n}{2}} \int_0^\infty dy \left\{(y-M_\pi)^2-i0^+\right\}^{\frac{n}{2}-2}.$$ Finally, performing a substitution $y=M_\pi x$ and using the analytic continuation of Eq. (\[5:6:intanc\]) with $\alpha=0$, we obtain $$\label{5:6:I0thr} I_{0,{\rm thr}}= \frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^\frac{n}{2} (n-3)}\frac{M_\pi^{n-3}}{m_N}.$$ Inserting Eq. (\[5:6:I0thr\]) into the series, the final result reads $$\label{5:6:series} \sum_{j=0}^\infty I_{j,\rm thr} =\frac{\Gamma\left(2-\frac{n}{2}\right)}{(4\pi)^\frac{n}{2}(n-3)} \frac{M_\pi^{n-3}}{m_N+M_\pi},$$ which is the same as $I_{\rm thr}$ of Eq. (\[5:6:ithr\]). This example shows that the infrared regularized amplitude is related to an [*infinite*]{} sum of heavy-baryon amplitudes with self-energy insertions in the heavy-baryon propagator, as depicted in Fig. \[5:6:fig:relprophb\]. The advantage of the relativistic approach is obvious, because for a general one-loop amplitude it may be very difficult, if not impossible, to obtain a closed expression for the sum of all insertions. To conclude this section, the method of infrared regularization provides a fully relativistic framework producing amplitudes having the relevant analytic properties and satisfying the chiral power-counting rules. At the moment, it is not yet clear, whether it can be generalized beyond the one-loop level. Summary and Concluding Remarks {#chap_cr} ============================== As we have discussed in great detail, the chiral $\mbox{SU(3)}_L\times\mbox{SU(3)}_R\times\mbox{U(1)}_V$ symmetry of QCD in the limit of vanishing $u$-, $d$-, and $s$-quark masses (Sec. \[subsec\_gsclqs\]), together with the assumption of its spontaneous breakdown to $\mbox{SU(3)}_V\times\mbox{U(1)}_V$ in the ground state (Sec.  \[sec\_ssbqcd\]), is one of the keys to understanding the phenomenology of the strong interactions in the low-energy regime. The importance of chiral symmetry was realized long before the formulation of QCD and led to a host of predictions within the current-algebra and PCAC approaches of the 1960’s [@Adler:1968]. Some of the consequences of an explicit symmetry breaking, yielding non-analytic terms in the perturbation, were worked out in the early 1970’s, but the development came to a halt [@Pagels:se] because it was not clear how to systematically organize a perturbative expansion. From the present point of view, the explicit symmetry breaking is due to the finite $u$-, $d$-, and $s$-quark masses, leading to divergences of the symmetry currents (Sec. \[subsec\_csbdqm\]). In 1979 Weinberg [@Weinberg:1978kz] laid the foundations for further progress with his observation that the constraints due to (chiral) symmetry may perturbatively be analyzed in terms of the most general effective field theory. A very important ingredient was the formulation of a consistent power-counting scheme (Secs. \[sec\_elwpcs\] and \[subsec\_pcs\]) which allowed for a systematic perturbative analysis in contrast to various commonly used [*ad hoc*]{} phenomenological approaches to the strong interactions at low energies. In particular, the inclusion of loop diagrams allowed for a perturbative restoration of unitarity which would be violated if only tree-level diagrams were used. Subsequently Gasser and Leutwyler [@Gasser:1983yg; @Gasser:1984gg] combined the ideas of Weinberg with other modern techniques of quantum field theory to analyze the Ward identities of QCD Green functions in terms of a local invariance of the generating functional under the chiral group (Sec. \[sec\_gfcwi\] and App. \[app\_gfwi\]). These papers were the starting point of what is nowadays called chiral perturbation theory. The mesonic sector has generated a host of successful applications, some of which have reached two-loop accuracy. Here, we have concentrated on a few elementary observables and processes, namely: masses of the Goldstone bosons (Secs. \[sec\_loel\] and \[subsec\_mgb\]), weak and electromagnetic $\pi$ decays (Secs. \[subsec\_pdpmn\] and \[sec\_ewzwa\]), $\pi\pi$ scattering (Secs. \[subsec\_pps\] and \[subsec\_eppsop6\]), and electromagnetic form factors (Sec. \[subsec\_emffp\]). Moreover, we have discussed in quite some detail how to construct the mesonic effective Lagrangian (Secs. \[sec\_tpgb\], \[sec\_clop4\], and \[subsec\_mclop6\]). At first sight, it might appear that the large number of low-energy parameters at ${\cal O}(p^6)$ would make any quantitative prediction at the two-loop level impossible. However, there are several reasons why this is not the case. To start with, there exist observables which do not depend on [*any*]{} new parameters at ${\cal O}(p^6)$, i.e., which can be predicted in terms of the ${\cal O}(p^2)$ and ${\cal O}(p^4)$ low-energy constants only. An example is given by the correction to Sirlin’s theorem discussed in Ref. [@Post:1997dk]. Clearly, such cases provide a natural testing ground for the convergence of the approach. Secondly, only a limited set of low-energy parameters contribute to any given process. It follows from the nature of the Ward identities that different physical processes are interrelated due to the underlying symmetries so that coefficients which have been fixed using one reaction can be used to [*predict*]{} another observable. In view of the ordinary implementation of symmetries, such as in the Wigner-Eckart theorem, this is not a surprise, because it is well-known that symmetries imply relations among $S$-matrix elements. However, the Ward identities provide [*additional*]{} constraints among Green functions of a different type and allow one to also include an explicit symmetry breaking (Sec. \[sec\_gfcwi\]). It is this second case which can systematically be studied in the framework of ChPT and which provides interesting new insights into our understanding of both spontaneous and explicit symmetry breaking within QCD. Finally, different methods exist which allow one to estimate the value of the parameters and thus, in combination with the ChPT result, test our physical picture of the strong interactions. In this work we have only considered elementary processes at an introductory level, not the extensions to and combinations with other methods. We omitted, for example, the weak interactions of kaons which are mediated by the exchange of $W$ bosons [*between*]{} the quark currents [@Ecker:1995gg; @deRafael:1995zv; @Pich:1995bw]. We also did not discuss the breaking of isospin symmetry which requires the inclusion of the electromagnetic interaction in terms of dynamical (virtual) photons [@Urech:1994hd; @Neufeld:1995mu; @Ananthanarayan:2002kj]. Chiral symmetry also dictates the interaction of the Goldstone bosons with other hadrons (Secs. \[sec\_tpf\] and \[sec\_loebl\]). By studying the axial-vector current matrix element (Sec. \[subsec\_gtravcme\]) and $\pi N$ scattering (Sec. \[subsec\_apnstl\]) we verified that a tree-level calculation using the lowest-order Lagrangian reproduces the Goldberger-Treiman relation and the Weinberg-Tomozawa result for the $s$-wave scattering lengths, respectively. As we have seen, the first systematic study in the pion-nucleon sector [@Gasser:1987rb] raised the question of a consistent power counting (Sec. \[sec\_eld\]). This problem was subsequently overcome in the framework of the heavy-baryon approach [@Jenkins:1990jv] (Sec. \[sec\_hbf\]) and most of the numerous applications in this sector have been performed in HBChPT [@Bernard:1995dp]. In the baryonic sector the chiral orders increase in units of one, because of the additional possibility of forming Lorentz invariants by contracting (covariant) derivatives with gamma matrices (Sec. \[sec\_loebl\]). As a result, in the SU(2)$\times$SU(2) baryonic sector at the one-loop level, up to and including ${\cal O}(p^4)$, one has in total $2+7+23+118=150$ [@Fettes:2000gb] low-energy constants as opposed to the $2+7=9$ free parameters of the corresponding mesonic sector [@Gasser:1983yg]. Nevertheless, numerous results have been obtained in the baryonic sector because, at the same time, a large amount of very precise experimental data are available due to the existence of a stable proton target. (Neutron data can also be extracted, e.g, from experiments on the deuteron.) The availability of new high-precision data in combination with the techniques of chiral perturbation theory have led to a considerable improvement of our understanding of the strong interactions at low energies, in particular since systematic corrections to the old current-algebra predictions could be worked out and (successfully) tested. We have not discussed the approach of the so-called small scale expansion to include the $\Delta(1232)$ resonance as an explicit degree of freedom [@Hemmert:1996xg]. Clearly this is an important issue in the baryonic sector because the first nucleon excitation is such a prominent feature of the low-energy spectrum as seen, e.g., in the total pion-nucleon scattering or the total photo absorption cross sections. Most recently, the method of infrared regularization [@Becher:1999he] has opened the possibility of reconciling the relativistic approach with a consistent chiral power counting scheme (Sec. \[sec\_mir\]). One may expect that this method will have a large impact insofar as many of the results obtained within the heavy-baryon framework will have to be checked with respect to relativistic corrections. The question regarding the radius of convergence in the baryonic sector remains a big challenge because, ultimately, a calculation at the two-loop level ${\cal O}(p^5)$ is, in general, required to quantitatively assess higher-order corrections [@McGovern:1998tm]. In comparison to two-loop calculations in the SU(2)$\times$SU(2) mesonic sector such an investigation in the (relativistic) nucleon sector is even more complicated for two reasons. First, due to the spin of the nucleon, the structure of vertices is richer than for spin-0 particles. Second, the nucleon mass introduces another mass in the propagators making the evaluation of the two-loop integrals more difficult than for a single mass. Finally, we would like to mention that a description of the nucleon-nucleon interaction within the framework of effective field theory has made tremendous progress and that a rigorous treatment of nuclei within field theory is no longer out of reach . In conclusion, chiral perturbation theory has added a new and unprecedented level of systematics to the description of strong-interaction processes at low energies and continues to be a very fruitful and rich field with promising perspectives. If this introductory review encourages students and newcomers to chiral perturbation theory to participate in this field of research, it has served its purpose. Acknowledgments {#acknowledgments .unnumbered} --------------- I am greatly indebted to David R. Harrington for carefully and critically reading the whole manuscript (!) and his uncountably numerous suggestions for improvement. He continuously forced me to explain the meaning of concepts instead of hiding behind the chiral jargon. I would like to thank my academic teachers Dieter Drechsel, Harold W. Fearing, and Justus H. Koch for sharing their deep insights into theoretical physics with me. Learning from them has been a pleasure! Numerous discussions with my collaborators and colleagues , , R. , R. , , , , , and are gratefully acknowledged. Special thanks go to Rolf Brockmann for many discussions on effective field theory, to Jambul Gegelia for his valuable discussions on the infrared regularization, to Martin Reuter for his kind help in preparing Appendix A on Green functions and Ward identities, and to Thomas Walcher for challenging discussions on spontaneous symmetry breaking. Last but not least, I would like to thank Erich Vogt for his incredible enthusiasm and his continuous encouragement to finish the manuscript. I dedicate this work to my family. I hope it was worth it! Green Functions and Ward Identities {#app_gfwi} =================================== In this appendix we will show how to derive Ward identities for Green functions in the framework of canonical quantization on the one hand, and quantization via the Feynman path integral on the other hand, by means of an explicit example. In order to keep the discussion transparent, we will concentrate on a simple scalar field theory with a global O(2) or U(1) invariance. To that end, let us consider the Lagrangian $$\begin{aligned} \label{app:gfwi:lphi4} {\cal L}&=&\frac{1}{2}(\partial_\mu \Phi_1\partial^\mu \Phi_1 +\partial_\mu \Phi_2\partial^\mu \Phi_2) -\frac{m^2}{2} (\Phi_1^2+\Phi_2^2) -\frac{\lambda}{4}(\Phi_1^2+\Phi_2^2)^2\nonumber\\ &=& \partial_\mu \Phi^\dagger \partial^\mu \Phi -m^2 \Phi^\dagger\Phi -\lambda (\Phi^\dagger\Phi)^2,\end{aligned}$$ where $$\Phi(x)=\frac{1}{\sqrt{2}}[\Phi_1(x)+i\Phi_2(x)], \quad \Phi^\dagger(x)=\frac{1}{\sqrt{2}}[\Phi_1(x)-i\Phi_2(x)],$$ with real scalar fields $\Phi_1$ and $\Phi_2$. Furthermore, we assume $m^2>0$ and $\lambda>0$, so there is no spontaneous symmetry breaking and the energy is bounded from below. Equation (\[app:gfwi:lphi4\]) is invariant under the global (or rigid) transformations $$\label{app:gfwi:inftrans1} \Phi'_1=\Phi_1-\epsilon \Phi_2,\quad \Phi'_2=\Phi_2+\epsilon \Phi_1,$$ or, equivalently, $$\label{app:gfwi:inftrans2} \Phi'=(1+i\epsilon)\Phi,\quad \Phi'^\dagger=(1-i\epsilon)\Phi^\dagger,$$ where $\epsilon$ is an infinitesimal real parameter. Applying the method of Gell-Mann and L[é]{}vy [@Gell-Mann:np], we obtain for a [*local*]{} parameter $\epsilon(x)$, $$\label{app:gfwi:dlphi4} \delta{\cal L}=\partial_\mu\epsilon(x)(i\partial^\mu\Phi^\dagger \Phi -i\Phi^\dagger\partial^\mu\Phi),$$ from which, via Eqs. (\[2:3:strom2\]) and (\[2:3:divergenz\]), we derive for the current corresponding to the global symmetry, $$\begin{aligned} J^{\mu}&=&\frac{\partial \delta\cal L}{\partial \partial_\mu \epsilon}=(i\partial^\mu\Phi^\dagger \Phi -i\Phi^\dagger\partial^\mu\Phi),\\ \partial_\mu J^{\mu}&=&\frac{\partial \delta\cal L}{\partial \epsilon}=0.\end{aligned}$$ Recall that the identification of Eq. (\[2:3:divergenz\]) as the divergence of the current is only true for fields satisfying the Euler-Lagrange equations of motion. We now extend the analysis to a [*quantum*]{} field theory. In the framework of canonical quantization, we first define conjugate momenta, $$\label{app:gfwi:conjmom} \Pi_i(x)=\frac{\partial \cal L}{\partial \partial_0\Phi_i},\quad \Pi(x)=\frac{\partial \cal L}{\partial \partial_0\Phi},\quad \Pi^\dagger(x)=\frac{\partial \cal L}{\partial \partial_0\Phi^\dagger},$$ and interpret the fields and their conjugate momenta as operators which, in the Heisenberg picture, are subject to the equal-time commutation relations $$[\Phi_i(\vec{x},t),\Pi_j(\vec{y},t)]=i\delta_{ij}\delta^3(\vec{x}-\vec{y}),$$ and $$\label{app:gfwi:eqtcr} {[}\Phi(\vec{x},t),\Pi(\vec{y},t)]= [\Phi^\dagger(\vec{x},t),\Pi^\dagger(\vec{y},t)] =i\delta^3(\vec{x}-\vec{y}).$$ The remaining equal-time commutation relations, involving fields or momenta only, vanish. For the quantized theory, the current operator then reads $$J^\mu(x)=:(i\partial^\mu\Phi^\dagger \Phi -i\Phi^\dagger\partial^\mu\Phi):,$$ where $:\quad :$ denotes normal or Wick ordering, i.e., annihilation operators appear to the right of creation operators. For a conserved current, the charge operator, i.e., the space integral of the charge density, is time independent and serves as the generator of infinitesimal transformations of the Hilbert space states, $$Q=\int d^3 x J^0(\vec{x},t).$$ Applying Eq. (\[app:gfwi:eqtcr\]), it is straightforward to calculate the equal-time commutation relations[^119] $$\begin{aligned} \label{app:gfwi:j0comrel} [J^0(\vec{x},t),\Phi(\vec{y},t)]&=&\delta^3(\vec{x}-\vec{y})\Phi(\vec{x},t), \nonumber\\ {[}J^0(\vec{x},t),\Pi(\vec{y},t)]&=&-\delta^3(\vec{x}-\vec{y}) \Pi(\vec{x},t),\nonumber\\ {[}J^0(\vec{x},t),\Phi^\dagger(\vec{y},t)]&=&-\delta^3(\vec{x}-\vec{y}) \Phi^\dagger(\vec{x},t),\nonumber\\ {[}J^0(\vec{x},t),\Pi^\dagger(\vec{y},t)]&=&\delta^3(\vec{x}-\vec{y}) \Pi^\dagger(\vec{x},t).\end{aligned}$$ In particular, performing the space integrals in Eqs.  (\[app:gfwi:j0comrel\]), one obtains $$\begin{aligned} \label{app:gfwi:Qcomrel} [Q,\Phi(x)]&=&\Phi(x), \nonumber\\ {[}Q,\Pi(x)]&=&-\Pi(x),\nonumber\\ {[}Q,\Phi^\dagger(x)]&=&-\Phi^\dagger(x),\nonumber\\ {[}Q,\Pi^\dagger(x)]&=&\Pi^\dagger(x).\end{aligned}$$ In order to illustrate the implications of Eqs. (\[app:gfwi:Qcomrel\]), let us take an eigenstate $|\alpha\rangle$ of $Q$ with eigenvalue $q_\alpha$ and consider, for example, the action of $\Phi(x)$ on that state, $$\begin{aligned} Q\left(\Phi(x)|\alpha\rangle\right)=\left([Q,\Phi(x)]+\Phi(x)Q\right) |\alpha\rangle =(1+q_\alpha)\left(\Phi(x)|\alpha\rangle\right).\end{aligned}$$ We conclude that the operators $\Phi(x)$ and $\Pi^\dagger(x)$ \[$\Phi^\dagger(x)$ and $\Pi(x)$\] increase (decrease) the Noether charge of a system by one unit. We are now in the position to discuss the consequences of the U(1) symmetry of Eq. (\[app:gfwi:lphi4\]) for the Green functions of the theory. To that end, let us consider as our prototype the Green function $$\label{app:gfwi:gmuxyz} G^\mu(x,y,z)=\langle 0|T[\Phi(x) J^\mu(y) \Phi^\dagger(z)]|0\rangle,$$ which describes the transition amplitude for the creation of a quantum of Noether charge $+1$ at $x$, propagation to $y$, interaction at $y$ via the current operator, propagation to $z$ with annihilation at $z$. First of all we observe that under the global infinitesimal transformations of Eq. (\[app:gfwi:inftrans2\]), $J^\mu(x)\mapsto J'^\mu(x)=J^\mu(x)$, or in other words $[Q,J^\mu(x)]=0$. We thus obtain $$\begin{aligned} \label{app:gfwi:gmutrans} G^\mu(x,y,z)\mapsto G'^\mu(x,y,z)&=& \langle 0|T[ (1+i\epsilon)\Phi(x) J'^\mu(y) (1-i\epsilon)\Phi^\dagger(z)]|0\rangle \nonumber\\ &=&\langle 0|T[\Phi(x) J^\mu(y) \Phi^\dagger(z)]|0\rangle\nonumber\\ &=&G^\mu(x,y,z),\end{aligned}$$ the Green function remaining invariant under the U(1) transformation. (In general, the transformation behavior of a Green function depends on the irreducible representations under which the fields transform. In particular, for more complicated groups such as SU($N$), standard tensor methods of group theory may be applied to reduce the product representations into irreducible components [@Balachandran:ab; @O'Raifeartaigh:vq; @Jones:ti]. We also note that for U(1), the symmetry current is charge neutral, i.e. invariant, which for more complicated groups, in general, is not the case.) Moreover, since $J^\mu(x)$ is the Noether current of the underlying U(1) there are further restrictions on the Green function beyond its transformation behavior under the group. In order to see this, we consider the divergence of Eq. (\[app:gfwi:gmuxyz\]) and apply the equal-time commutation relations of Eqs.  (\[app:gfwi:j0comrel\]) to obtain (see Sec. \[subsec\_cgf\]) $$\begin{aligned} \label{app:gfwi:gmuwt} \partial_\mu^y G^\mu(x,y,z)&=&[\delta^4(x-y)-\delta^4(z-y)] \langle 0|T[\Phi(x)\Phi^\dagger(z)]|0\rangle,\end{aligned}$$ where we made use of $\partial_\mu J^\mu=0$. Equation (\[app:gfwi:gmuwt\]) is the analogue of the Ward identity of QED \[see Eq. (\[2:4:qedwardidentity\])\] [@Ward:1950xp; @Fradkin:1955jr; @Takahashi:xn]. In other words, the underlying symmetry not only determines the transformation behavior of Green functions under the group, but also relates $n$-point Green functions containing a symmetry current to $(n-1)$-point Green functions \[see Eq. (\[2:4:gendmug\])\]. In principle, calculations similar to those leading to Eqs.(\[app:gfwi:gmutrans\]) and (\[app:gfwi:gmuwt\]), can be performed for any Green function of the theory. However, we will now show that the symmetry constraints can be compactly summarized in terms of an invariance property of a generating functional. The generating functional is defined as the vacuum-to-vacuum transition amplitude in the presence of external fields, $$\begin{aligned} \label{app:gfwi:genfunc1} \lefteqn{ W[j,j^\ast,j_\mu]=\langle 0,+\infty|0,-\infty\rangle_{j,j^\ast,j_\mu}} \nonumber\\ &=&\exp(iZ[j,j^\ast,j_\mu])\nonumber\\ &=&\langle 0|T\left(\exp\left\{i\int d^4x[j(x)\Phi^\dagger(x)+j^\ast(x) \Phi(x) +j_\mu(x) J^\mu(x)]\right\}\right)|0\rangle,\nonumber\\\end{aligned}$$ where $\Phi$ and $\Phi^\dagger$ are the field operators and $J^\mu(x)$ is the Noether current. Note that the field operators and the conjugate momenta are subject to the equal-time commutation relations and, in addition, must satisfy the Heisenberg equations of motion. Via this second condition and implicitly through the ground state, the generating functional depends on the dynamics of the system which is determined by the Lagrangian of Eq. (\[app:gfwi:lphi4\]). The Green functions of the theory involving $\Phi$, $\Phi^\dagger$, and $J^\mu$ are obtained through functional derivatives of Eq. (\[app:gfwi:genfunc1\]). For example, the Green function of Eq. (\[app:gfwi:gmuxyz\]) is given by $$\label{app:gfwi:gmuxyz1} G^\mu(x,y,z)=(-i)^3 \left.\frac{\delta^3 W[j,j^\ast,j_\mu]}{\delta j^\ast(x) \delta j_\mu(y) \delta j(z)}\right|_{j=0,j^\ast=0,j_\mu=0}.$$ In order to discuss the constraints imposed on the generating functional via the underlying symmetry of the theory, let us consider its path integral representation [@Zinn-Justin:1989mi; @Das:1993gd],[^120] $$\begin{aligned} \label{app:gfwi:genfunc2} W[j,j^\ast,j_\mu]&=&\int [d\Phi_1][d\Phi_2] e^{iS[\Phi,\Phi^\ast,j,j^\ast, j_\mu]},\end{aligned}$$ where $$\label{app:gfwi:actionsources} S[\Phi,\Phi^\ast,j,j^\ast,j_\mu]=S[\Phi,\Phi^\ast]+ \int d^4 x [\Phi(x)j^\ast(x)+\Phi^\ast(x) j(x)+J^\mu(x) j_\mu(x)]$$ denotes the action corresponding to the Lagrangian of Eq. (\[app:gfwi:lphi4\]) in combination with a coupling to the external sources. Let us now consider a [*local*]{} infinitesimal transformation of the fields \[see Eqs. (\[app:gfwi:inftrans2\])\] together with a [*simultaneous*]{} transformation of the external sources, $$\label{app:gfwi:inftranssources} j'(x)=[1+i\epsilon(x)]j(x),\quad j'^\ast(x)=[1-i\epsilon(x)] j^\ast(x),\quad j_\mu'(x)=j_\mu(x)-\partial_\mu\epsilon(x).$$ The action of Eq. (\[app:gfwi:actionsources\]) remains invariant under such a transformation, $$S[\Phi',\Phi'^\ast,j',j'^\ast,j_\mu']= S[\Phi,\Phi^\ast,j,j^\ast,j_\mu].$$ We stress that the transformation of the external current $j_\mu$ is necessary to cancel a term resulting from the kinetic term in the Lagrangian. We can now verify the invariance of the generating functional as follows, $$\begin{aligned} \label{app:gfwi:winvariance} W[j,j^\ast,j_\mu]&=&\int [d\Phi_1][d\Phi_2] e^{iS[\Phi,\Phi^\ast,j,j^\ast, j_\mu]}\nonumber\\ &=&\int [d\Phi_1][d\Phi_2] e^{iS[\Phi',\Phi'^\ast,j',j'^\ast, j'_\mu]}\nonumber\\ &=&\int [d\Phi_1'][d\Phi_2'] \left|\left(\frac{\partial\Phi_i}{\partial \Phi_j'}\right)\right| e^{iS[\Phi',\Phi'^\ast,j',j'^\ast, j'_\mu]}\nonumber\\ &=&\int [d\Phi_1][d\Phi_2] e^{iS[\Phi,\Phi^\ast,j',j'^\ast, j'_\mu]}\nonumber\\ &=&W[j',j'^\ast,j'_\mu].\end{aligned}$$ We made use of the fact that the Jacobi determinant is one and renamed the integration variables. In other words, given the [*global*]{} U(1) symmetry of the Lagrangian, Eq. (\[app:gfwi:lphi4\]), the generating functional is invariant under the [*local*]{} transformations of Eq. (\[app:gfwi:inftranssources\]). It is this observation which, for the more general case of the chiral group SU($N$)$\times$SU($N$), was used by Gasser and Leutwyler as the starting point of chiral perturbation theory. We still have to discuss, how this invariance allows us to collect the Ward identities in a compact formula. We start from Eq. (\[app:gfwi:winvariance\]), $$\begin{aligned} 0&=&\int[d\Phi_1] [d\Phi_2]\left(e^{iS[\Phi,\Phi^\ast,j',j'^\ast,j'_\mu]} -e^{iS[\Phi,\Phi^\ast,j,j^\ast,j_\mu]}\right)\\ &=&\int[d\Phi_1] [d\Phi_2]\int d^4x \left\{ \epsilon[\Phi j^\ast-\Phi^\ast j] -iJ^\mu\partial_\mu\epsilon\right\} e^{iS[\Phi,\Phi^\ast,j,j^\ast,j_\mu]}.\end{aligned}$$ Observe that $$\Phi(x) e^{iS[\Phi,\Phi^\ast,j,j^\ast,j_\mu]} =-i\frac{\delta}{\delta j^\ast(x)} e^{iS[\Phi,\Phi^\ast,j,j^\ast,j_\mu]},$$ and similarly for the other terms, resulting in $$\begin{aligned} 0&=&\int[d\Phi_1] [d\Phi_2]\int d^4 x\left\{ \epsilon(x)\left[-ij^\ast(x)\frac{\delta}{\delta j^\ast(x)} +ij(x)\frac{\delta}{\delta j(x)}\right]\right.\\ &&\left. -\partial_\mu\epsilon(x)\frac{\delta}{\delta j_\mu(x)}\right\} e^{iS[\Phi,\Phi^\ast,j,j^\ast,j_\mu]}.\end{aligned}$$ Finally we interchange the order of integration, make use of partial integration, and apply the divergence theorem: $$\label{app:gfwi:0fl} 0=\int d^4 x \epsilon(x)\left[i j(x)\frac{\delta}{\delta j(x)} -i j^\ast(x)\frac{\delta}{\delta j^\ast(x)} +\partial_\mu ^x \frac{\delta}{\delta j_\mu(x)}\right] W[j,j^\ast,j_\mu].$$ Since Eq. (\[app:gfwi:0fl\]) must hold for any $\epsilon(x)$ we obtain as the master equation for deriving Ward identities, $$\label{app:gfwi:mewi} \left[j(x)\frac{\delta}{\delta j(x)}- j^\ast(x)\frac{\delta}{\delta j^\ast(x)} -i\partial_\mu^x \frac{\delta}{\delta j_\mu(x)}\right] W[j,j^\ast,j_\mu] =0.$$ We note that Eqs. (\[app:gfwi:winvariance\]) and (\[app:gfwi:mewi\]) are equivalent. As a final illustration let us re-derive the Ward identity of Eq. (\[app:gfwi:gmuwt\]) using Eq. (\[app:gfwi:mewi\]). For that purpose we start from Eq. (\[app:gfwi:gmuxyz1\]), $$\partial_\mu^y G^\mu(x,y,z)= (-i)^3 \partial_\mu^y \left.\frac{\delta^3 W}{\delta j^\ast(x)\delta j_\mu (y)\delta j(z)}, \right|_{j=0,j^\ast=0,j_\mu=0},$$ apply Eq. (\[app:gfwi:mewi\]), $$=(-i)^2\left\{ \frac{\delta^2}{\delta j^\ast(x)\delta j(z)}\left[ j^\ast(y)\frac{\delta}{\delta j^\ast(y)}-j(y)\frac{\delta}{\delta j(y)} \right] W\right\}_{j=0,j^\ast=0,j_\mu=0},$$ make use of $\delta j^\ast(y)/\delta j^\ast(x)= \delta^4(y-x)$ and $\delta j(y)/\delta j(z)=\delta^4(y-z)$ for the functional derivatives, $$=(-i)^2\left\{\delta^4(x-y)\frac{\delta^2 W}{\delta j^\ast(y) \delta j(z)} -\delta^4(z-y)\frac{\delta^2 W}{\delta j^\ast(x) \delta j(y)}\right\}_{j=0,j^\ast=0,j_\mu=0},$$ and, finally, use the definition of Eq. (\[app:gfwi:genfunc1\]), $$\partial_\mu^y G^\mu(x,y,z) =[\delta^4(x-y)-\delta^4(z-y)]\langle 0| T[\Phi(x)\Phi^\dagger(z) ]|0\rangle$$ which is the same as Eq. (\[app:gfwi:gmuwt\]). In principle, any Ward identity can be obtained by taking appropriate higher functional derivatives of $W$ and then using Eq. (\[app:gfwi:mewi\]). Dimensional Regularization: Basics {#app_drb} ================================== For the sake of completeness we provide a simple illustration of the method of dimensional regularization. For a detailed account the interested reader is referred to Refs. . Let us consider the integral $$\label{app:drb:int} I=\int\frac{d^4k}{(2\pi)^4}\frac{i}{k^2-M^2+i0^+}$$ which appears in the calculation of the masses of the Goldstone bosons \[see Eq. (\[4:8:diag\])\]. We introduce $$a\equiv\sqrt{\vec{k}^2+M^2}>0$$ so that $$\begin{aligned} k^2-M^2+i0^+ &=&[k_0+(a-i0^+)][k_0-(a-i0^+)],\end{aligned}$$ and define $$f(k_0)=\frac{1}{[k_0+(a-i0^+)][k_0-(a-i0^+)]}.$$ In order to determine $\int_{-\infty}^{\infty} dk_0 f(k_0)$ as part of the calculation of $I$, we consider $f$ in the complex $k_0$ plane and make use of Cauchy’s theorem $$\label{app:drb:cauchy} \oint_C dz f(z)=0$$ for functions which are differentiable in every point inside the closed contour $C$. We choose the contour as shown in Fig. \[app:drb:wickrotation\_fig\], $$0=\sum_{i=1}^4 \int_{\gamma_i} dz f(z),$$ and make use of $$\int_\gamma f(z)dz=\int_a^b f[\gamma(t)]\gamma'(t)dt$$ to obtain for the individual integrals $$\begin{aligned} \int_{\gamma_1} f(z) dz &=& \int_{-\infty}^\infty f(t)dt,\\ \int_{\gamma_2} f(z) dz &=& \lim_{R\to\infty} \int_{0}^\frac{\pi}{2} f(Re^{it})iRe^{it}dt =0, \,\,\mbox{since}\,\, \lim_{R\to\infty} \underbrace{Rf(Re^{it})}_{\mbox{$\sim \frac{1}{R}$}}=0,\\ \int_{\gamma_3} f(z) dz &=& \int_{\infty}^{-\infty} f(it)idt,\\ \int_{\gamma_4} f(z) dz &=& \lim_{R\to\infty} \int_{\frac{3}{2}\pi}^\pi f(Re^{it})iRe^{it}dt =0.\end{aligned}$$ In combination with Eq. (\[app:drb:cauchy\]) we obtain the so-called Wick rotation $$\label{app:drb:wickrotation} \int_{-\infty}^\infty f(t)dt=-i\int_{\infty}^{-\infty}dt f(it) =i\int_{-\infty}^\infty dt f(it).$$ As an intermediate result the integral of Eq. (\[app:drb:int\]) reads $$I=\frac{1}{(2\pi)^4}i\int_{-\infty}^\infty dk_0\int d^3 k \frac{i}{(ik_0)^2-\vec{k}^2 -M^2 +i0^+} =\int \frac{d^4 l}{(2\pi)^4} \frac{1}{l^2+M^2-i0^+},$$ where $l^2=l_1^2+l_2^2+l_3^2+l_4^2$ denotes a Euclidian scalar product. In this [*special*]{} case, the integrand does not have a pole and we can thus omit the $-i0^+$ which gave the positions of the poles in the original integral consistent with the boundary conditions. The degree of divergence can be estimated by simply counting the powers of momenta [@Veltman:wz]. If the integral behaves asymptotically as $\int d^4 l /l^2$, $\int d^4 l /l^3$, $\int d^4 l /l^4$ the integral is said to diverge quadratically, linearly, and logarithmically, respectively. Thus, our example $I$ diverges quadratically. Various methods have been devised to regularize divergent integrals. We will make use of [*dimensional*]{} regularization, because it preserves algebraic relations between Green functions (Ward identities) if the underlying symmetries do not depend on the number of dimensions of space-time. In dimensional regularization, we generalize the integral from 4 to $n$ dimensions and introduce polar coordinates $$\begin{aligned} \label{app:drb:polkoord} l_1&=& l\cos(\theta_1),\nonumber\\ l_2&=& l\sin(\theta_1)\cos(\theta_2),\nonumber\\ l_3&=& l\sin(\theta_1)\sin(\theta_2)\cos(\theta_3),\nonumber\\ &\vdots&\nonumber\\ l_{n-1}&=&l\sin(\theta_1)\sin(\theta_2)\cdots\cos(\theta_{n-1}),\nonumber\\ l_{n}&=&l\sin(\theta_1)\sin(\theta_2)\cdots\sin(\theta_{n-1}),\end{aligned}$$ where $0\leq l$, $\theta_i\in[0,\pi], i=1,\cdots,n-2$, $\theta_{n-1}\in [0,2\pi]$. A general integral is then symbolically of the form $$\label{app:drb:volumenelement} \int d^n l\cdots = \int_0^\infty l^{n-1}dl\int_0^{2\pi}d\theta_{n-1} \int_0^\pi d\theta_{n-2}\sin(\theta_{n-2})\cdots\int_0^\pi d\theta_1 \sin^{n-2}(\theta_1)\cdots .$$ If the integrand does not depend on the angles, the angular integration can explicitly be carried out. To that end one makes use of $$\int_0^\pi \sin^m(\theta) d\theta=\frac{\sqrt{\pi} \Gamma\left(\frac{m+1}{2} \right)}{\Gamma\left(\frac{m+2}{2}\right)}$$ which can be shown by induction. We then obtain for the angular integration $$\begin{aligned} \label{app:drb:winkelintegration} \int_0^{2\pi}d\theta_{n-1}\cdots \int_0^\pi d\theta_1 \sin^{n-2}(\theta_1)&=& 2\pi\underbrace{\frac{\sqrt{\pi}\Gamma(1)}{\Gamma\left(\frac{3}{2}\right)} \frac{\sqrt{\pi}\Gamma\left(\frac{3}{2}\right)}{\Gamma(2)}\cdots \frac{\sqrt{\pi}\Gamma\left(\frac{n-1}{2}\right)}{\Gamma\left(\frac{n}{2} \right)}}_{\mbox{$(n-2)$ factors}}\nonumber\\ &=&2\frac{\pi^\frac{n}{2}}{\Gamma\left(\frac{n}{2}\right)}.\end{aligned}$$ We define the integral for $n$ dimensions ($n$ integer) as $$\label{app:drb:im2} I_n(M^2,\mu^2)=\mu^{4-n}\int\frac{d^nk}{(2\pi)^n}\frac{i}{k^2-M^2+i0^+},$$ where for convenience we have introduced the renormalization scale $\mu$ so that the integral has the same dimension for arbitrary $n$. (The integral of Eq. (\[app:drb:im2\]) is convergent only for $n=1$.) After the Wick rotation of Eq. (\[app:drb:wickrotation\]) and the angular integration of Eq. (\[app:drb:winkelintegration\]) the integral formally reads $$I_n(M^2,\mu^2)=\mu^{4-n}2\frac{\pi^\frac{n}{2}}{\Gamma\left(\frac{n}{2}\right)} \frac{1}{(2\pi)^n} \int_0^\infty dl \frac{l^{n-1}}{l^2+M^2}.$$ For later use, we investigate the (more general) integral $$\label{app:drb:mgint} \int_0^\infty \frac{l^{n-1}dl}{(l^2+M^2)^\alpha} =\frac{1}{(M^2)^\alpha}\int_0^\infty \frac{l^{n-1}dl}{(\frac{l^2}{M^2}+1 )^\alpha} =\frac{1}{2}(M^2)^{\frac{n}{2}-\alpha} \int_0^\infty \frac{ t^{\frac{n}{2}-1}dt}{(t+1)^\alpha},$$ where we made use of the substitution $t\equiv l^2/M^2$. We then make use of the Beta function $$\label{app:drb:betafunktion} B(x,y)=\int_0^\infty \frac{t^{x-1}dt}{(1+t)^{x+y}}=\frac{\Gamma(x)\Gamma(y)}{ \Gamma(x+y)},$$ where the [*integral*]{} converges for $x>0$, $y>0$ and diverges if $x\leq 0$ or $y\leq 0$. For non-positive values of $x$ or $y$ we make use of the analytic continuation in terms of the Gamma function to define the Beta function and thus the integral of Eq. (\[app:drb:mgint\]).[^121] Putting $x=n/2$, $x+y=\alpha$ and $y=\alpha-n/2$ our (intermediate) integral reads $$\label{app:drb:allgint} \int_0^\infty \frac{l^{n-1}dl}{(l^2+M^2)^\alpha}= \frac{1}{2}(M^2)^{\frac{n}{2}-\alpha}\frac{\Gamma\left(\frac{n}{2}\right) \Gamma\left(\alpha-\frac{n}{2}\right)}{\Gamma(\alpha)}$$ which, for $\alpha=1$, yields for our original integral $$\begin{aligned} \label{app:drb:iint} I_n(M^2,\mu^2)&=&\mu^{4-n}\underbrace{2\frac{\pi^\frac{n}{2}}{\Gamma\left( \frac{n}{2}\right)}}_{\mbox{angular integration}} \frac{1}{(2\pi)^n} \frac{1}{2} (M^2)^{\frac{n}{2}-1}\frac{\Gamma \left(\frac{n}{2}\right)\Gamma\left(1-\frac{n}{2}\right)}{ \underbrace{\Gamma(1)}_{\mbox{1}}}\nonumber\\ &=&\frac{\mu^{4-n}}{(4\pi)^{\frac{n}{2}}} (M^2)^{\frac{n}{2}-1} \Gamma\left(1-\frac{n}{2}\right).\end{aligned}$$ Since $\Gamma(z)$ is an analytic function in the complex plane except for poles of first order in $0,-1,-2,\cdots$, and $a^z=\exp[\ln(a)z]$, $a\in R^+$ is an analytic function in $C$, the right-hand side of Eq. (\[app:drb:iint\]) can be thought of as a function of a [*complex*]{} variable $n$ which is analytic in $C$ except for poles of first order for $n=2,4,6,\cdots$. Making use of $$\mu^{4-n}=(\mu^2)^{2-\frac{n}{2}},\quad (M^2)^{\frac{n}{2}-1}=M^2 (M^2)^{\frac{n}{2}-2}, \quad (4\pi)^\frac{n}{2}=(4\pi)^2(4\pi)^{\frac{n}{2}-2},$$ we define (for complex n) $$I(M^2,\mu^2,n)=\frac{M^2}{(4\pi)^2}\left(\frac{4\pi\mu^2}{M^2}\right)^{2- \frac{n}{2}} \Gamma\left(1-\frac{n}{2}\right).$$ Of course, for $n\to 4$ the Gamma function has a pole and we want to investigate how this pole is approached. The property $\Gamma(z+1)=z\Gamma(z)$ allows one to rewrite $$\Gamma\left(1-\frac{n}{2}\right)= \frac{\Gamma\left(1-\frac{n}{2}+1\right)}{1-\frac{n}{2}} =\frac{\Gamma\left(2-\frac{n}{2}+1\right)}{\left(1-\frac{n}{2}\right) \left(2-\frac{n}{2}\right)}=\frac{\Gamma\left(1+\frac{\epsilon}{2}\right)}{ (-1)\left(1-\frac{\epsilon}{2}\right)\frac{\epsilon}{2}},$$ where we defined $\epsilon\equiv 4-n$. Making use use of $a^x=\exp[\ln(a)x]=1+\ln(a)x+O(x^2)$ we expand the integral for small $\epsilon$ $$\begin{aligned} I(M^2,\mu^2,n) &=&\frac{M^2}{16\pi^2}\left[1+\frac{\epsilon}{2}\ln\left( \frac{4\pi\mu^2}{M^2}\right)+O(\epsilon^2)\right]\\ &&\times \left(-\frac{2}{\epsilon}\right)\left[1+\frac{\epsilon}{2}+O(\epsilon^2)\right] \left[\underbrace{\Gamma(1)}_{\mbox{1}} +\frac{\epsilon}{2}\Gamma'(1)+O(\epsilon^2)\right] \\ &=&\frac{M^2}{16\pi^2}\left[-\frac{2}{\epsilon}\underbrace{-\Gamma'(1)}_{ \mbox{$\gamma_E=0.5772\cdots$}}-1-\ln(4\pi)+\ln\left(\frac{M^2}{\mu^2}\right) +O(\epsilon)\right],\end{aligned}$$ where $\gamma_E$ is Euler’s constant. We finally obtain $$\label{app:drb:im22} I(M^2,\mu^2,n)=\frac{M^2}{16\pi^2}\left[ R+\ln\left(\frac{M^2}{\mu^2}\right)\right]+O(n-4),$$ where $$\label{app:constantR} R=\frac{2}{n-4}-[\mbox{ln}(4\pi)+\Gamma'(1)+1].$$ Using the same techniques one can easily derive a very useful expression for the more general integral $$\begin{aligned} \label{app:drb:moregenint} \lefteqn{\int \frac{d^n k}{(2\pi)^n} \frac{(k^2)^p}{(k^2-M^2+i0^+)^q}=} \nonumber\\ &&i(-)^{p-q} \frac{1}{(4\pi)^{\frac{n}{2}}}(M^2)^{p+\frac{n}{2}-q} \frac{\Gamma\left(p+\frac{n}{2}\right)\Gamma\left(q-p-\frac{n}{2}\right)}{ \Gamma\left(\frac{n}{2}\right)\Gamma(q)}.\end{aligned}$$ We first assume $M^2>0$, $p=0,1,\cdots$, $q=1,2,\cdots$, and $p<q$. The last condition is used in the Wick rotation to guarantee that the quarter circles at infinity do not contribute to the integral. The transition to the Euclidian metric produces the factor $i(-)^{p-q}$. The angular integral in $n$ dimensions is then performed as in Eq. (\[app:drb:winkelintegration\]). The remaining radial integration is done using Eq. (\[app:drb:mgint\]) with the substitution $n-1\to 2p+n-1$ and $\alpha \to q$. The analytic continuation of the right-hand side of Eq.  (\[app:drb:moregenint\]) is used to also define expressions with (integer) $q\leq p$ in dimensional regularization. In the context of combining propagators by using Feynman’s trick one encounters integrals of the type of Eq.  (\[app:drb:moregenint\]) with $M^2$ replaced by $A-i0^+$, where $A$ is a real number. In this context it is important to consistently deal with the boundary condition $-i0^+$ [@Veltman:wz]. For example, let us consider a term of the type $\ln(A-i0^+)$. To that end one expresses a complex number $z$ in its polar form $z=|z|\exp(i\varphi),$ where the argument $\varphi$ of $z$ is uniquely determined if, in addition, we demand $-\pi\leq\varphi < \pi$. For $A>0$ one simply has $\ln(A-i0^+)=\ln(A)$. For $A<0$ the infinitesimal imaginary part indicates that $-|A|$ is reached in the third quadrant from below the real axis so that we have to use the $-\pi$. We then make use of $\ln(ab)=\ln(a)+\ln(b)$ and obtain $$\ln(A-i0^+)=\ln(|A|)+\ln(e^{-i\pi})=\ln(|A|)-i\pi,\quad A<0.$$ Both cases can be summarized in a single expression $$\label{app:drb:lna} \ln(A-i0^+)=\ln(|A|)-i\pi\Theta(-A)\quad\mbox{for}\,A\in R.$$ The preceding discussion is of importance for consistently determining imaginary parts of loop integrals. Let us conclude with the general observation that (ultraviolet) divergences of one-loop integrals in dimensional regularization always show up as single poles in $\epsilon=4-n$. Loop Integrals {#app_li} ============== In Appendix \[app\_drb\] we discussed the basic ideas of the method of dimensional regularization. Here we outline the calculation of more complicated one-loop integrals of mesonic as well as heavy-baryon chiral perturbation theory. We restrict ourselves to the cases needed to reproduce the examples discussed in the main text and refer the interested reader to Refs.  for more details. One-Loop Integrals of the Mesonic Sector {#app_sec_olims} ---------------------------------------- In the mesonic sector we will use the following definition and nomenclature for the scalar loop integrals (i.e., no Lorentz indices) extended to $n$ dimensions: $$\begin{aligned} \label{app:olims:ipidef} \lefteqn{I_{\pi\cdots\pi}(q_1,\cdots,q_m)\equiv}\nonumber\\ &&i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{(k+q_1)^2-M_\pi^2+i0^+}\cdots \frac{1}{(k+q_m)^2-M_\pi^2+i0^+},\end{aligned}$$ where we omit an explicit reference to the scale $\mu$ and the “number of dimensions” $n$.[^122] In the SU(3)$\times$SU(3) case one also needs loop integrals with different masses such as, e.g., $$I_{\pi K}(q_1,q_2)=i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{(k+q_1)^2-M_\pi^2+i0^+}\frac{1}{(k+q_2)^2-M_K^2+i0^+}.$$ ### $I_\pi$ {#app_subsec_ipi} We define $$\label{app:ipi} I_\pi(q)\equiv i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{(k+q)^2-M_\pi^2+i0^+}.$$ Using a shift $k\to k-q$ (in the regularized integral) we obtain $$I_\pi(q)=I_\pi(0).$$ However, this is just the basic integral $I$ we discussed in detail in App. \[app\_drb\]: $$\label{app:ipib} I_\pi(0)=\frac{M^2_\pi}{16\pi^2}\left[R+\ln\left(\frac{M^2_\pi}{\mu^2}\right) \right]+O(n-4),$$ where $$\label{app:ipiR} R=\frac{2}{n-4}-[\ln(4\pi)+\Gamma'(1)+1].$$ Later on we will also use the common notation $A_0(M_\pi^2)$ for the integral $I_\pi(0)$. ### $I_{\pi\pi}$ {#app_subsec_ipipi} In the calculation of the one-loop contribution of Fig. \[4:9:pionffloop1\] to the electromagnetic form factor of the pion, Eq. (\[4:9:loopc1\]), we encounter an integral of the type $$\label{app:ipipi} I_{\pi\pi}(q_1,q_2)\equiv i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{(k+q_1)^2-M_\pi^2+i0^+}\frac{1}{(k+q_2)^2-M_\pi^2+i0^+}.$$ Using a shift $k\to k-q_2$ we obtain $$I_{\pi\pi}(q_1,q_2)=I_{\pi\pi}(q_1-q_2,0).$$ It is thus sufficient to consider $I_{\pi\pi}(q,0)$. To that end, we first combine the denominators using Feynman’s trick: $$\label{app:ipipiftrick} \frac{1}{ab}=\int_0^1 dz \frac{1}{[az+b(1-z)]^2},$$ with $a=(k+q)^2-M_\pi^2+i0^+$ and $b=k^2-M_\pi^2+i0^+$ to obtain $$I_{\pi\pi}(q,0)=i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \int_0^1 dz \frac{1}{[k^2+2k\cdot q z-(M_\pi^2-zq^2)+i0^+]^2},$$ and perform the shift $k\to k-zq$, resulting in $$I_{\pi\pi}(q,0)=\int_0^1 dz\left( i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{\left\{k^2-[M_\pi^2+z(z-1)q^2]+i0^+\right\}^2}\right).$$ Performing the Wick rotation, applying the results of Eqs. (\[app:drb:winkelintegration\]) and (\[app:drb:allgint\]), expanding the result in $(n-4)$, and finally performing the $z$ integration, we obtain $$\label{app:ipipib} I_{\pi\pi}(q,0)=\frac{1}{16\pi^2}\left[ R+\ln\left(\frac{M^2_\pi}{\mu^2}\right)+1+J^{(0)}\left(\frac{q^2}{M^2_\pi} \right)+O(n-4)\right],$$ where [@Unkmeir:1999md] $$\begin{aligned} J^{(0)}(x) &=&\int_0^1 dz \ln[1+x(z^2-z)-i0^+]\\ &=& \left \{ \begin{array}{ll} -2-\sigma\ln\left(\frac{\sigma-1}{\sigma+1}\right),&x<0,\\ -2+2\sqrt{\frac{4}{x}-1}\,\mbox{arccot} \left(\sqrt{\frac{4}{x}-1}\right),&0\le x<4,\\ -2-\sigma\ln\left(\frac{1-\sigma}{1+\sigma}\right)-i\pi\sigma,& 4<x, \end{array} \right.\end{aligned}$$ with $$\sigma(x)=\sqrt{1-\frac{4}{x}},\quad x\notin [0,4].$$ Note that Eq. (\[app:ipipib\]) represents a case where the $i0^+$ boundary condition has to be treated consistently, as discussed at the end of App. \[app\_drb\]. For later use we introduce the notation $$B_0(q^2,M_\pi^2)=I_{\pi\pi}(q,0),$$ where the subscript $0$ refers to the scalar character of the integral. Next we want to determine the tensor integrals appearing in Eq. (\[4:9:loopc1\]) by reducing them to already known integrals. The general idea consists of parameterizing the tensor structure in terms of the metric tensor and products of external four-vectors and multiplying the results by invariant functions of Lorentz scalars. We first consider $$\label{app:Ipipik} i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{k^\mu}{(k+q)^2-M_\pi^2+i0^+}\frac{1}{k^2-M_\pi^2+i0^+},$$ which must have the form $$\label{app:B1} q^\mu B_1(q^2,M_\pi^2),$$ where the subscript $1$ refers to one four-vector $k$ in the numerator of the integral. We contract Eq. (\[app:B1\]) with $q_\mu$ and make use of $q\cdot k=[(k+q)^2 -M_\pi^2-(k^2-M_\pi^2)-q^2]/2$ to obtain $$\begin{aligned} q^2 B_1(q^2,M_\pi^2)&=&\frac{1}{2} i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{k^2-M_\pi^2+i0^+}\\ &&-\frac{1}{2} i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{(k+q)^2-M_\pi^2+i0^+}\\ &&-\frac{1}{2}q^2 i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{(k+q)^2-M_\pi^2+i0^+}\frac{1}{k^2-M_\pi^2+i0^+}\\ &=&-\frac{1}{2}q^2 B_0(q^2,M_\pi^2),\end{aligned}$$ where we used the argument in Appendix \[app\_subsec\_ipi\] to show that the first two integrals cancel. We have thus reduced the determination of Eq. (\[app:Ipipik\]) to an already known integral: $$B_1(q^2,M_\pi^2)=-\frac{1}{2}B_0(q^2,M_\pi^2).$$ Finally, we also need $$\begin{aligned} \label{app:B20B21} \lefteqn{q^\mu q^\nu B_{20}(q^2,M_\pi^2) +g^{\mu\nu}q^2 B_{21}(q^2,M_\pi^2)=}\nonumber\\ && i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{k^\mu k^\nu}{(k+q)^2-M_\pi^2+i0^+}\frac{1}{k^2-M_\pi^2+i0^+},\end{aligned}$$ where the first subscript $2$ refers to two four-vectors $k$ in the numerator of the integral and the second subscripts $0$ and $1$ refer to the number of metric tensors in the parameterization, respectively. Contracting with $g_{\mu\nu}$ and making use of $g_{\mu\nu} g^{\mu\nu}=n$ in $n$ dimensions we obtain $$\label{app:cond1} q^2 B_{20} + n q^2 B_{21} =A_0+M_\pi^2 B_0.$$ Similarly, contracting with $q_\mu$ we obtain $$\label{app:cond2} q^2 B_{20} +q^2 B_{21}= \frac{1}{2} A_0+\frac{q^2}{4} B_0.$$ By subtracting Eq. (\[app:cond2\]) from (\[app:cond1\]) we find $$\begin{aligned} \label{app:q2B21} q^2 B_{21}&=&\frac{1}{n-1}\left[\frac{1}{2}A_0+(M_\pi^2-\frac{q^2}{4}) B_0\right]\nonumber\\ &=&\frac{1}{32\pi^2}\left\{\left(M_\pi^2-\frac{q^2}{6}\right)\left[ R+\ln\left(\frac{M_\pi^2}{\mu^2}\right)\right]\right.\nonumber\\ &&\left.+\frac{2}{3}J^{(0)}\left(\frac{q^2}{M_\pi^2}\right)\left( M_\pi^2-\frac{q^2}{4}\right)-\frac{q^2}{18}\right\}+O(n-4),\end{aligned}$$ where we made use of $$\frac{1}{n-1}=\frac{1}{3+(n-4)}=\frac{1}{3}\left(1-\frac{n-4}{3}+\cdots\right)$$ and Eqs. (\[app:ipib\]) and (\[app:ipipib\]). From Eq. (\[app:cond2\]) we obtain $$\begin{aligned} \label{app:B20} B_{20}&=&\frac{1}{q^2(n-1)}\left(\frac{n-2}{2}A_0 +n\frac{q^2}{4}B_0-M_\pi^2B_0\right)\nonumber\\ &=&\frac{1}{3q^2}\left[A_0+\frac{6M_\pi^2-q^2}{96\pi^2}+(q^2-M_\pi^2)B_0\right ]+O(n-4)\nonumber\\ &=&\frac{1}{48\pi^2}\left[R+\ln\left(\frac{M^2_\pi}{\mu^2}\right) +\frac{5}{6}+\left(1-\frac{M_\pi^2}{q^2}\right)J^{(0)}\left( \frac{q^2}{M_\pi^2}\right)\right]+O(n-4).\nonumber\\\end{aligned}$$ In working out Eqs. (\[app:q2B21\]) or (\[app:B20\]) it must be remembered that $R$ contains a term $2/(n-4)$ which, when multiplied by $(n-4)$, gives a term of order $(n-4)^0$. This must be done carefully to obtain, e.g., the $-q^2/18$ term in Eq. (\[app:q2B21\]) and the $5/6$ term in Eq. (\[app:B20\]). One-Loop Integrals of the Heavy-Baryon Sector {#sec_olinhbs} --------------------------------------------- ### Basic Loop Integral {#subsec_bli} The structure of the one-loop integrals in the heavy-baryon approach is slightly different from the integrals of the mesonic sector discussed in Sec. \[app\_sec\_olims\]. Here we will outline the calculation of a basic loop integral which serves as a starting point for more complicated calculations. Consider an integral of the type $$\label{app:int1} \int \frac{d^4k}{(2\pi)^4}\frac{1}{v\cdot k +\alpha +i0^+} \frac{1}{k^2-A+i0^+},$$ where $\alpha$ and $A$ are arbitrary real numbers and $v$ is the four-velocity of the heavy-baryon approach. Counting powers of the momenta, the integral is linearly divergent. An integral of this type appears in the calculation of the nucleon self energy in Sec. \[subsec\_aop3olcnm\]. We combine the denominators using the following Feynman trick: $$\label{app:feynmantricksphb} \frac{1}{ab}=2\int_0^\infty dy \frac{1}{(2ya+b)^2}.$$ Below, this choice will allow us most easily to combine the $y$ integration with the “radial” integration of the loop momentum after the Wick rotation. Inserting $a=v\cdot k +\alpha +i0^+$ and $b=k^2-A+i0^+$, we obtain $$\label{app:int2} 2\int \frac{d^4k}{(2\pi)^4}\int_0^\infty dy \frac{1}{(k^2+2yv\cdot k +2 y\alpha-A+i0^+)^2}.$$ We now generalize to $n$ dimensions: $$\label{app:int3} 2\mu^{4-n}\int_0^\infty dy \int \frac{d^nk}{(2\pi)^n} \frac{1}{(k^2+2yv\cdot k +2 y\alpha-A+i0^+)^2}$$ and perform a shift of integration variables $k\to k-yv$ so that there remain no terms linear in $k$ in the denominator: $$\label{app:int4} 2\mu^{4-n}\int_0^\infty dy \int \frac{d^n k}{(2\pi)^n} \frac{1}{(k^2-y^2-A+2y\alpha+i0^+)^2},$$ where we made use of $v^2=1$. Finally, we shift the integration variable $y\to y+\alpha$ in order to eliminate terms linear in $y$ in the denominator: $$\label{app:int5} 2\mu^{4-n}\int_{-\alpha}^\infty dy \int\frac{d^nk}{(2\pi)^n}\frac{1} {(k^2-y^2-A+\alpha^2+i0^+)^2}.$$ The $y$ integration is split into $[-\alpha,0]$ and $[0,\infty[$. Making use of a Wick rotation and Eqs. (\[app:drb:winkelintegration\]), (\[app:drb:allgint\]), and (\[app:drb:lna\]) we obtain for the first integral $$\begin{aligned} \label{app:int6a} \lefteqn{2\mu^{4-n}\int_{-\alpha}^0 dy \int\frac{d^nk}{(2\pi)^n}\frac{1} {(k^2-y^2-A+\alpha^2+i0^+)^2}}\nonumber\\ &=&-\frac{i}{8\pi^2}\int_{-\alpha}^0 dy\left[R+1+ \ln\left(\frac{|A+y^2-\alpha^2|}{\mu^2}\right) -i\pi\Theta(\alpha^2-A-y^2)\right]\nonumber\\ &&+O(n-4),\end{aligned}$$ where $$R=\frac{2}{n-4}-[\mbox{ln}(4\pi)+\Gamma'(1)+1].$$ Performing a Wick rotation and using Eq. (\[app:drb:winkelintegration\]), the second integral reads $$\begin{aligned} \label{app:int6b} \lefteqn{2\mu^{4-n}\int_0^\infty dy \int\frac{d^nk}{(2\pi)^n}\frac{1} {(k^2-y^2-A+\alpha^2+i0^+)^2}}\nonumber\\ &=&\frac{4i\mu^{4-n}}{(4\pi)^\frac{n}{2}\Gamma\left(\frac{n}{2}\right)} \int_0^\infty dy\int_0^\infty dx \frac{x^{n-1}}{(x^2+y^2+A-\alpha^2-i0^+)^2}.\end{aligned}$$ Using polar coordinates $x=r\cos(\varphi)$ and $y=r\sin(\varphi)$ together with $$\int_0^{\frac{\pi}{2}}\sin^{2\alpha+1}(\varphi)\cos^{2\beta+1}(\varphi)d\varphi = \frac{\Gamma(\alpha+1)\Gamma(\beta+1)}{2\Gamma(\alpha+\beta+2)},$$ we rewrite the second integral as ($\Gamma(1/2)=\sqrt{\pi}$) $$\frac{2i \mu^{4-n}\sqrt{\pi}}{(4\pi)^\frac{n}{2} \Gamma\left(\frac{n+1}{2}\right)} \int_0^\infty dr\frac{r^n}{(r^2+A-\alpha^2-i0^+)^2}.$$ Finally, applying again Eq. (\[app:drb:allgint\]) for the radial integral we obtain for the second integral (in four dimensions) $$\label{app:int7b} 2\int_0^\infty dy \int\frac{d^4k}{(2\pi)^4}\frac{1} {(k^2-y^2-A+\alpha^2+i0^+)^2} =\frac{-i}{8\pi}\sqrt{A-\alpha^2-i0^+},$$ where we made use of $\Gamma(2)=1$ and $\Gamma(-1/2)=-2\sqrt{\pi}$. The remaining $y$ integration of the first integral is elementary and we obtain as the final expression $$\begin{aligned} \label{app:intfinal} \lefteqn{\mu^{4-n}\int \frac{d^n k}{(2\pi)^4}\frac{1}{v\cdot k +\alpha +i0^+} \frac{1}{k^2-A+i0^+}}\nonumber\\ &=&\frac{-i}{8\pi^2}\Bigg( \alpha\left[R+\ln\left(\frac{|A|}{\mu^2}\right)-1\right] \nonumber\\ &&+ \left\{ \begin{array}{ll} 2\sqrt{\alpha^2-A}\,\mbox{arccosh}\left(\frac{\alpha}{\sqrt{A}}\right) &\\ -2\pi i\sqrt{\alpha^2-A},&A>0\wedge\alpha>\sqrt{A},\\ 2\sqrt{A-\alpha^2}\,\arccos\left(-\frac{\alpha}{\sqrt{A}}\right),& A>\alpha^2,\\ -2\sqrt{\alpha^2-A}\,\mbox{arccosh}\left(-\frac{\alpha}{\sqrt{A}}\right), &A>0\wedge\alpha<-\sqrt{A},\\ 2\sqrt{\alpha^2-A}\,\mbox{arcsinh}\left(\frac{\alpha}{\sqrt{-A}}\right)&\\ -i\pi\left(\alpha+\sqrt{\alpha^2-A}\,\right),&A<0, \end{array} \right\} \nonumber\\ &&+O(n-4)\Bigg),\end{aligned}$$ with $ R=\frac{2}{n-4}-[\mbox{ln}(4\pi)+\Gamma'(1)+1]$. ### $J_{\pi N}$ {#subsec_jpin} In analogy to the mesonic case we define $$\label{app:jpin} J_{\pi N}(q;\omega)\equiv i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{(k+q)^2-M_\pi^2+i0^+}\frac{1}{v\cdot k+\omega+i0^+}.$$ Using a shift $k\to k-q$ we obtain $$J_{\pi N}(q;\omega)=J_{\pi N}(0;\omega-v\cdot q).$$ It is thus sufficient to consider $J_{\pi N}(0;\omega)$ which, using the result of Eq. (\[app:intfinal\]), is given by[^123] $$\begin{aligned} \label{app:jpinb} \lefteqn{J_{\pi N}(0;\omega)}\nonumber\\ &=& \frac{\omega}{8\pi^2}\left[R+\ln\left(\frac{M^2_\pi}{\mu^2}\right)-1 \right]\nonumber\\ &&+\frac{1}{8\pi^2}\left\{ \begin{array}{ll} 2\sqrt{\omega^2-M_\pi^2}\, \mbox{arccosh}\left(\frac{\omega}{M_\pi} \right)-2\pi i\sqrt{\omega^2-M^2_\pi}, &\omega>M_\pi,\\ 2\sqrt{M^2_\pi-\omega^2}\,\arccos\left(-\frac{\omega}{M_\pi} \right), &\omega^2<M^2_\pi,\\ -2\sqrt{\omega^2-M_\pi^2}\, \mbox{arccosh}\left(-\frac{\omega}{M_\pi} \right), & \omega<-M_\pi, \end{array} \right\}\nonumber\\ &&+O(n-4).\end{aligned}$$ In the calculation of the nucleon self energy we also need tensor integrals which, as in the mesonic case, one may reduce to already known integrals. Let us introduce the notation $$C_0(\omega,M_\pi^2)=J_{\pi N}(0;\omega),$$ where the subscript $0$ refers to the scalar character of the integral. Once again, the general idea in the determination of tensor integrals consists of parameterizing the tensor structure in terms of the metric tensor and products of the four-velocity $v$. We first consider $$\label{app:JpiNk} i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{k^\mu}{k^2-M_\pi^2+i0^+}\frac{1}{v\cdot k+\omega +i0^+},$$ which must have the form $$\label{app:C1} v^\mu C_1(\omega,M_\pi^2),$$ where the subscript $1$ refers to one four-vector $k$ in the numerator of the integral. We contract Eq. (\[app:C1\]) with $v_\mu$, make use of $v^2=1$, and add and subtract $\omega$ in the numerator of the integral, obtaining $$\begin{aligned} \label{app:c1result} C_1(\omega,M_\pi^2)&=&i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{k^2-M_\pi^2+i0^+}\nonumber\\ &&-\omega i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{k^2-M_\pi^2+i0^+}\frac{1}{v\cdot k+\omega+i0^+}\nonumber\\ &=&I_\pi(0)-\omega J_{\pi N}(0;\omega),\end{aligned}$$ where $I_\pi(0)$ is defined in Eq. (\[app:ipib\]). We have thus reduced the determination of Eq. (\[app:JpiNk\]) to the already known integrals $I_\pi$ and $J_{\pi N}$. As our final example, let us discuss $$i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{k^\mu k^\nu}{k^2-M_\pi^2+i0^+}\frac{1}{v\cdot k+\omega+i0^+},$$ which must be of the form $$\label{app:C20C21} v^\mu v^\nu C_{20}(\omega,M_\pi^2) +g^{\mu\nu} C_{21}(\omega,M_\pi^2),$$ where the first subscript $2$ refers to two four-vectors $k$ in the numerator of the integral and the second subscripts $0$ and $1$ refer to the number of metric tensors in the parameterization, respectively. Contracting with $v_\mu$ and adding and subtracting $\omega$ in the numerator, we obtain $$\label{app:condt21} v^\nu C_{20}+v^\nu C_{21}= -\omega v^{\nu} C_1=-\omega v^\nu[I_\pi(0) -\omega J_{\pi N}(0;\omega)],$$ where we made use of Eq. (\[app:c1result\]) and $$i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{k^\nu}{k^2-M_\pi^2+i0^+}=0.$$ Finally, contracting with $g_{\mu\nu}$ and making use of $g_{\mu\nu} g^{\mu\nu}=n$ in $n$ dimensions we obtain $$\label{app:condt22} C_{20} + n C_{21} =M_\pi^2 C_0,$$ where we made use of $$\label{app:intvan} i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n}\frac{1}{v\cdot k+\omega+i0^+}=0$$ in dimensional regularization. In order to verify Eq. (\[app:intvan\]), one writes $$\begin{aligned} i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n}\frac{1}{v\cdot k+\omega+i0^+} &=& i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{v\cdot k}{(v\cdot k)^2-(\omega+i0^+)^2}\\ && -\omega i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n} \frac{1}{(v\cdot k)^2-(\omega+i0^+)^2}.\end{aligned}$$ The first term vanishes, because the integrand is odd in $k$. For the evaluation of the second term we choose $v^\mu=(1,0,\cdots,0)$. Applying the residue theorem, we obtain for the integral $$\int_{-\infty}^\infty dk_0 \frac{1}{k_0^2-(\omega+i0^+)^2}= \frac{\pi i}{\omega},\quad \omega\neq 0,$$ so that we may define for arbitrary $\omega$ $$\begin{aligned} i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n}\frac{1}{v\cdot k+\omega+i0^+} &=&-i\mu^{4-n} \int \frac{d^n k}{(2\pi)^n}\frac{\omega}{ k_0^2-(\omega+i0^+)^2}\\ =\frac{\pi}{(2\pi)^n}\mu^{4-n}\int d^{n-1}k. \end{aligned}$$ However, the last term vanishes in dimensional regularization. From Eqs. (\[app:condt21\]) and (\[app:condt22\]) we obtain $$\begin{aligned} \label{app:C21} C_{21}&=&\frac{1}{n-1}\left[(M_\pi^2-\omega^2)J_{\pi N}(0;\omega)+\omega I_\pi(0)\right]\nonumber\\ &=&\frac{1}{3}\left[(M_\pi^2-\omega^2)J_{\pi N}(0;\omega)+\omega I_\pi(0)\right] -\frac{\omega}{12\pi^2}\left(\frac{M_\pi^2}{2}-\frac{\omega^2}{3}\right), \nonumber\\\end{aligned}$$ and $$\begin{aligned} \label{app:C20} C_{20}&=&M_\pi^2 J_{\pi N}(0,\omega)-C_{21}\nonumber\\ &=&\frac{1}{3}(2M_\pi^2 +\omega^2)J_{\pi N}(0;\omega) -\frac{1}{3}\omega I_\pi(0) +\frac{\omega}{12\pi^2}\left(\frac{M_\pi^2}{2}-\frac{\omega^2}{3}\right), \nonumber\\\end{aligned}$$ where, as in the mesonic case, it is important to identify the finite terms resulting from the product of $O(n-4)$ terms and $R$. Finally, when discussing the relation between the infrared regularization method of Ref. [@Becher:1999he] and the heavy-baryon approach in Sec. \[sec\_mir\], we made use of the fact that an integral of the type $$\label{app:dnkk2pvkmq} \int d^n k \frac{(k^2)^p}{(v\cdot k+i0^+)^q}$$ vanishes in dimensional regularization. This can be seen by substituting $k=\lambda k'$ and relabeling $k'=k$ $$=\lambda^{n+2p-q}\int d^n k \frac{(k^2)^p}{(v\cdot k+i0^+)^q}.$$ Since $\lambda>0$ is arbitrary and, for fixed $p$ and $q$, the result is to hold for arbitrary $n$, Eq. (\[app:dnkk2pvkmq\]) is set to zero in dimensional regularization. We emphasize that the vanishing of Eq. (\[app:dnkk2pvkmq\]) has the character of a prescription [@Itzykson:rh]. The integral does not depend on any scale and its analytic continuation is ill defined in the sense that there is no dimension $n$ where it is meaningful. It is ultraviolet divergent for $n+2p-q\geq 0$ and infrared divergent for $n+2p-q\leq 0$. Different Forms of ${\cal L}_4$ in $\mbox{SU(2)}\times\mbox{SU(2)}$ =================================================================== GL Versus GSS {#app_sec_glvgss} ------------- The purpose of this appendix is to explicitly relate the two commonly used ${\cal O}(p^4)$ $\mbox{SU(2)}\times\mbox{SU(2)}$ Lagrangians of Refs.  [@Gasser:1983yg] (GL) and [@Gasser:1987rb] (GSS). In their pioneering work on mesonic $\mbox{SU(2)}\times\mbox{SU(2)}$ chiral perturbation theory [@Gasser:1983yg], Gasser and Leutwyler used a notation adopted from the O(4) nonlinear $\sigma$ model, because the two Lie groups $\mbox{SU(2)}\times\mbox{SU(2)}$ and O(4) are locally isomorphic, i.e., their Lie algebras are isomorphic. The effective Lagrangian was written in terms of invariant scalar products of real four-vectors in contrast to the nowadays standard trace form. The dynamical pion degrees of freedom were expressed in terms of a four-component real vector field of unit length with components $U^A(x)$, $A=0,1,2,3$. The connection to the SU(2) matrix $U(x)$ of Sec. \[subsec\_aqcd\] can be expressed as $$\begin{aligned} \label{app:glvgss:uau} U(x)&=&U^0(x)+i\vec{\tau}\cdot\vec{U}(x),\nonumber\\ U^0(x)&=&\frac{1}{2}\mbox{Tr}[U(x)],\nonumber\\ U^i(x)&=&-\frac{i}{2}\mbox{Tr}[\tau_i U(x)],\quad i=1,2,3,\end{aligned}$$ with $$\sum_{A=0}^3 [U^A(x)]^2=1,$$ so that $U$ is unitary. The lowest-order Lagrangian in the trace notation is given by the $\mbox{SU(2)}\times\mbox{SU(2)}$ version of Eq. (\[4:5:l2\]), and the transcription of the ${\cal O}(p^4)$ $\mbox{SU(2)}\times\mbox{SU(2)}$ Lagrangian \[see Eq. (5.5) of Ref. [@Gasser:1983yg]\] reads[^124] $$\begin{aligned} \label{app:glvgss:l4gl} {\cal L}^{\rm GL}_4 &=& \frac{l_1}{4} \left\{\mbox{Tr}[D_{\mu}U (D^{\mu}U)^{\dagger}] \right\}^2 +\frac{l_2}{4}\mbox{Tr}[D_{\mu}U (D_{\nu}U)^{\dagger}] \mbox{Tr}[D^{\mu}U (D^{\nu}U)^{\dagger}] \nonumber \\ &&+\frac{l_3}{16}\left[\mbox{Tr}(\chi U^\dagger+ U\chi^\dagger)\right]^2 +\frac{l_4}{4}\mbox{Tr}[D_\mu U(D^\mu\chi)^\dagger +D_\mu\chi(D^\mu U)^\dagger]\nonumber\\ &&+l_5\left[\mbox{Tr}(f^R_{\mu\nu}U f^{\mu\nu}_LU^\dagger) -\frac{1}{2}\mbox{Tr}(f_{\mu\nu}^L f^{\mu\nu}_L +f_{\mu\nu}^R f^{\mu\nu}_R)\right]\nonumber\\ &&+i\frac{l_6}{2}\mbox{Tr}[ f^R_{\mu\nu} D^{\mu} U (D^{\nu} U)^{\dagger} + f^L_{\mu\nu} (D^{\mu} U)^{\dagger} D^{\nu} U]\nonumber\\ &&-\frac{l_7}{16}\left[\mbox{Tr}(\chi U^\dagger-U\chi^\dagger)\right]^2 \nonumber\\ &&+\frac{h_1+h_3}{4}\mbox{Tr}(\chi\chi^\dagger) +\frac{h_1-h_3}{16}\left\{ \left[\mbox{Tr}(\chi U^\dagger + U\chi^\dagger)\right]^2\right. \nonumber\\ &&\left. +\left[\mbox{Tr}(\chi U^\dagger-U\chi^\dagger)\right]^2 -2\mbox{Tr}(\chi U^\dagger\chi U^\dagger + U\chi^\dagger U\chi^\dagger) \right\}\nonumber\\ &&-2h_2 \mbox{Tr}(f_{\mu\nu}^L f^{\mu\nu}_L +f_{\mu\nu}^R f^{\mu\nu}_R).\end{aligned}$$ When comparing with the SU(3)$\times$SU(3) version of Eq. (\[4:7:l4gl\]) we first note that Eq. (\[app:glvgss:l4gl\]) contains fewer independent terms which results from the application of the trace relations, as discussed in Sec. \[subsec\_mclop6\]. The bare and the renormalized low-energy constants $l_i$ and $l_i^r$ are related by $$\label{app:lilir} l_i=l_i^r+\gamma_i\frac{R}{32\pi^2},$$ where $R=2/(n-4)-[\ln(4\pi)+\Gamma'(1)+1]$ and $$\gamma_1=\frac{1}{3},\quad \gamma_2=\frac{2}{3},\quad \gamma_3=-\frac{1}{2},\quad \gamma_4=2,\quad \gamma_5=-\frac{1}{6},\quad \gamma_6=-\frac{1}{3},\quad \gamma_7=0.$$ In the SU(2)$\times$SU(2) sector one often uses the scale-independent parameters $\bar{l}_i$ which are defined by $$\label{app:barli} l_i^r=\frac{\gamma_i}{32\pi^2}\left[\bar{l}_i+\ln\left(\frac{M^2}{\mu^2}\right) \right],\quad i=1,\cdots,6,$$ where $M^2=B_0(m_u+m_d)$. Since $\ln(1)=0$, the $\bar{l}_i$ are proportional to the renormalized coupling constant at the scale $\mu=M$. Table \[app:tablebarli\] contains numerical values for the scale-independent low-energy coupling constants $\bar{l}_i$ as obtained in Ref. [@Gasser:1983yg] together with more recent determinations. Value Obtained from $\gamma_i$ ------------- ------------------------------------------ ------------------------------------------------------- ---------------- $-2.3\pm 3.7$ [@Gasser:1983yg] $\pi\pi$ $D$-wave scattering lengths ${\cal O}(p^4)$ $-1.7\pm 1.0$ [@Bijnens:1994ie] $\pi\pi$ and $K_{l4}$ $\bar{l}_1$ $-1.5$ [@Bijnens:1997vq] $\pi\pi$ $D$-wave scattering lengths ${\cal O}(p^6)$ $\frac{1}{3}$ $-1.8$ [@Colangelo:2001df] $\pi\pi$ scattering ${\cal O}(p^4)$ + Roy equations $-0.4\pm 0.6$ [@Colangelo:2001df] $\pi\pi$ scattering ${\cal O}(p^6)$ + Roy equations $6.0 \pm 1.3 $ [@Gasser:1983yg] $\pi\pi$ $D$-wave scattering lengths ${\cal O}(p^4)$ $6.1\pm 0.5$ [@Bijnens:1994ie] $\pi\pi$ and $K_{l4}$ $\bar{l}_2$ $4.5$ [@Bijnens:1997vq] $\pi\pi$ $D$-wave scattering lengths ${\cal O}(p^6)$ $\frac{2}{3}$ $5.4$ [@Colangelo:2001df] $\pi\pi$ scattering ${\cal O}(p^4)$ + Roy equations $4.3\pm 0.1$ [@Colangelo:2001df] $\pi\pi$ scattering ${\cal O}(p^6)$ + Roy equations $2.9\pm 2.4$ [@Gasser:1983yg] SU(3) mass formulae $\bar{l}_3$ $|\bar{l}_3|\leq 16$ [@Colangelo:2001sp] $K_{l4}$ decay $-\frac{1}{2}$ $4.3\pm 0.9$ [@Gasser:1983yg] $F_K/F_\pi$ $\bar{l}_4$ $4.4\pm 0.3$ [@Bijnens:1998fm] scalar form factor ${\cal O}(p^6)$ 2 $4.4\pm 0.2$ [@Colangelo:2001df] $\pi\pi$ scattering ${\cal O}(p^6)$ + Roy equations $13.9 \pm 1.3$ [@Gasser:1983yg] $\pi\to e\nu\gamma$ ${\cal O}(p^4)$ $\bar{l}_5$ $13.0\pm 0.9$ [@Bijnens:1998fm] $\pi\to e\nu\gamma$ ${\cal O}(p^6)$ [@Bijnens:1996wm] $-\frac{1}{6}$ $16.5\pm 1.1$ [@Gasser:1983yg] $\langle r^2\rangle_\pi$ ${\cal O}(p^4)$ $\bar{l}_6$ $16.0\pm 0.5\pm 0.7$ vector form factor ${\cal O}(p^6)$ $-\frac{1}{3}$ [@Bijnens:1998fm] $l_7$ $O(5\times 10^{-3})$ [@Gasser:1983yg] $\pi^0$-$\eta$ mixing 0 : \[app:tablebarli\] Scale-independent low-energy coupling constants $\bar{l}_i$. Secondly, the expression proportional to $(h_1-h_3)$ can be rewritten so that the $U$’s completely drop out, i.e., it contains only external fields. The trick is to use $$\begin{aligned} \label{app:glvgss:trick} \lefteqn{2\mbox{Tr}(\chi U^\dagger\chi U^\dagger+U\chi^\dagger U\chi^\dagger)=} \nonumber\\ &&[\mbox{Tr}(\chi U^\dagger+U\chi^\dagger)]^2 +[\mbox{Tr}(\chi U^\dagger- U\chi^\dagger)]^2\nonumber\\ &&+[\mbox{Tr}(\tau_i\chi)]^2 +[\mbox{Tr}(\tau_i\chi^\dagger)]^2 -[\mbox{Tr}(\chi)]^2-[\mbox{Tr}(\chi^\dagger)]^2.\end{aligned}$$ The terms proportional to the $l_i$ agree with Eq. (4.2) of Ref. [@Bellucci:1994eb] but the $l_4$ term is not yet in the form of the SU(3)$\times$SU(3) version of Eq. (\[4:7:l4gl\]). By means of a total-derivative argument in combination with a field transformation as discussed in Sec. \[sec\_clop4\] we will transform ${\cal L}_4^{\rm GL}$ of Eq. (\[app:glvgss:l4gl\]) into another form which is often used in the literature. To that end let us make use of $$\begin{aligned} \lefteqn{\mbox{Tr}[D_\mu\chi (D^\mu U)^\dagger+D^\mu U (D_\mu\chi)^\dagger] =}\\ &&\partial_\mu \mbox{Tr}[\chi (D^\mu U)^\dagger + D^\mu U \chi^\dagger] -\mbox{Tr}[\chi (D^2 U)^\dagger+D^2 U\chi^\dagger],\end{aligned}$$ rewrite the $D^2 U$ and $(D^2 U)^\dagger$ terms by using $$\begin{aligned} 2 D^2 U U^\dagger&=&D^2 U U^\dagger - U (D^2 U)^\dagger -2D_\mu U (D^\mu U)^\dagger,\\ 2 U (D^2 U)^\dagger&=&U (D^2 U)^\dagger -D^2 U U^\dagger-2 D_\mu U (D^\mu U)^\dagger,\end{aligned}$$ to obtain $$\begin{aligned} \label{app:glvgss:intmed1} \lefteqn{\mbox{Tr}[D_\mu\chi (D^\mu U)^\dagger+D^\mu U (D_\mu\chi)^\dagger]=} \nonumber\\ &&\mbox{tot. der.}+ \mbox{Tr}[D_\mu U (D^\mu U)^\dagger(\chi U^\dagger + U\chi^\dagger)] \nonumber\\ &&-\frac{1}{2}\mbox{Tr}\{(\chi U^\dagger-U\chi^\dagger) [U (D^2 U)^\dagger- D^2 U U^\dagger]\},\end{aligned}$$ where “tot. der.” refers to a total-derivative term which has no dynamical significance. We make use of a trace relation for arbitrary $2\times 2$ matrices $A_i$ \[see Eqs. (\[4:10:F2\]) and (\[4:10:Tracesu2\])\], $$\begin{aligned} \label{app:glvgss:tr2by2} \lefteqn{\mbox{Tr}(A_1 A_2 A_3 + A_1 A_3 A_2) -\mbox{Tr}(A_1)\mbox{Tr}(A_2 A_3) -\mbox{Tr}(A_2)\mbox{Tr}(A_3 A_1)}\nonumber\\ && -\mbox{Tr}(A_3)\mbox{Tr}(A_1 A_2) +\mbox{Tr}(A_1)\mbox{Tr}(A_2)\mbox{Tr}(A_3)=0,\end{aligned}$$ and $\mbox{Tr}(D_\mu U U^\dagger)=0$ \[see Eq.  (\[4:5:kauprop2\])\] to rewrite the first term of (\[app:glvgss:intmed1\]) as the product of two trace terms, $$\mbox{Tr}[D_\mu U (D^\mu U)^\dagger(\chi U^\dagger + U\chi^\dagger)] =\frac{1}{2}\mbox{Tr}[D_\mu U (D^\mu U)^\dagger] \mbox{Tr}(\chi U^\dagger + U\chi^\dagger).$$ By adding and subtracting appropriate $\chi$ terms to generate an expression proportional to the lowest-order equation of motion which, for SU(2)$\times$SU(2), reads \[see Eq. (\[4:5:eom\])\] $${\cal O}^{(2)}_{\rm EOM}(U)=D^2 U U^\dagger - U(D^2 U)^\dagger -\chi U^\dagger + U\chi^\dagger+\frac{1}{2}\mbox{Tr}(\chi U^\dagger -U\chi^\dagger)=0,$$ we re-express the last term of Eq. (\[app:glvgss:intmed1\]) as $$\begin{aligned} \label{app:glvgss:intmed2} \lefteqn{-\frac{1}{2}\mbox{Tr}\{(\chi U^\dagger-U\chi^\dagger) [U (D^2 U)^\dagger- D^2 U U^\dagger]\}=}\nonumber\\ &&+\frac{1}{2}\mbox{Tr}[(\chi U^\dagger-U\chi^\dagger) {\cal O}^{(2)}_{\rm EOM}(U)] \nonumber\\ &&+\frac{1}{2}\mbox{Tr}[(\chi U^\dagger- U\chi^\dagger) (\chi U^\dagger- U\chi^\dagger)]-\frac{1}{4}[\mbox{Tr} (\chi U^\dagger- U\chi^\dagger)]^2.\end{aligned}$$ The $l_4$ term can thus be written as $$\begin{aligned} \label{app:glvgss:intmed3} \lefteqn{\mbox{Tr}[D_\mu\chi (D^\mu U)^\dagger+D^\mu U (D_\mu\chi)^\dagger] =}\nonumber\\ &&\mbox{tot. der.}+ \frac{1}{2}\mbox{Tr}[D_\mu U (D^\mu U)^\dagger] \mbox{Tr}(\chi U^\dagger+ U\chi^\dagger) +\frac{1}{2}\mbox{Tr}(\chi U^\dagger\chi U^\dagger+U\chi^\dagger U\chi^\dagger) \nonumber\\ &&-\mbox{Tr}(\chi\chi^\dagger) -\frac{1}{4}[\mbox{Tr} (\chi U^\dagger- U\chi^\dagger)]^2 +\frac{1}{2}\mbox{Tr}[(\chi U^\dagger-U\chi^\dagger){\cal O}^{(2)}_{\rm EOM}],\end{aligned}$$ which, except for the total derivative and the equation-of-motion term, is the same as Eq. (5.9) of Gasser, Sainio, and Švarc (GSS) [@Gasser:1987rb]. The difference between the Lagrangians of [@Gasser:1983yg] and [@Gasser:1987rb] then reads $$\begin{aligned} \label{app:glvgss:diffl} {\cal L}_4^{\rm GL}-{\cal L}^{\rm GSS}_4&=&\frac{l_4}{4}\left\{\mbox{tot. der.} +\frac{1}{2}\mbox{Tr}\left[(\chi U^\dagger - U\chi^\dagger) {\cal O}^{(2)}_{\rm EOM} \right]\right\},\end{aligned}$$ which agrees with Eq. (26) of Ecker and Mojžiš [@Ecker:1995rk] once their expressions are rewritten in the above notation. Let us also specify the field transformation required to connect the two Lagrangians. For that purpose we rewrite Eq. (\[app:glvgss:diffl\]) in accord with Eq. (2.11) of [@Scherer:1994wi], $$\begin{aligned} \label{app:glvgss:diffl2} {\cal L}_4^{\rm GSS}(U)&=&{\cal L}_4^{\rm GL}(U) + \mbox{tot. der.} -\frac{l_4}{8}\mbox{Tr}[(\chi U^\dagger - U\chi^\dagger) {\cal O}^{(2)}_{\rm EOM}].\end{aligned}$$ According to Eqs. (\[4:7:dl2\]) and (\[4:7:addstruc2\]) we need to insert $\alpha_1=0$ and $\alpha_2=-l_4/(2F_0^2)$ in Eq.(\[4:7:s2\]) in order to relate the two Lagrangians. Finally, making use of Eqs. (\[app:glvgss:intmed3\]) and (\[app:glvgss:trick\]) and dropping the total derivative as well as the equation-of-motion term let us explicitly write out the GSS Lagrangian: $$\begin{aligned} \label{app:glvgss:l4gss} \lefteqn{{\cal L}^{\rm GSS}_4 = \frac{l_1}{4} \left\{\mbox{Tr}[D_{\mu}U (D^{\mu}U)^{\dagger}] \right\}^2 +\frac{l_2}{4}\mbox{Tr}[D_{\mu}U (D_{\nu}U)^{\dagger}] \mbox{Tr}[D^{\mu}U (D^{\nu}U)^{\dagger}] \nonumber} \\ &&+\frac{l_3+l_4}{16}\left[\mbox{Tr}(\chi U^\dagger+ U\chi^\dagger)\right]^2 +\frac{l_4}{8}\mbox{Tr}[D_\mu U(D^\mu U)^\dagger]\mbox{Tr}(\chi U^\dagger +U\chi^\dagger)\nonumber\\ &&+l_5\mbox{Tr}(f^R_{\mu\nu}U f^{\mu\nu}_LU^\dagger) +i\frac{l_6}{2}\mbox{Tr}[ f^R_{\mu\nu} D^{\mu} U (D^{\nu} U)^{\dagger} + f^L_{\mu\nu} (D^{\mu} U)^{\dagger} D^{\nu} U]\nonumber\\ &&-\frac{l_7}{16}\left[\mbox{Tr}(\chi U^\dagger-U\chi^\dagger)\right]^2 +\frac{h_1+h_3-l_4}{4}\mbox{Tr}(\chi\chi^\dagger)\nonumber\\ &&+\frac{h_1-h_3-l_4}{16}\left\{ \left[\mbox{Tr}(\chi U^\dagger + U\chi^\dagger)\right]^2 +\left[\mbox{Tr}(\chi U^\dagger-U\chi^\dagger)\right]^2\right.\nonumber\\ &&\left. -2\mbox{Tr}(\chi U^\dagger\chi U^\dagger + U\chi^\dagger U\chi^\dagger) \right\} -\frac{4h_2+l_5}{2}\mbox{Tr}(f_{\mu\nu}^L f^{\mu\nu}_L +f_{\mu\nu}^R f^{\mu\nu}_R).\end{aligned}$$ Different Parameterizations {#app_sec_dp} --------------------------- In App. \[app\_sec\_glvgss\] we saw that two versions of the ${\cal O}(p^4)$ $\mbox{SU(2)}\times\mbox{SU(2)}$ mesonic Lagrangian, Eqs. (\[app:glvgss:l4gl\]) and (\[app:glvgss:l4gss\]), are used in the literature. Since they are related by a field transformation, they must yield the same results for physical observables [@Chisholm; @Kamefuchi:sb]. Furthermore, in SU(2)$\times$SU(2) two different parameterizations of the SU(2) matrix $U(x)$ \[see Eqs. (\[4:6:u1\]) and (\[4:6:u2\])\] are popular, $$\begin{aligned} \label{app:dp:exp} U(x)&=&\exp\left[i\frac{\vec{\tau}\cdot\vec{\phi}(x)}{F_0}\right],\\ \label{app:dp:sqrt} U(x)&=&\frac{1}{F_0}\left[\sigma(x)+i\vec{\tau}\cdot\vec{\pi}(x) \right],\quad \sigma(x)=\sqrt{F^2_0-\vec{\pi}\,^2(x)},\end{aligned}$$ where the pion fields of the two parameterizations are non-linearly related \[see Eq. (\[4:6:ft\])\]. In this appendix we collect the pion wave function renormalization constants entering a calculation at ${\cal O}(p^4)$ depending on which Lagrangian and parameterization is used. The actual calculation parallels that of Sec. \[subsec\_mgb\] and will not be repeated here. The self energies up to ${\cal O}(p^4)$ can be written as $$\label{app:dp:selfenergy} \Sigma(p^2) = A + B p^2.$$ The renormalized mass and the wave function renormalization constant are, respectively to ${\cal O}(p^4)$ and ${\cal O}(p^2)$, given by $$\begin{aligned} \label{app:dp:mass} M^2_{\pi,4} &=& M^2_{\pi,2}(1+B)+A,\\ \label{app:dp:Z} Z&=&1+B,\end{aligned}$$ where $M^2_{\pi,2}=2B_0 m$ is the prediction at ${\cal O}(p^2)$. The different values for $A$, $B$, and $Z$ are given in Table \[app:dp:tab:abz\]. Note that the result for the pion mass is, as expected, independent of the Lagrangian and parameterization used: $$\label{app:dp:Mpi2} M_{\pi,4}^2=M^2\left(1+\frac{2}{3} \frac{I}{F_0^2}\right) -\frac{1}{6}\frac{M^2}{F_0^2}I +2 l_3 \frac{M^4}{F_0^2} =M^2-\frac{\bar{l}_3}{32\pi^2 F_0^2} M^4,$$ where $M^2=2 B_0 m$ and $$l_3=-\frac{1}{64\pi^2}\left[\bar{l}_3+\ln\left(\frac{M^2}{\mu^2}\right) +R\right]$$ \[see Eqs. (\[app:lilir\]) and (\[app:barli\])\]. On the other hand, the constants $A$, $B$, and $Z$ are auxiliary mathematical quantities and thus depend on both Lagrangian and parameterization. [|c|c|c|c|]{} &$A$&$B$&$Z$\ &&&\ GL, Eq. (\[app:dp:exp\])& $-\frac{1}{6}\frac{M^2}{F^2_0} I+2 l_3 \frac{M^4}{F^2_0}$& $\frac{2}{3}\frac{I}{F^2_0}$& $1+\frac{2}{3}\frac{I}{F^2_0}$\ &&&\ &&&\ GL, Eq. (\[app:dp:sqrt\])& $\frac{3}{2}\frac{M^2}{F^2_0} I+2 l_3 \frac{M^4}{F^2_0}$& $-\frac{I}{F^2_0}$& $1-\frac{I}{F^2_0}$\ &&&\ &&&\ GSS, Eq. (\[app:dp:exp\])& $-\frac{1}{6}\frac{M^2}{F^2_0} I+2 (l_3+l_4)\frac{M^4}{F^2_0}$& $\frac{2}{3}\frac{I}{F^2_0}-2l_4\frac{M^2}{F^2_0}$& $1+\frac{2}{3}\frac{I}{F^2_0}-2l_4\frac{M^2}{F^2_0}$\ &&&\ &&&\ GSS, Eq. (\[app:dp:sqrt\])& $\frac{3}{2}\frac{M^2}{F^2_0} I+2 (l_3+l_4)\frac{M^4}{F^2_0}$& $-\frac{I}{F^2_0}-2l_4\frac{M^2}{F^2_0}$& $1-\frac{I}{F^2_0}-2l_4\frac{M^2}{F^2_0}$\ &&&\ [Hem+ 97a]{} R. Akhoury and A. Alfakih, Annals Phys.  [**210**]{}, 81 (1991). S. L. Adler and W. A. Bardeen, Phys. Rev.  [**182**]{}, 1517 (1969). S. L. Adler and R. F. Dashen, [*Current Algebras and Applications to Particle Physics*]{} (Benjamin, New York, 1968). S. L. Adler, Phys. Rev. [**139**]{}, B1638 (1965). S. L. Adler, Phys. Rev.  [**177**]{}, 2426 (1969). S. L. Adler, in [*Lectures on Elementary Particles and Quantum Field Theory*]{}, 1970 Brandeis University Summer Institute in Theoretical Physics, Volume 1, edited by S. Deser, M. , and H. Pendleton (M.I.T. Press, Cambridge, Massachusetts, 1970). E. S. Abers and B. W. Lee, Phys. Rept.  [**9**]{}, 1 (1973). V. de Alfaro, S. Fubini, G. Furlan, and C. Rossetti, [*Currents in Hadron Physics*]{} (North-Holland, Amsterdam, 1973). A. Ali Khan [*et al.*]{} \[CP-PACS Collaboration\], Phys. Rev. D [**65**]{}, 054505 (2002). G. Altarelli, Phys. Rept.  [**81**]{}, 1 (1982). G. Altarelli, Ann. Rev. Nucl. Part. Sci.  [**39**]{}, 357 (1989). N. W. Ashcroft and N. D. Mermin, [*Solid State Physics*]{} ( College, Philadelphia, 1976) Chap. 32. B. Ananthanarayan and B. Moussallam, JHEP [**0205**]{}, 052 (2002). S. R. Amendolia [*et al.*]{}, Phys. Lett. B [**178**]{}, 435 (1986). S. R. Amendolia [*et al.*]{} \[NA7 Collaboration\], Nucl. Phys.  [**B277**]{}, 168 (1986). L. Ametller, J. Bijnens, A. Bramon, and F. Cornet, Phys. Lett. B [**276**]{}, 185 (1992). G. Amoros, J. Bijnens, and P. Talavera, Nucl. Phys.  [**B568**]{}, 319 (2000). G. Amoros, J. Bijnens, and P. Talavera, Phys. Lett. B [**480**]{}, 71 (2000). B. Ananthanarayan, G. Colangelo, J. Gasser, and H. Leutwyler, Phys. Rept.  [**353**]{}, 207 (2001). C. Arzt, Phys. Lett. B [**342**]{}, 189 (1995). M. Abramowitz and I. A. Stegun (Eds.), [*Handbook of Mathematical Functions*]{} (Dover, New York, 1972). A. P. Balachandran, G. Marmo, B. S. Skagerstam, and A. Stern, [*Classical Topology and Quantum States*]{} (World Scientific, , 1991) Chap. 12.2. S. Balk, J. G. Körner, and D. Pirjol, Nucl. Phys.  [**B428**]{}, 499 (1994). W. A. Bardeen, Phys. Rev.  [**184**]{}, 1848 (1969). S. Bellucci and C. Bruno, Nucl. Phys.  [**B452**]{}, 626 (1995). J. Bijnens and F. Cornet, Nucl. Phys.  [**B296**]{}, 557 (1988). J. D. Bjorken and S. D. Drell, [*Relativistic Quantum Mechanics*]{} (McGraw-Hill, New York, 1964). J. D. Bjorken and S. D. Drell, [*Relativistic Quantum Fields*]{} (McGraw-Hill, New York, 1964). S. R. Beane, P. F. Bedaque, M. J. Savage, and U. van Kolck, Nucl. Phys.  [**A700**]{}, 377 (2002). T. Becher, [*Lorentz Invariant Baryon CHPT*]{}, in [@Bernstein:2002]. M. A. B. Bég, B. W. Lee, and A. Pais, Phys. Rev. Lett.  [**13**]{}, 514 (1964). S. Bellucci, J. Gasser, and M. E. Sainio, Nucl. Phys.  [**B423**]{}, 80 (1994) \[Erratum, ibid.  [**B431**]{}, 413 (1994)\]. A. A. Belkov, A. V. Lanyov, A. Schaale, and S. Scherer, Acta Phys. Slov.  [**45**]{}, 121 (1995). A. A. Bel’kov, A. V. Lanyov, and S. Scherer, J. Phys. G [**22**]{}, 1383 (1996). J. Bernstein, Rev. Mod. Phys. [**46**]{}, 7 (1974) \[Erratum, ibid. [**47**]{}, 259 (1975)\]. V. Bernard, N. Kaiser and U.-G. Mei[ß]{}ner, Nucl. Phys. [**B383**]{}, 442 (1992). V. Bernard, N. Kaiser, J. Kambor, and U.-G. Mei[ß]{}ner, Nucl. Phys.  [**B388**]{}, 315 (1992). V. Bernard, N. Kaiser, and U.-G. Mei[ß]{}ner, Phys. Rev. Lett.  [**69**]{}, 1877 (1992). V. Bernard, N. Kaiser, J. Kambor, and U.-G. Mei[ß]{}ner, Phys. Rev. D [**46**]{}, 2756 (1992). V. Bernard, N. Kaiser, and U.-G. Mei[ß]{}ner, Z. Phys. C [**60**]{}, 111 (1993). V. Bernard, N. Kaiser, A. Schmidt, and U.-G. Mei[ß]{}ner, Phys. Lett. B [**319**]{}, 269 (1993). V. Bernard, N. Kaiser, T. S. Lee, and U.-G. Mei[ß]{}ner, Phys. Rept.  [**246**]{}, 315 (1994). V. Bernard, N. Kaiser, and U.-G. Mei[ß]{}ner, Phys. Rev. Lett.  [**74**]{}, 3752 (1995). V. Bernard, N. Kaiser, and U.-G. Mei[ß]{}ner, Int. J. Mod. Phys. E [**4**]{}, 193 (1995). V. Bernard, N. Kaiser, and U.-G. Mei[ß]{}ner, Phys. Rev. C [**52**]{}, 2185 (1995). V. Bernard, N. Kaiser, and U.-G. Mei[ß]{}ner, Phys. Lett. B [**383**]{}, 116 (1996). V. Bernard, N. Kaiser, and U. G. Mei[ß]{}ner, Nucl. Phys.  [**A611**]{}, 429 (1996). V. Bernard, N. Kaiser, and U.-G. Mei[ß]{}ner, Nucl. Phys.  [**A615**]{}, 483 (1997). A. M. Bernstein, D. Drechsel, and Th. Walcher (Eds.), [*Chiral Dynamics: Theory and Experiment*]{}. Proceedings, Workshop, Mainz, Germany, 1 - 5 September, 1997, (Springer, Berlin, 1998, Lecture Notes in Physics, Vol. 513). V. Bernard, H. W. Fearing, T. R. Hemmert, and U.-G. Mei[ß]{}ner, Nucl. Phys.  [**A635**]{}, 121 (1998) \[Erratum, ibid.  [**A642**]{}, 563 (1998)\]. V. Bernard, N. Kaiser, and U.-G. Mei[ß]{}ner, Phys. Rev. C [**62**]{}, 028201 (2000). A. M. Bernstein, J. L. Goity, and U.-G. Mei[ß]{}ner (Eds.), [*Chiral Dynamics: Theory and Experiment III*]{}. Proceedings, Workshop, Jefferson Laboratory, USA, 17 - 20 July, 2000 (World Scientific, Singapore, 2002). V. Bernard, T. R. Hemmert, and U.-G. Mei[ß]{}ner, Phys. Lett. B [**545**]{}, 105 (2002). H. A. Bethe and F. de Hoffmann, [*Mesons and Fields, Volume II, Mesons*]{} (Row, Peterson and Company, Evanston, 1955). A. M. Bernstein and B. R. Holstein (Eds.), [*Chiral Dynamics: Theory and Experiment*]{}. Proceedings, Workshop, Cambridge, USA, 25 - 29 July, 1994 (Springer, Berlin, 1995, Lecture Notes in Physics, Vol. 452). R. K. Bhaduri, [*Models of the Nucleon*]{} (Addison-Wesley, Redwood City, 1988). J. Bijnens, A. Bramon, and F. Cornet, Z. Phys. C [**46**]{}, 599 (1990). J. Bijnens, C. Bruno, and E. de Rafael, Nucl. Phys.  [**B390**]{}, 501 (1993). J. Bijnens, Int. J. Mod. Phys. A [**8**]{}, 3045 (1993). J. Bijnens, G. Colangelo, and J. Gasser, Nucl. Phys.  [**B427**]{}, 427 (1994). J. Bijnens, G. Ecker, and J. Gasser, in [*The Second DA$\Phi$NE Physics Handbook*]{}, edited by L. Maiani, G. Pancheri, and N. Paver (Frascati, Italy, 1995). J. Bijnens, J. Prades, and E. de Rafael, Phys. Lett. B [**348**]{}, 226 (1995). J. Bijnens, G. Colangelo, G. Ecker, J. Gasser, and M. E. Sainio, Phys. Lett. B [**374**]{}, 210 (1996). J. Bijnens, G. Colangelo, G. Ecker, J. Gasser, and M. E. Sainio, Nucl. Phys.  [**B508**]{}, 263 (1997) \[Erratum, ibid.  [**B517**]{}, 639 (1997)\]. J. Bijnens, G. Colangelo, and P. Talavera, JHEP [**9805**]{}, 014 (1998). J. Bijnens, G. Colangelo, and G. Ecker, JHEP [**9902**]{}, 020 (1999). J. Bijnens, G. Colangelo, and G. Ecker, Annals Phys.  [**280**]{}, 100 (2000). J. Bijnens, L. Girlanda, and P. Talavera, Eur. Phys. J. C [**23**]{}, 539 (2002). M. C. Birse, X. D. Ji, and J. A. McGovern, Phys. Rev. Lett.  [**86**]{}, 3204 (2001). J. S. Bell and R. Jackiw, Nuovo Cim. A [**60**]{}, 47 (1969). T. Becher and H. Leutwyler, Eur. Phys. J. C [**9**]{}, 643 (1999). T. Becher and H. Leutwyler, JHEP [**0106**]{}, 017 (2001). B. Borasoy and U.-G. Mei[ß]{}ner, Annals Phys.  [**254**]{}, 192 (1997). P. B[ü]{}ttiker and U.-G. Mei[ß]{}ner, Nucl. Phys. [**A668**]{}, 97 (2000). H.-J. Borchers, Nuovo Cim. [**15**]{}, 784 (1960). B. Borasoy, Phys. Rev. D [**59**]{}, 054021 (1999). B. Borasoy, Phys. Rev. D [**61**]{}, 114017 (2000). L. S. Brown, W. J. Pardee, and R. D. Peccei, Phys. Rev. D [**4**]{}, 2801 (1971). A. P. Balachandran and C. G. Trahern, [*Lectures On Group Theory For Physicists*]{} (Bibliopolis, Naples, 1984). J. Bijnens and P. Talavera, Nucl. Phys.  [**B489**]{}, 387 (1997). J. Bijnens and P. Talavera, JHEP [**0203**]{}, 046 (2002). U. Bürgi, Phys. Lett. B [**377**]{}, 147 (1996); Nucl. Phys.  [**B479**]{}, 392 (1996). C. P. Burgess, Phys. Rept.  [**330**]{}, 193 (2000). F. Butler, H. Chen, J. Sexton, A. Vaccarino, and D. Weingarten, Nucl. Phys. [**B430**]{}, 179 (1994). C. G. Callan, S. R. Coleman, J. Wess, and B. Zumino, Phys. Rev.  [**177**]{}, 2247 (1969). S. R. Coleman and D. J. Gross, Phys. Rev. Lett.  [**31**]{}, 851 (1973). G. F. Chew, M. L. Goldberger, F. E. Low, and Y. Nambu, Phys. Rev. [**106**]{}, 1337 (1957). Y.-Q. Chen, Phys. Lett. B [**318**]{}, 524 (1993). J. Chisholm, Nucl. Phys. [**26**]{}, 469 (1961). K.-c. Chou, H.-y. Guo, K. Wu, and X.-c. Song, Phys. Lett. B [**134**]{}, 67 (1984). T. P. Cheng and L. F. Li, [*Gauge Theory of Elementary Particle Physics*]{} (Clarendon, Oxford, 1984). See http://www.claymath.org/prizeproblems/index.htm. S. Coleman, J. Math. Phys. [**7**]{}, 787 (1966). J. C. Collins, [*Renormalization*]{} (Cambridge University Press, Cambridge, 1984). S. R. Coleman, J. Wess, and B. Zumino, Phys. Rev.  [**177**]{}, 2239 (1969). G. Colangelo, M. Finkemeier, and R. Urech, Phys. Rev. D [**54**]{}, 4403 (1996). G. Colangelo, J. Gasser, and H. Leutwyler, Phys. Rev. Lett.  [**86**]{}, 5008 (2001). G. Colangelo, J. Gasser, and H. Leutwyler, Nucl. Phys. [**B603**]{}, 125 (2001). R. E. Cutkosky, J. Math. Phys.  [**1**]{}, 429 (1960). E. B. Dally [*et al.*]{}, Phys. Rev. Lett.  [**45**]{}, 232 (1980). E. B. Dally [*et al.*]{}, Phys. Rev. Lett.  [**48**]{}, 375 (1982). R. F. Dashen, Phys. Rev.  [**183**]{}, 1245 (1969). A. K. Das, [*Field Theory: A Path Integral Approach*]{} (World Scientific, Singapore, 1993). A. K. Das, Mod. Phys. Lett. A [**9**]{}, 341 (1994). J. F. Donoghue and B. R. Holstein, Phys. Rev. D [**48**]{}, 137 (1993). S. Dürr and J. Kambor, Phys. Rev. D [**61**]{}, 114025 (2000). J. F. Donoghue, B. R. Holstein, and Y. C. Lin, Phys. Rev. D [**37**]{}, 2423 (1988). J. F. Donoghue, C. Ramirez, and G. Valencia, Phys. Rev. D [**39**]{}, 1947 (1989). J. F. Donoghue, E. Golowich, and B. R. Holstein, [*Dynamics of the Standard Model*]{} (Cambridge University Press, Cambridge, 1992). N. A. Doughty, [*Lagrangian Interaction*]{} (Addison Wesley, Sydney, 1990), Chaps. 3.3 and 20.7. D. Drechsel, G. Knöchlein, A. Metz, and S. Scherer, Phys. Rev. C [**55**]{}, 424 (1997). M. J. Dugan, M. Golden, and B. Grinstein, Phys. Lett. B [**282**]{}, 142 (1992). R. F. Dashen and M. Weinstein, Phys. Rev.  [**183**]{}, 1261 (1969). J. F. Donoghue and D. Wyler, Nucl. Phys.  [**B316**]{}, 289 (1989). T. Ebertshäuser, [*Mesonic Chiral Perturbation Theory: Odd Intrinsic Parity Sector*]{}, PhD thesis, Johannes Gutenberg-Universität, Mainz, Germany, 2001, http://archimed.uni-mainz.de/. D. Ebert, A. A. Belkov, A. V. Lanyov, and A. Schaale, Int. J. Mod. Phys. A [**8**]{}, 1313 (1993). T. Ebertshäuser, H. W. Fearing, and S. Scherer, Phys. Rev. D [**65**]{}, 054033 (2002). G. Ecker, Prog. Part. Nucl. Phys.  [**35**]{}, 1 (1995). G. Ecker, J. Gasser, H. Leutwyler, A. Pich, and E. de Rafael, Phys. Lett. B [**223**]{}, 425 (1989). G. Ecker, J. Gasser, A. Pich, and E. de Rafael, Nucl. Phys.  [**B321**]{}, 311 (1989). G. Ecker and M. Mojžiš, Phys. Lett. B [**365**]{}, 312 (1996). G. Ecker and M. Mojžiš, Phys. Lett. B [**410**]{}, 266 (1997). E. Epelbaum, W. Glöckle, and U.-G. Mei[ß]{}ner, Nucl. Phys.  [**A671**]{}, 295 (2000). D. Ebert and H. Reinhardt, Nucl. Phys.  [**B271**]{}, 188 (1986). D. Espriu, E. de Rafael, and J. Taron, Nucl. Phys.  [**B345**]{}, 22 (1990) \[Erratum, ibid.  [**B355**]{}, 278 (1990)\]. P. J. Ellis and H. B. Tang, Phys. Rev. C [**57**]{}, 3356 (1998). P. J. Ellis and K. Torikoshi, Phys. Rev. C [**61**]{}, 015205 (2000). T. E. Ericson and W. Weise, [*Pions and Nuclei*]{} (Clarendon, , 1988) Appendices 3 and 8. H. W. Fearing, G. I. Poulis, and S. Scherer, Nucl. Phys.  [**A570**]{}, 657 (1994). H. W. Fearing, R. Lewis, N. Mobed, and S. Scherer, Phys. Rev. D [**56**]{}, 1783 (1997). H. W. Fearing, T. R. Hemmert, R. Lewis, and C. Unkmeir, Phys. Rev. C [**62**]{}, 054006 (2000). N. Fettes, U.-G. Mei[ß]{}ner, and S. Steininger, Nucl. Phys. [**A640**]{}, 199 (1998). N. Fettes, U.-G. Mei[ß]{}ner, M. Mojžiš, and S. Steininger, Annals Phys.  [**283**]{}, 273 (2001) \[Erratum, ibid.  [**288**]{}, 249 (2001)\]. M. Finkemeier, H. Georgi, and M. McIrvin, Phys. Rev. D [**55**]{}, 6933 (1997). P. Finelli, N. Kaiser, D. Vretenar, and W. Weise, nucl-th/0205016. N. Fettes and U.-G. Mei[ß]{}ner, Nucl. Phys. [**A676**]{}, 311 (2000). N. Fettes and U.-G. Mei[ß]{}ner, Nucl. Phys.  [**A693**]{}, 693 (2001). C. D. Froggatt and J. L. Petersen, Nucl. Phys.  [**B129**]{}, 89 (1977). E. S. Fradkin, Zh. Eksp. Teor. Fiz.  [**29**]{}, 258 (1955) \[Sov. Phys. JETP [**2**]{}, 361 (1955)\]. H. Fritzsch, M. Gell-Mann, and H. Leutwyler, Phys. Lett. B [**47**]{}, 365 (1973). H. W. Fearing and S. Scherer, Phys. Rev. D [**53**]{}, 315 (1996). H. W. Fearing and S. Scherer, Phys. Rev. C [**62**]{}, 034003 (2000). L. L. Foldy and S. A. Wouthuysen, Phys. Rev.  [**78**]{}, 29 (1950). J. Gasser, M. E. Sainio, and A. Švarc, Nucl. Phys. [**B307**]{}, 779 (1988). J. Gasser, [*QCD at Low Energies*]{}, lectures given as part of the “Cours du Troisième Cycle de la Physique en Suisse Romande”, Lausanne, Switzerland, January and February 1989. J. Gegelia, G. Japaridze, and X. Q. Wang, hep-ph/9910260. M. Gell-Mann, Phys. Rev.  [**125**]{}, 1067 (1962). M. Gell-Mann, Phys. Lett.  [**8**]{}, 214 (1964). M. Gell-Mann, Physics [**1**]{}, 63 (1964). M. Gell-Mann, R. J. Oakes, and B. Renner, Phys. Rev.  [**175**]{}, 2195 (1968). G. C. Gellas, T. R. Hemmert, C. N. Ktorides, and G. I. Poulis, Phys. Rev. D [**60**]{}, 054022 (1999). G. C. Gellas, T. R. Hemmert, and U.-G. Mei[ß]{}ner, Phys. Rev. Lett.  [**85**]{}, 14 (2000). H. Georgi, [*Weak Interactions and Modern Particle Theory*]{} (Benjamin/Cummings, Menlo Park, 1984). H. Georgi, Nucl. Phys. [**B361**]{}, 339 (1991). S. Gasiorowicz and D. A. Geffen, Rev. Mod. Phys.  [**41**]{}, 531 (1969). H. W. Grie[ß]{}hammer and T. R. Hemmert, Phys. Rev. C [**65**]{}, 045207 (2002). M. M. Giannini, Rept. Prog. Phys.  [**54**]{}, 453 (1990). J. Gegelia and G. Japaridze, Phys. Rev. D [**60**]{}, 114038 (1999). E. Golowich and J. Kambor, Nucl. Phys.  [**B447**]{}, 373 (1995). M. Gell-Mann and F. Low, Phys. Rev.  [**84**]{}, 350 (1951). M. Gell-Mann and M. L[é]{}vy, Nuovo Cim.  [**16**]{}, 705 (1960). J. Gasser and H. Leutwyler, Phys. Rept.  [**87**]{}, 77 (1982). J. Gasser and H. Leutwyler, Annals Phys.  [**158**]{}, 142 (1984). J. Gasser and H. Leutwyler, Nucl. Phys. [**B250**]{}, 465 (1985). J. Gasser and H. Leutwyler, Nucl. Phys. [**B250**]{}, 517 (1985). J. Gasser and H. Leutwyler, Nucl. Phys. [**B250**]{}, 539 (1985). M. Gell-Mann and Y. Ne’eman, [*The Eightfold Way*]{} (Benjamin, New York, 1964). J. Goldstone, Nuovo Cim.  [**19**]{}, 154 (1961). J. Goldstone, A. Salam, and S. Weinberg, Phys. Rev.  [**127**]{}, 965 (1962). M. Golterman, [*Chiral perturbation theory, non-leptonic kaon decays, and the lattice*]{}, in [@Bernstein:2002]. O. W. Greenberg, Phys. Rev. Lett.  [**13**]{}, 598 (1964). W. Greiner, [*Theoretical Physics. Vol. 4a: Quantum Theory (In German)*]{} (Deutsch, Thun, 1985). C. Grosse-Knetter, Phys. Rev. D [**49**]{}, 6709 (1994). D. J. Gross, Nucl. Phys. Proc. Suppl.  [**74**]{}, 426 (1999). D. E. Groom [*et al.*]{} \[Particle Data Group Collaboration\], Eur. Phys. J. C [**15**]{}, 1 (2000). J. Gasser and M. E. Sainio, Eur. Phys. J. C [**6**]{}, 297 (1999). M. L. Goldberger and S. B. Treiman, Phys. Rev.  [**110**]{}, 1178 (1958). M. L. Goldberger and S. B. Treiman, Phys. Rev.  [**111**]{}, 354 (1958). P. A. Guichon, G. Q. Liu, and A. W. Thomas, Nucl. Phys.  [**A591**]{}, 606 (1995). S. L. Glashow and S. Weinberg, Phys. Rev. Lett.  [**20**]{}, 224 (1968). D. J. Gross and F. Wilczek, Phys. Rev. Lett.  [**30**]{}, 1343 (1973). D. J. Gross and F. Wilczek, Phys. Rev. D [**8**]{}, 3633 (1973). K. Gottfried and V. F. Weisskopf, [*Concepts of Particle Physics, Vol. II*]{} (Oxford University Press, New York, 1986). R. Haag, [*Local Quantum Physics: Fields, Particles, Algebras*]{} (Springer, Berlin, 1992). T. Hannah, Nucl. Phys.  [**B593**]{}, 577 (2001). P. G. Harris [*et al.*]{}, Phys. Rev. Lett.  [**82**]{}, 904 (1999). T. R. Hemmert, B. R. Holstein, G. Knöchlein, and S. Scherer, Phys. Rev. D [**55**]{}, 2630 (1997). T. R. Hemmert, B. R. Holstein, G. Knöchlein, and S. Scherer, Phys. Rev. Lett.  [**79**]{}, 22 (1997). T. R. Hemmert, B. R. Holstein, and J. Kambor, Phys. Lett. B [**395**]{}, 89 (1997); J. Phys. G [**24**]{}, 1831 (1998). T. R. Hemmert, B. R. Holstein, J. Kambor, and G. Knöchlein, Phys. Rev. D [**57**]{}, 5746 (1998). T. R. Hemmert, B. R. Holstein, G. Knöchlein, and D. Drechsel, Phys. Rev. D [**62**]{}, 014013 (2000). E. L. Hill, Rev. Mod. Phys. [**23**]{}, 253 (1951). M. Y. Han and Y. Nambu, Phys. Rev.  [**139**]{}, B1006 (1965). B. R. Holstein, Am. J. Phys. [**65**]{}, 519 (1997). G. ’t Hooft, Nucl. Phys.  [**B72**]{}, 461 (1974). G. Holzwarth and B. Schwesinger, Rept. Prog. Phys.  [**49**]{}, 825 (1986). G. ’t Hooft and M. J. Veltman, Nucl. Phys.  [**B44**]{}, 189 (1972). G. ’t Hooft and M. J. Veltman, Nucl. Phys.  [**B153**]{}, 365 (1979). E. D’Hoker and S. Weinberg, Phys. Rev. D [**50**]{}, 6050 (1994). D. Issler, SLAC-PUB-4943-REV (1990) (unpublished). C. Itzykson and J. B. Zuber, [*Quantum Field Theory*]{} (McGraw-Hill, New York, 1980). R. Jackiw, [*Field Theoretic Investigations in Current Algebra*]{}, in Ref. [@Treiman:1972]. F. Jegerlehner, Eur. Phys. J. C [**18**]{}, 673 (2001). E. Jenkins, Nucl. Phys. [**B368**]{}, 190 (1992). M. Jetter, Nucl. Phys.  [**B459**]{}, 283 (1996). E. Jenkins and A. V. Manohar, Phys. Lett. B [**255**]{}, 558 (1991). H. F. Jones, [*Groups, Representations and Physics*]{} (Hilger, , 1990). A. Jaffe and E. Witten, [*Quantum Yang-Mills Theory*]{}, http://www.claymath.org/prizeproblems/yangmills.htm. N. Kaiser, R. Brockmann, and W. Weise, Nucl. Phys.  [**A625**]{}, 758 (1997). S. Kamefuchi, L. O’Raifeartaigh, and A. Salam, Nucl. Phys.  [**28**]{}, 529 (1961). M. Kermani [*et al.*]{} \[CHAOS Collaboration\], Phys. Rev. C [**58**]{}, 3431 (1998). R. Kaiser and H. Leutwyler, Eur. Phys. J. C [**17**]{}, 623 (2000). J. Kambor and M. Mojžiš, JHEP [**9904**]{}, 031 (1999). B. Kubis and U.-G. Mei[ß]{}ner, Nucl. Phys. [**A679**]{}, 698 (2001). B. Kubis and U.-G. Mei[ß]{}ner, Eur. Phys. J. C [**18**]{}, 747 (2001). M. Knecht and A. Nyffeler, Eur. Phys. J. C [**21**]{}, 659 (2001). M. Knecht, B. Moussallam, and J. Stern, Nucl. Phys.  [**B429**]{}, 125 (1994). M. Knecht, B. Moussallam, J. Stern, and N. H. Fuchs, Nucl. Phys. [**B457**]{}, 513 (1995). M. Knecht, B. Moussallam, J. Stern, and N. H. Fuchs, Nucl. Phys.  [**B471**]{}, 445 (1996). P. Ko, Phys. Lett. B [**349**]{}, 555 (1995). R. Koch, Nucl. Phys.  [**A448**]{}, 707 (1986). U. van Kolck, Prog. Part. Nucl. Phys.  [**43**]{}, 337 (1999). R. Koch and E. Pietarinen, Nucl. Phys.  [**A336**]{}, 331 (1980). A. Krause, Helv. Phys. Acta [**63**]{}, 3 (1990). B. Kubis, T. R. Hemmert, and U.-G. Mei[ß]{}ner, Phys. Lett. B [**456**]{}, 240 (1999). K. B. Kumar, J. A. McGovern, and M. C. Birse, Phys. Lett. B [**479**]{}, 167 (2000). M. Le Bellac, [*Quantum and Statistical Field Theory*]{} (Clarendon, Oxford, 1991). H. Lehmann, K. Symanzik, and W. Zimmermann, Nuovo Cim.  [**1**]{}, 205 (1955). G. Leibbrandt, Rev. Mod. Phys.  [**47**]{}, 849 (1975). H. Leutwyler, in [*Perspectives in the Standard Model*]{}, Proceedings of the 1991 Advanced Theoretical Study Institute in Elementary Particle Physics, Boulder, Colorado, 2 - 28 June, 1991, edited by R. K. Ellis, C. T. Hill, and J. D. Lykken (World Scientific, Singapore, 1992). H. Leutwyler, Annals Phys.  [**235**]{}, 165 (1994). H. Leutwyler, in [*Hadron Physics 94: Topics on the Structure and Interaction of Hadronic Systems*]{}, Proceedings, Workshop, Gramado, Brasil, edited by V. E. Herscovitz (World Scientific, Singapore, 1995). H. Leutwyler, Phys. Lett. B [**378**]{}, 313 (1996). H. Leutwyler, hep-ph/0107332. S. Leupold, hep-ph/0111204. A. Liesenfeld [*et al.*]{} \[A1 Collaboration\], Phys. Lett. B [**468**]{}, 20 (1999). M. F. Lutz and E. E. Kolomeitsev, Nucl. Phys. [**A700**]{}, 193 (2002). R. F. Lebed and M. A. Luty, Phys. Lett. B [**329**]{}, 479 (1994). M. E. Luke and A. V. Manohar, Phys. Lett. B [**286**]{}, 348 (1992). L. F. Li and H. Pagels, Phys. Rev. Lett.  [**26**]{}, 1204 (1971). M. Lutz, Nucl. Phys. [**A677**]{}, 241 (2000). A. I. L’vov, S. Scherer, B. Pasquini, C. Unkmeir, and D. , Phys. Rev. C [**64**]{}, 015203 (2001). K. Maltman, Phys. Rev. D [**53**]{}, 2573 (1996). J. L. Manes, Nucl. Phys. [**B250**]{}, 369 (1985). T. Mannel, W. Roberts, and Z. Ryzak, Nucl. Phys.  [**B368**]{}, 204 (1992). A. V. Manohar, [*Lectures given at 35th Int. Universitätswochen für Kern- und Teilchenphysik: Perturbative and Nonperturbative Aspects of Quantum Field Theory*]{}, Schladming, Austria, 2 - 9 March, 1996, hep-ph/9606222. A. Manohar, [*Quark Masses*]{}, in Ref. [@Groom:in]. H. Marsiske [*et al.*]{} \[Crystal Ball Collaboration\], Phys. Rev. D [**41**]{}, 3324 (1990). E. Matsinos, Phys. Rev. C [**56**]{}, 3014 (1997). J. A. McGovern and M. C. Birse, Phys. Lett. B [**446**]{}, 300 (1999). J. A. McGovern and M. C. Birse, Phys. Rev. D [**61**]{}, 017503 (2000). S. Myint and C. Rebbi, Nucl. Phys. Proc. Suppl.  [**34**]{}, 213 (1994). U.-G. Mei[ß]{}ner, Rept. Prog. Phys.  [**56**]{}, 903 (1993). A. Manohar and H. Georgi, Nucl. Phys.  [**B234**]{}, 189 (1984). G. Müller and U.-G. Mei[ß]{}ner, Nucl. Phys.  [**B492**]{}, 379 (1997). M. Mojžiš, Eur. Phys. J. C [**2**]{}, 181 (1998). W. R. Molzon [*et al.*]{}, Phys. Rev. Lett.  [**41**]{}, 1213 (1978) \[Erratum, ibid.  [**41**]{}, 1523 (1978)\]. G. Morpurgo, Physics [**2**]{}, 95 (1965). U. Mosel, [*Fields, Symmetries, and Quarks*]{}. (McGraw-Hill, Hamburg, 1989). W. J. Marciano and H. Pagels, Phys. Rept.  [**36**]{}, 137 (1978). D. Morgan and M. R. Pennington, Phys. Lett. B [**272**]{}, 134 (1991). U.-G. Mei[ß]{}ner and I. Zahed, in [*Advances in Nuclear Physics, Vol. 17*]{}, edited by J. W. Negele and E. Vogt (Plenum, New York, 1986). M. M. Nagels [*et al.*]{}, Nucl. Phys.  [**B147**]{}, 189 (1979). Y. Nambu, Phys. Rev. Lett.  [**4**]{}, 380 (1960). Y. Ne’eman, Nucl. Phys.  [**26**]{}, 222 (1961). N. K. Nielsen, Am. J. Phys. [**49**]{}, 1171 (1981). Y. Nambu and G. Jona-Lasinio, Phys. Rev.  [**122**]{}, 345 (1961). Y. Nambu and G. Jona-Lasinio, Phys. Rev.  [**124**]{}, 246 (1961). H. Neufeld and H. Rupertsberger, Z. Phys. C [**71**]{}, 131 (1996). S. Okubo, Prog. Theor. Phys. [**12**]{}, 603 (1954). L. O’Raifeartaigh, [*Group Structure of Gauge Theories*]{} ( University Press, Cambridge, 1986). C. Ordonez, L. Ray, and U. van Kolck, Phys. Rev. C [**53**]{}, 2086 (1996). H. Pagels, Phys. Rept.  [**16**]{}, 219 (1975). A. Pich, Rept. Prog. Phys.  [**58**]{}, 563 (1995). A. Pich, in [*Probing the Standard Model of Particle Interactions*]{}, Proceedings of the Les Houches Summer School in Theoretical Physics, Session 68, Les Houches, France, 28 July - 5 September 1997, edited by R. Gupta, A. Morel, E. de Rafael, and F. David (Elsevier, Amsterdam, 1999). S. Pislak [*et al.*]{} \[BNL-E865 Collaboration\], Phys. Rev. Lett.  [**87**]{}, 221801 (2001). H. D. Politzer, Phys. Rev. Lett.  [**30**]{}, 1346 (1973). N. K. Pak and P. Rossi, Nucl. Phys. [**B250**]{}, 279 (1985). A. Pich and E. de Rafael, Nucl. Phys.  [**B367**]{}, 313 (1991). M. E. Peskin and D. V. Schroeder, [*An Introduction to Quantum Field Theory*]{} (Addison-Wesley, Reading, 1995). P. Post and K. Schilcher, Phys. Rev. Lett.  [**79**]{}, 4088 (1997). P. Post and K. Schilcher, Nucl. Phys. [**B599**]{}, 30 (2001). P. Post and K. Schilcher, Eur. Phys. J. C [**25**]{}, 427 (2002). P. Post and J. B. Tausk, Mod. Phys. Lett. A [**11**]{}, 2115 (1996). M. A. Preston, [*Physics of the Nucleus*]{} (Addison-Wesley, Reading, MA, 1962). E. de Rafael, in [*CP Violation and the Limits of the Standard Model*]{}, Proceedings of the 1994 Advanced Theoretical Study Institute in Elementary Particle Physics, Boulder, Colorado, 29 May - 24 June, 1994, edited by J. F. Donoghue (World Scientific, Singapore, 1995). R. J. Rivers, [*Path Integral Methods in Quantum Field Theory*]{} (Cambridge University Press, Cambridge, 1987). J. Roche [*et al.*]{} \[VCS Collaboration\], Phys. Rev. Lett.  [**85**]{}, 708 (2000). L. Rosselet [*et al.*]{}, Phys. Rev. D [**15**]{}, 574 (1977). S. M. Roy, Phys. Lett. B [**36**]{}, 353 (1971). T. E. Rudy, H. W. Fearing, and S. Scherer, Phys. Rev. C [**50**]{}, 447 (1994). L. H. Ryder, [*Quantum Field Theory*]{} (Cambridge University Press, Cambridge, 1985). SAID Program, R. A. Arndt, W. J. Briscoe, R. L. Workman, and I. I. Strakovsky, http://gwdac.phys.gwu.edu/. J. S. Schwinger, Phys. Rev. Lett.  [**3**]{}, 296 (1959). J. S. Schwinger, Phys. Lett. B [**24**]{}, 473 (1967). F. Scheck, [*Electroweak and Strong Interactions: An Introduction to Theoretical Particle Physics*]{} (Springer, Berlin, 1996), Chap. 3.5.2. H. C. Schröder [*et al.*]{}, Eur. Phys. J. C [**21**]{}, 473 (2001). S. Scherer and H. W. Fearing, Phys. Rev. D [**52**]{}, 6445 (1995). J. Stern, [*Light Quark Masses and Condensates in QCD*]{}, in [@Bernstein:pm]. S. Steininger, U.-G. Mei[ß]{}ner, and N. Fettes, JHEP [**9809**]{}, 008 (1998). Y. Takahashi, Nuovo Cim.  [**6**]{}, 371 (1957). H. B. Tang, hep-ph/9607436. M. V. Terentev, Yad. Fiz.  [**16**]{}, 162 (1972) \[Sov. J. Nucl. Phys. [**16**]{}, 87 (1973)\]. A. W. Thomas, in [*Advances in Nuclear Physics, Vol. 13*]{}, edited by J. W. Negele and E. Vogt (Plenum, New York, 1984). Y. Tomozawa, Nuovo Cim. [**46 A**]{}, 707 (1966). S. Treiman, R. Jackiw, and D. J. Gross, [*Lectures on Current Algebra and Its Applications*]{} (Princeton University Press, Princeton, 1972). A. W. Thomas and W. Weise, [*The Structure of the Nucleon*]{} (Wiley-VCH, Berlin, 2001). C. Unkmeir, S. Scherer, A. I. L’vov, and D. Drechsel, Phys. Rev. D [**61**]{}, 034002 (2000). C. Unkmeir, A. Ocherashvili, T. Fuchs, M. A. Moinester, and S. Scherer, Phys. Rev. C [**65**]{}, 015206 (2002). R. Urech, Nucl. Phys.  [**B433**]{}, 234 (1995). M. J. Veltman, [*Diagrammatica. The Path to Feynman Rules*]{} (Cambridge University Press, Cambridge, 1994). J. Volmer [*et al.*]{} \[The Jefferson Lab F(pi) Collaboration\], Phys. Rev. Lett.  [**86**]{}, 1713 (2001). C. Vafa and E. Witten, Nucl. Phys. [**B234**]{}, 173 (1984). U. Vogl and W. Weise, Prog. Part. Nucl. Phys.  [**27**]{}, 195 (1991). J. C. Ward, Phys. Rev.  [**78**]{}, 182 (1950). K. M. Watson, Phys. Rev.  [**95**]{}, 228 (1954). S. Weinberg, Phys. Rev.  [**112**]{}, 1375 (1958). S. Weinberg, Phys. Rev. Lett.  [**17**]{}, 616 (1966). S. Weinberg, Phys. Rev. Lett.  [**18**]{}, 188 (1967). S. Weinberg, Phys. Rev.  [**166**]{}, 1568 (1968). S. Weinberg, Phys. Rev. Lett.  [**31**]{}, 494 (1973). S. Weinberg, Physica A [**96**]{}, 327 (1979). S. Weinberg, Nucl. Phys.  [**B363**]{}, 3 (1991). S. Weinberg, [*The Quantum Theory Of Fields. Vol. 2: Modern Applications*]{} (Cambridge University Press, Cambridge, 1996). G. C. Wick, Phys. Rev.  [**80**]{}, 268 (1950). K. G. Wilson, Phys. Rev. D [**10**]{}, 2445 (1974). E. Witten, Nucl. Phys. [**B223**]{}, 422 (1983). J. Wess and B. Zumino, Phys. Lett. B [**37**]{}, 95 (1971). I. Zahed and G. E. Brown, Phys. Rept.  [**142**]{}, 1 (1986). A. Zee, Phys. Rev. D [**7**]{}, 3630 (1973). S. L. Zhu, S. Puglia, and M. J. Ramsey-Musolf, Phys. Rev. D [**63**]{}, 034002 (2001). J. Zinn-Justin, [*Quantum Field Theory And Critical Phenomena*]{} (Clarendon, Oxford, 1989). G. Zweig, CERN Report Nr. TH401, 4R12 (1964). [^1]: [email protected], http://www.kph.uni-mainz.de/T/ [^2]: For a prediction of hadron masses in the framework of lattice QCD see, e.g., Refs. [@Butler:1994em; @AliKhan:2001tx]. [^3]: They are not considered as “typical” hadrons due to their special role as the (approximate) Goldstone bosons of spontaneous chiral symmetry breaking. [^4]: The counting refers to ordinary chiral perturbation theory in the mesonic sector, where $D$ is an even number. [^5]: In this report we often adopt the convention that 1 stands for the unit matrix in $n$ dimensions. It should be clear from the respective context which dimensionality actually applies. [^6]: In our notation, the indices denoting group parameters and generators will appear as subscripts or superscripts depending on what is notationally convenient. We do not distinguish between upper and lower indices, i.e., we abandon the methods of tensor analysis. [^7]: We use the standard representation for the Dirac matrices (see, e.g., Ref. [@Bjorken_1964]). [^8]: We use natural units, i.e., $\hbar=c=1, e>0$, and $\alpha=e^2/4\pi \approx 1/137$. [^9]: Masses of gauge fields can be induced through a spontaneous breakdown of the gauge symmetry. [^10]: Historically, the color degree of freedom was introduced into the quark model to account for the Pauli principle in the description of baryons as three-quark states [@Greenberg:pe; @Han:pf]. [^11]: For the sake of clarity, the Gell-Mann matrices contain a superscript $C$, indicating the action in color space. [^12]: $$\epsilon_{\mu\nu\rho\sigma}=\left\{ \begin{array}{rl} +1& \mbox{if $\{\mu,\nu,\rho,\sigma\}$ is an even permutation of $\{0,1,2,3\}$} \\ -1& \mbox{if $\{\mu,\nu,\rho,\sigma\}$ is an odd permutation of $\{0,1,2,3\}$} \\ 0& \mbox{otherwise} \end{array} \right.$$ [^13]: Unless stated otherwise, we use the convention of Ref. [@Bjorken_1964]. [^14]: In case of fields, a transformation of the argument $\vec{x}\to -\vec{x}$ is implied. [^15]: Note that in the above sense, also $q$ is a chiral variable. However, the assignment of handedness does not have such an intuitive meaning as in the case of $q_L$ and $q_R$. [^16]: Here we adopt a covariant normalization of the spinors, $u^{(\alpha)\dagger}(\vec{p}\,)u^{(\beta)}(\vec{p}\,) = 2 E\delta_{\alpha\beta}$, etc. [^17]: By exponentiating elements of the Lie algebra u($N$) any element of U($N$) can be obtained. [^18]: Note that the transformation need not be realized linearly on the fields. [^19]: Normal ordering symbols are suppressed. [^20]: We have chosen to have the fields (field operators) rotate actively and thus must transform the states of Hilbert space in the opposite direction. [^21]: In the large $N_c$ (number of colors) limit of Ref. [@'tHooft:1973jz] the singlet axial-vector current is conserved, because the strong coupling constant behaves as $g^2\sim N_c^{-1}$. [^22]: Strictly speaking, we should also include the color indices. However, since we are only discussing color-neutral quadratic forms a summation over such indices is always implied, with the net effect that one can completely omit them from the discussion. [^23]: Here we assume that the dynamical system described by the Hamiltonian does not lead to a spontaneous symmetry breakdown. We will come back to this point later. [^24]: Later on, we will also refer to matrix elements of time-ordered products between states other than the vacuum as Green functions. [^25]: The singlet axial-vector current involves an anomaly such that the Green functions involving this current operator are related to Green functions containing the contraction of the gluon field-strength tensor with its dual. [^26]: The time ordering of $n$ points $x_1,\cdots,x_n$ gives rise to $n$! distinct orderings, each involving products of $n-1$ theta functions. [^27]: This matrix element will be dealt with in Sec. \[subsec\_gtravcme\]. [^28]: As in Refs.[@Gasser:1983yg; @Gasser:1984gg], we omit the coupling to the singlet axial-vector current which has an anomaly, but include a singlet vector current $v^\mu_{(s)}$ which is of some physical relevance in the two-flavor sector. [^29]: Many books on quantum field theory such as Refs.  [@Itzykson:rh; @Collins:xc; @Ryder:wq; @Rivers:hi] reserve the symbol $Z[v,a,s,p]$ for the generating functional of all Green functions as opposed to the argument of the exponential which denotes the generating functional of connected Green functions. [^30]: In order to obtain Green functions from the generating functional the simple rule $$\frac{\delta f(x)}{\delta f(y)}=\delta(x-y)$$ is extremely useful. Furthermore, the functional derivative satisfies properties similar to the ordinary differentiation, namely linearity, the product and chain rules. [^31]: In deriving these results we need to make use of $q_{\gamma,f}\bar{q}_{\delta,f'}=-\bar{q}_{\delta,f'}q_{\gamma,f}$, since the quark fields are anti-commuting field operators. [^32]: In Adler’s version, the right-hand side of Eq. (\[2:4:divasc\]) contains a renormalized field operator creating and destroying pions instead of $m_q P_i$. From a modern point of view, the combination $m_q P_i/(M_\pi^2 F_\pi)$ serves as an interpolating pion field (see Sec. \[subsec\_pps\]). Furthermore, the anomaly term is not yet present in Ref.[@Adler:1965]. [^33]: The case of a quantum field theory with an infinite volume $V$ has to be distinguished from, say, a nonrelativistic particle in a one-dimensional potential of a shape similar to the function of Fig. \[3:1:potng\]. For example, in the case of a symmetric double-well potential, the solutions with positive parity have always lower energy eigenvalues than those with negative parity (see, e.g., Ref. [@Greiner:mm]). [^34]: The field $\Phi'$ instead of $\Phi$ is assumed to vanish at infinity. [^35]: For continuous symmetry groups one may have a non-countably infinite number of ground states. [^36]: Of course, the Lagrangian is invariant under the full group O(3) which can be decomposed into its two components: the proper rotations connected to the identity, SO(3), and the rotation-reflections. For our purposes it is sufficient to discuss SO(3). [^37]: We say, somewhat loosely, that $T_1$ and $T_2$ do not annihilate the ground state or, equivalently, finite group elements generated by $T_1$ and $T_2$ do not leave the ground state invariant. This should become clearer later on. [^38]: The restriction to compact groups allows for a complete decomposition into finite-dimensional irreducible unitary representations. [^39]: The abbreviation $\sum_n\hspace{-1.4em}\int\hspace{0.5em} |n\rangle\langle n|$ includes an integral over the total momentum $\vec{p}$ as well as all other quantum numbers necessary to fully specify the states. [^40]: See Ref. [@Groom:in] for empirical limits on nucleon decay as well as baryon-number violating $Z$ and $\tau$ decays. [^41]: Recall that each quark is assigned a baryon number 1/3. [^42]: In this Section all physical quantities such as the ground state, the quark operators etc. are considered in the chiral limit. [^43]: The commutation relations also remain valid for [*equal*]{} times if the symmetry is explicitly broken. [^44]: Depending on the equations of motion, we will require more restrictive properties of the functions $\phi_i$. [^45]: An [invariant]{} subgroup has the additional property that the left and right cosets coincide for each $g$ which allows for a definition of the factor group $G/H$ in terms of the complex product. However, here we do not need this property. [^46]: Of course, the Goldstone boson fields are not constant vectors in $R^n$ but functions on Minkowski space \[see Eq. (\[4:2:m1\])\]. This is accomplished by allowing the cosets $gH$ to also depend on $x$. [^47]: Since the Goldstone bosons are pseudoscalars, a true parity transformation is given by $\phi_a(\vec{x},t)\mapsto -\phi_a(-\vec{x},t)$ or, equivalently, $U(\vec{x},t)\mapsto U^\dagger(-\vec{x},t)$. [^48]: At this stage, this is only a tree-level argument. We will see in Sec. \[subsec\_mgb\] that the Goldstone bosons remain massless in the chiral limit even when loop corrections have been included. [^49]: In the present case $\mbox{Tr}(\partial^\mu U U^\dagger)=0.$ [^50]: $\Phi$ and $\partial_\mu\Phi$ are matrices which, in general, do not commute. [^51]: Later on we will also see that the $\pi\pi$ scattering amplitude is effected by ${\cal L}_{\rm s.b.}$. [^52]: Note that the number of independent momenta is [*not*]{} the number of faces or closed circuits that may be drawn on the internal lines of a diagram. This may, for example, be seen using a diagram with the topology of a tetrahedron which has four faces but $N_L=6-(4-1)=3$ (see, e.g., Chap. 6-2 of Ref.  [@Itzykson:rh]). [^53]: Recall that the dimensions of a Lagrangian density and a field $\Phi$ are energy$^4$ and energy, respectively. [^54]: In principle, we could also “gauge” the U(1)$_V$ symmetry. However, this is primarily of relevance to the SU(2) sector in order to fully incorporate the coupling to the electromagnetic field \[see Eq. (\[2:4:rlasu2\])\]. Since in SU(3), the quark-charge matrix is traceless, this important case is included in our considerations. For further discussions, see Ref. [@Ebertshauser:2001nj]. [^55]: Under certain circumstances it is advantageous to introduce for each object with a well-defined transformation behavior a separate covariant derivative. One may then use a product rule similar to the one of ordinary differentiation \[see Eqs. (18) and (19) of Ref.[@Fearing:1994ga]\]. [^56]: There is a certain freedom in the choice of the elementary building blocks. For example, by a suitable multiplication with $U$ or $U^\dagger$ any building block can be made to transform as $V_R \cdots V_R^\dagger$ without changing its chiral order [@Fearing:1994ga]. The present approach most naturally leads to the Lagrangian of Gasser and Leutwyler [@Gasser:1984gg]. [^57]: At ${\cal O}(p^2)$ invariance under $C$ does not provide any additional constraints. [^58]: Some derivations in the literature neglect the second condition of Eq. (\[4:5:upcond\]) and thus obtain the wrong equations of motion. [^59]: Applying Eq. (\[4:5:invariants\]) one finds $\mbox{Tr}[D^2UU^\dagger-U(D^2 U)^\dagger]=0$. [^60]: Recall that the entries $V_{ud}$ and $V_{us}$ of the Cabibbo-Kobayashi-Maskawa matrix are real. [^61]: See Chap. 10.14 of Ref. [@Bjorken_1964] with the substitution $a/\sqrt{2}\to V_{ud} F_0$ in Eq. (10.140). [^62]: Of course, in the chiral limit, the pion is massless and, in such a world, the massive leptons would decay into Goldstone bosons, e.g., $e^-\to\pi^-\nu_e$. However, at ${\cal O}(p^2)$, the symmetry breaking term of Eq. (\[4:3:lqm\]) gives rise to Goldstone-boson masses, whereas the decay constant is not modified at ${\cal O}(p^2)$. [^63]: In the analysis of Ref. [@Groom:in] $f_\pi=\sqrt{2} F_\pi$ is used. [^64]: The first parameterization is popular, because the pion field appears only linearly in the term proportional to the Pauli matrices, leading to a substantial simplification when deriving Feynman rules. It is specific to SU(2) because, in contrast to the general case of SU($N$), in SU(2) the totally symmetric $d$ symbols vanish \[see Eq. (\[2:1:dabc\])\]. On the other hand, the exponential parameterization can be used for any $N$. [^65]: For a general proof of the equivalence of $S$-matrix elements evaluated at tree level (phenomenological approximation), see Sec. 2 of Ref. [@Coleman:sm]. [^66]: Since the whole procedure is rather technical, we will restrict ourselves, by means of the example to be discussed in Sec. \[subsec\_mgb\], to an explicit verification that the renormalization procedure indeed leads to finite predictions for physical observables. [^67]: The $\phi$ meson can decay into both $K^+K^-$ and $\pi^+\pi^-\pi^0$. [^68]: In order to conform with our previous convention of Eq. (\[4:2:utrafo\]), we need to substitute $U_{\rm W}\to U^\dagger$. Furthermore $F_\pi$ of Ref. [@Witten:tw] corresponds to $2F_0$. Finally, $$\partial^2 U U^\dagger- U\partial^2 U^\dagger=2\partial_\mu(\partial^\mu U U^\dagger).$$ [^69]: For pedagogical reasons, we make use of the physical fields. From a technical point of view, it is often advantageous to work with the Cartesian fields and, at the end of the calculation, express physical processes in terms of the Cartesian components. [^70]: Note that we work in SU(3) and thus with the exponential parameterization of $U$. [^71]: \[fnic\] The matrix element $\langle \pi^j(p')|\bar{q}\gamma^\mu q|\pi^i(p)\rangle$ must be of the form $\delta^{ij}(p'+p)^\mu f(q^2)$ which results in $(p+p')^\mu f(q^2)$ for the neutral pion ($i=j=3$). On the other hand, under charge conjugation $\bar{q}\gamma^\mu q\mapsto -\bar{q}\gamma^\mu q$ and $|\pi^0\rangle\mapsto |\pi^0\rangle$, and thus $f(q^2)=-f(q^2)=0$. [^72]: A second structure proportional to $q^\mu$ vanishes for on-mass-shell pions because of current conservation. [^73]: For example, in electron scattering reactions often the polarization vector $\epsilon_\mu=e\bar{u}(k_f)\gamma_\mu u(k_i)/q^2$ is used, with four-momentum transfer $q=k_i-k_f$. [^74]: The low-energy coupling constants of the SU(2)$\times$SU(2) Lagrangian are denoted by $l_i$ in distinction to the $L_i$ of the SU(3)$\times$SU(3) Lagrangian. [^75]: For neutral particles such as the neutron or the $K^0$ one has $e\int d^3 x \rho(r)=0$. [^76]: Breit-frame kinematics, i.e. $q^2=-\vec{q}\,^2$, comes closest to the nonrelativistic situation. [^77]: The numerical values in Ref. [@Gasser:1984ux] were obtained with $M_\pi= 135$ MeV, $M_K=495$ MeV, $F_0\approx F_\pi=93.3$ MeV, $\mu=M_\rho =770$ MeV, and $L^r_9(M_\rho)=(6.9\pm 0.7)\cdot 10^{-3}$. [^78]: A more recent value is given by $\langle r^2\rangle_\pi=(0.439\pm 0.008)\,\mbox{fm}^2$ [@Amendolia:1986wj]. Also (model-dependent) results have been obtained from pion-electroproduction experiments [@Liesenfeld:1999mv; @Volmer:2001ek]. [^79]: In Ref. [@Fearing:1994ga] the building blocks $[A]_\pm\equiv\frac{1}{2}(A U^\dagger\pm U A^\dagger)$ transforming as $V_R\cdots V_R^\dagger$ were used. The notation of Eq. (\[4:10:apm\]) has some advantages when implementing the total-derivative procedure to be discussed below. Moreover, it is more closely related to the conventions used in the baryonic sector (see Sec. \[sec\_tpf\]). [^80]: $K$ does not define an operation of SU(3)$\times$SU(3) on SU(3), because $K(1,1,U)=1\neq U\,\forall\, U$ (see Sec. \[subsec\_gc\]). [^81]: From an aesthetical point of view it would have been more satisfactory to introduce the covariant derivative as $\nabla_\mu(A)_\pm\equiv\partial_\mu(A)_\pm-i[\Gamma_\mu,(A)_\pm]$ to generate a closer formal correspondence to Eqs. (\[4:10:covder1\]). However, we follow the standard convention used in the literature. [^82]: In that sense, the final $\mbox{SU}(N_f)_L\times \mbox{SU}(N_f)_R$ set given in Ref. [@Bijnens:1999sh] for the even-intrinsic-parity sector is not minimal when setting all external fields to zero. In that case the structures $Y_1$ and $Y_4$ are not independent and can be eliminated. However, if one replaces these two terms by the terms (120) and (139) of Ref. [@Fearing:1994ga], then the entire set will remain independent whether or not there are external fields and both structures, (120) and (139), vanish explicitly when the fields are set to zero. [^83]: The definition differs by a factor of $(-M_\pi)$ [@Gasser:1983yg] from the conventional definition of scattering lengths in the effective range expansion (see, e.g., Ref. [@Preston:1962]). [^84]: Technically speaking the adjoint representation is faithful (one-to-one) modulo the center $Z$ of SU(3) which is defined as the set of all elements commuting with all elements of SU(3) and is given by $Z=\{1_{3\times 3}, \exp(2\pi i/3)1_{3\times 3}, \exp (4\pi i/3) 1_{3\times 3} \}$ [@O'Raifeartaigh:vq]. [^85]: The relation with the notation of Sec. \[subsec\_mclop6\] is given by $(D_\mu U)_-=-2iu_\mu$ [@Ebertshauser:2001nj]. [^86]: The power counting will be discussed below. [^87]: In fact, also the definition of the pion-nucleon form factor of Eq. (\[5:3:def\_gt\]) contains a sign opposite to the standard convention so that, in the end, the Goldberger-Treiman relation emerges with the conventional sign. [^88]: Using $m_N=938.3$ MeV, $g_A=1.267$, $F_\pi=92.4$ MeV, and $g_{\pi N}=13.21$ [@Schroder:rc], one obtains $\Delta_{\pi N}=2.6$ %. [^89]: The terminology “first and second classes” refers to the transformation property of strangeness-conserving semi-leptonic weak interactions under ${\cal G}$ conjugation [@Weinberg:1958ut] which is the product of charge symmetry and charge conjugation ${\cal G}={\cal C}\exp(i\pi I_2)$. A second-class contribution would show up in terms of a third form factor $G_T$ contributing as $$G_T(t) \bar{u}(p') i\frac{\sigma^{\mu\nu} q_\nu}{2 m_N} \gamma_5 \frac{\tau_i}{2} u(p).$$ Assuming a perfect ${\cal G}$-conjugation symmetry, the form factor $G_T$ vanishes. [^90]: One also finds the parameterization [@Becher:2001hv] $$T=\bar{u}(p')\left(D-\frac{1}{4m_N} [q\,'\hspace{-.9em}/\hspace{.2em},q\hspace{-.45em}/\hspace{.1em}]B\right)u(p)$$ with $D=A+\nu B$, where, for simplicity, we have omitted the isospin indices. [^91]: Recall that we use the normalization $\bar{u}u=2 m_N$. [^92]: The threshold parameters are defined in terms of a multipole expansion of the $\pi N$ scattering amplitude [@Chew:1957]. The sign convention for the $s$-wave scattering parameters $a_{0+}^{(\pm)}$ is opposite to the convention of the effective range expansion. [^93]: We do not expand the fraction $1/(1+\mu)$, because the $\mu$ dependence is not of dynamical origin. [^94]: The result, in principle, holds for a general target of isospin $T$ (except for the pion) after replacing 3/4 by $T(T+1)$ and $\mu$ by $M_\pi/M_T$. [^95]: For a recent discussion of the problem with $\gamma_5$ in dimensional regularization, see Ref. [@Jegerlehner:2000dz]. Since we are neither dealing with matrix elements containing anomalies nor considering closed fermion loops, we can safely make use of normal dimensional regularization [@Gasser:1987rb; @Jegerlehner:2000dz]. [^96]: It is straightforward to also determine the second-rank tensor integral of Eq. (\[5:4:integrals\]) using the methods described in Appendices \[app\_drb\] and \[app\_li\]. Regarding the analyticity properties we are interested in, one does not obtain any new information. [^97]: In order to facilitate the comparison with the Foldy-Wouthuysen result below, we make use of the “non-covariant” form of the Dirac equation. [^98]: The (second) quantization of the relevant fields will be discussed in Sec. \[subsec\_nfs\]. [^99]: The standard textbook treatment of the nonrelativistic reduction leading to the Pauli equation considers only terms of $1/m$ and thus does not yet generate non-Hermitian terms (see, e.g., Refs.  [@Bjorken_1964; @Itzykson:rh]). [^100]: In the framework of plane-wave solutions, Eq. (\[5:5:psilpsiltilde\]) already provides a hint that one may have to expect “unconventional normalization factors” when dealing with Feynman rules in the heavy-baryon approach. [^101]: It may be worthwhile to remember that $P_{v\pm}$ do not define orthogonal projectors in the mathematical sense, because they do not satisfy $P^\dagger_{v\pm}= P_{v\pm}$, with the exception of the special case $v^\mu=(1,0,0,0)$ used in Eq. (\[5:5:psiptwocomponent\]). [^102]: Of course, the decomposition of Eq. (\[5:5:pmvk\]) alone is not a sufficient criterion for $v\cdot k \ll m$. Taking, for example, $\vec{p}\perp \vec{v}$ one finds $v\cdot k=v\cdot p-m= Ev^0-m\gg m$ for large $v^0$. [^103]: For a derivation in the framework of the path-integral approach, see Ref. [@Mannel:1991mc] and Appendix A of Ref. [@Bernard:1992qa]. [^104]: In fact, setting all external fields to zero and dropping the interaction term proportional to $g_A$, it is easy to verify that the one-particle wave functions indeed satisfy the relation implied by Eq. (\[5:5:solutionhv\]). [^105]: Replacing ${\cal N}_v\to h^{(+)}$ and ${\cal H}_v\to h^{(-)}$, omitting all terms containing the chiral vielbein $u_\mu$, and interpreting the covariant derivative as that of QCD, the result of Eq. (\[5:5:lhneff\]) is identical with Eq. (7) of the discussion of heavy quark effective theory in Sec. 2 of Ref.  [@Balk:1993ev]. [^106]: In order to be able to invert the operator ${\cal C}$ of Ref. [@Bernard:1992qa], strictly speaking the projection operators $P_{v-}$ should not be included in the definition of ${\cal C}$. [^107]: For the classification of the irreducible representations of the Poincaré group, one makes use of the so-called Pauli-Lubanski vector $$W_\mu\equiv \frac{1}{2} \epsilon_{\mu\nu\rho\sigma}J^{\nu\rho} P^\sigma,$$ where $\epsilon_{\mu\nu\rho\sigma}$ is the completely antisymmetric tensor in four indices, $\epsilon_{0123}=1$, $J^{\mu\nu}$ denotes the generalized angular momentum operator, and $P^\mu$ is the four-momentum operator (see, e.g, Refs. [@Itzykson:rh; @Jones:ti]). Both $W^2$ and $P^2$ are Lorentz invariant and translationally invariant and are thus used as Casimir operators, where the eigenvalues are denoted by $m^2$ and $-m^2 s(s+1)$, $s=0,1/2,1,\cdots$. For the massive spin-1/2 case one obtains $$W_\mu=\frac{1}{4}\epsilon_{\mu\nu\rho\sigma}\sigma^{\nu\rho}P^\sigma.$$ Using (in four dimensions) $$\gamma_5\sigma_{\mu\nu}=-\frac{i}{2}\epsilon_{\mu\nu\rho\sigma}\sigma^{\rho \sigma},$$ together with the special choice $\tilde{v}^\mu=P^\mu/m$, one easily finds that the spin matrix is, for this special case, proportional to the Pauli-Lubanski vector, $W^\mu=m S^\mu_{\tilde{v}}$. [^108]: In evaluating Eq. (\[5:5:sss\]), we made use of $$\gamma_\mu a\hspace{-.5em}/ \gamma^\mu =(2-n)a\hspace{-.5em/},\quad \gamma_\mu a\hspace{-.5em}/\hspace{.2em} b\hspace{-.5em}/\hspace{.2em} \gamma^\mu=4 a\cdot b +(n-4)a\hspace{-.5em}/\hspace{.2em}b\hspace{-.5em}/\hspace{.2em},$$ in $n$ dimensions. [^109]: For the sake of simplicity, we consider one generic fermion field. [^110]: Strictly speaking we should also include the mesonic Lagrangian. [^111]: The overall factor $2\stackrel{\circ}{m}_N$ in Eq. (\[5:5:tnonrelred\]) is a result of our normalization of the spinors \[see Eq. (\[5:3:mthr\])\]. [^112]: In reality, the excitation energy of the $\Delta(1232)$ resonance very often provides the limit of convergence of the expansion. [^113]: The nomenclature of Refs.  [@Gasser:1987rb] and [@Bernard:1992qa] differs from the (more or less) standard convention of Eq. (\[5:5:l2pinhbb\]). The constants $c_i$ of Ecker and Mojžiš [@Ecker:1995rk] differ by a factor $1/\stackrel{\circ}{m}_N$ from those of Eq. (\[5:5:l2pinhbb\]). [^114]: Recall that ${\cal N}_{\tilde{v}}$ are two-component fields. [^115]: In the heavy-baryon formulation one has no closed fermion loops. In other words, in the single-nucleon sector exactly one fermion line runs through the diagram connecting the initial and final states. [^116]: In the remaining part of this section, we adopt the common practice of leaving out the projector $P_{v+}$ in the propagator and (possibly) in vertices with the understanding that all operators act only in the projected subspace. [^117]: Strictly speaking we should say $m_N-\stackrel{\circ}{m}_N={\cal O}[M^2/\stackrel{\circ}{m}_N, M^3/(4\pi F_0)^2]$, where the second result originates from the loop contribution. [^118]: In order to make it easier for the interested reader to follow Ref. [@Becher:1999he] we have used the notation there, omitting a factor $\mu^{n-4}$ and choosing the opposite overall sign in comparison with previous sections. [^119]: The transition to normal ordering involves an (infinite) constant which does not contribute to the commutator. [^120]: Up to an irrelevant constant the measure $[d\Phi_1][d\Phi_2]$ is equivalent to $[d\Phi][d\Phi^\ast]$, with $\Phi$ and $\Phi^\ast$ considered as independent variables of integration. [^121]: Recall that $\Gamma(z)$ is single valued and analytic over the entire complex plane, save for the points $z=-n$, $n=0,1,2,\cdots$, where it possesses simple poles with residue $(-1)^n/n!$ [@Abramowitz]. [^122]: If $m\geq 3$, the integral converges for $n=4$. [^123]: Our $J_{\pi N}(0;\omega)$ corresponds to $-\mu^{4-n}J_0(\omega)$ of Ref. [@Bernard:1995dp]. [^124]: Note that Gasser and Leutwyler called the ${\cal O}(p^2)$ and ${\cal O}(p^4)$ Lagrangians ${\cal L}_1$ and ${\cal L}_2$, respectively.
{ "pile_set_name": "ArXiv" }
--- abstract: | We investigate the representation of the so-called orthogonally $a$-Jensen mappings acting on $C^*$-modules. More precisely, let $\mathfrak{A}$ be a unital $C^*$-algebra with the unit $1$, let $a \in \mathfrak{A}$ be fixed such that $a, 1-a$ are invertible and let $\mathscr{E}, \mathscr{F}, \mathscr{G}$ be inner product $\mathfrak{A}$-modules. We prove that if there exist additive mappings $\varphi, \psi$ from $\mathscr{F}$ into $\mathscr{E}$ such that $\big\langle \varphi(y), \psi(z)\big\rangle=0$ and $a \big\langle \varphi(y), \varphi(z)\big\rangle a^\ast = (1 - a)\big\langle \psi(y), \psi(z)\big\rangle (1 - a)^\ast$ for all $y, z\in \mathscr{F}$, then a mapping $f: \mathscr{E} \to \mathscr{G}$ is orthogonally $a$-Jensen if and only if it is of the form $f(x) = A(x) + B(x, x) +f(0)$ for $x\in \mathscr{K} := \varphi(\mathscr{F})+\psi(\mathscr{F})$, where $A: \mathscr{E} \to \mathscr{G}$ is an $a$-additive mapping on $\mathscr{K}$ and $B$ is a symmetric $a$-biadditive orthogonality preserving mapping on $\mathscr{K}\times \mathscr{K}$. Some other related results are also presented. address: 'Department of Mathematics, Farhangian University, Tehran, Iran' author: - Ali Zamani title: 'Orthogonally $a$-Jensen mappings on $C^*$-modules' --- [**Introduction**]{} ==================== Orthogonally functionals on an inner product space when the orthogonality is the ordinary one have been considered by Pinsker [@pi]. Next Sundaresan [@Su] generalized the result of Pinsker to arbitrary Banach spaces equipped with the Birkhoff–James orthogonality. In the recent decades, mappings satisfying a functional equation under some orthogonality conditions have been investigated by several mathematicians, who have presented many interesting results and applications, see, e. g., [@C.L.Z; @C.L.W; @F.M.P; @I.T.Y; @M.Z; @P.P.V; @Ratz]. Jensen [@J] first studied the functions satisfying the condition $f(\frac{x + y}{2}) =\frac{f(x) + f(y)}{2}$. It is easy to see that every continuous Jensen function on $\mathbb{C}$ is affine in the sense that $f-f(0)$ is additive. The Jensen functional equation has been extensively studied from many aspects by many mathematicians, see, e.g., [@Ng; @S.K] and the references therein. Let us recall some definitions and introduce our notation. An inner product module over a $C^{*}$-algebra $\mathfrak{A}$ is a (right) $\mathfrak{A}$-module $\mathscr{E}$ equipped with an $\mathfrak{A}$-valued inner product $\langle\cdot,\cdot\rangle$, which is $\mathbb{C}$-linear and $\mathfrak{A}$-linear in the second variable and has the properties $\langle x, y\rangle^*=\langle y, x\rangle$ as well as $\langle x, x\rangle \geq 0$ with equality if and only if $x = 0$. An inner product $\mathfrak{A}$-module $\mathscr{E}$ is called a Hilbert $\mathfrak{A}$-module if it is complete with respect to the norm $\|x\|=\|\langle x, x\rangle\|^{\frac{1}{2}}$. Although inner product $C^{*}$-modules generalize inner product spaces by allowing inner products to take values in an arbitrary $C^{*}$-algebra instead of the $C^{*}$-algebra of complex numbers, but some fundamental properties of inner product spaces are no longer valid in inner product $C^{*}$-modules. For example, not each closed submodule of an inner product $C^{*}$-module is complemented. Therefore, when we are studying in inner product $C^{*}$-modules, it is always of some interest to find conditions to obtain the results analogous to those for inner product spaces. We refer the reader to [@Man] for more information on the theory $C^*$-algebras and the structure of Hilbert $C^{*}$-modules. Let $\mathscr{E}$ and $\mathscr{F}$ be two inner product $\mathfrak{A}$-modules. A morphism between inner product $\mathfrak{A}$-modules $\mathscr{E}$ and $\mathscr{F}$ is a mapping $\varphi:\mathscr{E}\longrightarrow \mathscr{F}$ satisfying $\langle \varphi(x), \varphi(y)\rangle=\langle x, y\rangle$ for all $x, y\in \mathscr{E}$. A mapping $t:\mathscr{E}\longrightarrow \mathscr{F}$ is called adjointable if there exists a mapping $s:\mathscr{F}\longrightarrow \mathscr{E}$ such that $\langle tx, y\rangle=\langle x, sy\rangle$ for all $x\in \mathscr{E}, y\in \mathscr{F}$. The unique mapping $s$ is denoted by $t^*$ and is called the adjoint of $t$. Furthermore, inner product $\mathfrak{A}$-modules $\mathscr{E}$ and $\mathscr{F}$ are unitarily equivalent ( and we write $\mathscr{E}\thicksim \mathscr{F}$) if there exists an adjointable mapping $u:\mathscr{E}\longrightarrow \mathscr{F}$ such that $u^* u=id_\mathscr{E}$ and $uu^*=id_\mathscr{F}$. A closed submodule $\mathscr{G}$ of an inner product $\mathfrak{A}$-module $\mathscr{E}$ is said to be orthogonally complemented if $\mathscr{G}\oplus \mathscr{G}^\perp=\mathscr{E}$, where $\mathscr{G}^\perp=\{x\in \mathscr{E}:\,\langle x, y\rangle=0 \,\,\mbox{for all}\,\,y\in \mathscr{G}\}$. A closed submodule $\mathscr{K}$ of an inner product $\mathfrak{A}$-module $\mathscr{E}$ is said to be fully complemented if $\mathscr{K}$ is orthogonally complemented and $\mathscr{K}^\perp\thicksim \mathscr{E}$. Note that the theory of inner product $C^{*}$-modules is quite different from that of inner product spaces. For example, not any closed submodule of an inner product $C^{*}$-module is complemented and there might exist bounded $\mathfrak{A}$-linear operators that are not adjointable. Throughout the paper let $\mathfrak{A}$ be a unital $C^*$-algebra with the unit $1$ and let $\mathscr{E}, \mathscr{F}, \mathscr{G}$ be inner product $\mathfrak{A}$-modules. We fix an element $a \in \mathfrak{A}$ such that $a, 1 - a$ are invertible. For instance, $a$ can be an element of $\mathfrak{A}$ satisfying $0<a<1$, where the order $c<d$ in $\mathfrak{A}$ means that $c, d$ are self-adjoint and the spectrum of $d-c$ is contained in $(m, \infty]$ for some positive number $m$. An additive mapping $A :\mathscr{E}\longrightarrow \mathscr{G}$ is called $a$-additive if $A(ax)=a A(x)$ for all $x\in \mathscr{E}$. A biadditive mapping $B: \mathscr{E} \times \mathscr{E}\to \mathscr{G}$ is called $a$-biadditive if $B(ax, ax) = a B(x, x)$ and $B\big((1-a)x, (1-a)x\big) = (1-a) B(x, x)$ for all $x\in \mathscr{E}$. It is symmetric if $B(x, y) = B(y, x)$ for all $x, y\in \mathscr{E}$. Furthermore, $B$ is said to be orthogonality preserving if for all $x, y\in \mathscr{E}$, $$\langle x, y\rangle=0\,\Longrightarrow\,B(x, y)=0.$$ A mapping $Q:\mathscr{E}\longrightarrow \mathscr{G}$ is said to be quadratic if it satisfies the so-called quadratic functional equation $$Q(x+y) + Q(x-y) = 2Q(x) + 2Q(y) \qquad (x, y\in \mathscr{E}).$$ Clearly any biadditive mapping is quadratic. A mapping $f: \mathscr{E} \to \mathscr{G}$ is called orthogonally $a$-Jensen if $$\begin{aligned} \label{id.200} \langle x, y\rangle=0\,\Longrightarrow\,f\big(ax + (1 - a)y\big) = a f(x) + (1 - a) f(y) \qquad (x, y\in \mathscr{E}).\end{aligned}$$ In particular if $p\in(0, 1)$, with $a = p1$ the mapping $f$ satisfying (\[id.200\]) is said to be orthogonally $p$-Jensen. Further if $p=\frac{1}{2}$ we say that $f$ is orthogonally Jensen. In this paper, we investigate the representation of the so-called orthogonally $a$-Jensen mappings acting on inner product $C^*$-modules. More precisely, we prove that if there exist additive mappings $\varphi, \psi$ from $\mathscr{F}$ into $\mathscr{E}$ such that $\big\langle \varphi(y), \psi(z)\big\rangle=0$ and $a \big\langle \varphi(y), \varphi(z)\big\rangle a^\ast = (1 - a)\big\langle \psi(y), \psi(z)\big\rangle (1 - a)^\ast$ for all $y, z\in \mathscr{F}$, then a mapping $f: \mathscr{E} \to \mathscr{G}$ is orthogonally $a$-Jensen if and only if it is of the form $f(x) = A(x) + B(x, x) +f(0)$ for $x\in \mathscr{K} := \varphi(\mathscr{F})+\psi(\mathscr{F})$, where $A: \mathscr{E} \to \mathscr{G}$ is an $a$-additive mapping on $\mathscr{K}$ and $B$ is a symmetric $a$-biadditive orthogonality preserving mapping on $\mathscr{K}\times \mathscr{K}$. In addition, we show that if $\mathscr{F}$ is a fully complemented submodule of $\mathscr{E}$ and $f$ is orthogonally Jensen, then $f$ is of the form $f(x) = A(x)+f(0)$ for $x\in \mathscr{F}$. [**Main results**]{} ==================== We start our work with the following lemmas. The first lemma follows immediately from (\[id.200\]). \[lm.201\] If $f:\mathscr{E}\longrightarrow \mathscr{F}$ is orthogonally $a$-Jensen, then - $a f( a^{-1}x) + (1 - a) f(0) = f(x)$ - $a f(0) + (1 - a) f\big( (1 - a)^{-1}x\big) = f(x)$ - $f( a^{-1}x) + a^{-1}(1 - a) f(0) = a^{-1} f(x)$ - $(1 - a)^{-1}a f(0) + f\big( (1 - a)^{-1}x\big) = (1 - a)^{-1} f(x)$ - $(1 - a)^{-1}a f(x) + f(0) = (1 - a)^{-1} f(ax)$ - $f(0) + a^{-1}(1 - a) f(x) = a^{-1} f\big((1 - a)x\big)$ for every $x \in \mathscr{E}$. \[lm.202\] Suppose that there exist additive mappings $\varphi, \psi:\mathscr{F}\longrightarrow \mathscr{E}$ such that $\big\langle \varphi(z), \psi(w)\big\rangle=0$ and $a \big\langle \varphi(z), \varphi(w)\big\rangle a^\ast = (1 - a) \big\langle \psi(z), \psi(w)\big\rangle (1 - a)^\ast$ for all $z, w\in \mathscr{F}$. If $f:\mathscr{E}\longrightarrow \mathscr{G}$ is orthogonally $a$-Jensen, then $$\begin{aligned} a f\big(\varphi(x) + \varphi(y)\big) &+ (1 - a) f\big(\psi(x) -\psi(y)\big) \\&= a \Big[f\big(\varphi(x)\big) + a^{-1}(1 - a) f\big(\psi(x)\big) - (1 - a)a^{-1} f(0)\Big] \\&\hspace{2cm} + (1 - a) \Big[(1 - a)^{-1}a f\big(\varphi(y)\big) - (1 - a)^{-1}a f(0) + f\big(\psi(-y)\big)\Big].\end{aligned}$$ for every $x, y\in \mathscr{F}$. We have $$\begin{aligned} \label{id.202.1} \Big\langle \varphi(x) &+ a^{-1}(1 - a)\psi(x), (1 - a)^{-1}a \varphi(y) - \psi(y)\Big\rangle\nonumber \\&= \big\langle \varphi(x), \varphi(y)\big\rangle \big((1 - a)^{-1}a\big)^\ast - \big\langle \varphi(x), \psi(y)\big\rangle \nonumber \\&\hspace{2.5cm} + a^{-1}(1 - a) \big\langle \psi(x), \varphi(y)\big\rangle \big((1 - a)^{-1}a\big)^\ast - a^{-1}(1 - a) \big\langle \psi(x) , \psi(y) \big\rangle \nonumber \\&= \big\langle \varphi(x), \varphi(y)\big\rangle \big((1 - a)^{-1}a\big)^\ast - a^{-1}(1 - a) \big\langle \psi(x) , \psi(y) \big\rangle \nonumber \\&\hspace{3cm}\Big(\mbox{since $\big\langle \varphi(x), \psi(y)\big\rangle=0$ and $\big\langle \psi(x), \varphi(y)\big\rangle = \big\langle \varphi(y), \psi(x)\big\rangle^* = 0$}\Big) \nonumber \\&= \big\langle \varphi(x), \varphi(y)\big\rangle \big((1 - a)^{-1}a\big)^\ast - a^{-1}(1 - a) \Big[(1 - a)^{-1}a \big\langle \varphi(x), \varphi(y)\big\rangle \big((1 - a)^{-1}a\big)^\ast\Big] \nonumber \\&\hspace{4.5cm}\Big(\mbox{since $(1 - a) \big\langle \psi(x), \psi(y)\big\rangle (1 - a)^\ast = a \big\langle \varphi(x), \varphi(y)\big\rangle a^\ast$}\Big) \nonumber \\&=0\end{aligned}$$ for every $x, y\in \mathscr{F}$. Therefore, we arrive at $$\begin{aligned} a f\big(\varphi(x) &+ \varphi(y)\big) + (1 - a) f\big(\psi(x)-\psi(y)\big) \\&= a f\big(\varphi(x+y)\big) + (1 - a) f\big(\psi(x-y)\big) \\&= f\Big(a \varphi(x+y) + (1 - a) \psi(x-y)\Big) \\&\hspace{2.7cm}\Big(\mbox{since $\big\langle \varphi(x+y), \psi(x-y)\big\rangle=0$ and $f$ is orthogonally $a$-Jensen}\Big) \\&= f\Big(a\big[\varphi(x) + a^{-1}(1 - a)\psi(x)\big] + (1 - a)\big[(1 - a)^{-1}a \varphi(y) - \psi(y)\big]\Big) \\&= a f\Big(\varphi(x) + a^{-1}(1 - a)\psi(x)\Big) + (1 - a) f\Big((1 - a)^{-1}a \varphi(y) - \psi(y)\Big) \\&\hspace{5cm}\Big(\mbox{since $f$ is orthogonally $a$-Jensen and (\ref{id.202.1}) holds}\Big) \\&= a f\Big(a\big[a^{-1}\varphi(x)\big] + (1 - a)\big[(1 - a)^{-1}a^{-1}(1 - a)\psi(x)\big]\Big) \\&\hspace{2cm}+ (1 - a) f\Big(a \big[a^{-1}(1 - a)^{-1}a\varphi(y)\big] + (1 - a)\big[-(1 - a)^{-1} \psi(y)\big]\Big) \\&= a\Big[a f\Big(a^{-1}\varphi(x)\Big) + (1 - a) f\Big((1 - a)^{-1}a^{-1}(1 - a)\psi(x)\Big)\Big] \\&\hspace{2cm}+ (1 - a)\Big[a f\Big(a^{-1}(1 - a)^{-1}a\varphi(y)\Big) + (1 - a) f\Big(-(1 - a)^{-1} \psi(y)\Big)\Big] \\&\hspace{1cm}\Big(\mbox{since $\big\langle a^{-1}\varphi(x), (1 - a)^{-1}a^{-1}(1 - a)\psi(x)\big\rangle = 0$}, \\&\hspace{5cm} \big\langle a^{-1}(1 - a)^{-1}a\varphi(y), -(1 - a)^{-1} \psi(y)\big\rangle = 0 \\&\hspace{8cm}\mbox{and $f$ is orthogonally $a$-Jensen}\Big) \\&= a \Big[f\big(\varphi(x)\big) - (1 - a) f(0) + f\big(a^{-1}(1 - a) \psi(x)\big) - a f(0)\Big] \\&\hspace{1cm}+ (1 - a) \Big[f\big((1 - a)^{-1}a \varphi(y)\big) - (1 - a) f(0) + f\big(-\psi(y)\big) - a f(0)\Big] \\&\hspace{8cm}\Big(\mbox{by Lemma \ref{lm.201}\,(i) and (ii)}\Big) \\&= a\Big[f\big(\varphi(x)\big) - (1 - a) f(0) + a^{-1} f\big((1 - a) \psi(x)\big) \\&\hspace{5cm}- a^{-1}(1 - a) f(0) - a f(0)\Big] \\&\hspace{3cm}+ (1 - a) \Big[(1 - a)^{-1} f\big(a \varphi(y)\big) - (1 - a)^{-1}a f(0) \\&\hspace{6cm} - (1 - a) f(0) + f\big(-\psi(y)\big) - a f(0)\Big] \\&\hspace{8cm}\Big(\mbox{by Lemma \ref{lm.201}\,(iii) and (iv)}\Big) \\&= a \Big[f\big(\varphi(x)\big) - (1 - a) f(0) + f(0) + a^{-1}(1 - a) f\big(\psi(x)\big) \\&\hspace{7cm} - (1 - a)a^{-1} f(0) - a f(0)\Big] \\&\hspace{2cm} + (1 - a) \Big[(1 - a)^{-1}a f\big(\varphi(y)\big) + f(0) \\&\hspace{4.5cm} - (1 - a)^{-1}a f(0) - (1 - a) f(0) + f\big(\psi(-y)\big) - a f(0)\Big] \\&\hspace{8cm}\Big(\mbox{by Lemma \ref{lm.201}\,(v) and (vi)}\Big).\end{aligned}$$ From this it follows that $$\begin{aligned} a f\big(\varphi(x) + \varphi(y)\big) &+ (1 - a) f\big(\psi(x) -\psi(y)\big) \\&= a \Big[f\big(\varphi(x)\big) + a^{-1}(1 - a) f\big(\psi(x)\big) - (1 - a)a^{-1} f(0)\Big] \\&\hspace{2cm} + (1 - a) \Big[(1 - a)^{-1}a f\big(\varphi(y)\big) - (1 - a)^{-1}a f(0) + f\big(\psi(-y)\big)\Big]\end{aligned}$$ and the lemma is proved. The condition that additive mappings $\varphi, \psi$ satisfying $\big\langle \varphi(x), \psi(y)\big\rangle=0$ and $a \big\langle \varphi(x), \varphi(y)\big\rangle a^\ast = (1 - a) \big\langle \psi(x), \psi(y)\big\rangle (1 - a)^\ast$ is not restrictive. In fact, there are non-trivial concrete examples of additive mappings satisfying this condition. A non-trivial example can be given in $\ell^2$ by $a = 1-p$ with $p\in(0, 1)$ and $$\begin{cases} \varphi, \psi:\ell^2\longrightarrow \ell^2 &\\ \varphi(\{a_n\}) = \big(\frac{1}{1-p}a_1, 0, \frac{1}{1-p}a_2, 0, \frac{1}{1-p}a_3, 0,\cdots\big) &\\ \psi(\{a_n\}) = \big(0, \frac{1}{p}a_1, 0, \frac{1}{p}a_2, 0, \frac{1}{p}a_3, 0,\cdots\big). \end{cases}$$ One can easily observe that $\big\langle \varphi(\{a_n\}), \psi(\{b_n\})\big\rangle=0$ and $$a \big\langle \varphi(\{a_n\}), \varphi(\{b_n\})\big\rangle a^\ast = (1 - a) \big\langle \psi(\{a_n\}), \psi(\{b_n\})\big\rangle (1 - a)^\ast = \sum_{n = 1}^\infty a_n b_n.$$ The following auxiliary results are needed in our investigation. \[pr.203\] Suppose that there exist additive mappings $\varphi, \psi:\mathscr{F}\longrightarrow \mathscr{E}$ such that $\big\langle \varphi(z), \psi(w)\big\rangle=0$ and $a \big\langle \varphi(z), \varphi(w)\big\rangle a^\ast = (1 - a) \big\langle \psi(z), \psi(w)\big\rangle (1 - a)^\ast$ for all $z, w\in \mathscr{F}$. If $f:\mathscr{E}\longrightarrow \mathscr{G}$ is an odd orthogonally $a$-Jensen mapping, then $f$ is additive on $\mathscr{K} := \varphi(\mathscr{F})+\psi(\mathscr{F})$. Since $f$ is odd $f(0)=0$. Thus for every $x, y\in \mathscr{F}$, by Lemma \[lm.202\] we conclude that $$\begin{gathered} \label{id.204} a f\big(\varphi(x) + \varphi(y)\big) + (1 - a) f\big(\psi(x)-\psi(y)\big) = \\a f\big(\varphi(x)\big) + (1 - a) f\big(\psi(x)\big) + a f\big(\varphi(y)\big) + (1 - a) f\big(-\psi(y)\big).\end{gathered}$$ Switching $x$ and $y$ in (\[id.204\]) we obtain $$\begin{gathered} \label{id.205} a f\big(\varphi(y) + \varphi(x)\big) + (1 - a) f\big(\psi(y)-\psi(x)\big) = \\a f\big(\varphi(y)\big) + (1 - a) f\big(\psi(y)\big) + a f\big(\varphi(x)\big) + (1 - a) f\big(-\psi(x)\big).\end{gathered}$$ Add (\[id.204\]) and (\[id.205\]) and use the fact that $f$ is odd to get $$\begin{aligned} 2a f\big(\varphi(x) + \varphi(y)\big) = 2a f\big(\varphi(x)\big) + 2a f\big(\varphi(y)\big),\end{aligned}$$ or equivalently, $$\begin{aligned} f\big(\varphi(x)+\varphi(y)\big)=f\big(\varphi(x)\big)+f\big(\varphi(y)\big).\end{aligned}$$ Hence $f$ is additive on $\varphi(\mathscr{F})$. Similarly $f$ is additive on $\psi(\mathscr{F})$. Now for every $z_1, z_2\in \mathscr{K}$ there exist $x_1, x_2, y_1, y_2\in \mathscr{F}$ such that $$z_1=\varphi(x_1)+\psi(y_1) \quad \mbox{and} \quad z_2=\varphi(x_2)+\psi(y_2).$$ We have $$\begin{aligned} f(z_1+z_2)&= f\big(\varphi(x_1+x_2)+\psi(y_1+y_2)\big) \\&=f\Big(aa^{-1} \varphi(x_1+x_2) + (1 - a)(1 - a)^{-1}\psi(y_1+y_2)\Big) \\&= a f\big(a^{-1} \varphi(x_1+x_2)\big) + (1 - a) f\big((1 - a)^{-1} \psi(y_1+y_2) \big) \\&\hspace{2cm}\Big(\mbox{since}\,\,\big\langle a^{-1} \varphi(x_1+x_2), (1 - a)^{-1} \psi(y_1+y_2)\big\rangle=0\,\, \\&\hspace{7cm}\mbox{and $f$ is orthogonally $a$-Jensen}\Big) \\&=f\big(\varphi(x_1+x_2)\big)+f\big(\psi(y_1+y_2)\big)\hspace{2cm}\Big(\mbox{by Lemma \ref{lm.201}\,(i) and (ii)}\Big) \\&= f\big(\varphi(x_1)\big)+f\big(\varphi(x_2)\big)+f\big(\psi(y_1)\big)+f\big(\psi(y_2)\big) \\&\hspace{5cm}\Big(\mbox{by the additivity of $f$ on $\varphi(\mathscr{F})$ and $\psi(\mathscr{F})$}\Big) \\&= a f\big(a^{-1} \varphi(x_1)\big)+ (1 - a) f\big((1 - a)^{-1} \psi(y_1)\big) \\&\hspace{3cm}+ a f\big(a^{-1} \varphi(x_2)\big) + (1 - a) f\big((1 - a)^{-1}\psi(y_2)\big) \\&\hspace{8cm}\Big(\mbox{by Lemma \ref{lm.201}\,(i) and (ii)}\Big) \\&= f\big(\varphi(x_1)+\psi(y_1)\big)+f\big(\varphi(x_2)+\psi(y_2)\big) \\&\hspace{3cm}\Big(\mbox{since}\,\,\big\langle a^{-1} \varphi(x_1), (1 - a)^{-1}\psi(y_1)\big\rangle = 0, \\&\hspace{5cm}\big\langle a^{-1}\varphi(x_2), (1 - a)^{-1}\psi(y_2)\big\rangle = 0 \\&\hspace{7cm}\mbox{and $f$ is orthogonally $a$-Jensen}\Big) \\&= f(z_1)+f(z_2).\end{aligned}$$ Thus $f$ is additive on $\mathscr{K}$. \[pr.210\] Suppose that there exist additive mappings $\varphi, \psi:\mathscr{F}\longrightarrow \mathscr{E}$ such that $\big\langle \varphi(z), \psi(w)\big\rangle=0$ and $a \big\langle \varphi(z), \varphi(w)\big\rangle a^\ast = (1 - a) \big\langle \psi(z), \psi(w)\big\rangle (1 - a)^\ast$ for all $z, w\in \mathscr{F}$. If $f:\mathscr{E}\longrightarrow \mathscr{G}$ is an even orthogonally $a$-Jensen mapping such that $f(0)=0$, then f is quadratic on $\mathscr{K} := \varphi(\mathscr{F})+\psi(\mathscr{F})$. Since $f$ is even and $f(0)=0$, putting $x=y$ in Lemma \[lm.202\] we infer that $$a f\big(2\varphi(x)\big) = 2a f\big(\varphi(x)\big) + 2(1 - a) f\big(\psi(x)\big) \,\qquad (x\in \mathscr{F}).$$ Similarly, we have $$(1 - a) f\big(2\psi(x)\big) = 2a f\big(\varphi(x)\big) + 2(1 - a) f\big(\psi(x)\big) \,\qquad (x\in \mathscr{F}).$$ Therefore, we conclude that $$\begin{aligned} \label{id.211} a f\big(2\varphi(x)\big) = (1 - a) f\big(2\psi(x)\big) \,\qquad (x\in \mathscr{F}).\end{aligned}$$ If we put $\frac{x}{2}$ instead of $x$ in (\[id.211\]) we get $$\begin{aligned} \label{id.212} a f\big(\varphi(x)\big) = (1 - a) f\big(\psi(x)\big) \,\qquad (x\in \mathscr{F}).\end{aligned}$$ Now for every $x, y\in \mathscr{F}$ we have $$\begin{aligned} f\big(\varphi(x)&+\varphi(y)\big) + f\big(\varphi(x)-\varphi(y)\big) \\&= f\big(\varphi(x)+\varphi(y)\big) + a^{-1}(1 - a) f\big(\psi(x)-\psi(y)\big) \hspace{4.2cm}\Big(\mbox{by}\,(\ref{id.212})\Big) \\&= f\big(\varphi(x)\big) + a^{-1}(1 - a) f\big(\psi(x)\big) + f\big(\varphi(y)\big) + a^{-1}(1 - a) f\big(\psi(y)\big) \\& \hspace{11cm}\Big(\mbox{by Lemma}\,\,\ref{lm.202}\Big) \\&= f\big(\varphi(x)\big) + f\big(\varphi(x)\big) + f\big(\varphi(y)\big) + f\big(\varphi(y)\big) \hspace{4.9cm}\Big(\mbox{by}\,\,(\ref{id.212})\Big) \\&= 2f\big(\varphi(x)\big) + 2f\big(\varphi(y)\big).\end{aligned}$$ Thus $$\begin{aligned} f\big(\varphi(x)+\varphi(y)\big)+f\big(\varphi(x)-\varphi(y)\big)=2f\big(\varphi(x)\big)+2f\big(\varphi(y)\big).\end{aligned}$$ So $f$ is quadratic on $\varphi(\mathscr{F})$. Similarly $f$ is quadratic on $\psi(\mathscr{F})$. By the same reasoning as in the last part of Proposition \[pr.203\] we conclude that $f$ is quadratic on $\mathscr{K}$. We are now in position to establish the main result. If $A:\mathscr{E}\longrightarrow \mathscr{G}$ is $a$-additive and $B:\mathscr{E}\times \mathscr{E}\longrightarrow \mathscr{G}$ is $a$-biadditive orthogonality preserving, then the mapping $f:\mathscr{E}\longrightarrow \mathscr{G}$ defined by $$f(x) = A(x) + B(x, x) + f(0) \qquad (x\in \mathscr{E})$$ is an orthogonally $a$-Jensen mapping. Namely, if $\langle x, y\rangle=0$ then $$\begin{aligned} f\big(ax + (1 - a)y\big)&= A\big(ax + (1 - a)y\big) + B\big(ax + (1 - a)y, ax + (1 - a)y\big) + f(0) \\&= a A(x ) + a B(x, x) + a f(0) \\&\hspace{4cm}+ (1 - a) A(y) + (1 - a)B(y, y) + (1 - a) f(0) \\&= a f(x) + (1 - a) f(x).\end{aligned}$$ The following theorem is a kind of converse of the previous discussion. \[th.217\] Let $a$ be an element of a unital $C^{*}$-algebra $\mathfrak{A}$ such that $a, 1-a$ are invertible and let $\mathscr{E}, \mathscr{F}, \mathscr{G}$ be inner product $\mathfrak{A}$-modules. Suppose that there exist additive mappings $\varphi, \psi:\mathscr{F}\longrightarrow \mathscr{E}$ such that $\big\langle \varphi(z), \psi(w)\big\rangle=0$ and $a \big\langle \varphi(z), \varphi(w)\big\rangle a^\ast = (1 - a) \big\langle \psi(z), \psi(w)\big\rangle (1 - a)^\ast$ for all $z, w\in \mathscr{F}$. Let $\mathscr{K} := \varphi(\mathscr{F})+\psi(\mathscr{F})$. If $f:\mathscr{E}\longrightarrow \mathscr{G}$ is an orthogonally $a$-Jensen mapping, then there exist unique mappings $A:\mathscr{E}\longrightarrow \mathscr{G}$ and $B:\mathscr{E}\times \mathscr{E}\longrightarrow \mathscr{G}$ such that $A$ is $a$-additive on $\mathscr{K}$, $B$ is symmetric $a$-biadditive orthogonality preserving on $\mathscr{K}\times \mathscr{K}$ and $$f(x) = A(x) + B(x, x) + f(0) \qquad (x\in \mathscr{K}).$$ Moreover, $B(x,y)=\frac{1}{8} \Big(f(x+y)+f(-x - y) - f(x-y) -f(-x + y)\Big)$ and $A(x)=\frac{1}{2} \Big(f(x)-f(-x)\Big)$ for every $x, y \in \mathscr{E}$. By passing to $f-f(0)$, if necessary, we may assume that $f(0)=0$. We decompose $f$ into its even and odd parts by $$f_o(x) =\frac{1}{2}\big(f(x)-f(-x)\big) \quad \mbox{and} \quad f_e(x) =\frac{1}{2}\big(f(x)+f(-x)\big)$$ for all $x\in \mathscr{E}$. Set $A(x) := f_o(x)$ and $B(x, y) :=\frac{1}{4}\big(f_e(x+y)-f_e(x-y)\big)$. It is easy to show that $f_o$ is odd orthogonally $a$-Jensen. So by Proposition \[pr.203\], $f_o$ is additive on $\mathscr{K}$. Also, by Lemma \[lm.201\] (v), we have $$A(ax) = f_o(ax) = a f_o(x) + b f_o(0) = a A(x) \qquad (x\in \mathscr{K}).$$ Hence $A$ is $a$-additive on $\mathscr{K}$. Furthermore, $f_e$ is even orthogonally $a$-Jensen. Since $f_e(0)= f(0)-f_o(0)=0$, $f_e$ is quadratic on $\mathscr{K}$ by Proposition \[pr.210\]. Thus $f_e(2x)=4f_e(x)$ and so, $$B(x, x)=\frac{1}{4}\big(f_e(2x)-f_e(0)\big)=f_e(x).$$ This implies $$f(x)=f_o(x)+f_e(x)=A(x)+B(x, x) \qquad (x\in \mathscr{K}).$$ For each $x, y, z\in \mathscr{E}$ we have $$\begin{aligned} \label{id.218} B(x+y, 2z)&= \frac{1}{4}\Big(f_e(x+y+2z)-f_e(x+y-2z)\Big)\nonumber \\&=\frac{1}{4}\Big(f_e\big((x+z)+(y+z)\big)+f_e\big((x+z)-(y+z)\big)\nonumber \\&\hspace{2cm}-f_e\big((x-z)+(y-z)\big)-f_e\big((x-z)-(y-z)\big)\Big)\nonumber \\&=\frac{1}{4}\Big(2f_e(x+z)+2f_e(y+z)-2f_e(x-z)-2f_e(y-z)\Big)\nonumber \\&=\frac{1}{2}\Big(f_e(x+z) - f_e(x-z) + f_e(y+z) - f_e(y-z)\Big)\nonumber \\&=2B(x, z)+2B(y, z).\end{aligned}$$ In particular, by choosing $y=0$, we get $$\begin{aligned} \label{id.219} B(x, 2z)&=2B(x, z)+2B(0, z)\nonumber \\&=2B(x, z)+\frac{1}{4}\Big(f(x)+f(-x)-f(x)-f(-x)\Big)\nonumber \\&=2B(x, z).\end{aligned}$$ If we replace $x$ by $x+y$ in (\[id.219\]), then by (\[id.218\]) we obtain $$B(x+y, z)=\frac{1}{2}B(x+y, 2z)=\frac{1}{2}[2B(x, z)+2B(y, z)]=B(x, z)+B(y, z),$$ and similarly $B(x, y+z)=B(x, y)+B(x, z)$. Therefore $B$ is biadditive on $\mathscr{K}\times \mathscr{K}$. Also, by Lemma \[lm.201\] (vi), we have $$B(ax, ax) = f_e(ax) = a f_e(x) + (1 - a) f_e(0) = a B(x, x) \qquad (x\in \mathscr{K})$$ and analogously $$B((1 - a)x, (1 - a)x) = (1 - a) B(x, x) \qquad (x\in \mathscr{K}).$$ Hence, $B$ is $a$-biadditive on $\mathscr{K}\times \mathscr{K}$. Further, for each $x, y\in \mathscr{K}$, it follows from $\langle x, y\rangle=0$ that $$\begin{aligned} B(x, y)&=\frac{1}{2}\big(B(x, y)+B(y, x)\big) \\&=\frac{1}{2}\big(B(x+y, x+y)-B(x, x)-B(y, y)\big) \\&=\frac{1}{2}\big(f(x+y)-A(x+y)\big)-\frac{1}{2}\big(f(x)-A(x)\big)-\frac{1}{2}\big(f(y)-A(y)\big) \\&=\frac{1}{2}\big(f(x+y)-f(x)-f(y)\big) \hspace{4.8cm}\Big(\mbox{$A$ is additive on $\mathscr{K}$}\Big) \\&=\frac{1}{2}\Big(f\big(aa^{-1} x + (1 - a)(1 - a)^{-1} y \big)-f(x)-f(y)\Big) \\&=\frac{1}{2}\Big(a f(a^{-1} x) + (1 - a) f((1 - a)^{-1}y)\Big) - \frac{1}{2}f(x)-\frac{1}{2}f(y) \\&\hspace{3.1cm}\Big(\mbox{since}\,\,\big\langle a^{-1} x, (1 - a)^{-1}y\big\rangle = 0 \,\,\mbox{and $f$ is orthogonally $a$-Jensen}\Big) \\&=\frac{1}{2}f(x)+\frac{1}{2}f(y)-\frac{1}{2}f(x)-\frac{1}{2}f(y)\hspace{3cm}\Big(\mbox{by Lemma \ref{lm.201}\,(i) and (ii)}\Big) \\&=0.\end{aligned}$$ Also, since $f_e$ is even, $B$ is symmetric. Thus $B$ is $a$-biadditive orthogonality preserving on $\mathscr{K}\times \mathscr{K}$. Finally suppose, $f(x) = A_1(x)+B_1(x,x)+f(0) = A_2(x)+B_2(x,x)+f(0)$ for any $x$ for the specified kind of mappings $A$ and $B$. Hence, $A_1(x)-A_2(x)=B_1(x,x)-B_2(x,x)$ for any $x$. However, the left part is an odd mapping, and the right part is an even mapping. So both these terms are equal to zero for any $x$. Thus we conclude that $A$ and $B$ are uniquely determined by $f$. The $a$-additive mappings $\varphi$, $\psi$ from $\mathscr {F}$ to $\mathscr {E}$ need not to be injective. Also, the linear span of their ranges need not coincide with $\mathscr {E}$. So, $a$ and $1-a$ might be assumed merely to admit generalized inverses inside the $C^*$-algebra $\mathfrak A$. By the requested equality $a \big\langle \varphi(z), \varphi(w) \big\rangle a^* = (1-a) \big\langle \psi(z), \psi(w) \big\rangle (1-a)^*$ for any $z, w \in \mathscr {F}$ the range projections of $a$ and of $1-a$ in the bidual von Neumann algebra ${\mathfrak A}^{**}$ of $\mathfrak A$ have to coincide. The domain projections of $a$ and of $1-a$ in the bidual von Neumann algebra ${\mathfrak A}^{**}$ of $\mathfrak A$ have to majorize the support projection of the subset $\big\langle \varphi(\mathscr {F}), \varphi(\mathscr {F})\big\rangle$ and of the subset $\big\langle \psi(\mathscr {F}), \psi(\mathscr {F})\big\rangle$ in ${\mathfrak A}^{**}$, respectively. For $a$ and $1-a$ admitting generalized inverses in $\mathfrak A$ the domain and the range projections belong to ${\mathfrak A} \subseteq {\mathfrak A}^{**}$. However, they might not belong to the center of ${\mathfrak A}$, so the Hilbert $C^*$-modules $\mathscr {F}$ and $\mathscr {E}$ cannot be reduced appropriately compatible with their module structure, in general. In the following result we obtain the representation of orthogonally $p$-Jensen mappings in inner product modules. \[cr.216\] Let $p\in(0, 1)$ be rational. Suppose that there exist additive mappings $\varphi, \psi:\mathscr{F}\longrightarrow \mathscr{E}$ such that $\big\langle \varphi(z), \psi(w)\big\rangle=0$ and $(1-p)^2\big\langle \varphi(z), \varphi(w)\rangle = p^2\big\langle \psi(z), \psi(w)\big\rangle$ for all $z, w\in \mathscr{F}$. If $f:\mathscr{E}\longrightarrow \mathscr{G}$ is orthogonally $p$-Jensen, then there exists a unique mapping $A:\mathscr{E}\longrightarrow \mathscr{G}$ such that $A$ is additive on $\mathscr{K} := \varphi(\mathscr{F})+\psi(\mathscr{F})$ and $$f(x) = A(x) + f(0) \qquad (x\in \mathscr{K}).$$ By Theorem \[th.217\], there exist unique mappings $A:\mathscr{E}\longrightarrow \mathscr{G}$ and $B:\mathscr{E}\times \mathscr{E}\longrightarrow \mathscr{G}$ such that $A$ is additive on $\mathscr{K}$, $B$ is symmetric biadditive orthogonality preserving on $\mathscr{K}\times \mathscr{K}$ and $$f(x) = A(x) + B(x, x) +f(0) \qquad (x\in \mathscr{K}).$$ We have $$\begin{aligned} A(x) + B(x, x) &+f(0) = f(x) \\&= (1-p) f(0) + p f(\frac{1}{p} x)\hspace{5cm}\Big(\mbox{by Lemma}\,\ref{lm.201}\Big) \\&= (1-p) f(0) + p \Big( A(\frac{1}{p} x) + B(\frac{1}{p} x,\frac{1}{p} x)+f(0)\Big) \hspace{1cm}\Big(\mbox{by Theorem}\,\ref{th.217}\Big) \\&=A(x)+\frac{1}{p}B(x,x)+f(0),\hspace{0.5cm}\Big(\mbox{by $\mathbb Q$-linearity A and $\mathbb{Q}$-bilinearity B}\Big)\end{aligned}$$ for any $x \in \mathscr{K}$. Consequently, $(1-\frac{1}{p}) B(x,x)=0$ and therefore, $B(x,x)=0$ for all $x \in \mathscr{K}$. Thus $B=0$ on $\mathscr{K}$ and $f(x) = A(x) + f(0)$ on $\mathscr{K}$. Note that if $A(x)=0$ for some $x \in \mathscr{K}$ then by $\mathbb{Q}$-linearity of $A$ we reach $A(qx) = 0$ for all rational numbers $q$. So, $f(qx) = A(qx) + f(0) = f(0)$ for all rational numbers $q$. \[cr.217\] Let $\mathscr{F}$ be a submodule of $\mathscr{E}$, and $\varphi:\mathscr{F}\longrightarrow \mathscr{E}$ be a morphism such that $\varphi(\mathscr{F} )\subseteq \mathscr{F}^\perp$. If $f:\mathscr{E}\longrightarrow \mathscr{G}$ is orthogonally Jensen, then there exists a unique mapping $A:\mathscr{E}\longrightarrow \mathscr{G}$ such that $A$ is additive on $\mathscr{K} := \mathscr{F}\oplus\varphi(\mathscr{F})$ and $$f(x) = A(x) + f(0) \qquad (x\in \mathscr{K}).$$ Let $id:\mathscr{F}\longrightarrow \mathscr{F}$ be the identity mapping. Since $\varphi$ is a morphism and $\varphi(\mathscr{F} )\subseteq \mathscr{F}^\perp$, for every $x, y\in \mathscr{F}$ we obtain $\big\langle \varphi(x), \varphi(y)\big\rangle = \big\langle id(x), id(y)\big\rangle$ and $\big\langle \varphi(x), id(y)\big\rangle=0$. It remains to apply Corollary \[cr.216\] for $p=\frac{1}{2}$. \[co.218\] Let $\mathscr{F}$ be a fully complemented submodule of $\mathscr{E}$. If $f:\mathscr{E}\longrightarrow \mathscr{G}$ be orthogonally Jensen, then there exists a unique mapping $A:\mathscr{E}\longrightarrow \mathscr{G}$ such that $A$ is additive on $\mathscr{F}$ and $f(x) = A(x) + f(0)$ for every $x\in \mathscr{F}$. Since $\mathscr{F}$ is a fully complemented submodule of $\mathscr{E}$, therefore $\mathscr{F}\oplus \mathscr{F}^\perp=\mathscr{E}$ and $\mathscr{F}^\perp\thicksim \mathscr{E}$. So, there exists an adjointable mapping $\phi:\mathscr{E}\longrightarrow \mathscr{F}^\perp$ such that $\phi^* \phi=id_\mathscr{E}$. We have $\big\langle \phi(x), \phi(y)\big\rangle = \big\langle \phi^* \phi(x), y\big\rangle = \big\langle x, y\big\rangle$ for all $x, y\in \mathscr{E}$. Let us define $\varphi := \phi|_\mathscr{F}$ . Then $\varphi(\mathscr{F} )\subseteq \mathscr{F}^\perp$ and $\big\langle \varphi(x), \varphi(y)\big\rangle = \big\langle x, y\big\rangle$ for all $x, y\in \mathscr{F}$. Thus $\varphi:\mathscr{F}\longrightarrow \mathscr{E}$ is a morphism such that $\varphi(\mathscr{F} )\subseteq \mathscr{F}^\perp$. Now the statement follows from Corollary \[cr.217\]. **Acknowledgement.** The author would like to thank the referee for her/his valuable suggestions and comments. He would also like to thank Professor M. S. Moslehian and Professor M. Frank for their invaluable suggestions while writing this paper. [99]{} D. Carando, S. Lassalle and S. I. Zalduendo, *Orthogonally additive holomorphic functions of bounded type over $C(K)$*, Proc. Edinb. Math. Soc. **53**(2)(2010), 609–618. J. Chmieliński, R. [Ł]{}ukasik and P. Wójcik, *On the stability of the orthogonality equation and the orthogonality-preserving property with two unknown functions*, Banach J. Math. Anal. **10** (2016), no. 4, 828–847. M. Frank, A. S. Mishchenko and A. A. Pavlov, *Orthogonality-preserving, $C^*$-conformal and conformal module mappings on Hilbert $C^*$-modules*, J. Funct. Anal. **260**(2011), 327–339. D. Ilišević, A. Turnšek and D. Yang, *Orthogonally additive mappings on Hilbert modules*, Studia Math. **221** (2014), no. 3, 209–229. J. L. W. V. Jensen, *Om konvekse Funktioner og Uligheder imellem Middelvaerdier*, Mat. Tidsskr. **B 16**(1905), 49–68. V. M. Manuilov and E. V. Troitsky, *Hilbert $C^*$-modules, Translations of Mathematical Monographs*, **226**, American Mathematical Society, Providence, RI, 2005. M.S. Moslehian and A. Zamani, *Mappings preserving approximate orthogonality in Hilbert $C*$-modules*, Math Scand. **122** (2018), 257–276. C. T. Ng, *Jensen’s functional equation on groups*, Aequationes Math. **39**(1990), 85–99. C. Palazuelos, A. M. Peralta and I. Villanueva, *Orthogonally additive polynomials on $C^*$-algebras*, Q. J. Math. **59**(2008), 363–374. A. G. Pinsker, *Sur une fonctionnelle dans l’espace de Hilbert*, C. R. (Dokl) Acad. Sci. URSS, N. Ser. **20**(1938), 411–414 (in French). J. Rätz, *Cauchy functional equation problems concerning orthogonality*, Aequationes Math. **62**(2001), 1–10. P. K. Sahoo and P. Kannappan, *Introduction to Functional Equations*, CRC Press, Boca Raton-London-New York 2011. K. Sundaresan, *Orthogonality and nonlinear functionals on Banach spaces*, Proc. Amer. Math. Soc. **34**(1972), 187–190.
{ "pile_set_name": "ArXiv" }
--- bibliography: - './references\_WhitePaper.bib' --- [**Detecting volcanically produced tori along orbits of exoplanets using UV spectroscopy**]{}\ [**Corresponding Author:**]{}\ Kristina G. Kislyakova – [email protected]\ Department of Astrophysics, University of Vienna\ Türkenschanzstrasse 17, 1180 Vienna\ Austria\ Space Research Institute, Austrian Academy of Sciences\ Schmiedlstrasse 6, 8042, Graz\ Austria\ [**Additional Authors:**]{} Luca Fossati (Space Research Institute, Austrian Academy of Sciences), Denis Shulyak (Max-Planck-Institute für Sonnensystemforschung), Eike Günther (Thüringer Landessternwarte Tautenburg, Tautenburg, Germany), Manuel Güdel (Department of Astrophysics, University of Vienna), Colin P. Johnstone (Department of Astrophysics, University of Vienna), Vladimir Airapetian (Goddard Space Flight Center, NASA, USA), Sudeshna Boro Saikia (Department of Astrophysics, University of Vienna), Allan Sacha Brun (Laboratoire AIM Paris-Saclay, France), Vera Dobos (Konkoly Observatory, Budapest, Hungary), Kevin France (Laboratory for Atmospheric and Space Physics, University of Colorado, Boulder, USA), Eric Gaidos (Department of Earth Sciences, University of Hawaii, Honolulu, USA), Maxim L. Khodachenko (Space Research Institute, Austrian Academy of Sciences), Antonino F. Lanza (Istituto Nazionale di AstroFisica, Catania, Italy), Helmut Lammer (Space Research Institute, Austrian Academy of Sciences), Lena Noack (Freie Universität Berlin, Germany), Rodrigo Luger (Center for Computational Astrophysics, Flatiron Institute, New York, USA), Antoine Strugarek (Laboratoire AIM Paris-Saclay, CEA/Irfu Universit$\acute{\rm e}$ Paris-Diderot CNRS/INSU), Aline Vidotto (School of Physics, Trinity College Dublin, Ireland), Allison Youngblood (Goddard Space Flight Center, NASA, USA) [**Detecting volcanically produced tori along orbits of exoplanets using UV spectroscopy**]{}\ 1.5 cm Introduction ============ Significant progress has been made since the discovery of the first exoplanet, 51 Peg b [@Mayor95], including collecting statistics on occurrence rates, masses, and radii of exoplanets (www.exoplanet.eu). Characterization is the next logical step. The Hubble Space Telescope ([*HST*]{}) is providing crucial observations for planet atmospheric characterization, particularly at ultraviolet (UV) wavelengths, which are not accessible from the ground. [*HST*]{} has been extremely effective in characterizing the upper atmospheres of giant exoplanets by observing escaping hydrogen (e.g., [@VM03; @Ehrenreich15; @Bourrier18]) and minor species present in the upper atmospheres of these planets [@Fossati10; @Haswell12; @VM13]. Characterizing the atmospheres of small rocky planets is more difficult due to their smaller sizes and pressure scale heights compared to those of giant planets. We suggest a new method for characterizing rocky exoplanets, namely, the detection of volcanically-produced tori lying along the planetary orbits. As we show below, [*HST*]{} is capable of detecting an Io-like plasma torus around a bright M dwarf in the solar neighborhood. Observations of this kind would be extremely timely as they would combine the unique observational capabilities of [*HST*]{}, particularly at far-UV (FUV) wavelengths, and new targets discovered by the [*TESS*]{} satellite. [*TESS*]{} is a whole-sky survey [@Ricker16] that is currently detecting a number of planets orbiting bright (V$<$10 mag) and close enough ($<$50pc) systems to enable characterization. Because of the large number of M dwarfs in the solar neighborhood and the red bandpass and observing windows of [*TESS*]{}, one can expect that [*TESS*]{} is likely to detect numerous close-in planets orbiting M dwarfs [@Barclay18]. Follow-up [*HST*]{} observations of these systems, particularly at FUV wavelengths, will give us important insights into atmospheres and even interior compositions of these planets. Status of research ================== An exoplanet requires internal heating sources to drive its volcanic activity. Several sources of internal energy are known for planets: radioactive decay, mantle differentiation, (inner) core formation, tidal heating, and induction heating. Due to the compactness of the planetary systems so far detected orbiting late M dwarfs and to the strong magnetic fields of the host stars, the latter two mechanisms, while insignificant for the Earth, can be very powerful for planets orbiting M dwarfs. Both of them are capable of generating enough heat to produce a global subsurface magma ocean [@K17; @Dobos19]. A sub-surface magma ocean can drive enormous volcanic activity, as is the case for the Jovian satellite Io due to the tidal interaction with other Galilean satellites [@Peale79]. Outgassed material is quickly lost to space, and forms a torus around Jupiter, along the moon’s orbit, populated mostly by oxygen and sulfur atoms and ions (Fig. 1; e.g., [@Murakami16]). Io’s torus has been observed both from space by the UVIS instrument on-board Cassini [@Steffl04] and by Voyager 1 [@Volwerk18], and from the ground (e.g., [@Thomas96]). Techniques for the removal of the geocoronal emission produced by neutral hydrogen and oxygen have been only recently developed [@Ben-Jaffel13], which is probably why Io’s torus has not yet been observed with [*HST*]{}. ![image](Murakami_et_al_SIII68nm_PlasmaTorus.jpg){width="0.8\columnwidth"} We suggest that similar tori can be produced by exoplanets in close orbit around late M dwarfs. A dense-enough torus can absorb the stellar light at the position of strong resonance lines of abundant elements, as in the case of the WASP-12 system [@Haswell12; @Fossati13]. @K18 studied the detectability of absorption signatures by O[i]{} superposed to the stellar FUV emission triplet at $\lambda\approx$1304Å taking the HST/STIS E140M observations of the active M dwarf AD Leo as a reference for the line strength and width. They have shown that at a spectral resolution of about 50000 (e.g., HST/STIS E140M grating) and a signal-to-noise ratio (S/N) of 10, which is currently reachable with [*HST*]{} for nearby M dwarfs, the torus would be detectable during one *HST* orbit if it had a column density larger than 10$^{13}$cm$^{-2}$ (Fig. 2). For comparison, the column density of Io’s torus is of the order of 10$^{13}$–10$^{14}$ cm$^{-2}$ [@Steffl04]. Planets orbiting M dwarfs are likely subject to efficient atmospheric escape (e.g., [@Airapetian17; @Garcia-Sage17]), which makes formation of an even denser plasma torus likely. Previously, @Demory16 have found indications of volcanism on 55Cnce by observing secondary eclipses and phase curves at mid-infrared wavelengths. [@Ridden-Harper16] have detected a tentative sodium signal for 55Cnce, which could potentially also have origin in volcanic activity. These studies suggest that volcanism can in principle be detectable on exoplanets. ![image](torus.png){width="0.6\columnwidth"} [*HST*]{} follow-up observations of planets orbiting M dwarfs can be used for detection of volcanic activity on these planets. Both tidal and induction heating are the most effective in planets orbiting late-type M dwarfs, due to compactness of these systems and strong stellar magnetic fields (e.g. [@Shulyak19; @Kochukov19]). However, these stars are also very dim, which requires very long integration times for observations. We suggest to focus on rocky planets orbiting M dwarfs of spectral types M4–M5, as they are still sufficiently bright, but also still likely to have planets with a sub-surface magma ocean due to tidal or induction heating. A possible target already available is the TRAPPIST-1 system. TTVs that are detected are indicative of tidal effects, including heating [@Luger17]. This system has been observed by the *HST* [@deWit18], however, one needs a brighter star to allow for the characterization of a torus. Therefore, a bigger and hotter star of a spectral class M4-M5 with a close-in rocky planet would be an ideal target. Observations of such tori have never been attempted before as only recently [*TESS*]{} has given us the possibility to find adequate targets for these novel observations. Science goals of a possible UV observational study ================================================== **Planet atmospheric composition and escape -** detecting a torus along the orbit will help us constraining the intensity of escape processes for planets orbiting M dwarfs. The torus will also enable us to identify the main constituents of the atmosphere. **Detection of exoplanet volcanic activity -** detecting a plasma torus would confirm the presence of active volcanoes on exoplanets. The material forming the torus should be constantly replenished by outgassing from the planetary interior, which in turn is supplied by the planet’s volcanic activity. **Characterization of planetary interiors -** [*HST*]{} observations of the torus will uniquely enable us to constrain the internal composition of exoplanets. This is because the composition of the outgassed material is directly connected to the composition and redox state of the planetary mantle [@Gaillard14]. Therefore, observations of this kind would give us the unique opportunity to reveal key properties of planetary interiors, which would be difficult or even impossible to obtain in any other way. Conclusions =========== The [*TESS*]{} mission is going to play a pivotal instrumental role in our future understanding of planets, particularly by providing numerous low-mass, transiting planets in close orbit to bright and nearby stars. Ultraviolet observations conducted with [*HST*]{} have successfully demonstrated their capability of detecting atomic and ionic species escaping from the atmospheres of giant planets. Furthermore, realistic predictions indicate that the same is going to be possible also for the tori of rocky planets in close orbit to nearby M dwarfs. [*HST*]{}’s remaining operational time is limited and it is of crucial importance to make best use of its capabilities, particularly at UV wavelengths, to gain as much observational material as possible to inform the science and guide the requirements of the next generation of space telescopes with UV capability.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We have made mapping observations of L1551 IRS 5, L1551NE, L723, and L43 and single-point observations of IRAS 16293-2422 in the submillimeter CS ($J$ = 7–6) and HCN ($J$ = 4–3) lines with ASTE. Including our previous ASTE observations of L483 and B335, we found a clear linear correlation between the source bolometric luminosities and the total integrated intensities of the submillimeter lines ($I_{CS}$ $\propto$ $L_{bol}^{0.92}$). The combined ASTE + SMA CS (7–6) image of L1551 IRS 5 exhibits an extended ($\sim$2000 AU) component tracing the associated reflection nebula at the west and southwest, as well as a compact ($\lesssim$500 AU) component centered on the protostellar position. The emission peaks of the CS and HCN emissions in L1551 NE are not located at the protostellar position but offset ($\sim$1400 AU) toward the associated reflection nebula at the west. With the statistical analyses, we confirmed the opposite velocity gradients of the CS (7–6) emission to those of the millimeter lines along the outflow direction, which we reported in our early paper. The magnitudes of the submillimeter velocity gradients are estimated to be (9.7$\pm$1.7) $\times$ 10$^{-3}$ km s$^{-1}$ arcsec$^{-1}$ in L1551 IRS 5 and (7.6$\pm$2.4) $\times$ 10$^{-3}$ km s$^{-1}$ arcsec$^{-1}$ in L483. We suggest that the “skewed” submillimeter molecular emissions toward the associated reflection nebulae at a few thousands AU scale trace the warm ($\gtrsim$40 K) walls of the envelope cavities, excavated by the associated outflows and irradiated by the central protostars directly. The opposite velocity gradients along the outflow direction likely reflect the dispersing gas motion at the wall of the cavity in the envelopes perpendicular to the outflow.' author: - 'Shigehisa <span style="font-variant:small-caps;">Takakuwa</span>' - 'Takeshi <span style="font-variant:small-caps;">Kamazaki</span>' title: | Skewed Distributions and Opposite Velocity Gradients of Submillimeter Molecular Lines\ in Low-Mass Protostellar Envelopes --- Introduction ============ Low-mass stars form in 3000 - 10000 AU scale condensations of molecular gas and dusts, so-called dense cores [@and00; @mye00]. Previous interferometric observations in millimeter molecular lines have revealed rotating and infalling gas motions in dense cores associated with protostars, “protostellar envelopes" [@oh96b; @oh97a; @oh97b; @mom98]. These millimeter observations have probed structures and kinematics of molecular gas with a temperature of $\sim$10 - 20 K and a density of 10$^{4-6}$ cm$^{-3}$ in the envelopes. On the other hand, submillimeter molecular lines can trace warmer ($\gtrsim$40 K) and denser ($\gtrsim$10$^{7}$ cm$^{-3}$) regions, and recent interferometric observations of protostellar envelopes in submillimeter molecular lines with the SMA have revealed compact ($\lesssim$500 AU) components associated with the central protostars, which often show rotational (possibly Keplerian) gas motion around the protostars [@tak04; @ta07b; @bri07; @lom08; @jor09; @yen11]. The SMA observations of protostellar envelopes, however, suffer from the effect of the missing flux. For example, the SMA observations of L1551 IRS 5 in the CS ($J$ = 7–6) line [@tak04] and those of IRAS 16293-2422 in the HCN ($J$ = 4–3) line [@ta07b] recovered only $\sim$11 $\%$ and $\sim$35 $\%$ of the total fluxes observed with CSO and JCMT, respectively, suggesting the presence of the extended ($\gtrsim$2000 AU) submillimeter components. In fact, the combined SMA + JCMT image of IRAS 16293-2422 in the HCN (4–3) line shows an extended ($\sim$3000 AU) envelope structure as well as a compact ($\sim$500 AU) disklike structure associated with the protostar. These results suggest that there is extended ($\gtrsim$2000 AU), warm ($\gtrsim$40K) and/or dense ($\gtrsim$10$^{7}$ cm$^{-3}$) gas in protostellar envelopes. In order to investigate the origin of the extended submillimeter molecular emissions in protostellar envelopes, we initiated mapping observations of protostellar envelopes in the CS (7–6) and HCN (4–3) lines with Atacama Submillimeter Telescope Experiment (ASTE)[^1], a 10-m submillimeter single-dish telescope at the Atacama Site in Chile ([@ta07a]; hereafter Paper I). In Paper I, we mapped the protostellar envelopes around L483 and B335. In L483, the HCN emission is slightly resolved and exhibits a western extension ($\sim$5000 AU) toward the direction of the associated reflection nebula and the blueshifted outflow. Furthermore, the position-velocity diagrams of the HCN and CS emissions along the axis of the associated molecular outflows in L483 and B335 exhibit possible velocity gradients opposite to those of the millimeter emissions or the associated molecular outflows. These results in Paper I are, however, still a little marginal and more robust confirmations of those results, including more observing samples and more sophisticated data analyses, are required. In the present paper, we have expanded our previous ASTE observations and have performed mapping observations of L1551 IRS 5, L1551 NE, L723, and L43. All of these and previous source samples are nearby ($D$ $\lesssim$300 pc) and representative ($L_{bol}$ $\gtrsim$3 $\LO$) low-mass protostellar objects suitable for the purpose of the present study. We have constructed data analysis tools by ourselves and have performed more objective, unambiguous data analyses of the previous and the present new data. From these observations and data analyses, we have revealed that the submillimeter molecular lines often show “skewed” emission distributions tracing the reflection nebula (L1551 IRS 5, L483, L1551 NE, and L723), and have found a linear correlation between the intensity of the submillimeter molecular lines and the protostellar luminosities. We have also verified the presence of the opposite velocity gradients of the submillimeter CS line to those of the millimeter lines and the associated outflows in L1551 IRS 5 and L483. With these results obtained from our ASTE observations and data analyses, we discuss the origin of the submillimeter molecular emissions in low-mass protostellar envelopes. The structure of the present paper is as follows. In $\S$2, we describe our new ASTE observations, and combination of the ASTE CS (7–6) data with the previously-published SMA CS (7–6) data in L1551 IRS 5 [@tak04]. In $\S$3.1., we discuss all the observed spectra toward the protostellar positions, while in $\S$3.2. we show the new results of the mapping observations. In $\S$4.1., we discuss the extents and distributions of the submillimeter molecular lines and their origin, including our simple radiative-transfer model. In $\S$4.2., we verify the presence of the opposite velocity gradients with our statistical analyses, and discuss a possible reason for the opposite submillimeter velocity gradients. ASTE Observations and Combining with the SMA Data ================================================= With the ASTE 10-m telescope we made mapping observations of L1551 IRS 5, L1551 NE, L723, and L43, in the CS ($J$ = 7–6; 342.882866 GHz) and HCN ($J$ = 4–3; 354.505480 GHz) lines on 2006 August 26, 28 – September 2. Remote observations were performed from an ASTE operation room of NAOJ at Mitaka, Japan, using the network observation system N-COSMOS3 developed by NAOJ ([@kam05]). Details of the ASTE telescope are presented by Ezawa et al. (2004). A cartridge-type 350 GHz receiver mounted on ASTE is a double sideband instrument with an IF frequency range of 5 to 7 GHz ([@koh05]), and the CS and HCN lines were observed simultaneously at different sidebands. The telescope beam size and the spectral resolution were $\sim$22$\arcsec$ and 125 kHz ($\sim$0.11 km s$^{-1}$), respectively. The typical DSB system noise temperature was $\sim$270 - 500 K. The telescope pointing was checked at the beginning and the middle of the observations by making continuum observations of Uranus or Jupiter, or five-point CO ($J$ = 3–2) observations of a late-type star, O-Cet, and was found to be better than 2$\arcsec$ during the whole observing period. As a standard source we also observed the central positions of IRAS 16293-2422 and L1551 IRS 5, and confirmed that the relative intensity was consistent within $\sim$30$\%$. The total on-source integration times for IRAS 16293-2422 and L1551 IRS 5 were 11.8 minutes and 23.8 minutes, respectively. We compared the observed CS and HCN spectra toward IRAS 16293-2422 to those obtained with the CSO telescope which has the same dish size as that of the ASTE telescope ([@bla94; @dis95]), and found that the main beam efficiency of ASTE was $\sim$0.6 both for the CS and HCN lines. No further correction for the sideband ratio was performed. Hereafter we show the observed line intensities in the unit of $T_{MB}$. In the mapping observations, typical rms noise levels per point were $\sim$0.16 K, $\sim$0.20 K, $\sim$0.20 K, and $\sim$0.27 K in L1551 IRS 5, L1551 NE, L723, and L43, respectively, with a typical on-source integration time of $\sim$4 - 12 minutes. The mapping regions cover most of the CS and HCN emission regions at a grid spacing of 10$\arcsec$, providing the Nyquist-sampled maps. Toward the center of L43 a better rms noise level ($\sim$0.18 K) was obtained with an on-source integration time of 13.3 minutes, to compare the central spectra to those toward other sources. In Paper I, we performed mapping observations of L483, B335, and single-point observations of L723 and IRAS 16293-2422, and in total we observed seven protostellar sources in the submillimeter CS and HCN lines with ASTE. The observed source properties and the ASTE observations from Paper I and the present paper (hereafter Paper II) are summarized in Table \[tab:source\]. Figure \[fig:I16spec\] shows the CS and HCN spectra toward IRAS 16293-2422 after combining both the Paper I and II data, and Figure \[fig:spall\] shows the observed CS ($J$ = 7–6) and HCN ($J$ = 4–3) spectra toward the central positions of all the other protostellar sources. We also combined the ASTE CS (7–6) image of L1551 IRS 5 with the SMA CS (7–6) image [@tak04], adopting the method described by @ta07b. Details of the SMA observations are described in @tak04. The conversion factor from $T_{MB}$ (K) to S (Jy beam$^{-1}$) was derived to be 46.6 as $$S = \frac{2k_{B}\Omega_{beam}}{\lambda^2}T_{MB},$$ where $k_{B}$ is the Boltzmann constant, $\lambda$ is the wavelength, and $\Omega_{beam}$ is the solid angle of the ASTE beam (= 22$\arcsec$). The SMA observations recovered $\sim$10$\%$ of the total CS (7–6) flux observed with ASTE. The resultant synthesized beam size and the rms noise level per channel in the combined image are 3$\farcs$6 $\times$ 2$\farcs$7 (P.A. = -65$^{\circ}$) and $\sim$1.1 Jy beam$^{-1}$, respectively, where the velocity resolution is the same as that of the SMA data ($\sim$0.179 km s$^{-1}$). [llccccccl]{} Source &IRAS &$L_{bol}$ &$T_{bol}$ &R. A. &Dec. &Distance &Ref. &ASTE Observation\ & &($\LO$) &(K) &(J2000.0) &(J2000.0) &(pc) & &\ L1551 IRS 5 &04287+1801 &22 &92 &04$^{h}$31$^{m}$34$^{s}.$14 &18$^{\circ}$08$\arcmin$05$\farcs$1 &140 &1,2,3,4 &Paper II (mapping)\ L1551 NE &04289+1802 &4.2 &91 &04 31 44.47 &18 08 32.2 &140 &1,2,4,5 &Paper II (mapping)\ IRAS 16293-2422 &16293-2422 &21 &42 &16 32 22.87 &-24 28 36.6 &160 &2,6,7 &Paper I & II (both one point)\ L43 &16316-1540 &2.7 &370 &16 34 29.30 &-15 47 01.7 &125 &8,9 &Paper II (mapping)\ L483 &18148-0440 &13 &52 &18 17 29.86 &-04 39 38.7 &200 &2,8,10 &Paper I (mapping)\ L723 &19156+1906 &3.3 &47 &19 17 53.62 &19 12 19.5 &300 &2,8,11 &Paper I (one point) & II (mapping)\ B335 &19345+0727 &3.1 &28 &19 37 00.94 &07 34 08.8 &150 &2,8,12 &Paper I (mapping)\ Results ======= Submillimeter CS and HCN Spectra -------------------------------- Figures \[fig:I16spec\] and \[fig:spall\] show the observed ASTE CS ($J$ = 7–6) and HCN ($J$ = 4–3) spectra toward the protostellar positions of all of our targets, and Table \[tab:spectra\] summarizes the observed line parameters derived from Gaussian fittings unless otherwise noted. Except for the CS and HCN lines toward L1551 NE and the CS line toward L43, we detected these submillimeter lines above 4$\sigma$ level. The CS line is stronger than the HCN line except for L43. The CS intensity ranges from $\lesssim$0.5 K (L1551 NE) to $\sim$9.1 K (IRAS 16293-2422) and the HCN intensity from $\sim$0.4 K (L723) to $\sim$5.8 K (IRAS 16293-2422). The line widths of these lines are typically $\sim$1 - 2 km s$^{-1}$. Toward L483 and L723 the HCN line widths are significantly wider than the CS line widths. The HCN ($J$ = 4–3) line consists of six hyperfine components, two of which ($F$ = 3–3 and 4–4) can be separated from the main line ($F$ = 4–3) by 1.977 MHz (-1.67 km s$^{-1}$) and -1.610 MHz (1.36 km s$^{-1}$), respectively ([@jew97]). The intensity ratio between the main and the other two hyperfine lines at the local thermal equilibrium condition is 0.0217 ([@jew97]), and these weaker hyperfine lines are unlikely to be detectable with the present observations. It is still possible, however, that the broader HCN line widths than the CS line widths found in L483 and L723 are due to the presence of the hyperfine lines, if there is any hyperfine anomaly. Toward L1551 IRS 5 ($L_{bol}$ = 22 $\LO$) and L483 ($L_{bol}$ = 13 $\LO$), the submillimeter intensities are $\sim$twice as high as those toward the other, less bright ($L_{bol}$ $\lesssim$4.2 $\LO$) sources, with an exception of the CS line toward B335 and the HCN line toward L43. Toward IRAS 16293-2422 ($L_{bol}$ = 21 $\LO$) the submillimeter CS and HCN intensities are much stronger than those toward the other sources (Fig. \[fig:I16spec\]). These results suggest that the intensities of the submillimeter molecular lines are related to the protostellar luminosities. We will discuss this point in $\S$4.1. (80mm,80mm)[fig1.eps]{} (160mm,160mm)[fig2.eps]{} [lccccccccccc]{} & & &\ Source &$\int T_{MB} dv$&$T_{MB}$&rms &$\Delta v$&$v_{\rm LSR}$& &$\int T_{MB} dv$&$T_{MB}$&rms &$\Delta v$&$v_{\rm LSR}$\ &(K km s$^{-1}$) &(K) &(K) &(km s$^{-1}$) &(km s$^{-1}$) & &(K km s$^{-1}$) &(K) &(K) &(km s$^{-1}$) &(km s$^{-1}$)\ L1551 IRS 5 &2.12&1.61&0.10 &0.92&6.45& &1.70&1.32&0.12 &0.88&6.29\ L1551 NE &1.26 &0.46 &0.19 &2.57 &7.18 & &- &- &0.20 &- &-\ IRAS 16293-2422 &31.05&9.12&0.12 &2.68&3.56& &19.41&5.79&0.08 &-&3.48\ L43 &0.68&0.52&0.18 &0.44&0.11& &0.85 &1.06 &0.19 &0.75 &0.41\ L483 &2.61 &1.80 &0.22 &1.36 &5.52 & &1.87 &0.84 &0.21 &2.09 &5.62\ L723 &1.21 &0.74 &0.12 &1.54 &11.24 & &1.18 &0.36 &0.09 &3.07 &11.09\ B335 &1.91 &1.20 &0.15 &1.49 &8.18 & &0.43 &0.58 &0.14 &0.69 &8.06\ Distribution of the Submillimeter CS and HCN Emission ----------------------------------------------------- ### L1551 IRS 5 L1551 IRS 5 is the brightest ($\sim$22 $\LO$) protostellar source in Taurus, and is classified as Class I [@eme84; @fro05]. L1551 IRS 5 was the first protostellar source where a bipolar molecular outflow was identified [@sne80]. Subsequent observations of L1551 IRS 5 in CO (1–0, 2–1, 3–2) lines have revealed structures and kinematics of the parsec-scale molecular outflow, with the blueshifted and redshifted lobes at the south-west and north-east of the protostar, respectively [@uch87; @sto06; @mor06]. High angular-resolution ($\sim$0$\farcs$04) 7-mm continuum observations of L1551 IRS 5 with VLA have found that L1551 IRS 5 is a triple protostellar system with a projected separation of $\sim$47 AU and $\sim$13 AU [@lim06]. SMA observations of L1551 IRS 5 in the CO (2–1) line within the $\sim$4000 AU region have found multiple molecular outflows, one of which shows precessing motion presumably due to the tidal interaction of the circumstellar disks around the multiple protostars [@wu09]. Interferometric observations of L1551 IRS 5 in millimeter molecular lines of CS (2–1), H$^{13}$CO$^{+}$ (1–0), and C$^{18}$O (1–0) have revealed a $\sim$2000 - 5000 AU-scale molecular envelope surrounding L1551 IRS 5, which is elongated perpendicularly to the outflow axis [@oh96b; @sai96; @mom98]. Along the outflow direction the south-western and north-eastern parts of the envelope are blueshifted and redshifted, whereas across the outflow direction the north-western and south-eastern parts are redshifted and blueshifted, respectively. The velocity gradients seen in the envelope along and across the outflow direction have been interpreted as an infalling and rotating gas motion in the envelope, respectively [@oh96b; @sai96; @mom98]. On the other hand, SMA observations of L1551 IRS 5 in the submillimeter CS (7–6) line have found a compact ($\sim$500 AU) molecular gas with mainly rotating motion around the protostar [@tak04]. With the present ASTE observations of L1551 IRS 5 we can investigate the extent and kinematics of the warm gas traced in the submillimeter molecular lines and the relation with the infalling envelope traced in the millimeter lines. Figure \[fig:irs5map\] shows total integrated intensity maps of the CS ($J$ = 7–6) ($left$ $panel$) and HCN ($J$ = 4–3) ($right$) lines in L1551 IRS 5. Figures \[fig:irs5cs\] and \[fig:irs5hcn\] show relevant line profile maps of the CS (7–6) and HCN (4–3) lines, respectively. There are intense blobs centered on the protostellar position both in the CS and HCN emissions. Both the CS and HCN emission distributions show a slight elongation toward the south-west of the protostar. The shapes of the emission distributions do not resemble the shape of the ASTE beam, suggesting that the submillimeter emissions are slightly resolved. Two-dimensional Gaussian fittings to the submillimeter emission distributions show that the deconvolved size of the HCN emission is $\sim$18$\arcsec$ $\times$ 15$\arcsec$ (= 2600 AU $\times$ 2100 AU; PA = 65$^{\circ}$ $\pm$ 12$^{\circ}$), and that in the CS emission only the major axis (PA = 70$^{\circ}$ $\pm$ 1$^{\circ}$) is resolved ($\sim$18$\arcsec$). The deconvolved sizes are smaller than the ASTE beam size and hence the estimated emission extents are not well-determined. In order to unambiguously measure the extent of the submillimeter molecular emission, we combined the present ASTE CS (7–6) data with our published SMA CS (7–6) data [@tak04]. The $middle$ and $right$ panels in Figure \[fig:l15comb\] show the combined ASTE + SMA total integrated intensity maps of the CS (7–6) emission in L1551 IRS 5 before and after the primary beam correction, respectively. The submillimeter CS emission in L1551 IRS 5 consists of two components; one is a compact ($\lesssim$500 AU) component centered on the protostellar positions, and the other a halo-like, extended ($\sim$2000 AU) component, delineated by brown dashed curves in Figure \[fig:l15comb\]. The central compact CS (7–6) component is also seen in the SMA-only image, which shows rotating gas motion inside the infalling envelope seen in the C$^{18}$O (1–0) line [@tak04]. The combined image shows another submillimeter CS component, which appears to extend toward the associated reflection nebula and hence the direction of the blueshifted molecular outflow, and the extension of this submillimeter emission component appears to be perpendicular to that of the C$^{18}$O (1–0) envelope (see Figure \[fig:l15comb\] $left$). The “asymmetric”, elongated distribution of the submillimeter molecular emissions toward the direction of the blueshifted molecular outflow is also seen in L483 (Paper I). The CS (7–6) line profile map shows that the emission peaks at the southwest of the protostar tend to be redshifted and at the northeast blueshifted (Figure \[fig:irs5cs\], see spectra with blue and red rectangles), although in the HCN (4–3) line profile map such a velocity structure is not clear (Figure \[fig:irs5hcn\]). This velocity feature in the submillimeter CS line appears to be opposite to that of the infalling envelope seen in the millimeter C$^{18}$O and H$^{13}$CO$^{+}$ (1–0) lines [@sai96; @mom98], and that of the CO outflows [@sto06; @mor06; @wu09]. Opposite velocity gradients of the submillimeter molecular lines from those of the millimeter lines and the associated outflows are also seen in L483 and B335 (Paper I). In $\S$4.2., we will present quantitative analyses of the submillimeter velocity gradients in these protostellar sources and discuss their origin. (160mm,160mm)[fig3.eps]{} (140mm,140mm)[fig4.eps]{} (140mm,140mm)[fig5.eps]{} (160mm,160mm)[fig6.eps]{} ### L1551 NE L1551 NE is the second brightest ($\sim$4.2 $\LO$) protostellar source in Taurus, located $\sim$2$\farcm$5 north-east of L1551 IRS 5 [@eme84]. L1551 NE is classified as a Class I protostar based on its spectral energy distribution [@fro05]. Radio continuum observations by @rod95 and @rei02 have revealed two compact sources with a separation of $\sim$70 AU and a position angle of $\sim$300$^{\circ}$, suggesting that L1551 NE is a protostellar binary. Optical and NIR observations of L1551 NE have found Herbig-Haro objects HH 28, HH 29, HH 454, and a collimated \[Fe II\] jet, and a bright reflection nebula [@dra85; @hod95; @dev99; @rei00; @hay09]. With JCMT, @mo95a detected blueshifted ($\sim$3 - 5 km s$^{-1}$ from the systemic velocity) emission in the wing of the CO (3–2) line toward L1551 NE that coincides spatially with its optical/infrared reflection nebula, suggesting that the blueshifted emission traces the molecular outflow driven by L1551 NE. Millimeter interferometric observations of L1551 NE in the CS (2–1, 3–2) lines revealed a compact ($\sim$700 AU) dense-gas component associated with the central protostar plus a NW-SE elongated ($\sim$6000 AU) component located toward $\sim$1000 AU west from the protostar [@yok03]. On the other hand, interferometric observations of the H$^{13}$CO$^{+}$ (1–0) line show a NW-SE elongated ($\sim$8000 AU) component located toward $\sim$700 AU east of the protostar [@sai01]. The prominent redshifted molecular outflow from L1551 IRS5 passes in projection through the location of L1551 NE, and likely impacts the molecular envelope around L1551 NE. The anti-correlation between the CS and H$^{13}$CO$^{+}$ emissions may be due to the shock chemistry [@pla95; @yok03]. Figures \[fig:necs\] and \[fig:nehcn\] show line profile maps of the CS (7–6) and HCN (4–3) emissions in L1551 NE, respectively. Figure \[fig:nemap\] shows the relevant total integrated intensity maps of the CS (7–6) ($left$ $panel$) and HCN (4–3) ($right$) lines in L1551 NE (red contours), superposed on the \[Fe II\] + 1.64-$\micron$ continuum image taken with SUBARU [@hay09]. Although the submillimeter molecular emissions are not spatially resolved with the ASTE 22$\arcsec$ beam, their peak positions appear to be $\sim$10$\arcsec$ west from the protostellar position. Along the east-west direction passing through the central protostellar position, the CS integrated intensity from -20$\arcsec$ to +10$\arcsec$ of the protostar exceeds above the 8$\sigma$ level, and the HCN integrated intensity from -10$\arcsec$ to +10$\arcsec$ above 4$\sigma$ level. Taking into account the detection levels, the ASTE pointing error ($\sim$2$\arcsec$) and the beam size, we consider that these $\sim$10$\arcsec$ offsets of the submillimeter emission peaks from the protostellar position are real. These offsets of the emission distributions resemble those of the millimeter CS (2–1, 3–2) emission distributions [@yok03]. We also note that the offsets of the submillimeter molecular emissions are toward the direction of the associated reflection nebula seen in the SUBARU image, which is similar to the case of L1551 IRS 5 (Fig. \[fig:l15comb\]). (140mm,140mm)[fig7.eps]{} (140mm,140mm)[fig8.eps]{} (140mm,140mm)[fig9.eps]{} ### L723 and L43 L723 is another Class 0 candidate with a luminosity of $\sim$3.3 $\LO$ at a distance of 300 pc [@dav87]. In L723 there are at least four centimeter sources; VLA 1, and VLA 2A, 2B, and 2C, and one 7-mm source; VLA 2D [@ang91; @car08]. VLA 1 is located at $\sim$4500 AU south-west from the VLA 2 complex, and only the VLA2 complex is associated with dense molecular gas seen in the millimeter CS (1–0, 2–1, 3–2) lines as well as the $X$-shaped CO outflow [@hay91; @hir98]. The peaks of the CS (2–1, 3–2) emission distributions are offset by $\sim$10$\arcsec$ east from the VLA 2 complex [@hir98]. SMA observations of L723 have found two compact ($\sim$600 AU) dusty cores at a projected separation of $\sim$880 AU (SMA 1 and SMA 2), where SMA 1 is associated with VLA 2D and SMA 2 VLA 2A, 2B, and 2C [@gir09]. Both SMA 1 and 2 show marginal evidence of infalling and rotating gas motion in the H$_{2}$CO (3$_{0,3}$–2$_{0,2}$) line, whereas only SMA 2 appears to be associated with HH objects and multiple (at least three) molecular outflows driven by the three different protostellar sources [@car08; @gir09]. Figures \[fig:l723cs\], \[fig:l723hcn\], and \[fig:l723map\] show line profile maps of the CS (7–6) and HCN (4–3) emission, and the corresponding total integrated intensity maps in L723, respectively. The CS and HCN total integrated intensities toward VLA 2 exceed $>$ 13$\sigma$, whereas those toward VLA 1 are below 3$\sigma$. These results suggest that only VLA 2 is associated with the dense and/or warm molecular gas traced by the submillimeter molecular lines, which is consistent with the previous millimeter observations [@hir98; @gir09]. The intensity of the CS emission at 10$\arcsec$ east of the protostar is comparable to that at the protostellar position ($\sim$14 $\sigma$), and is much higher than that at 10$\arcsec$ west ($\sim$6 $\sigma$). The HCN intensity at 10$\arcsec$ east ($\sim$10 $\sigma$) is also higher than that at 10$\arcsec$ west ($\sim$6 $\sigma$). Although the ASTE observations do not resolve the submillimeter emissions in L723, the differences of the submillimeter intensities between 10$\arcsec$ east and west of the protostar are probably real. Hence, the peaks of the submillimeter emission distributions in L723 are likely offset from the protostellar position, as in the case of L1551 NE. The millimeter CS (2–1, 3–2) emission distributions also show similar peak offsets by $\sim$10$\arcsec$ east from the protostar [@hir98]. The direction of these offsets is toward the associated blueshifted outflow. (140mm,140mm)[fig10.eps]{} (140mm,140mm)[fig11.eps]{} (120mm,120mm)[fig12.eps]{} Figures \[fig:l43cs\] and \[fig:l43hcn\] show line profile maps of the CS (7–6) and HCN (4–3) emission in L43, respectively. L43 ($or$ RNO 91, IRAS 16316-1540) is a Class I - II protostar with $L_{bol}$ $\sim$2.7 $\LO$ and $T_{bol}$ $\sim$370 K [@shi00; @ang02; @che09], and hence somewhat a more evolved source than the other sources of our sample. The protostar is associated with an $\sim$4000-AU scale dusty envelope [@shi00; @che09], and a bipolar molecular outflow along the SE - NW direction [@ben98; @lee00; @lee02]. Figures \[fig:l43cs\] and \[fig:l43hcn\] show that the signal-to-noise ratio of the submillimeter molecular lines is not high enough to discuss their spatial distributions. One interesting point is that L43 is the only source in our ASTE observations with the HCN (4–3) emission more intense than the CS (7–6) emission. This could be related to the later protostellar evolutionary stage of L43 than that of the other sources. (140mm,140mm)[fig13.eps]{} (140mm,140mm)[fig14.eps]{} Discussion ========== “Skewed” Distributions of the Submillimeter Molecular Emissions --------------------------------------------------------------- As described in the previous section, the submillimeter CS (7–6) and HCN (4–3) emissions toward low-mass protostars often show “skewed” distributions with respect to the protostellar positions. The combined ASTE + SMA image of the CS emission in L1551 IRS 5 illustrates that there is a halo-like component elongated ($\sim$2000 AU) toward the associated reflection nebula and the blueshifted outflow, as well as a central compact ($\lesssim$500 AU) component centered on the protostellar position. In L1551 NE, the emission peaks of the CS and HCN lines are likely offset from the protostellar position, and are shifted toward the associated reflection nebula. Similar peak offsets of the submillimeter CS and HCN emissions are also seen in L723. In L483, we found possible extensions of the CS ($\sim$2300 AU) and HCN ($\sim$5500 AU) emissions toward the direction of the associated reflection nebula and the blueshifted outflow (Paper I). The combined ASTE + SMA CS (7–6) image in B335, another Class 0 protostar, also shows an elongated ($\sim$2000 AU) component tracing the reflection nebula as well as a compact ($\lesssim$500 AU) component centered on the protostellar position [@yen11]. These results indicate that the asymmetric, or “skewed” distributions of the submillimeter molecular emissions toward the associated reflection nebulae are relatively frequent. One possible interpretation of these results is that the extended submillimeter molecular emissions at a few thousands AU scale trace “walls” of the cavity of the envelope evacuated by the associated outflows. In Figure \[fig:envconf\] we show a schematic diagram to explain the skewed distribution of the submillimeter molecular emissions. At the surface of the cavity wall molecular gas is irradiated by the stellar photons and hence the gas temperature is likely to be higher than that at the midplane [@nak03; @whi03]. As discussed in Paper I, the submillimeter CS and HCN emissions likely trace warmer ($\gtrsim$40 K) molecular gas than that traced by millimeter molecular emissions, and hence the warmer cavity walls are selectively traced by the submillimeter molecular emissions whereas the colder, midplane regions are traced by the millimeter emissions. Here, at the side of the associated reflection nebula the wall is directly seen along the line of sight, while at the other side the wall is likely obscured by the cold foreground envelope gas. The skewness of the submillimeter molecular emissions, where the submillimeter emissions are bright only at the side of the associated reflection nebula, may be explained by the foreground absorption along the line of sight. In order to illustrate this possibility, we performed simple calculations based on the Large Velocity Gradient (LVG) model [@gol74; @sco74]. The brightness temperature of the submillimeter emission at the obscured side ($\equiv$ $T_{obs}$) can be written as, $$T_{obs} = T_{re} e^{-\tau} + J(T_{\rm ex}) (1-e^{-\tau}) - J(T_{bg}),$$ where $$J(T)=\frac{\frac{h\nu}{k}}{\exp(\frac{h\nu}{kT})-1}.$$ In the above expressions, $h$ is the Planck constant, $k$ is Boltzmann’s constant, $\nu$ is the line frequency, $T_{\rm ex}$ and $\tau$ are the excitation temperature and the optical depth in the foreground cloud, $T_{bg}$ is the microwave background radiation temperature (= 2.725 K), and $T_{re}$ is the brightness temperature of the submillimeter emission from the warm cavity wall at the rear side, respectively. For simplicity, $T_{re}$ is fixed to be the peak brightness temperature of the halo-like CS emission in L1551 IRS 5 ($\equiv$ 9.2 K; see Fig. \[fig:l15comb\]). We calculated $T_{\rm ex}$ and $\tau$ in the foreground cloud with a gas kinetic temperature of 10 K for a different gas density ($\equiv$ $n_{H_{2}}$) and CS abundance per unit velocity gradient ($\equiv$ $X(CS)/dV/dR$) based on the LVG model, and then derived $T_{obs}$ as a function of the foreground gas density and the abundance. Details of our LVG calculation are described in Paper I. In Figure \[fig:lvg\] we show calculated $T_{obs}$ and $\tau$ as a function of $n_{H_{2}}$ for different $X(CS)/dV/dR$. A horizontal dashed line in the upper panel shows the 3 $\sigma$ upper limit of the combined ASTE + SMA CS (7–6) data ($\sim$3.6 K). This figure illustrates that if the foreground cloud is optically thick against the submillimeter CS emission, the emission from the rear side could be significantly absorbed, down to the sensitivity limit of the ASTE + SMA CS (7–6) data. The range of the foreground gas density where the line intensity along the line of sight is lower than the sensitivity limit depends on adopted $X(CS)/dV/dR$. The typical $X(C^{34}S)/dV/dR$ value in cold dark clouds was estimated to be $\sim$5.0 $\times$ 10$^{-11}$ km $^{-1}$ s pc from the multi-transitional C$^{34}$S (1–0, 2–1) observations [@tak00]. Then the typical $X(CS) / X(C^{34}S)$ ratio of 22 [@chi96] yields $X(CS)/dV/dR$ = $X(C^{34}S)/dV/dR$ $\times$ 22 $\sim$10$^{-9}$. With $X(CS)/dV/dR$ = 10$^{-9}$ the density range is $\sim$10$^{5.5-6.3}$ cm$^{-3}$, which is typical in cloud cores, and the optical depth ranges from $\sim$1 to $\sim$14. Two-dimensional radiative transfer models of protostellar envelopes [@spa95; @nak03; @whi03] are likely to support our interpretation that the extended submillimeter molecular emissions at a few thousands AU scale trace walls of the cavity of the envelopes evacuated by the associated outflows. @spa95 suggested that the narrow $^{12}$CO and $^{13}$CO $J$=6–5 emissions detected toward many low-mass young stellar objects are produced at the wall of the cavity evacuated by the bipolar outflow, heated by the radiation field generated in the inner part of the accretion disk. @nak03 and @whi03 demonstrated that at the surface of the wall of the cavity opened by the outflow in the envelope the gas temperature becomes higher than that predicted from spherically symmetric models, due to the heating by the protostellar photons. In fact, the intensities of the submillimeter molecular lines observed with the ASTE 22$\arcsec$ beam appear to correlate with the central protostellar luminosities, as discussed in $\S$3.1. To quantitatively show this trend, we compiled published CS (7–6) data [@bla95; @mo95b] and protostellar luminosities [@mor94; @fro05]. In Figure \[fig:corr\] we plot integrated intensities of the CS (7–6) emission versus bolometric luminosities of the protostellar sources. The data points with filled circles are from Tables \[tab:source\] and \[tab:spectra\] in the present paper. There is a clear linear correlation between the source bolometric luminosities and the submillimeter CS intensities, and the least square fitting to the data points yields $I_{CS}$ $\propto$ $L_{bol}^{0.92}$. The linear correlation between the source luminosities and the intensities of the submillimeter molecular emission may also support the idea that the submillimeter molecular emission at a few thousands AU scale traces the wall of the envelope cavities heated by the protostars. Detailed physical and chemical models of the envelopes and the 3-dimensional radiative transfer calculations should help to understand the distributions and the origins of the submillimeter molecular lines in protostellar envelopes. (100mm,100mm)[fig15.eps]{} (70mm,70mm)[fig16.eps]{} (100mm,100mm)[fig17.eps]{} Different Kinematics between the Millimeter and Submillimeter Molecular Lines in the Protostellar Envelopes ----------------------------------------------------------------------------------------------------------- In Paper I, we suggested that in L483 and B335 the submillimeter molecular lines show opposite velocity gradients to those of the millimeter lines and the outflows along the outflow axis. As mentioned in $\S$3.2.1, from our new ASTE data we found that the submillimeter CS emission in L1551 IRS 5 is also likely to show a similar, opposite velocity gradient. In this final subsection, we will show how significant these opposite velocity gradients in the submillimeter molecular line are with our statistical analyses, and discuss the origin of the opposite velocity gradients. To unambiguously assess the statistical significance of the velocity gradients, we adopted a method described by Goodman et al. (1993). We first fit a single Gaussian function to each spectrum whose peak intensity is above 3 $\sigma$, to derive the central $LSR$ velocity of each spectrum ($\equiv$ $v_{LSR}$). The error of the derived central velocity ($\equiv$ $\sigma_{v_{LSR}}$) can be estimated as $$\sigma_{v_{LSR}} = 1.15(\frac{\sigma_T}{T})(\delta_v \Delta v)^{\frac{1}{2}},$$ where $T$ is the peak of a Gaussian fit to the line profile, $\Delta v$ is the FWHM line width of a Gaussian fit, $\sigma_T$ is the rms noise in the spectrum, and $\delta_v$ is the ASTE velocity resolution [@lan82]. Then, we fit a plane function to the estimated central velocities with the errors at the different positions as, $$v_{LSR} = v_{0} + a\Delta \alpha + b\Delta \delta,$$ where ($\Delta \alpha$, $\Delta \delta$) denotes the positions of the spectra with respect to the central protostellar position along $R.A.$ and $Decl.$, $v_{0}$ is the central velocity of the spectrum toward the central protostellar position, or the systemic velocity, and $a$ and $b$ are the velocity gradients along $R.A.$ and $Decl.$, respectively. In the case of L483 and B335 the outflow axes are along the $R.A.$ direction, and hence $a$ and $b$ denote the velocity gradients along and across the outflow direction, respectively. In the case of L1551 IRS 5 we rotated the vector ($a$, $b$) by -18$^{\circ}$ to estimate the velocity gradients along and across the outflow axis. These fittings enable us to estimate the velocity gradients and the statistical errors along and across the outflow direction unambiguously. Figures \[fig:l1551csfit\], \[fig:l1551hcnfit\], \[fig:l483csfit\], \[fig:l483hcnfit\], and \[fig:b335csfit\] show the results of the Gaussian fittings to the CS and HCN spectra in L1551 IRS 5, CS and HCN spectra in L483, and the CS spectra in B335, respectively. Table \[tab:plane\] summarizes the results of the plane fittings. Figure \[fig:l1551csfit\] shows that the CS spectra in L1551 IRS 5 at the north-eastern side are blueshifted whereas those at the south-western side redshifted. This sense of the velocity gradient is opposite to the sense of the infalling envelope and the associated outflow. The estimated velocity gradient along the outflow direction is $\sim$(-9.7$\pm$1.7) $\times$ 10$^{-3}$ km s$^{-1}$ arcsec$^{-1}$ (here the negative sign denotes the opposite velocity gradient to that of the associated outflow and the millimeter molecular lines). The derived velocity gradient is $\sim$5.9 times the statistical error, and hence the presence of the velocity gradient in the CS emission opposite to that of the millimeter lines and the outflow is statistically significant. In L483, Figure \[fig:l483csfit\] shows that the CS emission tends to be blueshifted at the eastern side and redshifted at the western side. The estimated value of the velocity gradient along the outflow direction is (-7.6$\pm$2.4) $\times$ 10$^{-3}$ km s$^{-1}$ arcsec$^{-1}$, $\sim$3.2 times the statistical error. No statistical significance of the opposite velocity gradient in the CS (7–6) emission was obtained in B335. Contrary to the CS (7–6) emission, no significant velocity gradient along any direction was verified in the HCN (4–3) emission (Table \[tab:plane\]). One possible interpretation of the absence of the velocity gradients in the HCN (4–3) emission is the blending from the hyperfine components. As already mentioned in $\S$3.1., two of the six hyperfine components, $F$ = 3–3 and $F$ = 4–4, are shifted by -1.67 km s$^{-1}$ and 1.36 km s$^{-1}$ from the main component ($F$ = 4–3), respectively. The typical line widths measured from the CS (7–6) line in those protostellar envelopes are 1 - 2 km s$^{-1}$, and hence the $F$ = 3–3 and $F$ = 4–4 hyperfine components could totally smear the velocity gradient in the main $F$ = 4–3 component if the intensity of those satellite lines are comparable to that of the main line. The much broader HCN line widths than the CS line widths found in L483 and L723 could also be due to the presence of the hyperfine lines. From these statistical analyses, we suggest that at least the CS (7–6) line in L1551 IRS 5 and L483 shows opposite velocity gradients to those of the millimeter lines and the associated outflows. It is therefore intriguing to discuss the origin of these opposite velocity gradients in the submillimeter line. As discussed in the last subsection the submillimeter molecular emission likely traces the warm surface of the cavity opened by the outflow in the envelope, whereas the millimeter emission the colder, infalling region at the midplane. Therefore, the submillimeter molecular line should trace a distinct gas motion at the cavity wall in the envelopes. A possible interpretation of the observed velocity structures seen in the submillimeter line is an expanding gas motion which is perpendicular to the outflow axis. The surface of the cavity wall in the envelope traced by the submillimeter emission could be stripped away by some mechanism such as the stellar wind. As shown in Figure \[fig:envconf\], on the side of the blueshifted outflow the bulk of the expanding gas at the cavity wall must be observed as a redshifted component and on the side of the redshifted outflow as a blueshifted component. On the other hand, the infalling gas component shows the same velocity gradient as that of the outflow. Hence, this configuration could explain the opposite velocity gradient observed in the submillimeter line to that of the outflow or infalling gas traced by the millimeter lines. In L1551 IRS 5, Pyo et al. (2005) have observed low-velocity ($\sim$100 km s$^{-1}$) \[Fe II\] winds with a wide opening angle of $\sim$100$^{\circ}$, as well as collimated high-velocity ($\sim$300 km s$^{-1}$) jets along the polar axis. They have suggested that such a wind with a wide opening angle will be effective in sweeping up envelope material from the close vicinity of its driving source. The observed velocity structure of the submillimeter CS line in L1551 IRS 5 may trace such a gas motion in the envelope, although the spatial scale in our observations ($\sim$2500 AU) is much larger than that of the \[Fe II\] observations ($\sim$100 - 400 AU). On the other hand, previous SMA observations of protostellar envelopes in submillimeter molecular lines have found compact ($\lesssim$500 AU) components associated with the central protostars, and those components show mainly rotational gas motion around the protostars [@tak04; @ta07b; @bri07; @lom08; @jor09; @yen11]. As already mentioned in the last subsection, our combined ASTE + SMA images of the protostellar envelopes around L1551 IRS 5 and B335 clearly show both the compact components and the extended components tracing the reflection nebulae. These results suggest that there are two distinct submillimeter components around protostellar sources; one is a compact ($\lesssim$500 AU), rotating gas component in the vicinity of the central protostars, and the other extended ($\sim$2000 AU) feature probably tracing the warm cavity wall of the envelope irradiated by the central protostars. (150mm,150mm)[fig18.eps]{} (120mm,120mm)[fig19.eps]{} (120mm,120mm)[fig20.eps]{} (120mm,120mm)[fig21.eps]{} (100mm,100mm)[fig22.eps]{} [llcccc]{} Line &Source &$a$&$b$&$v_{0}$&$rms$\ & &$\times$ 10$^{-3}$ (km s$^{-1}$ arcsec$^{-1}$) &$\times$ 10$^{-3}$ (km s$^{-1}$ arcsec$^{-1}$) &(km s$^{-1}$) &(km s$^{-1}$)\ CS ($J$=7–6) &L1551 IRS 5 &-9.7$\pm$1.7 &-3.6$\pm$1.4 &6.51$\pm$0.03 &0.14\ &L483 &-7.6$\pm$2.4 &3.6$\pm$2.8 &5.52$\pm$0.05 &0.14\ &B335 &-6.5$\pm$5.8 &-3.1$\pm$6.5 &8.18$\pm$0.06 &0.07\ HCN ($J$=4–3) &L1551 IRS 5 &1.2$\pm$1.7 &6.1$\pm$2.1 &6.34$\pm$0.04 &0.09\ &L483 &1.4$\pm$3.4 &8.5$\pm$3.4 &5.62$\pm$0.14 &0.06\ Summary ======= We have conducted mapping observations of L1551 IRS 5, L1551 NE, L723, and L43, and single-point observations of IRAS 16293-2422 in the submillimeter CS ($J$ = 7–6) and HCN ($J$ = 4–3) lines with ASTE. We have analyzed the present new ASTE data as well as our previous ASTE data of L483, B335, L723, and IRAS 16293-2422 (Paper I) in a systematical way, and have obtained the following results. - We detected the CS and HCN lines toward all the protostellar positions above 4$\sigma$ level, except for the CS and HCN lines at the protostellar position of L1551 NE and the CS line at the protostellar position of L43. The CS intensity ranges from $\lesssim$0.5 K (L1551 NE) to $\sim$9.1 K (IRAS 16293-2422), and the HCN intensity from $\sim$0.4 K (L723) to $\sim$5.8 K (IRAS 16293-2422). The CS line is stronger than the HCN line toward all the sources except for L43. There is a linear correlation between the source luminosities and the intensities of these submillimeter lines ($I_{CS}$ $\propto$ $L_{bol}^{0.92}$). - Our mapping observations of the protostellar envelopes in the CS and HCN emissions show that the submillimeter emissions often exhibit “skewed” distributions toward the direction of the associated reflection nebulae. In the combined ASTE + SMA CS (7–6) image of L1551 IRS 5 there appears an extended ($\sim$2000 AU) component tracing the associated reflection nebula at the southwest, as well as a compact ($\lesssim$500 AU) component centered on the protostellar position. The peaks of the CS and HCN emissions in L1551 NE are not located at the protostellar position but offset by $\sim$1400 AU toward the west, where the associated reflection nebula resides. The CS emission in L723 is also skewed toward the direction of the blueshifted outflow. These results are consistent with our earlier result in L483, where the CS and HCN emissions are resolved and show elongation ($>$ 2000 AU) toward the direction of the associated reflection nebula (Paper I). We suggest that these skewed submillimeter molecular emissions at a few thousands AU scale trace the warm ($\gtrsim$40 K) walls of the envelope cavities, excavated by the associated blueshifted outflows and irradiated by the central protostars directly. On the other hand, the submillimeter emissions at the other side are obscured due to the absorption from the cold ($\sim$10 K) and dense ($\sim$10$^{6}$ cm$^{-3}$) foreground envelope material. The detected linear correlation between the protostellar luminosities and the intensities of the submillimeter molecular lines may also support this interpretation. - From our statistical analyses, we verified that along the outflow directions the CS (7–6) emission in L1551 IRS 5 and L483 shows opposite velocity gradients to those of the millimeter molecular emissions and the associated outflows. The velocity gradients are estimated to be (9.7$\pm$1.7) $\times$ 10$^{-3}$ km s$^{-1}$ arcsec$^{-1}$ in L1551 IRS 5 and (7.6$\pm$2.4) $\times$ 10$^{-3}$ km s$^{-1}$ arcsec$^{-1}$ in L483. One possible interpretation on the origin of the different velocity gradients is that the submillimeter molecular line traces the dispersing gas motion at the surface of of the cavity opened by the outflow in the envelope, which is perpendicular to the outflow direction. The absence of any clear velocity gradient in the HCN (4–3) emission may be due to the presence of the hyperfine components and the blending of the velocity features. We are grateful to N. Ohashi, P. T. P. Ho, and Masao Saito for their fruitful discussions. We thank all the ASTE staff for their dedicated support of the telescope and observatory operations. Observations with ASTE were carried out remotely from Japan by using NTT’s GEMnet2 and its partner R$\&$E (Research and Education) networks, which are based on AccessNova collaboration of University of Chile, NTT Laboratories, and National Astronomical Observatory of Japan. A part of this study was financially supported by the MEXT Grant-in-Aid for Scientific Research on Priority Areas No. 15071202. S. T. acknowledges the grant from the National Science Council of Taiwan (NSC 99-2112-M-001-009-MY3) in support of this work. André, P., Ward-Thompson, D., & Barsony, M. 2000, in Protostars and Planets IV, ed. V. Mannings, A. P. Boss, & S. S. Russell (Tucson, AZ: Univ. of Arizona Press), 59 Anglada, G., Estalella, R., Rodríguez, L. F., Torrelles, J. M., López, R., & Cantó, J. 1991, , 376, 615 Anglada, G., & Rodríguez, L. F. 2002, RMxAA, 38, 13 Bence, S. J., Padman, R., Isaak, K. G., Wiedner, M. C., & Wright, G. S. 1998, , 299, 965 Blake, G. A., van Dishoeck, E. F., Jansen, D. J., Groesbeck, T. D., & Mundy, L. G. 1994, , 428, 680 Blake, G. A., Sandell, G., van Dishoeck, E. F., Groesbeck, T. D., Mundy, L. G., & Aspin, C. 1995, , 441, 689 Brinch, C., Crapsi, A., J[ø]{}rgensen, J. K., Hogerheijde, M. R., & Hill, T. 2007, , 475, 915 Carrasco-González, C., Anglada, G., Rodríguez, L. F., Torrelles, J. M., Osorio, M., & Girart, J. M. 2008, , 676, 1073 Chen, J.-H., Evans, N. J., Lee, J.-E., & Bourke, T. L. 2009, , 705, 1160 Chin, Y.-N., Henkel, C., Whiteoak, J. B., Langer, N., & Churchwell, E. B. 1996, , 305, 960 Davidson, J. A. 1987, , 315, 602 Devine, D., Reipurth, B., & Bally, J. 1999, , 118, 972 Draper, P. W., Warren-Smith, R. F., & Scarrott, S., M. 1985, , 216, 7 Emerson, J. P., Harris, S., Jennings, R. E., Beichman, C. A., Baud, B., Beintema, D. A., Marsden, P. L., Wesselius, P. R. 1984, , 278, L49 Ezawa, H., Kawabe, R., Kohno, K., & Yamamoto, S. 2004, Proc. SPIE, 5489, 763 Froebrich D. 2005, , 156, 169 Girart, J. M., Rao, R., & Estalella, R. 2009, , 694, 56 Goldreich, P., & Kwan, J. 1974, , 189, 441 Goodman, A. A., Benson, P. J., Fuller, G. A., & Myers, P. C. 1993, , 406, 528 Gregersen, E. M., Evans II, N. J., Zhou, S., & Choi, M. 1997, , 484, 256 Hayashi, S. S., Hasegawa, T., & Kaifu, N. 1991, , 377, 492 Hayashi, M., & Pyo, T.-S. 2009, , 694, 582 Hirano, N., Hayashi, S. S., Umemoto, T., & Ukita, N. 1998, , 504, 334 Hodapp, K.-W., & Ladd, E. F. 1995, , 453, 715 Jewitt, D. C., Matthews, H. E., Owen, T., & Meier, R. 1997, Science, 278, 90 J[ø]{}rgensen, J. K., van Dishoeck, E. F., Visser, R., Bourke, T. L., Wilner, D. J., Lommen, D., Hogerheijde, M. R., & Myers, P. C. 2009, , 507, 861 Kamazaki, T., et al. 2005, Astronomical Society of the Pacific Conference Series, 347, 533 Kohno, K. 2005, in ASP Conf. Ser. 344, The Cool Universe: Observing Cosmic Dawn, ed C. Lidman & D. Alloin, (San Francisco: ASP), 242 Landman, D. A., Roussel-Dupre, R., & Tanigawa, G. 1982, , 261, 732 Lee, C.-F., Mundy, L. G., Reipurth, B., Ostriker, E. C., & Stone, J. M. 2000, , 542, 925 Lee, C.-F., Mundy, L. G., Stone, J. M., & Ostriker, E. C. 2002, , 576, 294 Lim, J., & Takakuwa, S. 2006, , 653, 425 Lommen, D., J[ø]{}rgensen, J. K., van Dishoeck, E. F., & Crapsi, A. 2008, , 481, 141 Momose, M., Ohashi, N., Kawabe, R., Nakano, T., & Hayashi, M. 1998, , 504, 314 Moriarty-Schieven, G. H., Wannier, P. G., Keene, J., & Tamura, M. 1994, , 436, 800 Moriarty-Schieven, G. H., Butner, H. M., & Wannier, P. G. 1995, , 445, L55 Moriarty-Schieven, G. H., Wannier, P. G., Mangum, J. G., Tamura, M., & Olmsted, V. K. 1995, , 455, 190 Moriarty-Schieven, G. H., Powers, J. A., Butner, H. M., Wannier, P. G., & Keene, J. 2000, , 533, L143 Moriarty-Schieven, G. H., Johnstone, D., Bally, J., & Jenness, T. 2006, , 645, 357 Myers, P. C., Evans, N. J., II, & Ohashi, N. 2000, in Protostars and Planets IV, ed. V. Mannings, A. P. Boss, & S. S. Russel (Tucson, AZ: Univ. of Arizona Press), 217 Nakazato, T., Nakamoto, T., & Umemura, M. 2003, , 583, 322 Ohashi, N., Hayashi, M., Ho, P. T. P, Momose, M., & Hirano, N. 1996, , 466, 957 Ohashi, N., Hayashi, M., Ho, P. T. P., & Momose, M. 1997a, , 475, 211 Ohashi, N., Hayashi, M., Ho, P. T. P., Momose, M., Tamura, M., Hirano, N., & Sargent, A. I. 1997b, , 488, 317 Park, Y.-S., Panis, J.-F., Ohashi, N., Choi, M., & Minh, Y. C. 2000, , 542, 344 Plambeck, P. L. & Snell, R. L. 1995, , 446, 234 Pyo, T.-S., Hayashi, M., Kobayashi, N., Tokunaga, A. T., Terada, H., Tsujimoto, M., Hayashi, S. S., Usuda, T., Yamashita, T., Takami, H., Takato, N., & Nedachi, K. 2005, , 618, 817 Reipurth, B., Yu, K. C., Heathcote, S., Bally, J., & Fodríguez, L. F. 2000, , 120, 1449 Reipurth, B., Rodríguez, L. F., Anglada, & G., Bally, J. 2002, , 124, 1045 Rodríguez, L. F., Anglada, G., & Raga, A. 1995, , 454, L149 Saito, M., Kawabe, R., Kitamura, Y., & Sunada, K. 1996, , 473, 464 Saito, M., Kawabe, R., Kitamura, Y., & Sunada, K. 2001, , 547, 840 Scoville, N. Z., & Solomon, P. M. 1974, , 187, L67 Shirley, Y. L., Evans II, N. J., Rawlings, J. M. C., & Gregersen, E. M. 2000, , 131, 249 Snell, R. L., Loren, R. B., & Plambeck, R. L. 1980, , 239, 17 Spaans, M., Hogerheijde, M. R., Mundy, L. G., & Van Dishoeck, E. F. 1995, , 455, L167 Stojimirović, I., Narayanan, G., Snell, R. L., Bally, J. 2006, , 649, 280 Takakuwa, S., Mikami, H., Saito, M., & Hirano, N. 2000, , 542, 367 Takakuwa, S., Ohashi, N., Ho, P. T. P., Chunhua, Q., Wilner, D. J., Qizhou, Z., Bourke, T. L., Hirano, N., Choi, M., & Yang, J. 2004, , 616, L15 Takakuwa, S., Kamazaki, T., Saito, M., Yamaguchi, N., & Kohno, K. 2007a, , 59, 1 (Paper I) Takakuwa, S., et al. 2007b, , 662, 431 Uchida, Y., Kaifu, N., Shibata, K., Hayashi, S. S., Hasegawa, T., & Hamatake, H. 1987, , 39, 907 van Dishoeck, E. F., Blake, G. A., Jansen, D. J., & Groesbeck, T. D. 1995, , 447, 760 Whitney, B. A., Wood, K., Bjorkman, J. E., & Wolff, M. J. 2003, , 591, 1049 Wu, P.-F., Takakuwa, S., & Lim, J. 2009, , 698, 184 Yen, H.-W., Takakuwa, S., & Ohashi, N. 2010, , 710, 1786 Yen, H.-W., Takakuwa, S., & Ohashi, N. 2011, , submitted Yokogawa, S., Kitamura, Y., Momose, M., & Kawabe, R. 2003, , 595, 266 [^1]: The ASTE project is driven by Nobeyama Radio Observatory (NRO), a branch of National Astronomical Observatory of Japan (NAOJ), in collaboration with University of Chile, and Japanese institutes including University of Tokyo, Nagoya University, Osaka Prefecture University, Ibaraki University., and Hokkaido University.
{ "pile_set_name": "ArXiv" }
--- author: - | Jay Bartroff\ \ title: '[**A New Characterization of Elfving’s Method for High Dimensional Computation**]{}' --- \[section\] \[section\] \[section\] \[section\] \[section\] \[section\] \[section\] \[section\] Introduction ============ @Elfving52 gave an elegant geometric characterization of $c$-optimal designs for linear regression models by associating the design’s support points and weights with a weighted and signed linear combination of those points, for a certain optimal choice of signs [see also @Chernoff72; @Chernoff99]. This can be used to immediately find $c$-optimal designs in 1 or 2 dimensions, but the state of the art for $k\ge 3$ dimensions is an algorithm of @Lopez-Fidalgo04 that involves searching over the unknown signs and support points. After introducing our notation and assumptions in Section \[sec:set-up\], in Section \[sec:result\] we give a new result on Elfving’s method which gives explicit formulae for the optimal signs and weights for any candidate set of support points, thus reducing the computational cost of the design, greatly when $k$ is large. Torsney and coauthors [@Torsney81; @Kitsos88; @Pukelsheim91] have given a different characterization of the optimal weights, but the novelty of the current method is (i) by giving both the optimal weights and signs, the current method allows fast computation by using Elfving’s geometric method along the lines of @Lopez-Fidalgo04 without needing to search over all signs, and (ii) the formulae for the weights and signs also hold for suboptimal support points, which are often used in practice. In Section \[sec:examples\] we discuss two examples: high-dimensional polynomial regression and estimating the turning point of a quadratic logistic regression model, in which we use Ford, Torsney, and Wu’s [-@Ford92] results for locally $c$-optimal designs in generalized linear models. In addition to the references already mentioned, there has been much recent research on Elfving’s problem, of which @Studden05 gives a review. @Dette93 showed that the popular $D$-optimal designs can be written as a sequence of $c$-optimal design problems and also considered a Bayesian approach. Later, @Dette97 proposed “standardized” optimality criteria to optimize estimation of a set of parameter values, in various senses, which involve $c$-optimal designs as their constituents. Elfving’s method has been generalized by Dette and coauthors in several other directions as well in @Dette93b, @Dette93d [@Dette93c; @Dette96], @Dette08, @Dette09, and @Dette10. @Fan03 characterized a certain generalized inverse of a candidate $c$-optimal design that can be used to assess optimality. In a recent application, @Bartroff10b [@Bartroff10d] have proposed using $c$-optimal designs as “base designs” in dynamic programming-based dose-finding clinical trial designs which are computed repeatedly in sequential Monte Carlo simulation, putting a premium on efficient computation. Notation and assumptions {#sec:set-up} ------------------------ We assume that the scalar-valued response variable $Y$ is normally distributed with mean $\theta'x$ and variance $\sigma^2$, where prime denotes transpose, $\theta\in\mathbb{R}^k$ is the unknown parameter, and $x$ is the design point taking values in the compact design space $\mathcal{X}\subseteq \mathbb{R}^k$. Our main assumption is that any collection of linearly independent (LI) vectors $x_1,\ldots,x_\ell\in\mathcal{X}$, $\ell<k$, can be completed to a LI set $x_1,\ldots,x_\ell,x_{\ell+1},\ldots,x_k\in\mathcal{X}$ of size $k$; this is not a strong assumption and any $\mathcal{X}$ not satisfying this can be reduced to a smaller dimensional case, in some sense. We define the information matrix of a given measure $\xi$ on $\mathcal{X}$ as $$M(\xi)=\int_{\mathcal{X}}xx'\xi(dx).$$ For a given nonzero vector $c\in\mathbb{R}^k$, the $c$-optimal design problem is to find a design measure $\xi$ achieving $$\label{uncon-problem} \min_{\xi\in\Xi} \Psi(\xi),{\quad\mbox{where}\quad} \Psi(\xi)=c'M(\xi)^{-1}c.$$ Here $\Xi$ is the class of designs for which $c'\theta$ is *estimable*, i.e., $$\Xi=\left\{\xi: \int_{\mathcal{X}}\xi(dx)=1{\quad\mbox{and}\quad} c=M(\xi)d {\quad\mbox{for some $d\in\mathbb{R}^k$}}\right\}.$$ The class $\Xi$ may admit singular designs in which case $M(\xi)^{-1}$ in means a generalized inverse [see @Seber03 p. 469] and the quadratic form $c' M(\xi)^{-1}c$ is invariant under which generalized inverse is chosen. Also define $$\Xi(x_1,\ldots,x_\ell)=\{\xi\in\Xi: \mbox{supp}(\xi)\subseteq \{x_1,\ldots,x_\ell\}\}.$$ @Elfving52 gave the following elegant solution to the problem (\[uncon-problem\]). Let $\mathcal{X}^-$ denote the reflection of $\mathcal{X}$ through the origin and $\mathcal{E}=\mbox{conv}(\mathcal{X}\cup \mathcal{X}^-)$, the convex hull of $\mathcal{X}\cup \mathcal{X}^-$. Points in $\mathcal{E}$, the so-called Elfving set, can be written in the form $\sum_{i=1}^{n} p_i{\varepsilon}_ix_i$, where $p_i\ge 0$, $\sum_{i=1}^n p_i=1$, ${\varepsilon}_i\in\{-1,+1\}$, and $x_i\in \mathcal{X}$. Let $z=\sum_{i=1}^{n} p_i{\varepsilon}_ix_i\in \mathcal{E}$ be the point on the ray $$\label{Rc} R(c)=\{\gamma c\in\mathbb{R}^k: \gamma\in\mathbb{R}\}$$ that is furthest from the origin. Then the design $\{(x_i,p_i)\}_{i=1}^n$, which puts weight $p_i$ on point $x_i$, $i=1,\ldots,n$, is $c$-optimal. Moreover, $n$ can be taken to be $k$ by Carathéodory’s theorem, and $$\label{ElfPsic/z} \Psi(\{(x_i,p_i)\})=||c||^2/||z||^2.$$ See @Chernoff72 [@Chernoff99]. A new characterization of Elfving’s method and an improved algorithm {#sec:result} ==================================================================== For LI vectors $x_1,\ldots,x_\ell\in\mathcal{X}$, $\ell\le k$, such that $c\in{\mbox{span}}(x_1,\ldots,x_\ell)$, define the design $\xi^*(x_1,\ldots,x_\ell)$ and the point $z^*(x_1,\ldots,x_\ell)\in \mathbb{R}^k$ as follows: $$\begin{aligned} \xi^*(x_1,\ldots,x_\ell)&=\{(x_i,p_i^*)\}_{i=1}^\ell\label{xi*}\\ z^*(x_1,\ldots,x_\ell)&=\sum_{i=1}^\ell {\varepsilon}_i^*p_i^*x_i,\label{z*}\end{aligned}$$ where $$\begin{aligned} {\varepsilon}_i^*&={\varepsilon}_i^*(x_1,\ldots,x_\ell)=\mbox{sign}\left(x_i'M^{-1}c\right),\quad i=1,\ldots,\ell,\label{epsi}\\ M&=M\left(\{(x_i,1/\ell)\}_{i=1}^\ell \right)\;\mbox{and $M^{-1}$ any generalized inverse,}\label{M}\\ p_i^*&=p_i^*(x_1,\ldots,x_\ell)={\left\vertx_i'M^{-1}c\right\vert}\left/\sum_{j=1}^\ell {\left\vertx_j'M^{-1}c\right\vert}\right.,\quad i=1,\ldots,\ell.\label{pi}\end{aligned}$$ Our main result, Theorem \[thm:phi\], is that solving the single optimization problem of maximizing the function $$\phi(x_1,\ldots,x_k)=||z^*(x_1,\ldots,x_k)||^2$$ over $\mathcal{S}=\{(x_1,\ldots,x_k)\in\mathcal{X}^k: x_1,\ldots,x_k\;\mbox{are LI}\}$ gives a $c$-optimal design. The proof of Theorem \[thm:phi\] is given in the Appendix, as well as proofs of Proposition \[prop:xi\*\], which establishes the properties of $\xi^*(x_1,\ldots,x_\ell)$ and $z^*(x_1,\ldots,x_\ell)$, and Lemma \[lem:span=C\] relating $\mbox{span}(x_1,\ldots,x_\ell)$ to the column space of $M(\{(x_i,q_i)\}_{i=1}^\ell)$. \[thm:phi\] Let $(x_1^*,\ldots,x_k^*)=\arg\max_\mathcal{S} \phi(x_1,\ldots,x_k)$. Then the design $$\xi^*(x_1^*,\ldots,x_k^*)=\{(x_i^*,p_i^*)\}_{i=1}^k$$ is $c$-optimal, where $p_i^*=p_i^*(x_1^*,\ldots,x_k^*)$ are given by . **Remarks.** 1. There is nothing special about the weights $1/\ell$ in . In fact, any strictly positive constants $a_1,\ldots,a_\ell$ (not necessarily summing to $1$) can be used if we replace $M$ in - by $\sum_{i=1}^\ell a_ix_ix_i'$ and $p_i^*$ in by $$p_i^*=a_i{\left\vertx_i'M^{-1}c\right\vert}\left/\sum_{j=1}^\ell a_j{\left\vertx_j'M^{-1}c\right\vert}\right..$$ 2. In practice, one is often constrained to apply a suboptimal design with support on some given set. Thus it is convenient to be able to compute the optimal weights for the given support set. Claim \[Psimin\] of Proposition \[prop:xi\*\], given in the Appendix, shows that can be used to find these. 3. For many models of interest, the vectors $x\in\mathcal{X}\subseteq\mathbb{R}^k$ are actually functions $x=x(u)$ of a scalar $u\in\mathcal{U}\subseteq\mathbb{R}$, for some closed interval $\mathcal{U}$. In this case the optimization problem of size $k^2$ (i.e., $k$ vectors in $\mathbb{R}^k$) in Theorem \[thm:phi\] reduces to an optimization problem of size $k$. Illustrative examples {#sec:examples} ===================== Polynomial regression --------------------- The method of Theorem \[thm:phi\] was used to compute the $c$-optimal designs in the polynomial model $Y\sim\mathcal{N}(\sum_{i=1}^{k}\theta_iu^{i-1},\sigma^2)$ for $k=6,\ldots,10$, $u\in[-1,1]$, and for $c$ equal to the $k$ standard basis vectors $e_1=(1,0,\ldots,0)',\ldots,e_k=(0,\ldots,0,1)'$. The corresponding designs for $k=3,4$, and $5$ appear in Lopez-Fídalgo and Rodriguez-Diaz[^1] [-@Lopez-Fidalgo04 Table 5], and the $k=1,2$ cases are well known. Here, in Table \[table:poly\], and in the next section, we abuse the above notation slightly by parametrizing designs $\xi^*$ by $u\in[-1,1]$ rather than $x=x(u)=(1,u,u^2,\ldots,u^{k-1})'\in{\mathbb{R}}^k$. A few interesting properties of these designs are (i) all are symmetric about $u=0$, (ii) for all $k$, the $e_1$-optimal design is $\xi^*=\{(0,1)\}$ with $\Psi(\xi^*)=1$, and (iii) for all $k$, the $e_j$-optimal design has positive weight at $u=0$ if and only if $j$ is odd. These and other properties were noted by @Studden68 who solved this problem explicitly using Tchebycheff systems. The algorithm was implemented in using its optimizer. On a 2.6 GHz laptop computer, the design in Table \[table:poly\] that took the longest to compute took roughly 3 seconds. We have computed these designs up to dimension $k=20$ using the same method with the longest taking roughly 10 seconds. Quadratic logistic regression {#sec:logistic} ----------------------------- We consider $c$-optimal designs for finding the turning point of a quadratic logistic regression model. Suppose $Y$ takes the values 0 or 1 according to $$\label{logit} P_\theta(Y=1|u)=1/(1+e^{-(\theta_1+\theta_2 u+\theta_3 u^2)}),$$ where $\theta=(\theta_1,\theta_2,\theta_3)'$ and the scalar $u$ takes values in some interval, taken here to be $[-1,1]$. For example, this quadratic logistic regression is popular for modeling the response to treatments for diseases other than cancer, for which efficacy is assumed to have a unimodal relationship to dose $u$ [e.g., @Thall04 p. 686]. In settings such as this it is assumed that $\theta_3<0$ and it is desired to estimate the dose $\eta=-\theta_2/(2\theta_3)$ maximizing probability of response $Y=1$. Given an estimate ${\widehat{\theta}}$ of $\theta$ with ${\widehat{\theta}}_3\ne 0$, we consider this as a locally (i.e., depending on $\theta$) $c$-optimal design problem by first writing $$\eta\approx -{\widehat{\theta}}_2/2{\widehat{\theta}}_3-(\theta_2-{\widehat{\theta}}_2)/(2{\widehat{\theta}}_3)+(\theta_3-{\widehat{\theta}}_3){\widehat{\theta}}_2/(2{\widehat{\theta}}_3^2),$$ and hence taking $c$ to be $(0,-1/(2{\widehat{\theta}}_3),{\widehat{\theta}}_2/(2{\widehat{\theta}}_3^2))'$, or equivalently, $$\label{cforeta} c=(0,-{\widehat{\theta}}_3,{\widehat{\theta}}_2)'.$$ @Chaloner89 considered a Bayesian approach to this problem in the linear model. Although the method described above in Section \[sec:result\] is for the linear model described in Section \[sec:set-up\], it can be applied to find locally $c$-optimal designs in the nonlinear model , and other generalized linear models, without further extension by applying Ford et al.’s [-@Ford92] results which show that, in this case, the locally $c$-optimal design can be found by applying Elfving’s method to a transformed version of the problem as if it were a linear model, described next. For a more general description of their method, which covers other models and optimality criteria, we refer readers to @Ford92. In the above notation, $x=x(u)=(1,u,u^2)'$ and ${\mathcal{X}}=\{x(u): u\in[-1,1]\}$. Given ${\widehat{\theta}}$, let $B=B({\widehat{\theta}})=||{\widehat{\theta}}||U$ where $U=U({\widehat{\theta}})$ is a $3\times 3$ orthonormal matrix with third row equal to ${\widehat{\theta}}'/||{\widehat{\theta}}||$. Then the locally $c$-optimal design can be found by finding the $(Bc)$-optimal design in the transformed design space $$\label{G} \mathcal{G}=\left\{w(z_3) z: z=Bx, x\in{\mathcal{X}}\right\}{\quad\mbox{where}\quad} w(\zeta)=e^{\zeta/2}/(1+e^\zeta),$$ and then transforming back to ${\mathcal{X}}$ via $B^{-1}$. We give an example of this transformation and calculation. Letting ${\widehat{\theta}}=(2,-6,-9)'$, a $B$ satisfying the above assumptions is $$\label{B} B=\left[ \begin{array}{ccc} 10.816 &1.110&1.664 \\ 0&9.153&-6.102\\ 2&-6&-9 \end{array} \right],$$ to 3 decimal places. Figure \[fig:X&Xtild\] shows ${\mathcal{X}}$, $\mathcal{G}$ for this $B$, and the reflection $\mathcal{G}^-$ of $\mathcal{G}$ through the origin, and Figure \[fig:convw/c\] shows $\mbox{conv}(\mathcal{G}\cup \mathcal{G}^-)$. For the sake of illustration we choose $c=(-.195,.1,-.243)'$, although this does not have first component equal to zero as in , and this gives $Bc=(-2.403,2.398,1.197)'$, the red arrow in Figure \[fig:convw/c\]. The method is then used to quickly compute the $(Bc)$-optimal design on $\mathcal{G}$ and transforming back to ${\mathcal{X}}$ gives, parametrizing by $u$, $$\xi^*=\{(-1,.135), (.181,.194), (.452,.671)\}.$$ From Figure \[fig:convw/c\] it is clear that the $(Bc)$-optimal design on $\mathcal{G}$ will be supported on the point $\overline{g}$ at the end of the red curve in Figure \[fig:X&Xtild\] with positive second coordinate, plus two interior points on the curve. Noting that $\overline{g}$ corresponds to the $u=-1$ end of the curve ${\mathcal{X}}$ under the transformation ${\mathcal{X}}{\rightarrow}\mathcal{G}$, it is not surprising then that $\xi^*$ is supported on $u=-1$ plus two interior points in $(-1,1)$. Discussion ========== We have given a new characterization of Elfving’s method which affords more efficient computation, particularly in high dimensions where visualization of the Elfving set $\mathcal{E}$ is difficult. Dette’s [-@Dette93] extension of Elving’s method to the popular $D$-optimality criterion holds the possibility extending the current computational method to find $D$-optimal designs. The approach of @Ford92 to generalized linear models applied in Section \[sec:logistic\] would apply there as well since the $D$-optimality criterion is also $B$-invariant. Another tantalizing area of extension is constrained $c$-optimal designs [@Cook94; @Cook95]. With linear constraints on the design measure, as in @Cook95 [Section 1.3], it can be shown that the constrained problem has similar ingredients to the standard $c$-optimal design problem. For example, the subset of points in $\mathcal{E}$ which satisfy the constraint is convex and symmetric about the origin, so one may suspect that, like the standard problem, the point where the ray $R(c)$ pierces this subset is a natural candidate for the constrained $c$-optimal design. However, difficulties remain in this approach, such as showing that an Elfving-like characterization like holds for the constrained problem. @Stigler71 has pointed out similar difficulties for the constrained $D$-optimal problem. Acknowledgements {#acknowledgements .unnumbered} ================ The author thanks two reviewers for their helpful comments. This work was partially supported by grants from the National Science Foundation (Number DMS-0907241) and the National Security Agency (Number H98230-11-1-0162). Appendix: Properties of the design and proof of Theorem \[thm:phi\] {#appendix-properties-of-the-design-and-proof-of-theoremthmphi .unnumbered} =================================================================== Before giving the proof of Theorem \[thm:phi\], we state and prove Proposition \[prop:xi\*\] which gives the properties of the design $\xi^*$, given by , and the “Elfving point” $z^*$, given by , needed to prove Theorem \[thm:phi\]. We state this as a separate proposition because it may be of interest on its own. In particular, Claim \[Psimin\] of Proposition \[prop:xi\*\] shows that the weights of the design $\xi^*(x_1,\ldots,x_\ell)$ are optimal for the given support set $x_1,\ldots,x_\ell$, showing that the formulae - can be used to compute sub-optimal designs with prescribed support. \[prop:xi\*\] Let $x_1,\ldots,x_\ell\in\mathcal{X}$, $\ell\le k$, be LI such that $c\in\mbox{span}(x_1,\ldots,x_\ell)$. Then the following hold. 1. \[invariant\] $\xi^*(x_1,\ldots,x_\ell)$ and $z^*(x_1,\ldots,x_\ell)$ are invariant with respect to the choice of generalized inverse in - 2. $z^*(x_1,\ldots,x_\ell)\in R(c)$\[zinR\] 3. $\xi^*(x_1,\ldots,x_\ell)\in \Xi$\[xiinXi\] 4. $\Psi(\xi^*(x_1,\ldots,x_\ell))=||c||^2/||z^*(x_1,\ldots,x_\ell) ||^2$\[Psi=c/z\^2\] 5. $\xi^*(x_1,\ldots,x_\ell)=\arg\min_{\xi\in\Xi(x_1,\ldots,x_\ell)} \Psi(\xi)$ \[Psimin\] **Proof.** For any design $\xi=\{(x_i,p_i)\}_{i=1}^\ell\in\Xi$, we have that $c=M(\xi)d$ for some $d$, hence $$\begin{gathered} \label{c=Md} c=M(\xi)d=M(\xi)M(\xi)^{-1}M(\xi)d=M(\xi)M(\xi)^{-1}c=\left(\sum_{i=1}^\ell p_ix_ix_i'\right)M(\xi)^{-1}c\\ =\sum_{i=1}^\ell p_i(x_i'M(\xi)^{-1}c)x_i.\end{gathered}$$ Letting $M$ be as in , we have that $c\in \mbox{span}(x_1,\ldots,x_\ell)\subseteq\mbox{colspace}(M)$ by Lemma \[lem:span=C\], which is further below in this appendix. Let $\gamma=\ell/\sum_{j=1}^\ell{\left\vertx_j'M^{-1}c\right\vert}$ and denote $z^*(x_1,\ldots,x_\ell)$ simply by $z^*$. Applying with $\xi=\{(x_i,1/\ell)\}_{i=1}^\ell$ gives $$\label{c=M*d} c=\ell^{-1}\sum_{i=1}^\ell(x_i'M^{-1}c)x_i=\gamma^{-1}\sum_{i=1}^\ell {\varepsilon}_i^*p_i^*x_i=\gamma^{-1}z^*.$$ This gives Claim \[zinR\], and also Claim \[xiinXi\], as follows. By relabeling if necessary, let $x_1,\ldots,x_m$, $m\le \ell$, be the $x_1,\ldots,x_\ell$ corresponding to nonzero $p_i^*$’s. Then shows that $c\in{\mbox{span}}(x_1,\ldots,x_m)\subseteq{\mbox{colspace}}(M(\xi))$ by Lemma \[lem:span=C\]. Denote $\xi^*(x_i,\ldots,x_\ell)=\{(x_i,p_i^*)\}_{i=1}^m$ simply by $\xi^*$. For Claim \[Psi=c/z\^2\], apply with $\xi=\xi^*$ to get $$c=\sum_{i=1}^mp_i^*(x_i'M(\xi^*)^{-1}c)x_i,$$ and subtracting this from the second-to-last expression in (with $m$ in place of $\ell$) gives $$\sum_{i=1}^m\left({\varepsilon}_i^*/\gamma-x_i'M(\xi^*)^{-1}c\right)p_i^*x_i=0.$$ Since the $x_i$ are LI and all the $p_i^*$ in the sum are positive, this gives $$x_i'M(\xi^*)^{-1}c={\varepsilon}_i^*/\gamma\quad\mbox{for all $i=1,\ldots,m$.}$$ Then $$\begin{gathered} \Psi(\xi^*)=c'M(\xi^*)^{-1}c=\left(\gamma^{-1}\sum_{i=1}^m {\varepsilon}_i^* p_i^*x_i\right)'M(\xi^*)^{-1}c= \gamma^{-1}\sum_{i=1}^m {\varepsilon}_i^*p_i^*(x_i'M(\xi^*)^{-1}c) \\= \gamma^{-1}\sum_{i=1}^m {\varepsilon}_i^*p_i^*({\varepsilon}_i^*/\gamma) =\gamma^{-2}=||c||^2/||z^*||^2,\end{gathered}$$ this last by , proving Claim \[Psi=c/z\^2\]. To show Claim \[Psimin\], subtract from the second-to-last expression in to get $$\sum_{i=1}^\ell\left[{\varepsilon}_i^*p_i^*/\gamma-p_i(x_i'M(\xi)^{-1}c)\right]x_i=0$$ which, by linear independence, gives $$\label{qipi} p_i(x_i'M(\xi)^{-1}c)={\varepsilon}_i^*p_i^*/\gamma\quad\mbox{for all $i=1,\ldots,\ell$.}$$ Then, using and , $$\begin{gathered} \Psi(\xi)=c'M(\xi)^{-1}c=\left[\sum_{i=1}^\ell p_i(x_i'M(\xi)^{-1}c)x_i\right]'M(\xi)^{-1}c=\sum_{i=1}^\ell p_i(x_i'M(\xi)^{-1}c)^2\\ =\sum_{i:\; p_i>0} p_i(x_i'M(\xi)^{-1}c)^2 =\gamma^{-2}\sum_{i:\; p_i>0} (p_i^*)^2/p_i.\label{Psi=p/q}\end{gathered}$$ Then the method of Lagrange multipliers easily shows that $p_i=p_i^*$ for all $i$ minimizes subject to $\sum_i p_i=1$, proving Claim \[Psimin\]. Finally, to prove Claim \[invariant\], suppose a different generalized inverse is used for $M^{-1}$, resulting in ${\widetilde{\xi}}, {\widetilde{{\varepsilon}}}_i, {\widetilde{p}}_i, {\widetilde{\gamma}}$, ${\widetilde{z}}$ instead of $\xi^*,{\varepsilon}_i^*,p_i^*,\gamma,z^*$. By the same argument leading to we would have $$\label{qipitilde} {\widetilde{{\varepsilon}}}_i{\widetilde{p}}_i/{\widetilde{\gamma}}={\varepsilon}_i^*p_i^*/\gamma\quad\mbox{for all $i=1,\ldots,\ell$.}$$ Also, using and Claims \[Psi=c/z\^2\] and \[Psimin\], $$\gamma=\frac{||z^*||}{||c||}=\left[\Psi(\xi^*)\right]^{-1/2}= \left[\Psi(\xi)\right]^{-1/2}=\frac{||{\widetilde{z}}||}{||c||}={\widetilde{\gamma}},$$ hence shows that ${\varepsilon}_i^*p_i^*={\widetilde{{\varepsilon}}}_i{\widetilde{p}}_i$ for all $i$, and it follows that $\xi^*={\widetilde{\xi}}$ and $z^*={\widetilde{z}}$. **Proof of Theorem \[thm:phi\].** Denote $\xi^*(x_1^*,\ldots,x_k^*)$ simply by $\xi^*$. Let $\xi=\{(x_i,p_i)\}_{i=1}^\ell$, $p_i>0$ and $\ell\le k$, be the design given by Elfving’s method. If $x_1,\ldots,x_\ell$ are LI then they can be completed to a LI set $x_1,\ldots,x_k\in\mathcal{X}$ by our assumption in the first paragraph of Section \[sec:set-up\]. Then $\xi\in\Xi(x_1,\ldots,x_k)$ and so $\Psi(\xi^*)\le \Psi(\xi)$ by Claim \[Psimin\] of Proposition \[prop:xi\*\]. Otherwise, assume that $x_1,\ldots,x_\ell$ are linearly dependent (LD). In this case we will construct a design ${\widetilde{\xi}}$ whose support is a strict subset of $x_1,\ldots,x_\ell$ and $\Psi({\widetilde{\xi}})\le\Psi(\xi)$. This argument can be repeated until only a LI support set remains, and the previous argument can be applied. By , $\Psi(\xi)=||c||^2/||\sum_{i=1}^\ell {\varepsilon}_ip_ix_i||^2$ for some choice of signs ${\varepsilon}_i$, and let $z=\sum_{i=1}^\ell {\varepsilon}_ip_ix_i$. Since $x_1,\ldots,x_\ell$ are LD, ${\varepsilon}_1x_1,\ldots,{\varepsilon}_\ell x_\ell$ are also LD so let $\alpha_1,\ldots,\alpha_\ell$ be constants, not all $0$, such that $\sum_{i=1}^\ell\alpha_i {\varepsilon}_ix_i=0$. Without loss of generality assume that $\sum_{i=1}^\ell\alpha_i\ge 0$, since otherwise each $\alpha_i$ could be replaced by $-\alpha_i$. Now set $x_0=0$, ${\varepsilon}_0=1$, $p_0=0$, and $\alpha_0=-\sum_{i=1}^\ell\alpha_i\le 0$. With this, we have that $\alpha_0,\ldots,\alpha_\ell$ are not all $0$ and satisfy $$\sum_{i=0}^\ell\alpha_i{\varepsilon}_ix_i= 0{\quad\mbox{and}\quad} \sum_{i=0}^\ell\alpha_i=0.$$ (In other words, ${\varepsilon}_0x_0,\ldots,{\varepsilon}_\ell x_\ell$ are affinely dependent.) Now set $$\delta=\min\{p_i/\alpha_i: i=0,\ldots,\ell,\;\alpha_i>0\}$$ and $q_i=p_i-\delta\alpha_i$, $i=0,\ldots,\ell$, and note that $q_i\ge 0$ for all $i$ and $$\sum_{i=0}^\ell q_i=\sum_{i=0}^\ell p_i-\delta\sum_{i=0}^\ell \alpha_i=1.$$ Note also that $q_i=0$ for some $0<i\le\ell$ by definition of $\delta$ and by virtue of the fact that $\alpha_0\le 0$. By relabeling, assume that $q_\ell=0$. We have $$\label{z=sumq} z=z-0=\sum_{i=0}^\ell p_i{\varepsilon}_i x_i- \delta\sum_{i=0}^\ell \alpha_i{\varepsilon}_ix_i=\sum_{i=0}^\ell q_i{\varepsilon}_ix_i =\sum_{i=0}^{\ell-1} q_i{\varepsilon}_ix_i.$$ It cannot be that $q_0=1$ since this would imply that $z=0$ by and it would follow that $c=0$, contradicting our assumption, hence $q_0<1$. Let ${\mathcal{X}}_0$ be any compact set containing ${\mathcal{X}}\cup\{0\}$, $${\widehat{\xi}}=\{(x_i,q_i)\}_{i=0}^{\ell-1},{\quad\mbox{and}\quad}{\widetilde{\xi}}=\{(x_i,q_i/(1-q_0))\}_{i=1}^{\ell-1}.$$ The former is a design on ${\mathcal{X}}_0$ by virtue of its inclusion of $x_0=0$ in its support. Then implies that $\Psi({\widehat{\xi}})\le\Psi(\xi)$. Also, $$M({\widetilde{\xi}})=(1-q_0)^{-1}\sum_{i=1}^{\ell-1}q_ix_ix_i'= (1-q_0)^{-1}\sum_{i=0}^{\ell-1}q_ix_ix_i'=(1-q_0)^{-1}M({\widehat{\xi}})$$ so that $$\Psi({\widetilde{\xi}})=c'M({\widetilde{\xi}})^{-1}c=(1-q_0)\Psi({\widehat{\xi}})\le \Psi({\widehat{\xi}})\le \Psi(\xi),$$ thus completing the desired reduction to the smaller support set $x_1,\ldots, x_{\ell-1}$. \[lem:span=C\] Let $x_1,\ldots,x_\ell\in\mathcal{X}$. Then $$\mbox{colspace}\left(M(\{(x_i,q_i)\}_{i=1}^\ell) \right) \subseteq\mbox{span}(x_1,\ldots,x_\ell)\;\mbox{for any $q_i\ge 0$, $\sum_{i=1}^\ell q_i=1$.}$$ Conversely, $$\mbox{span}(x_1,\ldots,x_\ell) \subseteq \mbox{colspace}\left(M(\{(x_i,q_i)\}_{i=1}^\ell) \right)\;\mbox{for any $q_i> 0$, $\sum_{i=1}^\ell q_i=1$.}$$ **Proof.** Let $M= M(\{(x_i,q_i)\}_{i=1}^\ell)$. For any $a\in\mathbb{R}^\ell$, $$Ma=\left(\sum_{i=1}^\ell q_ix_ix_i'\right)a=\sum_{i=1}^\ell \left(q_ix_i' a\right)x_i\in \mbox{span}(x_1,\ldots,x_\ell).$$ Conversely, given constants $a_1,\ldots,a_\ell$ and strictly positive $q_1,\ldots,q_\ell$, we will show that $\sum_{i=1}^\ell a_ix_i \in \mbox{colspace}(M)$. Let $b=(a_1/\sqrt{q_1},\ldots,a_\ell/\sqrt{q_\ell})'\in\mathbb{R}^\ell$. Let $X$ be the $\ell\times k$ matrix with rows $\sqrt{q_1}x_1',\ldots, \sqrt{q_\ell}x_{\ell}'$ and note that $X'X=\sum_{i=1}^\ell q_ix_ix_i'= M$. Let $v$ be the projection of $b$ onto $\mbox{colspace}(X)$ and let $d$ be any vector such that $Xd=v$. Then $(b-v)\perp\mbox{colspace}(X)$, or $X'b=X'v$, so we have $$\sum_{i=1}^\ell a_ix_i=X'b=X'v=X'Xd=M d\in \mbox{colspace}(M).$$ Bartroff, J. and Lai, T. L. (2010). Approximate dynamic programming and its applications to the design of phase [I]{} cancer trials. , 25:245–257. Bartroff, J. and Lai, T. L. (2011). Incorporating individual and collective ethics into phase [I]{} cancer trial designs. , 67:596–603. Chaloner, K. (1989). Bayesian design for estimating the turning point of a quadratic regression. , 18:1385–1400. Chernoff, H. (1972). . Society for Industrial and Applied Mathematics, Philadelphia. Chernoff, H. (1999). Gustav [E]{}lfving’s impact on experimental design. , 14(2):201–205. Cook, D. and Fedorov, V. (1995). Constrained optimization of experimental design. , 26:129–148. Cook, R. D. and Wong, W. K. (1994). On the equivalence of constrained and compound optimal designs. , 89:687–692. Dette, H. (1993a). Bayesian ${D}$-optimal and model robust designs in linear regression models. , 25:27–46. Dette, H. (1993b). Elfving’s theorem for ${D}$-optimality. , 21:753–766. Dette, H. (1993c). A new interpretation of optimality for ${E}$-optimal designs in linear regression models. , 40:37–50. Dette, H. (1996). A note on [B]{}ayesian $c$- and ${D}$-optimal designs in nonlinear regression models. , 24(3):1225–1234. Dette, H. (1997). Designing experiments with respect to ‘standardized’ optimality criteria. , 59:97–110. Dette, H., Bretz, F., Pepelyshev, A., and Pinheiro, J. (2008). Optimal designs for dose-finding studies. , 103(483):1225–1237. Dette, H. and Holland-Letz, T. (2009). A geometric characterization of $c$-optimal designs for heteroscedastic regression. , 37(6B):4088–4103. Dette, H., Kiss, C., Bevanda, M., and Bretz, F. (2010). Optimal designs for the [E]{}max, log-linear and exponential models. , 97(2):513–518. Dette, H. and Studden, W. J. (1993). Geometry of ${E}$-optimality. , 21:416–433. Elfving, G. (1952). Optimum allocation in linear regression theory. , 23:255–262. Fan, S. K. and Chaloner, K. (2003). A geometric method for singular c-optimal designs. , 113(1):249–257. Ford, I., Torsney, B., and Wu, C. F. J. (1992). The use of a canonical form in the construction of locally optimal designs for non-linear problems. , 54:569–583. Kitsos, C. P., Titterington, D. M., and Torsney, B. (1988). An optimal design problem in rhythmometry. , 44:657–671. López-Fidalgo, J. and Rodríguez-Díaz, J. (2004). lfving’s method for $m$-dimensional models. , 59:235–244. Pukelsheim, F. and Torsney, B. (1991). Optimal weights for experimental designs on linearly independent support points. , 19:1614–1625. Seber, G. A. F. and Lee, A. J. (2003). . Wiley Series in Probability and Statistics. Wiley-Interscience \[John Wiley & Sons\], Hoboken, NJ, second edition. Stigler, S. M. (1971). Optimal experimental design for polynomial regression. , 66:311–318. Studden, W. J. (1968). Optimal designs on [T]{}chebyscheff points. , 39:1435–1447. Studden, W. J. (2005). Elfving’s theorem revisited. , 130(1–2):85–94. Thall, P. F. and Cook, J. D. (2004). Dose-finding based on efficacy-toxicity trade-offs. , 60(3):684–693. Torsney, B. (1981). . PhD thesis, University of Glasgow. $k$ $j:\; c=e_j$ $\xi^*(0)$ $\Psi(\xi^*)$ ----- -------------- ------------ -------------- --------------- -------------- -------------- ----------- -------- 1 1 1 2 0 (.309, .419) (.809, .061) (1, 1/50) 25 3 3/8 (.707, 1/4) (1, 1/16) 64 6 4 0 (.309, .265) (.809, .175) (1, .06) 400 5 1/4 (.707, 1/4) (1, 1/8) 64 6 0 (.309, 1/5) (.809, 1/5) (1, 1/10) 256 1 1 1 2 0 (.309, .419) (.809, .061) (1, 1/50) 25 3 .352 (.5, 2/9) (.866, .074) (1, .028) 324 7 4 0 (.309, .265) (.809, .175) (1, .06) 400 5 2/9 (.5, .194) (.866, .175) (1, .06) 2304 6 0 (.309, 1/5) (.809, 1/5) (1, 1/10) 256 7 1/6 (.5, 1/6) (.866, 1/6) (1, 1/12) 1024 1 1 1 2 0 (.223, .412) (.623, .052) (.901, .025) (1, .010) 49 3 .352 (.5, 2/9) (.866, .074) (1, .028) 324 4 0 (.223, .248) (.623, .147) (.901, .075) (1, .031) 3136 8 5 2/9 (.5, .194) (.866, .139) (1, 1/18) 2304 6 0 (.223, .180) (.623, .152) (.901, .117) (1, .051) 12544 7 1/6 (.5, 1/6) (.866, 1/6) (1, 1/12) 1024 8 0 (.223, .143) (.623, .143) (.901, .143) (1, .071) 4096 1 1 1 2 0 (.223, .412) (.623, .052) (.901, .025) (1, .010) 49 3 .344 (.383, .213) (.707, .063) (.924, .037) (1, .016) 1024 4 0 (.223, .248) (.623, .147) (.901, .075) (1, .031) 3136 9 5 .213 (.383, .178) (.707, .113) (.924, .072) (1, .031) 25600 6 0 (.223, .180) (.623, .152) (.901, .117) (1, .051) 12544 7 .156 (.383, .147) (.707, 1/8) (.924, .103) (1, .047) 65536 8 0 (.223, .143) (.623, .143) (.901, .143) (1, .071) 4096 9 1/8 (.383, 1/8) (.707, 1/8) (.924, 1/8) (1, 1/16) 16384 1 1 1 2 0 (.174, .409) (.5, .049) (.766, .021) (.940, .014) (1, .006) 81 3 .344 (.383, .213) (.707, .063) (.924, .037) (1, .016) 1024 4 0 (.174, .241) (.5, .137) (.766, .062) (.940, .042) (1, .019) 14400 10 5 .213 (.383, .178) (.707, .113) (.924, .072) (1, .031) 25600 6 0 (.174, .172) (.5, .136) (.766, .093) (.940, .068) (1, .031) 186624 7 .156 (.383, .147) (.707, 1/8) (.924, .103) (1, .047) 65536 8 0 (.174, .134) (.5, .123) (.766, .107) (.940, .092) (1, .043) 331776 9 1/8 (.383, 1/8) (.707, 1/8) (.924, 1/8) (1, 1/16) 16384 10 0 (.174, 1/9) (.5, 1/9) (.766, 1/9) (.940, 1/9) (1, 1/18) 65536 : $c$-optimal designs $\xi^*$ for polynomial regression model $Y\sim\mathcal{N}(\sum_{i=1}^{k}\theta_iu^{i-1},\sigma^2)$ and $c$ equal to standard basis vectors $e_1,\ldots,e_k\in{\mathbb{R}}^k$ for $k=6,\ldots,10$. Parametrizing $\xi^*$ by $u\in[-,1,1]$, column 3 is $\xi^*(0)$ and the pairs $(u,p)$ in columns 4–8 are the positive $u$ values such that $\xi^*(u)=\xi^*(-u)=p$. The support points and weights are given to 3 decimal places. \[table:poly\] [^1]: We have verified the values in @Lopez-Fidalgo04 [Table 5] using our method, except for the value of $\Psi(\xi^*)$ in the case $c=e_2$, $k=3$ ($d=2$ in their notation), which we calculate to be 1 rather than the value 4 given there.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We propose a generative latent variable model for unsupervised scene decomposition. Our model, SPACE, provides a unified probabilistic modeling framework to combine the best of previous models. SPACE can explicitly provide factorized object representation per foreground object while also decomposing background segments of complex morphology. Previous models are good at either of these, but not both. With the proposed parallel-spatial attention, SPACE resolves the scalability problem of previous methods and thus makes the model applicable to scenes with a much larger number of objects without performance degradation. Besides, the foreground/background distinction of SPACE is more effective and intuitive than other methods because unlike other methods SPACE can detect static objects that have been difficult to detect as foreground. In experiments on Atari and 3D-Rooms, we show that SPACE achieves the above properties consistently in all experiments in comparison to SPAIR, IODINE, and GENESIS.' author: - | Sungjin Ahn$^1$, {Zhixuan Lin$^2$[^1] , Weihao Sun$^1$, Skand Vishnawath Peri$^1$}, Gautam Singh$^1$,\ **Yi-Fu Wu$^1$, Fei Deng$^1$, Jindong Jiang$^1$**\ \ $^1$Rutgers University\ $^2$Zhejiang University bibliography: - 'refs.bib' title: Spatially Parallel Attention and Component Extraction for Scene Decomposition --- Introduction ============ One of the unsolved key challenges in machine learning is unsupervised learning of structured representation for a visual scene containing many objects with occlusion, partial observability, and complex background. When properly decomposed into meaningful symbolic entities such as objects and spaces, this structured representation brings many advantages of symbolic representation that contemporary deep learning with continuous vector representation has not been successful. The advantages include sample efficiency of downstream tasks such as a deep reinforcement learning agent [@Mnih2013PlayingAW], ability of visual variable binding [@sun1992variable] for reasoning and causal inference over the relationship between the objects and agents in a scene, and compositionality and transferability for generalization. Recently, several methods have been proposed for this problem, [unsupervised object-oriented scene decomposition]{}, and these can be categorized into two approaches: *mixture-scene* models and *spatial-attention* models. In the mixture-scene models [@nem; @iodine; @monet; @genesis], a scene is explained by a mixture of $K$ component images. The main benefit is to provide flexible segmentation maps for objects and background segments with complex morphology for which spatial attention with bounding boxes may have difficulties. However, this approach has limitations in scalability to scenes with many objects, and thus currently has been applied to scenes with only less than ten components. This is because, to obtain a jointly complete scene, a component needs to refer to other components, and thus inference is inherently performed sequentially either by an RNN generating a component per step or by iterative inference [@marino2018iterative]. Besides, because each component corresponds to a full-scale image, important physical features of objects like position and scale are only implicitly encoded in the scale of a full image, and thus further disentanglement is required to extract these useful features. The spatial-attention models [@air; @spair] show the opposite properties to mixture-scene models. It detects objects using spatial attention and thus can obtain not only the appearance representation but also fully disentangled explicit geometric representation of an object such as position and scale. Such features grounded on the semantics of physics should be useful in many ways (e.g., sample efficiency, interpretability, geometric reasoning and inference, and transferability). However, representing areas such as complex background segments that have too flexible morphology to be captured by spatial attention (e.g., based on rectangular bounding boxes) is the main limitation of this approach. It also shows scalability issues as objects are processed sequentially. In this paper, we propose a method, called *Spatially Parallel Attention and Component Extraction* (SPACE), that combines the best of both approaches. learns to process foreground objects, which can be captured efficiently by bounding boxes, by using parallel spatial-attention while decomposing the remaining area that includes both morphologically complex objects and background segments by the component mixture. Thus, provides object-wise disentangled representation of foreground objects along with explicit properties like *where* per object while also providing decomposed representations of complex background components. Besides, we resolve the scalability issue of existing spatial attention methods by developing fully parallel foreground-object processing. In experiments on 3D-room scenes and Atari game scenes, we quantitatively and qualitatively compare the representation of to other models and show that combines the benefits of both approaches in addition to significant speed-ups due to the parallel foreground processing. The contributions of the paper are as follows. First, we introduce a model that unifies the benefits of two existing approaches in a principled framework of probabilistic latent variable modeling. Second, we introduce a spatially parallel multi-object processing module and demonstrate that it can significantly mitigate the scalability problem of previous methods. Third, we demonstrate SPACE can detect static objects (e.g., the key in Montezuma’s revenge) that always appear in the same position in all training images and thus are typically treated as a part of the background in other models. This problem has been a key challenge in unsupervised object detection. Lastly, we provide extensive analysis on and comparisons to previous models. The Proposed Model: ==================== In this section, we describe our proposed model, Spatially Parallel Attention and Component Extraction (SPACE). The main idea of SPACE is to propose a unified probabilistic generative model that combines the benefits of the spatial-attention models and mixture-scene models. Generative Process ------------------ SPACE assumes that a scene $\rvx$ is decomposed into two latents: foreground $\rvz^{\text{fg}}$ and background $\rvz^{\text{bg}}$. The foreground is further decomposed into a set of independent foreground objects $\rvz^{\text{fg}}= \{\rvz_i^{\text{fg}}\}$ and the background is also decomposed further into a sequence of background segments $\rvz^{\text{bg}}= \rvz^{\text{bg}}_{1:K}$. The foreground is first generated and then the generation of the background is conditioned on the foreground. The image distributions of the foreground objects and the background components are combined together into the following pixel-wise mixture model to produce the complete image distribution. Here, the foreground mixing probability $\alpha$ is computed as $\alpha = f_\alpha(\rvz^{\text{fg}})$. This way, the foreground is given precedence in assigning its own mixing weight and the remaining is apportioned to the background. The mixing weight assigned to the background is further sub-divided among the $K$ background components. These weights are computed as $\pi_k=f_{\pi_k}(\rvz^{\text{bg}}_{1:k})$ and $\sum_{k} \pi_k =1$. With these notations, the complete generative model can be described as follows. We now describe the foreground and background models in more detail. **Foreground.** implements $\rvz^{\text{fg}}$ as a structured latent. In this structure, an image is treated as if it were divided into $H\times W$ cells and each cell is tasked with modeling at most one (nearby) object in the scene. This type of structuring has been used in [@yolo; @rn; @spair]. Similarly to SPAIR, in order to model an object, each cell $i$ is associated with a set of latents $(\rvz^{\text{pres}}_i, \rvz^{\text{where}}_i, \rvz^{\text{depth}}_i, \rvz^{\text{what}}_i)$. In this notation, $\rvz^{\text{pres}}$ is a binary random variable denoting if the cell models any object or not, $\rvz^{\text{where}}$ denotes the size of the object and its location relative to the cell, $\rvz^{\text{depth}}$ denotes the *depth* of the object to resolve occlusions and $\rvz^{\text{what}}$ models the object appearance and its mask. These latents may then be used to compute the foreground image component $p(\rvx|\rvz^{\text{fg}})$ which is modeled as a Gaussian distribution $\mathcal{N}(\mu^{\text{fg}}, \sigma^2_{\text{fg}})$. In practice, we treat $\sigma^2_{\text{fg}}$ as a hyperparameter and decode only the mean image $\mu^{\text{fg}}$. In this process, reconstructs the objects associated to each cell having $\rvz^{\text{pres}}=1$. For each such cell, the model uses the $\rvz_i^{\text{what}}$ to decode the object glimpse and its mask and the glimpse is then positioned on a full-resolution canvas using $\rvz^{\text{where}}_i$ via the Spatial Transformer Network [@spatial_transformer]. Using the object masks and $\rvz^{\text{depth}}_i$, all the foreground objects are combined into a single foreground mean-image $\mu^{\text{fg}}$ and the foreground mask $\alpha$ (See Appendix \[ax:implementation\] for more details). imposes a prior distribution on these latents as follows: Here, only $\rvz^{\text{pres}}_i$ is modeled using a Bernoulli distribution while the remaining are modeled as Gaussian. **Background.** To model the background, implements $\rvz_k^{\text{bg}}$, similar to GENESIS, as $(\rvz_k^m, \rvz_k^c)$ where $\rvz_k^m$ models the mixing probabilities $\pi_k$ of the components and $\rvz^c_k$ models the RGB distribution $p(\rvx|\rvz^{\text{bg}}_k)$ of the $k^\text{th}$ background component as a Gaussian $\mathcal{N}(\mu_i^{\text{bg}}, \sigma^2_{\text{bg}})$. The following prior is imposed upon these latents. Inference and Training ---------------------- Since we cannot analytically evaluate the integrals in  due to the continuous latents $\rvz^{\text{fg}}$ and $\rvz^{\text{bg}}_{1:K}$, we train the model using a variational approximation. The true posterior on these variables is approximated as follows. This is used to derive the following ELBO to train the model using reparameterization trick and SGD [@vae]. () &= \_[q(\^, \^\_[1:K]{}|)]{} - (q(\^|)p(\^))\ &-\_[q(\^|)]{} \_[k=1]{}\^K \_[q(\^\_[&lt;k]{}|, \^)]{} (q(\^\_k|\^\_[&lt;k]{}, \^, )p(\^\_k|\^\_[&lt;k]{}, \^)) See Appendix \[ax:derivations\] for the derivation of the ELBO and the related details. **Parallel Inference of Cell Latents.** One of the attractive features of is the mean-field modeling of the approximate inference where $\rvz^{\text{fg}}_i = (\rvz^{\text{pres}}_i, \rvz^{\text{where}}_i, \rvz^{\text{depth}}_i, \rvz^{\text{what}}_i)$ for each cell do not depend on other cells.\[sec:parallel\] On the contrary, SPAIR’s inference for each cell’s latents auto-regressively depend on some or all of the previously traversed cells in a row-major order i.e., $q(\rvz^{\text{fg}}|\rvx) = \prod_{i=1}^{HW} q(\rvz^{\text{fg}}_i|\rvz^{\text{fg}}_{<i},\rvx)$. Although, in principle, it appears attractive to model such dependence between objects, it becomes prohibitively expensive in practice as the number of objects increases. On the other hand, ’s inference can be executed in parallel for all cells resulting in significant gains in training speeds. Consequently, SPACE can efficiently deal with scenes with a much larger number of objects. Our experiments demonstrate that this can be done without any adverse effects on the modeling performance. **Preventing Box-Splitting.** If the prior for the bounding box size is set to be too small, then the model could split a large object by multiple bounding boxes and when the size prior is too large, the model may not capture small objects in the scene, hence, causing a trade-off between the prior values of the bounding box size. To alleviate this problem, we found it helpful to introduce an auxiliary loss which we call the *boundary loss*. In the boundary loss, we construct a boundary of thickness $b$ pixels along the borders of each glimpse. Then, we restrict an object to be inside this boundary and penalize the model if an object’s mask overlaps with the boundary area. Thus, the model is penalized if it tries to split a large object by multiple smaller bounding boxes. A detailed implementation of the boundary loss is mentioned in Appendix \[ax:aux\_loss\]. Related Works ============= Our proposed model is inspired by several recent works in unsupervised object-oriented scene decomposition. The Attend-Infer-Repeat (AIR) [@air] framework uses a recurrent neural network to attend to different objects in a scene and each object is sequentially processed one at a time. An object-oriented latent representation is prescribed that consists of ‘what’, ‘where’, and ‘presence’ variables. The ‘what’ variable stores the appearance information of the object, the ‘where’ variable represents the location of the object in the image, and the ‘presence’ variable controls how many steps the recurrent network runs and acts as an interruption variable when the model decides that all objects have been processed. Since the number of steps AIR runs scales with the number of objects it attends to, it does not scale well to images with many objects. Spatially Invariant Attend, Infer, Repeat (SPAIR) [@spair] attempts to address this issue by replacing the recurrent network with a convolutional network. Similar to YOLO [@yolo], the locations of objects are specified relative to local grid cells rather than the entire image, which allow for spatially invariant computations. In the encoder network, a convolutional neural network is first used to map the image to a feature volume with dimensions equal to a pre-specified grid size. Then, each cell of the grid is processed *sequentially* to produce objects. This is done sequentially because the processing of each cell takes as input feature vectors and sampled objects of nearby cells that have already been processed. SPAIR therefore scales with the pre-defined grid size which also represents the maximum number of objects that can be detected. Our model uses an approach similar to SPAIR to detect foreground objects, but importantly we make the foreground object processing fully parallel to scale to large number of objects without performance degradation. For unsupervised mixture-scene models, several recent models have shown promising results. MONet [@monet] leverages a deterministic recurrent attention network that outputs pixel-wise masks for the scene components. A variational autoencoder (VAE) [@vae] is then used to model each component. IODINE [@iodine] approaches the problem from a spatial mixture model perspective and uses amortized iterative refinement of latent object representations within the variational framework. GENESIS [@genesis] also uses a spatial mixture model which is encoded by component-wise latent variables. Relationships between these components are captured with an autoregressive prior, allowing complete images to be modeled by a collection of components. Evaluation ========== ![Qualitative demonstration of trained on the 3D-room dataset.[]{data-label="fig:eval:3d-space"}](images/3d_space.png){width="0.9\linewidth"} ![Qualitative demonstration of trained jointly on a selection of 10 ATARI games. We show 6 games with complex background here.[]{data-label="fig:eval:atari-space"}](images/atari_space.png){width="1.0\linewidth"} [0.9]{} ![Qualitative comparison between , SPAIR, IODINE and GENESIS for the 3D-Room dataset.[]{data-label="fig:eval:qualitative-3d"}](images/3dlarge_compare.png "fig:"){width="1.0\linewidth"} [0.9]{} ![Qualitative comparison between , SPAIR, IODINE and GENESIS for the 3D-Room dataset.[]{data-label="fig:eval:qualitative-3d"}](images/3dsmall_compare.png "fig:"){width="1.0\linewidth"} [0.8]{} ![Qualitative comparison between , SPAIR, IODINE and GENESIS for Space Invaders, Air Raid, and River Raid.[]{data-label="fig:eval:qualitative-atari"}](images/spaceinvader_compare.png "fig:"){width="1.0\linewidth"} [1.0]{} ![Qualitative comparison between , SPAIR, IODINE and GENESIS for Space Invaders, Air Raid, and River Raid.[]{data-label="fig:eval:qualitative-atari"}](images/airraid_compare.png "fig:"){width="1.0\linewidth"} [0.9]{} ![Qualitative comparison between , SPAIR, IODINE and GENESIS for Space Invaders, Air Raid, and River Raid.[]{data-label="fig:eval:qualitative-atari"}](images/riverraid_compare.png "fig:"){width="1.0\linewidth"} [0.8]{} ![Case illustration of Montezuma’s Revenge comparing object-detection behaviour in and SPAIR.[]{data-label="fig:eval:case-illus-atari"}](images/montezuma_compare.png "fig:"){width="1.0\linewidth"} We evaluate our model on two datasets: 1) an Atari [@bellemare13arcade] dataset that consists of random images from a pretrained agent playing the games, and 2) a generated 3D-room dataset that consists of images of a walled enclosure with a random number of objects on the floor. In order to test the scalability of our model, we use both a small 3D-room dataset that has 4-8 objects and a large 3D-room dataset that has 18-24 objects. Each image is taken from a random camera angle and the colors of the objects, walls, floor, and sky are also chosen at random. Additional details of the datasets can be found in the Appendix \[ax:dataset\]. **Baselines.** We compare our model against two mixture-scene models (IODINE and GENESIS) and one spatial-attention model (SPAIR). Since SPAIR does not have an explicit background component, we add an additional VAE for processing the background. Additionally, we test against two implementations of SPAIR: one where we train on the entire image using a $16 \times 16$ grid and another where we train on random $32 \times 32$ pixel patches using a $4\times4$ grid. The latter is consistent with SPAIR’s training regime on Space Invaders. This kind of patch-based learning is employed because training SPAIR on the larger grid size is slow due to the sequential nature of the latent inference in SPAIR. Lastly, for performance reasons, unlike the original SPAIR implementation, we use parallel processing for rendering the objects from their respective latents onto the canvas[^2]. Thus, our SPAIR implementation can be seen as a stronger baseline in terms of the speed than the original SPAIR. The complete details of the architecture used is given in Appendix \[ax:implementation\]. Qualitative Comparison of Inferred Representations -------------------------------------------------- In this section, we provide a qualitative analysis of the generated representations of the different models. Figure \[fig:eval:3d-space\] and Figure \[fig:eval:atari-space\] show some samples of from the 3D-Room and Atari datasets. More qualitative results of can be found in Appendix \[ax:more\_results\]. Figure \[fig:eval:qualitative-3d\] shows sample scene decompositions of our baselines from the 3D-Room dataset and Figure \[fig:eval:qualitative-atari\] shows the results on Atari. Note that SPAIR does not use component masks and IODINE and GENESIS do not separate foreground from background, hence the corresponding cells are left empty. Additionally, we only show a few representative components for IODINE and GENESIS since we ran those experiments with larger $K$ than can be displayed. **IODINE & GENESIS.** In the 3D-Room environment, IODINE is able to segment the objects and the background into separate components. However, it occasionally does not properly decompose objects (see the orange ball in the Small 3D-Room experiment is missing from the reconstruction) and may generate blurry objects. GENESIS, on the other hand, is unable to capture foreground objects, instead segments the background walls, floor, and sky into multiple components. In Atari, for all games, both IODINE and GENESIS fail to capture the foreground properly. We believe this is because the objects in Atari games are smaller, less regular and lack the obvious latent factors like color and shape as in the 3D dataset, which demonstrates that detection-based approaches are more appropriate in this case. **SPAIR.** The $16 \times 16$ implementation of SPAIR is able to detect tight bounding boxes in both 3D-Room and most Atari games. When SPAIR is trained on patches, it often fails to detect the foreground objects in proper bounding boxes, frequently uses multiple bounding boxes for one object and redundantly detects parts of the background as foreground objects. This is a limitation of the patch training as the receptive field of each patch is limited to a $32 \times 32$, hence prohibiting it to detect objects larger than that and making it difficult to distinguish the background from foreground. These two properties are illustrated well in Space Invaders, where it is able to detect the small aliens, but it detects the long piece of background ground on the bottom of the image as foreground objects. **SPACE.** In 3D-Room, SPACE is able to accurately detect almost all objects despite the large variations in object positions, colors and shapes, while producing a clean segmentation of the background walls, ground, and sky. This is in contrast to the SPAIR model, while being able to provide a similar foreground detection quality, encodes the whole background into a single component, which makes the representation less disentangled and the reconstruction more blurry. Similarly in Atari, SPACE consistently captures all foreground objects while producing clean background segmentation across many different games. **Dynamic Backgrounds.** SPACE and SPAIR exhibit some very interesting behavior when trained on games with dynamic backgrounds. For the most static game - Space Invaders, both SPACE and SPAIR work well. For Air Raid, in which the background building moves, SPACE captures all objects accurately while providing a two-component segmentation, whereas SPAIR $16 \times 16$ and SPAIR $4 \times 4$ patch produce splitting and heavy re-detections. In the most dynamic games, SPAIR completely fails because of the difficulty to model dynamic background with a single VAE component, while SPACE is able to perfectly segment the blue racing track while accurately detecting all foreground objects. **Foreground vs Background.** Typically, foreground is the dynamic local part of the scene that we are interested in, and background is the relatively static and global part. This definition, though intuitive, is ambiguous. Some of the objects, such as the red shields near the bottom in Space Invaders and the key in Montezuma’s Revenge (Figure \[fig:eval:case-illus-atari\]) are detected as foreground objects in SPACE, but are considered background in SPAIR. Though these objects are static[^3], they are important elements of the games and should be considered as foreground objects. Similar behavior is observed in Atlantis, where SPACE detects some foreground objects from the middle base that is above the water. We believe this is an interesting property of SPACE and could be very important for providing useful representations for downstream tasks. By using a spatial broadcast network [@Watters2019SpatialBD] which is much weaker when compared to other decoders like sub-pixel convolutional nets ([@Shi_2016_CVPR]), we limit the capacity of background module, which favors modeling static objects as foreground rather than background. **Boundary Loss.** We notice SPAIR sometimes splits objects into two whereas SPACE is able to create the correct bounding box for the objects (for example, see Air Raid). This may be attributed to the addendum of the auxiliary boundary loss in the SPACE model that would penalize splitting an object with multiple bounding boxes. Quantitative Comparison ----------------------- In this section we compare SPACE with the baselines in several quantitative metrics. We first note that each of the baseline models has a different *decomposition capacity*, which we define as the capability of the model to decompose the scene into its semantic constituents such as the foreground objects and the background segmented components. For SPACE, the decomposition capacity is equal to the number of grid cells $H\times W$ (which is the maximum number of foreground objects that can be detected) plus the number of background components $K$. For SPAIR, the decomposition capacity is equal to the number of grid cells $H \times W$ plus 1 for background. For IODINE and GENESIS, it is equal to the number of components $K$. [0.24]{} ![Quantitative performance comparison between , SPAIR, IODINE and GENESIS in terms of batch-processing time during training, training convergence and converged pixel MSE. Convergence plots showing pixel-MSE were computed on a held-out validation set during training.[]{data-label="fig:eval:mse"}](tikz/training_latency_plot.pdf "fig:"){width="1.0\linewidth"} [0.25]{} ![Quantitative performance comparison between , SPAIR, IODINE and GENESIS in terms of batch-processing time during training, training convergence and converged pixel MSE. Convergence plots showing pixel-MSE were computed on a held-out validation set during training.[]{data-label="fig:eval:mse"}](tikz/convergence_plot_3droom_4x4.pdf "fig:"){width="1.0\linewidth"} [0.24]{} ![Quantitative performance comparison between , SPAIR, IODINE and GENESIS in terms of batch-processing time during training, training convergence and converged pixel MSE. Convergence plots showing pixel-MSE were computed on a held-out validation set during training.[]{data-label="fig:eval:mse"}](tikz/convergence_plot_3droom_8x8.pdf "fig:"){width="1.0\linewidth"} [0.24]{} ![Quantitative performance comparison between , SPAIR, IODINE and GENESIS in terms of batch-processing time during training, training convergence and converged pixel MSE. Convergence plots showing pixel-MSE were computed on a held-out validation set during training.[]{data-label="fig:eval:mse"}](tikz/convergence_plot_3droom_16x16.pdf "fig:"){width="1.0\linewidth"} [1.0]{} ![Quantitative performance comparison between , SPAIR, IODINE and GENESIS in terms of batch-processing time during training, training convergence and converged pixel MSE. Convergence plots showing pixel-MSE were computed on a held-out validation set during training.[]{data-label="fig:eval:mse"}](tikz/legend.png "fig:"){width="1.0\linewidth"} **Gradient Step Latency.** The leftmost chart of Figure \[fig:eval:mse\] shows the time taken to complete one gradient step for different decomposition capacities for each of the models. We see that SPAIR’s latency grows with the number of cells because of the sequential nature of its latent inference step. Similarly GENESIS and IODINE’s latency grows with the number of components $K$ because each component is processed sequentially in both the models. IODINE is the slowest overall with its computationally expensive iterative inference procedure. Furthermore, both IODINE and GENESIS require storing data for each of the $K$ components, so we were unable to run our experiments on 256 components or greater before running out of memory on our 22GB GPU. On the other hand, SPACE employs parallel processing for the foreground which makes it scalable to large grid sizes, allowing it to detect a large number of foreground objects without any significant performance degradation. Although this data was collected for gradient step latency, this comparison implies a similar relationship exists with inference time which is a main component in the gradient step. **Time for Convergence.** The remaining three charts in Figure \[fig:eval:mse\] show the amount of time each model takes to converge in different experimental settings. We use the pixel-wise mean squared error (MSE) as a measurement of how close a model is to convergence. We see that not only does SPACE achieve the lowest MSE, it also converges the quickest out of all the models. **Average Precision and Error Rate.** In order to assess the quality of our bounding box predictions, we measure the Average Precision and Object Count Error Rate of our predictions. Our results are shown in Table \[tbl:metrics\]. We only report these metrics for 3D-Room since we have access to the ground truth bounding boxes for each of the objects in the scene. has a slightly better error rate than our implementation of SPAIR but a comparable average precision. We can assert that, despite using a parallel foreground module and thus having much faster inference latency for large number of objects, can still find similar quality bounding boxes as SPAIR. Conclusion ========== We propose SPACE, a unified probabilistic model that combines the benefits of the object representation models based on spatial attention and the scene decomposition models based on component mixture. SPACE can explicitly provide factorized object representation per foreground object while also decomposing complex background segments. SPACE also achieves a significant speed-up and thus makes the model applicable to scenes with a much larger number of objects without performance degradation. Besides, the detected objects in SPACE are also more intuitive than other methods. We show the above properties of SPACE on Atari and 3D-Rooms. Interesting future directions are to replace the sequential processing of background by a parallel one and to improve the model for natural images. Our next plan is to apply SPACE for object-oriented model-based reinforcement learning. ### Acknowledgments {#acknowledgments .unnumbered} SA thanks to Kakao Brain, Center for Super Intelligence (CSI), and Element AI for their support. Additional Results of {#ax:more_results} ====================== [1.0]{} ![Object detection and background segmentation using on 3D-Room data set with small number of objects. Each row corresponds to one input image.[]{data-label="fig:eval:qualitative-3dsmall-more"}](images/3dsmall_showcase_1.png "fig:"){width="1.0\linewidth"} [1.0]{} ![Object detection and background segmentation using on 3D-Room data set with large number of objects.[]{data-label="fig:eval:qualitative-3dsmall-more"}](images/3dlarge_showcase_2.png "fig:"){width="1.0\linewidth"} ELBO Derivations ================ In this section, we derive the ELBO for the log-likelihood $\log p(\rvx)$. \[ax:derivations\] $$\begin{aligned} \log p(\rvx) &= \log \iint p(\rvx|\rvz^{\text{fg}},\rvz^{\text{bg}}_{1:K})\left[p(\rvz^{\text{fg}})\prod_{k=1}^K p(\rvz^{\text{bg}}_{k}|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}})\right] d\rvz_{1:K}d\rvz^{\text{fg}}\\ &= \log \eE_{q(\rvz^{\text{fg}}, \rvz^\bg_{1:K}|\rvx)} p(\rvx|\rvz^{\text{fg}},\rvz^{\text{bg}}_{1:K})\frac{p(\rvz^{\text{fg}})\prod_{k=1}^K p(\rvz^{\text{bg}}_{k}|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}})}{q(\rvz^{\text{fg}}|\rvx)\prod_{k=1}^K q(\rvz^{\text{bg}}_{k}|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}}, \rvx)}\\ &\geq \eE_{q(\rvz^{\text{fg}}, \rvz^\bg_{1:K}|\rvx)} \log \left[p(\rvx|\rvz^{\text{fg}},\rvz^{\text{bg}}_{1:K})\frac{p(\rvz^{\text{fg}})\prod_{k=1}^K p(\rvz^{\text{bg}}_{k}|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}})}{q(\rvz^{\text{fg}}|\rvx)\prod_{k=1}^K q(\rvz^{\text{bg}}_{k}|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}}, \rvx)}\right]\\ &= \eE_{q(\rvz^{\text{fg}}, \rvz^{\text{bg}}_{1:K}|\rvx)} \log\left[p(\rvx|\rvz^{\text{fg}}, \rvz^{\text{bg}}) \right] - \KL(q(\rvz^{\text{fg}}|\rvx)\parallel p(\rvz^{\text{fg}})) \\ &\quad -\eE_{q(\rvz^{\text{fg}}|\rvx)} \sum_{k=1}^K \eE_{q(\rvz^{\text{bg}}_{<k}|\rvx, \rvz^{\text{fg}})} \KL(q(\rvz^{\text{bg}}_k|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}}, \rvx)\parallel p(\rvz^{\text{bg}}_k|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}}))\end{aligned}$$ **KL Divergence for the Foreground Latents** Under the ’s approximate inference, the $\KL(q(\rvz^{\text{fg}}|\rvx)\parallel p(\rvz^{\text{fg}}))$, is evaluated as follows. $$\begin{aligned} &\KL(q(\rvz^{\text{fg}}|\rvx)\parallel p(\rvz^{\text{fg}})) \\ &\hspace{5mm}= \eE_{q(\rvz^{\text{fg}}|\rvx)} \log \frac{q(\rvz^{\text{fg}}|\rvx)}{p(\rvz^{\text{fg}})}\\ &\hspace{5mm}= \eE_{q(\rvz^{\text{fg}}|\rvx)} \log \frac{\prod_{i=1}^{HW} q(\rvz^{\text{pres}}_i|\rvx)\left(q(\rvz^{\text{where}}_i|\rvx)q(\rvz^{\text{what}}_i|\rvz^{\text{where}}_i,\rvx)q(\rvz^{\text{depth}}_i|\rvx)\right)^{\rvz^{\text{pres}}_i}}{\prod_{i=1}^{HW} p(\rvz^{\text{pres}}_i)\left(p(\rvz^{\text{where}}_i)p(\rvz^{\text{what}}_i)p(\rvz^{\text{depth}}_i)\right)^{\rvz^{\text{pres}}_i}}\\ &\hspace{5mm}= \sum_{i=1}^{HW}\Bigg[\KL(q(\rvz^{\text{pres}}_i|\rvx)\parallel p(\rvz^{\text{pres}}_i)) + \eE_{q(\rvz^{\text{pres}}_i|\rvx)}\rvz^{\text{pres}}_i\Big[ \KL(q(\rvz^{\text{where}}_i|\rvx)\parallel p(\rvz^{\text{where}}_i)) \\ &\hspace{5mm} +\eE_{q(\rvz^{\text{where}}_i|\rvx)}\KL(q(\rvz^{\text{what}}_i|\rvx)\parallel p(\rvz^{\text{what}}_i)) + \KL(q(\rvz^{\text{depth}}_i|\rvx)\parallel p(\rvz^{\text{depth}}_i))\Big]\Bigg]\end{aligned}$$ **KL Divergence for the Background Latents** Under our GENESIS-like modeling of inference for the background latents, the KL term for the background is evaluated as follows. $$\begin{aligned} &\KL(q(\rvz^{\text{bg}}_k|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}}, \rvx)\parallel p(\rvz^{\text{bg}}_k|\rvz^{\text{bg}}_{<k}, \rvz^{\text{fg}}))\\ &\hspace{5mm}= \KL(q(\rvz^m_k|\rvz^m_{<k}, \rvz^{\text{fg}}, \rvx)\parallel p(\rvz^m_k|\rvz^m_{<k}, \rvz^{\text{fg}})) + \eE_{q(\rvz^m_k|\rvz^m_{<k},\rvz^{\text{fg}}, \rvx)} \KL(q(\rvz^c_k|\rvz^m_k,\rvx)\parallel p(\rvz^c_k|\rvz^m_k))\end{aligned}$$ **Relaxed treatment of $\rvz^{\text{pres}}$** In our implementation, we model the Bernoulli random variable $\rvz^{\text{pres}}_i$ using the Gumbel-Softmax distribution [@jang2016categorical]. We use the relaxed value of $\rvz^{\text{pres}}$ in the entire training and use hard samples only for the visualizations. Boundary Loss {#ax:aux_loss} ============= In this section we elaborate on the implementation details of the *boundary loss*. We construct a kernel of the size of the glimpse, $gs\times gs$ (we use $gs=32$) with a boundary gap of $b=6$ having negative uniform weights inside the boundary and a zero weight in the region between the boundary and the glimpse. This ensures that the model is penalized when the object is outside the boundary. This kernel is first mapped onto the global space via STN [@spatial_transformer] to obtain the global kernel. This is then multiplied element-wise with global object mask $\alpha$ to obtain the *boundary loss map*. The objective of the loss is to minimize the mean of this *boundary loss map*. In addition to the ELBO, this loss is also back-propagated via *RMSProp* ([@rmsprop]). Implementation Details {#ax:implementation} ====================== Dataset Details {#ax:dataset} =============== **Atari.** For each game, we sample 60,000 random images from a pretrained agent [@wu2016tensorpack]. We split the images into 50,000 for the training set, 5,000 for the validation set, and 5,000 for the testing set. Each image is preprocessed into a size of $128\times128$ pixels with BGR color channels. We present the results for the following games: Space Invaders, Air Raid, River Raid, Montezuma’s Revenge. We also train our model on a dataset of 10 games jointly, where we have 8,000 training images, 1,000 validation images, and 1,000 testing images for each game. We use the following games: Asterix, Atlantis, Carnival, Double Dunk, Kangaroo, Montezuma Revenge, Pacman, Pooyan, Qbert, Space Invaders. **Room 3D.** We use MuJoCo [@mujoco] to generate this dataset. Each image consists of a walled enclosure with a random number of objects on the floor. The possible objects are randomly sized spheres, cubes, and cylinders. The small 3D-Room dataset has 4-8 objects and the large 3D-Room dataset has 18-24 objects. The color of the objects are randomly chosen from 8 different colors and the colors of the background (wall, ground, sky) are chosen randomly from 5 different colors. The angle of the camera is also selected randomly. We use a training set of 63,000 images, a validation set of 7,000 images, and a test set of 7,000 images. We use a 2-D projection from the camera to determine the ground truth bounding boxes of the objects so that we can report the average precision of the different models. [^1]: Work done while visiting Rutgers University. Authors named inside {} equally contributed. Correspondance to `[email protected]` [^2]: It is important to note that the worst case complexity of rendering is $\mathcal{O}(hw\times HW)$, (where $(h, w)$ is the image size) which is extremely time consuming when we have large image size and/or large number of objects. [^3]: The pretrained agent used to collect the game frames for Montezuma’s Revenge does not reach a state in which it can actually capture the key, hence the position of the key remains constant over time
{ "pile_set_name": "ArXiv" }
--- abstract: 'Knowledge base exchange is an important problem in the area of data exchange and knowledge representation, where one is interested in exchanging information between a source and a target knowledge base connected through a mapping. In this paper, we study this fundamental problem for knowledge bases and mappings expressed in , the profile of based on the description logic . More specifically, we consider the problem of computing universal solutions, identified as one of the most desirable translations to be materialized, and the problem of computing -representations, which optimally capture in a target TBox the information that can be extracted from a source TBox and a mapping by means of unions of conjunctive queries. For the former we provide a novel automata-theoretic technique, and complexity results that range from to , while for the latter we show -completeness.' author: - | Marcelo Arenas\ PUC Chile &\ Univ. of Oxford, U.K.\ [email protected] Elena Botoeva\ Free U. of Bolzano\ Italy\ [email protected] Diego Calvanese\ Free U. of Bolzano, Italy &\ TU Vienna, Austria\ [email protected] Vladislav Ryzhikov\ Free U. of Bolzano\ Italy\ [email protected] title: Exchanging OWL 2 QL Knowledge Bases --- Introduction {#sec-introduction} ============ Complex forms of information, maintained in different formats and organized according to different structures, often need to be shared between agents. In recent years, both in the data management and in the knowledge representation communities, several settings have been investigated that address this problem from various perspectives: in *information integration*, uniform access is provided to a collection of data sources by means of an ontology (or global schema) to which the sources are mapped [@Lenz02]; in *peer-to-peer systems*, a set of peers declaratively linked to each other collectively provide access to the information assets they maintain [@KeAM03; @ACGRS06; @FKMT06]; in *ontology matching*, the aim is to understand and derive the correspondences between elements in two ontologies [@EuSh07; @SE13]; finally, in *data exchange*, the information stored according to a source schema needs to be restructured and translated so as to conform to a target schema [@FKMP05; @Barc09]. The work we present in this paper is inspired by the latter setting, investigated in databases. We study it, however, under the assumption of incomplete information typical of knowledge representation [@ArPR11]. Specifically, we investigate the problem of *knowledge base exchange*, where a source knowledge base (KB) is connected to a target KB by means of a declarative mapping specification, and the aim is to exchange knowledge from the source to the target by exploiting the mapping. We rely on a framework for KB exchange based on lightweight Description Logics (DLs) of the family [@CDLLR07], recently proposed in [@ABCRS12; @ABCRS12b]: both source and target are KBs constituted by a DL TBox, representing implicit information, and an ABox, representing explicit information, and mappings are sets of DL concept and role inclusions. Note that in data and knowledge base exchange, differently from ontology matching, mappings are first-class citizens. In fact, it has been recognized that building schema mappings is an important and complex activity, which requires the designer to have a thorough understanding of the source and how the information therein should be related to the target. Thus, several techniques and tools have been developed to support mapping design, e.g., exploiting lexical information [@FHHM*09]. Here, similar to data exchange, we assume that for building mappings the target signature is given, but no further axioms constraining the target knowledge are available. In fact, such axioms are derived from the source KB and the mapping. We consider two key problems: computing *universal solutions*, which have been identified as one of the most desirable translations to be materialized; *-representability* of a source TBox by means of a target TBox that captures at best the intensional information that can be extracted from the source according to a mapping using union of conjunctive queries. Determining -representability is a crucial task, since it allows one to use the obtained target TBox to infer new knowledge in the target, thus reducing the amount of extensional information to be transferred from the source. Moreover, it has been noticed that in many data exchange applications users only extract information from the translated data by using specific queries (usually conjunctive queries), so query-based notions of translation specifically tailored to store enough information to answer such queries have been widely studied in the data exchange area [@MH03; @FKNP08; @APRR09; @FK12; @PSS13]. For these two problems, we investigate both the task of checking *membership*, where a candidate universal solution (resp., -representation) is given and one needs to check its correctness, and *non-emptiness*, where the aim is to determine the existence of a universal solution (resp., -representation). We significantly extend previous results in several directions. First of all, we establish results for [@OWL2QL], one of the profiles of the standard Web Ontology Language OWL 2 [@OWL2], which is based on the DL . To do so, we have to overcome the difficulty of dealing with null values in the ABox, since these become necessary in the target to represent universal solutions. Also, for the first time, we address disjointness assertions in the TBox, a construct that is part of . The main contribution of our work is then a detailed analysis of the computational complexity of both membership and non-emptiness for universal solutions and -representability. For the non-emptiness problem of universal solutions, previous known results covered only the simple case of $\dlliterdfs$, the RDFS fragment of , in which no new facts can be inferred, and universal solutions always exist and can be computed in polynomial time via a chase procedure (see [@CDLLR07]). We show that in our case, instead, the problem is -hard, hence significantly more complex, and provide an upper bound based on a novel approach exploiting two-way alternating automata. We provide also NP upper bounds for the simpler case of ABoxes without null values, and for the case of the membership problem. As for -representability, we adopt the notion of *-representability* introduced in [@ABCRS12; @ABCRS12b] and extend it to take into account disjointness of . For that case we show -completeness of both non-emptiness and membership, improving on the previously known upper bounds. The paper is organized as follows. In Section \[sec-preliminaries\], we give preliminary notions on DLs and queries. In Section \[sec-kbe\], we define our framework of KB exchange and discuss the problem of computing solutions. In Section \[sec-cont\], we overview our contributions, and then we provide our results on computing universal solutions in Section \[sec-univ\], and on UCQ-representability in Section \[sec-ucq-rep\]. Finally, in Section \[sec-conclusions\], we draw some conclusions and outline issues for future work. The proofs are available in an extended technical report accessible at <http://arxiv.org/abs/1304.5810>. Preliminaries {#sec-preliminaries} ============= The DLs of the family [@CDLLR07] of light-weight DLs are characterized by the fact that standard reasoning can be done in polynomial time. We adapt here , the DL underlying , and present now its syntax and semantics. Let $N_C$, $N_R$, $N_a$, $N_\ell$ be pairwise disjoint sets of *concept names*, *role names*, *constants*, and *labeled nulls*, respectively. Assume in the following that $A\in N_C$ and $P\in N_R$; in , $B$ and $C$ are used to denote basic and arbitrary (or complex) concepts, respectively, and $R$ and $Q$ are used to denote basic and arbitrary (or complex) roles, respectively, defined as follows: $$\begin{array}{r@{~~}c@{~~}l} R & ::= & P ~\mid~ P^- \\ Q & ::= & R ~\mid~ \NOT R \end{array} \qquad\qquad \begin{array}{r@{~~}c@{~~}l} B & ::= & A ~\mid~ \SOMET{R} \\ C & ::= & B ~\mid~ \NOT B \end{array}$$ From now on, for a basic role $R$, we use $R^-$ to denote $P^-$ when $R=P$, and $P$ when $R=P^-$. A TBox is a finite set of *concept inclusions* $B \ISA C$ and *role inclusions* $R\ISA Q$. We call an inclusion of the form $B_1 \ISA \neg B_2$ or $R_1 \ISA \neg R_2$ a *disjointness assertion*. An ABox is a finite set of *membership assertions* $B(a)$, $R(a,b)$, where $a, b \in N_a$. In this paper, we also consider extended ABoxes, which are obtained by allowing labeled nulls in membership assertions. Formally, an *extended ABox* is a finite set of membership assertions $B(u)$ and $R(u,v)$, where $u, v \in (N_a \cup N_\ell)$. Moreover, a(n *extended*) *KB* $\K$ is a pair $\tup{\T,\A}$, where $\T$ is a TBox and $\A$ is an (extended) ABox. A *signature* $\Sigma$ is a finite set of concept and role names. A KB $\K$ is said to be *defined over* (or simply, *over*) $\Sigma$ if all the concept and role names occurring in $\K$ belong to $\Sigma$ (and likewise for TBoxes, ABoxes, concept inclusions, role inclusions and membership assertions). Moreover, an *interpretation* $\I$ *of* $\Sigma$ is a pair $\tup{\dom,\Int{\cdot}}$, where $\dom$ is a non-empty domain and $\Int{\cdot}$ is an interpretation function such that: $\Int{A} \subseteq \dom$, for every concept name $A \in \Sigma$; $\Int{P} \subseteq \dom \times \dom$, for every role name $P \in \Sigma$; and \[it:int-obj\] $\Int{a} \in \dom$, for every constant $a \in N_a$. Function $\Int{\cdot}$ is extended to also interpret concept and role constructs: $$\begin{array}{r@{~~}c@{~~}l} \Int{(\SOMET{R})} &=& \{x \in \dom \st \exists y \in \dom\text{ such that }(x,y) \in \Int{R}\};\\ \Int{(P^-)} &=& \{(y,x) \in \dom \times \dom \st (x,y) \in \Int{P}\};\\ \Int{(\NOT B)} &=& \dom \setminus \Int{B}; \qquad \Int{(\NOT R)}~=~(\dom \times \dom) \setminus \Int{R}. \end{array}$$ Note that, consistently with the semantics of , we do *not* make the unique name assumption (UNA), i.e., we allow distinct constants $a,b\in N_a$ to be interpreted as the same object, i.e., $\Int{a}=\Int{b}$. Note also that labeled nulls are *not* interpreted by $\I$. Let $\I = \tup{\dom,\Int{\cdot}}$ be an interpretation over a signature $\Sigma$. Then $\I$ is said to satisfy a concept inclusion $B \ISA C$ over $\Sigma$, denoted by $\I \models B \ISA C$, if $\Int{B} \subseteq \Int{C}$; $\I$ is said to satisfy a role inclusion $R \ISA Q$ over $\Sigma$, denoted by $\I \models R \ISA Q$, if $\Int{R} \subseteq \Int{Q}$; and $\I$ is said to satisfy a TBox $\T$ over $\Sigma$, denoted by $\I \models \T$, if $\I \models \alpha$ for every $\alpha \in \T$. Moreover, satisfaction of membership assertions over $\Sigma$ is defined as follows. A *substitution* over $\I$ is a function $h : (N_a \cup N_\ell) \to \Int{\Delta}$ such that $h(a) = \Int{a}$ for every $a \in N_a$. Then $\I$ is said to satisfy an (extended) ABox $\A$, denoted by $\I \models \A$, if there exists a substitution $h$ over $\I$ such that: - for every $B(u) \in \A$, it holds that $h(u) \in \Int{B}$; and - for every $R(u,v) \in \A$, it holds that $(h(u),h(v)) \in \Int{R}$. Finally, $\I$ is said to *satisfy* a(n extended) KB $\kb$, denoted by $\I \models \K$, if $\I \models \T$ and $\I \models \A$. Such $\I$ is called a *model* of $\K$, and we use $\Mod(\K)$ to denote the set of all models of $\K$. We say that $\K$ is *consistent* if $\Mod(\K) \neq \emptyset$. As is customary, given an (extended) KB $\K$ over a signature $\Sigma$ and a membership assertion or an inclusion $\alpha$ over $\Sigma$, we use notation $\K \models \alpha$ to indicate that for every interpretation $\I$ of $\Sigma$, if $\I\models \K$, then $\I \models \alpha$. Queries and certain answers --------------------------- A $k$-ary query $q$ over a signature $\Sigma$, with $k \geq 0$, is a function that maps every interpretation $\tup{\dom, \Int{\cdot}}$ of $\Sigma$ into a $k$-ary relation $\Int{q} \subseteq (\dom)^k$. In particular, if $k = 0$, then $q$ is said to be a Boolean query, and $\Int{q}$ is either a relation containing the empty tuple $()$ (representing the value true) or the empty relation (representing the value false). Given a KB $\K$ over $\Sigma$, the set of *certain answers* to $q$ over $\K$, denoted by $\cert(q, \K)$, is defined as: $$\textstyle \bigcap_{\I\in\Mod(\K)} \{ \begin{array}[t]{@{}l} (a_1, \ldots, a_k) \ \mid\\ \{a_1, \ldots, a_k\} \subseteq N_a \text{ and } (a_1^\I, \ldots, a_k^\I) \in q^\I \}, \end{array}$$ Notice that the certain answer to a query does *not* contain labeled nulls. Besides, notice that if $q$ is a Boolean query, then $\cert(q, \K)$ evaluates to true if $\Int{q}$ evaluates to true for every $\I \in \Mod(\K)$, and it evaluates to false otherwise. A *conjunctive query $(\CQ)$ over a signature $\Sigma$* is a formula of the form $q(\vec x) = \exists \vec y.\, \varphi(\vec x, \vec y)$, where $\vec x$, $\vec y$ are tuples of variables and $\varphi(\vec x, \vec y)$ is a conjunction of atoms of the form $A(t)$, with $A$ a concept name in $\Sigma$, and $P(t,t')$, with $P$ a role name in $\Sigma$, where each of $t,t'$ is either a constant from $N_a$ or a variable from $\vec x$ or $\vec y$. Given an interpretation $\I = \tup{\dom, \Int{\cdot}}$ of $\Sigma$, the answer of $q$ over $\I$, denoted by $\Int{q}$, is the set of tuples $\vec a$ of elements from $\dom$ for which there exist a tuple $\vec b$ of elements from $\dom$ such that $\I$ satisfies every conjunct in $\varphi(\vec a, \vec b)$. A union of conjunctive queries ($\UCQ$) over a signature $\Sigma$ is a formula of the form $q(\vec{x}) = \textstyle \bigvee_{i=1}^n q_i(\vec{x})$, where each $q_i$ ($1 \leq i \leq n$) is a over $\Sigma$, whose semantics is defined as $\Int{q} = \bigcup_{i=1}^n \Int{q_i}$. Exchanging Knowledge Bases {#sec-kbe} ========================== We generalize now, in Section \[sec-sol\], the setting proposed in [@ArPR11] to , and we formalize in Section \[sec-problems\] the main problems studied in the rest of the paper. A knowledge base exchange framework for {#sec-sol} ---------------------------------------- Assume that $\Sigma_1$, $\Sigma_2$ are signatures with no concepts or roles in common. An inclusion $E_1 \ISA E_2$ is said to be *from* $\Sigma_1$ *to* $\Sigma_2$, if $E_1$ is a concept or a role over $\Sigma_1$ and $E_2$ is a concept or a role over $\Sigma_2$. A mapping is a tuple $\M = (\Sigma_1, \Sigma_2, \T_{12})$, where $\T_{12}$ is a TBox consisting of inclusions from $\Sigma_1$ to $\Sigma_2$ [@ABCRS12]. Recall that in this paper, we deal with $\dlliter$ TBoxes only, so $\T_{12}$ is assumed to be a set of $\dlliter$ concept and role inclusions. The semantics of such a mapping is defined in [@ABCRS12] in terms of a notion of satisfaction for interpretations, which has to be extended in our case to deal with interpretations not satisfying the UNA (and, more generally, the standard name assumption). More specifically, given interpretations $\I$, $\J$ of $\Sigma_1$ and $\Sigma_2$, respectively, pair $(\I,\J)$ *satisfies* TBox $\T_{12}$, denoted by $(\I, \J) \models \T_{12}$, if for every $a \in N_a$, it holds that $a^\I = a^\J$, for every concept inclusion $B \ISA C \in \T_{12}$, it holds that $\Int[\I]{B} \subseteq \Int[\J]{C}$, and for every role inclusion $R \ISA Q \in \T_{12}$, it holds that $\Int[\I]{R} \subseteq \Int[\J]{Q}$. Notice that the connection between the information in $\I$ and $\J$ is established through the constants that move from source to target according to the mapping. For this reason, we require constants to be interpreted in the same way in $\I$ and $\J$, i.e., they preserve their meaning when they are transferred. Besides, notice that this is the only restriction imposed on the domains of $\I$ and $\J$ (in particular, we require neither that $\Delta^\I = \Delta^\J$ nor that $\Delta^\I \subseteq \Delta^\J$). Finally, $\Sat_\M(\I)$ is defined as the set of interpretations $\J$ of $\Sigma_2$ such that $(\I,\J) \models \T_{12}$, and given a set ${\cal X}$ of interpretations of $\Sigma_1$, $\Sat_\M({\cal X})$ is defined as $\bigcup_{\I\in \mathcal{X}} \Sat_\M(\I)$. The main problem studied in the knowledge exchange area is the problem of translating a KB according to a mapping, which is formalized through several different notions of translation (for a thorough comparison of different notions of solutions see [@ABCRS12]). The first such notion is the concept of solution, which is formalized as follows. Given a mapping $\M = (\Sigma_1, \Sigma_2, \T_{12})$ and KBs $\K_1$, $\K_2$ over $\Sigma_1$ and $\Sigma_2$, respectively, $\K_2$ is a *solution* for $\K_1$ under $\M$ if $\Mod(\K_2) \subseteq \Sat_\M(\Mod(\K_1))$. Thus, $\K_2$ is a solution for $\K_1$ under $\M$ if every interpretation of $\K_2$ is a valid translation of an interpretation of $\K_1$ according to $\M$. Although natural, this is a mild restriction, which gives rise to the stronger notion of universal solution. Given $\M$, $\K_1$ and $\K_2$ as before, $\K_2$ is a *universal solution* for $\K_1$ under $\M$ if $\Mod(\K_2) = \Sat_\M(\Mod(\K_1))$. Thus, $\K_2$ is designed to exactly represent the space of interpretations obtained by translating the interpretations of $\K_1$ under $\M$ [@ABCRS12]. Below is a simple example demonstrating the notion of universal solutions. This example also illustrates some issues regarding the absence of the UNA, which has to be given up to comply with the standard, and regarding the use of disjointness assertions. \[exa-disj-source\] Assume $\M = (\{\concept{F}, \concept{G}\}, \{\concept{F'}, \concept{G'}\}$, $\T_{12})$, where $\T_{12}=\{F \ISA F', G \ISA G'\}$, and let $\K_1 = \tup{\T_1, \A_1}$, where $\T_1 = \{\}$ and $\A_1 = \{F(a), G(b)\}$. Then the ABox $\A_2 = \{F'(a), G'(b)\}$ is a universal solution for $\K_1$ under $\M$. Now, if we add a seemingly harmless disjointness assertion $\{F \ISA \NOT G\}$ to $\T_1$, we obtain that $\A_2$ is no longer a universal solution (not even a solution) for $\K_1$ under $\M$. The reason for that is the lack of the UNA on the one hand, and the presence of the disjointness assertion in $\T_1$ on the other hand. In fact, the latter forces $a$ and $b$ to be interpreted differently in the source. Thus, for a model $\J$ of $\A_2$ such that $\Int[\J]{a} = \Int[\J]{b}$ and $\Int[\J]{F'} = \Int[\J]{G'} = \{\Int[\J]{a}\}$, there exists no model $\I$ of $\K_1$ such that $(\I,\J) \models \T_{12}$ (which would require $\Int{a} = \Int[\J]{a}$ and $\Int{b} = \Int[\J]{b}$). In general, there exists no universal solution for $\K_1$ under $\M$, even though $\K_1$ and $\T_{12}$ are consistent with each other. A second class of translations is obtained in [@ABCRS12] by observing that solutions and universal solutions are too restrictive for some applications, in particular when one only needs a translation storing enough information to properly answer some queries. For the particular case of $\UCQ$, this gives rise to the notions of $\UCQ$-solution and universal $\UCQ$-solution. Given a mapping $\M = (\Sigma_1, \Sigma_2, \T_{12})$, a KB $\K_1 = \tup{\T_1,\A_1}$ over $\Sigma_1$ and a KB $\K_2$ over $\Sigma_2$, $\K_2$ is a *$\UCQ$-solution* for $\K_1$ under $\M$ if for every query $q \in \UCQ$ over $\Sigma_2$: $\cert(q, \tup{\T_1 \cup \T_{12}, \A_1}) \subseteq \cert(q, \K_2)$, while $\K_2$ is a *universal $\UCQ$-solution* for $\K_1$ under $\M$ if for every query $q \in \UCQ$ over $\Sigma_2$: $\cert(q, \tup{\T_1 \cup \T_{12}, \A_1}) = \cert(q, \K_2)$. Finally, a last class of solutions is obtained in [@ABCRS12] by considering that users want to translate as much of the knowledge in a TBox as possible, as a lot of effort is put in practice when constructing a TBox. This observation gives rise to the notion of $\UCQ$-representation [@ABCRS12], which formalizes the idea of translating a source TBox according to a mapping. Next, we present an alternative formalization of this notion, which is appropriate for our setting where disjointness assertions are considered.[^1] Assume that $\M = (\Sigma_1, \Sigma_2, \T_{12})$ and $\T_1$, $\T_2$ are TBoxes over $\Sigma_1$ and $\Sigma_2$, respectively. Then $\T_2$ is a [*$\UCQ$-representation*]{} of $\T_1$ under $\M$ if for every query $q \in \UCQ$ over $\Sigma_2$ and every ABox $\A_1$ over $\Sigma_1$ that is consistent with $\T_1$: $$\begin{gathered} \label{eq-ucq-rep} \tag{\dag} \cert(q, \tup{\T_1 \cup \T_{12}, \A_1}) \ = \\ \mathop{\bigcap_{\A_2 \,:\, \A_2 \text{ is an ABox over } \Sigma_2 \text{ that}}}_{\text{is a $\UCQ$-solution for } \A_1 \text{ under } \M} \cert(q, \tup{\T_2,\A_2}).\end{gathered}$$ Notice that in the previous definition, $\A_2$ is said to be a $\UCQ$-solution for $\A_1$ under $\M$ if the KB $\tup{\emptyset,\A_2}$ is a $\UCQ$-solution for the KB $\tup{\emptyset,\A_1}$ under $\M$. Let us explain the intuition behind the definition of the notion of $\UCQ$-representation. Assume that $\T_1$, $\T_2$, $\M$ satisfy . First, $\T_2$ captures the information in $\T_1$ that is translated by $\M$ and that can be extracted by using a , as for every ABox $\A_1$ over $\Sigma_1$ that is consistent with $\T_1$ and every $q$ over $\Sigma_2$, if we choose an arbitrary $\UCQ$-solution $\A_2$ for $\A_1$ under $\M$, then it holds that $\cert(q, \tup{\T_1 \cup \T_{12}, \A_1}) \subseteq \cert(q, \tup{\T_2,\A_2})$. Notice that $\A_1$ is required to be consistent with $\T_1$ in the previous condition, as we are interested in translating data that make sense according to $\T_1$. Second, $\T_2$ does not include any piece of information that can be extracted by using a and it is not the result of translating the information in $\T_1$ according to $\M$. In fact, if $\A_1$ is an ABox over $\Sigma_1$ that is consistent with $\T_1$ and $q$ is a over $\Sigma_2$, then it could be the case that $\cert(q, \tup{\T_1 \cup \T_{12}, \A_1}) \subsetneq \cert(q, \tup{\T_2,\A^\star_2})$ for some $\UCQ$-solution $\A^\star_2$ for $\A_1$ under $\M$. However, the extra tuples extracted by query $q$ are obtained from the extra information in $\A^\star_2$, as if we consider a tuple $\vec{a}$ that belong to $\cert(q, \tup{\T_2,\A_2})$ for every $\UCQ$-solution $\A_2$ for $\A_1$ under $\M$, then it holds that $\vec{a} \in \cert(q, \tup{\T_1 \cup \T_{12}, \A_1})$. \[exa-ucq-rep\] Assume that $\M = (\{\concept{F}, \concept{G}, \concept{H}, \concept{D}\}$, $\{\concept{F'}, \concept{G'}, \concept{H'}\}$, $\T_{12})$, where $\T_{12} = \{F \ISA F', G \ISA G', H \ISA H'\}$, and let $\T_1 = \{F \ISA G\}$. As expected, TBox $\T_2 = \{ F' \ISA G'\}$ is a $\UCQ$-representation of $\T_1$ under $\M$. Moreover, we can add the inclusion $D \ISA \NOT H'$ to $\T_{12}$, and $\T_2$ will still remain a -representation of $\T_1$ under $\M$. Notice that in this latter setting, our definition has to deal with some ABoxes $\A_1$ that are consistent with $\T_1$ but not with $\T_1 \cup \T_{12}$, for instance $\A_1 = \{H(a), D(a)\}$ for some constant $a$. In those cases, Equation  is trivially satisfied, since $\Mod(\tup{\T_1 \cup \T_{12}, \A_1}) = \emptyset$ and the set of $\UCQ$-solutions for $\A_1$ under $\M$ is empty. On the problem of computing solutions {#sec-problems} ------------------------------------- Arguably, the most important problem in knowledge exchange [@ArPR11; @ABCRS12], as well as in data exchange [@FKMP05; @K05], is the task of computing a translation of a KB according to a mapping. To study the computational complexity of this task for the different notions of solutions presented in the previous section, we introduce the following decision problems. The *membership* problem for universal solutions (resp. universal -solutions) has as input a mapping $\M = (\Sigma_1, \Sigma_2, \T_{12})$ and KBs $\K_1$, $\K_2$ over $\Sigma_1$ and $\Sigma_2$, respectively. Then the question to answer is whether $\K_2$ is a universal solution (resp. universal -solution) for $\K_1$ under $\M$. Moreover, the membership problem for $\UCQ$-representations has as input a mapping $\M = (\Sigma_1, \Sigma_2, \T_{12})$ and TBoxes $\T_1$, $\T_2$ over $\Sigma_1$ and $\Sigma_2$, respectively, and the question to answer is whether $\T_2$ is a $\UCQ$-representation of $\T_1$ under $\M$. In our study, we cannot leave aside the existential versions of the previous problems, which are directly related with the problem of computing translations of a KB according to a mapping. Formally, the *non-emptiness* problem for universal solutions (resp. universal $\UCQ$-solutions) has as input a mapping $\M = (\Sigma_1, \Sigma_2, \T_{12})$ and a KB $\K_1$ over $\Sigma_1$. Then the question to answer is whether there exists a universal solution (resp. universal -solution) for $\K_1$ under $\M$. Moreover, the non-emptiness problem for $\UCQ$-representations has as input a mapping $\M = (\Sigma_1, \Sigma_2, \T_{12})$ and a TBox $\T_1$ over $\Sigma_1$, and the question to answer is whether there exists a $\UCQ$-representation of $\T_1$ under $\M$. [cc]{} ABoxes extended ABoxes -- -------- ----------------- in -complete & ABoxes extended ABoxes -- -------- ----------------- in -hard, in Our contributions {#sec-cont} ================= In Section \[sec-problems\], we have introduced the problems that are studied in this paper. It is important to notice that these problems are defined by considering only KBs (as opposed to extended KBs), as they are the formal counterpart of . Nevertheless, as shown in Section \[sec-univ\], there are natural examples of specifications and mappings where null values are needed when constructing solutions. Thus, we also study the problems defined in Section \[sec-problems\] in the case where translations can be extended KBs. It should be noticed that the notions of solution, universal solution, $\UCQ$-solution, universal $\UCQ$-solution, and $\UCQ$-representation have to be enlarged to consider extended KBs, which is straightforward to do. In particular, given a mapping $\M = (\Sigma_1, \Sigma_2, \T_{12})$ and TBoxes $\T_1$, $\T_2$ over $\Sigma_1$ and $\Sigma_2$, respectively, $\T_2$ is said to be a $\UCQ$-representation of $\T_1$ under $\M$ in this extended setting if in Equation , $\A_2$ is an extended ABox over $\Sigma_2$ that is a $\UCQ$-solution for $\A_1$ under $\M$. The main contribution of this paper is to provide a detailed analysis of the complexity of the membership and non-emptiness problems for the notions of universal solution and $\UCQ$-representation. In Figure \[fig-results\], we provide a summary of the main results in the paper, which are explained in more detail in Sections \[sec-univ\] and \[sec-ucq-rep\]. It is important to notice that these results considerably extend the previous known results about these problems [@ABCRS12; @ABCRS12b]. In the first place, the problem of computing universal solutions was studied in [@ABCRS12] for the case of $\dlliterdfs$, a fragment of $\dlliter$ that allows neither for inclusions of the form $B \ISA \SOMET{R}$ nor for disjointness assertions. In that case, it is straightforward to show that every source KB has a universal solution that can be computed by using the chase procedure [@CDLLR07]. Unfortunately, this result does not provide any information about how to solve the much larger case considered in this paper, where, in particular, the non-emptiness problem is not trivial. In fact, for the case of the notion of universal solution, all the lower and upper bounds provided in Figure \[fig-results\] are new results, which are not consequences of the results obtained in [@ABCRS12]. In the second place, a notion of -representation that is appropriate for the fragment of not including disjointness assertions was studied in [@ABCRS12; @ABCRS12b]. In particular, it was shown that the membership and non-emptiness problems for this notion are solvable in polynomial time. In this paper, we considerably strengthen these results: by generalizing the definition of the notion of $\UCQ$-representation to be able to deal with , that is, with the entire language $\dlliter$ (which includes disjointness assertions); and by showing that the membership and non-emptiness problems are both $\NLOGSPACE$-complete in this larger scenario. It turns out that reasoning about universal $\UCQ$-solutions is much more intricate. In fact, as a second contribution of our paper, we provide a $\PSPACE$ lower bound for the complexity of the membership problem for the notion of universal $\UCQ$-solution, which is in sharp contrast with the $\NP$ and $\NLOGSPACE$ upper bounds for this problem for the case of universal solutions and $\UCQ$-representations, respectively (see Figure \[fig-results\]). Although many questions about universal $\UCQ$-solutions remain open, we think that this is an interesting first result, as universal $\UCQ$-solutions have only been investigated before for the very restricted fragment $\dlliterdfs$ of $\dlliter$ [@ABCRS12], which is described in the previous paragraph. Computing universal solutions {#sec-univ} ============================= In this section, we study the membership and non-emptiness problems for universal solutions, in the cases where nulls are not allowed (Section \[sec-univ-owlql\]) and are allowed (Section \[sec-non-emp-uni-sol\]) in such solutions. But before going into this, we give an example that shows the shape of universal solutions in . \[exa-null\] Assume that $\M = (\{\concept{F}, \role{S}\}, \{\concept{G'}\}$, $\{\SOMET{S^-} \ISA G'\})$, and let $\K_1 = \tup{\T_1, \A_1}$, where $\T_1 = \{F \ISA \SOMET{S}\}$ and $\A_1 = \{F(a)\}$. Then a natural way to construct a universal solution for $\K_1$ under $\M$ is to ‘populate’ the target with all implied facts (as it is usually done in data exchange). Thus, the ABox $\A_2 = \{G'(n)\}$, where $n$ is a labeled null, is a universal solution for $\K_1$ under $\M$ if nulls are allowed. Notice that here, a universal solution with non-extended ABoxes does not exist: substituting $n$ by any constant is too restrictive, ruining universality. Now, assume $\M = (\{\concept{F}, \role{S}, \role{T}\}$, $\{\role{S'}\}$, $\{S \ISA S', T \ISA S'\})$, and $\K_1 = \tup{\T_1, \A_1}$, where $\T_1 = \{F \ISA \SOMET{S}, \SOMET{S^-} \ISA \SOMET{S}\}$ and $\A_1 = \{F(a)$, $T(a,a)\}$. In this case, we cannot use the same approach as in Example \[exa-null\] to construct a universal solution, as now we would need of an infinite number of labeled nulls to construct such a solution. However, as $S$ and $T$ are transferred to the same role $S'$, it is possible to use constant $a$ to represent all implied facts. In particular, in this case $\A_2 = \{S'(a,a)\}$ is a universal solution for $\K_1$ under $\M$. Universal solutions without null values {#sec-univ-owlql} --------------------------------------- We explain here how the $\NP$ upper bound for the non-emptiness problem for universal solutions is obtained, when ABoxes are not allowed to contain null values. Assume given a mapping $\M = (\Sigma_1$, $\Sigma_2$, $\T_{12})$ and a KB $\K_1 = \tup{\T_1, \A_1}$ over $\Sigma_1$. To check whether $\K_1$ has a universal solution under $\M$, we use the following non-deterministic polynomial-time algorithm. First, we construct an ABox $\A_2$ over $\Sigma_2$ containing every membership assertion $\alpha$ such that $\tup{\T_1 \cup \T_{12}, \A_1} \models \alpha$, where $\alpha$ is of the form either $B(a)$ or $R(a,b)$, and $a,b$ are constants mentioned in $\A_1$. Second, we guess an interpretation $\I$ of $\Sigma_1$ such that $\I \models \K_1$ and $(\I, \Uni_{\A_2}) \models \T_{12}$, where $\Uni_{\A_2}$ is the interpretation of $\Sigma_2$ naturally corresponding[^2] to $\A_2$. The correctness of the algorithm is a consequence of the facts that: - there exists a universal solution for $\A_1$ under $\M$ if and only if $\A_2$ is a solution for $\A_1$ under $\M$; and - $\A_2$ is a solution for $\A_1$ under $\M$ if and only if there exists a model $\I$ of $\K_1$ such that $(\I, \Uni_{\A_2}) \models \T_{12}$. Moreover, the algorithm can be implemented in a non-deterministic polynomial-time Turing machine given that: $\A_2$ can be constructed in polynomial time; if there exists a model $\I$ of $\K_1$ such that $(\I, \Uni_{\A_2}) \models \T_{12}$, then there exists a model of $\K_1$ of polynomial-size satisfying this condition; and it can be checked in polynomial time whether $\I \models \K_1$ and $(\I, \Uni_{\A_2}) \models \T_{12}$. In addition, in this case, the membership problem can be reduced to the non-emptiness problem, thus, we have that: \[the:non-emp-uni-sol-owlql-aboxes-np\] The non-emptiness and membership problems for universal solutions are in . The exact complexity of these problems remains open. In fact, we conjecture that these problems are in . We conclude by showing that reasoning about universal $\UCQ$-solutions is harder than reasoning about universal solutions, which can be explained by the fact that TBoxes have bigger impact on the structure of universal -solutions rather than of universal solutions. In fact, by using a reduction from the validity problem for quantified Boolean formulas, similar to a reduction in [@KKLSWZ11], we are able to prove the following: \[the:memb-uni-UCQ-sol-pspace-hard\] The membership problem for universal -solutions is -hard. Universal solutions with null values {#sec-non-emp-uni-sol} ------------------------------------ We start by considering the non-emptiness problem for universal solutions with null values, that is, when extended ABoxes are allowed in universal solutions. As our first result, similar to the reduction above, we show that this problem is -hard, and identify the inclusion of inverse roles as one of the main sources of complexity. To obtain an upper bound for this problem, we use *two-way alternating automata on infinite trees (2ATA)*, which are a generalization of nondeterministic automata on infinite trees [@Vard98] well suited for handling inverse roles in . More precisely, given a KB $\K$, we first show that it is possible to construct the following automata: - $\acan$ is a 2ATA that accepts trees corresponding to the canonical model of $\K$ [^3] with nodes arbitrary labeled with a special symbol $G$; - $\asol$ is a 2ATA that accepts a tree if its subtree labeled with $G$ corresponds to a tree model $\I$ of $\K$ (that is, a model forming a tree on the labeled nulls); and - $\agfin$ is a (one-way) non-deterministic automaton that accepts a tree if it has a finite prefix where each node is marked with $G$, and no other node in the tree is marked with $G$. Then to verify whether a KB $\K_1 = \tup{\T_1, \A_1}$ has a universal solution under a mapping $\M=(\Sigma_1, \Sigma_2, \T_{12})$, we solve the non-emptiness problem for an automaton $\Bau$ defined as the product automaton of $\pi_{\Gamma_{\K}}(\acan)$, $\pi_{\Gamma_\K}(\asol)$ and $\agfin$, where $\K=\tup{\T_1 \cup \T_{12}, \A_1}$, $\pi_{\Gamma_\K}(\acan)$ is the projection of $\acan$ on a vocabulary $\Gamma_\K$ not mentioning symbols from $\Sigma_1$, and likewise for $\pi_{\Gamma_\K}(\asol)$. If the language accepted by $\Bau$ is empty, then there is no universal solution for $\K_1$ under $\M$, otherwise a universal solution (possibly of exponential size) exists, and we can compute it by extracting the ABox encoded in some tree accepted by $\Bau$ . Summing up, we get: \[the:non-emp-uni-sol-pspace-hard-and-exptime\] If extended ABoxes are allowed in universal solutions, then the non-emptiness problem for universal solutions is -hard and in . Interestingly, the membership problem can be solved more efficiently in this scenario, as now the candidate universal solutions are part of the input. In the following theorem, we pinpoint the exact complexity of this problem. \[the:memb-uni-sol-npcomplete\] If extended ABoxes are allowed in universal solutions, then the membership problem for universal solutions is -complete. Computing [$\UCQ$]{}-representations {#sec-ucq-rep} ==================================== In Section \[sec-univ\], we show that the complexity of the membership and non-emptiness problems for universal solutions differ depending on whether ABoxes or extended ABoxes are considered. On the other hand, we show in the following proposition that the use of null values in ABoxes does not make any difference in the case of $\UCQ$-representations. In this proposition, given a mapping $\M$ and TBoxes $\T_1$, $\T_2$, we say that $\T_2$ is a $\UCQ$-representation of $\T_1$ under $\M$ [*considering extended ABoxes*]{} if $\T_1$, $\T_2$, $\M$ satisfy Equation  in Section \[sec-sol\], but assuming that $\A_2$ is an extended ABox over $\Sigma_2$ that is a $\UCQ$-solution for $\A_1$ under $\M$. A TBox $\T_2$ is a $\UCQ$-representation of a TBox $\T_1$ under a mapping $\M$ if and only if $\T_2$ is a $\UCQ$-representation of $\T_1$ under $\M$ considering extended ABoxes. Thus, from now on we study the membership and non-emptiness problems for $\UCQ$-representations assuming that ABoxes can contain null values. We start by considering the membership problem for $\UCQ$-representations. In this case, one can immediately notice some similarities between this task and the membership problem for universal $\UCQ$-solutions, which was shown to be $\PSPACE$-hard in Theorem \[the:memb-uni-UCQ-sol-pspace-hard\]. However, the universal quantification over ABoxes in the definition of the notion of $\UCQ$-representation makes the latter problem computationally simpler, which is illustrated by the following example. \[ex:repres-pieces\] Assume that $\M = (\Sigma_1, \Sigma_2, \T_{12})$, where $\Sigma_1= \{ \concept{F}, \role{S_1}, \role{S_2}, \role{T_1}, \role{T_2}\}$, $\Sigma_2 = \{ \concept{F'}, \role{S'}, \role{T'}, \concept{G'}\}$ and $\T_{12} = \{ F \ISA F', S_1 \ISA S'$, $S_2 \ISA S', T_1 \ISA T', T_2 \ISA T', \SOMET{T_1^-} \ISA G'\}$. Moreover, assume that $\T_1 = \{F \ISA \SOMET{S_1}, F \ISA \SOMET{S_2}, \SOMET{S_1^-} \ISA \SOMET{T_1}, \SOMET{S_2^-} \ISA \SOMET{T_2}\}$ and $\T_2 = \{ F' \ISA \SOMET{S'}, \SOMET{{S'}^-} \ISA \SOMET{T'}, \SOMET{{T'}^-} \ISA G'\}$. If we were to verify whether $\tup{\T_2, \{F'(a)\}}$ is a universal $\UCQ$-solution for $\tup{\T_1, \{F(a)\}}$ under $\M$ (which it is in this case), then we would first need to construct the [*path*]{} $\pi= \tup{F'(a), S'(a,n), T'(n,m), G'(m)}$ formed by the inclusions in $\T_2$, where $n,m$ are fresh null values, and then we would need to explore the translations according to $\M$ of all paths formed by the inclusions in $\T_1$ to find one that matches $\pi$. On the other hand, to verify whether $\T_2$ is a $\UCQ$-representation of $\T_1$ under $\M$, one does not need to execute any “backtracking”, as it is sufficient to consider independently a polynomial number of pieces $\C$ taken from the paths formed by the inclusions in $\T_1$, each of them of polynomial size, and then checking whether the translation $\C'$ of $\C$ according to $\M$ matches with the paths formed from $\C'$ by the inclusions in $\T_2$. If any of these pieces does not satisfy this condition, then it can be transformed into a witness that Equation  is not satisfied, showing that $\T_2$ is not a $\UCQ$-representation of $\T_1$ under $\M$ (as we have a universal quantification over the ABoxes over $\Sigma_1$ in the definition of $\UCQ$-representations). In fact, one of the pieces considered in this case is $\C = \tup{T_2(n,m)}$, where $n$, $m$ are null values, which does not satisfy the previous condition as the translation $\C'$ of $\C$ according to $\M$ is $\tup{T'(n,m)}$, and this does not match with the path $\tup{T'(n,m), G'(m)}$ formed from $\C'$ by the inclusions in $\T_2$. This particular case is transformed into an ABox $\A_1=\{T_2(b, c)\}$ and a query $q = T'(b,c) \wedge G'(c)$, where $b$, $c$ are fresh constants, for which we have that Equation is not satisfied. Notice that disjointness assertions in the mapping may cause $\tup{\T_1 \cup \T_{12},\A_1}$ to become inconsistent for some source ABoxes $\A_1$ (which will make all possible tuples to be in the answer to every query), therefore additional conditions have to be imposed on $\T_2$. To give more intuition about how the membership problem for $\UCQ$-representations is solved, we give an example showing how one can deal with some of these inconsistency issues. Assume that $\M = (\Sigma_1, \Sigma_2, \T_{12})$, where $\Sigma_1 = \{ \concept{F}, \concept{G}, \concept{H} \}$, $\Sigma_2 = \{ \concept{F'}, \concept{G'}, \concept{H'}\}$ and $\T_{12}= \{F \ISA F', G \ISA G', H \ISA H'\}$. Moreover, assume that $\T_1 = \{ F \ISA G\}$ and $\T_2 =\{F' \ISA G'\}$. In this case, it is clear that $\T_2$ is a $\UCQ$-representation of $\T_1$ under $\M$. However, if we add inclusion $H \ISA \neg G'$ to $\T_{12}$, then $\T_2$ is no longer a $\UCQ$-representation of $\T_1$ under $\M$. To see why this is the case, consider an ABox $\A_1=\{ F(a), H(a)\}$, which is consistent with $\T_1$, and a query $q=F'(b)$, where $b$ is a fresh constant. Then we have that $\cert(q, \tup{\T_1 \cup \T_{12}, \A_1}) =\{()\}$ as KB $\tup{\T_1 \cup \T_{12}, \A_1}$ is inconsistent, while $\cert(q,\tup{\T_2, \A_2}) =\emptyset$ for $\UCQ$-solution $\A_2 = \{F'(a),H'(a)\}$ for $\A_1$ under $\M$. Thus, we conclude that Equation  is violated in this case. One can deal with the issue raised in the previous example by checking that on every pair $(B,B')$ of $\T_1$-consistent basic concepts over $\Sigma_1$,[^4] it holds that: $(B,B')$ is $(\T_1 \cup \T_{12})$-consistent if and only if $(B,B')$ is $(\T_{12} \cup \T_{2})$-consistent, and likewise for every pair of basic roles over $\Sigma_1$. This condition guarantees that for every ABox $\A_1$ over $\Sigma_1$ that is consistent with $\T_1$, it holds that: $\tup{\T_1 \cup \T_{12}, \A_1}$ is consistent if and only if there exists an extended ABox $\A_2$ over $\Sigma_2$ such that $\A_2$ is a $\UCQ$-solution for $\A_1$ under $\M$ and $\tup{\T_2, \A_2}$ is consistent. Thus, the previous condition ensures that the sets on the left- and right-hand side of Equation  coincide whenever the intersection on either of these sides is taken over an empty set. The following theorem, which requires of a lengthy and non-trivial proof, shows that there exists an efficient algorithm for the membership problem for $\UCQ$-representations that can deal with all the aforementioned issues. \[eq:memb-repr-nl\] The membership problem for $\UCQ$-representations is -complete. We conclude by pointing out that the non-emptiness problem for $\UCQ$-representations can also be solved efficiently. We give an intuition of how this can be done in the following example, where we say that $\T_1$ is *-representable under* $\M$ if there exists a -representation $\T_2$ of $\T_1$ under $\M$. \[ex:repres-non-emp\] Assume that $\M = (\Sigma_1, \Sigma_2, \T_{12})$, where $\Sigma_1 = \{ \concept{F}, \concept{G}, \concept{H} \}$, $\Sigma_2 = \{ \concept{F'}, \concept{G'}\}$ and $\T_{12}= \{F \ISA F', G \ISA G', H \ISA F'\}$. Moreover, assume that $\T_1 = \{ F \ISA G\}$. Then it follows that $\T_1 \cup \T_{12} \models F \ISA G'$, and in order for $\T_1$ to be $\UCQ$-representable under $\M$, the following condition must be satisfied: - there exists a concept $B'$ over $\Sigma_2$ s.t. $\T_{12} \models F \ISA B'$, and for each concept $B$ over $\Sigma_1$ with $\T_1 \cup \T_{12} \models B \ISA B'$ it follows that $\T_1 \cup \T_{12} \models B \ISA G'$. The idea is then to add the inclusion $B' \ISA G'$ to a -representation $\T_2$ so that $\T_{12} \cup \T_{2} \models F \ISA G'$ as well. In our case, concept $F'$ satisfies the condition $\T_{12} \models F \ISA F'$, but it does not satisfy the second requirement as $\T_1 \cup \T_{12} \models H \ISA F'$ and $\T_1 \cup \T_{12}\not\models H\ISA G'$. In fact, $F' \ISA G'$ cannot be added to $\T_2$ as it would result in $\T_{12} \cup \T_2 \models H \ISA G'$, hence in Equation , the inclusion from right to left would be violated. There is no way to reflect the inclusion $F \ISA G'$ in the target, so in this case . The proof of the following result requires of some involved extensions of the techniques used to prove Theorem \[eq:memb-repr-nl\]. \[eq:ne-repr-nl\] The non-emptiness problem for $\UCQ$-representations is -complete. The techniques used to prove Theorem \[eq:ne-repr-nl\], which is sketched in the example below. Consider $\M$ and $\T_1$ from Example \[ex:repres-non-emp\], but assuming that $\T_{12}$ does not contain the inclusion $H \ISA F'$. Again, $\T_1 \cup \T_{12} \models F \ISA G'$, but now condition $(\star)$ is satisfied. Then, an algorithm for computing a representation essentially needs to take any $B'$ given by condition ($\star$) and add the inclusion $B' \ISA F'$ to $\T_2$. In this case, $\T_2 = \{F' \ISA G'\}$ is a -representation of $\T_1$ under $\M$. Conclusions {#sec-conclusions} =========== In this paper, we have studied the problem of KB exchange for , improving on previously known results with respect to both the expressiveness of the ontology language and the understanding of the computational properties of the problem. Our investigation leaves open several issues, which we intend to address in the future. First, it would be good to have characterizations of classes of source KBs and mappings for which universal (-)solutions are guaranteed to exist. As for the computation of universal solutions, while we have pinned-down the complexity of membership for extended ABoxes as -complete, an exact bound for the other case is still missing. Moreover, it is easy to see that allowing for inequalities between terms (e.g., $a \neq b$ in Example \[exa-disj-source\]) and for negated atoms in the (target) ABox would allow one to obtain more universal solutions, but a full understanding of this case is still missing. Finally, we intend to investigate the challenging problem of computing universal -solutions, adopting also here an automata-based approach. Acknowledgements ================ This research has been partially supported by the EU IP project Optique (grant FP7-318338), by EU Marie Curie action FP7-PEOPLE-2009-IRSES (grant 24761), by ERC grant agreement DIADEM, no. 246858, and by Fondecyt grant 1131049. Diego Calvanese has been partially supported by the Wolfgang Pauli Institute Vienna. The authors are also grateful to Evgeny Sherkhonov and Roman Kontchakov for helpful comments and discussions. Philippe Adjiman, Philippe Chatalic, François Goasdou[é]{}, Marie-Christine Rousset, and Laurent Simon. Distributed reasoning in a peer-to-peer setting: [A]{}pplication to the [S]{}emantic [W]{}eb. , 25:269–314, 2006. Marcelo Arenas, Jorge P[é]{}rez, Juan L. Reutter, and Cristian Riveros. Composition and inversion of schema mappings. , 38(3):17–28, 2009. Marcelo Arenas, Jorge P[é]{}rez, and Juan L. Reutter. Data exchange beyond complete data. In [*PODS 2011*]{}, pages 83–94, 2011. Marcelo Arenas, Elena Botoeva, Diego Calvanese, Vladislav Ryzhikov, and Evgeny Sherkhonov. Exchanging description logic knowledge bases. In [*KR 2012*]{}, 2012. Marcelo Arenas, Elena Botoeva, Diego Calvanese, Vladislav Ryzhikov, and Evgeny Sherkhonov. Representability in *DL-Lite*$_r$ knowledge base exchange. In [*Proc. of the 25th Int. Workshop on Description Logic (DL 2012)*]{}, volume 846 of [*CEUR Electronic Workshop Proceedings, [<http://ceur-ws.org/>]{}*]{}, 2012. Alessandro Artale, Diego Calvanese, Roman Kontchakov, and Michael Zakharyaschev. The *DL-Lite* family and relations. , 36:1–69, 2009. Jie Bao et al. eb [O]{}ntology [L]{}anguage document overview (second edition). ecommendation, World Wide Web Consortium, December 2012. <http://www.w3.org/TR/owl2-overview/>. Pablo Barcel[ó]{}. Logical foundations of relational data exchange. , 38(1):49–58, 2009. Diego Calvanese, Giuseppe De Giacomo, Domenico Lembo, Maurizio Lenzerini, and Riccardo Rosati. Tractable reasoning and efficient query answering in description logics: The *DL-Lite* family. , 39(3):385–429, 2007. J[é]{}r[ô]{}me Euzenat and Pavel Shvaiko. . Springer, 2007. Ronald Fagin and Phokion G. Kolaitis. Local transformations and conjunctive-query equivalence. In [*PODS 2012*]{}, pages 179–190, 2012. Ronald Fagin, Phokion G. Kolaitis, Ren[é]{}e J. Miller, and Lucian Popa. Data exchange: [S]{}emantics and query answering. , 336(1):89–124, 2005. Ronald Fagin, Phokion G. Kolaitis, Alan Nash, and Lucian Popa. Towards a theory of schema-mapping optimization. In [*PODS 2008*]{}, pages 33–42, 2008. Ronald Fagin, Laura M. Haas, Mauricio A. Hern[á]{}ndez, Ren[é]{}e J. Miller, Lucian Popa, and Yannis Velegrakis. Clio: [S]{}chema mapping creation and data exchange. In [*Conceptual Modeling: Foundations and Applications – Essays in Honor of John Mylopoulos*]{}, volume 5600 of [*Lecture Notes in Computer Science*]{}, pages 198–236, 2009. Ariel Fuxman, Phokion G. Kolaitis, Ren[é]{}e J. Miller, and Wang Chiew Tan. Peer data exchange. , 31(4):1454–1498, 2006. Anastasios Kementsietsidis, Marcelo Arenas, and Ren[é]{}e J. Miller. Mapping data in peer-to-peer systems: Semantics and algorithmic issues. In [*Proc. of the ACM SIGMOD Int. Conf. on Management of Data*]{}, pages 325–336, 2003. Phokion G. Kolaitis. Schema mappings, data exchange, and metadata management. In [*PODS 2005*]{}, pages 61–75, 2005. Boris Konev, Roman Kontchakov, Michel Ludwig, Thomas Schneider, Frank Wolter, and Michael Zakharyaschev. Conjunctive query inseparability of [OWL 2 QL]{} [TBoxes]{}. In [*Proc. of the 25th AAAI Conference on Artificial Intelligence, (AAAI 2011)*]{}, 2011. Maurizio Lenzerini. Data integration: [A]{} theoretical perspective. In [*Proc. of the 21st ACM SIGACT SIGMOD SIGART Symp. on Principles of Database Systems (PODS 2002)*]{}, pages 233–246, 2002. Jayant Madhavan and Alon Y. Halevy. Composing mappings among data sources. In [*VLDB 2003*]{}, pages 572–583, 2003. Boris Motik, Bernardo Cuenca Grau, Ian Horrocks, Zhe Wu, Achille Fokoue, and Carsten Lutz. eb [O]{}ntology [L]{}anguage profiles (second edition). ecommendation, World Wide Web Consortium, December 2012. <http://www.w3.org/TR/owl2-profiles/>. Reinhard Pichler, Emanuel Sallinger, and Vadim Savenkov. Relaxed notions of schema mapping equivalence revisited. , 52(3):483–541, 2013. Pavel Shvaiko and J[é]{}r[ô]{}me Euzenat. Ontology matching: State of the art and future challenges. , 25(1):158–176, 2013. Moshe Y. Vardi. Reasoning about the past with two-way automata. In [*Proc. of the 25th Int. Coll. on Automata, Languages and Programming (ICALP’98)*]{}, volume 1443 of [*Lecture Notes in Computer Science*]{}, pages 628–641. Springer, 1998. 1.5cm 1.5cm -3cm [^1]: If disjointness assertions are not allowed, then this new notion can be shown to be equivalent to the original formalization of $\UCQ$-representation proposed in [@ABCRS12]. [^2]: Interpretation $\Uni_{\A_2}$ can be defined as the Herbrand model of $\A_2$ extended with fresh domain elements to satisfy assertions of the form $\SOMET{R}(a)$ in $\A_2$. [^3]: If $\K = \tup{\T,\A}$, then this model essentially corresponds to the chase of $\A$ with $\T$ (see  [@KKLSWZ11] for a formal definition). [^4]: A pair $(B, B)'$ is $\T$-consistent for a TBox $\T$, if the KB $\tup{\T, \{B(a), B'(a)\}}$ is consistent, where $a$ is an arbitrary constant.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study the relationship between two norms on the first cohomology of a hyperbolic 3-manifold: the purely topological Thurston norm and the more geometric harmonic norm. Refining recent results of Bergeron, , and Venkatesh as well as older work of Kronheimer and Mrowka, we show that these norms are roughly proportional with explicit constants depending only on the volume and injectivity radius of the hyperbolic 3-manifold itself. Moreover, we give families of examples showing that some (but not all) qualitative aspects of our estimates are sharp. Finally, we exhibit closed hyperbolic -manifolds where the Thurston norm grows exponentially in terms of the volume and yet there is a uniform lower bound on the injectivity radius.' address: - | Dept. of Math.\ Brown University\ Box 1917\ Providence, RI 02912, USA - | Dept. of Math., MC-382\ University of Illinois\ 1409 W. Green St.\ Urbana, IL 61801, USA author: - 'Jeffrey F. Brock' - 'Nathan M. Dunfield' bibliography: - 'big-IHS.bib' title: | Norms on the cohomology\ of hyperbolic 3-manifolds --- Introduction ============ Suppose $M$ is a closed oriented hyperbolic -manifold. Our main goal here is to understand the relationship between two norms on $H^1(M; \R)$: the purely topological Thurston norm and the more geometric harmonic norm. Precise definitions of these norms are given in Section \[sec:norms\], but for now here is an informal sketch. The *Thurston norm* ${\norm{\phi}_{\mathit{Th}}}$ of an integral class $\phi \in H^1(M; \R)$ measures the topological complexity of the simplest surface dual to $\phi$; it extends to all of $H^1(M;\R)$ where its unit ball is a finite-sided polytope with rational vertices. It makes sense for any -manifold, though unlike in the hyperbolic case where it is nondegenerate, there can be nontrivial $\phi$ of norm $0$. While it was introduced by Thurston in the 1970s [@Thurston1986], its roots go back to the early days of topology, to questions about the genus of knots in the -sphere, and it has been extensively studied in many contexts. Turning to geometry, as with any Riemannian manifold, the hyperbolic metric on $M$ gives a norm on $H^1(M; \R)$ which appears in the proof of the Hodge theorem. Specifically, if we identify $H^1(M; \R)$ with the space of harmonic -forms, then the *harmonic norm* ${\norm{\cdotspaced}_{L^2}}$ is the one associated with the usual inner product on the level of forms: $$\pair{\alpha, \beta} = \int_M \alpha \wedge *\beta$$ As it comes from a positive-definite inner product, the unit ball of ${\norm{\cdotspaced}_{L^2}}$ is a nice smooth ellipsoid. The harmonic norm appears, for example, in the Cheeger-Müller formula for the Ray-Singer analytic torsion of $M$ [@Cheeger1979; @Muller1978]. By Mostow rigidity, the hyperbolic metric on $M$ is unique, and so a posteriori the harmonic norm depends solely on the underlying topology of $M$. It is thus very natural to ask how these two norms are related. Kronheimer and Mrowka seem to have been the first to study this question in the more general context of arbitrary Riemannian metrics on $M$; specifically, using deep results from gauge theory, they characterized the Thurston norm as the infimum (over all possible metrics) of certain scaled harmonic norms [@KronheimerMrowka1997b]. Their results have specific consequences for any given metric including the hyperbolic one; see Section \[sec:gauge\] for a complete discussion. More recently, Bergeron, , and Venkatesh [@BergeronSengunVenkatesh2016] examined the relationship between these two norms in the case of the hyperbolic metric. There, motivated by questions about torsion growth in the homology of arithmetic groups, they proved the following beautiful result: \[prop:BSV\] Suppose $M_0$ is a closed orientable hyperbolic -manifold. There exist constants $C_1$ and $C_2$, depending on $M_0$, so that for every finite cover $M$ of $M_0$ one has $$\label{eq:norms} \frac{C_1}{\vol(M)} {\norm{\cdotspaced}_{\mathit{Th}}} \leq {\norm{\cdotspaced}_{L^2}} \leq C_2 {\norm{\cdotspaced}_{\mathit{Th}}} \mtext{on $H^1(M ; \R)$.}$$ In fact, their proof immediately gives that the constants $C_1$ and $C_2$ depend only on a lower bound on the injectivity radius $\inj( M_0)$, which is half the length of the shortest closed geodesic. Our main result is the following refinement of Theorem \[prop:BSV\]: [theorem]{}[thmmain]{}\[thm:main\] For all closed orientable hyperbolic -manifolds $M$ one has $$\label{eq:main} \frac{\pi}{\sqrt{\vol(M)}} {\norm{\cdotspaced}_{\mathit{Th}}} \leq {\norm{\cdotspaced}_{L^2}} \leq \frac{10 \pi}{\sqrt{\inj(M)}} {\norm{\cdotspaced}_{\mathit{Th}}} \mtext{on $H^1(M ; \R)$.}$$ We also give families of examples which show that some (but not all) qualitative aspects of Theorem \[thm:main\] are sharp. The first proves that the basic form of the first inequality in (\[eq:main\]) cannot be improved: [theorem]{}[thmlowerexs]{}\[thm:lowerexs\] There exists a sequence of $M_n$ and $\phi_n \in H^1(M_n; \R)$ so that 1. The quantities $\vol(M_n)$ and $\inj(M_n)$ both go to infinity as $n$ does. 2. The ratio $\frac{{\norm{\phi_n}_{\mathit{Th}}}}{{\norm{\phi_n}_{L^2}} \sqrt{\vol(M_n)}}$ is constant. The next result concerns the second inequality of (\[eq:main\]), and shows that the harmonic norm can blow up relative to the Thurston norm when the injectivity radius gets small. [theorem]{}[thminjsmallexs]{}\[thm:injsmallexs\] There exists a sequence of $M_n$ and $\phi_n \in H^1(M_n; \R)$ so that 1. The volumes of the $M_n$ are uniformly bounded and $\inj(M_n) \to 0$ as $n \to \infty$. 2. ${\norm{\phi_n}_{L^2}} \big/ {\norm{\phi_n}_{\mathit{Th}}} \to \infty$ like $\sqrt{-\log\big(\inj(M_n)\big)}$ as $n \to \infty$. The growth of ${\norm{\phi_n}_{L^2}} \big/ {\norm{\phi_n}_{\mathit{Th}}}$ in Theorem \[thm:injsmallexs\] is much slower than the most extreme behavior permitted by the second inequality in (\[eq:main\]). We suspect that the examples in Theorem \[thm:injsmallexs\] have the worst possible behavior, but we are unable to improve (\[eq:main\]) in that direction and believe doing so requires an entirely new approach. Manifolds with large norms -------------------------- A very intriguing conjecture of [@BergeronSengunVenkatesh2016] is that for congruence covers of a fixed *arithmetic* hyperbolic -manifold $M_0$, the size of the Thurston norm grows more slowly than one would naively expect. Specifically, any cover $M$ of $M_0$ should have a nontrivial $\phi \in H^1(M; \Z)$ where ${\norm{\phi}_{\mathit{Th}}}$ is bounded by a polynomial in $\vol(M)$; in contrast, the usual estimates using a natural triangulation of $M$ give only that there is a $\phi$ with ${\norm{\phi}_{\mathit{Th}}}$ is at most exponential in $\vol(M)$. More generally, such estimates give that for any $\epsilon > 0$ there is a contant $C$ so that any closed hyperbolic -manifold $M$ with $\inj(M) > \epsilon$ and $H^1(M; \Z) \neq 0$ has a nontrivial $\phi \in H^1(M; \Z)$ where ${\norm{\phi}_{\mathit{Th}}} < C^{\vol(M)}$. Our other contribution here is to show that such a priori estimates on the size of the Thurston norm cannot be substantially improved. Specifically, we give examples of closed hyperbolic -manifolds where the Thurston norm is exponentially large and yet there is a uniform lower bound on the injectivity radius. [theorem]{}[thmlargenorm]{}\[thm:largenorm\] There exist constants $C_1, C_2, \epsilon_1 > 0$ and a sequence of closed hyperbolic -manifolds $M_n$ with $\vol(M_n) \to \infty$ where for all $n$: 1. $\inj(M_n) > \epsilon_1$, 2. $b_1(M_n) = 1$, 3. ${\norm{\phi_n}_{\mathit{Th}}} > C_1 e^{C_2 \vol(M_n)}$ where $\phi_n$ is a generator of $H^1(M_n; \Z)$. \[item:expgrowth\] Presumably, though we do not show this, the examples in Theorem \[thm:largenorm\] are nonarithmetic, and so point toward a divergence in behavior of the regulator term in the analytic torsion in the arithmetic and nonarithmetic cases, which is consistent with e.g. the experiments of [@BrockDunfield2015 §4]. Proof highlights ---------------- The arguments for the two inequalities of (\[eq:main\]) in Theorem \[thm:main\] are mostly independent. For the first inequality, we take a quite different approach from that of [@BergeronSengunVenkatesh2016]. Namely, to mediate between the topological Thurston norm and the analytical harmonic norm, we study what we call the *least-area norm*. For an integral class $\phi \in H^1(M; \Z)$, this norm is simply the least area of any embedded smooth surface dual to $\phi$. Using results on minimal surfaces of Schoen and Uhlenbeck, we show this new norm is uniformly comparable to the Thurston norm in the context of hyperbolic -manifolds (Theorem \[thm:areavsthurston\]). Also, an argument using the coarea formula in geometric measure theory allows us to reinterpret the least-area norm as the $L^1$-analog of the $L^2$-based harmonic norm (Lemma \[lem:areavsone\]). With these connections in place, the first inequality of (\[eq:main\]) boils down to simply the Cauchy-Schwarz inequality. In contrast, our proof of the second inequality in (\[eq:main\]) follows the approach of [@BergeronSengunVenkatesh2016] closely. The key improvement is a refined upper bound on the $L^\infty$-norm of a harmonic -form in terms of its harmonic norm. The latter result is Theorem \[thm:nosobolev\] and is proved by a detailed analysis of the natural series expansion of a harmonic -form about a point in $\H^3$. For the examples, Theorem \[thm:lowerexs\] uses a simple construction involving finite covers, and Theorem \[thm:injsmallexs\] is based on Dehn filling a suitable 2-cusped hyperbolic -manifold. Finally, the proof of Theorem \[thm:largenorm\] involves gluing together two acylindrical homology handlebodies by a large power of a pseudo-Anosov; as with our prior work on integer homology spheres with large injectivity radius [@BrockDunfield2015], the key is controlling the homology of the resulting manifolds. Acknowledgements ---------------- Brock was partially supported by US NSF grant DMS-1207572 and the GEAR Network (US NSF grant DMS-1107452), and this work was partially conducted at ICERM. Dunfield was partially supported by US NSF grant DMS-1106476, the GEAR Network, a Simons Fellowship, and this work was partially completed while visiting ICERM, the University of Melbourne, and the Institute for Advanced Study. We gratefully thank Nicolas Bergeron, Haluk , Akshay Venkatesh, Anil Hirani, Mike Freedman, Tom Church, Tom Mrowka, and Peter Kronheimer for helpful conversations and comments. Finally, we thank the several referees for their helpful comments and suggestions. The harmonic and Thurston norms {#sec:norms} =============================== Conventions ----------- Throughout this paper, all manifolds of any dimension are orientable and moreover oriented. All cohomology will have $\R$ coefficients unless noted otherwise. The harmonic norm ----------------- For a closed Riemannian -manifold $M$, the natural positivedefinite inner product on the space $\Omega^k(M)$ of real-valued $k$-forms is given by $$\pair{\alpha, \beta} = \int_M \alpha \wedge *\beta \mtext{where $*$ is the Hodge star operator.}$$ The *harmonic representative* of a class $[\alpha] \in H^k(M)$ is the unique one that minimizes $\pair{\alpha, \alpha}$; equivalently, if $\Delta = d d^* + d^* d$ is the Hodge Laplacian, it is the representative where $\Delta \alpha = 0$. Thus the cohomology $H^k(M)$ inherits the above inner product via the identification of it with the subspace of harmonic forms. The corresponding norm on $H^k(M)$ is, by definition, the *harmonic norm* ${\norm{\cdotspaced}_{L^2}}$ discussed in the introduction. Equivalently, it is defined by $$\label{eq:harnormalt} {\norm{\phi}_{L^2}} = \inf \setdef{ {\norm{\alpha}_{L^2}}}{\mbox{$\alpha \in \Omega^k(M)$ represents $\phi$}}$$ For a 1-form $\alpha$ on $M$, a useful geometric viewpoint on ${\norm{\alpha}_{L^2}}$ is the following. For a point $p$ in $M$, denote the operator norm of the linear functional $\alpha_p \maps T_p M \to \R$ by $\abs{\alpha_p}$; equivalently $\abs{\alpha_p} = \sqrt{*\big(\alpha_p \wedge * \alpha_p\big)}$ which is also just the length of $\alpha_p$ under the metricinduced isomorphism of $T^p M \to T_p M$. The harmonic norm of $\alpha$ is then simply the $L^2$-norm of the associated function $\abs{\alpha} \maps M \to \R_{\geq 0}$ since $$\label{eq:L2asOp} {\norm{\alpha}_{L^2}} = \sqrt{\int_M \alpha \wedge *\alpha} = \sqrt{\int_M \abs{\alpha}^2 \ {\mathit{dVol}}}$$ Analogously, we define the $L^1$- and $L^\infty$-norms of the 1-form $\alpha$ as $$\label{eq:formnorms} {\norm{\alpha}_{L^1}} = \int_M \abs{\alpha} \ {\mathit{dVol}}\mtext{and} {\norm{\alpha}_{L^\infty}} = \max_{p \in M} \abs{\alpha_p}.$$ The Thurston norm ----------------- For a connected surface, define $\eulerminus{S} = \max\big( -\chi(S), 0 \big)$; extend this to all surfaces via $\eulerminus{S \sqcup S'} = \eulerminus{S} + \eulerminus{S'}$. For a compact irreducible -manifold $M$, the Thurston norm of $\phi \in H^1(M; \Z) \cong H_2(M, \partial M; \Z)$ is defined by $${\norm{ \phi }_{\mathit{Th}}} = \min \setdef{ \eulerminus{S} }{\mbox{$S$ is a properly embedded surface dual to $\phi$}}$$ The Thurston norm extends by continuity to all of $H^1(M)$, and the resulting unit ball is a finite-sided rational polytope [@Thurston1986]. When $M$ is hyperbolic, the Thurston norm is non-degenerate with ${\norm{\phi}_{\mathit{Th}}} > 0$ for all nonzero $\phi$; in general, it is only a seminorm. The least area norm =================== To mediate between the harmonic and Thurston norms, we introduce two additional norms on the first cohomology of a closed Riemannian -manifold $M$, namely the least-area norm and the $L^1$-norm. In fact, these two norms coincide, but the differing perspectives they offer are a key tool used to prove the lower bound in Theorem \[thm:main\]. For $\phi$ in $H^1(M; \Z)$, let $\cF_\phi$ be the collection of smooth maps $f \maps S \to M$ where $S$ is a closed oriented surface with $f_*([S])$ dual to $\phi$. The *least area norm* of $\phi$ is $${\norm{\phi}_{\mathit{LA}}} = \inf \setdef{\Area\left(f(S)\right)}{f \in \cF_\phi}$$ By standard results in geometric measure theory, the value ${\norm{\phi}_{\mathit{LA}}}$ is always realized by a smooth *embedded* surface $S \subset M$, whose components may be weighted by integer multiplicities; see e.g. [@Hass1988 Lemma 2.1] for details. We will show below that ${\norm{\cdotspaced}_{\mathit{LA}}}$ is a seminorm on $H^1(M; \Z)$ which extends continuously to a seminorm on all of $H^1(M)$. In analogy with (\[eq:harnormalt\]), we use the $L^1$-norm on 1-forms given in (\[eq:formnorms\]) to define the following function on $H^1(M)$: $${\norm{\phi}_{L^1}} = \inf \setdef{ {\norm{\alpha}_{L^1}}}{\mbox{$\alpha \in \Omega^1(M)$ represents $\phi$}}$$ Unlike the harmonic norm, the value ${\norm{\phi}_{L^1}}$ is typically not realized by any smooth form $\alpha$. Despite this, it is easy to show that ${\norm{\cdotspaced}_{L^1}}$ is a seminorm on $H^1(M)$. As promised, these two new norms are in fact the same: \[lem:areavsone\] ${\norm{\phi}_{\mathit{LA}}} = {\norm{\phi}_{L^1}}$ for all $\phi \in H^1(M; \Z)$. Note that one consequence of Lemma \[lem:areavsone\] is the promised fact that ${\norm{\cdotspaced}_{\mathit{LA}}}$ extends continuously from $H^1(M; \Z)$ to a seminorm on all of $H^1(M)$. To show ${\norm{\phi}_{\mathit{LA}}} \geq {\norm{\phi}_{L^1}}$, let $S$ be a smooth embedded surface dual to $\phi$ of area ${\norm{\phi}_{\mathit{LA}}}$. For each $\epsilon > 0$, consider a dual 1-form $\alpha_{\epsilon}$ which is supported in an $\epsilon$neighborhood of $S$ and is a slight smoothing of $\frac{1}{2 \epsilon} d \big(\mbox{signed distance to S}\big)$ there. An easy calculation in Fermi coordinates shows that ${\norm{\alpha_{\epsilon}}_{L^1}} \to \Area(S) ={\norm{\phi}_{\mathit{LA}}}$ as $\epsilon \to 0$, giving ${\norm{\phi}_{\mathit{LA}}} \geq {\norm{\phi}_{L^1}}$. To establish ${\norm{\phi}_{\mathit{LA}}} \leq {\norm{\phi}_{L^1}}$, let $\alpha$ be any representative of $\phi$. Since $\phi$ is an integral class, by integrating $\alpha$ we get a smooth map $f \maps M \to S^1$ so that $\alpha = f^*({\mathit{dt}})$, where we have parameterized $S^1 = \R/\Z$ by $t \in [0, 1]$. For almost all $t \in [0, 1]$, the set $S_t = f^{-1}(t)$ is a smooth surface. For all $t$, we define $\Area(S_t)$ to be the 2-dimensional Hausdorff measure of $S_t$. As the operator norm $\abs{\alpha}$ is equal to the 1-Jacobian of the map $f$, the Coarea Formula [@Morgan2009 Theorem 3.8] is precisely $$\label{eq:areas} \int_M \abs{\alpha} \ {\mathit{dVol}}= \int_0^1 \Area(S_t) \ {\mathit{dt}}$$ Since a function on $[0, 1]$ is less than or equal to its mean on a set of positive measure, there are many $t$ so that $S_t$ is smooth and $\Area(S_t) \leq {\norm{\alpha}_{L^1}}$. Taking the infimum over representatives $\alpha$ of $\phi$ gives ${\norm{\phi}_{\mathit{LA}}} \leq {\norm{\phi}_{L^1}}$ as desired. Relationship with the Thurston norm ----------------------------------- When $M$ is hyperbolic, the least area norm is very closely related to the Thurston norm: \[thm:areavsthurston\] For any closed hyperbolic -manifold $M$ and $\phi \in H^1(M )$ one has: $$\label{eq:thurvsone} \pi {\norm{\phi}_{\mathit{Th}}} \leq {\norm{\phi}_{\mathit{LA}}} \leq 2 \pi {\norm{\phi}_{\mathit{Th}}}$$ The moral behind this result is that *stable* minimal surfaces in hyperbolic -manifolds have uniformly bounded intrinsic curvature [@Schoen1983], and hence area and genus are essentially proportionate. Specifically, we will use the following fact, which was first observed by Uhlenbeck [@Uhlenbeck1980] in unpublished work. \[lem:stable\] For any stable closed minimal surface $S$ in a hyperbolic -manifold: $$\label{eq:stable} \pi \eulerminus{S} \leq \Area(S) \leq 2 \pi \eulerminus{S}.$$ The proof is essentially the same as [@Hass1995 Lemma 6], which you should see for details. As $S$ is minimal, its intrinsic curvature $K \maps S \to \R$ is bounded above by that of hyperbolic space, i.e. by $-1$. In particular, by Gauss-Bonnet, every component of $S$ has negative Euler characteristic and moreover $$2 \pi \eulerminus{S} = - 2 \pi \chi (S) = \int_S -K \ {\mathit{dA}}\geq \int_S 1 \ {\mathit{dA}}= \Area(S)$$ giving the righthand inequality in (\[eq:stable\]). For the other inequality, since $S$ is stable, the main argument in [@Hass1995 Lemma 6] with the test function $f=1$ gives that $\pi \eulerminus{S} \leq \Area(S)$ as desired. Pick a surface $S$ dual to $\phi$ which is incompressible and realizes the Thurston norm, i.e. ${\norm{\phi}_{\mathit{Th}}} = \eulerminus{S}$. Since $S$ is incompressible, by [@FreedmanHassScott1983] we can assume that $S$ has least area in its isotopy class and hence is a stable minimal surface. Thus by Lemma \[lem:stable\] we have $${\norm{\phi}_{\mathit{LA}}} \leq \Area(S) \leq 2 \pi \eulerminus{S} = 2 \pi {\norm{\phi}_{\mathit{Th}}}.$$ proving the second half of (\[eq:thurvsone\]). For the other inequality, suppose $S$ is a least area surface dual to $\phi$. Note again that $S$ must be a stable minimal surface, and so Lemma \[lem:stable\] gives $$\pi {\norm{\phi}_{\mathit{Th}}} \leq \pi \eulerminus{S} \leq \Area(S) = {\norm{\phi}_{\mathit{LA}}}$$ proving the rest of (\[eq:thurvsone\]). Pointwise bounds on harmonic 1-forms ==================================== Let $M$ be a closed hyperbolic -manifold. A key tool used by [@BergeronSengunVenkatesh2016] in their proof of both inequalities (\[eq:norms\]) in Theorem \[prop:BSV\] is that there is a constant $C$, depending somehow on the injectivity radius of the hyperbolic -manifold, so that $${\norm{\cdotspaced}_{L^\infty}} \leq C {\norm{\cdotspaced}_{L^2}}$$ In our parallel Theorem \[thm:main\], we will use this fact only in the second inequality in (\[eq:main\]), after first refining it into: \[thm:nosobolev\] If $\alpha$ is a harmonic 1-form on a closed hyperbolic manifold $M$ then $$\label{eq:supnorm} {\norm{\alpha}_{L^\infty}} \leq \frac{5}{\sqrt{\inj(M)}} {\norm{\alpha}_{L^2}}$$ While the $5$ in (\[eq:supnorm\]) can be improved, it seems unlikely that the exponent on $\inj(M)$ can be significantly reduced. If $T_{\epsilon}$ is a tube of volume $1$ around a core geodesic of length $2 \epsilon$, then there is a harmonic 1-form $\alpha_\epsilon$ on $T_\epsilon$ (namely ${\mathit{dz}}$ in cylindrical coordinates) so that $$\frac{{\norm{\alpha_{\epsilon}}_{L^\infty}}}{{\norm{\alpha_{\epsilon}}_{L^2}}} \asymp \left( \epsilon \log( \epsilon^{-1} ) \right)^{-1/2}$$ Our proof of Theorem \[thm:nosobolev\] starts by understanding how certain harmonic -forms behave on balls in $\H^3$ via \[lem:dfsharp\] If $f \maps \H^3 \to \R$ is harmonic and $B$ is a ball of radius $r$ centered about $p$ then $$\label{eq:dfbound} \abs{{\mathit{df}}_p} \leq \frac{1}{\sqrt{\nu(r)}} {\norm{{\mathit{df}}}_{L^2(B)}}$$ where $$\nu(r) = {6 \pi \left( r + 2 r \csch^2(r) - \coth(r)\left(r^2 \csch^2(r) + 1\right)\right)}$$ The function $\nu \maps \R_{\geq 0} \to \R_{\geq 0}$ is a monotone increasing bijection with $\nu(r) \sim 4 \pi r^3/3$ as $r \to 0$ and $\nu(r) \sim 6 \pi r$ as $r \to \infty$. The estimate in (\[eq:dfbound\]) is sharp; in fact, our proof gives a single harmonic function $f$ for which (\[eq:dfbound\]) is an equality for all $r$. Theorem \[thm:nosobolev\] will follow directly from Lemma \[lem:dfsharp\] when $r$ is large, but $\nu(r)$ goes to $0$ too fast as $r \to 0$ to immediately give (\[eq:supnorm\]) when $r$ is small. The missing ingredient needed to prove Theorem \[thm:nosobolev\] is the following notion. A *Margulis number* for $M = \Gamma \big\backslash \H^3$ with $\Gamma \leq \Isom^+(\H^3)$ is a $\mu > 0$ so that for all $p \in \H^3$ the subgroup $$\spandef{\gamma \in \Gamma}{ d\big(p, \gamma(p)\big) < \mu}$$ is abelian. For example, $\mu = 0.1$ is a Margulis number for any such $M$ [@Meyerhoff1987 Theorem 2], and here we will use that $\mu = 0.29$ is a Margulis number whenever $H^1(M) \neq 0$ by [@CullerShalen2012]. For any fixed Margulis number $\mu$, define $${M_{\mathit{thick}}}= \setdef{m \in M}{\inj_m M \geq \mu/2} \mtext{and} {M_{\mathit{thin}}}= \setdef{m \in M}{\inj_m M < \mu/2}$$ When $M$ is closed, the thin part ${M_{\mathit{thin}}}$ is a disjoint union of tubes about the finitely many closed geodesics of length less than $\mu$. We will need: \[lem:degree\] Suppose $M$ is a hyperbolic -manifold with Margulis number $\mu$, and set $\epsilon = \min\left\{ \inj(M), \, \mu/2 \right\}$. Let $\pi \maps \H^3 \to M$ be the universal covering map. If $B \subset \H^3$ is a ball of radius $\mu/2$, then $$\label{eq:degree} \max_{m \in M} \abs{B \cap \pi^{-1}(m)} \leq \frac{\mu}{\epsilon}$$ We first assemble the pieces and prove Theorem \[thm:nosobolev\] assuming Lemmas \[lem:dfsharp\] and \[lem:degree\]. We will assume that $H^1(M) \neq 0$ as otherwise the only harmonic 1-form is identically zero. Since $H^1(M) \neq 0$, Theorem 1.1 of [@CullerShalen2012] gives that $0.29$ is a Margulis number for $M$. Setting $\mu = 0.29$ and $\epsilon = \inj(M)$, there are two cases depending on how $\epsilon$ compares to $\mu/2$. First we do the easy case of when $\epsilon \geq \mu/2$. Let $m \in M$ be a point where $\abs{\alpha_m}$ is maximal. Take $\pi \maps \H^3 \to M$ to be the universal covering map and set $\alphatil = \pi^*(\alpha)$. Fix a ball $B \subset \H^3$ of radius $\epsilon$ centered at a point $p$ in $\pi^{-1}(m)$. As $\alpha$ is harmonic on the compact manifold $M$, it is both closed and coclosed; as these are local properties, the same is true for $\alphatil$. As $\H^3$ is contractible and $\alphatil$ is closed, we have $\alphatil = {\mathit{df}}$ for some $f \maps \H^3 \to \R$. Moreover $f$ is harmonic since $\Delta f = (d^*\circ d) f = d^* \alphatil = 0$. Using Lemma \[lem:dfsharp\] we get $${\norm{\alpha}_{L^\infty}} = \abs{\alpha_m} = \abs{\alphatil_p} \leq \frac{1}{\sqrt{\nu(\epsilon)}} {\norm{\alphatil|_B}_{L^2}} \leq \frac{1}{\sqrt{\nu(\epsilon)}} {\norm{\alpha}_{L^2}}$$ where the last inequality follows as $\pi|_B$ is injective. The inequality (\[eq:supnorm\]) now follows from the fact that $\sqrt{\epsilon/\nu(\epsilon)} < 3.5 < 5$ for $\epsilon \geq \mu/2 = 0.145$. Now suppose $\epsilon < \mu/2$. We take the same setup as before except that $B \subset \H^3$ will now have radius $\mu/2$. By Lemma \[lem:degree\] it follows that $${\norm{\alphatil |_B}_{L^2}} \leq \sqrt{\frac{\mu}{\epsilon}} {\norm{\alpha|_{\pi(B)}}_{L^2}} \leq \sqrt{\frac{\mu}{\epsilon}} {\norm{\alpha}_{L^2}}$$ Hence $${\norm{\alpha}_{L^\infty}} = \abs{\alpha_m} = \abs{\alphatil_p} \leq \frac{1}{\sqrt{\nu(\mu/2)}} {\norm{\alphatil|_B}_{L^2}} \leq \sqrt{\frac{\mu}{\nu(\mu/2)}} \frac{1}{\sqrt{\epsilon}} {\norm{\alpha}_{L^2}}$$ As $\sqrt{\mu/\nu(\mu/2)} \approx 4.78$ at $\mu = 0.29$, we have proved (\[eq:supnorm\]) in this case as well. Turning to the lemmas, we start with the easier one which follows from a simple geometric argument. We can assume $\epsilon = \inj(M) < \mu/2$ as otherwise $\pi|_B$ is injective and the result is immediate since the righthand side of (\[eq:degree\]) is $2$. The basic idea of the proof is that the worst-case senario is when $m$ is on a closed geodesic $C$ of minimal length $2 \epsilon$, and $B$ is centered at a point of $\pi^{-1}(C)$. Then $B$ can contain $n + 1$ points in $\pi^{-1}(C)$ where $n = \lfloor \mu/ 2\epsilon \rfloor$; using that $n+1 \leq \mu/\epsilon$ then gives (\[eq:degree\]). We now give the detailed proof. If $m \in {M_{\mathit{thick}}}$, then any pair of distinct points in $\pi^{-1}(m)$ are distance at least $\mu$ apart, and hence at most one is in the open ball $B$; as $\mu/\epsilon \geq 2$, we have proven (\[eq:degree\]) in this case. If $m \in {M_{\mathit{thin}}}$, it lies in some tube $T$ about a short closed geodesic $C$. The components of $\pi^{-1}(T)$ are radius $R$ neighborhoods about the various geodesic lines in $\pi^{-1}(C)$. First, note that $B \cap \pi^{-1}(M)$ must lie in a single component $\Ttil$ of $\pi^{-1}(T)$; let $\gamma \in \Gamma$ generate the stabilizer of $\Ttil$. Pick a $\mtil_0 \in \Ttil \cap \pi^{-1}(m)$; then $\pi^{-1}(m)$ consists of $\gamma^n \cdot \mtil_0$ for $n \in \Z$. Adjust $\mtil_0$ if necessary so that $\mtil_0 \in B$ and any $\mtil_n = \gamma^n \cdot \mtil_0$ in $B$ has $n \geq 0$. Since $d(\mtil_0, \mtil_n) \geq n \cdot \mathrm{len}(C) \geq 2 n \epsilon$, if $\mtil_n \in B$ we have $n \leq \mu / 2\epsilon$. So there are at most $(\mu / 2\epsilon) + 1$ of the $\mtil_i$ in $B$. Since $2 \epsilon \leq \mu$, we get $\abs{B \cap \pi^{-1}(m)} \leq \mu/\epsilon$ as desired. While the calculations in the proof of Lemma \[lem:dfsharp\] are somewhat involved (unsurprisingly given the formula for $\nu(r)$), the basic idea is simple and we sketch it now. Using the natural series expansion for harmonic functions about $p$, we show that after a suitable isometry of $\H^3$ fixing $p$ we have $${\mathit{df}}= a \omega + \beta$$ where $a \in \R_{\geq 0}$, the 1-form $\omega$ is a fixed and independent of $f$ with $\abs{\omega_p} = 1$, the 1-form $\beta$ vanishes at $0$, and $\beta$ is orthogonal to $\omega$ in $L^2\big(B_r(p)\big)$ for all $r$. Then ${\mathit{df}}_p = a \omega_p$ and for each such $B = B_r(p)$ the orthogonality of $\omega$ and $\beta$ implies $a {\norm{\omega}_{L^2(B)}} \leq {\norm{{\mathit{df}}}_{L^2(B)}}$. Hence $$\abs{{\mathit{df}}_p} = \abs{a \omega_p} = a \leq \frac{{\norm{{\mathit{df}}}_{L^2(B)}}}{{\norm{\omega}_{L^2(B)}}}$$ Thus we simply define $\nu(r)$ to be $\norm{\omega}_{L^2(B_r(p))}^2$ and Lemma \[lem:dfsharp\] will then follow by calculating $\nu(r)$ explicitly. Series expansions of harmonic functions --------------------------------------- We now describe in detail the key tool used to prove Lemma \[lem:dfsharp\]: that every harmonic function $f \maps \H^3 \to \R$ has a series expansion in terms of a certain basis $\{ {\Psi_{\ell m}}\}$ of harmonic functions centered around $p$; throughout, see [@Minemura1973] or [@ElstrodtGrunewaldMennicke1998 §3.5] for details. Consider the spherical coordinates $(r, \phi, \theta) \in [0, \infty) \times [0, \pi) \times [0, 2 \pi)$ on $\H^3$ centered about $p$; in these coordinates, the metric is: $${\mathit{ds}}^2_{\H^3} = {\mathit{dr}}^2 + \sinh^2(r) {\mathit{ds}}^2_{S^2} = {\mathit{dr}}^2 + \sinh^2(r) \left( {\mathit{d}\phi}^2 + \sin^2(\phi) {\mathit{d}\theta}^2 \right)$$ For $\ell \in \Z_{\geq 0}$, define ${\psi_\ell}\maps \R_{\geq 0} \to \R_{\geq 0}$ by $${\psi_\ell}(r) = \frac{\Gamma\left(\frac{3}{2}\right)\Gamma(\ell +2)}{\Gamma\left(\ell + \frac{3}{2}\right)} \tanh^{\ell} \left( \frac{r}{2} \right) \cdot \, {{}_2 F_1 \left(-\frac{1}{2}, \ell; \, \ell + \frac{3}{2};\, \tanh^2 \left(\frac{r}{2} \right) \right)}$$ where ${}_2 F_1$ is the usual hypergeometric function and $\Psi_0$ is simply the constant function $1$. If ${Y_{\ell m}}(\phi, \theta)$ for $\ell \geq 0$ and $-\ell \leq m \leq \ell$ are the usual basis for the *real* spherical harmonics on $S^2$, define: $${\Psi_{\ell m}}= {\psi_\ell}(r) {Y_{\ell m}}(\phi, \theta)$$ These are harmonic functions on all of $\H^3$, and moreover every harmonic function $f \maps \H^3 \to \R$ has unique ${a_{\ell m}}\in \R$ so that $$\label{eq:series} f = \sum_{\ell = 0}^\infty \sum_{m = -\ell}^{\ell} {a_{\ell m}}{\Psi_{\ell m}}$$ where the series converges absolutely and uniformly on compact subsets of $\H^3$. We will use the following elementary properties of these functions: \[lem:psilm\] 1. \[item:ordervanish\] ${\psi_\ell}(r)$ vanishes to order exactly $\ell$ at $r = 0$. Consequently, ${\Psi_{\ell m}}$ vanishes to order exactly $\ell - 1$ at $p$. 2. On any ball $B$ about $p$, the functions ${\Psi_{\ell m}}$ are orthogonal in $L^2(B)$. 3. The 1-forms ${\omega_{\ell m}}= d {\Psi_{\ell m}}$ are also orthogonal in each $\Omega^1(B)$. The claim (a) follows since $\tanh \frac{r}{2} = \frac{1}{2} r + O(r^2)$ for small $r$ and in addition ${{}_2 F_1 \left(a, b; \, c;\, 0 \right)} = 1$. The second claim (b) is an easy consequence of the fact that the ${Y_{\ell m}}$ are orthonormal as elements of $L^2\left(S^2\right)$. For part (c), we have $${\omega_{\ell m}}= d\left({\psi_\ell}{Y_{\ell m}}\right) = {Y_{\ell m}}\frac{\partial {\psi_\ell}}{\partial r} {\mathit{dr}}+ {\psi_\ell}d{Y_{\ell m}}$$ and then observing that the cross-terms vanish we get $$\label{eq:omegawedge} {\omega_{\ell m}}\wedge *\omega_{k n} = {Y_{\ell m}}Y_{k n} \frac{\partial {\psi_\ell}}{\partial r} \frac{\partial \psi_k}{\partial r} {\mathit{dVol}}+ {\psi_\ell}\psi_k d{Y_{\ell m}}\wedge * dY_{kn}.$$ To compute $\pair{{\omega_{\ell m}}, \ \omega_{k n}}$, we integrate the above over $B$ and show it vanishes unless $(\ell, m) = (k, n)$. In fact, we argue that the integral of (\[eq:omegawedge\]) over each $S^2$ where $r$ is fixed is $0$. For the first term on the right-hand side of (\[eq:omegawedge\]) this follows immediately from the orthogonality of the ${Y_{\ell m}}$ in $L^2\left(S^2\right)$. For the second term, note that $* dY_{kn} = \left({\overline{*}}dY_{kn} \right)\wedge {\mathit{dr}}$ where ${\overline{*}}$ is the Hodge star operator on $S^2$, and hence the real claim is that $\pair{d{Y_{\ell m}}, dY_{kn}}_{S^2}$ vanishes. This can be deduced from the orthogonality of ${Y_{\ell m}}$ and $Y_{kn}$ via $$\pair{d{Y_{\ell m}}, dY_{kn}}_{S^2} = \pair{{Y_{\ell m}}, d^{{\overline{*}}} d Y_{kn}}_{S^2} = \pair{{Y_{\ell m}}, \Delta_{S^2} Y_{kn}}_{S^2} = k(k+1) \pair{{Y_{\ell m}}, Y_{kn}}_{S^2}$$ where the last equality holds because $Y_{kn}$ is an eigenfunction of $\Delta_{S^2}$. This finishes the proof of (c). We now prove Lemma \[lem:dfsharp\] using the approach sketched earlier. Let the ${a_{\ell m}}$ be the coefficients in the expansion (\[eq:series\]) for $f$, and note we get a corresponding expansion $${\mathit{df}}= \sum_{\ell = 1}^\infty \sum_{m = -\ell}^{\ell} {a_{\ell m}}{\omega_{\ell m}}\mtext{where ${\omega_{\ell m}}= d {\Psi_{\ell m}}$.}$$ Here the series converges absolutely and uniformly on $B$, and henceforth we view $B$ as the domain of all our 1-forms. Defining $$\eta = a_{1,-1} \omega_{1,-1} + a_{1,0} \omega_{1,0} + a_{1,1} \omega_{1,1}$$ we see by Lemma \[lem:psilm\](\[item:ordervanish\]) that $\eta_p = {\mathit{df}}_p$. Because of the orthogonality of the ${\omega_{\ell m}}$ on $B$, we know ${\norm{\eta}_{L^2}} \leq {\norm{{\mathit{df}}}_{L^2}}$ and so it suffices to prove (\[eq:dfbound\]) for $\eta$, or indeed for the components of $\eta$. In fact, because we can use an isometry of $\H^3$ fixing $0$ to interchange the $Y_{1,m}$, it suffices to establish (\[eq:dfbound\]) for the single form $\omega_{1,0}$, or indeed any multiple of it. Thus Lemma \[lem:dfsharp\] will follow immediately from the next result. For the harmonic 1-form $\omega = \sqrt{3 \pi} \cdot \omega_{1,0}$ we have $\abs{\omega_p} = 1$ and $\norm{\omega}_{L^2(B_r(p))} = \nu(r)$. The ${\psi_\ell}$ actually have alternate expressions in terms of elementary functions; in the case of interest, using that $${{}_2 F_1 \left(1/2, 1; \, 3/2;\, x^2 \right)} = \sum_{n=0}^\infty \frac{(1/2)_n (1)_n}{(3/2)_n} \frac{x^{2n}}{n!} = \sum_{n = 0}^\infty \frac{x^{2n}}{2n+1} = \frac{\arctanh(x^2)}{x}$$ and applying two contiguous relations for hypergeometric functions yields $${{}_2 F_1 \left(-1/2, 1; \, 5/2;\, x^2 \right)} = \frac{3 \left(x^3-\left(x^2-1\right)^2 \arctanh(x)+x\right)}{8 x^3}$$ and hence $$\psi_1(r) = \coth(r) - r \csch^2(r) = \frac{\sinh(r)\cosh(r) - r}{\sinh^2(r)} = \frac{2}{3} r -\frac{4}{45} r^3 + O(r^5).$$ As $Y_{1,0}$ is $\sqrt{3/4\pi}\cos{\phi}$, writing $$\omega = \sqrt{3 \pi} \cdot \omega_{1,0} = \frac{3}{2} d\left( \psi_1 \cos \phi \right)$$ in terms of the orthonormal coframe $${\kernhat{1}{{\mathit{dr}}}{-0.5}}= {\mathit{dr}}\qquad {\kernhat{1}{{\mathit{d}\phi}}{-0.5}}= \sinh(r) {\mathit{d}\phi}\qquad {\kernhat{1.5}{{\mathit{d}\theta}}{-0.5}}= \sinh(r) \sin(\phi) {\mathit{d}\theta}$$ gives $$\omega = \frac{3}{2} \left( \left(\partial_r \psi_1\right) \cos(\phi) {\kernhat{1}{{\mathit{dr}}}{-0.5}}- \frac{\psi_1 \sin(\phi)}{\sinh(r)} {\kernhat{1}{{\mathit{d}\phi}}{-0.5}}\right) \mtext{where $\partial_r \psi_1 = \displaystyle \frac{\partial \psi_1}{\partial r} = 2 \cdot \frac{r\coth(r) -1}{\sinh^2(r)}$,}$$ and hence $$\abs{\omega}^2 = \frac{9}{4} \left( \left(\partial_r \psi_1\right)^2 \cos^2(\phi) + \frac{\psi_1^2 \sin^2 \phi}{\sinh^2(r)} \right)$$ Approaching the origin along the ray $\phi=0$ gives: $$\abs{\omega_p} = \frac{3}{2} \lim_{r \to 0} \frac{\partial \psi_1}{\partial r} = 3 \lim_{r \to 0} \frac{r \coth(r) -1}{\sinh^2(r)} = 1$$ Computing the $L^2$-norm of $\omega$ on $B$ gives $$\begin{split} {\norm{\omega}_{L^2}}^2 &= \int_B \abs{\omega}^2 {\mathit{dVol}}= \int_0^{R} \int_0^{\pi} \int_0^{2 \pi} \abs{\omega}^2 \sinh^2(r) \sin(\phi) \ {\mathit{d}\theta}{\mathit{d}\phi}{\mathit{dr}}\\ &= \frac{9 \pi}{2} \int_0^{R} \int_0^{\pi} \left(\partial_r \psi_1\right)^2 \sinh^2(r) \cos^2(\phi) \sin(\phi) + \psi_1^2 \sin^3(\phi) \ {\mathit{d}\phi}{\mathit{dr}}\\ &= 3 \pi \int_0^{R} \left(\partial_r \psi_1\right)^2 \sinh^2(r) + 2 \psi_1^2 \ {\mathit{dr}}\\ &= 6 \pi \int_0^{R} \coth ^2(r)+2\csch^2(r) -6 r \coth (r) \csch^2(r) \\ & \hspace{4cm}+ r^2 \csch^2(r) \left(2 \coth^2(r)+\csch^2(r)\right){\mathit{dr}}\\ &= {6 \pi \left( R + 2 R \csch^2(R) - \coth(R)\left(R^2 \csch^2(R) + 1\right)\right)} \end{split}$$ which proves the lemma. Proof of Theorem \[thm:main\] ============================= This section is devoted to the proof of We start with the lower bound in (\[eq:main\]), where we use the two guises of the least-area/$L^1$–norm to mediate between the Thurston and harmonic norms and thereby reduce the claim to the Cauchy-Schwarz inequality. Suppose $\phi \in H^1(M)$ and let $\alpha$ be the harmonic 1-form representing $\phi$. By Theorem \[thm:areavsthurston\] and Lemma \[lem:areavsone\] we have $$\pi {\norm{\phi}_{\mathit{Th}}} \leq {\norm{\phi}_{\mathit{LA}}} = {\norm{\phi}_{L^1}}$$ From the definition of the $L^1$-norm, we have ${\norm{\phi}_{L^1}} \leq {\norm{\alpha}_{L^1}}$, and applying Cauchy-Schwarz to the pair $\abs{\alpha} \maps M \to \R$ and the constant function 1 gives $$\label{eq:cs} \pi {\norm{\phi}_{\mathit{Th}}} \leq {\norm{\alpha}_{L^1}} = {\norm{ \abs{\alpha} \cdot 1 }_{L^1}} \leq {\norm{ \alpha }_{L^2}} {\norm{ 1 }_{L^2}} = {\norm{ \alpha }_{L^2}} \sqrt{\vol(M)}$$ Since ${\norm{\phi}_{L^2}} = {\norm{\alpha}_{L^2}}$ by definition, dividing (\[eq:cs\]) through by $\sqrt{\vol(M)}$ gives the first part of (\[eq:main\]). The proof of the upper-bound in (\[eq:main\]) is essentially the same as given in [@BergeronSengunVenkatesh2016] for the corresponding part of (\[eq:norms\]), but using the upgraded Theorem \[thm:nosobolev\] to relate ${\norm{\cdotspaced}_{L^\infty}}$ and ${\norm{\cdotspaced}_{L^2}}$ and so give a sharper result. By continuity of the norms, it suffices to prove the upper bound for $\phi \in H^1(M; \Z)$. Using Theorem \[thm:areavsthurston\], fix a surface $S$ dual to $\phi$ of area at most $2 \pi {\norm{\phi}_{\mathit{Th}}}$; by definition, this means that for every *closed* 2-form $\beta$ on $M$ one has $\int_M \beta \wedge \alpha = \int_S \beta$. If $\alpha$ is the harmonic representative of $\phi$, then $d^* \alpha = -*d*\alpha = 0$, and so it follows that $*\alpha$ is closed. Hence $$\begin{split} {\norm{\alpha}_{L^2}}^2 &= \int_M \alpha \wedge *\alpha = \int_M *\alpha \wedge \alpha = \int_S *\alpha \\ &\leq \int_S \abs{*\alpha} \ {\mathit{dA}}= \int_S \abs{\alpha} \ {\mathit{dA}}\leq \int_S {\norm{\alpha}_{L^\infty}} \, {\mathit{dA}}\\ &\leq {\norm{\alpha}_{L^\infty}} \Area(S) \leq 2 \pi {\norm{\alpha}_{L^\infty}} {\norm{\phi}_{\mathit{Th}}}. \end{split}$$ Applying (\[eq:supnorm\]) from Theorem \[thm:nosobolev\], we get $${\norm{\alpha}_{L^2}}^2 \leq \frac{10 \pi}{\sqrt{\inj(M)}} {\norm{\alpha}_{L^2}} {\norm{\phi}_{\mathit{Th}}}$$ Dividing through by ${\norm{\alpha}_{L^2}}$ gives the upper bound in (\[eq:main\]), proving the theorem. The gauge theory viewpoint {#sec:gauge} -------------------------- As mentioned in the introduction, Kronheimer and Mrowka found a striking relationship between the Thurston norm and harmonic norms as one varies the Riemannian metric. Specifically, if $h$ is a Riemannian metric on $M$, let ${\norm{\cdotspaced}_{h}}$ denote the induced harmonic norm on $H^1(M; \R)$. They proved: \[thm:KM\] Suppose $M$ is a closed oriented irreducible -manifold. Then for all $\phi \in H^1(M;\R)$ one has $$\label{eq:KM} 4\pi {\norm{\phi}_{\mathit{Th}}} = \inf \setdef{\ {\norm{s_h}_{h}} \cdot {\norm{\phi}_{h}} \ }{\mbox{$h$ is a Riemannian metric on $M$}}$$ where $s_h$ is the scalar curvature of $h$. Theorem \[thm:KM\] above is equivalent to Theorem 2 of [@KronheimerMrowka1997b], as we explain below. Kronheimer and Mrowka pointed out to us that, in this form, specializing (\[eq:KM\]) to a hyperbolic metric $h$ on $M$ gives $$\frac{2 \pi}{3 \sqrt{\vol(M)}} {\norm{\cdotspaced}_{\mathit{Th}}}\leq {\norm{\cdotspaced}_{L^2}}$$ since the scalar curvature of a hyperbolic metric is $-6$; this is only slightly weaker than the first inequality of (\[eq:main\]). Theorem \[thm:KM\] gives another perspective on Theorem \[thm:main\], namely that (\[eq:main\]) bounds, from above and below, how close the hyperbolic metric can be to realizing the infimum in (\[eq:KM\]). Theorem \[thm:KM\] is a corollary of previous deep work of Kronheimer and Mrowka [@KronheimerMrowka1997a] which in turn depends on results of Gabai [@Gabai1983], Eliashberg-Thurston [@EliashbergThurston1998], and Taubes [@Taubes1996]. In contrast, our proof here of the first inequality of (\[eq:main\]) uses only some basic facts about minimal surfaces. It would be very interesting to try to extend our approach to arbitrary metrics and prove results along the lines of Theorem \[thm:KM\]; if successful, this would provide a new perspective on their results. Here is why Theorem \[thm:KM\] is equivalent to Theorem 2 of [@KronheimerMrowka1997b], which is stated in terms of the *dual* Thurston norm on $H^2(M; \R)$. Recall that for a norm $x$ on a finite-dimensional real vector space $V$, the dual norm $x^*$ on $V^*$ is defined by $$x^*(\psi) = \sup \setdef{ \psi(v) }{\mbox{$v \in V$ with $x(v) \leq 1$}}$$ Theorem 2 of [@KronheimerMrowka1997b] says that for all $\psi \in H^2(M; \R)$ one has $$\label{eq:KMorig} {\norm{\psi}^*_{\mathit{Th}}} = 4 \pi \sup_h \frac{{\norm{\psi_h}_{h}}}{{\norm{s_h}_{h}}}$$ where ${\norm{\cdotspaced}^*_{\mathit{Th}}}$ on $H^2(M; \R)$ is the norm dual to the usual Thurston norm on $H_2(M; \R)$. It is not hard to show that the harmonic norms on $H^1(M; \R)$ and $H^2(M; \R)$ are dual for any fixed $h$. Hence if $c_h = {\norm{s_h}_{h}}/4 \pi$ the norm $c_h {\norm{\cdotspaced}_{h}}$ on $H^1(M; \R)$ is dual to $\frac{1}{c_h} {\norm{\cdotspaced}_{h}}$ on $H^2(M; \R)$. The equivalence of (\[eq:KM\]) and (\[eq:KMorig\]) now follows from the following simple fact: if $x_n$ are norms on $V$ where $x = \sup x_n$ is finite on $V$, then $x$ is a norm with $x^* = \inf x_n^*$. Families of examples ==================== This section is devoted to proving Theorems \[thm:lowerexs\] and \[thm:injsmallexs\], starting with the former as it is easier. The examples $M_n$ will be built from a tower of finite covers, where the $\phi_n$ are pullbacks of some fixed class $\phi_0 \in H^1(M_0)$. To analyze that situation, first consider a degree $d$ covering map $\pi: \Mtil \to M$ and some $\phi\in H^1(M)$; as we now explain, the norms of $\phi$ and $\pi^*(\phi)$ differ by factors depending only on $d$. If $\alpha$ is the harmonic representative of $\phi$, then as being in the kernel of the Laplacian is a local property, the form $\pi^*(\alpha)$ must be the harmonic representative of $\pi^*(\phi)$. It follows that $${\norm{\pi^*(\phi)}_{L^2}} = \sqrt{d} \cdot {\norm{\phi}_{L^2}}.$$ In contrast, it is a deep theorem of Gabai [@Gabai1983 Cor. 6.13] that $${\norm{\pi^*(\phi)}_{\mathit{Th}}} = d \cdot {\norm{\phi}_{\mathit{Th}}}$$ Thus, the ratio $$\frac{{\norm{\cdotspaced}_{\mathit{Th}}}}{{\norm{\cdotspaced}_{L^2}} \sqrt{\vol{(\cdotspaced)}}}$$ is the same for both $(M, \phi)$ and $\left(\Mtil, \pi^*(\phi)\right)$. To prove the theorem, let $M_0$ be any closed hyperbolic 3-manifold with a nonzero class $\phi_0\in H^1(M_0)$ and choose a tower of finite covers $$M_0 \leftarrow M_1 \leftarrow M_2 \leftarrow M_3 \leftarrow \cdots$$ where $\inj(M_n) \to \infty$; this can be done since $\pi_1(M_0)$ is residually finite, indeed residually simple, see e.g. [@LongReid1998]. Taking $\phi_n \in H^1(M_n)$ to be the pullback of $\phi$, we have constructed pairs $(M_n, \phi_n)$ which have all the claimed properties. Harmonic forms on tubes ----------------------- To prove Theorem \[thm:injsmallexs\], we will need: \[lem:tube\] Suppose $V$ is a tube in a closed hyperbolic -manifold $M$ with a core $C$ of length $\epsilon$ and depth $R$. If $\alpha$ is a harmonic 1-form on $M$ with $\int_C \alpha = 1$ then $${\norm{\alpha|_V}_{L^2}} \geq \sqrt{\frac{2 \pi}{\epsilon} \log\left(\cosh R \right)}$$ Before proving the lemma, we give coordinates on the tube $V$ and do some preliminary calculations. Specifically, consider cylindrical coordinates $(r, \theta, z) \in [0, R] \times [0, 2\pi] \times [0, \epsilon]$ with the additional identification $(r, \theta, \epsilon) \sim (r, \theta + \theta_0, 0)$, where the twist angle $\theta_0$ is determined by the geometry of the tube; the metric on $V$ is then given by $$g_{\H^3} = {\mathit{dr}}^2 + \sinh^2(r) {\mathit{d}\theta}^2 + \cosh^2(r) {\mathit{dz}}^2$$ Since $\left\{ {\mathit{dr}}, \sinh(r) {\mathit{d}\theta}, \cosh(r) {\mathit{dz}}\right\}$ gives an orthonormal basis of 1-forms at each point of $V$, we have $$\label{eq:voltube} \Vol(V) = \int_V {\mathit{dVol}}= \int_0^\epsilon \int_0^{2 \pi} \int_0^R \sinh(r) \cosh(r) {\mathit{dr}}{\mathit{d}\theta}{\mathit{dz}}= \pi \epsilon \sinh^2(R)$$ Note that the form ${\mathit{dz}}$ is compatible with the identification of $z = 0$ with $z = \epsilon$, giving a 1-form on $V$. Now the form ${\mathit{dz}}$ is closed and also coclosed since $$d^*({\mathit{dz}}) = -*d*{\mathit{dz}}= - * d \big(\tanh(r) {\mathit{dr}}\wedge {\mathit{d}\theta}\big) = - * 0 = 0$$ and hence ${\mathit{dz}}$ is harmonic. Notice $\omega = \frac{1}{\epsilon} {\mathit{dz}}$ is a reasonable candidate for $\alpha|_V$ in Lemma \[lem:tube\] as $\int_C \omega = 1$. For this form, we have $$\begin{split}\label{eq:normomega} {\norm{\omega}_{L^2}}^2 &= \int_V \omega \wedge * \omega = \frac{1}{\epsilon^2} \int_V {\mathit{dz}}\wedge \big(\tanh(r) {\mathit{dr}}\wedge {\mathit{d}\theta}\big) \\ &= \frac{1}{\epsilon^2} \int_0^\epsilon \int_0^{2\pi} \int_0^R \tanh(r) {\mathit{dr}}{\mathit{d}\theta}{\mathit{dz}}= \frac{2 \pi}{\epsilon} \log(\cosh R) \end{split}$$ Thus Lemma \[lem:tube\] can be interpreted as saying that the harmonic norm of $\alpha|_V$ is at least that of this explicit $\omega$, and we take this viewpoint in the proof itself. From now on, we denote $\alpha|_V$ by $\alpha$ and we use only that $\alpha$ is closed, coclosed, and $\int_C \alpha = 1$. There is an action of $G = S^1 \times S^1$ on $V$ by isometries, namely translations in the $\theta$ and $z$ coordinates. First, we show it suffices to prove the lemma for the average of $\alpha$ under this action, namely $$\alphabar = \int_{G} g^*(\alpha) {\mathit{dg}}\mtext{where ${\mathit{dg}}$ is Haar measure on $G$.}$$ The advantage of $\alphabar$ will be that it must be $G$-invariant. Note that $\alphabar$ can be $C^\infty$-approximated by finite averages over suitably chosen finite subsets $\{g_i\}$ of $G$: $$\label{eq:finiteave} \alphabar \approx \frac{1}{N} \sum^N_i g_i^*(\alpha)$$ and from this it follows that $\alphabar$ is closed, coclosed, and $\int_C \alphabar = 1$. Moreover, since all ${\norm{g_i^*(\alpha)}_{L^2}} = {\norm{\alpha}_{L^2}}$, the triangle inequality applied to (\[eq:finiteave\]) gives that ${\norm{\alphabar}_{L^2}} \leq {\norm{\alpha}_{L^2}}$. This shows we need only consider the $G$-invariant form $\alphabar$. We next show that $\alphabar$ is equal to $\omega = \frac{1}{\epsilon} {\mathit{dz}}$; combined with the calculation of ${\norm{\omega}_{L^2}}$ in (\[eq:normomega\]), this will establish the lemma. Since the -forms $\{ {\mathit{dr}}, {\mathit{d}\theta}, {\mathit{dz}}\}$ are $G$-invariant, it follows that $\alphabar$ can be expressed as $$\alphabar = a(r) {\mathit{dr}}+ b(r) {\mathit{d}\theta}+ c(r) {\mathit{dz}}$$ Using that ${\mathit{d}\alphabar}= 0$, we get that $b$ and $c$ must be constants; as $\abs{{\mathit{d}\theta}} = 1/\sinh(r)$, we must further have $b=0$ so that $\alphabar$ makes sense along the core $C$ where $r=0$. As $\alphabar$ is coclosed, we learn that $$\begin{split} 0 &= d(*\alphabar) = d\big( a(r) \sinh(r) \cosh(r) {\mathit{d}\theta}\wedge {\mathit{dz}}+ c \tanh(r) {\mathit{dr}}\wedge {\mathit{d}\theta}\big) \\ & = \frac{\partial}{\partial r} \big( a(r) \sinh(r) \cosh(r) \big) {\mathit{dr}}\wedge {\mathit{d}\theta}\wedge {\mathit{dz}}\end{split}$$ Consequently, $a(r) \sinh(r) \cosh(r)$ is constant. Moreover, that constant must be $0$ to prevent $a(r)$ from blowing up as $r \to 0$. So we know $\alphabar = c {\mathit{dz}}$ and finally that $c$ must be $1/\epsilon$ to ensure $\int_C \alphabar = 1$. So $\alphabar$ is equal to $\omega$ as claimed, proving the lemma. (4.72, 4.57) .. controls (4.27, 4.57) and (3.81, 4.73) .. (3.81, 5.12) .. controls (3.81, 5.58) and (4.39, 5.68) .. (4.92, 5.68); (4.92, 5.68) .. controls (5.23, 5.68) and (5.55, 5.68) .. (5.86, 5.68); (6.27, 5.68) .. controls (6.56, 5.68) and (6.86, 5.68) .. (7.15, 5.68); (7.15, 5.68) .. controls (7.69, 5.68) and (8.27, 5.58) .. (8.27, 5.12) .. controls (8.27, 4.73) and (7.81, 4.57) .. (7.36, 4.57); (6.95, 4.57) .. controls (6.65, 4.57) and (6.35, 4.57) .. (6.05, 4.57); (6.05, 4.57) .. controls (5.75, 4.57) and (5.44, 4.57) .. (5.13, 4.57); (6.05, 3.25) .. controls (6.04, 2.80) and (5.88, 2.34) .. (5.48, 2.34) .. controls (5.02, 2.34) and (4.92, 2.92) .. (4.92, 3.45); (4.92, 3.45) .. controls (4.92, 3.82) and (4.92, 4.19) .. (4.92, 4.57); (4.92, 4.57) .. controls (4.92, 4.87) and (4.92, 5.17) .. (4.92, 5.48); (4.92, 5.89) .. controls (4.92, 6.43) and (4.39, 6.80) .. (3.81, 6.80) .. controls (2.69, 6.80) and (2.69, 5.10) .. (2.69, 3.66); (2.69, 3.25) .. controls (2.69, 2.57) and (2.69, 1.90) .. (2.69, 1.22); (2.69, 1.22) .. controls (2.69, 0.61) and (2.19, 0.11) .. (1.58, 0.11) .. controls (0.46, 0.11) and (0.76, 2.62) .. (1.01, 4.66) .. controls (1.31, 7.17) and (3.40, 9.10) .. (5.88, 8.95); (6.29, 8.92) .. controls (7.27, 8.86) and (8.26, 8.80) .. (9.24, 8.74); (9.24, 8.74) .. controls (9.84, 8.70) and (10.38, 8.31) .. (10.43, 7.73) .. controls (10.47, 7.22) and (10.03, 6.80) .. (9.50, 6.80); (9.09, 6.80) .. controls (8.17, 6.80) and (7.15, 6.67) .. (7.15, 5.89); (7.15, 5.48) .. controls (7.15, 5.17) and (7.15, 4.87) .. (7.15, 4.57); (7.15, 4.57) .. controls (7.15, 4.19) and (7.15, 3.82) .. (7.15, 3.45); (7.15, 3.45) .. controls (7.15, 1.69) and (4.95, 1.22) .. (2.90, 1.22); (2.49, 1.22) .. controls (1.94, 1.22) and (1.58, 1.75) .. (1.58, 2.34) .. controls (1.58, 2.95) and (2.08, 3.45) .. (2.69, 3.45); (2.69, 3.45) .. controls (3.37, 3.45) and (4.04, 3.45) .. (4.72, 3.45); (5.13, 3.45) .. controls (5.44, 3.45) and (5.74, 3.45) .. (6.05, 3.45); (6.05, 3.45) .. controls (6.35, 3.45) and (6.65, 3.45) .. (6.95, 3.45); (7.36, 3.45) .. controls (8.78, 3.45) and (9.34, 5.15) .. (9.29, 6.80); (9.29, 6.80) .. controls (9.28, 7.37) and (9.26, 7.95) .. (9.25, 8.53); (9.24, 8.94) .. controls (9.22, 9.60) and (8.40, 9.78) .. (7.65, 9.82) .. controls (6.88, 9.86) and (6.09, 9.60) .. (6.08, 8.93); (6.08, 8.93) .. controls (6.08, 7.85) and (6.07, 6.76) .. (6.06, 5.68); (6.06, 5.68) .. controls (6.06, 5.38) and (6.06, 5.07) .. (6.06, 4.77); (6.05, 4.36) .. controls (6.05, 4.13) and (6.05, 3.89) .. (6.05, 3.66); Our examples are made by Dehn filling a certain -cusped hyperbolic -manifold. Let $W$ be a compact manifold with $\bdry W = T_1 \sqcup T_2$ where both $T_i$ are tori, whose interior is hyperbolic, which fibers over the circle, and where maps $H_1(T_i; \Z) \to H_1(W; \Z)$ are isomorphisms. (The last condition should be interpreted as saying that $W$ is homologically indistinguishable from $T \times I$.) Further, we require that $W$ has an orientation-reversing involution that interchanges the $T_i$ and acts on $H_1(W; \Z)$ by the identity. One such $W$ is described in Figure \[fig:link\]. Since $W$ fibers, there is a -dimensional face $F$ of the Thurston norm ball so that any $\phi \in H^1(W; \Z)$ in the cone $C_F = \R_{>0} \cdot F$ can be represented by a fibration. Choose $\alpha, \beta \in C_F$ that form an integral basis for $H^1(W; \Z)$, and let $a, b \in H_1(W; \Z)$ be the dual homological basis where $\alpha(a) = \beta(b) = 1$ and $\alpha(b) = \beta(a) = 0$. Let $M_n$ be the closed manifold obtained by Dehn filling $W$ along the curves in $T_i$ homologically equal to $a - n b$; since $W$ is a $\Z$-homology $T \times I$, it follows that $H^1(M_n; \Z)$ is $\Z$ and is generated by the extension $\phi_n$ of $\phitil_n = n \alpha + \beta$ to $M_n$. We will show there are constants $c_1, c_2 > 0$ so that (i) \[item:norm\] For all $n$, we have ${\norm{\phi_n}_{\mathit{Th}}} = n {\norm{\alpha}_{\mathit{Th}}} + {\norm{\beta}_{\mathit{Th}}} - 2$. (ii) \[item:hyper\] For large $n$, the manifold $M_n$ is hyperbolic. Moreover, we have $\vol(M_n) \to \vol(W)$ and $\inj(M_n) \sim \frac{c_1}{n^2}$ as $n \to \infty$. (iii) \[item:tube\] For large $n$, there is a tube $V_n$ in $M_n$ with core geodesic $\gamma_n$ of length $2 \inj(M_n)$, with depth $R_n \geq\arcsinh(c_2 n)$, and where $\int_{\gamma_n} \phi_n = 1$. Here is why these three claims imply the theorem. From (\[item:hyper\]) and (\[item:tube\]), an easy calculation with Lemma \[lem:tube\] gives a $c_4 > 0$ so that $$\label{eq:normgrowth} {\norm{\phi_n}_{L^2}} \geq {\norm{\phi_n |_{V_n}}_{L^2}} \geq c_4 n \sqrt{\log n} \mtext{for all large $n$.}$$ Combining (\[eq:normgrowth\]) with (\[item:norm\]-\[item:tube\]) now gives both parts of Theorem \[thm:injsmallexs\]. So it remains to prove the claims (\[item:norm\]-\[item:tube\]). For (\[item:norm\]), first note that since ${\norm{\cdotspaced}_{\mathit{Th}}}$ is linear on $C_F$ we have $${\norm{\phitil_n}_{\mathit{Th}}} = n {\norm{\alpha}_{\mathit{Th}}} + {\norm{\beta}_{\mathit{Th}}}$$ Let $\Stil_n$ be a fiber in the fibration dual to $\phitil_n$, and hence $\eulerminus{\Stil_n} = {\norm{\phitil_n}_{\mathit{Th}}}$ by [@Thurston1986 §3]. For homological reasons, the boundary of $\Stil_n$ has only one connected component in each $T_i$. Thus $\Stil_n$ can be capped off with two discs to give a surface $S_n \subset M_n$ which is a fiber in a fibration of $M_n$ over the circle. Hence $${\norm{\phi_n}_{\mathit{Th}}} = \eulerminus{S_n} = \eulerminus{\phitil_n} - 2 = {\norm{\phitil_n}_{\mathit{Th}}} - 2 = n {\norm{\alpha}_{\mathit{Th}}} + {\norm{\beta}_{\mathit{Th}}} - 2$$ proving claim (\[item:norm\]). Starting in on (\[item:hyper\]), the Hyperbolic Dehn Surgery Theorem [@ThurstonLectureNotes] shows that $M_n$ is hyperbolic for large $n$ and that $\vol(M_n)$ converges to $\vol(W)$. Moreover, for large enough $n$, the cores of the Dehn filling solid tori are isotopic to the two shortest geodesics in $M_n$. In fact, these core geodesics have the same length: the required involution of $W$ extends to $M_n$ and thus give an isometry of $M_n$ which interchanges them. Let $\gamma_n \subset M_n$ be the core geodesic inside $T_1$. Using [@NeumannZagier1985 Proposition 4.3] one can show that there is a constant $c_1 > 0$ so that $\operatorname{length}(\gamma_n) \sim \frac{2 c_1}{n^2}$ as $n \to \infty$. As $2 \inj(M_n) = \operatorname{length}(\gamma_n)$, we have shown (\[item:hyper\]). Finally, for (\[item:tube\]), first note that $\gamma_n$ meets the fiber surface $S_n$ in a single point, giving $\int_{\gamma_n} \phi_n = 1$ if we orient $\gamma_n$ suitably. So it remains only to estimate the depth $R_n$ of the Margulis tube $V_n$ about $\gamma_n$. Let $V_n'$ be the image of $V_n$ under the above involution of $M_n$. The picture from the Hyperbolic Dehn Surgery Theorem shows that the geometry of $M_n$ outside the $V_n \sqcup V_n'$ converges to that of a compact subset of $W$. As $\vol(M_n) \to \vol(W)$, we must have that the volume of $V_n$ converges as $n \to \infty$. A straightforward calculation using (\[eq:voltube\]) now gives the lower bound on $R_n$ claimed in (\[item:tube\]). Proof of Theorem \[thm:largenorm\] ================================== [images/genus2homshaded]{}\[font=\] at (9.5,5.5) [$e_1$]{}; at (40.5,16.3) [$e_2$]{}; at (60.7,16.3) [$e_4$]{}; at (90.0,6.0) [$e_3$]{}; at (49.6,1.1) [(a) Generators of $H_1(S; \Z)$]{}; [images/genus2altshaded]{}\[font=\] at (10.1,6.0) [$a$]{}; at (20.5,30.9) [$b$]{}; at (49.6,27.0) [$c$]{}; at (86.8,31.7) [$d$]{}; at (89.0,7.1) [$e$]{}; at (49.6,1.1) [(b) Curves for Dehn twisting]{}; This section is devoted to proving: We first outline the construction of the $M_n$, sketch why they should have the desired properties, and finally fill in the details. In contrast with the rest of this paper, in this section, all (co)homology groups will use $\Z$ coefficients. The construction ---------------- Let $S$ be a fixed surface of genus 2. We use the basis $e_i$ for $H_1(S)$ shown in Figure \[fig:genus2\](a); we take $e^i$ to be the algebraically dual basis of $H^1(S)$, that is, the one where $e^i(e_j) = \delta_{ij}$. Let $W$ be a compact hyperbolic -manifold with totally geodesic boundary, where $\partial W$ has genus 2 and the map $H_1(\partial W) \to H_1(W)$ is onto; one possible choice for $W$ is the tripus manifold of [@Thurston1997 §3.3.12]. Note that homologically $W$ is indistinguishable from a genus 2 handlebody. Let $W_1$ and $W_2$ be two copies of $W$ whose boundaries have been marked by $S$ so that $\ker\left( H_1(S) \to H_1(W_1) \right) = \pair{e_2, e_4}$ and $\ker\left( H_1(S) \to H_1(W_2) \right) = \pair{e_2, e_3}$. Thus $H_1(W_1)$ is spanned by the images of $e_1$ and $e_3$ and so $H^1(W_1)$ is spanned by natural extensions of $e^1$ and $e^3$. For a carefully chosen pseudo-Anosov $f \maps S \to S$, the examples used in Theorem \[thm:largenorm\] are $$M_n = W_1 \cup_{f^n} W_2 = \rightquom{W_1 \sqcup W_2}{(x \in \partial W_1)\sim (f^n(x) \in \partial W_2)}{3pt}{\big}$$ We first sketch the requirements on $f$ and the overall structure of the argument. Note that our particular markings mean that $H^1(M_0) = \Z = \pair{e^1}$, where we are taking $f^0$ to be the identity map on $S$. The key requirement is that $f^* \maps H^1(S) \to H^1(S)$ preserves the subspace $\pair{e^1, e^3}$ and acts on it as an Anosov matrix in $\SL{2}{\Z}$. We will also arrange that $H^1(M_n) = \Z = \pair{\phi_n}$ for all $n$. Since the action of $f^*$ on $\pair{e^1, e^3}$ is complicated, the coefficients of $\phi_n$ with respect to $\{e^1, e^3\}$ will grow exponentially in $n$. This will force the restriction of $\phi_n$ to $W_1$ to have Thurston norm which is exponential in $n$, and a standard lemma will then give that $\phi_n$ itself has Thurston norm which is exponential in $n$. We will also show that $\vol(M_n)$ grows (essentially) linearly in $n$, and combining these will give part (\[item:expgrowth\]) of Theorem \[thm:largenorm\]. The specific choice we make for $f$ is given in the next lemma. \[lem:pA\] There exists a pseudo-Anosov $f \maps S \to S$ whose action $f_*$ on $H_1(S)$ is given by $$B = \left(\begin{array}{rrrr} 3 & 0 & 1 & 0 \\ 1 & 0 & 0 & -1 \\ -1 & 0 & 0 & 0 \\ 1 & 1 & 1 & 3 \end{array}\right)$$ The homomorphism $\MCG(S) \to \Aut(H_1(S))$ induced by taking the action of a mapping class on homology is onto the symplectic group of the form given by $$J = \left(\begin{array}{rr|rr} 0 & -1 & 0 & 0 \\ 1 & 0 & 0 & 0 \\ \hline 0 & 0 & 0 & 1 \\ 0 & 0 & -1 & 0 \end{array}\right)$$ where we are following the conventions given in Figure \[fig:genus2\](a). It is easy to check that $B^t J B = J$, and so there is some $f \in \MCG(S)$ with $f_* = B$. Composing $f$ with a complicated element of the Torelli group if necessary, we can arrange that $f$ is pseudo-Anosov as required. Alternatively, consider the following product of Dehn twists: $$f = \tau_a \circ \tau_d^{-1} \circ \tau_c \circ \tau_b^{-1} \circ \tau_d \circ \tau_c^{-1} \circ \tau_e^{-1}$$ where the curve labeling conventions for the right-handed Dehn twists $\tau$ follow Figure \[fig:genus2\](b). An easy calculation gives $f_* = B$, and using [@Twister; @SnapPy; @hikmot2016] one can rigorously verify that the mapping torus of $f$ is hyperbolic with volume $\approx 7.51768989647$; in particular, $f$ is pseudo-Anosov. The precise technical statements needed to prove Theorem \[thm:largenorm\] are the following three lemmas. \[lem:alggengrows\] For all $n$, the group $H^1(M_n) \cong \Z$ is generated by $\phi_n = a_n e^1 + c_n e^3$ where both $a_n$ and $c_n$ grow exponentially in $n$ at a rate of $\lambda = \frac{3 + \sqrt{5}}{2} \approx 2.62$. \[lem:geomgrows\] The manifolds $M_n$ are all hyperbolic with injectivity radius bounded uniformly below, and $\vol(M_n) \asymp n$ as $n \to \infty$. \[lem:normsplitting\] Suppose $M^3$ is a closed irreducible -manifold. Suppose $F \subset M$ is an incompressible surface dividing $M$ into submanifolds $A$ and $B$. For all $\phi \in H^1(M)$ we have $${\norm{\phi}_{\mathit{Th}}} \geq {\norm{\phi_A}_{\mathit{Th}}} + {\norm{\phi_B}_{\mathit{Th}}}$$ where $\phi_A$ and $\phi_B$ are the images of $\phi$ in $H^1(A)$ and $H^1(B)$ respectively. We first prove Theorem \[thm:largenorm\] assuming the lemmas, and then establish each of them in turn. By Lemma \[lem:normsplitting\], if $\phibar_n$ denotes the restriction of $\phi_n$ to $H^1(W_1)$, we have $${\norm{\phi_n}_{\mathit{Th}}} \geq {\norm{\phibar_n}_{\mathit{Th}}}$$ Since $W_1$ is hyperbolic with totally geodesic boundary, it is atoroidal and acylindrical and hence the Thurston norm on $H^1(W_1)$ is nondegenerate. As any two norms on a finite-dimensional vector space are uniformly comparable, Lemma \[lem:alggengrows\] gives that ${\norm{\phibar_n}_{\mathit{Th}}} \asymp \lambda^n$ as $n \to \infty$. Since $\vol(M_n) \asymp n$ by Lemma \[lem:geomgrows\], we have that ${\norm{\phi_n}_{\mathit{Th}}}$ grows exponentially in $\vol(M_n)$ as required. In fact, working a little harder one can make the rate of exponential growth explicit, namely $$\label{eq:explicit} \log{{\norm{\phi_n}_{\mathit{Th}}}} > \ 0.348 \cdot \vol(M_n) \mtext{for large $n$.}$$ Specifically, take $f$ to be the map constructed in the second proof of Lemma \[lem:pA\]. If we use [@Tian1990] in the manner of [@Namazi2005 Chapter 12] to get a refined version of the model for $M_n$, it follows that $\vol(M_n)/n$ limits to $\vol(M_f) \approx 7.51768989$. Combining with the explicit formula for $\lambda$ in Lemma \[lem:alggengrows\] gives (\[eq:explicit\]). First, we show that $H^1(M_n) = \Z$ for all $n$, and moreover identify the generator $\phi_n$ in terms of the basis $e^i$ for $H^1(S)$. (The stronger claim $H_1(M_n) \cong \Z$ is also true, but we have no need for this here.) Let $$F = f^* = B^t = \left(\begin{array}{rrrr} 3 & 1 & -1 & 1 \\ 0 & 0 & 0 & 1 \\ 1 & 0 & 0 & 1 \\ 0 & -1 & 0 & 3 \end{array}\right)$$ By Mayer-Vietoris, we have $$\label{eq:MV} \begin{split} H^1(M_n) &= \operatorname{image}\big(H^1(W_1) \to H^1(S)\big) \cap (f^*)^n \Big( \operatorname{image}\big(H^1(W_2) \to H^1(S)\big)\Big) \\ &= \pair{e^1, e^3} \cap \pair{ F^n(e^1), F^n(e^4)} \end{split}$$ Notice that $F$ preserves the subspace $\pair{e^1, e^3}$ and acts there by the matrix $$\Fbar = \left(\begin{array}{rr} 3 & -1 \\ 1 & 0 \end{array}\right)$$ which is an Anosov matrix in $\SL{2}{\Z}$. For $n = 0$, the intersection in (\[eq:MV\]) is $$\pair{e^1, e^3} \cap \pair{e^1, e^4} = \pair{e^1}.$$ Hence, for general $n$ the intersection is spanned by $F^n(e^1)$ which is $a_n e^1 + c_n e^3$ where $$\Fbar^n = \left(\begin{array}{rr} a_n & b_n \\ c_n & d_n \end{array}\right)$$ Thus $H^1(M_n) = \Z$ as claimed, with generator $\phi_n$ restricting to $H^1(W_1)$ as $a_n e^1 + c_n e^3$, where $a_n$ and $c_n$ grow exponentially in $n$, specifically at a rate $\lambda = \frac{3 + \sqrt{5}}{2} \approx 2.618034$. Though a geometric limit argument could be given to verify the injectivity radius and volume claims, we refer for efficiency to the main result of [@BrockMinskyNamaziSouto2016] for a bi-Lipschitz model for the manifolds $M_n$. Following the terminology of [@BrockMinskyNamaziSouto2016], the *decorated manifolds* $\cM$ are the pair of acylindrical manifolds $W_1$ and $W_2$ with totally geodesic boundary, the *decorations* $\mu_1$ and $\mu_2$ on the boundaries $\bdry W_1$ and $\bdry W_2$ can be taken to be bounded length markings in the induced hyperbolic structure on each boundary component, and the gluing map is $f^n$ as above. The *large heights* condition in Theorem 8.1 of [@BrockMinskyNamaziSouto2016] is evidently satisfied since the curve complex distance $$d_{\cC(\bdry W_2)}\big(f^n(\mu_1),\mu_2\big)$$ grows linearly with $n$ and likewise for $\big(\mu_1, f^n(\mu_2)\big)$ in $\cC(\bdry W_1)$. Furthermore, the pair $\big(f^n(\mu_1) , \mu_2\big)$ has *$R$-bounded combinatorics*, where $R$ is independent of $n$, for the following reason. Picking points $X$ and $Y$ in the Teichmüller space of $S$ where $\mu_1$ and $\mu_2$ have bounded length, the quasi-fuchsian manifolds $Q\big(f^n(Y),X\big)$ converge strongly to a manifold $N$ with injectivity radius bounded below by [@McMullen1996 Corollary 3.13]. Thus there is a uniform lower bound on the injectivity radii of the approximates $Q(f^n(Y),X)$. Applying the Bounded Geometry Theorem [@Minsky2001] to the $Q(f^n(Y), X)$ gives the needed uniform upper bound on the distance between $f^n(\mu_1)$ and $\mu_2$ in any subsurface projection. The *model manifold* for the $(\cM,R)$-gluing $X_n$ determined by these data is as follows. Let $\Mtil_f$ be the fiber cover of the mapping torus $M_f$ and let $\Mtil_f[0,1]$ be a fundamental domain for the action of $f$ as an isometric covering translation $\alpha_f: \Mtil_f \to \Mtil_f$ bounded by a choice of fiber and its translate by $\alpha_f$. Defining $\Mtil_f[k,k+1]$ to be $\alpha_f^k (\Mtil_f[0,1])$, we use $\Mtil_f [0,n]$ to denote the union of $n$ successive such fundamental domains. Then the model manifold $\mathbb{M}_{X_n}$ is the gluing of $W_1$ and $W_2$ along their boundary to $\Mtil_f [0,n]$ in the manner described in [@BrockMinskyNamaziSouto2016 §2.15]. Given that $\Mtil_f$ is periodic, we know that $\inj(\mathbb{M}_{X_n})$ is bounded below independent of $n$ and that $\vol(\mathbb{M}_{X_n}) \sim \vol(M_f) \cdot n$ as $n \to \infty$. Now Theorem 8.1 of [@BrockMinskyNamaziSouto2016] gives a $K$ so that for all large $n$ there is a $K$–bi-Lipschitz diffeomorphism $$f_{X_n} \colon \mathbb{M}_{X_n} \to M_n$$ Combined with the above facts about the geometry of $\mathbb{M}_{X_n}$, this gives the claimed properties for $M_n$ and so proves the lemma. Let $S \subset M$ be a surface dual to $\phi$ which is *taut*, that is, the surface $S$ realizes ${\norm{\phi}_{\mathit{Th}}}$, is incompressible, and no union of components of $S$ is separating. As $F$ and $S$ are incompressible, we can isotope $S$ so that $F \cap S$ consists of curves that are essential in both $S$ and $F$; in particular, every component of $S \setminus F$ has non-positive Euler characteristic. As $S \cap A$ and $S \cap B$ are dual to $\phi_A$ and $\phi_B$ respectively, we have $${\norm{\phi}_{\mathit{Th}}} = -\chi(S) = -\chi(S \cap A) - \chi(S \cap B) = \eulerminus{S \cap A} + \eulerminus{S \cap B} \geq {\norm{\phi_A}_{\mathit{Th}}} + {\norm{\phi_B}_{\mathit{Th}}}$$ as desired.
{ "pile_set_name": "ArXiv" }
epsf Mixtures of liquid crystals with a small amount of polymer (polymer-stabilized liquid crystals[@crawford; @mariani; @broer], or PSLC’s) show promise for electro-optic devices such as light shutters and displays [@hikmet1; @yang1; @yang2; @held1], because the polymer tends to form a network that aligns the liquid crystal[@fung1]. Since polymers and liquid crystals tend to be immiscible, the dispersions are prepared by mixing a small amount of miscible monomer with the liquid crystal and photopolymerizing. As the polymers grow, the system phase separates into an ordered phase rich in liquid crystal and an isotropic phase rich in polymer. Long before the system reaches equilibrium, however, the polymerization “freezes” the mixture into a crosslinked network of polymer-rich domains. Thus, the fabrication of PSLC’s involves interplay among three kinetic processes: polymerization, phase separation, and phase ordering. Depending on the time scales that control these processes, a rich variety of morphologies have been observed[@fung2; @rajaram; @held2; @hikmet2]. Because of the number of nonequilibrium processes involved, however, there is little theoretical understanding of the factors that control the domain morphology. In this Letter, we focus on the interplay between phase separation (PS) and phase ordering (PO) kinetics in mixtures of short, rigid polymers ([*rods*]{}) and long, flexible polymers ([*coils*]{}), as a first step towards rational design and control of the network morphology. It is well known that thermodynamic factors such as the anisotropy of the isotropic/nematic interfacial tension can influence domain morphology, leading to anisotropic domain shapes. However, there are also kinetic factors that control domain morphology, such as the anisotropic diffusion coefficient of a rod. To capture these thermodynamic and kinetic effects, we use a Cahn-Hilliard framework that allows composition and orientational density to evolve in a coupled fashion as functions of position and time following a temperature quench[@liu96]. In contrast to earlier studies that treat orientational density as a scalar order parameter[@dorgan; @lansac], this framework includes the orientational density’s second-order tensorial nature[@prost]. Although it is instructive to study the case of two coupled scalar order parameters (Model C[@hohenberg]), a scalar cannot capture the direction of nematic order. Because a vector does not have head/tail symmetry, it is crucial to retain the tensor order parameter to obtain domain anisotropy[@sagui]. To assess the effects of phase ordering, we study two systems. The first (denoted RC) is a mixture of rods and coils. The second (denoted CC) is a polymer blend, identical to RC except that the rods are replaced by flexible chains of the same length that do not align. In both cases, we solve the linearized coupled partial differential equations of motion analytically, and the nonlinear equations numerically into the late-time regime, after quenching the isotropic, homogeneous mixture into the coexistence region. We find that phase ordering dramatically affects morphology, giving rise, for example, to interconnected networks or elongated domains depending on whether phase ordering or phase separation is the initially dominant process. We study two-dimensional, incompressible mixtures of chain molecules made up of monomers of the same size. In RC, we have short rods that are $N_{A}=10$ monomers long, and coils that are $N_{B}=100$ monomers long. The phase behavior of the system is governed by a bulk free energy[@holyst; @liu93] $F_{\rm RC}$ that couples the area fraction of rods, $\phi$, to the orientational density of the rods, $\tensor {S}$. This free energy consists of two parts: the Flory-Huggins free energy, which governs phase separation, and the Landau-de Gennes free energy, which governs the isotropic/nematic transition. The free energy was calculated within the random phase approximation[@liu93]. It depends on the rod and coil lengths, and contains only two free parameters. These are the two interaction parameters, namely the Flory parameter, $\chi$, and the Maier-Saupe parameter, $w$, which control the isotropic and anisotropic monomer-monomer interactions, respectively. Both of these parameters depend on temperature with entropic and enthalpic contributions: $\chi = \chi_{0} + \chi_{1}/T$ and $w = w_{0} + w_{1}/T$. We have chosen the parameters $\chi_{i}$ and $w_{i}$ so that the critical temperature for phase separation according to the Flory-Huggins free energy, $T_{c}$, and the isotropic/nematic transition temperature of the pure rod system, $T_{ni}$, are both of order unity in our (arbitrary) units, with $T_{ni}$ higher than $T_{c}$. We use $\chi_{0} = 0.055$, $\chi_{1} = 0.036$, $w_{0} = 0.55$, and $w_{1} = 0.26$, which yield $T_{c} = 1.14$ and $T_{ni} = 1.3$. The resulting RC phase diagram is shown in Fig. \[phasediagram\]. to Below a triple point, the system demixes into an isotropic phase rich in coils and a nematic phase rich in rods. Above the triple point, there is an isotropic-isotropic coexistence region and an isotropic-nematic (IN) coexistence region. The IN coexistence region ends in a critical point that is connected to the IN transition of a pure rod system by a second-order transition line. Note that in three dimensions, the IN transition is first order, so the phase diagram would be different. Below the triple point, however, we find that the IN coexistence curves are qualitatively identical in two and three dimensions. The dashed line represents the spinodal for phase separation (PS), while the dotted line is the spinodal for phase ordering (PO). In the CC system, we also use $N_{A}=10$ and $N_{B}=100$ monomers. Because the “rods” are now flexible, the free energy $F_{\rm CC}$ is given by the Flory-Huggins free energy only, without the Maier-Saupe contribution [@glotzer]. The parameter $\chi$ is exactly the same as in the RC system. Thus, CC is identical to RC except that there is no phase ordering. We study the morphology following four quenches from the isotropic, homogeneous phase into the IN coexistence region below the triple point, to the four points marked [**A**]{}-[**D**]{} in the RC phase diagram. We use a simplified version of the equations of motion derived within the dynamical random phase approximation[@liu96] $$\begin{aligned} {\partial \phi \over \partial t}& = & \Gamma_{\phi \phi} \nabla^2{{\delta F(\phi,\tensor {S})} \over \delta \phi} \label{peom}\\ {\partial S^{ij} \over\partial t} & = & -\Gamma^{ijkl}_{S S}{{\delta F(\phi,\tensor {S})} \over \delta S^{kl}} \label{Seom}\end{aligned}$$ where $\Gamma_{\phi \phi}$ and $\Gamma^{ijkl}_{S S}$ are calculated Onsager coefficients that depend on the single chain and single rod diffusion coefficients; expressions for these coefficients are provided in Ref. [@liu96]. The free energy functional $F(\phi,\tensor {S})$ consists of the bulk free energy and a nonlocal free energy that controls the cost of gradients in composition and orientational density[@liu93]. Note that the cost of gradients in orientational density are determined by the Frank elastic constants that characterize nematic elasticity. For CC, the time evolution is given by Eq. \[peom\] alone, since $\tensor S = 0$ and $\delta F_{\rm CC}/\delta S^{ij} = 0$. We emphasize that for both systems, the equations of motion are not phenomenological; they are derived from microscopic models. We solve the discretized equations on a square lattice of size $150^{2}$ using a variable timestep Runge-Kutta numerical integration scheme. We have also varied the number of lattice points and the size of the mesh to verify that our observations do not depend on system size or discretization. [*Quench [**A**]{}*]{}: Here we quench to a point where RC is initially unstable with respect to PS, but metastable with respect to PO. Thus, the point marked [**A**]{} in Fig. \[phasediagram\] lies below the spinodal for PS and above the one for PO. For CC, where phase ordering is absent, this quench leads to circular droplets rich in the shorter coils, in a matrix rich in the longer coils. RC initially displays almost identical behavior with circular droplets rich in rods. Once the concentration of rods in the droplets is comparable to the concentration at the IN spinodal, however, the rods in the droplets begin to order. At this time, the droplets abruptly expel more coils because the coils are less soluble when the rods are ordered. In addition, the droplets must develop defects because the rods want to be parallel to each other and parallel to the droplet interface. As the rods order, a pair of defects forms inside each droplet and separates, with the two defects moving along the director in opposite directions towards the edge of the droplet. The magnitude of orientational order is lower at the edges than in the center; thus, the defects migrate towards the edges to lower their energy. Since the overall system is isotropic, the long axis of each droplet is randomly oriented. Note that it is essential to retain the tensorial nature of the orientational order parameter in order to obtain randomly-oriented elongated domains. Fig. \[quenchA\] shows a snapshot of the RC system at a time at which there is significant phase separation and ordering. to [*Quench [**B**]{}*]{}: For this quench, CC is metastable with respect to PS, so isolated droplets rich in the long coils would form by nucleation if we had included thermal noise. Since we have not included thermal noise, the CC system does not evolve. In this region RC is also initially metastable with respect to PS, but it is unstable with respect to PO. The instability towards orientational ordering eventually drives the system to phase separate because the two order parameters are coupled. This makes physical sense: once the rods are strongly aligned, they expel the coils into isolated droplets. Although there is no orientational order within the coil-rich droplets, the droplets are anisotropic because of the nematic elasticity in the surrounding rod-rich matrix. The appearance of droplets even when the system is only metastable to phase separation is reminiscent of recent theoretical work on protein and polymer crystallization, which shows that a quench below the spinodal for one order parameter (density) can promote domain growth in another order parameter (crystallization)[@frenkel; @olmsted]. The advantage of our system is that we can calculate the coupled equations of motion for our two order parameters &gt;from a microscopic model of rods and coils. [*Quench [**C**]{}*]{}: In this quench, RC is initially unstable both to PS and to PO, but is more unstable with respect to PS. In both RC and CC, the system forms a bicontinuous network that rapidly breaks up into droplets rich in long coils. In RC, however, defects in the surrounding rod-rich domains give rise to droplets that are noncircular, as shown in Fig. \[quenchC\]. Furthermore, the coupling of PS and PO leads to a faster onset of phase separation in RC than in CC, because the rod-rich regions tend to expel coils as the rods order. to [*Quench [**D**]{}*]{}: As in Quench [**C**]{}, RC is initially unstable both to PS and to PO. However, in this quench it is more unstable with respect to PO. Once the degree of order is significant, then phase separation begins. Because of PO, PS begins much earlier in RC than in CC; by the time CC displays any noticeable phase separation, the concentration difference between the two phases in RC has already reached its equilibrium value. Fig. \[quenchD\] presents snapshots of the systems as they evolve in time. In CC, an interconnected network rich in the longer coils initially appears. However, it quickly breaks up into droplets that become increasingly circular over time. In RC, on the other hand, small highly-ordered rod-rich droplets initially form in a coil-rich matrix. As these rod-rich droplets grow, the surrounding coil-rich region shrinks into a network of interconnected domains that are strikingly fibrillar, despite the fact that the coil-rich phase is the minority phase. This is in contrast to systems with only a compositional order parameter (Model B [@hohenberg]), where only the majority phase can form networks. Eventually, this network breaks up to form coil-rich droplets; again, these are noncircular because of defects in the surrounding rod-rich regions. However, in experiments the crosslinking of the polymer network may arrest the phase separation at the fibrillar network stage. to In summary, we find that phase ordering can significantly influence domain morphology, especially when (1) the system is unstable to both phase separation and phase ordering and (2) orientational order is significant at the onset of phase separation. This is consistent with morphologies observed in PSLC’s, where fibrillar networks of the minority polymer-rich domains form when the system is initially ordered[@fung2; @rajaram; @held2]. Our results show that we can learn much about domain morphology from the phase diagram and spinodal lines, and that the coupling of phase separation and ordering kinetics leads to rich phenomena that warrant further study. We thank Abdullah Al Sunaidi, Rashmi Desai, Weinan E, Peter Palffy-Muhoray, and Rebecca M. Nyquist for stimulating and instructive discussions that led to this research. This work was supported by NIST through the Center for Theoretical and Computational Materials Science and by the NSF through Grant No. CHE-9624090 (AJL). G. P. Crawford and S. Zumer, [*Liquid Crystals in Complex Geometries Formed by Polymer and Porous Networks*]{} (Taylor & Francis, London, 1996). P. Mariani, B. Samoria, A. S. Angeloni, and P. Ferruti, Liq. Cryst. [**1**]{} 327 (1986). D. J. Broer, R. G. Gossink, and R. A. M. Hikmet, Angew. Makromol. Chem. [**183**]{}, 45 (1990). R. A. M. Hikmet, J. Appl. Phys. [**68**]{}, 4406 (1990). D. K. Yang, L. C. Chien, and J. W. Doane, Appl. Phys. Lett. [**60**]{}, 3102 (1992). D. K. Yang, J. L. West, L. C. Chien, and J. W. Doane, J. Appl. Phys. [**76**]{}, 1331 (1994). I. Dierking, L. L. Kosbar, A. Afzali-Ardakani, A. C. Lowe, and G. A. Held, J. Appl. Phys. [**81**]{}, 3007 (1997). Y. K. Fung, A. Borstnik, S. Zumer, D. K. Yang, and J. W. Doane, Phys. Rev. E [**55**]{}, 1637 (1997). Y. K. Fung, D. K. Yang, Y. Sun, L. C. Chien, S. Zumer, and J. W. Doane, Liq. Cryst. [**19**]{}, 797 (1995). C. V. Rajaram, S. D. Hudson, and L. C. Chien, Chem. Mater. [**7**]{}, 2300 (1996). G. A. Held, L. L. Kosbar, I. Dierking, A. C. Lowe, G. Grinstein, V. Lee, and R. D. Miller, Phys. Rev. Lett. [**79**]{}, 3443 (1997). R. A. M. Hikmet and R. Howard, Phys. Rev. E [**48**]{}, 2752 (1994). A. J. Liu and G. H. Fredrickson, Macromolecules [**29**]{}, 8000 (1996). J. R. Dorgan, J. Chem. Phys. [**98**]{}, 9094 (1993). Y. Lansac, F. Fried, and P. Maissa, Liq. Cryst. [**18**]{}, 829 (1995). P. G. de Gennes and J. Prost, [*The Physics of Liquid Crystals*]{} (Oxford University Press, New York, 1993). P. C. Hohenberg and B. I. Halperin, Rev. Mod. Phys. [**49**]{}, 435 (1977). C. Sagui, A. M. Somoza, and R. C. Desai, Phys. Rev. E [**50**]{}, 4865 (1994). R. Holyst and M. Schick, J. Chem. Phys. [**96**]{}, 721 (1992). A. J. Liu and G. H. Fredrickson, Macromolecules [**26**]{}, 2817 (1993). For a review of computer simulations of polymer blends, see S. C. Glotzer, [*Annual Reviews of Computational Physics*]{} [**2**]{}, 1 (1995). P. R. ten Wolde and D. Frenkel, Science [**277**]{}, 1975 (1997). P. D. Olmsted, W. C. K. Poon, T. C. B. McLeish, N. J. Terrill, and A. J. Ryan, Phys. Rev. Lett. [**81**]{}, 373 (1998).
{ "pile_set_name": "ArXiv" }
--- abstract: 'This paper describes the mixtures-of-trees model, a probabilistic model for discrete multidimensional domains. Mixtures-of-trees generalize the probabilistic trees of @chow:68 in a different and complementary direction to that of Bayesian networks. We present efficient algorithms for learning mixtures-of-trees models in maximum likelihood and Bayesian frameworks. We also discuss additional efficiencies that can be obtained when data are “sparse,” and we present data structures and algorithms that exploit such sparseness. Experimental results demonstrate the performance of the model for both density estimation and classification. We also discuss the sense in which tree-based classifiers perform an implicit form of feature selection, and demonstrate a resulting insensitivity to irrelevant attributes.' author: - | Marina Meilă [email protected]\ Department of Statistics\ University of Washington\ Seattle, WA 98195-4322, USA Michael I. Jordan [email protected]\ Division of Computer Science and Department of Statistics\ University of California\ Berkeley, CA 94720-1776, USA bibliography: - 'sample.bib' title: Learning with Mixtures of Trees --- Bayesian Networks, Mixture Models, Chow-Liu Trees Introduction ============ Probabilistic inference has become a core technology in AI, largely due to developments in graph-theoretic methods for the representation and manipulation of complex probability distributions [@pearl:88]. Whether in their guise as directed graphs (Bayesian networks) or as undirected graphs (Markov random fields), *probabilistic graphical models* have a number of virtues as representations of uncertainty and as inference engines. Graphical models allow a separation between qualitative, structural aspects of uncertain knowledge and the quantitative, parametric aspects of uncertainty...\ [*Remainder omitted in this sample. See http://www.jmlr.org/papers/ for full paper.*]{} Appendix A. {#app:theorem .unnumbered} =========== In this appendix we prove the following theorem from Section 6.2: [**Theorem**]{} [*Let $u,v,w$ be discrete variables such that $v, w$ do not co-occur with $u$ (i.e., $u\neq0\;\Rightarrow \;v=w=0$ in a given dataset ${{\cal D}}$). Let $N_{v0},N_{w0}$ be the number of data points for which $v=0, w=0$ respectively, and let $I_{uv},I_{uw}$ be the respective empirical mutual information values based on the sample ${{\cal D}}$. Then $$N_{v0} \;>\; N_{w0}\;\;\Rightarrow\;\;I_{uv} \;\leq\;I_{uw}$$ with equality only if $u$ is identically 0.*]{} [**Proof**]{}. We use the notation: $$P_v(i) \;=\;\frac{N_v^i}{N},\;\;\;i \neq 0;\;\;\; P_{v0}\;\equiv\;P_v(0)\; = \;1 - \sum_{i\neq 0}P_v(i).$$ These values represent the (empirical) probabilities of $v$ taking value $i\neq 0$ and 0 respectively. Entropies will be denoted by $H$. We aim to show that ${\frac{\partial I_{uv}}{\partial P_{v0}}} < 0$....\ [*Remainder omitted in this sample. See http://www.jmlr.org/papers/ for full paper.*]{} 0.2in
{ "pile_set_name": "ArXiv" }
--- author: - 'D. Moss' - 'V. V. Pipin' - 'D. Sokoloff' - 'J. T. Hoeksema' date: 'Received ..... ; accepted .....' title: Reversals of the solar magnetic dipole in the light of observational data and simple dynamo models --- =1 Introduction {#intro} ============ Our research presented in this paper starts from the following quite simple statements. The photospheric solar magnetic dipole reverses its orientation every 11 years, in the course of the solar activity cycle. It is believed that the solar cycle is driven by a dynamo mechanism operating deep in the convection zone. The evolution of the solar dipole is synchronized with sunspot activity. Details of this synchronization may depend on the underlying dynamo mechanism @PK13. Fig. \[inv1\] shows reversals of the dipolar component of the large-scale magnetic field of the Sun. ![Evolution of the dipole component of the radial magnetic field as obtained from measurements at the Wilcox Solar Observatory (WSO) from 1976 to 2013.[]{data-label="inv1"}](WSO_inv){width="0.99\columnwidth"} The Sun is basically axisymmetric so a natural expectation is that the solar magnetic field is also approximately axisymmetric. Correspondingly, a standard mean-field description of the solar dynamo assumes axial symmetry and that the solar dipole vanishes at the instant of reversal. In contrast, observational data (Livshits & Obridko 2006; DeRosa et al. 2012) show that the solar magnetic dipole rotates from pole to pole during the course of a reversal, even becoming orthogonal to the rotation axis at some instant, corresponding obviously to the vanishing of the polar dipole. At such a time the total magnetic field is strongly nonaxisymmetric. Moss, Kitchatinov & Sokoloff (2013) suggested that this apparent contradiction is explicable as a manifestation of contributions from [ fluctuations of the surface magnetic field (specifically its poloidal component)]{} near the moment of reversal. The aim of this paper is to develop and present a quantitative model that allows direct comparison of the interpretation of Moss et al. (2013) with observations. Using a simple simulation of this process, we produce models in which the trajectory of the dipole axis on the surface mimics well the observed behaviour. Other properties of the model also resemble those observed. In this paper, we concentrate on the fact (known for some time from solar activity studies, but little exploited) that the large-scale magnetic field becomes substantially nonaxisymmetric during a reversal. Quantifying this nonaxisymmetry in terms of [ fluctuations of the surface solar magnetic field]{}, we suggest that something can be learned about the relative contributions of mean-field and small-scale dynamos to the solar magnetic field (Sect. \[disc\]). We note that the relationship between dipolar and quadrupolar modes in solar dynamo action was addressed by DeRosa et al. (2012) and we do not discuss it here. The manifestation of a substantial nonaxisymmetric component of the solar dipolar magnetic field is interpreted in the framework of mean-field dynamo theory. One further point is that the apparent discrepancy between dynamo theory and the observations applies explicitly (in the form discussed above) only to mean-field dynamos. Of course, the solar magnetic field obtained from direct numerical simulations (DNS) is not strictly axisymmetric – even for hydrodynamics which is axisymmetric on average. In principle, one might prefer to use DNS and thus avoid any mean-field interpretation with its associated uncertainties, rather than involving magnetic fluctuations in the mean-field description. Of course, at the moment DNS models are expensive in terms of computing resources required. A hint that the sort of behaviour studied here may not be incompatible with DNS is given by Brown et al. (2011), [ where one model appears to show a latitudinal migration of the dipolar component]{} . We believe that the mean-field formulation of dynamo theory is a fruitful way to obtain a qualitative understanding of the solar cycle, and then compare it with observations and formulation. Thus the formulation of an adequate description of the mean-field behaviour of the solar cycle is a desirable undertaking. One outcome of the investigation presented here is the recognition that the observational analysis isolates some features of the solar dynamo that appear instructive and informative in various other contexts related to solar dynamos. We discuss such byproducts at the end of the paper (Sect. \[disc\]). Methodology {#method} =========== Our tactic in comparing theory and observations is as follows. First we must understand the capability of a mean-field dynamo model in explaining the phenomenon under discussion. A reasonable approach is to take the simplest dynamo model, compare it with observations and then add more and more realistic details until the desired agreement is achieved. We start with the illustrative Parker (1955) dynamo model and find that it is adequate to provide the mean-field component of the theoretical modelling. This makes using a detailed mean-field model unnecessary at this stage. Of course, we recognize the purely illustrative nature of this toy model (for example, the incorrect phase relation between toroidal and poloidal magnetic fields); however such issues are not relevant for the point being discussed. Note that the primitive nature of this mean-field model allows us [ to introduce a parameter which quantifies the link between]{} toroidal and poloidal magnetic field and so avoid any specific connection with the $\alpha$-effect introduced by Steenbeck, Krause and Rädler (see Krause and Rädler 1980). Thus the results should also be applicable to flux-transport models (e.g. Choudhuri 2008). Standard mean-field dynamo models do not contain explicit magnetic fluctuations and we have to include them somehow. This can be done in various forms and again we do it in the simplest way, i.e. by adding magnetic fluctuations “by hand” using a random number generator, as described in Sect. \[dyn\]. This is found to be sufficient to reproduce the observational phenomenology. The reversal phenomenon is quite robust and does not depend on details of the model. We see below that the magnetic fluctuation properties from our analysis are meaningful for understanding the surface manifestations of solar dynamo action. Thus there is no motivation for considering more realistic dynamo models. Of course the true solar dynamo is much more complicated than the toy model we use to mimic the observations. Various approaches have been discussed for determining the position of the solar magnetic pole, see, for example, [@Kukl88] and [@pat2013]. A delicate feature of our analysis is the fact that the observational data exploited are observations of the magnetic field at the solar surface. Such a field is mainly contributed by the magnetic fields that originate in solar active regions, which in turn are mainly determined by the toroidal magnetic field within the Sun. On the other hand, we are interested in the solar dipole magnetic moment and the toroidal field gives no contribution to this quantity. Of course, the fact that only a tiny part of the surface magnetic field determines the quantity being discussed requires some care in using the data. We note however that the situation is recognized by solar observers who have verified the reliability of the result with great care. ![Temporal evolution of the axisymmetric component of the solar magnetic dipole (red), its Gaussian filtering with a 1-year window (black) and the residual noise (blue), compared with sunspot number.[]{data-label="dip1"}](WSOD){width="0.95\columnwidth"} Observational data {#obs} ================== [ In the paper we analyze the components of the global magnetic field deduced from the daily magnetograph observations that have been made at the Wilcox Solar Observatory (WSO) at Stanford since May of 1976. The solar magnetograph measures the line-of-sight component of the photospheric magnetic field with three arc min resolution (see details in [@Sch77] and [@du77]). This homogeneous data set provides information about the evolution of the photospheric magnetic field through the past three Solar Cycles, 21, 22 and 23. The daily magnetograms were interpolated on to the Carrington grid and assembled into synoptic charts which were further used to determine the coefficients of the spherical harmonics of the potential magnetic field outside of the Sun. The procedure is described in detail by Hoeksema & Scherrer (1986). Below, we discuss some particular points which are essential for the current study.]{} a)![ a) The components of the solar equatorial dipole; b) the amplitude of the equatorial dipole.[]{data-label="orig"}](bxn "fig:"){width="0.85\columnwidth"}\ b)![ a) The components of the solar equatorial dipole; b) the amplitude of the equatorial dipole.[]{data-label="orig"}](fig5a "fig:"){width="0.95\columnwidth"}\ We consider the spherical harmonic decomposition of a scalar potential $$\begin{aligned} \psi=R\sum_{n=1}^{\infty}\sum_{m=0}^{n}\left(\frac{R}{r}\right)^{n+1}[g_{n}^{(m)}\cos\left(m\phi\right)+\nonumber \\ +h_{n}^{(m)}\sin\left(m\phi\right)]P_{n}^{m}\left(\theta\right).\end{aligned}$$ The potential magnetic field above the solar surface can be represented as a gradient of a scalar potential, ${\mathbf B}=-{\mathbf \nabla}\psi$. The data reduction uses the calculations of the radial harmonic coefficients of the photospheric magnetic field described by [@zhhx93], see also . Hoeksema & Scherrer (1986). [ Note, that the fitting procedure is made using the radial magnetic field as a boundary condition (see, [@zhhx93])]{}. Coefficients of the spherical harmonics are computed for each 10 degrees of Carrington rotation. In our analysis we use the components of the axial and equatorial dipoles provided by the first three coefficients of the potential field extrapolation of the photospheric magnetic field, namely, $g_{1}^{(0)}$, $g_{1}^{(1)}$ and $h_{1}^{(1)}$. These three coefficients give the dipole part of the radial component of magnetic field: $$\begin{aligned} B_{r}=2\left(\frac{R}{r}\right)^{3}[g_{1}^{(0)}\cos\theta-(g_{1}^{(1)}\cos\phi+\\ +h_{1}^{(1)}\sin\phi)\sin\theta].\label{Br}\end{aligned}$$ From these three coefficients we define the inclination of the effective dipole and its azimuth by $$\begin{aligned} g^{(h)} & = & \sqrt{ {g_{1}^{(1)}}^2+{h_{1}^{(1)}}^2}, \label{amp}\\ \Theta & = & \arctan\left(\frac{g_{1}^{(0)}}{g^{(h)}} \right),\label{tilt} $$ where $\Theta$ is the inclination of the dipole, and $g^{(h)}$ is amplitude of the equatorial dipole. The components $g_{1}^{(1)}$ and $h_{1}^{(1)}$ determine the Carrington longitude of the equatorial dipole as $\arctan\left(h_{1}^{(1)}/g_{1}^{(1)}\right)$, its azimuth $\Lambda$ is calculated as a sum of the Carrington longitude and the phase offset, $\Phi_{{\rm Car}}$. [ Figures \[dip1\] and \[orig\]a,b show the evolution of the components of the dipolar field of the Sun.]{} A dynamo model {#dyn} ============== Our idea is to start with a simple axisymmetric dynamo model and to add small-scale nonaxisymmetric injections of poloidal magnetic field, distributed randomly in area over the whole surface of the sphere. For the underlying (axisymmetric) dynamo model we use an extension of the 1D Parker model, as described by Moss, Sokoloff, Kuzanyan & Petrov (2004)[, see also Moss, Saar & Sokoloff (2008)]{}. The governing equations for the toroidal field $B(\theta)$ and potential $A(\theta)$ for the poloidal field are $$\frac{\partial B}{\partial t}=Dg\sin\theta\frac{\partial A}{\partial\theta}+\frac{\partial^{2}B}{\partial\theta^{2}}-\mu^{2}B, \label{eqB}$$ $$\frac{\partial A}{\partial t}=\alpha B+\frac{\partial^{2}A}{\partial\theta^{2}}-\mu^{2}A. \label{eqA}$$ Here $\theta$ is the polar angle, the factor $\mu$ is introduced to represent radial diffusion, and we take $\mu=3$ as appropriate for a dynamo region of about $30\%$ in radius. The factor $g(\theta)$ is introduced to allow a representation of latitudinal variations in the rotation law – we take $g=1$ – and take the dynamo number $D=-10^{3}$. Of course it would be possible to use a more sophisticated, two-dimensional, dynamo model. However we are not here studying details of the overall field behaviour as a function of radius, but rather we only need a mechanism to produce approximately sinusoidal variations of radial field at the surface. a)![a) Evolution of the magnetic dipole components from the model data (with time independent poloidal field injection amplitude of arbitrary sign). b) Evolution of azimuth (see comments in the text). [ If the half-period (0.176) of the oscillations in panel a) is identified with the 11 yr sunspot cycle, then the time unit is about 62.5 yr.]{}[]{data-label="mod"}](model_d0 "fig:"){width="0.94\columnwidth" height="0.16\paperheight"} b)![a) Evolution of the magnetic dipole components from the model data (with time independent poloidal field injection amplitude of arbitrary sign). b) Evolution of azimuth (see comments in the text). [ If the half-period (0.176) of the oscillations in panel a) is identified with the 11 yr sunspot cycle, then the time unit is about 62.5 yr.]{}[]{data-label="mod"}](fig6b "fig:"){width="0.95\columnwidth" height="0.16\paperheight"} Once our model has settled to a regular oscillatory behaviour [ we map the axisymmetric field from the mean field dynamo on to the surface of a sphere. Then]{} we add, at fixed time intervals $dt_{{\rm inj}}$, co-rotating patches of radial field, strength $B_{{\rm inj}}$ [ of arbitrary sign]{}, in circular regions at randomly located positions on the surface of the sphere. In the models described, these regions are of radius $20-30^{\circ}$. They are intended to represent injections of field from solar active regions - we are only interested in the resulting contributions to the global poloidal field. [The resulting global poloidal field is thus nonaxisymmetric, being the sum of the symmetric dynamo generated field and the injections which are random in longitude (and in latitude). In our modelling the injected field is just added formally to the poloidal field output from Eqs. (\[eqB\]), (\[eqA\]) – the injected field is not input back into the dynamo equations. We can justify this procedure by noting that the dynamo is the result of processes occurring deep within the Sun, and can hardly be affected by the superficial phenomena associated with surface activity (“the tail does not wag the dog”).]{} The injected field has strength of order the maximum dynamo poloidal field and decays with time constant $t_{{\rm dec}}$. The code allows multiple spots to coexist, with each generated successively after the fixed interval $dt_{\rm inj}$. If appropriate, we can modulate the injection amplitude with the phase of solar cycle. The evolution of the dipole components [ for a typical realization of the model]{} is illustrated in Fig. \[mod\]a. The result is not very sensitive to the input parameters. The azimuth of the dipole is random, because of the nature of the field injection mechanism. [ It is determined by the position of the instantaneous dipole, which in turn is given by the most recent surface field fluctuations. The evolution of the toroidal field follows the evolution of the axial dipole with phase shift of about $\pi/2$, and with amplitude of the toroidal field larger by a factor of about 40 compared to the poloidal. (This ratio could easily be adjusted by changing the parameters in Eqs. (\[eqB\]), (\[eqA\]), without changing our general conclusions.)]{} Fig. \[mod\]b illustrates the evolution of azimuth, smoothed using a Gaussian filter with a window which corresponds to one year in the model. [Any apparent correlation in the evolutionary paths in Fig. 6 appears to be coincidental.]{} Results ======= Basic analysis of the data {#basic} -------------------------- ![Spectrum of the equatorial dipole components.[]{data-label="spek"}](spektrn){width="0.95\columnwidth"} For the purpose of analysis we have to separate the mean and fluctuating parts of the field and then do the same for the inclination and azimuth of the dipole. For the axisymmetric part of the dipole, which is represented by $g_{1}^{(0)}$, this can be done by suitable filtering of the data in the time series. We use a Gaussian filter with a one year window, which is often used in analysis of the sunspot data set [@hath09]. Fig. \[dip1\] shows the evolution of $g_{1}^{(0)}$, its mean and noise components, and the smoothed sunspot number as given by the [@sidc] data base. [ Note that the noise is [*not*]{} the fluctuating part of the field (as defined in Sect. \[dyn\]), but the numerical residual after data analysis.]{} a)![a) The transformed components of the equatorial dipole, rotation period $P=26.90$ days; b) the upper curve shows the azimuth of the equatoial dipole in the frame rotating with the Carrington period P=27.253 days, the lower curve shows the azimuth for the mode P=26.90 days.[]{data-label="transf"}](byn "fig:"){width="0.85\columnwidth"} b)![a) The transformed components of the equatorial dipole, rotation period $P=26.90$ days; b) the upper curve shows the azimuth of the equatoial dipole in the frame rotating with the Carrington period P=27.253 days, the lower curve shows the azimuth for the mode P=26.90 days.[]{data-label="transf"}](fig5bn "fig:"){width="0.95\columnwidth"} Now we move to the analysis of the behaviour of the equatorial dipole, and our first aim is to determine the rotation rate of this dipole. For this purpose we determine Fourier spectra of $g_{1}^{(1)}$ and $h_{1}^{(1)}$ using the standard fast Fourier transform (FFT) routines provided by Scientific Python (www.scipy.org). Fig. \[spek\] shows the result. Very similar spectra were found earlier by Svalgaard & Wilcox (1975) and [@kotov83] from analysis of the rotation of the interplanetary magnetic field. We see that the spectra have peaks in a range of synodic rotation periods between 26 and 30 days, including the Carrington period of 27.2753 days. We recall that the synodic rotation period refers to motion as observed from Earth. Our interpretation of the plot is that the azimuthal dipole field is, of course, not just white noise and that its memory is sufficiently long to feel the dipole rotation and to identify at least the main peak in the spectra, which corresponds to the rotation period $P=26.90$ day. The scatter of the periods in Fig. \[spek\] could be a consequence of the differential rotation. In this paper we restrict our analysis to a particular mode with period $P=26.90$ day. We perform a filtering of the azimuthal dipole data in the frame rotating with the period $P=26.90$ day using the following transformation ![Evolution of the large-scale dipolar magnetic field components plotted with the sunspot number data. \[datam\]](fig9n){width="0.95\columnwidth"} $$\begin{aligned} g_{1}^{(1)\prime} & = & g_{1}^{(1)}\cos\Omega t+h_{1}^{(1)}\sin\Omega t\label{transfe}\\ h_{1}^{(1)\prime} & = & h_{1}^{(1)}\cos\Omega t-g_{1}^{(1)}\sin\Omega t.\nonumber\end{aligned}$$ [ Fig. \[transf\]a shows evolution of the components obtained in result of this transformation]{} for the case $\Omega=2\pi/P$, $P=26.90$ day. [ Of course, similar ]{}results can be obtained for other periods from the given range. The amplitude of the equatorial dipole remains the same under transformation. [ Fig. \[transf\](b) shows how different the azimuth can be if we follow the original equatorial dipole components (Fig. \[orig\]) and the components of the mode $P=26.90$ day. For example, we observe the quasi-regular large-scale fluctuation of this mode during the descending phase of cycle 23, when the mode was permanently visible around longitude $260^\circ$ for a period of about 3 years [**starting around the year 2002.**]{} [**The existence of such events is also suggested by results of thorough period analysis (see, e.g., Berdyugina et al 2006)**]{} We have to stress that the our particular procedure of analysis is employed because of our restriction to consideration of the first three modes of the decomposition given by the definitions of azimuth from Eq. (1). The azimuth can be uncertain if the signal has a complicated spectrum like that shown by Figure \[spek\]. In the literature, the reader can find some alternative possibilities for determination of the azimuth of the solar dipole (e.g. [@Kukl88; @pat2013]). However, our method allows direct comparison of results with predictions of the dynamo models.]{} The nonaxisymmetric modes have the largest variations in time. Using the modes in the rotating coordinate frame we will define the inclination and the azimuth of dipole for the same reference frame. For the remainder of our illustrations we use the mode corresponding to the period $P=26.90$ day. This value corresponds to the position of the highest peak in Fig. \[spek\]. Clearly, in our definition the azimuth of the global dipole has an arbitrary zero point. On the other hand, it is well-known that the given characteristics of the large-scale magnetic field of the Sun trace the sectorial structure of the interplanetary magnetic field in the heliosphere [@WSO2]. A problem in determination of the global dipole azimuth is that the two-sector structure, which is observed during much of the cycle, changes to a four sector structure during the maxima of the sunspot cycle. The sectorial structure reflects the total contribution of all the non-axisymmetric modes. It can be used for the azimuth determination only during the phases when the two-sector structure dominates. [ We show the long-term evolution of the magnitude of the total dipole strength of the solar field in Fig. \[datam\], together with the sunspot number data. The Sun currently appears to be presenting rather atypical behaviour in the context of the last century or so. The ongoing decline in activity emphasizes that we are not dealing with a strictly periodic system and that our modelling, by assuming an underlying regular periodicity, can only hope to give a partial representation.]{} ![The auto-correlation function of the magnetic dipole fluctuations (mean values are subtracted)[]{data-label="corr"}](avto_corN){width="0.9\columnwidth"} \[resul\] Axial dipole data {#axialdip} ----------------- We calculate the rms value of the noise $\delta d$, the amplitude of the magnetic dipole cycle variations $\bar{d}$ and its ratio $\delta d/\bar{d}$, both for the observations ($d$ is estimated for cycle 22) and the model to get the following results. The observations give $\delta d=0.24$, $\bar{d}=2.2$, $\delta d/\bar{d}=0.11$. For the amplitude of the observed equatorial dipole we get similar results (see Fig. \[orig\]b and Fig. \[datam\]). Here magnetic field is normalized to the solar radius $R$ and magnetic moment is measured in units of Gauss $R^{3}$. For illustration, we compare with [ one of our models]{} which has a half period (corresponding to the 11 yr sunspot cycle) of 0.176 in our dimensionless units, $dt_{{\rm inj}}=0.0015=dt_{{\rm dec}}$, and 5 spots are simultaneously present (although by the time a spot is removed from the simulation, it has been present for 5 decay times, and so its field is essentially reduced to zero). The model gives $\delta d=0.00085$, $\bar{d}=0.015$, $\delta d/\bar{d}=0.058$. Here magnetic field is measured in units of the equipartition magnetic field. For an order of magnitude estimate, we assume that the magnetic field unit in the model corresponds to $10^3$ G (typical field strength of sunspots) and then normalize to the solar radius. This gives $\delta = 15$ Gauss $R^3$ and $\delta d = 0.85$ Gauss $R^3$. Taking into account the very crude nature of the model, this seems to be in reasonable agreement with observations. [ In physical units, the injection interval $dt_{\rm inj}$ is about 0.094 yr (ca 1.1 months), which we chose rather arbitrarily to be equal to the decay time $t_{\rm dec}$.]{} We conclude that the model produces fluctuations of the magnetic dipole that are comparable to those found in the observations. The general nature of the results was not very sensitive to the choice of parameters[ , provided that the interval between injections is not too long or very short, or that the injected field strength is not much larger or smaller than the maximum dipole field strength produced by the dynamo. The results are quite insensitive to $t_{\rm dec}$, provided that it is not very long, or to the number of spots present provided that $dt_{\rm in j}$ and $dt_{\rm dec}$ are not very different in magnitude]{}. We can estimate the relative duration of the reversal epoch $t^{*}$ as follows. If the cyclic behaviour is sinusoidal, we can estimate the time during which $\bar{d}\sin t$ is lower than $\delta d$. A simple calculation gives $\delta d/\bar{d}=0.06$, and for an 11-year cycle $t^{*}\approx4$ month. The autocorrelation function of the noise is presented in Fig. \[corr\]. The correlation time of the fluctuations can be estimated from the plot as $\tau_{{\rm cor}}\approx 0.18$ yr$=65 {\rm d} \approx 2$ months, i.e. about half the duration of the reversal epoch. Note, that parameters of fluctuations of the axial dipole and the amplitude of the equatorial dipole are rather similar. We estimate $b/(B\sqrt{N})B_P/B\approx\delta d/\bar{d}$ [ (here $B$ is the mean magnetic field field and $B_P$ its poloidal component)]{} following Eq. (1) of Moss et al. (2013), where $B$ is the large-scale magnetic field strength in the near-surface layers within the Sun, $b$ is the corresponding strength of magnetic fluctuations and $N$ is number of convective cells in the domain of dynamo action. Following Moss et al. (2013) we take as a crude estimate $N=10^{3}$ [ and $B_P/B \approx 0.1$]{} and arrive at an estimate $b/B=2$, which is more or less in agreement with the estimate obtained by Sokoloff & Khlystova (2010) from statistics of the sunspot groups that violate the Hale polarity law. The direction of the magnetic dipole at the epoch of reversal ------------------------------------------------------------- We can follow the evolution of the total magnetic dipole during the epoch of reversal by plotting the [ position of the]{} end of the dipole vector as a point on a sphere of unit radius (Fig. \[inclin\]a). We conclude that the reversal track for the model (Fig. \[inclin\]c) is quite similar to that deduced from observations. Of course, a shorter filtering time gives a more complicated trajectory for the dipole. [ We give an alternative presentation of the dipole track in Fig. \[thetalamb\].]{} a)![ The track of the [ pole the of the magnetic dipole]{} on a sphere of unit radius. a) WSO data using a Gaussian filter with a one year window for the model described in Fig. \[mod\]; b) the same for the half year window; c) the same for the model – see Fig. \[mod\] – which, in contrast to the data, covers an interval with 6 reversals of the global dipole.[]{data-label="inclin"}](WSO3D1 "fig:"){width="0.85\columnwidth"}\ b)![ The track of the [ pole the of the magnetic dipole]{} on a sphere of unit radius. a) WSO data using a Gaussian filter with a one year window for the model described in Fig. \[mod\]; b) the same for the half year window; c) the same for the model – see Fig. \[mod\] – which, in contrast to the data, covers an interval with 6 reversals of the global dipole.[]{data-label="inclin"}](WSO3D2 "fig:"){width="0.85\columnwidth"}\ c)![ The track of the [ pole the of the magnetic dipole]{} on a sphere of unit radius. a) WSO data using a Gaussian filter with a one year window for the model described in Fig. \[mod\]; b) the same for the half year window; c) the same for the model – see Fig. \[mod\] – which, in contrast to the data, covers an interval with 6 reversals of the global dipole.[]{data-label="inclin"}](3d_mod0 "fig:"){width="0.85\columnwidth"} a)![The track of the [ pole of the magnetic dipole]{} in the $\theta-\lambda$ plane: a) WSO data, b) the same for the model (Fig. \[mod\]) which, [ in contrast to the 4 reversals in the WSO data]{}, covers an interval with 6 reversals of the global dipole.[]{data-label="thetalamb"}](hodWSO "fig:"){width="0.85\columnwidth"}\ b)![The track of the [ pole of the magnetic dipole]{} in the $\theta-\lambda$ plane: a) WSO data, b) the same for the model (Fig. \[mod\]) which, [ in contrast to the 4 reversals in the WSO data]{}, covers an interval with 6 reversals of the global dipole.[]{data-label="thetalamb"}](hodmod "fig:"){width="0.85\columnwidth"} We find that the trajectories of the [ pole of the magnetic dipole]{} depend on the method of smoothing applied to the WSO data as well as for the model data. These plots are quite similar to those shown in Fig. \[inclin\]b and we do not present them here. Using the model data, we demonstrated that the longitude of reversal tracks is distributed on the sphere quite randomly. Unfortunately, there are no corresponding long-term observational data to confirm this point observationally. Cyclic evolution of the amplitude of magnetic fluctuations ---------------------------------------------------------- According to the interpretation of the observational data suggested, the equatorial magnetic dipole is a large-scale manifestation of magnetic fluctuations. A notable feature of the plots describing the equatorial dipole (Figs. \[orig\],\[transf\]) is that the amplitude of the noise is modulated with the solar cycle period. We did not include such a feature in our model, so quite naturally the plot for the equatorial dipole for the model data (Fig. \[mod\]) does not show such an effect. If we include modulation of the fluctuations with the amplitude of the toroidal field variations, the desired modulation of the equatorial dipole appears. We conclude that the rms of the solar magnetic fluctuations has an 11-year periodicity. We find that this modulation survives if we introduce artificially a long-term evolution of the cycle amplitude, as has occurred during recent solar cycles (we omit the plot for the sake of brevity). We note that cyclic behaviour of the equatorial dipole was mentioned previously by Hoeksema (1995) and Obridko & Livshits (2006). Discussion and conclusions {#disc} ========================== Our analysis of the observational data, and comparison with a simple model of the solar dynamo, confirm that non-zero nonaxisymmetric values of the total magnetic dipole at the epoch of solar magnetic reversal are consistent with a large-scale contribution from [ fluctuations of the solar surface magnetic field]{}, as discussed by Moss, Kitchatinov & Sokoloff (2013). [ We envisage that these fluctuations, [**connected loosely**]{} to active regions, are driven by an instability of the underlying layers.]{} The analysis gives an estimate for the ratio of the rms value of the fluctuation to the amplitude of cyclic variations of the mean solar field of $b/B \approx 2$. This estimate is independent of the estimate of Sokoloff & Khlystova (2010), obtained from the statistics of sunspot groups that do not follow the Hale polarity law. These estimates differ by a factor of about 2, which appears quite satisfactory given that the amplitude of solar cycle varies substantially from cycle to cycle, and that the estimates come from analyses of quite different data. The random nature of the orientation of the solar magnetic dipole during the epoch of reversal is clearly visible from the trajectories of the [ pole of the magnetic dipole]{} on the unit sphere (Fig. \[inclin\]c). On the other hand, the epoch of reversal occupies a relatively small part of the solar cycle (about 4 months only). The duration of the reversal epoch is estimated as the time during which the fluctuating part of the dipole is larger than the mean-field contribution. In comparison, the memory time for the dipole fluctuations is estimated as about 2 months, i.e. about half the duration of the reversal epoch. This is why the dipole trajectory is far from just being noise, and some traces of regular behaviour are visible in the trajectory during a given reversal. This might be instructive for understanding of the phenomenon of active longitudes (e.g. Usoskin et al. 2007 and references therein). Our analysis gives further important information relevant to studies of the solar dynamo. We see that magnetic fluctuations are cyclically modulated. This means that they (at least mainly) originate in the periodic global dynamo action, rather than from small-scale dynamo action, as was suggested by Goode et al. (2012) and Abramenko (2013). Possibly this is because active regions are generated more readily in the parts of the cycle when toroidal fields are stronger – recall that the axisymmetric poloidal field is out of phase with the dynamo generated toroidal field. In other words, we confirm here the interpretation of Stenflo (2013) that a local dynamo does not give a significant contribution to the observed [small-scale]{} flux. Following this interpretation however, based on other observational data, we infer that, as far as we can confirm from the observations available, solar magnetic activity is a result of a single physical process which simultaneously generates the solar activity cycle as well as the magnetic fluctuations. We do not exclude in principle the possibility that an additional dynamo excitation mechanism for small-scale magnetic field is active somewhere in the solar interior. However we are unable to isolate its manifestations from the general cyclic behaviour associated with the main driver of solar magnetic activity. One more important point is that the small-scale magnetic fluctuations considered here result in weak nonaxisymmetric magnetic field components, as demonstrated by the equatorial dipole data (Figs. \[orig\], \[transf\]). An additional illustration of this fact is given in Fig. \[datam\] which plots the large-scale total dipole component of the solar activity, including both the axisymmetric and nonaxisymmetric parts. We see that the nonaxisymmetric components are stronger during maxima of sunspot activity, i.e. they are associated with the toroidal magnetic field. The lifetime of bursts of nonaxisymmetric components is of order 1 year. Any such large-scale components could be described by standard mean-field dynamo equations. Moss et al. (2013) estimated the lifetime of such nonaxisymmetric bursts as about 1 year, in reasonable agreement with observations. Our interpretation is that the nonaxisymmetric magnetic field components appear as a result of decay of the large-scale toroidal magnetic field. [ The typical correlation time of the nonaxisymmetric modes is estimated to be around 1 year. This agrees with results of the recent analysis of rotation of solar active regions made by [@pelt10]. They, following [@KR80], suggested the existence of non-oscillatory nonaxisymmetric modes which rotate rigidly with an angular velocity that is different from the overall rotation period. These modes are coupled to the global toroidal field and affected by differential rotation. This is consistent with the results presented in Fig. \[spek\]. Note that wavelet analysis by [@MPl00] reveals that nonaxisymmetric modes with different periods of rotations are excited during different epochs of the solar cycle, e.g. the mode with P=26.9 days is present during the decaying phase of activity, while modes with periods around 28 days are present during the rising phase of the solar cycle (cf. Fig. \[transf\]b).]{} Perhaps an equally valid interpretation is to link the rotation rates to the rotation rates of the remnant flux patterns. A more complicated approach might include the possibility of azimuthal dynamo wave excitation in spherical shell convection (Cole et al. 2013). Note, the rotational periods of the emerging nonaxisymmetric fields can be related to the “new magnetic flux” which might originate in a subsurface shear layer – @Beal1999. This is compatible with a distributed dynamo operating in the bulk of the convection zone and being shaped by the subsurface shear layer (). Obviously, our interpretation does not exclude the possibility that occasionally the equatorial magnetic dipole might vanish simultaneously with the strength of the axial dipole (which is part of the mean field) passing through zero during the course of an reversal. In such a case, at some instant during the reversal the total dipole will vanish. Remarkably (cf. Fig. \[datam\]), this seems to be happening just now (Obridko 2013). Of course, there are details of the data that we have not analyzed, which could affect our conclusions. These include, for example, the study of effects of the higher order components of the decomposition Eq(1) on reversals of polar field. However, such a study goes beyond the simple dynamo model discussed in this paper. [ In summary, we have presented a simple heuristic model – basically a cartoon – to illustrate a mechanism to reproduce the behaviour of the solar dipole during reversals. We have necessarily made a number of more-or-less arbitrary – but reasonable – assumptions. Our principle conclusion is that a model of this sort can quite satisfactorily represent the observed phenomena during a reversal. The mechanism seems quite robust, and quantitatively supports the idea proposed by Moss, Kitchatinov & Sokoloff (2013). Thus we feel confident that it captures the essence of the solar behaviour without performing excessive fine tuning. Moreover, we find that the detailed observational data concerning solar dipole reversals have, perhaps rather unexpectedly, the potential to reveal much about dynamo action in the solar interior. Clearly, there is considerable scope for a more detailed analysis of a more sophisticated model.]{} V.P. and D.S. are grateful to RFBR for financial support under grant 12-02-00170-a. Useful discussions with V.N.Obridko are acknowledged. [14]{} natexlab\#1[\#1]{} , V. I. 2013, in IAU Symposium, Vol. 294, IAU Symposium, ed. A. G. [Kosovichev]{}, E. [de Gouveia Dal Pino]{}, & Y. [Yan]{}, 289–300 , E. E., [Hoeksema]{}, J. T., [Kosovichev]{}, A. G., & [Scherrer]{}, P. H. 1999, , 517, L163 Berdyugina, S.V., Moss, D. , Sokoloff, D.D., Usoskin, I. G., 2006, A& A,445, 703 , 2011, , 731, 69 , A. 2005, , 625, 539 , A. R. 2008, Advances in Space Research, 41, 868 Cole, E., Käpylä, P.J., Mantere, M.J., Brandenburg, A., 2013, arXiv preprint 1309.6802, 2013 , M. L., [Brun]{}, A. S., & [Hoeksema]{}, J. T. 2012, , 757, 96 , Jr., T. L., [Wilcox]{}, J. M., [Svalgaard]{}, L., [Scherrer]{}, P. H., & [McIntosh]{}, P. S. 1977, , 55, 63 , Jr., T. L., [Scherrer]{}, P. H., [Svalgaard]{}, L., & [Wilcox]{}, J. M. 1979, , 61, 233 , P. R., [Abramenko]{}, V., & [Yurchyshyn]{}, V. 2012, , 86, 018402 Hathaway, D. 2009, Space Science Reviews, 144, 401, 10.1007/s11214-008-9430-4 Hoeksema, J. T. & Scherrer, P.H. 1986, Sol. Phys., 105, 205 Hoeksema, J. T. 1995, , 72, 137 , C. D., [Akasofu]{}, S. I., [Hoeksema]{}, J. T. & [Hakamada]{}, K., 1985, Planet.Space Sci., 33, 915 , V. A. & [Levitskii]{}, L. S. 1983, Izvestiya Ordena Trudovogo Krasnogo Znameni Krymskoj Astrofizicheskoj Observatorii, 68, 56 Krause, F. & Rädler, K.-H. 1980, Mean-Field Magnetohydrodynamics and Dynamo Theory (Berlin: Akademie-Verlag), 271 , I. M. & [Obridko]{}, V. N. 2006, Astr. Rep., 50, 926 , A. V. & [Plyusnina]{}, L. A. 2000, , 197, 1 , D., [Sokoloff]{}, D., [Kuzanyan]{}, K., & [Petrov]{}, A. 2004, Geophy. Astrophys. Fluid Dyn., 98, 257 Moss, D., Saar, S. & Sokoloff, D. 2008, MNRAS, 388, 416 , D., [Kitchatinov]{}, L. L. & [Sokoloff]{}, D. 2013, , 550, L9 , G. & [Kuklin]{}, G. V. 1988, Issledovaniia Geomagnetizmu Aeronomii i Fizike Solntsa, 79, 49 Obridko, V.N., All-Russian Astronomical meeting, St Petersburg, Abstract book, 2013, 202 Parker, E. 1955, , 122, 293 , J., [Inhester]{}, B., & [Cameron]{}, R. 2013, , 558, L4 , J., [Korpi]{}, M. J., & [Tuominen]{}, I. 2010, , 513, A48 , V. V. & [Kosovichev]{}, A. G. 2011, , 741, 1 Pipin, V. V. & Kosovichev, A. G. 2013, , 776, 36 , P. H., [Wilcox]{}, J. M., [Svalgaard]{}, L., [et al.]{} 1977, , 54, 353 , D. & [Khlystova]{}, A. I. 2010, Astr. Nachr., 331, 82 , J. O. 2013, in IAU Symposium, Vol. 294, IAU Symposium, ed. A. G. [Kosovichev]{}, E. [de Gouveia Dal Pino]{}, & Y. [Yan]{}, 119–130 Svalgaard, L. & Wilcox, J.M. 1975, Sol. Phys., 41, 461 Usoskin, I. G., Berdyugina, S. V., Moss, D. & Sokoloff, D.D., 2007, Adv. Sp. Res., 40, 951 . 2010, Monthly Report on the International Sunspot Number, online catalogue, http://www.sidc.be/sunspot-data/ , X. & [Hoeksema]{}, J. T. 1993, , 143, 41
{ "pile_set_name": "ArXiv" }
--- abstract: | In this paper we provide a ${\tilde{O}}(m\sqrt{n})$ time algorithm that computes a $3$-multiplicative approximation of the girth of a $n$-node $m$-edge directed graph with non-negative edge lengths. This is the first algorithm which approximates the girth of a directed graph up to a constant multiplicative factor faster than All-Pairs Shortest Paths (APSP) time, i.e. $O(mn)$. Additionally, for any integer $k \ge 1$, we provide a deterministic algorithm for a $O(k\log\log n)$-multiplicative approximation to the girth in directed graphs in ${\tilde{O}}(m^{1+1/k})$ time. Combining the techniques from these two results gives us an algorithm for a $O(k\log k)$-multiplicative approximation to the girth in directed graphs in ${\tilde{O}}(m^{1+1/k})$ time. Our results naturally also provide algorithms for improved constructions of roundtrip spanners, the analog of spanners in directed graphs. The previous fastest algorithms for these problems either ran in All-Pairs Shortest Paths (APSP) time, i.e. $O(mn)$, or were due Pachocki [*et al.*]{} [@PRSTV18] which provided a randomized algorithm that for any integer $k \ge 1$ in time $\tilde{O}(m^{1+1/k})$ computed with high probability a $O(k\log n)$ multiplicative approximation of the girth. Our first algorithm constitutes the first sub-APSP-time algorithm for approximating the girth to constant accuracy, our second removes the need for randomness and improves the approximation factor in Pachocki [*et al.*]{} [@PRSTV18], and our third is the first time versus quality trade-off for obtaining constant approximations. author: - | Shiri Chechik\ Tel Aviv University\ [[email protected]]{} [^1] - | Yang P. Liu\ Stanford University\ `[email protected]` [^2] - | Omer Rotem\ Tel Aviv University\ [[email protected]]{} - | Aaron Sidford\ Stanford University\ `[email protected]` [^3] bibliography: - 'refs.bib' title: | Constant Girth Approximation for Directed Graphs in\ Subquadratic Time --- Introduction ============ The *girth* of a graph $G$ is the length of the shortest cycle in $G$. It is an important graph quantity that has been studied extensively in both combinatorial settings (see Bollobás’s book [@Bollobas] for a discussion) and computational settings. In particular, exact algorithms for the girth running in time $O(mn)$ in weighted directed graphs [@OS17] are known. On the other hand, a result of Vassilevska W. and Williams show that a truly subcubic algorithm for girth (i.e. running in time $n^{3-{\varepsilon}}$ for some ${\varepsilon}> 0$) implies a truly subcubic algorithm for the All Pairs Shortest Path (APSP) problem [@VW10]. As it is a longstanding open problem whether APSP admits a truly subcubic time algorithm, exact computation of the girth in truly subcubic time would be a major breakthrough. This has motivated the study of efficient *approximation* algorithms for the girth. There has been extensive work on approximating the girth in *undirected* graphs [@IR77; @LL09; @RW12; @DKMS17]. Many such algorithms use the concept of a $\alpha$-*spanner* of a graph $G$, a fundamental combinatorial object which was introduced by Chew [@Chew89]. An $\alpha$-spanner of a graph $G$ is a subgraph of $G$ which multiplicatively preserves distances up to a factor of $\alpha$. It is well-known that $(2k-1)$-spanners with $O(n^{1+ 1/k})$ edges exist for any undirected weighted graph [@ADDJ93], and work on the efficient construction of such spanners [@TZ05; @LTZ05; @BS03] implies a $O(mn^\frac1k)$ time algorithm for $(2k-1)$-multiplicative girth approximation in undirected graphs. There has also been work on improved spanner constructions in the case of undirected unweighted graphs [@LL09; @RW12], and these algorithms also immediately imply algorithms for girth approximation in undirected unweighted graphs. Therefore, in order to obtain efficient constant factor girth approximations in directed graphs, it is natural to study an analog of spanners in directed graphs. Unfortunately, approximately computing all pairs distances in directed graphs is a notoriously difficult problem and while sparse spanners do exist in all undirected graphs, they do not exist in all directed graphs. For example, any directed spanner for the “directed" complete bipartite graph with $n$ vertices on the left directed towards $n$ vertices on the right clearly requires all $n^2$ edges. This problem seems to arise from the fact that the distance metric $d(u, v)$ in directed graphs is asymmetric. Therefore, if we want to construct sparse spanners, it is natural to work instead with the symmetric *roundtrip distance* metric, defined as $d(u {\leftrightarrows}v) := d(u, v) + d(v, u)$ [@CowenW04] and similarly define an $\alpha$-*roundtrip spanner* of a directed graph $G$ to be a subgraph that multiplicatively preserves roundtrip distances up to a factor of $\alpha$. Interestingly, there do exist roundtrip spanners for directed graphs with comparable sparsity as spanners for undirected graphs. A result of Roditty, Thorup, and Zwick [@RTZ08] shows that for any $k \ge 1$ and ${\varepsilon}> 0$, every graph has a $(2k+{\varepsilon})$-roundtrip spanner with $O(k^2 n^{1+ 1/k}\log(nW) {\varepsilon}^{-1})$ edges, where $W$ is the maximum edge weight. Unforunately, this algorithm ran in time $\Omega(mn)$, as it requires the computation of all pairs distances in the graph. Recent work Pachocki [*et al.*]{} [@PRSTV18] gave a randomized algorithm running in time $\tilde{O}(m^{1+1/k})$ which on weighted directed graphs $G$ returns a $O(k\log n)$-roundtrip spanner with $\tilde{O}(n^{1+1/k})$ edges and an $O(k\log n)$ approximation to the girth. Up to a logarithmic approximation factor, this matches the sparsity and runtime known for spanners on undirected weighted graphs and girth on sparse graphs. The result of Pachocki [*et al.*]{} [@PRSTV18] constitutes one of small, but rapidly growing [@CohenKPPRSV17; @CohenKKPPRS18], set of instances where it is possible to obtain robust nearly linear time approximations to fundamental quantities of directed graphs in nearly linear time, overcoming typical running time gaps between solving problems on directed and undirected graphs. However, a fundamental open problem left open by this work is whether it is possible to achieve subquadratic algorithms for constant factor approximation of the girth in directed graphs, and more ambitiously to fully close this gap and provide algorithms for $O(k)$ girth approximation and $O(k)$ roundtrip spanners in directed graphs that fully match the runtime and sparsity of those in undirected graphs. This is the primary problem this paper seeks to address and this paper provides multiple new girth approximation algorithms with improved runtime, approximation quality, and dependency on randomness. Our Results ----------- In this paper we provide a subquadratic algorithm for constant factor girth approximation in directed graphs and in turn show several improvements on the girth approximation algorithms and roundtrip spanner constructions in the work of Pachocki [*et al.*]{} [@PRSTV18]. Here and throughout the remainder of the paper we use $\tilde{O}(\cdot)$ notation to hide factors polylogarithmic in $n$, where $n$ is the number of vertices in the graph. In \[sec:constantapprox\] we consider obtaining constant approximations to the girth. In particular we provide a randomized algorithm that obtains a $3$-approximation to the girth on graphs with non-negative integer edge weights in ${\tilde{O}}(m \sqrt{n})$ time. Up to logarithmic factors this matches the runtime that would be predicted from the fact that $(2k - 1)$-undirected spanners with ${\tilde{O}}(n^{1+1/k})$ edges can be constructed in ${\tilde{O}}(mn^{1/k})$ time for $k = 2$. Further, we show that this procedure can be used to with high probability obtain constant multiplicative roundtrip spanners in directed graphs with arbitrary edge weights in ${\tilde{O}}(m \sqrt{n})$ time. \[thm:3girth\] For any directed graph $G$ with $n$ vertices, $m$ edges, integer non-negative edge weights, and unknown girth $g$ we can compute in $\tilde{O}(m \sqrt{n})$ time an estimate $g'$ such that $g \le g' \le 3g$ with high probability in $n$. \[thm:const\_spanner\] For any directed graph $G$ with $n$ vertices, $m$ edges, integer non-negative edge weights, we can compute in $\tilde{O}(m\sqrt{n})$ time an $8$-multiplicative roundtrip spanner with ${\tilde{O}}(n^{3/2})$ edges with high probability in $n$. Then, in \[sec:algo\] we give algorithms for a $O(k\log\log n)$-multiplicative approximation of the the girth and construct $O(k \log\log n)$ multiplicative roundtrip spanners with ${\tilde{O}}(n^{1+1/k})$ edges for a weighted directed graph $G$ with $n$ vertices and $m$ edges in $\tilde{O}(m^{1 + 1/k})$ time. These algorithms are deterministic and constitute the first deterministic nearly linear time algorithms for $\tilde{O}(1)$ multiplicative approximation of the girth and $\tilde{O}(1)$ multiplicative roundtrip spanners with $\tilde{O}(n)$ edges. \[thm:girth\] For any integer $k \ge 1$ and weighted directed graph $G$ with $n$ vertices, $m$ edges, and unknown girth $g$ we can compute in $\tilde{O}(m^{1 + 1/k})$ time an estimate $g'$ such that $g \le g' \le O(k \log\log n) \cdot g$. \[thm:spanner\] For any integer $k \ge 1$ and any weighted directed graph $G$ with $n$ vertices and $m$ edges, we can compute in $\tilde{O}(m^{1 + 1/k})$ time an $O(k \log\log n)$ multiplicative roundtrip spanner with $\tilde{O}(n^{1 + 1/k})$ edges. Setting $k = \frac{\log n}{\log\log n}$ yields the following corollaries. For $k = \Omega(\log n)$ these results nearly match the optimal algorithms in undirected graphs for $O(k)$ girth approximation and the construction of $O(k)$ spanners. For any weighted directed graph $G$ with $n$ vertices, $m$ edges, and unknown girth $g$ we can compute in $\tilde{O}(m)$ time an estimate $g'$ such that $g \le g' \le O(\log n) \cdot g$. For any weighted directed graph $G$ with $n$ vertices and $m$ edges, we can compute in $\tilde{O}(m)$ time an $O(\log n)$ multiplicative roundtrip spanner with $\tilde{O}(n)$ edges. Interestingly, our results for constant factor randomized approximations and our results for deterministic approximations are achieved in different ways. Highlighting this, in \[sec:klogk\_approx\] we show how to combine the techniques of these algorithms to obtain both $O(k \log k)$ multiplicative approximations to the girth and $O(k \log k)$ multiplicative roundtrip spanners of size ${\tilde{O}}(n^{1 + 1/k})$ in ${\tilde{O}}(m n^{1/k})$ time with high probability in $n$. \[thm:const\_girth\] For any integer $k \ge 1$ and any weighted directed graph $G$ with $n$ vertices, $m$ edges, and unknown girth $g$ we can compute in $\tilde{O}(m^{1 + 1/k})$ time an estimate $g'$ such that $g \le g' \le O(k \log k) \cdot g$ with high probability in $n$. \[thm:arb\_const\_spanner\] For any integer $k \ge 1$ and any weighted directed graph $G$ with $n$ vertices and $m$ edges, we can compute in $\tilde{O}(m^{1 + 1/k})$ time an $O(k \log k)$ multiplicative roundtrip spanner with $\tilde{O}(n^{1 + 1/k})$ edges with high probability in $n$. This shows that for any fixed ${\varepsilon}> 0$, that there is an algorithm running in time $m^{1+{\varepsilon}}$ that approximates the girth of a directed graph to within a constant depending on ${\varepsilon}$, but not on $m$ or $n$. Additionally, this almost matches the $O(k)$-multiplicative girth approximation algorithms running in $m^{1+1/k}$ time in undirected graphs. Comparison to previous work --------------------------- While the existence of roundtrip spanners matching the quality in undirected graphs was shown in [@RTZ08], the runtime was $O(mn)$ and required an APSP computation. Our results, \[thm:3girth\], \[thm:const\_spanner\] are the first to show that constant factor girth approximation and construction of constant factor roundtrip spanners with ${\tilde{O}}(n\sqrt{n})$ edges can be built in subquadratic ${\tilde{O}}(m\sqrt{n})$ time. This algorithm leverages new randomized techniques for testing a notion we call *similarity* between vertices not present in previous girth approximation and roundtrip spanner algorithms and we believe is of independent interest. Our \[thm:girth\] and \[thm:spanner\] offer direct improvements over the analogous results in [@PRSTV18]. Specifically, our algorithms provide a tighter multiplicative girth approximation and multiplicative spanner stretch in the same runtime as the algorithms in [@PRSTV18], which produce a $O(k \log n)$ girth approximation and $O(k\log n)$ roundtrip spanner with $\tilde{O}(n^{1 + 1/k})$ edges in time $\tilde{O}(m^{1 + 1/k}).$ Additionally, our algorithm is deterministic and in our opinion, simpler. The algorithm of Pachocki [*et al.*]{} [@PRSTV18] involved the following pieces. First, they use a method of Cohen to estimate ball sizes [@Cohen97] and resolve the case where there is a vertex whose inball and outball (of some small radius) intersect in a significant fraction of the vertices. In the other case, they use exponential clustering (see [@MPX13]) to partition the graph and recurse. Finally, they rerun the algorithm $n^{1/k}$ times. On the other hand, our algorithm simply grows inballs and outballs from various vertices, and uses a delicate cutting conditition to decide when to cut and recurse. Additionally, \[thm:const\_girth\] and \[thm:arb\_const\_spanner\] further improve upon Pachocki [*et al.*]{} [@PRSTV18] by combining the ideas from the constant factor girth approximation algorithm and the deterministic ball-growing algorithm, completely removing the dependence on $n$ in the approximation factor while still running in time ${\tilde{O}}(m^{1+1/k})$. We remark that the ideas for our deterministic algorithm are essential in obtaining this last result, and that more directly combining the ideas of [@PRSTV18] with our constant factor approximation algorithm does not seem to give an $O(k\log k)$ multiplicative girth approximation in ${\tilde{O}}(m^{1+1/k})$ time. Overview of Approach {#sec:approach_overview} -------------------- #### Summary of randomized $O(1)$ approach. Our approach to obtaining a $3$-approximation in \[sec:constantapprox\] to the girth is rooted in the simple insight that if a vertex $v$ is in a cycle of length $R$ then every vertex in the ball of radius $\alpha$ from $v$ is at distance at most $\alpha + R$ from every vertex in the cycle. Consequently, for each vertex if we repeatedly prune vertices from its outball of radius $R$ if they do not have the property that they can reach every vertex in this ball by traversing a distance at most $2R$, then we will never prune away vertices in a cycle of length $R$ from that vertex. Leveraging these insights, we can show that if we randomly compute distances to and from a random ${\tilde{O}}(\sqrt{n})$ vertices and if a cycle of length $O(R)$ is not discovered immediately then we can efficiently implement a pruning procedure so that each vertex only has in expectation ${\tilde{O}}(\sqrt{n})$ vertices that could possibly be in a cycle of length $O(R)$ through that vertex. By then checking each of these sets for a cycle and being careful about the degrees of the vertices (and therefore the cost of the algorithm) this approach yields essentially a $4$-approximation to the girth in ${\tilde{O}}(m\sqrt{n})$ time with high probability in $n$. Our $3$-approximation is then obtained by carefully applying this argument to both outballs and inballs and leveraging the simple fact that if a vertex $v$ is on a cycle $C$ of length $R$ then for every $c \in C$ either $d(v,c) \leq R/2$ or $d(c,v) \leq R/2$. #### Overview of deterministic $O(k \log \log n)$ results: Our deterministic algorithm in \[sec:algo\] is based on a different approach than our randomized constant approximation algorithms in \[sec:constantapprox\]. We think this approach is of independent interest and further demonstrate its utility in \[sec:klogk\_approx\] by showing how to combine the insights that underly it with the algorithm from \[sec:algo\] to achieve arbitrary constant approximations. For the sake of simplicity, we focus on unweighted directed graphs $G$ and for a parameter $R$, construct a subgraph (roundtrip spanner) $H$ so that if the roundtrip distance between $u$ and $v$ is at most $R$ in $G$, then their roundtrip distance is at most $O(Rk\log\log n)$ in $H$. The key insight of guiding our algorithm is the following: instead of partitioning the graph into disjoint pieces and recursing (as is done in [@PRSTV18]), we instead allow the pieces to overlap on the boundaries. This is justified by the following observation. Consider a subgraph $W$ of $G$, and let $W'$ denote the subgraph consisting of all vertices within distance $R$ of $W$. Then if we recursively build a roundtrip spanner on $W'$, then we are guaranteed that we can delete $W$ from our graph. Indeed, if $u \in W$ and the roundtrip distance between $u$ and $v$ is at most $R$, then $u, v \in W'$. This simple observation allows us to overcome the critical challenge in [@PRSTV18], arguing that that graph can be broken apart, while nevertheless preserving roundtrip distance. This observation also forms the basis of an optimal spanner construction on unweighted undirected graphs, which appears in a book of Peleg (exercise 3 on page 188 in [@Peleg00]). Specifically, for any integer $k \ge 1$, we can construct a $(2k-1)$-spanner with $O(n^{1 + 1/k})$ edges in time $O(m).$ The construction works as follows. Start at any vertex $v$, let $B_i$ denote the ball of radius $i$ centered at $v$, and let $|B_i|$ denote the number of vertices in $B_i$. Grow such balls around $v$ until we find an index $i$ with $|B_{i+1}| \le n^{1/k} |B_i|.$ We can clearly guarantee that $i \le k$. At this point, add a spanning tree on $B_{i+1}$ to your spanner and delete all vertices in $B_i.$ Now, recurse on the remaining graph. It is easy to check that the resulting spanner is as desired. Our algorithm for directed graphs is similar, and we give a more specific overview in \[sec:4overview\]. We gain an $O(\log\log n)$ dependence over the undirected spanner algorithm presented because we must recurse on the balls we grew instead of simply building a spanning tree on them. The precise condition for recursion and corresponding calculation are performed in the algorithms [<span style="font-variant:small-caps;">GoodCut</span>]{} (\[algo:goodcut\]) and \[lemma:inbound\]. Further, our $O(k\log k)$ approximations of \[sec:klogk\_approx\] are then achieved by using the techniques of the algorithms in \[sec:constantapprox\] to better control the size of the outballs and inballs in an invocation of the deterministic algorithm of \[sec:algo\]. Preliminaries {#sec:prelim} ============= For weighted directed graph $G$, we let $V(G)$ and $E(G)$ denote the vertex and edge sets of $G$. We assume all edge lengths are nonnegative. For a subgraph $S \subseteq G$ (not necessarily vertex induced), let $V(S)$ denote the set of vertices of $G$, and let $E(S)$ denote the set of edges. For a subset $W \subseteq V(G)$, we define $G[W]$ to be subgraph induced by $W$. When the graph $G$ is clear from context, we let $n$ and $m$ denote $|V(G)|$ and $|E(G)|$ respectively. For a weighted directed graph $G$ with non-negative edge lengths, we let $d_G(u, v)$ denote the (shortest path) distance from $u$ to $v$ in $G$. When the graph $G$ is clear from context, we simply denote this as $d(u, v).$ If there is no path from $u$ to $v$, we let $d(u, v) = \infty.$ When $S$ is a subgraph of $G$, we let $d_S(u, v)$ denote the (shortest path) distance from $u$ to $v$ only using the edges in $E(S).$ We denote the *roundtrip distance* between $u$ and $v$ as $d_G(u {\leftrightarrows}v) := d_G(u, v) + d_G(v, u)$ and define a roundtrip spanner. We say that a subgraph $S \subseteq G$ is an *$\alpha$-roundtrip spanner* if $d_S(u {\leftrightarrows}v) \le \alpha \cdot d_G(u {\leftrightarrows}v)$ for all $u, v \in V(G).$ For weighted directed graph $G$ we define the *inball* and *outball* of radius $r$ around a vertex $v$ as $${B^{\mathrm{in}}}_v(r) {:=}G[\{ u : d(u, v) \le r \}] \text{ and } {B^{\mathrm{out}}}_v(r) {:=}G[\{ u : d(v, u) \le r \}]$$ respectively. In other words, the inball of radius $r$ around $v$ is the subgraph induced by vertices $u$ with $d(u, v) \le r.$ The outball is defined similarly. We define the *ball* of radius $r$ around vertex $v$ as $$B_v(r) {:=}G[\{ u : d(u {\leftrightarrows}v) \le r \}].$$ In other words, the ball of radius $r$ around $v$ is the subgraph induced by vertices $u$ within roundtrip distance $r$ of $v$. Randomized Constant Approximations {#sec:constantapprox} ================================== Here we provide algorithms for efficiently computing a 3-approximation to the girth \[sec:girth3\]. To simplify our algorithm and analysis we assume that the maximum degree of $G$ is bounded by $O(m/n)$, i.e. we assume it is only a constant larger than the average degree, which is $2m/n$. We justify this assumption by showing that we can always reduce to this case as is formalized in the following lemma. We defer the proof to \[proofs:regular\]. \[lemma:regular\] Given a directed weighted graph $G=(V,E)$ of $n$ vertices and $m$ edges with non negative edge weights, one can construct a graph $H$ in $O(m)$ time of $O(n)$ vertices and $O(m)$ edges with non negative edge weights and of maximum degree $O(m/n)$ such that 1. All roundtrip distances (between pairs of vertices in $G$) in $H$ and in $G$ are the same. 2. Given a cycle in $H$, one can find in $O(m)$ time a cycle in $G$ of the same length. 3. Given a subgraph $H'$ of $H$, one can find in $O(m)$ time a subgraph $G'$ of $G$ such that the number of edges in $G'$ is at most the number of edges in $H'$ and the roundtrip distances in $H'$ and $G'$ are the same. An $\tilde{O}(m \sqrt{n})$ Time $3$-approximation to Girth {#sec:girth3} ----------------------------------------------------------- In this section we show a procedure that given a directed weighted graph $G$ and a girth estimate $R$, returns a cycle of length at most $3R$ if the girth in $G$ is at most $R$. The algorithm is given by ${\textsc{GirthApprox}}$ (See \[algo:girth3\]) which in turn invokes the subroutine ${\textsc{SimilarSet}}$ (See \[alg:similar\_vertices\]). In order to approximate the girth of $G$ we invoke this procedure for every $r=(1+\epsilon)^i$ for $1\leq i \leq \log_{1+\epsilon}{n W}$ and stop once the procedure returns a cycle. If $g$ is the girth of $G$ this incurs an additional $\log_{1+\epsilon}{g}$ factor to the running time (as for the first index $i$ such that $(1+\epsilon)^i >g$ the algorithm will return a cycle w.h.p.) and an additional $(1+\epsilon)$ factor in the approximation ratio. The additional $(1+\epsilon)$ factor in the approximation ratio can be avoided if the weights are integers by simply using binary search on the range between 1 and $nW$ (where $W$ is the maximum edge weight in $G$) and finding two consecutive integers $i$ and $i+1$ such that the procedure returned a cycle of length at most $3(i+1)$ when invoked on $i+1$ but not a cycle when invoked on $i$. This incurs a $\log{nW}$ factor in the running time that can be improved to $O(\log{n})$ by the same method as done in [@PRSTV18] of contracting small weight strongly connected components and deleting large weight edges (see Section 5.1 in [@PRSTV18] for more details). Let $G =(V,E)$ be a directed graph with $n$ vertices and $m$ edges. We assume the graph $G$ is of average degree $\delta = 2m/n$ and that also the maximum degree in the graph is also $O(\delta)$. The subroutine ${\textsc{SimilarSet}}$ gets as an input the graph $G$ and the target distance $R$ and either returns a cycle of length at most $3R$ or returns a subset $A_v$ of vertices for every $v\in V$. The subset $A_v$ for a vertex $v\in V$ consists of vertices at distance at most $R/2$ from $v$ with the guarantee that $A_v$ contains all vertices that are 1. At distance at most $R/2$ from $v$ and 2. On a cycle of length $R$ with $v$. Procedure ${\textsc{GirthApprox}}$ invokes the Procedure ${\textsc{SimilarSet}}$ twice, once on $G$ and once on the reversed graph of $G$ (the graph obtained by reversing every edge of $G$). If a cycle of length $3R$ is returned in one of these calls then procedure ${\textsc{GirthApprox}}$ returns such a cycle. Otherwise, let $\{A_v^{\mathrm{in}}\}_{v \in V}$ be the sets returned from invoking ${\textsc{SimilarSet}}$ on the graph $G$ and $\{A_v^{\mathrm{out}}\}_{v \in V}$ on the reversed graph. Next, the procedure for every $v\in V$ checks if there is a cycle containing $v$ of length at most $R$ in the induced graph of $A_v^{\mathrm{in}}\cup A_v^{\mathrm{out}}$. If such a cycle exists then the procedure returns such a cycle. Procedure ${\textsc{SimilarSet}}$ works as follows. The algorithms starts by sampling $O(\log{n})$ independent subsets $S_i$ of expected size $O(\sqrt{n})$ each for $1 \le i \le M$ where $M=50\log{n}$. From every vertex $w \in \bigcup_{1 \le i \le M}{S_i}$ the algorithm runs Dijkstra from and to $w$ in $G$. If a cycle of length $3R$ is detected then the algorithm returns it. Next for every vertex $v\in V$ and index $1 \le i \le M$ the algorithm defines a set $T_i(v) \subseteq S_i$. The sets $T_i(v)$ will be used to reduce the number of potential vertices that can be on a cycle of length at most $R$ with $v$. First, the set $T_0(v)$ consists of all vertices in $S_0$ that are at distance at most $R/2$ from $v$. Let $R_0(v)$ be a sampled set of $O(\log{n})$ vertices from $T_0(v)$. Now, the sets $T_{i}(v)$ and $R_{i}(v)$ are defined as follows. The set $T_{i}(v)$ is the set of all vertices $s \in S_i$ such that $d(v,s) \le R/2$ and $d(s, t) \le 3R/2 \text{ for all } t\in \bigcup_{0 \le j \le i-1}{R_{j}(v)}$. Again, define $R_i(v)$ as a sampled set of $O(\log n)$ vertices from $T_i(v)$. To gain intuition for the definition of $T_i(v)$ and $R_i(v)$, consider the set $G_i(v)$ of all vertices $s \in V(G)$ such that $d(v,s) \le R/2$ and $d(s, t) \le 3R/2 \text{ for all } t\in \bigcup_{0 \le j \le i-1}{R_{j}(v)}$. We remark that our algorithm does not compute $G_i(v)$, but its definition is essential for the analysis. Intuitively, the set $G_i(v)$ consists of the vertices after $i$ rounds that the algorithm still believes could be in a cycle of length $R$ with $v$. If $|G_i(v)| \ge 100\sqrt{n}\log n$, then by the choice of $S_i$ as an independent random set of expected size $O(\sqrt{n})$, we have that $T_i(v)$ is a random sample of $G_i(v)$ of expected size at least $100\log n.$ In this way, $R_i(v)$ is just a random sample of $G_i(v)$ of size $O(\log n).$ As we show in \[lem:SubsetShrinks\], if $|G_M(v)| \ge 100\sqrt{n}\log n$, our algorithm discovers w.h.p. a cycle of length at most $3R$ sometime during the shortest path computations done at the beginning. On the other hand, if $|G_M(v)| \le 100\sqrt{n}\log n$, then we can grow a shortest path tree from $v$ but only include vertices in $G_M(v)$ to search for a cycle of length $R$, only paying runtime $|G_M(v)| = {\tilde{O}}(\sqrt{n})$ for that vertex $v$. Formalizing this final step, the algorithm computes a shortest path tree $T(v)$ from $v$ up to depth $R/2$, keeping only vertices $s \in V$ such that $d(v,s) \le R/2$ and $d(s, t) \le 3R/2 \text{ for all } t\in \bigcup_{0 \le j \le M}{R_{j}(v)}$. The set $A_v$ is the set of vertices in $T(v)$. Invoke ${\textsc{SimilarSet}}(G,R)$ to either find a cycle of length at most $3R$ or set $A_v^{\mathrm{out}}\subseteq V$ for each $v \in V(G)$. \[line:similar\_search\_1\] Invoke ${\textsc{SimilarSet}}({G^{\mathrm{rev}}},R)$ where ${G^{\mathrm{rev}}}$ is the graph where the direction of every edge is reversed to either find a cycle of length at most $3R$ or sets $A_v^{\mathrm{in}}\subseteq V$ for each $v \in V(G)$. If a cycle of length at most $3R$ has yet to be found for each $v \in V(G)$ perform Dijsktra from $v$ in the graph induced by $A_v^{{\mathrm{out}}} \cup A_v^{{\mathrm{in}}}$ to find a cycle of length at most $R$ through one of the $v$. \[line:search\_union\] Return any cycle of length at most $3R$ found. \[line:final\_ball-search\] For $M = 50\log n$, sample sets $S_0, S_1, \cdots, S_M \subseteq V(G)$, each of expected size $O(n^{1/2})$ by sampling every vertex $v \in V$ independently with probability $p = n^{-1/2}$. Run Dijkstra to/from each vertex $v\in S_i$ for every $1 \leq i \leq M$. If there exists a vertex $v \in \cup_{1 \le i \le M}{S_i}$ such that $v$ is on a cycle of length $3R$ then return the shortest such cycle. \[line:returncycledetected\] Set $T_0(v) \gets \{s \in S_0 \mid d(v, s) \le R/2 \}$. Let $R_{i-1}(v)$ be $100 \log{n}$ vertices chosen independently at random from $T_{i-1}(v)$ Let $R_{i-1}(v) = T_{i-1}(v)$. $T_i(v) \gets \{ s \in S_{i} \mid d(v, s) \le R/2 \text{ and } d(s, t) \le 3R/2 \text{ for all } t\in \cup_{0 \le j \le i-1}{R_{j}(v)}\}$ Compute a shortest path tree $T(v)$ up to depth $R/2$ keeping only vertices $s$ such that $d(s, t) \le 3R/2 \text{ for all } t\in \cup_{0 \le j \le M}{R_{j}(v)}\}$. Set $A_v$ to be the set of vertices in $T(v)$. \[line:GrowSmallBall\] $A_v$ for all $v \in V$ Next we prove the correctness of our girth computation algorithm ${\textsc{GirthApprox}}$ (\[algo:girth3\]) and bound its running time. First we prove the following lemma which provides a fairly straightforward argument that the algorithm always outputs the correct result. The more challenging part of the analysis will be to bound its running time. \[lem:correctness\] If $G$ contains a cycle of length at most $R$ then ${\textsc{GirthApprox}}(G,R)$ (\[algo:girth3\]) returns a cycle of length at most $3R$. Assume $G$ contains a cycle $C$ of length at most $R$. Let $v$ be a vertex in $C$. If the algorithm returns a cycle in line \[line:returncycledetected\] of ${\textsc{SimilarSet}}$ (\[alg:similar\_vertices\]) then since this cycle has length at most $3R$, the algorithm works as desired. Consequently, we assume that this is not the case. Our goal is now to show that $A_v^{\mathrm{out}}$ contains all vertices $c \in C$ such that $d(v,c)\leq R/2$ and that $A_v^{\mathrm{in}}$ contains all vertices $c \in C$ such that $d(c,v)\leq R/2$. Since for all $c \in C$ either $d(v,c)\leq R/2$ or $d(c,v)\leq R/2$ this will imply that $C \subseteq A_v^{\mathrm{out}}\cup A_v^{\mathrm{in}}$ and therefore a cycle of length at most $R$ will be found in Line \[line:search\_union\] of ${\textsc{GirthApprox}}(G,R)$ (\[algo:girth3\]) and the algorithm works as desired. Further, note that it suffices to show that $A_v^{\mathrm{out}}$ contains all vertices $c \in C$ such that $d(v,c)\leq R/2$ as this will imply the desired claimed regarding $A_v^{\mathrm{in}}$ by symmetry. Consider the execution of ${\textsc{SimilarSet}}$ (\[alg:similar\_vertices\]) from Line \[line:similar\_search\_1\] of ${\textsc{GirthApprox}}(G,R)$ (\[algo:girth3\]). Further, consider a vertex $t \in T_i(v)$ for some $0 \le i \le M-1$. Recall that $d(v,t) \leq R/2$ (by definition and construction of $T_i(v)$). Consider a vertex $c$ in $C$. As $v$ and $c$ are on a cycle of length $R$ we have $d(c,v) \leq R$ and therefore $d(c,t) \leq d(c,v) + d(v,t) \leq 3R/2$ by triangle inequality. It follows by construction that each vertex $c \in C$ with $d(v,c)\leq R/2$ will be added to $A_v$ as desired. With the correctness of ${\textsc{GirthApprox}}$ (\[algo:girth3\]) established, in the remainder of this section we focus on analyzing its running time. To do this we will consider an invocation of ${\textsc{SimilarSet}}$ (\[alg:similar\_vertices\]) and both bound its running time and the size of the sets $A_v$ it computes. Before setting up the proofs, for each vertex $v \in V$ we define $$G_0(v) = \{s \in V \mid d(v, s) \le R/2 \}$$ and $$G_i(v) = \{s \in V \mid d(v, s) \le R/2 \text{ and } d(s, t) \le 3R/2 \text{ for all } t\in \cup_{0 \le j \le i-1}{R_{j}(v)}\} ~.$$ Notice that the distribution of $T_i(v)$ is the distribution on vertices that results from taking each $s \in G_i(v)$ and including it in $T_i(v)$ with probability $p = 1/\sqrt{n}$. Loosely speaking, the analysis of the running time is roughly as follows. The main non trivial part is to show that the expected size of the sets $A_v^{\mathrm{in}}$ and $A_v^{\mathrm{out}}$ is $\tilde{O}(\sqrt{n})$. This, together with the assumption that the maximum degree is $O(m/n)$, will imply that the running time of our algorithm is $\tilde{O}(m \sqrt{n})$. We roughly speaking show the following for the set $A_v^{\mathrm{out}}$ (similarly for the set $A_v^{\mathrm{in}}$ ). We want to claim that w.h.p. the sets $G_i(v)$ are decreasing by at least a constant factor until there is a set $G_i(v)$ of $\tilde{O}(\sqrt{n})$ size. As $A_v^{\mathrm{out}}$ is a subset of $G_M(v) \subseteq G_{i}(v)$, the claim follows. Assume this is not case, i.e., there exists an index $i$ such that $|G_{i+1}(v)| > 0.8 |G_{i}(v)|$. Note that by construction for every vertex $s$ in $G_{i+1}(v)$ all vertices in $R_{i}(v)$ are at distance at most $3R/2$ from it. As $R_{i}(v)$ is a sampled set of $G_{i}(v)$, we can show that w.h.p. most vertices in $G_i(v)$ (say 0.9 fraction of them) are at distance at most $3R/2$ from $s$. As $|G_{i+1}(v)| > 0.8 |G_{i}(v)|$, this means that this is also true for most vertices in $G_{i}(v)$. That is, most vertices in $G_{i}(v)$ are at distance at most $3R/2$ to most of the other vertices in $G_{i}(v)$. We show by counting argument that in this case there must be many pairs of vertices $u$ and $v$ such that $u,v \in G_i(v)$ and $d_G(u,v) \leq 3R/2$ and $d_G(v,u) \leq 3R/2$ (hence $u$ and $v$ are both on a cycle of length at most $3R$). That is, w.h.p. $G_i(v)$ contains many vertices that are on cycles of length at most $3R$. W.h.p. we can show that such a vertex will belong to $S_i(v)$ and therefore the algorithm will detect a cycle of length $3R$ and will not continue to computing the sets $A_v^{\mathrm{out}}$. \[lem:ManyEdges\] Consider a vertex $v$, index $i \in [M]$ such that $|G_{i}(v)| \geq 200 \sqrt{n} \log{n}$ and a vertex $u \in V$. If there are less than $0.9|G_i(v)|$ vertices $s \in G_i(v)$ such that $d(u,s) \leq 3R/2$ then with probability at least $1-2/n^{10}$, $u\notin G_{i+1}(v)$. Note that the distribution of obtaining $T_j(v)$ is equivalent to the distribution of picking every vertex in $G_j(v)$ with probability $p$ for every $1 \le j \le M$. We first show that with high probability $T_{i}(v)$ contains at least $100 \log{n}$ vertices (and therefore also $R_{i}(v)$). As $|G_{i}(v)| \geq 200 \sqrt{n} \log{n}$ then the expected size of $|T_{i}(v)|$ is at least $200 \log{n}$. Therefore, by Chernoff Bound the probability that $|T_{i}(v)| \le 100 \log{n}$ is at most $\left(\frac{e^{-1/2}}{1/2^{1/2}}^{100 \log{n}}\right) < 1/n^{10}$. Assume this is indeed the case, that is, $T_{i}(v)$ contains at least $100 \log{n}$ vertices. The set $R_{i}(v)$ is a sampled set of $100 \log{n}$ vertices from $T_{i}(v)$. As the distribution of obtaining the set $T_{i}(v)$ is equivalent to distribution of picking every vertex in $G_{i}(v)$ with probability $p$ then the distribution of $R_{i}(v)$ is equivalent to picking $100 \log{n}$ vertices from $G_{i}(v)$ (every vertex in $G_{i}(v)$ has the same probability appearing in $R_{i}(v)$). Consider a uniformly random vertex $s$ from $G_{i}(v)$. With probability at least $1/10$ we have $d(u,s) > 3R/2$. In other words with probability at most $9/10$ we have $d(u,s) \leq 3R/2$. Therefore, the probability that for every vertex $s$ in $R_j(v)$ we have $d(u,s) \leq 3R/2$ is at most $(9/10)^{100 \log{n}}\leq 1/n^{10}$. The lemma follows by union bound over the events that either $|T_{i}(v)|$ is smaller than $100 \log{n}$ or for all $s \in R_i(v)$ we have $d(u,s) \leq 3R/2$. \[lem:SubsetShrinks\] If there exists a vertex $v$ and an index $i$ such that $|G_{i}(v)| \geq 200 \sqrt{n} \log{n}$ and $|G_{i+1}(v)| \geq 0.8 |G_{i}(v)|$ then with probability at least $1-1/n^{8}$ there exists a vertex in $T_i(v)$ that is contained in a cycle of length at most $3R$. Assume such a vertex $v$ and index $i$ exist. We say that a vertex $u$ is $(v,i)$-dense if there are at least $0.9|G_i(v)|$ vertices $s \in G_i(v)$ such that $d(u,s) \leq 3R/2$. By union bound on all vertices $v\in V$ on \[lem:ManyEdges\], with probability at least $1-2/n^{9}$, all vertices in $G_{i+1}(v)$ are $(v,i)$-dense. As $G_{i+1}(v) \subseteq G_{i}(v)$ and $|G_{i+1}(v)| \geq 0.8|G_{i}(v)|$, we also have that with probability at least $1-2/n^{9}$, $0.8|G_{i}(v)|$ vertices in $G_{i}(v)$ are $(v,i)$-dense. Assume this is indeed the case. Imagine constructing the following directed graph $H$ whose set of vertices is $G_{i}(v)$ and set of edges is the following. For every vertex $u$ in $G_{i}(v)$ add an outgoing edge for every vertex $s$ such that $d(u,s) \leq 3R/2$. Note that if there exists two edges in the graph $(u,s)$ and $(s,u)$ then both $u$ and $s$ are on a cycle of length at most $3R$. We next show that by counting argument there are many vertices in $G_{i}(v)$ that are on a cycle of length at most $3R$. Every $(v,i)$-dense vertex $u$ has $0.9|G_{i}(v)|$ outgoing edges in $H$. There are at least $0.9|G_{i}(v)|$ $(v,i)$-dense vertices in $H$. We get that the number of edges $E(H)$ is at least $0.71|G_{i}(v)|^2$, that is, $|E(H)| \geq 0.71|G_{i}(v)|^2$. On the other hand let $\alpha$ be the fraction of vertices in $G_{i}(v)$ that do appear on a cycle of length at most $3R$. For every edge in $H$ give a credit of 1/2 for each of its endpoints vertices. Note that every vertex $x$ that do not belong to a cycle of length at most $3R$ can get a credit of less than $|G_{i}(v)|/2$. To see this, note that there is no other vertex with both incoming and outgoing edge to $x$ (as otherwise $x$ is on a cycle of length at most $3R$) so the total number of incoming and outgoing edges of $x$ is at most $|G_{i}(v)|-1 < |G_{i}(v)|$. Hence, the total credit of $x$ is less than $|G_{i}(v)|/2$. The total credit of a vertex $x$ that do participate in a cycle of length at most $3R$ is less than $|G_{i}(v)|$. We get that the total credit of all vertices, which is also equal to the total number of edges in $H$, is less than $\alpha |G_{i}(v)| |G_{i}(v)|/2 + (1-\alpha) |G_{i}(v)|^2$. It follows that $0.71|G_{i}(v)|^2 \leq \alpha |G_{i}(v)| |G_{i}(v)|/2 + (1-\alpha) |G_{i}(v)|^2$. Straight forward calculation show that $\alpha < 0.58$ and thus $1-\alpha > 0.42$. In other words, at least $0.42|G_{i}(v)|$ vertices in $G_{i}(v)$ belong to a cycle of length at most $3R$. Next, we claim that w.h.p. there is such a vertex in $T_i(v)$. Recall that the distribution of $T_i(v)$ is equivalent to picking every vertex in $G_i(v)$ with probability $p$. Consider one vertex that participates in a cycle of length at most $3R$ the probability it does not belong to $T_i(v)$ is $1-p$. The probability that none of the $0.42|G_{i}(v)|$ vertices belong to $T_i(v)$ is at most $(1-p)^{0.42|G_{i}(v)|} \leq (1-p)^{84 \log{n}/p} \leq 1/n^{10}$. The lemma follows (as $ 1/n^{10} + 2/n^{9} < 1/n^{8}$ for large enough $n$). Finally, the following concludes the running time of our algorithm. \[lem:totaltime\] The expected running time of \[algo:girth3\] is $O(m\sqrt{n}\log{n}+n\sqrt{n}\log^3{n}) = \tilde{O}(m \sqrt{n})$. Consider one of the executions of ${\textsc{SimilarSet}}$ (\[alg:similar\_vertices\]) by ${\textsc{GirthApprox}}$ (\[algo:girth3\]). This algorithm computes Dijkstra to/from each vertex $w\in S_i$ for every $1 \leq i \leq M$ in $O(m+n\log{n})$ time. The expected size of each $S_i$ is $O(n^{1/2})$. Thus, the expected time of this computation for $S_i$ is $O(m\sqrt{n}+n\sqrt{n}\log{n})$. There are $O(\log{n})$ sets $S_i$ and therefore there is at most $O(m\sqrt{n}\log{n}+n\sqrt{n}\log^2{n})$ expected time for the computation of all Dijkstra’s. Next, for every vertex $v$ the algorithm computes the sets $T_i(v)$ for every $i\in [M]$. The set $T_0(v)$ can be computed easily in $O(|S_0|)$ time which is $O(n^{1/2})$ in expectation. In order to compute $T_i(v)$ for $i >0$, the algorithm considers every vertex $s \in S_i$ and it check if $s$ is at distance at most $3R/2$ from every vertex in $t\in \cup_{j\in[0,...,i-1]}{R_{j}(v)}$. There are $O(\log^2{n})$ vertices $t$ in $\cup_{j\in[0,...,i-1]}{R_{j}(v)}$. The distance $d(s,v)$ is already computed and thus can be retrieved in $O(1)$ time. Overall, computing the set $T_i(v)$ takes $O(n^{1/2} \log^2{n})$ in expectation. Therefore, $O(n^{1/2} \log^3{n})$ for all indices $i \in [M]$. Hence, for all vertices $v$ $O(n^{3/2} \log^3{n})$ expected time for this part. Next, we bound the cost of computing the balls, $A_v$, and we bound their size. By a slight abuse of notation we call a vertex $s$ $(v,M)$-dense if it satisfies $$d(v, s) \le R/2 \text{ and } d(s, t) \le 3R/2 \text{ for all } t\in \cup_{j\in[0,...,M]}{R_{j}(v)}.$$ The algorithm grows a ball from every vertex $v\in V$ by only keeping vertices $s$ that are $(v,M)$-dense to compute $A_v$. We first show that if there is no index $i$ such that $|G_{i}(v)| \geq 200 \sqrt{n} \log{n}$ and $|G_{i+1}(v)| \geq 0.8 |G_{k}(v)|$ then the expected time to compute the ball of $v$ is $O(n^{1/2}\log^3{n} + n^{1/2} \log{n} \cdot \delta)$. We do that by showing that the expected number of vertices in $G_M(v)$ is $O(n^{1/2} \log{n})$. As the maximum degree in $G$ is $O(\delta)$ and checking if a vertex $s$ is $(v,M)$-dense takes $O(\log^2{n})$ time, then the claim follows. As for every $i$ such that $|G_{i}(v)| \geq O(\sqrt{n}\log{n})$ we have $|G_{i+1}(v)| \leq 0.8 |G_{i}(v)|$ then straight forward calculation shows that there exists an index $M' \in[1..M]$ such that $|G_{M'}(v)| < O(\sqrt{n}\log{n})$. Note that the ball of $v$ contains only vertices from $G_{M'}(v)$ and thus the claim follows. We now assume that there exists a vertex $v$ and index $i$ such that $|G_{i}(v)| \geq O(\sqrt{n} \log{n})$ and $|G_{i+1}(v)| \geq 0.8 |G_{i}(v)|$. By claim \[lem:SubsetShrinks\] in this case with probability at least $1-1/n^8$ the algorithm finds a cycle of length $3R$ and returns it in Line \[line:returncycledetected\]. Therefore, in this case the algorithm does not compute the balls in Line \[line:GrowSmallBall\]. With probability at most $1/n^8$ the algorithm does not find a cycle in Line \[line:returncycledetected\] and therefore continues to computing the balls in Line \[line:GrowSmallBall\]. The computation of all balls in Line \[line:GrowSmallBall\] is bounded by $O(mn)$ in this case. As this happens with very small probability this does not effect the asymptotic bound of the expected running time. The lemma follows. We conclude this section with the proof of \[thm:3girth\]. The algorithm calls Algorithm ${\textsc{GirthApprox}}$ using a binary search on the range $[1, nW]$ to find a parameter $R$ such that Algorithm ${\textsc{GirthApprox}}$ returns a cycle (of length at most $3(R+1)$) when invoked on $R+1$ but not on $R$. As mentioned above the dependency on $\log{nW}$ can be improved to $\log{n}$ using the method used in [@PRSTV18] (Section 5.1). Roughly speaking this method constructs in $O(m \log{n})$ time a set of graphs such that the number of vertices in all these graphs together is $O(n \log{n})$, the number of edges is $O(m \log{n})$, the ratio between the maximum edge weight and the minimum edge weight in all these graphs is $O(n)$ and the shortest cycle is contained in one of these graphs. Instead of running binary search on $G$, we run it in each of these graphs. Now using \[lem:correctness\] and \[lem:totaltime\] the theorem follows. We give our result on constant approximation roundtrip spanners in ${\tilde{O}}(m\sqrt{n})$ time and show \[thm:const\_spanner\] in \[sec:constspanner\]. Deterministic $O(k \log \log n)$ Approximation Algorithms {#sec:algo} ========================================================= In this section we present our deterministic algorithms for computing a $O(k \log \log n)$ approximation to the girth and computing $O(k \log \log n)$ multiplicative roundtrip spanners. Our main result will be showing how to compute improved roundtrip covers as defined originally in [@RTZ08]. Leveraging this result we will prove \[thm:girth\] and \[thm:spanner\]. First, leveraging the definitions of balls in \[sec:prelim\] we define roundtrip covers. Intuitively, roundtrip covers are a union of balls of radius $kR$ such that if vertices $u, v \in V(G)$ satisfy $d(u {\leftrightarrows}v) \le R$ then $u, v$ are both in some ball in the cover. A collection $C$ of balls is a *$(k, R)$ roundtrip cover* of a weighted directed graph $G$ if and only if every ball in $C$ has radius at most $kR$, and for any $u, v \in V(G)$ with $d(u {\leftrightarrows}v) \le R$ there is a ball $B \in C$ such that $u, v \in B$. Specifically, we show the following theorem. \[thm:cover\] For an $n$-vertex $m$-edge graph $G$, an execution of ${\textsc{RoundtripCover}}(G, k, R)$ returns a collection $C$ of balls that forms a $(O(k\log\log n), R)$ roundtrip cover of a weighted directed graph $G$ in time $m^{1+O(1/k)}$ where $\sum_{B \in C} |V(B)| = n^{1+O(1/k)}.$ To show \[thm:girth\] from \[thm:cover\], we can compute $(k,2^i)$ roundtrip covers for all $0 \le i \le O(\log n)$, and set our girth estimate as the minimum radius of any ball in the cover that has a cycle. To compute a roundtrip spanner, simply take the union of all the balls in the $(k,2^i)$ roundtrip covers for all $i = O(\log n).$ The rest of the section is organized as follows. In \[sec:mainalgo\] we state our main algorithm. In \[sec:analysis\] we analyze the algorithm and prove \[thm:cover\]. In \[sec:mainappl\] we use \[thm:cover\] to formally prove \[thm:girth\] and \[thm:spanner\]. Technical Overview {#sec:4overview} ------------------ We focus on unweighted directed graphs $G$ and for a parameter $R$, construct a roundtrip spanner $H$ so that if the roundtrip distance between $u$ and $v$ is at most $R$ in $G$, then their roundtrip distance is at most $O(Rk\log\log n)$ in $H$. Our approach is based on growing inballs and outballs in the graph $G$. Fix a vertex $v$, and let ${B^{\mathrm{in}}}_i, {B^{\mathrm{out}}}_i$ denote the inball and outball of radius $iR$ around $v$, and let $|{B^{\mathrm{in}}}_i|, |{B^{\mathrm{out}}}_i|$ denote the number of vertices in the balls and fix $d = O(k \log\log n).$ We start by growing and inball and outball around $v$. First, if $|{B^{\mathrm{in}}}_d \cap {B^{\mathrm{out}}}_d| \ge \frac{n}{2}$, then we can build a roundtrip ball of radius $2dR+R$ and delete ${B^{\mathrm{in}}}_d \cap {B^{\mathrm{out}}}_d$ from our graph. This is safe essentially by our observation above. Otherwise, we find an index $i$ such that $|{B^{\mathrm{in}}}_{i+1}|$ isn’t much larger than $|{B^{\mathrm{in}}}_i|$, we recursively build a roundtrip cover on ${B^{\mathrm{in}}}_{i+1}$ and then delete ${B^{\mathrm{in}}}_i$. This is safe to do by our observation above. Similarly, if there is an index $i$ such that $|{B^{\mathrm{out}}}_{i+1}|$ isn’t much larger than $|{B^{\mathrm{out}}}_i|$, we recursively build a roundtrip cover on ${B^{\mathrm{out}}}_{i+1}$ and then delete ${B^{\mathrm{out}}}_i$. Through standard ball cutting inequalities we can show that such an index $i$ exists (\[lemma:inbound\]). We would like to elaborate on a few points. First, when we compare the sizes of $|{B^{\mathrm{in}}}_{i+1}|$ and $|{B^{\mathrm{in}}}_i|$, we compare both the number of vertices *and* edges, the former to control the size of the roundtrip spanner constructed, and the latter to control runtime. Second, we grow the inball and outball at the same rate, i.e. we alternately add an edge at a time to the inball and outball to maintain that the work spent on each is the same. Main Algorithm {#sec:mainalgo} -------------- We first give a high-level description of our algorithm for computing Roundtrip Covers, [<span style="font-variant:small-caps;">RoundtripCover</span>]{}, which is presented formally as \[algo:roundtripcover\]. #### High-level Description of Algorithm. As discussed in \[sec:approach\_overview,sec:4overview\], our algorithm is based on ball growing along with the following observation: if for a radius $r'$ we compute a roundtrip cover of ${B^{\mathrm{in}}}_v(r'+R)$ and add all the balls in the computed roundtrip cover on ${B^{\mathrm{in}}}_v(r'+R)$ to our final cover, then we can safely delete all vertices $u \in {B^{\mathrm{in}}}_v(r')$ from our graph and recurse on the rest of graph; the deleted vertices are already satisfied in the sense that for every $u' \in V(G)$ with $d(u {\leftrightarrows}u') \le R$ there is a ball $B$ in the cover such that $u, u' \in B$. Indeed, if $u \in {B^{\mathrm{in}}}_v(r')$ and $d(u {\leftrightarrows}u') \le R$ then $u, u' \in {B^{\mathrm{in}}}_v(r'+R)$ and therefore we are guaranteed that the roundtrip cover on ${B^{\mathrm{in}}}_v(r'+R)$ contains a ball $B$ such that $u,u' \in B$. Using this observation, we grow inballs and outballs around vertices in our graph $G$ to “partition" our graph into pieces that possibly overlap, where the overlap corresponds to the boundary ${B^{\mathrm{in}}}_v(r'+R) \backslash {B^{\mathrm{in}}}_v(r')$ in our example. We describe our algorithm in more detail now. Consider any vertex $v$. We grow an inball and outball around $v$ at the same rate, spending the same time on the inball and outball. First, we consider the case that $|V({B^{\mathrm{in}}}_v(r))|, |V({B^{\mathrm{out}}}_v(r))| \ge \frac{3n}{4}$ for some $r = O(Rk \log\log n)$, as was done in Pachocki [*et al.*]{} [@PRSTV18]. Then we know that $|V({B^{\mathrm{in}}}_v(r)) \cap V({B^{\mathrm{out}}}_v(r))| \ge \frac{n}{2}.$ By our observation above, we can add the ball $B_v(2r+R)$ to our roundtrip cover, delete ${B^{\mathrm{in}}}_v(r) \cap {B^{\mathrm{out}}}_v(r)$ from $G$, and recurse on the remainder. Otherwise, if we find a radius $r'$ such that say ${B^{\mathrm{in}}}_v(r')$ and ${B^{\mathrm{in}}}_v(r'+R)$ satisfy the conditions of [<span style="font-variant:small-caps;">GoodCut</span>]{} (\[algo:goodcut\]), then we recurse on ${B^{\mathrm{in}}}_v(r'+R)$ and delete ${B^{\mathrm{in}}}_v(r')$ from our graph and recurse on the remaining graph. This is safe to do by our observation above. We can also do an analogous process on ${B^{\mathrm{out}}}_v(r')$ and ${B^{\mathrm{out}}}_v(r'+R).$ By a variant of the standard ball-growing inequality (\[lemma:inbound\]) we can show that a good cut always exists. We now will give some intuition about the condition in [<span style="font-variant:small-caps;">GoodCut</span>]{} and the (somewhat strange) appearance of the $O(\log\log n)$ in our algorithm. First, we remark that the condition in [<span style="font-variant:small-caps;">GoodCut</span>]{} must track both the number of vertices and edges in the ball: the former to control recursion depth and roundtrip cover size, and the latter to control runtime. Now we give intuition for why we require an $O(k \log\log n)$ approximation factor in our algorithm. Consider growing inballs ${B^{\mathrm{in}}}_v(r)$ from $v$ for various radii $r$, and recall that we make a cut depending on the relative sizes of $|V({B^{\mathrm{in}}}_v(r))|$ and $|V({B^{\mathrm{in}}}_v(r + R))|$. Now, note that if for example $|V({B^{\mathrm{in}}}_v(r))| = O(1)$, we can afford to have $|V({B^{\mathrm{in}}}_v(r + R))| = O(n^{1/k})$, as we can simply run a naive algorithm on ${B^{\mathrm{in}}}_v(r + R)$ now. On the other hand, if for example $|V({B^{\mathrm{in}}}_v (r))| = \Omega(n)$, we can essentially only afford to have $|V({B^{\mathrm{in}}}_v(r + R))| \le \left(1 + \frac{1}{k}\right)|V({B^{\mathrm{in}}}_v(r))|.$ To see the latter, note that the recurrence $T(m) = \left(1 + \frac1k\right)(T(m/2) + T(m/2))$ has solution $T(m) = m^{1+O(1/k)}.$ Now, interpolating between these two extremes allows us to compute the optimal way to do ball cutting (which is done in [<span style="font-variant:small-caps;">GoodCut</span>]{}). This leads to a ball cutting procedure with $O(k\log\log n)$ levels, and thus results in an $O(k \log \log n)$ approximation ratio. ${i_{\mathrm{in}}}, {i_{\mathrm{out}}}{\leftarrow}0$. $r {\leftarrow}5kR\log\log n$. Take any $v \in V(G).$ *\\\\some condition below in lines \[line:bothbig\], \[line:goodcutin\], \[line:goodcutout\] will trigger eventually* \[line:bothbig\] \[line:cuthalf\] $\{B_v( 2r+R) \} \cup$ [<span style="font-variant:small-caps;">RoundtripCover</span>]{}$(G\backslash ({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1) R) \cap {B^{\mathrm{out}}}_v( ({i_{\mathrm{out}}}+1) R)), R, k)$. \[line:goodcutin\] \[line:cutin\] [<span style="font-variant:small-caps;">RoundtripCover</span>]{}$({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1) R), R, k) \cup $[<span style="font-variant:small-caps;">RoundtripCover</span>]{}$(G\backslash {B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R), R, k)$. \[line:goodcutout\] \[line:cutout\] [<span style="font-variant:small-caps;">RoundtripCover</span>]{}$({B^{\mathrm{out}}}_v( ({i_{\mathrm{out}}}+1) R), R, k) \cup $[<span style="font-variant:small-caps;">RoundtripCover</span>]{}$(G\backslash {B^{\mathrm{out}}}_v( {i_{\mathrm{out}}}R), R, k)$. \[line:growin\] ${i_{\mathrm{in}}}{\leftarrow}{i_{\mathrm{in}}}+ 1$ ${i_{\mathrm{out}}}{\leftarrow}{i_{\mathrm{out}}}+ 1$ \[algo:roundtripcover\] **true** **false** \[algo:goodcut\] #### Explanation of \[algo:roundtripcover\]: We now explain what each piece of \[algo:roundtripcover\] is doing. Here, ${i_{\mathrm{in}}}$ and ${i_{\mathrm{out}}}$ track the radius of the inball and outball that we are growing. We grow the balls at the same rate. If we notice that at any point we are in position to make a good cut (see lines \[line:goodcutin\], \[line:goodcutout\]) then we do so. Otherwise, we know that both balls will eventually contain many vertices (see line \[line:bothbig\]). In this case, we add $B_v(2r+R)$ to our roundtrip cover, delete ${B^{\mathrm{in}}}_v({i_{\mathrm{in}}}R) \cap {B^{\mathrm{out}}}_v({i_{\mathrm{out}}}R)$ from our graph, and recurse. To grow the inball and outball at the same rate, we run Dijkstra to grow the inball and outball, alternately processing an edge at a time from the inball and outball. We check the condition of [<span style="font-variant:small-caps;">GoodCut</span>]{} on a ball when we have certified that we have processed all vertices up to distance ${i_{\mathrm{in}}}R$ or ${i_{\mathrm{out}}}R$ respectively. Analysis of [<span style="font-variant:small-caps;">RoundtripCover</span>]{} and proof of \[thm:cover\] {#sec:analysis} ------------------------------------------------------------------------------------------------------- In this section we prove \[thm:cover\], bounding the performance of our roundtrip cover algorithm \[algo:roundtripcover\]. We start by showing that $1+\max({i_{\mathrm{in}}}, {i_{\mathrm{out}}}) \le 5k\log\log n$ at all points in the algorithm, hence some condition in lines \[line:bothbig\], \[line:goodcutin\], \[line:goodcutout\] will trigger eventually. \[lemma:inbound\] At all points during \[algo:roundtripcover\], we have that $1+\max({i_{\mathrm{in}}}, {i_{\mathrm{out}}}) \le 5k\log\log n.$ We show $1 + {i_{\mathrm{in}}}\le 5k\log\log n$, and the bound on $1 + i_{{i_{\mathrm{out}}}}$ is analogous. To prove this we assume that none of the conditions in the inner loop of the algorithm trigger, and compute the resulting vertex and edge sizes of ${B^{\mathrm{in}}}_v({i_{\mathrm{in}}}R)$ and ${B^{\mathrm{out}}}_v({i_{\mathrm{out}}}R)$. To this end, assume that $|V({B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R))| \le \frac{3n}{4}$ and $|E({B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R))| \le m.$ By the conditions of lines \[line:bothbig\], \[line:goodcutin\], and \[line:growin\] we know that each time we increment ${i_{\mathrm{in}}}$ either $$\label{eq:vert} |V({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1)R))| \ge |V({B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R))|^\frac{k-1}{k} n^\frac1k$$ or $$\begin{aligned} \label{eq:edge1}|E({B^{\mathrm{in}}}_v(({i_{\mathrm{in}}}+1)R))| &\ge |E({B^{\mathrm{in}}}_v({i_{\mathrm{in}}}R))|^\frac{k-1}{k} m^\frac1k \text{ and }\\ \label{eq:edge2} |E({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1)R))| &\ge \left(1+\frac{1}{k}\right)|E({B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R))|.\end{aligned}$$ We first show that \[eq:vert\] can only hold for $2k \log 4\log n$ values of ${i_{\mathrm{in}}}$. To this end, define a sequence $\{x_i\}_{i \ge 0}$ as $x_0 = 1$ and $x_{i+1} = x_i^\frac{k-1}{k} n^\frac1k.$ By induction it follows that $x_i = n^{1 - \left(\frac{k-1}{k}\right)^i}.$ In particular, $$x_{2k\log4\log n} = n^{1 - \left(\frac{k-1}{k}\right)^{2k\log 4\log n}} \ge \frac{3}{4}n.$$ This shows that the condition in \[eq:vert\] can only hold at most $2k \log 4\log n$ times. Similarly, after \[eq:edge1\] holds for $2k \log 4\log n$ different ${i_{\mathrm{in}}}$, we will have that $|E({B^{\mathrm{in}}}_v({i_{\mathrm{in}}}R))| \ge \frac{3m}{4}.$ At this point, \[eq:edge2\] can hold at most $k$ times. This gives us that in total $$1+{i_{\mathrm{in}}}\le 1 + 2k \log 4\log n + 2k \log 4\log n + k \le 5k \log\log n$$ as desired. Now we proceed to proving \[thm:cover\]. We first show that the algorithm indeed returns a $(O(k\log\log n), R)$ roundtrip cover. Then we bound the total size of balls in the roundtrip cover, as well as the runtime. #### Returns a $(O(k\log\log n), R)$ roundtrip cover. We analyze lines \[line:cuthalf\], \[line:cutin\], and \[line:cutout\]. In line \[line:cuthalf\], note that by \[lemma:inbound\], we know that $({i_{\mathrm{in}}}+1)R, ({i_{\mathrm{out}}}+1)R \le r.$ Therefore, we know that ${B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R) \cap {B^{\mathrm{out}}}_v( {i_{\mathrm{out}}}R) \subseteq B_v( 2r).$ Additionally, it is clear that for any vertex $u \in {B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R) \cap {B^{\mathrm{out}}}_v( {i_{\mathrm{out}}}R)$, if another vertex $u'$ satisfies $d(u {\leftrightarrows}u') \le R$ then $u' \in B_v( 2r+R).$ Therefore, the ball $B_v( 2r+R)$ contains both $u$ and $u'$, so we can safely delete ${B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R) \cap {B^{\mathrm{out}}}_v( {i_{\mathrm{out}}}R)$ from $G$ and recurse. This is exactly what is happening in line \[line:cuthalf\]. In line \[line:cutin\], note that for any vertex $u \in {B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R)$, if another vertex $u'$ satisfies $d(u {\leftrightarrows}u') \le R$ then $u' \in {B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1)R).$ Therefore, if we construct a roundtrip cover on ${B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1)R)$, then we can safely delete ${B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R)$ from $G$ and recurse. This is exactly what occurs in line \[line:cutin\]. The same argument now applies to line \[line:cutout\]. Finally, note that all balls we create are of radius $2r+R = O(Rk \log\log n).$ #### Total sizes of balls is $n^{1+O(1/k)}$. We show by induction that the total number of vertices among all balls in the rountrip cover computed is at most $10n^\frac{k}{k-1}$ for an input graph $G$ with $n$ vertices. We show this by analyzing lines \[line:cuthalf\], \[line:cutin\], and \[line:cutout\]. For line \[line:cuthalf\], note that because $\min(|V({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1) R))|, |V({B^{\mathrm{out}}}_v( ({i_{\mathrm{out}}}+1) R))|) \ge \frac{3n}{4}$, we know that $|V({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1) R)) \cap V({B^{\mathrm{out}}}_v( ({i_{\mathrm{out}}}+1) R))| \ge \frac{n}{2}$. Therefore, it suffices to verify $$2n + 10\left(\frac{n}{2}\right)^\frac{k}{k-1} \le 10n^\frac{k}{k-1}$$ which is clear. For line \[line:cutin\], for simplicity let $s = |V({B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R))|.$ Then by the condition of [<span style="font-variant:small-caps;">GoodCut</span>]{}, it suffices to note that $$\begin{aligned} &10|V({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1)R))|^\frac{k}{k-1} + 10(n-s)^\frac{k}{k-1} \le 10(s^\frac{k-1}{k} n^\frac{1}{k})^\frac{k}{k-1} + 10(n-s)^\frac{k}{k-1} \\ &\le 10s n^\frac{1}{k-1} + 10(n-s) n^\frac{1}{k-1} = 10n^\frac{k}{k-1}. \end{aligned}$$ The same argument now applies to line \[line:cutout\]. #### Can be implemented to run in time $m^{1+O(1/k)}$. We can implement the algorithm to grow ${B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R)$ and ${B^{\mathrm{out}}}_v( {i_{\mathrm{out}}}R)$ at the same rate, i.e., we process a single inedge and outedge at a time, and increment ${i_{\mathrm{in}}}$ and ${i_{\mathrm{out}}}$ when we are sure that we’ve processed the whole inball or outball. This can be done with Dijkstra’s algorithm. We stop growing a ball once it contains at least $\frac{3n}{4}$ vertices. This way, any time we recurse, the total amount of work we have done to this point is at most twice the number of edges in the piece we are recursing on in lines \[line:cuthalf\], \[line:cutin\], and \[line:cutout\]. To bound the runtime, we imagine lines \[line:cutin\] and \[line:cutout\] as partitioning the graph into pieces of the form ${B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R)$ or ${B^{\mathrm{out}}}_v( {i_{\mathrm{out}}}R)$ and then recursing on ${B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1)R)$ or ${B^{\mathrm{out}}}_v( ({i_{\mathrm{out}}}+1) R)$. This way, the depth of the recursion is at most $O(\log n)$ because we know that $|V({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1)R))|, |V({B^{\mathrm{out}}}_v( ({i_{\mathrm{out}}}+1) R))| \le \frac{3n}{4}$ when we recurse. We will now show that the total number of edges in level $\ell$ of the recursion is bounded by $\left(1+\frac{2}{k}\right)^\ell m^\frac{k}{k-1}$, where the top level is level $0$. We proceed by induction on $\ell.$ Say that the algorithm partitions $G$ into $G = G_1 \cup G_2 \cup \dots \cup G_j,$ where each $G_i$ is either of the form ${B^{\mathrm{in}}}_v( {i_{\mathrm{in}}}R)$ or ${B^{\mathrm{out}}}_v( {i_{\mathrm{out}}}R).$ For simplicity, let $s_i = |E(G_i)|$ and let $t_i = |E({B^{\mathrm{in}}}_v( ({i_{\mathrm{in}}}+1) R))|$ or $t_i = |E({B^{\mathrm{out}}}_v( ({i_{\mathrm{out}}}+1) R))|$ corresponding to what $G_i$ was. We know by the condition of [<span style="font-variant:small-caps;">GoodCut</span>]{} that $t_i \le \max(\left(1+\frac{1}{k}\right)s_i, s_i^\frac{k-1}{k} m^\frac{1}{k}).$ By induction, we know that the total number of edges processed in level $\ell$ is at most $$\begin{aligned} &\sum_i \left(1+\frac{2}{k}\right)^{\ell-1} t_i^\frac{k}{k-1} \le \left(1+\frac{2}{k}\right)^{\ell-1} \sum_i \max\left(\left(1+\frac{1}{k}\right)s_i, s_i^\frac{k-1}{k} m^\frac{1}{k}\right)^\frac{k}{k-1} \\ &\le \left(1+\frac{2}{k}\right)^{\ell-1} \sum_i \left(1+\frac{2}{k}\right)s_i m^\frac{1}{k-1} \le \left(1+\frac{2}{k}\right)^\ell m^\frac{k}{k-1} \end{aligned}$$ as $\sum_i s_i \le m$ obviously. Now, it is clear that the total work done on a graph $G$ at some node of the recursion tree is $\tilde{O}(|E(G)|)$ as line \[line:cuthalf\] only occurs $O(\log n)$ times. Now taking $\ell = O(\log n)$ in the above claim completes the proof. Proofs of \[thm:girth\] and \[thm:spanner\] {#sec:mainappl} ------------------------------------------- Both theorems follow easily from \[thm:cover\]. We first show the result for unweighted graphs. To show this, run $$\text{{\textsc{RoundtripCover}}}(G, O(k), 2^i) \text{ for } 0 \le i \le O(\log n).$$ Now, set our estimate $g'$ of the girth to be the smallest radius of any nontrivial ball that we had in a roundtrip cover. By the guarantees of [<span style="font-variant:small-caps;">RoundtripCover</span>]{}, it is clear that $g \le g' \le O(k\log\log n) \cdot g$ as desired. It is clear that the algorithm runs in time $\tilde{O}(m^{1+\frac1k})$ by \[thm:cover\]. We can extend this to weighted graphs by instead taking $0 \le i \le O(\log nW)$, where $W$ is the maximum edge weight. This can be improved to $O(\log n)$ by the same method as done in [@PRSTV18], where they give a general reduction by contracting small weight strongly connected components and deleting large weight edges (see Section 5.1 in [@PRSTV18] for more details). We first show the result for unweighted graphs. It is easy to see that $$\bigcup_{i = 0}^{O(\log n)} \text{{\textsc{RoundtripCover}}}(G, O(k), 2^i)$$ is an $O(k \log\log n)$ spanner with $\tilde{O}(n^{1+ 1/k})$ edges by \[thm:cover\]. It is clear that the algorithm runs in time $\tilde{O}(m^{1+\frac1k}).$ The extension to weighted graphs follows as in the above paragraph (proof of \[thm:girth\]). An $O(k \log k)$ Approximation in ${\tilde{O}}(m^{1+1/k})$ Time {#sec:klogk_approx} =============================================================== In this section we explain how to combine the ideas from \[algo:roundtripcover\] and \[algo:girth3\] to give an algorithm for $(O(k \log k), R)$-roundtrip covers with ${\tilde{O}}(n^{1+ 1/k})$ edges in time ${\tilde{O}}(m^{1 + 1/k})$. Then \[thm:const\_girth\] and \[thm:arb\_const\_spanner\] follow from this in the same way that \[thm:girth\] and \[thm:spanner\] followed from \[thm:cover\]. \[thm:roundtripcover2\] For an $n$-vertex $m$-edge graph $G$, an execution of ${\textsc{RoundtripCover2}}(G, k, R)$ returns a collection $C$ of balls that form a $(O(k\log k), R)$ roundtrip cover of a weighted directed graph $G$ in time $m^{1+O(1/k)}$ where $\sum_{B \in C} |V(B)| = n^{1+O(1/k)}.$ The remainder of the section is organized as follows. We first give an overview for our approach, which combines the complementary approaches of sections \[sec:algo\] and \[sec:constantapprox\]. We then state our main algorithm, \[algo:roundtripcover2\]. Afterwards, we analyze \[algo:roundtripcover2\] to prove \[thm:roundtripcover2\] in \[sec:analysis2\]. Finally, we apply \[thm:roundtripcover2\] to prove \[thm:const\_girth\] and \[thm:arb\_const\_spanner\]. #### Overview of approach. Throughout this section, we assume that we have applied \[lemma:regular\] to make our graph $G$ approximately regular. Here we give a high level overview for the ideas behind the algorithm. Let $G$ be an $n$-vertex $m$-edge graph and let $K {:=}10k \log k$ for integer $k$. We start by generalizing \[algo:girth3\] and \[alg:similar\_vertices\] slightly, where we consider the case where the sampled sets $S_i$ have size ${\tilde{O}}(n^{1/k})$ instead of ${\tilde{O}}(n^{1/2}).$ To elaborate, we first view \[algo:girth3\] and \[alg:similar\_vertices\] as algorithms with the following guarantees. They add ${\tilde{O}}(n^{3/2})$ edges towards a spanner, and then for each vertex $v$ which is not yet in a cycle of length $4R$ using the current spanner edges builds a data structure $D_v$ (corresponding to \[alg:similar\_vertices\]) which certifies that for all but at most $O(n^{1/2})$ other vertices $u$ we have that $d(v {\leftrightarrows}u) > 4R$. We can generalize this as follows. There is a corresponding algorithm (\[algo:buildsimilar\]) which has the following guarantees. It adds ${\tilde{O}}(n^{1+1/k})$ edges towards a spanner, and then for each vertex $v$ which is not yet in a cycle of length $2KR$ using the current spanner edges builds a data structure $D_v$ which certifies that for all but at most $O(n^\frac{k-1}{k})$ other vertices $u$ we have that $d(v {\leftrightarrows}u) > KR$. After running this generalized algorithm (\[algo:buildsimilar\]), for a vertex $v$, we can define *$i$-similar vertices* to $v$, which are intuitively the vertices that the data structure $D_v$ thinks could still possibly be in a cycle of length $kR$ with $v$ and which are within distance $iR$ of $v$. Then we define a sequence $E_v^0, E_v^1, \cdots, E_v^K$ of “balls" centered at $v$, where $E_v^i$ is the outball from $v$ consisting of $i$-similar vertices. The following important conditions hold: $v \in E_v^0,$ and $E_v^i \subseteq E_v^{i+1}$ for all $0 \le i < K.$ Finally, if $u \in E_v^i$ and $d(u {\leftrightarrows}u') \le R$, then $u' \in E_v^{i+1}.$ This allows us to apply the ball-growing procedure in \[algo:roundtripcover\] but using the balls $E_v^i$. Note that by our choice of $K$ and a variant of \[lemma:inbound\], there exists a good cut. This is because $$n^{1 - \left(\frac{k-1}{k}\right)^K} \ge n^{1 - \frac{1}{k}} = n^\frac{k-1}{k}.$$ Hence, we can make this good cut and then recurse. Here, our cutting condition is simpler (only checks vertices, not edges) because we have reduced to the case of regular graphs through \[lemma:regular\]. $C {\leftarrow}\emptyset$. $(G', C', D) {\leftarrow}{\textsc{BuildSimilar}}(G, k, R)$ $C {\leftarrow}C'.$ $C {\leftarrow}C \cup {\textsc{BallGrow}}(G', k, R, D).$ \[line:callballgrow\] $C$. \[algo:roundtripcover2\] $C {\leftarrow}\emptyset.$ $K {\leftarrow}10k \log k.$ Select uniformly random subsets $S^1, S^2, \cdots, S^{100\log {\hat{n}}} \subseteq V(G)$ where $|S^i| = 100n^\frac1k \log^2 {\hat{n}}{\text{ for all }}i$. For all vertices $u \in S^i$ for some $i$, build a shortest path tree to and from $u$. $C {\leftarrow}\bigcup_{i=1}^{100\log {\hat{n}}} \bigcup_{u \in S^i} B_u((4K+1)R).$ \[line:updcover\] ${\mathrm{on}}[v] {\leftarrow}\textbf{false}.$ \[line:markdead\] \[line:searchv\] $T_v^i = \{ u \in S^i : d(v, u) \le KR \text{ and } d(u, w) \le 2KR {\text{ for all }}w \in S_v^j {\text{ for all }}1 \le j < i. \}$ \[line:maketi\] $S_v^i {\leftarrow}$ uniform sample of $T_v^i$ of size $50\log {\hat{n}}$. Return to line \[line:searchv\]. Have $D$ store all the $S_v^i$ and shortest path trees from all vertices $u \in S^i$ for some $i$. $\left(G\left[\{v : {\mathrm{on}}[v] = \textbf{true}\}\right], C, D \right).$ \[algo:buildsimilar\] $C {\leftarrow}\emptyset$. $K {\leftarrow}10k \log k$. ${\mathrm{on}}[v] {\leftarrow}\textbf{true} {\text{ for all }}v \in V(G).$ \[line:findv\] $E_v^i {\leftarrow}\{ u \in V(G) : {\mathrm{on}}[u] \text{ and } {\textsc{Similar}}(G, u, v, D, i, R) \text{ and } u \text{ reachable from } v \text{ through } E_v^i\}$. \[line:makeevi\] $E_v^{i+1} {\leftarrow}\{ u \in V(G) : {\mathrm{on}}[u] \text{ and } {\textsc{Similar}}(G, u, v, D, i+1, R) \text{ and } u \text{ reachable from } v \text{ through } E_v^{i+1}\}$. \[line:goodcuttrue\] $C {\leftarrow}C \cup {\textsc{RoundtripCover2}}(E_v^{i+1}, k, R)$. ${\mathrm{on}}[v] {\leftarrow}\textbf{false} {\text{ for all }}v \in E_v^i.$ Break loop and return to line \[line:findv\]. $C$. \[algo:ballgrow\] $K {\leftarrow}10k \log k.$ **false** **false** **true** \[algo:similartest\] **true** **false** \[algo:goodcut2\] Explanation of algorithms {#sec:explanation} ------------------------- #### Explanation of \[algo:roundtripcover2\], \[algo:buildsimilar\], \[algo:ballgrow\], \[algo:similartest\], \[algo:goodcut2\]. Throughout, we let ${\hat{n}}$ be the number of vertices at the top level of recursion in the algorithms and we let $K {:=}10k \log k$ for integer $k$. We start by explaining \[algo:buildsimilar\] ([<span style="font-variant:small-caps;">BuildSimilar</span>]{}), which builds a data structure which allows efficient similarity queries. It follows the same blueprint as \[alg:similar\_vertices\]. The algorithm first selects sets $S^i$ for $1 \le i \le 100 \log {\hat{n}}$, where $|S^i| = 100n^{1/k} \log^2 {\hat{n}}$ for all $i$. It then computes shortest path trees to and from all vertices in all $S^i$. The algorithm then adds roundtrip balls of radius $O(KR)$ centered at each $u \in S^i$ to our roundtrip cover. The algorithm then marks all vertices $v$ from the graph that are within distance $2KR$ both to and from some vertex $u$ in some $S^i$ as not turned on anymore. Then for all vertices $v \in V(G)$ the algorithm builds sets $T_v^i$ and a uniform sample $S_v^i$ of $T_v^i$ of size $O(\log {\hat{n}})$ that allow us to “test" whether another vertex $u$ is similar to $v$, i.e. could potentially be in a cycle of length $O(KR)$ with $v$. Eventually, $|T_v^i|$ gets small, and the algorithm stops processing vertex $v$. Finally, it returns the graph $G'$ of all still on vertices, the updated roundtrip cover, and the data structure $D$ for similarity testing consisting of all the $S_v^i$ for each vertex $v$ and shortest path trees from all $u \in S^i$. Now we explain \[algo:similartest\] ([<span style="font-variant:small-caps;">Similar</span>]{}), which uses the data structure $D$ computed by [<span style="font-variant:small-caps;">BuildSimilar</span>]{} to decide whether vertex $u$ is $i$-similar to vertex $v$. It returns true if and only if $$d(v, u) \le iR \text{ and } d(u, w) \le (i+K)R {\text{ for all }}w \in S_v^j {\text{ for all }}1 \le j \le 100\log {\hat{n}}.$$ Intuitively, this contains a ball around $v$ of distance $iR$ that contains all vertices which could potentially be in a cycle of length $KR$ with $v$, according to the algorithm. Now we explain \[algo:goodcut2\] ([<span style="font-variant:small-caps;">GoodCut2</span>]{}), which decides whether cutting out ball $B_1$ and recursing on $B_2$ constitutes good enough progress. This simply takes as input two balls $B_1$ and $B_2$ and decides whether recursing on $B_2$ and then deleting $B_1$ is good enough progress in trying to achieve a ${\tilde{O}}(n^{1+O(k^{-1})})$ total size of roundtrip covers. Here, we check only the vertex condition instead of the edge condition (different from \[algo:goodcut\] [<span style="font-variant:small-caps;">GoodCut</span>]{}) because we have already reduced to the case where our graph $G$ is approximately regular (\[lemma:regular\]). Now we explain \[algo:ballgrow\] ([<span style="font-variant:small-caps;">BallGrow</span>]{}), which grows the balls $E_v^i$. ${\mathrm{on}}[v] = \textbf{false}$ if vertex $v$ has been resolved, i.e. we can ensure that for any $u$ with $d(v {\leftrightarrows}u) \le R$, that $v$ and $u$ are in a roundtrip ball of diameter $O(KR).$ Otherwise, ${\mathrm{on}}[v] = \textbf{true}.$ We now grow balls $E_v^0, E_v^1, \cdots, E_v^K$ around $v$, up until line \[line:goodcuttrue\] is satisfied. Our main claim is that when we recurse on $E_v^{i+1}$, then we can safely remove all vertices in $E_v^i$. While the definition in line \[line:makeevi\] $$E_v^i {\leftarrow}\{ u \in V(G) : {\mathrm{on}}[u] \text{ and } {\textsc{Similar}}(G, u, v, D, i, R) \text{ and } u \text{ reachable from } v \text{ through } E_v^i\}$$ may seem recursive, all we mean is to say that we run a search from $v$, only keeping vertices which are $i$-similar, i.e. ${\textsc{Similar}}(G, u, v, D, i, R)$ is true. Finally, our main algorithm \[algo:roundtripcover2\] ([<span style="font-variant:small-caps;">RoundtripCover2</span>]{}) first calls [<span style="font-variant:small-caps;">BuildSimilar</span>]{} to build the similarity data structure needed for [<span style="font-variant:small-caps;">BallGrow</span>]{}. It also removes vertices from $G$ that were already resolved (i.e. in cycles of length $2KR$) to get a graph $G'$. Then it grows balls to partition $G'$ and recurse. Analysis {#sec:analysis2} -------- In this section we analyze the above algorithms. We first show that the number of similar vertices to any vertex $v$ in $G'$ (line \[line:callballgrow\]) is at most $n^\frac{k-1}{k}$ with high probability. \[lemma:Ksizebound\] Consider an execution of ${\textsc{RoundtripCover2}}(G_0, k, r)$ on an ${\hat{n}}$-vertex vertex graph $G_0$. Consider a recursive execution of ${\textsc{RoundtripCover2}}(G, k, r)$ on an $n$-vertex $m$-edge graph $G$. Consider the resulting execution ${\textsc{BallGrow}}(G', k, R, D)$ (line \[line:callballgrow\]). With probability at least $1 - {\hat{n}}^{-7}$ we have that for all $v \in V(G')$ that the number of vertices $u \in V(G')$ satisfying $$d(v, u) \le KR \text{ and } d(u, w) \le 2KR {\text{ for all }}w \in S_v^j {\text{ for all }}1 \le j \le 100\log{\hat{n}}$$ is at most $n^\frac{k-1}{k}.$ We follow the same approach as the proofs in \[sec:constspanner\]. Consider a vertex $v \in V(G)$. Define $$H_v^i := \{ u \in V(G) : d(v, u) \le KR \text{ and } d(u, w) \le 2KR {\text{ for all }}w \in S_v^j {\text{ for all }}1 \le j < i \},$$ i.e. all vertices $u \in V(G)$ which would “pass" the $i$-th level similarity test for $v$. Our main claim is that if $|H_v^i| \ge n^\frac{k-1}{k}$, then we have that $\frac{|H_v^{i+1}|}{|H_v^i|} \le \frac{9}{10}$ with high probability. This implies the result, because if $|H_v^{100 \log {\hat{n}}}| \ge n^\frac{k-1}{k}$ still, then we have that $$|H_v^{100 \log {\hat{n}}}| \le \left(\frac{9}{10}\right)^{100\log{\hat{n}}} n < 1,$$ an obvious contradiction. Now we show that if $|H_v^i| \ge n^\frac{k-1}{k}$, then we have that $\frac{|H_v^{i+1}|}{|H_v^i|} \le \frac{9}{10}$ with high probability. Note that by definition that $T_v^i = S^i \cap H_v^i$. It is direct to verify by a Chernoff bound that $|T_v^i| \ge 50 \log {\hat{n}}$ with probability at least $1 - {\hat{n}}^{-10}$ assuming that $|H_v^i| \ge n^\frac{k-1}{k}.$ By the definition of $S_v^i$ (a uniformly random subset of $T_v^i$ of size $50 \log {\hat{n}}$) and symmetry we can think of $S_v^i$ simply as a uniformly random subset of $H_v^i$ of size $50 \log {\hat{n}}.$ We now argue that for at least $\frac{9}{10}$ fraction of vertices in $w \in H_v^i$ we have that $$\Pr_{w' \in H_v^i}\left[d(w, w') \le 2KR\right] \le \frac{4}{5},$$ i.e. only $\frac{4}{5}$ fraction of vertices $w' \in H_v^i$ satisfy $d(w, w') \le 2KR$. Assume the contrary for contradiction. By the Pigeonhole principle, there are at least $$\left(\frac{9}{10}\cdot \frac{4}{5} - \frac{1}{2}\right)|H_v^i|^2 = .22|H_v^i|^2$$ (unordered) pairs of vertices $w, w' \in H_v^i$ such that both $d(w, w') \le 2KR$ and $d(w', w) \le 2KR$. By the Pigeonhole principle again, there must be a vertex $w \in H_v^i$ for which at least $.44 |H_v^i|$ vertices $w'$ satisfy both $d(w, w') \le 2KR$ and $d(w', w) \le 2KR$, so $d(w, w') \le 4KR.$ Now, note that $.44 |H_v^i| \ge .44 n^\frac{k-1}{k}$ by our condition. We argue that this is impossible because $v$ should have been marked as not on with high probability in line \[line:markdead\]. Indeed, the probability that $v$ failed to get marked as not on is at most $$\left(1 - \frac{.44n^\frac{k-1}{k}}{n}\right)^{\sum_{i=1}^{100\log {\hat{n}}} |S^i|} \le 1 - {\hat{n}}^{-20}$$ as desired. Now, consider the $\frac{9}{10}$ fraction of vertices $w \in H_v^i$ with $$\Pr_{w' \in H_v^i}\left[d(w, w') \le 2KR\right] \le \frac{4}{5}.$$ For each of these vertices, the probability that $d(w, w')$ for all $w' \in S_v^i$ is at most $\left(\frac{4}{5}\right)^{50\log{\hat{n}}} \le 1-{\hat{n}}^{-10}.$ By definition then, we have that $\frac{|H_v^{i+1}|}{|H_v^i|} \le \frac{1}{10}$ by definition, as the $\frac{9}{10}$ fraction of vertices in $H_v^i$ discussed in this paragraph will with high probability not be in $H_v^{i+1}.$ We next claim that the ball growing scheme of \[algo:ballgrow\] satisfies some important conditions, which intuitively make the $E_v^i$ look like balls of radius $iR$. \[lemma:ballgrowworks\] Consider an execution of ${\textsc{RoundtripCover2}}(G_0, k, R)$ on $n$-vertex $m$-edge graph $G_0$. Now, consider the resulting execution of ${\textsc{BallGrow}}(G, k, R, D)$ on graph $G$. We have that in the execution for all $v \in V(G)$ that 1. $v \in E_v^0$. 2. $E_v^i \subseteq E_v^{i+1}$ for $0 \le i \le K-1$. 3. For $0 \le i \le K-1$, if $u \in E_v^i$ and $d(u {\leftrightarrows}u') \le R$, then $u' \in E_v^{i+1}$. For the first claim, note that vertices $w \in S_v^j$ for all $j$ satisfy $d(v, w) \le KR$ by line \[line:maketi\] of [<span style="font-variant:small-caps;">BuildSimilar</span>]{}. Therefore, $v$ satisfies all conditions of being in $E_v^i$ on line \[line:makeevi\] of [<span style="font-variant:small-caps;">BallGrow</span>]{}. The second claim is obvious from looking at the the definition of $E_v^i$ in line \[line:makeevi\] of [<span style="font-variant:small-caps;">BallGrow</span>]{}. For the third claim, note that if $u \in E_v^i$ and $d(u {\leftrightarrows}u') \le R$ then $d(v, u') \le d(v, u) + R = (i+1)R$. Also, for any $w \in S_v^j$ for $1 \le j \le 100\log {\hat{n}}$ we have that $d(u', w) \le d(u, w) + R \le (i+1+K)R.$ Hence $u' \in E_v^{i+1}$ as desired. We now show the analogue to \[lemma:inbound\], specifically that for some iteration $0 \le i \le K-1$ in [<span style="font-variant:small-caps;">BallGrow</span>]{}, we have that the condition in line \[line:goodcuttrue\] triggers. Consider an execution of ${\textsc{RoundtripCover2}}(G_0, k, R)$ on $n$-vertex $m$-edge graph $G_0$. Now, consider the resulting execution of ${\textsc{BallGrow}}(G, k, R, D)$ on graph $G$. For some $0 \le i \le K-1$ we have that the condition in line \[line:goodcuttrue\] is true, i.e. we get a good cut. \[lemma:inbound2\] The computation proceeds the same way as in \[lemma:inbound\]. Assume for contradiction that the condition in line \[line:goodcuttrue\] is never true, so $$|V(E_v^{i+1})| > n^\frac1k |V(E_v^i)|^\frac{k-1}{k}.$$ By \[lemma:ballgrowworks\], we know that $|V(E_v^0)| \ge 1$, as $v \in E_v^0.$ Therefore, one can check by induction that $$|V(E_v^i)| \ge n^{1-\left(\frac{k-1}{k}\right)^i}.$$ For $K = 10k\log k$ we have that $$|V(E_v^K)| \ge n^{1-\left(\frac{k-1}{k}\right)^K} > n^\frac{k-1}{k},$$ which contradicts \[lemma:Ksizebound\]. We turn to proving \[thm:roundtripcover2\]. We break the analysis into pieces. We show that executing ${\textsc{RoundtripCover2}}(G, k, R)$ on a $n$-vertex $m$-edge graph $G$ returns a $(O(k\log k), R)$ roundtrip cover $C$ with total size of all balls at most $n^{1+O\left(\frac1k\right)}$ in time $m^{1+O\left(\frac1k\right)}$ with high probability. #### Returns a $(O(k\log k), R)$ roundtrip cover. We first argue that in a call to ${\textsc{BuildSimilar}}(G, k, R)$ that for all vertices $v$ where we marked ${\mathrm{on}}[v] = \textbf{false}$ that $v$ is properly resolved, i.e. for any vertex $u$ with $d(v {\leftrightarrows}u) \le R$ that $u$ and $v$ are in a roundtrip ball of radius at most $(4K+1)R.$ Indeed, note that if we mark ${\mathrm{on}}[v] = \textbf{false}$, then there must have been a vertex $w$ for which $w \in S^j$ for some $j$, and $d(v {\leftrightarrows}w) \le 2KR + 2KR = 4KR.$ Then $d(u {\leftrightarrows}w) \le (4K+1)R$. We have added the ball $B_w((4K+1)R)$ to our roundtrip cover $C$, as desired (line \[line:updcover\]). The only other piece to verify is that when we mark ${\mathrm{on}}[v] = \textbf{false}$ in an execution of ${\textsc{BallGrow}}(G, k, R, D)$ that $v$ is properly resolved, i.e. that in some recursive subproblem we have that for all $u$ with $d(v {\leftrightarrows}u) \le R$ that $v$ and $u$ are in a roundtrip ball of radius at most $O(KR).$ But this holds immediately by \[lemma:ballgrowworks\]: if $v \in E_w^i$ for some $w$, and $d(v {\leftrightarrows}u) \le R$, then $u \in E_w^{i+1}$ as desired. #### Total size of balls in $C$ is $n^{1+O\left(\frac1k\right)}$ with high probability. The analysis here follows closely to the corresponding paragraph in \[sec:algo\]. We will show that the total size of all graphs processed in a single level of recursion the algorithm is at most $n^\frac{k}{k-1}$, where our initial call was ${\textsc{RoundtripCover2}}(G, k, R)$ for a $n$-vertex $m$-edge graph $G$. Then, the total size of all graphs processed in the recursion is ${\tilde{O}}(n^\frac{k}{k-1})$, as the recursion depth is at most logarithmic. Then the bound on total size of balls in $C$ follows as for a graph $G$ with $n$ vertices, the total size of balls added to $C$ during ${\textsc{BuildSimilar}}(G, k, R)$ is at most $$\sum_{j=1}^{100\log n} |S^j| = {\tilde{O}}(n^{1+\frac1k}).$$ To show that the total size of all graphs processed at recursion depth $\ell$ in the algorithm is at most $n^\frac{k}{k-1}$, we use induction. Indeed, this holds at the bottom level of recursion. Consider an execution of ${\textsc{BallGrow}}(G, k, R, D)$ on an $n$-vertex $m$-edge graph $G$. Let $F_1^1, F_2^1, \cdots, F_t^1$ be all the balls $E_v^i$ for which line \[line:goodcuttrue\] was satisfied (and we know that line \[line:goodcuttrue\] is satisfied for some $i$ by \[lemma:inbound2\]). Let $F_1^2, F_2^2, \cdots, F_t^2$ be the corresponding balls $E_v^{i+1}$. We have that $\sum_{i=1}^t |V(F_i^1)| \le n$, as we marked all vertices in $E_v^i$ as not on anymore if line \[line:goodcuttrue\] was satisfied for $E_v^i$ and $E_v^{i+1}$. Additionally, by the condition of [<span style="font-variant:small-caps;">GoodCut2</span>]{}, we have that $|V(F_i^2)| \le n^\frac1k |V(F_i^1)|^\frac{k-1}{k}.$ By induction, the total sizes of all graphs processed at depth $\ell$ through recursion on $F_1^2, F_2^2, \cdots, F_t^2$ is at most $$\sum_{i=1}^t |V(F_i^2)|^\frac{k}{k-1} \le \sum_{i=1}^t \left(n^\frac1k |V(F_i^1)|^\frac{k-1}{k}\right)^\frac{k}{k-1} \le \sum_{i=1}^t n^\frac{1}{k-1} |V(F_i^1)| \le n^\frac{k}{k-1}$$ as desired. #### Can be implemented to run in time $m^{1+O\left(\frac1k\right)}$ with high probability. The analysis in the above section on the total size of balls, we know that the total number of vertices in all graphs processed during the algorithm ${\textsc{RoundtripCover2}}(G, k, R)$ is at most $n^{1+O\left(\frac1k \right)}.$ As we have reduced to the case of regular graphs through \[lemma:regular\], the total number of edges in all graphs processed during ${\textsc{RoundtripCover2}}(G, k, R)$ is at most ${\tilde{O}}(\delta n^{1+O\left(\frac1k \right)}) \le {\tilde{O}}(m^{1+O\left(\frac1k\right)})$ for $\delta = O(m/n).$ We now argue that the non-recursive runtime of ${\textsc{RoundtripCover2}}(G, k, R)$ on a graph $G$ with $n$ vertices and $m$ edges is ${\tilde{O}}(m^{1+1/k})$. We start by analyzing \[algo:buildsimilar\] (${\textsc{BuildSimilar}}$). We have that in ${\textsc{BuildSimilar}}(G, k, R)$ that $\sum_{i=1}^{100\log {\hat{n}}} |S^i| \le {\tilde{O}}(n^{1/k})$, where ${\hat{n}}$ is the number of vertices in the graph at the top level of recursion. Therefore, building a shortest path tree to and from all vertices in $\bigcup_i S^i$ takes ${\tilde{O}}(m^{1+1/k})$ time using Dijkstra’s algorithm. Computing all the sets $T^i_v$ clearly takes time $$n \cdot \sum_{i=1}^{100\log {\hat{n}}} |S^i| \cdot K = {\tilde{O}}(m^{1+1/k}).$$ We proceed to analyze \[algo:similartest\] ([<span style="font-variant:small-caps;">Similar</span>]{}). This clearly takes ${\tilde{O}}(1)$ time per call, as we have precomputed all shortest path trees and distances to and from all vertices $u \in S^i$ in \[algo:buildsimilar\]. Also, \[algo:goodcut2\] ([<span style="font-variant:small-caps;">GoodCut2</span>]{}) also obviously takes $O(1)$ time per call. Now we analyze \[algo:ballgrow\] ([<span style="font-variant:small-caps;">BallGrow</span>]{}). We can build the sets $E_v^i$ by running any search from $v$, only keeping vertices $u$ that satisfy ${\textsc{Similar}}(G, u, v, D, i, R).$ This takes time proportional to ${\tilde{O}}(\delta |E_v^{i+1}|)$, where we have used that each call to [<span style="font-variant:small-caps;">Similar</span>]{} takes ${\tilde{O}}(1)$ time. In accounting for this runtime, we can push the contribution to the next recursion level (as we are recursing on $E_v^{i+1}$). Therefore, the total non-recursive runtime used is ${\tilde{O}}(m^{1+1/k})$ as claimed for an input graph with $m$ edges. Now, as the total number of edges over all graphs is ${\tilde{O}}(m^{1+O\left(\frac1k\right)})$, the total runtime would also be ${\tilde{O}}(m^{1+O\left(\frac1k\right)})$ as desired. We now use \[thm:roundtripcover2\] to get multiplicative girth approximation and roundtrip spanners, proving \[thm:const\_girth\] and \[thm:arb\_const\_spanner\]. We first show the result for unweighted graphs. To show this, run $$\text{{\textsc{RoundtripCover2}}}(G, O(k), 2^i) \text{ for } 0 \le i \le O(\log n).$$ Now, set our estimate $g'$ of the girth to be the smallest radius of any nontrivial ball that we had in a roundtrip cover. By the guarantees of [<span style="font-variant:small-caps;">RoundtripCover2</span>]{}, it is clear that $g \le g' \le O(k\log k) \cdot g$ as desired. It is clear that the algorithm runs in time $\tilde{O}(m^{1+\frac1k})$ by \[thm:roundtripcover2\]. We can extend this to weighted graphs by instead taking $0 \le i \le O(\log nW)$, where $W$ is the maximum edge weight. This can be improved to $O(\log n)$ by the same method as done in [@PRSTV18], where they give a general reduction by contracting small weight strongly connected components and deleting large weight edges (see Section 5.1 in [@PRSTV18] for more details). We first show the result for unweighted graphs. It is easy to see that $$\bigcup_{i = 0}^{O(\log n)} \text{{\textsc{RoundtripCover2}}}(G, O(k), 2^i)$$ is an $O(k\log k)$ spanner with $\tilde{O}(n^{1+\frac{1}{k}})$ edges by \[thm:roundtripcover2\]. It is clear that the algorithm runs in time $\tilde{O}(m^{1+\frac1k}).$ The extension to weighted graphs follows as in the above paragraph (proof of \[thm:const\_girth\]). Conclusion and Open Problems {#sec:open} ============================ In this paper we provided multiple results on computing round-trip spanners and multiplicative approximations to the girth of an arbitrary directed graph. Our results all either improve running times, decrease the use of randomness, or improve the approximation quality of previous results. Ultimately, this work brings the state-of-the art performance of roundtrip spanners algorithms on directed graphs closer to matching that for undirected graphs. An immediate open problem left open by our work is to fully close the gap between algorithmic guarantees for spanners of undirected graphs and roundtrip spanners of directed graphs and provide a deterministic algorithm which for all $k$ in $\tilde{O}(m n^{1/k})$ time computes a $O(k)$ roundtrip spanner with $\tilde{O}(n^{1 + 1/k})$ edges. This paper resolves this problem for $k = \Omega(\log n)$ and makes progress on it for smaller values of $k$; it is still open to resolve it for all $k$. Another key open problem is to further clarify the complexity of approximating the girth of a directed graph. Currently the only algorithms which provably outperform APSP for approximating the girth of a graph are Pachocki et. al. [@PRSTV18] and this paper. Consequently, all known girth approximation algorithms for directed graphs leverage techniques immediately applicable for spanner computation (with the sole possible exception of the algorithms of Section \[sec:constantapprox\]). Therefore, beyond improving roundtrip spanner routines to obtain an algorithm which can compute an $O(k)$-multiplicative approximation to the girth in $\tilde{O}(m n^{1/k})$, this suggests the even more challenging open problem of circumventing this “spanner barrier” to obtaining even faster running times. For undirected graphs, it possible to overcome this barrier in certain cases [@IR77; @LL09; @RW12; @DKMS17]. However, some of the techniques used in these results are known not to extend to directed graphs, see e.g. [@RW12; @PRSTV18]. Consequently, further clarifying the complexity of girth approximation beyond the spanner barrier with either improved algorithms or new conditional lower bounds remains an difficult and interesting frontier. One final open problem is to improve the parallel complexity of these routines. Previous work on the efficient construction of roundtrip spanners [@PRSTV18] provided such a result. Further, there have been recent advances in the efficient parallel computation of reachability in directed graphs [@Fineman18; @JLS19arxiv] and commute times of random walks. The combination of the ideas from these works with the results of this paper could be useful for obtaining further improvements for the efficient parallel computation of girth and roundtrip spanners [@CohenKPPRSV17]. Acknowledgements {#sec:acknowledgements} ================ We thank the anonymous reviewers for helpful feedback and suggestion of many of the open problems discussed in . We thank Jakub Pachocki, Liam Roditty, Roei Tov, and Virginia Vassilevska Williams for helpful discussions. [^1]: This project has received funding from the European Union’s Horizon 2020 research grant agreement 803118. [^2]: Research supported by the U.S. Department of Defense via an NDSEG fellowship. [^3]: Research supported by NSF CAREER Award CCF-1844855
{ "pile_set_name": "ArXiv" }
--- abstract: 'We reexamine the branching ratios, $CP$-asymmetries, and other observables in a large number of $B_q\to VV(q=u,d,s)$ decays in the perturbative QCD (PQCD) approach, where $V$ denotes a light vector meson $(\rho, K^*, \omega, \phi)$. The essential difference between this work and the earlier similar works is of parametric origin and in the estimates of the power corrections related to the ratio $r_i^2=m_{V_i}^2/m_B^2(i=2,3)$ ($m_V$ and $m_B$ denote the masses of the vector and $B$ meson, respectively). In particular, we use up-to-date distribution amplitudes for the final state mesons and keep the terms proportional to the ratio $r_i^2$ in our calculations. Our updated calculations are in agreement with the experimental data, except for a limited number of decays which we discuss. We emphasize that the penguin annihilation and the hard-scattering emission contributions are essential to understand the polarization anomaly, such as in the $B\to \phi K^*$ and $B_s \to \phi\phi$ decay modes. We also compare our results with those obtained in the QCD factorization (QCDF) approach and comment on the similarities and differences, which can be used to discriminate between these approaches in future experiments.' author: - 'Zhi-Tian Zou$^a$[^1], Ahmed Ali$^b$[^2], Cai-Dian Lü $^c$[^3], Xin Liu$^d$, Ying Li$^a$[^4]' title: | Improved Estimates of The $B_{(s)}\rightarrow V V$ Decays\ in Perturbative QCD Approach --- Introduction ============ Exclusive $B_q$ $(q=u,d,s)$ meson decays, especially $B_q\to VV$ modes, where $V$ stands for a light vector meson ($\rho, K^*, \omega, \phi$), have aroused a great deal of interest for both theorists [@qcdf1; @qcdf2; @qcdf3; @qcdfbtovv; @qcdfbsvv; @yang; @btovv; @bsvv; @jpg32] and in experiments [@hfag]. In contrast to the scalar and pseudoscalar mesons, vector mesons can be produced in several polarization states. Thus, the fraction of a given polarization state is an interesting observable, apart from the decay widths. Phenomenology of the $B_q\to VV$ decays offers rich opportunities for our understanding of the mechanism for hadronic weak decays and $CP$ asymmetry and searching for the effect of new physics beyond the standard model. In general, the underlying dynamics for such decays is extremely complicated, but in the heavy quark limit ($m_b\to \infty$), it is greatly simplified due to the factorization of the hadronic matrix elements in terms of the decay constants and form factors. Based on this, a number of two-body hadronic $B$ decays had been calculated in this so-called naive factorization approach [@nfa]. However these calculations encounter three major difficulties: (i) for the so-called penguin-dominated, and also for the color-suppressed tree-dominated decays, the predicted branching ratios are systematically below the measurements, (ii) this approach can not account for the direct $CP$ asymmetries measured in experiments, and (iii) the predictions of transverse polarization fraction in penguin-dominated charmless $B_q\to VV$ decays are too small to explain the data, in which large such fractions are measured. All these indicate that this factorization approach needs improvements, for example by including some more perturbative QCD contributions [@gaijin]. In the current market, there are essentially three approaches to implement perturbative improvements: QCD factorization (QCDF) [@qcdf; @pdm], perturbative QCD approach(PQCD) [@pqcd], and the soft-collinear effective theory (SCET) [@sect]. All these frameworks have been employed in the literature to quantitatively study the dynamics of the $B_q \to PP, VP, VV$ decays, having light pseudoscalar ($P$)and/or Vector ($V$) mesons in the final states. In the $B_q\to VV$ decays, as the $B_q$ meson is heavy, the vector mesons are energetic with $E_{V}\simeq m_{B}/2$. As the spectator quark ($u,d$ or $s$) in the $B_q$ meson is soft, a hard gluon exchange is needed to kick it into an energetic one to form a fast moving light vector meson. The theoretical picture here is that a hard gluon from the spectator quark connects with the other quarks of the four-quark operators of the weak interaction [@pqcd]. The underlying theory is thus a six-quark effective theory, and can be perturbatively calculated [@lv23275]. In contrast to the other two approaches (QCDF and SCET), the PQCD approach is based on the $k_T$ factorization formalism [@pqcd1; @pqcd2; @pqcd3]. The basic idea here is to take into account the transverse momentum $k_T$ of the valence quarks in the hadrons, as a result of which the end-point singularity in the collinear factorization (employed in the QCDF approach) can be avoided. On the other hand, the transverse momentum dependence introduces an additional energy scale leading to double logarithms in QCD corrections. These terms could be resummed through the renormalization group approach, which results in the appearance of the Sudakov form factor. This form factor effectively suppresses the end-point contribution of the distribution amplitude of the mesons in the small transverse momentum region, making the calculation in the PQCD approach reliable. It is worth mentioning that in this framework, the so-called annihilation diagrams are also perturbatively calculable without introducing additional parameters [@anni1; @anni2]. The PQCD approach has been successfully used to study a number of pure annihilation type decays, and these predictions were confirmed subsequently in experiments [@bsvv; @pipi; @anni2; @dk; @anniexp]. Thus, in our view, this method is reliable in dealing with the pure annihilation-type and annihilation-dominated decays as well. Several years ago, H. Y. Cheng and C. K. Chua updated [@qcdfbtovv; @qcdfbsvv] the previous predictions [@qcdf1; @qcdf2; @qcdf3] for $B_q \to VV$ decays in the QCDF factorization approach by taking the transverse polarization contributions into account, and using the updated values of the parameters in the input wave functions and the form factors. In the PQCD framework, although many studies of the two-body $B_q$-decays are available [@btovv; @bsvv; @jpg32], a reappraisal is needed for the following reasons: (i) In the previous studies, the terms proportional to “$r_{i}^2=m_{V_i}^2/m_B^2\, (i=2,3)$" have been omitted in the amplitudes, especially in the denominator of the propagators of virtual quarks and gluons. As we point out later, these terms do bring the earlier PQCD predictions in better accord in terms of the measured observables in some problematic cases, such as the $B\to \phi K^*$ and $B_s \to \phi\phi$ decays, (ii) recent progress in the study of the distribution amplitudes of the vector meson, especially for the $\phi$ meson, undertaken in the context of the QCD sum rules, may significantly impact on some of the calculations done earlier, and (iii) Experimental data for some of the $B_q\to VV$ decays, such as the branching ratio and the polarization fractions of $B_s \to \phi\phi$, are now available. In addition, we work out a number of observables, such as $\phi_{\parallel}$, $\phi_{\perp}$, $A_{CP}^0$, $A_{CP}^{\perp}$, $\Delta\phi_{\parallel}$ and $\Delta\phi_{\perp}$ for the first time in PQCD. Among others, we revisit the $B\to \rho~(\omega)\phi$ decay modes, the direct $CP$ asymmetry of which could help us distinguish the PQCD and competing approaches. A related issue is the large fraction of the transverse polarization observed in some of these decays. In the PQCD framework, penguin-annihilation contribution is the key to understanding this phenomenon. Especially, the chirally enhanced (S-P)(S+P) penguin-annihilation gives rise to large transverse polarizations. Together with the hard spectator-scattering contributions, this could help solve the transverse polarization puzzle in the penguin-dominated $B_q \to VV$ decays. This work is organized as follows. In Sec. \[sec:function\], we outline the framework of the PQCD approach and specify the various input parameters, such as the wave functions and decay constants. Details of the perturbative calculations for the $B_q\to VV$ decays are presented in in Sec. \[jiexi\], and the various input functions are given in the Appendix. Numerical results of our calculations are presented in Sec. \[result\] and compared in detail with the available experiments and earlier theoretical works. Finally, a short summary is given in Sec. \[summary\]. FORMALISM AND WAVE FUNCTION {#sec:function} =========================== Our goal is to calculate the transition matrix elements: $$\begin{aligned} \mathcal{M}\propto \langle VV|\mathcal{H}_{eff}|B_{q}\rangle~,\end{aligned}$$ with the weak effective Hamiltonian $\mathcal{H}_{eff}$ written as [@rmp681125] $$\begin{aligned} \mathcal{H}_{eff}=\frac{G_{F}}{\sqrt{2}}\left\{ V_{ub}^{*}V_{uX}\left[C_{1}(\mu)O_{1}^{q}(\mu)+C_{2}(\mu)O_{2}^{q}(\mu)\right] -V_{tb}^{*}V_{tX}\left[\sum_{i=3}^{10}C_{i}(\mu)O_{i}(\mu)\right]\right\}~. \label{eq:heff}\end{aligned}$$ Here, $V_{ub(X)}$ and $V_{tb(X)}$ ($X=d,s$) are the CKM matrix elements, $C_{i}(\mu)$ are the effective Wilson coefficient calculated at the scale $\mu$, and the local four-quark operators $O_{j}\,(j=1,...,10)$ are defined and classified as follows: - Current-current (tree) operators, $$\begin{aligned} O_{1}^{u}=(\bar{b}_{\alpha}u_{\beta})_{V-A}(\bar{u}_{\beta}X_{\alpha})_{V-A}, \;\;\;O_{2}^{u}=(\bar{b}_{\alpha}u_{\alpha}) _{V-A}(\bar{u}_{\beta}X_{\beta})_{V-A},\end{aligned}$$ - QCD penguin operators, $$\begin{aligned} &&O_{3}=(\bar{b}_{\alpha}X_{\alpha})_{V-A}\sum_{q^{\prime}}(\bar{q}^{\prime}_{\beta}q^{\prime}_{\beta})_{V-A},\;\;\; O_{4}=(\bar{b}_{\alpha}X_{\beta})_{V-A}\sum_{q^{\prime}}(\bar{q}^{\prime}_{\beta}q^{\prime}_{\alpha})_{V-A},\\ &&O_{5}=(\bar{b}_{\alpha}X_{\alpha})_{V-A}\sum_{q^{\prime}}(\bar{q}^{\prime}_{\beta}q^{\prime}_{\beta})_{V+A},\;\;\; O_{6}=(\bar{b}_{\alpha}X_{\beta})_{V-A}\sum_{q^{\prime}}(\bar{q}^{\prime}_{\beta}q^{\prime}_{\alpha})_{V+A},\end{aligned}$$ - Electroweak penguin operators, $$\begin{aligned} &&O_{7}=\frac{3}{2}(\bar{b}_{\alpha}X_{\alpha})_{V-A}\sum_{q^{\prime}}e_{q^{\prime}}(\bar{q}^{\prime}_{\beta}q^{\prime}_{\beta})_{V+A}, \;\;O_{8}=\frac{3}{2}(\bar{b}_{\alpha}X_{\beta})_{V-A}\sum_{q^{\prime}}e_{q^{\prime}}(\bar{q}^{\prime}_{\beta}q^{\prime}_{\alpha})_{V+A},\\ &&O_{9}=\frac{3}{2}(\bar{b}_{\alpha}X_{\alpha})_{V-A}\sum_{q^{\prime}}e_{q^{\prime}}(\bar{q}^{\prime}_{\beta}q^{\prime}_{\beta})_{V-A},\;\; O_{10}=\frac{3}{2}(\bar{b}_{\alpha}X_{\beta})_{V-A}\sum_{q^{\prime}}e_{q^{\prime}}(\bar{q}^{\prime}_{\beta}q^{\prime}_{\alpha})_{V-A},\end{aligned}$$ with the SU(3) color indices $\alpha$ and $\beta$ and the active quarks $q^{\prime}=(u,d,s,c)$. The left-handed (right-handed) current $V\pm A$ are defined as $\gamma_{\mu}(1\pm\gamma_{5})$. Following [@prd58094009], we introduce the following combinations $a_{i}$ of the Wilson coefficients: $$\begin{aligned} &&a_{1}=C_{2}+C_{1}/3,\;\;\;\;\;\;a_{2}=C_{1}+C_{2}/3,\nonumber\\ &&a_{i}=C_{i}+C_{i\pm 1}/3,\,i=3,5,7,9\, / \,4,6,8,10.\end{aligned}$$ In the perturbative approach to hadronic $B_q$ decays, several typical scales are encountered with large logarithms involving the ratios of these scales. They are resummed using the renormalization group (RG) techniques. Standard model specifies the Wilson coefficients at the electroweak scale $m_W$, the W boson mass, and the RG equations enable us to evaluate the dynamical effects in scaling the Wilson coefficients in Eq. (\[eq:heff\]) from $m_W$ to $m_b$, the $b$-quark mass. The physics between the scale $m_b$ and the factorization scale $\Lambda_h$, taken typically as $\Lambda_h \simeq \sqrt{m_b\Lambda_{\rm QCD}}$, can be calculated perturbatively and included in the so-called hard kernel in the PQCD approach. The soft dynamics below the factorization scale $\Lambda_h$ is nonperturbative and is described by the hadronic wave functions of the mesons involved in the decays $B_q \to VV$. Finally, based on the factorization ansatz, the decay amplitudes are described by the convolution of the Wilson coefficients $C(t)$, the hard scattering kernel $H(x_i,b_i,t)$ and the light-cone wave functions $\Phi_{M_{i},B}(x_j,b_j)$ of the mesons [@prd55and56]: $$\begin{aligned} \mathcal {A}\;\sim\;&&\int\,dx_{1}dx_{2}dx_{3}b_{1}db_{1}b_{2}db_{2}b_{3}db_{3}\nonumber\\ &&\times{\rm Tr}\left[C(t)\Phi_{B}(x_{1},b_{1})\Phi_{M_{2}}(x_{2},b_{2})\Phi_{M_{3}} (x_{3},b_{3})H(x_{i},b_{i},t)S_{t}(x_{i})e^{-S(t)}\right],\end{aligned}$$ where ${\rm Tr}$ denotes the trace over Dirac and color indices, $b_{i}$ are the conjugate variables of the quark transverse momenta $k_{iT}$, $x_{i}$ are the longitudinal momentum fractions carried by the quarks, and $t$ is the largest scale in the hard kernel $H(x_{i},b_{i},t)$. The jet function $S_{t}(x_{i})$ coming from the threshold resummation of the double logarithms $\ln^2 x_i$ smears the end-point singularities in $x_{i}$ [@prd66094010]. The Sudakov form factor $e^{-S(t)}$ from the resummation of the double logarithms suppresses the soft dynamics effectively i.e. the long distance contributions in the large-$b$ region [@prd57443; @lvepjc23275]. In the PQCD approach, both the initial and the final state meson wave functions are important non-perturbative inputs. For $B_{q}~(q=u,d,s)$ meson, the light-cone matrix element could be decomposed as [@bwave1; @bwave2] $$\begin{aligned} \int d^4z e^{ik\cdot z}\langle0|q_{\beta}(z)\bar{b}_{\alpha}(0)|B_{q}(P_{B_q})\rangle= \frac{i}{\sqrt{6}}\left\{(\makebox[-1.5pt][l]{/}P_{B_q}+M_{B_q})\gamma_{5}\left[\phi_{B_q}(k)- \frac{\makebox[-1.5pt][l]{/}n-\makebox[-1.2pt][l]{/}v}{\sqrt{2}\bar{\phi}_{B_q}}(k)\right]\right\} _{\beta\alpha},\end{aligned}$$ where $n=(1,0,\vec{0}_T)$ and $v=(0,1,\vec{0}_T)$ are the unit vectors of the light-cone coordinate system. Corresponding to the two Lorentz structures in the $B_{q}$ meson distribution amplitudes, there are two wave functions $\phi_{B_q}(k)$ and $\bar{\phi}_{B_q}(k)$, obeying the following normalization conditions: $$\begin{aligned} \int \frac{d^4k}{(2\pi)^4}\phi_{B_q}(k)=\frac{f_{B_q}}{2\sqrt{6}},\;\;\int\frac{d^4k}{(2\pi)^4}\bar{\phi}_{B_q}(k)=0,\end{aligned}$$ where $f_{B_q}$ is the decay constant of the $B_q$ meson. Due to the numerical suppression, the contribution of $\bar{\phi}_{B}$ is often neglected. Finally, for convenience, the wave function of $B$ meson can be expressed as: $$\begin{aligned} \Phi_{B_{q}}(x,b)=\frac{i}{\sqrt{6}}(\makebox[-1.5pt][l]{/}P_{B_q}+M_{B_q})\gamma_{5} \phi_{B_q}(x,b),\end{aligned}$$ with the light-cone distribution amplitude $$\begin{aligned} \phi_{B_q}(x,b)=N_{B_q}x^2(1-x^2)\exp\left[-\frac{M_{B_q}^2x^2}{2\omega_{q}} -\frac{1}{2}w_{q}^2b^2\right],\end{aligned}$$ where $N_{B_q}$ is a normalization factor and $\omega_{q}$ is a shape parameter. For $B^0$($B^{\pm}$) meson, we use $\omega_q=0.4\pm0.04$ GeV, which is determined by the calculation of form factor and other well known decay modes [@pqcd1; @pqcd2; @omg]. Taking into account the small SU(3) breaking and the fact that the $s$ quark is heavier than the $u$ or $d$ quark, we use the shape parameter $\omega_s=0.5\pm0.05$ GeV for the $B_s$ meson, indicating that the $s$ quark momentum fraction is larger than that of the $u$ or $d$ quark in the $B^{\pm}$ or $B^0$ meson [@bsvv]. The light vector meson is treated as a light-light quark-antiquark system with the momentum $P^2=M_V^2$, and its polarization vectors $\epsilon$ include one longitudinal polarization vector $\epsilon_L$ and two transverse polarization vectors $\epsilon_T$, which are defined in [@btovv; @bksphi]. Up to twist-3, the vector meson wave functions are given by [@vwave]: $$\begin{aligned} &&\Phi_{V}^{L}\,=\,\frac{1}{\sqrt{6}}\left[M_{V}\makebox[0pt][l]{/} \epsilon_{L}\phi_{V}(x)\,+\,\makebox[0pt][l]{/}\epsilon_{L}\makebox[-1.5pt][l] {/}P\phi_{V}^{t}(x)+M_{V}\phi_{V}^{s}(x)\right]\nonumber\\ &&\Phi_{V}^{\perp}\,=\,\frac{1}{\sqrt{6}}\left[M_{V}\makebox[0pt][l]{/} \epsilon_{T}\phi_{V}^{v}(x)\,+\,\makebox[0pt][l]{/}\epsilon_{T}\makebox[-1.5pt][l] {/}P\phi_{V}^{T}(x)\,+\,M_{V}i\epsilon_{\mu\nu\rho\sigma}\gamma_{5} \gamma^{\mu}\epsilon_{T}^{\nu}n^{\rho}v^{\sigma}\phi_{V}^{a}(x)\right],\end{aligned}$$ for the longitudinal polarization and the transverse polarization, respectively. Here $\epsilon_{\mu\nu\rho\sigma}$ is Levi-Civita tensor with the convention $\epsilon^{0123}=1$. The twist-2 distribution amplitudes are given by $$\begin{aligned} &&\phi_{V}(x)\,=\,\frac{3f_{V}}{\sqrt{6}}x(1-x)\left[1+a^{\|}_{1V}C_{1}^{3/2}(t) +a_{2V}^{\|}C_2^{3/2}(t)\right],\\ &&\phi_{V}^{T}(x)\,=\,\frac{3f_{V}^T}{\sqrt{6}}x(1-x)\left[1+a^{\perp}_{1V}C_{1}^{3/2}(t) +a_{2V}^{\perp}C_2^{3/2}(t)\right],\end{aligned}$$ with $t=2x-1$, and $f_{V}^{(T)}$ are the decay constants of the vector meson, which for $V=\rho, \omega, K^*, \phi$ are shown numerically in Table \[tb:fv\]. For the Gegenbauer moments, we use the following values[@vwave; @jhep03]: $$\begin{aligned} &&a_{1\rho}^{\|(\perp)}=a_{1\omega}^{\|(\perp)}=a_{1\phi}^{\|(\perp)}=0, \;\;a_{1K^*}^{\|(\perp)}=0.03\pm0.02\,(0.04\pm0.03)~,\nonumber\\ &&a_{2\rho}^{\|(\perp)}=a_{2\omega}^{\|(\perp)}=0.15\pm0.07\,(0.14\pm0.06)\; a_{2\phi}^{\|(\perp)}=0\;(0.20\pm0.07)~,\nonumber\\ &&a_{2K^*}^{\|(\perp)}=0.11\pm0.09\;(0.10\pm0.08)~.\end{aligned}$$ For the twist-3 distribution amplitudes, for simplicity, we adopt the asymptotic forms $$\begin{aligned} &&\phi_{V}^t(x)=\frac{3f_V^T}{2\sqrt{6}}t^2,\;\;\phi_{V}^s(x)=\frac{3f_{V}^T}{2\sqrt{6}}(-t),\nonumber\\ &&\phi_V^v(x)=\frac{3f_V}{8\sqrt{6}}(1+t^2),\;\;\phi_V^a(x)=\frac{3f_V}{4\sqrt{6}}(-t).\end{aligned}$$ . [c!c!c]{} & [$f_{V}$(MeV)]{} & [$f_{V}^{T}$(MeV)]{}\ & [$216\,\pm\,3$]{}& [$165\,\pm\,9$]{}\ & [$187\,\pm\,5$]{}& [$151\,\pm\,9$]{}\ & [$220\,\pm\,5$]{}& [$185\,\pm\,10$]{}\ & [$215\,\pm\,5$]{}& [$186\,\pm\,9$]{}\ \[tb:fv\] Perturbative calculation {#jiexi} ======================== At leading order, there are eight types of Feynman diagrams contributing to the $B_q\rightarrow VV$ decays, which are presented in Fig.\[fig:diagram\]. The first row shows the emission-type diagrams, with the first two contributing to the usual form factor; the last two are the so-called hard-scattering emission diagrams. In fact, the first two diagrams are the only contributions calculated in the naive factorization approach. The second row shows the annihilation-type diagrams, with the first two factorizable and the last two nonfactorizable. In the following, we shall give the general factorization amplitudes for these $B_q\rightarrow VV$ decays. We use the symbol $LL$ to describe the amplitude of the $(V-A)(V-A)$ operators, $LR$ denotes the amplitude of the $(V-A)(V+A)$ operators and $SP$ denotes that of $(S-P)(S+P)$ operators resulting from the Fierz transformation of the $(V-A)(V+A)$ operators. For the $B_q\to VV$ decays, both the longitudinal polarization and the transverse polarization contribute. The amplitudes can be decomposed as follows: $$\begin{aligned} \mathcal {A}(\epsilon_{2},\epsilon_{3})=i\mathcal {A}^{L}+i(\epsilon_{2}^{T*}\cdot\epsilon_{3}^{T*})\mathcal {A}^{N}+(\epsilon_{\mu\nu\alpha\beta}n^{\mu}v^{\nu}\epsilon_{2}^{T*\alpha}\epsilon_{3}^{T*\beta})\mathcal {A}^{T}, \label{amplitude}\end{aligned}$$ where $\mathcal{A}^{L}$ is the longitudinally polarized decay amplitude, $\mathcal {A}^{T}$ and $\mathcal {A}^{N}$ are the transversely polarized contributions, and $\epsilon^T$ is the transverse polarization vector of the vector meson. =11 cm The longitudinal polarization amplitudes for the factorizable emission diagrams in Fig.(a) and (b) are as follows: $$\begin{aligned} \mathcal{A}_{ef}^{LL(LR),L}&=&-8\pi C_F M_{B}^4f_{V_2}\int_{0}^{1}dx_1dx_3\int_{0}^{1/\Lambda}b_1db_1b_3db_3\phi_{B}(x_1,b_1)\left\{\left[(-1+x_3)\phi_3(x_3)\right.\right.\nonumber\\ &&\left.\left.+r_3(2x_3-1)(\phi_3^s(x_3)+\phi_3^t(x_3))\right] E_{ef}(t_a)h_{ef}(x_1,x_3(1-r_2^2),b_1,b_3)\right.\nonumber\\ &&\left.+2r_3\phi_3^s(x_3) E_{ef}(t_b)h_{ef}(x_3,x_1(1-r_2^2),b_3,b_1)\right\},\end{aligned}$$ where $r_{i}=\frac{M_{V_i}}{M_B}$ and $C_F=4/3$ is a color factor. The functions $h_{ef}$, $t_{a,b}$, and $E_{ef}$ can be found in Appendix \[app:a\]. There is no $(S-P)(S+P)$ type amplitude, as a vector meson can not be produced through this type of operators. In the PQCD approach, the traditional emission contribution is also the dominant one. Unknown higher order perturbative QCD corrections will influence the emission contributions as well as those from other topologies. At present, although the next-to-leading order (NLO) contributions have not been completed, the vertex correction has been done and is used to improve the predictions for the decays $B \to \pi \rho(\omega)$ and the $B \to \pi$ form factors [@zhourui], which allows us to estimate the stability of the emission diagram in NLO. The results quoted below are based on the leading order calculations, but we also estimate the uncertainties from the partial NLO contributions based on the available results, as explained in Sec. \[result\]. The last two diagrams in the first row in Fig.\[fig:diagram\] are the hard-scattering emission diagrams, whose contributions are given below: $$\begin{aligned} \mathcal{M}_{enf}^{LL,L}&=&-16\sqrt{\frac{2}{3}}\pi C_{F}M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\phi_{2}(x_{2})\nonumber\\ &&\times \left\{\left[(x_2-1)\phi_3(x_3)+r_3x_3(\phi_3^s(x_3)-\phi_3^t(x_3))\right] E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[(x_2+x_3)\phi_3(x_3)-r_3x_3(\phi_3^s(x_3)+\phi_3^t(x_3))\right] E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{enf}^{LR,L}&=&16\sqrt{\frac{2}{3}}\pi C_{F}r_{2}M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_{1},b_1)\nonumber\\ &&\times\left\{\left[r_3((x_2-x_3-1)(\phi_2^s(x_2)\phi_3^s(x_3)-\phi_2^t(x_2)\phi_3^t(x_3))\right.\right.\nonumber\\ &&\left.\left.+(x_2+x_3-1)(\phi_2^t(x_2)\phi_3^s(x_3)-\phi_2^s(x_2)\phi_3^t(x_3)))\right.\right.\nonumber\\ &&\left.\left.+(x_2-1)\phi_{3}(x_3)(\phi_2^s(x_2)+\phi_2^t(x_2))\right] E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[r_3((x_3-x_2)(\phi_2^t(x_2)\phi_3^s(x_3)+\phi_2^s(x_2)\phi_3^t(x_3))\right.\right.\nonumber\\ &&\left.\left.+(x_2+x_3)(\phi_2^s(x_2)\phi_3^s(x_3)+\phi_2^t(x_2)\phi_3^t(x_3)))\right.\right.\nonumber\\ &&\left.\left.+x_2\phi_{3}(x_3)(\phi_2^s(x_2)-\phi_2^t(x_2))\right]E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{enf}^{SP,L}&=&-16\sqrt{\frac{2}{3}}\pi C_{F}M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\phi_{2}(x_{2})\nonumber\\ && \times \left\{\left[\phi_{3}(x_3)(x_2-x_3-1)+r_3x_3(\phi_{3}^s(x_3)+\phi_3^t(x_3))\right]E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[\phi_3(x_3)x_2+r_3x_3(\phi_3^t(x_3)-\phi_3^s(x_3))\right]E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\}.\end{aligned}$$ The functions $t_{c(d)}$, $E_{enf}$, $h_{enf}$, $\alpha$, $\beta_{i}$ for the nonfactorizable emission diagrams are also listed in Appendix \[app:a\]. As is well known, the hard-scattering emission diagrams with a light meson (pseudoscalar or vector) are suppressed. This can be seen from the figures (c) and (d), which are symmetrical. But, compared with the figure (d), the anti-quark propagator in figure (c) has an additional negative sign. As a result, the two contributions cancel each other. Figures (e) and (f) are the factorizable annihilation diagrams, whose factorizable contributions are listed below: $$\begin{aligned} \mathcal{A}_{af}^{LL(LR),L}&=&8C_F\pi f_{B}M_{B}^{4}\int_{0}^{1}dx_2dx_3\int_{0}^{1/\Lambda}b_2db_2b_3db_3\nonumber\\ &&\times\left\{\left[\phi_{2}(x_2)\phi_3(x_3)(x_3-1)+2r_2r_3\phi_2^s(x_2)(x_3\phi_{3}^t(x_3)-(x_3-2)\phi_3^s(x_3))\right]\right.\nonumber\\ &&\left.\cdot E_{af}(t_e)h_{af}(\alpha_1,\beta,b_2,b_3)\right.\nonumber\\ &&-\left.\left[-x_2\phi_2(x_2)\phi_3(x_3)+2r_2r_3\phi_3^s(x_3)((x_2-1)\phi_2^t(x_2)+(x_2+1)\phi_2^s(x_2))\right]\right.\nonumber\\ &&\left.\cdot E_{af}(t_f)h_{af}(\alpha_2,\beta,b_3,b_2)\right\}, \label{eq:anlll}\end{aligned}$$ $$\begin{aligned} \mathcal{A}_{af}^{SP,L}&=&-16C_{F}f_{B}\pi M_{B}^{4}\int_{0}^{1}dx_2dx_3\int_{0}^{1/\Lambda}b_2db_2b_3db_3\nonumber\\ &&\times\left\{\left[2r_2\phi_3(x_3)\phi_{2}^{s}(x_2)+r_3(x_3-1)\phi_2(x_2)(\phi_{3}^s(x_3)+\phi_3^t(x_3))\right]\right.\nonumber\\ &&\left.\cdot E_{af}(t_e)h_{af}(\alpha_1,\beta,b_2,b_3)\right.\nonumber\\ &&\left.+\left[2r_3\phi_{2}(x_2)\phi_3^s(x_3)+r_2x_2\phi_3(x_3)(\phi_{2}^{t}(x_2)-\phi_2^s(x_2))\right]\right.\nonumber\\ &&\left.\cdot E_{af}(t_f)h_{af}(\alpha_2,\beta,b_3,b_2)\right\}, \label{eq:anspl}\end{aligned}$$ and the related scales and the hard functions listed in Appendix \[app:a\] The last two diagrams in Fig.1 are the nonfactorizable annihilation diagrams. The expressions for the corresponding amplitudes are as follows: $$\begin{aligned} \mathcal{M}_{anf}^{LL,L}&=&16\sqrt{\frac{2}{3}}C_F\pi M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times\left\{\left[r_2r_3(\phi_2^t(x_2)(\phi_3^t(x_3)(1-x_2+x_3)+\phi_3^s(x_3)(x_2+x_3-1))\right.\right.\nonumber\\ &&\left.\left.+\phi_2^s(x_2)(\phi_3^t(x_3)(1-x_2-x_3)+\phi_3^s(x_3)(x_2-x_3+3)))\right.\right.\nonumber\\ &&\left.\left.-x_2\phi_2(x_2)\phi_3(x_3)\right]E_{anf}(t_g)h_{anf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&-\left.\left[r_2r_3(\phi_{2}^{s}(x_2)(\phi_3^s(x_3)(1+x_2-x_3)+\phi_{3}^{t}(x_3)(x_2+x_3-1))\right.\right.\nonumber\\ &&\left.\left.+\phi_{2}^t(x_2)(\phi_3^s(x_3)(1-x_2-x_3)+\phi_3^t(x_3)(x_3-x_2-1)))\right.\right.\nonumber\\ &&\left.\left.+(x_3-1)\phi_2(x_2)\phi_3(x_3)\right]E_{anf}(t_h)h_{anf}(\alpha,\beta_2,b_1,b_2)\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{anf}^{LR,L}&=&16\sqrt{\frac{2}{3}}C_F\pi M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times\left\{\left[\phi_3(x_3)(\phi_{2}^{t}(x_{2})+\phi_{2}^{s}(x_2))r_{2}(x_{2}-2)\right.\right.\nonumber\\ &&\left.\left.-\phi_{2}(x_{2})(\phi_3^s(x_3)-\phi_3^t(x_3))r_{3}(x_{3}+1)\right]E_{anf}(t_g)h_{anf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[\phi_2(x_2)(\phi_{3}^{s}(x_{3})-\phi_{3}^{t}(x_3))r_{3}(x_{3}-1)\right.\right.\nonumber\\ &&\left.\left.-\phi_{3}(x_{3})(\phi_2^s(x_2)+\phi_2^t(x_2))r_{2}x_{3}\right]E_{anf}(t_h)h_{anf}(\alpha,\beta_2,b_1,b_2)\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{anf}^{SP,L}&=&16\sqrt{\frac{2}{3}}C_F\pi M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times\left\{\left[r_2r_3(\phi_2^t(x_2)(\phi_3^t(x_3)(1-x_2+x_3)-\phi_3^s(x_3)(x_2+x_3-1))\right.\right.\nonumber\\ &&\left.\left.+\phi_2^s(x_2)(\phi_3^t(x_3)(x_2+x_3-1)+\phi_3^s(x_3)(x_2-x_3+3)))\right.\right.\nonumber\\ &&\left.\left.+(x_3-1)\phi_2(x_2)\phi_3(x_3)\right]E_{anf}(t_g)h_{anf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&-\left.\left[r_2r_3(\phi_{2}^{s}(x_2)(\phi_3^s(x_3)(1+x_2-x_3)+\phi_{3}^{t}(x_3)(1-x_2-x_3))\right.\right.\nonumber\\ &&\left.\left.+\phi_{2}^t(x_2)(\phi_3^s(x_3)(x_2+x_3-1)+\phi_3^t(x_3)(x_3-x_2-1)))\right.\right.\nonumber\\ &&\left.\left.-x_2\phi_2(x_2)\phi_3(x_3)\right]E_{anf}(t_h)h_{anf}(\alpha,\beta_2,b_1,b_2)\right\}.\end{aligned}$$ For the $B_{(s)}\to VV$ decays, the transverse polarization amplitudes of the two factorizable emission diagrams yield: $$\begin{aligned} \mathcal{A}_{ef}^{LL(LR),N}&=&8\pi C_F M_{B}^4f_{V_2}r_2\int_{0}^{1}dx_1dx_3\int_{0}^{1/\Lambda}b_1db_1b_3db_3\phi_{B}(x_1,b_1)\left\{\left[\phi_3^T(x_3)\right.\right.\nonumber\\ &&\left.\left.+r_3((x_3+2)\phi_3^v(x_3)-x_3\phi_3^a(x_3))\right] E_{ef}(t_a)h_{ef}(x_1,x_3(1-r_2^2),b_1,b_3)\right.\nonumber\\ &&\left.+r_3(\phi_3^a(x_3)+\phi_3^v(x_3)) E_{ef}(t_b)h_{ef}(x_3,x_1(1-r_2^2),b_3,b_1)\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{A}_{ef}^{LL(LR),T}&=&-8\pi C_F M_{B}^4f_{V_2}r_2\int_{0}^{1}dx_1dx_3\int_{0}^{1/\Lambda}b_1db_1b_3db_3\phi_{B}(x_1,b_1)\left\{\left[\phi_3^T(x_3)\right.\right.\nonumber\\ &&\left.\left.+r_3((x_3+2)\phi_3^a(x_3)-x_3\phi_3^v(x_3))\right] E_{ef}(t_a)h_{ef}(x_1,x_3(1-r_2^2),b_1,b_3)\right.\nonumber\\ &&\left.+r_3(\phi_3^a(x_3)+\phi_3^v(x_3)) E_{ef}(t_b)h_{ef}(x_3,x_1(1-r_2^2),b_3,b_1)\right\}.\end{aligned}$$ The transverse polarization amplitudes of the two hard-scattering emission diagrams fig.(c) and (d) are given below: $$\begin{aligned} \mathcal{M}_{enf}^{LL,N}&=&16\sqrt{\frac{2}{3}}\pi C_{F}M_{B}^{4}r_2\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times \left\{\left[(1-x_2)\phi_3^T(x_3)(\phi_2^a(x_2)+\phi_2^v(x_2))\right] E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.-\left[2r_3(x_2+x_3)(\phi_2^a(x_2)\phi_3^a(x_3)+\phi_2^v(x_2)\phi_3^v(x_3))\right.\right.\nonumber\\ &&\left.\left.-x_2\phi_3^T(x_3)(\phi_2^a(x_2)+\phi_2^v(x_2))\right] E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{enf}^{LL,T}&=&16\sqrt{\frac{2}{3}}\pi C_{F}M_{B}^{4}r_2\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times \left\{\left[(x_2-1)\phi_3^T(x_3)(\phi_2^a(x_2)+\phi_2^v(x_2))\right] E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[2r_3(x_2+x_3)(\phi_2^a(x_2)\phi_3^v(x_3)+\phi_2^v(x_2)\phi_3^a(x_3))\right.\right.\nonumber\\ &&\left.\left.-x_2\phi_3^T(x_3)(\phi_2^a(x_2)+\phi_2^v(x_2))\right] E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{enf}^{LR,N}&=&16\sqrt{\frac{2}{3}}\pi C_{F}M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\phi_2^T(x_2)\nonumber\\ &&\times \left\{\left[r_3x_3(\phi_3^a(x_3)-\phi_3^v(x_3))-\phi_3^T(x_3)(r_2^2(x_2-1)-x_3r_3^2)\right]\right.\nonumber\\ &&\left.\cdot E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[r_3x_3(\phi_3^a(x_3)-\phi_3^v(x_3))+\phi_3^T(x_3)(r_2^2x_2+r_3^2x_3))\right]\right.\nonumber\\ &&\left. E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{enf}^{LR,T}&=&16\sqrt{\frac{2}{3}}\pi C_{F}M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\phi_2^T(x_2)\nonumber\\ &&\times \left\{\left[r_3x_3(\phi_3^v(x_3)-\phi_3^a(x_3))-\phi_3^T(x_3)(r_2^2(x_2-1)+x_3r_3^2)\right]\right.\nonumber\\ &&\left.\cdot E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[r_3x_3(\phi_3^v(x_3)-\phi_3^a(x_3))+\phi_3^T(x_3)(r_2^2x_2-r_3^2x_3))\right]\right.\nonumber\\ &&\left. E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{enf}^{SP,N}&=&16\sqrt{\frac{2}{3}}\pi C_{F}M_{B}^{4}r_2\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times \left\{\left[2r_3(x_3+1-x_2)(\phi_2^v(x_2)\phi_3^v(x_3)-\phi_2^a(x_2)\phi_3^a(x_3))\right.\right.\nonumber\\ &&\left.\left.+(x_2-1)\phi_3^T(x_3)(\phi_2^v(x_2)-\phi_2^a(x_2))\right]E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[x_2\phi_3^T(x_3)(\phi_2^a(x_2)-\phi_2^v(x_2))\right]E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{enf}^{SP,T}&=&16\sqrt{\frac{2}{3}}\pi C_{F}M_{B}^{4}r_2\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times \left\{\left[2r_3(x_2-x_3-1)(\phi_2^v(x_2)\phi_3^a(x_3)-\phi_2^a(x_2)\phi_3^v(x_3))\right.\right.\nonumber\\ &&\left.\left.+(x_2-1)\phi_3^T(x_3)(\phi_2^a(x_2)-\phi_2^v(x_2))\right]E_{enf}(t_c)h_{enf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&\left.+\left[x_2\phi_3^T(x_3)(\phi_2^v(x_2)-\phi_2^a(x_2))\right]E_{enf}(t_d)h_{enf}(\alpha,\beta_2,b_{1},b_{2})\right\}.\end{aligned}$$ The transverse polarization amplitudes for the factorizable annihilation diagrams are: $$\begin{aligned} \mathcal{A}_{af}^{LL(LR),N}&=&8C_F\pi f_{B}r_2r_3M_{B}^{4}\int_{0}^{1}dx_2dx_3\int_{0}^{1/\Lambda}b_2db_2b_3db_3\nonumber\\ &&\times\left\{\left[(\phi_{2}^v(x_2)\phi_3^v(x_3)+\phi_2^a(x_2)\phi_3^a(x_3))(x_3-2)-\right.\right.\nonumber\\ &&\left.\left. x_3(\phi_2^v(x_2)\phi_{3}^a(x_3)+\phi_2^a(x_2)\phi_3^v(x_3))\right]E_{af}(t_e)h_{af}(\alpha_1,\beta,b_2,b_3)\right.\nonumber\\ &&+\left.\left[(x_2-1)(\phi_2^a(x_2)\phi_3^v(x_3)+\phi_2^v(x_2)\phi_3^a(x_3))+\right.\right.\nonumber\\ &&\left.\left.(x_2+1)(\phi_2^a(x_2)\phi_3^a(x_3)+\phi_2^v(x_2)\phi_3^v(x_2))\right]E_{af}(t_f)h_{af}(\alpha_2,\beta,b_3,b_2)\right\}, \label{eq:anlln}\end{aligned}$$ $$\begin{aligned} \mathcal{A}_{af}^{LR,T}&=&-\mathcal{A}_{af}^{LL,T}=8C_F\pi f_{B}r_2r_3M_{B}^{4}\int_{0}^{1}dx_2dx_3\int_{0}^{1/\Lambda}b_2db_2b_3db_3\nonumber\\ &&\times\left\{\left[(\phi_{2}^a(x_2)\phi_3^v(x_3)+\phi_2^v(x_2)\phi_3^a(x_3))(x_3-2)-\right.\right.\nonumber\\ &&\left.\left. x_3(\phi_2^v(x_2)\phi_{3}^v(x_3)+\phi_2^a(x_2)\phi_3^a(x_3))\right]E_{af}(t_e)h_{af}(\alpha_1,\beta,b_2,b_3)\right.\nonumber\\ &&+\left.\left[(x_2-1)(\phi_2^v(x_2)\phi_3^v(x_3)+\phi_2^a(x_2)\phi_3^a(x_3))+(x_2+1)\right.\right.\nonumber\\ &&\left.\left.\cdot (\phi_2^v(x_2)\phi_3^a(x_3)+\phi_2^a(x_2)\phi_3^v(x_2))\right]E_{af}(t_f)h_{af}(\alpha_2,\beta,b_3,b_2)\right\}, \label{eq:anllt}\end{aligned}$$ $$\begin{aligned} \mathcal{A}_{af}^{SP,N}&=&-\mathcal{A}_{af}^{SP,T}=16C_F\pi f_{B}M_{B}^{4}\int_{0}^{1}dx_2dx_3\int_{0}^{1/\Lambda}b_2db_2b_3db_3\nonumber\\ &&\times\left\{\left[r_2\phi_3^T(x_3)(\phi_{2}^a(x_2)+\phi_2^v(x_2))\right]E_{af}(t_e)h_{af}(\alpha_1,\beta,b_2,b_3)\right.\nonumber\\ &&-\left.\left[r_3\phi_2^T(x_2)(\phi_3^a(x_3)-\phi_3^v(x_3))\right]E_{af}(t_f)h_{af}(\alpha_2,\beta,b_3,b_2)\right\}. \label{eq:anspt}\end{aligned}$$ From Eqs. (\[eq:anlll\]) and (\[eq:anlln\]), we find that large cancellations between the two annihilation type diagrams ( (e) and (f)) take place, as a result of which they are highly power suppressed. These two symmetric diagrams will cancel each other due to the relative negative sign introduced by the anti-quark propagator in diagram (e). This agrees with the naive argument that the annihilation contributions are negligible, especially for two identical final state mesons. For Eq. (\[eq:anllt\]), although the cancellations are not as severe as in Eq. (\[eq:anlll\]) and Eq. (\[eq:anlln\]), the contribution is also highly suppressed being proportional to $r_2r_3$. From Eqs. (\[eq:anspl\]) and (\[eq:anspt\]), it is interesting to see that no cancellations or power suppression are involved. The chiral enhancement here is important to explain the large direct $CP$ asymmetry, generated by the strong phase and the transverse polarization fraction in the penguin-dominated $B$ decays [@cptr]. The chirally enhanced penguin annihilation contribution will be discussed in Sec.\[result\]. For the nonfactorizable annihilation diagrams ((g) and (h)), we get: $$\begin{aligned} \mathcal{M}_{anf}^{LL,N}&=&\mathcal{M}_{anf}^{SP,N}=16\sqrt{\frac{2}{3}}C_F\pi M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times\left\{\left[-2r_2r_3(\phi_2^a(x_2)\phi_3^a(x_3)+\phi_2^v(x_2)\phi_3^v(x_3))\right.\right.\nonumber\\ &&\left.\left.-\phi_2^T(x_2)\phi_3^T(x_3)(r_2^2(x_2-1)-r_3^2x_3)\right]E_{anf}(t_g)h_{anf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&+\left.\left[\phi_2^T(x_2)\phi_3^T(x_3)(r_2^2x_2-r_3^2(x_3-1))\right]E_{anf}(t_h)h_{anf}(\alpha,\beta_2,b_1,b_2)\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{anf}^{LL,T}&=&-\mathcal{M}_{anf}^{SP,T}=16\sqrt{\frac{2}{3}}C_F\pi M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times\left\{\left[2r_2r_3(\phi_2^a(x_2)\phi_3^v(x_3)+\phi_2^v(x_2)\phi_3^a(x_3))\right.\right.\nonumber\\ &&\left.\left.-\phi_2^T(x_2)\phi_3^T(x_3)(r_2^2(x_2-1)+r_3^2x_3)\right]E_{anf}(t_g)h_{anf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&+\left.\left[\phi_2^T(x_2)\phi_3^T(x_3)(r_2^2x_2+r_3^2(x_3-1))\right]E_{anf}(t_h)h_{anf}(\alpha,\beta_2,b_1,b_2)\right\},\end{aligned}$$ $$\begin{aligned} \mathcal{M}_{anf}^{LR,N}&=&-\mathcal{M}_{anf}^{LR,T}=16\sqrt{\frac{2}{3}}C_F\pi M_{B}^{4}\int_{0}^{1}d[x]\int_{0}^{1/\Lambda}b_1db_1b_2db_2\phi_{B}(x_1,b_1)\nonumber\\ &&\times\left\{\left[r_2(2-x_2)\phi_3^T(x_3)(\phi_2^a(x_2)+\phi_2^v(x_2))\right.\right.\nonumber\\ &&\left.\left.r_3(x_3+1)\phi_2^T(x_2)(\phi_3^a(x_2)-\phi_3^v(x_3))\right] E_{anf}(t_g)h_{anf}(\alpha,\beta_1,b_1,b_2)\right.\nonumber\\ &&+\left.\left[r_3(x_3-1)\phi_2^T(x_2)(\phi_3^v(x_3)-\phi_3^a(x_3))+\right.\right.\nonumber\\ &&\left.\left.r_2x_2\phi_3^T(x_3)(\phi_2^a(x_2)+\phi_2^v(x_2))\right] E_{anf}(t_h)h_{anf}(\alpha,\beta_2,b_1,b_2)\right\},\end{aligned}$$ This completes the derivation of the various contributions in the $B_{(s)} \to VV$ decays. We now turn to the presentation of our numerical results in the next section. NUMERICAL RESULTS AND DISCUSSIONS {#result} ================================= We start this section by listing the input parameters used in our numerical calculations. The vector meson decay constants have been summarized in Table \[tb:fv\]. Other parameters, such as the CKM matrix elements, QCD scale (GeV), the masses (GeV) and the decay constant of the $B_{(s)}$ mesons (GeV) and the corresponding lifetimes (in $ps$) are taken from the PDG review [@pdg] and are given below: $$\begin{aligned} &&\Lambda_{\overline{MS}}^{f=4}=0.25\pm0.05,\;M_{B}=5.279,\;M_{B_s}=5.366,\;f_{B}=0.21\pm0.02,\;f_{B_s}=0.24\pm0.03,\nonumber\\ &&\tau_{B^{\pm/0}}= 1.641/1.519,\; \tau_{B_s}= 1.497,\;m_{b}({\rm pole})=4.8,\nonumber\\ &&V_{ud}=0.97427\pm0.00015,\;V_{us}=0.22534\pm0.00065,\;V_{ub}=0.00351^{+0.00015}_{-0.00014},\nonumber\\ &&V_{td}=0.00867^{+0.00029}_{-0.00031},\;V_{ts}=0.0404_{-0.0005}^{+0.0011},\;V_{tb}=0.999146^{+0.000021}_{-0.000046},\nonumber\\ &&\alpha=(89_{-4.2}^{+4.4})^\circ,\;\gamma=(68_{-11}^{+10})^\circ. \label{para}\end{aligned}$$ With three polarization amplitudes, $\mathcal{A}^{L}$, $\mathcal{A}^{N}$, and $\mathcal{A}^{T}$, the decay width is expressed as $$\begin{aligned} \Gamma(B_{(s)}\rightarrow VV)\,=\,\frac{|\overrightarrow{P}|}{8\pi M_{B}^{2}}\left[\mid\mathcal{A}^{L}\mid^{2}+2(\mid\mathcal{A}^{N}\mid^{2}+\mid\mathcal{A}^{T}\mid^{2})\right],\end{aligned}$$ where the analytic formulas for the amplitudes $\mathcal{A}^{L}$, $\mathcal{A}^{N}$ and $\mathcal{A}^{T}$ can be found in [@bsvv; @btovv; @jpg32]. In our convention, given in Eq. (\[amplitude\]), the helicity amplitudes are defined as follows: $$\begin{aligned} A_0=A_{L}=-\mathcal{A}^{L},\;\;\;A_{\parallel}=\sqrt{2}\mathcal{A}^{N},\;\;\;A_{\perp}=\sqrt{2}\mathcal{A}^{T},\end{aligned}$$ where the definitions of $A_{0,\parallel,\perp}$ are the same as those in [@bsvv]. In this work, we also predict their relative phases $\phi_{\parallel}=arg(A_{\parallel}/A_{0})$ and $\phi_{\perp}=arg(A_{\perp}/A_0)$. The polarization fractions $f_{L,\parallel,\perp}$ are defined as $$\begin{aligned} f_{L,\parallel,\perp}=\frac{\mid A_{L,\parallel,\perp}\mid^2}{\mid A_{0}\mid^{2}+\mid A_{\parallel}\mid^{2}+\mid A_{\perp}\mid^{2}}~.\end{aligned}$$ In addition to the direct $CP$ asymmetry parameters, we also evaluate the following observables: $$\begin{aligned} A_{CP}^0=(\bar{f}_L-f_L)/(\bar{f}_L+f_L),\;\;\; A_{CP}^{\perp}=(\bar{f}_{\perp}-f_{\perp})/(\bar{f}_{\perp}+f_{\perp}),\;\;\; \Delta\phi_{\parallel}=(\bar{\phi}_{\parallel}-\phi_{\parallel})/2~.\end{aligned}$$ Of these, $\Delta\phi_{\perp}=(\bar{\phi}_{\perp}-\phi_{\perp})/2$ is being worked out for the $B_{q}\to VV$ decays for the first time. Modes $Br(10^{-6})$ $f_{L}$(%) $f_{\perp}$ (%) $\phi_{\parallel}$(rad) $\phi_{\perp}$(rad) --------------------------------- ---------------------------- ----------------------------- ---------------------------- ----------------------------------- ---------------------------- -- -- -- -- -- [$B^0\rightarrow K^{*0}\phi$]{} [$9.8_{-3.8}^{+4.9}$]{} [$56.5_{-5.9}^{+5.8}$]{} [$21.3_{-2.9}^{+2.8}$]{} [$2.15_{-0.19}^{+0.22}$]{} [$2.14_{-0.19}^{+0.23}$]{} [$Exp$]{} [$9.8\pm0.6$]{} [$48\pm3$]{} [$24\pm5$]{} [$2.40\pm0.13$]{} [$2.39\pm0.13$]{} [$B^+\to K^{*+}\phi$]{} [$10.3_{-3.8}^{+4.9}$]{} [$57.0_{-5.9}^{+6.3}$]{} [$21.0_{-3.0}^{+3.0}$]{} [$2.18_{-0.19}^{+0.23}$]{} [$2.19_{-0.20}^{+0.22}$]{} [$10.0\pm2.0$]{} [$50\pm5$]{} [$20\pm5$]{} [$2.34\pm0.18$]{} [$2.58\pm0.17$]{} [$16.7_{-7.1}^{+8.9}$]{} [$34.7_{-7.1}^{+8.9}$]{} [$31.6_{-4.4}^{+3.5}$]{} [$2.01_{-0.23}^{+0.23}$]{} [$2.00_{-0.21}^{+0.24}$]{} [$19\pm5$]{} [$34.8\pm4.6$]{} [$36.5\pm4.4\pm2.7$ ]{} [$2.71^{+0.31}_{-0.36}\pm0.22$]{} [$$]{} [$0.39_{-0.17}^{+0.20}$]{} [$50.0_{-7.2}^{+8.1}$]{} [$24.2_{-3.9}^{+3.6}$]{} [$1.95_{-0.22}^{-0.21}$]{} [$1.95_{-0.22}^{+0.21}$]{} [$1.10\pm0.29$]{} [$51\pm15\pm7$]{} [$28\pm11\pm2$]{} [$1.75\pm0.58\pm0.30$]{} [$$]{} [$5.4_{-2.4}^{+3.0}$]{} [$38.3_{-10.5}^{+12.1}$]{} [$30.0_{-6.1}^{+5.3}$]{} [$2.12_{-0.25}^{+0.21}$]{} [$2.15_{-0.23}^{+0.22}$]{} [$28.1\pm4.6\pm5.6$]{} [$31\pm12\pm4$]{} [$38\pm11\pm4$]{} [$$]{} [$$]{} $A^{dir}_{CP}$(%) $A^{0}_{CP}$(%) $A^{\perp}_{CP}$(%) $\Delta\phi_{\parallel}(rad)$ $\Delta\phi_{\perp}(rad)$ [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$$]{} [$4\pm6$]{} [$-11\pm12$]{} [$0.11\pm0.22$]{} [$0.08\pm0.22$]{} [$-1.0_{-0.26}^{+0.18}$]{} [$-0.60_{-0.14}^{+0.12}$]{} [$0.75_{-0.11}^{+0.23}$]{} [$-0.05_{-0.33}^{+0.12}$]{} [$-0.01$]{} [$-1\pm8$]{} [$17\pm11\pm2$]{} [$22\pm24\pm8$]{} [$0.07\pm0.2\pm0.05$]{} [$0.19\pm0.20\pm0.07$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} [$0.0$]{} : Updated branching ratios, percentage of the longitudinal polarization $f_{L}$ and the transverse polarizations $f_{\perp}$, relative phases, and the $CP$ asymmetry parameters $A^{0}_{CP}$ and $A^{\perp}_{CP}$ in the $B \rightarrow K^{*0}\phi$, $B_s \rightarrow \bar{K}^{*0}\phi$, $B_s\to \phi\phi$ and $B_s\to \bar{K}^{*0}K^{*0}$ decays calculated in the PQCD approach. \[tb:zui\] [Class]{} [This work]{} [ QCDF]{} [ PQCD(former)]{} [Exp]{} -- ----------- ------------------------------------------------------- -------------------------------------- ---------------------------------- ------------------------- [C]{} [$0.27_{-0.09-0.04-0.01}^{+0.10+0.06+0.00}$]{} [$0.9_{-0.4-0.2}^{+1.5+1.1}$]{} [$0.9\pm0.1\pm0.1$]{} [$0.73\pm0.28$]{} [T]{} [$26.0_{-8.1-1.4-1.2}^{+10.1+1.4+1.5}$]{} [$25.5_{-2.6-1.5}^{+1.5+1.1}$]{} [$35\pm5\pm4$]{} [$24.2\pm3.1$]{} [E,P]{} [$0.40_{-0.12-0.08-0.01}^{+0.15+0.09+0.01}$]{} [$0.08_{-0.02-0.00}^{+0.02+0.36}$]{} [$1.9\pm0.2\pm0.2$]{} [$<1.6$]{} [C,P]{} [$0.50_{-0.18-0.07-0.05}^{+0.21+0.09+0.05}$]{} [$0.7_{-0.3-0.2}^{+0.9+0.7}$]{} [$1.2\pm0.2\pm0.2$]{} [$<4.0$]{} [P]{} [$3.3_{-1.1-0.9-0.1}^{+1.3+1.1+0.0}$]{} [$4.6_{-0.5-3.5}^{+0.6+3.5}$]{} [$5.9$]{} [$3.4^{+1.7}_{-1.3}$]{} [P]{} [$8.4_{-2.8-1.9-0.9}^{+3.1+2.2+0.6}$]{} [$8.9_{-1.0-5.5}^{+1.1+4.8}$]{} [13]{} [$<12.0$]{} [P]{} [$4.7_{-1.5-1.3-0.3}^{+2.1+1.6+0.2}$]{} [$2.5_{-0.4-1.5}^{+0.4+2.5}$]{} [9.6]{} [$2.0\pm0.5$]{} [P]{} [$0.34_{-0.11-0.09-0.03}^{+0.13+0.10+0.02}$]{} [$0.6_{-0.1-0.3}^{+0.1+0.2}$]{} [0.35]{} [$0.8\pm0.5$]{} [E]{} [$0.21_{-0.09-0.05-0.02}^{+0.09+0.03+0.01}$]{} [$0.1_{-0.0-0.1}^{+0.0+0.1}$]{} [0.11]{} [$<2.0$]{} [P]{} [$0.013_{-0.006-0.002-0.001}^{+0.007+0.001+0.001}$]{} [$<0.33$]{} [P]{} [$0.010_{-0.004-0.002-0.001}^{+0.005+0.001+0.001}$]{} [$<1.2$]{} [P]{} [$0.012_{-0.002-0.004-0.001}^{+0.003+0.005+0.001}$]{} [$0.0189_{-0.0021}^{+0.0061}$]{} [$<0.2$]{} [T]{} [$13.5_{-3.9-0.7-1.0}^{+5.0+0.4+1.1}$]{} [$20.0_{-1.9-0.9}^{+4.0+2.0}$]{} [$17\pm2\pm1$]{} [$24.0\pm1.9$]{} [T]{} [$12.1_{-3.7-0.4-0.1}^{+4.5+0.1+0.1}$]{} [$16.9_{-1.6-0.9}^{+3.2+1.7}$]{} [$17\pm2\pm1$]{} [$15.9\pm2.1$]{} [P]{} [$9.9_{-3.3-2.4-0.5}^{+3.8+2.7+0.3}$]{} [$9.2_{-1.1-5.4}^{+1.2+3.6}$]{} [$17$]{} [$9.2\pm1.5$]{} [P]{} [$6.1_{-1.9-1.3-0.5}^{+2.5+1.3+0.3}$]{} [$5.5_{-0.5-2.5}^{+0.6+1.3}$]{} [$9.0$]{} [$4.6\pm1.1$]{} [P]{} [$4.0_{-1.3-0.9-0.3}^{+1.7+1.3+0.3}$]{} [$3.0_{-0.3-1.5}^{+0.4+2.5}$]{} [$7.9$]{} [$<7.4$]{} [P]{} [$0.56_{-0.19-0.12-0.02}^{+0.23+0.13+0.02}$]{} [$0.6_{-0.1-0.3}^{+0.1+0.3}$]{} [$0.40$]{} [$1.2\pm0.5$]{} [P]{} [$0.028_{-0.012-0.004-0.002}^{+0.015+0.003+0.002}$]{} [$<3.0$]{} : Updated branching ratios (in units of $10^{-6}$) of $B\rightarrow VV$ decays calculated in the PQCD approach. For comparison, we also give the updated theoretical predictions in the QCD factorization (QCDF) approach [@qcdfbtovv] and the previous predictions in the PQCD approach[@btovv]. Experimental data are from the Particle Data Group [@pdg] \[tb:br\] [ This work ]{} [QCDF]{} [PQCD(former)]{} [Exp.]{} -- ------------------------------------------------ -------------------------------------- ------------------ --------------------- [$0.12_{-0.02-0.01-0.00}^{+0.04+0.15+0.00}$]{} [$0.92_{-0.04-0.37}^{+0.03+0.06}$]{} [$0.60$]{} [$0.75\pm0.14$]{} [$0.95_{-0.01-0.01-0.00}^{+0.01+0.01+0.00}$]{} [$0.92_{-0.02-0.02}^{+0.01+0.01}$]{} [$0.94$]{} [$0.977\pm0.026$]{} [$0.67_{-0.06-0.04-0.06}^{+0.04+0.03+0.06}$]{} [$0.52_{-0.25-0.36}^{-0.11+0.50}$]{} [$0.87$]{} [$0.66_{-0.10-0.02-0.04}^{+0.07+0.04+0.06}$]{} [$0.94_{-0.01-0.20}^{+0.01+0.04}$]{} [$0.82$]{} [$0.65_{-0.03-0.04-0.00}^{+0.03+0.03+0.00}$]{} [$0.39_{-0.00-0.31}^{+0.00+0.60}$]{} [$0.74$]{} [$0.57\pm0.10$]{} [$0.68_{-0.03-0.03-0.02}^{+0.04+0.03+0.02}$]{} [$0.53_{-0.03-0.32}^{+0.02+0.45}$]{} [$0.78$]{} [$$]{} [$0.65_{-0.05-0.02-0.00}^{+0.05+0.02+0.00}$]{} [$0.58_{-0.10-0.14}^{+0.07+0.43}$]{} [$0.82$]{} [$0.69\pm0.13$]{} [$0.58_{-0.08-0.02-0.01}^{+0.07+0.02+0.02}$]{} [$0.52_{-0.07-0.48}^{+0.04+0.48}$]{} [$0.78$]{} [$0.80\pm0.13$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$0.99$]{} [$$]{} [$0.95_{-0.01-0.01-0.00}^{+0.01+0.01+0.00}$]{} [$$]{} [$$]{} [$$]{} [$0.94_{-0.02-0.02-0.00}^{+0.02+0.01+0.00}$]{} [$$]{} [$$]{} [$$]{} [$0.97_{-0.01-0.01-0.00}^{+0.01+0.01+0.00}$]{} [$$]{} [$0.65$]{} [$$]{} [$0.98_{-0.01-0.01-0.00}^{+0.01+0.01+0.00}$]{} [$0.96_{-0.01-0.02}^{+0.01+0.02}$]{} [$0.94$]{} [$0.95\pm0.016$]{} [$0.97_{-0.01-0.00-0.00}^{+0.01+0.00+0.00}$]{} [$0.96_{-0.01-0.03}^{+0.01+0.02}$]{} [$0.97$]{} [$0.90\pm0.06$]{} [$0.75_{-0.03-0.03-0.02}^{+0.03+0.02+0.02}$]{} [$0.67_{-0.03-0.48}^{+0.02+0.31}$]{} [$0.85$]{} [$0.78\pm0.12$]{} [$0.70_{-0.03-0.04-0.01}^{+0.03+0.04+0.00}$]{} [$0.48_{-0.04-0.40}^{+0.03+0.52}$]{} [$0.82$]{} [$0.48\pm0.08$]{} [$0.64_{-0.06-0.02-0.03}^{+0.06+0.02+0.04}$]{} [$0.67_{-0.04-0.39}^{+0.03+0.32}$]{} [$0.81$]{} [$0.41\pm0.19$]{} [$0.74_{-0.04-0.03-0.02}^{+0.03+0.02+0.01}$]{} [$0.45_{-0.04-0.38}^{+0.02+0.55}$]{} [$0.75$]{} [$0.75\pm0.25$]{} [$0.95_{-0.01-0.02-0.00}^{+0.01+0.01+0.00}$]{} [$$]{} [$$]{} [$$]{} : Updated percentage of the longitudinal polarizations $f_{L}$ of $B\rightarrow VV$ decays calculated in the PQCD approach compared with the updated theoretical predictions in the QCD factorization (QCDF) approach [@qcdfbtovv] and the previous predictions in the PQCD approach [@btovv]. Experimental data are from the Particle Data Group[@pdg]. \[tb:fl\] [ This work ]{} [QCDF]{} [Exp.]{} -- ------------------------------------------------ ----------------------------- ------------------- -- [$70.7_{-5.2-5.4-6.0}^{+2.9+0.8+3.8}$]{} [$30_{-16-26}^{+17+14}$]{} [$$]{} [$0.83_{-0.59-0.31-0.00}^{+0.50+0.66+0.00}$]{} [$-4_{-0-3}^{+0+3}$]{} [$$]{} [$59.4_{-8.3-5.5-6.3}^{+11.9+6.3+5.0}$]{} [$3_{-6-76}^{+2+51}$]{} [$$]{} [$-73.7_{-6.2-6.0-0.9}^{+6.7+2.6+3.3}$]{} [$-30_{-14-18}^{+15+16}$]{} [$$]{} [$-8.9_{-0.6-2.8-1.0}^{+0.6+2.8+1.1}$]{} [$-15_{-8-14}^{+4+16}$]{} [$-6\pm9\pm2$]{} [$24.5_{-1.5-3.4-0.6}^{+1.2+2.9+0.0}$]{} [$32_{-3-14}^{+1+2}$]{} [$21\pm15\pm2$]{} [$5.6_{-0.3-1.3-0.9}^{+0.3+1.2+0.8}$]{} [$23_{-5-18}^{+9+5}$]{} [$45\pm25$]{} [$0.0$]{} [$-14_{-1-2}^{+1+6}$]{} [$$]{} [$29.8_{-5.7-9.5-4.7}^{+2.0+6.4+4.6}$]{} [$0$]{} [$$]{} [$0.0$]{} [$$]{} [$$]{} [$0.0$]{} [$$]{} [$$]{} [$0.0$]{} [$$]{} [$$]{} [$0.0$]{} [$0.8_{-0-0.5}^{+0+0.4}$]{} [$$]{} [$0.05_{-0.01-0.03-0.00}^{-0.03+0.05+0.00}$]{} [$0.06$]{} [$-5\pm5$]{} [$-11.2_{-2.0-2.5-0.6}^{+1.8+2.4+0.9}$]{} [$-8_{-1-4}^{+1+3}$]{} [$-20\pm9$]{} [$22.7_{-1.5-2.5-1.2}^{+1.1+2.6+0.4}$]{} [$43_{-2-28}^{+6+12}$]{} [$31\pm13$]{} [$-1.0_{-0.3-0.0-0.2}^{+0.2+0.2+0.1}$]{} [$-0.3_{-0-0}^{+0+2}$]{} [$-1\pm16$]{} [$9.1_{-3.2-3.5-0.3}^{+3.3+1.3+0.0}$]{} [$56_{-4-43}^{+3+4}$]{} [$29\pm35$]{} [$23.0_{-4.2-2.2-1.4}^{+4.6+0.2+0.7}$]{} [$16_{-3-34}^{+1+17}$]{} [$$]{} [$-1.0$]{} [$0.05$]{} [$-1\pm8$]{} [$0.0$]{} [$$]{} [$$]{} : Direct $CP$ asymmetries (%) in the $B \to VV$ decays and comparison with the predictions from QCDF[@qcdfbtovv]. Experimental data are from the Particle Data Group[@pdg]. For $B^0\rightarrow K^{*0(+)}\rho^{0(-)}$, the data is from the ref.[@cpks+rho-] \[tb:cp\] [$f_{\perp}$ ]{} [$\phi_{\parallel}$]{} [$\phi_{\perp}$]{} [$A^{0}_{CP}$]{} [$A^{\perp}_{CP}$]{} [$\Delta\phi_{\parallel}$]{} [$\Delta\phi_{\perp}$]{} -- ---------------------------- ---------------------------- ---------------------------- ------------------------------- ----------------------------- ------------------------------ ----------------------------- [$45.9_{-8.2}^{+1.1}$]{} [$2.68_{-1.09}^{+1.90}$]{} [$2.81_{-1.95}^{+0.95}$]{} [$88.9_{-120.7}^{+9.0}$]{} [$-11.6_{-2.9}^{+16.2}$]{} [$-98.9_{-69.6}^{+251.9}$]{} [$-105_{-41}^{+266}$]{} [$2.42_{-0.19}^{+0.21}$]{} [$3.12_{-0.06}^{+0.06}$]{} [$3.16_{-0.05}^{+0.06}$]{} [$-2.05_{-0.55}^{+0.53}$]{} [$39.0_{-8.4}^{+7.6}$]{} [$10.2_{-3.1}^{+3.0}$]{} [$9.58_{-3.19}^{+2.93}$]{} [$16.7_{-3.6}^{+5.0}$]{} [$3.13_{-0.19}^{+0.17}$]{} [$3.13_{-0.19}^{+0.17}$]{} [$26.6_{-12.2}^{+19.8}$]{} [$-60.0_{-12.1}^{+11.8}$]{} [$-87.8_{-15.3}^{+13.7}$]{} [$-98.4_{-15.1}^{+12.9}$]{} [$18.2_{-5.3}^{+6.1}$]{} [$3.20_{-0.20}^{+0.25}$]{} [$3.21_{-0.22}^{+0.24}$]{} [$-5.70_{-16.2}^{+11.8}$]{} [$17.0_{-22.1}^{+19.1}$]{} [$105_{-10.4}^{+13.2}$]{} [$108_{-11.1}^{+13.8}$]{} [$16.9_{-1.8}^{+2.7}$]{} [$4.67_{-3.06}^{+0.02}$]{} [$4.66_{-3.06}^{+0.01}$]{} [$3.64_{-1.07}^{+1.20}$]{} [$-7.71_{-1.86}^{+1.97}$]{} [$-0.12_{-1.79}^{+1.72}$]{} [$0.22_{-1.65}^{+1.85}$]{} [$15.6_{-2.5}^{+2.5}$]{} [$3.31_{-0.21}^{+0.23}$]{} [$3.30_{-0.21}^{+0.22}$]{} [$23.8_{-5.1}^{+4.7}$]{} [$-50.9_{-3.9}^{+4.9}$]{} [$128_{-4.4}^{+4.1}$]{} [$127_{-4.3}^{+43}$]{} [$18.3_{-2.3}^{+2.6}$]{} [$2.18_{-0.20}^{+0.21}$]{} [$2.14_{-0.19}^{+0.21}$]{} [$1.46_{-1.62}^{+1.44}$]{} [$-8.92_{-4.01}^{+5.01}$]{} [$-2.28_{-1.89}^{+1.79}$]{} [$-12.0_{-4.9}^{+3.5}$]{} [$19.7_{-3.6}^{+4.0}$]{} [$2.26_{-0.16}^{+0.20}$]{} [$2.31_{-0.15}^{+0.19}$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$3.34_{-0.06}^{+0.08}$]{} [$3.37_{-0.09}^{+2.60}$]{} [$0.02_{-0.01}^{+0.02}$]{} [$-75.3_{-10.5}^{+21.1}$]{} [$56.4_{-9.7}^{+10.9}$]{} [$-129_{-2.0}^{+258}$]{} [$2.36_{-0.76}^{+1.08}$]{} [$3.76_{-0.31}^{+0.22}$]{} [$3.77_{-0.27}^{+0.24}$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$2.78_{-0.86}^{+1.08}$]{} [$3.77_{-0.28}^{+0.20}$]{} [$3.78_{-0.25}^{+0.20}$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$0.05_{-0.02}^{+0.02}$]{} [$3.26_{-0.14}^{+0.20}$]{} [$3.50_{-0.17}^{+0.17}$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$0.46_{-0.06}^{+0.08}$]{} [$3.20_{-0.09}^{+0.07}$]{} [$3.18_{-0.10}^{+0.07}$]{} [$0.002_{-0.003}^{+0.003}$]{} [$-0.32_{-0.64}^{+0.25}$]{} [$-0.11_{-0.32}^{+0.10}$]{} [$-0.79_{-0.45}^{+0.11}$]{} [$1.18_{-0.29}^{+0.38}$]{} [$2.64_{-0.15}^{+0.14}$]{} [$2.57_{-0.15}^{+0.16}$]{} [$-2.02_{-0.74}^{+0.69}$]{} [$76.2_{-14.7}^{+11.0}$]{} [$70.7_{-16.1}^{+16.8}$]{} [$83.9_{-19.7}^{+17.3}$]{} [$11.9_{-2.0}^{+2.3}$]{} [$1.94_{-0.14}^{+1.44}$]{} [$1.94_{-0.15}^{+1.43}$]{} [$11.3_{-2.4}^{+2.3}$]{} [$-34.0_{-2.8}^{+3.7}$]{} [$-26.4_{-4.0}^{+157}$]{} [$-27.3_{-4.0}^{+158}$]{} [$13.7_{-1.9}^{+2.1}$]{} [$1.81_{-0.18}^{+0.20}$]{} [$1.81_{-0.18}^{+0.19}$]{} [$-0.36_{-0.11}^{+0.12}$]{} [$0.98_{-0.25}^{+0.20}$]{} [$-1.19_{-0.36}^{+0.38}$]{} [$-1.54_{-0.49}^{+0.41}$]{} [$17.2_{-3.5}^{+3.4}$]{} [$2.18_{-0.20}^{+0.23}$]{} [$2.18_{-0.20}^{+0.22}$]{} [$11.2_{-4.3}^{+3.9}$]{} [$-19.9_{-3.6}^{+5.5}$]{} [$-37.9_{-6.1}^{+7.0}$]{} [$-38.7_{-6.1}^{+7.2}$]{} [$12.9_{-2.4}^{+1.7}$]{} [$1.98_{-0.17}^{+0.20}$]{} [$1.99_{-0.19}^{+0.18}$]{} [$7.21_{-2.50}^{+2.54}$]{} [$-19.1_{-2.6}^{+4.2}$]{} [$20.2_{-5.8}^{+4.8}$]{} [$28.4_{-6.2}^{-7.7}$]{} [$2.36_{-0.76}^{+1.08}$]{} [$3.76_{-0.31}^{+0.22}$]{} [$3.77_{-0.27}^{+0.23}$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} [$\sim0.0$]{} : Updated percentage of the transverse polarizations $f_{\perp}$(%), relative phases $\phi_{\parallel}(rad)$, $\phi_{\perp}(rad)$, $\Delta\phi_{\parallel}(10^{-2} rad)$, $\Delta\phi_{\perp}(10^{-2}rad)$ and the $CP$ asymmetry parameters $A^{0}_{CP}$(%) and $A^{\perp}_{CP}$(%) in $B\rightarrow VV$ decays calculated in the PQCD approach. \[tb:ft\] [Class]{} [This work]{} [ QCDF]{} [ PQCD(former)]{} [Exp]{} -- ----------- ------------------------------------------------------- --------------------------------------- ------------------------------------------------------- ------------ [T]{} [$24.0_{-8.7-1.4-2.4}^{+10.9+1.2+0.0}$]{} [$21.6_{-2.8-1.5}^{+1.3+0.9}$]{} [$20.9_{-6.2-1.4-1.1}^{+8.2+1.4+1.2}$]{} [$$]{} [C]{} [$0.40_{-0.15-0.07-0.03}^{+0.19+0.11+0.00}$]{} [$1.3_{-0.6-0.3}^{+2.0+1.7}$]{} [$0.33_{-0.07-0.09-0.01}^{+0.09+0.14+0.00}$]{} [$<767$]{} [C]{} [$0.35_{-0.14-0.08-0.08}^{+0.16+0.09+0.04}$]{} [$1.1_{-0.5-0.3}^{+1.5+1.3}$]{} [$0.31_{-0.07-0.06-0.02}^{+0.10+0.12+0.04}$]{} [$$]{} [P]{} [$5.4_{-1.7-1.4-0.5}^{+2.7+1.8+0.3}$]{} [$7.6_{-1.0-1.8}^{+1.0+2.3}$]{} [$6.7_{-1.2-1.4-0.2}^{+1.5+3.4+0.5}$]{} [$$]{} [P]{} [$0.23_{-0.05-0.01-0.02}^{+0.15+0.03+0.01}$]{} [$0.18_{-0.01-0.04}^{+0.01+0.09}$]{} [$0.23_{-0.07-0.01-0.01}^{+0.09+0.03+0.00}$]{} [$<617$]{} [P]{} [$0.17_{-0.07-0.04-0.01}^{+0.10+0.05+0.00}$]{} [$0.18_{-0.12-0.04}^{+0.44+0.47}$]{} [$0.16_{-0.05-0.04-0.00}^{+0.09+0.10+0.01}$]{} [$$]{} [P]{} [$1.5_{-0.6-0.2-0.1}^{+0.7+0.2+0.0}$]{} [$0.68_{-0.04-0.53}^{+0.04+0.73}$]{} [$1.0_{-0.2-0.2-0.0}^{+0.2+0.3+0.0}$]{} [$$]{} [P]{} [$0.74_{-0.24-0.14-0.00}^{+0.39+0.22+0.00}$]{} [$0.34_{-0.02-0.26}^{+0.02+0.36}$]{} [$0.51_{-0.11-0.10-0.01}^{+0.12+0.17+0.01}$]{} [$<320$]{} [E]{} [$0.009_{-0.003-0.002-0.001}^{+0.003+0.001+0.000}$]{} [$0.004_{-0.0-0.003}^{+0.0+0.005}$]{} [$0.007_{-0.001-0.001-0.000}^{+0.002+0.001+0.000}$]{} [$$]{} [P]{} [$0.40_{-0.18-0.10-0.01}^{+0.16+0.10+0.00}$]{} [$0.19_{-0.02-0.15}^{+0.02+0.21}$]{} [$0.39_{-0.08-0.07-0.00}^{+0.09+0.13+0.01}$]{} [$$]{} : Updated branching ratios (in units of $10^{-6}$) of $B_s\rightarrow VV$ decays calculated in the PQCD approach. For comparison, we also cite the updated theoretical predictions in the QCDF approach [@qcdfbsvv] and the previous predictions in the PQCD approach [@bsvv]. Experimental data are from the Particle Data Group [@pdg] \[tb:brbs\] [This work]{} [ QCDF]{} [ PQCD(former)]{} -- ------------------------------------------------ -------------------------------------- ------------------------------------------------------- [$0.95_{-0.01-0.01-0.00}^{+0.01+0.01+0.00}$]{} [$0.92_{-0.02-0.03}^{+0.01+0.01}$]{} [$0.937_{-0.002-0.003-0.002}^{+0.001+0.002+0.000}$]{} [$0.57_{-0.10-0.08-0.00}^{+0.06+0.06+0.01}$]{} [$0.90_{-0.05-0.23}^{+0.04+0.03}$]{} [$0.455_{-0.003-0.043-0.009}^{+0.004+0.069+0.006}$]{} [$0.50_{-0.08-0.15-0.01}^{+0.07+0.11+0.01}$]{} [$0.90_{-0.04-0.23}^{+0.03+0.03}$]{} [$0.532_{-0.002-0.029-0.013}^{+0.003+0.035+0.023}$]{} [$0.42_{-0.09-0.03-0.06}^{+0.13+0.03+0.05}$]{} [$0.52_{-0.05-0.21}^{+0.03+0.20}$]{} [$0.438_{-0.040-0.023-0.015}^{+0.051+0.021+0.037}$]{} [$0.86_{-0.01-0.01-0.00}^{+0.01+0.01+0.00}$]{} [$0.88_{-0.00-0.18}^{+0.01+0.02}$]{} [$0.870_{-0.002-0.003-0.004}^{+0.002+0.009+0.009}$]{} [$0.69_{-0.09-0.09-0.02}^{+0.08+0.08+0.02}$]{} [$0.95_{-0.02-0.42}^{+0.01+0.00}$]{} [$0.443_{-0.075-0.061-0.004}^{+0.000+0.054+0.009}$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} [$\sim1.0$]{} : Percentage of the longitudinal polarizations $f_{L}$ in $B_s \to VV $ decays and comparison with the QCDF approach [@qcdfbsvv] and the previous predictions in the PQCD approach [@bsvv]. \[tb:flbs\] [This work]{} [ QCDF]{} [ PQCD(former)]{} -- -------------------------------------------- --------------------------------- --------------------------------------------- [$-9.1_{-1.5-1.2-0.3}^{+1.4+1.0+0.2}$]{} [$-11_{-1-1}^{+1+4}$]{} [$-8.2_{-1.2-1.7-1.1}^{+1.0+1.2+0.4}$]{} [$62.7_{-5.9-16.0-7.9}^{+6.4+10.5+7.5}$]{} [$46_{-17-25}^{+15+10}$]{} [$61.8_{-4.7-22.8-2.3}^{+3.2+17.1+4.4}$]{} [$-78.1_{-2.2-7.4-8.3}^{+2.9+13.1+8.1}$]{} [$-50_{-15-6}^{+20+21}$]{} [$-62.1_{-3.9-12.6-1.9}^{+4.8+19.7+5.5}$]{} [$8.8_{-8.9-2.9-0.2}^{+2.5+0.5+0.0}$]{} [$21_{-2-4}^{+1+2}$]{} [$9.3_{-0.7-3.6-0.2}^{+0.4+3.3+0.3}$]{} [$-4.3_{-0.5-0.5-1.0}^{+0.6+0.6+1.2}$]{} [$83_{-0.0-36}^{+1.0+10}$]{} [$10.1_{-0.9-1.8-0.5}^{+0.9+1.6+1.3}$]{} [$28.0_{-3.2-2.3-5.1}^{+1.3+0.5+3.4}$]{} [$-8_{-1-15}^{+3+20}$]{} [$3.6_{-0.6-2.4-0.2}^{+0.6+2.4+0.6}$]{} [$0.0$]{} [$0.2_{-0.3-0.2}^{+0.4+0.5}$]{} [$0.0$]{} [$-2.9_{-1.1-1.3-0.2}^{+0.7+1.5+0.2}$]{} [$0$]{} [$-2.1_{-0.1-1.3-0.1}^{+0.2+1.7+0.1}$]{} [$-2.9_{-1.1-1.3-0.2}^{+0.7+1.5+0.2}$]{} [$0$]{} [$-2.1_{-0.1-1.3-0.1}^{+0.2+1.7+0.1}$]{} [$11.1_{-1.5-4.4-1.4}^{+1.0+1.9+1.2}$]{} [$0$]{} [$6.0_{-0.5-3.9-0.4}^{+0.7+2.7+1.0}$]{} [$-3.3_{-1.0-1.4-0.2}^{+0.8+1.5+0.5}$]{} [$0$]{} [$-2.0_{-0.1-1.3-0.1}^{+0.1+1.7+0.1}$]{} : Direct $CP$ asymmetries (%) in the $B_s\to VV$ decays and comparison with the QCDF approach [@qcdfbsvv] and the previous predictions in the PQCD approach [@bsvv]. \[tb:cpbs\] [$f_{\perp}$]{} [$\phi_{\parallel}$]{} [$\phi_{\perp}$]{} [$A^{0}_{CP}$]{} [$A^{\perp}_{CP}$]{} [$\Delta\phi_{\parallel}$]{} [$\Delta\phi_{\perp}$]{} -- ---------------------------- ---------------------------- ---------------------------- ------------------------------- ------------------------------ ------------------------------ ----------------------------- [$2.31_{-0.21}^{+0.22}$]{} [$3.07_{-0.09}^{+0.07}$]{} [$3.07_{-0.08}^{+0.08}$]{} [$-2.71_{-0.72}^{+0.68}$]{} [$55.0_{-10.5}^{+10.3}$]{} [$12.4_{-4.7}^{+4.8}$]{} [$12.5_{-4.8}^{+4.5}$]{} [$22.5_{-4.7}^{+7.3}$]{} [$1.94_{-0.10}^{+2.52}$]{} [$1.99_{-0.10}^{+2.53}$]{} [$-17.5_{-13.0}^{+21.2}$]{} [$22.0_{-31.4}^{+29.9}$]{} [$-31.5_{-16.2}^{+274}$]{} [$-36.5_{-15.8}^{+222}$]{} [$26.1_{-7.0}^{+9.8}$]{} [$2.18_{-0.28}^{+0.33}$]{} [$2.23_{-0.27}^{+0.32}$]{} [$-5.99_{-50.21}^{+23.52}$]{} [$6.95_{-32.14}^{+27.91}$]{} [$30.7_{-24.3}^{+30.9}$]{} [$36.5_{-24.2}^{+31.3}$]{} [$27.7_{-7.0}^{+5.2}$]{} [$3.53_{-0.25}^{+0.33}$]{} [$3.54_{-0.24}^{+0.36}$]{} [$45.4_{-23.4}^{+19.0}$]{} [$-32.9_{-4.0}^{+5.6}$]{} [$93.7_{-14.1}^{+11.1}$]{} [$93.4_{-13.8}^{+11.1}$]{} [$8.89_{-1.06}^{+0.80}$]{} [$3.11_{-0.09}^{+0.10}$]{} [$3.29_{-0.09}^{+0.09}$]{} [$3.27_{-1.19}^{+1.07}$]{} [$-32.8_{-5.8}^{+7.4}$]{} [$-43.7_{-9.5}^{+9.9}$]{} [$-63.9_{-9.6}^{+10.8}$]{} [$16.1_{-5.8}^{+7.3}$]{} [$3.38_{-0.17}^{+0.20}$]{} [$3.35_{-0.23}^{+0.30}$]{} [$-2.24_{-5.45}^{+6.67}$]{} [$4.38_{-15.93}^{+17.52}$]{} [$-36.7_{-11.9}^{+12.5}$]{} [$-32.7_{-18.6}^{+16.5}$]{} [$\sim0.0$]{} [$3.40_{-0.04}^{+0.04}$]{} [$3.27_{-0.15}^{+0.16}$]{} [$0.0$]{} [$30.5_{-16.3}^{+15.0}$]{} [$2.87_{-0.59}^{+0.44}$]{} [$-27.4_{-6.0}^{+6.9}$]{} [$\sim0.0$]{} [$3.40_{-0.04}^{+0.04}$]{} [$3.27_{-0.15}^{+0.16}$]{} [$0.0$]{} [$30.5_{-16.3}^{+15.0}$]{} [$2.87_{-0.59}^{+0.44}$]{} [$-27.4_{-6.0}^{+6.9}$]{} [$\sim0.0$]{} [$3.48_{-0.05}^{+0.04}$]{} [$2.63_{-0.22}^{+0.18}$]{} [$0.0$]{} [$27.9_{-9.9}^{+9.3}$]{} [$-9.30_{-5.23}^{+1.50}$]{} [$-30.4_{-23.4}^{+19.1}$]{} [$\sim0.0$]{} [$3.40_{-0.04}^{+0.04}$]{} [$3.27_{-0.11}^{+0.16}$]{} [$0.0$]{} [$30.8_{-15.3}^{+14.0}$]{} [$2.71_{-0.52}^{+0.42}$]{} [$-26.7_{-5.7}^{+6.3}$]{} : Updated percentage of the transverse polarizations $f_{\perp}$(%), relative phases $\phi_{\parallel}({\rm rad})$, $\phi_{\perp}({\rm rad})$, $\Delta\phi_{\parallel}(10^{-2} {\rm rad})$, $\Delta\phi_{\perp}(10^{-2} {\rm rad})$ and the $CP$ asymmetry parameters $A^{0}_{CP}$ and $A^{\perp}_{CP}$ in $B_s\rightarrow VV$ decays calculated in the PQCD approach. \[tb:ftbs\] For the charmless $B_{q} \to VV$ decays, it is naively expected that the helicity amplitudes $H_{i}$ (with helicity $i=0,-,+$) satisfy the hierarchy pattern $$\begin{aligned} H_0:H_-:H_+=1: \frac{\Lambda_{QCD}}{m_b}:(\frac{\Lambda_{QCD}}{m_b})^2. \label{eq:hierarchy}\end{aligned}$$ In the naive factorization approach, longitudinal polarizations dominate the branching ratios of $B$ decays. In sharp contrast to these expectations, large transverse polarization of order 50$\%$ is observed in $B\to K^*\phi$, $B\to K^*\rho$ and $B_s\to \phi\phi$ decays, which poses an interesting challenge for the theory. This shows that the scaling behavior shown in Eq. (\[eq:hierarchy\]) is violated. In order to interpret this large transverse polarization many mechanisms have been proposed, such as the penguin-induced annihilation contributions [@flpenguin], final-state interactions [@flfsi], form-factor tuning [@flfft], and even onset of new physics [@flnp]. As pointed out in the context of QCDF [@qcdf3], after taking into account the NLO effects, e.g., vertex-, penguin- and hard spectator-scattering contributions, the effective Wilson coefficients $a_i^h$ become helicity dependent. Including these effects, for some penguin-dominant modes, the constructive (destructive) interference in the transverse (longitudinal) amplitudes of $B \to VV$ decays makes the total transverse contribution comparable to the longitudinal one, and the transverse polarization fraction may reach as high as $50\%$. In order to interpret the observed large transverse polarization fraction in the penguin-dominated $B\to VV$ decays, e.g., $B\to K^*\phi$, $B\to K^*\rho$, both the PQCD and the QCDF frameworks rely on penguin annihilation. However, in QCDF, the penguin-annihilation amplitude involves a troublesome endpoint divergence, which is fudged by introducing non-perturbative parameters. Hence, in QCDF, one can fit the existing experimental data on the branching ratios, $f_L$ and the $CP$ asymmetries by adjusting the annihilation parameters $\rho_A$ and $\phi_A$, which reduces the predictive power of the theory. In contrast, in the PQCD approach, the annihilation type diagrams can be perturbatively calculated without introducing any fudge factor (or parameter), which allows us to predict the direct $CP$ asymmetry and transverse polarization. The large transverse polarization fraction can be interpreted on the basis of the chirally enhanced annihilation diagrams, especially the $(S-P)(S+P)$ penguin annihilation, introduced by the QCD penguin operator $O_6$ [@flpqcd]. A nice feature of the $(S-P)(S+P)$ penguin annihilation operator is that the light quarks in the final states are not produced through chiral currents. So, there is no suppression caused by the helicity flip. As a result, the polarization fractions satisfy $$\begin{aligned} f_L\approx f_{\parallel} \approx f_{\perp}.\end{aligned}$$ Thus, in the PQCD approach, the penguin-annihilation together with the hard-scattering emission diagrams can explain the large transverse polarization fraction measured in experiments. We present our numerical results for the branching ratios, direct $CP$ asymmetries, and some other observables introduced earlier in the text, in Tables \[tb:zui\]-\[tb:ftbs\]. The dominant topologies contributing to these decays are also indicated in the tables through the symbols $T$ (the color-allowed tree contributions), $C$ (the color-suppressed tree contributions), $P$ (penguin contributions), and $E$ ($W$-exchange annihilation contributions). Theoretical uncertainties quoted in the tables are estimated from three sources: the first error quoted is from the input hadronic parameters, such as the decay constants of the initial $B_q$ and the final vector-mesons and the parameters in the distribution amplitudes of the initial and final states, which can be found in sec. \[sec:function\] and Eq. (\[para\]). The second error arises from the scale uncertainties, characterized by $\Lambda_{QCD}=(0.25\pm0.05)$ GeV and the variations of the factorization scales $t$ (from $0.8t$ to $1.2t$) detailed in Appendix \[app:a\]. The scale-dependent uncertainty can be reduced only if the next-to-leading order contributions are known. The last error is the combined uncertainty in the CKM matrix elements and the angles of the unitary triangle. In Tables \[tb:zui\],\[tb:ft\] and \[tb:ftbs\], we have combined these uncertainties by adding them in quadrature and show the resulting uncertainty, due to to the space limitations in the Tables. We now discuss these results. For the branching ratios, the most important theoretical uncertainty is the first error caused by the nonperturbative input parameters. In the PQCD approach, the wave functions are the primary important input parameters and they heavily influence the predictions of the branching ratios, as also discussed in [@xiaowang]. We have adopted the new updated wave functions. While, for the direct $CP$ asymmetry parameters, the dominant uncertainty arises from the second error, which is caused by the unknown higher order QCD corrections. From the definition: $$\begin{aligned} A_{CP}^{dir}&\equiv&\frac{BR(\bar{B}\to f)-BR(B\to \bar{f})}{BR(\bar{B}\to f)+BR(B\to \bar{f})}\nonumber\\ &=&\frac{\mid A(\bar{B}\to f)\mid^2-\mid A(B\to \bar{f})\mid^2}{\mid A(\bar{B}\to f)\mid^2+\mid A(B\to \bar{f})\mid^2},\end{aligned}$$ it is apparent that the wave functions of the initial $B_{q}$ meson and the final vector mesons are overall factors, hence they drop out in the ratio and do not provide significant contributions in the estimates of the direct $CP$ asymmetries. Direct $CP$ asymmetries are proportional to the strong phases originated from the hard part, and the NLO QCD corrections will influence the strong phases significantly. Not having these corrections at our disposal, we can only estimate them by varying the scales. The resulting theoretical uncertainty is larger than the one from the wave functions, and we assume that the variation of the scales is an adequate account of the NLO corrections at this stage. For comparison, the updated results of the QCDF approach[@qcdfbtovv; @qcdfbsvv] and the earlier PQCD predictions[@bsvv; @btovv; @jpg32] are also presented. We have updated the PQCD computations in this work and the main improvements are: (i) Use of the updated vector mesons distribution amplitudes with new estimates of the Gegenbauer moments and decay constants, and (ii) the treatment of the terms in the decay amplitude proportional to the ratio $r_{i}^2=m_{V_i}^2/m_B^2(i=2,3)$. Since wave functions are the most important inputs in PQCD, their improved knowledge is expected to yield improved estimates of the branching ratios, polarization fractions, and other observables. We recall that in the earlier PQCD computations, $r_{i}^2$-dependent terms in the denominator of the propagators of the virtual quarks and gluons were omitted. From Appendix A, we find that, although their contribution is formally power suppressed, it can numerically change the real and imaginary parts of amplitudes and enhance the transverse polarization component, especially for the penguin-dominant decays. To quantify this, we have listed the amplitudes, branching ratios and transverse polarization fractions of the penguin-dominant decays $B^0\to K^{*0}\phi$, $B_s \to \phi\phi$ and the tree-dominant decay $B^+\to \rho^+\rho^0$ with and without the $r_i^2$-terms in Table\[tb:amp\]. We note that for the two penguin-dominant decays, the impact of the $r_i^2$-dependent terms in the amplitudes of the annihilation part, as well as in the imaginary part of the emission diagrams, is numerically significant. Taking the factorizable annihilation diagrams as an example, in the range near $x_3 \to 1$ or $x_2 \to 0$, the nonzero $r_i^2$ contributes a non-negligible imaginary part. So by keeping the $r_i^2$-terms, the branching ratios are reduced, while the transverse polarization fractions rise. The two main improvements go in the right direction in explaining the observed branching ratios and the large transverse polarization fractions in $B \to K^*\phi$ and $B_s\to \phi\phi$ decays in the PQCD approach. For the tree-dominant decay $B^+\to \rho^+\rho^0$, however, the effect on the traditional emission diagrams produced by the $r_i^2$-terms is tiny, as expected. This is further discussed in Appendix A. Thus, the improved PQCD treatment presented here yields better consistency with the data. Modes $A^{L}$ $A^{N}$ $A^T$ Br $f_{L}$ ------------------------------------------ -------------- -------------- ---------------- --------------- ---------- ---------- [$B^0\rightarrow K^{*0}\phi(r_i^2)$]{} emission -3.3+0.67$i$ -0.66+0.06$i$ 0.64-0.05$i$ [9.8]{} [56]{} annihilation 0.32-1.6$i$ -0.43+0.84$i$ 0.42-0.83$i$ [$B^0\rightarrow K^{*0}\phi$]{} emission -3.0-0.09$i$ -0.71-0.012$i$ 0.69+0.03$i$ [15]{} [70]{} annihilation -0.42-1.95i 0.05+1.28$i$ -0.11-1.38$i$ [$B_s\rightarrow \phi\phi(r_i^2)$]{} emission -2.8+0.37$i$ -0.60+0.10$i$ 0.60-0.08$i$ [16.7]{} [34.7]{} annihilation 0.68-1.2$i$ -0.53+1.0$i$ 0.53-1.0$i$ [$B_s\rightarrow \phi\phi$]{} emission -2.6-0.02$i$ -0.64+0.03$i$ 0.63-0.005$i$ [26.6]{} [45]{} annihilation -0.04-1.8$i$ 0.18+1.8$i$ -0.15-1.7$i$ [$B^+\rightarrow \rho^+\rho^0(r_i^2)$]{} emission 3.0+5.9$i$ 0.28+0.33$i$ 0.27-0.29$i$ [13.5]{} [98]{} annihilation $\sim0$ $\sim0$ $\sim0$ [$B^+\rightarrow \rho^+\rho^0$]{} emission 2.8+5.8$i$ 0.12+0.33$i$ -0.11-0.29$i$ [13.3]{} [99]{} annihilation $\sim0$ $\sim0$ $\sim0$ : Amplitudes ($10^{-3}$), branching ratios ($10^{-6}$) and the polarization fractions ($\%$) with (and without) the $r_i^2$-dependent terms in the $B^0\to K^{*0}\phi$, $B_s \to \phi\phi$ and $B^0\to \rho^+\rho^0$ decays. \[tb:amp\] In Table \[tb:zui\], we list the current experimental measurements in $B^0(B^+) \to K^* (K^{*+}) \phi$ and $B_s \to \overline K^{*0}\phi$, $B_s \to \overline K^{*0}K^{*0}$ and $B_s \to \phi \phi$ decays and compare them with our theoretical results worked out in this paper. These decays are all penguin-dominated and are measured with large fraction of transverse polarization. For these decays, the naive factorization approach predicts too small branching ratios by a factor of $2\sim 3$ [@qcdfbtovv], due to the small contribution from the penguin operators. In [@btovv; @bsvv; @jpg32; @flfft], the authors have studied these $B_q\to VV$ decays, but those predictions are not in good agreement with the currently available experimental data. The primary task is to bring up the branching ratios and explain the polarization anomaly in these decays. In our update, we explain the bulk of the data. However, we note that for $B_s \to K^{*0}\overline{K}^{*0}$ and $B_s \to \overline{K}^{*0}\phi$ modes, our calculated branching ratios are $(5.4_{-2.4}^{+3.0})\times 10^{-6}$ and $(0.39_{-0.19}^{+0.20})\times 10^{-6}$ respectively, which are much smaller than the data, though they are compatible with the QCDF predictions $(6.6_{-1.4-1.7}^{+1.1+1.9})\times 10^{-6}$ and $(0.37_{-0.05-0.20}^{+0.06+0.24})\times10^{-6}$ respectively. In Table \[tb:br\], we have given our estimates of the $B \to VV$ branching ratios for different topologies. For the penguin dominant decay modes (indicated by $P$ in the tables), our updated predictions basically agree with the QCDF predictions, except for $B^0\to K^{*0}\omega$. Due to the constructive interference between the penguin emission contributions and the penguin annihilation contributions, our prediction for this decays is almost twice as large as that of QCDF, and it also comes out larger than the current experimental data. As the experimental error is still large, we wait for consolidated date from Belle-II experiment. For the color-suppressed decay $B^0 \to\rho^0\rho^0$, the calculated branching fraction in this work is $(0.27_{-0.09-0.04-0.01}^{+0.10+0.06+0.00})\times 10^{-6}$, while BaBar and Belle obtained $(0.9\pm 0.32\pm0.14)\times10^{-6}$ [@babar] and $(0.4\pm0.4_{-0.3}^{+0.2})\times10^{-6}$ [@belle], respectively, with the current world average being $(0.73\pm0.28)\times 10^{-6}$. Our result, within errors, agrees with the Belle data. Judged from the isospin triangle, since the decay rate of $B^0\to \rho^0\rho^0$ is so small, the rate for the decay $B^0 \to \rho^+\rho^-$ ought to be double that of $B^+\to \rho^+\rho^0$. In experiment, however, within errors, these two rates are equal to each other, which is puzzling. Thus, the experimental situation is still in a state of flux. In Table \[tb:fl\], discussed in more detail later, we show that for the $B^0\to \rho^0\rho^0$ decay the longitudinal polarization fraction is as small as $12\%$. As is well known, for the color-suppressed decays, the longitudinal polarization contributions from two hard-scattering emission diagrams largely cancel against each other. What’s worse, the remaining longitudinal polarization contributions are nearly canceled by those from the annihilation diagrams. On the other hand, the chiral enhanced annihilation diagrams and hard-scattering emission diagrams provide a large transverse polarizable contribution. In the end, the $B^0\to \rho^0\rho^0$ is almost totally dominated by the transverse polarization component. In Table \[tb:fl\], we adopt the BaBar data [@babar:rhorho], but note that Belle has provided a new measurement $0.21^{+0.18}_{-0.22}\pm{0.13}$ [@belle:rhorho], which supports our theoretical calculations. Thus, it is important to have a refined measurement of the branching fractions and the longitudinal polarization fractions for $B\to \rho\rho$ to draw definitive conclusion. It should be noted that if the next leading order corrections are included, the branching fraction of $B\to \rho^0\rho^0$ might be enlarged while its transverse polarization fraction $f_\perp$ will become smaller [@Li:2006cva]. The previous PQCD estimates for the $B^0 \to \rho^0 \omega$ decay rate exceeded the current experimental upper bound. In this work, this branching ratio is now lower than the upper experimental bound but is about a factor five larger than the QCDF prediction due to a near cancelation of the color-suppressed tree amplitudes. In the framework of PQCD, although the color-suppressed tree amplitudes also almost mutually cancel, the decay can get significant contributions from the annihilation type diagrams so that the decay rate comes back up and is not as small as in the QCDF prediction. This, together with some other predictions, provides an experimental check on these two competing frameworks. In Table. \[tb:fl\], we have given the fraction of the longitudinal polarization component, $f_L$ for $B \to VV$ decays, where we have compared them with the available data, and also with the previous PQCD [@btovv] and QCDF [@qcdfbtovv] approaches. Of these, the predictions for the decays $B\to \phi\rho(\omega)$ are worked out for the first time. For the $B^0\to \rho^0\omega$ decay, we predict the longitudinal polarization fraction as small as $67\%$, which is due to a significant transverse polarization component, $f_\perp$, from the penguin annihilation diagrams. The $f_L$ for this decay is in agreement with the QCDF prediction [@qcdfbtovv] but is significantly less than the previous PQCD prediction ($87\%$). From Table. \[tb:fl\], one also sees that for $B^0\to K^{*0}\omega$, our estimate for the longitudinal polarization fraction is in excellent agreement with the experimental data. We note that, for $B^0 \to \omega \omega$, our predicted longitudinal fraction is $66\%$, while the QCDF approach yields $94\%$ [@qcdfbtovv], where the longitudinal contributions highly dominate the amplitude. In QCDF, as in $B^0\to \rho^0\omega$, the penguin annihilation contributions are also tiny in $B^0 \to \omega\omega$. In PQCD, together with the hard scattering contributions, the considerable penguin annihilation contributions yield a different result. For the $B^0\to \phi\phi$, the previous PQCD prediction of the longitudinal polarization fraction is $65\%$ [@btovv], while our updated longitudinal polarization fraction is given by $f_L\sim 1$, which is confirmed also in [@YL]. For the $B^+\to K^{*0}\rho^+$ mode, our result is larger than the data, while in the QCDF framework, the central value is the same as the data, as this mode is used to extract the input parameters [@qcdfbtovv]. In this paragraph we shall discuss direct $CP$-asymmetries in the decays $B \to VV$ shown in Table \[tb:cp\] and their current measurements. Though none of the current experimental measurements for the $CP$ asymmetries shown in Table \[tb:cp\] is conclusive, they are in accord with our theoretical calculations. This, in turn, implies that the dominant strong phases in these channel estimated in our approach are in the right ball-park. From Table. \[tb:cp\], one also notes that the $CP$ asymmetries are large for the penguin dominant decays, but they are small for the color allowed tree-dominant decays and almost pure penguin-dominant processes, such as $K^{*0}\rho^+$ and $K^*\phi$. For $B^0\to \rho^0\omega$ decays, our prediction is about 60$\%$, while that of QCDF is only $3\%$. In PQCD, since the emission diagrams nearly cancel each other, the annihilation diagrams provide the dominant contributions. As direct $CP$ asymmetry is proportional to the interference between the tree and penguin contributions, the sizable interference makes the $CP$ asymmetry parameter large, reaching $60\%$. For the $B^0\to \rho^0\rho^0/\omega\omega$ modes, the large penguin contributions from the chirally enhanced annihilation diagrams, which are at the same level as the tree contributions from the emission diagrams, make the the $CP$ asymmetry parameter as large as $70\%$. On the other hand, for pure annihilation type decay $B^0 \to K^{*0}\bar{K}^{*0}$, since there are no contributions from tree operators, it is natural to expect that the direct $CP$ asymmetry is practically zero. In summary, the entries in Tables \[tb:br\], \[tb:fl\] and \[tb:cp\] show that for these $B \to VV$ decays our updated predictions are in good agreement with experiment, and, broadly speaking, are also in agreement with the QCDF predictions [@qcdfbtovv]. In Table \[tb:ft\], we give the predictions for the perpendicular polarization fraction, $f_\perp$, the relative phases, $\phi_{\parallel}({\rm rad})$, $\phi_{\perp}({\rm rad})$, $\Delta\phi_{\parallel}(10^{-2}~{\rm rad})$, $\Delta\phi_{\perp}(10^{-2}~{\rm rad})$, and the $CP$ asymmetry parameters $A^{0}_{CP}$ and $A^{\perp}_{CP}$ for the $B \to VV$ decays for the first time in the PQCD framework. These remain to be confronted with the data. In fact, these variables are already experimentally measured in five channels: $B^0(B^+) \to K^* (K^{*+}) \phi$ and $B_s \to \overline K^{*0}\phi$, $B_s \to \overline K^{*0}K^{*0}$ and $B_s \to \phi \phi$, which are shown in Table \[tb:zui\]. Our results are in good agreement with the data. We now discuss the results for the $B_s\to VV$ decays. Since the initial and the final state distribution amplitudes (DAs) are the most important input parameters in the PQCD approach, our predictions for $B_s\to VV$ decays in Tables \[tb:brbs\], \[tb:flbs\], \[tb:cpbs\], \[tb:ftbs\] are almost the same as the predictions in [@bsvv], as the DAs we adopted here are similar to those used in [@bsvv], except for the DAs of the $\phi$ meson. For the $B_s\to \phi\phi$ decay, the central values of the branching ratio and the longitudinal polarization fraction estimated in [@bsvv] are $35.3\times 10^{-6}$ and 61.9$\%$, respectively. It is apparent that neither the branching ratio nor the polarization fraction are in conformity with the experimental data, posted as $(19\pm5)\times 10^{-6}$ and $(34.8\pm4.6)\%$, respectively. With the updated DAs of $\phi$ meson, the current predictions of all the observables listed in Table \[tb:zui\] agree better with the data. This can be confirmed by the similar updates in $B_s\to \pi^+\pi^-$ and $B^0\to K^+K^-$ decays[@xiaowang]. Also, due to the terms proportional to the ratio $r_i^2=m_{\phi}^2/M_{B_s}^2$ in the denominators, which we keep, their influence is expected to be more pronounced, as $m_{\phi}$ is larger than the other light vector-meson masses. For the rest of the decay modes, the numerical values of the polarization fraction are basically consistent with the former PQCD predictions  [@bsvv]. From Tables \[tb:brbs\] and \[tb:flbs\], we note that for the color-suppressed decays $B_s\to \overline{K}^{*0}\rho(\omega)$, the branching ratios in PQCD are smaller than in QCDF by a factor of 3 due to the near cancellation of the hard scattering contributions. On the other hand, chirally enhanced annihilation and the hard scattering diagrams enhance the transverse polarization contribution, making it comparable to the longitudinal polarization one. For $B_s\to \omega\phi$, the pure emission mode, the $(S-P)(S+P)$ densities in the hard scattering diagrams also contribute a sizable transverse polarization component. For $B_s\to K^{*+}K^{*-}$, due to the large transverse polarization contribution from chirally enhanced annihilation diagrams, the longitudinal polarization fraction is as small as $40\%$, which is similar to $B_s\to K^{*0}\overline{K}^{*0}$. We also emphasize the measurements of the modes $B_s\to \overline{K}^{*0}\rho(\omega)$ and $B_s\to \omega \phi$ to distinguish among the competing dynamical models in the interpretation of the polarization anomaly. Direct $CP$ asymmetries of $B_s\to VV$ decays are listed in Table \[tb:cpbs\]. We note that they are small for the penguin-dominant processes, since the interference between tree and penguin contributions due to the former are too small, which is opposite to the tree-dominant process $B_s\to K^{*-}\rho^+$, which also has a small direct $CP$ asymmetry. For $B_s \to \rho^{0}\phi$, QCDF predicts about $83\%$ for direct $CP$ asymmetry with large charming penguin contributions. But in the framework of PQCD, it is only $-4.3\%$ because this mode belongs to the pure emission-type processes. Hence, measurement of direct CP asymmetry in this mode will help us to distinguish the PQCD and the QCDF approaches. As is well known, SU(3) symmetry relates a number of $B_s\to VV$ and $B_{u,d}\to VV$ processes, such as $B_s\to K^{*-}\rho^+$ and $B^0\to \rho^+\rho^- $. In the PQCD approach, presented here, this relation is well satisfied: $$\begin{aligned} &&\mathcal{B}(B_s\to K^{*-}\rho^+)=(24.0_{-8.7-1.4-2.4}^{+10.9+1.2+0.0})\sim \mathcal{B}(B^0\to \rho^+\rho^-)=(26.0_{-8.1-1.4-1.2}^{+10.1+1.4+1.5}),\end{aligned}$$ in units of $10^{-6}$. On the other hand, SU(3)-breaking in the decay rates for $B\to K^*\phi$ and $B_s\to \phi\phi$ is significant, as can be seen in Table \[tb:zui\]. In the PQCD approach, the SU(3)-breaking effects are caused by the differences between the initial and final state wave functions, such as the shape parameter $\omega_{B}$ and $\omega_{B_s}$, as well as the decay constants of the $B$ and $B_s$ mesons, along with the Gegenbauer moments and the decay constants of the final vector mesons. They conspire to yield a cumulative 60% SU(3)-breaking effect. Other SU(3)-breaking effects lie in between these two cases, as can be numerically calculated from the entries in various tables presented here. $U$-spin symmetry, relating a number of $B_{(s)} \to h_1 h_2$ ($h_i$ are light mesons) has been advocated in the literature  [@rf]. For $B_{(s)} \to VV$ decays, it has been studied in [@qcdfbsvv] and checked against the explicit QCDF estimates, and seems to hold well. Since we have calculated the $B$ and $B_s$ decays to $VV$ in this work in the PQCD approach, we also check the $U$-spin symmetry in some representative decays studied in [@qcdfbsvv]: $$\begin{aligned} &&A_{CP}(B_s\to K^{*-}\rho^+)=-A_{CP}(B^0\to K^{*+}\rho^-)\frac{\mathcal{B}(B^0\to K^{*+}\rho^-)}{\mathcal{B}(B_s\to K^{*-}\rho^+)}\frac{\tau(B_s)}{\tau(B)},\nonumber\\ &&A_{CP}(B_s\to \bar{K}^{*0}\rho^0)=-A_{CP}(B^0\to K^{*0}\rho^0)\frac{\mathcal{B}(B^0\to K^{*0}\rho^0)}{\mathcal{B}(B_s\to \bar{K}^{*0}\rho^0)}\frac{\tau(B_s)}{\tau(B)},\nonumber\\ &&A_{CP}(B_s\to \rho^+\rho^-)=-A_{CP}(B^0\to K^{*+}K^{*-})\frac{\mathcal{B}(B^0\to K^{*+}K^{*-})}{\mathcal{B}(B_s\to \rho^+\rho^-)}\frac{\tau(B_s)}{\tau(B)},\nonumber\\ &&A_{CP}(B_s\to K^{*+}K^{*-})=-A_{CP}(B^0\to \rho^+\rho^-)\frac{\mathcal{B}(B^0\to \rho^+\rho^-)}{\mathcal{B}(B_s\to K^{*+}K^{*-})}\frac{\tau(B_s)}{\tau(B)}.\end{aligned}$$ Using these $U$-spin relations as well as the branching ratios, the lifetimes of $B$ and $B_s$ mesons and the direct $CP$ asymmetries in $B$ decays, we can get the relevant direct $CP$ asymmetries in $B_s$ decays. This can be then compared with the explicit calculations in the PQCD approach to check whether the $U$-spin symmetry works well or not. We show this comparison in Table \[tb:u-spin\], where the entries in the last two columns have to be compared with each other. We find that, within the calculational errors, the $U$-spin symmetry works well in the direct CP asymmetries in the PQCD approach as well. modes Br $A_{CP}$($\%$) modes Br $A_{CP}$($\%$)($U$) $A_{CP}$(PQCD) ----------------------- ------ -------------------------------------------- ----------------------------- ------ --------------------- ---------------------------------------- $B^0\to K^{*+}\rho^-$ 8.4 $24.5_{-1.5-3.4-0.6}^{+1.2+2.9+0.0}$ $B_s\to K^{*-}\rho^+$ 24.0 -8.4 -$9.1_{-1.5-1.2-0.3}^{+1.4+1.0+0.2}$ $B^0\to K^{*0}\rho^0$ 3.3 $-8.9_{-0.6-2.8-1.0}^{+0.6+2.8+1.}$ $B_s\to \bar{K}^{*0}\rho^0$ 0.40 72.3 $62.7_{-5.9-16.0-7.9}^{+6.4+10.5+7.5}$ $B^0\to K^{*+}K^{*-}$ 0.21 $29.8_{-5.7-9.5-4.7}^{+2.0+6.4+4.6}$ $B_s\to \rho^+\rho^-$ 1.5 -4.1 -$2.9_{-1.1-1.3-0.2}^{+0.7+1.5+0.2}$ $B^0\to \rho^+\rho^-$ 26.0 $0.83_{-0.59-0.31-0.00}^{+0.50+0.66+0.00}$ $B_s\to K^{*+}K^{*-}$ 5.4 -3.9 $8.8_{-8.9-2.9-0.2}^{+2.5+0.5+0.0}$ \[tb:u-spin\] SUMMARY ======= In this paper, we have reexamined the branching ratios, polarization fractions, relative phases, and direct $CP$ asymmetries in $B_q\to VV$ ($q=u,d,s$) decays in the PQCD approach. Compared to the previous PQCD calculations, the updated longitudinal and transverse decay constants as well as the Gegenbauer moments in the vector mesons wave functions have been adopted, which allows us to reduce the parametric uncertainties in the branching ratios and other observables. What concerns the predictions of the polarization fractions and their relative phases, we have kept track of the terms proportional to the ratio $r_i^2=m_{V_i}^2/m_B^2 (i=2,3)$, which have been ignored in some earlier estimations. In addition, we have studied the decay modes $B \to \rho(\omega)\phi$ that have not been explored before. For the observables $f_{\perp}$, $\phi_{\parallel}$, $\phi_{\perp}$, $\Delta \phi_{\parallel}$, $\Delta\phi_{\perp}$, $A^{0}_{CP}$, and $A_{CP}^{\perp}$, we have provided the first PQCD predictions. So, this work updates and goes beyond what is already known in this approach. Our numerical results are listed in the Tables in the preceding section. For the well-measured decay modes, such as $B\to K^*\phi$ and $B_s\to \phi\phi$, the updated PQCD predictions for all the experimental observables fare better than the previous predictions in this approach, improving comparison with experiments. In addition, in many $B(B_s)\to VV$ decays, our results agree with the updated QCDF predictions [@qcdfbtovv; @qcdfbsvv], as well as with the experimental data. Yet, in some other cases, our predictions and those in QCDF differ and we have discussed some of these decays, such as $B^0\to \rho^0(K^{*0})\omega$ involving the annihilation contributions. For the tree dominated $B\to \rho\rho$ processes, our results respect the isospin triangle relations, while the experimental data, taken on the face value, shows significant isospin-violation. Our estimated decay rate and the polarization fraction in $B^0 \to \rho^0\rho^0$ are in good agreement with the Belle measurement, but not so compared to the BABAR data. This calls for a refined measurement of $B\to \rho\rho$ decays in the future. From the entries in Tables \[tb:fl\] and \[tb:flbs\], we note that our updated longitudinal polarization fractions are in good agreement with the data and the predictions in the QCDF approach [@qcdfbtovv; @qcdfbsvv] in some topologies. But for the color-suppressed decay modes, $B^0\to\rho^0\rho^0$, $B^0\to \rho^0\omega$, $B_s\to\overline K^{*0}\rho^0$ and $B_s\to \overline K^{*0}\omega$, the longitudinal contributions dominate the decay amplitudes in the QCDF approach, while in this work, the transverse polarization contributions are comparable to the longitudinal polarization contributions, and are even dominant in the amplitude for $B^0\to\rho^0\rho^0$. This provides the possibility of distinguishing between these two approaches. Table \[tb:cp\] and \[tb:cpbs\] list predictions of the $CP$ asymmetry parameters,which agree with the experimental data, wherever available, and, generally, also with the QCDF predictions [@qcdfbtovv; @qcdfbsvv] in some topologies. For the color-suppressed decays, $B^0\to \rho^0\rho^0$, $B^0\to \omega\omega$, $B_s\to\overline K^{*0}\rho^0$ and $B_s\to \overline K^{*0}\omega$, both PQCD and QCDF predict large direct $CP$ asymmetries. But for $B^0\to \rho^0\omega$, the central value of the QCDF prediction is only 3$\%$, while the prediction of this work is about 60$\%$ due to the large annihilation contributions. For $B^+\to K^{*+}\omega$ and $B_s\to \rho\phi$ decays, which are almost purely dominated by penguin contributions, we predict very small $CP$ asymmetries, but QCDF predicts them to be of orders 0.56 and 0.83 respectively due to the charming penguins, which needs to be confirmed by experiments. Our predictions for many $B_s^0\to VV$ decays basically agree with the previous PQCD predictions [@bsvv]. But for a few penguin dominant decay modes, for example, $B_s\to \phi\phi$, $B_s\to \overline {K}^{*0}\phi$ and $B_s\to K^{*0}\overline K ^{*0}$, the improvements are significant, especially in the polarization fractions. Acknowledgment {#acknowledgment .unnumbered} ============== We are grateful to Yue-Long Shen for useful discussions. This research was supported in part by the National Science Foundation of China under the Grant Nos. 11447032, 11175151, 11235005, 11205072, 11375208, 11228512, the Natural Science Foundation of Shandong province (ZR2014AQ013) and the Program for New Century Excellent Talents in University (NCET) by Ministry of Education of P. R. China (Grant No. NCET-13-0991). Related Hard Functions {#app:a} ====================== In this appendix, we summarize the functions that appear in the analytic formulas in the Section \[jiexi\]. The first two diagrams in Fig.\[fig:diagram\] are factorizable emission diagrams, whose hard scales $t_{a(b)}$ can be determined by $$\begin{aligned} t_a=\max\{\sqrt{x_3(1-r_2^2)}M_B,\;1/b_1,\;1/b_3\},\end{aligned}$$ $$\begin{aligned} t_b=\max\{\sqrt{x_1(1-r_2^2)}M_B,\;1/b_1,\;1/b_3\}.\end{aligned}$$ The function $h_{ef}$ consists of two parts: the jet function $S_{t}(x)$ and the propagator of virtual quarks and gluons. $$\begin{aligned} h_{ef}(x_1,x_3,b_1,b_3)&=&K_0(\beta b_1)\left[\theta(b_1-b_3)I_0(\alpha b_3)K_0(\alpha b_1)\right.\nonumber\\ &&\left.+\theta(b_3-b_1)I_0(\alpha b_1)K_0(\alpha b_3)\right]S_t(x_3),\end{aligned}$$ with $\alpha=\sqrt{x_3}M_B$ and $\beta=\sqrt{x_1x_3}M_B$. The jet function in the factorization formulas can be given as[@66094010]: $$\begin{aligned} S_t(x)=\frac{2^{1+2c}\Gamma(3/2+c)}{\sqrt{\pi}\Gamma(1+c)}\left[x(1-x)\right]^c,\end{aligned}$$ with $c=0.4$. In the nonfactorizable contributions, due to the small numerical effect, we drop the jet function in the nonfactorizable emission diagrams and nonfactorizable annihilation diagrams[@plb555]. The evolution factors $E_{ef}(t_{a})$ and $E_{ef}(t_{b})$ in the matrix elements are given by $$\begin{aligned} E_{ef}(t)\,=\,\alpha_{s}(t)\exp[-S_{B}(t)-S_{3}(t)].\end{aligned}$$ The Sudakov exponents are defined as $$\begin{aligned} S_{B}(t)\,=\,s\left(x_{1}\frac{M_{B}}{\sqrt{2}},b_{1}\right)\,+\,\frac{5}{3}\int_{1/b_{1}}^{t}\frac{d\bar{\mu}}{\bar{\mu}}\gamma_{q}(\alpha_{s}(\bar{\mu})),\end{aligned}$$ $$\begin{aligned} S_{i}(t)\,=\,s\left(x_{i}\frac{M_{B}}{\sqrt{2}},b_i\right)\,+\,s\left((1-x_{i})\frac{M_{B}}{\sqrt{2}},b_i\right) \,+\,2\int_{1/b_i}^{t}\frac{d\bar{\mu}}{\bar{\mu}}\gamma_{q}(\alpha_{s}(\bar{\mu})),\end{aligned}$$ where the $s(Q,b)$ can be found in the Appendix A in the Ref.[@pqcd1]. $x_{i}$ is the momentum fraction of “quark" in vector meson, with $i=2,3$. For the rest of diagrams, the related functions are summarized as follows: $$\begin{aligned} t_{c}=&&\max\{\sqrt{(1-r_2^2)x_3x_1}\,M_{B},\sqrt{\mid[(x_2-1)(1-r_3^2)+x_1)][r_2^2+x_3(1-r_2^{2})]\mid}\,M_{B},\nonumber\\ &&1/b_{1},1/b_{2}\},\end{aligned}$$ $$\begin{aligned} t_{d}=&&\max\{\sqrt{(1-r_2^2)x_3x_1}\,M_{B},\sqrt{\mid[x_2(r_3^2-1)+x_1)]x_3(1-r_2^{2})\mid}\,M_{B},\nonumber\\ &&1/b_{1},1/b_{2}\}.\end{aligned}$$ $$\begin{aligned} E_{enf}(t)\,=\,\alpha_{s}(t)\exp[-S_{B}(t)-S_{2}(t)-S_{3}(t)]| \,_{b_{1}=b_{3}}.\end{aligned}$$ $$\begin{aligned} h_{enf}(\alpha,\beta_i,b_{1},b_{2})\,&=&\,\left[\theta(b_{2}-b_{1})I_{0}(\alpha b_1)K_{0}(\alpha b_2)+\theta(b_{1}-b_{2})I_{0}(\alpha b_2)K_{0}(\alpha b_1)\right]\nonumber\\ &&\times \left\{\begin{array}{ll} \frac{i\pi}{2}H_{0}^{(1)}\left(\sqrt{|\beta^2_i|}M_Bb_{2}\right),& \;\;\beta_i^2<0;\\ K_{0}\left(\beta_i M_B b_{2}\right),&\;\;\beta_i^{2}>0, \end{array}\right.\end{aligned}$$ with $i=1,2$ and $$\begin{aligned} \alpha &=&\sqrt{(1-r_2^2)x_3x_1}M_B,\\ \beta_{1}^{2}&=&[(x_2-1)(1-r_3^2)+x_1)][r_2^2+x_3(1-r_2^{2})],\\ \beta_2^{2}&=&[x_2(r_3^2-1)+x_1)]x_3(1-r_2^{2}),\end{aligned}$$ The hard functions and the scales for factorizable annihilation diagrams Fig.(e) and (f) are $$\begin{aligned} &&t_{e}\,=\,\max\{ \alpha_1 M_{B},\beta M_B,1/b_{2},1/b_{3}\},\nonumber\\ &&t_{f}\,=\,\max\{\alpha_2 M_{B},\beta M_B,1/b_{2},1/b_{3}\},\\ &&E_{af}(t)\,=\,\alpha_{s}(t)\cdot \exp[-S_{2}(t)-S_{3}(t)],\end{aligned}$$ $$\begin{aligned} h_{af}(\alpha_i,\beta,b_{2},b_{3})\,&=&\,(\frac{i\pi}{2})^{2}H_{0}^{(1)}\left(\beta M_{B}b_{2}\right)\left[\theta(b_{2}-b_{3})H_{0}^{(1)}\left(\alpha_i M_{B}b_{2}\right)J_{0}\left(\alpha_i M_{B}b_{3}\right)\right.\nonumber\\ &&\left.+\theta(b_{3}-b_{2})H_{0}^{(1)}\left(\alpha_i M_{B}b_{3}\right)J_{0}\left(\alpha_i M_{B}b_{2}\right)\right]\cdot S_{t}(x_{3}),\end{aligned}$$ with $$\begin{aligned} \alpha_1&=&\sqrt{1-x_3(1-r_2^2)}\\ \alpha_2&=&\sqrt{(1-r_2^2)[r_3^2+x_2(1-r_3^2)]},\\ \beta&=&\sqrt{[(1-r_2^2)(1-x_3)][r_3^2+x_2(1-r_3^2)]}.\end{aligned}$$ For the nonfactorizable annihilation diagrams, the scales and the hard functions are $$\begin{aligned} t_g&=&\max\{\alpha M_{B}, \sqrt{|\beta_1|}M_{B}, 1/b_1,1/b_2\},\\ t_h&=&\max\{\alpha M_{B},\sqrt{|\beta_2|} M_{B},1/b_1,1/b_2\},\\ E_{anf}(t)\,&=&\,\alpha_{s}(t)\cdot \exp[-S_{B}(t)-S_{2}(t)-S_{3}(t)]\mid\,_{b_{2}=b_{3}},\end{aligned}$$ $$\begin{aligned} h_{anf}(\alpha,\beta_i,b_{1},b_{2})\,&=&\,\frac{i\pi}{2}\left[\theta(b_{1}-b_{2})H_{0}^{(1)}\left(\alpha M_{B}b_{1}\right)J_{0}\left(\alpha M_{B}b_{2}\right)\right.\nonumber\\ &&\left.+\theta(b_{2}-b_{1})H_{0}^{(1)}\left(\alpha M_{B}b_{2}\right)J_{0}\left(\alpha M_{B}b_{1}\right)\right]\nonumber\\ &&\times \left\{\begin{array}{ll} \frac{i\pi}{2}H_{0}^{(1)}\left(\sqrt{|\beta_i|}M_{B}b_{1}\right),& \beta_{i}<0,\\ K_{0}\left(\sqrt{\beta_{i}}M_{B}b_{1}\right),& \beta_{i}>0, \end{array}\right.\end{aligned}$$ with $i=1,2$. $$\begin{aligned} \alpha&=&\sqrt{(1-x_3)(1-r_2^2)[r_3^2+x_2(1-r_3^2)]},\\ \beta_{1}&=&1-[(1-r_3^2)(1-x_2)-x_1][r_2^2+x_3(1-r_2^2)],\\ \beta_{2}&=&(1-r_2^2)(1-x_3)[x_1-x_2(1-r_3^2)-r_3^2].\end{aligned}$$ [99]{} M. Beneke, J. Rohrer and D. S. Yang, Nucl. Phys. **B774**, 64 (2007). M. Bartsch, G. Buchalla and C. Kraus, arXiv:0810.0249 (2008). H. Y. Cheng and K. C. Yang, Phys. Rev. D **78** (2008) 094001, Erratum-ibid. D **79** (2009) 039903. H. Y. Cheng and C. K. Chua, Phys. Rev. D **80**, 114008 (2009). H. Y. Cheng and C. K. Chua, Phys. Rev. D **80**, 114026 (2009). X. Q. Li, G. R. Lu and Y. D. Yang, Phys. Rev. D **68**, 114015 (2003), Erratum-ibid. D **71** 019902 (2005). Y. Li and C. D. Lü, Phys. Rev. D **73**, 014024 (2006); H. W. Huang, C. D. Lü, T. Morii, Y. L. Shen, G. L. Song, and J. Zhu, Phys. Rev. D **73**, 014011 (2006); J. Zhu, Y. L Shen, and C. D. Lü, Phys. Rev. D **72**, 054015 (2005); C. D. Lü, Y. L. Shen, J. Zhu, Eur. Phys. J. C **41**, 311-317 (2005). A. Ali, G. Kramer, Y. Li, C. D. Lü, Y. L. Shen, W. Wang, and Y. M. Wang, Phys. Rev. D **76**, 074018 (2007). J. Zhu, Y. L. Shen and C. D. Lü, J. Phys. G **32**, 101-110 (2006). Heavy Flavour Averaging Group (HFAG); Y. Amhis [*et al.*]{}, arxiv:1412.7515, and online update at http://www.slac.stanford.edu/xorg/hfag. M. Wirbel, B. Stech and M. Bauer, Z. Phys. C **29**, 637 (1985); B. Stech and M. wirbel, Z. Phys. C **34**, 103 (1987). A. Ali and C. Greub, Phys. Rev. D **57**, 2996 (1998); G. Kramer, W. E. Palmer and H. Simma, Nucl. Phys. **B428**, 77 (1994); Z. Phys. C **66**, 429 (1995); A. Ali, G. Kramer and C. D. Lü, Phys. Rev. D **58**, 094009 (1998); Phys. Rev. D **59**, 014005 (1999); Y. H. Chen, H. Y. Cheng, B. Tseng and K. C. Yang, Phys. Rev. D **60** 094014 (1999). M. Beneke, G. Buchalla, M. Neubert and C. T. Sachrajda, Phys. Rev. Lett. **83**, 1914 (1999); Nucl. Phys. **B591**, 313 (2000). M. Beneke and M. Neubert, Nucl. Phys. **B675**, 333 (2003). Y. Y. Keum, H. N. Li and A. I. Sanda, Phys. Lett. B **504**, 6 (2001); C. D. Lü, K. Ukai and M. Z. Yang, Phys. Rev. D **63**, 074009 (2001). C. W. Bauer, D. Pirjol and I. W. Stewart, Phys. Rev. Lett. **87**, 201806 (2001). C. D. Lü and M. Z. Yang, Eur. Phys. J. C **23**, 275 (2002). Y.Y. Keum, H. N. Li, and A. I. Sanda, Phys. Lett. B **504**, 6 (2001); Phys. Rev. D **63**, 054008 (2001). C. D. Lü, K. Ukai, and M. Z. Yang, Phys. Rev. D **63**, 074009. H. N. Li, Prog. Part. Phys. **51**, 85 (2003), and reference therein. C. D. Lü and K. Ukai, Eur. Phys. J. C **28**, 305 (2003). Y. Li and C. D. Lü, J. Phys. G **29**, 2115 (2003); High Energy Phys. Nucl. Phys.**27**, 1062 (2003). Y. Li, C. D. Lü, Z. J. Xiao, and X. Q. Yu, Phys. Rev. D **70**, 034009 (2004). R. H. Li, C. D. Lü, and H. Zou, Phys. Rev. D **78**, 014018 (2008). M. J. Morello *et* *al*. (CDF Collaboration), CDF public note, Report No. 10498,2011; K. Nakamura *et* *al*. (Particle Data Group), J. Phys. G **37**, 075021 (2010). G. Buchalla, A. J. Buras and M. E. Lautenbacher, Rev. Mod. Phys. **68**, 1125 (1996). A. Ali, G.Kramer and C. D. Lü in Ref. [@gaijin]. C. H. Chang and H. N. Li, Phys. Rev. D **55**, 5577 (1997); T. W. Teh and H. N. Li, Phys. Rev. D **56**, 1615 (1997). H. N. Li, Phys. Rev. D **66**, 094010 (2002). H. N. Li and B. Tseng, Phys. Rev. D **57**, 443 (1998) C. D. Lü and M. Z. Yang, Eur. Phys. J. C **23**, 275-287 (2002). A. G. Grozin and M. Neubert, Phys. Rev. D **55**, 272 (1997); M. Beneke and T. Feldmann, Nucl. Phys. **B592**, 3 (2001). H. Kawamura, J. Kodaira, C. F. Qiao, and K. Tanaka, Phys. Lett. B **523**, 111 (2001); **536**, 344(E) (2002); Mod. Phys. Lett. A **18**, 799 (2003). Y. Y. Keum, H. N. Li, Phys. Rev. D **63**, 074006 (2001); C. D. Lü, M. Z. Yang, Eur. Phys. J. C **23**, 275 (2002); H. N. Li and H. L. Yu, Phys. Rev. D **53**, 2480 (1996). H. N. Li, Phys. Lett. B **622**, 63 (2005). P. Ball, V. M. Braun, Y. Koike, and K. Tannka, Nucl. Phys. **B529**, 323 (1998); P. Ball and V. M. Braun, Nucl. Phys. **B543**, 201 (1999); P. Ball and R. Zwicky, Phys. Rev. D **71**, 014029 (2005). P. Ball and G. W. Jones, J. High Energy Phys. 03 (2007) 069. Zhou Rui, Gao Xiang-Dong, C. D. Lü, Eur. Phys. J. C **72**, 1923 (2012); Hsiang-nan Li, Yue-Long Shen, and Yu-Ming Wang, Phys. Rev. D **85**, 074004 (2012). B. H. Hong and C. D. Lü, Sci. China G**49**, 357 (2006); H. W. Huang *et al*.,Phys. Rev. D **73**, 014011 (2006); H. N. Li and S. Mishima, Phys. Rev. D **71**, 054025 (2005). J. Beringer *et al*.,(Particle Data Group), Phys. Rev. D **86**, 010001 (2012). R. Aaij *et al*. (LHCb Collaboration), JHEP11(2013)092. A. L. Kaan, Phys. Lett. B **601**, 151 (2004). H. Y. Cheng, C. K. Chua, and A. Soni, Phys. Rev. D **71**, 014030 (2005); P. Colangelo, F. De Fazio, and T. N. Pham, Phys. Lett. B **597**, 291 (2004). H. N. Li, Phys. Lett. B **622**, 63 (2005). C. S. Kim and Y. D. Yang, arXiv:hep-ph/0412364; S. Baek, A. Datta, P. Hamel, O. F. Hernandez, and D. London, Phys. Rev. D **72**, 094008 (2005); Q. Chang, X. Q. Li, and Y. D. Yang, J. High Energy Phys. 06 (2007) 038; B. Aubert *et al*. (BABAR Collaboration), Phys. Rev. D **78**, 071104(R) (2008). P. Vanhoefer *et al*. (Belle Collaboration), arXiv:1212.4015 \[hep-ex\]. P. Vanhoefer *et al*. (Belle Collaboration), arXiv:1212.4015 \[hep-ex\]. H. N. Li and S. Mishima, Phys. Rev. D **71**, 054025 (2005). J. P. Lees *et al*. (BABAR Collaboration) ,Phys. Rev. D **85**, 072005 (2012). Zhen-Jun Xiao, Wen-Fei Wang, Y.-Y. Fan, Phys. Rev. D **85**,094003 (2012). A. L. Kagan, Phys. Lett. B **601**, 151 (2004). B. Aubert *et al*. (BABAR Collaboration), Phys. Rev. D **78**, 071104 (2008). C. C. Chiang *et al*. (Belle Collaboration), Phys. Rev. D **78**, 111102 (2008). H. n. Li and S. Mishima, Phys. Rev. D [**73**]{}, 114014 (2006) \[hep-ph/0602214\]; H. n. Li, Private communications. Y. Li, Phys. Rev. D **89**, 014003 (2014). R. Fleisher, R. Knegjens, Eur. Phys. J. C **71**, 1789 (2011); R. Fleischer, Eur. Phys. J. C **10**, 299 (1999); K. D. Bruyn, R. Fleischer, arXiv:1412.6834 \[hep-ph\]; K. D. Bruyn, R. Fleischer, R. Knegjens, M. Merk, Nucl. Phys. **B868**, 351-367 (2013). H.-n. Li, Phys. Rev. D **66**, 094010 (2002). H.-n. Li and K. Ukai, Phys. Lett. B **555**, 197 (2003). [^1]: [email protected] [^2]: [email protected] [^3]: [email protected] [^4]: [email protected]
{ "pile_set_name": "ArXiv" }
--- abstract: 'We investigate top anti-top quark pair production in lead-lead collisions at the Large Hadron Collider with nucleon-nucleon center of mass energy of 5.5 TeV. Due to the very high temperature and energy density created in heavy ion collision, a new state of QCD matter known as Quark-Gluon Plasma (QGP) is expected to be produced. Top decay products loose energy inside the QGP medium. Therefore, we also study the medium modifications of different kinematic distributions. We observe significant modification in the dijets and trijets invariant mass distributions.We also found that the peak position and shape of the distributions could be used to characterize the nature of jet energy loss in the QGP.' author: - 'Lusaka Bhattacharya$^1$, Kirtiman Ghosh$^2$ and Katri Huitu $^3$' title: 'Top anti-top pairs at the LHC heavy ion collision: a new interesting probe of quark gluon plasma' --- The primary goal of heavy ion collisions (HIC) at Relativistic Heavy Ion Collider (RHIC) at BNL, and the Large Hadron Collider (LHC) at CERN is to produce and study the properties of a hot/dense state of QCD matter known as [*Quark-Gluon Plasma*]{} (QGP) [@dk1]. QGP is a deconfined state of matter where quarks and gluons are the effective degrees of freedom rather than nucleons or hadrons [@dk2]. Fast partons propagating in a hot/dense nuclear medium are expected to loose a large fraction of their energy [@jq]. The observation of the suppression of energetic partons in the QGP, that is [*jet quenching*]{} [@jq_RHIC], and centrality-dependent [*dijet asymmetry*]{} [@da_LHC] are the most important results from the HIC at RHIC and LHC experiment, respectively. $W/Z$-bosons are massive weakly interacting Standard Model (SM) particles. The larger LHC HIC energies open the possibility to probe the nucleus-nucleus collisions via the $W/Z$-bosons. The vector bosons are produced early ($1/M_{W(Z)} \sim 10^{-3}$ fm/c) and their decay time is small ($\tau_Z \sim 0.08$ fm/c and $\tau_W \sim 0.09$ fm/c [@pdg]). Whereas, in the most accepted picture of the QGP formation and evolution, at the LHC, QGP is expected to form after $1/\Lambda_{\rm QCD}\sim 1$ fm/c of the initial hard scattering, thermalize quickly and it might last $\sim 10$ fm/c. Therefore, weak bosons are produced and decay before the formation of QGP and the decay products of pass through the QGP. $W/Z$-boson dominantly decays to a pair of quarks. Quarks loose energy in the QGP and thus the hadronic decays of $W/Z$-boson could be an interesting probe to characterize QGP. However, in presence of huge QCD dijet background, it is extremely challenging to study the hadronic decays of $W/Z$-bosons. The leptonic BF of $W/Z$-boson is small however, due to small background, the signature could be easily detected at the LHC HIC. Unlike jets, leptons interact electromagnetically with the QGP and loose experimentally insignificant amount of energy within the QGP [@ConesadelValle:2009vp]. Therefore, the leptonic decays of weak bosons behave as a medium blind reference. ATLAS collaboration has already measured the $W/Z$-boson yield in the leptonic decay channels for $\sqrt {s_{NN}}=2.76$ TeV [@ATLAS_WZ]. A new regime of heavy ion physics will be reached at the LHC with $\sqrt {s_{NN}} = 5.5$ TeV where hard and semi-hard particle production can dominate over the underlying soft events. The higher LHC energies open the possibility to study top quarks for the first time at the HIC. The top quark was discovered at the Tevatron experiment [@top]. After the discovery, properties of top quark have been extensively studied at the Tevatron and LHC. Presently, different properties of top quark is known with good precision. As an example, top quark mass and full decay width are determined to be $m_{t}=173.5\pm 0.6$ GeV and $\Gamma_t=2.0^{+0.7}_{-0.6}$ GeV [@pdg]. Different decay channels and branching ratios of top quark have also been observed and measured. As a result, top quark can now be considered as standard benchmark for other experimental observations. In view of this fact, it is important to investigate top quarks at the LHC HIC and study the medium (QGP) influence on different kinematic distributions which are precisely known from the previous $pp$ and $p\bar p$ collider experiments. In this letter, we have for the first time studied $t\bar t$ production at the LHC HIC and proposed few kinematic distributions for the study of the QGP created in the LHC HIC. The main source of top quarks at the LHC HIC is the top anti-top ($t \bar t$) pair production. At leading order in perturbation theory there are two processes that contribute to $t\bar t$ production: quark-antiquark annihilation, $q\bar q \to t\bar t$ and gluon-gluon fusion, $gg \to t\bar t$. With $5.5$ TeV center-of-mass energy per nucleon, the NLO+NNLL $t\bar t$ production cross-section per nucleon-nucleon collision in Pb-Pb reaction is given by $\sigma_{NN}(t\bar t)=80.6$ pb [@top_cross]. Therefore, the total $t\bar t$ production cross-section in minimum bias Pb-Pb scattering is estimated to be 3.5 $\mu b$ in the frame work of the Glauber model [@GM]. The instantaneous luminosity of the LHC Pb-Pb collision at $\sqrt s_{NN}=5.5$ TeV is expected to be $10^{27}{\rm cm^2 s^{-1}}$. Therefore, $1$ nb$^{-1}$ integrated luminosity data will be accumulated within one month ($10^6$ second) of Pb-Pb collision at $\sqrt s_{NN}=5.5$ TeV. These estimations indicate that about 3500 $t\bar t$ events are expected to be produced in one month of LHC HIC running at $\sqrt s_{NN}=5.5$ TeV. Due to the large top mass, $t \bar t$ pairs are produced early in the HIC. The decay width of top quark is large and thus the decay time is small. As a result, top quarks are produced and decay before formation of QGP. Top quark decays to a bottom quark ($b$) and $W$-boson with almost 100% branching fraction: $t\to b W^+$. $W$-boson subsequently decays hadronically ($W^\pm \to q \bar q^{\prime}$) with 67.7% BF or leptonically ($W^\pm \to l \nu_l$) with $32.3$% BF [@pdg]. Therefore, for $t \bar t$ production, there are only three possible final state topologies: [*(i) Hadronically decaying $t \bar t$ pairs:*]{} Both the top quarks decay hadronically ($t\to b q \bar q^{\prime}$) and give rise to 2-$b$ jets and four light quarks jets in the final state with 46% effective branching ratio. [*(ii) Semi-leptonically decaying $t \bar t$ pairs:*]{} One top quark decays hadronically and the other decays leptonically ($t\to b l \nu_l$). The final state is characterized by $2$-$b$ jets+$2$-light quark jets+one charged lepton + one neutrino. The effective branching fraction of semi-leptonic decay mode of $t \bar t$ pairs is $43.7$%. [*(iii) Leptonically decaying $t \bar t$ pairs:*]{} In this case, both the top quarks decay leptonically giving rise to $2$-$b$ jets + $2$-lepton + $2$-neutrino final state with 10.4% effective branching fraction. The jets ($b$-jets as well as light jets) and leptons can be observed at the LHC. The neutrinos remain invisible at the detectors and give rise to a imbalance in the transverse momentum known as missing transverse momentum. However, in the HIC environment, faithful measurement of missing transverse momentum will be challenging in the presence of a continuum energy deposit from the produced QGP. Hadronically decaying $t \bar t$ events suffer from huge QCD background. Moreover, due to a large combinatorial background it is difficult to reconstruct top quark and $W$-boson from hadronically decaying $t\bar t$ pairs. Leptonically decaying $t \bar t$ channel is a clean signal channel due to less background. However, the rate of this channel is suppressed by the top quark leptonic branching fraction. Moreover, in the absence of proper knowledge about missing transverse momentum, the reconstruction of top quark and $W$-boson mass will be difficult from $t \bar t$ leptonic decay channel. As a result, in this letter, we have investigated the semi-leptonic decay channel of $t \bar t$ pairs due to the following advantages: (i) The rate of semi-leptonic $t \bar t$ final state is relatively large. If we consider only electron and muon decay modes, Pb-Pb collision with $\sqrt s_{NN}=5.5$ TeV gives rise to about 1000 semi-leptonically decaying $t \bar t$ events for $1$ nb$^{-1}$ integrated luminosity. (ii) Due to the presence of a lepton, the semi-leptonic final state of $t\bar t$ pairs suffers less from the QCD background. (iii) At the parton level, the semi-leptonic $t\bar t$ final state contains only $2$-light quark jets arising from the $W$-decay. Therefore, $W$ invariant mass can be constructed with out any ambiguity. However, there is a two fold ambiguity in the reconstruction of top quark mass. After the decay of $t \bar t$ pairs, the decay products pass through the hot and dense QGP medium and thus, loose energy. The energy loss of energetic partons, so called “jet quenching”, leads to a number of phenomena which are already seen at RHIC and LHC. In this work, we have investigated the quenching of $t \bar t$ decay products and proposed some new phenomena which could be seen at the LHC HIC. We have used PYTHIA [@pythia] to simulate the production and decay of $t \bar t$ pairs. Subsequently, the PYTHIA generated $t \bar t$ events are passed in to a fast Monte-Carlo simulator PYQUEN [@pyquen] for simulating the energy loss (quenching) of $t \bar t$ decay products. Finally, quarks and gluons are hadronized according to the Lund string mode . PYQUEN simulates the radiative and collisional (Coll.) energy loss [@coll_loss] of hard partons in longitudinally expanding QGP taking into account the realistic nuclear geometry. The radiative energy loss is calculated in the framework of BDMS model [@bdms] with the simple generalization to a massive quark case using the “dead-cone” approximation [@dead-cone]. Measuring jet energy as a sum of the energies of final hadrons moving inside an angular cone with a given finite size allow some of the radiated gluons to belong to the jet and thus some part of the radiated energy to be reconstructed. Therefore, the knowledge of angular structure of medium-induced radiation is very important for any phenomenological prediction using jets. In this analysis, we have used the simple parametrizations of the gluon distribution over the emission angle $\theta$ available in PYQUEN: (i) Small-angular radiation (SAR): $dN^g/d\theta \propto {\rm sin}\theta {\rm exp}[-(\theta-\theta_0)^2/2\theta_0^2]$, where $\theta_0\sim 5^0$ is the typical angle of the coherent gluon radiation as estimated in Ref. [@SA]; (ii) Wide-angular radiation (WAR): $dN^g/d\theta \propto 1/\theta$. The strength of the energy loss in PYQUEN is determined mainly by the initial maximal temperature $T_0^{max}$ of hot matter in Pb-Pb collisions. The energy loss also depends on the number $N_f$ of active flavors in the medium. In our analysis, we have used $T_0^{max}=1$ GeV and $N_f=2$. In this letter, we have investigated $4$-jets out of which $2$-jets are $b$-tagged plus one lepton signature as signal of $t\bar t$ production in HIC. Therefore, before going into the details of our analysis, it is important to discuss the status of jet reconstruction and $b$-tagging in the context of HIC. The main obstacle to studying jets in HIC is the presence of the huge background given by the underlying event (UE) This UE needs to be properly subtracted from the momentum of a given jet in order to reconstruct its “true” momentum. It was shown in Ref. [@Cacciari:2010te] that in presence of QGP, faithful reconstruction of jets are possible using anti-$k_T$ algorithm. Jets have already been successfully used as an observable in HIC by the ATLAS collaboration [@da_LHC]. Moreover, in Ref. [@vitev], different jet shape variables have been studied for the jets passing through QGP. In Ref [@White:2005au], $b$-tagging in the environment of HIC has been studied by examining the reconstruction efficiency and rejection power (against light quark jets) using secondary vertex finding. Their study suggests that $\epsilon_b=50\%$ $b$-tagging efficiency can be achieved at the HIC for a rejection power of 50. -15pt ------------- ------- ------- ------------ ------------ ------- ------- ------------ ------------ Energy loss Scenario $a_0$ $m_0$ $\sigma_1$ $\sigma_2$ $a_0$ $m_0$ $\sigma_1$ $\sigma_2$ nb GeV GeV GeV nb GeV GeV GeV Only Coll. 8.6 75.6 6.1 3.8 5.3 162.8 8.6 6.2 Only SAR 19.6 78.9 2.6 2.2 14.8 170.4 3.7 2.8 Only WAR 9.5 78.7 4.0 2.4 6.2 169.9 9.9 3.0 Coll.+SAR 7.16 75 6.4 4.9 4.5 161.9 8.9 7.5 Coll.+WAR 4.25 76.9 12.1 3.3 2.6 160.9 14.9 9.25 ------------- ------- ------- ------------ ------------ ------- ------- ------------ ------------ : Parameters used for the Gaussian fitting of Fig \[wmass\] and \[tmass\].[]{data-label="fit"} -15pt In our analysis, we have introduced a set of basic selection criteria to identify electrons, muons, jets etc. The object selection is described in brief in the following: (i) Jets are constructed using anti-$k_T$ algorithm with $R=0.4$ and only jets with $p_T > 20$ GeV and $|\eta| < 2.5$ are considered for further analysis. To take into account the effects of finite detector resolution, we have smeared the jet energies with Gaussian functions in Ref. [@smear]. (ii) We demand that lepton candidates (both electron and muon) have $p_T > 20$ GeV and are separated from jets by at least $\Delta R = 0.5$. After reconstructing different objects, we consider events with one lepton (electron or muon) and $\ge$ 4-jets for further analysis. We also demand that out of the 4-jets, two jets are $b$-tagged. The dominant background for semileptonic $t \bar t$ signal arises from $W/Z$+jets production followed by the leptonic decay of $W/Z$-boson. Here, $b$-jets results from the mistagging of light jets. Since the mistagging efficiency of light quark jets to be tagged as $b$-jets is small, $b$-tagging significantly reduce this background. Production of $W/Z~b\bar b$+jets also contributes to the background. However, we have estimated that these cross-sections are very small compared to the $t\bar t$ cross-section. Therefore, semileptonic decay products of $t\bar t$ could be easily detected over the background. To study the influence of the hot/dense QGP, we have constructed the following kinematic distributions.\ [*Di-jets invariant mass distribution:*]{} We have ordered the non $b$ jets according to their $p_T$ hardness ($p_T^{j_1}>p_T^{j_2}>...$) and constructed the invariant mass of the hardest ($j_1$) and the second hardest ($j_2$) jet, $m_{jj}$. In the semileptonic decay of $t\bar t$ pairs, non $b$-jets arise from the decay of one $W$-boson. Therefore, in absence of QGP, $m_{jj}$ distribution should be peaked at the $W$-mass, $m_W=80.4$ GeV. However, in presence of QGP, $W$-decay product suffers energy loss. In Fig. \[wmass\], we have presented the $m_{jj}$ distributions for different energy loss scenarios. To determine the peak position of the distributions, we have fitted the distributions with asymmetric Gaussian functions:\ $$f(m) = a_0 \left\{ \begin{array}{rl} {\rm exp}\left (-\frac{(m-m_0)^2}{\sigma_1}\right) &\mbox{ if $m<m_0$} \\ {\rm exp}\left (-\frac{(m-m_0)^2}{\sigma_2}\right ) &\mbox{ otherwise} \end{array} \right.$$ where, $a_0$, $m_0$ and $\sigma_{1,2}$ are the parameters of fitting. In Fig. \[wmass\] (left panel), we have presented $m_{jj}$ distributions for collisional and radiative (SAR as well as WAR) energy loss scenario separately. Corresponding fitting parameters are presented in Table \[fit\]. Fig. \[wmass\] (left panel) and Table \[fit\] show that collisional energy loss significantly (about 5 GeV) shifts the peak position of $m_{jj}$ distribution from $W$-mass. Whereas, radiative energy loss (both SAR and WAR) gives rise to small change in the peak position of $m_{jj}$ distribution. Due to the high boost of hard partons, the radiated gluons shift towards the parent parton direction and thus, resulting jets include large part of radiated gluon energies. As a result, radiative energy loss of hard partons has negligible impact on the peak position of $m_{jj}$ distribution. However, Fig. \[wmass\] (left panel) shows that WAR significantly changes the shape (which could be quantified by the fitting parameters $a_0,~\sigma_1~{\rm and}~\sigma_2$ in Table \[fit\]) of $m_{jj}$ distribution. In the passage of a fast parton through QGP, collisional and radiative energy loss occur simultaneously. Therefore, in Fig. \[wmass\] (right panel), we have presented $m_{jj}$ distribution in presence of both collisional and radiative energy loss. Corresponding fitting parameters are presented in the last two rows of Table \[fit\]. Table \[fit\] shows that the peak position of $m_{jj}$ distribution is determined by the collisional energy loss. Whereas, the shape parameters are governed by the nature of radiative loss. In Fig. \[wmass\] (right panel), we have also presented the background. The background contributions are substantially small compared to the $t \bar t$ contribution.\ [*Tri-jets ($bjj$) invariant mass distribution:*]{} Hadronic decay of one top quark gives rise to one $b$-jet and two light quark jets. Therefore, it is important to study invariant mass distribution of $b$ jet and di-jets system arising from the hadronic top quark decay. However, it is difficult to identify the $b$ jet and di-jets arising from the same top decay. There are several algorithm available in the literature for the reconstruction of top quark. However, most of the algorithms rely on the knowledge of $m_{bjj}$ peak position (top mass: $m_t=173.2$ GeV). In presence of QGP, energy loss of top decay products shift the peak position of $m_{bjj}$ distribution. Therefore, most of the top reconstruction algorithm are not applicable for Pb-Pb collision in their present form. In our analysis, we have used the following simplified algorithm for reconstructing $m_{bjj}$ peak position. We first order $b$ jets according to their $p_T$ hardness ($p_T^{b_1}>p_T^{b_2}$) and constructed two invariant masses: $m_{b_1 jj}~{\rm and}~m_{b_2 jj}$. Out of these two invariant masses, we consider the invariant mass which is close to the $m_t=173.2$ GeV for plotting $m_{bjj}$ distribution. In Fig. \[tmass\], we have presented $m_{bjj}$ distributions. In Fig. \[tmass\] (left panel), we have shown the effect of collisional and radiative energy loss on $m_{bjj}$ distribution separately. Whereas, right panel of Fig. \[tmass\] shows the resulting $m_{bjj}$ distribution if we consider both collisional and radiative energy loss simultaneously. We have also fitted these distributions with asymmetric Gaussian functions and the fitting parameters are presented in Table \[fit\]. Due to the energy loss, the position of $m_{bjj}$ peak shifts about 12 GeV from $m_t$. The shift could be easily observed at the LHC as a signature of QGP. To summarize, we have investigated semileptonic $t \bar t$ signature at the LHC HIC. Semileptonic $t \bar t$ signature could be easily observed at the LHC HIC with $\sqrt s_{NN}=5.5$ TeV. However, due to the presence of QGP $t \bar t$ decay products suffer energy loss and thus shape of different kinematic distributions are modified significantly. As for example, we have studied dijets and trijets invariant mass distributions and predicted significant change and shift in the shape and peak position of these distributions, respectively. [499]{} B. Muller, [*The Physics of Quark Gluon Plasma*]{}, Springer, Heidelberg, 1985. C. Y. Wong, [*Introduction of High Energy Heavy Ion Collisions*]{}, World Scientific, Singapore, 1994; R. C. Hwa (ed.), [*Quark Gluon Plasma*]{}, Vol. I II, World Scientific, Singapore, 1990, 1995. J. D. Bjorken, FERMILAB-PUB-82-059-THY (1982). J. Adams [*et al.*]{} \[STAR Collaboration\], Nucl. Phys. A [**757**]{}, 102 (2005); K. Adcox [*et al.*]{} \[PHENIX Collaboration\], Nucl. Phys. A [**757**]{}, 184 (2005). G. Aad [*et al.*]{} \[Atlas Collaboration\], Phys. Rev. Lett.  [**105**]{}, 252303 (2010); S. Chatrchyan [*et al.*]{} \[CMS Collaboration\], Phys. Lett. B [**712**]{}, 176 (2012). C. Amsler [*et al.*]{} (Particle Data Group), Phys.Lett.B 667, (2008). Z. Conesa del Valle, Eur. Phys. J. C [**61**]{}, 729 (2009). The ATLAS Collaboration, ATLAS-CONF-2011-078; The ATLAS Collaboration, ATLAS-CONF-2012-052. F. Abe [*et al.*]{} \[CDF Collaboration\], Phys. Rev. Lett.  [**80**]{}, 2779 (1998); F. Abe [*et al.*]{} \[CDF Collaboration\], Phys. Rev. Lett.  [**82**]{}, 271 (1999) \[Erratum-ibid.  [**82**]{}, 2808 (1999)\]; B. Abbott [*et al.*]{} \[D0 Collaboration\], Phys. Rev. Lett.  [**80**]{}, 2063 (1998). http://www.lpthe.jussieu.fr/ cacciari/ttbar/ M. L. Miller, K. Reygers, S. J. Sanders and P. Steinberg, Ann. Rev. Nucl. Part. Sci.  [**57**]{}, 205 (2007); B. Alver, M. Baker, C. Loizides and P. Steinberg, arXiv:0805.4411 \[nucl-ex\]. T. Sjostrand[*et al.*]{}, JHEP [**0605**]{}, 026 (2006). I. P. Lokhtin, A.M. Snigirev, Eur. Phys. J. C 45, (2006) 211. J. D. Bjorken, Fermilab publication Pub-82/29-THY, 1982; E. Braaten, M. Thoma, Phys. Rev. D 44, (1991) 1298; I.P. Lokhtin, A.M. Snigirev, Eur. Phys. J. C 16, (2000). R. Baier, Yu. L. Dokshitzer, A.H. Mueller, D. Schiff, Phys. Rev. C 60, (1999) 064902; R. Baier, Yu. L. Dokshitzer, A.H. Mueller, D. Schiff, Phys. Rev. C 64, (2001) 057902. Yu.L. Dokshitzer, D. Kharzeev, Phys. Lett. B 519, (2001) 199 I.P. Lokhtin, A.M. Snigirev, Phys. Lett. B 440, (1998) 163. M. Cacciari, J. Rojo, G. P. Salam and G. Soyez, Eur. Phys. J. C [**71**]{}, 1539 (2011). I. Vitev, S. Wicks and B. -W. Zhang, JHEP [**0811**]{}, 093 (2008); B. -W. Zhang, Y. He, R. B. Neufeld, I. Vitev and E. Wang, arXiv:1207.6558 \[nucl-th\]; I. Vitev, J. Phys. G [**38**]{}, 124087 (2011); Y. He, I. Vitev and B. -W. Zhang, Phys. Lett. B [**713**]{}, 224 (2012). S. N. White, Acta Phys. Hung. A [**25**]{}, 531 (2006). G. Aad [*et al.*]{}, \[ATLAS Collaboration\], arXiv:0901.0512.
{ "pile_set_name": "ArXiv" }
--- abstract: 'The abstract should appear at the top of the left-hand column of text, about 0.5 inch (12 mm) below the title area and no more than 3.125 inches (80 mm) in length. Leave a 0.5 inch (12 mm) space between the end of the abstract and the beginning of the main text. The abstract should contain about 100 to 150 words, and should be identical to the abstract text submitted electronically along with the paper cover sheet. All manuscripts must be in English, printed in black ink.' address: 'Author Affiliation(s)' bibliography: - 'strings.bib' - 'refs.bib' title: AUTHOR GUIDELINES FOR ICIP 2013 PROCEEDINGS MANUSCRIPTS --- One, two, three, four, five Introduction {#sec:intro} ============ These guidelines include complete descriptions of the fonts, spacing, and related information for producing your proceedings manuscripts. Please follow them and if you have any questions, direct them to Conference Management Services, Inc.: Phone +1-979-846-6800 or email to\ `[email protected]`. Formatting your paper {#sec:format} ===================== All printed material, including text, illustrations, and charts, must be kept within a print area of 7 inches (178 mm) wide by 9 inches (229 mm) high. Do not write or print anything outside the print area. The top margin must be 1 inch (25 mm), except for the title page, and the left margin must be 0.75 inch (19 mm). All [*text*]{} must be in a two-column format. Columns are to be 3.39 inches (86 mm) wide, with a 0.24 inch (6 mm) space between them. Text must be fully justified. PAGE TITLE SECTION {#sec:pagestyle} ================== The paper title (on the first page) should begin 1.38 inches (35 mm) from the top edge of the page, centered, completely capitalized, and in Times 14-point, boldface type. The authors’ name(s) and affiliation(s) appear below the title in capital and lower case letters. Papers with multiple authors and affiliations may require two or more lines for this information. Please note that papers should not be submitted blind; include the authors’ names on the PDF. TYPE-STYLE AND FONTS {#sec:typestyle} ==================== To achieve the best rendering both in printed proceedings and electronic proceedings, we strongly encourage you to use Times-Roman font. In addition, this will give the proceedings a more uniform look. Use a font that is no smaller than nine point type throughout the paper, including figure captions. In nine point type font, capital letters are 2 mm high. [**If you use the smallest point size, there should be no more than 3.2 lines/cm (8 lines/inch) vertically.**]{} This is a minimum spacing; 2.75 lines/cm (7 lines/inch) will make the paper much more readable. Larger type sizes require correspondingly larger vertical spacing. Please do not double-space your paper. TrueType or Postscript Type 1 fonts are preferred. The first paragraph in each section should not be indented, but all the following paragraphs within the section should be indented as these paragraphs demonstrate. MAJOR HEADINGS {#sec:majhead} ============== Major headings, for example, “1. Introduction”, should appear in all capital letters, bold face if possible, centered in the column, with one blank line before, and one blank line after. Use a period (“.”) after the heading number, not a colon. Subheadings {#ssec:subhead} ----------- Subheadings should appear in lower case (initial word capitalized) in boldface. They should start at the left margin on a separate line. ### Sub-subheadings {#sssec:subsubhead} Sub-subheadings, as in this paragraph, are discouraged. However, if you must use them, they should appear in lower case (initial word capitalized) and start at the left margin on a separate line, with paragraph text beginning on the following line. They should be in italics. PRINTING YOUR PAPER {#sec:print} =================== Print your properly formatted text on high-quality, 8.5 x 11-inch white printer paper. A4 paper is also acceptable, but please leave the extra 0.5 inch (12 mm) empty at the BOTTOM of the page and follow the top and left margins as specified. If the last page of your paper is only partially filled, arrange the columns so that they are evenly balanced if possible, rather than having one long column. In LaTeX, to start a new column (but not a new page) and help balance the last-page column lengths, you can use the command “$\backslash$pagebreak” as demonstrated on this page (see the LaTeX source below). PAGE NUMBERING {#sec:page} ============== Please do [**not**]{} paginate your paper. Page numbers, session numbers, and conference identification will be inserted when the paper is included in the proceedings. ILLUSTRATIONS, GRAPHS, AND PHOTOGRAPHS {#sec:illust} ====================================== Illustrations must appear within the designated margins. They may span the two columns. If possible, position illustrations at the top of columns, rather than in the middle or at the bottom. Caption and number every illustration. All halftone illustrations must be clear black and white prints. Colors may be used, but they should be selected so as to be readable when printed on a black-only printer. Since there are many ways, often incompatible, of including images (e.g., with experimental results) in a LaTeX document, below is an example of how to do this [@Lamp86]. FOOTNOTES {#sec:foot} ========= Use footnotes sparingly (or not at all!) and place them at the bottom of the column on the page on which they are referenced. Use Times 9-point type, single-spaced. To help your readers, avoid using footnotes altogether and include necessary peripheral observations in the text (within parentheses, if you prefer, as in this sentence). ![Example of placing a figure with experimental results.[]{data-label="fig:res"}](image1){width="8.5cm"} \(a) Result 1 ![Example of placing a figure with experimental results.[]{data-label="fig:res"}](image3){width="4.0cm"} \(b) Results 3 ![Example of placing a figure with experimental results.[]{data-label="fig:res"}](image4){width="4.0cm"} \(c) Result 4 COPYRIGHT FORMS {#sec:copyright} =============== You must include your fully completed, signed IEEE copyright release form when form when you submit your paper. We [**must**]{} have this form before your paper can be published in the proceedings. REFERENCES {#sec:ref} ========== List and number all bibliographical references at the end of the paper. The references can be numbered in alphabetic order or in order of appearance in the document. When referring to them in the text, type the corresponding reference number in square brackets as shown at the end of this sentence [@C2]. An additional final page (the fifth page, in most cases) is allowed, but must contain only references to the prior literature.
{ "pile_set_name": "ArXiv" }
--- author: - 'Phoolendra K. Mishra and Kristopher L. Kuhlman' bibliography: - 'review\_paper.bib' title: 'Unconfined Aquifer Flow Theory - from Dupuit to present' --- Abstract ======== Analytic and semi-analytic solution are often used by researchers and practicioners to estimate aquifer parameters from unconfined aquifer pumping tests. The non-linearities associated with unconfined (i.e., water table) aquifer tests makes their analysis more complex than confined tests. Although analytical solutions for unconfined flow began in the mid-1800s with Dupuit, Thiem was possibly the first to use them to estimate aquifer parameters from pumping tests in the early 1900s. In the 1950s, Boulton developed the first transient well test solution specialized to unconfined flow. By the 1970s Neuman had developed solutions considering both primary transient storage mechanisms (confined storage and delayed yield) without non-physical fitting parameters. In the last decade, research into developing unconfined aquifer test solutions has mostly focused on explicitly coupling the aquifer with the linearized vadose zone. Despite the many advanced solution methods available, there still exists a need for realism to accurately simulate real-world aquifer tests. Introduction ============ Pumping tests are widely used to obtain estimates of hydraulic parameters characterizing flow and transport processes in subsurface (e.g., @kruseman90 [@batu1998aquifer]). Hydraulic parameter estimates are often used in planning or engineering applications to predict flow and design of aquifer extraction or recharge systems. During a typical pumping test in a horizontally extensive aquifer, a well is pumped at constant volumetric flow rate and head observations are made through time at one or more locations. Pumping test data are presented as time-drawdown or distance-drawdown curves, which are fitted to idealized models to estimate aquifer hydraulic properties. For unconfined aquifers properties of interest include hydraulic conductivity, specific storage, specific yield, and possibly unsaturated flow parameters. When estimating aquifer properties using pumping test drawdown data, one can use a variety of analytical solutions involving different conceptualizations and simplifiying assumptions. Analytical solutions are impacted by their simplifiying assumptions, which limit their applicability to characterize certain types of unconfined aquifers. This review presents the historical evolution of the scientific and engineering thoughts concerning groundwater flow towards a pumping well in unconfined aquifers (also referred to variously as gravity, phreatic, or water table aquifers) from the steady-state solutions of Dupuit to the coupled transient saturated-unsaturated solutions. Although it is sometimes necessary to use gridded numerical models in highly irregular or heterogeneous systems, here we limit our consideration to analytically derived solutions. Early Well Test Solutions ========================= Dupuit’s Steady-State Finite-Domain Solutions --------------------------------------------- [@dupuit1857] considered steady-state radial flow to a well pumping at constant volumetric flowrate $Q$ \[L$^3$/T\] in a horizontal homogeneous confined aquifer of thickness $b$ \[L\]. He used Darcy’s law [@darcy1856] to express the velocity of groundwater flow $u$ \[L/T\] in terms of radial hydraulic head gradient $\left(\partial h/\partial r\right)$ as $$\label{darcys-law} u=K\frac{\partial h}{\partial r},$$ where $K=kg/\nu$ is hydraulic conductivity \[L/T\], $k$ is formation permeability \[L$^2$\], $g$ is the gravitational constant \[L/T$^2$\], $\nu$ is fluid kinematic viscosity \[L$^2$/T\], $h=\psi+z$ is hydraulic head \[L\], $\psi$ is gage pressure head \[L\], and $z$ is elevation above an arbitrary datum \[L\]. Darcy derived a form equivalent to for one-dimensional flow through sand-packed pipes. Dupuit was the first to apply to converging flow by combining it with mass conservation $Q=\left(2\pi rb \right)u$ across a cylindrical shell concentric with the well, leading to $$\label{dupuit_1} Q=K\left( 2\pi rb\right)\frac{\partial h}{\partial r}.$$ Integrating between two radial distances $r_1$ and $r_2$ from the pumping well, Dupuit evaluated the confined steady-state head difference between the two points as $$\label{dupuit_confined} h(r_{2})-h(r_{1})=\frac{Q}{2\pi Kb}\log\left( \frac{r_2}{r_1}\right).$$ This is the solution for flow to a well at the center of a circular island, where a constant head condition is applied at the edge of the island ($r_2$). @dupuit1857 also derived a radial flow solution for unconfined aquifers by neglecting the vertical flow component. Following a similar approach to confined aquifers, @dupuit1857 estimated the steady-state head difference between two distances from the pumping well for unconfined aquifers as $$\label{dupuit_unconfined} h^{2}(r_{2}) - h^{2} (r_{1}) = \frac{Q}{\pi K} \log\left( \frac{r_2}{r_1}\right).$$ These two solutions are only strictly valid for finite domains; when applied to domains without a physical boundary at $r_2$, the outer radius essentially becomes a fitting parameter. The solutions are also used in radially infinite systems under pseudo-static conditions, when the shape of the water table does not change with time. Equations and are equivalent when $b$ in is average head $\left( h(r_1)+h(r_2)\right)/2$. In developing , [@dupuit1857] used the following assumptions (now commonly called the Dupuit assumptions) in context of unconfined aquifers: - the aquifer bottom is a horizontal plane; - groundwater flow toward the pumping wells is horizontal with no vertical hydraulic gradient component; - the horizontal component of the hydraulic gradient is constant with depth and equal to the water table slope; and - there is no seepage face at the borehole. These assumptions are one of the main approaches to simplifying the unconfined flow problem and making it analytically tractable. In the unconfined flow problem both the head and the location of the water table are unknowns; the Dupuit assumptions eliminate one of the unknowns. Historical Developments after Dupuit ------------------------------------ @narasimhan98 and @vries2007 give detailed historical accounts of groundwater hydrology and soil mechanics; only history relevant to well test analysis is given here. @forchheimer1886 first recognized the Laplace equation $\nabla^2 h = 0$ governed two-dimensional steady confined groundwater flow (to which is a solution), allowing analogies to be drawn between groundwater flow and steady-state heat conduction, including the first application of conformal mapping to solve a groundwater flow problem. @slichter1898 also arrived at the Laplace equation for groundwater flow, and was the first to account for a vertical flow component. Utilizing Dupuit’s assumptions, @forchheimer1898 developed the steady-state unconfined differential equation (to which is a solution), $\nabla^2 h^2=0$. @boussinesq1904 first gave the transient version of the confined groundwater flow equation $\alpha_s \nabla^2 h = \partial h/\partial t$ (where $\alpha_s=K/S_s$ is hydraulic diffusivity \[L$^2$/T\] and $S_s$ is specific storage \[1/L\]), based upon analogy with transient heat conduction. In Prague, @thiem1906 was possibly the first to use for estimating $K$ from pumping tests with multiple observation wells [@simmons2008]. Equation (commonly called the Thiem equation) was tested in the 1930’s both in the field (@wenzel1932recent performed a 48-hour pumping test with 80 observation wells in Grand Island, Nebraska) and in the laboratory (@wyckoff1932dupuitflow developed a 15-degree unconfined wedge sand tank to simulate converging flow). Both found the steady-state solution lacking in ability to consistently estimate aquifer parameters. @wenzel1942 developed several complex averaging approaches (e.g., the “limiting” and “gradient” formulas) to attempt to consistently estimate $K$ using steady-state confined equations for a finite system from transient unconfined data. @muskat1932partialpen considered partial-penetration effects in steady-state flow to wells, discussing the nature of errors associated with assumption of uniform flux across the well screen in a partially penetrating well. Muskat’s textbook on porous media flow [@muskat1937book] summarized much of what was known in hydrology and petroleum reservoir engineering around the time of the next major advance in well test solutions by Theis. Confined Transient Flow ----------------------- [@theis1935] utilized the analogy between transient groundwater flow and heat conduction to develop an analytical solution for confined transient flow to a pumping well (see Figure \[fig:diagram\]). He initially applied his solution to unconfined flow, assuming instantaneous drainage due to water table movement. The analytical solution was based on a Green’s function heat conduction solution in an infinite axis-symmetric slab due to an instantaneous line heat source or sink [@carslaw1921]. With the aid of mathematician Clarence Lubin, Theis extended the heat conduction solution to a continuous source, motivated to better explain the results of pumping tests like the 1931 test in Grand Island. [@theis1935] gave an expression for drawdown due to pumping a well at rate $Q$ in a homogeneous, isotropic confined aquifer of infinite radial extent as an exponential integral $$\label{theis} s(r,t)=\frac{Q}{4\pi T}\int_{r^2 /(4 \alpha_s t)}^{\infty}\frac{e^{-u}}{u} \;\mathrm{d}u,$$ where $s=h_0(r)-h(t,r)$ is drawdown, $h_0$ is pre-test hydraulic head, $T=Kb$ is transmissivity, and $S=S_s b$ is storativity. Equation is a solution to the diffusion equation, with zero-drawdown inital and far-field conditions, $$s(r,t=0) = s(r\rightarrow \infty,t) = 0.$$ The pumping well was approximated by a line sink (zero radius), and the source term assigned there was based upon , $${\label{boulton_sink_well}} \lim_{r \rightarrow 0} r\frac{\partial s}{\partial r}=-\frac{Q}{2 \pi T}.$$ ![Unconfined well test diagram[]{data-label="fig:diagram"}](well-diagram.eps){width="60.00000%"} Although the transient governing equation was known through analogy with heat conduction, the transient storage mechanism (analogous to specific heat capacity) was not completely understood. Unconfined aquifer tests were known to experience slower drawdown than confined tests, due to water supplied by dewatering the zone near the water table, which is related to the formation specific yield (porosity less residual water). @muskat1934transient and [@hurst1934unsteady] derived solutions to confined transient radial flow problems for finite domains, but attributed transient storage solely to fluid compressibility. @jacob1940 derived the diffusion equation for groundwater flow in compressible elastic confined aquifers, using mass conservation and Darcy’s law, without recourse to analogy with heat conduction. @terzaghi1923 developed a one-dimensional consolidation theory which only considered the compressibility of the soil (in his case a clay), unknown at the time to most hydrologists [@batu1998aquifer]. @meinzer1928 studied regional drawdown in North Dakota, proposing the modern storage mechanism related to both aquifer compaction and the compressiblity of water. @jacob1940 formally showed $S_s=\rho_w g(\beta_p + n\beta_w)$, where $\rho_w$ and $\beta_w$ are fluid density \[M/L$^3$\] and compressibility \[LT$^2$/M\], $n$ is dimensionless porosity, and $\beta_p$ is formation bulk compressibility. The axis-symmetric diffusion equation in radial coordinates is $$\label{diffusion} \frac{\partial ^2 s}{\partial r^2}+\frac{1}{r}\frac{\partial s}{\partial r}= \frac{1}{\alpha_s}\frac{\partial s}{\partial t}.$$ When deriving analytical expressions, the governing equation is commonly made dimensionless to simplify presentation of results. For flow to a pumping well, it is convenient to use $L_C = b$ as a characteristic length, $T_C = Sb^2/T$ as a characteristic time, and $H_C = Q/(4 \pi T)$ as a characteristic head. The dimensionless diffusion equation is $$\label{diffusion} \frac{\partial ^2 s_D}{\partial r_D^2}+\frac{1}{r_D}\frac{\partial s_D}{\partial r_D}=\frac{\partial s_D}{\partial t_D},$$ where $r_D=r/L_C$, $s_D=s/H_c$, and $t_D=t/T_C$ are scaled by characteristic quantities. The [@theis1935] solution was developed for field application to estimate aquifer hydraulic properties, but it saw limited use because it was difficult to compute the exponential integral for arbitrary inputs. [@wenzel1942] proposed a type-curve method that enabled graphical application of the [@theis1935] solution to field data. [@cooperjacob1946] suggested for large values of $t_D$ ($t_D \geq 25$), the infinite integral in the [@theis1935] solution can be approximated as $$\label{JacobCooper} s_D(t_D,r_D) = \int_{r^2/(4\alpha_st)}^{\infty} \frac{e^{-u}}{u} \; \mathrm{d}u \approx \log_e \left(\frac{4 Tt}{r^2S}\right) - \gamma$$ where $\gamma \approx 0.57722$ is the Euler-Mascheroni constant. This leads to Jacob and Cooper’s straight-line simplification $$\Delta s \approx 2.3 \frac{Q}{4 \pi T}$$ where $\Delta s$ is the drawdown over one log-cycle (base 10) of time. The intercept of the straight-line approximation is related to $S$ through This approximation made estimating hydraulic parameters much simpler at large $t_D$. @hantush1961 later extended Theis’ confined solution for partially penetrating wells. Observed Time-drawdown Curve ---------------------------- Before the time-dependent solution of @theis1935, distance drawdown was the diagnostic plot for aquifer test data. Detailed distance-drawdown plots require many observation locations (e.g., the 80 observation wells of @wenzel1936TheimTest). Re-analyzing results of the unconfined pumping test in Grand Island, [@wenzel1942] noticed that the [@theis1935] solution gave inconsistent estimates of $S_s$ and $K$, attributed to the delay in the yield of water from storage as the water table fell. The @theis1935 solution corresponds to the Dupuit assumptions for unconfined flow, and can only re-create the a portion of observed unconfined time-drawdown profiles (either late or early). The effect of the water table must be taken into account through a boundary condition or source term in the governing equation to reproduce observed behavior in unconfined pumping tests. ![Drawdown data from Cape Cod [@moenchetal2001], observation well F377-037. Upper dashed curve is confined model of @hantush1961 with $S=S_sb$, lower dotted curve is same with $S=S_sb + S_y$. Solid curve is unconfined model of @neuman1974 using $S_y=0.23$.[]{data-label="fig:capecod"}](cape_cod_F377-037.eps){width="80.00000%"} [@walton1960] recognized three distinct segments characterizing different release mechanisms on time-drawdown curve under water table conditions (see Figure \[fig:capecod\]). A log-log time-drawdown plot in an unconfined aquifer has a characteristic shape consisting of a steep early-time segment, a flatter intermediate segment and a steeper late-time segment. The early segment behaves like the @theis1935 solution with $S=S_s b$ (water release due to bulk medium relaxation), the late segment behaves like the @theis1935 solution with $S=S_s b + S_y$ [@gambolati1976transient] (water release due to water table drop), and the intermediate segment represents a transition between the two. Distance-drawdown plots from unconfined aquifer tests do not show a similar inflection or change in slope, and do not produce good estimates of storage parameters. Early Unconfined Well Test Solutions ==================================== Moving Water Table Solutions Without Confined Storage ----------------------------------------------------- The [@theis1935] solution for confined aquifers can only reproduce either the early or late segments of the unconfined time-drawdown curve (see Figure \[fig:capecod\]). [@boulton1954a] suggested it is theoretically unsound to use the @theis1935 solution for unconfined flow because it does not account for vertical flow to the pumping well. He proposed a new mechanism for flow towards a fully penetrating pumping well under unconfined conditions. His formulation assumed flow is governed by $\nabla^2 s = 0$, with transient effects incorporated through the water table boundary condition. He treated the water table (where $\psi=0$, located at $z=\xi$ above the base of the aquifer) as a moving material boundary subject to the condition $h\left( r,z=\xi,t\right)=z$. He considered the water table without recharge to be comprised of a constant set of particles, leading to the kinematic boundary condition $$\label{dynamic} \frac{D}{Dt}\left(h - z \right) = 0$$ which is a statement of conservation of mass, for an incompressible fluid. @boulton1954a considered the Darcy velocity of the water table as $u_z=-\frac{K_z}{S_y}\frac{\partial h}{\partial z}$ and $u_r=-\frac{K_r}{S_y}\frac{\partial h}{\partial r}$, and expressed the total derivative as $$\label{material_derivative} \frac{D}{Dt}=\frac{\partial}{\partial t}- \frac{K_r}{S_y}\frac{\partial h}{\partial r}\frac{\partial}{\partial r}- \frac{K_z}{S_y}\frac{\partial h}{\partial z}\frac{\partial}{\partial z},$$ where $K_r$ and $K_z$ are radial and vertical hydraulic conductivity components. Using , the kinematic boundary condition in terms of drawdown is $${\label{free_surface}} \frac{\partial s}{\partial t}-\frac{K_r}{S_y} \left( \frac{\partial s}{\partial r} \right)^2- \frac{K_z}{S_y} \left( \frac{\partial s}{\partial z} \right)^2=- \frac{K_z}{S_y}\frac{\partial s}{\partial z}.$$ @boulton1954a utilized the wellbore and far-field boundary conditions of @theis1935. He also considered the aquifer rests on an impermeable flat horizontal boundary $\left. \partial h/\partial z \right|_{z=0}= 0$; this was also inferred by @theis1935 because of his two-dimensional radial flow assumption. [@dagan1967] extended Boulton’s water table solution to the partially penetrating case by replacing the wellbore boundary condition with $$\lim_{r \rightarrow 0} r\frac{\partial s}{\partial r}= \begin{cases}\frac{Q}{2\pi K (\ell - d)} & b-\ell < z < b-d \\ 0 & \text{otherwise}\end{cases},$$ where $\ell$ and $d$ are the upper and lower boundaries of the pumping well screen, as measured from the initial top of the aquifer. The two sources of non-linearity in the unconfined problem are: 1) the boundary condition is applied at the water table, the location of which is unknown *a priori*; 2) the boundary condition applied on the water table includes $s^2$ terms. In order to solve this non-linear problem both Boulton and Dagan linearized it by disregarding second order components in the free-surface boundary condition and forcing the free surface to stay at its initial position, yielding $$\label{boulton_linear_water_table} \frac{\partial s}{\partial t}=- \frac{K_z}{S_y}\frac{\partial s}{\partial z} \qquad\qquad z=h_0,$$ which now has no horizontal flux component after neglecting second-order terms. Equation can be written in non-dimensional form as $$\label{dimensionless_MWT_bc} \frac{\partial s_D}{\partial t_D}=- K_D \sigma^{\ast} \frac{\partial s_D}{\partial z_D} \qquad\qquad z_D=1,$$ where $K_D=K_z/K_r$ is the dimensionless anisotropy ratio and $\sigma^{\ast}=S/S_y$ is the dimensionless storage ratio. Both [@boulton1954a] and [@dagan1967] solutions reproduce the intermediate and late segments of the typical unconfined time-drawdown curve, but neither of them reproduces the early segment of the curve. Both are solutions to the Laplace equation, and therefore disregard confined aquifer storage, causing pressure pulses to propagate instantaneously through the saturated zone. Both solutions exhibit an instantaneous step-like increase in drawdown when pumping starts. Delayed Yield Unconfined Response --------------------------------- [@boulton1954b] extended Theis’ transient confined theory to include the effect of delayed yield due to movement of the water table in unconfined aquifers. Boulton’s proposed solutions ([-@boulton1954b; -@boulton1963]) reproduce all three segments of the unconfined time-drawdown curve. In his formulation of delayed yield, he assumed as the water table falls water is released from storage (through drainage) gradually, rather than instantaneously as in the free-surface solutions of [@boulton1954a] and [@dagan1967]. This approach yielded an integro-differential flow equation in terms of vertically averaged drawdown $s^\ast$ as $$\label{boulton_solution} \frac{\partial ^2 s^\ast}{\partial r^2}+ \frac{1}{r}\frac{\partial s^\ast}{\partial r}= \left[\frac{S}{T}\frac{\partial s^\ast}{\partial t} \right]+ \left\lbrace\alpha S_y \int _0^t \frac{\partial s^\ast}{\partial \tau} e^{-\alpha (t-\tau )}\;\mathrm{d}\tau \right \rbrace$$ which Boulton linearized by treating $T$ as constant. The term in square brackets is instantaneous confined storage, the term in braces is a convolution integral representing storage released gradually since pumping began, due to water table decline. [@boulton1963] showed the time when delayed yield effects become negligible is approximately equal to $\frac{1}{\alpha}$, leading to the term “delay index”. [@prickett1965] used this concept and through analysis of large amount of field drawdown data with [@boulton1963] solution, he established an empirical relationship between the delay index and physical aquifer properties. Prickett proposed a methodology for estimation of $S$, $S_y$, $K$, and $\alpha$ of unconfined aquifers by analyzing pumping tests with the [@boulton1963] solution. Although Boulton’s model was able to reproduce all three segment of the unconfined time-drawdown curve, it failed to explain the physical mechanism of the delayed yield process because of the non-physical nature of the “delay index” in the [@boulton1963] solution. [@streltsova1972a] developed an approximate solution for the decline of the water table and $s^\ast$ in fully penetrating pumping and observation wells. Like @boulton1954b, she treated the water table as a sharp material boundary, writing the two-dimensional depth-averaged flow equation as $$\label{streltsova_equation} \frac{\partial ^2 s^\ast}{\partial r^2}+ \frac{1}{r}\frac{\partial s^\ast}{\partial r}= \frac{S}{T}\left(\frac{\partial s^\ast}{\partial t}- \frac{\partial \xi}{\partial t} \right).$$ The rate of water table decline was assumed to be linearly proportional to the difference between the water table elevation $\xi$ and the vertically averaged head $b-s^\ast$, $$\label{streltsova_watertable} \frac{\partial \xi}{\partial t}=\frac{K_z}{S_yb_z} \left( s^\ast-b+\xi \right)$$ where $b_z=b/3$ is an effective aquifer thickness over which water table recharge is distributed into the deep aquifer. Equation can be viewed as an approximation to the zero-order linearized free-surface boundary condition of [@boulton1954a] and [@dagan1967]. Streltsova considered the initial condition $\xi (r,t=0)=b$ and used the same boundary condition at the pumping well and the outer boundary $(r\rightarrow \infty )$ used by [@theis1935] and [@boulton1963]. Equation has the solution [@streltsova1972b] $$\label{strelstsova_sol} \frac{\partial \xi}{\partial t} = -\alpha_T \int _0^t e^{-\alpha_T (t-\tau)} \frac{\partial s^\ast}{\partial \tau}\;\mathrm{d}\tau$$ where $\alpha_T = K_z/(S_yb_z)$. Substituting into produces solution of [@boulton1954b; @boulton1963]; the two solutions are equivalent. Boulton’s delayed yield theory (like that of Streltsova) does not account for flow in unsaturated zone but instead treats water table as material boundary moving vertically downward under influence of gravity. [@streltsova1973] used field data collected by [@meyer1962] to demonstrate unsaturated flow had virtually no impact on the observed delayed process. Although Streltsova’s solution related Boulton’s delay index to physical aquifer properties, it was later found to be a function of $r$ [@neuman1975; @herrera1978]. The delayed yield solutions of Boulton and Streltsova do not account for vertical flow within the unconfined aquifer through simplifying assumptions; they cannot be extended to account for partially penetrating pumping and observation wells. Prickett’s pumping test in the vicinity of Lawrenceville, Illinois [@prickett1965] showed that specific storage in unconfined aquifers can be much greater than typically observed values in confined aquifers – possibly due to entrapped air bubbles or poorly consolidated shallow sediments. It is clear the elastic properties of unconfined aquifers are too important to be disregarded. Delayed Water Table Unconfined Response --------------------------------------- Boulton’s ([-@boulton1954b; -@boulton1963]) models encountered conceptual difficulty explaining the physical mechanism of water release from storage in unconfined aquifers. [@neuman1972] presented a physically based mathematical model that treated the unconfined aquifer as compressible (like @boulton1954b [@boulton1963] and @streltsova1972a [@streltsova1972b]) and the water table as a moving material boundary (like @boulton1954a and @dagan1967). In Neuman’s approach aquifer delayed response was caused by physical water table movement, he therefore proposed to replace the phrase “delayed yield” by “delayed water table response”. [@neuman1972] replaced the Laplace equation of [@boulton1954a] and [@dagan1967] by the diffusion equation; in dimensionless form it is $$\label{neuman_diffusion} \frac{\partial ^2 s_D}{\partial r_D^2}+ \frac{1}{r_D}\frac{\partial s_D}{\partial r_D}+ K_D\frac{\partial ^2s_D}{\partial z_D^2} = \frac{\partial s_D}{\partial t_D}.$$ Like @boulton1954a and @dagan1967, Neuman treated the water table as a moving material boundary, linearized it (using ), and treated the anisotropic aquifer as three-dimensional axis-symmetric. Like @dagan1967, @neuman1974 accounted for partial penetration. By including confined storage in the governing equation , Neuman was able to reproduce all three parts of the observed unconfined time-drawdown curve and produce parameter estimates (including the ability to estimate $K_z$) very similar to the delayed yield models. Compared to the delay index models, Neuman’s solution produced similar fits to data (often underestimating $S_y$, though), but [@neuman1975; @neuman1979] questioned the physical nature of Boulton’s delay index. He performed a regression fit between the @boulton1954b and @neuman1972 solutions, resulting in the relationship $$\label{alpha-regression} \alpha = \frac{K_z}{S_yb}\left[3.063-0.567 \log\left( \frac{K_Dr^2}{b^2}\right) \right]$$ demonstrating $\alpha$ decreases linearly with $\log r$ and is therefore not a characteristic aquifer constant. When ignoring the logarithmic term in the relationship $\alpha=3K_z/(S_yb)$ proposed by [@streltsova1972a] is approximately recovered. After comparative analysis of various methods for determination of specific yield, [@neuman1987] concluded the water table response to pumping is a much faster phenomenon than drainage in the unsaturated zone above it. @malama2011 recently proposed an alternative linearization of , approximately including the effects of the neglected second-order terms, leading to the alternative water table boundary condition of $$\label{malama} S_y \frac{\partial s}{\partial t} =- K_z \left( \frac{\partial s}{\partial z} + \beta \frac{\partial^2 s}{\partial z^2} \right) \qquad z=h_0$$ where $\beta$ is a linearization coefficient \[L\]. The parameter $\beta$ provides additional adjustment of the shape of the intermediate portion of the time-drawdown curve (beyond adjustments possible with $K_D$ and $\sigma^\ast$ alone), leading to improved estimates of $S_y$. When $\beta=0$ simplifies to . Hybrid Water Table Boundary Condition ------------------------------------- The solution of [@neuman1972; @neuman1974] was accepted by many hydrologists “as the preferred model ostensibly because it appears to make the fewest simplifying assumptions” [@moenchetal2001]. Despite acceptance, [@nwankwor1984] and [@moench1995] pointed out that significant difference might exist between measured and model-predicted drawdowns, especially at locations near the water table, leading to significantly underestimated $S_y$ using Neuman’s models (e.g., see Figure \[fig:capecod\]). @moench1995 attributed the inability of Neuman’s models to give reasonable estimates of $S_y$ and capture this observed behavior near the water table due to the later’s disregard of “gradual drainage”. In an attempt to resolve this problem, [@moench1995] replaced the instantaneous moving water table boundary condition used by Neuman with one containing a @boulton1954b delayed yield convolution integral, $$\label{moench_hybrid} \int _0^t\frac{\partial s}{\partial \tau} \sum _{m=1}^M \alpha _m e^{-\alpha _m (t-\tau )}\;\mathrm{d}\tau =- \frac{K_z}{S_y}\frac{\partial s}{\partial z}$$ This hybrid boundary condition ($M=1$ in @moench1995) included the convolution source term @boulton1954b [@boulton1963] and @streltsova1972a [@streltsova1972b] used in their depth-averaged governing flow equations. In addition to this new boundary condition, [@moench1995] included a finite radius pumping well with wellbore storage, conceptually similar to how @papadopulosandcooper1967 modified the solution of @theis1935. In all other respects, his definition of the problem was similar to [@neuman1974]. Moench’s solution resulted in improved fits to experimental data and produced more realistic estimates of specific yield [@moenchetal2001], including the use of multiple delay parameters $\alpha_m$ [@moench2003]. @moenchetal2001 used with $M = 3$ to estimate hydraulic parameters in the unconfined aquifer at Cape Cod. They showed that $M=3$ enabled a better fit to the observed drawdown data than obtained by $M<3$ or the model of [@neuman1974]. Similar to the parameter $\alpha $ in Boulton’s model, the physical meaning of $\alpha _m$ is not clear. Unconfined Solutions Considering Unsaturated Flow ================================================= As an alternative to linearizing the water table condition of @boulton1954a, the unsaturated zone can be explicitly included. The non-linearity of unsaturated flow is substituted for the non-linearity of . By considering the vadose zone, the water table is internal to the domain, rather than a boundary condition. The model-data misfit in Figure \[fig:capecod\] at “late intermediate” time is one of the motivations for considering the mechanisms of delayed yield and the effects of the unsaturated zone. Unsaturated Flow Without Confined Aquifer Storage ------------------------------------------------- [@kroszynskidagan1975] were the first to account analytically for the effect of the unsaturated zone on aquifer drawdown. They extended the solution of [@dagan1967] by accounting for unsaturated flow above the water table. They used Richards’ equation for axis-symmetric unsaturated flow in a vadose zone of thickness $L$ $$\label{richards} K_r\frac{1}{r}\frac{\partial}{\partial r}\left( k(\psi )r\frac{\partial \sigma}{\partial r}\right)+ K_z\frac{\partial}{\partial z}\left( k(\psi )\frac{\partial \sigma}{\partial z}\right) = C(\psi)\frac{\partial \sigma }{\partial t} \qquad \xi < z <b+L$$ where $\sigma = b + \psi_a - h$ is unsaturated zone drawdown \[L\], $\psi _a$ is air-entry pressure head \[L\], $0 \leq k(\psi )\leq 1$ is dimensionless relative hydraulic conductivity, $C(\psi)=d\theta/d \psi$ is the moisture retention curve \[1/L\], and $\theta$ is dimensionless volumetric water content (see inset in Figure \[fig:diagram\]). They assumed flow in the underlying saturated zone was governed by the Laplace equation (like @dagan1967). The saturated and unsaturated flow equations were coupled through interface conditions at the water table expressing continuity of hydraulic heads and normal groundwater fluxes, $$\begin{aligned} s=\sigma\qquad \nabla s \cdot \textbf{n}=\nabla \sigma \cdot \textbf{n} \qquad z= \xi\end{aligned}$$ where $\textbf{n}$ is the unit vector perpendicular to the water table. To solve the unsaturated flow equation , @kroszynskidagan1975 linearized by adopting the [@gardner1958] exponential model for the relative hydraulic conductivity, $k(\psi )=e^{\kappa_a (\psi -\psi_a)}$, where $\kappa_a$ is the sorptive number \[1/L\] (related to pore size). They adopted the same exponential form for the moisture capacity model, $\theta (\psi )=e^{\kappa_k (\psi -\psi_k)}$, where $\psi_k$ is the pressure at which $k(\psi)=1$, $\kappa_a=\kappa_k$, and $\psi_a=\psi_k$, leading to the simplified form $C(\psi)=S_y\kappa_a e^{\kappa_a (\psi -\psi_a)}$. In the limit as $\kappa_k=\kappa_a \rightarrow \infty$ their solution reduces to that of [@dagan1967]. The relationship between pressure head and water content is a step function. [@kroszynskidagan1975] took unsaturated flow above the water table into account but ignored the effects of confined aquifer storage, leading to early-time step-change behavior similar to @boulton1954a and @dagan1967. Increasingly Realistic Saturated-Unsaturated Well Test Models ------------------------------------------------------------- [@mathiasbutler2006] combined the confined aquifer flow equation with a one-dimensional linearized version of for a vadose zone of finite thickness. Their water table was treated as a fixed boundary with known flow conditions, decoupling the unsaturated and saturated solutions at the water table. Although they only considered a one-dimensional unsaturated zone, they included the additional flexibility provided by different exponents $(\kappa_a \neq \kappa_k)$. [@mathiasbutler2006] did not consider a partially penetrating well, but they did note the possibility of accounting for it in principle by incorporating their uncoupled drainage function in the solution of @moench1997, which considers a partially penetrating well of finite radius. [@tartakovskyneuman2007] similarly combined the confined aquifer flow equation , but with the original axis-symmetric form of considered by @kroszynskidagan1975. Also like @kroszynskidagan1975, their unsaturated zone was characterized by a single exponent $\kappa_a=\kappa_k$ and reference pressure head $\psi_a=\psi_k$. Unlike @kroszynskidagan1975 and @mathiasbutler2006, [@tartakovskyneuman2007] assumed an infinitely thick unsaturated zone. [@tartakovskyneuman2007] demonstrated flow in the unsaturated zone is not strictly vertical. Numerical simulations by @moench2008 showed groundwater movement in the capillary fringe is more horizontal than vertical. [@mathiasbutler2006] and @moench2008 showed that using the same exponents and reference pressure heads for effective saturation and relative permeability decreases model flexibility and underestimates $S_y$. @moench2008 predicted an extended form of @tartakovskyneuman2007 with two separate exponents, a finite unsaturated zone, and wellbore storage would likely produce more physically realistic estimates of $S_y$. [@mishraneuman2010] developed a new generalization of the solution of [@tartakovskyneuman2007] that characterized relative hydraulic conductivity and water content using $\kappa_a \neq \kappa_k$, $\psi_a \neq \psi_k$ and a finitely thick unsaturated zone. @mishraneuman2010 validated their solution against numerical simulations of drawdown in a synthetic aquifer with unsaturated properties given by the model of [@vangenuchten1980]. They also estimated aquifer parameters from Cape Cod drawdown data [@moenchetal2001], comparing estimated @vangenuchten1980 parameters with laboratory values [@maceetal1998]. [@mishraneuman2011] further extended their [-@mishraneuman2010] solution to include a finite-diameter pumping well with storage. @mishraneuman2010 [@mishraneuman2011] were the first to estimate non-exponential model unsaturated aquifer properties from pumping test data, by curve-fitting the exponential model to the [@vangenuchten1980] model. Analyzing pumping test data of @moenchetal2001 (Cape Cod, Massachusetts) and @nwankwor1984 [@nwankwor1992] (Borden, Canada), they estimated unsaturated flow parameters similar to laboratory-estimated values for the same soils. Future Challenges ================= The conceptualization of groundwater flow during unconfined pumping tests has been a challenging task that has spurred substantial theoretical research in the field hydrogeology for decades. Unconfined flow to a well is non-linear in multiple ways, and the application of analytical solutions has required utilization of advanced mathematical tools. There are still many additional challenges to be addressed related to unconfined aquifer pumping tests, including: - Hysteretic effects of unsaturated flow. Different exponents and reference pressures are needed during drainage and recharge events, complicating simple superposition needed to handle multiple pumping wells, variable pumping rates, or analysis of recovery data. - Collecting different data types. Validation of existing models and motivating development of more realistic ones depends on more than just saturated zone head data. Other data types include vadose zone water content [@meyer1962], and hydrogeophysical data like microgravity [@damiata2006] or streaming potentials [@malama2009]. - Moving water table position. All solutions since @boulton1954a assume the water table is fixed horizontal $\xi(r,t)=h_0$ during the entire test, even close to the pumping well where large drawdown is often observed. Iterative numerical solutions can accommodate this, but this has not been included in an analytical solution. - Physically realistic partial penetration. Well test solutions suffer from the complication related to the unknown distribution of flux across the well screen. Commonly, the flux distribution is simply assumed constant, but it is known that flux will be higher near the ends of the screened interval that are not coincident with the aquifer boundaries. - Dynamic water table boundary condition. A large increase in complexity comes from explicitly including unsaturated flow in unconfined solutions. The kinematic boundary condition expresses mass conservation due to water table decline. Including an analogous dynamic boundary condition based on a force balance (capillarity vs. gravity) may include sufficient effects of unsaturated flow, without the complexity associated with the complete unsaturated zone solution. - Heterogeneity. In real-world tests heterogeneity is present at multiple scales. Large-scale heterogeneity (e.g., faults or rivers) can sometimes be accounted in analytical solutions using the method of images, or other types of superpostion. A stochastic approach [@neuman2004type] could alternatively be developed to estimate random unconfined aquifer parameter distribution parameters. Despite advances in considering physically realistic unconfined flow, most real-world unconfined tests (e.g., @wenzel1942, @nwankwor1984 [@nwankwor1992], or @moenchetal2001) exhibit non-classical behavior that deviates from the early-intermediate-late behavior predicted by the models summarized here. We must continue to strive to include physically relevant processes and representatively linearize non-linear phenomena, to better understand, simulate and predict unconfined flow processes. Acknowledgments {#acknowledgments .unnumbered} =============== This research was partially funded by the Environmental Programs Directorate of the Los Alamos National Laboratory. Los Alamos National Laboratory is a multi-program laboratory managed and operated by Los Alamos National Security (LANS) Inc. for the U.S. Department of Energy’s National Nuclear Security Administration under contract DE-AC52-06NA25396. Sandia National Laboratories is a multi-program laboratory managed and operated by Sandia Corporation, a wholly owned subsidiary of Lockheed Martin Corporation, for the U.S. Department of Energy’s National Nuclear Security Administration under contract DE-AC04-94AL85000.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We calculate the quantum noise limited displacement sensitivity of a Michelson-Fabry-Perot (MFP) with detuned cavities, followed by phase-sensitive homodyne detection. We show that the standard quantum limit can be surpassed even with resonant cavities and without any signal-recycling mirror nor additional cavities. Indeed, thanks to the homodyne detection, the output field quadrature can be chosen in such a way to cancel the effect of input amplitude fluctuations, i.e., eliminating the force noise. With detuned cavities, the modified opto-mechanical susceptivity allows to reach unlimited sensitivity for large enough (yet finite) optical power. Our expressions include mirror losses and cavity delay effect, for a realistic comparison with experiments. Our study is particularly devoted to gravitational wave detectors and we consider both an interferometer with free-falling mirrors, and a MFP as readout for a massive detector. In the latter case, the sensitivity curve of the recently conceived ’DUAL’ detector, based on two acoustic modes, is obtained.' author: - 'J. Belfi' - 'F. Marin' title: 'Sensitivity below the standard quantum limit in gravitational wave detectors with Michelson-Fabry-Perot readout ' --- Introduction ============ The sensitivity of interferometers used for the measurement of strain or displacement is commonly referred to a so-called standard quantum limit (SQL), calculated considering independent fluctuations of the radiation pressure acting on the sensing mirrors and of the detected light. Both noise terms are derived from the quantum fluctuations of the electromagnetic field. Several studies, starting at least from the beginning of the 80ies, have shown that an apparatus can beat the SQL, and accurate quantum calculations are in general necessary to find the actual sensitivity. The seminal work by Unruh [@Unruh] shows that the SQL can be surpassed if quantum correlation characterizes the measuring electromagnetic field, and Jaekel and Reynaud [@Jaekel] show that in this case an ultimate limit is imposed by the dissipative part of the mechanical susceptivity. Several schemes have been proposed to reach the goal of a sensitivity beyond the SQL, often using additional optical cavities, such as the quantum locking [@Courty], the local meter [@Brag97; @Kha02; @Dan06], the detuned signal recycling (studied firstly in Ref. [@Meers88] and later analyzed with a deeper attention to quantum noise in Refs. [@Buonanno; @Harms]). It should be noticed that a simple detuning from resonance of an optical cavity allows to rotate the field quadratures [@Levenson; @Lugiato] and create a correlation between amplitude and phase fluctuations, that are commonly related respectively to radiation pressure and detected field fluctuations [@Caves]. This effect is exploited in several proposals of schemes for the generation of ponderomotive squeezing [@mancini94; @Fabre; @Corbitt] and for quantum non-demolition measurements [@Heidmann97]. A recent work by Arcizet *et al.* [@Arcizet] clearly explains that a detuned Fabry-Perot cavity can indeed provide a sensitivity well beyond the SQL, with a frequency behavior very similar to that foreseen for interferometers with signal-recycling mirror [@Buonanno; @Harms]. Most of the mentioned studies have been stimulated by the development of large interferometric detectors of gravitational waves (gw). Recently, a new class of gw detectors has been conceived, based on huge masses kept at cryogenic temperature and called DUAL detectors [@Cerdonio; @Bonaldi; @Marin]. Differently from previous massive cryogenic antennas (such as Weber bars), the DUAL system do not exploit particularly a mechanical resonance of a solid body, but it takes advantage of elastic forces to achieve a useful sensitivity in a wide frequency range. At this purpose, it has to give up to the usual resonant mechanical amplifier, and it needs a very sensitive readout. One possibility is using a Michelson interferometer with suitable Fabry-Perot cavities in the two arms (Michelson-Fabry-Perot, MFP)[@Marin1]. The readout would be in principle similar to the large gw interferometers like VIRGO[@VIRGO] and LIGO[@LIGO], but a more complicated mechanical susceptivity and response function to gw must be taken into account. The definition of SQL is less obvious than in usual interferometric detectors, as well as the possibility to surpass it, and a study of such a system fully including quantum noise is still lacking. In this article we calculate the sensitivity of a MFP interferometer with detuned cavities. The calculation is very similar to the one described in Ref. [@Arcizet] for a simple cavity, and we find indeed spectral curves well beyond the SQL, with shapes comparable to those typical of signal-recycled interferometers. In addition, we include cavity losses for a better comparison with realistic experimental schemes, we introduce as additional degree of freedom the choice of the final detected field quadrature, and we apply the results to both a standard free-falling masses interferometer and to a DUAL detector. Theoretical model ================= Simple cavity {#sec2a} ------------- Before describing our complete model, we analyze the paradigmatic case of a Fabry-Perot cavity with a movable mirror, neglecting mirror losses and cavity field delay (short cavity regime). Such a calculation is reported in details in Ref. [@Arcizet], and we only add the choice of the detected field quadrature that can be performed by using a local oscillator with tunable phase. Such simplified scheme is useful to understand the physical meaning of the phenomena that will be observed in the complete system. We use the semi-classical formalism described in Ref. [@Fabre], valid in the limit of strong fields, where quantum field fluctuations are treated as classic stochastic variables. In the limit of high finesse and nearly resonant conditions, the equation for the cavity field $\alpha$ reads $$\label{eq1} (-\gamma+i\psi)\alpha+ \sqrt{2\gamma}\,\alpha_{in}=0$$ where $\psi=2k L\,\textrm{Mod}[2\pi]$ is the phase detuning from the closest resonance ($k$ is the laser field wavenumber and $L$ is the cavity length), $2\gamma$ is the input mirror intensity transmission, $\alpha_{in}$ is the input field. The electric fields are normalized such that $|\alpha|^2$ is a flux of photons. The input/output coupler boundary conditions are $$\label{eq2} \alpha_{out}=-\alpha_{in}+\sqrt{2\gamma}\,\alpha \, .$$ In a linearized analysis, the general electric field can be considered as a sum of a steady state $\bar{\alpha}$ (which is null for vacuum fields) and the fluctuations $\delta\alpha(t)$ around it. The steady state of the intracavity and reflected fields are respectively $$\label{eq3} \bar{\alpha}=\frac{\sqrt{2\gamma}}{\gamma-i\bar{\psi}}\,\bar{\alpha}_{in}$$ and $$\label{eq4} \bar{\alpha}_{out}=\frac{\gamma+i\bar{\psi}}{\gamma-i\bar{\psi}}\, \bar{\alpha}_{in} \, .$$ The steady state of the cavity detuning $\bar{\psi}$ is $$\label{eq5} \bar{\psi} = \bar{\psi}_{0}+4\,\hbar k^2 \chi |\bar{\alpha}|^2 \, .$$ Here $\bar{\psi}_{0}$ is the cold-cavity detuning (for vanishing laser field) and the last term in Eq. (\[eq5\]) is the radiation pressure effect, where $\chi$ is the movable mirror susceptivity. The linearized fluctuations $\delta\psi(t)$ of $\psi$ around its steady state and the field fluctuations can be written in the Fourier space defining $\delta\alpha(t)=\delta \tilde{\alpha}(\Omega)e^{-i\Omega t}$, $\delta\alpha^*(t)=\delta \tilde{\alpha}^*(\Omega)e^{-i\Omega t}$ and $\delta\psi(t)=\delta\tilde{\psi}(\Omega)e^{-i\Omega t}$. The equations for such fluctuations read $$\label{eq6} \delta\tilde{\psi}(\Omega) = \delta\tilde{\psi}_{0}+ 4\,\hbar k^2 \chi(\bar{\alpha}^*\delta\tilde{\alpha}+\bar{\alpha}\delta\tilde{\alpha}^*)$$ $$\label{eq7} (\gamma - i\bar{\psi})\delta\tilde{\alpha} - i\bar{\alpha}\delta\tilde{\psi} = \sqrt{2\gamma}\,\delta\tilde{\alpha}_{in}$$ where $\delta\tilde{\psi}_0$ is the signal to be detected, and the equation for $\delta\tilde{\alpha}^*$ is the conjugate of Eq. (\[eq7\]). It is useful to use the quadratures of the field fluctuations, defined as $$\delta p=\delta\tilde{\alpha}+\delta\tilde{\alpha}^* \, \,; \,\, \, \delta q=i(\delta\tilde{\alpha}^*-\delta\tilde{\alpha}) \, .$$ $\delta p$ and $\delta q$ correspond respectively to the amplitude and phase fluctuations, referred to the input mean field that is taken as real (i.e., $\bar{\alpha}_{in}^* = \bar{\alpha}_{in}$). Using such quadratures, the equation (\[eq7\]) becomes $$\label{eq8} \gamma\,\delta p+\bar{\psi}\,\delta q -i(\bar{\alpha}-\bar{\alpha}^*) \delta\tilde{\psi}=\sqrt{2\gamma}\,\delta p_{in}$$ $$\label{eq9} \gamma\,\delta q-\bar{\psi}\,\delta p -(\bar{\alpha}+\bar{\alpha}^*) \delta\tilde{\psi}=\sqrt{2\gamma}\,\delta q_{in} \, .$$ To write clearer expressions, we define $\Psi = \bar{\psi}/\gamma$ (detuning normalized to the half cavity linewidth) and we use a normalized input laser power $p$ (with the dimensions of a force divided by a length, i.e., the inverse of a susceptivity) defined by $$p=16 \frac{\hbar k^2 \bar{\alpha}_{in}^2}{\gamma^2}=16 \frac{k P_{in}}{\gamma^2 c}$$ where $P_{in} = \hbar kc\bar{\alpha}_{in}^2$ is the real input power ($c$ is the speed of light). With these definitions, the expressions (\[eq3\]) and (\[eq4\]) for the fields steady states become $$\label{eq10} \bar{\alpha}=\sqrt{\frac{2}{\gamma}}\frac{e^{i\eta}}{\sqrt{1+\Psi^2}}\,\bar{\alpha}_{in}$$ $$\label{eq11} \bar{\alpha}_{out}=e^{2i\eta}\,\bar{\alpha}_{in}$$ where $\,\,\eta=\arctan\Psi$. The quadrature fluctuations, according to Eqs. (\[eq8\]) and (\[eq9\]), are rotated by the same angle as the average field. Indeed, we have inside the cavity $$\label{eq12} \left( \begin{array}{c} \delta p \\ \delta q \end{array} \right)=\sqrt{\frac{2}{\gamma}}\frac{1}{\sqrt{1+\Psi^2}} \left( \begin{matrix} \cos\eta & -\sin\eta \cr \sin\eta & \cos\eta \end{matrix} \right)\, \left( \begin{array}{c} \delta p_{in} \\ \delta q_{in} \end{array} \right)+\left(\frac{2}{\gamma}\right)^{\frac{3}{2}}\frac{\bar{\alpha}_{in}}{1+\Psi^2} \left( \begin{array}{c} -\sin2\eta \\ \cos2\eta \end{array} \right)\delta\tilde{\psi}$$ where we have used Eq. (\[eq10\]) for replacing $\bar{\alpha}$. For the reflected fields, we find relations similar to Eq. (\[eq12\]), where the rotation angle for the field quadratures is again the same as for the steady state field: $$\label{eq13} \left( \begin{array}{c} \delta p_{out} \\ \delta q_{out} \end{array} \right)= \left( \begin{matrix} \cos2\eta & -\sin2\eta \cr \sin2\eta & \cos2\eta \end{matrix} \right)\, \left( \begin{array}{c} \delta p_{in} \\ \delta q_{in} \end{array} \right)+\frac{4}{\gamma}\frac{\bar{\alpha}_{in}}{1+\Psi^2} \left( \begin{array}{c} -\sin2\eta \\ \cos2\eta \end{array} \right)\delta\tilde{\psi} \,\, .$$ The equation for $\delta\tilde{\psi}$, in the simplified notation, reads $$\label{eq14} \delta\tilde{\psi}=\delta\tilde{\psi}_0+4\,\hbar k^2\chi\sqrt{\frac{2}{\gamma}} \frac{\bar{\alpha}_{in}}{\sqrt{1+\Psi^2}}[\cos\eta\,\delta p+\sin\eta\,\delta q]$$ and, replacing Eq. (\[eq12\]) for $\delta p$ and $\delta q$, we obtain $$\label{eq15} \delta\tilde{\psi}=\delta\tilde{\psi}_0+4\,\hbar k^2\chi\frac{2}{\gamma} \frac{\bar{\alpha}_{in}}{1+\Psi^2}\left[\delta p_{in}- \frac{2}{\gamma} \frac{\Psi}{1+\Psi^2}\,\bar{\alpha}_{in}\,\delta\tilde{\psi}\right] \,\, .$$ The detected quadrature can be chosen at will by tuning the homodyne angle $w$, and the corresponding fluctuations are $$\label{eq16} \delta E_{out}=\delta p_{out} \cos w+\delta q_{out} \sin w \,\, .$$ We remark that, with respect to the input field, an amplitude detection is obtained for $w=0$ while a phase detection corresponds to $\,w=\pi/2$. On the other hand, with respect to the output field (see Eq. (\[eq11\])), $w=2\eta\,$ corresponds to a pure amplitude detection and $\,w=2\eta+\pi/2\,$ to a pure phase detection. With $\,w=2\eta$ we have $\,\,\delta E_{out}=\delta p_{in}\,\,$ while for $\,\,w=2\eta+\pi/2\,\,$ we obtain $$\label{eq17} \delta E_{out}=\delta q_{in}+\frac{4}{\gamma}\frac{\bar{\alpha}_{in}}{1+\Psi^2}\delta\tilde{\psi} \, .$$ In general, for a given homodyne phase $\,w$, the detected field fluctuations are $$\label{eq18} \delta E_{out}=\delta p_{in}\cos(w-2\eta)+\left(\delta q_{in}+\frac{4}{\gamma}\frac{\bar{\alpha}_{in}}{1+\Psi^2}\delta\tilde{\psi}\right)\sin(w-2\eta) \,\, .$$ Replacing the expression (\[eq15\]) for $\delta\psi$ in Eq. (\[eq18\]), we get $$\label{eq19} \delta E_{out}=\delta p_{in}\cos(w-2\eta)+\sin(w-2\eta)\left[\delta q_{in}+ \frac{4}{\gamma}\frac{\bar{\alpha}_{in}}{1+\Psi^2}\frac{\delta\tilde{\psi}_0}{1+A} +\frac{2}{\Psi}\frac{A}{1+A}\delta p_{in}\right]$$ with $$A=\frac{\Psi}{(1+\Psi^2)^2}\,p\,\chi \,\,\, .$$ The last term in Eq. (\[eq19\]) is the contribution of the radiation pressure noise, proportional to the input field amplitude fluctuations $\,\delta p_{in}$. For $\,w=2\eta+\pi/2$, the amplitude fluctuations only enter into the detection noise through such radiation pressure term. Changing the detection angle from ($w=2\eta+\pi/2$), the contribution of the phase fluctuations $\,\,\delta q_{in}\,$ to the signal-to-noise ratio (SNR) remains the same (both $\,\delta q_{in}\,$ and $\,\delta\psi_0\,$ multiply the angle-dependent factor $\,\sin(w-2\eta)\,\,$). On the other hand, a further contribution of the amplitude fluctuations comes into play, that can compensate the radiation pressure fluctuations. The optimal SNR is obtained when the contribution of $\,\,\delta p_{in}\,$ is completely cancelled, a situation occurring for an angle $\,w_{opt}\,$ such that $$\tan(w_{opt}-2\eta)=-\frac{1+A}{A}\,\frac{\Psi}{2} \,\, .$$ In this case, the SNR is only limited by the phase fluctuations, and it increases with the coefficient of $\,\psi_0\,$ within square brackets in Eq. (\[eq19\]). In particular, the sensitivity increases for negative detuning (i.e., for $A<0$) and one can get in principle unlimited SNR if the denominator of this coefficient vanishes, i.e., for $\,\,A\to-1.\,\,$ This can happen for large enough power and/or susceptivity, more precisely if $\,\,p\chi>16/3\sqrt{3}\,.\,$ Such increase in sensitivity at negative detuning is usually interpreted as due to a modified effective susceptivity (‘optical spring‘) originated by the position-dependent radiation pressure force. Complete system --------------- The optical scheme we consider, reported in Fig. (\[schema1\]), is a MFP with the addition of an homodyne balanced detection at the output. In this way, one can choose the quadrature of the output field to be detected as in the paradigmatic case just analyzed. Vacuum fluctuations are introduced through the mirrors losses in the cavities and through the output port of the Michelson beam splitter, while laser field fluctuations enter through the beam splitter input port. ![Left: the optical configuration considered is a Michelson-Fabry-Perot with the additional free choice of the detected field quadrature. Right: mechanical scheme, with self- and cross- susceptivities.[]{data-label="schema1"}](figu1.eps){width="0.9\columnwidth"} Concerning the mechanics of the system, we consider a susceptivity matrix to include the possibility of changing the length of one arm by acting with a force on the other arm. Such a possibility is important in the case of an interferometer mounted on a solid body. A sketch of the mechanical scheme is included in Fig. (\[schema1\]) (right). The equations for the electric fields in the two cavities (labeled by $i,j=(1,2)$) read $$\label{dadt} \tau_i \frac{d\alpha_i}{dt}=-(\gamma_i-i\psi_i)\alpha_i+ \sqrt{2\gamma_{mi}}\alpha^{in}_i+\sqrt{2 \gamma_{li}}\delta\alpha_{li}$$ where $\tau_i=2L_i/c$ is the cavity roundtrip time, c is the speed of light, $2\gamma_{mi}$ is the input mirror intensity transmission, $2\gamma_{li}$ are the roundtrip intensity losses (including transmission from the back mirror, absorption and scattering in both mirrors, diffraction losses, etc.), $\gamma_i=\gamma_{mi}+\gamma_{li}$, $\delta\alpha_{li}$ are the vacuum fluctuations entering through cavity losses which are mimic by a partially transmitting output mirror. Assuming an ideal 50$\%$ beam splitter, the input fields $\alpha^{in}_i$ of the two cavities are $$\label{ain1} \alpha^{in}_1=\frac{\alpha_{in}-\delta\alpha_V}{\sqrt{2}}$$ $$\label{ain2} \alpha^{in}_2=\frac{\alpha_{in}+\delta\alpha_V}{\sqrt{2}}$$ where $\alpha_{in}$ is the laser input field and $\delta\alpha_V$ are the vacuum fluctuations entering through the beam splitter output port. In Eqs. (\[ain1\]-\[ain2\]), to simplify the notation, we have neglected the phase difference between $\alpha^{in}_1$ and $\alpha^{in}_2$ introduced by the length difference in the paths from the beam splitter to the two input cavity mirrors. This phase difference will be re-considered at the output port of the beam splitter (Eq. (\[BSout\])). The cavity input/output coupler boundary conditions are $$\label{aout} \alpha^{out}_i=-\alpha^{in}_i+\sqrt{2\gamma_{mi}}\alpha_i \, .$$ At the output port of the beam splitter, fields are recombined giving: $$\label{BSout} \alpha_{BS}=\frac{1}{\sqrt{2}}(-\alpha^{out}_1+\alpha^{out}_2 e^{i\theta})$$ where the phase $\theta$ accounts for the double path difference between the beam splitter and the two input mirrors. Finally, the observed quadrature of the output field can be chosen by changing the detection phase $w$. The steady state of the intracavity field is obtained by zeroing the time derivative in Eq. (\[dadt\]), using Eqs. (\[ain1\]-\[ain2\]) and neglecting the field fluctuations: $$\bar{\alpha}_i=\frac{\sqrt{\gamma_{mi}}}{\gamma_i-i\bar{\psi}_i}\bar{\alpha}_{in} \, .$$ Using Eq. (\[aout\]), we find for the steady state of the reflected field: $$\bar{\alpha}^{out}_i=\frac{\gamma_{mi}-\gamma_{li}+i\bar{\psi}_i}{\gamma_{mi}+\gamma_{li}-i\bar{\psi}_i} \frac{\bar{\alpha}_{in}}{\sqrt{2}} \, .$$ The cavity length is sensitive to several kinds of forces acting on the system, including classic deterministic (e.g., the gravitational wave effect), stochastic (thermal noise) and quantum forces (the radiation pressure acting on the mirrors). The steady state cavity detuning $\bar{\psi}_i$ can be written as $$\bar{\psi}_i = \bar{\psi}_{0i}+4\hbar k^2 \chi^0_{ii}|\bar{\alpha}_i|^2+ 4\hbar k^2 \chi^0_{ij}|\bar{\alpha}_j|^2$$ where $\chi^0_{ij}$ is the stationary (zero-frequency) susceptivity matrix. The equation for the fluctuations of the cavity phase detuning $\delta\tilde{\psi}_i(\Omega)$ (expressed in the Fourier space) is $$\label{psi} \delta\tilde{\psi}_i(\Omega) = \delta\tilde{\psi}_{0i}+ 4\hbar k^2 [\chi_{ii}(\Omega)(\bar{\alpha}^*_i\delta\tilde{\alpha}_i+\bar{\alpha}_i\delta\tilde{\alpha}^*_i)+\chi_{ij}(\Omega)(\bar{\alpha}^*_j\delta\tilde{\alpha}_j+\bar{\alpha}_j\delta\tilde{\alpha}^*_j)]$$ where $\delta\tilde{\psi}_0$ contains the effects of thermal and external noise, and of the gw signal and $\chi_{ij}(\Omega)$ is the susceptivity matrix. The complete set of equations for the fields and cavities displacements are reported in Appendix A. The complete expressions with different parameters for the two cavities are useful for a future numerical analysis of the effects of the asymmetries and of the allowed tolerances in the parameters. However in this work, for a simpler understanding of the physical phenomena, we will take identical cavities with $\gamma_{m1} = \gamma_{m2} = \gamma_m$, $\gamma_{l1} = \gamma_{l2} = \gamma_l$, $\tau_1 = \tau_2 = \tau$. As a further restriction to our analysis, we will consider that: a\) the Michelson interferometer working point gives a dark fringe at the beam splitter output port. This corresponds to setting equal distances between the beam splitter and the two cavities, i.e., $e^{i\theta}=1$; b\) the two cavities have the same detuning from the laser frequency: $\bar{\psi}_1 = \bar{\psi}_2 = \bar{\psi}$; c\) the mechanical system is symmetric: $\chi_{11}=\chi_{22}=\chi_s$ and $\chi_{12}=\chi_{21}=\chi_c$. These three conditions determine the cancellation of the effect of the input laser field fluctuations in the output. This requirement is important for a system working in the acoustic frequency range. Indeed, while it is very difficult to reduce the laser field amplitude fluctuations at the shot noise level [@Conti], for phase fluctuations the reduction to the quantum limit is even more difficult and far from being demonstrated in strong power laser fields. The field fluctuations $\delta \tilde{E}_{out}$ (seen by the homodyne detection) are described by a vector of coefficients $V_{out}$, multiplying the input fluctuations $X_{in}$: $$\label{Eout} \delta \tilde{E}_{out}(\Omega) = V_{out} \cdot X_{in}(\Omega)$$ with $X_{in}$ and $V_{out}$ given respectively in equations (\[Xin\]) and (\[Vout\]) of the Appendix, where we have used for convenience the quadratures of the field fluctuations. As already remarked, the coefficients $V_{out}[\delta p_{in}]$ and $V_{out}[\delta q_{in}]$ that multiply the input laser field fluctuations $\delta p_{in}$ and $\delta q_{in}$ are null in the completely symmetric case that we are considering. Besides the previous defined normalized detuning $\Psi=\bar{\psi}/\gamma$, we use for a more compact notation $\Gamma_m = \gamma_m/\gamma$ (in the case of loss-less cavities, $\Gamma_m = 1$); $\Omega_{cav}=\gamma/\tau$ (cutoff angular frequency of the cavity) and the normalized input laser power $p$ is now $$p=16 \frac{\Gamma_m^{3/2}k P_{in}}{\gamma^2 c} \, .$$ The expression of the sensitivity $S_L(\Omega)$ in the detection of $\delta(L_1-L_2)$ (defined as the signal spectral power with unitary signal-to-noise spectral density) is given in Appendix A, Eq. (\[NSsimm\]). We see that in $S_L(\Omega)$ the susceptivities only appear as a difference $\chi_s - \chi_c$. We define in the following $\chi_s - \chi_c = \chi$. General discussion ================== The first situation that we consider is with the cavities at resonance ($\bar{\psi} = 0$) and pure phase quadrature detection ($w = \pi/2$). Such configuration corresponds to the present operative gw interferometers VIRGO and LIGO, and we will define it in the following as ’normal case’. The $S_L(\Omega)$ of Eq. (\[NSsimm\]) becomes: $$S_L = \frac{\hbar}{\sqrt{\Gamma_m}}\frac{1+(\frac{\Omega}{\Omega_{cav}})^2}{p}\left[1+\left(\frac{p|\chi|}{1+(\frac{\Omega}{\Omega_{cav}})^2}\right)^2\right] \, .$$ It can be seen that $S_L$ can be written as a sum of a ’displacement noise’ term $S_{xx}$ and a term proportional to a ’force noise’ $S_{FF}$: $$S_L = S_{xx} + \left|\chi\right|^2 S_{FF}$$ with $$S_{xx} S_{FF} = \hbar^2/\Gamma_m \, .$$ The origin of the two terms can be found respectively in the intracavity field phase noise and amplitude noise. The first term limits the detection sensitivity of a phase signal created by the mirrors displacement; the second term is due to the fluctuations of the radiation pressure acting on the cavity mirrors. For each detection frequency $\Omega$, the optimal $S_L(\Omega)$ is reached for $S_{xx}= \hbar\left|\chi\right|/\sqrt{\Gamma_m}$. With this condition, $S_L$ is equal to what we call standard quantum limit: $$\label{SQL} SQL(\Omega) = \frac{2\hbar\left|\chi(\Omega)\right|}{\sqrt{\Gamma_m}} \, .$$ With long cavities (i.e., when $\Omega/\Omega_{cav}$ is not negligible) and/or a frequency-dependent susceptivity, the $SQL$ defines an envelop of possible spectral density curves, each one determined by the choice of the input power. Already with resonant cavities, tuning the homodyne phase $w$ allows to change the sensitivity significantly. $S_L$ still assumes a simple form: $$\begin{aligned} S_L(\Omega)&=& \frac{\hbar}{\sqrt{\Gamma_m}}\frac{1+(\frac{\Omega}{\Omega_{cav}})^2}{p}\Big[1+( 1-\Gamma_m)\left(\frac{p Re\chi}{1+(\frac{\Omega}{\Omega_{cav}})^2}\right)^2\\\nonumber&+&\left( \frac{\cos w}{\sin{w}}+\sqrt{\Gamma_m}\frac{p Re\chi}{1+(\frac{\Omega}{\Omega_{cav}})^2}\right)^2+\left(\frac{p Im\chi}{1+(\frac{\Omega}{\Omega_{cav}})^2}\right)^2\Big] \label{NSresonant}\end{aligned}$$ that, for loss-less cavities ($\Gamma_m = 1$) and real $\chi$, can be further simplified to $$S_L(\Omega)= \hbar\frac{1+(\frac{\Omega}{\Omega_{cav}})^2}{p}\left[1+\left(\frac{\cos w}{\sin{w}}+\frac{p \chi}{1+(\frac{\Omega}{\Omega_{cav}})^2}\right)^2\right] \, . \label{NSreslossless}$$ The physical interpretation of Eq. (\[NSreslossless\]) is simple, as already explained for a simple cavity. Like in the ’normal case’, for resonant cavities there is no quadrature rotation, the vacuum fluctuations $\delta p_V$ and $\delta q_V$ are transferred respectively to the amplitude ($\delta p_{BS}$) and phase ($\delta q_{BS}$) fluctuations of the output field, and the radiation pressure is proportional simply to $\delta p_V$. The cavity length fluctuations, besides the signal, contain a term due to the radiation pressure, and a term proportional to the length fluctuations is present in the output field phase. As a consequence, the amplitude fluctuations $\delta p_V$ are found both in the output field amplitude quadrature $\delta p_{BS}$ (giving the first term in round brackets of Eq. (\[NSreslossless\])) and in the output phase $\delta q_{BS}$ (second term in round brackets of Eq. (\[NSreslossless\])). A suitable choice of the homodyne phase, and therefore of the detected quadrature, brings these terms to cancel each other. Also for $\Gamma_m \ne 1$, choosing $$w = w_0 = \arctan\left(-\frac{1+(\Omega/\Omega_{cav})^2}{\sqrt{\Gamma_m}\,p\, Re\chi}\right) \, , \label{w0}$$ the contribution of amplitude fluctuations is minimized and the $S_L$ becomes $$S_L = S_L^0 = S_{xx}+S_{FF} [(1-\Gamma_m) Re\chi^2 + Im\chi^2] \, .$$ For loss-less cavities ($\Gamma_m = 1$) and real $\chi$, the radiation pressure noise can be completely cancelled and we have $S_L^0 = S_{xx}$. For frequency-dependent susceptivity and/or long cavities, $S_L^0$ defines the locus of the minima of spectral curves that can be tuned by changing $w$. The value of $w$ necessary to achieve the best sensitivity at a chosen frequency is given by Eq. (\[w0\]). We remark that in the ’normal case’ the $SQL$ is an absolute limit, while in the general resonant case (in particular, for $\Gamma_m = 1$) the sensitivity limit given by $S_{xx}$ can be decreased at will by increasing the laser power. For detuned cavities, both the radiation pressure and the output field fluctuations contain a mixture of $\delta p_V$ and $\delta q_V$ and the force noise is not due any more solely to the pure amplitude fluctuations $\delta p_V$. A clear discussion of the phenomena involved in this case is given in Ref. [@Arcizet]. ![Sensitivity $S_L$ as a function of input power $p$, for $\Omega_{cav}\to\infty$, $\Gamma_m=1$, and constant, real $\chi$. Solid line: ’normal case’ ($\Psi=0$, $w=\pi/2$); dashed line: general resonant case ($\Psi=0$, $w=\arctan(-1/p\chi)$); dotted lines: $\Psi=-0.4$ and $w=\pi/2$ (a), $w=2.0$ (b), $w=2.3$ (c).[]{data-label="costantchi"}](figu2.eps){width="0.9\columnwidth"} For a better understanding of the physics, we analyze the case of a constant, real, positive $\chi$ and very short cavities ($\Omega_{cav} \gg \Omega$) with negligible losses ($\Gamma_m \simeq 1$). The ’normal case’ sensitivity is shown in Fig. (\[costantchi\]) with a solid line, as a function of the input power. The $SQL$ is reached for $p=p_{SQL}=1/\chi$. For $p > p_{SQL}$ the sensitivity is worse due to strong radiation pressure effect, for $p < p_{SQL}$ it is deteriorated by phase noise. The sensitivity $S_L^0$ for the general resonant case is shown in Fig. (\[costantchi\]) with a dashed line. In this case, $S_L^0$ coincides with $S_{xx}$. For low power it approaches the ’normal case’ sensitivity (that is here dominated by $S_{xx}$), but for $p > p_{SQL}/2$ it surpasses the $SQL$. If we allow for different values of the detuning $\Psi$, $S_L$ can decrease well below the $SQL$ and even below $S_L^0$, as shown in Fig. (\[costantchi\]) with dotted lines, and it is unlimited if $p$ is strong enough. The physics behind this effect was previously explained in Section \[sec2a\]. In short terms, it is the result of: a) the cancellation of the input amplitude fluctuations (thanks to a good choice of the homodyne detection angle); b) the modified effective susceptivity, that increases the sensitivity to the mirror motion. Considering now a complex susceptivity $\chi$, taking into account mechanical dissipation, it can be shown that the minimal sensitivity is $\,\,\hbar \, Im\chi$, as already found in Ref. [@Jaekel] for a MFP with resonant cavities and squeezed input fields, and by Arcizet *et al.* [@Arcizet] for a detuned Fabry-Perot cavity. This phenomenon can be understood as follows. In the detected field, the fluctuations $\delta p_V$ and $\delta q_V$ are present for two different reasons: directly in the field reflected by the interferometer (purely optical effect), and because of the length fluctuations induced by radiation pressure (opto-mechanical effect). As we have already seen, the same situation is found in the case of resonant cavities, but only for $\delta p_V$. The two effects give different linear combinations of $\delta p_V$ and $\delta q_V$, with real coefficients depending on $\Psi$ and $w$ (we are considering $\Omega_{cav}\to\infty$). However, the opto-mechanical effect is mediated by the susceptivity $\chi$. If $\chi$ is real, an appropriate choice of $\Psi$, $w$ and $p$ (and therefore of the coefficients multiplying $\delta p_V$ and $\delta q_V$) can bring to a complete cancellation between the purely optical and opto-mechanical effects. As already explained, the same situation, occuring in the case of resonant cavities only for $\delta p_V$, defines $S^0_L$. However, if $\chi$ has an imaginary component, the radiation pressure fluctuations cannot be completely cancelled. In other words, the total detected fluctuations cannot be completely deleted because of the de-phasing between intracavity intensity changes and cavity length changes introduced by the complex $\chi$. In spite of this interesting physical result, a region with significant imaginary part of the susceptivity is of limited practical interest and in the following we will assume a real $\chi$. Free-falling mirrors ==================== We apply now our results to find the sensitivity of an interferometer with free-falling mirrors, still with loss-less short cavities. Such a scheme is a good approximation for mirrors suspended to a pendulum with low oscillation frequency. The susceptivity can be written as $$\label{chifree} \chi = - \chi_0 \left(\frac{\Omega_0}{\Omega}\right)^2$$ where $\chi_0$ and $\Omega_0$ are constants. We remark that a finite zero-frequency susceptivity (and therefore a defined pendulum oscillation frequency) is necessary in the calculation of the steady state and stability of the system, but it can be neglected in the evaluation of the spectra. ![Sensitivity $S_L$ in the case of free-falling mirrors, as a function of the frequency $\Omega$, for $\Omega_{cav}\to\infty$, $\Gamma_m=1$, $p=1$. Solid line: ’normal case’ ($\Psi=0$, $w=\pi/2$); dashed line: resonant case limit $S_L^0$; dash-dotted line: SQL; dotted lines: $\Psi=0$ and $w=1.2$ (a), $w=0.8$ (b), $w=0.4$ (c), $w=0.2$ (d).[]{data-label="free1"}](figu3.eps){width="0.9\columnwidth"} As well known, in the ’normal case’ for each particular choice of the power $P_{in}$ the $SQL$ is reached at a corresponding frequency $\Omega_{SQL}$ given by $$\Omega_{SQL}^2=p\chi_0\Omega_0^2 \, ,$$ where $S_L$ is: $\,S_L(\Omega_{SQL})=2\hbar/(\sqrt{\Gamma_m}\,p)$. At high frequencies, $S_L$ tends to the asymptotic value $\,S_L(\Omega\to\infty)=\frac{1}{2}S_L(\Omega_{SQL})$, limited by phase noise, while below $\Omega_{SQL}$ the $S_L$ increases as $\Omega^{-4}$ due to radiation pressure fluctuations. An example is shown in Fig. \[free1\] (solid line). Keeping resonant cavities but changing $w$, the sensitivity curves can surpass the $SQL$, with minima lying on the horizontal line given by $S_{xx}$ (Fig. \[free1\]). We remark that in our scheme (with completely symmetric cavities), the average field at the output port vanishes, therefore the local oscillator power can be kept low. At first order, the balanced homodyne detection scheme is not sensitive to the local oscillator noise. Therefore, the phase tuning element (e.g., a phase mirror or an electro-optic modulator) is not critical for the noise budget. As a consequence, this sensitivity tuning technique can be easier to be implemented than other schemes (e.g., tuning the MFP cavities or a signal recycling mirror). ![Sensitivity $S_L$ in the case of free-falling mirrors, as a function of the frequency $\Omega$, for $\Omega_{cav}\to\infty$, $\Gamma_m=1$, $p=1$. Solid line: ’normal case’ ($\Psi=0$, $w=\pi/2$); dashed line: resonant case limit $S_L^0$; dash-dotted line: SQL; dotted lines: $w=\pi/2$ and $\Psi=0.4$ (a), $\Psi=0.6$ (b), $\Psi=0.8$ (c), $\Psi=0.9$ (d), $\Psi=1.1$ (e), $\Psi=1.2$ (f), $\Psi=1.4$ (g).[]{data-label="free2"}](figu4.eps){width="0.9\columnwidth"} If we now allow for detuned cavities we see that even $S_L^0$ can be largely surpassed. Some examples are shown in Fig. (\[free2\]) for a detection phase $w$ kept at $\pi/2$. We remark that the position of the minimum $S_L$ shifts toward low frequencies at increasing detuning. Therefore, the input power should be increased to keep the same optimal frequency range. Similar results are described in Ref. [@Arcizet] (see their Fig. 3)[@nota1]. ![Sensitivity $S_L$ in the case of free-falling mirrors, as a function of the frequency $\Omega$, for $\Omega_{cav}\to\infty$, $\Gamma_m=1$, $p=1$. Black solid line: ’normal case’ ($\Psi=0$, $w=\pi/2$); long-dashed straight line: resonant case limit $S_L^0$; dash-dotted line: SQL; dashed lines: the phase $w$ is optimized for each frequency, $\Psi=0.06$ (a), $\Psi=0.1$ (b), $\Psi=0.2$ (c), $\Psi=0.3$ (d), $\Psi=0.4$ (e); gray solid line: both $w$ and $\Psi$ are optimized for each frequency. In the inset: phase $w$ optimized for each frequency, for the values of $\Psi$ used in the main plot.[]{data-label="free3"}](figu5.eps){width="0.9\columnwidth"} Tuning the phase $w$ allows to obtain families of sensitivity curves, one for each fixed value of detuning, whose envelopes are very broad and low. Some examples are shown in Fig. (\[free3\]), where the gray solid curve represents the absolute sensitivity limit that can be reached optimizing, for each frequency, both the phase $w$ and the detuning. In the inset of the figure is shown the optimal phase $w$, as a function of the frequency, that is necessary to obtain the envelope curves. For resonant cavities, inserting the susceptivity (\[chifree\]) in equation (\[w0\]), we find that the optimal phase is $w=w_0=\arctan(\Omega/\Omega_{SQL})^2$. The other curves differ from $w_0$ for a very flat additional phase. We notice that, besides choosing a particular detection phase to optimize the detection at the preferred frequency, one can also obtain a sensitivity curve corresponding to one of the above envelopes by adding a suitable frequency-dependent quadrature rotation at the output, before the homodyne detection, like in the so-called variational-output interferometers ([@Vyatchanin95; @Kimble01]). ![Sensitivity $S_L$ in the case of free-falling mirrors, as a function of the frequency $\Omega$, for $\Omega_{cav}\to\infty$, $p=1$. Solid lines: ’normal case’ ($\Psi=0$, $w=\pi/2$, from lower to upper curve $\Gamma_m=1,\, 0.8,\, 0.5$); dashed line: resonant case limit $S_L^0$ for $\Gamma_m=0$; dash-dotted line: SQL for $\Gamma_m=0$; dotted lines: $w=\pi/2$, $\Psi=0.6$, and from lower to upper curve $\Gamma_m=1,\, 0.8,\, 0.5$.[]{data-label="losses"}](figu6.eps){width="0.9\columnwidth"} Moderate cavity losses do not change qualitatively the above features, and also quantitatively the enhanced sensitivity is well preserved. An example is shown in Fig. (\[losses\]). Even at $\Gamma_m = 0.5$ (input mirror transmission equal to the other roundtrip losses) the dip in the $S_L$ is well pronounced, below the $SQL$. The expression of $S_L$ including the cavity response is reported in the Appendix A (Eq. (\[NSsimm\])). The simplified expression for the ’normal case’ is $$S_L^0 (\Omega)= S_L(\Omega_{SQL})\cdot\frac{1}{2}\left[1+\left(\frac{\Omega}{\Omega_{cav}}\right)^2+\frac{\left(\frac{\Omega_{SQL}}{\Omega}\right)^4}{1+\left(\frac{\Omega}{\Omega_{cav}}\right)^2}\right] \, .$$ Two examples are shown in Fig. (\[cut\]), for $\Omega_{cav}= 0.3\,\Omega_{SQL}$ and $\Omega_{cav}= 3\,\Omega_{SQL}$ (dark, thin lines). In any case, the $SQL$ remains the same (as in Eq. (\[SQL\])), and at high frequency (above $\Omega_{cav}$ and $\Omega_{SQL}$) the $S_L$ now increases as $\Omega^2$. ![Sensitivity $S_L$ in the case of free-falling mirrors, as a function of the frequency $\Omega$, for $\Gamma_m=1$, $p=1$. Thin, dark lines (black online): ’normal case’ ($\Psi=0$, $w=\pi/2$); thick, light lines (red online): $\Psi=0.6$, $w=1.3$. Solid lines: $\Omega_{cav}\to\infty$; dotted lines: $\Omega_{cav}=3 \,\Omega_{SQL}$; dashed lines: $\Omega_{cav}=0.3 \,\Omega_{SQL}$. Dash-dotted line: SQL.[]{data-label="cut"}](figu7.eps){width="0.9\columnwidth"} For detuned cavities, the most interesting situation is when $\Omega_{cav}$ is around or above $\Omega_{SQL}$. In this case, for $\Omega<\Omega_{SQL}$ the $S_L$ is not significantly modified with respect to the previously studied situation of $\tau \simeq 0$. At high frequency (well above $\Omega_{cav}$) we can write $$S_L (\Omega)\simeq S_L(\Omega_{SQL})\cdot\frac{1}{2}\frac{1+\Psi^2}{\sin^2(w-\arctan\Psi)}\left(\frac{\Omega}{\Omega_{cav}}\right)^2 \, .$$ $S_L$ increases as $\Omega^2$ like in the ’normal case’, and is always above it, with an optimal detection phase of $w = \frac{\pi}{2}+\arctan(\Psi)$. What is more peculiar is the behavior at intermediate frequencies. Here a further increase in sensitivity is found (see the ’bump’ in the light (read online) dotted curve of Fig. (\[cut\])). The $S_L$ is here lower than in the case of negligible $\tau$, yet remaining above the $SQL$. DUAL detector ============= A DUAL gw detector exploits two oscillation modes of a mechanical system with a readout symmetric with respect to the center of mass. Do to the geometry, the responses to the readout force of the two modes must be summed, while the responses to the tidal force of the gw are subtracted. In the frequency region between the two resonance frequencies, the susceptivities of the two modes are in anti-phase, giving a reduced response to the readout force and an enhanced response to the gw. Different configurations have been proposed and studied: two nested spheres [@Cerdonio], where the relevant modes are the first quadrupolar mode of the inner and outer sphere; two nested cylinders [@Bonaldi], again acting on the first quadrupolar mode of the nested bodies; a single hollow cylinder, exploiting its first and second quadrupolar modes; a symmetric set of cylinders, where the first DUAL mode is given by the link between them and the second one by their first oscillation mode [@Marin]. For our analysis, we consider the simple case of very low (vanishing) first resonance frequency. The susceptivity (i.e., the difference between $\chi_s$ and $\chi_c$) can be written as $$\chi(\Omega) = \left[-\frac{1}{\Omega^2}+\frac{1}{\mu (\Omega_R^2-\Omega^2)}\right]\chi_0\Omega_0^2$$ where $\Omega_R$ is the (second) DUAL resonance frequency, $\mu$ is an adimensional modal mass and the frequency is normalized to $\Omega_0$. A typical choice of $\Omega_0$ is $\Omega_0 = v_s/R$, where $R$ is a typical dimension of the detector and $v_s$ is the sound velocity in the detector material. $\chi_0$ depends on the material properties and geometrical configuration of the detector. The response to the gw is $$H_{GW}(\Omega) = 1+\frac{\Omega^2}{\mu_{GW} (\Omega_R^2-\Omega^2)}$$ where $\mu_{GW}$ is an adimensional gw sensitive modal mass. Both $\mu$ and $\mu_{GW}$ are normalized to the detector physical mass and can be calculated from the detector geometry: $\mu$ is a surface overlap integral between the radiation pressure of the field interrogating the surface and the considered mechanical mode; $\mu_{GW}$ is a volume overlap integral between the quadrupolar gw force and the mechanical mode. We take for example in this article the ’QUAD’ detector described in Ref. [@Marin], with four parallel molybdenum cylinders. In this case, we have $\mu = 0.28$, $\mu_{GW} = 1.1$, $\Omega_R = 1.4$. ![Sensitivity to a gravitational wave $S_{GW}$ of a DUAL detector, as a function of the frequency $\Omega$, in arbitrary units. Dark, solid line (black online): $SQL_{GW}$. Light solid line (red online): ‘normal case‘ (resonant cavities, $w=\pi/2$) with $p=0.4$. Dotted lines: ‘normal case‘ with $p=2$ (a), $p=0.1$ (b).[]{data-label="dual1"}](figu8.eps){width="1.0\columnwidth"} Being interested in the sensitivity to a gw signal, we can define it as $$S_{GW}(\Omega) = \frac{S_L(\Omega)}{|H_{GW}(\Omega)|^2} \, .$$ In the case of the free-falling masses MFP considered above, we have $H_{GW}=1$ and therefore $S_{GW}=S_L$. The standard quantum limit is defined similarly to Eq. (\[SQL\]): $$SQL_{GW}= \frac{2 \hbar|\chi(\Omega)|}{\sqrt{\Gamma_m}|H_{GW}(\Omega)|^2}$$ and is shown in Fig. (\[dual1\]) with a dark, solid line. In the ’normal case’, varying the input power gives a family of sensitivity curves, limited by the $SQL_{GW}$. A proper choice of the power gives a rather flat curve. For lower power, the sensitivity is peaked around $\Omega_R$ where both the gw signal and the radiation pressure are amplified. A deep in the radiation pressure effect is found around a particular frequency $\Omega_{rp}$ where the interference between the two detector modes gives $\chi(\Omega_{rp})=0$ (for the parameters here considered it happens at $\Omega_{rp}\simeq0.655$). For high input power, around this frequency we have the best sensitivity (Fig. \[dual1\]). ![Sensitivity to a gravitational wave $S_{GW}$ of a DUAL detector, as a function of the frequency $\Omega$, in arbitrary units. Dark, solid line (black online): $SQL_{GW}$, the other curves are for $p=0.4$; dashed line: resonant case limit; light solid line (red online): ‘normal case’; dotted lines: $w=\pi/2$ and $\Psi=-0.5$ (a), $\Psi=0.5$ (b).[]{data-label="dual2"}](figu9.eps){width="1.0\columnwidth"} Also for a DUAL detector, keeping the cavities at resonance and varying $w$ one obtains a set of curves whose envelop, shown in Fig. (\[dual2\]) with a dashed line, is well below the $SQL_{GW}$ and allows to widen the detector sensitivity around $\Omega_R$. Changing the detuning $\Psi$, we find a $S_{GW}$ getting even lower. As shown in Fig. (\[dual2\]), depending on the parameters, the reduction below $SQL_{GW}$ can occur either in a range between $\Omega_{rp}$ and $\Omega_R$ (where $\chi>0$), or below $\Omega_{rp}$ and above $\Omega_R$ (where $\chi<0$). It can be shown as a general property that, for fixed detuning and phase $w$, it is not possible to obtain a sensitivity curve falling below the $SQL$ both in frequency regions with positive $\chi$ and in regions with negative $\chi$. Conclusions =========== We have presented the analysis of a Michelson-Fabry-Perot interferometer with the addition of the free choice of the detected field quadrature. Our study fully accounts for quantum noise, including radiation pressure effects and possible losses in the cavities, and we consider in particular the case of detuned optical cavities. The use of a susceptivity matrix allows to extend the applicability of our results from the usual, free-falling mirrors to the most general mechanical system, including any kind of DUAL detectors with mirrors installed on elastic, mechanically coupled test masses. In view of the development of extended numerical evaluations, considering construction tolerances and variations of system parameters, we have given complete general expressions of the expected output signals. A physical analysis of simplified expressions shows that, thanks to the possible choice of the detected quadrature, the sensitivity can surpass the SQL even with resonant cavities. This is due to the complete cancellation of the radiation pressure fluctuations in the detected output field quadrature. With detuned cavities, in the case of real susceptivity and loss-less cavities, the peak spectral sensitivity is unlimited if the susceptivity and/or the laser power are strong enough: the ultimate sensitivity is only limited by the reactive component of the susceptivity. We also show that reasonable optical losses do not critically modify the performance. The scheme that we consider, with the choice of the detected field quadrature, is easier to be implemented than other proposed and tested configurations (e.g., the resonant sideband extraction[@Heinzel] or the signal-recycling cavity[@Buonanno; @Harms]). It must be remarked, however, that here the sensitivity enhancement occurs in a region of high susceptivity (corresponding in the case of free-falling mirrors to the low-frequency region). While, for a chosen value of the detection phase, the sensitivity is optimized at a particular frequency (i.e., for a particular value of the susceptivity), one can implement a frequency-dependent rotation of the output field before detection, as proposed and analyzed in several works[@Vyatchanin95; @Kimble01], in order to enlarge the useful bandwidth. Without discussing and detailing the possible experimental schemes suitable at this purpose, we give the obtainable sensitivity curves that are indeed broad and deep with respect to the SQL. The use of detuned cavities is again somehow simpler than other schemes that are based on additional mirrors and cavities or on the use of squeezed light, yet allowing to observe interesting noise reduction effects similar to those reported in the literature concerning the mentioned configurations. We remark that detuned cavities imply in general critical stability problems due, e.g., to radiation pressure and photo-thermal effects[@bel; @Marino07]. The analysis of this topic is beyond the purpose of this article, and it should be developed for each particular system including the full frequency response and the possible active stabilization. Appendix A ========== The equations for the field and cavity phase linearized fluctuations in the Fourier space can be written in a compact matrix form. Defining an input fluctuations vector $A_{in}(\Omega)$ and a cavity fluctuations vector $A(\Omega)$ as $$A(\Omega)= \left( \begin{array}{c} \delta\tilde{\alpha}_1(\Omega) \\ \delta\tilde{\alpha}_1^*(\Omega) \\ \delta\tilde{\psi}_1(\Omega)\\ \delta\tilde{\alpha}_2(\Omega) \\ \delta\tilde{\alpha}_2^*(\Omega) \\ \delta\tilde{\psi}_2(\Omega) \end{array} \right);\;\;A_{in}(\Omega)=\left( \begin{array}{c} \sqrt{\gamma_{m1}}(\delta\tilde{\alpha}_{in}-\delta\tilde{\alpha}_V)+\sqrt{2 \gamma_{l1}}\delta\tilde{\alpha}_{l1} \\ \sqrt{\gamma_{m1}}(\delta\tilde{\alpha}^*_{in}-\delta\tilde{\alpha}^*_V)+\sqrt{2 \gamma_{l1}}\delta\tilde{\alpha}^*_{l1} \\ \delta\tilde{\psi}_{01}\\ \sqrt{\gamma_{m2}}(\delta\tilde{\alpha}_{in}+\delta\tilde{\alpha}_V)+\sqrt{2 \gamma_{l2}}\delta\tilde{\alpha}_{l2} \\ \sqrt{\gamma_{m2}}(\delta\tilde{\alpha}^*_{in}+\delta\tilde{\alpha}^*_V)+\sqrt{2 \gamma_{l2}}\delta\tilde{\alpha}^*_{l2} \\ \delta\tilde{\psi}_{02} \end{array}\right)$$ equations (\[dadt\]-\[ain2\]), (\[psi\]) can be written as $$\label{eqM} \mathbf{M}(\Omega)\cdot A(\Omega)=A_{in}(\Omega)$$ where $$\mathbf{M}(\Omega)= \left(\begin{array}{cccccc} -i\,\tau \,\Omega - i\bar{\psi}_1 + \gamma_1 & 0 & - i\,{{{\bar{\alpha}}}_1} & 0 & 0 & 0 \cr 0 & - i\,\tau \,\Omega + {{{i\bar{\psi} }}_1} + {{\gamma }_1} & i\, {{{{{\bar{\alpha}}}_1}}^*} & 0 & 0 & 0 \cr 4\,k^2\,\hbar \,{{\chi }_{11}}\,{{{{{\bar{\alpha}}}_1}}^*} & 4\,k^2\,\hbar \, {{\chi }_{11}}\,{{{\bar{\alpha}}}_1} & 1 & 4\,k^2\,\hbar \,{{\chi }_{12}}\,{{{{{\bar{\alpha}}}_2}}^*} & 4\,k^2\,\hbar \, \chi_{12}\,{{{\bar{\alpha}}}_2} & 0 \cr 0 & 0 & 0 & - i\,\tau \,\Omega - {{{i\bar{\psi} }}_2} + {{\gamma }_2} & 0 & - i\,{{{\bar{\alpha}}}_2} \cr 0 & 0 & 0 & 0 & - i\,\tau \,\Omega + {{{i\bar{\psi} }}_2} + {{\gamma }_2} & i\,{{{{{\bar{\alpha}}}_2}}^*} \cr 4\,k^2\,\hbar \,{{\chi }_{21}}\, {{{\bar{\alpha}}}_1} & 4\,k^2\,\hbar \,{{\chi }_{21}}\,{{{{{\bar{\alpha}}}_1}}^*} & 0 & 4\,k^2\,\hbar \,\chi_{22}\, {{{{{\bar{\alpha}}}_2}}^*} & 4\,k^2\,\hbar \,{{\chi }_{22}}\,{{{\bar{\alpha}}}_2} & 1 \cr \end{array}\right)\\$$ It is useful to write the field fluctuations in terms of amplitude and phase quadratures, defined as $$\delta p=\delta\tilde{\alpha}+\delta\tilde{\alpha}^* \, \,; \,\, \, \delta q=i(\delta\tilde{\alpha}^*-\delta\tilde{\alpha}) \, .$$ Eq. (\[eqM\]) becomes $$\label{sistemaN} \mathbf{N}(\Omega)\cdot X(\Omega)=X_{in}(\Omega)$$ with fluctuation vectors $X(\Omega)$ and $X_{in}(\Omega)$ and system matrix $\mathbf{N}(\Omega)$ given by $$\label{Xin} X(\Omega)= \left( \begin{array}{c} {{{\delta p}}_1}\\{{{\delta q}}_1}\\ {{{\delta \tilde{\psi} }}_1}\\{{{\delta p}}_2}\\ {{{\delta q}}_2}\\{{{\delta \tilde{\psi} }}_2}\end{array} \right);\;\;X_{in}(\Omega)=\left( \begin{array}{c} {\sqrt{{{2\gamma }_{l1}}}} {{{\delta p}}_{{l1}}} + {\sqrt{{{\gamma }_{m1}}}} \left( {{{\delta p}}_{{in}}} - {{{\delta p}}_V} \right) \\ {\sqrt{{{2\gamma }_{l1}}}} {{{\delta q}}_{{l1}}} + {\sqrt{{{\gamma }_{m1}}}} \left( {{{\delta q}}_{{in}}} - {{{\delta q}}_V} \right) \\ {{{\delta \tilde{\psi} }}_{{01}}}\\ {\sqrt{{{2\gamma }_{l2}}}} {{{\delta p}}_{{l2}}} + {\sqrt{{{\gamma }_{m2}}}} \left( {{{\delta p}}_{{in}}} + {{{\delta p}}_V} \right) \\ {\sqrt{{{2\gamma }_{l2}}}} {{{\delta q}}_{{l2}}} + {\sqrt{{{\gamma }_{m2}}}} \left( {{{\delta q}}_{{in}}} + {{{\delta q}}_V} \right) \\ {{{\delta \tilde{\psi} }}_{{02}}} \end{array}\right)$$ $$\mathbf{N}(\Omega)=\left(\begin{matrix} -i \tau_1 \Omega + {{\gamma }_1} & {{\bar{\psi} }_1} & \frac{2 {{\bar{\alpha} }_{ {in}}} {\sqrt{{{\gamma }_{ {m1}}}}} {{\bar{\psi} }_1}}{{{{\gamma }_1}}^2 + {{{\bar{\psi} }_1}}^2} & 0 & 0 & 0 \cr -{{\bar{\psi} }_1} & -i \tau_1 \Omega + {{\gamma }_1} & \frac{-2 {{\bar{\alpha} }_{ {in}}} {{\gamma }_1} {\sqrt{{{\gamma }_{ {m1}}}}}}{{{{\gamma }_1}}^2 + {{{\bar{\psi} }_1}}^2} & 0 & 0 & 0 \cr \frac{-4 k^2 \hbar {{\bar{\alpha} }_{ {in}}} {{\gamma }_1} {\sqrt{{{\gamma }_{ {m1}}}}} {{\chi }_1}}{{{{\gamma }_1}}^2 + {{{\bar{\psi} }_1}}^2} & \frac{-4 k^2 \hbar {{\bar{\alpha} }_{ {in}}} {\sqrt{{{\gamma }_{ {m1}}}}} {{\chi }_1} {{\bar{\psi} }_1}}{{{{\gamma }_1}}^2 + {{{\bar{\psi} }_1}}^2} & 1 & \frac{-4 k^2 \hbar {{\bar{\alpha} }_{ {in}}} {{\gamma }_2} {\sqrt{{{\gamma }_{ {m2}}}}} {{\chi }_{12}}}{{{{\gamma }_2}}^2 + {{{\bar{\psi} }_2}}^2} & \frac{-4 k^2 \hbar {{\bar{\alpha} }_{ {in}}} {\sqrt{{{\gamma }_{ {m2}}}}} {{\chi }_{12}} {{\bar{\psi} }_2}}{{{{\gamma }_2}}^2 + {{{\bar{\psi} }_2}}^2} & 0 \cr 0 & 0 & 0 & -i \tau_2 \Omega + {{\gamma }_2} & {{\bar{\psi} }_2} & \frac{2 {{\bar{\alpha} }_{ {in}}} {\sqrt{{{\gamma }_{ {m2}}}}} {{\bar{\psi} }_2}}{{{{\gamma }_2}}^2 + {{{\bar{\psi} }_2}}^2} \cr 0 & 0 & 0 & -{{\bar{\psi} }_2} & -i \tau_2 \Omega + {{\gamma }_2} & \frac{-2 {{\bar{\alpha} }_{ {in}}} {{\gamma }_2} {\sqrt{{{\gamma }_{ {m2}}}}}}{{{{\gamma }_2}}^2 + {{{\bar{\psi} }_2}}^2} \cr \frac{-4 k^2 \hbar {{\bar{\alpha} }_{ {in}}} {{\gamma }_1} {\sqrt{{{\gamma }_{ {m1}}}}} {{\chi }_{21}}}{{{{\gamma }_1}}^2 + {{{\bar{\psi} }_1}}^2} & \frac{-4 k^2 \hbar {{\bar{\alpha} }_{ {in}}} {\sqrt{{{\gamma }_{ {m1}}}}} {{\chi }_{21}} {{\bar{\psi} }_1}}{{{{\gamma }_1}}^2 + {{{\bar{\psi} }_1}}^2} & 0 & \frac{-4 k^2 \hbar {{\bar{\alpha} }_{ {in}}} {{\gamma }_2} {\sqrt{{{\gamma }_{ {m2}}}}} {{\chi }_2}}{{{{\gamma }_2}}^2 + {{{\bar{\psi} }_2}}^2} & \frac{-4 k^2 \hbar {{\bar{\alpha} }_{ {in}}} {\sqrt{{{\gamma }_{ {m2}}}}} {{\chi }_2} {{\bar{\psi} }_2}}{{{{\gamma }_2}}^2 + {{{\bar{\psi} }_2}}^2} & 1 \cr \end{matrix}\right)\\$$ From the boundary conditions (\[aout\]) we obtain the quadrature fluctuations of the reflected fields, according to $$\begin{aligned} \delta p_{1}^{out}&=&-\frac{{\delta p_{in}-\delta p_V}}{\sqrt{2}}+\sqrt{2\gamma_{m1}}\delta p_1\\\delta q_{1}^{out}&=&-\frac{{\delta q_{in}-\delta q_V}}{\sqrt{2}}+\sqrt{2\gamma_{m1}}\delta q_1\\ \delta p_{2}^{out}&=&-\frac{{\delta p_{in}+\delta p_V}}{\sqrt{2}}+\sqrt{2\gamma_{m2}}\delta p_2\\\delta q_{2}^{out}&=&-\frac{{\delta q_{in}+\delta q_V}}{\sqrt{2}}+\sqrt{2\gamma_{m2}}\delta q_2 \label{boundaryphase}\end{aligned}$$ Finally, at the output port of the beam splitter, field quadratures are recombined according to $$\begin{aligned} \label{outBS} \delta p_{BS}&=& \frac{1}{\sqrt{2}}(-\delta p_{1}^{out}+\delta p_{2}^{out}\cos{\theta}+\delta q_{2}^{out}\sin{\theta})\\\nonumber \delta q_{BS}&=& \frac{1}{\sqrt{2}}(-\delta q_{1}^{out}+\delta q_{2}^{out}\cos{\theta}+\delta p_{2}^{out}\sin{\theta})\end{aligned}$$ where $\theta$ is the phase shift introduced by the interferometer unbalance. Symmetric interferometer ------------------------ We will solve the above equations in the case of identical cavities and balanced Michelson interferometer. Therefore we replace $\gamma_{m1} = \gamma_{m2} = \gamma_m$, $\gamma_{l1} = \gamma_{l2} = \gamma_l$, $\tau_1 = \tau_2 = \tau$, and we set $\theta = 0$. The field quadratures fluctuations at the output port of the beam splitter can be expressed by the product between input fluctuation $X_{in}$ and two vectors of coefficients $P_{BS}[\delta x_i]$ and $Q_{BS}[\delta x_i]$, where $\delta x_i$ is the generic input fluctuation, with $i=$1 to 10: $\delta p_{BS}=P_{BS}\cdot X_{in}$ and $\delta q_{BS}=Q_{BS}\cdot X_{in}$. The vectors $P_{BS}$ and $Q_{BS}$, obtained from Eqs. (\[sistemaN\]-\[outBS\]), are the following: $$\begin{aligned} \Delta \cdot P_{BS}[\delta p_{in}]&=&\frac{1}{{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\Big\{ -64 k^4 \gamma {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^3 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} ( {{{\bar{\psi} }_1}}^2 - {{{\bar{\psi} }_2}}^2 ) \\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 \Big[ {{\chi }_1} {{\bar{\psi} }_1} { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 [ ( \gamma - i \tau \Omega ) {{{\bar{\psi} }_1}}^2 + \gamma ( ( \gamma - i \tau \Omega ) ( 2 \gamma - i \tau \Omega ) + {{{\bar{\psi} }_2}}^2 ) ] \\\nonumber &-&{{\chi }_2} {{\bar{\psi} }_2} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 [ \gamma {{{\bar{\psi} }_1}}^2 + ( \gamma - i \tau \Omega ) ( 2 {\gamma }^2 - i \gamma \tau \Omega + {{{\bar{\psi} }_2}}^2 ) ] \Big]\\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 )( 2 \gamma - i \tau \Omega ) [ {{\chi }_{12}} {{\bar{\psi} }_1} ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_2}}^2 ) \\\nonumber &-& {{\chi }_{21}} {{\bar{\psi} }_2} ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_1}}^2 ) ] \\\nonumber &+& {{\gamma }_m}( \gamma - i \tau \Omega ) { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{ ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 ( {{{\bar{\psi} }_1}}^2 - {{{\bar{\psi} }_2}}^2 ) \Big\}\end{aligned}$$ $$\begin{aligned} \Delta \cdot P_{BS}[\delta p_{V}]&=&\frac{1}{{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\times\\\nonumber&\times&\Big\{ -64 k^4 {\hbar}^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^2 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} [ ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) \\\nonumber &+& \gamma {{\gamma }_m} ( 2 {\gamma }^2 + {{{\bar{\psi} }_1}}^2 + {{{\bar{\psi} }_2}}^2 ) ]\\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} \Big[ {{\chi }_2}{{\bar{\psi} }_2} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 [ - \gamma {{\gamma }_m} ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ) + ( -{ ( \gamma - i \tau \Omega ) }^2 \\\nonumber &+& ( \gamma - i \tau \Omega ) {{\gamma }_m} - {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ] - {{\chi }_1} {{\bar{\psi} }_1} { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 [ ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) \\\nonumber &-& {{\gamma }_m} ( i \gamma \tau \Omega ( \gamma -i \tau \Omega ) + ( \gamma - i \tau \Omega ) {{{\bar{\psi} }_1}}^2 - \gamma {{{\bar{\psi} }_2}}^2 ) ] \Big] \\\nonumber&+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( 2 \gamma - i \tau \Omega ) [ {{\chi }_{21}} {{\bar{\psi} }_2} ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_1}}^2 ) \\\nonumber &+& {{\chi }_{12}} {{\bar{\psi} }_1} ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_2}}^2 ) ] \\\nonumber &-& { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 \Big[ ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) \\\nonumber &-& {{\gamma }_m} ( \gamma - i \tau \Omega ) [2 { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 + {{{\bar{\psi} }_2}}^2 ] \Big] \Big\}\end{aligned}$$ $$\begin{aligned} \Delta \cdot P_{BS}[\delta q_{in}]&=&\frac{1}{{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\times\\\nonumber&\times&\Big\{ -64k^4{\hbar }^2{{{\bar{\alpha} }_{{in}}}}^4{{{\gamma }_m}}^3 ( {{\chi }_1}{{\chi }_2} - {{\chi }_{12}}{{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} ( {{\bar{\psi} }_1} - {{\bar{\psi} }_2} )( {{\bar{\psi} }_1}{{\bar{\psi} }_2} -{\gamma }^2 ) \\\nonumber&-& 8k^2\hbar {{{\bar{\alpha} }_{{in}}}}^2{{{\gamma }_m}}^2 \Big[ {{\chi }_1}{{\bar{\psi} }_1}{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 [ {\gamma }^2{{\bar{\psi} }_2} + {{{\bar{\psi} }_1}}^2{{\bar{\psi} }_2} - {{\bar{\psi} }_1}( {( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) ]\\\nonumber&-& {{\chi }_2}{{\bar{\psi} }_2}{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 [ {\gamma }^2{{\bar{\psi} }_1} + {{\bar{\psi} }_1}{{{\bar{\psi} }_2}}^2- {{\bar{\psi} }_2} ( {( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ) ] \Big] \\\nonumber&-& 8i k^2\hbar {{{\bar{\alpha} }_{{in}}}}^2 {{{\gamma }_m}}^2\tau \Omega ( {{\chi }_{12}} - {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2}( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( 2\gamma - i \tau \Omega ) \\\nonumber&+& {{\gamma }_m}{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 {( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2( {{\bar{\psi} }_1} - {{\bar{\psi} }_2} ) ( {( \gamma - i \tau \Omega ) }^2 - {{\bar{\psi} }_1}{{\bar{\psi} }_2} ) \Big\}\end{aligned}$$ $$\begin{aligned} \Delta \cdot P_{BS}[\delta q_{V}]&=&\frac{1}{{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\times\\\nonumber&\times&\Big\{-64 k^4 {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^3 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} ( {{\bar{\psi} }_1} + {{\bar{\psi} }_2} ) ( {\gamma }^2 + {{\bar{\psi} }_1} {{\bar{\psi} }_2} ) \\\nonumber &-& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 \Big[{{\chi }_1} {{\bar{\psi} }_1} { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 [ {\gamma }^2 {{\bar{\psi} }_2} + {{{\bar{\psi} }_1}}^2 {{\bar{\psi} }_2} + {{\bar{\psi} }_1} ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) ]\\\nonumber &+& {{\chi }_2} {{\bar{\psi} }_2} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 [ {\gamma }^2 {{\bar{\psi} }_1}+ {{\bar{\psi} }_1} {{{\bar{\psi} }_2}}^2 + {{\bar{\psi} }_2} ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ) ] \Big] \\\nonumber &-& 8 i k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 \tau \Omega ( {{\chi }_{12}} + {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} ( {\gamma }^2 +{{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( 2 \gamma - i \tau \Omega ) \\\nonumber &-& {{\gamma }_m} { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ( {{\bar{\psi} }_1} + {{\bar{\psi} }_2} ) ( { ( \gamma - i \tau \Omega ) }^2 + {{\bar{\psi} }_1} {{\bar{\psi} }_2} ) \Big\}\end{aligned}$$ $$\begin{aligned} \Delta \cdot P_{BS}[\delta p_{l1}]&=&\frac{ {\sqrt{\gamma - {{\gamma }_m}}} {\sqrt{2\gamma_m}} ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 )} {{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\Big\{64 k^4 \gamma {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^2 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} \\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} [ {{\chi }_1} {{\bar{\psi} }_1} \gamma ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) - {{\chi }_2} {{\bar{\psi} }_2} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 )^2 ( \gamma - i \tau \Omega ) ] \\\nonumber &-& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} {{\chi }_{21}} {{\bar{\psi} }_2} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 )( 2 \gamma - i \tau \Omega ) ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_1}}^2 ) \\\nonumber &-& ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ( { ( \gamma - i \tau \Omega ) }^3 + ( \gamma - i \tau \Omega ) {{{\bar{\psi} }_2}}^2 ) \Big\}\end{aligned}$$ $$P_{BS}[\delta p_{l2}]\,=\,-P_{BS}[\delta p_{l1}] \,\,\,\,\,\,\,\,\, (1\,\leftrightarrow \,2)$$ $$\begin{aligned} \Delta \cdot P_{BS}[\delta q_{l1}]&=&\frac{ {\sqrt{\gamma - {{\gamma }_m}}} {\sqrt{2\gamma_m}} {{\bar{\psi} }_1}( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 )} {{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\Big\{64 k^4 {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^2 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2}\\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} [ {{\chi }_1} {{\bar{\psi} }_1} ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) + {{\chi }_2} {{\bar{\psi} }_2} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ] \\\nonumber &+& 8 i k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 \tau \Omega {{\gamma }_m} {{\chi }_{21}} {{\bar{\psi} }_2} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( 2 \gamma - i \tau \Omega ) \\\nonumber &+& ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) \Big\}\end{aligned}$$ $$P_{BS}[\delta q_{l2}]\,=\,-P_{BS}[\delta q_{l1}] \,\,\,\,\,\,\,\,\, (1\,\leftrightarrow \,2)$$ $$\begin{aligned} \Delta \cdot P_{BS}[\delta \psi_{01}]&=&\frac{2 \bar{\alpha}_{in}\gamma_m} {{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }}\Big\{8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} ( {{\chi }_2} + {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2}( 2 \gamma - i \tau \Omega ) \\\nonumber &+& {{\bar{\psi} }_1} ( 2 \gamma - i \tau \Omega ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 )\Big\}\end{aligned}$$ $$P_{BS}[\delta \psi_{02}]\,=\,-P_{BS}[\delta \psi_{01}] \,\,\,\,\,\,\,\,\, (1\,\leftrightarrow \,2)$$ $$\begin{aligned} \Delta \cdot Q_{BS}[\delta p_{in}]&=&\frac{1}{{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\Big\{ 64 k^4 {\gamma }^2 {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^3 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) \\\nonumber &\times&[ {{\bar{\psi} }_1}({\gamma }^2 + {{{\bar{\psi} }_1}}^2) - {{\bar{\psi} }_2} ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ] \\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 \Big[ {{\chi }_2} [ {\gamma }^2 ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 )^2 ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ) \\\nonumber &-& {{\bar{\psi} }_1} {{\bar{\psi} }_2}{ ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ] - {{\chi }_1} [ {\gamma }^2 ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 )^2 ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) \\\nonumber &-&{{\bar{\psi} }_1} {{\bar{\psi} }_2} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2] \Big]\\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) [ {{\chi }_{21}} ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_1}}^2 ) ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_2}}^2 ) \\\nonumber &-& {{\chi }_{12}} ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_1}}^2 ) ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_2}}^2 ) ] \\\nonumber &+& {{\gamma }_m} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 ( {{\bar{\psi} }_1} - {{\bar{\psi} }_2} ) ( -{ ( \gamma - i \tau \Omega ) }^2 + {{\bar{\psi} }_1} {{\bar{\psi} }_2} ) \Big\}\end{aligned}$$ $$\begin{aligned} \Delta \cdot Q_{BS}[\delta p_{V}]&=&\frac{1}{{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\Big\{ 64 k^4 {\gamma }^2 {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^3 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) \\\nonumber &\times&[ {{\bar{\psi} }_1} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) + {{\bar{\psi} }_2} ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ] \\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 \Big[ {{\chi }_2} [ {\gamma }^2 { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ) + {{\bar{\psi} }_1} {{\bar{\psi} }_2}{ ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ] \\\nonumber &+& {{\chi }_1} [ {\gamma }^2 { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 )+{{\bar{\psi} }_1} {{\bar{\psi} }_2} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 ] \Big]\\\nonumber &-& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 [ {{\chi }_{21}} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_1}}^2 ) ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_2}}^2 ) \\\nonumber &+& {{\chi }_{12}} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_1}}^2 ) ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_2}}^2 ) ] \\\nonumber &+& {{\gamma }_m} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 ( {{\bar{\psi} }_1} + {{\bar{\psi} }_2} ) ( { ( \gamma - i \tau \Omega ) }^2 + {{\bar{\psi} }_1} {{\bar{\psi} }_2} ) \Big\}\end{aligned}$$ $$\begin{aligned} \Delta \cdot Q_{BS}[\delta q_{in}]&=&\frac{1}{{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\Big\{64 k^4 \gamma {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^3 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} ( {{{\bar{\psi} }_1}}^2 - {{{\bar{\psi} }_2}}^2 ) \\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 \Big[ {{\chi }_1} {{\bar{\psi} }_1} { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 [ ( \gamma - i \tau \Omega ) (i \gamma \tau \Omega+ {{{\bar{\psi} }_1}}^2) - \gamma {{{\bar{\psi} }_2}}^2 ] \\\nonumber &+& {{\chi }_2}{{\bar{\psi} }_2} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 [ \gamma {{{\bar{\psi} }_1}}^2 - ( \gamma - i \tau \Omega ) ( i \gamma \tau \Omega + {{{\bar{\psi} }_2}}^2 ) ] \Big] \\\nonumber &+& 8 i k^2 \tau \Omega \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) [ {{\chi }_{12}} {{\bar{\psi} }_2} ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_1}}^2 )\\\nonumber&-& {{\chi }_{21}} {{\bar{\psi} }_1} ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_2}}^2 ) ] + {{\gamma }_m} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 ( \gamma - i \tau \Omega ) ( {{{\bar{\psi} }_1}}^2 - {{{\bar{\psi} }_2}}^2 ) \Big\}\end{aligned}$$ $$\begin{aligned} \Delta \cdot Q_{BS}[\delta q_{V}]&=&\frac{1}{{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\times\\\nonumber&\times&\Big\{ -64 k^4 {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^2 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} \\\nonumber &\times&[ ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) - \gamma {{\gamma }_m} ( 2 {\gamma }^2 + {{{\bar{\psi} }_1}}^2 + {{{\bar{\psi} }_2}}^2 ) ]\\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} \Big[ - {{\chi }_2}{{\bar{\psi} }_2} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 [ ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ) \\\nonumber &-& {{\gamma }_m} ( \gamma {{{\bar{\psi} }_1}}^2 + ( \gamma - i \tau \Omega ) ( 2 {\gamma }^2 - i \gamma \tau \Omega + {{{\bar{\psi} }_2}}^2 ) ) ] \\\nonumber &-& {{\chi }_1} {{\bar{\psi} }_1} { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 [ ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) \\\nonumber &-& {{\gamma }_m} ( \gamma {{{\bar{\psi} }_1}}^2 + ( \gamma - i \tau \Omega ) ( 2 {\gamma }^2 - i \gamma \tau \Omega + {{{\bar{\psi} }_2}}^2 ) ) - i \tau \Omega ( {{{\bar{\psi} }_1}}^2 - {{{\bar{\psi} }_2}}^2 ) ] \Big] \\\nonumber &+& 8 i k^2 \tau \Omega \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{{\gamma }_m}}^2 ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) [ {{\chi }_{12}} {{\bar{\psi} }_2} ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_1}}^2 ) \\\nonumber &+& {{\chi }_{21}} {{\bar{\psi} }_1} ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_2}}^2 ) ] \\\nonumber &-& { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2 [ ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) \\\nonumber &-& {{\gamma }_m}( \gamma - i \tau \Omega ) ( 2 { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 + {{{\bar{\psi} }_2}}^2 ) ]\Big\}\end{aligned}$$ $$\begin{aligned} \Delta \cdot Q_{BS}[\delta p_{l1}]&=&\frac{ {\sqrt{\gamma - {{\gamma }_m}}} {\sqrt{2\gamma_m}} ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 )} {{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\Big\{-64 k^4 {\gamma }^2 {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^2 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_2}\\\nonumber &-& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} [ {{\chi }_1} {\gamma }^2 ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) +{{\chi }_2} {{\bar{\psi} }_1} {{\bar{\psi} }_2} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2] \\\nonumber &+& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} {{\chi }_{21}}( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_2}}^2 ) ( \gamma ( \gamma - i \tau \Omega ) + {{{\bar{\psi} }_1}}^2 ) \\\nonumber &+& {{\bar{\psi} }_1} ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ( { ( i \gamma + \tau \Omega ) }^2 - {{{\bar{\psi} }_2}}^2 ) \Big\}\end{aligned}$$ $$Q_{BS}[\delta p_{l2}]\,=\,-Q_{BS}[\delta p_{l1}] \,\,\,\,\,\,\,\,\, (1\,\leftrightarrow \,2)$$ $$\begin{aligned} \Delta \cdot Q_{BS}[\delta q_{l1}]&=&\frac{ {\sqrt{\gamma - {{\gamma }_m}}} {\sqrt{2\gamma_m}} ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 )} {{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }^2}\Big\{-64 k^4 \gamma {\hbar }^2 {{{\bar{\alpha} }_{ {in}}}}^4 {{{\gamma }_m}}^2 ( {{\chi }_1} {{\chi }_2} - {{\chi }_{12}} {{\chi }_{21}} ) {{\bar{\psi} }_1} {{\bar{\psi} }_2} \\\nonumber &-& 8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} [ {{\chi }_1} {{\bar{\psi} }_1} \gamma ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ) + {{\chi }_2}{{\bar{\psi} }_2} { ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 ( \gamma - i \tau \Omega ) ] \\\nonumber &-& 8 i k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 \tau \Omega {{\gamma }_m} {{\chi }_{21}} {{\bar{\psi} }_1} ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) ( {\gamma }^2 - i \gamma \tau \Omega - {{{\bar{\psi} }_2}}^2 ) \\\nonumber &-& { ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 )( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }^2 [ { ( \gamma - i \tau \Omega ) }^3 + ( \gamma - i \tau \Omega ) {{{\bar{\psi} }_2}}^2 ] \Big\}\end{aligned}$$ $$Q_{BS}[\delta q_{l2}]\,=\,-Q_{BS}[\delta q_{l1}] \,\,\,\,\,\,\,\,\, (1\,\leftrightarrow \,2)$$ $$\begin{aligned} \Delta \cdot Q_{BS}[\delta \psi_{01}]&=&\frac{2 \bar{\alpha}_{in}\gamma_m} {{( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) }{( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) }}\times\\\nonumber&\times&\Big\{ -8 k^2 \hbar {{{\bar{\alpha} }_{ {in}}}}^2 {{\gamma }_m} [ {{\chi }_2} {{\bar{\psi} }_2} ( \gamma ( \gamma - i \tau \Omega ) -{{{\bar{\psi} }_1}}^2 ) + {{\chi }_{21}} {{\bar{\psi} }_1} ( \gamma ( \gamma - i \tau \Omega ) - {{{\bar{\psi} }_2}}^2 ) ]\\\nonumber&-& ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) ( { ( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 )( \gamma ( \gamma -i \tau \Omega ) - {{{\bar{\psi} }_1}}^2) \Big\}\end{aligned}$$ $$Q_{BS}[\delta \psi_{02}]\,=\,-Q_{BS}[\delta \psi_{01}] \,\,\,\,\,\,\,\,\, (1\,\leftrightarrow \,2)$$ where $\Delta$ is the determinant of $\mathbf{N}$, expressed by $$\begin{aligned} \Delta&=&\frac{1}{(\gamma^2 + \bar{\psi}_1^2)(\gamma^2 + \bar{\psi}_2^2 )}\Big\{64k^4\hbar^2\bar{\alpha}_{in}^4\gamma_m^2 (\chi_1\chi_2 - \chi_{12}\chi_{21}) {{\bar{\psi} }_1}{{\bar{\psi} }_2}\\\nonumber&+&( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) [ {( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ] ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) [ {( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ]+\\\nonumber&& 8k^2\hbar {{\bar{\alpha} }_{{in}}}^2 \gamma_m \big[\chi_2 \bar{\psi}_2 ( {\gamma }^2 + {{{\bar{\psi} }_1}}^2 ) [ {( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_1}}^2 ]+\\\nonumber&& {{\chi }_1}{{\bar{\psi} }_1} ( {\gamma }^2 + {{{\bar{\psi} }_2}}^2 ) [ {( \gamma - i \tau \Omega ) }^2 + {{{\bar{\psi} }_2}}^2 ]\big]\Big\}.\end{aligned}$$ The homodyne detection after the beam splitter allows to choose the output quadrature $\delta E_{out}=\delta p_{BS} \cos{w} + \delta q_{BS} \sin{w} $, according to the detection phase $w$. It is useful to describe also $\delta E_{out}$ as product $V_{out} \cdot X_{in}$, with a coefficients vector $V_{out}$ defined as $$\label{Vout} V_{out} = P_{BS} \cos{w} + Q_{BS} \sin{w} \, .$$ The input fluctuations $\delta \psi_{01}$ and $\delta \psi_{02}$ contain the gw signal. They also include thermal noise and any kind of classical fluctuations of the cavities length, that we are not considering in this article since in our approach they are not distinguishable from the gw signal. Taking an optimal interferometer orientation, the gw signal is proportional to $\delta (L_1-L_2) = (2k)^{-1}(\delta\psi_{01}-\delta\psi_{02})$. Therefore we define a sensitivity $S_L(\Omega)$ as $$S_L(\Omega)=\frac{\Sigma_{i=1,8} |V_{out}[\delta x_i]|^2 \cdot S_{x_i x_i}}{(2k)^2|V_{out}[\delta\psi_{01}]-V_{out}[\delta\psi_{02}]|^2}$$ where $S_{x_i x_i}$ are the spectral densities of the input fluctuations $\delta x_i$, which are assumed uncorrelated. We take all noise spectra double-sided and normalized to shot noise. Therefore, for all the vacuum fluctuations (i.e., for $\delta p_V, \delta q_V, \delta p_{l1}, \delta q_{l1}, \delta p_{l2}, \delta q_{l2})$ the spectra are $S_{x_i x_i}=1$. In totally symmetric conditions, we have $\bar{\psi}_1=\bar{\psi}_2=\bar{\psi}$, $\chi_{11}=\chi_{22}=\chi_s$ and $\chi_{12}=\chi_{21}=\chi_c$. To write clearer expressions, we define $\Psi = \bar{\psi}/\gamma$ (detuning normalized to the half cavity linewidth); $P_{in} = \hbar kc\bar{\alpha}_{in}^2$ (input laser power); $\Gamma_m = \gamma_m/\gamma$; $\Omega_{cav}=\gamma/\tau$ (cutoff angular frequency of the cavity). The expression of $S_L(\Omega)$ is $$\begin{aligned} \label{NSsimm} &&S_L= \left\{\sqrt{\Gamma_m}\,p \left[ {( 2\Psi \cos{w} + ({\Psi }^2-1)\sin{w} ) }^2 + \left(\frac{\Omega}{\Omega_{cav}}\right)^2(\sin{w} -\Psi \cos{w})^2 \right] \right\}^{-1} \\\nonumber&\times&\hbar \Big\{ \frac{p^2}{2}|(\chi_s-\chi_c)|^2 \left[ 1 + \frac{3{\Psi }^2-1}{1 + {\Psi }^2} \cos{2w}+ \frac{\Psi ( {\Psi }^2-3)} {1 + {\Psi }^2} \sin{2w} + \frac{{\Psi }^2}{2{{\Gamma }_m}} \right] \\\nonumber &+& \frac{p}{{\sqrt{{{\Gamma }_m}}}}\Big[ {Re(\chi_s-\chi_c) }\Big( \Psi ( 1 + {\Psi }^2 ) \left(1+{\Psi }^2-\left(\frac{\Omega}{\Omega_{cav}}\right)^2\right) \\\nonumber &+& 2\Psi {{{\Gamma }}_m}\left( 2{\Psi }^2 -2 - \left(\frac{\Omega}{\Omega_{cav}}\right)^2 \right) \cos{2w} \\\nonumber &+& {{{\Gamma }}_m} \left( 1 + {\Psi }^4 - {\Psi }^2 \left(6+\left(\frac{\Omega}{\Omega_{cav}}\right)^2\right)+ \left(\frac{\Omega}{\Omega_{cav}}\right)^2 \right) \sin{2w} \Big) \\\nonumber &+& {Im(\chi_s-\chi_c) }\Big( 2\Psi ( 1 + {\Psi }^2 ) \frac{\Omega} {{{\Omega }_{{cav}}}} ({{\Gamma }_m} -1 ) \Big) \Big] \\\nonumber &+& {( 1 + {\Psi }^2 ) }^2\left[ {\Psi }^4 + 2{\Psi }^2\left( 1 - \left(\frac{\Omega}{\Omega_{cav}}\right)^2\right) + \left(1+\left(\frac{\Omega}{\Omega_{cav}}\right)^2 \right)^2 \right]\Big\}.\end{aligned}$$ We notice that, in this case, $V_{out}[\delta\psi_{01}]=V_{out}[\delta\psi_{02}]$ and the detection is not sensitive to the common mode of the interferometer. In the short cavity regime, with $\Omega/\Omega_{cav}\simeq 0$, the Eq. (\[NSsimm\]) reduces to $$\begin{aligned} \label{NS0tau} S_{L}^{0}&=&\Big\{{\sqrt{{{\Gamma }_m}}}\,p \left[2\Psi \cos{w} + ({\Psi }^2-1)\sin{w} \right]^2 \Big\}^{-1}\\\nonumber&\times&\hbar \Big\{ \frac{p^2}{2}|(\chi_s-\chi_c)|^2 \left[ 1 + \frac{3{\Psi }^2-1}{1 + {\Psi }^2} \cos{2w}+ \frac{\Psi ( {\Psi }^2-3)}{ 1 + {\Psi }^2} \sin{2w} + \frac{{\Psi }^2}{2{{\Gamma }_m}} \right] \\\nonumber &+& \frac{p}{{\sqrt{{{\Gamma }_m}}}}{Re(\chi_s-\chi_c) }\Big[ \Psi {( 1 + {\Psi }^2 ) }^2 + {{{\Gamma }}_m} \sin{2w} \big( 1 + {\Psi }^4 - 6{\Psi }^2 \big) \\\nonumber &+& 4{{{\Gamma }}_m}\Psi ({\Psi }^2-1) \cos{2w} \Big]+ {( 1 + {\Psi }^2 ) }^4 \Big\} \,\,\,. \end{aligned}$$ [99]{} W. G. Unruh, in [*Quantum Optics, Experimental Gravitation, and Measurement Theory*]{}, edited by P. Meystre and M.O. Scully (Plenum, New York, 1982). M. T. Jaekel and S. Reynaud, Europhys. Lett. [**13**]{}, 301 (1990). J. M. Courty, A. Heidmann, and M. Pinard, Phys. Rev. Lett. [**90**]{}, 083601 (2003). V.B. Braginsky, M.L. Gorodetsky, F.Y. Khalili, Phys. Lett. A [**232**]{}, 340 (1997). F.Y. Khalili, Phys. Lett. A [**298**]{}, 308 (2002). S.L. Danilishin and F.Y. Khalili, Phys. Rev. D [**73**]{}, 022002 (2006). B. J. Meers, Phys. Rev. D [**38**]{}, 2317 (1988). A. Buonanno and Y. Chen, Phys. Rev. D [**64**]{}, 042006 (2001); Class. Quantum Grav. [**18**]{}, L95 (2001); Phys. Rev. D [**65**]{}, 042001 (2002). J. Harms, Y. Chen, S. Chelkowski, A. Franzen, H. Vahlbruch, K. Danzmann, R. Schnabel, Phys. Rev. D [**68**]{}, 042001 (2003). M. Levenson, R. Shelby, and S. Perlmutter, Opt. Lett. [**10**]{}, 514 (1985). P. Galatola, L. Lugiato, M. Porreca, P. Tombesi, and G. Leuchs, Opt. Commun. [**85**]{}, 95 (1991). C. M. Caves, Phys. Rev. D [**23**]{}, 1693 (1981). S. Mancini, P. Tombesi, Phys. Rev. A [**49**]{}, 4055 (1994). C. Fabre, M. Pinard, S. Bourzeix, A. Heidmann, E. Giacobino, S. Reynaud, Phys. Rev. A [**49**]{}, 1337 (1994). T. Corbitt, Y. Chen, F. Khalili, D. Ottaway, S. Vyatchanin, S. Whitcomb, and N. Mavalvala, Phys. Rev. A [**73**]{}, 023801 (2006). A. Heidmann, Y. Hadjar, M. Pinard, Appl. Phys. B [**64**]{}, 173 (1997). O. Arcizet, T. Briant, A. Heidmann and M. Pinard, Phys. Rev. A 73, 033819 (2006). M. Cerdonio, L. Conti, J. A. Lobo, A. Ortolan, L. Taffarello and J. P. Zendri, Phys. Rev. Lett. [**87**]{}, 031101 (2001). M. Bonaldi, M. Cerdonio, L. Conti, M. Pinard, G. A. Prodi, L. Taffarello, and J. P. Zendri, Phys. Rev. D [**68**]{}, 102004 (2003). F. Marin, in the Proceedings of the GWADW, Elba, Italy, 2006. F. Marin, L. Conti, M. De Rosa, Phys. Lett. A [**309**]{}, 15 (2003); Class. Quantum Grav. [**21**]{}, S1237 (2004). F. Acernese [*et al.*]{}, Class. Quantum Grav. [**23**]{}, S63 (2006). D. Sigg, Class. Quantum Grav. [**23**]{}, S51 (2006). L. Conti, M. De Rosa, F. Marin, Appl. Opt. [**39**]{}, 5732 (2000). The authors of Ref. [@Arcizet], besides a particular choice of detection phase, plot sensitivity curves maintaining a fixed product $P_{in}/(1+\Psi^2)^2$, thus considering an input power increasing with the detuning. S.P. Vyatchanin, E.A. Zubova, Phys. Lett. A [**201**]{}, 269 (1995). H. J. Kimble, Y. Levin, A. B. Matsko, K. S. Thorne, and S. P. Vyatchanin, Phys. Rev. D [**65**]{}, 022002 (2001). J. Mizuno, K.A. Strain, P.G. Nelson, J.M. Chen, R. Schilling, A. Rüdiger, W. Winkler and K. Danzmann, Phys. Lett. A [**175**]{}, 273 (1993); G. Heinzel, J. Mizuno, R. Schilling, W. Winkler, A. Rüdiger and K. Danzmann, Phys. Lett. A [**217**]{}, 305 (1996). J. M. Aguirregabiria, L. Bel, Phys. Rev. A [**36**]{}, 3768 (1987). F. Marino, F. Marin, Phys. Lett. A [**364**]{}, 441 (2007).
{ "pile_set_name": "ArXiv" }
--- abstract: 'A number of recent work studied the effectiveness of feature selection using Lasso. It is known that under the restricted isometry properties (RIP), Lasso does not generally lead to the exact recovery of the set of nonzero coefficients, due to the looseness of convex relaxation. This paper considers the feature selection property of nonconvex regularization, where the solution is given by a multi-stage convex relaxation scheme. Under appropriate conditions, we show that the local solution obtained by this procedure recovers the set of nonzero coefficients without suffering from the bias of Lasso relaxation, which complements parameter estimation results of this procedure in [@Zhang09-multistage].' author: - | Tong Zhang\ Statistics Department\ Rutgers University\ Piscataway, NJ 08854\ `[email protected]` bibliography: - 'L1.bib' - 'learning.bib' - 'pubj.bib' title: 'Multi-stage Convex Relaxation for Feature Selection' --- Introduction ============ We consider the linear regression problem, where we observe a set of input vectors ${{\mathbf x}}_1, \ldots, {{\mathbf x}}_n \in R^p$, with corresponding desired output variables $y_1, \ldots, y_n$. In a statistical linear model, it is common to assume that there exists a target coefficient vector $\bar{{{\mathbf w}}} \in R^p$ such that $$y_i = \bar{{{\mathbf w}}}^\top {{\mathbf x}}_i + \epsilon_i \qquad (i=1,\ldots,n) , \label{eq:noise-def}$$ where $\epsilon_i$ are zero-mean independent random noises (but not necessarily identically distributed). Moreover, we assume that the target vector $\bar{{{\mathbf w}}}$ is sparse. That is, $\bar{k}=\|\bar{{{\mathbf w}}}\|_0$ is small. Here we use the standard notation $${{\mathrm{supp}}}({{\mathbf w}})= \{j: {{\mathbf w}}_j \neq 0\} \qquad \|{{\mathbf w}}\|_0 = |{{\mathrm{supp}}}({{\mathbf w}})|$$ for any vector ${{\mathbf w}}\in R^p$. This paper focuses on the feature selection problem, where we are interested in estimating the set of nonzero coefficients ${{\mathrm{supp}}}(\bar{{{\mathbf w}}})$ (also called support set). Let ${{\mathbf y}}$ denote the vector of $[y_i]$ and $X$ be the $n \times d$ matrix with each row a vector ${{\mathbf x}}_i$. The standard statistical method is subset selection ($L_0$ regularization), which computes the following estimator $$\hat{{{\mathbf w}}}_{L_0} = \arg\min_{{{\mathbf w}}\in R^p} \|X {{\mathbf w}}- {{\mathbf y}}\|_2^2 \qquad \text{subject to } \|{{\mathbf w}}\|_0 \leq k , \label{eq:L0}$$ where $k$ is a tuning parameter. This method is arguably a natural method for feature selection because if noise $\epsilon_i$ are iid Gaussian random variables, then (\[eq:L0\]) can be regarded as a Bayes procedure with an appropriately defined sparse prior over ${{\mathbf w}}$. However, because the optimization problem in (\[eq:L0\]) is nonconvex, the global solution of this problem cannot be efficiently computed. In practice, one can only find an approximate solution of (\[eq:L0\]). The most popular approximation to $L_0$ regularization is the $L_1$ regularization method which is often referred to as Lasso [@Tib96]: $$\hat{{{\mathbf w}}}_{L_1} = \arg\min_{{{\mathbf w}}\in R^p} \left[ \frac{1}{n} \|X {{\mathbf w}}- {{\mathbf y}}\|_2^2 + \lambda \|{{\mathbf w}}\|_1 \right] , \label{eq:L1}$$ where $\lambda>0$ is an appropriately chosen regularization parameter. The global optimum of (\[eq:L1\]) can be easily computed using standard convex programming techniques. It is known that in practice, $L_1$ regularization often leads to sparse solutions (although often suboptimal). Moreover, its performance has been theoretically analyzed recently. For example, it is known from the compressed sensing literature (e.g., [@CandTao05-rip]) that under certain conditions referred to as [*restricted isometry property*]{} (RIP), the solution of $L_1$ relaxation (\[eq:L1\]) approximates the solution of the $L_0$ regularization problem (\[eq:L0\]). The prediction and parameter performance of this method has been considered in [@BunTsyWeg07; @BiRiTs07; @Koltchinskii08; @ZhangHuang06; @Zhang07-l1; @GeerBuhlmann09-conditions]. Exact support recovery was considered by various authors such as [@MeinBuh06; @ZhaoYu06; @Wainwright06]. It is known that under some more restrictive conditions referred to as [*irrepresentable conditions*]{}, $L_1$ regularization can achieve exact recovery of the support set. However, the $L_1$ regularization method (\[eq:L1\]) does not achieve exact recovery of the support set under the RIP type of conditions, which we are interested in here. Although it is possible to achieve exact recovery using post-processing by thresholding the small coefficients of Lasso solution, this method is suboptimal under RIP in comparison to the $L_0$ regularization method (\[eq:L0\]) because it requires the smallest nonzero coefficients to be $\sqrt{\bar{k}}$ times larger than the noise level instead of only requiring the nonzero coefficients to be larger than the noise level with $L_0$ regularization in (\[eq:L0\]). This issue, referred to as the [*bias*]{} of Lasso for feature selection, was extensively discussed in [@Zhang10-mc+]. Detailed discussion can be found after Theorem \[thm:multi-stage-featsel\]. It is worth mentioning that under a stronger mutual coherence condition (similar to irrepresentable condition), this post-processing step does not give this bias factor $\sqrt{\bar{k}}$ as shown in [@Lounici08] (also see [@Zhang07-l1]). Therefore the advantage of bias removal for the multi-stage procedure discussed here is only applicable when RIP holds but when the irrepresentable condition and mutual incoherence conditions fail. A thorough discussion of various conditions is beyond the scope of the current paper, and we would like to refer the readers to [@GeerBuhlmann09-conditions]. Nevertheless, it is worth pointing out that even in the classical $d<n$ setting with the design matrix $X$ being rank $d$, the irrepresentable condition or the mutual incoherence condition can still be violated while the RIP type sparse-eigenvalue condition used in this paper holds trivially. In fact, this was pointed out in [@Zou06] as the main motivation of adaptive Lasso. Adaptive Lasso behaves similarly to the above mentioned post-processing, and thus suffers from the same bias problem. The bias of Lasso is due to the looseness of convex relaxation for $L_0$ regularization. Therefore the remedy is to use a non-convex regularizer that is close to $L_0$ regularization. One drawback of using nonconvex optimization formulation is that we can only find a local optimal solution and different computational procedure may lead to a different local solution. Therefore the theoretical analysis has to be integrated with specific computational procedure to show that the local minimum obtained by the procedure has desirable properties (e.g., exact support recovery). Several nonconvex computational procedures have been analyzed in the literature, including an adaptive forward backward greedy procedure (referred to as FoBa) to approximately solve the regularization method (\[eq:L0\]) considered in [@Zhang08-foba], and the MC+ method in [@Zhang10-mc+] to solve a non-convex regularized problem using a path-following procedure. Both methods can achieve unbiased feature selection. Related to the above mentioned work, a different procedure, referred to as [*multi-stage convex relaxation*]{}, was analyzed in [@Zhang09-multistage]. This procedure solves a nonconvex problem using multiple stages of Lasso relaxations, where convex formulations are iteratively refined based on solutions obtained from the previous stages. However, only parameter estimation performance was analyzed in [@Zhang09-multistage]. Unfortunately, the result in [@Zhang09-multistage] does not directly imply that multi-stage convex relaxation achieves unbiased recovery of the support set. The purpose of this paper is to prove such a support recovery result analogous to related result in [@Zhang10-mc+] (which is for a different procedure), and this result complements the parameter estimation result of [@Zhang09-multistage]. Multi-Stage Convex Relaxation with Capped-$L_1$ Regularization {#sec:theory} ============================================================== We are interested in recovering $\bar{{{\mathbf w}}}$ from noisy observations ${{\mathbf y}}$ using the following nonconvex regularization formulation: $$\hat{{{\mathbf w}}} = \arg\min_{{{\mathbf w}}} \left[ \frac{1}{n}\| X {{\mathbf w}}- {{\mathbf y}}\|_2^2 + \lambda \sum_{j=1}^p g(|{{\mathbf w}}_j|) \right] , \label{eq:sparse-reg}$$ where $g(|{{\mathbf w}}_j|)$ is a regularization function. For simplicity, this paper only considers the specific regularizer $$g(u) = \min(u,\theta) , \label{eq:capped-L1-reg}$$ which is referred to as capped-$L_1$ regularization in [@Zhang09-multistage]. The parameter $\theta$ is a thresholding parameter which says that we use $L_1$ penalization when a coefficient is sufficiently small, but the penalty does not increase when the coefficient is larger than a threshold $\theta$. Detailed discussions can be found in [@Zhang09-multistage]. Similar to [@Zhang09-multistage], one can analyze general regularization function $g(u)$. However, some of such functions (such as adaptive Lasso) do not completely remove the bias. Therefore we only analyze the simple function (\[eq:capped-L1-reg\]) in this paper for clarity. While a theoretical justification has been given in [@Zhang09-multistage] for multi-stage convex relaxation, similar procedure has been shown to work well empirically without theoretical justification [@CanWakBoy08; @WipfNaga10]. Moreover, a two-stage version was proposed in [@ZouLi08], which does not remove the bias issue discussed in this paper. Since the regularizer (\[eq:capped-L1-reg\]) is nonconvex, the resulting optimization problem (\[eq:sparse-reg\]) is a non-convex regularization problem. However the regularizer in (\[eq:capped-L1-reg\]) is continuous and piecewise differentiable, and thus its solution is easier to compute than the $L_0$ regularization method in (\[eq:L0\]). For example, standard numerical techniques such as sub-gradient descent lead to local minimum solutions. Unfortunately, it is difficult to find the global optimum, and it is also difficult to analyze the quality of the local minimum obtained from the gradient descent method. As a matter of fact, results with non-convex regularization are difficult to reproduce because different numerical optimization procedures can lead to different local minima. Therefore the quality of the solution heavily depend on the numerical procedure used. In the following, we consider a specific numerical procedure referred to as multi-stage convex relaxation in [@Zhang09-multistage]. The algorithm is given in Figure \[fig:multi-stage-sparse\]. The procedure converges to a local optimal solution of (\[eq:sparse-reg\]) due to a simple concave duality argument, where (\[eq:sparse-reg\]) is rewritten as $$\hat{{{\mathbf w}}} = \arg\min_{{{\mathbf w}}} \min_{\{\lambda_j \geq 0\}}\left[ \frac{1}{n}\| X {{\mathbf w}}- {{\mathbf y}}\|_2^2 + \sum_{j=1}^p \lambda_j |{{\mathbf w}}_j| + \sum_{j=1}^p g^*(\lambda_j) \right] ,$$ with $g^*(\lambda_j)=\max((\lambda-\lambda_j)\theta,0)$. The procedure of Figure \[fig:multi-stage-sparse\] can be regarded as an alternating optimization method to solve this joint optimization problem of ${{\mathbf w}}$ and $\{\lambda_j\}$, where the first step solves for ${{\mathbf w}}$ with $\{\lambda_j\}$ fixed, and the second step is the closed form solution of $\{\lambda_j\}$ with ${{\mathbf w}}$ fixed. A more detailed discussion can be found in [@Zhang09-multistage]. Our goal is to show that this procedure can achieve unbiased feature selection as described in [@Zhang10-mc+]. Initialize $\lambda_j^{(0)}=\lambda$ for $j=1,\ldots,d$\ For $\ell=1,2,\ldots$ - Let $$\hat{{{\mathbf w}}}^{(\ell)} = \arg\min_{{{\mathbf w}}\in R^p} \left[ \frac{1}{n}\|X {{\mathbf w}}- {{\mathbf y}}\|_2^2 + \sum_{j=1}^p \lambda_j^{(\ell-1)} |{{\mathbf w}}_j| \right] . \label{eq:convex-relax-L1}$$ - Let $\lambda_j^{(\ell)} = \lambda I(|\hat{{{\mathbf w}}}_j^{(\ell)}| \leq \theta )$ ($j=1,\ldots,d$) Theoretical Analysis ==================== We require some technical conditions for our analysis. First we assume sub-Gaussian noise as follows. \[assump:fixed\] Assume that $\{\epsilon_i\}_{i=1,\ldots,n}$ in (\[eq:noise-def\]) are independent (but not necessarily identically distributed) sub-Gaussians: there exists $\sigma \geq 0$ such that $\forall i$ and $\forall t \in R$, $${{\mathbf E}}_{\epsilon_i} e^{t \epsilon_i} \leq e^{\sigma^2 t^2/2} .$$ Both Gaussian and bounded random variables are sub-Gaussian using the above definition. For example, if a random variable $\xi \in [a,b]$, then ${{\mathbf E}}_\xi e^{t (\xi - {{\mathbf E}}\xi)} \leq e^{(b-a)^2 t^2/8}$. If a random variable is Gaussian: $\xi \sim N(0,\sigma^2)$, then ${{\mathbf E}}_\xi e^{t \xi} \leq e^{\sigma^2 t^2/2}$. We also introduce the concept of sparse eigenvalue, which is standard in the analysis of $L_1$ regularization. Given $k$, define $$\begin{aligned} \rho_+(k)=&\sup \left\{\frac{1}{n}\|X {{\mathbf w}}\|_2^2/\|{{\mathbf w}}\|_2^2 : \|{{\mathbf w}}\|_0 \leq k \right\} ,\\ \rho_-(k)=&\inf \left\{\frac{1}{n}\|X {{\mathbf w}}\|_2^2/\|{{\mathbf w}}\|_2^2 : \|{{\mathbf w}}\|_0 \leq k \right\} .\end{aligned}$$ The following result for parameter estimation was obtained in [@Zhang09-multistage], under the Assumption \[assump:fixed\]. If we assume that the target $\bar{{{\mathbf w}}}$ is sparse, with ${{\mathbf E}}y_i = \bar{{{\mathbf w}}}^\top {{\mathbf x}}_i$, and $\bar{k}=\|\bar{{{\mathbf w}}}\|_0$, and we choose $\theta$ and $\lambda$ such that $$\lambda \geq 20 \sigma \sqrt{2\rho_+(1) \ln (2p/\eta)/n}$$ and $$\theta \geq 9 \lambda /\rho_-(2\bar{k}+s) .$$ Assume that $\rho_+(s)/\rho_-(2\bar{k}+2s) \leq 1+ 0.5 s/\bar{k}$ for some $s \geq 2 \bar{k}$, then with probability larger than $1-\eta$: $$\|\hat{{{\mathbf w}}}^{(\ell)} - \bar{{{\mathbf w}}}\|_2 \leq \frac{17}{\rho_-(2\bar{k}+s)} \left[ 2 \sigma \sqrt{\rho_+(\bar{k})} \left(\sqrt{\frac{7.4 \bar{k}}{n}} + \sqrt{\frac{2.7 \ln (2/\eta)}{n}} \right) +\lambda \sqrt{k_\theta} \right] + \frac{0.7^{\ell} \cdot \sqrt{\bar{k}} \lambda}{\rho_-(2 \bar{k}+s)} , \label{eq:multi-stage-param-est-bound}$$ where $\hat{{{\mathbf w}}}^{(\ell)}$ is the solution of (\[eq:convex-relax-L1\]), and $k_\theta = \left|\{j \in \bar{F}: |\bar{{{\mathbf w}}}_j| \leq 2\theta\}\right|$. The condition $\rho_+(s)/\rho_-(2\bar{k}+2s) \leq 1+ 0.5 s/\bar{k}$ requires the eigenvalue ratio $\rho_+(s)/\rho_-(s)$ to grow sub-linearly in $s$. Such a condition, referred to as [*sparse eigenvalue condition*]{}, is also needed in the standard analysis of $L_1$ regularization [@ZhangHuang06; @Zhang07-l1]. It is related but slightly weaker than the RIP condition in compressive sensing [@CandTao05-rip], which requires the condition $$1 - \delta_{s'} \leq \rho_-(s') \leq \rho_+(s') \leq 1 + \delta_{s'} ,$$ for some $\delta_{s'} \in (0,1)$ and $s'>\bar{k}$. For example, with $s'=6\bar{k}$, and the restricted isometry constant $\delta_{s'} \leq 1/3$, then the sparse eigenvalue condition above holds with $s=2\bar{k}$. For simplicity, in this paper we do not make distinctions between RIP and sparse eigenvalue condition. Note that in the traditional low-dimensional statistical analysis, one assumes that $\rho_+(s)/\rho_-(2\bar{k}+2s)< \infty$ as $s \to \infty$, which is significantly stronger than the condition we use here. Although in practice it is often difficult to verify the sparse eigenvalue condition for real problems, the parameter estimation result in (\[eq:multi-stage-param-est-bound\]) nevertheless provides important theoretical insights for multi-stage convex relaxation. For standard Lasso, we have the following bound $$\|\hat{{{\mathbf w}}}_{L_1}- \bar{{{\mathbf w}}}\|_2 =O(\sqrt{k} \lambda) ,$$ where $\hat{{{\mathbf w}}}_{L_1}$ is the solution of the standard $L_1$ regularization. This bound is tight for Lasso, in the sense that the right hand side cannot be improved except for the constant—this can be easily verified with an orthogonal design matrix. It is known that in order for Lasso to be effective, one has to pick $\lambda$ no smaller than the order $\sigma \sqrt{\ln p/n}$. Therefore, the parameter estimation error of the standard Lasso is of the order $\sigma \sqrt{\bar{k} \ln p/n}$, which cannot be improved. In comparison, if we consider the capped-$L_1$ regularization with $g(|{{\mathbf w}}_j|)$ defined in (\[eq:capped-L1-reg\]), the bound in (\[eq:multi-stage-param-est-bound\]) can be significantly better when most non-zero coefficients of $\bar{{{\mathbf w}}}$ are relatively large in magnitude. In the extreme case where $k_\theta = \left|\{j: |\bar{{{\mathbf w}}}_j| \in (0,2\theta]\}\right|=0$, which can be achieved when all nonzero components of $\bar{{{\mathbf w}}}$ are larger than the order $\sigma \sqrt{\ln p/n}$, we obtain the following better bound $$\|\hat{{{\mathbf w}}}^{(\ell)}- \bar{{{\mathbf w}}}\|_2 = O(\sqrt{\bar{k}/n} + \sqrt{\ln (1/\eta)/n})$$ for the multi-stage procedure for a sufficiently large $\ell$ at the order of $\ln k + \ln \ln p$. This bound is superior to the standard one-stage $L_1$ regularization bound $\|\hat{{{\mathbf w}}}_{L_1}- \bar{{{\mathbf w}}}\|_2 = O(\sqrt{\bar{k} \ln (p/\eta)/n})$. In the literature, one is often interested in two types of results, one is parameter estimation bound as in (\[eq:multi-stage-param-est-bound\]), and the other is feature selection consistency: that is, to identify the set of nonzero coefficients of the truth. Although the parameter estimation bound in (\[eq:multi-stage-param-est-bound\]) is superior to Lasso, the result does not imply that one can correctly select all variables under this condition. Moreover, the specific proof presented in [@Zhang09-multistage] does not directly imply such a result. Therefore it is important to know whether the multi-stage convex relaxation can achieve unbiased feature selection as studied in [@Zhang10-mc+]. In the following, we present such a result which supplements the parameter estimation bound of (\[eq:multi-stage-param-est-bound\]). While the main high-level argument follows that of [@Zhang09-multistage], there are many differences in the details, and hence a full proof (which is included in Section \[apx:proof\]) is still needed. This theorem is the main result of the paper. It is worth mentioning that although we only consider the simple capped-$L_1$ regularizer, similar results can be obtained for other regularizers (with virtually the same proof) such that $g'(u) \in [0, \infty)$, $g'(u)>0$ when $u$ belongs to a neighbor of $0$, and $g'(u)=0$ when $u \geq \theta$, with a threshold $\theta>0$ appropriately chosen at the order of the noise level — the condition of $g'(u)=0$ when $u \geq \theta$ ensures the removal of feature selection “bias” of Lasso which we discussed above. As an example, very similar result can be obtained for the MC+ penalty of [@Zhang10-mc+] or SCAD penalty of [@FanLi01] using the multi-stage convex relaxation procedure here. In fact, in practice there may be additional advantages of using a smooth nonconvex penalty such as MC+ due to the extra smoothness, although such advantage is not revealed in our theoretical analysis. \[thm:multi-stage-featsel\] Let Assumption \[assump:fixed\] hold. Assume also that the target $\bar{{{\mathbf w}}}$ is sparse, with ${{\mathbf E}}y_i = \bar{{{\mathbf w}}}^\top {{\mathbf x}}_i$, and $\bar{k}=\|\bar{{{\mathbf w}}}\|_0$. Let $\bar{F}={{\mathrm{supp}}}(\bar{{{\mathbf w}}})$. Choose $\theta$ and $\lambda$ such that $$\lambda \geq 7 \sigma \sqrt{2\rho_+(1) \ln (2p/\eta)/n}$$ and $$\theta > 9 \lambda /\rho_-(1.5\bar{k}+s) .$$ Assume that $$\min_{j \in \bar{F}} |\bar{{{\mathbf w}}}_j| > 2 \theta$$ and $\rho_+(s)/\rho_-(1.5\bar{k}+2s) \leq 1+ 2s/(3\bar{k})$ for some $s \geq 1.5 \bar{k}$, then with probability larger than $1-\eta$: $${{\mathrm{supp}}}(\hat{{{\mathbf w}}}^{(\ell)})={{\mathrm{supp}}}(\bar{{{\mathbf w}}})$$ when $\ell>L$, where $\hat{{{\mathbf w}}}^{(\ell)}$ is the solution of (\[eq:convex-relax-L1\]) and $$L = \left\lfloor \frac{0.5 \ln \bar{k} }{\ln (\rho_-(1.5\bar{k}+s)\theta/(6\lambda))} \right\rfloor +1 .$$ Theorem \[thm:multi-stage-featsel\] is the main result of this paper. If $$\min_{{{\mathbf w}}_j \in \bar{F}} |{{\mathbf w}}_j| \geq c \sigma\sqrt{\ln p/n}$$ for a sufficiently large constant $c$ that is independent of $\bar{k}$ (but could depend on the RIP condition), then we can pick both parameters $\lambda=O(\sigma \sqrt{\ln p/n})$ and $\theta=O(\sigma \sqrt{\ln p/n})$ at the noise level, so that Theorem \[thm:multi-stage-featsel\] can be applied. In this case, Theorem \[thm:multi-stage-featsel\] implies that multi-stage capped-$L_1$ regularization achieves exact recovery of the support set ${{\mathrm{supp}}}(\bar{{{\mathbf w}}})$. In comparison, Lasso does not achieve exact sparse recovery under RIP conditions. While running Lasso followed by thresholding small coefficients to zero (or using adaptive Lasso of [@Zou06] or the two-stage procedure of [@ZouLi08]) may achieve exact recovery, such a procedure requires the condition that $$\min_{{{\mathbf w}}_j \in \bar{F}} |{{\mathbf w}}_j| \geq c' \sigma \sqrt{\bar{k}\ln p/n} \label{eq:min-bias}$$ for some constant $c'$ (also depends on the RIP condition). This extra $\sqrt{\bar{k}}$ factor is referred to as the bias of the Lasso procedure in [@Zhang10-mc+]. Moreover, it is known that for exact recovery to hold, the requirement of $\min_{{{\mathbf w}}_j \in \bar{F}} |{{\mathbf w}}_j| \geq c \sigma \sqrt{\ln p/n}$ (up to a constant) is necessary for all statistical procedures, in the sense that if $\min_{{{\mathbf w}}_j \in \bar{F}} |{{\mathbf w}}_j| \leq c' \sigma \sqrt{\ln p/n}$ for a sufficiently small constant $c'$ (under appropriate RIP conditions), then no statistical procedure can achieve exact recovery with large probability. Therefore statistical procedures that can achieve exact support recovery under (\[eq:min-bias\]) are referred to as (nearly) unbiased feature selection methods in [@Zhang10-mc+]. Theorem \[thm:multi-stage-featsel\] shows that multi-stage convex relaxation with capped-$L_1$ regularization achieves unbiased feature selection. Results most comparable to what we have obtained here are that of the FoBa procedure in [@Zhang08-foba] and that of the MC+ procedure in [@Zhang10-mc+]. Both can be regarded as (approximate) optimization methods for nonconvex formulations. The former is a forward backward greedy algorithm, which does not optimize (\[eq:sparse-reg\]), while the latter is a path-following algorithm for solving formulations similar to (\[eq:sparse-reg\]). Although results in [@Zhang10-mc+] are comparable to ours, we should note that unlike our procedure, which is efficient due to the finite number of convex optimization, there is no proof showing that the path-following strategy in [@Zhang10-mc+] is always efficient (in the sense that there may be exponentially many switching points). Simulation Study ================ Numerical examples can be found in [@Zhang09-multistage] that demonstrate the advantage of multi-stage convex relaxation over Lasso. Therefore we shall not repeat a comprehensive study. Nevertheless, this section presents a simple simulation study to illustrate the theoretical results. The $n \times p$ design matrix $X$ is generated with iid random Gaussian entries and each column is normalized with 2-norm $\sqrt{n}$. Here $n=100$ and $p=250$. We then generate a vector $\bar{{{\mathbf w}}}$ with $\bar{k}=30$ nonzero coefficients, and each nonzero coefficient is uniformly generated from the interval $(1,10)$. The observation is $y=X \bar{{{\mathbf w}}} + \epsilon$, where $\epsilon$ is zero-mean iid Gaussian noise with standard deviation $\sigma=1$. We study the feature selection performance of Multi-stage convex relaxation method in Figure \[fig:multi-stage-sparse\] using various configurations of $\lambda=\tau \sigma \sqrt{\ln(p)/n}$ (with $\tau=1,2,4,8,16,32$), and $\theta=\mu \lambda$ for various constants $\mu=0.5,1,2,4$. The experiments are repeated for 100 times, and Table \[tab:performance\] reports the probability (percentage in the 100 runs) of exact support recovery for each configuration at various stages $\ell$. Note that $\ell=1$ corresponds to Lasso and $\ell=2$ is an adaptive Lasso like two stage method [@Zou06; @ZouLi08]. The main purpose of this study is to illustrate that it is beneficial to use more than two stages, as predicted by our theory. However, since only $O(\ln(\bar{k}))$ is sufficient, optimal results can be achieved with relatively small number of stages. These conclusions can be clearly seen from Table \[tab:performance\]. Specifically the results for $\ell=2$ are better than those of $\ell=1$ (standard Lasso), while results of $\ell=4$ are better than those of $\ell=2$. Although the performance of $\ell=8$ is even better, the improve over $\ell=4$ is small at the optimal configuration of $\lambda$ and $\theta$. This is consistent with our theory, which implies that a relatively small number of stages is needed to achieve good performance. [|c|c|c|c|c|c|c|c|]{}\ $\lambda$ & $ 0.23$ & $ 0.47$ & $ 0.94$ & $ 1.9$ & $ 3.8$ & $ 7.5$\ $\ell=1$ & $ 0$ & $ 0$ & $ 0$ & $ 0$ & $ 0$ & $ 0$\ $\ell=2$ & $ 0$ & $ 0$ & $ 0.02$ & $ 0$ & $ 0$ & $ 0$\ $\ell=4$ & $ 0$ & $ 0.05$ & $ 0.63$ & $ 0.18$ & $ 0$ & $ 0$\ $\ell=8$ & $ 0$ & $ 0.12$ & $ 0.83$ & $ 0.25$ & $ 0$ & $ 0$\ \ $\lambda$ & $ 0.23$ & $ 0.47$ & $ 0.94$ & $ 1.9$ & $ 3.8$ & $ 7.5$\ $\ell=1$ & $ 0$ & $ 0$ & $ 0$ & $ 0$ & $ 0$ & $ 0$\ $\ell=2$ & $ 0$ & $ 0.04$ & $ 0.15$ & $ 0.06$ & $ 0$ & $ 0$\ $\ell=4$ & $ 0$ & $ 0.33$ & $ 0.86$ & $ 0.13$ & $ 0$ & $ 0$\ $\ell=8$ & $ 0$ & $ 0.38$ & $ 0.93$ & $ 0.16$ & $ 0$ & $ 0$\ \ $\lambda$ & $ 0.23$ & $ 0.47$ & $ 0.94$ & $ 1.9$ & $ 3.8$ & $ 7.5$\ $\ell=1$ & $ 0$ & $ 0$ & $ 0$ & $ 0$ & $ 0$ & $ 0$\ $\ell=2$ & $ 0$ & $ 0.14$ & $ 0.22$ & $ 0$ & $ 0$ & $ 0$\ $\ell=4$ & $ 0$ & $ 0.29$ & $ 0.6$ & $ 0.02$ & $ 0$ & $ 0$\ $\ell=8$ & $ 0$ & $ 0.3$ & $ 0.62$ & $ 0.02$ & $ 0$ & $ 0$\ \ $\lambda$ & $ 0.23$ & $ 0.47$ & $ 0.94$ & $ 1.9$ & $ 3.8$ & $ 7.5$\ $\ell=1$ & $ 0$ & $ 0$ & $ 0$ & $ 0$ & $ 0$ & $ 0$\ $\ell=2$ & $ 0$ & $ 0.01$ & $ 0.01$ & $ 0$ & $ 0$ & $ 0$\ $\ell=4$ & $ 0$ & $ 0.06$ & $ 0.06$ & $ 0$ & $ 0$ & $ 0$\ $\ell=8$ & $ 0$ & $ 0.06$ & $ 0.06$ & $ 0$ & $ 0$ & $ 0$\ Proof of Theorem \[thm:multi-stage-featsel\] {#apx:proof} ============================================ The analysis is an adaptation of [@Zhang09-multistage]. While the main proof structure is similar, there are nevertheless subtle and important differences in the details, and hence a complete proof is still necessary. The main technical differences are as follows. The proof of [@Zhang09-multistage] tracks the progress from one stage $\ell-1$ to the next stage $\ell$ using a bound on 2-norm parameter estimate, while in the current proof we track the progress using the set of variables that differ significantly from the true variables. Moreover, in [@Zhang09-multistage], we compare the current estimated parameter to the true parameter $\bar{{{\mathbf w}}}$, which is sufficient for parameter estimation. However, in order to establish feature selection result of this paper, it is necessary to compare the current estimated parameter to the least squares solution $\tilde{{{\mathbf w}}}$ within the true feature set $\bar{F}$ as defined below in (\[eq:ls\]). These subtle technical differences mean that many details in the proofs presented below differ from that of [@Zhang09-multistage]. Auxiliary lemmas ---------------- We first introduce some definitions. Consider the positive semi-definite matrix $A= n^{-1} X^\top X \in {{\mathbb R}}^{d \times d}$. Given $s,k \geq 1$ such that $s + k \leq d$. Let $I,J$ be disjoint subsets of $\{1,\ldots,d\}$ with $k$ and $s$ elements respectively. Let $A_{I,I} \in R^{k \times k}$ be the restriction of $A$ to indices $I$, $A_{I,J} \in R^{k \times s}$ be the restriction of $A$ to indices $I$ on the left and $J$ on the right. Similarly we define restriction ${{\mathbf w}}_I$ of a vector ${{\mathbf w}}\in R^p$ on $I$; and for convenience, we allow either ${{\mathbf w}}_I \in R^k$ or ${{\mathbf w}}_I \in R^p$ (where components not in $I$ are zeros) depending on the context. We also need the following quantity in our analysis: $$\pi(k,s)=\sup_{{{\mathbf v}}\in R^{k}, {{\mathbf u}}\in R^s ,I,J} \frac{{{\mathbf v}}^\top A_{I,J} {{\mathbf u}}\|{{\mathbf v}}\|_2}{{{\mathbf v}}^\top A_{I,I} {{\mathbf v}}\|{{\mathbf u}}\|_\infty } .$$ The following two lemmas are taken from [@Zhang07-l1]. We skip the proof. \[lem:gamma-bound\] The following inequality holds: $$\pi(k,s) \leq \frac{s^{1/2}}{2} \sqrt{\rho_+(s)/\rho_-(k+s)-1} ,$$ \[lem:inner-prod\] Consider $k,s > 0$ and $G \subset \{1,\ldots,d\}$ such that $|G^c|=k$. Given any ${{\mathbf w}}\in R^p$. Let $J$ be the indices of the $s$ largest components of ${{\mathbf w}}_G$ (in absolute values), and $I= G^c \cup J$. Then $$\max(0,{{\mathbf w}}_I^\top A {{\mathbf w}}) \geq \rho_-(k+s) (\|{{\mathbf w}}_I\|_2 - \pi(k+s,s)\|{{\mathbf w}}_G\|_1/s) \|{{\mathbf w}}_I\|_2 .$$ Our analysis requires us to keep track of progress with respect to the least squares solution $\tilde{{{\mathbf w}}}$ with the true feature set $\bar{F}$, which we define below: $$\tilde{{{\mathbf w}}} = \arg\min_{{{\mathbf w}}\in R^p} \|X {{\mathbf w}}- {{\mathbf y}}\|_2^2 \qquad \text{subject to } {{\mathrm{supp}}}({{\mathbf w}}) \subset \bar{F} , \label{eq:ls}$$ where $\bar{F}={{\mathrm{supp}}}(\bar{{{\mathbf w}}})$. The following lemmas require varying degrees of modifications from similar lemmas in [@Zhang09-multistage], and thus the proofs are included for completeness. \[lem:sub-Gaussian-infnorm\] Define $\hat{\epsilon}= \frac{1}{n} X^\top (X \tilde{{{\mathbf w}}}-{{\mathbf y}})$. Under the conditions of Assumption \[assump:fixed\], with probability larger than $1-\eta$: $$\forall j \in \bar{F}: \quad |\hat{\epsilon}_j|=0 , \quad |\tilde{{{\mathbf w}}}_j- \bar{{{\mathbf w}}}_j| \leq \sigma \sqrt{2 \rho_-(\bar{k})^{-1} \ln (2p/\eta)/ n} ,$$ and $$\forall j \notin \bar{F}: \quad |\hat{\epsilon}_j|\leq \sigma \sqrt{2 \rho_+(1) \ln (2p/\eta)/ n} .$$ Let $\tilde{P}$ be the projection matrix to the subspace spanned by columns of $X$ in $\bar{F}$, then we know that $$X \tilde{{{\mathbf w}}} = \tilde{P} {{\mathbf y}}$$ and $$(I-\tilde{P}) {{\mathbf E}}{{\mathbf y}}= {{\mathbf E}}{{\mathbf y}}- X \bar{{{\mathbf w}}} = 0 .$$ Therefore for each $j$ $$n |\hat{\epsilon}_j| = |X_j^\top (X \tilde{{{\mathbf w}}}-{{\mathbf y}}) | = |X_j^\top (I-\tilde{P}) ({{\mathbf y}}- {{\mathbf E}}{{\mathbf y}})) | .$$ It implies that $\hat{\epsilon}_j=0$ if $j \in \bar{F}$. Since for each $j$: the column $X_j$ satisfies $\|X_j^\top (I-\tilde{P})\|_2^2 \leq n \rho_+(1)$, we have from sub-Gaussian tail bound that for all $j \notin \bar{F}$ and $\epsilon>0$: $$P \left[ |\hat{\epsilon}_j| \geq \epsilon \right] \leq 2 \exp [ - n\epsilon^2/(2 \sigma^2 \rho_+(1)) ] .$$ Moreover, for each $j \in \bar{F}$, we have $$|\tilde{{{\mathbf w}}}_j-\bar{{{\mathbf w}}}_j|= e_j^\top (X_{\bar{F}}^\top X_{\bar{F}})^{-1} X_{\bar{F}}^\top ({{\mathbf y}}- {{\mathbf E}}{{\mathbf y}}) .$$ Since $\|e_j^\top (X_{\bar{F}}^\top X_{\bar{F}})^{-1} X_{\bar{F}}^\top\|_2^2= e_j^\top (X_{\bar{F}}^\top X_{\bar{F}})^{-1} e_j \leq n^{-1} \rho_-(\bar{k})^{-1}$, we have for all $\epsilon>0$: $$P \left[ |\tilde{{{\mathbf w}}}_j-\bar{{{\mathbf w}}}_j| \geq \epsilon \right] \leq 2 \exp [ - n \rho_-(\bar{k})\epsilon^2/(2 n\sigma^2) ] .$$ Taking union bound for $j=1,\ldots,d$ (each with probability $\eta/d$) we obtain the desired inequality. \[lem:L1-nonsparse-dr\] Consider $G\subset \{1,\ldots,d\}$ such that $\bar{F}\cap G = \emptyset$. Let $\hat{{{\mathbf w}}}=\hat{{{\mathbf w}}}^{(\ell)}$ be the solution of (\[eq:convex-relax-L1\]), and let $\Delta \hat{{{\mathbf w}}}= \hat{{{\mathbf w}}}-\tilde{{{\mathbf w}}}$. Let $\lambda_G = \min_{j \in G} \lambda_j^{(\ell-1)}$ and $\lambda_0 = \max_{j} \lambda_j^{(\ell-1)}$. If $2\|\hat{\epsilon}\|_\infty\| < \lambda_G$, then $$\sum_{j \in G} |\hat{{{\mathbf w}}}_j | \leq \frac{2 \|\hat{\epsilon}\|_\infty}{\lambda_G - 2\|\hat{\epsilon}\|_\infty} \sum_{j \notin \bar{F}\cup G} |\hat{{{\mathbf w}}}_j| + \frac{\lambda_0}{\lambda_G - 2\|\hat{\epsilon}\|_\infty} \sum_{j \in \bar{F}} |\Delta \hat{{{\mathbf w}}}_j| \leq \frac{\lambda_0}{\lambda_G - 2\|\hat{\epsilon}\|_\infty}\|\Delta \hat{{{\mathbf w}}}_{G^c}\|_1 .$$ For simplicity, let $\lambda_j=\lambda_j^{(\ell-1)}$. The first order equation implies that $$\frac{1}{n} \sum_{i=1}^n 2 ({{\mathbf x}}_i^\top \hat{{{\mathbf w}}} - y_i) {{\mathbf x}}_{i,j} + \lambda_j {{\mathrm{sgn}}}(\hat{{{\mathbf w}}}_j) = 0 ,$$ where ${{\mathrm{sgn}}}({{\mathbf w}}_j)=1$ when ${{\mathbf w}}_j>0$, ${{\mathrm{sgn}}}({{\mathbf w}}_j)=-1$ when ${{\mathbf w}}_j<0$, and ${{\mathrm{sgn}}}({{\mathbf w}}_j) \in [-1,1]$ when ${{\mathbf w}}_j=0$. This implies that for all ${{\mathbf v}}\in {{\mathbb R}}^p$, we have $$2 {{\mathbf v}}^\top A \Delta \hat{{{\mathbf w}}} \leq - 2 {{\mathbf v}}^\top \hat{\epsilon} - \sum_{j=1}^p \lambda_j {{\mathbf v}}_j {{\mathrm{sgn}}}(\hat{{{\mathbf w}}}_j) . \label{eq:dr}$$ Now, let ${{\mathbf v}}=\Delta \hat{{{\mathbf w}}}$ in (\[eq:dr\]), and notice that $\hat{\epsilon}_{\bar{F}}=0$, we obtain $$\begin{aligned} 0 \leq& 2 \Delta \hat{{{\mathbf w}}}^\top A \Delta \hat{{{\mathbf w}}} \leq 2 |\Delta \hat{{{\mathbf w}}}^\top \hat{\epsilon} | - \sum_{j=1}^p \lambda_j \Delta \hat{{{\mathbf w}}}_j {{\mathrm{sgn}}}(\hat{{{\mathbf w}}}_j) \\ \leq& 2 \|\Delta \hat{{{\mathbf w}}}_{\bar{F}^c}\|_1 \|\hat{\epsilon} \|_\infty - \sum_{j \in \bar{F}} \lambda_j \Delta \hat{{{\mathbf w}}}_j {{\mathrm{sgn}}}(\hat{{{\mathbf w}}}_j) - \sum_{j \notin \bar{F}} \lambda_j \Delta \hat{{{\mathbf w}}}_j {{\mathrm{sgn}}}(\hat{{{\mathbf w}}}_j) \\ \leq& 2 \|\Delta \hat{{{\mathbf w}}}_{\bar{F}^c}\|_1 \|\hat{\epsilon} \|_\infty + \sum_{j \in \bar{F}} \lambda_j |\Delta \hat{{{\mathbf w}}}_j| - \sum_{j \notin \bar{F}} \lambda_j |\hat{{{\mathbf w}}}_j| \\ \leq& \sum_{j \in G} (2\|\hat{\epsilon}\|_\infty -\lambda_G) |\hat{{{\mathbf w}}}_j | + \sum_{j \notin G\cup \bar{F}} 2 \|\hat{\epsilon}\|_\infty |\hat{{{\mathbf w}}}_j| + \sum_{j \in \bar{F}} \lambda_0 |\Delta \hat{{{\mathbf w}}}_j| .\end{aligned}$$ By rearranging the above inequality, we obtain the first desired bound. The second inequality uses $2\|\hat{\epsilon}\|_\infty \leq \lambda_0$. \[lem:L1-nonsparse3\] Using the notations of Lemma \[lem:L1-nonsparse-dr\], and let $J$ be the indices of the largest $s$ coefficients (in absolute value) of $\hat{{{\mathbf w}}}_G$. Let $I= G^c \cup J$ and $k=|G^c|$. If $0 \leq \lambda_0 /(\lambda_G -2\|\hat{\epsilon}\|_\infty) \leq 3$, then $$\|\Delta \hat{{{\mathbf w}}}\|_2 \leq (1+ (3 k/s)^{0.5}) \|\Delta \hat{{{\mathbf w}}}_I\|_2 .$$ Using $\lambda_0/(\lambda_G-2\|\hat{\epsilon}\|_\infty) \leq 3$, we obtain from Lemma \[lem:L1-nonsparse-dr\] $$\|\hat{{{\mathbf w}}}_G\|_1 \leq 3 \|\Delta \hat{{{\mathbf w}}} - \hat{{{\mathbf w}}}_G \|_1 .$$ Therefore $$\begin{aligned} \|\Delta \hat{{{\mathbf w}}}-\Delta \hat{{{\mathbf w}}}_I\|_\infty \leq& \|\Delta \hat{{{\mathbf w}}}_J\|_1/s \\ =& s^{-1} [\|\Delta \hat{{{\mathbf w}}}_G\|_1 - \|\Delta \hat{{{\mathbf w}}}-\Delta \hat{{{\mathbf w}}}_I\|_1] \\ \leq& s^{-1} [3\|\Delta \hat{{{\mathbf w}}} - \hat{{{\mathbf w}}}_G\|_1 - \|\Delta \hat{{{\mathbf w}}}-\Delta \hat{{{\mathbf w}}}_I\|_1],\end{aligned}$$ which implies that $$\begin{aligned} \|\Delta \hat{{{\mathbf w}}}-\Delta \hat{{{\mathbf w}}}_I\|_2 \leq & (\|\Delta \hat{{{\mathbf w}}}-\Delta \hat{{{\mathbf w}}}_I\|_1 \|\Delta \hat{{{\mathbf w}}}-\Delta \hat{{{\mathbf w}}}_I\|_\infty)^{1/2} \\ \leq& \left[\|\Delta \hat{{{\mathbf w}}}-\Delta \hat{{{\mathbf w}}}_I\|_1 (3\|\Delta \hat{{{\mathbf w}}} - \hat{{{\mathbf w}}}_G\|_1 - \|\Delta \hat{{{\mathbf w}}}-\Delta \hat{{{\mathbf w}}}_I\|_1)\right]^{1/2} s^{-1/2} \\ \leq& \left[ (3\|\Delta \hat{{{\mathbf w}}} - \hat{{{\mathbf w}}}_G\|_1/2)^2\right]^{1/2} s^{-1/2} \\ \leq& (3/2) s^{-1/2} \|\Delta \hat{{{\mathbf w}}} - \hat{{{\mathbf w}}}_G\|_1 \\ \leq& (3/2) s^{-1/2} \bar{k}^{1/2} \|\Delta \hat{{{\mathbf w}}} - \hat{{{\mathbf w}}}_G\|_2 \leq (3 k/s)^{1/2} \|\Delta \hat{{{\mathbf w}}}_I\|_2 .\end{aligned}$$ The third inequality uses the simple algebraic inequality $a(3b-a) \leq (3b/2)^2$. By rearranging this inequality, we obtain the desired bound. Note that in the above derivation, we have used the fact that $\bar{F} \cap G=\emptyset$, which implies that $\Delta \hat{{{\mathbf w}}}_G=\hat{{{\mathbf w}}}_G$, and thus $\Delta \hat{{{\mathbf w}}} - \hat{{{\mathbf w}}}_G=\Delta\hat{{{\mathbf w}}}_{G^c}$. \[lem:L1-nonsparse1\] Let the conditions of Lemma \[lem:L1-nonsparse-dr\] and Lemma \[lem:L1-nonsparse3\] hold, and let $k=|G^c|$. If $t=1-\pi(k+s,s) k^{1/2} s^{-1} \in (0,4/3)$, and $0 \leq \lambda_0/(\lambda_G -2\|\hat{\epsilon}\|_\infty) \leq (4-t)/(4-3t)$, then $$\|\Delta \hat{{{\mathbf w}}}\|_2 \leq (1+ (3k/s)^{0.5}) \|\Delta \hat{{{\mathbf w}}}_I\|_2 \leq \frac{1+ (3k/s)^{0.5}}{t \rho_-(k+s)} \left[ 2 \|\hat{\epsilon}_{G^c}\|_2 + \left(\sum_{j \in \bar{F}} (\lambda_j^{(\ell-1)})^2\right)^{1/2}\right] .$$ Let $J$ be the indices of the largest $s$ coefficients (in absolute value) of $\hat{{{\mathbf w}}}_G$, and $I= G^c \cup J$. The conditions of the lemma imply that $$\begin{aligned} \max(0,\Delta \hat{{{\mathbf w}}}_I^\top A \Delta \hat{{{\mathbf w}}}) \geq& \rho_-(k+s) [\|\Delta \hat{{{\mathbf w}}}_I\|_2 - \pi(k+s,s) \|\hat{{{\mathbf w}}}_G\|_1/s] \|\Delta \hat{{{\mathbf w}}}_I\|_2 \\ \geq& \rho_-(k+s) [1 - (1-t)(4-t)(4-3t)^{-1} ] \|\Delta \hat{{{\mathbf w}}}_I\|_2^2 \\ \geq& 0.5 t \rho_-(k+s) \|\Delta \hat{{{\mathbf w}}}_I\|_2^2 .\end{aligned}$$ In the above derivation, the first inequality is due to Lemma \[lem:inner-prod\]; the second inequality is due to the conditions of this lemma plus Lemma \[lem:L1-nonsparse-dr\], which implies that $$\|\hat{{{\mathbf w}}}_G\|_1 \leq \frac{\lambda_0}{\lambda_G -2\|\hat{\epsilon}\|_\infty} \|\hat{{{\mathbf w}}}_{G^c}\|_1\leq \frac{\lambda_0}{\lambda_G -2\|\hat{\epsilon}\|_\infty} \sqrt{k}\|\hat{{{\mathbf w}}}_{I}\|_2 ;$$ and the last inequality follows from $1-(1-t)(4-t)(4-3t)^{-1} \geq 0.5 t$, which holds for $t \in (0,4/3)$. If $\Delta \hat{{{\mathbf w}}}_I^\top A \Delta \hat{{{\mathbf w}}} \leq 0$, then the above inequality, together with Lemma \[lem:L1-nonsparse3\], imply the lemma. Therefore in the following, we can assume that $$\Delta \hat{{{\mathbf w}}}_I^\top A \Delta \hat{{{\mathbf w}}} \geq 0.5 t \rho_-(k+s) \|\Delta \hat{{{\mathbf w}}}_I\|_2^2 .$$ Moreover, let $\lambda_j=\lambda_j^{(\ell-1)}$. We obtain from (\[eq:dr\]) with ${{\mathbf v}}=\Delta \hat{{{\mathbf w}}}_I$ the following: $$\begin{aligned} & 2 \Delta \hat{{{\mathbf w}}}_I^\top A \Delta \hat{{{\mathbf w}}} \leq - 2 \Delta \hat{{{\mathbf w}}}_I^\top \hat{\epsilon} - \sum_{j \in I} \lambda_j \Delta \hat{{{\mathbf w}}}_j {{\mathrm{sgn}}}(\hat{{{\mathbf w}}}_j) \\ = & - 2 \Delta \hat{{{\mathbf w}}}_I^\top \hat{\epsilon}_{G^c} - 2 \Delta \hat{{{\mathbf w}}}_I^\top \hat{\epsilon}_{G} - \sum_{j \in \bar{F}} \lambda_j \Delta \hat{{{\mathbf w}}}_j {{\mathrm{sgn}}}(\hat{{{\mathbf w}}}_j) -\sum_{j \in G} \lambda_j |\Delta \hat{{{\mathbf w}}}_j| -\sum_{j \in \bar{F}^c \cap G^c} \lambda_j |\Delta \hat{{{\mathbf w}}}_j| \\ \leq & 2 \|\Delta \hat{{{\mathbf w}}}_I\|_2 \|\hat{\epsilon}_{G^c}\|_2 +2 \|\hat{\epsilon}_{G}\|_\infty \sum_{j \in G} |\Delta \hat{{{\mathbf w}}}_j| + \sum_{j \in \bar{F}} \lambda_j |\Delta \hat{{{\mathbf w}}}_j| -\sum_{j \in G} \lambda_j |\Delta \hat{{{\mathbf w}}}_j| \\ \leq & 2 \|\Delta \hat{{{\mathbf w}}}_I\|_2 \|\hat{\epsilon}_{G^c}\|_2 + (\sum_{j \in \bar{F}} \lambda_j^2)^{1/2} \|\Delta \hat{{{\mathbf w}}}_I\|_2 .\end{aligned}$$ Note that the equality uses the fact that $G \subset \bar{F}^c$, and $\Delta \hat{{{\mathbf w}}}_j {{\mathrm{sgn}}}(\hat{{{\mathbf w}}}_j) =|\hat{{{\mathbf w}}}_j|$ for $j \in \bar{F}^c$. The last inequality uses the fact that $\forall j \in G$: $\lambda_j \geq \lambda_G \geq 2 \|\hat{\epsilon}_G\|_\infty$. Now by combining the above two estimates, we obtain $$\|\Delta \hat{{{\mathbf w}}}_I\|_2 \leq \frac{1}{t \rho_-(k+s)} \left[ 2 \|\hat{\epsilon}_{G^c}\|_2 + (\sum_{j \in \bar{F}} \lambda_j^2)^{1/2} \right] .$$ The desired bound follows from Lemma \[lem:L1-nonsparse3\]. \[lem:lambda-bound\] Let $\lambda_j= \lambda I(|{{{\mathbf w}}}_j|\leq \theta)$ for some ${{{\mathbf w}}} \in R^p$, then $$\left(\sum_{j \in \bar{F}} \lambda_j^2\right)^{1/2} \leq \lambda \sqrt{\sum_{j \in \bar{F}} I(|\bar{{{\mathbf w}}}_j| \leq 2\theta)} + \lambda \left|\{j \in \bar{F}: |\bar{{{\mathbf w}}}_j-{{{\mathbf w}}}_j|\geq \theta \}\right|^{1/2} .$$ By assumption, if $|\bar{{{\mathbf w}}}_j-{{{\mathbf w}}}_j| \geq \theta$, then $$I(|{{{\mathbf w}}}_j| \leq \theta) \leq 1 \leq I(|\bar{{{\mathbf w}}}_j-{{{\mathbf w}}}_j|\geq \theta) ;$$ otherwise, $I(|{{{\mathbf w}}}_j| \leq \theta) \leq I(|\bar{{{\mathbf w}}}_j| \leq 2\theta)$. It follows that the following inequality always holds: $$I(|{{{\mathbf w}}}_j|\leq \theta) \leq I(|\bar{{{\mathbf w}}}_j| \leq 2\theta) + I(|\bar{{{\mathbf w}}}_j-{{{\mathbf w}}}_j|\geq \theta) .$$ The desired bound is a direct consequence of the above result and the 2-norm triangle inequality $$(\sum_j (x_j + \Delta x_j)^2)^{1/2} \leq (\sum_j x_j^2)^{1/2} + (\sum_j \Delta x_j^2)^{1/2} .$$ \[lem:bound-ell=1\] Define $F^{(\ell)} = \{j: |\hat{{{\mathbf w}}}_j^{(\ell)} -\bar{{{\mathbf w}}}_j| \geq \theta\}$. Under the conditions of Theorem \[thm:multi-stage-featsel\], we have for all $s \geq 2 \bar{k}$: $$\|\hat{{{\mathbf w}}}^{(\ell)}- \tilde{{{\mathbf w}}}\|_2 \leq \frac{5.7 \lambda}{\rho_-(1.5\bar{k}+s)} \sqrt{|F^{(\ell-1)}|} ,$$ and $$\sqrt{|F^{(\ell)}|}\leq \frac{6 \lambda \theta^{-1}}{\rho_-(1.5\bar{k}+s)} \sqrt{|F^{(\ell-1)}|} .$$ For all $t \in [0.5,4/3)$, by using Lemma \[lem:sub-Gaussian-infnorm\], we know that the condition of the theorem implies that $$\frac{\lambda} {\lambda - 2 \|\hat{\epsilon}\|_\infty} \leq 7/5 \leq \frac{4-t}{4-3t} .$$ Moreover, Lemma \[lem:gamma-bound\] implies that the condition $$0.5 \leq t=1 - \pi(1.5\bar{k}+s,s) (1.5\bar{k})^{0.5} /s$$ is also satisfied. This means that the conditions of Lemma \[lem:L1-nonsparse1\] (with $\lambda_0=\lambda_G=\lambda$) are satisfied. Now, we assume that at some $\ell \geq 1$, $$|G_\ell^c| \leq 1.5 \bar{k}, \quad \text{where } G_\ell=\{j \notin \bar{F}: \lambda_j^{(\ell-1)} =\lambda \} , \label{eq:G}$$ then it is easy to verify that $G_\ell^c \setminus \bar{F} \subset F^{(\ell-1)}$. Moreover, with the definition of $G=G_\ell$ in Lemma \[lem:L1-nonsparse1\] and Lemma \[lem:lambda-bound\], we can set $\lambda_0=\lambda_G=\lambda$ and obtain (note also that $\hat{\epsilon}_{\bar{F}}=0$) $$\begin{aligned} \|\hat{{{\mathbf w}}}^{(\ell)}-\tilde{{{\mathbf w}}}\|_2 \leq& \frac{1+\sqrt{3}}{t \rho_-(1.5\bar{k}+s)} \left[ 2\|\hat{\epsilon}_{G_\ell^c \setminus\bar{F}}\|_2 + \left(\sum_{j \in \bar{F}} (\lambda_j^{(\ell-1)})^2\right)^{1/2} \right]\\ \leq& \frac{1+\sqrt{3}}{t \rho_-(1.5\bar{k}+s)} \left[ 2 \sqrt{| F^{(\ell-1)}\setminus \bar{F}|} \|\hat{\epsilon}\|_\infty + \sqrt{|F^{(\ell-1)}\cap \bar{F}|} \lambda \right]\\ \leq& \frac{1+\sqrt{3}}{t \rho_-(1.5\bar{k}+s)} \left[ (2/7) \sqrt{| F^{(\ell-1)}\setminus \bar{F}|} + \sqrt{|F^{(\ell-1)}\cap \bar{F}|} \right] \lambda \\ \leq& \frac{1+\sqrt{3}}{0.5 \rho_-(1.5\bar{k}+s)} \left[ \sqrt{1.082 |F^{(\ell-1)}|} \right]\lambda \\ \leq& \frac{5.7 \lambda}{\rho_-(1.5\bar{k}+s)} \sqrt{|F^{(\ell-1)}|} ,\end{aligned}$$ where the first inequality is due to Lemma \[lem:L1-nonsparse1\]. The second inequality uses the facts that $G_\ell^c \setminus \bar{F} \subset F^{(\ell-1)} \setminus \bar{F}$, and Lemma \[lem:lambda-bound\] with $I(|\bar{{{\mathbf w}}}_j| \leq 2\theta)=0$ (for all $j \in \bar{F}$). The third inequality uses $2\|\hat{\epsilon}\|_\infty \leq (2/7)\lambda$, and the fourth inequality uses $(2/7)a + b \leq \sqrt{1.082(a^2+b^2)}$. Since Lemma \[lem:sub-Gaussian-infnorm\] implies that $$\|\tilde{{{\mathbf w}}}-\bar{{{\mathbf w}}}\|_\infty \leq (1/7) \lambda /\sqrt{\rho_+(1) \rho_-(\bar{k})} ,$$ we know that $j \in F^{(\ell)}$ implies that $$|\tilde{{{\mathbf w}}}_j-\hat{{{\mathbf w}}}^{(\ell)}_j|\geq \theta - (1/7) \lambda /\sqrt{\rho_+(1) \rho_-(\bar{k})} \geq (41/42) \theta .$$ Therefore $$\begin{aligned} \sqrt{|F^{(\ell)}|} \leq& (41\theta/42)^{-1} \|\tilde{{{\mathbf w}}}-\hat{{{\mathbf w}}}^{(\ell)}\|_2 \\ \leq& \frac{5.7 \lambda (41\theta/42)^{-1}}{\rho_-(1.5\bar{k}+s)} \sqrt{|F^{(\ell-1)}|} \\ \leq& \frac{6 \lambda \theta^{-1}}{\rho_-(1.5\bar{k}+s)} \sqrt{|F^{(\ell-1)}|} .\end{aligned}$$ That is, under the assumption of (\[eq:G\]), the lemma holds at $\ell$. Therefore next we only need to prove by induction on $\ell$ that (\[eq:G\]) holds for all $\ell=1, 2,\ldots$. When $\ell=1$, we have $G_1^c=\bar{F}$, which implies that (\[eq:G\]) holds. Now assume that (\[eq:G\]) holds at $\ell$ for some $\ell \geq 1$. Then by the induction hypothesis we know that the lemma holds at $\ell$. This means that $$\begin{aligned} \sqrt{|G_{\ell+1}^c \setminus \bar{F}|} \leq & \sqrt{|F^{(\ell)}|} \\ \leq& \frac{6 \lambda \theta^{-1}}{\rho_-(1.5\bar{k}+s)} \sqrt{|F^{(\ell-1)}|} \\ \leq& \sqrt{0.5 |F^{(\ell-1)}|} \\ \leq& \cdots \leq 0.5^{\ell/2} |F^{(0)}| .\end{aligned}$$ The first inequality is due to the fact $G_{\ell+1}^c \setminus \bar{F} \subset F^{(\ell)}$. The second inequality uses the assumption of $\theta$ in the theorem. The last inequality uses induction. Now note that $F^{(0)}=\bar{F}$, we thus have $|G_{\ell+1}^c \setminus \bar{F}| \leq 0.5\bar{k}$. This completes the induction step. Proof of Theorem \[thm:multi-stage-featsel\] {#proof-of-theoremthmmulti-stage-featsel} -------------------------------------------- Define $$\beta=\frac{6 \lambda \theta^{-1}}{\rho_-(1.5\bar{k}+s)} ,$$ We have $\beta<1$ by the assumption of the theorem. Using induction, we have from Lemma \[lem:bound-ell=1\] that $$\begin{aligned} \sqrt{|F^{(L)}|}\leq& \beta \sqrt{|F^{(L-1)}|} \\ \leq& \cdots \\ \leq& \beta^{L} \sqrt{|F^{(0)}|} \\ \leq& \beta^{L} \sqrt{\bar{k}} < 1 .\end{aligned}$$ This means that when $\ell > L$, $|F^{(\ell-1)}|=0$. Therefore by applying Lemma \[lem:bound-ell=1\] again we obtain $$\|\hat{{{\mathbf w}}}^{(\ell)}- \tilde{{{\mathbf w}}}\|_2 =0 .$$ Since Lemma \[lem:sub-Gaussian-infnorm\] implies that $$\|\tilde{{{\mathbf w}}}-\bar{{{\mathbf w}}}\|_\infty \leq (1/7) \lambda /\sqrt{\rho_+(1) \rho_-(\bar{k})} < \theta ,$$ we have $${{\mathrm{supp}}}(\tilde{{{\mathbf w}}})={{\mathrm{supp}}}(\bar{{{\mathbf w}}}) .$$ This implies that ${{\mathrm{supp}}}(\hat{{{\mathbf w}}}^{(\ell)})={{\mathrm{supp}}}(\bar{{{\mathbf w}}})$. Discussion ========== This paper investigated the performance of multi-stage convex relaxation for feature selection, where it is shown that under RIP, the procedure can achieve unbiased feature selection. This result complements that of [@Zhang09-multistage] which studies the parameter estimation performance of multi-stage convex relaxation. It also complements similar results obtained in [@Zhang08-foba] and [@Zhang10-mc+] for different computational procedures. One advantage of our result over that in [@Zhang10-mc+] is that the multi-stage convex relaxation method is provably efficient because the correct feature set can be obtained after no more than $O(\log \bar{k})$ number of iterations. In comparison, a computational efficiency statement for the path-following method of [@Zhang10-mc+] remains open.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We report x-ray diffraction, resistivity, thermopower, and magnetization of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$, in which A-site ordered tetragonal phase appears above $x=1$, and reveal that the system exhibits typical properties seen in the antiferromagnetic insulator with Mn$^{3+}$. We succeed in preparing both A-site ordered and disordered phases for $x=1$ in different preparation conditions, and observe a significant decrease of the resistivity in the disordered phase. We discuss possible origins of the decrease focusing on the dimensionality and the disordered effect.' address: - 'Waseda Institute for Advanced Study, Waseda University, Tokyo 169-8050, Japan' - 'Department of Physics, Waseda University, Tokyo 169-8555, Japan' author: - 'W. Kobayashi$^{\dagger}$' - 'T. Ishibashi' title: 'Structure-related transport properties of A-site ordered perovskite Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ' --- Introduction ============ Perovskite oxide is denoted as $AB$O$_3$, where $A$ and $B$ represent lanthanide (and/or alkaline earth elements) and transition-metal elements, respectively. $B$ ion is surrounded by six oxygen ions, and $B$O$_6$ octahedron is formed. The octahedron mainly contributes electrical and magnetic proerties of the perovskite oxide. According to the ionic radius of $A$ ion, the bond angle $\angle$ $B$-O-$B$ deviates from 180$^{\circ}$, which causes a change of the bandwidth. In a substituted system of $R^{3+}_{1-x}$$A^{2+}_{x}$$B$O$_3$, the carrier concentration is also controlled as well as the bandwidth. The A-site ordered manganese oxide $R$BaMn$_2$O$_6$ ($R:$ lanthanide) has attracted much attention because of significant physical phenomena such as the large magnetoresistance of 1,000 % at room temperature [@nakajima0], charge and orbital orderings at high temperatures [@nakajima1]. These properties attribute to A-site ordering working as “a periodic” Coulomb potential which stabilizes charge, spin, and/or orbital orderings of the electrons on B-site [@motome]. Owing to the randomness, A-site disordered phase of $R_{0.5}$Ba$_{0.5}$MnO$_3$ displays the spin-glass state or the itinerant ferromagnetic state instead of the charge ordered state seen in the ordered phase [@nakajima1; @akahoshi]. The A-site ordered cobalt oxide Sr$_3$YCo$_4$O$_{10.5}$ also exhibits peculiar magnetic properties with a high ferromagnetic (ferrimagnetic) transition temperature of 335 K [@kobayashi2], which is in contrast with a spin state crossover near 100 K in LaCoO$_3$ [@asai]. In this compound, the A-site ordering stabilizes oxygen deficient ordering, which makes a volume of the CoO$_6$ octahedron larger than that of LaCoO$_3$ [@ishiwata] and gives a different coordination number of Co$^{3+}$ from LaCoO$_3$. These modifications stabilize high-spin and/or intermediate-spin states of Co$^{3+}$ even at low temperatures, which causes the peculiar magnetic properties. Hence, partial substitution for the A-site or B-site cations strongly affects the spin state leading to significant suppression of the magnetic order; Ca substitution for Sr site [@yoshida] and 6%-Mn doping in the B site [@kobayashi4] destroys the room-temperature ferromagnetism of Sr$_3$YCo$_4$O$_{10.5}$. Recently, a new A-site ordered perovskite Y$_{0.8}$Sr$_{2.2}$Mn$_2$GaO$_{7.9}$ (Sr$_{2.93}$Y$_{1.07}$ Mn$_{2.66}$Ga$_{1.34}$O$_{10.53}$) was reported by Gillie [*et al.*]{} [@gillie], which is isostructural to Sr$_{3}$YCo$_4$O$_{10.5}$ [@istomin1; @withers]. As shown in Fig. 1, this compound has an octahedral site where Mn$^{3+}$ ions mainly occupy with about 10%-Ga ions intermixed and a tetrahedral site where both Mn$^{3+}$ and Ga$^{3+}$ ions occupy. They found an antiferromagnetic state below 100 K in this material showing that Y$_{0.8}$Sr$_{2.2}$Mn$_2$GaO$_{7.9}$ was a antiferromagnetic insulator, however, they did not report the transport properties. We have prepared polycrystalline samples of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=$ 0, 0.5, 1, 2) where $d$ represents oxygen deficiency and investigated the transport and magnetic properties in relation to the strcture. We have succeeded in preparing both A-site ordered and disordered phases for $x=1$ using different preparation conditions and observe a significant decrease of the resistivity in the disordered phase. We attribute this to a change of dimensionality in conduction and/or A-site disordered effect. This material can be a good playground to study order-disorder effect on the electronic states of the antiferromagnetic insulator with Mn$^{3+}$. ![(Color online) Crystal structure of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x\geq 1$). ](fig1.eps){width="4cm"} Experiments =========== Polycrystalline samples of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2) were prepared by a solid state reaction. Stoichiometric amounts of SrCO$_3$, Er$_2$O$_3$, Mn$_3$O$_4$, and Ga$_2$O$_3$ were mixed, and the mixture was sintered at 1250 $^{\circ}$C for 12 h for $x=0$ and 0.5, 6 h for $x=1$ and 2 in N$_2$ flow ($100-200$ ml/min). Then, the product was finely ground, pressed into a pellet, and sintered at 1250 $^{\circ}$C for 12 h for $x=0$, 24 h for $x=0.5$, and 10 h for $x=1$ and 2 in N$_2$ flow ($100-200$ ml/min). The second process was repeated 2 times for $x=0.5$, and once for $x=1$ and 2 with intermediate grindings and pelletizings. As shown in Fig. 2(c), $x=1$ sample exhibits A-site ordered structure. We sintered the $x=1$ sample again using the second process, and found the sample shows a disordered structure shown in Fig. 3. The x-ray diffraction of the sample was measured using a standard diffractometer with Cu K$\alpha$ radiation as an x-ray source in the $\theta -2\theta $ scan mode. The structural simulations were performed using a RIETAN-2000 program [@izumi]. The resistivity was measured by a four-probe method in a liquid He cryostat. The thermopower was measured using a steady-state technique in a liquid He cryostat with copper-constantan thermocouple to detect a small temperature gradient of about 1 K/cm. The magnetization was measured from 5 to 400 K by a commercial superconducting quantum interference device (SQUID, Quantum Design MPMS). Results and discussion ====================== ![(Color online) (a)-(d) X-ray diffraction patterns of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2). The insets represent the magnified x-ray patterns at low angles. The red line of the inset of Fig. 2(d) represents a simulated pattern. ](fig2.eps){width="6cm"} Figure 2 shows the x-ray diffraction patterns of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2). All the peaks are indexed as a cubic cell of the space group $Pm3m$ with the lattice parameter of $a\sim $ 3.8 and 3.85 Å for $x=$ 0 and 0.5, respectively. This cubic cell is also seen in Sr$_{1-x}$Y$_{x}$CoO$_{3-\delta }$ with small $x$ [@kobayashi2]. With increasing Ga content $x$, crystal structure changes from the cubic perovskite to a tetragonal A-site ordered perovskite (space group: $I4/mmm$, Fig. 1) with the lattice parameter of $a\sim $ 7.63 and 7.65 Å, and $c\sim $ 15.58 and 15.56 Å for $x=$ 1 and 2, respectively. This result is consistent with the structural analysis by Gillie [*et al.*]{} [@gillie]. As shown in the inset of Fig. 2(d), superstructure peaks corresponding to the A-site ordering are observed, while they do not appear for $x=$ 0 and 0.5 samples. Ga ions selectively occupy the tetrahedral site to stabilize the A-site ordered structure as shown in Fig. 1 [@gillie]. Thus, $x=1$ is a minimal amount to stabilize the structure. Gillie [*et al.*]{} reported that the oxygen content of Sr$_{2}$YMn$_{2}$GaO$_{7.9}$ is 7.9 showing formal valence of Mn ion is almost 3+. Thus, it is assumed that the formal valence of Mn ion is also 3+ in the ordered compounds presented here. Figure 3 shows the x-ray diffraction patterns of Sr$_{3}$ErMn$_{3}$GaO$_{10.5-d}$ with different preparation conditions. Obviously, two patterns are different; one sample was identified to the tetragonal ordered phase (hereafter this is denoted by O sample), and the other was identified to the disordered-cubic perovskite phase (D sample). Similar structures are originally found in $R$BaMn$_{2}$O$_{6}$/$R_{0.5}$Ba$_{0.5}$MnO$_3$ systems [@nakajima1]. We would like to say that the O sample is metastable so that the longer sintering stabilizes the disordered phase. Thus, this composition is just on the verge of order and disorder. As shown in the inset of Fig. 3, the superstructure peaks of D sample are hardly visible. ![(Color online) X-ray diffraction patterns of ordered and disordered Sr$_{3}$ErMn$_{2}$Ga$_2$O$_{10.5-d}$. The inset represents the magnified x-ray pattern at low angles. ](fig3.eps){width="6.5cm"} Figure 4 (a) shows the thermopower of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2). The magnitude of the thermopower is between $60$ and $110$ $\mu $V/K, and the sign is negative showing that the carriers are electrons. Assuming that the formal Mn valence is 3+, we expect that a tiny amount of electrons on Mn$^{2+}$ moves in the background of Mn$^{3+}$ as shown in the inset of Fig. 4(a). Using an extended Heikes formula [@koshibae], the valence of Mn ion was evaluated to be 2.71+ at 300 K corresponding to $d=0.43$ for $x=1$ sample with spin degeneracy term of $g_{\rm Mn^{2+}}=$6 and $g_{\rm Mn^{3+}}=$5. Figure 4 (b) shows the resistivity of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2). Semiconducting temperature dependence is observed for all the samples. With $x$, the magnitude of the resistivity decreases mainly due to decrease of scattering centers of the Ga ions. As seen in the inset of Fig. 4(b), the temperature dependence is described by an activation-type conduction $\rho =\rho _0$exp($\frac{E_{\rm g}}{k_{\rm B}T}$) where $E_{\rm g}$ represents activation energy above 200 K. The activation energy $E_{\rm g}$ was evaluated to be 0.133, 0.146, 0.149, and 0.246 eV for $x=$0, 0.5, 1, and 2, respectively. ![(Color online) (a) Thermopower and (b) resistivity of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=$0, 0.5, 1, and 2). ](fig4.eps){width="6cm"} Figure 5(a) shows the magnetization of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2). As shown in Figs. 5(b) and (c), the data was fitted by the Curie-Weiss law described by $\chi =\frac{C}{T-\theta _{\rm W}}+\chi _0$, where $C$, $\theta _{\rm W}$, and $\chi _0$ represent Curie constant, Weiss temperature and temperature independent term of the magnetic susceptibility, respectively. $C$ was evaluated to be 0.023, 0.019, 0.026, and 0.018 emu/K$\cdot $g for $x=0, 0.5, 1$, and 2, corresponding to 12.34, 11.14, 13.17, and 11.07 $\mu _{\rm B}$/f.u., respectively. These values are roughly explained by coexistence of 9.6 $\mu _{\rm B}$ of Er$^{3+}$ with $g=\frac{6}{5}$ and $J=\frac{15}{2}$ and 3.87 $\mu _{\rm B}$ and 4.90 $\mu _{\rm B}$ in the high-spin states of Mn$^{4+}$ and Mn$^{3+}$. $\chi _0$ was evaluated to be 1.39$\times$10$^{-5}$, 1.57$\times$10$^{-5}$, 0, and 0 emu/g for $x=0, 0.5, 1$, and 2 samples, respectively. Since $x=0$ and 0.5 samples show relatively good electric conductance compared with those of $x=1$ and 2 samples, this contribution may come from conducting electrons. All the samples exhibit negative $\theta _{\rm W}$ ($-$45, $-$29, $-$52, and $-$30 K for $x=0, 0.5, 1$, and 2, respectively) implying antiferromagnetic interaction in this system. At around 60 K, the slope of $\chi ^{-1}$ changes showing an existence of magnetic anomaly for all the samples. Since the antiferromagnetism was observed in Sr$_2$YMn$_2$GaO$_{8-d}$ below 100 K [@gillie], the anomaly at around 60 K in Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ can be also related to antiferromagnetic order of Mn$^{3+}$. ![(Color online) Magnetic susceptibility of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=$0, 0.5, 1, and 2). The inset shows inverse susceptibility. ](fig5.eps){width="6cm"} ![(Color online) (a) Resistivity and (b) thermopower and magnetization of the ordered and disordered phases. ](fig6.eps){width="7cm"} Lastly, we will discuss a difference of the transport and magnetic properties between the ordered (O sample) and disordered samples (D sample). Figure 6 (a) shows the temperature dependences of the resistivity of the O and D samples. The magnitude of the resistivity of the D sample is one order of magnitude smaller than that of the O sample, while the thermopower and magnetization of the D sample quite resemble those of the O sample as seen in Fig. 6(b). This strongly contrasts with the difference between the ferromagnetic-metal state of the disordered $R_{0.5}$Ba$_{0.5}$MnO$_3$ and charge-ordered insulating state of the ordered $R$BaMn$_2$O$_6$ [@nakajima1]. This variety of the properties comes from Mn$^{3.5+}$ with the charge degree of freedom, while Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ has Mn$^{3+}$ without the charge degree of freedom causing the small difference between the properties. The difference of the resistivity shown in Fig. 6(a) is explained by several possible origins as follows: (1) a decrease of the carrier concentration, (2) an increase of the scattering time, (3) a change of dimensionality in conduction and (4) a change of electronic structure induced by the disordering of A site. As seen in Fig. 6(b), two samples exhibit almost the same magnitude of the thermopower of $-100$ $\mu$V/K. Since thermopower is a function of carrier concentration, the result shows that carrier concentration does not differ so much in the two samples. In addition, since the content $x$ of Ga ions which can be scattering centers is 1 in both samples, a possibility of (2) can be also denied. As stated in the introduction, Ga ion selectively occupies tetrahedral sites in the O phase, while Ga and Mn ions randomly occupies in the D phase. Thus, a dimensionality in conduction can change from 2D in the O sample to 3D in the D sample, which can be a possible origin of the decrease of the resistivity. Another possibility is a change of electronic structure induced by the disordering of A site. A-site disordering generally induces a random potential in a perovskite oxide, which may affect the electronic structure of the material. Indeed, it is theoretically found that the introduced random potential in the Hubbard Hamiltonian causes an antiferromagnetic metallic phase instead of the antiferromagnetic insulating state [@shinaoka]. Although the origin of the difference of the resistivity is not clear at present, this system can be a good playground for investigating A-site disorder effect on antiferromagnetic insulator with Mn$^{3+}$. summary ======= In summary, we have measured x-ray diffraction, resistivity, thermopower, and magnetization of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ system, in which A-site ordered tetragonal phase appears above $x=1$, and observed large negative thermopower, semiconducting conduction, and magnetic susceptibility with a kink at 60 K implying antiferromagnetism. We succeed to prepare both A-site ordered and disordered phases for $x=1$ sample and observe a significant decrease of the resistivity in the disordered phase. We attribute this to a change of dimensionality in conduction and/or a change of electronic structure induced by the disordering of A sites. acknowledgements ================ We acknowledge I. Terasaki for fruitful discussion. This study was supported by the program entitled “Promotion of Environmental Improvement for Independence of Young Researchers” under the Special Coordination Funds for Promoting Science and Technology provided by MEXT, Japan. [99]{} email address: [email protected] T. Nakajima and Y. Ueda, J. Appl. Phys. [**98**]{} (2005) 46108. T. Nakajima, H. Kageyama, H. Yoshizawa and Y. Ueda, J. Phys. Soc. Jpn. [**71**]{} (2002) 2843. Y. Motome, N. Furukawa, and N. Nagaosa, Phys. Rev. Lett. [**91**]{} (2003) 167204. D. Akahoshi, M. Uchida, Y. Tomioka, T. Arima, Y. Matsui, and Y. Tokura, Phys. Rev. Lett. [**90**]{} (2003) 177203. W. Kobayashi, S. Ishiwata, I. Terasaki, M. Takano, I. Grigoraviciute, H. Yamauchi, and M. Karppinen, Phys. Rev. B [**72**]{} (2005) 104408. K. Asai, O. Yokokura, N. Nishimori, H. Chou, J. M. Tranquada, G. Shirane, S. Higuchi, Y. Okajima, and K. Kohn, J. Phys. Soc. Jpn [**50**]{} (1994) 3025. S. Ishiwata, W. Kobayashi, I. Terasaki, K. Kato, and M. Takata, Phys. Rev. B [**75**]{} (2007) 220406. S. Yoshida, W. Kobayashi, T. Nakano, I. Terasaki, K. Matsubayashi, Y. Uwatoko, I. Grigoraviciute, M. Karppinen, and H. Yamauchi, (submitted). W. Kobayashi, S. Yoshida, and I. Terasaki, Prog. Solid State Chem. [**35**]{} (2007) 355. L. J. Gillie, H. M. Palmer, A. J. Wright, J. Hadermann, G. Van Tendeloo, and C. Greaves, J. Phys. Chem. Solids [**65**]{} (2004) 87. S. Ya. Istomin, J. Grins, G. Svensson, O. A. Drozhzhin, V. L. Kozhevnikov, E. V. Antipov, and J. P. Attfield, Chem. Mater. [**15**]{} (2003) 4012 R. L. Withers, M. James, and D. J. Goosens, J. Solid State. Chem. [**174**]{} (2003) 198. F. Izumi and T. Ikeda, Mater. Sci. Forum [**198**]{} (2000) 321. W. Koshibae, K. Tsutsui, and S. Maekawa, Phys. Rev. B [**62**]{} (2000) 6869. H. Shinaoka, and M. Imada, Phys. Rev. Lett. [**102**]{} (2009) 016404.
{ "pile_set_name": "ArXiv" }
--- author: - 'Fernando Ramiro-Manzano' - Nikola Prtljaga - Lorenzo Pavesi - Georg Pucker - Mher Ghulinyan title: 'A fully integrated high-Q Whispering-Gallery Wedge Resonator' --- **Microresonator devices which posses ultra-high quality factors are essential for fundamental investigations and applications. Microsphere and microtoroid resonators support remarkably high Q’s at optical frequencies, while planarity constrains preclude their integration into functional lightwave circuits. Conventional semiconductor processing can also be used to realize ultra-high-Q’s with planar wedge-resonators. Still, their full integration with side-coupled dielectric waveguides remains an issue. Here we show the full monolithic integration of a wedge-resonator/waveguide vertically-coupled system on a silicon chip. In this approach the cavity and the waveguide lay in different planes. This permits to realize the shallow-angle wedge while the waveguide remains intact, allowing therefore to engineer a coupling of arbitrary strength between these two. The precise size-control and the robustness against post-processing operation due to its monolithic integration makes this system a prominent platform for industrial-scale integration of ultra-high-Q devices into planar lightwave chips.** Confining photons in a tiny dielectric volume of an ultra-high-Q (UHQ) cavity increases dramatically light-matter interactions. For this reason, high-Q resonators have been used for a number of fundamental investigations [@Aoki; @forces; @comb; @coptomech; @particle] and applications [@cuttedge; @lasers1; @lasers2]. In particular, UHQ resonators have been employed for a vast spectrum of studies in quantum photonics [@Aoki; @Vernooy], lasing [@lasers1; @lasers2], nonlinear optics [@comb; @opo-hydex; @gaeta], telecommunications [@Murugan; @flipflop], and sensing [@sensing1; @sensing2]. Engineering of UHQ’s in whispering-gallery type resonators has become an enticing objective. One of the challenges to reach such $Q$ values relies on reducing the light scattering at the interface between the resonator and the environment. This way UHQ’s were achieved by reflowing the resonator material, as it has been demonstrated for silica-based microspheres[@sphere] and microtoroids[@toroid]. The non-planarity, however, is the drawback which limits their integration with dielectric waveguides and, consequently, into planar photonic circuits. Another approach for achieving UHQ’s is to engineer the geometry of the planar microdisk resonators by realizing a shallow-angle wedge at its rim[@1stwedge; @2ndwedge]. As a result, the fundamental modes of the wedge-resonator are pushed far away from the scattering edge of the device and, hence, suffer less from surface-induced losses. Very recently, record UHQ’s of $\sim10^9$ have been reported for large diameter (5 mm) and relatively thick (10 $\mu$m) silica wedge resonators[@leearxiv]. ![ **Evanescent field coupling schemes in integrated photonics.** (a) A nanometric gap, $d_1$, between a microdisk resonator and a dielectric waveguide defines a side-coupling rate of $\gamma_1\sim\exp(-d_1)$. (b) Such coupling becomes increasingly inefficient when the pre-defined cavity/waveguide distance, $d_2$, increases during the formation of the shallow-angle wedge. In a realistic fabrication process also the waveguide retracts forming a wedge-shape (the dashed inclined line), thus, quenching completely the coupling. Because of this, typically, an off-chip technique (tapered-fiber) is used. (c) The vertical coupling allows for both engineering arbitrarily the evanescent field coupling by controlling the vertical gap, $d_v$, and aligning horizontally the waveguide to the resonator mode. (d) More importantly, this configuration enables the integrated photonics technology to access UHQ wedge resonators with buried dielectric waveguides. The vertical coupling scheme, in fact, permits to realize independently the wedge resonator, maintaining an appropriate coupling rate, $\gamma_v\sim\exp(-d_v)$. It allows for a direct access of devices through integrated waveguides and has the flexibility to engineer free-standing devices[@i3e-ptl]. []{data-label="scheme"}](scheme2a.jpg){width="6cm"} ![image](imagesa.jpg){width="15cm"} The planarity of wedge resonators, their CMOS-compatibility and accurate control of the processing entail an important step forward towards a full integration of UHQ devices into planar lightwave circuits. However, a last quest in this direction – *the integration with on-chip dielectric waveguides* – is still open. The mode retraction from the cavity rim precludes intrinsically the realization of side-coupled waveguides in the close proximity of the resonator in order to provide an appropriate mode coupling (Fig. \[scheme\]a,b). In this work, we show that this important milestone can be reached by opting for a vertical coupling scheme between the wedge microresonator and a bus-integrated waveguide (Fig. \[scheme\]c,d). Thereby, this study focuses on the proof-of-concept demonstration of the all-on-chip complete integration of wedge resonators. Recently, the feasibility of this technology in realizing free-standing microdisk and spiderweb resonators coupled vertically to integrated waveguides through an air gap has been demonstrated[@i3e-ptl]. The resonator-waveguide vertical coupling does not require expensive lithographic techniques for the gap definition at a desired precision and allows for an independent choice of materials and thicknesses for optical components. Such advantages can be of great utility in a number of applications to high-speed integrated photonics as well as to cavity optomechanics[@coptomech], requiring high ($\sim10^4$) or ultra-high Q’s, respectively. Owing to the separation of the resonator and the waveguide into different planes, our approach enables, on one side, to realize the waveguide and the shallow-angle wedge resonator in different technological steps, and on the other side, to define arbitrary and independently the vertical coupling gap size. We realized 400 nm-thick and $50~\mu$m-diameter silicon nitride (SiN$_x$) wedge resonators vertically coupled to silicon oxynitride (SiO$_x$N$_y$) waveguides using standard silicon microfabrication tools (Fig. \[sem\]a,c, see also Methods). In order to prove the suppression of surface-induced losses in a fully integrated device, we also realized devices identical to wedge resonators with the only difference of using a conventional dry ion-etching step for the resonator definition (Fig. \[sem\]b,d). Spectroscopic characterization and numerical mode analysis were thus performed for both types of devices. ![image](spectra_newa.jpg){width="16cm"} The wedge resonator was formed during the transfer of the circular photoresist pattern into the SiN$_x$ layer. Due to the adhesion properties of the photoresist a $\theta=7^{\circ}$ sharp angle wedge is formed by the end of etching. Numerical calculations show that the fundamental mode is retracted to a 2.2 $\mu$m smaller effective radius, $r_{eff}$, with respect to the dry-etched (microdisk, hereafter) resonator (Fig. \[sem\]e,g). The devices were characterized in typical waveguide transmission experiments in a broad near-infrared wavelength range between 1350 nm and 1600 nm (Fig. \[spectra\]a). Figure  \[spectra\]b shows a series of sharp and broad resonances corresponding to first- (fundamental) and second-order radial mode families of the wedge resonator. A blow-up of the spectrum around a fundamental mode with $-15$ dB of transmission suppression is shown in the top panel of Fig. \[spectra\]c. By accounting for the critical coupling of this mode to the waveguide, an intrinsic Q$_i\approx7.6\times10^4$ is found from a Lorentzian fit. Interestingly, the fundamental mode of the disk device at the very similar wavelength shows an intrinsic Q$_i$ of only $1.8\times10^4$ (bottom panel in Fig. \[spectra\]c). A similar three to fourfold difference in quality factors of the fundamental modes in the wedge and the disk resonators is observed over a broad spectral range (Fig. \[spectra\]d). Contrary, the second order mode families show much lower and close $Q$’s over the analyzed spectral range. At a shorter wavelength ($\sim1370.5$ nm) we observe that the transmission curve at the wedge resonator’s undercoupled fundamental mode $M_w=166$ is split into a doublet of symmetric and antisymmetric modes due to scattering. The individual resonances have $Q_{1}=2\times10^5$ and $Q_{2}=1.56\times10^5$, respectively (Fig. \[spectra\]e, top panel). A fit to the spectrum, using the model for a coherent sum of Lorentzian lineshapes[@borselliOE], shows that the Q’s of the individual modes are narrower than the scattering-related $Q_s=\omega_0/\gamma_s$ ($\omega_0$ is the uncoupled mode frequency and $\gamma_s$ is the scattering rate). This indicates to the formation of largely separated standing wave modes. Contrary, no doublet was observed for the disk’s (similarly undercoupled) fundamental mode at $\lambda\approx$1368.7 nm ($M_d=186$) (Fig. \[spectra\]e, bottom panel). In this case, an important linewidth broadening ($Q_{1,2}\sim3.5\times10^4$) and no observable doublet splitting ($Q_{1,2}<<Q_s$) were found from the sum-Lorentzian fit. ![**Numerical analysis of mode confinement and intensity profiles.** (a) Numerically calculated peak wavelengths of the first (M1) and the second (M2) order radial modes of the wedge resonator show an excellent agreement with the experimentally measured ones. (b) The calculated mode confinement factors are reported both for the wedge and the dry-etched resonators. (c) The horizontal and (d) the vertical profiles (through field maxima) of the first order radial modes of two types of resonators around $\lambda=1576$ nm. In particular, the mode maximum in the wedge resonator is situated at a $\sim1\mu$m smaller radius with respect to the dry-etched case, therefore, is more isolated form the physical edge of the device. The vertical profiles show an identical confinement in both cases. Panels (e) and (f) show the situation for the second order radial modes. In this case, the mode M2 in the wedge resonator is closely situated near the physical edges of the resonator. Insets in (c) and (e) show a close-up of the intensity profiles at the resonator/air interface. []{data-label="fem"}](analysis2a.jpg){width="8.5cm"} In Fig. \[fem\]a we compare the measured resonance positions of two radial families of the wedge resonator to the numerically calculated ones. The observed sub-nanometer-precision matching of mode wavelengths permits us to use the numerical model for further analysis of the modal characteristics of the devices. The total intrinsic loss of a whispering-gallery resonator is given as $1/Q_i=1/Q_r+1/Q_m+1/Q_s$, where $Q_r^{-1}$ is the radiative loss, $Q_m^{-1}$ is the modal loss related to the material absorption and $Q_s^{-1}$ is the loss contribution accounting for both the surface scattering and surface absorption. The radiative Q’s can be estimated using the ray-optics approach[@Levi] and for $50~\mu$m-diameter devices are beyond $10^{19}$, having therefore an increasingly negligible contribution to the overall $Q_i$ for both types of resonators. The quality factor, related to the material and modal loss is $Q_m=\lambda(\Gamma\alpha\Lambda r_{eff})^{-1}$, where $\alpha$ is the bulk material’s absorption coefficient, $\Lambda$ is the free-spectral range of the cavity modes, $\lambda$ is the wavelength and $\Gamma$ is the confinement factor of the mode. Figure \[fem\]b shows that the first radial family modes of the wedge resonator have a confinement of $\sim80\%$ while the $\Gamma$’s in the disk are slightly ($\sim4\%$) larger. Considering that the free-spectral ranges are of about 9.1 nm and 8 nm in the wedge and the disk resonators, respectively, the estimated $Q_m$’s for both devices are very similar. The last contribution $Q_s^{-1}$ is taking into account the sidewall effects in terms of mode’s interaction with the resonators external rim by scattering due to roughness and absorption by specimen on the surface. The horizontal profiles across the fundamental modes brightest spot (Fig. \[fem\]c) show that the mode ($M_w=141$) in the wedge resonator is retracted from the etched interface by almost 1 $\mu$m with respect to the disk’s case ($M_d=158$) and, consequently, is better isolated from scattering and absorption. On the other hand, the vertical profiles of these modes coincide (Fig. \[fem\]d), suggesting that $Q_s$–contribution from planar interfaces is nearly identical for both cases. This analysis shows that under the conditions of negligible radiative and very similar material losses the main contribution to the intrinsic $Q_i$ comes from the surface-related $Q_s$. In view of the experimental results (Fig. \[spectra\]d), the observed differences in $Q$’s confirm the effective suppression of surface-induced losses in wet-etched devices[@1stwedge; @2ndwedge; @leearxiv]. This conclusions are further supported by the analysis of horizontal (Fig. \[fem\]e) and vertical (Fig. \[fem\]f) profiles of the second-order mode families. In this case, the modes in the wedge resonator have a significant overlap with the interfaces as compared to the disk’s case, hence, larger surface-related losses are expected to significantly limit the $Q$ (see the experimental M2 data-sets in Fig. \[spectra\]d). In conclusion, these results show the feasibility of using conventional silicon microfabrication tools for the realization of planar high-Q wedge resonators monolithically integrated with vertically coupled dielectric waveguides. The integration of waveguides simplifies significantly the access to the resonator, avoiding thus the use of complicated tapered-fiber free-space coupling schemes and guaranteeing stable operation of the device. By opting for a vertical evanescent coupling scheme, several critical issues are resolved; (i) the resonator and the waveguide are fabricated in different fabrication steps without influencing one another, (ii) the wedge angle of the resonator can be controlled without interfering with underlaying waveguiding components, (iii) the waveguide can be freely aligned to the mode position in the resonator, and (iv) the vertical gap between the resonator and the waveguide can be defined for engineering an evanescent coupling of arbitrary strength. The results of this study will boost an intensive research towards the complete integration of fully functional ultra-high quality factor planar resonators into planar photonic circuits. For this ultimate scope, all necessary ingredients, such as the transparent silicon-based materials[@leearxiv] as well as an easily accessible technology to realize integrated free-standing devices[@i3e-ptl] are already available.\ **Methods**\ [99]{} Aoki, T., Dayan, B., Wilcut, E., Bowen, W. P., Parkins, A., Kippenberg, T., Vahala, K., and Kimble, H. Observation of strong coupling between one atom and a monolithic microresonator, Nature **443**, 671-674 (2006). Rakich, P.T., Popovic, M.A., Soljačic, M. and Ippen, E.P. Trapping, corralling and spectral bonding of optical resonances through optically induced potentials. Nature Photon. **1**, 658-665 (2007). Del’Haye, P., Schliesser, A., Arcizet, O., Wilken, T., Holzwarth, R., and Kippenberg, T. J. Optical frequency comb generation from a monolithic microresonator, Nature 450, 1214-1217 (2007). Kippenberg, T.J., and Vahala, K.J. Cavity Optomechanics: Back-Action at the Mesoscale. Science **321**, 1172 (2008). Zhu, J., Ozdemir, S. K., Xiao, Y.-F., Li, L., He, L., Chen, D.-R., and Yang, L. On-chip single nanoparticle detection and sizing by mode splitting in an ultrahigh-Q microresonator, Nature Photon. **4**, 46-49 (2010). *“Practical Applications of Microresonators in Optics and Photonics,"* Edt. A.B. Matsko (CRC Press, 2009). Michler, P., Kiraz, A., Zhang, L., Becher, C., Hu, E., and Imamoglu, A. Laser emission from quantum dots in microdisk structures. Appl. Phys. Lett. **77**, 184-186 (2000). Zhang, Zh., Yang, L., Liu, V., Hong, T., Vahala, K., and Scherer, A. Visible submicron microdisk lasers, Appl. Phys. Lett. **90**, 111119 (2007). Vernooy, D., Furusawa, A., Georgiades, N., Ilchenko, V., and Kimble, H. Cavity QED with high-Q whispering gallery modes, Phys. Rev. A **57**, 2293-2296 (1998). Razzari, L., Duchesne, D., Ferrera, M., Morandotti, R., Chu, S., Little, B. E., and Moss, D. J. CMOS-compatible integrated optical hyper-parametric oscillator, Nature Photon. **4**, 41-45 (2009). Johnson, A.R., Okawachi, Y., Levy, J.S., Cardenas, J., Saha, K., Lipson, M. and Gaeta, A.L., Chip-based frequency combs with sub-100 GHz repetition rates, Opt. Lett. **37**, 875-877 (2012). Murugan, G. S., Wilkinson, J. S. and Zervas, M. N. Optical excitation and probing of whispering gallery modes in bottle microresonators: potential for all-fiber add-drop filters, Opt. Lett. **35**, 1893-1895, (2010). Liu, L., Kumar, R., Huybrechts, K., Spuesens, T., Roelkens, G., Geluk, E.-J., de Vries, T., Regreny, P., Van Thourhout, D., Baets, R., and Morthier, G. An ultra-small, low-power, all-optical flip-flop memory on a silicon chip, Nature Photon. **4**, 182–187 (2010). Vollmer, F., Braun, D., Libchaber, A., Khoshsima, M.,Teraoka, I., and Arnold, S. Protein detection by optical shift of a resonant microcavity, Appl. Phys. Lett. **80**, 4057-4059 (2002). Hanumegowda, N. M., Stica, C. J., Patel, B. C., White, I., and Fan, X. Refractometric sensors based on microsphere resonators, Appl. Phys. Lett. **87**, 201107 (2005). Vernooy, D. W., Ilchenko, V. S., Mabuchi, H., Streed, E. W. & Kimble, H. J. High-Q measurements of fused-silica microspheres in the near infrared. Opt. Lett. **23**, 247–249 (1998). Armani, D.K., Kippenberg, T.J., Spillane S.M., and Vahala, K.J. Ultra-high-Q toroid microcavity on a chip. Nature **421**, 925-928 (2003). Kippenberg, T. J., Spillane, S. M., Armani, D. K., and Vahala, K. J. Fabrication and coupling to planar high-Q silica disk microcavities, Appl. Phys. Lett. **83**, 797-799 (2003). Kippenberg, T. J., Kalkman, J., Polman, A., and Vahala, K. J. Demonstration of an erbium-doped microdisk laser on a silicon chip, Phys. Rev. A **74**, 051802 (2006). Lee, H., Chen, T., Li, J., Yang, K. Y., Jeon, S., Painter, O., and Vahala, K.J. Chemically etched ultrahigh-Q wedge-resonator on a silicon chip, Nature Photon. **6**, 369-373 (2012). Ghulinyan, M., Guider, R., Pucker G., and Pavesi, L. Monolithic whispering-gallery mode resonators with vertically coupled integrated bus waveguides, IEEE Photon. Technol. Lett. **23**, 1166-1168 (2011). Borselli, M., Johnson, T. J., and Painter, O. Beyond the Rayleigh scattering limit in high-Q silicon microdisks: theory and experiment, Opt. Express **13**, 1515-1530 (2005). Frateschi, N. C., and Levi, A. F. J. The spectrum of microdisk lasers, J. Appl. Phys. **80**, 644-653 (1996).
{ "pile_set_name": "ArXiv" }
ULB-TH/13-02 **** [Glenn Barnich$^{a}$ and Pierre-Henry Lambert$^{b}$]{} ** Physique Théorique et Mathématique\ Université Libre de Bruxelles\ and\ International Solvay Institutes\ Campus Plaine C.P. 231, B-1050 Bruxelles, Belgium <span style="font-variant:small-caps;">Abstract</span>. We show that the symmetry algebra of asymptotically flat four dimensional spacetimes at null infinity in the sense of Newman and Unti is isomorphic to the direct sum of the abelian algebra of infinitesimal conformal rescalings with $\mathfrak{bms}_4$. We then work out the local conformal properties of the relevant Newman-Penrose coefficients, as well as the surface charges and their algebra. *Proceedings of the* **International Conference on Quantum Field Theory and Gravity***, July 31–August 4, 2012, Center of Theoretical Physics, Tomsk State Pedagogical University, Russia, the* **Workshop on Gravity, Quantum, and Black Holes** *at the International Conference on Mathematical Modeling in Physical Sciences, September 3–7, 2012, Budapest, Hungary, and the **8th edition of the Workshop on Quantum Field Theory and Hamiltonian Systems**, September 19-22, 2012, Craiova, Romania.* Introduction ============ This conference proceedings summarizes the results of paper [@Barnich:2011nu] to which we refer for detailed computations and discussions. The definitions of asymptotically flat four dimensional space-times at null infinity by Bondi-Van der Burg-Metzner-Sachs [@Bondi:1962px; @Sachs:1962wk] (BMS) and Newman-Unti (NU) [@newman:891] in 1962 merely differ by the choice of the radial coordinate. Such a change of gauge should not affect the asymptotic symmetry algebra if, as we contend, this concept is to have a major physical significance. The problem of comparing the symmetry algebra in both cases is that, besides the difference in gauge, the very definitions of these algebras are not the same. Indeed, NU allow the leading part of the metric induced on Scri to undergo a conformal rescaling. When this generalization is considered in the BMS setting, it turns out that the symmetry algebra is the direct sum of the BMS algebra $\mathfrak{bms}_4$ [@Sachs2] with the abelian algebra of infinitesimal conformal rescalings [@Barnich:2009se], [@Barnich:2010eb]. In this note we show that, as expected, the asymptotic symmetry algebra in the NU framework is again the direct sum of $\mathfrak{bms}_4$ with the abelian algebra of infinitesimal conformal rescalings of the metric on Scri and thus coincides, as it should, with the generalized symmetry algebra in the BMS approach. We then discuss the transformation properties of the Newman-Penrose coefficients parametrizing solution space in the NU approach, focussing on the inhomogeneous terms in the transformation laws that contain the information on the central extensions of the theory, and we finally study the associated surface charges and their algebra by following the analysis in the BMS gauge [@Barnich:2011mi]. NU metric ansatz and asymptotic symmetries {#sec:from-bondi-van} ========================================== The metric ansatz of NU can be written as $$\label{eq:10} ds^2=Wdu^2-2dr du+ g_{AB}(dx^A-V^Adu)(dx^B-V^Bdu)\,,$$ with coordinates $u,r,x^A$ and where $ g_{AB}dx^Adx^B=r^2\bar\gamma_{AB}dx^Adx^B +r C_{AB}dx^Adx^B+o(r)\,,\label{eq:11} $ with $\bar\gamma_{AB}$ conformally flat. Below, we will use standard stereographic coordinates $\zeta=\cot{\frac{\theta}{2}}e^{i\phi},\bar \zeta$, $\bar\gamma_{AB}dx^Adx^B=e^{2{{\widetilde}\varphi}} d\zeta d\bar\zeta$, ${{\widetilde}\varphi}={{\widetilde}\varphi}(u,x)$. There is also an additional condition, related to the fixing of the origin of the affine parameter of the null geodesic generators of the null hypersurfaces used to build the metric [@newman:891], which yields here $C^A_A=0$ [@Barnich:2011nu]. In the following we denote by $\bar D_A$ the covariant derivative with respect to $\bar \gamma_{AB}$ and by $\bar\Delta$ the associated Laplacian. The fall-off conditions are $ V^A=O(r^{-2})$ and $W=-2 r{\partial}_u {{\widetilde}\varphi}+\bar\Delta {{\widetilde}\varphi}+O(r^{-1})\,, $ where $\bar \Delta {{\widetilde}\varphi}=4 e^{-2{{\widetilde}\varphi}}{\partial}\bar{\partial}{{\widetilde}\varphi}$ with ${\partial}= {\partial}_\zeta,\bar{\partial}={\partial}_{\bar\zeta}$. The infinitesimal NU transformations are defined as those infinitesimal transformations that leave the form of the metric and the fall-off conditions invariant, up to a rescaling of the conformal factor $\delta{{\widetilde}\varphi}(u,x^A)={{\widetilde}\omega}(u,x^A)$, and are in this case generated by $$\begin{gathered} \label{eq:26} \left\{\begin{array}{l} \xi^u=f,\\ \xi^A=Y^A+I^A, \quad I^A=- {\partial}_B f\int_r^\infty dr^\prime g^{AB},\\ \xi^r=-r {\partial}_u f+Z+ J,\quad J={\partial}_A f \int_r^\infty dr^\prime V^A, \end{array}\right.\end{gathered}$$ with ${\partial}_r f=0={\partial}_r Y^A={\partial}_r Z$, $Z={\frac{1}{2}}\bar\Delta f$, ${\partial}_u Y^A=0$, with $Y^A$ a conformal Killing vector of $\bar\gamma_{AB}$, i.e. $Y^\zeta \equiv Y=Y(\zeta),\quad Y^{\bar \zeta} \equiv \bar Y=\bar Y(\bar \zeta)$ in the coordinates ($\zeta,\bar\zeta$), and also with $$\label{eq:14} f=e^{ {{\widetilde}\varphi}}\big[ {\widetilde}T+ {\frac{1}{2}}\int_0^u du^\prime e^{- {{\widetilde}\varphi}}{\widetilde}\psi\big],\quad {\widetilde}T= {\widetilde}T(\zeta,\bar \zeta),$$ with $\psi=\bar D_AY^A$ and ${\widetilde}\psi=\psi-2{\widetilde}\omega$. Asymptotic Killing vectors thus depend on $Y^A,{\widetilde}T,{{\widetilde}\omega}$ and the metric, $\xi=\xi[Y,{\widetilde}T,{{\widetilde}\omega};g]$. For such metric dependent vector fields, consider the suitably modified Lie bracket taking the metric dependence of the spacetime vectors into account, $ [\xi_1,\xi_2]_M=[\xi_1,\xi_2]-\delta^g_{ \xi_1}\xi_2+\delta^g_{ \xi_2}\xi_1, $ where $\delta^g_{\xi_1}\xi_2$ denotes the variation in $\xi_2$ under the variation of the metric induced by $\xi_1$, $\delta^g_{ \xi_1}g_{\mu\nu}=\cL_{\xi_1}g_{\mu\nu}$. Consider now the extended $\mathfrak{bms}_4$ algebra, i.e., the semi-direct sum of the algebra of conformal Killing vectors of the Riemann sphere with the abelian ideal of infinitesimal supertranslations, trivially extended by infinitesimal conformal rescalings of the conformally flat degenerate metric on Scri. The commutation relations are given by $[(Y_1,{\widetilde}T_1,{{\widetilde}\omega}_1),(Y_2,{\widetilde}T_2,{{\widetilde}\omega}_2)]=({\widehat}Y,{\widehat}{{\widetilde}T},{\widehat}{{{\widetilde}\omega}})$ where $$\left\{\begin{array}{l} \label{eq:5}{\widehat}Y^A= Y^B_1{\partial}_B Y^A_2-Y^B_2{\partial}_B Y^A_1,\\ {\widehat}{{\widetilde}T}=Y^A_1{\partial}_A {\widetilde}T_2-Y^A_2{\partial}_A {\widetilde}T_1 +{\frac{1}{2}}({\widetilde}T_1{\partial}_AY^A_2-{\widetilde}T_2{\partial}_AY^A_1),\\ {\widehat}{{{\widetilde}\omega}}=0\,. \end{array}\right.$$ In these terms, one can show the following: The spacetime vectors $\xi[Y,{\widetilde}T,{{\widetilde}\omega};g]$ realize the extended $\mathfrak{bms}_4$ algebra in the modified Lie bracket, $$\Big[\xi[Y_1,{\widetilde}T_1,{{\widetilde}\omega}_1;g],\xi[Y_2,{\widetilde}T_2,{{\widetilde}\omega}_2;g]\Big]_M= \xi[{\widehat}Y,{\widehat}{{\widetilde}T},{\widehat}{{{\widetilde}\omega}};g]\,,\label{eq:1}$$ in the bulk of an asymptotically flat spacetime in the sense of Newman and Unti. Note in particular that for two different choices of the conformal factor ${{\widetilde}\varphi}$ which is held fixed, ${{\widetilde}\omega}=0$, the asymptotic symmetry algebras are isomorphic to $\mathfrak{bms}_4$, which is thus a gauge invariant statement. Explicit relations between the NU and the BMS gauges\ and local conformal transformation laws of the NU coefficients ==============================================================   The choice of the radial coordinate in the definition of asymptotically flat space-times in the BMS [@Bondi:1962px], [@Sachs:1962wk], [@Sachs2] and the NU [@newman:891] approaches differs but the relation between the two radial coordinates does not involve constant terms [@Barnich:2011nu] and is of the form $ r^\prime=r +O(r^{-1})\label{eq:3}\,. $ This change of coordinates only affects lower order terms in the asymptotic expansion of the metric that play no role in the definition of asymptotic symmetries and explains a posteriori why the asymptotic symmetry algebras in both approaches are isomorphic. In the BMS set-up, the general solution to Einstein’s field equations is parametrized by some functions [@Bondi:1962px], [@Sachs:1962wk], [@Barnich:2010eb] among which are the mass and angular momentum aspects, and the news tensors. In the NU case instead [@newman:891], the free data characterizing solution space are described in terms of the spin coefficient $\sigma^0$ and its time derivative, and also in terms of the $\Psi_\alpha^0$ (with $\alpha=0,1,2,3,4$), five complex scalars representing all the components of the Weyl tensor. The explicit relations between the free data characterizing asymptotic solution space in both approaches were established for instance in [@Barnich:2011nu]. Using the “eth” operators [@newman:863] defined for a field $\eta^s$ of spin weight $s$ according to the conventions of [@Penrose:1986] through $ \eth \eta^s= P^{1-s}\bar {\partial}(P^s\eta^s)\,, \bar \eth \eta^s=P^{1+s}{\partial}(P^{-s}\eta^s)$ with $ P=\sqrt 2 e^{-{\widetilde}\varphi}\,, $ where $\eth,\bar\eth$ raise respectively lower the spin weight by one unit and let $\cY=P^{-1} \bar Y$ and $\bar \cY=P^{-1} Y$. The conformal Killing equations and the conformal factor then become $\eth \bar \cY=0=\bar\eth \cY$ and $\psi=(\eth \cY+\bar\eth \bar \cY)$. Using the notation $S=(Y,{\widetilde}T,{\widetilde}\omega)$, we have $-\delta_S \bar \gamma_{AB}=2{\widetilde}\omega \bar\gamma_{AB}$ for the background metric. To work out the transformation properties of the NU coefficients characterizing asymptotic solution space, one needs to evaluate the subleading terms in the Lie derivative of the metric on-shell. This can also be done by translating the results from the BMS gauge, using the dictionary of [@Barnich:2011nu], which yields in this case $$\begin{split} \label{eq:16b} -\delta_S \sigma^0 & = [f{\partial}_u+\cY\eth+ \bar \cY\bar\eth+\frac{3}{2}\eth \cY-\frac{1}{2} \bar\eth \bar \cY-{\widetilde}\omega] \sigma^0-\eth^2 f\,,\\ -\delta_S \dot\sigma^0 & = [f{\partial}_u+ \cY\eth + \bar \cY\bar\eth+2\eth \cY-2{\widetilde}\omega]\dot\sigma^0-{\frac{1}{2}}\eth^2{\widetilde}\psi\,,\\ -\delta_S\Psi^0_i&=[f{\partial}_u+\cY\eth+\bar\cY\bar\eth+\frac{5-i}{2}\eth\cY +\frac{1+i}{2}\bar\eth\bar\cY-3{\widetilde}\omega]\Psi^0_i+(4-i)\eth f\Psi^0_{i+1}\,,\\ \end{split}$$ with $i=1,2,3,4$. Surface charge algebra {#sec:surf-charge-algebra} ====================== In this section, ${\widetilde}\omega=0$ so that $f=T+{\frac{1}{2}}u\psi$ and we use the notation $s=(\cY,\bar\cY,T)$ for elements of the symmetry algebra, which is given in these terms by $[s_1,s_2]={\widehat}s$ where $$\begin{gathered} \label{eq:57} {\widehat}\cY=\cY_1\eth \cY_2 -(1\leftrightarrow 2),\qquad {\widehat}{\bar\cY}=\bar \cY_1\bar \eth \bar \cY_2 -(1\leftrightarrow 2),\\ {\widehat}T= (\cY_1\eth +\bar \cY_1\bar \eth)T_2-{\frac{1}{2}}\psi_1 T_2 -(1\leftrightarrow 2)\,. \end{gathered}$$ The translation of the charges, the non-integrable piece due to the news and the central charges computed in [@Barnich:2011mi] gives here Q\_[s]{}\[\]&=-d\^2\^,\ \_[s]{}\[,\]&=d\^2 \^ f , \[eq:19\]\ K\_[s\_1,s\_2]{}\[\]&= d\^2 \^ . We recognize all the ingredients of the surface charges described in [@0264-9381-1-1-005]. More precisely the angular (super-)momentum that we get is $$Q_{\cY,0,0}=-\frac{1}{8\pi G}\int d^2\Omega^\varphi\, \cY\Big[\Psi^0_1 +\sigma^0\eth\bar\sigma^0+{\frac{1}{2}}\eth(\sigma^0\bar\sigma^0) - \frac{u}{2}\eth\big(\Psi^0_2+\bar\Psi^0_2+ {\partial}_u(\sigma^0\bar\sigma^0)\big)\Big]\,. \label{eq:49}$$ and differs from $Q_{\eta_c}$ given in equation (4) of [@0264-9381-1-1-005] by the explicitly $u$-dependent term of the second line. It thus has a similar structure to Penrose’s angular momentum as described in equations (11), (12), and (17a) of [@0264-9381-1-1-005] in the sense that it also differs by a specific amount of linear supermomentum, but the amount is different and explicitly $u$-dependent, $ Q_{\cY,0,0}= Q^{u=0}_{\cY,0,0}+{\frac{1}{2}}u Q_{0,0,\eth\cY}\,. $ The main result derived in [@Barnich:2011mi] states that if one is allowed to integrate by parts, and if one defines the “Dirac bracket” through $ \label{eq:45} \{Q_{s_1},Q_{s_2}\}^*[\cX]=-\delta_{s_2} Q_{s_1}[\cX]+\Theta_{s_2}[-\delta_{s_1}\cX,\cX], $ then the charges define a representation of the $\mathfrak{bms}_4$ algebra, up to a field dependent central extension, $ \{Q_{s_1},Q_{s_2}\}^*=Q_{[s_1,s_2]} +K_{s_1,s_2}, $ where $K_{s_1,s_2}$ satisfies the generalized cocycle condition $ K_{[s_1,s_2],s_3}-\delta_{s_3} K_{s_1,s_2}+{\rm cyclic} (1,2,3)=0\,. $ This representation theorem can be verified directly in the present context [@Barnich:2011nu]. To the best of our knowledge, except for the previous analysis in the BMS gauge, the above representation result does not exist elsewhere in the literature. A major issue in these considerations is whether one uses the globally well-defined version of the $\mathfrak{bms}_4$ algebra or a local version which contains the Virasoro algebra and involves an expansion in terms of Laurent series. The formulas presented above generally apply to both cases, except for divergences in the charges that appear in the second case and have to be handled properly. This is discussed in more details in [@Barnich:2011nu]. Acknowledgements {#sec:acknowledgements .unnumbered} ================ The authors thank Cédric Troessaert for useful discussions. G.B. is Research Director of the Fund for Scientific Research-FNRS (Belgium). This work is supported in part by the Belgian Federal Science Policy Office through the Interuniversity Attraction Pole P6/11, by IISN-Belgium, by “Communauté française de Belgique - Actions de Recherche Concertées” and by Fondecyt Projects No. 1085322 and No. 1090753. [10]{} url \#1[[\#1]{}]{}urlprefix\[2\]\[\][[\#2](#2)]{} Barnich G and Lambert P-H 2012 Advances in Mathematical Physics, vol. 2012, Article ID 197385, [[1102.0589v3]{}](http://arxiv.org/abs/1102.0589). Bondi H, van der Burg M G and Metzner A W 1962 [*Proc. Roy. Soc. Lond. A*]{} [**269**]{} 21 Sachs R K 1962 [*Proc. Roy. Soc. Lond. A*]{} [**270**]{} 103. Newman E T and Unti T W J 1962 [*Journal of Mathematical Physics*]{} [**3**]{} 891–901 <http://link.aip.org/link/?JMP/3/891/1>. Sachs R K 1962 [*Phys. Rev.*]{} [**128**]{} 2851–2864. Barnich G and Troessaert C 2010 [*Phys. Rev. Lett.*]{} [**105**]{} 111103, [[0909.2617]{}](http://arxiv.org/abs/0909.2617). Barnich G and Troessaert C 2010 [*JHEP*]{} [**05**]{} 062, [[1001.1541]{}](http://arxiv.org/abs/1001.1541). Barnich G and Troessaert C 2011 [*JHEP*]{} [**1112**]{} 105, [[1106.0213]{}](http://arxiv.org/abs/1106.0213). Newman E T and Penrose R 1966 [*Journal of Mathematical Physics*]{} [**7**]{} 863–870 <http://link.aip.org/link/?JMP/7/863/1>. Penrose R and Rindler W 1986 [*[Spinors and Space-Time, Volume 2: Spinor and Twistor Methods in Space-Time Geometry]{}*]{} (Cambridge University Press). Dray T and Streubel M 1984 [*Classical and Quantum Gravity*]{} [**1**]{} 15–26 <http://stacks.iop.org/0264-9381/1/15>.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Estimates of the bulk metal abundance of the Sun derived from the latest generation of model atmospheres are significantly lower than the earlier standard values. In Paper I we demonstrated that a low solar metallicity is inconsistent with helioseismology if the quoted errors in the atmospheres models (of order 0.05 dex) are correct. In this paper we undertake a critical analysis of the solar metallicity and its uncertainty from a model atmospheres perspective, focusing on CNO. We argue that the non-LTE corrections for abundances derived from atomic features are overestimated in the recent abundance studies, while systematic errors in the absolute abundances are underestimated. If we adopt the internal consistency between different indicators as a measure of goodness of fit, we obtain intermediate abundances \[C/H\] = 8.44 +/- 0.06, \[N/H\] = 7.96 +/- 0.10 and \[O/H\] = 8.75 +/- 0.08. The errors are too large to conclude that there is a solar abundance problem, and permit both the high and low scales. However, the center-to-limb continuum flux variations predicted in the simulations appear to be inconsistent with solar data, which would favor the traditional thermal structure and lead to high CNO abundances of (8.52, 7.96, 8.80) close to the seismic scale. We argue that further empirical tests of non-LTE corrections and the thermal structure are required for precise absolute abundances. The implications for beryllium depletion and possible sources of error in the numerical simulations are discussed.' author: - | M.H. Pinsonneault$^1$ & Franck Delahaye $^{1,2}$\ 1 Department of Astronomy, The Ohio State University, Columbus OH 43210 USA\ 2 LUTH, (UMR 8102 associée au CNRS et à l’Université Paris 7), Observatoire de Paris, F-92195 Meudon, France. title: ' The Solar Heavy Element Abundances: II. Constraints from Stellar Atmospheres.' --- Introduction ============ The uncertainty in the absolute chemical composition of stars is the limiting factor in our ability to do high precision stellar astrophysics. Traditionally, we have had to rely on a small database of fundamental stellar parameters such as mass, distance, and radius. However, current and upcoming space missions promise a wealth of astrometric and photometric data. Large surveys undertaken primarily for other purposes (microlensing, planet searches and cosmology) have discovered thousands of eclipsing binaries, yielding numerous precise mass estimates. The rapidly developing field of optical interferometry has also permitted a growing number of direct radius estimates. Asteroseismology is also growing in importance, and missions such as COROT promise a wealth of detailed information on the pulsational properties of solar-like stars. Our stellar interiors models have become highly sophisticated and successful when compared with observational diagnostics. In particular, the resolution of the solar neutrino problem in favor of the solar model predictions and the agreement between theoretical predictions and helioseismic data are both encouraging signs. The combination of better observations and theory has opened the prospect of a new era of precision stellar astrophysics, which could have broad consequences for diverse subfields of astronomy. Stellar atmospheres theory has traditionally employed a series of approximations when deriving abundances. Classical models assume an ad hoc turbulent velocity field adjusted to yield abundances independent of excitation potential and line strength. Convection is usually treated in an approximate fashion, with the mixing length theory. Horizontal temperature fluctuations (granulation) are not included. The models also typically assume that the molecular and atomic levels are described by local thermodynamic equilibrium (LTE), e.g. by the local temperature alone. The compilations of solar abundances used for theoretical solar models [@ag1989; @gn1993; @gs1998] employed model atmospheres with approximations at the level described above. The mean thermal structure employed was semi-empirical [@hm1974 hereafter HM74]. When these approximations are relaxed, different conclusions about the abundances are obtained. Departures from LTE are expected at a modest level for solar conditions, and have been investigated by a number of authors [for example @c1986; @sh1990; @k1993; @stb2001; @w2001]. Numerical simulations of convection have matured to the level where they can be used to predict velocity fields, temperature fluctuations, and changes in the mean thermal structure of the upper atmosphere [@sn1998]. Abundances derived from these simulations yield a very different pattern, which has been developed in a series of papers [@pla2001; @pla2002; @asplund2000a; @asplund2000b; @agsak2004; @agsak2005]; papers by Lodders (2003) and @ags2005 summarize the revised abundance scale. The net effect is in the sense of systematically lower metal abundances. The downward revisions for the heavier elements (e.g. Fe and Si) are small, while the claimed reduction in the abundances of lighter species (especially CNO) is more dramatic. Models employing different treatments of granulation and non-LTE corrections [@h2001; @sh2002] predict smaller abundance reductions. The central temperature predicted by interiors models is sensitive to the abundances of the heavier elements, but not the lighter ones. As a result, the new abundance scale does not disturb the agreement between interiors models and observational data for purposes such as the mass-luminosity relationship and solar neutrino fluxes. However, the inferred solar sound speed profile, and the radii of interiors models, is sensitive to the bulk metallicity. Serious problems have emerged when comparing interiors models with the revised abundance scale. These discrepancies are evidence for problems in our understanding of stellar interiors, stellar atmospheres, or both. In Paper I [@dp2006] we investigated the errors in solar abundances predicted by the combination of stellar interiors models and helioseismic data. In this paper we examine the uncertainties in the abundance predictions from stellar atmospheres theory. We begin with a brief summary of the results from Paper I, and follow with a discussion of the motivation and main results from the revised stellar atmospheres models. In Section 2, we perform a critical analysis of the precision of the solar CNO abundances and discuss the implications for Be. We demonstrate in that section that the errors in the abundances are larger than previously estimated, and that there is evidence that the “best” current solar CNO abundances are intermediate between the new and old scales, with errors permitting both. We discuss the implications of our finding and future tests in Section 3. In particular, we argue that inconsistencies between the solar thermal structure and that predicted by the simulations would favor a higher abundance scale closer to the seismic value and discuss uncertainties in the numerical convection simulations. Constraints from Helioseismology -------------------------------- Helioseismology provides two powerful constraints on the solar composition: diagnostics of the internal solar temperature gradient and diagnostics of the equation of state. Inversions of the observed solar pulsation frequencies yield accurate measures of the sound speed as a function of depth. In turn, the gradient in the sound speed can be directly tied to the temperature gradient. Since the temperature gradient is related to the opacity, and thus the composition, information on the solar abundances is encoded in the seismic data for the radiative interior. One can even obtain meaningful constraints on the solar age from the helium abundance profile deduced in the deep interior. In addition to the vector information on the sound speed profile, there are also precise scalar quantities that can be extracted. The thermal structure at the base of the solar convection zone is nearly adiabatic, while the temperature gradient in the interior is radiative. As discussed in Paper I, Section 2.2 the resulting discontinuity in $\nabla$ generates a distinct signal that can be used to precisely localize the base of the solar convection zone [$R_{cz}= 0.7133 \pm 0.0005 R_{sun}$ @ba2004]. Seismology also sets strict limits on convective overshooting [$< 0.05 H_P$, @cdmt1995]. The depth of the solar convection zone is sensitive to the light metal abundances in the Sun but insensitive to most of the other uncertainties in solar interiors models (see Paper I for a detailed error budget). Ionization also induces a depression in the adiabatic temperature gradient, and the absolute abundances of the species in question can be inferred from the magnitude of the perturbation in the surface convection zone. An extremely precise surface helium abundance can be deduced from this effect ($Y_{surf}=0.2483 \pm 0.0046$, see Paper I, Section 2.3 for the sources used in this estimate). More recently, @ab2006 have demonstrated that the ionization signal of metals in the convection zone can be detected in the seismic data, leading to a bulk metallicity $Z=0.017 \pm 0.002$. Because the majority of the solar metals are in the form of CNO, this is primarily a constraint on their abundance. In principle one might be able to use this technique to solve for individual heavy element abundances by fitting the strength of distinct ionization stages. However, it is not yet clear that there is sufficient spatial resolution in the seismic data to permit such a detailed analysis. In Paper I we demonstrated that the combination of the surface convection zone depth and surface helium abundance constraints was a powerful diagnostic of the solar heavy element abundances. The surface helium abundance is tied to the initial solar helium abundance with a correction for gravitational settling. The initial helium is sensitive to the central opacity and the abundances of the heavier metals (especially iron). The convection zone depth is sensitive to the opacity at temperatures   2 million K, where bound-free opacity from light metals (CNONe) is an important contributor. The most significant new finding in Paper I was that the combination of the two scalar constraints could be used to rule out some abundance combinations with high statistical significance. The detailed sound speed profile adds additional information; we found that models consistent with the scalar constraints could be constructed with low oxygen and very high neon, but such models exhibited substantial sound speed deviations relative to solar data in the deep interior. The inferred solar oxygen and iron abundances ($[O/H]=8.86 \pm 0.05, [Fe/H]= 7.50 \pm 0.05$) are consistent with the @gs1998 absolute abundances, but strongly inconsistent with the new abundance scale within its quoted errors [@l2003; @ags2005]. Although there are potentially positive chemical evolution consequences for the revised abundance scale [@turck2004], it is not easy to generate interiors models that are consistent with both seismology and the low abundance scale. The most commonly cited possible explanations on the interiors side (high neon, enhanced gravitational settling, and errors in the high temperature radiative opacities) are all strongly disfavored. As previously mentioned, solutions with high neon degrade agreement with the sound speed profile, and are also problematic from a stellar atmospheres perspective [@s2005]. An increase in the degree of gravitational settling increases the convection zone depth but decreases surface helium, trading improved agreement with one diagnostic for worse agreement in another. Enhanced differential settling of metals with respect to helium is inconsistent with the underlying physics and would have to be extreme [@gwc2005]. Three independent quantum mechanical calculations yield extremely similar Rosseland mean opacities at the temperatures of interest for the base of the solar convection zone. As discussed in Paper I, both the atomic physics and equation of state are relatively simple in this regime, and the concordance between different calculations is thus not a surprise. Physical processes neglected in classical stellar models (such as rotational mixing and radiative acceleration) can be independently constrained by other data, and in any case they would tend to induce higher rather than lower surface abundances. The scalar constraints are insensitive to the other theoretical ingredients in standard solar models (e.g. convection theory, surface boundary conditions, low temperature opacities, equation of state, and nuclear reaction rates). The considerations above indicate that it is extremely challenging to reconcile a low solar metal abundance with current stellar interiors models and seismic data. This does not imply that a metal-poor Sun is impossible, but it certainly motivates an investigation of the uncertainty in the atmospheres models used to derive the abundances. Model Atmosphere Ingredients ---------------------------- The new solar abundance estimates are derived from a variety of changes in the model atmospheres. Changes in oscillator strengths and equivalent widths of spectral lines contribute for some diagnostics, and as discussed below we largely concur with the revised values. The magnitude of non-LTE corrections depends on the atomic model and the relative importance of photo-excitation and collisions on the level populations in the model atmosphere. The comprehensive re-examinations of the solar oxygen [@agsak2004 hereafter AGSAK04] and carbon [@agsab2005 hereafter AGSAB05] adopted a particular model for NLTE corrections, and we assess its accuracy and uncertainties by comparison with limb darkening data and other published calculations. In @pla2001 [@pla2002] the abundances derived from forbidden lines have been reduced by the application of blending corrections; AGSAK04 and AGSAB05 used the revised equivalent widths for C and O abundance studies. Both the uncertainty in these corrections and their central value is incorporated in our error analysis. Three other coupled changes in the atmosphere models are the treatment of convective velocity fields (“macro/microturbulence”), horizontal temperature fluctuations (granulation), and the impact of convective overshooting on the mean thermal stratification. All of these features are derived from numerical convection simulations; a good discussion can be found in @sn1998. Their combined impact can be deduced by the comparison of the results in the 3D case in the published studies with the results from the semi-empirical 1D HM74 thermal structure and the theoretical 1D MARCS models. Since the initial 3D model atmosphere is derived with physics similar to that in the MARCS code, the impact of the convection treatment can be indirectly inferred by comparing MARCS and 3D abundances. A comparison of HM74 with MARCS and 3D is a measure of the impact of different choices of the thermal structure. The numerical convection simulations predict line profiles in excellent agreement with the data for iron and silicon lines [@asplund2000a; @asplund2000b]. There are some trends with excitation potential that may be related to NLTE corrections [@stb2001] and issues with the initial generation of simulations when compared with line profiles in the outer solar photosphere [@scott2006]. The concordance of the predicted amplitude of horizontal temperature fluctuations with the solar granulation pattern is encouraging [@sn1998; @asplund2000a]. The validity of the steep solar temperature gradient in the simulations is not as clear; there is an apparent conflict between the simulations and the mean thermal structure of the solar atmosphere, as well as the degree of convective penetration for the upper atmospheric layers [@ay2006]. Since these effects are all tied together in the abundance studies, we focus on the agreement between different diagnostics of abundance as a valuable test of the precision of the results obtained from different atmospheric models. The Solar CNOBe Abundances ========================== Revised solar abundances have been derived for a number of species. In this paper we focus on CNO for several reasons. First, these are the elements where the difference in abundance is largest, and they are also the cases where there is the largest variety of distinct abundance indicators. The difference between interiors and atmospheres based abundances for heavier species, such as Fe and Si, are not statistically significant. Furthermore, details of the new abundance estimates and internally consistent comparison with prior work are published for only some of the heavier elements. The solar Be abundance is an important diagnostic of mixing, and the photospheric abundance is linked to the solar O (Balachandran & Bell 1998). We therefore also briefly discuss the implications of our result for Be. We begin with a description of our overall approach, and then follow with individual sections on oxygen, carbon, and nitrogen. Our primary references for O, C, N are respectively AGSAK04, AGSAB05, and @ags2005. We include other studies for external comparisons, and discuss newer results based on other models and diagnostics when available. Overall Approach ---------------- Our primary metric for the accuracy of the abundances derived from the different model atmospheres is the consistency of the estimates derived from distinct classes of indicators. One might justifiably apply a different standard, noting for example the greater degree of sophistication in the input physics for the hydrodynamic simulations. However, it is not clear that the ingredients that induce the abundance changes (such as changes in the temperature gradient) are actually a necessary consequence of these improvements in the model, as the absolute errors in the first-principles theoretical models have not been quantified. We proceed in two steps. In the first, we compute relative abundances and errors within a given assumed model atmosphere for the atomic and molecular features. Since the abundances from atomic lines have smaller model-to-model differences (and thus smaller systematic errors), we adopt the atomic abundances for our base estimate. The difference between atomic and molecular abundances is then used to infer which of the different atomic scales should be adopted for the central value, and the uncertainty in the differential scales is used as a measure of the systematic error arising from the choice of model atmospheres. For the atomic line diagnostics, we include random errors from the dispersion of results from single permitted lines about the adopted mean. For the forbidden lines, uncertainties in oscillator strengths and blending corrections become the dominant random error source. We also include systematic errors (NLTE corrections and zero-point shifts in the average oscillator strength) by comparing the study values with external constraints and other published calculations. This is important when comparing the atomic and molecular indicators, since NLTE corrections are usually included in the former but not the latter. As a result, changes in the degree of NLTE corrections have a direct impact on goodness of fit. When available, we adopt a weighted mean of the permitted and forbidden atomic indicators and the error in the mean when comparing with the molecular data. For the molecular indicators, we include all measured lines of a given diagnostic and use the total dispersion (rather than the error in the mean) as a measure of goodness of fit. We adopt a weighted mean of the various indicators for an average molecular abundance. However, the error in the mean will understate the true uncertainty; it is frequently the case that the mean values for different molecular diagnostics differ by more than the dispersion within each indicator. We therefore compute the dispersion of each indicator about the adopted mean and average these values to obtain an error in the molecular abundances. One could alternately compute the dispersion in the molecular abundances and treat this as a measure of the systematic uncertainty, adding it to the error in the mean in quadrature; this procedure yields somewhat smaller errors. Although the latter approach may be practical when there are numerous molecular probes available, we prefer the former method for situations like oxygen (where there are 2 values, and thus an unreliable estimate of the uncertainty in the mean). We derive final abundance estimates for each species by comparing the mean abundances derived within each class of models for atomic and molecular indicators. In the case of oxygen, the 3D and HM74 models exhibit comparable differences with opposite sign, while the MARCS models have an internally consistent intermediate abundance. We therefore adopt a mean of the different derived oxygen abundances and an uncertainty from the scatter. In the case of nitrogen, the HM74 abundances are preferred. The case of carbon depends on the origin of systematic differences in abundances inferred from atomic features. If the low zero point of AGSAB05 is adopted, the 3D models are the favored solution. If the higher zero point of previous work is adopted, the situation is similar to that for oxygen. Oxygen Abundance Indicators --------------------------- AGSAK04 derived a low solar oxygen abundance ($8.66 \pm 0.05$) from four distinct indicators: atomic lines (forbidden and permitted) and two different classes of infrared molecular lines ((v,r) and (r,r)). All four indicators had formerly been used to obtain higher absolute oxygen abundance (8.83 to 8.90). Abundance estimates from all indicators are reduced in the theoretical atmospheres that include substantial overshooting because lines become stronger for a fixed abundance in the presence of a steeper temperature gradient. Molecular abundances are reduced more than atomic ones because the cooler atmospheric structure of the 3D hydro models changes the chemical equilibrium. AGSAK04 included other effects that reduced the abundances derived from atomic features without impacting the molecular indicators: a combination of changes in oscillator strengths, the inclusion of blending features, and their claim of large non-LTE effects for the permitted lines. In this section we discuss the uncertainties in each of these cases. We advocate a substantial decrease in the magnitude of the NLTE corrections for the \[O/H\] derived from the permitted atomic oxygen lines, and an increased error in both the abundance derived from the forbidden line (from uncertainties in the oscillator strength) and the permitted lines (from uncertainties in the NLTE corrections). We also derive an increased error from the \[O/H\] derived from the IR molecular lines from the internal scatter and trends with excitation potential, and argue that the correspondence between the trends in the MARCS and 3D models is evidence for errors in the common underlying model. Our basic results are summarized in Table 1. In Table 1, the upper part of the table repeats the mean abundances and errors for the 4 different indicators presented in AGSAK04. We present our revised estimates for the same three cases in the lower part. The last three rows in each sub-table give the mean discrepancy between atomic and molecular indicators. At the end of the section we synthesize this information to obtain our best estimate for the solar oxygen abundance. In each of the following subsections, the error estimates derived are internal ones. The differences between the results from the three classes of models are prima facia evidence that systematic errors are important. The systematic errors are discussed in the final subsection. ### Forbidden Oxygen Lines The forbidden oxygen line at 6300.3 Å  has traditionally yielded high oxygen abundances. @pla2001 argued that a nearby blended Ni line contributed significantly to the oxygen feature. They treated the continuum level, log (gfNi), and the oxygen abundance as free parameters, but assumed that the line profiles as given from the simulations were exact. The inclusion of the Ni feature induced a direct reduction of 0.13 dex in the inferred oxygen abundance. In addition, the usage of a 3D model atmosphere structure led to a further reduction of 0.08 dex in the oxygen abundance to an estimated \[O/H\] = 8.69 +/- 0.05. In AGSAK04 another forbidden line at 6363.7 Å  was considered as a second indicator. The authors reduced the equivalent width by 0.5 mÅ  for an estimated contribution from a blended CN feature to obtain an oxygen equivalent width of 1.4 mÅ , also implying a low abundance. We begin our analysis by noting that abundances derived from blended features are usually treated with great caution. The most conservative procedure is to ignore the blending feature and treat the derived abundance as an upper limit. When this is done for the two forbidden lines, the maximum abundance obtained for 3D, HM, and MARCS are (8.82, 8.8), (8.9, 8.88), (8.86, 8.84) respectively. In the section that follows, we include the reduction in abundances from estimates of the blending contribution. We adopt the AGSAK04 values for \[O/H\] derived from the 6300.3 Å  line, subject to the caution on the strength of the Ni blending feature below. For the 6363.7 Å  line, @m2004 argued that the (10,5) Q$_2$ 25.5 CN line is unblended in the solar spectrum and has the same oscillator strength as the feature blended with the forbidden line. He derived a smaller correction for the blended CN line (0.35 mÅ  rather than the 0.5 mÅ  value used in AGSAK04), which we adopt here. This leads to a modest 0.04 dex increase in the derived abundance from that line, which we also treat as an uncertainty in the \[O/H\] derived from this feature from the uncertainty in the contribution of CN to the blended feature. The error analysis for a blended feature is more complex than the one that can be employed for an isolated line. The derived \[O/H\] is sensitive to the continuum level, and an error component for this should be included; @pla2001 estimate this uncertainty at 0.02 dex, which we adopt for both lines. AGSAK04 adopted lower values for the oscillator strengths than those found in the NIST database, but their choice is well-supported by the improved atomic physics [see @sz2000] . However, the errors in individual theoretical log(gf) values are higher than those assigned in @pla2001 and subsequent papers; we adopt 0.04 dex for individual lines. @pla2001 also estimated that uncertainties in the underlying equation of state induces errors of 0.02 dex. Especially for the 6300 Å  line, the results depend heavily on the detailed line profiles, particularly in cases where the individual components cannot be directly disentangled. @pla2001 estimated uncertainties of 0.02 dex from the central wavelength of the Ni feature and a 0.04 dex uncertainty from the central wavelength of the \[O I\] line. We treat these errors as representative of the uncertainties in the line profiles, and adopt them for both forbidden lines. The treatment of the Ni line in the main forbidden line is more problematic. In the initial study, the oscillator strength was highly uncertain. @pla2001 treated log (gf) for Ni as a free parameter. The continuum level, $log (gf_{Ni})$, and \[O/H\] were treated as free parameters and the combination that produced the minimum $\chi^2$ was adopted. However, @j2003 have measured the oscillator strength of the Ni feature (log gf = -2.11), and the Ni abundance of the solar mixture is well-constrained by knowledge of the solar Si/Fe and the relative meteoritic abundances. With the new gf value and \[Ni/H\]=6.25 the blending feature would be 0.23 dex stronger than the best $\chi^2$ value obtained in the 2001 paper. AGSAK04 did not report the inferred strength of the blending feature in their fit, but the similarity in the absolute abundance suggests that it would be comparable. We believe that it is no longer appropriate to treat this as a free parameter, and the same method should be used as is done for other blended features: namely, the strength of the Ni line should be held fixed (and varied within its uncertainty) while the free parameters are the oxygen abundance and continuum level. The direct effect of increasing a blending contribution is usually to decrease the abundance, but the present case is more complicated. For the quoted values in the AGSAK04 fit, the Ni line contributes 25% of the total equivalent width of the line. An increase of 0.23 dex in the strength of the feature would imply a total Ni contribution of 43% of the blended equivalent width at a fixed continuum level; such a combination would be a poor fit to the line shape and would yield a reduction in the inferred oxygen of 0.12 dex. An increase in the continuum level would be required to restore the agreement with the line profile, which would in turn lead to an increased total equivalent width. The net impact on the derived oxygen is not obvious, and not necessarily in the negative direction. @r1998 invoked log gf=-1.95 for the Ni feature and obtained \[O/H\]=8.75 for the forbidden line when his oscillator strength for the forbidden line was adjusted to the same value as that employed by Allende Prieto et al. In the absence of other information, we assign an additional error component of 0.04 dex for the strength of the Ni feature, slightly higher than the value advocated by @m2004. Adding the errors in quadrature, we obtain an uncertainty of 0.078 per line (0.055 in the average) and abundances derived from the forbidden lines systematically 0.02 dex higher than those found in AGSAK04. ### Permitted Atomic Oxygen Lines The permitted atomic oxygen lines have relatively high oscillator strengths, but very high excitation potentials. The most commonly used features are the OI triplet at 7771.8, 7774.2, 7775.4 Å , with an excitation potential of 9.15 eV. AGSAK04 also considered three other atomic features (6158.1 Å , 8446.7 Å , and 9266 Å ). The primary reason for the low oxygen abundance inferred by AGSAK04 from the triplet is a large non-LTE correction. Because these lines arise from such a high energy state, non-LTE effects must be included. However, the quoted values of the non-LTE corrections in the literature vary drastically. For the triplet, the average is -0.06 dex for @h2001, -0.22 to -0.28 for AGSAK04, and -0.16 dex for @pafb2004. These variations can be partially traced to different assumptions about the importance of collisional excitation (as opposed to photoionization), but larger differences at the 0.10 dex level remain even for cases that make similar assumptions about hydrogen collisions. AGSAK04 neglected hydrogen collisions in their estimate of the non-LTE effects. They justified this by noting that for some well-studied lines, the classical @d1968 formulism overestimates the collision rate. However, an inspection of their Figure 6 indicates that the neglect of collisional effects in their model yields changes as a function of limb darkening that differ from the observed solar values. This impression is confirmed by the more detailed study of @pafb2004, who found that models including hydrogen collisions ( their $S_H=1$ case) were a better fit to the solar data. We therefore conclude that the non-LTE corrections in AGSAK04 are overestimated, which has a significant effect on the concordance of the different oxygen indicators. AGSAK04 applied larger downward reductions to the HM model than to the 3D and MARCS models, which we do not believe to be justified. @pafb2004 indicate that similar corrections are obtained for Kurucz and 3D models in their detailed study of the triplet as a function of limb darkening. This is particularly important because the discordance between the oxygen derived for the 1D models from atomic and molecular lines was used as a primary argument for the superiority of the 3D models, and this discrepancy can be directly traced to the assignment of very large NLTE corrections to the HM model. By very similar logic, the internal dispersion in abundance for the atomic lines in the 1D case arises from the assignment of large NLTE corrections to some of the lines; the internal agreement of the HM case is improved (and that of the 3D and MARCS cases degraded) with smaller NLTE effects. We illustrate this point in Figure 1. For the range of NLTE corrections that we consider reasonable (shaded band) the internal dispersion for the 3D models exceeds that of either 1D model. This emphasizes the role of substantial NLTE corrections to the scientific conclusions of AGSAK04. In order to quantify this effect, we normalized the non-LTE corrections in Table 3 of AGSAK04 to an average for the triplet of three values: 0.16 dex (the best fit from @pafb2004 and 0.11 dex (the mean between Holweger 2001 and @pafb2004), and 0.06 dex [@h2001]. We adopt the 0.11 dex level as our best case for reasons outlines below. The NLTE corrections for the other lines were scaled by the linear ratio of the average NLTE corrections for the triplet and the target values. We present revised abundance estimates for the permitted lines in AGSAK04 computed in this manner in Table 2. In Table 2, the first set of values contains the LTE results. The next three sets represent calculations where the NLTE corrections were normalized to obtain a mean triplet correction of 0.06 dex, 0.11 dex (our adopted mean), and 0.16 dex. The final set of results includes the NLTE corrections originally applied in AGSAK04. This procedure yields non-LTE \[O/H\] abundances of (8.69, 8.73, and 8.76) for 3D models, 1D HM74, and 1D MARCS respectively in the 0.16 dex case and (8.73, 8.76, 8.80) in the 0.11 dex case. Since the NLTE corrections are significant for the triplet, the uncertainty in these corrections is a major ingredient in the error budget. Even the reduced NLTE corrections of Allende Prieto et al. (2004) for the triplet are substantially larger than the corrections used by Holweger (2001), who found an average NLTE correction of -0.06 dex, 0.10 dex lower than the value reported by AGSAK04. Surprisingly, none of the authors involved commented on the origin of the difference. Holweger (2001) obtained average LTE and NLTE triplet abundances of 8.78 and 8.72 respectively; his NLTE abundance is close to that obtained for the two 1D models in AGSAK04. From @pafb2004, the case with no hydrogen collisions was ruled out at the $3 \sigma$ level, and LTE models were ruled out with high confidence. However, the authors did not consider whether even lower NLTE corrections than their $S_H = 1$ case would have provided improved fits to the data. In cases such as this, we see no justification for simply adopting one NLTE correction (0.16 dex) over another published value (0.06 dex), and adopt the average of the two (0.11 dex). We note that our central values are close to what we would infer if we simply took the triplet alone as an oxygen abundance indicator and assigned a 0.16 dex NLTE correction. We used the dispersion in the abundances derived from individual lines as a base random error. A change of 0.05 dex in the triplet NLTE correction yields an average change in \[O/H\] of 0.035 dex, which we include as an additional systematic error. Finally, the log (gf) values from AGSAK04 are lower than previously published values by @bhgvf1991; we therefore add another 0.025 dex systematic error, following a similar error analysis by @m2004. The net effect is a total error estimate (summarized in Table 1) of 0.064 to 0.074 dex. ### Infrared Molecular Lines The IR molecular oxygen lines have been the primary abundance indicator used in previous compilations of solar abundances [e.g. @gs1998], and 1D model atmospheres yield relatively high absolute oxygen abundances. The absolute abundances are a strong function of the thermal structure of the model atmospheres, and the different thermal structure of the 3D model of AGSAK04 yields much lower predicted abundances than the 1D models. We note that @h2001 discounted CNO abundances derived from molecular lines because of their high temperature sensitivity. AGSAK04 considered two sets of OH lines: (v,r) and (r,r). The abundances predicted as a function of excitation potential from the 3D hydro simulations are compared with those from the 1D Holweger-Mueller and MARCS codes in Figures 2 and 3. On these figures we have also indicated the average atomic abundances for each class of models. We note the presence of striking trends with excitation potential in the 3D and MARCS models for the (r,r) lines. The correspondence between the MARCS and 3D trends indicates that the origin of these features is common to both models, which indicates that it is a feature of the base atmospheric treatment rather than being induced by the convection simulation. In the context of 1D models, trends such as those seen in the 3D models would be interpreted as a problem with the thermal structure or the assumed microturbulence. AGSAK04 noted this trend, and claimed that it could be removed by invoking an outer atmosphere structure even cooler than the one predicted by the simulations. Figure 1 suggests an alternate explanation, namely that the thermal structure in the outer layers is closer to the hotter semi-empirical HM74 model. We will return to this point when we consider more recent work on CO abundances in the outer solar atmosphere. AGSAK04 discarded the (r,r) data for the weaker and stronger lines, in effect deriving an abundance from the valley in Figures 2 and 3. There is no better justification for discarding the high than the low points in this figure; such a procedure is not required for the 1D models. We therefore derived average abundances using all of the features for all three models and both indicators; the standard deviation about the mean is an indicator of the quality of the fit for the individual bands. Our results are summarized in Table 1. The average molecular abundances were obtained with a weighted mean, but a simple averaging of the errors underestimates the dispersion. We therefore computed $\sigma$ for each band around the weighted molecular mean in Table 1 and averaged the (v,r) and (r,r) values to obtain the total error in the molecular abundances presented there. @m2004 considered a third molecular oxygen abundance indicator, and derived relative abundance patterns comparable to what AGSAK04 found. We do not include this in Table 1 because it is not clear that systematic errors between model atmospheres codes can be properly accounted for in a differential analysis. If we had included the @m2004 the mean molecular abundances would have been minimally altered for 3D and HM. The average would be reduced for MARCS and the internal error in the MARCS \[O/H\] would be dramatically increased. This provides further evidence that there is an underlying issue in the thermal structure of the MARCS model. Melendez also computed abundances with the same indicators as AGSAK04 for a Kurucz model atmosphere, and derived similar abundances as would be found for the HM74 model. @ay2006 present evidence for a high solar oxygen derived from CO studies; we postpone a discussion of this interesting result to our conclusion, in the context of tests of the solar thermal structure. ### Oxygen Abundance and Error Analysis Our overall result from the reanalysis of the AGSAK04 oxygen indicators is that the abundances derived from the atomic indicators are systematically increased for all models. In the original paper, the 3D abundance estimators were found to yield consistent abundances, while the 1D abundances from different methods were highly discordant. This conclusion no longer holds when the reduced NLTE corrections inferred from limb darkening studies are employed. In fact, one would obtain very similar conclusions to those presented in Table 1 from the triplet abundance of 8.72 presented in Allende Prieto et al. (2004) for the 3D model (e.g. the 0.16 dex case presented in Table 2). Rather than simply adopting one model or another as correct, we interpret the difference between the HM74 and 3D abundances as evidence that the thermal structure of the Sun is intermediate between the two. The difference between the atomic and molecular abundances is roughly equal in magnitude and opposite in sign between these models; the MARCS model yields an intermediate abundance where the two classes of indicators give the same abundance, but with a larger error. The error in (Atomic-Molecular) is substantial; these differences are formally significant only at slightly more than $1 \sigma$. We therefore argue that the mean of the derived atomic abundances (8.75) is a reasonable estimator of what one would obtain from a model with a thermal structure capable of reproducing the atomic and molecular data; one would obtain 8.74 from comparing HM and 3D, and 8.76 from MARCS alone (which is already internally consistent). We have a random error of 0.05 dex for the mean atomic abundance, but this is insufficient for a total error because of the presence of strong systematic differences. Adopting the consistency between atomic and molecular indicators as a measure of goodness of fit, $~1 \sigma$ deviations could make either the 3D model (\[O/H\]=8.68) or the HM model (8.80) consistent. We treat this as a $1 \sigma = 0.06 $ systematic error, and note that it is comparable to the zero-point shift that we obtain for the atomic indicators relative to AGSAK04 estimated below. We can also examine systematic errors by comparing the AGSAK04 values with abundance estimates by other authors. These are most easily analyzed by comparing LTE abundances for the triplet. Holweger (2001) derived an average LTE triplet abundance of 8.78 for his standard model and 8.85 for the alternate VAL model. @bhgvf1991 reported 8.84 for the triplet for a HM model and 8.78 for a MACKKL atmosphere. These should be compared with 8.89, 8.87, and 8.93 for the three LTE cases in AGSAK04. We note that the 1D cases in AGSAK04 used the equivalent widths obtained with the line broadening of the 3D hydro models, so there will be differences between their results and those obtained with other 1D codes. The average LTE abundance is 8.85, with $\sigma = 0.055 dex$. Systematic differences at the 0.06 dex level are thus a reasonable estimate of the current state of the art for oxygen abundances when estimated with different techniques. Adding systematic (0.06) and random (0.05) errors in quadrature, we obtain 8.75 +/- 0.08 as our final oxygen abundance estimate for the Sun. This is less than $1.6 \sigma$ below the helioseismic abundance, and therefore we conclude that the existence of a solar oxygen problem has not been demonstrated with high statistical significance. Carbon Abundance Indicators --------------------------- The overall story for carbon follows a similar path to the changes in the inferred oxygen abundance, and the comprehensive reanalysis of AGSAB05 for carbon has a similar logical structure to the 2004 oxygen paper. Although both the carbon and oxygen are reduced, the C/O ratio is preserved. A cool outer solar atmosphere in the 3D models yields substantially reduced abundances from molecular indicators, while blending features and non-LTE effects reduce the carbon abundance inferred from atomic features. What distinguishes the carbon from the oxygen case is that the NLTE effects are smaller, and as a result one might anticipate a smaller offset in the atomic line abundances than the change in atomic oxygen abundance indicators. However, this is not the actual published result; if anything, the derived carbon from atomic features is lower than the molecular value for all of the models presented in AGSAB05. Furthermore, we will demonstrate that this effect cannot be explained by any of the effects used to explain the differences in the comparison with prior work given by AGSAB05. Until the origin of this difference is understood, we therefore have to consider two different systematic sets of abundances for atomic features, and the best choice of model hinges on which set is correct. In this section, we discuss the three classes of indicators (forbidden and allowed atomic lines, and molecular) in turn, and as for the oxygen abundance synthesize our final best estimate and error in the fourth subsection. Our overall estimates are presented in Table 3. The top set of values represents the original AGSAB05 values for the different indicators. The middle set is what we would obtain with the low AGSAB05 normalization of the atomic abundances, while the bottom set is what we obtain with the higher Biemont/Holweger normalization. ### Forbidden Lines The \[CI\] line at 8727 Å  was the subject of a detailed analysis by @pla2002. They incorporated blending from a nearby Si feature to reduce the equivalent width attributable to carbon from 6.5 to 5.3 mÅ , with a corresponding reduction in the inferred abundance. They also employ a lower oscillator strength than previous studies. Unlike the case of oxygen, the carbon abundance derived from the forbidden line is almost as temperature sensitive as that derived from molecular features, so the atomic versus molecular diagnostic is less powerful for carbon than for oxygen. We adopt the AGSAB05 central values for our base case, but note that there may be explained systematics in the atomic carbon abundances in AGSAB05 which we discuss below. We include their error estimate for uncertainties in the equation of state (0.02 dex), but assign a larger uncertainty to the atomic physics (0.04 dex) in accord with the quoted theoretical uncertainties. Our principle reservation on the error budget is the uncertainty in the continuum level and the contribution to the equivalent width of the blend from the wing of the Si feature. Their reduced $\chi^2$ permits only small deviations (of order 0.01 dex) in the derived carbon abundance, but the base model relies upon the assumption that the underlying velocity field is exact. Although the overall agreement with Fe [@asplund2000a] and Si [@a2000] line profiles is good, it is not errorless. We cannot evaluate this ingredient directly, but an estimate based upon the mean deviation observed in clean lines would seem to be a worthwhile exercise. For the present, we therefore assign the same blending uncertainty of 0.03 dex adopted by @pla2002 to obtain a total uncertainty of 0.054 dex. ### Permitted Atomic Lines AGSAB05 considered a subset of the permitted atomic features used in previous solar abundance studies [e.g. @bhgv1993; @sh1990]. @sh1990 found that small NLTE corrections are required for CI, and that the strength of the correction depends on equivalent width. They found an average of -0.05 dex; if restricted to the weaker lines included in AGSAB05, their average NLTE correction would be -0.02 dex. AGSAB05 computed Non-LTE corrections for 1D models, and the MARCS corrections were applied to the 3D models. Hydrogen collisions were not included in the NLTE corrections; this resulted in larger downward abundance revisions (an average of -0.08 to -0.09) than @sh1990. AGSAB05 note that the case of carbon should be an analog of oxygen, and we concur. As a result, we contend that the case with hydrogen collisions should be included in the base model. Both sources indicate that including hydrogen collisions roughly halves the expected NLTE correction. We therefore considered two cases for NLTE corrections: a maximum of half the AGSAB05 value (corresponding to their hydrogen collision case, average -0.04 dex) and a minimum of one quarter of the AGSAB05 value (corresponding to the SH90 case, average of -0.02 dex). Our best value is the average between the two (a mean of -0.03 dex), and the error induced by uncertainties in NLTE corrections is 0.01 dex; adopting the AGSAB05 hydrogen collision case would only have changed our mean value by 0.01 dex. We applied these proportional NLTE corrections to the AGSAB05 LTE results for their three classes of models (middle values, Table 3). There is a small reduction in the dispersion (and mean trend with equivalent width) for the 1D models and a corresponding increased for both in the 3D model; none of these features, however, are drastic. The internal dispersion in the permitted atomic abundances is of order 0.03 dex. A more substantial issue emerges when we compare the AGSAB05 abundances with prior work, and this is true even for the LTE estimates. The mean LTE abundance for the @bhgv1993 sample is 8.56 for the lines in common with AGSAB05; this should be compared with a HM LTE value of 8.48 for the latter compilation. This offset of 0.08 dex is comparable to the average difference between atomic and molecular abundance indicators. The mean difference in equivalent width and oscillator strength for the lines in common is negligible, and would yield an offset of less than 0.01 dex if applied under the assumption that all of the lines are on the linear part of the curve of growth. We illustrate the differences in Figure 4, defined in the sense (Biemont- AGSAB05). In this figure we have corrected the Biemont abundances to the AGSAB05 equivalent width and oscillator strengths. The differences are significant even for weak lines, suggesting that differences in the classical line broadening are probably not responsible. A similar, but smaller, effect is present in the forbidden line. @pla2002 inferred a HM abundance of 8.48, which would also be obtained from @sh1990 when a blending correction is made to the equivalent width. AGSAB05 could not trace a comparable difference (0.06 dex) relative to the earlier work of Lambert. The only obvious source that we can derive is a note by Sturenburg & Holweger that they corrected their atomic abundances for the fraction of C tied up in CO, which could be of the right order to explain the differences. Until the origin of this discrepancy (which is not present for oxygen) is explained, we have to treat this as a systematic uncertainty in the atomic abundance scale. Abundances derived under this scale are the last set of values in Table 3. ### Molecular Lines We consider the same four molecular indicators that were included in AGSAB05. They chose to disregard one of them (CH electronic lines) in their derived mean abundances, on the grounds that they are located in a crowded portion of the spectrum and sensitive to the treatment of line broadening. However, the formal errors in the CH electronic abundances are similar to those for the other molecular species, and as such we see no obvious reason to exclude them. We do treat the CH (v,r) abundances as being more reliable, as they are based on many more lines than the other diagnostics. We therefore assigned double weight to the CH values and single weight to both the C2 electronic and CH electronic values. As for oxygen, the mean was derived by a weighted average of the carbon obtained with different molecular indicators, and the scatter of the individual line measurements for all diagnostics around the adopted mean was taken as a measure of the random error. We did not include carbon (or oxygen) abundances derived from CO line studies, because there are complex correlated errors. Had we included them, the net effect would have been to increase the molecular abundances relative to the atomic values. ### Carbon Abundance and Error Analysis Our final inference concerning carbon depends on which atomic abundance scale is adopted. If we take the low scale of AGSAB05, the 3D model atomic and molecular abundances (8.40, 8.42) are closer than those for HM74 (8.45, 8.55); 1D MARCS abundances are also consistent (8.40, 8.44). We would estimate a mean value of 8.41-8.42, with a random error of 0.04 dex. The HM74 average of 8.50 would be an effective $2 \sigma$ internal inconsistency, implying a 0.04 dex systematic uncertainty for a total abundance of 8.41 +/- 0.06. Adopting the higher scale would give pairwise results of (8.44, 8.42), (8.50, 8.55), (8.44, 8.44); all three models are internally consistent within the errors, and a mean abundance would be 8.47 with a total error of 0.05 (0.04 random, 0.03 systematic). We adopt the mean of these approaches (8.44), and estimate an error of 0.04 (random) and 0.04 (systematic) for a total of 0.06 dex when combined in quadrature. Nitrogen Abundance Indicators ----------------------------- Our discussion of nitrogen is necessarily briefer than that of oxygen and carbon, largely because the published results are preliminary and incomplete. Holweger (2001) derived a non-LTE \[N/H\] = 8.0 +/- 0.11, comparable to results in previous compilations of solar abundances from @gs1998. The compilation of models in @ags2005 yields atomic and molecular nitrogen abundance estimates of (7.85 +/-0.08, 7.73 +/- 0.05), (7.97 +/- 0.08, 7.95 +/- 0.05), (7.94 +/- 0.08, 7.82 +/- 0.05) for 3D, HM, and MARCS respectively. The same correspondence between 3D and MARCS that was seen in oxygen is replicated in nitrogen, but the internal consistency in the HM model is higher than that in the other models. The formal significance of the disagreement in the 3D models is under $2 \sigma$, however, so we cannot exclude the possibility that they may be consistent. We therefore adopt the HM result as the central value (7.96 +/- 0.06), and treat the difference with the 3D result (7.78 +/- 0.06) as a $2 \sigma$ systematic error. This yields a total uncertainty in \[N/H\] of 0.10 dex dominated by systematic uncertainties. The Solar Beryllium Abundance ----------------------------- There is an interesting linkage between the solar O and Be abundances. In stellar interiors Be is destroyed at modest temperatures (of order 3.5 million K). It can therefore be used as a diagnostic of mixing in stars, especially in conjunction with the more fragile light element Li [@mhp1997]. Traditional model atmospheres studies [@cbm1975] yield a solar photospheric beryllium abundance roughly half of the meteoritic abundance. However, the only accessible Be feature is located in a crowded portion of the spectrum in the near UV, and the continuum opacity is uncertain in this regime (largely from the contribution of numerous weak iron lines). Since the strength of a line is a function of the ratio of the line to the continuous opacity, a higher photospheric Be could be derived if the continuous opacity background was higher than that of the model. In an important paper, @bb1998 pointed out that nearby OH lines could be used to test the continuous opacity close to beryllium. They derived a UV OH lines that were too strong if they used the absolute oxygen abundance obtained from the IR OH lines, and interpreted this as evidence that the continuous opacity is underestimated in the spectral window relevant for Be. Similar conclusions for 3D models were obtained by @a2004. Following @l2003, we note that the uncertainties in the ad hoc corrections are substantial. Asplund (2004) quotes photospheric and meteoritic abundance errors of 0.09 and 0.08 dex respectively, implying that his zero net photospheric depletion has a $1 \sigma$ uncertainty of 0.12 dex. Even if the Balachandran and Bell argument is entirely correct, the data sets a $2 \sigma$ limit of 0.24 dex on beryllium depletion and does not require that it be zero. There is also the possibility of substantial NLTE corrections to the UV OH lines, which would reduce or even eliminate the requirement for a mechanism to reduce the strength of the lines. We also note that the value of the oxygen used by @bb1998 for the IR lines (8.91) in the HM74 model is larger than the value derived from other molecular and atomic indicators, and even slightly larger than the value from the (v,r) transitions in the same model from the work of Asplund and collaborators. We contend that this promising approach still has substantial errors, including large uncertainties in the absolute solar oxygen abundance. We therefore believe that the approach of @l2003 is the best current picture of the degree of beryllium depletion in the Sun: namely, there is a substantial uncertainty in the degree of solar beryllium depletion, and that further work is required before powerful observational bounds can be used to constrain interiors calculations. Conclusions and Future Tests ============================ Our basic conclusion is simple: the difference between the solar CNO abundances as derived from model atmospheres and model interiors considerations is not statistically significant. The systematic errors in photospheric abundance indicators will have to be reduced before a “solar abundance problem” can be established (or ruled out) with confidence. However, the disagreement between the solar thermal structure and that of the simulations would favor the higher abundance scale, and there is some recently published evidence to that effect. If this is confirmed, it switches the nature of the problem from being a question of the correct abundance scale to a question of the uncertainties in numerical convection simulations. We begin with a synthesis and explanation of our findings. We then divide our conclusions into two parts. We recommend steps to more firmly establish the photospheric abundance scale, and contend that accurate solar abundances require tests of the thermal structure of the models and the magnitude of non-LTE abundance corrections. In our final subsection we then gather together evidence that the atmospheric abundance scale problem may be tied to the limited resolution in the convection simulations or errors in the underlying model atmosphere treatment. The consequences for the solar beryllium abundance, which is a useful diagnostic of internal mixing, are also explored. The two main justifications for the superiority of the 3D hydro atmospheres are the treatment of line broadening and the inclusion of granulation. Both of these represent genuine improvements in the atmospheric physics. However, neither of these effects is actually primarily responsible for the difference in the solar abundance scale. Many of the abundance indicators are insensitive to the effective microturbulence. If temperature fluctuations are imposed on a semi-empirical Holweger-Mueller atmosphere, the resulting granulation corrections are usually smaller than the 3D convection effects reported by Asplund and collaborators, and frequently opposite in sign [@h2001]. The main driver behind the systematic reductions in abundance derived from the 3D models is a theoretically predicted change in the thermal structure, coupled with large assumed non-LTE corrections for atomic features. Neither of these changes is directly supported by observational tests. Instead, the argument for the superiority of the abundances derived from the newer model atmospheres is an indirect one, focused on the concordance of abundances derived from different indicators. A consistent chain of logic emerges from the comprehensive studies of oxygen (AGSAK04) and carbon (AGSAB05). Classical LTE model atmospheres tend to yield internally consistent, and high, carbon and oxygen abundances for atomic and molecular indicators. The application of a different thermal structure in the 3D hydro atmospheres drastically reduces the abundances inferred from highly temperature sensitive molecular indicators, but has a smaller effect on atomic features. Large NLTE corrections are then applied to the abundances derived from permitted atomic features for both 1D and 3D models. The net result is that the abundance estimates from 1D models become internally inconsistent (atomic indicators yield lower abundances than molecular ones), while abundances derived from the 3D models are internally consistent. The abundances derived from forbidden lines are insensitive to NLTE effects, but they are reduced in the newer generation of models by the inclusion of blending features. As a secondary argument, the fits to individual indicators are argued to be superior in the 3D models when compared to the fits to individual indicators in the 1D models. This approach is appealing on the surface, but when examined in detail the picture is decidedly more ambiguous. If anything, the hints from the data would lean towards the opposite conclusion. The abundances derived from forbidden lines have the smallest systematic errors, but errors in both the theoretical oscillator strengths and the treatment of blending features result in non-negligible random errors. More to the point, the internal consistency of abundances derived from forbidden and molecular lines is actually similar in the 3D and 1D cases. From Table 1, the forbidden and molecular oxygen abundances are (8.71, 8.64) for 3D and (8.77, 8.84) for the HM; the differences are identical. Given the errors, neither discrepancy is statistically significant with high confidence. The abundances reported for permitted atomic features in AGSAK04 and AGSAB05 are significantly lower for 1D models than the corresponding molecular abundances, while the reported 3D results are in agreement. In the case of oxygen, this rests completely on the assignment of large NLTE corrections. These corrections were obtained under the assumption that hydrogen collisions were unimportant. Detailed studies of the response of the triplet to limb-darkening indicate that models including hydrogen collisions are favored, and the inferred NLTE corrections decrease. As a result, the internal consistency of the oxygen indicators is comparable for the different classes of atmospheres. Nitrogen is consistent for HM74 models and inconsistent (but at less than $2 \sigma$) for the 3D case. In the case of carbon, the situation is made more complex by significant zero-point offsets between earlier studies of carbon abundances that are not explained. Again, the assignment of larger NLTE corrections is uncertain (and, unlike the case of oxygen, not directly tested against limb-darkening data). A clean distinction between models on the basis of consistency is not obtained. However, the 3D models do yield different molecular and atomic abundances for both N and O, and might also do so for C. One might then hope to find distinct differences in the quality of the fits to different molecular indicators. The usual patterns, unfortunately, manifest themselves as simple zero-point shifts. For every case where there are issues with the 1D models (e.g. small trends with excitation potential in the \[O/H\] derived from (v,r) OH transitions in the HM model) there are comparable or larger effects for the 3D models (e.g. substantial trends in the \[O/H\] derived from (r,r) OH transitions). In a recent preprint, @scott2006 examined CO indicators, and the resulting pattern is illustrative. The 3D models yielded similar results for two of the three features studied, while the 1D models performed better in a different pair of indicators. The $C^{12}/C^{13}$ ratio from the 1D models ranges from 69 to 84, while the same ratio for the 3D models ranges from 83 to 108. These values should be contrasted with the expected terrestrial ratio of 89. Scott et al. (2006) choose comparisons that favor the 3D models, while an advocate of the traditional models might reasonably stress the other cases. In our view, the best choice of models is not clearly distinguishable from the CNO abundance studies. We recommend caution when extrapolating these model results to other stars, where the differential effects can be even more drastic. Establishing the Absolute Photospheric Abundance Scale ------------------------------------------------------ The single most important test that is required for atmospheres theory is a discriminant between the different proposed thermal structures of the solar atmosphere. The recent paper by Ayres et al. (2006) makes an important contribution by making direct comparisons of solar data with the thermal properties of the simulations. They present evidence that the solar center-to-limb variations in continuum flux are inconsistent with the predictions of the 3D hydro simulations. They also note that the predicted magnitude of fluctuations in the upper atmosphere from the simulations appears to be larger than the observed pattern. Ayres et al. then construct an empirical model of the atmosphere and derive a high oxygen abundance (8.85) from CO molecular features under the assumption of a fixed C/O ratio. In retrospect this conclusion is not surprising. The HM model is not a purely theoretical exercise; it was constructed to reproduce the mapping of the source function as a function of optical depth inferred from limb darkening studies of continuum flux and strong lines [see also @arg1998]. The relative trends we have inferred from atomic and molecular abundance indicators support the conclusions of Ayres et al., but the current errors make our evidence in this matter suggestive but not conclusive. It would also be highly beneficial to repeat the HM exercise with the full 3D models as opposed to the restricted form of them that Ayres et al. (2006) had available to them. Ultimately, the absolute accuracy of photospheric abundances is directly tied to the absolute accuracy of the thermal structure. This suggests that an approach similar to that of @sh2002 may be the optimal one. In their paper they examined the impact of temperature fluctuations around an assumed mean empirical thermal structure, which in their case was the HM74 model. Interestingly, the abundance corrections that they derive would act in the sense of increasing the concordance between abundance indicators. Oxygen abundances from atomic indicators would be slightly increased; although they did not consider molecular features directly, the net effect would certainly have the same sign as that obtained from 3D hydro models, namely a decrease in the inferred abundance. In such a differential approach, deviations between the mean structure of the simulations and the empirical data would be used as guidance concerning the underlying physics. In contrast, the 3D model abundances assume that the ab initio profile is correct. A similar approach could be employed for the velocity field that replaces the microturbulence and macroturbulence in traditional 1D atmospheres. A second ingredient that must be tested empirically, rather than by theoretical assertion, is the magnitude of NLTE corrections. The available evidence suggests that NLTE corrections are in general small for the Sun, but for the level of precision required in the absolute abundance scale these small corrections are significant. Studies of different spectral features yield different conclusions about the physical model employed in NLTE studies. This implies that there are significant uncertainties in absolute theoretical calculations. Fortunately, NLTE corrections can be constrained by the response of line strength to limb darkening in the Sun. It should be possible to develop improved theoretical models with a sufficient database of information developed in this fashion. One other stringent test of NLTE effects may be to focus on the species whose relative abundances can be reliably inferred from meteoritic data. For example, NLTE effects may be significant for iron [@stb2001] but less so for Si [@w2001]. @h2001 noted that there may be a conflict between the photospheric and meteoritic Fe/Si ratio, albeit one of marginal significance. A similar situation may exist for Na [@ags2005]. Another tractable problem is the absolute error for the forbidden C and O lines. In these cases, uncertainties in the line profiles and continuum levels should be included. Better atomic data (such as oscillator strengths for both the lines and the blending features) would also be useful. The accuracy of the theoretically predicted turbulent velocity field as a function of optical depth should also be subjected to a more rigorous analysis. @scott2006 present evidence that the generation of simulations used for the abundance analysis yielded poor fits to the line bisectors of CO lines. Higher resolution simulations gave better line profile fits, but for (unspecified) unrealistically high C/O abundances. The higher resolution simulations were not employed in the CO abundance analysis in that paper. It is worth keeping in mind that line profiles are integral quantities, and as a result the uniqueness of the solutions is not established by individual cases of good fits. This is particularly true when the abundance itself is treated as a free parameter. It would be extremely useful if future papers on abundances derived using numerical simulations illustrated individual line fits, as well as quantifying the actual impact of the “effective microturbulence” on the abundance estimates. It is useful to separate out the impact of velocity broadening from the effect of granulation and temperature gradient changes. This can be done by using the mean thermal structure and temperature fluctuations from the simulations and a more traditional micro/macroturbulence model to infer abundances, and comparing the results with the full 3D models. @scott2006 constructed such a test case (their 1DAV model), and found only small abundance offsets, of order 0.04 dex for oxygen derived from IR OH lines. They also inferred carbon abundances from CO; in this case O was held fixed and the carbon was adjusted to fit different molecular indicators. The deviations in the derived carbon abundances relative to the 3D case ranged from small (0.01 dex for the LE lines) to modest (0.06 dex for the weak $\Delta \nu = 1$ lines) to large (0.14 dex for the $\Delta \nu = 2$ lines). These deviations may explain the changes in excitation potential that @ay2006 needed to obtain consistent abundances within a 1D framework. This exercise implies that the impact of the improved microphysics varies substantially for different indicators, and is worth quantifying across the board. An alternate exercise (using the revised velocity field and relative temperature fluctuations while adopting a HM74 mean thermal structure) might also be illuminating. Uncertainties in Numerical Convection Simulations ------------------------------------------------- First-principles theoretical model atmosphere calculations have undeniable strengths. The ability to naturally reproduce line widths and include granulation is a powerful addition to our ability to reliably interpret stellar and solar spectra. The principal difficulty with such models is that errors in the input physics generate absolute errors in the inferred atmospheric structure that cannot be calibrated away in the absence of explicit free parameters. This phenomenon is the major reason why numerical convection simulations have not replaced the simple mixing length theory in stellar interiors calculations. Interiors models that can reproduce observed stellar radii are simply more useful for most purposes than models with a better physical treatment of convection that fail to do so. Before the results from such models are adopted as the new abundance standard, it will be necessary to perform an extensive theoretical error analysis and to compare the models with the strongest observational constraints. We believe that accurate solar abundance calculations must reproduce the observed solar thermal structure, and from the Ayres et al. (2006) paper the Asplund models employing numerical convection simulations appear to yield a temperature gradient steeper than the real Sun. This could be caused by errors in the background (1D) stellar atmospheres treatment; for example, uncertainties in the equation of state and continuous opacities will induce absolute errors in the thermal structure. An approach similar to that employed in interiors models would be useful for assessing the uncertainties in the thermal structure and abundance predictions, and this should be included in the error budget for abundances. It is more likely, however, that the major error source in 3D hydro model atmospheres is related to uncertainties in the numerical convection simulations. The approximations in hydro simulations of giant planet atmospheres are have been demonstrated to be strongly affected by the quality of the assumed physics [@eg2004]. @zs2006 also provides a good summary of the uncertainties in the related problem of terrestrial and solar dynamo models. Another phenomenon that could be related is the issue of convective overshooting below surface convection zones. Numerical simulations have tended to favor extensive overshooting, and the early models had a substantial nearly adiabatic overshoot region, in conflict with the stringent limits set by seismology (less than $ 0.05 H_p$). More recent 3D @bct2002 and 2D [@rg2005; @rg2006] calculations found that the filling factor for plumes is smaller than previously thought, which led to an overestimate in earlier models of the changes induced by overshooting in the thermal structure. The newer simulations predict strongly subadiabatic overshooting (effectively, overmixing without changing the thermal structure), which is consistent with the seismic limits. However, they still produce a substantial mixed region below the surface convection zone of order $0.4 H_p$. Since even a small overmixing of $0.05 H_p$ drastically increases pre-MS lithium depletion [@mhp1997], which is already too efficient relative to stellar data [@ptc2002], it is likely that even this reduced overshooting is too large to be compatible with stellar constraints. We argue that there is a common pattern in both “undershooting” and “overshooting” above and below convective regions. In both cases, the numerical simulations may be overestimating the degree of mixing and the impact on the thermal structure of convection outside the formal bounds set by the Schwartzschild criterion. There are two plausible error sources that should be investigated. The treatment of heat transfer in the atmosphere convection simulations is necessarily simplified, and this may be leading to an artificial inhibition in energy transport between turbulent cells projected into the radiative atmosphere and their surroundings. Resolution effects, however, may be even more important. Even the highest resolution simulations available today are many orders of magnitude away from being able to reproduce the characteristic Reynolds numbers in the Sun. Scott et al. (2006) found significant changes in line bisectors for the outer layers of their solar model when they increased their resolution, and these changes were in the sense of reducing the temperature contrast in the upper atmosphere and improving the shape of the bisectors relative to data. Numerical tests with substantially increased resolution may shed some interesting light on the sensitivity of the predictions to the underlying numerics; 2D convection simulations may be useful in this regard. We are optimistic that the net effect of such testing will be a greatly improved understanding of the strengths and weaknesses of theoretical atmospheres models, just as we are confident that the net result of the solar abundance controversy will be a far more secure knowledge of stellar abundances. We would like to thank Martin Asplund for providing tables of the abundances derived from molecular oxygen abundance indicators. We would also like to thank Don Terndrup, Jennifer Johnson, Andreas Korn, Chris Sneden, and Hans Ludwig for helpful discussions on stellar abundance determinations. FD would like to thank Claude Zeippen for discussions on the uncertainties in the atomic data. Allende Prieto, C., Asplund, M., & Fabiani Bendicho, P. 2004, , 423, 1109 Allende Prieto,C., Lambert, D. L., & Asplund, M. 2001, , 556, L63 Allende Prieto,C., Lambert, D. L., & Asplund, M. 2002, , 573, L137 Allende Prieto, C., Ruiz Cobo, B., & Garcia Lopez, R. J. 1998, , 502, 951 Anders, E., & Grevesse, N. 1989, , 53, 197 Antia, H. M., & Basu, S. 2006, , in press (astro-ph/0603001) Asplund, M. 2004, , 417, 769 Asplund, M., Nordlund, [Å]{}., Trampedach, R., & Stein, R. F. 2000, , 359, 743 Asplund, M., Nordlund, [Å]{}., Trampedach, R., Allende Prieto, C., & Stein, R. F. 2000, , 359, 729 Asplund, M. 2000, , 359, 755 Asplund, M., Grevesse, N., & Sauval, A. J. 2005, ASP Conf. Ser. 336: Cosmic Abundances as Records of Stellar Evolution and Nucleosynthesis, 336, 25 Asplund, M., Grevesse, N., Sauval, A. J., Allende Prieto, C., & Kiselman, D. 2005, , 435, 339 Asplund, M., Grevesse, N., Sauval, A. J., Allende Prieto, C., & Blomme, R. 2005, , 431, 693 Asplund, M., Grevesse, N., Sauval, A. J., Allende Prieto, C., & Kiselman, D. 2004, , 417, 751 Ayres, T. R., Plymate, C. & Keller, C. U. 2006, , (in press) Balachandran, S. C., & Bell, R. A. 1998, , 392, 791 Basu, S., & Antia, H. M. 2004, , 606, L85 Biemont, E., Hibbert,A., Godefroid, M., & Vaeck, N. 1993, , 412, 431 Biemont, E., Hibbert, A., Godefroid, M., Vaeck, N., & Fawcett, B. C. 1991, , 375, 818 Brummell, N. H., Clune, T. L., & Toomre, J. 2002, , 570, 825 Carlsson,M., Uppsala Astronomical Observatory Repport No 33 Chmielewski, Y., Brault, J. W., & Mueller, E. A. 1975, , 42, 37 Christensen-Dalsgaard, J., Monteiro, M. J. P. F. G., & Thompson, M. J. 1995, , 276, 283 Delahaye, F., & Pinsonneault, M. 2006, , 647, in press Drawin, H. W., Z. Phys, 211,404 Evonak, M. & Glatzmaier, G. 2004 Geophysical and Astrophysical Fluid Dynamics 98, 241 Grevesse, N., & Noels, A. 1993, Physica Scripta Volume T, 47, 133 Grevesse, N., & Sauval, A. J. 1998, Space Science Reviews, 85, 161 Guzik, J. A., Watson, L. S., & Cox, A. N. 2005, , 627, 1049 Holweger, H. 2001, AIP Conf. Proc. 598: Joint SOHO/ACE workshop ”Solar and Galactic Composition”, 598, 23 Holweger, H., & Mueller, E. A. 1974, , 39, 19 Johansson, S., Litz[é]{}n, U., Lundberg, H., & Zhang, Z. 2003, , 584, L107 Kiselman, D. 1993, , 275, 269 Lodders, K. 2003, , 591, 1220 Mel[é]{}ndez, J. 2004, , 615, 1042 Piau, L., & Turck-Chi[è]{}ze, S. 2002, , 566, 419 Pinsonneault, M. 1997, , 35, 557 Reetz, J. 1998, PhD thesis, Ludwig-Maximilians Univ. Rogers, T. M., & Glatzmaier, G. A. 2005, , 620, 432 Rogers, T. M., & Glatzmaier, G. A. 2006 submitted to ApJ - astro-ph/0601668 Schmelz, J. T., Nasraoui, K., Roames, J. K., Lippner, L. A., & Garst, J. W. 2005, , 634, L197 Scott,P. C., Asplund, M., Grevesse, N. ,A., Sauval, J.  2006, in press, astro-ph/0605116 Shchukina, N., & Trujillo Bueno, J. 2001, , 550, 970 Steffen, M., & Holweger, H. 2002, , 387, 258 Stein, R. F., & Nordlund, A. 1998, , 499, 914 Storey, P. J., & Zeippen, C. J. 2000, , 312, 813 Stuerenburg, S., & Holweger, H. 1990, , 237, 125 Turck-Chi[è]{}ze, S., Couvidat, S., Piau, L., Ferguson, J., Lambert, P., Ballot, J., Garc[í]{}a, R. A., & Nghiem, P. 2004, Physical Review Letters, 93, 211102 Wedemeyer, S. 2001, , 373, 998 Zhang, K., & Schubert, G. 2006, Reports of Progress in Physics, 69, 1581 Table1.tex Table2.tex Table3.tex
{ "pile_set_name": "ArXiv" }
--- abstract: 'Let $ X_1, \ldots , X_n $ be an i.i.d. sample in $ \R^p $ with zero mean and the covariance matrix $ \St $. The classical PCA approach recovers the projector $ \Pt $ onto the principal eigenspace of $ \St $ by its empirical counterpart $ \Pe $. Recent paper [@Koltchinskii_NAACOSPOSC] investigated the asymptotic distribution of the Frobenius distance between the projectors $ \|\Pe - \Pt\|_2 $, while [@Naumov_BCSFSPOSC] offered a bootstrap procedure to measure uncertainty in recovering this subspace $ \Pt $ even in a finite sample setup. The present paper considers this problem from a Bayesian perspective and suggests to use the credible sets of the pseudo-posterior distribution on the space of covariance matrices induced by the conjugated Inverse Wishart prior as sharp confidence sets. This yields a numerically efficient procedure. Moreover, we theoretically justify this method and derive finite sample bounds on the corresponding coverage probability. Contrary to [@Koltchinskii_NAACOSPOSC; @Naumov_BCSFSPOSC], the obtained results are valid for non-Gaussian data: the main assumption that we impose is the concentration of the sample covariance $ \Se $ in a vicinity of $ \St $. Numerical simulations illustrate good performance of the proposed procedure even on non-Gaussian data in a rather challenging regime.' address: - | Moscow Institute of Physics and Technology,\ 141701, Dolgoprudny, RF.\ - | Weierstrass Institute and Humboldt University,\ Mohrenstr. 39, 10117 Berlin, Germany.\ - | National Research University Higher School of Economics,\ 20 Myasnitskaya ulitsa, 101000, Moscow, RF. - | Skolkovo Institute of Science and Technology (Skoltech),\ 143026, Moscow, RF. - | Institute for Information Transmission Problems RAS,\ Bolshoy Karetny per. 19, 127051, Moscow, RF. author: - - title: Bayesian inference for spectral projectors of the covariance matrix --- pp\_main\_march.tex ./source/Numerical.tex pp\_proofs\_march.tex [99]{} ./source/Bibliography.tex
{ "pile_set_name": "ArXiv" }
--- author: - Tommi Tenkanen - and Eemeli Tomberg bibliography: - 'Initial\_conditions.bib' title: Initial conditions for plateau inflation --- Introduction ============ In this paper we study the initial conditions that are needed to start a phase of accelerated expansion, i.e. cosmic inflation, in the case where inflation is driven by a scalar field whose potential exhibits a plateau regime. We do not confine ourselves to models where inflation occurs necessarily at large scales but study also scenarios where the scale of inflation and the corresponding field value $\phi$ can be much smaller than the Planck scale $M_{\rm P}$. Scenarios like this are encountered, for example, in various models where the Standard Model (SM) Higgs field drives inflation [@Salopek:1988qh; @Bezrukov:2007ep; @Bauer:2008zj] and which are therefore of particular interest in the search for the connection between low energy particle physics and cosmic inflation (for a recent review, see Ref. [@Rubio:2018ogq]), or in the context of so-called $\alpha$-attractor models [@Ferrara:2013rsa; @Kallosh:2013yoa]. Furthermore, plateau models are precisely those most favored by the Cosmic Microwave Background (CMB) data [@Akrami:2018odb; @Martin:2013nzq]. In order for (slow-roll) inflation to happen, the kinetic energy stored in the inflaton field must remain (much) smaller than the potential energy. However, it is a non-trivial question of how generically this happens in a given model of inflation. This was studied in the seminal papers [@Kung:1990; @Goldwirth:1990; @Goldwirth:1992] (see also Refs. [@Feldman:1989; @East:2015ggf; @Clough:2016ymm]), in which it was shown that “chaotic” models of inflation where inflation occurs at $\phi > M_{\rm P}$ are not very sensitive to the initial conditions but the occurence of inflation in the so-called small-field or hilltop models with $\phi \lesssim M_{\rm P}$ can depend heavily on the initial conditions. However, as recently shown in Ref. [@Chowdhury:2019otk], models where the inflaton potential exhibits a plateau are not “small-field” models in the usual sense, and inflation can occur also at $\phi \sim M_{\rm P}$ without a need for fine-tuning[^1]. In this study we will further consolidate this claim by studying scenarios where the field range can be considerably smaller than the Planck scale and also where the inflaton kinetic term can be non-canonical. In this way, we extend the study of Ref. [@Chowdhury:2019otk] to address the claims made by Ijjas et al. in Ref. [@Ijjas:2013vea] concerning fine-tuning in the initial conditions needed to start inflation and how the Planck data favor models for which this issue is, according to Ref. [@Ijjas:2013vea], most problematic. We will show that this is generically not the case in a homogeneous and isotropic situation even in scenarios where inflation occurs at very small field values, $\phi \ll M_{\rm P}$. In addition to presenting this important result, we will also comment on the claims made in Ref. [@Bedroya:2019tba] regarding the required amount of fine-tuning in initial conditions for inflation in the context of the “Trans-Planckian Censorship” conjecture and show that generically no fine-tuning is needed in models like those studied in this paper. The paper is organized as follows: in Sec. \[inflation\], we present the class of inflationary models we study in this paper and discuss the scaling laws that the inflaton equation of motion and the Friedmann equation exhibit in our scenario. In Sec. \[results\], we present our main results regarding the initial conditions and predictions for CMB observables in these models and then conclude in Sec. \[conclusions\]. Cosmic inflation {#inflation} ================ As an interesting example of a model where the inflaton potential exhibits a plateau, we consider the action $$\label{SJ} S_J = \int {\rm d}^4x \sqrt{-g}\left(\frac12 F(R,\phi)- \frac12g^{\mu\nu}\partial_\mu\phi\partial_\nu\phi - V(\phi)\right) ,$$ where $g$ is the determinant of the space-time metric $g_{\mu\nu}$, $R=g^{\mu\nu}R_{\mu\nu}(\Gamma,\partial\Gamma)$ is the curvature scalar, $R_{\mu\nu}$ is the Ricci tensor which is constructed by contraction from the Riemann tensor $R\indices{^\lambda_{\mu\nu\sigma}}$ which, in turn, is constructed from the space-time connection $\Gamma$ and its first derivatives, and $\phi$ is the inflaton field. In the following, we take $$\label{F} F(R,\phi) = M_{\rm P}^2 R + \alpha R^2 +G(\phi)R,$$ where $\alpha$ is a dimensionless parameter and $G(\phi)$ encapsulates the possible non-minimal couplings between the inflaton and gravity. In the following, we will take $G(\phi)=\xi \phi^2$ for simplicity, and allow the connection to depend on the inflaton field value too, in addition to the space-time metric. That is, we do [*not*]{} require that the connection should be the usual Levi-Civita one but allow for a more general starting point[^2]. We do, however, require the connection to be symmetric in its two lower indices, $\Gamma^\lambda_{\mu\nu}=\Gamma^\lambda_{\nu\mu}$, so that the torsion tensor vanishes (for scenarios with non-vanishing torsion, see Refs. [@Rasanen:2018ihz; @Shimada:2018lnm]). Thus, $\alpha=0\,, \xi\neq 0$ corresponds to the usual non-minimal inflation (see e.g. [@Spokoiny:1984bd; @Futamase:1987ua; @Fakir:1990eg; @Amendola:1990nn; @Kaiser:1994vs; @Komatsu:1999mt; @Bezrukov:2007ep; @Lerner:2009xg; @Kaiser:2013sna; @Kallosh:2013maa; @Kallosh:2013tua; @Chiba:2014sva; @Boubekeur:2015xza]) and $\alpha\neq 0$ to the scenario studied in Refs. [@Enckell:2018hmo; @Antoniadis:2018ywb; @Antoniadis:2018yfq; @Tenkanen:2019jiq; @Tenkanen:2019wsd; @Gialamas:2019nly], while $\alpha=0=\xi$ represents the minimally coupled case where only the form of $V(\phi)$ is important. A Weyl transformation $$g_{\mu\nu}\to \frac{\varphi + G(\phi)}{M_{\rm P}^2} g_{\mu\nu} ,$$ where $\varphi$ is an auxiliary field shows that the action is equivalent to [@Enckell:2018hmo] $$\label{SE} S_E = \int {\rm d}^4x \sqrt{-g}\bigg[\frac12 M_{\rm P}^2 R- \frac12\partial^\mu\chi\partial_\mu\chi + \frac{\alpha}{2 M_{\rm P}^4}\left(1+8\alpha\frac{\bar{U}}{M_{\rm P}^4}\right)\left(\partial^\mu\chi\partial_\mu\chi\right)^2 - U(\chi)\bigg] ,$$ where the re-defined inflaton field $\chi$ is given by $$\label{field_redefinition} \frac{{\rm d}\phi}{{\rm d}\chi} =\sqrt{\left(1+\frac{G(\phi)}{M_{\rm P}^2}\right)\left(1+8\alpha\frac{\bar{U}}{M_{\rm P}^4}\right)} ,$$ and the potential is $$\label{U} U(\chi) \equiv \frac{\bar{U}(\chi)}{1+8\alpha\bar{U}(\chi)/M_{\rm P}^4}\,, \quad \bar{U}(\chi) \equiv \frac{V(\phi(\chi))}{\qty[1+G(\phi(\chi))/M_{\rm P}^2 ]^2} .$$ Note that only one field is dynamical; this is due to the fact that the Jordan frame connection depends on both $\phi$ and $g_{\mu\nu}$ and not just the metric. This renders the $\alpha R^2$ term non-dynamical, unlike in the case of the famous Starobinsky model [@Starobinsky:1980te], which is based on metric gravity (where the connection is precisely the Levi-Civita one). However, a non-zero $\alpha R^2$ term will still affect the dynamics of $\chi$, as we will see more explicitly in the following. In the Friedmann–Robertson–Walker (FRW) case with zero spatial curvature the Friedmann equation and the equation of motion for the field read [@Enckell:2018hmo] $$\begin{aligned} \label{FRW_eqs} 3 M_{\rm P}^2 H^2 &=& \frac12 \bigg[1+3\alpha(1+8\alpha\frac{\bar{U}}{M_{\rm P}^4})\frac{\dot{\chi}^2}{M_{\rm P}^4} \bigg]\dot{\chi}^2+U\,,\\ \nonumber 0 &=& \bigg[1+6\alpha(1+8\alpha\frac{\bar{U}}{M_{\rm P}^4})\frac{\dot{\chi}^2}{M_{\rm P}^4}\bigg]\ddot{\chi}+3\bigg[1+2\alpha(1+8\alpha\frac{\bar{U}}{M_{\rm P}^4})\frac{\dot{\chi}^2}{M_{\rm P}^4}\bigg]H\dot{\chi} + 12\alpha^2\frac{\dot{\chi}^4}{M_{\rm P}^8}\bar{U}' + U'\, ,\end{aligned}$$ and inflation happens when the first slow-roll parameter $$\label{epsilon_H} \epsilon_H \equiv -\frac{\dot{H}}{H^2} = \frac{\dot{\chi}^2}{2M_{\rm P}^2H^2}\qty[ 1 + 2\alpha\qty(1+8\alpha\frac{\bar{U}}{M_{\rm P}^4})\frac{\dot{\chi}^2}{M_{\rm P}^4} ]$$ is smaller than one. ![A sketch of the Einstein frame potential $U(\chi)$ with the quartic Jordan frame potential . The potential is exponentially flat for $\chi/M_{\rm P} > (\xi^2 + 8\alpha_{\rm eff})^{-1/4}$. []{data-label="plateau_potential"}](plateau_potential.pdf) In Ref. [@Enckell:2018hmo], the equations were studied in the slow-roll limit, where $\epsilon_H \ll 1$ and the $\ddot{\chi}$-term in is negligible. It was shown that the leading slow-roll predictions for the power spectrum amplitude and the spectral index do not depend on $\alpha$ but are equal to the standard $\alpha=0$ case, that is, $$\label{As_and_ns} 24 \pi^2 M_{\rm P}^4 A_s = \frac{U}{\epsilon_U} = \frac{\bar{U}}{\epsilon_{\bar{U}}} \, , \qquad n_s = 1 + 2\eta_U - 6\epsilon_U = 1 + 2\eta_{\bar{U}} - 6\epsilon_{\bar{U}} \, ,$$ where the potential slow-roll parameters are defined as $$\begin{aligned} \epsilon_U &\equiv \frac{1}{2}M_{\rm P}^2\qty(\frac{U'}{U})^2 \, , \qquad & \epsilon_{\bar{U}} &\equiv \frac{1}{2}M_{\rm P}^2\qty(\frac{\bar{U}'}{\bar{U}})^2 = \epsilon_U \large|_{\alpha=0} \, , \\ \eta_U &\equiv M_{\rm P}^2\frac{U''}{U} \, , \qquad & \eta_{\bar{U}} &\equiv M_{\rm P}^2\frac{\bar{U}''}{\bar{U}} = \eta_U\large|_{\alpha=0} \, . \end{aligned}$$ The number of e-folds of inflation does not depend on $\alpha$ either[^3]. The main effect of a non-zero $\alpha$ is to lower the predicted tensor-to-scalar ratio: $$\label{r} r = 16\epsilon = \frac{16}{1+8\alpha\bar{U}/M_{\rm P}^4}\epsilon_{\bar{U}} \, ,$$ so that in case of large $\alpha$, the scenario yields negligible primordial gravitation waves. The Planck and BICEP2/Keck Array limits for the aforementioned observables are [@Akrami:2018odb; @Ade:2018gkx] $$\label{planck} A_s = 2.1 \times 10^{-9} \, , \quad n_s = 0.9625 \pm 0.0048 \, , \quad r < 0.06 \, ,$$ measured at the pivot scale $k=0.05\,{\rm Mpc}^{-1}$. In this paper, we will show by studying few example potentials that the slow-roll solution is still an attractor in the presence of the non-minimal kinetic terms, so the results and are general and practically independent of the initial conditions. An example: Higgs inflation --------------------------- As a representative example, we study the potential $$\label{V} V(\phi) = \frac{\lambda}{4}\phi^4 \, ,$$ which is the (Jordan frame) potential encountered in e.g. the Higgs inflation model, where the SM Higgs is the field responsible for driving inflation [@Salopek:1988qh; @Bezrukov:2007ep; @Bauer:2008zj]. The potential $U(\chi)$ is then exponentially flat for large field values as long as either $\xi$ or $\alpha$ is non-zero, see Fig. \[plateau\_potential\]. If $\xi \ll 1$, we have slow-roll inflation with the quartic potential at the CMB scales, and a non-zero $\alpha$ only modifies $r$ as discussed above. The observables in terms of the number of e-folds $N$ are easily calculated and read $$\label{phi4_observables} A_s = \frac{2\lambda N^3}{3\pi^2} \, , \quad n_s = 1 - \frac{3}{N} \, , \quad r = \frac{16}{N(1 + 128\alpha\lambda N^2)} \, .$$ If $\xi \gg 1$, we have the Palatini-Higgs inflation [@Bauer:2008zj], so $$\label{higgs_inf_observables} A_s = \frac{\lambda N^2}{12\pi^2 \xi} \, , \quad n_s = 1 - \frac{2}{N} \, , \quad r = \frac{2}{\xi N^2(1 + 2\alpha\lambda/\xi^2)} \, .$$ Here $N$ depends on the energy scale of inflation, and using the observed value of $A_s$ and assuming instant reheating it can be written for our pivot scale $k=0.05\,{\rm Mpc}^{-1}$ as $$\label{N} N \approx 56 - \frac{1}{4}{\rm ln}\left( \frac{0.06}{r}\right) \, .$$ We will consider the compatibility of these predictions with the Planck observations in section \[cmb\_predictions\]. Before analyzing the time evolution in this model, we note that the model obeys a scaling law: when the coordinates and couplings in the Jordan frame action are scaled as $$\label{scaling} x^\mu \to sx^\mu \, , \quad \alpha \to s^2\alpha \, , \quad \lambda \to s^{-2}\lambda \, ,$$ where $s$ is a scaling parameter, then the scalar curvature changes as $R \to s^{-2} R$, and the action stays the same apart from a constant scaling[^4]: $$\label{scaling_SJ} S_J \to s^2 S_J \, .$$ In this scaling, the parameter $$\label{alpha_eff} \alpha_{\rm eff} \equiv \alpha\lambda$$ remains constant. The classical equations of motion do not change in such a rescaling of the action. There is then no change in some of the dimensionless parameters, such as the slow-roll parameters, $r$, or $n_s$ as functions of the number of e-folds of inflation, as can be directly seen from , . There is a scaling in the dimensionful parameters, however, such as the energy scale of inflation, as well as in $A_s$. In the next section, we will consider the time evolution of the model in phase-space with different parameter values. Because of the scaling symmetry, it is enough to consider different values of $\xi$ and $\alpha_{\rm eff}$. The phase-space flow diagrams we present do not depend on the values of $\lambda$ and $\alpha$ separately. These can always be fixed for any $\alpha_{\rm eff}$ by, for example, fixing the power spectrum amplitude $A_s$ at the CMB scale. Results ======= Attractor behaviour {#attractor} ------------------- We study the phase-space flow of solutions to with the potential in two cases. First, Figs. \[alpha\_10to9\_chi\] and \[alpha\_10to9\_phi\] show the inflaton trajectories in phase-space for the parameter values $$\xi=0 \, , \qquad \alpha_{\rm eff}=10^9 \, .$$ As emphasized above, these trajectories only depend on $\alpha_{\rm eff}$, not $\alpha$ and $\lambda$ separately. For this choice, the canonical field $\chi$ has a maximum value $$\chi_{\rm max} \approx 8.9 \times 10^{-3} M_{\rm P} \,,$$ corresponding to $\phi \to \infty$, which comes from integrating with $G=0$. The inflationary plateau is compressed to values close to $\chi_{\rm max}$, so the phase-space diagrams in Fig. \[alpha\_10to9\_chi\] presented in $(\chi,\dot{\chi})$ coordinates are not very informative. For this reason, Fig. \[alpha\_10to9\_phi\] shows the diagrams in $(\phi, \dot{\phi})$ coordinates with a long, explicit plateau at $\phi \gg M_{\rm P}(8\alpha_{\rm eff})^{-1/4}$. ![Phase-space flow of the solutions to Eq. with $\xi=0$, $\alpha_{\rm eff}=10^9$. The red dashed lines correspond to $\epsilon_H = 1$, so inflation occurs in points between them. Slow-roll inflation happens near the points $(\pm\chi_{\rm max},0)$. Trajectories in the light-coloured upper region eventually end up to the slow-roll regime at $\chi \approx \chi_{\rm max}$ and the trajectories in the dark-coloured lower region eventually end up to the slow-roll regime at $\chi \approx -\chi_{\rm max}$. Only the trajectories confined in the middle region end up oscillating near the potential minimum without ever entering slow-roll. **Left:** The whole $\chi$-range, with $\dot{\chi}_\epsilon$ defined to be the maximum $\dot{\chi}$-value on the dashed lines. **Right:** A zoomed-in version near the positive $\chi_{\rm max}$ showing the convergence of trajectories to the horizontal slow-roll line from above, and the divergence of trajectories below which will overshoot the minimum and end up into slow-roll on the other side of the potential.[]{data-label="alpha_10to9_chi"}](alpha_10to9_chi_1.pdf "fig:") ![Phase-space flow of the solutions to Eq. with $\xi=0$, $\alpha_{\rm eff}=10^9$. The red dashed lines correspond to $\epsilon_H = 1$, so inflation occurs in points between them. Slow-roll inflation happens near the points $(\pm\chi_{\rm max},0)$. Trajectories in the light-coloured upper region eventually end up to the slow-roll regime at $\chi \approx \chi_{\rm max}$ and the trajectories in the dark-coloured lower region eventually end up to the slow-roll regime at $\chi \approx -\chi_{\rm max}$. Only the trajectories confined in the middle region end up oscillating near the potential minimum without ever entering slow-roll. **Left:** The whole $\chi$-range, with $\dot{\chi}_\epsilon$ defined to be the maximum $\dot{\chi}$-value on the dashed lines. **Right:** A zoomed-in version near the positive $\chi_{\rm max}$ showing the convergence of trajectories to the horizontal slow-roll line from above, and the divergence of trajectories below which will overshoot the minimum and end up into slow-roll on the other side of the potential.[]{data-label="alpha_10to9_chi"}](alpha_10to9_chi_2.pdf "fig:") ![Same as Fig. \[alpha\_10to9\_chi\] but for $\alpha_{\rm eff} = 0$, $\xi = 9.5 \times 10^8$. Here $\chi_{100}$ is the minimum value of $\chi$ where there is $\mathcal{O}(100)$ e-folds of slow-roll inflation left. **Left:** A wide view; the trajectories continue almost horizontally outside of the figure. **Right:** A zoomed-in version near $\chi_{100}$, where the curving of the trajectories to the slow-roll attractor near $\dot{\chi}=0$ can be seen.[]{data-label="xi_10to9_chi"}](xi_10to9_chi_1.pdf "fig:") ![Same as Fig. \[alpha\_10to9\_chi\] but for $\alpha_{\rm eff} = 0$, $\xi = 9.5 \times 10^8$. Here $\chi_{100}$ is the minimum value of $\chi$ where there is $\mathcal{O}(100)$ e-folds of slow-roll inflation left. **Left:** A wide view; the trajectories continue almost horizontally outside of the figure. **Right:** A zoomed-in version near $\chi_{100}$, where the curving of the trajectories to the slow-roll attractor near $\dot{\chi}=0$ can be seen.[]{data-label="xi_10to9_chi"}](xi_10to9_chi_2.pdf "fig:") ![Same as Fig. \[alpha\_10to9\_chi\] but with the Jordan frame field $\phi$ instead of $\chi$. Slow-roll happens near $\dot{\phi}=0$, and $\dot{\phi}_\epsilon$ is defined as the $\dot{\phi}$-value with $\epsilon_H=1$ when $\phi=10^4 M_{\rm P} \times (8\alpha_{\rm eff})^{-1/4}$. **Left:** A representative image of the phase-space, with the maximum value of $\phi$ chosen so that it corresponds to $\mathcal{O}(100)$ e-folds of slow-roll inflation. **Middle:** A zoomed-in version showing the flow near the origin. **Right:** A zoomed-out version, which shows the eventual convergence of trajectories towards slow-roll.[]{data-label="alpha_10to9_phi"}](alpha_10to9_phi_1.pdf "fig:") ![Same as Fig. \[alpha\_10to9\_chi\] but with the Jordan frame field $\phi$ instead of $\chi$. Slow-roll happens near $\dot{\phi}=0$, and $\dot{\phi}_\epsilon$ is defined as the $\dot{\phi}$-value with $\epsilon_H=1$ when $\phi=10^4 M_{\rm P} \times (8\alpha_{\rm eff})^{-1/4}$. **Left:** A representative image of the phase-space, with the maximum value of $\phi$ chosen so that it corresponds to $\mathcal{O}(100)$ e-folds of slow-roll inflation. **Middle:** A zoomed-in version showing the flow near the origin. **Right:** A zoomed-out version, which shows the eventual convergence of trajectories towards slow-roll.[]{data-label="alpha_10to9_phi"}](alpha_10to9_phi_2.pdf "fig:") ![Same as Fig. \[alpha\_10to9\_chi\] but with the Jordan frame field $\phi$ instead of $\chi$. Slow-roll happens near $\dot{\phi}=0$, and $\dot{\phi}_\epsilon$ is defined as the $\dot{\phi}$-value with $\epsilon_H=1$ when $\phi=10^4 M_{\rm P} \times (8\alpha_{\rm eff})^{-1/4}$. **Left:** A representative image of the phase-space, with the maximum value of $\phi$ chosen so that it corresponds to $\mathcal{O}(100)$ e-folds of slow-roll inflation. **Middle:** A zoomed-in version showing the flow near the origin. **Right:** A zoomed-out version, which shows the eventual convergence of trajectories towards slow-roll.[]{data-label="alpha_10to9_phi"}](alpha_10to9_phi_3.pdf "fig:") ![Same as Fig. \[alpha\_10to9\_phi\] but for $\alpha_{\rm eff} = 0$, $\xi = 9.5 \times 10^8$. The field value $\phi = 10^6 M_{\rm P}/\sqrt{\xi}$ corresponds to $\mathcal{O}(100)$ e-folds of slow-roll inflation and $\dot{\phi}_\epsilon$ is the $\dot{\phi}$-value with $\epsilon_H=1$ when $\phi = 10^6 M_{\rm P}/\sqrt{\xi}$. The behaviour is almost identical to Fig. \[alpha\_10to9\_phi\] despite the fact that the middle region without slow-roll is almost too narrow to be seen in the left and right panels.[]{data-label="xi_10to9_phi"}](xi_10to9_phi_1.pdf "fig:") ![Same as Fig. \[alpha\_10to9\_phi\] but for $\alpha_{\rm eff} = 0$, $\xi = 9.5 \times 10^8$. The field value $\phi = 10^6 M_{\rm P}/\sqrt{\xi}$ corresponds to $\mathcal{O}(100)$ e-folds of slow-roll inflation and $\dot{\phi}_\epsilon$ is the $\dot{\phi}$-value with $\epsilon_H=1$ when $\phi = 10^6 M_{\rm P}/\sqrt{\xi}$. The behaviour is almost identical to Fig. \[alpha\_10to9\_phi\] despite the fact that the middle region without slow-roll is almost too narrow to be seen in the left and right panels.[]{data-label="xi_10to9_phi"}](xi_10to9_phi_2.pdf "fig:") ![Same as Fig. \[alpha\_10to9\_phi\] but for $\alpha_{\rm eff} = 0$, $\xi = 9.5 \times 10^8$. The field value $\phi = 10^6 M_{\rm P}/\sqrt{\xi}$ corresponds to $\mathcal{O}(100)$ e-folds of slow-roll inflation and $\dot{\phi}_\epsilon$ is the $\dot{\phi}$-value with $\epsilon_H=1$ when $\phi = 10^6 M_{\rm P}/\sqrt{\xi}$. The behaviour is almost identical to Fig. \[alpha\_10to9\_phi\] despite the fact that the middle region without slow-roll is almost too narrow to be seen in the left and right panels.[]{data-label="xi_10to9_phi"}](xi_10to9_phi_3.pdf "fig:") Second, Figs. \[xi\_10to9\_chi\] and \[xi\_10to9\_phi\] show the trajectories for $$\xi=9.5\times 10^8 \, , \qquad \alpha_{\rm eff}=0 \, ,$$ which gives the correct CMB predictions for $\lambda=0.1$, see Eqs. , . This time, there is no maximum value for $\chi$ but at the CMB scales, $N \approx 50$, we have $\phi \gg M_{\rm P}$ and the canonical field value $$\chi_{50} \approx 4.6 \times 10^{-4} M_{\rm P} \, ,$$ well below the Planck scale. As can be seen from the figures, most trajectories end up in the regime where slow-roll inflation will eventually take place. Either the trajectories fall into slow-roll directly, or they pass $\chi=0$, slow down, and enter slow-roll on the other side of the potential. Only the trajectories with fine-tuned initial conditions end up oscillating around the origin without yielding a significant amount of slow-roll inflation. This is generically true for all potentials where at least one of the parameters $\xi$, $\alpha_{\rm eff}$ is large. This statement of course depends on the choice of the field parameters and the phase-space probability measure and should therefore be taken in a qualitative manner only. Predictions for observables {#cmb_predictions} --------------------------- Finally, let us see when the predictions of the model comply with the CMB observations , in particular with the measured value of $n_s$, as for large $\xi$ or $\alpha$ the tensor-to-scalar ratio is always within the Planck bounds and the value of $\lambda$ can always be chosen such that the predicted amplitude $A_s$ matches with observations. From , we see that for $N \lesssim 55$, the quartic case with small $\xi$ gives too small predictions for the spectral index regardless of the value of $\alpha_{\rm eff}$, $n_s \lesssim 0.945$ (see Eq. ), which are ruled out by CMB observations. However, the large-$\xi$ case gives $n_s \lesssim 0.964$ (see Eq. ), which is compatible with observations. Note that the non-minimal coupling does not need to be very large, as we only need to require $\xi \gtrsim 1$. Therefore, for large enough $\xi$ and $N$, the Higgs-like $\phi^4$ model is generally compatible with observations. Let us then consider a case where the scale of inflation is small and where, consequently, the number of e-folds between the horizon exit of the pivot scale and the end of inflation is also small. This happens in scenarios with a large $\xi$ or $\alpha_{\rm eff}$, as then inflation takes place at small field values $\chi$ and the tensor-to-scalar ratio $r$ is small. To see how the value of $\chi$ at the CMB scales maps to $\xi$ and $\alpha_{\rm eff}$ in our model, see Fig. \[alpha\_xi\]. By choosing a large enough $\xi$ or $\alpha_{\rm eff}$, the excursion of the field during inflation $\Delta \chi$ and $r$ can be made small enough to agree with the so-called trans-Planckian censorship conjecture limits [@Bedroya:2019tba] $$\label{TCC_limits} r \lesssim 10^{-30}, \, \qquad |\Delta \chi| \lesssim 10^{-13}M_{\rm P}\,,$$ needed to ensure that no scale observed in today’s (classical) universe originated from super-Planckian energy scales during inflation. With Eqs. , and $N \sim$ 40–50, these translate to $$\label{TCC_parameter_limits} \begin{split} \alpha_{\rm eff} &\gtrsim 10^{24} \, , \quad \xi \ll 1 \, , \\ \alpha_{\rm eff}/\xi &\gtrsim 10^{26} , \quad 1 \lesssim \xi^2 \ll \alpha_{\rm eff} \, , \\ \xi &\gtrsim 10^{27} \, , \quad \alpha_{\rm eff} \ll \xi^2 . \end{split}$$ Demanding perturbativity, $\lambda < 1$, and the observed value for $A_s$ (see Eqs. and ), gives an upper limit for $\xi$: $$\label{xi_upper_limit} \xi \lesssim 10^{10} \, .$$ Thus, to satisfy the limits , $\alpha_{\rm eff}$ must be large. Note that when $r$ is very small, this starts to affect the CMB value of $N$, see Eq. , and thus also the predicted value of $n_s$. In the limit $r<10^{-30}$, we have $N < 39$ and therefore $n_s < 0.95$, which is in tension with observations even in the case of a large $\xi$. Therefore, while there is no problem with initial conditions, the Higgs-like inflation model with an $R^2$ term in Palatini gravity is in tension with observations if we also demand it to satisfy the limits . However, for a shallower potential, for instance for $\phi^{2n}$ with $n<2$ and a non-minimal coupling $\xi\phi^n$, the prediction for the spectral index becomes $n_s = 1-(1+n/2)/N$ at the limit of large $\xi$ [@Jarv:2017azx], which can be made compatible with Planck even when $N \lesssim 40$; see an example of this in Ref. [@Tenkanen:2019wsd]. Therefore, while the Higgs-like $\phi^4$ scenario is in tension with the CMB observations when requiring it to satisfy the proposed trans-Planckian censorship conjecture limits, this is not, in general, the case for scenarios belonging to the same model class. While we have not studied initial conditions for general $\phi^{2n}$ potentials in this paper, we do not expect them to differ from the Higgs-like $\phi^4$ case in scenarios where also an $R^2$ term in Palatini gravity is added to the Lagrangian. ![The canonical field value $\chi$ at CMB scales corresponding to a time when there was 50 e-folds of inflation left, $\chi_{50}$, as a function of $\xi$ and $\alpha_{\rm eff}$. Roughly, the value of $\chi_{50}$ is dictated by $\xi^2 + 8\alpha_{\rm eff}$. Note that the field excursion during inflation, $\Delta \chi$, can be much smaller than the actual field value $\chi_{50}$, especially for $\alpha_{\rm eff} \gg \xi^2$, where all inflation happens close to a constant $\chi$-value, as discussed in section \[attractor\]. []{data-label="alpha_xi"}](alpha_xi.pdf) Conclusions =========== We have studied initial conditions for inflation in scenarios where the inflaton potential has a plateau shape. Our motivation for this was two-fold: such models can be obtained in a large number of model classes and are, most importantly, among those most favored by Planck data. As a representative example, we considered a Higgs inflation model where the SM Higgs couples non-minimally to gravity, and also allowed the action to contain an $R^2$ term in the context of Palatini gravity. We showed that inflation with a large number of e-folds generically occurs in a large part of the model parameter space, without any need for fine-tuning of parameters even when the scale of inflation and the inflaton field value during inflation are much smaller than the Planck scale. As shown in the paper, the phase-space trajectories end up in the regime where slow-roll inflation eventually takes place; either because the trajectories fall into the slow-roll regime directly, or because they pass $\chi=0$, slow down, and enter slow-roll on the other side of the potential. Only the trajectories with fine-tuned initial conditions end up oscillating around the origin without yielding a significant amount of slow-roll inflation. Our findings hold generically for all Higgs-like potentials where at least one of the parameters $\xi$, $\alpha_{\rm eff}$ is large. The above statements, however, depend on the choice of the coordinates and the corresponding phase-space probability measure, and should therefore be taken in a qualitative fashion only. Likewise, our conclusions apply only in the case where inflation occurs in a homogeneous and isotropic FRW universe. Future studies are needed to extend our study to scenarios where the initial conditions may not be homogeneous nor isotropic and to quantify the probability of inflation in such cases. Finally, we discussed the compatibility of the models studied in this paper with the CMB observations and the recently proposed trans-Planckian censorship conjecture, and showed that while the latter can easily be satisfied for large enough $\alpha_{\rm eff}$ (without fine-tuning of the initial conditions), the predicted value for the spectral index $n_s$ is, in this model, in tension with the Planck data. However, if the requirement of satisfying the conjecture is lifted, the models are in perfect agreement with data. This reinforces the motivation for future studies on this topic. Acknowledgements {#acknowledgements .unnumbered} ================ TT is funded by the Simons foundation. ET acknowledges support from the Academy of Finland project 320123. [^1]: Like Ref. [@Chowdhury:2019otk], by “fine-tuned” initial conditions we mean a set of initial conditions that occupies only a small fraction of the inflaton phase-space. For discussion on the choice of a phase-space measure, see Section VII of Ref. [@Chowdhury:2019otk]. [^2]: This is an example of what is often called a Palatini or metric-affine theory; see Ref. [@Bauer:2008zj] for original work and Ref. [@Tenkanen:2020dge] for an introduction to the topic, as well as Refs. [@Bauer:2010jg; @Tamanini:2010uq; @Enqvist:2011qm; @Azri:2017uor; @Rasanen:2017ivk; @Tenkanen:2017jih; @Racioppi:2017spw; @Markkanen:2017tun; @Jarv:2017azx; @Racioppi:2018zoy; @Enckell:2018kkc; @Carrilho:2018ffi; @Aoki:2018lwx; @Enckell:2018hmo; @Antoniadis:2018ywb; @Rasanen:2018fom; @Kannike:2018zwn; @Rasanen:2018ihz; @Almeida:2018oid; @Antoniadis:2018yfq; @Shimada:2018lnm; @Takahashi:2018brt; @Jinno:2018jei; @Tenkanen:2019jiq; @Rubio:2019ypq; @Jinno:2019und; @Tenkanen:2019xzn; @Tenkanen:2019wsd; @Gialamas:2019nly; @Racioppi:2019jsp; @Shaposhnikov:2020geh] for some other recent studies on inflation in this context. [^3]: When the quantities with and without a bar are compared here, it is done with a fixed value of the Jordan frame field $\phi$. [^4]: Note that this does not depend on the form of $V(\phi)$; in general, the $\lambda$-scaling can be replaced by a scaling of $V(\phi)$.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We follow a low-energy effective theory approach to identify the general class of theories that describes a real vector field (of unconstrained norm) coupled to gravity. The resulting set may be regarded as a generalization of the conventional vector-tensor theories, and as a high-momentum completion of aether models. We study the conditions that a viable cosmology, Newtonian limit and absence of classical and quantum instabilities impose on the parameters of our class of models, and compare these constraints with those derived in previously studied and related cases. The most stringent conditions arise from the quantum stability of the theory, which allows dynamical cosmological solutions only for a non-Maxwellian kinetic term. The gravitational constant in the Newtonian limit turns to be scale dependent, suggesting connections to dark matter and degravitation. This class of theories has a very rich gravitational phenomenology, and offers an ample but simple testing ground to study modifications of gravity and their cosmological implications.' author: - 'Cristian Armendariz-Picon' - 'Alberto Diez-Tejedor' title: Aether Unleashed --- Introduction ============ The symmetries of a spacetime severely constrain the transformation properties of its matter content. In Minkowski spacetime for instance Lorentz symmetry allows scalar fields to have a non-vanishing expectation value, forbidding non-scalar fields from being non-zero. It is perhaps because of these constraints that most cosmologists have focused on cosmological solutions with scalar fields. Nevertheless, in an expanding universe no symmetry principle prevents the existence of non-vanishing tensor fields, as long as they are constant in space and invariant under *spatial* rotations. Since only bosonic fields can acquire large occupation numbers, and if we restrict ourselves to fields of spin less than two, this leaves vector fields (or one-forms) as essentially the only additional type of field that may be relevant cosmologically. If the vector field only couples to gravity, the presence of a preferred Lorentz frame will remain unobservable in the matter sector, while, of course, Lorentz symmetry is broken anyway by the expansion of the universe and is not of particular relevance on large scales. Massless vector-tensor theories were considered long time ago to explore different modifications of gravity, and as alternatives to the more prevalent scalar-tensor theories at that time [@Will:1972zz; @Hellings:1973zz]. Indeed, as we shall see, vector-tensor theories have a much richer gravitational phenomenology than scalar-tensor theories, and thus offer a wider framework to study departures from general relativity. Theories with vectors have also experienced a modest revival in connection with inflation and the accelerated expansion of the universe [@Ford:1989me; @ArmendarizPicon:2004pm; @Wei:2006tn; @Kanno:2006ty; @Novello:2006ng; @Boehmer:2007qa; @Koivisto:2007bp; @Jimenez:2008au; @Koivisto:2008ig; @Golovnev:2008cf; @Koivisto:2008xf; @Yokoyama:2008xw; @Kanno:2008gn; @Watanabe:2009ct; @Koh:2009ne] and studies of the spontaneous breaking of Lorentz invariance in the so-called Einstein-aether models of Jacobson and Mattingly [@Jacobson:2000xp] (see [@Gasperini:1987nq] for earlier discussions, and [@Jacobson:2008aj] for a status report.) In this article we study a class of theories that somewhat differs from previous models. We explore the most general vector-tensor theory expected to describe our universe at low energies, large distances and long timescales. We thus follow a low energy effective action approach and consider all generally covariant terms with the lowest number of vector field derivatives. Since we deal with massive vectors, a consistent quantum description at low energies does not require gauge invariance, so many different kinetic terms for the vector field are possible. Conversely, if we do not assume gauge invariance, no symmetry principle prevents the appearance of mass terms for the vector, or vector self-interactions for that matter. Our effective action thus contains all generally covariant terms with two vector field derivatives and a general quartic potential[^1] $V$ that general covariance forces to depend on the squared norm of the vector, $A_\mu A^\mu$. Up to a constant term, any scalar quartic potential can be written in the form $$\label{eq:potential} V(A_\mu A^\mu)=\lambda (A_\mu A^\mu+ M^2)^2,$$ where $M$ is a mass scale and $\lambda$ a dimensionless parameter. In fact, around an extremum, any potential can be locally approximated by (\[eq:potential\]). Such a dependence opens up a quite interesting possibility, because the vacuum of the theory has a non-vanishing vector field $A_\mu A^\mu=-M^2$, and thus in flat space Lorentz symmetry is spontaneously broken.[^2] In fact, in some modern approaches to quantum gravity Lorentz invariance is expected to be broken at high energies [@Kostelecky:1988zi; @Horava:2009uw]. However, if broken, it should be done in a very subtle way. If the breaking relies in the existence of a fundamental preferred frame at the Planck-scale, some unnatural fine tuning issues appear [@Collins:2004bp], or additional ill-behaved modes appear in the spectrum [@Charmousis:2009tc; @Blas:2009yd]. Moreover, a fixed external structure goes drastically against our currently accepted principles (i.e. general covariance). This suggests that, probably, the more reasonable way in which Lorentz invariance could be broken is by the existence of a vacuum expectation value (vev) for some non-scalar field, which, as in the ghost condensate [@ArkaniHamed:2003uy], may be the gradient of a scalar. A widely studied class of models in which Lorentz invariance is spontaneously broken involves the Einstein-aether models introduced by Jacobson and Mattingly in [@Jacobson:2000xp]. One of the most distinctive properties of these models is that the norm of the vector field is constrained to have a constant value. This constraint may appear artificial, but it could be naively justified as the limit in which the coupling constant $\lambda$ in the potential of equation (\[eq:potential\]) tends to infinity. Hence, aether models may be regarded as a particular case of the class of theories we consider here. Because we drop the fixed-norm constraint we call this class of theories “unleashed aether" models. Similar kind of actions have been considered for instance in [@Bluhm:2004ep; @Bailey:2006fd; @Bluhm:2007bd; @Kostelecky:2009zr] under the name of “bumblebee models." In general, the latter are models in which a vector field acquires a vev due to the presence of a potential. In that sense, the unleashed aether is a bumblebee model. The real difference between the unleashed aether and the most commonly studied bumblebee models lies in the parameters in our action that are assumed to be zero in the bumblebee case. The main goal of this article is to study the conditions that self-consistency and phenomenological viability imposes on the parameters of our class of theories, and to determine to what extent these differ from massless vector fields and aether models. In Section \[sec:generalities\] we present and discuss general properties of our class of models. In Section \[sec:cosmology\] we study the properties of cosmological solutions at early and late times, and derive basic constraints on the value of the scale $M$ in the potential. In Section \[sec:PPN\] we study the Newtonian limit of these theories, and explore their relation with massless theories and aether models. Finally, in Section \[sec:stability\], we derive the conditions that classical and quantum-mechanical stability places on the parameter space of unleashed aether models. Generalities {#sec:generalities} ============ We want to identify the most general action that describes the dynamics of a real vector field $A^\mu$ coupled to gravity. Without additional restrictions this action will contain all possible terms compatible with the symmetries of the theory, which we assume to consist just of general covariance and a $\mathbb{Z}_2$ symmetry $A^\mu\to -A^\mu$. If we are interested in low energies however, we expect only those terms with the least number of derivatives to be relevant in the limit of long distances and times. We can thus concentrate on the most general generally covariant action with at most two derivatives acting on the metric and the vector. In principle, such an action would also contain arbitrary powers of the vector field, but since a vector has dimensions of mass, terms with too many powers of $A^\mu$ must be suppressed by corresponding powers of an energy scale $\Lambda$. If the background value of the field is smaller than $\Lambda$, the dominant terms are hence given by those with no suppression, that is, by the operators of dimension four. Therefore, we just have to consider the most general action with two derivatives acting on the metric and vector field, but with no vector field operators of dimension higher than four: $$\begin{aligned} \label{eq:action} S=\int d^{4}x \, \sqrt{-g}\Bigg[\frac{R}{16\pi G} -\frac{\beta_1}{2} F_{\mu\nu} F^{\mu\nu}-\beta(\nabla_{\mu}A^{\mu})^{2} &+&\beta_{13}R_{\mu\nu}A^{\mu}A^{\nu}+\nonumber \\ &+&\beta_4 R A_\mu A^\mu-V(A_\mu A^\mu)\Bigg],\end{aligned}$$ where the $\beta_i$ are dimensionless coefficients, $F_{\mu\nu}\equiv \partial_\mu A_\nu-\partial_\nu A_\mu$, and $V$ is a quartic potential of the form (\[eq:potential\]), which we plot in Figure \[fig:potential\]. This is the most general action with two derivatives acting on $A^\mu$ because a term proportional to $\nabla_\mu A_\nu \nabla^\nu A^\mu$ can be eliminated integrating by parts and using the identity ${[\nabla_\mu,\nabla_\nu]A^\rho=R_{\mu\nu}{}^\rho{}_\sigma A^\sigma}$. The only dimension four operator with one vector derivative, $A_\mu A^\mu \nabla_\nu A^\nu$, is excluded by the discrete symmetry $A^\mu\to -A^\mu$. Observe that we do not assume gauge invariance, which would only allow the Maxwell term in the vector sector. Although gauge invariance seems to be necessary for a consistent description of *massless* vector field interactions [@Weinberg:gauge], it is not required for massive vector fields, at least at sufficiently low energies. The stability of the theory and phenomenological considerations then set constraints on the values of the so far arbitrary coefficients $\beta_i$ and the derivatives of the potential $V$. When we explore these constraints in Section \[sec:stability\], we shall see that the term proportional to the Maxwell Lagrangian describes a vector field (under Lorentz-transformations), whereas the term proportional to the divergence squared of the field actually introduces an additional scalar (under Lorentz transformations). In that sense, it is debatable whether this is actually just a vector field Lagrangian. Of course, we should also add to the action (\[eq:action\]) all the matter terms of the standard model of particle physics. A general vector field action would contain couplings of the vector field $A^\mu$ to this sector, as in bumblebee models. Such “Lorentz-violating" interactions have been thoroughly considered and constrained in the Standard Model Extension program initiated by Kostelecký (see e.g. [@Colladay:1998fq; @Kostelecky:2008ts]). In this article we assume that such couplings are negligible. Since interactions between the vector field and the matter sector would be mediated by gravitation, it is consistent to assume that they are small. For a review of modern tests of Lorentz invariance, see [@Mattingly:2005re]. The validity of the effective theory proposed here requires that the background value of the vector field components satisfy ${|A^\mu| \ll \Lambda}$. This is the assumption that allows us to neglect terms in the Lagrangian such as, for instance, $$\mathcal{L}\supset \frac{\gamma}{\Lambda^2}A^\mu A^\nu \nabla_\mu A^\rho \nabla_\nu A_\rho.$$ Because for a potential like $(\ref{eq:potential})$ we expect the field to be in the vicinity of ${A_\mu A^\mu=-M^2}$, this implies that $M^2\ll \Lambda^2$. This raises a mild fine-tuning reminiscent of the hierarchy problem—why is the Higgs vev much smaller than the Planck scale. Up to an overall constant, a generic potential would be of the form $$V(A_\mu A^\mu)=2\lambda_1\Lambda^2 A_\mu A^\mu+\lambda_2 (A_\mu A^\mu)^2+\frac{\lambda_3}{\Lambda^2} (A_\mu A^\mu)^3+\cdots,$$ with dimensionless coefficients $\lambda_i$ of order one. Comparing with equation (\[eq:potential\]) we find that $\lambda=\lambda_2$ and $M^2=(\lambda_1/\lambda_2) \Lambda^2$. Thus, $M^2\ll \Lambda^2$ requires $\lambda_1\ll \lambda_2$. This fine-tuning does not have to be severe though. It suffices for instance that ${\lambda_1/\lambda_2\equiv M^2/\Lambda^2<10^{-2}}$ in order for our expression to be relatively accurate. On a more practical side, the restriction $M<\Lambda$ is almost inescapable: If $M$ were of the order of $\Lambda$, we would have to include all powers of $A^\mu$ in our effective theory. Incidentally, a very small value of $\lambda$ can be easily justified if $\beta_{13}=\beta_4=0$. In that case, as we discuss in Section \[sec:stability\], the theory has additional symmetries in the limit $\lambda\to 0$. For a review about effective field theories see [@Burgess:2007pt]. (220,200) ![A plot of the potential (\[eq:potential\]) for $M^2>0$ and $\lambda>0$. Because of the indefinite signature of spacetime, the potential does not have the shape of a Mexican hat. Instead, around the two local minima along the $A_0$ direction, the potential looks like a horse saddle and thus has negative Gaussian curvature. If $M^2<0$ the roles of $A_0$ and $A_i$ are reversed.[]{data-label="fig:potential"}](figura1.pdf "fig:"){height="8cm"} (-200,18)[$A_{i}$]{} (-33,50)[$A_{0}$]{} (-296,100)[$V$]{} \[\] The action (\[eq:action\]) is to some extent a generalization of previously considered theories. Setting $V\equiv 0$ one recovers the class of vector-tensor theories studied in [@Will:1981cz], while replacing the potential (\[eq:potential\]) by a fixed-norm constraint leads to a subclass of the Einstein-aether models considered in [@Jacobson:2000xp], and the vector field theories studied by Carroll and Lim [@Carroll:2004ai].[^3] Although a fixed normed constraint seems to be difficult to justify from a particle physics perspective, general potentials $V(A_\mu A^\mu)$ have been argued to appear in theories in which gauge invariance is spontaneously broken [@Bjorken:1963vg; @Kraus:2002sa; @Ford:1989me], and a fixed-norm constraint can be naively recovered from a potential of the form (\[eq:potential\]), at least classically, by taking the limit $\lambda\to \infty$ (we shall make this connection more explicit below.) We are ready now to explore some of the physical consequences of a vector field described by the action (\[eq:action\]). Varying with respect to the field $A^\mu$ leads to the equations of motion $$\label{eq:motionA} \beta_1 \nabla_\nu F^\nu{}_\mu+\beta \nabla_\mu (\nabla_\nu A^\nu)+ \beta_{13} R_{\mu\nu} A^\nu+\beta_4 R A_\mu-\frac{dV}{dA^2}A_\mu=0.$$ Because the covariant derivative and the metric commute, the equations we obtain by varying with respect to $A^\mu$ are equivalent to the ones we get varying with respect to $A_\mu$. In fact, because the transition from a vector to a form involves just a field redefinition, ${A^\mu\to A_\mu=g_{\mu\nu} A^\nu}$, assuming that the action (\[eq:action\]) is a functional of a vector $A^\mu$ or a form $A_\mu$ leads to physically equivalent results both classically and at the quantum level, even though, curiously, the corresponding expressions for the energy-momentum tensors $T_{\mu\nu}\equiv-(2/\sqrt{-g})(\delta S_A/\delta g^{\mu\nu})$ have different forms: If the action is taken to be a functional of a form $A_\mu$, the energy-momentum tensor is $$\begin{aligned} \label{eq:EMT} T^\mathrm{form}_{\mu\nu}&=& 2\beta_1 F_{\mu\rho}F_\nu{}^\rho +2\,\beta \left[\left(A^\rho\nabla_\rho(\nabla_\sigma A^\sigma)+ (\nabla_\rho A^\rho)^2\right) g_{\mu\nu}-2A_{(\mu} \nabla_{\nu)}(\nabla_\rho A^\rho)\right]+\nonumber \\ &+&\beta_{13}\left[2\nabla_\rho \nabla_{(\mu}(A_{\nu)}A^\rho)-\nabla_\rho\nabla_\sigma(A^\rho A^\sigma) g_{\mu\nu}-\Box(A_\mu A_\nu)-4 A^\rho R_{\rho(\mu}A_{\nu)}\right]+\nonumber \\ &-&2\beta_4 \left[R_{\mu\nu}A_\rho A^\rho+R A_\mu A_\nu-\nabla_\mu\nabla_\nu(A_\rho A^\rho)+ \Box (A_\rho A^\rho)g_{\mu\nu}\right]+ \nonumber \\ &+&2\frac{dV}{dA^2}A_\mu A_\nu+ \mathcal{L} \, g_{\mu\nu},\end{aligned}$$ while if the action is a functional of a vector $A^\mu$, the energy momentum tensor turns out to be $$\begin{aligned} T^{\text{vect}}_{\mu\nu}&=& T^\mathrm{form}_{\mu\nu} +4\beta_1 (\nabla_\rho F^\rho{}_{(\mu})A_{\nu)} +4\,\beta A_{(\mu} \nabla_{\nu)}(\nabla_\rho A^\rho) +4\beta_{13}A^\rho R_{\rho(\mu}A_{\nu)}+ \nonumber \\ &{}&+4\beta_4R A_\mu A_\nu-4\frac{dV}{dA^2}A_\mu A_\nu.\end{aligned}$$ The additional terms in the energy-momentum tensor of a vector are proportional to the equation of motion (\[eq:motionA\]), and thus vanish identically. In the following, we work with the energy-momentum tensor of a form, equation (\[eq:EMT\]). Note that $dA^2$ stands for $d(A_\mu A^\mu)$. Cosmological Solutions {#sec:cosmology} ====================== On large scales, our universe seems to be well described by a (spatially flat) Friedmann-Robertson-Walker (FRW) spacetime, $$\label{eq:FRW} ds^2=-dt^2+a^2(t) \, d\vec{x}\,^2.$$ Its symmetries, homogeneity and isotropy, already constrain the possible configurations of a non-vanishing cosmic vector field quite strongly. Homogeneity (invariance under spatial translations) implies that a cosmic vector field must be spatially constant, while isotropy (invariance under spatial rotations) implies that its spatial components must vanish.[^4] Hence, a cosmic vector field must be of the form $$\label{eq:A} A^\mu=(u(t),0,0,0).$$ Because of homogeneity and isotropy, the spatial components of the vector field (\[eq:A\]) satisfy the equations of motion (\[eq:motionA\]), while the $\mu=0$ component of equation (\[eq:motionA\]) dictates that $u$ has to obey $$\label{eq:motion} \beta \left(\ddot{u}+3H\dot{u}\right)+ \left[V'-3(\beta_{13}+4\beta_4)H^2+3(\beta-\beta_{13}-2\beta_4)\dot{H}\right]u=0,$$ where a dot denotes a derivative with respect to cosmic time $t$, a prime a derivative with respect to $A_\mu A^\mu$, and $H\equiv \dot{a}/a$ is Hubble’s constant. It follows automatically from equation (\[eq:motion\]) that if $\beta=0$, $u$ is non-dynamical. Indeed, in that case the “kinetic" term of the vector is the one of electromagnetism, which does not contain any time derivatives of $A_0\equiv -u$. This is why there are no terms proportional to $\beta_1$ in the equation of motion for $u$. In aether models, the fixed norm constraint simply demands that the value of $u$ be constant. The energy-momentum tensor (\[eq:EMT\]) also explains why isotropy requires that the vector field have zero spatial components. Otherwise, say, the term proportional to $dV/dA^2$ would yield an energy-momentum tensor which would not be invariant under spatial rotations, and the FRW metric (\[eq:FRW\]) would not be a solution of Einstein’s equations. Certainly, if the vector field is subdominant, it is not strictly necessary that $A^\mu$ satisfy equation (\[eq:A\]) in order for equation (\[eq:FRW\]) to be an *approximate* solution of the Einstein equations. Ultimately, equation (\[eq:A\]) should be justified dynamically, though we shall not explore this issue here. If the vector does point in the time direction, the energy density $\rho\equiv -T^0{}_0$ and pressure $p\, \delta^i_j \equiv T^i{}_j$ are given, from (\[eq:EMT\]), by $$\begin{aligned} \rho&=& \beta (2u\ddot{u}-\dot{u}^2) +6(\beta_{13}+2\beta_4)Hu \dot{u}+ \left[2V'+6(\beta-\beta_{13}-2\beta_4)\dot{H}-9(\beta+2\beta_4)H^2\right]u^2+V, \nonumber \\ \\ p&=&2(\beta-\beta_{13}-2\beta_4)(u\ddot{u}+\dot{u}^2)-\beta \dot{u}^2 +(3\beta-2\beta_{13}-2\beta_4)\left[4 H u \dot{u}+(2\dot{H}+3H^2)u^2\right]-V.\nonumber \\ \label{eq:pressure}\end{aligned}$$ All the non-diagonal components of the energy-momentum tensor vanish, as demanded by isotropy. Because the Maxwell term does not contain $u(t)$, the energy density and pressure do not have factors proportional to $\beta_1$. It will be useful to work with a simplified form of the energy density. The equation of motion (\[eq:motion\]) can be used to eliminate the term proportional to $\ddot{u}$ in $\rho$, leading to the simpler expression $$\rho=-\beta\dot{u}^2-6(\beta-\beta_{13}-2\beta_4)H u \dot{u}-3(3\beta-2\beta_{13}-2\beta_4)H^2 u^2+V. \label{eq:energy}$$ Using the equation of motion again, it can be easily shown that the energy density (\[eq:energy\]) satisfies the conservation equation $\dot{\rho}+3H(\rho+p)=0$, which is just an expression of the conservation law $\nabla_\mu T^\mu{}_\nu=0$. Note that $\rho$ may become negative for certain values of the parameters. In principle one should not be scared by the possible appearance of negative energies, as long as the Hamiltonian has a local minimum. We explore the classical and quantum stability of this class of theories in Section \[sec:stability\]. The executive summary is: Unleashed aether theories are stable only if $\beta=0$ or $\beta_1=0$. Since cosmological solutions with $\beta=0$ have no dynamics, we shall thus set $\beta_1=0$.[^5] In any case, this does not have any effect on our equations, because, as mentioned above, $\beta_1$ does not appear in the field equation of motion nor the energy-momentum tensor. For $\beta_{1}\neq 0\neq\beta$ the theory is unstable, at least in the ultraviolet. However, by an appropriate choice of the parameters we can get the dangerous modes out of the spectrum, recovering a sensible low energy effective field theory: the aether model. Solutions at early times ------------------------ We now turn our attention to the solutions of the vector field equation and its cosmological implications. The equation of motion (\[eq:motion\]) resembles that of a scalar field, with additional contributions to its effective mass stemming from the expansion of the universe. We shall study the case in which this mass is dominated by the expansion, $H^2\gg V'$, which we generally expect to happen at early times, and also the case in which the mass is dominated by the potential term, $H^2\ll V',$ which we typically expect to happen at late times. We take the expansion history as given, which in general requires the vector field to be subdominant. In some cases, if the equation of state of the vector field is compatible with the assumed expansion of the universe, our solutions also apply to periods in which the vector field is the dominant component in the universe. In the limit $H^2\gg V'$, the equation of motion of the field takes the form $$\ddot{u}+3H\dot{u}+3\left(\frac{\beta-\beta_{13}-2\beta_4}{\beta} \dot{H} -\frac{\beta_{13}+4 \beta_4}{\beta} H^2\right)u=0.$$ The behavior of the solution crucially depends on the size of the term in parenthesis, which is typically of order $H^2$. To find explicit solutions it is convenient and realistic to assume that the universe expands as a power law, $a\propto t^p$, which leads to the exact solution $$\begin{aligned} \label{eq:ut} u(t)&=&C_1 \, t^{(1-3 p+f)/2}+C_2\, t^{(1-3p-f)/2}, \quad \text{where}\\ f&=&\sqrt{(1+3p)^2+\frac{12p}{\beta}\left[\beta_{13}(p-1)+2\beta_4(2p-1)\right]},\end{aligned}$$ and $C_1$ and $C_2$ are two integration constants. To be more specific, let us consider for example radiation domination, $p=1/2$. Keeping the dominant mode only we get $$\label{eq:early evolution} u(t)\approx C_1 \, t^{(q-1)/4}, \quad \text{where}\quad q\equiv \sqrt{25-12\beta_{13}/\beta}.$$ Thus, the field grows for $q>1$, decays for $q<1$, and oscillates logarithmically with decaying amplitude if $q$ is purely imaginary. Its equation of state can then be calculated by substituting (\[eq:ut\]) into equations (\[eq:energy\]) and (\[eq:pressure\]), though we shall not do so, as the resulting expression is not particularly illuminating. A quite interesting dark energy model in which a massless vector field evolves at early times following equation (\[eq:early evolution\]) has been recently proposed in [@Jimenez:2008au]. Solutions at intermediate times ------------------------------- Generally we expect the condition $H^2\gg V'$ to break down at late times, either because $H$ steadily decreases, or because the field evolves to regions of larger potential slope as the universe expands. In the limit $H^2\ll V'$, the equation of motion (\[eq:motion\]) resembles that of a minimally coupled scalar field, $$\ddot{u}+3H \dot{u}+\frac{V'}{\beta}u=0.$$ Thus, even if the field is initially frozen, during a small time $\Delta t$ it changes by an amount $$\label{eq:roll} \frac{\Delta u}{u} \approx -\frac{V'}{2\beta} \Delta t^2.$$ It follows that during a Hubble time $\Delta t=H^{-1}$ the relative change of the field is of order $V'/(\beta H^2)$. If the latter is large, the field rolls down until the potential slope is sufficiently small. Solutions at late times {#sec:extremum} ----------------------- If the potential $V$ has the form (\[eq:potential\]), the equations of motion in the absence of matter have flat space and $\bar{A}^\mu=(M,0,0,0)$ as a solution. We would speak of this configuration as the (classical) “vacuum" of the theory. In an expanding universe however, $\bar{A}^\mu=(M,0,0,0)$ does not solve equation (\[eq:motion\]), unless spacetime is de Sitter and $\beta_{13}=-4\beta_4$. To study the properties of cosmological solutions around the vacuum, let us substitute into equation (\[eq:motion\]) the ansatz $u=M+\delta u$, with $\delta u\ll M$ and $V$ given by equation (\[eq:potential\]). Keeping only the leading terms we get $$\label{eq:linearized cosmo} \beta [\delta\ddot{u}+3 H \delta\dot{u}] -4\lambda M^2 \delta u= 3M[(\beta_{13}+4\beta_4) H^2-(\beta-\beta_{13}-2\beta_4) \dot{H}].$$ The configuration $u=M$ is stable if $-4\lambda M^2/ \beta>0$, which is what we assume in the following. In this case, the equation describes a damped harmonic oscillator with a time-dependent external force. To solve the equation we shall consider the limit of large masses or times, that is $H/M\ll 1$ (we assume that $\dot{H}\approx H^2$.) In this limit, one can think of the driving force as being constant, so an approximate particular solution of the inhomogeneous equation is $$\label{eq:adiabatic} \delta u_\mathrm{ad} =\frac{3}{4}\frac{(\beta-\beta_{13}-2\beta_4) \dot{H}-(\beta_{13}+4\beta_4) H^2}{\lambda M^2} M.$$ This is the leading correction to the solution $u=M$ in powers of the adiabaticity parameter $H/M$. Substituting equation (\[eq:adiabatic\]) back into (\[eq:linearized cosmo\]) one can indeed verify that corrections to this solution are suppressed by two additional powers of $H/M$. The derivatives of this adiabatic solution are also suppressed by powers of $H/M$. Not coincidentally, in the limit $H/M\ll 1$, the inequality $-4\lambda M^2/ \beta>0$ follows from one of the conditions for classical stability at low momenta that we derive in Section \[sec:stability aether\]. Having found a particular solution of the inhomogeneous equation, we can come back to the general solution of the homogeneous equation, $$\label{eq:oscillations} \delta u_\mathrm{osc}\approx \frac{B}{a^{3/2}} \cos\left(2 \sqrt{-\frac{\lambda}{\beta}} M t+\varphi \right),$$ where $B$ is an arbitrary amplitude and $\varphi$ an arbitrary phase. During a Hubble time the field performs $ \sqrt{-\lambda/\beta} M/H\gg 1$ oscillations, while its amplitude remains essentially constant. We have seen that the full solution of equation (\[eq:linearized cosmo\]) has the form $$\label{eq:full} u=u_\mathrm{ad}+\delta u_\mathrm{osc},$$ where $\delta u_\mathrm{osc}$ is given by equation (\[eq:oscillations\]), and, to lowest order, $u_\mathrm{ad}\approx M$. In order to find the energy density and pressure associated with the full solution (\[eq:full\]), we substitute it into equations (\[eq:energy\]) and (\[eq:pressure\]). Because we are interested on cosmological timescales, we consider time averages of $\rho$ and $p$ over Hubble times. Among the terms quadratic in $u_\mathrm{ad}$, those without derivatives dominate. Terms that involve mixed terms $u_\mathrm{ad} \, \delta u_\mathrm{osc}$ average to zero, while the averages of terms quadratic in $\delta u_\mathrm{osc}$ give non-zero contributions proportional to the square of the amplitude $B$. We can thus split the contributions to the energy density in an adiabatic component, and an oscillatory one. To leading order, the energy density and pressure of the adiabatic solution are, $$\begin{aligned} \rho_\mathrm{ad}&=&-3 (3\beta-2\beta_{13}-2\beta_4) H^2 M^2 ,\\ p_\mathrm{ad}&=& (3\beta-2\beta_{13}-2\beta_4) (3 H^2+2 \dot{H}) M^2.\end{aligned}$$ The subleading terms are suppressed at least by a factor of order $H^2/M^2$, and, hence, their contributions to $\rho$ and $p$, of order $H^4$, are negligible. Because the Friedmann equations in the presence of matter can be cast as $$3 H^2= 8 \pi G_\mathrm{c} \rho_m \quad \text{and} \quad -(3H^2+2\dot{H})=8\pi G_\mathrm{c} p_m,$$ an observer in such a universe trying to determine $G_\mathrm{c}$ through the expansion of the universe would find a renormalized value of Newton’s constant, $$\label{eq:cosmic G} G_\mathrm{c}=\frac{G}{1+8\pi G M^2 (3\beta-2\beta_{13}-2\beta_4)}.$$ As we discuss in Section \[sec:stability aether\], stability requires that $3\beta-2\beta_{13}>0$. Exactly the same renormalization occurs in aether models [@Carroll:2004ai]. We show in Section \[sec:PPN\] that for generic parameter values $G_\mathrm{c}$ also differs from the gravitational constant measured in “local" solar system measurements $\mathrm{G}_\mathrm{N}$. Following the analysis of [@Carroll:2004ai], one can then derive a not too stringent order of magnitude limit on the scale $M$ based on nucleosynthesis constraints, $$\label{eq:looseBBNbound} 8 \pi G M^2 \lesssim 1.$$ Since our analysis assumes that $M<\Lambda$, and presumably $\Lambda$ is smaller than the scale of quantum gravity $M_G\equiv (8\pi G)^{-1/2}$, this is also a necessary condition for the validity of our approximations. Hence, we typically expect $G_c\approx G$. Nevertheless, for some parameter choices $G$ and $G_c$ can differ significantly. If $8\pi G M^2 (3\beta-2\beta_{13}-2\beta_4)$ is fine-tuned to be close to $-1$, the Planck scale one would derive from $G_c$ would be much lower than the actual quantum gravity scale $M_G$. On the other hand, if $8\pi GM^2 \gg 1$, the actual scale of quantum gravity would be much lower than the one guessed from the value of $G_c$. Of course, this would require $\Lambda\gg M_G$. The density and pressure of the oscillating solutions are dominated by those terms with the maximum number of derivatives acting on $u$. Thus, substituting equation (\[eq:oscillations\]) into (\[eq:energy\]) and (\[eq:pressure\]) and averaging over time we find that the oscillating vector field behaves like non-relativistic matter, $$\rho_\mathrm{osc}\approx 4\lambda M^2 \frac{B^2}{a^3} \quad \text{and} \text \quad p_\mathrm{osc}\approx 0.$$ The effective pressure is consistent with the scale factor dependence of the energy density. Since we assume that the vector field only couples to gravity, there is no additional decay channel through which the vector field could dissipate its energy (if $\beta_{1}=0$ the spectrum of the model consists of a single scalar.) Thus, an oscillating vector field is a potential dark matter candidate, somewhat similar to the axion, though we would have to study how structure develops to confirm that the vector actually mimics dark matter. These oscillations are absent in aether models, because the constraint on the vector field norm only allows the lowest order adiabatic component. Note that in the limit in which $\lambda\to \infty$, for fixed $M$, the adiabatic approximation $u_\mathrm{ad}\approx M$ becomes exact, and the energy density of the oscillating solution diverges, unless $B=0$. The energy density of the oscillating field depends on the value of the amplitude $B$, and therefore, constraints on the scale $M$ based on this solution depend on initial conditions. Nevertheless, based on the adiabatic solution (\[eq:adiabatic\]), which does not depend on initial conditions, it appears reasonable to assume that when $H\approx M$, the amplitude of the oscillations was $(B/a_M)\approx M/\lambda$. Under this assumption, demanding that the energy density of the oscillating field during nucleosynthesis be no more than the total energy density we arrive at the bound $$M\lesssim ( \lambda^2 M_P^3\, T^2_\mathrm{nucl})^{1/5}\sim \lambda^{2/5}\, 10^{10}\, \textrm{GeV},$$ where $T_\mathrm{nucl}\sim \textrm{MeV}$ is the temperature during nucleosynthesis, and $M_P\approx 10^{18}\, \mathrm{GeV}$ is the reduced Planck mass. For $\lambda$ of order one this limit improves the bound (\[eq:looseBBNbound\]) by eight orders of magnitude. Newtonian Limit {#sec:PPN} =============== The presence of the vector field not only affects gravity on cosmological scales, but also locally. In a massless vector field theory for instance, the vector field determines the gravitational constant that enters the Newtonian limit of the theory, and the post-Newtonian parameters of the theory generically differ from those of general relativity [@Will:1981cz]. In a massive theory, one would expect the effects of the vector field to be negligible at sufficiently large scales, but it turns out that, even on large scales, the vector field leaves its imprint on the effective gravitational constant that appears in the Poisson equation. In this section we study some aspects of the Newtonian and Post-Newtonian limits of the unleashed aether. A general discussion of the Post-Newtonian limit in the gravitational sector of Kostelecký’s standard model extension [@Colladay:1998fq; @Kostelecky:2008ts] can be found in [@Bailey:2006fd]. We would like to determine the gravitational field created by a non-relativistic and static source sitting in otherwise empty space. For a potential like (\[eq:potential\]) the equations of motion in the absence of matter have Minkowski spacetime as a solution, $$\label{eq:background solution} ds^2=-dt^2+d\vec{x}^2, \quad \bar{A}^\mu=(M,0,0,0).$$ We expect the presence of a sufficiently small and static matter source, with $T_{\mu\nu}=\rho_m \delta_{\mu 0} \delta_{\nu 0}$, to modify this solution slightly, so we can use first order perturbation theory to calculate its gravitational field. Because the energy density $\rho_m$ is a scalar under spatial rotations, in this limit it suffices to restrict our attention to the scalar sector. Let us hence perturb $$ds^2=-(1+2\Phi) dt^2+(1-2\Psi)d\vec{x}\,^2, \quad A^\mu=\bar{A}^\mu+(u,\partial_i s),$$ and solve the equations of motion to first order in the perturbations. If we were interested, say, in gravitational radiation we would have to study the vector and tensor sectors as well. In Fourier space the $\mu=0$ and $\mu=i$ linearized field equations (\[eq:motionA\]) are $$\begin{aligned} (\beta_1 k^2-4\lambda M^2)(u+M \Phi)&=&(\beta_{13}-\beta_1-2\beta_4) k^2 M\Phi -4 \beta_4 k^2 M (\Psi-\Phi), \label{eq:linearized} \\ \beta k^2 k_i s&=&0 \label{eq:linearized i}.\end{aligned}$$ Because we are interested in solutions that vanish at infinity, it follows from equation (\[eq:linearized i\]) that $s=0$. As we discuss in Section \[sec:stability aether\], stability in this background requires either $\beta_1=0$, or $\beta_1>0$ and $\lambda<0$. Hence, the relative signs of the operator multiplying $u+M \Phi$ in equation (\[eq:linearized\]) are correct. The same equation can be combined with the linearized $i-j$ Einstein equations $$\label{eq:linearized ij} (\Phi-\Psi)(1-16\pi G M^2 \beta_4)=32\pi G \beta_4 M(u+M\Phi),$$ to solve for the value of Eddington parameter $\gamma$, implicitly defined by the relation $\Psi=\gamma \Phi$. If $\beta_4=0$, $\gamma=1$ follows directly from (\[eq:linearized ij\]), but in the general case the solution is not that simple, $$\label{eq:gamma scale} \gamma=\frac{(\beta_1 k^2 -4\lambda M^2)(1-16\pi G M^2 \beta_4)-32\pi GM^2k^2\beta_4(\beta_{13}-\beta_1+2\beta_4)}{(\beta_1 k^2 -4\lambda M^2)(1-16\pi G M^2 \beta_4)-128\pi G M^2 k^2\beta_4^2}.$$ Inspection of equation (\[eq:gamma scale\]) reveals that the Eddington parameter is scale-dependent, and that the scale in the problem is set by the squared mass of the scalar, $-\lambda M^2$. Therefore, for coefficients $\beta_i$ of order one there are two relevant regimes: $k^2 \ll |\lambda M^2|$ (large scales), and $k^2 \gg |\lambda M^2|$ (small scales). The value of $\gamma$ in both regimes is $$\label{eq:gamma} \gamma= \left\{ \begin{array}{cr} \displaystyle 1 & \text{large scales},\\ \displaystyle \frac{\beta_1+16\pi G M^2 \beta_4(\beta_1-2\beta_{13} -4\beta_4)}{\beta_1-16\pi G M^2\beta_4(\beta_1+8\beta_4)} &\text{small scales}. \end{array} \right.$$ On short scales the value of $\gamma$ agrees with the one calculated for massless vectors in [@Will:1981cz], while on large scales it approaches the value in aether models [@Eling:2003rd; @Foster:2005dk] and general relativity $\gamma=1$ (the case $\beta_1=\beta_4=0$ requires special consideration, and is analyzed below.) At present, the best measurement of $\gamma$ comes from Doppler tracking of the Cassini spacecraft [@Bertotti:2003rm], which gives $\gamma-1=(2.1\pm 2.3)\times 10^{-5}$. In the limit $\lambda\to \infty$, all lengths fall in the long wavelength regime. It is also instructive to study how $u$ is related to the potential $\Phi$. Combining again equations (\[eq:linearized\]) and (\[eq:linearized ij\]) we can solve for $u$ in terms of $\Phi$. To keep the expression manageable, let us write down the solution in terms of the linear perturbation of a quantity proportional to the perturbation of the squared norm $A_\mu A^\mu$, $$\label{eq:delta AA} u+M\Phi=\frac{ (\beta_1-\beta_{13}+2\beta_4) \left(16 \pi G M^2 \beta_4-1\right)} {k^2[\beta_1-16 \pi G M^2 \beta_4 (\beta_1+8 \beta_4)] -4 \lambda M^2(1-16\pi G M^2\beta4)} M k^2 \Phi.$$ In the massive case the vector field perturbation $u$ is still non-negligible even far away from the matter source. In fact, at long distances, $u$ approaches $-M\Phi$, which typically decays as $1/r$, where $r$ is the distance to the source. This is quite different from a massive scalar-tensor theory, in which to first order the scalar field perturbation $\delta\varphi$ satisfies the equation $(k^2+m^2)\delta \varphi \propto \rho_m$, and thus vanishes exponentially far away from matter, $\delta \phi\propto \exp(-m r)/r$. Clearly, the origin of the difference is that $u+M\Phi$ plays the role of the short-ranged scalar $\delta\varphi$. Once we know how $\Psi$ and $u$ are related to $\Phi$ we can proceed to solve for the gravitational potential $\Phi$. The linearized 00 Einstein equation is $$\label{eq:linearized 00} -k^2 \Psi=4\pi G \left\{\left[(2\beta_{13}+4\beta_4)k^2 -8\lambda M^2\right]M(u+M\Phi)-2\beta_{13}k^2 M^2\Phi+4\beta_4k^2 M^2(\Psi-\Phi)+\rho_m\right\}.$$ Inserting $\Psi=\gamma \Phi$ and equation (\[eq:delta AA\]) into (\[eq:linearized 00\]) leads to the Poisson equation $\Delta \Phi =4\pi G_\mathrm{eff} \rho_m$. However, the effective gravitational constant $G_\mathrm{eff}$ is actually not constant, but scale dependent, $$\label{eq:Newtonian G} \frac{G_\mathrm{eff}}{G}= \left\{ \begin{array}{cl} [1-8 \pi G M^2(\beta_1+2\beta_4)]^{-1} & \text{large scales}, \\ \displaystyle \frac{\beta_1}{\beta_1 \gamma -8\pi G M^2 [2\beta_1(\beta_{13}-\beta_4(\gamma-2))-(\beta_{13}+2\beta_4)(\beta_{13}+2\beta_4(1-2\gamma))]} & \text{small scales}. \end{array} \right.$$ The vector field again renormalizes $G$ on large and small scales, but by a different amount than in the cosmological case (\[eq:cosmic G\]), even on large scales. This disagreement in the local and cosmological values of $G$ points to a violation of Birkhoff’s theorem in these theories, and also is what led to the nucleosynthesis constraint discussed in Section \[sec:cosmology\]. Note that the change in the gravitational constant caused by the term proportional to $\beta_4$ is the one that follows from replacing $A^2\to -M^2$ in the action (\[eq:action\]), and is the same both for the cosmological and Newtonian gravitational constants. If we set $\beta_4=0$, our value of $G_\mathrm{eff}$ on large scales agrees with the one found for aether models in [@Carroll:2004ai]. Note that the value of $\beta$ does not have any impact on the Newtonian limit of the theory. As an illustration of these results, let us consider a massive vector with $\beta_4=0$. It follows automatically from equation (\[eq:gamma\]) that $\gamma=1$, so in this case the theory agrees with the current measurements of $\gamma$ both on short and large scales. The effective Newton constant takes the simpler form $$\frac{G_\mathrm{eff}}{G}= \left\{ \begin{array}{cc} [1-8\pi G M^2 \beta_1]^{-1}& \text{large scales},\\ \displaystyle [1-8\pi G M^2 \beta_{13}(2-\beta_{13}/\beta_1)]^{-1} & \text{small scales}. \end{array} \right.$$ This equation is remarkable because it says that the strength of gravity changes with scale. It is hence tempting to speculate a connection with dark matter or even dark energy. For instance, if $\beta_1/\beta_{13}>0$ gravity is stronger at large scales than at small scales, so if the vector had a Compton wavelength of a few kiloparsecs this could eventually explain the flatness of galactic rotation curves without dark matter. Recall that if $\beta_1\neq 0$, stability imposes $\beta_1>0$ and $\beta_{13}>0$. Such a scale-dependent behavior of the gravitational coupling does not occur in massless vector field theories or aether models, though it does happen in theories with more complicated Lagrangians [@Zlosnik:2006zu]. In connection with modified Newtonian dynamics [@Milgrom:1983ca], a specific model that involves a fixed-norm vector has been proposed by Bekenstein in [@Bekenstein:2004ne]. Ideas about degravitation on large scales have been explored in [@Dvali:2007kt] and references therein. \* Although our expressions (\[eq:gamma\]) and (\[eq:Newtonian G\]) still hold when $\beta_1=0$ or $\beta_4=0$, they become ill-defined when both $\beta_1$ and $\beta_4$ vanish. This case however is particularly simple. It follows directly from equation (\[eq:linearized ij\]) that $\gamma=1$, while the remaining linearized equations lead to $$\frac{G_\mathrm{eff}}{G}=\frac{4\lambda M_G^2}{4\lambda M_G^2-\beta_{13}^2 k^2}.$$ Therefore, on short scales the theory does not have a Newtonian limit (the gravitational potential satisfies the equation $k^4\Phi\propto \rho$), while if the field is massive, $G_\mathrm{eff}\approx G$ on long scales. Ghosts and Instabilities {#sec:stability} ======================== The strongest constraints on the parameters in the action (\[eq:action\]) are set by the requirement that the theory be classically and quantum-mechanically stable. In this somewhat technical section we proceed to analyze the conditions that classical and quantum-mechanical stability place on the parameters of the action (\[eq:action\]). Analogous conditions for vector fields of constrained norm have been derived in [@Lim:2004js; @Carroll:2008em], while a somewhat similar analysis of massless vector fields of arbitrary norm has been carried out in [@Jimenez:2008sq]. Instabilities in models of inflation supported by vector fields have been recently reported in [@Himmetoglu:2008zp; @Himmetoglu:2008hx; @Himmetoglu:2009qi]. Our stability analysis closely resembles the one for massive gravity in [@Dubovsky:2004sg] In order to analyze the stability of unleashed aether models, we consider vector field fluctuations $b_\mu$ around the background value $\bar{A}_\mu$ of the field, $$\label{eq:fluctuations} A_\mu=\bar{A}_\mu+b_\mu.$$ Because we are mostly interested in cosmological solutions we assume that the background vector $\bar{A}^\mu$ has the form (\[eq:A\]), and leave the case of a field that points along a spatial direction to Appendix \[sec:appendix\]. We neglect metric perturbations, which relies on the assumption that $\Lambda\ll M_G$. In conformal time coordinates, the metric and background field are hence given by $$\label{eq:background} ds^2=a^2(\eta)(-d\eta^2+d\vec{x}\,^2), \quad \bar{A}^\mu=a^{-1}\cdot (\bar{A},0,0,0).$$ Note that $\bar{A}$ plays the role of $u$ in Section \[sec:cosmology\], and is hence dynamical. To obtain the Lagrangian of the fluctuations $b^\mu$ we insert the expansion (\[eq:fluctuations\]) and the background (\[eq:background\]) into the action (\[eq:action\]). Expanding to quadratic order in the fluctuations we arrive at $$\begin{aligned} \label{eq:action a} \mathcal{L}_b=&-&\beta_{1}\partial_{\mu}b_{\nu}\partial^{\mu}b^{\nu}-(\beta-\beta_1)(\partial_{\mu}b^{\mu})^{2} -4\beta(\mathcal{H}^2 b_0^2 -\mathcal{H}b_0\partial_\mu b^\mu) +\nonumber \\ {}&+&\beta_{13}R_{\mu\nu}b^\mu b^\nu+a^2\beta_4R \,b_\mu b^\mu-a^2 V' b_\mu b^\mu-2a^4V'' \cdot (\bar{A}^\mu b_\mu)^2 ,\end{aligned}$$ where indices are raised with the Minkowski metric $\eta^{\mu\nu}$, $\mathcal{H}=d\log a/d \eta$, and a prime denotes a derivative with respect to $A_\mu A^\mu$. For arbitrary parameters $\beta_1$ and $\beta$, the four components of the vector field are dynamical, because the determinant of the Hessian matrix $$\label{eq:Hessian} \frac{\partial^2 \mathcal{L}_b}{\partial (db_\mu/d\eta) \partial(db_\nu/d\eta)}$$ is non-zero [@Henneaux:1992ig]. However, as we shall see below, for $\beta=0$ or $\beta_1=0$ the determinant of the Hessian vanishes, signaling that some of the vector field components are constrained. This is what happens for instance for the gauge invariant Maxwell theory ($\beta=0$). We thus study the degenerate cases $\beta=0$ and $\beta_1=0$ separately. In aether models, the fixed norm constrain requires $b^0=0$ [@Lim:2004js], and thus eliminates one of the four dynamical fields. Note by the way that if we expand the potential of equation (\[eq:potential\]) to quartic order in the perturbations we obtain an additional term $$\mathcal{L}_b\supset \lambda (b_\mu b^\mu)^2.$$ Thus, for $\lambda>1$, the theory is strongly coupled. Stability in the vector sector ------------------------------ The background value of the vector field determines the set of unbroken spacetime symmetries. Because $\bar{A}_\mu$ points along the time direction, the action (\[eq:action a\]) is symmetric under spatial translations and rotations. It is convenient to decompose the fields in irreducible representations of this unbroken symmetry group. Expanding as usual $$\label{eq:action fluctuations} b_0\equiv -u \quad \text{and} \quad b_i=\partial_i r+v_i, \quad \text{with} \quad \partial_i v^i=0,$$ decouples the two scalar modes $u$ and $r$ from the two vector modes in $\vec{v}$. The Lagrangian of the vector sector is then given by $$\label{eq:vector L} \mathcal{L}_v=-\beta_1 \partial_\mu \vec{v}\cdot \partial^\mu \vec{v}- a^2 m_v^2 \, \vec{v}\cdot \vec{v},$$ where $$\label{eq:mv} m_v^2\equiv V'-3(\beta_{13}+4\beta_4)H^2-(\beta_{13}+6\beta_4)\dot{H}.$$ The factor of $a^2$ in the mass term of equation (\[eq:vector L\]), as well as the contributions to the mass proportional to the curvature explicitly break time translation invariance. Therefore, the quantity one would naively identify as the frequency of the modes in the vector sector $$\label{eq:omega v} \omega_v^2=\vec{p}\,^2+\frac{a^2}{\beta_1}m_v^2$$ also depends on time (in this regard one can consider $a^2 m^2_v/\beta_1$ as the “mass” of these excitations.) We can nevertheless quantize the theory under the approximation that the frequency is constant if the relative change in $\omega$ during the characteristic time $\Delta \eta \approx \omega^{-1}$ is small [@Mukhanov:2007zz], $$\label{eq:adiabaticity} \frac{1}{\omega^2}\frac{d\omega}{d\eta}\ll 1.$$ In the short wavelength regime, $\vec{p}\, ^2/a^2 \gg m_v^2$, this adiabaticity condition is automatically satisfied, but in the long wavelength regime $\vec{p}\, ^2/a^2 \ll m_v^2$, it imposes restrictions on the magnitude of $V'$ and its time derivatives. For instance, if $V'\ll H^2$, the mass $m_v$ is of the order of $H^2$, and hence $\omega^{-2}_v (d\omega_v/d\eta)$ is of order one. Our analysis in this section exclusively applies in the adiabatic regime (\[eq:adiabaticity\]). In this regime, we can set all explicit factors of $a$ to one. Returning now to equation (\[eq:vector L\]), and expressing the vector in terms of two orthogonal transverse polarizations immediately reveals that the vector sector is ghost free for ${\beta_1\geq 0}$, and free of classical instabilities at low momenta ($\vec{p}\,^2\ll |m_v^2|$) for $m_v^2/\beta_1\geq 0.$ These conditions are summarized in Table \[table:vector stability\]. We should also point out that tachyonic, long-wavelength instabilities in a theory may be allowed, as long as they are not too severe. Their severity depends on the magnitude of the tachyonic “mass" $\omega^2\equiv -m_t^2$, which determines the rate of growth of the mode, proportional to $\exp(m_t \, \eta)$. Since we mostly assume that all masses are much larger than the Hubble scale, the presence of a tachyon would imply instabilities on timescales much shorter than the Hubble time, which is likely to be problematic. On the other hand, tachyonic masses of the order of $H$ (or smaller) should not be much of a problem. Note that if $\beta_1=0$, the vector sector is not dynamical. [**Vector**]{} Low $p$ High $p$ ---------------- ----------------------- ------------------ Classical $\beta_1 m_v^2\geq 0$ Quantum $\beta_1 \geq 0$ $\beta_1 \geq 0$ : Stability conditions in the vector sector. The checkmark means that the condition is automatically satisfied. The mass $m_v$ is defined in equation (\[eq:mv\]).[]{data-label="table:vector stability"} Stability in the scalar sector ------------------------------ To proceed further, we expand the scalar fields in appropriately normalized Fourier modes, $$\label{eq:expansion} u=\int \frac{d^4p}{(2\pi)^{4}} \, u(p) \exp(i p_\mu x^\mu), \quad r= \int \frac{d^4p}{(2\pi)^{4}} \, \frac{r(p)}{|\vec{p}\,|} \exp(i p_\mu x^\mu).$$ Since the fields are real, it follows that $u^*(p)=u(-p)$ and $r^*(p)=r(-p)$. Substituting the expansions (\[eq:expansion\]) in (\[eq:action a\]) and assuming as above that all explicitly time-dependent quantities are constant we arrive at the action (in Fourier space) $$\label{eq:timelike L} \mathcal{S}_s=-\int \frac{d^4p}{(2\pi)^{4}} \left(\begin{array}{l}u \\ r\end{array}\right)^\dag D \left(\begin{array}{l}u\\ r\end{array} \right),$$ where the matrix $D$ is given by $$D=\left(\begin{array}{lr} \beta \, \omega^2-\beta_1\vec{p}\,{}^2+ m_u^2 & -i\,\omega\, |\vec{p}\,| (\beta-\beta_1)-2\beta |\vec{p}\,| H \\ i\,\omega\, |\vec{p}\,| (\beta-\beta_1)-2\beta |\vec{p}\,|H & -\beta_1\,\omega^2+\beta\,\vec{p}\,{}^2+m_v^2 \end{array} \right)$$ and we have also defined $$\label{eq:mu} m_u^2\equiv 2V''\bar{A}^2-V'+(2\beta+3\beta_{13}+12\beta_4)H^2-(2\beta-3\beta_{13}-6\beta_4)\dot{H}.$$ In spite of our notation, the mass parameter $m_u^2$ is not the mass of the field $u$, because the Lagrangian is not diagonal in that field. Note that in an expanding universe, the terms proportional to $R_{\mu\nu}$ and $R$ in equation (\[eq:action\]) also contribute to $m_u$ and $m_v$. These contributions are typically negligible, unless the derivatives of the potential are sufficiently small. The inverse of the matrix $D$ is the field propagator. Thus, in order to find the propagating modes we just have to find the values of $\omega^2$ at which its eigenvalues have poles, or, equivalently, the values of $\omega^2$ at which the eigenvalues of $D$ have zeros. The determinant of $D$ is $$\label{eq:determinant} \text{det}\, D=-\beta\beta_1 \omega^4+(2\beta\beta_1 \vec{p}\,^2-\beta_1 m_u^2+\beta m_v^2)\omega^2-(\beta\beta_1\vec{p\,}^4-\beta m_u^2 \, \vec{p}\,^2+\beta_1 m_v^2\, \vec{p}\,^2+4\beta^2 H^2 \vec{p}\,^2 -m_u^2 m_v^2).$$ Requiring that $\det D$ vanish we thus arrive at the frequencies of the *two* propagating scalar modes $$\begin{aligned} \omega^2&=&\vec{p}\,^2+\frac{m_v^2}{2\beta_1}-\frac{m_u^2}{2\beta}\pm \Delta, \quad \text{where} \label{eq:omega}\\ \Delta&\equiv& \sqrt{\left[\left(\frac{1}{\beta_1}-\frac{1}{\beta}\right) (m_u^2+m_v^2)-4\frac{\beta}{\beta_1} H^2\right]\vec{p}\,^2 +\frac{1}{4}\left(\frac{m_u^2}{\beta}+\frac{m_v^2}{\beta_1}\right)^2}. \label{eq:delta}\end{aligned}$$ Observe that some of the factors of $\omega^2$ in the determinant (\[eq:determinant\]) are multiplied by the mass terms $m_v^2$ and $m_u^2$. Since factors of $\omega$ arise from time derivatives, and because the mass terms contain derivatives of the potential, we may say that the potential is also part of the vector field kinetic terms. The theory is free of classical instabilities (exponentially growing modes) if the frequency $\omega$ in equation (\[eq:omega\]) is real. In the limit of low momenta, in which $\vec{p}\,^2$ is negligible, the two frequencies approach[^6] $$\begin{aligned} \label{eq:omega low p} \omega^2_+ \to &-&\frac{m_u^2}{\beta} +\left[1-\frac{(\beta-\beta_1)(m_u^2+m_v^2)-4 \beta^2H^2}{\beta_1m_u^2+\beta m_v^2}\right]\vec{p}\,^2 +\mathcal{O}\left(\vec{p}\,^4\right),\\ \omega^2_-\to &{}&\frac{m_v^2}{\beta_1} +\left[1+\frac{(\beta-\beta_1)(m_u^2+m_v^2)-4 \beta^2H^2}{\beta_1m_u^2+\beta m_v^2}\right]\vec{p}\,^2+\mathcal{O}\left(\vec{p}\,^4\right),\end{aligned}$$ in which, for later convenience, we have included terms that are typically subdominant at low momenta. In the limit of high momenta, in which mass terms can be neglected, the frequencies become $$\label{eq:omega high p} \omega_\pm^2\to \vec{p}\,^2\pm |\vec{p}| \sqrt{\left(\frac{1}{\beta_1}-\frac{1}{\beta}\right) (m_u^2+m_v^2)-4\frac{\beta}{\beta_1}H^2}+\mathcal{O}\left(|\vec{p}\,|^0\right).$$ Demanding that the mode frequencies be real at low and high momenta we arrive at the conditions on Table \[table:scalar stability\], which if satisfied actually imply that the system is classically stable for all momenta (this follows from a detailed analysis of equations (\[eq:omega\]) and (\[eq:delta\]).) Inspection of equations (\[eq:omega\]) and (\[eq:delta\]) shows that the regimes of high and low momentum not only depend on the masses of the fields, but also on the values of $\beta$ and $\beta_1$. Note that in the high-momentum limit, the leading term in the frequency gives $\omega_\pm\approx \vec{p}\,^2$. We consider the subleading contributions, proportional to $|\vec{p}\,|$, because if the latter are imaginary there are instabilities on arbitrarily short timescales. The residues at the poles of the eigenvalues of $D^{-1}$ determine whether the propagating modes are ghosts. If these residues have negative signs, the corresponding modes have negative energies. In the presence of ghosts, the vacuum can then decay into positive energy quanta and negative energy ghosts, leading to a quantum-mechanically unstable theory. How severe this instability is depends on the way the ghosts couple to positive energy quanta, and on the volume of phase space available for the decay [@Carroll:2003st; @Cline:2003gs]. The coupling of ghosts to positive energy quanta is essentially determined by the coupling constants of the theory and the magnitude of the residues, while the phase space available for decay typically depends on the cut-off up to which the theory is valid. Choosing any of these parameters appropriately might tame ghost instabilities, although we shall not explore this possibility here. In this context, it is worthwhile pointing out however that with Lorentz-invariance spontaneously broken, the cutoff of the theory involves *spatial* momenta rather than Lorentz-covariant four-momenta. To illustrate this let us consider one of the many higher-dimensional operators that we neglected in (\[eq:action\]), $$\label{eq:corrections} \frac{\delta}{\Lambda^6}A^\mu A^\nu A^\rho A^\sigma (\nabla_\mu \nabla_\nu A^\tau) (\nabla_\rho\nabla_\sigma A_\tau),\label{eq:higher dimensional}$$ where $\delta$ is a dimensionless parameter of order one. Expanding (\[eq:corrections\]) to quadratic order in the fluctuations $b_\mu$ in Minkowski space we obtain $$\label{eq:correction} \mathcal{L}_s\supset \delta\frac{\bar{A}^4 \omega^4}{\Lambda^6}u^2.$$ Hence, this term would make an effective contribution to the matrix $D$ which would become comparable to the lowest order term $\beta\omega^2 u^2$ if $$\omega^2\approx \frac{\Lambda^6}{\bar{A}\,^4}.$$ As we have seen, at low momenta the dispersion has the form $\omega^2=m^2$, where, by definition, $m$ is the mass of the field, while at high momenta the dispersion relation typically approaches $\omega^2=c_s^2\, \vec{p}\,^2$, where $c_s^2$ is a coefficient of order one. Thus, our effective description is certainly not valid for masses and spatial momenta beyond $\Lambda^3/\bar{A}^2$. In fact, higher-dimensional operators analogous to (\[eq:corrections\]), of the form $$\frac{1}{\Lambda^{4n-2}}A^{\mu_1}\cdots A^{\mu_n} A^{\rho_1}\cdots A^{\rho_n} (\nabla_{\mu_1}\cdots \nabla_{\mu_m} A^\tau) (\nabla_{\rho_1}\cdots \nabla_{\rho_m} A_\tau),$$ point to a breakdown scale of the order $\Lambda^2/\bar{A}$. Note that in the limit $\bar{A}\to 0$, the corrections due to the new term goes to zero. In that limit Lorentz invariance is restored, and large spatial momenta can be made small by a Lorentz transformation.[^7] If $\lambda_+(\omega^2)$ and $\lambda_-(\omega^2)$ are the two eigenvalues of the matrix $D$, and the former have zeros at, respectively, $\omega_+$ and $\omega_-$, the residues at the poles of $D^{-1}$ are[^8] $$\label{eq:Z} Z_\pm=\lim_{\omega^2 \to\omega_{\pm}^2} \frac{(-\omega^2+\omega_{\pm}^2)}{\lambda_{\pm}(\omega)}\quad \Rightarrow \quad \frac{1}{Z_\pm}=-\frac{d\lambda_{\pm}}{d\omega^2}\Bigg|_{\omega^2_\pm}.$$ The theory is free of ghosts if both residues are positive. Since the two eigenvalues of the matrix $D$ are $$\lambda_{1, 2}= \frac{\text{tr}\, D \pm \sqrt{\text{tr}^2 D-4\, \text{det}\, D}}{2}$$ the condition for the absence of ghosts is then $$\label{eq:no ghosts} \frac{1}{Z_{\pm}}=-\left(\frac{1}{\text{tr}\, D}{\frac{d\, \text{det}\, D}{d\omega^2}}\right)\Bigg|_{\omega^2_{\pm}} >0.$$ The trace of the matrix $D$ at $\omega_\pm$ is $$\text{tr} D\big|_{\omega^2_{\pm}}=2(\beta-\beta_1) \vec{p}\,^2+\frac{\beta+\beta_1}{2}\left(\frac{m_u^2}{\beta}+\frac{m_v^2}{\beta_1}\right)\pm (\beta-\beta_1) \Delta,$$ whereas the derivative of the determinant is $$\label{eq:derivative D} \frac{d\, \text{det}\, D}{d\omega^2}\Bigg|_{\omega^2_{\pm}}=\mp 2\beta\beta_1 \Delta.$$ From the previous equations it immediately follows that the theory always contains a high-momentum ghost. Indeed, in the limit of high momenta, the trace approximately equals the same value both at $\omega_+$ and $\omega_-$, while the value of $d \det D/d\omega^2$ changes sign. Therefore, the residues cannot be positive for both frequencies. More precisely, in the limit of high-momentum the residues approach $$\label{eq:Z generic} \frac{1}{Z_\pm}\to \pm \frac{\sqrt{(\beta_1^{-1}-\beta^{-1})(m_v^2+m_u^2)-4\beta\beta_1^{-1} H^2}}{(\beta_1^{-1}-\beta^{-1})|\vec{p}\,|},$$ which implies that one linear combination of fields is a ghost, while the remaining orthogonal combination is stable (classical stability forces the term inside the square root to be positive.) Because at low momenta the residues approach $$\label{eq:Z low p} \frac{1}{Z_+}\to -\beta,\quad \frac{1}{Z_-}\to \beta_1,$$ it is possible to avoid the existence of low-momentum ghosts for appropriate choices of $\beta$ and $\beta_1$. Classical and quantum stability conditions are summarized in Table \[table:scalar stability\]. Note by the way that there is a ghost at high momenta for $\beta=\beta_1$, even though equation (\[eq:Z generic\]) breaks down for those values. In this case, the trace of $D$ has always the same value, irrespectively of $\vec{p}$. If $H$ or any of the two scalar masses is non-zero, $\Delta$ is non-zero, and, as before, equation (\[eq:no ghosts\]) cannot be satisfied. This is immediately apparent in the Lagrangian (\[eq:action a\]). For $\beta=\beta_1$, the Lagrangian is already diagonal in the fields $b^\mu$, and because of the indefinite signature of the Minkowski metric, at least one of the fields has to be a ghost. [**Scalar**]{} ($\beta_1\neq 0, \, \beta\neq 0$) Low $p$ High $p$ -------------------------------------------------- ------------------------------------------------ ------------------------------------------------------------------- Classical $ \beta_1\cdot m_v^2>0,\, \beta\cdot m_u^2<0$ $(\beta_1^{-1}-\beta^{-1})(m_u^2+m_v^2)-4\beta\beta_1^{-1} H^2>0$ Quantum $ \beta_1>0, \beta<0$ : Stability conditions in the scalar sector for $\beta_1\neq 0$ and $\beta\neq 0$. The cross means that the condition cannot be satisfied. The parameters $m_v$ and $m_u$ are defined, respectively, in equations (\[eq:mv\]) and (\[eq:mu\]).[]{data-label="table:scalar stability"} ### Massless case It is important to realize that the high-momentum limit is different from the strict massless case $$H^2=m_u^2=m_v^2=0.$$ Indeed, in the massive case there are always two propagating modes in the scalar sector at high momenta, with frequencies that approach the relativistic dispersion relations ${\omega^2=\vec{p}\,^2(1\pm \mathcal{O}(m/|\vec{p}\,|))}$. In contrast, in the massless case, $\Delta$ in equation (\[eq:delta\]) vanishes, and equation (\[eq:derivative D\]) implies that there is a double pole at $\omega^2=\vec{p}\,^2$ instead of two single poles at $\omega^2=\vec{p}\,^2$. Thus, in the massless case there is a scalar with higher time derivatives plus the two degrees of freedom of the vector sector. Lagrangians with higher time derivatives are usually quantized using a procedure developed by Ostrogradski, and are typically unstable (see [@Woodard:2006nt] for a review.) The same conclusion follows by substituting the Lorentz-covariant decomposition ${b_\mu=\partial_\mu \phi+v_\mu}$, with $\partial_\mu v^\mu=0$ into the action $$\label{eq:S_M} S_M=\int d^{4}x \,\left[ -\frac{\beta_{1}}{2}F_{\mu\nu}F^{\mu\nu}-\beta(\nabla_{\mu}A^{\mu})^{2}-m^2 \,A_\mu A^\mu\right].$$ This yields a Lagrangian which describes massive electrodynamics in “Lorentz gauge” and a scalar $\phi$ whose kinetic term contains higher order derivatives, $$\label{eq:Lorentz L} \mathcal{L}_M=-\frac{\beta_1}{2} (\partial_\mu v_\nu-\partial_\nu v_\mu)(\partial^\mu v^\nu-\partial^\nu v^\mu)-\beta (\Box \phi)^2-m^2 \partial_\mu \phi \partial^\mu \phi -m^2 v^\mu v_\mu.$$ In this form, it is manifest that the Lagrangian not only describes a vector field, but also a scalar field (under Lorentz transformations.) By inspecting the scalar kinetic term in equation (\[eq:Lorentz L\]) one would naively conclude that in the massive case there are two scalar poles, at $p^2=-m^2/\beta$ and $p^2=0$, but this is illusory. Because $\partial_\mu v^\mu=0$, the propagator of the original vector field $A_\mu$ is $$\label{eq:propagator A} \Delta_A=\Delta_v+p_\mu p_\nu \Delta_\phi= \frac{1}{\beta_1 p^2+m^2}\left(\eta_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\right)+ \frac{1}{\beta \, p^2+m^2}\frac{p_\mu p_\nu}{p^2}.$$ The propagator of $v$ is just that of electrodynamics in $R_\xi$ gauge [@Weinberg:Rxi], with $\xi\to 0$ and an added mass term. It can be obtained by inverting the kinetic term in equation (\[eq:S\_M\]) and letting $\beta$, which plays the role of $1/\xi$, tend to infinity. This propagator is manifestly transverse (projects onto the subspace perpendicular to $p_\mu$), while the contribution from $\phi$ to $\Delta_A$ is clearly longitudinal (projects onto the subspace parallel to $p_\mu$.) In the form (\[eq:propagator A\]) it is now clear that there are no poles at $p^2=0$. Instead, there is a pole at $p^2=-m^2/\beta$ in the longitudinal sector, and three poles at $p^2=-m^2/\beta_1$ in the transverse sector. These correspond to the same frequencies that we found in our previous analysis, as the reader can readily verify by setting $H=0$, $m_v^2=m^2$ and $m_u^2=-m^2$ in equations (\[eq:omega\]) and (\[eq:delta\]). Note that we could have also obtained (\[eq:propagator A\]) by directly inverting the differential operator in equation (\[eq:Lorentz L\]). Let us finally discuss $m^2=0$, as we intended to do. In this case, the poles at $p^2=0$ that canceled in the massive case remain, and give rise to factors proportional to $1/p^4$. There is then a double pole at $p^2=0$, and two single poles at $p^2=0$. These considerations also illustrate that a free ($V\equiv 0$) vector field in flat space can behave very differently from a free vector field in an expanding universe if $\beta, \beta_{13}$ or $\beta_4$ are different from zero. ### Degenerate Limits Our conclusions do not apply to the two degenerate limits in which equation (\[eq:omega\]) breaks down namely, for $\beta=0$ and $\beta_1=0$. In both cases, the determinant of the Hessian matrix (\[eq:Hessian\]) vanishes, or equivalently, the determinant of $D$, equation (\[eq:determinant\]), is not quartic in $\omega$, but quadratic. The Lagrangian therefore describes only one scalar propagating mode, instead of two. When $\beta=0$, the “kinetic" part of the Lagrangian is that of the electromagnetism and the Proca Lagrangian. The determinant of the matrix $D$ only has one zero at $$\label{eq:b0 omega} \omega_0^2=\frac{m_v^2}{\beta_1}-\frac{m_v^2}{m_u^2}\,\vec{p}\,^2,$$ and the residue at this pole is $$\label{eq:b0 r} Z_0=\frac{1}{\beta_1}+\frac{m_v^2-m_u^2}{m_u^4}\, \vec{p}\,^2.$$ Thus, equations (\[eq:b0 omega\]) and (\[eq:b0 r\]) result in the stability conditions in Table \[table:b0b1 stability\]. Note that some of these conditions are the same as in the vector sector, Table \[table:vector stability\], and that if the theory is stable both in the low and high momentum limits, it will be stable for all momenta. If $m_u^2=0$, the equation $\text{det}\, D=0$ has no solution and the theory does not have any propagating scalar degrees of freedom. From the point of view of flat space, equation (\[eq:propagator A\]), setting $\beta=0$ eliminates the longitudinal polarization associated with the scalar $\phi$. In the massless case, one of the the remaining transverse polarizations disappears, leaving just the two modes of the vector sector. In this case, the theory has an additional symmetry, gauge invariance. As illustration, let us consider the familiar example of the Proca Lagrangian, in which $\beta_1=1/2$, $\beta=\beta_{13}=\beta_4=0$ and $V=m^2 A_\mu A^\mu/2$. For a vanishing background field this implies $m_v^2=m^2/2$ and $m_u^2=-m^2/2$. It follows from Table \[table:b0b1 stability\], that the Proca field is stable for $m^2>0$ and both classically and quantum-mechanically unstable for $m^2<0$. Again, the massless limit is not continuous: for $m^2=0$ there are no propagating scalars. Massive versions of electromagnetism with mass terms that violate Lorentz invariance ($m_u^2\neq -m_v^2$) have been studied in [@Dvali:2005nt; @Gabadadze:2004iv] . [**Scalar**]{} ($\beta_1= 0$) Low $p$ High $p$ ------------------------------- ------------------ ----------------------------- Classical $ \beta m_u^2<0$ $m_v^2(m_u^2-4\beta H^2)<0$ Quantum $\beta<0$ $ m_u^2-m_v^2-4\beta H^2<0$ : Stability conditions in the scalar sector for timelike background fields in the degenerate cases $\beta_1=0$ or $\beta= 0$. If the theory is stable both in the low and high momentum limits, it follows from equations (\[eq:b0 omega\]), (\[eq:b0 r\]), (\[eq:b1 omega\]) and (\[eq:b1 r\]) that it is actually stable for all momenta. The parameters $m_v$ and $m_u$ are defined, respectively, in equations (\[eq:mv\]) and (\[eq:mu\]).[]{data-label="table:b0b1 stability"} If $\beta_1=0$ the determinant of $D$ vanishes at $$\label{eq:b1 omega} \omega_0^2=-\frac{m_u^2}{\beta}-\frac{m_u^2-4\beta H^2}{m_v^2}\vec{p}\,^2,$$ while the residues at the poles are $$\label{eq:b1 r} Z_0=-\frac{1}{\beta}-\frac{m_v^2-m_u^2+4\beta H^2}{m_v^4}\vec{p}\,^2.$$ Combining both (\[eq:b1 omega\]) and (\[eq:b1 r\]) we arrive at the stability conditions in Table \[table:b0b1 stability\]. If these conditions are satisfied, the theory is stable for all momenta. Once again, if $m_v^2=0$ the inverse of $D$ has no poles, and the theory does not contain any dynamical scalar (and, since $\beta_1=0$, there is no dynamical vector either.) In a Lorentz-covariant language, setting $\beta_1=0$ eliminates the three transverse poles in the propagator of equation (\[eq:propagator A\]). In the massless case, the longitudinal polarization also disappears, leaving a theory without dynamics. In fact, in this limit the theory has an additional symmetry $b_\mu\to b_\mu+\varepsilon_\mu{}^{\nu\rho\sigma}\partial_\nu c_{\rho\sigma}$, where $c_{\rho\sigma}$ is an antisymmetric tensor (a two-form), which can be used to gauge all the components of $b_\mu$ away. Stability around the minimum of the potential {#sec:stability aether} --------------------------------------------- As an important application of this analysis, let us consider stability in the case where the vector field sits at the minimum of a potential given by equation (\[eq:potential\]). As discussed in Subsection \[sec:extremum\] this is generically possible only in flat space. Substituting $H\equiv0$ and equation (\[eq:potential\]) into equations (\[eq:mv\]) and (\[eq:mu\]) we find $$m_v^2=0\quad \text{and} \quad m_u^2=4\lambda M^2.$$ Let us consider the case $\beta_1\neq 0$ and $\beta\neq 0$ first. The spectrum then contains two vector and two scalar degrees of freedom, with dispersion relations at low momenta given, respectively, by equations (\[eq:omega v\]) and (\[eq:omega low p\]), $$\label{eq:spectrum} \omega^2_{v_1}=\vec{p}\,^2, \quad \omega^2_{v_2}=\vec{p}\,^2,\quad \omega^2_{s_+}\approx-\frac{4\lambda M^2}{\beta}, \quad \omega^2_{s_-}\approx\frac{\beta}{\beta_1}\vec{p}\,^2.$$ The three “massless" excitations $v_1, v_2$ and $s_-$ can thus be interpreted as the Nambu-Goldstone bosons associated with the spontaneous breaking of the three generators of Lorentz boosts. At low momenta, the corresponding residues are, from equations (\[eq:vector L\]) and (\[eq:Z low p\]), $$\label{eq:residues} Z_{v_1}=1/\beta_1, \quad Z_{v_2}=1/\beta_1,\quad Z_{s_+}=-1/\beta, \quad Z_{s_-}=1/\beta_1.$$ In consequence, at low momenta the theory is either classically or quantum-mechanically stable, but not both. The absence of ghosts requires $\beta_1>0$ and $\beta<0$, which then implies that $s_-$ is classically unstable. Conversely, classical stability demands $\lambda/\beta<0$ and $\beta/\beta_1>0$, which then implies that either $v_1, v_2$ and $s_-$, or $s_+$, are ghosts. The only way to escape this conclusion is to choose a well-behaved set of massless modes ($\beta>0$ and $\beta_1>0$) and push the massive mode—either a tachyon or a ghost—out of the domain of validity of the effective theory. Indeed, as mentioned above, if the mass of $s_+$ is of the order $\Lambda^2/M$, we cannot trust our analysis anyway, so the presence of a tachyon or ghost in the spectrum is not necessarily fatal. In principle, this pathology could be solved by the ultraviolet completion of the theory. Note that these instabilities cannot be avoided by lowering the mass scale $M$ and “squeezing" the low-momentum regime out of the spectrum of the theory; at high momenta, $\vec{p}\,^2\gg 4\lambda M^2$, according to Table \[table:scalar stability\] and equation (\[eq:Z generic\]), either $s_+$ or $s_-$ are ghost-like too. Formally, we can decouple the massive mode $s_+$ by sending $|\lambda|$ to infinity, which is the limit in which we expect to recover aether models. In fact, the low momentum spectrum (\[eq:spectrum\]) and residues (\[eq:residues\]) of the massless modes $v_1, v_2$ and $s_-$ agree with those of aether theories [@Lim:2004js]. In that sense, unleashed aether models can be regarded as a high-momentum completions of the former, which, as we have argued, are inconsistent for generic values of $\beta$ and $\beta_1$, unless $\lambda$ is larger than $(\Lambda/M)^4$. This is not just a fine-tuning problem. Because we expect $\Lambda/M\gg 1$, it also means that the theory is strongly coupled. The only way in which strong coupling issues could be circumvented (without avoiding fine-tuning) in a sensible low energy effective theory is by considering $0<\beta\lesssim(M/\Lambda)^{4}$. In any case, from this perspective it is also easy to understand the agreement on large scales of the Newtonian limits of both constrained and unleashed aether models. At low momenta, both are essentially the same theory. These instabilities can be averted by setting $\beta_1=0$ or $\beta=0$. Since no non-trivial cosmological solutions with $\beta=0$ exist, let us hence consider $\beta_1=0$ next. In this case the vector sector is non-dynamical, and because $m_v^2=0$, there are no dynamical scalars in the theory either. This is however a degenerate limit. In an expanding universe, the field is slightly displaced from the minimum, as discussed in Section \[sec:cosmology\]. Assuming that $H^2\ll 4\lambda M^2$ and inserting equation (\[eq:adiabatic\]) into equation (\[eq:mv\]) we find, to leading order, $$\label{eq:mv minimum} m_v^2=-(2\beta_{13}-3\beta)\dot{H}.$$ Hence, $m_v^2$ is of order $H^2$ and generally different from zero. The spectrum of the theory then consists of a single scalar $s_0$ with dispersion relation and residues given by, respectively, $$\omega_0^2\approx -\frac{4\lambda M^2}{\beta}\left(1-\frac{\beta}{2\beta_{13}-3\beta} \frac{\vec{p}\,^2}{\dot{H}}\right) \quad \text{and} \quad \quad Z_0\approx -\frac{1}{\beta}+\frac{4\lambda M^2 \vec{p}\,^2}{(2\beta_{13}-3\beta)^2 \dot{H}^2}.$$ Thus, the scalar is stable at low momenta if $\beta<0$ and $\lambda>0$. Because $\dot{H}<0$ whenever the null energy condition[^9] is satisfied, the scalar is stable at high momenta if in addition $2\beta_{13}-3\beta<0$. Comparing with the general case above, equation (\[eq:spectrum\]), we see that at low momenta the effect of setting $\beta_1=0$ is to remove the (classically unstable) scalar Goldstone mode $s_-$ from the spectrum. The remaining scalar is massive, and thus is absent from the low-momentum description provided by aether theories. In fact, aether models with $\beta_1=0$ do not have any propagating degrees of freedom. The case $\beta=0$ is somewhat similar. Using equation (\[eq:mv minimum\]) we find that the dispersion relation and residues of the only scalar mode are $$\omega_0^2\approx-2\beta_{13}\dot{H}\left(\frac{1}{\beta_1}-\frac{\vec{p}\,^2}{4\lambda M^2}\right), \quad \text{and} \quad Z_0\approx \frac{1}{\beta_1}- \frac{\vec{p}\,^2}{4\lambda M^2}$$ Thus, using $\dot{H}<0$ again, stability at low momenta requires $\beta_1>0$ and $\beta_{13}>0,$ while stability at high momenta then imposes $\lambda<0$. Relation between energy conditions, instabilities and superluminality --------------------------------------------------------------------- A violation of the standard energy conditions is often associated with superluminal propagation and instabilities in a theory [@Carroll:2003st; @Dubovsky:2005xd]. In k-field models for example, the absence of ghosts implies the null energy condition $\rho+p\geq 0$, and also places at the same time restrictions on the value of the speed of sound (see for instance [@ArmendarizPicon:2005nz]). It is therefore instructive to check whether these relations also extend to unleashed aether models. From equation (\[eq:omega\]) one could immediately calculate, say, the group velocity of the perturbations, $$v_g=\frac{d\omega}{dp}.$$ For instance, it follows from equation (\[eq:spectrum\]) that the propagation of $s_-$ is subluminal for $\beta/\beta_1<1$, and superluminal for $\beta/\beta_1>1$. As discussed in Section \[sec:stability aether\], the stability of that mode (in our restricted sense) only requires $\beta>0$ and $\beta_1>0$. Similarly, from equation (\[eq:omega high p\]), at high momenta the dispersion relations of the scalar modes approach the relativistic expressions $\omega_{\pm}=p$, whilst our stability analysis indicates that in that limit there is always a ghost in the spectrum. Thus, subluminal propagation seems to be independent from stability considerations (in this respect, the unleashed aether is analogous to k-fields.) The requirement of subluminal propagation is sometimes used to constrain the parameters in this type of theories, but we believe that there is nothing inconsistent with faster than light propagation in this background. The constant vector field defines a preferred frame and explicitly breaks Lorentz invariance. Even if a signal propagates superluminally, it will move forward in our preferred time coordinate, it will not be possible to construct closed timelike curves, and no conflict with causality will arise. For different viewpoints on the issue see e.g. [@Adams:2006sv; @Bruneton:2006gf; @Babichev:2007dw; @Dubovsky:2008bd]. Readers concerned about superluminal propagation should impose appropriate conditions on the parameters of the theory. For instance, the additional constraint $\beta/\beta_1\leq 1$ implies that the mode $s_-$ propagates subluminally (at low momenta). Analogous conditions in the gravitational wave sector have been discussed in [@Lim:2004js]. Note that signal propagation in the vector sector is never superluminal. In the limit in which the expansion of the universe is negligible the energy density (\[eq:energy\]) is $$\rho=-\beta\dot{u}^2+V.$$ In the same limit, the sum of pressure and energy density is $$\rho+p=2\left(\frac{\beta_{13}+2\beta_4}{\beta}-1\right)V'u^2-2(\beta_{13}+2\beta_4)\dot{u}^2.$$ Therefore, the energy density is positive for $\beta<0$ and $V>0$, while $\rho+p$ is positive if $\beta_{13}+2\beta_4<0$ and $V'(\beta_{13}+2\beta_4-\beta)/\beta>0$. Examination of Tables \[table:vector stability\] to \[table:b0b1 stability\] shows that these conditions bear a marginal resemblance with the conditions for the absence of ghosts in the limit of low momentum, which is the one that applies for an homogeneous field. In particular, the absence of ghosts implies $\beta<0$, which is the condition for the “kinetic” part of the energy density to be positive. Otherwise, at least in this particular case, it appears that the weak and null energy conditions have little to do with the presence or absence of ghosts, at least in this class of models. Summary and Conclusions ======================= At sufficiently long timescales and large distances, any generally covariant theory that contains a real vector field coupled to gravity, and is invariant under $A_\mu\to -A_\mu$, should be described by an action of the form (\[eq:action\]). Massless vector-tensor theories [@Will:1981cz] and aether models [@Jacobson:2000xp] can be regarded as particular limits of this “unleashed aether" class. We have identified some of their common features, and pointed out in what other aspects they differ significantly. Stability imposes severe conditions on the parameters of the action (\[eq:action\]). The spectrum of the theory generally consists of a massive excitation and three massless fields, which can be interpreted as the Nambu-Goldstone bosons associated with the spontaneous symmetry breaking of invariance under boosts. At least one of these four fields is always a tachyon or a ghost. For appropriate choices of parameters however, it is possible to decouple the dangerous modes. Setting $\beta_1=0$ eliminates all the transverse components (in a Lorentz covariant way) of the vector. And setting $\beta=0$ eliminates just the longitudinal polarization, which is dynamical only for a massive vector or in an expanding universe. Both choices result in a well-behaved theory. Non-trivial cosmological solutions exist only if $\beta_1=0$, that is, if the kinetic term of the vector field is proportional to the squared divergence of the vector. We have also extended our stability analysis to models in which the vector field points along a spatial direction (see the Appendix \[sec:appendix\]) and found that they also suffer generically from ghosts. There is yet another alternative to eliminate the unstable fields from the theory. For appropriate choices of parameters, the only unstable field in the theory is massive. If its mass is sufficiently heavy, its excitations are beyond the regime of validity of our low-energy description and can be ignored. The remaining three fields, the Nambu-Goldstone bosons, are the only fields present in the spectrum of aether models. In that sense the aether can be regarded as a low-momentum description of unleashed aether models. Since the symmetries of the universe only encompass spatial translations and rotations, it is possible for a vector field to have a non-vanishing expectation value along its time direction. In unleashed aether models such a non-vanishing cosmic vector field leads to a host of interesting gravitational phenomena. In the Newtonian limit of the theory Newton’s constant is typically scale-dependent, and for some parameter values the Eddington parameter $\gamma$ differs from its value in general relativity. On cosmological scales, the vector field leaves an imprint on the expansion of the universe, even in the vicinity of a potential minimum. The field renormalizes Newton’s constant, like in aether models, but it also typically oscillates around it. Under relatively mild assumptions, these oscillations strongly constrain the value of the field at the minimum of the potential. Perhaps the most important conclusion is that self-consistent and phenomenologically acceptable modifications of gravity are difficult to come by. In this article we have added a massive vector field of unknown origin to the gravitational sector, and we have employed a low-energy effective approach to study the implications of this addition. We have seen that general vector field theories have a very rich gravitational phenomenology, but that self-consistency and agreement with basic cosmological and solar system observations impose severe constraints on this class of models. Nevertheless, a large volume of parameter space is still allowed by our requirements, and many interesting phenomenological aspects remain to be investigated. Unleashed aether models can wander far away. We thank Riccardo Penco for useful comments on an earlier version of this manuscript, and Quentin G. Bailey for pointing out to us a few missing references. ADT acknowledges the kind hospitality of the members of the Physics Department at Syracuse University, where this work was initiated. The work of CAP was supported in part by the US National Science Foundation under grant PHY-0604760. The work of ADT was supported by a UNAM postdoctoral fellowship and funds from the Basque Government GICO7/51-IT-221-07. Stability of Spacelike Vectors {#sec:appendix} ============================== The stability analysis in Section \[sec:stability\] assumes that the vector field points in the time direction, as required by isotropy. But even if the field is aligned along a spatial direction it is possible in some cases to restore the isotropy of spacetime by either considering appropriate configurations of multiple vector fields [@ArmendarizPicon:2004pm; @Golovnev:2008cf], or by simply assuming that the vector is subdominant. Our stability analysis can also be extended to this type of configurations, for which the background value of $A^\mu$ is spacelike. If the vector field points along a spatial direction, we can always pick our coordinates so that $\bar{A}^\mu$ points in the $z$ direction, $$\label{eq:spacelike A} \bar{A}^\mu=a^{-1}(0,0,0,\bar{A}).$$ In order to determine whether such a configuration is stable, we consider again the fluctuations around the background value of the field, $A_\mu=\bar{A}_\mu+b_\mu$. But instead of expanding the fluctuations as in (\[eq:action fluctuations\]) we decompose the perturbations in irreducible representations of $SO(2)$, the group of rotations that leaves $\bar{A}^\mu$ in equation (\[eq:spacelike A\]) invariant, $$\label{eq:spacelike expansion} b_0\equiv -\tilde{u}, \quad b_\alpha\equiv\partial_\alpha \tilde{r}+\tilde{v}_\alpha \,\, (\text{with} \, \partial_\alpha \tilde{v}^\alpha=0), \quad b_z\equiv \tilde{t},$$ where $\alpha$ runs over the transverse spatial coordinates, $\alpha=x,y$. Thus, under $SO(2)$, $\tilde{u}$, $\tilde{r}$ and $\tilde{t}$ are scalars, while $\tilde{v}$ is a vector. Substituting the expansion (\[eq:spacelike expansion\]) into equation (\[eq:action a\]) we obtain the Lagrangian for the perturbations. The Lagrangian for the transverse vector modes $\tilde{v}^\alpha$ takes the simple form $$\mathcal{L}_{\tilde{v}}=-\beta_1 \partial_\mu \tilde{v}_\alpha \partial^\mu \tilde{v}^\alpha-m_v^2\tilde{v}_\alpha \tilde{v}^\alpha,$$ where $m_v^2$ is given by equation (\[eq:mv\]) (recall that we assume that $a^2 m_v^2$ is adiabatically constant, and thus set $a= 1$.) Thus, the absence of instabilities leads again to the conditions in Table \[table:vector stability\]. For spacelike background fields the scalar sector in this case is slightly more complicated, because there are three coupled scalar fields, instead of two. In Fourier space, their action is $$\begin{aligned} \label{eq:spacelike L} \mathcal{S}_{\tilde{s}}=-\frac{1}{2}\int \frac{d^4p}{(2\pi)^{4}}&& \Big\{\left[\beta \, \omega^2 -\beta_1 \vec{p}\,^2+\tilde{m}_u^2\right] \tilde{u}^*\tilde{u}+ \left[-\beta_1 \omega^2 \vec{p}_{\perp}{}^2 +\beta_1 \vec{p}\,^2 p_{\perp}^2+(\beta-\beta_1)\vec{p}\,^4_{\perp}+p_{\perp}^2 m_v^2\right] \tilde{r}^*\tilde{r} + \nonumber \\ &&+\left[-\beta_1 \omega^2 +\beta_1\vec{p}\,^2+(\beta-\beta_1)p_{\parallel}^2+\tilde{m}_t^2\right] \tilde{t}^*\tilde{t} -4\beta H \vec{p}_{\perp}\!^2 \tilde{r}^*\tilde{u}+\\ &&-4i\beta H p_\parallel\,\tilde{t}^*\tilde{u} -2(\beta-\beta_1)\left[i\omega \vec{p}_{\perp}\!^2\tilde{r}^* \tilde{u}-\omega p_{\parallel} \tilde{t}^*\tilde{u} -i \vec{p}_\perp\!^2 p_{\parallel}\, \tilde{t}^*\tilde{r}\right]+h.c. \Big\}, \nonumber\end{aligned}$$ where we have defined $$\begin{aligned} \tilde{m}_u^2&=&-V'+(2\beta+3\beta_{13}+12\beta_4)H^2-(2\beta-3\beta_{13}-6\beta_4)\dot{H}, \label{eq:mu tilde} \\ \tilde{m}_t^2&=&V'+2V''\bar{A}^2-3(\beta_{13}+4\beta_4)H^2-(\beta_{13}+6\beta_4)\dot{H}, \label{eq:mt tilde}\end{aligned}$$ and $p_{\parallel}=p_z$, $\vec{p}_{\perp}=(p_x,p_y,0)$. Note that $\tilde{m}_u^2$ differs from the $m_u^2$ in equation (\[eq:mu\]). For later convenience, we have not normalized $s$ canonically. A general analysis of the stability of this system would be rather cumbersome, because we would have to diagonalize the $3\times3$ matrix. For our purposes however it suffices to focus on particular values of the momentum $\vec{p}$, for which some of the scalars decouple from the others. Specifically, setting $\vec{p}_{\perp}=0$ in equation (\[eq:spacelike L\]) we arrive at the action $$\begin{aligned} \mathcal{S}_{\tilde{s}}=-\int \frac{d^4p}{(2\pi)^{4}} &\Big\{&\left[\beta \, \omega^2 -\beta_1 p_\parallel^2+\tilde{m}_u^2\right] \tilde{u}^*\tilde{u} +\left[-\beta_1 \omega^2 +\beta p_{\parallel}^2+\tilde{m}_t^2\right] \tilde{t}^*\tilde{t}+\nonumber \\ {}&+&2i\beta H p_\parallel\left[\tilde{u}^* \tilde{t}- \tilde{t}^* \tilde{u}\right] +(\beta-\beta_1) \omega p_{\parallel}\left[\tilde{u}^* \tilde{t}+\tilde{t}^* \tilde{u}\right] \Big\}.\end{aligned}$$ This has the same form as equation (\[eq:timelike L\]). The field $\tilde{r}$ is non-dynamical and has disappeared from the Lagrangian, while $\tilde{t}$ plays the role of $r$ in equation (\[eq:timelike L\]). The only difference is an overall factor of $i$ in the non-diagonal terms, which does not affect the trace nor the determinant of the matrix $D$. Thus, the stability of perturbations that only depend on $z$ (and time) is determined by the same conditions we derived in the previous section, with mass parameters given by equations (\[eq:mu tilde\]) and (\[eq:mt tilde\]). It follows for instance that for generic values of $\beta, \beta_1$, the theory has a high-momentum ghost. The presence of a ghost at high momenta is not exclusive of $z$ dependent perturbations, and also appears for perturbations that only depend on $x$ and $y$. Indeed, setting $p_{\parallel}=0$ into the action (\[eq:spacelike L\]) gives $$\begin{aligned} \mathcal{S}=-\int \frac{d^4p}{(2\pi)^{4}} &\Big\{&\left[\beta \, \omega^2-\beta_1 \vec{p}_{\perp}\!^2 +\tilde{m}_u^2 \right]\tilde{u}^*\tilde{u} +\left[-\beta_1 \omega^2 \vec{p}_{\perp}\!^2 + \beta\vec{p}_{\perp}\!^4 +\vec{p}_{\perp}\!^2 m_v^2\right] \tilde{r}^*\tilde{r} - \nonumber \\ {}&&-2\beta H \vec{p}_{\perp}\!^2\left[\tilde{u}^*\tilde{r}+\tilde{r}^*\tilde{u}\right] -i\omega (\beta-\beta_1) \vec{p}_{\perp}\!^2(\tilde{r}^* \tilde{u}-\tilde{u}^* \tilde{r})+\nonumber \\ &&+\left[-\beta_1 \omega^2+\beta_1 \vec{p}_{\perp}\!^2+ \tilde{m}_t^2\right]\tilde{t}^*\tilde{t}\Big\}.\end{aligned}$$ The field $\tilde{t}$ has decoupled from the other two scalars, whose action has again the form of (\[eq:timelike L\]) (after canonically normalizing $\tilde{r}$.) Thus, as before, the theory has ghosts at sufficiently high values of $\vec{p}\,^2_{\perp}$ for generic values of $\beta, \beta_1$. For $\beta=0$ or $\beta_1=0$ these instabilities disappear if appropriate conditions are satisfied. These are listed in Tables \[table:b0 spacelike stability\] and \[table:b1 spacelike stability\]. As an application, let us consider for instance Lagrangians with $\beta=0$, as in the “new inflation" vector models of [@Ford:1989me], or as in the dark energy model of [@ArmendarizPicon:2004pm]. Using equations (\[eq:mv\]), (\[eq:mu tilde\]) and (\[eq:mt tilde\]), and neglecting curvature terms we find $$\begin{aligned} \tilde{m}_t^2-\tilde{m}_u^2&\approx& 2V'+2V'' \bar{A}^2\\ m_v^2-\tilde{m}_u^2&\approx&2V' \\ m_v^2&\approx&V'.\end{aligned}$$ It follows automatically from our analysis that all models with “run-away" potentials with $V'<0$, in which the derivatives of the potential dominate over curvature terms, contain ghosts in the scalar sector at high momenta, and are thus quantum-mechanically unstable [@Dubovsky]. [**Scalar**]{} ($\beta=0$) Low $p$ High $p$ ---------------------------- ----------------------------------------------- ----------------------------------------------------------------------- Classical $ \beta_1 \tilde{m}_t^2>0,\, \beta_1 m_v^2>0$ $\tilde{m}_t^2\tilde{m}_u^2<0,\, m_v^2\tilde{m}_u^2<0$ Quantum $\beta_1>0$ $\beta_1>0,\, \tilde{m}_t^2-\tilde{m}_u^2>0,\, m_v^2-\tilde{m}_u^2>0$ : Stability conditions in the scalar sector for spacelike background fields and $\beta=0$.[]{data-label="table:b0 spacelike stability"} [**Scalar**]{} ($\beta_1=0$) Low $p$ High $p$ ------------------------------ -------------------------- ----------------------------------------------------------------------------------- Classical $\beta \tilde{m}_u^2<0$ $ \tilde{m}_r^2(\tilde{m}_u^2-4\beta H^2)<0,\, m_v^2(\tilde{m}_u^2-4\beta H^2)<0$ Quantum $\beta<0$ $\tilde{m}_u^2-\tilde{m}_t^2-4\beta H^2>0,\, \tilde{m}_u^2-m_v^2-4\beta H^2>00$ : Stability conditions in the scalar sector for spacelike background fields and $\beta_1=0$.[]{data-label="table:b1 spacelike stability"} Let us conclude by emphasizing that the stability conditions we have derived in this Appendix are necessary but not sufficient. Because we have not analyzed the stability of all Fourier modes, a violation of our conditions points to the instability of certain field modes, but their satisfaction does not imply that all the modes are free of instabilities. [99]{} C. M. Will and K. J. Nordtvedt, “Conservation Laws and Preferred Frames in Relativistic Gravity. I. Preferred-Frame Theories and an Extended PPN Formalism,” Astrophys. J.  [**177**]{}, 757 (1972). R. W. Hellings and K. Nordtvedt, “Vector-Metric Theory of Gravity,” Phys. Rev.  D [**7**]{}, 3593 (1973). L. H. Ford, “Inflation Driven by a Vector Field,” Phys. Rev.  D [**40**]{}, 967 (1989). C. Armendariz-Picon, “Could dark energy be vector-like?,” JCAP [**0407**]{}, 007 (2004) \[arXiv:astro-ph/0405267\]. H. Wei and R. G. Cai, “Interacting vector-like dark energy, the first and second cosmological coincidence problems,” Phys. Rev.  D [**73**]{}, 083002 (2006) \[arXiv:astro-ph/0603052\]. S. Kanno and J. Soda, “Lorentz violating inflation,” Phys. Rev.  D [**74**]{}, 063505 (2006) \[arXiv:hep-th/0604192\]. M. Novello, E. Goulart, J. M. Salim and S. E. Perez Bergliaffa, “Cosmological effects of nonlinear electrodynamics,” Class. Quant. Grav.  [**24**]{}, 3021 (2007) \[arXiv:gr-qc/0610043\]. C. G. Boehmer and T. Harko, “Dark energy as a massive vector field,” Eur. Phys. J.  C [**50**]{}, 423 (2007) \[arXiv:gr-qc/0701029\]. T. Koivisto and D. F. Mota, “Accelerating Cosmologies with an Anisotropic Equation of State,” Astrophys. J.  [**679**]{}, 1 (2008) \[arXiv:0707.0279 \[astro-ph\]\]. J. B. Jimenez and A. L. Maroto, “A cosmic vector for dark energy,” Phys. Rev.  D [**78**]{}, 063005 (2008) \[arXiv:0801.1486 \[astro-ph\]\]. T. Koivisto and D. F. Mota, “Anisotropic Dark Energy: Dynamics of Background and Perturbations,” JCAP [**0806**]{}, 018 (2008) \[arXiv:0801.3676 \[astro-ph\]\]. A. Golovnev, V. Mukhanov and V. Vanchurin, “Vector Inflation,” JCAP [**0806**]{}, 009 (2008) \[arXiv:0802.2068 \[astro-ph\]\]. T. S. Koivisto and D. F. Mota, “Vector Field Models of Inflation and Dark Energy,” JCAP [**0808**]{}, 021 (2008) \[arXiv:0805.4229 \[astro-ph\]\]. S. Yokoyama and J. Soda, “Primordial statistical anisotropy generated at the end of inflation,” JCAP [**0808**]{}, 005 (2008) \[arXiv:0805.4265 \[astro-ph\]\]. S. Kanno, M. Kimura, J. Soda and S. Yokoyama, “Anisotropic Inflation from Vector Impurity,” JCAP [**0808**]{}, 034 (2008) \[arXiv:0806.2422 \[hep-ph\]\]. M. a. Watanabe, S. Kanno and J. Soda, “Hairy Inflation,” arXiv:0902.2833 \[hep-th\]. S. Koh, “Vector Field and Inflation,” arXiv:0902.3904 \[hep-th\]. T. Jacobson and D. Mattingly, “Gravity with a dynamical preferred frame,” Phys. Rev.  D [**64**]{}, 024028 (2001) \[arXiv:gr-qc/0007031\]. M. Gasperini, “Singularity Prevention and Broken Lorentz Symmetry," Class. Quant. Grav.  [**4**]{}, 485 (1987). T. Jacobson, “Einstein-aether gravity: a status report,” PoS [**QG-PH**]{}, 020 (2007) \[arXiv:0801.1547 \[gr-qc\]\]. V. A. Kostelecky and S. Samuel, “Spontaneous Breaking of Lorentz Symmetry in String Theory,” Phys. Rev.  D [**39**]{}, 683 (1989). P. Horava, “Quantum Gravity at a Lifshitz Point,” Phys. Rev.  D [**79**]{}, 084008 (2009) \[arXiv:0901.3775 \[hep-th\]\]. J. Collins, A. Perez, D. Sudarsky, L. Urrutia and H. Vucetich, “Lorentz invariance and quantum gravity: an additional fine-tuning problem?,” Phys. Rev. Lett.  [**93**]{}, 191301 (2004) \[arXiv:gr-qc/0403053\]. C. Charmousis, G. Niz, A. Padilla and P. M. Saffin, “Strong coupling in Horava gravity,” JHEP [**0908**]{}, 070 (2009) \[arXiv:0905.2579 \[hep-th\]\]. D. Blas, O. Pujolas and S. Sibiryakov, “On the Extra Mode and Inconsistency of Horava Gravity,” JHEP [**0910**]{}, 029 (2009) \[arXiv:0906.3046 \[hep-th\]\]. N. Arkani-Hamed, H. C. Cheng, M. A. Luty and S. Mukohyama, “Ghost condensation and a consistent infrared modification of gravity,” JHEP [**0405**]{}, 074 (2004) \[arXiv:hep-th/0312099\]. R. Bluhm and V.A. Kostelecky, “Spontaneous Lorentz violation, Nambu-Goldstone modes, and gravity,” Phys. Rev.  D [**71**]{}, 065008 (2005) \[arXiv:hep-th/0412320\]. Q.G. Bailey and V.A. Kostelecky, “Signals for Lorentz violation in post-Newtonian gravity,” Phys. Rev.  D [**74**]{}, 045001 (2006) \[arXiv:gr-qc/0603030\]. R. Bluhm, S-H. Fung and V.A. Kostelecky, “Spontaneous Lorentz and Diffeomorphism Violation, Massive Modes, and Gravity,” Phys. Rev.  D [**77**]{}, 065020 (2008) \[arXiv:0712.4119 \[hep-th\]\]. V.A. Kostelecky and R. Potting, “Gravity from spontaneous Lorentz violation,” Phys. Rev.  D [**79**]{}, 065018 (2009) \[arXiv:0901.0662 \[gr-qc\]\]. S. Weinberg, “The Quantum theory of fields. Vol. 1: Foundations,” Section 5.9, [*Cambridge, UK: Univ. Pr. (1995) 609 p*]{}. D. Colladay and V. A. Kostelecky, “Lorentz-violating extension of the standard model,” Phys. Rev.  D [**58**]{}, 116002 (1998) \[arXiv:hep-ph/9809521\]. V. A. Kostelecky and N. Russell, “Data Tables for Lorentz and CPT Violation,” arXiv:0801.0287 \[hep-ph\]. D. Mattingly, “Modern tests of Lorentz invariance,” Living Rev. Rel.  [**8**]{}, 5 (2005) \[arXiv:gr-qc/0502097\]. C. P. Burgess, “Introduction to effective field theory,” Ann. Rev. Nucl. Part. Sci.  [**57**]{}, 329 (2007) \[arXiv:hep-th/0701053\]. C. M. Will, “Theory And Experiment In Gravitational Physics,” [*Cambridge, Uk: Univ. Pr. ( 1981) 342p*]{}. S. M. Carroll and E. A. Lim, “Lorentz-violating vector fields slow the universe down,” Phys. Rev.  D [**70**]{}, 123525 (2004) \[arXiv:hep-th/0407149\]. J. D. Bjorken, Annals Phys.  [**24**]{}, 174 (1963). P. Kraus and E. T. Tomboulis, “Photons and gravitons as Goldstone bosons, and the cosmological constant,” Phys. Rev.  D [**66**]{}, 045015 (2002) \[arXiv:hep-th/0203221\]. C. Eling and T. Jacobson, “Static post-Newtonian equivalence of GR and gravity with a dynamical preferred frame,” Phys. Rev.  D [**69**]{}, 064005 (2004) \[arXiv:gr-qc/0310044\]. B. Z. Foster and T. Jacobson, “Post-Newtonian parameters and constraints on Einstein-aether theory,” Phys. Rev.  D [**73**]{}, 064015 (2006) \[arXiv:gr-qc/0509083\]. B. Bertotti, L. Iess and P. Tortora, “A test of general relativity using radio links with the Cassini spacecraft,” Nature [**425**]{}, 374 (2003). T. G. Zlosnik, P. G. Ferreira and G. D. Starkman, “Modifying gravity with the Aether: an alternative to Dark Matter,” Phys. Rev.  D [**75**]{}, 044017 (2007) \[arXiv:astro-ph/0607411\]. M. Milgrom, “A Modification Of The Newtonian Dynamics As A Possible Alternative To The Hidden Mass Hypothesis,” Astrophys. J.  [**270**]{}, 365 (1983). J. D. Bekenstein, “Relativistic gravitation theory for the MOND paradigm,” Phys. Rev.  D [**70**]{}, 083509 (2004) \[Erratum-ibid.  D [**71**]{}, 069901 (2005)\] \[arXiv:astro-ph/0403694\]. G. Dvali, S. Hofmann and J. Khoury, “Degravitation of the cosmological constant and graviton width,” Phys. Rev.  D [**76**]{}, 084006 (2007) \[arXiv:hep-th/0703027\]. E. A. Lim, “Can we see Lorentz-violating vector fields in the CMB?,” Phys. Rev.  D [**71**]{}, 063504 (2005) \[arXiv:astro-ph/0407437\]. S. M. Carroll, T. R. Dulaney, M. I. Gresham and H. Tam, “Instabilities in the Aether,” arXiv:0812.1049 \[hep-th\]. J. B. Jimenez and A. L. Maroto, “Viability of vector-tensor theories of gravity,” arXiv:0811.0784 \[astro-ph\]. B. Himmetoglu, C. R. Contaldi and M. Peloso, “Instability of anisotropic cosmological solutions supported by vector fields,” arXiv:0809.2779 \[astro-ph\]. B. Himmetoglu, C. R. Contaldi and M. Peloso, “Instability of the ACW model, and problems with massive vectors during inflation,” arXiv:0812.1231 \[astro-ph\]. B. Himmetoglu, C. R. Contaldi and M. Peloso, “Ghost instabilities of cosmological models with vector fields nonminimally coupled to the curvature,” arXiv:0909.3524 \[astro-ph.CO\]. S. L. Dubovsky, “Phases of massive gravity,” JHEP [**0410**]{}, 076 (2004) \[arXiv:hep-th/0409124\]. M. Henneaux and C. Teitelboim, “Quantization of gauge systems,” [*Princeton, USA: Univ. Pr. (1992) 520 p*]{} V. Mukhanov and S. Winitzki, “Introduction to quantum effects in gravity,” [*Cambridge, UK: Cambridge Univ. Pr. (2007) 273 p*]{} S. M. Carroll, M. Hoffman and M. Trodden, “Can the dark energy equation-of-state parameter w be less than -1?,” Phys. Rev.  D [**68**]{}, 023509 (2003) \[arXiv:astro-ph/0301273\]. J. M. Cline, S. Jeon and G. D. Moore, “The phantom menaced: Constraints on low-energy effective ghosts,” Phys. Rev.  D [**70**]{}, 043543 (2004) \[arXiv:hep-ph/0311312\]. C. Armendariz-Picon, M. Fontanini, R. Penco and M. Trodden, “Where does Cosmological Perturbation Theory Break Down?,” arXiv:0805.0114 \[hep-th\]. R. P. Woodard, “Avoiding dark energy with 1/R modifications of gravity,” Lect. Notes Phys.  [**720**]{}, 403 (2007) \[arXiv:astro-ph/0601672\]. S. Weinberg, “The quantum theory of fields. Vol. 2: Modern applications,” Section 15.2, [*Cambridge, UK: Univ. Pr. (1996) 489 p*]{} G. Dvali, M. Papucci and M. D. Schwartz, “Infrared Lorentz violation and slowly instantaneous electricity,” Phys. Rev. Lett.  [**94**]{}, 191602 (2005) \[arXiv:hep-th/0501157\]. G. Gabadadze and L. Grisa, “Lorentz-violating massive gauge and gravitational fields,” Phys. Lett.  B [**617**]{}, 124 (2005) \[arXiv:hep-th/0412332\]. S. Dubovsky, T. Gregoire, A. Nicolis and R. Rattazzi, “Null energy condition and superluminal propagation,” JHEP [**0603**]{}, 025 (2006) \[arXiv:hep-th/0512260\]. C. Armendariz-Picon and E. A. Lim, “Haloes of k-essence,” JCAP [**0508**]{}, 007 (2005) \[arXiv:astro-ph/0505207\]. A. Adams, N. Arkani-Hamed, S. Dubovsky, A. Nicolis and R. Rattazzi, “Causality, analyticity and an IR obstruction to UV completion,” JHEP [**0610**]{}, 014 (2006) \[arXiv:hep-th/0602178\]. J. P. Bruneton, “On causality and superluminal behavior in classical field theories. Applications to k-essence theories and MOND-like theories of gravity,” Phys. Rev.  D [**75**]{}, 085013 (2007) \[arXiv:gr-qc/0607055\]. E. Babichev, V. Mukhanov and A. Vikman, “k-Essence, superluminal propagation, causality and emergent geometry,” JHEP [**0802**]{}, 101 (2008) \[arXiv:0708.0561 \[hep-th\]\]. S. Dubovsky and S. Sibiryakov, “Superluminal Travel Made Possible (in two dimensions),” JHEP [**0812**]{}, 092 (2008) \[arXiv:0806.1534 \[hep-th\]\]. The existence of instabilities in vector field theories with run-away potentials was suggested long ago to one of us by S. Dubovsky in a private communication. [^1]: As we shall see, the “potential" $V$ actually contains kinetic terms for some of the vector degrees of freedom. [^2]: We consider here positive values of $M^2$. In that case, $A^\mu$ is timelike in the broken phase. If $M^2$ is negative, the vector vev is spacelike. [^3]: Our parameters are related to those of [@Carroll:2004ai] by $\beta_{13}\equiv \beta_1+\beta_3$, $\beta\equiv \beta_1+\beta_2+\beta_3$, and by those of [@Will:1981cz] by $\beta_1\equiv 2\epsilon-\tau$, $\beta\equiv -\tau$, $\beta_{13}\equiv \eta-\tau$ and $\beta_4\equiv\omega$. [^4]: Except for some especial cases in which non-homogeneous field configurations are pure gauge so do not contribute to the energy-momentum tensor. [^5]: It is important to point out here that we are only considering *classical* (coherent) excitations for the vector field. Theories with $\beta=0$ but $\beta_{1}\neq 0$ could also have cosmological relevance (as it happens for instance with the electromagnetic field) if they do not behave classically. (For instance, in the case of the Cosmic Microwave Background, what we have is a thermal distribution of photons). [^6]: We are assuming that $(m_u^2/\beta+m_v^2/\beta_1)<0$. Otherwise the values of $\omega_+$ and $\omega_-$ become interchanged. [^7]: A similar phenomenon was described in the context of a scalar field theory in an inflating universe in [@ArmendarizPicon:2008yv]. There, the scale of Lorentz symmetry breaking is set by the Hubble constant, which plays the role of $\bar{A}$ here. [^8]: In some cases, one of the eigenvalues may have multiple zeros. In that case, evaluation of equation (\[eq:Z\]) at those zeroes gives the corresponding residues. [^9]: $\rho\geq 0$ and $\rho+p\geq 0$.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We interpret heterotic M-theory in terms of $h$-cobordism, that is the eleven-manifold is a product of the ten-manifold times an interval is translated into a statement that the former is a cobordism of the latter which is a homtopy equivalence. In the non-simply connected case, which is important for model building, the interpretation is then in terms of $s$-cobordism, so that the cobordism is a simple-homotopy equivalence. This gives constraints on the possible cobordisms depending on the fundamental groups and hence provides a characterization of possible compactification manifolds using the Whitehead group– a quotient of algebraic K-theory of the integral group ring of the fundamental group– and a distinguished element, the Whitehead torsion. We also consider the effect on the dynamics via diffeomorphisms and general dimensional reduction, and comment on the effect on F-theory compactifications.' --- [ **Constraints on heterotic M-theory from s-cobordism** ]{} Hisham Sati [^1] Department of Mathematics\ University of Maryland\ College Park, MD 20742 Introduction ============ A major goal of string theory is to provide a unification of fundamental interactions. This includes constructing the standard model via string compactifications [@GSW2], most notably via heterotic M-theory [@HW1] [@HW2]. Eleven-dimensional spacetime is taken to be an interval $I$ times a ten-manifold $M^{10}$, and with the two boundaries each supporting an $E_8$ gauge theory. One of the boundaries is called the hidden sector and the other is the visible sector, in which the structure group is broken down to a realistic symmetry group. The ten-manifold $M^{10}$ is typically taken to be Minkowski space ${{\mathbb R}}^{1,3}$ times a Calabi-Yau threefold $X^6$. In the visible sector one usually works with SU(5) or SO(10) $\subset E_8$ and breaks this group further (at least in principle) to the standard model group (ideally) SU(3)$\times$SU(2)$\times$U(1). Physical and mathematical constraints on the (Calabi-Yau) manifold $X^6$ and bundles on $X^6$ are recently reviewed in [@He]. Wilson lines are needed to break the gauge group from the grand unified (GUT) group to the standard model group [@BOS] [@W85]. In order to introduce Wilson lines, the manifolds $M^{10}$ must have a nontrivial fundamental group. Starting with a simply connected Calabi-Yau manifold, one gets a smooth non-simply connected Calabi-Yau manifold by dividing by a freely acting discrete symmetry $X^6 \mapsto X^6/\Gamma$, where $\Gamma$ is a discrete group of finite order $|\Gamma|$, and the resulting fundamental group is $\pi_1(X^6/\Gamma) =\Gamma$. Important choices for the finite group include $\Gamma={{\mathbb Z}}_2$, which breaks SU(5) down to SU(3)$\times$SU(2)$\times$U(1), and $\Gamma= {{\mathbb Z}}_3\times {{\mathbb Z}}_3$ or ${{\mathbb Z}}_6$, which break SO(10) down to SU(3)$\times$SU(2)$\times {\rm U(1)}^2$ [@GSW2]. A major area of research involves choosing $\Gamma$ so that one gets the standard model, not just as far as the symmetry groups are concerned but also accounting for example for correct generations and spectra of particles. A sampler of fundamental groups of Calabi-Yau threefolds $X^6$ applied in the heterotic setting include: ${{\mathbb Z}}_2$ [@BD], ${{\mathbb Z}}_2 \times {{\mathbb Z}}_2$ [@DOPR], ${{\mathbb Z}}_3 \times {{\mathbb Z}}_3$ [@BOPR] [@BHOP], ${{\mathbb Z}}_8 \times {{\mathbb Z}}_8$ constructed in [@GP] on which rank 5 bundles are constructed in [@BBD], abelian surface fibrations over ${{\mathbb C}}P^1$ with (abelianization of) fundamental group ${{\mathbb Z}}_n \times {{\mathbb Z}}_n$ are considered in [@Sch] [@DGS], complete intersection Calabi-Yau manifolds with fundamental groups which include [@CD] ${{\mathbb Z}}_3$, ${{\mathbb Z}}_3 \times {{\mathbb Z}}_2$, ${{\mathbb Z}}_3 \times {{\mathbb Z}}_3$, ${{\mathbb Z}}_5$, ${{\mathbb Z}}_5 \times {{\mathbb Z}}_2$, ${{\mathbb Z}}_5 \times {{\mathbb Z}}_5$, and the quaternion group $Q_8$, the latter being closely related to construction in [@BH] of Calabi-Yau threefolds with nonabelian fundamental groups, roughly speaking a semidirect product of ${{\mathbb Z}}_8$ with a quaternion group. Torsion curves, important for instanton corrections to the heterotic minimally supersymmetric standard model (MSSM), are studied in [@BKOS] for the quintic as well as for threefolds with fundamental groups ${{\mathbb Z}}_3 \times {{\mathbb Z}}_3$. Almost all known Calabi-Yau threefolds are simply connected. For example, only 16 out of about 500 million hypersurfaces in complex 4-dimensional toric varieties have nontrivial fundamental groups, and the only groups which occur are ${{\mathbb Z}}_2$, ${{\mathbb Z}}_3$ or ${{\mathbb Z}}_5$ [@BK]. All elliptically fibered Calabi-Yau threefolds are simply connected, with the exception of fibrations over an Enriques base. In [@DOPW] elliptic fibrations without section, i.e. torus bundles, with nontrivial fundamental group are constructed. Another class of examples with no section is the Schoen family [@Sc] which are fiber products of two rational elliptic surfaces. Free finite group actions on these are classified (under certain conditions) [@BD2] giving fundamental groups $\pi_1(X) \in \{ {{\mathbb Z}}_2, {{\mathbb Z}}_3, {{\mathbb Z}}_4, {{\mathbb Z}}_2\times{{\mathbb Z}}_2, {{\mathbb Z}}_5, {{\mathbb Z}}_6, {{\mathbb Z}}_2\times{{\mathbb Z}}_4, {{\mathbb Z}}_3\times{{\mathbb Z}}_3\ \}$. [^2] In another class of threefolds, the complete intersections in products of projective spaces, an exhaustive search [@Br] of the 7890 such threefolds leads to many interesting fundamental groups including ${{\mathbb Z}}_i$ (for $i=2,3,4,5,6,8,10,12$), ${{\mathbb Z}}_2 \times {{\mathbb Z}}_j$ (for $j=2,4,8, 10$), ${{\mathbb Z}}_4 \times {{\mathbb Z}}_k$ (for $k=4,8$), ${{\mathbb Z}}_5 \times {{\mathbb Z}}_5$, ${{\mathbb Z}}_8 \times {{\mathbb Z}}_8$, as well as semidirect products ${{\mathbb Z}}_3 \rtimes {{\mathbb Z}}_4$, ${{\mathbb Z}}_4 \rtimes {{\mathbb Z}}_4$, ${{\mathbb Z}}_5 \times {{\mathbb Z}}_{10}$ and groups involving the quaternion group $Q_8$, namely ${{\mathbb Z}}_2 \times Q_8$, ${{\mathbb Z}}_2 \times {{\mathbb Z}}_2 \times Q_8$, ${{\mathbb Z}}_4 \rtimes Q_8$, ${{\mathbb Z}}_8 \rtimes Q_8$ (for a complete list see [@Br]). In this paper we seek constraints on the possible fundamental groups coming from global considerations, namely from looking at the relation between the heterotic boundary and bounding M-theory. We first interpret this relation as a cobordism which connects one boundary component to the other through the eleven-dimensional bulk. We take one of the two boundary components and the bulk to be of the same homotopy type. It is natural to ask when such cobordisms are trivial, that is when are they of product (or “cylinder") form, as is usually the case in heterotic M-theory. When the fundamental groups of both the eleven-manifold $Y^{11}$ and the ten-manifold $M^{10}$ are trivial then we consider the cobordism as an $h$-cobordism ($h$ is for homotopy). When the fundamental groups are equal but nontrivial then we view heterotic M-theory as an $s$-cobordism ($s$ is for simple homotopy). Since the dimension of the nontrivial part of $M^{10}$, namely the Calabi-Yau threefold, is six then the h-cobordism [@Mil0] and the s-cobordism [@Ker] [@Maz] [@Sta] theorems can be applied. In both cases we are assuming that inclusions of the boundaries in $Y^{11}$ are homotopy equivalences. The case when $\pi_1(Y^{11})$ is nontrivial is discussed extensively in [@DMW-Spinc] in relation the partition functions and to type IIA string theory. The obstruction to finding a cobordism that is of the cylinder type is the Whitehead torsion of the inclusion $\tau (Y^{11}, M^{10})$, which is an element of the Whitehead group of the fundamental group ${\rm Wh}(\pi_1(M^{10}))$. The Whitehead group is extensively studied and is well-known for finite groups (see [@Ol]), which is the case we mainly study as such groups seem to be the most interesting for model building. Given our identification of heterotic M-theory as an $s$-cobordism, we are able to identify fundamental groups that allow trivial cobordisms from the ones which do not. We view this as providing global consistency constraints on heterotic compactifications in view for model building. We summarize the main point of this article with Consider heterotic M-theory with the $E_8$ heterotic string theory on each of the two the boundary components. Then $(i)$. M-theory is an $s$-cobordism for one of the two connected components of the boundary. $(ii)$. Consistency requires the Whitehead torsion, in the Whitehead group of the integral group ring of the fundamental group of the boundary component, to vanish. Since the use of $h$- and $s$-cobordism and the Whitehead torsion is novel in the context of heterotic M-theory and is perhaps not widely known in theoretical physics in general, we choose to take an expository route to arrive at our conclusions. We provide the description of heterotic M-theory in terms of $h$ and $s$-cobordism in section \[sec cob\]. Then in section \[sec ww\] we look at constraints on the fundamental group, coming from the Whitehead group in section \[sec Wh\] and from the Whitehead torsion in section \[sec tau\]. We provide many examples along the way and then in section \[sec ex\] we consider representative examples explicitly appearing in model building. We then consider the dynamical aspects in section \[sec dyn\], emphasizing the main points of this article. We first consider automorphisms, including diffeomorphisms and issues of orientation, in section \[sec aut\], and then we consider dynamical aspects of compactifications in section \[sec com\]. Heterotic M-theory as an $h$-cobordism and $s$-cobordism {#sec cob} ======================================================== In this section we set up heterotic M-theory as a cobordism, first as an $h$-cobordism and then as $s$-cobordism. Viewed from M-theory the data involves an eleven-dimensional manifold $Y^{11}$ which is a product $[0,1]\times M^{10}$ together with an $E_8$ bundle on each of $M^{10} \times \{0\}$ and $M^{10}\times \{1\}$. We will consider this from a ten-dimenional point of view, where we will have a cobordism taking one boundary component to the other. #### $H$-cobordism. A compact connected eleven-manifold $Y^{11}$ whose boundary $\partial Y^{11}$ is the disjoint union of two closed manifolds $M^{10}$ and $M'^{10}$, $\partial Y^{11}=M^{10} \cup M'^{10}$, is called an $h$-cobordism, provided the inclusions of $M^{10}$ into $Y^{11}$ and of $M'^{10}$ into $Y^{11}$ are both homotopy equivalences. The pair $(Y^{11}, M^{10})$ is called a $h$-cobordism with base $M^{10}$ and top $M'^{10}$. A smooth $h$-cobordism is one where $Y^{11}$ is a smooth manifold. A trivial or product $h$-cobordism is of the form $M^{10} \times [0,1]$. If $Y^{11}$ is simply-connected, then the $h$-cobordism theorem can be applied (see [@Mil0]) to give that $Y^{11}$ is diffeomorphic to the product $M^{10}\times [0,1]$. This is the configuration that is usually considered in heterotic M-theory [@HW1] [@HW2]. We can consider a more detailed description, which will be useful in section \[sec tau\] and section \[sec dyn\]. An eleven-dimensional cobordism $(Y^{11}; M_0^{10}, f_0, M_1^{10}, f_1)$ consists of a compact oriented eleven-manifold $Y^{11}$, two closed ten-manifolds $M_0^{10}$ and $M_1^{10}$, a disjoint decomposition $\partial Y^{11}=\partial_0 Y^{11} \coprod \partial_1Y^{11}$ of the boundary $\partial Y^{11}$ of $Y^{11}$ and orientation-preserving diffeomorphisms $f_0: M_0^{10} \to \partial_0Y^{11}$ and $f_1: (M_1^{10})^- \to \partial_1Y^{11}$. By $X^{-}$ we mean the manifold $X$ taken with the opposite orientation. On the boundary $\partial Y^{11}$ we use the orientation with respect to the decomposition $TY^{11}=T\partial Y^{11} \oplus {{\mathbb R}}$ coming from an inward normal field to the boundary. If $\partial_0Y^{11}=M_0^{10}$, $\partial_1Y^{11}=(M_1^{10})^-$, and $f_0$ and $f_1$ are the identity maps, then the $h$-cobordism can be referred to as $(Y^{11};\partial_0Y^{11}, \partial_1Y^{11})$. An $h$-cobordism over $M_0^{10}$ is trivial if it is diffeomorphic relative $M_0^{10}$ to the trivial $h$-cobordism $(M_0^{10}\times [0,1]; M_0^{10}\times \{0\}, (M_0^{10}\times \{1\})^{-})$. #### The fundamental group. The $h$-cobordism theorem can be applied only when the fundamental group is trivial. Next we consider the more interesting case when the fundamental group is not necessarily trivial. We will assume that $\pi_1(Y^{11})\cong \pi_1(M^{10})$. The fundamental group functor takes products to products, that is, the fundamental group is multiplicative. For $M^{10}={{\mathbb R}}^{1,3}\times X$, we have $ \pi_1(M^{10})\cong \pi_1({{\mathbb R}}^{1,3}) \times \pi_1(X) \cong \pi_1(X)$, so that the fundamental group of $M^{10}$ is determined by that of the Calabi-Yau threefold $X$. The generalization from Minkowski to other four-dimensional spacetimes gives an obvious modification, which depends on whether or not the latter is simply connected. Next we consider the appropriate description of heterotic M-theory when $\pi_1(M^{10})\neq 0$. By our assumption, this is equivalent to taking $\pi_1(Y^{11}) \neq 0$, considered in [@DMW-Spinc]. #### $S$-cobordism. Let $M^{10}$ be a connected compact 10-manifold with fundamental group ${\Gamma}$, and consider the family ${\ensuremath{\mathcal F}}$ of all $h$-cobordisms built on $M^{10}$. These are connected compact 11-manifolds $Y^{11}$ with exactly two boundary components, one of which is $M^{10}$ and the other of which is some other manifold $M'^{10}$ such that $Y^{11}$ is homotopy equivalent to both $M^{10}$ and $M'^{10}$. There is a map $\tau : \mathcal{F} \to {\rm Wh}({\Gamma})$ called the [*Whitehead torsion*]{} which induces a natural one-to-one correspondence from $\mathcal{F}/\sim$ to ${\rm Wh}({\Gamma})$, where $\sim$ is the equivalence relation induced by diffeomorphisms $Y^{11} \to Y'^{11}$ which are the identity on $M^{10}$. If $Y^{11}$ is the “trivial" $h$-cobordism $Y^{11}=M^{10} \times [0,1]$, then $\tau (Y^{11})=1$. This is an application of the Barden-Mazur-Stallings theorem [@Ker] [@Maz] [@Sta] (see [@Ro] for a review). If the fundamental group ${\Gamma}$ is such its Whitehead group ${\rm Wh}({\Gamma})$ is trivial, then certainly the Whitehead torsion will vanish and we are back to the case of an $h$-cobordism. Consequently, $Y^{11}$ is diffeomorphic (relative $M^{10}$) to a product $M^{10} \times [0,1]$. In particular, the other boundary component $M'^{10}$ is diffeomorphic to $M^{10}$. There is a bijection, given by the Whitehead torsion $\tau (Y^{11}, M^{10})$, between the set of diffeomorphism classes of $h$-cobordisms $(Y^{11},M^{10})$ with a given base $M^{10}$ and the set ${\rm Wh}(\pi_1(M^{10}))$. The cylinder corresponds to 0 under this bijection. We will consider this in much more detail in the following sections. Note that there are several versions of the the $s$-cobordism (and $h$-cobordism) theorem depending on the category of spaces within which we are working; for example we could work with homeomorphisms rather than diffeomorphisms (but here we are assuming all spaces to be smooth). However, if we start with a homotopy equivalence then we might not be able to extend it to a diffeomorphism. Consider a large class of manifolds called [*aspherical*]{}, which are ones for which all homotopy groups vanish except the first one, i.e. the fundamental group. Let $Y$ and $M$ be aspherical spaces and let $\alpha: \pi_1( M) \to \pi_1(Y)$ be an isomorphism. Then, by the Theorem of Hurewicz, $\alpha$ is induced by a homotopy equivalence. It is an open conjecture of Borel from 1955 that this can be extended to a homeomorphism. The strengthening to smooth manifolds fails [@DH]. Note that we can work in category of spaces other than that of smooth manifolds, since the $h$- and $s$-cobordism arguments work for piecewise linear (PL) and topological spaces. This implies, for example, that orbifolds are also included in our discussion, for which we would choose the category of topological spaces. The Whitehead group and Whitehead torsion {#sec ww} ========================================= We now consider the Whitehead group and Whitehead torsion in our setting of heterotic M-theory via algebraic K-theory of the group ring of the fundamental group and give the main properties which are useful for us. The Whitehead group {#sec Wh} ------------------- Algebraic K-theory roughly characterizes how, in passing from a field to an arbitrary ring, notions of linear algebra related to the general linear group and vector spaces might extend. One measure of failure of such an extension is $K_1(R)$, the algebraic K-theory of an associative ring $R$. Let $\widetilde{K}_1(R)$ be the cokernel of the map $K_1({{\mathbb Z}}) \to K_1(R)$ induced by the canonical ring homomorphism ${{\mathbb Z}}\to R$. Since ${{\mathbb Z}}$ is a ring with Euclidean algorithm then the homomorphism det: $K_1({{\mathbb Z}}) \to \{ \pm\}$, given by $[A] \mapsto {\rm det}(A)$, is a bijection. Hence $\widetilde{K}_1(R)$ is the quotient of $K_1(R)$ by a cyclic group of order two generated by the class of the $1\times1$-matrix $(-1)$. We are interested in the case when $R$ is a group algebra ${{\mathbb Z}}[\Gamma]$ of the fundamental group $\Gamma=\pi_1(M^{10})$, that is in integer linear combinations of elements of ${\Gamma}$. Define the [*Whitehead group*]{} ${\rm Wh}(\Gamma)$ of a group $\Gamma$ to be the cokernel of the map $\Gamma \times \{\pm \} \to K_1({{\mathbb Z}}[\Gamma])$ which sends $({{\gamma}}, \pm 1)$ to the class of the invertible $1\times1$-matrix $(\pm {{\gamma}})$. In other words, ${\rm Wh}({\Gamma})$ is the quotient of $K_1({{\mathbb Z}}[{\Gamma}])$ by the image of $\{ \pm {{\gamma}}: {{\gamma}}\in {\Gamma}\}$, that is $ {\rm Wh}({\Gamma})= K_1({{\mathbb Z}}[{\Gamma}]) / \{ \pm {{\gamma}}: {{\gamma}}\in {\Gamma}\}\;. $ The zero element $0 \in {\rm Wh}({\Gamma})$ is represented by the identity matrix $I_n$ for any positive integer $n$. Note that one can define the Whitehead group of the fundamental group by choosing a base point, as is usual in the fundamental group. However, the end result will be independent of the choice of the base point. Therefore, one should think of $\pi_1(M)$ in ${\rm Wh}(\pi_1(M))$ as the [*fundamental groupoid*]{} of $M$. Note also that the Whitehead group can be viewed either additively or multiplicatively. In the first point of view, this corresponds to adding two cobordisms by connecting one ‘cylinder’ to another over a ten-dimensional section, while an instance of the second point of view is a ’flip’. #### Example 1. Trivial case. Consider the case when the fundamental group is trivial. Then the group algebra ${{\mathbb Z}}[1]={{\mathbb Z}}$ is a ring with a Gaussian algorithm, so that the determinant induces an isomorphism $K_1({{\mathbb Z}}) \buildrel{\cong}\over{\to} \{\pm\}$ and the Whitehead group ${\rm Wh}(\{1\})$ of the trivial group vanishes. Hence any $h$-cobordism over a simply-connected closed $M^{10}$ is trivial. Thus, as expected in this case, $s$-cobordism reduces to $h$-cobordism. #### Example 2. Finite cyclic groups. ${\rm Wh}({\Gamma})$ is torsion-free for a finite cyclic group. For example, ${\rm Wh}({{\mathbb Z}}_p)$, $p$ odd prime, is the free abelian group of rank $(p-3)/2$ and ${\rm Wh}({{\mathbb Z}}_2)=0$. We will consider many more examples in section \[sec ex\]. #### Properties of the Whitehead group. We are interested in the case when the fundamental group $\Gamma$ is a finite group. For such a group the following useful properties hold [@Mil] [@ADS] [@Ol] 1. [*Functoriality:*]{} ${\rm Wh}(\Gamma)$ is a covariant functor of $\Gamma$, that is, any homomorphism $f: \Gamma_1 \to \Gamma_2$ induces a homomorphism $f_*: {\rm Wh}(\Gamma_1) \to {\rm Wh}(\Gamma_2)$. 2. [*Trivial group:*]{} Let ${\Gamma}=\pi_1(M)$ be trivial. Then from $K_1({{\mathbb Z}})={{\mathbb Z}}_2$ one gets ${\rm Wh}(\pi_1(M)) = {\rm Wh}(1)=1$. This is example 1 above. 3. [*Low rank:*]{} Whitehead showed that ${\rm Wh}(\Gamma)=1$ if $|\Gamma| \leq 4$. This implies, for instance, that ${{\mathbb Z}}_2$, ${{\mathbb Z}}_3$, ${{\mathbb Z}}_4$ and ${{\mathbb Z}}_2 \times {{\mathbb Z}}_2$ have trivial Whitehead group and hence lead to (desirable) trivial $h$-cobordisms. 4. [*Rank:*]{} By a result of Bass, ${\rm Wh}(\Gamma)$ is a finitely generated abelian group of rank $r(\Gamma)-q(\Gamma)$, where $r(\Gamma)$ is the number of irreducible real representations of $\Gamma$ and $q(\Gamma)$ is the number of irreducible rational representations of $\Gamma$. Explicitly, $q(\Gamma)$ is the number of conjugate classes of cyclic subgroups of $\Gamma$ and $r(\Gamma)$ is the number of conjugate classes of unordered pairs $\{ {{\gamma}}, {{\gamma}}^{-1}\}$. 5. [*Free product:*]{} The Whitehead group of a free product is multiplicative, ${\rm Wh}(G \ast H)={\rm Wh}(G) \oplus {\rm Wh}(H)$. Unfortunately, there is no corresponding formula for Cartesian products. For example, ${\rm Wh}({{\mathbb Z}}_3)=0$ and ${\rm Wh}({{\mathbb Z}}_4)=0$ but ${\rm Wh}({{\mathbb Z}}_3 \times {{\mathbb Z}}_4) \cong {{\mathbb Z}}$. More on this will be discussed in section \[sec ex\]. #### The torsion subgroup. We have seen above that the Whitehead group of a cyclic group ${{\mathbb Z}}_p$ of prime order $p$ is torsion-free. While these groups form an important class of fundamental groups we are considering, we should consider other cases as well. In particular, there could be groups ${\Gamma}$ for which ${\rm Wh}({\Gamma})$ is torsion. The torsion in the algebraic K-group is ${\rm Tor}(K_1({{\mathbb Z}}[{\Gamma}]))=(\pm) \times {\Gamma}^{\rm ab} \times SK_1({{\mathbb Z}}[{\Gamma}])$, where ${\Gamma}^{\rm ab}$ is the abelianization of ${\Gamma}$ (that is the first homology group $H_1(M^{10})$) and $SK_1({{\mathbb Z}}[{\Gamma}])=\ker (K_1({{\mathbb Z}}[{\Gamma}]) \to K_1({{\mathbb Q}}[{\Gamma}]))$. This kernel of the change of coefficients homomorphism is the full torsion subgroup of ${\rm Wh}({{\mathbb Z}}[{\Gamma}])$. #### Properties of the torsion subgroup. The torsion subgroup $SK_1({{\mathbb Z}}[{\Gamma}])$ of ${\rm Wh}({\Gamma})$ is highly nontrivial [@ADS] [@Wa74] [@Ol]. Some of the useful properties are 1. The torsion subgroup of ${\rm Wh}(\Gamma)$ is isomorphic to $SK_1({{\mathbb Z}}[\Gamma])$. 2. The torsion in ${\rm Wh}({\Gamma})$ comes from $SL(2,{{\mathbb Z}}[{\Gamma}])$. 3. $SK_1({{\mathbb Z}}[\Gamma])$ is non-vanishing for all groups of the form $\Gamma \cong ({{\mathbb Z}}_p)^n$, $n\geq 3$ and $p$ an odd prime. 4. $SK_1({{\mathbb Z}}[\Gamma])=1$ if $\Gamma \cong {{\mathbb Z}}_{p^n}$ or ${{\mathbb Z}}_{p^n}\times {{\mathbb Z}}_p$ (for any prime $p$, and any $n$), if $\Gamma\cong ({{\mathbb Z}}_2)^n$ (any $n$), or if $\Gamma$ is any dihedral, quaternion, or semidihedral 2-group. 5. The classes of finite groups $\Gamma$ for which ${\rm Wh}(\Gamma)=1$, or $SK_1({{\mathbb Z}}[\Gamma])=1$, are [*not*]{} closed under products. This provides many nontrivial examples using products. For a finitely generated fundamental group ${\Gamma}$ the vanishing of the Whitehead group ${\rm Wh}({\Gamma})$ is equivalent to the statement that each $h$-cobordism over a closed connected $M^{10}$ is trivial. Knowing that all $h$-cobordisms over a given manifold are trivial is useful, but strong. Alternatively, we could have ${\rm Wh}({\Gamma})$ nontrivial yet the distinguished element, the Whitehead torsion $\tau$ is zero. Whitehead torsion {#sec tau} ----------------- The Whitehead torsion, which is essentially a linking matrix for handles in the handle decomposition of the manifold, serves as an obstruction to the reduction of an $h$-cobordism to a product. We have encountered above many situations where the Whitehead group is not trivial. In certain cases these elements, including the distinguished element given by the Whitehead torsion, can be characterized. This characterization can be geometric due to the realization theorem which says that every Whitehead torsion comes from a manifold (see [@Ker]). First, note that a map $f: Y^{11}\to M_0^{10}$ induces a homomorphism $f_*: {\rm Wh}(\pi_1(Y^{11})) \to {\rm Wh}(\pi_1(M_0^{10}))$ on the corresponding Whitehead groups such that ${\rm id}_*={\rm id}$, $(g \circ f)_*=g_* \circ f_*$, and $f \simeq g$ implies that $f_*=g_*$. Next, the Whitehead torsion of our eleven-dimensional $h$-cobordism $(Y^{11}; M_0^{10}, f_0, M_1^{10}, f_1)$ over $M_0^{10}$, $ \tau (Y^{11}, M_0^{10}) \in {\rm Wh}(\pi_1(M_0^{10}))\;, $ is defined to be the preimage of the Whitehead torsion $\tau (M_0^{10}\buildrel{f_0}\over{\longrightarrow} \partial_0Y^{11} \buildrel{\iota_0}\over{\longrightarrow} Y^{11})\in {\rm Wh}(\pi_1(Y^{11})$ under the isomorphism $(\iota_0 \circ f_0)_*: {\rm Wh}(\pi_1(M_0^{10})) \buildrel{\cong}\over{\longrightarrow} {\rm Wh}(\pi_1(Y^{11}))$, where $\iota_0: \partial_0Y^{11} \hookrightarrow Y^{11}$ is the inclusion (see [@KL]). Next we will consider the simple situation when the diffeomorphisms are the identity. #### Geometric definition of Whitehead torsion. There is a description of Whitehead torsion at the level of chain complexes [@Mil] [@Co]. Let ${\ensuremath{\mathcal W}}(M^{10})$ be the collection of all pairs of finite complexes $(Y^{11}, M^{10})$ such that $M^{10}$ is a strong deformation retract of $Y^{11}$. For any two objects $(Y^{11}_1, M^{10})$, $(Y^{11}_2, M^{10}) \in {\ensuremath{\mathcal W}}$ define an equivalence $(Y^{11}_1, M^{10}) \sim (Y^{11}_2, M^{10})$ if and only if $Y^{11}_1$ and $Y^{11}_2$ are simple homotopically equivalent relative to the subcomplex $M^{10}$. Define ${\rm Wh}(M^{10})= {\ensuremath{\mathcal W}}/\sim$ and let $[Y^{11}_1, M^{10}]$ and $[Y^{11}_2, M^{10}]$ be two classes in ${\rm Wh}(M^{10})$. For $Y^{11}_1 \bigsqcup_{M^{10}} Y^{11}_2$, the disjoint union of $Y^{11}_1$ and $Y^{11}_2$ identified along the common subcomplex $M^{10}$, an abelian group structure can be defined on the Whitehead group ${\rm Wh}(M^{10})$ by $[Y^{11}, M^{10}] \oplus [Y^{11}_2, M^{10}]= [Y^{11}_1 \bigsqcup_{M^{10}} Y^{11}_2, M^{10}]$. The universal cover $(\widetilde{Y}^{11}, \widetilde{M}^{10})$ of an element $(Y^{11}, M^{10})$ in ${\ensuremath{\mathcal W}}$ can be equipped with the CW-complex structure lifted from the CW-structure of $(Y^{11}, M^{10})$. The inclusion $\widetilde{M}^{10} \subset \widetilde{Y}^{11}$ is a homotopy equivalence. Let $C_*(\widetilde{Y}^{11}, \widetilde{M}^{10})$ be the cellular chain complex of $(\widetilde{Y}^{11}, \widetilde{M}^{10})$. The covering action of $\pi_1(Y^{11})$ on $(\widetilde{Y}^{11}, \widetilde{M}^{10})$ induces an action on $C_*(\widetilde{Y}^{11}, \widetilde{M}^{10})$ and makes it a finitely generated free acyclic chain complex of ${{\mathbb Z}}[\pi_1(Y^{11})]$-modules. In addition to the boundary map $\partial$, there is a contraction map $\delta$ of degree $+1$ on $C_*(\widetilde{Y}^{11}, \widetilde{M}^{10})$ such that $\partial \delta + \delta \partial ={\rm id}$ and $\delta^2=0$. The module homomorphism $\partial + \delta: \bigoplus_{i=0}^\infty C_{2i+1}(\widetilde{Y}^{11}, \widetilde{M}^{10}) \to \bigoplus_{i=0}^\infty C_{2i}(\widetilde{Y}^{11}, \widetilde{M}^{10}) $ is an isomorphism of ${{\mathbb Z}}[\pi_1(Y^{11})]$-modules. The image and the range of this homomorphism are finitely generated free modules with a basis we choose coming from the CW-structure on $(\widetilde{Y}^{11}, \widetilde{M}^{10})$. Consider the matrix of this homomorphism $\partial + \delta$ which is an invertible matrix with entries in ${{\mathbb Z}}[\pi_1(Y^{11})]$ and hence lies in $GL(n, {{\mathbb Z}}[\pi_1(Y^{11})])$ for some $n$. Now take the image of this matrix in ${\rm Wh}(\pi_1(Y^{11}))$ via an isomorphism $\tau$, sending $(Y^{11}, M^{10})$ to ${\rm Wh}(\pi_1(Y^{11}))$. More explicitly, let $ \cdots \buildrel{\partial}\over{\longrightarrow} C_{i+1} \buildrel{\partial}\over{\longrightarrow} C_{i} \buildrel{\partial}\over{\longrightarrow} \cdots C_{0} \buildrel{\partial}\over{\longrightarrow} 0 $ be the complex which calculates the homology $H_*(Y^{11}, M^{10};{{\mathbb Z}}[{\Gamma}])$ of the inclusion $M^{10} \subset Y^{11}$. Each $C_i$ is a finitely generated free ${{\mathbb Z}}[{\Gamma}]$-module. Up to orientation and translation by an element in ${\Gamma}$, each $C_i$ has a preferred basis over ${{\mathbb Z}}[{\Gamma}]$ coming from the $i$-simplices added to get from $M^{10}$ to $Y^{11}$ in some triangulation of the universal covering spaces. The group $Z_i$ of $i$-cycles is the kernel of $\partial : C_i \to C_{i-1}$ and the group $B_i$ of $i$-boundaries is the image of $\partial: C_{i+1} \to C_i$. Since $M^{10} \subset Y^{11}$ is a deformation retract, homotopy invariance of homology gives that $H_*=0$, so that $B_*=Z_*$. Let ${\ensuremath{\mathcal M}}_i \in GL({{\mathbb Z}}[\pi_1(M^{10})])$ be the matrices representing the isomorphism $B_i \oplus B_{i-1} \cong C_i$ coming from a choice of section $0 \to B_i \to C_i \to B_{i-1} \to 0$. Let $[{\ensuremath{\mathcal M}}_i] \in {\rm Wh}(\pi_1(M^{10}))$ be the corresponding equivalence classes. The Whitehead torsion is then $ \tau(Y^{11}, M^{10})= \sum (-1)^i [{\ensuremath{\mathcal M}}_i] \in {\rm Wh}(\pi_1(M^{10}))\;. $ Note that the Whitehead group is identified as a quotient of $K_1({{\mathbb Z}}[{\Gamma}])$ by the subgroup generated by the units of the form $\pm {{\gamma}}$ for ${{\gamma}}\in {\Gamma}=\pi_1(M_0^{10})$. In the present context, this ensures the independence of the choice of ${{\mathbb Z}}[{\Gamma}]$-basis within the cellular equivalence class of ${{\mathbb Z}}[{\Gamma}]$-bases. #### Properties of Whitehead torsion. The Whitehead torsion has existence and uniqueness properties. 1\. [*Existence.*]{} Given $\alpha \in {\rm Wh}(\pi_1(M^{10})$, there exists an $h$-cobordism $Y^{11}$ with $\tau (Y^{11})=\alpha$. This implies that if the Whitehead group is nontrivial then we can find a cobordism for every element in that group. In order to get a trivial $h$-cobordism, that is one of cylinder type, we have to make sure that the element ${\rm Wh}({\Gamma})$ we identify for our spaces will be the zero element. This is of course not guaranteed to occur. 2\. [*Uniqueness.*]{} $\tau (Y^{11})= \tau (Y'^{11})$ if and only if there exists a diffeomorphism $f: Y^{11} \to Y'^{11}$ such that $f|_M={\rm id}_M$. This tells us that we are allowed to “deform" $Y^{11}$ in a nice way and still be able to get the same type of cobordism. In particular, for $Y^{11}$ with $\tau(Y^{11})=0$ we can always find a diffeomorphic $Y'^{11}$ for which the property that the Whitehead torsion is zero is preserved. #### Elements of finite order in the Whitehead group. We have seen that the Whitehead group of products of finite cyclic groups may contain torsion. Elements of finite order can be characterized as follows [@Mil]. Consider an orthogonal representation ${\Gamma}\to O(n)$ of the finite group ${\Gamma}$. This representation gives rise to a ring homomorphism $\rho: {{\mathbb Z}}[{\Gamma}] \to {\rm \mathbb{M}}_n ({{\mathbb R}})$, where ${\rm \mathbb{M}}_n({{\mathbb R}})$ is the algebra of $n \times n$ matrices over the real numbers. This induces a group homomorphism $\rho_*: \widetilde{K}_1({{\mathbb Z}}[\pi]) \to \widetilde{K}_1({\rm \mathbb{M}}_n({{\mathbb R}}))\cong \widetilde{K}_1({{\mathbb R}}) \cong {{\mathbb R}}^+$. Since ${{\mathbb R}}^+$ has no elements of finite order then there is the corresponding homomorphism ${\rm Wh}({\Gamma}) \to {{\mathbb R}}^+$. Therefore, an element $\omega \in {\rm Wh}({\Gamma})$ has finite order if and only if $\rho_*(\omega)=1$ for every orthogonal representation $\rho$ of ${\Gamma}$. #### Elements of ${\rm Wh}({\Gamma})$ as matrices and the representation dimension. Nontrivial elements of the Whitehead group can be represented by matrices, usually of small size. The [*representation dimension*]{} of a group ${\Gamma}$ is said to be less than or equal to $m$, with notation $r$-$\dim {\Gamma}\leq m$, if every element of ${\rm Wh}({\Gamma})$ can be realized as a matrix in $GL(m, {{\mathbb Z}}[{\Gamma}])$. If ${\Gamma}$ is finite then $r$-$\dim {\Gamma}\leq 2$. Furthermore, the representation dimension of the finite group ${\Gamma}$ satisfies $r$-$\dim {\Gamma}\leq 1$ if and only if ${\Gamma}$ admits no epimorphic mapping onto the following (see [@Sh]) [*1.*]{} the generalized quaternion group, [*2.*]{} the binary tetrahedral, octahedral, or icosahedral groups, [*3.*]{} and the groups ${{\mathbb Z}}_{p^2} \times {{\mathbb Z}}_{p^2}$, ${{\mathbb Z}}_p \times {{\mathbb Z}}_p \times {{\mathbb Z}}_p$, $Z_p \times {{\mathbb Z}}_2 \times {{\mathbb Z}}_2 \times {{\mathbb Z}}_2$, ${{\mathbb Z}}_4 \times {{\mathbb Z}}_2 \times {{\mathbb Z}}_2$, and ${{\mathbb Z}}_4 \times {{\mathbb Z}}_4$, for $p$ a prime. Thus $r$-$\dim {\Gamma}\leq 1$ for all finite simple groups. However, if we take products then the size of the matrix can grow (see expression for an explicit matrix). Further examples in heterotic M-theory {#sec ex} ======================================= We have already seen many classes of examples both for the Whitehead group in section \[sec Wh\] and for the Whitehead torsion in section \[sec tau\]. we now provide more examples and in particular ones which appear explicitly in model building (cf. the introduction). #### Tori and free abelian groups. The fundamental group of the circle is the free abelian group ${{\mathbb Z}}$, so that the corresponding Whitehead torsion is zero, ${\rm Wh} ({{\mathbb Z}})=0$. For the $n$-torus $T^n$, the fundamental group $\pi_1(T^n)={{\mathbb Z}}^n$. This free abelian group of rank $n$ has a trivial Whitehead torsion ${\rm Wh}(\pi_1(T^n))=0$, since ${\rm Wh}({{\mathbb Z}}\oplus \cdots \oplus {{\mathbb Z}})=0$ by the multiplicative property of Whitehead torsion under free product (section \[sec Wh\]). It follows from the theorem of Bass about the rank of the Whitehead group that ${\rm Wh}(\Gamma)$ of a free abelian group $\Gamma$ is zero if and only if $\Gamma$ has exponent 1, 2, 3, 4, or 6 [@Mil]. #### Cyclic groups. Suppose ${\Gamma}$ is a finite group. Then ${\rm Wh}({\Gamma})$ is finitely generated, and ${\rm rank}({\rm Wh}({\Gamma}))$ is the difference between the number of irreducible representations of ${\Gamma}$ over ${{\mathbb R}}$ and the number of irreducible representations of ${\Gamma}$ over ${{\mathbb Q}}$. For $\Gamma$ a cyclic group ${{\mathbb Z}}_p$ of order $p$, an odd prime, the numbers of representations are $q({{\mathbb Z}}_p)=2$ and $r({{\mathbb Z}}_p)=\frac{1}{2}(p+1)$, respectively. This implies that ${\rm Wh}({{\mathbb Z}}_p)$ is the free abelian group of rank $(p-3)/2$ and that ${\rm Wh}({{\mathbb Z}}_2)=0$. Alternatively, note that ${{\mathbb Z}}_p$ has $(p-1)/2$ inequivalent two-dimensional irreducible representations over ${{\mathbb R}}$, but one $(p-1)$-dimensional irreducible representation over ${{\mathbb Q}}$ (since ${{\mathbb Q}}[{{\mathbb Z}}_p] \cong {{\mathbb Q}}\times {{\mathbb Q}}(\zeta)$, $\zeta$ a primitive $p$-th root of unity, and $[{{\mathbb Q}}(\zeta) : {{\mathbb Q}}]=p-1$), so ${\rm rank}({\rm Wh}({{\mathbb Z}}_p))=\frac{p-1}{2}+1 -2= (p-3)/2$. Note that we have already seen that ${\rm Wh}({{\mathbb Z}}_k)=0$ for $k=2,3,4,6$. #### Units in the group ring. Consider the integral group ring ${{\mathbb Z}}[{{\mathbb Z}}_p]$ of the finite cyclic group ${{\mathbb Z}}_p$ and let $\zeta$ be a primitive $p$th root of unity with corresponding group ring ${{\mathbb Z}}[\zeta]$. The pullback square of rings $ \xymatrix{ {{\mathbb Z}}[{{\mathbb Z}}_p] \ar[rr] \ar[d] && {{\mathbb Z}}[\zeta] \ar[d] \\ {{\mathbb Z}}\ar[rr] && \mathbb{F}_p }\;, $ where $\mathbb{F}_p$ is the field with $p$ elements, implies that the $(p-1)$st power of any unit in ${{\mathbb Z}}[\zeta]$ comes from a unit in ${{\mathbb Z}}[{{\mathbb Z}}_p]$. An example of a unit in ${{\mathbb Z}}[\zeta]$ is $(\zeta + \zeta^{-1})^r$. This is invariant under complex conjugation in ${{\mathbb Z}}[\zeta]$ (this corresponds to invariance under the orientation duality discussed in section \[sec aut\]). #### The quintic and the cyclic group of order 5. The quintic threefold plays an important role as a prototype example of compactification on Calabi-Yau manifolds. Consider the one-parameter family of quintic threefolds ${\ensuremath{\mathcal Q}}:= \{ z_1^5 + z_2^5 +z_3^5 + z_4^5 + z_5^5 + \psi^5 z_1z_2 z_3 z_4 z_5=0\} \subset {{\mathbb C}}P^4$. The defining equation is invariant under the ${{\mathbb Z}}_5 \times {{\mathbb Z}}_5 \subset {\rm PGL}(5,{{\mathbb C}})$ group action $ [z_1 : z_2 : z_3 : z_4 : z_5]\mapsto [z_2 : z_3 : z_4 : z_5 : z_1]\;, \qquad [z_1 : z_2 : z_3 : z_4 : z_5]\mapsto [\zeta z_1 : \zeta^2 z_2 : \zeta^3 z_3 : \zeta^4 z_4 : z_5]\;, $ where $\zeta=e^{2\pi i/5}$. The fixed points lie on ${{{\mathbb C}}\text{P}}^4-{\ensuremath{\mathcal Q}}$, so that ${\ensuremath{\mathcal Q}}/{{\mathbb Z}}_5$ and ${\ensuremath{\mathcal Q}}/{{\mathbb Z}}_5\times {{\mathbb Z}}_5$ are smooth Calabi-Yau threefolds. The six different ${{\mathbb Z}}_5$ subgroups in ${{\mathbb Z}}_5 \times {{\mathbb Z}}_5$ can be used. The Whitehead group of ${{\mathbb Z}}_5=\{t~|~t^5=1\}$ is ${\rm Wh}({{\mathbb Z}}_5)={{\mathbb Z}}$ with generator the torsion $\tau (u)$ of the unit $u=1-t+t^2 \in {{\mathbb Z}}[{{\mathbb Z}}_5]$ [@Mil]. The identity $(t+t^{-1} -1)(t^2 + t^{-2} -1)=1$ indeed shows that $u$ is a unit. The homomorphism $\alpha: {{\mathbb Z}}[{{\mathbb Z}}_5] \to {{\mathbb C}}$, sending $t$ to $\zeta$, also sends $\{\pm {{\gamma}}: {{\gamma}}\in {\Gamma}\}$ to the roots of unity in ${{\mathbb C}}$, and hence $x \mapsto |\alpha (x)|$ defines a homomorphism from ${\rm Wh}({{\mathbb Z}}_5)$ into ${{\mathbb R}}^*_+$, the nonzero positive real numbers. Then the map $u \mapsto 1- \zeta - \zeta^{-1} =1- 2\cos(2\pi/5)$ can be used to show that no power of $u$ is equal to 1. Indeed, $|\alpha (u)|= | 1- 2\cos(2\pi/5)| \approx 0.4$, so that $\alpha$ defines an element of infinite order in ${\rm Wh}({{\mathbb Z}}_5)$. Note that the unit $u$ is self-conjugate, and that the automorphism $t \mapsto t^2$ of ${{\mathbb Z}}_5$ carries $u$ to $u^{-1}$. In fact, for $\Gamma$ finite abelian, every element of ${\rm Wh}(\Gamma)$ is self-conjugate [@Mil] (see the last paragraph in section \[sec tau\]). We see from the example of the quintic that, a priori, there are countably infinitely many elements in the Whitehead group of the fundamental group of the quintic. Unless the Whitehead torsion is the zero element, there will be an obstruction to having a trivial $h$-cobordism and hence to a consistent relation to heterotic M-theory. Therefore, it is an interesting problem to compute the Whitehead torsion of the quintic. Recall from the end of section \[sec Wh\] that the full torsion subgroup of the Whitehead group is given by $SK_1({{\mathbb Z}}[{\Gamma}])$. Therefore, one way to tell that ${\rm Wh}({\Gamma})$ is nontrivial is to detect torsion via $SK_1({{\mathbb Z}}[{\Gamma}])$. #### Products of abelian groups. We now consider products of abelian groups, in particular of cyclic groups. [*1. Products of groups of even order.*]{} For even order, we have already seen that the Whitehead group of the lowest rank non-simple group, ${{\mathbb Z}}_2 \times {{\mathbb Z}}_2$, is zero. Next we consider products of ${{\mathbb Z}}_2$ with ${{\mathbb Z}}_4$ and so on. We use the following two general formulae [@Ol] for the torsion part of the Whitehead group $ SK_1({{\mathbb Z}}[ ({{\mathbb Z}}_2)^k \times {{\mathbb Z}}_{2^n}])\cong \left[ \oplus_{r=1}^k \binom{k}{r} \cdot ({{\mathbb Z}}_{2^{r-1}}) \right] \oplus \left[ \oplus_{s=2}^n ({{\mathbb Z}}_{2^s})\right]$ and $SK_1({{\mathbb Z}}[({{\mathbb Z}}_2)^2 \times {{\mathbb Z}}_{2^n}]) \cong {{\mathbb Z}}_2^{n-1}$. For instance, the following cases can then be deduced: [*1.*]{} $SK_1({{\mathbb Z}}[{{\mathbb Z}}_4 \times {{\mathbb Z}}_4])\cong {{\mathbb Z}}_2$. [*2.*]{} $SK_1({{\mathbb Z}}[{{\mathbb Z}}_2 \times {{\mathbb Z}}_2 \times {{\mathbb Z}}_4])\cong {{\mathbb Z}}_2$. [*3.*]{} $SK_1({{\mathbb Z}}[({{\mathbb Z}}_2)^3 \times {{\mathbb Z}}_4])\cong ({{\mathbb Z}}_2)^3 \times {{\mathbb Z}}_4$. This last case is curious in that ${\rm Wh}(\Gamma)=\Gamma$. We can also use the general formula $SK_1({{\mathbb Z}}[{{\mathbb Z}}_{4} \times {{\mathbb Z}}_{2^n}]) \cong ({{\mathbb Z}}_2)^{(n-1)}$ to deduce other relevant groups. For example, $SK_1({{\mathbb Z}}[{{\mathbb Z}}_4 \times {{\mathbb Z}}_8])\cong ({{\mathbb Z}}_2)^2$, $SK_1({{\mathbb Z}}[{{\mathbb Z}}_4 \times {{\mathbb Z}}_{16}])\cong ({{\mathbb Z}}_2)^3$, $SK_1({{\mathbb Z}}[{{\mathbb Z}}_4 \times {{\mathbb Z}}_{32}])\cong ({{\mathbb Z}}_2)^4$, etc. [*2. Products of groups of odd order.*]{} Next we consider the case when the orders of the groups in the products are odd. We will look at groups of the form $({{\mathbb Z}}_p)^k$, ${{\mathbb Z}}_{p^2} \times {{\mathbb Z}}_{p^n}$ and $({{\mathbb Z}}_p)^2 \times {{\mathbb Z}}_{p^n}$, as well as combinations involving three factors, using general results from reference [@Ol]. $(i)$ The torsion subgroup $SK_1({{\mathbb Z}}[{\Gamma}])$ is trivial if ${\Gamma}$ is cyclic or an elementary 2-group, or of type ${{\mathbb Z}}_p \oplus {{\mathbb Z}}_{p^n}$. However, $SK_1({{\mathbb Z}}[{\Gamma}])$ is nontrivial form most abelian groups [@ADS]. If ${\Gamma}=({{\mathbb Z}}_p)^k$, $p$ odd, then $SK_1({{\mathbb Z}}[{\Gamma}])$ is a ${{\mathbb Z}}_p$-vector space of dimension $(p^k-1)/(p-1) - \binom{p+k+1}{p}$. For example, for ${\Gamma}=({{\mathbb Z}}_3)^3$, the torsion subgroup is $SK_1({{\mathbb Z}}[({{\mathbb Z}}_3)^3]) \cong ({{\mathbb Z}}_3)^3$. $(ii)$ For $p$ an odd prime, $SK_1({{\mathbb Z}}[{{\mathbb Z}}_{p^2} \times {{\mathbb Z}}_{p^n}]) \cong ({{\mathbb Z}}/p)^{(p-1)(n-1)}$. $(iii)$ For $p$ an odd prime, $SK_1({{\mathbb Z}}[({{\mathbb Z}}_p)^2\times {{\mathbb Z}}_{p^n} ]) \cong ({{\mathbb Z}}_p)^{np(p-1)/2} $. Let $p$ be an odd prime and $\Gamma$ an elementary abelian $p$-group of rank $k$. Then [@Ste] $SK_1({{\mathbb Z}}[\Gamma])$ is an elementary abelian $p$-group of rank $(p^k-1)/(p-1) - \binom{p+k-1}{p}$. In particular $SK_1({{\mathbb Z}}[\Gamma])\neq 0$ for $k\geq 3$. For example, the following table can be formed (see also [@Ste]) $ \begin{array}{|c|c|} \hline \Gamma & SK_1({{\mathbb Z}}[\Gamma])\\ \hline \hline {{\mathbb Z}}_{p^2}\times {{\mathbb Z}}_{p^2} ~(p=3,5,7) & ({{\mathbb Z}}_p)^{p-1}\\ \hline {{\mathbb Z}}_{p^2}\times {{\mathbb Z}}_{p}\times {{\mathbb Z}}_p ~(p=3,5,7) & ({{\mathbb Z}}_p)^{p(p-1)}\\ \hline {{\mathbb Z}}_{27} \times {{\mathbb Z}}_9 & ({{\mathbb Z}}_3)^4\\ \hline {{\mathbb Z}}_{27}\times {{\mathbb Z}}_3 \times {{\mathbb Z}}_3 & ({{\mathbb Z}}_3)^9\\ \hline {{\mathbb Z}}_9 \times {{\mathbb Z}}_9 \times {{\mathbb Z}}_3 & ({{\mathbb Z}}_3)^{15} \times ({{\mathbb Z}}_9)^2\\ \hline \end{array} $ #### Nonabelian groups. We have already seen examples of nonabelian groups in section \[sec Wh\]. In addition, [*1. Crystallographic groups.*]{} $SK_1({{\mathbb Z}}[\Gamma])=0$ for $\Gamma$ a dihedral, the binary tetrahedral or icosahedral group [@Mag] [@Ste]. [*2. The quaternion group.*]{} The Whitehead group ${\rm Wh}({{\mathbb Z}}[Q_8])$ of the quaternion group $Q_8$ of order 8 is isomorphic to $\pm V$, where $V={{\mathbb Z}}_2\times {{\mathbb Z}}_2$ is Klein’s 4-group [@Ke]. Note that $V$ is the factor group $Q_8/\{\pm\}$, where $\{\pm\}$ is the commutator subgroup of $Q_8$. [*3. Products with abelian groups.*]{} If $\Gamma$ is any (nonabelian) quaternion or semidihedral 2-group, then for all $k \geq 0$, the torsion subgroup is $ SK_1({{\mathbb Z}}[\Gamma \times ({{\mathbb Z}}_2)^k]) \cong ({{\mathbb Z}}_2)^{2^k-k-1}$. [*4. Nonabelian groups with specified abelianization.*]{} For instance, if for order $|\Gamma|=16$ the torsion subgroup is given by $ SK_1({{\mathbb Z}}[\Gamma]) \cong \left\{ \begin{array}{ll} 1 & {\rm if~} \Gamma^{\rm ab} \cong {{\mathbb Z}}_2 \times {{\mathbb Z}}_2 {\rm ~~or~~} {{\mathbb Z}}_2 \times {{\mathbb Z}}_2 \times {{\mathbb Z}}_2\\ {{\mathbb Z}}_2 & {\rm if~} \Gamma^{\rm ab} \cong {{\mathbb Z}}_4 \times {{\mathbb Z}}_2. \end{array} \right. $ #### Finding Whitehead groups via transfer. Looking at inclusions tells us about the corresponding Whitehead groups. We will consider several situations. [*1.*]{} Consider the cyclic group ${{\mathbb Z}}_{2k+1}$ of order $2k+1$ as a subgroup of the cyclic group ${{\mathbb Z}}_{4k+2}$ of order $4k+2$. Then the transfer $i^*: {\rm Wh}({{\mathbb Z}}_{4k+2}) \to {\rm Wh}({{\mathbb Z}}_{2k+1})$, corresponding to $i: {{\mathbb Z}}_{2k+1} \hookrightarrow {{\mathbb Z}}_{4k+2}$, is onto for all $k$ [@Kw]. [*2.*]{} Now consider the inclusion $i: {{\mathbb Z}}_{2k} \hookrightarrow {{\mathbb Z}}_{2k}\oplus {{\mathbb Z}}_2$. Then the transfer $i^*: {\rm Wh}({{\mathbb Z}}_{2k} \oplus Z_2) \to {\rm Wh}({{\mathbb Z}}_{2k})$ is onto if and only if $k=1,2$ or 3 [@Kw2]. Since ${\rm Wh}({{\mathbb Z}}_{2k})=0$ for $k=1,2$ and 3, then this means that ${\rm Wh}({{\mathbb Z}}_{2k} \oplus {{\mathbb Z}}_2)$ is trivial for these values of $k$. [*3.*]{} Now let ${\Gamma}$ be a finite abelian group of odd order. Then $i^*: {\rm Wh}({\Gamma}\oplus {{\mathbb Z}}_2) \to {\rm Wh}({\Gamma})$ is onto [@Kw2]. This then can tell us whether ${\Gamma}\oplus {{\mathbb Z}}_2$ is trivial from whether or not the Whitehead group of ${\Gamma}$ itself is trivial. In general, if ${\Gamma}\to {\Gamma}'$ is a surjection of finite abelian groups induces a surjection $SK_1({{\mathbb Z}}[{\Gamma}]) \to SK_1({{\mathbb Z}}[{\Gamma}'])$ [@ADS]. #### Semidirect products. For finite ${\Gamma}$, the torsion subgroup of the Whitehead group is trivial $SK_1(R[{\Gamma}])=1$ for all rings of integers in number fields if and only if ${\Gamma}$ is a semidirect product of two cyclic groups of relatively prime orders [@ADOS]. In general, we can determine the ranks of the (torsion-free part) of these groups using Bass’ theorem. Given the above rules and results, it is a straightforward exercise to find the Whitehead groups of the fundamental groups appearing in the literature of model building (reviewed partially in the introduction). This includes, for instance, the groups appearing in [@Br]. #### Whitehead torsion. The approach in this paper can also guide us to anticipate conditions on cobordisms when constructing Calabi-Yau threefolds with fundamental groups of certain types. Recall that just because the Whitehead group is nontrivial does not mean that the particular element, the Whitehead torsion, is a nontrivial element. That is, one still has to compute the Whitehead torsion (geometrically), which we do not do here. We consider examples where elements in the torsion subgroup of the Whitehead group can be explicitly characterized (see [@Ol]). $(i)$ For $\Gamma={{\mathbb Z}}_4 \times {{\mathbb Z}}_2 \times {{\mathbb Z}}_2= \langle g \rangle \langle h_1 \rangle \langle h_2 \rangle$, the torsion subgroup is $SK_1({{\mathbb Z}}[\Gamma]) \cong {{\mathbb Z}}/2$, and the nontrivial element is represented by the matrix $ \left[ \begin{array}{cc} 1+ 8(1-g^2)(1+h_1)(1+h_2)(1-g) & -(1-g^2)(1+h_1)(1+h_2)(3+g)\\ &\\ -13(1-g^2)(1+h_1)(1+h_2)(3-g) & 1+ 8(1-g^2)(1+h_1)(1+h_2)(1+g) \end{array} \right] \in {\rm GL}(2, {{\mathbb Z}}[\Gamma])\;. \label{22} $ In this case, one would have to check for a given $h$-cobordism built out of $Y^{11}$ and $M^{10}$ whether the corresponding Whitehead torsion is the zero element or the nontrivial element represented by matrix . $(ii)$ For $\Gamma={{\mathbb Z}}_3 \times Q_8=\langle g \rangle \times \langle a, b \rangle$, where $Q_8$ is a quaternion group of order 8, the torsion subgroup is $SK_1({{\mathbb Z}}[\Gamma])\cong {{\mathbb Z}}/2$, and the nontrivial element is represented by the unit $ 1+ (2-g-g^2)(1-a^2)\left( 3g + a + 4g^2a+4(g^2-g)b + 8ab \right) \in ({{\mathbb Z}}[\Gamma])^*\;. $ Again, one would check the geometry to see which of the two elements one gets. It would be very interesting to calculate the Whitehead torsion explicitly for interesting classes of non-simply connected Calabi-Yau manifolds. As far as we know, no such calculations exist. One approach could be to find an explicit Morse function (which seems not easy). Dynamical aspects {#sec dyn} ================= In this section we consider some dynamical aspects of heterotic M-theory as they arise in connection to the Whitehead group and Whitehead torsion. We consider the effect of diffeomorphisms as well as orientation characters in section \[sec aut\] and then consider dynamical constraints on general compactifications in heterotic M-theory in section \[sec com\]. Automorphisms {#sec aut} ------------- #### Diffeomorphism. We study the effect of diffeomorphisms on our cobordisms, starting with a visible sector $M_0^{10}$. Two eleven-dimensional cobordisms $(Y^{11}; M_0^{10}, f_0, M_1^{10}, f_1)$ and $(Y'^{11}; M_0^{10}, f'_0, M_1'^{10}, f'_1)$ over $M_0^{10}$ are diffeomorphic relative $M_0^{10}$ if there is an orientation preserving diffeomorphism $F: Y^{11} \to Y'^{11}$ such that $F \circ f_0=f'_0$. Indeed in [@FM] the quantum integrand in the M-theory effective action is shown to be invariant under the group of Spin diffeomorphisms of $Y^{11}$ which act freely on the space of metrics. On the other hand, the effective action of the heterotic string is invariant under diffeomorphisms $\varphi: M^{10} \to M^{10}$ which lift to the Spin bundle and to the $E_8$ vector bundles [@W-tool]. The global anomaly is absent for arbitrary choices of the Spin $M^{10}$ and the two $E_8$ vector bundles. In addition to the many examples that we have considered so far, one might be able to generate others using diffeomorphism. In a sense, constructing manifolds with cobordisms for which the Whitehead torsion is nontrivial would be easier than calculating the Whitehead torsion for a given fixed cobordism. The idea is to take a cobordism and and glue it to another after a ‘twist’ via an automorphism, i.e. a diffeomorphism in our case. This may give rise to a nonzero Whitehead torsion. This requires the study of the mapping torus as is done with the global anomalies in the heterotic effective action, e.g. in [@W-tool]. #### Scale and intervals. In the discussion so far we have used unit intervals $[0,1]$ to characterize the cobordism. In the physical set-up of Horava-Witten [@HW1] [@HW2] we have a length scale imposed by the dynamics in the theory. In the above formulation, we can introduce this length scale by simply replacing the unit interval by the interval $[0, L]$ or $[-L, L]$, with $L$ the (dynamical) length in the eleventh direction. #### Manifolds with non-positive sectional curvature. It is interesting to note that ${\rm Wh}(\Gamma)$ is trivial for $\Gamma$ the fundamental group of closed manifolds with all the sectional curvatures $\leq 0$ [@FJ]. Therefore, although not Calabi-Yau (see [@HLW] Theorem 2.3), such spaces are admissible for $s$-cobordism (see [@Rares]). #### The Whitehead torsion relative to left vs. right boundary. We ask whether it makes a difference to take the Whitehead torsion relative to the left boundary vs. taking it relative to the right boundary. There is a duality theorem which relates the Whitehead torsion relative to one boundary to that of the second boundary [@Mil]. For any orientable $h$-cobordism $(Y^{11}, M^{10}, M'^{10})$ we have the relation between $\tau (Y^{11}, M^{10})$ and $\tau (Y^{11}, M^{10})$ as $ \tau (Y^{11}, M'^{10})= \overline{\tau} (Y^{11}, M^{10})\;, $ where $\overline{\tau}$ is the conjugate of $\tau$, defined as follows. If $a=\sum n_i {{\gamma}}_i$ is an element of ${{\mathbb Z}}[{\Gamma}]$, with $n_i\in {{\mathbb Z}}, {{\gamma}}_i\in{\Gamma}$, then the conjugate of $a$ is the element $\sum n_i {{\gamma}}_i^{-1}$. This conjugation operation is an anti-automorphism of the group ring with corresponding automorphism on $GL({{\mathbb Z}}[{\Gamma}])$ given by sending each matrix to its conjugate transpose. Passing to the abelianized group $K_1({{\mathbb Z}}[{\Gamma}])$ gives an automorphism and hence an automorphism also of the quotient ${\rm Wh}({\Gamma})$. We see that ‘reversing’ the direction of the cobordism, that is taking $M'^{10}$ to $M^{10}$ instead of going from $M^{10}$ to $M'^{10}$, will result only in a mild modification in having to deal with the conjugate torsion. For large classes of examples in which we are interested, there is even a simplification. If ${\Gamma}$ is finite abelian then every element $\omega$ of ${\rm Wh}({\Gamma})$ is self-conjugate, $\omega= \overline{\omega}$. This in particular holds for the distinguished element, the Whitehead torsion. Therefore, for finite abelian fundamental groups working with the Whitehead torsion relative to $M^{10}$ is equivalent to working with the Whitehead torsion relative to $M'^{10}$. #### Remark on the $E_8$ gauge bundles. General boundary conditions for M-theory on a manifold with boundary are considered in [@DFM] [@DMW-boundary]. The left and right boundaries in heterotic M-theory each carries an $E_8$ bundle which, in the process of model building is desired to be broken down to a realistic group. Each of the two bundles is characterized with a degree four characteristic class, $a_L$ for left and $a_R$ for right. As explained in [@DFM], when $a_L=a_R$ then the eleven-dimensional spacetime provides a homotopy of the left and right connections so that the $E_8$ bundles on the boundaries necessarily have $a_L=a_R$, which is the case in the non-supersymmetric model in [@FH]. However, in (the supersymmetric) Horava-Witten theory, $a_L + a_R=\frac{1}{2}p_1(Y^{11})$. In order to overcome this difficulty, the authors of [@DFM] give a parity-invariant formulation of the C-field in M-theory by passing from $Y^{11}$ to $Y^{11}_d$, the orientation double cover of $Y^{11}$, and defining the C-field to be a parity invariant $E_8$ cocycle on $Y^{11}_d$. This is done via a nontrivial deck transformation $\sigma$ on $Y^{11}_d$, so that a parity-invariant $E_8$ cocycle is one for which the differential character corresponding to the C-field satisfies $\sigma^* ([\check{C}])=[\check{C}]^{\cal{P}}$, where the action of the parity ${\cal P}$ is $[\check{C}]^{\cal P}= [\check{C}]^*$. While this solves the parity problem it uses boundary conditions which lead to a Bianchi identity for the C-field which is different from the one in [@HW2]. We should keep these subtleties in mind when dealing with bundles, which are always there (but we do not directly deal with them in this paper). Nevertheless, next we provide an explanation of this in our current context. #### Orientation characters and twisted group algebras of the fundamental group. The orientation character $\omega(M_0^{10}): \pi_1(M_0^{10}) \to {{\mathbb Z}}_2=\{\pm 1\}$ sends a loop ${{\gamma}}: S^{1} \to M_0^{10}$ to $\omega({{\gamma}})=+1$ (respectively, -1) if ${{\gamma}}$ is orientation-preserving (respectively, orientation-reversing). Thus, in the oriented case $\omega ({{\gamma}})=+1$ for all ${{\gamma}}\in {\Gamma}$, that is $\omega$ is trivial if and only if $M_0^{10}$ is orientable. This has the following effect on the integral group ring of the fundamental group. The orientation character defines a twisted involution (an anti-automorphism) on the group ring ${{\mathbb Z}}[{\Gamma}]$ given by $a \mapsto \omega(a) a^{-1}$, i.e. $\pm a$ according to whether $a$ is orientation preserving or reversing. The resulting group ring is denoted ${{\mathbb Z}}[{\Gamma}]^\omega$. Let us consider this in more detail. An involution on ${{\mathbb Z}}[{\Gamma}]$ is a function ${{\mathbb Z}}[{\Gamma}] \to {{\mathbb Z}}[{\Gamma}]$, taking an element $a$ to an element $\overline{a}$ satisfying: $\overline{(a + b)}=\overline{a} + \overline{b}$, $\overline{(ab)}=\overline{b}\cdot \overline{a}$,  $\overline{\overline{(a)}}=a$, and $\overline{1}=1\in {{\mathbb Z}}[{\Gamma}]$. This gives rise to the $\omega$-twisted involution on ${{\mathbb Z}}[{\Gamma}]$, defined as the map from ${{\mathbb Z}}[{\Gamma}]$ to ${{\mathbb Z}}[{\Gamma}]$ given by $ a=\sum_{{{\gamma}}\in {\Gamma}} n_{{{\gamma}}}{{\gamma}}\longmapsto \overline{a}=\sum_{{{\gamma}}\in {\Gamma}} \omega({{\gamma}}) n_{{\gamma}}{{\gamma}}^{-1}\;, \quad n_{{\gamma}}\in {{\mathbb Z}}. $ In this case we have to use $\omega$-twisted cohomology and fundamental class in evaluating expressions in the theory. Starting from the cellular ${{\mathbb Z}}[\pi_1(M_0^{10})]$-module chain complex $C(\widetilde{M}_0^{10})$, the $\omega(M_0^{10})$-twisted involution on ${{\mathbb Z}}[\pi_1(M_0^{10})]$ can be used to define the left ${{\mathbb Z}}[\pi_1(M_0^{10})]$-module structure on the dual cochain complex $C(\widetilde{M}_0^{10})={\rm Hom}_{{{\mathbb Z}}[\pi_1(M_0^{10})]}\left( C(\widetilde{M}_0^{10}, {{\mathbb Z}}[\pi_1(M_0^{10})] \right)$. When $M_0^{10}$ is compact, [^3] the fundamental class is given by $[M_0^{10}] \in H_{10}(M_0^{10}; {{\mathbb Z}}^{\omega(M)})$ such that the cap product defines ${{\mathbb Z}}[\pi_1]$-module isomorphisms $ [M_0^{10}] \cap - ~: H^*_{\omega(M)} (\widetilde{M}_0^{10}) \buildrel{\cong}\over{\longrightarrow} H_{10-*}(\widetilde{M}^{10}_0) $ with $\widetilde{M}^{10}_0$ the universal cover of $M_0^{10}$. Quantities, e.g. ones appearing the effective action and the corresponding partition function, should be formulated using this fundamental class. There is a duality formula for the Whitehead torsion which takes into account the orientation character. Let $Y^{11}$ be an eleven-dimensional $h$-cobordism and let $\omega: \Gamma \to \{\pm 1\}$ be the orientation character. This gives rise to an anti-involution on the integral group ring ${{\mathbb Z}}{\Gamma}$ by sending a group element $g$ to $\omega g^{-1}$, as above, and hence leads to an involution $*$ on the Whitehead group ${\rm Wh}({\Gamma})$. Then Milnor’s duality formula is cast as $ \tau (Y^{11}, M'^{10})= \tau (Y^{11}, M^{10})^*\;. $ #### Effect on F-theory. Recently there has been a lot of research activity in model building using F-theory (see [@De] and references therein). F-theory can be considered as a limit of M-theory on a 2-torus when the volume of the two-torus becomes very small. This means that constraints on the possible fundamental groups of $Y^{11}$, assumed to have a 2-torus factor, will have an effect on the possible fundamental groups on the space on which F-theory is considered. Nontrivial fundamental groups in this context are considered in [@Br2]. Therefore, we expect that our discussion in the heterotic/M-theory setting will have, via duality, consequences for fundamental groups in F-theory. This is strengthened by the fact that in a class of models which admit perturbative heterotic duals, the F-theory and heterotic computations match [@DW]. It would be interesting to perform explicit checks of this in relevant examples. Compactification {#sec com} ---------------- We have considered in general the relation between M-theory on a general eleven-manifold and heterotic string theory on a general ten-manifold $M_0^{10}$. There are two aspects to this. First, for consistency the theory should make sense on any admissible manifold and so studying this might give insight into understanding the theory further. Second, there are certain favorable types of spaces for model building. We have in mind that $M_0^{10}$ is a product (or a bundle) of a Calabi-Yau threefold $X^6$ with a four-dimensional spacetime. In general, the latter can be taken to be a general four-manifold that solves the equations of motion and does/does not break supersymmetry according to the goal one has in mind. It can be taken to be flat Minkowski or something close. We study such situations in this section and consider whether the choice of four-dimensional spacetime changes the discussion we have had so far. We take M-theory on an eleven-manifold $Y^{11}=Z^7 \times N^4$, where $N^4$ is spacetime and $Z^7$ is a seven-dimensional cobordism of the Calabi-Yau threefold $X^6$. This always exists because the Stiefel-Whitney numbers of a Calabi-Yau threefold are zero: $w_1=0$ because of orientation, $w_2=0$ because of Spin, and $w_3=0$ because both $w_1$ and $w_2$ are zero; then the Stiefel-Whitney numbers $w_1w_5[X^6]$, $w_2w_4[X^6]$, and $w_3w_3[X^6]$ are all zero. The heterotic ten-manifold is of the form $M_0^{10}=X^6 \times N^4$. #### The $h$-cobordism of a product. Let $(Z^7; X_0^6, X_1^6)$ be a seven-dimensional $h$-cobordism for the Calabi-Yau threefold $X_0^6$, and let $N^4$ be a closed four-manifold. Then we can form an eleven-dimensional $h$-cobordism $(Z^7 \times N^4; X_0^6 \times N^4, X_1^6\times N^4)$. From the cut and paste properties of the Whitehead torsion (see [@Mil] [@We] [@KL]), we get that the torsion are related as follows $ \tau (Z^7 \times N^4, X_0^6 \times N^4)= \tau (Z^7, X_0^6)~ \chi (N^4)\;, \label{eq zcs} $ where $\chi(N^4)$ is the Euler characteristic of $N^4$. Thus the value of this invariant will determine whether there we can relate the discussion of torsion in eleven/ten dimensions to that in seven/six dimensions. The former is the global picture we have built so far, and the latter correspond to the actual situation studied in model building, that is the fundamental groups appearing as examples are those of $X^6$ and not (necessarily) of $M^{10}$. If spacetime were compact and odd-dimensional then the Euler characteristic would vanish identically. In that case, the torsion would vanish. For example, if we take spacetime to be the circle $S^1$ then $Z^7 \times S^1 \approx X_0^6 \times S^ \times [0,1] \approx X_1^6 \times S^1 \times [0,1]$, i.e. the torsion vanishes. In particular, this gives $X_0^6 \times S^1 \approx X_1^6 \times S^1$. #### Product with a torus and Wall’s finiteness obstruction. The circle $S^1$ has fundamental group $\pi_1(S^1)\cong {{\mathbb Z}}$. If we consider the product $S^1 \times Y$, then what is the corresponding Whitehead group in terms of that of the factors? There is in fact a direct sum decomposition [@BHS] $ {\rm Wh}({{\mathbb Z}}\times \Gamma) \cong {\rm Wh}(\Gamma) \oplus \widetilde{K}_0({{\mathbb Z}}[\Gamma]) \oplus N$, for some Nil-group $N$. For the 2-torus with fundamental group ${{\mathbb Z}}^2$, the process can be repeated. It might seem that for this product we can have nonzero Whitehead group for the product manifold even though that group for the factors might not be zero. However, elements in $\widetilde{K}_0({{\mathbb Z}}[\pi_1(X)])$, called Wall’s finiteness obstruction, detects whether or not $X^6$ has the homotopy type of a CW-complex. If we are within the category of such spaces then this element within the class group vanishes. #### Spacetime with flat structure. A manifold admits a flat structure if the tangent bundle is isomorphic to a flat vector bundle, i.e. admits a flat connection. Even for such manifolds, one can have nonzero Euler-characterstic. For example, if we take take the connected sum $N^4=(\Sigma_3 \times \Sigma_3)~ \#_{i=1}^6(S^1 \times S^3)$, where $\Sigma_3$ is a surface of genus 3. The product $\Sigma_3 \times \Sigma_3$ is almost parallelizable and the product of spheres $S^1 \times S^3$ is parallelizable. Then the Euler characteristic is $\chi(N^4)=4$ (see [@Sm]). In this example, the fundamental group is the free product $\pi_1(N^4)={\Gamma}_1 \ast {\Gamma}_2$, where ${\Gamma}_1$ is the direct product of two copies of a non-abelian surface group and ${\Gamma}_2$ is of rank 6. In fact, $S^1 \times S^3$ can be replaced by any parallelizable four-manifold. #### Compact vs. noncompact spacetime. So far we have taken $N^4$ to be compact. For compact manifolds, the existence of a smooth Lorentzian metric is equivalent to the manifold having a vanishing Euler characteristic (see [@St]). However, the situation gets modified in the presence of singularities (see [@Ma]). What if it is not compact? Noncompact spacetimes are more desirable for the purpose of equipping spacetime with a Lorentzian structure; all noncompact manifolds admit a Lorentzian metric. On the other hand, every noncompact manifold admits vector fields with any specified set of isolated zeros. This suggests that noncompact manifolds with nonzero (appropriate notion of) [^4] Euler characteristic are abundant. Note that for noncompact Riemann surfaces, the Cohn-Vossen theorem gives the inequality $\int_\Sigma K dA \leq 2\pi \chi(\Sigma)$ (see e.g. [@Ku]). In general one works with $L^2$-Euler characteristics. For example, the Euler characteristic of an Asymptotically Locally Euclidean (ALE) space corresponding to the Lie algebra of type $A_n$ is $n+1$. It is important to note that it should be checked whether equation extends to the noncompact case. Furthermore, strictly speaking, in the noncompact case we have to use the noncompact version of the $s$-cobordism theorem, for which the Whitehead torsion lives a new group, which fits into an exact sequence involving the Whitehead group and algebraic $K_0$, as well as information about the ends [@Si2]. Some aspects of behavior of ends in M-theory are discussed in [@DMW-corner]. In the following few paragraphs we describe a way for studying the Whitehead torsion via other invariants, namely the Reidemeister torsion [@Mil] and the Ray-Singer torsion [@RS]. This then provides a setting for making some direct connections to phenomenology. #### Relation of the Whitehead torsion to Reidemeister torsion. The Whitehead torsion $\tau$ is closely related to Reidemeister torsion (or R-torsion) $\Delta$; the former generalizes the latter but is a more delicate invariant. Algebraically, the Whitehead torsion is more general than R-torsion in that it is also defined for noncommutative rings (such as the group ring of the fundamental group ${{\mathbb Z}}[\pi_1(X)]$) for which the determinant, needed for the R-torsion, is not defined. The R-torsion is a topological invariant which distinguishes spaces which are homotopy equivalent but not homeomorphic, and is defined for spaces whose fundamental group $\pi$ is finite and for which the homology with coefficients in a certain $\pi$-representation vanishes. The R-torsion is defined in more general situations than Whitehead torsion, since any homotopy equivalence is a homology equivalence. Furthermore, R-torsion has two advantages over the Whitehead torsion: \(i) It is more likely to be defined. \(ii) Its value is an honest real number, instead of being an element of a somewhat esoteric group. On the other hand, when defined, the Whitehead torsion is a sharper invariant. When they are both defined, the R-torsion is a function of the Whitehead torsion. That is, for each unitary (orthogonal) representation $\rho$ of the fundamental group $\pi$, the R-torsion is the real part of the determinant of the complex (real) matrix induced by $\rho$ from any matrix representation of the Whitehead torsion. One can find a useful criterion for when the Whitehead torsion is zero by studying the R-torsion. For concreteness, let $h : \pi_1(M^{10}) \to O(n)$ be an orthogonal representation of the fundamental group $\pi=\pi_1(M^{10})$. Then $h$ extends to a unique homomorphism from the group ring ${{\mathbb Z}}[\pi]$ to the ring ${\cal M}_n({{\mathbb R}})$ of all real $n \times n$ matrices and determines a homomorphism $h_*: {\rm Wh}(\pi) \to \overline{K}_1({{\mathbb R}}) \cong {{\mathbb R}}^+$. Suppose that the Whitehead torsion $\tau(Y^{11}; M^{10}) \in {\rm Wh}(\pi)$ is defined and suppose that $\pi$ is a finite group. Then it follows from the identity relating the two torsions [@Mil] $ \Delta_h(Y^{11}; M^{10})= h_*\tau (Y^{11}; M^{10}) $ that $\tau(Y^{11}; M^{10})$ is an element of finite order in ${\rm Wh}(\pi)$ if and only if the R-torsion is $\Delta_h(Y^{11}; M^{10})=1$ for all possible orthogonal representations $h$ of $\pi$. If $\pi$ is finite abelian, then $\tau (Y^{11}; M^{10})=0$ if and only if $\Delta_h(Y^{11}; M^{10})=1$ for all possible such representations $h$. Since the R-torsion is easier to calculate, this gives a concrete way of checking whether the Whitehead torsion vanishes without having to go through the difficult task of calculating it explicitly. #### Examples of when R-torsion is defined and the Whitehead torsion is not. There are examples in which the Whitehead torsion cannot be defined but the R-torsion can (see [@Mil]). For instance, the Whitehead torsion $\tau (S^1)$ of the circle $S^1$ cannot be defined since the module $H_0(\hat{S}^1)$ for the universal cover $\hat{S}^1$ is not zero, and is not a free ${{\mathbb Z}}[\pi]$-module. On the other hand, the R-torsion is defined; if the homomorphism $h$ from the fundamental group $\pi_1(S^1)$ to the units $\mathbb{F}^\times$ in a field $\mathbb{F}$ maps a generator into the field element $x\neq 1$, then the associated R-torsion $\Delta_h(S^1) \in \mathbb{F}^\times /\pm h(\pi_1)$ is well-defined and equal to $1-h$, up to multiplication by $h^{m}$ for some $m\in {{\mathbb Z}}^\times$. Another example is a knot complement $X$ in the 3-sphere with $h: \pi_1(X) \to \mathbb{F}^\times$ mapping each loop with linking number $+1$ into the field element $x \neq 1$. Then the R-torsion is well-defined, and is equal to $(1-h)/A(h)$, where $A(h)$ is the Alexander polynomial of the knot. #### Effect on phenomenology. The Ray-Singer torsion, which is an analytic analog of R-torsion and which coincides with it for Riemannian manifolds, has direct physical applications. The Ray-Singer torsion can be defined using determinants of Laplacians. In this form it has natural connection to one-loop amplitudes. For example, this torsion governs the threshold corrections for the heterotic string [@BCOF]. In M-theory compactifications on manifolds with $G_2$ holonomy, the GUT scale $M_{\rm GUT}$ is essentially given by the Ray-Singer torsion $\Delta_{RS}(\Sigma)$ via $M_{\rm GUT}^3= \Delta_{RS}(\Sigma)/V_\Sigma$, where $V_\Sigma$ is the volume of the corresponding 3-cycle $\Sigma$ [@FW]. For example, when $\Sigma=S^3/{{\mathbb Z}}_q$ is a lens space, on which there is a Wilson line of eigenvalues $\left( e^{2\pi i(2m/q)}, e^{2\pi i(2m/q)}, e^{2 \pi i(m/q)}, e^{-2\pi i(3m/q)}, e^{-2\pi i(3m/q)} \right)$ with $m$ and $q$ coprime integers, then the Ray-Singer torsion for the lens space is $\Delta_{RS}(\Sigma)=4q\sin^2(5\pi m/q)$. Now, the more delicate Whitehead torsion can be partially studied by considering the R-torsion (or Ray-Singer torsion) as above. It should be an obstruction to supersymmetry in heterotic M-theory. The breaking scale would be the intermediate 5-dimensional scale, and only gravitationally mediate to the visible sector. It would be interesting to see how this works explicitly. #### Higher-dimensional compactifications. If we take our eleven-manifold $Y^{11}$ to be a product of two manifolds, where the internal manifold is of dimension lower than 6 then we can no longer apply the $s$-cobordism arguments we have been using. In particular, the $s$-cobordism theorem fails in dimensions five and it is an open problem in dimension four (see [@CS] [@Ch]). For example, there exists an $h$-cobordism $(W^5, T^4, T^4)$, where $T^4$ is the four-dimensional torus, for which there is no diffeomorphism from $W^5$ to $T^4 \times [0,1]$. Since ${\rm Wh}(\pi_1(T^4))=0$, the $s$-cobordism indeed fails in five dimensions. For topological spaces, the theorem fails in both four and five dimensions [@Si]; one might say that we could apply the $s$-cobordism in this case to the spacetime part rather than the internal part, now that spacetime has grown to admissible dimensions. This certainly can be done and will give consistency conditions depending on fundamental groups of spacetime (the arguments we have outlined will go through with the obvious changes). However, we would then not be studying fundamental groups for purposes of particle physics but rather for purposes of cosmology. The author would like to thank Jonathan Rosenberg for useful discussions on the Whitehead torsion and Kenji Fukaya and the referee for useful comments. He also acknowledges the hospitality of the Department of Physics and the Department of Mathematics at the National University of Singapore where part of this work was done. [99]{} R. C. Alperin, R. K. Dennis, R. Oliver, and M. R. Stein, [*$SK_1$ of finite abelian groups II*]{}, Invent. Math. [**87**]{} (1987), no. 2, 253–302. R. C. Alperin, R. K. Dennis and M. R. Stein, [*The nontriviality of $SK_1({{\mathbb Z}}\pi)$*]{}, in Orders, Group Rings and Related Topics, Lecture Notes in Math., vol. 353, Springer-Verlag, New York, 1973, 1–7. A. Bak, V. Bouchard, and R. Donagi, [*Exploring a new peak in the heterotic landscape*]{}, J. High Energy Phys. [**06**]{} (2010) 108, 1–31, \[[arXiv:0811.1242]{}\] \[[hep-th]{}\]. H. Bass, A. Heller, and R. Swan, [*The Whitehead group of a polynomial extension*]{}, Publ. de I’lnst. des Hautes Etudes Sci. [**22**]{} (1964) 61–79. V. Batyrev and M. Kreuzer, [*Integral cohomology and mirror symmetry for Calabi-Yau 3-folds*]{}, Mirror symmetry V, 255–270, AMS/IP Stud. Adv. Math., 38, Amer. Math. Soc., Providence, RI, 2006, \[[math.AG/0505432]{}\]. M. Bershadsky, S. Cecotti, H. Ooguri, and C. Vafa, [*Kodaira-Spencer theory of gravity and exact results for quantum string amplitudes*]{}, Commun. Math. Phys. [**165**]{} (1994) 311–428, \[[arXiv:hep-th/9309140]{}\]. L. Borisov and Z. Hua, [*On Calabi-Yau threefolds with large nonabelian fundamental groups*]{}, Proc. Amer. Math. Soc. [**136**]{} (2008), no. 5, 1549–1551, \[[arXiv:math/0609728]{}\] \[[math.AG]{}\]. V. Bouchard and R. Donagi, [*An SU(5) heterotic standard model*]{}, Phys. Lett. [**B633**]{} (2006) 783–791, \[[arXiv:hep-th/0512149]{}\]. V. Bouchard and R. Donagi, [*On a class of non-simply connected Calabi-Yau threefolds*]{}, Comm. Numb. Theor. Phys. [**2**]{} (2008) 1–61, \[[arXiv:0704.3096]{}\] \[[math.AG]{}\]. V. Braun, [*On free quotients of complete intersection Calabi-Yau manifolds*]{}, \[[arXiv:1003.3235]{}\] \[[hep-th]{}\]. V. Braun, [*Discrete Wilson lines in F-theory*]{} , \[[arXiv:1010.2520]{}\] \[[hep-th]{}\]. V. Braun, Y.-H. He, B. A.Ovrut, and T. Pantev, [*The exact MSSM spectrum from string theory*]{}, J. High Energy Phys. [**0605**]{} (2006) 043, \[[arXiv:hep-th/0512177]{}\]. V. Braun, M. Kreuzer, B. A. Ovrut, and E. Scheidegger, [*Worldsheet instantons and torsion curves, part A: Direct computation*]{}, J. High Energy Phys. [**0710**]{} (2007) 022, \[[arXiv:hep-th/0703182]{}\]. V. Braun, B. A.Ovrut, T. Pantev, and R. Reinbacher, [*Elliptic Calabi-Yau threefolds with ${{\mathbb Z}}_3 \times {{\mathbb Z}}_3$ Wilson lines*]{}, J. High Energy Phys. [**0412**]{} (2004) 062, \[[arXiv:hep-th/0410055]{}\]. J. D. Breit, B. A. Ovrut, and G. C. Segre, [*$E_6$ symmetry breaking in the superstring theory*]{}, Phys. Lett. [**B158**]{} (1985) 33–39. P. Candelas and R. Davies, [*New Calabi-Yau manifolds with small Hodge numbers*]{}, \[[arXiv:0809.4681]{}\] \[[hep-th]{}\]. S. Cappell and J. Shaneson, [*On 4-dimensional $s$-cobordisms*]{}, J. Differential Geom. [**22**]{} (1985), no. 1, 97–115. W. Chen, [*Smooth $s$-cobordisms of elliptic 3-manifolds*]{}, J. Differential Geom. [**73**]{} (2006), no. 3, 413–490. M. M. Cohen, A Course in Simple-Homotopy Theory, GTM 10, Springer- Verlag, New York-Berlin, 1973. M. Davis and J.-C. Hausmann, [*Aspherical manifolds without smooth or PL structure*]{}, Lect. Notes in Math., Vol. 1370, Springer-Verlag, New York, 1989, pp. 135-142. F. Denef, [*Les Houches lectures on constructing string vacua*]{}, \[[arXiv:0803.1194]{}\] \[[hep-th]{}\]. E. Diaconescu, D. S. Freed and G. Moore, [*The M-theory 3-form and $E_8$ gauge theory*]{}, Elliptic cohomology, 44–88, Cambridge Univ. Press, Cambridge, 2007, \[[arXiv:hep-th/0312069]{}\]. R. Donagi, P. Gao, and M. B. Schulz, [*Abelian fibrations, string junctions, and flux/geometry duality*]{}, \[[arXiv:0810.5195]{}\] \[[hep-th]{}\]. R. Donagi, B. A.Ovrut, T. Pantev, and R. Reinbacher, [*SU(4) instantons on Calabi-Yau threefolds with ${{\mathbb Z}}_2 \times {{\mathbb Z}}_2$ fundamental group*]{}, J. High Energy Phys. [**0401**]{} (2004) 022, \[[arXiv:hep-th/0307273]{}\]. R. Donagi, B. A.Ovrut, T. Pantev, and D. Waldram, [*Standard models from heterotic M-theory*]{}, Adv. Theor. Math. Phys. [**5**]{} (2002) 93–137, \[[arXiv:hep-th/9912208]{}\]. R. Donagi and M. Wijnholt, [*Model building with F-theory*]{}, \[[arXiv:0802.2969]{}\] \[[hep-th]{}\]. M. Fabinger and P. Horava, [*Casimir effect between world-branes in heterotic M-theory*]{}, Nucl. Phys. [**B580**]{} (2000) 243–263, \[[arXiv:hep-th/0002073]{}\]. F.T. Farrell and L.E. Jones, [*Topological rigidity for compact nonpositively curved manifolds*]{}, Proc. Sympos. Pure Math. [**54**]{}, Part 3, American Mathematical Society, Providence, R.I. (1993), 229–274. D. S. Freed and G. W. Moore, [*Setting the quantum integrand of M-theory*]{}, Commun. Math. Phys. [**263**]{} (2006) 89–132, \[[arXiv:hep-th/0409135]{}\]. T. Friedmann and E. Witten, [*Unification scale, proton decay, and manifolds of $G_2$ holonomy*]{}, Adv. Theor. Math. Phys. [**7**]{} (2003) 577–617, \[[arXiv:hep-th/0211269]{}\]. M. B. Green, J. H. Schwarz, and E. Witten, [Superstring Theory]{}, vol 2, Cambridge Univ. Press, Cambridge, 1988. M. Gross and S. Popescu, [*Calabi-Yau threefolds and moduli of abelian surfaces I*]{}, Compositio Math. [**127**]{} (2001), 169–228. Y.-H. He, [*An algorithmic approach to heterotic string phenomenology*]{}, Mod. Phys. Lett. A [**25**]{} (2010) 79–90, \[[arXiv:1001.2419]{}\] \[[hep-th]{}\]. G. Heier, S. S. Y. Lu, and B. Wong, [*On the canonical line bundle and negative holomorphic sectional curvature*]{}, Math. Res. Lett. [**17**]{} (2010) 1101–1110. P. Horava and E. Witten, [*Heterotic and type I string dynamics from eleven dimensions*]{}, Nucl. Phys. [**B460**]{} (1996) 506–524, \[[arXiv:hep-th/9510209]{}\]. P. Horava and E. Witten, [*Eleven-dimensional supergravity on a manifold with boundary*]{}, Nucl. Phys. [**B475**]{} (1996) 94–114, \[[arXiv:hep-th/9603142]{}\]. M. E. Keating, [*On the K-theory of the quaternion group*]{}, Mathematika [**20**]{} (1973) 59–62. M. Kervaire, [*Le théorème de Barden-Mazur-Stallings*]{}, Comment. Math. Helv. [**40**]{} (1965) 31–42. M. Kreck and W. Lück, The Novikov Conjecture: Geometry and Algebra, Birkhäuser, Basel, 2005. W. Kühnel, Differential Geometry: Curves– Surfaces– Manifolds, Amer. Math. Soc., Providence, RI, 2006. K. W. Kwun, [*Transfer homomorphisms of Whitehead groups of some cyclic groups*]{}, Amer. J. Math. [**93**]{} (1971) 310–316. K. W. Kwun, [*Transfer homomorphisms of Whitehead groups of some cyclic groups II*]{}, Lecture Notes in Math. [**298**]{}, 437–440, Springer, Berlin, 1972. B. Magurn, [*$SK_1$ of dihedral groups*]{}, J. Algebra [**51**]{} (1978), no. 2, 399–415. L. Markus, [*Line element fields and Lorentz structures on differentiable manifolds*]{}, Ann. of Math. (2) [**62**]{} (1955), 411–417. B. Mazur, [*Relative neighborhoods and the theorems of Smale*]{}, Ann. of Math. (2) [**77**]{} (1963) 232–249. J. Milnor, Lectures on the $h$-cobordism theorem, Princeton Univ. Press, Princeton, NJ, 1965. J. Milnor, [*Whitehead torsion*]{}, Bull. Amer. Math. Soc. [**72**]{} (1966) 358–426. R. Oliver, Whitehead groups of finite groups, Cambridge Univ. Press, Cambridge, 1988. R. Rares, [*On the topology and differential geometry of Kähler threefolds*]{}, PhD Dissertation, Stony Brook University, 2005. D. B. Ray and I. M. Singer, [*Analytic torsion for complex manifolds*]{}, Ann. Math. (2) [**98**]{} (1973) 154–177. J. Rosenberg, Algebraic K-theory and its Applications, Springer-Verlag, New York, 1994. H. Sati, [*Geometry of Spin and Spin${}^c$ structures in the M-theory partition function*]{} \[[arXiv:1005.1700]{}\] \[[hep-th]{}\]. H. Sati, [*Duality and cohomology in M-theory with boundary*]{}, \[[arXiv:1012.4495]{}\] \[[hep-th]{}\]. H. Sati, [*Corners in M-theory*]{}, J. Phys. [**A44**]{} (2011) 255402, \[[arXiv:1101.2793]{}\] \[[hep-th]{}\]. C. Schoen, [*On fiber products of rational elliptic surfaces with section*]{}, Math. Zeitschrift [**197**]{}(2) (1988) 177–199. M. B. Schulz, [*Calabi-Yau duals of torus orientifolds*]{}, J. High Energy Phys. [**0605**]{} (2006) 023, \[[arXiv:hep-th/0412270]{}\]. V. V. Sharko, Functions on Manifolds: Algebraic and Geometric Aspects, American Mathematical Society, Providence, RI, 1993. L. C. Siebenmann, [*Disruption of low-dimensional handlebody theory by Rohlin’s theorem*]{}, Topology of Manifolds, 57–76, Markham, Chicago, Ill, 1970. L. C. Siebenmann, [*Infinite simple homotopy types*]{}, Indag. Math. [**32**]{} (1970) 479–495. J. Smillie, [*Flat manifolds with non-zero Euler characteristic*]{}, Comment. Math. Helv. [**52**]{} (1977), no. 3, 453–455. J. Stallings, [*On infinite processes leading to differentiability in the complement of a point*]{}, in Differential and Combinatorial Topology, 245–254, Princeton Univ. Press, Princeton, NJ, 1965. N. Steenrod, The Topology of Fiber Bundles, Princeton Univ. Press, Princeton, NJ, 1951. M. R, Stein, [*Whitehead groups of finite groups*]{}, Bull. Amer. Math. Soc. [**84**]{} (1978) 201–212. C. T. C. Wall, [*Norms of units in group rings*]{}, Proc. London Math. Soc. [**29**]{} (1974), 593–632. S. Weinberger, The Topological Classification of Stratified Spaces, Univ. of Chicago Press, Chicago, Ill 1994. E. Witten, [*Symmetry breaking patterns in superstring models*]{}, Nucl. Phys. [**B258**]{} (1985) 75–100. E. Witten, [*Topological tools in 10-dimensional physics*]{}, Int. J. Mod. Phys. A [**1**]{} (1986) 39–64. [^1]: e-mail: [[email protected]]{}\ Current address: Department of Mathematics, University of Pittsburgh, 139 University Place, Pittsburgh, PA 15260. [^2]: A free quotient of the manifold corresponding to ${{\mathbb Z}}_3 \times {{\mathbb Z}}_3$ by the quaternion group is given in [@CD]. [^3]: $M_0^{10}$ does not necessarily have to be a manifold, but just a Poincaré duality complex. [^4]: Note that there are various definitions and versions of the Euler characteristic in the noncompact setting.
{ "pile_set_name": "ArXiv" }
--- abstract: | The recent temperature measurements of the two older isolated neutron stars PSR 1929+10 and PSR 0950+08 (ages of $3\times 10^6$ and $2\times 10^7$ yr, respectively) indicate that these objects are heated. A promising candidate heat source is friction between the neutron star crust and the superfluid it is thought to contain. We study the effects of superfluid friction on the long-term thermal and rotational evolution of a neutron star. Differential rotation velocities between the superfluid and the crust (averaged over the inner crust moment of inertia) of $\bar\omega\sim 0.6$ rad s$^{-1}$ for PSR 1929+10 and $\sim 0.02$ rad s$^{-1}$ for PSR 0950+08 would account for their observed temperatures. These differential velocities could be sustained by pinning of superfluid vortices to the inner crust lattice with strengths of $\sim$ 1 MeV per nucleus. Pinned vortices can creep outward through thermal fluctuations or quantum tunneling. For thermally-activated creep, the coupling between the superfluid and crust is highly sensitive to temperature. If pinning maintains large differential rotation ($\sim 10$ rad s$^{-1}$), a feedback instability could occur in stars younger than $\sim 10^5$ yr causing oscillations of the temperature and spin-down rate over a period of $\sim 0.3 t_{\rm age}$. For stars older than $\sim 10^6$ yr, however, vortex creep occurs through quantum tunneling, and the creep velocity is too insensitive to temperature for a thermal-rotational instability to occur. These older stars could be heated through a steady process of superfluid friction. author: - 'Michelle B. Larson' - Bennett Link title: 'Superfluid Friction and Late-time Thermal Evolution of Neutron Stars' --- Introduction ============ A cooling neutron star cools initially through neutrino emission before making a transition to photon cooling at an age of $\sim 10^5$ yr (see Tsuruta 1998 for a comprehensive review; see Fig. 1). Internal heating processes, if they occur, could affect when and how abruptly the star makes the transition to photon cooling. Later, when the heating power begins to exceed the luminosity from residual heat, the heat source would control the star’s thermal evolution. Internal heating processes that could occur include superfluid frictional heating (see, , [@greenstein75]; [@HGG]; Alpar  1987; Shibazaki & Lamb 1989; [@VEM]; Umeda  1993; Van Riper, Link & Epstein 1995), structural readjustment through plastic flow or “starquakes” ([@starquakes]; [@ruderman76]; Cheng   1992), chemical disequilibrium driven by the star’s spin-down (Reisenegger 1995), and magnetic field decay (Thompson & Duncan 1996; Heyl & Kulkarni 1998). Temperature measurements of neutron stars older than $\sim 10^5$ yr provide strong tests of cooling models in the photon cooling era, and offer the possibility of constraining the heating processes that might occur. In particular, the recent temperature measurements of PSRs 1929+10 and 0950+08 by Pavlov, Stringfellow & Córdova (1996) pose a serious challenge to standard cooling models, and seem in fact to demand internal heating at rates of $\sim 10^{-4} L_\odot$ and $\sim 10^{-5} L_\odot$, respectively (see Fig. 1.) In this paper we explore the possibility that heat generated by friction between the neutron star crust and the interior neutron fluid is the dominant heating processes taking place in old neutron stars with conventional magnetic fields ($\sim 10^{12}$ G). Most of the mass of a neutron star is expected to be in the form of a neutron superfluid that condenses shortly after the star’s birth (Migdal 1959). Large velocity differences between the superfluid and the crust could develop in the star’s inner crust, where the vortices that thread the rotating superfluid pin to nuclei ([@AI]; [@ruderman76]; Alpar 1977; Epstein & Baym 1988; [@pizzo]). Differential rotation between the stellar crust and the superfluid would lead to frictional heat generation, while variations in the frictional coupling would affect the star’s spin behavior. Studies of superfluid friction in the inner crust indicate that the coupling can be highly temperature-dependent, scaling with temperature as ${\rm e}^{-E/kT}$, where $E$ is an energy $\gg kT$ (see, , Alpar   1984; Link, Epstein & Baym 1993; Chau & Cheng 1993). The star’s thermal and rotational evolution are thus coupled and must be considered together. Usually the core superfluid is regarded as corotating with the stellar crust, though Sedrakian & Sedrakian (1995; see also Sedrakian & Cordes 1998) have suggested that interactions between superfluid vorticity in the core with the London current near the crust-core interface could maintain differential rotation between the crust and core. In this case, friction between the core superfluid and the normal matter could also be a heat source. The goal of this paper is to describe the role played by superfluid friction in late-time neutron star thermal evolution. Though heat generation in the core is a possibility, we will focus primarily on friction in the crust, as the coupling there has been studied in detail. We consider first (§3) the case in which the superfluid is in rotational equilibrium, spinning down at the same rate as the crust. This case was originally considered by Alpar  (1987) and applied to the upper limit on the luminosity of PSR 1929+10 available at the time to obtain a constraint on the excess angular momentum residing in the superfluid. In later work, Shibazaki & Lamb (1989), Umeda  (1993), and Van Riper, Link & Epstein (1995) included the effects of superfluid friction in simulations of neutron star thermal evolution to obtain further constraints on the heating through this process. This study is largely motivated by the recent measurements of the temperatures of the old pulsars PSRs 1929+10 and 0950+08 (Pavlov, Stringfellow & Córdova 1996), which, as we show, provide the most stringent constraints to date on the rotation of the superfluid interior. A crucial issue in neutron star thermal evolution is the possible development of thermal instabilities. Shibazaki and Mochizuki (1994; hereafter SM) have shown that under certain circumstances a feedback instability between the star’s thermal and rotational states can occur if the frictional coupling of the superfluid to the crust is sufficiently sensitive to temperature; as heat is generated, the frictional coupling is increased, creating more heat. The star executes a limit cycle in which its temperature oscillates about the temperature at which it is marginally stable, accompanied by oscillations in the rotation rate. This case we study in detail in § 4. We conclude that the coupling of the superfluid to the crust is nearly independent of temperature in a star older than $\sim 10^6$ yr, effectively decoupling the star’s rotational and thermal evolution. In §5 we use x-ray and optical data from cooling neutron stars to obtain constraints on the excess angular momentum residing in the superfluid. Candidate Heating Processes in Old Pulsars ========================================== Fig. 1 shows a comparison of thermal evolution calculations (neglecting possible internal heating) with the surface temperature measurements of eight neutron stars. Residual heat is adequate to explain the temperatures of the younger objects, but cannot account for the temperatures of PSR 1929+10 and PSR 0950+08. Pavlov, Stringfellow & Córdova (1996) have measured temperatures of $T_{s}^{\infty} = 1.0 - 3.0 \times 10^{5}$ K for PSR 1929+10 and $6.6 - 7.4 \times 10^{4}$ K for PSR 0950+08 (see §5.1 for further discussion of the observations). The discrepancy between these measurements and the predictions of cooling theory is far too large to be accounted for by atmospheric uncertainties or modification of the energy transport by a magnetic field, and internal heating is required. We now discuss several candidate heating processes: superfluid friction, structural relaxation, chemical disequilibrium and magnetic field decay. Differential rotation between the neutron star crust and the neutron superfluid would generate heat through friction. If the superfluid and crust are in rotational equilibrium with respect to each other (both slowing down at the same rate), the heating power is (Shibazaki & Lamb 1989; Van Riper, Link & Epstein 1995; Umeda  1993; see eq. \[\[pp\]\] below) $$\label{rehe} H (t) = {\Delta J_s} \big\vert {\dot\Omega}_0 \big\vert \equiv I_s {\bar\omega} \big\vert {\dot\Omega}_0\big\vert,$$ where $\Delta J_s$ is the excess angular momentum residing in the superfluid, $I_s$ is the moment of inertia of the portion of the superfluid that is differentially rotating, $\bar\omega$ is the angular velocity difference between the two components averaged over the superfluid moment of inertia and $\vert {\dot\Omega}_0\vert$ is the spin-down rate. In principle, the heat could be generated anywhere in the star in which there are superfluid neutrons. Usually, however, the superfluid in the core is regarded as tightly coupled to the rotation of the solid through Fermi liquid effects ([@ALS]). Analyses of glitch data ([@AEO]) and spin variations in accreting neutron stars ([@Boyntonetal]) and in isolated pulsars ([@boynton]; [@deeter]) support this picture. In the inner crust, however, interaction between crustal nuclei and superfluid vorticity could lead to substantial differential rotation (see, e.g., [@AI]) and heating. The velocity difference that can develop is determined by uncertain microphysics, however, $\bar\omega$ could exceed $\sim 10$ rad s$^{-1}$ ([@EB]; [@pizzo]). In a star with a stiff equation of state and a thick crust, $\Delta J_s$ could then be as large as $\sim 6\times 10^{45}$ ergs s, in which case the heating power is approximately $$\label{heatmag} H (t) = 0.8 \,\left ({\Delta J_s}\over {6\times 10^{45} \ {\rm ergs \ s}} \right ) \left ( {P\over {0.2\ {\rm s}}}\right )^{-1} \left ( { {t_{\rm age}}\over {10^6\ {\rm yr}} }\right )^{-1}\ L_\odot,$$ where $P$ is the star’s spin period, $t_{\rm age}$ is the spindown age $\equiv P/2\dot P$, and a rapid initial spin rate was assumed. A heating rate this large would begin to play an important role in thermal evolution after $\sim 10^5$ yr, when $H(t)$ becomes comparable to the luminosity from residual heat. Another process that heats the star is structural relaxation occurring as the star spins down and becomes less oblate. The neutron star crust probably becomes brittle when its temperature drops below $\sim 10^8$ K (Ruderman 1976) at an age of $\sim 10^4$ yr, and subsequently suffers structural relaxation through violent starquakes. The rate of heat generation is then of order (Cheng  1992), $$\dot E_{\rm quake} \sim {B \theta_c^2\over t_{\rm age}} = 10^{-5} \left ({B\over 10^{48}\ {\rm erg}}\right ) \left ({\theta_c\over 10^{-3}}\right )^2 \left ({t_{\rm age}\over 10^6\ {\rm yr}}\right )^{-1} L_\odot,$$ where $B$ is the shear modulus of the crust and $\theta_c$ is the critical strain angle at which the crust breaks. The critical strain angle is quite uncertain. Cheng  (1992) have shown that structural relaxation could constitute an important heat source in older stars if $\theta_c$ is $\gap 10^{-2}$. However, a value of $\theta_c$ this large is close to that for a perfect Coulomb lattice, while lattice imperfections most likely make $\theta_c$ considerably smaller (Smolukowski 1970). We assume that superfluid friction dominates starquake heating. Reisenegger (1995) has studied heating arising as a neutron star spins down and compression of the matter drives it from chemical equilibrium. For the $\sim 10^{12}$ G fields expected for most isolated neutron stars this process appears to be relatively unimportant, but could be relevant in stars with small magnetic fields ($\lap 10^{10}$ G). Reisenegger cautions, however, that chemical disequilibrium in superfluid neutron matter could give larger heating rates than those he estimates for normal matter. Thompson & Duncan (1996) have studied heat generation by the decay of a strong magnetic field, and argue that the dominant decay process of a $\gap 10^{14}$ G field is the irrotational mode of ambipolar diffusion (Goldreich & Reisenegger 1992). Equating the rate of energy loss by field decay to the photon luminosity gives a surface temperature of $$4\pi\sigma R^{2}T_{s}^{4} \simeq \left(\frac{4\pi R^{3}}{3}\right)\left(\frac{B_{p}^{2}}{8\pi t_{ambip}^{irr}}\right)$$ with $$t_{ambip}^{irr} \simeq \frac {5 \times 10^{15}}{T_{8}^{6}B_{12}^{2}}{\rm yr}.$$ Relating the surface temperature to the internal temperature using the results of Gudmundsson, Pethick & Epstein (1982; see eq. \[\[gpe\]\] below) we obtain $$T_{s} = 5.8 \times 10^{5}\left(\frac{B} {10^{16}{\rm G}}\right)^{-0.57}.$$ The fields required to account for the observed temperatures of PSRs 1929+10 and 0950+08 are $\sim 10^{16}$ G and $\sim 10^{17}$ G, respectively. By contrast, the [*dipole*]{} fields of these objects inferred from the vacuum dipole model are only $\sim 10^{11}$ G. Thus, heating by magnetic field decay appears an unlikely heat source to power the emission of PSRs 1929+10 and 0950+08. Steady Heating From Superfluid Friction ======================================= The first situation we consider is internal heating arising from steady slow-down of the neutron superfluid. In the next section, we study perturbations to this equilibrium state. We describe a neutron star as consisting of two components - a solid crust and an interior liquid (the superfluid). The crust is acted upon by an external torque, and the crust and superfluid are coupled through friction. Regardless of the system’s initial spin state, eventually an equilibrium will be reached in which the two components are both spinning down at the same rate, with the liquid spinning more rapidly than the solid to a degree determined by the external torque and the strength of the frictional coupling. In this state of rotational equilibrium, friction between the two components generates heat at a rate that is nearly constant (for an external torque that is changing slowly). The rate of heat production is given by the difference between the rate of change of the total rotational energy of the star and the rate at which work is done by the external torque ([@SL]; Van Riper, Link & Epstein 1995): $$\begin{aligned} H (t) &=& N_{ext}\Omega_c(t) -\frac{d}{dt}\left[\frac{1}{2}I_c\Omega_c^2(t) + \frac{1}{2} \int dI_{s}\,\Omega_s^2({\hbox{\bf r}},t)\right] \nonumber \\ &=& \int dI_{s}\vert\dot\Omega_{s}({\mathbf r},t)\vert\omega({\mathbf r},t), \label{ee}\end{aligned}$$ where $N_{ext}$ is the external braking torque, $\Omega_c (t)$ is the angular velocity of the crust and any components of the star tightly coupled to it, $\Omega_s ({\mathbf r},t)$ is the superfluid angular velocity at position ${\mathbf r}$ in the star, $\omega({\mathbf r},t) \equiv \Omega_s({\mathbf r},t) - \Omega_c(t)$ is the angular velocity [*lag*]{} between the superfluid and the crust, and $I_s$ is the superfluid moment of inertia. Eq. \[\[ee\]\] gives the heating rate whether the two components are in rotational equilibrium or not. In rotational equilibrium, the crust superfluid is everywhere spinning down at the rate of the crust $\vert\dot\Omega_0 (t)\vert$, and the heating rate is $$H = \Delta J_s |\dot\Omega_0 (t)|, \label{pp}$$ where $\Delta J_s\equiv \int dI_s \omega_0 ({\mathbf r},t)$ is the excess angular momentum in the superfluid and $\omega_0 ({\mathbf r},t)$ is the lag in equilibrium. After $\sim 10^6$ yr, when the star has lost most of its residual heat, the heating rate is approximately balanced by cooling through the emission of surface photons, , $$4\pi\sigma R_{\infty}^{2}T_{s,\infty}^{4} = \Delta J_s |\dot\Omega_{0,\infty} (t)|, \label{steadyrate}$$ where $R_{\infty}$ is the radius, $T_{s,\infty}$ is the surface temperature and $|\dot\Omega_{0,\infty} (t)|$ is the spin-down rate (subscript $\infty$ indicates quantities seen by a distant observer; $\Delta J_s$ is evaluated at the stellar surface). Observed stellar quantities are related to their values at the surface through the redshift $e^{-\Phi} \equiv (1 - 2GM/Rc^2)^{-1/2}$ as $$T_{s,\infty} = e^{\Phi}T_{s} \qquad |\dot\Omega_{0,\infty} (t)| = e^{2\Phi}|\dot\Omega_{0} (t)| \qquad R_{\infty} = e^{-\Phi}R.$$ In §5, we will apply eq. \[\[steadyrate\]\] to PSRs 1929+10 and 0950+08 to obtain estimates for the values of $\Delta J_s$ required to heat these sources to their observed temperatures. Thermal-rotational Instability ============================== We now study the stability of the rotational equilibrium described above by determining how a neutron star containing a pinned inner crust superfluid responds to a perturbation of its thermal and rotational state. This problem was originally considered by SM under the simplifying assumption that the superfluid angular velocity has no gradients. We extend their work to account for gradients in the superfluid angular velocity lag (as would arise from vortex pinning) and the effects of quantum tunneling and vortex self-energy on the superfluid dynamics. We find (as did SM) that under some circumstances a feedback instability that couples the star’s thermal and rotational states can occur. In contrast to SM however, we find that the thermal-rotational instability cannot occur in stars older than $\sim 10^{6}$ yr. While our discussion will be formulated with coupling to the inner crust superfluid in mind, many of our results are general and can be applied to coupling with a core liquid. Perturbation Analysis --------------------- The star’s total angular momentum changes under an external torque as (neglecting general relativistic effects) $$\dot J_{\rm tot}(t) = I_c\dot\Omega_c(t) + \int dI_s\,\dot\Omega_s({\mathbf r},t) = N_{\rm ext}(t) \equiv -I\vert\dot\Omega_0 (t)\vert, \label{jdot}$$ where $I_c$ is the moment of inertia of the crust plus any other component(s) to which it is effectively coupled (, the core), $I_s$ is the superfluid component in differential rotation, and $I \equiv I_c + I_s$ is the total moment of inertia. In rotational equilibrium, the crust superfluid, the crust and the core are all spinning down at a rate $\vert\dot\Omega_0(t)\vert$. A neutron star becomes isothermal within $\sim 10^4$ yr after its birth (Van Riper 1991; Umeda  1993). The star’s thermal evolution is then governed by $$C_v\dot T(t) = H + \Lambda_{c}, \label{FF}$$ where $T$ is the internal temperature, $C_v$ is the heat capacity, $H$ is the internal heating rate and $\Lambda_{c}$ is the cooling rate. The star cools through emission processes that depend on age and composition. We parameterize the cooling rate as $$\Lambda_{c} = -BT^n(t), \label{lambda}$$ where $B$ and $n$ are constants which depend on the cooling mechanism. For the first $\sim 10^5$ yr of a neutron star’s thermal evolution, the dominant mode of energy loss is through modified ($n=8$) or direct ($n=6$) URCA reactions. Later, the star cools primary through the emission of photons from the surface ($n\simeq 2.2$). Combining eq. \[\[pp\]\] for the equilibrium heating rate with eqs. \[\[FF\]\] and \[\[lambda\]\] gives the temperature evolution equation for the equilibrium state: $$C_{v,0} \dot T_0 = - B T_0^n + \Delta J_s\vert\dot{\Omega}_0(t)\vert, \label{Tss}$$ where here and henceforth the subscript “0” denotes equilibrium quantities, themselves functions of time. The heating rate in eq. \[\[ee\]\] is determined by the rotational state of the superfluid. The superfluid obeys the equation of motion ([@bc]), $${\partial {\mathbf w}({\mathbf r}, t) \over \partial t} + {\mathbf\nabla}\times\left [{\mathbf w}({\mathbf r}, t) \times {\mathbf v_v}({\mathbf r}, t)\right ] = 0, \label{vorticity}$$ where ${\mathbf w}\equiv {\mathbf\nabla} \times {\mathbf v_s}$ is the superfluid vorticity, ${\mathbf v_s}$ is the fluid velocity and ${\mathbf v_v}$ is the vortex velocity. All quantities are averaged over regions containing many vortices. The circulation around any contour is given by $$\oint d{\mathbf l}\cdot {\mathbf v_s} = \int_A d{\mathbf A}\cdot {\mathbf w} = \kappa N, \label{circ}$$ where $N$ is the number of vortices surrounded by the contour of area $A$ and $\kappa$ is the quantum of circulation ($h/2m_n$; $m_n=$ neutron mass). The component of the vorticity along the axis of rotation is related to the areal density $n$ of vortices in the perpendicular plane. For rotation along the $z$-axis, eq. \[\[circ\]\] gives $${\rm w}_z ({\mathbf r}, t) = \kappa n ({\mathbf r}, t).$$ The $z-$component of eq. \[\[vorticity\]\] gives the conservation law $${\partial n ({\mathbf r}, t)\over\partial t} + {\mathbf\nabla}\cdot \left [n ({\mathbf r}, t){\mathbf v_v}({\mathbf r}, t) \right ] = 0. \label{continuity}$$ We shall focus on axisymmetric superfluid rotation. In this case the $z$-component $\Omega_s$ of the superfluid angular velocity a distance $r_p$ from the rotation axis is given by $$\oint d{\mathbf l}\cdot {\mathbf v_s} = 2\pi r_p^2 \Omega_s({\mathbf r}, t) = \kappa \int_0^{r_p} dr_p^\prime\, 2\pi r_p^\prime n ({\mathbf r}, t).$$ From eq. \[\[continuity\]\], the equation of motion is $$\dot\Omega_s({\mathbf r},t) = -{w_z\over r_p} v(\omega,T) = -v(\omega,T)\left(\frac{2}{r_p} + \frac{\partial} {\partial r_p}\right)\Omega_s({\mathbf r},t), \label{GG}$$ where $v(\omega,T)$, the average radial velocity of vortex lines, is determined by the microscopic processes that govern the vortex mobility. In the absence of pinning, the superfluid would approximate rigid body rotation so that $\partial\Omega_s/\partial r_p\simeq 0$. For simplicity, we assume that pinning introduces gradients in $\Omega_s$ that are negligible compared to $2\Omega_s/r_p$. We assume that thermal conduction maintains isothermality and neglect gradients in the perturbed temperature. We examine the response of the system to the following [*axisymmetric*]{} perturbations of its thermal and rotational states: $$\Omega_s({\mathbf r},t) = \Omega_{s,0}({\mathbf r}) + \delta\Omega_s({\mathbf r},t)$$ $$\Omega_c(t) = \Omega_{c,0} + \delta\Omega_c(t)$$ $$\omega({\mathbf r},t) = \omega_0({\mathbf r}) + \delta\omega({\mathbf r},t)$$ $$T(t) = T_0 + \delta T(t),$$ where unperturbed quantities are evaluated at the time of the initial perturbation. Equation (\[jdot\]) becomes $$I_c\delta\dot\Omega_c(t) + \int dI_s\,\delta\dot\Omega_s({\mathbf r},t) = 0. \label{aa}$$ To linear order in the perturbations, eq. \[\[FF\]\] becomes the following integro-differential equation: $$\begin{aligned} C_{v,0}\delta\dot T(t) = - C_{v,0}\frac{\dot T_0}{T_0}\delta T(t) &-& Bn T^{n-1}_0 \delta T(t) \nonumber \\ &+& \vert\dot{\Omega}_0\vert \int dI_{s}\delta\omega ({\mathbf r},t) - \int dI_{s}\omega_0 ({\mathbf r})\delta\dot\Omega_{s}({\mathbf r},t). \label{bb}\end{aligned}$$ Equation (\[GG\]) becomes $$\delta\dot\Omega_s({\mathbf r},t) = -\vert\dot{\Omega}_0\vert\left(\eta_{\omega}\frac{\delta \omega({\mathbf r},t)}{\omega_0({\mathbf r})} + \eta_{_T}\frac{\delta T(t)}{T_0} + \frac {\delta\Omega_s({\mathbf r},t)}{\Omega_{s,0}({\mathbf r})}\right), \label{cc}$$ where $$\eta_\omega = \frac{\partial\ln v}{\partial\ln\omega} \qquad\qquad \eta_{_T} = \frac{\partial\ln v}{\partial\ln T},$$ are evaluated in equilibrium and measure the sensitivity of the vortex radial velocity to changes in the lag and temperature. The quantities $\eta_\omega$ and $\eta_{_T}$ could have spatial dependence, but for the vortex velocity we will adopt below they are nearly constant. We seek separable solutions for the perturbed quantities of the form $\delta A({\mathbf r}, t) = \delta A({\mathbf r})e^{-i\omega t}$. Combining eqs. \[\[aa\]\] - \[\[cc\]\], we obtain $$\begin{aligned} i\omega \left[\frac{\eta_{\omega}\eta_{_T}}{T_{0}}I_0^{2}\right. &+& \left. \frac {\eta_{_T}I_{c}}{T_{0}}I_1 - C_{v,0}\eta_{\omega}I_{-1} - \frac{\eta_{\omega}\eta_{_T}}{T_{0}}I_1I_{-1} + I_{c}C_{v,0}\right] - C_{v,0}I_{c} {\dot T_{0}\over T_0} \nonumber \\ &-& BnT_{0}^{n-1}I_{c} + \frac {\vert\dot\Omega_{0}\vert I \eta_{_T}}{T_{0}}I_0 + \frac{C_{v,0}\dot T_{0}\eta_{\omega}}{T_{0}}I_{-1} + BnT_{0}^{n-1}\eta_{\omega}I_{-1} = 0, \label{intdisp}\end{aligned}$$ where $$I_n \equiv \int dI_s \omega_0^n \left(\frac{i\omega}{\vert\dot\Omega_{0}\vert} - \frac{\eta_{\omega}}{\omega_{0}({\mathbf r})} - \frac{1}{\Omega_{s,0}({\mathbf r})}\right)^{-1}$$ Evaluation of eq. \[\[intdisp\]\] requires a form for $\omega_{0}({\mathbf r})$. Pinning calculations indicate that the pinning energy per nucleus peaks near a density of $\sim 10^{14}$ g cm$^{-3}$ ([@EB]; [@pizzo]); in these regions, $\omega_0$ will be the largest. To model this behavior, we treat $\omega_0$ as taking a value $\omega_0=\omega_{0,{\rm max}}$ through a region of total moment of inertia $\Delta I_s$, and zero otherwise. Since the pinning energy is largest in the densest regions of the crust, $\Delta I_{s}$ probably accounts for most of the crust’s moment of inertia. Evaluating the integrals in eq. \[\[intdisp\]\], we find, $$\omega^{2} + i\omega Y - Z = 0, \label{mm}$$ with $$Y = \frac{\dot T_0}{T_{0}} + \frac{B n T_0^{n-1}}{C_{v,0}} + \tilde \Omega - \frac{\vert\dot\Omega_{s,0}\vert \eta_{_T} \omega_{0,{\rm max}}\Delta I_{s}}{C_{v,0} T_{0}}, \label{OO}$$ $$Z = \left ({\dot{T_{0}}\over T_0} + {B n T_0^{n-1}\over C_{v,0}} \right ) \tilde{\Omega} + {\eta_{_T}\vert\dot\Omega_{0}^{2}\vert I \Delta I_{s}\over C_{v,0}T_{0}I_{c}},$$ where $$\tilde \Omega \equiv \vert\dot\Omega_{s,0}\vert \left(\frac{1}{\Omega_{s,0}}+\frac{(I_{c}+\Delta I_{s})\eta_\omega}{I_c\omega_{0,{\rm max}}}\right).$$ Perturbations are damped when $Y>0$, unstable when $Y<0$, and marginally stable (or oscillatory) when $Y = 0$. The condition for a star to be unstable is, $$\frac{\dot T_0}{T_{0}} + \frac{B n T_0^{n-1}}{C_{v,0}} + \tilde \Omega - \frac{\vert\dot\Omega_{s,0}\vert \eta_{_T} \omega_{0,{\rm max}}\Delta I_{s}}{C_{v,0} T_{0}} \leq 0.$$ Using the equilibrium equation, eq. \[\[Tss\]\], we make the replacement $$BT_0^{n} = \vert\dot\Omega_{0}\vert\Delta I_s \omega_{0,{\rm max}} - C_{v,0} \dot T_0,$$ to obtain a quadratic equation for the critical temperature $T_c$, $$T_{c}^{2} + \frac {(n-1)I_{c}\omega_{0,{\rm max}}\vert\dot T\vert}{ \eta_\omega I \vert\dot\Omega_{0}\vert}T_{c} + \frac {(n-\eta_{_T})I_{c}\omega_{0,{\rm max}}^{2} \Delta I_{s}}{a I \eta_\omega} = 0, \label{quad}$$ where $a\equiv C_{v,0}/T_0$ is independent of temperature for degenerate matter. We have assumed $\tilde\Omega \sim \vert\dot\Omega_{0}\vert\eta_{\omega}I/\omega_{0,{\rm max}}I_{c}$, which is valid as long as $\omega_{0,{\rm max}} \ll \eta_\omega\Omega_{s}$. Eq. \[\[quad\]\] has a positive solution: $$T_{c} = - \frac {(n-1)\omega_{0,{\rm max}}I_{c}\vert\dot T\vert}{ 2\eta_\omega I \vert\dot\Omega_{0}\vert} + \frac{1}{2}\left(\frac {(n-1)^{2}\omega_{0,{\rm max}}^{2}I_{c}^{2}\vert\dot T\vert^{2}}{\eta_\omega^{2}I^{2} \vert\dot\Omega_{0}\vert^{2}} + \frac{4(\eta_{_T}-n)\omega_{0,{\rm max}}^{2}I_{c} \Delta I_{s}}{a \eta_\omega I}\right)^{1/2}.$$ Below we estimate $\eta_{_T} = \eta_\omega \simeq 30$. For a star of age $\sim 10^4$ yr, $\vert\dot T\vert$ is $\sim 10^{-7}{\rm K \ s^{-1}}$ and $\vert\dot\Omega_{0}\vert$ is $\sim 10^{-10}{\rm rad \ s^{-2}}$. Taking $\omega_{0,{\rm max}}\sim 10$ rad s$^{-1}$ and $n = 8$ (modified URCA cooling), we see that the first term under the square root is negligible compared to the second. The critical temperature is thus approximately $$T_c^2 \simeq \frac{(\eta_{_T} - n)\omega_{0,{\rm max}}^{2}I_{c}\Delta I_{s}}{a\eta_\omega I}. \label{JJ}$$ From eq. \[\[JJ\]\], we see that a minimum sensitivity of the vortex velocity to temperature is required for a thermal-rotational instability to occur; for $\eta_{_T}<n$, the star is stable at any temperature. Eq. \[\[JJ\]\] agrees with eq. \[24\] of SM in the limit $\Delta I_s=I_s$, $\eta_{_T} \gg n$ and $I_{c} \simeq I$. Eq. \[\[JJ\]\] can be applied for any coupling of the superfluid and crust that depends on $T$ and $\omega$. In principle, the superfluid component with excess angular momentum $\omega_{0,{\rm max}}\Delta I_s$ could be anywhere in the star. In the inner crust, however, pinning of vortices to the nuclear lattice ([@AI]; [@ruderman76]; Alpar 1977; Epstein & Baym 1988; [@pizzo]) could sustain significant differential rotation between the superfluid and the crust. The mobility of vortices in the presence of pinning determines the superfluid’s ability to respond to changes in temperature and rotation rate. Vortex dynamics in the presence of pinning has been studied in detail by Link & Epstein (1991; hereafter LE) and LEB under the assumption that vortex stresses do not break the nuclear lattice. We now review the key results of this work, and study the implications of vortex pinning for the thermal-rotational instability. Vortex Dynamics in the Presence of Pinning ------------------------------------------ Were superfluid vortices perfectly pinned to the inner crust lattice, the superfluid velocity would be fixed. As the solid crust slows under the external torque, a velocity difference between the crust and superfluid would develop exerting a Magnus force on the pinned vortices directed radially outward. If the lag $\omega$ locally exceeds a critical value $\omega_c$, vortices cannot remain pinned in the presence of the Magnus force; the vortices unpin and flow outward, spinning down the superfluid. The critical lag is determined by the condition that the Magnus force per unit length of vortex equal the pinning force per unit length (see, , LE) $${F_p\over l} = \rho_{s}\kappa r_p \omega_c, \label{pinforce}$$ where $F_p$ is the pinning force per nucleus, $l$ is the lattice spacing and $\rho_{s}$ is the superfluid mass density. For $\omega<\omega_c$, pinned vortices can still move outward as thermal or quantum excitations allow them to overcome their pinning barriers. The resulting average velocity of [*vortex creep*]{} is determined by the pinning strength, the properties of vortices, the characteristic energy of excitations on a pinned vortex, and the velocity difference between a pinned vortex and the superfluid flowing past it. Accounting for quantum effects and the vortex self-energy, LEB and LE obtain a creep velocity of the form (see eq. 6.9 of LEB) $$v_{cr} = v_0 \exp (- A (\omega)/T_{\rm eff}). \label{ddd}$$ Here $A$ is the [*activation energy*]{} for a segment of vortex line to overcome its pinning barrier; it decreases with $\omega$, becoming zero for $\omega=\omega_c$. The prefactor $v_{0}$ is a microscopic velocity comparable to the radial component of the velocity of an unpinned vortex segment. The radial velocity is determined by the dissipative processes associated with vortex motion. Epstein & Baym (1992) have shown that drag arising from the excitation of [*Kelvin modes*]{} causes free vortices to move radially outward at velocities comparable to the velocity difference between the superfluid and normal matter. For the pinning energies estimated by Epstein & Baym (1988), this velocity difference could be $\sim 10^6$ cm s$^{-1}$; we estimate $v_0\simeq 10^6$ cm s$^{-1}$. The effective temperature, $T_{\rm{eff}}$, is (eq. 4.10, LEB) $$T_{\rm eff} = T_q {\rm coth} {T_q\over T},$$ where $T_q$ is the [*cross-over temperature*]{} that determines the transition from vortex motion through thermal activation to quantum tunneling. For $T\gg T_q$, vortices move primarily through classical thermal activation. For $T\ll T_q$, the dominant process is quantum tunneling. In these two limits $$T_{\rm eff} \rightarrow \left\{ \begin{array}{ll} T & \mbox{for $T\gg T_q$ (classical)} \\ T_q & \mbox{for $T\ll T_q$ (quantum)} \end{array} \right.$$ In equilibrium, the superfluid and the crust are both spinning down at a rate $\vert\dot\Omega_0(t)\vert$. From eqs. \[\[GG\]\] and \[\[ddd\]\], the equilibrium state satisfies $${A(\omega_0)\over T_{\rm eff}} = \ln {4 v_0 t_{\rm age}\over r_p}, \label{steadystate}$$ where we took $\Omega_s/\vert\dot\Omega_0\vert\simeq 2 t_{\rm age}$. The pinning force is a function of density alone, and so is constant on spherical shells. From eq. \[\[steadystate\]\], we see that $\omega_0$ is nearly constant on such shells except near the rotational poles. For purposes of obtaining estimates, we henceforth take $r_p\simeq R$. For $v_0=10^6$ cm s$^{-1}$ and $t_{\rm age}=10^6$ yr, $$\ln {4 v_0 t_{\rm age}\over R} \simeq 30. \label{sslag}$$ In equilibrium, $A(\omega_0)$ must take a particular local value for given $T_{\rm eff}$ and $t_{\rm age}$. If the pinning force per nucleus is relatively small, $\omega_0$ must be $\ll\omega_c$ to ensure that $A(\omega)$ is sufficiently large. On the other hand, if the pinning force per nucleus is relatively large, $\omega_0$ must be close to $\omega_c$ to allow the vortices to overcome their pinning barriers. We will refer to these two limiting cases as [*strong*]{} and [*weak pinning*]{}, respectively. \[These cases correspond to the limits of [*flexible*]{} and [*stiff*]{} vortices discussed by LE\]. In the equilibrium state we have assumed, the local lag has everywhere adjusted to the value required to satisfy eq. \[\[steadystate\]\]. Evaluated about equilibrium, the sensitivity of the vortex velocity in eq. \[\[ddd\]\] to temperature is $$\eta_{_T} = \left [\ln {4 v_0 t_{\rm age}\over R}\right ] \left ({T_q\over T}\right )^2 {\rm csch}^2\,{T_q\over T}.$$ In the classical limit ($T\gg T_q$), $\eta_{_T}\simeq 30$. Below $T=T_q$, both $\eta_{_T}$ and $T_c$ quickly drop to zero as $T$ is reduced. Hence, for the limit cycle to be relevant, $T\gap T_q$ is required. The cross-over temperature $T_q$ is equal to half the ground state energy of excitations on a pinned vortex line, and depends sensitively on the pinning energy and density. For weakly-pinned vortices, LEB estimate (eq. 3.11, LEB) $$T_q \simeq 0.2 \left ({\Lambda\over 3}\right )\left ({l\over 50\ {\rm fm}}\right )^{-2}\ {\rm keV}, \label{Tqweak}$$ where $\Lambda$ is a weak function of density and we have chosen fiducial values appropriate to the denser regions of the inner crust. At lower density, , near the neutron drip density, $\Lambda$ is $\simeq 7$. For strongly-pinned vortices, pinning drives the ground state energy of vortex excitations up to a considerably higher value (eq. 3.13, LEB): $$T_q \simeq 60 \left ({\Lambda\over 3}\right )\left ({l\over 50\ {\rm fm}}\right )^{-2}\ {\rm keV}.$$ To compare these temperatures to those of cooling neutron stars, we convert from surface temperature to internal temperature using the results of Gudmundsson, Pethick & Epstein (1982): $$T_8 = 1.288\left(\frac{T^4_{s,6}}{g_{s,14}}\right)^{0.455} \label{gpe}$$ where $$g_s = \frac{GM}{R^{2}}e^{-\Phi}.$$ Here $T_8 \equiv T/10^8$, $T_{s,6} \equiv T_s/10^6$, $g_{s,14} \equiv g_s/10^{14}$ is the surface gravity, and $T_{s,\infty} = e^{\Phi}T_{s}$. For typical neutron star parameters, the internal temperature is $$T = 12\ T_{s,6,\infty}^{1.82}\ {\rm keV}.$$ For PSR 0950+08, we estimate 0.09 keV $<T<$ 0.11 keV. This object is thus well into the quantum creep regime, for which $\eta_{_T}\simeq 0$. The limit cycle cannot occur, and steady slowdown of the crust and superfluid is a stable state. Our conclusion regarding the stability of old stars differs from that of SM, who assumed that thermal creep is always the dominant process. For PSR 1929+10 the case is less clear; we find 0.18 keV $<T<$ 1.3 keV, compared to $T_q\simeq 0.2$ keV estimated in eq. \[\[Tqweak\]\]. However, this estimate applies only to the special case of extremely weak pinning; $T_q$ could be substantially higher than $0.2$ keV. It thus appears that PSR 1929+10 is also in the quantum creep regime (or borderline), undergoing steady slow-down. Stars younger than PSR 1929+10 ($t_{\rm age}=3\times 10^6$ yr), could be in the thermal creep regime, and hence could be subject to the thermal-rotational instability. Such a star becomes unstable if the temperature falls below the critical temperature given by eq. \[\[JJ\]\]. The quantity $\eta_\omega$, which measures the sensitivity of the creep rate to changes in $\omega$ is, from eqs. \[\[ddd\]\] and \[\[steadystate\]\], $$\eta_\omega = - \left [\ln {4 v_0 t_{\rm age}\over R}\right ] {d\ln A\over d\ln\omega}.$$ The form for $A(\omega)$ depends on the strength of pinning. In the weak and strong pinning limits, the activation energy is (eq. B.12, LE, in the limit $\omega\ll\omega_c$; eq. 3.15, LE) $$A (\omega) \rightarrow \left\{ \begin{array}{ll} {\beta\over\omega} & \mbox{weak pinning} \\ U_0 (1 - {\omega\over\omega_c})^{3/2} & \mbox{strong pinning}, \end{array} \right. \label{activation}$$ where $U_0$ is the pinning energy per nucleus, and $\beta$ is a parameter that measures the strength of the coupling and is related to $U_0$. For strong pinning, $U_0$ is relatively large ($\gap$ 1 MeV), and $T_q$ is up to $\sim 60$ keV. A neutron star is expected to cool below this temperature within $\sim 100$ years of its birth. If strong pinning occurs, therefore, quantum tunneling is the dominant creep process during most of the star’s thermal evolution, and the thermal-rotational instability cannot occur. Hence, the weak pinning case is the relevant one for the thermal-rotational instability. In this case $$\eta_\omega = \ln {4 v_0 t_{\rm age}\over R} \simeq 30.$$ Taking $\eta_{_T}=\eta_\omega\simeq 30$, we estimate the internal temperature at which the star becomes unstable from eq. \[\[JJ\]\]. In the limit $I_{s}\ll I_{c}$, appropriate if it is the inner crust superfluid that drives the instability, we obtain $$T_c = 11 \left ({a\over 3.3\times 10^{29}\mbox{ erg s$^{-1}$}}\right )^{-1/2} \left ({\omega_{0,{\rm max}}\over 10\mbox{ rad s$^{-1}$}} \right ) \left ({\Delta I_s\over 7.3\times 10^{43} \mbox{ g cm$^2$}}\right )^{1/2} \,{\rm keV}, \label{Tcest}$$ where $n=8$ for modified URCA cooling in a young neutron star. Here and in the following, we estimate stellar parameters such as $I_s$, $R$ and $a$ using a $1.4 M_\odot$ stellar model based on the Friedman & Pandharipande (1981) equation of state (see Table 1). A cooling neutron star reaches a temperature of $\sim 10$ keV after $\sim 10^4$ yr. While the critical temperature depends sensitively on the uncertain pinning parameters $\omega_{0,{\rm max}}$ and $\Delta I_s$, this estimate suggests that young, cooling neutron stars [*could*]{} be unstable to perturbations in temperature and rotation rate. If, on the other hand, $\omega_{0,{\rm max}}$ or $\Delta I_s$ are significantly smaller than estimated in eq. \[\[Tcest\]\], the star will cool into the quantum creep regime before it can become unstable. As a star cools and reaches its critical temperature, temperature perturbations begin to grow. Stability is restored as the star is heated to slightly above its critical temperature. The star is again able to cool, eventually becoming unstable again. A limit cycle ensues, wherein the star oscillates about its $T_c$ as originally demonstrated by SM. To evaluate the characteristic period of the oscillations about this marginally-stable state, we solve eq. \[\[mm\]\] with $Y = 0$. Using eq. \[\[Tss\]\] and assuming $I_{s} \ll I_{c}$, we obtain $$\tau_{osc} \simeq 2\pi\left[\frac{\eta_{_{T}}(1+\eta_{\omega})(n-1)BT_{0}^{n-1} \vert\dot\Omega_{0}\vert}{a T_{0}(\eta_{_T}-1)\omega_{0,{\rm max}}}+\frac{\eta_{\omega}(\eta_{_T}+\eta_{\omega}) \vert\dot\Omega_{0}\vert^{2}}{(\eta_{_T}-1)\omega_{0,{\rm max}}^{2}}\right]^{-1/2}. \label{tosc}$$ For a young star cooling through the modified URCA process, the first term is negligible for $\omega_{0,{\rm max}}\simeq 1-10$ rad s$^{-1}$. Taking $\eta_{_T}=\eta_\omega=30$, $\Omega_{s,0}/|\dot\Omega_0| \simeq 2 t_{age}$ and the spin period as $P\simeq 2 \pi/\Omega_{s,0}$ gives $$\tau_{osc} \simeq 0.26 \left({P\over 0.1\ {\rm s}}\right) \left( {\omega_{0,{\rm max}}\over 10 \ {\rm rad \ s^{-1}}}\right) t_{age}.$$ Constraints from Surface Temperature Measurements ================================================= Old Pulsars ----------- Pavlov, Stringfellow & Córdova (1996) have recently detected thermal emission from PSRs 1929+10 and 0950+08 in the UV-optical band using the COSTAR corrected Faint Object Camera on the Hubble Space Telescope. Assuming the observed flux arises from the entire surface of a neutron star with radius 10 km, they obtain surface temperatures of $T_{s}^{\infty} = 1.0 - 3.0 \times 10^{5}$ K for PSR 1929+10 and $6.6 - 7.4 \times 10^{4}$ K for PSR 0950+08. Previous X-ray observations of PSRs 1929+10 (Yancopoulos, Hamilton, & Helfand, 1994) and 0950+08 (Manning & Willmore, 1994) produced blackbody fits for emitting regions of only $\sim$ 20-30 meters in diameter, suggesting that the observed X-ray emission originates from a hot polar cap. However, for both of these objects, extension of the blackbody spectra into the UV-optical range predicts a flux which is several orders of magnitude smaller than that observed by Pavlov, Stringfellow & Córdova (1996), consistent with the interpretation that the UV-optical emission originates from the entire neutron star surface. The analysis in §4 indicates that these two sources are too cold for the thermal-rotational instability to occur. Assuming steady heating by superfluid friction, we determine from eq. \[\[steadyrate\]\] the values of the excess angular momentum $\Delta J_s$ and average lag $\overline\omega\equiv\Delta J_s/I_s$ required to heat these objects to their observed temperatures. We find $\Delta J_{s} \sim 4 \times 10^{43}$ ergs s and $\overline\omega \sim 0.6$ rad s$^{-1}$ for PSR 1929+10, and $\Delta J_{s} \sim 1 \times 10^{42}$ ergs s and $\overline\omega \sim 0.02$ rad s$^{-1}$ for PSR 0950+08 (see Table 2). Our estimates were obtained for an FP equation of state, which gives a radius $R\simeq 11$ km, close to that assumed for the surface temperature determinations. Our results are consistent with earlier upper limits on $\bar\omega$ obtained for PSR 1929+10. Shibazaki & Lamb (1989) obtained $\bar\omega < 0.02$ rad s$^{-1}$ and $\bar\omega < 4.0$ rad s$^{-1}$ for stiff and soft equations of state. Alpar  (1987) found $\bar\omega < 0.7$ rad s$^{-1}$ for a moderate equation of state. Some amount of internal heating also appears to be required for PSR 1055-52 (see Fig. 1). Ögelman & Finley (1993) obtained a temperature of $T_{s}^{\infty} = 6.9 - 8.1 \times 10^{5}$ K for this pulsar using ROSAT PSPC data. This temperature was obtained by interpreting the soft blackbody component of the spectrum as originating from a cooling neutron star of radius $\simeq$ 10 km. If we assume that steady heating by the internal superfluid provides the heat for this pulsar, we obtain $\Delta J_{s} \sim 5 \times 10^{44}$ ergs s and $\overline\omega \sim 7$ rad s$^{-1}$. Young Pulsars ------------- Surface temperature measurements and upper limits for young pulsars are given in Table 3, all fits are blackbody fits with a stellar radius of 10 km. These pulsars could be subject to the thermal-rotational instability discussed in §4. We consider PSR 1055-52 along with the younger pulsars because its internal temperature is high enough that the limit cycle cannot be ruled out. For most neutron stars younger than $\sim 10^{6}$ yr, the observed temperatures can be accounted for by their residual heat content. As discussed in §4, if the thermal-rotational instability occurs, a cooling neutron star cannot cool below its critical temperature. By requiring the internal temperature deduced from the observed surface temperature to be greater than or equal to the critical temperature, we obtain from eq. \[\[JJ\]\] the constraint, $$4.1 \times 10^{4}\left(\frac{(\eta_{_T} - n) \omega_{0,{\rm max}}^{2}\Delta I_{s}}{a \eta_\omega R^{1.82} e^{-2.73\Phi}} \right)^{0.275} \leq T_{s,\infty}, \label{fff}$$ giving an upper limit on $\omega_{0,{\rm max}}^{2}\Delta I_{s}$ inasmuch as $R$, $\Phi$, $\eta_{_T}$, $a$ and $n$ are known. Superfluid pinning is expected to be strongest in the densest regions of the inner crust; the characteristic lag $\omega_{0}({\mathbf r})$ will be largest in these regions. It is reasonable to expect then that $\omega_{0,{\rm max}}>\bar\omega\equiv I_s^{-1}\Delta J_s$. For this situation, we obtain the constraint, $$\overline\omega \leq \left(\frac {\omega_{0,{\rm max}}^{2}\Delta I_{s}}{I_{s}}\right)^{1/2}. \label{ooo}$$ Where $\omega_{0,{\rm max}}^{2}\Delta I_{s}$ is obtained with eq. \[\[fff\]\]. For stars younger than $\sim 10^5$ yr, the dominant cooling process is neutrino emission; we assume $n=8$ (modified URCA process). We take $\eta_{_T} = \eta_{\omega}=30$, as estimated in § 4, and the stellar parameters of Table 1 for $R$, $\Phi$ and $a$. In Table 3 we list constraints from eqs. \[\[fff\]\] and \[\[ooo\]\] for young neutron stars. We obtain upper limits on $\bar\omega$ of $\sim 30$ rad s$^{-1}$, typically. Note that these constraints apply [ *only*]{} if the dominant creep process is classical thermal activation; if quantum tunneling is the dominant process, the star is stable at any temperature. Discussion and Conclusions ========================== PSRs 1929+10 and 0950+08 require significant internal heating to account for their observed temperatures. A promising candidate heat source is friction between the neutron star crust and the superfluid it contains. In this paper we have studied the effects of superfluid friction on the long-term thermal and rotational evolution of a neutron star. We conclude that average differential rotation between the superfluid and the crust of $\bar\omega\sim 0.6$ rad s$^{-1}$ and $\sim 0.02$ rad s$^{-1}$ would account for the temperatures of PSRs 1929+10 and 0950+08 respectively. A larger lag, $\bar\omega\sim 7$ rad s$^{-1}$, is compatible with the temperature of PSR 1055-52. These differential velocities could be sustained by the pinning of superfluid vortices to the inner crust lattice. Pinned vortices can creep outward through thermal fluctuations or quantum tunneling, depending on the pinning strength and stellar temperature. For thermally-activated creep, the coupling between the superfluid and crust is highly sensitive to temperature. Under some circumstances, a feedback instability can occur that brings the superfluid and crust closer to corotation and heats the star until stability is restored. A hysteresis develops in which the star oscillates about its critical temperature. For stars older than $\sim 10^6$ yr, however, vortex creep occurs through quantum tunneling, and the creep velocity is too insensitive to temperature for a thermal-rotational instability to occur; these stars are stable. Our conclusion regarding the stability of old stars differs from that of Shibazaki & Mochizuki (1994), who assumed that thermal creep is always the dominant process. The thermal-rotational instability could, however, occur in younger stars. Assuming that young stars are stable or marginally stable leads to upper limits on the superfluid differential velocity of $\sim 10$ rad s$^{-1}$. These upper limits are consistent with the estimates for $\bar\omega$ obtained for the older PSRs 1929+10 and 0950+08. The estimates we obtain for $\bar\omega$ are consistent with first-principles calculations of the maximum lag sustainable by vortices before unpinning. Based on the pinning calculations of Alpar, Cheng & Pines (1989) and Ainsworth, Pines & Wambach (1989), Van Riper, Link & Epstein (1995) obtain an upper limit to the average lag velocity of $\bar\omega\sim 10$ rad s$^{-1}$. From the pinning calculations of Epstein & Baym (1988), Van Riper, Link & Epstein (1995) obtain an upper limit of $\bar\omega\sim 10^2$ rad s$^{-1}$. The recent calculations of Pizzochero, Viverit & Broglia (1997) give a pinning force of $F_p=0.63$ MeV fm$^{-1}$ at a density $\rho_s=8\times 10^{13}$ g cm$^{-3}$, and a pinning energy of 7.5 MeV; the corresponding critical lag is $\omega_c=16$ rad s$^{-1}$. To estimate the pinning strength required to sustain differential rotation of the magnitudes estimated above, we take $F_p$ appearing in eq. \[\[pinforce\]\] to be $U_0/r_0$, where $r_0$ is the effective range of the pinning potential. With the lag from PSR 1929+10 we obtain, $$U_{0} \gap 0.5\, {\rm MeV} \left(\frac{\bar\omega}{0.6\ {\rm rad \ s^{-1}}}\right)\left(\frac{\rho_{s}}{10^{14} \ {\rm g \ cm^{-3}}}\right)\left(\frac{R}{10\ {\rm km}}\right)\left(\frac{r_{0}}{10\ {\rm fm}}\right)\left(\frac{l}{50\ {\rm fm}}\right). \label{pinenergy}$$ Eq. \[\[pinenergy\]\] represents a lower limit since $\bar\omega < \bar{\omega}_{c}$, however, $\bar\omega$ could be close to $\bar{\omega}_c$ for pinning strengths this large (LEB). Our estimates of $\bar\omega$ for PSR 0950+08 imply a pinning energy $\gap$ 0.02 MeV. The thermal-rotational instability described in this paper might produce oscillations in the temperature and spin-down rate of younger pulsars with a characteristic period of $\sim 0.3\ t_{\rm age}$. Oscillations in the Crab pulsar, for example, could occur over a timescale of $\tau_{osc}\sim 10^2$ yr. Detection of such long-period oscillations, especially in the presence of timing irregularities (, glitches), would be problematic. Observational evidence for the thermal-rotational instability in any neutron star would offer valuable insight into the manner in which the neutron star crust is coupled to its superfluid interior. Our estimates for the excess angular momentum $\Delta J_s$ required to heat PSRs 1929+10 and 0950+08 to their observed temperatures hold whether the frictional heat is generated in the crust or in the core. In obtaining constraints on the average lag $\bar\omega$, we assumed coupling to the inner crust superfluid. If the coupling is elsewhere, these estimates scale as $\Delta J_s/I_s$, where $I_s$ is the moment of inertia of the component that possesses differential rotation. The upper limits on $\Delta I_s\omega^2_{0,{\rm max}}$ obtained for young pulsars apply for the crust since the coupling parameters $\eta_{_T}$ and $\eta_\omega$ were determined for vortex creep. An issue that complicates all interpretations of surface temperatures from cooling neutron stars is the uncertainty in atmospheric composition. We used the results of blackbody fits in our analysis. Temperature measurements are also available for several pulsars using model atmospheres. Heavy-element atmospheric models produce temperatures similar to the blackbody results. However, non-magnetic, light-element atmospheres can give temperatures up to three times lower than those obtained with blackbody fits (Romani 1987). If the temperature were in fact three times lower than the blackbody value, $\overline\omega$ estimated for PSRs 1929+10, 0950+08 and 1055-52 would decrease by almost two orders of magnitude. The upper limits on $\overline\omega$ obtained for younger stars would decrease by a factor of three. Hence, the results presented in Tables 2 and 3 are conservative and could decrease with improved atmospheric considerations and temperature measurements. Sudden increases in pulsar rotation rates ([*glitches*]{}) have been observed in many younger pulsars and are thought to represent angular momentum transfer from the superfluid to the crust. In the Vela pulsar, for example, fractional changes in the rotation rate of the crust of $\sim 10^{-6}$ are observed every few years ([@CDKP]). The maximum angular momentum available for a glitch is $\Delta J_s$, giving a maximum glitch magnitude of $$\frac{\Delta \Omega_{c}}{\Omega_{c}} \simeq \frac{\Delta J_s}{I_{c}\Omega_{c}} \simeq 2.6 \times 10^{-2}\left(\frac{\Delta J_s}{2 \times 10^{45}\ {\rm ergs\ s}}\right) \left(\frac{I_{c}}{1.1 \times 10^{45}\ {\rm g \ cm^{2}}}\right)^{-1} \left(\frac{\Omega_{c}}{70 \ {\rm s^{-1}}}\right)^{-1}.$$ The upper limit of $\bar\omega\simeq 30$ rad s$^{-1}$ obtained for Vela gives $\Delta J_s\simeq 2\times 10^{45}$ ergs s for the inner crust angular momentum excess, easily compatible with the angular momentum requirements of glitches. The smaller values obtained for PSRs 1929+10 and 0950+08 are also adequate to produce Vela-sized glitches, though none has been observed. We thank K. Van Riper for providing us with the results of cooling simulations and G. Pavlov and D. Page for helpful discussions. MBL would like to thank the Patricia Roberts Harris Graduate Fellowship for support. This work was supported by NASA EPSCoR grant \#291748. Abney, M., Epstein, R. I., & Olinto, A. 1996, , 466, L91. Ainsworth, T., Pines, D., & Wambach, J. 1989, Phys. Lett. B, 222, 173. Alpar, M. A. 1977, , 213, 527. Alpar, M. A., Anderson, P.W., Pines, D. & Shaham, J. 1984, , 276, 325. Alpar, M. A., Brinkmann, W., Kizilo$\breve{\hbox{\rm g}}$lu, Ü., Ögelman, H., & Pines, D. 1987, , 177, 101. Alpar, M. A., Cheng, K.S., & Pines, D. 1989, , 346, 823. Alpar, M. A., Langer, S. A., & Sauls, J. A. 1984, , 282, 533. Anderson, P.W., & Itoh, N. 1975, Nature, 256, 25. Baym, G. & Chandler, E. 1983, J. Low Temp. Phys, 50, 57. Baym, G., Pethick, C. J., & Sutherland, P. 1971, , 170, 299. Baym, G. & Pines, D. 1971, Ann. Phys., 66, 816. Becker, W., & Ashenbach, B. 1995, in Lives of the Neutron Stars, ed M.A. Alpar, U. Kiziloglu, & J. van Paradijs (Dordrecht:Kluwer), 47. Boynton, P. E. 1981, in [*IAU Symposium 95, Pulsars*]{}, ed. R. Wielebinski & W. Sieber (Dordrecht: Reidel), p. 279. Boynton, P. E., Deeter, J. E., Lamb, F. K., Zylstra G., Pravdo, S. H., White, N. E., Wood. K. S., & Yentis, D. J. 1984, , 283, L53. Chau, H. F. & Cheng, K. S. 1993, , 47, 2707. Cheng, K. S., Chau, W. Y., Zhang, J. L., & Chau, H. F. 1992, , 396, 135. Cordes, J. M., Downs, G. S., & Krause-Polstorff, J. 1988, , 330, 847. Deeter, J. E. 1981, PhD thesis, University of Washington. Epstein, R. I., Baym, G. 1988, , 328, 680.  —1992, , 387, 276. Finley, J. P., Ögelman H., Kiziloğlu, Ü. 1992, , 394, L21. Friedman, B., & Pandharipande, V. R. 1981, Nucl. Phys. A, 361, 502. Goldreich, P., & Reisenegger, A. 1992, 395, 250. Greenstein, G. 1975, , 200, 281. Gudmundsson, E., Pethick, C., Epstein, R. I. 1982, , 259, L19. Hailey, C.J. & Craig, W.W., 1995, , 455, L151. Halpern, J. P. & Ruderman, M., 1993, , 415, 286. Harding, D., Guyer, R. A., & Greenstein, G. 1978, , 222, 991. Heyl, J.S., & Kulkarni, S.R., 1998, 506, L61. Link, B., & Epstein, R.I., 1991, , 373, 592 (LE). Link, B., Epstein, R. I., Baym, G. 1993, , 403, 285 (LEB). Manning, R.A. & Willmore, A.P. 1994, MNRAS, 266, 635. Migdal, A.B. 1959, , 13, 655. Ögelman H., Finley, J.P. 1993, , 413, L31. Ögelman H., Finley, J.P., Zimmermann, H.U. 1993, Nature, 361, 136. Pandharipande, V. R., & Smith, R. A. 1975, Nucl. Phys. A, 237, 507. Pavlov, G.G., Stringfellow, G.S., Córdova, F.A. 1996, , 467, 370. Pizzochero, P. M., Viverit, L. & Broglia, R. A. 1997, , 79, 3347. Reisenegger, A. 1995, , 442, 749. Romani, R.W., 1987, , 313, 718. Ruderman, M. 1976, , 203, 213. Sedrakian, A. O., Sedrakian, D. M. 1995, , 447, 305. Sedrakian, A. O., Cordes, J. M. 1998, astro-ph/9806042. Shibazaki, N. & Lamb, F. K. 1989, , 346, 808. Shibazaki N. & Mochizuki Y. 1994, , 438, 288 (SM). Smolukowski, R. 1970, , 24, 923. Thompson, C. & Duncan, R.C. 1996, , 473, 322. Tsuruta, S. 1998, Phys. Rep., 291, 1. Umeda, H., Shibazaki, N., Nomoto, K., & Tsuruta, S. 1993, , 408, 186. Van Riper, K. A. 1991, , 75, 449. Van Riper, K. A., Epstein, R. I., & Miller, G. S. 1991, , 381, L47. Van Riper, K. A., Link, B., Epstein, R.I. 1995, , 448, 294. Yancopoulos, S., Hamilton, T.T., & Helfand, D. 1994, , 429, 832. [ll]{}\ $M$ (${\rm M_{\odot}}$) & 1.4\ $R$ (km) & 10.9\ $e^{-\Phi}$ & 1.27\ $I_s$ (g  ${\rm cm^2}$) & $7.3 \times 10^{43}$\ $a$ (${\rm ergs \ K^{-2}}$) & $3.3 \times 10^{29}$\ \ \[bbbb\] [lcccll]{} 1929+10 & 6.49 & 5.00 - 5.48 & 1.4 $\times 10^{-13}$ & 9.8 $\times 10^{41}$ - 8.1 $\times 10^{43}$ & 0.01 - 1.1 0950+08 & 7.23 & 4.82 - 4.87 & 2.3 $\times 10^{-14}$ & 1.1 $\times 10^{42}$ - 1.8 $\times 10^{42}$ & 0.01 - 0.02 1055-52 & 5.73 & 5.84 - 5.91 & 9.4 $\times 10^{-13}$ & 3.3 $\times 10^{44}$ - 6.3 $\times 10^{44}$ & 4.6 - 8.6 \[dddd\] [lcclll]{} 0531+21 & 3.10 & $<$ 6.19 & $ < 5.2\times 10^{46}$ & $<$ 35 & Becker & Aschenback (1995) 0833-45 & 4.05 & 6.20 & $ < 5.4\times 10^{46}$ & $<$ 36 & Ögelman, Finley & Zimmermann (1993) 0002+6246 & 4.50 & 6.26 & $ < 8.9\times 10^{46}$ & $<$ 46 & Hailey & Craig (1995) 0656+14 & 5.04 & 5.97 & $ < 7.7\times 10^{45}$ & $<$ 14 & Finley, Ögelman & Kizoloğlu (1992) 0630+178 & 5.48 & 5.80 & $< 1.8 \times10^{45}$ & $<$ 6.6 & Halpern & Ruderman (1993) 1055-52 & 5.73 & 5.91 & $< 3.6 \times 10^{45}$ & $<$ 9.3 & Ögelman & Finley (1993) \[cccc\]
{ "pile_set_name": "ArXiv" }
--- abstract: 'We perform one- and two-points magnitude cumulant analysis of one-dimensional longitudinal velocity profiles stemming from three different experimental set-ups and covering a broad range of Taylor scaled Reynolds numbers from $R_\lambda =$ 89 to 2500. While the first-order cumulant behavior is found to strongly depend on Reynolds number and experimental conditions, the second-order cumulant and the magnitude connected correlation functions are shown to display respectively universal scale and space-lag behavior. Despite the fact that the Extended Self-Similarity (ESS) hypothesis is not consistent with these findings, when extrapolating our results to the limit of infinite Reynolds number, one confirms the validity of the log-normal multifractal description of the intermittency phenomenon with a well defined intermittency parameter $C_2 = 0.025 \pm 0.003$. But the convergence to zero of the magnitude connected correlation functions casts doubt on the asymptotic existence of an underlying multiplicative cascading spatial structure.' address: 'Centre de Recherche Paul Pascal, Avenue Schweitzer, 33600 Pessac, France' author: - 'J. Delour, J.F. Muzy and A. Arneodo' title: 'Should we abandon cascade models to describe the spatial complexity of fully developed turbulence velocity profiles ?' --- Since Richardson’s original work [@bRic22], a common “mental image” of fully developed turbulence is a dynamical cascading process in which large eddies split up into smaller ones which themselves blow up into even smaller ones and so forth [@bFri95]. According to this picture, energy propagates from the integral scale, where eddies are generated, down to the dissipative scale, where they vanish by viscous dissipation, through a multiplicative process, each eddy inheriting a fraction of its parent’s energy. Since this early intuitive description, the notion of [*cascade*]{} has remained the creed of many models proposed in the literature [@bFri95] to mimic the statistical properties of turbulent signals. Among the available experimental data, a lot of effort has been devoted to the study of the longitudinal velocity component recorded in directional flows such as jets or wind tunnels [@bFri95; @bMon75; @Ans84; @Arn96]. In these configurations, the Taylor hypothesis that considers the spatial structure of the flow as globally advected upon the probe, enables us to interpret temporal time series as spatial profiles. In 1941, Kolmogorov [@Kol41] resumed Richardson’s picture in his statistical analysis of the spatial fluctuations of velocity profile in the sense that he linked the one-point statistics of the velocity increments $\delta v_l = v(x+l)-v(x)$ over different distances $l$, by some dimensional analysis which predicted the remarkable scaling behavior of the moments of $\delta v_l$ : $M_{q,l} = <\delta v_l^q> \sim l^{\zeta_q}$ where $\zeta_q = q/3$. Actually, $\zeta_q$ turned out to be a non-linear function of $q$ in most experiments [@bFri95; @bMon75; @Ans84; @Arn96] and many studies inspired from Kolmogorov and Obukhov second theory [@Kol62] tried to explain and to predict the analytical shape of this non-linearity. The controversial situation at the origin of some disagreement between models ([*e.g.*]{} the log-normal [@Kol62; @Cas90; @Arn98a] as opposed to log-Poisson models [@Dub94]) results from the experimental observation that the moments $M_{q,l}$ do not really scale perfectly. Indeed, there is a persistent curvature when one plots $\ln(M_{q,l})$ [*vs*]{} $\ln(l)$, which means that, rigorously speaking, there is no scale invariance. In order to give a sense to the exponents $\zeta_q$, Benzi [*et al.*]{} [@Ben95] defined the “Extended Self-Similarity” (ESS) hypothesis by proposing the following behavior for the moments of the velocity increments : $$\label{ESS} M_{q,l} \sim f(l)^{\zeta_q},$$ where $f(l)$ would be some $q$-independent function of $l$. Along this line, $\ln(M_{q_1,l})$ [*vs*]{} $\ln(M_{q_2,l})$ curves look definitely more linear than using standard log-scale representations and, by assuming that $\zeta_3 = 1$ [@bFri95], some experimental consensus has apparently been reached on the nonlinearity of the $\zeta_q$ spectrum [@Arn96]. In a recent theoretical work [@aArad99], Arad [*et al.*]{} suggest that at low Reynolds number, structure functions scaling properties are “poluted” by anisotropic effects that can be mastered using the irreductible representations of the rotation group. Unfortunately, this analysis is not tractable with single point data. Hopefully, as shown below, some statistical quantities turn out to display universal behavior that are likely to be insensitive to anisotropic effects. In the early nineties, Castaing [*et al.*]{} [@Cas90] recasted the cascade picture and the ESS hypothesis in a probabilistic description that accounts for the continuous deformation of the probability density functions (pdf) of $\delta v_l$ with $l$, by the mean of a propagator $G_{l l'}$ ($l'>l$): $$\label{Castaing1} P_l(\delta v) = \int_{-\infty}^{\infty} G_{l l'}(u)e^{-u}P_{l'}(e^{-u} \delta v ) du.$$ This equation can be related to a cascade process in which the variable $\delta v_l$ is continuously decomposed as $\delta v_l = \prod_{i=1}^{n} W_{l_{i+1},l_i} \delta v_{l'}$, where $\ln(|W_{l_{i+1},l_i}|)$ are independent random variables of law $G_{l_{i+1}l_i}$. Within this framework, the $\zeta_q$ and the $f(l)$ ESS functions can be related to the shape of $G_{ll'}$ thanks to the cumulant generating function of $G$ [@Cas95; @Arn98a], $$\label{cumexp} \ln \hat G_{ll'}(-iq) = \ln \left( M_{q,l}/M_{q,l'} \right) = \zeta_q \left(f(l)-f(l')\right),$$ the classical scale invariant case corresponding to $f(l) = \ln(l)$. In this paradigm, the successive terms of the polynomial development of $\ln \left( M_{q,l}/M_{q,l'} \right)$ as a function of $q$, involve the cumulants $C_q(l)= - C_q f(l)$ of the logarithm of $|\delta v_l|$ (“magnitude”) from which one can express the $\zeta_q$ spectrum as $\zeta_q = - \sum^{+\infty}_{k=1} C_k \frac{q^k}{k!}$. The question whether turbulent velocity signals do (or do not) present a space-scale cascade structure thus amounts to checking that the successive velocity magnitude cumulants possess the same scale behavior. This perspective can be extended from structure functions to unfused correlations functions [@Lvo96] (e.g., $<\delta v_l^q(x) \delta v_{l'}^p(x+\Delta x)>$), by studying the behavior of the magnitude connected correlation functions (MCCF) : $C_{ll'}(\Delta x) = \linebreak <\ln(|\delta v_l (x)|) \ln(|\delta v_{l'} (x+\Delta x)|)>_c$. These correlation functions have been introduced in Ref. [@Arn98c] and play the same role for the multivariate multi-point velocity law as the previous magnitude cumulants for the one-point law. For multiplicative cascade processes, they are expected to display a logarithmic dependence on the spatial distance $\Delta x$ [@Arn98c]. Our goal in this Letter is to push further the analysis carried out in Refs [@Arn98a; @Arn98c] by studying both the cumulants and the connected correlation functions associated to $\ln(|\delta v_l|)$ at different scales $l$, with the aim to establish a clear diagnostic about (i) the scaling properties of experimental turbulent velocity records and (ii) the validity of multiplicative cascade models in the limit of infinite Reynolds number. We will work with signals gracefully supplied by three different experimental groups. The signals labelled with Taylor scaled Reynolds number $R_\lambda =$ 89, 208, 463, 703 and 929 stem from Castaing’s group and were recorded in a gaseous helium jet at very low temperature in CRTB (Grenoble) [@thCha]; the signal labelled $R_\lambda =$ 570 stems from Baudet’s group and was recorded in an air jet in LEGI (Grenoble); the signal labelled $R_\lambda =$ 2500 was recorded at the ONERA wind tunnel in Modane by Gagne and collaborators [@thMal]. Let us start investigating one-point statistics via the computation of the three first cumulants $C_q(l)$ of $\ln(|\delta v_l|)$. In Fig. 1a, $C_1(l)$ is plotted [*vs*]{} $\ln(l)$ for all signals. They are all characterized by an integral scale $L$ corresponding to a decorrelation of the increments $|\delta v_l|$ for $l>L$. Below this saturated regime, there is no well defined “inertial range” since one observes a continuous cross-over across scales towards the smooth dissipative regime down the Kolmogorov scale $\eta$. Let us note that the integral scale does not depend on the Reynolds number but only on geometrical considerations. As an illustration, one gets the same value for $L$ for all the signals stemming from Castaing’s group which were recorded in the same apparatus, $R_\lambda$ being controlled either by the mean speed of the flow or by the value of viscosity that is changed by tuning temperature. Furthermore, the numerical values obtained for $L$ : $1.82$ cm for Castaing’s group, 0.91 m for Baudet’s group and 10.2 m for Gagne’s group, correspond to the characteristic sizes of the experimental set-ups. $L$ is thus a geometric parameter and when one increases $R_{\lambda}$, the inertial range increases via the decrease of $\eta$. Let us notice that as proposed in Ref. [@thCha], the shape of the cross-over from the saturated regime to the inertial one is rather well reproduced by a simple Langevin equation (inset of Fig. 1d) : $\delta v(x) = - \gamma v(x) + \sigma W(x)$, where $\gamma = 1/L$ and $W(x)$ is an appropriate fractional Brownian noise ($B_{H} = 1/3$). However the precise functional dependence of $C_1(l)$ is likely to depend on the experimental parameters and to be strongly influenced by anistropic effects [@aArad99]. The situation is very different for the second magnitude cumulant $C_2(l)$ when plotted [*vs*]{} $\ln(l)$ as shown in Fig. 1b. We first notice that the decorrelation scale where $C_2(l)$ saturates, is even better defined with $C_2(l)$ than with $C_1(l)$, since a linear behavior is observed up to this scale without any cross-over. The most important feature is that, whatever the set-up and whatever $R_{\lambda}$, all the data fall on [*linear*]{} curves which all have the [*same*]{} slope $- C_2 = - 0.025 \pm 0.003$. As a matter of fact, if we superimpose all the curves by setting all $L$ to 1, as shown in Fig.1e, we cannot distinguish anymore one curve from the others. This consideration is particularly relevant in the case of the signals of Castaing’s group where $R_\lambda$ is varied without changing $L$. The results concerning the third magnitude cumulant $C_3(l)$ are reported in Figs 1c and 1f. The slope of the curves obtained when plotting $C_3(l)$ [*vs*]{} $\ln(l)$ systematically decreases when increasing $R_\lambda$. For small $R_\lambda$, $C_3$ is significantly different from zero which means that the log-normal paradigm is not valid. In fact, from the curvature of the corresponding experimental $C_1(l)$ curves, we believe that this is rather a confirmation of the absence of a well defined inertial range for these low values of $R_\lambda$. For the largest values of $R_\lambda$ ($\geq 800$), $C_3$ becomes small enough (up to finite sample effects as previously reported in Ref. [@Arn98a]) to be neglected [@note]. At high Reynolds numbers, one can thus suppose that $C_3=0$ (and thus so the higher order cumulants) that implies a normal shape for the propagator $G_{ll'}$ in agreement with log-normal models [@Kol62; @Cas90; @Arn98a]. To summarize, the scale behavior of structure functions is well described by the first two terms in the cumulant expansion of $G_{ll'}$ and the differences observed in the behavior of $C_1(l)$ (Figs 1a and 1d) and $C_2(l)$ (Figs 1b and 1e) bring the experimental demonstration of the inconsistency of the ESS hypothesis which rigorously requires an identical scale behavior for the two cumulants (Eq. (\[cumexp\])). The full $\zeta_q$ spectrum cannot thus be defined in the range of Reynolds numbers we have investigated but the scaling exponent $C_2 = 0.025 \pm 0.003$ of $C_2(l)$ is likely to be a universal (model independent) characteristic of intermittency. Let us now focus on the MCCF : $C_{ll'}(\Delta x) = <\ln(|\delta v_l (x)|) \ln(|\delta v_{l'} (x+\Delta x)|)>_c$. As pointed out in Ref.[@Arn98c], we have checked that the MCCF computed at different scales $l$ and $l'$ all collapse on a single curve provided $\Delta x > \max(l,l')$. Let us recall that for a multiplicative cascade, we expect the MCCF to behave as $- C_2 \ln(\Delta x / L)$. In Fig.2 are shown the MCCF computed at a single small scale ($l=l'$) for each experimental signal. Three main observations must be raised from these curves. First, the MCCF do not behave linearly but as the square of the logarithm of the spatial distance $\Delta x$ : $C_{ll} (\Delta x) = \alpha(R_{\lambda}) \ln^2(\Delta x / L)$. This quadratic behavior is universal as far as experimental set-ups are concerned and has never been observed before. Second, the respective integral scales at which the magnitudes are decorrelated are nearly the same as the ones observed on the corresponding cumulants $C_q(l)$ ($q =$ 1,2), which is not a trivial result. Finally, as shown in Fig. 3, the prefactor $\alpha(R_{\lambda})$ has a systematic decreasing behavior as a function of $R_\lambda$. As far as the analytic shape of this decrease is concerned, there is no certainty. As illustrated in Fig. 3a, a power-law behavior : $\alpha \sim R_{\lambda}^{-a}$ with $a \simeq 0.20 \pm 0.03$ provides a reasonable good fit of the data. But as shown in Fig. 3b, one can also fit the data by $\alpha \sim K/\ln(R_{\lambda})$ with $K = 0.027 \pm 0.006$, which is a particular Reynolds number dependance that allows us to define a characteristic scale of the flow under study. Indeed, if one considers that the decorrelation scale is the same for the magnitude cumulants $C_q(l)$ and the MCCF, from the behavior of the variance as $-C_2 \ln(l / L)$ and of the MCCF as $\alpha \ln^2(\Delta x / L)$, one can define a characteristic scale $l_c$ where the two curves meet (when identifying scale and space-lag according to the multiplicative cascade picture [@Arn98c]) : $\ln(l_c/L) = -C_2/\alpha$. With the numerical values extracted from Figs 1b, 2 and 3b, one gets $(l_c/L) \sim R_\lambda^{-b}$ with $b = 0.91 \pm 0.3$. This new scale $l_c$ might well be the Taylor scale $\lambda$ for which $b = 1$. In that respect, our approach could be a way to measure objectively $\lambda$ and this even when the dissipative scale $\eta$ is not resolved. This universal quadratic behavior of the MCCF [*vs*]{} $\ln(\Delta x)$ and the systematic decrease of $\alpha(R_{\lambda})$ when increasing $R_{\lambda}$, are new observations which enables us to make some conjecture about the asymptotic limit of infinite Reynolds number. As pointed out in Refs [@Arn98a; @Arn98c], correlations in the magnitude are the signature of a multiplicative structure. If we extrapolate the observed behavior to infinite $R_{\lambda}$, we predict the convergence to zero of the MCCF on any finite range of scales. That is to say, in the limit of infinite $R_{\lambda}$, the intermittent spatial fluctuations of the longitudinal velocity component are not likely to display a multiplicative hierarchical structure. To summarize, we have advocated in this paper the study of scaling properties of longitudinal velocity increments by means of one- and two-points magnitude cumulant analysis. The results of our measurements for seven flows, at seven different $R_\lambda$ and stemming from different experimental set-ups, are two-fold. Concerning the one-point (fused) statistics, we mainly observe that the log-normal approximation is pertinent at sufficiently high Reynolds number and that the first two cumulants have not the same scale behavior as assumed by ESS hypothesis. While $C_1(l)$ displays some Reynolds and set-up dependent departure from scale invariance, $C_2(l)$ exhibits scale invariance behavior with a universal intermittency coefficient $C_2 = 0.025 \pm 0.003$. Concerning the two-points (unfused) statistics, we observe that the behavior of the MCCF [*vs*]{} $\ln(\Delta x)$ is quadratic whatever the considered experiment but with a prefactor $\alpha(R_\lambda)$ which decreases when increasing $R_\lambda$. These new observations lead us to conjecture that in the limit of infinite Reynolds number, the space-scale structure underlying the fluctuations of the longitudinal velocity is not of a multiplicative type but more likely in the spirit of the log-normal multifractal description pioneered by Kolmogorov and Obukhov in 1962 [@Kol62], i.e. with an intermittency coefficient $C_2 = 0.025 \pm 0.003$ but without any correlations across scales. This is not at all shocking from a physical point of view since the “mental image” of a dynamical cascading process [@bFri95] does not [*a priori*]{} imply that there should be some multiplicative (cascading) spatial organization. Furthermore, one must realize that our conclusions come out from the study of 1D cuts of the velocity field only. In a forthcoming publication, we plan to carry out a similar analysis of a 3D velocity field issued from direct numerical simulations, with the specific goal of testing the validity of this non multiplicative log-normal multifractal picture to account for the intermittent nature of fully developed turbulent 3D velocity fields. We are very grateful to B.Castaing’s, C.Baudet’s and Y.Gagne’s groups for the permission to use their experimental signals. [18]{} L.F. Richardson, [*Weather Prediction by Numerical Process*]{} (Cambridge University Press, Cambridge, 1922). U. Frisch, [*Turbulence*]{} (Cambridge University Press, Cambridge, 1995). A.S. Monin and A.M. Yaglom, [*Statistical Fluid Mechanics*]{} (MIT Press, Cambridge, MA, 1975), Vol 2. F. Anselmet [*et al.*]{}, J. Fluid Mech. [**140**]{}, 63 (1984). A. Arneodo [*et al.*]{}, Europhys. Lett. [**34**]{}, 411 (1996) and references therein. A.N. Kolmogorov, C.R. Acad. Sci. USSR [**30**]{}, 301 (1941). \(a) A.N. Kolmogorov, J. Fluid Mech. [**13**]{}, 82 (1962); (b) A.M. Obukhov, J. Fluid Mech. [**13**]{}, 77 (1962). B. Castaing, Y. Gagne and E.J. Hopfinger, Physica D [**46**]{}, 177 (1990); B. Castaing, Y. Gagne and M. Marchand, Physica D [**68**]{}, 387 (1993). A. Arneodo, S. Manneville and J.F. Muzy, Eur. Phys. J. B [**1**]{}, 129 (1998); A. Arneodo [*et al.*]{}, Phil. Trans. R. Soc. Lond. A [**357**]{}, 2415 (1999). B. Dubrulle, Phys. Rev. Lett. [**73**]{}, 959 (1994); Z.S. She and E.C. Waymire, Phys. Rev. Lett. [**74**]{}, 262 (1995). R. Benzi [*et al.*]{}, Physica D [**80**]{}, 385 (1995). I. Arad [*et al.*]{}, Phys. Rev. Lett. [**81**]{}, 5330 (1998); I. Arad, V. L’vov and I. Procaccia, Phys. Rev. E [**59**]{}, 6753 (1999). B. Castaing and B. Dubrulle, J. Phys. II France [**5**]{}, 895 (1995); F. Chillà, J. Peinke and B. Castaing, J. Phys. II France [**6**]{}, 455 (1996). V. L’vov and I. Procaccia, Phys. Rev. Lett. [**76**]{}, 2898 (1996); A.L. Fairhall [*et al.*]{}, Phys. Rev. Lett. [**79**]{}, 3174 (1997); R. Benzi, L. Biferale and F. Toschi, Phys. Rev. Lett. [**80**]{}, 3244 (1998). A. Arneodo [*et al.*]{}, Phys. Rev. Lett. [**80**]{}, 708 (1998). O. Chanal [*et al.*]{}, Eur. Phys. J. B [**17**]{}, 309 (2000). Y. Malécot [*et al.*]{}, Eur. Phys. J. B [**16**]{}, 549 (2000). Note that statistical convergence is faster in the estimation of $C_3$ when using the propagator as in Ref. [@Arn98a] than when computing directly $C_3(l)$ as reported in Fig. 1c.
{ "pile_set_name": "ArXiv" }
--- abstract: 'In order to effectively analyze information regarding ongoing events that impact local communities across language and country borders, researchers often need to perform multilingual data analysis. This analysis can be particularly challenging due to the rapidly evolving event-centric data and the language barrier. In this abstract we present preliminary results of a case study with the goal to better understand how researchers interact with multilingual event-centric information in the context of cross-cultural studies and which methods and features they use.' author: - 'Simon Gottschalk\*, Elena Demidova\*, Viola Bernacchi\*\*, Richard Rogers\*\*\*' title: 'Ongoing Events in Wikipedia: A Cross-lingual Case Study' --- Introduction {#sec:introduction} ============ In case of events concerning several communities, event-centric information available in different languages can reflect community-specific bias [@Rogers:2013]. As events unfold, it therefore becomes particularly important to analyze language-specific differences across the local event representations to better understand the community-specific perception of the events, the propagation of event-centric information across languages and communities as well as its impact. Wikipedia with its large multicultural user community and editions in over 290 languages is a rich source of multilingual information with respect to ongoing and past events. Information about ongoing events of global importance can evolve quickly in different language editions as the event unfolds [@Fetahu:2015], making Wikipedia an important and up-to-date source for cross-cultural studies. The potentially large amount of multilingual information to be analyzed, event dynamics and the language barrier limit the analysis methods, in particular with respect to the features accessible for the analysis, making cross-cultural studies of ongoing events particularly challenging. Although existing tools facilitate automatic cross-lingual article comparisons in Wikipedia with respect to selected features, including for example a cross-lingual text alignment for articles pairs by MultiWiki [@Gottschalk2017], a comparison of the topics covered by the articles by Omnipedia [@Bao:2012] and a computation of the article similarity in Manypedia [@Massa:2012], there is still a significant room for further developments in this area. In this abstract we present preliminary results of a case study concerning an analysis of cross-lingual event-centric information in Wikipedia exemplified through the Brexit referendum in June 2016. In this study, we observed an interdisciplinary, multicultural group of information professionals (communication design and sociology researchers) who performed the task of cross-cultural analysis of the Brexit representation in the multilingual Wikipedia in June 2016. In the study, six participants worked for 14 hours together as a team, resulting in a total of 84 person-hours spent. Objectives & Methodology {#sec:challenges} ======================== The key study objective was to *observe which methods and features can be efficiently used to perform cross-lingual analytics on multilingual Wikipedia*. As a first step, we introduced the participants to the topic of Brexit in Wikipedia and presented them with several example tools that can assist Wikipedia analytics in general. We asked the participants to: 1) define their own research questions and analysis methods related to Brexit and its international perception from the social and cultural perspectives; and 2) conduct an analysis of the Brexit-related articles across different Wikipedia language editions with respect to these questions. We focus the discussion on two Brexit-related articles selected by the participants: the “United Kingdom European Union membership referendum, 2016” (*referendum*) article and the “United Kingdom withdrawal from the European Union” (*withdrawal*) article. [images/timeline01c\_top.pdf]{} Observations {#sec:feature_selection} ============ Content Analysis ---------------- Due to the language barrier, the participants focused on the content features involving less text, e.g. the tables of contents and images. **Textual features**: Most Wikipedia articles contain a table of contents (TOC) that is arranged individually in each language edition. As the effort to read through whole Wikipedia articles – especially when the reader is not proficient in the particular language – was estimated too high, the participants decided to focus on the TOCs as an approximation of textual analysis. The participants arranged the TOC entries from the English, German, French and Italian article revisions by their frequency across languages. This TOC comparison has shown that the articles differed in many aspects: While the German article focused on the referendum results in the different regions, the English Wikipedia covered much more aspects, especially on economical and societal implications of the withdrawal. **Multimedia features**: Wikipedia articles can contain images that can be shared across different language editions. It became evident that images containing the map of the UK and the referendum ballot paper were shared most frequently across languages. Smaller deviations were observed for several languages, e.g. the German article contained many photos of the politicians. Temporal Analysis ----------------- The description of ongoing events may vary substantially over time. **Textual features:** The TOC of a Wikipedia article is subject of change over time. To observe this behavior, the participants extracted the TOC three times per day, in the time from the 22nd to the 24th of June 2016 (with the 23rd of June being the referendum date). The TOCs of two of these points in time made visible that the French version did not have the referendum results on the 23rd of June as opposed to other language versions under consideration. Within the following day, the English TOC stayed nearly unchanged in contrast to the other languages where especially the German article showed a major increase of new sections. The analysis on the withdrawal article revealed that the German version contained detailed historical and legislative entries while the French version focused on the referendum’s impact and the public opinion. **Edit-based features:** A valuable temporal aspect with respect to the cross-lingual information propagation comes from the editing process by the Wikipedia editors and the origin of the first revision per language. To explore this feature, the lists of all language editions for the referendum ($48$ language editions) and the withdrawal articles ($59$ language editions) have been analyzed using the Wikipedia Edits Scraper and IP Localizer tool. This way, the participants gained the edit histories of $107$ article language editions. They manually created two annotations for the initial revision of each article: “Creation date”, i.e. the date of the first edit, and “Origin”, i.e. the reference to other language editions that were utilized to create the respective edition. A visualization presented in Figure \[fig:timeline\] has been created by the participants and illustrates the English articles’ development including the number of edits and the creation of the other language editions. Article editions directly translated from other languages are marked and important events related to Brexit are added to the timeline. Edit peaks correlate with the sub-events in particular during the referendum held on the 23rd of June 2016. Network & Controversy Analysis ------------------------------ The interlinking between Wikipedia articles across languages can provide useful insights into the coherence of the language editions. **Category-based features:** In Wikipedia, editors can assign categories like “Politics” or “EU-skepticism in the UK” to the articles. To gain more insights into the perception of Brexit within the network of Wikipedia articles, the participants analyzed this categorization in all available language editions.The majority of the categories was related to basic concepts like “Referendum”. The Scottish and the English Wikipedia were rather disconnected from the network. In the withdrawal articles, “Eurocentrism” was identified as a main category connecting languages. **Links-based features:** Each Wikipedia article can contain links to external sources that the participants extracted and compared using the MultiWiki tool. For most of the language editions, the overlap of the linked web pages was rather low , and just in some cases it reached higher values: the English and German withdrawal articles shared $17.32\%$ of links, mainly related to news pages like the BBC or The Guardian. **Edit-based features:** To observe controversies, the participants reviewed the English discussion pages on the articles using their TOCs. A major finding facilitated by this feature was a discussion among the Wikipedia editors on the question to which article the search term “Brexit” should link to. Having initially redirected to the referendum article, the term lead to the withdrawal article a few days after the actual referendum. Conclusion {#sec:conclusion} ========== As our case study illustrates, certain types of cross-lingual analysis on Wikipedia mostly relying on non-textual features (e.g. links, images and category structures) can be performed efficiently, manually or by using existing tools, and can already provide interesting insights. In contrast, analysis of textual content and discussion pages appears to be limited through the language barrier.
{ "pile_set_name": "ArXiv" }
--- abstract: 'In the context of high-frequency quasi-periodic oscillation (HF QPOs) we further explore the appearance of an observable signal generated by hot spots moving along quasi-elliptic trajectories close to the innermost stable circular orbit in the Schwarzschild spacetime. The aim of our investigation is to reveal whether observable characteristics of the Fourier power-spectral density can help us to distinguish between the two competing models, namely, the idea of bright spots orbiting on the surface of an accretion torus versus the scenario of intrinsic oscillations of the torus itself. We take the capabilities of the present observatories (represented by the Rossi X-ray Timing Explorer, RXTE) into account, and we also consider the proposed future instruments (represented here by the Large Observatory for X-ray Timing, LOFT).' author: - 'V. Karas, P. Bakala, G. Török, M. Dovčiak, M. Wildner, D. Wzientek, E. Šrámková, M. Abramowicz, K. Goluchová, G. P. Mazur, F. H. Vincent' title: 'Distinguishing between spot and torus models of high-frequency quasiperiodic oscillations' --- [2]{} Introduction ============ Low-mass X-ray binaries (LMXBs) represent a particular example of astronomical sources in which puzzling quasi-periodic modulation of the observed signal can develop and reach very high (kilohertz) frequencies [@Kli:2006:CompStelX-Ray:]. Orbiting inhomogeneities, a.k.a. spots residing on the surface of an accretion disk, have been introduced as a tentative explanation for persistant high-frequency oscillations (HF QPOs) observed in X-rays from accreting black holes and neutron stars in LMXBs. These QPOs have been discussed, in particular, in a series of papers [@kar:1999; @mor-ste:1999; @ste-vie:1998a; @ste-vie:1998b; @ste-vie:1999; @ste-vie:2002] that explain QPOs as a direct manifestation of modes of relativistic epicyclic motion of blobs at various radii $r$ in the inner parts of the accretion disc. Within the model, the two observed HF QPO peaks arise due to the Keplerian and periastron precession of the relativistic orbits. In a similar manner, the authors of refs. [@cad-etal:2008; @ger-etal:2009; @kos-etal:2009] introduced a concept according to which the QPOs are generated by a tidal disruption of large accreting inhomogeneities. Recently, [@ger:2012] argued that the behaviour of the azimuth phase $\phi(t)$ for non-closed quasi-elliptic orbits in the curved spacetime can be responsible for the observed pairs of HF QPOs. In this contribution we investigate the behaviour of the observable signal produced by radiating small circular hotspots. We discuss the detectability of the produced signal propagated from the strongly curved spacetime region, and in particular we concentrate of the properties of Fourier power-spectral density (PDS), as predicted in different scenarios. It has been known that PDS reflects the physical properties of the source, however, it does not provide a complete description. We thus explore a possibility to discriminate via PDS between the two specific models that are thought to provide a promising scheme for HF QPOs in black hole sources. In fact, according to the Wiener-Khinchin theorem, the power spectral function $S(\omega)$ can be calculated from the autocovariance function $C(t)$ as [@bendat:2000] $$S(\omega)={\cal F}[C](\omega).$$ It can be seen that only the Gaussian processes are completely determined by their power spectra. Nonetheless, it is interesting to investigate the predicted profiles of the power spectrum for well-defined processes and to reveal specific features, their differences and similarities. In our discussion we consider the capabilities of the present X-ray observatories represented by Rossi X-ray Timing Explorer (RXTE), as well as the proposed future instruments represented by the Large Observatory for X-ray Timing (LOFT). We also compare the signal produced by spots to the signal obtained from another specific kind of simulations assuming axisymmetric epicyclic disc-oscillation modes. Our present paper is based on a recent work [@bakala:2013] where further details can be found. In particular, in the cited paper we explore the width of the QPO peaks with 3:2 frequency ratio that has been reported to occur in a number of sources. Set-up of the model =================== Qualtitative mathematical properties of Fourier power spectra, as predicted by the hot-spot scenario, have been studied in our refs. [@pec:2013; @pec:2008]. Next, in ref. [@kar:1999] a quantitative study of the expected PDS was presented and the orbital motion of the spots and their distribution within a narrow range of radii was discussed. The adopted phenomenological description of the source employs spots or clumps orbiting in the Schwarzschild metric as an approximation to more realistic models of inhomogeneous disc–type accretion flows around black holes and neutron stars. In order to reproduce the quality factor of the observed QPOs, it was found that spots must be distributed in a zone of only several gravitational radii from the central black hole, with their observed luminosity influenced by Doppler and lensing effects. Furthermore, we found that the expected values of the quality factor of the oscillations can reach about several hundred (typically, $Q\approx3\times10^2$). However, the assumption about strictly circular orbits was clearly a simplification which calls for further discussion in the present paper. Furthermore, the effect of background noise has to be taken properly into account [@Wit:2012]. The left panel of Figure \[figure:3to2models\] illustrates the spot scenario which we investigate. The radial distribution or drifting of the spots can clearly result in various levels of signal coherence. Nevertheless, small circular spots related to a single preferred radius cannot reproduce the often observed 3:2 frequency ratio. We also consider a more elaborate scheme where the multiple spots are created and drifted around radii close to two preferred orbits with Keplerian frequencies roughly in the 3:2 ratio. These orbits are set as $r_{2}=8M$ (radius where the radial epicyclic frequency reaches the maximum value) and $r_{3}=6M$ (ISCO). Spots are then created within the regions $[r_{i}-\delta{r},\,r_{i}]$ with the size given by $\delta{r}=0.75M$. Since we assume the black hole mass $M\doteq11M_\odot$, our setup leads to the main observable frequencies around 110Hz and 160Hz (see Figure \[figure:PDSfit\] drawn for the signal fraction $n=10\%$ and $\mathcal{D}=65^{\circ}$). In the following consideration we compare the PDS obtained for this setup to the PDS resulting from the model of the oscillating optically thin torus slowly drifting through the resonant radius $r_{3:2}$. To define the torus kinematics, we assume the $m=0$ radial and vertical oscillations with equal intrinsic amplitudes. The possible QPO origin in the resonances between this or similar disc oscillation modes has been extensively discussed in works of [@abr-klu:2001; @abr-etal:2003a; @abr-etal:2003b; @hor:2005; @klu-etal:2004; @tor-etal:2005], and several other authors. Here we adopt the concept previously investigated by [@bur-etal:2004] who focused on optically thin torus with slender geometry. The visual appearance of torus influenced by lensing and Doppler effects is illustrated in the right panel of Figure \[figure:3to2models\]. Within the adopted concept the periodic changes of the observed luminosity are partially governed by the radial oscillations due to changes of the torus volume while the vertical oscillations modulate the flux just due to lensing effects in the strong gravitational field. The contribution of the two individual oscillations to the variations of the observed flux thus strongly depends on the inclination angle [see also ref. @maz-etal:2013]. Here we set $\mathcal{D}=65^{\circ}$ where the fractions of the power in the two observed peaks are comparable. We set the black hole mass $M=5.65M_{\odot}$ and $a=0$ $(r_{3:2}=10.8M)$, implying that the two oscillatory frequencies are $\nu_\theta(r_{3:2})=160$Hz and $\nu_\mathrm{r}(r_{3:2})\doteq110$Hz. Assuming this setup we produce torus drift lightcurves for the interval $r/r_{3:2}\in[0.97,1.03]$. The resulting PDS drawn for the signal fraction $n=10\%$ is included in Figure \[figure:PDSfit\]. We note that similar PDS can be reached assuming e.g. a near extreme rotating black hole with $a=0.98$ and $M\doteq18M_{\odot}$. Using Figure \[figure:PDSfit\] we can finally confront the predictions obtained for spots drifted around preferred radii to those expected for the oscillating torus slowly passing the resonant orbit $r_{3:2}$. Inspecting this Figure, we can find that the RXTE PDS obtained for the given setup of the two models are rather similar. On the other hand, the LOFT PDS clearly reveal presence/absence of the harmonics additional to the 3:2 peaks representing the signature of spot motion. ![image](models_csX.eps){width="\hsize"} ![image](pds_fit4r.eps){width="\hsize"} We assume the global source flux described by the approximations of the spectral distribution $N(E)$ and power density spectra $P(\nu)$, $$\begin{aligned} \label{equation:global:spectrum} &&N(E)= kE^{-2.5},\\ \label{equation:global:variability} &&P(\nu)= p_0\nu^{p_1}+ \frac{1}{\pi}\frac{p_3p_4}{(\nu-p_2)^2+p_3^2}, $$ where $k$ is chosen to normalize the assumed countrate roughly to $1$Crab and $p_{\mathrm{i}}=[0.001,-1.3,2.5,0.8,0.002]$. This setup roughly corresponds to the so-called high steep power law (HSPL) state in GRS 1915+105. Expected properties of PDS ========================== ![image](single_spot_dec80_E1comp.eps){width="\hsize"} ![image](e_xpected80.eps){width="\hsize"} ![image](e_xpected30.eps){width="\hsize"} ![image](single_spot_dec30_comp_r.eps){width="\hsize"} Figure \[figure:PDSD80\] shows PDS resulting from the RXTE and LOFT simulations assuming various levels of $n$ and the inclination $\mathcal{D}=80^{\circ}$ (i.e., nearly equatorial view). Since the signal from the spot strongly depends on the source inclination $D$ [e.g. ref. @sch-ber:2004a], we will compare the results for two representative values of $D$ corresponding to the nearly equatorial view $D=80^{\circ}$ and the view close to vertical axis $D=30^{\circ}$. In the following, we assume a spot orbiting at $r=6.75M$ with constant angular velocity of the Keplerian value $\Omega_\mathrm{K}$. The spot trajectory slightly deviates from the circular shape due to the radial epicyclic oscillation, having small amplitude $\mathcal{E}>0$. Nearly equatorial (edge-on) view -------------------------------- In Figure \[figure:D30\] (upper panels) we include amplitude spectra and time dependent energy spectra of the net spot signal calculated for the distant observer at $D=80^{\circ}$. The spot signal is dominated by the Keplerian frequency and its harmonics amplified by relativistic effects which is well illustrated by the amplitude spectrum on the upper left panel of Figure \[figure:D30\]. The eccentricity corresponding to the amplitude of the radial epicyclic oscillation $\mathcal{E}=0.1M$ causes only negligible modulation at the radial and precession frequencies. The increased eccentricity corresponding to the amplitude of the radial epicyclic oscillation $\mathcal{E}=1M$ can be well recognized in the amplitude spectra but the signal is still dominated by the Keplerian frequency and its harmonics. The time dependent energy spectra of the spot are depicted in the upper right panel of Figure \[figure:D30\]. We can see that they clearly reveal the signatures of relativistic redshift effects. So far we have reproduced just the variability and spectra of the net spot flux. In order to assess the observable effects we have to study the total composition of the net spot flux together with the global source flux given by Equations (\[equation:global:spectrum\]) and (\[equation:global:variability\]). Assuming this composed radiation, we can consider the capabilities of the RXTE and LOFT instruments using their response matrices and provided software tools. The time-dependent spectra describing the composed radiation are then convolved with the appropriate response matrix giving an estimate of the observed data. These are Fourier transformed to the resulting power spectra. Within such consideration the detectability of the spot signatures depends obviously on the fraction of photons from the spot in the total flux. We referred to this [*signal to noise ratio*]{} shortly as the signal fraction $n$. Figure \[figure:PDSD80\] includes PDS resulting from the RXTE and LOFT simulations assuming various levels of $n$ and the inclination $\mathcal{D}=80^{\circ}$. It includes the cases when the signal is weak for the RXTE and there are no significant features within its PDS as well as the high signal fraction when first two harmonics of the Keplerian frequency can be seen. Comparing the both panels of this Figure we can deduce that when the weak QPO signal corresponding to the hot-spot Keplerian frequency is around the limits of the RXTE detectability, the LOFT observations can clearly reveal its first and second harmonics. We checked that there is, in practice, no qualitative difference between the cases of $\mathcal{E}=0.1M$ and $\mathcal{E}=1M$. It is therefore unlikely that the periastron precession or radial epicyclic frequency can be detected in addition to the harmonics when the inclination angle is close to the equatorial plane. View close to vertical axis --------------------------- For $D=30^{\circ}$ (lower panels), the signal is dominated by the Keplerian frequency but the harmonics are much less amplified in comparison to the nearly equatorial view (see the bottom left panel of Figure \[figure:D30\]). Eccentricity corresponding to the amplitude of radial epicyclic oscillation $\mathcal{E}=0.1M$ causes again rather negligible modulation at the radial and precession frequencies. Nevertheless, we can see that the increased eccentricity of $\mathcal{E}=1M$ affects the variability more than for the large inclination angle. Furthermore, its influence is comparable to those of second harmonics of the Keplerian frequency. The time dependent energy spectra are again depicted in the lower right panel of Figure \[figure:D30\]. Finally, figure \[figure:PDSD30\] shows PDS resulting from RXTE and LOFT simulations assuming various levels of $n$. It is drawn for $\mathcal{E}=1M$ and includes the few cases when the signal is weak for the RXTE and there are no significant features within its PDS, plus one case when some feature at the Keplerian frequency can be seen. Comparing the both panels of this Figure we can deduce that when the weak QPO signal corresponding to the hot-spot Keplerian frequency is around the limits of the RXTE detectability, the LOFT observations can clearly reveal its first and second harmonics but also the radial epicyclic frequency $\nur$. Although it is not directly shown, we check that decreasing of eccentricity to $\mathcal{E}=0.1M$ leads to similar PDS but with the missing feature at $\nur$. Conclusions =========== We can identify the signatures of the spot motion mostly with the harmonic content of the observable signal. For large inclination angles, the LOFT observations could easily reveal the Keplerian frequency of the spot together with its first and second harmonics when the strongest (but weak) single signal is around the limits of the RXTE detectability. Nevertheless, radial epicyclic frequency could be also found providing that the inclination is small. In our analysis we have paid attention to the timing signatures of the motion of small circular spots radiating isotropically from the slightly eccentric geodesic orbits. The case of highly eccentric orbits and/or spot having large azimuthal shear will be presented elsewhere. We studied the comparison between spot and torus scenarios serving as fiducial representations of two very specific kinematic models. Obviously, any general validity of this discussion is limited. For instance, a consideration of resonance driven effects or the role of torus geometrical thickness could give rise to some harmonic content in the signal from the oscillating tori. Despite uncertainties, the elaborated comparison indicates clearly that the increased sensitivity of the proposed LOFT mission can be crucial for resolving the QPO nature. We refer the reader to ref. [@bakala:2013] for additional details. We acknowledge the project CZ.1.07/2.3.00/20.0071 Synergy supporting the international collaboration of Institute of Physics in Opava, the Czech Science Foundation project GAČR 13-00070J in Prague (VK), GAČR 209/12/P740 (GT, EŠ), and the European Union 7th Framework Programme No. 312789 ‘StrongGravity’ (MD). We also acknowledge the Polish grant NCN UMO-2011/01/B/ST9/05439 and the Swedish VR grant (MAA). Astronomical Institute is supported by the research program RVO:67985815. [99]{} Abramowicz, M. A., & Klu[ź]{}niak, W. 2001, A&A, 374, L19 Abramowicz, M. A., Bulik, T., Bursa, M., & Klu[ź]{}niak, W. 2003a, A&A, 404, L21 Abramowicz, M. A., Karas, V., Klu[ź]{}niak, W., Lee, W. H., & Rebusco, P. 2003b, PASJ, 55, 466 Bakala, P., Török, G., Karas,  V. et al. 2014, MNRAS, accepted (arXiv:1401.4468) Bendat, J. S., & Piersol, A. G. 2000, Random Data: Analysis and Measurement Procedures (New York: Wiley) Bursa, M., Abramowicz, M. A., Karas, V., & Kluzniak, W. 2004, ApJ, 617, L45 Čadež, A., Calvani, M., & Kostić, U. 2008, A&A, 487, 527 Germana, C., Kostič, U., Čadež, A., & Calvani, M. 2009, in AIP Conf. Proc. 1126, SIMBOL-X: Focusing on the Hard X-ray Universe: Proc. 2nd International Simbol-X Symposium, eds. J. Rodriguez & P. Ferrando (New York: Melville), 367 Germana, C. 2013, MNRAS, 430, L1 Horák, J. 2005, Astronomische Nachrichten, 326, 824 Karas V. 1999a, ApJ, 526, 953 Karas V. 1999b, PASJ, 51, 317 Kluźniak, W., Abramowicz, M. A., Kato, S., Lee, W. H., & Stergioulas, N. 2004, ApJ, 603, L89 Kostić, U., Čadež, A., Calvani, M., & Gomboc, A. 2009, A&A, 496, 307 Mazur, G., P., Vincent, F., H., Johansson, M., Šrámková, E., Török, G., Bakala, P., Abramowicz, M. A. 2013, Astronomy & Astrophysics, Volume 554, id.A57 (arXiv:1303.3834) Morsink, S. M., & Stella, L. 1999, ApJ, 513, 827 Pecháček, T., Goosmann, R. W., Karas, V., Czerny, B., & Dovčiak, M. 2013, A&A, 556, id. A77 Pecháček, T., Karas, V., & Czerny, B. 2008, A&A, 487, 815 Schnittman, J. D., Bertschinger, E. 2004a, ApJ, 606, 1098 Stella, L., & Vietri, M. 1998a, in Abstracts of the 19th Texas Symposium on Relativistic Astrophysics and Cosmology, eds. J. Paul, T. Montmerle, & E. Aubourg (Saclay, France: CEA) Stella, L., & Vietri, M. 1998b, ApJ, 492, L59 Stella, L., & Vietri, M. 1999, Physical Review Letters, 82, 17 Stella, L., & Vietri, M. 2002, in The Ninth Marcel Grossmann Meeting, Proc. MGIXMM Meeting, eds. V. G. Gurzadyan, R. T. Jantzen, & R. Ruffini, Part A, 426 Török, G., Abramowicz, M. A., Klu[ź]{}niak, W., & Stuchl[í]{}k, Z. 2005, A&A, 436, 1 van der Klis, M. 2006, in Compact Stellar X-Ray Sources, ed. W. H. G. Lewin & M. van der Klis (Cambridge: Cambridge Univ. Press), p. 39 Witzel, G., Eckart, A., Bremer, M., Zamaninasab, M., Shahzamanian, B., Valencia-S., M., Schödel, R., Karas, V., Lenzen, R., Marchili, N., Sabha, N., García-Marín, M., Buchholz, R. M., Kunneriath, D., & Straubmeier, C. 2012, ApJ, 203, id. 18
{ "pile_set_name": "ArXiv" }
--- author: - '[A. L[ó]{}pez Ariste]{}, [P. Mathias]{}, [B. Tessore]{}, [A. Lèbre]{}, [M. Aurière]{}, [P. Petit]{}, [N. Ikhenache]{}, [E. Josselin]{}, [J. Morin]{}, [M. Montargès]{}.' date: 'Received ...; accepted ...' title: 'Convective cells in Betelgeuse: imaging through spectropolarimetry[^1]' --- [We assess the ability to image the photosphere of red supergiants and, in particular Betelgeuse, through the modelling of the observed linear polarization in atomic spectral lines. We also aim to analyse the resulting images over time, to measure the size and dynamics of the convective structures in these stars.]{} [Rayleigh scattering polarizes the continuum and spectral lines depolarize it. This depolarization is seen as a linear polarization signal parallel to the radial direction on the stellar disk. Integrated over the disk, it would result in a null signal, except if brightness asymmetries/inhomogeneities are present. This is the basic concept behind our imaging technique. Through several tests and comparisons, we have tried to assess and extend its validity, and to determine what can be learnt unambiguously through it.]{} [The several tests and comparisons performed prove that our technique reliably retrieves the salient brightness structures in the photosphere of Betelgeuse, and should be relevant to other red supergiants. For Betelgeuse, we demonstrate that these structures we infer are convective cells, with a characteristic size of more than 60% of the stellar radius. We also derive the characteristic upflow and downflow speeds, 22 and 10km.s$^{-1}$ respectively. We find weak magnetic fields concentrated in the downflow lanes in between granules, similarly to the quiet sun magnetism. We follow those convective structures in time. Changes happen on timescales of one week, but individual structures can be tracked over four years of observations.]{} [The measured characteristics of the convection in Betelgeuse confirm the predictions of numerical simulations in both the strong, supersonic upflows as well as the size of the convective cells. They also concur in the presence of weak magnetic fields that are completely dominated by the convective flows and constrained to the dark lanes of down-flowing plasma.]{} Introduction ============ The chemical enrichment of the interstellar medium is mostly due to the late stages of stellar evolution; this is particularly true for red supergiants (RSG) because of their mass loss. However, the mechanisms that drive mass loss from these stars are not well understood. Mechanisms that are often invoked include thermal gas and radiation pressure, acoustic and shock waves, Alfvén waves, magnetism, and most probably other additional phenomena. The strong wind of these stars is probably triggered by dust, formed at a great distance from the photosphere, and so has no direct connection with the atmosphere dynamics. It is thus essential to characterize the phenomena that take place close to the stellar surface. [@Josselin2007] suggested that high velocities and steep velocity gradients, possibly caused by convective motion, generate line asymmetries, that turbulent pressure decreases the effective gravity, and that this decrease combined with radiative pressure on lines initiates the mass loss. Convection seems indeed to be a key component to understand the evolution of massive, cool evolved stars. [@Schwarzschild1975] suggested that the outer envelope of RSG could host a small number of large convective cells. These gigantic convection cells have also been predicted in simulations [e.g., @Freytag2002; @Chiavassa2011] and suggested in observations of Betelgeuse as bright photospheric spots observed with direct UV imaging [@Gilliland1996] or with interferometric techniques [e.g., @Haubois2009]. Indeed, because of its relative proximity (about 200pc), and as an MIab supergiant, Betelgeuse offers the largest angular diameter of any star (except for the Sun), and has been extensively studied in interferometry [e.g., @Haubois2009; @Montarges2016]. Because interferometry may suffer from modelling hypothesis (limb-darkening, spot(s) shape...) and cannot resolve velocity fields, it is interesting to develop alternative approaches that complement and reinforce these results. Since 2013 the instrument Narval@TBL has been monitoring linear polarization on Betelgeuse with roughly one observation per month during the visibility period of the year, leading up to 43 observations in total through April 2018. [@Auriere2016] first interpreted linear polarization in terms of Rayleigh scattering in the continuum and depolarization during the formation of spectral lines. Those authors also proposed a first modelling of linear polarization through bright spots that would result in the net linear polarization we observe. A tentative confirmation of such bright spots, in addition to the linear polarization interpretation, was seen in the images inferred from interferometry with the PIONEER instrument [@Montarges2016]. Beyond that first success, it was already obvious that the profiles could not be interpreted as simply being due to one or two bright spots on top of an otherwise homogeneous photosphere. Further insight came from [@Mathias2018] who added to the linear polarization the analysis of the circular polarization signatures, due to the Zeeman effect, monitored over more than eight years. Two results were crucial for two reasons. First, the signals presented a rough periodicity, as seen in the Fourier power maps, of about 1800 days, a timescale usually associated with the convective one for a red supergiant as Betelgeuse [@Stothers2010]. Second, the linear polarization signals were predominantly found blue-shifted with respect to the intensity signals, while Zeeman-induced circular polarization was red-shifted. Both results pointed towards a conclusion that is the subject of the present paper: the bright spots identified through linear polarization were the convective cells of Betelgeuse seen at the photospheric level whereas the tiny magnetic fields were dragged around by the convective flows and were thus concentrated in the intergranular dark lanes where the plasma sinks into the stellar interior. The path towards imaging the photosphere of Betelgeuse is based upon the interpretation that we are seeing convective cells. This allows us to propose models of the brightness of the photosphere of Betelgeuse in terms of spherical harmonics and to link the Doppler shift of the local spectral profile to this brightness. This is introduced in quite some detail in Sect. 2. The model parameters are the coefficients of the decomposition of the brightness in terms of spherical harmonics. We present in Sect. 3 a simple inversion algorithm based upon the Marquardt-Levemberg iteration that allows us to fit the observed linear polarization profiles in terms of an image of the photosphere of Betelgeuse. In Sect.4, we present the data that allowed us to conclude on the convective model and which provide some preliminary estimates of the velocities involved in the convection of Betelgeuse. The size of the convection cells is analysed in Sect.5, and a characteristic length scale of $0.62\,R_*$ is found for these structures. Finally, in Sect.6, we discuss the four years of convective dynamics of Betelgeuse. Most of the data presented here have already been described by [@Auriere2016] and [@Mathias2018] but, since, the monitoring observations of Betelgeuse continues, with new observations corresponding to 2017-2018. Origin and modelisation of linear polarization profiles ======================================================= Observational facts and previous basic model -------------------------------------------- At high polarization sensitivity, the atomic photospheric lines in the spectra of Betelgeuse show linear polarization signals. [@Auriere2016] were the first to show these signals over three years of observations of Betelgeuse. In this paper we present the continuation of this series of measurements up to April 2018. Since then linear polarization signatures have been detected in the spectra of other red supergiants [@Tessore2017]. The amplitude of those signals is smaller than 0.1% of the continuum intensity and, characteristically, several peaks with different signs are seen across and sometimes beyond the width of the intensity profile. [@Auriere2016] produced the basic explanation of the origin of that signal that we now summarize. The photospheric continuum in Betelgeuse is polarized by Rayleigh scattering. The local amplitude of this polarization can be estimated using a MARCS atmospheric model for Betelgeuse with $T_{eff}=3750K$, $\log g=0.5$, $M=15 M_{sun}$, $v_t=5km s^{-1}$ and solar composition [@Gustafsson2008]. The wavelength and position dependence of this Rayleigh polarization is illustrated in Fig.\[J20\]. With respect to the integrated intensity, the maximum contribution comes from regions at $\mu=\cos \theta=0$ (with $\mu=1$, $\theta =0$ corresponding to disk centre and $\mu =0$, $\theta=\pi/2$ to the photospheric limb, where the line of sight is perpendicular to the stellar radius) and, at this distance from disk centre, increases from about 1% in the red to up to 8% in the blue. These numbers update and correct previous estimates , mostly because of the improved atmospheric model. The polarization amplitudes so computed are about 100 times larger than the corresponding ones measured in the limb of the Sun [@Leroy1972; @Stenflo2005; @Fluri1999] at similar position and wavelength. This polarized continuum is then absorbed by atoms in the regions above the continuum forming region. The de-excitation of these atoms leads to a decrease of the local polarization since the photon, once absorbed, is simply re-emitted unpolarized. Thus, atomic spectral lines depolarize the continuum Rayleigh polarization when forming in the atmosphere of Betelgeuse. ![Expected linear polarization rates in the continuum due to Rayleigh scattering in a model atmosphere adequate for Betelgeuse ($T_{eff}=3750K$, $\log g=0.5$, $M=15 M_{sun}$, $v_t=5 km s^{-1}$ and solar composition). Top: Polarization rates at the fixed wavelentgh of 600nm as a function of the angular distance $\mu =\cos \theta$ to disk centre. The envelope, beyond the stellar radius, is artificially shown with negative values of $\mu = \cos \theta$. The thin line gives the actual polarization rate respect to the local intensity. The thick line gives the rate normalized to the intensity at the disk centre, providing a measure of the relative contribution to polarization the integrated polarization signal. Bottom: Dependence of the polarization rate with wavelength at a fixed distance to disk centre $\mu =\cos \theta=0.$.[]{data-label="J20"}](J20_figure.png){width="50.00000%"} On top of this mechanism that is common to all spectral lines, specific lines may be sensitive to atomic alignment induced by the anisotropy of the radiation field. For those specific lines, the re-emitted photons are polarized. The details of these two competing mechanisms can be found in [@LandiDeglInnocenti2004]. However, [@Auriere2016] demonstrated that the spectra of Betelgeuse were compatible with a scenario in which, unlike the solar case, the intrinsic line polarization was negligible, all lines just depolarizing the continuum. The observation of the linear polarization in the $D_1$ and $D_2$ lines was the key to this conclusion. While both lines are equally able to depolarize the continuum (an ability that mostly depends on the depth of the line), the $D_2$ line produces 10 to 100 times more intrinsic polarization than the $D_1$ line. Unlike the solar spectrum, where the $D_2$ line is 10 times more polarized than $D_1$, in Betelgeuse both lines showed a very similar signal both in amplitude and shape, an observation only compatible with the depolarization process. This conclusion could be easily extended to most other lines, with smaller intrinsic polarization levels thus leading to the conclusion that all lines were depolarizing the continuum, unlike the solar case. The reasons behind the differences between this scenario in Betelgeuse and the solar case are left for future work. In these conditions, were we able to observe a local point at the limb of Betelgeuse, we would see a polarized continuum along the direction tangent to the local limb. Spectral lines that depolarize this continuum would decrease this polarization signal at their respective wavelengths. This local polarization spectrum would look very much like the intensity spectra. However, in an unresolved star, the observed global polarized spectrum is integrated over the whole stellar disk. Since the linear polarization direction rotates over $2\pi$, linear polarization signals should cancel out, except if brightness or polarization inhomogeneities are present over the stellar disk. This is the basic idea behind the techniques of imaging through spectropolarimetry. The ratio of Stokes $Q$ to Stokes $U$ in the continuum could eventually provide the polar angle at which the maximum brightness is found. But this is hardly sufficient for imaging and, furthermore, our instruments do not measure the polarization in the continuum (while relative polarization, e.g. the polarization in the line respect to the continuum, is measured down to sensitivities of 1 in $10^{5}$, absolute polarization, as measuring the continuum polarization requires, is limited by photometric and hardly goes beyond sensitivities of 1 in 1000). The situation is improved by the analysis of the polarization signals within the spectral lines. The observed signals inside atomic lines are found at different Doppler shifts. Assuming a fixed radial velocity, the Doppler shift of each signal provides a distance to the disk centre. Simultaneously, and as for the continuum, the ratio of $Q$ to $U$ Stokes parameters fixes the polar angle, but this time we obtain one angle per signal. Thus, the combination of Doppler shift, due to a hypothetically constant radial velocity, with the ratio $Q/U$, due to the depolarization of the continuum with polarization tangent to the local limb, leads to a two dimensional mapping of brightness inhomogeneities over the stellar disk. As the polarization of all lines is due to continuum depolarization, the spectral shape of this polarization is that of the spectral line and proportional to the depth of the line. Under these circumstances, we can add up several thousand lines over the spectra after scaling wavelength to a velocity respect to line centre for each line and weighting the signals by the depth of the intensity line. This is a basic description of LSD [@Donati1997] which will effectively increase the signal-to-noise ratio of the resulting pseudo-line. Applying the previous basic mapping schemeto the LSD pseudo-line, [@Auriere2016] computed for each wavelength position across the spectroscopic line the polar angle at which brightness was maximum and plotted that as a spot. The resulting maps consistently show two to three bright spots. At two particular dates, Jan 9th, 2014 and Dec 9th, 2015, interferometric observations were available that concurred on the presence of hot spots at at least one of the selected positions and up to a 180 degrees ambiguity in the azimuth angle of the spots, inherent to the polarization measurements [@Montarges2016; @Auriere2016]. Despite its crudeness, this basic method was able to pinpoint some of the brightest spots over Betelgeuse. Among the many simplifications onto which a simple first such model is based, one drives our present improvement in modelling the linear polarization of Betelgeuse. At any distance of the disk centre, determined by the Doppler shift of the polarization signal, the basic method assigns the polar angle to a unique source of brightness, independently of other sources. However, it is obvious that the combination of several spots may result in a signal peak at the correct wavelength and with the right $Q/U$ ratio. But how could one find the information to localize two spots that combine their signals into a single maximum of linear polarization? While at first sight it appears that such improvement is beyond hope, further thought reveals that such information exists in the data: if the bright spot was a single, isolated one, both $Q$ and $U$ would produce a maximum at the same wavelength. But this is not always the case, as can be seen by inspecting in detail the profiles shown in [@Auriere2016], [@Mathias2018]. The maxima of $Q$ and $U$ often occur at different wavelengths. This indicates the presence of ad minima asymmetrical spots and, in general, of multiple spots. [@Mathias2018] adds a second, indirect argument to that described above. In that work, [@Mathias2018] present simultaneous circular and linear measurements. The Stokes $V$ circular polarization is shown to be due to the Zeeman effect and attributed to photospheric magnetic fields with net longitudinal fluxes of 1G or less. In addition, most of the Zeeman signatures are redshifted, while the linear polarization signatures are predominantly blueshifted (see also Sect.5). The basic, simplest interpretation of these observations is the same that explains the similar behaviour of the very weakest fields found in the solar photosphere. These weak fields are unable to impact the plasma dynamics and are therefore dragged by the main convection flows over the surface. Hot plasma rises in the middle of hot granules and moves horizontally to the granular edges while cooling; during this advective motion, it drags any weak magnetic field into the intergranular dark lanes where the plasma sinks. Plasma flows converge into these intergranular lanes and magnetic fields dragged into those places cannot escape. In summary, the brightest places that, we assume, would be the largest contributors to the net linear polarization correspond to the centre of convection cells and are blue-shifted, whereas magnetic fields are expected at the intergranular dark lanes where plasma sinks into the star and are thus redshifted. This corresponds neatly to the observational data from the polarization signals in Betelgeuse. We must conclude that linear polarization in Betelgeuse comes from the rising plasma in the centre of convection cells and that these convection cells cover the surface of Betelgeuse, leaving only narrow, cold, redshifted intergranular lanes in which weak magnetic fields are found. In view of this evidence, the surface of Betelgeuse cannot be modelled with a few bright spots on top of a homogeneously bright surface, but rather as a collection of convective cells with bright broad centres and dark narrow lanes. The new model ------------- Mathematically, the best way to do this modelling is by describing the surface brightness of Betelgeuse as the absolute value of a linear combination of real spherical harmonics up to a maximum value of the index $\ell$. By real spherical harmonics we mean that we handle the real and imaginary parts of the usual complex spherical harmonics separately and with different coefficients. Specifically, the star brightness model is defined by a set of parameters $a_n$ which result in the brightness at a point at distance $\mu$ from disk centre and polar angle $\chi$ (see Fig.\[cartoon\]) given by: $$B(\mu,\chi)=\left \| \sum_{\substack{\ell=0,\ell_{max} \\ m=-\ell,+\ell}} a_{\ell}^m y_{\ell}^m(\mu,\chi) \right \| \label{Beq}$$ ![image](cartoon.png){width="\textwidth"} For completeness, we have defined the function $y_{\ell}^m(\mu,\chi)$ as $$\begin{aligned} y_{\ell}^m(\mu,\chi)=\sqrt{2} (-1)^m \Re(Y_{\ell}^m(\mu,\chi)) \quad \textrm{if} \quad m>0 \\ y_{\ell}^m(\mu,\chi)=\sqrt{2} (-1)^m \Im(Y_{\ell}^m(\mu,\chi)) \quad \textrm{if} \quad m<0 \\ y_{\ell}^m(\mu,\chi)=\Re(Y_{\ell}^m(\mu,\chi)) \quad \textrm{if} \quad m=0 \end{aligned}$$ with $\Re$ and $\Im$ meaning the real and imaginary parts respectively, and, as usual, $$Y^m_n(\mu,\chi) = \sqrt{\frac{2n+1}{4\pi} \frac{(n-m)!}{(n+m)!}} e^{i m \chi} P^m_n(\mu)$$ where $P^m_n(\mu)$ is the associated Legendre polynomial. The velocity at each point over the disk is radial and proportional to the brightness. In the presence of convection, this proportionality can be justified by arguing that the radiative losses, hence the brightness, have to be roughly balanced by the enthalpy flux (see, e.g., [@Nordlund2009] discussing the solar granulation case). Brightness therefore acquires a dependence also on $v_z$ as $$B(\mu,\chi,v_z) \propto \sigma T_{eff}^4 \approx \rho v_z \left( \frac{5}{2}k_BT+x\Xi \right)$$ where the left side recalls the black body relationship between brightness and effective temperature, and on the right side, the gas temperature $T$ and the ionization potential of hydrogen $\Xi$ times the ionization fraction $x$, times the density $\rho$ are transported upwards with a vertical velocity $v_z$. Our brightness distribution is adimensional and normalized. To provide a velocity with the right units but proportional to this brightness we define a global velocity $v_0$ that gives the maximum upflow velocity at the brightest point. This velocity $v_0$ must be large enough in order for the maximum Doppler shift (at disk centre) to encompass the most blue-shifted polarization signal observed in the series of data. For Betelgeuse, this velocity is at least 40km.s$^{-1}$, but is rather 60km.s$^{-1}$ for $\mu$Cep and just 30km.s$^{-1}$ for CETau, two comparative M-type RSGs for which we also have spectropolarimetric Narval observations . All put together, the Doppler velocity at any particular point over the disk will be modelled as $$v=v_0\cos(\mu) B(\mu,\chi,v_z)$$ The dark intergranular lanes are assigned here a near zero velocity, rather than a negative one. This only offsets the zero velocity definition. This assumption that the brightness distribution corresponds to convective cells (and hence that velocity and brightness are related) is demonstrated in Sect. 4. Finally, even if present, horizontal velocities are not considered in our model. We also have to fix the Doppler width of the local profile from a single point in the atmosphere of Betelgeuse. This width is actually the convolution of the physical broadening of the profile emitted by Betelgeuse (combination of the thermal broadening, micro- and macroturbulence and any other physical broadening mechanism) with the PSF of our spectrograph. This combined width is found again by examination of the data. The individual peaks in the polarization profiles, as well as the Stokes $I$ profiles, are found to have an average FWHM of 20km.s$^{-1}$ in our data. This width is probably the combination of the instrumental width and various sources of stellar broadening that will limit our ability to resolve stellar structures. In an ad hoc attempt to separate these two sources of broadening, we set the FWHM of a point source in our model to 10km.s$^{-1}$. The profile emitted by each point is therefore described by a Gaussian of FWHM 10km.s$^{-1}$ centered at $v_z$: $$e^{-(v-v_z)^2/6^2}.$$ Integrated over the stellar disk, the total polarization profile is just the addition of these profiles for all points with amplitude given by the brightness weighted by the Rayleigh scattering geometry factor $\sin^2 \mu$ and projected into Stokes $Q$ and $U$ for the position angle $$\begin{aligned} Q(v)=\sum_{\mu,\chi,v_z} B(\mu,\chi) \sin^2 \mu \cos 2\chi e^{-(v-v_z)^2/6^2}\\ U(v)=\sum_{\mu,\chi,v_z} B(\mu,\chi) \sin^2 \mu \sin 2\chi e^{-(v-v_z)^2/6^2} \label{QU}\end{aligned}$$ Last scattering approximation ----------------------------- The simple addition of profiles is justified by assuming that the polarization is fully due to the last scattering of the photon on each point over the disk. This assumption has been used in the past to model the continuum polarization of the Sun (Stenflo 2005) or to attempt to interpret the details of the second solar spectrum [e.g., @Anusha2010; @Belluzzi2011]. Applied to the Sun or to the present case of Betelgeuse, this approximation assumes that the degree of linear polarization is weak everywhere so that instead of being created through multiple scattering events, the emergent linear polarization at each point of the stellar surface is created by only one scattering event, the last one. Through this assumption, the only contribution of limb darkening onto the polarization signal is through the amplification of the anisotropy in the radiation field at the height of scattering. We also hypothesise that the amplitude of linear polarization is determined uniquely by the brightness. This is of course not rigorously true. Locally, the amount of polarization in the continuum will be given by the local brightness, but also by the anisotropy of the radiation field on which, for example, the height of the scattering point over the atmosphere of Betelgeuse plays a non-negligible role. Thus, under the symbol $B(\mu,\chi,v_z)$, referred to as brightness above, we may also find height variations in the atmosphere. Only direct comparisons of our inferred images with those obtained by other techniques permit the assumption that actual brightness dominates our adimensional parameter $B(\mu,\chi,v_z)$. Imaging through inversion ========================= Given a set of Stokes $Q$ and $U$ profiles as those published by [@Auriere2016] or [@Mathias2018], our purpose is to find the right set of $a_n$ parameters in Eq.\[Beq\] that better fit the observed profiles. A simple Marquardt-Levemberg minimisation technique suffices to find that optimal fit with no further regularisation than the squared difference between model and observation $\chi^2$. Figure\[plotSp\_0109\] shows one example of observations plotted as dots for both $Q$ and $U$ and the best fit as a continuous line. The brightness image that produces the best fit profiles is represented on the left. ![image](Y5Specta_09012014.png){width="\textwidth"} The quality of the fit allows one to conclude that the brightness distribution is certainly one of the possible brightness distributions on Betelgeuse, though perhaps not the only one. One obvious ambiguity is the 180-degrees symmetry inherent to linear polarization dependence on geometry visible in the double angles of Eq.\[QU\]. The image in Fig.\[plotSp\_0109\] can thus be rotated by 180 degrees with totally identical resulting profiles. It cannot be excluded that such 180-degrees ambiguities may also be at work in other more subtle manner. For example it can be seen that any linear combination of the original image and its 180-degrees ambiguous counterpart may also be a solution. Another source of multiple solutions is the choice of $l_{max}$, which in Eq.\[Beq\] sets the maximum quantum number $l$ of the the spherical harmonics used in the decomposition of the brilliance. This value sets the size of the smallest structures that the code will retrieve. Therefore we discuss it at length in Sect. 5. But a first bound can be set by noticing that, given the number of wavelength bins with polarization signal, the maximum number of degrees of freedom in the data must be around 30. This corresponds roughly to the number of coefficients for $l_{max}=5$ (35), and this sets a limit to the highest $l_{max}$ we can afford with the present data. One is left with the multiple solutions that $l_{max}$ values of one to five may provide. Rather than a lengthy and cumbersome exploration of ambiguities and multiple solutions we have preferred to resort to comparisons of the images inferred with our present method with the images obtained through other observational techniques as interferometry. Such comparisons are also produced by [@Auriere2016] and [@Montarges2016] between spectropolarimetric images with the basic model and images inferred from the data obtained by the PIONEER interferometer at similar dates. Using the same data, we expect that the general agreement between our new model and their inferred images is still valid. Figure\[ALMA\] (left) shows that this is the case for the date of January, 9th 2014 up to a 180 degrees rotation with respect to the solution shown in Fig.\[plotSp\_0109\]. More recently, on November 2015, ALMA also observed Betelgeuse with high resolution, though looking at high atmospheric layers [@Kervella2018]. Our data from December 9th, 2015 shows a correlation if one accepts that it is the brightest north-east granule in our image (Fig.\[ALMA\] right) that raises higher and thus becomes their single hot spot. ![image](fig3a.png){width="50.00000%"}![image](VLT_Jan2014.png){width="40.00000%"} ![image](fig3b.png){width="50.00000%"}![image](aa31761-17-fig4.png){width="40.00000%"} A final comparison has been made with the other M type RSG, CETau, which has been observed by [@Montarges2018] with PIONEER on the VLT, and which is also monitored in spectropolarimetry with Narval at TBL [@Tessore2017a] . The reconstructed images from interferometry show a dominating central bright spot. Figure\[CETau\] shows an image inferred from inversion of the linear polarization signals of this star with the same model described above except for the value of the $v_0$ velocity which is reduced to 30km.s$^{-1}$ to better fit the observed Doppler span of the signals. A bright granule appears well centred in our image as well. This result is quite rewarding in two senses. First, a central granule is insensitive to the 180-degrees ambiguity. Therefore, there are no ambiguous images that could also be solution of our inversion problem. The present image with the central granule is the only solution our model offers and it nicely corresponds with the results from interferometry. Second, since the amplitude of linear polarization due to scattering diminishes towards the disk centre, our technique is less sensitive to features near disk centre and will tend to disregard them. Despite that, a dominant central granule is inferred. Those two facts increase the confidence in the inference of such a central bright granule. The comparison with the interferometric observations is a final demonstration that our technique recovers the most salient features of the photospheric brightness distribution of these stars and, in particular, of Betelgeuse. ![image](fig4.png){width="50.00000%"}![image](aa31471-17-fig4.png){width="50.00000%"} To determine if modelling as complex as this in terms of spherical harmonics was necessary just to reproduce the main brightness features captured by other techniques, we refer the reader to Fig. \[muCep\]. This single observation of $\mu$Cep on July 10th, 2017 is interpreted through the basic model of [@Auriere2016] and also with the new model based upon spherical harmonics. ![Images of $\mu$Cep from July 10th, 2015. They are the result of the inversion of LSD profiles with the basic model (left) or the new model based on spherical harmonics (right).[]{data-label="muCep"}](fig5a.png "fig:"){width="25.00000%"}![Images of $\mu$Cep from July 10th, 2015. They are the result of the inversion of LSD profiles with the basic model (left) or the new model based on spherical harmonics (right).[]{data-label="muCep"}](muCeph_Y5.png "fig:"){width="25.00000%"} The original basic method suffers from the fact that Stokes $Q$ and $U$ profiles have peaks at quite different wavelengths. The only available interpretation of such misaligned peaks in the basic model is in terms of elongated structures. But they are difficult to explain or justify in stellar terms. However, the very same profiles interpreted with the present method result in a characteristic image of several bright spots of round shapes, in agreement with the interpretation of the brightness distribution as convective cells. This is one further confidence test supporting the conclusion that we are imaging true brightness structures in the photosphere of these supergiants. The positive comparison of our inferred images with images retrieved through other techniques must be tainted, to end this section, with some negative remarks. The model presented above is able to reproduce all signals blue-shifted with respect to the star reference. But no redshifted signal can be explained with our model. However, some observed profiles do show small red-shifted signals which, while showing amplitudes three times smaller than the blue-shifted ones, are shifted to the red with up to 15% of the maximum Doppler shifts to the blue. One possible explanation for these red-shifted signals is that they are coming from the dark intergranular lanes where plasma may be sinking back towards the stellar interior (we assume here the convective explanation of these bright spots that will be demonstrated in the next section). However these lanes are dark and occupy a small area over the star, so that their contribution to the integrated signal cannot be large. Another possibility is that those redshifted signals are due to horizontal velocities when seen near the limb. Horizontal flows such as advection must diverge from the centres of the granules with velocities of the order of the speed of sound ($\sim 6$km.s$^{-1}$). Near the limb, those horizontal velocities would project onto the line of sight with positive or negative sign. The blueshifted projection would just add and modify the rest of the signal, and the redshifted projection would stand prominently with maximum redshifts comparable to the speed of sound. A third explanation is that the convective cells of Betelgeuse may rise high enough that we are able to see the top of those cells beyond the stellar limb. None of these explanations has been implemented in our model and thus those red-shifted signatures are left unexplained. Since they are small amplitude signals and completely independant from the rest of the signals, we argue that they are not affecting the images we present. However this may not be the case for all red supergiants. These signals are a clear indication that our simple model should evolve towards more sophisticated and realistic scenari. Convective cells ================ The central result of this work is that we are imaging convective structures on the photosphere of RSG through our spectropolarimetric technique. All the images shown in this paper’s figures do indeed show what can be easily interpreted as large convective cells in both shapes and sizes expected for red supergiants. However, we can provide further evidence that this is the right interpretation of these structures. In convective cells, one expects hot bright plasma rising over the cell. As density drops, the hot plasma loses flotability and stops rising. Plasma movements are then due exclusively to any horizontal velocity present. As long as plasma is hot, and while it cools down through radiation, these horizontal movements will dominate the dynamics. Finally, the cold plasma sinks in the dark, cold lanes between cells. These are the signatures that we must see in our data and images in order to confirm the convective nature of the imaged structures. To produce this evidence we must add to the picture the Stokes $V$ profiles observed often simultaneously with the Stokes $Q$ and $U$ profiles. These Stokes $V$ profiles were introduced and discussed by [@Mathias2018] who demonstrated also the Zeeman origin of these signatures. Through those signatures, the presence of weak magnetic fields in the photosphere of Betelgeuse was confirmed. A time-Fourier decomposition of these Stokes $V$ signals, observed since 2009, showed a quasi-periodicity of about 2000 days that has been linked to the characteristic convection scales [e.g. @Stothers2010]. In the convective atmosphere of the Sun one observes two regimes of magnetism. The strongest fields are able to maintain their topology against the convective motions of the plasma and even succeed in stopping convection altogether. They form active regions and sunspots. Weaker fields cannot, however, impact plasma movements and are dragged around by the flows. They end up in the intergranular lanes and, in particular, in the network that marks the boundaries of supergranulation. These weak fields trace the borders between convective cells where the cold plasma sinks into the solar interior. ![Histograms of the Doppler velocities (respect to the centre of mass of the star) of the peaks in the profiles of linear polarization (blue) and circular polarization (red) for the whole dataset available for Betelgeuse (2013-2018). Each peak in the profiles has been given a weight proportional to its amplitude to stress the large signals over small, mostly noise-induced, peaks.[]{data-label="histogram"}](histogram_velos.png){width="50.00000%"} Figure\[histogram\] shows histograms of the velocities of the linear polarization and of the Stokes $V$ peaks. Since the algorithm for the determination of peaks may be easily mislead by noise, we have weighted the histogram by the amplitude of the peaks (a peak with an amplitude twice that of another peak is counted twice for the statistics). Linear polarization, arising from the brightest places, that is, the centre of the convection cells, is blueshifted by about 15km.s$^{-1}$ in about 70% of the weighted cases. A secondary peak of linear polarization appears at +10km.s$^{-1}$, thus redshifted. This secondary peak corresponds to those low amplitude peaks left largely unexplained, as said in the previous section. The Stokes $V$ signals on the other hand are redshifted by about 10km.s$^{-1}$. This may be easily compared with those weak fields in the solar case, that are dragged around by plasma motions and found in the dark lanes between convective cells where the cold plasma sinks. Numerical simulations of magnetic fields in Betelgeuse [@Freytag2002] confirm this general scenario. We can interpret these histograms as follows. The linear polarization comes from the brightest places, that is from the centre of convection cells where hot plasma rises into the atmosphere and blue-shifts the profile. Stokes $V$ comes from photospheric magnetic fields in Betelgeuse that would be too weak to alter plasma motions and tend to concentrate in the intergranular lanes. Thus, Fig.\[histogram\] confirms that the bright spots imaged with our technique in Betelgeuse are actual convection cells. And, by extension, those bright spots observed with other techniques and which coincide with our spots also are due to convective cells, though perhaps those that overshoot into higher atmospheric layers, depending on the observation wavelength. The average velocities measured are around 15km.s$^{-1}$, both blue and red-shifted. This must correspond to the typical radial velocities of convection in Betelgeuse weighted by the $\cos \mu$ integration factor over the stellar disk, or two thirds of the typical radial velocity of those convective movements. We have to conclude that the typical convective velocity in Betelgeuse is of the order of 22km.s$^{-1}$. This matches nicely with the velocities found by [@Freytag2002] in numerical simulation codes, velocities which are up to five or seven times the speed of sound. More classical values, as those of 7km.s$^{-1}$ found by [@Stothers2010] from basic convection theory, are too small to fit the profiles, and have to be understood as velocities occurring deeper in the convective interior, before the density and temperature drops convert the rise velocities into shocks. Size of convective cells ======================== ![Histograms of the final value of the merit function $\chi^2$ after inversion of the whole dataset of Betelgeuse with $\ell_{max}=3,4,$ and 5 respectively. A marginal improvement of the inversion results is seen with larger values of $\ell_{max}$.[]{data-label="Chi2"}](Chi2.png){width="50.00000%"} The inversion code described in Sect.2.2 requires, among other parameters, to fix the maximum value of the quantum number $\ell$. By fixing this number we constrain the size of the structures imaged on the surface of Betelgeuse or in other words, the size of the convective cells. Figure\[Chi2\] shows a histogram of the merit function $\chi^2$ when inverting the full data set with $\ell_{max}=3,4,$ and 5 respectively. There is a marginal diminution of the merit function for $\ell=5$. We found a better way to determine the right value for $\ell_{max}$ through the distribution of the number of peaks in the profiles as a function of $\ell_{max}$. The basic idea is that, in general, the larger $\ell_{max}$, the more granules will be visible in the image and, since each granule produces a peak in the profiles at the right wavelength, more peaks will be present in the profiles. This statement should however be tempered since misinterpretation may arise from the location of the granules on the stellar disk, the geometry of which may possibly lead to mutual cancellation if two granules merge into a single peak. Also, the spectrograph resolution may mask thin peaks Keeping in mind all these arguments, we counted the peaks in both Stokes $U$ and Stokes $Q$ in the observed profiles. The histogram on the number of peaks is seen in red bars in the 4 plots of Fig.\[L16\]. Next, we created random stars with $\ell_{max}\in [1;6]$ and computed the Stokes $Q$ and $U$ profiles arising from them in the framework of our new model. The histograms of the number of peaks in those sets of random stars are seen in the blue bars of the plots of Fig.\[L16\]. Before comparing the observed and simulated histograms, we must keep in mind that the observations are a series of 43 dates over more than four years. Hence, a temporal continuity is expected in that data set, and we shall not expect the observed histogram to represent the full dynamics of the convection of a star as Betelgeuse, but rather a snapshot. The counts from the random stars may, on its side, produce patterns that are not expected in real stars. With these further warnings in mind one can compare both histrograms for different values of $\ell$. As expected, $\ell=1$ results in too many cases with no polarization at all, something that has not been observed yet in Betelgeuse. On the other extreme, $\ell_{max}=6$ shows that two and three peaks are equally probable, contrary to observations that favour two peaks. Our dataset contains two profiles with four peaks, but we cannot conclude whether these are real or due to noise of some other bias. This advices against considering $\ell_{max}$ larger than five. The right value of $\ell_{max}$ appears therefore to be between $\ell_{max}=3$ and $\ell_{max}=5$. ![image](Histograms_L.png){width="100.00000%"} In order to decide between $\ell_{max}=3, 4,$ or 5 we run and show here two more tests. Figure\[Y5\_345\] shows the inversion for a given date with $\ell_{max}$=5. But only the right-most image shows the full result. In the leftmost image we show the reconstruction with just those coefficients up to $\ell=3$, and the centre image up to $\ell=4$. The main granule, here shown in the ambiguous solution that places it to the east, is the only structure recovered up to $\ell=3$. A transition image is seen in $\ell=4$ that adds small-scale structures which take granular shape with $\ell=5$. The series of images show that $\ell=3$ recovers the brightest granule, but that $\ell=5$ is required to recover the right shape and size of all but the largest granule. The value $\ell=4$, while recovering also the smaller structures, corrupts their shape and produces elongated granules. ![Inversion of Betelgeuse profiles from January 9th, 2014, also shown in Fig. 1. From left to right the image has been made using only the spherical harmonic coefficients up to $\ell=3,4,$ and 5 respectively. As expected, the main spot appears from $l=3$ while higher coefficients keep adding more and more detail. []{data-label="Y5_345"}](fig9.png){width=".5\textwidth"} A similar conclusion can be reached in the second and complementary test shown in Fig.\[Y345\_3\]. Here, three different inversions of the same data used in Fig.\[Y5\_345\] are presented with $\ell_{max}=3$ (left) to 5(right). But in each case, only the reconstruction up to $\ell=3$ is shown. This means that the image at left shows the final solution with $\ell_{max}=3$, but the centre and right images show only the result of the low-$\ell$ coefficients (spherical harmonics with $\ell=4$ and $\ell=4,5$ respectively are dropped). The main granule is seen in all three images, and we have used the 180-degrees ambiguity to place it always towards the east. At $\ell=3$ the inversion tries to fit the rest of the information available in the Stokes profiles by adding granules at the limb. This can be seen as a strategy to make small structures with low $\ell$ : Although the full structure is large, placed at the limb, we only see one half or one quarter of it. This behaviour disappears as $\ell$ grows, the higher $\ell$ coefficients allow for the smaller structures and there is no need for the coefficients up to $\ell=3$ to reproduce the available information with granules at the limb, the coefficients of larger $\ell$ recover those small structures. ![Another different reconstruction of Betelgeuse on January 9th, 2014. This time the inversion with $\ell_{max}=3,4,$ and 5 are shown from left to right, but the reconstructions have been made using only the spherical harmonics coefficients up to $\ell=3$. This means that the left figure with $\ell_{max}=3$ is final for that inversion, but in the centre and right images the high order coefficients are missing from the reconstruction. The images $\ell_{max}=4$ and 5 show only the biggest granule, the fine details being missed with the high order coefficients. The $\ell_{max}=3$ case however tries to reproduce the profiles by adding large granules on the edges thus making them small by projection. This shows the necessity for $\ell_{max}>3$ to correctly invert the observed profiles.[]{data-label="Y345_3"}](fig10.png){width=".5\textwidth"} All these arguments combined together lead to the conclusion that $\ell_{max}=5$ recovers the actual size and number of the convective cells in Betelgeuse. The typical size of convective cells for $\ell_{max}=5$ can be readily computed as $\pi/\ell=0.62$ times the disk radius, and their number in about 3 to 5 over the visible hemisphere of Betelgeuse Photospheric dynamics of Betelgeuse during 4.5 years ==================================================== Figure\[all\] shows the inferred images of the convective cells in Betelgeuse from November 2013 to August 2018 in 44 observations. The brightness is normalized for each image individually and all inversions have been made using $\ell_{max}=5$. The inversion process also takes advantage of the continuity of the observed profiles: the solution found for the previous date was used as initial solution of the next date in the Marquardt-Levemberg algorithm. Consequently, the 180 degrees ambiguity is solved once, in the first image and then the choice is carried on by continuity to all the other images. Most of the data inverted for these images have already been presented by [@Auriere2016] (November 2013 – April 2015) and by [@Mathias2018] (September 2015 – April 2017). The data covering September 2017 – August 2018 are used here for the first time. It extends the initial monitoring already presented in [@Auriere2016] and [@Mathias2018]. Fig. \[profiles\] shows the actual LSD profiles for Stokes $Q$ and $U$ as well as the total lineal polarization $\sqrt{Q^2+U^2}$ for every observation date. As in previous cases, observations were obtained with Narval on the Telescope Bernard Lyot at the Observatoire du Pic du Midi [@Auriere2003]. While the linear polarization signal is too small to be reliably detected in individual lines we have used the LSD technique [@Donati1997] to add-up the signal of several thousand lines (see details in [@Auriere2016]). The noise level on the resulting LSD profile is a few times $10^{-5}$ the continuum intensity, while the peak amplitudes of linear polarization are at least ten times larger and reach, in some particular occasions, 0.1% of the continuum intensity. ![image](fig11.png){width="\textwidth"} After inversion of all the available data set, we obtain a series of images dating back to the end of 2013. The series starts with a dominating granule towards the eastern hemisphere and roughly centred on the equator. A secondary granule is seen in the western hemisphere but close to the South. As said before, the main granule coincides with the bright spot seen in interferometric data with PIONIER by [@Montarges2016] on January 2014 and also put in evidence by [@Auriere2016]. It dominates the inferred images of Betelgeuse until the end of 2014. During this period, it keeps moving around, stretching and coming back. There is a nice agreement between these motions and deformations and what numerical simulations show [e.g. @Freytag2002; @Chiavassa2011]. ![image](fig_12.png){width="\textwidth"} By the end of 2014, the south granule grows in importance and two other granules appear in the north-western quarter. This new scenario lasts until the end of 2016. At this point, we return to a situation similar to that of 2014, with the two granules, one in the southeast, and the other towards the south. This coming back suggests two possibilities. It could be that these convective cells are long-lasting structures of the atmosphere of Betelgeuse. Even when the star goes through a period of more vigorous convection, they keep their presence so that the star always shows at least the two of them. Another possibility is that during the period 2014-2016 these granules lost predominance but new granules in the north-west and north respectively took the leading role. In such scenario, our choice between the two ambiguous solutions would be responsible for the granules being kept at the southern hemisphere. New interferometric observations could help us distinguish between such ambiguous solutions. The main granule lasts for at least two years. If we take the images at face value, we see a small granule in the north-western part already in November 2013. It keeps growing in brightness till 2016 then dwindles back to the end of the series. This sets a lifetime of four years for the convective cells of Betelgeuse. [@Mathias2018] produced a temporal Fourier analysis of the signals and found considerable power for timescales around 2000 days. We find that these periods, that well correspond to the time scales associated to convection [@Stothers2010] are well represented by the evolution seen in our images. Nevertheless, it is also important to stress that the profiles and the images keep changing all the time. At intervals of one week to one month the variations are well beyond noise levels. Although the number and average position of the granules are quite long-lasting, their actual shape and position keep changing in time spans of one month or even less, changes confirmed by the variability of the observed Stokes $Q$ and $U$ profiles. We also searched for patterns in the time series of the harmonic coefficients. We do not see any indications of periods in individual coefficients. Some of them appear to show trends: they are either diminishing or growing during the full series. But in general we could not extract any clear information from individual coefficients. Conclusion ========== We observe convective cells in the photosphere of Betelgeuse by inference from the inversion of linear polarization spectral profiles observed with the Narval instrument. The key to the interpretation is the blue-shift of the linear polarization signals, that we associate with the brighter parts, those places where hot plasma rises. Simultaneously, the circular polarization signals, having its origin in the Zeeman effect, are attributed to weak magnetic fields. In analogy to the solar atmosphere, such weak fields are dragged by convective flows to the dark intergranular lanes where the cold plasma sinks into the stellar interior. They will in consequence be redshifted with respect to the linear polarization signals. This key fact is confirmed by a series of other results, as the timescales of the signals, the aspect of the bright structures seen and their dynamics. We have been able to give a characteristic size of 0.6 stellar radii to these convective cells and to associate velocities to those plasma movements. The characteristic upflow velocities measured are of the order of 20km.s$^{-1}$ and we also have a suggestion as to the values of the horizontal advection velocities that would be slightly smaller, but of the same order. Downflows velocities from circular polarization are also of the order of 20km.s$^{-1}$ once the observed values are corrected from projection onto the line of sight. We have monitored Betelgeuse in linear polarization over more than four years. The shape and position of the convection cells change from month to month and even from week to week, but individual structures can be tracked over years. Among the several flaws of the imaging method we developed and used throughout this work, one of the most obvious is the presence of ambiguities. We could rotate by 180 degrees every single image presented here and the linear polarization would fit identically well the observations. Because of this it is unclear at given times whether the chosen solution to this ambiguity problem is the right one. It is clear that periodic checks with other imaging methods, in particular interferometry, is desirable to fix these ambiguities. This work was supported by the “Programme National de Physique Stellaire” (PNPS) of CNRS/INSU co-funded by CEA and CNES. [25]{} natexlab\#1[\#1]{} Anusha, L. S., Nagendra, K. N., Stenflo, J. O., [et al.]{} 2010, The Astrophysical Journal, 718, 988 Auri[è]{}re, M. 2003, EAS Publications Series, 9, 105 Auri[è]{}re, M., [L[ó]{}pez Ariste]{}, A., Mathias, P., [et al.]{} 2016, Astronomy and Astrophysics, 591, A119 Belluzzi, L. & [Trujillo Bueno]{}, J. 2011, The Astrophysical Journal, 743, 3 Chiavassa, A., Freytag, B., Masseron, T., & Plez, B. 2011, Astronomy and Astrophysics, 535, A22 Doherty, L. R. 1986, The Astrophysical Journal, 307, 261 Donati, J.-F., Semel, M., Carter, B. D., Rees, D. E., & [Collier Cameron]{}, A. 1997, Monthly Notices of the Royal Astronomical Society, 291, 658 Fluri, D. M. & Stenflo, J. O. 1999, Astronomy and Astrophysics, 341, 902 Freytag, B., Steffen, M., & Dorch, B. 2002, Astronomische Nachrichten, 323, 213 Gilliland, R. L. & Dupree, A. K. 1996, The Astrophysical Journal, 463, L29 Gustafsson, B., Edvardsson, B., Eriksson, K., [et al.]{} 2008, Astronomy and Astrophysics, 486, 951 Haubois, X., Perrin, G., Lacour, S., [et al.]{} 2009, Astronomy and Astrophysics, 508, 923 Josselin, E. & Plez, B. 2007, Astronomy and Astrophysics, 469, 671 Kervella, P., Decin, L., Richards, A. M. S., [et al.]{} 2018, Astronomy and Astrophysics, 609, A67 , E. & Landolfi, M. 2004, Astrophysics and Space Science Library, 307 Leroy, J. L. 1972, Astronomy and Astrophysics, 19, 287 Mathias, P., Auri[è]{}re, M., [L[ó]{}pez Ariste]{}, A., [et al.]{} 2018, ArXiv e-prints, arXiv:1804.01831 Montarg[è]{}s, M., Kervella, P., Perrin, G., [et al.]{} 2016, Astronomy and Astrophysics, 588, A130 Montarg[è]{}s, M., Norris, R., Chiavassa, A., [et al.]{} 2018, Astronomy and Astrophysics, 614, A12 Nordlund, [Å]{}., Stein, R. F., & Asplund, M. 2009, Living Reviews in Solar Physics, 6, 2 Schwarzschild, M. 1975, The Astrophysical Journal, 195, 137 Stenflo, J. O. 2005, Astronomy and Astrophysics, 429, 713 Stothers, R. B. 2010, The Astrophysical Journal, 725, 1170 Tessore, B. 2017, PhD thesis, Universit[é]{} de Montpellier Tessore, B., L[è]{}bre, A., Morin, J., [et al.]{} 2017, Astronomy and Astrophysics, 603, A129 [^1]: Based on observations obtained at the Télescope Bernard Lyot (TBL) at Observatoire du Pic du Midi, CNRS/INSU and Université de Toulouse, France.
{ "pile_set_name": "ArXiv" }
--- abstract: 'In this work we continue our research on nonharmonic analysis of boundary value problems as initiated in [@Ruzhansky-Tokmagambetov:IMRN]. There, we assumed that the eigenfunctions of the model operator on which the construction is based do not have zeros. In this paper we have weakened this condition extending the applicability of the developed pseudo-differential analysis. Also, we do not assume that the underlying set $\Omega$ is bounded.' address: - ' Michael Ruzhansky: Department of Mathematics Imperial College London 180 Queen’s Gate, London, SW7 2AZ United Kingdom ' - ' Niyaz Tokmagambetov: al–Farabi Kazakh National University 71 al–Farabi ave., Almaty, 050040 Kazakhstan, and Department of Mathematics Imperial College London 180 Queen’s Gate, London, SW7 2AZ United Kingdom ' author: - Michael Ruzhansky - Niyaz Tokmagambetov title: Nonharmonic analysis of boundary value problems without WZ condition --- [^1] Introduction ============ In [@Ruzhansky-Tokmagambetov:IMRN] the authors developed pseudo-differential calculus in terms of the ‘model’ densely defined operator L. The main examples are operators in $\Omega\subset\mathbb R^n$ equipped with (arbitrary) boundary conditions on $\partial\Omega$ for which the global Fourier analysis in terms of its eigenfunctions was introduced. Such a ‘model’ operator L does not have to be self-adjoint, so the construction is based on biorthogonal systems rather than on an orthonormal basis (to take into account the non-self-adjointness). Also, the operator L does not have to be elliptic. The ‘model’ operator ${\rm L}$ was considered as a differential operator of order $m$ with smooth coefficients on an open bounded set $\Omega\subset\mathbb R^n$ equipped with some boundary conditions which one can denote as (BC). In [@Ruzhansky-Tokmagambetov:IMRN] one worked with discrete sets of eigenvalues and eigenfunctions indexed by a countable set, and one developed elements of the symbolic calculus assuming that the system of eigenfunctions is the without zeros in $\Omega$ (so called WZ-system). We refer to [@Ruzhansky-Tokmagambetov:IMRN] for examples and an extensive list of references in this subject. In this paper we will drop some conditions of the ‘model’ operator ${\rm L}$. Let us consider the case when ${\rm L}$ is an arbitrary operator in $\Omega\subseteq\mathbb R^n$ with the discrete spectrum and the system of eigenfunctions which is a Riesz basis in $L^{2}(\Omega)$. Denote the corresponding countable index set by $\ind$. However, in different problems it may be more convenient to make different choices for this set, e.g. $\ind=\mathbb N$ or $\mathbb Z$ or $\mathbb Z^k$, etc. In order to allow different applications we will be denoting it by $\ind$, and without loss of generality we will assume that $$\label{EQ:ind} \ind \textrm{ is a subset of } \mathbb Z^{K} \textrm{ for some } K\geq 1.$$ For simplicity, one can think of $\ind=\mathbb Z$ or $\ind=\mathbb N\cup\{0\}$ throughout this paper. Thus, in this paper we will be always working in the following setting: \[Assumption\_1\] Let $\Omega\subseteq\mathbb R^{n}$, $n\geq 1$, be an open set. Assume that ${\rm L}$ is a densely defined operator with a discrete spectrum $\{\lambda_{\xi}\in\mathbb C: \, \xi\in\ind\}$ on $L^{2}(\Omega)$, and the system of corresponding eigenfunctions $\{u_{\xi}: \; \xi\in\ind\}$ is a Riesz basis in $L^{2}(\Omega)$ (i.e. for every $f\in L^{2}(\Omega)$ there exists a unique series $\sum_{\xi\in\ind} a_\xi u_\xi(x)$ that converges to $f$ in $L^{2}(\Omega)$), where $\ind$ is a countable set as in [(\[EQ:ind\])]{}, and we order the eigenvalues with the occurring multiplicities in the ascending order: $$\label{EQ: EVOrder} |\lambda_{j}|\leq|\lambda_{k}| \quad\textrm{ for } |j|\leq |k|.$$ We denote by $u_{\xi}$ the eigenfunction of ${\rm L}$ corresponding to the eigenvalue $\lambda_{\xi}$ for each $\xi\in\ind$, so that $$\label{SpecPr} {\rm L}u_{\xi}=\lambda_{\xi}u_{\xi} \,\,\,\,\,\, \textrm{ in } \Omega,\quad \textrm{ for all } \xi\in\ind.$$ The conjugate spectral problem is $$\label{ConjSpecPr} {\rm L^{\ast}}v_{\xi}=\overline{\lambda}_{\xi}v_{\xi}\,\,\,\,\,\, \textrm{ in } \Omega\quad \textrm{ for all } \xi\in\ind.$$ Let $\|u_{\xi}\|_{L^{2}}=1$ and $\|v_{\xi}\|_{L^{2}}=1$ for all $\xi\in\ind.$ Here, we can take biorthogonal systems $\{u_{\xi}\}_{\xi\in\ind}$ and $\{v_{\xi}\}_{\xi\in\ind}$, i.e. $$\label{BiorthProp} (u_{\xi},v_{\eta})_{L^2}=0 \,\,\,\, \hbox{for} \,\,\,\, \xi\neq\eta, \,\,\,\, \hbox{and} \,\,\,\, (u_{\xi},v_{\eta})_{L^2}=1 \,\,\,\, \hbox{for} \,\,\,\, \xi=\eta,$$ where $$(f, g)_{L^{2}}:=\int_{\Omega}f(x)\overline{g(x)}dx$$ is the usual inner product of the Hilbert space $L^{2}(\Omega)$. From N.K. Bari’s work [@bari] it follows that the system $\{u_{\xi}: \,\,\, \xi\in\ind\}$ is a basis in $L^{2}(\Omega)$ if and only if the system $\{v_{\xi}: \,\,\, \xi\in\ind\}$ is a basis in $L^{2}(\Omega)$. Therefore, by Bari [@bari], the system $\{v_{\xi}: \,\,\, \xi\in\ind\}$ is also a basis in $L^{2}(\Omega)$. Also, Assumption \[Assumption\_1\] will imply that the spaces $C^\infty_{{\rm L}}(\overline\Omega)$ and $C^\infty_{{\rm L}^*}(\overline\Omega)$ of test functions introduced in Subsection \[SEC:TD\] are dense in $L^2(\Omega)$. Define the weight $$\label{EQ:angle} \langle\xi\rangle:=(1+|\lambda_{\xi}|^2)^{\frac{1}{2m}},$$ which will be instrumental in measuring the growth/decay of Fourier coefficients and of symbols. Here $m>0$ is an arbitrary number that we fix throughout the paper. For simplicity we can take $m=1$. However, if L is, for example, a differential operator, it is convenient to take $m$ to be equal to its order. To give the interpretation for $\langle\xi\rangle$ in terms of the operator analysis, we can define the operator ${\rm L}^{\circ}$ by setting its values on the basis $u_{\xi}$ by $$\label{EQ:Lo-def} {\rm L}^{\circ} u_{\xi}:=\overline{\lambda_{\xi}} u_{\xi},\quad \textrm{ for all } \xi\in\ind.$$ If L is self-adjoint, we have ${\rm L}^{\circ}={\rm L}^*={\rm L}$. Consequently, we can informally think of $\langle\xi\rangle$ as of the eigenvalues of the positive (first order) operator $({\rm I}+{\rm L^\circ\, L})^{\frac{1}{2m}}.$ With a similar definition for $({\rm L}^{*})^{\circ}$, we can observe that $({\rm L}^{*})^{\circ}=({\rm L}^{\circ})^{*}$. Simplest examples of non-periodic boundary conditions were considered in [@Kanguzhin_Tokmagambetov_Tulenov] and [@Kanguzhin_Tokmagambetov] in the case of $\Omega=[0,1]$ being the segment. This extends to the non-periodic case the periodic analysis developed in [@RT07; @Ruzhansjy-Turunen:NFA; @RT; @Ruzhansky-Turunen-JFAA-torus] on the torus which can be viewed as analysis on $\Omega=[0,1]$ with periodic boundary conditions. We refer to [@Ruzhansky-Tokmagambetov:IMRN] for further examples. Preliminary {#SEC:TD} =========== In this section we collect some results on ${\rm L}$–distributions, ${\rm L}$–Fourier transform, Plancherel formula and Sobolev spaces $\mathcal H^{s}_{{\rm L}}(\Omega)$, and we omit the proofs because they are a straightforward extension of those in [@Ruzhansky-Tokmagambetov:IMRN]. Global distributions generated by the boundary value problem ------------------------------------------------------------ In this subsection we describe the spaces of distributions generated by ${\rm L}$ and by its adjoint ${\rm L}^*$ and the related global Fourier analysis. The more far-reaching aim of this analysis is to establish a version of the Schwartz kernel theorem for the appearing spaces of distributions. We first define the space $C_{{\rm L}}^{\infty}(\overline{\Omega})$ of test functions. \[TestFunSp\] The space $C_{{\rm L}}^{\infty}(\overline{\Omega}):={\rm Dom}({\rm L}^{\infty})$ is called the space of test functions for ${\rm L}$. Here we define $${\rm Dom}({\rm L}^{\infty}):=\bigcap_{k=1}^{\infty}{\rm Dom}({\rm L}^{k}),$$ where ${\rm Dom}({\rm L}^{k})$ is the domain of the operator ${\rm L}^{k}$, in turn defined as $${\rm Dom}({\rm L}^{k}):=\{f\in L^{2}(\Omega): \,\,\, {\rm L}^{j}f\in {\rm Dom}({\rm L}), \,\,\, j=0, \,1, \, 2, \ldots, k-1\}.$$ The Fréchet topology of $C_{{\rm L}}^{\infty}(\overline{\Omega})$ is given by the family of norms $$\label{EQ:L-top} \|\varphi\|_{C^{k}_{{\rm L}}}:=\max_{j\leq k} \|{\rm L}^{j}\varphi\|_{L^2(\Omega)}, \quad k\in\mathbb N_0, \; \varphi\in C_{{\rm L}}^{\infty}(\overline{\Omega}).$$ Analogously to the ${\rm L}$-case, we introduce the space $C_{{\rm L^{\ast}}}^{\infty}(\overline{\Omega})$ corresponding to the adjoint operator ${\rm L}_\Omega^*$ by $$C_{{\rm L^{\ast}}}^{\infty}(\overline{\Omega}):= {\rm Dom}(({\rm L^{\ast}})^{\infty})=\bigcap_{k=1}^{\infty}{\rm Dom}(({\rm L^{\ast}})^{k}),$$ where ${\rm Dom}(({\rm L^{\ast}})^{k})$ is the domain of the operator $({\rm L^{\ast}})^{k}$, $${\rm Dom}(({\rm L^{\ast}})^{k}):=\{f\in L^{2}(\Omega): \,\,\, ({\rm L^{\ast}})^{j}f\in {\rm Dom}({\rm L^{\ast}}), \,\,\, j=0, \ldots, k-1\},$$ which satisfy the adjoint boundary conditions corresponding to the operator ${\rm L}_\Omega^*$. The Fréchet topology of $C_{{\rm L}^*}^{\infty}(\overline{\Omega})$ is given by the family of norms $$\label{EQ:L-top-adj} \|\psi\|_{C^{k}_{{\rm L}^*}}:=\max_{j\leq k} \|({\rm L}^*)^{j}\psi\|_{L^2(\Omega)}, \quad k\in\mathbb N_0, \; \psi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega}).$$ Since we have $u_\xi\in C^\infty_{{\rm L}}(\overline\Omega)$ and $v_\xi\in C^\infty_{{\rm L}^*}(\overline\Omega)$ for all $\xi\in\ind$, we observe that Assumption \[Assumption\_1\] implies that the spaces $C^\infty_{{\rm L}}(\overline\Omega)$ and $C^\infty_{{\rm L}^*}(\overline\Omega)$ are dense in $L^2(\Omega)$. We note that if ${\rm L}$ is self-adjoint, i.e. if ${\rm L}^*={\rm L}$ with the equality of domains, then $C_{{\rm L^{\ast}}}^{\infty}(\overline{\Omega})=C_{{\rm L}}^{\infty}(\overline{\Omega}).$ In general, for functions $f\in C_{{\rm L}}^{\infty}(\overline{\Omega})$ and $g\in C_{{\rm L}^*}^{\infty}(\overline{\Omega})$, the $L^2$-duality makes sense in view of the formula $$\label{EQ:duality} ({\rm L}f, g)_{L^2(\Omega)}=(f,{\rm L}^*g)_{L^2(\Omega)}.$$ Therefore, in view of the formula [(\[EQ:duality\])]{}, it makes sense to define the distributions $\mathcal D'_{{\rm L}}(\Omega)$ as the space which is dual to $C_{{\rm L}^*}^{\infty}(\overline{\Omega})$. \[DistrSp\] The space $$\mathcal D'_{{\rm L}}(\Omega):=\mathcal L(C_{{\rm L}^*}^{\infty}(\overline{\Omega}), \mathbb C)$$ of linear continuous functionals on $C_{{\rm L}^*}^{\infty}(\overline{\Omega})$ is called the space of ${\rm L}$-distributions. We can understand the continuity here either in terms of the topology [(\[EQ:L-top-adj\])]{} or in terms of sequences, see Proposition \[TH: UniBdd\]. For $w\in\mathcal D'_{{\rm L}}(\Omega)$ and $\varphi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega})$, we shall write $$w(\varphi)=\langle w, \varphi\rangle.$$ For any $\psi\in C_{{\rm L}}^{\infty}(\overline{\Omega})$, $$C_{{\rm L}^*}^{\infty}(\overline{\Omega})\ni \varphi\mapsto\int_{\Omega}{\psi(x)} \, \varphi(x)\, dx$$ is an ${\rm L}$-distribution, which gives an embedding $\psi\in C_{{\rm L}}^{\infty}(\overline{\Omega})\hookrightarrow\mathcal D'_{{\rm L}}(\Omega)$. We note that in the distributional notation formula [(\[EQ:duality\])]{} becomes $$\label{EQ:duality-dist} \langle{\rm L}\psi, \varphi\rangle=\langle \psi,\overline{{\rm L}^* \overline{\varphi}}\rangle.$$ With the topology on $C_{{\rm L}}^{\infty}(\overline{\Omega})$ defined by [(\[EQ:L-top\])]{}, the space $$\mathcal D'_{{\rm L^{\ast}}}(\Omega):=\mathcal L(C_{{\rm L}}^{\infty}(\overline{\Omega}), \mathbb C)$$ of linear continuous functionals on $C_{{\rm L}}^{\infty}(\overline{\Omega})$ is called the space of ${\rm L^{\ast}}$-distributions. \[TH: UniBdd\] A linear functional $w$ on $C_{{\rm L}^*}^{\infty}(\overline{\Omega})$ belongs to $\mathcal D'_{{\rm L}}(\Omega)$ if and only if there exists a constant $c>0$ and a number $k\in\mathbb N_0$ with the property $$\label{EQ: UnifBdd-s1} |w(\varphi)|\leq c \|\varphi\|_{C^{k}_{{\rm L}^*}} \quad \textrm{ for all } \; \varphi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega}).$$ The space $\mathcal D'_{{\rm L}}(\Omega)$ has many similarities with the usual spaces of distributions. For example, suppose that for a linear continuous operator $D:C_{{\rm L}}^{\infty}(\overline{\Omega})\to C_{{\rm L}}^{\infty}(\overline{\Omega})$ its adjoint $D^*$ preserves the domain of ${\rm L}^*$ and is continuous on the space $C_{{\rm L}^*}^{\infty}(\overline{\Omega})$, i.e. that the operator $D^*:C_{{\rm L}^*}^{\infty}(\overline{\Omega})\to C_{{\rm L}^*}^{\infty}(\overline{\Omega})$ is continuous. Then we can extend $D$ to $\mathcal D'_{{\rm L}}(\Omega)$ by $$\langle Dw,{\varphi}\rangle := \langle w, \overline{D^* \overline{\varphi}}\rangle \quad (w\in \mathcal D'_{{\rm L}}(\Omega),\; \varphi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega})).$$ This extends [(\[EQ:duality-dist\])]{} from L to other operators. The convergence in the linear space $\mathcal D'_{{\rm L}}(\Omega)$ is the usual weak convergence with respect to the space $C_{{\rm L}^*}^{\infty}(\overline{\Omega})$. The following principle of uniform boundedness is based on the Banach–Steinhaus Theorem applied to the Fréchet space $C_{{\rm L}^*}^{\infty}(\overline{\Omega})$. \[LEM: UniformBoundedness\] Let $\{w_{j}\}_{j\in\mathbb N}$ be a sequence in $\mathcal D'_{{\rm L}}(\Omega)$ with the property that for every $\varphi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega})$, the sequence $\{w_{j}(\varphi)\}_{j\in\mathbb N}$ in $\mathbb C$ is bounded. Then there exist constants $c>0$ and $k\in\mathbb N_0$ such that $$\label{EQ: UniformBoundedness} |w_{j}(\varphi)|\leq c \|\varphi\|_{C^{k}_{{\rm L}^*}} \quad \textrm{ for all } \; j\in\mathbb N, \,\, \varphi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega}).$$ The lemma above leads to the following property of completeness of the space of ${\rm L}$-distributions. \[TH: Com-nessDistr\] Let $\{w_{j}\}_{j\in\mathbb N}$ be a sequence in $\mathcal D'_{{\rm L}}(\Omega)$ with the property that for every $\varphi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega})$ the sequence $\{w_{j}(\varphi)\}_{j\in\mathbb N}$ converges in $\mathbb C$ as $j\rightarrow\infty$. Denote the limit by $w(\varphi)$. [(i)]{} Then $w:\varphi\mapsto w(\varphi)$ defines an ${\rm L}$-distribution on $\Omega$. Furthermore, $$\lim_{j\rightarrow\infty}w_{j}=w \,\,\,\,\,\,\,\, \hbox{in} \,\,\,\,\,\,\, \mathcal D'_{{\rm L}}(\Omega).$$ [(ii)]{} If $\varphi_{j}\rightarrow\varphi$ in $\in C_{{\rm L}^*}^{\infty}(\overline{\Omega})$, then $$\lim_{j\rightarrow\infty}w_{j}(\varphi_{j})=w(\varphi) \,\,\,\,\,\,\,\, \hbox{in} \,\,\,\,\,\,\, \mathbb C.$$ Similarly to the previous case, we have analogues of Proposition \[TH: UniBdd\] and Theorem \[TH: Com-nessDistr\] for ${\rm L^{\ast}}$-distributions. ${\rm L}$-Fourier transform {#SEC:FT} --------------------------- In this subsection we define the ${\rm L}$-Fourier transform generated by our operator ${\rm L}$ and its main properties. The main difference between the self-adjoint and non-self-adjoint problems ${\rm L}$ is that in the latter case we have to make sure that we use the right functions from the available biorthogonal families of $u_\xi$ and $v_\xi$. We start by defining the spaces that we will obtain on the Fourier transform side. Let $\mathcal S(\ind)$ denote the space of rapidly decaying functions $\varphi:\ind\rightarrow\mathbb C$. That is, $\varphi\in\mathcal S(\ind)$ if for any $M<\infty$ there exists a constant $C_{\varphi, M}$ such that $$|\varphi(\xi)|\leq C_{\varphi, M}\langle\xi\rangle^{-M}$$ holds for all $\xi\in\ind$. Here $\langle\xi\rangle$ is already adapted to our case since it is defined by [(\[EQ:angle\])]{}. The topology on $\mathcal S(\ind)$ is given by the seminorms $p_{k}$, where $k\in\mathbb N_{0}$ and $$p_{k}(\varphi):=\sup_{\xi\in\ind}\langle\xi\rangle^{k}|\varphi(\xi)|.$$ Continuous linear functionals on $\mathcal S(\ind)$ are of the form $$\varphi\mapsto\langle u, \varphi\rangle:=\sum_{\xi\in\ind}u(\xi)\varphi(\xi),$$ where functions $u:\ind \rightarrow \mathbb C$ grow at most polynomially at infinity, i.e. there exist constants $M<\infty$ and $C_{u, M}$ such that $$|u(\xi)|\leq C_{u, M}\langle\xi\rangle^{M}$$ holds for all $\xi\in\ind$. Such distributions $u:\ind \rightarrow \mathbb C$ form the space of distributions which we denote by $\mathcal S'(\ind)$. We now define the L-Fourier transform on $C_{{\rm L}}^{\infty}(\overline{\Omega})$. \[FT\] We define the ${\rm L}$-Fourier transform $$(\mathcal F_{{\rm L}}f)(\xi)=(f\mapsto\widehat{f}): C_{{\rm L}}^{\infty}(\overline{\Omega})\rightarrow\mathcal S(\ind)$$ by $$\label{FourierTr} \widehat{f}(\xi):=(\mathcal F_{{\rm L}}f)(\xi)=\int_{\Omega}f(x)\overline{v_{\xi}(x)}dx.$$ Analogously, we define the ${\rm L}^{\ast}$-Fourier transform $$(\mathcal F_{{\rm L}^{\ast}}f)(\xi)=(f\mapsto\widehat{f}_{\ast}): C_{{\rm L}^{\ast}}^{\infty}(\overline{\Omega})\rightarrow\mathcal S(\ind)$$ by $$\label{ConjFourierTr} \widehat{f}_{\ast}(\xi):=(\mathcal F_{{\rm L}^{\ast}}f)(\xi)=\int_{\Omega}f(x)\overline{u_{\xi}(x)}dx.$$ The expressions [(\[FourierTr\])]{} and [(\[ConjFourierTr\])]{} are well-defined by the Cauchy-Schwarz inequality, for example, $$\label{EQ: Ineq1} |\widehat{f}(\xi)|=\left|\int_{\Omega}f(x)\overline{v_{\xi}(x)}dx\right|\leq\|f\|_{L^{2}} \|v_{\xi}\|_{L^{2}}=\|f\|_{L^{2}}<\infty.$$ Moreover, we have \[LEM: FTinS\] The ${\rm L}$-Fourier transform $\mathcal F_{{\rm L}}$ is a bijective homeomorphism from $C_{{\rm L}}^{\infty}(\overline{\Omega})$ to $\mathcal S(\ind)$. Its inverse $$\mathcal F_{{\rm L}}^{-1}: \mathcal S(\ind) \rightarrow C_{{\rm L}}^{\infty}(\overline{\Omega})$$ is given by $$\label{InvFourierTr} (\mathcal F^{-1}_{{\rm L}}h)(x)=\sum_{\xi\in\ind}h(\xi)u_{\xi}(x),\quad h\in\mathcal S(\ind),$$ so that the Fourier inversion formula becomes $$\label{InvFourierTr0} f(x)=\sum_{\xi\in\ind}\widehat{f}(\xi)u_{\xi}(x) \quad \textrm{ for all } f\in C_{{\rm L}}^{\infty}(\overline{\Omega}).$$ Similarly, $\mathcal F_{{\rm L}^{\ast}}:C_{{\rm L}^{\ast}}^{\infty}(\overline{\Omega})\to \mathcal S(\ind)$ is a bijective homeomorphism and its inverse $$\mathcal F_{{\rm L}^{\ast}}^{-1}: \mathcal S(\ind)\rightarrow C_{{\rm L}^{\ast}}^{\infty}(\overline{\Omega})$$ is given by $$\label{ConjInvFourierTr} (\mathcal F^{-1}_{{\rm L}^{\ast}}h)(x):=\sum_{\xi\in\ind}h(\xi)v_{\xi}(x), \quad h\in\mathcal S(\ind),$$ so that the conjugate Fourier inversion formula becomes $$\label{ConjInvFourierTr0} f(x)=\sum_{\xi\in\ind}\widehat{f}_{\ast}(\xi)v_{\xi}(x)\quad \textrm{ for all } f\in C_{{\rm L^*}}^{\infty}(\overline{\Omega}).$$ By dualising the inverse ${\rm L}$-Fourier transform $\mathcal F_{{\rm L}}^{-1}: \mathcal S(\ind) \rightarrow C_{{\rm L}}^{\infty}(\overline{\Omega})$, the ${\rm L}$-Fourier transform extends uniquely to the mapping $$\mathcal F_{{\rm L}}: \mathcal D'_{{\rm L}}(\Omega)\rightarrow \mathcal S'(\ind)$$ by the formula $$\label{EQ: FTofDistr} \langle\mathcal F_{{\rm L}}w, \varphi\rangle:= \langle w,\overline{\mathcal F_{{\rm L}^*}^{-1}\overline{\varphi}}\rangle, \quad\textrm{ with } w\in\mathcal D'_{{\rm L}}(\Omega),\; \varphi\in\mathcal S(\ind).$$ It can be readily seen that if $w\in\mathcal D'_{{\rm L}}(\Omega)$ then $\widehat{w}\in\mathcal S'(\ind)$. The reason for taking complex conjugates in [(\[EQ: FTofDistr\])]{} is that, if $w\in C_{{\rm L}}^{\infty}(\overline{\Omega})$, we have the equality $$\begin{gathered} \langle \widehat{w},\varphi\rangle = \sum_{\xi\in\ind} \widehat{w}(\xi) \varphi(\xi)= \sum_{\xi\in\ind} \left( \int_\Omega w(x) \overline{v_\xi(x)}dx\right) \varphi(\xi)\\ = \int_\Omega w(x) \overline{\left( \sum_{\xi\in\ind} \overline{\varphi(\xi)} v_\xi(x)\right)} dx = \int_\Omega w(x) \overline{\left( \mathcal F_{{\rm L}^*}^{-1} \overline{\varphi} \right)} dx =\langle w,\overline{\mathcal F_{{\rm L}^*}^{-1}\overline{\varphi}}\rangle.\end{gathered}$$ Analogously, we have the mapping $$\mathcal F_{{\rm L}^*}: \mathcal D'_{{\rm L}^*}(\Omega)\rightarrow \mathcal S'(\ind)$$ defined by the formula $$\label{EQ: FTofDistr2} \langle\mathcal F_{{\rm L}^*}w, \varphi\rangle:= \langle w,\overline{\mathcal F_{{\rm L}}^{-1}\overline{\varphi}}\rangle, \quad\textrm{ with } w\in\mathcal D'_{{\rm L}^*}(\Omega),\; \varphi\in\mathcal S(\ind).$$ It can be also seen that if $w\in\mathcal D'_{{\rm L}^*}(\Omega)$ then $\widehat{w}\in\mathcal S'(\ind)$. We note that since systems of $u_\xi$ and of $v_\xi$ are Riesz bases, we can also compare $L^2$-norms of functions with sums of squares of Fourier coefficients. The following statement follows from the work of Bari [@bari Theorem 9]: \[LEM: FTl2\] There exist constants $k,K,m,M>0$ such that for every $f\in L^{2}(\Omega)$ we have $$m^2\|f\|_{L^{2}}^2 \leq \sum_{\xi\in\ind} |\widehat{f}(\xi)|^2\leq M^2\|f\|_{L^{2}}^2$$ and $$k^2\|f\|_{L^{2}}^2 \leq \sum_{\xi\in\ind} |\widehat{f}_*(\xi)|^2\leq K^2\|f\|_{L^{2}}^2.$$ However, we note that the Plancherel identity can be also achieved in suitably defined $l^2$-spaces of Fourier coefficients, see Proposition \[PlanchId\]. Plancherel formula, Sobolev spaces $\mathcal H^{s}_{{\rm L}}(\Omega)$, and their Fourier images {#SEC:Sobolev} -------------------------------------------------- In this subsection we discuss Sobolev spaces adapted to ${\rm L}$ and their images under the L-Fourier transform. We start with the $L^2$-setting, where we can recall inequalities between $L^2$-norms of functions and sums of squares of their Fourier coefficients, see Lemma \[LEM: FTl2\]. However, below we show that we actually have the Plancherel identity in a suitably defined space $l^{2}_{{\rm L}}$ and its conjugate $l^{2}_{{\rm L}^{*}}$. Let us denote by $$l^{2}_{{\rm L}}=l^2({\rm L})$$ the linear space of complex-valued functions $a$ on $\ind$ such that $\mathcal F^{-1}_{{\rm L}}a\in L^{2}(\Omega)$, i.e. if there exists $f\in L^{2}(\Omega)$ such that $\mathcal F_{{\rm L}}f=a$. Then the space of sequences $l^{2}_{{\rm L}}$ is a Hilbert space with the inner product $$\label{EQ: InnerProd SpSeq-s} (a,\ b)_{l^{2}_{{\rm L}}}:=\sum_{\xi\in\ind}a(\xi)\ \overline{(\mathcal F_{{\rm L^{\ast}}}\circ\mathcal F^{-1}_{{\rm L}}b)(\xi)}$$ for arbitrary $a,\,b\in l^{2}_{{\rm L}}$. The reason for this choice of the definition is the following formal calculation: $$\begin{aligned} \label{EQ:PL-prelim} \nonumber (a,\ b)_{l^{2}_{{\rm L}}}& =\sum_{\xi\in\ind}a(\xi)\ \overline{(\mathcal F_{{\rm L^{\ast}}}\circ\mathcal F^{-1}_{{\rm L}}b)(\xi)}\\ \nonumber &=\sum\limits_{\xi\in\ind }a(\xi)\int_{\Omega}\overline{(\mathcal F^{-1}_{{\rm L}}b)(x)}u_{\xi}(x)dx\\ \nonumber &=\int_{\Omega}\left[\sum\limits_{\xi\in\ind}a(\xi)u_{\xi}(x)\right]\overline{(\mathcal F^{-1}_{{\rm L}}b)(x)}dx\\ \nonumber &=\int_{\Omega}(\mathcal F^{-1}_{{\rm L}}a)(x)\overline{(\mathcal F^{-1}_{{\rm L}}b)(x)} dx\\ &=(\mathcal F^{-1}_{{\rm L}}a,\,\mathcal F^{-1}_{{\rm L}}b)_{L^{2}},\end{aligned}$$ which implies the Hilbert space properties of the space of sequences $l^{2}_{{\rm L}}$. The norm of $l^{2}_{{\rm L}}$ is then given by the formula $$\label{EQ:l2norm} \|a\|_{l^{2}_{{\rm L}}}=\left(\sum_{\xi\in\ind}a(\xi)\ \overline{(\mathcal F_{{\rm L^{\ast}}}\circ\mathcal F^{-1}_{{\rm L}}a)(\xi)}\right)^{1/2}, \quad \textrm{ for all } \; a\in l^{2}_{{\rm L}}.$$ We note that individual terms in this sum may be complex-valued but the whole sum is real and nonnegative due to formula [(\[EQ:PL-prelim\])]{}. Analogously, we introduce the Hilbert space $$l^{2}_{{\rm L^{\ast}}}=l^{2}({\rm L^{\ast}})$$ as the space of functions $a$ on $\ind$ such that $\mathcal F^{-1}_{{\rm L^{\ast}}}a\in L^{2}(\Omega)$, with the inner product $$\label{EQ: InnerProd SpSeq-s_2} (a,\ b)_{l^{2}_{{\rm L^{\ast}}}}:=\sum_{\xi\in\ind}a(\xi)\ \overline{(\mathcal F_{{\rm L}}\circ\mathcal F^{-1}_{{\rm L^{\ast}}}b)(\xi)}$$ for arbitrary $a,\,b\in l^{2}_{{\rm L^{\ast}}}$. The norm of $l^{2}_{{\rm L^{\ast}}}$ is given by the formula $$\|a\|_{l^{2}_{{\rm L^{\ast}}}}=\left(\sum_{\xi\in\ind}a(\xi)\ \overline{(\mathcal F_{{\rm L}}\circ\mathcal F^{-1}_{{\rm L^{\ast}}}a)(\xi)}\right)^{1/2}$$ for all $a\in l^{2}_{{\rm L^{\ast}}}$. The spaces of sequences $l^{2}_{{\rm L}}$ and $l^{2}_{{\rm L^{\ast}}}$ are thus generated by biorthogonal systems $\{u_{\xi}\}_{\xi\in\ind}$ and $\{v_{\xi}\}_{\xi\in\ind}$. The reason for their definition in the above forms becomes clear again in view of the following Plancherel identity: [(Plancherel’s identity)]{}\[PlanchId\] If $f,\,g\in L^{2}(\Omega)$ then $\widehat{f},\,\widehat{g}\in l^{2}_{{\rm L}}, \,\,\, \widehat{f}_{\ast},\, \widehat{g}_{\ast}\in l^{2}_{{\rm L^{\ast}}}$, and the inner products [(\[EQ: InnerProd SpSeq-s\]), (\[EQ: InnerProd SpSeq-s\_2\])]{} take the form $$(\widehat{f},\ \widehat{g})_{l^{2}_{{\rm L}}}=\sum_{\xi\in\ind}\widehat{f}(\xi)\ \overline{\widehat{g}_{\ast}(\xi)}$$ and $$(\widehat{f}_{\ast},\ \widehat{g}_{\ast})_{l^{2}_{{\rm L^{\ast}}}}=\sum_{\xi\in\ind}\widehat{f}_{\ast}(\xi)\ \overline{\widehat{g}(\xi)}.$$ In particular, we have $$\overline{(\widehat{f},\ \widehat{g})_{l^{2}_{{\rm L}}}}= (\widehat{g}_{\ast},\ \widehat{f}_{\ast})_{l^{2}_{{\rm L^{\ast}}}}.$$ The Parseval identity takes the form $$\label{Parseval} (f,g)_{L^{2}}=(\widehat{f},\widehat{g})_{l^{2}_{{\rm L}}}=\sum_{\xi\in\ind}\widehat{f}(\xi)\ \overline{\widehat{g}_{\ast}(\xi)}.$$ Furthermore, for any $f\in L^{2}(\Omega)$, we have $\widehat{f}\in l^{2}_{{\rm L}}$, $\widehat{f}_{\ast}\in l^{2}_{{\rm L^{\ast}}}$, and $$\label{Planch} \|f\|_{L^{2}}=\|\widehat{f}\|_{l^{2}_{{\rm L}}}=\|\widehat{f}_{\ast}\|_{l^{2}_{{\rm L^{\ast}}}}.$$ Now we introduce Sobolev spaces generated by the operator ${\rm L}$: \[SobSp\] For $f\in\mathcal D'_{{\rm L}}(\Omega)\cap \mathcal D'_{{\rm L}^{*}}(\Omega)$ and $s\in\mathbb R$, we say that $$f\in\mathcal H^{s}_{{\rm L}}(\Omega)\; \textrm{ if and only if }\; \langle\xi\rangle^{s}\widehat{f}(\xi)\in l^{2}_{{\rm L}}.$$ We define the norm on $\mathcal H^{s}_{{\rm L}}(\Omega)$ by $$\label{SobNorm} \|f\|_{\mathcal H^{s}_{{\rm L}}(\Omega)}:=\left(\sum_{\xi\in\ind} \langle\xi\rangle^{2s}\widehat{f}(\xi)\overline{\widehat{f}_{\ast}(\xi)}\right)^{1/2}.$$ The Sobolev space $\mathcal H^{s}_{{\rm L}}(\Omega)$ is then the space of ${\rm L}$-distributions $f$ for which we have $\|f\|_{\mathcal H^{s}_{{\rm L}}(\Omega)}<\infty$. Similarly, we can define the space $\mathcal H^{s}_{{\rm L^{\ast}}}(\Omega)$ by the condition $$\label{SobNorm2} \|f\|_{\mathcal H^{s}_{{\rm L^{\ast}}}(\Omega)}:=\left(\sum_{\xi\in\ind}\langle\xi\rangle^{2s}\widehat{f}_{\ast}(\xi)\overline{\widehat{f}(\xi)}\right)^{1/2}<\infty.$$ We note that the expressions in [(\[SobNorm\])]{} and [(\[SobNorm2\])]{} are well-defined since the sum $$\sum_{\xi\in\ind} \langle\xi\rangle^{2s}\widehat{f}(\xi)\overline{\widehat{f}_{\ast}(\xi)}= (\langle\xi\rangle^{s}\widehat{f}(\xi),\langle\xi\rangle^{s}\widehat{f}(\xi))_{l^{2}_{\rm L}}\geq 0$$ is real and non-negative. Consequently, since we can write the sum in [(\[SobNorm2\])]{} as the complex conjugate of that in [(\[SobNorm\])]{}, and with both being real, we see that the spaces $\mathcal H^{s}_{{\rm L}}(\Omega)$ and $\mathcal H^{s}_{{\rm L^{\ast}}}(\Omega)$ coincide as sets. Moreover, we have \[SobHilSpace\] For every $s\in\mathbb R$, the Sobolev space $\mathcal H^{s}_{{\rm L}}(\Omega)$ is a Hilbert space with the inner product $$(f,\ g)_{\mathcal H^{s}_{{\rm L}}(\Omega)}:=\sum_{\xi\in\ind }\langle\xi\rangle^{2s}\widehat{f}(\xi)\overline{\widehat{g}_{\ast}(\xi)}.$$ Similarly, the Sobolev space $\mathcal H^{s}_{{\rm L^{\ast}}}(\Omega)$ is a Hilbert space with the inner product $$(f,\ g)_{\mathcal H^{s}_{{\rm L^{\ast}}}(\Omega)}:=\sum_{\xi\in\ind}\langle\xi\rangle^{2s}\widehat{f}_{\ast}(\xi)\overline{\widehat{g}(\xi)}.$$ For every $s\in\mathbb R$, the Sobolev spaces $\mathcal H^{s}(\Omega)$, $\mathcal H^{s}_{{\rm L}}(\Omega)$, and $\mathcal H^{s}_{{\rm L}^*}(\Omega)$ are isometrically isomorphic. Spaces $l^{p}({\rm L})$ and $l^{p}({\rm L}^*)$ {#SEC:lp} ---------------------------------------------- In this subsection we describe the $p$-Lebesgue versions of the spaces of Fourier coefficients. These spaces can be considered as the extension of the usual $l^p$ spaces on the discrete set $\ind$ adapted to the fact that we are dealing with biorthogonal systems. Thus, we introduce the spaces $l^{p}_{\rm L}=l^{p}({\rm L})$ as the spaces of all $a\in\mathcal S'(\ind)$ such that $$\label{EQ:norm1} \|a\|_{l^{p}({\rm L})}:=\left(\sum_{\xi\in\ind}| a(\xi)|^{p} \|u_{\xi}\|^{2-p}_{L^{\infty}(\Omega)} \right)^{1/p}<\infty,\quad \textrm{ for }\; 1\leq p\leq2,$$ and $$\label{EQ:norm2} \|a\|_{l^{p}({\rm L})}:=\left(\sum_{\xi\in\ind}| a(\xi)|^{p} \|v_{\xi}\|^{2-p}_{L^{\infty}(\Omega)} \right)^{1/p}<\infty,\quad \textrm{ for }\; 2\leq p<\infty,$$ and, for $p=\infty$, $$\|a\|_{l^{\infty}({\rm L})}:=\sup_{\xi\in\ind}\left( |a(\xi)|\cdot \|v_{\xi}\|^{-1}_{L^{\infty}(\Omega)}\right)<\infty.$$ \[REM:lps\] We note that in the case of $p=2$, we have already defined the space $l^{2}({\rm L})$ by the norm [(\[EQ:l2norm\])]{}. There is no problem with this since the norms [(\[EQ:norm1\])]{}-[(\[EQ:norm2\])]{} with $p=2$ are equivalent to that in [(\[EQ:l2norm\])]{}. Indeed, by Lemma \[LEM: FTl2\] the first one gives a homeomorphism between $l^{p}({\rm L})$ with $p=2$ just defined and $L^{2}(\Omega)$ while the space $l^{2}({\rm L})$ defined by [(\[EQ:l2norm\])]{} is isometrically isomorphic to $L^{2}(\Omega)$ by the Plancherel identity in Proposition \[PlanchId\]. Therefore, both norms lead to the same space which we denote by $l^{2}({\rm L})$. The norms [(\[EQ:norm1\])]{}-[(\[EQ:norm2\])]{} with $p=2$ and the one in [(\[EQ:l2norm\])]{} are equivalent, but there are advantages in using both of them. Thus, the norms [(\[EQ:norm1\])]{}-[(\[EQ:norm2\])]{} allow us to view $l^{2}({\rm L})$ as a member of the scale of spaces $l^{p}({\rm L})$ for $1\leq p\leq \infty$ with subsequent functional analytic properties, while the norm [(\[EQ:l2norm\])]{} is the one for which the Plancherel identity [(\[Planch\])]{} holds. Analogously, we also introduce spaces $l^{p}_{{\rm L^{\ast}}}=l^{p}({\rm L^{\ast}})$ as the spaces of all $b\in\mathcal S'(\ind)$ such that the following norms are finite: $$\|b\|_{l^{p}({\rm L^{\ast}})}=\left(\sum_{\xi\in\ind}| b(\xi)|^{p} \|v_{\xi}\|^{2-p}_{L^{\infty}(\Omega)} \right)^{1/p},\quad \textrm{ for }\; 1\leq p\leq2,$$ $$\|b\|_{l^{p}({\rm L^{\ast}})}=\left(\sum_{\xi\in\ind}| b(\xi)|^{p} \|u_{\xi}\|^{2-p}_{L^{\infty}(\Omega)} \right)^{1/p},\quad \textrm{ for }\; 2\leq p<\infty,$$ $$\|b\|_{l^{\infty}({\rm L^{\ast}})}=\sup_{\xi\in\ind}\left(|b(\xi)|\cdot \|u_{\xi}\|^{-1}_{L^{\infty}(\Omega)}\right).$$ Before we discuss several basic properties of the spaces $l^{p}({\rm L})$, we recall a useful fact on the interpolation of weighted spaces from Bergh and Löfström [@Bergh-Lofstrom:BOOK-Interpolation-spaces Theorem 5.5.1]: \[TH: IWS\] Let us write $d\mu_{0}(x)=\omega_{0}(x)d\mu(x),$ $d\mu_{1}(x)=\omega_{1}(x)d\mu(x),$ and write $L^{p}(\omega)=L^{p}(\omega d\mu)$ for the weight $\omega$. Suppose that $0<p_{0}, p_{1}<\infty$. Then $$(L^{p_{0}}(\omega_{0}), L^{p_{1}}(\omega_{1}))_{\theta, p}=L^{p}(\omega),$$ where $0<\theta<1$, $\frac{1}{p}=\frac{1-\theta}{p_{0}}+\frac{\theta}{p_{1}}$, and $\omega=\omega_{0}^{\frac{p(1-\theta)}{p_{0}}}\omega_{1}^{\frac{p\theta}{p_{1}}}$. From this it is easy to check that we obtain: For $1\leq p\leq2$, we have $$(l^{1}({\rm L}), l^{2}({\rm L}))_{\theta,p}=l^{p}({\rm L}),$$ $$(l^{1}({\rm L}^{\ast}), l^{2}({\rm L}^{\ast}))_{\theta,p}=l^{p}({\rm L}^{\ast}),$$ where $0<\theta<1$ and $p=\frac{2}{2-\theta}$. The reason that the interpolation above is restricted to $1\leq p\leq2$ is that the definition of $l^p$-spaces changes when we pass $p=2$, in the sense that we use different families of biorthogonal systems $u_\xi$ and $v_\xi$ for $p<2$ and for $p>2$. We note that if ${\rm L}={\rm L}^*$ is self-adjoint, so that we can take $u_\xi=v_\xi$ for all $\xi\in\ind$, then the scales $l^{p}({\rm L})$ and $l^{p}({\rm L}^{\ast})$ coincide and satisfy interpolation properties for all $1\leq p<\infty$. Using these interpolation properties we can establish further properties of the Fourier transform and its inverse: \[TH: HY\] Let $1\leq p\leq2$ and $\frac{1}{p}+\frac{1}{p'}=1$. There is a constant $C_{p}\geq 1$ such that for all $f\in L^{p}(\Omega)$ and $a\in l^{p}({\rm L})$ we have $$\label{EQ:HY} \|\widehat{f}\|_{l^{p'}({\rm L})}\leq C_{p}\|f\|_{L^{p}(\Omega)}\quad \textrm{ and }\quad \|\mathcal F_{{\rm L}}^{-1}a\|_{L^{p'}(\Omega)}\leq C_{p}\|a\|_{l^{p}({\rm L})}.$$ Similarly, we also have $$\label{EQ:HYast} \|\widehat{f}_*\|_{l^{p'}({\rm L}^*)}\leq C_{p}\|f\|_{L^{p}(\Omega)}\quad \textrm{ and } \quad \|\mathcal F_{{\rm L}^{*}}^{-1}b\|_{L^{p'}(\Omega)}\leq C_{p}\|b\|_{l^{p}({\rm L}^*)},$$ for all $b\in l^{p}({\rm L}^*)$. It follows from the proof that if ${\rm L}$ is self-adjoint, then the $l^{2}_{L}$-norms discussed in Remark \[REM:lps\] coincide, and so we can put $C_{p}=1$ in inequalities [(\[EQ:HY\])]{} and [(\[EQ:HYast\])]{}. If ${\rm L}$ is not self-adjoint, $C_{p}$ may in principle depend on ${\rm L}$ and its domain through constants from inequalities in Lemma \[LEM: FTl2\]. We now turn to the duality between spaces $l^{p}({\rm L})$ and $l^{q}({\rm L}^{\ast})$: \[TH:Duality lp\] Let $1\leq p<\infty$ and $\frac{1}{p}+\frac{1}{p'}=1$. Then $$\left(l^{p}({\rm L})\right)'=l^{p'}({\rm L}^{\ast}) \quad \textrm{ and }\quad \left(l^{p}({\rm L}^{\ast})\right)'=l^{p'}({\rm L}).$$ Schwartz’ kernel theorem {#SEC:Schwartz} ------------------------ In our case the Schwartz kernel theorem is also valid and here we will briefly discuss it. So, from now on we will make the following: \[Assumption\_4\] Assume that the number $s_0\in\mathbb R$ is such that we have $$\sum_{\xi\in\ind} \langle\xi\rangle^{-s_0}<\infty.$$ Recalling the operator ${\rm L}^{\circ}$ in [(\[EQ:Lo-def\])]{} the Assumption \[Assumption\_4\] is equivalent to assuming that the operator $({\rm I}+{\rm L^\circ L})^{-\frac{s_0}{4m}}$ is Hilbert-Schmidt on $L^2(\Omega)$. Indeed, recalling the definition of $\langle\xi\rangle$ in [(\[EQ:angle\])]{}, namely that $\langle\xi\rangle$ are the eigenvalues of $({\rm I}+{\rm L^\circ L})^{-\frac{s_0}{2m}}$, the condition that the operator $({\rm I}+{\rm L^\circ L})^{-\frac{s_0}{4m}}$ is Hilbert-Schmidt is equivalent to the condition that $$\label{EQ:HS-conv} \|({\rm I}+{\rm L^\circ L})^{-\frac{s_0}{4m}}\|_{\tt HS}^2\cong \sum_{\xi\in\ind} \langle\xi\rangle^{-s_0}<\infty.$$ If L is elliptic, we may expect that we can take any $s_0>n$ but this depends on the domain. The order $s_0$ will enter the regularity properties of the Schwartz kernels. We will use the notations $$C^{\infty}_{{\rm L}}(\overline{\Omega}\times \overline{\Omega}):= C^{\infty}_{{\rm L}}(\overline{\Omega})\bar\otimes C^{\infty}_{{\rm L}}(\overline{\Omega})$$ and $$C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega}\times \overline{\Omega}):= C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})\bar\otimes C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})$$ with the Fréchet topologies given by the family of tensor norms $$\label{EQ:L-top-Tensor} \|\varphi\otimes\psi\|_{C^{k}_{{\rm L}}(\overline{\Omega}\times \overline{\Omega})}:=\max_{j+l\leq k} \|{\rm L}^{j}\varphi\|_{L^2(\Omega)}\|{\rm L}^{l}\psi\|_{L^2(\Omega)}, \quad k\in\mathbb N_0, \; \varphi, \psi\in C_{{\rm L}}^{\infty}(\overline{\Omega})$$ and $$\label{EQ:L-top-Tensor-2} \|\varphi\otimes\psi\|_{C^{k}_{{\rm L}^*}(\overline{\Omega}\times \overline{\Omega})}:=\max_{j+l\leq k} \|({\rm L}^*)^{j}\varphi\|_{L^2(\Omega)}\|({\rm L}^*)^{l}\psi\|_{L^2(\Omega)}$$ for all $k\in\mathbb N_0, \; \varphi, \psi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega}),$ respectively, and for the corresponding dual spaces we write $$\mathcal D'_{{\rm L}}(\Omega\times\Omega):= \left(C^{\infty}_{{\rm L}}(\overline{\Omega}\times \overline{\Omega})\right)^\prime,$$ $$\mathcal D'_{{\rm L^{\ast}}}(\Omega\times\Omega):= \left(C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega}\times \overline{\Omega})\right)^\prime.$$ For any linear continuous operator $$A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$$ there exists a kernel $K_{A}\in \mathcal D'_{{\rm L}}(\Omega\times\Omega)$ such that for all $f\in C^{\infty}_{{\rm L}}(\overline{\Omega})$, we can write, in the sense of distributions, $$\label{EQ:int1} Af(x)=\int_{\Omega}K_{A}(x,y)f(y)dy.$$ As usual, $K_{A}$ is called the Schwartz kernel of $A$. For $f\in C^{\infty}_{{\rm L}}(\overline{\Omega})$, using the Fourier series formula $$f(y)=\sum\limits_{\eta\in\ind}\widehat{f}(\eta) u_{\eta}(y),$$ we can also write $$\label{EQ:int2} Af(x)=\sum\limits_{\eta\in\ind}\widehat{f}(\eta)\int_{\Omega}K_{A}(x,y)u_{\eta}(y)dy.$$ Also, for any linear continuous operator $$A:C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L^{\ast}}}(\Omega)$$ there exists a kernel $\widetilde{K}_{A}\in \mathcal D'_{{\rm L^{\ast}}}(\Omega\times\Omega)$ such that for all $f\in C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})$, we can write, in the sense of distributions, $$\label{EQ:int1} Af(x)=\int_{\Omega}\widetilde{K}_{A}(x,y)f(y)dy.$$ ${\rm L}$–admissible operators and ${\rm L}$-quantization {#SEC:admissible operators} ========================================================= In this section we describe the ${\rm L}$-quantization of the ${\rm L}$–admissible operator induced by the operator ${\rm L}$. We say that the linear continuous operator $$A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$$ belongs to the class of L–admissible operators if $$\sum\limits_{\eta\in\ind}u_{\eta}^{-1}(x)\, u_{\eta}(z)\, \int_{\Omega}K_{A}(x,y)u_{\eta}(y)dy$$ is in $\mathcal D'_{{\rm L}}(\Omega\times\Omega)$. \[RM:L-admissible: simple case\] In the case when ${\rm L}$ is the Laplace operator with periodic boundary conditions on the torus $\mathbb T^{n}$ the class of L–admissible operators coincides with the class of all periodic pseudo-differential operators as in [@Ruzhansky-Turunen-JFAA-torus]. So, from now on we will assume that operators $A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$ are from the class of L–admissible operators. \[RM:L-admissible\] Note, that the expression $$u_{\eta}^{-1}(x)\, \int_{\Omega}K_{A}(x,y)u_{\eta}(y)dy$$ exists for any operator $A$ from the class of L–admissible operators. Moreover, it is in $\mathcal D'_{{\rm L}}(\Omega)\otimes\mathcal S'(\ind).$ Indeed, since $\sum\limits_{\eta\in\ind}u_{\eta}^{-1}(x)\, u_{\eta}(z)\, \int_{\Omega}K_{A}(x,y)u_{\eta}(y)dy$ is in $\mathcal D'_{{\rm L}}(\Omega\times\Omega)$, by taking Fourier transform in z, we get this statement. We now define the L-symbol of an L-admissible operator. \[$L$–Symbols\] The ${\rm L}$-symbol of a linear continuous L–admissible operator $$A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$$ is defined by $$\sigma_{A}(x, \xi):=u_{\xi}^{-1}(x)\, \int_{\Omega}K_{A}(x,y)u_{\xi}(y)dy.$$ This is well-defined as an element of $\mathcal D'_{{\rm L}}(\Omega)\otimes\mathcal S'(\ind)$ in view of Remark \[RM:L-admissible\]. Indeed, we have $$Au_{\xi}=\int_{\Omega}K_{A}(x,y)u_{\xi}(y)dy,$$ and for $f\in C^{\infty}_{{\rm L}}(\overline{\Omega})$ from the expansion $$f(x)=\sum_{\xi\in\ind}\widehat{f}(\xi)u_{\xi}(x)$$ and by the operator $A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$ acting on $f$, we get $$Af(x)=\sum_{\xi\in\ind}\widehat{f}(\xi)Au_{\xi}(x)=\sum_{\xi\in\ind}\widehat{f}(\xi)\int_{\Omega}K_{A}(x,y)u_{\xi}(y)dy.$$ Now, if we define $$u_{\xi}(x)\sigma_{A}(x, \xi):=\int_{\Omega}K_{A}(x,y)u_{\xi}(y)dy,$$ we have the implication $$\label{EQ:KofA} K_{A}(x,y)=\sum_{\xi\in\ind}u_{\xi}(x)\sigma_{A}(x, \xi)\overline{v_{\xi}(y)}.$$ Therefore we obtain the following representation of the operator $A$ by its symbol: \[QuanOper\] Let $$A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$$ be a linear continuous L–admissible operator with [L]{}-symbol $\sigma_{A}\in\mathcal D'_{{\rm L}}(\Omega)\otimes\mathcal S'(\ind).$ Then the ${\rm L}$–quantization $$\label{Quantization} Af(x)=\sum_{\xi\in\ind} \widehat{f}(\xi)\,\sigma_{A}(x, \xi)\,u_{\xi}(x)$$ is true for every $f\in C^{\infty}_{{\rm L}}(\overline{\Omega})$ . The [L]{}-symbol $\sigma_{A}$ can be written as $$\label{FormSymb} \sigma_{A}(x,\xi)=u_{\xi}^{-1}(x)(Au_{\xi})(x).$$ \[COR: SymFor2\] We have the following equivalent formulae for [L]{}-symbols: $$\begin{aligned} {\rm (i)} \,\,\,\,\, &\sigma_{A}(x, \xi)=u_{\xi}^{-1}(x)(Au_{\xi})(x);\\ {\rm (ii)} \,\,\,\,\, &\sigma_{A}(x,\xi)=u_{\xi}^{-1}(x)\int_{\Omega}K_{A}(x,y)u_{\xi}(y)dy.\end{aligned}$$ Similarly, we can introduce an analogous notion of the ${\rm L^{\ast}}$-quantization. We say that the continuous operator $$A:C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L^{\ast}}}(\Omega)$$ belongs to the class of ${\rm L^{\ast}}$–admissible operators if $$\sum\limits_{\eta\in\ind}v_{\eta}^{-1}(x)\, v_{\eta}(z)\, \int_{\Omega}\widetilde{K}_{A}(x,y)v_{\eta}(y)dy$$ is in $\mathcal D'_{{\rm L}^{\ast}}(\Omega\times\Omega)$. \[RM:L\*-admissible\] Similarly to Remark \[RM:L-admissible\], note that the expression $$v_{\xi}^{-1}(x)\, \int_{\Omega}\widetilde{K}_{A}(x,y)v_{\xi}(y)dy$$ exists for any operator $A$ from the class of ${\rm L^{\ast}}$–admissible operators. Moreover, it is in $\mathcal D'_{{\rm L^{\ast}}}(\Omega)\otimes\mathcal S'(\ind).$ We also can define the ${\rm L^{\ast}}$-symbol of an ${\rm L^{\ast}}$–admissible operator. \[$L$–Symbols2\] The ${\rm L^{\ast}}$-symbol of a linear continuous ${\rm L}^{*}$–admissible operator $$A:C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L^{\ast}}}(\Omega)$$ is defined by $$\tau_{A}(x, \xi):=v_{\xi}^{-1}(x)\, \int_{\Omega}\widetilde{K}_{A}(x,y)v_{\xi}(y)dy.$$ Similarly to the case of L-symbols we have $$Av_{\xi}=\int_{\Omega}\widetilde{K}_{A}(x,y)v_{\xi}(y)dy,$$ and for $f\in C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})$ from the expantion $$f(x)=\sum_{\xi\in\ind}\widehat{f}_{\ast}(\xi)v_{\xi}(x)$$ and by the operator $A:C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L^{\ast}}}(\Omega)$ acting on $f$, we get $$Af(x)=\sum_{\xi\in\ind}\widehat{f}_{\ast}(\xi)Av_{\xi}(x)=\sum_{\xi\in\ind}\widehat{f}_{\ast}(\xi)\int_{\Omega}\widetilde{K}_{A}(x,y)v_{\xi}(y)dy.$$ Now, we have $$v_{\xi}(x)\tau_{A}(x, \xi):=\int_{\Omega}\widetilde{K}_{A}(x,y)v_{\xi}(y)dy,$$ hence also the implication $$\label{EQ:KofA2} \widetilde{K}_{A}(x,y)=\sum_{\xi\in\ind}v_{\xi}(x)\tau_{A}(x, \xi)\overline{u_{\xi}(y)}.$$ We also record the resulting representation of the operator $A$ by its symbol: \[QuanOper2\] Let $$A:C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L^{\ast}}}(\Omega)$$ be a linear continuous ${\rm L}^{\ast}$–admissible operator with ${\rm L}^{\ast}$-symbol $\tau_{A}\in\mathcal D'_{{\rm L^{\ast}}}(\Omega)\otimes\mathcal S'(\ind).$ Then the ${\rm L}^{\ast}$–quantization is given by $$\label{Quantization2} Af(x)=\sum_{\xi\in\ind} \widehat{f}_{\ast}(\xi)\,\tau_{A}(x, \xi)\,v_{\xi}(x)$$ for every $f\in C^{\infty}_{{\rm L}}(\overline{\Omega})$ . The ${\rm L}^{\ast}$-symbol $\tau_{A}$ can be written as $$\label{FormSymb} \tau_{A}(x,\xi)=v_{\xi}^{-1}(x)(Av_{\xi})(x).$$ \[COR: SymFor\] We have the following equivalent formulae for ${\rm L}^{\ast}$-symbols: $$\begin{aligned} {\rm (i)} \,\,\,\,\, &\tau_{A}(x, \xi)=v_{\xi}^{-1}(x)(Av_{\xi})(x);\\ {\rm (ii)} \,\,\,\,\, &\tau_{A}(x,\xi)=v_{\xi}^{-1}(x)\int_{\Omega}\widetilde{K}_{A}(x,y)v_{\xi}(y)dy.\end{aligned}$$ We now briefly describe the notion of Fourier multipliers which is a natural name for operators with symbols independent of $x$. In [@Delgado-Ruzhansky-Togmagambetov:nuclear] the analysis of this paper is applied to investigate the spectral properties of such operators, so we can be brief here. \[Lfm\] Let $A:C_L^{\infty}{(\overline{\Omega})}\rightarrow C_L^{\infty}{(\overline{\Omega})}$ be a continuous linear operator. We will say that $A$ is an $L$-Fourier multiplier if it satisfies $${\mathcal{F}_L}(Af)(\xi)=\sigma(\xi){\mathcal{F}_L}(f)(\xi),\; f\in C_{L}^{\infty}{(\overline{\Omega})},$$ for some $\sigma:\ind\rightarrow\mathbb C$. Analogously we define $L^*$-Fourier multipliers: Let $B:C_{L^*}^{\infty}{(\overline{\Omega})}\rightarrow C_{L^*}^{\infty}{(\overline{\Omega})}$ be a continuous linear operator. We will say that $B$ is an $L^*$-Fourier multiplier if it satisfies $${\mathcal{F}_{L^*}}(Bf)(\xi)=\tau(\xi){\mathcal{F}_{L^*}}(f)(\xi),\, f\in C_{L^*}^{\infty}{(\overline{\Omega})},$$ for some $\tau:\ind\rightarrow\mathbb C$. As used in [@Delgado-Ruzhansky-Togmagambetov:nuclear], we have the following simple relation between the symbols of an operator and its adjoint. \[admu\] The operator $A$ is an $L$-Fourier multiplier by $\sigma(\xi)$ if and only if $A^*$ is an $L^*$-Fourier multiplier by $\overline{\sigma(\xi)}$. Difference operators {#SEC:differences} ==================== In this section we discuss difference operators that will be instrumental in defining symbol classes for the symbolic calculus of operators. Let $q_{j}\in C^{\infty}({\Omega}\times{\Omega})$, $j=1,\ldots,l$, be a given family of smooth functions. We will call the collection of $q_j$’s [*[L]{}-strongly admissible*]{} if the following properties hold: - The multiplication by $q_{j}(\cdot,\cdot)$ is a continuous linear mapping on $C^{\infty}_{{\rm L^*}}(\overline{\Omega}\times \overline{\Omega})$, for all $j=1,\ldots,l$; - $q_{j}(x,x)=0$ and $\nabla_{y}q_{j}(x,y)|_{y=x}\not=0$ for all $j=1,\ldots,l$ and all $x\in\Omega$; - ${\rm rank}(\nabla_{y}q_{1}(x,y), \ldots, \nabla_{y}q_{l}(x,y))|_{y=x}=n $ for all $x\in\Omega$; - the diagonal in $\Omega\times\Omega$ is the only set when all of $q_j$’s vanish: $$\bigcap_{j=1}^l \left\{(x,y)\in\Omega\times\Omega: \, q_j(x,y)=0\right\}=\{(x,x):\, x\in\Omega\}.$$ We note that the first property above implies that for every $x\in\Omega$, the multiplication by $q_{j}(\cdot,\cdot)$ is also well-defined and extends to a continuous linear mapping on $\mathcal D'_{{\rm L}}(\Omega\times\Omega)$. Also, the last property above contains the second one but we chose to still give it explicitly for the clarity of the exposition. The collection of $q_j$’s with the above properties generalises the notion of a strongly admissible collection of functions for difference operators introduced in [@Ruzhansky-Turunen-Wirth:JFAA] in the context of compact Lie groups. We will use the multi-index notation $$q^{\alpha}(x,y):=q^{\alpha_1}_{1}(x,y)\cdots q^{\alpha_l}_{l}(x,y).$$ Analogously, the notion of an ${\rm L}^{*}$-strongly admissible collection suitable for the conjugate problem is that of a family $\widetilde{q}_{j}\in C^{\infty}({\Omega}\times{\Omega})$, $j=1,\ldots,l$, satisfying the properties: - The multiplication by $\widetilde{q}_{j}(\cdot,\cdot)$ is a continuous linear mapping on $C^{\infty}_{{\rm L}}(\overline{\Omega}\times\overline{\Omega})$, for all $j=1,\ldots,l$; - $\widetilde{q}_{j}(x,x)=0$ and $\nabla_{y}\widetilde{q}_{j}(x,y)|_{y=x}\not=0$ for all $j=1,\ldots,l$ and all $x\in\Omega$; - $ {\rm rank}(\nabla_{y}\widetilde{q}_{1}(x,y), \ldots, \nabla_{y}\widetilde{q}_{l}(x,y))|_{y=x}=n $ for all $x\in\Omega$; - the diagonal in $\Omega\times\Omega$ is the only set when all of $\widetilde{q}_j$’s vanish: $$\bigcap_{j=1}^l \left\{(x,y)\in\Omega\times\Omega: \, \widetilde{q}_j(x,y)=0\right\}=\{(x,x):\, x\in\Omega\}.$$ We also write $$\widetilde{q}^{\alpha}(x,y):=\widetilde{q}^{\alpha_1}_{1}(x,y)\cdots \widetilde{q}^{\alpha_l}_{l}(x,y).$$ For an operator $A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow\mathcal D'_{{\rm L}}(\Omega)$ with Schwartz kernel $K_{A}$, let us define $A_{q^{\alpha}}:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow\mathcal D'_{{\rm L}}(\Omega)$ as an operator with the kernel $$q^{\alpha}(x,y)K_{A}(x,y),$$ We understand this formula in the sense of distributions, i.e. $$\langle q^{\alpha} K_{A},{\varphi}\rangle := \langle K_{A}, q^{\alpha} \varphi \rangle \quad (K_{A}\in \mathcal D'_{{\rm L}}(\Omega\times\Omega),\; \varphi\in C_{{\rm L}^*}^{\infty}(\overline{\Omega}\times\overline{\Omega})).$$ Then, we have $$A_{q^{\alpha}}f(x)=\int_{\Omega}q^{\alpha}(x,y)K_{A}(x,y)f(y)dy.$$ Also analogously, for an operator $B:C^{\infty}_{{\rm L}^{*}}(\overline{\Omega})\rightarrow\mathcal D'_{{\rm L}^{*}}(\Omega)$ with Schwartz kernel $K_{B}$, we define $B_{\widetilde{q}^{\alpha}}:C^{\infty}_{{\rm L}^{*}}(\overline{\Omega})\rightarrow\mathcal D'_{{\rm L}^{*}}(\Omega)$ as an operator with the kernel $$\widetilde{q}^{\alpha}(x,y)K_{B}(x,y).$$ We understand this formula in the sense of distributions, i.e. $$\langle \widetilde{q}^{\alpha} K_{B},{\varphi}\rangle := \langle K_{B}, \widetilde{q}^{\alpha} \varphi \rangle \quad (K_{B}\in \mathcal D'_{{\rm L}^*}(\Omega\times\Omega),\; \varphi\in C_{{\rm L}}^{\infty}(\overline{\Omega}\times\overline{\Omega})).$$ Then, we get $$B_{\widetilde{q}^{\alpha}}f(x)=\int_{\Omega}\widetilde{q}^{\alpha}(x,y)K_{B}(x,y)f(y)dy.$$ \[DEF: DifferenceOper\]\[DEF: DifferenceOper\_2\] Let $$A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$$ be an ${\rm L}$–admissible operator with the symbol $a\in\mathcal D'_{{\rm L}}(\Omega)\otimes\mathcal S'(\ind)$ and with the Schwartz kernel $K_{A}\in\mathcal D'_{{\rm L}}(\Omega\times\Omega)$. Then we define the difference operator $$\Delta_{q}^{\alpha}:\mathcal D'_{{\rm L}}(\Omega)\otimes\mathcal S'(\ind)\rightarrow\mathcal D'_{{\rm L}}(\Omega)\otimes\mathcal S'(\ind)$$ acting on ${\rm L}$–symbols by $$\begin{aligned} \Delta_{q}^{\alpha}\,a(x,\xi) & := u_{\xi}^{-1}(x) \int_{\Omega} K_{A_{q^{\alpha}}}(x,y)u_{\xi}(y)dy {}\\ & = u_{\xi}^{-1}(x) \int_{\Omega} q^{\alpha}(x,y)K_{A}(x,y)u_{\xi}(y)dy,\end{aligned}$$ where $K_{A_{q^{\alpha}}}\in\mathcal D'_{{\rm L}}(\Omega\times\Omega)$ is the Schwartz kernel of the ${\rm L}$–admissible operator $A_{q^{\alpha}}:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega).$ Analogously, for the ${\rm L^{\ast}}$–admissible operator $$B:C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L^{\ast}}}(\Omega)$$ with the symbol $b\in\mathcal D'_{{\rm L^{\ast}}}(\Omega)\otimes\mathcal S'(\ind)$ and with the Schwartz kernel $\widetilde{K}_{B}\in\mathcal D'_{{\rm L^{\ast}}}(\Omega\times\Omega)$ we define the difference operator $$\widetilde{\Delta}_{q}^{\alpha}:\mathcal D'_{{\rm L^{\ast}}}(\Omega)\otimes\mathcal S'(\ind)\rightarrow\mathcal D'_{{\rm L^{\ast}}}(\Omega)\otimes\mathcal S'(\ind)$$ acting on ${\rm L^{\ast}}$–symbols by $$\begin{aligned} \widetilde{\Delta}_{q}^{\alpha}\,b(x,\xi) & := v_{\xi}^{-1}(x) \int_{\Omega} \widetilde{K}_{B_{q^{\alpha}}}(x,y)v_{\xi}(y)dy {}\\ & = v_{\xi}^{-1}(x) \int_{\Omega} \widetilde{q}^{\alpha}(x,y)\widetilde{K}_{B}(x,y)v_{\xi}(y)dy,\end{aligned}$$ where $K_{B_{q^{\alpha}}}\in\mathcal D'_{{\rm L^{\ast}}}(\Omega\times\Omega)$ is the Schwartz kernel of the ${\rm L^{\ast}}$–admissible operator $B_{q^{\alpha}}:C^{\infty}_{{\rm L^{\ast}}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L^{\ast}}}(\Omega).$ We now record the Taylor expansion formula with respect to a family of $q_j$’s, which follows from expansions of functions $g$ and $q^{\alpha}(e,\cdot)$ by the common Taylor series: \[TaylorExp\] Any smooth function $g\in C^{\infty}({\Omega})$ can be approximated by Taylor polynomial type expansions, i.e. for any $e\in\Omega$, we have $$g(x)=\sum_{|\alpha|< N}\frac{1}{\alpha!}D^{(\alpha)}_{x}g(x)|_{x=e}\, q^{\alpha}(e,x)+\sum_{|\alpha|= N}\frac{1}{\alpha!}q^{\alpha}(e,x)g_{N}(x)$$ $$\sim\sum_{\alpha\geq 0}\frac{1}{\alpha!}D^{(\alpha)}_{x}g(x)|_{x=e}\, q^{\alpha}(e,x) \label{TaylorExpFormula}$$ in a neighborhood of $e\in\Omega$, where $g_{N}\in C^{\infty}({\Omega})$ and $D^{(\alpha)}_{x}g(x)|_{x=e}$ can be found from the recurrent formulae: $D^{(0,\cdots,0)}_{x}:=I$ and for $\alpha\in\mathbb N_0^l$, $$\mathsf \partial^{\beta}_{x}g(x)|_{x=e}=\sum_{|\alpha|\leq|\beta|}\frac{1}{\alpha!} \left[\mathsf \partial^{\beta}_{x}q^{\alpha}(e,x)\right]\Big|_{x=e}D^{(\alpha)}_{x}g(x)|_{x=e},$$ where $\beta=(\beta_1, \ldots, \beta_n)$ and $ \partial^{\beta}_{x}=\frac{\partial^{\beta_{1}}}{\partial x_{1}^{\beta_{1}}}\cdots \frac{\partial^{\beta_{n}}}{\partial x_{n}^{\beta_{n}}}. $ Analogously, any function $g\in C^{\infty}({\Omega})$ can be approximated by Taylor polynomial type expansions corresponding to the adjoint problem, i.e. we have $$g(x)=\sum_{|\alpha|< N}\frac{1}{\alpha!}\widetilde{D}^{(\alpha)}_{x}g(x)|_{x=e}\, \widetilde{q}^{\alpha}(e,x)+\sum_{|\alpha|= N}\frac{1}{\alpha!}\widetilde{q}^{\alpha}(e,x)g_{N}(x)$$ $$\sim\sum_{\alpha\geq 0}\frac{1}{\alpha!}\widetilde{D}^{(\alpha)}_{x}g(x)|_{x=e}\, \widetilde{q}^{\alpha}(e,x) \label{TaylorExpFormula}$$ in a neighborhood of $e\in\Omega$, where $g_{N}\in C^{\infty}({\Omega})$ and $\widetilde{D}^{(\alpha)}_{x}g(x)|_{x=e}$ are found from the recurrent formula: $\widetilde{D}^{(0,\cdots,0)}:=I$ and for $\alpha\in\mathbb N_{0}^{l}$, $$\partial^{\beta}_{x}g(x)|_{x=e}=\sum_{|\alpha|\leq|\beta|}\frac{1}{\alpha!} \left[ \partial^{k}_{x}\widetilde{q}^{\alpha}(e,x)\right]\Big|_{x=e}\widetilde{D}^{(\alpha)}_{x}g(x)|_{x=e},$$ where $\beta=(\beta_1, \ldots, \beta_n)$, and $\partial^{\beta}$ is defined as in Proposition \[TaylorExp\]. It can be seen that operators $D^{(\alpha)}$ and $\widetilde{D}^{(\alpha)}$ are differential operators of order $|\alpha|$. We will understand them in distributions sense, i.e. for the L–admissible (${\rm L}^{\ast}$–admissible) operator $A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$ ($B:C^{\infty}_{{\rm L}^{\ast}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}^{\ast}}(\Omega)$) define the operator $D^{(\alpha)}A$ ($\widetilde{D}^{(\alpha)}B$) as an operator with the Schwartz kernel $D^{(\alpha)}_{x}K_{A}(x,y)$ ($\widetilde{D}^{(\alpha)}_{x}K_{B}(x,y)$). Then we can act on L–symbols (${\rm L}^{\ast}$–symbols) by $D^{(\alpha)}$ ($\widetilde{D}^{(\alpha)}$). Symbolic calculus {#SEC:differences2} ================= Using such difference operators and derivatives $D^{(\alpha)}$ from Proposition \[TaylorExp\] we can now define classes of symbols. \[DEF: SymClass\] Let $m\in\mathbb R$, $0\leq\delta,\rho\leq 1$. Let $$A:C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$$ be an ${\rm L}$–admissible operator with the symbol $a\in\mathcal D'_{{\rm L}}(\Omega)\otimes\mathcal S'(\ind)$ and with the Schwartz kernel $K_{A}\in\mathcal D'_{{\rm L}}(\Omega\times\Omega)$. Then the ${\rm L}$-symbol class $S^m_{\rho,\delta}(\overline{\Omega}\times\ind)$ consists of such symbols $a(x,\xi)$ which are smooth in $x$ for all $\xi\in\ind$, and which satisfy $$\label{EQ:symbol-class} \left|\Delta_{q}^\alpha D^{(\beta)}_{x} a(x,\xi) \right| \leq C_{a\alpha\beta m} \ \langle\xi\rangle^{m-\rho|\alpha|+\delta|\beta|}$$ for all $x\in\overline{\Omega}$, for all $\alpha,\beta\geq 0$, and for all $\xi\in\ind$. Here we understand $D^{(\beta)}_{x} a(x,\xi)$ as the symbol of the operator $D^{(\beta)}_{x}A$, where the operators $D^{(\beta)}_{x}$ are defined in Proposition \[TaylorExp\]. We will often denote them simply by $D^{(\beta)}$. The class $S^m_{1,0}(\overline{\Omega}\times\ind)$ will be often denoted by writing simply $S^m(\overline{\Omega}\times\ind)$. In [(\[EQ:symbol-class\])]{}, we assume that the inequality is satisfied for $x\in\Omega$ and it extends to the closure $\overline\Omega$. Furthermore, we define $$S^{\infty}_{\rho,\delta}(\overline{\Omega}\times\ind):=\bigcup\limits_{m\in\mathbb R}S^{m}_{\rho,\delta}(\overline{\Omega}\times\ind)$$ and $$S^{-\infty}(\overline{\Omega}\times\ind):=\bigcap\limits_{m\in\mathbb R}S^{m}(\overline{\Omega}\times\ind).$$ When we have two L-strongly admissible collections, expressing one in terms of the other similarly to Proposition \[TaylorExp\] and arguing similarly to [@Ruzhansky-Turunen-Wirth:JFAA], we can convince ourselves that for $\rho>\delta$ the definition of the symbol class does not depend on the choice of an L-strongly admissible collection. Analogously, we define for the ${\rm L}^*$–admissible operator $$B:C^{\infty}_{{\rm L}^*}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}^*}(\Omega)$$ with the symbol $b\in\mathcal D'_{{\rm L}^*}(\Omega)\otimes\mathcal S'(\ind)$ and with the Schwartz kernel $\widetilde{K}_{B}\in\mathcal D'_{{\rm L}^*}(\Omega\times\Omega)$ the ${\rm L^{\ast}}$-symbol class $\widetilde{S}^m_{\rho,\delta}(\overline{\Omega}\times\ind)$ as the space of such symbols $b(x,\xi)$ which are smooth in $x$ for all $\xi\in\ind$, and which satisfy $$\left|\widetilde{\Delta}_{(x)}^\alpha \widetilde{D}^{(\beta)} b(x,\xi) \right| \leq C_{a\alpha\beta m} \ \langle\xi\rangle^{m-\rho|\alpha|+\delta|\beta|}$$ for all $x\in\overline{\Omega}$, for all $\alpha,\beta\geq 0$, and for all $\xi\in\ind$. Here we understand $\widetilde{D}^{(\beta)} b(x,\xi)$ as the symbol of the operator $\widetilde{D}^{(\beta)} B$. Similarly, we can define classes $\widetilde{S}^{\infty}_{\rho,\delta}(\overline{\Omega}\times\ind)$ and $\widetilde{S}^{-\infty}(\overline{\Omega}\times\ind)$. If $a\in S^m_{\rho,\delta}(\overline{\Omega}\times\ind)$, it is convenient to denote by $a(X,D)={\rm Op_L}(a)$ the corresponding ${\rm L}$-pseudo-differential operator defined by $$\label{EQ: L-tor-pseudo-def} {\rm Op_L}(a)f(x)=a(X,D)f(x):=\sum_{\xi\in\ind} u_{\xi}(x)\ a(x,\xi)\widehat{f}(\xi).$$ The set of operators ${\rm Op_L}(a)$ of the form (\[EQ: L-tor-pseudo-def\]) with $a\in S^m_{\rho,\delta}(\overline{\Omega}\times\ind)$ will be denoted by ${\rm Op_L}(S^m_{\rho,\delta} (\overline{\Omega}\times\ind))$, or by $\Psi^m_{\rho,\delta} (\overline{\Omega}\times\ind)$. If an operator $A$ satisfies $A\in{\rm Op_L}(S^m_{\rho,\delta}(\overline{\Omega}\times\ind))$, we denote its ${\rm L}$-symbol by $\sigma_{A}=\sigma_{A}(x, \xi), \,\, x\in\overline{\Omega}, \, \xi\in\ind$. Naturally, $\sigma_{a(X,D)}(x,\xi)=a(x,\xi)$. Analogously, if $a\in \widetilde{S}^m_{\rho,\delta}(\overline{\Omega}\times\ind)$, we denote by $a(X,D)={\rm Op_{L^*}}(a)$ the corresponding ${\rm L^{\ast}}$-pseudo-differential operator defined by $$\label{EQ: L-tor-pseudo-def_2} {\rm Op_{L^*}}(a)f(x)=a(X,D)f(x):=\sum_{\xi\in\ind} v_{\xi}(x)\ a(x,\xi)\widehat{f}_{\ast}(\xi).$$ The set of operators ${\rm Op_{L^*}}(a)$ of the form (\[EQ: L-tor-pseudo-def\_2\]) with $a\in \widetilde{S}^m_{\rho,\delta}(\overline{\Omega}\times\ind)$ will be denoted by ${\rm Op_{L^*}}(\widetilde{S}^m_{\rho,\delta} (\overline{\Omega}\times\ind))$, or by $\widetilde{\Psi}^m_{\rho,\delta} (\overline{\Omega}\times\ind)$. \[REM: Topology of SymClass\] [(Topology on $S^{m}_{\rho, \delta}(\overline{\Omega}\times\ind)$ ($\widetilde{S}^{m}_{\rho, \delta}(\overline{\Omega}\times\ind)$)).]{} The set $S^{m}_{\rho, \delta}(\overline{\Omega}\times\ind)$ ($\widetilde{S}^{m}_{\rho, \delta}(\overline{\Omega}\times\ind)$) of symbols has a natural topology. Let us consider the functions $p_{\alpha\beta}^{l}: S^{m}_{\rho, \delta}(\overline{\Omega}\times\ind)\rightarrow\mathbb R$ ($\widetilde{p}_{\alpha\beta}^{l}: \widetilde{S}^{m}_{\rho, \delta}(\overline{\Omega}\times\ind)\rightarrow\mathbb R$) defined by $$p_{\alpha\beta}^{l}(\sigma):={\rm sup}\left[\frac{\left|\Delta_{(x)}^{\alpha}D^{(\beta)}\sigma(x, \xi)\right|}{\langle\xi\rangle^{l-\rho|\alpha|+\delta|\beta|}}:\,\, (x, \xi)\in\overline{\Omega}\times\ind\right]$$ $$\left(\widetilde{p}_{\alpha\beta}^{l}(\sigma):={\rm sup}\left[\frac{\left|\widetilde{\Delta}_{(x)}^{\alpha}\widetilde{D}^{(\beta)}\sigma(x, \xi)\right|}{\langle\xi\rangle^{l-\rho|\alpha|+\delta|\beta|}}:\,\, (x, \xi)\in\overline{\Omega}\times\ind\right]\right).$$ Now $\{p_{\alpha\beta}^{l}\}$ ($\{\widetilde{p}_{\alpha\beta}^{l}\}$) is a countable family of seminorms, and they define a Fréchet topology on $S^{m}_{\rho, \delta}(\overline{\Omega}\times\ind)$ ($\widetilde{S}^{m}_{\rho, \delta}(\overline{\Omega}\times\mathbb Z)$). Due to the bijective correspondence of ${\rm Op_L}(S^{m}_{\rho, \delta}(\overline{\Omega}\times\ind))$ and $S^{m}_{\rho, \delta}(\overline{\Omega}\times\ind)$ (${\rm Op_{L^*}}(\widetilde{S}^{m}_{\rho, \delta}(\overline{\Omega}\times\ind))$ and $\widetilde{S}^{m}_{\rho, \delta}(\overline{\Omega}\times\mathbb Z)$), this directly topologises also the set of operators. These spaces are not normable, and the topologies have but a marginal role. The notion of a symbol can be naturally extended to that of an amplitude. \[DEF: Amplitude\] The class $\mathcal A^m_{\rho,\delta}(\overline{\Omega})$ of ${\rm L}$-amplitudes consists of the functions $a(x,y,\xi)$ which are smooth in $x$ and $y$ for all $\xi\in\ind$, and $a(x,x,\xi)$ is an L–symbol for some L–admissible operator and which satisfy $$\left|\Delta_{(x)}^\alpha \Delta_{(y)}^{\alpha'} D^{(\beta)}_{x} D^{(\gamma)}_{y} a(x,y,\xi) \right| \leq C_{a\alpha\alpha'\beta\gamma m} \ \langle\xi\rangle^{m-\rho(|\alpha|+|\alpha'|)+\delta(|\beta|+|\gamma|)}$$ for all $x,y\in\overline{\Omega}$, for all $\alpha,\alpha', \beta,\gamma\geq 0$, and for all $\xi\in\ind$. Such a function $a$ will be also called an ${\rm L}$-amplitude of order $m\in\mathbb R$ of type $(\rho,\delta)$. Formally we may also define $$({\rm Op_L}(a)f)(x) := \sum_{\xi\in\ind} \int_{\Omega} u_{\xi}(x)\ \overline{v_{\xi}(y)}\ a(x,y,\xi)\ f(y)\ dy$$ for $f\in C_{{\rm L}}^\infty(\overline{\Omega})$. Sometimes we may denote ${\rm Op_L}(a)$ by $a(X,Y,D).$ We also write $\mathcal A^{m}(\overline{\Omega}):=\mathcal A^{m}_{1, 0}(\overline{\Omega})$ as well as $$\mathcal A^{-\infty}(\overline{\Omega}):=\bigcap\limits_{m\in\mathbb R}\mathcal A^{m}(\overline{\Omega}) \,\,\,\, \hbox{and} \,\,\,\, \mathcal A^{\infty}_{\rho,\delta}(\overline{\Omega}):=\bigcup\limits_{m\in\mathbb R}\mathcal A^{m}_{\rho,\delta}(\overline{\Omega}).$$ Clearly we can regard the ${\rm L}$-symbols as a special class of ${\rm L}$-amplitudes, namely the ones independent of the middle argument. Analogously, the class $\widetilde{\mathcal A}^m_{\rho,\delta}(\overline{\Omega})$ of ${\rm L^{\ast}}$-amplitudes consists of the functions $a(x,y,\xi)$ which are smooth in $x$ and $y$ for all $\xi\in\ind$, and $a(x,x,\xi)$ is an ${\rm L}^{\ast}$–symbol for some ${\rm L}^{\ast}$–admissible operator and which satisfy $$\left|\widetilde{\Delta}_{(x)}^\alpha \widetilde{\Delta}_{(y)}^{\alpha'} \widetilde{D}^{(\beta)}_{x} \widetilde{D}^{(\gamma)}_{y} a(x,y,\xi) \right| \leq C_{a\alpha\beta\gamma m} \ \langle\xi\rangle^{m-\rho(|\alpha|+|\alpha'|)+\delta(|\beta|+|\gamma|)}$$ for all $x,y\in\overline{\Omega}$, for all $\alpha, \alpha', \beta,\gamma\geq 0$, and for all $\xi\in\ind$. Formally we may also write $$({\rm Op_{L^*}}(a)f)(x) := \sum_{\xi\in\ind} \int_{\Omega} v_{\xi}(x)\ \overline{u_{\xi}(y)}\ a(x,y,\xi)\ f(y)\ dy$$ for $f\in C_{{\rm L^{\ast}}}^\infty(\overline{\Omega})$. We also write $\widetilde{\mathcal A}^{m}(\overline{\Omega}):=\widetilde{\mathcal A}^{m}_{1, 0}(\overline{\Omega})$ as well as $ \widetilde{\mathcal A}^{-\infty}(\overline{\Omega}):=\bigcap\limits_{m\in\mathbb R}\widetilde{\mathcal A}^{m}(\overline{\Omega})$ and $\widetilde{\mathcal A}^{\infty}_{\rho,\delta}(\overline{\Omega}):=\bigcup\limits_{m\in\mathbb R}\widetilde{\mathcal A}^{m}_{\rho,\delta}(\overline{\Omega}). $ \[DEF: EquivAmplit\] We say that amplitudes $a, a'$ are $m(\rho, \delta)$-equivalent $(m\in\mathbb R)$, $a\stackrel{m,\rho,\delta}{\sim} a'$, if $a-a'\in\mathcal A^{m}_{\rho,\delta}(\overline{\Omega})$; they are asymptotically equivalent, $a\sim a'$ (or $a\stackrel{-\infty}{\sim} a'$ if we need additional clarity), if $a-a'\in\mathcal A^{-\infty}(\overline{\Omega})$. For the corresponding operators we also write ${\rm Op}(a)\stackrel{m,\rho,\delta}{\sim} {\rm Op}(a')$ and ${\rm Op}(a)\sim {\rm Op}(a')$ (or ${\rm Op}(a)\stackrel{-\infty}{\sim} {\rm Op}(a')$ if we need additional clarity), respectively. It is obvious that $\stackrel{m,\rho,\delta}{\sim}$ and $\sim$ are equivalence relations. From the algebraic point of view, we could handle the amplitudes, symbols, and operators modulo the equivalence relation $\sim$, because the ${\rm L}$-pseudo-differential operators form a $\ast$-algebra with ${\rm Op}(S^{-\infty}(\overline{\Omega}\times\ind))$ as a subalgebra. The next theorem is a prelude to asymptotic expansions, which are the main tool in the symbolic analysis of ${\rm L}$-pseudo-differential operators. Let $(m_{j})_{j=0}^{\infty}\subset\mathbb R$ be a sequence such that $m_{j}>m_{j+1}$, and $m_{j}\rightarrow-\infty$ as $j\rightarrow\infty$, and $\sigma_{j}\in S^{m_{j}}_{\rho,\delta}(\overline{\Omega}\times\ind)$ for all $j\in\ind$. Then there exists an ${\rm L}$-symbol $\sigma\in S^{m_{0}}_{\rho,\delta}(\overline{\Omega}\times\ind)$ such that for all $N\in\ind$, $$\sigma\stackrel{m_{N},\rho,\delta}{\sim}\sum_{j=0}^{N-1}\sigma_{j}.$$ We will now look at the formula for the symbol of the adjoint operator. Let $A\in {\rm Op_L} (S^m_{\rho,\delta}(\overline{\Omega}\times\ind))$. By the definition of the adjoint operator we have $$(Au_{\xi}, v_{\eta})_{L^2}=(u_{\xi}, A^{*}v_{\eta})_{L^2}$$ or $$\int_{\Omega}Au_{\xi}(x)\overline{v_{\eta}(x)}dx=\int_{\Omega}u_{\xi}(x)\overline{A^{\ast}v_{\eta}(x)}dx$$ for $\xi, \eta\in\ind$. Plugging in the integral expressions, we get $$\begin{aligned} \int_{\Omega}{\left[\int_{\Omega}K_{A}(x,y)u_{\xi}(y)dy\right]}\overline{v_{\eta}(x)}dx & = \int_{\Omega}{u_{\xi}(x)}\overline{\left[\int_{\Omega}K_{A^{\ast}}(x,y)v_{\eta}(y)dy\right]}dx \\ & = \int_{\Omega}{u_{\xi}(y)}{\left[\int_{\Omega}\overline{K_{A^{\ast}}(y,x)} \overline{v_{\eta}(x)}dx\right]}dy\end{aligned}$$ for $\xi, \eta\in\ind$, where we swapped $x$ and $y$ in the last formula. Consequently, we get the familiar property $$K_{A^{\ast}}(x,y)=\overline{K_{A}(y,x)}.$$ Now, using this and the equation (\[EQ:KofA\]), and formula (ii) in Corollary \[COR: SymFor2\], and then formula (ii) in Corollary \[COR: SymFor\] and the Taylor expansion in Proposition \[TaylorExp\], we can write for the ${\rm L}^*$-symbol $\tau_{A^*}$ of $A^*$ that $$\begin{aligned} \tau_{A^*}(x,\xi) & = v_{\xi}^{-1}(x) \int_\Omega K_{A^{\ast}}(x,y)v_\xi(y) dy \\ & = v_{\xi}^{-1}(x) \int_\Omega \overline{K_{A}(y,x)} v_\xi(y) dy \\ & = v_{\xi}^{-1}(x)\int_\Omega \sum_{\eta\in\ind} \overline{u_\eta(y) \sigma_A(y,\eta)} v_\eta(x) v_\xi(y) dy \\ & \sim v_{\xi}^{-1}(x) \int_\Omega \sum_{\eta\in\ind} \overline{u_\eta(y)} \sum_\alpha \frac{1}{\alpha!} \overline {D_x^{(\alpha)} \sigma_A(x,\eta) q^\alpha(x,y)} v_\eta(x) v_\xi(y) dy\end{aligned}$$ as an asymptotic sum. Formally regrouping terms for each $\alpha$, we obtain $$\tau_{A^*}(x,\xi) \sim \sum_\alpha \frac{1}{\alpha!} v_{\xi}^{-1}(x) \int_\Omega\sum_{\eta\in\ind} \overline {u_\eta(y) q^\alpha(x,y) D_x^{(\alpha)} \sigma_A(x,\eta)} v_\eta(x) v_\xi(y) dy.$$ Using the formula (\[EQ:KofA2\]), and taking $$\widetilde{q}(x,y):=\overline{q(x,y)}$$ we can write this as $$\tau_{A^*}(x,\xi) \sim \sum_\alpha \frac{1}{\alpha!} \widetilde \Delta_{\widetilde{q}}^\alpha D_x^{(\alpha)}\overline{\sigma_A(x,\xi)}.$$ Making rigorous estimates for the remainder in a routine way, and assuming in the following theorem that for every $x\in\Omega$, the multiplication by $q_{j}(x,\cdot)$ preserves both spaces $C_{{\rm L}}^\infty(\overline{\Omega})$ and $C_{{\rm L}^{*}}^\infty(\overline{\Omega})$, we obtained: Let $0\leq\delta<\rho\leq 1$. Let $A\in {\rm Op_L} (S^m_{\rho,\delta}(\overline{\Omega}\times\ind))$. Assume that the conjugate symbol class $\widetilde{S}^{m}_{\rho,\delta}(\overline{\Omega}\times\ind)$ is defined with strongly admissible functions $\widetilde{q}_{j}(x,y):=\overline{q_{j}(x,y)}$ which are ${\rm L}^{*}$-strongly admissible. Then the adjoint of $A$ satisfies $A^{\ast}\in {\rm Op_{L^*}}(\widetilde{S}^{m}_{\rho,\delta}(\overline{\Omega}\times\ind))$, with its ${\rm L}^*$-symbol $\tau_{A^*}\in \widetilde{S}^{m}_{\rho,\delta}(\overline{\Omega}\times\ind)$ having the asymptotic expansion $$\tau_{A^*}(x,\xi) \sim \sum_\alpha \frac{1}{\alpha!} \widetilde \Delta_x^\alpha D_x^{(\alpha)}\overline{\sigma_A(x,\xi)}.$$ We now treat symbols of the amplitude operators. Let $0\leq\delta<\rho\leq 1$ and let $a\in \mathcal A^m_{\rho,\delta}(\overline{\Omega})$ be such that ${\rm Op_L}(a)$ is $L$-admissible. Then there exists a unique [L]{}-symbol $\sigma\in S^m_{\rho,\delta}(\overline{\Omega}\times\ind)$ satisfying ${\rm Op_L}(a)={\rm Op_L}(\sigma)$, where $$\sigma(x,\xi) \sim \sum_{\alpha\geq 0 } \frac{1}{\alpha!} \ \Delta_{(x)}^{\alpha} \ D_y^{(\alpha)} a(x,y,\xi)|_{y=x}.$$ As a linear operator on $C_{{\rm L}}^{\infty}(\overline{\Omega})$, the operator ${\rm Op_L}(a)$ possesses the unique L-symbol $\sigma=\sigma_{{\rm Op_L}(a)}$, but at the moment we do not yet know whether $\sigma\in S^m_{\rho,\delta}(\overline{\Omega}\times\ind)$. By Theorem \[QuanOper\] the L-symbol is computed from $$\sigma(x,\xi)=u_{\xi}^{-1}(x)({\rm Op_L}(a)u_{\xi})(x) =u_{\xi}^{-1}(x)\sum_{\eta\in\ind} \int_{\Omega} u_{\eta}(x)\ \overline{v_{\eta}(y)}\ a(x,y,\eta)\ u_{\xi}(y) dy.$$ Now we approximate the function $a(x,\cdot,\eta)\in C^{\infty}(\Omega)$ by Taylor polynomial type expansions, by using Proposition \[TaylorExp\], we have $$\begin{aligned} \sigma(x,\xi)&\sim u_{\xi}^{-1}(x) \sum_{\alpha\geq 0}\frac{1}{\alpha!}\sum_{\eta\in\ind} \int_{\Omega} u_{\eta}(x)\ \overline{v_{\eta}(y)} q^{\alpha}(x,y) \big[ D^{(\alpha)}_{y}a(x,y,\eta)\big]_{y=x}\ u_{\xi}(y) dy \\ &\sim \sum_{\alpha\geq 0} \frac{1}{\alpha!} \ \Delta_{(x)}^\alpha \ D_y^{(\alpha)} a(x,y,\xi)|_{y=x},\end{aligned}$$ Omitting a routine verification of the properties of the remainder, this yields the statement. We now formulate the composition formula. \[Composition\] Let $m_{1}, m_{2}\in\mathbb R$ and $\rho>\delta\geq0$. Let $A, B:C_{{\rm L}}^{\infty}(\overline{\Omega})\rightarrow C_{{\rm L}}^{\infty}(\overline{\Omega})$ be linear continuous and L–admissible operators, and assume that their [L]{}-symbols satisfy $$\begin{aligned} |\Delta_{(x)}^{\alpha}\sigma_{A}(x,\xi)|&\leq C_{\alpha}\langle\xi\rangle^{m_{1}-\rho|\alpha|},\\ |D^{(\beta)}\sigma_{B}(x,\xi)|&\leq C_{\beta}\langle\xi\rangle^{m_{2}+\delta|\beta|},\end{aligned}$$ for all $\alpha,\beta\geq 0$, uniformly in $x\in\overline{\Omega}$ and $\xi\in\ind$. Then $$\sigma_{AB}(x,\xi)\sim\sum_{\alpha\geq 0} \frac{1}{\alpha!}(\Delta_{(x)}^{\alpha}\sigma_{A}(x,\xi))D^{(\alpha)}\sigma_{B}(x,\xi), \label{CompositionForm}$$ where the asymptotic expansion means that for every $N\in\mathbb N$ we have $$|\sigma_{AB}(x,\xi)-\sum_{|\alpha|<N}\frac{1}{\alpha!}(\Delta_{(x)}^{\alpha}\sigma_{A}(x,\xi))D^{(\alpha)}\sigma_{B}(x,\xi)|\leq C_{N}\langle\xi\rangle^{m_{1}+m_{2}-(\rho-\delta)N}.$$ First, by the Schwartz kernel theorem from Subsection \[SEC:Schwartz\], we have $$\begin{aligned} ABf(x)&=\int_{\Omega}K_{A}(x,y) (Bf)(y)dy \\ &=\int_{\Omega}K_{A}(x,y)\Big[\int_{\Omega}K_{B}(y,z)f(z)dz\Big]dy \\ &=\int_{\Omega}\int_{\Omega}K_{A}(x,y) K_{B}(y,z)f(z)dzdy.\end{aligned}$$ Hence $$\begin{aligned} \sigma_{AB}(x,\xi)&=u_{\xi}^{-1}(x)(A(Bu_{\xi}))(x) \\ &=u_{\xi}^{-1}(x)\int_{\Omega}K_{A}(x,y)\Big[\int_{\Omega} K_{B}(y,z)u_{\xi}(z)dz\Big]dy \\ &=u_{\xi}^{-1}(x)\int_{\Omega}K_{A}(x,y)u_{\xi}(y)\sigma_{B}(y,\xi)dy.\end{aligned}$$ Now we approximate the function $\sigma_{B}(\cdot,\xi)\in C_{{\rm L}}^{\infty}(\overline{\Omega})$ by Taylor polynomial type expansions. By using Proposition \[TaylorExp\], we get $$\begin{aligned} \sigma_{AB}(x,\xi)&\sim u_{\xi}^{-1}(x)\int_{\Omega}K_{A}(x,y) \Big[\sum_{\alpha\geq 0} \frac{1}{\alpha!}q^{\alpha}(x,y)D^{(\alpha)}_{x}\sigma_{B}(x,\xi)\Big]u_{\xi}(z)dy \\ &=\sum_{\alpha\geq 0} \frac{1}{\alpha!} \Big[ u_{\xi}^{-1}(x)\int_{\Omega} q^{\alpha}(x,y) K_{A}(x,y) u_{\xi}(y) dy \Big]D^{(\alpha)}_{x}\sigma_{B}(x,\xi)\end{aligned}$$ Using Definition \[DEF: DifferenceOper\], we have $$\begin{aligned} \sigma_{AB}(x,\xi)\sim \sum_{\alpha\geq 0} \frac{1}{\alpha!}(\Delta_{(x)}^{\alpha}\sigma_{A}(x,\xi))D^{(\alpha)}_{x}\sigma_{B}(x,\xi).\end{aligned}$$ Omitting a routine treatment of the remainder, this completes the proof. On further results ================== Properties of integral kernels {#SEC:kernels} ------------------------------ We now establish some properties of Schwartz kernels of pseudo-differential operators with symbols in the introduced Hörmander-type classes. In the following Theorem \[TH: KernelofPDO\], let us make the assumption on the growth of $L^\infty$-norms of the eigenfunctions $u_\xi$. Finding estimates for the norms $\|u_{\xi}\|_{L^{\infty}}$ in terms of the corresponding eigenvalues of L is a challenging problem even for self-adjoint operators L, see e.g. Sogge and Zelditch [@Sogge-Zelditch:max-ef-growth-Duke] and references therein. Thus, on tori or, more generally, on compact Lie groups, the eigenfunctions of the Laplacian can be chosen to be uniformly bounded. However, even for the Laplacian, on more general manifolds, such growth depends on the geometry of the manifold. We refer to [@Delgado-Ruzhansky:invariant Remark 8.9] for a more thorough discussion of this topic as well as for a list of relevant references. \[TH: KernelofPDO\] Let $\mu_0$ be a constant such that there is $C>0$ such that for all $\xi\in\ind$ we have $$\|u_{\xi}\|_{L^{\infty}}\leq C \langle\xi\rangle^{\mu_0}.$$ Let $a\in S^{\mu}_{\rho,\delta}(\overline{\Omega}\times\ind)$, $\rho>0$. Then the kernel $K(x,y)$ of the pseudo-differential operator ${\rm Op_L}a$ satisfies $$\label{EQ:ests-L0} ({\rm L}^{\ast}_{y})^{k}(q^{\alpha}(x,y)K(x,y))\in L^\infty,$$ for all $|\alpha|>(\mu+mk+2\mu_0+s_0)/\rho$ and $x\neq y$, where $m$ is the order from [(\[EQ:angle\])]{} and $s_0$ is the constant from Assumption \[Assumption\_4\]. If ${\rm L}$ is a differential operator it follows that $$\label{EQ:ests-L} |({\rm L}^{\ast}_{y})^{k}K(x,y)|\leq C_{Nk}|x-y|^{-N}$$ for any $N>(\mu+mk+2\mu_0+s_0)/\rho$ and $x\neq y$. By Corollary \[COR: SymFor2\] we have $$u_{\xi}(x)a(x,\xi)= \int_{\Omega}K(x,y)u_{\xi}(y)dy,$$ and from Definition \[DEF: DifferenceOper\] we get $$\begin{aligned} u_{\xi}(x)\Delta_{(x)}^{\alpha}a(x,\xi)&=\int_{\Omega}q^{\alpha}(x,y)K(x,y)u_{\xi}(y)dy,\end{aligned}$$ and also $$\begin{gathered} u_\xi(x)\lambda_\xi^k \Delta_{(x)}^{\alpha} a(x,\xi)=\int_{\Omega}q^{\alpha}(x,y)K(x,y)\lambda_{\xi}^{k}u_{\xi}(y)dy \\ = \int_{\Omega}q^{\alpha}(x,y)K(x,y){\rm L}_{y}^{k}u_{\xi}(y)dy =\int_{\Omega} ({\rm L}^{\ast}_{y})^{k}(q^{\alpha}(x,y)K(x,y))u_{\xi}(y)dy.\end{gathered}$$ This means that $$({\rm L}^{\ast}_{y})^{k}(q^{\alpha}(x,y)K(x,y))= {\mathcal F}_{\rm L}^{-1} (u_\xi(x)\lambda_\xi^k \Delta_{(x)}^{\alpha} a(x,\xi))(y).$$ Since it follows from assumptions that $$\lambda_{\xi}^{k}\Delta_{(x)}^{\alpha} a(x,\xi)\in S^{\mu+mk-\rho|\alpha|}(\overline{\Omega}\times\ind),$$ we have $$\lambda_{\xi}^{k} |\Delta_{(x)}^{\alpha}a(x,\xi)|\leq C\langle\xi\rangle^{\mu+mk-\rho|\alpha|}.$$ We recall now the norm $$\|a(x,\cdot)\|_{l^{1}({\rm L})}=\sum_{\xi\in\ind}| a(x,\xi)| \|u_{\xi}\|_{L^{\infty}(\Omega)}$$ from Subsection \[SEC:lp\]. It follows that $$\|u_{\xi}(x)\lambda_{\xi}^{k}\Delta_{(x)}^{\alpha}a(x,\xi)\|_{l^{1}({\rm L})}\leq C\sum_{\xi\in\ind}\langle\xi\rangle^{\mu+mk-\rho|\alpha|} \|u_{\xi}\|_{L^{\infty}(\Omega)}^2\leq C\sum_{\xi\in\ind}\langle\xi\rangle^{\mu+mk-\rho|\alpha|+2\mu_0}.$$ Consequently, if $$|\alpha|>(\mu+mk+2\mu_0+s_0)/\rho,$$ where $s_0$ is the constant from Assumption \[Assumption\_4\], we have that $u_{\xi}(x)\lambda_{\xi}^{k}\Delta_{(x)}^{\alpha}a(x,\xi)$ is in $l^{1}({\rm L})$ with respect to $\xi$, and hence $({\rm L}^{\ast}_{y})^{k}(q^{\alpha}(x,y)K(x,y))$ is in $L^\infty$ by the Hausdorff-Young inequality in Theorem \[TH: HY\]. Since ${\rm L}^{\ast}_{y}$ is a differential operator for differential operators L, in this case we also have $$q^\alpha(x,y)({\rm L}^{\ast}_{y})^{k}K(x,y)\in L^{\infty}(\Omega\times\Omega)$$ for such $\alpha$. By the properties of $q^\alpha$ it implies the statement of the theorem. In particular, if L is for example locally elliptic, [(\[EQ:ests-L\])]{} implies that for $x\neq y$, the kernel $K(x,y)$ is a smooth function. And, if $a\in S^{-\infty}(\overline{\Omega}\times\ind)$, then the integral kernel $K(x,y)$ of ${\rm Op_L}a$ is smooth in $x$ and $y$. The singular support of $w\in\mathcal D'_{{\rm L}}(\Omega)$ is defined as the complement of the set where $w$ coincides with a test function. Namely, $x\notin {\rm sing\,supp}\,\, w$ if there is an open neighbourhood $U$ of $x$ and a smooth function $f\in C_{{\rm L}}^{\infty}(\overline{\Omega})$ such that $w(\varphi)=f(\varphi)$ for all $\varphi\in C_{{\rm L}}^{\infty}(\overline{\Omega})$ with ${\rm supp} \,\varphi\subset U$. As an immediate consequence of Theorem \[TH: KernelofPDO\] we obtain the information on how the singular support is mapped by a pseudo-differential operator: \[COR: SingularSupp\] Let $\sigma_{A}\in S^{\mu}_{\rho,\delta}(\overline{\Omega}\times\ind)$, $1\geq \rho>\delta\geq 0$. Then for every $w\in\mathcal D'_{{\rm L}}(\Omega)$ we have $$\label{EQ: SingSupp} {\rm sing\,supp}\,\, Aw\subset {\rm sing\,supp}\,\, w.$$ For elliptic operators, in Corollary \[COR:SingularSupp-ell\] we state also the inverse inclusion. ${\rm L}$-elliptic pseudo–differential operators {#SEC:elliptic} ------------------------------------------------ In this subsection we discuss operators that are elliptic in the symbol classes generated by L. For such operators we can obtain parametrix and then also a-priori estimates by the properties of pseudo-differential operators in, for example, Sobolev spaces, once they are established in Section \[SEC:L2\], see Theorem \[L2-Bs and El-ty\]. Thus, from the asymptotic expansion for the composition of pseudo-differential operators, we get an expansion for a parametrix of an elliptic operator: \[El-ty\] Let $1\geq \rho>\delta\geq 0$. Let $\sigma_A\in S^\mu_{\rho,\delta}(\overline{\Omega}\times\ind)$ be elliptic in the sense that there exist constants $C_0>0$ and $N_0\in\mathbb N$ such that $$\label{elliptic} |\sigma_A(x,\xi)| \geq C_0 \langle\xi\rangle^\mu$$ for all $(x,\xi)\in\overline{\Omega}\times\ind$ for which $\xi\geq N_0$; this is equivalent to assuming that there exists $\sigma_B\in S^{-\mu}_{\rho,\delta}(\overline{\Omega}\times\ind)$ such that $I-BA,I-AB$ are in ${\rm Op_L} S^{-\infty}$. Let $$A \sim \sum_{j=0}^\infty A_j,$$ with $\sigma_{A_j}\in S^{\mu-(\rho-\delta)j}_{\rho,\delta}(\overline{\Omega}\times\ind)$. Then $$B \sim \sum_{k=0}^\infty B_k,$$ where $B_k\in S^{-\mu-(\rho-\delta)k}_{\rho,\delta}(\overline{\Omega}\times\ind)$ is such that $$\sigma_{B_0}(x,\xi)= 1/\sigma_{A_0}(x,\xi)$$ for large enough $\xi$, and recursively $$\sigma_{B_N}(x,\xi) = \frac{-1}{\sigma_{A_0}(x,\xi)} \sum_{k=0}^{N-1} \sum_{j=0}^{N-k} \sum_{|\alpha|=N-j-k} \frac{1}{\alpha!} \left[ \Delta_{(x)}^{\alpha} \sigma_{A_j}(x,\xi) \right] D_x^{(\alpha)} \sigma_{B_k}(x,\xi).$$ Theorem \[TH: KernelofPDO\] applied to the parametrix from in Theorem \[El-ty\], implies the inverse inclusion to the singular supports from Corollary \[COR: SingularSupp\] for elliptic operators: \[COR:SingularSupp-ell\] Let $1\geq \rho>\delta\geq 0$ and assume that $\sigma_{A}\in S^{\mu}_{\rho,\delta}(\overline{\Omega}\times\ind)$ is [L]{}-elliptic. Then for every $w\in\mathcal D'_{{\rm L}}(\Omega)$ we have $$\label{EQ: SingSupp} {\rm sing\,supp}\,\, Aw={\rm sing\,supp}\,\, w.$$ Sobolev embedding theorem {#SEC:embeddings} ------------------------- In this subsection we give an example of a Sobolev embedding theorem for Sobolev spaces $\mathcal H^s_{\rm L}$ associated to L, considered in Section \[SEC:Sobolev\]. However, only limited conclusions are possible in the abstract setting when no further specifics about L are available. Now, let $C({\Omega})$ be the Banach space under the norm $$\|f\|_{C({\Omega})}:=\sup\limits_{x\in{\Omega}} |f(x)|.$$ We recall that we have a differential operator L of order $m$ with smooth coefficients in the open set $\Omega\subset\mathbb R^n$, and also the operator ${\rm L}^\circ$ from [(\[EQ:Lo-def\])]{}. The following theorem is conditional to the local regularity estimate [(\[EQ:Sob-as\])]{}. It is satisfied with $\varkappa=1$ if, for example, L is locally elliptic, i.e. elliptic in the classical sense of $\mathbb R^n$. However, if L is for example a sum of squares satisfying Hörmander’s commutator condition, the number $\varkappa\geq 1$ may depend on the order to which the Hörmander condition is satisfied, see e.g. [@Garetto-Ruzhansky:sum-JDE] in the context of compact Lie groups. \[TH:SETh\] Let $k$ be an integer such that $k>n/2$. Let $\varkappa$ be such that the operators ${\rm L}$ and ${\rm L}^{\circ}$ satisfy the inequality $$\label{EQ:Sob-as} \Big\|\frac{\partial^{\alpha}f}{\partial x^{\alpha}}\Big\|_{L^{2}(\Omega)}\leq C\Big\|({\rm I}+{\rm L}^\circ{\rm L})^{\frac{\varkappa k}{2m}}f\Big\|_{L^{2}(\Omega)}$$ for all $f\in C^{\infty}({\Omega})$, for all $\alpha\in\mathbb N_0^{n}$ with $|\alpha|\leq k$. Then we have the continuous embedding $$\mathcal H^{\varkappa k}_{{\rm L}}(\Omega)\hookrightarrow C({\Omega}).$$ The proof is similar to [@Ruzhansky-Tokmagambetov:IMRN] so we omit it. Conditions for $L^{2}$-boundedness {#SEC:L2} ---------------------------------- In this subsection we will discuss what conditions on the ${\rm L}$-symbol $a$ guarantee the $L^{2}$-boundedness of the corresponding pseudo-differential operator ${\rm Op_L}(a):C^{\infty}_{{\rm L}}(\overline{\Omega})\rightarrow \mathcal D'_{{\rm L}}(\Omega)$. The proofs of the following results are similar to [@Ruzhansky-Tokmagambetov:IMRN] so we omit them. \[L2-Bs\] Let $k$ be an integer $>n/2$. Let $a:\overline{\Omega}\times\ind\rightarrow\mathbb C$ be such that $$\label{EQ:torus-L2-pse} \left| \partial^{\alpha}_{x} a(x,\xi) \right| \leq C \quad\textrm{ for all } (x,\xi)\in\overline{\Omega}\times\ind,$$ and all $|\alpha|\leq k$, all $x\in\Omega$ and $\xi\in\ind$. Then the operator ${\rm Op_L}(a)$ extends to a bounded operator from $L^{2}(\Omega)$ to $L^{2}(\Omega)$. From a suitable adaption of the composition Theorem \[Composition\], using that by Proposition \[TaylorExp\] the operators $\partial^{\alpha}_{x}$ and $D^{(\alpha)}_{x}$ can be expressed in terms of each other as linear combinations with smooth coefficients, we immediately obtain the result in Sobolev spaces: \[Hs-Bs\] Let $k$ be an integer $>n/2$. Let $\mu\in\mathbb R$ and let $a:\overline{\Omega}\times\mathbb Z\rightarrow\mathbb C$ be such that $$\label{Hs-cond} \left| \partial^{\alpha}_{x} a(x,\xi) \right| \leq C \langle\xi\rangle^{\mu} \quad\textrm{ for all } (x,\xi)\in\overline{\Omega}\times\ind,$$ and for all $\alpha$. Then operator ${\rm Op_L}(a)$ extends to a bounded operator from $\mathcal H^{s}_{L}(\Omega)$ to $\mathcal H^{s-\mu}_{L}(\Omega),$ for any $s\in\mathbb R.$ By using Theorem \[El-ty\] and Corollary \[Hs-Bs\], we get \[L2-Bs and El-ty\] Let $A$ be an [L]{}-elliptic pseudo-differential operator with [L]{}-symbol $\sigma_A\in S^{\mu}(\overline{\Omega}\times\ind)$, $\mu\in\mathbb R$, and let $Au=f$ in $\Omega$, $u\in \mathcal H^{\infty}_{{\rm L}}(\Omega)$. Then we have the estimate $$\|u\|_{\mathcal H^{s+\mu}_{{\rm L}}(\Omega)}\leq C_{sN} (\|f\|_{\mathcal H^{s}_{{\rm L}}(\Omega)}+ \|u\|_{\mathcal H^{-N}_{{\rm L}}(\Omega)}).$$ for any $s, N\in\mathbb R$. [RTW14]{} N. K. Bari. Biorthogonal systems and bases in [H]{}ilbert space. , 148(4):69–107, 1951. J. Bergh and J. L[[ö]{}]{}fstr[[ö]{}]{}m. . Springer-Verlag, Berlin, 1976. Grundlehren der Mathematischen Wissenschaften, No. 223. J. Delgado and M. Ruzhansky. Fourier multipliers, symbols and nuclearity on compact manifolds. , to appear in [*J. Anal. Math.*]{} J. Delgado, M. Ruzhansky, and N. Tokmagambetov. Schatten classes, nuclearity and nonharmonic analysis on compact manifolds with boundary. , to appear in [*J. Math. Pures Appl.*]{} C. Garetto and M. Ruzhansky. Wave equation for sums of squares on compact [L]{}ie groups. , 258(12):4324–4347, 2015. B. Kanguzhin and N. Tokmagambetov. The [F]{}ourier transform and convolutions generated by a differential operator with boundary condition on a segment. In [*Fourier Analysis: Trends in Mathematics*]{}, pages 235–251. Birkhäuser Basel AG, Basel, 2014. B. Kanguzhin, N. Tokmagambetov, and K. Tulenov. Pseudo-differential operators generated by a non-local boundary value problem. , 60(1):107–117, 2015. M. Ruzhansky and N. Tokmagambetov. Nonharmonic analysis of boundary value problems. , 2016 (12), 3548–3615, 2016. M. Ruzhansky and V. Turunen. On the [F]{}ourier analysis of operators on the torus. In [*Modern trends in pseudo-differential operators*]{}, volume 172 of [*Oper. Theory Adv. Appl.*]{}, pages 87–105. Birkh[ä]{}user, Basel, 2007. M. Ruzhansky and V. Turunen. On the toroidal quantization of periodic pseudo-differential operators. , 30(9-10):1098–1124, 2009. M. Ruzhansky and V. Turunen. , volume 2 of [*Pseudo-Differential Operators. Theory and Applications*]{}. Birkh[ä]{}user Verlag, Basel, 2010. M. Ruzhansky and V. Turunen. Quantization of pseudo-differential operators on the torus. , 16(6):943–982, 2010. M. Ruzhansky, V. Turunen, and J. Wirth. H[ö]{}rmander class of pseudo-differential operators on compact [L]{}ie groups and global hypoellipticity. , 20(3):476–499, 2014. C. D. Sogge and S. Zelditch. Riemannian manifolds with maximal eigenfunction growth. , 114(3):387–437, 2002. [^1]: The first author was supported in parts by the EPSRC grant EP/K039407/1 and by the Leverhulme Grant RPG-2014-02. The second author was supported in parts by the MESRK grant 0773/GF4. No new data was collected or generated during the course of this research.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study a deSitter/Anti-deSitter/Poincare Yang-Mills theory of gravity in d-space-time dimensions in an attempt to retain the best features of both general relativity and Yang-Mills theory: quadratic curvature, dimensionless coupling and background independence. We derive the equations of motion for Lie algebra valued scalars and show that in the geometric optics limit they traverse geodesics with respect to the Lorentzian geometry determined by the frame fields. Mixing between components appears to next to leading order in the WKB approximation. We then restrict to two space-time dimensions for simplicity, in which case the theory reduces to the well known Katanaev-Volovich model. We complete the Hamiltonian analysis of the vacuum theory and use it to prove a generalized Birkhoff theorem. There are two classes of solutions: with torsion and without torsion. The former are parametrized by two constants of motion, have event horizons for certain ranges of the parameters and a curvature singularity. The latter yield a unique solution, up to diffeomorphisms, that describes a space constant curvature .' author: - Jack Gegenberg - Gabor Kunstatter title: 'Two Dimensional Gravity as a modified Yang-Mills Theory' --- Introduction ============ General Relativity is often called a gauge theory of the gravitational field, but it is not a gauge theory of the Yang-Mills type. In the latter, the action functional $S_{YM}[A]$ is quadratic in the curvature of a connection $A$ of principle bundle over the spacetime manifold $(M_D,{\bf g})$, where ${\bf g}$ is a [*given*]{} non-dynamical Lorentzian metric on $M_D$: S\_[YM]{}\[A\]=\_[M\_D]{} d\^D x g\^g\^F\^A\_ F\^B\_ h\_[AB]{}. \[eq:action1\] In the above, $g_{YM}$ is the gauge coupling constant. It has dimension Length${}^{\frac{D}{2}-2}$, and hence is dimensionless in 4D; the $g^{\mu\nu}$ are the contravariant components of the metric tensor ${\bf g}$; the $F^A_{\mu\alpha}$ are the components of the curvature of the connection; the indices $A,B=1,2,...,n$ are in the adjoint representation of the gauge group; and finally $h_{AB}$ are components of the Cartan-Killing metric of the group. It is the fact that $g_{YM}^2$ is dimensionless in 4-D which permits Yang-Mills gauge theories to be perturbatively renormalizable By contrast, the Einstein-Hilbert action of General Relativity is linear in the curvature of the Christoffel connection of the Lorentzian metric. There is no background: the metric ${\bf g}$ is dynamical. Moreover, the coupling constant in Einstein gravity has dimension Length${}^{-2}$ in four spacetime dimensions. It is this which has stalled progress in constructing quantum gravity starting from Einstein’s theory. It was perhaps Townsend [@townsend] who first highlighted the fact that the gravitational constant has the dubious distinction of being the only dimensionful fundamental constant (the others being $\hbar$ and $c$) that is tied to a specific dynamical theory. He therefore suggested that the gravitational constant $G$ should somehow be linked directly to the structure of spacetime. This could be achieved by replacing the Poincare group as a potential local gauge symmetry of gravity by the deSitter group, which necessarily entails a dimensionful constant. With this as motivation, he proceeded to consider a Yang-Mills type Lagrangian for gravity with the deSitter group as gauge group. Besides those of Townsend, there have in fact been many attempts to construct a Yang-Mills type gravitational theory. The first was by Weyl almost one hundred years ago, and the goal was to unify gravity with electromagnetism [@weyl]. Work in the 70’s and 80’s, inspired by the work of Utiyama, Yang and Mills on non-Abelian gauge theories, constructed Yang-Mills type theories with gauge groups associated with gravity, for example the Poincare, DeSitter/anti DeSitter and Conformal groups [@early]. More recently, J. T. Wheeler[@wheeler] and collaborators have worked on 4D Yang-Mill gravity, with the conformal group SO(4,2) as the gauge group, while H.-Y. Guo[@guo] and his collaborators have tackled the de Sitter case. In the first order formalism of Einstein gravity- the so-called Einstein-Cartan action- the equations of motion force the torsion to be zero. In Yang-Mills gravity this does not happen. Generically, the spacetime geometry has non-vanishing torsion as well as (quasi-)Riemannian curvature [@early; @guo]. The consequences of this for the viability of such theories is still an open question. We note here the result of [@ymflow] that torsion de-stabilizes anti-de Sitter 2+1 dimensional spacetime. In spite of its quantum motivation, little progress has been made in quantizing Yang-Mills gravity. In fact, to date, there has been no canonical analysis of such theories, a necessary first step towards understanding the quantum theory. In this paper we begin to close this gap. After a general discussion that includes a discussion of the coupling to matter showing that, to leading order in the geometric optics limit, Higgs fields propagate along geodesics of the Lorentzian geometry, we will undertake the construction of the canonical form of a toy model of Yang-Mills gravity, wherein the spacetime is two dimensional, and the gauge group is the lineland version of de Sitter/anti-de Sitter/Poincare gravity, that is, SO(2,1)/SO(1,2)/ISO(1,1). In this case, the Lagrangian reduces to a special case of the Katanaev-Volkov model[@kv86], which was extensively studied in a somewhat different context in the 1990’s[@kv90s]. We will solve the Hamiltonian equations of motion for the vacuum case, finding in the case of zero torsion that the solutions are equivalent to those of Jackiw-Teitelboim dilaton gravity [@jt]. General Formalism ================= Algebra and Action ------------------ In this section we outline the general procedure for constructing a gauge theory of gravity in a $D$-dimensional spacetime. We note here the record of such attempts sampled in [@townsend; @weyl; @early; @wheeler; @guo]. The ‘kinematical’ gauge group associated with such a theory is one SO(D,1)/SO(D-1,2)/ISO(D-1,1), corresponding to positive/negative/zero cosmological constant. The generators $J_A=(J_a,F_{ab})$ with $ A=0,1,...,D;{}{} a,b=0,1,...D-1$ obey =-2\_[DD]{}J\_[ab]{}; =-(\_[ac]{}J\_b-\_[ab]{}J\_c); =-(\_[ac]{}J\_[bd]{}+\_[bd]{}J\_[ac]{}-\_[bc]{}J\_[ad]{}-\_[ad]{}J\_[bc]{}), where $\eta_{ab}$ is the $(D)$-dimensional Minkowski metric. If $\eta_{DD}=1,-1,0$, then the gauge group is, respectively, SO(D,1)/SO(D-1,2)/ISO(D-1,1). The gauge potential is decomposed according to A\_=e\^a\_J\_a +\_\^[ab]{} J\_[ab]{}, where the constant $\lambda$ has dimension $L^{-1}$, so that with the vielbein $e^a_\mu$ dimensionless, the spin-connection $\omega^{ab}_\mu$ and the gauge potential $A_\mu$ have dimension $L^{-1}$. The generator of translations, i.e. the 2-momentum, $P_a$ has dimension $L^{-1}$ and is related to the above via: P\_a = J\_a When working in terms of $P_a$, $\lambda$ appears directly in the commutator algebra as opposed to the definition of the gauge potential. It is for this reason that Townsend[@townsend] considered it to be a property of the spacetime structure, rather than a coupling constant.\ Note that although it would be natural to identify $\lambda$ with the dimensionally appropriate power of the $g_{YM}$, up to a dimensionless number of order unity, in order to keep things as general as possible we keep them distinct in what follows. We note here that the structure constants for any of the DS/ADS/Poincare gauge groups have structure constants f\^[\[cd\]]{}\_[ab]{}&=&-2\_[DD]{}\^[\[c]{}\_a\^[d\]]{}\_b;[\ ]{}f\^d\_[a\[bc\]]{}&=&-\^d\_[\[b]{}\_[c\]a]{};[\ ]{}f\^[ef]{}\_[\[ab\]\[cd\]]{}&=&-2\^[\[e]{}\_[\[a]{}\_[b\]\[d]{}\^[f\]]{}\_[c\]]{}. The Cartan-Killing metrics on the gauge groups, defined by $h_{ij}:=2 f^k{}_{il} f^l{}_{jk}$ are all of the form h\_[ab]{}=-2D\_[DD]{}\_[ab]{}; h\_[\[ab\]\[cd\]]{}=-D(\_[ac]{}\_[bd]{}-\_[bc]{}\_[ad]{}). The field strength is defined as usual: F=F\^[ij]{}\_dx\^dx\^:&=&dA+\[A,A\][\ ]{}&=& T\^a J\_a+\^[ab]{} J\_[ab]{}. Thus $[F]=L^{-2}$. The Lie algebra valued 2-forms $\Omega$ and $T$ are respectively the ‘curvature plus volume element’ and torsion of the spin connection $\omega$: \^[ab]{}&:=& d\^[ab]{}+\^[ac]{}\_c\^b-\^2\_[DD]{}e\^ae\^b;\[curv\]\ T\^a&:=&de\^a +\^a\_be\^b.\[torsion\] In the above, indices $a,b,...=0,1,2,D-2$ are raised and lowered by the Minkowski metric $\eta_{ab}$. and e.g. $F^a:=\frac{1}{2} F^a_{ij} dx^i\wedge dx^j$. Thus $\left[T^a_{\mu\nu}\right]=L^{-1}$ and $\left[\Omega^{ab}_{\mu\nu}\right]=L^{-2}$. Most importantly, the ‘background metric’, g\_:= \_[ab]{} e\^a\_e\^b\_, is not fixed, but is subject to the dynamics determined by the equations of motion for the gauge field. The action can be written explicitly in the form $S=S_{EH}+S_1$, where S\_[EH]{}:&=&d\^D x(R-\^2\_[ DD]{});\ S\_1:&=&-d\^D x(+ \^2\_[DD]{}T\_[a]{}T\^[a]{}), where $K:=R^{ab\mu\nu}R_{ab\mu\nu}$.\ Comparing the term $S_{EH}$ to the usual Einstein-Hilbert action, we find that the Newton gravitational constant in $D$ dimensions, $G_D$, (in units where the speed of light is one) is related to the gauge coupling constant $G$ and the scale factor $\lambda$ by G\_D=. In general $G_D$ has dimensions of $L^{D-2}$, so that it is dimensionless in 2 spacetime dimensions. We also remark again that $D=4$ is also special in that $g_{YM}$ is dimensionless. To close this section, we consider the issue of the background metric ${\bf g}$. In the following, as in most of the literature on Yang-Mill gravity, the ‘background’ $g_{\mu\nu}$ will not really be a background, but is rather, dynamical, via $g_{\mu\nu}:=\eta_{ab} e^a_\mu e^b_\nu$. One pays a price for this, however, in that the gauge transformations generated by the ‘translations’ $J_a$ no longer preserve the action: some of the gauge symmetry is broken. Adding Matter ------------- Torsion theories are distinguished by the fact that the field content describes more than one kind of geometry. There is curvature associated with the Riemannian metric used to raise and lower indices, and there is also the Riemann-Cartan connection and associated curvature. When the torsion is non-zero, the metric compatible with the Riemann-Cartan connection is not the same as the Riemannian metric constructed out of the fierbeins/vielbeins. The only way to decide which geometry is relevant in a particular physical context is to look at matter couplings. It is straightforward to add most forms of matter using the principle of minimal couple. Only spinors will couple directly to the torsion, whereas all other matter Lagrangians will just depend on $e^a{}_\mu$. Here we consider a Higgs-like scalar $\phi^A(x)$ that takes its values in the adjoint representation and couples to the vacuum action via $S=S_{YM}+S_{higgs}$, where S\_[higgs]{}=\_[M\_2]{} d\^2x g\^h\_[ij]{} D\_\^i D\_\^j. . Thus the matter field obeys the gauge covariant wave equation D\_D\^\^i=0. In the geometric optics limit, a wave field has approximately constant amplitude, but varying phase. Thus for a Higgs type of matter we write \^j=R\^j e\^[iS\^j/]{} Note that the Lie algebra index $j=0,1,2$ is not summed over here, or subsequently. Now the geometric optics limits has particles traveling with momenta $k^{(i)}_\mu=\partial_\mu S^i$ orthogonal to the constant surfaces $S^i(t,x)=const$. Also, $\nabla_\mu$ is the Lorentzian covariant derivative with respect to the background metric $g_{\mu\nu}=\eta_{ab}e^a_\mu e^b_\nu$. We assume that the amplitudes $R^j$ are slowly varying compared to the phase. The wave equation $D^\mu D_\mu \phi^j=0$ becomes, after dropping terms in $\partial_\mu R^j$ and keeping only terms of leading and subleading orders ($1/\hbar^2, 1/\hbar$, respectively) 0=-e\^[iS\^j/]{} R\^j k\^[(j)]{} k\^[(j)]{}\_+(e\^[i S\^j/]{}R\^j\^k\^[(j)]{}\_+ 2 e\^[iS\^l/]{}f\^j\_[kl]{}A\^[(k)]{} k\^[(l)]{}\_R\^l). Hence, to leading order C\^j:=-R\^j k\^[(j)]{} k\^[(j)]{}\_=0 \[eq:ray\] Note that this expression is real. The gauge covariant derivative of $C^j$ reduces to the partial derivative. Thus, to leading order $k^{j\nu}\nabla_\mu k^j_\nu=0$. Using the smoothness of the phase $S^j$ in order to change the order of partial differentiation we find that $k^{j\nu}\nabla_\nu k^j_\mu=0$. Thus to leading order the trajectories are null geodesics of the Lorentzian geometry compatible with the frame-field $e^i_\mu$ on spacetime. The subleading terms are pure imaginary terms, and hence e\^[i S\^j/]{}R\^j\^k\^[(j)]{}\_+2 e\^[iS\^l/]{}f\^j\_[kl]{}A\^[(k)]{} k\^[(l)]{}\_R\^l=0 The latter are more complicated because of the relative phase factor. 1+1 Dimensions: Action and Covariant Equations of Motion ======================================================== Things simplify quite a bit in 1+1 dimensions. The group is $SO(2,1),SO(1,2),ISO(1,1)$ respectively, for $k:=-\eta_{22}=-1,+1,0$ with generators $J_a,J$, and algebra: &=& 0\ &=& k \_[ab]{}J\ &=&\_a\^b J\_b Note that $J$ generates an Abelian one dimensional subalgebra. A = e\^a J\_a + J As before we split the curvature into F= J + T\^a J\_a where $\Omega:=d\omega+\frac{k}{2}\lambda^2\epsilon_{ab}e^a\wedge e^b$ and $T^a:=de^a-\epsilon^a{}_b\omega\wedge e^b$. That is: \_&=& \_\_- \_\_+ k \^2 V\_\ T\^a\_&=&\_e\^a\_-\_e\^a\_-\^a\_b(\_e\^b\_-\_e\^b\_) with $V_{\mu\nu}:= \epsilon_{ab} e^a{}_\mu e^b{}_\nu$. To recover the expressions from the previous section, in an arbitrary number of dimensions, we replace $\omega={\frac{1}{2}}\epsilon^{ab}\omega_{ab}$. Note that $V^{\mu\nu}V_{\mu\nu} = -2$ and $k:=-\eta_{22}$. The cartan metric $h_{ij}=h_{ji}$ is defined to be: h\_[ij]{} := -2 f\_[ki]{}\^l f\_[lj]{}\^k with components: h\_[22]{} &=& -4\ h\_[ab]{} &=& -4k\_[ab]{}\ h\_[a2]{} &=& 0 The action Eq.(\[eq:action1\]) becomes: S\_[YM]{} = d\^2 x (-\^2 - k\^2 T\^a \_[ab]{} T\^b + 2k\^2 \^4 + 2k\^2 V\^ \_),\[eq:covaction\] with \_ := \_\_- \_\_The last term in the action corresponds to the usual Einstein-Cartan term. In 2-dimensions it is a total divergence and will be dropped. Note that when $k=0$, the above action reduces simply to a single term, namely the curvature-squared term. The action (\[eq:covaction\]) corresponds is of the same form as the Katanaev-Volovich model of 2-D gravity with torsion[@kv86; @kv90s], albeit with a specific ratio of coefficients determined by the gauge coupling parameter. The equations of motion are the critical points of the action functional (\[eq:covaction\]). That is, since the the spin-connection $\omega_\mu$ and the frame-fields $e^a_\mu$ are functionally independent and S=d\^2x(W\^\_+E\^[a]{}e\_[a]{}), we have that a necessary condition for a critical point is that W\^&:=&-\_\^-C\_[ab]{} e\^a\_T\^[b]{}=0;\[eq:omegaeom\]\ E\^[a]{}&:=&-C D\_T\^[a]{}+e\^a\_\^+e\^[a]{}=0.\[eq:eeom\] where we have defined $C:= k\lambda^2$. As well, the above spacetime tensor indices $\mu,\nu,...$ are raised and lowered by the ‘background metric’ $g_{\mu\nu}:=\eta_{ab}e^a_\mu e^b_\nu$. There are two covariant derivatives. The first, $\nabla_\nu$, is with respect to the background Lorentzian metric $g_{\mu\nu}$, while the second, $D_\nu$ is with respect to the spin-connection. That is D\_T\^[a]{}:=\_T\^[a]{}-\^a\_b\_T\^[b]{}. Finally, the tensor $\tau^{\mu\nu}$ is defined as \^&:=&\_\^\^+C\_[ab]{}T\^a\_\^T\^[b]{}[\ ]{}&-&g\^(\^2+C T\^2), where ${\tilde R}^2:=\tilde{R}_{\mu\nu}\tilde{R}^{\mu\nu}$ and $T^2:=\eta_{ab} T^a{}_{\mu\nu}T^{b\mu\nu}$. Hamiltonian Analysis ==================== We parametrize the ‘background metric’ $g_{\mu\nu}=\eta_{ab}e^a_\mu e^b_\nu$, where the frame-field components are the $J_a$ components of the gauge potential $A_\mu=\lambda e^a_\mu J_a +\omega_\mu J$. We write: e\^0=n dt+pdx;e\^1=qN\^1 dt +qdx. We note that the metric is: g\_= ( [c c]{} -(n\^2-q\^2(N\^1)\^2)&-np+q\^2N\^1\ -np+q\^2N\^1&q\^2-p\^2 ) =q N, where $N:=n-N^1 p$. Note also that g\^= ( [c c]{} -(q\^2-p\^2)&-np+q\^2N\^1\ -np+q\^2N\^1&(n\^2-q\^2(N\^1)\^2) ) Another potentially useful form of the metric is: ds\^2= -dt\^2+ (q\^2-p\^2)(dx + dt)\^2 As before, we define: F=\_0\_1-\_1\_0 , and T\^0 &=&-n’+q(\_1 N\^1-\_0);[\ ]{}T\^1 &=&-(qN\^1)’-\_0 p +\_1 n. The action is S\_[YM]{}=d\^2x. Note that for the group ISO(1,1) (i.e. $k=0$) only the $F^2$ term remains. In two dimensions this gives a rather trivial solution space so we henceforth consider only $k=\pm1$.\ The momenta canonically conjugate to $p,q,n,N^1,\omega_0,\omega_1$ are respectively \_p&=&-;\[eq:conp\]\ \_q&=&;\[eq:conq\]\ \_n&=&0;\ \_1&=&0;\ P\_0&=&0;\ P\_1&=&. \[eq:P1 F\] The total Hamiltonian density is ‘pure constraint’: H=N H\_s +N\^1 D +\_0 M, where the Hamiltonian constraint $H_s$ is H\_s:=(-k\_p\^2+k\_q\^2+ \^2 P\_1\^2) - q -D\_p, the diffeo constraint $D$ is D:=-q D\_q-p D\_p, where $D\Pi_q:=\Pi_q'+\omega_1 \Pi_p$ and $D\Pi_p:=\Pi_p'+\omega_1 \Pi_q$. Finally the ‘Gauss law constraint’ is M:=-P\_1’+q\_p+p\_q. Note that above, $\Pi_q':=\partial_x \Pi_q$ is the spatial derivative. The self-consistency of these constraints, that is $0\approx\dot H_s(x)=[H_s(x),\int dy H(y)]$, etc., must be checked. We smear the constraints: $H_s[u]:=\int dy u(y) H_s(y)$, etc., and find =0; =dx (p H\_s - D)0; &=&dx{\^2 u vq P\_1 M + D - (qu’+u\_1 p)H\_s}[\ ]{}&&0; =0; and this is a strong equality. =D\[u v’-v u’\]0. And finally =0, strongly. We see that the constraint algebra closes, and the constraints are self-consistent. The equations of motion are: \_1 &=& \^2 N q P\_1 +’\_0;\ &=& k N q \_q-N\_1 +(qN\^1)’-N\_1p\_1+\_0 p;\ &=&-Nk q \_p + N’ +(N\^1p)’-N\^1q\_1+\_0q;\ \_1 &=& N\_q+N\^1(q\_p +p\_q)\[eq:P1dot\]\ \_q&=& - N + N\^1 D\_q-\_0 \_p;\[eq:Piqdot\]\ \_p &=& N\^1 D\_p - \_0 \_q.\[eq:Pipdot\] In the above we have defined: := (-k\^2\_p+k\^2\_q+ \^2 P\_1\^2 -\^2) Solutions ========= The gauge is fixed by p0,\_10,Q:=q-10. The consistency conditions for this choice are, respectively: &N’+\_0-[k]{} N \_p0;\[eq:pdot\]\ &\_0’+[k]{} CN P\_10;\[eq:omega1dot\]\ &(N\^1)’+[k]{} N \_q0.\[eq:qdot\] The constraints reduce to H\_s&=&-\_p’0;\[eq:Hsgf\]\ D&=&-\_q’0;\[eq:Dgf\]\ M&=&\_p-P\_1’0.\[eq:Mgf\] Now from Eq.(\[eq:Dgf\]) we have that $\Pi_q=\Pi_q(t)$ is an integration (spatial) constant. We use this and Eq.(\[eq:Mgf\])(which allows us to replace $\Pi_p$ by $P_1'$) in Eq.(\[eq:Hsgf\]) to get the second order differential equation P\_1”+(P\_1’)\^2-P\_1\^2-B=0, \[eq:P1de\] where B:=\_q\^2-. There are two classes of solutions to (\[eq:P1de\]). This can be seen as follows. Define: C\_1:= e\^[[k]{} P\_1]{}\[eq:C1\] It is easy to verify that C\_1\^= e\^[[k]{} P\_1]{} P\_1\^ Thus the solutions bifurcate into two classes: P\_1\^=0 && P\_1\^2-B\ P\_1\^0 && C\_1 = C\_1(t) As we will see, the first condition requires that the torsion be zero. It leads to a solution-space of lower dimension. The second condition allows for non-zero torsion. Torsion-less Solutions ---------------------- This class of solutions of the Hamiltonian constraint has $P_1'=0$, and hence P\_1\^2+=1. \[eq:P1 TF\] In this case the Gauss Law constraint $M=0$ implies $\Pi_p=0$. If we now use these in the equation of motion (\[eq:Pipdot\]) for $\dot{\Pi}_p$, we find that either $\omega_0=0$ or $\Pi_q=0$. If we use the former in the consistency condition $0=\dot{\omega}_1=\omega_0'+\lambda^2 N P_1$, then either $N=0$, which leads to a degenerate geometry, or $P_1=0$. But $P_1=0$ implies $\dot{P}_1=0$, and hence the equation of motion for $\dot{P}_1$ implies $\Pi_q=0$. Hence we must have that both $\Pi_p$ and $\Pi_q$ are zero; that is, the metric is torsion-free. Since $\Pi_q=0$ we now find from (\[eq:P1 TF\]) above that $P_1^2=k^2=1$. We now find that from the consistency conditions (\[eq:pdot\]) and (\[eq:omega1dot\]) that &N’+\_00;\[eq:pdot2\]\ &\_0’+k\^2 N0;\[eq:omega1dot2\] that $N''=k\lambda^2 N$ and $N^1=N^1(t)$. The function $N$ is then of the form $N_0(t)\sin{{\lambda}(x-x_0(t))}$, respectively $N_0(t)\sinh{{\lambda}(x-x_0(t))}$, as $k>0$, respectively $k<0$. In these expressions, $N_0(t),x_0(t)$ are integration constants. The metric is then of the form (with $k<0$): ds\^2=-(N\_0(t))\^2dt\^2+(dx+N\^1(t)dt)\^2. We have not completely fixed the coordinate invariance. One can choose: dy = dx + N\^1(t)dt As well, the lapse can be set to one using the residual time reparameterization invariance so that the metric becomes: ds\^2=-\^2[((y-y\_0(t)))]{}dt\^2+dy\^2. where y\_0(t) := x\_0(t) + dt N\_1(t) The remaining free function $y_0(t)$ is related to the fact that we have not completely fixed the gauge invariance. Indeed, the nontrivial consistency conditions, (60, 61), for the torsionless case, where $\Pi_p=0,P_1=\pm 1/\alpha$ boil down to (\[eq:pdot2\]) and (\[eq:omega1dot2\]), respectively, which can be written as a matrix equation ’=Awhere $\phi=[N,\omega_0]^T$ and $$A: = \left( \begin{array}{cc} 0 & -1\\ -k\lambda^2 & 0 \end{array}\right)$$ The system is preserved under linear transformations $\phi\to \bar{\phi}=L\phi$, where $L$ is a 2x2 matrix: $$L = \left( \begin{array}{cc} b_1& b_2\\ k\lambda^2 b_2& b_1 \end{array}\right)$$ which has unit determinant if $b_1^2-k\lambda^2 b_2^2=1$. Note that $b_i$ are functions of $t$. This is an $O(1,1)$ transformation. We can use such a transformation to transform $\tilde{x}_0(t)$ away. Indeed, such a transformation is given by $b_1(t)=-b_0(t)\lambda\cosh{(\lambda y_0(t))},b_2(t)=b_0(t)\sinh{(\lambda y_0(t))}$. This gives $N\propto \sinh{(\lambda y)}$. Note that in this case $b_1^2-k\lambda^2 b_2^2=k\lambda^2 b_0^2(t)$. On the other hand, if we choose to transform so that $N\propto \cosh{(\lambda y)}$, then we would have different $b_1,b_2$ satisfying $b_1^2-k\lambda^2 b_2^2=-k\lambda^2 b_0^2(t)$. Thus only one of these transformations is continuously connected to the identity transformation. Now consider the stationary metric ds\^2=-n\^2(z)d\^2+q\^2(z) dz\^2. For this to have constant curvature, that is, for $R=-2 \lambda^2$, it is required that q n”-q’ n’- k\^2 q\^3 n=0. If you solve this for $n(z)$ with $k=-1$ you get n(z)=A(())+B (()), where in general the integration constants $A,B$ can be $\tau$-dependent. In the above, $\theta'=q(z)$. Now choose $y$ as a new coordinate, so that $\int dz q(z)=y - y_0(\tau)$. The special case $A=0$ is then ds\^2=- dt\^2 +dy\^2. where we have scaled $B$ away by a trivial coordinate transformation $dt=B(\tau)d\tau$. The Ricci scalar for this metric is $R=-2\lambda^2$, a metric with constant negative curvature. For $k>0$ we get the above, but with $\sinh$ replaced by $\sin$. In this case $R=2\lambda^2$, giving us a space of constant positive curvature. This solution, as we have seen is torsionless, and of constant curvature. In fact, the solution has a flat Yang-Mills connection, that is, $F^i_{\mu\nu}=0$. Furthermore, the expression $\tau^{\mu\nu}$, which is quadratic in the Yang-Mills curvature, satisfies $\tau^{\mu\nu}+\frac{\lambda^4}{2} g^{\mu\nu}=0$. Torsion-full Solution ---------------------- One can verify that in the case $P_1^\prime \neq0$, the following is the first integral of Eq.(\[eq:P1de\]) (P\_1\^)\^2 &=&\ &=:&f\^2(t,P\_1) \[eq:solx\] where $ C_1(t)$ is an integration constant. We note also that as we have seen above, $\Pi_q'=0$ and hence $\Pi_q=\Pi_q(t)$. Now it is easy see that $\Pi_p$ is given as a function of $P_1$ from Eq.(\[eq:Mgf\]) by \_p=P\_1’=f(t,P\_1)=\^. Differentiate (\[eq:pdot\]) with respect to $x$. We write, for notational ease $r:=P_1$. We get N”+\_0’-[k]{}(Nf)’=0. We replace $\omega_0'$ according to (\[eq:omega1dot\]) and write $f'=\partial_x r f_r$ where $f_r=\partial_r f$ to get f\^2N\_[rr]{}+(f f\_r-[k]{} f\^2)N\_r-[k]{}(f f\_r+Cr)N=0. The general solution is (according to MAPLE) N(t,r)=f e\^[[k]{} r]{}(B\_1(t) g+B\_2(t)), where $B_1(t),B_2(t)$ are integration constants and $g(t,r)$ is given by $g_r=e^{-{k} r}f^{-3}$. We can now compute $\omega_0(t,r)$ from (\[eq:pdot\]) to get \_0(t,r)= -f\_r N-. From (\[eq:qdot\]) we find (N\^1)\_r=-[k]{}\_q e\^[[k]{} r]{} (B\_1(t) g+B\_2(t)). Now consider the equations of motion for the time derivatives of the momenta: &=&N \_q+N\^1 f; \[eq:rdot\]\ \_q&=&-f(N f\_r+ \_0); \[eq:piqdot\]\ &=&N\^1 f f\_r-\_0 \_q.\[eq:pipdot\] Consider the linear combination f\_r-=(N f\_r+\_0)\_q, by (\[eq:rdot\]) and (\[eq:pipdot\]). Use (\[eq:piqdot\]) on the right hand side to get \_q\_q+f(f\_r-)=0\[eq:sys\]. According to (\[eq:solx\]) f\^2=(f\^2)\_r+\_1(t) e\^[-[k]{} r]{}+2\_q\_q. Thus using this in (\[eq:sys\]) we get $\dot{C}_1=0$ and hence $C_1$ is both a space and time constant. From (\[eq:piqdot\]) we then get $\dot{\Pi}_q=B_1(t)$. Consider again (\[eq:pipdot\]). After multiplying by $f$. The left hand side becomes, after using (\[eq:solx\]) =2[k]{}\_q\_q. Hence (\[eq:pipdot\]) becomes 2[k]{}\_q\_q=(f N\^1+N\_q)(f\^2)\_r +2B\_1\_q, so that after canceling the left side with the last term on the right, we get using the expressions obtained above for $N,\omega_0,N^1$: 0&=&f(f\^2)\_r\_q[\ ]{}&=&f(f\^2)\_r\_q[\ ]{}&=& f(f\^2)\_r\_q\^r du f\^[-3]{}(u). where we integrated by parts to get the last equality. Since we have already seen that $B_1=\dot{\Pi}_q$, the most general nontrivial solution is $B_1=0$, so that $\Pi_q$ is a spatial and temporal constant. To compute the shift vector $N^1$, we solve (\[eq:qdot\]). The solution contains an arbitrary function of time, but this in turn is required to be zero by (\[eq:rdot\]). We now change the spatial coordinate from $x$ to $r$, so that $\dot{r}=0$. We also change to an new time coordinate $\tau$ by $d\tau=B_2(t) dt$. Then all the equations of motion and constraints are satisfied by f\^2&=&C\_1 e\^[-[k]{} r]{}+k\^2([k]{} r-1)\^2+ \_q\^2;\ N(,r)&=&f e\^[[k]{} r]{};\ \_0(,r)&=&-f f\_re\^[[k]{} r]{};\ N\^1(,r)&=&-\_q e\^[[k]{} r]{}. The metric is ds\^2=-N\^2d\^2+(dr+f N\^1 d)\^2. The Ricci scalar and torsion of the above are, respectively: R&=&-2k\^2( r\^2+[k]{} r -1)\ T\^0&=&-kNq\_p= -ke\^[kr]{}( C\_1 e\^[-[k]{} r]{}+k\^2([k]{} r-1)\^2+ \_q\^2)\ T\^1&=&Nq \_q= \_q e\^[kr]{} Thus the general solution with torsion depends on two integration constants, $C_1$ and $\Pi_q$. Event horizons exist for negative $C_1$. Conclusions =========== We have presented a study of DS/ADS/Poincare Yang-Mills gravity. As in general relativity, the theory is background independent, although this is done at the expense of reducing the symmetry group. We have shown that test ‘Higgs particles’ traverse geodesics with respect to the Lorentzian geometry determined by $e^i_\mu$. In two spacetime dimensions the action is a special case of the Katanaev-Volovich model. We completed the Hamiltonian analysis of the vacuum theory, confirming the existence of a generalized Birkhoff theorem: the solutions are static and parametrized by two parameters. In addition one of us (JG) is working on Yang-Mills gravity in 4D with gauge group SO(4,2), in collaboration with S. Rahmati and S. Seahra. In most work along these lines (see e.g. [@guo; @wheeler]), the torsion is forced to be zero [*ab initio*]{}. We will relax this, and explore implications, especially for cosmology. We close by mentioning that there is another possibility in principle allows the construction of an action that is invariant under the full gauge group. One can introduce two metrics: one is dynamical, and determined by the gauging $e^a_\mu$ of the generator $J_a$. The other metric, the background $g_{\mu\nu}$ is chosen in a way informed by the uniformization theorems in 2 and 3D. That is, given the topology, the manifold will admit a particular ‘round’ or homogeneous metric. The idea is to choose the background to be precisely that round geometry. This procedure is well defined in two or in three dimensions, but there is a problem in four or higher dimensions, for which there is no known uniformization theorem. Recall that the 3D uniformization theorem was proved using the Ricci flow [@thurston]. The latter exists in any dimension, and always converges to its fixed points, the homogeneous geometries. So one could require that the background geometry is such that the Ricci flow of the geometry determined by the frame fields and spin connection converges to it in the infinite limit of the flow parameter. However, given that we are really interested in working with YM type actions, it is more sensible to postulate that the consistency is provided by requiring that the Yang-Mills flow of the gauge potential $A$ determined by the frame fields and spin connection flows to the background ‘round geometry’. In future work we will therefore consider an alternate theory wherein the background a metric which is the appropriate homogeneous geometry for some topology. Acknowledgments {#acknowledgments .unnumbered} ================ The authors are grateful to Daniel Grumiller for drawing our attention to the Katanaev-Volovich model. We thank Sanjeev Seahra for useful conversations. One of, JG, acknowledges the support of the CECS in Valdivia, Chile, where part of this work was done. We also acknowledge the partial financial support of NSERC as well as the Perimeter Institute for Theoretical Physics (funded by Industry Canada and the Province of Ontario Ministry of Research and Innovation).\ [99]{} P.K. Townsend [*Small-scale structure of spacetime as the origin of the gravitational constant*]{}, Phys. Rev. [**D15**]{}, 2795 (1977). H. Weyl, [*Space-Time-Matter*]{}, Methuen, London (1918). A sample, not exhaustive, list of papers on Yang-Mills gravity from the 70’s and 80’s: R.  Utiyama, ‘Invariant theoretical interpretation of interaction’, Physical Review [**101**]{},1597 (1956);doi:10.1103/PhysRev.101.1597;\ T.  W.  B.  Kibble, ‘Lorentz invariance and the gravitationa l field’, J. Math. Phys. [**2**]{},212 (1961); gravitational constant’, Phys. Rev. D [**15**]{}, 2795 (1977). S.  W.  MacDowell and F.  Mansouri, ‘Unified geometric theor y of gravity and su- pergravity’, Phys. Rev. Lett. [**38**]{} , 739–742 (1977);\ K.  Hayashi and T.  Shirafuji, ‘Gravity from Poincar ´e gaug e theory of fundamental interactions’, Prog. Theor. Phys. [**64**]{} , 866–882 (1980);\ E. A. Ivanov and J. Niederle, ‘Gauge formulation of gravitational theories. I. The Poincare, de Sitter, and conformal cases’, Phys. Rev. D [**25**]{}, 976 (1981). Again a sample of the literature in this area would include: J. T. Wheeler, ‘Auxiliary field in conformal gauge theory’, Phys. Rev. d [**44**]{}, 1769 (1991); de Sitter gravity’, Class. Quant. Grav. [**24**]{}, 4009 (2007). or on ‘conformal gravity’. H.-Y  Guo, et. al., ‘Snyder’s model-de Sitter special relativity duality and de Sitter gravity’, Class. Quant. Grav. [**24**]{}, 4009 (2007); C.-G. Huang, H.-Q. Zhang, H.-Y. Guo: [*Cosmological solutions with torsion in a model of the de Sitter gauge theory of gravity*]{}. JCAP 10 (2008) 010; C.-G. Huang, M.-S. Ma: On torsion-free vacuum solutions of the model of de Sitter gauge theory of gravity (II). Front. Phys. China, 4 (2009) 525–529; C.-G. Huang, M.-S. Ma: [*de Sitter spacetimes with torsion in the model of de Sitter gauge theory of gravity*]{}. Phys. Rev. D 80 (2009) 084033; C.-G. Huang, M.-S. Ma: [*A new solution with torsion in model of dS gauge theory of gravity*]{}. Commun. Theor. Phys. 55 (2011) 65–68. J.  Gegenberg, A. C.  Day, H.  Liu and S.  S.  Seahra ‘An instability of hyperbolic space under the Yang-Mills flow’ Journal of Mathematical Physics, [**55**]{} , 042501 arXiv: 1210.0839 \[hep-th\]. M. O. Katanaev and I. V. Volovich, “String model with dynamical geometry and torsion,” Phys. Lett. B175 (1986) 413–416. \[arXiv:hep-th/0209014\]. P. Schaller and T. Strobl, “Canonical Quantization of Non-Einsteinian Gravity and the Problem of Time”, Class. Quant. Grav.11 (1994) 331-346 \[hep-th/9211054\] ; Noriaki Ikeda and Ken-Iti Izawa, “Quantum Gravity with Dynamical Torsion in Two Dimensions” Prog. Theor. Phys. 89 (1993) 223-230; T. Strobl, “Comment on Gravity and the Poincare Group”, Phys.Rev. D48 (1993) 5029-5031 \[arXiv:hep-th/9302041\]; W. Kummer and D. J. Schwarz, “General analytic solution of R\*\*2 gravity with dynamical torsion in two-dimensions,” Phys. Rev. D45 (1992) 3628–3635. M. O. Katanaev, W. Kummer, and H. Liebl, “Geometric interpretation and classification of global solutions in generalized dilaton gravity,” Phys. Rev. D53 (1996) 5609–5618 \[gr-qc/9511009\]. R. Jackiw, in Quantum Theory of Gravity , edited by S. Christensen (Hilger, Bristol, 1984), p. 403;C. Teitelboim, in Quantum Theory of Gravity , edited by S. Christensen (Hilger, Bristol, 1984), p.327; M . Hen- neaux, Phys. Rev. Lett. 54 , 959 (1985). See also D. Louis-Martinez, J. Gegenberg and G. Kunstatter, Phys . Letts. B [**321**]{} , 193 (1994). E.  Inönü, E.P.  Wigner, [*On the Contraction of Groups and Their Representations*]{}, Proc. Nat. Acad. Sci. [**39**]{} (6): 510–24 (1953). O.  Brodbeck and N. Straumannm [*A generalized Birkhoff theorem for the Einstein-Yang-Mills system*]{}, J. Math. Phys. [**34**]{}, 2412-2423 (1993); T. A.  Oliynyk and H. P.  Kunzle, [*On all possible static spherically symmetric EYM solitons and black holes*]{}, Class. Quant. Grav. [**19**]{}, 457 (2002), \[arXiv:gr-qc/0109075\]. The original geometrization conjecture of W.P. Thurston was proved by J. Hamilton and G. Pereleman. For references, see J.  W.  Morgan. [*Recent progress on the Poincaré conjecture and the classification of 3-manifolds*]{}, Bulletin Amer. Math. Soc. 42 (2005) no. 1, 57-78.
{ "pile_set_name": "ArXiv" }
--- abstract: 'For homeland and transportation security applications, 2D X-ray explosive detection system (EDS) have been widely used, but they have limitations in recognizing 3D shape of the hidden objects. Among various types of 3D computed tomography (CT) systems to address this issue, this paper is interested in a stationary CT using fixed X-ray sources and detectors. However, due to the limited number of projection views, analytic reconstruction algorithms produce severe streaking artifacts. Inspired by recent success of deep learning approach for sparse view CT reconstruction, here we propose a novel image and sinogram domain deep learning architecture for 3D reconstruction from very sparse view measurement. The algorithm has been tested with the real data from a prototype 9-view dual energy stationary CT EDS carry-on baggage scanner developed by GEMSS Medical Systems, Korea, which confirms the superior reconstruction performance over the existing approaches.' author: - - - title: 'Deep Learning Reconstruction for 9-View Dual Energy CT Baggage Scanner' --- Explosive detection system (EDS), sparse-view X-ray CT, convolutional neural network (CNN) Introduction {#sec:intro} ============ In homeland and aviation security applications, there has been increasing demand for X-ray CT EDS system for carry-on baggage screening. A CT-EDS can produce an accurate 3D object structure for segmentation and threat detection, which is often not possible when a 2D-EDS system captures projection views in only one or two angular directions. There are currently two types of CT EDS systems: gantry-based CT and stationary CT. While gantry-based CT EDS is largely the same as medical CT, baggage screening should be carried out continuously, so it is often difficult to continuously screen carry-on bags because of the possible mechanical overloading of the gantry system. On the other hand, a stationary CT EDS system uses fixed X-ray sources and detectors, making the system suitable for routine carry-on baggage inspection. For example, Fig. \[fig:ct\_system\] shows source and detector geometry of the prototype stationary CT-EDS system developed by GEMSS Medical Systems, Korea. As shown in Fig. \[fig:ct\_system\](a), nine pairs of X-ray source and dual energy detector in the opposite direction are distributed at the same angular interval. For seamless screening without stopping convey belt, each pair of source and detectors are arranged along the z-direction as shown in Fig. \[fig:ct\_system\](b) so that different projection view data can be collected while the carry-on baggages moves continuously on the conveyor belt. Then, 9-view fan beam projection data is obtained for each z-slice by rebinning the measurement data. This type of stationary CT system is suitable for EDS applications because it does not require a rotating gantry, but with only 9 projection views it is difficult to use a conventional filtered backprojection (FBP) algorithm due to severe streaking artifacts. Therefore, advanced reconstruction algorithms with fast reconstruction time are required. For sparse-view CT EDS, model-based iterative reconstruction (MBIR) with the total variation (TV) penalty have been extensively investigated [@mandava2017image; @kisner2012limited]. Inspired by the recent success of deep learning approach for sparse view and limited angle CT [@han2016deep; @han2017framing; @jin2017deep; @gu2017multi] that outperform the classical MBIR approach, this paper aims at developing a deep learning approach for real-world sparse view CT EDS. However, neural network training using the retrospective angular subsampling as in the existing works [@han2016deep; @han2017framing; @jin2017deep; @gu2017multi] is not possible for our prototype system, since there are no ground-truth data for the real world sparse view CT EDS. We therefore propose a novel deep learning approach composed of image domain and sinogram domain learning that compensate for the imperfect label data. Theory ====== Problem Formulation ------------------- Recall that the forward model for sparse view CT EDS system can be represented by $$\begin{aligned} \label{eq:fwd} g_\Theta = {{\mathcal P}}_\Theta {{\mathcal R}}f\end{aligned}$$ where ${{\mathcal R}}$ denotes the 3D projection operator from an $x-y-z$ volume image to a $s-\theta-z$ domain sinogram data with $s,\theta$ and $z$ denoting the detector, projection angle, and the direction of the conveyor belt travel, respectively. See Fig. \[fig:flow\] for the coordinate systems. In , ${{\mathcal P}}_\Theta$ denotes the view sampling operator for the measured angle set $\Theta$, and $g_\Theta$ refers to the measured sinogram data. For each projection view data, we use the notation $g_\theta$ and ${{\mathcal P}}_\theta$, where $\theta$ denotes the specific view. The main technical issue of the sparse view CT reconstruction is the non-uniqueness of the solution for . More specifically, there exists a null spacce ${{\mathcal N}}_\Theta$ such that $${{\mathcal P}}_\Theta {{\mathcal R}}h=0,\quad \forall h \in {{\mathcal N}}_\Theta,$$ which leads to infinite number of feasible solutions. To avoid the non-uniqueness of the solution, constrained form of the penalized MBIR can be formulated as : $$\begin{aligned} \label{eq:MBIR} \min_{f\in {{\mathbb R}}^3} \|\mathrm{L} f\|_{1}, \quad \mbox{subject to} \quad g_\Theta = {{\mathcal P}}_\Theta{{\mathcal R}}\ ,\end{aligned}$$ where $\mathrm{L}$ refers to a linear operator and $\|\cdot\|_1$ denotes the $l_1$ norm. For the case of the TV penalty, $\mathrm{L}$ corresponds to the derivative. Then, the uniqueness of is guaranteed that if the ${{\mathcal N}}_{\mathrm{L}} \cap {{\mathcal N}}_\Theta = \{0\}$, where ${{\mathcal N}}_{\mathrm{L}}$ denotes the null space of the operator $\mathrm{L}$. Instead of designing a linear operator ${\mathrm{L}}$ such that the common null space of ${{\mathcal N}}_\Theta$ and ${{\mathcal N}}_{\mathrm{L}}$ to be zero, we can design a frame ${{\mathcal W}}$, its dual $\tilde {{\mathcal W}}$, and shrinkage operator $S_\lambda$ such that $\tilde {{\mathcal W}}^\top {{\mathcal W}}=I$ and $$\tilde {{\mathcal W}}^\top S_\lambda {{\mathcal W}}(f^*+g) = f^*\quad \forall g \in {{\mathcal N}}_\Theta$$ for the ground-truth image $f^*$. This frame-based regularization is also an active field of research for image denoising, inpainting, etc [@cai2008framelet]. One of the most important contributions of the deep convolutional framelet theory [@ye2017deep] is that ${{\mathcal W}}$ and $\tilde {{\mathcal W}}^\top$ correspond to the encoder and decoder structure of a convolutional neural network (CNN), respectively, and the shrinkage operator $S_\lambda$ emerges by controlling the number of filter channels and nonlinearities. More specifically, a convolutional neural network can be designed such that ${{\mathcal Q}}=\tilde {{\mathcal W}}^\top S_\lambda {{\mathcal W}}$ and $$\begin{aligned} \label{eq:Qc1} {{\mathcal Q}}(f^*+h) = f^* ,\quad \forall h \in {{\mathcal N}}_\Theta \quad .\end{aligned}$$ In other word, directly removes the null space component. Eq.  is the constraint we use for training our neural network. Derivation of Image and Projection Domain CNNs ---------------------------------------------- More specifically, our sparse view reconstruction algorithm finds the unknown $f\in {{\mathbb R}}^3$ that satisfy both data fidelity and the so-called frame constraints [@ye2017deep]: $$\begin{aligned} \label{eq:constraint} g_\Theta = {{\mathcal P}}_\Theta{{\mathcal R}}f ,\quad {{\mathcal Q}}^I (f) = f^*\quad , $$ where ${{\mathcal Q}}^I$ is the image domain CNN that satisfies and $f^*$ denotes the ground-truth images that are available for training data. Now, by defining ${{\mathcal M}}$ as a right-inverse of ${{\mathcal P}}_\Theta {{\mathcal R}}$, i.e. $({{\mathcal P}}_\Theta{{\mathcal R}}){{\mathcal M}}g_\Theta =g_\Theta, \forall g_\Theta$, we have $${{\mathcal M}}g_\Theta = f^*+h$$ for some $h\in {{\mathcal N}}_\Theta$, since the right inverse is not unique due to the existence of the null space. Thus, we can show that $ {{\mathcal M}}g_\Theta$ is the feasible solution for , since we have $$\begin{aligned} \label{eq:Q2} {{\mathcal Q}}^I {{\mathcal M}}g_\Theta = {{\mathcal Q}}^I \left(f^*+h\right) = f^*\quad ,\end{aligned}$$ for the training data, and $$\begin{aligned} \label{eq:fidelity} {{\mathcal P}}_\Theta{{\mathcal R}}{{\mathcal M}}g_\Theta = {{\mathcal P}}_\Theta{{\mathcal R}}(f^*+h) = g_\Theta\quad. \end{aligned}$$ Therefore, the neural network training problem to satisfy can be equivalently represented by $$\begin{aligned} \label{eq:opt} \min_{{{\mathcal Q}}^I} \sum_{i=1}^N\|f^{*(i)} - {{\mathcal Q}}^I {{\mathcal M}}g_\Theta^{(i)} \|^2\end{aligned}$$ where $\{(f^{*(i)},g_\Theta^{(i)})\}_{i=1}^N$ denotes the training data set composed of ground-truth image an its sparse view projection. Since a representative right inverse for the sparse view projection is the inverse Radon transform after zero padding to the missing view, ${{\mathcal M}}g_\Theta^{(i)}$ in can be implemented using the standard FBP algorithm. In fact, this is the main theoretical ground for the success of image domain CNN when the ground-truth data is available [@han2016deep; @han2017framing; @jin2017deep; @gu2017multi]. Moreover, the fan-beam rebinning makes the problem separable for each $z$ slices, so we use the 2D FBP for each slice as shown in Fig. \[fig:flow\]. However, the main technical difficulties in our 9-view CT EDS system is that we do not have ground-truth image $\{f^{*(i)}\}_{i=1}^N$. One could use physical phantoms and atomic number to form a set of ground-truth images, but those data set may be different from the real carry-on bags, so we need a new method to account for the lack of ground-truth for neural network training. Thus, to overcome the lack of the ground-truth data, the approximate label images are generated using an MBIR with TV penalty. Then, using MBIR reconstruction as label data $\{f^{*(i)}\}_{i=1}^N$, an 2D image domain network ${{\mathcal Q}}^I$ is trained to learn the mapping between the artifact-corrupted 2D image and MBIR reconstruction in $x-y$ domain. One downside of this approach is that the network training by is no more optimal, since the label data is not the ground-truth image. Thus, the generated sinogram data from the denoised 3D volume may be biased. Thus, we impose additional frame constraint to the sinogram data in addition to : $$\begin{aligned} \label{eq:constraint2} g_\theta^* = {{\mathcal Q}}^S\left( g_\theta\right) , $$ for the measured angle $\theta$, where ${{\mathcal Q}}^S$ is the $s-z$ sinogram domain CNN and $g_\theta^*$ denotes the ground-truth sinogram data measured at $\theta$. Then, Eq.  leads to the following network training: $$\begin{aligned} \label{eq:opt2} \min_{{{\mathcal Q}}^S} \sum_{\theta\in \Theta}\sum_{i=1}^N \|g_\theta^{*(i)} - {{\mathcal Q}}^S \left({{\mathcal P}}_\theta {{\mathcal R}}{{\mathcal Q}}^I{{\mathcal M}}g_\Theta^{(i)}\right) \|^2\end{aligned}$$ More specifically, as shown in Fig. \[fig:flow\], 3D sinogram data is generated in the $s-\theta-z$ domain by applying the forward projection operator along 720-projection views after stacking the image domain network output over multiple slices to form 3D reconstruction volume in the $x-y-z$ domain. Next, a 2D sinogram domain network ${{\mathcal Q}}^S$ is trained so that it can learn the mapping between the synthetic $s-z$ sinogram data and the real projection data in the $s-z$ domain. Since the real projection data is available only in 9 views, this sinogram network training is performed using synthetic and real projection data in the measured projection views. The optimization problems and can be solved sequentially or simultaneously, and in this paper we adopt the sequential optimization approach for simplicity. After the neural networks ${{\mathcal Q}}^I$ and ${{\mathcal Q}}^S$ are trained, the inference can be done simply by obtaining $x-y-z$ volume images from the 9 view projection data by slice-by-slice FBP algorithm, which are then fed into ${{\mathcal Q}}^I$ to obtain the denoised 3D volume data. Then, by applying projection operator, we generate 720 projection view data in $s-\theta-z$ domain, which are fed into the ${{\mathcal Q}}^S$ to obtain denoised sinogram data for each $\theta$ angle. Then, the final reconstruction is obtained by applying FBP algorithms. One could use post-processing using additional TV-based denosing. This algorithmic flow is illustrated in Fig. \[fig:flow\]. Methods ======= Real CT EDS data Acquisition ---------------------------- We collected CT EDS data using the prototype stationary 9 view dual energy CT-EDS system developed by GEMSS Medical Systems, Korea as shown in Fig. \[fig:ct\_system\]. the distance from source to detector (DSD) and the distance from source to object (DSO) are $1202.6 \rm{mm}$ and $648.2 \rm{mm}$, respectively. The number of detector is $384$ with a pitch of $1.5 \rm{mm}$. The region of interest (ROI) is $256\times256$ and the pixel size is $2 \rm{mm}^2$. The detectors collect low and high energy X-ray at 80KVp and 120KVp, respectively. We collect 47 sets of projection data from the prototype CT EDS baggage scanner. Among the 47 sets, 32 dataset are simple-objects and the other set are realistic carry-on bags. The 47 set of 28 simple- and 13 baggage-objects was used during the training phase, and the validation was performed by two simple- and one baggage-object. The other set was used for test. Network Architecture and Training ---------------------------------- Fig. \[fig:network\] illustrates modified the U-Net structure [@ronneberger2015u] for the image domain and the sinogram domain networks. To account for the multi-energy image and sinogram data, the input for the network is two channel multi-energy image and sinogram data. The proposed network consists of convolution layer, batch normalization, rectified linear unit (ReLU) [@krizhevsky2012imagenet], and contracting path connection with concatenation [@ronneberger2015u]. A detail parameters are illustrated as shown in Fig. \[fig:network\]. The proposed networks were trained by stochastic gradient descent (SGD). The regularization parameter was $\lambda = 10^{-4}$. The learning rate has been set from $10^{-3}$ to $10^{-5}$, which has been reduced step by step in each epoch. The number of epoch was 200. The batch size was 12 and the patch size for image and projection data are $256\times256\times2$ and $768\times384\times2$, respectively. The network was implemented using MatConvNet toolbox (ver.24) [@vedaldi2015matconvnet] in the MATLAB 2015a environment (MathWorks, Natick). Central processing unit (CPU) and graphic processing unit (GPU) specification are i7-7700 CPU (3.60 GHz) and GTX 1080 Ti GPU, respectively. Experimental Results ==================== To evaluate the performance of the proposed method, we perform image reconstruction from real 9-view CT EDS prototype system. Fig. \[fig:result\_pht\] illustrates image reconstruction results of bag using various methods such as FBP, MBIR with TV penalty, image domain CNN [@han2016deep; @jin2017deep], and the proposed method. The FBP reconstruction results suffered from severe streaking artifacts, so it was difficult to see the threats in the tomographic reconstruction and 3D rendering. The MBIR and image domain CNN were slight better in their reconstruction quality, but the detailed 3D structures were not fully recovered and several objects were not detected as indicated by the red arrow in Fig. \[fig:result\_pht\]. Moreover, the 3D rendering results in Fig. \[fig:result\_pht\] correctly identify the shape of grenade and knife as well as the frame of the bag, which was not possible using other methods. Because we do not have the ground-truth in the image domain, we perform quantitative evaluation using normalized mean squares error (NMSE) in the sinogram domain. More specifically, after obtaining the final reconstruction, we perform the forward projection to generate the sinogram data in the measured projection view and calculated the normalized mean square errors. Table \[tab\_quality\] showed that the proposed method provides the most accurate sinogram data compared to the other methods. Moreover, the $s-z$ projection data in Fig. \[fig:result\_r\_pht\] showed that the projection data from the proposed method is much closer to the ground-truth measurement data. \[tab\_quality\] Energy level FBP MBIR-TV Image CNN Ours -------------- ----------- ----------- ----------- ----------- 80 kvP 1.6647e+1 5.8247e-1 3.3207e-1 0.6845e-1 120 kvP 1.0536e+1 6.0440e-1 3.2249e-1 0.5450e-1 : NMSE value comparison of various methods. Conclusion ========== In this paper, we proposed a novel deep learning reconstruction algorithm for a prototype 9-view dual energy CT EDS for carry-on baggage scanner. Even though the number of projection view was not sufficient for high equality 3D reconstruction, our method learns the relationships between the 2D tomographic slices in $x-y$ domain as well as the 2D projections in $s-z$ domain such that the artifact-corrupted image and sinogram data can be successively refined to obtain high quality images. Using real data from our prototype 9-view CT EDS system, we demonstrated that the proposed method outperforms the existing algorithms, delivering high quality three reconstruction for threat detection. Acknowledgment {#acknowledgment .unnumbered} ============== This work is supported by Korea Agency for Infrastructure Technology Advancement, Grant number 17ATRP-C071164-05-000000. [10]{} Sagar Mandava, David Coccarelli, Joel A Greenberg, Michael E Gehm, Amit Ashok, and Ali Bilgin, “Image reconstruction for view-limited x-ray ct in baggage scanning,” in [*Anomaly Detection and Imaging with X-Rays (ADIX) II*]{}. International Society for Optics and Photonics, 2017, vol. 10187, p. 101870F. Sherman J Kisner, Eri Haneda, Charles A Bouman, Sondre Skatter, Mikhail Kourinny, and Simon Bedford, “Limited view angle iterative ct reconstruction,” in [*Computational Imaging X*]{}, 2012, vol. 8296. Yoseop Han, Jaejoon Yoo, and Jong Chul Ye, “Deep residual learning for compressed sensing ct reconstruction via persistent homology analysis,” , 2016. Yoseob Han and Jong Chul Ye, “Framing [U]{}-net via deep convolutional framelets: Application to sparse-view [CT]{},” , 2017. Kyong Hwan Jin, Michael T McCann, Emmanuel Froustey, and Michael Unser, “Deep convolutional neural network for inverse problems in imaging,” , vol. 26, no. 9, pp. 4509–4522, 2017. Jawook Gu and Jong Chul Ye, “Multi-scale wavelet domain residual learning for limited-angle [CT]{} reconstruction,” , 2017. Jian-Feng Cai, Raymond H Chan, and Zuowei Shen, “A framelet-based image inpainting algorithm,” , vol. 24, no. 2, pp. 131–149, 2008. Jong Chul Ye, Yo Seob Han, and Eunjoo Cha, “Deep convolutional framelets: A general deep learning framework for inverse problems,” , 2017. Olaf Ronneberger, Philipp Fischer, and Thomas Brox, “U-net: Convolutional networks for biomedical image segmentation,” in [*International Conference on Medical Image Computing and Computer-Assisted Intervention*]{}. Springer, 2015, pp. 234–241. Alex Krizhevsky, Ilya Sutskever, and Geoffrey E Hinton, “Imagenet classification with deep convolutional neural networks,” in [*Advances in neural information processing systems*]{}, 2012, pp. 1097–1105. Andrea Vedaldi and Karel Lenc, “Matconvnet: Convolutional neural networks for matlab,” in [*Proceedings of the 23rd ACM international conference on Multimedia*]{}. ACM, 2015, pp. 689–692.
{ "pile_set_name": "ArXiv" }
--- abstract: | Tone is a prosodic feature used to distinguish words in many languages, some of which are endangered and scarcely documented. In this work, we use unsupervised representation learning to identify probable clusters of syllables that share the same phonemic tone. Our method extracts the pitch for each syllable, then trains a convolutional autoencoder to learn a low-dimensional representation for each contour. We then apply the mean shift algorithm to cluster tones in high-density regions of the latent space. Furthermore, by feeding the centers of each cluster into the decoder, we produce a prototypical contour that represents each cluster. We apply this method to spoken multi-syllable words in Mandarin Chinese and Cantonese and evaluate how closely our clusters match the ground truth tone categories. Finally, we discuss some difficulties with our approach, including contextual tone variation and allophony effects.\ [**Index Terms**]{}: representation learning, unsupervised learning, low-resource languages, computational phonology, tone address: | $^1$University of Toronto, Canada\ $^2$Vector Institute for Artificial Intelligence, Canada\ $^3$St Michael’s Hospital, Canada\ `{bai,frank}@cs.toronto.edu, [email protected]` bibliography: - 'refs.bib' title: Representation Learning for Discovering Phonemic Tone Contours --- Introduction ============ Tonal languages use pitch to distinguish different words, for example, [*yi*]{} in Mandarin may mean ‘one’, ‘to move’, ‘already’, or ‘art’, depending on the pitch contour. Of over 6000 languages in the world, it is estimated that as many as 60-70% are tonal [@ethnologue; @yip-tone]. A few of these are national languages (e.g., Mandarin Chinese, Vietnamese, and Thai), but many tonal languages have a small number of speakers and are scarcely documented. There is a limited availability of trained linguists to perform language documentation before these languages become extinct, hence the need for better tools to assist linguists in these tasks. One of the first tasks during the description of an unfamiliar language is determining its phonemic inventory: what are the consonants, vowels, and tones of the language, and which pairs of phonemes are contrastive? Tone presents a unique challenge because unlike consonants and vowels, which can be identified in isolation, tones do not have a fixed pitch, and vary by speaker and situation. Since tone data is subject to interpretation, different linguists may produce different descriptions of the tone system of the same language [@yip-tone]. In this work, we present a model to automatically infer phonemic tone categories of a tonal language. We use an unsupervised representation learning and clustering approach, which requires only a set of spoken words in the target language, and produces clusters of syllables that probably have the same tone. We apply our method on Mandarin Chinese and Cantonese datasets, for which the ground truth annotation is used for evaluation. Our method does not make any language-specific assumptions, so it may be applied to low-resource languages whose phonemic inventories are not already established. Tone in Mandarin and Cantonese ------------------------------ ![Pitch contours for the four Mandarin tones and six Cantonese tones in isolation, produced by native speakers. Figure adapted from [@mandarin-cantonese-tones].[]{data-label="fig:mandarin-cantonese-tones"}](figures/mandarin-cantonese-tones.png){width="\linewidth"} Mandarin Chinese (1.1 billion speakers) and Cantonese (74 million speakers) are two tonal languages in the Sinitic family [@ethnologue]. Mandarin has four lexical tones: high (55), rising (25), low-dipping (214), and falling (51)[^1]. The third tone sometimes undergoes sandhi, addressed in section \[sec:third-tone-sandhi\]. We exclude a fifth, neutral tone, which can only occur in word-final positions and has no fixed pitch. Cantonese has six lexical tones: high-level (55), mid-rising (25), mid-level (33), low-falling (21), low-rising (23), and low-level (22). Some descriptions of Cantonese include nine tones, of which three are [*checked*]{} tones that are flat, shorter in duration, and only occur on syllables ending in /p/, /t/, or /k/. Since each one of the checked tones are in complementary distribution with an unchecked tone, we adopt the simpler six tone model that treats the checked tones as variants of the high, mid, and low level tones. Contours for the lexical tones in both languages are shown in Figure \[fig:mandarin-cantonese-tones\]. Related Work ============ Many low-resource languages lack sufficient transcribed data for supervised speech processing, thus unsupervised models for speech processing is an emerging area of research. The Zerospeech 2015 and 2017 challenges featured unsupervised learning of contrasting phonemes in English and Xitsonga, evaluated by an ABX phoneme discrimination task [@zerospeech-challenge]. One successful approach used denoising and correspondence autoencoders to learn a representation that avoided capturing noise and irrelevant inter-speaker variation [@zerospeech-autoencoders]. Deep LSTMs for segmenting and clustering phonemes in speech have also been explored in [@muller1] and [@muller2]. In Mandarin Chinese, deep neural networks have been successful for tone classification in isolated syllables [@mandarin-classification-cnn] as well as in continuous speech [@ryant1; @ryant2]. Both of these models found that Mel-frequency cepstral coefficients (MFCCs) outperformed pitch contour features, despite the fact that MFCC features do not contain pitch information. In Cantonese, support vector machines (SVMs) have been applied to classify tones in continuous speech, using pitch contours as input [@cantonese-svm]. Unsupervised learning of tones remains largely unexplored. Levow [@levow2006] performed unsupervised and semi-supervised tone clustering in Mandarin, using average pitch and slope as features, and $k$-means and asymmetric $k$-lines for clustering. Graph-based community detection techniques have been applied to group $n$-grams of contiguous contours into clusters in Mandarin [@mandarin-tone-shapes]. Our work appears to be the first model to use unsupervised deep neural networks for phonemic tone clustering. Data and Preprocessing ====================== ![Diagram of our model architecture, consisting of a convolutional autoencoder to learn a latent representation for each pitch contour, and mean shift clustering to identify groups of similar tones.[]{data-label="fig:architecture-conv"}](figures/architecture-conv-simple.pdf){width="\linewidth"} We use data from Mandarin Chinese and Cantonese. For each language, the data consists of a list of spoken words, recorded by the same speaker. The Mandarin dataset is from a female speaker and is provided by Shtooka[^2], and the Cantonese dataset is from a male speaker and is downloaded from Forvo[^3], an online crowd-sourced pronunciation dictionary. We require all samples within each language to be from the same speaker to avoid the difficulties associated with channel effects and inter-speaker variation. We randomly sample 400 words from each language, which are mostly between 2 and 4 syllables; to reduce the prosody effects with longer utterances, we exclude words longer than 4 syllables. \[sec:third-tone-sandhi\] We extract ground-truth tones for evaluation purposes. In Mandarin, the tones are extracted from the pinyin transcription; in Cantonese, we reference the character entries on Wiktionary[^4] to retrieve the romanized pronunciation and tones. For Mandarin, we correct for third-tone sandhi (a phonological rule where a pair of consecutive third-tones is always realized as a second-tone followed by a third-tone). We also exclude the neutral tone, which has no fixed pitch and is sometimes thought of as a lack of tone. Pitch extraction and syllable segmentation ------------------------------------------ We use Praat’s autocorrelation-based pitch estimation algorithm to extract the fundamental frequency (F0) contour for each sample, using a minimum frequency of 75Hz and a maximum frequency of 500Hz [@praat-f0]. The interface between Python and Praat is handled using Parselmouth [@parselmouth]. We normalize the contour to be between 0 and 1, based on the speaker’s pitch range. Next, we segment each speech sample into syllables, which is necessary because syllable boundaries are not provided in our datasets. This is done using a simple heuristic that detects continuously voiced segments, and manual annotation where the heuristic fails. To obtain a constant length pitch contour as input to our model, we sample the pitch at 40 equally spaced points. Note that by sampling a variable length contour to a constant length, information about syllable length is lost; this is acceptable because we consider tones which differ on length as variations of the same tone. ![image](figures/representations-contours.pdf){width="0.8\linewidth"} Model ===== Convolutional autoencoder ------------------------- We use a convolutional autoencoder (Figure \[fig:architecture-conv\]) to learn a two-dimensional latent vector for each syllable. Convolutional layers are widely used in computer vision and speech processing to learn spatially local features that are invariant of position. We use a low dimensional latent space so that the model learns to generate a representation that only captures the most important aspects of the input contour, and also because clustering algorithms tend to perform poorly in high dimensional spaces. Our encoder consists of three layers. The first layer applies 2 convolutional filters (kernel size 4, stride 1) followed by max pooling (kernel size 2) and a tanh activation. The second layer applies 4 convolutional filters (kernel size 4, stride 1), again with max pooling (kernel size 2) and a tanh activation. The third layer is a fully connected layer with two dimensional output. Our decoder is the encoder in reverse, consisting of one fully connected layer and two deconvolution layers, with the same layer shapes as the encoder. We train the autoencoder using PyTorch [@pytorch], for 500 epochs, with a batch size of 60. The model is optimized using Adam [@adam] with a learning rate of 5e-4 to minimize the mean squared error between the input and output contours. Mean shift clustering --------------------- We run the encoder on each syllable’s pitch contour to get their latent representations; we apply principal component analysis (PCA) to remove any correlation between the two dimensions. Then, we run mean shift clustering [@meanshift; @ghassabeh], estimating a probability density function in the latent space. The procedure performs gradient ascent on all the points until they converge to a set of stationary points, which are local maxima of the density function. These stationary points are taken to be cluster centers, and points that converge to the same stationary point belong to the same cluster. Unlike $k$-means clustering, the mean shift procedure does not require the number of clusters to be specified, only a bandwidth parameter (set to 0.6 for our experiments). The cluster centers are always in regions of high density, so they can be viewed as prototypes that represent their respective clusters. Another advantage is that unlike $k$-means, mean shift clustering is robust to outliers. Although the mean shift procedure technically assigns every point to a cluster, not all such clusters are linguistically plausible as phonemic tones, because they contain very few points. Thus, we take only clusters larger than a threshold, determined empirically from the distribution of cluster sizes; the rest are considered spurious clusters and we treat them as unclustered. Finally, we feed the remaining cluster centers into the decoder to generate a prototype pitch contour for each cluster. Results ======= --------------------------------------------------------------------------------- [**Cluster**]{} [**T1**]{} [**T2**]{} [**T3**]{} [**T4**]{} ----------------------------- ------------ ------------ ------------ ------------ **A & 1 & 163 & 12 & 4\ **B & 108 & 0 & 0 & 1\ **C & 0 & 5 & 53 & 31\ **D & 1 & 0 & 0 & 97\ **N/A & 47 & 30 & 53 & 129\ ********** --------------------------------------------------------------------------------- : Cluster and tone frequencies for Mandarin.[]{data-label="tab:confusion-matrix-mandarin"} -------------------------------------------------------------------------------------------------------------------- [**Cluster**]{} [**T1**]{} [**T2**]{} [**T3**]{} [**T4**]{} [**T5**]{} [**T6**]{} -------------------------------------- ------------ ------------ ------------ ------------ ------------ ------------ **A & 5 & 5 & 59 & 109 & 7 & 105\ **B & 102 & 3 & 36 & 2 & 2 & 7\ **C & 93 & 0 & 0 & 2 & 0 & 0\ **D & 0 & 64 & 4 & 3 & 2 & 11\ **E & 0 & 28 & 2 & 4 & 30 & 2\ **N/A & 70 & 39 & 51 & 45 & 15 & 49\ ************ -------------------------------------------------------------------------------------------------------------------- : Cluster and tone frequencies for Cantonese.[]{data-label="tab:confusion-matrix-cantonese"} Figure \[fig:representations-contours\] shows the latent space learned by the autoencoders and the clustering output. Our model found 4 tone clusters in Mandarin, matching the number of phonemic tones (Table \[tab:confusion-matrix-mandarin\]) and 5 in Cantonese, which is one fewer than the number of phonemic tones (Table \[tab:confusion-matrix-cantonese\]). In Mandarin, the 4 clusters correspond very well with the the 4 phonemic tone categories, and the generated contours closely match the ground truth in Figure \[fig:mandarin-cantonese-tones\]. There is some overlap between tones 3 and 4; this is because tone 3 is sometimes realized a low-falling tone without the final rise, a process known as half T3 sandhi [@tone-sandhi-book], thus, it may overlap with tone 4 (falling tone). In Cantonese, the 5 clusters A-E correspond to low-falling, mid-level, high-level, mid-rising, and low-rising tones. Tone clustering in Cantonese is expected to be more difficult than in Mandarin because of 6 contrastive tones, rather than 4. The model is more effective at clustering the higher tones (1, 2, 3), and less effective at clustering the lower tones (4, 5, 6), particularly tone 4 (low-falling) and tone 6 (low-level). This confirms the difficulties in prior work, which reported worse classification accuracy on the lower-pitched tones because the lower region of the Cantonese tone space is more crowded than the upper region [@cantonese-svm]. Two other sources of error are carry-over and declination effects. A carry-over effect is when the pitch contour of a tone undergoes contextual variation depending on the preceding tone; strong carry-over effects have been observed in Mandarin [@carry-over]. Prior work [@levow2006] avoided carry-over effects by using only the second half of every syllable, but we do not consider language-specific heuristics in our model. Declination is a phenomenon in which the pitch declines over an utterance [@yip-tone; @cantonese-svm]. This is especially a problem in Cantonese, which has tones that differ only on pitch level and not contour: for example, a mid-level tone near the end of a phrase may have the same absolute pitch as a low-level tone at the start of a phrase. [**First Syllable**]{} [**All Syllables**]{} ------------------- ------------------------ ----------------------- [**Mandarin**]{} 0.738 0.641 [**Cantonese**]{} 0.515 0.464 : Normalized mutual information (NMI) between cluster assignments and ground truth tones, considering only the first syllable of each word, or all syllables.[]{data-label="tab:nmi"} To test this hypothesis, we evaluate the model on only the first syllable of every word, which eliminates carry-over and declination effects (Table \[tab:nmi\]). In both Mandarin and Cantonese, the clustering is more accurate when using only the first syllables, compared to using all of the syllables. Conclusions and future work =========================== We propose a model for unsupervised clustering and discovery of phonemic tones in tonal languages, using spoken words as input. Our model extracts the F0 pitch contour, trains a convolutional autoencoder to learn a low-dimensional representation for each contour, and applies mean shift clustering to the resulting latent space. We obtain promising results with both Mandarin Chinese and Cantonese, using only 400 spoken words from each language. Cantonese presents more difficulties because of its larger number of tones, especially at the lower half of the pitch range, and also due to multiple contrastive level tones. Finally, we briefly explore the influence of contextual variation on our model. A limitation of this study is that our model only considers pitch, which is only one aspect of tone. In reality, pitch is determined not only by tone, but by a complex mixture of intonation, stress, and other prosody effects. Tone is not a purely phonetic property – it is impossible to determine on a phonetic basis whether two pitch contours have distinct underlying tones, or are variants of the same underlying tone (perhaps in complementary distribution). Instead, two phonemic tones can be shown to be contrastive only by providing a minimal pair, where two semantically different lexical items are identical in every respect other than their tones. The last problem is not unique to tone: similar difficulties have been noted when attempting to identify consonant and vowel phonemes automatically [@phoneme-is-hard]. In future work, we plan to further explore these issues and develop more nuanced models to learn tone from speech. Acknowledgments =============== We thank Prof Gerald Penn for his help suggestions during this project. Rudzicz is a CIFAR Chair in AI. [^1]: The numbers are Chao tone numerals, where 1 is the lowest and 5 is the highest pitch. [^2]: `http://shtooka.net/`. We use the `cmn-caen-tan` dataset. [^3]: `https://forvo.com/` [^4]: `https://en.wiktionary.org/`
{ "pile_set_name": "ArXiv" }
--- abstract: 'Particle motion is considered in incompressible two-dimensional flows consisting of a steady background gyre on which an unsteady wave-like perturbation is superimposed. A dynamical systems point of view that exploits the action–angle formalism is adopted. It is argued and demonstrated numerically that for a large class of problems one expects to observe a mixed phase space, i.e., the occurrence of “regular islands” in an otherwise “chaotic sea.” This leads to patchiness in the evolution of passive tracer distributions. Also, it is argued and demonstrated numerically that particle trajectory stability is largely controlled by the background flow: trajectory instability, quantified by various measures of the “degree of chaos,” increases on average with increasing $\left|\mathrm{d}\omega/\mathrm{d}I\right|$, where $\omega (I)$ is the angular frequency of the trajectory in the background flow and $I$ is the action.' author: - 'Francisco J. Beron-Vera, María J. Olascoaga and Michael G. Brown' date: 'Manuscript version from .' title: 'Passive tracer patchiness and particle trajectory stability in incompressible two-dimensional flows' --- Introduction ============ This paper deals with the kinematics of fluid particles in unsteady incompressible flows on the Cartesian plane. Namely, we study properties of trajectories $(x(t),y(t))$ that satisfy equations of the form \[sys\] $$\dot{x}=\partial _{y}\psi ,\quad \dot{y}=-\partial _{x}\psi ,$$where the overdot stands for time derivative and $\psi (x,y,t)$ is the streamfunction. Furthermore, we consider the latter to be split into a steady background component and an unsteady perturbation component, i.e., $$\psi =\psi ^{(0)}(x,y)+\varepsilon \psi ^{(1)}(x,y,t),$$ where $\varepsilon $ is a dimensionless parameter. Equations (\[sys\]) constitute a canonical Hamiltonian system with $\psi $ the Hamiltonian and $(x,y)$ the generalized coordinate–conjugate momentum pair. Two related issues are addressed in this paper. First, we investigate a cause of “patchiness” in passive tracer distributions, i.e., distributions that are mostly vigorously stirred but include poorly stirred regions (Sect. \[pat\]). Second, we study the influence of the background flow on particle trajectory stability (Sect. \[sta\]). Prior to discussing these issues, the kinematic models that we use to illustrate our results are briefly described in Sect. \[kin\]. The conclusions of the paper are given in Sect. \[con\]. Kinematic models\[kin\] ======================= Two background flow structures in a region $[0,L]\times \lbrack 0,W]$ of the $\beta $ plane are considered here. One is chosen to represent a large-scale single-gyre wind-driven ocean circulation with streamfunction given by [@Stommel-66]$$\text{S}:\psi ^{(0)}=a\left[ b\mathrm{e}^{b_{+}x}+(1-b)\mathrm{e}^{b_{-}x}-1\right] \sin \frac{\pi y}{W},$$where $a:=\tau W/(\pi \lambda D),$ $b:=(1-\mathrm{e}^{b_{-}L})/(\mathrm{e}^{b_{+}L}-\mathrm{e}^{b_{-}L})$, and $b_{\pm }:=-\frac{1}{2}\beta /\lambda \pm \frac{1}{2}[(\beta /\lambda )^{2}+(\pi /W)^{2}]^{\frac{1}{2}}$. Here, $D$ is the depth, $\tau $ the wind stress amplitude (per unit density), and $\lambda $ the bottom friction. The other background streamfunction chosen corresponds to solid body rotation,$$\text{R}:\psi ^{(0)}=\frac{\omega _{\mathrm{R}}}{2}\left[ \left( x-L/2\right) ^{2}+\left( y-W/2\right) ^{2}\right] .$$The reason for this highly idealized choice will be discussed below. Parameter values used in our numerical work are listed in Table \[par\]. \[par\] Parameter Value ----------- ---------------------------------- $10$ $2\pi $ $200$ $9.8$ $^{-2}$ $10^{-4}$ $^{-1}$ $10^{-11}$ $^{-1}$$^{-1}$ $2\times 10^{-3}$ $^{-2}$$^{-2}$ $10^{-5}$ $^{-1}$ $2\pi $ $^{-1}$ : Background flow parameters. The perturbation streamfunction is constructed by superposing standing Rossby-like modes with a power-law spectrum, namely, $$\psi ^{(1)}\hspace{-0.01in}=\hspace{-0.01in}a\sum_{k,l}A\mathrm{e}^{-\gamma x}\hspace{-0.01in}\sin (kx\hspace{-0.01in}+\hspace{-0.01in}\phi _{k})\sin (ly\hspace{-0.01in}+\hspace{-0.01in}\phi _{l})\cos (\sigma t\hspace{-0.01in}+\hspace{-0.01in}\phi _{\sigma }), \label{per}$$where $$\begin{gathered} A(k,l):=\frac{\pi ^{2}(L^{-2}+W^{-2})}{k^{2}+l^{2}}, \\ \sigma (k,l):=-\frac{\beta k}{k^{2}+l^{2}+f_{0}^{2}/(gD)},\end{gathered}$$and the $\phi (k,l)$’s are random numbers uniformly distributed between $0$ and $2\pi .$ Here, $Lk/\pi $ and $Wl/\pi \ $are positive integers; $\gamma $ is a constant; $f_{0}$ is the reference Coriolis parameter; and $g$ is the acceleration of gravity. Dashed lines in Figs. \[str\]a,b and Fig. \[str\]c are streamlines for background flows S and R, respectively. Solid lines in these figures are total flow streamlines corresponding to a snapshot of the flow at $t\approx 9 $ $\mathrm{y}$. The perturbation in each case involves $10\times 10=100$ modes. In Figs. \[str\]a,c the perturbation has $\varepsilon =0.05$ and $\gamma =0.$ In Fig. \[str\]b the amplitude of the $Lk/\pi =1=Wl/\pi $ mode is set to zero, $\phi _{k}=0=\phi _{l}$ so the flow vanishes at the boundary, $\varepsilon =0.25,$ and $\gamma =0.4$ $\mathrm{Mm}^{-1}$. The flows used to produce Fig. \[str\] and all of the numerical particle trajectory simulations presented in this paper were chosen to illustrate important aspects of Lagrangian dynamics; the flows are in many ways not representative of realistic oceanic flows. We* *note, however, that we focus on flows with complicated time dependence, and that the strong perturbations to the background are considered. In Fig. \[str\]b, for example, it is seen that the perturbation leads to the presence of an eddy-like structure in the flow. Also, we note that in the flows that we have described, particle trajectories are periodic in the limit of zero perturbation strength with typical periods of about $1$ $\mathrm{y}$. Thus in an integration time of $10$ $\mathrm{y}$ most trajectories will have made approximately $10$ revolutions around the gyre. The phenomena described below are not limited to gyre-scale flows. In general, the trends that we describe should be evident after times in excess of a few periods of particle revolution in any background gyre flow on which a perturbation field with a broad band of frequencies is superimposed. Passive tracer patchiness\[pat\] ================================ In this section we present numerical evidence and a theoretical argument that suggest that for a large class of systems of the form (\[sys\]) phase space $(x,y)$ should be partitioned into “regular islands” in a “chaotic sea.” Such a mixed phase space leads to patchiness in passive tracer distributions. Numerical results are presented for a time-periodic flow \[$n=1 $ term in the sum in (\[per\])\] and subsequently for flows with complicated time dependence ($n$ large). Figure \[poi\] shows, for the time-periodic case ($n=1$), a Poincaré section and, in the same environment, two additional trajectory diagnostics whose applicability is not restricted to time-periodic flows. The Poincaré section was constructed by plotting the $(x,y)$ coordinates of several trajectories at integer multiples of the period of the streamfunction; it shows the usual mixture of “regular islands” in an otherwise “chaotic sea” [cf., e.g., @Tabor-89]. The middle panel shows, for a dense set of trajectories with $x(0)=x_{0}$ fixed and $y(0)=y_{0}$ variable, a plot of $y$ vs. $y_{0}$ at a fixed value of $t.$ The initial conditions chosen fall inside the region of the Poincaré section shown, and it is seen that both regular islands and the chaotic sea evident in the Poincaré section can be identified in the $y$ vs. $y_{0}$ plot. The same structures can also be seen in the lower panel of Fig. \[poi\] which shows, for the same trajectories used to produce the middle panel, finite time estimates of Lyapunov exponents (described in more detail below), $\nu $ vs. $y_{0}.$ Plots of $y$ vs. $y_{0}$ and $\nu $ vs. $y_{0}$ are used below to distinguish between apparently regular and apparently chaotic trajectories for flows with complicated ($n$ large) time dependence i.e., in flows for which a Poincaré section cannot be constructed. Figure \[alp\] shows plots of $y$ vs. $y_{0}$ and $\nu $ vs. $y_{0}$ for the nonperiodic flows used to produce Fig. \[str\]. Trajectories in Fig. \[alp\]b are generally more unstable than in Fig. \[alp\]a. The enhanced stability in Fig. \[alp\]a is reflected in a relatively unstructured $y(y_{0})$ plot and smaller (on average) Lyapunov exponents than are seen in Fig. \[alp\]b. In both cases the background flow structure is the same; the difference in the stability behavior is due to the difference in the strength of the perturbation. As expected, trajectory instability is seen to increase with increasing perturbation strength. The difference in trajectory stability seen in Figs. \[alp\]a,c has a different explanation. The same perturbation was used in both cases, so this cannot be the cause of the difference. The cause is the influence of the background flow; this topic will be discussed in detail in the following section. We return our attention now to Fig. \[alp\]b which corresponds to the strongly perturbed flow shown in Fig. \[str\]b. It is seen in Fig. [alp]{}b that embedded among mostly chaotic trajectories are bands of apparently nonchaotic trajectories. These nonchaotic bands are most readily* *identified among the trajectories whose initial positions are near the center of the gyre; the reason for this will be discussed in the following section. Bands of nonchaotic trajectories far from the gyre center are also present, however. This is seen in Fig. \[alp-blo\] where two regions of Fig. \[alp\]b are blown up. These apparently nonchaotic bands of trajectories are the counterparts of the “regular islands” seen in Fig. \[poi\]. Trajectories in these bands diverge only very slowly (power law dependence on time) from neighboring trajectories while chaotic trajectories diverge at an exponential rate from neighboring trajectories. The nonchaotic regions of flows are important in applications because they correspond to regions where the concentration of a passive tracer will remain high for a long duration. The existence of these regions leads to a large variance of tracer concentration or “tracer patchiness” [cf., e.g., @Pasmanter-88; @Malhotra-Mezic-Wiggins-98]. Another way to visualize passive tracer patchiness is offered in Figs. [lag]{} and \[lag-blo\]. Both figures show discrete samples of a material line of fluid at $t=0$ (vertical line segments in the figures) and at $t=12$ $\mathrm{y,}$ in the environments shown in Fig. \[str\]. Initial conditions in Fig. \[lag\] are as in Fig. \[alp\], whereas those in Fig. \[lag-blo\] are as in Fig. \[alp-blo\]. These figures again show that while most of the initial material line segment is vigorously stirred, there are small portions of the initial segment, corresponding to the island-like structures seen in Figs. \[alp\] and \[alp-blo\], that are poorly stirred. We turn our attention now to explaining the occurrence of island-like structures in Figs. \[poi\]–\[alp-blo\]. First, we note that in the background flow, particle motion is describable using action–angle variables, reviewed below, and trajectories fall on tori. For perturbed systems with periodic time dependence, as in Fig. \[poi\], it is well-known that particle trajectory dynamics are constrained by the KAM theorem [cf., e.g., @Arnold-89] which guarantees that for sufficiently small $\varepsilon $ some of the original tori—and associated nonchaotic motion—are preserved. Related theoretical results, generally known as KAM theory [cf., e.g., @Tabor-89], describe how the nonsurviving tori break up to form chains of “islands” surrounded by a “chaotic sea” as seen in Fig. \[poi\]. For a large perturbation strength $\varepsilon $ all of the original tori will have been broken up, but the secondary islands that are formed in the process are robust and persist even when the magnitude of the perturbation exceeds that of the background flow. It has been shown [@Brown-98; @Beigie-etal-91] that for multiply-periodic perturbations the situation is essentially the same as for perturbations with simple periodic time dependence. This follows from the observation that (\[sys\]), with $\psi (x,y,$* *$\sigma _{1}t,\cdots ,\sigma _{n}t)$ where $\sigma _{i}t$ is defined modulo $2\pi $, can be transformed to an autonomous Hamiltonian system with a bounded phase space with* *$(n+1)$ degrees of freedom that is constrained by $n$ integrals. KAM theory (the KAM theorem and related results) applies to the transformed system, so phase space is generically partitioned into nonintersecting regular and chaotic regions. A Poincaré section could, in principle, be constructed for such a system by using a multiple slicing technique [cf. @Parker-Chua-89] but slicing is practical only when $n=1$. The significance of the extension of KAM theory to multiply-periodic systems is that in the system defined by (\[sys\]) with $\psi ^{(1)}$ given by (\[per\]), phase space $(x,y)$ is expected to be partitioned into “regular islands” in an otherwise “chaotic sea.” The numerical evidence presented in Figs. \[alp\]–[lag-blo]{} supports this expectation. The coexistence of regular and chaotic fluid particle trajectories in mesoscale and large-scale oceanic flows has been suggested in some analyses of surface drifters and submerged floats [@Osborne-etal-86a; @Osborne-etal-89; @Richardson-etal-89; @Brown-Smith-90]. The preceding discussion provides an explanation of the underlying physics. Particle trajectory stability\[sta\] ==================================== In this section we describe the important influence of the background flow on particle trajectory stability that was mentioned above in our discussion of Figs. \[alp\]a,c. The ideas presented here apply to any canonical Hamiltonian system in which the Hamiltonian consists of a superposition of an integrable component and a nonintegrable perturbation. Other applications are described in Beron-Vera and Brown (2003a,b). [Beron-Brown-03,Beron-Brown-03b]{} The explanation of this behavior makes use of the action–angle description of the motion of particles in the background flow [cf., e.g., @Abdullaev-Zaslavsky-91]. Let $$I:=\frac{1}{2\pi }\oint \mathrm{d}x\,Y(x;\psi ^{(0)}),$$where $Y$ is the meridional coordinate of an isoline of $\psi ^{(0)}$, be the action variable, and consider the canonical transformation $(x,y)\mapsto (\vartheta ,I),$ defined implicitly by $$y=\partial _{x}G,\;\vartheta =\partial _{I}G,\;G(x,I):=\int \mathrm{d}x\,Y(x;\psi ^{(0)}),$$where $\vartheta $ is the angle variable. According to the above transformation,$$\psi (x,y,t)\mapsto \bar{\psi}^{(0)}(I)+\varepsilon \bar{\psi}^{(1)}(I,\vartheta ,t)$$and Eqs. (\[sys\]a) take the form $$\dot{I}=-\varepsilon \partial _{\vartheta }\bar{\psi}^{(1)},\quad \dot{\vartheta}=\omega +\varepsilon \partial _{I}\bar{\psi}^{(1)}, \label{act}$$where $$\omega (I):=\mathrm{d}\bar{\psi}^{(0)}/\mathrm{d}I. \label{ome}$$ When $\varepsilon =0$, Eqs. (\[act\]), which have one degree of freedom, are autonomous and the corresponding Hamiltonian, $\bar{\psi}^{(0)},$ is an integral of motion that constrains the dynamics. As a consequence, the equations can be solved by quadratures and the motion is periodic with angular frequency $\omega $. Namely, $I=I_{0}$ and $\vartheta =\vartheta _{0}+\omega t$ $\func{mod}2\pi $, where $I_{0}$ and $\vartheta _{0}$ are constants. Every solution curve is thus a line that winds around an invariant one-dimensional torus $\{I_{0}\}\times T^{1}\subset \mathbb{R}\times T^{1}$, whose representation in $(x,y)$-space is the closed curve given by the isoline $\psi ^{(0)}=\bar{\psi}^{(0)}(I_{0}).$ With the perturbation term, the corresponding Hamiltonian, $\bar{\psi}^{(0)}+\varepsilon \bar{\psi}^{(1)},$ is no longer an integral of motion (the equations are nonautonomous) and the system may be sensitive to initial* *conditions, thereby leading to chaotic motion. The distinction between regular and chaotic trajectories is commonly quantified by the* *Lyapunov exponent [cf., e.g., @Parker-Chua-89], a measure of the rate at which neighboring trajectories diverge,$$\nu _{\infty }:=\lim\limits_{t\rightarrow \infty }\dfrac{1}{t}\ln |\nu ^{\mathsf{Q}}|\text{\textit{,}} \label{lya}$$where $\nu ^{\mathsf{Q}}(t)$ is the largest of the two eigenvalues of the so-called stability matrix $\mathsf{Q}(t)$, which is given by $$\mathsf{Q}:=\left[ \begin{array}{cc} \partial _{I_{0}}I & \partial _{\vartheta _{0}}I \\ \partial _{I_{0}}\vartheta & \partial _{\vartheta _{0}}\vartheta\end{array}\right] .$$ Because of the area preservation property of Eqs. (\[sys\]a) or (\[act\]) the product of the two eigenvalues of $\mathsf{Q}$ is unity, so there is no loss of generality in considering only the largest eigenvalue. Each column of $\mathsf{Q}$ corresponds to a vector perturbation $(\delta I,\delta \vartheta )$ to a trajectory in the nonautonomous system (\[act\]), and satisfies the so-called variational equations,$$\left( \begin{array}{c} \delta \dot{I} \\ \delta \dot{\vartheta}\end{array}\right) \hspace{-0.02in}\hspace{-0.02in}=\hspace{-0.02in}\hspace{-0.02in}\left[ \begin{array}{cc} 0 & 0 \\ \omega ^{\prime } & 0\end{array}\right] \hspace{-0.02in}\hspace{-0.02in}\left( \begin{array}{c} \delta I \\ \delta \vartheta\end{array}\right) \hspace{-0.02in}\hspace{-0.02in}+\hspace{-0.02in}\varepsilon \hspace{-0.02in}\hspace{-0.02in}\left[ \begin{array}{rr} -\partial _{I\vartheta }\bar{\psi}^{(1)} & -\partial _{\vartheta \vartheta }\bar{\psi}^{(1)} \\ \partial _{II}\bar{\psi}^{(1)} & \partial _{I\vartheta }\bar{\psi}^{(1)}\end{array}\right] \hspace{-0.02in}\hspace{-0.02in}\left( \begin{array}{c} \delta I \\ \delta \vartheta\end{array}\right) , \label{var}$$where $\omega ^{\prime }:=\mathrm{d}\omega /\mathrm{d}I.$ Equations ([var]{}) and (\[act\]) constitute a system of four coupled equations. Variational equations that describe the growth of perturbations using Cartesian coordinates $(\delta x,\delta y)$ have the same form as (\[var\]) except that in the Cartesian form all four elements of the first matrix on the r.h.s. of (\[var\]) are generally nonzero. Our numerical finite-time Lyapunov exponent estimates are based on the Cartesian equivalent of Eqs. (\[var\]) and (\[act\]), which is generally more convenient for numerical calculations. We have chosen to show the $(\delta I,\delta \vartheta )$ form of these equations to highlight the important role played by $\omega ^{\prime }$. An example of a closely related study which does not exploit action–angle variables, and which consequently overlooks the critical importance of $\omega ^{\prime }$, is @Richards-etal-95. A simple but very important observation follows from the action–angle formalism. Dependence of both particle (\[act\]) and variational (\[var\]) equations on the background flow enters only through the function $\omega (I)$. Equations (\[var\]) strongly suggests that trajectory stability and $\omega ^{\prime }$ are closely linked. The following heuristic argument explains the mechanism by which $\omega ^{\prime }$ is expected to control trajectory stability. If one assumes that $\varepsilon $ is small and the second derivatives of $\bar{\psi}^{(1)}$ are zero-mean random variables, then when $\omega ^{\prime }=0$ these terms should lead to slow (power-law) growth of $\delta \vartheta $ and $\delta I.$ If $\left| \omega ^{\prime }\right| $ is large, this term will cause a rapid growth of $\left| \delta \vartheta \right| $ for any nonzero $\left| \delta I\right| $. The perturbation terms will then lead to a mixing of $\left| \delta \vartheta \right| $ and $\left| \delta I\right| $. The term $\omega ^{\prime }$ will lead, in turn, to further growth of $\left| \delta \vartheta \right| .$ As this process repeats itself, both $\left| \delta I\right| $ and $\left| \delta \vartheta \right| $ are expected to grow rapidly. The role played by $\omega ^{\prime }$ in this process is to amplify small perturbations caused by the second term on the r.h.s. of Eqs. (\[var\]). Thus when $\varepsilon $ is small, trajectory instability is expected to be significantly enhanced when $\left| \omega ^{\prime }\right| $ is large. When $\varepsilon $ is sufficiently large that the two terms on the r.h.s. of Eqs. (\[var\]) have comparable magnitude, the role played by $\left| \omega ^{\prime }\right| $ in amplifying perturbations is expected to be much less important. Increased trajectory instability should result in larger numerical estimates of Lyapunov exponents. A dynamical-systems-based argument on the role of $\omega ^{\prime }$ in controlling trajectory stability is given below; that argument is consistent with the above heuristic argument. The lower panels of Fig. \[alp\] show the absolute value of the stability parameter [@Zaslavsky-98; @Beron-Brown-03; @Beron-Brown-03b]$$\alpha (I):=\frac{I}{\omega }\frac{\mathrm{d}\omega }{\mathrm{d}I} \label{def}$$as a function of trajectory initial condition; recall that these initial conditions correspond to variable $y_{0}$ with $x_{0}$ fixed at the gyre center. Comparison of the middle and lower panels of Fig. \[alp\]a suggests that when the perturbation to the background steady flow is weak, trajectory instability increases, on average, with increasing $\left| \alpha \right| $. Figure \[alp\]b shows that for a strong perturbation this trend is less strong, although the most stable trajectories are clearly those in the region of the flow where $\left| \alpha \right| $ is small. The background flow R used to produce Fig. \[alp\]c was chosen because it has the property $\alpha =0$ $\forall I$. Because the same perturbation flows were used to produce Fig. \[alp\]a and \[alp\]c the difference between these figures is entirely due to the difference in the background flows. The remarkable stability of trajectories in Fig. \[alp\]c is due to the property $\alpha =0$ $\forall I$ in flow R. The same comment applies to the difference between Fig. \[lag\]a and \[lag\]c, which were produced using the same flows that were used to produce Fig. \[alp\]a and \[alp\]c. All of the aforementioned observations relating to Figs. \[alp\] and \[lag\] are consistent with the heuristic argument given in the preceding paragraph describing how $\left| \omega ^{\prime }\right| $ is expected to control trajectory stability. The physical interpretation of the stability parameter $\alpha $ is illustrated in Fig. \[she\]. Figures \[she\]a,b show the evolution of a material line of fluid in background flows S and R, respectively. The material line is shown at $t=0$ and at $t=12$ Also shown is a plot of both $2\pi /\omega $ and $\left| \alpha \right| $ as a function of $y_{0}$ (for $y_{0}>y_{\mathrm{C}},$ the meridional coordinate of the center gyre) in each environment. As a consequence of the uniqueness of solutions to Eq. (\[sys\]) and continuity of the velocity field, the material line of fluid cannot break or intersect itself but it can increase in complexity with time. Because the motion in Fig. \[she\] is integrable (i.e., each point of the material line is constrained to lie on a surface of constant $I$ $\forall t$) and because attention is restricted to background flows for which $\psi ^{(0)}$ has compact and closed level sets, i.e., gyre flows, the length of the material line can grow with time, at most, following a power law. Background flow R has a special property. In that background flow the material line just rotates clockwise at a constant rate $\omega =\omega _{\mathrm{R}}$ $(=2\pi $ $\mathrm{y}^{-1})$, independent of $I$, so $\alpha =0$ $\forall I.$ In contrast, $\omega $ varies with $I$ in background flow S. The monotonic decay of $\omega $ as a function of $I$ in background flow S induces a* *shear in phase space which causes the outermost points of the material line to rotate more slowly than the innermost ones and, hence, causes the material line to spiral. In background flow R there is no shear. In polar coordinates radial shear can be defined as $$r\partial _{r}\left( r^{-1}u_{\theta }\right) , \label{rad}$$where $r$ is the radial coordinate and $u_{\theta }$ is the $\theta $-component of the velocity field. More correctly, this quantity is twice the $r\theta $-component of the strain-rate tensor for rotational motion [cf., e.g., @Batchelor-64]. The connection with motion in phase space can be accomplished by identifying $I$ with $r$ and $\omega I$ with $u_{\theta }$. The replacements $r\mapsto I$ and $u_{\theta }\mapsto \omega I$ in (\[rad\]) thus give the analogous expression $I\omega ^{\prime }$ for the shear in phase space. Notice that this expression is (apart from the $\omega ^{-1}$-factor) the stability parameter $\alpha $. We have chosen to include the $\omega ^{-1}$-factor in the definition of $\alpha $ because of precedent [@Zaslavsky-98; @Beron-Brown-03; @Beron-Brown-03b] and because it is convenient to make $\alpha $ dimensionless. To see the importance of the shear in the background flow, compare Figs. [she]{}a,b with Figs. \[lag\]a,c, which show the evolution of the same initial material line segments in the total (background plus perturbation) flows. Notice the highly complicated structure of the segment in the perturbed flow S (Fig. \[lag\]a) as compared to that in the unperturbed one (Fig. \[she\]a). (Note that the number of particles used to produce Fig. \[lag\]a is far too small to resolve what should be an unbroken smooth curve which does not intersect itself.) In contrast, observe that in environment R the perturbation has only a very minor effect on the evolution of the material line (Figs. \[lag\]c and \[she\]b). Additional insight into why $\alpha $ should be expected to control trajectory stability comes from the following argument. The perturbation streamfunction $\varepsilon \psi ^{(1)}$ has the effect of introducing perturbations to the action $I$ of a given particle by the amount $\delta I$. If $\delta I$ is assumed to be small and of the same order as the perturbation streamfunction \[$\delta I=O(\varepsilon ),$ say\], then $\omega $ experiences the change $$\omega \mapsto \left( 1+\alpha \delta I/I\right) \omega$$$+$ $O(\varepsilon ^{2}).$ The perturbation to $\omega $ depends on both the perturbation $\delta I$ and the background flow via $\alpha $. Under the change $I\mapsto I+\delta I,$ a sufficient condition for $\omega $ to remain invariant at $O(\varepsilon )$ is $\alpha =0.$ This provides an explanation for the remarkable stability of the particle trajectories in flow R. To $O(\varepsilon )$ a nonvanishing shear $(\alpha \neq 0)$ appears as a necessary condition to sustain the successive stretching and folding of the material line of fluid after it gets distorted by the perturbation. (Of course, chaotic motion is still possible when $\alpha =0$ provided that $\varepsilon $ is sufficiently large.) It is thus expected that where $|\alpha |$ is small (resp., large) there will be less (resp., more) sensitivity to initial conditions and, hence, the motion be more regular (resp., chaotic). Support for this conjecture is given in the numerical simulations presented in this paper. Finally, the role of $\omega ^{\prime }$ in dynamical systems theory deserves further comment. A nondegeneracy condition, $\omega ^{\prime }\neq 0,$ must be satisfied in order for the KAM theorem to apply and, hence, to guarantee that some trajectories are nonchaotic provided the strength of the time-dependent perturbation is sufficiently weak. This theorem does not imply, however, that trajectories are unstable when $\omega ^{\prime }=0;$ the KAM theorem does not address this limit. The mechanism that leads to chaos is the excitation of resonances at discrete frequencies. For a sufficiently strong perturbation, neighboring resonances overlap and chaotic motion results [cf., e.g., @Tabor-89]. The width in frequency of each resonance is proportional to $|\omega ^{\prime }|^{1/2},$ so one expects, on average, motion to become increasingly chaotic as $\left| \omega ^{\prime }\right| $ increases. This expected trend is consistent with the arguments and numerical simulations that we have presented above. The trend toward increasingly chaotic motion with increasing $\left| \omega ^{\prime }\right| $ does not, of course, rule out some nonchaotic motion for fixed but large $\left| \omega ^{\prime }\right| .$ Note also that the trends that we have described apply on average; details depend on details of the flow, both background and perturbation. Concluding remarks\[con\] ========================= In this paper we considered particle motion in unsteady incompressible two-dimensional flows consisting of a steady background gyre on which a highly structured unsteady wave-like perturbation is superimposed. The numerical simulations presented strongly suggest that: (i) phase space is mixed, characteristic of near-integrable one-and-a-half-degree-of-freedom Hamiltonian systems; and (ii) particle trajectory stability strongly depends on the structure of the background (steady) component of the flow. The mixed phase space structure, in which “islands” of stability emerge from an otherwise chaotic “sea,” was explained as a consequence of the applicability of KAM theory. The mixed phase space provides an explanation for the occurrence of patches of poorly stirred fluid in a mostly vigorously stirred flow. Trajectory instability was shown to increase with increasing magnitude of $\alpha :=I\omega ^{\prime }/\omega ,$ where $2\pi /\omega (I)$ is the period of revolution of a particle in the background gyre flow and $I$ is the particle’s action variable in the background flow. These results provide important insight into the physics underlying Lagrangian ocean dynamics. In addition to this insight, the results described are potentially important in a variety of practical problems. The occurrence of Lagrangian “islands of stability” has important implications for the transport and dispersal of tracers ranging from nutrients to toxic pollulants. Knowledge that such “islands” are smaller and less abundant, on average, in regions of flows where $\left| \omega ^{\prime }\right| $ is large might be exploited when deciding where to place a sewage outfall, for example. The comments of an anonymous reviewer have led to improvements in the manuscript. This work has been supported by Code 321OA of the US Office of Naval Research. [18]{} natexlab\#1[\#1]{}url \#1[[\#1]{}]{}urlprefix Abdullaev, S. S. and Zaslavsky, G. M., Classical nonlinear dynamics and chaos of rays in wave propagation problems in inhomogeneous media, [*Usp. Phys. Nauk*]{}, [*161*]{}, 1–43, 1991. Arnold, V. I., [*Mathematical Methods of Classical Mechanics*]{}, Springer, 2nd edn., 1989. Batchelor, G. K., [*An Introduction to Fluid Dynamics*]{}, Cambridge University, 1964. Beigie, D., Leonard, A., and Wiggins, S., Chaotic transport in the homoclinic and heteroclinic tangle regions of quasiperiodically forced two-dimensional dynamical systems, [*Nonlinearity*]{}, [*4*]{}, 775–819, 1991. Beron-Vera, F. J. and Brown, M. G., Ray stability in weakly range-dependent sound channels, [*J. Acoust. Soc. Am.*[,]{} in press (e-print nlin.CD/0208038)** ]{}, 2003. Beron-Vera, F. J. and Brown, M. G., Travel time stability in weakly range-dependent sound channels, [*J. Acoust. Soc. Am.*[,]{} submitted (e-print nlin.CD/0307002)** ]{}, 2003. Brown, M. and Smith, K., Are sofar trajectories chaotic?, [*J. Phys. Oceanogr.*]{}, [*20*]{}, 139–149, 1990. Brown, M. G., Phase space structure and fractal trajectories in $1\frac{1}{2}$ degree of freedom [H]{}amiltonian systems whose time dependence is quasiperiodic, [*Nonlin. Proc. Geophys.*]{}, [*5*]{}, 69–74, 1998. Malhotra, N., Mezić, I., and Wiggins, S., Patchiness: A new diagnostic for [L]{}agrangian trajectory analysis in time-dependent fluid flows, [*Int. J. Bif. Chaos*]{}, [*8*]{}, 1053–1093, 1998. Osborne, A. R., Kirwan, A. D., Provenzale, A., and Bergamasco, L., A search for chaotic behavior in large and mesoscale motions in the pacific ocean, [ *Physica*]{}, [*23D*]{}, 75–83, 1986. Osborne, A. R., Kirwan, A. D., Provenzale, A., and Bergamasco, L., Fractal drifter trajectories in the kuroshio extension, [*Tellus*]{}, [*41A*]{}, 416–435, 1989. Parker, T. S. and Chua, L. O., [*Practical Numerical Algorithms for Chaotic Systems*]{}, Springer, 1989. Pasmanter, R., Anomalous diffusion and anomalous stretching in vortical flows, [*Fluid Dyn. Res.*]{}, [*3*]{}, 320–326, 1988. Richards, K. J., Jia, Y., and Rogers, C. F., Dispersion of tracers by ocean gyres, [*J. Phys. Oceanogr.*]{}, [*25*]{}, 873–887, 1995. Richardson, P. L., Walsh, D., Armi, L., Schr[ö]{}der, M., and Price, J. F., Tracking three meddies with [SOFAR]{} floats, [*J. Phys. Oceanogr.*]{}, [ *19*]{}, 371–383, 1989. Stommel, H., [*The [G]{}ulf [S]{}tream*]{}, University of California, 2nd edn., 1966. Tabor, M., [*Chaos and Integrability in Nonlinear Dynamics*]{}, John Wiley and Sons, 1989. Zaslavsky, G. M., [*Physics of Chaos in Hamiltonian Systems*]{}, Imperial College, 1998.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study the formation and final structure of the rare earth peak ($A\sim160$) of the $r$-process nucleosynthesis. The rare earth peak forms at late times in the $r$-process after neutron exhaustion (neutron-to-seed ratio unity or $R=1$) as matter decays back to stability. Since rare earth peak formation does not occur during [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium it is sensitive to the strong interplay between late time thermodynamic evolution and nuclear physics input. Depending on the conditions the peak forms either because of the pattern of the neutron capture rates or because of the pattern of the separation energies. We analyze three mass models under different thermodynamic conditions. We find that the subtleties of each mass model, including separation energies and neutron capture rates, influence not only the final shape of the peak but also when it forms. We identify the range of nuclei which are influential in rare earth peak formation.' author: - 'Matthew R. Mumpower' - 'G. C. McLaughlin' - Rebecca Surman bibliography: - 'repf.bib' title: 'Formation Of The Rare Earth Peak: Gaining Insight Into Late-Time r-Process Dynamics' --- Introduction {#intro} ============ Approximately half of the elements beyond $A>100$ are made in the rapid neutron capture process, or $r$-process, in which successive neutron captures occur on timescales faster than $\beta$-decays. At the present time, there is significant uncertainty with the astrophysical environment responsible for this synthesis event [@Arnould200797; @Qian2007237]. The leading candidate site [@argast2004] is believed to be core-collapse supernovae e.g [@1992ApJ...399..656M; @1996ApJ...471..331Q; @2000PASJ...52..601S; @2000ApJ...533..424O; @2001ApJ...562..887T; @2005NuPhA.752..550Q; @2011PrPNP..66..346T] even though most recent simulations do not yield favorable conditions for the $r$-process . Other candidate sites include compact object mergers [@1977ApJ...213..225L; @1999ApJ...525L.121F; @2005NuPhA.758..587G; @2008ApJ...679L.117S; @2010MNRAS.406.2650M; @2011ApJ...738L..32G; @2011arXiv1106.6142W; @Caballero2011arXiv1105.6371C], gamma-ray burst outflows [@2004ApJ...603..611S; @2005NuPhA.758..189M; @2006ApJ...643.1057S], neutrino induced nucleosynthesis in He shells [@PhysRevLett.106.201104; @2011PhRvL.106t1104B], supernova fallback [@2006ApJ...646L.131F], and collapse of O-Ne-Mg cores [@1998ApJ...493L.101W; @2003ApJ...593..968W; @2007ApJ...667L.159N]. Experimentally, it is difficult to measure the properties of the short-lived nuclei far from stability that participate in the $r$-process. Recent developments using radioactive beams show promise (e.g. [@PhysRevLett.94.112501; @2009AIPC.1098..153J]), but current experimental data on neutron-rich isotopes is limited. Thus $r$-process studies must rely not only on model calculations of the environment, but also on theoretical mass models, e.g. [@Moller1995; @1996PhLB..387..455P; @2009PhRvL.102o2503G]. Despite these difficulties, much has been learned about the $r$-process over the past 50 years. The most prominent features in the $r$-process abundance distribution above atomic mass number of $A=100$ are two distinct peaks occurring at $A=130$ and $A=195$. It was hypothesized very early that the formation of these peaks should be associated with the long $\beta$-decay rates of closed neutron shells [@1957RvMP...29..547B]. Since this seminal paper much effort has been put in to researching the conditions for a sufficient initial neutron-to-seed ratio, a key requirement in order to produce a main $r$-process out to the third peak ($A=195$). For reviews see [@Arnould200797; @2003PrPNP..50..153Q; @Cowan:1991zz; @Qian2007237]. After the two main peaks, the second most prominent feature above A=100 is the smaller peak near $A\sim160$ known as the rare earth peak. While less abundant than the other peaks, the rare earth peak can in principle be used as a powerful tool and offers an alternative way to probe the $r$-process. This is due to the following properties: (1) Observational data from metal-poor stars show very consistent trends among the rare earth and heavier elements. This suggests that these elements were created in the same type of synthesis event . Thus, the rare earth peak provides a natural diagnostic of $r$-process models. (2) The rare earth peak forms away from closed neutron or proton shells and therefore by a different mechanism than the main peaks. This means it is a different and unique probe of late-time $r$-process conditions. (3) The rare earth peak is extremely sensitive not only to late-time thermodynamic behavior, but also to nuclear physics input [@Surman:1997; @Mumpower2010ncr; @2011PhRvC..83d5809A]. Typical variations in final rare earth abundance patterns from simulations with different mass models are highlighted in Figure \[fig:rep\]. To date the rare earth region has received relatively little attention. Fission cycling has been suggested as a mechanism for obtaining the rare earth peak [@1957PASP...69..201C; @1971Natur.231..103S], but it is not favored . Large uncertainties found in fission probabilities and fragment distributions of current nuclear models further compound difficulties with a successful description of rare earth peak formation by fission cycling [@2011ApJ...738L..32G]. Surman et al. [@Surman:1997] investigated the formation of the peak in a hot $r$-process environment with temperatures high enough to support [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium. The formation of the rare earth peak under these conditions was attributed to the co-action of nuclear deformation and $\beta$-decay as the free neutrons are quickly captured during freeze-out. This was followed by a study of late-time abundances changes among the major peaks [@2001PhRvC..64c5801S]. Otsuki et al. [@2003NewA....8..767O] investigated a range of r-process models and found similar rare earth elemental abundance patterns, provided the temperature was constant during freeze-out. Most recently, Arcones et al. [@2011PhRvC..83d5809A] studied the sensitivity of late-time abundance fluctuations to changes in the nuclear physics inputs. Arcones et al. [@2011PhRvC..83d5809A] pointed out that the rare earth peak is sensitive to changes at late-times, e.g. to non-equilibrium effects such as neutron capture even when the abundance of free neutrons can become very low ($\sim10^{-5}$). This manuscript presents a more complete picture of rare earth peak formation. We explore the sensitivity of the peak formation mechanism to late-time thermodynamic behavior and nuclear physics input. The funneling formation mechanism of [@Surman:1997] is reviewed for hot evolutions. We introduce a different trapping mechanism for peak formation in cold evolutions where the temperatures and densities decline relatively quickly and therefore photo-dissociation plays no role in the late-time dynamics after $R=1$. We study the effects of three different mass models and show how large uncertainties in this region stem from nuclear physics. Lastly, we show that the nuclei which contribute to peak formation are approximately 10 to 15 neutrons from stability, and thus represent prime candidates to be measured in future radioactive ion beam facilities (FRIB [^1] or FAIR [^2]). r-Process Conditions and Calculations ===================================== Abundance weighted lifetimes are used throughout the text to characterize the late-time dynamics of the $r$-process. These are provided below for the reader’s convenience: $$\label{eqn:tau-ng} \tau_{n\gamma} \equiv \frac{\sum_{Z\geqslant8,A}Y(Z,A)}{\sum_{Z\geqslant8,A}N_{n}\langle \sigma v\rangle_{Z,A} Y(Z,A)}$$ $$\label{eqn:tau-gn} \tau_{\gamma n} \equiv \frac{\sum_{Z\geqslant8,A}Y(Z,A)}{\sum_{Z\geqslant8,A}\lambda_{\gamma n}(Z,A)Y(Z,A)}$$ $$\label{eqn:tau-beta} \tau_{\beta} \equiv \frac{\sum_{Z\geqslant8,A}Y(Z,A)}{\sum_{Z\geqslant8,A}\lambda_{\beta}(Z,A)Y(Z,A)}$$ where $N_{n}$ is the neutron number density, $\langle \sigma v\rangle_{Z,A}$ the thermally averaged neutron capture cross section for nuclei $(Z,A)$, $\lambda_{\gamma n}(Z,A)$ the photo-dissociation rate for nuclei $(Z,A)$, $\lambda_{\beta}(Z,A)$ the full $\beta$-decay rate (including $\beta$-delayed neutron emission channels) for nuclei $(Z,A)$ and $Y(Z,A)$ the abundance of nuclei $(Z,A)$. A reduced sum denoted with a superscript “REP” is taken over the rare earth region, $A=150$ to $A=180$, when applicable. The neutron-to-seed ratio or $R$ is defined as: $$\label{eqn:r} R \equiv \frac{Y_{n}}{\sum_{Z\geqslant8,A}Y(Z,A)}$$ where $Y_{n}$ is the abundance of free neutrons. Since rare earth peak formation is highly dependent on the rate of decrease in the temperature and density, we consider rare earth peak formation under two different thermodynamic evolutions. One scenario is a classical hot $r$-process which operates under high temperatures ($T_{9}\gtrsim1$) at the time in which neutron captures are important for peak formation. A second scenario is a cold $r$-process which operates under low temperatures ($T_{9}\sim0.5$) at the time in which neutron captures are important for peak formation [@2007ApJ...666L..77W]. The classical $r$-process begins with a phase of [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium marked by an abundance weighted lifetime ratio of neutron capture to photo-dissociation of [$\tau_{n\gamma}/\tau_{\gamma n}$]{}$=1$. During this phase the temperature is still sufficiently high so that neutron captures dominate $\beta$-decays ([$\tau_{\beta}/\tau_{n\gamma}$]{}$\gg1$) and the Saha equation can be used to determine abundances along an isotopic chain [@Cowan:1991zz]. The second phase, known as the freeze-out epoch, is marked by the weakening of the [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium ([$\tau_{n\gamma}/\tau_{\gamma n}$]{}$\lesssim1$) and the abundance weighted lifetime ratio of $\beta$-decay versus neutron capture falls to [$\tau_{\beta}/\tau_{n\gamma}$]{}$\approx1$. It is during this phase that the formation of the rare earth peak proceeds with competition between neutron captures, photo-disintegrations and $\beta$-decays. In the cold $r$-process the first phase [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium is dramatically shorter than the first phase of the classical scenario. Freeze-out is now caused by a rapid drop in temperature rather than the consumption of free neutrons (as in the classical case). The bulk of the cold $r$-process operates in the second phase, under low temperatures ($T_{9}\sim0.5$), where photo-disintegrations have frozen out [@2007ApJ...666L..77W]. Once neutron exhaustion ($R=1$) occurs in the cold $r$-process the free neutrons available to the system must come from the recapture of $\beta$-delayed emitted neutrons. The importance of this effect on the final abundance distribution was noted in [@2010ApJ...712.1359F; @2011PhRvC..83d5809A]. This recapture effect is crucial to peak formation as can be seen from the fact that malformed abundance distributions result if $\beta$-delayed neutron emission is artificially turned off (see [@2011PhRvC..83d5809A]). Our calculations consists of a nuclear reaction network containing $r$-process relevant nuclides as described in [@Surman:1997; @2001PhRvC..64c5801S]. Previous versions of this network code have been used in the studies of Beun [@Beun:2008gn] and Surman [@Surman:2008ef]. The primary reaction channels for nuclides in this section of the reaction network are beta-decay, neutron capture, and photo-dissociation. Our fully implicit $r$-process reaction network handles consistently neutron capture rates at low temperatures and calculations with low abundances of free neutrons, both important for simulations with cold evolutions. For the initial abundances we use self-consistent output from an intermediate reaction network [@Hix:1999yd] with PARDISO solver [@Schenk2004]. Our $r$-process calculations start at $T_{9}=2$ with densities $\rho\approx9\cdot10^{8}$ g/cm$^{3}$ for hot evolutions with $Y_{e}=.30$ and $\rho\approx5\cdot10^{8}$ g/cm$^{3}$ for cold evolutions with $Y_{e}=.40$. At this time the neutron-to-seed ratios are $R\approx45$ and $R\approx35$ respectively. We study the late-time hot and cold $r$-process evolutions in the context of a monotonically decreasing temperature with density parameterized as: $$\label{eqn:Rho-late} \rho(t)\propto t^{-n}$$ where $n$ controls the type of late-time $r$-process evolution (the time when rare earth peak formation occurs). For hot $r$-process evolutions we set $n=2$ and for cold $r$-process evolutions we set $n=6$. A decaying density of $n=2$ is characteristic of wind models at late times while $n=6$ represents a faster decline. We use three different mass models in our nucleosynthesis calculations: Finite Range Droplet Model (FRDM) [@Moller1995], Extended Thomas-Fermi with Strutinsky Integral and Quenching (ETFSI-Q) [@1996PhLB..387..455P] and version 17 of the Hartree Fock Bogoliubov masses (HFB-17) [@2009PhRvL.102o2503G]. The FRDM and ETFSI-Q neutron capture rates are from [@2000ADNDT..75....1R] and were computed with the statistical model code NON-SMOKER [@1998sese.conf..519R]. The HFB-17 neutron capture rates are from the publicly available Brusslib online-database [^3] and were computed with the statistical model code TALYS , which is also publicly available. The HFB mass model is under constant development and is therefore updated with the latest experimental data and theoretical techniques [^4]. The $\beta$-decay rates used in our $r$-process network come from [@2003PhRvC..67e5802M]. Peak Formation in Hot Environments ================================== The mechanism for rare earth peak formation in hot environments was first described in [@Surman:1997]. We review the basic physical arguments in this section. Under hot conditions the $r$-process path (time ordered set of most abundant isotopes) traverses the NZ-plane between the line of stability and the neutron drip line. The path is initially constrained by [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium and is thus found to lie on a line of constant separation energy via the Saha equation. As the free neutrons are consumed, the path moves back toward stability and [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium begins to break down. During this freeze-out from equilibrium, rare earth peak formation can potentially occur. The necessary and sufficient conditions for peak formation are as follows: (1) a deformation maximum or other nuclear structure effect must produce a kink in the lines of constant neutron separation energy around $A\sim160$, and (2) the $r$-process path must traverse this kink region during freeze-out, before $\beta$-decay takes over in the region. The latter allows for the interplay of neutron capture, photo-dissociation and $\beta$-decay as the $r$-process path crosses the region which contains the separation energy kink. During peak formation, the $r$-process path moves toward stability at a rate approximately equal to the average $\beta$-decay rate along the path. The separation energy kink causes a corresponding kink in the $r$-process path as material moves through this important region. This provides a mismatch between the $\beta$-decay rates of material below and above the kink. Due to the kink in the path, nuclei below the peak ($A=150$ to $A=158$) are farther from stability and so $\beta$-decay faster than the average nuclei along the path. Since the nuclei below the peak decay faster than the path moves, these nuclei then proceed to capture neutrons in an attempt to return the $r$-process path back to equilibrium. Conversely, due to the kink in the path, nuclei above the peak ($A=168$ to $A=180$) are closer to stability and so $\beta$-decay slower than average along the path. The path therefore moves before these nuclei have a chance to decay and so they photo-dissociate to shift the $r$-process path back to equilibrium. In the peak region ($A=159$ to $A=167$) some nuclei are still in [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium which limits the amount of material flowing out of the peak region in either direction. The net result causes material to funnel into the peak region, creating the local maximum. The essence of this effect is shown in Figure \[fig:hpf\]. At neutron exhaustion, $R=1$ (left panel), the $r$-process is just beginning to break from [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium. Here the path lies along a line of constant separation energy ($\sim3.0$ MeV) and the abundances show an odd-even effect due to the population of primarily even-N nuclei in equilibrium. No peak exists at this time. Later in the simulation (right panel), peak formation occurs as the path encounters the region with the separation energy kink. The separation energy kink causes the kink in the $r$-process path. Nuclei along the path in the peak region have $\beta$-decay rates which range from 1 $s^{-1}$ (above the kink) to 10 $s^{-1}$ (below the kink). The resultant photo-dissociation above the kink and $\beta$-decay followed by neutron capture below the kink causes material to funnel into the peak region. Peak Formation in Cold Environments {#repf:cold} =================================== In the previous section we analyzed rare earth peak formation in hot evolutions and found that photo-dissociation was crucial in peak formation. However, we also find well formed solar-like rare earth peaks in simulations of cold environments where photo-dissociation plays no role in the dynamics after $R=1$. After $R=1$, the cold $r$-process path is controlled on average by the competition between neutron captures and $\beta$-decays [$\tau_{\beta}/\tau_{n\gamma}$]{}$\approx1$. Locally, over the rare earth region, the exact position of the path is more complicated due to the variation among individual rates. As the material decays back to stability peak formation will ensue if the path encounters a peak region where neutron capture rates are slow relative to the above and below regions. The essence of the effect is that slow neutron capture rates in the peak region cause a bow (inwards towards stability) in the lines of constant neutron capture rates relative to the lines of constant $\beta$-decay rates thus causing material to become trapped in the peak region. The cold formation mechanism is shown in Figure \[fig:cpf\]. The left panel shows a snapshot of the abundance pattern and rates at neutron exhaustion, $R=1$. At this point in time the $r$-process path is still influenced by residual photo-dissociation flows. This is reflected in an odd-even effect in the abundances and flat $r$-process path (similar to hot evolutions). However, the photo-dissociation rates are decreasing so rapidly they play no further role in the dynamics after this point. Shortly, the neutron capture rates will become comparable to the $\beta$-decay rates and large odd-even behavior of the abundances will be washed-out [@1999ApJ...516..381F]. In fact, this has already begun to happen as can be seen with the slight bowing of the neutron capture rate lines in the peak region ($A=159$ to $A=167$). At a slightly later time in the simulation (right panel of Figure \[fig:cpf\]) the system has moved closer to stability and the $r$-process path now encounters the slower capture rates in the peak region. Below the peak ($A=150$ to $A=158$) neutron captures occur much faster than $\beta$-decay rates along the $r$-process path, so the net result is material shifting towards the peak region. In the peak region the path encounters the slow capture rates (note the bowing of the neutron capture rate lines) so that any material being shifted into the peak region becomes hung up. Above the peak ($A=168$ to $A=180$) the flow of material is again dominated by the relatively faster neutron capture rates. The net result is trapping of material into the peak region. Another interesting feature in the right panel of Figure \[fig:cpf\] is the trough to the left of the peak. A trough can occur if a gap in the $r$-process path proceeds for long periods of time as matter decays back to stability. Along a gap in the $r$-process path the neutron capture rates are relatively fast resulting in movement of material to more neutron-rich isotopes and a depletion of material in the gap region. In our figures, the lines of constant neutron capture rates have been averaged over even-N nuclei. Even-N neutron capture rates are more important to rare earth peak formation because at a given temperature, odd-N nuclei have faster neutron capture rates which causes material to pass through the odd-N nuclei quickly. Thus material builds up (or stays) in even-N nuclei which sets the $r$-process path. The importance of individual neutron capture rates in the rare earth peak was highlighted in [@Mumpower2010ncr]. Influence of Mass Model on Rare Earth Peak Formation ==================================================== From the previous two sections it is clear that the details of the late-time thermodynamic evolution are critical in setting the relevant nuclear physics and thus determine the mechanism for peak formation. Despite the differences in peak formation mechanisms, we find that the final abundances among simulations with the same mass model yet differing late-time thermodynamic behavior can be remarkably similar. This is in contrast to the differences found in the final abundance pattern when comparing between mass models with similar thermodynamic conditions. In this section we focus on the influence of different separation energies and neutron capture rates on rare earth peak formation. A successful peak formation is imprinted on the final abundances in a cold evolution when the $r$-process path encounters structure in the neutron capture rates and this structure lasts until the point at which $\beta$-decays take over neutron captures in the region ([$\tau^{REP}_{n\gamma}\approx\text{ a few }\tau^{REP}_{\beta}$]{}). A successful peak formation occurs in a hot evolution when the $r$-process path encounters a local deformation maximum leading to a well-defined kink structure in the separation energies in the rare earth region. For a given mass model, the structure of neutron capture rates and the structure of the separation energies may not align in the NZ-plane. This in turn can affect the timing and location of peak formation and hence the nuclei which are relevant. Odd-even effects in the abundances can accumulate or persist through the decay back to stability resulting in visible features in the final abundances. Smoothing of the abundances typically occurs in between neutron capture freeze-out ([$\tau^{REP}_{n\gamma}=\tau^{REP}_{\beta}$]{}) and the time in which $\beta$-decays fully take over neutron captures in the region ([$\tau^{REP}_{n\gamma}\approx\text{ a few }\tau^{REP}_{\beta}$]{}). We now discuss three different mass models in this context. Since separation energies vary among mass models we instead (for consistency) use $\langle\delta N\rangle$, the abundance weighted average neutrons from stability, to measure the $r$-process path’s progression. Compared to the other mass models studied here, we find that simulations which use the FRDM mass model best match the solar data in the rare earth peak region in both hot and cold evolutions. In fact we find (in agreement with previous studies [@Surman:1997; @2011PhRvC..83d5809A]) that the FRDM mass model is the only model to show a well-formed rare earth peak consistently in the final abundance pattern. Simulations with the FRDM mass model do not consistently form rare earth peaks far from stability ($\langle\delta N\rangle>20$). Instead, peak formation ensues when the path is much closer; on average in between 15 and 20 neutrons away from stability. We can see the evolution of the peak region for a cold FRDM evolution in Figure \[fig:form-cold-frdm\]. At $\langle\delta N\rangle\sim20$ (top panel) the structure in the capture rates has yet to manifest itself resulting in relatively flat abundances. As the path moves back to stability, $\langle\delta N\rangle\sim15$ (middle panel), it encounters nuclei in the peak region with relatively slower neutron capture rates than the surrounding regions (note the bending in the red lines). These conditions persist all the way back to stability resulting in a well-formed rare earth peak. A similar scenario occurs in hot evolutions; see Figure \[fig:form-hot-frdm\]. FRDM shows a slight overlap between neutron capture structure and separation energy structure. The structure in the separation energies occurring between $\langle\delta N\rangle\sim12\text{ to }20$ and the structure in the capture rates occurring between $\langle\delta N\rangle\sim10\text{ to }15$. This delays peak formation in cold scenarios until around 15 neutrons from stability, while hot evolutions typically begin peak formation approximately 20 neutrons from stability. Simulations with the ETFSI mass model consistently form a solar-like rare earth peak far from the stable nuclei ($\langle\delta N\rangle\gtrsim20$). This is most apparent in colder simulations (see right panel of Figure \[fig:cpf\]). However, this is not the end of the story as the material must decay back to stability. Figure \[fig:form-cold-etfsi\] highlights this transition at an abundance weighted average of $\langle\delta N\rangle\sim20$ (top panel), 15 (middle panel) and 10 (bottom panel) neutrons from stability. As the decay back to stability proceeds the $r$-process path encounters nuclei whose neutron capture rates become homogeneous around the peak region. This slowly dissolves the structure, flattening the lines of constant neutron capture rates (compare top and middle panels). By the time the path is on average 15 neutrons away from stability (middle panel) the cold trapping mechanism can not continue because neutron capture rates in the peak region are no longer slower than the surrounding regions. These conditions persist back to stability resulting in a final abundance pattern with a more modest rare earth peak. Note that a small odd-even effect reappears since in this model $\beta$-delayed neutron emission is still relevant as neutron capture freezes out in the rare earth region ([$\tau^{REP}_{n\gamma}=\tau^{REP}_{\beta}$]{}). Solar-like rare earth peaks form far from the stable nuclei in ETFSI models under hot evolutions as well. Far from stability, the structure (kink) in the separation energies results in the hot peak formation mechanism. Like the FRDM case, the separation energy kink in ETFSI disappears as one moves closer to stability. However, the kink disappears while neutron captures are still dominant ([$\tau^{REP}_{n\gamma}\lesssim\tau^{REP}_{\beta}$]{}) far from stability ($\langle\delta N\rangle\geqslant20$) resulting in a flattened final abundance distribution; see Figure \[fig:form-hot-etfsi\]. In this mass model the structure in the separation energies occurs farther from stability ($\langle\delta N\rangle\geqslant20$) than the structure seen in the neutron capture rates ($\langle\delta N\rangle\sim20$) influencing peak formation in a similar fashion to the FRDM case. The gross separation energy structure occurs very early on “before” the top panel of Figure \[fig:form-hot-etfsi\] and has already dissolved by $\langle\delta N\rangle\sim20$. Version 17 of the HFB mass model is optimized to over 2000 measured masses from [@2003NuPhA.729....3A] corresponding to a root mean square error of $\lesssim0.6$MeV. This data set features detailed structure in the separation energies but little overall structure in the neutron capture rates for the nuclei relevant to rare earth peak formation. These features are reflected in our $r$-process abundances. Figure \[fig:form-cold-hfb17\] shows the decay back to stability of a cold $r$-process using HFB-17. At every snapshot, highlighting the $r$-process path’s decay back to stability, we do not find the structure in the neutron capture rates as is found in the other two mass models. It is this relative homogeneity in the neutron capture rates throughout the rare earth region which prevents the trapping mechanism from occurring in cold evolutions. In hot $r$-process evolutions the situation is more intricate than for the corresponding cases of the other two mass models. The detailed structure in the separation energies results in a complex separation energy kink structure in the rare earth region. However, due to the lack of gross structure as the separation energy increases (i.e. during the decay back to stability) the funneling mechanism cannot operate. This can be seen in Figure \[fig:form-hot-hfb17\] and illustrates the subtleties involved in forming the rare earth peak. The discussion in this section showcases the need for nuclear structure measurements far from stability. As we have seen, the nuclei that are important for rare earth peak formation lie in between 10 and 20 neutrons away from stability. Furthermore, it is the nuclei which are the closest to stability, those in between 10 and 15 neutrons from stability, which are most influential to peak formation as they set or potentially dissolve the peak structure all together. In Figure \[fig:rep-in\] we highlight these influential nuclei together with recent experimental mass measurements ([^5] [^6] green and [@2003NuPhA.729....3A] gray) and known neutron capture rates ([^7] red). Summary and Conclusions ======================= We have studied the formation and evolution of the rare earth peak at late-times during the $r$-process. To take into account uncertainties with nuclear physics in the region our calculations employed three mass models (FRDM, ETFSI, and HFB-17). Two late-time evolutions were considered: A hot $r$-process with temperatures high enough to support [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} equilibrium and a cold $r$-process with lower temperatures where there are no photo-dissociation flows, only competition between neutron captures and $\beta$-decays after $R=1$. Both of these evolutions are similar at early times so that the changes in abundances are not due to to the physics that sets the neutron-to-seed ratio, but instead due to the changes in the nuclear physics input or changes in the late-time behavior of the evolution. The differences in late-time evolution (hot vs cold) determine which nuclear physics input is important (separation energies vs neutron capture rates respectively) during the final stages of the $r$-process. In hot evolutions the combination of photo-dissociation, neutron capture and beta-decay results in a mechanism which funnels material into the peak region. A successful peak formation in hot evolutions is imprinted on the abundance pattern when the structure in the separation energies, the kink, is well defined *and* the $r$-process path crosses the kink region during the [$(n,\gamma)\rightleftarrows(\gamma,n)$]{} freeze-out. We contrast this with the peak formation mechanism which occurs in cold $r$-process environments. Here the important nuclear physics for peak formation lies in the local structure of the neutron capture rates. When the neutron capture rates are slow in the peak region relative to the surrounding regions (creating the characteristic bow in the lines of constant neutron capture rates) material can become trapped in the peak region, thus forming the peak. A successful peak formation in cold evolutions is imprinted on the abundance pattern when the structure in the neutron capture rates lasts until the point at which $\beta$-decays take over neutron captures in the region ([$\tau^{REP}_{n\gamma}\approx\text{ a few }\tau^{REP}_{\beta}$]{}). The rare earth peak is extremely sensitive to the subtleties of nuclear physics input. Neutron capture is particularly important in both hot and cold evolutions. For instance, we find that neutron capture can play two competing roles in peak formation: it can be responsible for creating the peak, but also for potentially dissolving the peak (wash-out). Neutron capture rate structure and separation energy structure in the same mass model may not overlap in the NZ-plane. This in turn can affect the timing and location of peak formation in different thermodynamic conditions. We have shown that the rare earth peak in principle offers unique insight into the late-time behavior of the $r$-process because it forms away from the closed shells during freeze-out while material decays back to stability. Rare earth peak formation is sensitive to the structure of separation energies and / or neutron capture rates about 10 to 15 neutrons away from the stable rare earth peak. Future measurements at radioactive ion beam facilities should reach this important region and will be critical in placing constraints on nuclear models. This in turn will lead to improved $r$-process predictions; allowing the rare earth peak to evolve into a powerful tool for understanding the $r$-process. Acknowledgements ================ We thank A. Arcones and T. Rauscher for valuable discussions. We thank North Carolina State University for providing the high performance computational resources necessary for this project. This work was supported in part by U.S. DOE Grant No. DE-FG02-02ER41216, DE-SC0004786, and DE-FG02-05ER41398. ![image](rep.pdf){width="85mm" height="61.8mm"} ![image](hpf.pdf){width="180mm" height="130mm"} ![image](cpf.pdf){width="180mm" height="130mm"} ![image](form-cold-frdm.pdf){width="180mm" height="140mm"} ![image](form-hot-frdm.pdf){width="180mm" height="140mm"} ![image](form-cold-etfsi.pdf){width="180mm" height="140mm"} ![image](form-hot-etfsi.pdf){width="180mm" height="140mm"} ![image](form-cold-hfb17.pdf){width="180mm" height="140mm"} ![image](form-hot-hfb17.pdf){width="180mm" height="140mm"} ![image](rep-in.pdf){width="100mm" height="90mm"} [^1]: http://www.frib.msu.edu [^2]: http://www.gsi.de/portrait/fair.html [^3]: http://www.astro.ulb.ac.be/ [^4]: http://www.astro.ulb.ac.be/ [^5]: http://isoltrap.web.cern.ch [^6]: https://www.jyu.fi/fysiikka/en/research/accelerator/igisol/trap [^7]: http://www.nndc.bnl.gov/exfor
{ "pile_set_name": "ArXiv" }
--- abstract: 'Bipartite entanglement entropies, calculated from the reduced density matrix of a subsystem, provide a description of the resources available within a system for performing quantum information processing. However, these quantities are not uniquely defined on a system of non-Abelian anyons. This paper describes how reduced density matrices and bipartite entanglement entropies (such as the von Neumann and Renyi entropies) may be constructed for non-Abelian anyonic systems, in ways which reduce to the conventional definitions for systems with only local degrees of freedom.' author: - 'Robert N. C. Pfeifer' bibliography: - 'EntEnt.bib' date: 'February 5, 2014' title: 'Measures of entanglement in non-Abelian anyonic systems' --- Introduction ============ Entanglement represents a fundamental resource for the performance of quantum information processing protocols, corresponding to the existence of non-classical correlations between subsystems. Demonstrations of these non-classical correlations have been performed as long ago as [-@bell1964],[@bell1964] and with modern experimental techniques it is becoming increasingly possible to systematically exploit this behaviour.[@steane1998; @bennett2000; @ladd2010] One area of particular interest is the generation of entanglement in non-Abelian anyonic systems. As well as representing an intriguing and novel phase of matter, these systems are capable of implementing quantum computing protocols with extremely low rates of decoherence, and acting as robust quantum memories.[@kitaev2003; @preskill2004; @nayak2008] A number of candidates have been proposed for the experimental realisation of such systems.[@read1999; @xia2004; @pan2008; @kumar2010; @stern2010; @sanghun-an2011] An *entanglement monotone* is a mapping ${\mathcal{M}}$ between the reduced density matrix description of a physical (sub)system and the real numbers, such that ${\mathcal{M}}[L(\hat{\rho})]\leq {\mathcal{M}}(\hat\rho)$ for all $\hat\rho$, where $L$ is an arbitrary LOCC (Local Operations and Classical Communication) operation over $\hat\rho$.[@bennett1996; @vidal2001; @vedral1997] As entanglement cannot be created by LOCC operations but can be eliminated (by collapse of the wavefunction), entanglement monotones provide partial orderings over the space of states, and are sometimes loosely referred to as a *measure of the entanglement* between the system described by $\hat\rho$ and its environment. Multiple such monotones exist.[vedral1997,vedral1997a,hamma2005,vidal2000,plenio2007,flammia2009]{} Of particular interest in this paper will be the von Neumann and Renyi *bipartite entanglement entropies*. On dividing an isolated physical system into two regions A and B, a bipartite entanglement entropy is an entanglement monotone computed from the reduced density matrix of either subsystem A or B (denoted ${\hat\rho_{\mathrm{A}}}$ and ${\hat\rho_{\mathrm{B}}}$ respectively) which describes the entanglement between regions A and B. Specifically, the Renyi entropies comprise a family of entanglement monotones conventionally defined as $$S_n(\hat\rho_{\mathrm{A}})=\frac{1}{1-n}\log_2{\left[{\mathrm{Tr}}{\left(\hat\rho_{\mathrm{A}}^{{\phantom{_{\mathrm{A}}}}n}\right)}\right]}, \quad n\not=1,\label{eq:Renyi}$$ and the von Neumann entropy $$S_1({\hat\rho_{\mathrm{A}}})={\mathrm{Tr}}\left(\hat\rho_{\mathrm{A}}\log_2{\hat\rho_{\mathrm{A}}}\right)\label{eq:vonNeumann}$$ corresponds to the limit $n\rightarrow 1$. A previous attempt has been made to define entanglement monotones for systems of anyons from an information theoretic viewpoint,[@hikami2008] with attention primarily to systems of anyons on the disc, and is adequate for Abelian anyonic systems and for appropriately ordered abstract systems of non-Abelian anyons. However, for non-Abelian anyons the formulation becomes ambiguous when applied to real physical systems. The present paper builds on this previous work to describe the full family of von Neumann and Renyi bipartite entanglement entropies for non-Abelian anyonic systems on surfaces of arbitrary genus. The reduced density matrix\[sec:rdm\] ===================================== In order to compute bipartite entanglement entropies between two regions A and B, it is first necessary to construct a reduced density matrix on either region A or region B. Although construction of the density matrix for an anyonic system is relatively straightforward, requiring only knowledge of the Hilbert space of the system, the construction of a reduced density matrix for an arbitrary subsystem involves a surprising number of subtleties which impact the subsequent calculation of entanglement entropy. Consequently, it is first necessary to discuss in some detail the construction of a reduced density matrix for a subsystem admitting non-Abelian excitations. General properties of the density matrix\[sec:general\] ------------------------------------------------------- Consider first the familiar case of a system of spins $i_1,\ldots,i_n$ on a lattice ${\mathcal{L}}$, where the total Hilbert space ${\mathcal{H}}$ admits a tensor product structure $${\mathcal{H}}=\bigotimes_{i=1}^n {\mathcal{H}}_i.\label{eq:tproddec}$$ The space of operators on the lattice ${\mathcal{L}}$ is then given by ${\mathcal{O}}={\mathcal{H}}\otimes\overline{{\mathcal{H}}}$. The density matrix $\hat\rho$ may be used to compute the expectation value of any operator $\hat{O}\in{\mathcal{O}}$ and thus occupies the space dual to ${\mathcal{O}}$ under the action of trace, such that $$\begin{split} F_{\hat\rho}&:{\mathcal{O}}\longrightarrow{\mathbb{C}}\\ F_{\hat\rho}&(\bullet) = {\mathrm{Tr}}(\hat\rho~\bullet).\label{eq:dmtrace} \end{split}$$ As ${\mathcal{O}}$ is self-dual under the action of trace, it follows that $\hat\rho\in{\mathcal{O}}$. For such a system, a subsystem ${\mathcal{A}}$ may be defined as a subset of the spins making up ${\mathcal{L}}$. If lattice ${\mathcal{L}}$ is then placed on a manifold $M$ then there exists a submanifold $A\subseteq M$, possibly disjoint, such that $A$ contains only the spins in ${\mathcal{A}}$. The Hilbert space of the portion of the spin system local to submanifold $A$ is then precisely the Hilbert space of the subset of spins ${\mathcal{A}}$, $${\mathcal{H}}_A = {\mathcal{H}}_{\mathcal{A}} = \bigotimes_{i\in{\mathcal{A}}} {\mathcal{H}}_i,$$ and it is unnecessary to distinguish between the subset of spins ${\mathcal{A}}$ and the enclosing submanifold $A$, both of which may consequently be referred to as “region A”. The space of operators on region A is then ${\mathcal{O}}_{\mathrm{A}}={\mathcal{H}}_A\otimes\overline{{\mathcal{H}}_A}$. The reduced density matrix ${\hat\rho_{\mathrm{A}}}$ occupies the space dual to ${\mathcal{O}}_{\mathrm{A}}$ under the action of the trace operation, $$\begin{split} F_{{\hat\rho_{\mathrm{A}}}}&:{\mathcal{O}}_{\mathrm{A}}\longrightarrow{\mathbb{C}}\\ F_{{\hat\rho_{\mathrm{A}}}}&(\bullet) = {\mathrm{Tr}}({\hat\rho_{\mathrm{A}}}~\bullet) \end{split}\label{eq:rhoinO}$$ and thus also ${\hat\rho_{\mathrm{A}}}\in{\mathcal{O}}_{\mathrm{A}}$. For a system of non-Abelian anyons, ${\mathcal{O}}_{\mathrm{A}}$ will once again be defined as the space of operators local to region A. Knowledge of the reduced density matrix ${\hat\rho_{\mathrm{A}}}$ on region A must permit evaluation of the expectation value of any operator in ${\mathcal{O}}_{\mathrm{A}}$, so once again ${\hat\rho_{\mathrm{A}}}$ must inhabit the space dual to ${\mathcal{O}}_{\mathrm{A}}$. However, the Hilbert space of a system of non-Abelian anyons does not in general admit the tensor product structure of [Eq. (\[eq:tproddec\])]{} and thus the definition of ${\mathcal{O}}_{\mathrm{A}}$ must proceed with greater care. The anyonic density matrix\[sec:adm\] ------------------------------------- In preparation for the construction of a reduced density matrix for a subsystem containing non-Abelian anyons, consider first the definition of the density matrix on the entirety of the system. Let a system of anyons inhabit a two-dimensional manifold $M$. Following the prescription given in [Ref. ]{}, a fusion tree diagram may be constructed by treating the anyons as punctures and performing a pairs-of-pants decomposition of $M$, then constructing the tree to fit inside the pants. It is well-known that valid labellings of this fusion tree then represent the basis vectors of a Hilbert space describing a system of anyons on $M$ (see e.g. Refs. ); it is less widely appreciated that different pairs-of-pants decompositions may yield fusion trees which appear equivalent but which differ by braiding operations, as illustrated (for example) in [Fig. \[fig:projection\]]{}. ![When constructing a fusion tree basis on a system of non-Abelian anyons, in order to unambiguously describe the state of a physical system it is necessary to specify the relevant pairs-of-pants decomposition. For example, when the fusion tree in diagram (i) is taken in conjunction with the pairs-of-pants decomposition indicated by shading in diagram (ii), this suffices to specify a state $|\psi_1{\rangle}$ on manifold $M_1$ (which is a disc with trivial total charge). If the same sum over weighted fusion trees (i) is instead associated with the decomposition of $M_1$ given in diagram (iii), then this specifies a different a different state $|\psi_2{\rangle}$, where $|\psi_2{\rangle}=\hat B_{234}|\psi_1{\rangle}$ for $\hat B_{234}$ given in diagram (iv). In diagrams (ii) and (iii), the dashed line represents the projection of the fusion tree onto the manifold when viewed from above, following its initial construction from the pairs-of-pants decomposition but prior to its flattening onto the page in the form of diagram (i). Vertex indices have been suppressed throughout this paper for simplicity, but may easily be reintroduced if required. \[fig:projection\]](projection){width="246.0pt"} To uniquely specify the state of a physical system it is therefore necessary not only to give a weighted sum over labellings of the fusion tree, but also the relevant pairs-of-pants decomposition. For a surface of genus 0 this may be achieved by specifying the projection of the fusion tree onto the manifold as in [Fig. \[fig:projection\]]{}(ii)-(iii), as the pairs-of-pants decomposition may then be reconstructed from the combination of this projection and the fusion tree. Having obtained a basis of states $\{|\psi^i_M{\rangle}\}$ which spans the Hilbert space ${\mathcal{H}}$ for a system of anyons on an arbitrary manifold $M$, operators acting on $M$ once again inhabit the space ${\mathcal{O}}={\mathcal{H}}\otimes\overline{{\mathcal{H}}}$, which for anyonic systems will also be denoted ${\mathcal{O}}_M$. For the example system of [Fig. \[fig:projection\]]{}(i)-(ii) this construction is illustrated in [Fig. \[fig:operatorM\]]{}. ![Graphical representation of an arbitrary operator $\hat{O}\in{\mathcal{O}}_M$ for the system of anyons described in (i)-(ii). Projection onto $M_1$ is taken to be as per (ii). \[fig:operatorM\]](operatorM){width="246.0pt"} The matrix trace of [Eq. (\[eq:dmtrace\])]{} is replaced by the (graphical) quantum trace $$\raisebox{-26.5pt}{\includegraphics[width=27.33pt]{qtrace1}}\textrm{~~~~~~or~~~~~~}\raisebox{-26.5pt}{\includegraphics[width=27.33pt]{qtrace2}}$$ which is equivalent to the matrix trace on Abelian systems, and the density matrix $\hat\rho$ is once again seen to inhabit ${\mathcal{O}}_M$. Non-trivial boundary charges ---------------------------- The method of constructing a basis for an anyonic system described in [Sec. \[sec:adm\]]{} makes use of a powerful duality between Unitary Braided Tensor Categories (UBTCs) and Schwarz-type Topological Quantum Field Theories (TQFTs). If an anyon model is considered as a collection of quasiparticle excitations in a quantum spin liquid, with the nature of the liquid selecting the UBTC which describes the anyon model and the behaviours of these excitations being described by the UBTC using the usual diagrammatic formalism, then the corresponding TQFT description is obtained by identifying the quantum spin liquid with the manifold of the TQFT and replacing the quasiparticle excitations with punctures. Consequently, whereas the UBTC picture incorporates both anyonic excitations and possible boundary charges of the quantum spin liquid, in the TQFT picture all anyonic charges are associated with a boundary of the manifold. Construction of a fusion tree basis treats all boundaries equivalently, and thus the leaves of the tree may be associated not only with anyons but also with the boundaries of the quantum spin liquid. For example, if the disc of [Fig. \[fig:projection\]]{}(ii) were permitted to carry a boundary charge then an appropriate basis would be given by the fusion tree in [Fig. \[fig:totalcharge\]]{}(i), as opposed to the one employed in [Fig. \[fig:projection\]]{}(i) where the boundary charge was fixed to be trivial. ![(i) Fusion tree basis for states of four anyons on manifold $M_1$ with a possible boundary charge $a_{\mathrm{b}}$. If $a_{\mathrm{b}}$ is trivial, the corresponding leaf may be deleted from the fusion tree to recover the basis of [Fig. \[fig:projection\]]{}(i). Note that the projection of the fusion tree now extends to the boundary of $M_1$. (ii) General form of an operator acting on a 3-anyon disc $D_1$ with boundary charge. Label $a_{D_1}$ represents the charge on the boundary of $D_1$. (iii) General form of an operator on $D_1$ which leaves the boundary charge invariant. The total boundary charge is denoted $a_{\mathrm{t}}$, and in this example corresponds to $\overline{a_{D_1}}$ for the value of $a_{D_1}$ given in diagram (ii), because the boundary is non-disjoint. Note that the projection of this fusion tree no longer extends to the boundary as the total boundary charge is not associated with a leaf of the fusion tree. \[fig:totalcharge\]](totalcharge){width="246.0pt"} ![image](ODinO){width="492.0pt"} As a consequence of this identification between anyons and punctures, it is always possible to identify an anyon model on an $n$-punctured manifold having $m$ disjoint sections of boundary with an anyon model on an equivalent $n+m$-punctured manifold which does not have a boundary. A simple example is the mapping between the $n$-punctured disc and the $n+1$-punctured sphere described in [Ref. ]{}. For any manifold (or, indeed, submanifold) with boundary charges, the Hilbert space of states on that manifold is spanned by the valid labellings of a fusion tree which includes the boundary charges. For example, if disc $D_1$ is a manifold containing three anyons $a_1$, $a_2$, and $a_3$, then the space of all possible operators acting on $D_1$ (including those which may change the boundary charge), denoted ${\mathcal{O}}_{D_1}$, may be represented by all diagrams of the form of [Fig. \[fig:totalcharge\]]{}(ii). The space of all operators acting on a general discoid manifold $D$ will be denoted ${\mathcal{O}}_D$. It is also useful to describe the space of all operators which do not modify the boundary charge, which will be denoted ${\mathcal{O}}_{D}^{\prime}$. A diagrammatic basis for ${\mathcal{O}}^{\prime}_D$ may be obtained by starting with a diagrammatic basis for ${\mathcal{O}}_D$ and performing the quantum trace over all charges associated with boundaries of the quantum spin liquid. Regardless of the number of disjoint sections making up the boundary, the result following simplification is a diagram such as that given in [Fig. \[fig:totalcharge\]]{}(iii) where a single charge label $a_{\mathrm{t}}$ connects the upper and lower halves of the diagram and represents the total charge on the boundary. The anyonic reduced density matrix\[sec:ardm\] ---------------------------------------------- ### Genus 0\[sec:genus0\] Consider next the definition of an operator on a disc $D$ explicitly satisfying $D\subset M$, where $M$ (ignoring any anyon-associated punctures) is a connected manifold of genus 0. The space of operators on $D$ is ${\mathcal{O}}_D$, and includes operators which modify the boundary charge of $D$. However, on $M$ this boundary is not physical. Consequently, any operator modifying the boundary charge of $D$ must in fact be a truncated description of an operator in ${\mathcal{O}}_M$ which transfers charge between $D$ and $M-D$. The action of such an operator on $M-D$ is therefore non-trivial. Referring back to [Sec. \[sec:general\]]{}, it is desireable now to define “operators local to $D$” for a system of non-Abelian anyons. Adopting a physical motivation for this definition, an operator is defined as “local to $D$” if that operator is trivial everywhere except on $D$. Under this definition, the space of operators local to $D$ corresponds to ${\mathcal{O}}_D^{\prime}$. Given specific diagrammatic bases for ${\mathcal{O}}_M$ and ${\mathcal{O}}^{\prime}_D$, comparison of the projection of an operator $\hat{O}_D\in{\mathcal{O}}^{\prime}_D$ on $D\subset M$ with the projection of an operator $\hat{O}_M\in{\mathcal{O}}_M$ on $M$ yields an explicit embedding of ${\mathcal{O}}^{\prime}_D$ into ${\mathcal{O}}_M$. This is illustrated for the example of $D_1\subset M_1$ in [Fig. \[fig:ODinO\]]{}(i)-(iii). The embedding of ${\mathcal{O}}^{\prime}_D$ in ${\mathcal{O}}_M$ is defined in terms of a braiding operator $\hat B\in{\mathcal{O}}_M$, and knowledge of this operator permits the calculation of expectation values for operators $\hat{O}_D\in{\mathcal{O}}^{\prime}_D$ using $\hat\rho$ \[[Fig. \[fig:ODinO\]]{}(iv)\]. Applying diagrammatic isotopy to this calculation, it is seen that applying the adjoint action $\hat B^\dagger(\bullet)\hat B$ to $\hat\rho$ then tracing out 1. all anyons not in $D$, and 2. the boundary charges of $D$ yields a reduced density matrix $\hat\rho_D\in{\mathcal{O}}^{\prime}_D$. A basis for ${\mathcal{O}}_M$ will be termed *compatible* with a given basis for ${\mathcal{O}}^{\prime}_D$ if $\hat B$ is trivial. The construction of $\hat\rho_{D_1}$ for $D_1\subset M_1$ is shown in [Fig. \[fig:ODinO\]]{}(v). ![image](threehandles){width="492.0pt"} ### Surfaces of higher genus\[sec:highergenus\] Generalisation of this approach to manifolds $M$ and submanifolds $A$ of arbitrary genus permits the construction of a reduced density matrix $\hat\rho_A$ for any manifold $M$ and submanifold $A\subset M$. This construction may be achieved as follows: 1. Introduce a fusion tree basis $F_A$ for operators on $A$ which has leaves for all anyons and all sections of boundary. Let ${\mathcal{C}}$ be the set of charges associated with boundaries on $A$ which are not also boundaries on $M$ (i.e. boundaries which would be created on cutting around $A$). Basis $F_A$ must be chosen so that tracing over all charges in ${\mathcal{C}}$ may be performed without braiding. The space of operators expressed in basis $F_A$ will be termed ${\mathcal{O}}_A$ and is analogous to ${\mathcal{O}}_D$ in [Sec. \[sec:genus0\]]{}. 2. In basis $F_A$, trace out all boundary charges which are not also associated with a boundary on $M$ (i.e. all charges in set ${\mathcal{C}}$). The resulting space of operators will be termed ${\mathcal{O}}_A^{\prime}$ and is analogous to ${\mathcal{O}}_D^{\prime}$ in [Sec. \[sec:genus0\]]{}. Let the fusion tree basis of ${\mathcal{O}}_A^{\prime}$ be denoted $F_A^{\prime}$. 3. Let $M^{\prime}$ be the disjoint manifold obtained on cutting $M$ around the boundary of $A$. Construct a fusion tree basis $F_{M^{\prime}}$ for operators on $M^{\prime}$ which is compatible with basis $F_A$. 4. Construct a fusion tree basis $F_M$ for operators on $M$ which yields basis $F_{M^{\prime}}$ on $M^{\prime}$ when cut around the boundary of $A$. 5. Specify $\hat\rho$ in basis $F_M$. 6. Cut around the boundary of $A$ to compute $\hat\rho$ on $M^{\prime}$ in basis $F_{M^{\prime}}$. 7. Trace out all anyons not in $A$, all punctures or boundaries created by cutting around the boundary of $A$,[^1] and all degrees of freedom associated with handles in $M$ but not in $A$ (and thus with non-trivial loops in $F_M$ but not in $F_A^{\prime}$), to yield ${\hat\rho_A}$ in basis $F_A^{\prime}$. An example of constructing a reduced density matrix for a more complicated manifold using this procedure is illustrated in [Fig. \[fig:highergenus\]]{}. Note 1: When the genus of submanifold $A$ is greater than zero, the embedding of ${\mathcal{O}}_A^\prime$ into ${\mathcal{O}}_{M'}$ in arbitrarily-selected bases $F'_A$ and $F_{M'}$ may involve not only braiding but also transformations acting on topological degrees of freedom. Including these transformations in $\hat{B}$, the definition of compatible bases in Step 3 remains unchanged as $\hat{B}={\mathbb{I}}$. Note 2: In the construction given above, the quantum trace is always performed without braiding and without requiring any transformations on the topological degrees of freedom. For this to be possible, the density matrix is first written in a basis $F_M$ which, when cut to yield $\hat\rho$ in basis $F'_M$, is compatible by construction with a basis on . More generally one may construct a reduced density matrix ${\hat\rho_A}$ from a density matrix $\hat\rho$ in *any* basis, including ones not compatible with $F^{\prime}_A$, provided the procedure of taking a quantum trace is augmented by a description $\hat{B}$ of the transformations corresponding to the embedding of ${\mathcal{O}}'_A$ into ${\mathcal{O}}_{M'}$. However, such transformations may always be interpreted as first performing a change of basis on $\hat\rho$ into one which *is* compatible with $F^{\prime}_A$, followed by performing the quantum trace *without* transforming or braiding. Consequently there is no disadvantage to requiring that the quantum trace always be performed without braiding or transforming the topological degrees of freedom, having first expressed the density matrix $\hat\rho$ in a basis which, after cutting, will be compatible with $F'_A$. Notes on disjoint manifolds --------------------------- It is noted that the computation of ${\hat\rho_A}$ presented in [Sec. \[sec:ardm\]]{} remains valid even when one or both out of manifold $M$ and submanifold $A$ are disjoint. Like the manifold itself, the fusion tree basis for a disjoint manifold is made up of two or more separate pieces, with one piece corresponding to each disjoint section. If a submanifold $A$ is disjoint, then a fusion tree basis on $M$ is said to be compatible with a basis on $A$ if that basis on $M$ is compatible with the basis for each section of $A$ taken independently. Entanglement entropies\[sec:entent\] ==================================== Given the fusion tree representation of a reduced density matrix, the corresponding matrix representation may be constructed by writing the coefficients in matrix form and absorbing the normalisation factors associated with the fusion tree.[@bonderson2007; @bonderson2008; @pfeifer2010] Once this is done, it is possible to calculate entanglement entropies from the matrix representation of ${\hat\rho_A}$ in the usual manner [(\[eq:Renyi\]–\[eq:vonNeumann\])]{}. However, when working with non-Abelian anyons, once again there are additional considerations which must be taken into account. Pairs-of-pants decomposition\[sec:ententpop\] --------------------------------------------- It is well-known that braiding operations are capable of generating entanglement between the non-local degrees of freedom of a system of non-Abelian anyons; indeed, this property is exploited when using anyons to implement topological quantum computation.[@nayak2008] Furthermore, bases having different projections onto the manifold of the anyonic system, and which therefore linearise the anyons according to different trajectories, may be related to one another by braiding (see e.g. Figs. \[fig:projection\] and \[fig:ODinO\]). Given a bipartition which divides a manifold $M$ into submanifolds $A$ and $B$ and a reduced density matrix on one of these submanifolds, $\hat\rho_X|_{X\in\{A,B\}}$, any unitary operation $\hat{U}_X$ satisfying $$\hat{U}_X\in{\mathcal{O}}'_X$$ will leave the eigenvalues of the reduced density operator $\hat{\rho}_X$ unchanged, as will any operator local to the other submanifold $$\hat{U}_Y\in{\mathcal{O}}'_Y\qquad Y\in\{A,B\},~Y\not=X,$$ or any operator which may be reduced to the sequential application of operators of these two types. Such operators will be denoted as belonging to ${\mathcal{O}}'_A\times{\mathcal{O}}'_B$. In contrast, operators which do not satisfy these criteria may potentially entangle the degrees of freedom associated with regions $A$ and $B$ and thus change the eigenvalues of $\hat\rho_X$. A simple example of such an operation is given in [Fig. \[fig:sigmabraid\]]{}. ![(i) Fusion tree basis for a chain of six Ising anyons on the disc with trivial boundary charge. Note the ordering of anyons on the fusion tree. Fusion products $a_1$ and $a_2$ may either be the fermion charge $\psi$ or the trivial charge ${\mathbb{I}}$. (ii) Braiding operator $\hat{B}_{34}$ affecting entanglement between regions $A$ and $B$. Under the action of this braid, basis elements satisfying $a_1=a_2$ acquire a phase of ${\mathrm{e}}^{-\pi{\mathrm{i}}/4}$ whereas those satisfying $a_1\not=a_2$ acquire a phase of ${\mathrm{e}}^{3\pi{\mathrm{i}}/4}$. This transformation is inherently nonlocal, belonging to neither ${\mathcal{O}}'_A$ nor ${\mathcal{O}}'_B$, and so is capable of generating entanglement between regions $A$ and $B$. (iii) Alternative projection of the fusion tree of diagram (i) onto the disc, differing from that of diagram (i) precisely by the braiding operator of diagram (ii). Note that a density matrix in this basis does not admit the construction of a reduced density matrix on region $A$ or region $B$ without first applying an appropriate braiding operator. For example, applying $(\hat{B}_{34})^{-1}$ to diagram (iii) permits construction of ${\hat\rho_A}$ or ${\hat\rho_B}$ by recovering the basis of diagram (i). It is, however, compatible with the construction of a reduced density matrix for region $C$ or region $D$ as shown in diagram (iv). \[fig:sigmabraid\]](sigmabraid){width="246.0pt"} Now let ${\mathrm{P}}_{1A}$ and ${\mathrm{P}}_{1B}$ be pairs-of-pants decompositions of submanifolds $A$ and $B$, and let ${\mathrm{P}}_1$ be a decomposition of $M$ comprising the union of ${\mathrm{P}}_{1A}$ and ${\mathrm{P}}_{1B}$. If ${\mathrm{P}}_2$ is a second such decomposition comprising the union of ${\mathrm{P}}_{2A}$ and ${\mathrm{P}}_{2B}$ for the same submanifolds $A$ and $B$, then because the boundary between submanifolds $A$ and $B$ remains unchanged, the operator $\hat{B}$ mapping between these two bases necessarily admits a decomposition $$\hat{B}=\hat{B}_A\times\hat{B}_B,\qquad\hat{B}_A\in{\mathcal{O}}'_A,\quad\hat{B}_B\in{\mathcal{O}}'_B$$ where $\hat{B}_A$ maps between ${\mathrm{P}}_{1A}$ and ${\mathrm{P}}_{2A}$, and $\hat{B}_B$ maps between ${\mathrm{P}}_{1B}$ and ${\mathrm{P}}_{2B}$.[^2] In [Sec. \[sec:ardm\]]{} it was shown that to compute a reduced density matrix on a general submanifold $A$, it was first necessary to transform the density matrix into a basis permitting the construction of ${\hat\rho_A}$. The pairs-of-pants decomposition ${\mathrm{P}}$ associated with such a basis necessarily comprises the union of decompositions ${\mathrm{P}}_{A}$ and ${\mathrm{P}}_{B}$ for submanifolds $A$ and $B$ respectively, and consequently specification of regions $A$ and $B$ suffices to fix the basis up to a unitary transformation in ${\mathcal{O}}'_A\times{\mathcal{O}}'_B$. A transformation belonging to ${\mathcal{O}}'_A\times{\mathcal{O}}'_B$ may not modify the eigenvalues of the reduced density matrix, and therefore when computing a bipartite entanglement entropy it is unnecessary to give a specific pairs-of-pants decomposition of the manifold. It does remain necessary, however, to specify the decomposition of manifold $M$ into submanifolds $A$ and $B$. This is clearly seen in [Fig. \[fig:sigmabraid\]]{} where diagrams (i) and (iv) differ by the entanglement-generating operation of [Fig. \[fig:sigmabraid\]]{}(ii) but admit fusion tree bases distinguishable only in terms of their projection onto the manifold. Merely stating that the bipartition separates the anyons into two groups {1,2,4} and {3,5,6} fails to distinguish between these two distinct scenarios with different entanglement entropies. Summary of considerations ------------------------- Systems of non-Abelian anyons admit larger families of entanglement monotones than do systems governed by Abelian statistics. When computing a bipartite entanglement entropy for non-Abelian anyons inhabiting a two-dimensional manifold $M$, in order to fully specify which entanglement entropy has been computed it is necessary to explicitly specify the bipartitioning of the manifold, not just the separation of the anyonic excitations into two groups. It is not, however, necessary to provide a specific pairs-of-pants decomposition of the manifold. In contrast, the pairs-of-pants decomposition *must* be given in order to unambiguously describe a state or operator on a physical system, including the density matrix. This consideration has, in the past, been frequently overlooked for anyon chains as the ordering of the anyons on the fusion tree and the natural ordering of anyons on the manifold (e.g. {1,2,3,4,5,6} in [Fig. \[fig:sigmabraid\]]{}) are assumed to coincide. The mapping between the fusion tree and the pairs-of-pants decomposition of the manifold attains greater significance in systems which are not quasi-one-dimensional, and in contexts which force a choice of basis that does not correspond to the natural ordering of the chain. An example of this is the construction of a reduced density matrix suitable for computing bipartite entanglement entropies between regions $A$ and $B$ in [Fig. \[fig:sigmabraid\]]{}(i). Discussion ========== The study of entanglement entropies in topologically ordered systems is currently a hot topic[@castelnovo2007; @chen2010; @eisert2010; @levin2006; @furukawa2007; @hamma2008; @haque2007; @jiang2012; @zhang2011; @nussinov2009; @papanikolaou2007; @li2008; @kato2013] with a major focus being the calculation of the topological entanglement entropy[@kitaev2006a] as an aid to the classification of different topologically ordered phases of matter. Fortunately this quantity depends only upon the topology of the manifold and the nature of the anyon model describing the quasiparticle excitations, and thus even in studies of non-Abelian anyons (e.g. [Ref. ]{}) its calculation is unaffected by the considerations raised in the present discussion. As interest increases in the condensed matter physics of anyonic systems, however, attention will turn to the behaviours of the entanglement monotones themselves, and it is vital that these quantities be well-defined if successful study is to be made of entanglement in systems of non-Abelian anyons. Previous work on the subject[@hikami2008] has concentrated on defining entanglement entropy with respect to bipartitions on an abstract fusion tree, separating the anyons into two groups (denoted $A$ and $B$) without attention to the manifold, and requiring that the anyons from both groups be contiguous on that tree. For the study of real physical systems, this definition is insufficient as the relationship of the physical system to the fusion tree is ambiguous, resulting in bipartite entanglement entropies which are not well-defined. To resolve this ambiguity it is necessary to specify the bipartition of the system in terms of the manifold rather than simply grouping the anyons on the fusion tree. ![The prescription given in permits calculation of entanglement entropies between the centre and edge of a chain of anyons even when anyons in one region are not contiguous on the fusion tree. This is essential when working with disjoint submanifolds, as in diagram (i), and may also simplify the calculation of entanglement entropies for central regions when employing a naturally-ordered fusion tree basis on the chain, as shown in diagram (ii). \[fig:expandingregion\]](compbasis){width="246.0pt"} Furthermore, it has proved possible to relax the requirement that anyons from both regions $A$ and $B$ be contiguous on the fusion tree. The most stringent requirements of contiguousness arise when a contiguous manifold $M$ is divided into two contiguous submanifolds $A$ and $B$. Even then, in order to permit construction of bipartite entanglement entropies it suffices that the anyons inhabiting at least one of these two regions (but not necessarily both) be consecutive in the fusion tree on $M$. If constructing the reduced density matrix on submanifold $A$, then relaxing the requirement that anyons in submanifold $B$ be consecutive in the fusion tree basis of $M$ leads to a construction for bipartite entanglement entropies not considered in [Ref. ]{}. This construction is essential for computing entanglement entropies where one submanifold is made up of two disjoint parts as shown in [Fig. \[fig:expandingregion\]]{}(i), and may also be convenient when comparing entanglement between the centre and ends of a chain as shown in [Fig. \[fig:expandingregion\]]{}(ii). It is also interesting to compare the current work with that of @kato2013,[@kato2013] in which entanglement entropies for non-Abelian anyonic systems are computed by embedding these systems into the Hilbert space of a spin system (see also [Ref. ]{}). In choosing a basis in which to perform this embedding, and in maintaining a notion of locality during the mapping from the manifold to the spin system, the embedding process and bipartition on the spin system may be understood to implicitly define submanifolds $A$ and $B$ on the original manifold (up to the equivalence class of deformations which are trivial when expressed in the fusion tree basis). Consequently the work of these authors is entirely consistent with, and may be understood in terms of, the framework presented herein. Finally, it is noted that the considerations raised in the present paper are unique to non-Abelian anyons. When applied to systems of fermions, bosons, or Abelian anyons, the definitions presented above always reduce to the usual definitions of entanglement entropies for these systems. The author thanks the Ontario Ministry of Research and Innovation Early Researcher Awards for financial support. Research at Perimeter Institute is supported by the Government of Canada through Industry Canada and by the Province of Ontario through the Ministry of Research and Innovation. [^1]: For situations where the boundary of intersects the boundary of , a puncture is not traced if any part of its boundary coincides with a boundary on . As before, operators coupling the charge on the boundary of to other charges in will belong to if they are trivial on . [^2]: It may be useful to note that the total charge of submanifold $B$ is accessible from within manifold . Consequently, operators involving the braid of anyons in around the entirety of , and vice versa, may be found in and respectively.
{ "pile_set_name": "ArXiv" }
--- author: - | M. V. Kompaniets\ St. Petersburg State University\ 7/9 Universitetskaya nab., St. Petersburg 199034, Russia\ E-mail: - | \ All Souls College, University of Oxford, OX1 4AL Oxford, UK\ E-mail: bibliography: - 'refs.bib' title: 'Renormalization group functions of $\field^4$ theory in the $\MS$-scheme to six loops' --- Introduction {#sec:intro} ============ A scalar field with quartic self-interaction is one of the simplest models in quantum field theory. With the discovery of the Higgs boson, it is now part of the standard model of elementary particles. However, it also appears elsewhere and in particular its application as a mean-field approximation to statistical systems is one way to study phase transitions [@Vasilev; @ZinnJustin; @KleinertSchulteFrohlinde:CriticalPhi4]. Critical exponents can be inferred from the renormalization group functions. Therefore, perturbative calculations in $\field^4$ theory have a long history and they reached a high loop order. The 5-loop field anomalous dimension and beta function were announced already in [@ChetyrkinKataevTkachov:5loopPhi4] and [@ChetyrkinGorishnyLarinTkachov:5loopPhi4; @Kazakov:MethodOfUniqueness]. It took ten years until a different group reproduced this calculation and revealed some inaccuracies in [@ChetyrkinGorishnyLarinTkachov:5loopPhi4]; corrected results were published in [@KNFCL:5loopPhi4]. A complete numeric check of all these analytic results was completed only much later in [@AdzhemyanKompaniets:5loopNumerical]. Very recently, the first 6-loop result for the field anomalous dimension was published [@BatkovichKompanietsChetyrkin:6loop]. In this note we present the remaining renormalization group functions of $\field^4$ theory at $6$-loop order, for the minimal subtraction (MS) scheme in dimensional regularization. The results for the $O(N)$-symmetric model and the corresponding critical exponents will be published elsewhere. Instead, we focus here on the technical aspects and the novel tools used in our calculation. Note that a completely independent calculation of the $6$-loop beta function with a very different method has recently been finished by Oliver Schnetz [@Schnetz:NumbersAndFunctions] and confirmed our result. We start with the well-known representation of $Z$-factors in terms of massless propagators (also known as *$p$-integrals*). The challenge is then to compute the $\eps$-expansions of such integrals and we refer to [@Vasilev; @ZinnJustin; @KleinertSchulteFrohlinde:CriticalPhi4] for a comprehensive discussion of traditional techniques. The general philosophy has hitherto been to exploit various relations, in particular the infrared $\Rstar$-operation [@ChetyrkinTkachov:InfraredR; @ChetyrkinSmirnov:Rcorrected; @Chetyrkin:RRR; @ChetyrkinGorishnyLarinTkachev:preprint], to relate the counterterms to simpler integrals. Recently, the cumbersome $\Rstar$-operation was automated in [@BatkovichKompaniets:Toolbox]. Augmented with the results [@BaikovChetyrkin:FourLoopPropagatorsAlgebraic; @SmirnovTentyukov:FourLoopPropagatorsNumeric; @LeeSmirnov:FourLoopPropagatorsWeightTwelve] for $4$-loop $p$-integrals and integration by parts (IBP) [@ChetyrkinTkachov:IBP; @VasilevPismakHonkonen:largeNeta], this traditional approach was used for the $6$-loop calculation of the field anomalous dimension [@BatkovichKompanietsChetyrkin:6loop]. This already required additional tricks for two propagators (see section 5 in [@BatkovichKompanietsChetyrkin:6loop]), but the situation is much more severe for the $4$-point diagrams contributing to the $6$-loop beta function: the method fails for $22$ graphs. Of these, $10$ are primitive and have been known for long [@BroadhurstKreimer:KnotsNumbers; @Schnetz:Census], but the remaining $12$ graphs contain subdivergences and require new techniques. Initially we calculated these diagrams via parametric integration [@Brown:TwoPoint; @Panzer:HyperIntAlgorithms] and a new subtraction method for subdivergences (different to the one we elaborate on here), which we will discuss elsewhere. Instead, in this paper we present a simple new approach based on parametric integration using hyperlogarithms [@Brown:TwoPoint; @Brown:PeriodsFeynmanIntegrals; @Panzer:PhD] which allows us to compute the counterterms of all $6$-loop $2$- and $4$-point graphs analytically, with the sole exception of a single graph which is, however, well-known [@Schnetz:K34; @Schnetz:GraphicalFunctions; @BroadhurstKreimer:KnotsNumbers]. We use the implementation [@Panzer:HyperIntAlgorithms] of this method, which was presented at the preceding conference [@Panzer:LL2014].Our new ingredient is an efficient and general procedure to generate convergent integrands for integrals which initially have subdivergences. This is achieved with a BPHZ-like renormalization scheme with one-scale counterterms introduced in [@BrownKreimer:AnglesScales]. A great advantage of this approach is that it is easily automatized and applicable to all integrals, contrary to the multitude of tricks and tools for special classes of diagrams combined in the traditional calculations. This simplifies the computation, reduces the possibilities for errors in the programs and ensures the reproducibility of our results.[^1] Also our calculation provides a confirmation of the $5$-loop results [@KNFCL:5loopPhi4] and the $6$-loop field anomalous dimension [@BatkovichKompanietsChetyrkin:6loop] with very different methods. Field theory and renormalization {#sec:phi4theory} ================================ The renormalized Lagrangian of the scalar $\field^4$ model in $\D=4-2\eps$ Euclidean dimensions is $$\lagrangian =% \int \dd[\D] x \left( \frac{1}{2}m^2 Z_1\field^2 +\frac{1}{2}Z_2\left(\partial\field \right)^2 +\frac{16\, \pi^2}{4!}Z_4 \, g\, \mu^{2\eps}\, \field^4 %\right) , % \label{eq:action}% \label{eq:lagrangian}%$$ with an arbitrary mass scale $\mu$. The $Z$-factors relate the renormalized field $\field$, mass $m$ and coupling $g$ to the bare field $\field_0$, bare mass $m_0$ and bare coupling $g_0$ via $$\begin{gathered} \label{eq:Z-factors}% Z_{\field} = \frac{\field_0}{\field} = \sqrt{Z_2} ,\quad Z_{m^2} = \frac{m_0^2}{m^2} = \frac{Z_1}{Z_2} \quad\text{and}\quad Z_g = \frac{g_0}{\mu^{2\eps} g} = \frac{Z_4}{Z_2^2} . % Z_1 = Z_{m^2}Z_{\field}^2, \qquad % Z_2 = Z_{\field}^2, \qquad % Z_4 = Z_{g} Z_{\field}^4 % \\ % \field_0 = \field Z_\field, \qquad % m_0^2 = m^2 Z_{m^2}, \qquad % g_0 = g\mu^{2\eps} Z_g. % \label{eq:bare-to-renormalized}%\end{gathered}$$ In dimensional regularization [@tHooftVeltman:RegularizationGaugeFields] and minimal subtraction, the $Z$-factors depend only on $\eps$ and $g$ as $$Z_i = Z_i(g,\eps) = 1 + \sum_{k=1}^{\infty} \frac{Z_{i,k}(g)}{\eps^k} \label{eq:Z-factor-expansion}%$$ and determine the renormalization group functions [@tHooft:DimRegRG; @Collins:CountertermsDimReg]. We are interested in the beta function $$\beta(g) = \restrict{ \mu\frac{\partial g}{\partial \mu} }{g_0} = - 2\eps \left( \frac{\partial \log (g Z_g)}{\partial g} \right)^{-1} = -2\eps g + 2 g^2 \frac{\partial Z_{g,1}(g)}{\partial g} %=-g\left(2\varepsilon + \gamma_g(g)\right) \label{eq:beta-def}%$$ and the anomalous dimensions for the field and mass, defined by $$\gamma_{i}(g) = \restrict{ \mu \frac{\partial \log Z_{i}}{\partial \mu} }{g_0,m_0,\field_0} = \beta(g) \frac{\partial \log Z_i(g)}{\partial g} = -2 g \frac{\partial Z_{i,1}(g)}{\partial g} \quad\text{for}\quad i=m^2, \field%, g . \label{eq:anom-dims}%$$ Counterterms from massless propagators {#sec:counterterms} ====================================== The $Z$-factors arise as the counterterms for the one-particle irreducible correlation functions $\Gamma_N$ of $N=2$ and $N=4$ fields. In terms of the Bogoliubov-Parasiuk $\Rbar$-operation [@BogoliubovParasiuk:Kausalfunktionen; @BogShirk], which subtracts UV subdivergences from Feynman integrals, the $Z$-factors can be expressed as $$\label{eq:Z-from-R} \begin{split} Z_1 &= 1 + \partial_{m^2} \Poles \Rbar {\Gamma}_2 (p,m^2,g,\mu) , \\ Z_2 &= 1 + \partial_{p^2} \Poles \Rbar {\Gamma}_2 (p,m^2,g,\mu) \quad\text{and} \\ Z_4 &= 1 + \Poles \Rbar {\Gamma}_4 (p,m^2,g,\mu)/g . \end{split}$$ The operator $\Poles$ is the minimal subtraction scheme (), meaning the projection $$\Poles \left( \sum_n c_n \eps^n \right) \defas \sum_{n<0} c_n \eps^n \label{eq:Poles-def}%$$ onto the pole part with respect to the dimensional regulator $\eps=(4-\D)/2$ and it ensures the form of the $Z$-factors. Recall that these depend only on $g$ and $\eps$, which allows us to simplify the calculation tremendously [@Vasilev; @ZinnJustin; @KleinertSchulteFrohlinde:CriticalPhi4]: - The action of $\partial_{m^2}$ turns a propagator diagram into a sum of diagrams with a squared propagator, $$\frac{\partial}{\partial m^2} \frac{1}{k^2+m^2} = - \frac{1}{k^2+m^2} \frac{1}{k^2+m^2}$$ which may be interpreted as a $4$-point graph with two additional, vanishing external momenta entering at a common vertex. This means that $Z_{m^2}$ can be expressed in terms of contributions to $Z_4$ of a subset of the $4$-point diagrams with modified symmetry factors. - We can set all internal masses to zero in the $2$-point diagrams for the computation of $Z_2$, such that $\Gamma_2(p,0,g,\mu)$ is given by $p$-integrals. - Also in $\Gamma_4$ we may set all internal masses to zero. Furthermore, we may set external momenta to zero until only a propagator ($p$-integral) with two external legs remains. This is called infrared rearrangement (IRR) [@Vladimirov:ManyLoopPhi4; @ChetyrkinKataevTkachov:Gegenbauer], see figure \[fig:one-scale\].[^2] These standard techniques express all $Z$-factors in terms of $p$-integrals without non-physical infrared divergences. In the next section, we will explain how these integrals can be computed by parametric integration, at least to the sixth loop order. As mentioned in the introduction, we will dispose of $\Rstar$ and IBP completely. Parametric integration {#sec:parametric} ====================== Of the many recent advances made in the evaluation of Feynman integrals, parametric integration is one of the most powerful methods for $p$-integrals; surpassed only by the position-space approach of graphical functions [@Schnetz:GraphicalFunctions] and the combination [@GolzPanzerSchnetz:GfParam] of both techniques used in [@PanzerSchnetz:Phi4Coaction; @Schnetz:NumbersAndFunctions]. The starting point is the representation of a Feynman integral $\FI(G)$ associated to a Feynman graph $G$ in terms of variables $\alpha_e$ associated to each edge $e\in E(G)$ of $G$ (these are called Schwinger-, Feynman- or $\alpha$-parameters). Let us write $\loops(G)$ for the number of loops of $G$ and $$\sdc(G) = \sum_{e\in E(G)} \pe_e - \loops(G) \frac{\D}{2} = \sum_{e\in E(G)} \pe_e - 2 \loops(G) + \eps \loops(G) \label{eq:sdc}%$$ for the superficial degree of convergence of $G$ given by power counting. The variables $\pe_e$ encode the exponents to which the momentum space propagators $1/(k_e^2 + m_e^2)^{\pe_e}$ are raised. Choose an arbitrary edge $e_0\in E(G)$.[^3] Then the parametric representation for $\FI(G)$ is $$\FI(G) = \Gamma(\sdc(G)) \left( \prod_{e\in E(G)} \int_0^{\infty} \frac{\alpha_e^{\pe_e-1} \dd \alpha_e}{\Gamma(\pe_e)} \right) \frac{\delta(1-\alpha_{e_0})}{\U^{\D/2-\sdc(G)} \F^{\sdc(G)}} \label{eq:alpha-rep}%$$ in terms of the Symanzik polynomials $\U$ and $\F$ [@Nakanishi:GraphTheoryFeynmanIntegrals; @Nakanishi:ParametricAnalyticPerturbation]. These can be expressed as $$\U = \sum_{T} \prod_{e \notin T} \alpha_e \quad\text{and}\quad \F = \sum_{C} p^2(C) \prod_{e\in C} \alpha_e + \U \sum_{e\in E(G)} m_e^2 \alpha_e \label{eq:UF}%$$ in terms of spanning trees $T$ and cuts $C$ which separate $G$ into 2 components; $p(C)$ is the momentum flowing through the cut edges $C$ [@BognerWeinzierl:GraphPolynomials]. In our case, all masses $m_e$ are zero and only a single external momentum $p$ is flowing through the graph (it has just $2$ external legs), such that $\F$ (and $\U$) are linear in all edge variables $\alpha_e$. For example, the graph $$G = \Graph[0.5]{dunce_1s} \qquad\text{gives}\qquad \begin{aligned} \U &= (\alpha_1+\alpha_2)(\alpha_3 + \alpha_4) + \alpha_3 \alpha_4 \quad\text{and}\\ \F &= p^2 \alpha_1 (\alpha_2 \alpha_3 + \alpha_2 \alpha_4 + \alpha_3 \alpha_4). \end{aligned} \label{eq:UF-dunce}%$$ Note that the momentum dependence factors out of the integral by simple power counting: $$\FI(G,p^2) = p^{-2\sdc(G)} \FI(G, 1). \label{eq:momentum-dependence}%$$ It was noted in [@Brown:TwoPoint] that for many massless propagators, the integrals over $\alpha_e$ in can be performed one after the other in terms of multiple polylogarithms if one chooses a good order for these integrations. Graphs which admit such a good order are called *linearly reducible* and can be computed, order by order in $\eps$, with the algorithm described in [@Brown:TwoPoint; @Brown:PeriodsFeynmanIntegrals; @BognerBrown:GenusZero] and implemented in [@Bogner:MPL; @Panzer:HyperIntAlgorithms; @Panzer:PhD]. In [@Panzer:MasslessPropagators] it was found that all massless propagators with $\leq 4$ loops are indeed linearly reducible, and the same method was even applied to some $6$-loop $p$-integrals [@Panzer:LL2014; @Panzer:PhD]. The remaining challenge to the straightforward application of parametric integration in the linearly reducible case is the presence of subdivergences. Note that the $\eps$-expansion needs to be performed on the integrand in . Subdivergences imply that the resulting integrals are divergent and not defined. Therefore we must find integrands which are free of subdivergences. Sector decomposition [@BinothHeinrich:SectorDecomposition; @BognerWeinzierl:ResolutionOfSingularities] is the standard approach to solve this problem and by now very well established and powerful [@BHJKSZ:SecDec3.0; @Smirnov:FIESTA4]. However, it is best suited for numerical calculations; the huge number of sectors that are generated makes it very inefficient for the high loop orders under consideration here, and also it introduces changes of variables which make the analytic integration of each sector much more complicated. These problems can in principle be avoided with the help of IBP relations, because it is always possible to write a Feynman integral in terms of master integrals without subdivergences [@ManteuffelPanzerSchabinger:QuasiFinite]. Unfortunately, such IBP reductions are too complicated in our case. Only at this conference, their solution at $4$-loops was achieved [@RUV:ForcerLL2016] and highlighted by the impressive computation [@BaikovChetyrkinKuehn:5loopQCD] of the $5$-loop QCD $\beta$-function (25 years after the $\field^4$-result). An extension of IBP to the next loop order seems out of reach with current technologies and some new ideas like [@ManteuffelSchabinger:Novel] are being investigated. Luckily, it is possible to avoid IBP altogether by the method explained below. Note that this is feasible only because there are only very few graphs in $\field^4$ theory (just $627$ graphs need to be computed for $\Gamma_4$ at six loops). In contrast, scalar $\field^3$ theory (in six dimensions) has many more diagrams and was therefore evaluated at mere $4$ loops only very recently [@Gracey:4loopPhi3; @Pismensky:etaphi3in4loop]. Subdivergences and one-scale renormalization scheme {#sec:one-scale-scheme} =================================================== The $\Rbar$-operation subtracts all UV-subdivergences of a given Feynman integral $\FI(G)$ and a final overall subtraction renders the integral itself finite [@Hepp:BP]. The famous forest formula [@Zimmermann:Bogoliubov] $$\R \FI(G) = (1-\K) \Rbar \FI(G) = \sum_{F} (-1)^{\abs{F}} \left[ \FI(G/F) - \K \FI(G/F) \right] \prod_{\gamma \in F} \K \FI(\gamma/F) \label{eq:forest-formula}%$$ expresses the renormalized integral $\R\FI$ explicitly as a sum over all forests $F$ (sets of proper subdivergences which are nested or disjoint).[^4] The operator $\K$ determines the renormalization scheme and is given by for . While guarantees that all poles in $\eps$ cancel in the sum of the regularized *integrals*, we do not obtain subdivergence-free *integrands* this way. The reason is that prior to the integration of the $\alpha$-parameters in , there are no poles in $\eps$ and so we clearly cannot commute the subtraction $\Poles$ with the integration $\FI$. Put differently, by definition of the -scheme, its counterterms depend on the regularization. This is not the case for schemes like the original BPHZ where the counterterms are defined by expansion, in the masses and momenta, around a fixed set of values (the renormalization point). It is well-known that in this case one may perform the subtractions under the integral sign; these subtracted integrands give convergent integrals even in $\D=4$ dimensions ($\eps=0$). Such schemes are therefore ideal for parametric integration and were studied in this context in detail in [@BrownKreimer:AnglesScales]. In particular, the authors suggested a scheme $\OneScale$ where all counterterms are *one-scale*—in other words, $p$-integrals. For logarithmically divergent graphs $G$ they set $$\OneScale \FI(G) \defas \begin{cases} \restrict{\FI(G)}{p^2=1} & \text{if $G$ is a $p$-integral and} \\ \restrict{\FI(G_{\onescale})}{p^2=1} & \text{if $G$ has more than two external legs,} \end{cases} \label{eq:onescale-def}%$$   $G=\Graph[0.4]{dunceleft}$ $G_{\onescale} = \Graph[0.4]{dunceleft1} \quad\text{or}\quad \Graph[0.4]{dunceleft2} $ $G_{\onescale}^{\text{IR-unsafe}} = \Graph[0.4]{dunceleft3}$  where $G_{\onescale}$ is any infrared-safe rearrangement of $G$ with only two external legs. We already mentioned this method in section \[sec:counterterms\] as the crucial simplification in the computation of $\Rbar \Gamma_4$: There is always at least one way to nullify external momenta to obtain a $p$-integral without introducing infrared divergences, see figure \[fig:one-scale\]. Crucially, the subtractions for $\K=\OneScale$ may be performed under the integral sign. We can thus safely expand the subtracted integrand in $\eps$ and integrate each term individually as described in section \[sec:parametric\] to obtain the renormalized Feynman integral $\R_{\onescale} \FI(G)$ in this scheme. It depends on $p^2$ and vanishes at $p^2=1$ by construction. With , we find $$\restrict{ \partial_{p^2} \R_{\onescale} \FI(G) }{p^2=1} = -\sum_{F} (-1)^{\abs{F}} \sdc(G/F) \OneScale \FI(G/F) \prod_{\gamma \in F} \OneScale \FI(\gamma/F) = - \sdc(G) \OneScale \FI(G) + \ldots$$ and solve for the unrenormalized Feynman integral in dimensional regularization: $$\OneScale\FI(G) = - \frac{\restrict{\left(\partial_{p^2} \R_{\onescale}\FI(G)\right)}{p^2=1}}{\sdc(G)} - \sum_{\emptyset \neq F} (-1)^{\abs{F}} \frac{\sdc(G/F)}{\sdc(G)} \OneScale\FI(G/F) \prod_{\gamma \in G} \OneScale \FI(\gamma/F) . \label{eq:p-integral-from-one-scale}%$$ Note how this formula expresses any given $p$-integral $\FI(G)$ (at $p^2=1$) in terms of - an integral $\R_{\onescale} \FI(G)$ without subdivergences (directly integrable in $\alpha$-parameters) and - products of lower-loop $p$-integrals which can be computed recursively with the same method. An example for the $\Rbar$-operation in this scheme is $$\begin{split}\label{eq:onescale-example}% \Rbar_{\onescale} \FI\left( \Graph[0.3]{sauron} \right) &= \FI \left( \Graph[0.3]{sauron} \right) - \restrict{\FI\left( \Graph[0.4]{bubblevert} \right)}{p^2=1} \FI\left( \Graph[0.4]{glasses}\right) \\&\quad - \restrict{\FI\left( \Graph[0.3]{dunceleft1} \right)}{p^2=1} \FI\left( \Graph[0.4]{bubblehorz}\right) - \restrict{\FI\left( \Graph[0.3]{dunceright1} \right)}{p^2=1} \FI\left( \Graph[0.4]{bubblehorz}\right) \\&\quad + 2 \restrict{\FI\left( \Graph[0.35]{bubblevert}\right)}{p^2=1} \restrict{\FI\left( \Graph[0.4]{bubblehorz} \right)}{p^2=1} \FI\left( \Graph[0.4]{bubblehorz}\right) \end{split}$$ where the terms in the last line come from the forests $F=\set{\Graph[0.2]{bubblevert}, \Graph[0.15]{dunceleft}}$ and $F=\set{\Graph[0.2]{bubblevert}, \Graph[0.15]{dunceright}}$. Note that $\partial_{p^2} \R_{\onescale} = \partial_{p^2} \Rbar_{\onescale}$ because $\R_{\onescale} = \Rbar_{\onescale} - \restrict{\Rbar_{\onescale}}{p^2=1}$ differ only by a constant. So we can apply $\partial_{p^2}$ to the left hand side and compute this integral parametrically to the desired order in $\eps$, and then use to express the unrenormalized $\FI(\Graph[0.15]{sauron})$ in terms of this quantity and the products of one- and two-loop integrals on the right hand side of . In section \[sec:product-polynomials\] we discuss a simpler example. For the overall quadratically divergent 2-point integrals contributing to $\Gamma_2$, the overall subtraction becomes $\OneScale\FI(G) = \restrict{\FI(G)}{p^2=1} + (p^2-1) \partial_{p^2}\restrict{\FI(G)}{p^2=1}$. We just take one additional derivative with respect to $p^2$ and compute $\partial_{p^2}^2 \R_{\onescale} = \partial_{p^2}^2 \Rbar_{\onescale}$ in this case, replacing all $\sdc(G)$ in by $-\sdc(G) (\sdc(G)+1)$ and similarly for $\sdc(G/F)$. Note that we can always factor off quadratic *sub*divergences, see section \[sec:factorization\]. Hence our above discussion of the simple subtractions for logarithmic subdivergences (together with at most one additional overall derivative) is sufficient for our calculation. In principle though, the method can also be applied more generally [@BrownKreimer:AnglesScales]. Remarks on the calculation {#sec:calculation} ========================== As was mentioned in the introduction, we wrote computer programs to fully automate the entire calculation. The main program is written in using the / library, which provide a very convenient way to manipulate Feynman graphs [@GraphState; @BatkovichKompaniets:Toolbox]. It generates the graphs, computes their symmetry factors and combines them into the counterterms to calculate the anomalous dimensions and the beta function via and . In order to evaluate the corresponding integrals $\FI(G)$, the program computes the forest formula and chooses one-scale structures for each graph, implementing the scheme . The resulting expressions of $\R_{\onescale} \FI(G)$ as linear combinations of products of integrals are passed on to a script. It computes the parametric representation according to section \[sec:product-polynomials\], performs the $\eps$-expansion and calls [@Panzer:HyperIntAlgorithms] to perform the integration of $\restrict{\left(\partial_{p^2} \R_{\onescale}\FI(G)\right)}{p^2=1}$. This is the most time-consuming step. Finally, the program reconstructs the value of the unsubtracted $p$-integral according to . A detailed discussion and publication of these programs is under preparation. Factorization {#sec:factorization} ------------- Due to the trivial momentum dependence of $p$-integrals, every $2$-point subgraph $\gamma$ can be integrated out separately if one replaces it with a propagator of index $\pe_e=\sdc(\gamma)$. For example, $$\FI\left( \Graph[0.8]{g1c} \right) = \restrict{\FI\left( \Graph{g2c} \right)^2}{p^2=1} \cdot \restrict{ \FI \left(\Graph{g4c} \right)}{p^2=1} \cdot \FI \left( \Graph{g5c} \right)$$ Therefore, we only need to compute $p$-integrals for $2$-connected graphs (graphs which do not have any $p$-integral as a subgraph). In particular this means that all quadratic propagator subdivergences factor out and we only have to deal with logarithmic subdivergences, where the simple subtractions suffice. Note that this factorization procedure introduces $\eps$-dependent propagator exponents $\pe_e$. With traditional methods it was very difficult to compute such integrals. For parametric integration, the effect of such exponents is just that the $\eps$-expansion of the integrand in also includes logarithms $\log (\alpha_e)$ of individual Schwinger parameters, in addition to $\log(\U)$ and $\log(\F)$. This causes no problem for the integration algorithms. For example, comprehensive results for *arbitrary* $\pe_e$ in the case of $4$-loop $p$-integrals were computed this way in [@Panzer:MasslessPropagators]. Parametric representation for products {#sec:product-polynomials} -------------------------------------- The forest formula for $\R_{\onescale} \FI(G)$ contains products of Feynman integrals. Note that the graph polynomials as defined in vanish for products (disjoint unions) of graphs. Instead, for the parametric integral representation for a product $G = \prod_{i=1}^n G_i$ one has to set [@BrownKreimer:AnglesScales] $$\U_G = \prod_{i=1}^n \U_{G_i} \quad\text{and}\quad \F_G = \sum_{i=1}^n \F_{G_i} \prod_{j\neq i}\U_{G_j} \quad\text{such that}\quad \frac{\F_G}{\U_G} = \sum_{i=1}^n \frac{\F_{G_i}}{\U_{G_i}} . \label{eq:UF-product}%$$ In order to ensure the cancellation of all subdivergences in the parametric integral representation for $\R_{\onescale} \FI(G)$, it is crucial that one tracks correctly the individual edges of the sub- and quotient graphs in . This is illustrated in the following example of a single logarithmic UV-subdivergence from , where we explicitly show the edge labels: $$\Rbar_{\onescale} \FI \left( \Graph[0.45]{dunce_1s} \right) = \FI\left( \Graph[0.45]{dunce_1s} \right) - \restrict{\FI\left( \Graph[0.45]{dunce_1s_sub} \right)}{p^2=1} \FI\left( \Graph[0.45]{dunce_1s_co} \right) . \label{eq:dunce-labels}%$$ The graph polynomials for the graph $G=\Graph[0.2]{dunce_1s}$ were given in and the prescription for the product of the subgraph $\gamma=\Graph[0.2]{dunce_1s_sub}$ at $p^2=1$ and the quotient $G/\gamma=\Graph[0.2]{dunce_1s_co}$ yields $$\U_{\gamma\cdot G/\gamma} = (\alpha_1+\alpha_2)(\alpha_3+\alpha_4) \quad\text{and}\quad \F_{\gamma \cdot G/\gamma} = \alpha_3 \alpha_4 (\alpha_1+\alpha_2) + p^2 \alpha_1 \alpha_2 (\alpha_3 + \alpha_4) . \label{eq:UF-dunce-product}%$$ The parametric representation for $\restrict{\partial_{p^2} \R_{\onescale} \FI(G)}{p^2=1}$ in this case is $$\Gamma(2\eps) \int_0^\infty \dd \alpha_1 \cdots \int_0^{\infty} \dd \alpha_4 \delta(1-\alpha_{e_0}) \left[ \frac{-2\eps}{\U_G^{2-3\eps} \F_G^{2\eps}} - \frac{-\eps \alpha_1 \alpha_2(\alpha_3 + \alpha_4)}{\U_{\gamma \cdot G/\gamma}^{2-2\eps} \F_{\gamma \cdot G/\gamma}^{1+\eps}} \right]_{p^2=1} \label{eq:dunce-parametric}%$$ which is finite at $\eps\rightarrow 0$ and may be integrated order by order in $\eps$. Many more examples are discussed in [@BrownKreimer:AnglesScales]. Simplification at leading order ------------------------------- For the leading order in $\eps$, we evaluate the parametric integrand in at $\eps=0$. If a graph $G$ is logarithmically divergent, that is $\restrict{\sdc(G)}{\eps=0}=0$, this means that $\F$ drops out completely and only the polynomial $\U$ plays a role for the integration. It determines the pole in $\eps$ coming from the $\Gamma(\sdc)$-prefactor; higher orders in $\eps$ are not needed for the determination of the $Z$-factors. This simplifies the parametric integration considerably. Since $\U$ is linear in all $\alpha_e$, the first integration is elementary and the result can be interpreted as the $p$-integral $G\setminus e$ (with the external momenta entering at the vertices that were incident to $e$) [@Panzer:MasslessPropagators; @Brown:TwoPoint].[^5] All $4$-point $6$-loop graphs contributing to $\Gamma_4$ are therefore effectively expressed in terms of $5$-loop $p$-integrals. In the traditional approach, this simplification is achieved with the $\Rstar$-operation [@ChetyrkinTkachov:InfraredR; @ChetyrkinSmirnov:Rcorrected; @Chetyrkin:RRR; @ChetyrkinGorishnyLarinTkachev:preprint]. Linear reducibility ------------------- Not all $6$-loop $p$-integrals are linearly reducible, but as explained we are essentially only computing $5$-loop $p$-integrals. It turns out that indeed all integrals needed for the $6$-loop calculation of the $Z$-factors are linearly reducible, with just a single exception. This is the primitive graph $$\FI\left( \Graph[0.2]{k_34legs} \right) = \frac{288}{30\eps} \Big( 58 \mzv{8} - 45 \mzv{3} \mzv{5} - 24 \mzv{3,5} \Big) + \bigo{\eps^0}, \label{eq:P64}%$$ $P_{6,4}$ in the notation of [@Schnetz:Census]. It has been known numerically since [@Broadhurst:5loopsbeyond] and in [@BroadhurstKreimer:KnotsNumbers] it was identified in terms of Riemann zeta values $\mzv{k} = \sum_{n=1}^{\infty} 1/n^k$ and the double zeta value $$\mzv{3,5} \defas \sum_{1\leq n < m} \frac{1}{n^3 m^5} \approx 0.037707673 . \label{eq:zeta35}%$$ It is conjectured that $\mzv{3,5}$ cannot be expressed as a rational linear combination of products of Riemann zeta values. An analytic calculation confirming the result was first provided in [@Schnetz:K34] and recently with the beautifully elegant method of graphical functions [@Schnetz:GraphicalFunctions]. While $P_{6,4}$ is not linearly reducible in the strict sense and thus not computable with , this is in fact only a limitation of this implementation which could be lifted: It is known how $P_{6,4}$ can be integrated parametrically by splitting the integrand as described in [@Brown:TwoPoint]. Results and checks {#sec:result} ================== The beta function of $\field^4$-theory in the $\MS$ scheme in $\D=4-2\eps$ dimensions to order $g^7$ is $$\begin{split} \label{eq:beta-result}% \beta^{\MS}(g) &= -2 \varepsilon g +3 g^2 - \lfrac{17}{3} g^3 +\left( 12 \mzv{3} +\lfrac{145}{8} \right) g^4 -\left( 120 \mzv{5} -18 \mzv{4} +78 \mzv{3} +\lfrac{3499}{48} \right) g^5 \\&\quad +\left( \numprint{1323} \mzv{7} +45 \mzv[2]{3} -\lfrac{675}{2} \mzv{6} +987 \mzv{5} -\lfrac{1189}{8} \mzv{4} +\lfrac{7965}{16} \mzv{3} +\lfrac{764621}{2304} \right) g^6 \\&\quad -\left( \lfrac{46112}{3} \mzv{9} +768 \mzv[3]{3} +\lfrac{51984}{25} \mzv{3,5} -\lfrac{264543}{25} \mzv{8} +\numprint{4704} \mzv{3}\mzv{5} +\lfrac{63627}{5} \mzv{7} -162 \mzv{3}\mzv{4} \right.\\&\left.\qquad\ +\lfrac{8678}{5} \mzv[2]{3} -\lfrac{6691}{2} \mzv{6} +\lfrac{63723}{10} \mzv{5} -\lfrac{16989}{16} \mzv{4} +\lfrac{779603}{240} \mzv{3} +\lfrac{18841427}{11520} \right) g^7 +\bigo{g^8} \\&\approx -2 \epsilon g +3 g^2 -5.667 g^3 +32.55 g^4 -271.6 g^5 +2849 g^6 -34776 g^7 +\bigo{g^8} . \end{split}$$ For the anomalous dimension of the mass to order $g^6$ we find $$\begin{aligned} \gamma^{\MS}_{m^2}(g) &= -g +\lfrac{5}{6} g^2 -\lfrac{7}{2} g^3 +\left( 3 \mzv{4} +\lfrac{3}{2} \mzv{3} +\lfrac{477}{32} \right) g^4 %\\&\quad -\left( \lfrac{75}{2} \mzv{6} -9 \mzv[2]{3} + \mzv{5} +\lfrac{65}{4} \mzv{4} +\lfrac{1519}{48} \mzv{3} +\lfrac{158849}{2304} \right) g^5 \nonumber\\&\quad +\left( \lfrac{55701}{100} \mzv{8} -288 \mzv{3}\mzv{5} -\lfrac{972}{25} \mzv{3,5} +54 \mzv{3}\mzv{4} -\lfrac{4629}{20} \mzv{7} \right.\nonumber\\&\left.\qquad\ +\lfrac{446}{5} \mzv[2]{3} +\lfrac{1141}{4} \mzv{6} +\lfrac{4019}{40} \mzv{5} +\lfrac{1695}{32} \mzv{4} +\lfrac{472891}{1440} \mzv{3} +\lfrac{7915913}{23040} \right) g^6 +\bigo{g^7} \label{eq:gamma-m2-result}% \\& \approx -g +0.8333 g^2 -3.5 g^3 +19.96 g^4 -150.8 g^5 +1355 g^6 +\bigo{g^7} .\nonumber\end{aligned}$$ Let us summarize various checks of our result. First note that we computed the renormalization group functions for the $O(N)$-symmetric $\field^4$ model with their full $N$-dependence (shown above for $N=1$). Our result agrees with the known $5$-loop results [@ChetyrkinKataevTkachov:5loopPhi4; @ChetyrkinGorishnyLarinTkachov:5loopPhi4; @Kazakov:MethodOfUniqueness; @KNFCL:5loopPhi4] and the 6-loop field anomalous dimension [@BatkovichKompanietsChetyrkin:6loop]. It also confirms the leading and subleading terms in the large $N$ expansion of the $6$-loop beta function obtained almost 20 years ago by John Gracey in [@Gracey:LargeNf]. Additional checks were provided by Dmitrii Batkovich, who calculated $575$ out of the $627$ six-loop $\Gamma_4$-counterterms using the traditional approach, which combines IBP for four-loop massless propagators (known from [@BaikovChetyrkin:FourLoopPropagatorsAlgebraic; @SmirnovTentyukov:FourLoopPropagatorsNumeric; @LeeSmirnov:FourLoopPropagatorsWeightTwelve]) and IRR with $\Rstar$ to compute the diagrams [@BatkovichKompaniets:6loop-Rstar]. The most complicated diagrams (that could not be checked this way) were additionally computed numerically with sector decomposition [@BinothHeinrich:SectorDecomposition], using a custom-made program by the first author, to at least 3 significant digits. Also note that in our first calculation, we resolved subdivergences in a different way than presented in section \[sec:one-scale-scheme\], namely by constructing primitive (i.e. subdivergence-free) linear combinations of graphs like in [@Panzer:MasslessPropagators; @Panzer:PhD]. This technique will be explained in detail elsewhere. The results obtained by both methods agree. Probably the strongest check comes from the completely independent computation of the $6$-loop renormalization group functions of $\field^4$ theory ($N=1$) by Oliver Schnetz [@Schnetz:NumbersAndFunctions]. His computation is very different both conceptually and technically; being carried out in position space with graphical functions [@Schnetz:GraphicalFunctions] and exploiting the powerful theory of generalized single-valued hyperlogarithms [@Schnetz:GeneralizedSV]. Outlook {#sec:outlook} ======= While the primitive $6$-loop graphs had been computed already thirty years ago [@Broadhurst:5loopsbeyond], subdivergences have hitherto been the obstacle for calculations in $\field^4$ theory. In this article, we showed how recently developed techniques (parametric integration and the one-scale BPHZ scheme) can overcome the limitations of traditional methods. This approach can principle also be used at $7$ loops, but the calculation using graphical functions promises to be much more efficient and is already underway [@Schnetz:NumbersAndFunctions]. The contributions from $7$-loop graphs without subdivergences (these are expected to give the most complicated transcendental numbers) have been completed already [@PanzerSchnetz:Phi4Coaction]. An obvious question is to which extent the techniques presented here are will be of use in other theories, like $\field^3$ in six dimensions (only known to four loops [@Gracey:4loopPhi3]) and gauge theories like QCD (known to five loops [@BaikovChetyrkinKuehn:5loopQCD]). One problem is the much higher number of graphs at a given loop order. A further complication might arise from the integrals themselves, because cubic graphs have more edges per loop order than graphs of $\field^4$. This means that the integrations will be more demanding. We also hope that graph-theoretic methods of desingularization, for example the BPHZ-derived approach we described here, might be useful in more general situations when the method [@ManteuffelPanzerSchabinger:QuasiFinite] via IBP is not an option. Both authors wish to express their admiration and sincere thanks towards Kostja Chetyrkin. Apart from his numerous impressive achievements, he is a very helpful and open researcher who shares his knowledge, expertise and experience. He kindly invited the second author to [KIT](https://www.ttp.kit.edu/) in 2014 where he introduced him to the first author and encouraged us to undertake the $6$-loop calculation. We are deeply indebted to Oliver Schnetz for providing us with the result of his own, independent computation in February of this year. At that time, the method presented here was not yet implemented and hence we had computed the majority of diagrams only via the traditional IBP/IRR/$\Rstar$-method (back then, we used parametric integration only for the most complicated diagrams). We were thus anxious to find a second method to evaluate the diagrams to exclude the possibility of errors. The comparison with Oliver’s result convinced us of the correctness, since the only discrepancy was localized in the coefficient of $\mzv{9}$ and could be tracked down to a copy error when we transferred the results for the primitive graphs from [@Schnetz:Census] into our program.[^6] Very valuable checks of plenty of $6$-loop integrals via the IBP/IRR/$\Rstar$-method where provided by Dmitrii Batkovich [@BatkovichKompaniets:6loop-Rstar]. Also we thank John Gracey for providing us with his results [@Gracey:LargeNf] on the large $N$ expansion. This check was performed during this conference, and we like to thank the organizers of Loops & Legs 2016 for the chance to present our result and bringing together so many exceptional scientists in this forum. The participants made it a very stimulating event, fostering scientific exchange. David Broadhurst provided us with a copy of [@Broadhurst:WithoutSubtractions], in which he computes the $\overline{\MS}$-renormalized $5$-loop propagator of $O(N)$-symmetric $\field^4$ theory with ingenuity and extreme economy. It provided additional strong checks of our result, since it amounts to an evaluation of the finite parts of $5$-loop propagator diagrams, which in turn give a subset of the counterterms at $6$ loops. Most figures in this article were created with [@BinosiTheussl:JaxoDraw] and [@Vermaseren:Axodraw]. [^1]: The publication of the and programs used for our calculations is under preparation. [^2]: One may also introduce new (auxiliary) external momenta and it is possible to apply IRR to 2-point diagrams themselves, see the discussion in [@BatkovichKompanietsChetyrkin:6loop]. However, we did not use any of these extensions in our calculation. [^3]: The value of the Feynman integral $\FI(G)$ in does not depend on the choice of $e_0$. It can be interpreted as a projective integral and we can actually replace $\delta(1-\alpha_{e_0})$ by $\delta(1-f(\alpha))$ with an arbitrary function $f(\alpha)$ that is homogeneous of degree one and positive when all $\alpha_e>0$. [^4]: Remarkably, this is essentially just a formula for the antipode in the Hopf algebra of Feynman graphs [@Kreimer:HopfAlgebraQFT]. [^5]: These relations between different $p$-integrals and vacuum integrals are also known as *cut-and-glue* [@BaikovChetyrkin:FourLoopPropagatorsAlgebraic]. [^6]: Since then we implemented the new method described in this article and recomputed all integrals this way, finding perfect agreement. This highly automated program computes all integrals from scratch, with the sole exception of .
{ "pile_set_name": "ArXiv" }
--- abstract: 'The small atmosphereless objects of our solar system, such as asteroids, the moon are covered by layer of dust particles known as regolith, formed by meteoritic impact. The light scattering studies of such dust layer by laboratory experiment and numerical simulation are two important tools to investigate their physical properties. In the present work, the light scattered from a layer of dust particles, containing 0.3$\mu$m $Al_{2}O_{3}$ at wavelength 632.8 nm is analysed. This work has been performed by using a light scattering instrument ’ellipsometer’, at the Department of Physics, Assam Universiy, Silchar, India. Through this experiment, we generated in laboratory the photometric and polarimetric phase curves of light scattered from such a layer. In order to numerically simulate this data, we used Hapke’s model combined with Mie’s single particle scattering properties. The perpendicular and parallel components of single particle albedo and the phase function were derived from Mie theory. By using the Hapke’s model combined with Mie theory, the physical properties of the dust grain such as grain size, optical constant $(n,k)$ and wavelength can be studied through this scheme. In literature, till today no theoretical model to represent polarisation caused due to scattering from rough surface is available, which can successfully explain the scattering process. So the main objective of this work is to develop a model which can theoretically estimate polarisation as caused due to scattering from rough surface and also to validate our model with the laboratory data generated in the present work.' address: 'Department of Physics, Assam University, Silchar, Assam, India, Pin-788011' author: - 'S. Deb' - 'A. K. Sen' bibliography: - 'mybibfile.bib' nocite: '[@*]' title: Laboratory light scattering from regolith surface and simulation of data by Hapke model --- [Regolith]{},[Hapke model]{},[Polarisation]{},[Mie theory]{},[Reflectance]{} Introduction ============ The unpolarised solar radiation scattered by asteroidal regolith surface (a particulate medium) is partially polarised (Hapke 1993). The polarisation phase curve, i.e. the variation of polarisation (P) as a function of phase angle $(\alpha)$ provides the main source of information about the physical properties of regolith, such as grain size, optical constant, albedo, surface roughness and porosity. The polarimetric parameters such as maximum and minimum polarisation; the phase angles corresponding to these extrema; inversion angle and slope at inversion can be retrieved from the polarisation phase curve. In order to account for different branches of polarisation phase curve, semiemperical relationships have been developed between polarisation values (such as maximum polarisation $P_{max}$ , slope at inversion) and albedo of the scattering medium (Umov 1905, KenKnight et al. 1967, Widorn 1967, Zellner 1977). Due to observational constrain, asteroids can only be observed in limited phase angle (Shestopalov 2004). In laboratory simulation, this constrain can be eliminated and observation can be carried out at different phase angles. The laboratory simulation obtained for layer of dust particles with multiple scattering, can be compared with remote sensing data of the moon and asteroids to retrieve the physical properties of regolith present on them. The present simulation work is being reported from the same laboratory, where in past the Hapke bidirectional reflectance model (Hapke 1993) has been widely used for modeling the photometric laboratory data (Deb et al. 2011, Bhattacharjee et al. 2011, Deb et al. 2012). A specific single particle scattering phase function p(g) is needed in Hapke model. The regolith consists of irregular particles, whose single particle phase function is generally unknown. So empirical phase functions such as one term or two term Henyey Greenstein phase function is incorporated in Hapke model. But the empirical phase functions do not explicitly contain the physical parameter, such as grain size and optical constant. In order to directly study the effects of these parameters, an exact theoretical single scattering phase function of these parameters is required. The exact solution of scattering by a smooth and homogeneous sphere is provided in Mie theory (Mie 1908). The solution of Mie theory depends on single particle properties such as grain size, optical constant and also on wavelength of incident radiation. So by incorporating a phase function $p(g)$ from Mie theory in Hapke Model, the effect of physical parameter can be studied. The Hapke model provides good agreement with laboratory data. Deb et al. 2011 used Hapke model (Hapke 1993) in combination with Mie theory to study the bidirectional reflectance of 0.3 $\mu$m alumina at 632.8 nm. The author found the best fit between the experimental data and model curve for absorption coefficient k=0.000009 for Alumina particles, which was in close agreement with the value obtained in Piatek et al. 2004. Deb et al. 2011 also compared the variation of relative intensity with particle size of alumina samples at 13 different diameters ranging from 0.1 to 30.09 $\mu$m (Nelson et al. 2000), with the model curve for absorption coefficient k=0.000009, based on above method. The close fit in above comparison confirmed the correctness of method used by Deb et al. 2011. Deb et al. 2012 used the same method as described in Deb et al. 2011, for fitting the variation of relative intensity with particle size of a given sample ( containing basalt and olivine) (Kaasalainen et al. 2003) and dunite particles (Kamei et al. 2002). For this given sample, Deb et al. 2012 also obtained a satisfactory fit between an empirical relation $I=D^{-a}+b$ ($D$=diameter of dust grain) and relative intensity variation with particle size, with suitable values of the parameters a and b. The authors concluded that the average particle size of actual regolith with known composition can be estimated by using this empirical relation. Bhattacharjee et al. 2011 studied the bidirectional reflectance of 0.3 $\mu$m and 1.0 $\mu$m alumina at two wavelengths 543.5 nm and 632.8 nm. The author used slightly different approach from Deb et al. 2011, 2012. The single particle albedo $w$ and asymmetry parameter $\xi$ were computed from Mie theoy. The asymmetry parameter was used to calculate an empirical phase function, Known as Henyey-Greenstein phase function $p(g)$. The authors theoretically computed the bidirectional reflectance of alumina samples by using $w$ and $p(g)$ in Hapke model (Hapke 1993). For fitting the bidirectional reflectance data with model curve, the absorption coefficient k was used as a free fitting parameter. The authors obtained best fitted values of k as 0.000001 and 0.00001 at 543.5 nm and 632.8 nm respectively. The value k=0.000001 at 543.5 nm was comparable with the value k=7.86$\times10^{-6}$ at 635 nm obtained in Piatek et al. 2004. In addition to the above photometric work, some researchers have also worked on the polarisation resulting from rough surface scattering. Hadamcik et al. 1996,2002,2003 had investigated the polarisation phase curves of regolith analogue of different terrestrial, meteorite and synthetic samples by using $PROGRA^{2}$ experimental setup. The variation of degree of polarisation with phase angle for basalt of different grain size range was obtained by Hadamcik et al. 1996. It was found that the value of maximum polarisation $P_{max}$ increases with grain size and corresponding phase angle $\alpha_{max}$ decreases with grain size. For both fluffy and compact particles, Hadamcik et al. 2002 found that the variation of $P_{max}$ as fuction of albedo follows the Umov law in case of both sifted (just sieved and not pressd) and packed (the surface is pressed with spatula) samples. The authors also obtained the wavelength dependence of $P_{max}$ for samples with two sets of fluffy particles and concluded that this fluffy particles could be good candidate for cometary dust. Hadamcik et al. 2003 found that the polarisation decreases with increase in interaction between the glass beads from single bead to agglomerates of beads and finally to deposited beads, which is due to the increase in multiple scattering. Worms et al. 1996,1999a,1999b had done the polarimetric laboratory simulation of deposited sample by $PROGRA^{2}$ experimental setup. The polarisation phase curve of 8-10 $\mu$m and 11-15 $\mu$m sized deposited boron carbide particles in sifted and packed form, were studied by Worms et al. 1996. It was found that in both size ranges, the polarisation is higher in packed sample than in sifted sample and also polarisation is higher for bigger grains as compared to smaller grains. For deposited samples of basaltic glass with grain sizes 10$\mu$m and 45$\mu$m in sifted and packed form, Worms et al. 1999a found that $P_{max}$ is higher for packed samples than sifted samples. Worms et al. 1999b noticed for $SiC$ and $B_{4}C$ samples, $P_{max}$ increases with increase in grain size, which confirmed the findings of Worms et al. 1996 and Hadamcik et al. 1996. It was also found that for 9 $\mu$m $B_{4}C$, 7 $\mu$m and 13 $\mu$m $SiC$ deposited samples, $P_{max}$ is higher for packed as compared to sifted sample. In Penttila et al. 2003, the authors theoretically analysed the polarimetric data of (9$\pm$1) $\mu m$,(13$\pm$2) $\mu m$ and (88$\pm$5) $\mu m$ boron carbide ($B_{4}C$) generated by $PROGRA^{2}$ microgravity experiment. For theoretical analysise, authors used a new model to describes the shapes of stochastic polyhedra with ray-tracing code and obtained (n,k)=(2,0.04), as the best fitted optical constant for boron carbide. Mcguire et al. 1995 experimentally studied the light scattering by large spherical and irregular particles (approximately 1 cm in size). The authors found that for phase angle between $30^{0}$ and $70^{0}$, the difference (between two orthogonal components of polarisation) $\Delta I(g)$ of scattered radiation from the particles is a proportional to the difference between Fresnels reflection coefficients. These coefficients depend only on refractive index for dielectric material. The authors concluded that this method may be applied to determine average value of refractive index of regolith present on solar system objects. In literature, a large amount of laboratory gererated polarimetric data on scattering from rough surface are available from experiment (Hadamcik et al. 1996,2002,2003; Worms et al. 1996,1999a,1999b). For simulating these data, no appropriate theoretical model which include the physical properties of the constituent particles, is available. In the present work, a model has been developed based on Mie theory (Mie 1908) and Hapke model (Hapke 1993) to describe the polarisation of the light scattered from rough surface, in terms of physical properties of the constituent particles. For verifying the model’s correctness, the model curve is compared with laboratory data generated in the present work by the instrument ellipsometer. ![image](fig1_ellipsometer)\ Instrumental details and sample preparation =========================================== The instrument ’L117 Gaertner ellipsometer’ is a polarimeter used in the present work to simulate the process of light scattering from regolith surface. The phase angle dependence of reflectance and polarisation of scattered ligth can be estimated by this instrument. The ellipsometer with different components is shown in fig.1. The light source in the ellipsometer is a linearly polarised He-Ne laser having wavelength 632.8 nm (red) and beam diameter 15$\mu m$. A photo diode is used as a detector to measure the intensity of scattered light from sample surface. A rotatable glan-thompson prism attached in front of the detector is used as an analyser. In the ellipsometer, samples are placed horizontally on a sample tray. The light source and the detector are rotated in a vertical scattering plane to produce different scattering geometries. The angle of incidence and emergence in ellipsometer can be varied from $35^{o}$ to $70^{o}$ in the steps of $5^{o}$, so that the phase angle coverage is $70^{o}$-$140^{o}$. In the present case, two different sets of scattering geometries are used to produce the phase curves, viz keeping incident angle fixed at i=$45^{o}$ and $60^{o}$, the emergent angle is varied from $35^{o}$ to $70^{o}$; and keeping emergent angle fixed at e=$35^{o}$, $45^{o}$, $50^{o}$, $55^{o}$ and $60^{o}$,the incident angle is varied from $35^{o}$ to $70^{o}$. The powdered material of 0.3$\mu$m sized $Al_{2}O_{3}$ is used as sample to produce regolith surface. The sample is poured gently in a sample tray. The top surface of sample is pressed with a glass spatula to make it smooth. Generation of photometric and polarimetric data =============================================== For the polarimetric measurement, a procedure is followed from the work of Sen et al. 1900 and Clarke 1971. The authors used this method for cometary polarimetric measurements. In the present case, this method is used for laboratory polarimetric measurements. The intensities $I_{1}$, $I_{2}$ and $I_{3}$ of the scattered light from the sample surface at three orientation $0^{0}$,$120^{0}$ and $240^{0}$ of the glan-thompson are recorded at each phase angle. From these intensities, the actual scattered intensity and the polarisation can be calculated by the relations, $$\begin{aligned}  I &=\frac{2}{3}(I_{1}+I_{2}+I_{3})\label{1a}\\[10pt]  P_{observe} &=\frac{2 \sqrt{I_{1}(I_{1}-I_{2})+I_{2}(I_{2}-I_{3})+I_{3}(I_{3}-I_{1})}}{I_{1}+I_{2}+I_{3}}\label{1b}\end{aligned}$$ Since the polarisation $P_{observe}$ values are expressed as ratios of intensities, so no absolute photometric callibration is required for our measurements. Modeling of normalised reflectance and polarisation =================================================== The well known reflectance model by Hapke 1993 is mostly used to interpret the laboratory simulated bidirectional reflectance data. To theoretically estimate the reflectance, we shall follow here an approach as described in our previous work Deb et al. 2011, 2012. The model requires two unknown parameters viz. the opposition surge amplitude $B_{0}$ and opposition angular width $h$, which are important mainly at phase angle smaller than $15^{0}$. Their effect can be neglected at large phase angles. This model also requires another parameter, single particle albedo *w* and a single particle phase function $p(g)$. The values of *w* and $p(g)$ can be obtained by using Mie theory as a function of three physical parameters viz. grain size of sample, optical constant $(n,k)$ of the sample and monochromatic wavelength of incident radiation. With the help of Mie theory, one can introduce *w* and $p(g)$ in Hapke model, and thus the physical properties of the grains can be included for numerical modeling. The formula for bidirectional reflectance as described by Hapke 1993 is reproduced below: $$\label{3} r(i,e,g)=\frac{1}{4\pi}(\frac{\mu_{o}}{\mu+\mu_{o}})\Big\{[1+B(g)]\textit{w}p(g)+\textit{w}[H(\mu_{o})H(\mu)-1]\Big\}$$ where $\mu_{0}=\cos i$ and $\mu=\cos e$, $B(g)=\frac{B_{0}}{1+(\frac{1}{h})tan (\frac{g}{2})}$ $B_{o}$= opposition surge amplitude, $h$= opposition surge width *w* = single-particle scattering albedo, $p(g)$= single particle scattering phase function at phase angle $g$ and in the present study Mie phase function is used, $$\label{4} H(x)=\Big[1-(1-\surd(1-\textit{w}))x\{r_{o}+(1-\frac{1}{2}r_{o}-r_{o}x)\}\ln\frac{1+x}{x}\Big]^{-1} %H(x)=\frac{1+2x}{1+2\gamma x}$$ and $$\label{5} r_{o}=\frac{2}{1+\surd(1-\textit{w})}-1$$ As described in the present work the phase angle is always greater than $15^{0}$, so the the opposition effect can be neglected (Hapke 1993). Using $B(g)$=0 in Equation (\[3\]), we have $$\label{6} r(i,e,g)=\frac{1}{4\pi}(\frac{\mu_{o}}{\mu+\mu_{o}})\Big\{\textit{w}p(g)+\textit{w}[H(\mu_{o})H(\mu)-1]\Big\}$$ From Equation (\[6\]), the perpendicular and parallel components of bidirectional reflectance can be written as, $$\label{7} r_{\bot}(i,e,g)=\frac{1}{4\pi}(\frac{\mu_{o}}{\mu+\mu_{o}})\Big\{[\textit{w}p(g)]_{\bot}+\textit{w}[H(\mu_{o})H(\mu)-1]\Big\}$$ $$\label{8} r_{\parallel}(i,e,g)=\frac{1}{4\pi}(\frac{\mu_{o}}{\mu+\mu_{o}})\Big\{[\textit{w}p(g)]_{\parallel}+\textit{w}[H(\mu_{o})H(\mu)-1]\Big\}$$ where $[{w}p(g)]_\bot$ and $[{w}p(g)]_{\parallel}$ are the radiance scattered once by a particle of the regolith, for incoming radiation with electric vector perpendicular and parallel to the scattering plane respectively. In laboratory experiment, the value of bidirectional reflectance can be calculated from the relation below (Hapke 1993), $$\label{9} r(i,e,g)=\frac{I(i,e,g)}{J}$$ where $I(i,e,g)$ and $J$ are respectively the intensities of scattered and incident radiation. In the laboratory we shall calculate $I(i,e,g)$, by using the Equation (\[1a\]), after recording scattered intensities at three positions of glan-thomson prism (or the analyser). By using proper normalisation (as well be discussed later), we can calculate the value of $r(i,e,g)$ for our simulation work. For polarised incident radiation, with electric vector making an angle $45^{0}$ with plane of reflection (or scattering). the total irradiance $J$ can be divided into two equal parts, each of value $\frac{J}{2}$, which are incident on the sample surface with electric vector perpendicular and parallel to the scattering plane. So we can write, $$\label{10} J_{\bot}=\frac{J}{2}$$ $$\label{11} J_{\parallel}=\frac{J}{2}$$ where $J_{\bot}$ and $J_{\parallel}$ are the components incident irradiance with electric vector perpendicular and parallel to the scattering plane respectively. Substituting values from Equation (\[9\]), (\[10\]) in Equation (\[7\]) and Equation (\[9\]), (\[11\]) in Equation (\[8\]), the perpendicular and parallel components of scattered intensities $I_{\bot}(i,e,g)$ and $I_{\parallel}(i,e,g)$ can be written as follows: $$\label{12} I_{\bot}(i,e,g)=\frac{J}{2}r_{\bot}(i,e,g)$$ $$\label{13} I_{\parallel}(i,e,g)=\frac{J}{2}r_{\parallel}(i,e,g)$$ The linear polarisation of the scattered light from the sample surface can be expressed in terms of the stokes parameter $I$, $Q$, $U$ (Stokes 1952), $$\label{14} P_{model}=\frac{\sqrt{Q^{2}+U^{2}}}{I}$$ where $$I=<E^{2}_{xo}>+<E^{2}_{yo}>$$ $$Q=<E^{2}_{xo}>-<E^{2}_{yo}>$$ $$U=2<E_{xo}E_{yo}cos\delta>$$ $E_{xo}$ and $E_{yo}$ are amplitudes of orthogonal components of light beam; $\delta$ is the phase difference between orthogonal components of light beam and $< >$ indicates the time average. Substituting the values of $I$, $Q$ and $U$ in Equation (\[14\]) and omitting the $< >$ sign for our convenience everywhere, we have $$\label{15} P_{model}=\frac{\sqrt{(E^{2}_{xo}-E^{2}_{yo})^{2}+4E^{2}_{xo}E^{2}_{yo}cos^{2}\delta}}{E^{2}_{xo}+E^{2}_{yo}}$$ The intensities $I_{\bot}$ and $I_{\parallel}$ are actually proportional to the amplitudes $E_{xo}$ and $E_{yo}$ of orthogonal components of light beam and we may write $$\label{16} I_{\bot}=KE^{2}_{xo}$$ and $$\label{17} I_{\parallel}=KE^{2}_{yo}$$ where $K$ is some arbitrary constant. Substituting Equation (\[16\]) and (\[17\]) in Equation (\[15\]), we have $$\label{18} P_{model}=\frac{\sqrt{(I_{\bot}-I_{\parallel})^{2}+4I_{\bot}I_{\parallel}cos^{2}\delta}}{(I_{\bot}+I_{\parallel})}$$ Substituting Equation (\[12\]) and (\[13\]) in Equation (\[18\]), we have $$\label{19} P_{model}=\frac{\sqrt{(r_{\bot}-r_{\parallel})^{2}+4r_{\bot}r_{\parallel}cos^{2}\delta}}{(r_{\bot}+r_{\parallel})}$$ In this modeling, we used the following expressions, which can be proved through some logical steps, $$\label{20} [{w}p(g)]_{\bot}=4{w}_{\bot}p_{\bot}(g)$$ $$\label{21} [{w}p(g)]_{\parallel}=4{w}_{\parallel}p_{\parallel}(g)$$ $$\label{22} {w}={w}_{\bot}+{w}_{\parallel}$$ $$\label{23} p(g)=p_{\bot}(g)+p_{\parallel}(g)$$ where ${w}_{\bot}$ and ${w}_{\parallel}$ are the contribution of total single particle albedo $w$ to $I_{\bot}(i,e,g)$ and $I_{\parallel}(i,e,g)$ respectively, $p_{\bot}(g)$ and $p_{\parallel}(g)$ are components single particle scattering phase function $p(g)$ with electric vector perpendicular and parallel to the scattering plane respectively.\ Substituting Equation (\[20\]) in Equation (\[7\]) and Equation (\[21\]) in Equation (\[8\]), we have $$\label{24} r_{\bot}(i,e,g)=\frac{1}{4\pi}(\frac{\mu_{o}}{\mu+\mu_{o}})\Big\{4\textit{w}_{\bot}p_{\bot}(g)+\textit{w}[H(\mu_{o})H(\mu)-1]\Big\}$$ $$\label{25} r_{\parallel}(i,e,g)=\frac{1}{4\pi}(\frac{\mu_{o}}{\mu+\mu_{o}})\Big\{4\textit{w}_{\parallel}p(g)_{\parallel}+\textit{w}[H(\mu_{o})H(\mu)-1]\Big\}$$ The values of $p_{\bot}(g)$ and $p_{\parallel}(g)$ can be obtained from Mie theory. In appendix, the calculation of ${w}_{\bot}$ and ${w}_{\parallel}$ based on Mie theory is provided, which is done in the present work. The values of $r_{\bot}(i,e,g)$ and $r_{\parallel}(i,e,g)$ can be computed by substituting values of $p_{\bot}(g)$, $p_{\parallel}(g)$, ${w}_{\bot}$ and ${w}_{\parallel}$ from Mie theory and $H(\mu)$,$H(\mu_{0})$ from Equation (\[4\]) in Equation (\[24\]) and (\[25\]) respectively. From Equation (\[24\]) and (\[25\]), the total theoretical bidirectional reflectance $r_{th}(i,e,g)$ can be expressed as: $$\label{26} r_{th}(i,e,g)=\frac{1}{2}\Big[r_{\bot}(i,e,g)+r_{\parallel}(i,e,g)\Big]$$ The above value of $r_{th}(i,e,g)$ as in Equation (\[26\]), can be identified to be same as $r(i,e,g)$ as expressed in Equation (\[3\]). This is a theoretically or model calculated value, which will be divided by theoretically calculated reflectance $r(i=45^{o},e=45^{o})$ for normalisation. Thus the normalised reflectance can be expressed as: $$\label{27} nr_{th}(i,e,g)=\frac{r_{th}(i,e,g)}{r_{th}(45^{o},45^{o},90^{o})}$$ This theoretical value will be compared with the observed value as obtained through laboratory experiment. However, this experimentally observed value should be also normalised in the same way. The experimental value of $r(i,e,g)$ is simply proportional to the intensity value $I$ recorded by our light detector in the laboratory in some arbitrary unit, with the scattering geometry defined by the parameter $(i,e,g)$. The value of $I$ is calculated from $I_{1}$, $I_{2}$ and $I_{3}$ by the relation (\[1a\]) . The observed normalised reflectance of the scattered light, which is normalised at $(i=45^{o}, e=45^{o}, g=90^{o})$ can be expressed as: $$nr_{ob}(i,e,g)=\frac{r_{ob}(i,e,g)}{r_{ob}(45^{o},45^{o},90^{o})}\times\frac{\cos i}{\cos e}$$ $$\label{28} =\frac{I(i,e,g)}{I(45^{o},45^{o},90^{o})}\times\frac{\cos i}{\cos e}$$ where $(\frac{cosi}{cose})$ takes care of the geometric correction (Deb et al. 2011). The present work consist of two simulation works, which are as follows: 1. Photometric simulation : Here we calculate theoretically the normalised reflectance of scattered light from the sample surface, by using equation (\[27\]). For experimental values of normalised reflectance, Equation (\[28\]) is used. We shall compare the two sets of reflectance values. 2. Polarimetric simulation : Here we calculate theoretically the polarisation of the scattered light from sample surface, by the Equation (\[19\]) with an unknown parameter $\delta$, which will be used as a free parameter. The experimental values of polarisation are estimated using Equation (\[1b\]). Further, here it may be noted that the quantity $(cos\delta)$ as mentioned in Equation (\[19\]), depends on the phase difference $\delta$ between the two orthogonal components of scattered radiation. The value of $\delta$ depend on some unknown sample parameters which should include roughness, texture and porosity among others. So we prefer to keep it as a free parameter during model fitting. The modeling routine used for the polarimetric data is a least-squares search over the parameter $\delta$ that minimises $\chi$, which is defined as $$\label{29} \chi=\frac{\sum(P_{observe}-P_{model})^{2}}{N}$$ where $P_{observe}$ and $P_{model}$ are computed from Equation (\[1b\]) and (\[19\]) repsectively, N is the number of observations. Error associated with estimated values of reflectance and polarisation ====================================================================== In our experimental work, the intensity values are measured as a function of controlled geometry. We set the angle of incidence and emergence manually and carefully, by noting values on the graduated scale. After that we measure the intensity $I$, on a scale where we have an uncertainty $\delta I=0.05\times I$. Thus we can assume that, the only uncertainty in our experiment is associated with the measurement of intensity $I$. Thus from Equation (\[28\]), it can be shown that the uncertainty in normalised reflectance can be expressed as: $$\delta nr=\sqrt{2}\frac{\delta I}{I}\times nr$$ Since the nr value can be as large as 1 and $\frac{\delta I}{I}\sim0.05$, we can conclude that the maximum uncertainty in measurement of nr can be $\delta nr=0.07$. The uncertainity in polarimetric measurement can be calculated from the relation given below (Sen et al. 1900): $$\label{2} \delta P \approx \big(\frac{2+P^{2}}{3}\big)^{\frac{1}{2}}(\frac{\delta I}{I})$$ Since $\frac{\delta I}{I}\sim0.05$, as discussed above and the maximum value of $P$ as observed in our experiment is of the order of 10%, we can safely take the upper limit on $\delta P$ as $\delta P< 4.18\%$ ![image](fig2){width="15cm" height="10cm"}\ ![image](fig3){width="15cm" height="10cm"}\ ![image](fig4){width="15cm" height="10cm"}\ ![image](fig5){width="15cm" height="10cm"}\ ![image](fig6){width="15cm" height="10cm"}\ ![image](fig7){width="15cm" height="10cm"}\ ![image](fig8){width="15cm" height="10cm"}\ Result and discussion ===================== The data points in the figure 2-8 are computed from the data generated by Ellipsometer. Each figure consists of two panels. The upper panel represents the variation of normalised reflectance of the scattered light as a function of phase angle for 0.3$\mu$m sized powder of $Al_{2}O_{3}$. The model curve in upper panels is obtained by using Equation (\[27\]). The lower panel represents the variation of percentage of polarisation $P$ of the scattered light as a function of phase angle, which is calculated from same data, under identical conditions. The model curve in lower panel is obtained by using Equation (\[19\]). Here we noted that for reflectance data, no curve fitting algorithm was used. The experimental data and theoretical curve were plotted as they are. However, for polarisation curve, a $\chi^{2}$ minimisation technique, for fitting the data was applied with phase difference $\delta$ as free parameter as discussed below. The refractive index of $Al_{2}O_{3}$ is n=1.766 at 632.8$\mu$m (Gervais1991). In the work of Deb et al. 2011, the bidirectional reflectance data for 0.3$\mu$m $Al_{2}O_{3}$ is studied by combining Hapke model with Mie theory. The best fitted value of absorption coefficient k=0.000009 at 632.8$\mu$m was obtained by Deb et al. 2011, for matching observed data with Hapke model. In the present work, the complex optical constant (n=1.766,k=0.000009) at 632.8$\mu$m for $Al_{2}O_{3}$ (as obtained by Deb et al. 2011)is used to compute $r_{\bot}(i,e,g)$ and $r_{\parallel}(i,e,g)$ from Equation (\[24\]) and (\[25\]). These values of reflectance are used to compute the theoretical normalised reflectance from Equation (\[27\]) to model the experimental normalised reflectance data for different scattering geometries. So no free parameter is used for fitting of the modeling for normaliased reflectance. The values of $r_{\bot}(i,e,g)$ and $r_{\parallel}(i,e,g)$, which were computed previously, are now used in Equation (\[19\]) to theoretically calculate percentage of polarisation. In order to fit the experimental data to this theoretical polarisation curve from the model (as in Equation (\[18\])), the phase difference $\delta$ is used as a free parameter. Towards this fitting procedure, the $\chi^{2}$- minimisation technique was used, as was previously done by Deb et al. 2011, for fitting reflectance data. The best fitted values of $\delta$ for different scattering geometries are lying in the range $[85.5^{0}-90^{0}]$, which are listed in table 1. In both types of modeling viz. for normalised reflectance $nr$ and percentage of polarisation $P$; the grain size, optical constant and monochromatic wavelength of incident radiation are taken as input parameters. Angle of incidence i Angle of emergence e Best fitted phase difference $\delta$ ---------------------- ---------------------- --------------------------------------- $(35^{0}-70^{0})$ $35^{0}$ $86^{0}$ $(35^{0}-70^{0})$ $45^{0}$ $89.5^{0}$ $(35^{0}-70^{0})$ $50^{0}$ $90^{0}$ $(35^{0}-70^{0})$ $55^{0}$ $90^{0}$ $(35^{0}-70^{0})$ $60^{0}$ $85.5^{0}$ $45^{0}$ $(35^{0}-70^{0})$ $87^{0}$ $60^{0}$ $(35^{0}-70^{0})$ $90^{0}$ : Table of best fitted values of free parameter $\delta$[]{data-label="table:1"} The agreement between the normalised reflectance data with the corresponding model curve is obtained for 0.3$\mu$m $Al_{2}O_{3}$, has been shown in upper panels of figure 2-8. This agreement proves the correctness of measured intensities, from which the normalised reflectance is computed by using Equation (\[28\]), in other words, using Hapke’s model combined with Mie Theory. From the same set of intensities, the percentage of polarisation is computed by using Equation (\[1b\]). Whereas, the theoretical values of polarisation has been estimated by using Equation (\[19\]), which has been derived in the present work based on our proposed model. For modeling this polarimetric data, however in literature no suitable procedure is available till today, which can successfully explain the data in terms of the physical properties of constituent grains of rough surface from which the scattering is taking place. For this reason, in the present work a new polarimetric model is developed based on Mie theory and Hapke model, to the explain polarimetric data obtained for 0.3$\mu$m $Al_{2}O_{3}$. The single particle physical properties viz. grain size, complex optical constant of grains and size parameter are introduced in present modeling by Mie theory. The agreement between the percentage of polarisation and corresponding model curve based on new polarimetric model, is obtained by taking phase difference $\delta$ as a free fitting parameter. In lower panels of figure 2-8, this comparison has been shown. This agreement proves the correctness of the model developed in the present work. It should be also noted that, though phase difference ($\delta $) has been used as a free parameter, but for the seven independent data sets the value of $\delta$ has been found to lie within a narrow range viz. $85^{0}- 90^{0}$. This fact further strengthens the consistency of our model. In future, this model can be tested for other samples to know further the range of applicability of our model. Finally, this model can be used to predict the polarization caused due to rough surface scattering. Alternately, form the observed polarization data from rough surface, one would be able to infer about the properties of constituent grains, by using this model. Acknowledgements {#acknowledgements .unnumbered} ================ This work is supported and financed by UGC-SAP grant ( 530/12/DRS/2010 (SAP-1) dtd. 01 Mar 2013). The authors are thankful to Amritaksha Kar, Dept. of Physics, Assam University, Silchar for his immense support and valuable suggestions in performing the laboratory work.The authors are also thankful to Dr. A. Deshmukhya, Dept. of Physics, Assam University, Silchar for her moral support and encouragement during this work. APPENDIX {#appendix .unnumbered} ======== In Mie theory, when an electromagnetic radiation of wavelength $\lambda$ interacts with a spherical particle of radius $a$ with complex refractive index $m$, the total scattering efficiency can be expressed as (Van de Hulst 1981): $$\label{A19} Q_{sca}=\frac{1}{x^{2}}\int_{0}^{\pi}[i_{1}(\theta)+i_{2}(\theta)]\sin\theta d\theta$$ where $\theta$ is the scattering angle, $x=\frac{2\pi a}{\lambda}$ is the size parameter; $i_{1}(\theta)$ and $i_{2}(\theta)$ are intensity functions associated with the electric vector perpendicular and parallel to the scattering plane respectively. From Equation (\[A19\]), the perpendicular and parallel components of total scattering efficiency $Q_{sca}$ can be expressed as: $$\label{A20} Q_{sca\perp}=\frac{1}{x^{2}}\int_{0}^{\pi}i_{1}(\theta)\sin\theta d\theta$$ $$\label{A21} Q_{sca\parallel}=\frac{1}{x^{2}}\int_{0}^{\pi}i_{2}(\theta)\sin\theta d\theta$$ The expressions of intensity functions $i_{1}(\theta)$ and $i_{2}(\theta)$ can be expressed as (Van de Hulst 1981): $$\label{A22} i_{1}(\theta)=|S_{1}(\theta)|^{2}$$ $$\label{A23} i_{2}(\theta)=|S_{2}(\theta)|^{2}$$ where $S_{1}(\theta)$ and $S_{2}(\theta)$ are amplitude functions associated with $i_{1}(\theta)$ and $i_{2}(\theta)$ respectively. The expressions of amplitude functions $S_{1}(\theta)$ and $S_{2}(\theta)$ (Van de Hulst 1981) can be expressed as, $$\label{A24} S_{1}(\theta)=\sum_{n=1}^{\infty}\frac{2n+1}{n(n+1)}[a_{n}\pi_{n}(\cos \theta)+b_{n}\tau_{n}(\cos \theta)]$$ $$\label{A25} S_{2}(\theta)=\sum_{n=1}^{\infty}\frac{2n+1}{n(n+1)}[b_{n}\pi_{n}(\cos \theta)+a_{n}\tau_{n}(\cos \theta)]$$ where $a_{n}$ and $b_{n}$ are function of size parameter $x$ and complex refractive index $m$ of the sphere, $$\label{A26} \pi_{n}(\cos\theta)=\frac{1}{\sin \theta}P^{1}_{n}(\cos \theta)$$ $$\label{A27} \tau_{n}(\cos\theta)=\frac{d}{d\theta}P^{1}_{n}(\cos \theta)$$ and $P^{1}_{n}$ is associated Legendre function of first kind of order 1 and degree n. Substituting Equation (\[A24\]) in Equation (\[A22\]), $i_{1}(\theta)$ can be written as, $$i_{1}(\theta)=|S_{1}(\theta)|^{2}=\sum_{n=1}^{\infty}\frac{2n+1}{n(n+1)}[a_{n}\pi_{n}+b_{n}\tau_{n}]\times\sum_{m=1}^{\infty}\frac{2m+1}{m(m+1)}[a_{m}\pi_{m}+b_{m}\tau_{m}]$$ $$\label{A28} \Rightarrow i_{1}(\theta)=\sum_{n=1}^{\infty}\sum_{m=1}^{\infty}\frac{2n+1}{n(n+1)}\frac{2m+1}{m(m+1)}(a_{n}a_{m}\pi_{n}\pi_{m}+a_{n}b_{m}\pi_{n}\tau_{m}+a_{m}b_{n}\pi_{m}\tau_{n}+b_{n}b_{m}\tau_{n}\tau_{m})$$ Also substituting Equation (\[A25\]) in Equation (\[A23\]), $i_{2}(\theta)$ can be written as, $$i_{2}(\theta)=|S_{2}(\theta)|^{2}=\sum_{i=1}^{\infty}\frac{2i+1}{i(i+1)}[b_{i}\pi_{i}+a_{i}\tau_{i}]\times\sum_{j=1}^{\infty}\frac{2j+1}{j(j+1)}[b_{j}\pi_{j}+a_{j}\tau_{j}]$$ $$\label{A29} \Rightarrow i_{2}(\theta)=\sum_{i=1}^{\infty}\sum_{j=1}^{\infty}\frac{2i+1}{i(i+1)}\frac{2j+1}{j(j+1)}(b_{i}b_{j}\pi_{i}\pi_{j}+b_{i}a_{j}\pi_{i}\tau_{j}+a_{i}b_{j}\tau_{i}\pi_{j}+a_{i}a_{j}\tau_{i}\tau_{j})$$ Now combining Equation (\[A28\]) and Equation (\[A20\]), the perpendicular component of scattering efficiency $Q_{sca\perp}$ can be written as, $$Q_{sca\perp}=\frac{1}{x^{2}}\int_{0}^{\pi}\sum_{n=1}^{\infty}\sum_{m=1}^{\infty}\frac{2n+1}{n(n+1)}\frac{2m+1}{m(m+1)}\Bigg(a_{n}a_{m}\pi_{n}\pi_{m}+a_{n}b_{m}\pi_{n}\tau_{m}+a_{m}b_{n}\pi_{m}\tau_{n}+b_{n}b_{m}\tau_{n}\tau_{m}\Bigg)\sin\theta d\theta$$ $$=\frac{1}{x^{2}}\sum_{n=1}^{\infty}\sum_{m=1}^{\infty}\frac{2n+1}{n(n+1)}\frac{2m+1}{m(m+1)}\Bigg(a_{n}a_{m}\int_{0}^{\pi}\pi_{n}\pi_{m}\sin\theta d\theta+a_{n}b_{m}\int_{0}^{\pi}\pi_{n}\tau_{m}\sin\theta d\theta$$ $$\label{A30} +a_{m}b_{n}\int_{0}^{\pi}\pi_{m}\tau_{n}\sin\theta d\theta+b_{n}b_{m}\int_{0}^{\pi}\tau_{n}\tau_{m}\sin\theta d\theta\Bigg)$$ Also combining Equation (\[A29\]) and Equation (\[A21\]), the parallel component of scattering efficiency $Q_{sca\parallel}$ can be written as, $$Q_{sca\parallel}=\frac{1}{x^{2}}\int_{0}^{\pi}\sum_{i=1}^{\infty}\sum_{j=1}^{\infty}\frac{2i+1}{i(i+1)}\frac{2j+1}{j(j+1)}\Bigg(b_{i}b_{j}\pi_{i}\pi_{j}+b_{i}a_{j}\pi_{i}\tau_{j}+a_{i}b_{j}\tau_{i}\pi_{j}+a_{i}a_{j}\tau_{i}\tau_{j}\Bigg)\sin\theta d\theta$$ $$=\frac{1}{x^{2}}\sum_{i=1}^{\infty}\sum_{j=1}^{\infty}\frac{2i+1}{i(i+1)}\frac{2j+1}{j(j+1)}\Bigg(b_{i}b_{j}\int_{0}^{\pi}\pi_{i}\pi_{j}\sin\theta d\theta+b_{i}a_{j}\int_{0}^{\pi}\pi_{i}\tau_{j}\sin\theta d\theta$$ $$\label{A31} +a_{i}b_{j}\int_{0}^{\pi}\tau_{i}\pi_{j}\sin\theta d\theta+a_{i}a_{j}\int_{0}^{\pi}\tau_{i}\tau_{j}\sin\theta d\theta\Bigg)$$ From Equation (\[A30\]) and (\[A31\]), it is found that the following four integrations have to be performed in order to calculate the values of $Q_{sca\perp}$ and $Q_{sca\parallel}$, $$\label{A32} \int_{0}^{\pi}\pi_{p}\pi_{q}\sin\theta d\theta$$ $$\label{A33} \int_{0}^{\pi}\pi_{p}\tau_{q}\sin\theta d\theta$$ $$\label{A34} \int_{0}^{\pi}\pi_{q}\tau_{p}\sin\theta d\theta$$ $$\label{A35} \int_{0}^{\pi}\tau_{p}\tau_{q}\sin\theta d\theta$$ where the index $p$ represents $n$ (in Equation (\[A30\])) and $i$ (in Equation (\[A31\])). Similarly the index $q$ represents $m$ (in Equation (\[A30\])) and $j$ (in Equation (\[A31\])). Using Equation (\[A26\]) in the expression (\[A32\]), we have $$\int_{0}^{\pi}\pi_{p}\pi_{q}\sin\theta d\theta=\int_{0}^{\pi}\frac{1}{\sin \theta}P^{1}_{p}(\cos \theta)\frac{1}{\sin \theta}P^{1}_{q}(\cos \theta)\sin\theta d\theta$$ $$\label{A1} =\int_{0}^{\pi}\frac{1}{\sin^{2}\theta}P^{1}_{p}(\cos \theta)P^{1}_{q}(\cos \theta)\sin\theta d\theta$$ Let $x=cos\theta$ $\Rightarrow dx=-sin\theta d\theta$ When $\theta=\pi$, $x=-1$ and $\theta=0$, $x=1$ Using the above substitutions in Equation (\[A1\]), we have $$\label{A36} \int_{0}^{\pi}\pi_{p}\pi_{q}\sin\theta d\theta=\int_{-1}^{1}\frac{P^{1}_{p}(x)P^{1}_{q}(x)}{1-x^{2}}dx$$ For $p=q$, Equation (\[A36\]) reduces to, $$\label{A37} \int_{0}^{\pi}\pi^{2}_{p}\sin\theta d\theta=\int_{-1}^{1}\frac{[P^{1}_{p}(x)]^{2}}{1-x^{2}}dx$$ From Riley et al. 2006, we have $$\label{A38} \int_{-1}^{1}\frac{[P^{q}_{p}(x)]^{2}}{1-x^{2}}dx=\frac{(p+q)!}{q(p-q)!}$$ For $q$=1, Equation (\[A38\]) reduces to, $$\int_{-1}^{1}\frac{[P^{1}_{p}(x)]^{2}}{1-x^{2}}dx=\frac{(p+1)!}{(p-1)!}$$ $$=\frac{(p+1)p(p-1)!}{(p-1)!}$$ $$\label{A39} =p(p+1)$$ Using Equation (\[A39\]) in Equation (\[A37\]), we have $$\label{A40} \int_{0}^{\pi}\pi^{2}_{p}\sin\theta d\theta=p(p+1)$$ The relation between $\pi_{p}$ and $\tau_{q}$, from Bohren and Huffman 1983 is reproduced below, $$\label{A41} \int_{0}^{\pi}(\pi_{p}\pi_{q}+\tau_{p}\tau_{q})\sin{\theta} d\theta=\delta_{pq}\frac{2p^{2}(p+1)^{2}}{2p+1}$$ where $\delta_{pq}$= kronecker delta And also we have, $$\int_{0}^{\pi}(\pi_{p}\tau_{q}+\pi_{q}\tau_{p})\sin{\theta} d\theta=\int_{0}^{\pi}(\frac{1}{\sin\theta}P^{1}_{p}\frac{dP^{1}_{q}}{d\theta}+\frac{1}{\sin\theta}P^{1}_{q}\frac{dP^{1}_{p}}{d\theta})\sin\theta d\theta$$ $$=\int_{0}^{\pi}(P^{1}_{p}\frac{dP^{1}_{q}}{d\theta}+P^{1}_{q}\frac{dP^{1}_{p}}{d\theta}) d\theta$$ $$=\int_{0}^{\pi}\frac{d}{d\theta}(P^{1}_{p}P^{1}_{q})d\theta$$ $$=[P^{1}_{p}P^{1}_{q}]^{\pi}_{0}$$ $$=0$$ $$\label{A42} \Rightarrow\int_{0}^{\pi}\pi_{p}\tau_{q}\sin{\theta} d\theta=-\int_{0}^{\pi}\pi_{q}\tau_{p}\sin{\theta} d\theta$$ For $p=q$, the Equation (\[A41\]) reduces to, $$\int_{0}^{\pi}(\pi^{2}_{p}+\tau^{2}_{p})\sin{\theta} d\theta=\frac{2p^{2}(p+1)}{2p+1}$$ $$\label{A43} \Rightarrow\int_{0}^{\pi}\tau^{2}_{p}\sin{\theta} d\theta=\frac{2p^{2}(p+1)^{2}}{2p+1}-\int_{0}^{\pi}\pi^{2}_{p}\sin{\theta} d\theta$$ Combining Equation (\[A40\]) and Equation (\[A43\]), we have $$\int_{0}^{\pi}\tau^{2}_{p}\sin{\theta} d\theta=\frac{2p^{2}(p+1)^{2}}{2p+1}-p(p+1)$$ $$=p(p+1)[\frac{2p(p+1)-2p-1}{2p+1}]$$ $$=p(p+1)[\frac{2p^{2}+2p-2p-1}{2p+1}]$$ $$\label{A44} \Rightarrow\int_{0}^{\pi}\tau^{2}_{p}\sin{\theta} d\theta=\frac{p(p+1)(2p^{2}-1)}{2p+1}$$ Combining Equation (\[A42\]) and Equation (\[A30\]), we have $$Q_{sca\perp}=\frac{1}{x^{2}}\sum_{n=1}^{\infty}\sum_{m=1}^{\infty}\frac{2n+1}{n(n+1)}\frac{2m+1}{m(m+1)} \Bigg(a_{n}a_{m}\int_{0}^{\pi}\pi_{n}\pi_{m}\sin\theta d\theta+a_{n}b_{m}\int_{0}^{\pi}\pi_{n}\tau_{m}\sin\theta d\theta$$ $$-a_{m}b_{n}\int_{0}^{\pi}\pi_{n}\tau_{m}\sin\theta d\theta+b_{n}b_{m}\int_{0}^{\pi}\tau_{n}\tau_{m}\sin\theta d\theta\Bigg)$$ $$\Rightarrow Q_{sca\perp}=\frac{1}{x^{2}}\sum_{n=1}^{\infty}\sum_{m=1}^{\infty}\frac{2n+1}{n(n+1)}\frac{2m+1}{m(m+1)}\Bigg[a_{n}a_{m}\int_{0}^{\pi}\pi_{n}\pi_{m}\sin\theta d\theta+(a_{n}b_{m}-a_{m}b_{n})\int_{0}^{\pi}\pi_{n}\tau_{m}\sin\theta d\theta$$ $$\label{A45} +b_{n}b_{m}\int_{0}^{\pi}\tau_{n}\tau_{m}\sin\theta d\theta\Bigg]$$ For $n=m$, Equation (\[A45\]) reduces to, $$\label{A46} Q_{sca\perp}=\frac{1}{x^{2}}\sum_{n=1}^{\infty}\Bigg[\frac{2n+1}{n(n+1)}\Bigg]^{2}\Bigg[\mid a_{n}\mid^{2}\int_{0}^{\pi}\pi^{2}_{n}\sin\theta d\theta+\mid b_{n}\mid^{2}\int_{0}^{\tau}\tau^{2}_{n}\sin\theta d\theta\Bigg]$$ Combing Equation (\[A42\]) and Equation (\[A31\]), we have $$Q_{sca\parallel}=\frac{1}{x^{2}}\sum_{i=1}^{\infty}\sum_{j=1}^{\infty}\frac{2i+1}{i(i+1)}\frac{2j+1}{j(j+1)}\Bigg(b_{i}b_{j}\int_{0}^{\pi}\pi_{i}\pi_{j}\sin\theta d\theta-b_{i}a_{j}\int_{0}^{\pi}\tau_{i}\pi_{j}\sin\theta d\theta$$ $$+a_{i}b_{j}\int_{0}^{\pi}\tau_{i}\pi_{j}\sin\theta d\theta+a_{i}a_{j}\int_{0}^{\pi}\tau_{i}\tau_{j}\sin\theta d\theta\Bigg)$$ $$\label{A47} \begin{split} \Rightarrow Q_{sca\parallel}=\frac{1}{x^{2}}\sum_{i=1}^{\infty}\sum_{j=1}^{\infty}\frac{2i+1}{i(i+1)}\frac{2j+1}{j(j+1)}\Bigg[b_{i}b_{j}\int_{0}^{\pi}\pi_{i}\pi_{j}\sin\theta d\theta \\ +(a_{i}b_{j}-b_{i}a_{j})\int_{0}^{\pi}\tau_{i}\pi_{j}\sin\theta d\theta+a_{i}a_{j}\int_{0}^{\pi}\tau_{i}\tau_{j}\sin\theta d\theta\Bigg] \end{split}$$ For $i=j$, Equation (\[A47\]) reduces to, $$\label{A48} Q_{sca\parallel}=\frac{1}{x^{2}}\sum_{i=1}^{\infty}\Bigg[\frac{2i+1}{i(i+1)}\Bigg]^{2}\Bigg[\mid b_{i}\mid^{2}\int_{0}^{\pi}\pi^{2}_{i}\sin\theta d\theta+\mid a_{i}\mid^{2}\int_{0}^{\pi}\tau^{2}_{i}\sin\theta d\theta\Bigg]$$ Using the values of integration from Equation (\[A40\]) and Equation (\[A44\]) in Equation (\[A46\]), we have $$Q_{sca\perp}=\frac{1}{x^{2}}\sum_{n=1}^{\infty}\Bigg[\frac{2n+1}{n(n+1)}\Bigg]^{2}\Bigg[n(n+1)\mid a_{n}\mid^{2}+\frac{n(n+1)(2n^{2}-1)}{2n+1}\mid b_{n}\mid^{2}\Bigg]$$ $$\label{A49} =\frac{1}{x^{2}}\sum_{n=1}^{\infty}\Bigg[\frac{(2n+1)^{2}}{n(n+1)}\mid a_{n}\mid^{2}+\frac{(2n+1)(2n^{2}-1)}{n(n+1)}\mid b_{n}\mid^{2}\Bigg]$$ Also the using values of integration from Equation (\[A40\]) and Equation (\[A44\]) in Equation (\[A48\]), we have $$Q_{sca\parallel}=\frac{1}{x^{2}}\sum_{i=1}^{\infty}\Bigg[\frac{2i+1}{i(i+1)}\Bigg]^{2}\Bigg[i(i+1)\mid b_{i}\mid^{2}+\frac{i(i+1)(2i^{2}-1)}{(2i+1)}\mid a_{i}\mid^{2}\Bigg]$$ $$\label{A50} =\frac{1}{x^{2}}\sum_{i=1}^{\infty}\Bigg[\frac{(2i+1)^{2}}{i(i+1)}\mid b_{i}\mid^{2}+\frac{(2i+1)(2i^{2}-1)}{i(i+1)}\mid a_{i}\mid^{2}\Bigg]$$ The total extinction efficiency (Van de Hulst 1981) can be expressed as, $$\label{A51} Q_{ext}=\frac{4}{x^2}Re\{S(0)\}$$ where $$S_{1}(0)=S_{2}(0)=S(0)=\frac{1}{2}\sum_{n=1}^{\infty}(2n+1)(a_{n}+b_{n})$$ For $\theta=0^{\circ}$, the perpendicular component $S_{1}(0)$ and parallel component $S_{2}(0)$ of amplitude function are equal, so the perpendicular component $Q_{ext\perp}$ and parallel component $Q_{ext\parallel}$ of total extinction efficiency $Q_{ext}$ are also equal. So from Equation (\[A51\]), the expressions for $Q_{ext\perp}$ and $Q_{ext\parallel}$ can be expressed as follows : $$\label{A52} Q_{ext\perp}=Q_{ext\parallel}=\frac{4}{x^2}Re\{S(0)\}=\frac{1}{2}\sum_{n=1}^{\infty}(2n+1)(a_{n}+b_{n})$$ The single particle albedo (Van de Hulst 1981, Hapke 1993) can be expressed as : $$\label{A53} w=\frac{Q_{sca}}{Q_{ext}}$$ From Equation (\[A53\]), the perpendicular and parallel components of single particle albedo can be defined as, $$\label{A54} w_{\perp}=\frac{Q_{sca\perp}}{Q_{ext\perp}}$$ $$\label{A55} w_{\parallel}=\frac{Q_{sca\parallel}}{Q_{ext\parallel}}$$ By substituting Equations (\[A49\]), (\[A50\]) and (\[A52\]) in Equations (\[A54\]) and (\[A55\]), $w_{\perp}$ and $w_{\parallel}$ can be now computed. References {#references .unnumbered} ==========
{ "pile_set_name": "ArXiv" }
--- abstract: 'The nucleation and propagation of hydrofractures by injection of over pressured fluids in an elastic and isotropic medium are studied experimentally. Non-Newtonian fluids are injected inside a gelatine whose mechanical properties are assumed isotropic at the experimental strain rates. Linear elastic theory predicts that plastic deformation associated to breakage of gelatin bonds is limited to a small zone ahead of the tip of the propagating fracture and that propagation will be maintained while the fluid pressure exceeds the normal stress to the fracture walls (Chávez-Álvarez,2008) (i.e., the minimum compressive stress), resulting in a single mode I fracture geometry. However, we observed the propagation of fractures type II and III as well as nucleation of secondary fractures, with oblique to perpendicular trajectories with respect to the initial fracture. In the [Video](http://ecommons.library.cornell.edu/handle/1813/14122) experimental evidence shows that the fracture shape depends on the viscoelastic properties of gelatine coupled with the strain rate achieved by fracture propagation.' author: - | Mariano Cerca, Jazmin Chavez Alvarez, Bernardino Barrientos,\ Enrique Soto and Carlos Mares\ \ Universidad Nacional Autónoma de México,\ Blvd Juriquilla 3001, Juriquilla, Querétaro, 76230, México\ Centro de investigaciones en Óptica,\ Loma del Bosque \#115, León, 37150, México. title: 'Geometry of elastic hydrofracturing by injection of an over pressured non-Newtonian Fluid' --- 1. M.J. Chávez Álvarez, L.M. Cerca Martínez, *Analogue simulation of magama rheology during dike emplacement: A preliminary study based on field observations and rheological determinations of materials, Bolletino di Geofisica,**49**, 29-34 (2008).*
{ "pile_set_name": "ArXiv" }
--- abstract: | As exemplified by the Kuramoto model, large systems of coupled oscillators may undergo a transition to phase coherence with increasing coupling strength. It is shown that below the critical coupling strength for this transition such systems may be expected to exhibit ‘echo’ phenomena: a stimulation by two successive pulses separated by a time interval $\tau $ leads to the spontaneous formation of response pulses at a time $\tau$, $2\tau $, $3\tau \ldots$, after the second stimulus pulse. Analysis of this phenomenon, as well as illustrative numerical experiments, are presented. The theoretical significance and potential uses of echoes in such systems are discussed. author: - 'Edward Ott, John H. Platig, Thomas M. Antonsen and Michelle Girvan' title: Echo Phenomena in Large Systems of Coupled Oscillators --- [**Large systems consisting of many coupled oscillators for which the individual natural oscillator frequencies are different naturally occur in a wide variety of interesting applications. As shown by Kuramoto, such systems can undergo a type of dynamical phase transition such that as the coupling strength is raised past a critical value, global synchronous collective behavior results. In this paper we show that another interesting, potentially useful, behavior of these systems also occurs [*below*]{} the critical coupling strength. Namely, we demonstrate that these systems exhibit [*echo*]{} phenomena: If a stimulus pulse is applied at time $t=0$, followed by a second stimulus pulse at time $t=\tau $, then pulse echo responses can appear at $t=2\tau ,3\tau ,\ldots$. This phenomenon depends on both nonlinearity and memory inherent in the oscillator system, the latter being a consequence of the continuous spectrum of the linearized system.**]{} I. Introduction =============== Due to their occurrence in a wide variety of circumstances, systems consisting of a large number of coupled oscillators with different natural oscillation frequencies have been the subject of much scientific interest[@pikovsky01; @strogatz04]. Examples where the study of such systems is thought to be relevant are synchronous flashing of fireflies[@buck88] and chirping of crickets[@walker69], synchronous cardiac pacemaker cells[@michaels87], brain function[@singer93], coordination of oscillatory neurons governing circadian rhythms in mammals[@yamaguchi02], entrainment of coupled oscillatory chemically reacting cells[@kiss05], Josephson junction circuit arrays[@wiesenfeld95], etc. The globally-coupled, phase-oscillator model of Kuramoto[@kuramoto84; @acebron05] exemplifies the key generic feature of large systems of coupled oscillators. In particular, Kuramoto considered the case where the distribution function of oscillator frequencies was monotonically decreasing away from its peak value, and he showed that, as the coupling strength $K$ between the oscillators is increased through a critical coupling strength $K_c$, there is a transition to sustained global cooperative behavior. In this state $(K>K_c)$ a suitable average over the oscillator population (this average is often called the ‘order parameter’) exhibits steady macroscopic oscillatory behavior. For $K<K_c$ a stimulus may transiently induce macroscopic oscillations, but the amplitude of these coherent oscillations (i.e., the magnitude of the order parameter) decays exponentially to zero with increasing time[@acebron05]. In the present paper we consider the Kuramoto model in the parameter range $K<K_c$, and we demonstrate that ‘echo’ phenomena occur for this system. The basic echo phenomenon can be described as follows: A first stimulus is applied at time $t=0$, and the response to it dies away; next, a second stimulus is applied at a later time, $t=\tau $, and its response likewise dies away; then at time $t=2\tau $ (also possibly at $n\tau $, for $n=3,4,\ldots$) an echo response spontaneously builds up and then decays away. An illustrative example is shown in Fig. 1, which was obtained by ![Illustration of the echo phenomenon. Stimuli at times $t=0$ and $t=\tau $ lead to direct system responses which rapidly decay away followed by echo responses that can arise at times $2\tau $, $3\tau , \ldots $. The ‘response’ plotted on the vertical axis is the magnitude of the complex valued order parameter, Eq. (8). See Sec. IV for details of this computation.](fig1) numerical simulation (see Sec. IV for details). In order for this phenomenon to occur, the system must have two fundamental attributes, nonlinearity and memory. Nonlinearity is necessary because the response seen in Fig. 1 is not the same as the sum of the responses to each of the individual stimulus pulses in the absence of the other pulse (which is simply the decay that occurs immediately after the individual stimuli, without the echo). Memory is necessary in the sense that the system state after the decay of the second pulse must somehow encode knowledge of the previous history even though the global average of the system state, as represented by the order parameter, is approximately the same as before the two pulses were applied. Echo phenomena of this type, occurring in systems of many oscillators having a spread in their natural oscillation frequencies, have been known for a long time. The first example was the ‘spin echo’ discovered in 1950 by Hahn[@hahn50], where the distribution of frequencies resulted from the position dependence of the precession frequency of nuclear magnetic dipoles in an inhomogeneous magnetic field. \[The spin echo forms the basis for modern magnetic resonance imaging (MRI).\] Subsequently, echoes for cyclotron orbits of charged particles in a magnetic field have been studied for the cases in which the distribution in frequency was due to magnetic field inhomogeneity[@gould65], relativistic dependence of the particle mass on its energy[@ott70], and Doppler shifts of the cyclotron frequency[@porkolab68]. Another notable case is that of plasma waves, where the frequency distribution results from the Doppler shift of the wave frequency felt by charged particles with different streaming velocities[@oneil68]. Although echo phenomena are well-known in the above settings, they have so far not received attention in the context of the Kuramoto model and its many related situations. It is our purpose in the present paper to investigate that problem. Two possible motivations for our study of echoes in the Kuramoto model are that they provide increased basic understanding of the model and also that they may be of potential use as a basis for future diagnostic measurements of related systems (see Sec. V). In what follows, Sec. II will give a formulation of the model problem that will be analyzed in Sec. III and numerically simulated in Sec. IV, while Sec. V will provide a discussion of the implications of the results obtained. II. Formulation =============== We consider the basic Kuramoto model supplemented by the addition of a $\delta $-correlated noise term $n(t)$ and two impulsive stimuli, one at time $t=0$, and the other at time $t=\tau $, $$d\theta _i/dt=\omega _i+K/N\sum ^N_{j=1}\sin (\theta _j-\theta _i)-h(\theta _i)\Delta (t)+n(t) \ ,$$ $$\Delta (t)=\hat d_0\delta (t)+\hat d_1\delta (t-\tau ) \ ,$$ $$\langle n(t)n(t')\rangle =2\xi \delta (t-t') \ ,$$ $$h(\theta )=\sum _nh_ne^{in\theta } \ , \ \ h_n=h^*_{-n} \ , \ \ h_0=0 \ ,$$ where $h^*_{-n}$ denotes the complex conjugate of $h_{-n}$. In the above $\theta _i(t)$ represents the angular phase of oscillator $i$, where $i=1,2,\ldots ,N\gg 1$; and $\omega _i$ is the natural frequency of oscillator $i$ where we take $\omega _i$ for different oscillators (i.e., different $i$) to be distributed according to some given, time-independent distribution function $g(\omega )$, where $g(\omega )$ has an average frequency $\bar \omega =\int \omega g(\omega )d\omega $, is symmetric about $\omega =\bar \omega $, and monotonically decreases as $|\omega -\bar \omega |$ increases. To motivate the impulsive stimuli term, consider the example of a population of many fireflies, and imagine that the stimuli at $t=0$ and at $t=\tau $ are external flashes of light at those times, where the constants $\hat d_0$ and $\hat d_1$ in Eq. (2) represent the intensity of these flashes. We hypothesize that a firefly will be induced by a stimulus flash to move its flashing phase toward synchronism with the stimulus flash. Thus a firefly that has just recently flashed will reset its phase by retarding it, while a firefly that was close to flashing will advance its phase. The amount of advance or retardation is determined by the ‘reset function’, $h(\theta )$. Since the reset function $h(\theta )$ depends on properties of the fireflies, we do not specify it further. Let $\theta ^+_i$ and $\theta ^-_i$ represent the phases of oscillator $i$ just after and just before a stimulus flash at $t=0$ or $t=\tau$. Then we have from Eq. (1) that $$\int ^{\theta ^{+}_{i}}_{\theta _i^-} \frac{d\theta }{h(\theta )}=\hat d_p ; \ \ p=0,1 \ .$$ Letting $F(\theta )=\int ^\theta d\theta /h(\theta )$, we obtain $$\theta ^+_i=F^{-1}(\hat d_p+F(\theta ^-_i)) \ .$$ In our subsequent analysis in Sec. III, we will for convenience assume that $\hat d_p$ is small, in which case $(\theta ^+_i-\theta ^-_i)$ is small, and we can use the approximation, $$\theta _i^+\cong \theta ^-_i+\hat d_ph(\theta ^-_i); \ p=0,1 \ .$$ Following Kuramoto we introduce the complex valued order parameter $R(t)$, $$R(t)=\frac{1}{N} \sum ^N_{j=1} e^{i\theta _j(t)} \ ,$$ in terms of which Eq. (1) can be rewritten as $$d\theta _i/dt=\omega _i+(K/N)Im[e^{-i\theta _i}R(t)]-h(\theta _i)\Delta (t)+n(t) \ .$$ In our analysis in Sec. III we will take the limit $N\rightarrow \infty$ useful for approximating the situation where $N\gg 1$. In that limit it is appropriate to describe the system state by a continuous distribution function $f(\theta ,\omega ,t)$, where $$\int ^{2\pi }_0f(\theta ,\omega ,t)\frac{d\theta}{2\pi }=1 \ ,$$ and the fraction of oscillators with angles and natural frequencies in the ranges $(\theta , \theta +d\theta )$ and $(\omega ,\omega +d\omega )$ is $f(\theta ,\omega ,t)g(\omega )d\omega d\theta /2\pi $. The conservation of the number of oscillators then gives the time evolution equation for $f(\theta ,\omega ,t)$, $$\frac{\partial f}{\partial t}+\frac{\partial}{\partial \theta }\left\{ f\left[ \omega +KIm(R(t)e^{-i\theta })-h(\theta )\Delta (t)\right]\right\}=\xi \frac{\partial ^2f}{\partial \theta ^2} \ ,$$ $$R^*(t)=\int d\omega f_1(\omega ,t)g(\omega ) \ ,$$ where $R^*$ denotes the complex conjugate of $R$, $f(\omega ,\theta ,t)\equiv 1$ for $t<0$, and, in writing Eq. (12), $f_1$ represents the $e^{i\theta }$ component of the Fourier expansion of $f(\omega ,\theta ,t)$ in $\theta $, $$f(\omega ,\theta ,t)=\sum ^{+\infty}_{n=-\infty }f_n(\omega ,t)e^{in\theta } \ ,$$ with $f_0=1$, $f_n=f^*_{-n}$. As seen in Eq. (11), the effect of the noise term in Eq. (1) is to introduce diffusion in the phase angle $\theta $ whose strength is characterized by the phase diffusion coefficient $\xi $. In Sec. III we will solve Eqs. (11) and (12) for the case $d_p\ll 1$, thus demonstrating the echo phenomenon as described in Sec. I. In Sec. IV we will present numerical solutions of Eq. (1) for large $N$. III. Analysis ============= A. Amplitude expansion ---------------------- In order to proceed analytically we use a small amplitude expansion and obtain results to second order (i.e., up to quadratic in the small amplitude). This will be sufficient to obtain the echo phenomenon. We introduce a formal expansion parameter $\epsilon $, as follows, $$f=1+\epsilon f^{(1)}+\epsilon ^2f^{(2)}+\mathcal{O}(\epsilon ^3) \ ;$$ $\hat d_p=\epsilon d_p $ for $p=0$, $1$; $R=\epsilon R^{(1)}+\epsilon ^2R^{(2)}+\mathcal{O}(\epsilon ^3)$; $R^{(m)*}=\int gf_1^{(m)}d\omega $; where $f^{(m)}=\Sigma _nf_n^{(m)}\exp (in\theta )$. (Although we formally take $\epsilon \ll 1$, when we finally get our answers, the results will apply for $\epsilon =1$ and $d_p=\hat d_p$, if $\hat d_p\ll 1$.) B. Order $\epsilon $ -------------------- In linear order (i.e., $\mathcal{O}(\epsilon))$, by multiplying Eq. (11) by $\exp (-i\theta )d\theta $ and integrating over $\theta $, we have for the component of $f^{(1)}$ varying as $e^{i\theta }$, $$\frac{\partial f_1^{(1)}}{\partial t}+(i\omega +\xi )f_1^{(1)}=\frac{K}{2} R^{(1)*}+ih_1\Delta (t) \ , \ \ R^{(1)*}(t)=\int f^{(1)}gd\omega \ ,$$ where $f_1^{(1)}(\omega ,t)=0$ for $t<0$ and $R^{(1)*}$ is the complex conjugate of $R^{(1)}$. Due to the delta function term on the right hand side of Eq. (15), $ih_1d_0\delta (t)$, at the instant just after the first delta function (denoted $t=0^+$), $f_1^{(1)}$ jumps from zero just before the delta function (denoted $t=0^-$) to the value $f_1^{(1)}(\omega ,0^+)=ih_1d_0$. Making use of this observation, in Appendix I we solve Eq. (15) for $0<t<\tau $, with the result that, for $K<K_c$, $$f_1^{(1)}(\omega ,t)=A(\omega )e^{-(i\omega +\xi )t}+\ \ ({\rm a\ more\ rapidly\ exponentially\ decaying\ component)} \ ,$$ where $$A(\omega )=ih_1d_0/D[-(i\omega +\xi )] \ ,$$ $$D(s)=1-\frac{K}{2}\int ^{+\infty}_{-\infty} \frac{g(\omega )d\omega }{s+\xi +i\omega } \ , \ {\rm for} \ Re(s)>0 \ ,$$ and $D(s)$ for $Re(s)\leq 0$ is defined from Eq.(18) by analytic continuation. Since Eq. (16) applies for $0<t<\tau $, we have that just before the application of the second delta function stimulus $(t=\tau ^-)$, $$f_1^{(1)}(\omega ,\tau ^-)\cong A(\omega )e^{-(i\omega +\xi )\tau } \ ,$$ where we have neglected the second term on the right hand side of Eq. (16) on the basis that, due to its more rapid exponential decay, it is small compared to the first term. Solutions of $D(s)=0$ govern the stability of the state with $R^{(1)}=0$. Let $s=s_0$ denote the solutions of $D(s)=0$ with the largest real part. If $Re(s_0)<0$ the state $R^{(1)}=0$ is stable, and a perturbation away from $R^{(1)}=0$ decays to zero with increasing $t$ at the exponential rate $Re(s_0)$. If $Re(s_0)>0$, then the perturbation grows and $R^{(1)}$ eventually saturates into a sustained nonlinear state of coherent cooperative oscillatory behavior[@kuramoto84; @acebron05]. In general, $Re(s_0)$ is an increasing function of the coupling constant $K$, and $Re(s_0)\stackrel{>}{<}0$ for $K\stackrel{>}{<}K_c$, where $K_c$ is a critical value that depends on $\xi$ and $g(\omega )$. Throughout this paper we shall be considering only the case $K<K_c$ for which $Re(s_0)<0$. It is instructive to consider $\xi =0$. In that case, the first term in Eq. (16) is of constant magnitude in time, but, as time $t$ increases, it oscillates more and more rapidly as a function of $\omega $. Because of this increasingly rapid variation in $\omega $, the contribution of this term to $R^{(1)*}(t)=\int gf_1^{(1)}d\omega $ decays in time (see Appendix I), and it does so at the same time-asymptotic rate as the contribution from the second more rapidly exponentially decaying contribution in Eq. (16). Thus the order parameter magnitude decays away, but the distribution function $f_1^{(1)}$ can still have a component (the first term in Eq. (16)) due to the pulse that has not decayed away. A similar conclusion applies for $\xi >0$ provided that $\xi $ is substantially less than the damping for the second term in Eq. (16). This is the source of the ‘memory’ referred to in Sec. I. It is also worth noting that the first term in Eq. (16) can be thought of as the manifestation of the continuous spectrum of the Kuramoto problem, discussed in detail in Ref.  [@strogatz92]. Thus the echo phenomenon that we derive subsequently can be regarded as an observable macroscopic consequence of the continuous spectrum, where by ‘macroscopic’ we mean that the effect can be seen through monitoring of the order parameter without the necessity of other more detailed knowledge of the distribution function. It is also of interest to consider $f_n^{(1)}$ for $n\geq 2$. From Eq. (11) we obtain for $|n|\geq 2$ $$\frac{\partial f_n^{(1)}}{\partial t}+(in\omega +n^2\xi )f_n^{(1)}=inh_n\Delta (t) \ ,$$ which does not have any contribution from the order parameter, $R$. For $\tau >t>0$, Eq. (20) yields $$f_n^{(1)}(\omega ,t)=inh_nd_0\exp [-(in\omega +n^2\xi )t]\ ,$$ for $0 < t < \tau $, which, similar to the first term on the right hand side of Eq. (16), also oscillates increasingly more rapidly with $\omega $ as $t$ increases. At time $t=\tau ^-$ Eq. (21) yields $$f_n^{(1)}(\omega ,\tau ^-)=inh_nd_0\exp [-(in\omega +n^2\xi )\tau ] \ ,$$ for $|n|\geq 2$. C. Order $\epsilon ^2$ ---------------------- Now proceeding to $\mathcal{O}(\epsilon ^2)$ and again (as done in obtaining Eq. (15)) taking the $e^{i\theta }$ component of Eq. (11), we have $$\frac{\partial f_1^{(2)}}{\partial t}+(i\omega +\xi )f_1^{(2)}-\frac{1}{2}KR^{(2)*}=-i\left\{ \frac{K}{2i}f_2^{(1)}R^{(1)}-\Delta (t)\sum ^{+\infty}_{n=-\infty}h_{-(n-1)}f^{(1)}_n\right\}$$ where $R^{(1,2)*}(t)=\int ^{+\infty}_{-\infty}g(\omega )f_1^{(1,2)}(\omega ,t)d\omega $. The above equation is linear in $f_1^{(2)}$ and is driven by several inhomogeneous terms appearing on the right hand side of Eq. (23) that are quadratic in first order quantities. Since we are interested in the components of $f_1^{(2)}$ that result in echoes, and since, by our previous discussion, we expect that the echoes depend on the presence of [*both*]{} stimulus delta functions (i.e., the delta function $\delta (t)$ of strength $d_0$ and the delta function $\delta (t-\tau )$ of strength $d_1$), we are interested in the component of $f_1^{(2)}$ that is proportional to the product $d_0d_1$ for $t>\tau $. We denote this component $f^{(2)}_{1,e}$, where the subscript $e$ stands for ‘echo’. From Eq. (23) we see that for $t>\tau $, the $f^{(2)}_{1,e}$ component of $f_1^{(2)}$ satisfied the following initial value problem $$\frac{\partial f_{1,e}^{(2)}}{\partial t}+(i\omega +\xi )f^{(2)}_{1,e}-\frac{1}{2}KR_e^{(2)*}=0 \ ,$$ $$f^{(2)}_{1,e}(\omega ,\tau ^+)=id_1\sum ^{+\infty}_{n=-\infty}h_{-(n-1)}f^{(1)}_n(\omega ,\tau ^-) \ ,$$ $$R_e^{(2)*}(t)=\int ^{+\infty}_{-\infty} g(\omega )f^{(2)}_{1,e}(\omega ,t)d\omega \ .$$ Since $f^{(1)}_n(\omega ,\tau^- )$ is proportional to $d_0$ (see Eqs. (19) and (22)), we see that the solution of Eqs. (24)–(26) for $f_{1,e}^{(2)}$ and $R_e^{(2)}$ will indeed be proportional to $d_0d_1$ as desired. We solve Eqs. (24)–(26) by taking Laplace transforms, $$\hat f^{(2)}_{1,e}(\omega ,s)=\int ^\infty _\tau e^{-st}f_{1,e}^{(2)}(\omega ,t)dt \ ,$$ $$\hat R^{(2)}_{e*}(s)\equiv \int ^\infty _\tau e^{-st}R_e^{(2)*}(t)dt \ ,$$ in terms of which we obtain from Eq. (24) $$\hat f^{(2)}_{1,e}(\omega ,s)=\hat R^{(2)}_{e*}\frac{K/2}{s+\xi +i\omega }+\frac{f^{(2)}_{1,e}(\omega ,\tau ^+)e^{-s\tau }}{s+\xi +i\omega } \ .$$ Multiplying Eq. (29) by $g(\omega )d\omega $ and integrating from $\omega =-\infty$ to $\omega =+\infty$, then yields $$\hat R^{(2)}_{e*}(s)=\frac{e^{-s\tau }}{D(s)}\int ^{+\infty}_{-\infty}\frac{f^{(2)}_{1,e}(\omega ,\tau ^+)}{s+\xi +i\omega }g(\omega )d\omega \ .$$ To find $R_e^{(2)*}(t)$ we take the inverse Laplace transform, $$R_e^{(2)*}(t)=\frac{1}{2\pi i}\int ^{+i\infty +\eta }_{-i\infty+\eta }e^{st}\hat R_{e*}^{(2)}(s)ds \ , \eta >0 \ .$$ For the purposes of evaluating the integral (31), we recall that $D(s)=0$ has roots whose real parts correspond to the exponential decay rate of a response to an initial stimulus toward the $R=0$ state. Thus, as before in our discussion of the linear response (see Eq. (16)), any poles at the roots of $D(s)=0$ give contributions that we assume decay substantially faster with increasing $t>\tau $ than the diffusion induced exponential decay rate $\xi $. Since we are interested in echoes that we will find occur for $t=2\tau ,3\tau ,\ldots $, we neglect contributions to Eq. (31) from such poles. Thus it suffices to consider only the contribution to Eq. (31) from the pole at $s+\xi +i\omega =0$. Hence Eqs. (30) and (31) yield $$R_e^{(2)*}(t)\cong \int ^{+\infty}_{-\infty}e^{-(i\omega +\xi )(t-\tau )}\frac{f^{(2)}_{1,e}(\omega ,\tau ^+)}{D[-(i\omega +\xi )]} g(\omega )d\omega \ .$$ D. Echoes --------- In order to see how Eq. (32) results in echoes, we recall our previous results, Eqs. (25), (19) and (21) for $f_{1,e}^{(2)}(\omega ,\tau ^+)$, and combine them to obtain $$f^{(2)}_{1,e}(\omega ,\tau ^+)=d_0d_1h_2h_1^*\frac{\exp (i\omega \tau -\xi \tau )}{D^*[-(i\omega +\xi )]}-d_0d_1\sum _{|n|\geq 2}nh_nh^*_{n-1}\exp [-(in\omega +n^2\xi )\tau ] \ ,$$ where we have used $h_0=0$,  $h_n=h^*_n$, $f^{(1)}_{-1}=f_1^{(1)*}$, and the first term on the right side of Eq. (33) corresponds to $n=-1$ in Eq. (25). Putting Eq. (33) into Eq. (32), we see that we have an integral of a sum over terms with exponential time variations of the form $$\exp\{-i\omega [t-(1-n)\tau ]\}\exp \{ -\xi [t+(n^2-1)\tau \} \ .$$ Considering the first exponential in Eq. (34), we see that, for large values of $|t-(1-n)\tau |$, there is rapid oscillation of the integrand with $\omega $, and the integral can therefore be expected to be near zero. However, such rapid oscillation is absent near the times $t=(1-n)\tau $, at which a large value of $R_e^{(2)*}$ will occur. Since $t>\tau $, the relevant times occur for $n<-1$; e.g., for $n=-1$, we get an echo at $t=2\tau$; for $n=-2$, we get an echo at $t=3\tau $; etc. Therefore, we henceforth replace the summation over $|n|\geq 2$ in Eq. (33) by a summation from $n=-\infty$ to $n=-2$. E. Evaluation for Lorentzian frequency distribution functions ------------------------------------------------------------- We now consider the case of a Lorentzian frequency distribution, $$g(\omega )=g_L(\omega )\equiv \frac{1}{\pi }\frac{\Delta}{(\omega -\bar \omega )^2+\Delta ^2}=\frac{1}{2\pi i}\left\{ \frac{1}{\omega -(\bar \omega +i\Delta )}-\frac{1}{\omega -(\bar \omega -i\Delta )}\right\} \ .$$ The right-most expression for $g_L(\omega )$ makes clear that, when the previously real variable $\omega $ is analytically continued into the complex plane, the function $g_L(\omega )$ results from the sum of two pole contributions, one at $\omega =\bar \omega +i\Delta $, and one at $\omega =\bar \omega -i\Delta $. The quantity $\bar \omega $ represents the average frequency of the distribution, while $\Delta $ represents the width of the distribution. Consideration of the Lorentzian will be particularly useful to us because the integral (32) can be explicitly evaluated, and also because our numerical experiments in Sec. IV will be for the case of a Lorentzian frequency distribution function. As a first illustration we consider the $n=-1$ term which results in an echo at $t=2\tau $. We first evaluate $D(s)$ by inserting the pole-form for $g_L(\omega )$ into Eq. (18) and closing the integration path with a large semicircle of radius approaching infinity. This yields a single residue contribution to $D(s)$, $$D(s)=1-\frac{K}{2}[s+\xi +i(\bar \omega -i\Delta )]^{-1} \ .$$ Note that the solution of $D(s)=0$ occurs at $$s=-i\omega -\left(\xi +\Delta -\frac{K}{2}\right) \ .$$ According to our previous assumptions, we require $K<K_c\equiv 2(\Delta +\xi )$ so that the $R=0$ state is stable, and $(\Delta -K/2)\tau - \xi \tau \gg 1$ so that we can neglect contributions from the pole at the root $D(s)=0$ in our approximation of (31) by (32). Using Eq. (36) and the $n=-1$ contribution to $f^{(2)}_{1,e}$ (i.e., the first term in (33)) in Eq. (32) we obtain for the echo term at $t=2\tau $ (denoted $R^{(2)*}_{2\tau }(\epsilon )$), $$R^{(2)*}_{2\tau }(t)=2ih^*_1h_2d_0d_1\Delta \int^{+\infty}_{-\infty} \frac{d\omega }{2\pi i}\cdot \frac{\exp[-i\omega (t-2\tau )-\xi t]}{[(\omega -\bar \omega )-i(\Delta -\frac{K}{2})][(\omega -\bar \omega )+i( \Delta -\frac{K}{2})]}\ .$$ For $t>2\tau $ $(t<2\tau )$ the integrand exponentially approaches zero as $Im(\omega )\rightarrow -\infty $ $(Im(\omega )\rightarrow +\infty)$, and we can therefore close the integration path with a large semicircle in the lower half $\omega $-plane (upper half $\omega $-plane). Thus the integral (38) is evaluated from the pole enclosed by the resulting path \[i.e., the pole $\omega =\omega _0-i(\Delta -\frac{K}{2})$ for $t>2\tau $, and the pole $\omega =\omega _0+i(\Delta -\frac{K}{2})$ for $t<2\tau $\], $$R^{(2)*}_{2\tau}(t)=\frac{h^*_1h_2d_0d_1\Delta }{\Delta -(K/2)}e^{-i\bar \omega (t-2\tau )-\xi t}e^{-(\Delta -\frac{K}{2})|t-2\tau |} \ .$$ From Eq. (39) we see that we obtain an echo that is approximately symmetric in shape about $t=2\tau $ (i.e., the envelope $\exp [-(\Delta -K/2)|t-2\tau |]$) for $\xi \ll (\Delta -\frac{1}{2}K)$. We can similarly evaluate the contribution $R^{(2)*}_{m\tau }(t)$ of echoes at $t=m\tau $ for $m=3,4,\ldots$. For example, the result for the echo at $t=3\tau $ is $$R^{(2)*}_{3\tau }=\frac{2h^*_2h_3d_0d_1\Delta }{\Delta -(K/4)}e^{-\xi (3\tau +t)}e^{-i\bar \omega (t-3\tau )}E(t-3\tau ) \ ,$$ $$E(t-3\tau )= \left\{ \begin{array}{ll} \exp [\Delta (t-3\tau )] \ , & {\rm for} \ t<3\tau \ , \\ \exp -[(\Delta -\frac{1}{2}K)(t-3\tau )] \ , & {\rm for} \ t>3\tau \ . \end{array} \right.$$ Thus, in the case $\xi =0$, the shape of the pulse envelope $E(t-3\tau )$ is asymmetric about $t=3\tau $, increasing at a more rapid exponential rate (namely, $\Delta )$ as $t$ increases toward $3\tau $, than the slower exponential rate of decrease (namely, $\Delta -(K/2)$) as $t$ increases away from $3\tau $. This is in contrast to the symmetrically shaped envelope $\exp [-(\Delta -\frac{1}{2}K)|t-2\tau |]$ for the echo at $t=2\tau $. In Appendix II we present an evaluation of $R^{(2)*}_{2\tau }(t)$ for the case of a Gaussian frequency distribution function, $$g(\omega )=g_G(\omega )\equiv [2\pi \Delta ^2]^{-1/2}\exp [-(\omega -\bar \omega )^2/(2\Delta ^2)] \ .$$ F. The small coupling limit --------------------------- We now consider a general frequency distribution function $g(\omega )$ but for the case where the coupling between oscillators is small. That is, $K\ll \Delta $, where $\Delta $ denotes the frequency width of $g(\omega )$ about its mean value $\omega =\bar \omega $. In this case a good approximation is provided by setting $K=0$. Thus $D[-(i\omega +\xi )]\cong 1$ and Eq. (33) yields $$f^{(2)}_{1,e}(\omega ,\tau ^+)=d_0d_1\sum ^\infty _{n=1}nh^*_nh_{n+1}\exp [-(-in\omega +n^2\xi )\tau ] \ ,$$ where we have replaced $n$ by $-n$ and used $h_n=h^*_{-n}$. Inserting Eq. (42) into Eq. (32) we obtain $$R_e^{(2)*}(t)=\sum ^\infty _{n=2}(n-1)d_0d_1h^*_{n-1}h_n\tilde g(t-n\tau )e^{-[(n^{2}-1)\tau +t]\xi } \ ,$$ where $\tilde g(t)$ is defined by $$\tilde g(t)=\int ^{+\infty}_{-\infty}d\omega e^{-i\omega t}g(\omega ) \ .$$ Thus, for $K\ll \Delta $, the shape of the echoes at $t=2\tau ,3\tau ,\ldots$ is directly given by the Fourier transform (44) of the frequency distribution function $g(\omega )$. Another point is that with $K\rightarrow 0$, Eq. (1) shows that the oscillators do not interact, and the nonlinearity needed to produce the echo phenomenon comes entirely from the stimulus function $h(\theta )$. IV. Simulations =============== We have performed direct numerical simulations of the system (1) with a Lorentzian oscillator distribution (see Eq. (35)), $\bar \omega =0$, $\Delta =1$ (corresponding to $K_c=2$), $\hat d_0=\hat d_1$, $K=1$, $\tau =50$, and $\xi =0$. At $t=0^{-}$ we initialize each phase $\theta _i$ for $i=1,2,\ldots ,N$ randomly and independently with a uniform distribution in the interval $(0,2\pi )$. We then apply the mapping given by Eq. (7) with $\hat d_p=\hat d_0$ to each $\theta _i$ in order to simulate the effect of the delta function at $t=0$. Next we integrate Eq. (1) for each $i=1,2,\ldots ,N$ forward in time to $t=\tau ^{-}$, again apply the mapping Eq. (7) (but now with $\hat d_p=\hat d_1$), and we then continue the integration. At each time step we also calculate $R(t)$ using Eq. (8). Figure 2 shows results for $\hat d_0=\hat d_1=1/4$, and ![$|R(t)|$ versus $t$ for (a) $N=10^6$, (b) $N=10^5$, (c) $N=10^4$, and (d) $N=10^3$, showing the echo at $t\cong 2\tau $ and the increase of fluctuations at lower $N$.](fig2) ![$|R(t)|$ versus $t$ blown up around $t\cong 2\tau =200$, for $N=10^6$, $10^5$, $10^4$, $10^3$ (solid curves) showing the increase of fluctuations at lower $N$. The dotted curve is the theoretical result from Eq. (39) with $\xi =0$.](fig3) $$h(\theta )=\sin \theta +\sin 2\theta \ ,$$ for several different system sizes, $N=10^6$, $10^5$, and $10^4$. Figure 2(a–c) shows $|R(t)|$ versus $t$ for $0\leq t\leq 125$. The responses to the delta functions at $t=0$ and $\tau $, as well as the echo at time $t=2\tau $ are clearly illustrated. The effect of lower $N$ is to increase the fluctuations making the echo somewhat less distinct. We do not see any echo at $t=3\tau $. This is in agreement with Eq. (40), since $h_3=0$ for the $h(\theta )$ employed in these computations. Figure 3 shows a blow-up of the numerically computed echo around the time $t=2\tau $ for $N=10^6$, $10^5$, and $10^4$. Also, plotted in Fig. 3 as asterisks is the result from our theoretical calculation Eq. (39). Reasonable agreement between the theoretical and computed echo shapes is obtained, although the agreement is somewhat obscured by fluctuation effects at the smaller system sizes $(N)$. While our choice $\hat d_0=\hat d_1=1/4$ might be regarded as questionable for applicability of the small amplitude approximation $(\hat d_p\ll 1$, for $p=0,1$) employed by Eq. (7) and by our theory of Sec. III, we have nonetheless evidently obtained good agreement between the theory and numerical experiment. Figure 4 illustrates the effect of varying the driving amplitude for a network of size $N=10^4$. For $\hat d_0=\hat d_1=1/8$ (Fig. 4(a)) the echo is swamped by the noise and is not seen. For $\hat d_0=\hat d_1=1/4$ (Fig. 4(b), same as 2(a)) the echo seems to have appeared, but because of the noise, this conclusion is somewhat questionable. Finally, at the larger driving of $\hat d_0=\hat d_1=1/2$, the echo is clearly present. Figures 5(a) and 5(b) show the effect of changing $h(\theta )$. In particular, Fig. 5(a) shows numerical results for $\hat d_0=\hat d_1=1/4$, $N=10^5$, and $h(\theta )=\sin \theta $, with all other parameters the same as before. Since $h_2$ is now zero, Eq. (39) now predicts that there is no echo, in agreement with Fig. 5(a). Figure 5(b) shows numerical results for $\hat d_0=\hat d_1=1/4$, $N=10^5$, and $$h(\theta )=\sin \theta +\sin 2\theta +\sin 3\theta,$$ with all other parameters the same as before. Since $h_1$, $h_2$ and $h_3$ are all nonzero, Eqs. (39) and (40) now predict echoes at both $t\cong 2\tau $ and at $t\cong 3\tau $, and this is confirmed by Fig. 5(b). ![Simulation of $10^5$ oscillators for $\tau =100$, $\hat d_0=\hat d_1=1/3$, $K=1=\frac{1}{2}K_c$, $h(\theta )=\sin \theta $. In this case, no echo at $t=2\tau =200$ is observed.](fig4) ![Simulation of $10^5$ oscillators for $\tau =100$, $\hat d_0=\hat d_1=??$, $K=1=\frac{1}{2}K_c$, $h(\theta )=\sin \theta +\sin 2\theta +\sin 3\theta $. In this case, echoes are seen at $t=2\tau =200$ and at $t=3\tau =300$. The inset shows a blow-up of the numerical result for the echo shape at $t=3\tau $ with the theoretical result, Eq. (41), superposed (dotted curve).](fig5) Finally, we note that similar numerical experiments to all of the above have been repeated using a Gaussian $g(\omega )$, and these yield similar results (not shown). V. Discussion ============= Echo phenomena as used for MRI provide a powerful medical diagnostic tool. Echoes in plasmas have also been used as a basis for measuring velocity space diffusion of plasma particles[@jensen69]. Thus it is of interest to consider whether there are potential diagnostic measurement uses of echoes in the context of situations that can be described by the Kuramoto model and its variants. For example, we note that the amplitude of the echo varies exponentially with $\xi $, providing a possible means of determining the phase diffusion coefficient $\xi $. For example, the amplitude of the echo at $t=2\tau $ varies as $e^{-\xi \tau }$. Thus the log of the ratio of measurements of the echo amplitude using two different values of $\tau $, divided by the difference in the $\tau $ values, provides a potential means of estimating $\xi $. Also, as indicated by Eq. (43), if one can lower the coupling $K$ sufficiently, then echoes provide a potential way of determining the oscillator frequency distribution function $g(\omega )$. In particular, for low $K$ the distribution $g(\omega )$ is directly given by the inverse Fourier transform of the echo profile. On the other hand, we have seen from the simulations in Sec. IV that finite $N$ leads to noise-like behavior that may compromise such attempts. We also note that the Kuramoto model is an idealization, and application to any given situation may require modifications of the model and theory to more closely correspond to the situation at hand. We, nevertheless, feel that consideration of echoes for diagnostics may be of potential use. Furthermore, these phenomena are of theoretical interest from at least two points of view. First, as mentioned in Sec. IIIb, the memory required by the echo phenomenon can be thought of as leading to a macroscopically observable consequence of the continuous spectrum[@strogatz92] of the Kuramoto model. A second point of theoretical interest relates to the recent work in Ref.  [@ott]. In that paper it was shown for a general class of initial conditions that are on a certain manifold of the infinite dimensional state space of the Kuramoto system, that the future time evolution of the order parameter is determined by the current value of the order parameter. In particular, there is an ordinary differential equation describing the order parameter evolution. The echo phenomenon provides an example showing that, if initial conditions do not lie on the specified manifold of Ref.  [@ott], other behavior can occur. In particular, well after the second stimulus (at $t=\tau $) and well before the occurrence of the first echo (at $t=2\tau $), the order parameter is essentially zero, yet it does not remain zero as would be predicted for initial conditions on the manifold of Ref.[@ott] for $K<K_c$. This is discussed further in Appendix III. In conclusion, we hope that our work will stimulate experimental groups to investigate the type of situations we have addressed. This work was supported by ONR (N00014-07-1-0734) and NSF (PHY0456249). Appendix I: Linear Analysis {#appendix-i-linear-analysis .unnumbered} =========================== In this Appendix we solve Eq. (15) for $0<t<\tau $ to obtain the solution (16) and (17) for $K<K_c$. Taking the Laplace transform, $\hat u(s)=\int ^\infty _0u(t)e^{-st}dt$, Eq. (15) yields $$\hat f_1^{(1)}(\omega ,s) =\left( \frac{K}{2}\hat R_*^{(1)}(s)+ih_1d_0\right) /(s+\xi +i\omega )$$ where $\hat R_*^{(1)}(s)$ denotes the Laplace transform of $R^{(1)*}(t)$. Multiplying Eq. (45) by $g(\omega )d\omega $ and integrating from $\omega =-\infty$ to $\omega =+\infty$, we obtain $$\hat R_*^{(1)}(s)=ih_1d_0I(s)/D(s) \ ,$$ where $I(s)=\int ^{+\infty}_{-\infty}d\omega g(\omega )/(s+\xi +i\omega )$ and $D(s)=1-(K/2)I(s)$. Inserting Eq. (46) in (45) gives $$\hat f_1^{(1)}(\omega ,s)=ih_1d_0[D(s)(s+\xi +i\omega )]^{-1} \ .$$ As noted in Sec. IIIb, $\hat f_1^{(1)}(\omega ,s)$ has poles in $s$ at the zeros of $D(s)$ and at $s=-(i\omega +\xi )$. These yield time dependences of the inverse Laplace transform of $\hat f_1^{(1)}$ (see Eq. (27)) that vary as $e^{s_{0}t}$ and as $e^{-(i\omega +\xi )t}$, respectively, where $s_0$ denotes the root of $D(s)=0$ with the least negative real part. For $t\approx \tau $ and $-[Re(s_0)+\xi ]\tau \gg 1$, we can neglect the contributions from poles arising from roots of $D(s)=0$, and use only the contribution from the pole at $s=-(i\omega +\xi )$. From Eqs. (27) and (47) this yields $$f_1^{(1)}(\omega ,t)\cong ih_1d_0e^{-(i\omega +\xi )t}/D[-(i\omega +\xi )] \ ,$$ thus confirming Eqs. (16) and (17). Appendix II: Echo at $t=2\tau $ for Gaussian $g(\omega )$ {#appendix-ii-echo-at-t2tau-for-gaussian-gomega .unnumbered} ====================================================== We consider the case $g(\omega )=g_G(\omega )\equiv (2\pi \Delta ^2)^{-1/2}\exp [-(\omega -\bar \omega )^2/(2\Delta ^2)]$. Putting this expression for $g(\omega )$ and the $n=-1$ contribution to $f_{1,e}^{(2)}$ (i.e., the first term in Eq. (33)) into Eq. (32) we have, $$R^{(2)*}_{2\tau }(t)=\frac{h_1h_2d_0d_1}{\sqrt{2\pi \Delta ^2}}\int ^{+\infty}_{-\infty}d\omega \frac{\exp-\left\{ \frac{[(\omega -\bar \omega )+i\Delta ^2(t-2\tau )]^2}{2\Delta ^2}+\frac{\Delta ^2}{2}(t-2\tau )^2+i\bar \omega (t-2\tau )-\xi t\right\}}{D[-(i\omega +\xi )]D^*[-(i\omega +\xi )]} \ .$$ The collective damping rate is determined by the root of $D(s)=0$ with the least negative real part. Denote this root $s=s_0$ where $$s_0=-(i\bar \omega +\xi +\gamma _0)=-(i\omega _0+\xi ) \ ,\ \omega _0=\bar \omega -i\gamma _0 \ ,$$ where $\gamma _0>0$ is real. Letting $F(\omega )\equiv D[-(i\omega +\xi )]$, continuing this function from real $\omega $ into the complex $\omega $-plane, and expanding around $\omega =\omega _0$, we have $$F(\omega )=(\omega -\omega _0)\eta +\mathcal{O}[(\omega -\omega _0)^2] \ ,$$ where $\eta $ is a complex constant. Letting $F_*(\omega )$ denote the continuation of the function of the real variable $\omega $ in Eq. (49), $D^*[-(i\omega +\xi )]$, into the complex $\omega $-plane, we have that this function has a zero at $\omega =\omega ^*_0$, $$F_*(\omega )=(\omega -\omega ^*_0)\eta ^*+\mathcal{O}[(\omega -\omega ^*_0)^2] \ .$$ Considering the oscillatory $\omega $ variation in the numerator of the integrand of Eq. (49) to be rapid (valid for $K\ll \gamma _0$), we can approximate the integral by the saddle point method, where the saddle point is at $$\omega _{sp}=\bar \omega -i\Delta ^2(t-2\tau ) \ ,$$ and the steepest descent path through $\omega =\omega _{sp}$ runs along the horizontal line $Im(\omega )=-\Delta ^2(t-2\tau )$ from $Re(\omega )=-\infty$ to $Re(\omega )=+\infty$ (see Fig. 6). From Fig. 6(a) we see that for $\Delta ^2|t-2\tau |<\gamma _0$, the poles at $\omega =\bar \omega \pm i\gamma _0$ are not intercepted by the steepest descent path, while for $\Delta ^2|t-2\tau |>\gamma _0$ one of the poles is intercepted (e.g., Fig. 6(b)). In the case where a pole is intercepted, its contribution dominates the contribution from the saddle point by virtue of its time dependence, $e^{-\gamma _{0}|t-2\tau |}$, as opposed to the saddle point contribution time dependence, $e^{-\frac{1}{2}\Delta ^2(t-2\tau )^2}$. Thus we obtain $$R^{(2)*}_{2\tau }(t)\sim e^{-i\bar \omega (t-2\tau )-\xi t}\times \left\{ \begin{array}{ll} e^{-\frac{\Delta ^2}{2}(t-2\tau )^2} & {\rm for} \ |t-2\tau |<2\gamma _0/\Delta ^2 \ , \\ e^{-\gamma _0|t-2\tau |} & {\rm for} \ |t-2\tau | >2\gamma _0/\Delta ^2 \ . \end{array} \right.$$ Near $\gamma _0=\Delta ^2|t-2\tau |/2$, the pole is near the saddle point, and a uniform asymptotic expansion of the integral (49) is necessary to obtain the transition between the two forms in Eq. (53). ![(a) Steepest descent path (dashed) through the saddle point $\omega =\omega _{sp}$ for $\Delta ^2|t-2\tau |<\gamma _*$. (b) The steepest descent path (dashed) for $\Delta ^2(t-2\tau )<-\gamma _*$. The dominant poles at the roots of $F(\omega )=0$ and $F_*(\omega )=0$ are shown as crosses, where in (b) the steepest descent path has intercepted the pole $\omega =\omega _0$ resulting in a pole contribution to $R^{(2)*}_{2\tau }(t)$.](fig6) ![(a) Steepest descent path (dashed) through the saddle point $\omega =\omega _{sp}$ for $\Delta ^2|t-2\tau |<\gamma _0$. (b) The steepest descent path (dashed) for $\Delta ^2(t-2\tau )<-\gamma _0$. The dominant poles at the roots of $F(\omega )=0$ and $F_*(\omega )=0$ are shown as crosses, where in (b) the steepest descent path has intercepted the pole $\omega =\omega _0$ resulting in a pole contribution to $R^{(2)*}_{2\tau }(t)$.](fig7) Appendix III:  Further Discussion of Ref.[@ott] {#appendix-iiifurther-discussion-of-ref. .unnumbered} =============================================== In Ref.  [@ott], a broad class of noiseless (e.g., $\xi =0$ in Eqs. (3) and (11)) globally coupled systems of phase oscillators was studied. The simplest example of this class is the Kuramoto model. Reference [@ott] considered Lorentzian $g(\omega )$ and a special class of initial conditions. Referring to Eq. (12), these initial conditions are of the form, $$f_n(\omega ,0)= \alpha ^n(\omega ) \ , \ \ {\rm for } \ \ n\geq 0 \ , \\$$ and $f_n(\omega ,0)=f^*_{-n}(\omega ,0)\ , \ \ {\rm for} \ \ n\leq 0 $ , where $|\alpha (\omega )|<1$ for $\omega $ on the real axis, $\alpha (\omega )$ is analytic in ${\rm Im} (\omega )< 0$, and $|\alpha (\omega )|\rightarrow 0$ as ${\rm Im} (\omega )\rightarrow -\infty $. Under these conditions, Ref. [@ott] shows that the order parameters (or parameter), see Eq. (12), that describe the nonlinear, macroscopic time evolution of the given system satisfy a finite set of ordinary differential equations in time. Thus the order parameter dynamics is low dimensional, while the dynamics of the full system determining the evolution of the distribution function $f(\omega ,\theta ,t)$ is infinite dimensional[@ott]. For example, for the Kuramoto problem with the above conditions satisfied, Ref.  [@ott] shows that $$dR/dt+\left( \Delta -\frac{1}{2}K\right) R + \frac{1}{2}K\Delta |R|^2R=0 \ ,$$ where we have taken $\bar \omega =0$ in Eq. (35). A consequence of Eq. (55) is that for $K<2\Delta \equiv K_c$, $|R(t)|$ [*decreases monotonically*]{} to zero. This behavior is not followed in the echo phenomena we discuss in the present paper. In particular, in Fig. 1, $|R(t)|$ is small between $t=\tau $ and $t=2\tau $, but then [*increased*]{} to form the echo in the vicinity of time $t=2\tau $. Referring to Eq. (34) and our subsequent discussion, we see that this is because there is a component of $f_1(\omega ,t)$ that varies as $\exp [-i\omega (t-2\tau )]$. Identifying $f_1(\omega ,0)$ in the linear problem with $\alpha (\omega )$ in the nonlinear problem \[Eq. (54)\] and considering $t_0$ as a new initial time (shift time so that $t_0$ goes to $t=0$), we see that $\alpha (\omega )\sim \exp [-i\omega (t_0-2\tau )]$. If we take $t_0$ to be such that $\tau <t_0<2\tau $ and $|R(t_0)|$ is small, then $\alpha (\omega )$ does not satisfy the condition of Ref.  [@ott] that $\alpha (\omega )\rightarrow 0$ as $Im(\omega )\rightarrow -\infty$. However, if $t_0>2\tau $, then it does. Thus the increase of $|R(t)|$ occurs only when the hypothesis under which Eq. (55) was derived does not hold. More generally, consider an initial condition for the original Kuramoto problem (without stimuli or noise) where $f_1(\omega ,0)$ is analytic on the real $\omega $-axis. Expressing $f_1(\omega ,0)$ as a Fourier integral transform, we have $$f_1(\omega ,0)=\int ^{+\infty}_{-\infty} e^{i\omega \eta }k(\eta ) d\eta \ ,$$ where $k(\eta )$ is the Fourier transform of $f_1(\omega ,0)$. Since $f_1(\omega ,0)$ is analytic in $\omega $, $k(\eta )$ decreases exponentially for sufficiently large $\eta $, $$|k(\eta )|<He^{-\beta \eta } \ , \ \ {\rm if}\ \ \eta >\eta _0 \ ,$$ for some set of positive constants $H,\beta ,\eta _0$. Using the Laplace transform technique (as in Appendix I), it can be shown that the solution to the linearized initial value Kuramoto problem contains a component of $f_1(\omega ,t)$ of the form $\exp (-i\omega t)f_1(\omega ,0)$, which we can express using Eq. (56) as $$\exp(-i\omega t)f_1(\omega ,0)=\int ^t_{-\infty}e^{-i\omega (t-\eta )}k(\eta )d\eta +\int ^\infty _te^{i\omega (\eta -t)}k(\eta )d\eta \ .$$ Setting $t=t_0$ and regarding $t=t_0$ as a new initial condition time, we note that the initial condition consists of two terms, namely the first and second integrals on the right hand side of Eq. (58). For $t_0>\eta _0$ sufficiently large, the second integral is smaller than the first by a factor of order $\exp (-\beta t_0)$. Furthermore, the first integral satisfies the condition $f_1(\omega ,t_0)\rightarrow 0$ as ${\rm Im}(\omega )\rightarrow -\infty$ \[because $(\eta -t_0)>0$ for the first integral\], while the second integral does not. Thus, if we choose to shift what we designate as the initial time to sufficiently large $t_0$, then aside from an exponentially small component of order $\exp (-\beta t_0)$, the initial condition obeys the requirement of Ref. [@ott] that $f_1(\omega ,t_0)$ goes to zero as ${\rm Im} (\omega )\rightarrow -\infty$. A. Pikovsky, M. Rosenblum and J. Kurths, [*Synchronization: A Universal Concept in Nonlinear Science*]{} (Cambridge University Press, 2001). S. H. Strogatz, [*Sync: The Emerging Science of Spontaneous Order*]{} (Penguin Science Press, 2004). J. Buck, Q. Rev. Biology [**63**]{}, 265 (1988). T. J. Walker, Science [**166**]{}, 891 (1969). D. C. Michaels, Circulation Research [**61**]{}, 704 (1987). W. Singer, Ann. Rev. Physiology [**55**]{}, 349 (1993); R. Eckhorn et al., Biological Cymbernetics [**60**]{}, 121 (1988); C. M. Gray, Nature [**338**]{}, 334 (1989). S. Yamaguchi et al., Science [**302**]{}, 1408 (2002); T. M. Antonsen et al., arXiv:0711.4135. I. Z. Kiss, Y. Zhai and J. L. Hudson, Science [**296**]{}, 1676 (2005). K. Wiesenfeld and J. W. Swift, Phys. Rev. E [**51**]{}, 1020 (1995). Y. Kuramoto, [*Chemical Oscillations, Waves and Turbulence*]{} (Springer, 1984); and in [*International Symposium on Mathematical Problems in Theoretical Physics*]{}, [**39**]{}, edited by H. Araki (Springer-Verlag, Berlin, 1975). For reviews of the Kuramoto model see J. A. Acebron, et al., Rev. Mod. Phys. [**77**]{}, 137 (2005); S. H. Strogatz, Physica D [**143**]{}, 1 (2000); and E. Ott, [*Chaos in Dynamical Systems*]{}, second edition, chapter 6, section 6. (Cambridge University Press, 2002). E. L. Hahn, Phys. Rev. [**80**]{}, 580 (1950). R. W. Gould, Phys. Lett. [**19**]{}, 477 (1965); F. W. Crawford and R. S. Harp, J. Appl. Phys. [**37**]{}, 4405 (1966). E. Ott, J. Plasma Phys. [**4**]{}, 471 (1970). M. Porkolab and J. Sinnis, Phys. Rev. Lett. [**21**]{}, 1227 (1968). T. M. O’Neil and R. W. Gould, Phys. Fluids [**11**]{}, 134 (1968); J. H. Malmberg, C. B. Wharton, R. W. Gould and T. M. O’Neil, Phys. Fluids [**11**]{}, 1147 (1968). S. H. Strogatz, R. E. Mirollo and P. C. Matthews, Phys. Rev. Lett. [**68**]{}, 2730 (1992). T. H. Jensen, J. H. Malmberg and T. M. O’Neil, Phys. Fluids [**12**]{}, 1728 (1969); C. H. Su and C. Oberman, Phys. Rev. Lett. [**20**]{}, 427 (1968); T. M. O’Neil, Phys. Fluids [**11**]{}, 2420 (1968). E. Ott and T. M. Antonsen, Phys. Rev. Lett. (submitted).
{ "pile_set_name": "ArXiv" }
It is common to distinguish two different types of quantum Hall effects: the integer and the fractional [@Girvin]. When an integer number of Landau levels are filled and the kinetic energy cost for occupying higher Landau levels is large enough, the occupation of higher Landau levels encouraged by many-body interactions can be treated perturbatively. In the fractional quantum Hall effect, many-body interactions stabilize a ground state with a gap to excitations even though only a fraction of a Landau level is occupied. It is assumed that only a single Landau level is partially occupied, because of the large kinetic energy cost for occupying higher Landau levels, and that this state is adiabatically connected to the state with higher Landau level occupation. We will argue that this traditional picture is incomplete. When the energy gap between Landau levels is small enough, there is a new even integer quantum Hall effect that is distinct from the old one, and whose stability relies on the existence of many-particle interactions and higher Landau level occupation. This work is motivated by the fact that the topology of the experimental phase diagram is inconsistent with an integral quantum Hall effect insensitive to many-particle interactions [@Jiang1; @Jiang2; @Pepper]. Of particular interest is the existence of a direct second order phase transition from $\nu=2$ to $\nu=0$ with the usual quantum Hall critical exponents (See Fig. 1; $\nu$ is defined as the value of the Hall conductivity $h\sigma_{xy}/e^2$ in fundamental units). In the non-interacting picture, the system is required to pass through an intermediate $\nu=1$ phase in the transition from $\nu=2$ to $\nu=0$ [@Girvin]. However, there is no experimental evidence for even a very thin region separating $\nu=2$ and $\nu=0$. We therefore postulate the existence of a new quantum Hall state, which we call the $2b$ state, that is distinct from the non-interacting spin-unresolved $2a$ state. The transition from the $2a$ to the $2b$ state would be first order in the absence of disorder and is probably second order in the presence of disorder. The global phase diagram implies that there are direct transitions between even-integer quantum Hall states in low magnetic fields. In fact, at low enough magnetic fields, the odd-integer quantum Hall effect is known to disappear [@Jiang1]. In addition, the transition directly from $\nu=0$ to $\nu=2$ and back to $\nu=0$ at low magnetic fields, shown in the phase diagram, has also been observed experimentally [@Jiang2; @Pepper]. epsf =8.5 cm The essential idea behind the $2b$ quantum Hall state is that up and down spin electrons pair to form a spinless boson that condenses into a bosonic many-body fractional quantum Hall state. It is clear this state must have significant occupation of higher Landau levels because the conventional spin unpolarized $2a$ state is the unique state at $\nu=2$ in the lowest Landau level. For purposes of illustration, we describe a spin-singlet many electron wavefunction [@Herbut] for the $2b$ state given by $$\label{wfn} \Phi=\prod_{k<l,\sigma}(z^\sigma_k-z^\sigma_l)\ {\rm Per} \left[ F(|z_i-z_j|) \right] {\large e}^{-\sum_i |z_i|^2/4l^2}$$ where $z_i$ is the coordinate of the i$^{th}$ electron, $z^\sigma_k$ is the coordinate of the k$^{th}$ spin $\sigma$ electron, $l$ is the magnetic length, and Per denotes the Permanent of the symmetric matrix whose $(i,j)^{th}$ component is $F(|z_i-z_j|)$. Given an order N matrix M$_{ij}$, the permanent Per$(M)=H_1(M)$ where $H_\alpha(M)=\sum_{P}(\alpha)^{S\{P\}}\prod_{ab}M_{ab}$, the product is over the $N/2$ pairs ${ab}$, the sum is over the possible permutations of these pairs, and $S\{P\}$ is $-1$ or $+1$ if the permutation $P$ is odd or even, respectively [@Wilczek; @Read]. The Pfaffian of a matrix M is Pf$(M)=H_{-1}(M)$. The behavior of the $2b$ state described in Eq. (\[wfn\]) is governed by the function $F$. For $F=1$, $\Phi$ is the conventional $2a$ state. On the other hand, if we take $F$ to be a delta function then we can define a pair coordinate and $\Phi$ vanishes as the second power of the pair coordinate as two pairs approach each other. This wavefunction is therefore equivalent to a $\nu'=1/2$ Laughlin state for bosonic pairs of charge $e^*=2e$. Its Hall conductivity is $\sigma_{xy}=(e^*)^2\nu'/h=2e^2/h$. If $F$ is short-ranged then it can be thought of as a pair wavefunction and its effective size is the coherence length. For distances much longer than this coherence length, $F$ acts like a delta-function and $\Phi$ appears to be a $\nu'=1/2$ state. The $2a$ to $2b$ transition is thus characterized by the divergence of this coherence length. At first, the existence of a pairing state stabilized by repulsive Coulomb interactions seems antintuitive. Moreover, because $\Phi$ in Eq. (\[wfn\]) is constructed by occupying higher Landau levels, we lose kinetic energy that must be compensated by a gain in interaction energy. However, we only lose interaction energy among the two electrons in a pair; we gain repulsive energy from pairing because the $2b$ state acts like a Laughlin $\nu'=1/2$ state for pairs. As a first approximation, let us ignore the energy cost in the $2b$ state for forming tightly bound pairs and calculate the interaction energy for all electrons not in the same pair. This is simply the Coulomb energy of a $\nu'=1/2$ state of spinless charge $e^*=2e$ bosons. Laughlin’s interpolation formula for $E_m$ [@Girvin], the Coulomb energy of a projected $\nu=1/m$ state, gives an energy per electron for the $2b$ state of $E^*_{2b}=-.49 (e^*)^2/\epsilon l^*=-1.39 e^2/\epsilon l$. In contrast, the energy of the $2a$ state is $E_{2a}=-\sqrt{\pi/8}\ e^2/\epsilon l$. Forming a paired state has yielded an energy gain per pair of $2(E_{2a}-E^*_{2b})=1.53 e^2/\epsilon l$ which must more than compensate the Coulomb and kinetic energy lost by pairing if the $2b$ state is to be energetically stable. This requirement is made easier by the softening of the Coulomb repulsion at short distances due to the finite thickness of the electron gas [@Girvin; @Wilczek]. As the magnetic field decreases, Coulomb energy becomes more important and the Coulomb energy gain of a $2b$ state with respect to a $2a$ state can more easily outweigh the loss of kinetic energy. To get a feeling for the magnitude of the magnetic field at which such a transition could occur, we study the properties of the $2a$ state perturbatively in powers of the ratio of the interaction energy to the kinetic energy, $$y={e^2/\epsilon l\over \hbar \omega_c}=l/a^*=\sqrt{B^*\over B}\ ,$$ where $a^*=\epsilon \hbar^2 /m^*$ is the effective Bohr radius and $m^*$ is the band effective mass. The $2a$ state is exact in the infinite magnetic field or $y=0$ limit and is perturbatively stable for sufficiently small $y$. Failure of perturbation theory, which we take to signal a transition, would occur at a critical value of $y_c\sim 1$. For concreteness, we take $y_c=2.3$, the value at which the second order perturbative expression for the gap of the $2a$ state vanishes [@Skyrmions]. Given material parameters for $GaAs-Al_xGa_{1-x}As$ of $a^*=100 \AA$ and $B^*=6.6 T$, this value of $y_c$ corresponds to a magnetic field of $B=B^*/y_c^2=1.3T$ which is in the same range as $.5T$, the experimental magnetic field at which the $\nu=0$,$1$, and $2$ quantum Hall states coincide (see Fig. 1) [@Jiang1]. [**BCS Instability:**]{} We will employ the technique of flux attachment to show that a spin degenerate electron gas at $\nu=2$ has the pairing instability necessary to stabilize the $2b$ state [@Zhang; @dhlee]. This procedure has the advantage that it implicitly describes wavefunctions not projected into the lowest Landau level and provides a motivation, other than energetics, for constructing a paired wavefunction. We choose as our basis a set of quasiparticles that are defined as an electron attached to one unit of spin-relative flux. This construction is accomplished by coupling $j^\sigma_\mu$, the $\mu^{\rm th}$ component of the current of spin $\sigma$ electrons, to the gauge field $a^\mu_\sigma$, and requiring that $$\label{constraint} \nabla \times {\bf a}_\sigma({\bf r})=2\pi \rho_{-\sigma}({\bf r'})\$$ where $\rho_\sigma({\bf r})$ denotes the density of spin $\sigma$ electrons. This flux attachment makes unlike spins relative bosons and keeps like spins relative fermions. The spin $\sigma$ quasiparticles see a mean magnetic field $$B_{eff,\sigma}= \nabla \times ({\bf A}-<{\bf a}_\sigma>)\ ,$$ which is zero since a $\nu=2$ spin-unpolarized state satisfies $\nabla \times {\bf A}=2\pi<\rho_\sigma({\bf r})>$. At the mean-field level, the quasiparticles form a spin-degenerate Fermi sea because like spins are non-interacting relative fermions in zero magnetic field. We can calculate the effect of fluctuation corrections on both the system’s linear-response function and the dressed interparticle interaction [@HLR]. First of all, we define the linear-response function $\Pi$ by the statement that a spin-dependent gauge field $A_{ext}^{\sigma'}$ of frequency $\omega$ and wavevector ${\bf q}$ induces a current of spin $\sigma$ quasiparticles $$J^{\sigma}_\mu(\omega,{\bf q})=\Pi^{\sigma,\sigma'}_{\mu\nu} (\omega,{\bf q})A^{\sigma'}_{ext,\nu}(\omega,{\bf q})$$ where $\mu,\nu$ take on the values (0,x,y). We follow the standard convention of working in Coulomb gauge so that ${\bf q}\cdot{\bf A^\sigma}={\bf q}\cdot{\bf a^\sigma}=0$. The linear-response kernel of the system can therefore be described by using only the 4x4 submatrix $\Pi^{\sigma,\sigma'}_{\alpha\beta}$ in which the subindices $\alpha$ and $\beta$ take on the values 0 and 1 corresponding to the time and transverse component of a vector, respectively. In this notation, the response function of the mean-field degenerate Fermi sea is $$K^{\sigma,\sigma'}_{\alpha\beta}(q,\omega) =\delta^{\sigma,\sigma'}\delta_{\alpha\beta}P_{\alpha\alpha}(q,\omega)\$$ while the bare momentum space interaction is $$\label{propagator} V^{\sigma,\sigma'}_{\alpha\beta}(q,\omega) ={2\pi e^2\over \epsilon q}\delta_{\alpha,o}\delta_{\beta,o} +{2\pi i\over q}\epsilon_{\alpha\beta}(1-\delta^{\sigma,\sigma'})\ .$$ We have assumed a spin-independent Coulomb interaction between electrons. In the RPA [@Zhang; @HLR], the screened momentum space interaction $D$ and response kernel $\Pi$ satisfy the matrix equations $$\label{rpa} D = V+V K D\ ,\ {\rm and}\qquad \Pi = K+K V \Pi\ .$$ We now use this flux attachment formalism to show that the system has an instability towards singlet BCS pairing. Because we are interested in spin-singlet pairing and have transmuted the particles into relative bosons, the pairing must be odd parity. The BCS instability can be shown to be strongest in the p-wave channel, the lowest allowed angular momentum channel. The screened interaction $D$, given in Eq. (\[rpa\]), yields the BCS pair interaction in the singlet channel $$\begin{aligned} V({\bf k},{\bf k'},\omega)&=&D^{\uparrow\downarrow}_{00}(q,\omega)+ ({\bf k} \times {\bf k'}) D^{\uparrow\downarrow}_{01}(q,\omega)/q\nonumber\\ &+&|{\bf k} \times {\bf k'}|^2 D^{\uparrow\downarrow}_{11}(q,\omega)/q^2\ .\end{aligned}$$ where $q=|{\bf k}-{\bf k'}|$ [@Wilczek; @Bonesteel]. We assume a gap function with odd $m$-wave pairing, $\Delta({\bf k})=|\Delta(k)|e^{-im\phi_{\bf k}}$. Because $D^{\sigma,\sigma'}_{01}$ is pure imaginary and $V({\bf k},{\bf k'},\omega)$ thus has a non-vanishing imaginary component, the pairing instability in the BCS gap equation, for a fixed $|m|$, is strongest for $m<0$. However, the dominant attraction in all angular momentum channels comes from $D^{\uparrow\downarrow}_{11}(q,\omega)$ [@Bonesteel; @unpublished]. A spin-singlet wavefunction corresponding to the flux attachment calculations is given by $$\label{wfn2} \Phi'=\prod_{k,l}{1\over(z^\uparrow_k-z^\downarrow_l)^{\bf *}} \ {\rm Pf} \left[ (z_i-z_j)G_{ij} \right] {\large e}^{-\sum_i |z_i|^2/4l^2}$$ instead of Eq. (\[wfn\]) where $G_{ij}=G(|z_i-z_j|)$ and the Pfaffian is defined in conjunction with Eq. (\[wfn\]). The prefactor in Eq. (\[wfn2\]) implements the singular gauge transformation described in Eq. (\[constraint\]) and can be shown to make $\Phi'$ into a spin-singlet wavefunction. (If we had chosen to implement the gauge transformation via the prefactor $\prod_{k,l}(z^\uparrow_k-z^\downarrow_l)$ then $\Phi'$ would not be a spin-singlet.) In contrast to the gauge theory description of $\Phi'$, a similar description of $\Phi$, given in Eq. (\[wfn\]), would transmute the electrons into bosonic quasiparticles. Therefore, a pairing mechanism describing $\Phi$ would have to bind two charge $e$ bosons into a charge $2e$ boson [@Rice]. Nevertheless, we believe that both $\Phi$ in Eq. (\[wfn\]) and $\Phi'$ describe the same physics in the thermodynamic limit. The physical behavior of the paired state is most easily expressed in terms of the gauge field and current combinations [@Zee]: $$a^\pm={1\over2}(a^\uparrow\pm a^\downarrow)\ ,\qquad j^\pm_\mu =j^\uparrow_\mu\pm j^\downarrow_\mu\ .$$ The $j^+$ current only couples to the $a^+$ gauge field while the $j^-$ current only couples to the $a^-$ gauge field. A singlet paired state has a gap $\Delta_s$ for single particle excitations in both the $j^+$ and $j^-$ channels. In addition, the $j^+$ channel has a gapless Goldstone mode whose spectrum is given by the poles of $a^+$’s propagator. While fluctuations of the statistical gauge field $a$ do not affect the low-energy spectrum of the $j^-$ channel [@Hall-Insulator], they gap the Goldstone mode of the $j^+$ channel and give it a quantized $\nu=2$ Hall effect [@Zhang; @dhlee; @HLR]. Both of these effects are apparent in the screened response functions $\Pi^\pm=2(\Pi^{\uparrow\uparrow}\pm\Pi^{\uparrow\downarrow})$ where $\Pi$ is defined in Eq. (\[rpa\]) and spin-symmetry assumed. [**Implications:**]{} The $2b$ state has a number of different quasiparticle excitations. Because the paired state describes spinless bosons at filling fraction $\nu'=1/2$, we can construct the usual quantum Hall quasiparticles with charge $\pm e^*\nu'=\pm e$ and statistics $\pm\pi\nu'=\pm\pi/2$. We refer to these charge 1 spinless semions as holons. When we bind a single electron to a holon, we make a charge 0 spin 1/2 semion which we refer to as a spinon. In addition, electrons are also quasiparticle excitations and may be viewed as a bound spinon-holon pair. Let us now construct quantum Hall hierarchies by using the $2b$ paired state as a parent state. Because electrons are well-defined quasiparticles, one can imagine constructing conventional integral quantum Hall states of up and down spin electrons on top of the parent $2b$ state. These are called $nb$ states in the figure. On the other hand, we can pair up and down spin electrons into a new boson and form a $\nu'=1/2$ bosonic quantum Hall states of these new pairs. The $4c$ state shown in Fig. 1 is an example of such a state. We can construct yet another hierarchy of paired spin-unpolarized states based on the condensation of holons at filling fractions $\nu=4\nu'=8p/(4p+1)$ where $p$ is an integer [@Girvin]. In analogy to the $2b$ state, we have constructed other singlet paired states at filling fraction $\nu=2/(2p+1)$ where p is an integer. These states correspond to bosonic quantum Hall states of charge $2e$ bosons at filling fraction $\nu'=1/2(2p+1)$. If we try to construct triplet paired states, spin-rotation symmetry requires them to be spin-polarized. However, there is no BCS instability for spin-polarized states; in fact, the $\omega=0$ BCS interaction is repulsive in all angular momentum channels [@Bonesteel]. Therefore, these states are unlikely to exist in nature unless spin symmetry is broken. In fact, there is evidence for triplet pairing in bilayer systems which do not have pseudospin-rotation (layer index) symmetry [@bilayer]. Let us now address the nature of the transitions between the various states shown in Fig. 1, in the presence of dirt. It is well established that the transitions between daughter and parent quantum Hall states are second order and are all similar to each other. These transitions can be viewed as the quantum percolation transition of a single edge state with the selection rule [@Kivelson] that only one state can undergo this transition at a time. In this language, the $2a$ state has 2 edge modes, one composed of each spin polarization, and the $2a$ to $1$ transition is associated with the percolation of the down spin edge state. Alternatively, for small Zeeman energy, the $2a$ state can be viewed as having a charged holon edge mode and a neutral spinon edge mode. The transition from $2a$ to $2b$ can therefore be viewed as spinon percolation. We postulate that this transition is similar to the $2a$ to $1$ transition but that the critical behavior is visible in the spin degrees of freedom and that the system is a spin-metal at the critical point. In addition, we identify the tetracritical point in Fig. 1 as the intersection of the phase boundaries associated with the two order parameters: the Girvin-MacDonald order parameter defined by attaching one unit of ${\bf a}^\uparrow$ flux to the up spin electron creation operator is nonzero in both the $1$ and $2a$ phases; the order parameter defined by attaching one unit of $({\bf a^\uparrow}+{\bf a^\downarrow})/2$ flux to the charge $2e$ pair creation operator is nonzero in both the $2a$ and $2b$ phases [@Girvin]. [**Other Experimental Implications:**]{} It has been conjectured that there is a universal value for the resistance at quantum hall phase transitions [@Kivelson]. Recent experiments have in fact found universal values for the resistance at quantum hall liquid to insulator transitions [@Jiang1; @Jiang2; @Pepper; @Tsui]. According to the pairing theory, the transition between $\nu=2$ and $\nu=0$ corresponds to a $\nu'=1/m=1/2$ quantum hall liquid to insulator transition of charge $e^*=2e$ quasiparticle with the corresponding [*universal resistances*]{} [@Kivelson]: $\rho_{xx}=h/(e^*)^2=h/4e^2$ and $\rho_{xy}=mh/(e^*)^2=h/2e^2$. This prediction is consistent with the observed values for the resistances in the experiments of Wong [*et. al.*]{} [@Jiang1]. (It is important to note that there are experiments at higher magnetic fields that find different resistances at the $\nu=2$ to insulator phase transition or a different critical exponent [@Pepper]. This discrepancy is not yet understood.) Our theory predicts a spin gap $\Delta_s$ associated with pairing. In the presence of dirt, this may become a pseudogap. Nonetheless, this gap should be experimentally observable by studying the density of free spins as a function of $\delta B=B-\pi\rho$ [@Skyrmion-exp]. The density of free spins should grow linearly with $|\delta B|$ in the $2a$ state while it should be strongly suppressed in the $2b$ state for temperatures below $\Delta_s$, even for $\delta B\ne 0$. In addition, the Zeeman splitting should act as a pair breaker and decrease the spin gap $\Delta_s$. We therefore expect the $2b$ state to be destroyed when the Zeeman energy exceeds the spin gap $\Delta_s$. Indeed, the triple point, where $\nu=0$,$1$, and $2$ coexist in Fig. 1, is pushed to lower fields as the Zeeman energy is increased by tilting the magnetic field [@Jiang1], The point contact conductance of the $2b$ state will be very different than that of a $2a$ state. The $2a$ state has 2 Fermi-liquid edge modes with nonzero conductance at zero temperature [@Wen]. In contrast, the $2b$ state has only a single $\nu'=1/2$ edge mode of charge $e^*$ carriers. For sufficiently low temperatures, the point contact tunneling will be non-Fermi-liquid like and the conductance will vanish as $T^2$ at very low temperatures [@Wen; @qhe-tunneling]. In addition, the scaling function for the resonance line shape, in the limit of small temperature, is known exactly for $\nu'=1/2$. In the limit of strong backscattering (large gate voltage), the conductance is dominated by single electron tunneling and hence is thermally activated because of the gap $\Delta_s$ to single particle excitations. This should provide an experimental measure of $\Delta_s$. In future work, we will explore in detail the energetic properties of the proposed paired wavefunctions for finite-size systems. We are grateful to H. W. Jiang for sharing his results with us prior to publication. We have also benefited from many useful conversations with N. E. Bonesteel, A. H. Castro-Neto, M.P.A. Fisher, D. Morse, S. L. Sondhi, and A. Zee. This work has been supported by the National Science Foundation under grant PHY94-07194 at the Institute for Theoretical Physics and grant DMR93-12606 at UCLA. Permanent Address: Dept. of Physics, University of\ California, Los Angeles, CA 90024. R. Prange and S. M. Girvin, [*The Quantum Hall Effect*]{} (Springer-Verlag, New York, 1987). L. W. Wong [*et. al.*]{}, Phys. Rev. B[**52**]{}, 18033 (1995); H.W. Jiang, [*unpublished*]{}. H. W. Jiang [*et. al.*]{}, Phys. Rev. Lett. [**71**]{}, 1439 (1993); T. Wang [*et. al.*]{}, Phys. Rev. Lett. [**72**]{}, 709 (1994); R. J. F. Hughes [*et. al.*]{}, J. Physics: Cond. Mat. [**6**]{}, 4763 (1994). I. Herbut, Phys. Rev. B[**46**]{}, 15582 (1992). M. Greiter [*et. al.*]{}, Nuc. Phys. B[**374**]{}, 567 (1992). G. Moore and N. Read, Nuc. Phys. B[**360**]{}, 362 (1991). S. L. Sondhi [*et. al.*]{}, Phys. Rev. B[**47**]{}, 16419 (1993). S. C. Zhang [*et. al.*]{}, Phys. Rev. Lett. [**62**]{}, 82 (1989); A. L. Fetter [*et. al.*]{}, Phys. Rev. B[**39**]{}, 9679 (1989); A. Lopez and E. Fradkin, Phys. Rev. B[**47**]{}, 7080 (1993). D. H. Lee and C. L. Kane, Phys. Rev. Lett. [**64**]{}, 1313 (1990). B. I. Halperin [*et. al.*]{}, Phys. Rev. B[**47**]{}, 7312 (1993). N. E. Bonesteel, Phys. Rev. B[**48**]{}, 11484 (1993); N. E. Bonesteel, [*in preparation*]{}. The details of this calculation are discussed in ref. [@Bonesteel] as well as A. M. Tikofsky, [*unpublished*]{}. M. J. Rice and Y. R. Wang, Phys. Rev. B[**37**]{}, 5893 (1988). X. G. Wen and A. Zee, Phys. Rev. B[**47**]{}, 2265 (1993). S. C. Zhang [*et. al.*]{}, Phys. Rev. Lett. [**69**]{}, 1252 (1992). M. Greiter [*et. al.*]{}, Phys. Rev. B [**46**]{}, 9586 (1992). S. Kivelson [*et. al.*]{}, Phys. Rev. B[**46**]{}, 2223 (1992). X. G. Wen, Phys Rev. B[**44**]{}, 5708 (1991). C. L. Kane and M. P. A. Fisher, Phys. Rev. B[**46**]{}, 15233 (1992); [*ibid.*]{}, Phys. Rev. B[**51**]{}, 13449 (1995). D. Shahar [*et. al.*]{}, Phys. Rev. Lett. [**74**]{}, 4511 (1995). S. E. Barrett [*et. al.*]{}, Phys. Rev. Lett. [**74**]{}, 5112 (1995).
{ "pile_set_name": "ArXiv" }
--- author: - | David [Coupier]{}\ [Laboratoire Paul Painlevé U.M.R. CNRS 8524]{}\ [Université Lille 1, France]{}\ [e-mail : [email protected]]{}\ David [Dereudre]{}\ [Laboratoire Paul Painlevé U.M.R. CNRS 8524]{}\ [Université Lille 1, France]{}\ [e-mail : [email protected]]{} bibliography: - 'biblioGibbs.bib' title: '[**Continuum Percolation for Quermass Model** ]{}' --- **Abstract** The continuum percolation for Markov (or Gibbs) germ-grain models is investigated. The grains are assumed circular with random radii on a compact support. The morphological interaction is the so-called Quermass interaction defined by a linear combination of the classical Minkowski functionals (area, perimeter and Euler-Poincaré characteristic). We show that the percolation occurs for any coefficient of this linear combination and for a large enough activity parameter. An application to the phase transition of the multi-type Quermass model is given. KEY-WORDS: Stochastic geometry, Gibbs point process, germ-grain model, Quermass interaction, percolation, phase transition. Introduction ============ The *germ-grain model* is built by unifying random convex sets– *the grains* –centered at the points– *the germs* –of a spatial point process. It is used for modelling random surfaces and interfaces, geometrical structures growing from germs, etc. For such models, the *continuum percolation* refers mainly to the existence of an unbounded connected component. This phenomenon expresses some macroscopic properties of materials as permeability, conductivity, etc. Moreover, it turns out to be an efficient tool to exhibit phase transition in Statistical Mechanics [@CCK; @GH]. For theses reasons, the continuum percolation has been abundantly studied since the eighties and the pioneer paper of Hall [@Hall]. When the grains are independent and identically distributed, and the germs are given by the locations of a Poisson point process (PPP), the germ-grain model is known as the *Boolean model*. In this context, the continuum percolation is well understood; see the book of Meester and Roy [@MR] for a very complete reference. One of the first results is the existence of a percolation threshold $z^*$ for the intensity parameter $z$ of the stationary PPP: provided the mean volume of the grain is finite, percolation occurs for $z>z^*$ and not for $z<z^*$. Because of the independence properties of the PPP, the Boolean model is sometimes caricatural for the applications in Biology or Physics. Mecke and its coauthors [@LMW; @M] have mentioned the need of developing, via Markov or Gibbs process, an interacting germ-grain model in which the interaction would locally depend on the geometry of the set. For this purpose, let us cite the Widom-Rowlinson model [@WR], the area interaction process [@BV] and the morphological model [@M]. Thus, Kendall, Van Lieshout and Baddeley suggested in [@KVB] a generalization of the previous models, called the *Quermass Interaction Process*. In this model, the formal Hamiltonian is a linear combination of the fundamental Minkowski functionals, namely in ${\ensuremath{{\mathbb{R}^2}}}$ the area ${\ensuremath{{\mathcal{A}}}}$, the perimeter ${\ensuremath{{\mathcal{L}}}}$ and the Euler-Poincaré characteristic ${\ensuremath{\chi}}$: $$H = \theta_1 {\ensuremath{{\mathcal{A}}}}+ \theta_2 {\ensuremath{{\mathcal{L}}}}+ \theta_3 {\ensuremath{\chi}}~.$$ The existence of infinite volume Gibbs point processes for the Hamiltonian $H$ has been recently proved in [@Der09]. This paper focuses on the continuum percolation for such processes.\ The existence of a percolation threshold $z^*$ for the Boolean model relies on a basic (but essential) monotonicity argument: see [@MR], Chapter 2.2. This argument fails in the case of Gibbs point processes with Hamiltonian $H$. So, no percolation threshold can be expected in our context. However, other stochastic arguments as stochastic domination or FKG lead to percolation results. In [@CCK], Chayes et al prove that percolation occurs for $z$ large enough and $\theta_2=\theta_3=0$. To our knowledge, the percolation phenomenon for other values of parameters $\theta_1,\theta_2,\theta_3$ has not been investigated yet. Our main result (Theorem \[mainTH\]) states that, for any $\theta_1,\theta_2,\theta_3$ (positive or negative), percolation occurs with probability $1$ for $z$ large enough. The only assumption bears on the random radii of the circular grains: they have to belong to a compact set not containing $0$. The proof of this theorem is relatively easy in the case $\theta_3=0$. Indeed, the *local energy* $h((x,R),{\ensuremath{\omega}})$– the energy variation when the grain $\bar B(x,R)$ is added to the configuration ${\ensuremath{\omega}}$ – is uniformly bounded (see Lemma \[Borneenergie\]) and by classical stochastic comparison with respect to the Poisson Process the result follows. So the main challenge of the present paper is the proof of Theorem \[mainTH\] when $\theta_3\neq 0$. In this setting, the local energy becomes unbounded from above and below and classical stochastic comparison arguments for point processes fail. In interpreting the percolation in our model via a site percolation model (see the beginning of Section \[SectionProofMainTH\]), we prove the result thanks to a stochastic domination result for graphs due to Liggett et al (Theorem 1.3 in [@LSS] or Lemma \[StochDomLiggett\] below). An arduous control of the hole number variation, when a new grain is added, is the main technical issue. We prove essentially that this variation is moderate for a large enough set of admissible locations of grains. Let us mention that our proof is inspired by the one of Proposition 3.1 in [@GH]. Following [@CCK; @GH], we use our percolation result (Theorem \[mainTH\]) to exhibit a phase transition phenomenon for Quermass model with several type of particles (Theorem \[theo:transition\]).\ Our paper is organized as follows. In Section \[QuermassModelsection\], the Quermass model and the main notations are introduced. The local energy $h((x,R),{\ensuremath{\omega}})$ is defined in (\[localenergy\]). Section \[Resultssection\] contains the results of the paper. Section \[stochasticargument\] is devoted to the case $\theta_3=0$ and Section \[sectiontransition\] to the phase transition result. The proof of Theorem \[mainTH\] is developed in Section \[SectionProofMainTH\]. Quermass Model {#QuermassModelsection} ============== Notations --------- We denote by ${\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$ the set of bounded Borel sets in ${\ensuremath{{\mathbb{R}^2}}}$ with a positive Lebesgue measure. For any ${\ensuremath{\Lambda}}$ and $\Delta$ in ${\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$, ${\ensuremath{\Lambda}}\oplus\Delta$ stands for the Minkoswki sum of these sets. Let $0\le R_0\le R_1$ be some positive reals and ${\ensuremath{{\mathcal{E}}}}$ be the product space ${\ensuremath{{\mathbb{R}^2}}}\times [R_0,R_1]$ endowed with its natural Euclidean Borel $\sigma$-algebra $\sigma({\ensuremath{{\mathcal{E}}}})$. For any ${\ensuremath{\Lambda}}\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$, ${\ensuremath{{\mathcal{E}}}}_{\ensuremath{\Lambda}}$ denotes the space ${\ensuremath{\Lambda}}\times [R_0,R_1]$. A [*configuration*]{} ${\ensuremath{\omega}}$ is a subset of ${\ensuremath{{\mathcal{E}}}}$ which is locally finite with respect to its first coordinate: $\#({\ensuremath{\omega}}\cap{\ensuremath{{\mathcal{E}}}}_{\ensuremath{\Lambda}})$ is finite for any ${\ensuremath{\Lambda}}$ in ${\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$. The configuration set ${\ensuremath{\Omega}}$ is endowed with the $\sigma$-algebra ${\ensuremath{{\mathcal{F}}}}$ generated by the functions ${\ensuremath{\omega}}\mapsto\#({\ensuremath{\omega}}\cap A)$ for any $A$ in $\sigma({\ensuremath{{\mathcal{E}}}})$.\ We will merely denote by ${\ensuremath{\omega}}_{\ensuremath{\Lambda}}$ instead of ${\ensuremath{\omega}}\cap{\ensuremath{{\mathcal{E}}}}_{\ensuremath{\Lambda}}$ the restriction of the configuration ${\ensuremath{\omega}}$ (with respect to its first coordinate) to ${\ensuremath{\Lambda}}$. Moreover, for any $(x,R)$ in ${\ensuremath{{\mathcal{E}}}}$, we will write ${\ensuremath{\omega}}\cup (x,R)$ instead of ${\ensuremath{\omega}}\cup\{(x,R)\}$. A configuration ${\ensuremath{\omega}}\in{\ensuremath{\Omega}}$ can be interpreted as a marked configuration on ${\ensuremath{{\mathbb{R}^2}}}$ with marks in $[R_0,R_1]$. To each $(x,R)\in{\ensuremath{\omega}}$ is associated the closed ball $\bar B(x,R)$ (the grain) centered at $x$ (the germ) with radius $R$. The germ-grain surface $\bar{\ensuremath{\omega}}$ is defined as $$\bar{\ensuremath{\omega}}= \bigcup_{(x,R)\in{\ensuremath{\omega}}} \bar B(x,R) ~.$$ Quermass interaction -------------------- Let us define the Quermass interaction as in Kendall et al. [@KVB]. The energy (or Hamiltonian) of a finite configuration ${\ensuremath{\omega}}$ in ${\ensuremath{\Omega}}$ is defined by $$\label{energy} H({\ensuremath{\omega}})= {\ensuremath{\theta}}_1 {\ensuremath{{\mathcal{A}}}}(\bar {\ensuremath{\omega}}) + {\ensuremath{\theta}}_2 {\ensuremath{{\mathcal{L}}}}(\bar {\ensuremath{\omega}}) +{\ensuremath{\theta}}_3 {\ensuremath{\chi}}(\bar{\ensuremath{\omega}}) ~,$$ where ${\ensuremath{\theta}}_1$, ${\ensuremath{\theta}}_2$ and ${\ensuremath{\theta}}_3$ are three real numbers, and ${\ensuremath{{\mathcal{A}}}}$, ${\ensuremath{{\mathcal{L}}}}$ and ${\ensuremath{\chi}}$ are the three fundamental Minkowski functionals, respectively area, perimeter and Euler-Poincaré characteristic. This last one is the difference between the number of connected components and the number of holes. Recall that a hole of $\bar{\ensuremath{\omega}}$ is a bounded connected component of $\bar{\ensuremath{\omega}}^c$. Hadwiger’s Theorem ensures that any functional $F$ defined on the space of finite unions of convex compact sets, which is continuous for the Hausdorff topology, invariant under isometric transformations and additive (i.e. $F(A\cup B)=F(A)+F(B)-F(A\cap B)$) can be decomposed as in (\[energy\]). This universal representation justifies the choice of the Quermass interaction for modelling mesoscopic random surfaces [@LMW; @M]. The energy inside ${\ensuremath{\Lambda}}\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$ of any given configuration ${\ensuremath{\omega}}$ in ${\ensuremath{\Omega}}$ (finite or not) is defined by $$\label{energyL} H_{\ensuremath{\Lambda}}({\ensuremath{\omega}}) = H({\ensuremath{\omega}}_{\Delta}) - H({\ensuremath{\omega}}_{\Delta\backslash {\ensuremath{\Lambda}}}) ~,$$ where $\Delta$ is any subset of ${\ensuremath{{\mathbb{R}^2}}}$ containing ${\ensuremath{\Lambda}}\oplus B(0,2R_1)$. By additivity of functionals ${\ensuremath{{\mathcal{A}}}}$, ${\ensuremath{{\mathcal{L}}}}$ and ${\ensuremath{\chi}}$, the difference $H_{\ensuremath{\Lambda}}({\ensuremath{\omega}})$ does not depend on the chosen set $\Delta$. Let us end with defining the local energy $h((x,R),{\ensuremath{\omega}})$ of the marked point $(x,R)\in{\ensuremath{{\mathcal{E}}}}$ (or of the associated ball $\bar B(x,R)$) with respect to the configuration ${\ensuremath{\omega}}$: $$\label{localenergy} h((x,R),{\ensuremath{\omega}}) = H_{\ensuremath{\Lambda}}({\ensuremath{\omega}}\cup(x,R)) - H_{\ensuremath{\Lambda}}({\ensuremath{\omega}}) ~,$$ for any ${\ensuremath{\Lambda}}\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$ containing $x$. Remark this definition does not depend on the choice of the set ${\ensuremath{\Lambda}}$. The local energy $h((x,R),{\ensuremath{\omega}})$ represents the energy variation when the ball $\bar B(x,R)$ is added to the configuration ${\ensuremath{\omega}}$. The Gibbs property ------------------ Let $Q$ be a reference probability measure on $[R_0,R_1]$. Without loss of generality, $R_0$ and $R_1$ can be chosen such that, for every ${\ensuremath{\varepsilon}}>0$, $$\label{optimalQ} Q([R_0+{\ensuremath{\varepsilon}},R_1]) < 1 \; \text{ and } \; Q([R_0,R_1-{\ensuremath{\varepsilon}}]) < 1 ~.$$ Let $z>0$. Let us denote by ${\ensuremath{\lambda}}$ the Lebesgue measure on ${\ensuremath{{\mathbb{R}^2}}}$ and by $\pi^z$ the PPP on ${\ensuremath{{\mathcal{E}}}}$ with intensity measure $z{\ensuremath{\lambda}}\otimes Q$. Under $\pi^z$, the law of the random surface $\bar {\ensuremath{\omega}}$ is the stationary boolean model with intensity $z>0$ and distribution of radius $Q$. Finally, for any ${\ensuremath{\Lambda}}\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$, let us denote by $\pi^z_{\ensuremath{\Lambda}}$ the PPP on ${\ensuremath{{\mathcal{E}}}}_{\ensuremath{\Lambda}}$ with intensity measure $z{\ensuremath{\lambda}}_{\ensuremath{\Lambda}}\otimes Q$, where ${\ensuremath{\lambda}}_{\ensuremath{\Lambda}}$ is the restriction of the Lebesgue measure ${\ensuremath{\lambda}}$ to ${\ensuremath{\Lambda}}$. \[Gibbs\] A probability measure $P$ on ${\ensuremath{\Omega}}$ is a Quermass Process for the intensity $z>0$ and the parameters $\theta_1, \theta_2, \theta_3$ if $P$ is stationary and if for every ${\ensuremath{\Lambda}}$ in ${\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$, for every bounded positive measurable function $f$ from ${\ensuremath{\Omega}}$ to ${\ensuremath{{\mathbb{R}}}}$, $$\label{DLR} \int f({\ensuremath{\omega}}) P(d{\ensuremath{\omega}}) = \int\int f({\ensuremath{\omega}}'_{\ensuremath{\Lambda}}\cup {\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c}) \frac{1}{Z_{\ensuremath{\Lambda}}({\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c})} e^{-H_{\ensuremath{\Lambda}}({\ensuremath{\omega}}'_{{\ensuremath{\Lambda}}}\cup{\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c})}\pi^z_{\ensuremath{\Lambda}}(d{\ensuremath{\omega}}'_{\ensuremath{\Lambda}}) P(d{\ensuremath{\omega}}) ~,$$ where $Z_{\ensuremath{\Lambda}}({\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c})$ is the partition function $$Z_{\ensuremath{\Lambda}}({\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c}) = \int\int e^{-H_{\ensuremath{\Lambda}}({\ensuremath{\omega}}'_{{\ensuremath{\Lambda}}}\cup{\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c})}\pi^z_{\ensuremath{\Lambda}}(d{\ensuremath{\omega}}'_{\ensuremath{\Lambda}}) ~.$$ The equations (\[DLR\])– for all ${\ensuremath{\Lambda}}\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$ –are called DLR for Dobrushin, Landford and Ruelle. They are equivalent to: for any ${\ensuremath{\Lambda}}\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$, the law of ${\ensuremath{\omega}}_{\ensuremath{\Lambda}}$ under $P$ given ${\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c}$ is absolutely continuous with respect to the Poisson Process $\pi^z_{\ensuremath{\Lambda}}$ with the local density $$\label{localdensity} g_{\ensuremath{\Lambda}}({\ensuremath{\omega}}'_{{\ensuremath{\Lambda}}}|{\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c}) = \frac{1}{Z_{\ensuremath{\Lambda}}({\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c})} e^{-H_{\ensuremath{\Lambda}}({\ensuremath{\omega}}'_{{\ensuremath{\Lambda}}}\cup{\ensuremath{\omega}}_{{\ensuremath{\Lambda}}^c})} ~.$$ See [@P] for a general presentation of Gibbs measures and DLR equations. The existence, the uniqueness or non-uniqueness (phase transition) of Quermass processes are difficult problems in statistical mechanics. The existence has been proved recently in [@Der09], Theorem 2.1 for any parameters $z>0$ and $\theta_1, \theta_2, \theta_3$ in ${\ensuremath{{\mathbb{R}}}}$ . A phase transition result is proved in [@CCK; @GLM; @WR] for $R_0=R_1$, $\theta_2=\theta_3=0$ and for $\theta_1=z$ large enough. Results {#Resultssection} ======= Percolation occurs ------------------ We say that *percolation occurs* for a given configuration ${\ensuremath{\omega}}\in{\ensuremath{\Omega}}$ if the subset $\bar{\ensuremath{\omega}}$ of ${\ensuremath{{\mathbb{R}^2}}}$ contains at least one unbounded connected component. The set of configurations such that percolation occurs is a translation invariant event. Its probability, called *the percolation probability*, equals to $0$ or $1$ for any ergodic Quermass process. However, the Quermass processes are not necessarily ergodic (they are only stationary) and their percolation probabilities may be different from $0$ and $1$. Besides, in [@CCK], Chayes et al have built two Quermass processes, both corresponding to $\theta_2=\theta_3=0$ and $\theta_1=z$ large enough, whose percolation probabilities respectively equal to $0$ and $1$. Since any mixture of these two processes is still a Quermass process, the authors obtain Quermass processes whose percolation probabilities equal to any value between $0$ and $1$.\ Our main result states that percolation occurs with probability $1$ for any (ergodic or not) Quermass process whenever the intensity $z$ is large enough. \[mainTH\] Let $R_0>0$ and $\theta_1,\theta_2,\theta_3\in{\ensuremath{{\mathbb{R}}}}$. There exists $z^{\ast}>0$ such that for any Quermass process $P$ associated to the parameters $\theta_1,\theta_2,\theta_3$ and $z>z^{\ast}$, percolation occurs $P$-almost surely. The proof of Theorem \[mainTH\] is based on a discretization argument which allows to reduce the percolation problem from the (continuum) Quermass model to a site percolation model on the lattice ${\ensuremath{{\mathbb{Z}}}}^{2}$ (up to a scale factor). This proof is rather long and technical so it is addressed in Section \[SectionProofMainTH\].\ Let us point out here that our theorem does not claim $z^{\ast}$ is a percolation threshold. In other words, for $z<z^{\ast}$, the percolation may be lost and recovered on different successive ranges.\ Another natural question involves the number of unbounded connected components. Following the classical arguments for continuum percolation, we prove that this number is almost surely equal to zero or one. \[numberInfCC\] For any Quermass process $P$ the number of unbounded connected component is a random variable in $\{0,1\}$. It is well-known that any Gibbs measure is a mixture of extremal ergodic Gibbs measures. For each ergodic Quermass process $P$, the number of connected component is almost surely a constant in ${\ensuremath{{\mathbb{N}}}}\cup\{+\infty\}$. For any ${\ensuremath{\Lambda}}\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$, thanks to the DLR equations (\[DLR\]), it is easy to prove that the law of ${\ensuremath{\omega}}_{\ensuremath{\Lambda}}$ under $P$ is equivalent to $\pi^{z}_{\ensuremath{\Lambda}}$. Therefore, in following the general scheme of the proof of Theorem 2.1 in [@MR], we show that the number of connected components is necessary $0$ or $1$. Percolation when $\theta_3=0$ {#stochasticargument} ----------------------------- In the particular case $\theta_3=0$, Theorem \[mainTH\] can be completed and proved in a simple way. First, let us recall the definitions involving the stochastic domination for point processes. We follow the notations given in [@GK]. An event $A$ in ${\ensuremath{{\mathcal{F}}}}$ is called increasing if for every ${\ensuremath{\omega}}\in A$ and any ${\ensuremath{\omega}}'\in{\ensuremath{\Omega}}$ containing ${\ensuremath{\omega}}$ then ${\ensuremath{\omega}}'\in A$ too. Let $P$ and $P'$ be two probability measures on ${\ensuremath{\Omega}}$. We say that $P$ is dominated by $P'$, denoted by $P\preceq P'$, if for every increasing event $A\in{\ensuremath{{\mathcal{F}}}}$, $P(A)\le P'(A)$. In this section, we focus our attention on the increasing event “there exists an unbounded connected component”. Let $P$ be any Quermass process and assume $\theta_3=0$ and $R_0>0$. Thanks to Lemma \[Borneenergie\], the local energy can be uniformly bound: there exist constants $C_0$ and $C_1$ such that for any $(x,R)\in{\ensuremath{{\mathcal{E}}}}$ and ${\ensuremath{\omega}}\in{\ensuremath{\Omega}}$, $$\label{borneenergie} C_0 \le h((x,R),{\ensuremath{\omega}}) \le C_1 ~.$$ Combining (\[borneenergie\]) and Theorem 1.1 in [@GK], we get the following stochastic dominations: $$\pi^{ze^{-C_1}} \preceq P \preceq \pi^{ze^{-C_0}} ~.$$ Now, the (stationary) Boolean models corresponding to $\pi^{ze^{-C_1}}$ and $\pi^{ze^{-C_0}}$ admit positive and finite percolation thresholds (see [@MRbook], Chapter 3). It follows : \[Propdomination\] Let $R_0>0$. For every $\theta_1, \theta_2$ in ${\ensuremath{{\mathbb{R}}}}$, there exist constants $z_0,z_1$ such that for any Quermas Process $P$ associated to parameters $z, \theta_1, \theta_2$ and $\theta_3=0$, the percolation occurs $P$-almost surely if $z>z_1$ and does not occur $P$-almost surely if $z<z_0$. Proposition \[Propdomination\] improves Theorem \[mainTH\] in the case $\theta_3=0$ since it ensures the existence of a subcritical regime. It is worth pointing out here that the uniform bounds in (\[borneenergie\]) do not hold whenever $\theta_3\neq 0$. Precisely, this is the hole number variation which cannot be uniformly bounded. Phase transition for multi-type Quermass Process {#sectiontransition} ------------------------------------------------ In this section, the multi-type Quermass model is introduced and a phase transition is exhibited, i.e. the existence of several Gibbs processes for the same parameters is proved. Let $K$ be a positive integer. The $K$-type Quermass model is defined on the space $\Omega_K$ of configurations in ${\ensuremath{{\mathcal{E}}}}_K={\ensuremath{{\mathbb{R}^2}}}\times [R_0,R_1]\times\{1,2,\ldots,K\}$. Each disc is now marked by a number specifying its type. We don’t give the natural extension of the notations involving the sigma-field and so on.\ The following Quermass energy function is defined such that all discs of a connected component have the same number. This is a non-overlapping multi-type germ-grain model. Precisely the energy of a finite configuration ${\ensuremath{\omega}}$ is now given by $$\label{energymc} H({\ensuremath{\omega}})= {\ensuremath{\theta}}_1 {\ensuremath{{\mathcal{A}}}}(\bar {\ensuremath{\omega}}) + {\ensuremath{\theta}}_2 {\ensuremath{{\mathcal{L}}}}(\bar {\ensuremath{\omega}}) +{\ensuremath{\theta}}_3 {\ensuremath{\chi}}(\bar{\ensuremath{\omega}})+ \sum_{ \begin{subarray}{c} (x,R,i),(y,R',j)\in{\ensuremath{\omega}}\\ i\neq j \end{subarray} } \phi(|x-y|-R-R') ~,$$ where $\phi$ is an hardcore potential equals to infinity on $]-\infty,0]$ and zero on $]0,+\infty]$. The energy inside ${\ensuremath{\Lambda}}\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$ of any finite or infinite configuration ${\ensuremath{\omega}}$ is defined as in (\[energyL\]) with the convention $+\infty-\infty=+\infty$. The definition of the $K$-type Quermass process via the DLR equations follows as in Definition \[Gibbs\]. The proof of the existence of such processes is similar to the one of the existence of Quermass process. See Theorem 2.1 of [@Der09] for more details. Here is our phase transition result: \[theo:transition\] Let $R_0>0$. For any $\theta_1$, $\theta_2$ and $\theta_3$ in ${\ensuremath{{\mathbb{R}}}}$, there exists $z_0>0$ such that, for any $z>z_0$, there exist several $K$-type Quermass Processes. The phase transition occurs. The proof essentially follows the scheme of the one of Theorem 2.2 of [@CCK] or Theorem 1.1 of [@GH]. It is based on a random-cluster representation (or Gray Representation) analogous to the Fortuin-Kasteleyn representation of the Potts model. The existence of an unbounded connected component allows to prove the existence of a $K$-type Quermass process in which the density of particles of a given type is larger than the ones of the other types. By symmetry of the types, we prove the existence of at least $K$ different $K$-type Quermass processes. Proof of Theorem \[mainTH\] {#SectionProofMainTH} =========================== General scheme -------------- In the following, $P$ denotes a stationary Quermass process on ${\ensuremath{\Omega}}$ associated to the intensity $z>0$ and the parameters $\theta_1,\theta_2,\theta_3\in{\ensuremath{{\mathbb{R}}}}$.\ Let $\ell$ be a real number such that $\ell>2R_{1}+2R_{0}$. Let us define the diamond box $\Delta$ as the interior of the convex hull of the eight points $(3\ell,0)$, $(6\ell,0)$, $(9\ell,3\ell)$, $(9\ell,6\ell)$, $(6\ell,9\ell)$, $(3\ell,9\ell)$, $(0,6\ell)$ and $(0,3\ell)$. This large octagon contains four smaller boxes $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ and $B_{\texttt{W}}$ with side length $\ell$; precisely $B_{\texttt{N}}=(4\ell,7\ell)+[0,\ell]^{2}$, $B_{\texttt{S}}=(4\ell,\ell)+[0,\ell]^{2}$, $B_{\texttt{E}}=(7\ell,4\ell)+[0,\ell]^{2}$ and $B_{\texttt{W}}=(\ell,4\ell)+[0,\ell]^{2}$. The subscripts $\texttt{N}$, $\texttt{S}$, $\texttt{E}$ and $\texttt{W}$ refer to the cardinal directions. See Figure \[fig:diamondbox\]. Thus, let us introduce the indicator function $\xi$ defined on ${\ensuremath{\Omega}}$ and equal to $1$ if and only if the two following conditions are satisfied: - Each box $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ and $B_{\texttt{W}}$, contains at least one point of ${{\ensuremath{\omega}}}_{\Delta}$; - The number $N_{cc}^{\Delta}({\ensuremath{\omega}})$ of connected components of $\bar{{\ensuremath{\omega}}}_{\Delta}$ having at least one ball centered in one of the boxes $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ or $B_{\texttt{W}}$, is equal to $1$. In other words, $\xi({\ensuremath{\omega}})=1$ means the boxes $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ and $B_{\texttt{W}}$ are connected through $\bar{{\ensuremath{\omega}}}_{\Delta}$. ![\[fig:diamondbox\] ](diamondbox.eps){width="7cm" height="7cm"} For any $x\in(6\ell\,{\ensuremath{{\mathbb{Z}}}})^{2}$, let $\tau_{x}$ be the translation operator on the configuration set ${\ensuremath{{\mathcal{E}}}}$ defined by $(y,R)\in\tau_{x}{\ensuremath{\omega}}$ if and only if $(y+x,R)\in{\ensuremath{\omega}}$. Hence, we can define the translated indicator function $\xi_{x}$ of $\xi$ on the translated box $\Delta_{x}=x+\Delta$ by $\xi_{x}({\ensuremath{\omega}})=\xi(\tau_{x}{\ensuremath{\omega}})$. Let us remark that $\xi_{x}({\ensuremath{\omega}})$ only depends on the restriction of the configuration ${\ensuremath{\omega}}$ to the box $\Delta_{x}$. Moreover, thanks to the stationary character of the Quermass process $P$, the random variables $\xi_{x}$, $x\in(6\ell\,{\ensuremath{{\mathbb{Z}}}})^{2}$, are identically distributed. They are dependent too.\ Let us consider $x,y\in(6\ell\,{\ensuremath{{\mathbb{Z}}}})^{2}$ such that $y=(6\ell,0)+x$. The boxes $\Delta_{x}$ and $\Delta_{y}$ have in common a cardinal box, i.e. $x+B_{\texttt{E}}=y+B_{\texttt{W}}$. So, the condition $\xi_{x}({\ensuremath{\omega}})=\xi_{y}({\ensuremath{\omega}})=1$ ensures that the cardinal boxes of $\Delta_{x}$ and $\Delta_{y}$ are connected together through the restriction of $\bar{{\ensuremath{\omega}}}$ to $\Delta_{x}\cup\Delta_{y}$. The same is true when $y=(0,6\ell)+x$. This induces a graph structure on the vertex set $V=(6\ell\,{\ensuremath{{\mathbb{Z}}}})^{2}$: for any $x,y\in V$, $\{x,y\}$ belongs to the edge set $E$ if and only if $$y-x \in \{ \pm(6\ell,0), \pm(0,6\ell) \} ~.$$ The graph $(V,E)$ is merely the square lattice ${\ensuremath{{\mathbb{Z}}}}^{2}$ with the scale factor $6\ell$. The family $\{\xi_{x},x\in V\}$ provides a site percolation process on the graph $(V,E)$. It has been built so as to satisfy the following statement. \[siteperco\] Let ${\ensuremath{\omega}}\in{\ensuremath{\Omega}}$ such that percolation occurs in the site percolation process $\{\xi_{x},x\in V\}$. Then, so does for ${\ensuremath{\omega}}$. Let $\Pi_{p}$ be the Bernoulli (with parameter $p$) product measure on $\{0,1\}^{V}$. A stochastic domination result of Liggett et al [@LSS] (Theorem 1.3) allows to compare the site percolation processes induced by the family $\{\xi_{x},x\in V\}$ and $\Pi_{p}$. Here is an adaptated version to our context. Basic definitions about stochastic domination for lattice state spaces are not recall here. They are similar to the ones presented in Section \[stochasticargument\] for point processes. See also [@Grim]. \[StochDomLiggett\] Let $p\in[0,1]$. Assume that, for any vertex $x\in V$, $$\label{UnifEspCond} P \left( \xi_{x}=1 \,|\, \xi_{y} : \{x,y\}\notin E \right) \geq p \; \mbox{ a.s.}$$ Then the distribution of the family $\{\xi_{x},x\in V\}$ stochastically dominates the probability measure $\Pi_{f(p)}$, where $f:[0,1]\to[0,1]$ is a deterministic function such that $f(p)$ tends to $1$ as $p$ tends to $1$. Actually, Theorem \[mainTH\] straight derives from Lemmas \[siteperco\] and \[StochDomLiggett\]. Let us first recall that in the site percolation model on the graph $(V,E)$, there exists a threshold value $p^{\ast}<1$ such that percolation occurs with $\Pi_{p}$-probability $1$ whenever $p>p^{\ast}$. See the book of Grimmett [@Grim], p. 25. So, let $p$ be a real number in $[0,1]$ such that $$\label{f(p)} f(p) > p^{\ast} ~.$$ Whenever the Quermass process $P$ satisfies (\[UnifEspCond\]) for that $p$, then combining Lemmas \[siteperco\] and \[StochDomLiggett\] percolation occurs $P$-a.s. Therefore it remains to show that for any $p>0$, hypothesis (\[UnifEspCond\]) holds for $z$ large enough.\ The next result claims that each Borel set of ${\ensuremath{{\mathbb{R}^2}}}$, sufficiently thick in some sense, contains at least one element of the configuration ${\ensuremath{\omega}}$ with a probability tending to $1$ as the intensity $z$ tends to infinity. It will be proved at the end of this section. \[LemmeNotempty\] Let $V\subset{\ensuremath{{\mathbb{R}^2}}}$ such that there exist $U\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$ with positive Lebesgue measure and ${\ensuremath{\varepsilon}}>0$ satisfying $U\oplus\bar{B}(0,R_{1}+R_{0}+{\ensuremath{\varepsilon}})\subset V$. Then there exists a constant $C>0$, depending on ${\ensuremath{\lambda}}(U)$ and ${\ensuremath{\varepsilon}}$, such that for any configuration ${\ensuremath{\omega}}\in{\ensuremath{\Omega}}$ and for any $z>0$, $$P \left( {\ensuremath{\omega}}_{V} = \emptyset \,|\, {\ensuremath{\omega}}_{V^{c}} \right) \leq C z^{-1} ~.$$ Since the Quermass process $P$ is stationary, it is sufficient to prove (\[UnifEspCond\]) with $x=(0,0)$. So, we focus our attention on the diamond box $\Delta=\Delta_{(0,0)}$ and use Lemma \[LemmeNotempty\] to check that condition (C1) is fulfilled in this box. Since $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ and $B_{\texttt{W}}$ are sufficiently thick (with side length $\ell>2R_{1}+2R_{0}$), it follows $$P \left( {\ensuremath{\omega}}_{B_{i}} = \emptyset \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) = P \left( P \left( {\ensuremath{\omega}}_{B_{i}} = \emptyset \,|\, {\ensuremath{\omega}}_{B_{i}^{c}} \right) \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) \leq C z^{-1} ~,$$ for any $i\in\{\texttt{N},\texttt{S},\texttt{E},\texttt{W}\}$. So the conditional probability that ${\ensuremath{\omega}}$ satisfies (C1) is larger than $1-4Cz^{-1}$.\ The equation $N_{cc}^{\Delta}({\ensuremath{\omega}})=0$ forces the box $B_{\texttt{N}}$ (for instance) to be empty of points of the configuration ${\ensuremath{\omega}}$. Hence, $$P \left( N_{cc}^{\Delta}({\ensuremath{\omega}}) = 0 \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) \leq C z^{-1} ~.$$ Checking that condition (C2) is fulfilled in the diamond box $\Delta$ needs what we call the Connection Lemma (Lemma \[ConnectionLemma\]). This result states the conditional probability that $N_{cc}^{\Delta}({\ensuremath{\omega}})$ is larger than $2$ converges to $0$ uniformly on the configuration outside $\Delta$. This is the heart of the proof of Theorem \[mainTH\]. Its technical proof is given in Section \[SectionConnectLemma\]. \[ConnectionLemma\] There exists a constant $C'>0$ such that for any configuration ${\ensuremath{\omega}}\in{\ensuremath{\Omega}}$ and for any $z>0$, $$\label{Ncc>=2} P \left( N_{cc}^{\Delta}({\ensuremath{\omega}}) \geq 2 \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) \leq C' z^{-1} ~.$$ The above inequalities and the Connection Lemma imply that conditions (C1) and (C2) are fulfilled in $\Delta$ with a probability tending to $1$ as $z$ tends to $\infty$: $$P \left( \xi_{(0,0)}({\ensuremath{\omega}}) = 1 \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) \geq 1 - (5C+C') z^{-1} ~.$$ The hypothesis (\[UnifEspCond\]) then follows. Let $x$ be a vertex of the graph $(V,E)$ which is not a neighbor of $(0,0)$. By construction, the box $\Delta_{x}$ is included in $\Delta^{c}=\Delta_{(0,0)}^{c}$ (since $\Delta$ is an open set). This means the random variable $\xi_{x}$ is measurable with respect to the $\sigma$-algebra induced by the configurations restricted to $\Delta_{(0,0)}^{c}$. So, $$P \left( \xi_{(0,0)} = 1 \,|\, \xi_{x} : \{(0,0),x\}\notin E \right) \geq 1 - (5C+C') z^{-1} ~,$$ and the hypothesis (\[UnifEspCond\]) holds with $x=(0,0)$ and any $p\in[0,1[$, provided the intensity $z$ is large enough. This ends the proof of Theorem \[mainTH\]. (Lemma \[LemmeNotempty\]) Let $U\in{\ensuremath{{\mathcal{B}}}}({\ensuremath{{\mathbb{R}^2}}})$ be a bounded Borel set with positive Lebesgue measure and $V\supset U\oplus\bar{B}(0,R_{1}+R_{0}+{\ensuremath{\varepsilon}})$. First, let us write: $$\begin{aligned} \label{probavide} P \left( {\ensuremath{\omega}}_{V} = \emptyset \,|\, {\ensuremath{\omega}}_{V^{c}} \right) & = & \frac{1}{Z_{V}({\ensuremath{\omega}}_{V^{c}})} \int_{{\ensuremath{\Omega}}_{V}} {\ensuremath{\mbox{\rm 1\kern-0.23em I}}}_{{\ensuremath{\omega}}_{V}=\emptyset} \, e^{-H_{V}({\ensuremath{\omega}}_{V}\cup{\ensuremath{\omega}}_{V^{c}})} \, \pi^{z}_{V}(d{\ensuremath{\omega}}_{V}) \nonumber \\ & = & \frac{e^{-z{\ensuremath{\lambda}}(V)}}{Z_{V}({\ensuremath{\omega}}_{V^{c}})} ~,\end{aligned}$$ since the empty configuration has a null energy, i.e. $H_{V}({\ensuremath{\omega}}_{V^{c}})=0$. A configuration ${\ensuremath{\omega}}$ whose restriction to $V$ satisfies $$\#{\ensuremath{\omega}}_{U\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]}=1 \; \mbox{ and } \; {\ensuremath{\omega}}_{V\setminus U}=\emptyset$$ is reduced to a ball $\bar{B}(x,R)$ centered at a $x$ in $U$ and with a radius $R_{0}<R<R_{0}+{\ensuremath{\varepsilon}}$. Since the ball $\bar{B}(x,R)$ does not overlap $\bar{{\ensuremath{\omega}}}_{V^{c}}$, its energy $H_{V}((x,R)\cup{\ensuremath{\omega}}_{V^{c}})$ is easy to compute; $$H_{V}((x,R)\cup{\ensuremath{\omega}}_{V^{c}}) = {\ensuremath{\theta}}_{1} 2\pi R + {\ensuremath{\theta}}_{2} \pi R^{2} + {\ensuremath{\theta}}_{3}$$ (it is not worth using inequalities of Lemma \[Borneenergie\] here). So, $H_{V}((x,R)\cup{\ensuremath{\omega}}_{V^{c}})$ is bounded by a positive constant $K$ only depending on parameters ${\ensuremath{\theta}}_{1}$, ${\ensuremath{\theta}}_{2}$, ${\ensuremath{\theta}}_{3}$ and radius $R_{1}$. Henceforth, $$\begin{aligned} \lefteqn{P \left( \#{\ensuremath{\omega}}_{U\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]} = 1 , \; {\ensuremath{\omega}}_{V\setminus U} = \emptyset \,|\, {\ensuremath{\omega}}_{V^{c}} \right) }\hspace*{2cm} \\ & & = \frac{1}{Z_{V}({\ensuremath{\omega}}_{V^{c}})} \int_{U\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]} e^{-H_{V}((x,R)\cup{\ensuremath{\omega}}_{V^{c}})} \, z e^{-z{\ensuremath{\lambda}}(V)} \, {\ensuremath{\lambda}}(dx) Q(dR) \\ & & \geq \frac{e^{-z{\ensuremath{\lambda}}(V)}}{Z_{V}({\ensuremath{\omega}}_{V^{c}})} z e^{-K} {\ensuremath{\lambda}}(U) Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}]) ~.\end{aligned}$$ Recall that $Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}])$ is positive by (\[optimalQ\]). Using the identity (\[probavide\]), we finally upperbound the conditional probability $P({\ensuremath{\omega}}_{V}=\emptyset | {\ensuremath{\omega}}_{V^{c}})$ by $$\left( z e^{-K} {\ensuremath{\lambda}}(U) Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}]) \right)^{-1} P \left( \#{\ensuremath{\omega}}_{U\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]} = 1 , \; {\ensuremath{\omega}}_{V\setminus U} = \emptyset \,|\, {\ensuremath{\omega}}_{V^{c}} \right) ~.$$ This proves Lemma \[LemmeNotempty\] with $C=(e^{-K} {\ensuremath{\lambda}}(U) Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}]))^{-1}$. Proof of the Connection Lemma {#SectionConnectLemma} ----------------------------- ### Outline Let us recall that $N_{cc}^{\Delta}({\ensuremath{\omega}})$ denotes the number of connected components of $\bar{{\ensuremath{\omega}}}_{\Delta}$ having at least one ball centered in one of the four cardinal boxes $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ or $B_{\texttt{W}}$. In this section, we assume $N_{cc}^{\Delta}({\ensuremath{\omega}})\geq 2$. Our strategy consists in exhibiting a subset $B$ of the diamond box $\Delta$ in which ${\ensuremath{\omega}}_{B}=\emptyset$. Moreover, for $x\in B$, if we are able to control uniformly the energy $H_{B}((x,R)\cup{\ensuremath{\omega}}_{B^{c}})$ on ${\ensuremath{\omega}}_{B^{c}}$, then the configuration ${\ensuremath{\omega}}$ should contain a point centered in $B$ with large probability as $z$ tends to infinity. This leads to the Connection Lemma. For $x\in B$, let us denote by $\mathcal{N}_{hol}((x,R),{\ensuremath{\omega}}_{B^{c}})$ the hole number variation when the ball $\bar{B}(x,R)$ is added to the configuration ${\ensuremath{\omega}}_{B^{c}}$. This quantity is central in our proof. Indeed, a first upperbound for the energy $H_{B}((x,R)\cup{\ensuremath{\omega}}_{B^{c}})$ is given by Lemma \[Borneenergie\]: $$\label{firstUB} H_{B}((x,R)\cup{\ensuremath{\omega}}_{B^{c}}) = h((x,R),{\ensuremath{\omega}}_{B^{c}}) \leq K - {\ensuremath{\theta}}_{3} \, \mathcal{N}_{hol}((x,R),{\ensuremath{\omega}}_{B^{c}}) ~,$$ where $K$ is a positive constant only depending on parameters ${\ensuremath{\theta}}_{1}$, ${\ensuremath{\theta}}_{2}$, ${\ensuremath{\theta}}_{3}$ and radii $R_{0}$, $R_{1}$. So, to upperbound the energy $H_{B}((x,R)\cup{\ensuremath{\omega}}_{B^{c}})$ it suffices to upperbound the number of created holes (resp. deleted holes) when $\theta_{3}$ is negative (resp. positive). This is the reason why the proof of the Connection Lemma differs according to the sign of the parameter $\theta_{3}$. ### When $\theta_{3}$ is negative {#section:theta3>0} Let ${\ensuremath{\omega}}$ be a configuration and $\alpha$ be a positive real number. A couple $(x,R)\in{\ensuremath{{\mathbb{R}}}}^{2}\times[R_{0},R_{1}]$ is said *good* if all the connected components of the set $\bar{{\ensuremath{\omega}}}_{\Delta}\cap\bar{B}(x,R)$ have an area larger than $\alpha$. These couples are well-named because adding a ball $\bar{B}(x,R)$ to the configuration ${\ensuremath{\omega}}_{\Delta}$, with a good couple $(x,R)$, does not create too many holes. \[BorneN\_t\] Let $(x,R)\in{\ensuremath{{\mathbb{R}}}}^{2}\times[R_{0},R_{1}]$ be a good couple. Then, $$\mathcal{N}_{hol}((x,R),{\ensuremath{\omega}}_{\Delta}) \leq \frac{\pi R_{1}^{2}}{\alpha} ~.$$ The number of created holes when the ball $\bar{B}(x,R)$ is added to ${\ensuremath{\omega}}_{\Delta}$ is smaller than the number of connected components of the set $\bar{{\ensuremath{\omega}}}_{\Delta}\cap\bar{B}(x,R)$. Since $(x,R)$ is good, all these connected components have an area larger than $\alpha$. So, there are at most $\pi R^{2}/\alpha$ such connected components. Let us denote by $\textrm{Bad}({\ensuremath{\omega}}_{\Delta},\alpha)$ the following set: $$\textrm{Bad}({\ensuremath{\omega}}_{\Delta},\alpha) = \{ x \in {\ensuremath{{\mathbb{R}}}}^{2} , \; \exists R \in [R_{0},R_{0}+{\ensuremath{\varepsilon}}] , \; (x,R) \; \mbox{ is not good }\} ~.$$ \[limiteBad\] The area of the set $\textrm{Bad}({\ensuremath{\omega}}_{\Delta},\alpha)$ tends to $0$ as $\alpha$ and ${\ensuremath{\varepsilon}}$ tend to $0$, uniformly on the configuration ${\ensuremath{\omega}}_{\Delta}$. Lemma \[limiteBad\] will be proved at the end of this section. Thanks to Lemmas \[BorneN\_t\] and \[limiteBad\], we are now able to prove the Connection Lemma. First, we need a family of (small) non-overlapping squared boxes whose union covers the convex hull of the boxes $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ and $B_{\texttt{W}}$. Precisely, for $\kappa>0$, let us consider a subset $\mathcal{B}$ of $\{v+[0,\kappa[^{2},v\in{\ensuremath{{\mathbb{R}}}}^{2}\}$ such that for any $B,B'$ in $\mathcal{B}$, $B\cap B'$ is empty, and $$\textrm{Conv}\left( B_{\texttt{N}}, B_{\texttt{S}}, B_{\texttt{E}}, B_{\texttt{W}} \right) \subset \bigcup_{B\in\mathcal{B}} B \subset \Delta ~.$$ The family $\mathcal{B}$ is made up of at most $c_{\kappa}=\kappa^{-2}{\ensuremath{{\mathcal{A}}}}(\Delta)$ elements.\ The hypothesis $N_{cc}^{\Delta}({\ensuremath{\omega}})\geq 2$ ensures the existence of two elements $(x_{1},\cdot)$ and $(x_{2},\cdot)$ of ${\ensuremath{\omega}}$, whose centers $x_{1}$ and $x_{2}$ are in the union of the four cardinal boxes $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ and $B_{\texttt{W}}$, and whose balls $\bar{B}(x_{1},\cdot)$ and $\bar{B}(x_{2},\cdot)$ belong to two different connected components of $\bar{{\ensuremath{\omega}}}$, say respectively $C_{1}$ and $C_{2}$. Let $[x_{1},x_{2}]$ be the segment in ${\ensuremath{{\mathbb{R}}}}^{2}$ linking $x_{1}$ with $x_{2}$ and $d$ be the euclidean distance on ${\ensuremath{{\mathbb{R}}}}^{2}$. The continuous application $$f : x \in [x_{1},x_{2}] \, \mapsto \, d(x,C_{1}) - d(x,\bar{{\ensuremath{\omega}}}\setminus C_{1})$$ satisfies $f(x_{1})<0$ and $f(x_{2})>0$. So there exists a point $x$ in $[x_{1},x_{2}]$ such that $d(x,C_{1})$ and $d(x,\bar{{\ensuremath{\omega}}}\setminus C_{1})$ are equal (and positive). Hence, the ball $\bar{B}(x,R_{0})$ does not contain any point of ${{\ensuremath{\omega}}}_{\Delta}$. Moreover, since $x$ is in the convex hull of the boxes $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$ and $B_{\texttt{W}}$, then it belongs to one box of the family $\mathcal{B}$, say $B$. With $\kappa<R_{0}/\sqrt{2}$, the box $B$ is contained in $\bar{B}(x,R_{0})$. Consequently, ${\ensuremath{\omega}}_B$ is empty : $$\label{choixB} P \left( N_{cc}^{\Delta}({\ensuremath{\omega}}) \geq 2 \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) \leq \sum_{B\in\mathcal{B}} P \left( {\ensuremath{\omega}}_{B} = \emptyset \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) ~.$$ For a given box $B\in\mathcal{B}$, let us consider the (random) set $U({\ensuremath{\omega}}_{\Delta\setminus B})$ of points $x\in B$ such that, for any radius $R\in [R_{0},R_{0}+{\ensuremath{\varepsilon}}]$, the couple $(x,R)$ is good: $$U({\ensuremath{\omega}}_{\Delta\setminus B}) = B \setminus \textrm{Bad}({\ensuremath{\omega}}_{\Delta\setminus B},\alpha) ~.$$ Let $x\in U({\ensuremath{\omega}}_{\Delta\setminus B})$ and $R\in [R_{0},R_{0}+{\ensuremath{\varepsilon}}]$. On the one hand, using (\[firstUB\]), ${\ensuremath{\theta}}_{3}\leq 0$ and Lemma \[BorneN\_t\], we get $$\label{secondUBpositif} H_{B}((x,R)\cup{\ensuremath{\omega}}_{B^{c}}) \leq K - {\ensuremath{\theta}}_{3} M ~,$$ where $M=M(R_{1},\alpha)$ denotes the upperbound given by Lemma \[BorneN\_t\]. On the other hand, Lemma \[limiteBad\] implies the area of $U({\ensuremath{\omega}}_{\Delta\setminus B})$ is larger than $\kappa^{2}/2$ for $\alpha$ and ${\ensuremath{\varepsilon}}$ small enough, uniformly on the configuration ${\ensuremath{\omega}}_{\Delta\setminus B}$. It follows: $$\begin{aligned} \lefteqn{P \left( \#{\ensuremath{\omega}}_{B\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]} = 1 \,|\, {\ensuremath{\omega}}_{B^{c}} \right) }\hspace*{2cm} \\ & & = \frac{1}{Z_{B}({\ensuremath{\omega}}_{B^{c}})} \int_{B\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]} e^{-H_{B}((x,R)\cup{\ensuremath{\omega}}_{B^{c}})} \, z e^{-z{\ensuremath{\lambda}}(B)} \, {\ensuremath{\lambda}}(dx) Q(dR) \\ & & \geq \frac{z e^{-z \kappa^{2}}}{Z_{B}({\ensuremath{\omega}}_{B^{c}})} \int_{U({\ensuremath{\omega}}_{\Delta\setminus B})\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]} e^{-H_{B}((x,R)\cup{\ensuremath{\omega}}_{B^{c}})} \, {\ensuremath{\lambda}}(dx) Q(dR) \\ & & \geq \frac{z e^{-z \kappa^{2}}}{Z_{B}({\ensuremath{\omega}}_{B^{c}})} e^{-K+{\ensuremath{\theta}}_{3}M} \, \frac{\kappa^{2}}{2} \, Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}]) ~.\end{aligned}$$ In the previous inequality, replacing $e^{-z \kappa^{2}} Z_{B}({\ensuremath{\omega}}_{B^{c}})^{-1}$ with the conditional probability $P({\ensuremath{\omega}}_{B}=\emptyset | {\ensuremath{\omega}}_{B^{c}})$, we obtain $$P \left( {\ensuremath{\omega}}_{B} = \emptyset \,|\, {\ensuremath{\omega}}_{B^{c}} \right) \leq \frac{2 \, e^{K-{\ensuremath{\theta}}_{3}M}}{z \, \kappa^{2} \, Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}])} ~.$$ Finally, the Connection Lemma derives from the above upperbound, (\[choixB\]) and with $$C' = \frac{2 \, c_{\kappa} \, e^{K-{\ensuremath{\theta}}_{3}M}}{\kappa^{2} \, Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}])} ~.$$ In order to prove Lemma \[limiteBad\], we have to locate the set $\textrm{Bad}({\ensuremath{\omega}}_{\Delta},\alpha)$. Lemma \[LocalBad\] says that the points $(x,.)$ in $\textrm{Bad}({\ensuremath{\omega}}_{\Delta},\alpha)$ are at distance around $R_{0}$ from $\bar{{\ensuremath{\omega}}}_{\Delta}$. We need some notations. Let $\bar{B}(x,R)$ and $\bar{B}(y,\cdot)$ be two balls satisfying $R_{0}\leq R\leq R_{0}+{\ensuremath{\varepsilon}}$ and $$0 < {\ensuremath{{\mathcal{A}}}}(\bar{B}(x,R)\cap \bar{B}(y,\cdot)) \leq \alpha ~.$$ Then there exists a positive function $g({\ensuremath{\varepsilon}},\alpha)$ tending to $0$ as $\alpha$ and ${\ensuremath{\varepsilon}}$ tend to $0$, such that $$\label{DistBad} |\, d(x,\bar{B}(y,\cdot)) - R_{0} \,| \leq g({\ensuremath{\varepsilon}},\alpha) ~.$$ The function $g$ is also allowed to depend on radii $R_{0}$ and $R_{1}$. The topological boundary $\partial \bar{{\ensuremath{\omega}}}_{\Delta}$ is composed of a finite number of arcs. Let $a$ be one of them. This arc is generated by an element of the configuration ${\ensuremath{\omega}}_{\Delta}$, say $(y,\cdot)$. Now, we can define the circular strip $S_{g}(a)$ of width $2g({\ensuremath{\varepsilon}},\alpha)$ by $$S_{g}(a) = \left\{ x \in {\ensuremath{{\mathbb{R}}}}^{2} ; \; \exists y' \in a \mbox{ s.t. }\begin{array}{c} x = y + \mu (y'-y) \, \mbox{ with } \, \mu > 0 \, \mbox{ and} \\ |\, d(x,y') - R_{0} \,| \leq g({\ensuremath{\varepsilon}},\alpha) \end{array} \right\} ~.$$ \[LocalBad\] The following inclusion holds; $$\label{InclusionBad} \textrm{Bad}({\ensuremath{\omega}}_{\Delta},\alpha) \subset \bigcup_{a, \, \mbox{\footnotesize{arc of} } \partial \bar{{\ensuremath{\omega}}}_{\Delta}} S_{g}(a) ~.$$ Let us consider a point $x$ in $\textrm{Bad}({\ensuremath{\omega}}_{\Delta},\alpha)$. Let $R\in [R_{0},R_{0}+{\ensuremath{\varepsilon}}]$ such that $(x,R)$ is not good. So there exists a connected component of $\bar{{\ensuremath{\omega}}}_{\Delta}\cap\bar{B}(x,R)$ of area smaller than $\alpha$. The boundary of this connected component through the open ball $B(x,R)$ is composed of a finite number of arcs, say $a_{1},\ldots,a_{n}$. Let $a$ be one of them realizing the minima $$d(x,a) = \min_{1\leq i\leq n} d(x,a_{i}) ~.$$ Let $(y,\cdot)$ be the element of the configuration ${\ensuremath{\omega}}_{\Delta}$ generating the arc $a$. Let $S(a)$ be the semi-infinite cone centered at $y$ and with arc $a$ (i.e. the union of semi-line $[y,y')$ for $y'\in a$). Then, $x$ necessarily belongs to $S(a)$. Indeed, the opposite situation could lead to the existence of another arc $a'$ satisfying $d(x,a')<d(x,a)$. To sum up, $x$ is in the semi-infinite cone $S(a)$ and the area of $\bar{B}(x,R)\cap \bar{B}(y,\cdot)$ is positive and smaller than $\alpha$. So $x$ satisfies (\[DistBad\]) and then belongs to $S_{g}(a)$. Let $a$ be an arc of the boundary $\partial \bar{{\ensuremath{\omega}}}_{\Delta}$. Some geometrical considerations allow to bound the area of the circular strip $S_{g}(a)$: $${\ensuremath{{\mathcal{A}}}}(S_{g}(a)) \leq 4 g({\ensuremath{\varepsilon}},\alpha) \textrm{length}(a) ~,$$ where $\textrm{length}(a)$ denotes the length of the arc $a$. We deduce from this bound and Lemmas \[LocalBad\] and \[BornePerimetre\]: $$\begin{aligned} {\ensuremath{{\mathcal{A}}}}( \textrm{Bad}({\ensuremath{\omega}}_{\Delta},\alpha) ) & \leq & \sum_{a \, \mbox{\footnotesize{arc of} } \partial \bar{{\ensuremath{\omega}}}_{\Delta}} {\ensuremath{{\mathcal{A}}}}( S_{g}(a) ) \\ & \leq & 4 g({\ensuremath{\varepsilon}},\alpha) \sum_{a \, \mbox{\footnotesize{arc of} } \partial \bar{{\ensuremath{\omega}}}_{\Delta}} \textrm{length}(a) \\ & \leq & 4 g({\ensuremath{\varepsilon}},\alpha) {\ensuremath{{\mathcal{L}}}}_{\Delta'}( \bar{{\ensuremath{\omega}}}_{\Delta} ) \; \; \mbox{ with } \Delta'=\Delta\oplus B(0,R_{1}) \\ & \leq & 4 g({\ensuremath{\varepsilon}},\alpha) \frac{2 {\ensuremath{{\mathcal{A}}}}( \Delta' \oplus B(0,R_{0}) )}{R_{0}} ~.\end{aligned}$$ This latter upperbound does not depend on the configuration ${\ensuremath{\omega}}_{\Delta}$. So, this ends the proof of Lemma \[limiteBad\]. ### When $\theta_{3}$ is positive In this section, we still assume that $N_{cc}^{\Delta}({\ensuremath{\omega}})$ is larger than $2$. But this time, our aim consists in upperbounding the number of deleted holes when the ball $\bar{B}(x,R)$, $x\in B$, is added to the configuration ${\ensuremath{\omega}}_{B^{c}}$. The existence of a suitable set $B$ derives from Lemma \[ExistenceBoule\]. Its proof is rather long and technical, mainly because of the uniformity of $\rho>0$ with respect to the configuration ${\ensuremath{\omega}}$. \[ExistenceBoule\] Assume $N_{cc}^{\Delta}({\ensuremath{\omega}})\geq 2$. There exist $\rho>0$ (which does not depend on ${\ensuremath{\omega}}$) and $O=O({\ensuremath{\omega}})\in\Delta$ such that: - $O$ is in $\textrm{Conv}\left( B_{\texttt{N}}, B_{\texttt{S}}, B_{\texttt{E}}, B_{\texttt{W}} \right) \oplus B(0,\frac{3}{2} R_{0})$; - $B(O,\rho R_{0}) \cap {\ensuremath{\omega}}$ is empty; - $B(O,(1+\rho) R_{0})$ does not (totally) contain any hole of $\bar{{\ensuremath{\omega}}}$. Let us first explain how the Connection Lemma straight derives from Lemma \[ExistenceBoule\]. As in Section \[section:theta3&gt;0\], we need the family $\mathcal{B}$ of non-overlapping squared boxes of length side $\kappa$. But here, $\mathcal{B}$ is required to cover a little bit more, i.e. $$\label{B2} \textrm{Conv}\left( B_{\texttt{N}}, B_{\texttt{S}}, B_{\texttt{E}}, B_{\texttt{W}} \right) \oplus B(0,\frac{3}{2} R_{0}) \subset \bigcup_{B\in\mathcal{B}} B ~,$$ and parameters $\kappa$ and ${\ensuremath{\varepsilon}}$ are chosen small enough so that $$\label{KappaEpsRho} \sqrt{2} \kappa + {\ensuremath{\varepsilon}}< \rho R_{0}$$ (where $\rho$ is given by Lemma \[ExistenceBoule\]). Thanks to statement $(i)$ and (\[B2\]), the point $O$ belongs to a box $B\in\mathcal{B}$. Thanks to $(ii)$, $(iii)$ and (\[KappaEpsRho\]), ${\ensuremath{\omega}}_{B}$ is empty and $\bar{{\ensuremath{\omega}}}_{B^{c}}$ has no hole in $\mathbf{B}:=B\oplus B(0,R_{0}+{\ensuremath{\varepsilon}})$. Hence, $$\label{choixB2} P \left( N_{cc}^{\Delta}({\ensuremath{\omega}}) \geq 2 \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) \leq \sum_{B\in\mathcal{B}} P \left( P \left( {\ensuremath{\omega}}_{B} = \emptyset \,|\, {\ensuremath{\omega}}_{B^{c}} \right) {\ensuremath{\mbox{\rm 1\kern-0.23em I}}}_{\bar{{\ensuremath{\omega}}}_{B^{c}} \mbox{\footnotesize{ has no hole in }} \mathbf{B}} \,|\, {\ensuremath{\omega}}_{\Delta^{c}} \right) ~.$$ Let us pick a box $B\in\mathcal{B}$, a couple $(x,R)\in B\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]$ and assume that $\bar{{\ensuremath{\omega}}}_{B^{c}}$ has no hole in $\mathbf{B}$. Then, no hole is deleted when $\bar{B}(x,R)$ is added to ${\ensuremath{\omega}}_{B^{c}}$. So, the hole number variation $\mathcal{N}_{hol}((x,R),{\ensuremath{\omega}}_{B^{c}})$ is nonnegative. Combining with ${\ensuremath{\theta}}_{3}\geq 0$ and (\[firstUB\]), the energy $H_{B}((x,R)\cup{\ensuremath{\omega}}_{B^{c}})$ is smaller than $K$ and we finish the proof of the Connection Lemma as in Section \[section:theta3&gt;0\]. First, $$P \left( \#{\ensuremath{\omega}}_{B\times[R_{0},R_{0}+{\ensuremath{\varepsilon}}]} = 1 \,|\, {\ensuremath{\omega}}_{B^{c}} \right) \geq \frac{z e^{-z \kappa^{2}}}{Z_{B}({\ensuremath{\omega}}_{B^{c}})} e^{-K} \, \kappa^{2} \, Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}]) ~.$$ Thus, replacing $e^{-z \kappa^{2}} Z_{B}({\ensuremath{\omega}}_{B^{c}})^{-1}$ by the conditional probability $P({\ensuremath{\omega}}_{B}=\emptyset | {\ensuremath{\omega}}_{B^{c}})$, we get $$P \left( {\ensuremath{\omega}}_{B} = \emptyset \,|\, {\ensuremath{\omega}}_{B^{c}} \right) \leq \frac{e^{K}}{z \, \kappa^{2} \, Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}])} ~.$$ Finally, the Connection Lemma derives from the above upperbound, (\[choixB2\]) and with $$C' = \frac{c_{\kappa} \, e^{K}}{\kappa^{2} \, Q([R_{0},R_{0}+{\ensuremath{\varepsilon}}])} ~,$$ where $c_{\kappa}$ still denotes the number of boxes contained in the family $\mathcal{B}$.\ Now, let us find a point $O$ and a radius $\rho>0$ satisfying the three properties of Lemma \[ExistenceBoule\]. A first applicant for the point $O$ can be obtained following the same method as in Section \[section:theta3&gt;0\]. Based on the hypothesis $N_{cc}^{\Delta}({\ensuremath{\omega}})\geq 2$, this method ensures the existence of a point $O'$ in the convex hull of the $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$, $B_{\texttt{W}}$’s, such that $$\label{O'} \mathrm{\mathbf{d}} := d(O',C_{1}) = d(O',\bar{{\ensuremath{\omega}}}_{\Delta}\setminus C_{1}) >0$$ where $C_{1}$ denotes a connected component of $\bar{{\ensuremath{\omega}}}_{\Delta}$ counting by $N_{cc}^{\Delta}({\ensuremath{\omega}})$. Two cases will be considered in the following. In the first one– $\mathrm{\mathbf{d}}\geq \frac{1}{2} R_{0}$ –the connected components of $\bar{{\ensuremath{\omega}}}_{\Delta}$ are far away from $O'$. So do their holes. Then, the choice $O=O'$ is appropriate. In the second case– $\mathrm{\mathbf{d}}\leq\frac{1}{2} R_{0}$ –we exhibit a region close to $O'$ without hole and choose a suitable point $O$ inside. About the radius $\rho$, it will be proved in the sequel that any positive real number such that $$\label{rhoC1} \left( 1 + \rho \right)^{2} \, < \, 1 + \frac{1}{4} ~,$$ $$\label{rhoC3} \sqrt{7} ( 1+\rho ) - \frac{7}{4}\, < \, 1$$ and $$\label{rhoC2} \left( 1-\frac{\sqrt{7}}{4}+\rho \right)^{2} + \left( \frac{3}{2}-\sqrt{\left( 1-\rho \right)^{2} - \left( 1-\frac{\sqrt{7}}{4}+\rho \right)^{2}} \right)^{2} \, < \, \left( \sqrt{3} - 1 - \rho \right)^{2} ~,$$ is suitable. For instance, $\rho=0.01$ satisfies these three conditions.\ **Case 1:** $\mathrm{\mathbf{d}}\geq \frac{1}{2} R_{0}$.\ By construction, $O'$ is in the convex hull of the boxes $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$, $B_{\texttt{W}}$ and is at distance at least $R_{0}+\mathrm{\mathbf{d}}$ from any point $x$ in ${\ensuremath{\omega}}_{\Delta}$. So, it satisfies properties $(i)$ and $(ii)$ of Lemma \[ExistenceBoule\]. Now, let us consider a hole $T$ of $\bar{{\ensuremath{\omega}}}_{\Delta}$. Assume in a first time that $O'$ does not belong to $T$. By (\[rhoC1\]) and Lemma \[DistTrou1\], $$d(O',T)^{2} \geq \left( 1+\frac{1}{4} \right) R_{0}^{2} \geq (1+\rho)^{2} R_{0}^{2} ~.$$ This means that the hole $T$ is outside the ball $B(O',(1+\rho) R_{0})$. Now, assume that $O'$ is in $T$. Since $O'$ is equidistant from two connected components of $\bar{{\ensuremath{\omega}}}_{\Delta}$ then one of them is inside the hole $T$. Hence, $T$ is too large to be totally covered by the ball $B(O',(1+\rho) R_{0})$. Consequently, $O'$ also satisfies $(iii)$.\ **Case 2:** $\mathrm{\mathbf{d}}\leq\frac{1}{2} R_{0}$.\ Let $\bar{B}(x_{1},R_{x_{1}})$ be a ball of the connected component $C_{1}$ on which the distance $d(O',C_{1})$ is reached. Let us consider the point $y_{1}$ on the segment $[O',x_{1}]$ satisfying $\bar{B}(y_{1},R_{0})$ is included in $\bar{B}(x_{1},R_{x_{1}})$ and $$d(O',\bar{B}(y_{1},R_{0})) = d(O',\bar{B}(x_{1},R_{x_{1}})) = d(O',C_{1}) = \mathrm{\mathbf{d}} ~.$$ In the same way, let us consider a point $y_{2}$ such that $\bar{B}(y_{2},R_{0})$ is included in $\bar{{\ensuremath{\omega}}}_{\Delta}\setminus C_{1}$ and $$d(O',\bar{B}(y_{2},R_{0})) = d(O',\bar{{\ensuremath{\omega}}}_{\Delta}\setminus C_{1}) = \mathrm{\mathbf{d}} ~.$$ The region without hole, mentioned at the beginning of the current section and which we need, is built from points $y_{1}$ and $y_{2}$. See Figure \[fig:nohole\]. Let $\mathcal{D}$ be the infinite line passing by $y_{1}$ and $y_{2}$. Thus, let us consider two infinite lines $\mathcal{D}'$ and $\mathcal{D}''$ parallel to $\mathcal{D}$ and such that $$d(\mathcal{D}',\mathcal{D}) = d(\mathcal{D}'',\mathcal{D}) = \frac{\sqrt{7}}{4} R_{0}$$ (say $O'$ and $\mathcal{D}'$ are on the same side of the line $\mathcal{D}$). We denote by $\mathcal{H}$ the intersection of the convex hull of balls $\bar{B}(y_{1},R_{0})$ and $\bar{B}(y_{2},R_{0})$ with the strip delimited by $\mathcal{D}'$ and $\mathcal{D}''$. On Figure \[fig:nohole\], the border of $\mathcal{H}$ is drawn in bold. \[noholeH\] With the previous notations and hypotheses, there is no hole in $\mathcal{H}$. ![\[fig:nohole\] ](nohole.eps){width="12cm" height="6cm"} The closest hole $T$ to the segment $[y_{1},y_{2}]$ is obtained by pressing a ball with radius $R_{0}$ against $\bar{B}(y_{1},R_{0})$ and $\bar{B}(y_{2},R_{0})$. If $l$ denotes the distance between $T$ and $[y_{1},y_{2}]$ then $2l$ is the distance between the center of this pressing ball and $[y_{1},y_{2}]$. Pythagoras Theorem gives $(2l)^{2}+(R_{0}+h)^{2}=(2R_{0})^{2}$ in which $h$ denotes $$h := \frac{1}{2} d(y_{1},y_{2}) - R_{0} \, \leq \, \mathrm{\mathbf{d}} ~.$$ In the worst case, $h=\frac{1}{2} R_{0}$. Hence, $l$ is always larger than $\frac{\sqrt{7}}{4} R_{0}$, which is the distance between $\mathcal{D}$ and $\mathcal{D}'$. To complete the proof, let us add there is no hole in the balls $\bar{B}(y_{1},R_{0})$ and $\bar{B}(y_{2},R_{0})$ since they are totally covered by $\bar{{\ensuremath{\omega}}}_{\Delta}$. The idea to conclude the proof can be sum up as follows. The region $\mathcal{H}$ is sufficiently thick to contain strictly more than half of a ball with radius $(1+\rho)R_{0}$. Hence, the part of this ball outside $\mathcal{H}$ (this is the hatched region on Figure \[fig:nohole\]) has a diameter smaller than $2R_{0}$. Thanks to Lemma \[DistTrou3\], it is possible to choose the center $O$ of this ball so that $\bar{B}(O,(1+\rho)R_{0})\cap \mathcal{H}^{c}$ does not contain any hole.\ Let $\mathcal{D}_{O}$ be the infinite line parallel to $\mathcal{D}''$, at distance $(1+\rho)R_{0}$ from $\mathcal{D}''$ and on the same side as $\mathcal{D}$ of the line $\mathcal{D}''$. It derives from (\[rhoC3\]) that the line $\mathcal{D}_{O}$ is trapped between $\mathcal{D}$ and $\mathcal{D}'$. Let $M$ be the center of the segment $[y_{1},y_{2}]$. Let us denote by $[z_{1},z_{2}]$ the following segment: $$[z_{1},z_{2}] := \bar{B}( M,(1-\rho)R_{0} ) \cap \mathcal{D}_{O} ~.$$ See Figure \[fig:nohole\]. We are going to choose the point $O$ on the segment $[z_{1},z_{2}]$. To do it, some geometrical results about the previous construction are needed. They will be proved at the end of the section: \[3results\] With the previous notations and hypotheses, the following statements hold: - for $i=1,2$, $d(O',z_{i})\leq \frac{3}{2}R_{0}$; - $[z_{1},z_{2}]\oplus B(0,\rho R_{0}) \subset B(O',R_{0}+\mathrm{\mathbf{d}})$; - for $i=1,2$, $d(y_{i},z_{i}) \leq (\sqrt{3}-1-\rho) R_{0}$. By convexity and statement $(a)$, any point of the segment $[z_{1},z_{2}]$ is at distance from $O'$ larger than $\frac{3}{2}R_{0}$. Moreover, $O'$ is in the convex hull of the $B_{\texttt{N}}$, $B_{\texttt{S}}$, $B_{\texttt{E}}$, $B_{\texttt{W}}$’s. Then, any point of $[z_{1},z_{2}]$ satisfies the property $(i)$ of Lemma \[ExistenceBoule\].\ By construction of the point $O'$, the ball $B(O',R_{0}+\mathrm{\mathbf{d}})$ does not contain any point of ${\ensuremath{\omega}}$. So does the set $[z_{1},z_{2}]\oplus B(0,\rho R_{0})$ thanks to statement $(b)$. This means that any point of the segment $[z_{1},z_{2}]$ satisfies the property $(ii)$ of Lemma \[ExistenceBoule\].\ Combining statement $(c)$ with $i=1$ and Lemma \[DistTrou2\], we check there is no hole of $\bar{{\ensuremath{\omega}}}_{\Delta}\setminus C_{1}$ in the ball $B(z_{1},(1+\rho) R_{0})$. Let us run the center of a ball with radius $(1+\rho) R_{0}$ along the segment $[z_{1},z_{2}]$ from $z_{1}$ to $z_{2}$ until that ball meets a hole of $\bar{{\ensuremath{\omega}}}_{\Delta}\setminus C_{1}$. Two cases can be distinguished. - This meet does not happen. Then, the ball $B(z_{2},(1+\rho) R_{0})$ does not contain any hole of $\bar{{\ensuremath{\omega}}}_{\Delta}\setminus C_{1}$. It does not contain any hole of $C_{1}$ either thanks to statement $(c)$ with $i=2$ and Lemma \[DistTrou2\]. In this case, $O=z_{2}$ satisfies the property $(iii)$ of Lemma \[ExistenceBoule\]. - This meet happens: let $O$ be the corresponding center and $T$ be the corresponding hole of $\bar{{\ensuremath{\omega}}}_{\Delta}\setminus C_{1}$. As just before, the ball $B(O,(1+\rho) R_{0})$ does not still contain any hole of $\bar{{\ensuremath{\omega}}}_{\Delta}\setminus C_{1}$. Denote by $\mathcal{C}$ the part of this ball outside $\mathcal{H}$: $$\mathcal{C} := B(O,(1+\rho) R_{0}) \cap \mathcal{H}^{c} ~.$$ On the one hand, the diameter of $\mathcal{C}$ is smaller than $2R_{0}$ thanks to (\[rhoC3\]). On the other hand, $\mathcal{C}$ is pressed against the hole $T$ (there is no hole in $\mathcal{H}$); see Figure \[fig:nohole\]. By Lemma \[DistTrou3\], the holes of the connected component $C_{1}$ are at distance from $T$ at least $2R_{0}$. So they cannot belong to the set $\mathcal{C}$. Therefore, this point $O$ satisfies the property $(iii)$ of Lemma \[ExistenceBoule\]. The infinite line $\mathcal{D}$ divides $\bar{B}(M,R_{0})$ into two half-balls; let $\mathcal{V}$ be the one containing the segment $[z_{1},z_{2}]$. Since $$d(O',y_{1}) = d(O',y_{2}) = R_{0}+\mathrm{\mathbf{d}} ~,$$ the half-ball $\mathcal{V}$ is included in the ball with center $O'$ and radius $R_{0}+\mathrm{\mathbf{d}}$. This inclusion admits two consequences. First, the points $z_{1}$ and $z_{2}$ which are in $\mathcal{V}$, are also in the ball $\bar{B}(O',R_{0}+\mathrm{\mathbf{d}})$. This implies, for $i=1,2$ $$d(O',z_{i}) \leq \frac{3}{2} R_{0} ~,$$ i.e. statement $(a)$. Second, the balls $\bar{B}(z_{i},\rho R_{0})$ which are included in $\mathcal{V}$, are also included in $\bar{B}(O',R_{0}+\mathrm{\mathbf{d}})$. So does the set $[z_{1},z_{2}]\oplus B(0,\rho R_{0})$ by convexity. Statement $(b)$ is proved. It remains to prove statement $(c)$. Let us introduce the orthogonal projection $h_{1}$ of $z_{1}$ over the infinite line $\mathcal{D}$ (see Figure \[fig:nohole\]). Using $d(M,z_{1})=(1-\rho)R_{0}$, $d(h_{1},z_{1})=(1+\rho-\frac{\sqrt{7}}{4})R_{0}$ and $\mathrm{\mathbf{d}}\leq\frac{1}{2} R_{0}$, we get $$d(y_{1},z_{1}) \leq \sqrt{\left( 1-\frac{\sqrt{7}}{4}+\rho \right)^{2} + \left( \frac{3}{2}-\sqrt{\left( 1-\rho \right)^{2} - \left( 1-\frac{\sqrt{7}}{4}+\rho \right)^{2}} \right)^{2}} \, R_{0} ~.$$ Thanks to (\[rhoC2\]), statement $(c)$ follows. Proofs of geometrical lemmas {#SectionGeoLemma} ---------------------------- \[BornePerimetre\] Let $\Delta$ be a bounded closed convex set. For any configuration ${\ensuremath{\omega}}$, let us denote by ${\ensuremath{{\mathcal{L}}}}_\Delta(\bar{\ensuremath{\omega}})$ the perimeter of $\bar{\ensuremath{\omega}}$ viewed through $\Delta$: $${\ensuremath{{\mathcal{L}}}}_\Delta(\bar{\ensuremath{\omega}}) = {\ensuremath{{\mathcal{L}}}}(\bar {\ensuremath{\omega}}\cap \Delta) - \textrm{length}(\partial \Delta \cap \bar {\ensuremath{\omega}}),$$ where $\textrm{length}(\partial \Delta \cap \bar {\ensuremath{\omega}})$ denotes the lentgh of the boundary of $\Delta$ which is inside the set $\bar {\ensuremath{\omega}}$. Then, $${\ensuremath{{\mathcal{L}}}}_\Delta(\bar{\ensuremath{\omega}}) \leq \frac{2 {\ensuremath{{\mathcal{A}}}}(\Delta\oplus B(0,R_0))}{R_0} ~.$$ The boundary of $\bar {\ensuremath{\omega}}$ viewed through $\Delta$ corresponds to a finite union of arcs, say $(a_i)_{1\le i \le n}$. For each arc $a_i$, coming from the ball $B(x_i,R_i)$, we consider the circular strip $S(a_i)$ of width $R_{0}$ defined by $$S(a_i) = \left\{ x \in {\ensuremath{{\mathbb{R}}}}^{2} ; \; \exists x' \in a_i \mbox{ s.t. }\begin{array}{c} x = x' + \mu (x_i-x') \, \mbox{ with } \, \mu > 0 \\ \mbox{ and } \; d(x,x')< R_{0} \end{array} \right\} ~.$$ Let us notice that the sets $(S(a_i))_{1\le i \le n}$ are disjoint. Indeed, let suppose that there exists $x\in S(a_i)\cap S(a_j)$ for some $i\neq j$. Without restriction, we can assume that the distance between $x$ and $a_i$ is smaller than or equal to the distance between $x$ and $a_j$. Let $y$ be the point on $a_i$ such that this distance is equal to $|y-x|$. Then, $y$ has to be strictly included in the ball $B(x_j,R_j)$ which contradicts the fact that $y$ is on the boundary of $\bar {\ensuremath{\omega}}$.\ This allows to compare the sum of the areas of $(S(a_i))_{1\le i \le n}$ with ${\ensuremath{{\mathcal{A}}}}(\bar{\ensuremath{\omega}})$: $$\begin{aligned} {\ensuremath{{\mathcal{L}}}}_\Delta(\bar{\ensuremath{\omega}}) = \sum_{i=1}^n \textrm{length}(a_i) & \leq & \frac{2}{R_0} \sum_{i=1}^n {\ensuremath{{\mathcal{S}}}}(a_i) \\ & \leq & \frac{2}{R_0} {\ensuremath{{\mathcal{A}}}}(\bar{\ensuremath{\omega}}) \\ & \leq & \frac{2 {\ensuremath{{\mathcal{A}}}}(\Delta\oplus B(0,R_0))}{R_0} ~.\end{aligned}$$ \[Borneenergie\] Let $\Delta$ be a bounded subset of ${\ensuremath{{\mathbb{R}}}}^{2}$, ${\ensuremath{\omega}}$ be a configuration on $\Delta$ and $(x,R)$ be an element of $\Delta\times[R_0,R_1]$. Let us denote by ${\ensuremath{{\mathcal{A}}}}((x,R),{\ensuremath{\omega}})$ the area variation when the ball $\bar{B}(x,R)$ is adding to the configuration $\bar{{\ensuremath{\omega}}}$: $${\ensuremath{{\mathcal{A}}}}((x,R),{\ensuremath{\omega}}) = {\ensuremath{{\mathcal{A}}}}((x,R)\cup{\ensuremath{\omega}}) - {\ensuremath{{\mathcal{A}}}}({\ensuremath{\omega}}) ~.$$ In the same way, we consider the perimeter variation ${\ensuremath{{\mathcal{L}}}}((x,R),{\ensuremath{\omega}})$ and the connected component number variation $\mathcal{N}_{cc}((x,R),{\ensuremath{\omega}})$. The following inequalities hold. $$\label{diffaire} 0 \leq {\ensuremath{{\mathcal{A}}}}((x,R),{\ensuremath{\omega}}) \leq \pi R_{1}^{2} ~.$$ $$\label{diffperimetre} - \frac{2 \pi (R_{1}+R_{0})^{2}}{R_{0}} \leq {\ensuremath{{\mathcal{L}}}}((x,R),{\ensuremath{\omega}}) \leq 2 \pi R_{1} ~.$$ $$\label{diffcc} - \pi \left( 1 + \frac{R_{1}}{R_{0}} \right) \leq \mathcal{N}_{cc}((x,R),{\ensuremath{\omega}}) \leq 1 ~.$$ Inequalities (\[diffaire\]), upperbounds of (\[diffperimetre\]) and (\[diffcc\]) are obvious. The border length of $\bar{{\ensuremath{\omega}}}$ which is lost when the ball $\bar{B}(x,R)$ is adding can be interpreted as the perimeter of $\bar{\ensuremath{\omega}}$ viewed through $\bar{B}(x,R)$ , i.e. as ${\ensuremath{{\mathcal{L}}}}_{\bar{B}(x,R)}(\bar{\ensuremath{\omega}})$. Thanks to Lemma \[BornePerimetre\], it is smaller than $$\frac{2 {\ensuremath{{\mathcal{A}}}}(\bar{B}(x,R)\oplus B(0,R_0))}{R_0} \leq \frac{2 \pi (R_{1}+R_{0})^{2}}{R_{0}} ~.$$ This gives the lowerbound of (\[diffperimetre\]). It remains to lowerbound $\mathcal{N}_{cc}((x,R),{\ensuremath{\omega}})$. For that purpose, the number of deleted connected components when $\bar{B}(x,R)$ is adding to $\bar{\ensuremath{\omega}}$, is smaller than the number of non-overlapping balls which overlap $\bar{B}(x,R)$. This number is at most $$\frac{2 \pi (R_{1}+R_{0})}{2 R_{0}} ~.$$ \[DistTrou1\] Let $\mathcal{C}$ be a connected component of $\bar{{\ensuremath{\omega}}}_{\Delta}$ and $T$ be a hole of $\mathcal{C}$. Any point $x\in{\ensuremath{{\mathbb{R}}}}^{2}$ such that $x\notin\mathcal{C}$ and $x\notin T$ satisfies $$d(x,T)^{2} \geq d(x,\mathcal{C})^{2} + 2 d(x,\mathcal{C}) R_{0} ~.$$ Let us consider a connected component $\mathcal{C}$, a hole $T$ and a point $x$ satisfying the assumptions of the lemma. Let $y$ be a point of the closure of $T$ such that $d(x,T)=|x-y|$. Necessarily, $y$ is on the boundary of two balls $B(z,R)$ and $B(z',R')$ of $\mathcal{C}$. Since $x$ belongs neither to $\mathcal{C}$ nor to $T$, at least one of $z-y$ or $z'-y$ has a nonnegative scalar product with $x-y$. Say $z-y$. Given $|x-z|$ and $|y-z|$, the distance $|x-y|$ is minimal when the vectors $z-y$ and $x-y$ are orthogonal. Hence, using $|x-z|\ge d(x,\mathcal{C})+R_0$ and $|y-z|\ge R_0$, it follows from Pythagoras Theorem that $$d(x,T)^2 \ge (d(x,\mathcal{C})+R_0)^2 - R_0^2 ~,$$ which concludes the proof. The following result is a straight consequence of Lemma \[DistTrou1\]. \[DistTrou2\] Let $\mathcal{C}$, $\mathcal{C}'$ be two connected components of $\bar{{\ensuremath{\omega}}}_{\Delta}$. Let $\bar{B}(x,R)$ be a ball of $\mathcal{C}$ and $T'$ be a hole of $\mathcal{C}'$ which does not contain $\bar{B}(x,R)$. Then, $$d(x,T') \geq \sqrt{3} R_{0} ~.$$ \[DistTrou3\] Let $T$ and $T'$ be two holes respectively of two connected components $\mathcal{C}$ and $\mathcal{C}'$ of $\bar{{\ensuremath{\omega}}}_{\Delta}$. If $T\not\subset T'$ and $T'\not\subset T$ then $$d(T,T') \geq 2 R_{0} ~.$$ Let $T$ and $T'$ be two holes satisfying the assumption of the lemma. We denote by $x$ and $y$ two points belonging respectively to the closure of $T$ and $T'$ such that $d(T,T')=|x-y|$. The point $x$ (respectively $y$) belongs to the boundary of two balls $B(z,R)$ and $B(z',R')$ of $\mathcal{C}$ (respectively $B(w,r)$ and $B(w',r')$ of $\mathcal{C}'$). An analysis, as in the proof of Lemma \[DistTrou1\], shows that the distance $|x-y|$ is minimal in the situation where $R=R'=r=r'=R_0$ and $z$, $z'$, $w$ and $w'$ form a parallelogram with length side $2R_0$. Then the points $x$ and $y$ are at the middle of two opposite sides and the result follows.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We show that a first order problem can approximate solutions of a robust optimization problem when the uncertainty set is scaled, and explore further properties of this first order problem.' author: - 'C.H. Jeffrey Pang' title: First order dependence on uncertainty sets in robust optimization --- Introduction ============ Robust optimization is the methodology of handling optimization problems with uncertain data. In practice, the presence of uncertainties in optimization problems can make nominal solutions meaningless. Such uncertainties can come from data uncertainty in measurement and estimation, or from uncertainty in implementation. We refer to the recent text [@BGN09] for more details. Consider the linear program: $$\begin{aligned} & \min_{x} & \bar{c}^{T}x+\bar{d}\\ & \mbox{s.t.} & \bar{A}x\leq\bar{b}.\end{aligned}$$ To account for the uncertainties in the data $(\bar{A},\bar{b})$, one instead considers a point $x$ to be feasible if it satisfies $$Ax\leq b\mbox{ for all }(A,b)\in\mathcal{U}.$$ Here, $\mathcal{U}$ is a set containing the nominal data $(\bar{A},\bar{b})$. We can consider the translation $\Delta\mathcal{U}=\mathcal{U}-(\bar{A},\bar{b})$ and ask: What is the behavior of optimal solutions to the robust optimization problem if the set $\Delta\mathcal{U}$ were to be scaled by some factor $\epsilon$? A large value of $\epsilon$ corresponds to a more robust solution, and a small value of $\epsilon$ places more importance in the objective function. Understanding the dependence of $\epsilon$ allows one to find a balance between optimization and robustness. The first order dependence on $\epsilon$ is addressed in Corollary \[cor:lin-approx\] for linear programs and Theorem \[thm:TFO-approx\] for nonlinear programs. The outline of this paper is as follows. We introduce robust linear programming in Section \[sec:Robust-lin\]. Before we introduce robust nonlinear programming in Section \[sec:Robust-nonlin\], we recall some topics in variational analysis (or nonsmooth analysis) as presented in the texts [@RW98; @Mor06; @Cla83] in Section \[sec:var-analysis\]. In Section \[sec:Robust-nonlin\], we also define the tangential problem, which will be important in Theorem \[thm:TFO-approx\], our main result. We present first order properties of the tangential problem in Section \[sec:First-order\], and study the effects of sums of uncertainty sets in the tangential problem in Section \[sec:uncertain-sets\]. \[sec:Robust-lin\]Robust linear programming =========================================== We keep our presentation compatible with [@BGN09], and begin with the definition of the robust counterpart of a linear program. (Robust counterpart) For $\bar{A}\in\mathbb{R}^{m\times n}$, $\bar{b}\in\mathbb{R}^{m}$, $\bar{c}\in\mathbb{R}^{n}$ and $\bar{d}\in\mathbb{R}$, where $m\geq n$, consider the linear program with parameters $(\bar{A},\bar{b},\bar{c},\bar{d})$ $$\begin{aligned} & & \min_{x}\bar{c}^{T}x+\bar{d}\label{eq:original-LP}\\ & & \mbox{s.t. }\bar{A}x\leq\bar{b}.\nonumber \end{aligned}$$ The *robust counterpart* (written as RC) of the above linear program is $$\begin{aligned} & & \min_{x}\left\{ \hat{c}(x)=\sup_{(A,b,c,d)\in\mathcal{U}}[c^{T}x+d]\mid Ax\leq b\mbox{ for all }(A,b,c,d)\in\mathcal{U}\right\} \\ & = & \min_{x,t}\{t\mid t\geq c^{T}x+d\mbox{, }Ax\leq b\mbox{ for all }(A,b,c,d)\in\mathcal{U}\},\end{aligned}$$ where $\mathcal{U}$ is an uncertainty set for the parameters $(A,b,c,d)$, with $(\bar{A},\bar{b},\bar{c},\bar{d})\in\mathcal{U}$. In a typical linear program, the variable $d$ does not affect the minimizer, but one has to take perturbations in $d$ into account in a robust optimization problem. The second formulation in the RC shows that we can rewrite the linear program so that $c$ stays constant at $\bar{c}$ and $d=0$. This is the approach we will take for the rest of this section, and we define $\mathcal{U}$ to be a set containing elements of the form $(A,b)$, where $(A,b)$ are close enough to $(\bar{A},\bar{b})$. For more details, we refer to [@BGN09]. We define $\Delta A$, $\Delta b$ and the set $\Delta\mathcal{U}$ by the relations$$\begin{aligned} \Delta A & := & A-\bar{A},\\ \Delta b & := & b-\bar{b},\\ \mbox{ and }\Delta\mathcal{U} & := & \mathcal{U}-(\bar{A},\bar{b}).\end{aligned}$$ The vector $x$ can be chosen so that it stays feasible under these first order perturbations. We write $x=\bar{x}+\Delta x$. The RC is therefore simplified to $$\begin{aligned} & & \min_{\Delta x}c^{T}(\bar{x}+\Delta x)+d\nonumber \\ & & \mbox{s.t. }[\bar{A}_{i}+\Delta A_{i}](\bar{x}+\Delta x)\leq[\bar{b}_{i}+\Delta b_{i}]\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i},\label{eq:RC-final}\end{aligned}$$ where $\Delta\mathcal{U}_{i}$ is the uncertainty in the $i$th row. When $\Delta\mathcal{U}$ is a small set , we seek to use a first order approximation to determine a robustly feasible $x$. Letting $x=\bar{x}+\Delta x$, and removing the second order term $(\Delta A_{i})(\Delta x)$ in gives$$\begin{aligned} \bar{A}_{i}(\bar{x}+\Delta x)+(\Delta A_{i})\bar{x}-(\bar{b}_{i}+\Delta b_{i}) & \leq & 0\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i},\label{eq:robust-feasibility}\\ \mbox{or }[\bar{A}_{i}\bar{x}-\bar{b}]+\bar{A}_{i}(\Delta x)+[(\Delta A_{i})\bar{x}-\Delta b_{i}] & \leq & 0\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}.\nonumber \end{aligned}$$ If $\bar{A}_{i}\bar{x}-\bar{b}_{i}<0$ and if $\Delta\mathcal{U}$ were small enough, this constraint will not be tight in the optimization problem. With these in mind, we define the first order problem of a linear program. \[def:linear-FO-robust\](First order problem) Let $\bar{x}$ be an optimal solution to . The *first order problem* is the problem$$\begin{aligned} & & \min_{\gamma}c^{T}\gamma\label{eq:FO-feasible}\\ & & \mbox{s.t. }\bar{A}_{i}\gamma+[(\Delta A_{i})\bar{x}-\Delta b]\leq0\nonumber \\ & & \mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}\mbox{ whenever }\bar{A}_{i}\bar{x}-\bar{b}_{i}=0.\nonumber \end{aligned}$$ The first order problem can also be written as $$\begin{aligned} & & \min_{\gamma}c^{T}\gamma\\ & & \mbox{s.t. }\bar{A}_{i}\gamma\leq-\max_{(\Delta A,\Delta b)\in\Delta\mathcal{U}}[(\Delta A_{i})\bar{x}-\Delta b_{i}]\\ & & \mbox{ whenever }\bar{A}_{i}\bar{x}-\bar{b}_{i}=0.\end{aligned}$$ In the case where the optimal solution $\bar{x}$ is nondegenerate, i.e., when $B=\{i\mid\bar{A}_{i}\bar{x}=\bar{b}_{i}\}$ is of size $n$ and $\bar{A}_{B}$ is invertible, the optimal solution $\bar{\gamma}$ of the first order problem is just $\bar{\gamma}=\bar{A}_{B}^{-1}w$, where $w$ is the vector$$w=\left(\begin{array}{c} -\max_{(\Delta A,\Delta b)\in\Delta\mathcal{U}}[(\Delta A_{i_{1}})\bar{x}-\Delta b_{i_{1}}]\\ \vdots\\ -\max_{(\Delta A,\Delta b)\in\Delta\mathcal{U}}[(\Delta A_{i_{n}})\bar{x}-\Delta b_{i_{n}}]\end{array}\right),\label{eq:LP-easy}$$ where $i_{1},\dots i_{n}$ are the $n$ elements in $B$. When $\bar{x}$ is a degenerate solution, the first order problem is still easy to solve. We illustrate with a particular example that the tangential constraints are easily obtained for rectangular uncertainty sets. (Rectangular uncertainty) Suppose that the uncertainty set $\Delta\mathcal{U}$ is rectangular, that is$$\begin{aligned} \Delta\mathcal{U} & := & \big\{(\Delta A,\Delta b):|\Delta A_{j,k}|\leq\epsilon_{j,k}\mbox{, }|\Delta b_{j}|\leq\delta_{j}\\ & & \qquad\mbox{ for all }j\in\{1,\dots,m\}\mbox{, }k\in\{1,\dots,n\}\big\}.\end{aligned}$$ Then for each $i\in B$, $$\max\left\{ (\Delta A_{i})\bar{x}-\Delta b_{i}\mid(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}\right\} =\delta_{i}+\sum_{k=1}^{n}\epsilon_{i,k}|\bar{x}_{k}|.$$ In Theorem \[thm:TFO-approx\], we will discuss how an adapted first order problem gives a first order approximation of the solution to a robust optimization problem in a general setting of nonlinear programs. For now, we shall present the corollary in the simpler setting of linear programming. \[cor:lin-approx\](to Theorem \[thm:TFO-approx\]) (First order approximation in linear programming) Consider the robust optimization problem$$\begin{aligned} & & \min_{x}c^{T}x+d\nonumber \\ & & \mbox{s.t. }(\bar{A}_{i}+\Delta A_{i})x\leq(\bar{b}_{i}+\Delta b_{i})\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\epsilon\Delta\mathcal{U}_{i}\mbox{ for all }i,\label{eq:small-e-RO}\end{aligned}$$ and the first order problem $$\begin{aligned} & & \min_{\gamma}c^{T}\gamma\nonumber \\ & & \mbox{s.t. }\bar{A}_{i}\gamma\leq-\max_{(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}}[(\Delta A_{i})\bar{x}-\Delta b_{i}]\label{eq:FO_approx_lin}\\ & & \phantom{\mbox{s.t. }}\mbox{ for all }i\mbox{ s.t. }\bar{A}_{i}\bar{x}=\bar{b}_{i}.\nonumber \end{aligned}$$ Let $\bar{\Gamma}$ be the set of optimal solutions to . Suppose 1. $\Delta\mathcal{U}_{i}$ are compact convex sets. 2. $\bar{\Gamma}$ is bounded. 3. There is some $\gamma^{\prime}$ such that $\bar{A}_{i}\gamma^{\prime}<0$ whenever $\bar{A}_{i}\bar{x}=\bar{b}_{i}$. 4. $\bar{x}$ is the unique minimizer of the nominal problem $\min\{c^{T}x\mid\bar{A}x\leq\bar{b}\}$. Then the set of cluster points of any sequence $\{\frac{1}{\epsilon}(\bar{x}_{\epsilon}-\bar{x})\}$, where $\bar{x}_{\epsilon}$ is an optimal solution to and $\epsilon\to0$, is a subset of $\bar{\Gamma}$. The objective value of , say $\bar{v}_{\epsilon}$, has an approximation $\bar{v}_{\epsilon}=\bar{v}+\epsilon\tilde{v}+o(\epsilon)$, where $\tilde{v}$ is the objective value of . In particular, if $\bar{\Gamma}$ contains only one element, say $\bar{\gamma}$, then $\lim_{\epsilon\to0}\frac{1}{\epsilon}(\bar{x}_{\epsilon}-\bar{x})=\bar{\gamma}$, or $\bar{x}_{\epsilon}\in\bar{x}+\epsilon\bar{\gamma}+o(\epsilon)$. The condition that $Q$ is Clarke regular at $\bar{A}\bar{x}-\bar{b}$ holds in this case because $\mathbb{R}_{-}^{m}$ is Clarke regular everywhere. The condition that $\bar{x}$ is the unique minimizer in (4) suffices because the domain is convex. The affine function $x\mapsto c^{T}x+d$ is locally Lipschitz and subdifferentially regular everywhere. \[sec:var-analysis\]Preliminaries in variational analysis ========================================================= In this section, we recall the definitions of some nonsmooth objects in variational analysis that will be necessary for the rest of the paper. We recall the definition of normal cones and Clarke regularity. \[def:normal\](Normal cones and Clarke regularity) Let $C\subset\mathbb{R}^{n}$. For a point $\bar{x}\in C$, a vector $v$ is *normal to $C$ at $\bar{x}$ in the regular sense,* or a *regular normal*, written $v\in\hat{N}_{C}(\bar{x})$, if $$v^{T}(x-\bar{x})\leq o(|x-\bar{x}|)\mbox{ for all }x\in C.$$ It is *normal to $C$ in the general sense*, or simply a *normal vector*, written $v\in N_{C}(\bar{x})$, if there are sequences $x_{i}\to\bar{x}$ and $v_{i}\to v$ with $v_{i}\in\hat{N}_{C}(x_{i})$. The set $C$ is *Clarke regular* at $\bar{x}$ if $N_{C}(\bar{x})=\hat{N}_{C}(\bar{x})$. We refer the reader to [@RW98 Corollary 6.29] for equivalent definitions of Clarke regularity. The sets we will encounter in this paper are all Clarke regular, so this does not cause difficulties. We recall the definition of the tangent cone, which will be important in our main result. \[def:tangent\](Tangent cones) The *tangent cone* of a set $C\subset\mathbb{R}^{m}$ at some $\bar{x}\in C$ is defined by$$T_{C}(\bar{x}):=\left\{ w\mid\frac{x_{i}-\bar{x}}{t_{i}}\to w\mbox{ for some }x_{i}\in C\mbox{, }t_{i}\searrow0\mbox{ and }x_{i}\to\bar{x}\right\} .$$ Next, we recall sublinearity and equivalent definitions of subdifferential regularity that will also be useful for our main result. We take the definitions of subdifferential regularity from [@RW98 Definition 7.25, Exercise 9.15, Corollary 8.19]. (positive homogeneity and sublinearity) A function $h:\mathbb{R}^{n}\to\mathbb{R}$ is *positively homogeneous* if $h(\lambda x)=\lambda h(x)$ for all $x$ and $\lambda>0$. It is *sublinear* if in addition $$h(x+x^{\prime})\leq h(x)+h(x^{\prime})\mbox{ for all }x\mbox{ and }x^{\prime}.$$ It is clear that sublinear functions are convex. \[def:equiv-subdif-reg\](Subdifferential regularity) Let $f:\mathbb{R}^{n}\to\mathbb{R}$ be locally Lipschitz at $\bar{x}$. \(a) We say that the function $f$ is *(subdifferentially) regular at $\bar{x}$* if the epigraph ${\mbox{\rm epi}}f:=\{(x,t)\mid t\geq f(x)\}$ is Clarke regular at $(x,f(x))$ as a subset of $\mathbb{R}^{n}\times\mathbb{R}$. \(b) Define the *subderivative* $df(\bar{x}):\mathbb{R}^{n}\to\mathbb{R}$ by $$df(\bar{x})(w):=\liminf_{\tau\searrow0}\frac{f(\bar{x}+\tau w)-f(\bar{x})}{\tau}.\label{eq:subderivative}$$ and the *regular subderivative* $\hat{d}f(\bar{x}):\mathbb{R}^{n}\to\mathbb{R}$ by$$\hat{d}f(\bar{x})(w):=\limsup_{\scriptsize{\begin{array}{c} \tau\searrow0\\ x\to\bar{x}\end{array}}}\frac{f(x+\tau w)-f(x)}{\tau}.$$ In general, the regular subderivative is sublinear. The function $f$ is (subdifferentially) regular at $\bar{x}$ if and only if $df(\bar{x})=\hat{d}f(\bar{x})$. Under subdifferential regularity, it is clear that the liminf in can be taken to be a full limit. Also, ${T_{\scriptsize{\mbox{\rm epi}(f)}}}(\bar{x},f(\bar{x}))={\mbox{\rm epi}}(df(\bar{x}))$. Since the tangent cone will play a major role in our main result, we now recall some calculus rules for tangent cones, highlighting a constraint qualification condition similar to that of condition in Theorem \[thm:TFO-approx\]. The rest of this section will not be essential to the development of the paper, so one may skip to the next section in a first reading. We now recall a formula for tangent cones under intersections. \[pro:tangent-NMFCQ\](Tangent cones to intersections) Let $C=C_{1}\cap\cdots\cap C_{m}$ for closed sets $C_{i}\subset\mathbb{R}^{n}$, and let $\bar{x}\in C$. Suppose $\bar{x}$ is Clarke regular at $C_{j}$ for all $j$. Assume either$$\begin{aligned} & & \sum_{j=1}^{m}\lambda_{j}v_{j}=0\mbox{, }v_{j}\in N_{C_{j}}(\bar{x})\mbox{ and }\lambda_{j}\geq0\mbox{ for all }j\in\{1,\dots,m\}\label{eq:NMFCQ1}\\ & & \mbox{ implies }\lambda_{j}=0\mbox{ for all }j\in J^{\prime},\nonumber \end{aligned}$$ or equivalently: 1. there are no vectors $\{y_{j}\}_{j=1}^{m}$ such that $y_{j}\perp T_{C_{j}}(\bar{x})$ and $y_{1}+\cdots+y_{m}=0$ other than $y_{j}=0$ for all $j\in\{1,\dots,m\}$, and there is a vector $w$ such that $w\in\mathbb{R}^{n}\backslash\{0\}$ such that $w\in{\mbox{\rm rint}}(T_{C_{j}}(\bar{x}))$ for all $j\in\{1,\dots,m\}$. Then one has$$T_{C}(\bar{x})=T_{C_{1}}(\bar{x})\cap\cdots\cap T_{C_{m}}(\bar{x}),$$ and $C$ is Clarke regular at $\bar{x}$. Other than the equivalence of and (a), this result is stated in a more general case in [@RW98 Theorem 6.42]. This result is obtained by consider the set $D:=C_{1}\times\cdots\times C_{2}\subset(\mathbb{R}^{n})^{m}$ and the mapping $F:x\mapsto(x,\dots,x)\in(\mathbb{R}^{n})^{m}$ with $X=\mathbb{R}^{n}$ and applying [@RW98 Theorems 6.31 and 6.41]. The constraint qualification condition required is . By [@RW98 Exercise 6.39(b)], is equivalent to the existence of a $w^{\prime}$ such that $F(w^{\prime})\in{\mbox{\rm rint}}(T_{D}(\bar{x},\dots,\bar{x}))$ and having $$y\perp T_{D}(\bar{x},\dots,\bar{x})\mbox{ and }F^{*}(y)=0\mbox{ implies }y=0.$$ These conditions are equivalent to that in (a). We recall the Mangasarian-Fromovitz constraint qualification. \[def:MFCQ\](Mangasarian-Fromovitz constraint qualification) For $\mathcal{C}^{1}$ functions $f_{j}:\mathbb{R}^{n}\to\mathbb{R}$ and $j\in\{1,\dots,m\}$, let $$Q:=\big\{x\in\mathbb{R}^{n}\mid f_{j}(x)\leq0\mbox{ for all }j\in\{1,\dots,m\}\big\}.$$ For $\bar{x}\in C$, let $J^{\prime}:=\{j\mid f_{j}(\bar{x})=0\}$. The *Mangasarian-Fromovitz constraint qualification* (MFCQ) is satisfied at $\bar{x}$ if there is a vector $w\in\mathbb{R}^{n}$ such that $$\nabla f_{j}(\bar{x})^{T}w<0\mbox{ for all }j\in J^{\prime}.$$ Another equivalent definition of the MFCQ is the following “positive linear independence” condition $$\sum_{j\in J^{\prime}}\lambda_{j}\nabla f_{j}(\bar{x})=0\mbox{ and }\lambda_{j}\geq0\mbox{ for all }j\in J^{\prime}\mbox{ implies }\lambda_{j}=0\mbox{ for all }j\in J^{\prime}.$$ The classical definition of the MFCQ also takes into account equality constraints in the set $Q$, which we omit since they are not of immediate interest. To handle sets defined by nonsmooth constraints, we need to recall the subdifferential. (Subdifferentials)\[def:subdifferential\] Consider a function $f:\mathbb{R}^{n}\rightarrow\mathbb{R}$ such that $f$ is locally Lipschitz at $\bar{x}$. For a vector $v\in\mathbb{R}^{n}$, one says that \(a) $v$ is a *regular subgradient* (also known as a *Fréchet* *subgradient*) of *$f$* at $\bar{x}$, written $v\in\hat{\partial}f(\bar{x})$, if $$f(x)\geq f(\bar{x})+\left\langle v,x-\bar{x}\right\rangle +o(\left|x-\bar{x}\right|);$$ \(b) $v$ is a *(general) subgradient* of $f$ at $\bar{x}$, written $v\in\partial f(\bar{x})$, if there are sequences $x_{i}\rightarrow\bar{x}$ and $v_{i}\to v$ such that $f(x_{i})\rightarrow f(\bar{x})$ and $v_{i}\in\hat{\partial}f(x_{i})$. \(c) The set $\hat{\partial}f(\bar{x})$ ** is the *regular subdifferential,* and the set $\partial f(\bar{x})$ is the *(general) subdifferential*. \(d) The function $f$ is (subdifferentially) regular at $\bar{x}$ ** if and only if $\partial f(\bar{x})=\hat{\partial}f(\bar{x})$. This characterization of subdifferentially regular functions is slightly different from the earlier definitions, but is equivalent in the case of locally Lipschitz functions in view of [@RW98 Corollary 8.11, Theorem 9.13 and Theorem 8.6]. We shall only be concerned with subdifferentially regular functions throughout this paper, so there is no need to distinguish between $\partial f(\bar{x})$ and $\hat{\partial}f(\bar{x})$. We conclude with results on the intersections of tangent cones described by constraints. \[pro:tangent-examples\](Tangent cone under constraints) Suppose $C=\{x\mid f_{j}(x)\leq0,j\in J\}$, and $J$ is a finite set. At the point $\bar{x}\in C$, let $J^{\prime}\subset J$ be the set of all $j$’s such that $f_{j}(\bar{x})=0$. If $f_{j}$ are continuous at $\bar{x}$ for all $j\in J$, $f_{j}$ are continuously differentiable at $\bar{x}$ for all $j\in J^{\prime}$ and the MFCQ is satisfied at $\bar{x}\in C$, then$$T_{C}(\bar{x})=\{z\mid\nabla f_{j}(\bar{x})^{T}z\leq0\mbox{ for all }j\in J^{\prime}\}.$$ In the nonsmooth case, if $f_{j}$ were locally Lipschitz and subdifferentially regular at $\bar{x}$ for all $j\in J^{\prime}$ and $$\begin{aligned} & & \sum_{j\in J^{\prime}}\lambda_{j}v_{j}=0\mbox{, }v_{j}\in\partial f_{j}(\bar{x})\mbox{ and }\lambda_{j}\geq0\mbox{ for all }j\in J^{\prime}\label{eq:nonsmooth-MFCQ}\\ & & \mbox{ implies }\lambda_{j}=0\mbox{ for all }j\in J^{\prime},\nonumber \end{aligned}$$ then $\bar{x}$ is Clarke regular at $C$, and $$\begin{aligned} T_{C}(\bar{x}) & = & \bigcap_{j\in J^{\prime}}\{z\mid v^{T}z\leq0\mbox{ for all }v\in\partial f_{j}(\bar{x})\}.\label{eq:Tangent_Qi_formula}\\ & = & \{z\mid v^{T}z\leq0\mbox{ for all }v\in\bigcup_{j\in J^{\prime}}\partial f_{j}(\bar{x})\}.\nonumber \end{aligned}$$ We prove the general nonsmooth case for this theorem, which implies the smooth case. There is a neighborhood $U$ of $\bar{x}$ such that $C\cap U=[\cap_{j\in J^{\prime}}C_{j}]\cap U$, where $C_{j}$ is defined by $C_{j}=\{x\mid f_{j}(x)\leq0\}$. Furthermore, $0\notin\partial f_{j}(\bar{x})$ for all $j\in J^{\prime}$. By [@RW98 Theorem 10.3] (normal cones to level sets) and [@RW98 Corollary 6.29(d)] (tangent-normal relations in regular sets), $C_{j}$ is Clarke regular at $\bar{x}$, and the tangent cones $T_{C_{j}}(\bar{x})$ and normal cones $N_{C_{j}}(\bar{x})$ are given by $$\begin{aligned} N_{C_{j}}(\bar{x}) & = & \{\lambda v\mid\lambda\geq0\mbox{, and }v\in\partial f_{j}(\bar{x})\},\\ \mbox{ and }T_{C_{j}}(\bar{x}) & = & \{w\mid w^{T}v\leq0\mbox{ for all }v\in\partial f_{j}(\bar{x})\}.\end{aligned}$$ Therefore, condition becomes$$\sum_{j\in J^{\prime}}v_{j}=0\mbox{, }v_{j}\in N_{C_{j}}(\bar{x})\mbox{ implies }v_{j}=0\mbox{ for all }j\in J^{\prime}.$$ By Proposition \[pro:tangent-NMFCQ\], the tangent cone $T_{C}(\bar{x})$ is $$\begin{aligned} T_{C}(\bar{x}) & = & \bigcap_{i\in J^{\prime}}T_{C_{j}}(\bar{x}),\end{aligned}$$ which gives the formula for the tangent cone in the statement. It is well known that for the sets $$\begin{aligned} C_{1} & := & \big\{(x_{1},x_{2})\in\mathbb{R}^{2}:x_{2}\geq x_{1}^{2}\big\}\\ C_{2} & := & \big\{(x_{1},x_{2})\in\mathbb{R}^{2}:x_{2}\leq-x_{1}^{2}\big\},\end{aligned}$$ we have $T_{C_{1}\cap C_{2}}(0)\subsetneq T_{C_{1}}(0)\cap T_{C_{2}}(0)$ but the MFCQ is not satisfied. The constraint qualification condition can be checked by another equivalent condition when $T_{C_{j}}(\bar{x})$ have nonempty interior. \[pro:CQ\](Constraint qualification) Assume the conditions of Proposition \[pro:tangent-examples\]. If $T_{C_{j}}(\bar{x})$ have nonempty interior and $0\notin\partial f_{j}(\bar{x})$ for all $j\in J^{\prime}$, then the condition is equivalent to the existence of a vector $w$ such that $w\in{\mbox{\rm int}}(T_{C_{j}}(\bar{x}))$ (or equivalently $w^{T}v<0$ for all $v\in\partial f_{j}(\bar{x})$) for all $j\in J^{\prime}$. Recall that by [@RW98 Theorem 10.3], if $0\notin\partial f(\bar{x})$, then the tangent cone $T_{C_{j}}(\bar{x})$ is equal to $\{z\mid z^{T}v\leq0\mbox{ for all }v\in\partial f_{j}(\bar{x})\}$, and the interior ${\mbox{\rm int}}(T_{C_{j}}(\bar{x}))$ is $\{z\mid z^{T}v<0\mbox{ for all }v\in\partial f_{j}(\bar{x})\}$, which gives the equivalence on the conditions on $w$. Next, the equivalence of and the condition in this result follow from Proposition \[pro:tangent-NMFCQ\]. \[sec:Robust-nonlin\]Robust nonlinear programming ================================================= We look at nonlinear programs of the form$$\min_{x}\{c^{T}x+d\mid\bar{A}x-\bar{b}\in Q\},\label{eq:RO_1}$$ where $Q\subset\mathbb{R}^{k}$ is a closed set. Specifically, we consider problems of the form$$\min_{x}\{c^{T}x+d\mid\bar{A}_{i}x-\bar{b}_{i}\in Q_{i},\,1\leq i\leq m\},\label{eq:RO_2}$$ where $Q_{i}\subset\mathbb{R}^{k_{i}}$ are nonempty closed sets, $\bar{A}_{i}\in\mathbb{R}^{k_{i}\times n}$, and $\bar{b}_{i}\in\mathbb{R}^{k_{i}}$. We may write $\bar{A}$ as a concatenation of the matrices $\bar{A}_{i}$ and $\bar{b}$ as a concatenation of the vectors $\bar{b}_{i}$, and this would make equivalent to for $Q=Q_{1}\times\cdots\times Q_{m}$ and $k=k_{1}+\cdots+k_{m}$. One case of interest is the set $Q_{i}=\{y\mid f_{i,j}(y)\leq0\mbox{ for all }j\in J\}$ for some $f_{i,j}:\mathbb{R}^{k_{i}}\to\mathbb{R}$ and the set $J$ is finite. Another case of interest is *conic programs*, which arise when all $Q_{i}$’s are closed convex pointed cones with nonempty interior. We now recall the definition of robust feasibility from [@BGN09]. \[def:Rob-feas\](Robust feasibility) Let an uncertain problem be given and $\Delta\mathcal{U}=\Delta\mathcal{U}_{1}\times\cdots\times\Delta\mathcal{U}_{m}$ be a perturbation set. A candidate solution $x\in\mathbb{R}^{n}$ is *robustly feasible* if it remains feasible for all realizations of the perturbation vector from the perturbation set, that is $$[\bar{A}_{i}+\Delta A_{i}]x-[\bar{b}_{i}+\Delta b_{i}]\in Q_{i}\,\forall(i,1\leq i\leq m,(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}),\label{eq:conic-feasible}$$ where $\Delta\mathcal{U}_{i}\subset(\mathbb{R}^{k_{i}\times n}\times\mathbb{R}^{k_{i}})$ is the uncertainty set in $(\bar{A}_{i},\bar{b}_{i})$. (Decomposing uncertainty sets) In the case where $\Delta\mathcal{U}$ is not a direct product of uncertainty sets, the uncertainty sets $\Delta\mathcal{U}_{i}$ can be defined as $$\Delta\mathcal{U}_{i}:=\{(\Delta A_{i},\Delta b_{i}):(\Delta A_{i},\Delta b_{i})=\Pi_{i}(\Delta A,\Delta b)\mbox{ for some }(\Delta A,\Delta b)\in\Delta\mathcal{U}\},$$ where $\Pi_{i}$ is the relevant projection from $\mathbb{R}^{k\times n}\times\mathbb{R}^{k}$ to $\mathbb{R}^{k_{i}\times n}\times\mathbb{R}^{k_{i}}$. It is clear that is equivalent to $$[\bar{A}+\Delta A]x-[\bar{b}+\Delta b]\in Q\,\forall(\Delta A,\Delta b)\in\Delta\mathcal{U}.$$ The definition for robust nonlinear programs encompasses nonlinear objective functions. \[exa:nonlin-obj\](Nonlinear objective) Consider the robust optimization problem$$\begin{aligned} & & \min_{x}f(x)\\ & & \mbox{s.t. }[\bar{A}_{i}+\Delta A_{i}]x-[\bar{b}_{i}+\Delta b_{i}]\in Q_{i}\,\forall(i,1\leq i\leq m,(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}).\end{aligned}$$ We can rewrite this robust problem as $$\begin{aligned} & & \min_{x,t}t\\ & & \mbox{s.t. }[\bar{A}_{i}+\Delta A_{i}]x-[\bar{b}_{i}+\Delta b_{i}]\in Q_{i}\,\forall(i,1\leq i\leq m,(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i})\\ & & \mbox{and }f(x)\leq t.\end{aligned}$$ The function $f$ is convex if and only if the epigraph ${\mbox{\rm epi}}(f)=\{(x,t)\mid f(x)\leq t\}$ is convex. Similarly, for a function $f$ locally Lipschitz at $\bar{x}$, the function $f$ is subdifferentially regular at $\bar{x}$ if and only if ${\mbox{\rm epi}}(f)$ is Clarke regular at $\bar{x}$. To prove our results for nonlinear functions, we can prove the result for linear objective functions and then appeal to the second formulation to obtain the result we need. The formula in the robust optimization constraint can be rewritten as $$\begin{aligned} & & [\bar{A}_{i}+\Delta A_{i}](\bar{x}+\Delta x)-[\bar{b}_{i}+\Delta b_{i}]\in Q_{i}\\ & \iff & [\bar{A}_{i}\bar{x}-\bar{b}_{i}]+\bar{A}_{i}(\Delta x)+[(\Delta A_{i})\bar{x}-\Delta b_{i}]+(\Delta A_{i})(\Delta x)\in Q_{i}.\end{aligned}$$ As in linear programming, we eliminate the second order term $(\Delta A)(\Delta x)$ to obtain a first order approximation. For nonlinear programs, we also need to approximate the set $Q_{i}$ at $\bar{A}_{i}\bar{x}-\bar{b}_{i}$ by the tangential approximation $T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})+[\bar{A}_{i}\bar{x}-\bar{b}_{i}]$ at $\bar{A}_{i}\bar{x}-\bar{b}_{i}$. This gives our definition of the tangential problem. \[def:tangent-pblm\](Tangential problem) Let $\bar{x}$ be an optimal solution to a nonlinear programming problem with parameters $(\bar{A},\bar{b})$ so that $Q_{i}$ is Clarke regular at $\bar{A}_{i}\bar{x}-\bar{b}_{i}$ for all $i$. The *tangential problem* to the robust optimization problem obtained with constraints as explained in Definition \[def:Rob-feas\] is$$\begin{aligned} & & \min_{\gamma}c^{T}\gamma\\ & & \mbox{s.t. }[\bar{A}_{i}\bar{x}-\bar{b}_{i}]+\bar{A}_{i}\gamma+[(\Delta A_{i})\bar{x}-\Delta b_{i}]\in T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})+[\bar{A}_{i}\bar{x}-\bar{b}_{i}]\\ & & \phantom{\mbox{s.t. }}\mbox{ for all }(i,1\leq i\leq m,(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}),\end{aligned}$$ or equivalently$$\begin{aligned} & & \min_{\gamma}c^{T}\gamma\label{eq:tangential-FO}\\ & & \mbox{s.t. }\bar{A}_{i}\gamma+[(\Delta A_{i})\bar{x}-\Delta b_{i}]\in T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})\nonumber \\ & & \phantom{\mbox{s.t. }}\mbox{ for all }(i,1\leq i\leq m,(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}),\nonumber \end{aligned}$$ which is also equivalent to $$\begin{aligned} & & \min_{\gamma}c^{T}\gamma\\ & & \mbox{s.t. }\bar{A}_{i}\gamma+L_{i}(\Delta\mathcal{U}_{i})\in T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})\\ & & \phantom{\mbox{s.t. }}\mbox{ for all }(i,1\leq i\leq m),\end{aligned}$$ where $L_{i}:\mathbb{R}^{m_{i}\times n}\times\mathbb{R}^{m_{i}}\to\mathbb{R}^{m_{i}}$ is defined by $L_{i}(\Delta A_{i},\Delta b_{i})=(\Delta A_{i})\bar{x}-\Delta b_{i}$. We call the corresponding constraints to the tangential problem the *tangential constraints*. (Clarke regularity assumption) The assumption that each $Q_{i}$ is Clarke regular at $\bar{A_{i}}\bar{x}-\bar{b}_{i}$ in Definition \[def:tangent-pblm\] comes about because the set $Q_{1}\times\cdots\times Q_{m}$ is Clarke regular at $\bar{A}\bar{x}-\bar{b}$ if and only if $Q_{i}$ is Clarke regular at $\bar{A}_{i}\bar{x}-\bar{b}_{i}$ for all $i$, and in this case,$$T_{Q_{1}\times\cdots\times Q_{m}}(\bar{A}\bar{x}-\bar{b})=T_{Q_{1}}(\bar{A}_{1}\bar{x}-\bar{b}_{1})\times\cdots\times T_{Q_{m}}(\bar{A}_{m}\bar{x}-\bar{b}_{m}).$$ (see [@RW98 Proposition 6.41].) This property makes the tangential problem independent of how we decompose the set $Q$ as a direct product of sets. We give some examples of tangential constraints. \[exa:TFO\](Examples of tangential constraints) (a) When $\bar{A}_{i}\bar{x}-\bar{b}_{i}=0$ and $Q_{i}$ is a closed convex cone, then $T_{Q_{i}}(0)=Q_{i}$. In this case, the corresponding tangential constraint is obtained by just removing the second order term $(\Delta A_{i})(\Delta x)$. \(b) When $\bar{A}_{i}\bar{x}-\bar{b}_{i}\in{\mbox{\rm int}}(Q_{i})$, then $T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})=\mathbb{R}^{k_{i}}$ and the corresponding tangential constraint vanishes. In view of Example \[exa:TFO\], we see that for linear programming, the tangential constraints and first order constraints are equivalent. When $\bar{A}_{i}\bar{x}-\bar{b}_{i}\in\partial Q_{i}\backslash\{0\}$, we may still be able to calculate the tangential constraints using the material recalled in Section \[sec:var-analysis\]. We illustrate the tangential problem with the example on second order cone programming (SOCP). \[exa:SOCP-easy\](Tangential problem in SOCP) Consider the SOCP problem$$\begin{aligned} & & \min_{x}c^{T}x\\ & & \mbox{s.t. }\bar{A}_{i}x-\bar{b}_{i}\in Q_{k_{i}}\mbox{ for all }1\leq i\leq m,\end{aligned}$$ where $Q_{d}\subset\mathbb{R}^{d}$ is the *second order cone* $$Q_{d}:=\{w=(w_{0},\dots,w_{d-1})\in\mathbb{R}^{d}\mid\|(w_{1},\dots,w_{d-1})\|_{2}-w_{0}\leq0\}.$$ Given an optimal solution $\bar{x}$, we show how to obtain the tangential constraint. If $\bar{A}_{i}\bar{x}-\bar{b}_{i}\in{\mbox{\rm int}}(Q_{k_{i}})$, then $T_{Q_{k_{i}}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})=\mathbb{R}^{k_{i}}$ by Example \[exa:TFO\], and so the tangential constraint vanishes. If $\bar{A}_{i}\bar{x}-\bar{b}_{i}=0$, then $T_{Q_{k_{i}}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})=Q_{k_{i}}$ by Example \[exa:TFO\], so the tangential constraint is $$\bar{A}_{i}\gamma+L_{i}(\Delta\mathcal{U}_{i})\subset Q_{k_{i}}.$$ We now consider the case $\bar{A}_{i}\bar{x}-\bar{b}_{i}\in\partial Q_{k_{i}}\backslash\{0\}$. Let $\bar{z}=\bar{A}_{i}\bar{x}-\bar{b}_{i}$. In this case, $\bar{z}_{0}=\|(\bar{z}_{1},\dots,\bar{z}_{k_{i}-1})\|_{2}$. The gradient of the map $(w_{0},w_{1},\dots,w_{k_{i}-1})\mapsto\|(w_{1},\dots,w_{k_{i}-1})\|_{2}-w_{0}$ at $\bar{z}$ is $(-1,\frac{(\bar{z}_{1},\dots,\bar{z}_{k_{i}-1})}{\|(\bar{z}_{1},\dots,\bar{z}_{k_{i}-1})\|_{2}})$. Let $R:\mathbb{R}^{k_{i}}\to\mathbb{R}^{k_{i}}$ be the reflection map $$R(w_{0},w_{1},\dots,w_{k_{i}-1}):=(-w_{0},w_{1},\dots,w_{k_{i}-1}),$$ i.e., $R$ multiplies the $0$th coordinate by $-1$. The gradient at $\bar{z}$ can also be written as $\frac{1}{\bar{z}_{0}}R\bar{z}$. Therefore, by Proposition \[pro:tangent-examples\], $$T_{Q_{k_{i}}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})=\{w\in\mathbb{R}^{k_{i}}\mid w^{T}R\bar{z}\leq0\}.$$ Therefore, the tangential constraint is$$\bar{A}_{i}\gamma+[(\Delta A_{i})\bar{x}-\Delta b_{i}]\in T_{Q_{k_{i}}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}.$$ This can be written equivalently as $$\begin{aligned} & & \bar{z}^{T}R\bar{A}_{i}\gamma+\bar{z}^{T}R[(\Delta A_{i})\bar{x}-\Delta b_{i}]\leq0\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i},\\ & \mbox{or } & \bar{z}^{T}R\bar{A}_{i}\gamma\leq-\max_{(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}}\bar{z}^{T}R[(\Delta A_{i})\bar{x}-\Delta b_{i}].\end{aligned}$$ \[sec:Main-result\]Main result: Approximation using the tangential problem ========================================================================== In Theorem \[thm:TFO-approx\] we prove that if the uncertainty set in a robust optimization problem is dilated or expanded, then the robust optimal solution can be predicted from the exact solution of the nonrobust problem and the tangential problem. We now prove a lemma needed for the proof of our main result. \[lem:compex-sets-in-intr\] (Compact sets in convex cones) Let $D\subset\mathbb{R}^{n}$ be Clarke regular at $0$ and $C\subset\mathbb{R}^{n}$ be a compact convex set such that ${\mbox{\rm int}}(T_{D}(0))\neq\emptyset$ and $C\subset T_{D}(0)$. Let $v\in{\mbox{\rm int}}(T_{D}(0))$. Then for all sufficiently small $\delta>0$, there exists $\bar{\epsilon}>0$ such that $C+\delta v+\delta^{2}\mathbb{B}\subset\frac{1}{\epsilon}D$ for all $\epsilon\in(0,\bar{\epsilon}]$. Since $v\in{\mbox{\rm int}}(T_{D}(0))$, $v+\delta\mathbb{B}\subset{\mbox{\rm int}}(T_{D}(0))$ for all sufficiently small $\delta>0$, and therefore $C+\delta v+\delta^{2}\mathbb{B}\subset{\mbox{\rm int}}(T_{D}(0))$. For every point $w\in C+\delta v+\delta^{2}\mathbb{B}$, we can find a convex polyhedral set $P_{w}$ such that ${\mbox{\rm int}}(P_{w})\neq\emptyset$ and $P_{w}\subset{\mbox{\rm int}}(T_{D}(0))$. A compactness argument shows that the set $C+\delta v+\delta^{2}\mathbb{B}$ is contained in the interior of finitely many of these convex polyhedral sets, so there is a convex polyhedral set $P$ such that $C+\delta v+\delta^{2}\mathbb{B}\subset P\subset{\mbox{\rm int}}(T_{D}(0))$. By the recession properties of tangent cones and the Clarke regularity of $D$, there is an $\bar{\epsilon}>0$ such that $\epsilon\mbox{conv}(\{0\}\cup P)\subset D$ for all $\epsilon\in[0,\bar{\epsilon}]$ (see [@RW98 Exercise 6.34(a)]. The roots of this result on local recession vectors can be traced back to [@R79].). Therefore $C+\delta v+\delta^{2}\mathbb{B}\subset\mbox{conv}(\{0\}\cup P)\subset\frac{1}{\epsilon}D$ for all $\epsilon\in[0,\bar{\epsilon}]$ as needed. We also need material in set-valued analysis as presented in [@RW98 Chapters 4 and 5] for the proof of Theorem \[thm:TFO-approx\]. [@RW98 Definition 5.4] (Set-valued continuity) We say that $S$ is a *set-valued map*, denoted by $S:\mathbb{R}^{n}\rightrightarrows\mathbb{R}^{m}$, if $S(x)\subset\mathbb{R}^{m}.$ A set-valued map $S$ is *outer semicontinuous (osc)* at $\bar{x}$ if $$\limsup_{x\to\bar{x}}S(x)\subset S(\bar{x}),$$ or equivalently $\limsup_{x\to\bar{x}}S(x)=S(\bar{x})$, but *inner semicontinuous (isc)* at $\bar{x}$ if $$\liminf_{x\to\bar{x}}S(x)\supset S(\bar{x}),$$ or equivalently when $S$ is closed-valued, $\liminf_{x\to\bar{x}}S(x)=S(\bar{x})$. It is called *continuous* at $\bar{x}$ if both conditions hold, i.e., if $S(x)\to S(\bar{x})$ as $x\to\bar{x}$. Here, the *outer limit* $\limsup_{x\to\bar{x}}S(x)$ and the *inner limit* $\liminf_{x\to\bar{x}}S(x)$ are defined by $$\begin{aligned} \limsup_{x\to\bar{x}}S(x) & := & \left\{ u\mid\exists x_{i}\to\bar{x},\exists u_{i}\to u\mbox{ with }u_{i}\in S(x_{i})\right\} \\ \liminf_{x\to\bar{x}}S(x) & := & \big\{u\mid\forall x_{i}\to\bar{x},\exists\{u_{i}\}\mbox{ with }u_{i}\in S(x_{i})\\ & & \quad\mbox{ s.t. }u\mbox{ is the limit of a subsequence of }\{u_{i}\}\big\}.\end{aligned}$$ If $S$ maps to compact sets, continuity as defined by inner and outer limits above is equivalent to continuity in the Pompieu-Hausdorff distance, which is a metric in the subset of compact sets. We refer to [@RW98] for more details. We also need to recall the definition of epi-convergence. A sequence of functions $h_{i}:\mathbb{R}^{n}\to\mathbb{R}$ is said to *epi-converge* to a function $h:\mathbb{R}^{n}\to\mathbb{R}$, written $h_{i}\xrightarrow{e}h$, if ${\mbox{\rm epi}}(h_{i})\to{\mbox{\rm epi}}(h)$. The history of epi-convergence can be traced back to the 1960’s, and the result we need for our proof ([@RW98 Theorem 7.33]) can be traced back to Salinetti (unpublished, but reported in [@RW84]) and [@AW81]. See [@RW98 Chapter 7]. Here is our theorem on the approximation properties of the tangential problem. \[thm:TFO-approx\](Approximation properties of Tangential problem) Consider the robust optimization problem$$\begin{aligned} & & \min_{x}\, f(x)\nonumber \\ & & \mbox{s.t. }[\bar{A}_{i}+\Delta A_{i}]x-[\bar{b}_{i}+\Delta b_{i}]\in Q_{i}\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\epsilon\Delta\mathcal{U}_{i}\mbox{ for all }i,\label{eq:RO_approx_1}\end{aligned}$$ and the tangential problem $$\begin{aligned} & & \min_{\gamma}\, df(\bar{x})(\gamma)\nonumber \\ & & \mbox{s.t. }\bar{A}_{i}\gamma+[(\Delta A_{i})\bar{x}-\Delta b_{i}]\in T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})\label{eq:FO_approx_1}\\ & & \mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}\mbox{ for all }i.\nonumber \end{aligned}$$ Let $\bar{\Gamma}$ be the set of optimal solutions to , $\Phi=\{x\mid\bar{A}x-\bar{b}\in Q\}$, and $\bar{x}$ be a solution of the nominal problem $\min\{f(x)\mid\bar{A}x-\bar{b}\in Q\}$. Suppose 1. $Q_{i}$ are closed sets that are Clarke regular at $\bar{A}_{i}\bar{x}-\bar{b}_{i}$, 2. $\Delta\mathcal{U}_{i}$ are compact convex sets 3. $\bar{\Gamma}$ is bounded. 4. \[enu:CQ\]There is some $\gamma^{\prime}$ such that $\bar{A}_{i}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i}))$ for all $i$. 5. (Compactness) If $\{x_{i}\}\subset\Phi$ and $f(x_{i})\to f(\bar{x})$, then $x_{i}\to\bar{x}$. 6. $f$ is locally Lipschitz and subdifferentially regular at $\bar{x}$. Then the set of cluster points of any sequence $\{\frac{1}{\epsilon}(\bar{x}_{\epsilon}-\bar{x})\}$, where $\bar{x}_{\epsilon}$ is an optimal solution to and $\epsilon\searrow0$, is a subset of $\bar{\Gamma}$. The objective value of , say $\bar{v}_{\epsilon}$, has an approximation $\bar{v}_{\epsilon}=\bar{v}+\epsilon\tilde{v}+o(\epsilon)$, where $\tilde{v}$ is the objective value of . In particular, if $\bar{\Gamma}$ contains only one element, say $\bar{\gamma}$, then $\lim_{\epsilon\searrow0}\frac{1}{\epsilon}(\bar{x}_{\epsilon}-\bar{x})=\bar{\gamma}$, or $\bar{x}_{\epsilon}\in\bar{x}+\epsilon\bar{\gamma}+o(\epsilon)$. The proof of this result is broken up into four steps. In steps 1 to 3, we prove this result for the affine function $f(x)=c^{T}x+d$, and $df(\bar{x})(\gamma)=c^{T}x$. In step 4, we use the observation in Example \[exa:nonlin-obj\] to treat the case where $f$ is locally Lipschitz and subdifferentially regular at $\bar{x}$. **Step 1: Rewriting the robust optimization problem .** We rewrite the constraint in the robust optimization problem.$$\begin{aligned} & & [\bar{A}_{i}+\Delta A_{i}]x-[\bar{b}_{i}+\Delta b_{i}]\in Q_{i}\\ & \Leftrightarrow & [\bar{A}_{i}+\Delta A_{i}]x-\bar{A}_{i}\bar{x}-\Delta b_{i}\in Q_{i}-[\bar{A}_{i}\bar{x}-\bar{b}_{i}]\\ & \Leftrightarrow & \bar{A}_{i}(x-\bar{x})+[(\Delta A_{i})\bar{x}-\Delta b_{i}]+(\Delta A_{i})(x-\bar{x})\in Q_{i}-[\bar{A}_{i}\bar{x}-\bar{b}_{i}].\end{aligned}$$ Hence, $$\begin{aligned} & & A_{i}x-b_{i}\in Q_{i}\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\epsilon\Delta\mathcal{U}_{i}\\ & \Leftrightarrow & \bar{A}_{i}(x-\bar{x})+[(\Delta A_{i})\bar{x}-\Delta b_{i}]+(\Delta A_{i})(x-\bar{x})\in Q_{i}-[\bar{A}_{i}\bar{x}-\bar{b}_{i}]\\ & & \quad\quad\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\epsilon\Delta\mathcal{U}_{i}.\end{aligned}$$ The next step is to scale the variables $\Delta A_{i}$ and $\Delta b_{i}$ so that the $\epsilon$ vanishes from the expression $\epsilon\Delta\mathcal{U}_{i}$. This gives$$\begin{aligned} & & \bar{A}_{i}(x-\bar{x})+[(\Delta A_{i})\bar{x}-\Delta b_{i}]+(\Delta A_{i})(x-\bar{x})\in Q_{i}-[\bar{A}_{i}\bar{x}-\bar{b}_{i}]\nonumber \\ & & \quad\quad\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\epsilon\Delta\mathcal{U}_{i}\nonumber \\ & \Leftrightarrow & \bar{A}_{i}\Big[\frac{1}{\epsilon}(x-\bar{x})\Big]+[(\Delta A_{i})\bar{x}-\Delta b_{i}]+\epsilon(\Delta A_{i})\Big[\frac{1}{\epsilon}(x-\bar{x})\Big]\nonumber \\ & & \quad\quad\in\frac{1}{\epsilon}\big[Q_{i}-[\bar{A}_{i}\bar{x}-\bar{b}_{i}]\big]\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}\nonumber \\ & \Leftrightarrow & \bar{A}_{i}\gamma_{\epsilon}+[(\Delta A_{i})\bar{x}-\Delta b_{i}]+\epsilon(\Delta A_{i})\gamma_{\epsilon}\label{eq:rob_approx_1}\\ & & \quad\quad\in\frac{1}{\epsilon}\big[Q_{i}-[\bar{A}_{i}\bar{x}-\bar{b}_{i}]\big]\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i},\nonumber \end{aligned}$$ where $\gamma_{\epsilon}:=\frac{1}{\epsilon}(x-\bar{x})$ in the final expression. We see that as $\epsilon\searrow0$, the expressions in converge to the corresponding expressions for the tangential constraints. Let $\Gamma_{\epsilon}$ denote the set of all feasible $\gamma_{\epsilon}$ for the robust problem with parameter $\epsilon$, and $\Gamma$ denote the set of all feasible $\gamma$ for the tangential problem. Similarly, let $\bar{\Gamma}_{\epsilon}$ and $\bar{\Gamma}$ denote the set of optimal solutions to the corresponding problems. It is elementary to check that the sets $\Gamma_{\epsilon}$, $\Gamma$, $\bar{\Gamma}_{\epsilon}$ and $\bar{\Gamma}$ are all closed. **Step 2: $\lim_{\epsilon\searrow0}\Gamma_{\epsilon}=\Gamma$.** Suppose that $\{\gamma_{j}\}$ is a sequence of feasible solutions to the robust problem with parameter $\epsilon_{j}$, that is $\gamma_{j}\in\Gamma_{\epsilon_{j}}$. Then each $\gamma_{j}$ satisfies the formula in with parameter $\epsilon_{j}$. It is clear that any limit of $\{\gamma_{j}\}$ is a feasible solution of the tangential problem , so $\limsup_{\epsilon\searrow0}\Gamma_{\epsilon}\subset\Gamma$. Next, we show that $\liminf_{\epsilon\searrow0}\Gamma_{\epsilon}\supset\Gamma$. Suppose $\tilde{\gamma}\in\Gamma$. We need to show that for any choice of $\epsilon_{j}\searrow0$, we can find $\gamma_{j}\in\Gamma_{\epsilon_{j}}$ such that $\tilde{\gamma}=\lim_{j\to\infty}\gamma_{j}$. Recall that $\tilde{\gamma}$ satisfies $$\bar{A}_{i}\tilde{\gamma}+L_{i}(\Delta\mathcal{U}_{i})\subset T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i}),$$ where the linear map $L_{i}:\mathbb{R}^{m_{i}\times n}\times\mathbb{R}^{m_{i}}\to\mathbb{R}^{m_{i}}$ by $L_{i}(\Delta A_{i},\Delta b_{i})=(\Delta A_{i})\bar{x}-\Delta b_{i}$. It follows from the convexity of $\Delta\mathcal{U}_{i}$ that $L_{i}(\Delta\mathcal{U}_{i})$ is convex. Since $\bar{A}_{i}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i}))$, we can apply Lemma \[lem:compex-sets-in-intr\] to tell us that for all sufficiently small $\delta>0$, there is some $\bar{\epsilon}>0$ such that $[\bar{A}_{i}(\tilde{\gamma}+\delta\gamma^{\prime})+L_{i}(\Delta\mathcal{U}_{i})+\delta^{2}\mathbb{B}]\subset\frac{1}{\epsilon}[Q_{i}-(\bar{A}_{i}\bar{x}-\bar{b}_{i})]$ for all $\epsilon\in[0,\bar{\epsilon}]$. Therefore $$\bar{A}_{i}(\tilde{\gamma}+\delta\gamma^{\prime})+L_{i}(\Delta\mathcal{U}_{i})+\delta^{2}\mathbb{B}\subset\frac{1}{\epsilon}\big[Q_{i}-[\bar{A}_{i}\bar{x}-\bar{b}_{i}]\big].$$ If $\epsilon<\bar{\epsilon}$ and $\epsilon\max\{\|\Delta A_{i}\|\mid(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}\}\|(\tilde{\gamma}+\delta\gamma^{\prime})\|<\delta^{2}$ , then for all $(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}$, $$\begin{aligned} & & \bar{A}_{i}(\tilde{\gamma}+\delta\gamma^{\prime})+[(\Delta A_{i})\bar{x}-\Delta b_{i}]+\epsilon(\Delta A_{i})(\tilde{\gamma}+\delta\gamma^{\prime})\label{eq:Aw-chain}\\ & \in & \bar{A}_{i}(\tilde{\gamma}+\delta\gamma^{\prime})+L_{i}(\Delta\mathcal{U}_{i})+\epsilon\|\Delta A_{i}\|\|(\tilde{\gamma}+\delta\gamma^{\prime})\|\mathbb{B}\nonumber \\ & \subset & \bar{A}_{i}(\tilde{\gamma}+\delta\gamma^{\prime})+L_{i}(\Delta\mathcal{U}_{i})+\delta^{2}\mathbb{B}\nonumber \\ & \subset & \frac{1}{\epsilon}\big[Q_{i}-[\bar{A}_{i}\bar{x}-\bar{b}_{i}]\big].\nonumber \end{aligned}$$ With this observation, we can choose a sequence $\delta_{j}\searrow0$ such that $(\tilde{\gamma}+\delta_{j}\gamma^{\prime})\in\Gamma_{\epsilon_{j}}$, which gives $\liminf_{\epsilon\searrow0}\Gamma_{\epsilon}\supset\Gamma$ as needed. **Step 3: $\limsup_{\epsilon\searrow0}\bar{\Gamma}_{\epsilon}\subset\bar{\Gamma}$.** Recall that for a closed set $D\subset\mathbb{R}^{n}$, the *indicator function* $\delta_{D}:\mathbb{R}^{n}\to\mathbb{R}\cup\{\infty\}$ is defined by $$\delta_{D}(x):=\begin{cases} 0 & \mbox{if }x\in D\\ \infty & \mbox{otherwise.}\end{cases}$$ Define $h_{\epsilon}:\mathbb{R}^{n}\to\mathbb{R}$ and $h:\mathbb{R}^{n}\to\mathbb{R}$ by $h_{\epsilon}(\gamma):=c^{T}\gamma+\delta_{\Gamma_{\epsilon}}(\gamma)$ and $h(\gamma):=c^{T}\gamma+\delta_{\Gamma}(\gamma)$. Since $\Gamma_{\epsilon}\to\Gamma$, we have ${\mbox{\rm epi}}(\delta_{\Gamma_{\epsilon}})=\Gamma_{\epsilon}\times[0,\infty)\to\Gamma\times[0,\infty)={\mbox{\rm epi}}(\delta_{\Gamma})$ (by [@RW98 Exercise 4.29(a)]), so in other words $\delta_{\Gamma_{\epsilon}}\xrightarrow{e}\delta_{\Gamma}$. By [@RW98 Exercise 7.8(a)] we have $h_{\epsilon}\xrightarrow{e}h$. We seek to apply [@RW98 Theorem 7.33], which gives us the result we need. Before we can do so, we have to check that $\{h_{\epsilon}\}$ is eventually level bounded, that is, for any $\alpha\in\mathbb{R}$, we have $\cup_{[0,\epsilon]}\{\gamma\mid h_{\epsilon}(\gamma)\leq\alpha\}$ being bounded for some $\epsilon>0$. Recall $\Phi=\{x\mid\bar{A}x-\bar{b}\in Q\}$. By Proposition \[pro:Condn-Phi\], the boundedness of $\bar{\Gamma}$ is equivalent to $\{c\}^{\perp}\cap T_{\Phi}(\bar{x})=\{0\}$. Suppose on the contrary that $\{h_{\epsilon}\}$ is not eventually level bounded. Then there is some $\alpha$ and sequences $\{\epsilon_{i}\}$ and $\{\gamma_{i}\}$ such that $\epsilon_{i}\searrow0$, $\{\gamma_{i}\}$ is unbounded, and $h_{\epsilon_{i}}(\gamma_{i})\leq\alpha$ (or equivalently, $\gamma_{i}\in\Gamma_{\epsilon_{i}}$ and $c^{T}\gamma_{i}\leq\alpha$). Let us write $x_{i}=\bar{x}+\epsilon_{i}\gamma_{i}$. Since $x_{i}\in\Phi$ and $$c^{T}\bar{x}\leq c^{T}x_{i}=c^{T}\bar{x}+\epsilon_{i}c^{T}\gamma_{i}\leq c^{T}\bar{x}+\epsilon_{i}\alpha,$$ we have $c^{T}x_{i}\to c^{T}\bar{x}$. Note that this also gives us $\alpha\geq0$. By our compactness assumption, $x_{i}\to\bar{x}$, which means that $|\epsilon_{i}\gamma_{i}|\to0$. This means that $\frac{\gamma_{i}}{|\gamma_{i}|}$ converges to a vector $\gamma_{\infty}$ in $T_{\Phi}(0)\backslash\{0\}$. Observe that $c^{T}\gamma_{\infty}=\lim_{i\to\infty}c^{T}\frac{\gamma_{i}}{|\gamma_{i}|}\leq\lim_{i\to\infty}\frac{\alpha}{|\gamma_{i}|}=0$, and that $c^{T}\gamma_{i}\geq0$, so $c^{T}\gamma_{\infty}=0$. This is a contradiction to $\{c\}^{\perp}\cap T_{\Phi}(\bar{x})=\{0\}$. We can thus apply [@RW98 Theorem 7.33] to conclude that $\min\{c^{T}\gamma\mid\gamma\in\Gamma_{\epsilon}\}$ converges to $\min\{c^{T}\gamma\mid\gamma\in\Gamma\}$ and $\limsup_{\epsilon\searrow0}\bar{\Gamma}_{\epsilon}\subset\bar{\Gamma}$, ending the proof of the theorem for the linear case. **Step 4: Locally Lipschitz subdifferentially regular $f$ at $\bar{x}$.** Consider the problem $$\begin{aligned} & & \min_{x,t}\, t\nonumber \\ & & \mbox{s.t. }[\bar{A}_{i}+\Delta A_{i}]x-[\bar{b}_{i}+\Delta b_{i}]\in Q_{i}\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\epsilon\Delta\mathcal{U}_{i}\mbox{ for all }i,\label{eq:nonlin-robust}\\ & & \mbox{and }(x,t)\in{\mbox{\rm epi}}(f)=\{(x,t)\mid f(x)\leq t\}.\nonumber \end{aligned}$$ and the tangential problem $$\begin{aligned} & & \min_{\gamma,s}\, s\nonumber \\ & & \mbox{s.t. }\bar{A}_{i}\gamma+[(\Delta A_{i})\bar{x}-\Delta b_{i}]\in T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i})\label{eq:nonlin-tangent}\\ & & \phantom{\mbox{s.t. }}\mbox{ for all }(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}\mbox{ for all }i.\nonumber \\ & & \mbox{and }(\gamma,s)\in{T_{\scriptsize{\mbox{\rm epi}(f)}}}(\bar{x},f(\bar{x})).\nonumber \end{aligned}$$ The robust and tangential problems are equivalent to the respective problems and in the statement of the theorem. We now show that if conditions (1) to (5) in the theorem statement are satisfied for and , then these conditions hold for and as well. For condition (1), we need only to check that ${\mbox{\rm epi}}(f)$ is Clarke regular at $(\bar{x},f(\bar{x}))$, which is immediate from the subdifferential regularity of $f$ at $\bar{x}$. Condition (2) is straightforward. For condition (3), we note that the set of minimizers of is just $\bar{\Gamma}\times\{df(\bar{x})(\hat{\gamma})\}$, where $\hat{\gamma}$ is any element in $\bar{\gamma}$. The set $\bar{\Gamma}\times\{df(\bar{x})(\hat{\gamma})\}$ is bounded if and only if $\bar{\Gamma}$ is bounded. We further assume that $f$ is locally Lipschitz at $\bar{x}$ with Lipschitz modulus $\kappa$. For condition (4), suppose $\gamma^{\prime}$ is a vector such that $\bar{A}_{i}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i}))$ for all $i$. Since $(\mathbf{0},1)\in{\mbox{\rm int}}({T_{\scriptsize{\mbox{\rm epi}(f)}}}(\bar{x},f(\bar{x})))$, we have $(\gamma^{\prime},(\kappa+1)|\gamma^{\prime}|)\in{\mbox{\rm int}}({T_{\scriptsize{\mbox{\rm epi}(f)}}}(\bar{x},f(\bar{x})))$, which verifies condition (4). We also need to check that given $f(x_{i})\to f(\bar{x})$ and $\{x_{i}\}\subset\Phi$ implies $x_{i}\to\bar{x}$, we have the compactness condition that $t_{i}\to f(\bar{x})$, $(x_{i},t_{i})$ satisfies $f(x_{i})\leq t_{i}$ and $\{x_{i}\}\subset\Phi$ implies $(x_{i},t_{i})\to(\bar{x},f(\bar{x}))$. Since $f(\bar{x})\leq f(x_{i})\leq t_{i}$ and $t_{i}\to f(\bar{x})$, we have $f(x_{i})\to f(\bar{x})$, which gives $x_{i}\to\bar{x}$, and thus $(x_{i},t_{i})\to(\bar{x},f(\bar{x}))$ as needed. We now take a closer look at step 4 of the proof of Theorem \[thm:TFO-approx\]. Consider the general case where $\bar{\Gamma}_{\epsilon}=\arg\min\{c^{T}\gamma\mid\gamma\in\Gamma_{\epsilon}\}$, $\bar{\Gamma}=\arg\min\{c^{T}\gamma\mid\gamma\in\Gamma\}$ and $\Gamma=\lim_{\epsilon\searrow0}\Gamma_{\epsilon}$. It may turn out that $\limsup_{\epsilon\searrow0}\bar{\Gamma}_{\epsilon}\subsetneq\bar{\Gamma}$, as the example in Figure \[fig:V\_e\_V\] shows. Example \[exa:square-eg\] shows that it is possible for $\bar{\Gamma}$ to be bounded but not be a singleton set. In such cases, it is possible that $\bar{\Gamma}_{\epsilon}$ is a singleton set, which occurs when the function $f$ is strictly convex for example. ![\[fig:V\_e\_V\]$\lim_{\epsilon\to0}\Gamma_{\epsilon}=\Gamma$, but $\limsup_{\epsilon\to0}\bar{\Gamma}_{\epsilon}\subsetneq\bar{\Gamma}$.](V_e_V) We make an observation on the condition $\bar{A}_{i}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i}))$ for all $i$ in Theorem \[thm:TFO-approx\]. (Constraint qualification in tangential problem) The optimization problem$$\begin{aligned} & & \min_{x\in\mathbb{R}^{2}}x_{2}\\ & & \mbox{s.t. }x\in Q_{1}:=\{x\mid-x_{1}+x_{2}^{2}\leq0\}\\ & & \phantom{\mbox{s.t. }}x\in Q_{2}:=\{x\mid x_{1}+x_{2}^{2}\leq0\}\end{aligned}$$ can be written equivalently as $$\begin{aligned} & & \min_{x\in\mathbb{R}^{2}}x_{2}\\ & & \mbox{s.t. }x\in Q_{1}\cap Q_{2}=\{(0,0)\},\end{aligned}$$ and the solution for both problems is $\bar{x}=(0,0)$. The tangential approximation for the first problem is $$\begin{aligned} & & \min_{\gamma\in\mathbb{R}^{2}}\gamma_{2}\\ & & \left.\begin{array}{c} \mbox{s.t. }\gamma\in T_{Q_{1}}(\bar{x})=\mathbb{R}_{+}\times\mathbb{R}\\ \phantom{\mbox{s.t. }}\gamma\in T_{Q_{2}}(\bar{x})=\mathbb{R}_{-}\times\mathbb{R}\end{array}\right\} \implies\gamma\in\{0\}\times\mathbb{R},\end{aligned}$$ while the tangential approximation for the second problem is $$\begin{aligned} & & \min_{\gamma\in\mathbb{R}^{2}}\gamma_{2}\\ & & \mbox{s.t. }\gamma\in T_{Q_{1}\cap Q_{2}}(\bar{x})=\{(0,0)\}.\end{aligned}$$ Clearly, the solutions for the two problems are different. This example shows that depending on how the optimization problem is written, the tangential problems may not be equivalent and may have different solutions. But note that in this case, $T_{Q_{1}\cap Q_{2}}(\bar{x})\subsetneq T_{Q_{1}}(\bar{x})\cap T_{Q_{2}}(\bar{x})$, which implies that the MFCQ does not hold, which in turn implies that there is no vector $\gamma^{\prime}$ such that $\bar{A}_{i}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i}))$ for all $i$. The condition $\bar{A}_{i}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i}))$ for all $i$ in Theorem \[thm:TFO-approx\] ensures that the MFCQ holds whenever a group of the $\{(A_{i},b_{i})\}_{i=1}^{m}$ contains repetitions, ensuring that the tangent cones of the intersections is the intersections of the corresponding tangent cones through Proposition \[pro:tangent-NMFCQ\]. Theorem \[thm:TFO-approx\] shows that the decrease in objective function of the robust optimization problem is differentiable in the size of the uncertainty set at $\epsilon=0$. This observation can help give an approximate of the maximum robustness one can afford if the objective is to be above a certain value. The tangential problem also shows that the variables that we should estimate or measure more accurately are those which make the set $L(\Delta\mathcal{U})=\{(\Delta A)\bar{x}-\Delta b\mid(\Delta A,\Delta b)\in\Delta\mathcal{U}\}$ small. For example, if $\bar{x}=0$, then more effort should be spent on determining $\bar{b}$ accurately rather than entries in $\bar{A}$. Likewise, $\bar{x}$ determines which variables in $\bar{A}$ should be measured more accurately than others. Besides these analytical properties, Theorem \[thm:TFO-approx\] shows that solving the tangential problem can help to obtain a good approximate of the robust solution. The robust optimization problem is known to be more computationally expensive than the original problem, so it will take more effort to obtain a desired level of accuracy. With the tangential problem, we can make use of the previously calculated optimization problem to obtain an approximate of the robust solution. Such an approximation is likely to be simpler than the robust optimization problem (for example, for LP in and for SOCP in Example \[exa:SOCP-easy\], though it still may be computationally difficult), and need not be computed to very high accuracy to obtain a good approximate of the robust optimization problem. (Relaxing the constraint qualification in Theorem \[thm:TFO-approx\]) The existence of $\gamma^{\prime}$ such that $\bar{A}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q}(\bar{A}\bar{x}-\bar{b}))$ can be relaxed slightly if more structure is known about the set $Q$. All we need is for the chain of inclusions to hold. For example, if $Q_{i}$ is polyhedral and there is some vector $\gamma^{\prime}$ such that $$\bar{A}_{i}\gamma^{\prime}\in{\mbox{\rm rint}}\big({\mbox{\rm span}}(\Pi_{i}(\Delta\mathcal{U}_{i}))\big)\cap T_{Q_{i}}(\bar{A}_{i}\bar{x}-\bar{b}_{i}),$$ where $\Pi_{i}(\Delta\mathcal{U}_{i})=\{\Delta A_{i}\mid(\Delta A_{i},\Delta b_{i})\in\Delta\mathcal{U}_{i}\mbox{ for some }\Delta b_{i}\}$, then the chain of inclusions would hold as well. A finer analysis on $\Pi_{i}(\Delta\mathcal{U})$ and local recession vectors can give stronger results. \[sec:First-order\]First order optimality conditions of the tangential problem ============================================================================== In this section, we discuss first order optimality conditions of the tangential problem, which can be useful for designing specialized numerical methods for the tangential problem. In view of Theorem \[thm:TFO-approx\], we also give sufficient conditions for $\bar{\Gamma}$ to be bounded and for $\bar{\Gamma}$ to be a singleton. As explained in [@BGN09 Chapters 5-8], the robust optimization problem is computationally tractable if either $Q$ is polyhedral or $\Delta\mathcal{U}$ is polyhedral, while most other problems encountered in practice are not computationally tractable. Recall that a typical constraint in a robust optimization problem whose nominal solution is $\bar{x}$ is $$[\bar{A}+\Delta A](\bar{x}+\Delta x)-[\bar{b}+\Delta b]\in Q\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}.$$ If there were no uncertainty in the matrix $\bar{A}$, then the constraint can be written as$$\bar{A}(\Delta x)-\Delta b\in Q-[\bar{A}\bar{x}-\bar{b}]\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}.\label{eq:robust-fixed-A}$$ Recall that the tangential constraint is of the form $$\bar{A}\gamma+[(\Delta A)\bar{x}-\Delta b]\in T_{Q}(\bar{A}\bar{x}-\bar{b})\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}.$$ If the variable $\Delta x$ in were replaced by $\gamma$, then we see that is similar to the tangential constraint. In other words, the uncertainty in $\bar{A}$ is transferred to $\bar{b}$ in the tangential constraint through $(\Delta A)\bar{x}$. Tangential problems are still often hard to compute efficiently, but the additional structure may be exploited for designing specialized methods. The case where $T_{Q}(\bar{A}\bar{x}-\bar{b})$ is polyhedral is just robust linear programming and is easy, while the tangential problem for $L(\Delta\mathcal{U})$ being polyhedral reduces to optimizing over the cone $T_{Q}(\bar{A}\bar{x}-\bar{b})$, as illustrated below. (Polyhedral $L(\Delta\mathcal{U})$) Suppose $L(\Delta\mathcal{U})$ is a polyhedral compact set of the form $L(\Delta\mathcal{U})={\mbox{\rm conv}}(v_{1},\dots,v_{J})$. The tangential problem $$\begin{aligned} & & \min_{\gamma}h(\gamma)\\ & & \mbox{s.t. }\bar{A}\gamma+L(\Delta\mathcal{U})\subset T_{Q}(\bar{A}\bar{x}-\bar{b}),\end{aligned}$$ where $h$ is sublinear, is equivalent to $$\begin{aligned} & & \min_{\gamma}h(\gamma)\\ & & \mbox{s.t. }\bar{A}\gamma+v_{j}\in T_{Q}(\bar{A}\bar{x}-\bar{b})\mbox{ for all }1\leq j\leq J.\end{aligned}$$ For this section, we define the sets $\Phi$ and $\Psi$ by $$\begin{aligned} \Phi & := & \{x:\bar{A}x-\bar{b}\in Q\}.\label{eq:Phi}\\ \Psi & := & \{y:y+L(\Delta\mathcal{U})\subset T_{Q}(\bar{A}\bar{x}-\bar{b})\}.\nonumber \end{aligned}$$ Hence $\Gamma$ can be written similarly as $$\begin{aligned} \Gamma & = & \{\gamma:\bar{A}\gamma+L(\Delta\mathcal{U})\subset T_{Q}(\bar{A}\bar{x}-\bar{b})\}\label{eq:Gamma}\\ & = & \{\gamma:\bar{A}\gamma\in\Psi\}.\nonumber \end{aligned}$$ Recall also that $$\bar{\Gamma}=\arg\min\{df(\bar{x})(\gamma)\mid\gamma\in\Gamma\}.\label{eq:barGamma}$$ \[pro:T\_Phi\](Tangent space of feasible set) Suppose $Q$ is Clarke regular at $\bar{A}\bar{x}-\bar{b}$. If there is a vector $\gamma^{\prime}$ such that $\bar{A}\gamma^{\prime}\in{\mbox{\rm rint}}(T_{Q}(\bar{A}\bar{x}-\bar{b}))$, then $T_{\Phi}(\bar{x})=\{\gamma:\bar{A}\gamma\in T_{Q}(\bar{A}\bar{x}-\bar{b})\}$, where $\Phi$ is defined in . This follows directly from [@RW98 Theorem 6.31]. The constraint qualification in that Theorem is equivalent to the condition in our statement through [@RW98 Exercise 6.39(b)]. The normal cone $N_{\Gamma}(\gamma)$ can be estimated from the image $N_{\Psi}(\bar{A}\gamma)$. \[pro:normal-of-tangent-feasible\](Normal cone of tangent feasible set) Suppose there is a vector $\gamma^{\prime}$ such that $\bar{A}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q}(\bar{A}\bar{x}-\bar{b}))$, and $Q$ is Clarke regular at $\bar{A}\bar{x}-\bar{b}$. Then the normal cone $N_{\Gamma}(\gamma)$ equals $\{\bar{A}^{T}v\mid v\in N_{\Psi}(\bar{A}\gamma)\}$, where $\Gamma$ and $\Psi$ are defined in and . Note that $\Gamma$ can be written in terms of $\Psi$ as $\Gamma=\{\gamma:\bar{A}\gamma\in\Psi\}$. The result follows directly from [@RW98 Theorem 6.14], though we still have to check the constraint qualification condition there. Through [@RW98 Exercise 6.39(b)], the constraint qualification condition required is that there is a vector $\gamma^{\prime}$ such that $\bar{A}\gamma^{\prime}\in{\mbox{\rm int}}(T_{\Psi}(\bar{A}\gamma))$. Note that the recession cone of $\Psi$ is $T_{Q}(\bar{A}\bar{x}-\bar{b})$. This means that $T_{Q}(\bar{A}\bar{x}-\bar{b})\subset T_{\Psi}(\bar{A}\gamma)$, which shows that $\bar{A}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q}(\bar{A}\bar{x}-\bar{b}))$ implies the constraint qualification condition. The conclusion is straightforward. One notices that if there is a vector $\gamma^{\prime}$ such that $\bar{A}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q}(\bar{A}\bar{x}-\bar{b}))$ and $0\in{\mbox{\rm int}}(L(\Delta\mathcal{U}))$ (which holds when $0\in{\mbox{\rm int}}(\Delta\mathcal{U})$), then the tangential problem is feasible. If a vector $\bar{x}$ is a solution of the problem $\min\{f(x)\mid x\in D\}$, where $f$ is locally Lipschitz and subdifferentially regular at $\bar{x}$ then it is well known that there is a $c\in\partial f(\bar{x})$ such that $-c\in N_{D}(\bar{x})$ (see [@RW98 Theorem 8.15] for example). We prove the following lemmas, whose proofs do not seem easy to find. \[lem:strict-loc-min\](Strict minimizers) Let $D\subset\mathbb{R}^{n}$ be Clarke regular at $\bar{x}$, $f:\mathbb{R}^{n}\to\mathbb{R}$ be locally Lipschitz and subdifferentially regular at $\bar{x}$. Then $-c\in{\mbox{\rm int}}(N_{D}(\bar{x}))$ for some $c\in\partial f(\bar{x})$ implies that $\bar{x}$ is a strict local minimizer of $\min\{f(x)\mid x\in D\}$. Since $-c\in{\mbox{\rm int}}(N_{D}(\bar{x}))$, there is some $\delta>0$ such that $-c+\delta\mathbb{B}\subset N_{D}(\bar{x})$, and so there is a neighborhood $O$ of $\bar{x}$ such that for $x\in D\cap O$, $$\begin{aligned} \Big[-c+\delta\frac{x-\bar{x}}{|x-\bar{x}|}\Big]^{T}(x-\bar{x}) & \leq & \frac{\delta}{2}|x-\bar{x}|\\ \Rightarrow c^{T}(x-\bar{x}) & \geq & \frac{\delta}{2}|x-\bar{x}|.\end{aligned}$$ By restricting $O$ if necessary, for $x\in D\cap O$, we have $f(x)\geq f(\bar{x})+df(\bar{x})(x-\bar{x})-\frac{\delta}{4}|x-\bar{x}|$. The well known characterization of the subderivative in terms of support functions (see for example [@Cla83; @R70; @RW98]) gives $$\begin{aligned} f(x) & \geq & f(\bar{x})+df(\bar{x})(x-\bar{x})-\frac{\delta}{4}|x-\bar{x}|\\ & = & f(\bar{x})+\max_{v\in\partial f(\bar{x})}v^{T}(x-\bar{x})-\frac{\delta}{4}|x-\bar{x}|\\ & \geq & f(\bar{x})+c^{T}(x-\bar{x})-\frac{\delta}{4}|x-\bar{x}|\\ & \geq & f(\bar{x})+\frac{\delta}{4}|x-\bar{x}|.\end{aligned}$$ Therefore $\bar{x}$ is a strict local minimizer. \[lem:equiv-strict\](Equivalence of strict minimizer condition) Let $D$ be Clarke regular at $\bar{x}$. When $f$ is $\mathcal{C}^{1}$ at $\bar{x}$, $\nabla f(\bar{x})\neq0$ and $-\nabla f(\bar{x})\in N_{D}(\bar{x})$ (which holds when $\bar{x}$ is a local minimizer of $\min\{f(x)\mid x\in D\}$), then the conditions $-\nabla f(\bar{x})\in{\mbox{\rm int}}(N_{D}(\bar{x}))$ and $\{\nabla f(\bar{x})\}^{\perp}\cap T_{D}(\bar{x})=\{0\}$ are equivalent. The fact that $-\nabla f(\bar{x})\in N_{D}(\bar{x})$ when $\bar{x}$ is a local minimizer of $\min\{f(x)\mid x\in D\}$ is well known (see [@RW98 Theorem 6.12] for example). Suppose that $-\nabla f(\bar{x})\in{\mbox{\rm int}}(N_{D}(\bar{x}))$, and $v\in\{\nabla f(\bar{x})\}^{\perp}\cap T_{D}(\bar{x})$. Then $v\in T_{D}(\bar{x})=[N_{D}(\bar{x})]^{*}$ (the polar cone of $N_{D}(\bar{x})$). In other words, $v^{T}s\leq0$ for all $s\in N_{D}(\bar{x})$. If $v\neq0$, since $-\nabla f(\bar{x})\in{\mbox{\rm int}}(N_{D}(\bar{x}))$, there is some $\epsilon>0$ such that $\nabla f(\bar{x})+\epsilon v\in N_{D}(\bar{x})$, which gives $v^{T}(\nabla f(\bar{x})+\epsilon v)\leq0$, and thus $v^{T}\nabla f(\bar{x})<0$. Since $v\in\{\nabla f(\bar{x})\}^{\perp}$, this means $v=0$, so $\{\nabla f(\bar{x})\}^{\perp}\cap T_{D}(\bar{x})=\{0\}$. Next, suppose $-\nabla f(\bar{x})\in N_{D}(\bar{x})\backslash{\mbox{\rm int}}(N_{D}(\bar{x}))$. Then $N_{N_{D}(\bar{x})}(-\nabla f(\bar{x}))\supsetneq\{0\}$. (See for example [@RW98 Exercise 6.19].) Let $w\in N_{N_{D}(\bar{x})}(-\nabla f(\bar{x}))\backslash\{0\}$. Firstly, $-\lambda\nabla f(\bar{x})\in N_{D}(\bar{x})$ for all $\lambda\geq0$, so $w^{T}(-\lambda\nabla f(\bar{x})-(-\nabla f(\bar{x})))\leq0$ for all $\lambda\geq0$, that is $(1-\lambda)w^{T}\nabla f(\bar{x})\leq0$ for all $\lambda\geq0$. This implies that $w^{T}\nabla f(\bar{x})=0$, or $w\in\{\nabla f(\bar{x})\}^{\perp}$. Secondly, $w\in N_{N_{D}(\bar{x})}(-\nabla f(\bar{x}))$ means that $w^{T}(s+\nabla f(\bar{x}))\leq0$ for all $s\in N_{D}(\bar{x})$. Since $w^{T}\nabla f(\bar{x})=0$, this means that $w^{T}s\leq0$ for all $s\in N_{D}(\bar{x})$, which means that $w\in T_{D}(\bar{x})$. Therefore, $w\in[\{\nabla f(\bar{x})\}^{\perp}\cap T_{D}(\bar{x})]\backslash\{0\}$, which gives us the equivalence between the two conditions. In Proposition \[pro:Condn-Phi\] below, we show that the conditions in Lemma \[lem:strict-loc-min\] can give us boundedness information on the robust problem. We also give conditions for which $\bar{\Gamma}$ is a singleton. \[pro:Condn-Phi\](Conditions for $\bar{\Gamma}$ and optimality) Suppose that $Q$ is Clarke regular at $\bar{A}\bar{x}-\bar{b}$, and there is a vector $\gamma^{\prime}$ such that $\bar{A}\gamma^{\prime}\in{\mbox{\rm int}}(T_{Q}(\bar{A}\bar{x}-\bar{b}))$. Let $f$ be locally Lipschitz and subdifferentially regular at $\bar{x}$, and $\bar{x}$ be a minimizer of the tangential problem . Recall also $\Phi$, $\Psi$, $\Gamma$ and $\bar{\Gamma}$ as defined in , and . 1. The set $\bar{\Gamma}$ is bounded if there is some $c\in\partial f(\bar{x})$ such that $-c\in{\mbox{\rm int}}(N_{\Phi}(\bar{x}))$. 2. If $f$ is $\mathcal{C}^{1}$ at $\bar{x}$ and $\nabla f(\bar{x})\neq0$, then $\bar{\Gamma}$ is bounded if and only if $-\nabla f(\bar{x})\in{\mbox{\rm int}}(N_{\Phi}(\bar{x}))$, which is also equivalent to $\{\nabla f(\bar{x})\}^{\perp}\cap T_{\Phi}(\bar{x})=\{0\}$. 3. A feasible $\bar{\gamma}\in\Gamma$ is in $\bar{\Gamma}$ if we can find $c\in\partial f(\bar{x})$ such that $-c\in N_{T_{\Phi}(\bar{x})}(\bar{\gamma})$. The condition $-c\in N_{T_{\Phi}(\bar{x})}(\bar{\gamma})$ holds when we can find $\{u_{i}\}_{i=1}^{k}$ and $\{v_{i}\}_{i=1}^{k}$ such that $v_{i}\in N_{T_{Q}(\bar{A}\bar{x}-\bar{b})}(\bar{\gamma}+u_{i})$, $u_{i}\in L(\Delta\mathcal{U})$ and $-c=\sum_{i=1}^{k}\lambda_{i}\bar{A}^{T}v_{i}$ for some $\lambda_{i}\geq0$, $1\leq i\leq k$. Here, $L$ is the linear map $L(\Delta A,\Delta b)=(\Delta A)\bar{x}-\Delta b$. 4. The set $\bar{\Gamma}$ is a singleton if we can find some $c\in\partial f(\bar{x})$ such that $-c\in{\mbox{\rm int}}(N_{T_{\Phi}(\bar{x})}(\bar{\gamma}))$ for some $\bar{\gamma}\in\bar{\Gamma}$. The condition $-c\in{\mbox{\rm int}}(N_{T_{\Phi}(\bar{x})}(\bar{\gamma}))$ holds when we can find $\{u_{i}\}_{i=1}^{k}$ and $\{v_{i}\}_{i=1}^{k}$ such that $v_{i}\in N_{T_{Q}(\bar{A}\bar{x}-\bar{b})}(\bar{A}\bar{\gamma}+u_{i})$, $u_{i}\in L(\Delta\mathcal{U})$, $k=\dim(Q)$, $\{\bar{A}^{T}v_{i}\}_{i=1}^{k}$ is linearly independent, and $-c=\sum_{i=1}^{k}\lambda_{i}\bar{A}^{T}v_{i}$, where $\lambda_{i}>0$ for all $1\leq i\leq k$. **Part (a):** Seeking a contradiction, suppose that $\Gamma$ is unbounded, so there is a sequence $\{\gamma_{i}\}$ of solutions of minimizers of such that $|\gamma_{i}|\to\infty$, with $\frac{\gamma_{i}}{|\gamma_{i}|}\to\gamma_{\infty}$. Note $$\begin{aligned} \gamma_{i} & \in & \Gamma\\ & = & \{\gamma:\bar{A}\gamma+L(\Delta\mathcal{U})\subset T_{Q}(\bar{A}\bar{x}-\bar{b})\}\\ & \subset & \{\gamma:\bar{A}\gamma\in T_{Q}(\bar{A}\bar{x}-\bar{b})\}\\ & = & T_{\Phi}(\bar{x}),\end{aligned}$$ so $\gamma_{\infty}\in T_{\Phi}(\bar{x})$. On the other hand, since $df(\bar{x})(\cdot)$ is continuous and positively homogeneous, we have $$\begin{aligned} df(\bar{x})(\gamma_{\infty}) & = & \lim_{i\to\infty}df(\bar{x})\Big(\frac{\gamma_{i}}{|\gamma_{i}|}\Big)\\ & = & \lim_{i\to\infty}\frac{1}{|\gamma_{i}|}df(\bar{x})(\gamma_{i})\\ & = & 0.\end{aligned}$$ The well known characterization of the subderivative in terms of support functions (see for example [@Cla83; @R70; @RW98]) gives $$\begin{aligned} \max_{v\in\partial f(\bar{x})}v^{T}\gamma_{\infty} & = & df(\bar{x})(\gamma_{\infty})=0\\ \Rightarrow c^{T}\gamma_{\infty} & \leq & 0.\end{aligned}$$ But $-c\in{\mbox{\rm int}}(N_{\Phi}(\bar{x}))$, so $c^{T}\gamma_{\infty}>0$. This contradiction tells us that the set $\bar{\Gamma}$ is bounded as needed. **Part (b):** In view of part (a) and Lemma \[lem:equiv-strict\], we just need to prove that if $\bar{\Gamma}$ is bounded, then $\{\nabla f(\bar{x})\}^{\perp}\cap T_{\Phi}(\bar{x})=\{0\}$. We prove the contrapositive. Suppose $w\in[\{\nabla f(\bar{x})\}^{\perp}\cap T_{\Phi}(\bar{x})]\backslash\{0\}$. Then $\nabla f(\bar{x})^{T}w=0$, and $w\in T_{\Phi}(\bar{x})$, that is $\bar{A}w\in T_{Q}(\bar{A}\bar{x}-\bar{b})$ by Proposition \[pro:T\_Phi\]. Let $\bar{\gamma}$ be some element in $\bar{\Gamma}$. Then $$\begin{aligned} \bar{A}(\lambda w+\bar{\gamma})+L(\Delta\mathcal{U}) & = & \bar{A}\lambda w+[\bar{A}\bar{\gamma}+L(\Delta\mathcal{U})]\\ & \subset & T_{Q}(\bar{A}\bar{x}-\bar{b})+T_{Q}(\bar{A}\bar{x}-\bar{b})\\ & = & T_{Q}(\bar{A}\bar{x}-\bar{b}).\end{aligned}$$ Therefore $\bar{\gamma}+\lambda w\in\bar{\Gamma}$ for all $\lambda\geq0$, which shows that $\bar{\Gamma}$ is unbounded, concluding our proof. **Parts (c), (d):** Note that $$\begin{aligned} \Psi & = & \{\gamma:\gamma+u\in T_{Q}(\bar{A}\bar{x}-\bar{b})\mbox{ for all }u\in L(\Delta\mathcal{U})\}\\ & \subset & \{\gamma:\gamma+u_{i}\in T_{Q}(\bar{A}\bar{x}-\bar{b})\mbox{ for }1\leq i\leq k\}\\ & = & \bigcap_{i=1}^{k}[T_{Q}(\bar{A}\bar{x}-\bar{b})-u_{i}],\end{aligned}$$ so $$\begin{aligned} N_{\Psi}(\bar{A}\bar{\gamma}) & \supset & N_{\cap_{i=1}^{k}[T_{Q}(\bar{A}\bar{x}-\bar{b})-u_{i}]}(\bar{A}\bar{\gamma})\\ & \supset & [N_{T_{Q}(\bar{A}\bar{x}-\bar{b})-u_{1}}(\bar{A}\bar{\gamma})]+\cdots+[N_{T_{Q}(\bar{A}\bar{x}-\bar{b})-u_{k}}(\bar{A}\bar{\gamma})]\\ & & \qquad\mbox{ (by [\ref{RW98}, Theorem 6.42])}\\ & = & [N_{T_{Q}(\bar{A}\bar{x}-\bar{b})}(\bar{A}\bar{\gamma}+u_{1})]+\cdots+[N_{T_{Q}(\bar{A}\bar{x}-\bar{b})}(\bar{A}\bar{\gamma}+u_{k})]\\ & \supset & \left\{ \sum_{i=1}^{k}\lambda_{i}v_{i}\mid\lambda_{i}\geq0\right\} .\end{aligned}$$ Therefore, by Proposition \[pro:normal-of-tangent-feasible\], $$\begin{aligned} N_{\Gamma}(\bar{\gamma}) & = & \{\bar{A}^{T}y:y\in N_{\Psi}(\bar{A}\bar{\gamma})\}\\ & \supset & \left\{ \sum_{i=1}^{k}\lambda_{i}\bar{A}^{T}v_{i}:\lambda_{i}\geq0\right\} .\end{aligned}$$ For part (c), the condition stated is equivalent to the existence of $c\in\partial f(\bar{x})$ such that $-c\in\{\sum_{i=1}^{k}\lambda_{i}v_{i}\mid\lambda_{i}\geq0\}$, which implies $-c\in N_{\Gamma}(\bar{\gamma})$, which in turn implies $\bar{\gamma}\in\bar{\Gamma}$. Part (d) follows by applying Lemma \[lem:strict-loc-min\]. The conditions (c) and (d) in Proposition \[pro:Condn-Phi\] can be helpful for designing numerical methods for solving the tangential problem. Due to Clarke regularity, the tangential problem is convex. However, the problem of determining whether a point is feasible is not necessarily easy. The result corresponding to conditions (c) and (d) in Proposition \[pro:Condn-Phi\] can also be generalized for robust optimization in general. The following result on normal cones in robust optimization combined with results on optimality of nonlinear programs (in Lemma \[lem:strict-loc-min\] for example) gives us the optimality conditions. The proof of the following result is a direct application of [@RW98 Theorems 6.14 and 6.42] similar to the proofs of Propositions \[pro:normal-of-tangent-feasible\] and \[pro:Condn-Phi\], so we shall only state the result. (Normal cones in robust optimization) For $f:\mathbb{R}^{n}\to\mathbb{R}$, consider the robust optimization problem$$\begin{aligned} & & \min_{x}f(x)\\ & & \mbox{s.t. }[\bar{A}+\Delta A]x-[\bar{b}-\Delta b]\in Q\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}.\end{aligned}$$ Let the sets $\Omega(\Delta A,\Delta b)\subset\mathbb{R}^{n}$ and $\Xi\subset\mathbb{R}^{n}$ be defined by $$\begin{aligned} \Omega(\Delta A,\Delta b) & := & \{x\mid[\bar{A}+\Delta A]x-[\bar{b}-\Delta b]\in Q\},\\ \mbox{ and }\Xi & := & \{x\mid[\bar{A}+\Delta A]x-[\bar{b}-\Delta b]\in Q\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}\}\\ & = & \bigcap_{(\Delta A,\Delta b)\in\Delta\mathcal{U}}\Omega(\Delta A,\Delta b).\end{aligned}$$ Let $\bar{x}\in\Xi$. 1. If $Q$ is Clarke regular at $[\bar{A}+\Delta A]\bar{x}-[\bar{b}-\Delta b]$ and the only vector $y\in N_{Q}([\bar{A}+\Delta A]\bar{x}-[\bar{b}-\Delta b])$ for which $[\bar{A}+\Delta A]^{T}y=0$ is $y=0$, then $$N_{\Omega(\Delta A,\Delta b)}(\bar{x})=\{[\bar{A}+\Delta A]^{T}y\mid y\in N_{Q}([\bar{A}+\Delta A]\bar{x}-[\bar{b}-\Delta b])\}.$$ 2. For any finite set $\{(\Delta A_{i},\Delta b_{i})\}_{i\in I}\subset\Delta\mathcal{U}$ such that $\bar{x}$ is Clarke regular at $\Omega(\Delta A_{i},\Delta b_{i})$ for all $i\in I$, the normal cone $N_{\Xi}(\bar{x})$ satisfies $N_{\Xi}(\bar{x})\supset\sum_{i\in I}N_{\Omega(\Delta A_{i},\Delta b_{i})}(\bar{x})$. \[sec:uncertain-sets\]Addition of uncertainty sets in the tangential problem ============================================================================ For much of this section we focus on the tangential robust problem on addition of uncertainty sets. More specifically, we ask what we can say about the tangential problem with uncertainty set $\lambda_{1}\Delta\mathcal{S}_{1}+\lambda_{2}\Delta\mathcal{S}_{2}$ given knowledge of the optimal solutions of the tangential problem for the uncertainty sets $\Delta\mathcal{S}_{1}$ and $\Delta\mathcal{S}_{2}$. Such a problem can arise from having to considering robust optimization problems with errors which are a sum of two or more unknown sources. We begin with some elementary properties. \[pro:Trans+incl\](Elementary properties of uncertainty sets) Suppose $Q$ is Clarke regular at $\bar{A}\bar{x}-\bar{b}$. For an uncertainty set $\Delta\mathcal{U}$, suppose the solution of the tangential problem is defined by $$\begin{aligned} v(\Delta\mathcal{U}) & := & \min_{\gamma}h(\gamma)\\ & & \mbox{s.t. }\bar{A}\gamma+[(\Delta A)\bar{x}-\Delta b]\in T_{Q}(\bar{A}\bar{x}-\bar{b})\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U},\end{aligned}$$ where $h$ is sublinear. Then - If $\Delta\mathcal{S}^{\prime}=\lambda\Delta\mathcal{S}$, then $v(\Delta\mathcal{S}^{\prime})=\lambda v(\Delta\mathcal{S})$. - If $\Delta\mathcal{S}^{\prime}\supset\Delta\mathcal{S}$, then $v(\Delta\mathcal{S})\leq v(\Delta\mathcal{S}^{\prime})$. - If $h(\gamma)=c^{T}\gamma$ and $\Delta\mathcal{S}^{\prime}=\Delta\mathcal{S}+\{\bar{s}\}$, then $v(\Delta\mathcal{S}^{\prime})=v(\Delta\mathcal{S})+c^{T}\bar{s}$. We have the following result to study how set addition affects the solution to the tangential problem. \[pro:set-add\](Set addition) Recall the definition of $v(\Delta\mathcal{U})$ in Proposition \[pro:Trans+incl\]. Suppose $\Delta\mathcal{S}=\lambda_{1}\Delta\mathcal{S}_{1}+\lambda_{2}\Delta\mathcal{S}_{2}$, where $\lambda_{1},\lambda_{2}\geq0$. Then $v(\Delta\mathcal{S})\leq\lambda_{1}v(\Delta\mathcal{S}_{1})+\lambda_{2}v(\Delta\mathcal{S}_{2})$. In view of Proposition \[pro:Trans+incl\], we only need to prove the case for $\lambda_{1}=\lambda_{2}=1$. Suppose $\bar{\gamma}_{1}$ and $\bar{\gamma}_{2}$ are solutions to $$\begin{aligned} & & \min_{\gamma}h(\gamma)\\ & & \mbox{s.t. }\bar{A}\gamma+[(\Delta A)\bar{x}-\Delta b]\in T_{Q}(\bar{A}\bar{x}-\bar{b})\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{S}_{i}\end{aligned}$$ for $i=1,2$. If $(\Delta A,\Delta b)\in\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2}$, then $$[(\Delta A)\bar{x}-\Delta b]=[(\Delta A_{1})\bar{x}-\Delta b_{1}]+[(\Delta A_{1})\bar{x}-\Delta b_{2}],$$ for some $[(\Delta A_{i})\bar{x}-\Delta b_{i}]\in\Delta\mathcal{S}_{i}$ for $i=1,2$. This means that $$\begin{aligned} & & \bar{A}(\bar{\gamma}_{1}+\bar{\gamma}_{2})-[(\Delta A)\bar{x}-\Delta b]\\ & = & \big[\bar{A}\bar{\gamma}_{1}-[(\Delta A_{1})\bar{x}-\Delta b_{1}]\big]+\big[\bar{A}\bar{\gamma}_{2}-[(\Delta A_{2})\bar{x}-\Delta b_{2}]\big]\\ & \in & T_{Q}(\bar{A}\bar{x}-\bar{b})+T_{Q}(\bar{A}\bar{x}-\bar{b})\\ & = & T_{Q}(\bar{A}\bar{x}-\bar{b}).\end{aligned}$$ So $\bar{\gamma}_{1}+\bar{\gamma}_{2}$ is a feasible, though not necessarily optimal, solution to the tangential problem where the uncertainty set is $\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2}$, which shows that $v(\Delta\mathcal{S})\leq h(\bar{\gamma}_{1}+\bar{\gamma}_{2})\leq h(\bar{\gamma}_{1})+h(\bar{\gamma}_{2})=v(\Delta\mathcal{S}_{1})+v(\Delta\mathcal{S}_{2})$. In a nondegenerate linear programming problem, we do have equality in Proposition \[pro:set-add\]. The assumption that $\bar{A}$ is an invertible square matrix below is not restrictive, because this is exactly what happens in a nondegenerate linear program. \[pro:set-add-nice\](Set addition in nondegenerate linear programming) Consider the tangential problem having a linear programming structure$$\begin{aligned} v(\Delta\mathcal{U}) & := & \min_{\gamma}c^{T}\gamma\\ & & \mbox{s.t. }\bar{A}\gamma+[(\Delta A)\bar{x}-\Delta b]\leq0\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}.\end{aligned}$$ Assume that $\bar{A}$ is square and invertible, and $-c=\sum_{i=1}^{k}\lambda_{i}\bar{A}_{i}$ for some $\lambda_{i}>0$, $1\leq i\leq k$. Then $v(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})=v(\Delta\mathcal{S}_{1})+v(\Delta\mathcal{S}_{2})$. Recall that feasibility can be rewritten as $$\bar{A}_{i}\gamma+\max_{(\Delta A,\Delta b)\in\Delta\mathcal{U}}[(\Delta A_{i})\bar{x}-\Delta b_{i}]\leq0\mbox{ for all }i.$$ Form the vector $w(\Delta\mathcal{U})\in\mathbb{R}^{m}$ by $$[w(\Delta\mathcal{U})]_{i}:=-\max_{(\Delta A,\Delta b)\in\Delta\mathcal{U}}[(\Delta A_{i})\bar{x}-\Delta b_{i}].$$ The value $v(\Delta\mathcal{U})$ is equal to $c^{T}\bar{A}^{-1}w(\Delta\mathcal{U})$. The condition on $-c$ in the statement assures that this minimizer is unique through convexity and Lemma \[lem:strict-loc-min\]. It is clear that for each $i$, we have $$\begin{aligned} & & \max_{(\Delta A,\Delta b)\in\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2}}[(\Delta A_{i})\bar{x}-\Delta b_{i}]\\ & = & \max_{(\Delta A,\Delta b)\in\Delta\mathcal{S}_{1}}[(\Delta A_{i})\bar{x}-\Delta b_{i}]+\max_{(\Delta A,\Delta b)\in\Delta\mathcal{S}_{2}}[(\Delta A_{i})\bar{x}-\Delta b_{i}],\end{aligned}$$ which implies that $w(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})=w(\Delta\mathcal{S}_{1})+w(\Delta\mathcal{S}_{2})$. This immediately gives $v(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})=v(\Delta\mathcal{S}_{1})+v(\Delta\mathcal{S}_{2})$ as needed. The next step is to ask whether the property in Proposition \[pro:set-add-nice\] is satisfied for problems of the form$$\begin{aligned} & & \min_{\gamma}c^{T}\gamma\\ & & \mbox{s.t. }\bar{A}\gamma+[(\Delta A)\bar{x}-\Delta b]\in Q\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U},\end{aligned}$$ where $Q$ is a convex cone. If $Q=\{y\mid\tilde{A}y\leq0\}$, where $\tilde{A}$ is an invertible square matrix, then the constraint above can be transformed into $$\tilde{A}\big[\bar{A}\gamma+[(\Delta A)\bar{x}-\Delta b]\big]\leq0\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U},$$ and we can apply Proposition \[pro:set-add-nice\] on $\tilde{A}\bar{A}$. In the general case, the equality $v(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})=v(\Delta\mathcal{S}_{1})+v(\Delta\mathcal{S}_{2})$ may not hold. We give two examples to illustrate this. In the first example, we have a degenerate linear program, while in the second example, we have different cones for a conic programming problem. \[exa:degen\](Inequality in sums of uncertainty sets 1) Consider the problem $$\begin{aligned} \bar{v}(\Delta\mathcal{U}) & := & \min_{x\in\mathbb{R}^{2}}x_{2}\\ & & \mbox{s.t. }(\bar{A}+\Delta A)x-(\mathbf{0}+\Delta b)\leq0\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U},\\ & & \mbox{,where }\bar{A}=\left(\begin{array}{cc} -1 & -1\\ 1 & -1\\ 0.5 & -1\end{array}\right)\mbox{, }\bar{b}=\mathbf{0}.\end{aligned}$$ The optimal solution to the nonrobust problem is $\bar{x}=0$, and the tangential problem is$$\begin{aligned} v(\Delta\mathcal{U}) & := & \min_{\gamma\in\mathbb{R}^{2}}\gamma_{2}\\ & & \mbox{s.t. }\bar{A}\gamma+[(\Delta A)\mathbf{0}-\Delta b]\leq0\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}.\end{aligned}$$ We illustrate this example in Figure \[fig:degen\]. Let the uncertainty sets $\Delta\mathcal{S}_{1}$ and $\Delta\mathcal{S}_{2}$ be defined by $$\begin{aligned} \Delta\mathcal{S}_{1} & := & \{\mathbf{0}\}\times\{\Delta b:\Delta b_{1}=0,|\Delta b_{2}|\leq4\epsilon,\Delta b_{3}=0\}\\ \Delta\mathcal{S}_{2} & := & \{\mathbf{0}\}\times\{\Delta b:\Delta b_{1}=0,\Delta b_{2}=0,|\Delta b_{3}|\leq3\epsilon\}\\ \Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2} & = & \{\mathbf{0}\}\times\{\Delta b:\Delta b_{1}=0,|\Delta b_{2}|\leq4\epsilon,|\Delta b_{3}|\leq3\epsilon\}.\end{aligned}$$ To find $v(\Delta\mathcal{S}_{1})$, we note that tangential problem becomes $$\left(\begin{array}{cc} -1 & -1\\ 1 & -1\\ 0.5 & -1\end{array}\right)\gamma\leq\left(\begin{array}{c} 0\\ -4\epsilon\\ 0\end{array}\right).$$ The first two rows are the active constraints, which gives a solution of $\bar{\gamma}(\Delta\mathcal{S}_{1})=(-2\epsilon,2\epsilon)$ and $v(\Delta\mathcal{S}_{1})=2\epsilon$. To find $v(\Delta\mathcal{S}_{2})$, a similar set of calculations shows that the first and third constraints are the active constraints, which gives $\bar{\gamma}(\Delta\mathcal{S}_{2})=(-2\epsilon,2\epsilon)$ and $v(\Delta\mathcal{S}_{2})=2\epsilon$. Similarly, $v(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})=2\epsilon$, and all constraints are active. We have $v(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})<v(\Delta\mathcal{S}_{1})+v(\Delta\mathcal{S}_{2})$ as needed. ![\[fig:degen\]Illustration of Example \[exa:degen\]. ](degen) Here is a second example for the case when the cone $Q$ is slightly more complicated. \[exa:square-eg\](Inequality in sums of uncertainty sets 2) Consider the problem $$\begin{aligned} \bar{v}(\Delta\mathcal{U}) & := & \min_{\gamma\in\mathbb{R}^{3}}x_{3}\\ & & \mbox{s.t.}(I+\Delta A)x-(\mathbf{0}+\Delta b)\in Q\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}.\end{aligned}$$ Here, the matrix $\bar{A}=I$ is the $3\times3$ identity matrix, $\bar{b}\in\mathbb{R}^{3}$ is the zero vector, and the convex cone $Q$ is defined by$$Q:=\{x\in\mathbb{R}^{3}:x_{3}\geq\max(|x_{1}|,|x_{2}|)\}.$$ The optimal solution to the nonrobust problem is $\bar{x}=\mathbf{0}$, and the tangential problem is $$\begin{aligned} v(\Delta\mathcal{U}) & := & \min_{\gamma\in\mathbb{R}^{3}}\gamma_{3}\\ & & \mbox{s.t.}\gamma+[(\Delta A)\mathbf{0}-\Delta b]\in T_{Q}(\mathbf{0})\mbox{ for all }(\Delta A,\Delta b)\in\Delta\mathcal{U}.\end{aligned}$$ Let $a<b$ and the sets $\mathcal{S}_{1}$ and $\mathcal{S}_{2}$ be defined by $$\begin{aligned} \Delta\mathcal{S}_{1} & := & \{\mathbf{0}\}\times\Big\{\Delta b\in\mathbb{R}^{3}:|\Delta b_{1}|\leq\frac{a}{2},|\Delta b_{2}|\leq\frac{b}{2},|\Delta b_{3}|\leq\frac{\delta}{2}\Big\}\\ \Delta\mathcal{S}_{2} & := & \{\mathbf{0}\}\times\Big\{\Delta b\in\mathbb{R}^{3}:|\Delta b_{1}|\leq\frac{b}{2},|\Delta b_{2}|\leq\frac{a}{2},|\Delta b_{3}|\leq\frac{\delta}{2}\Big\}\\ \Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2} & = & \{\mathbf{0}\}\times\Big\{\Delta b\in\mathbb{R}^{3}:|\Delta b_{1}|\leq\frac{a+b}{2},|\Delta b_{2}|\leq\frac{a+b}{2},|\Delta b_{3}|\leq\delta\Big\}.\end{aligned}$$ See Figure \[fig:square-eg\] for an illustration of the convex cone $Q$, and the projection of $\Delta\mathcal{S}_{1}$, $\Delta\mathcal{S}_{2}$ and $\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2}$ onto the 2-dimensional space corresponding to the first 2 coordinates in $\mathbb{R}^{3}$. It is elementary that $v(\Delta\mathcal{S}_{1})=v(\Delta\mathcal{S}_{2})=\frac{b}{2}+\frac{\delta}{2}$, and $v(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})=\frac{a+b}{2}+\delta$. This gives $v(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})<v(\Delta\mathcal{S}_{1})+v(\Delta\mathcal{S}_{2})$. If the calculations had been performed with the second order cone $Q^{\prime}$ defined by $$Q^{\prime}:=\Big\{x\in\mathbb{R}^{3}:x_{3}\geq\sqrt{x_{1}^{2}+x_{2}^{2}}\Big\}$$ instead, then we also get the conclusion $v(\Delta\mathcal{S}_{1}+\Delta\mathcal{S}_{2})<v(\Delta\mathcal{S}_{1})+v(\Delta\mathcal{S}_{2})$. ![\[fig:square-eg\]These diagrams illustrate Example \[exa:square-eg\]. The figure on the left illustrates the convex cone $Q$ in $\mathbb{R}^{3}$, while the figure on the right illustrates the first 2 coordinates for the shapes $\Delta\mathcal{S}_{1}$ and $\Delta\mathcal{S}_{2}$.](box "fig:") ![\[fig:square-eg\]These diagrams illustrate Example \[exa:square-eg\]. The figure on the left illustrates the convex cone $Q$ in $\mathbb{R}^{3}$, while the figure on the right illustrates the first 2 coordinates for the shapes $\Delta\mathcal{S}_{1}$ and $\Delta\mathcal{S}_{2}$.](square "fig:") -------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- Acknowledgments {#acknowledgments .unnumbered} =============== I am grateful to Henry Wolkowicz for conversations and his initial ideas in [@WY09] that led to me studying this problem, and to the Natural Sciences Engineering Research Council (Canada) for supporting the research. The research for this paper was carried out in the Department of Combinatorics and Optimization, University of Waterloo, when I was on a postdoctoral position there, and I gratefully acknowledge them for providing a splendid working environment. [10]{} F. Alizadeh and D. Goldfarb, *Second-order cone programming,* Mathematical Programming, Ser B 95:3–51 (2003). H. Attouch and R. J.-B. Wets, *Approximation and convergence in nonlinear optimization*, in Nonlinear Programming 4, edited by O. Mangasarian, R. Meyer and S. Robinson, pp. 367–395, Academic Press, New York, 1981. A. Ben-Tal, L. El Ghaoui and A. Nemirovski. *Robust Optimization*, Princeton Series in Applied Mathematics, Princeton, 2009. F.H. Clarke. *Optimization and Nonsmooth Analysis*. Wiley, New York, 1983. Republished as Vol. 5, Classics in Applied Mathematics, SIAM, 1990. B.S. Mordukhovich. *Variational Analysis and Generalized Differentiation I and II*., Grundlehren der mathematischen Wissenschaften, Vols 330 & 331, Springer, Berlin, 2006. R. T. Rockafellar, *Convex Analysis*, Princeton, 1970. R. T. Rockafellar, *Clarke’s tangent cones and the boundaries of closed sets in $\mathbb{R}^{n}$*, Nonlinear Analysis: Theory, Methods and Applications, 3, 145–154, 1979. R. T. Rockafellar and R. J.-B. Wets, *Variational Systems, an introduction*, in Multifunctions and Integrands: Stochastic Analysis, Approximation and Optimization, edited by G. Salinetti, Lecture Notes in Mathematics, 1091, pp. 1–54, Springer Verlag, Berlin, 1984. \[RW98\]R.T. Rockafellar and R.J.-B. Wets. *Variational Analysis*, Grundlehren der mathematischen Wissenschaften, Vol 317, Springer, Berlin, 1998. H. Wolkowicz and W.L.N. Yeung, *An alternative approach to sensitivity analysis in linear programming*, 2009 (unpublished).
{ "pile_set_name": "ArXiv" }
--- author: - 'Shihang <span style="font-variant:small-caps;">Shen</span>,$^{1,2}$ Haozhao <span style="font-variant:small-caps;">Liang</span>$^*$,$^{2,3}$ Jie <span style="font-variant:small-caps;">Meng</span>,$^{1,4,5}$ Peter <span style="font-variant:small-caps;">Ring</span>,$^{1,6}$ and Shuangquan <span style="font-variant:small-caps;">Zhang</span>$^{1}$' title: 'Effects of Tensor Force in the Relativistic Scheme: A Case Study of Neutron Drops' --- Introduction ============ This proceeding and the presentation given in the conference are mainly based on our recent progress in the self-consistent relativistic Brueckner-Hartree-Fock (RBHF) theory for finite nuclei [@Shen2016; @Shen2017], the discovery of the tensor effects on the spin-orbit (SO) splittings in neutron drops [@Shen2018; @Shen1802.08108], and the discussion on the spin symmetry in the Dirac sea [@Shen1802.08110]. To understand the nuclear energy density functionals in terms of nucleon-nucleon ($NN$) interaction is one of the present frontiers in nuclear physics. As manifested by the quadrupole moment of deuteron, the tensor force is one of the crucial components in the $NN$ interaction [@Bethe1940]. For heavier nuclei, the shell-model calculations showed that the tensor force plays an indispensable role in the shell-structure evolution in exotic nuclei far away from stability [@Otsuka2005]. Also since around 10 years ago, intensive investigations on the impact of tensor forces have been carried out in the scheme of nuclear density functional theory (DFT), including both the non-relativistic and relativistic frameworks. However, it is still difficult to pin down significant features in experimental data which are only connected to the tensor forces and therefore suitable for an adjustment of their parameters, as reviewed in Ref. [@Sagawa2014]. For example, it was found in Ref. [@Long2007] that the two-proton separation energies of the $N = 82$ isotones can be reproduced only by the relativistic Hartree-Fock (RHF) theory with the effective interaction PKA1, which includes the meson-nucleon tensor interactions generated by both the $\pi$-pseudovector and $\rho$-tensor couplings, whereas such two-proton separation energies cannot be reproduced by the relativistic mean-field (RMF) theory, i.e., the relativistic Hartree theory, which excludes both tensor related couplings. This implies fingerprints of tensor effects could be found in such observable related to the nuclear mass. In contrast, it was found in Ref. [@Lalazissis2009] that if one lets the $\pi N$ coupling strength as a free parameter to fit the nuclear masses on the nuclear chart globally, one obtains the best fit with vanishing tensor forces. Furthermore, evolution of the single-particle spectra was the main motivation for most of the studies on the tensor effects during the past decade. Famous examples are the energy differences between the $1h_{11/2}$ and $1g_{7/2}$ single-proton states along the $Z = 50$ isotopes as well as the energy differences between the $1i_{13/2}$ and $1h_{9/2}$ single-neutron states along the $N = 82$ isotones [@Schiffer2004]. These empirical shell-structure evolutions could be described by the energy density functionals including the tensor force, e.g., by the non-relativistic Skyrme Hartree-Fock (SHF) theory with the effective interaction SLy5+T [@Colo2007], and qualitatively by the RHF theory with PKO1 or PKO3 [@Long2008]. The tensor forces were found to play important roles there, although in terms of tensor force the relativistic scheme uses slightly different language from that in the conventional non-relativistic scheme [@Tarpanov2008; @Moreno-Torres2010]. However, in the context of DFT, the single-particle energies are only defined as auxiliary quantities. In experiment, they are often fragmented and therefore only indirectly accessible. The fragmentation is caused by the effects beyond mean field, i.e., by the admixture of complicated configurations, such as couplings to the low-lying surface vibrations [@Litvinova2011]. Another kind of observable deemed to be sensitive to the tensor force is the nuclear spin-isospin excitation. For example, the spin-dipole resonances in $^{208}$Pb show a very special energy ordering among different angular-momentum components [@Wakasa2012], i.e., $E_x(1^-) \lesssim E_x(2^-) < E_x(0^-)$, instead of the normal one $ E_x(2^-) < E_x(1^-) < E_x(0^-)$. This could be described by the self-consistent SHF + random phase approximation (RPA) calculations with the effective interaction SLy5+T$_W$ or T43 [@Bai2010]. Nevertheless, such particular spin-dipole resonances in $^{208}$Pb cannot be reproduced by any self-consistent relativistic RPA calculation so far, although the fully self-consistent RHF+RPA calculations reproduced nicely the spin-dipole resonances in another nucleus $^{16}$O [@Liang2008; @Liang2012]. This remains a puzzle for how to further improve the relativistic energy density functionals, in particular, the properties of their tensor component. Because of the above aspects, obviously, how to determine precise values for the strength parameters of the tensor forces in universal nuclear energy density functionals by a phenomenological fit to experimental data in finite nuclei is still an open problem. In such a situation we propose to determine these strengths from microscopic *ab initio* calculations based on the well known bare $NN$ forces. Very recently, we achieved for the first time the fully self-consistent RBHF theory for finite nuclei [@Shen2016; @Shen2017], and then we chose the systems of neutron drop to investigate and identify the essential tensor effects [@Shen2018]. Since the proton-neutron interaction is not involved, the equations for neutron drops are much easier to be solved than those for finite nuclei. Therefore they have been investigated in the literature by many different *ab initio* methods [@Pudliner1996; @Smerzi1997; @Pederiva2004; @Gandolfi2011; @Bogner2011; @Maris2013; @Potter2014; @Tews2016] and also by nuclear DFT [@Zhao2016]. Results and Discussion ====================== Starting from the relativistic bare $NN$ interaction Bonn A [@Machleidt1989], the neutron drops confined in an external harmonic-oscillator potential are investigated by using the RBHF theory. Special attention is paid on the possible signature of the tensor force in the neutron-neutron interaction, i.e., the evolution of the SO splitting with the neutron number. See Ref. [@Shen2017] for the detailed formalism and Ref. [@Shen2018] for the numerical details. ![(Color online) From top to bottom, spin-orbit splittings of the $1p,\,1d,\,1f$, and $2p$ doublets of $N$-neutron drops in a HO trap with $\hbar\omega = 10$ MeV, calculated by the RBHF theory with the relativistic bare nucleon-nucleon interaction Bonn A. For comparison, (left) the results obtained by the RMF theory with the effective interactions NL3, PK1, PKDD, DD-ME2, and PC-PK1, and (right) the results obtained by the RHF theory with the effective interaction PKO1 with different $\pi N$ coupling strengths characterized by $\lambda$. Taken from Ref. [@Shen2018].[]{data-label="fig1"}](fig1a.eps "fig:"){width="7cm"}           ![(Color online) From top to bottom, spin-orbit splittings of the $1p,\,1d,\,1f$, and $2p$ doublets of $N$-neutron drops in a HO trap with $\hbar\omega = 10$ MeV, calculated by the RBHF theory with the relativistic bare nucleon-nucleon interaction Bonn A. For comparison, (left) the results obtained by the RMF theory with the effective interactions NL3, PK1, PKDD, DD-ME2, and PC-PK1, and (right) the results obtained by the RHF theory with the effective interaction PKO1 with different $\pi N$ coupling strengths characterized by $\lambda$. Taken from Ref. [@Shen2018].[]{data-label="fig1"}](fig1b.eps "fig:"){width="7cm"} First of all, we show in Fig. \[fig1\] the SO splittings of the $1p,\,1d,\,1f$, and $2p$ doublets of $N$-neutron drops in a HO trap calculated by the RBHF theory. We can find a clear pattern from these *ab initio* results. Each SO splitting decreases as the SO aligned orbital with $j_>= l + 1/2$ is filled and reaches a minimum when this orbital is fully occupied, and then the SO splitting increases as the SO anti-aligned orbital with $j_< = l - 1/2$ begins to be occupied and reaches a maximum when this orbital is fully occupied. This pattern repeats as the number of neutron continues to increase. Otsuka *et al.* [@Otsuka2005] have found a similar effect between neutrons and protons in nuclei. They explained it in terms of the monopole effect of the tensor force, which produces an attractive interaction between a proton in a SO aligned orbital with $j_>=l + 1/2$ and a neutron in a SO anti-aligned orbital with $j'_<=l' - 1/2$, and a repulsive interaction between the same proton and a neutron in a SO aligned orbit with $j'_>=l'+ 1/2$. They also pointed out that a similar mechanism, but with smaller amplitude, exists also for the tensor interaction between neutrons in the isovector $T=1$ channel [@Otsuka2005]. That is exactly what is seen in the figure. For instance, the SO splitting of the $1d$ doublets decreases from $N=20$ to $N=28$, because the interaction with the neutrons filling into the $1f_{7/2}$ shell above $N=20$ shifts the $1d_{5/2}$ orbital upward and the $1d_{3/2}$ orbital downward. Above $N=28$, the neutrons filling into $2p_{1/2}$ and $1f_{5/2}$ interact with the $1d$-neutrons in the opposite way, and thus increase the SO splitting. In other words, a clear signature of the tensor effect on the SO splitting in neutron drops is identified, for the first time, in the pure relativistic scheme. Such a clear tensor effect can be served as a benchmark for the existing relativistic energy density functionals as well as as an important guide for their future developments. In the left panel of Fig. \[fig1\], we show the SO splittings of the same neutron drops calculated by various RMF energy density functionals, including the nonlinear meson-exchange models NL3 [@Lalazissis1997] and PK1 [@Long2004], the density-dependent meson-exchange models PKDD [@Long2004] and DD-ME2 [@Lalazissis2005], and the nonlinear point-coupling model PC-PK1 [@Zhao2010]. It is seen that none of the results is able to reproduce the specific evolution pattern of the SO splitting generated by the *ab initio* RBHF calculations, because it is known that there is no tensor component in the RMF energy density functionals. In the right panel of Fig. \[fig1\], we show the same results calculated by the RHF theory with the density-dependent meson-exchange model PKO1 [@Long2006], which includes the tensor interaction generated by the $\pi$-pseudovector coupling. It is seen that the specific evolution pattern of the SO splitting by the RBHF theory can now be clearly reproduced, although the size of the effect is somewhat too small. This can be understood by the fact that in the original fit of PKO1 [@Long2006] only the information of nuclear mass was used, and thus the strength of the tensor force was basically out of control [@Lalazissis2009]. In order to have more quantitative ideas on the proper strength of the tensor force, we multiplied a factor $\lambda$ in front of the $\pi N$ coupling, whereas other parameters in PKO1 remain unchanged. It is seen that the size of the tensor effect does significantly depend on the value of $\lambda$, where the cases of $\lambda = 0.7,\, 1.0$, and $1.3$ are shown in the figure. With $\lambda = 1.3$, the specific evolution pattern of the SO splitting generated by the *ab initio* RBHF calculations can be nicely reproduced. In such a way, the *ab initio* calculations will guide the complete fit of the RHF energy density functionals in the near future. Summary and Perspectives ======================== We have studied the systems of neutron drop confined in an external HO potential using the RBHF theory with the relativistic bare $NN$ interaction. It was found that the SO splitting decreases as the SO aligned orbital with $j_>= l + 1/2$ is being occupied, and increases again as the next SO anti-aligned orbital with $j_< = l - 1/2$ being occupied. This is similar to the effects of the tensor forces between neutrons and protons found in Ref. [@Otsuka2005]. Such a specific pattern is not restricted to a specific mass region. It is generally valid for all the neutron numbers $4 \leq N \leq 50$ under investigation. Therefore, we can expect that a similar feature is valid also for realistic nuclei all over the nuclear chart. This evolution pattern of the SO splitting can be reproduced by the RHF energy density functional with the tensor force, or even perfectly by readjusting a single parameter $\lambda$ for the tensor coupling strength, but cannot be reproduced by any RMF energy density functional. In addition, it is important to remark that in the RBHF theory there are no higher-order configurations, which means the effects like particle-vibrational coupling [@Litvinova2011] are not included in these meta-data. This allows to adjust the tensor force to these meta-data without the ambiguity of additional effects, because these effects are neither included in the present concept of nuclear DFT, nor in the present RBHF calculations. Based on these recent progress, the ongoing research with different strategies include: (1) to carry out a complete fit of relativistic energy density functional by taking into account the experimental data of nuclear masses and radii, as well as these *ab initio* meta-data of the SO splittings in neutron drops to control the strength of the tensor force; (2) to carry out a fit in each spin-isospin channel by using each spin-isospin density distribution generated by the *ab initio* calculations, where the Kohn-Sham single-particle potentials and orbitals can be uniquely determined from the density distributions by the inverse Kohn-Sham method [@Wang1993]; (3) to derive the energy density functional directly from the bare $NN$ interactions in an *ab initio* way [@Liang2018]. Acknowledgements {#acknowledgements .unnumbered} ================ This work was partly supported by the Major State 973 Program of China No. 2013CB834400, the NSFC under Grants No. 11335002, No. 11375015, and No. 11621131001, the Overseas Distinguished Professor Project from Ministry of Education of China No. MS2010BJDX001, the Research Fund for the Doctoral Program of Higher Education of China under Grant No. 20110001110087, and the DFG (Germany) cluster of excellence “Origin and Structure of the Universe” (www.universe-cluster.de). S.S. would like to thank the short-term Ph.D. student exchange program of Peking University and the RIKEN IPA project, and H.L. would like to thank the RIKEN iTHES project and iTHEMS program. [19]{} S.H. Shen, J.N. Hu, H.Z. Liang, J. Meng, P. Ring, and S.Q. Zhang, Chin. Phys. Lett. **33**, 102103 (2016). S.H. Shen, H.Z. Liang, J. Meng, P. Ring, and S.Q. Zhang, Phys. Rev. C **96**, 014316 (2017). S.H. Shen, H.Z. Liang, J. Meng, P. Ring, and S.Q. Zhang, Phys. Lett. B **778**, 344 (2018). S.H. Shen, H.Z. Liang, J. Meng, P. Ring, and S.Q. Zhang, arXiv:1802.08108 \[nucl-th\]. S.H. Shen, H.Z. Liang, J. Meng, P. Ring, and S.Q. Zhang, arXiv:1802.08110 \[nucl-th\]. H.A. Bethe, Phys. Rev. **57**, 390 (1940). T. Otsuka, T. Suzuki, R. Fujimoto, H. Grawe, and Y. Akaishi, Phys. Rev. Lett. **95**, 232502 (2005). H. Sagawa and G. Colò, Prog. Part. Nucl. Phys. **76**, 76 (2014). W.H. Long, H. Sagawa, N. Van Giai, and J. Meng, Phys. Rev. C **76**, 034314 (2007). G.A. Lalazissis, S. Karatzikos, M. Serra, T. Otsuka, and P. Ring, Phys. Rev. C **80**, 041301 (2009). J.P. Schiffer *et al.*, Phys. Rev. Lett. **92**, 162501 (2004). G. Colò, H. Sagawa, S. Fracasso, and P.F. Bortignon, Phys. Lett. B **646**, 227 (2007). W.H. Long, H. Sagawa, J. Meng, and N. Van Giai, Europhys. Lett. **82**, 12001 (2008). D. Tarpanov, H.Z. Liang, N. Van Giai, and C. Stoyanov, Phys. Rev. C **77**, 054316 (2008). M. Moreno-Torres, M. Grasso, H.Z. Liang, V. De Donno, M. Anguiano, and N. Van Giai, Phys. Rev. C **81**, 064327 (2010). E.V. Litvinova and A.V. Afanasjev, Phys. Rev. C **84**, 014305 (2011). T. Wakasa *et al.*, Phys. Rev. C **85**, 064606 (2012). C.L. Bai, H.Q. Zhang, H. Sagawa, X.Z. Zhang, G. Colò, and F.R. Xu, Phys. Rev. Lett. **105**, 072501 (2010). H.Z. Liang, N. Van Giai, and J. Meng, Phys. Rev. Lett. **101**, 122502 (2008). H.Z. Liang, P.W. Zhao, and J. Meng, Phys. Rev. C **85**, 064302 (2012). B.S. Pudliner, A. Smerzi, J. Carlson, V.R. Pandharipande, S.C. Pieper, and D.G. Ravenhall, Phys. Rev. Lett. **76**, 2416 (1996). A. Smerzi, D.G. Ravenhall, and V.R. Pandharipande, Phys. Rev. C **56**, 2549 (1997). F. Pederiva, A. Sarsa, K. Schmidt, and S. Fantoni, Nucl. Phys. A **742**, 255 (2004). S. Gandolfi, J. Carlson, and S.C. Pieper, Phys. Rev. Lett. **106**, 012501 (2011). S.K. Bogner, R.J. Furnstahl, H. Hergert, M. Kortelainen, P. Maris, M. Stoitsov, and J.P. Vary, Phys. Rev. C **84**, 044306 (2011). P. Maris, J.P. Vary, S. Gandolfi, J. Carlson, and S.C. Pieper, Phys. Rev. C **87**, 054318 (2013). H.D. Potter, S. Fischer, P. Maris, J.P. Vary, S. Binder, A. Calci, J. Langhammer, and R. Roth, Phys. Lett. B **739**, 445 (2014). I. Tews, S. Gandolfi, A. Gezerlis, and A. Schwenk, Phys. Rev. C **93**, 024305 (2016). P.W. Zhao and S. Gandolfi, Phys. Rev. C **94**, 041302 (2016). R. Machleidt, Adv. Nucl. Phys. **19**, 189 (1989). G.A. Lalazissis, J. König, and P. Ring, Phys. Rev. C **55**, 540 (1997). W.H. Long, J. Meng, N. Van Giai, and S.G. Zhou, Phys. Rev. C **69**, 034319 (2004). G.A. Lalazissis, T. Nikšić, D. Vretenar, and P. Ring, Phys. Rev. C **71**, 024312 (2005). P.W. Zhao, Z.P. Li, J.M. Yao, and J. Meng, Phys. Rev. C **82**, 054319 (2010). W.H. Long, N. Van Giai, and J. Meng, Phys. Lett. B **640**, 150 (2006). Y. Wang and R.G. Parr, Phys. Rev. A **47**, R1591 (1993). H.Z. Liang, Y.F. Niu, and T. Hatsuda, Phys. Lett. B **779**, 436 (2018).
{ "pile_set_name": "ArXiv" }
--- address: | Theoretical Physics Group, The Blackett Laboratory\ Imperial College London, Prince Consort Road\ London, SW7 2AZ, UK\ author: - 'A. Hanany, G. Torri' title: Brane tilings and supersymmetric gauge theories --- Supersymmetric gauge theories ============================= In the past 40 years, gauge theories have played a crucial role in High Energy Physics. In particular, of their biggest successes, has been the formulation of the Standard Model from a gauge principle. In this vast sea of theories, the class of supersymmetric gauge theories deserves special attention. In fact, the constraints that supersymmetry imposes, make a gauge theory much more well-behaved and easier to investigate from a theoretical point of view. Furthermore, supersymmetric gauge theories are shown to arise naturally in Superstring Theory, possibly one of the preferred candidates to solve the problems that were left open by the Standard Model. In particular, the gauge degrees of freedom are brought into the theory by the open strings that are stretched between two D-branes, and are confined precisely on them. Roughly speaking, we could say that, in String Theory, supersymmetric gauge theories arise on the world-volume of D-branes. The interesting thing about this is that the structure of these gauge theories is highly sensitive to the geometry that the D-branes are probing, thus leading to a connection between the shape of space-time and the nature of the interactions of the Universe. Studying supersymmetric gauge theories in String Theory is not only an interesting endeavour *per se*, but it could also shed some light on the fascinating duality between gauge theories and gravity conjectured by Maldacena more than 10 years ago. In the following we will be mostly concerned about super-Yang-Mills gauge theories in $(3+1)$ dimensions, although similar considerations can be made for Chern-Simons theories in $(2+1)$ dimensions, *mutatis mutandis*. Quivers ======= As we mentioned above, the structure of supersymmetric gauge theories is highly constrained precisely by supersymmetry. In particular, a supersymmetric lagrangian is unambiguously identified by specifying the number and types of gauge groups, the matter content and a superpotential. Furthermore, given the simplicity of the information needed to characterise a supersymmetric gauge theory, it turns out that many of them can be represented by a graph called *quiver*. This is essentially a directed graph containing arrows and nodes with the convention that: - each node represents a gauge group $U(N)$; - each arrow going from a node $a$ to a different node $b$ represents a field $X_{ab}$ in the bifundamental representation $(\mathbf{N},\overline{\mathbf{N}})$ of $U(N_a) \times U(N_b)$. - each loop on a node $a$ represents a field $\phi_a$ in the adjoint representation of $U(N_a)$ . - each superpotential term corresponds to a closed loop in the quiver[^1]; As an example, Figure \[fig:quiverconi\] shows the quiver of the well known conifold theory. ![Quiver of the conifold theory[]{data-label="fig:quiverconi"}](quiverconi.pdf){height="1.5cm"} The theories that can be represented by a quiver, usually called *quiver gauge theories*, constitute a class of theories which is quite easy to investigate. In fact, they are usually characterised by a very simple superpotential and they contain matter that transform in bifundamental or adjoint representations of the gauge groups. Brane tilings ============= Another way of representing supersymmetric gauge theories is through a type of graph called *brane tiling*. It is a bi-periodic and bi-partite graph with a repeating structure which is called *fundamental domain*. In order to read off from the brane tiling the necessary information to reconstruct the lagrangian, one must keep in mind that: - each face corresponds to a $U(N)$ gauge group; - each edge corresponds to a bi-fundamental field; - each node corresponds to a superpotential term; The nice feature about brane tilings is that they allow one to write down the lagrangian of a supersymmetric gauge theory in a completely unambiguous way[^2]. As an example, Figure \[fig:tilingconi\] represents the brane tiling of the conifold theory. ![Tiling of the conifold theory[]{data-label="fig:tilingconi"}](tilingconi.pdf){height="6cm"} Once a supersymmetric gauge theory is specified through a brane tilings, it is possible to investigate relevant features of this theory. In particular, through a technique called *forward algorithm*[^3], it is possible to study the vacuum moduli space, i.e. the space of zero-energy solutions to the F-terms and the D-terms of the gauge theory. This algorithm also makes it easy to determine the R-charges of the operators of the theory one is studying. The brane tilings have been extensively used in the past to investigate a phenomenon called *toric duality* for D3-branes. Roughly speaking, this corresponds to the situation where two (or more) different gauge theories have the same vacuum moduli space. This duality has been shown to be equivalent to Seiberg duality. Brane tilings seem also like a very natural set-ups to study the *Higgs mechanism* in for supersymmetric gauge theories. In fact, on the tiling this phenomenon is implemented simply as the removal of one or more edges. This makes it very simple also to understand how different gauge theories can be connected through the Higgs mechanism. Finally, a very fascinating application of brane tilings is the study of supersymmetric Chern-Simons theories in $(2+1)$ dimensions. In 2008, Aharony, Bergman, Jafferis and Maldacena provided the first example of conformal field theory living on an M2-brane probing flat space. With brane tilings it has been possible to extend this duality to more complicated backgrounds and to investigate fascinating features of such theories. [6]{} O. Aharony, S. S. Gubser, J. M. Maldacena, H. Ooguri and Y. Oz, Phys. Rept.  [**323**]{} (2000) 183 \[arXiv:hep-th/9905111\]. I. R. Klebanov, arXiv:hep-th/0009139. J. M. Maldacena, Adv. Theor. Math. Phys.  [**2**]{} (1998) 231 \[Int. J. Theor. Phys.  [**38**]{} (1999) 1113\] \[arXiv:hep-th/9711200\]. A. Hanany and K. D. Kennaway, hep-th/0503149. S. Franco, A. Hanany, K. D. Kennaway, D. Vegh and B. Wecht, JHEP [**0601**]{}, 096 (2006) \[arXiv:hep-th/0504110\]. S. Franco, A. Hanany, D. Martelli, J. Sparks, D. Vegh and B. Wecht, JHEP [**0601**]{}, 128 (2006) \[arXiv:hep-th/0505211\]. I. R. Klebanov and G. Torri, arXiv:0909.1580 \[hep-th\]. A. Hanany and A. Zaffaroni, arXiv:0808.1244 \[hep-th\]. A. Hanany, D. Vegh, A. Zaffaroni, arXiv:0809.1440. J. Davey, A. Hanany, N. Mekareeya and G. Torri, arXiv:0903.3234 \[hep-th\]. J. Davey, A. Hanany, N. Mekareeya and G. Torri, JHEP [**0911**]{} (2009) 028 \[arXiv:0908.4033 \[hep-th\]\]. [^1]: However not all the loops correspond to a superpotential term! [^2]: With a quiver this is not quite true since not all the closed loops correspond to terms in the superpotential. [^3]: The name suggests the existence of an *inverse algorithm* which, starting from a certain Calabi-Yau, gives the brane tiling of the theory that has that manifold as its vacuum moduli space.
{ "pile_set_name": "ArXiv" }
--- abstract: 'The excess low-frequency vibrational spectrum, called boson peak, and non-affine elastic response are the most important particularities of glasses. Herein, the vibrational and mechanical properties of polymeric glasses are examined by using coarse-grained molecular dynamics simulations, with particular attention to the effects of the bending rigidity of the polymer chains. As the rigidity increases, the system undergoes a glass transition at a higher temperature (under a constant pressure), which decreases the density of the glass phase. The elastic moduli, which are controlled by the decrease of the density and the increase of the rigidity, show a non-monotonic dependence on the rigidity of the polymer chain that arises from the non-affine component. Moreover, a clear boson peak is observed in the vibrational density of states, which depends on the macroscopic shear modulus $G$. In particular, the boson peak frequency is scaled as $\omega_\mathrm{BP} \propto \sqrt{G}$. These results provide a positive correlation between the boson peak, shear elasticity, and the glass transition temperature.' author: - Naoya Tomoshige - Hideyuki Mizuno - Tatsuya Mori - Kang Kim - Nobuyuki Matubayasi title: | Boson peak, elasticity, and glass transition temperature in polymer glasses:\ Effects of the rigidity of chain bending --- Introduction ============ Glasses show vibrational and mechanical properties that are markedly different from other crystalline materials [@Phillips_1981; @Alexander_1998]. Thermal measurements and scattering experiments have been performed to study the properties of various glassy systems, such as covalent-bonding [@Zeller_1971; @Buchenau_1984; @Nakayama_2002; @Monaco_2006; @Baldi_2010; @Chumakov_2011], molecular [@Yamamuro_1996; @Ramos_2003; @Monaco_2009; @Shibata_2015; @Kabeya_2016], metallic [@Berg_1983; @Yong_2008; @Bruna_2011; @Huang_2014], and polymeric [@Niss_2007; @Hong_2008; @Caponi_2011; @Perez-Castaneda2_2014; @Terao_2018; @Zorn_2018] glasses. For instance, the excess vibrational modes at low frequencies and the excess heat capacity at low temperatures exceeding the Debye predictions, which describe the corresponding crystalline values, have been observed universally in various glassy materials. This phenomenon, which is referred to as the boson peak (BP), has been widely studied. The ideas of elastic heterogeneities [@Schirmacher_2006; @Schirmacher_2007; @Schirmacher_2015] and criticality near isostatic state and marginally stable state [@Wyart_2005; @Wyart_2006; @Wyart_2010; @DeGiuli_2014] have been introduced, following the recent theoretical advances for understanding the origin of anomalies in glasses. Based on these theories, the mean-field formulations have been developed by using the effective medium technique [@Schirmacher_2006; @Schirmacher_2007; @Schirmacher_2015; @Wyart_2010; @DeGiuli_2014]. In addition, more recent studies [@Milkus_2016; @Krausser_2017] have focused on the local inversion-symmetry breaking, which can explain the microscopic origin of the BP. Molecular dynamics (MD) simulations play an essential role for studying the vibrational and mechanical properties of glasses. Firstly, MD simulations enable to assess the theoretical predictions. In fact, various MD simulations have been performed on simple atomic glasses, e.g., Lennard-Jones (LJ) systems [@Schober_1996; @Mazzacurati_1996; @Shintani_2008; @Monaco2_2009; @Mizuno2_2013; @Lerner_2016; @Wang_2019]. Concerning the isostaticity and marginal stability [@Wyart_2005; @Wyart_2006; @Wyart_2010; @DeGiuli_2014], the systems with a finite-ranged, purely repulsive potential have also been studied [@Silbert_2005; @Silbert_2009; @Vitelli_2010; @Xu_2010], and are considered as the simplest model of glasses. In particular, it is crucial for MD simulations to solve finite-dimensional effects that are not captured by the mean-field treatments [@Mizuno_2017; @Shimada_2018; @Mizuno_2018]. Secondly, MD simulations perform quasi-experiments on well-defined systems and access data that cannot be examined experimentally. Relevant systems to experiments and applications have been simulated, including covalent-bonding [@Taraskin_1997; @Taraskin_1999; @Horbach_2001; @Leonforte_2006; @Beltukov_2016; @Beltukov_2018], metallic [@Derlet_2012; @Fan_2014; @Crespo_2016; @Brink_2016], polymeric [@Jain_2004; @Schnell:2011he; @Ness:2017cc; @Milkus:2018ha; @Giuntoli:2018fv] glasses. These simulation studies complete theoretical understandings based on simple systems and experimental observations of more complex systems. The vibrational properties and the BP of polymeric glasses have been studied by both of experiments [@Niss_2007; @Hong_2008; @Caponi_2011; @Perez-Castaneda2_2014; @Terao_2018; @Zorn_2018] and MD simulations [@Jain_2004; @Schnell:2011he; @Ness:2017cc; @Milkus:2018ha; @Giuntoli:2018fv]. The effects caused by non-covalent bonds including bending forces and chain length represent an important feature of polymer glasses. Previous experiments [@Niss_2007; @Hong_2008] have investigated the effects of the pressure or densification on the frequency and intensity of the BP in polymeric glasses. It was demonstrated that the evolution of the BP with pressure cannot be scaled by the Debye values (i.e., the Debye frequency and the Debye level). Therefore, the pressure effects cannot be explained only by the variation of macroscopic elasticity. In contrast, another experiment [@Caponi_2011] has shown that the polymerization effects on the BP is explained by the change in macroscopic elasticity as the frequency and intensity variations of the BP are both scaled by the Debye values. In addition, Zaccone *et al.* have recently performed MD simulations to calculate the vibrational density of states (vDOS) in polymeric glasses by changing the chain length and the rigidity of the chain bending [@Milkus:2018ha]. This work studied the vibrational eigenstates in a wide range of frequencies and the effects of the chain length and bending rigidity on the high-frequency spectra. Furthermore, Giuntoli and Leporini studied the BP of polymeric glasses having chains with highly rigid bonds [@Giuntoli:2018fv]. It was demonstrated that the BP decouples with macroscopic elasticity and arises from non-bonding interactions only. Although these studies [@Milkus:2018ha; @Giuntoli:2018fv] have helped understand polymeric glass properties, the effects of bending rigidity and chain length on the low-frequency spectra and BP need to be further studied. Herein, the vDOS and the elastic moduli of polymeric glasses are analyzed through coarse-grained MD simulations. In particular, the connection between the BP and elasticity as well as the glass transition temperature is explored by systematically changing the bending stiffness of short and long polymer chains. The contributions of the present study are given as follows. We demonstrate that polymeric glasses can exhibit extremely-large non-affine elastic response (compared to atomic glasses), whereas the BP is simply scaled by the behavior of macroscopic shear modulus. This behavior of the BP can be explained by the theory of elastic heterogeneities [@Schirmacher_2006; @Schirmacher_2007; @Schirmacher_2015]. Our results indicate that effects of the bending rigidity on the BP are encompassed in change of macroscopic elasticity, which is in contrast to effects of pressure [@Niss_2007; @Hong_2008], but instead is similar to effects of polymerization [@Caponi_2011]. Furthermore, we show the positive correlation among the BP, elasticity, and the glass transition temperature. Finally, we will discuss the relaxation dynamics in the liquid state, in relation to our results of low-frequency vibrational spectra. System Description ================== Coarse-grained MD simulations are performed by using the Kremer–Grest model [@Kremer:1990iv], which treats polymer chains as linear series of monomer beads (particles) of mass $m$. Each polymer chain is composed of $L$ monomer beads, and two cases are considered in this study: long chain length with $L=50$ and short chain length with $L=3$. In a three-dimensional cubic simulation box under periodic boundary conditions, $N_\mathrm{p} = 5000$ and $4998$ is defined as the total number of monomers for $L=50$ and $L=3$ respectively, which means that the number of polymeric chains is $N_\mathrm{p}/L = 100$ for $L=50$ and $1666$ for $L=3$. The polymer chain is modeled by three types of inter-particle potentials as follows. Firstly, all the monomer particles interact via the LJ potential: $$U_\mathrm{LJ}(r) = 4\varepsilon_\mathrm{LJ} \left[ \left( \frac{\sigma}{r} \right)^{12} - \left( \frac{\sigma}{r} \right)^{6} \right], \label{eq:LJ}$$ where $r$ is the distance between two monomers, $\sigma$ is the diameter of monomer, and $\varepsilon_\mathrm{LJ}$ is the energy scale of the LJ potential. The LJ potential is truncated at the cut-off distance of $r_c = 2.5 \sigma$, where the potential and the force (first derivative of the potential) are shifted to zero continuously [@Shimada:2018fp]. Throughout this study, the mass, length, and energy scales are measured in units of $m$, $\sigma$, $\varepsilon_\mathrm{LJ}$, respectively. The temperature is measured by $\varepsilon_\mathrm{LJ}/k_\mathrm{B}$ ($k_\mathrm{B}$ is the Boltzmann constant). Secondly, sequential monomer-beads along the polymeric chain are connected by a finitely extensible nonlinear elastic (FENE) potential: $$U_\mathrm{FENE}(r) = \left\{ \begin{aligned} & -\frac{\varepsilon_\mathrm{FENE}}{2} R_0^2 \ln \left[ 1 - \left( \frac{r}{R_0} \right)^{2} \right] & (r \le R_0), \\ & \infty & (r > R_0), \end{aligned} \right. \label{eq:FENE}$$ where $\varepsilon_\mathrm{FENE}$ is the energy scale of the FENE potential, and $R_0$ is the maximum length of the FENE bond. Their values are defined as $\varepsilon_\mathrm{FENE} = 30$ and $R_0 = 1.5$, according to Ref. [@Milkus:2018ha]. Finally, three consecutive monomer beads along the chain interact via the bending potential defined as follows: $$U_\mathrm{bend}(\theta) = \varepsilon_\mathrm{bend} \left[ 1 - \cos(\theta - \theta_0) \right], \label{eq:bend}$$ where $\theta$ is the angle formed by three consecutive beads, and $\varepsilon_\mathrm{bend}$ is the associated energy scale. This potential intends to stabilize the angle $\theta$ at $\theta_0$ that we set as $\theta_0 = 109.5^\circ$. Here, the value of $\varepsilon_\mathrm{bend}$ in a wide range from $\varepsilon_\mathrm{bend} = 10^{-3}$ to $10^4$, and the effects of the bending rigidity on the vibrational and mechanical properties of the polymeric system are studied. MD simulations are performed by using the LAMMPS [@Plimpton_1995; @lammps]. The polymeric system is first equilibrated in the melted, liquid state at a temperature $T=1.0$. Further, the system is cooled down under a fixed pressure condition of [$P =0$]{} and with a cooling rate of $dT/dt = 10^{-4}$. During the cooling process, the glass transition occurs at a particular temperature, i.e., the glass transition temperature. After the glass transition, the system is quenched down towards the zero temperature, i.e., $T=0$ state. ![\[fig1\] [Glass transition temperature and density in the glass state.]{} (a) The specific volume $v$ versus the temperature $T$ during the process that the system is cooled down from the liquid state to the glass state. The color of line indicates the value of bending rigidity $\varepsilon_\mathrm{bend}$ according to the color bar. (b) Glass transition temperature $T_g$ (triangles) and density $\rho$ at zero temperature after the glass transition (circles) are plotted against $\varepsilon_\mathrm{bend}$. The chain length is $L=50$. ](fig1.pdf){width="48.00000%"} ![\[fig2\] [Conformation of polymeric chains.]{} (a) Probability distribution of angle formed by three consecutive beads along the chain, $P(\theta)$, is presented for several different rigidities $\varepsilon_\mathrm{bend}$. The color of line indicates the value of bending rigidity $\varepsilon_\mathrm{bend}$ according to the color bar. (b) Radius of inertia $R_g$ is plotted as a function of $\varepsilon_\mathrm{bend}$. The chain length is $L=50$. ](fig2.pdf){width="48.00000%"} ![image](fig3.pdf){width="96.00000%"} Results ======= Glass transition temperature ---------------------------- When the polymeric system is cooled down from the liquid state under a constant pressure, the volume of the system monotonically decreases with decreasing the temperature. Figure \[fig1\]a shows the specific volume $v$ as a function of the temperature $T$ for several different bending rigidities $\varepsilon_\mathrm{bend}$ and the chain length $L=50$. For each value of $\varepsilon_\mathrm{bend}$, the slope of the $v$-$T$ curve clearly presents a discontinuous change at a certain temperature, which is defined as the glass transition temperature $T_g$. Figure \[fig1\]b (triangles) presents the value of $T_g$ as a function of $\varepsilon_\mathrm{bend}$. As the rigidity increases from $\varepsilon_\mathrm{bend} = 1$ to $10^3$, $T_g$ progressively increases from $T_g \simeq 0.45$ to $0.75$. Below $\varepsilon = 1$ and above $\varepsilon = 10^3$, the variation of $T_g$ is low or even negligible. In addition, Figure \[fig1\]b (circles) plots the density $\rho(=1/v)$ of the system that is quenched down to $T=0$. The density decreases from $\rho \simeq 1.09$ to $0.97$ as the rigidity increases from $\varepsilon_\mathrm{bend} =1$ to $10^3$. As the chain bending becomes rigid, the glass transition occurs at a higher temperature, and as a result, the density in the glass state becomes lower. The similar observation was obtained by Milkus *et al* [@Milkus:2018ha]. These behaviors of $T_g$ and $\rho$ can be understood by studying the microscopic conformation of the polymeric chains. Figure \[fig2\]a presents the probability distribution of the angle formed by three consecutive beads along the chain, $P(\theta)$, when changing the rigidity $\varepsilon_\mathrm{bend}$. Two peaks are observed at approximately $\theta \simeq 70^\circ$ and $\theta \simeq 120^\circ$ for a low rigidity ($\varepsilon_\mathrm{bend}\le 1$). A similar distribution $P(\theta)$ was also reported in Ref. [@Milkus:2018ha]. As the rigidity increases, the peak position in $P(\theta)$ shifts towards $\theta_0=109.5^\circ$. It is noted that the bending potential $U_\mathrm{bend}(\theta)$ in Eq. (\[eq:bend\]) tends to stabilize the angle $\theta$ at $\theta_0 = 109.5^\circ$. In addition, Figure \[fig2\]b presents the radius of gyration $R_g$ as a function of $\varepsilon_\mathrm{bend}$. It can be observed that $R_g$ increases from $R_g \simeq 11.5$ to $16.5$ with an increasing $\varepsilon_\mathrm{bend}$. Importantly, these variations of conformation are induced intensively when the rigidity increases from $\varepsilon_\mathrm{bend} = 1$ to $10^3$, which exactly matches the region where variations of $T_g$ and $\rho$ are observed in Fig. \[fig1\]b. Therefore, it can be concluded that the conformation changes of the polymeric chains control the glass transition temperature and the density. In fact, as the rigidity of the chain bending increases, the angle $\theta$ of the polymer chains tends to be stabilized at $\theta_0 = 109.5^\circ$ and the radius of inertia increases. As a result, the glass transition occurs at a higher temperature and the lower density (larger volume). At $\varepsilon_\mathrm{bend} \ls 1$, the effect of the bending interaction of Eq. (\[eq:bend\]) is weak compared to those of the LJ and FENE components of Eqs. (\[eq:LJ\]) and (\[eq:FENE\]). However, at $\varepsilon_\mathrm{bend} \gs 10^3$, the opposite phenomenon occurs. It is noted that the glass transition occurs at a lower temperature for $L=3$ than for $L=50$, which is consistent with a previous report [@Durand_2010]. Correspondingly, the values of $\rho$ for $L=3$ becomes larger than that of $L=50$. However, common results were observed between $L=3$ and $50$ with respect to the dependences on the rigidity $\varepsilon_\mathrm{bend}$. Specifically, $T_g$ and $\rho$, as well as the conformation of the polymeric chains progressively change when the rigidity increases from $\varepsilon_\mathrm{bend} = 1$ to $10^3$, which also occurs for $L=50$. Elastic properties ------------------ The elastic properties of polymer glasses are studied by changing the strength of bending rigidity. An external strain is applied to the system at $T=0$, which enables to measure the corresponding elastic moduli. Specifically, the volume-changing bulk deformation and the volume-conserving shear deformation are applied, which provide the bulk modulus $K$ and the shear modulus $G$, respectively [@Mizuno_2013]. Figure \[fig3\] presents the values of $K$ and $G$ as functions of $\varepsilon_\mathrm{bend}$. Disordered systems exhibit large non-affine elastic responses [@Alexander_1998]. The elastic moduli, $M=K$ and $G$, are decomposed into affine moduli $M_\mathrm{A}$ and non-affine moduli $M_\mathrm{NA}$, i.e., $M = M_\mathrm{A} - M_\mathrm{NA}$ [@Tanguy_2002; @Lemaitre_2006]. In Fig. \[fig3\], these affine and non-affine components are also presented. First, the bulk modulus $K$ is analyzed for $L=50$ and presented Fig. \[fig3\]a. The affine component $K_\mathrm{A}$ decreases from $K_\mathrm{A} \simeq 155$ to $130$ as $\varepsilon_\mathrm{bend}$ changes from $1$ to $10^3$. The reduction of $K_\mathrm{A}$ is caused by the decrease of the density $\rho$ with the increasing $\varepsilon_\mathrm{bend}$ (see Fig. \[fig1\]b). In contrast, the non-affine component $K_\mathrm{NA}$ shows a non-monotonic dependence on the $\varepsilon_\mathrm{bend}$. In particular, $K_\mathrm{NA}$ slightly increases from $\varepsilon_\mathrm{bend} = 1$ to $30$, which is induced by the decrease of the density $\rho$. As $\varepsilon_\mathrm{bend}$ is further increased above $\varepsilon_\mathrm{bend} = 30$, $K_\mathrm{NA}$ decreases. This is because the non-affine relaxation process is constrained due to the large rigidity of $\varepsilon_\mathrm{bend}$. As a result, the total modulus of $K=K_\mathrm{A}-K_\mathrm{NA}$ also presents a non-monotonic behavior, which is demonstrated in Fig. \[fig3\]a. From $\varepsilon_\mathrm{bend}=1$ to $10^2$, $K$ decreases from $K \simeq 80$ to $60$, which is caused by the reduction of $K_\mathrm{A}$. Moreover, $K$ increases from $K \simeq 60$ to $65$ above $\varepsilon_\mathrm{bend} = 100$, which is caused by the reduction of $K_\mathrm{NA}$. Therefore, the $\varepsilon_\mathrm{bend}$ dependence of the bulk modulus $K$ is determined by the competition between the density reduction and the increase in the bending rigidity. Further, the shear modulus $G$ is analyzed for $L=50$ and presented Fig. \[fig3\]c. It can be observed that the bending rigidity strongly affects the shear modulus compared to the bulk modulus. Particularly, above $\varepsilon_\text{bend} = 10^2$, both of the affine $G_\mathrm{A}$ and non-affine $G_\mathrm{NA}$ components considerably increase. As the shear deformation is anisotropic and causes deformations of the angles $\theta$ of polymeric chains, its response is expected to be highly affected by the bending rigidity. Interestingly, contrary to the important increases of $G_\mathrm{A}$ and $G_\mathrm{NA}$, the total shear modulus $G = G_\mathrm{A} - G_\mathrm{NA}$ shows a low variation (by comparing $G_\mathrm{A} \simeq G_\mathrm{NA} \simeq 900$ with $G \simeq 24$ at $\epsilon_\text{bend}=10^4$). The bending rigidity increases the affine shear modulus but, at the same time, the non-affine component also increases to cancel the increase in $G_\mathrm{A}$, and as a result, the total shear modulus presents a low increase. The elasticity of the shear deformation is therefore different from that of the bulk deformation, which is obvious when the elastic moduli are decomposed into affine and non-affine components. Figure \[fig3\] also shows $K$ in (b) and $G$ in (d) for $L=3$. The values of $K$ and $G$ of $L=3$ are smaller than those of $L=50$, due to the bonding energy, $\varepsilon_\mathrm{FENE}$, connecting the monomers along the polymeric chains. The responses of $K$ and $G$ to the variation of $\varepsilon_\mathrm{bend}$ are also weaker for $L = 3$. However, $K$ and $G$, as well as affine $K_\mathrm{A}$ and $G_\mathrm{A}$ and non-affine $K_\mathrm{NA}$ and $G_\mathrm{NA}$, exhibit overall common dependences on $\varepsilon_\mathrm{bend}$ between $L=3$ and $50$. Therefore, the decrease in $\rho$ and increase in $\varepsilon_\mathrm{bend}$ engenders similar effects on the elasticity for $L=3$ and $50$. Finally, it is remarked that the polymer glasses present larger non-affine elastic components than the atomic (LJ) glasses [@Leonforte_2005; @Mizuno_2013]. Even under an isotropic bulk deformation, the non-affine $K_\mathrm{NA}$ ($\simeq 80$ for $L=50$ and $\simeq 50$ for $L=3$, at $\varepsilon_\mathrm{bend}\le 1$) is approximately half of the magnitude of the affine $K_\mathrm{A}$ ($\simeq 155$ for $L=50$ and $\simeq 120$ for $L=3$, at $\varepsilon_\mathrm{bend}\le 1$). This result is different from that of the LJ glasses, where a negligible value of $K_\mathrm{NA} \simeq 0.5$ (whereas $K_\mathrm{A} \simeq 60.2$) was obtained [@Mizuno_2013]. Larger non-affine moduli reflect various elastic responses due to the multiple degrees of conformations in polymeric chains. Therefore, the non-affine deformation process must be considered to characterize the elastic property of polymeric systems. ![\[fig4\] [Low-frequency vibrational spectra.]{} We plot the vDOS $g(\omega)$ divided by $\omega^2$, i.e., the reduced vDOS $g(\omega)/\omega^2$, with changing the strength of bending rigidity $\varepsilon_\mathrm{bend}$. The chain length is (a) $L=50$ and (b) $L=3$. The horizontal lines indicate the Debye level $A_\mathrm{D}$. The color of line indicates the value of bending rigidity $\varepsilon_\mathrm{bend}$ according to the color bar. Black lines present value of the LJ glass which is taken from Ref. [@Shimada:2018fp]. ](fig4.pdf){width="48.00000%"} ![\[fig5\] [Debye frequency and Debye level.]{} Plots of the Debye frequency $\omega_\mathrm{D}$ (circles) and the Debye level $A_\mathrm{D} = 3/\omega_\mathrm{D}^3$ (triangles) as functions of the strength of bending rigidity $\varepsilon_\mathrm{bend}$. The chain length is (a) $L=50$ and (b) $L=3$. The values of $\omega_\mathrm{D}$ and $A_\mathrm{D}$ are calculated from the elastic moduli of $K$ and $G$ that are presented in Fig. \[fig3\]. The arrows indicate values of atomic LJ glasses that are taken from Ref. [@Shimada:2018fp]. ](fig5.pdf){width="48.00000%"} ![\[fig6\] [Scaled vibrational spectra.]{} We present the data presented in Fig. \[fig4\], in the scaled form: we scale the reduced vDOS $g(\omega)/\omega^2$ and the frequency $\omega$ by the Debye level $A_\mathrm{D}$ and the Debye frequency $\omega_\mathrm{D}$. Here the values of $A_\mathrm{D}$ and $\omega_\mathrm{D}$ are presented in Fig. \[fig5\]. The chain length is (a) $L=50$ and (b) $L=3$. The color of line indicates the value of bending rigidity $\varepsilon_\mathrm{bend}$ according to the color bar. Black lines present value of the LJ glass which is taken from Ref. [@Shimada:2018fp]. ](fig6.pdf){width="48.00000%"} ![\[fig7\] [Localization nature of vibrational states.]{} Plots of participation ratio $P^k$ as a function of the scaled frequency $\omega/\omega_\mathrm{D}$, for several different bending rigidities of $\varepsilon_\mathrm{bend}$. The chain length is (a) $L=50$ and (b) $L=3$. The color of line indicates the value of bending rigidity $\varepsilon_\mathrm{bend}$ according to the color bar. Data are shown as the average values over bins in the frequency domain of $\left[ \omega - \Delta\omega/2, \omega + \Delta\omega/2\right]$ with $\Delta \omega \simeq 0.06$. The vertical line indicates the position of $\omega_\mathrm{BP}/\omega_\mathrm{D}$ averaged over the examined systems with varied $\varepsilon_\mathrm{bend}$. ](fig7.pdf){width="48.00000%"} Low-frequency vibrational spectra --------------------------------- ### Reduced vDOS Finally, the spectra of vibrational eigenmodes in polymer glasses are studied. The vibrational mode analysis is performed on the configuration of the polymeric system at $T=0$, which corresponds to the inherent structure [@Kittel_1996; @Ashcroft]. The Hessian matrix is diagonalized to obtain the eigenfrequencies $\omega^k$ that corresponds to the square root of the eigenvalues $\lambda^k$, i.e., $\omega^k = \sqrt{\lambda^k}$ ($k=1,2,...,3N_\mathrm{p}$). The specific expression of the Hessian matrix is given in Supplementary Material [^1]. The statistics of the eigenfrequency provide the vDOS, $g(\omega)$. Figure \[fig4\] presents the reduced version of the vDOS, $g(\omega)/\omega^2$, when changing the rigidity $\varepsilon_\mathrm{bend}$ and for $L=50$ in (a) and $L=3$ in (b). The reduced vDOS, $g(\omega)/\omega^2$, of the Debye theory is the so-called Debye level $A_\mathrm{D}$ [@Kittel_1996; @Ashcroft]. $A_\mathrm{D}$ is calculated from the elastic moduli, $K$ and $G$, as follows: $A_\mathrm{D} = 3/\omega_\mathrm{D}^3$, where $\omega_\mathrm{D}$ is the Debye frequency defined as $\omega_\mathrm{D} = \left[ 18 \pi^2 \rho/ ({2{c_\mathrm{T}}^{-3} + {c_\mathrm{L}}^{-3}}) \right]^{1/3}$, and $c_\mathrm{L} =\sqrt{(K+4G/3)/\rho}$ and $c_\mathrm{T} = \sqrt{G/\rho}$ are the longitudinal and transverse sound speeds, respectively. Figure \[fig5\] presents the values of $\omega_\mathrm{D}$ and $A_\mathrm{D}$ as functions of $\varepsilon_\mathrm{bend}$. As the bulk modulus is approximately four times larger than the shear modulus, $\omega_\mathrm{D}$ and $A_\mathrm{D}$ are mostly determined with the shear modulus, i.e, $\omega_\mathrm{D} \approx \left( 9 \pi^2 \rho \right)^{1/3} c_\mathrm{T}$ and $A_\mathrm{D} \approx 1/ \left(3 \pi^2 \rho c_\mathrm{T}^3 \right)$. As shown in Fig. \[fig4\], the polymer glasses present clear excess peaks over the Debye level, i.e., the BP. The BP frequency, $\omega_\mathrm{BP}$, is defined as the frequency at which $g(\omega)/\omega^2$ is maximal. As $\varepsilon_\mathrm{bend}$ increases, $\omega_\mathrm{BP}$ shifts to a higher frequency. In addition, the height of the reduced vDOS, $g(\omega_\mathrm{BP})/\omega_\mathrm{BP}^2$, becomes lower. These shifts are observed in the region from $\varepsilon_\mathrm{bend}=10$ to $10^3$ for $L=50$ and $3$. Importantly, this region corresponds to the shear modulus $G$ variations, as shown in Figs. \[fig3\]c and \[fig3\]d. As the bulk modulus is much larger than the shear modulus, the bulk modulus should only have minor effects on the low-frequency spectra. Therefore, the BP of the proposed system should only be controlled by the shear elasticity. To confirm this hypothesis, the scaled vDOS $g(\omega)/(\omega^2 A_\mathrm{D})$ is plotted as a function of the scaled frequency $\omega/\omega_\mathrm{D}$ and presented in Fig. \[fig6\]. As discussed above, $A_\mathrm{D}$ and $\omega_\mathrm{D}$ are determined mostly by the shear modulus $G$. Although deviations are observed for $L=50$, the scaled vDOSs collapse for different values of $\varepsilon_\mathrm{bend}$. In particular, an exact collapse is obtained for $L=3$. This result indicates that the effects engendered by the bending rigidity on the low-frequency spectra are comprised of the the shear modulus changes. A same collapse was observed in effects of pressure on the BP in the covalent-bonding network glass (Na$_2$FeSi$_3$O$_8$) [@Monaco_2006]. In addition, a previous experiment [@Caponi_2011] demonstrated that the effects of the polymerization are also comprised by the macroscopic elasticity changes. The collapsed results for (a) $L=50$ and (b) $L=3$ are consistent with the experimental observation. According to the collapses observed in Fig. \[fig6\], $\omega_\mathrm{BP} / \omega_\mathrm{D}$ does not depend on $\varepsilon_\mathrm{bend}$. As stated previously, when $\varepsilon_\mathrm{bend}$ varies, $\omega_\mathrm{D} \propto \rho^{1/3} c_\mathrm{T} \propto \rho^{-1/6} \sqrt{G}$. As $\rho$ varies in a range of 15%, as shown in Fig. \[fig1\], the effect of $\rho$ on $\omega_\mathrm{D}$ is weak. Thus, $\omega_\mathrm{D}$ is approximately proportional to $\sqrt{G}$, which leads to $\omega_\mathrm{BP} \propto \sqrt{G}$ in the variation of $\varepsilon_\mathrm{bend}$. The $\varepsilon_\mathrm{bend}$ dependence of the BP frequency is determined by the shear modulus, which is a macroscopic quantity describing the entire system in an averaged manner. It is noted that the recent study [@Baggioli_2019] predicts $\omega_\mathrm{BP} \propto \sqrt{G}$ from the phonon Green’s function with diffusive damping. It might be interesting to study effects of $\varepsilon_\mathrm{bend}$ on phonon transport and the phonon’s Green function. According to the heterogeneous elasticity theory [@Schirmacher_2006; @Schirmacher_2007; @Schirmacher_2015], the spatial fluctuations of the local shear modulus $\delta G$ control nature of the BP [^2]. The collapse of $g(\omega)/( \omega^2 A_\mathrm{D})$ as a function of $\omega/\omega_\mathrm{D}$ indicates that the shear modulus fluctuations relative to the macroscopic value, $\delta G/G$, are constant for all the cases of different bending rigidities. Therefore, the results of this study can be explained as follows. The increase in bending rigidity does not affect the shear modulus fluctuations (relative to the macroscopic moduli) but only affects the macroscopic shear modulus, which leads to the collapse of the scaled vDOS. ### Participation ratio To further study the vibrational eigenstates, the participation ratio $P^k$ that measures the extent of localization of the eigenmodes $k$ is calculated as follows [@Schober_1996; @Mazzacurati_1996]: $$P^k = \frac{1}{N_\mathrm{p}}\left[ \sum_{i=1}^{N_\mathrm{p}} (\boldsymbol{e}^k_i\cdot \boldsymbol{e}^k_i)^2\right]^{-1},$$ where $\boldsymbol{e}^k_i$ $(i=1, 2, \cdots , N_\mathrm{p})$ are the eigenvectors associated with the eigenfrequencies $\omega^k$ ($i$ is the index of the monomer particle and $N_\mathrm{p}$ is the number of monomer particles). The $\boldsymbol{e}^k_i$ represents the displacements of each monomer bead $i$ in the eigenmode $k$. It is noted that $\boldsymbol{e}^k_i$ is obtained from the diagonalization of the Hessian matrix and is orthonormalized as $\sum_{i=1}^{N_\mathrm{p}} \boldsymbol{e}^k_i\cdot \boldsymbol{e}^l_i = \delta_{kl}$ ($\delta_{kl}$ is the Kronecker delta). The following extreme cases can occur: ${P}^k = 2/3$ for an ideal sinusoidal plane wave, ${P}^k = 1$ for an ideal mode in which all constituent particles vibrate equally, and ${P}^k = 1/N_\mathrm{p} \ll 1$ for a perfect localization, which indicates that each vibrational state is associated only with a single atom and that $e^k_i \cdot e^k_i=1$ for a single $i$, otherwise $e^k_i \cdot e^k_i=0$. Figure \[fig7\] presents the value of $P^k$ as a function of the scaled frequency $\omega/\omega_\mathrm{D}$, for different $\varepsilon_\mathrm{bend}$. It is noted that the presented data are the binned average values. Below the BP frequency $\omega_\mathrm{BP}$, $P^k$ progressively decreases when $\omega$ decreases due to the spatially localized vibrations. The low-frequency localization below $\omega_\mathrm{BP}$ has also been observed in multiple glasses [@Schober_1996; @Mazzacurati_1996; @Taraskin_1997; @Taraskin_1999]. Importantly, $P^k$ below $\omega_\mathrm{BP}$ collapses between different values of $\varepsilon_\mathrm{bend}$. This result indicates that the variations of not only the vDOS and the vibrational states due to $\varepsilon_\mathrm{bend}$ can be characterized by the macroscopic shear modulus changes. However, $P^k$ does not collapse above $\omega_\mathrm{BP}$, as also shown in Fig. \[fig7\]. This result is attributed to the fact that the high-frequency modes above $\omega_\mathrm{BP}$ reflect microscopic vibrations that cannot be captured by the macroscopic elasticity. ### Comparison with LJ glasses The low-frequency spectra are comparable to that of atomic LJ glasses reported in Ref. [@Shimada:2018fp]. As observed in Fig. \[fig4\], the height of $g(\omega_\mathrm{BP})/\omega_\mathrm{BP}^2$ of LJ glasses is higher than that of polymer glasses, and $\omega_\mathrm{BP}$ is lower than that of polymer glasses. These observations are different from the study reported in Ref. [@Giuntoli:2018fv], which demonstrated that the low-frequency spectra of polymer glasses correspond to those atomic LJ glasses. In Ref. [@Giuntoli:2018fv], the bonded monomers interact via a harmonic potential with a large bonding energy scale of $k = 2500$. This value is two orders of magnitude larger than $\varepsilon_\mathrm{FENE} = 30$, investigated in this study. With respect to the large bonding energy, the rigidity of the polymeric chains has a smaller effect on the low-frequency spectra. Therefore, the low-frequency spectra are mainly determined by the non-bonding LJ interactions, whereas the elasticity is mainly determined mainly by the bonding rigidity. As a results, the BP decouples with the macroscopic elasticity, as demonstrated in the previous study [@Giuntoli:2018fv]. In contrast to the the results presented in Ref. [@Giuntoli:2018fv], the rigidity of the polymeric chains is necessary to determine the elasticity and the low-frequency spectra with respect to the bonding energy scale of $\varepsilon_\mathrm{FENE} = 30$. In fact, the $\varepsilon_\mathrm{bend}$ reduces the height of $g(\omega_\mathrm{BP})/\omega_\mathrm{BP}^2$, as shown in Fig \[fig4\]. In this case, the BP couples with the macroscopic elasticity. However, the plot of the scaled $g(\omega)/( \omega^2 A_\mathrm{D})$ as a function of$\omega/\omega_\mathrm{D}$ does not collapse between the polymer glasses and LJ glass, as shown in Fig. \[fig6\]. The height of $g(\omega)/ (\omega^2 A_\mathrm{D})$ is consistent between the polymer glasses and LJ glass, but $\omega/\omega_\mathrm{D}$ of the LJ glass is lower than that of the polymer glasses. This result indicates that vibrational states differences between polymer glasses and LJ glasses cannot be described only by changes in macroscopic elasticity, changes in the local elastic properties should be considered as well [@Monaco_2006; @Niss_2007; @Hong_2008; @Mizuno2_2013]. In addition, the length scale of collective vibrational modes in the BP region is discussed. For atomic LJ glasses, the length scale was evaluated as $\xi_\mathrm{BP}=2 \pi c_\mathrm{T} /\omega_\mathrm{BP}$, which corresponds to the size of approximately $23$ particle [@Leonforte_2005]. This length scale diverges near the isostatic point or the marginally stable point, theoretically [@Wyart_2005; @Wyart_2006; @Wyart_2010; @DeGiuli_2014] as well as numerically [@Silbert_2005; @Lerner_2014; @Karimi_2015; @Shimada_2018; @Mizuno_2018]. The present study evaluates the length scale of collective vibrational modes in polymeric glasses as $\xi_\mathrm{BP}=2 \pi c_\mathrm{T} /\omega_\mathrm{BP} \approx 12$, which corresponds to half of that for LJ glasses. The vibrational modes in the BP region are more localized nature due to the polymerization. Moreover, the value of $\xi_\mathrm{BP}$ is independent of the bending rigidity $\varepsilon_\mathrm{bend}$ because of $\omega_\mathrm{BP} \propto \omega_\mathrm{D} \propto c_\mathrm{T}$. In other words, the bending rigidity does not affect the length scale of the collective vibrational motions in the BP region. Discussion ========== The glass transition temperature, elastic properties, and the low-frequency vibrational spectra were studied in polymeric glasses. In particular, the bending energy scale was highly varied for long chains ($L=50$) and short chains ($L=3$). As the system becomes rigid by increasing the bending rigidity, the glass transition occurs at a higher temperature, leading to a lower density in the glass phase. The lowering density directly affects the isotropic bulk deformation, but does not affect the shear elasticity. The shear elasticity is controlled by only the bending rigidity only. The non-affinity of polymeric glasses is much larger than that of atomic LJ glasses. This is due to the more complex conformational relaxations of the polymeric chains during non-affine deformation. Even under an isotropic elastic deformation, the non-affine relaxation process should be considered to describe the elastic response. In addition, it is demonstrated that the BP frequency and its intensity are simply scaled by the Debye frequency and the Debye level which are mainly determined by the macroscopic shear modulus. This result indicates that the BP is controlled by macroscopic shear modulus and that the bending rigidity has a small impact on heterogeneities of local elasticity properties. The effects of the bending rigidity on the BP is similar to that of the polymerization, which has also been explained by macroscopic elasticity changes [@Caponi_2011]. The presented results provide a simple relationship between the BP and the elasticity as well as the glass transition temperature. As the system becomes more rigid by increasing the bending rigidity, the glass transition temperature and the shear modulus are increased. On the contrary, the bulk modulus $K$ decreases due to the decrease in the density $\rho$ caused by the increase in the glass transition temperature $T_g$. However, the BP is mainly determined by the shear modulus $G$: $\omega_\mathrm{BP} \propto \omega_\mathrm{D} \propto \sqrt{G}$. Therefore, the glass transition temperature, the shear elasticity, and the boson peak frequency are positively correlated. A similar relationship between $T_g$ and $\omega_\mathrm{BP}$ was observed experimentally in ionic liquids systems [@Kofu_2015] and also numerically in LJ glasses [@Wang_2015]. It is noted that the studies of Refs. [@Kofu_2015; @Wang_2015] provided the relationship of $T_g \propto \omega_\mathrm{BP}^2$, but a clear power-law like relationship between $T_g$ and $\omega_\mathrm{BP}$ was not observed in polymeric glasses. Finally, it is worthwhile to discuss the structural relaxation in the liquid state above the glass transition temperature. A previous study [@Larini_2008] has demonstrated the scaling relationship between the structural relaxation time $\tau_\alpha$ and the Debye-Waller factor $\langle u^2\rangle $ as $\tau_\alpha \propto \exp\left( a \langle u^2\rangle ^{-1} + b \langle u^2\rangle ^{-2} \right)$ (where $a, b$ are constants) for multiple glass-forming liquids including polymeric glasses. Here, the Debye-Waller factor in the harmonic approximation [@Shiba_2016] is estimated as $\langle u^2\rangle = 3T \int_0^\infty {g(\omega)}/{\omega^2} d\omega \propto T \omega_\mathrm{BP}^{-2} \propto T G^{-1}$. It is naturally expected that the relaxation dynamics become drastically slow by increasing the bending rigidity because of the following relationship: $$\tau_\alpha \propto \exp\left( \alpha \frac{\omega_\mathrm{BP}^{2}}{T} + \beta \frac{\omega_\mathrm{BP}^{4}}{T^2} \right) \propto \exp\left( \alpha' \frac{G}{T} + \beta' \frac{G^{2}}{T^2} \right),$$ where $\alpha, \beta, \alpha', \beta'$ are constants. This simple relationship demonstrates that the BP below $T_g$ and the structural relaxation above $T_g$ are well correlated in the polymeric glasses with varying the bending rigidity. Further work is necessary to evaluate its validity by calculating $\tau_\alpha$. The authors thank Atsushi Ikeda for useful discussions and suggestions. This work was supported by JSPS KAKENHI Grant Numbers: JP19K14670 (H.M.), JP17K14318 (T.M.), JP18H04476 (T.M.), JP18H01188 (K.K.), JP15K13550 (N.M.), and JP19H04206 (N.M.). This work was also partially supported by the Asahi Glass Foundation and by the Post-K Supercomputing Project and the Elements Strategy Initiative for Catalysts and Batteries from the Ministry of Education, Culture, Sports, Science, and Technology. The numerical calculations were performed at Research Center of Computational Science, Okazaki Research Facilities, National Institutes of Natural Sciences, Japan. [81]{}ifxundefined \[1\][ ifx[\#1]{} ]{}ifnum \[1\][ \#1firstoftwo secondoftwo ]{}ifx \[1\][ \#1firstoftwo secondoftwo ]{}““\#1””@noop \[0\][secondoftwo]{}sanitize@url \[0\][‘\ 12‘\$12 ‘&12‘\#12‘12‘\_12‘%12]{}@startlink\[1\]@endlink\[0\]@bib@innerbibempty , ed., @noop [**]{}, , Vol.  (, , ) @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****, ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****, ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****, ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [**]{},  ed. (, ) @noop [**]{} (, ) @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} \ Formalism of the Hessian Matrix =============================== The Hessian matrix of the interaction potential $U(\bm{r})$ is generally expressed as follows: $$\label{general_hessian} H_{nm}^{ab} = \frac {\partial^2 U\left(\bm{r}\right)} {\partial r_n^a \partial r_m^b} \quad (a, b=x, y, x)$$ where $n$ and $m$ denote the particle number index ($n$, $m$=1, 2, $\cdots$, $N_\mathrm{p}$). As given in Ref. [@Milkus:2018ha], the following expressions are useful using a generic argument $z$ for the first and second derivatives of $U(z)$: $$\label{for_a_generic_argument_1} \frac {\partial U\left(z\right)} {\partial x} = \frac {\partial U\left(z\right)} {\partial z} \frac {\partial z} {\partial x},$$ $$\begin{aligned} \label{for_a_generic_argument_2} \frac {\partial U^2\left(z\right)} {\partial x\partial y} = \frac {\partial U^2\left(z\right)} {\partial^2z} \frac {\partial z} {\partial x} \frac {\partial z} {\partial y} + \frac {\partial U\left(z\right)} {\partial z} \frac {\partial^2 z} {\partial x\partial y} = c\frac {\partial z} {\partial x} \frac {\partial z} {\partial y} + t\frac {\partial^2 z} {\partial x\partial y}.\end{aligned}$$ Two-body interaction -------------------- For two-body interactions (FENE and LJ potentials), the distance between particles $i$ and $j$, $z = | \bm{r}_j - \bm{r}_i | = r_{ij}$ is used and the following relationships are obtained: $$H_{nm}^{ab} = \frac {\partial^2 U\left(r_{ij}\right)} {\partial r_n^a \partial r_m^b} = c_{ij}\frac {\partial r_{ij}} {\partial r_n^a} \frac {\partial r_{ij}} {\partial r_m^b} + t_{ij} \frac {\partial^2r_{ij}} {\partial r_n^a\partial r_m^b},$$ with $$c_{ij} = \frac {\partial^2U\left(r_{ij}\right)} {\partial r_{ij}^2},\quad t_{ij} = \frac {\partial U\left(r_{ij}\right)} {\partial r_{ij}},$$ and $$\frac{\partial r_{ij}} {\partial r_n^a} = \left(\delta_{nj}-\delta_{ni}\right)\hat n_{ij}^a,$$ $$\frac{\partial^2 r_{ij}} {\partial r_n^a\partial r_m^b} = \frac{1}{r_{ij}} \left( \delta_{nj}-\delta_{ni} \right) \left( \delta_{mj}-\delta_{mi} \right) \left( \delta_{ab}-\hat{n}_{ij}^a\hat{n}_{ij}^b \right), \label{eq1:Milkus}$$ where, $\hat n_{ij} = \bm{r}_{ij}/r_{ij}$ is the unit vector between the particles $i$ and $j$. These expressions are same as those presented in Ref. [@Milkus:2018ha]. Three-body interaction ---------------------- For three-body interactions (bending potential), the bond angle of particles $i$, $j$, and $k$ is used as follows: $$z = \theta_{ijk} = \arccos\frac {\left( \bm{r}_j - \bm{r}_i \right) \cdot \left( \bm{r}_k - \bm{r}_i \right)} {r_{ij}r_{ki}} = \arccos A_{ijk},$$ hence, $$H_{nm}^{ab} = \frac {\partial^2 U\left(\theta_{ijk}\right)} {\partial r_n^a\partial r_m^b} = \tilde{c}_{ijk}\frac {\partial \theta_{ijk}} {\partial r_n^a} \frac {\partial \theta_{ijk}} {\partial r_m^b} + \tilde{t}_{ijk} \frac {\partial^2\theta_{ijk}} {\partial r_n^a\partial r_m^b}$$ with $$\tilde{c}_{ijk} = \frac {\partial^2U\left(\theta_{ijk}\right)} {\partial \theta_{ijk}^2}, \quad \tilde{t}_{ijk} = \frac {\partial U\left(\theta_{ijk}\right)} {\partial \theta_{ijk}}.$$ This following expression is obtained: $$H_{nm}^{ab} = \frac {\tilde{c}_{ijk}} {\sin^2 \theta_{ijk}} \frac {\partial A_{ijk}} {\partial r_n^a} \frac {\partial A_{ijk}} {\partial r_m^b} - \frac {\tilde{t}_{ijk}} {\sin\theta_{ijk}} \left[ \frac {\cos\theta_{ijk}} {\sin^2 \theta_{ijk}} \frac {\partial A_{ijk}} {\partial r_n^a} \frac {\partial A_{ijk}} {\partial r_m^b} + \frac {\partial^2 A_{ijk}} {\partial r_n^a\partial r_m^b} \right], \label{eq2:Milkus}$$ with $$\frac {\partial A_{ijk}} {\partial r_n^a} = \frac{1}{r_{ij}} \left( \delta_{nj}-\delta_{ni} \right) \left( \hat{n}_{ik}^a - \hat{n}_{ij}^a\cos\theta_{ijk} \right) + \frac{1}{r_{ik}} \left( \delta_{nk}-\delta_{ni} \right) \left( \hat{n}_{ij}^a - \hat{n}_{ik}^a\cos\theta_{ijk} \right), \nonumber$$ $$\begin{aligned} \frac {\partial^2 A_{ijk}} {\partial r_n^a\partial r_m^b} &= \frac {\delta_{ji}^n\delta_{ji}^m}{r_{ij}^2} \left[ \left( 3\hat{n}_{ij}^a\hat{n}_{ij}^b - \delta_{ab} \right) \cos\theta_{ijk} - \left( \hat{n}_{ik}^a\hat{n}_{ij}^b + \hat{n}_{ij}^a\hat{n}_{ik}^b \right) \right] &\nonumber\\ & \quad+ \frac {\delta_{ji}^n\delta_{ki}^m}{r_{ij}r_{ik}} \left[ \delta_{ab} + \hat{n}_{ij}^a\hat{n}_{ik}^b\cos\theta_{ijk} - \left( \hat{n}_{ik}^a\hat{n}_{ik}^b + \hat{n}_{ij}^a\hat{n}_{ij}^b \right) \right] &\nonumber\\ & \quad+ \frac {\delta_{ki}^n\delta_{ji}^m}{r_{ij}r_{ik}} \left[ \delta_{ab} + \hat{n}_{ik}^a\hat{n}_{ij}^b\cos\theta_{ijk} - \left( \hat{n}_{ik}^a\hat{n}_{ik}^b + \hat{n}_{ij}^a\hat{n}_{ij}^b \right) \right] &\nonumber\\ & \quad+ \frac {\delta_{ki}^n\delta_{ki}^m}{r_{ik}^2} \left[ \left( 3\hat{n_{ik}^a}\hat{n_{ik}^b} - \delta_{ab} \right) \cos\theta_{ijk} - \left( \hat{n_{ij}^a}\hat{n_{ik}^b} + \hat{n_{ik}^a}\hat{n_{ij}^b} \right) \right].\end{aligned}$$ The differences between the proposed calculation and the expression defined in Ref. [@Milkus:2018ha] arise from Eq. (\[eq1:Milkus\]) and the second term in the r.h.s. of Eq. (\[eq2:Milkus\]). The overall profile of the vDOS $G(\omega)$ is not affected by implementing the diagonalization of the Hessian matrix using the expressions in Ref. [@Milkus:2018ha]. A certain number of negative frequency eigenmodes that have been reported in Ref. [@Milkus:2018ha] have also been observed. On the contrary, the presented results of $g(\omega)$ using Eqs. (\[eq1:Milkus\]) and (\[eq2:Milkus\]) do not exhibit any negative eigenfrequency modes (see Fig. 4 in the main text). [1]{}ifxundefined \[1\][ ifx[\#1]{} ]{}ifnum \[1\][ \#1firstoftwo secondoftwo ]{}ifx \[1\][ \#1firstoftwo secondoftwo ]{}““\#1””@noop \[0\][secondoftwo]{}sanitize@url \[0\][‘\ 12‘\$12 ‘&12‘\#12‘12‘\_12‘%12]{}@startlink\[1\]@endlink\[0\]@bib@innerbibempty @noop [****,  ()]{} [^1]: The expression of the Hessian matrix is already described in Ref. [@Milkus:2018ha]. However, the expression includes errors. Therefore, the corrected expression is provided in the Supplementary Material. [^2]: The value of $\delta G$ is quantified by the standard deviation of probability distribution function of the local shear modulus [@Mizuno_2013].
{ "pile_set_name": "ArXiv" }
--- abstract: | Two-stage randomization is a powerful design for estimating treatment effects in the presence of interference; that is, when one individual’s treatment assignment affects another individual’s outcomes. Our motivating example is a two-stage randomized trial evaluating an intervention to reduce student absenteeism in the School District of Philadelphia. In that experiment, households with multiple students were first assigned to treatment or control; then, in treated households, one student was randomly assigned to treatment. Using this example, we highlight key considerations for analyzing two-stage experiments in practice. Our first contribution is to address additional complexities that arise when household sizes vary; in this case, researchers must decide between assigning equal weight to households or equal weight to individuals. We propose unbiased estimators for a broad class of individual- and household-weighted estimands, with corresponding theoretical and estimated variances. Our second contribution is to connect two common approaches for analyzing two-stage designs: linear regression and randomization inference. We show that, with suitably chosen standard errors, these two approaches yield identical point and variance estimates, which is somewhat surprising given the complex randomization scheme. Finally, we explore options for incorporating covariates to improve precision. We confirm our analytic results via simulation studies and apply these methods to the attendance study, finding substantively meaningful spillover effects. [**Key Words**]{}: two-stage randomization; randomization inference; causal inference under interference; student attendance. author: - | Guillaume Basse\ Harvard - | Avi Feller\ UC Berkeley bibliography: - 'ref.bib' title: | **Analyzing two-stage experiments\ in the presence of interference[^1]** --- Introduction {#section:intro} ============ A common assumption in causal inference is “no interference” between units: one individual’s outcomes are unaffected by another individual’s treatment assignment [@cox1958planning; @rubin1980comment]. However, this assumption does not hold in settings ranging from infectious diseases to education to labor markets [for a recent review, see, @Halloran:2016dc]. In many of these cases, such interference is of direct substantive interest. Two-stage randomization is a powerful design for estimating causal effects involving interference. In the setting we consider, first whole clusters (e.g., households, schools, or graph partitions) are assigned to treatment or control. Second, units within each treated cluster are randomly assigned to treatment or control, as if each treated cluster were a separate, individually-randomized experiment. This design allows researchers to assess spillover effects either by comparing untreated units in treated clusters with pure control units in control clusters or by comparing units across clusters with different proportions assigned to treatment [@hudgens2012toward]. Our motivating example is a large randomized evaluation of an intervention targeting student absenteeism among elementary and high school students in the School District of Philadelphia [@Rogers_Feller_SDP]. In the original study, parents of at-risk students were randomly assigned to a direct mail intervention with tailored information about their students’ attendance over the course of the year. In treated households with multiple eligible students, one student was selected at random to be the subject of the mailings, following a two-stage randomization. Substantively, this is a rare opportunity to study intra-household dynamics around student behavior. Methodologically, this presents a rich test case for understanding how to analyze two-stage experiments in practice. There has been substantial interest in two-stage randomization in recent years, with prominent examples in economics [@crepon2013labor], education [@somers2010enhanced], political science [@sinclair2012detecting], and public health [@hudgens2012toward], as well as closely related variants in the context of large-scale social networks [@ugander2013graph]. Such designs have become especially common in development economics [@angelucci2016programme]. There is also a small but growing methodological literature on analyzing two-stage experiments, including @hudgens2012toward, @liu2014large, and @Rigdon:2015cv in statistics; and @sinclair2012detecting and @baird2014designing in the social sciences. We build on this literature by addressing three practical issues that arise in analyzing the attendance study. First, school districts are typically interested in the intervention’s impact on students rather than on households; that is, districts give equal weight to each individual rather than equal weight to each household. Similarly, public health researchers administering treatment to villages of different sizes might be interested in the impact on the overall population rather than on village-level averages, especially if the treatment is more effective in larger villages. With the exception of @sinclair2012detecting, however, existing approaches focus either on equal weights for households [e.g., @hudgens2012toward] or side-step the issue by assuming households are of equal size [e.g., @baird2014designing]. We propose unbiased estimators for a broad class of individual- and household-weighted estimands, with corresponding theoretical and estimated variances. We also derive the bias of a simple difference in means for estimating individual-weighted estimands. Since researchers typically estimate these two estimands with different precision [see, e.g., @Athey:2016wn], we recommend that researchers report both in practice. Second, we connect two common approaches for analyzing two-stage designs: linear regression, which is more common in the social sciences, and randomization inference, which is more common in epidemiology and public health. We show that, with suitably chosen standard errors, regression and randomization inference yield identical point and variance estimates. These results hold for a broad class of weighted estimands. We believe this equivalence will be important in practice, since the vast majority of applied papers in this area take a “regression first” approach to analysis that can obfuscate key inferential issues. Lastly, we explore options for incorporating covariates to improve precision, with a focus on post-stratification and model-assisted estimation. We then confirm our analytic results via simulation studies and apply these methods to the attendance study. Overall, we find strong evidence of a spillover effect that is (depending on the scale of the outcome) roughly 60 to 80 percent as large as the primary effect. This holds across different estimands as well as with and without covariate adjustment. Accounting for spillovers therefore dramatically improves the cost effectiveness of the intervention, from around \$5 per additional day to around \$3 per additional day. This paper proceeds as follows. Section \[section:setup\] defines the two-stage randomization, sets up the notation, and discusses the relevant assumptions. Section \[section:estimands\] defines the estimands of interest both for constant and varying household sizes. Sections \[section:estimation\] and \[section:variance\] deal with unbiased estimators and their variance. Section \[sect:connection-with-regression\] demonstrates how we can use regression with appropriate standard errors to obtain the randomization-based estimators. Section \[section:covariate-adjustment\] explores covariate adjustment. Section \[section:simulations\] reports the results of extensive simulation studies. Section \[section:data\] analyzes the student attendance experiment. Section \[section:discussion\] concludes and offers directions for future work. The supplementary materials contains additional technical material and all proofs. Motivating example: Student absenteeism --------------------------------------- More than 10 percent of public school students in the United States—over five million students—are chronically absent each year, defined as missing 18 or more days of the roughly 180-day school year [@crdc2016]. @Rogers_Feller_SDP recently conducted the first randomized evaluation of an intervention aimed at reducing student absenteeism for this population. This intervention delivered targeted information to parents of at-risk students in the School District of Philadelphia via five pieces of direct mail over the 2014–2015 School Year. The mailing clearly stated the student’s number of absences that year (“Your student has been absent 16 days this school year”), included a simple bar chart showing the same information graphically, and gave additional text on the importance of attending school.  @Rogers_Feller_SDP find that the treatment reduces chronic absenteeism by over 10 percent relative to control. The approach is extremely cost-effective, costing around \$5 per additional day of student attendance—more than an order of magnitude more cost-effective than the current best-practice intervention. A key practical challenge in implementing the original study was that the mailings were designed to provide information about a single student. Students were eligible to be the target of the intervention if they met certain pre-specified criteria, including type of school, home language, and no perfect attendance in the previous year. In households with multiple eligible students, one student was randomly selected to be the focal student. (The study excludes other members of the household, such as non-eligible siblings.) @Rogers_Feller_SDP addressed possible spillover only briefly in the original study, largely because the focus was on the primary effect of the intervention and because households with multiple students were around 15 percent of all households in the sample. In this paper, we consider a subset of $N = 3,876$ households with between $n_i = 2$ and $n_i = 7$ eligible students in each household and $n^+ = 8,654$ total students. Table \[tbl:num\_students\_hh\] shows the distribution of household size. The vast majority of these households (82 percent) have only two students; only one percent (35 households) have five or more students. Our goal is to estimate the primary and spillover effects on attendance for this finite sample. We are also interested in the extent to which these estimates differ depending on whether we give equal weight to each household or to each individual. The original study estimated these effects with a simple difference-in-means estimator, which could be biased in practice; see Section \[section:sd-bias\]. 2 3 4–7 ---------------------------------- ------- ------ ------ Total $N$ 3,169 557 150 Proportion assigned to treatment 0.66 0.65 0.65 : Number of households by size, and proportion of treated households for each size.[]{data-label="tbl:num_students_hh"} This experimental design presents a rare opportunity to assess intra-household spillovers. There is substantial evidence across fields that such intra-household spillovers are meaningful in magnitude. For example, several voter mobilization studies have found spillover effects that are between one-third and two-thirds as large as the primary effect [@nickerson2008voting; @sinclair2012detecting]. We are interested in spillover in the attendance study for two key reasons. First, ignoring the spillover effect under-states the overall impact of the intervention. For example, an important metric is the cost of each additional student day; ignoring spillover artificially lowers the corresponding cost-effectiveness estimates. Second, the research team faced a practical question of whether to develop a separate intervention for households with multiple eligible students, which would be costly to implement and test. If the spillover effect is comparable in magnitude to the primary effect, such development is unnecessary. This is similar to decisions around interventions targeting infectious diseases [@hudgens2012toward] and in economics [@baird2014designing]. Setup and assumptions\[section:setup\] ====================================== We now review the setup and assumptions for a two-stage experiment in the presence of interference. The discussion closely follows @hudgens2012toward, modifying their terminology slightly to better fit our applied example and to recognize some small differences in emphasis in the social science literature. We first define potential outcomes and state the relevant assumptions, then follow with a description of two-stage randomized designs. We postpone the formal introduction of our estimands to Section \[section:estimands\]. For additional reviews on causal inference under interference, see, among others, @sinclair2012detecting [@bowers2013reasoning; @VanderWeele:2014gr; @Halloran:2016dc; @aronow2012estimating; @Athey:2016wn]. Potential outcomes and relaxing SUTVA {#section:assumptions} ------------------------------------- We use the potential outcomes framework to describe the problem [@neyman::1923; @rubin1974estimating]. Consider N households $i = 1, \ldots, N$ with $n_i$ individuals in household $i$, and where $n^+ \equiv \sum n_i$ is the total number of individuals. To be consistent with the existing literature on two-stage experiments, we use the double-index notation, such that $\cdot_{ij}$ denotes the individual $j$ in household $i$. For household $i$, let ${\pmb{Z}}_i = (Z_{i1}, \ldots, Z_{in_i})$ denote the assigment vector for the $n_i$ units in that household, where $Z_{ij} = 1$ if the $j^{th}$ individual in household $i$ is assigned to treatment, and $Z_{ij} = 0$ otherwise. Similarly, define ${\pmb{Z}}_{i, -j}$ as the sub-vector of ${\pmb{Z}}_i$ that excludes the $j^{th}$ value. Finally, aggregate all household-level assignments via ${\pmb{Z}}= \{{\pmb{Z}}_1, \ldots, {\pmb{Z}}_N\}$. Let $Y_{ij}^{\text{obs}}$ denote the observed outcome for individual $j$ in household $i$, which will be either binary or continuous in our motivating example. In general, let ${\pmb{Y}}_i({\pmb{Z}}) = ({\pmb{Y}}_{i1}({\pmb{Z}})$, $\ldots$, ${\pmb{Y}}_{in_i} ({\pmb{Z}}))$ be the vector of potential outcomes for household $i$, and ${\pmb{Y}}({\pmb{Z}}) = \{ {\pmb{Y}}_1({\pmb{Z}})$, $\ldots$, ${\pmb{Y}}_N({\pmb{Z}})\}$ be the list of potential outcome vectors for all households. At this stage, practical inference is infeasible without imposing additional restrictions on the structure of potential outcomes. However, the standard Stable Unit Treatment Value Assumption [SUTVA; @rubin1980comment], which implies that there is no interference between units, is inappropriate in our context. We instead focus on putting structure on two types of interference: between-household and within-household interference. ### Between-household interference First, we assume that there is no between-household interference, which @sobel2006randomized refers to as *partial interference*. \[asst:partial-interference\] Interference occurs only within a household. That is, ${\pmb{Y}}_i({\pmb{Z}}) = {\pmb{Y}}_i({\pmb{Z}}_i)$. This is effectively a “between household SUTVA” assumption and greatly reduces the complexity of the problem. In the context of the attendance study, this assumption states that students in different households do not affect each others’ attendance. This assumption is violated if, for instance, friends skip school together. Nonetheless, we view spillovers within households as far more important than spillovers between households and thus consider Assumption \[asst:partial-interference\] to be a useful approximation. ### Within-household interference Even with the partial interference assumption, practical inference remains challenging. The key complication is that, without additional restrictions, the potential outcomes depend on the identity of the treated individual. To see this, consider household $i$ with three students in which only one student is assigned to treatment. Under partial interference, the oldest student, $j = 1$, has three potential outcomes, $Y_{i1}(1,0,0)$, $Y_{i1}(0,1,0)$, and $Y_{i1}(0,0,1)$, which correspond to assigning the oldest, middle, and youngest student to treatment, respectively, as well as $Y_{i1}(0,0,0)$ if none receive treatment. Thus, the oldest student actually has *two* different spillover effects, $Y_{i1}(0,1,0) - Y_{i1}(0,0,0)$ and $Y_{i1}(0,0,1) - Y_{i1}(0,0,0)$, depending on which other student in the household receives the treatment. As @hudgens2012toward argue, this makes inference difficult, especially with respect to variance calculations [see also @tchetgen2012estimation]. Instead, @hudgens2012toward propose the *stratified interference* assumption, which states that the precise identity of the treated individual in the treated cluster does not matter for untreated individuals in the same cluster [see also @manski2013public]. \[asst:stratified-interference\] $$Y_{ij}({\pmb{Z}}_{i,-j},Z_{ij}=0)=Y_{ij}({\pmb{Z}}'_{i,-j},Z_{ij}=0) \,\,\,\,\, \forall {\pmb{Z}}_{i,-j}, {\pmb{Z}}'_{i,-j} \,\,\,\, \mbox{s.t.}\,\,\, \sum_j^{n_i} Z_{ij}=\sum_j^{n_i} Z'_{ij}=1$$ Heuristically, this imposes additional structure on the problem by assuming that potential outcomes are only a function of the number (or, depending on context, proportion) of individuals assigned to treatment within each household. Assignment mechanism and observed outcomes ------------------------------------------ Two-stage randomization is a special case of a multi-stage, nested randomization  that is used to assign treatment to units in a nested structure (throughout, we will refer to individuals nested within household). Figure \[fig:multilevel\_schematic\] highlights the sequential nature of the two-stage design we consider. Specifically, we consider designs in which each stage follows complete randomization; that is, randomizations in which a fixed number of units are assigned to treatment at each stage [@imbens2015causal]. See also @liu2014large, who refer to this as a permutation randomization, and @tchetgen2012estimation, who contrast completely randomized and Bernoulli designs in the second stage. Formally, let $\pmb{H} = (H_1, \ldots, H_N)$ be the vector of treatment assignments at the household level, such that $H_i = 1$ if household $i$ is assigned to treatment and $H_i = 0$ otherwise. For the first stage of randomization, we assume that a fixed integer of households $N_1 \in \{1, \ldots, N-1\}$ are assigned to treatment, with $P(\pmb{H}) = 1/{N \choose N_1}$ for all $\pmb{H}$ such that $\sum_i^N H_i = N_1$ and $P(\pmb{H}) = 0$ otherwise. Analogously define $N_0 = N - N_1$ as the number of households assigned to control. For the second stage, individuals are assigned to treatment conditional on the realized value of the first stage of randomization. For individuals in households with $H_i = 1$, we assume that exactly one individual is assigned to treatment, with $P({\pmb{Z}}_i \mid H_i = 1) = \frac{1}{n_i}$ if $\sum_j^{n_i} Z_{ij} = 1$ and $P({\pmb{Z}}_i \mid H_i = 1) = 0$ otherwise. For individuals in households with $H_i = 0$, $Z_{ij} = 0$ for all $j$. Thus, all individuals assigned to treatment are in households assigned to treatment. Given the assumptions in Section \[section:assumptions\] and the type of assignments allowed by the two-stage randomization mechanism we consider, the potential outcomes for individual $j$ in household $i$ simplify to $Y_{ij}({\pmb{Z}}) = Y_{ij}(H_i = h, Z_{ij}=z)$. There are three possible combinations: $Y_{ij}(1,1)$, $Y_{ij}(1,0)$, and $Y_{ij}(0,0)$. We regard these potential outcomes as fixed and define the observed outcome as a deterministic function of the treatment assignment and potential outcomes: $$Y_{ij}^{{\text{obs}}} = H_i Z_{ij} Y_{ij}(1,1) + H_i(1 - Z_{ij})Y_{ij}(1,0) + (1 - H_i)Y_{ij}(0,0),$$ where the randomness is entirely due to $\pmb{H}$ and ${\pmb{Z}}$. That is, unless otherwise stated, all expectations and variances are with respect to the randomization distribution; inference is fully justified by the randomization itself [@fisher1960design]. Finally, we introduce the sets ${\mathcal{T}}_{hz} = \{(i,j) : H_i = h \,\, \mbox{ and } \,\, Z_{ij} = z\}$ to denote the set of households and individuals who are assigned to $H_i = h$ and $Z_{ij} = z$. Estimands {#section:estimands} ========= We next discuss estimands of interest, closely following @hudgens2012toward. We start with the setting in which all households or clusters are of equal size and then turn to households of varying size. \[section:estimands-equal\] Constant household size ----------------------- ### Primary and Spillover Effects We first assume that all households are of the same size. That is, $n_i = n$ for all households $i = 1, \ldots, N$. Next, we define average potential outcomes at the household level, $\overline{Y}_i(h,z) = \frac{1}{n}\sum_j^n Y_{ij}(h,z)$, which average across all individuals $j = 1, \ldots, n$ within each household. Since all households are the same size, we can analogously define the average potential outcomes for the sample, $\overline{Y}(h,z) = \frac{1}{N}\sum_i^N \overline{Y}_{i}(h,z)$, which is the average household-level potential outcome across all households. Estimands are contrasts between these sample average potential outcomes. Unless otherwise stated, all estimands we consider here are finite sample estimands; that is, they are defined for the units in our sample. \[def:equal-size-fp-effects\] Define the average [*primary effect*]{} as follows: $$\tau^P = \frac{1}{Nn} \sum_{i}^N\sum_j^n (Y_{ij}(1,1) - Y_{ij}(0,0)) = \frac{1}{N}\sum_i^N (\overline{Y}_{i}(1,1) - \overline{Y}_{i}(0,0)) = \overline{Y}(1,1) - \overline{Y}(0,0),$$ and the average [*spillover effect*]{} as: $$\tau^S = \frac{1}{Nn} \sum_{i}^N\sum_j^n (Y_{ij}(1,0) - Y_{ij}(0,0)) = \frac{1}{N}\sum_i^N (\overline{Y}_{i}(1,0) - \overline{Y}_{i}(0,0)) = \overline{Y}(1,0) - \overline{Y}(0,0).$$ These estimands have various names in the literature. We take the terminology *primary effect* from @toulis2013estimation, but @baird2014designing use the term *treatment on the uniquely treated* for an analogous quantity. @hudgens2012toward call these estimands the *total* and *indirect* effects, respectively. ### Other Estimands\[section:overall-effect\] @hudgens2012toward propose two additional estimands, the *direct* and *overall* effects. The direct effect is essentially the impact of the second stage, within-household randomization, which is $\tau^D = \overline{Y}(1,1) - \overline{Y}(1,0)$ with equal-sized households. Following the effect decomposition in @hudgens2012toward, the direct effect can be defined as the difference between the primary and spillover effects. Since the spillover effect is a more natural estimand in our setting, we do not discuss the direct effect further. The overall effect is the impact of household-level random assignment. From a policy perspective, this quantity is of obvious interest to decision makers, providing a single number for the intervention’s impact. From a statistical perspective, however, the overall effect has the somewhat awkward feature that it is only defined with respect to a given assignment mechanism: $$\tau^O = \frac{1}{Nn} \sum_{i}^N\sum_j^n \left(E\left[Y_{ij}(1,Z_{ij})\right] - Y_{ij}(0,0)\right),$$ where $E\left[Y_{ij}(1,Z_{ij})\right]$ is the expected potential outcome for individual $j$ in household $i$: $$E\left[Y_{ij}(1,Z_{ij})\right] = P(Z_{ij} = 1 \mid H_i = 1)~Y_{ij}(1,1) + P(Z_{ij} = 0 \mid H_i = 1)~Y_{ij}(1,0).$$ Similar to @VanderWeele:2011fd, we can therefore re-write the overall effect as a weighted average of the primary and spillover effects: $$\tau^O = P(Z_{ij} = 1 \mid H_i = 1)~\tau^P + P(Z_{ij} = 0 \mid H_i = 1)~\tau^S,$$ with weights equal to the second stage treatment probability. As a result, we focus on the primary and spillover effects throughout the main text and defer corresponding results for the overall effect to the supplementary materials. Varying household size {#sec:overall-weights} ---------------------- We now generalize these results to allow for varying household size. Broadly, there are now two types of estimands, *household-weighted* estimands (‘HW’) that assign equal weight to households, regardless of the number of individuals in each household; and *individual-weighted estimands* (‘IW’) that assign equal weight to individuals, regardless of the distribution across households. A substantial literature on cluster-randomized trials addresses related questions; see, among others, @donner2000design [@imai2009essential; @schochet2013estimators; @aronow_middleton_cluster]. Specifically, we generalize the equal-sized household estimands to allow for *two-stage weights*. \[def:unequal-size-fp-effects\] Define the average [*primary effect*]{} as follows: $$\tau^P_W = \sum_{i=1}^N w_i^\ast \sum_{j=1}^{n_i} (Y_{ij}(1,1) - Y_{ij}(0,0)),$$ and the average [*spillover effect*]{} as: $$\tau^S_W = \sum_{i=1}^N w_i^\ast \sum_{j=1}^{n_i} (Y_{ij}(1,0) - Y_{ij}(0,0)),$$ where $w_i^\ast = \frac{1}{Nn_i}$ corresponds to household-weighted estimands and $w_i^\ast = \frac{1}{n^+}$ corresponds to individual-weighted estimands. When $n_i = n$ for all $i$, both HW and IW estimands are identical to the equal-sized household estimands. When $n_i$ is not constant, the resulting household weighted estimands are: $$\tau^P_{HW} = \frac{1}{N} \sum_{i=1}^N \frac{1}{n_i} \sum_{j=1}^{n_i} (Y_{ij}(1,1) - Y_{ij}(0,0)) \,\,\,\,\, \mbox{and} \,\,\,\,\, \tau^S_{HW} = \frac{1}{N} \sum_{i=1}^N \frac{1}{n_i} \sum_{j=1}^{n_i} (Y_{ij}(1,0) - Y_{ij}(0,0)).$$ These are the estimands in @hudgens2012toward. The corresponding individual weighted estimands are: $$\tau^P_{IW} = \frac{1}{n^+} \sum_i^N \sum_j^{n_i} (Y_{ij}(1,1) - Y_{ij}(0,0)) \,\,\,\,\,\,\, \mbox{ and } \,\,\,\,\,\,\, \tau^S_{IW} = \frac{1}{n^+} \sum_i^N \sum_j^{n_i} (Y_{ij}(1,0) - Y_{ij}(0,0)),$$ where $n^+$ is the total number of individuals. Estimation {#section:estimation} ========== Next, we turn to estimating these quantities of interest. First, we generalize the results of @hudgens2012toward to allow for unbiased estimation of any two-stage weighted estimand. We then discuss additional complications that arise when estimating individual-weighted effects. Unbiased estimation {#section:unbiased-estimation} ------------------- We now define two-stage weights for estimation, of which household and individual weights are special cases. \[prop:eq-est-unbiasedness\] Define two-stage inverse probability weights $w_i^{(00)}, w_i^{(10)}$, and $w_i^{(11)}$ as follows: $$\begin{aligned} w_i^{(11)} &= \frac{1}{P(H_i = 1)}\frac{1}{P(Z_{ij}=1 | H_i=1)}, \\ w_i^{(10)} &= \frac{1}{P(H_i = 1)}\frac{1}{P(Z_{ij}=0 | H_i=1)}, \\ w_i^{(00)} &= \frac{1}{P(H_i = 0)}. \end{aligned}$$ Consider two-stage estimand weights, $w_i^\ast$, as in Definition \[def:unequal-size-fp-effects\]. The weighted primary and spillover effect estimators $\widehat{\tau}^P_W$ and $\widehat{\tau}^S_W$, $$\begin{aligned} \widehat{\tau}^P_W &=& \sum_{(i,j)\in {\mathcal{T}}_{11}} w_i^{(11)} w_i^\ast Y_{ij}^{{\text{obs}}}(1,1) - \sum_{(i,j) \in {\mathcal{T}}_{00}} w_i^{(00)} w_i^\ast Y_{ij}^{{\text{obs}}}(0,0),\\ \widehat{\tau}^S_W &=& \sum_{(i,j)\in {\mathcal{T}}_{10}} w_i^{(10)} w_i^\ast Y_{ij}^{{\text{obs}}}(1,0) - \sum_{(i,j) \in {\mathcal{T}}_{00}} w_i^{(00)} w_i^\ast Y_{ij}^{{\text{obs}}}(0,0), \end{aligned}$$ are unbiased for their corresponding estimands with respect to the randomization distribution. That is $\mathbb{E}[\widehat{\tau}^P_W] = \tau^P_W$ and $\mathbb{E}[\widehat{\tau}^S_W] = \tau^S_W.$ The proof is given in the supplementary materials and follows closely from @hudgens2012toward. These estimators have a simple difference-in-means form either when household size is constant or with equal weight on households, as in @hudgens2012toward: $$\widehat{\tau}^P_{HW} = \frac{1}{N_1} \sum_{i\in {\mathcal{T}}_{11}}\overline{Y}_i^{{\text{obs}}}(1,1) - \frac{1}{N_0} \sum_{i \in {\mathcal{T}}_{00}} \overline{Y}_i^{{\text{obs}}}(0,0) \,\,\,\,\, \mbox{and} \,\,\,\,\, \widehat{\tau}_{HW}^S = \frac{1}{N_1} \sum_{i\in {\mathcal{T}}_{10}}\overline{Y}_i^{{\text{obs}}}(1,0) - \frac{1}{N_0} \sum_{i \in {\mathcal{T}}_{00}} \overline{Y}_i^{{\text{obs}}}(0,0).$$ The unbiased, individual weighted estimators have the form of Horvitz-Thompson estimators modified for our two-stage randomization. Thus the unbiased estimators for $\tau^P_{IW}$ and $\tau^S_{IW}$ are: $$\begin{aligned} \widehat{\tau}^P_{IW} &= \frac{1}{n^+}\frac{N}{N_1} \sum_{(i,j) \in {\mathcal{T}}_{11}} \frac{n_i}{1} \, Y_{ij}^{{\text{obs}}}(1,1) \quad\quad\;\; - \quad \frac{1}{n^+}\frac{N}{N_0} \sum_{(i,j) \in {\mathcal{T}}_{00}} Y_{ij}^{{\text{obs}}}(0,0), \\[1em] \widehat{\tau}^S_{IW} &= \frac{1}{n^+}\frac{N}{N_1} \sum_{(i,j) \in {\mathcal{T}}_{10}} \frac{n_i}{n_i - 1} \, Y_{ij}^{{\text{obs}}}(1,0) \quad - \quad \frac{1}{n^+}\frac{N}{N_0} \sum_{(i,j) \in {\mathcal{T}}_{00}} Y_{ij}^{{\text{obs}}}(0,0),\end{aligned}$$ with household-level assignment probabilities $N_1/N$ and $N_0/N$ and (conditional) individual-level probabilities $1/n_i$ and $(n_i-1)/n_i$. Since these are inverse probability weight estimators and since the probability of treatment assignment is a function of household size, the primary effect estimator up-weights larger households, with weights proportional to $n_i = \{1/n_i\}^{-1}$. Similarly, the spillover effect estimator down-weights larger households, with weights proportional to $n_i/(n_i - 1) = \{(n_i-1)/n\}^{-1}$. As is common with Horvitz-Thompson estimators, unbiased estimation typically comes at the price of additional variance. In practice, researchers can often reduce this variance by first normalizing the weights (i.e., H[á]{}jek weights), which introduces some small bias. See the supplementary materials for more details on the H[á]{}jek estimator in this context. Bias of the simple difference estimator {#section:sd-bias} --------------------------------------- Contrast the unbiased estimator with a simple difference estimator, that is, the difference-in-means across individuals ignoring households: $$\begin{aligned} \widehat{\tau}^{P}_{sd} &= \frac{1}{n_{11}^+} \sum_{(i,j) \in {\mathcal{T}}_{11}} Y_{ij}^{{\text{obs}}}(1,1) - \frac{1}{n_{00}^+} \sum_{(i,j) \in {\mathcal{T}}_{00}} Y_{ij}^{{\text{obs}}}(0,0) \label{eq:primary_sd}\\ \widehat{\tau}^{S}_{sd} &= \frac{1}{n_{10}^+} \sum_{(i,j) \in {\mathcal{T}}_{10}} Y_{ij}^{{\text{obs}}}(1,0) - \frac{1}{n_{00}^+} \sum_{(i,j) \in {\mathcal{T}}_{00}} Y_{ij}^{{\text{obs}}}(0,0),\label{eq:spillover_sd}\end{aligned}$$ where $n_{11}^+ = \sum_i {\mathds{1}}(H_i=1) \sum_j {\mathds{1}}(Z_{ij}=1)$, $n_{10}^+ = \sum_i {\mathds{1}}(H_i=1) \sum_j {\mathds{1}}(Z_{ij}=0)$, and $n_{00}^+ = \sum_i {\mathds{1}}(H_i=0)n_i$. These are essentially the estimators used in @Rogers_Feller_SDP. Despite its intuitive appeal, this estimator can be biased in practice. There are two main sources of bias. First, echoing results from @aronow_middleton_cluster, when household sizes vary, the quantities $n_{11}^+$, $n_{10}^+$, and $n_{00}^+$ are themselves random variables. Thus, both the numerator and denominator of each group average are random; and the mean of a ratio is not, in general, equal to the ratio of means. Second, individual-level treatment probabilities vary by household size; in the design we consider here, the probability of treatment assignment conditional on being in a treated household is $\mathbb{P}\{Z_{ij} = 1 \mid H_i = 1\} = 1/n_i$. Thus, ignoring $n_i$—and, by extension, the varying treatment probability—can lead to biased estimates. We derive the exact form of the bias in the following proposition. \[prop:bias-sd-iw\] The simple difference estimators, $\widehat{\tau}^{P}_{sd}$ and $\widehat{\tau}^{S}_{sd}$, defined in Equation \[eq:primary\_sd\] and \[eq:spillover\_sd\], have the following bias for their respective estimands. $$\text{bias}\left(\widehat{\tau}^{P}_{sd}\right) = \frac{1}{N\overline{n}} \sum_i \left( \frac{\overline{n}}{n_i} - 1\right) \sum_j Y_{ij}(1,1) + \frac{1}{N_0\overline{n}} \text{cov}\left( \frac{\sum_{{\mathcal{T}}_{00}} Y_{ij}(0,0)}{ n_{00}^+ } , n_{00}^+\right)$$ and $$\begin{aligned} \text{bias}\left(\widehat{\tau}^{S}_{sd}\right) &= \frac{1}{N\overline{n}} \sum_i \left(\frac{ \frac{\overline{n}}{\overline{n}-1} }{\frac{n_i}{n_i-1}} - 1\right) \sum_j Y_{ij}(1,0) + \bigg( \frac{1}{N_0\overline{n}} \text{cov}\left( \frac{\sum_{{\mathcal{T}}_{00}}Y_{ij}(0,0)}{n_{00}^+}, n_{00}^+\right) -\\ & \qquad\qquad \frac{1}{N_1(\overline{n}-1)} \text{cov}\left( \frac{\sum_{{\mathcal{T}}_{10}}Y_{ij}(1,0)}{n_{10}^+}, n_{10}^+\right) \bigg). \nonumber \end{aligned}$$ If household size is constant, all of these terms are zero. If the covariance between household size and potential outcomes is zero, only the first term of each equation remains. In simulations in Section \[section:simulations\], we show that the overall bias can be large if household sizes vary and treatment effects also vary by household size. Stratification and post-stratification by household size {#section:stratified-randomization} -------------------------------------------------------- Finally, we consider stratification and post-stratification by household size. If household-level randomization is stratified by household size, inference for the individual-weighted estimand is immediate. In particular, let $\tau^P_{k}$ and $\tau^S_{k}$ be the stratum-specific estimands for the stratum with household size $n_i = k$, $$\tau^P_k = \frac{1}{N^{(k)}} \sum_i^{N} {\mathds{1}}(n_i = k) \frac{1}{k} \sum_j^k (Y_{ij}(1,1) - Y_{ij}(0,0)) \,\,\,\,\,\, \mbox{and} \,\,\,\,\,\, \tau^S_k = \frac{1}{N^{(k)}} \sum_i^{N} {\mathds{1}}(n_i = k) \frac{1}{k} \sum_j^k (Y_{ij}(1,0) - Y_{ij}(0,0)),$$ where $N^{(k)}$ is the number of households of size $k$. Since household size is constant within each stratum, the corresponding household- and individual-weighted estimands are equivalent. We can therefore re-write the overall individual-weighted estimands as weighted averages of the stratum-specific effects, $$\tau^P_{IW} = \sum_{k=2}^K \frac{n^{(k)+}}{n^+} \tau^{P}_k \,\,\,\,\,\, \mbox{ and } \,\,\,\,\,\, \tau^S_{IW} = \sum_{k=2}^K \frac{n^{(k)+}}{n^+} \tau^{S}_k,$$ where $n^{(k)+} = \sum_{i:~n_i = k} n_i$ and where we assume (without essential loss of generality) that household sizes range from $k=2,\ldots, K$. Plugging in $\widehat{\tau}^{P}_k$ and $\widehat{\tau}^{S}_k$ gives the corresponding unbiased estimate. To modify the above results for household-weighted estimates, simply replace the weight $n^{(k)+}/n^+$ with $N^{(k)}/N$. In other words, weight each stratum by the number of households in that stratum, rather than the number of individuals. While stratification is not necessary to obtain unbiased estimates of household-weighted estimands, stratification will generally improve precision so long as household size is predictive of the outcome. In practice, it is not always possible or feasible to stratify randomization by household size. Fortunately, researchers can often post-stratify by household size; that is, the researcher can analyze the experiment as if randomization had been stratified by size. In the supplementary materials, we extend the theoretical guarantees from @miratrix2013adjusting to two-stage randomization, and include additional discussion of the technical details. Finally, if there are relatively small samples or the distribution of household sizes varies widely, researchers might want to “mix and match” among possible strategies. In the attendance study, there are 3,169 households of size $n_i = 2$ but only two households of size $n_i = 7$. Thus, it is unreasonable to post-stratify precisely on household size. Instead, we post-stratify by dividing household size into $n_i \in \{2, 3, 4-7\}$, using the unbiased IW estimator for households of size four to seven. This is inherently a bias-variance tradeoff and will depend on the particular context. If desired, we could also adjust for $n_i$ via regression, as discussed in Section \[section:covariate-adjustment\] [see also @aronow_middleton_cluster]. Of course, researchers should pre-specify such procedures whenever possible. Variance {#section:variance} ======== We next provide the theoretical variance of the unbiased, weighted estimators of Section \[section:unbiased-estimation\] as well as a conservative estimator of that variance. We conclude with a brief discussion of inference given a point estimate and its estimated variance. Theoretical and estimated variance ---------------------------------- We give general results for the variance of the unbiased, weighted estimator, of which the household and individual weights are special cases. Given estimand weights $w_i^*$, define the transformed potential outcomes as $Y_{ij}^w(h,z) \equiv N n_i w_i^\ast Y_{ij}(h,z)$. For the household weights, the transformed and original potential outcomes are identical, $Y_{ij}^w(h,z) = Y_{ij}(h,z)$. For the individual weights, the transformed potential outcomes are re-scaled by the relative household size, $Y_{ij}^w(h,z) = (n_i/\bar{n}) \cdot Y_{ij}(h,z)$, where $\bar{n}$ is the average household size. We now define several useful terms, effectively decomposing the overall variance of the transformed potential outcomes into a within- and between-household variance. Let $\sigma^{2,w}_{i,hz} = 1/n_i\sum_j (Y^w_{ij}(h,z) - \overline{Y}^w_i(h,z))^2$ be the within-household potential outcome variances for $Y^w_{ij}(h,z)$, and let $\Sigma_{11}^w = \frac{1}{N} \sum_i \sigma^{2,w}_{i,11}$ and $\Sigma^w_{10} = \frac{1}{N}\sum_i\frac{1}{(n_i-1)^2}\sigma^{2,w}_{i,10}$ be the (re-scaled) average within-cluster variances for $Y^w_{ij}(1,1)$ and $Y^w_{ij}(1,0)$ respectively. Finally, define the between-cluster variance of cluster-level averages: $$\begin{aligned} V_{hz}^w &=& \frac{1}{N-1} \sum_i^N (\overline{Y}^w_i(h,z) - \overline{Y}^w(h,z))^2, \\ V_P^w &=& \frac{1}{N-1} \sum_i^N ([\overline{Y}^w_i(1,1) - \overline{Y}^w_i(0,0)] - [\overline{Y}^w(1,1) - \overline{Y}^w(0,0)])^2,\\ V_S^w &=& \frac{1}{N-1} \sum_i^N ([\overline{Y}^w_i(1,0) - \overline{Y}^w_i(0,0)] - [\overline{Y}^w(1,0) - \overline{Y}^w(0,0)])^2,\end{aligned}$$ where $V_{hz}^w$ is the between-cluster variance of the average cluster-level potential outcome, $\overline{Y}^w_i(h,z)$. $V_P^w$ and $V_S^w$ are the (unidentifiable) cluster-level treatment effect variation for the primary and spillover effects, respectively. \[prop:eq-size-var\] The two-stage weighted estimators have the following variances under the randomization distribution: $${\mbox{\text{Var}}}(\widehat{\tau}_{W}^P) = \frac{\Sigma_{11}^w + V_{11}^w}{N_1} + \frac{V_{00}^w}{N_0} - \frac{V_P^w}{N}$$ and $${\mbox{\text{Var}}}(\widehat{\tau}_{W}^S) = \frac{\Sigma_{10}^w + V_{10}^w}{N_1} + \frac{V_{00}^w}{N_0} - \frac{V_S^w}{N}.$$ This variance has the same form as the standard Neymanian variance. However, the increased variance due to the two-level randomization is reflected in the first numerator, which has two terms instead of one. Intuitively, this is a decomposition of the marginal variance of potential outcomes into $\Sigma^w_{hz}$, the average of the within-household variances, and $V^w_{hz}$, the variance of the household-level average potential outcomes. We can obtain an estimated variance that is a “conservative” estimate for the true variance (in the sense of being too wide in expectation) with respect to the randomization distribution. Let $s^{2,w}_{hz}$ be the cluster-level sample variance for the cluster-level average transformed potential outcomes, $\overline{Y}^{{\text{obs}},w}_i(h,z)$. That is, $$s^{2,w}_{hz} = \frac{1}{N_h - 1} \sum_{i}^N {\mathds{1}}(H_i = h)\left(\overline{Y}^{{\text{obs}},w}_i(h,z) - \overline{Y}^{{\text{obs}},w}(h,z)\right)^2,$$ where $\overline{Y}^{{\text{obs}},w}(h,z)$ is the average transformed observed outcome for the set $\mathcal{T}_{hz}$, and where $\overline{Y}^{w,{\text{obs}}}_i(h,z)$ is the average observed outcome for the set $\mathcal{T}_{hz}$ in household $i$. \[th:eq-size-var-est-conservative\] Consider the variance estimators ${\widehat{\mbox{Var}}}(\widehat{\tau}_{W}^P)$ and ${\widehat{\mbox{Var}}}(\widehat{\tau}_{W}^S)$: $$\begin{aligned} {\widehat{\mbox{Var}}}(\widehat{\tau}_{W}^P) &=& \frac{s^{2,w}_{11}}{N_1} + \frac{s^{2,w}_{00}}{N_0}, \\ {\widehat{\mbox{Var}}}(\widehat{\tau}_{W}^S) &=& \frac{s^{2,w}_{10}}{N_1} + \frac{s^{2,w}_{00}}{N_0}. \end{aligned}$$ The proposed estimators are conservative estimates of their respective estimands. That is, $\mathbb{E}({\widehat{\mbox{Var}}}(\widehat{\tau}_{W}^P)) \geq {\mbox{\text{Var}}}(\widehat{\tau}_{W}^P)$ and $\mathbb{E}({\widehat{\mbox{Var}}}(\widehat{\tau}_{W}^S)) \geq {\mbox{\text{Var}}}(\widehat{\tau}_{W}^S)$. ${\widehat{\mbox{Var}}}(\widehat{\tau}_{W}^P)$ and ${\widehat{\mbox{Var}}}(\widehat{\tau}_{W}^S)$ are unbiased if $V_{P}^w = 0$ and $V_{S}^w = 0$, respectively. The results in Proposition \[prop:eq-size-var\] and Theorem \[th:eq-size-var-est-conservative\] can be applied to HW and IW estimators by simply plugging the appropriated weights defined in Proposition \[prop:eq-est-unbiasedness\]. In particular, plugging in the HW weights recovers the results of @hudgens2012toward. Consistent with their results, the estimated variance for the weighted estimator is unbiased if the treatment effects are constant. Inference --------- We briefly discuss inference for these quantities given an estimator and its estimated variance, following the setup in @liu2014large. For additional discussion, see @tchetgen2012estimation and @Rigdon:2015cv. For general outcomes, @liu2014large derive both Chebyshev and Wald confidence intervals (CIs) for household weighted estimands under two asymptotic regimes. In the first regime, the number of households (i.e., $N$) remains fixed while the size of each household grows large, i.e., $\text{min}(n_1, \ldots, n_N) \rightarrow \infty$. Inference in this regime either relies on Chebyshev CIs, which are typically too wide to be practically useful, or Wald CIs, which require several additional conditions that are unlikely to hold in our example, including homogeneity of impacts across households. In the second regime, the size of each hold (i.e., $n$) remains fixed while the number of households grows large, i.e., $N \rightarrow \infty$. At a high level, valid Wald CIs in this regime require a standard Lindeberg condition on the estimators and some restrictions on the within- and between-household variances, as well as requiring that $N_0/N_1$ remains relatively constant as $N\rightarrow \infty$. We rely on this approach here, as it seems like a reasonable asymptotic approximation in our setting. Thus, an asymptotic $1-\gamma$ CI for $\tau^P_{HW}$ is: $$\widehat{\tau}^P_{HW} \pm z_{1-\gamma/2}\sqrt{ {\widehat{\mbox{Var}}}(\widehat{\tau}^P_{HW}) },$$ where $z_{1-\gamma/2}$ is the $1-\gamma/2$ quantile of the standard Normal distribution. We caution that this asymptotic approximation might have poor performance with a small number of households [see, for example, @clubSandwich_package]. In the supplementary materials we discuss the corresponding assumptions for individual weighted estimands. We argue that we can obtain valid Wald CIs via the $N \rightarrow \infty$ asymptotic regime of @liu2014large separately for each household size stratum. Regression-based estimation {#sect:connection-with-regression} =========================== We now connect these randomization-based results with more familiar regression-based methods. Our key result is that, with the appropriate standard errors, conventional linear regression estimates are equivalent to the randomization-based estimates. This approach regards regression as a convenient tool [sometimes known as a *derived linear model*; see @hinkelmann2012design], and does not equate the regression with a specific generative model. In other words, this approach does not impose a model for the potential outcomes. See @baird2014designing for additional discussion of regression for two-stage randomized designs. We consider two basic regression approaches, an individual-level regression and a household-level regression. For simplicity, we start with the equal-sized household case and then show that these results generalize to any two-stage weights. We then demonstrate the dangers of using standard errors that ignore the two-stage structure. This section builds on existing results for robust and cluster-robust standard errors, especially @McCaffrey2001generalizations and @Bell2002bias. See also @cameron2015practitioner [@imbens2012robust; @clubSandwich_package]. Individual-level regression {#section:individual-level-regression} --------------------------- First, we construct the individual-level linear model, $$\label{eq:equal-linear-model-bad} Y_{ij}^{{\text{obs}}} = \alpha + \beta^P H_i Z_{ij} + \beta^S H_i (1 - Z_{ij}) + \varepsilon_{ij},$$ where the uncertainty in $\varepsilon_{ij}$ is entirely due to randomization. It is straightforward to show that standard OLS estimates for $\beta^P$ and $\beta^S$ are identical to the randomization-based estimators in Section \[section:unbiased-estimation\]. The theorem below states that the the randomization-based standard errors are equivalent to a particular cluster-robust generalization of heteroskedasticity-consistent standard errors, known as *HC2* [@mackinnon1985some]. \[th:reg-cluster-robust\] For equal-sized households, let $Y$ be the vector containing all the observed outcomes $Y^{\text{obs}}_{ij}$; let $X$ denote the appropriate design matrix (formally defined in the supplementary materials) with columns corresponding to the intercept, $HZ$, and $H(1-Z)$; and let $\beta = (\alpha, \beta^P, \beta^S)$. The linear model in Equation \[eq:equal-linear-model-bad\] can therefore be re-written as $Y = X\beta + \varepsilon$, with corresponding least squares estimate, $\widehat{\beta}^{{\text{ols}}}$. These estimates are unbiased for their corresponding estimands. Further, define the cluster-robust generalization of HC2 standard errors as: $$\widehat{Var}^{clust}_{hc2}(\widehat{\beta}^{{\text{ols}}}) = (X^tX)^{-1} \sum_{s=1}^S X_s^t(I_{N_s} - P_{ss})^{-1/2} \widehat{\varepsilon}_s \widehat{\varepsilon}_s (I_{N_s} - P_{ss})^{-1/2} X_s (X^tX)^{-1}$$ where $X_s$ and $\widehat{\varepsilon}_{s}$ are the subsets of $X$ and $\widehat{\varepsilon}$ corresponding to household $s$, and $P_{ss}$ is defined as $P_{ss} = X_s(X^tX)^{-1} X_s^t.$ Then $$\widehat{Var}^{clust}_{hc2}(\widehat{\beta}^{P,{\text{ols}}}) = \widehat{Var}(\widehat{\tau}^P) \,\,\,\,\,\, \mbox{ and } \,\,\,\,\,\, \widehat{Var}^{clust}_{hc2}(\widehat{\beta}^{S,{\text{ols}}}) = \widehat{Var}(\widehat{\tau}^S)$$ where $\widehat{Var}^{clust}_{hc2}(\widehat{\beta}^{P,{\text{ols}}}) = (\widehat{Var}^{clust}_{hc2}(\widehat{\beta}^{{\text{ols}}}))_{22}$ and $\widehat{Var}^{clust}_{hc2}(\widehat{\beta}^{S,{\text{ols}}}) = \widehat{Var}^{clust}_{hc2}(\widehat{\beta}^{ols}))_{33}$. In short, Theorem \[th:reg-cluster-robust\] confirms that we can obtain the same randomization-based point- and variance-estimators via the individual-level linear model in Equation \[eq:equal-linear-model-bad\] with HC2 cluster-robust standard errors. This is similar to results obtained with heteroskedastic-robust standard errors in simpler designs [e.g., @samii2012equivalencies; @imbens2015causal]. In Section \[section:no\_clustering\], we demonstrate the effect of failing to account for clustering on standard errors. Researchers can estimate these standard errors directly in `R` via, for example, the `clubSandwich` package. See @clubSandwich_package for additional discussion on the performance of clustered standard errors with a relatively small number of clusters. Household-level regression {#section:cluster-level-regression} -------------------------- We now consider regression at the household level. This is a common strategy in cluster-randomized trials and yields identical inference to individual-level regression with clustered standard errors [see, e.g., @cameron2015practitioner; @Athey:2016wn]. We separately aggregate treated and control units within each treated household, thus considering three types of household-level aggregates for household $i$. Each treated household has two household-level averages, $\overline{Y}_i^{\text{obs}}(1,1)$ and $\overline{Y}_i^{\text{obs}}(1,0)$; each control household has one household-level average, $\overline{Y}_i^{\text{obs}}(0,0)$. We can therefore assemble a vector of household-average outcomes, $\overline{Y}_k^{\text{obs}}$ of length $2N_1 + N_0$. We introduce the indicators $H^{(11)}_k$ and $H^{(10)}_k$; $H_k^{(11)} = 1$ if the aggregate is over the treated units in treated households and $H_k^{(11)} = 0$ otherwise; $H_k^{(10)} = 1$ if the aggregate is over the control units in treated households and $H_k^{(10)} = 0$ otherwise. We then consider the following linear model: $$\label{eq:equal-linear-model-good} \overline{Y}_k^{{\text{obs}}} = \alpha + \beta^P H_k^{(11)} + \beta^S H_k^{(10)} + \varepsilon'_k.$$ We now show that we can obtain the randomization-based point and variance estimates via the linear model estimates with standard (i.e., non-cluster) heteroskedastic-robust standard errors. \[th:reg\] For equal-sized households, the OLS estimates for $\beta^P$ and $\beta^S$ in Equation \[eq:equal-linear-model-good\] are unbiased estimators of the corresponding estimands. Define the heteroskedastic-robust estimator of the variances: $${\widehat{\mbox{Var}}_{hc2}}(\widehat{\tau}^P) \equiv {\widehat{\mbox{Var}}_{hc2}}(\widehat{\beta}^P) = \frac{\sum_{k: H^{(10)}_k=0} \widehat{\varepsilon}_{k}^2 (H^{(11)}_k -\overline{H})^2 }{(\sum_{k: H_k^{(10)}=0}(H^{(11)}_k - \overline{H}^{(11)})^2)^2}$$ and $${\widehat{\mbox{Var}}_{hc2}}(\widehat{\tau}^S) \equiv {\widehat{\mbox{Var}}_{hc2}}(\widehat{\beta}^S) = \frac{\sum_{k: H^{(11)}_k=0} \widehat{\varepsilon}_{k}^2 (H^{(10)}_k -\overline{H})^2 }{(\sum_{k: H_k^{(11)}=0}(H^{(10)}_k - \overline{H}^{(10)})^2)^2}$$ We have: $${\widehat{\mbox{Var}}_{hc2}}(\widehat{\tau}^P) = {\widehat{\mbox{Var}}}(\widehat{\tau}^P) \,\,\,\,\,\, \mbox{ and } \,\,\,\,\,\, {\widehat{\mbox{Var}}_{hc2}}(\widehat{\tau}^S) = {\widehat{\mbox{Var}}}(\widehat{\tau}^S),$$ where the $\widehat{\varepsilon}_{k}$’s are the $HC2$ residuals (see supplementary materials for exact definition). In short, Theorem \[th:reg\] states that we can aggregate to the household level and proceed as if this were a standard completely randomized trial [@imbens2015causal]. Intuitively, the aggregation at the household level is another way of accounting for the household structure in the two-stage randomization scheme. Since the definition for the heteroskedastic-robust estimator of the variance is standard, it is straightforward to fit Equation \[eq:equal-linear-model-good\] and the corresponding variance in `R` via, for example, the `vcovHC` function with the `HC2` option in the `sandwich` package [@zeileis2004sandwich]. Results for weighted estimands ------------------------------ Finally, we can modify the household-level regression to estimate any two-stage weighted estimand. Thus, we can use this approach to recover the results in Sections \[section:estimation\] and \[section:variance\]. The key idea is to run an unweighted regression on transformed outcomes. \[th:reg-weight\] Let $w_i^*$ be two-stage estimand weights, with transformed potential outcomes, $Y^{obs,w}_{ij}(h,z) = N n_i w_i^* Y^{obs}_{ij}(h,z)$. Then the results of Theorem \[section:cluster-level-regression\] hold when applied to the transformed potential outcomes. That is: $\widehat{\beta}^P = \widehat{\tau}_W^P$, $\widehat{\beta}^S = \widehat{\tau}_W^S$, $\widehat{Var}_{hc2}(\widehat{\beta}^P) = \widehat{Var}(\widehat{\tau}_W^P)$, and $\widehat{Var}_{hc2}(\widehat{\beta}^S) = \widehat{Var}(\widehat{\tau}_W^S)$. This approach is subtly different from using Weighted Least Squares, which would also reweight the design matrix. Failing to account for clustering {#section:no_clustering} --------------------------------- It is instructive to consider the consequences of “naively” analyzing a two-stage experiment as if it were a completely randomized experiment, ignoring the household structure. With equal-sized households, the point estimates will be the same as for the appropriate analysis, but the standard errors will differ. In particular, let $\widehat{{\mbox{\text{Var}}}}^{\text{het}}_{\text{hc2}}(\widehat{\tau}^P)$ be the (non-cluster) HC2 robust standard error for the primary effect; that is, these are the variances from Equation (\[eq:equal-linear-model-good\]) for the household aggregates incorrectly applied to the individual level. We show in the supplementary materials that, $$\mathbb{E}\bigg[\widehat{{\mbox{\text{Var}}}}^{\text{het}}_{\text{hc2}}(\widehat{\tau}^P)\bigg] - {\mbox{\text{Var}}}(\widehat{\tau}^P)=(\Sigma_{00} + V_{00})\left\{\frac{1}{nN_0-1}\bigg(1 - n\rho_{00}\bigg) - \frac{1}{N}\frac{V_p}{\Sigma_{00} + V_{00}}\right\},$$ where $V_{00}$ and $\Sigma_{00}$ are the between- and within-household variances, respectively, of the control potential outcomes, $Y_{ij}(0,0)$, and $\rho_{00} \equiv V_{00}/(\Sigma_{00} + V_{00})$ is the intraclass correlation (ICC). This quantity is negative—that is, the variance is anti-conservative in expectation—if: $$\rho_{00} > \frac{1}{n} - \left(\frac{nN_0 - 1}{nN}\right) \left(\frac{ V_P }{\Sigma_{00} + V_{00}}\right).$$ Since the last term is non-negative, we can build intuition by setting that term to zero. For example, consider the special case of $V_P = 0$, which would occur if there is a constant additive effect. Under these conditions, the estimated variance is anti-conservative if $\rho_{00} > 1/n$. In the social sciences, typical values of ICC range from 0.1 to 0.3 [e.g., @gelman2006data]. Thus, even with households of size 4 or 5, the estimated variance could be anti-conservative. We see this behavior in the simulations in Section \[section:simulations\]. Covariate adjustment {#section:covariate-adjustment} ==================== Finally, we explore how to incorporate individual- and cluster-level covariates in a two-stage experiment. There is an extensive literature on the use of covariates in randomized trials [see, for example, @imbens2015causal]. In this section, we briefly address stratification, post-stratification, and model-assisted estimation. #### Stratification and post-stratification. As with household size in Section \[section:stratified-randomization\], the simplest way to account for covariates is to incorporate them into the randomization by stratifying on them. In general, the researcher can partition households into discrete strata, regard each stratum as a separate “mini experiment,” and estimate a pooled effect by averaging across strata. This will improve the precision of the treatment effect estimate so long as the stratifying covariate is predictive of the outcome. Researchers cannot, however, stratify by a covariate that varies within household, as this could destroy the nested structure in the data by assigning different individuals in the same household to different “household-level” treatments. For example, we cannot stratify the two-stage randomization by gender, as some houses will have both boys and girls; instead, we could stratify by whether all students in the household are boys, which is an aggregate version of the individual-level covariate. Finally, as with household size, researchers can also post-stratify on household-level covariates. Of course, it is possible to combine stratification and post-stratification; for example, first stratify by household size and then, for each household size, post-stratify by whether the household speaks English as the primary language. #### Model-assisted estimation. We also consider model-assisted estimation to incorporate covariates [@cochran::1977; @rosenbaum2002covariance; @hansen2009attributing; @aronow2013class]. Following the setup in @hansen2009attributing, consider $K$ covariates $x^{(1)}, \ldots, x^{(K)}$ (which typically include a constant) with corresponding coefficient vector, $\gamma = (\gamma_1, \ldots, \gamma_K)$, such that $r(\gamma) = \{r_{ij}(\{x^{(k)}\}_k, \gamma)\}$ for $i=1,\ldots,N$ and $j=1,\ldots,n_i$ is a function mapping covariates to predictions. To simplify notation, we will let $x = \{x^{(k)}\}_k$. In practice, the coefficients in $\gamma$ are typically coefficients from a linear regression. In this case, let $$r_{ij}(x, \gamma) = \sum_k^K x_{ij}^{(k)} \gamma_k,$$ where $x_{ij}$ is a vector of covariates associated with unit $j$ in cluster $i$. We regard $r_{ij}(x, \gamma)$ as fixed and known for all units, rather than estimated from the data. We then define the (residualized) potential outcome as: $$e^\gamma_{ij}(h,z) = Y_{ij}(h,z) - r_{ij}(x, \gamma).$$ As with the corresponding potential outcomes, $Y_{ij}(h,z)$, the residualized potential outcomes, $e^\gamma_{ij}(h,z)$, are assumed to be fixed and are only observed if $H_i = h$ and $Z_{ij} = z$. We can then substitute $e^\gamma_{ij}(h,z)$ for $Y_{ij}(h,z)$ in defining the primary effect: $$\begin{aligned} \tau^P &= \overline{Y}(1,1) - \overline{Y}(0,0) \\ &= \left[\overline{e}^\gamma(1,1) + \overline{r}(x, \gamma)\right] - \left[\overline{e}^\gamma(0,0) + \overline{r}(x, \gamma)\right] = \overline{e}^\gamma(1,1) - \overline{e}^\gamma(0,0),\end{aligned}$$ re-writing $\tau^S$ in an analogous way. Given $r_{ij}(\gamma)$, model-assisted estimation of $\tau^P$ is immediate via substituting the observed values of the residualized outcomes, $e^{\gamma, {\text{obs}}}_{ij}$, in place of the unadjusted outcomes, $Y_{ij}^{\text{obs}}$. The resulting difference-in-means estimator is unbiased regardless of the exact values of $r_{ij}(\gamma)$; that is, there is no need to appeal to a “correctly specified” linear model to obtain the particular coefficient vector, $\gamma$. So long as the covariates are predictive of the outcome, the variance of $e^{\gamma, {\text{obs}}}_{ij}$ will generally be smaller than the variance of $Y^{\text{obs}}_{ij}$, and the resulting model-assisted estimator will also have smaller estimated variance (so long as $\gamma$ does not include extreme values). Finally, the above derivations assume that $\gamma$ is fixed and known; in practice, we must find some way to determine $\gamma$. The most straightforward approach is to generate a random hold-out sample, and estimate $\gamma$ via a regression of $Y$ on $X$ for this group. While not always possible, the attendance study effectively has a hold-out sample that we use for this purpose; see Section \[section:data\]. See @hansen2009attributing and @aronow2013class for additional discussion. Simulations {#section:simulations} =========== Failing to account for the cluster structure -------------------------------------------- We now turn to simulations assessing the importance of accounting for the cluster structure. Reflecting the nested structure in the actual experiment, we generate potential outcomes in two stages. First, we simulate household-level average potential outcomes via: $$\begin{aligned} \overline{Y}_{i}(0,0) &\sim N\left(\mu_{00}, \sigma^2_{\mu}\right) \\ \tau_i^P &\sim N\left(\overline{\tau}^P, \sigma^2_{\tau^P}\right)\\ \tau_i^S &\sim N\left(\overline{\tau}^S, \sigma^2_{\tau^S}\right),\end{aligned}$$ where $\overline{Y}_{i}(1,1) = \overline{Y}_{i}(0,0) + \tau_i^P$ and $\overline{Y}_{i}(1,0) = \overline{Y}_{i}(0,0) + \tau_i^S$. Then, conditional on these values, we generate individual-level potential outcomes via $Y_{ij}(h,z) \sim N\left(\overline{Y}_i(h,z), \sigma^2_y\right),$ for each $h$ and $z$. Across all simulations, we fix the mean potential outcomes, with $\mu_{00} = 2$, $\bar{\tau}^S = 0.7$, and $\bar{\tau}^P = 1.5$, and fix household size at $n_i = 4$ for all households. For convenience, we also restrict the household-level variance terms to be equal to each other, such that $\sigma_\mu = \sigma_{\tau^P} = \sigma_{\tau^S} = \sigma_c$, where $\sigma_c$ is the common standard deviation. Thus, we vary three main parameters, $N \in \{50, 100, 500, 1000\}$ (we always set $N_1=N/2$) and $\sigma_{c}, \sigma_y \in \{0.1, 0.2, 0.3, 0.4, 0.5\}$. For each combination of parameters, we consider three methods: cluster-robust standard errors; non-cluster, robust standard errors; and nominal standard errors. For each, we compute the coverage of the associated 95% confidence interval, averaging over 2,000 random draws of the assignment vector. $\widehat{\tau}^P$ $\widehat{\tau}^S$ ------------------------ -------------------- -------------------- Cluster robust SEs 0.97 0.98 Non-cluster robust SEs 0.93 0.87 Nominal SEs 0.95 0.86 : Average coverage of 95% confidence intervals over 2,000 replications.[]{data-label="tbl:sim_overview"} Table \[tbl:sim\_overview\] shows the overall coverage for 95% confidence intervals, averaged across all values of the simulation parameters. As expected, coverage with the cluster-robust standard errors is slightly larger than 95% coverage. By contrast, the non-cluster and nominal standard errors have below 95% coverage. First, this coverage pattern is quite stable across different sample sizes. At the same time, coverage strongly depends on the within- and between-household variances. Figure \[fig:coverage\_icc\] shows the relationship between coverage and the intraclass correlation among control potential outcomes, defined as $\sigma^2_c/(\sigma^2_c + \sigma^2_y)$ in this simulation. Consistent with the results in Section \[section:no\_clustering\], the coverage for non-clustered standard errors grows increasingly poor as the ICC increases. [0.5]{} ![Coverage for 95% confidence intervals for clustered SEs, non-cluster robust SEs, and nominal SEs, with respect to the intraclass correlation of control potential outcomes, $\sigma^2_c/(\sigma^2_c + \sigma^2_y)$.[]{data-label="fig:coverage_icc"}](coverage_icc_primary "fig:"){width="\textwidth"} [0.5]{} ![Coverage for 95% confidence intervals for clustered SEs, non-cluster robust SEs, and nominal SEs, with respect to the intraclass correlation of control potential outcomes, $\sigma^2_c/(\sigma^2_c + \sigma^2_y)$.[]{data-label="fig:coverage_icc"}](coverage_icc_spillover "fig:"){width="\textwidth"} Comparing the three estimators for the IW estimand -------------------------------------------------- We now focus on the IW estimand and compare the unbiased estimator, the difference-in-means estimator, and the post-stratified estimator. We consider two scenarios: (a) when treatment effects are uncorrelated with household size; and (b) when treatment effects are correlated with household size. The data generating process is the same as above, with a balanced household-level randomization with $N = 200$, $N_1 = 100$, and fixed $\sigma_c = \sigma_y = 0.3$. We generate households of size 2, 3 and 4 with equal probability, and introduce the parameters $\overline{\tau}^P_k, \overline{\tau}^S_k, \mu_{00}^{(k)}$ for $k = 2,3,4$. For scenario (a), we set $\overline{\tau}^P_k = 1.5, \overline{\tau}^S_k=0.7, \mu_{00}^{(k)}=2$ for all $k=2,3,4$. For scenario (b), we allow the effects to vary by household size, as follows: $\overline{\tau}^P_2 = 1.5, \overline{\tau}^P_3 = 0.75, \overline{\tau}^P_4 = 0.37$, $\overline{\tau}^S_2 = 0.7, \overline{\tau}^S_3 = 0.35, \overline{\tau}^S_4 = 0.17$, and $\mu_{00}^{(2)} = 2, \mu_{00}^{(3)}=1, \mu_{00}^{(4)} = 0.5$. The results are presented in Table \[table:iw-results\]. We see that when the treatment effect is uncorrelated with household size, the bias of all three estimators is negligible, but the Monte Carlo standard error of the unbiased estimator is an order of magnitude larger than that of the other two estimators. When treatment effect is correlated with household size, the biases of the unbiased and the post-stratified estimators are still very small, but the bias of the simple difference estimator is substantial—roughly the same size as the standard error. Again, the standard errors are smallest for the post-stratified estimator; overall, the post-stratified estimator clearly dominates in terms of RMSE. --------- ------------------- ---------------- ------ ----------------- ---------------- ------ Avg. $|$bias$|$ Monte Carlo SE Avg. $|$bias$|$ Monte Carlo SE Unbiased 0.11 0.07 Simple difference 0.04 0.18 0.12 Post-stratified 0.04 0.04 \[1em\] Unbiased 0.12 0.07 Simple difference 0.04 0.12 0.13 Post-stratified 0.04 0.04 --------- ------------------- ---------------- ------ ----------------- ---------------- ------ : Bias and SE for different estimators for the IW estimand over 2,000 replications. ‘’ denotes $\leq 10^{-3}$.[]{data-label="table:iw-results"} Student absenteeism in the School District of Philadelphia {#section:data} ========================================================== Setup and covariate balance --------------------------- We now apply these methods to our motivating example, the @Rogers_Feller_SDP attendance intervention in the School District of Philadelphia. The original study included three active treatment arms and one control arm, with $N = 28,080$ total households. The first treatment arm merely reminded parents of the importance of attendance and did not provide any student-specific information. The second and third arms provided parents with different types of student-specific information. @Rogers_Feller_SDP found weak impacts of assignment to the first arm and large impacts for the second and third arms (all relative to control). They also found minimal differences between the second and third arms. We now make three modifications that preserve the substantive questions from the original study but allow us to focus on the two-stage randomized design. First, we exclude households from the weak first treatment arm, instead using this group as the holdout sample for estimating the covariate adjustment model (see Section \[section:covariate-adjustment\]). Second, we combine the second and third treatment arms, since these are substantively very similar and have virtually identical impacts. Together, these modifications essentially create a two-arm trial: control vs. combined second and third arms from the original study. Finally, we restrict the universe to the subset of households with two or more eligible students, which yields $N = 3,876$ total households, of which $N_1 = 2,568$ (66 percent) were assigned to treatment and $N_0 = 1,308$ (34 percent) were assigned to control, and $n^+ = 8,654$ total students. We next assess covariate balance for this sample. While household-level randomization was not stratified by household size (see Section \[section:stratified-randomization\]), the balance by household size is excellent, as shown in Table \[tbl:num\_students\_hh\]. Table \[tbl:cov\_balance\] shows the covariate balance for each stage of randomization. The left bank shows balance for the first stage, household-level randomization, with covariates aggregated to the household level. The right bank shows balance for the second stage randomization, with individual-level covariates among households assigned to treatment ($H_i = 1$). Statistically, Table \[tbl:cov\_balance\] shows that covariate balance is excellent for both stages of randomization, with all normalized differences [@imbens2015causal] below 0.05 in absolute value. Substantively, Table \[tbl:cov\_balance\] emphasizes that the students come from largely disadvantaged households. Over three-quarters of these students qualify for Free or Reduced Price Lunch, which is only available to families at or near the Federal Poverty Line. Over 15 percent of households speak a language other than English at home, with 7 percent of students designated as Limited English Proficiency (LEP). Moreover, this is a very high-absence group, with an average of around 13 days absent in the previous school year (out of roughly 180 possible days); we also include number of absences prior to randomization in early October. We observe the grade for each student, which we treat as a discrete covariate and which ranges from first grade to high school senior. While we do not show balance by grade to conserve space, there is excellent balance across this covariate as well. -------------------------------- --------------- --------------- ---------- -- ------------------ ------------------ ---------- For $H_i = 1$ For $H_i = 0$ $\Delta$ For $Z_{ij} = 1$ For $Z_{ij} = 0$ $\Delta$ Female 0.53 0.54 -0.03 0.53 0.53 -0.01 Black/African-American 0.51 0.50 0.03 0.52 0.52 -0.01 English spoken at home 0.84 0.83 0.04 0.84 0.83 0.02 Limited English Proficiency 0.07 0.07 -0.01 0.07 0.08 -0.04 Free or Reduced Price Lunch 0.78 0.79 -0.03 0.78 0.79 -0.02 Prior year absences (days) 16.70 16.55 0.02 16.65 16.93 -0.03 Start-of-year absences (days) 1.20 1.14 0.05 1.22 1.21 0.01 Students per household ($n_i$) 2.2 2.2 -0.01 -------------------------------- --------------- --------------- ---------- -- ------------------ ------------------ ---------- : Covariate means and normalized differences ($\Delta$) by stage of randomization. For the first stage, balance is assessed for covariates aggregated to the household level. For the second stage, balance is assessed for individual-level covariates, restricted to households assigned to treatment ($H_i = 1$).[]{data-label="tbl:cov_balance"} Finally, we are broadly interested in days absent as the outcome of interest. However, the distribution of absences has a long right tail; for example, several students in the sample are absent for over half the school year. As this greatly increases the variance, we consider two transformed outcomes of interest. First, we consider an indicator for whether a student is chronically absent, defined as missing 18 or more days during the school year, i.e., ${\mathds{1}}(\text{days} \geq 18)$; among students in the control group, 36 percent are chronically absent. Second, we consider log-absences, defined as $\text{log}(\text{days} + 1)$, to allow for a continuous outcome without the very heavy right tail; baseline absences among students in the control group are around 13 days or $\text{log}(13 + 1) \approx 2.6$. For interpretability, we also report key point estimates in terms of raw days. Results ------- Figure \[fig:SDP\_main\_effects\] shows the estimated impacts and corresponding 95% confidence intervals for the primary and spillover effects for both household- and individual-weighted estimands. In terms of chronic absenteeism (i.e., the binary outcome), the estimates for the HW and IW estimands are nearly identical: the unbiased estimates for the primary effects are around -4 percentage points (SE of 1.5 percentage points) for both $\tau_{HW}^P$ and $\tau_{IW}^P$; the unbiased estimates for the spillover effects are around -3 percentage points (SE of 1.5 percentage points) for both $\tau_{HW}^S$ and $\tau_{IW}^S$. These results are virtually unchanged when post-stratifying by household size, using the (conditionally) unbiased estimator within each post-stratification cell, defined by $n_i = 2$, $n_i = 3$, and $n_i \in \{4, \ldots, 7\}$. As discussed in Section \[section:overall-effect\] and in the supplementary materials, the overall effect is essentially a weighted average of the primary and spillover effects. Unsurprisingly, the unadjusted estimate of the overall effect is around 3.5 percentage points (SE of 1.4 percentage points) for both the HW and IW weighted estimands, with nearly identical results for the post-stratified estimate. The results are somewhat more variable for the impact on log-absences (i.e., the continuous outcome). The point estimates are quite close for the household-weighted and individual-weighted estimands: $\widehat{\tau}^P_{HW} = -0.085$ log-days and $\widehat{\tau}^P_{IW} = -0.093$ log-days for the primary effect, and $\widehat{\tau}^S_{HW} = -0.051$ log-days and $\widehat{\tau}^S_{IW} = -0.058$ log-days for the spillover effect. In terms of raw days, these are roughly -1.2 days for the primary effect and -0.7 days for the spillover effect. The point estimates are similarly close for the post-stratified estimator. The standard errors, however, are considerably larger for the unadjusted IW estimates: roughly 0.033 log-days for the IW estimands compared to 0.023 log-days for the HW estimands. Thus, the corresponding confidence intervals are roughly 50 percent larger for the IW estimands than for the HW esitmands. Post-stratification greatly reduces the standard errors for the IW estimand: for both IW and HW estimates, the standard errors are roughly 0.023 log-days, comparable to the standard errors for the HW estimate without post-stratification. As with the impact on chronic absenteeism, the estimates for the overall effect fall between the primary and spillover effect estimates, with estimates around -0.07 log-days (SE of 0.023 log-days) or around 1 day. Despite minor differences between the sets of estimates, the pattern of effects is fairly clear: in general, we find that the spillover effect is between 60 and 80 percent as large as the primary effect, depending on the outcome. We also find few differences between the HW and IW estimates. [0.5]{}   [0.5]{} [0.5]{}   [0.5]{} Next, Figure \[fig:SDP\_cov\_adj\] shows covariate-adjusted estimates for individual-weighted estimands. First, we take advantage of the fact that there is a natural holdout sample in the experiment as analyzed; see Section \[section:data\]. To obtain $\widehat{r}_{ij}(\gamma)$, we regress the outcome on covariates listed in Table \[tbl:cov\_balance\] as well as student grade (categorical). Results do not appear sensitive to the particular choice of model. The resulting point estimates in Figure \[fig:SDP\_cov\_adj\] are largely unchanged, if slightly larger in magnitude than the unadjusted estimates. The standard errors, however, are meaningfully smaller, especially for the continuous outcome: 0.018 log-days for the model-assisted estimator versus 0.033 log-days for the unadjusted estimator. Next, we can combine model-assisted estimation with post-stratification, though the results are essentially identical. Finally, while do not have theoretical guarantees for covariate adjustment in a two-stage experiment, classical regression adjustment is nearly identical to the model-assisted estimation with the holdout sample. In the end, we find strong evidence of intra-household spillover for the attendance intervention. This pattern holds with and without covariate adjustment, though the covariate-adjusted estimates are more precise. This underscores that merely focusing on the primary effect significantly under-estimates the impact and cost effectiveness of the intervention. @Rogers_Feller_SDP report costs of around \$6.60 per household. In our sample, the primary effect is around 1.2 days and the overall effect (i.e., the weighted average of the primary and spillover effects) is around 1 day. Thus, if we only consider the primary effect, the cost for each additional student day is \$6.60 / 1.2 days $\approx$ \$5.50 / day. By contrast, if we also consider spillovers, the cost for each additional student day is \$6.60 / ($2.2 \cdot 1$ day) $\approx$ \$3 / day, where $\bar{n} = n^+/N \approx 2.2$ is the average number of students per household in our sample. Discussion\[section:discussion\] ================================ Two-stage randomizations are increasingly common designs in settings with interactions between units. This paper addresses several practical issues in analyzing such designs. First, we address issues that arise when household sizes vary. Second, we demonstrate that regression can yield identical point- and variance-estimates to those derive from fully randomization-based methods. Methodologically, we believe that this is a useful addition to the literatures on both causal inference with interference and randomization-based inference. Substantively, we find convincing evidence of spillover effects of a large-scale attendance intervention. There are several directions for future work. First, we are actively exploring covariate adjustment in this and other settings with more complex randomization schemes. The model-assisted approach is one such option, but many are possible [@lin2013agnostic; @aronow2013class]. In particular, extending the asymptotic results of @liu2014large to incorporate covariates would be fruitful. Second, there is an open question of how to separately test the null hypotheses for no primary and no spillover effects in this type of design. Recent work from @athey2015exact offers one promising direction. Third, it will be useful to extend these results to other, related designs. For example, @Weiss:2016dm discuss an interesting setting in which random assignment occurs at the individual level but individuals are then administered treatment in groups (such as in group therapy). @kang_imbens_peer_encourage propose a “peer encouragement” design, which extends the two-stage randomization considered here to consider noncompliance. Fourth, we anticipate additional connections with non-randomized studies that mimic a two-stage randomized design, such as @hong2006evaluating and @perez2014assessing. Finally, we believe one promising direction for future work is to relax the stratified interference assumption while retaining partial interference. For example, @paluck2016changing collect detailed social network data for students within schools; the authors assume no interference between schools but leverage the within-school network structure for inference. @Arpino:2016du offer another possibility, using covariates to construct interference patterns between units. In the end, we hope that the results we give here will lead to increased use of two-stage randomized designs in practice. [^1]: Email: `[email protected]`. We thank Peter Aronow, Peng Ding, Winston Lin, Joel Middleton, Caleb Miles, Luke Miratrix, James Pustejovsky, Todd Rogers, Shruthi Subramanyam, John Ternovksi, and Elizabeth Tipton for helpful comments and discussion. We also thank the excellent research partners at the School District of Philadelphia, especially Adrienne Reitano and Tonya Wolford.
{ "pile_set_name": "ArXiv" }
--- bibliography: - 'diet\_ramses.bib' title: Cosmological Simulations using Grid Middleware --- Introduction {#sec:intro} ============ One way to access the aggregated power of a collection of heterogeneous machines is to use a grid middleware, such as [<span style="font-variant:small-caps;">Diet</span>]{}[@CDL+02a], GridSolve [@GridSolve06] or Ninf [@Ninf99]. It addresses the problem of monitoring the resources, of handling the submissions of jobs and as an example the inherent transfer of input and output data, in place of the user. In this paper we present how to run cosmological simulations using the [<span style="font-variant:small-caps;">Ramses</span>]{}application along with the [<span style="font-variant:small-caps;">Diet</span>]{}middleware. We will describe how to write the corresponding [<span style="font-variant:small-caps;">Diet</span>]{}client and server. The remainder of the paper is organized as follows: Section \[sec:diet\_overview\] presents the [<span style="font-variant:small-caps;">Diet</span>]{}middleware. Section \[sec:ramses\_overview\] describes the [<span style="font-variant:small-caps;">Ramses</span>]{}cosmological software and simulations, and how to interface it with [<span style="font-variant:small-caps;">Diet</span>]{}. We show how to write a client and a server in Section \[sec:interfacing\]. Finally, Section \[sec:experiments\] presents the experiments realized on Grid’5000, the French Research Grid, and we conclude in Section \[sec:ccl\]. DIET overview {#sec:diet_overview} ============= DIET architecture ----------------- [<span style="font-variant:small-caps;">Diet</span>]{}[@CDL+02a] is built upon the client/agent/server paradigm. A **Client** is an application that uses [<span style="font-variant:small-caps;">Diet</span>]{}to solve problems. Different kinds of clients should be able to connect to [<span style="font-variant:small-caps;">Diet</span>]{}: from a web page, a PSE such as Matlab[^1] or Scilab[^2], or from a program written in C or Fortran. Computations are done by servers running a **Server Daemons ([[<span style="font-variant:small-caps;">SeD</span>]{}]{})**. A [<span style="font-variant:small-caps;">SeD</span>]{}encapsulates a computational server. For instance it can be located on the entry point of a parallel computer. The information stored by a [<span style="font-variant:small-caps;">SeD</span>]{}is a list of the data available on its server, all information concerning its load (for example available memory and processor) and the list of problems that it can solve. The latter are declared to its parent agent. The hierarchy of scheduling agents is made of a **Master Agent (MA)** and **Local Agents (LA)** (see Figure \[fig:archi\_appli\]). When a Master Agent receives a computation request from a client, agents collect computation abilities from servers (through the hierarchy) and chooses the best one according to some scheduling heuristics. The MA sends back a reference to the chosen server. A client can be connected to a MA by a specific name server or by a web page which stores the various MA locations (and the available problems). The information stored on an agent is the list of requests, the number of servers that can solve a given problem and information about the data distributed in its subtree. For performance reasons, the hierachy of agents sould be deployed depending on the underlying network topology. Finally, on the opposite of GridSolve and Ninf which rely on a classic socket communication layer (nevertheless several problems to this approach have been pointed out such as the lack of portability or the limitation of opened sockets), [<span style="font-variant:small-caps;">Diet</span>]{}uses Corba. Indeed, distributed object environments, such as *Java*, *DCOM* or Corba have proven to be a good base for building applications that manage access to distributed services. They provide transparent communications in heterogeneous networks, but they also offer a framework for the large scale deployment of distributed applications. Moreover, Corba systems provide a remote method invocation facility with a high level of transparency which does not affect performance [@DPP01]. How to add a new grid application within [<span style="font-variant:small-caps;">Diet</span>]{}? ------------------------------------------------------------------------------------------------ [r]{}[.4]{} The main idea is to provide some integrated level for a grid application. Figure \[fig:archi\_appli\] shows these different kinds of level. The **application server** must be written to give [<span style="font-variant:small-caps;">Diet</span>]{}the ability to use the application. A simple API is available to easily provide a connection between the [<span style="font-variant:small-caps;">Diet</span>]{}server and the application. The main goals of the [**[<span style="font-variant:small-caps;">Diet</span>]{}server**]{} are to answer to monitoring queries from its responsible Local Agent and launch the resolution of a service, upon an application client request. The **application client** is the link between high-level interface and the [<span style="font-variant:small-caps;">Diet</span>]{}client, and a simple API is provided to easily write one. The main goals of the **[<span style="font-variant:small-caps;">Diet</span>]{}client** are to submit requests to a scheduler (called Master Agent) and to receive the identity of the chosen server, and final step, to send the data to the server for the computing phase. [<span style="font-variant:small-caps;">Ramses</span>]{}overview {#sec:ramses_overview} ================================================================ [<span style="font-variant:small-caps;">Ramses</span>]{}[^3] is a typical computational intensive application used by astrophysicists to study the formation of galaxies. [<span style="font-variant:small-caps;">Ramses</span>]{}is used, among other things, to simulate the evolution of a collisionless, self-gravitating fluid called “dark matter” through cosmic time (see Figure \[fig:ramses\_result\]). Individual trajectories of macro-particles are integrated using a state-of-the-art “N body solver”, coupled to a finite volume Euler solver, based on the Adaptive Mesh Refinement technics. The computational space is decomposed among the available processors using a *mesh partitionning* strategy based on the Peano–Hilbert cell ordering ([@Teyssier02; @Teyssier06]). Cosmological simulations are usually divided into two main categories. Large scale periodic boxes (see Figure \[fig:ramses\_result\]) requiring massively parallel computers are performed on very long elapsed time (usually several months). The second category stands for much faster small scale “zoom simulations”. One of the particularity of the HORIZON project is that it allows the re-simulation of some areas of interest for astronomers. For example in Figure \[fig:zoom\], a supercluster of galaxies has been chosen to be re-simulated at a higher resolution (highest number of particules) taking the initial information and the boundary conditions from the larger box (of lower resolution). This is the latter category we are interested in. Performing a zoom simulation requires two steps: the first step consists of using [<span style="font-variant:small-caps;">Ramses</span>]{}on a low resolution set of initial conditions [*i.e.,* ]{}with a small number of particles) to obtain at the end of the simulation a catalog of “dark matter halos”, seen in Figure \[fig:ramses\_result\] as high-density peaks, containing each halo position, mass and velocity. A small region is selected around each halo of the catalog, for which we can start the second step of the “zoom” method. This idea is to resimulate this specific halo at a much better resolution. For that, we add in the Lagrangian volume of the chosen halo a lot more particles, in order to obtain more accurate results. Similar “zoom simulations” are performed in parallel for each entry of the halo catalog and represent the main resource consuming part of the project. [l]{}[5.5cm]{} [<span style="font-variant:small-caps;">Ramses</span>]{}simulations are started from specific initial conditions, containing the initial particle masses, positions and velocities. These initial conditions are read from Fortran binary files, generated using a modified version of the <span style="font-variant:small-caps;">Grafic</span>[^4] code. This application generates Gaussian random fields at different resolution levels, consistent with current observational data obtained by the WMAP[^5] satellite observing the cosmic microwave background radiation. Two types of initial conditions can be generated with <span style="font-variant:small-caps;">Grafic</span>: -0.15cm - single level: this is the “standard” way of generating initial conditions. The resulting files are used to perform the first, low-resolution simulation, from which the halo catalog is extracted. - multiple levels: this initial conditions are used for the “zoom simulation”. The resulting files consist of multiple, nested boxes of smaller and smaller dimensions, as for Russian dolls. The smallest box is centered around the halo region, for which we have locally a very high accuracy thanks to a much larger number of particles. The result of the simulation is a set of ”snaphots”. Given a list of time steps (or expansion factor), [<span style="font-variant:small-caps;">Ramses</span>]{}outputs the current state of the universe ([*i.e.,* ]{}the different parameters of each particules) in Fortran binary files. These files need post-processing with <span style="font-variant:small-caps;">Galics</span> softwares: HaloMaker, TreeMaker and GalaxyMaker. These three softwares are meant to be used sequentially, each of them producing different kinds of information: - HaloMaker: detects dark matter halos present in [<span style="font-variant:small-caps;">Ramses</span>]{}output files, and creates a catalog of halos - TreeMaker: given the catalog of halos, TreeMaker builds a merger tree: it follows the position, the mass, the velocity of the different particules present in the halos through cosmic time - GalaxyMaker: applies a semi-analytical model to the results of TreeMaker to form galaxies, and creates a catalog of galaxies Interfacing [<span style="font-variant:small-caps;">Ramses</span>]{}within [<span style="font-variant:small-caps;">Diet</span>]{} {#sec:interfacing} ================================================================================================================================= Architecture of underlying deployment ------------------------------------- The current version of [<span style="font-variant:small-caps;">Ramses</span>]{}requires a NFS working directory in order to write the output files, hence restricting the possible types of solving architectures. Each [<span style="font-variant:small-caps;">Diet</span>]{}server will be in charge of a set of machines (typically 32 machines to run a $256^3$ particules simulation) belonging to the same cluster. For each simulation the generation of the initial conditions files, the processing and the post-processing are done on the same cluster: the server in charge of a simulation manages the whole process. Server design ------------- The [<span style="font-variant:small-caps;">Diet</span>]{}server is a library. So the [<span style="font-variant:small-caps;">Ramses</span>]{}server requires to define the `main()` function, which contains the problem profile definition and registration, and the solving function, whose parameter only consists of the profile and named after the service name, `solve_serviceName`. The [<span style="font-variant:small-caps;">Ramses</span>]{}solving function contains the calls to the different programs used for the simulation, and which will manage the MPI environment required by [<span style="font-variant:small-caps;">Ramses</span>]{}. It is recorded during the profile registration. The [<span style="font-variant:small-caps;">SeD</span>]{}is launched with a call to `diet_SeD()` in the `main()` function, which will never return (except if some errors occur). The [<span style="font-variant:small-caps;">SeD</span>]{}forks the solving function when requested. ### Defining services {#sec:defservice} To match client requests with server services, clients and servers must use the same problem description. A unified way to describe problems is to use a name and define its arguments. The [<span style="font-variant:small-caps;">Ramses</span>]{}service is described by a profile description structure called `diet_profile_desc_t`. Among its fields, it contains the name of the service, an array which does not contain data, but their characteristics, and three integers `last_in, last_inout` and `last_out`. The structure is defined in `DIET_server.h`. The array is of size $last\_out + 1$. Arguments can be: [IN:]{} Data are sent to the server. The memory is allocated by the user. [INOUT:]{} Data, allocated by the user, are sent to the server and brought back into the same memory zone after the computation has completed, [*w*ithout any copy]{}. Thus freeing this memory while the computation is performed on the server would result in a segmentation fault when data are brought back onto the client. [OUT:]{} Data are created on the server and brought back into a newly allocated zone on the client. This allocation is performed by [<span style="font-variant:small-caps;">Diet</span>]{}. After the call has returned, the user can find its result in the zone pointed at by the *value* field. Of course, [<span style="font-variant:small-caps;">Diet</span>]{}cannot guess how long the user needs these data for, so it lets him/her free the memory with `diet_free_data()`. The fields *last\_in*, *last\_inout* and *last\_out* of the structure respectively point at the indexes in the array of the last IN, last INOUT and last OUT arguments. Functions to create and destroy such profiles are defined with the prototypes below. Note that if a server can solve multiple services, each profile should be allocated. ``` {.c language="C"} diet_profile_desc_t *diet_profile_desc_alloc( const char* path, int last_in, int last_inout, int last_out ); diet_profile_desc_t *diet_profile_desc_alloc( int last_in, int last_inout, int last_out ); int diet_profile_desc_free(diet_profile_desc_t *desc); ``` The cosmological simulation is divided in two services: `ramsesZoom1` and `ramsesZoom2`, they represent the two parts of the simulation. The first one is used to determine interesting parts of the universe, while the second is used to study these parts in details. The `ramsesZoom2` service uses nine data. The seven firsts are IN data, and contain the simulation parameters: - a file containing parameters for [<span style="font-variant:small-caps;">Ramses</span>]{} - resolution of the simulation (number of particules) - size of the initial conditions (in $Mpc.h^{-1}$) - center’s coordinates of the initial conditions (3 coordinates: $c_x$, $c_y$ and $c_z$) - number of zoom levels (number of nested boxes) The last two are an integer for error controls, and a file containing the results obtained from the simulation post-processed with <span style="font-variant:small-caps;">Galics</span>. This conducts to the following inclusion in the server code (note: the same allocation must be performed on the client side, with the `diet_profile_t` structure): ``` {.c language="C"} /* arg.profile is a diet_profile_desc_t * */ arg.profile = diet_profile_desc_alloc("ramsesZoom2", 6, 6, 8); ``` Every argument of the profile must then be set with `diet_generic_desc_set()` defined in `DIET_server.h`, like: ``` {.c language="C"} diet_generic_desc_set(diet_parameter(pb,0), DIET_FILE, DIET_CHAR); diet_generic_desc_set(diet_parameter(pb,1), DIET_SCALAR, DIET_INT); ``` ### Registering services Every defined service has to be added in the service table before the [<span style="font-variant:small-caps;">SeD</span>]{}is launched. The complete service table API is defined in `DIET_server.h`: ``` {.c language="C"} typedef int (* diet_solve_t)(diet_profile_t *); int diet_service_table_init(int max_size); int diet_service_table_add(diet_profile_desc_t *profile, NULL, diet_solve_t solve_func); void diet_print_service_table(); ``` The first parameter, *profile*, is a pointer on the profile previously described (section \[sec:defservice\]). The second parameter concerns the convertor functionality, but this is out of scope of this paper and never used for this application. The parameter *solve\_func* is the type of the `solve_serviceName()` function: a function pointer used by [<span style="font-variant:small-caps;">Diet</span>]{}to launch the computation. Here, the prototype is then: ``` {.c language="C"} int solve_ramsesZoom2(diet_profile_t* pb) { /* Data downloading */ /* Computation */ /* Data uploading */ } ``` ### Data management The first part of the solve function (called `solve_ramsesZoom2()`) is to receive data. The API provides useful functions to help coding the solve function, [*e.g.,* ]{}get IN arguments, set OUT ones, with `diet__get()` functions defined in `DIET_data.h`. Do not forget that the necessary memory space for OUT arguments is allocated by [<span style="font-variant:small-caps;">Diet</span>]{}. So the user should call the `diet__get()` functions to retrieve the pointer to the zone his/her program should write to. To set INOUT and OUT arguments, one should use the `diet__desc_set()` defined in `DIET_server.h`. These should be called within “solve” functions only. ``` {.c language="C"} diet_file_get(diet_parameter(pb,0), NULL, &arg_size, &nmlPath); diet_scalar_get(diet_parameter(pb,1), &resol, NULL); diet_scalar_get(diet_parameter(pb,2), &size, NULL); diet_scalar_get(diet_parameter(pb,3), &cx, NULL); diet_scalar_get(diet_parameter(pb,4), &cy, NULL); diet_scalar_get(diet_parameter(pb,5), &cz, NULL); diet_scalar_get(diet_parameter(pb,6), &nbBox, NULL); ``` The results of the simulation are packed into a tarball file if it succeeded. Thus we need to return this file and an error code to inform the client whether the file really contains results or not. In the following code, the `diet_file_set()` function associate the [<span style="font-variant:small-caps;">Diet</span>]{}parameter with the current file. Indeed, the data should be available for [<span style="font-variant:small-caps;">Diet</span>]{}, when it sends the resulting file to the client. ``` {.c language="C"} char* tgzfile = NULL; tgzfile = (char*)malloc(tarfile.length()+1); strcpy(tgzfile, tarfile.c_str()); diet_file_set(diet_parameter(pb,7), DIET_VOLATILE, tgzfile); ``` Client {#sec:client} ------ In the [<span style="font-variant:small-caps;">Diet</span>]{}architecture, a client is an application which uses [<span style="font-variant:small-caps;">Diet</span>]{}to request a service. The goal of the client is to connect to a Master Agent in order to dispose of a [<span style="font-variant:small-caps;">SeD</span>]{}which will be able to solve the problem. Then the client sends input data to the chosen [<span style="font-variant:small-caps;">SeD</span>]{}and, after the end of computation, retrieve output data from the [<span style="font-variant:small-caps;">SeD</span>]{}. [<span style="font-variant:small-caps;">Diet</span>]{}provides libraries containing functions to easily and transparently access the [<span style="font-variant:small-caps;">Diet</span>]{}platform. ### Structure of a client program {#sec:cl_struct} Since the client side of [<span style="font-variant:small-caps;">Diet</span>]{}is a library, a client program has to define the `main()` function: it uses [<span style="font-variant:small-caps;">Diet</span>]{}through function calls. ``` {.c language="C"} #include "DIET_client.h" int main(int argc, char *argv[]) { diet_initialize(configuration_file, argc, argv); // Successive DIET calls ... diet_finalize(); } ``` The client program must open its [<span style="font-variant:small-caps;">Diet</span>]{}session with a call to `diet_initialize()`. It parses the configuration file given as the first argument, to set all options and get a reference to the [<span style="font-variant:small-caps;">Diet</span>]{}Master Agent. The session is closed with a call to `diet_finalize()`. It frees all resources, if any, associated with this session on the client, servers, and agents, but not the memory allocated for all INOUT and OUT arguments brought back onto the client during the session. Hence, the user can still access them (and still has to free them !). The client API follows the GridRPC definition [@gridRPC:02]: all `diet_` functions are “duplicated” with `grpc_` functions. Both `diet_initialize()`/`grpc_initialize()` and `diet_finalize()`/`grpc_finalize()` belong to the GridRPC API. A problem is managed through a *function\_handle*, that associates a server to a service name. The returned *function\_handle* is associated to the problem description, its profile, during the call to `diet_call()`. ### Data management The API to the [<span style="font-variant:small-caps;">Diet</span>]{}data structures consists of modifier and accessor functions only: no allocation function is required, since `diet_profile_alloc()` allocates all necessary memory for all argument **descriptions**. This avoids the temptation for the user to allocate the memory for these data structures twice (which would lead to [<span style="font-variant:small-caps;">Diet</span>]{}errors while reading profile arguments). Moreover, the user should know that arguments of the `_set` functions that are passed by pointers are **not** copied, in order to save memory. Thus, the user keeps ownership of the memory zones pointed at by these pointers, and he/she must be very careful not to alter it during a call to [<span style="font-variant:small-caps;">Diet</span>]{}. An example of prototypes: ``` {.c language="C"} int diet_scalar_set(diet_arg_t* arg, void* value, diet_persistence_mode_t mode, diet_base_type_t base_type); int diet_file_set(diet_arg_t* arg, diet_persistence_mode_t mode, char* path); ``` Hence the arguments used in the `ramsesZoom2` simulation are declared as follows: // IN parameters if (diet_file_set(diet_parameter(arg.profile,0), DIET_VOLATILE, namelist)) { cerr << "diet_file_set error on the <namelist.nml> file" << endl; return 1; } diet_scalar_set(diet_parameter(arg.profile,1), &resol, DIET_VOLATILE, DIET_INT); diet_scalar_set(diet_parameter(arg.profile,2), &size, DIET_VOLATILE, DIET_INT); diet_scalar_set(diet_parameter(arg.profile,3), &arg.cx, DIET_VOLATILE, DIET_INT); diet_scalar_set(diet_parameter(arg.profile,4), &arg.cy, DIET_VOLATILE, DIET_INT); diet_scalar_set(diet_parameter(arg.profile,5), &arg.cz, DIET_VOLATILE, DIET_INT); diet_scalar_set(diet_parameter(arg.profile,6), &arg.nbBox, DIET_VOLATILE, DIET_INT); // OUT parameters diet_scalar_set(diet_parameter(arg.profile,8), NULL, DIET_VOLATILE, DIET_INT); if (diet_file_set(diet_parameter(arg.profile,7), DIET_VOLATILE, NULL)) { cerr << "diet_file_set error on the OUT file" << endl; return 1; } It is to be noticed that the OUT arguments should be declared even if their values is set to NULL. Their values will be set by the server that will execute the request. Once the call to [<span style="font-variant:small-caps;">Diet</span>]{}is done, we need to access the OUT data. The 8[^th^]{}parameter is a file and the 9[^th^]{}is an integer containing the error code of the simulation (`0` if the simulation succeeded): ``` {.c language="C"} int* returnedValue; size_t tgzSize = 0; char* tgzPath = NULL; diet_scalar_get(diet_parameter(simusZ2[reqID].profile,8), &returnedValue, NULL); if (!*returnedValue) { diet_file_get(diet_parameter(simusZ2[reqID].profile,7), NULL, &tgzSize, &tgzPath); } ``` Experiments {#sec:experiments} =========== Experiments description {#sec:platform_desc} ----------------------- Grid’5000[^6] is the French Research Grid. It is composed of 9 sites spread all over France, each with 100 to 1000 PCs, connected by the RENATER Education and Research Network (1Gb/s or 10Gb/s). For our experiments, we deployed a [<span style="font-variant:small-caps;">Diet</span>]{}platform on 5 sites (6 clusters). - 1 MA deployed on a single node, along with omniORB, the monitoring tools, and the client - 6 LA: one per cluster (2 in Lyon, and 1 in Lille, Nancy, Toulouse and Sophia) - 11 [[<span style="font-variant:small-caps;">SeD</span>]{}]{}s: two per cluster (one cluster of Lyon had only one [<span style="font-variant:small-caps;">SeD</span>]{}due to reservation restrictions), each controling 16 machines (AMD Opterons 246, 248, 250, 252 and 275) We studied the possibility of computing a lot of low-resolution simulations. The client requests a $128^3$ particles $100Mpc.h^{-1}$ simulation (first part). When he receives the results, he requests simultaneously 100 sub-simulations (second part). As each server cannot compute more than one simulation at the same time, we won’t be able to have more than 11 parallel computations at the same time. Results ------- The experiment (including both the first and the second part of the simulation) lasted 16h 18min 43s (1h 15min 11s for the first part and an average of 1h 24min 1s for the second part). After the first part of the simulation, each [<span style="font-variant:small-caps;">SeD</span>]{}received 9 requests (one of them received 10 requests) to compute the second part (see Figure \[fig:comp\_time\_expe\], left). As shown in Figure \[fig:comp\_time\_expe\] (right) the total execution time for each [<span style="font-variant:small-caps;">SeD</span>]{}is not the same : about 15h for Toulouse and 10h30 for Nancy. Consequently, the schedule is not optimal. The equal distribution of the requests does not take into account the machines processing power. In fact, at the time when [<span style="font-variant:small-caps;">Diet</span>]{}receives the requests (all at the same time) the second part of the simulation has never been executed, hence [<span style="font-variant:small-caps;">Diet</span>]{}doesn’t know anything on its processing time, the best it can do is to share the total amount of requests on the available [[<span style="font-variant:small-caps;">SeD</span>]{}]{}s. A better makespan could be attained by writing a plug-in scheduler[@CCDS06]. The benefit of running the simulation in parallel on different clusters is clearly visible: it would take more than 141h to run the 101 simulation sequentially. Furthermore, the overhead induced by the use of [<span style="font-variant:small-caps;">Diet</span>]{} is extremely low. Figure \[fig:lat\_find\_expe\] shows the time needed to find a suitable [<span style="font-variant:small-caps;">SeD</span>]{}for each request, as well as in log scale, the latency ([*i.e.,* ]{}the time needed to send the data from the client to the chosen [<span style="font-variant:small-caps;">SeD</span>]{}, plus the time needed to initiate the service). The finding time is low and nearly constant (49.8ms on average). The latency grows rapidly. Indeed, the client requests 100 sub-simulations simultaneously, and each [<span style="font-variant:small-caps;">SeD</span>]{}cannot compute more than one of them at the same time. Requests cannot be proceeded until the completion of the precedent one. This waiting time is taken into account in the latency. Note that the average time for initiating the service is 20.8ms (taken on the 12 firsts executions). The average overhead for one simulation is about 70.6ms, inducing a total overhead for the 101 simulations of 7s, which is neglectible compared to the total processing time of the simulations. Conclusion {#sec:ccl} ========== In this paper, we presented the design of a [<span style="font-variant:small-caps;">Diet</span>]{}client and server based on the example of cosmological simulations. As shown by the experiments, [<span style="font-variant:small-caps;">Diet</span>]{}is capable of handling long cosmological parallel simulations: mapping them on parallel resources of a grid, executing and processing communication transfers. The overhead induced by the use of [<span style="font-variant:small-caps;">Diet</span>]{}is neglectible compared to the execution time of the services. Thus [<span style="font-variant:small-caps;">Diet</span>]{}permits to explore new research axes in cosmological simulations (on various low resolutions initial conditions), with transparent access to the services and the data. [^1]: <http://www.mathworks.fr/> [^2]: <http://www.scilab.org/> [^3]: <http://www.projet-horizon.fr/Codes> [^4]: [ http://web.mit.edu/edbert]( http://web.mit.edu/edbert) [^5]: <http://map.gsfc.nasa.gov> [^6]: <http://www.grid5000.fr>
{ "pile_set_name": "ArXiv" }
--- abstract: 'We derive a system of coupled flow equations for the proper-vertices of the background effective average action and we give an explicit representation of these by means of diagrammatic and momentum space techniques. This explicit representation can be used as a new computational technique that enables the projection of the flow of a large new class of truncations of the background effective average action. In particular, these can be single– or bi–field truncations of local or non–local character. As an application we study non–abelian gauge theories. We show how to use this new technique to calculate the beta function of the gauge coupling (without employing the heat kernel expansion) under various approximations. In particular, one of these approximations leads to a derivation of beta functions similar to those proposed as candidates for an “all–orders” beta function. Finally, we discuss some possible phenomenology related to these flows.' author: - 'Alessandro Codello[^1]' title: | RG flow equations for the proper-vertices\ of the background effective average action --- *SISSA*\ *via Bonomea 265,* *I-34136* *Trieste, Italy* Introduction ============ The effective average action (EAA) formalism is a promising approach to QFT which recently has seen important developments and applications [@Berges_Tetradis_Wetterich_2002; @Gies_2006]. The EAA is a $k$ dependent functional that interpolates smoothly between the bare action for $k=\Lambda$ and the standard effective action for $k=0$; from this point of view it offers a new approach to quantization. When applied to theories with local symmetries, as non-abelian gauge theories [@Reuter_Wetterich_1994a; @Reuter_1995], non-linear sigma models [@Codello_Percacci_2009], gravity [@Reuter_1996], or membranes [@Codello_Zanusso_2011] it can be implemented using the background field method [@Reuter_Wetterich_1994a]. This defines the background EAA (bEAA). In this formalisms one usually makes a truncation ansatz for the bEAA and then performs calculations with the aid of heat kernel techniques [@Codello_Percacci_Rahmede_2009]. Generally, the range of applicability of heat kernel methods is confined to special cases, where the Hessian of the bEAA is of generalized Laplacian type [@Benedetti_Groh_Machado_Saueressig_2010]. This technology permits the study of truncations where the bEAA is the sum of local operators; these are interesting from the point of view of renormalizability, but are not enough to capture the full form of the bEAA, which in general contains non-local operators. These non-local terms are of fundamental importance, if one wants to construct RG trajectories that reach the far infrared (IR) $k=0$. It has been shown in a workable examples [@Codello_2010] that only by employing these kind of truncations it is possible to construct an EAA that correctly reproduces the effective action in the limit $k\rightarrow0$. Another important and complementary issue that requires stronger computational tools is the one related to “bi–field” truncations [@Reuter_Bi_Field]: since (as we will review in section 2) the full bEAA is a functional of both fluctuation and background fields, its RG flow takes place in the enlarged theory space where this functional naturally lives. Bi–field truncations generally give rise to operators that are difficult to treat using heat kernel methods. In this paper we present a novel formalisms based on the hierarchy of coupled flow equations satisfied by the one-particle-irreducible (1PI) proper-vertices of the bEAA. We give an explicit representation of these equations by means of diagrammatic and momentum space techniques. This explicit representation can be used as a new computational technique that enables the projection of the flow of a large new class of truncations of the bEAA. These include both the non-local and bi-field truncations mentioned above. The hierarchy of flow equations for the proper-vertices contains vertices with both fluctuation and background legs. The non-trivial part of the formalism is the explicit momentum space representation of vertices with background legs, since these contain legs attached to the cutoff action (which is function of the background field). These terms are not present in the non-background formalism of the standard EAA and constitute the necessary contributions to make the RG flow covariant. A key element in the derivation of this representation is the perturbative expansion of the (un-traced) heat kernel [@Codello_Zanusso_2012]. As an application we study non-abelian gauge theories. First we show how standard heat kernel computation of the one–loop beta function of the gauge coupling is reproduced by our formalism, then we extend the study under two approximations, a single– and a bi–field one, to obtain RG improved beta functions. One of these turns out to be very similar to the “all-orders” beta function proposed by Ryttov and Sannino [@Ryttov_Sannino_2008]. Finally, we discuss some possible phenomenology related to these RG flows. In the second section of the paper we review the basic properties of the bEAA and we derive the hierarchy of coupled flow equations for the proper-vertices and we expose the relative momentum space representation by introducing the diagrammatic rules. The details of the derivation of the non-trivial momentum space representation of vertices with background legs are given in appendix A. In the third section we apply our formalisms to non-abelian gauge theories to show how it works in practice. Flow equations for proper-vertices of the bEAA ============================================== In this section we first review the construction of the bEAA and the derivation of the exact RG flow equation it satisfies; successively we derive the hierarchy of flow equations for the proper-vertices of bEAA and we expose their momentum space representation and relative diagrammatic rules. Construction of bEAA -------------------- The crucial point in the construction of the EAA for theories with local gauge symmetries is, obviously, the preservation of gauge invariance during the RG coarse-graining procedure. A possible way to cut–off field modes covariantly is to define the cutoff using covariant differential operators. But if we try to introduce a cutoff by simply taking it as a function of a covariant differential operator, constructed with the quantum fields, we will spoil the simple one-loop structure of the exact flow that the EAA obeys. To obtain a one-loop like flow equation, the cutoff action has to be quadratic in the quantum fields. Still, the EAA will not be gauge invariant because of the non-covariant coupling of the quantum fields to the source, a problem that affects also the standard effective action. A way out of this is to employ the background field method, as was first done in [@Reuter_Wetterich_1994a; @Reuter_Wetterich_1993a], to define what we will call the background effective average action (bEAA). The theories that we have in mind are non-abelian gauge theories, gravity and non-linear sigma models. In this paper we will use the language of non-abelian gauge theories to keep the notation as light as possible, but it is clear that everything is readily translated to the other cases. In the background field formalism [@Abbott_1981] the quantum field $A_{\mu}$ is linearly split between the background field $\bar{A}_{\mu}$ and the fluctuation field $a_{\mu}$: $$A_{\mu}=\bar{A}_{\mu}+a_{\mu}\,.\label{gauge_1}$$ The cutoff action $\Delta S_{k}$ is taken to be quadratic in the fluctuation field, while the cutoff kernel $R_{k}[\bar{A}]$ is constructed employing the background field alone: $$\Delta S_{k}[\varphi;\bar{A}]=\frac{1}{2}\int d^{d}x\,\varphi\, R_{k}[\bar{A}]\varphi\,,\label{gauge_2}$$ where $\varphi=(a_{\mu},\bar{c},c)$ is the fluctuation multiplet, combining the fluctuation field $a_{\mu}$ and the ghost fields $\bar{c}$ and $c$. To construct the bEAA we introduce the cutoff action (\[gauge\_2\]) into the integro-differential definition of the standard background effective action[^2]: $$e^{-\Gamma_{k}[\varphi;\bar{A}]}=\int D\chi\exp\left(-S[\chi+\varphi;\bar{A}]-\Delta S_{k}[\chi;\bar{A}]+\int d^{d}x\,\Gamma_{k}^{(1;0)}[\varphi;\bar{A}]\chi\right)\,,\label{gauge_3}$$ where the fluctuation field multiplet $\chi$ has vanishing vacuum expectation value $\left\langle \chi\right\rangle =0$. The role of the cutoff action is to suppress field modes with momenta smaller than the RG scale $k$, the shape of the cutoff kernel must be constructed in a way consistent with this property; more details on this can be found in [@Berges_Tetradis_Wetterich_2002; @Gies_2006]. With these definitions the bEAA is a functional that interpolates smoothly between the bare action (\[gauge\_4\]) for $k\rightarrow\infty$ (fact that can be easily checked) and the full background effective action for $k\rightarrow0$. This fact, together with the exact flow equation bEAA satisfies (that we derive in the next subsection), offers the starting point to develop a formalisms that can be used to define and construct QFT characterized by local gauge symmetries. The computational technique introduced in this paper has to be seen as a tool to pursue this research route. In (\[gauge\_3\]) $S[\varphi;\bar{A}]$ is the bare action which is the sum of an invariant action (which can be the classical non-abelian gauge theory action), the background gauge-fixing action and the background ghost action: $$S[\varphi;\bar{A}]=S[\bar{A}+a]+S_{gf}[a;\bar{A}]+S_{gh}[a,\bar{c},c;\bar{A}]\,.\label{gauge_4}$$ The background gauge-fixing action is $$S_{gf}[a;\bar{A}]=\frac{1}{2\alpha}\int d^{d}x\,\bar{D}_{\mu}a^{\mu}\bar{D}_{\nu}a^{\nu}\,,\label{gauge_5}$$ where $\alpha$ is the gauge-fixing parameter, while the background ghost action is $$S_{gh}[a,\bar{c},c;\bar{A}]=\int d^{d}x\,\bar{D}_{\mu}\bar{c}\, D^{\mu}c=\int d^{d}x\,\bar{D}_{\mu}\bar{c}\left(\bar{D}^{\mu}+ga^{\mu}\right)c\,.\label{gauge_6}$$ With these definitions, the bEAA (\[gauge\_3\]) is invariant under combined physical[^3], $$\delta_{\theta}A_{\mu}=D_{\mu}\theta\qquad\delta_{\theta}\bar{c}=[\bar{c},\theta]\qquad\delta_{\theta}c=[c,\theta]\qquad\delta_{\theta}\bar{A}_{\mu}=0\,,\label{gauge_6A}$$ and background, $$\bar{\delta}_{\theta}A_{\mu}=\delta_{\theta}\bar{c}=\bar{\delta}_{\theta}c=0\qquad\qquad\bar{\delta}_{\theta}\bar{A}_{\mu}=\bar{D}_{\mu}\theta\,,\label{gauge_6B}$$ gauge transformations: $$(\delta_{\theta}+\bar{\delta}_{\theta})\Gamma_{k}[\varphi;\bar{A}]=0\,.\label{gauge_6.01}$$ For a proof of (\[gauge\_6.01\]) and for more details about the (background) gauge invariance of the bEAA we refer to the literature [@Ellwanger_1994]. See also [@Lavrov_Shapiro_2013]. It is possible to define a gauge invariant functional of the background field, that we will call “gauge invariant effective average action” (gEAA), by setting to zero the fluctuation multiplet $\varphi=0$ in the bEAA: $$\bar{\Gamma}_{k}[\bar{A}]\equiv\Gamma_{k}[0;\bar{A}]\,.\label{gauge_6.1}$$ This definition is equivalent to a parametrization of the bEAA as the sum of a functional of the full quantum field $A_{\mu}=\bar{A}_{\mu}+a_{\mu}$, the gEAA (\[gauge\_6.1\]), and a “remainder functional” $\hat{\Gamma}_{k}[\varphi;\bar{A}]$ (rEAA) which remains a functional of both the fluctuation multiplet and the background field: $$\Gamma_{k}[\varphi;\bar{A}]=\bar{\Gamma}_{k}[\bar{A}+a]+\hat{\Gamma}_{k}[\varphi;\bar{A}]\,.\label{gauge_6.2}$$ To recover (\[gauge\_6.1\]) we must have $\hat{\Gamma}_{k}[0;\bar{A}]=0$. The gEAA defined in this way is a functional invariant under physical gauge transformations (\[gauge\_6A\]): $$\delta_{\theta}\bar{\Gamma}_{k}[\bar{A}]=0\,,\label{gauge_6.3}$$ while the rEAA remains a functional invariant under combined physical and background gauge transformations, as the full bEAA, and is subject to modified Ward-Takahashi identities. We refer to [@Reuter_Wetterich_1994a] for more details on this point. What is important is that the gEAA flows, in the IR $k\rightarrow0$, to the standard gauge invariant effective action of the background field formalism, which can thus be computed using the bEAA formalism. ### Exact flow equation for the bEAA We derive now the exact flow equation satisfied by the bEAA [@Reuter_Wetterich_1994a]. Differentiating with respect to the “RG time” $t=\log k/k_{0}$ both sides of the integro-differential equation (\[gauge\_3\]) we find: $$\begin{aligned} e^{-\Gamma_{k}[\varphi;\bar{A}]}\partial_{t}\Gamma_{k}[\varphi;\bar{A}] & = & \int D\chi\left(\partial_{t}\Delta S_{k}[\chi;\bar{A}]-\int d^{d}x\,\partial_{t}\Gamma_{k}^{(1;0)}[\varphi;\bar{A}]\chi\right)\times\nonumber \\ & & \times e^{-S[\varphi+\chi;\bar{A}]-\Delta S_{k}[\chi;\bar{A}]+\int\Gamma_{k}^{(1;0)}[\varphi;\bar{A}]\chi}\,.\label{gauge_7}\end{aligned}$$ We can re-express the terms on the rhs of (\[gauge\_7\]) as expectation values and use (\[gauge\_2\]) to rewrite (\[gauge\_7\]) as: $$\begin{aligned} \partial_{t}\Gamma_{k}[\varphi;\bar{A}] & = & \left\langle \partial_{t}\Delta S_{k}[\chi;\bar{A}]\right\rangle -\int d^{d}x\,\partial_{t}\Gamma_{k}^{(1;0)}[\varphi;\bar{A}]\left\langle \chi\right\rangle \nonumber \\ & = & \frac{1}{2}\int d^{d}x\left\langle \chi\chi\right\rangle \partial_{t}R_{k}\,,\label{gauge_8}\end{aligned}$$ where we used $\left\langle \chi\right\rangle =0$. The two-point function of $\chi$ in (\[gauge\_8\]) can be written in terms of the inverse Hessian of the bEAA plus the cutoff action (where functional derivatives are taken with respect to $\varphi$): $$\left\langle \chi\chi\right\rangle =\left(\Gamma_{k}^{(2;0)}[\varphi;\bar{A}]+\Delta S_{k}^{(2;0)}[\varphi;\bar{A}]\right)^{-1}=\left(\Gamma_{k}^{(2;0)}[\varphi;\bar{A}]+R_{k}[\bar{A}]\right)^{-1}\,,\label{gauge_9}$$ where in the second step we used the fact that the Hessian of cutoff action (\[gauge\_2\]) is the cutoff kernel $R_{k}[\bar{A}]$. Inserting (\[gauge\_9\]) back into (\[gauge\_8\]), and writing a functional trace in place of the integral, finally gives: $$\partial_{t}\Gamma_{k}[\varphi;\bar{A}]=\frac{1}{2}\mathrm{Tr}\left(\Gamma_{k}^{(2;0)}[\varphi;\bar{A}]+R_{k}[\bar{A}]\right)^{-1}\,\partial_{t}R_{k}[\bar{A}]\,.\label{gauge_10}$$ Equation (\[gauge\_10\]) is the exact RG flow equation for the bEAA [@Reuter_Wetterich_1994a]. The flow equation (\[gauge\_10\]) has the same general properties as the flow equation for the standard effective average action [@Berges_Tetradis_Wetterich_2002]. In the following it will be useful to define the field dependent “regularized propagator”: $$G_{k}[\varphi;\bar{A}]=\left(\Gamma_{k}^{(2;0)}[\varphi;\bar{A}]+R_{k}[\bar{A}]\right)^{-1}\,,\label{gauge_12}$$ to rewrite the flow equation (\[gauge\_10\]) in the following compact form: $$\partial_{t}\Gamma_{k}[\varphi;\bar{A}]=\frac{1}{2}\mathrm{Tr}\, G_{k}[\varphi;\bar{A}]\partial_{t}R_{k}[\bar{A}]\,.\label{gauge_13}$$ The flow equation satisfied by the gEAA is readily obtained combining (\[gauge\_6.2\]) and (\[gauge\_13\]): $$\partial_{t}\bar{\Gamma}_{k}[\bar{A}]=\partial_{t}\Gamma_{k}[0;\bar{A}]=\frac{1}{2}\textrm{Tr}\, G_{k}[0;\bar{A}]\partial_{t}R_{k}[\bar{A}]\,.\label{gauge_14}$$ By construction the flow equation for the gEAA respects gauge symmetry, in the sense that the trace on the rhs of (\[gauge\_14\]), the “beta functional”, is a gauge invariant functional of $\bar{A}_{\mu}$. Still, it is important to realize that equation (\[gauge\_14\]) is not a closed equation since it involves the Hessian of the bEAA $\Gamma_{k}^{(2;0)}[0;\bar{A}]$ taken with respect to the fluctuation field. This fact implies that, even if in the limit $k\rightarrow0$ the gEAA flows to the gauge invariant effective action $\bar{\Gamma}[\bar{A}]\equiv\Gamma[0;\bar{A}]$, for $k\neq0$ it is necessary to consider the flow in the extended theory space where the bEAA lives, composed of all functionals of the fields $\varphi$ and $\bar{A}_{\mu}$ invariant under simultaneous physical and background gauge transformations. When we set $\varphi=0$ the Hessian $\Gamma_{k}^{(2;0)}[0,\bar{A}]$ in (\[gauge\_14\]) becomes “super-diagonal”, since the ghost action is bilinear, and we can immediately perform the multiplet trace in the flow equation. Using the shorthands $\Gamma_{k,aa}=\Gamma_{k}^{(2,0,0;0)}[0,0,0;\bar{A}]$, $\Gamma_{k,\bar{c}c}=\Gamma_{k}^{(0,1,1;0)}[0,0,0;\bar{A}]$, $\Delta S_{k,aa}=\Delta S_{k}^{(2,0,0;0)}[0,0,0;\bar{A}]$ and $\Delta S_{k,\bar{c}c}=\Delta S_{k}^{(0,1,1;0)}[0,0,0;\bar{A}]$ we can write the flow equation in the following matrix form: $$\begin{aligned} \partial_{t}\bar{\Gamma}_{k}[\bar{A}] & = & \frac{1}{2}\textrm{Tr}\left(\begin{array}{ccc} \Gamma_{k,aa}+\Delta S_{k,aa} & 0 & 0\\ 0 & 0 & \Gamma_{k,\bar{c}c}+\Delta S_{k,\bar{c}c}\\ 0 & -\left(\Gamma_{k,\bar{c}c}+\Delta S_{k,\bar{c}c}\right) & 0 \end{array}\right)^{-1}\nonumber \\ & & \times\left(\begin{array}{ccc} \partial_{t}\Delta S_{k,aa} & 0 & 0\\ 0 & 0 & \partial_{t}\Delta S_{k,\bar{c}c}\\ 0 & -\partial_{t}\Delta S_{k,\bar{c}c} & 0 \end{array}\right)\nonumber \\ & = & \frac{1}{2}\textrm{Tr}\left(\begin{array}{ccc} G_{k,aa} & 0 & 0\\ 0 & 0 & -G_{k,\bar{c}c}\\ 0 & G_{k,\bar{c}c} & 0 \end{array}\right)\left(\begin{array}{ccc} \partial_{t}\Delta S_{k,aa} & 0 & 0\\ 0 & 0 & \partial_{t}\Delta S_{k,\bar{c}c}\\ 0 & -\partial_{t}\Delta S_{k,\bar{c}c} & 0 \end{array}\right)\nonumber \\ & = & \frac{1}{2}\textrm{Tr}\left(\begin{array}{ccc} G_{k,aa}\partial_{t}\Delta S_{k,aa} & 0 & 0\\ 0 & G_{k,\bar{c}c}\partial_{t}\Delta S_{k,\bar{c}c} & 0\\ 0 & 0 & G_{k,\bar{c}c}\partial_{t}\Delta S_{k,\bar{c}c} \end{array}\right)\nonumber \\ & = & \frac{1}{2}\textrm{Tr}\, G_{k,aa}\partial_{t}\Delta S_{k,aa}-\textrm{Tr}\, G_{k,\bar{c}c}\partial_{t}\Delta S_{k,\bar{c}c}\,,\label{gauge_14.1}\end{aligned}$$ where we also defined $G_{k,aa}=(\Gamma_{k,aa}+\Delta S_{k,aa})^{-1}$ and $G_{k,\bar{c}c}=(\Gamma_{k,\bar{c}c}+\Delta S_{k,\bar{c}c})^{-1}$. In (\[gauge\_14.1\]) we used the property that the trace over anti-commuting fields carries an additional minus sign. The last line of equation (\[gauge\_14.1\]) gives the flow equation for the gEAA in its commonly used form [@Reuter_Wetterich_1994a]: $$\begin{aligned} \partial_{t}\bar{\Gamma}_{k}[\bar{A}] & = & \frac{1}{2}\textrm{Tr}\left(\Gamma_{k}^{(2,0,0;0)}[0;\bar{A}]+\Delta S_{k}^{(2,0,0;0)}[0,0,0;\bar{A}]\right)^{-1}\partial_{t}\Delta S_{k}^{(2,0,0;0)}[0,0,0;\bar{A}]\nonumber \\ & & -\textrm{Tr}\left(\Gamma_{k}^{(0,1,1;0)}[0;\bar{A}]+\Delta S_{k}^{(0,1,1;0)}[0,0,0;\bar{A}]\right)^{-1}\partial_{t}\Delta S_{k}^{(0,1,1;0)}[0,0,0;\bar{A}]\,.\label{gauge_15}\end{aligned}$$ One can obtain equation (\[gauge\_15\]) as an RG improvement of the one–loop gEAA obtained from (\[gauge\_3\]) by saddle point expansion. One finds: $$\begin{aligned} \bar{\Gamma}_{k}[\bar{A}] & = & S[\bar{A}]+\frac{1}{2}\textrm{Tr}\log\left(S^{(2)}[\bar{A}]+S_{gf}^{(2;0)}[0;\bar{A}]+S_{gh}^{(2,0,0;0)}[0,0,0;\bar{A}]+\Delta S_{k}^{(2,0,0;0)}[0,0,0;\bar{A}]\right)\nonumber \\ & & -\textrm{Tr}\log\left(S_{gh}^{(0,1,1;0)}[0,0,0;\bar{A}]+\Delta S_{k}^{(0,1,1;0)}[0,0,0;\bar{A}]\right)\,;\label{gauge_15.1}\end{aligned}$$ if we now substitute in the traces on the rhs the Hessian of bare action $S[\varphi;\bar{A}]$ with the Hessians of the bEAA $\Gamma_{k}[\varphi;\bar{A}]$ and we differentiate with respect to $t=\log k/k_{0}$ on both sides, we recover the exact RG flow equation (\[gauge\_15\]). We note that one may just substitute $S[\bar{A}]$ with $\bar{\Gamma}_{k}[\bar{A}]$ to obtain directly a closed equation involving only the gEAA; this seams to suggests that truncations where the rEAA is approximated with the classical gauge-fixing and ghost actions may be a consistent approximation when considering the flow of the gEAA. The flow equation for rEAA can be deduced differentiating equation (\[gauge\_6.2\]): $$\partial_{t}\hat{\Gamma}_{k}[\varphi;\bar{A}]=\partial_{t}\bar{\Gamma}_{k}[\bar{A}+a]-\partial_{t}\Gamma_{k}[\varphi;\bar{A}]\,,\label{gauge_16}$$ where we used (\[gauge\_10\]) and (\[gauge\_15\]). Flow equations for the proper-vertices of the bEAA -------------------------------------------------- In this section we derive the system of equations governing the RG flow of the proper-vertices of both the bEAA and the gEAA. To obtain these equations we take functional derivatives of the flow equation (\[gauge\_10\]) satisfied by the bEAA with respect to the fields $\varphi$ and $\bar{A}_{\mu}$. When we differentiate with respect to the background field, we have to remember that the cutoff terms present in the flow equation depend explicitly on it: this adds additional terms to the flow equations for the proper-vertices that are not present in the non-background formalism. We will see that these terms are crucial to preserve gauge covariance of the gEAA along the flow. ### Derivation To start, we take a functional derivative of the flow equation (\[gauge\_10\]) with respect to the fluctuation field multiplet, or with respect to the background field, to obtain the following flow equations for the one-vertices of the bEAA: $$\begin{aligned} \partial_{t}\Gamma_{k}^{(1;0)}[\varphi;\bar{A}] & = & -\frac{1}{2}\textrm{Tr}\, G_{k}[\varphi;\bar{A}]\,\Gamma_{k}^{(3;0)}[\varphi;\bar{A}]G_{k}[\varphi;\bar{A}]\partial_{t}R_{k}[\bar{A}]\nonumber \\ \partial_{t}\Gamma_{k}^{(0;1)}[\varphi;\bar{A}] & = & -\frac{1}{2}\textrm{Tr}\, G_{k}[\varphi;\bar{A}]\left(\Gamma_{k}^{(2;1)}[\varphi;\bar{A}]+R_{k}^{(1)}[\bar{A}]\right)G_{k}[\varphi;\bar{A}]\partial_{t}R_{k}[\bar{A}]\nonumber \\ & & +\frac{1}{2}\textrm{Tr}\, G_{k}[\varphi;\bar{A}]\partial_{t}R_{k}^{(1)}[\bar{A}]\,.\label{gauge_27}\end{aligned}$$ Note that in the second equation of (\[gauge\_27\]), where we differentiated with respect to the background field, there are additional terms containing functional derivatives of the cutoff kernel $R_{k}[\bar{A}]$. Taking further derivatives of equation (\[gauge\_27\]), with respect to both the fluctuation field multiplet and background field, gives the following flow equations for the two-vertices[^4]: $$\begin{aligned} \partial_{t}\Gamma_{k}^{(2;0)} & = & \textrm{Tr}\, G_{k}\,\Gamma_{k}^{(3;0)}G_{k}\,\Gamma_{k}^{(3;0)}G_{k}\partial_{t}R_{k}-\frac{1}{2}\textrm{Tr}\, G_{k}\,\Gamma_{k}^{(4;0)}G_{k}\partial_{t}R_{k}\nonumber \\ \partial_{t}\Gamma_{k}^{(1;1)} & = & \textrm{Tr}\, G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\,\Gamma_{k}^{(3;0)}G_{k}\partial_{t}R_{k}\nonumber \\ & & +\textrm{Tr}\, G_{k}\,\Gamma_{k}^{(3;0)}G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\partial_{t}R_{k}\nonumber \\ & & -\frac{1}{2}\textrm{Tr}\, G_{k}\Gamma_{k}^{(3;1)}G_{k}\partial_{t}R_{k}-\frac{1}{2}\textrm{Tr}\, G_{k}\,\Gamma_{k}^{(3;0)}G_{k}\partial_{t}R_{k}^{(1)}\nonumber \\ \partial_{t}\Gamma_{k}^{(0;2)} & = & \textrm{Tr}\, G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\partial_{t}R_{k}\nonumber \\ & & -\frac{1}{2}\textrm{Tr}\, G_{k}\left(\Gamma_{k}^{(2;2)}+R_{k}^{(2)}\right)G_{k}\partial_{t}R_{k}\nonumber \\ & & -\textrm{Tr}\, G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\partial_{t}R_{k}^{(1)}+\frac{1}{2}\textrm{Tr}\, G_{k}\partial_{t}R_{k}^{(2)}\,.\label{gauge_28}\end{aligned}$$ Proceeding in this way we generate a hierarchy of flow equations for the proper-vertices $\Gamma_{k}^{(n;m)}[\varphi;\bar{A}]$ of the bEAA. In general the flow equation for $\Gamma_{k}^{(n;m)}[\varphi;\bar{A}]$ involves proper-vertices up to $\Gamma_{k}^{(n+2;m)}[\varphi;\bar{A}]$ and functional derivatives of the cutoff kernel up to $R_{k}^{(m)}[\bar{A}]$. Note that, as they stand in (\[gauge\_27\]) and (\[gauge\_28\]), every equation of the hierarchy contains the same information as the original flow equation (\[gauge\_10\]). To profit of the vertex expansion we perform a Taylor expansion of the functional $\Gamma_{k}[\varphi;\bar{A}]$ around $\varphi=0$ and $\bar{A}_{\mu}=0$: $$\Gamma_{k}[\varphi;\bar{A}]=\sum_{n,m=0}^{\infty}\frac{1}{n!m!}\int_{x_{1}...x_{n}y_{1}...y_{m}}\gamma_{k,x_{1}...x_{n}y_{1}...y_{m}}^{(n;m)}\varphi_{x_{1}}...\varphi_{x_{n}}\bar{A}_{y_{1}}...\bar{A}_{y_{m}}\,,\label{gauge_29}$$ where we defined the zero-field proper-vertices as follows: $$\gamma_{k,x_{1}...x_{n}y_{1}...y_{m}}^{(n;m)}\equiv\Gamma_{k}^{(n;m)}[0;0]_{x_{1}...x_{n}y_{1}...y_{m}}\,.\label{gauge_29.1}$$ If we evaluate the hierarchy of flow equations at $\varphi=0$ and $\bar{A}_{\mu}=0$ it becomes an infinite system of coupled integro-differential equations for the zero-field proper-vertices $\gamma_{k}^{(n;m)}$. From the Taylor expansion (\[gauge\_29\]) that defines these vertices, we see that this system can be used to project the RG flow of all terms of the bEAA which are analytic in the fields $\varphi$ and $\bar{A}_{\mu}$. In particular these terms can be of non-local character. Note that in the above considerations is not necessary to expand around the zero-background configuration $\bar{A}_{\mu}=0$; one can choose to expand around any background, preferably where one is able to perform computations. An example is a constant magnetic field configuration or, in the gravitational case, a sphere or an Einstein space. As for the bEAA, we can derive a hierarchy of flow equations for the proper-vertices of the gEAA. In this case the functional depends only on the background field. Taking a functional derivative of (\[gauge\_14\]) with respect to this field gives the following flow equation for the one-vertex of the gEAA: $$\begin{aligned} \partial_{t}\bar{\Gamma}_{k}^{(1)}[\bar{A}] & = & -\frac{1}{2}\textrm{Tr}\, G_{k}[0,\bar{A}]\left(\Gamma_{k}^{(2;1)}[0,\bar{A}]+R_{k}^{(1)}[\bar{A}]\right)G_{k}[0,\bar{A}]\partial_{t}R_{k}[\bar{A}]\nonumber \\ & & +\frac{1}{2}\textrm{Tr}\, G_{k}[0,\bar{A}]\partial_{t}R_{k}^{(1)}[\bar{A}]\,.\label{gauge_30}\end{aligned}$$ Note that since $\partial_{t}\Gamma_{k}^{(0;1)}[0;\bar{A}]=\partial_{t}\bar{\Gamma}_{k}^{(1)}[\bar{A}]$, equation (\[gauge\_30\]) is the same as the second equation in (\[gauge\_27\]), but with $\varphi=0$. A further derivative of (\[gauge\_30\]) with respect to $\bar{A}_{\mu}$ gives the flow equation for the two-vertex of the gEAA: $$\begin{aligned} \partial_{t}\bar{\Gamma}_{k}^{(2)} & = & \textrm{Tr}\, G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\partial_{t}R_{k}\nonumber \\ & & -\frac{1}{2}\textrm{Tr}\, G_{k}\left(\Gamma_{k}^{(2;2)}+R_{k}^{(2)}\right)G_{k}\partial_{t}R_{k}\nonumber \\ & & -\textrm{Tr}\, G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\partial_{t}R_{k}^{(1)}+\frac{1}{2}\textrm{Tr}\, G_{k}\partial_{t}R_{k}^{(2)}\,.\label{gauge_31}\end{aligned}$$ As for (\[gauge\_30\]), this equation is equal to the last equation in (\[gauge\_28\]) if we set $\varphi=0$ and use $\partial_{t}\Gamma_{k}^{(0;2)}[0;\bar{A}]=\partial_{t}\bar{\Gamma}_{k}^{(2)}[\bar{A}]$. As previously stated, the terms coming from functional derivatives of the cutoff kernel (that are present in the background formalism but not in the non-background one) are crucial in preserving the covariant character of the flow of the gEAA and its vertices. As for the bEAA, we can perform a Taylor expansion of the gEAA analogous to (\[gauge\_29\]) and define the zero-field proper-vertices $$\bar{\gamma}_{k,x_{1}...x_{n}}^{(n)}\equiv\bar{\Gamma}_{k}^{(n)}[0]_{x_{1}...x_{n}}\,,\label{gauge_31.1}$$ to turn the hierarchy of flow equations for the proper-vertices of the gEAA into an infinite system of coupled integro-differential equations for the vertices $\bar{\gamma}_{k}^{(m)}$. ### Compact form We introduce now a compact notation to rewrite the flow equations for the proper-vertices just derived. If we introduce the formal operator $$\tilde{\partial}_{t}=(\partial_{t}R_{k}-\eta_{k}R_{k})\frac{\partial}{\partial R_{k}}\,,\label{gauge_32.1}$$ where $\eta_{k}$ is the multiplet matrix of anomalous dimensions, we can rewrite the flow equation for the bEAA (\[gauge\_10\]) as: $$\partial_{t}\Gamma_{k}[\varphi;\bar{A}]=\frac{1}{2}\textrm{Tr}\, G_{k}[\varphi;\bar{A}]\partial_{t}R_{k}[\bar{A}]=-\frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\log G_{k}[\varphi;\bar{A}]\,.\label{gauge_32}$$ In (\[gauge\_32\]) we used the following simple relations: $$\tilde{\partial}_{t}G_{k}=-G_{k}\partial_{t}R_{k}G_{k}\qquad\qquad\tilde{\partial}_{t}\log G_{k}=G_{k}^{-1}\tilde{\partial}_{t}G_{k}=G_{k}\partial_{t}R_{k}\,.$$ In this way, we can rewrite the flow equation for the one-vertices of the bEAA (\[gauge\_27\]) in the following compact form: $$\begin{aligned} \partial_{t}\Gamma_{k}^{(1;0)}[\varphi;\bar{A}] & = & \frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\left\{ \Gamma_{k}^{(3;0)}[\varphi;\bar{A}]G_{k}[\varphi;\bar{A}]\right\} \nonumber \\ \partial_{t}\Gamma_{k}^{(0;1)}[\varphi;\bar{A}] & = & \frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\left\{ \left(\Gamma_{k}^{(2;1)}[\varphi;\bar{A}]+R_{k}^{(1)}[\bar{A}]\right)G_{k}[\varphi;\bar{A}]\right\} \,,\label{gauge_33}\end{aligned}$$ while the flow equations for the two-vertices of the bEAA (\[gauge\_28\]) read now: $$\begin{aligned} \partial_{t}\Gamma_{k}^{(2;0)} & = & -\frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\left\{ \Gamma_{k}^{(3;0)}G_{k}\,\Gamma_{k}^{(3;0)}G_{k}\right\} +\frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\left\{ \Gamma_{k}^{(4;0)}G_{k}\right\} \nonumber \\ \partial_{t}\Gamma_{k}^{(1;1)} & = & -\frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\left\{ \left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\,\Gamma_{k}^{(3;0)}G_{k}\right\} +\frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\left\{ \Gamma_{k}^{(3;1)}G_{k}\right\} \nonumber \\ \partial_{t}\Gamma_{k}^{(0;2)} & = & -\frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\left\{ \left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\left(\Gamma_{k}^{(2;1)}+R_{k}^{(1)}\right)G_{k}\right\} \nonumber \\ & & +\frac{1}{2}\textrm{Tr}\,\tilde{\partial}_{t}\left\{ \left(\Gamma_{k}^{(2;2)}+R_{k}^{(2)}\right)G_{k}\right\} \,.\label{gauge_34}\end{aligned}$$ This compact form turns out to be very useful since the flow equations (\[gauge\_33\]) and (\[gauge\_34\]) contain fewer terms than their counter-parts (\[gauge\_27\]) and (\[gauge\_28\]), and are thus much more manageable when employed in actual computations. The same reasoning applies to all subsequent equations of the hierarchy and extend to the flow equations for the zero-field proper-vertices $\gamma_{k}^{(n;m)}$. The flow equations for the proper-vertices of the gEAA are just those for the bEAA evaluated at $\varphi=0$ and we won’t repeat them here. We will see in the next section how the flow equation for the zero-field proper-vertices just defined can be turned into a powerful computational device to perform computations in the bEAA framework. Diagrammatic and momentum space techniques ------------------------------------------ In this section we introduce a useful diagrammatic representation to organize the various contributions to the flow equations for the zero-field proper-vertices $\gamma_{k}^{(n;m)}$ and we expose the momentum space rules that enables us to calculate these contributions explicitly. As we said before, all these results are valid for any theory with local gauge invariance. For this reason we try to maintain our notation as general as possible. ![Diagrammatic representation of the flow equations for $\partial_{t}\gamma_{k}^{(1;0)}$ and $\partial_{t}\gamma_{k}^{(0;1)}$ as given in (\[gauge\_27\]).](Flow_Eqn_bEAA_1.pdf) As we will see, there are some non-trivial technical steps that have to be made in order to write explicit momentum space flow equations for zero-field proper-vertices with some background legs, i.e. containing the vertices $\gamma^{(n;m)}$ with $m>0$. This issue is related to the dependence of the cutoff kernel $R_{k}[\bar{A}]$ on the cutoff operator which is constructed using the background field. The functional derivatives of the cutoff kernel $R_{k}^{(m)}[\bar{A}]$ can be calculated as terms of a Taylor series expansion of the cutoff kernel with respect to the background field. We will perform this expansion using the perturbative expansion for the (un-traced) heat kernel developed in [@Codello_Zanusso_2012] and reviewed in appendix A. ### Diagrammatic rules We start to introduce the diagrammatic rules used to represent the hierarchy of flow equations for the zero-field proper-vertices. In particular, we show how to represent equations (\[gauge\_27\]) and (\[gauge\_28\]) graphically. The virtue of diagrammatic techniques is that they allow to switch from coordinate space to momentum space straightforwardly. When considering vertices with background lines, it is useful to introduce the “tilde” bEAA[^5] defined by: $$\tilde{\Gamma}_{k}[\varphi;\bar{A}]=\Gamma_{k}[\varphi;\bar{A}]+\Delta S_{k}[\varphi;\bar{A}]\,,\label{gauge_D_1.01}$$ and to define the related “tilde” zero-field proper-vertices $$\tilde{\gamma}_{k}^{(n;m)}=\gamma_{k}^{(n;m)}+\Delta S_{k}^{(n;m)}[0;0]\,.\label{gauge_D_1.02}$$ Obviously, for every $n$ the relation $\tilde{\gamma}_{k}^{(n;0)}=\gamma_{k}^{(n;0)}$ holds. Hence, if we want, we can re-phrase the flow equations for the zero-field proper-vertices solely in terms of the tilde vertices $\tilde{\gamma}^{(n;m)}$. We represent the zero-field regularized propagator $G_{k}[0;0]$ with an internal continuous line, the cutoff insertions $\partial_{t}R_{k}[0]$ are indicated with a crossed circle and the zero-field proper-vertices $\tilde{\gamma}_{k}^{(n;m)}$ are represented as vertices with $n$ external continuous lines (fluctuation legs) and $m$ external thick wavy lines (background legs). Note that $\Delta S_{k}^{(n;m)}[0;0]=0$ if $n>2$ since the cutoff action is quadratic in the fluctuation fields. This diagrammatic rules are summarized graphically as follows: ![Diagrammatic representation of the flow equations for the vertices $\partial_{t}\gamma_{k}^{(2;0)}$, $\partial_{t}\gamma_{k}^{(1;1)}$ and $\partial_{t}\gamma_{k}^{(0;2)}$ as in equation (\[gauge\_28\]).](Flow_Eqn_bEAA_2_bis.pdf) ![image](Feynman_Rules.pdf) Finally, to every closed loop we associate a coordinate or a momentum[^6] $\Omega\int_{q}$ integral ($\Omega$ is the space-time volume), together with the factor $\partial_{t}R_{k}-\eta R_{k}$. Here the anomalous dimension $\eta_{k}$ pertains to the fields present in the cutoff action. The application of these diagrammatic rules to the flow equations (\[gauge\_27\]) for the zero-field one-vertices $\partial_{t}\gamma_{k}^{(1;0)}$ and $\partial_{t}\gamma_{k}^{(0;1)}$ gives the representation of Figure 1, while the flow equations (\[gauge\_28\]) for the zero-field two-vertices $\partial_{t}\gamma_{k}^{(2;0)}$, $\partial_{t}\gamma_{k}^{(1;1)}$ and $\partial_{t}\gamma_{k}^{(0;2)}$ can be represented as in Figure 2. As explained in the previous section, it is sometimes useful to work with the set of flow equations for the bEAA (\[gauge\_33\]) and (\[gauge\_34\]) written employing the formal operator $\tilde{\partial}_{t}$ defined in (\[gauge\_32.1\]). In this case there is no explicit insertion of the cutoff term $\partial_{t}R_{k}[0]$ in the loops, but to every loop is now associated an integration together with the action of this formal operator, i.e. $\int_{x}\tilde{\partial}_{t}$ in coordinate space or $\Omega\int_{q}\tilde{\partial}_{t}$ in momentum space. In this way, the flow equations (\[gauge\_33\]) for the zero-field one-vertices $\partial_{t}\gamma_{k}^{(1;0)}$ and $\partial_{t}\gamma_{k}^{(0;1)}$ can be represented as in Figure 3, while equations (\[gauge\_34\]) for the zero-field two-vertices $\partial_{t}\gamma_{k}^{(2;0)}$, $\partial_{t}\gamma_{k}^{(1;1)}$ and $\partial_{t}\gamma_{k}^{(0;2)}$ can be represented as in Figure 4. Note that in this last representation, the flow equations for the different one-vertices or two-vertices assume a more symmetric form with respect to each other. This reflects in a computational advantage, especially when considering the flow equation for the two-vertex $\gamma_{k}^{(0;2)}$, where the additional terms involving cutoff kernel vertices are accounted for by the use of the tilde vertices and by the action of the operator $\tilde{\partial}_{t}$. We will see in section 3.2.2 that this fact is very useful. ![Diagrammatic representation of the flow equation for the one-vertices $\partial_{t}\gamma_{k}^{(1;0)}$ and $\partial_{t}\gamma_{k}^{(0;1)}$ as given in (\[gauge\_33\]). ](Flow_Eqn_bEAA_1_tilde.pdf) ### Momentum space representation We are now in the position to write down the flow equations for the zero-field proper-vertices $\gamma_{k}^{(n;m)}$ of the bEAA in momentum space. The only non-standard step is to write the momentum space representation of the terms involving functional derivatives of the cutoff kernel $R_{k}[\bar{A}]$ with respect to the background field, i.e. the momentum representation of the vertices $\gamma_{k}^{(2;1)}$ and $\gamma_{k}^{(2;2)}$. The zero-field proper-vertex $\gamma_{k}^{(3;0)}$ is represented graphically by the following diagram: ![image](Vertex_FFF.pdf) Here we use capital letters to indicate general composite indices, in the case of non-abelian gauge theories these have to be interpreted as $A=a\,\alpha,B=b\,\beta,C=c\,\gamma$, while, for example, in the gravitational context they have to interpreted as $A=\alpha\beta,B=\gamma\delta,C=\epsilon\kappa$. Note that each index is associated with a momentum variable, so that $A,B,C$ are the indices of the related momenta $q,p,-q-p$ respectively. Note also that we always define ingoing momenta as being positive. ![Diagrammatic representation of the flow equations for the two-vertices $\partial_{t}\gamma_{k}^{(2;0)}$, $\partial_{t}\gamma_{k}^{(1;1)}$ and $\partial_{t}\gamma_{k}^{(0;2)}$ as in equation (\[gauge\_34\]).](Flow_Eqn_bEAA_2_tilde.pdf) The two-fluctuations one-background zero-field proper-vertex $\tilde{\gamma}_{k}^{(2;1)}=\gamma_{k}^{(2;1)}+\Delta S_{k}^{(2;1)}[0;0]$ is represented graphically by the diagram: ![image](Vertex_FFB.pdf) The non-trivial technical step is to write the explicit momentum space representation for the zero-field proper-vertex $\tilde{\gamma}_{k}^{(2;1)}$. We start by stating the final result and we postpone the details of the derivations to appendix A: $$[\tilde{\gamma}_{q,-q-p,p}^{(2;1)}]^{ABC}=[\gamma_{q,-q-p,p}^{(2;1)}]^{ABC}+[l_{q,-q-p,p}^{(2;1)}]^{ABC}R_{q+p,q}^{(1)}\,,\label{gauge_D_1.1}$$ In (\[gauge\_D\_1.1\]) we introduced the “cutoff operator action” $L[\varphi;\bar{A}]$, defined as that action whose Hessian with respect to $\varphi$ is the cutoff operator and we defined its vertices as $$l_{x_{1}...x_{n}y_{1}...y_{m}}^{(n;m)}\equiv L^{(n;m)}[0;0]_{x_{1}...x_{n}y_{1}...y_{m}}\,.$$ Furthermore, $$R_{q+p,q}^{(1)}\equiv\frac{R_{q+p}-R_{q}}{(q+p)^{2}-q^{2}}\label{gauge_D_1}$$ represents the first finite-difference derivative of the cutoff shape function $R_{q}\equiv R_{k}(q^{2})$. The momentum space representation (\[gauge\_D\_1.1\]) is of crucial importance since it gives, together with the generalization to higher vertices given in the following, access to the computational use of the flow equations for the zero-field proper-vertices $\gamma_{k}^{(n;m)}$. The four-fluctuation vertex $\gamma_{k}^{(4;0)}$ is represented graphically as: ![image](Vertex_FFFF.pdf) Note that here we are giving the four-vertex only for a particular combination of momenta, which is not the most general one, since this will be the case that we will use in section 3. The two-fluctuations two-backgrounds vertex $\tilde{\gamma}_{k}^{(2;2)}=\gamma_{k}^{(2;2)}+\Delta S_{k}^{(2;2)}[0;0]$ is represented instead by the following diagram: ![image](Vertex_FFBB.pdf) The momentum space representation of $\tilde{\gamma}_{k}^{(2;2)}$, derived in appendix B, turns out to be: $$\begin{aligned} [\tilde{\gamma}_{q,-q,p,-p}^{(2;2)}]^{ABCD} & = & [\gamma_{q,-q,p,-p}^{(2;2)}]^{ABC}+[l_{q,-q,p,-p}^{(2;2)}]^{ABCD}R_{q}'\nonumber \\ & & +[l_{q,-q-p,p}^{(2;1)}]^{ABM}[l_{q+p,-q,-p}^{(2;1)}]^{MCD}R_{q+p,q}^{(2)}\,.\label{gauge_D_1.2}\end{aligned}$$ Note that now the cutoff operator action $L[\varphi;\bar{A}]$ enters both as a four-vertex and as a product of two three-vertices, and that this time we need to consider the second finite-difference derivative of the cutoff shape function $R_{q}$: $$R_{q+p,q}^{(2)}\equiv\frac{2}{(q+p)^{2}-q^{2}}\left[\frac{R_{q+p}-R_{q}}{(q+p)^{2}-q^{2}}-R_{q}'\right]\,.\label{gauge_D_2}$$ The version of (\[gauge\_D\_1.2\]) with general momenta is given in appendix B. Equations (\[gauge\_D\_1.1\]) and (\[gauge\_D\_1.2\]) are the example of the kind of relations needed to obtain the explicit momentum space representation of the flow equations for the zero-field proper-vertices of the bEAA. With the technique of appendix B one can find the explicit representation for all the other zero-field proper-vertices $\tilde{\gamma}_{k}^{(n;m)}$. The finite-difference derivatives (\[gauge\_D\_1\]) and (\[gauge\_D\_2\]) can be expanded for small external momenta as follows: $$\begin{aligned} R_{q+p,q}^{(1)} & = & R_{q}'+p\cdot q\, R_{q}''+\frac{1}{2}p^{2}\, R_{q}''+\frac{2}{3}(p\cdot q)^{2}\, R_{q}^{(3)}+O(p^{3})\nonumber \\ R_{q+p,q}^{(2)} & = & R_{q}''+\frac{2}{3}p\cdot q\, R_{q}^{(3)}+\frac{1}{3}p^{2}\, R_{q}^{(3)}+\frac{1}{3}(p\cdot q)^{2}\, R_{q}^{(4)}+O(p^{3})\,.\label{gauge_D_3}\end{aligned}$$ As we will see in section 3, the “correction terms” in (\[gauge\_D\_3\]), proportional to $p$, $p^{2}$, or higher, are those needed to make the flow of the zero-field proper-vertices $\bar{\gamma}_{k}^{(m)}=\gamma_{k}^{(0;m)}$ fully transverse, as they should be by construction. We are now ready to write the flow equations for the zero-field proper-vertices of the bEAA in their full momentum space form. We will present solely the flow equations for the two-vertices, equations (\[gauge\_28\]) and (\[gauge\_34\]), since it is from these equations that in the next section we will extract the beta functions of the gauge coupling. All other equations can be derived using the rules stated in the previous subsection and their generalizations. We find the following momentum space representation for the first equation in (\[gauge\_28\]), describing the flow of the zero-field fluctuation-fluctuation two-vertex: $$\begin{aligned} [\partial_{t}\gamma_{p,-p}^{(2;0)}]^{AB} & = & \Omega\int_{q}(\partial_{t}R_{q}-\eta R_{q})[G_{q}]^{12}[\gamma_{q,-q-p,p}^{(3;0)}]^{2A3}[G_{q+p}]^{34}[\gamma_{q+p,-q,-p}^{(3;0)}]^{4B5}[G_{q}]^{51}\nonumber \\ & & -\frac{1}{2}\Omega\int_{q}(\partial_{t}R_{q}-\eta R_{q})[G_{q}]^{12}[\gamma_{q,-q,p,-p}^{(4;0)}]^{2AB3}[G_{q}]^{31}\,.\label{gauge_D_4}\end{aligned}$$ In (\[gauge\_D\_4\]) and in the following relations $\eta$ is the multiplet matrix of anomalous dimensions of the fluctuation fields in $\varphi$. Note also that we are using the generalized notation for the indices introduced before and we use just integers to denote dummy indices. With respect to the first equation in Figure 2, the first line in (\[gauge\_D\_4\]) is the contribution from the first diagram, while the second line is the contribution from the second one. The second equation in (\[gauge\_28\]), describing the flow of the fluctuation-background zero-field two-vertex, takes the following form: $$\begin{aligned} [\partial_{t}\gamma_{p,-p}^{(1;1)}]^{AB} & = & \Omega\int_{q}(\partial_{t}R_{q}-\eta R_{q})[G_{q}]^{12}[\tilde{\gamma}_{q,-q-p,p}^{(2;1)}]^{2A3}[G_{q+p}]^{34}[\gamma_{q+p,-q,-p}^{(3;0)}]^{4B5}[G_{q}]^{51}\nonumber \\ & & -\frac{1}{2}\Omega\int_{q}(\partial_{t}R_{q}-\eta R_{q})[G_{q}]^{12}[\tilde{\gamma}_{q,-q,p,-p}^{(3;1)}]^{2AB3}[G_{q}]^{31}\nonumber \\ & & -\Omega\int_{q}[l_{q,-q-p,p}^{(2;1)}(\partial_{t}R_{q+p,q}^{(1)}-\eta R_{q+p,q}^{(1)})]^{4A1}\nonumber \\ & & \times[G_{q+p}]^{12}[\gamma_{q+p,-q,-p}^{(3;0)}]^{2B3}[G_{q}]^{34}\,.\label{gauge_D_5}\end{aligned}$$ In (\[gauge\_D\_5\]) there are now two tilde vertices since, referring to the second equation in Figure 2, there is a background line attached to, respectively, a three-vertex, a four-vertex and to the factor $\partial_{t}R_{k}[\bar{A}]$. The contribution from these three diagrams are respectively the first, second and third lines of (\[gauge\_D\_5\]). The last equation of (\[gauge\_28\]) takes the following form: $$\begin{aligned} [\partial_{t}\gamma_{p,-p}^{(0;2)}]^{AB} & = & \Omega\int_{q}(\partial_{t}R_{q}-\eta R_{q})[G_{q}]^{12}[\tilde{\gamma}_{q,-q-p,p}^{(2;1)}]^{2A3}[G_{q+p}]^{34}[\tilde{\gamma}_{q+p,-q,-p}^{(2;1)}]^{4B5}[G_{q}]^{51}\nonumber \\ & & -\frac{1}{2}\Omega\int_{q}(\partial_{t}R_{q}-\eta R_{q})[G_{q}]^{12}[\tilde{\gamma}_{q,-q,p,-p}^{(2;2)}]^{2AB3}[G_{q}]^{31}\nonumber \\ & & -\Omega\int_{q}[l_{q,p,-q-p}^{(2;1)}(\partial_{t}R_{q+p,q}^{(1)}-\eta R_{q+p,q}^{(1)})]^{1A2}[G_{q+p}]^{23}[\tilde{\gamma}_{q+p,-q,-p}^{(2;1)}]^{3B4}[G_{q}]^{41}\nonumber \\ & & +\frac{1}{2}\Omega\int_{q}\left\{ [l_{q,p,-p,-q}^{(2;2)}(\partial_{t}R_{q}'-\eta R_{q}')]^{1AB2}[G_{q}]^{21}\right.\nonumber \\ & & \left.+[l_{q,-q-p,p}^{(2;1)}]^{1A3}[l_{q+p,-q,-p}^{(2;1)}]^{3B2}(\partial_{t}R_{q+p,q}^{(2)}-\eta R_{q+p,q}^{(2)})[G_{q}]^{21}\right\} \,.\label{gauge_D_6}\end{aligned}$$ Equation (\[gauge\_D\_6\]) represents the flow of the zero-field background-background two-vertex of the bEAA and thus every term is written in terms of the tilde vertices and of the cutoff action $L[\varphi;\bar{A}]$. As we explained earlier, these equations are very general and can be adapted to every theory with local gauge symmetry. In terms of the compact representation introduced in subsection 2.2.2 using the formal operator $\tilde{\partial}_{t}$ defined in (\[gauge\_32.1\]), the flow equations for the zero-field two-vertices of the bEAA are given in equation (\[gauge\_34\]). These are represented graphically in Figure 4 and all three have the same overall structure. The flow of the zero-field fluctuation-fluctuation two-vertex has the following momentum space representation: $$\begin{aligned} [\partial_{t}\gamma_{p,-p}^{(2;0)}]^{AB} & = & -\frac{1}{2}\Omega\int_{q}\tilde{\partial}_{t}\left\{ [\gamma_{q,-q-p,p}^{(3;0)}]^{4A1}[G_{q+p}]^{12}[\gamma_{q+p,-q,-p}^{(3;0)}]^{2B3}[G_{q}]^{34}\right\} \nonumber \\ & & +\frac{1}{2}\Omega\int_{q}\tilde{\partial}_{t}\left\{ [\gamma_{q,-q,p,-p}^{(4;0)}]^{1AB2}[G_{q}]^{21}\right\} \,.\label{gauge_D_7}\end{aligned}$$ The second equation in (\[gauge\_34\]), shown in Figure 4, expresses the flow of the zero-field fluctuation-background two-vertex and differs from (\[gauge\_D\_7\]) in two tilde vertices: $$\begin{aligned} [\partial_{t}\gamma_{p,-p}^{(1;1)}]^{AB} & = & -\frac{1}{2}\Omega\int_{q}\tilde{\partial}_{t}\left\{ [\tilde{\gamma}_{q,-q-p,p}^{(2;1)}]^{4A1}[G_{q+p}]^{12}[\gamma_{q+p,-q,-p}^{(3;0)}]^{2B3}[G_{q}]^{34}\right\} \nonumber \\ & & +\frac{1}{2}\Omega\int_{q}\tilde{\partial}_{t}\left\{ [\tilde{\gamma}_{q,-q,p,-p}^{(3;1)}]^{1AB2}[G_{q}]^{21}\right\} \,.\label{gauge_D_8}\end{aligned}$$ Finally, the compact form for the flow of the zero-field background-background two-vertex is: $$\begin{aligned} [\partial_{t}\gamma_{p,-p}^{(0;2)}]^{AB} & = & -\frac{1}{2}\Omega\int_{q}\tilde{\partial}_{t}\left\{ [\tilde{\gamma}_{q,-q-p,p}^{(2;1)}]^{4A1}[G_{q+p}]^{12}[\tilde{\gamma}_{q+p,-q,-p}^{(2;1)}]^{2B3}[G_{q}]^{34}\right\} \nonumber \\ & & +\frac{1}{2}\Omega\int_{q}\tilde{\partial}_{t}\left\{ [\tilde{\gamma}_{q,-q,p,-p}^{(2;2)}]^{1AB2}[G_{q}]^{21}\right\} \,.\label{gauge_D_9}\end{aligned}$$ Note that in equation (\[gauge\_D\_9\]) all zero-field proper-vertices are tilde vertices. As already said, equation (\[gauge\_D\_9\]), as equation (\[gauge\_D\_6\]), represents also the flow of the zero-field proper-vertex $\bar{\gamma}_{k}^{(2)}$ of the gEAA since we have that $\partial_{t}\bar{\gamma}_{k}^{(2)}=\partial_{t}\gamma_{k}^{(0;2)}$. Thus the flow equation for the zero-field proper-vertices of the bEAA are formally as those of the standard EAA when written in terms of the formal operator $\tilde{\partial}_{t}$ but with tilde vertices in place of the standard vertices. All the non-trivial dependence on the cutoff kernel is in this way hidden in the dependence of the tilde vertices on it. This turns out to be a very useful property in actual computations. One can thus naively draw the diagrams as in the non-background formalism with the only caveat that the vertices are tilde vertices and that these can be represented using the rules derived in subsection 2.3.1. Remember also that $\tilde{\gamma}^{(n;0)}=\gamma^{(n;0)}$ and $\bar{\gamma}^{(m)}=\gamma^{(0;m)}.$ Clearly, these are only the first equations of the respective hierarchies and the results exposed in this section are valid for all the subsequent equations of the hierarchy, for both the zero-field proper-vertices of the bEAA and of the gEAA. Equations (\[gauge\_D\_4\]-\[gauge\_D\_9\]), together with the momentum space rules of subsection 2.3.1, are the main result of this paper. A lot of information about the flow of both the bEAA and of the gEAA can be extracted already from the flow of the zero-field two-vertices described by these equations. The results of this section, when combined with the flow equations of the previous one, constitute the basis for a concrete framework in which all truncations of the bEAA which are analytic in the fields can be treated. In particular, one can consider a “curvature expansion” where the gEAA is expanded in powers of the (generalized) curvatures, and where the running is encoded in $k$–dependent structure functions[^7]. As we said, the methods presented in this section can be extended to gravity, non-linear sigma models and membranes. Finally, we remark that truncations like $\bar{\Gamma}_{k}[A]=\int W(F^{2})$, where $W$ is an arbitrary function of the invariant $F^2\equiv F_{\mu\nu}F^{\mu\nu}$, in non-abelian gauge theories [@ReuterWetterich_1994c], or like $\bar{\Gamma}_{k}[g]=\int\sqrt{g}f(R)$, in gravity [@Codello_Percacci_Rahmede_2008], cannot be treated with the present method and one needs to resort to other techniques to treated them. To be more precise, one can treat only polynomial approximations to these functions since we ultimately set the background gauge field, or the background metric, to zero. An application: non-abelian gauge theories ========================================== In this section we apply the formalism developed in section 2 to non-abelian gauge theories with gauge group $SU(N)$. We derive the beta function for the gauge coupling and we show how standard results obtained by heat kernel methods can be recovered in our new framework. Then we calculate the anomalous dimensions of the fluctuation and ghost fields, $\eta_{a,k}$ and $\eta_{c,k}$, and we use them to “close” the beta function of $g_{k}$, obtaining new RG improved forms for this. Finally, we study the phenomenology associated to these RG flows. Truncation ansatz ----------------- As a simple application of the formalism, we consider a truncation ansatz for the gEAA which is the RG improvement of the classical action: $$\bar{\Gamma}_{k}[A]=I[A]\equiv\frac{1}{4}\int d^{d}x\, F_{\mu\nu}^{a}F^{a\mu\nu}\,.\label{gauge_A_1}$$ The quantum gauge field can be expressed in terms of the background and fluctuation fields as follows: $$A_{\mu}=Z_{\bar{A},k}^{1/2}\bar{A}_{\mu}+Z_{a,k}^{1/2}a_{\mu}\,,\label{gauge_A_2}$$ where we rescaled the gauge fields to account for their renormalization. $Z_{a,k}$ is the running wave-function renormalization of the fluctuation field while $Z_{\bar{A},k}$ is the running wave-function renormalization of the background field. In the background field formalism the gauge coupling is related to the wave-function renormalization of the background gauge field [@Abbott_1981]: $$g_{k}=Z_{\bar{A},k}^{-1/2}\,.\label{gauge_A_4}$$ It is thus sufficient to determine the scale dependence of $Z_{\bar{A},k}$ to find the running of $g_{k}$. One defines the anomalous dimension of the background field: $$\eta_{\bar{A},k}=-\partial_{t}\log Z_{\bar{A},k}=-g_{k}^{2}\partial_{t}Z_{\bar{A},k}\,;\label{gauge_A_2.1}$$ then the beta function of the gauge coupling is: $$\partial_{t}g_{k}=\frac{1}{2}\eta_{\bar{A},k}g_{k}\,.\label{gauge_A_2.2}$$ Inserting (\[gauge\_A\_2\]) into equation (\[gauge\_V\_2\]) from appendix B gives the following expansion for the gEAA (\[gauge\_A\_1\]): $$\begin{aligned} \bar{\Gamma}_{k}[Z_{\bar{A},k}^{1/2}\bar{A}+Z_{a,k}^{1/2}a] & = & Z_{\bar{A},k}\bar{\Gamma}_{k}[\bar{A}]+Z_{\bar{A},k}^{1/2}Z_{a,k}^{1/2}\int d^{d}x\,\bar{F}^{a\mu\nu}(\bar{D}_{\mu}a_{\nu})^{a}+\nonumber \\ & & +\frac{1}{2}Z_{a,k}\int d^{d}x\, a_{\mu}^{a}\left[(-\bar{D}^{2})^{ab}g^{\mu\nu}+2f^{abc}F^{c\mu\nu}+\bar{D}^{ac\mu}\bar{D}^{cb\nu}\right]a_{\nu}^{b}+\nonumber \\ & & +g_{k}Z_{a,k}^{3/2}f^{abc}\int d^{d}x\,(\bar{D}_{\mu}a_{\nu})^{a}\, a_{\mu}^{b}a_{\nu}^{c}+\nonumber \\ & & +\frac{1}{4}g_{k}^{2}Z_{a,k}^{2}f^{abc}f^{ade}\int d^{d}x\, a^{b\mu}a^{c\nu}a_{\mu}^{d}a_{\nu}^{e}\,.\label{gauge_A_3}\end{aligned}$$ We used (\[gauge\_A\_4\]) to set $g_{k}Z_{\bar{A},k}^{1/2}=1$ in the covariant derivatives and in the second term of the second line. Quantities with a bar are constructed with the background gauge field of (\[gauge\_A\_3\]). We approximate the rEAA to be the sum of the bare background gauge-fixing action (\[gauge\_5\]) and of the bare background ghost action (\[gauge\_6\]) with rescaled fluctuation and ghost fields[^8]: $$\hat{\Gamma}_{k}[Z_{a,k}^{1/2}a,Z_{c,k}^{1/2}c,Z_{c,k}^{1/2}\bar{c};Z_{\bar{A},k}^{1/2}\bar{A}]=Z_{a,k}S_{gf}[a;\bar{A}]+Z_{c,k}S_{gh}[a,c,\bar{c};\bar{A}]\,.\label{gauge_A_3.1}$$ With these definitions our truncation comprises three running couplings $\{g_{k},Z_{a,k},Z_{c,k}\}$, or three anomalous dimensions $\{\eta_{\bar{A},k},\eta_{a,k},\eta_{c,k}\}$, where the anomalous dimensions of the fluctuation and ghost fields are defined as: $$\eta_{a,k}=-\partial_{t}\log Z_{a,k}\qquad\qquad\eta_{c,k}=-\partial_{t}\log Z_{c,k}\,.\label{gauge_A_3.2}$$ We can say that $g_{k}$ parametrizes the RG evolution of the gEAA, while $Z_{a,k}$ and $Z_{c,k}$ account for the influence of the RG evolution of the full bEAA. This truncation is an example of *bi–field* truncation in the nomenclature of [@Reuter_Bi_Field] since the ansatz has a non-trivial $k$–dependence in both $\bar{\Gamma}_{k}[\bar{A}]$ and $\hat{\Gamma}_{k}[a,\bar{c},c;\bar{A}]$. More general bi–field truncations will include, for example, terms like the fluctuation gluon mass. Even if these truncations can be treated with the present methods we will not treat these cases for the moment. Derivation of the beta functions -------------------------------- In this subsection we calculate the anomalous dimensions of the background, fluctuation and ghost fields. In subsection 3.2.1 we review the standard heat kernel calculation of $\eta_{\bar{A},k}$, while in subsection 3.2.2 we re-derive it by the methods presented in section 2. In subsection 3.2.3 we use the methods of section 2 to calculate $\eta_{a,k}$ and $\eta_{c,k}$. ### Heat kernel calculation of $\eta_{\bar{A},k}$ We review how to calculate the beta function $\partial_{t}Z_{\bar{A},k}$ of the wave-function renormalization of the background field, using the standard method based on the local heat kernel expansion [@Gies_2006; @Reuter_Wetterich_1994a]. If we take a scale derivative of the expansion (\[gauge\_A\_3\]) and we evaluate at $a_{\mu}=0$, we find: $$\partial_{t}\bar{\Gamma}_{k}[Z_{\bar{A},k}^{1/2}\bar{A}]=\partial_{t}Z_{\bar{A},k}\,\frac{1}{4}\int d^{d}x\,\bar{F}_{\mu\nu}^{a}\bar{F}^{a\mu\nu}\,.\label{gauge_ZA_7}$$ From (\[gauge\_ZA\_7\]) we see that to extract $\partial_{t}Z_{\bar{A},k}$ we need to consider the flow equation (\[gauge\_15\]) for the gEAA, which we rewrite here for convenience $$\begin{aligned} \partial_{t}\bar{\Gamma}_{k}[\bar{A}] & = & \frac{1}{2}\textrm{Tr}\left(\Gamma_{k}^{(2,0,0,0)}[0,0,0;\bar{A}]+\Delta S_{k}^{(2,0,0;0)}[0,0,0;\bar{A}]\right)^{-1}\partial_{t}\Delta S_{k}^{(2,0,0;0)}[0,0,0;\bar{A}]\nonumber \\ & & -\textrm{Tr}\left(\Gamma_{k}^{(0,1,1,0)}[0,0,0;\bar{A}]+\Delta S_{k}^{(0,1,1;0)}[0,0,0;\bar{A}]\right)^{-1}\partial_{t}\Delta S_{k}^{(0,1,1;0)}[0,0,0;\bar{A}]\,,\label{gauge_ZA_1}\end{aligned}$$ and extract, from the functional traces on rhs, the coefficient of the invariant $\frac{1}{4}\int\bar{F}^{2}$. As a first step to evaluate the rhs of (\[gauge\_ZA\_1\]), we make the following choices for the cutoff kernels of the fluctuation and ghost fields: $$\begin{aligned} \Delta S_{k}^{(2,0,0;0)}[0,0,0;\bar{A}] & = & Z_{a,k}R_{k}[\bar{A}]\nonumber \\ \Delta S_{k}^{(0,1,1;0)}[0,0,0;\bar{A}] & = & Z_{c,k}R_{k}^{gh}[\bar{A}]\,.\label{gauge_ZA_1.1}\end{aligned}$$ Note that, consistently with the previous discussion, we introduced in the respective cutoff kernels (\[gauge\_ZA\_1.1\]) the wave-function renormalization of the fluctuation fields $Z_{a,k}$ and $Z_{c,k}$. We specify the differential operators that will be the arguments of the cutoff kernels $R_{k}[\bar{A}]$ and $R_{k}^{gh}[\bar{A}]$ in a moment. Second, we calculate the Hessians present in the denominators of the traces on the rhs of the flow equation (\[gauge\_ZA\_1\]). Using the general decomposition of the bEAA given in (\[gauge\_6.2\]), these can be written as follows: $$\begin{aligned} \Gamma_{k}^{(2,0,0;0)}[a,\bar{c},c;\bar{A}] & = & \bar{\Gamma}_{k}^{(2)}[\bar{A}+a]+\hat{\Gamma}_{k}^{(2,0,0;0)}[a,\bar{c},c;\bar{A}]\nonumber \\ \Gamma_{k}^{(0,1,1;0)}[a,\bar{c},c;\bar{A}] & = & \hat{\Gamma}_{k}^{(0,1,1;0)}[a,\bar{c},c;\bar{A}]\,.\label{gauge_ZA_3}\end{aligned}$$ From (\[gauge\_A\_3\]) one can read-off the Hessian of the gEAA: $$\bar{\Gamma}_{k}^{(2)}[\bar{A}+a]^{ab\,\mu\nu}=Z_{a,k}\left[(-D^{2})^{ab}g^{\mu\nu}+2f^{abc}F^{c\mu\nu}+D^{ac\mu}D^{cb\nu}\right]\,,\label{gauge_ZA_3.01}$$ where $(-D^{2})^{ab}\equiv-D_{\mu}^{ac}D^{cb\mu}$. Note that we can write the second term as $2f^{abc}F^{c\mu\nu}=-2(F^{\mu\nu})^{ab}$. The Hessians of the rEAA (\[gauge\_A\_3.1\]) are just the Hessians of the bare background gauge-fixing and background ghost actions: $$\begin{aligned} \hat{\Gamma}_{k}^{(2,0,0;0)}[a,\bar{c},c;\bar{A}] & = & Z_{a,k}S_{gf}^{(2;0)}[a;\bar{A}]\nonumber \\ \hat{\Gamma}_{k}^{(0,1,1;0)}[a,\bar{c},c;\bar{A}] & = & Z_{c,k}S_{gh}^{(0,1,1;0)}[a,\bar{c},c;\bar{A}]\,,\label{gauge_ZA_3.1}\end{aligned}$$ from (\[gauge\_5\]) and (\[gauge\_6\]) we can easily read off the following forms: $$\begin{aligned} \hat{\Gamma}_{k}^{(2,0,0;0)}[a,\bar{c},c;\bar{A}]^{ab\,\mu\nu} & = & -\frac{1}{\alpha}Z_{a,k}\,\bar{D}^{ac\mu}\bar{D}^{cb\nu}\nonumber \\ \hat{\Gamma}_{k}^{(0,1,1;0)}[a,\bar{c},c;\bar{A}]^{ab} & = & -Z_{c,k}\,\bar{D}_{\mu}^{ac}D^{cb\mu}\,.\label{gauge_ZA_4}\end{aligned}$$ In the flow equation (\[gauge\_ZA\_1\]) we need the Hessian (\[gauge\_ZA\_3\]) evaluated at $a_{\mu}=\bar{c}=c=0$; combining (\[gauge\_ZA\_3.01\]) and (\[gauge\_ZA\_4\]) we find: $$\begin{aligned} \Gamma_{k}^{(2,0,0;0)}[0,0,0;\bar{A}]^{ab\,\mu\nu} & = & Z_{a,k}\left[(-\bar{D}^{2})^{ab}g^{\mu\nu}+2f^{abc}\bar{F}^{c\mu\nu}+\left(1-\frac{1}{\alpha}\right)\bar{D}^{ac\mu}\bar{D}^{cb\nu}\right]\nonumber \\ \Gamma_{k}^{(0,1,1;0)}[0,0,0;\bar{A}]^{ab} & = & Z_{c,k}\delta^{ab}(-\bar{D}^{2})\,.\label{gauge_ZA_5}\end{aligned}$$ From now on we set the gauge-fixing parameter to $\alpha=1$ in order to make the first Hessian in (\[gauge\_ZA\_5\]) proportional to the Laplace-type differential operator $$(D_{T}^{\mu\nu})^{ab}=(-D^{2})^{ab}g^{\mu\nu}-2(F^{\mu\nu})^{ab}\,.\label{gauge_ZA_5.1}$$ We need now to choose the cutoff operator, the eigenvalues of which, we compare to the RG scale $k$ to separate the fast field modes from the slow field modes. Without introducing any running coupling in the cutoff action, apart for the wave-function renormalization of the fluctuation fields, there are two possible choices in the gauge sector. Looking back at equation (\[gauge\_ZA\_5\]), we see that we can take as cutoff operator the covariant Laplacian $-\bar{D}^{2}$ or the full Laplace-type differential operator $\bar{D}_{T}$. We call cutoff actions constructed in this way, respectively, type I and type II. In the ghost sector there is no such freedom and we choose as cutoff operator, in both cases, the covariant Laplacian $-\bar{D}^{2}$. We start by considering the type II case. In view of (\[gauge\_ZA\_5\]) we can rewrite the flow equation (\[gauge\_ZA\_1\]) as follows: $$\partial_{t}\bar{\Gamma}_{k}[\bar{A}]=\frac{1}{2}\textrm{Tr}_{1c}\frac{\partial_{t}R_{k}(\bar{D}_{T})-\eta_{a,k}R_{k}(\bar{D}_{T})}{\bar{D}_{T}+R_{k}(\bar{D}_{T})}-\textrm{Tr}_{0c}\frac{\partial_{t}R_{k}(-\bar{D}^{2})-\eta_{c,k}R_{k}(-\bar{D}^{2})}{-\bar{D}^{2}+R_{k}(-\bar{D}^{2})}\,,\label{gauge_ZA_6}$$ where we emphasized that the traces are also over space-time as well as color indices. We proceed by employing the local heat kernel expansion (for the operators $\bar{D}_{T}$ and $-\bar{D}^{2}$) to expand the traces on the rhs side of equation (\[gauge\_ZA\_6\]) in terms of gauge invariant operators. The functional trace of a Laplace-type second order differential operator $\mathcal{O}=-D^{2}+U$ can be expanded as: $$\textrm{Tr}\, h(\mathcal{O})=\frac{1}{(4\pi)^{d/2}}\sum_{n=0}^{\infty}B_{2n}(\Delta)Q_{\frac{d}{2}-n}[h]\,,\label{gauge_ZA_6.1}$$ where the first heat kernel coefficients $B_{2n}(\mathcal{O})$ are: $$\begin{aligned} B_{0}(\mathcal{O}) & = & \int d^{d}x\,\textrm{tr}\,1\nonumber \\ B_{2}(\mathcal{O}) & = & \int d^{d}x\,\textrm{tr}\, U\nonumber \\ B_{4}(\mathcal{O}) & = & \int d^{d}x\,\textrm{tr}\left(\frac{1}{2}U^{2}+\frac{1}{6}D^{2}U+\frac{1}{12}[D_{\mu},D_{\nu}][D^{\mu},D^{\nu}]\right)\,,\label{gauge_ZA_6.2}\end{aligned}$$ and the $Q$-functionals are defined as: $$Q_{n}[h]=\left\{ \begin{array}{ccc} \frac{1}{\Gamma(n)}\int_{0}^{\infty}dz\, z^{n-1}h(z) & & n>0\\ (-1)^{n}h^{(n)}(0) & & n\leq0 \end{array}\right..\label{gauge_ZA_6.3}$$ For more details on the applications of the heat kernel expansion to the calculation of functional traces see [@Codello_Percacci_Rahmede_2009]. The invariant $\frac{1}{4}\int\bar{F}^{2}$ is contained in the heat kernel coefficients $B_{4}(\bar{D}_{T})$ and $B_{4}(-\bar{D}^{2})$; a use of equation (\[gauge\_ZA\_6.1\]) gives the following: $$\begin{aligned} \left.\partial_{t}\bar{\Gamma}_{k}[\bar{A}]\right|_{\int\bar{F}^{2}} & = & \frac{1}{(4\pi)^{d/2}}\left\{ \frac{1}{2}B_{4}(\bar{D}_{T})Q_{\frac{d}{2}-2}\left[\left(\partial_{t}R_{k}-\eta_{a,k}R_{k}\right)G_{k}\right]\right.\nonumber \\ & & \left.-B_{4}(-\bar{D}^{2})Q_{\frac{d}{2}-2}\left[\left(\partial_{t}R_{k}-\eta_{c,k}R_{k}\right)G_{k}\right]\right\} \,,\label{gauge_ZA_7.1}\end{aligned}$$ where we introduced the regularized propagator $$G_{k}(z)=\frac{1}{z+R_{k}(z)}\,.\label{gauge_ZA_7.2}$$ For the differential operator $\bar{D}_{T}$ we find the following heat kernel coefficient[^9]: $$\begin{aligned} B_{4}(\bar{D}_{T}) & = & \int d^{d}x\left\{ \frac{1}{2}\textrm{tr}\, U^{2}+\frac{1}{12}\textrm{tr}\,[D_{\mu},D_{\nu}][D^{\mu},D^{\nu}]\right\} \nonumber \\ & = & \int d^{d}x\left\{ \frac{1}{2}\left(2f^{abc}\bar{F}^{c\mu\nu}\right)\left(2f^{bad}\bar{F}_{\nu\mu}^{d}\right)+\frac{1}{12}\left(-f^{abc}\bar{F}_{\mu\nu}^{c}\right)\left(-f^{bad}\bar{F}_{\mu\nu}^{d}\right)\right\} \nonumber \\ & = & \frac{24-d}{12}N\int d^{d}x\,\bar{F}_{\mu\nu}^{a}\bar{F}^{a\mu\nu}\,,\label{gauge_ZA_8}\end{aligned}$$ where we used $U^{ab\mu\nu}=2f^{abc}\bar{F}^{c\mu\nu}$, $[\bar{D}_{\mu},\bar{D}_{\nu}]^{ab}=-f^{abc}\bar{F}_{\mu\nu}^{c}$ and $f^{abc}f^{abd}=N\delta^{cd}$. The ghost operator $-\bar{D}^{2}$, i.e. the Laplacian, has instead the following heat kernel coefficient: $$B_{4}(-\bar{D}^{2})=\int d^{d}x\left\{ \frac{1}{12}\textrm{tr}\,[D_{\mu},D_{\nu}][D^{\mu},D^{\nu}]\right\} =-\frac{N}{12}\int d^{d}x\,\bar{F}_{\mu\nu}^{a}\bar{F}^{a\mu\nu}\,.\label{gauge_ZA_9}$$ Inserting (\[gauge\_ZA\_8\]) and (\[gauge\_ZA\_9\]) in (\[gauge\_ZA\_7.1\]), comparing with (\[gauge\_ZA\_7\]) and using (\[gauge\_A\_2.1\]) finally gives: $$\eta_{\bar{A},k}=-\frac{g_{k}^{2}N}{(4\pi)^{d/2}}\left\{ \frac{24-d}{6}Q_{\frac{d}{2}-2}\left[\left(\partial_{t}R_{k}-\eta_{a,k}R_{k}\right)G_{k}\right]+\frac{1}{3}Q_{\frac{d}{2}-2}\left[\left(\partial_{t}R_{k}-\eta_{c,k}R_{k}\right)G_{k}\right]\right\} \,.\label{gauge_ZA_10}$$ Equation (\[gauge\_ZA\_10\]) gives the type II anomalous dimension of the background field, within the truncation specified by equations (\[gauge\_A\_3\]) and (\[gauge\_A\_3.1\]), in the gauge $\alpha=1$, in arbitrary dimension and for general cutoff shape function $R_{k}(z)$. When instead we employ the type I cutoff, the gauge trace in the flow equation (\[gauge\_ZA\_6\]) becomes: $$\frac{1}{2}\textrm{Tr}_{1c}\frac{\partial_{t}R_{k}(-\bar{D}^{2})-\eta_{a,k}R_{k}(-\bar{D}^{2})}{-\bar{D}^{2}-2\bar{F}+R_{k}(-\bar{D}^{2})}\,.\label{gauge_ZA_12}$$ To collect all terms proportional to $\frac{1}{4}\int\bar{F}^{2}$ we expand the denominator in powers of the curvature: $$\begin{aligned} \left[-\bar{D}^{2}-2\bar{F}+R_{k}(-\bar{D}^{2})\right]_{\mu\nu}^{-1} & = & g_{\mu\nu}G_{k}(-\bar{D}^{2})+2\, G_{k}(-\bar{D}^{2})\bar{F}_{\mu\nu}\, G_{k}(-\bar{D}^{2})\nonumber \\ & & +4G_{k}(-\bar{D}^{2})\bar{F}_{\mu\alpha}\, G_{k}(-\bar{D}^{2})\bar{F}_{\;\nu}^{\alpha}\, G_{k}(-\bar{D}^{2})+O(\bar{F}^{3})\,.\label{gauge_ZA_13}\end{aligned}$$ When we insert back the expansion (\[gauge\_ZA\_13\]) in the trace (\[gauge\_ZA\_12\]), the third term we obtain is already proportional to $\frac{1}{4}\int\bar{F}^{2}$: $$\left.\textrm{Tr}_{1c}\left\{ \left[\partial_{t}R_{k}(-\bar{D}^{2})-\eta_{a,k}R_{k}(-\bar{D}^{2})\right]G_{k}^{3}(-\bar{D}^{2})\bar{F}_{\mu\alpha}\bar{F}_{\;\nu}^{\alpha}\right\} \right|_{\int\bar{F}^{2}}=\qquad\qquad$$ $$\qquad\qquad=\frac{1}{(4\pi)^{d/2}}Q_{\frac{d}{2}}\left[(\partial_{t}R_{k}-\eta_{a,k}R_{k})G_{k}^{3}\right]\left(N\int d^{d}x\,\bar{F}_{\mu\nu}^{a}\bar{F}^{a\mu\nu}\right)\,,\label{gauge_ZA_13.1}$$ the second term we obtain is zero when traced over space-time indices, while the first term we obtain generates a contribution proportional to $\frac{1}{4}\int\bar{F}^{2}$ when expanded using the heat kernel: $$\left.\textrm{Tr}_{1c}\left[\left(\partial_{t}R_{k}(-\bar{D}^{2})-\eta_{a,k}R_{k}(-\bar{D}^{2})\right)G_{k}(-\bar{D}^{2})\right]\right|_{\int\bar{F}^{2}}=\qquad\qquad$$ $$\qquad\qquad=\frac{1}{(4\pi)^{d/2}}Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta_{a,k}R_{k})G_{k}\right]\left(-\frac{d}{12}N\int d^{d}x\,\bar{F}_{\mu\nu}^{a}\bar{F}^{a\mu\nu}\right)\,.\label{gauge_ZA_13.2}$$ Note that with respect to (\[gauge\_ZA\_9\]) the heat kernel coefficient in (\[gauge\_ZA\_13.2\]) contains an additional factor $d$, since the trace is over spin one fields. Combining the contributions (\[gauge\_ZA\_13\]) and (\[gauge\_ZA\_13.2\]) gives the type I anomalous dimension of the background field: $$\begin{aligned} \eta_{\bar{A},k} & = & -\frac{g_{k}^{2}N}{(4\pi)^{d/2}}\left\{ -\frac{d}{6}Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta_{a,k}R_{k})G_{k}\right]+8Q_{\frac{d}{2}}\left[(\partial_{t}R_{k}-\eta_{a,k}R_{k})G_{k}^{3}\right]\right.\nonumber \\ & & \left.+\frac{1}{3}Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta_{c,k}R_{k})G_{k}\right]\right\} \,.\label{gauge_ZA_14}\end{aligned}$$ As (\[gauge\_ZA\_10\]), this relation is valid within the truncation (\[gauge\_A\_3\]) and (\[gauge\_A\_3.1\]), in the gauge $\alpha=1$, in arbitrary dimension and for general cutoff shape function $R_{k}(z)$. Equation (\[gauge\_ZA\_10\]) and (\[gauge\_ZA\_14\]) are the main results of this subsection and represent the RG running of the wave-function renormalization of the background field in the two schemes, type I and II. In subsection 3.3 we will specify the cutoff shape function $R_{k}(z)$ in order to obtain explicit forms for the anomalous dimensions. As a final comment, we remark that to calculate $\partial_{t}Z_{\bar{A},k}$ for general values of the gauge-fixing parameter $\alpha$ (possibly scale dependent) using the heat kernel expansion, we need to be able to evaluate functional traces containing the following non-minimal operator: $$(D_{T}^{\mu\nu})^{ab}=(-D^{2})^{ab}g^{\mu\nu}+\left(1-\frac{1}{\alpha}\right)D^{ac\mu}D^{cb\nu}-2(F^{\mu\nu})^{ab}\,.$$ A way to handle this is to use the re-summation technique proposed in [@Benedetti_Groh_Machado_Saueressig_2010]; otherwise, one can perform the calculation by decomposing the gauge field into its spin components as was done in [@ReuterWetterich_1994c]. ### Calculation of $\eta_{\bar{A},k}$ from $\partial_{t}\bar{\gamma}_{k}^{(2)}$ We show how to calculate the anomalous dimension of the background field from the flow equation for the zero-field two-point function $\bar{\gamma}_{k}^{(2)}$ of the gEAA employing the techniques introduced in section 2. In particular, we are going to work with the flow equation for $\bar{\gamma}_{k}^{(2)}$ in its compact form given in equation (\[gauge\_D\_9\]). After the field multiplet decomposition, this equation can be represented diagrammatically as shown in Figure 5. In formulas, we can write the flow equation as[^10]: $$\begin{aligned} [\partial_{t}\bar{\gamma}_{p,-p}^{(2)}]^{mn\,\mu\nu} & = & -\frac{1}{2}\Omega\int_{q}\tilde{\partial}_{t}[a_{p,q}]^{mn\,\mu\nu}+\frac{1}{2}\Omega\int_{q}\tilde{\partial}_{t}[b_{p,q}]^{mn\,\mu\nu}+\Omega\int_{q}\tilde{\partial}_{t}[c_{p,q}]^{mn\,\mu\nu}\nonumber \\ & & -\Omega\int_{q}\tilde{\partial}_{t}[d_{p,q}]^{mn\,\mu\nu}\,.\label{gauge_ZA_22.1}\end{aligned}$$ Here $\Omega=(2\pi)^{d}\delta(0)$ is a volume factor coming from the momentum integrals. Tensor products in the integrands are: ![Diagrammatic representation of the compact form of the flow equation for the zero-field two-point function $\bar{\gamma}_{k}^{(2)}$ of the gEAA after the field multiplet decomposition within the truncation specified by equations (\[gauge\_A\_3\]) and (\[gauge\_A\_3.1\]). Thick wavy lines represent the background field $\bar{A}_{\mu}$, light wavy lines represent the fluctuation field $a_{\mu}$, while dotted lines represent the ghost fields $\bar{c}$ and $c$. ](Flow_Eqn_gEAA_2_bis.pdf) $$\begin{aligned} [a_{p,q}]^{mn\,\mu\nu} & = & [\tilde{\gamma}_{q,-q-p,p}^{(2,0,0;1)}]^{abm\,\alpha\beta\mu}[G_{q+p}]^{bc\,\beta\gamma}[\tilde{\gamma}_{q+p,-q,-p}^{(2,0,0;1)}]^{cdn\,\gamma\delta\nu}[G_{q}]^{da\,\delta\alpha}\nonumber \\ {}[b_{p,q}]^{mn\,\mu\nu} & = & [\tilde{\gamma}_{q,-q,p,-p}^{(2,0,0;2)}]^{abmn\,\alpha\beta\mu\nu}[G_{q}]^{ba\,\beta\alpha}\nonumber \\ {}[c_{p,q}]^{mn\,\mu\nu} & = & [\tilde{\gamma}_{q,-q-p,p}^{(0,1,1;1)}]^{abm\,\mu}[G_{q+p}^{c}]^{bc}[\tilde{\gamma}_{q+p,-q,-p}^{(0,1,1;1)}]^{cdn\,\nu}[G_{q}^{c}]^{da}\nonumber \\ {}[d_{p,q}]^{mn\,\mu\nu} & = & [\tilde{\gamma}_{q,-q,p,-p}^{(0,1,1;2)}]^{abmn\,\mu\nu}[G_{q}^{c}]^{ba}\,.\label{gauge_ZA_22.2}\end{aligned}$$ The “tilde” gauge vertices in (\[gauge\_ZA\_22.2\]) are the following: $$\begin{aligned} [\tilde{\gamma}_{q,-q-p,p}^{(2,0,0;1)}]^{abm\,\alpha\beta\mu} & = & Z_{a}[I_{q,-q-p,p}^{(3)}]^{abm\,\alpha\beta\mu}+Z_{a}[S_{gf\, q,-q-p,p}^{(2;1)}]^{abm\,\alpha\beta\mu}\nonumber \\ & & +Z_{a}[l_{q,-q-p,p}^{(2;1)}]^{abm\,\alpha\beta\mu}R_{q+p,p}^{(1)}\,,\label{gauge_ZA_22.3}\end{aligned}$$ (note that $\tilde{\gamma}_{q,-q-p,p}^{(2,0,0;1)}=\tilde{\gamma}_{q+p,-q,-p}^{(2,0,0;1)}$) and: $$\begin{aligned} [\tilde{\gamma}_{q,-q,p,-p}^{(2,0,0;2)}]^{abmn\,\alpha\beta\mu\nu} & = & Z_{a}[I_{q,-q,p,-p}^{(4)}]^{abmn\,\alpha\beta\mu\nu}+Z_{a}[S_{gf\, q,-q,p,-p}^{(2;2)}]^{abmn\,\alpha\beta\mu\nu}\nonumber \\ & & +Z_{a}[l_{q,-q,p,-p}^{(2;2)}]^{abmn\,\alpha\beta\mu\nu}R'_{q}\nonumber \\ & & +Z_{a}[l_{q,-q-p,p}^{(2;1)}]^{abc\,\alpha\beta\gamma}[l_{q+p,-q,-p}^{(2;1)}]^{cmn\,\gamma\mu\nu}R_{q+p,q}^{(2)}\,.\label{gauge_ZA_22.4}\end{aligned}$$ The “tilde” ghost vertices are instead: $$[\tilde{\gamma}_{q,-q-p,p}^{(0,1,1;1)}]^{abm\,\mu}=Z_{c}[S_{gh\, q,-q-p,p}^{(0,1,1;1)}]^{abm\,\mu}+Z_{c}[l_{q,-q-p,p}^{(1,1;1)}]^{abm\,\mu}R_{q+p,q}^{(1)}\label{gauge_ZA_22.5}$$ and $$\begin{aligned} [\tilde{\gamma}_{q,-q,p,-p}^{(0,1,1;2)}]^{abmn\,\mu\nu} & = & Z_{c}[S_{gh\, q,-q,p,-p}^{(0,1,1;2)}]^{abmn\,\mu\nu}+Z_{c}[l_{q,-q,p,-p}^{(1,1;2)}]^{abmn\,\mu\nu}R'_{q}\nonumber \\ & & +Z_{c}[l_{q,-q-p,p}^{(1,1;1)}]^{abc\,\mu}[l_{q+p,-q,-p}^{(1,1;1)}]^{cmn\,\nu}R_{q+p,q}^{(2)}\,.\label{gauge_ZA_22.6}\end{aligned}$$ The momentum space representations of the vertices of the functionals $I$, $S_{gf}$ and $S_{gh}$ appearing in these equations can be found in appendix B. The vertices $l$ of the cutoff action will be given after we make the cutoff action specifications. We will soon see that these vertices, when reinserted in equation (\[gauge\_ZA\_22.1\]), will correctly conspire to make the flow of $\partial_{t}\bar{\gamma}^{(2)}$ correctly transverse. To proceed we first need to construct the regularized propagators that enter the integrands in (\[gauge\_ZA\_22.2\]). The fluctuation field regularized propagator is: $$[G_{p}^{-1}]^{\alpha\beta\, ab}=[\gamma_{p,-p}^{(2,0,0;0)}+R_{p}]^{\alpha\beta\, ab}\,.\label{gauge_P_0}$$ Within our truncation, given by equations (\[gauge\_A\_3\]) and (\[gauge\_A\_3.1\]), we have: $$\begin{aligned} [\gamma_{p,-p}^{(2,0,0;0)}]^{ab\,\alpha\beta} & = & [\bar{\gamma}_{p,-p}^{(2)}+\hat{\gamma}_{p,-p}^{(2,0,0;0)}]^{ab\,\alpha\beta}\nonumber \\ & = & Z_{a}[I_{p,-p}^{(2)}]^{ab\,\alpha\beta}+Z_{a}[S_{gf\, p,-p}^{(2;0)}]^{ab\,\alpha\beta}\nonumber \\ & = & \Omega\, Z_{a}\,\delta^{ab}\, p^{2}\left[(1-P)^{\alpha\beta}+\frac{1}{\alpha}P^{\alpha\beta}\right],\label{gauge_P_1}\end{aligned}$$ where we used the two-vertices of $I$ and $S_{gf}$ from appendix B and we introduced the longitudinal $P^{\alpha\beta}=p^{\alpha}p^{\beta}/p^{2}$ and transverse $(1-P)^{\alpha\beta}$ projectors. Inserting (\[gauge\_P\_1\]) in (\[gauge\_P\_0\]) gives the following form: $$[G_{p}^{-1}]^{ab\,\alpha\beta}=\Omega\, Z_{a}\,\delta^{ab}\, p^{2}\,\left[(1-P)^{\alpha\beta}+\frac{1}{\alpha}P^{\alpha\beta}\right]+[R_{p}]^{ab\,\alpha\beta}\,.\label{gauge_P_2}$$ We need now to specify the tensor structure of the cutoff kernel; there are two basic choices: $$\begin{aligned} [R_{p}]^{ab\,\alpha\beta} & = & \Omega\, Z_{a}\,\delta^{ab}\left[(1-P)^{\alpha\beta}+\frac{1}{\alpha}P^{\alpha\beta}\right]R_{p}\label{gauge_P_3}\\ {}[R_{p}]^{ab\,\alpha\beta} & = & \Omega\, Z_{a}\,\delta^{ab}g^{\alpha\beta}R_{q}\,.\label{gauge_P_4}\end{aligned}$$ Using the following relation that can be easily verified, $$\left[\mathbf{1}a+(\mathbf{1}-\mathbf{P})b+\mathbf{P}c\right]^{-1}=\frac{1}{a+b}(\mathbf{1}-\mathbf{P})+\frac{1}{a+c}\mathbf{P}\,,$$ we can invert equation (\[gauge\_P\_2\]) to obtain the explicit expression for the regularized propagator. In the case that we are using the cutoff kernel as defined in (\[gauge\_P\_3\]), we find the form: $$[G_{p}]^{ab\,\alpha\beta}=(\Omega\, Z_{a})^{-1}\left[\delta^{ab}\frac{1}{p^{2}+R_{p}}(1-P)^{\alpha\beta}+\delta^{ab}\frac{\alpha}{p^{2}+R_{p}}P^{\alpha\beta}\right]\,,\label{gauge_P_5}$$ while in the case we are using the cutoff kernel as defined in (\[gauge\_P\_4\]), we get instead: $$[G_{p}]^{ab\,\alpha\beta}=(\Omega\, Z_{a})^{-1}\left[\delta^{ab}\frac{1}{p^{2}+R_{p}}(1-P)^{\alpha\beta}+\delta^{ab}\frac{\alpha}{p^{2}+\alpha R_{p}}P^{\alpha\beta}\right]\,.\label{gauge_P_6}$$ In this paper we will consider the choice corresponding to (\[gauge\_P\_4\]), since this is the minimal cutoff kernel choice we can make[^11]. It’s useful to define the transverse and longitudinal regularized propagators: $$G_{T,k}(z)=\frac{1}{z+R_{k}(z)}\qquad\qquad G_{L,k}(z)=\frac{\alpha}{z+\alpha R_{k}(z)}\,.\label{gauge_P_7}$$ Note that the transverse regularized propagator is equal to the regularized propagator defined in (\[gauge\_ZA\_7.2\]), i.e. $G_{T,k}(z)=G_{k}(z)$. We can now write (\[gauge\_P\_6\]) as follows: $$[G_{p}]^{ab\,\alpha\beta}=(\Omega\, Z_{a})^{-1}\left[\delta^{ab}(1-P)^{\alpha\beta}G_{p}^{T}+\delta^{ab}P^{\alpha\beta}G_{p}^{L}\right]\,.\label{gauge_P_7.3}$$ Note that $G_{L,k}(z)=0$ if $\alpha=0$ and that $G_{L,k}(z)=G_{T,k}(z)=G_{k}(z)$ if $\alpha=1$; from now on we set $\alpha=1$ so that the tensor structure of the regularized propagator is proportional to the identity. The ghost regularized propagator, defined by $$[G_{p}^{-1}]^{ab}=[\gamma_{p,-p}^{(0,1,1;0)}+R_{p}]^{ab}\,,$$ is easily obtained from $$[\gamma_{p,-p}^{(0,1,1;0)}]^{ab}=Z_{c}[S_{gh\; p,-p}^{(0,1,1;0)}]^{ab}=\Omega\, Z_{c}\,\delta^{ab}p^{2}\,,\label{gauge_P_8}$$ together with the minimal cutoff kernel choice $$[R_{p}^{c}]^{ab}=\Omega\, Z_{c}\,\delta^{ab}R_{p}\,,\label{gauge_P_9}$$ in the form: $$[G_{p}^{c}]^{ab}=(\Omega\, Z_{c})^{-1}\delta^{ab}\frac{1}{p^{2}+R_{p}}=(\Omega\, Z_{c})^{-1}\delta^{ab}G_{p}\,.\label{gauge_P_10}$$ This completes the construction of the regularized propagators needed in the flow equation (\[gauge\_ZA\_22.1\]). We need now to specify the cutoff operator we employ to cutoff the field modes. Remember from section 2 that the the cutoff operator action, here $L[a;\bar{A}]$ in the gauge sector and $L[\bar{c},c;\bar{A}]$ in the ghost sector, is just the action whose Hessian is the cutoff operator. As in the previous subsection, there are two basic options in the gauge sector, which we called type I and type II, where the cutoff operator is taken to be, respectively, the gauge Laplacian $-\bar{D}^{2}$ or the operator $\bar{D}_{T}$. We start to consider the type I case, where the gauge cutoff operator action is the following: $$L[a;\bar{A}]=\frac{1}{2}\int d^{d}x\,\bar{D}_{\mu}a_{\nu}\bar{D}^{\mu}a^{\nu}\,;\label{gauge_CA_1}$$ one can easily check that the Hessian of this action is the gauge Laplacian: $$L^{(2;0)}[0;\bar{A}]_{xy}^{ab\,\alpha\beta}=g^{\alpha\beta}\int d^{d}z\,\bar{D}_{z\mu}^{ac}\delta_{zx}\bar{D}_{z}^{cb\mu}\delta_{zy}=-g^{\alpha\beta}\bar{D}_{x\mu}^{ac}\bar{D}_{y}^{cb\mu}\delta_{xy}=(-\bar{D}_{x}^{2})^{ab}g^{\alpha\beta}\delta_{xy}\,.\label{gauge_CA_2}$$ The ghost cutoff operator is the gauge Laplacian in both type I and II cases; the ghost cutoff operator action is then simply[^12]: $$L[\bar{c},c;\bar{A}]=\int d^{d}x\,\bar{D}_{\mu}\bar{c}\,\bar{D}^{\mu}c\,.\label{gauge_CA_2.1}$$ Again we have $L^{(1,1;0)}[0,0;\bar{A}]_{xy}^{ab}=(-\bar{D}_{x}^{2})^{ab}\delta_{xy}$. The vertices of the actions (\[gauge\_CA\_1\]) and (\[gauge\_CA\_2.1\]) can be found in appendix B. Equation (\[gauge\_ZA\_22.1\]) is a tensor equation from which we can obtain two independent scalar equations after we contract it with the projectors $(1-P)^{\mu\nu}$ and $P^{\mu\nu}$. We start by considering the projection in the transverse direction; in particular, we find the following results for the integrands (\[gauge\_ZA\_22.2\]): $$\begin{aligned} (1-P)_{\mu\nu}[a_{p,q}]^{mn\,\mu\nu} & = & (\Omega\, Z_{a})^{-2}(1-P)_{\mu\nu}[\tilde{\gamma}_{q,-q-p,p}^{(2,0,0;1)}]^{abm\,\alpha\beta\mu}[\tilde{\gamma}_{q+p,-q,-p}^{(2,0,0;1)}]^{ban\,\beta\alpha\nu}G_{q}G_{q+p}\nonumber \\ & = & N\delta^{mn}\left\{ 8(d-1)p^{2}-4dq^{2}(1-x^{2})\left(1+R_{q+p,q}^{(1)}\right)^{2}\right\} G_{q}G_{q+p}\nonumber \\ \nonumber \\ (1-P)_{\mu\nu}[b_{p,q}]^{mn\,\mu\nu} & = & (\Omega\, Z_{a})^{-1}(1-P)_{\mu\nu}[\tilde{\gamma}_{q,-q,p,-p}^{(2,0,0;2)}]^{aamn\,\alpha\alpha\mu\nu}G_{q}\nonumber \\ & = & N\delta^{mn}\left\{ 2d(d-1)(1+R_{q}')+4dq^{2}(1-x^{2})R_{q+p,q}^{(2)}\right\} G_{q}\nonumber \\ \nonumber \\ (1-P)_{\mu\nu}[c_{p,q}]^{mn\,\mu\nu} & = & (\Omega\, Z_{c})^{-2}(1-P)_{\mu\nu}[\tilde{\gamma}_{q,-q-p,p}^{(0,1,1;1)}]^{abm\,\mu}[\tilde{\gamma}_{q+p,-q,-p}^{(0,1,1;1)}]^{ban\,\nu}G_{q}G_{q+p}\nonumber \\ & = & N\delta^{mn}\left\{ 4q^{2}(1-x^{2})\left(1+R_{q+p,q}^{(1)}\right)^{2}\right\} G_{q}G_{q+p}\nonumber \\ \nonumber \\ (1-P)_{\mu\nu}[d_{p,q}]^{mn\,\mu\nu} & = & (\Omega\, Z_{c})^{-1}(1-P)_{\mu\nu}[\tilde{\gamma}_{q,-q,p,-p}^{(0,1,1;2)}]^{aamn\,\mu\nu}G_{q}\nonumber \\ & = & N\delta^{mn}\left\{ 2(d-1)(1+R_{q}')+4q^{2}(1-x^{2})R_{q+p,q}^{(2)}\right\} G_{q}\,.\label{gauge_ZA_22.7}\end{aligned}$$ In the first line of every contraction we used the fact that in the gauge $\alpha=1$ the regularized propagators are proportional to the identity, while in the second line of every contraction we performed the tensor products between the “tilde” vertices. Here $x$ is the cosine of the angle between the vectors $p$ and $q$. We evaluated group factors as usual using $f^{abc}f^{abd}=N\delta^{cd}$. Note also that all factors $\Omega\, Z_{a}$ and $\Omega\, Z_{c}$ simplified in the final result. The transverse component of the lhs of the flow equation (\[gauge\_ZA\_22.1\]) is: $$(1-P)_{\alpha\beta}[\partial_{t}\bar{\gamma}_{p,-p}^{(2)}]^{ab\,\alpha\beta}=\Omega\,\delta^{ab}(d-1)\partial_{t}Z_{\bar{A}}\, p^{2}\,;\label{gauge_ZA_23}$$ inserting the contractions (\[gauge\_ZA\_22.7\]) into the transverse component of the flow equation (\[gauge\_ZA\_22.1\]), gives the following explicit relation: $$\begin{aligned} (d-1)\,\partial_{t}Z_{\bar{A}}\, p^{2} & = & -4(d-1)N\int_{q}\tilde{\partial}_{t}\left\{ G_{q}G_{q+p}\right\} \nonumber \\ & & -2Nd\int_{q}q^{2}(1-x^{2})\tilde{\partial}_{t}\left\{ G_{q}\left[G_{q+p}\left(1+R_{q+p,q}^{(1)}\right)^{2}-R_{q+p,q}^{(2)}\right]\right\} \nonumber \\ & & +Nd(d-1)\int_{q}\tilde{\partial}_{t}\left\{ G_{q}(1+R_{q}')\right\} \nonumber \\ & & +4N\int_{q}q^{2}(1-x^{2})\tilde{\partial}_{t}\left\{ G_{q}\left[G_{q+p}\left(1+R_{q+p,q}^{(1)}\right)^{2}-R_{q+p,q}^{(2)}\right]\right\} \nonumber \\ & & -2N(d-1)\int_{q}\tilde{\partial}_{t}\left\{ G_{q}(1+R_{q}')\right\} \,.\label{gauge_ZA_24}\end{aligned}$$ In (\[gauge\_ZA\_24\]) the first three lines can be traced back to the gauge diagrams $(a)$ and $(b)$ of Figure 5, while the last two lines come from the ghost diagrams $(c)$ and $(d)$. This equation is the projected form of the flow equation for $\bar{\gamma}^{(2)}$ within the truncation of the bEAA we are considering and with the functional trace explicitly evaluated. The rhs of (\[gauge\_ZA\_24\]) contains contributions of arbitrary order in $p^{2}$: from the heat kernel perspective this is equivalent to the re-summation of all contributions proportional to the invariants of the form $\int\bar{F}_{\mu\nu}(-\bar{D}^{2})^{n}\bar{F}^{\mu\nu}$ present in the coefficients $B_{2n}$ for any $n\in\mathbb{N}$. To extract the beta function $\partial_{t}Z_{\bar{A}}$ we need those terms on the rhs of (\[gauge\_ZA\_24\]) proportional to $p^{2}$. The first integral in equation (\[gauge\_ZA\_24\]) is needed only in the limit $p\rightarrow0$, since its coefficient is already of order $p^{2}$; it’s easy to see that it can be rewritten as a $Q$-functional in the following way: $$\begin{aligned} \int_{q}\tilde{\partial}_{t}\left\{ G_{q}G_{q+p}\right\} & = & -2\int_{q}(\partial_{t}R_{q}-\eta R_{q})G_{q}^{2}G_{q+p}\nonumber \\ & = & -\frac{2}{(4\pi)^{d/2}}Q_{\frac{d}{2}}\left[(\partial_{t}R_{k}-\eta R_{k})G_{k}^{3}\right]+O(p^{2})\,.\label{gauge_ZA_25}\end{aligned}$$ The integrals in the second and fourth lines of (\[gauge\_ZA\_24\]) have the same form; this is an important fact since it can be shown that the following relation holds in arbitrary dimension: $$\int_{q}q^{2}(1-x^{2})\tilde{\partial}_{t}\left\{ G_{q}\left[G_{q+p}\left(1+R_{q+p,q}^{(1)}\right)^{2}-R_{q+p,q}^{(2)}\right]\right\} =\qquad\qquad\qquad\qquad\qquad$$ $$=\frac{d-1}{(4\pi)^{d/2}}\left\{ -\frac{1}{2}Q_{\frac{d}{2}-1}\left[(\partial_{t}R_{k}-\eta R_{k})G_{k}\right]+\frac{1}{12}Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta R_{k})G_{k}\right]p^{2}\right.$$ $$\qquad\qquad\qquad\left.-\frac{1}{120}Q_{\frac{d}{2}-3}\left[(\partial_{t}R_{k}-\eta R_{k})G_{k}\right]p^{4}\right\} +O(p^{6})\,.\label{gauge_ZA_27}$$ Relation (\[gauge\_ZA\_27\]) can be checked by inserting a sufficiently smooth cutoff shape function and calculating both sides of it in a Taylor expansion in $p^{2}$. Relations like (\[gauge\_ZA\_25\]) and (\[gauge\_ZA\_27\]) serve as a dictionary to transform the explicit integrals on the rhs of the flow equation (\[gauge\_ZA\_24\]) into $Q$-functionals. With the aid of these relations we can extract from the rhs of equation (\[gauge\_ZA\_24\]) all the terms of order $p^{2}$ and a comparison with the lhs finally leads to the type I anomalous dimension of the background field: $$\begin{aligned} \eta_{\bar{A}} & = & -\frac{g^{2}N}{(4\pi)^{d/2}}\left\{ 8Q_{\frac{d}{2}}\left[(\partial_{t}R_{k}-\eta_{a}R_{k})G_{k}^{3}\right]-\frac{d}{6}Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta_{a}R_{k})G_{k}\right]+\right.\\ & & \left.+\frac{1}{3}Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta_{c}R_{k})G_{k}\right]\right\} \,,\end{aligned}$$ which is precisely equation (\[gauge\_ZA\_14\]). With this result we succeeded in showing that the $p^{2}$ terms of (\[gauge\_ZA\_24\]) indeed correspond exactly to the coefficient of the $\frac{1}{4}\int\bar{F}^{2}$ term of the functional trace that we calculated previously using the heat kernel expansion. We consider now the type II cutoff. This means that in the above derivation the gauge cutoff operator action has to be replaced by the following action: $$L[a;\bar{A}]=I[\bar{A}+a]+S_{gf}[a;\bar{A}]\,,\label{gauge_ZA_28}$$ where the action $I[A]$ has been defined in (\[gauge\_A\_1\]) and the gauge fixing action in (\[gauge\_5\]). One can check that $L^{(2;0)}[0;\bar{A}]=\bar{D}_{T}$ as it should. Diagrams $(a)$ and $(b)$ now contract to: $$\begin{aligned} (1-P)_{\mu\nu}[a_{p,q}]^{mn\,\mu\nu} & = & N\delta^{mn}\left[8(d-1)p^{2}-4dq^{2}(1-x^{2})\right]\left(1+R_{q+p,q}^{(1)}\right)^{2}G_{q}G_{q+p}\nonumber \\ \nonumber \\ (1-P)_{\mu\nu}[b_{p,q}]^{mn\,\mu\nu} & = & N\delta^{mn}\left\{ 2d(d-1)(1+R_{q}')\right.\nonumber \\ & & \left.+\left[8(d-1)p^{2}-4dq^{2}(1-x^{2})\right]R_{q+p,q}^{(2)}\right\} G_{q}\,.\label{gauge_ZA_28.1}\end{aligned}$$ In the ghost sector we don’t make any changes. The first two terms of equation (\[gauge\_ZA\_24\]) are now replaced by: $$-N\int_{q}\left[4(d-1)p^{2}+2dq^{2}(1-x^{2})\right]\tilde{\partial}_{t}\left\{ G_{q}\left[G_{q+p}\left(1+R_{q+p,q}^{(1)}\right)^{2}-R_{q+p,q}^{(2)}\right]\right\} \,.\label{gauge_ZA_29}$$ By using the following relation, that can be checked as before, $$\int_{q}\tilde{\partial}_{t}\left\{ G_{q}\left[G_{q+p}\left(1+R_{q+p,q}^{(1)}\right)^{2}-R_{q+p,q}^{(2)}\right]\right\} =\qquad\qquad\qquad\qquad\qquad\qquad$$ $$\qquad=\frac{1}{(4\pi)^{d/2}}\left\{ -Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta R_{k})G_{k}\right]+\frac{1}{6}Q_{\frac{d}{2}-3}\left[(\partial_{t}R_{k}-\eta R_{k})G_{k}\right]\frac{p^{2}}{k^{2}}\right\} +O\left(\frac{p^{4}}{k^{4}}\right)\,,\label{gauge_ZA_29.1}$$ to expand the rhs of (\[gauge\_ZA\_29\]) one arrives at the expression for the type II anomalous dimension of the background field: $$\eta_{\bar{A}}=-\frac{g^{2}N}{(4\pi)^{d/2}}\left\{ \frac{24-d}{6}Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta_{a}R_{k})G_{k}\right]+\frac{1}{3}Q_{\frac{d}{2}-2}\left[(\partial_{t}R_{k}-\eta_{c}R_{k})G_{k}\right]\right\} \,.$$ Again, as a further confirmation of our methods, we have re-derived the result (\[gauge\_ZA\_10\]) obtained before employing the heat kernel expansion. Along the same lines one can study the longitudinal equation obtained from (\[gauge\_ZA\_22.1\]) by contraction with $P^{\mu\nu}$ and make an important check. In fact one finds, in both type I and II cases, that the equation (in particular the rhs) turns out to be identically zero; this is expected, since, by construction, the flow of the zero-field proper-vertices $\bar{\gamma}^{(2)}$ of the gEAA is transverse. Another check one can make is to control that the terms of order $p^{0}$ in equation (\[gauge\_ZA\_24\]) cancel each other, since otherwise they will generate a non-gauge invariant mass term for the background field. Using (\[gauge\_ZA\_27\]) and the following relation, $$\int_{q}\tilde{\partial}_{t}\left\{ G_{q}(1+R_{q}')\right\} =-\frac{1}{(4\pi)^{d/2}}Q_{\frac{d}{2}-1}\left[(\partial_{t}R_{k}-\eta R_{k})G_{k}\right]\,,\label{gauge_ZA_30}$$ to evaluate equation (\[gauge\_ZA\_24\]) at $p=0$, gives: $$\begin{aligned} & -2Nd\left(-\frac{1}{2}\frac{d-1}{(4\pi)^{d/2}}Q_{\frac{d}{2}-1}\left[(\partial_{t}R_{k}-\eta_{a}R_{k})G_{k}\right]\right)+Nd(d-1)\left(-\frac{1}{(4\pi)^{d/2}}Q_{\frac{d}{2}-1}\left[(\partial_{t}R_{k}-\eta_{a}R_{k})G_{k}\right]\right)\\ & +4N\left(-\frac{1}{2}\frac{d-1}{(4\pi)^{d/2}}Q_{\frac{d}{2}-1}\left[(\partial_{t}R_{k}-\eta_{c}R_{k})G_{k}\right]\right)-2N(d-1)\left(-\frac{1}{(4\pi)^{d/2}}Q_{\frac{d}{2}-1}\left[(\partial_{t}R_{k}-\eta_{c}R_{k})G_{k}\right]\right)\\ & =\left[-2\left(-\frac{1}{2}\right)+(-1)\right]\frac{Nd(d-1)}{(4\pi)^{d/2}}Q_{\frac{d}{2}-1}\left[(\partial_{t}R_{k}-\eta_{a}R_{k})G_{k}\right]\\ & +\left[4\left(-\frac{1}{2}\right)-(-2)\right]\frac{Nd(d-1)}{(4\pi)^{d/2}}Q_{\frac{d}{2}-1}\left[(\partial_{t}R_{k}-\eta_{c}R_{k})G_{k}\right]=0\,.\end{aligned}$$ As we see the $p^{0}$ contributions in both the gauge and ghost sectors correctly cancel each other. The same can be checked in case of the type II cutoff. We remark that these are nontrivial checks of the formalism. It this framework it is not difficult to relax the condition $\alpha=1$. One does not need to change anything apart considering the regularized propagator in its general form (\[gauge\_P\_7.3\]) and using the $\alpha$ dependent gauge-fixing vertices from appendix B. Work in this direction will be material for a future studies [@Codello_in_preparation]. The aim of this subsection was to show how the method introduced in section 2 can be used to reproduce the results obtained with the aid of the local heat kernel expansion, and to show how the different cutoff choices are reflected in this formalism. In this way we made an important test of the method. We remark that this way of evaluating the flow equation for the gEAA is very general and can be applied to any general truncation ansatz, in particular to those that cannot be treated with the aid of the local heat kernel expansion. We move now to perform a new application of the technique by calculating $\eta_{a,k}$ and $\eta_{c,k}$ in the next subsection. ### Calculation of $\eta_{a,k}$ and $\eta_{c,k}$ In this subsection we calculate the anomalous dimensions of the fluctuation and ghost fields; these are related to the scale derivatives of wave-function renormalization constants, $\partial_{t}Z_{a}$ and $\partial_{t}Z_{c}$, by the relations: $$\eta_{a}=-\partial_{t}\log Z_{a}\qquad\qquad\qquad\eta_{c}=-\partial_{t}\log Z_{c}\,.\label{gauge_Za_0}$$ We will extract $\partial_{t}Z_{a}$ and $\partial_{t}Z_{c}$ form, respectively, the flow equations for the zero-field two-point functions $\gamma^{(2,0,0;0)}$ and $\gamma^{(0,1,1;0)}$ obtained from the flow equation (\[gauge\_D\_4\]) of section 2 after performing the field multiplet decomposition. ![Diagrammatic representation of the flow equation for the zero-field proper-vertex $\gamma_{k}^{(2,0,0;0)}$ of the bEAA used to calculate $\partial_{t}Z_{a,k}$.](Flow_Eqn_bEAA_2_tris.pdf) The flow equation for $\gamma^{(2,0,0;0)}$, within the truncation represented by equations (\[gauge\_A\_3\]) and (\[gauge\_A\_3.1\]), is given in Figure 6. In formulas this equation can be written as follows: $$\begin{aligned} [\partial_{t}\gamma_{p,-p}^{(2,0,0;0)}]^{mn\,\mu\nu} & = & \Omega\int_{q}(\partial_{t}R_{q}-\eta_{a}R_{q})\,[a_{p,q}]^{mn\,\mu\nu}-\frac{1}{2}\Omega\int_{q}(\partial_{t}R_{q}-\eta_{a}R_{q})\,[b_{p,q}]^{mn\,\mu\nu}\nonumber \\ & & -2\Omega\int_{q}(\partial_{t}R_{q}-\eta_{c}R_{q})\,[c_{p,q}]^{mn\,\mu\nu}\,,\label{gauge_Za_2.10}\end{aligned}$$ where the tensor products entering it are: $$\begin{aligned} [a_{p,q}]^{mn\,\mu\nu} & = & [G_{q}]^{ab\,\alpha\beta}[\gamma_{q,-q-p,p}^{(3,0,0;0)}]^{bcm\,\beta\gamma\mu}[G_{q+p}]^{cd\,\gamma\delta}[\gamma_{q+p,-q,-p}^{(3,0,0;0)}]^{den\,\delta\epsilon\nu}[G_{q}]^{ea\,\epsilon\alpha}\nonumber \\ {}[b_{p,q}]^{mn\,\mu\nu} & = & [G_{q}]^{ab\,\alpha\beta}[\gamma_{q,-q,p,-p}^{(4,0,0;0)}]^{bcmn\,\beta\gamma\mu\nu}[G_{q}]^{ca\,\gamma\alpha}\nonumber \\ {}[c_{p,q}]^{mn\,\mu\nu} & = & [G_{q}^{c}]^{ab}[\gamma_{p,q,-q-p}^{(1,1,1;0)}]^{mbc\,\mu}[G_{q}^{c}]^{cd}[\gamma_{p,q+p,-q}^{(1,1,1;0)}]^{nde\,\nu}[G_{q}^{c}]^{ea}\,,\label{gauge_Za_2.12}\end{aligned}$$ and the vertices the following: $$\begin{aligned} [\gamma_{p_{1},p_{2},p_{3}}^{(3,0,0;0)}]^{abm\,\alpha\beta\mu} & = & gZ_{a}^{3/2}[I_{p_{1},p_{2},p_{3}}^{(3)}]^{abm\,\alpha\beta\mu}\nonumber \\ {}[\gamma_{p_{1},p_{2},p_{3},p_{4}}^{(4,0,0;0)}]^{abmn\,\alpha\beta\mu\nu} & = & g^{2}Z_{a}[I_{p_{1},p_{2},p_{3},p_{4}}^{(4)}]^{abmn\,\alpha\beta\mu\nu}\nonumber \\ {}[\gamma_{p_{1},p_{2},p_{3}}^{(1,1,1;0)}]^{mab\,\mu} & = & Z_{c}^{1/2}[S_{gh\, p_{1},p_{2},p_{3}}^{(1,1,1;0)}]^{mab\,\mu}\,.\label{gauge_Za_2.13}\end{aligned}$$ Diagram $(d)$ in Figure 6 is identically zero, since in our truncation there is no term bilinear both in the ghost and in the gauge fields, and we omitted to write it in (\[gauge\_Za\_2.10\]). The vertices of the actions in (\[gauge\_Za\_2.13\]) are given in appendix B. To deal with scalar equations we project equation (\[gauge\_Za\_2.10\]) using the orthogonal projectors $(1-P)^{\mu\nu}$ and $P^{\mu\nu}$. In this way we obtain two independent equations describing, respectively, the flow of the transverse and longitudinal components of $\gamma^{(2,0,0;0)}$. We will use the transverse equation to extract $\partial_{t}Z_{a}$. We mention here that the longitudinal equation can be used to extract the running of $\alpha$, if one considers a truncation with running gauge parameter [@Codello_in_preparation]. Applying the transverse projector to the lhs of the flow equation (\[gauge\_Za\_2.10\]) and using (\[gauge\_P\_1\]) gives: $$(1-P)_{\mu\nu}[\partial_{t}\bar{\gamma}_{p,-p}^{(2)}+\partial_{t}\hat{\gamma}_{p,-p}^{(2,0,0;0)}]^{mn\,\mu\nu}=\Omega\,\delta^{mn}(d-1)\,\partial_{t}Z_{a}\, p^{2}\,,\label{gauge_Za_1}$$ where we used the trace $(1-P)_{\alpha}^{\alpha}=d-1$. Equation (\[gauge\_Za\_1\]) shows that we can extract $\partial_{t}Z_{a}$ as the coefficient of the $p^{2}$ term of the transverse projection of equation (\[gauge\_Za\_2.10\]). Acting on the first integrand in (\[gauge\_Za\_2.12\]) with the transverse projector gives: $$\begin{aligned} (1-P)_{\mu\nu}[a_{p,q}]^{mn\,\mu\nu} & = & g^{2}\,\Omega\, Z_{a}(-N\delta^{mn})\left\{ -5(d-1)p^{2}-2(d-1)p\cdot q\right.\nonumber \\ & & \left.+2\left[(2d-3)x^{2}-3d+4\right]q^{2}\right\} (G_{q})^{2}G_{q+p}\,.\label{gauge_Za_2.2}\end{aligned}$$ The group factor here is $f^{amb}f^{bna}=-N\delta^{mn}$, while, as before, the variable $x$ is the cosine of the angle between $p$ and $q$. The transverse contribution from the second integrand in (\[gauge\_Za\_2.12\]) is: $$(1-P)_{\mu\nu}[b_{p,q}]^{mn\,\mu\nu}=g^{2}\,\Omega\, Z_{a}(-2N\delta^{mn})\left\{ -(d-1)^{2}\right\} (G_{q})^{2}\,,\label{gauge_Za_2.3}$$ here the group factor is $f^{abm}f^{anb}+f^{abn}f^{amb}=-2N\delta^{mn}$. From (\[gauge\_Za\_2.12\]) we find the transverse contribution from the ghost diagram $(c)$: $$(1-P)_{\mu\nu}[c_{p,q}]^{mn\,\mu\nu}=g^{2}\,\Omega\, Z_{a}(-N\delta^{mn})\left\{ -(1-x^{2})q^{2}\right\} (G_{q})^{2}G_{q+p}\,,\label{gauge_Za_2.4}$$ where the group factor is $f^{abm}f^{ban}=-N\delta^{mn}$. Note that each integrand (\[gauge\_Za\_2.2\]-\[gauge\_Za\_2.4\]) (or diagram) is proportional to $g^{2}Z_{a}$ since the fluctuation three-vertex comes with a factor $gZ_{a}^{3/2}$, the four-vertex with a factor $g^{2}Z_{a}^{2}$, while the regularized gauge propagators come with a power of $Z_{a}^{-1}$ and a gauge cutoff kernel insertion with a power of $Z_{a}$. In the ghost diagrams the three-vertex gives a factor $gZ_{a}^{1/2}Z_{c}$, there is no four-vertex, the regularized ghost propagator has a factor of $Z_{c}^{-1}$ and the ghost cutoff kernel insertion has a power of $Z_{c}$. Also, all the space-time volume factors $\Omega$ on both sides of equation (\[gauge\_Za\_2.10\]) delete each other. Once we insert equations (\[gauge\_Za\_2.2\]), (\[gauge\_Za\_2.3\]) and (\[gauge\_Za\_2.4\]) back in (\[gauge\_Za\_2.10\]) we obtain the explicit flow of the vertex $\gamma_{p,-p}^{(2,0,0;0)}$, to all orders in the external momenta $p$. The momentum integrals in (\[gauge\_Za\_2.10\]) can be performed using spherical coordinates with the $z$-axis along the vector $p$: $$\int_{q}\rightarrow\frac{S_{d-1}}{(2\pi)^{d}}\int_{0}^{\infty}dq\, q^{d-1}\int_{-1}^{1}dx\left(1-x^{2}\right)^{\frac{d-3}{2}}\,,\label{gauge_Za_2.5}$$ where $S_{d}=\frac{2\pi^{\frac{d}{2}}}{\Gamma(\frac{d}{2})}$ is the volume of the $d$-dimensional sphere. We also change variable $z=q^{2}$ in the radial integral so that: $$\int_{0}^{\infty}dq\, q^{d-1}\rightarrow\frac{1}{2}\int_{0}^{\infty}dz\, z^{\frac{d}{2}-1}\,.\label{gauge_Za_2.51}$$ After expanding the transverse projection of equation (\[gauge\_Za\_2.10\]) in powers of $p$ the $x$-integrals can be easily performed; from the terms proportional to $p^{2}$ we can extract $\partial_{t}Z_{a}$ and from these we obtain the anomalous dimension of the fluctuation field: $$\begin{aligned} \eta_{a} & = & \frac{g^{2}N}{(4\pi)^{d/2}}\left\{ -5Q_{\frac{d}{2}}\left[\left(\partial_{t}R_{k}-\eta_{a}R_{k}\right)G_{k}^{3}\right]-(3d-1)Q_{\frac{d}{2}+1}\left[\left(\partial_{t}R_{k}-\eta_{a}R_{k}\right)G_{k}^{2}G_{k}'\right]\right.\nonumber \\ & & -\left(3d-1\right)Q_{\frac{d}{2}+2}\left[\left(\partial_{t}R_{k}-\eta_{a}R_{k}\right)G_{k}^{2}G_{k}''\right]+Q_{\frac{d}{2}+1}\left[\left(\partial_{t}R_{k}-\eta_{c}R_{k}\right)G_{k}^{2}G_{k}'\right]\nonumber \\ & & \left.+Q_{\frac{d}{2}+2}\left[\left(\partial_{t}R_{k}-\eta_{c}R_{k}\right)G_{k}^{2}G_{k}''\right]\right\} \,,\label{gauge_Za_5}\end{aligned}$$ where we wrote everything in terms of $Q$-functionals. Equation (\[gauge\_Za\_5\]) is valid in arbitrary dimension and for every cutoff shape function, and gives implicitly the anomalous dimension of the fluctuation field, in the gauge $\alpha=1$, as part of a linear system for the variables $\eta_{a}$ and $\eta_{c}$. In the following we derive an analogous equation for the anomalous dimension $\eta_{c}$ of the ghost fields, that together with (\[gauge\_Za\_5\]), can be used to solve for $\eta_{a}$ and $\eta_{c}$ in function of $g$. We now calculate the scale derivative of the ghost wave-function renormalization $\partial_{t}Z_{c}$. The only term in our truncation (\[gauge\_A\_3.1\]) that contains the ghost fields and the related wave-function renormalization is the following: $$Z_{c}\int d^{d}x\,\bar{D}_{\mu}\bar{c}\left(\bar{D}^{\mu}+gZ_{a}^{1/2}a^{\mu}\right)c\,.\label{gauge_gh_1}$$ We can extract $\partial_{t}Z_{c}$ from the flow equation for the ghost-ghost zero-field proper-vertex $\gamma^{(0,1,1;0)}$. This equation is obtained from the flow equation (\[gauge\_D\_4\]) for the zero-field proper-vertex $\gamma^{(2;0)}$ after the field multiplet decomposition; it is represented graphically in Figure 7 and reads as follows: $$[\partial_{t}\gamma_{p,-p}^{(0,1,1;0)}]^{mn}=\Omega\int_{q}(\partial_{t}R_{q}-\eta_{a}R_{q})\,[e_{p,q}]^{mn}+\Omega\int_{q}(\partial_{t}R_{q}-\eta_{c}R_{q})\,[f_{p,q}]^{mn}\,,\label{gauge_gh_1.1}$$ where the integrands are: ![Graphical representation of the flow equation for the ghost-ghost zero-field proper vertex $\gamma_{k}^{(0,1,1;0)}$ from (\[gauge\_gh\_1.1\]).](Flow_Eqn_ghost.pdf) $$\begin{aligned} [e_{p,q}]^{mn} & = & [G_{q}]^{ab\,\alpha\beta}[\gamma_{q,p,-q-p}^{(1,1,1;0)}]^{bmc\,\beta}[G_{q+p}]^{cd}[\gamma_{-q,q+p,-p}^{(1,1,1;0)}]^{den\,\gamma}[G_{q}]^{ea\,\gamma\alpha}\nonumber \\ {}[f_{p,q}]^{mn} & = & [G_{q}]^{ab}[\gamma_{-q-p,,p,q}^{(1,1,1;0)}]^{bmc\,\gamma}[G_{q+p}]^{cd\,\gamma\delta}[\gamma_{q+p,-q,-p}^{(1,1,1;0)}]^{den\,\delta}[G_{q}]^{ea}\,.\label{gauge_gh_2.1}\end{aligned}$$ The only vertex we need is thus the following: $$[\gamma_{p_{1},p_{2},p_{3}}^{(1,1,1;0)}]^{mab\,\mu}=gZ_{c}^{1/2}[S_{gh\, p_{1},p_{2},p_{3}}^{(1,1,1;0)}]^{mab\,\mu}\,,$$ which can be found in appendix B. In equation (\[gauge\_gh\_1.1\]) only the ghost-ghost-gluon vertex appears since the action (\[gauge\_gh\_1\]) is at most trilinear in the fluctuation and ghost fields, thus the second term in the flow equation (\[gauge\_D\_4\]) vanishes. Inserting (\[gauge\_gh\_1\]) in the lhs of (\[gauge\_gh\_1.1\]) gives: $$[\partial_{t}\gamma_{p,-p}^{(0,1,1;0)}]^{mn}=\Omega\,\delta^{mn}\,\partial_{t}Z_{c}\, p^{2}\,.\label{gauge_gh_2}$$ As before we perform the tensor products in (\[gauge\_gh\_2.1\]) and we evaluate the integrals in (\[gauge\_gh\_1.1\]); we then extract $\partial_{t}Z_{c}$ from the term of order $p^{2}$; after writing everything in terms of $Q$-functionals, we obtain the anomalous dimension of the ghost fields in the following form: $$\begin{aligned} \eta_{c} & = & \frac{g^{2}N}{(4\pi)^{d/2}}\left\{ -Q_{\frac{d}{2}}\left[(\partial_{t}R-\eta_{a}R)G_{k}^{3}\right]-Q_{\frac{d}{2}+1}\left[(\partial_{t}R-\eta_{a}R)G_{k}^{2}G_{k}'\right]\right.\nonumber \\ & & \left.+Q_{\frac{d}{2}+1}\left[(\partial_{t}R-\eta_{c}R)G_{k}^{2}G_{k}'\right]\right\} \,.\label{gauge_gh_7}\end{aligned}$$ Equation (\[gauge\_gh\_7\]) is the ghost anomalous dimensions in the gauge $\alpha=1$ (within the truncation we are considering) and is valid for general cutoff shape function and dimension. Equations (\[gauge\_Za\_5\]) and (\[gauge\_gh\_7\]) are the main results of this subsection. Beta functions and non-perturbative predictions ----------------------------------------------- In this subsection we study the different ways one has to “close” the beta function for the gauge coupling derived in the previous subsection and we make a comparison with other approaches. Then we study some possible phenomenology related to these RG flows. ### Beta functions To obtain an explicit form for the anomalous dimension of the background field, $$\eta_{\bar{A},k}=-\partial_{t}\log\, Z_{\bar{A},k}\,,\label{BF_0}$$ given in equation (\[gauge\_ZA\_10\]) for type II cutoff or in equation (\[gauge\_ZA\_14\]) for type I cutoff, we need to specify the cutoff shape function $R_{k}(z)$. We will consider the theta– or optimized–cutoff shape function: $$R_{k}(z)=Z_{k}(k^{2}-z)\theta(k^{2}-z)\,,\label{BF_0.01}$$ (where $Z_{k}$ represents $Z_{a,k}$ or $Z_{c,k}$) since it allows us to perform the integrals in the $Q$-functionals analytically for every value of the dimension $d$. A simple integration gives the following form: $$Q_{n}[(\partial_{t}R_{k}-\eta R_{k})G_{k}^{m}]=\frac{k^{2(n+1-m)}}{\Gamma(n+2)}\left[2(n+1)-\eta\right]\label{BF_0.1}$$ (where $\eta$ stands for $\eta_{a,k}$ or $\eta_{c,k}$). If we insert (\[BF\_0.1\]) in the type I anomalous dimension (\[gauge\_ZA\_14\]), we find: $$\eta_{\bar{A},k}=\frac{g_{k}^{2}N\, k^{d-4}}{(4\pi)^{d/2}\Gamma\left(\frac{d}{2}\right)}\left[-\frac{192-d(d-2)^{2}}{6d}+\frac{192-d^{2}(d+2)}{6d(d+2)}\eta_{a,k}+\frac{1}{3}\eta_{c,k}\right]\,;\label{BF_1}$$ while if we insert (\[BF\_0.1\]) in the type II anomalous dimension (\[gauge\_ZA\_10\]) we get: $$\eta_{\bar{A},k}=\frac{g_{k}^{2}N\, k^{d-4}}{(4\pi)^{d/2}\Gamma\left(\frac{d}{2}\right)}\left[-\frac{(26-d)(d-2)}{6}+\frac{24-d}{6}\eta_{a,k}+\frac{1}{3}\eta_{c,k}\right]\,.\label{BF_2}$$ Note that only the type II anomalous dimension has the one–loop term proportional to $26-d$, while type I becomes positive already at $d\approx7.174$; for $d\rightarrow\infty$ both coefficients go as $d^{2}/12$; at $d=2$ the type II coefficient is zero while the type I is negative. See Figure 8 for a comparison. Equation (\[BF\_1\]) and (\[BF\_2\]) show that the anomalous dimension of the background field $\eta_{\bar{A},k}$ is determined by the anomalous dimensions of the fluctuation and ghost fields $$\eta_{a,k}=-\partial_{t}\log\, Z_{a,k}\qquad\qquad\eta_{c,k}=-\partial_{t}\log\, Z_{c,k}\,.\label{BF_3}$$ This reflects the fact that the flow of the gEAA (here represented by $\eta_{\bar{A},k}$) is not closed but depends on the flow of the full bEAA (here represented by $\eta_{a,k}$ and $\eta_{c,k}$). The anomalous dimensions of the fluctuation and of the ghost fields are given, respectively, in equations (\[gauge\_Za\_5\]) and (\[gauge\_gh\_7\]). The $Q$-integrals that we need now are: ![The one-loop coefficient $a$ in the cutoff schemes type I (lower curve at $d=2$) and type II (upper curve at $d=2$). For reference we plotted the line $-\frac{11}{3}$.](Beta_gauge_a.pdf) $$\begin{aligned} Q_{n}[(\partial_{t}R_{k}-\eta R_{k})G_{k}^{m}G'_{k}] & = & 0\nonumber \\ Q_{n}[(\partial_{t}R_{k}-\eta R_{k})G_{k}^{m}G_{k}''] & = & -\frac{k^{2(n+1-m)}}{\Gamma(n)}\,;\label{BF_3.1}\end{aligned}$$ with these we find the following system: $$\begin{aligned} \eta_{a,k} & = & \frac{g_{k}^{2}N\, k^{d-4}}{(4\pi)^{d/2}\Gamma\left(\frac{d}{2}\right)}\left[-\frac{8(d+6)}{d(d+2)}+\frac{20}{d(d+2)}\eta_{a,k}\right]\label{BF_4}\\ \eta_{c,k} & = & \frac{g_{k}^{2}N\, k^{d-4}}{(4\pi)^{d/2}\Gamma\left(\frac{d}{2}\right)}\left[-\frac{4}{d}+\frac{4}{d(d+2)}\eta_{a,k}\right]\,.\label{BF_5}\end{aligned}$$ We want to remember to the reader that we are working in the gauge $\alpha=1$ and that (\[BF\_4\]) and (\[BF\_5\]) generally depend on the gauge-fixing parameter $\alpha$. Contrary to $\eta_{\bar{A},k}$, the anomalous dimensions $\eta_{a,k}$ and $\eta_{c,k}$ are the same for both cutoff types. Note also that the one–loop terms in (\[BF\_4\]) and (\[BF\_5\]) are always negative. In the background field formalism the gauge coupling is related to the wave-function renormalization of the background field by (\[gauge\_A\_4\]), i.e. $g_{k}=Z_{\bar{A},k}^{-1/2}$. Thus the beta function for the gauge coupling can be obtained from the anomalous dimensions of the background field by the following simple relation: $$\partial_{t}g_{k}=\frac{1}{2}\eta_{\bar{A},k}g_{k}\,.\label{BF_6}$$ One can now shift to dimensionless coupling $\tilde{g}_{k}=k^{\frac{d-4}{2}}g_{k}$ to find: $$\partial_{t}\tilde{g}_{k}=\frac{1}{2}(d-4+\eta_{\bar{A},k})\tilde{g}_{k}\,.\label{BF_6.1}$$ Even if the previous equations are valid in any dimension, in the following we will consider the physical interesting case $d=4$ where $g_{k}=\tilde{g}_{k}$. Owning to the structure of both (\[BF\_1\]) and (\[BF\_2\]) we can write the following general form for the background field anomalous dimension in a given scheme: $$\eta_{\bar{A},k}=\left(2a+b\,\eta_{a,k}+b'\,\eta_{c,k}\right)g_{k}^{2}\,,\label{BF_7}$$ together with the following values for the coefficients in the two cutoff schemes: $$a_{I}=-\frac{11}{3}\frac{N}{(4\pi)^{2}}\qquad\qquad b_{I}=\frac{2}{3}\frac{N}{(4\pi)^{2}}\qquad\qquad b_{I}'=\frac{1}{3}\frac{N}{(4\pi)^{2}}\,,\label{BF_8}$$ $$a_{II}=-\frac{11}{3}\frac{N}{(4\pi)^{2}}\qquad\qquad b_{II}=\frac{10}{3}\frac{N}{(4\pi)^{2}}\qquad\qquad b_{II}'=\frac{1}{3}\frac{N}{(4\pi)^{2}}\,.\label{BF_9}$$ As expected we recover the scheme independent one–loop result $a_{I}=a_{II}=-\frac{11}{3}\frac{N}{(4\pi)^{2}}$. The coefficients $b$ depend instead on the cutoff type; however, these coefficients are gauge dependent and may have closer values in another gauge such as the Landau $\alpha=0$. Note also that $b'_{I}=b'_{II}$ since the ghost contributions are the same in both schemes. Since the anomalous dimensions of the fluctuation and ghost fields (\[BF\_4\]) and (\[BF\_5\]), to lowest order, are proportional to $g_{k}^{2}$, the only term on the lhs of (\[BF\_7\]) of order $g_{k}^{2}$ is the first. To this order we recover the standard one–loop beta function of the gauge coupling: $$\partial_{t}g_{k}=a\, g_{k}^{3}+O(g_{k}^{5})=-\frac{11}{3}\frac{N}{(4\pi)^{2}}g_{k}^{3}+O(g_{k}^{5})\,.\label{BF_10}$$ Since the beta function is negative at small coupling, non-abelian gauge theories are asymptotically free in $d=4$ [@Gross_Wilczek_1973]. From this we see that the one–loop approximation is equivalent to set $\eta_{a,k}=\eta_{c,k}=0$ in (\[BF\_7\]). For a discussion of the physical mechanism behind (\[BF\_10\]) asymptotic freedom we refer to [@Reuter_Nink_2012]. In the general case, to determine $\eta_{\bar{A},k}$ from equation (\[BF\_7\]) we need to calculate or specify the anomalous dimensions $\eta_{a,k}$ and $\eta_{c,k}$ as functions of $g_{k}$. As we said, this is the actual manifestation of the fact that the flow of the gEAA ($g_{k}$) is not closed but is given in terms of the flow of the bEAA ($\eta_{a,k}$ and $\eta_{c,k}$). We discuss now two different approximations that allow us to obtain a closed form for $\eta_{\bar{A},k}$ and thus an improved form for the beta function of the gauge coupling. The first approximation, or improvement, consists in identifying the anomalous dimension of the background field with the anomalous dimension of the fluctuation field, and by setting the anomalous dimension of the ghost fields to zero: $$\eta_{\bar{A},k}=\eta_{a,k}\qquad\qquad\eta_{c,k}=0\,.\label{BF_11}$$ Previous applications of the bEAA to non-abelian gauge theories [@Reuter_Wetterich_1994a; @ReuterWetterich_1994c] used this identification. As we will see in a moment, this identification becomes exact in the supersymmetric limit. If we insert now (\[BF\_11\]) in (\[BF\_7\]) we find the following linear system for the variable $\eta_{\bar{A},k}$: $$\eta_{\bar{A},k}=\left(2a+b\,\eta_{\bar{A},k}\right)g_{k}^{2}\,,\label{BF_12}$$ which is easily solved to yield: $$\eta_{\bar{A},k}=\frac{2a}{1-b\, g_{k}^{2}}g_{k}^{2}\,,\label{BF_13}$$ or, using (\[BF\_6\]), the following beta function for the gauge coupling: $$\partial_{t}g_{k}=\frac{a}{1-b\, g_{k}^{2}}g_{k}^{3}\,.\label{BF_14}$$ The beta function (\[BF\_14\]) is a rational function of the gauge coupling; this shows how the identification (\[BF\_11\]) implements the re–summation of an infinite number of perturbative contributions. One clearly sees the presence of a singularity in the beta function (\[BF\_14\]) at $g_{k}^{2}=1/b$. We will see in the next subsection a possible physical consequence of this fact. The beta function (\[BF\_14\]) in the two schemes is shown in Figure 9. In the spirit of [@Reuter_Bi_Field] this is a single–field truncation. ![Single–field beta functions from (\[BF\_14\]). Asymptotes from the left: type II, RS and type I beta functions. The one-loop beta function (without asymptote) is shown for comparison. Note the similarity between the type II and the RS beta functions.](Beta_gauge_I.pdf) Ryttov and Sannino (RS) [@Ryttov_Sannino_2008] proposed an “all orders” beta function for the gauge coupling that has exactly the structure found in (\[BF\_14\]), but with the following coefficients: $$a_{RS}=-\frac{11}{3}\frac{N}{(4\pi)^{2}}\qquad\qquad b_{RS}=\frac{34}{11}\frac{N}{(4\pi)^{2}}\,,\label{BF_14.01}$$ chosen in order to match the one– and two–loop results for the beta function upon Taylor expansion in $g_{k}$. The inspiration behind the ansatz of RS comes from the knowledge of the exact beta function found by Novikov-Shifman-Vainshtein-Zakharov (NSVZ) [@Novikov_Shifman_Vainshtein_Zakharov_1983] in the supersymmetric $\mathcal{N}=1$ version of the model. The NSVZ beta function is also of the form (\[BF\_14\]) but with coefficients $a_{NSVZ}=-3\frac{N}{(4\pi)^{2}}$ and $b_{NSVZ}=2\frac{N}{(4\pi)^{2}}$. RS postulated the existence of a perturbative massless scheme where all the perturbative orders are reproduced by the choice (\[BF\_14.01\]); here we note the interesting fact that the RS beta function is very similar to our type II beta function, as clearly shown in Figure 9. We see that the bEAA formalism offers a framework where beta functions of the form (\[BF\_14\]) can be obtained from first principles. The linear relation for the anomalous dimension of the background field postulated in [@Pica_Sannino_2011] to derive the RS beta function, is here a consequence of the structure of the exact RG flow equation for the gEAA. As a future work, it will be interesting to reproduce the exact results for supersymmetric beta functions of [@Novikov_Shifman_Vainshtein_Zakharov_1983] using functional RG methods. When the coupling is small $g_{k}\ll1$ we can expand the beta function (\[BF\_14\]) to find: $$\partial_{t}g_{k}=a\, g_{k}^{3}+ab\, g_{k}^{5}+O\left(g_{k}^{7}\right)\,.\label{BF_14.1}$$ As we already noticed, the one-loop contributions are scheme independent $a_{I}=a_{II}=a_{RS}$ and equal to the perturbative result, while the two–loop contributions are instead different. The RS coefficient is, by construction, equal to the perturbative result $a_{RS}b_{RS}=-\frac{g_{k}^{5}}{(4\pi)^{4}}\frac{102}{9}N^{2}$ [@Abbott_1981], while we find $a_{I}b_{I}=-\frac{g_{k}^{5}}{(4\pi)^{4}}\frac{22}{9}N^{2}$ and $a_{II}b_{II}=-\frac{g_{k}^{5}}{(4\pi)^{4}}\frac{110}{9}N^{2}$. Note that the type I coefficient is $79\%$ smaller than the two–loop coefficient, while type II coefficient is just $8\%$ bigger. These two–loop coefficients are expected not to be equal to the perturbative result, since this is scheme independent only in massless schemes, while the bEAA formalisms implements a kind of mass dependent regularization [@Gies_2006]. The second way we can obtain a closed beta function for the gauge coupling, is to first calculate the anomalous dimensions $\eta_{a,k}$ and $\eta_{c,k}$ and then reinsert them back in (\[BF\_7\]). This means that we are considering the flow of the full bEAA which takes place in the enlarged theory space of all functionals of the fluctuation fields $a_{\mu}$, $\bar{c}$, $c$, and of the background field $\bar{A}_{\mu}$. In the truncation we are considering, given by equations (\[gauge\_A\_3\]) and (\[gauge\_A\_3.1\]), this is the three-dimensional space parametrized by the coupling constants $\{g_{k},Z_{a,k},Z_{c,k}\}$. In the spirit of [@Reuter_Bi_Field] this is a bi–field truncation; see also [@Rosten_2011] for a discussion of this enlarged class of truncations. We have seen that the anomalous dimensions of the fluctuation field and of the ghost fields are determined by a linear system; we will solve this to obtain $\eta_{a,k}$ and $\eta_{c,k}$ as functions of the gauge coupling $g_{k}$ and then we will reinsert them back in equation (\[BF\_7\]) to obtain $\eta_{\bar{A},k}$; from this we then obtain a new closed form for $\partial_{t}g_{k}$. In the most general case (compatible with our truncation), the anomalous dimensions $\eta_{a,k}$ and $\eta_{c,k}$ are determined by the following linear system; $$\begin{aligned} \eta_{a,k} & = & (A+B\,\eta_{a,k}+C\,\eta_{c,k})g_{k}^{2}\nonumber \\ \eta_{c,k} & = & (A'+B'\,\eta_{a,k}+C'\,\eta_{c,k})g_{k}^{2}\,.\label{BF_16}\end{aligned}$$ The coefficients in (\[BF\_16\]) do not depend on the scheme I or II, but they depend on the gauge and on $R_{k}(z)$. In particular, setting $d=4$ in (\[BF\_4\]) and (\[BF\_5\]) gives the following values for the coefficients: $$A=-\frac{10}{3}\frac{N}{(4\pi)^{2}}\qquad\qquad B=\frac{5}{6}\frac{N}{(4\pi)^{2}}\qquad\qquad C=0\,,$$ $$A'=-\frac{N}{(4\pi)^{2}}\qquad\qquad B'=\frac{1}{6}\frac{N}{(4\pi)^{2}}\qquad\qquad C'=0\,.\label{BF_16.1}$$ As a note, we mention that the scheme independent coefficients $A$ and $A'$ become, respectively, $-\frac{13-3\alpha}{3}$ and $-\frac{3-\alpha}{2}$ when $\alpha\neq1$ and the coefficients $C$ and $C'$ become non-zero. A system analogous to (\[BF\_16\]) was studied within the non-background EAA approach to non-abelian gauge theories in [@Ellwanger_Hirsch_Weber_1996]. Considering that the anomalous dimensions in the rhs of (\[BF\_16\]) are at least of order $g_{k}^{2}$ we find, to lowest order, the following forms: $$\begin{aligned} \eta_{a,k} & = & -\frac{10}{3}\frac{g_{k}^{2}N}{(4\pi)^{2}}+O(g_{k}^{4})\nonumber \\ \eta_{c,k} & = & -\frac{g_{k}^{2}N}{(4\pi)^{2}}+O(g_{k}^{4})\,.\label{BF_15.1}\end{aligned}$$ The terms on the rhs of (\[BF\_15.1\]) are scheme independent and agree with the ones of [@Ellwanger_Hirsch_Weber_1996]. To solve the system (\[BF\_16\]) we rewrite it as a matrix equation: $$\bar{\eta}_{k}=\left(\bar{A}+\mathbf{M}\,\bar{\eta}_{k}\right)g_{k}^{2}\,,\label{BF_17}$$ where we defined: $$\bar{\eta}_{k}=\left(\begin{array}{c} \eta_{a,k}\\ \eta_{c,k} \end{array}\right)\qquad\qquad\bar{A}=\left(\begin{array}{c} A\\ A' \end{array}\right)\qquad\qquad\mathbf{M}=\left(\begin{array}{cc} B & C\\ B' & C' \end{array}\right)\,.\label{BF_18}$$ The linear system (\[BF\_17\]) is easily solved: $$\bar{\eta}_{k}=\left(1-g_{k}^{2}\,\mathbf{M}\right)^{-1}\bar{A}\,,\label{BF_20}$$ or more explicitly: $$\begin{aligned} \eta_{a,k} & = & \frac{A(1-C'g_{k}^{2})+ACg_{k}^{2}}{(1-Bg_{k}^{2})(1-C'g_{k}^{2})+B'Cg_{k}^{4}}g_{k}^{2}\nonumber \\ \eta_{c,k} & = & \frac{A'(1-Bg_{k}^{2})+AB'g_{k}^{2}}{(1-Bg_{k}^{2})(1-C'g_{k}^{2})+B'Cg_{k}^{4}}g_{k}^{2}\,.\label{BF_21}\end{aligned}$$ We can now turn back to the beta function for the gauge coupling and “close” it by inserting the functions (\[BF\_21\]) in (\[BF\_7\]). Under the condition $C=C'=0$, we find the following RG improved result for the beta function of the gauge coupling: ![Bi–field beta functions from (\[BF\_22\]). Asymptotes from the left: Padè, type II and type I beta functions. Now the type I and type II beta function have the asymptotes in the same place. The one–loop beta function (without asymptote) is shown for comparison.](Beta_gauge_II.pdf) $$\partial_{t}g_{k}=\frac{a+c\, g_{k}^{2}+d\, g_{k}^{4}}{1-b\, g_{k}^{2}}g_{k}^{3}\,,\label{BF_22}$$ with coefficients: $$c=\frac{1}{2}(bA+b'A'-2aB)\qquad\qquad d=\frac{b'}{2}(AB'-A'B)\,.\label{BF_23}$$ Using the values (\[BF\_8\]) and (\[BF\_16.1\]) we find: $$c_{I}=\frac{16}{9}\frac{N^{2}}{(4\pi)^{4}}\qquad\qquad d_{I}=\frac{5}{108}\frac{N^{3}}{(4\pi)^{6}}\,,$$ $$c_{II}=-\frac{8}{3}\frac{N^{2}}{(4\pi)^{4}}\qquad\qquad d_{II}=\frac{5}{108}\frac{N^{3}}{(4\pi)^{6}}\,.\label{BF_24}$$ Note that only the coefficients $c$ are different in the two schemes. The beta function (\[BF\_22\]) clearly shows the influence of the fluctuation couplings, here represented by $Z_{a,k}$ and $Z_{c,k}$, on the flow of the gauge coupling $g_{k}$. We remember that the anomalous dimensions we used to obtain the closed form for the beta function of $g_{k}$ was calculated in the gauge $\alpha=1$, in a future work [@Codello_in_preparation] we will study the dependence on the gauge-fixing parameter, considering in particular the Landau gauge *$\alpha=0$.* Beta functions of the form (\[BF\_22\]) are obtained as Padé approximations of the perturbative beta function [@Gockeler_Horsley_Irving_Pleiter_Rakow_Schierholz_Stuben_2006]. By employing data at three–loops one finds a beta function of the form (\[BF\_22\]) with coefficients given by: $$a_{\textrm{Padè}}=-b_{1}\qquad\qquad b_{\textrm{Padè}}=\frac{b_{2}}{b_{1}}\qquad\qquad c_{\textrm{Padè}}=-b_{1}+\frac{b_{0}b_{2}}{b_{1}}\qquad\qquad d_{\textrm{Padè}}=0\,,$$ where $b_{1},b_{2},b_{3}$ are the one–, two– and three–loop coefficients. The bi–field beta functions are compared with the Padè re-summed one in Figure 10. When making this comparisons one has to remember that starting from the three–loop coefficients these become scheme dependent even when employing massless renormalization schemes. Finally, the coefficient $d$ becomes non-zero when one considers data at four–loops, thus our beta function (\[BF\_22\]) contains re-summed information up to four–loops. ### Flow and physical observables We now integrate the beta functions derived in the previous subsection. We will be able to do these steps analytically. We start to consider the single–field beta function (\[BF\_14\]); a simple integration of this gives: ![Implicit plot for the single–field flow of $g_{k}$ from (\[IF\_2\]). From top: RS, type II, type I and one-loop flows. ](Beta_g_t_std.pdf) $$t=-\frac{1}{2a}\left(\frac{1}{g_{k}^{2}}-\frac{1}{g_{0}^{2}}\right)-\frac{b}{a}\log\frac{g_{k}}{g_{0}}\,,\label{IF_2}$$ where $t=\log\frac{k}{k_{0}}$, with $k_{0}$ a reference scale where $g_{k}=g_{k_{0}}$. Equation (\[IF\_2\]) is transcendental and cannot be inverted analytically except in the one–loop case $b=0$. In this last case one obtains: ![The mass gap as a function of $g_{0}$ (with $k_{0}=91.19$ GeV) for the single–field flows compared to the flow proposed by RS from (\[IF\_2.3\]). From top-left the RS, type II and type I curves.](Beta_Mass_Gap_Std.pdf) $$g_{k}^{2}=\frac{g_{0}^{2}}{1-2a\log\frac{k}{k_{0}}}\,.\label{IF_2.2}$$ Since $a=-\frac{11}{3}\frac{N}{(4\pi)^{2}}<0$, in the UV limit $k\rightarrow\infty$ the theory is asymptotically free, i.e. $g_{k}\rightarrow0$. When $b\neq0$ we can make an implicit plot of $t(g_{k})$ using (\[IF\_2\]); the result for the three cases (type I, type II and RS) together with the one–loop case are shown in Figure 11. One sees that the type II and RS flows are very similar. The improved flows appear to be convex functions (contrary to the perturbative result) and have a minimum at the value where the beta function (\[BF\_14\]) has the pole, i.e. at $g_{k}^{2}=\frac{1}{b}$. Note the important fact that even if the beta functions have a pole, the flow they generate is instead smooth. But one cannot invert $t(g_{k})$ to obtain $g_{k}(t)$ for every value of $t\in\mathbb{R}$. If we choose to be continuously connected to the asymptotic free regime, when we lower the RG scale toward the IR we reach the point $t(b^{-1/2})$ after which we cannot proceed further, i.e. the curve $g_{k}(t)$ cannot be extended. This signals that our single coupling truncation breaks down (and one needs to consider more sophisticated truncations to proceed further toward the IR [@Ellwanger_Hirsch_Weber_1998]). Following [@Sannino_Schechter_2010] we can interpret $t(b^{-1/2})$ as the point where bound states are formed and identify the relative mass scale $M$ as (twice) the mass gap. With this identification, using (\[IF\_2\]), we find the following relation for the mass gap as a function of $k_{0}$ and $g_{0}$: $$M=k_{0}\left(g_{0}^{2}b\right)^{\frac{b}{2a}}e^{\frac{1}{2a}\left(\frac{1}{g_{0}^{2}}-b\right)}\,.\label{IF_2.3}$$ This estimate for the mass gap is plotted in Figure 12. For $k_{0}=91.19$ GeV and $N=3$ one finds $M_{I}=0.61$ GeV and $M_{II}=1.79$ GeV, to be compared with the estimate $M_{RS}=1.69$ GeV of [@Sannino_Schechter_2010]. As for the beta functions, the type II estimate and the one based on the RS beta function are similar, while a strong scheme dependence is observed in the type I case. In the same way one can integrate the bi–field beta function (\[BF\_22\]) to obtain the following analytical relation: ![Implicit plot for the bi–field flow of $g_{k}$ from (\[IF\_4\]). From the top: RS, type II, type I and one-loop flows. Note that the type I and II are closer than in Figure 11.](Beta_g_t_new.pdf) $$\begin{aligned} t & = & -\frac{1}{2a}\left(\frac{1}{g_{k}^{2}}-\frac{1}{g_{0}^{2}}\right)-\frac{ab+c}{a^{2}}\log\frac{g_{k}}{g_{0}}+\frac{ab+c}{4a^{2}}\log\frac{a+cg_{k}^{2}+dg_{k}^{4}}{a+cg_{0}^{2}+dg_{0}^{4}}\nonumber \\ & & +\frac{abc-2ad+c^{2}}{4a^{2}\sqrt{c^{2}-4ad}}\log\frac{\left(c+2dg_{k}^{2}-\sqrt{c^{2}-4ad}\right)\left(c+2dg_{0}^{2}+\sqrt{c^{2}-4ad}\right)}{\left(c+2dg_{k}^{2}+\sqrt{c^{2}-4ad}\right)\left(c+2dg_{0}^{2}-\sqrt{c^{2}-4ad}\right)}\,.\label{IF_4}\end{aligned}$$ The implicit plot of (\[IF\_4\]) is shown in Figure 13. Obviously if we set $c=d=0$ we recover (\[IF\_2\]). As before, we can estimate the mass gap $M$, the results is now: $$\begin{aligned} M & = & k_{0}\left(g_{0}^{2}b\right)^{\frac{ab+c}{2a^{2}}}e^{\frac{1}{2a}\left(\frac{1}{g_{0}^{2}}-b\right)}\left(\frac{a+c/b+d/b^{2}}{a+cg_{0}^{2}+dg_{0}^{4}}\right)^{\frac{ab+c}{4a^{2}}}\nonumber \\ & & \times\left[\frac{\left(c+2d/b-\sqrt{c^{2}-4ad}\right)\left(c+2dg_{0}^{2}+\sqrt{c^{2}-4ad}\right)}{\left(c+2d/b+\sqrt{c^{2}-4ad}\right)\left(c+2dg_{0}^{2}-\sqrt{c^{2}-4ad}\right)}\right]^{\frac{abc-2ad+c^{2}}{4a^{2}\sqrt{c^{2}-4ad}}}\,.\label{IF_5}\end{aligned}$$ Evaluating (\[IF\_5\]) for the different cutoff choices gives Figure 14. For $k_{0}=91.19$ GeV and $N=3$ one now finds the estimates $M_{I}=0.54$ GeV and $M_{II}=0.85$ GeV, which are nearer to each other but smaller than the RS or the lattice predictions. ![The mass gap $M$ as a function of $g_{0}$ (with $k_{0}=91.19$ GeV) for the (from top-left) RS and bi–field flows.](Beta_Mass_Gap_New.pdf) To conclude, we discuss the RG scale dependence of a physical observable $\mathcal{O}_{k}(g_{k})$. In particular, if this is RG invariant $\frac{d}{dt}\mathcal{O}_{k}(g_{k})=0$, then it’s flow is governed by the beta function $\beta(g_{k})=\partial_{t}g_{k}$ of the coupling constant. If the observable has dimension $d_{\mathcal{O}}$ the we can set $\mathcal{O}_{k}(g_{k})=k^{d_{\mathcal{O}}}\tilde{\mathcal{O}}(g_{k})$ and obtain: $$d_{\mathcal{O}}\tilde{\mathcal{O}}(g_{k})+\beta(g_{k})\partial_{g}\tilde{\mathcal{O}}(g_{k})=0\,.\label{NP_1}$$ It’s easy to integrate (\[NP\_1\]) to find: $$\mathcal{O}_{k}=\mathcal{O}_{k_{0}}e^{-d_{\mathcal{O}}\int_{g_{0}}^{g_{k}}\frac{dg}{\beta(g)}}\,,\label{NP_2}$$ where $t=\int_{g_{0}}^{g_{k}}\frac{dg}{\beta(g)}$. If we insert (\[IF\_2\]) into (\[NP\_2\]) we find: $$\mathcal{O}_{k}=\mathcal{O}_{k_{0}}\left(\frac{g_{k}}{g_{0}}\right)^{d_{\mathcal{O}}\frac{b}{a}}e^{\frac{d_{\mathcal{O}}}{2a}\left(\frac{1}{g_{k}^{2}}-\frac{1}{g_{0}^{2}}\right)}\,.\label{NP_3}$$ One notices that for any value of $d_{\mathcal{O}}$ the observable is a smooth function of $g_{k}$ even if the beta function has a pole. A similar, but more complex relation, can be obtained starting from (\[IF\_4\]): $$\begin{aligned} \mathcal{O}_{k} & = & \mathcal{O}_{k_{0}}\left(\frac{g_{k}}{g_{0}}\right)^{d_{\mathcal{O}}\frac{ab+c}{a^{2}}}e^{\frac{d_{\mathcal{O}}}{2a}\left(\frac{1}{g_{k}^{2}}-\frac{1}{g_{0}^{2}}\right)}\left(\frac{a+cg_{k}^{2}+dg_{k}^{4}}{a+cg_{0}^{2}+dg_{0}^{4}}\right)^{d_{\mathcal{O}}\frac{ab+c}{4a^{2}}}\nonumber \\ & & \times\left[\frac{\left(c+2dg_{k}^{2}-\sqrt{c^{2}-4ad}\right)\left(c+2dg_{0}^{2}+\sqrt{c^{2}-4ad}\right)}{\left(c+2dg_{k}^{2}+\sqrt{c^{2}-4ad}\right)\left(c+2dg_{0}^{2}-\sqrt{c^{2}-4ad}\right)}\right]^{d_{\mathcal{O}}\frac{abc-2ad+c^{2}}{4a^{2}\sqrt{c^{2}-4ad}}}\,.\label{NP_4}\end{aligned}$$ Even in this case observables are smooth functions of the coupling constant. The simplest observable is the invariant scale $M$; we can recover (\[IF\_5\]) by setting $d_{\mathcal{O}}=1$, $g_{k}^{2}=1/b$, $\mathcal{O}_{k}=M$ and $\mathcal{O}_{k_{0}}=k_{0}$ in (\[NP\_4\]). Other possible observables have different canonical dimensions, but in general observables turn out to be smooth functions of the coupling constant in spite of the singular behavior of the RG improved (or re-summed) beta functions. Thus this kind of beta functions are consistent approximations to the RG flow that can be applied to the study of the IR physics of non-abelian gauge theories. Discussion and future perspectives ================================== In this paper we introduced a method that enables the projection of the RG flow equations of a large new class of truncations of the background effective average action (bEAA). Our method is based on the explicit momentum space representation of the hierarchy of flow equations satisfied by the proper-vertices of the bEAA. The key step in our construction is the determination of the explicit momentum space form of vertices with background legs, since these are related to functional derivatives of the cutoff action and must be treated with care. We showed how these vertices indeed have a simple form related to the vertices of the cutoff action, the action defined by having the cutoff operator as Hessian, and contains finite difference derivatives of the cutoff shape function. We also gave simple diagrammatic rules that provide a representation of the flow equations for the proper-vertices of the bEAA that can be used in actual computations. Our method is very general and can be applied to a variety of new interesting truncations of the bEAA, with applications to non-abelian gauge theories, nonlinear sigma models, gravity and membranes; or any other theory characterized by local gauge symmetries. As a first application, we studied a bi–field truncation of the bEAA for $SU(N)$ non-abelian gauge theories. Employing bi–field truncations we can explore the dependency of the bEAA on all its arguments: fluctuation and background fields. We considered the three-dimensional subspace of theory space parametrized by the three anomalous dimensions $\{\eta_{\bar{A},k},\eta_{a,k},\eta_{c,k}\}$. We showed how all the results obtained with the aid of the local heat kernel expansion, usually employed in this framework, are reproduced by our technique. The structure of the exact flow equation for the bEAA implies that $\eta_{\bar{A},k}$ is linearly related to both $\eta_{a,k}$ and $\eta_{c,k}$; in the background field method, the beta function of the gauge coupling is related to $\eta_{\bar{A},k}$, thus to obtain $\partial_{t}g_{k}$ one needs $\eta_{a,k}$ and $\eta_{c,k}$. Usually one imposes $\eta_{a,k}=\eta_{\bar{A},k}$ and $\eta_{c,k}=0$, in order to treat the problem by using only a single–field truncation. Instead, we calculated $\eta_{a,k}$ and $\eta_{c,k}$ directly and showed how they can be determined as a function of $g_{k}$ by solving a linear system; in this way we obtained a new RG improved form for the beta function of the gauge coupling. We then discussed some phenomenological implications of the flow so obtained, which has similarities with re-summed perturbative beta functions. As a future extension of this study, one can consider more general bi–field truncations by allowing the gauge fluctuation mass and gauge-fixing parameter to run [@Codello_in_preparation]. In another application one can study bi–field truncations in the context of quantum gravity and check how Asymptotic Safety, up to now observed in single–field truncations, is robust to the bi–field extension [@Codello_D'Odorico_Pagani_2013]. Still, as work done in the non-background formalism [@Ellwanger_Hirsch_Weber_1998] has shown, the non-trivial part of the bEAA has a non-local structure. With the method presented here we can project a large class of non-local truncations. In a gauge invariant formalism, the finite part of the effective action is encoded in “structure functions”, which are generally functions of the covariant Laplacian, and act on local invariants. As an example, in non-abelian gauge theories, one can consider the following ansatz for the gauge invariant part of the bEAA: $$\bar{\Gamma}_{k}[A]=\int d^{d}x\, F_{\mu\nu}^{a}f_{k}(-D^{2})F^{a\mu\nu}+O(F^{3})\,,\label{C_1}$$ where $f_{k}$ is a running structure function which encodes non-trivial physical information. By applying our algorithm we can obtain a RG flow equation for the running structure function, which will be of the form: $$\partial_{t}f_{k}=\mathcal{F}_{R_{k}}^{d}\left(g_{k},f_{k},f_{k}',f_{k}''\right)\,.\label{C_1.1}$$ By following the flow of the running structure function $f_{k}$ from the UV, where the theory is asymptotically free, down to the IR, we can obtain the full finite non-local effective action, from which we can extract relevant physical information. Similar applications can be made in the context of gravity, extending the results obtained using the non-local heat kernel expansion [@Codello_2010], in the context of non-linear sigma models or in the context membrane theory. For these reasons, among others, the computational strategy based on the hierarchy of flow equations for the zero-field proper-vertices of the bEAA, when combined with the momentum space rules that we derived, is a promising tool for further studies of the bEAA in its full generality. Acknowledgments {#acknowledgments .unnumbered} =============== I would like to thank M. Reuter for stimulating discussions and O. Zanusso, M. Demmel, R. Percacci for careful reading the manuscript. Momentum space representation of background vertices ==================================================== In this appendix we give the details of the derivation of the momentum space reprentation of vertices of the bEAA with background legs. Perturbative expansion of the heat kernel ----------------------------------------- We shortly review the perturbative expansion for the heat kernel as developed in [@Codello_Zanusso_2012], to which we refer for more details. This expansion will be used in the next subsection to derive the momentum space representation of background vertices. The heat kernel $K^{s}(x,y)$ satisfies the following partial differential equation with boundary condition: $$\left(\partial_{s}+\Delta_{x}\right)K^{s}(x,y)=0\qquad\qquad K^{0}(x,y)=\delta(x-y)\,,\label{HK_1}$$ where $\Delta=-D_{\mu}D^{\mu}+U$ is the Laplacian operator constructed using the covariant derivatives $D_{\mu}$ and $U$ is a potential term. We decompose the the Laplacian as the sum of a “non-interacting” Laplacian $-\partial^{2}$ and of an interaction $V$ in the following way: $$\Delta=-\partial^{2}+V\,.\label{HK_PT_1}$$ The potential $V$ contains $U$ and all terms proportional to the gauge connection $A_{\mu}$. Two examples are the flat space Laplacian $\Delta=-\partial^{2}+U$, where simply $V=U$, or the abelian gauge Laplacian where $V$ contains all terms that vanish for $A_{\mu}=0$: $$\Delta=-D_{\mu}D^{\mu}=-\left(\partial_{\mu}+iA_{\mu}\right)\left(\partial^{\mu}+iA^{\mu}\right)=-\partial^{2}\underbrace{-2A_{\mu}\partial^{\mu}-\partial_{\mu}A^{\mu}+A_{\mu}A^{\mu}}_{V}\,.$$ To derive the perturbative series for the heat kernel we need first to calculate the heat kernel[^13] $K_{0,xy}^{s}$ of the operator $-\partial^{2}$, around which we will perform the expansion. From equation (\[HK\_1\]) we see that it satisfies the following equation with boundary condition: $$\left(\partial_{s}-\partial_{x}^{2}\right)K_{0,xy}^{s}=0\qquad\qquad K_{0,xy}^{0}=\delta_{xy}\,;\label{HK_PT_2}$$ the solution of (\[HK\_PT\_2\]) is the standard Gaussian: $$K_{0,xy}^{s}=\frac{1}{(4\pi s)^{d/2}}e^{-\frac{(x-y)^{2}}{4s}}\,.\label{HK_PT_5}$$ Equation (\[HK\_PT\_5\]) is the solution around which we will construct the perturbative expansion. Note that the heat kernel (\[HK\_PT\_5\]) satisfies: $$K_{0,xy}^{s_{1}+s_{2}}=\int_{z}K_{0,xz}^{s_{1}}K_{0,zy}^{s_{2}}\,.\label{HK_PT_5.1}$$ To derive the perturbative expansion of $K_{xy}^{s}$ around $K_{0,xy}^{s}$ we define the operator[^14] $U_{xy}^{s}=\int_{z}K_{0,xz}^{-s}K_{zy}^{s}$. Using (\[HK\_1\]) and (\[HK\_PT\_2\]) we find it satisfies the following equation: $$\begin{aligned} \partial_{s}U_{xy}^{s} & = & \int_{z}\left[\partial_{s}K_{0,xz}^{-s}K_{zy}^{s}+K_{0,xz}^{-s}\partial_{s}K_{zy}^{s}\right]\nonumber \\ & = & \int_{z}\left[K_{0,xz}^{-s}(-\partial_{z}^{2})K_{zy}^{s}-K_{0,xz}^{-s}\Delta_{z}K_{zy}^{s}\right]\nonumber \\ & = & -\int_{z}K_{0,xz}^{-s}V_{z}K_{zy}^{s}\nonumber \\ & = & -\int_{zw}K_{0,xz}^{-s}V_{z}K_{0,zw}^{s}U_{wy}^{s}\,.\label{HK_PT_6}\end{aligned}$$ We know that equation (\[HK\_PT\_6\]) is solved by Dyson’s series: $$U_{xy}^{s}=T\exp\left\{ -\int_{0}^{s}dt\,\int_{z}K_{0,xz}^{-t}V_{z}K_{0,zy}^{t}\right\} \,,$$ where the exponential is time-ordered with respect to $s$. Using (\[HK\_PT\_5.1\]) we immediately find: $$K_{xy}^{s}=\int_{z}K_{0,xz}^{s}\, T\exp\left\{ -\int_{0}^{s}dt\,\int_{w}K_{0,zw}^{-t}V_{w}K_{0,wy}^{t}\right\} \,.\label{HK_PT_7}$$ Rescaling the integration variable in (\[HK\_PT\_7\]) as $t\rightarrow t/s$ and using (\[HK\_PT\_5.1\]) gives the final formula for the perturbative expansion of the heat kernel: $$\begin{aligned} K_{xy}^{s} & = & \int_{z}K_{0,xz}^{s}\, T\exp\left\{ -s\int_{0}^{1}dt\int_{w}K_{0,zw}^{-st}V_{w}K_{0,wy}^{st}\right\} \nonumber \\ & = & K_{0,xy}^{s}-s\int_{0}^{1}dt\,\int_{z}K_{0,xz}^{s(1-t)}V_{z}\, K_{0,zy}^{st}+\nonumber \\ & & +s^{2}\int_{0}^{1}dt_{1}\int_{0}^{t_{1}}\, dt_{2}\,\int_{zw}K_{0,xz}^{s(1-t_{1})}V_{z}\, K_{0,zw}^{s(t_{1}-t_{2})}V_{w}\, K_{0,wy}^{st_{2}}+O(V^{3})\,.\label{HK_PT_8}\end{aligned}$$ In the following section we will use this relation to find an explicit momentum space representation for the functional derivatives of the cutoff action. Derivation of momentum space representation ------------------------------------------- In this section we derive the momentum space representation of background vertices, in particular we show how to obtain relations (\[gauge\_D\_1.1\]) and (\[gauge\_D\_1.2\]) used in section 2.3 to obtain the momentum space representation of the flow equations for the zero-field proper-vertices of the bEAA. We need to calculate the momentum space representation of the cutoff vertices $\Delta S_{k}^{(2;1)}[0;0]$ and $\Delta S_{k}^{(2;2)}[0;0]$. From the definition of the cutoff action (\[gauge\_2\]) we see that[^15]: $$\Delta S_{k}^{(2;0)}[0;\bar{A}]_{xy}^{AB}=\int_{zw}\delta_{xz}R_{k}[\bar{A}]_{zw}^{AB}\delta_{wy}=R_{k}[\bar{A}]_{xy}^{AB}\,.\label{gauge_MSR_1}$$ The cutoff kernel $R_{k}[\bar{A}]$ is a function of the cutoff operator $L^{(2;0)}[0;\bar{A}]$, constructed as the Hessian of the cutoff operator action $L[a;\bar{A}]$. For example, if the cutoff operator is the gauge Laplacian, then: $$L^{(2;0)}[0;\bar{A}]_{xy}=\int_{z}\bar{D}_{z\mu}\delta_{zx}\bar{D}_{z}^{\mu}\delta_{zy}=-\int_{z}\delta_{zx}\bar{D}_{z\mu}\bar{D}_{z}^{\mu}\delta_{zy}=-\bar{D}_{\mu x}\bar{D}^{\mu y}\delta_{xy}=-\bar{D}_{x}^{2}\delta_{xy}\,.$$ Thus, after making a Laplace transform, we can write the cutoff kernel in terms of the heat kernel of the cutoff operator[^16]: $$R_{k}[\bar{A}]_{xy}^{AB}=\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)\, K^{s}[\bar{A}]_{xy}^{AB}\,,\label{gauge_MSR_2}$$ where the heat kernel can be written in terms of the Hessian of the cutoff operator action: $$K^{s}[\bar{A}]=\exp\left\{ -sL^{(2;0)}[0;\bar{A}]\right\} \,.\label{gauge_MSR_3}$$ Inserting (\[gauge\_MSR\_2\]) in equation (\[gauge\_MSR\_1\]) and setting the background field to zero after differentiating with respect to it one time, gives the following representation for the cutoff vertex with one background leg in terms of the heat kernel: $$\Delta S_{k}^{(2;1)}[0;0]_{xyz}^{ABC}=\left.\frac{\delta R_{k}[\bar{A}]_{xy}^{AB}}{\delta\bar{A}_{z}^{C}}\right|_{\bar{A}=0}=\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)\left.\frac{\delta K^{s}[\bar{A}]_{xy}^{AB}}{\delta\bar{A}_{z}^{C}}\right|_{\bar{A}=0}\,.\label{gauge_MSR_4}$$ We can now use the perturbative expansion for the heat kernel obtained in the previous section, equation (\[HK\_PT\_8\]), to write the last term of (\[gauge\_MSR\_4\]) as follows: $$\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)\left.\frac{\delta K^{s}[\bar{A}]_{xy}^{AB}}{\delta\bar{A}_{z}^{C}}\right|_{\bar{A}=0}=\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)(-s)\int_{0}^{1}dt\, K_{0,xw}^{s(1-t)}L^{(2;1)}[0;0]_{wuz}^{ABC}K_{0,uy}^{st}\,.\label{gauge_MSR_5}$$ In (\[gauge\_MSR\_5\]) we omitted to explicitly write the coordinate integrals and we wrote the flat space heat kernels as $K^{s}[0]_{xy}^{AB}=K_{0,xy}^{s}\delta^{AB}$, where $K_{0,xy}^{s}$ is given in equation (\[HK\_PT\_5\]). Going to momentum space and inserting (\[gauge\_MSR\_5\]) in (\[gauge\_MSR\_4\]) gives the following representation for the cutoff vertex: $$\begin{aligned} \Delta S_{k}^{(2;1)}[0;0]_{p_{1},p_{2},p_{3}}^{ABC} & = & \int_{0}^{\infty}ds\,\tilde{R}_{k}(s)(-s)\int_{0}^{1}dt\, K_{0,p_{1}}^{s(1-t)}[l_{p_{1},p_{2},p_{3}}^{(2;1)}]^{ABC}K_{0,p_{2}}^{st}\nonumber \\ & = & \int_{0}^{\infty}ds\,\tilde{R}_{k}(s)(-s)\int_{0}^{1}dt\, e^{-s(1-t)p_{1}^{2}}[l_{p_{1},p_{2},p_{3}}^{(2;1)}]^{ABC}e^{-stp_{2}^{2}}\nonumber \\ & = & [l_{p_{1},p_{2},p_{3}}^{(2;1)}]^{ABC}\int_{0}^{1}dt\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)(-s)e^{-s(1-t)p_{1}^{2}-stp_{2}^{2}}\,,\label{gauge_MSR_6}\end{aligned}$$ where the cutoff operator vertices are defined as: $$l_{x_{1}...x_{n}y_{1}...y_{m}}^{(n;m)}\equiv L^{(n;m)}[0;0]_{x_{1}...x_{n}y_{1}...y_{m}}\,.$$ Here we used the following simple momentum space representation for the flat space heat kernel $K_{0,p}^{s}=e^{-sp^{2}}$. It is left to evaluate the double integral in (\[gauge\_MSR\_6\]). This can be done with the aid of the $Q$-functionals (see the appendix A of [@Codello_Percacci_Rahmede_2009]): $$Q_{n}[h](z+a)=\int_{0}^{\infty}ds\,\tilde{h}(s)s^{-n}e^{-sa}\,,\label{gauge_MSR_6.01}$$ $$Q_{n}[h](z+a)=\left\{ \begin{array}{ccc} \frac{1}{\Gamma(n)}\int_{0}^{\infty}dz\, z^{n-1}h(z+a) & & n>0\\ (-1)^{n}h^{(n)}(a) & & n\leq0 \end{array}\right..\label{gauge_MSR_6.1}$$ Using (\[gauge\_MSR\_6.1\]) we find: $$\begin{aligned} \int_{0}^{\infty}ds\,\tilde{R}_{k}(s)(-s)e^{-s(1-t)p_{1}^{2}-stp_{2}^{2}} & = & -Q_{-1}[R_{k}(z+s(1-t)p_{1}^{2}+stp_{2}^{2})]\nonumber \\ & = & R_{k}'(s(1-t)p_{1}^{2}+stp_{2}^{2})\,.\label{gauge_MSR_7}\end{aligned}$$ Now the parameter integral is easily evaluated: $$\int_{0}^{1}dt\, R_{k}'(s(1-t)p_{1}^{2}+stp_{2}^{2})=\frac{R_{k}(p_{2}^{2})-R_{k}(p_{1}^{2})}{p_{2}^{2}-p_{1}^{2}}\,.\label{gauge_MSR_8}$$ If we introduce the first finite-difference derivative, defined as $$f_{p_{1},p_{2}}^{(1)}=\frac{f(p_{2}^{2})-f(p_{1}^{2})}{p_{2}^{2}-p_{1}^{2}}\,,$$ we can finally write, for the cutoff vertex with one external background leg (\[gauge\_MSR\_6\]), the following momentum space representation: $$\Delta S_{k}^{(2;1)}[0;0]_{p_{1},p_{2},p_{3}}^{ABC}=[l_{p_{1},p_{2},p_{3}}^{(2;1)}]^{ABC}R_{p_{1},p_{2}}^{(1)}\,.\label{gauge_MSR_9}$$ We just need now to consider (\[gauge\_MSR\_9\]) with the momentum values $p_{1}=q$, $p_{2}=-q-p$ and $p_{3}=p$ to prove the relation given in equation (\[gauge\_D\_1.1\]): $$\Delta S_{k}^{(2;1)}[0;0]_{q,-q-p,p}^{ABC}=[l_{q,-q-p,p}^{(2;1)}]^{ABC}R_{q+p,q}^{(1)}\,.\label{gauge_MSR_10}$$ Along the same lines we can derive the momentum space representation for the cutoff vertex with two external background legs. In place of (\[gauge\_MSR\_4\]) we have now $$\Delta S_{k}^{(2;2)}[0;0]_{xyzw}^{ABCD}=\left.\frac{\delta^{2}R_{k}[\bar{A}]_{xy}^{AB}}{\delta\bar{A}_{w}^{D}\delta\bar{A}_{z}^{C}}\right|_{\bar{A}=0}=\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)\left.\frac{\delta^{2}K^{s}[\bar{A}]_{xy}^{AB}}{\delta\bar{A}_{w}^{D}\delta\bar{A}_{z}^{C}}\right|_{\bar{A}=0}\,.\label{gauge_MSR_11}$$ Using the perturbative expansion (\[HK\_PT\_8\]) gives the following expansion for the second functional derivative of the cutoff kernel: $$\begin{aligned} \left.\frac{\delta^{2}R_{k}[\bar{A}]_{xy}^{AB}}{\delta\bar{A}_{w}^{D}\delta\bar{A}_{z}^{C}}\right|_{\bar{A}=0} & = & \int_{0}^{\infty}ds\,\tilde{R}_{k}(s)\,\left.\frac{\delta^{2}K^{s}[\bar{A}]_{xy}^{AB}}{\delta\bar{A}_{w}^{D}\delta\bar{A}_{z}^{C}}\right|_{\bar{A}=0}\nonumber \\ & = & \int_{0}^{\infty}ds\,\tilde{R}_{k}(s)(-s)\int_{0}^{1}dt\, K_{0,xu}^{s(1-t)}L^{(2;2)}[0;0]_{uvzw}^{ABCD}K_{0,vy}^{st}\nonumber \\ & & +2\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)\, s^{2}\int_{0}^{1}dt_{1}\int_{0}^{t_{1}}dt_{2}\, K_{0,xu}^{s(1-t_{1})}L^{(2;1)}[0;0]_{uvz}^{AMB}\nonumber \\ & & \times K_{0,vr}^{s(t_{1}-t_{2})}L^{(2;1)}[0;0]_{rtw}^{CMD}K_{0,ty}^{st_{2}}\,.\label{gauge_MSR_12}\end{aligned}$$ When we insert (\[gauge\_MSR\_12\]) into (\[gauge\_MSR\_11\]) and shift to momentum space, the first contribution in (\[gauge\_MSR\_12\]), like in the previous case, takes the form: $$[l_{p_{1},p_{2},p_{3},p_{4}}^{(2;2)}]^{ABCD}R_{p_{4},p_{1}}^{(1)}\,.\label{gauge_MSR_13}$$ The second contribution takes instead the following form: $$[l_{p_{1},-p_{1}-p_{2},p_{2}}^{(2;1)}]^{AMB}[l_{p_{3},p_{1}+p_{2},p_{4}}^{(2;1)}]^{CMD}\qquad\qquad\qquad$$ $$\qquad\qquad\qquad\times\int_{0}^{1}dt_{1}\int_{0}^{t_{1}}dt_{2}\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)\, s^{2}\, e^{-s(1-t_{1})p_{1}^{2}-s(t_{1}-t_{2})(p_{1}+p_{2})^{2}-s_{2}t_{2}p_{4}^{2}}\,.\label{gauge_MSR_14}$$ We can calculate the double integral in (\[gauge\_MSR\_14\]) using the properties of the $Q$-functionals as before: $$\int_{0}^{\infty}ds\,\tilde{R}_{k}(s)\, s^{2}\, e^{-s(1-t_{1})p_{1}^{2}-s(t_{1}-t_{2})(p_{1}+p_{2})^{2}-s_{2}t_{2}p_{4}^{2}}=\qquad\qquad$$ $$\qquad\qquad=Q_{-2}[R_{k}(z+s(1-t_{1})p_{1}^{2}+s(t_{1}-t_{2})(p_{1}+p_{2})^{2}+s_{2}t_{2}p_{4}^{2})]$$ $$\qquad=R_{k}''(s(1-t_{1})p_{1}^{2}+s(t_{1}-t_{2})(p_{1}+p_{2})^{2}+s_{2}t_{2}p_{4}^{2})\,,\label{gauge_MSR_15}$$ and $$2\int_{0}^{1}dt_{1}\int_{0}^{t_{1}}dt_{2}\, R_{k}''(s(1-t_{1})p_{1}^{2}+s(t_{1}-t_{2})(p_{1}+p_{2})^{2}+s_{2}t_{2}p_{4}^{2})=\qquad\qquad$$ $$\qquad\qquad=\frac{2}{(p_{1}+p_{2})^{2}-p_{4}^{2}}\left[\frac{R_{k}((p_{1}+p_{2})^{2})-R_{k}(p_{1}^{2})}{(p_{1}+p_{2})^{2}-p_{1}^{2}}-\frac{R_{k}(p_{4}^{2})-R_{k}(p_{1}^{2})}{p_{4}^{2}-p_{1}^{2}}\right]\,.\label{gauge_MSR_16}$$ Finally, inserting (\[gauge\_MSR\_16\]) in (\[gauge\_MSR\_14\]) and combining with (\[gauge\_MSR\_13\]), gives the following momentum space representation for (\[gauge\_MSR\_11\]): $$\Delta S_{k}^{(2;2)}[0;0]_{p_{1},p_{2},p_{3},p_{4}}^{ABCD}=[l_{p_{1},p_{2},p_{3},p_{4}}^{(2;2)}]^{ABCD}R_{p_{4},p_{1}}^{(1)}+\qquad\qquad$$ $$\qquad\qquad+[l_{p_{1},-p_{1}-p_{2},p_{2}}^{(2;1)}]^{AMB}[l_{p_{3},p_{1}+p_{2},p_{4}}^{(2;1)}]^{CMD}\frac{2}{(p_{1}+p_{2})^{2}-p_{4}^{2}}\left[R_{p_{1}+p_{2},p_{1}}^{(1)}-R_{p_{4},p_{1}}^{(1)}\right]\,.\label{gauge_MSR_17}$$ To recover relation (\[gauge\_D\_1.2\]) we set $p_{1}=-p_{2}=q$ and $p_{3}=-p_{4}=p$ in (\[gauge\_MSR\_17\]) so that: $$\Delta S_{k}^{(2;2)}[0;0]_{q,-q,p,-p}^{ABCD}=[l_{q,-q,p,-p}^{(2;2)}]^{ABCD}R_{p}'+[l_{q,-q-p,p}^{(2;1)}]^{AMB}[l_{q+p,-q,-p}^{(2;1)}]^{CMD}R_{p+q,q}^{(2)}\,,\label{gauge_MSR_18}$$ where $R_{p}'\equiv R'(p^{2})$. In (\[gauge\_MSR\_18\]) we defined the second finite-difference derivative of a function as: $$f_{p+q,q}^{(2)}=\frac{2}{(p+q)^{2}-p^{2}}\left[f_{p+q,p}^{(1)}-f'_{p}\right]\,.\label{gauge_MSR_19}$$ This concludes the derivation of relations (\[gauge\_D\_1.1\]) and (\[gauge\_D\_1.2\]) needed in section 2.3 to write the explicit momentum space representation for the flow equations of the zero-field proper-vertices of the bEAA. Basic relations for non-abelian gauge theories ============================================== In this appendix we review our conventions and we collect the relevant formulae for functional variations and derivatives that are used in the paper. Definitions and conventions --------------------------- We consider the Lie groups $SU(N)$; we pick up a representation such that the group elements are represented by matrices $R=e^{-\theta}$, where $\theta=-i\theta^{a}t^{a}$ are the (infinitesimal) group parameters, the indices $a,b,...$ run from one to $\dim SU(N)=N^{2}-1$ and the $\dim R\times\dim R$ matrices $t^{a}$ are the generators of the Lie algebra of $SU(N)$ in the given representation. The generators satisfy the following commutation relations $\left[t^{a},t^{b}\right]=if^{abc}t^{c}$; in a general representation the structure constants $f^{abc}$ are antisymmetric in the first two indices $f^{abc}=-f^{bac}$. The covariant derivative is defined as $D_{\mu}=\partial_{\mu}+gA_{\mu}$, where $g$ is the gauge coupling constant and $A_{\mu}$ is the Lie algebra valued connection. The components of the connection are defined by $A_{\mu}=-iA_{\mu}^{a}t^{a}$ and one may write the covariant derivative as: $$D_{\mu}=\partial_{\mu}-iA_{\mu}^{a}t^{a}\,.\label{BNA_1.1}$$ The gauge field strength $F_{\mu\nu}$ is the curvature of the gauge connection; it can be defined as the commutator of covariant derivatives acting on matter fields $\phi$, $[D_{\mu},D_{\nu}]\phi=F_{\mu\nu}\phi$, with $F_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}+[A_{\mu},A_{\nu}]$ the field strength. In component form $F_{\mu\nu}=-iF_{\mu\nu}^{a}t^{a}$ and we have: $$F_{\mu\nu}^{a}=\partial_{\mu}A_{\nu}^{a}-\partial_{\nu}A_{\mu}^{a}+g\, f^{abc}A_{\mu}^{b}A_{\nu}^{c}\,.\label{BNA_4}$$ In the adjoint representation the structure constants are related to the generators $f^{abc}=i(t_{\textrm{ad}}^{c})^{ab}$ and are completely antisymmetric and we have $\dim\textrm{ad}=\dim SU(N)$. In this representation the covariant derivative (\[BNA\_1.1\]) becomes: $$D_{\mu}^{ab}=\partial_{\mu}\delta^{ab}-iA_{\mu}^{c}(t_{\textrm{ad}}^{c})^{ab}=\partial_{\mu}\delta^{ab}+g\, f^{acb}A_{\mu}^{c}\,.\label{BNA_5}$$ We also have the following relation relating the commutator of the covariant derivatives and the field strength $[D_{\mu},D_{\nu}]^{ab}=-f^{abc}F_{\mu\nu}^{c}$. Under an infinitesimal gauge transformation $R\approx1-\theta$ matter fields and the connection transform as: $$\delta_{\theta}\phi=[\phi,\theta]\qquad\delta_{\theta}A_{\mu}=D_{\mu}\theta\qquad\delta_{\theta}F_{\mu\nu}=\left[F_{\mu\nu},\theta\right]\,.\label{BNA_6}$$ In components we have instead[^17]: $$\delta_{\theta}\phi^{i}=i\theta^{a}(t_{\textrm{R}}^{a})^{ij}\phi^{j}\qquad\delta_{\theta}A_{\mu}^{a}=D_{\mu}^{ab}\theta^{b}=\partial_{\mu}\theta^{a}+f^{abc}A_{\mu}^{b}\theta^{c}\qquad\delta_{\theta}F_{\mu\nu}^{a}=i\theta^{c}(t_{\textrm{ad}}^{c})^{ab}F_{\mu\nu}^{b}=f^{abc}F_{\mu\nu}^{b}\theta^{c}\,.\label{BNA_7}$$ where $i,j,...$ to the dimension of the representation of the matter fields. Variations and functional derivatives ------------------------------------- We consider the functional: $$I[A]=\frac{1}{4}\int d^{d}x\, F_{\mu\nu}^{a}F^{a\mu\nu}\,,\label{gauge_V_1}$$ which can be easily shown to be gauge invariant $\delta_{\theta}I[A]=0$. Expanding (\[gauge\_V\_1\]) around a background configuration $A_{\mu}=\bar{A}_{\mu}+a_{\mu}$ gives the following: $$\begin{aligned} I[\bar{A}+a] & = & I[\bar{A}]+\int d^{d}x\,\bar{F}^{a\mu\nu}(\bar{D}_{\mu}a_{\nu})^{a}\nonumber \\ & & +\frac{1}{2}\int d^{d}x\, a_{\mu}^{a}\left[(-\bar{D}^{2})^{ab}g^{\mu\nu}+2f^{abc}F^{c\mu\nu}+\bar{D}^{ac\mu}\bar{D}^{cb\nu}\right]a_{\nu}^{b}\nonumber \\ & & +g\, f^{abc}\int d^{d}x\,(\bar{D}_{\mu}a_{\nu})^{a}\, a_{\mu}^{b}a_{\nu}^{c}+\frac{1}{4}g^{2}f^{abc}f^{ade}\int d^{d}x\, a^{b\mu}a^{c\nu}a_{\mu}^{d}a_{\nu}^{e}\,.\label{gauge_V_2}\end{aligned}$$ One may use $2f^{abc}F^{c\mu\nu}=2iF^{c\mu\nu}(t_{\textrm{ad}}^{c})^{ab}=-2F^{\mu\nu}$ to write the differential operator in the quadratic term of (\[gauge\_V\_2\]) as $-\bar{D}^{2}g^{\mu\nu}-2F^{\mu\nu}+\bar{D}^{\mu}\bar{D}^{\nu}$. Note that the background gauge-fixing action (\[gauge\_5\]) and the background ghost action (\[gauge\_6\]) are already in their varied form since they are by construction quadratic in the fields. We now calculate the functional derivatives of the functional (\[gauge\_V\_1\]), of the gauge-fixing action (\[gauge\_5\]) and of the ghost action (\[gauge\_6\]) that are needed in the flow equations for both the bEAA and the gEAA. The Hessian of $I[A]$ at $A_{\mu}=0$ becomes: $$[I_{p_{1},p_{2}}^{(2)}]^{\alpha\beta\, ab}=(2\pi)^{d}\delta_{p_{1}+p_{2}}\delta^{ab}\left[-g^{\alpha\beta}(p_{1}\cdot p_{2})+\frac{1}{2}\left(p_{1}^{\alpha}p_{2}^{\beta}+p_{1}^{\beta}p_{2}^{\alpha}\right)\right]\,.\label{gauge_FD_4}$$ Note that in both (\[gauge\_FD\_4\]), as in all the following formulas, the indices and the momentum variables are related in the precise order they appear. The third functional derivative of (\[gauge\_V\_1\]), which gives the gauge three-vertex, is: $$[I_{p_{1},p_{2},p_{3}}^{(3)}]^{\alpha\beta\gamma\, abc}=(2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}}igf^{abc}\left[g^{\alpha\beta}(p_{2}-p_{1})^{\gamma}+g^{\beta\gamma}(p_{3}-p_{2})^{\alpha}+g^{\gamma\alpha}(p_{1}-p_{3})^{\beta}\right]\,.\label{gauge_FD_5}$$ The fourth functional derivative of (\[gauge\_V\_1\]), which gives the gauge four-vertex, is: $$\begin{aligned} [I_{p_{1},p_{2},p_{3},p_{4}}^{(4)}]^{\alpha\beta\gamma\delta\, abcd} & = & (2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}+p_{4}}g^{2}\left[f^{eab}f^{ecd}\left(g^{\alpha\gamma}g^{\beta\delta}-g^{\alpha\delta}g^{\beta\gamma}\right)\right.\nonumber \\ & & +f^{eac}f^{ebd}\left(g^{\alpha\beta}g^{\gamma\delta}-g^{\alpha\delta}g^{\beta\gamma}\right)\nonumber \\ & & \left.+f^{ead}f^{ebc}\left(g^{\alpha\beta}g^{\gamma\delta}-g^{\alpha\gamma}g^{\beta\delta}\right)\right]\,.\label{gauge_FD_6}\end{aligned}$$ We will use the vertices (\[gauge\_FD\_5\]) and (\[gauge\_FD\_6\]) to construct the zero-field proper vertices of the bEAA with both fluctuation and background legs. This is possible, since for a gauge invariant action as is (\[gauge\_V\_1\]), the following property holds: $$\frac{\delta I[A]}{\delta A_{\mu}^{a}}=\frac{\delta I[\bar{A}+a]}{\delta a_{\mu}^{a}}=\frac{\delta I[\bar{A}+a]}{\delta\bar{A}_{\mu}^{a}}\,.\label{gauge_FD_7}$$ Next, we need the vertices coming from the gauge-fixing action (\[gauge\_5\]), which in component form reads: $$S_{gf}[a;\bar{A}]=\frac{1}{2\alpha}\int d^{d}x\,(\partial_{\mu}a^{a\mu}+g\, f^{abc}\bar{A}_{\mu}^{b}a^{c\mu})(\partial_{\nu}a^{a\nu}+g\, f^{ade}\bar{A}_{\nu}^{d}a^{e\nu})\,,\label{gauge_FD_7.1}$$ The second functional derivative of (\[gauge\_FD\_7.1\]) with respect to the fluctuation field gives rise to: $$[S_{gf\; p,-p}^{(2;0)}]^{\alpha\beta\, ab}=-(2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}}\frac{1}{\alpha}\delta^{ab}p_{1}^{\alpha}p_{2}^{\beta}\,.\label{gauge_FD_8}$$ The mixed functional derivatives give the fluctuation-fluctuation-background vertex: $$[S_{gf\; p_{1},p_{2},p_{3}}^{(2;1)}]^{\alpha\beta\gamma\, abc}=(2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}}\,\frac{1}{\alpha}igf^{abc}\left(g^{\alpha\gamma}p_{2}^{\beta}-g^{\beta\gamma}p_{1}^{\alpha}\right)\,.\label{gauge_FD_9}$$ Four mixed functional derivatives give the fluctuation-fluctuation-background-background vertex: $$[S_{gf\; p_{1},p_{2},p_{3},p_{4}}^{(2;2)}]^{\alpha\beta\gamma\delta\, abcd}=(2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}+p_{4}}\frac{1}{\alpha}g^{2}\left(f^{ace}f^{bde}g^{\alpha\gamma}g^{\beta\delta}+f^{ade}f^{bce}g^{\alpha\delta}g^{\beta\gamma}\right)\,.\label{gauge_FD_10}$$ The vertices (\[gauge\_FD\_9\]) and (\[gauge\_FD\_10\]) will be used in section 3.2.2. The ghost action (\[gauge\_6\]), when written out explicitly, reads: $$S_{gh}[a,\bar{c},c;\bar{A}]=\int d^{d}x\left(\partial_{\mu}\bar{c}^{a}+f^{abc}\bar{A}_{\mu}^{b}\bar{c}^{c}\right)\left(\partial^{\mu}c^{a}+f^{ade}\bar{A}^{d\mu}c^{e}+gf^{ade}a^{d\mu}c^{e}\right)\,.\label{gauge_FD_11}$$ Note that (\[gauge\_FD\_11\]) generates the three-vertices ghost-ghost-fluctuation and ghost-ghost-background but only the four-vertex ghost-ghost-background-background. The two three-vertices differ by a factor of two since the background field enters both covariant derivatives while the fluctuation field does not. We have: $$[S_{gh\; p_{1},p_{2},p_{3}}^{(1,1,1;0)}]^{\alpha\, abc}=-(2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}}igf^{abc}p_{2}^{\alpha}\label{gauge_FD_12}$$ and $$[S_{gh\; p_{1},p_{2},p_{3}}^{(0,1,1;1)}]^{\gamma\, abc}=(2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}}if^{abc}(p_{2}-p_{1})^{\gamma}\,.\label{gauge_FD_13}$$ Finally the four-vertex is: $$[S_{gh\; p_{1},p_{2},p_{3},p_{4}}^{(0,1,1;2)}]^{\gamma\delta\, abcd}=(2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}+p_{4}}g^{\gamma\delta}\left(f^{eac}f^{ebd}+f^{ead}f^{ebc}\right)\,.\label{gauge_FD_14}$$ The gauge cutoff operator action in the type I case is $$L[a;\bar{A}]=\frac{1}{2}\int d^{d}x\,\bar{D}_{\mu}a_{\nu}\,\bar{D}^{\mu}a^{\nu}\,;\label{gauge_FD_15}$$ its vertices are: $$\begin{aligned} [L_{p_{1},p_{2},p_{3}}^{(2;1)}]^{\alpha\beta\gamma\, abc} & = & (2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}}if^{abc}g^{\alpha\beta}(p_{2}-p_{1})^{\gamma}\nonumber \\ {}[L_{p_{1},p_{2},p_{3},p_{4}}^{(2;2)}]^{\alpha\beta\gamma\delta\, abcd} & = & (2\pi)^{d}\delta_{p_{1}+p_{2}+p_{3}+p_{4}}g^{\alpha\beta}g^{\gamma\delta}\left(f^{eac}f^{ebd}+f^{ead}f^{ebc}\right)\,.\label{gauge_CA_3}\end{aligned}$$ The ghost cutoff operator action is: $$L[\bar{c},c;\bar{A}]=\int d^{d}x\,\bar{D}_{\mu}\bar{c}\,\bar{D}^{\mu}c\,;\label{gauge_CA_4}$$ since $L[\bar{c},c;\bar{A}]=S_{gh}[0,\bar{c},c;\bar{A}]$ we have: $$\begin{aligned} [L_{p_{1},p_{2},p_{3}}^{(1,1;1)}]^{\alpha\, abc} & = & [S_{gh\; p_{1},p_{2},p_{3}}^{(0,1,1;1)}]^{\alpha\, abc}\nonumber \\ {}[L_{p_{1},p_{2},p_{3},p_{4}}^{(1,1;2)}]^{\gamma\delta\, abcd} & = & [S_{gh\; p_{1},p_{2},p_{3},p_{4}}^{(0,1,1;2)}]^{\gamma\delta\, abcd}\,.\label{gauge_CA_5}\end{aligned}$$ With these we derived all variations and vertices that we use in sections 3.2.2 and 3.2.3. [10]{} J. Berges, N. Tetradis and C. Wetterich, Phys. Rep. 363 (2002) 223. H. Gies, Lect. Notes Phys. 852 (2012) 287, hep-ph/0611146. M. Reuter and C. Wetterich, Nucl. Phys. B 417 (1994) 181. M. Reuter, Phys. Rev. D 53 (1996) 4430, hep-th/9511128; M. Reuter, hep-th/9602012; M. Reuter, Mod. Phys. Lett. A 12 (1997) 2777, hep-th/9604124. A. Codello and R. Percacci, Phys. Lett. B 672 (2009) 280, arXiv:0810.0715; R. Percacci and O. Zanusso, Phys. Rev. D 81 (2010) 065012, arXiv:0910.0851; M. Fabbrichesi, R. Percacci, A. Tonero and O. Zanusso, Phys. Rev. D 83 (2011) 025016, arXiv:1010.0912; R. Flore, A. Wipf and O. Zanusso, arXiv:1207.4499. M. Reuter, Phys. Rev. D 57 (1998) 971, hep-th/9605030. A. Codello and O. Zanusso, Phys. Rev. D 83 (2011) 125021, arXiv:1103.1089; A. Codello, N. Tetradis and O. Zanusso, arXiv:1212.4073. A. Codello, R. Percacci and C. Rahmede, Annals Phys. 324 (2009) 414, arXiv:0805.2909. D. Benedetti, K. Groh, P.F. Machado and F. Saueressig, JHEP 1106 (2011) 079, arXiv:1012.3081. A. Codello, Annals Phys. 325 (2010) 1727, arXiv:1004.2171; A. Satz, A. Codello and F.D. Mazzitelli, Phys. Rev. D 82 (2010) 084011, arXiv:1006.3808; A. Codello, New J. Phys. 14 (2012) 015009, arXiv:1108.1908. E. Manrique and M. Reuter, Annals Phys. 325 (2010) 785, arXiv:0907.2617; E. Manrique, M. Reuter and F. Saueressig, Annals Phys. 326 (2011) 463, arXiv:1006.0099; E. Manrique, M. Reuter and F. Saueressig, Annals Phys. 326 (2011) 440, arXiv:1003.5129. A. Codello and O. Zanusso, J. Math. Phys. 54 (2013) 013513, arXiv:1203.2034. T.A. Ryttov and F. Sannino, Phys. Rev. D 78 (2008) 065001, arXiv:0711.3745. M. Reuter and C. Wetterich, Nucl. Phys. B 391 (1993) 147; M. Reuter and C. Wetterich, Nucl. Phys. B 408 (1993) 91; M. Reuter and C. Wetterich, Nucl. Phys. B 427 (1994) 291. L.F. Abbott, Nucl. Phys. B 185 (1981) 189; L.F. Abbott, Acta Phys. Polon. B 13 (1982) 33. U. Ellwanger, Phys. Lett. B 335 (1994) 364, hep-th/9402077; D.F. Litim and J.M. Pawlowski, Nucl. Phys. Proc. Suppl. 74 (1999) 325, hep-th/9809020; D.F. Litim and J.M. Pawlowski, hep-th/9901063; D.F. Litim and J.M. Pawlowski, Phys. Lett. B 546 (2002) 279, hep-th/0208216; F. Freire, D.F. Litim and J.M. Pawlowski, Phys. Lett. B 495 (2000) 256, hep-th/0009110. P.M. Lavrov and I.L. Shapiro, arXiv:1212.2577. M. Reuter and C. Wetterich, hep-th/9411227; M. Reuter and C. Wetterich, Phys. Rev. D 56 (1997) 7893, hep-th/9708051; H. Gies, Phys. Rev. D 66 (2002) 025006, hep-th/0202207; H. Gies, Phys. Rev. D 68 (2003) 085015, hep-th/0305208. A. Codello, R. Percacci and C. Rahmede, Int. J. Mod. Phys. A 23 (2008) 143-150, arXiv:0705.1769; P. Machado and F. Saueressig, Phys. Rev. D 77 (2008) 124045, arXiv:0712.0445. U. Ellwanger, M. Hirsch and A. Weber, Z. Phys. C 69 (1996) 687, hep-th/9506019. A. Codello, in preparation. D. J. Gross and F. Wilczek, Phys. Rev. Lett. 30 (1973) 1343; H. D. Politzer, Phys. Rev. 30 (1973) 1346. A. Nink and M. Reuter, JHEP 1301 (2013) 062, arXiv:1208.0031. V.A. Novikov, M.A. Shifman, A.I. Vainshtein and V.I. Zakharov, Nucl. Phys. B 229 (1983) 381. C. Pica and F. Sannino, Phys. Rev. D 83 (2011) 116001, arXiv:1011.3832. O.J. Rosten, arXiv:1106.2544. M. Gockeler, R. Horsley, A.C. Irving, D. Pleiter, P.E.L. Rakow, G.Schierholz and H.Stuben, Phys. Rev. D 73 (2006) 014513, hep-ph/0502212. F. Sannino and J. Schechter, Phys. Rev. D 82 (2010) 096008, arXiv:1009.0265. A. Codello, G. D’Odorico and C. Pagani, Phys. Rev. D 89 (2014) 8, 081701, arXiv:1304.4777. U. Ellwanger, M. Hirsch and A. Weber, Eur. Phys. J. C 1 (1998) 563, hep-ph/9606468; L. von Smekal, R. Alkofer and A. Hauck, Phys. Rev. Lett. 79 (1997) 3591, hep-ph/9705242; R. Alkofer and L. von Smekal, Phys. Rept. 353 (2001) 281, hep-ph/0007355; J.M. Pawlowski, D.F. Litim, S. Nedelko and L. von Smekal, Phys. Rev. Lett. 93 (2004) 152002, hep-th/0312324; C.S. Fischer and H. Gies, JHEP 0410 (2004) 048, hep-ph/0408089; C.S. Fischer and J.M. Pawlowski, Phys. Rev. D 75 (2007) 025012, hep-th/0609009; C.S. Fischer, A. Maas and J.M. Pawlowski, Annals Phys. 324 (2009) 2408, arXiv:0810.1987; C.S. Fischer and J.M. Pawlowski, Phys. Rev. D 80 (2009) 025023, arXiv:0903.2193; L. Fister and J.M. Pawlowski, arXiv:1301.4163. [^1]: email: [email protected] [^2]: Functional derivatives of a functionals depending on many arguments are indicated by the notation $\Gamma^{(n_{1},n_{2},...)}[....]$. [^3]: See appendix B for our conventions. [^4]: Here and in other equations of this section we omit, for clarity, to write explicitly the arguments of the functionals when these are understood. [^5]: Note that $\tilde{\Gamma}_{k}[\varphi;\bar{A}]$ is the actual Legendre transform of the functional generator of connected correlation functions $W_{k}[J;\bar{A}]$. [^6]: We define $\int_{x}\equiv\int d^{d}x$ and $\int_{q}\equiv\int\frac{d^{d}q}{(2\pi)^{d}}$. [^7]: See [@Codello_2010] for a first one–loop application in the context of gravity. [^8]: Again we used $g_{k}Z_{\bar{A},k}^{1/2}=1$. [^9]: Note that $\textrm{tr}\, U^{2}\equiv U^{ab\mu\nu}U_{\nu\mu}^{ba}$. [^10]: In this subsection, as in the next we omit to write the scale index $k$ on functionals and on couplings to simplify the notation. [^11]: In (\[gauge\_P\_3\]), we are introducing in the cutoff kernel the gauge-fixing parameter, if this is running it will give rise to additional terms on the rhs of the flow equation, generated by $\partial_{t}R_{k}$ , proportional to $\partial_{t}\alpha_{k}$. A similar choice has been made in [@Ellwanger_Hirsch_Weber_1996]. [^12]: Note that $L[\bar{c},c;\bar{A}]=S_{gh}[0,\bar{c},c;\bar{A}]$. [^13]: Here and in the following we use the compact notation $K_{xy}^{s}\equiv K^{s}(x,y)$ and $\delta_{xy}\equiv\delta^{(d)}(x-y)$. We also define $\int_{x}\equiv\int d^{d}x$ and $\int_{q}\equiv\int\frac{d^{d}q}{(2\pi)^{d}}$. [^14]: Not to be confused with the potential $U$. [^15]: In the gravitational case one must additionally care about factors of $\sqrt{g}$. [^16]: For simplicity we consider $R_{k}^{AB}$ having tensor structure proportional to $\delta^{AB}$. [^17]: Note that $-iA_{\mu}^{c}(t_{\textrm{ad}}^{c})^{ab}\theta^{b}=i\theta^{c}(t_{\textrm{ad}}^{c})^{ab}A_{\mu}^{b}$.
{ "pile_set_name": "ArXiv" }
--- abstract: 'The spectra of plasma and magnetoplasma excitations in a two-dimensional system of anisotropic heavy fermions were investigated for the first time. The spectrum of microwave absorption by disk-like samples of stressed AlAs quantum wells at low electron densities showed two plasma resonances separated by a frequency gap. These two plasma resonances correspond to electron mass principle values of $(1.10 \pm 0.05) m_0$ and $(0.20 \pm 0.01) m_0$. The observed results correspond to the case of a single valley strongly anisotropic Fermi surface. It was established that electron density increase results in population of the second valley, manifesting itself as a drastic modification of the plasma spectrum. We directly determined the electron densities in each valley and the inter-valley splitting energy from the ratio of the two plasma frequencies.' author: - 'V. M. Muravev$^{a}$, A. R. Khisameeva$^{a,b}$, V. N. Belyanin$^{a,b}$, I. V. Kukushkin$^{a}$, L. Tiemann$^{c}$, C. Reichl$^{c}$, W. Dietsche$^{c}$, W. Wegscheider$^{c}$' date: - - title: 'Magnetoplasma excitations of two-dimensional anisotropic heavy fermions in AlAs quantum wells' --- The last few decades have witnessed a surge in research on the fascinating and often unexpected collective states arising from strong electron-electron interaction. Selectively doped semiconductor heterostructures appeared to be nearly ideal systems for such research owing to their strongly reduced disorder. Examples of phenomena caused by electron correlations include the fractional quantum Hall effect [@Tsui:82], metal-insulator transition [@Pudalov], and spin-textured structures [@Spin]. The electron-electron interaction strength is characterized by the ratio of the Coulomb interaction energy to the Fermi energy, and is proportional to the effective mass of charge carriers. This has motivated interest in new two-dimensional electron systems (2DES) that show heavy fermions behavior. One of the most promising materials of choice is n-AlAs [@Shayegan:06]. AlAs is an indirect gap semiconductor in which the electrons occupy three equivalent valleys at the $X$ points of the Brillouin zone. This degeneracy is lifted when the electrons are confined to a 2D layer. In a quantum well grown on a GaAs $(001)$ wafer, only the in-plane $[100]$ ($X$) and $[010]$ ($Y$) valleys are occupied for well widths greater than $5$ nm [@Maezawa:92]. This differs from Si $(001)$ metal oxide semiconductor field-effect transistors (MOSFETs), in which the two valleys with the out-of-plane major axes are occupied. Such unusual behavior stems from biaxial compression of the AlAs layer, induced by lattice mismatch between the AlAs and GaAs. Moreover, the residual in-plane strain lifts the $X$ and $Y$ valley degeneracy, leading to inter-valley energy splitting $\Delta E$ (Fig. 1) [@Ando; @Shayegan:02; @Shayegan:05; @Grayson:11]. This splitting modifies the plasma spectrum, as observed in the present paper. Transport measurements revealed large and anisotropic AlAs electron effective masses $m_{\rm l}=(1.1 \pm 0.1) m_0$ and $m_{\rm tr}=(0.20 \pm 0.02) m_0$ [@Smith:87; @CR:93; @Mass:04], corresponding to the longitudinal and transverse Fermi ellipsoid axes directions, respectively. The effective Lande $g$-factor of electrons in bulk AlAs ($g^{\ast} = 2$) is much larger than in GaAs ($g^{\ast} = -0.44$). These characteristics make AlAs heterostructure 2DES unique and versatile subject of many-body and valleytronics phenomena investigations. Microwave magnetospectroscopy is the most direct method to characterize Fermi surfaces and determine effective masses [@Dresselhaus:54]. This method has revealed well-studied 2D plasma excitations in isotropic single component GaAs heterostructure 2DES [@Stern:67; @Allen:77]. All attempts to study plasma dynamics in anisotropic 2DES, however, have been limited to experiments in which an applied magnetic field creates a small anisotropy in the an initially isotropic 2DES [@Barke:86; @Kozlov1; @Kozlov2]. The only example of collective behavior in a multi-component 2DES was found in GaAs double quantum wells occupied by electrons in one well and holes in the other [@Kukushkin:11]. Heretofore, 2D plasmon physics in systems with native strong mass anisotropy, as well as multi-component 2DES, has not been well explored, despite a number of interesting physical predictions [@Chaplik]. AlAs 2DES provide an important research opportunity by combining strong anisotropy with the ability to tune carrier density in each valley. Measurements were carried out on high-quality $15$ nm AlAs quantum well heterostructures fabricated by molecular beam epitaxy (MBE) on a $(001)$ GaAs substrate. The electron density $n_s$ and electron mobility $\mu$ were in the ranges of $1{.}7\times 10^{11} - 2{.}4\times 10^{11}\,\text{cm}^{-2}$ and $1{.}2\times 10^{5} - 2{.}0\times 10^{5} \,\text{cm}^{2}/\text{V} \cdot \text{s}$, respectively. A variation of the electron density was achieved by short illumination of the sample. The illumination was performed by a green light emitting diode ($2.2$ eV) at $T = 1.6$ K. A coplanar waveguide (CPW) was fabricated on top of the crystal surface using standard photolithography tools. The waveguide contained a central $1{.}1$ mm wide stripe spaced $0{.}6$ mm from the grounded planes (Fig. 1 inset). The total length of the coplanar waveguide was $4$ mm. The parameters of the waveguide were chosen to provide a characteristic impedance of $Z_0 = 50$ $\Omega$. Six equidistant 2DES disks of diameter $d=0{.}5$ mm were fabricated in the slots of the CPW. The disk centers were spaced $1{.}5$ mm apart to minimize cross-talk effects. Arrows indicate the basic crystallography directions in Fig. 1. We detected the resonant absorption of microwave probe radiation ($f=1-40$ GHz) propagating along the CPW. An alternating electric field concentrated in the slots of the CPW oscillates the 2D plasmas in the disks. Resonant absorption of microwaves occurs whenever a plasmon is excited in a disk. The sample was immersed in a cryostat with a superconducting coil. $50$ $\Omega$ coaxial cables connected the sample in series between a microwave generator ($f=0 - 40$ GHz) and a tunnel diode with a preamplifier placed outside the cryostat. The output power of the generator did not exceed $100$ nW and the output signal was detected by a standard lock-in technique. The magnetic field was applied normal to the surface of the sample. Helium vapor was pumped out to attain a temperature of $T = 1.6$ K. Figure 1 shows the magnetic-field dependencies of the coplanar waveguide transmission for several microwave frequencies. The horizontal axis is located at the signal level when no microwave radiation is supplied to the CPW. Each curve shows a resonance with respect to zero magnetic field. Most of the resonances are symmetric, the rest have an asymmetric line-shape. Observed asymmetric resonances in the transmission can be treated and analyzed as Fano-type resonances (see Supplementary Material  [@Fano]). The resonance shifts to lower magnetic fields with increasing microwave frequency $f$, indicating the edge magnetoplasma (EMP) nature of the observed resonance. For frequencies above $15$ GHz, a second plasma resonance arises (inset in Fig. 1). The resonance behavior exhibits the positive magnetodispersion characteristic of a cyclotron magnetoplasmon. For more transmission curves we refer to Fig. 1S in the Supplementary Material . The resonance origins are best identified by plotting the resonant magnetic-fields versus microwave frequencies, as shown in Fig. 2. The data were obtained at an electron density of $1{.}7\times 10^{11}\,\text{cm}^{-2}$. The magnetodispersion has two branches separated by a frequency gap. The low-frequency branch corresponds to an edge magnetoplasmon propagating along the edge of the disk. This is a mode with anomalously weak attenuation that propagates in a narrow strip near the edge of the 2DES [@Volkov:88; @Allen:83]. The EMP frequency decreases as $\omega_{-} \approx \sigma_{xy} q \propto n_s q/B$ in the strong magnetic field limit. The high-frequency branch has a positive magnetodispersion. The electric field $\vec{E}$ aligned along the $[1\bar{1}0]$ (Fig. 1) crystal direction can be factorized into two components along the Fermi ellipsoid axes as $\vec{E}=\vec{E}_{\rm l} + \vec{E}_{\rm tr}$. In the $B=0$ T limit, each of these components excites a separate 2D plasma wave with corresponding masses $m_{\rm l}$ and $m_{\rm tr}$. Therefore, the gap in the magnetoplasmon spectrum of the disk vividly demonstrates the highly anisotropic nature of the Fermi surface in AlAs 2DES [@Dahl; @Kozlov2]. For the sake of comparison, we performed the same measurements on a geometrically identical sample made from GaAs quantum well ($d=0.5$ mm, $n_s = 1{.}4\times 10^{11}\,\text{cm}^{-2}$). The inset in Fig. 2 shows that for $B=0$ T the magnetic field edge and cyclotron magnetoplasma modes are degenerate, highlighting the mass $m^{\ast}=0.067 m_0$ isotropism in GaAs. The plasma excitation spectrum in a 2DES with mass anisotropy can be described using the dipole approximation as [@Dahl; @Heitmann; @Margulis]: $$\omega_{\rm l,\rm tr}=\frac{1}{2} \left[ \sqrt{(\Omega_{\rm tr}+\Omega_{\rm l})^2+\omega_c^2} \pm \sqrt{(\Omega_{\rm tr}-\Omega_{\rm l})^2+\omega_c^2} \right], \label{1}$$ where $\Omega_{\rm l}$ and $\Omega_{\rm tr}$ are plasma frequencies along the main crystallographic directions for $B=0$ T, and $\omega_c=eB/m_c$ is the cyclotron frequency. The cyclotron mass is determined as a geometric mean of effective masses along the crystallographic axes, $m_c=\sqrt{m_{\rm l} m_{\rm tr}}$. The frequencies $\Omega_{\rm l, \rm tr}$ obey the 2D-plasmon dispersion [@Stern:67]: $$\Omega_{\rm l,\rm tr}^2=\frac{n_s e^2}{2 m_{\rm l,\rm tr} \varepsilon_0 \varepsilon^{\ast}} q, \label{2}$$ where $\varepsilon^{\ast}=(\varepsilon_{\rm GaAs}+1)/2$ is the effective dielectric permittivity of the surrounding medium and $q=2{.}4/d$ is the wave vector for the disk geometry [@Kukushkin:03]. From our experiments, we measure zero-field plasma frequencies $\Omega_{\rm l}=(6.5 \pm 0.2)$ GHz and $\Omega_{\rm tr}=(15.3 \pm 0.5)$ GHz. Using Eq. (2) we find the effective masses in the AlAs quantum well along the main crystallographic directions to be $m_{\rm l}=(1.10 \pm 0.05) m_0$ and $m_{\rm tr}=(0.20 \pm 0.01) m_0$. These mass values agree with results obtained from commensurability oscillation measurements [@Shayegan:06]. Figure 3(a) shows CPW microwave transmission as a function of magnetic field for the identical samples with 2DES densities of $1{.}7\times 10^{11}\,\text{cm}^{-2}$ and $2{.}4\times 10^{11}\,\text{cm}^{-2}$. A short light flash from a light emitting diode varied the electron density. The magnetoplasma resonance shifted to larger magnetic field values with increased electron density. This is consistent with Eq ($1-2$). However, the zero-field plasma frequencies determined from the detailed magnetodispersion curve of the $2{.}4\times 10^{11}\,\text{cm}^{-2}$ 2DES have a ratio $\Omega_{\rm tr}/\Omega_{\rm l} = (1.80 \pm 0.05)$. This number is inconsistent with Eq. (\[2\]), which predicts $\Omega_{\rm tr}/\Omega_{\rm l} = \sqrt{m_{\rm l}/m_{\rm tr}} = (2.3 \pm 0.1)$. This suggests that the plasma dynamics undergo a qualitative metamorphosis when the electron density changes. We attribute observed phenomenon to the energy splitting between the $X$ and $Y$ valleys. Indeed, the residual in-plane strain lifts the $X$ and $Y$ valley degeneracy, leading to an inter-valley energy splitting $\Delta E$ (Fig. 1). For a 2DES where $n_s=1{.}7\times 10^{11}\,\text{cm}^{-2}$, we find that all electrons occupy only the $X$ valley, leaving the $Y$ valley empty. As we increase the density, some of the electrons begin to fill the $Y$ valley (Fig. 3(b)). The total density is then defined as $n_s=n_x+n_y$, where $n_x$ and $n_y$ are the charge carrier concentrations in the $X$ and $Y$ valleys respectively. The collective plasma excitations in such a system could be considered using a two-component anisotropic plasma model [@Vitlina]. The plasma frequencies along the $[100]$ and $[010]$ directions are described by the following expressions: $$\Omega_{[100]}^2=\frac{e^2 q}{2 \varepsilon_0 \varepsilon^{\ast}} \left( \frac{n_x}{m_{\rm l}}+\frac{n_{y}}{m_{\rm tr}} \right), \label{3}$$ $$\Omega_{[010]}^2=\frac{e^2 q}{2 \varepsilon_0 \varepsilon^{\ast}} \left( \frac{n_x}{m_{\rm tr}}+\frac{n_{y}}{m_{\rm l}} \right). \label{4}$$ Using these expressions with our obtained plasma frequencies $\Omega_{[100]}$, $\Omega_{[010]}$ and AlAs masses $m_{\rm l}$, $m_{\rm tr}$, we deduced the densities $n_x$ and $n_y$ in each of the valleys to be $n_x=(2{.}10 \pm 0.05)\times 10^{11}\,\text{cm}^{-2}$ and $n_y=(0{.}30 \pm 0.05)\times 10^{11}\,\text{cm}^{-2}$. From the difference of densities $\Delta n = n_x - n_y$, we can directly determine the inter-valley energy splitting $\Delta E$ using the 2D density of states: $$\Delta E=\frac{\pi \hbar^2 \Delta n}{\sqrt{m_{\rm l}m_{\rm tr}}}. \label{5}$$ This calculation gives $\Delta E=(0.90 \pm 0.05)$ meV, which is consistent with all previous studies of valley splitting in AlAs [@Ando; @Shayegan:02; @Shayegan:05]. However, these studies were conducted in strong magnetic fields ($B > 1$ T), leaving the low magnetic field range unexplored. Our experiments directly determine the valley populations from the plasma frequencies in the weak magnetic field limit. The solid lines in Figs. 3(b) and (c) represent theoretical predictions of Eq. (1) with $\Omega_{\rm l} = \Omega_{[100]}$ and $\Omega_{\rm tr} = \Omega_{[010]}$ for the mode magnetodispersion. Some discrepancy exists between the experimental data and theory at moderate magnetic fields. The difference is especially pronounced for the high-frequency cyclotron magnetoplasma mode (Fig. 3(c)). One possible explanation could be hybridization between the cyclotron magnetoplasma mode and an inter-component cyclotron mode. The additional inter-component mode is a dispersionless excitation with frequency $\omega=eB/\sqrt{m_{\rm l} m_{\rm tr}}$, which coincides with the cyclotron resonance [@Chaplik]. Another possible explanation of the observed discrepancy is that $\Omega_{[100]}$, $\Omega_{[010]}$ and the corresponding $\Delta E$ are not constant with the $B$-field. This discrepancy should motivate further research. The agreement between theory and experiment returns in the limit of strong magnetic fields. Our results suggest opportunities for future applications of plasmonics in AlAs 2DESs. They can be used to study the inter-valley energy spacing at $B=0$ T using the discovered plasma resonance method. Such experiments could unveil roles of electron-electron interaction in semiconductor valley-splitting [@Ando; @Shayegan:02; @Shayegan:05]. They could also be used to study novel relativistic plasma excitations [@Muravev]. Relativistic plasma waves in highly anisotropic two-component electron liquids of AlAs may reveal unpredicted physical phenomena. The authors gratefully acknowledge financial support from the Russian Scientific Fund (Grant No. 14-12-00693). [14]{} D. C. Tsui, H. L. Stormer and A. C. Gossard, Phys. Rev. Lett. [**48**]{}, 1559 (1982). S. V. Kravchenko, G. V. Kravchenko, J. E. Furneaux, V. M. Pudalov and M. D’lorio, Phys. Rev. B [**50**]{}, 8039 (1994). X. Z. Yu, Y. Onose, N. Kanazawa *et.* *al*, Nature [**465**]{}, 901 (2010). M. Shayegan, E. P. De Poortere, O. Gunawan, Y. P. Shkolnikov, E. Tutuc and K. Vakili, Phys. Status Solidi B [**243**]{}, 3629 (2006). K. Maezawa, T. Mizutani, and S. Yamada, J. Appl. Phys. [**71**]{}, 296 (1992). T. Ando, A. B. Fowler, and F. Stern, Rev. Mod. Phys. [**54**]{}, 437 (1982). Y. P. Shkolnikov, E. P. De Poortere, E. Tutuc, and M. Shayegan, Phys. Rev. Lett. [**89**]{}, 226805 (2002). Y. P. Shkolnikov, S. Misra, N. C. Bishop, E. P. De Poortere, and M. Shayegan, Phys. Rev. Lett. [**95**]{}, 066809 (2005). S. Prabhu-Gaunkar, S. Birner, S. Dasgupta, C. Knaak, and M. Grayson, Phys. Rev. B [**84**]{}, 125319 (2011) T. P. Smith, III, W. I. Wang, F. F. Fang, and L. L. Chang, Phys. Rev. B [**35**]{}, 9349 (1987). T. S. Lay, J. J. Heremans, Y. W. Suen, M. B. Santos, K. Hirakawa, M. Shayegan, and A. Zrenner, Appl. Phys. Lett. [**62**]{}, 3120 (1993). O. Gunawan, Y. P. Shkolnikov, E. P. De Poortere, E. Tutuc, and M. Shayegan, Phys. Rev. Lett. [**93**]{}, 246603 (2004). G. Dresselhaus, A. F. Kip, and C. Kittel, Phys. Rev. B [**98**]{}, 368 (1955). Frank Stern, Phys. Rev. Lett. [**18**]{}, 546 (1967). S. J. Allen, Jr., D. C. Tsui, and R. A. Logan, Phys. Rev. Lett. [**38**]{}, 980 (1977). E. Batkc and C. W. Tu, Phys. Rev. B [**34**]{}, 3027 (1986). V. E. Kozlov, S. I. Gubarev, I. .V. Kukushkin, JETP Lett. [**94**]{}, 397 (2011). V. E. Kozlov, S. I. Gubarev, A. A. Dremin, I. .V. Kukushkin, JETP Lett. [**96**]{}, 525 (2012). I. V. Kukushkin, A. V. Rossokhaty, S. Schmult, and K von Klitzing, Semicond. Sci. Technol. [**26**]{}, 014023 (2011). A. V. Chaplik, Surface Science Reports [**5**]{}, 289 (1985). Ugo Fano, Phys. Rev. B [**124**]{}, 1866 (1961). V. A. Volkov and S. A. Mikhailov, Zh. Eksp. Teor. Fiz. [**94**]{}, 217-241 (1988) \[JETP Lett. [**94**]{}, 217-241 (1988)\]. S. J. Allen and H. L. Stormer and J. C. M. Hwang, Phys. Rev. B [**28**]{}, 8 (1983). C. Dahl, F. Brinkop, A. Wixforth, J. P. Kotthaus, J. H. English, and M. Sundaram, Solid State Commun. [**80**]{}, 673 (1991). V. Shikin, S. Nazin, D. Heitmann, and T. Demel, Phys. Rev. B [**43**]{}, 11903 (1991). V. A. Geyler, V. A. Margulis, and A. V. Shorokhov, Phys. Rev. B [**63**]{}, 245316 (2001). I. V. Kukushkin, J. H. Smet, S. A. Mikhailov, D. V. Kulakovskii, K. von Klitzing, and W. Wegscheider, Phys. Rev. Lett. [**90**]{}, 156801 (2003). R. Z. Vitlina and A. V. Chaplik, Soviet Phys. JETP [**54**]{}, 536 (1981). V. M. Muravev, P. A. Gusikhin, I. V. Andreev, and I. V. Kukushkin, Phys. Rev. Lett. [**114**]{}, 106805 (2015).
{ "pile_set_name": "ArXiv" }
--- abstract: 'I. Raptis and R. Zapatrin in the quant-ph/0010104 show possibility to express general state of $l$-qubits quantum register as sum at most $2^l-l$ product states. In the comment is suggested more simple construction with possibility of generalization for decomposition of tensor product of Hilbert spaces with arbitrary dimension $n$ (here simplicial complexes used in the article mentioned above would not be applied directly). In this case it is decomposition with $n^l-(n^2-n)l/2$ product states.' address: 'FRC/IRH, 197101, Mira Street 8, St.-Petersburg, Russia' author: - 'A. Yu. Vlasov' date: 10 November 2000 title: 'Comment on ‘Decomposition of pure states of a quantum register’' --- [2]{} Decomposition of qubits and qu$n$its ==================================== Result of [@RZ] is proved with using simplicial complexes, etc., but it can be shown also with less special constructions. Such simplification may be useful because it makes possible to apply similar construction not only for qubit spaces ${{\Bbb C}}^2 \otimes \cdots \otimes {{\Bbb C}}^2$, but also for ‘qu$n$it’ spaces ${{\Bbb C}}^n \otimes \cdots \otimes {{\Bbb C}}^n$. Here is represented only short introduction to construction with arbitrary $n$, because [*the result is already known and described in* ]{}[@Sudbery]. The result of [@RZ] is related with possibility of exclusion by local unitary transformations of $l$ terms with one ‘unit’ like ${{|100\ldots0\rangle}}$, ${{|010\ldots0\rangle}}$, etc., from general $l$-qubit state with $2^l$ terms. To show it, let us consider some state ${{|h\rangle}} = \alpha{{|00\ldots0\rangle}} + \alpha_1 {{|10\ldots0\rangle}} + \cdots$, where $\alpha$ may be zero. It is possible with using a local unitary transformation in first qubit $U_1\colon (\alpha,\alpha_1) \to (\alpha',0)$ to produce state ${{|h'\rangle}} = \alpha'{{|00\ldots0\rangle}} + 0 {{|10\ldots0\rangle}} + \cdots$, where $|\alpha'|^2 = |\alpha|^2 + |\alpha_1|^2$. Next, because ${{|h'\rangle}} = \alpha'{{|00\ldots0\rangle}} + \alpha_2 {{|01\ldots0\rangle}} + \cdots$, it is possible to eliminate $\alpha_2$ by a local unitary transformation of second qubit and with $l$ similar steps we can produce some state ${{|h^{(1)}\rangle}}$. If ${{|h\rangle}}$ is product state, then ${{|h^{(1)}\rangle}}$ has only one nonzero term, but in general case it is possible to produce sum with no more than $2^l-l$ terms only by iterating the process: ${{|\tilde{h}\rangle}}=\lim_{N \to \infty} {{|h^{(N)}\rangle}}$. To check convergence let us consider a slightly different algorithm, when $N$’th step is elimination of term with one unit in position $i_N$ where coefficient $\alpha_{i_N}^{(N)} = \alpha_{\max}^{(N)}$ is maximal between $l$ similar terms. Absolute value of coefficient $\alpha^{(N)}$ of term ${{|00\ldots0\rangle}}$ is always limited $|\alpha^{(N)}| \le |h|\equiv 1$ and meets an equation: $|\alpha^{(N)}|^2 = |\alpha|^2+\sum_{K=1}^N |\alpha_{\max}^{(K)}|^2$. So $\sum_{N=1}^\infty |\alpha_{\max}^{(N)}|^2 = |\tilde\alpha|^2 - |\alpha|^2$ and then $\lim\limits_{N \to \infty} \alpha_{\max}^{(N)}~=~0$. To generalize the method to arbitrary $n > 2$ (‘qu$n$it’ [@qunit]) let us consider first $n = 3$ (‘qutrit’ [@qutrit]). If we start with some state ${{|h\rangle}} = \alpha{{|00\ldots0\rangle}} + \alpha_1 {{|10\ldots0\rangle}} + \beta_1 {{|20\ldots0\rangle}} + \cdots$, it is possible with using a local unitary $SU(3)$ transformation in first qutrit $U_1\colon (\alpha,\alpha_1,\beta_1) \to (\alpha',0,0)$ to produce state ${{|h'\rangle}} = \alpha'{{|00\ldots0\rangle}} + 0 {{|10\ldots0\rangle}} + 0 {{|20\ldots0\rangle}} + \cdots$ with $|\alpha'|^2 = |\alpha|^2 + |\alpha_1|^2 + |\beta_1|^2$ and algorithms similar with discussed above may eliminate $2l$ components with one ‘1’ or ‘2’ and $l-1$ ‘0’. But in the case it is not all components appropriate for exclusion. Let us write state ${{|g\rangle}}$ without the $2l$ term as: ${{|g\rangle}} = \gamma{{|11\ldots1\rangle}} + \gamma_1 {{|21\ldots1\rangle}} + \cdots$. Because it is possible to transform qutrit space without change of component ${{|0\rangle}}$, i.e., $U_1\colon (\alpha,\gamma,\gamma_1) \to (\alpha,\gamma',0)$ we can exclude all $l$ components with one ‘2’ and $l-1$ ‘1’ and so it is possible to have no more than $3^l-3l$ components. The similar process is possible to apply to arbitrary qu$n$it ${{\Bbb C}}^n$. First, we can exclude $(n-1)l$ terms with one ‘1’ and $l-1$ ‘0’. Second, $(n-2)l$ terms with one ‘2’ and $l-1$ ‘1’. After $n-1$ iterations we can neglect $((n-1)+(n-2)+\cdots+1)l = n (n-1) l /2$ terms and so we have no more than $$n^l - \case1/2 n (n-1) l$$ components. For $l=2$ we have $n^2 - n (n-1) 2 /2 = n$ terms of usual Schmidt decomposition. Acknowledgements {#acknowledgements .unnumbered} ================ Author is grateful to R. Zapatrin and C. Zalka for explanations and comments. Many thanks to A. Sudbery for reference to the work [@Sudbery]. I. Raptis and R. Zapatrin, E-print: quant-ph/0010104 (2000). H. A. Carteret, A. Higuchi and A. Sudbery, E-print: quant-ph/0006125 (2000). D. Kaszlikowski, P. Gnacinski, [*et al*]{}, E-print: quant-ph/0005028 (2000). C. M. Caves and G. J. Milburn, E-print: quant-ph/9910001 (1999).
{ "pile_set_name": "ArXiv" }
--- abstract: 'We calculate the Kadowaki-Woods ratio (KWR) in Fermi liquids with arbitrary band structures. We find that, contrary to the single band case, the ratio is not generally independent of the effects of electronic correlations (universal). This is very surprising given the experimental findings of a near universal KWR in many multiband metals. We identify a limit where the universality of the ratio, which has been observed experimentally in many strongly correlated electron systems, is recovered. We discuss the KWR in Dirac semimetals in two and three dimensions. In the two-dimensional case we also generalize the KWR to account for the logarithmic factor in the self-energy. In both cases we find that the KWR is independent of correlations, but strongly dependent on the doping of the system: for massless fermions the KWR is proportional to the inverse square of the carrier density, whereas the KWR for systems with massive quasiparticles is proportional to the inverse of the carrier density.' author: - 'D. C. Cavanagh' - 'A. C. Jacko' - 'B. J. Powell' bibliography: - 'KWRRef.bib' title: 'When is the Kadowaki-Woods ratio universal?' --- Introduction ============ Fermi liquid theory describes the low temperature behavior of the vast majority of metals extremely well [@AM; @GV; @Landau1; @Landau2; @Schofield]. One of the beauties of Fermi liquid theory is that it reduces the description of the interacting electron fluid to a small number of (Landau) parameters. Therefore, ratios in which these parameters cancel, such as the Wilson-Sommerfeld ratio and Wiedemann-Franz law [@AM; @Hewson] provided important tests of Fermi liquid theory. In a Fermi liquid the electronic contributions to the resistivity \[$\rho_{el}\left(T\right)=AT^2$\] and heat capacity \[$C_{el}\left(T\right)=\gamma T$\] are both governed by the effective mass, $m^*$ – roughly speaking $A\propto m^{*2}$ and $\gamma\propto m^*$. So the Kadowaki-Woods ratio, $A/\gamma^2$, should be constant in a Fermi liquid [@AK; @PAH; @Coleman; @Auerbach; @Miyake]. More precisely one might expect correlations to leave the Kadowaki-Woods ratio (KWR) unrenormalized because a Kramers-Kronig transformation relates the real and imaginary parts of the self-energy [@Luttinger; @Miyake; @Jacko], which determine the electronic contributions to the heat capacity and resistivity respectively. This means that the KWR is somewhat similar to a fluctuation-dissipation theorem. First Rice [@Rice] and later Kadowaki and Woods [@KW] found that $A/\gamma^2$ is approximately constant within classes of materials (transition metals and heavy fermion compounds, respectively). However, the ratio differs by two orders of magnitude between these two classes. It was subsequently discovered that the Kadowaki-Woods ratio (KWR) in transition metals and organic charge transfer salts can be even larger than in the heavy fermions (see Refs. and references therein.) It was long believed [@Miyake; @Li] that the size of the KWR gave an indication of the strength of the electron-electron scattering [^1]. However, this has been shown to be incorrect [@Jacko; @Hussey]. Rather, the large variations in the KWR between different classes of materials can be explained almost entirely by taking into account non-interacting properties of the materials (e.g. electron density and dimensionality) [@Jacko; @Hussey]. Furthermore, it has been shown [@Jacko] that the modified KWR $A/\gamma^2f$, where $f$ is a material specific function of the non-interacting band structure (defined below), takes the same predicted value, $81/4\pi\hbar k_B^2e^2$, in a large range of transition metals, charge transfer salts, heavy fermion compounds, and elemental metals, a result which has since been verified in many other materials [@Jref1; @Jref2; @Jref3; @Jref4; @Jref5]. Deviations from this universal predicted value of the KWR could provide an indication of non-Fermi liquid behavior. Most previous calculations of the (modified) KWR [@Jacko; @PAH; @Auerbach] have focused on simple, single-band models with toy dispersion relations, e.g. spherical Fermi surfaces. Understanding the effects of more complicated and realistic band structures is important if the modified KWR is to be used to identify deviations from Fermi liquid theory. Previous studies of the KWR in systems with orbital degeneracy [@Kontani; @Tsuji] or multiple bands [@Hussey] have found that either of these can cause significant variation in the KWR. In this paper, we derive a modifed KWR for Fermi liquids with arbitrary dispersion relations, including band structures with multiple bands. We find that the universality of the KWR evident in the single band expression is not a general feature of the multiple band case. In particular in the most general case the strength of electronic correlations [*does*]{} affect the value of the KWR. This is extremely puzzling as the KWR is found to be close to its universal value in many multiband systems. However, if the renormalization is the same on all bands (in a sense made precise below) correlations do cancel from the KWR. In the case of uniform renormalization across all bands, we demonstrate, that the KWR $R_{KW}\propto1/nN_b^2$ for massive quasiparticles in a system of $N_b$ bands. This is particularly relevant to semimetals where the low career density opens the possibility of large variations in the career density, $n$. This is further enhanced in Dirac semimetals – we find that the massless fermion dispersion relations lead to $R_{KW}\propto1/n^2N_b^2$ in three dimensions. In two dimensions a similar result holds once the KWR is generalized to account for the logarithmic factor in the imaginary part of the self-energy. Furthermore, we show that the logarithmic factor in the imaginary part of the self-energy leads to an increase in the KWR, which may provide an experimental signature of this factor in the self energy. The remainder of this paper is laid out as follows. In the following section, we calculate the resistivity of an arbitrary multi-band Fermi liquid by calculating the conductivity from the Kubo formula of linear response. In Section \[heat\], we calculate the effect of multiple bands on the heat capacity. In Section \[KWRSec\], we combine these results to determine the form of the Kadowaki-Woods ratio in arbitrary band structuress. In Section \[graphenesec\], we then apply this form for the KWR to simple models of Dirac semimetals in two and three dimensions. Conductivity from the Kubo Formula ================================== In general, the contribution to the intraband self-energy from interband terms scales quadratically with the intraband self-energy, such that, as long as the self-energy is small, the intraband contribution will dominate [^2]. When intraband scattering is the dominant contribution to the scattering rate, the diagonal component of the conductivity tensor, $\sigma_{xx}$, for a material with $N_b$ bands crossing the Fermi surface in the low temperature limit is [^3] $$\begin{aligned} \sigma_{xx}&=&e^2\hbar\int\limits_{-\infty}^{\infty}\frac{d^3\textbf{k}}{\left(2\pi\right)^3} \int\limits_{-\infty}^{\infty} \frac{d\omega}{4\pi} \sum\limits_{\tilde{a}}^{2N_b} \frac{2\pi Z_{\tilde{a}}\delta\left(\omega-Z_{\tilde{a}} \omega_{\textbf{k}\tilde{a}}\right)} {\Sigma_{\tilde{a}\tilde{a}}''\left(\omega\right)}v_{\textbf{k}x\tilde{a}}^2 \left(\frac{dn_f(\omega)}{d\omega}\right), \label{eqn:dc}\end{aligned}$$ where $\Sigma_{\tilde{a}\tilde{a}}\left(\omega\right)=\Sigma_{\tilde{a}\tilde{a}}'\left(\omega\right)+i\Sigma_{\tilde{a}\tilde{a}}''\left(\omega\right) $ is the self-energy, $Z_{\tilde{a}}=|1-[\partial\Sigma'_{\tilde{a}\tilde{a}}\left(\omega\right)/\partial\omega]_{\omega=0}|^{-1}$ is the quasiparticle weight, $v_{\textbf{k}x\tilde{a}} $ the $x$-component of the group velocity of an electron in spin-band $\tilde{a}=\left(a,\sigma\right)$, where $a$ denotes the band and $\sigma$ the spin (giving $2N_b$ spin-bands), and we have neglected vertex corrections (cf. Ref. ). The sharply peaked derivative of the Fermi-Dirac distribution at low temperatures implies that $$\begin{aligned} \sigma_{xx}&\approx &\sum\limits_{\tilde{a}=1}^{2N_b}\frac{e^2\hbar\langle v_{\textbf{k}x\tilde{a}}^2\rangle}{2} \int\limits_{-\infty}^{\infty}\frac{d^3\textbf{k}}{\left(2\pi\right)^3} Z_{\tilde{a}}\delta\left(Z_{\tilde{a}}\mu-Z_{\tilde{a}} \omega_{\textbf{k}\tilde{a}}\right) \int\limits_{-\infty}^{\infty} d\omega \left(\frac{-1}{\Sigma_{\tilde{a}\tilde{a}}''\left(\omega\right)} \right)\left(-\frac{dn_f(\omega)}{d\omega}\right),\label{2bc}\end{aligned}$$ where $\langle ... \rangle$ indicates an average over the Fermi surface. The conductivity is clearly then the sum, in series, of the conductivities of the individual bands: $$\begin{aligned} \sigma_{xx}=\sum\limits_{\tilde{a}=1}^{2N_b} \sigma_{xx\tilde{a}}&=&\sum\limits_{\tilde{a}=1}^{2N_b}\frac{e^2\hbar\langle v_{\textbf{k}x\tilde{a}}^2\rangle}{2} D_{0;\tilde{a}} \int\limits_{-\infty}^{\infty} d\omega \left(\frac{1}{\Sigma_{\tilde{a}\tilde{a}}''\left(\omega\right)}\right) \left(\frac{dn_f(\omega)}{d\omega}\right),\label{2bcf}\end{aligned}$$ where the bare density of states of band $\tilde{a}$ at the Fermi level is $D_{0;\tilde{a}}= (2\pi)^{-3}\int_{-\infty}^{\infty}d^3\textbf{k} \left[\delta\left(\omega_{\textbf{k}\tilde{a}}-\mu\right)\right]$ and $\sigma_{xx\tilde{a}}$ is the conductivity of due to spin-band $\tilde{a}$. It immediately follows that, if interband scattering is neglected, the resistivities of the bands must add in parallel, consistent with Matthiessen’s rule [@AM]. ### The Quadratic Contribution to the Resistivity To determine the explicit form of the conductivity in a Fermi liquid, we generalise a local (momentum-independent) phenomenological self-energy model proposed by Miyake, Matsuura and Varma, [@Miyake] to include multiple Fermi surfaces. We assume that in each band the self-energy takes the phenomenological form $$\begin{aligned} \Sigma''_{\tilde{a}\tilde{a}}\left(\omega\right)&=& -\frac{\hbar}{2\tau_{0;\tilde{a}}} -s_{\tilde{a}}\frac{\omega^2+\left(\pi k_BT\right)^2}{\left(\omega^*_{\tilde{a}}\right)^2} \hspace*{2cm} \text{ for } \left|\omega^2+\left(\pi k_BT\right)^2\right|<\left(\omega^*_{\tilde{a}}\right)^2 \text{ and } \\ &=& -\left(\frac{\hbar}{2\tau_{0;\tilde{a}}} +s_{\tilde{a}}\right)F\left[\left(\frac{\omega^2+\left(\pi k_BT\right)^2}{\left(\omega^*_{\tilde{a}}\right)^2}\right)^{\frac{1}{2}}\right] \text{ for } \left|\omega^2+\left(\pi k_BT\right)^2\right|>\left(\omega^*_{\tilde{a}}\right)^2, \end{aligned}$$ \[phenom\] where the energy $\omega$ is measured from the Fermi level, $\tau_{0;\tilde{a}}^{-1}$ is the impurity scattering rate, $s_{\tilde{a}}=n_{\tilde{a}}/3\pi D_{0;\tilde{a}}$ is the electron-electron scattering rate in the unitary limit [@Miyake; @Jacko] (with $n_{\tilde{a}}$ the density of charge carriers in band $\tilde{a}$), $F(x)$ is a monotonically decreasing function with $F(1)=1$ and $F(\infty)=0$ and $\omega^*_{\tilde{a}}$ is an energy scale characterising the strength of the many-body correlations (we will show below that $Z_{\tilde{a}} \approx \omega^*_{\tilde{a}} / 4s_{\tilde{a}}$). This is the natural form for a local Fermi liquid provided interband interactions are weak compared to the intraband interactions. If the two are comparable (as in some multiorbital models) the self energy is proportional to $N_b-1$ [@Kontani]. In the limit of vanishing impurity scattering, $\tau^{-1}_{0;\tilde{a}}\rightarrow 0$, the conductivity is then given by $$\begin{aligned} \sigma_{xx}&=&-\sum\limits_{\tilde{a}=1}^{2N_b}\frac{3\pi e^2\hbar }{2}\frac{D_{0;\tilde{a}}^2\langle v_{0x\tilde{a}}^2\rangle\left(\omega^*_{\tilde{a}}\right)^2}{n_{\tilde{a}}} \int\limits_{-\infty}^{\infty} d\omega \frac{1}{\omega^2+\left(\pi k_BT\right)^2}\frac{dn_f(\omega)}{d\omega}.\label{2bcondlowT}\end{aligned}$$ After computing the energy integral, the $N_b$-band $A$ coefficient is $$\begin{aligned} A&=&\frac{8 k_B^2}{\pi e^2\hbar}\left(\sum\limits_{\tilde{a}=1}^{2N_b}\langle v_{0x\tilde{a}}^2\rangle\frac{D_{0;\tilde{a}}^2\left(\omega^*_{\tilde{a}}\right)^2}{n_{\tilde{a}}} \right)^{-1}=\left(\sum\limits_{\tilde{a}=1}^{2N_b} A_{\tilde{a}}^{-1}\right)^{-1}, \label{Ainparallel}\end{aligned}$$ thus, we see that the coefficients of the individual bands, $A_{\tilde{a}}$, add in parallel. The Heat Capacity via Kramers-Kronig transform {#heat} ============================================== It follows from the extensivity of the heat capacity that, in a multi-band system, the total heat capacity is given by the sum in series of the heat capacity due to each individual band [@AM],$$\begin{aligned} \gamma&=&\frac{C_{el; N_b}\left(T\right)}{T} =\frac{\pi^2 k_B^2}{3} \sum\limits_{\tilde{a}=1}^{2N_b} \frac{D_{0;\tilde{a}}}{Z_{\tilde{a}}}=\frac{\pi^2 k_B^2}{3}\sum\limits_{\tilde{a}=1}^{2N_b} D_{0;\tilde{a}}\left(1-\frac{\partial \Sigma'_{\tilde{a}\tilde{a}}\left(\omega\right)}{\partial \omega}\right). \label{2bC}\end{aligned}$$ To determine how the heat capacity is influenced by interactions, and calculate the relevant $\gamma$ coefficients, we need to first find the real part of the self-energy in each band. The Fermi liquid self-energy, being a causal response function in the time domain, satisfies the conditions for the Kramers-Kronig relations in the frequency domain [@Luttinger] – in particular $\lim\limits_{\omega\rightarrow\infty} \Sigma''\left(\omega\right)\rightarrow 0$. Knowledge of the form of the imaginary part of the self-energy is therefore sufficient to determine the real part, which appears in the definition of the quasiparticle weight and therefore in the expression for the heat capacity. The real part of the self-energy within each band is then $$\begin{aligned} \Sigma'_{\tilde{a}\tilde{a}}\left(\omega\right)&=&\frac{1}{\pi} P \int\limits_{-\infty}^{\infty} d\omega' \frac{\Sigma''_{\tilde{a}\tilde{a}}(\omega')}{\omega'-\omega}\nonumber\\ &=& \frac{-s_{\tilde{a}}}{\pi} \left[\vphantom{P \int\limits_{-\omega^*_{\tilde{a}}}^{\omega^*_{\tilde{a}}} \frac{d\omega'}{\omega'-\omega} \left(\frac{\omega'^2}{\left(\omega^*_{\tilde{a}}\right)^2}\right)}P\int\limits_{-\infty}^{-\omega^*_{\tilde{a}}} \frac{d\omega'}{\omega'-\omega} F\left[\frac{\left|\omega'\right|}{\omega^*_{\tilde{a}}}\right]+P \int\limits_{-\omega^*_{\tilde{a}}}^{\omega^*_{\tilde{a}}} \frac{d\omega'}{\omega'-\omega} \left(\frac{\omega'^2}{\left(\omega^*_{\tilde{a}}\right)^2}\right) +P\int\limits_{\omega^*_{\tilde{a}}}^{\infty} \frac{d\omega'}{\omega'-\omega} F\left[\frac{\left|\omega'\right|}{\omega^*_{\tilde{a}}}\right]\right] , \label{KKint}\end{aligned}$$ where we have again taken the limit of vanishing impurity scattering, and have restricted the pole of the integral, $\omega'=\omega$, to occur below the cut-off energy scale, $|\omega'|\le|\omega^*_{\tilde{a}}|$ (i.e., only considered low-energy excitations). The first term in Eq. (\[KKint\]) contributes a logarithmic term to the result, which we approximate by the lowest order terms in a Taylor series expansion, while the second and third terms contribute linearly to the self-energy. Neglecting terms of order ${\cal O}\left(\omega/\omega^*_{\tilde{a}}\right)^3$ and higher, we find that the real part of the self-energy for a low-energy quasiparticle in band $\tilde{a}$ is $$\Sigma'_{\tilde{a}\tilde{a}}\left(\omega\right)=-\frac{4n_{\tilde{a}}\omega} {3\pi^2 D_{0;\tilde{a}}\omega^*_{\tilde{a}}} \xi, \label{realse}$$ where $$\begin{aligned} \xi&=&\frac{1}{2}\left(1+\int\limits_{1}^{\infty}dy \frac{F(y)}{y^{2}}\right) \label{xidef}\end{aligned}$$ it follows straightforwardly from the definition of $F(y)$ that $1/2\leq\xi\leq 1$, henceforth we take $\xi=1$ for simplicity. Inserting this expression for the real parts of the self-energies into the heat capacity expression Eq. (\[2bC\]), and taking the strong scattering ($4s_{p\tilde{a}}/\pi\omega^*_{\tilde{a}} \gg1$) limit, which corresponds physically to $m^*\gg m_0$, we obtain $$\begin{aligned} \gamma&=&\frac{4k_B^2}{9} \sum\limits_{\tilde{a}=1}^{2N_b} \frac{n_{\tilde{a}}}{\omega^*_{\tilde{a}}}. \label{2by}\end{aligned}$$ Note that Eqs. (\[2bC\]) and (\[2by\]) imply that $\omega^*_{\tilde{a}}\approx4s_{\tilde{a}}Z_{\tilde{a}}=4Z_{\tilde{a}}n_{\tilde{a}}/3\pi D_{0;\tilde{a}}$, which gives a straightforward interpretation of this energy scale. The Kadowaki-Woods Ratio {#KWRSec} ======================== Given the above calculations of $A$ and $\gamma$ for the multiple band system, Eqs. (\[Ainparallel\]) and (\[2by\]), we find that the KWR is given by $$\begin{aligned} R_{KW;N_b}=\frac{A}{\gamma^2} &=&\frac{81}{2\pi e^2\hbar k_B^2 }\left(\sum\limits_{\tilde{b}=1}^{2N_b}\langle v_{0x\tilde{b}}^2\rangle \frac{D_{0;\tilde{b}}^2\left(\omega^*_{\tilde{b}}\right)^2}{n_{\tilde{b}}} \right)^{-1}\left[\sum\limits_{\tilde{a}=1}^{2N_b}\frac{n_{\tilde{a}}}{\omega^*_{\tilde{a}}}\right]^{-2}.\label{NbKWR}\end{aligned}$$ Alternatively, one may write $$\begin{aligned} \frac{A}{\gamma^2} f_{x;N_b}\left(\left\lbrace\omega^*_{\tilde{a}}\right\rbrace, \left\lbrace n_{\tilde{a}}\right\rbrace\right)&=&\frac{81}{4\pi e^2 \hbar k_B^2} ,\label{KWRNband}\end{aligned}$$ where we have defined the material specific function for an $N_b$-band Fermi liquid with the resistivity measured in the $x$ direction, $$\begin{aligned} f_{x;N_b}\left(\left\lbrace\omega^*_{\tilde{a}}\right\rbrace, \left\lbrace n_{\tilde{a}}\right\rbrace\right)&=& \frac{1}{2}\left[\sum\limits_{\tilde{a}=1}^{2N_b} \frac{n_{\tilde{a}}}{\omega^*_{\tilde{a}}}\right]^{2} \left[\sum\limits_{\tilde{b}=1}^{2N_b}\langle v_{0x\tilde{b}}^2\rangle \frac{D_{0;\tilde{b}}^2\left(\omega^*_{\tilde{b}}\right)^2}{n_{\tilde{b}}} \right].\label{Nbfdx}\end{aligned}$$ In the one-band limit, this expression simplifies to that calculated in Ref. (with $n_\uparrow=n_\downarrow=n/2$ and $D_{0,\uparrow}=D_{0,\downarrow}=D_0/2$) $$\begin{aligned} f_{x;1}\left(n\right)&=& n \langle v_{0x}^2\rangle D_0^2 \label{1bfdxcomp}.\end{aligned}$$ The inclusion of multiple Fermi surface sheets significantly complicates the form of $f_{x;N_b}\left(\left\lbrace\omega^*_{\tilde{a}}\right\rbrace, \left\lbrace n_{\tilde{a}}\right\rbrace\right)$. Most importantly $\omega^*_{\tilde{a}}$, which describes the electronic correlations, does not cancel out of the multiband expression as it does for the single band KWR [@Jacko]. Therefore, our calculation predicts that the Kadowaki-Woods ratio is not, in general, independent of electronic correlations. This is rather surprising as observed values of the KWR (including the values for many multiband systems) are in almost universal agreement with the prediction from the single band calculation, that electronic correlations do not influence the KWR [@Jacko]. It is therefore important to ask how renormalization effects might cancel in the multiband case and hence universality might be recovered. The simplest limiting case for which the effects of many-body correlations cancel out of the KWR is when $\omega_{\tilde{a}}^*$ is independent of the band index, $\tilde{a}$. This is a straightforward extension of the earlier assumption of the locality of the self-energy, by assuming that it is independent of band index as well as momentum. This assumption yields $$\begin{aligned} f_{x;N_b}\left(\left\lbrace n_{\tilde{a}}\right\rbrace\right)&=& \frac{1}{2}\left[\sum\limits_{\tilde{a}=1}^{2N_b} n_{\tilde{a}}\right]^{2} \left[\sum\limits_{\tilde{b}=1}^{2N_b}\langle v_{0x\tilde{b}}^2\rangle \frac{D_{0;\tilde{b}}^2}{n_{\tilde{b}}} \right]. \label{Nbfsimple}\end{aligned}$$ Though this calculation has been performed with exactly uniform correlation strengths for simplicity, the result will hold approximately while the correlation strengths are close to uniform. Other limits do produce a universal KWR, for example, for free fermions in three dimensions $D_{0;\tilde{b}}\propto n_{\tilde{b}}$, then if the carrier density in one band is much larger than all others the correlations cancel from the KWR. But it seems unlikely that this is relevant the behaviour of a broad range range of materials. Note, in particular, that if we have a single heavy band it will dominate the heat capacity, but be shorted out of the resistivity. Therefore, the limit of a single heavy band is far from universal. Therefore, the above calculation seems to suggest the correlation strength (as measured by $\omega_{\tilde{a}}^*=4Z_{\tilde{a}}n_{\tilde{a}}/3\pi D_{0;\tilde{a}}$) does not vary strongly between different bands in strongly correlated systems. If, further, all of the bands are identical, i.e., if the carrier density $n$, the Fermi velocity $\langle v_{0x}^2\rangle$, and density of states $D_0$ are equal for all bands, we have $$\begin{aligned} f_{x;N_b}\left(n\right)&=& N_b^2\langle v_{0x}^2\rangle n D_0^2=N_b^2f_{dx;1}(n),\label{funif}\end{aligned}$$ and $$\begin{aligned} R_{KW;N_b}&=&\frac{81}{4N_b^2\pi e^2\hbar k_B^2\langle v_{0x}^2\rangle n D_0^2}=\frac{R_{KW;1}}{N_b^2}. \label{2bKWRnDeq} \end{aligned}$$ Thus we see that the expression for the KWR in the single-band case is modified by a simple factor of $1/N_b^2$. At first glance this expression appears rather similar to the finding of Kontani [[*et al*]{}. ]{}[@Kontani; @Tsuji] that in the multiorbital periodic Anderson model with $N_o$ impurity states the KWR is reduced by a factor of $1/N_o(N_o-1)$. However, closer examination reveals that the results are actually very different. In particular Kontani’s factor arises because of interorbital terms in the self energy whereas our $1/N_b^2$ factor arises purely from the electronic structure. We do not obtain Kontani’s factor because of our assumption that interband interactions are irrelevant at low energies. In contrast the model Hamiltonian studied by Kontani [@Kontani] explicitly sets the intra- and inter-orbital interactions to the same strength. Which approach is appropriate will depend on the material. This therefore adds another layer of non-universality to the KWR. Dirac Semimetals {#graphenesec} ================ The general expression, Eq. (\[KWRNband\]), can be applied to systems of arbitrary band structure to calculate the generalised Kadowaki-Woods ratio, taking into account the effects of multiple bands. Even for materials where the quasiparticle weight is similar for all bands such efforts will, in general, involve first principles band structure calculations. In this section, we calculate the KWR for a simple, linear dispersion ($\varepsilon_{\textbf{k}\tilde{a}}=\hbar v_F \left|\textbf{k}\right|$) appropriate for Dirac semimetals. These models present analytically tractable and instructive examples of complicated band structures for which the presence of multiple bands is important (here, the number of bands may be taken as equivalent to the number of Dirac cones, as each cone will form a sheet in the Fermi surface). Furthermore, these materials are an example of materials where the different sheets of the Fermi surface where we expect the correlations to affect all bands equally and the low carrier density suggests that correlations will play an important role. We also determine an expression for the KWR in two dimensions, accounting for the logarithmic factor arising in the self-energy [@FLdopedgraphene; @GrRMP1; @GrRMP2; @GrRMP3; @gbandrenorm; @gproplong; @Chaplik; @Wilkins; @SE2D; @UC]. 3D Dirac Semimetals ------------------- We first apply the KWR expression to a simple model of a two-band three-dimensional Dirac semimetal, e.g. Cd$_3$As$_2$, which possesses properties similar to doped graphene [@Cd3As2-1; @Cd3As2-2]. The spatial symmetry between the two bands and the spin symmetry simplifies the calculation greatly; applying the expression for $N_b$ identical bands, Eq. (\[funif\]), we find $$\frac{A}{\gamma^2} = \frac{1}{\tilde{f}_{3x,2}}\frac{81}{4\pi e^2 \hbar k_B^2} = \frac{81\hbar}{8e^2k_B^2 n^2}, \label{grapheneKWR} $$ where $\tilde{f}_{3x,2} = 2 n^2/\pi \hbar^2$ is the material specific function for a 3D Dirac semimetal and is larger by a factor of 4 than the corresponding $f$ in a single-band calculation with the same dispersion. We note that the Kadowaki-Woods ratio for these materials depends straightforwardly on the electronic density, which is tunable via chemical doping [@Cd3As2-1; @Cd3As2-2], providing a potential experimental test of this expression. Doped Graphene and the Self-Energy in Two Dimensions ---------------------------------------------------- In calculating Eq. (\[grapheneKWR\]), we utilised the form of self-energy given in Eq. (\[phenom\]), which implicitly assumes a three-dimensional material. In two dimensions, the self-energy differs from the three dimensional case, with the inclusion of an additional logarithmic factor [@2DEG1; @2DEG2; @2DEG3; @dasSarmaSE; @dasSarmaSE2; @FLTin2D; @GQ; @GV]. In order to calculate the KWR for doped graphene, we introduce the following model for the imaginary part of the self-energy at low energies and temperatures for 2D systems $$\begin{aligned} \Sigma''_{\tilde{a}\tilde{a}}\left(\omega\right)&=& s_{\tilde{a}}\frac{\omega^2+\left(\pi k_B T\right)^2}{\left(\omega^*_{\tilde{a}}\right)^2}\log\left(\frac{\sqrt{\omega^2+\left(\pi k_B T\right)^2}}{B_{\tilde{a}}\omega^*_{\tilde{a}}}\right),\label{gSim}\end{aligned}$$ where $B_{\tilde{a}}$ is a constant of order unity (for example the calculation in Ref. gives $B_{\tilde{a}}\approx1/\pi$ for graphene; but other calculations give slightly different results). We stress that, in the zero and high temperature ($k_BT\gg\omega$) limits, Eq. (\[gSim\]) reproduces the known results for those limits [@2DEG1; @dasSarmaSE; @dasSarmaSE2]. We further assume that, above the relevant energy scale, $\omega^*$, the self-energy decreases monotonically, as in the 3D case. From this expression for imaginary part of the self-energy, we calculate the real part using the Kramers-Kronig transformation, and find that $$\begin{aligned} \Sigma'_{\tilde{a}\tilde{a}}\left(\omega\right)&=&-\frac{4s}{\pi}\left[\xi\log\left(B_{\tilde{a}}\right)+1\right]\omega+ {\cal O}\left(\omega^3\right)\label{gSre}\end{aligned}$$ where the logarithmic factor arises due to the requirement that the self-energy be continuous at $\omega=\omega^*$. In defining the 2D KWR, we must account for the fact that the logarithmic contribution to the imaginary part of the self-energy (and scattering rate) results in a corresponding logarithmic factor in the resistivity, $\rho=-\tilde{A}T^2\log\left(\pi k_BT/B\omega^*\right)$. The coefficient of this logarithmically adjusted quadratic term will be used in our expression. Calculating the conductivity from the Kubo formula gives $$\begin{aligned} \tilde{A}_{\tilde{a}}&=& \frac{8 k_B^2}{\pi e^2\hbar n_{\tilde{a}}\langle v_{x\tilde{a}}^2\rangle\omega^*_{\tilde{a}} D_{0;\tilde{a}}^2}\label{2DA},\end{aligned}$$ where, in evaluating the energy integral (see Eq. (\[2bcondlowT\])), we have approximated the logarithmic contribution by its value at the Fermi surface. From the derivative of the real part of the self-energy, we find the value for the quasiparticle weight and therefore the linear coefficient of the heat capacity: $$\begin{aligned} \gamma_{\tilde{a}} &=& \frac{2\pi k_B^2n_{\tilde{a}}}{9\pi} \left( \log\left(B_{\tilde{a}}\right) +\frac{1}{2}\right).\end{aligned}$$ Defining the two-dimensional Kadowaki-Woods ratio as $\tilde{A}/\gamma^2$ we find $$\begin{aligned} R_{KW}=\frac{\tilde{A}}{\gamma^2}&=&\frac{81}{4\pi e^2\hbar k_B^2} \frac{1}{\tilde{f}_{2x;2}\left(n\right)}\end{aligned}$$ where $$\begin{aligned} \tilde{f}_{2x;2}\left(n\right)&=& \frac{1}{2}\left[\sum\limits_{\tilde{a}=1}^{2N_b} \frac{n_{\tilde{a}}\left( \log\left(B_{\tilde{a}}\right) +\frac{1}{2}\right)}{\omega^*_{\tilde{a}}}\right]^{2} \left[\sum\limits_{\tilde{b}=1}^{2N_b} \frac{\langle v_{0x\tilde{b}}^2\rangle D_{0;\tilde{b}}^2\left(\omega^*_{\tilde{b}}\right)^2}{n_{\tilde{b}}C_{\tilde{b}}} \right]\nonumber\end{aligned}$$ is the two-dimensional material specific function, which for graphene takes the form $$\begin{aligned} \tilde{f}_{2x;2}\left(n\right)=4\tilde{f}_{2x;1}\left(n\right)&=&\frac{2n^2}{\pi\hbar^2}\left[\log\left(B\right)+\frac{1}{2}\right]^2.\end{aligned}$$ Here the Kadowaki-Woods ratio again depends on the electron density in a straightforward manner, but differs from the previously derived expression (Eq. (\[grapheneKWR\])) only by a factor of $\left[ \log\left(B\right)+1/2\right]^2$. Assuming $B \simeq 1/\pi$, this term more than doubles the expected Kadowaki-Woods ratio. Measuring the KWR of graphene presents a number of technical challenges, chiefly the difficulty of performing calorimetric measurements on a material of single atom thickness. It has been a long standing problem to observe the logarithmic factor in the resistivity of a two-dimensional Fermi liquid because it would require data over many orders of magnitude in temperature. However, observation of this correction offers a potential test of the dimensionality of the Fermi liquid self energy without the need to take data over multiple decades. Conclusions =========== We have shown that, in general, the Kadowaki-Woods ratio of a multiband local Fermi liquid is changed by electronic correlations. This is in marked contrast to the single band case, where the KWR is independent of the strength of the electronic correlations. It is therefore puzzling that the experimental data suggest the within classes of materials the KWR is remarkably consistent, and that the modified KWR is remarkably consistent across many chemically diverse strongly correlated metals. The simplest explanation is that the correlations are indeed very similar across all bands in these materials. We have also shown that a non-parabolic dispersion does not significantly alter the form of the KWR, provided the fermions remain massive. In the case of uniform renormalization across bands, we have further demonstrated that $R_{KW}\propto1/nN_b^2$ for massive quasiparticles in a system of $N_b$ bands. This is particularly interesting in semimetals where the low career density opens the possibility of large variations in the career density, $n$. This is further enhanced in Dirac semimetals, where we have shown that the massless fermion dispersion relations lead to $R_{KW}\propto1/n^2N_b^2$ in three dimensions. In two dimensions a similar result holds once the KWR is generalized to account for the logarithmic factor in the imaginary part of the self-energy. Furthermore, we have shown that the logarithmic factor in the imaginary part of the self-energy leads to an increase in the KWR, which may provide an experimental signature of this factor in the self energy. Acknowledgments {#acknowledgments .unnumbered} =============== This work was supported by the Australian Research Council (ARC) under grant DP130100757. BJP is supported by the ARC under grant FT130100161. [^1]: Miyake [[*et al*]{}. ]{}[@Miyake] argued that variations in the KWR arise from differences in the quasiparticle weight. [^2]: For a derivation see the supplementary material. [^3]: The derivation of Eq. (\[eqn:dc\]) is a straightforward generalization of the textbook one band case, but we could not find it in the literature and therefore we give the derivation in the supplementary information.
{ "pile_set_name": "ArXiv" }
--- abstract: 'The next-to-leading order (NLO) QCD radiative corrections to $W^+W^-$ production at hadron colliders are well understood. We combine NLO perturbative QCD calculations with soft-gluon resummation of threshold logarithms to find a next-to-next-to leading logarithmic (NNLL) prediction for the total cross section and the invariant mass distribution at the LHC. We also obtain approximate next-to-next-to-leading order (NNLO) results for the total $W^+W^-$ cross section at the LHC which includes all contributions from the scale dependent leading singular terms. Our result for the approximate NNLO total cross section is the most precise theoretical prediction available. Uncertainties due to scale variation are shown to be small when the threshold logarithms are included. NNLL threshold resummation increases the $W^+W^-$ invariant mass distribution by $\sim 3-4\%$ in the peak region for both $\sqrt{S}=8$ and $14$ TeV. The NNLL threshold resummed and approximate NNLO cross sections increase the NLO cross section by $0.5-3\%$ for $\sqrt{S}=7,\,8,\,13,$ and $14$ TeV.' author: - 'S. Dawson$^{a}$, Ian M. Lewis$^{a}$, and Mao Zeng$^{b}$' bibliography: - 'paper.bib' title: 'Threshold Resummed and Approximate NNLO results for $W^+W^-$ Pair Production at the LHC' --- Introduction ============ Exploring the Higgs and electroweak sector of the Standard Model is one of the primary goals of the LHC. The pair production of gauge bosons is important both as a test of the $SU(2)\times U(1)$ gauge structure and as a background for Higgs boson searches. Precise predictions for both total and differential cross sections are needed in order to understand the shape of the background to the Higgs signal and to search for anomalous three gauge boson couplings. Understanding the theoretical prediction for $pp\rightarrow W^+W^-$ is particularly important for the measurement of the Higgs decay channel, $H\rightarrow W^+W^-\rightarrow2\ell2\nu$, where there is no resonant structure. The $W^+W^-$ background is estimated by a sideband analysis, where the cross section is normalized via a control region with a minimum dilepton invariant mass. Using Monte Carlo, the line shapes of the $W^+W^-$ distributions are then extrapolated into the Higgs signal region [@CMS:bxa; @ATLAS:2013wla]. A change in the $W^+W^-$ invariant mass distribution will alter the dilepton invariant mass distribution, and consequently change the extrapolation of the background estimates in the Higgs signal region. The production of $W^+W^-$ pairs with a subsequent leptonic decay has been studied at the Tevatron [@Aaltonen:2009aa; @Abazov:2009ys], while both ATLAS [@ATLAS:2012mec] and CMS [@Chatrchyan:2013oev; @Chatrchyan:2013yaa] have reported results at the LHC. The LHC results for the total $W^+W^-$ cross section, $$\begin{aligned} \hbox{ATLAS},~\sqrt{S}=7~{\rm TeV}~~&\sigma=&51.9\pm 2.0 (\hbox{stat})\pm 3.9 (\hbox{syst})\pm 2.0 (\hbox{lumi})~\hbox{pb} \nonumber \\ {\hbox{CMS}},~\sqrt{S}=7~{\rm TeV}~~&\sigma=&52.4\pm 2.0 (\hbox{stat})\pm 4.5 (\hbox{syst})\pm 1.2 (\hbox{lumi})~ \hbox{pb} \nonumber \\ {\hbox{CMS}},~\sqrt{S}=8~{\rm TeV}~~&\sigma=&69.9\pm 2.8 (\hbox{stat})\pm 5.6 (\hbox{syst})\pm 3.1 (\hbox{lumi})~ \hbox{pb}\, ,\end{aligned}$$ are slightly higher than the Standard Model predictions at next-to-leading order (NLO) in QCD [@Campbell:2011bn],[^1] $$\begin{aligned} &\sigma(\sqrt{S}=7~{\rm TeV})_{\hbox{Theory}}&= 47.04^{+2.02}_{-1.51}~\hbox{pb} \nonumber \\ &\sigma(\sqrt{S}=8~{\rm TeV})_{\hbox{Theory}}&=57.25^{+2.347}_{-1.60}~\hbox{pb}\, . \label{theorypred}\end{aligned}$$ The slight differences between the measured LHC values and the MCFM NLO predictions have led to speculation that the measured $W^+W^-$ cross section is a subtle sign of new physics [@Feigl:2012df; @Curtin:2012nn; @Curtin:2013gta]. The NLO QCD corrections to $pp\rightarrow W^+W^-$ were computed in Refs. [@Frixione:1993yp; @Ohnemus:1991kk], and then extended to include leptonic decays in Ref. [@Dixon:1998py]. The NLO predictions for the total cross section have a $3-4\%$ uncertainty at the LHC due to the choice of parton distribution functions (PDFs) and renormalization/factorization scale variation [@Campbell:2011bn]. The contribution from $gg\rightarrow W^+W^-$ is formally NNLO, but numerically contributes $\sim 3\%$ at $\sqrt{S}=7$ TeV and $\sim 4\%$ at $\sqrt{S}=14$ TeV [@Dicus:1987dj; @Glover:1988fe; @Binoth:2006mf; @Binoth:2005ua]. The NLO results have been interfaced with a shower Monte Carlo using the formalism of the POWHEG box [@Melia:2011tj; @Hamilton:2010mb; @Hoche:2010pf]. The electroweak corrections and the contribution from the $\gamma\gamma$ initial state are also known and contribute less than $1-2\%$ to the total cross section at the LHC [@Bierweiler:2012kw; @Baglio:2013toa]. These corrections are enhanced at large values of the $W^+W^-$ invariant mass, but have opposite signs and largely cancel. In this paper, we extend these results by including a resummation of threshold logarithms in the prediction of $W^+W^-$ production. Previously, the resummation of large logarithms associated with gluon emission at low transverse momentum, $p_T$, in $W^+W^-$ production was considered [@Grazzini:2005vw]. Unlike $p_T$ resummation which is normalized to the NLO cross section, the emission of soft gluons near threshold can potentially enhance the rate. We consider the next-to-leading logarithmic (NLL) and next-to-next-to-leading logarithmic (NNLL) resummation of threshold corrections. To accomplish this, we utilize the formalism of soft collinear effective theory (SCET) [@Bauer:2000ew; @Bauer:2000yr; @Bauer:2001yt; @Beneke:2002ph] which allows the resummation to be performed directly in momentum space [@Becher:2006mr; @Becher:2006nr]. This formalism has been used for processes with colorless final states such as Drell-Yan [@Becher:2007ty], Higgs production [@Ahrens:2008qu; @Ahrens:2008nc; @Ahrens:2010rs], associated W/Z plus Higgs production [@Dawson:2012gs], direct photon production [@Becher:2009th] and SUSY slepton pair production [@Broggio:2011bd]. The SCET formalism has also been applied to top quark pair production to resum the threshold corrections to the invariant mass distribution and to the total cross section [@Ahrens:2010zv; @Ahrens:2011mw; @Ahrens:2011px]. The total cross section for top quark pair production using threshold resummation has been obtained using two different sets of threshold limits: one starting from the invariant mass distribution of the top quark pair (pair invariant mass kinematics) and the other beginning from the transverse momentum or rapidity distribution of the top quark[@Ahrens:2010zv; @Ahrens:2011mw; @Kidonakis:2010dk]. The total cross section is then obtained by integrating over the resummed distributions. Within the theoretical uncertainties, the total cross sections for top quark pair production obtained with the different starting points are in reasonable agreement[@Ahrens:2011mw; @Ahrens:2011px]. The total cross section can also be obtained in the threshold limit, $\beta\rightarrow 0$, where $\beta$ is the top quark velocity, and the terms of ${\cal O}(\alpha_s^n\ln^m\beta)$ resummed[@Beneke:2009ye]. At the LHC and the Tevatron, however, the largest contributions to the total cross section for top quark pair production are not from the $\beta\rightarrow 0$ region. In this work, we will use pair invariant mass kinematics for the $W^+W^-$ final state to obtain our resummed results. The calculation of differential cross sections involves several scales. We consider the threshold limit $z\equiv {M_{WW}^2\over {s}} \rightarrow 1$ which dynamically becomes important due to the fast decline of the parton luminosity function as $\tau$ increases [@Becher:2007ty], with $M_{WW}$ the invariant mass of the $W^+W^-$ pair and ${s}$ the partonic center of mass energy. Near the partonic threshold, up to subleading powers of $(1-z)$, the cross section factorizes into a soft function which describes the soft gluon emissions and a hard function which includes the virtual corrections to the cross section. We can combine the NLO soft and hard functions with their renormalization group (RG) evolution equations to give NNLL resummed results which resum large logarithms of the form $\alpha_s^n\biggl({\ln^m(1-z)\over 1-z}\biggr)_+$ with $ m \leq 2n-1$. Alternatively, the RG evolution of the hard function, known to NNLO, can be matched with exact NLO results for the hard function to obtain the approximate NNLO hard function which includes the leading scale dependent contributions. Combined with the known NNLO soft function [@Becher:2007ty] for color singlet production, we are able to obtain the approximate NNLO result as an alternative to the NLO+NNLL resummed result. The advantage of the NLO+NNLL resummed results is that they contain powers of $\alpha_s$ to all orders, while the advantage of the approximate NNLO results is that we used the soft function to one order higher (NNLO) and the results do not contain higher orders of $\alpha_s$ which are sometimes not desired. In any case, the two results are extremely close to each other, and we recommend our approximate NNLO result as the most precise theoretical results available to be compared with future experiments, because it turns out to have smaller scale variations. In Section \[basics.sec\], we review the formalism and SCET resummation in the threshold region. This follows closely the approach of Ref. [@Ahrens:2010zv]. Section \[results.sec\] contains results for the NNLL differential and total cross sections, along with approximate results for the NNLO cross section for $pp\rightarrow W^+W^-$. Brief conclusions are presented in Section \[conc.sec\]. Basics {#basics.sec} ====== In this section, we review the fixed order results for $pp\rightarrow W^+W^-$ (\[born\] and \[nlo\]), the SCET formalism used to derive the RG improved NNLL results for the differential and total cross sections, including the matching to the fixed order NLO result (\[thresh\_res\]), and the derivation of an approximate NNLO result(\[approxnnlo\]). Born Level Result {#born} ----------------- The Born level process arises through the annihilation process $$\begin{aligned} q(p_1)+{\overline q}(p_2)\rightarrow W^+(p_3)+W^-(p_4)\, .\end{aligned}$$ This annihilation proceeds via $s$-channel $\gamma/Z$ exchange and a $t$-channel contribution, as shown in Figs. \[feyn.FIG\](a) and (b), respectively. The partonic cross section is $$\begin{aligned} {\hat \sigma}^0_{q{\overline q}} &=& {1\over 2 s}\int d\Phi_2 \mid A^0_{q{\overline q}}(s,t)\mid^2\, ,\end{aligned}$$ where the partonic level invariants are $$\begin{aligned} s&=& (p_1+p_2)^2\nonumber \\ t&=& (p_1-p_3)^2=M_W^2-{s\over 2}(1-\beta\cos\theta)\nonumber \\ u&=& (p_1-p_4)^2\, ,\end{aligned}$$ with $s+t+u=2M_W^2$ and $\beta=\sqrt{1-{4M_W^2\over s}}$. At the Born level we have $M_{WW}^2=s$. The two body phase space is $$\begin{aligned} d\Phi_2&=& {1\over 8 \pi s} dt={\beta\over 16\pi } d\cos\theta\, .\end{aligned}$$ Finally, the color-averaged and spin-summed and averaged matrix element squared is $$\mid A^0_{q{\overline q}}\mid^2={1\over 4 N_C} \biggl\{ c_{q}^{tt}F_q^0(s,t)+c_{q}^{ss}(s)K_q^0(s,t)-c_q^{ts}(s)J_q^0(s,t)\biggr\} \, , \label{lores}$$ where $c_{q}^{ss}K_q^0$ is the $s$-channel contribution, $c_{q}^{tt}F_q^0$ is the $t$-channel contribution, and $c_q^{ts}J_q^0$ is from $s$- and $t$-channel interference. The results have been found in Ref. [@Frixione:1993yp] and are given in Appendix  \[appa\] for convenience. Due to the collinear factorization [@Collins:1989gx; @Bodwin:1984hc; @Collins:1985ue; @Collins:1988ig; @Aybat:2008ct], the hadronic level cross section is obtained by convolving the partonic level cross section with PDFs. In general, the hadronic cross section can be written as $$\begin{aligned} \displaystyle {d^2\sigma\over dM_{WW} d\cos\theta} =\sum_{ij} {\beta_W\over 16 \pi M_{WW} S} \int _\tau^1 {dz\over z} {\cal{ L}}_{ij} \biggl({\tau\over z}, \mu_f\biggr) C_{ij}(z,M_{WW},\cos\theta,\mu_f) \label{had_born}\end{aligned}$$ where the sum runs over all possible initial state partons, $\mu_f$ is the factorization scale, $$\tau = \frac{M^2_{WW}}{S},\quad z=\frac{M^2_{WW}}{s},\quad{\rm and}\quad \beta_W=\sqrt{1-\frac{4 M^2_W}{M^2_{WW}}}\,.$$ The long-distance collinear physics is described by the parton luminosity $${\cal L}_{ij}\biggl(y,\mu_f\biggr)= \int_y^1{dx\over x}f_i(x,\mu_f)f_j\biggl({y\over x},\mu_f\biggr),$$ where $f_i$ is the PDF of a parton with flavor $i$. The coefficient functions describe the hard partonic process and can be written as a power series in $\alpha_s$, $$C_{ij}=C_{ij}^0+{\alpha_s\over 4\pi}C_{ij}^1+\ldots \, . \label{coeffs}$$ We have chosen the normalization of Eq. (\[had\_born\]) such that the leading order coefficient is $$C_{ij}^0=\delta(1-z)\mid A_{ij}^0\mid^2\, ,$$ and the Born level cross section is recovered. NLO result {#nlo} ---------- At NLO, the scattering coefficients of Eq. (\[had\_born\]) receive corrections from virtual loops and real gluon emission in the $q {\overline q}$ channel, along with tree level contributions from $qg\rightarrow q W^+W^-$. In dimensional regularization with $N=4-2\epsilon$, the one-loop virtual diagrams contribute $$\sigma_{VIRT}={1\over 2 s}\int d\Phi_2\mid A^1_{q{\overline q}}(s,t)\mid^2\, ,$$ where $$\begin{aligned} \mid A_{q{\overline q}}^1\mid^2&=& \mid A_{q{\overline q},reg}^1\mid^2+\mid A_{q{\overline q},div}^1\mid^2 \\ \mid A_{q{\overline q},div}^1\mid^2 &=&- {\alpha_s\over 4\pi}C_F \biggl({4\pi\mu^2\over s}\biggr)^\epsilon \Gamma(1+\epsilon) \biggl({4\over \epsilon^2}+{6\over \epsilon}\biggr)\mid A^0_{q{\overline q}}(s,t)\mid^2\nonumber \\ \mid A_{q{\overline q},reg}^1\mid^2&=&{\alpha_s\over 16 \pi N_C}C_F \biggl\{ c_{q}^{tt}F_q^1(s,t)+c_{q}^{ss}(s)K_q^1(s,t)-c_q^{ts}(s)J_q^1(s,t)\biggr\} \, . \label{nlores}\end{aligned}$$ We note that since QCD does not renormalize electroweak couplings, all the UV divergences cancel leaving only IR divergences in $A^1_{q\bar{q},div}$. The one-loop corrections to the $t$-channel exchange are given by $c_{q}^{tt}F_q^1(s,t)$, to the $s$-channel exchange by $c_{q}^{s s}K_q^1(s,t)$, and the interference between the $s$- and $t$-channels by $c_q^{ts}J_q^1(s,t)$. Expressions for these terms can be found in Appendix \[appa\]. As will be discussed in the next section, the real hard gluon emission contribution is not relevant for the resummation of the threshold logarithms and we therefore do not give it here, although it can be found in Ref. [@Frixione:1993yp]. Threshold Resummation and Matching {#thresh_res} ---------------------------------- We now discuss the resummation of the large logarithms in the partonic threshold limit, $z\rightarrow 1$. In this limit, since there is no phase space available for hard gluon emission, the total phase space is well described by the Born $2\rightarrow 2$ process and Eq. (\[had\_born\]) can be used. In addition to the collinear factorization, in the threshold limit the coefficient functions can be further factorized into hard, $H$, and soft, $S$, functions:[^2] $$C_{q{\overline q}}(z,M_{WW},\cos\theta,\mu_f)\equiv H(M_{WW},\cos\theta,\mu_f)S(\sqrt{s}(1-z),\mu_f)+ {\cal O}(1-z)\, . \label{hsdef}$$ The soft function is given by the vacuum expectation values of soft Wilson loops [@Becher:2006mr; @Becher:2007ty] and the hard function is calculated by matching the full QCD result onto the relevant SCET operator. It is this matching that integrates out the hard QCD modes and leaves the soft and collinear modes that comprise SCET. The hard function is given by[^3] $$\begin{aligned} H(M_{WW},\cos\theta,\mu)=|C_{WW}(M_{WW},\cos\theta,\mu)\mathcal{O}_{WW}|^2\, , \label{wilscoeff}\end{aligned}$$ where $C_{WW}$ is the Wilson coefficient of the relevant SCET operator $\mathcal{O}_{WW}$. The Wilson coefficient is calculated by matching the renormalized QCD and SCET amplitudes: $$\begin{aligned} \mathcal{M}^{\rm ren}(\epsilon,M_{WW},\cos\theta)=Z(\epsilon,M_{WW},\mu)\, C_{WW}(M_{WW},\cos\theta,\mu)\mathcal{O}_{WW}, \label{zren}\end{aligned}$$ where $\mathcal{M}^{\rm ren}$ is the renormalized QCD amplitude and $Z$ is the SCET renormalization constant. In dimensional regularization, SCET loops are scaleless and vanish. This implies that the UV and IR singularities of SCET coincide and cancel. Since SCET and QCD describe the same low-scale physics and have the same IR pole structure, $Z$ can be determined by the behavior of IR singularities in QCD [@Giele:1991vf; @Kunszt:1994np; @Catani:1996jh; @Catani:1996vz; @Catani:1998bh; @Becher:2009cu; @Becher:2009qa]. In the $\overline{\rm MS}$ scheme, we have $$\begin{aligned} Z(\epsilon,M_{WW},\cos\theta,\mu) = 1-\frac{\alpha_s C_F}{2\pi}\left(4\pi\right)^\epsilon e^{-\epsilon\gamma_E}\left\{\frac{1}{\epsilon^2}+\frac{1}{\epsilon}\left(\ln\frac{\mu^2}{-M^2_{WW}}+\frac{3}{2}\right)\right\}\, . \label{SCETren}\end{aligned}$$ The poles in $Z$ and the NLO QCD squared amplitudes in Eq. (\[nlores\]) cancel. Hence, the one-loop contribution to the hard function is just the finite terms of Eq. (\[nlores\]) [@Manohar:2003vb; @Becher:2006nr]. Since the hard function is calculated in the perturbative region of QCD it can be expanded in powers of $\alpha_s$: $$H(M_{WW},\cos\theta,\mu_h)= H^0(M_{WW},\cos\theta)+{\alpha_s(\mu_h)\over 4\pi} \biggl[H^1_{reg}(M_{WW},\cos\theta,\mu_h)+ H^1_{extra}(M_{WW},\cos\theta,\mu_h)\biggr], \label{HNLO}$$ where $\mu_h$, termed the hard scale, is the scale at which QCD and SCET are matched. The normalization of the hard function is such that $$H^0(M_{WW},\cos\theta)=\mid A^0_{q{\overline q}}\mid^2\, , \label{H0}$$ and $$\begin{aligned} {\alpha_s\over 4\pi} H^1_{reg}(M_{WW},\cos\theta,\mu_h)&=&\mid A_{q{\overline q},reg}^1\mid^2 \nonumber \\ H^1_{extra}(M_{WW},\cos\theta,\mu_h)&=&- C_F H^0(M_{WW},\cos\theta)\nonumber\\ &&\times\biggl\{{\pi^2\over 3}+ 2\log^2\biggl({M_{WW}^2\over \mu_h^2}\biggr) -6\log\biggl({M_{WW}^2\over \mu_h^2}\biggr)\biggr\}\, . \label{H1}\end{aligned}$$ Now we have all the pieces to resum the threshold logarithms. As mentioned before, the hard function is calculated at the matching scale $\mu_h$. Since the soft function describes the soft physics, it is evaluated at a soft scale, $\mu_s$, associated with the scale of soft gluon emission. By using the RG equations (RGEs), the hard and soft functions can be evolved to the factorization scale. The RG evolution of the soft function resums logs of the form $\alpha_s^n\ln^m(\mu_s/M_{WW})$. By choosing the soft scale $\mu_s\sim M_{WW}(1-\tau)$, the RGE running resums the large threshold logarithms. In Table \[scetorders\] we list the accuracy of the resummation at a given order. The resulting coefficient function is [@Becher:2007ty] $$\begin{aligned} C(z,M_{WW},\cos\theta,\mu_f)&=& U(M_{WW},\mu_h,\mu_s,\mu_f)\,H(M_{WW},\cos\theta,\mu_h) \nonumber\\ &&\times{\tilde s}\biggl( \ln\biggl({M_{WW}^2\over \mu_s^2}\biggr)+\partial_\eta,\mu_s\biggr) \, {e^{-2\gamma_E\eta}\over\Gamma(2\eta)} {z^{-\eta}\over (1-z)^{1-2\eta}}\end{aligned}$$ where $\eta=2a_\Gamma(\mu_s,\mu_f)$ and $$\begin{aligned} U(M_{WW},\mu_h,\mu_s,\mu_f) = \left(\frac{M^2_{WW}}{\mu^2_h}\right)^{-2a_\Gamma(\mu_h,\mu_s)}\exp\left[4S(\mu_h,\mu_s)-2a_{\gamma^V}(\mu_h,\mu_s)+4a_{\gamma^\phi}(\mu_s,\mu_f)\right].\end{aligned}$$ Finally, ${\tilde s}$ is can be expressed as a power expansion in logarithms, $$\begin{aligned} {\tilde s}(L,\mu)&=&1+ {\alpha_s\over 4 \pi}\sum_{n=0}^2s^{(1,n)}L^n +\biggl({\alpha_s\over 4 \pi}\biggr)^2 \sum_{n=0}^4 s^{(2,n)}L^n.\end{aligned}$$ Expressions for the $s^{(i,n)}$ are given in Appendix \[appb\] and are identical to those found for Drell-Yan [@Becher:2007ty]. We will present results both at NLL and NNLL. The corresponding order of the needed functions is given in Table \[scetorders\]. Explicit expressions for the functions $a$ and $S$, and the anomalous dimensions $\Gamma$, $\gamma^V$, and $\gamma^\phi$ can be found in the appendices of Ref. [@Becher:2007ty]. The choice of the soft scale, $\mu_s$ ,is discussed in Section \[difs\]. The resummed results are only valid in the region $z\rightarrow 1$. To extend these results to all $z$, the resummed cross section needs to be matched with the full fixed order cross section. This allows the inclusion of the non-singular terms in $(1-z)$ which are present in the fixed order result but not the resummed result. For NNLL resummation, this means matching with the NLO cross section: $$\begin{aligned} d\sigma^{NLO+NNLL}&\equiv & d\sigma^{NNLL}(\mu_h,\mu_f,\mu_s)+\biggl(d\sigma^{NLO}(\mu_f)-d\sigma^{NLO}(\mu_f)\mid_{leading~singularity}\biggr) \label{matchdef}\end{aligned}$$ where $d\sigma^{NNLL}$ is the threshold resummed cross section and $d\sigma^{NLO}\mid_{leading~singularity}$ contains only the ${\cal O}(\alpha_s)$ NLO terms which are singular as $z\rightarrow 1$, $$d\sigma^{NLO}\mid_{leading~singularity}\equiv d\sigma^{NNLL}\mid_{\mu_h=\mu_f=\mu_s}\, . \label{matching}$$ Subtracting $d\sigma^{NLO}\mid_{leading~singularity}$ prevents double counting of terms common to the resummed and fixed order results. Also, in the limit $z\rightarrow 1$, the matched cross section corresponds to the resummed results, as desired. Order Accuracy: $\alpha_s^n \ln^m(\mu_s/M_{WW})$ $\Gamma_{cusp}$ $\gamma^h,\gamma^\phi$ $H, {\tilde s}$ ------- -------------------------------------------- ----------------- ------------------------ ----------------- NLL $2n-1 \leq m \leq 2n$ 2-loop 1-loop tree NNLL $2n-3 \leq m \leq 2n$ 3-loop 2-loop 1-loop : Accuracy of the SCET resummation at a given order and required accuracy of SCET inputs. \[scetorders\] Approximate NNLO Results {#approxnnlo} ------------------------ The full NNLO cross section can only be determined from a complete calculation. However, the scale dependent terms that are singular as $z\rightarrow1$ can be determined to NNLO accuracy via the known hard and soft functions and their respective RGEs. As will be shown in the next section, most of the NLO correction comes from the leading singular piece. Hence, we expect that including the scale dependent, leading singular pieces of the NNLO cross section is a good estimate of the full NNLO result. The inclusion of these pieces is known as approximate NNLO. The coefficient function in Eq. (\[hsdef\]) can be expanded in a power series, $$\begin{aligned} C(z,M,\cos\theta,\mu)&=&C^{0}(z,M,\cos\theta,\mu)+\frac{\alpha_s}{4\pi}C^{1}(z,M,\cos\theta,\mu) \nonumber \\ && +\left(\frac{\alpha_s}{4\pi}\right)^2C^{2}(z,M,\cos\theta,\mu) \, .\end{aligned}$$ The leading order, $C^{0}$, and NLO, $C^{1}$, contributions are fully known analytically. The NNLO contribution, $C^{2}$, can only be approximately determined from the hard and soft functions. The soft function is known fully to NNLO [@Becher:2007ty]; hence the only approximation comes from the unknown scale-independent NNLO piece of the hard function. The approximate NNLO cross section is found by calculating the scale-dependent, leading singular pieces of $C^{2}$ and adding this contribution to the full NLO result. We expand the hard function as a power series in $\alpha_s$: $$\begin{aligned} H_{approx}(M_{WW},\cos\theta,\mu_f)=H(M_{WW},\cos\theta,\mu_f)+\left(\frac{\alpha_s}{4\pi}\right)^2H^{2}_{approx}(M_{WW},\cos\theta,\mu_f), \label{Happrox}\end{aligned}$$ where the full NLO hard function, $H$, is given in Eq. (\[HNLO\]). The NNLO hard piece, $H^{2}_{approx}$, contains only the scale dependent pieces at NNLO. We further expand $H^{2}_{approx}$ in a power series of logs: $$\begin{aligned} H^{2}_{approx}(M_{WW},\cos\theta,\mu_f)=\sum^3_{n=1}h^{(2,n)}L_{WW}^n, \label{H2}\end{aligned}$$ where $L_{WW}=\ln(M^2_{WW}/\mu^2_f)$. The coefficients $h^{(2,n)}$ can be found in Appendix \[appb\]. The approximate NNLO hard function in Eq. (\[Happrox\]) is independent of scale to order $\alpha_s^3$. Hence, scale variation only contains the $\mathcal{O}(\alpha_s^3)$ uncertainties, not taking into account the unknown NNLO scale independent and nonsingular in $(1-z)$ pieces at $\mathcal{O}(\alpha_s^2)$. Variation of the factorization scale may therefore underestimate the total uncertainty in the approximate NNLO result. However, this uncertainty can be further estimated by noting that there is an ambiguity in the logs used to expand $H^{2}$. For example, introduce a scale $Q_h\sim M_{WW}$. Then $H^{2}$ can be expanded in logs of the form $L_Q=\ln(Q^2_h/\mu^2_f)$ instead of $L_{WW}$. The difference between these two schemes will be order one contributions to the scale independent piece. Hence, in addition to the variation of the factorization scale, the uncertainty associated with the unknown NNLO scale independent piece is estimated by making the replacement $L_{WW}\rightarrow L_Q$ in Eq. (\[H2\]) and varying the new scale $Q_h$ around the central value $M_{WW}$, which is the natural choice from the RGEs. We note that when including the scale independent pieces, the full NNLO hard function is independent of the scale $Q_h$. Similarly, since the soft function is known fully to NNLO there is no ambiguity in the choice of scales used in the log expansion. Results {#results.sec} ======= Soft Scale Choice {#scale} ----------------- In the process of performing resummation in the SCET formalism, two additional scales are introduced: the hard scale $\mu_h$, where the hard function is evaluated; and the soft scale, $\mu_s$, where the soft function is evaluated. Since the hard function is calculated from matching QCD onto SCET at the scale of the hard scattering process, the central value of $\mu_h$ is naturally chosen to be the scale of the hard scattering process, $\mu_h=M_{WW}$. The soft scale is chosen to be associated with the scale of the soft gluon emissions, such that the RG evolution resums large logs associated with soft gluon emission at threshold. Following Refs. [@Becher:2006nr; @Becher:2006mr; @Becher:2007ty], we choose the soft scale to be related to the hadronic energy scale, avoiding the Landau poles that plague the traditional perturbative QCD resummation [@Catani:1996yz; @Becher:2006nr]. Hence, in the hadronic threshold limit $\tau\rightarrow1$ we want $\mu_s\sim M_{WW}(1-\tau)$. However, at hadron colliders most of the cross section is accumulated far from $\tau=1$ and the choice of soft scale away from this limit is less clear. Another constraint on $\mu_s$ is that the soft function should be a well-behaved perturbative series. Hence, away from the threshold region, we choose $\mu_s$ such that the $\mathcal{O}(\alpha_s)$ piece of the soft function is minimized relative to the LO piece. Figure \[musdep.FIG\] shows the ratio of the NNLL-resummed invariant mass distribution evaluated with only the $\mathcal{O}(\alpha_s)$ piece of the soft function, $(d\sigma^{NNLL}_{WW})_{\alpha_s}$, to the NNLL-resummed distribution evaluated with only the LO piece of the soft function, $(d\sigma^{NNLL}_{WW})_0$. In both distributions, all other pieces of the resummed cross sections (aside from the soft function) are evaluated to full NNLL order. This ratio is shown for $\sqrt{S}=8$ TeV at various values of $M_{WW}$ with $\mu_f=\mu_h=M_{WW}$ and MSTW2008nnlo PDFs. The soft scale is chosen to correspond to the minimum of these ratios, which is found to be well described by the parameterization $$\begin{aligned} \mu_s = M_{WW} \frac{(1-\tau)}{(a+b\sqrt{\tau})^c}.\end{aligned}$$ For $\sqrt{S}=8$ TeV, it is found that $a=1.542$, $b=6.27$, and $c=1.468$. Performing a similar fit at $\sqrt{S}=14$ TeV it is found that $a=1.544$, $b=6.123$, and $c=1.499$. With this parameterization $\mu_s$ has the correct dependence on $\tau$ in the threshold region. Figure \[musChoice.FIG\] shows the minimum value of the ratio $\mu_s/M_{WW}$ at $\sqrt{S}=8$ TeV (solid) and $\sqrt{S}=14$ TeV (dashed) as a function of $\tau$. As can be clearly seen, the hadronic energy makes little difference to the choice of the soft scale. For simplicity, independent of the hadronic energy scale, all results presented here use the central value of the soft scale corresponding to the 8 TeV solution: $$\begin{aligned} \mu_s^{\rm min} = M_{WW} \frac{(1-\tau)}{(1.542+6.27\sqrt{\tau})^{1.468}}.\end{aligned}$$ Differential Cross Section {#difs} -------------------------- ![Ratio of the contribution of the leading singularity to the fixed order NLO cross section at $\sqrt{S}=8$ and $14$ TeV using MSTW2008 PDFs. NNLO (NLO) PDFs are used for the NNLL leading contribution (fixed order NLO contribution). []{data-label="leadingfig"}](MWW_leading_NLO_ratio.eps){width=".6\textwidth"} We begin by considering the validity of the matching of the NNLL results to the fixed order NLO results. In order for the matching of Eq. (\[matchdef\]) to be valid, the sub-leading terms in $(1-z)$ must be small. In Fig. \[leadingfig\], we show the contribution of the leading singularity to the fixed order NLO differential cross section for $\sqrt{S}=8$ TeV and $\sqrt{S}=14$ TeV. We fix the central scale to be $\mu_f=2M_W$. From this figure, we see that the leading singularity captures $\sim 90\%$ of the NLO fixed order cross section. Hence, the threshold singularities contribute most of the NLO cross section and we may expect that by resumming the higher order logarithms we capture most of the higher order cross section. In Fig. \[mww\_14\_all\], we show $d \sigma/dM_{WW}$ versus $M_{WW}$ for $\sqrt{S}= 14$ TeV, with MSTW2008 PDFs. The curves are LO, NLO, NLL (matched), and NNLL (matched), with $\mu_h=M_{WW}$, $\mu_s=\mu_s^{\rm min}$ and $\mu_f$ varied up and down by a factor of $2$ from the central value of $\mu_f^0=2M_W$. It is apparent that the NNLL resummation slightly increases the rate at the peak. A change in the $W^+W^-$ invariant mass distribution may be consequential to the analysis of the $H\rightarrow W^+W^-\rightarrow 2\ell 2\nu$ decay channel. In the zero jet bin, the major background is the SM (non-Higgs) production of $W^+W^-$ [@CMS:bxa; @ATLAS:2013wla]. To estimate this background a sideband analysis is performed. A control region is defined with a minimum dilepton invariant mass, where the $W^+W^-$ background strongly dominates the Higgs signal. The control region is used to normalize the cross section and then Monte Carlo is used to extrapolate the line shapes into the signal region. If higher order corrections alter the $M_{WW}$ distribution, the dilepton invariant mass distribution will be changed and the extrapolation to the signal region will need to take this into account. In Fig. \[NLO\_ratio\] we explore the effect of higher order corrections on the invariant mass distribution by plotting the ratio of the NLO+NNLL matched and NLO $M_{WW}$ distributions. The scales are set to be $\mu_h=M_{WW}$, $\mu_s=\mu_s^{\rm min}$, and $\mu_f = 2 M_W$. For the NLO cross section evaluated with NLO PDFs (solid), the resummation increases the invariant mass distribution by $\sim 3-4\%$ in the peak region for both $\sqrt{S}=8$ and $14$ TeV and decreases it by $\sim 2\%$ and $\sim 1\%$ in the high mass region for $\sqrt{S}=8$ and $14$ TeV, respectively. However, most of this change in the $M_{WW}$ distribution is from the different PDFs used for the resummed and NLO results, as can be seen when the NLO cross section is evaluated with NNLO PDFs (dashed). In this case, the resummation only alters the invariant mass distribution by $\lesssim 1\%$ for a wider range $M_{WW}$. This indicates that the calculation of the $W^+W^-$ cross section is firmly under theoretical control. The factorization scale dependence of the invariant mass distributions is shown in Fig. \[matched\_nnll\] for (a) the NNLL resummed and NLO leading, and (b) the NLO and NNLL matched results. Here we present the factorization scale dependence as a percent difference from the central value. Using the definition of matching in Eq. (\[matchdef\]), Fig. \[NNLL\_lead\] indicates that there is a cancellation between the $\mu_f$ dependencies the NNLL resummed and NLO leading results for $M_{WW}\lesssim 400$ GeV. Comparing Figs. \[matched\_nnll\] (a) and (b), it is apparent that for $M_{WW}\gtrsim190$ GeV the NLO $\mu_f$ dependence also cancels against the NNLL resummed dependence. Hence, although the cancellation is not as efficient at lower $M_{WW}$, as the invariant mass increases the $\mu_f$ dependence of the NNLL matched result is less than that of the NLO result. This can be seen in Fig. \[NNLL\_NLO\], where we see that that for $M_{WW}\gtrsim 220-230$ GeV the $\mu_f$ dependence of the NNLL matched result is lower than the NLO result. Hence, although the scale dependence of the NNLL matched result is larger than that of the NLO result at the peak of the invariant mass distribution, one can show that the resummation and matching procedure decreases the factorization scale dependence of the total cross section relative to the NLO result. ![Scale dependence of the $\sigma^{NNLL}$ differential cross sections at $\sqrt{S}=14$ TeV using MSTW2008nnlo PDFs. The scales are varied by a factor of 2 up and down from the central scales, $\mu_h^0=M_{WW}$, $\mu_s^0=\mu_s^{\rm min}$ and $\mu_f^0=2M_W$.[]{data-label="mww_14_scales"}](MWW_NNLL_Scale_14TeV.eps){width=".6\textwidth"} In Fig. \[mww\_14\_scales\], we show the deviation from the central scales for the NNLL resummed differential cross section, $d \sigma^{NNLL}/dM_{WW}$, versus $M_{WW}$ for $\sqrt{S}= 14$ TeV, with MSTW2008nnlo PDFs. Again, we present the scale dependence as a percent difference from the central value. The central scales are $\mu_h^0=M_{WW}$, $\mu_s^0=\mu_s^{\rm min}$ and $\mu_f^0=2M_W$ and are separately varied up and down by a factor of $2$. The hard and soft scale variations are of the order of $\sim 1-2\%$ and are relatively independent of $M_{WW}$. The factorization scale dependence near the peak is $\sim \pm 6\%$ and is always greater than the hard and soft scale dependencies for the invariant mass range presented. Finally, in Fig. \[tot\_14\_scales\] we show the variation of the NNLL matched cross section with the central scale choices $\mu_h^0=M_{WW}$, $\mu_s^0=\mu_s^{\rm min}$ and $\mu_f^0=2M_W$ at $\sqrt{S}=14$ TeV. The cross section varies by less than $\pm 2\%$ as the scales are varied from the central values. Again, we see that there is a large cancellation of the factorization scale when computing the matched cross section. In contrast to the NNLL results in Fig. \[mww\_14\_scales\], the $\mu_f$ dependence of the matched result is less than (similar to) the hard and soft scale dependencies for $\mu<\mu^0$ ($\mu>\mu^0)$. Also, note that unlike the soft and factorization scales, the hard scale dependence, with a minimum near the central value, actually never decreases but always increases the total cross section as it is varied from the central value. This explains why in Fig. \[mww\_14\_scales\] the effect of the hard scale variation was always to increase the differential cross section value above the central value. ![Scale dependence of the matched $\sigma^{NLO+NNLL}$ cross sections at $\sqrt{S}=14$ TeV using MSTW2008nnlo PDFs. The scales are varied by a factor of 2 up and down from the central scales, $\mu_h^0=M_{WW}$, $\mu_s^0=\mu_s^{\rm min}$ and $\mu_f^0=2M_W$. []{data-label="tot_14_scales"}](ScaleVar_14TeV.eps){width=".6\textwidth"} Total Cross Section ------------------- .2in $\sigma({\rm pb})$ $\sqrt{S}=7$ TeV $\sqrt{S}=8$ TeV $\sqrt{S}=13$ TeV $\sqrt{S}$=14 TeV --------------------------------- ------------------------ ------------------------ ----------------------- ------------------------ $\sigma^{NLO}$ $45.7^{+1.5}_{-1.1}$ $55.7^{+1.7}_{-1.2}$ $110.6^{+2.5}_{-1.6}$ $122.2^{+2.5}_{-1.8}$ $\sigma^{gg}$ $1.0^{+0.3}_{-0.2}$ $1.3^{+0.4}_{-0.3}$ $3.5^{+0.9}_{-0.7}$ $4.1^{+0.9}_{-0.7}$ $\sigma^{NLO+NNLL}$ $44.9_{-0.6}^{+0.6}$ $54.8^{+0.7}_{-0.8}$ $108.2^{+1.3}_{-1.5}$ $119.5^{+1.5}_{-1.6}$ $\sigma^{\prime NLO+NNLL}$ $45.9^{+0.5}_{-0.6}$ $56.1^{+0.7}_{-0.8}$ $111.7^{+1.8}_{-1.6}$ $123.6^{+2.0}_{-1.8}$ $\sigma^{NNLO}_{approx}$ $45.0^{+0.4}_{-0.1} $ $54.9^{+0.5}_{-0.05} $ $108.3_{-0.4}^{+1.0}$ $119.6_{-0.5}^{+1.2} $ $\sigma^{\prime NNLO}_{approx}$ $46.0^{+0.4}_{-0.047}$ $56.2^{+0.6}_{-0.1}$ $111.8^{+1.7}_{-1.1}$ $123.7^{+1.8}_{-1.2}$ : Total cross sections for $pp\rightarrow W^+W^-$ with $\mu_f^0=2M_{W}$, $\mu_h^0=M_{WW}$, $\mu_s^0=\mu_s^{\rm min}$, and $Q_h^0=M_{WW}$. The NLO $\sigma^{NLO}$ includes the $gg$ contribution and is evaluated with NLO PDFs and the remaining entries are evaluated with MSTW2008nnlo PDFs. The primed cross section $\sigma^{\prime NLO+NNLL}$ is the sum of $gg$ contribution $\sigma^{gg}$ and the matched $\sigma^{NLO+NNLL}$, while $\sigma^{\prime NNLO}_{approx}$ is the sum of $\sigma^{gg}$ and approximate NNLO, $\sigma^{NNLO}_{approx}$. The last row, $\sigma^{\prime NNLO}$, is our best prediction for the cross section. \[sigtab1\] In this section, we compile our final results for the total $W^+W^-$ cross sections at the LHC. In Tables \[sigtab1\] and \[sigtab2\], we show successively improved results for the total cross sections using MSTW2008 PDFs. Both tables fix the central value of $\mu_h^0=M_{WW}$ and $\mu^0_s=\mu_s^{\rm min}$. Table \[sigtab1\] takes the central factorization scale to be fixed at $\mu_f^0=2M_W$, while Table \[sigtab2\] uses a dynamical central scale, $\mu_f^0=M_{WW}$. The top line is the NLO result obtained from MCFM [@Campbell:2011bn] (which includes the $gg$ initial state) and is calculated using NLO PDFs. The second line of the tables is the $gg$ contribution, $\sigma^{gg}$, calculated using MCFM, but with NNLO PDFs (as is appropriate for combining with the NNLL and approximate NNLO results). The third and fourth rows contain the NNLL matched and approximate NNLO cross sections evaluated with NNLO PDFs but without the $gg$ contribution, $\sigma^{NLO+NNLL}$ and $\sigma^{NLO}_{approx}$, respectively. The fifth and sixth rows are the same as the third and fourth, but with the $gg$ contribution now included. The uncertainties in the matched cross section correspond to taking the central values of the hard, soft, and factorization scales and varying each separately up and down by a factor of $2$. The uncertainties in the approximate NNLO cross section correspond to varying the factorization and $Q_h$ scales by a factor of two around their central values. The resulting uncertainties are added in quadrature. As noted previously in the discussion of the previous subsection, the factorization scale dependence of the matched cross section is less than that of the NLO cross section. This cancellation is more extreme at $\sqrt{S}=7$ and $8$ TeV. Hence, even with hard and soft scale variation taken into account the scale dependence of $\sigma^{\prime NLO+NNLL}$ is less than that of $\sigma^{NLO}$. At $\sqrt{S}=13$ and $14$ TeV, once the uncertainties associated with hard and soft scale variation are taken into account, the scale uncertainty of $\sigma^{\prime NLO+NNLL}$ is similar to or greater than that of $\sigma^{NLO}$. .2in $\sigma({\rm pb})$ $\sqrt{S}=7$ TeV $\sqrt{S}=8$ TeV $\sqrt{S}=13$ TeV $\sqrt{S}$=14 TeV --------------------------------- ----------------------- ----------------------- ----------------------- ----------------------- $\sigma^{NLO}$ $44.8^{+1.2}_{-0.9}$ $54.7^{+1.4}_{-1.0}$ $108.8^{+1.2}_{-1.3}$ $120.3^{+2.0}_{-1.3}$ $\sigma^{gg}$ $0.9^{+0.2}_{-0.2}$ $1.2^{+0.3}_{-0.1}$ $3.3^{+0.8}_{-0.6}$ $3.7^{+0.7}_{-0.6}$ $\sigma^{NLO+NNLL}$ $44.7^{+0.5}_{-0.6}$ $54.6^{+0.6}_{-0.8}$ $108.1^{+1.4}_{-1.5}$ $119.4^{+1.6}_{-1.7}$ $\sigma^{\prime NLO+NNLL}$ $45.6^{+0.6}_{-0.6}$ $55.8^{+0.7}_{-0.8}$ $111.4^{+2.0}_{-1.8}$ $123.1^{+2.1}_{-2.0}$ $\sigma^{NNLO}_{approx}$ $44.8^{+0.4}_{-0.1}$ $54.7^{+0.4}_{-0.04}$ $108.2_{-0.4}^{+1.0}$ $119.6_{-0.6}^{+1.2}$ $\sigma^{\prime NNLO}_{approx}$ $45.7^{+0.4}_{-0.04}$ $55.9^{+0.5}_{-0}$ $111.5^{+1.6}_{-1.0}$ $123.3^{+1.7}_{-1.2}$ : Total cross sections for $pp\rightarrow W^+W^-$ with $\mu_f^0=M_{WW}$, $\mu_h^0=M_{WW}$, $\mu_s^0=\mu_s^{\rm min}$, and $Q_h^0 = M_{WW}$. The NLO $\sigma^{NLO}$ includes the $gg$ contribution and is evaluated with NLO PDFs and the remaining entries are evaluated with MSTW2008nnlo PDFs. The primed cross section $\sigma^{\prime NLO+NNLL}$ is the sum of $gg$ contribution $\sigma^{gg}$ and the matched $\sigma^{NLO+NNLL}$, while $\sigma^{\prime NNLO}_{approx}$ is the sum of $\sigma^{gg}$ and approximate NNLO, $\sigma^{NNLO}_{approx}$. The last row, $\sigma^{\prime NNLO}$, is our best prediction for the cross section. \[sigtab2\] The scale dependence of the approximate NNLO cross sections, $\sigma^{\prime NNLO}_{approx}$, at $\sqrt{S}=7$ and $8$ TeV is reduced by at least a factor of three relative to the NLO cross section, while at $\sqrt{S}=13$ and $14$ TeV the uncertainties of the NLO and $\sigma^{\prime NNLO}_{approx}$ cross sections are more similar. This is due to a cancellation of the factorization scale dependence of $\sigma^{gg}$ and $\sigma^{NNLO}_{approx}$ at $\sqrt{S}=7$ and $8$ TeV that is not present at $\sqrt{S}=13$ and $14$ TeV. This also explains why the scale variation of approximate NNLO cross section without $gg$ contribution is similar to (less than) that with the $gg$ contribution at $7$ and $8$ TeV ($13$ and $14$ TeV.) Finally, with $\mu^0_f=M_{WW}$ at $\sqrt{S}=8$ TeV, the zero in the scale uncertainty of $\sigma^{\prime NNLO}_{approx}$ indicates that the factor of $2$ of both the factorization and $Q_h$ scales increase the cross section. Although there is no significant difference between the cross section prediction of $\sigma^{\prime NNLO}_{approx}$ and $\sigma^{\prime NLO+NNLL}$, we consider $\sigma^{\prime NNLO}_{approx}$ to be our best prediction for the LHC $W^+W^-$ production cross sections, since $\sigma^{\prime NNLO}_{approx}$ for the most part has less scale variation than $\sigma^{\prime NLO+NNLL}$. By comparing the NLO and approximate NNLO, $\sigma^{\prime NNLO}_{approx}$, cross sections, it is apparent that the effect of the higher order corrections is to increase the cross section less than $\sim 1$ pb at $\sqrt{S}=8$ TeV and less than $\sim 3$ pb at $\sqrt{S}=14$ TeV, while reducing the theoretical uncertainty from scale variations. The matched NNLL cross section increases the NLO cross section by similar amounts, with a slightly increased scale uncertainty. There is very little difference between using a fixed factorization scale and a dynamic factorization scale. It appears that the prediction for the $W^+W^-$ cross section is under good theoretical control. Conclusion {#conc.sec} ========== Now that the Higgs boson is discovered, a full exploration of the electroweak sector has begun. An important signal for this exploration is the pair production of gauge bosons, in particular $W^+W^-$ production. This signal is a major background to $H\rightarrow W^+W^-$ and is sensitive to the electroweak gauge boson triple coupling, which directly probes the mechanism of electroweak symmetry breaking and the $SU(2)\times U(1)$ gauge structure, respectively. In order to be sensitive to new physics in the $W^+W^-$ signal and measure the $H\rightarrow W^+W^-$ decay channel well, it is important to have accurate and precise theoretical predictions for $W^+W^-$ production cross section and differential distributions. In this paper we resummed large logarithms associated with soft gluon emission at partonic threshold, $z=M^2_{WW}/s\rightarrow 1$, at NNLL order for $W^+W^-$ pair production. This resummation was performed using the formalism of SCET [@Bauer:2000ew; @Bauer:2000yr; @Bauer:2001yt; @Beneke:2002ph] which allows for the resummation directly in momentum space [@Becher:2006mr; @Becher:2006nr]. The NNLL resummed results were then matched onto the known NLO results [@Frixione:1993yp; @Ohnemus:1991kk]. We also calculated the approximate NNLO $W^+W^-$ cross section. We thus obtain the most accurate cross sections and invariant mass distributions for $W^+W^-$ production that have been calculated to date. We found that the effect of the threshold resummation on the invariant mass distribution was to increase the differential cross section $\sim 3-4\%$ in the peak region for both $\sqrt{S}=8$ and $14$ TeV. The matched NLO+NNLL and approximate NNLO cross section both increased the NLO cross section by $\sim 0.5-1.5\%$ for a factorization scale central value $\mu^0_f=2M_W$ and $\sim 2-3\%$ for a central scale of $\mu^0_f = M_{WW}$, within the theoretical uncertainties. The theoretical uncertainties of the approximate NNLO cross section were generally decreased relative to those of the NLO cross section. These results indicate that the sideband analysis used for $W^+W^-$ background estimation to $H\rightarrow W^+W^-$ signal [@CMS:bxa; @ATLAS:2013wla] is not significantly altered by higher order corrections. Also, the strong coupling constant perturbative expansion of the $W^+W^-$ production cross section is firmly under theoretical control. Acknowledgements {#acknowledgements .unnumbered} ================ S.D. and I.L. are supported by the U.S. Department of Energy under grant No. DE-AC02-98CH10886. M.Z. is supported by the National Science Foundation, grant PHY-0969739. We thank George Sterman and Andrea Ferroglia for helpful conversations. Fixed Order Results {#appa} =================== Lowest Order Results {#appa:LO} -------------------- The coefficients of Eq. (\[lores\]) are [@Frixione:1993yp] $$\begin{aligned} c_q^{tt}&=&{\pi^2\alpha^2_{EM}\over s_W^2}\nonumber \\ c_q^{ts}(s)&=&{4\pi^2\alpha^2_{EM}\over s_W^2} {1\over s} \biggl(Q_q+{s\over s-M_Z^2} {1\over s_W^2}(T_{3,q}-Q_qs_W^2)\biggr)\nonumber \\ c_q^{ss}(s)&=&{16\pi^2\alpha^2_{EM}\over s^2}\biggl\{ \biggl(Q_q+{1\over 2 s_W^2}(T_{3,q}-2Q_qs_W^2){s\over s-M_Z^2}\biggr)^2 +\biggl({T_{3,q}\over 2 s_W^2}{s\over s-M_Z^2}\biggr)^2\biggr\}\end{aligned}$$ with $T_{3,q}=\pm {1\over 2}$ and $s_W=\sin\theta_W$. The functions occurring in the lowest order amplitudes are, $$\begin{aligned} F_u^{0}(s,t)&=&F_d^{0}(s,u) \nonumber \\ &=&16\biggl( {ut\over M_W^4}-1 \biggr) \biggl({1\over 4}+{M_W^4\over t^2} \biggr)+16{s\over M_W^2} \nonumber \\ J_u^{0}(s,t)&=&-J_d^{0}(s,u) \nonumber \\ &=& 16\biggl( {ut\over M_W^4}-1\biggr)\biggl({s\over 4}-{M_W^2\over 2}-{M_W^4\over t}\biggr) +16s\biggl({s\over M_W^2}-2+{2M_W^2\over t}\biggr) \nonumber \\ K_u^{0}(s,t)&=& K_d^{0}(s,u)\nonumber \\ &=& 8\biggl({ut\over M_W^4}-1\biggr) \biggl({s^2\over 4}-sM_W^2+3M_W^4\biggr) +8s^2\biggl({s\over M_W^2}-4\biggr) \, .\end{aligned}$$ NLO Results {#appa:NLO} ----------- The functions occurring in the one-loop virtual amplitude are[@Frixione:1993yp], $$\begin{aligned} F_u^{1}(s,t)&=&{4(80t^2+73st-140M_W^2t+72M_W^4) \over t^2} -{4(4t+s)^2\over s\beta^2 t} -{128(t+2s)\over M_W^2}\nonumber \\&& +{64(t+s)\over M_W^4} -\biggl({32(t^2-3st-3M_W^4)\over t^2}+{128s\over t-M_W^2}\biggr)\log\biggl({-t\over M_W^2}\biggr) \nonumber \\ &&+\biggl({8(6t^2+8st-19M_W^2t+12M_W^4)\over t^2}-{32t^2-128st-26s^2\over s\beta^2t} +{6(4t+s)^2\over s\beta^4t}\biggr)\log\biggl({s\over M_W^2}\biggr) \nonumber \\ &&+32s\biggl({2M_W^4\over t}-u\biggr)I_4 -64 (t-M_W^2)\biggl( {2M_W^4\over t^2}-{u\over t}\biggr)I_{3t} \nonumber \\ &&+\biggl( {16t(4M_W^2-u)-49s^2+72M_W^2s-48M_W^4\over 2t} +{2(8t^2-14st-3s^2)\over \beta^2t} -{3(4t+s)^2\over 2\beta^4t}\biggr)I_{3l} \nonumber \\&& +{32\pi^2\over 3}\biggl({2(t+2s)\over M_W^2}-{3t+2s-4M_W^2\over t} -{t(t+s)\over M_W^4}\biggr) \nonumber \\ J_u^{1}(s,t) &=& -{128(t^2+2st+2s^2)\over M_W^2} -{16(t^2-21st-26M_W^2t+34M_W^2s+17M_W^4)\over t}\nonumber\\ && +{64st(t+s)\over M_W^4}+{32s^2\over t-M_W^2} \nonumber \\ && +\biggl(16(t-5s+2M_W^2)-{48M_W^2(2s+M_W^2)\over t} +{64s(2t+s)\over t-M_W^2} -{32s^2t\over (t-M_W^2)^2}\biggr)\log\biggl({-t\over M_W^2}\biggr) \nonumber \\ && + \biggl( {16(4t+s)\over\beta^2} -16(3t-2s)+{48M_W^2(2t-2s-M_W^2)\over t}\biggr) \log\biggl({s\over M_W^2}\biggr) \nonumber \\ &&+16s\biggl(t(2s+u)-2M_W^2(2s+M_W^2)\biggr)I_4 +32(t-M_W^2)\biggl({2M_W^2(2s+M_W^2)\over t}-2s-u\biggr)I_{3t} \nonumber \\ && +\biggl(32st-12s^2+32M_W^4-16M_W^2(2t+7s)-{4s(4t+s)\over\beta^2}\biggr)I_{3l} \nonumber \\ &&+{32\pi^2\over 3}\biggl({2(t^2+2st+2s^2)\over M_W^2} -{st(t+s)\over M_W^4}-{2M_W^2(2t-2s-M_W^2)\over t} -t-4s\biggr) \nonumber \\ K_u^{1}(s,t)&=& 16\biggl\{12t^2+20st-24M_W^2t+17s^2-4M_W^2s+12M_W^4+ {s^2t(t+s)\over M_W^4} \nonumber \\ && -{2s(2t^2+3st+2s^2)\over M_W^2}\biggr\}(2-{\pi^2\over 3})\end{aligned}$$ with $F^1_d(s,t)=F^1_u(s,u)$, $J^1_d(s,t)=-J^2_u(s,u)$, and $K^1_d(s,t)=K^1_u(s,u)$. The integrals are given by $$\begin{aligned} I_4&=&{1\over st}\biggl(2\log^2\biggl({-t\over M_W^2}\biggr)-4\log\biggl({M_W^2-t\over M_W^2}\biggr)\log\biggl({-t\over M_W^2}\biggr) -4{\rm Li}_2\biggl({t\over M_W^2}\biggr)\biggr) \nonumber \\ I_{3t}&=&{1\over M_W^2-t}\biggl({1\over 2}\log^2\biggl({M_W^2\over s}\biggr)-{1\over 2}\log^2\biggl({-t\over s}\biggr) -{\pi^2\over 2}\biggr) \nonumber \\ I_{3l}&=&{1\over s\beta} \biggl(4{\rm Li}_2\biggl( {\beta-1\over 1+\beta}\biggr) +\log^2\biggl({1-\beta\over 1+\beta}\biggr) +{\pi^2\over 3}\biggr)\, .\end{aligned}$$ Approximate NNLO Results {#appb} ======================== The hard scattering kernel is expanded in a power series: $$\begin{aligned} C(z,M,\cos\theta,\mu)&=&C^{0}(z,M,\cos\theta,\mu)+\frac{\alpha_s}{4\pi}C^{1}(z,M,\cos\theta,\mu) \nonumber \\ && +\left(\frac{\alpha_s}{4\pi}\right)^2C^{2}(z,M,\cos\theta,\mu). \, .\end{aligned}$$ Similarly, the hard function is expanded in a power series: $$\begin{aligned} H(M_{WW},\cos\theta,\mu_f)&=&H^{0}(M_{WW},\cos\theta) +\frac{\alpha_s}{4\pi}H^{1}(M_{WW},\cos\theta,\mu_f)\nonumber\\&&+\left(\frac{\alpha_s}{4\pi}\right)^2H^{2}(M_{WW},\cos\theta,\mu_f)\label{NNLOexact},\end{aligned}$$ where $H^1=H^1_{reg}+H^1_{extra}$ where $H^1_{reg}$ and $H^1_{extra}$ are defined in Eq. (\[H1\]). The approximate NNLO cross section is found by calculating the scale dependent pieces of the leading singular contribution to $C^{2}$ and adding this contribution to the total NLO cross section.Using the results for the hard and soft functions to NNLO, an approximate formula for the NNLO piece, $C^{2}$, can be determined which includes the leading singular pieces. The result is written as an expansion of $C^{2}$ in “plus"-functions: $$\begin{aligned} C^{2}(z,M,\cos\theta,\mu_f)=\sum^3_{n=0}D^{(n)}\left[\frac{\ln^n(1-z)}{1-z}\right]_+ + R^{(0)}\delta(1-z),\end{aligned}$$ where $$\begin{aligned} D^{(3)}&=&64 H^{0} s^{(2,4)}\\ D^{(2)}&=&24 H^{0} \left[s^{(2,3)}+4 L_s s^{(2,4)}\right]\\ D^{(1)}&= &8 H^{0}\left[s^{(2,2)}+3 L_s s^{(2,3)}+6\left(L^2_s-\frac{2\pi^2}{3}\right)s^{(2,4)}\right]+8 H^{(1)} s^{(1,2)} \\ D^{(0)}&=&2H^{0}\left[s^{(2,1)}+2L_s s^{(2,2)}+3\left(L^2_s-\frac{2\pi^2}{3}\right)s^{(2,3)}+4\left(L^3_s-2L_s\pi^2+16\zeta_3\right)s^{(2,4)}\right]\\ &&+2 H^{1}\left[s^{(1,1)}+2L_s s^{(1,2)}\right]\nonumber\end{aligned}$$ and $$\begin{aligned} R^{(0)}&=& H^{0}\left[ s^{(2,0)}+ L_M s^{(2,1)}+\left(L^2_M-\frac{2\pi^2}{3}\right)s^{(2,2)} \right. \nonumber \\ && \left. +\left(L^3_M-2L_M\pi^2+16\zeta_3\right)s^{(2,3)}\right.\\ &&\left.+\left(L^4_M-4L^2_M\pi^2+\frac{4\pi^4}{15}+64L_M\zeta_3\right)s^{(2,4)}\right] \nonumber \\ &&+H^{1}\left[s^{(1,0)}+L_Ms^{(1,1)}+\left(L^2_M-\frac{2\pi^2}{3}\right)s^{(1,2)}\right]+H^{2}\, .\end{aligned}$$ The logarithms are defined as $$\begin{aligned} L_M&=&\log\biggl({M^2\over \mu_f^2}\biggr)\nonumber \\ L_s&=&\log\biggl({s\over \mu_f^2}\biggr)\end{aligned}$$ The soft contributions are found from the RG evolution and explicit calculation of the soft function [@Becher:2007ty], $$\begin{aligned} s^{(1,0)}&=&\frac{C_F\pi^2}{3} \nonumber \\ s^{(2,0)}&=&C_F\left[C_F\frac{\pi^4}{18}+C_A\left(\frac{2428}{81}+\frac{67\pi^2}{54}-\frac{\pi^4}{3}-\frac{22}{9}\zeta_3\right)-T_Fn_f\left(\frac{656}{81}+\frac{10\pi^2}{27}-\frac{8}{9}\zeta_3\right)\right] \nonumber \\ s^{(1,2)}&=&\frac{\Gamma_0}{2} \nonumber \\ s^{(1,1)} &=& \gamma^s_0 \nonumber \\ s^{(2,4)}&=&\frac{\Gamma^2_0}{8} \nonumber \\ s^{(2,3)}&=&\frac{\Gamma_0}{6}(3\gamma^s_0-\beta_0) \nonumber \\ s^{(2,2)}&=&\frac{1}{2}\left(\Gamma_0 s^{(1,0)}+\Gamma_1 +(\gamma^s_0)^2-\beta_0\gamma^s_0\right) \nonumber \\ s^{(2,1)}&=&s^{(1,0)}(\gamma^s_0-\beta_0)+\gamma^s_1,\end{aligned}$$ where $C_F=4/3$, $C_A=3$, $T_F=1/2$, $n_f=5$, and $\zeta_3$ is a Riemann zeta function. Expressions for $\Gamma_0,\Gamma_1,\gamma_0^s, \gamma_1^s$ and $\beta_0$ can be found in Ref. [@Becher:2007ty] (where the soft anomalous dimension is written as $\gamma^W$ instead of $\gamma^s$). Similarly, the hard coefficients can be expanded as a power series in logs, $$\begin{aligned} H^{0}(M_{WW},\cos\theta)&=&h^{(0,0)}(M_{WW},\cos\theta) \\ H^{1}(M_{WW},\cos\theta,\mu_f)&=&\sum^2_{n=0}h^{(1,n)}\left(M_{WW},\cos\theta,\frac{M_{WW}}{Q_h}\right)L^n_Q \nonumber \\ H^{2}(M_{WW},\cos\theta,\mu_f)&=&\sum^4_{n=0}h^{(2,n)}\left(M_{WW},\cos\theta,\frac{M_{WW}}{Q_h}\right)L^n_Q\nonumber\, ,\end{aligned}$$ and $$L\equiv \ln\biggl({Q_h^2\over \mu_f^2}\biggr) \, .$$ We have introduced an additional arbitrary scale $Q_h$. Using the RGEs of the hard function, we can solve for the hard coefficients: $$\begin{aligned} h^{(1,2)}&=&-\frac{\Gamma_0}{2}h^{(0,0)} \nonumber \\ h^{(1,1)}&=&-\left(\gamma^V_0+\Gamma_0\ln\frac{M^2_{WW}}{Q^2_h}\right)h^{(0,0)} \nonumber \\ h^{(2,4)}&=&~~\frac{\Gamma^2_0}{8}h^{(0,0)} \nonumber \\ h^{(2,3)}&=&\frac{\Gamma_0}{6}\left[3\gamma^V_0+\beta_0+3\Gamma_0\ln\frac{M^2_{WW}}{Q^2_h}\right]h^{(0,0)} \nonumber \\ h^{(2,2)}&=&\frac{1}{2}\left[-\Gamma_0 h^{(1,0)}-\Gamma_1 h^{(0,0)}+\left(\gamma^V_0 +\Gamma_0\ln\frac{M^2_{WW}}{Q^2_h}\right)\left(\gamma^V_0+\beta_0+\Gamma_0\ln\frac{M^2_{WW}}{Q^2_h}\right)h^{(0,0)}\right] \nonumber \\ h^{(2,1)}&=&-\left(\gamma^V_0+\beta_0+\Gamma_0\ln\frac{M^2_{WW}}{Q^2_h}\right)h^{(1,0)}-\left(\gamma^V_1 +\Gamma_1\ln\frac{M^2_{WW}}{Q^2_h}\right)h^{(0,0)},\end{aligned}$$ where the arguments of the hard coefficients have been suppressed. The anomalous dimension of the hard Wilson coefficient $C_V$, $\gamma^V$, can be found in Ref. [@Becher:2007ty]. The coefficients $h^{(0,0)}$ and $h^{(1,0)}$ can be calculated from the known LO and NLO hard functions given in Eqs. (\[H0\]) and (\[H1\]). Additionally, since an additional arbitrary scale $Q_h$ was introduced, the $Q_h$ dependence of $h^{(1,0)}$ and $h^{(2,0)}$ can be solved for: $$\begin{aligned} h^{(1,0)}&=&\sum^2_{n=0}h^{(1,n)}_{Q_h=M_{WW}}\ln^n\frac{M^2_{WW}}{Q^2_h}\\ h^{(2,0)}&=&\sum^4_{n=0}h^{(2,n)}_{Q_h=M_{WW}}\ln^n\frac{M^2_{WW}}{Q^2_h}\nonumber\, ,\end{aligned}$$ where the subscript $Q_h=M_{WW}$ indicates the value of $Q_h$ at which the coefficients on the RHS are evaluated at. Using these coefficients, the NNLO result in Eq. (\[NNLOexact\]) is independent of the scale $Q_h$. However, without a full calculation, it is not possible to know $h^{(2,0)}$. Since the other NNLO coefficients, $h^{(2,n)}$ for $n=1,2,3$, are independent of $h^{(2,0)}$, then $h^{(2,0)}$ can be set to zero and an approximate NNLO result is obtained. The purpose of introducing $Q_h$ is now clear, as discussed in the Section \[approxnnlo\]. [^1]: The theoretical predictions have been evaluated at NLO using MCFM with MSTW2008nlo PDFs and a central scale choice of $\mu_f=M_W$. The uncertainties shown in Eq. (\[theorypred\]) result from varying the scale up and down by a factor of $2$. The predictions of Eq. (\[theorypred\]) include the next-to-next-to-leading order (NNLO) contribution from the $gg$ initial state [@Campbell:2011bn]. [^2]: Since we are interested in a color singlet final state, the soft function $S$ has no $\cos\theta$ dependence. [^3]: In Eqs. (\[wilscoeff\]) and (\[zren\]), the sum over Dirac structures is implied. See Ref. [@Ahrens:2011mw] for an example of the relevant notation.
{ "pile_set_name": "ArXiv" }
--- abstract: | We classify certain integrable (both classical and quantum) generalisations of Dirac magnetic monopole on topological sphere $S^2$ with constant magnetic field, completing the previous local results by Ferapontov, Sayles and Veselov. We show that there are two integrable families of such generalisations with integrals, which are quadratic in momenta. The first family corresponds to the classical Clebsch systems, which can be interpreted as Dirac magnetic monopole in harmonic electric field. The second family is new and can be written in terms of elliptic functions on sphere $S^2$ with very special metrics. address: - 'Department of Mathematical Sciences, Loughborough University, Loughborough LE11 3TU, UK; Moscow State University and Steklov Mathematical Institute, Moscow, Russia' - 'Department of Mathematical Sciences, Loughborough University, Loughborough LE11 3TU, UK' author: - 'A.P. Veselov' - 'Y. Ye' title: Integrable generalisations of Dirac magnetic monopole --- Introduction ============ The history of quantum integrable systems with magnetic fields goes back to the pioneering work in the 1930s by Dirac [@Dirac] on the celebrated magnetic monopole and by Landau [@Landau], who considered the case of constant magnetic field on the plane (Landau problem). Since then this area was of a substantial interest of the mathematical and theoretical physicists (see e.g. [@Fer-Ves; @KV; @Mlad; @NP; @Prieto; @Sol; @WY]). In spite of this, the general problem of quantum integrability in two dimensions in the presence of a magnetic field is still far from complete solution. Some important results in this direction have been found, in particular, by Winternitz and his collaborators in [@ber-wint; @DGRW; @MW]. Ferapontov and Fordy [@FF] derived the classical integrability conditions in the case, when the integral is quadratic in momenta. Ferapontov, Sayles and Veselov [@FSV] considered the quantum case and showed that the conditions of quantum integrability are different from classical case (see the details in the next section). However, remarkably they coincide in the case when the density of the magnetic field $B$ is constant. This case was studied in [@FSV], where a local classification of such systems under some additional assumptions was found. The final list consists of two families (see the next section). The first one contains the Dirac magnetic monopoles in the external harmonic field (and their hyperbolic versions), which are known to be equivalent to the classical Clebsch integrable cases of the free rigid body in infinite ideal fluid [@Clebsch]. The second family is more mysterious and is the main object of our study. We show that under certain assumptions on the parameters the corresponding systems can be extended to the smooth systems on the topological sphere $\mathbb S^2$, which can be described in terms of elliptic functions. More precisely, we represent the sphere $\mathbb S^2$ as the quotient of a real torus $\mathbb T$ by the involution $\sigma: u\to -u,\, u \in \mathbb T.$ Consider the elliptic function $\mathcal Q(z)$ defined as the inversion $w=\mathcal Q(z)$ of the elliptic integral $$z=\int_{\beta_2}^{w}\frac{2d\xi}{\sqrt{P(\xi)}},$$ where $$P(x)=a_3(x-\beta_1)(x-\beta_2)(x-\beta_3)(x-\beta_4)$$ is a polynomial with $a_3<0$ and 4 real roots: $\beta_1>\beta_2>0>\beta_3>\beta_4,$ such that $$\beta_1+\beta_2+\beta_3+\beta_4=0, \, \beta_1+\beta_4<0,\,\, \beta_2+\beta_3>0.$$ The elliptic function $\mathcal Q(z)$ is even and has two periods: real $2K_1$ and pure imaginary $2iK_2$, where $$K_1=\int_{\beta_2}^{\beta_1}\frac{2d\xi}{\sqrt{P(\xi)}}, \,\,\, K_2=\int_{\beta_2}^{\beta_3}\frac{2d\xi}{\sqrt{-P(\xi)}}.$$ It satisfies the differential equation $ 4\mathcal Q'^2=P(\mathcal Q) $ and can be expressed via the standard Weierstrass elliptic function $\wp(z)$. In the limiting case when $\beta_1+\beta_4=\beta_2+\beta_3=0$ (so $P(x)$ is even), $\mathcal Q$ can be written in terms of the Jacobi’s elliptic $sn$-function [@WW] as $$\mathcal Q=\beta_2 \, \textit{sn}(\alpha (z-\beta_2);k),\,\, \alpha=\sqrt{a_3}\beta_1/2, \,\, k=\beta_2/\beta_1.$$ Introduce two real-valued functions $$Q_1(u_1):=\mathcal Q(u_1),\,\,\,\,\, Q_2(u_2):=\mathcal Q(iu_2)$$ with periods $2K_1$ and $2K_2$ respectively, and consider the torus $$\mathbb T^2=\mathbb R^2(u_1,u_2)/4K_1 \mathbb Z\oplus 4K_2 \mathbb Z.$$ On this torus the corresponding classical Hamiltonian $H$ and integral $F$ have the following explicit form with $Q_1=Q_1(u_1)$ and $Q_2=Q_2(u_2)$ $$\label{H1} H=\frac{1}{Q_1^2-Q_2^2}\left[(p_1-A_1)^2+(p_2-A_2)^2\right]+\frac{\mu}{Q_1+Q_2},$$ $$\label{F1} F=\frac{1}{Q_1^2-Q_2^2}\left[Q_2^2(p_1-A_1)^2+Q_1^2(p_2-A_2)^2\right]$$ $$+\frac{2kQ_2'}{Q_1-Q_2}(p_1-A_1)+\frac{2kQ_1'}{Q_2-Q_1}(p_2-A_2)-\frac{\mu Q_1Q_2}{Q_1+Q_2}-kB(Q_1+Q_2)^2,$$ where $B$ is the density of magnetic field assumed to be constant and $k=-4B/a_3.$ The magnetic potential $A=A_1du_1+A_2du_2$ is determined by the relation $$\label{A1} dA=B(Q_1^2(u_1)-Q_2^2(u_2))du_1\wedge du_2.$$ There is a problem with these systems on the torus, because $Q_1-Q_2=0$ at the half-periods of the torus, which creates singularities in the formulas. However, we show that on the quotient $\mathbb S^2=\mathbb T^2/\sigma$ of the torus by involution $\sigma$ having exactly these points fixed, this problem disappears and we have regular smooth systems on $\mathbb S^2.$ In the limiting even case we do have two singularities in the potential $h$, but the metric becomes the standard metric on the round sphere, so we have the new integrable electric perturbation of Dirac magnetic monopole (and new integrable two-centre problem) on the standard sphere (see [@VY]). The plan of the paper is following. In the next two sections we describe the classical and quantum integrability conditions in 2D in the presence of magnetic field and prove the local classification result in the case of non-zero constant magnetic field, mainly following unpublished work of Ferapontov, Sayles and Veselov [@FSV]. Then we show that under certain condition on the parameters these systems can be extended to the regular analytic integrable systems on the topological sphere $S^2$ with some very special metrics. Integrable magnetic fields in 2D: local classification ====================================================== In two dimensions it is always possible to reduce both Hamiltonian $H$ and integral $F$ to a diagonal form: $$\begin{aligned} H&=&g^{11} \left(p_1-A_1\right)^2+ g^{22} \left(p_2-A_2\right)^2+h,\nonumber\\[-4mm] \label{classic_hams}\\[-2mm] F&=& g^{11} v^1\left(p_1-A_1\right)^2+ g^{22}v^2 \left(p_2-A_2\right)^2 +\phi^1 \left(p_1-A_1\right) +\phi^2\left(p_2-A_2\right) + \varphi, \nonumber\end{aligned}$$ in which metric $g^{ii}$ and all the other coefficients $v^i$, $A_i$, $\phi^i$, $h$, $\varphi$ are functions depending on the coordinates $(q^1,q^2)$. Ferapontov and Fordy [@FF] showed that Poisson commutativity of $H$ and $F$ is equivalent to the following integrability conditions $$\begin{aligned} && (C1) \ \ \ \partial_i v^i = 0, \ \ \ \ \,\, i=1,2, \nonumber\\[3mm] && (C2) \ \ \ \partial_j v^i= \left(v^j-v^i\right) \partial_j \ln(g^{ii}) \ \ \ \ \mbox{for all} \ i \neq j, \nonumber\\[3mm] && (C3) \ \ \ \partial_i \phi^i = \frac{1}{2 g^{ii}} \left( \phi^1 \partial_1 g^{ii} + \phi^2 \partial_2 g^{ii}\right), \,\, \ i=1,2, \qquad\qquad\qquad \qquad\qquad\quad \ \; \nonumber\\[4mm] && (C4) \ \ \ 2 \sqrt{g^{11} g^{22}}(v^2-v^1) B = g^{22} \partial_2 \phi^1 +g^{11} \partial_1 \phi^2 , \label{full_prob_with_B}\\[4mm] && (C5) \ \ \ \partial_1 \varphi - v^1 \partial_1 h - \frac{\phi^2}{\sqrt{g^{11} g^{22}}} B =0, \quad \partial_2 \varphi- v^2 \partial_2 h + \frac{\phi^1}{\sqrt{g^{11} g^{22}}} B =0, \nonumber\\[2mm] && (C6) \ \ \ \phi^1 \partial_1 h + \phi^2 \partial_2 h=0, \nonumber\end{aligned}$$ where $$\label{mag} B:=\sqrt{g^{11} g^{22}}(\partial_1 A_2 -\partial_2 A_1)$$ is the magnetic field density. Consider now the following quantum analogue of the Hamiltonian and the integral: $$\label{q_ops} \hat{H} = \sqrt{g^{11}g^{22}}\nabla_1\frac{g^{11}}{\sqrt{g^{11}g^{22}}}\nabla_1 + \sqrt{g^{11}g^{22}}\nabla_2\frac{g^{22}}{\sqrt{g^{11}g^{22}}}\nabla_2 + h,$$ $$\hat{F}= v^1\sqrt{g^{11}g^{22}}\nabla_1\frac{g^{11}}{\sqrt{g^{11}g^{22}}}\nabla_1 + v^2\sqrt{g^{11}g^{22}}\nabla_2\frac{g^{22}}{\sqrt{g^{11}g^{22}}}\nabla_2 +\phi^1 \nabla_1 +\phi^2\nabla_2 + \varphi,$$ where $\nabla_j=i\partial_j-A_j, \, j=1,2$. Ferapontov, Sayles and Veselov [@FSV] derived the necessary and sufficient conditions for commutativity $[\hat H, \hat F]=0$ and showed that the first conditions (C1)-(C5) are the same, but the last condition (C6) in quantum case is replaced by $$\label{qcond} (C6)^*\ \ \ \phi^1 \partial_1 h + \phi^2 \partial_2 h+ \sqrt{g^{11} g^{22}} \left(v^2-v^1\right)\left( \frac{\partial_2 g^{11}}{g^{11}} \partial_1 B +\frac{\partial_1 g^{22}}{g^{22}}\partial_2 B-\partial_1 \partial_2 B \right) =0.$$ In particular, we see that if the magnetic density $B$ is constant then the extra term $$\frac{\partial_2 g^{11}}{g^{11}} \partial_1 B +\frac{\partial_1 g^{22}}{g^{22}}\partial_2 B-\partial_1 \partial_2 B =0$$ vanishes and the quantum and classical integrability conditions coincide. Local classification of all such systems (under some additional assumptions) was done by Ferapontov, Sayles and Veselov [@FSV], who proved in the quantum case the following Suppose that the quantum system with the Hamiltonian $\hat{H}$ of the form (\[q\_ops\]) has magnetic field with a constant non-zero density $B$, a non-constant electric potential $h$ and assume that the system has no integrals, which are linear in momenta. Then the system has a second order integral $\hat{F}$ if and only if it can be locally reduced to one of the forms specified below, where in each case the metric is of Stäckel form $$\label{metric} ds^2=\frac{q^1-q^2}{f(q^1)}(dq^1)^2+\frac{q^2-q^1}{f(q^2)}(dq^2)^2$$ with $$\begin{aligned} \label{I} \mbox{\textit{\textrm{(I)}}} && f(q)=a_3 q^3+a_2 q^2+a_1 q+a_0,\,\,\,\, h=\mu(q^1+q^2);\\ \label{II} \mbox{\textit{\textrm{(II)}}} && f(q)=a_3 q^3+a_2 q^2+a_1q+a_0 q^{\frac32},\,\,\, h=\frac{\mu}{\sqrt{q^1}+\sqrt{q^2}}\end{aligned}$$ depending on real parameters $a_0, a_1, a_2, a_3\neq 0$ and $\mu.$ The Gaussian curvature of the metrics respectively is $$\label{GK} {\textit{\textrm{(I)}}} \,\,\, K=-\frac{a_3}{4} \quad {\textrm{and \,\,\, (II)}} \,\,\, K=-\frac{a_3}{4}+\frac{a_0}{( \sqrt{q^1}+\sqrt{q^2})^3}.$$ The corresponding quantum integral $\hat{F}$ can be chosen in the form (\[q\_ops\]) with $v^1=q^2, \, v^2=q^1$ and $$\begin{aligned} \mbox{\textrm{(I)}} && \phi^1=k\sqrt{-\frac{f(q^1) f(q^2)}{(q^1-q^2)^2}}\, , \ \ \ \ \phi^2=-k \sqrt{-\frac{f(q^1) f(q^2)}{(q^1-q^2)^2}}\, , \nonumber\\[3mm] && \varphi=\mu q^1 q^2 -k B (q^1+q^2);\\ \mbox{\textrm{(II)}} && \phi^1=k \frac{ \sqrt{-f(q^1) f(q^2)}}{\sqrt{q^1 q^2}-q^2}\, , \ \ \ \ \phi^2=k \frac{ \sqrt{-f(q^1) f(q^2)}}{\sqrt{q^1 q^2}-q^1}\, , \nonumber\\[2mm] && \varphi= -\frac{\mu\sqrt{q^1 q^2}}{\sqrt{q^1}+\sqrt{q^2}}-kB\left( \sqrt{q^1}+\sqrt{q^2} \right)^2,\end{aligned}$$ where $k=-4B/a_3.$ The proof is rather lengthy and technical. We present it now with all the details, mainly following the unpublished work [@FSV]. Proof of the local classification ================================= Since the classical and quantum integrability conditions coincide in our case, we will consider for simplicity the classical case, assuming that the Hamiltonian $H$ and integral $F$ are reduced to the diagonal form $$H=g^{11} \left(p_1-A_1\right)^2+ g^{22} \left(p_2-A_2\right)^2+h,$$ $$F= g^{11} v^1\left(p_1-A_1\right)^2+ g^{22}v^2 \left(p_2-A_2\right)^2 +\phi^1 \left(p_1-A_1\right) +\phi^2\left(p_2-A_2\right) + \varphi,$$ where all the coefficients are functions of the local coordinates $(q^1,q^2)$. We assume also that the magnetic density $$B=\sqrt{g^{11} g^{22}} \left( \partial_1 A_2 -\partial_2 A_1\right), \label{mag}$$ is a non-zero constant. As we have seen in that case the classical and quantum integrability conditions coincide. Without loss of generality locally we can take $v^1=q^2$, $v^2=q^1$. By integrability condition (C2), we must have metric of Stäckel form $$\label{staekel} ds^2=\frac{q^1-q^2}{f_1(q^1)}(dq^1)^2+\frac{q^2-q^1}{f_2(q^2)}(dq^2)^2.$$ In order to make the metric positive definite, we require that $f_1(q^1)$ and $f_2(q^2)$ have different sign. Now, we use condition (C5) $$\partial_1 \varphi - v^1 \partial_1 h - \frac{\phi^2}{\sqrt{g^{11} g^{22}}} B =0,$$ $$\partial_2 \varphi - v^2 \partial_2 h + \frac{\phi^1}{\sqrt{g^{11} g^{22}}} B =0.$$ The consistency condition gives $$\phi^1 \partial_1 B + \phi^2 \partial_2 B+ \sqrt{g^{11} g^{22}} \left(v^2-v^1\right)\left( \frac{\partial_2 g^{11}}{g^{11}} \partial_1 h +\frac{\partial_1 g^{22}}{g^{22}}\partial_2 h-\partial_1 \partial_2 h \right) =0. \label{consistency}$$ *Note that this condition coincides with the quantum integrability condition $(C6)^*$ given by (\[qcond\]) with the roles of $h$ and $B$ interchanged. Thus we see an interesting [*duality*]{} between the potential $h$ and the magnetic field density $B$, which holds [*only in the quantum case*]{}.* It is interesting that the self-duality conditions $B=\pm h$ appear as the factorisability condition for the Hamiltonian in the work by Ferapontov and Veselov [@Fer-Ves]. Since we assumed that $B$ is constant, this relation reduces to $$\left(q^1-q^2\right) \partial_1 \partial_2 h-\partial_1 h+\partial_2 h=0,$$ which can be simplified to $$\partial_1 \partial_2 \left[ \left(q^1-q^2 \right) h \right]=0.\label{h-pde}$$\ Now assume that $h$ is not a constant. Solving (\[h-pde\]) we get $$h=\frac{a(q^1)-b(q^2)}{q^1-q^2}, \label{h-solution}$$ where $a$ and $b$ are arbitrary functions, and $$\phi^j=-\frac{\partial_i h}{\partial_j h} \phi^i,\ \ \ \ \ \ i\neq j,\ \ \ \ \ i,j=1,2.\label{phi-relation}$$ From condition (C3): $$\partial_1 \phi^1 = \frac{1}{2 g^{11}} \left( \phi^1 \partial_1 g^{11} + \phi^2 \partial_2 g^{11}\right), \ \ \ \ \ \partial_2 \phi^2 = \frac{1}{2 g^{22}} \left( \phi^1 \partial_1 g^{22} + \phi^2 \partial_2 g^{22}\right),$$ after rearranging terms and using relation (\[phi-relation\]), we have $$\phi^1=\exp \left[ \int\frac{1}{2 g^{11}}\left( \partial_1g^{11}-\frac{\partial_1 h \ \partial_2 g^{11}}{\partial_2 h} \right) dq^1 \right],$$ $$\phi^2=\exp \left[ \int\frac{1}{2 g^{22}}\left( \partial_2g^{22}-\frac{\partial_2 h \ \partial_1 g^{22}}{\partial_1 h} \right) dq^2 \right].$$ Using the Stäckel form of the metric (\[staekel\]) we deduce that $$\label{ph1} \phi^1=\beta(q^2) \sqrt{ \frac{f_1(q^1)}{a(q^1)-b(q^2)-(q^1-q^2) b^\prime(q^2) }}$$ $$\label{ph2} \phi^2=\alpha(q^1) \sqrt{ \frac{f_2(q^2)}{b(q^2)-a(q^1)+(q^1-q^2) a^\prime(q^1) }}$$ for some arbitrary functions $\alpha$ and $\beta$. Substituting this back to condition (C6), we have $$- \frac{\alpha(q^1)}{\beta(q^2)}= \sqrt{\frac{f_1(q^1) \left( -a(q^1)+b(q^2)+(q^1-q^2) a^\prime(q^1) \right)^3}{f_2(q^2) \left( a(q^1)-b(q^2)-(q^1-q^2) b^\prime(q^2) \right)^3}},$$ which after taking logarithm and differentiating by $q^1$ and $q^2$, gives $$\label{ab} a^{\prime \prime}(q^1)(a(q^1)-b(q^2)-(q^1-q^2) b^\prime(q^2))^3 = b^{\prime \prime}(q^2) (b(q^2)-a(q^1)+(q^1-q^2) a^\prime(q^1))^3,$$ and thus $$- \frac{\alpha(q^1)}{\beta(q^2)}= \sqrt{\frac{f_1(q^1) a^{\prime\prime}(q^1)}{f_2(q^2) b^{\prime\prime}(q^2)}}.$$ Rearranging terms and separating $q^1$ and $q^2$, we arrive at the final relation for $\alpha$ and $\beta$ $$-\frac{ \alpha(q^1)}{\sqrt{f_1(q^1) a^{\prime\prime}(q^1)}}=\frac{\beta(q^2)}{\sqrt{f_2(q^2) b^{\prime\prime}(q^2)}}=constant.\label{alpha-beta relation}$$ Substituting $q^1=q^2=q$, we have $$\begin{aligned} \big(a^{\prime \prime}(q)+b^{\prime \prime}(q) \big) \big( a(q)-b(q) \big)^3 = 0.\end{aligned}$$ Hence we have the following two cases: 1. $a(q)= b(q)$ 2. $a^{\prime \prime}(q)=-b^{\prime \prime}(q)$ Case A: $a(q)= b(q)$ Denote $f(q):=a(q)= b(q)$ and substitute this into equation (\[ab\]) to have $$f^{\prime \prime}(q^1) \left(\frac{ f(q^1)-f(q^2)}{q^1-q^2}-f^\prime(q^2) \right)^3 = f^{\prime \prime}(q^2) \left( - \frac{f(q^1)-f(q^2)}{q^1-q^2}+ f^\prime(q^1) \right)^3.$$ Then we fix $q^2$ and assume $q^1$ is near to $q^2$. Using Taylor expansion up to $5^{th}$ order derivatives of $f$, we have $$\begin{aligned} \frac{ f(q^1)-f(q^2)}{q^1-q^2}&=&f^\prime(q^2) +\frac12 f^{\prime \prime}(q^2) (q^1-q^2)+\frac16 f^{\prime \prime \prime}(q^2) (q^1-q^2)^2 \nonumber \\ &&+\frac{1}{24} f^{(4)}(q^2) (q^1-q^2)^3+\frac{1}{120} f^{(5)}(q^2) (q^1-q^2)^4+ \cdots , \nonumber\\[5mm] f^\prime(q^1)&=&f^\prime(q^2) + f^{\prime \prime}(q^2) (q^1-q^2)+\frac12 f^{\prime \prime \prime}(q^2) (q^1-q^2)^2 \nonumber \\ &&+\frac{1}{6} f^{(4)}(q^2) (q^1-q^2)^3 +\frac{1}{24} f^{(5)}(q^2) (q^1-q^2)^4+ \cdots , \nonumber\\[5mm] f^{\prime\prime}(q^1)&=& f^{\prime \prime}(q^2)+ f^{\prime \prime \prime}(q^2) (q^1-q^2)+\frac{1}{2} f^{(4)}(q^2) (q^1-q^2)^2 \nonumber\\ &&+\frac{1}{6} f^{(5)}(q^2) (q^1-q^2)^3+ \cdots . \nonumber\end{aligned}$$ After the substitution the first coefficients are cancelled, while the cancellation of $(q^1-q^2)^6$ term gives the following necessary condition for $f$: $$f^{\prime \prime}(q) \Big( 40 f^{\prime \prime \prime}(q)^3-45 f^{\prime \prime}(q) f^{\prime \prime \prime}(q) f^{(4)}(q)+9 f^{\prime \prime}(q)^2 f^{(5)}(q) \Big)=0. \label{f-caseA}$$ First we notice that $f^{\prime \prime}(q)$ can not be zero since this will give to a constant potential $h,$ which contradicts our assumption. This means that $g(q):=f^{\prime \prime}(q)$ satisfies the equation $$\frac{40}{9} (g^\prime)^3-5 g\, g^\prime g^{ \prime \prime}+ g^2 g^{ \prime \prime \prime}=0. \label{g-caseA}$$ Remarkably this happens to be $n=-\frac23$ case of the following solvable equation: $$(n-1) (n-2) (y^\prime)^3 +3(n-1) y\, y^\prime y^{ \prime \prime}+y^2 y^{ \prime \prime \prime} =0 \label{PZ}$$ with the general solution of the form $$[y(x)]^n=c_0+c_1x+c_2 x^2$$ (see Equation 27 in [@PZ], Section 3.5.3.). Thus we have $$f^{\prime \prime}(q)= g(q) = (c_0+c_1 q+c_2 q^2)^{-\frac32}. \label{fprimeprime} \nonumber$$ Integrating this twice, we arrive at the following general formula for $f$: $$f(q)=\frac{4}{4 c_0 c_2 -c_1^2} \sqrt{c_0+c_1 q+c_2 q^2}+C_1 q+C_0, \nonumber$$ where $C_0$ and $C_1$ are constants and we assumed that $4 c_0 c_2 -c_1^2\neq0$. Ignoring the linear term, which only gives a constant shift of the potential, and relabelling the constants we have $$a(q)=b(q)= \sqrt{c_0+c_1 q+c_2 q^2}. \nonumber$$ In the case, when $4 c_0 c_2 -c_1^2=0$, modulo linear terms we have two subcases: $$a(q)=b(q)= c q^2, \nonumber$$ and $$a(q)=b(q)= \frac{c}{q+d}, \nonumber$$ where $c$ and $d$ are some constants. Case B: $a^{\prime \prime}(q)=-b^{\prime \prime}(q)$ In that case we have that $$a(q)=-b(q)+C_1 q+C_0.$$ Denote $f(q):=b(q), \, g(q):=2f(q)-C_1 q-C_0$ then, similarly to the previous case, Taylor expansion in the equation (\[ab\]) leads to the following differential equation for $g$: $$3g^2 g^\prime g^{\prime\prime}+g^3 g^{\prime\prime\prime}=0.\label{g-relation1}$$ Trivial solution $g\equiv0$ leads to the constant potential, so we can divide equation (\[g-relation1\]) by $g^2$ to get $$3g^\prime g^{\prime\prime}+g g^{\prime\prime\prime}=0,$$ which has the general solution $$g(q)= \sqrt{c_0+c_1 q+c_2 q^2}, \nonumber$$ where $c_0$, $c_1$, $c_2$ are arbitrary constants. Hence modulo linear terms $$a(q)=- \frac12 \sqrt{c_0+c_1 q+c_2 q^2}=-b(q).$$ One can check that this case does not lead to any new solutions compared to case A. Thus we have the following three different cases to analyse: 1. $ a(q)=b(q)= \sqrt{c_0+c_1 q+c_2 q^2}$, 2. $ a(q)=b(q)=c q^2$, 3. $ a(q)=b(q)=\displaystyle \frac{c}{q+d}.$ Case (1): $\ a(q)=b(q)=\sqrt{c_0+c_1 q+c_2 q^2}$ Without loss of generality we can reduce this case to 2 subcases 1. $a(q)=b(q)=\mu\sqrt{q}$, 2. $a(q)=b(q)=\sqrt{c_0+c_2 q^2}$. Subcase (i): $\ \ a(q)=b(q)=\mu\sqrt{q}$ We have $$a(q^1)=\mu\sqrt{q^1},\, b(q^2)=\mu\sqrt{q^2},\, h=\frac{\mu\sqrt{q^1}-\mu\sqrt{q^2}}{q^1-q^2}=\frac{\mu}{\sqrt{q^1}+\sqrt{q^2}}.$$ From equation (\[alpha-beta relation\]) $$\alpha (q^1)=-\frac{\tilde{k}}{2} \sqrt{- (q^1)^{-\frac32} f_1(q^1)}\ , \ \ \ \ \beta(q^2)= \frac{\tilde{k}}{2} \sqrt{- (q^2)^{-\frac32} f_2(q^2)}\ ,$$ with some constant $\tilde{k}$. Moreover, from (\[ph1\]) and (\[ph2\]) we have $$\phi^1=k \frac{ \sqrt{-f_1 f_2}}{\sqrt{q^1 q^2}-q^2}, \ \ \ \ \phi^2=k \frac{ \sqrt{-f_1 f_2}}{\sqrt{q^1 q^2}-q^1}, \quad k=\tilde k/\sqrt{\mu}.\nonumber$$ Substituting them into the last unused condition (C4), we have $$\begin{aligned} \frac{4 B}{k} (q^1-q^2) \left( \sqrt{q^1}-\sqrt{q^2} \right)^2 &=&-\left( \sqrt{q^1}-\sqrt{q^2} \right) \left( \frac{f_1^\prime(q^1)}{\sqrt{q^1}}+\frac{f_2^\prime(q^2)}{\sqrt{q^2}} \right) \nonumber\\ & -&\!\!\!\! \left( \sqrt{\frac{q^2}{q^1}}-2 \right) \frac{f_1(q^1)}{q^1}+\left( \sqrt{\frac{q^1}{q^2}}-2 \right) \frac{f_2(q^2)}{q^2}. \nonumber\end{aligned}$$ Changing coordinate $\sqrt{q^1}=x ,\sqrt{q^2}=y$ and $f(z^2)=z^3 F(z)$, we have $$\begin{aligned} \ \ \ \frac{8 B}{k} (x+y)(x-y)^3 =- (x-y) \left( x F^\prime(x)+y F^\prime(y) \right)+(x+y) (F(x)-F(y)). \nonumber\end{aligned}$$ Applying the operator $\frac{\partial^3}{\partial x^2 \partial y}$ to this relation yields $$\frac{d}{dx}\left(x^3 F^{\prime \prime} (x)\right) = -\frac{96 B}{k} x^3.$$ Solving this ODE gives $$F(x)=-\frac{4 B}{k} x^3+a_2 x+\frac{a_1}{x}+a_0,$$ where $a_i$ are constants. Therefore we can derive the functions $f_1$ and $f_2$ in the metric: $$\label{f} f_1(q)=f_2(q)=-\frac{4 B}{k} q^3+a_2 q^2+a_1q+a_0 q^{\frac32}.$$ To find the potential in the integral $F$ we use condition (C5): $$\begin{aligned} \partial_1 \varphi=v^1 \partial_1 h+\frac{\phi^2}{\sqrt{g^{11} g^{22}}} B=q^2 \partial_1 h -k B\left( 1+\sqrt{\frac{q^2}{q^1}} \right), \nonumber\\ \partial_2 \varphi=v^2 \partial_2 h-\frac{\phi^1}{\sqrt{g^{11} g^{22}}} B=q^1\partial_2 h-k B \left( 1+\sqrt{\frac{q^1}{q^2}} \right), \nonumber\end{aligned}$$ which now has solution $$\varphi(q^1,q^2)= -\frac{\mu\sqrt{q^1 q^2}}{\sqrt{q^1}+\sqrt{q^2}}-kB\left( \sqrt{q^1}+\sqrt{q^2} \right)^2.$$ Subcase (ii): $\ \ a(q)=b(q)=\sqrt{c_0+c_2q^2}$ We assume for simplicity that $c_0=c$, $c_2=1$, so $$a(q^1)=\sqrt{c+(q^1)^2}, \, b(q^2)=\sqrt{c+(q^2)^2},\, h=\frac{\sqrt{c+(q^1)^2}-\sqrt{c+(q^2)^2}}{q^1-q^2}.$$ Again from (\[ph1\]) (\[ph2\]) and (\[ab\]) we have $$\begin{aligned} \phi^1=&&k \sqrt{ \frac{-f_1 f_2}{\left(c+(q^2)^2\right) \left( c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \right)}}\ , \nonumber\\ \phi^2=&-&k \sqrt{ \frac{-f_1 f_2}{\left(c+(q^1)^2\right) \left( c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \right)}}\ ,\nonumber\end{aligned}$$ Similarly to the previous subcase we have $$\begin{aligned} && \frac{4B}{k} (q^1-q^2) \sqrt{c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2}} \nonumber\\ =&&-\frac{f_1^\prime(q^1)}{\sqrt{c+(q^1)^2}}-\frac{f_2^\prime(q^2)}{\sqrt{c+(q^2)^2}}\nonumber\\ +&&\frac {c(q^2-q^1)+3q^1\left( c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \right)}{(c+(q^1)^2)^{\frac32}\left( c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \right)} \,f_1(q^1) \nonumber\\ +&&\frac{c(q^1-q^2)+3q^2\left( c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \right)}{\left(c+(q^2)^2\right)^{\frac32}\left( c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \right)} \,f_2(q^2).\nonumber\end{aligned}$$ Making the substitution $$f_1(q^1)=(c+(q^1)^2) F_1(q^1), \ \ \ f_2(q^2)=(c+(q^2)^2) F_2(q^2),$$ we can simplify above equation to be $$\begin{aligned} && \frac{4B}{k} (q^1-q^2) \sqrt{c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2}} \nonumber\\[4mm] =&&-\sqrt{c+(q^1)^2} \,F_1^\prime(q^1)-\sqrt{c+(q^2)^2} \,F_2^\prime(q^2)\nonumber\\[3mm] &+&\frac {q^2 \sqrt{c+(q^1)^2}-q^1 \sqrt{c+(q^2)^2}}{ \ c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \ } \,F_1(q^1)\nonumber\\[2mm] &+&\frac{q^1 \sqrt{c+(q^2)^2}-q^2 \sqrt{c+(q^1)^2}}{ \ c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \ } \,F_2(q^2).\nonumber\end{aligned}$$ We now differentiate this with respect to $q^1$ and $q^2$ to have $$\begin{aligned} -&&\!\!\!\!\!\!\!\!\!\!\!\! \frac{B}{ck}\frac{ (q^1-q^2) \left[3c(q^1-q^2)^2+2(c+3q^1q^2)\left(c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2}\right)\right]}{\sqrt{c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2}}} \nonumber\\ =&&-\sqrt{c+(q^1)^2} \,F_1^\prime(q^1)-\sqrt{c+(q^2)^2} \,F_2^\prime(q^2)\nonumber\\ &+&\frac {q^2 \sqrt{c+(q^1)^2}-q^1 \sqrt{c+(q^2)^2}}{ \ c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \ } \,F_1(q^1)\nonumber\\ &+&\frac{q^1 \sqrt{c+(q^2)^2}-q^2 \sqrt{c+(q^1)^2}}{ \ c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \ } \,F_2(q^2).\nonumber\end{aligned}$$ Note that these two equations are only different in the left hand side, so we can subtract them to get $$B\bigg[c(q^1-q^2)^2+2(c+q^1q^2)\left( c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2}\right)\bigg]=0. \nonumber$$ Since we assumed that $B\neq 0,$ we have $$c(q^1-q^2)^2+2(c+q^1q^2)\left( c+q^1q^2-\sqrt{c+(q^1)^2} \sqrt{c+(q^2)^2} \right)=0,$$ which after simplification reduces to $$0=c^2 (q^1-q^2)^4.$$ From this we have $c=0,$ which leads to a constant potential $h$. Thus in the subcase (ii) we have no required integrable cases. Case (2): $\ \ a(q)=b(q)=c q^2$ Without loss of generality we can assume that $c=1$, so $$a(q^1)=(q^1)^2, \, b(q^2)=(q^2)^2, \, h=q^1+q^2.$$ From equations (\[ph1\]), (\[ph2\]) and (\[alpha-beta relation\]) we have $$\phi^1=k\sqrt{-\frac{f_1 f_2}{(q^1-q^2)^2}}\, , \ \ \ \ \phi^2=-k \sqrt{-\frac{f_1 f_2}{(q^1-q^2)^2}}, \nonumber$$ where $f_1=f_1(q^1), \, f_2=f_2(q^2)$ and $k$ is an arbitrary constant. From condition 4), we have $$\frac{B}{k}= \frac{f_1(q^1)-f_2(q^2)}{2 (q^1-q^2)^3}-\frac{f_1^\prime(q^1)+f_2^\prime(q^2)}{4 (q^1-q^2)^2}. \label{Gaussian curvature}$$ Rearranging and putting $q^1=q^2=q$ implies that $f_1(q)=f_2(q):=f(q)$. Hence equation (\[Gaussian curvature\]) reduces to $$\frac{4B}{k} (q^1-q^2)^3= 2 \left( f(q^1)-f(q^2) \right)-(q^1-q^2) \left(f^\prime (q^1)+f^\prime (q^2) \right). \nonumber$$ Applying $\frac{\partial^3}{\partial q^{1^2} \partial q^{2}}$ to the above equation gives $ f^{\prime \prime \prime}=-24B/k, $ implying that $$f_1(q)=f_2(q)=- \frac{4B}{k} q^3+a_2 q^2+a_1 q+a_0.$$ The Gaussian curvature $K$ in this case is a constant equals to $\frac BK.$\ As before, to find the integral $F$ we use condition 5): $$\partial_1 \varphi=v^1 \partial_1 h+\frac{\phi^2}{\sqrt{g^{11} g^{22}}} B=q^2-k B,\,\, \partial_2 \varphi=v^2 \partial_2 h-\frac{\phi^1}{\sqrt{g^{11} g^{22}}} B=q^1-k B,$$ so that $$\varphi(q^1,q^2)=q^1 q^2 -k B (q^1+q^2). \nonumber$$ Case (3): $\ \ a(q)=b(q)=\frac{c}{q+d}$ We can assume for simplicity that $c=1$ and $d=0$, so $$a(q)=b(q)=1/q, \, h=-1/q^1 q^2.$$ From (\[ph1\]) and (\[ph2\]) we have $$\phi^1=k \sqrt{-\frac{q^1 f_1 f_2}{q^2 (q^1-q^2)^2}}\, , \ \ \ \ \phi^2=-k \sqrt{-\frac{q^2 f_1 f_2}{q^1 (q^1-q^2)^2}}\, . \nonumber$$ Applying the operator $\frac{\partial^4}{\partial q^{1^2} \partial q^{2^2}}$ to condition (C4) in this case, we have $$\Big(5 (q^1)^3+(q^1)^2 q^2 -q^1 (q^2)^2-5(q^2)^3\Big) B =0,$$ which means that in this case magnetic field is zero. Thus we have shown that only cases (1)(i) and (2) lead to the integrable systems with non-zero constant magnetic field and non-constant potential. This completes the proof of Theorem 1. We should emphasize that this is a local classification and all these metrics are incomplete. We are going to show now that under certain assumptions on the parameters these systems can be extended to the analytic integrable systems on a topological sphere, thus presenting some integrable generalisations of the Dirac magnetic monopole. Case I: Dirac magnetic monopole in harmonic field ================================================= To understand the global geometry of the case I we should consider two major subcases, when the cubic polynomial $f(q)=a_3 q^3+a_2 q^2+a_1 q+a_0$ has I a) three distinct real roots; II b) one real root and two complex conjugated roots. It is easy to check that the metric (\[metric\]) is positive definite and has positive Gaussian curvature $K$ only in the case I a) with $a_3<0.$ Let us show that in this case this metric is simply the standard metric on a round sphere $S^2 \subset \mathbb R^3.$ Without loss of generality we can restrict ourselves to the case $a_3=-4$ corresponding to the unit sphere. Consider a sphere given in Cartesian coordinates $x_1$, $x_2$, $x_3$ in $\mathbb R^3$ by the equation $$x_1^{\,2}+x_2^{\,2}+x_3^{\,2}=1, \nonumber$$ and introduce, following C. Neumann, the spherical elliptic coordinates as the roots $q^1$, $q^2$ of the quadratic equation $$\phi(q)=\frac{x_1^{\,2}}{\alpha_1-q}+\frac{x_2^{\,2}}{\alpha_2-q}+\frac{x_3^{\,2}}{\alpha_3-q}=0, \label{quadforq}$$\ where $\alpha_1$, $\alpha_2$, $\alpha_3$ are arbitrary constants (see [@Moser; @N]). Rewrite the quantity $\phi$ in terms of the roots $q^1$, $q^2$ as follows:\ $$\phi(q) =\frac{(q-q^1)(q-q^2)}{(\alpha_1-q)(\alpha_2-q)(\alpha_3-q)} \nonumber$$\ and computing the residues we come to the following expression of the Cartesian coordinates $x_1$, $x_2$, $x_3$ and the spherical elliptic coordinates $q^1$, $q^2$:\ $$\begin{aligned} \ \ x_1^{\,2}= \frac{(\alpha_1-q^1)(\alpha_1-q^2)}{(\alpha_1-\alpha_2)(\alpha_1-\alpha_3)}, \, x_2^{\,2}= \frac{(\alpha_2-q^1)(\alpha_2-q^2)}{(\alpha_2-\alpha_1)(\alpha_2-\alpha_3)}, \, x_3^{\,2}= \frac{(\alpha_3-q^1)(\alpha_3-q^2)}{(\alpha_3-\alpha_1)(\alpha_3-\alpha_2)}. \nonumber\end{aligned}$$ A simple calculation shows then that in the elliptic coordinates $q^1$, $q^2$ the metric takes the form $$ds^2= \frac{q^1-q^2}{4(\alpha_1-q^1)(\alpha_2-q^1)(\alpha_3-q^1)} (dq^1)^2+ \frac{q^2-q^1}{4(\alpha_1-q^2)(\alpha_2-q^2)(\alpha_3-q^2)} (dq^2)^2, \nonumber$$ which is of Stäckel type (\[metric\]) with cubic polynomial $$\begin{aligned} f(x)= 4(\alpha_1-x)(\alpha_2-x)(\alpha_3-x) \nonumber\end{aligned}$$\ having 3 real roots. Note that if we order the roots and the elliptic coordinates by $$\alpha_1>q^1>\alpha_2>q^2>\alpha_3,$$ then we have general case of metrics in class I a) with $x=q^1, \, y=q^2.$ The degenerate case, when two of the roots of cubic $f$ collide, corresponds to the usual spherical coordinates on sphere. Let us show now that in terms of Cartesian coordinates the potential $h=\mu(q^1+q^2)$ is quadratic. We have by definition $$q^2-\Big[ (\alpha_2+\alpha_3) x_1^{\,2}+(\alpha_1+\alpha_3) x_2^{\,2}+(\alpha_1+\alpha_2) x_3^{\,2} \Big]q +(\alpha_2 \alpha_3 x_1^{\,2}+\alpha_1 \alpha_3x_2^{\,2}+\alpha_1 \alpha_2 x_3^{\,2})=0,$$ which implies that $ q^1+q^2= (\alpha_2+\alpha_3) x_1^{\,2}+(\alpha_1+\alpha_3) x_2^{\,2}+(\alpha_1+\alpha_2) x_3^{\,2}. $ Thus the potential $h=\mu(q^1+q^2)$ is a quadratic function of $x_1$, $x_2$, $x_3$, which could be chosen arbitrary. Integrable systems of type I a) with $a_3<0$ can be extended to the Dirac magnetic monopoles on the round sphere in the external harmonic field with arbitrary quadratic potential. They are equivalent to the classical integrable Clebsch systems considered on the co-adjoint orbits of the Euclidean group $E(3).$ Indeed, it is well-known that the Dirac magnetic monopole in the external harmonic field is equivalent to a special Clebsch integrable case of the rigid body motion in the infinite ideal fluid (see [@V]). Recall that the Kirchhoff equations for such a motion are simply Euler equations on the dual space $e(3)^*$ of the Lie algebra of the isometry group $E(3)$ of Euclidean space $\mathbb R^3$ (see e.g. Perelomov [@Perelomov]). The corresponding variables $M_i, x_i, \, i=1,2,3$ have the canonical Lie-Poisson brackets $$\label{e3} \left\{M_i, M_j\right\} = \epsilon_{ijk} M_k, \ \ \left\{M_i, x_j\right\} = \epsilon_{ijk} x_k, \ \ \left\{x_i, x_j \right\}=0.$$ We have two Casimir functions $$C_1=|x|^2,\quad C_2=(M,x).$$ As it was first pointed out by S.P. Novikov and Schmelzer [@NS], the symplectic leaf with $C_1=|x|^2=1, C_2=(M,x)=\nu$ is symplectically isomorphic to the cotangent bundle of the unit sphere $T^*S^2$ with additional Dirac magnetic field with density $B=\nu$. In the coordinates $M,x$ the Hamiltonian and the integral of the corresponding Clebsch system have the form $$\label{clebsch} H=|M|^2-\mu(\alpha_1 x_1^2+\alpha_2 x_2^2+\alpha_3 x_3^2),$$ $$F=\alpha_1 M_1^2+\alpha_2 M_2^2+\alpha_3 M_3^2+\mu(\alpha_2\alpha_3 x_1^2+\alpha_1\alpha_3 x_2^2+\alpha_1\alpha_2 x_3^2).$$ To get the quantum version one should simply replace $M,x$ by $\hat M, \hat x$ with the commutation relations $$[\hat M_i, \hat M_j] = \epsilon_{ijk} \hat M_k, \ \ [\hat M_i, \hat x_j] = \epsilon_{ijk} \hat x_k, \ \ [\hat x_i, \hat x_j]=0.$$ Note that there is no ordering problem since both Hamiltonian and integral written only in terms of the squares of variables. In the remaining cases of type I we have different versions of elliptic coordinates on the hyperbolic plane in external harmonic field, see e.g. [@V2]. Let us consider here only the most degenerate case when $f(x)=4x^3.$ Making change of variables $X=(q^1)^{-1/2},\,Y=(q^2)^{-1/2},$ we have $$ds^2=(\frac{1}{X^2}+\frac{1}{Y^2})(dX^2+dY^2).$$ Denote $w=X+iY$ and $z=w^2=X^2-Y^2+2iXY=u+iv$, then $$ds^2=\frac{X^2+Y^2}{X^2Y^2}(dX^2+dY^2)=\frac{4w\bar{w}}{\Im(w^2)^2}dwd\bar{w}=\frac{dzd\bar{z}}{\Im(z)^2}=\frac{du^2+dv^2}{v^2},$$ which is the canonical hyperbolic metric on the upper half plane. The potential $h$ in $u,v$-coordinates is $$h=\mu(q^1+q^2)=\mu\frac{Y^2-X^2}{X^2Y^2}=-\frac{4\mu u}{v^2}.$$ Case II: new integrable generalisations of Dirac monopole ========================================================= Let us first of all rewrite the formulas in more convenient variables $$x_1=\sqrt{q^1}, \, x_2=\sqrt{q^2}.$$ Then metric (\[metric\]) takes the form $$\label{metric1} ds^2=4\frac{x_1^2-x_2^2}{P(x_1)}dx_1^2+4\frac{x_2^2-x_1^2}{P(x_2)}dx_2^2$$ with $$\label{P} P(x)=a_3x^4+a_2x^2+a_0x+a_1$$ (note an unusual order of the coefficients). The Gaussian curvature in the new coordinates is $$\label{K}K=-\frac{a_3}{4}+\frac{a_0}{(x_1+x_2)^3}.$$ The electric potential $h$ becomes $$\label{h} h=\frac{\mu}{x_1+x_2},$$ while the magnetic potential $A$ is determined by $$\label{A} \partial_1 A_2-\partial_2 A_1=4B\frac{x_1^2-x_2^2}{\sqrt{-P(x_1)P(x_2)}}.$$ The integral $F$ has the form (\[q\_ops\]) with $$\label{phi} \phi^1=k \frac{\sqrt{-P(x_1)P(x_2)}}{2(x_1-x_2)}=-\phi^2, \quad \varphi= -\frac{\mu x_1x_2}{x_1+x_2}-kB(x_1+x_2)^2,$$ where as before $k=-4B/a_3.$ To study the regularity condition we can assume without loss of generality that $a_3<0$ and $a_0\leq 0$. For the analysis of the special case $a_0=0$ we refer to our paper [@VY], so let us assume now that $a_0<0.$ One can show that in order to define regular system on a sphere the polynomial $P(x)$ must have 4 real roots, which we denote $\beta_i, \, i =1,2,3,4:$ $$P(x)=a_3x^4+a_2x^2+a_0x+a_1=a_3(x-\beta_1)(x-\beta_2)(x-\beta_3)(x-\beta_4).$$ We assume also that there are no multiple roots and that $\beta_1>\beta_2>\beta_3>\beta_4,$ such that $$\beta_1+\beta_2+\beta_3+\beta_4=0.$$ Simple arguments show that we have that actually $\beta_1>\beta_2>0>\beta_3>\beta_4$ and that $$\label{roots} \beta_1+\beta_4<0, \quad \beta_2+\beta_3>0$$ (see Figure 1). ![Graph and zeroes of $P(x)$[]{data-label="Case-II"}](graph1 "fig:"){width="2in" height="1.85in"} ![Graph and zeroes of $P(x)$[]{data-label="Case-II"}](graph2 "fig:"){width="2in" height="1.85in"} The algebraic conditions on the coefficients of the quartic polynomial (\[P\]) for having 4 distinct real roots are $$\Delta>0, \,\, a_2 a_3<0, \,\, 4a_1a_3-a_2<0,$$ where $\Delta$ is the discriminant of $P(x)=0$: $$\Delta=256a_1^3a_3^3-128a_1^2a_2^2a_3^2+144a_0^2a_1a_2a_3^2-27a_0^4a_3^2+16a_1a_2^4a_3-4a_0^2a_2^3a_3,$$ or, under our assumption that $a_3<0$, $$\label{condroots1} a_2>0, \,\, a_3<0, \,\, a_1<\frac{a_2}{4a_3}<0,$$ $$\label{condroots2} 256a_1^3a_3^2-128a_1^2a_2^2a_3+144a_0^2a_1a_2a_3-27a_0^4a_3+16a_1a_2^4-4a_0^2a_2^3<0.$$ Under these assumptions we can make change of variables $$\label{u} u_1=\int_{\beta_2}^{x_1}\frac{2dx}{\sqrt{P(x)}},\,\,\,\,\, u_2=\int_{\beta_2}^{x_2} \frac{2dx}{\sqrt{-P(x)}}$$ with $x_1\in[\beta_2,\beta_1], \, x_2 \in[\beta_3,\beta_2].$ We can express the variables $x_1,x_2$ via $u_1,u_2$ using the elliptic function $\mathcal Q(z)$ defined as the inversion $w=\mathcal Q(z)$ of the elliptic integral $$\label{Q} z=\int_{\beta_2}^{w}\frac{2d\xi}{\sqrt{P(\xi)}}=\int_{\beta_2}^{w}\frac{2d\xi}{\sqrt{a_3\xi^4+a_2\xi^2+a_0\xi+a_1}},$$ as follows $$\label{u2} x_1=Q_1(u_1):=\mathcal Q(u_1),\,\,\,\,\, x_2=Q_2(u_2):=\mathcal Q(iu_2).$$ The elliptic function $\mathcal Q(z)$ is even, of order 2 and has two periods: real $2K_1$ and pure imaginary $2iK_2$, where $$\label{periods} K_1=\int_{\beta_2}^{\beta_1}\frac{2d\xi}{\sqrt{P(\xi)}}, \,\,\, K_2=\int_{\beta_2}^{\beta_3}\frac{2d\xi}{\sqrt{-P(\xi)}}.$$ It satisfies the differential equation $$4\mathcal Q'^2=P(\mathcal Q)=a_3\mathcal Q^4+a_2\mathcal Q^2+a_0\mathcal Q+a_1$$ and can be expressed via the standard Weierstrass elliptic function $\wp(z)$. In particular, when $a_0=0$ we have $$4\mathcal Q'^2=a_3(\mathcal Q^2-\beta_1^2)(\mathcal Q^2-\beta_2^2)$$ and $\mathcal Q$ can be written as one of the Jacobi’s elliptic functions [@WW]: $$\mathcal Q=\beta_2 \, \textit{sn}(\alpha (z-\beta_2);k),\,\, \alpha=\sqrt{a_3}\beta_1/2, \,\, k=\beta_2/\beta_1.$$ In the new coordinates the metric (\[metric1\]) takes the form $$\label{metric2} ds^2=(Q_1^2(u_1)-Q_2^2(u_2))(du_1^2+du_2^2),$$ and the potential is $$\label{pot2} h=\frac{\mu}{Q_1(u_1)+Q_2(u_2)}.$$ Consider now the real torus $$\mathbb T^2=\mathbb R^2(u_1,u_2)/4K_1 \mathbb Z\oplus 4K_2 \mathbb Z,$$ identifying the points $(u_1,u_2)$ and $(u_1+4K_1m,u_2+4K_2n), \,\, m,n \in \mathbb Z.$ Formula (\[metric2\]) defines a semi-positive metric on $\mathbb T^2$. Indeed, $$Q_1^2(u_1)-Q_2^2(u_2)=x_1^2-x_2^2=(x_1+x_2)(x_1-x_2)\geq 0,$$ since $x_1+x_2\geq \beta_2+\beta_3>0$ by (\[roots\]) and $x_1\geq x_2.$ The potential $h$ is regular everywhere on the torus, since the denominator $Q_1(u_1)+Q_2(u_2)=x_1+x_2$ is always positive. Thus (\[metric2\]) fails to be a Riemannian metric on $\mathbb T^2$ only at the points when $x_1=x_2=\beta_2,$ which correspond to $(u_1,u_2)=(0,0)$ and three half-periods $(2K_1,0), (0,2K_2), (2K_1,2K_2)$ of the torus. Note that the functions $Q_1$ and $Q_2$ are even, so the metric and the potential are invariant under the involution $$\sigma:(u_1,u_2)\rightarrow (-u_1,-u_2),$$ having exactly those 4 points fixed. The quotient $\mathbb T^2/\sigma=\mathbb S^2$ is a topological sphere (see Fig. 2, where we are using octahedron to represent it). ![Octahedron as a quotient of torus by involution[]{data-label="topsphere"}](square){width="5.6in" height="1.85in"} We claim that the projection $p: \mathbb T^2 \rightarrow \mathbb S^2$ maps the semi-positive metric (\[metric2\]) to a proper Riemannian metric on $\mathbb S^2$ with induced smooth structure. Indeed, we need to check only that this works in the vicinity of the 4 fixed points. Let us check this at the point $(0,0).$ If $x\approx \beta_2$ then $P(x)\approx c(x-\beta_2),\, c=P'(\beta_2),$ $$u_1=\int_{\beta_2}^{x_1}\frac{2dx}{\sqrt{P(x)}}\approx \int_{\beta_2}^{x_1}\frac{2dx}{\sqrt{c(x-\beta_1)}}=\frac{4}{\sqrt{c}}\sqrt{x_1-\beta_2}.$$ Thus near $(0,0)$ we have $ x_1\approx \beta_2+Cu_1^2, \,\, x_2\approx \beta_2-Cu_2^2, \,\, C=\sqrt{c}/4, $ and thus $ x_1+x_2\approx 2\beta_2, \,\, x_1-x_2\approx C(u_1^2+u_2^2),\,\, x_1^2-x_2^2 \approx 2C\beta_2(u_1^2+u_2^2). $ Thus locally metric (\[metric2\]) has the form $ ds^2\approx 2C\beta_2(u_1^2+u_2^2)(du_1^2+du_2^2)=2C\beta_2z\bar z dzd\bar z, $ where we introduced complex coordinate $z=u_1+iu_2.$ The involution $\sigma$ acts by $z\to -z$, so the complex coordinate on the quotient is $w=z^2=v_1+iv_2$, in which metric takes regular form $ ds^2\approx \frac{1}{2}C\beta_2dwd\bar w=\frac{1}{2}C\beta_2(dv_1^2+dv_2^2). $ The situation near 3 other fixed points is similar. Thus we have proved Local integrable systems of type II given by (\[II\]) with parameters, satisfying the conditions (\[condroots1\]),(\[condroots2\]), can be extended to smooth generalisations of Dirac magnetic monopole (\[H1\]),(\[F1\]),(\[A1\]) on topological sphere $\mathbb S^2$ with special metric given in terms of elliptic functions by (\[metric2\]), (\[pot2\]). In the quantum case we should add the usual quantisation conditions for the total magnetic flux $$\label{Quant} \frac{1}{2\pi}B \int_{\mathbb S^2} d\sigma \in \mathbb Z,$$ where $d\sigma$ is the area form on sphere with metric (\[metric2\]). Geometrically this is the integrality of the first Chern class of the corresponding line bundle [@WY]. In the limiting case $a_0=0$ the metric on the sphere becomes standard, but the potential becomes singular at two points. The corresponding system can be viewed as a new integrable version of Euler two-centre problem and was studied in [@VY]. [@VY] The system of type II given by (\[II\]) with $a_0=0$ can be written, similarly to type I, on the dual Lie algebra $e(3)^*$, where the Hamiltonian and integral have the following form $$\label{H} H=\frac{1}{2}|M|^2-\mu\frac{|q|}{\sqrt{R(q)}},$$ $$\label{F} F=A M_1^2+B M_2^2+\frac{2 \sqrt{A B}}{|q|}(M,q) M_3-2\mu\sqrt{AB}\frac{q_3}{\sqrt{R(q)}},$$ where $ R(q)=Aq_2^2+Bq_1^2+(A+B)q_3^2-2\sqrt{AB}|q|q_3 $ and $\mu, A, B$ are parameters satisfying $A>B>0$. The corresponding electric potential has two Coulomb-like singularities, so this system can be considered as new integrable two-centre problem on the sphere in the external Dirac magnetic field. Let us consider now another limiting case when $\beta_1=\beta_2$, assuming for simplicity that $a_3=-1.$ The function $\mathcal Q(u)$ satisfies the equation $$4\mathcal Q'^2=-(\mathcal Q-\beta_1)^2 R(\mathcal Q),\quad R(\xi)=(\xi-\beta_3)(\xi-\beta_4).$$ Solving this equation and putting $u=iu_2$, we have $$\label{limit2} x_2=Q_2(u_2)=\beta_1-\frac{4ce^{\frac{1}{2}\sqrt{c}u_2}}{(b+e^{\frac{1}{2}\sqrt{c}u_2})^2-4c},$$ where $$b=2\beta_1-\beta_3-\beta_4=4\beta_1>0, \,\, c=R(\beta_1)=(\beta_1-\beta_3)(\beta_1-\beta_4)>0.$$ Note that since $b^2-4c>0$ the denominator in (\[limit2\]) is always positive, $\beta_3\leq Q_2(u_2) < \beta_1=\beta_2$ and $Q_2(u_2)\to \beta_1$ as $u_2 \to \pm \infty.$ Since $$\beta_1-\frac{4ce^{\frac{1}{2}\sqrt{c}u}}{(b+e^{\frac{1}{2}\sqrt{c}u})^2-4c}=\beta_1-\frac{4c}{\sqrt{D}\cosh\frac12\sqrt{c}(u-\delta)+2b},$$ where $D=b^2-4c, \delta=\frac{\ln D}{\sqrt{c}},$ we see that $Q_2$ has the symmetry $$Q_2(2\delta-u)=Q_2(u).$$ We have a problem with the first coordinate $x_1$ though, since the second solution of the equation is $Q_1(u_1)\equiv \beta_1$. To deal with this issue we consider the limit $\beta_2 \to \beta_1$ more carefully. Namely, let us introduce $\varepsilon=\frac12 (\beta_1-\beta_2), \bar \beta=\frac12 (\beta_1+\beta_2),$ so that $$-(x-\beta_1)(x-\beta_2)=\varepsilon^2-(x-\bar \beta)^2.$$ Define now coordinate $u_1$ as the integral $$u_1=\int_{\bar\beta}^{x_1} \frac{dx}{\sqrt{\varepsilon^2-(x-\bar \beta)^2}}=\arcsin \frac{x_1-\bar\beta}{\varepsilon},$$ so that the inversion gives $$\label{limit1} x_1=\bar\beta+\varepsilon \sin u_1.$$ Since we have $$du_1^2=\frac{dx_1^2}{\varepsilon^2-(x_1-\bar \beta)^2}$$ we see that in coordinates $u_1,u_2$ when $\varepsilon \to 0$ the metric (\[metric1\]) has the following limit on the cylinder $0\leq u_1 \leq 2\pi, \, u_2 \in \mathbb R$ : $$\label{metric3} ds^2=\frac{4(\beta_1^2-Q_2^2(u_2))}{c}\left [du_1^2+\frac{c}{4}du_2^2\right ].$$ where $c=R(\beta_1)=(\beta_1-\beta_3)(\beta_1-\beta_4).$ We claim that this metric can be extended to the sphere. To show consider first the central projection $p$ of the cylinder $x^2+y^2=1$ to the unit sphere $S^2$ given by $x^2+y^2+z^2=1.$ Parametrising the cylinder as $x=\cos v_1, y=\sin v_1, z=\sinh v_2$ after a simple calculation we have the following form of the metric on the cylinder, induced from the standard metric on the $S^2$: $$\label{metric4} ds^2=\frac{1}{\cosh^2 v_2}\left [dv_1^2+ dv_2^2\right].$$ Now let us change variables in (\[metric3\]) as follows $$\label{change2} u_1=2\tilde{u}_1, \, u_2=\frac{4}{\sqrt{c}}\tilde{u}_2+\delta,$$ so that the metric takes the form $$\label{metric5} ds^2=\frac{16(\beta_1^2-\tilde{Q}_2^2(\tilde{u}_2))}{c}\left [d\tilde{u}_1^2+d\tilde{u}_2^2\right ]$$ with $$\tilde{Q}_2(\tilde{u})=\beta_1-\frac{4c}{\sqrt{D}\cosh 2\tilde u+2b}.$$ Since $\beta_1^2-\tilde{Q}_2^2(\tilde{u}_2)$ decays as $A e^{-2 \tilde{u}_2}, \, A=\frac{8\beta_1 c}{\sqrt{D}}$ when $\tilde{u}_2 \to \infty$ (and as $A e^{2\tilde{u}_2}$ when $\tilde{u}_2 \to -\infty$), we see that the asymptotic behaviour of the metric (\[metric5\]) is the same as the standard metric on the unit sphere (in cylindrical version (\[metric4\])). Note that the change $u_1=2\tilde{u}_1$ corresponds to the double covering of the sphere by the cylinder (which is the degeneration of the torus). The electric potential $h$ in the coordinates $u_1,u_2$ has the form $$h=\frac{\mu}{\beta_1+Q_2(u_2)},$$ while the magnetic potential satisfies $$\partial_2 A_1-\partial_1 A_2=B\frac{\beta_1^2-Q_2^2(u_2)}{R(\beta_1)}.$$ Note that since the right-hand side is independent on $u_1$, we can choose $A_2\equiv 0$ and $$A_1=\int_{-\infty}^{u_2} B\frac{\beta_1^2-Q_2^2(\xi)}{R(\beta_1)}d\xi.$$ Since all the coefficients in the Hamiltonian $$H=\frac{R(\beta_1)}{4(\beta_1^2-Q_2^2(u_2))}\left [(p_1-A_1(u_2))^2+\frac{4}{R(\beta_1)}p_2^2\right ]+\frac{\mu}{\beta_1+Q_2(u_2)}$$ do not depend on $u_1,$ the system has an obvious linear integral $F=p_1,$ and thus is not covered by Theorem 1. Concluding remarks ================== There are several natural questions about new integrable case II, which are still to be answered. We have an interesting metric on topological $\mathbb S^2$ defined by (\[metric2\]). Can it be induced from the Euclidean metric by a suitable embedding of $\mathbb S^2$ into $\mathbb R^3$? If yes, is there an explicit realisation of such a surface? To study the orbits in the classical version of new system and especially the spectrum of the corresponding quantum problem seems to be a very difficult problem. Part of the reasons is the non-zero magnetic field, which is prevent the standard use of the separation of variables (see although recent interesting progress in this direction in [@MagriS; @S]). A limiting even case with $a_0=0$ would be easier to study since in that case we have the usual Dirac magnetic monopole with additional electric field [@VY]. Acknowledgements ================ We are very grateful to Alexey Bolsinov and Jenya Ferapontov for many useful and stimulating discussions. The work of A.P. Veselov was supported by the Russian Science Foundation grant no. 20-11-20214. [99]{} J. Bérubé and P. Winternitz [*Integrable and superintegrable quantum systems in a magnetic field.*]{} J. Math. Phys. [**45**]{} (2004), 1959-1973. A. Clebsch [*Über die Bewegung eines Körpers in einer Flüssigkeit.*]{} Math. Annalen [**3**]{} (1870), 238-262. P.A.M. Dirac [*Quantised singularities in the electromagnetic field.*]{} Proc. Roy. Soc. [**A 133**]{} (1931), 60–72. B. Dorizzi, B. Grammaticos, A. Ramani and P. Winternitz [*Integrable Hamiltonian systems with velocity-dependent potentials.*]{} J. Math. Phys. [**26**]{} (1985), 3070-3079. E. V. Ferapontov and A. P. Fordy [*Non-homogeneous systems of hydrodynamic type, related to quadratic Hamiltonians with electromagnetic term.*]{} Phys. D [**108**]{} (1997), 350-364. E.V. Ferapontov, M. Sayles and A.P. Veselov [*Integrable Schrödinger operators with magnetic fields.*]{} Unpublished, 2005. E. V. Ferapontov and A. P. Veselov [*Integrable Schrödinger operators with magnetic fields: factorization method on curved surfaces.*]{} J. Math. Phys. [**42**]{} (2001), 590-607. G.M. Kemp and A.P. Veselov [*On geometric quantization of the Dirac magnetic monopole.*]{} J. Nonlin. Math. Phys. [**21:1**]{} (2014), 34-42. L. D. Landau and E. M. Lifshitz [*Quantum Mechanics - Non-relativistic Theory.*]{} (Pergamon, New York, 1965). F. Magri and T. Skrypnyk [*Clebsch system.*]{} arXiv 1512.04872 (2015) E. McSween and P. Winternitz [*Integrable and superintegrable Hamiltonian systems in magnetic fields.*]{} J. Math. Phys. [**41**]{} (2000), 2957-2967. I.M. Mladenov and V.V. Tsanov [*Geometric quantisation of the MIC-Kepler problem.*]{} J. Phys. A [**20**]{} (1987), no. 17, 5865–5871. J. Moser [*Various aspects of integrable Hamiltonian systems.*]{} Progr. Math. [**8**]{} (1980), 233-289. V. P. Nair and A. P. Polychronakos [*Quantum mechanics on the noncommutative plane and sphere.*]{} Phys. Lett. [**B 505**]{} (2001), 267. C. Neumann [*De problemate quodam mechanico, quod ad primam integralium ultraellipticorum classem revocatur.*]{} J. Reine Angew. Math. [**3**]{} (1859), 54-66. S.P. Novikov and I. Schmelzer [*Periodic solutions of Kirchhoff’s equations for the free motion of a rigid body in a fluid and the extended theory of Lyusternik-Shnirelman-Morse. I.*]{} Funct. Anal. Appl. [**15:3**]{} (1981), 54–66. A.M. Perelomov [*Integrable Systems of Classical Mechanics and Lie Algebras.*]{} Birkhäuser, 1989. A.D. Polyanin and V.F. Zaitsev [*Handbook of Ordinary Differential Equations: Exact Solutions, Methods, and Problems.*]{} 2nd Edition, Chapman and Hall/CRC, 2003. C. Tejero Prieto [*Quantization and spectral geometry of a rigid body in a magnetic monopole field.*]{} Differential Geom. Appl. [**14**]{} (2001), no. 2, 157-179. T. Skrypnyk [*“Symmetric" separation of variables for the Clebsch system.*]{} J. Geom. Physics [**135**]{} (2019), 204-218. M. A. Soloviev [*Dirac’s magnetic monopole and the Kontsevich star product.*]{} J. Phys. A [**51:9**]{} (2018), 095205. A.P. Veselov [*Landau-Lifschitz equation and integrable systems of classical mechanics.*]{} Dokl. AN SSSR [**270:5**]{} (1983), 1094-1097. A.P. Veselov [*Confocal surfaces and integrable billiards on the sphere and in the Lobachevsky space.*]{} J. Geom. Physics [**7**]{}(1990), 81-107. A.P. Veselov and Y. Ye [*New integrable two-centre problem on sphere in Dirac magnetic field.*]{} arXiv:1907.06174 E.T. Whittaker and G.N. Watson [*A Course in Modern Analysis*]{}, 4th ed., Cambridge University Press, 1990. T.T. Wu and C.N. Yang [*Dirac monopole without strings: monopole harmonics.*]{} Nuclear Physics B [**107**]{} (1976), 365–380.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We consider nonlinear elliptic systems satisfying componentwise coercivity condition. The nonlinear terms have controlled growths with respect to the solution and its gradient, while the behaviour in the independent variable is governed by functions in Morrey spaces. We firstly prove essential boundedness of the weak solution and then obtain Morrey regularity of its gradient.' address: | Department of Civil Engineering,\ Design, Construction and Environment\ Second University of Naples\ Via Roma 29\ 81031 Aversa\ Italy author: - 'Lubomira G. Softova' title: Boundedness of the solutions to nonlinear systems with Morrey data --- 1362 \#1 \#1\#2\#3[[0= -.60]{}]{} Ø[[O]{}]{} ¶[[P]{}]{} [H]{} Ł[[L]{}]{} ł[ ]{} [[const]{}]{}[[const]{}]{} [[loc]{}]{}[[loc]{}]{} \[thm\][Lemma]{} \[thm\][Corollary]{} \[thm\][Proposition]{} \[thm\][Definition]{} \[thm\][Remark]{} \[thm\][Example]{} Introduction ============ Let $\Omega\subset{{\Bbb R}}^n, n\geq 2$ be a bounded domain satisfying the -condition. We are interested in boundedness and Morrey regularity of the weak solutions to nonlinear elliptic systems of the type $$\label{NS0} -\div\bA(x,\bu, D\bu)+\bb(x,\bu,D\bu)=\ff(x),\qquad x\in \Omega$$ where the nonlinear terms are Carathéodory maps $$\begin{aligned} \bA(x,\bu,\bz): &\ \Omega\times{{\Bbb R}}^N\times{{\Bbb M}}^{N\times n}\to {{\Bbb R}}^{N\times n},\\ \bb(x,\bu,\bz): &\ \Omega\times{{\Bbb R}}^N\times{{\Bbb M}}^{N\times n}\to {{\Bbb R}}^{N}.\end{aligned}$$ The celebrated result of De Giorgi [@DeG] and Nash [@N] implies that any weak solution $u\in W_0^{1,2}(\Omega)$ of the linear elliptic equation $D_i(A_{ij}(x)D_ju +g_i(x))=f(x)$ is locally Hölder continuous when $g_i\in L^p$ with $p>n$ and $f\in L^q$ with $q>n/2,$ even if the coefficients are only $L^\infty.$ Unfortunately the De Giorgi-Nash result does not hold anymore if we consider a system of uniformly elliptic equations because of the lack of [*Maximum principle*]{}. This was shown by De Giorgi himself almost ten years later, constructing a counterexample [@DG]. Precisely, the function $\bu=1-x/|x|^\gamma\in W_0^{1,2}(\B_1(0);{{\Bbb R}}^n)$ is a solution to $$D_i(A_{ij}^{\alpha\beta}(x)D_ju^\beta(x))=0\qquad \text{ in } \B_1(0)$$ with suitably chosen coefficients $A_{ij}^{\alpha\beta}\in L^\infty(\B_1(0)).$ Moreover, the result of De Giorgi-Nash cannot be extended to quasilinear systems even if the coefficients are analytic functions, as it was shown by Giusti and Miranda in [@GMir]. In order to get a maximum principle for elliptic systems we need to impose some quite restrictive structural conditions. The simplest one requires the system to be in diagonal form, or [*decoupled.*]{} Consider the operator $ \div (\bA(x,D\bu))=0$ in $\Omega $ with coefficients $$A_i^\alpha(x,D\bu)=\sum_{j=1}^n\sum_{\beta=1}^N\delta_{\alpha\beta}A^{\alpha\beta}_{ij}(x)D_ju^\beta$$ where $\delta_{\alpha\beta}$ is the Kronecker delta. Then $u^\alpha$ solves a single elliptic equation and $ \sup_\Omega u^\alpha\leq \sup_{\partial \Omega} u^\alpha, $ for each $\alpha=1,\ldots,N.$ One more example was given by Nečas and Stará in [@NS]. Consider the system $ \div\bA(x,\bu,D\bu)=0$ in $ \Omega $ that is diagonal for large values of $u^\alpha,$ that is, $$\label{exNS} 0<\theta_\alpha\leq u^\alpha\ \Longrightarrow\ A_i^\alpha(x,\bu,D\bu)= \sum_{j=1}^n\sum_{\beta=1}^N\delta_{\alpha\beta} A_{ij}^{\alpha\beta}(x,\bu)D_ju^\beta$$ with bounded and elliptic $A_{ij}^{\alpha\beta}.$ It turns out that $$\sup_\Omega u^\alpha\leq \max\big\{ \theta_\alpha; \sup_{\partial\Omega} u^\alpha \big\}$$ also in this case. The situation becomes more complicated if we consider [*general nonlinear system*]{} $$\label{NS} \div\bA(x,\bu, D\bu)=\bb(x,\bu,D\bu)\,.$$ Along with the Carathéodory conditions on the maps $\bA(x,\bu,\bz)$ and $\bb(x,\bu,\bz)$ we need to control also the growths of $\bA$ and $\bb$ with respect to $\bu$ and $\bz.$ These additional [*controlled growth conditions*]{} ensure the convergence of the integrals in the definition of [*weak solution*]{} to (see ). In [@LP] Leonetti and Petricca assume [*componentwise coercivity condition*]{} on $\bA$ and positivity of $\bb$ for large values of $u^\alpha,$ that is, there exist positive constants $\theta_\alpha$ such that $$\label{exLP} \theta_\alpha\leq u^\alpha\quad \Longrightarrow\quad \begin{cases} \ds \nu|\bz^\alpha|^p -M_\alpha\leq \sum_{i=1}^n A_i^\alpha(x,\bu,\bz)z_i^\alpha\\ \ds 0\leq b^\alpha(x,\bu,\bz)\,. \end{cases}$$ Combining the [*Sobolev inequality*]{} with the [*Stampacchia Lemma [@Stamp]*]{} they get a componentwise bound of the solution, covering this way also the systems studied in [@NS], since is a special case of . Let us note that getting essential boundedness of the weak solution to is a starting point for a further study of its regularity in various function spaces. In [@DK; @P; @PS2] the authors obtain better integrability and Hölder regularity of the bounded solutions to quasilinear elliptic equations $(N=1)$ under controlled growth conditions on the nonlinear terms. Further this result has been extended in [@Sf] to semilinear uniformly elliptic systems of the form $$\label{QLS} \div(\bA(x)D\bu+\ba(x,\bu))=\bb(x,\bu,D\bu)\qquad \text{ in } \Omega$$ with minimal regular assumptions on the coefficients and the underlying domain. Precisely, it is shown that if the nonlinear terms satisfy the controlled growth conditions with $\varphi\in L^p(\Omega),$ $p>2$ and $\psi\in L^q(\Omega), $ $q>\frac{2n}{n+2}$ then any bounded weak solution to belongs to $W_0^{1,r}$ with $r=\min\{p,q^*\}.$ The natural question that arises is what kind of regularity of the solution to we can expect if the given functions $\varphi$ and $\psi$ belong to some Morrey spaces. In the case of a single equation we count with the result of Byun and Palagachev [@BP]. Combining the Gehring-Giaquinta-Modica lemma, the Adams trace inequality and the Hartmann-Stampacchia maximum principle they obtain $L^\infty$ estimate of the solution. Further, the Morrey-type estimate of the gradient permits the authors to show also Hölder regularity of the solution. Our goal is to obtain a componentwise maximum principle for any solution of supposing that the operators $\bA$ and $\bb$ satisfy structural conditions expressed in terms of Morrey functions. As a consequence we obtain also Morrey regularity of the gradient of $\bu$ extending such a way the regularity results obtained in [@BP; @DK; @LP; @NS; @PSesyst; @Sf] to nonlinear systems with Morrey data. Recall that a real valued function $f\in L^p(\Omega)$ belongs to the Morrey space $L^{p,\lambda}(\Omega)$ with $p\in[1,\infty),$ $\lambda\in(0,n),$ if $$\label{defMorrey} \|f\|_{p,\lambda;\Omega}=\left( \sup_{\B_r(x) }\frac1{r^\lambda} \int_{\Omega\cap\B_r(x)}|f(y)|^p\,dy \right)^{1/p}<\infty$$ where the supremum is taken over all balls $\B_r(x),$ $r\in(0,\diam\Omega]$ and $x\in \ol\Omega.$ Working in the framework of the Morrey spaces we note that the Sobolev trace inequality is not enough anymore. For this goal we will use the following result due to Adams. \[AdamsTrace\] Let $m$ be a positive Radon measure with support in $\Omega$ and such that for each ball $\B_\rho$ it holds $$\label{tau0} m(\B_{\rho}) \leq K \rho^{\tau_0},\quad \tau_0=\frac{s}{r}(n-r),\quad 1<r<s<\infty, \quad r<n$$ with an absolute constant $K>0.$ Then $$\label{Adams1} \left(\int_\Omega |v(x)|^s\, dm \right)^{\frac1s}\leq C(n,s,r)K^{\frac1s}\left( \int_\Omega |Dv(x)|^r\, dx \right)^{\frac1r}$$ for each function $ v\in W_0^{1,r}(\Omega).$ In what follows we suppose that $\Omega\subset {{\Bbb R}}^n, n\geq2,$ is a bounded domain satisfying the -condition, that is, there exists a constant $A_\Omega>0$ such that $$\tag{A}\label{A} |\Omega_r(x)|\geq A_\Omega\, r^n\qquad \forall\ x\in\ol\Omega, \ r\in(0,\diam\Omega]$$ where $\Omega_r(x)=\Omega\cap \B_r(x).$ It is worth noting that the -condition excludes interior cusps at each point of the boundary and guarantees the validity of the Sobolev embedding theorem in $W^{1,p}(\Omega).$ This geometric property is surely satisfied when $\partial\Omega$ has the uniform interior cone property (e.g. $C^1$-smooth or Lipschitz continuous boundaries), but it holds also for the Reifenberg falt domains boundaries (cf. [@PS2]). Throughout the text the standard summation convention on the repeated indexes is adopted. The letter $C>0$ is used for various constants and may change from one occurrence to another. Maximum principle ================= Consider the nonlinear system $$\label{NonlSyst} - D_i\big(A_{i}^{\alpha}(x,\bu, D\bu)\big)+ b^\alpha(x,\bu,D\bu) =f^\alpha(x) \ \text{ in } \Omega$$ where $ \bA=\{A^\alpha_i(x,\bu,\bz)\}_{i\leq n}^{\alpha\leq N} $ and $ \bb=(b^1(x,\bu,\bz),\ldots, b^N(x,\bu,\bz)) $ are measurable in $x\in\Omega$ and continuous in $(\bu,\bz)$ for almost all (a.a.) $x\in\Omega.$ Suppose that for each $(x,\bu,\bz)\in\Omega\times {{\Bbb R}}^N\times {{\mathbb M}}^{N\times n}$ the following [*controlled growth conditions*]{} hold. Namely, $$\label{contr1} \begin{cases} \ds |\bA(x,\bu,\bz) |\leq \Lambda\big(\varphi(x)+|\bu|^{\frac{\nu}2}+|\bz|\big)\\[5pt] \ds |\bb (x,\bu,\bz)|\leq \Lambda\big(\psi(x) + |\bu|^{\nu-1}+|\bz|^{2\frac{\nu-1}{\nu})} \big) \end{cases}$$ as $ |\bu|, |\bz|\to\infty,$ with some positive constant $\Lambda.$ Here $\nu$ is the Sobolev conjugate of $2,$ that is, $$\label{conjugate} \nu=\begin{cases} \ds \frac{2n}{n-2} & \text{ if } n\geq 3\\ \text{any number } > 2 & \text{ if } n =2, \end{cases}$$ and the given functions $\varphi,$ $\psi$ and $f^\alpha$ satisfy $$\label{regularity} \begin{cases} \ds \varphi\in L^{p,\lambda}(\Omega), & p>2, \ \lambda\in(0,n), \ p+\lambda>n\\[5pt] \ds \psi, f^\alpha \in L^{q,\mu}(\Omega), & q>\frac{\nu}{\nu-1}, \ \mu\in(0,n), \ 2q+\mu>n. \end{cases}$$ In the particular case $n=2$ the powers of $|\bu|$ could be arbitrary positive numbers greater then 1, while the growth of $|\bz|$ is strictly sub-quadratic (cf. [@G; @LU]). Under a [*weak solution*]{} of we mean a function $\bu\in W^{1,2}(\Omega;{{\Bbb R}}^N)\cap L^\nu(\Omega;{{\Bbb R}}^N),$ satisfying $$\begin{aligned} \label{weak} \nonumber \sum_{i=1}^n \int_\Omega A^\alpha_i(x,\bu, D\bu)D_i\phi^\alpha(x)\, dx & +\int_\Omega b^\alpha(x,\bu(x),D \bu(x))\phi^\alpha(x)\, dx\\ & = \int_{\Omega} f^\alpha(x) \phi^\alpha(x)\,dx\end{aligned}$$ for all $\phi=(\phi^1,\ldots,\phi^N)\in W_0^{1,2}(\Omega;{{\Bbb R}}^N).$ The conditions - are the natural ones that ensure the convergence of the integrals in . Moreover, they are optimal as it is seen from the following example in the case of single equation (cf. [@LSU; @P]). The function $ u(x)=|x|^{\frac{r-2}{r-1}}\in W^{1,2}(\B_1(0)),$ with $ n\geq3 $ and $ \frac{n+2}{n}<r<2$ is a solution to the equation $\Delta u=C|Du|^r$ in $\B_1(0).$ Note that $u\not\in L^\infty(\B_1(0)).$ Generally we cannot expect boundedness of the solutions to unless we add some restrictions on the structure of the operator (see for example [@JS; @LP]). For this goal we impose componentwise coercivity on $A_i^\alpha$ and a sign condition on $b^\alpha.$ For every $\alpha\in\{1,\ldots,N\}$ there exist positive constants $\theta_\alpha,$ $ \gamma $ and a function $ \varphi $ such that for each $u^\alpha\geq \theta_\alpha$ we have $$\label{coercivity} \begin{cases} \ds \gamma|\bz^\alpha|^2 - \Lambda \varphi(x)^2\leq \sum_{i=1}^n A_i^\alpha (x,\bu,\bz)z_i^\alpha\\ \ds\varphi\in L^{p,\lambda}(\Omega), \ p>2, \ p+\lambda>n\\ \ds 0\leq b^\alpha(x,\bu,\bz)\quad \text{ for a.a. } x\in\Omega, \ \forall \ \bz\in {{\mathbb M}}^{N\times n}\,. \end{cases}$$ \[thm1\] Let $\Omega$ be -type domain and $\bu\in W^{1,2}(\Omega;{{\Bbb R}}^N)\cap L^\nu(\Omega; {{\Bbb R}}^N)$ be a weak solution to under the conditions , and and such that $\sup_{\partial\Omega}u^\alpha<\infty$. Then $$\sup_\Omega u^\alpha \leq\max\{ \theta_\alpha, \sup_{\partial\Omega} u^\alpha \}+ M_\alpha\qquad \alpha\in \{1,\ldots,N\}$$ where $M_\alpha$ depends on $ n, p, \lambda, \Lambda, \gamma,$ $\|\varphi \|_{p,\lambda;\Omega},$ $\| f^\alpha \|_{q,\mu;\Omega}$ and $ |\Omega|.$ We choose a constant $L>0$ such that $ L\geq \max\{ \theta_\alpha; \sup_{\partial\Omega}u^\alpha \} $ and define the set $ \cA_L^\alpha=\{ x\in\Omega:\ u^\alpha(x)-L>0 \}\,. $ Then we take a vector function $\bv$ as follows $$v^\beta= \begin{cases} \ds \max\,\{u^\alpha-L; 0\} &\text{if } \beta=\alpha\\ 0 &\text{if } \beta\not=\alpha \end{cases},\qquad D v^{\beta}=\begin{cases} D u^\alpha\chi_{\cA_L^\alpha} &\text{if } \beta=\alpha\\ 0 &\text{if } \beta\not=\alpha \end{cases}.$$ It is clear that $ \bv \in W_0^{1,2}(\Omega;{{\Bbb R}}^N)$ and hence $\bv\in L^\nu(\Omega;{{\Bbb R}}^N)$ by the Sobolev embedding. Choosing $\phi^\alpha=v^\alpha$ as a test function we obtain $$\begin{aligned} \int_{\cA^\alpha_L} A_i^\alpha(x,\bu,D\bu)D_iu^\alpha(x) \,dx \ +& \int_{\cA^\alpha_L} b^\alpha(x,\bu,D\bu)(u^\alpha(x)-L)\,dx\\ =& \int_{\cA^\alpha_L}f^\alpha(x)(u^\alpha(x)-L)\,dx\,.\end{aligned}$$ We start with the case $ n\geq 3$ when $\nu=2n/(n-2).$ Define the Radon measure $dm$ supported in $\Omega$ by $$dm:=(\chi_\Omega(x) +\varphi(x)^2+|f^\alpha(x)|)\,dx,$$ where $\chi_\Omega$ is the characteristic function of $\Omega.$ Then by we get the estimate $$\begin{aligned} \label{eq1} \nonumber \int_{\cA^\alpha_L} |Du^\alpha(x)|^2\,dx\leq &\ \frac{\Lambda}{\gamma}\int_{\cA^\alpha_L} \varphi(x)^2\,dx+\frac1{\gamma} \int_{\cA^\alpha_L}|f^\alpha(x)|(u^\alpha(x)-L)\,dx\\ \nonumber \leq &\ \frac{\Lambda}{\gamma} \int_{\cA^\alpha_L} (\chi_\Omega(x) +\varphi(x)^2 +|f^\alpha(x)|)\,dx\\ &\ + \frac{1}{\gamma} \int_{\cA^\alpha_L}|f^\alpha(x)|(u^\alpha(x)-L)\,dx\\ \nonumber \leq &\ C(\Lambda,\gamma)\big( m(\cA^\alpha_L )+ J \big).\end{aligned}$$ In order to estimate the integral $J= \int_{\cA^\alpha_L}|f^\alpha(x)|(u^\alpha(x)-L)\,dx $ we make use of the Lemma \[AdamsTrace\] applied to the Radon measure $dm'=|f^\alpha(x)|dx.$ Hence $$J=\int_{\cA^\alpha_L}(u^\alpha(x)-L)dm'\leq \left(\int_{\cA_L^\alpha}|u^\alpha(x)-L|^{s'}\,dm' \right)^{\frac1{s'}} m'(\cA^\alpha_L)^{1-\frac1{s'}}\,.$$ Evaluating the measure $m'$ over a ball $\B_\rho$ we get $$\begin{aligned} m'(\B_\rho)=& \ \int_{\B_\rho} |f^\alpha(x)|\,dx \leq C(n)\rho^{n-\frac{n-\mu}q}\left(\frac1{\rho^\mu} \int_{\B_\rho} |f^\alpha(x)|^q\,dx \right)^{1/q}\\ \leq&\ C(n)\rho^{n-\frac{n-\mu}q} \|f^\alpha\|_{q,\mu;\Omega}=K \rho^{n-\frac{n-\mu}q}\end{aligned}$$ with $K=K(n,q,\diam\Omega, \|f^\alpha\|_{q,\mu;\Omega} ).$ We apply now the Lemma \[AdamsTrace\] with $r'=2,$ $\tau'_0=n-\frac{n-\mu}{q}$ and $s'=\frac2{n-2}\big(n-\frac{n-\mu}q \big)>2,$ calculated via . Hence $$\begin{aligned} \label{eq2} \nonumber J\leq &\ C K^{\frac1{s'}}\left( \int_{\cA^\alpha_L} |Du^\alpha(x)|^2\,dx\right)^{\frac12} m'(\cA^\alpha_L)^{1-{\frac1{s'}}}\\ \leq &\ C\left[ \varepsilon \int_{\cA_L^\alpha} |Du^\alpha(x)|^2\,dx+\frac1{\varepsilon} m'(\cA_L^\alpha)^{2(1- {\frac1{s'}})} \right]\,.\end{aligned}$$ Combining and , taking $\varepsilon $ small enough, moving the integral of the gradient on the left-hand side, and keeping in mind that $2(1-\frac1{s'})>1$ and $m'(\cA^\alpha_L)\leq m(\cA^\alpha_L)$ we obtain $$\label{eq3} \int_{\cA^\alpha_L}|Du^\alpha(x)|^2\,dx\leq C\big( m(\cA^\alpha_L) +m(\cA^\alpha_L)^{2(1-\frac1{s'})} \big)\leq C m(\cA^\alpha_L)$$ where the constant $C$ depends on known quantities. To complete the estimate we will use once again the Lemma \[AdamsTrace\]. It is immediate that $m(\B_\rho)$ of a ball $\B_\rho\Subset\Omega$ is $$\begin{aligned} \label{measure_bound} \nonumber m(\B_{\rho}) & = \int_{\B_\rho}\big(\chi_\Omega(x)+\varphi(x)^2+|f^\alpha(x)| \big)\,dx\\ &\leq C(n) \rho^n+ \rho^{n-\frac{2(n-\lambda)}p} \|\varphi\|^2_{p,\lambda;\Omega}+ \rho^{n-\frac{n-\mu}{q}}\|f^\alpha\|_{q,\mu;\Omega} \leq K \rho^{\tau_0}\end{aligned}$$ with $K=K(n,p,q,\diam \Omega, \|\varphi\|_{p,\lambda;\Omega},\|f^\alpha\|_{q,\mu;\Omega} )$ and $$\tau_0=\min\left\{n-\frac{2(n-\lambda)}p; n-\frac{n-\mu}q \right\} >n-2.$$ Applying with $r=2<n$ and calculating $s$ from we get $$\begin{aligned} \label{estimate1} \nonumber \int_{\cA^\alpha_L} (u^\alpha(x)-L)\,dm &\leq \left( \int_{\cA^\alpha_L}|u^\alpha(x)-L|^s\,dm \right)^{\frac1s} m(\cA^\alpha_L)^{1-\frac1s}\\ &\leq C K^{\frac1s} \left( \int_{\cA^\alpha_L}|D u^\alpha(x)|^2\,dx \right)^{\frac12} m(\cA^\alpha_L)^{1-\frac1s} \\ \nonumber &\leq C(n,p,K,\gamma,\Lambda)m(\cA^\alpha_L)^{1+\frac12-\frac1s}\end{aligned}$$ with $s=\min\left\{\frac{2np-4(n-\lambda)}{p(n-2)} ; \frac{2nq-2(n-\mu)}{q(n-2)} \right\}>2.$ A similar bound holds also in the case $n=2.$ In fact, for any ball $\B_\rho\subset{{\Bbb R}}^2$ we have $$m(\B_\rho)\leq C\rho^2+ \rho^{2-\frac{2(2-\lambda)}p} \|\varphi\|^2_{p,\lambda;\Omega} + \rho^{2-\frac{2-\mu}{q}} \|f^\alpha\|_{q,\mu;\Omega} \leq K \rho^{\tau_0}$$ with $\tau_0=\min\left\{ 2-\frac{2(2-\lambda)}{p}; 2-\frac{2-\mu}{q} \right\}>0.$ Choosing $s=2$ we calculate $r$ from $$r=\max\left\{ \frac{2p}{2p-2+\lambda}; \frac{4q}{4q-2+\mu} \right\}\in(1,2).$$ Then by the Hölder and the Adams trace inequalities we obtain $$\begin{aligned} \label{estimate2} \nonumber \int_{\cA^\alpha_L} (u^\alpha(x)&-L)\,dm \leq\left(\int_{\cA^\alpha_L} (u^\alpha(x)-L)^2\,dm \right)^{\frac12}m(\cA^\alpha_L)^{\frac12}\\ & \leq C K^{\frac12}\left( \int_{\cA^\alpha_L} |Du^\alpha(x)|^r\,dx \right)^{\frac1r} m(\cA^\alpha_L)^{\frac12}\\ \nonumber & \leq C K^{\frac12}\left( \int_{\cA^\alpha_L}|Du^\alpha(x)|^{2}\,dx \right)^{\frac{1}2} \left(\int_{\cA^\alpha_L}\chi_\Omega(x)\, dx \right)^{\frac1r-\frac{1}2}m(\cA^\alpha_L)^{\frac12}\\ \nonumber & = CK^{\frac12} \left( \int_{\cA^\alpha_L}|Du^\alpha(x)|^{2}\,dx \right)^{\frac{1}2} m(\cA^\alpha_L)^{\frac1r}\,.\end{aligned}$$ In order to estimate the integral in the last term we go back to . Consider again the Radon measure $dm'=|f^\alpha(x)|dx$ and calculate $m'(\B_\rho)\leq K\rho^{2-\frac{2-\mu}{q}}.$ Then choosing $s'=2$ we get $r'=\frac{4q}{4q-(2-\mu)}\in(1,2)$ from . This way, the Lemma \[AdamsTrace\] and the Hölder inequality give $$\begin{aligned} \label{eq4} \nonumber J\leq &\ \left(\int_{\cA_L^\alpha}|u^\alpha(x)-L|^2\,dm' \right)^{\frac12} m'(\cA^\alpha_L)^{\frac12}\\ \leq &\ C K^{\frac12} \left(\int_{\cA_L^\alpha}|Du^\alpha(x)|^{r'}\,dx \right)^{\frac1{r'}} m'(\cA^\alpha_L)^{\frac12}\\ \nonumber \leq &\ C K^{\frac12} \left(\int_{\cA_L^\alpha}|Du^\alpha(x)|^2\,dx \right)^{\frac12} m(\cA^\alpha_L)^{\frac1{r'}-\frac12} m(\cA^\alpha_L)^{\frac12}\\ \nonumber \leq &\ C\left[\varepsilon \int_{\cA_L^\alpha}|Du^\alpha(x)|^2\,dx +\frac1{\varepsilon} m(\cA_L^\alpha)^{\frac2{r'}} \right]\,.\end{aligned}$$ Unifying and , taking $\varepsilon$ small enough and keeping in mind that $\frac2{r'}>1,$ we get $$\int_{\cA_L^\alpha}|Du^\alpha(x)|^2\,dx\leq C m(\cA_L^\alpha)$$ where the constant depends on the same quantities as in . Then the estimate becomes $$\label{eq5} \int_{\cA^\alpha_L}(u^\alpha(x)-L)\,dm\leq C m(\cA^\alpha_L)^{1+\frac1r-\frac12}.$$ Unifying the estimates and we obtain $$\label{estimate3} \int_{\cA^\alpha_L}(u^\alpha(x)-L)\, dm\leq C m(\cA^\alpha_L)^{1+\sigma_0}$$ where $$\sigma_0=\begin{cases} \ds\frac12-\frac1s=\max\big\{\frac{p+\lambda-n}{np-2(n-\lambda)};\frac{\mu+2q-n}{2nq-2(n-\mu)} \big\} & \text{ if } n>2\\[10pt] \ds\frac1r-\frac12= \min\big\{\frac{p+\lambda-2}{2p}; \frac{2q+\mu-2}{4q} \big\} & \text{ if } n=2\,. \end{cases}$$ Suppose now that $ m(\cA^\alpha_L)>0, $ otherwise $\sup_{\Omega}u^\alpha(x)\leq L.$ For any $L_1>L$ we have $\cA^\alpha_{L_1}\subset \cA^\alpha_L$ and therefore yields $$\begin{aligned} (L_1-L)m(\cA^\alpha_{L_1})&\leq \int_{\cA^\alpha_{L_1}}(u^\alpha(x)-L)\,dm\\ &\leq \int_{\cA^\alpha_L}(u^\alpha(x)-L)\,dm\leq C m(\cA^\alpha_L)^{1+\sigma_0}\,.\end{aligned}$$ Hence $$m(\cA^\alpha_{L_1})\leq \frac{C}{L_1-L}m(\cA^\alpha_L)^{1+\sigma_0}\,.$$ In order to estimate the measure of the set $\cA_L^\alpha$ we will apply the following Maximum Principle due to Stampacchia [@Stamp Lemma 4.1]. \[lemMaximum\] Let $\Theta:[L_0,\infty)\to [0,\infty)$ be a decreasing function. Assume that there exist $c,a\in (0,\infty)$ and $b\in (1,\infty)$ such that $$L_1>L\geq L_0 \ \Longrightarrow \ \Theta(L_1)\leq \frac{c}{(L_1-L)^\alpha}(\Theta(L))^b.$$ Then $$\Theta(L_0+d)=0 \quad \text{ where }\quad d=\big[c\Theta(L_0)^{b-1} 2^{\frac{ab}{b-1}} \big]^{\frac1a}\,.$$ The application of the Lemma \[lemMaximum\] to the function $\Theta(L)=m(\cA^\alpha_L)$ with $a=1,$ $b=1+\sigma_0$ and $L_0=\max\{\theta_\alpha,\sup_{\partial\Omega} u^\alpha \}$ yields $$\label{estimate4} m(\cA^\alpha_{L_0+d_\alpha})=0 \qquad \text{ where } \qquad d_\alpha\leq C m(\Omega)^{\sigma_0} 2^{1+\frac1{\sigma_0}}\,.$$ The last assertion means that for each $\alpha=1,\ldots,N$ there exists a constant $M_\alpha$ depending on $n,$ $p,$ $\lambda,$ $q,$ $\mu,$ $\gamma,$ $\Lambda,$ $|\Omega|,$ $\|\varphi\|_{p,\lambda;\Omega}$ and $\|f^\alpha\|_{q,\mu;\Omega}$ such that $$\label{Maximum} \sup_{\Omega}u^\alpha<\max\{\theta_\alpha; \sup_{\partial\Omega}u^\alpha\}+M_\alpha$$ and this completes the proof of Theorem \[thm1\] The Dirichlet Problem ===================== We study the boundedness and the Morrey regularity of the weak solutions to the following Dirichlet problem $$\label{DP3} \begin{cases} - \div\bA(x,\bu(x) D\bu(x))+ \bb(x,\bu,D\bu) =\ff(x) & x\in\Omega \\ \bu(x)=0\quad & x\in \partial\Omega \end{cases}$$ in a bounded domain $\Omega\subset{{\Bbb R}}^n.$ \[crlr1\] Let $\bu\in W_0^{1,2}(\Omega,{{\Bbb R}}^N)$ be a solution to and assume , , and . Suppose in addition that $$\label{coercivity2} \begin{cases} \ds \gamma |\bz^\alpha|^2-\Lambda \varphi(x)^2\leq \sum_{i=1}^n A_i^\alpha(x,\bu,\bz)z_i^\alpha\\[6pt] \ds \varphi(x)\in L^{p,\lambda}(\Omega), \ p>2, \ \lambda\in(0,n), \ \lambda+p>n,\\[6pt] \ds 0\leq b^\alpha(x,\bu,\bz)\, \sign u^\alpha(x) \end{cases}$$ for $|u^\alpha|\geq \theta_\alpha>0.$ Then there exists a constant $M$ depending on known quantities such that $$\|\bu\|_{\infty;\Omega}\leq M\,.$$ Take a positive constant $L$ such that $L\geq \theta_\alpha$ and consider the set $\bar \cA^\alpha_L=\{x\in\Omega: u^\alpha(x)+L<0 \}.$ Then the Theorem \[thm1\] applied to $-u^\alpha$ gives $$\label{eq6} \inf_\Omega u^\alpha>-\theta_\alpha - M_\alpha\,.$$ Unifying and we get boundedness of $\|u^\alpha\|_{\infty;\Omega}$ for each $\alpha=1,\ldots,N.$ Then $$\|\bu\|_{\infty;\Omega}=\max_{1\leq \alpha\leq N}\|u^\alpha\|_{\infty;\Omega}=: M<\infty\,.$$ \[MorreyEst\] Let $\Omega $ be a bounded -type domain in ${{\Bbb R}}^n, n\geq3,$ and $\bu\in W_0^{1,2}(\Omega,{{\Bbb R}}^N) $ be a weak solution to under the assumptions , , and . Then $D\bu\in L^{2,n-2}(\Omega,{{\Bbb R}}^N)$ and $$\label{Gradient4} \int_{\Omega_\rho(x)}|Du(y)|^2\,dy\leq C \rho^{n-2}\qquad \forall x\in\Omega, \ \rho\in(0,\diam\Omega]$$ with a constant depending on known quantities. Fix $x_0\in\Omega$ and $\rho>0$ be such that $\B_\rho(x_0)\subset \B_{2\rho}(x_0)\Subset \Omega,$ $\rho>1.$ Define a cut-off function $\zeta(x)\in C^1({{\Bbb R}}^n)$ $$\zeta(x)=\begin{cases} 1 & x\in \B_{\rho}(x_0),\\ 0 & x\not\in \B_{2\rho}(x_0), \end{cases}\qquad |D\zeta|\leq \frac{C}{\rho}\,.$$ For any fixed $\alpha$ take $\phi^\alpha(x)=e^{u^\alpha(x)}\zeta(x)^2 $ as a test function in to get $$\begin{aligned} \sum_{i=1}^n \int_{\Omega}& A_i^\alpha(x,\bu, D\bu) e^{u^\alpha(x)}D_iu^\alpha(x) \zeta(x)^2\,dx\\ &= \int_{\Omega} f^\alpha(x) e^{u^\alpha(x)}2\zeta(x) D_i\zeta(x)\,dx\\ &\quad - \sum_{i=1}^n \int_{\Omega} A_i^\alpha(x,\bu, D\bu)e^{u^\alpha(x)}2\zeta(x) D_i\zeta(x)\,dx\\ &\quad -\int_{\Omega}b^\alpha(x,\bu,D\bu)e^{u^\alpha(x)}\zeta(x)^2\,dx\,.\end{aligned}$$ The left-hand side can be estimated by while for the right-hand side we use and $$\begin{aligned} \sum_{i=1}^n& e^{-M}\int_{\Omega} \big(\gamma |D u^\alpha(x)|^2 -\Lambda \varphi(x)^2 \big)\zeta(x)^2 \,dx\\ & \leq 2e^{M} \int_{\Omega} | f^\alpha(x)| \zeta(x) |D\zeta(x)|\,dx\\ &\quad + 2 n \Lambda e^{M}\int_{\Omega} \big(\varphi(x)+|\bu|^{\frac{n}{n-2}} +|D\bu| \big)\zeta(x) |D\zeta|\,dx\\ &\quad + \Lambda e^{M} \int_{\Omega}\big( \psi(x)+|\bu|^{\frac{n+2}{n-2} }+|D\bu|^{\frac{n+2}{n}} \big) \zeta(x)^2\,dx.\end{aligned}$$ To proceed further, we use the Young inequality $ ab\leq \varepsilon a^p+\frac{b^{p/(p-1)}}{\varepsilon^{1/(p-1)}}, $ whence $$\begin{aligned} \int_{\Omega}|f^\alpha(x)| \zeta(x) |D\zeta(x)|\,dx\leq &\ \frac12\int_{\Omega}| f^\alpha(x)|^2 \zeta(x)^2\,dx +\frac12\int_{\Omega}|D\zeta(x)|^2\,dx\\ \int_{\Omega}\varphi(x) \zeta(x) |D\zeta(x)|\,dx\leq&\ \frac12\int_{\Omega} \varphi(x)^2 \zeta(x)^2\,dx +\frac12\int_{\Omega}|D\zeta(x)|^2\,dx\\ \int_{\Omega} |\bu|^{\frac{n}{n-2}}\zeta(x) |D\zeta(x)|\,dx \leq&\ \frac12 M^{\frac{n}{n-2}}\left( \int_{\Omega}\zeta(x)^2\,dx+ \int_{\Omega}|D\zeta(x)|^2\,dx \right)\\ \int_{\Omega}|D\bu|\zeta(x) |D\zeta|\,dx\leq&\ \varepsilon\int_{\Omega}|D\bu|^2\zeta(x)^2\,dx +\frac1{\varepsilon} \int_{\Omega}|D\zeta(x)|^2\,dx\\ \int_{\Omega} |D\bu|^{\frac{n+2}{n}}\zeta(x)^2\,dx\leq&\ \varepsilon\int_{\Omega} |D\bu|^2\zeta(x)^2\,dx +\varepsilon^{-\frac{n+2}{n-2}}\int_{\Omega}\zeta(x)^2\,dx\,.\end{aligned}$$ Unifying the above estimates we get $$\begin{aligned} \label{Gradient} \nonumber \int_{\Omega}& |Du^\alpha(x)|^2\zeta(x)^2\,dx\\ &\ \leq C\int_{\Omega}(1+|f^\alpha(x)|+\psi(x)+\varphi(x)^2)\zeta(x)^2\,dx\\ \nonumber &\quad +C\int_{\Omega}|D\zeta(x)|^2\,dx+ \varepsilon C \int_{\Omega} |D\bu(x)|^2\zeta(x)^2\,dx\end{aligned}$$ with constants depending on $n, \Lambda,\gamma, M,$ and $ \varepsilon.$ Summing up over $\alpha$ from $1$ to $N,$ fixing $\varepsilon$ small enough and moving the last term to the left-hand side we obtain $$\begin{aligned} \label{Gradient2} \nonumber \int_{\Omega} & |D\bu(x)|^2\zeta(x)^2\,dx \leq C\int_{\Omega}(1+\psi(x)+\varphi(x)^2) \zeta(x)^2\, dx\\ &+C \sum_{\alpha=1}^N\int_{\Omega} |f^\alpha(x)| \zeta(x)^2\,dx +C\int_{\Omega}|D\zeta(x)|^2\,dx\,.\end{aligned}$$ Then, by the definition of $\zeta$ and by we have $$\begin{aligned} & \int_{\B_{2\rho}} \big( 1+ \psi(x)+ \varphi(x)^2 \big)\,dx \leq C \big[\rho^n+ \rho^{n-\frac{n-\mu}{q}}\|\psi\|_{q,\mu;\Omega}\\ &\qquad\qquad + \rho^{n-\frac{2(n-\lambda)}{p}}\|\varphi\|^2_{p,\lambda;\Omega}\big]\\ &\sum_{\alpha=1}^N \int_{\B_{2\rho}} |f^\alpha(x)| \,dx \leq C \rho^{n-\frac{n-\mu}{q}}\|\ff\|_{q,\mu;\Omega} \\ &\int_{\B_{2\rho}}|D\zeta(x)|^2\,dx\leq C \rho^{n-2}\,.\end{aligned}$$ Hence $$\label{Gradient3} \int_{\B_\rho}|D\bu |^2\,dx \leq C \rho^{\lambda_0}$$ with $\lambda_0=\min\{n-2, n-\frac{2(n-\lambda)}{p}, n-\frac{n-\mu}{q} \}=n-2$ and the constant depends on known quantities. Let $\B_{\rho}(x_0)\cap \partial\Omega\not=\emptyset.$ Then we extend $u^\alpha$ and the given functions $f^\alpha,\varphi,$ and $\psi$ as zero in $\Omega^c$ and consider the test functions $$\phi^\alpha(x)=(e^{|u^\alpha(x)|}-1)\zeta(x)^2\sign u^\alpha(x)$$ where $\zeta(x)$ is the cut-off function defined above. Thus gives $$\begin{aligned} \sum_{i=1}^n& \int_{\Omega} A_i^\alpha(x,\bu,D\bu) e^{|u^\alpha(x)|} D_iu^\alpha(x)\zeta(x)^2\,dx\\ &\ = \int_{\Omega} f^\alpha(x) (e^{|u^\alpha(x)|}-1)\zeta(x)^2 \sign u^\alpha(x)\,dx\\ &\quad - \sum_{i=1}^n \int_{\Omega} A_i^\alpha(x,\bu,D\bu) (e^{|u^\alpha(x)|}-1)2\zeta(x)D_i\zeta(x) \sign u^\alpha(x)\,dx\\ &\quad - \int_{\Omega} b^\alpha(x,\bu,D\bu) (e^{|u^\alpha(x)|}-1)\zeta(x)^2\sign u^\alpha(x)\,dx\,.\end{aligned}$$ Hence the conditions and give $$\begin{aligned} \gamma \int_{\Omega}& |D u^\alpha(x)|^2\zeta(x)^2\,dx\leq \Lambda e^{M}\int_{\Omega} \varphi(x)^2 \zeta(x)^2 \,dx\\ &\ +e^{M}\int_\Omega|f^\alpha(x)|\zeta(x)^2\,dx\\ &\ +2 n\Lambda e^{M} \int_{\Omega} (\varphi(x)+ |\bu|^{\frac{n}{n-2}}+|D\bu|)\zeta(x) |D\zeta(x)|\,dx\end{aligned}$$ and to get the desired estimate we argue as above. [12]{} Adams, D., Traces of potentials. II., [*Indiana Univ. Math. J.*]{} [**22**]{} (1973), 907–918. Byun, S.-S., Palagachev, D., Boundedness of the weak solutions to quasilinear elliptic equations with Morrey data, Indiana Univ. Math. J., [**62**]{} (5) (2013), 1565–1585. Campanato, S., [*Sistemi ellittici in forma divergenza. Regolarità all’interno,*]{} Pubblicazioni della Classe di Scienze: [ Quaderni, Scuola Norm. Sup., Pisa,]{} 1980. Chiarenza, F., Regularity for solutions of quasilinear elliptic equations under minimal assumptions, [*Pot. Anal.,*]{} [**4**]{} (4) (1995), 325–334. De Giorgi, E., Sulla differenziabilitá e l’analiticitá delle estremali degli integrali multipli regolary, [*Mem. Accad. Sci. Torino Cl. Sci. Fis. Mat. Natur.,*]{} [**3**]{} (3) (1957), 25–43. De Giorgi, E., [Un esempio di estremali discontinue per un problema variazionale di tipo ellittico,]{} [*Bull. Unione Mat. It.,*]{} [**4**]{} (1968), 135–137. Dong, H., Kim, D., Global regularity of weak solutions to quasilinear elliptic and parabolic equations with controlled growth, [*Commun. Part. Differ. Equ.,*]{} [**36**]{} (2011), 1750–1777. Giaquinta, M., *Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic Systems,* Annals of Mathematics Studies, 105, Princeton University Press, Princeton, NJ, 1983. Gilbert, G., Trudinger, N.S., *Elliptic Partial Differential Equations of Second Order,* 2nd ed. Springer-Verlag, 1983. Giusti, E., Miranda, M., Sulla regolarità delle soluzioni deboli di una classe di sistemi ellittici quasi-lineari, [*Arch. Rat. Mech. Anal.,*]{} [**31**]{} (1968), 173–184. John, O., Stará, J., On the regularity and nonregularity of elliptic and parabolic systems, Equadiff-7, Proc. Conf. Prague, 1989; J. Kurzweil ed., Teubner-Texte Math. [**118**]{}, Teubner, Leipzig, 1990, 28–36. Ladyzhenskaya, O.A., Solonnikov, V.A., Ural’tseva, N.N., [*Linear and Quasilinear Equations of Parabolic Type,*]{} Translations of Mathematical Monographs, **23**, Amer. Math. Soc., Providence, R.I., 1968. Ladyzhenskaya, O.A., Ural’tseva, N.N., [*Linear and Quasilinear Equations of Elliptic Type,*]{} 2nd Edition, Nauka, Moscow, 1973, (in Russian). Leonetti, F., Petricca, P.V., Regularity for solutions to some nonlinear elliptic systems, [*Complex Var. Ell. Eq.,*]{} [**56**]{} (12) (2011), 1099–1113. Morrey, C.B., [*Multiple Integrals in the Calculus of Variations,*]{} Springer-Verlag, Berlin, 1966. Nash, J., [Continuity of solutions of parabolic and elliptic equations,]{} *Amer. J. Math.* **80** (1958), 931–954. Nečas, J., J. Stará, [ Principio di massimo per i sistemi ellittici quasi-lineari non diagonali,]{} [*Boll. Unione Mat. Ital.,*]{} [**6**]{} (1972), 1–10. Palagachev, D.K., [Global Hölder continuity of weak solutions to quasilinear divergence form elliptic equations,]{} *J. Math. Anal. Appl.* **359** (2009), 159–167. Palagachev, D.K., Softova, L.G., Fine regularity for elliptic systems with discontinuous ingredients, [*Arch. Math.*]{} **86** (2006), 145–153. Palagachev, D.K., Softova, L.G., The Calderón-Zygmund property for quasilinear divergence form equations over Reifenberg flat domains, [*Nonl. Anal.,*]{} [**74**]{} (2011), 1721–1730. Rokotson, M., Equivalence between the growth of $\int_{B(x,r)} |\nabla u|^pdy$ and $T$ in the equation $P[u]=T,$ [*J. Diff. Equ.,*]{} [**86**]{} (1990), 102–122. Softova, L., $L^p$-integrability of the gradient of solutions to quasilinear systems with discontinuous coefficients, [*Differ. Int. Equ.,*]{} [**26**]{} (9-10) (2013), 1091–1104. Stampacchia, G., [ Equations elliptiques du second ordre a coefficients discontinuous,]{} [*Séminaire de Mathématiques Supérieures,*]{} Université de Montréal, Vol. [**16**]{}, 1966, 326pp.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present the study of the $WH$ and $ZH$ search with the Higgs Boson decayed to $b\bar{b}$ at the Large Hadron Collider. The Higgs Boson and the Vector Boson are required to be boosted, and the Higgs Boson is reconstructed with Jet Trimming Technique. The statistical significance for 30$fb^{-1}$ data is 4.5 $\sigma$, which is comparable to the previous result [@Butterworth:2008iy].' author: - | Wenhan Zhu\ Department of Physics, Princeton University, Princeton, NJ 08544 title: 'Application of Jet Trimming in Boosted Higgs Search[^1]' --- Introduction ============ The Higgs Boson search is the most important search at the Large Hadron Collider (LHC), because it is an essential part of the standard-model electroweak symmetry breaking. Current electroweak fits, together with the LEP and Tevatron exclusion limit, favour a light Higgs boson one with mass around 120 GeV [@Grunewald:2007pm] [@Aaltonen:2010yv]. It is a challenging task for the discovery of Higgs Boson at this mass region [@Aad:2009wy] [@Ball:2007zza]. The Mass Dropping and Filtering method introduced in [@Butterworth:2008iy] makes the Higgs Boson production channel associated with a vector boson very promising. In this channel, the Higgs Boson will decay hadronically into two b-tagged jet with the vector boson decay leptonically. The dominant background for this process is $VV$,$Vj$ and $t\bar{t}$. We will employ a similar kinematic selection to [@Butterworth:2008iy], but we will reconstruct the Higgs Boson using the jet trimming technique [@Krohn:2009th]. Jet Trimming ============ Jet Trimming [@Krohn:2009th] is a designated procedure for removing the ISR/MI/Pileup from the FSR. The intrinsic idea is that ISR/MI/Pileup will be much softer than the FSR, therefore, we will form a fat jet using a larger cone and then recluster the fat jet with a smaller cone and throw away the softer subjets. We will make some change to the original algorithm described in [@Krohn:2009th] for the jet substructure of the boosted Higgs Boson. First, we will find the two b-tagged jets by clustering the jet constituents of the fat jet. Also, the Higgs jet, different from a QCD jet, is a dipole itself, so we expect to have more radiation between the two b quarks. Therefore, we will use a dynamical $f_{\rm cut}$, which is proportional to the distance between the subjet and the fat jet. The jet trimming algorithm proceeds as follows: - Cluster all the final state particles with Fastjet 2.4.2 [@Cacciari:2005hq],anti$k_T$ algorithm with a cone size 1.2. - Cluster the particles in the hardest jet with a smaller cone size 0.3 with anti$k_T$ algorithm to find out the hardest two subjets, and we will require each of the two subjets to be b tagged. We assume a 60% tagging efficiency and 2% of mistagging efficiency. - Cluster the remaining particles with an even smaller cone size 0.2 with $k_T$ algorithm to form the subjets. - If $p_{T}^{i} > f_{\rm cut} p_{T} \Delta R$ the subjet is kept else it is trimmed, $p_{T}$ is the $p_T$ of the fat jet and $\Delta R$ is the distance between the subjet and the fat jet. The $f_{\rm cut}$ is chosen to be 0.03 in this analysis. - Now we have the Higgs Candidate. We will require the Higgs Candidate $p_T$ larger than 200 GeV and $\eta$ less than 2.5. Results ======= The events are generated by Pythia 6.403 [@Sjostrand:2006za], fully showered and hadronized. The underlying event is incorporated by Pythia “DW” tune. For this analysis, signal samples of $WH, ZH$ were generated, as well as $WW,ZW,ZZ,Z+{\rm jet},W+{\rm jet},t\bar{t}$ to study backgrounds. There are three search channels in this analysis and the channel specific cuts are very similar to [@Butterworth:2008iy]: - Leptonic channel: two opposite sign lepton ($e$ or $\mu$) with $p_{T}>30$ GeV and $|\eta|<2.5$, with an invariant mass between 80 and 100 GeV. - Missing $E_T$ channel: Missing Transverse momentum$>200$GeV. - Semi-leptonic Channel: Missing transverse momentum$>30$GeV plus a lepton ($e$ or $\mu$) with $p_T>30$GeV. Veto event if there is jet with $p_{T}>30$ GeV and $|\eta|<3.0$. - all channel: no more lepton with $p_{T}>30$ GeV and $|\eta|<2.5$ except to reconstruct the vector boson, no more b-tagged jets with $p_{T}>30$ and $|\eta|<2.5$. The mass spectrum of the Higgs Candidate with $m_H=$115 GeV is shown in Fig \[fig:mass\] for the three sub-channel and combined channel. The number of both signal and background for Higgs Mass between 112-128 GeV for 30$fb^{-1}$ data is shown in Table \[tab:all\], the significance is 4.5 $\sigma$(8.2$\sigma$ for 100 $fb^{-1}$). The result is comparable to  [@Butterworth:2008iy] and offer an alternative strategy for search for boosted Higgs. ![The Signal and Background for a 115 GeV SM Higgs Boson for 30 $fb^{-1}$. On the top, the left plot is or leptonic channel and the right plot is for missing Et channel. On the bottom, the left plot is for semi-leptonic channel and the right plot is the total signal and background for all the channel. []{data-label="fig:mass"}](channel1 "fig:") ![The Signal and Background for a 115 GeV SM Higgs Boson for 30 $fb^{-1}$. On the top, the left plot is or leptonic channel and the right plot is for missing Et channel. On the bottom, the left plot is for semi-leptonic channel and the right plot is the total signal and background for all the channel. []{data-label="fig:mass"}](channel2 "fig:") ![The Signal and Background for a 115 GeV SM Higgs Boson for 30 $fb^{-1}$. On the top, the left plot is or leptonic channel and the right plot is for missing Et channel. On the bottom, the left plot is for semi-leptonic channel and the right plot is the total signal and background for all the channel. []{data-label="fig:mass"}](channel3 "fig:") ![The Signal and Background for a 115 GeV SM Higgs Boson for 30 $fb^{-1}$. On the top, the left plot is or leptonic channel and the right plot is for missing Et channel. On the bottom, the left plot is for semi-leptonic channel and the right plot is the total signal and background for all the channel. []{data-label="fig:mass"}](all "fig:") Conclusion and Outlook ====================== Here we have applied jet trimming technique to the boosted Higgs Boson search for a low mass(115 GeV) SM Higgs Boson. The statistical significance for 30$fb^{-1}$ data is 4.5 $\sigma$, which is comparable to the previous result  [@Butterworth:2008iy]. This could be considered as an alternative search strategy for the high-$p_T$ $WH$,$ZH$ channel at the LHC. Acknowledgement =============== The author would like to acknowledge useful discussions with Valerie Halyo, David Krohn, Gavin Salam and Lian-tao Wang. [99]{} M. W. Grunewald, J. Phys. Conf. Ser.  [**110**]{}, 042008 (2008). \[arXiv:0709.3744 \[hep-ph\]\]. T. Aaltonen [*et al.*]{} \[CDF and D0 Collaborations\], Phys. Rev. Lett.  [**104**]{}, 061802 (2010) \[arXiv:1001.4162 \[hep-ex\]\]. G. Aad [*et al.*]{} \[ The ATLAS Collaboration \], \[arXiv:0901.0512 \[hep-ex\]\]. G. L. Bayatian [*et al.*]{} \[ CMS Collaboration \], J. Phys. G [**G34**]{}, 995-1579 (2007). D. Krohn, J. Thaler and L. T. Wang, JHEP [**1002**]{}, 084 (2010) \[arXiv:0912.1342 \[hep-ph\]\]. J. M. Butterworth, A. R. Davison, M. Rubin and G. P. Salam, Phys. Rev. Lett.  [**100**]{}, 242001 (2008) \[arXiv:0802.2470 \[hep-ph\]\]. T. Sjostrand, S. Mrenna and P. Z. Skands, JHEP [**0605**]{}, 026 (2006) \[arXiv:hep-ph/0603175\]. M. Cacciari and G. P. Salam, Phys. Lett.  B [**641**]{}, 57 (2006) \[arXiv:hep-ph/0512210\]. [^1]: Talk presented by Wenhan Zhu at Pheno 2011, Madison, Wisconsin, May 2011.
{ "pile_set_name": "ArXiv" }
\#1[to 0pt[\#1]{}]{} \#1[0= -0.25em0-0 .05em0-0 -0.025em.0433em0]{} \#1[\#1]{} \#1 0= =0 =1 =cmr5 =cmr6 =cmr7 =cmr8 =cmr9 =cmr10 =cmr12 =cmr17 =cmmi5 =cmmi6 =cmmi7 =cmmi8 =cmmi9 =cmmi10 =cmmi12 =cmsy5 =cmsy6 =cmsy7 =cmsy8 =cmsy9 =cmsy10 =cmsy10 at 12pt =cmbx5 =cmbx6 =cmbx7 =cmbx8 =cmbx9 =cmbx10 =cmbx12 =cmti7 =cmti8 =cmti9 =cmti10 =cmti12 =cmsl8 =cmsl9 =cmsl10 =cmsl12 =cmtt8 =cmtt9 =cmtt10 =cmtt12 =cmex10 =cmss10 =cmss12 =22.6 true cm =17 true cm =5000 =5000 =17.7cm =23.9cm =cmti12 =cmr12 =cmcsc10 at 12pt =cmr8 =cmr10 =cmti12 =cmti10 \#1 \#1  \#1 \#1 \#1 \#1\#2 1ex  \#2 \#1\#2  \#1 1ex \#2 =0 =0 \#1[99=99 by \#1 99]{} \#1[\#1]{} =0 \#1[99=99 by \#1 Figure 99]{} \#1[\#1]{} =0 \#1 by 1 [^1] \#1 [^1]: $^{\the\notecount}$
{ "pile_set_name": "ArXiv" }
--- abstract: 'Power-meter measurements are used to study a model that accounts for the use of power by a cyclist. The focus is on relations between rates of change of model quantities, such as power and speed, both in the context of partial derivatives, where other quantities are constant, and Lagrange multipliers, where other quantities vary to maintain the imposed constraints.' author: - 'Tomasz Danek [^1], Michael A. Slawinski [^2], Theodore Stanoev [^3]' bibliography: - 'DSSbici2\_arXiv.bib' date: 'May 12, 2020' title: | On modelling bicycle power-meter measurements: Part II\ [Relations between rates of change of model quantities]{} --- Introduction ============ Using power meters to study cycling performance allows us to gain quantitative information about relations whose qualitative aspects are known based on observations; for instance, riding with a given speed with a tailwind requires less effort than riding with the same speed against a headwind. The quantification of such a relation, however, is necessary to proceed with an optimization to achieve—under constraints imposed by the capacity of a cyclist—the least time to cover a distance that is subject to winds and contains flats, hills and descents. Many studies examine the physics of cycling. This article is the second part of @DSSbici1, which also contains the pertinent bibliography. Herein—using a model relating power-meter measurements to the motion of a bicycle, examined by @DSSbici1—we formulate expressions that allow us to quantify relations between the rates of change of parameters contained within this model. We begin this paper by presenting the expression to account, by modelling, for the values measured by a power meter. Using this expression and the implicit function theorem, we derive explicit expressions of the rates between the model parameters. We complete this paper by interpreting and comparing quantitive results based on power-meter and GPS measurements collected on a flat course and an inclined course. Both are in Northwestern Italy; the former is between Rivalta Bormida and Pontechino, in Piemonte; the latter is between Rossiglione and Tiglieto, in Liguria. In the appendices, we compare the flat-course results to optimizations based on the Lagrange multipliers, and comment on the bijection between the generated power and the bicycle speed. Formulation =========== Measurements and model {#sec:MM} ---------------------- To account for the cyclist’s use of the power measured by a power meter, we consider, $$\label{eq:formula} P = F_{\!\leftarrow}\,V_{\!\rightarrow} \,,$$ where $P$ stands for the value of the required power, $F_{\!\leftarrow}$ for the forces opposing the motion, and $V_{\!\rightarrow}$ for the ground speed of the bicycle. Explicitly, we assume that [e.g., @DSSbici1] $$\label{eq:P} P = \underbrace{ \frac{ mg\sin\theta + m\,a + {\rm C_{rr}}mg\cos\theta + \tfrac{1}{2}\,\eta\,{\rm C_{d}A}\,\rho\, \left( V_{\!\rightarrow} + w_{\leftarrow} \right)^{2} }{ 1-\lambda } }_{F_{\!\leftarrow}} V_{\!\rightarrow}\,;$$ herein, $m$ is the mass of the cyclist and the bicycle, $g$ is the acceleration due to gravity, $\theta$ is the slope, $a$ is the change of speed, $\rm C_{rr}$ is the rolling-resistance coefficient, $\rm C_{d}A$ is the air-resistance coefficient, $\rho$ is air density, $V_{\!\rightarrow}$ is the ground-speed of the bicycle, $w_{\leftarrow}$ is the wind component opposing the motion, $\lambda$ is the drivetrain-resistance coefficient, $\eta$ is a quantity that ensures the proper sign for the tailwind effect, $w_{\leftarrow}<-V_{\!\rightarrow}\iff\eta=-1$, otherwise, $\eta=1$; throughout this work, $\eta=1$. To estimate quantities that appear on the right-hand side of equation (\[eq:P\])—specifically, $\rm C_{d}A$, $\rm C_{rr}$ and $\lambda$—given the measurement, $P$, we write $$\label{eq:model} f = P - \underbrace{ \frac{ mg\sin\theta\,V_{\!\rightarrow} + m\,a\,V_{\!\rightarrow} + {\rm C_{rr}}mg\cos\theta\,V_{\!\rightarrow} + \tfrac{1}{2}\,\eta\,{\rm C_{d}A}\,\rho \left(V_{\!\rightarrow}+w_{\leftarrow}\right)^{2} V_{\!\rightarrow} }{ 1-\lambda } }_{F_{\!\leftarrow} V_{\!\rightarrow}} \,,$$ and minimize the misfit, $\min f$, as discussed by @DSSbici1. Implicit function theorem ------------------------- We seek relations between the ratios of quantities on the right-hand side of equation (\[eq:P\]). To do so—since $f$, stated in expression (\[eq:model\]), possesses continuous partial derivatives in all its variables at all points, except at ${\lambda=1}$, which is excluded by mechanical considerations, and since ${f=0}$, as a consequence of equation (\[eq:formula\])—we invoke the implicit function theorem to write $$\label{eq:Thm} \dfrac{\partial y}{\partial x} = -\dfrac{ \dfrac{\partial f}{\partial x} }{ \dfrac{\partial f}{\partial y} } =: -\dfrac{\partial_{x}f}{\partial_{y}f} \,,$$ where $x$ and $y$ are any two quantities among the arguments of $$\label{eq:f} f(P,m,g,\theta,a,{\rm C_{rr}},{\rm C_{d}A},\rho,V_{\!\rightarrow},w_{\leftarrow},\lambda) \,.$$ Expressions of partial derivatives {#sub:Partials} ---------------------------------- To use formula (\[eq:Thm\]), in the context of expression (\[eq:model\]), we obtain all partial derivatives of $f$, with respect to its arguments. $$\partial_{P}f = 1$$ $$\partial_{m}f = -\dfrac{ \left( a + g\left({\rm C_{rr}}\cos\theta+\sin\theta\right) \right) V_{\!\rightarrow} }{ 1-\lambda }$$ $$\partial_{\theta}f = -\dfrac{ m\,g\left(\cos\theta-{\rm C_{rr}}\sin\theta\right) V_{\!\rightarrow} }{ 1-\lambda }$$ $$\partial_{a}f = -\dfrac{m\,V_{\!\rightarrow}}{1-\lambda}$$ $$\partial_{\rm C_{rr}}f = -\dfrac{m\,g\,\cos\theta\,V_{\!\rightarrow}}{1-\lambda}$$ $$\partial_{\rm C_{d}A}f = -\dfrac{ \eta\,\rho \left(V_{\!\rightarrow}+w_{\leftarrow}\right)^{2} V_{\!\rightarrow} }{ 2\,(1-\lambda) }$$ $$\partial_{\rho}f = -\dfrac{ \eta\,{\rm C_{d}A} \left(V_{\!\rightarrow}+w_{\leftarrow}\right)^{2} V_{\!\rightarrow} }{ 2\,(1-\lambda) }$$ $$\partial_{V_{\!\rightarrow}}f = -\dfrac{ 2\,m\,a + \eta\,\rho\,{\rm C_{d}A} \left(V_{\!\rightarrow}+w_{\leftarrow}\right) \left(3\,V_{\!\rightarrow}+w_{\leftarrow}\right) + 2\,m\,g\left({\rm C_{rr}}\cos\theta+\sin\theta\right) }{ 2\,(1-\lambda) }$$ $$\partial_{w_{\leftarrow}}f = -\dfrac{ \eta\,\rho\,{\rm C_{d}A} \left(V_{\!\rightarrow}+w_{\leftarrow}\right) V_{\!\rightarrow} }{ 1-\lambda }$$ $$\partial_{\lambda}f = -\dfrac{ \left( 2\,m\,a + \eta\,\rho\,{\rm C_{d}A} \left(V_{\!\rightarrow}+w_{\leftarrow}\right)^{2} + 2\,m\,g\left( {\rm C_{rr}}\cos\theta+\sin\theta \right) \right) V_{\!\rightarrow} }{ 2\,(1-\lambda)^2 }$$ In accordance with the definition of a partial derivative, all variables in expression (\[eq:f\]) are constant, except the one with respect to which the differentiation is performed. This property is apparent in Appendix \[app:ConstraintPower\], where we examine a relation between differences and derivatives. Values of partial derivatives ----------------------------- ### Common input values To use the partial derivatives stated in Section \[sub:Partials\], we consider measurements collected during two rides. The flat-course measurements correspond to a nearly flat course. The inclined-course measurements correspond to an uphill with a nearly constant inclination. For both the flat and inclined course, we let $m=111$ and $g=9.81$.[^4] Other values are stated in Sections \[sec:FlatCourseVals\] and \[sec:InclCourseVals\]. ### Flat-course input values {#sec:FlatCourseVals} According to @DSSbici1, the flat-course input values are as follows. The average measured power, $\overline P=258.8\pm57.3$, and speed, $\overline V_{\!\rightarrow}=10.51\pm0.9816$. The values inferred by modelling are ${\rm C_{d}A}=0.2607\pm0.002982$, ${\rm C_{rr}}=0.00231\pm0.005447$ and $\lambda=0.03574\pm0.0004375$. We set $w_{\leftarrow}=0\implies\eta=1$ and $\rho=1.20406\pm0.000764447$. The average slope is $\overline\theta=0.002575\pm0.04027$, which indicates a flat course. The change of speed is $\overline a=0.006922\pm0.1655$, which indicates a steady tempo. Thus, we set $\overline\theta = \overline a = 0$. The corresponding values of partial derivatives, formulated in Section \[sub:Partials\], are listed in the left-hand column of Table \[table:Partials\]. ### Inclined-course input values {#sec:InclCourseVals} ![ Left-hand plot: ground speed, from GPS measurements, $\overline V_{\!\rightarrow}=4.138\pm0.2063$; right-hand plot: power, from power meters, $\overline{P}=286.6\pm33.07$ []{data-label="fig:FigSpeedPowerSteep"}](FigSpeedSteep.pdf "fig:") ![ Left-hand plot: ground speed, from GPS measurements, $\overline V_{\!\rightarrow}=4.138\pm0.2063$; right-hand plot: power, from power meters, $\overline{P}=286.6\pm33.07$ []{data-label="fig:FigSpeedPowerSteep"}](FigPowerSteep.pdf "fig:") ![ Optimal values; left-hand plot: ${\rm C_{d}A}=0.2702\pm0.002773$; middle plot: ${\rm C_{rr}}=0.01298\pm0.011$; right-hand plot: $\lambda=0.02979\pm0.004396$[]{data-label="fig:FigParSteep"}](FigParSteep.pdf) Following the method outlined by @DSSbici1, we calculate the inclined-course values. The data are grouped in eleven speed intervals whose centres range from $3.7$ to $4.7$, and which contain three-hundred-and-ninety-two data points. The mode is $4.2$, and is represented by seventy-eight data points. The distributions of the ground speed and power are illustrated in Figure \[fig:FigSpeedPowerSteep\]; their means are $\overline V_{\!\rightarrow}=4.138\pm0.2063$ and $\overline P=286.5783\pm33.11394$, respectively. The distributions of the inferred parameters are illustrated in Figure \[fig:FigParSteep\]; their values are ${\rm C_{d}A}=0.2702\pm0.002773$, ${\rm C_{rr}}=0.01298\pm0.011$, $\lambda=0.02979\pm0.004396$. The average slope of the inclined course is $\overline\theta=0.04592\pm0.1106$, which is $2.63^\circ$ and $4.60\%$. This course is known among cyclists of the region as a particularly constant incline. The change of speed throughout the ride is $\overline a=0.001011\pm0.1015$, which indicates a steady pace. In view of the constantness and steadiness, we set, $\overline\theta=0.04592$ and $\overline a = 0$. Since the change of altitude is negligible—in the context of air density—we set $\overline\rho=1.168\pm0.001861$, which—under standard meteorological conditions—corresponds to the altitude of $400$. We set $w_{\leftarrow}=0\implies\eta=1$. The corresponding values of partial derivatives, formulated in Section \[sub:Partials\], are listed in the right-hand column of Table \[table:Partials\]. ### Theorem requirements ![Misfit of equation (\[eq:model\]): left-hand plot: flat course, $f=0.4137\pm6.321$; right-hand plot: inclined course, $f=2.03\pm4.911$[]{data-label="fig:Figf"}](Figfflat.pdf "fig:") ![Misfit of equation (\[eq:model\]): left-hand plot: flat course, $f=0.4137\pm6.321$; right-hand plot: inclined course, $f=2.03\pm4.911$[]{data-label="fig:Figf"}](Figfsteep.pdf "fig:") As required by the implicit function theorem and as shown in Figure \[fig:Figf\], $f=0$, in the neighbourhood of the maxima of the distributions, for both the flat and inclined courses. Also, as required by the theorem, in formula (\[eq:Thm\]), and as shown in Table \[table:Partials\], $\partial_{y}f\neq0$, in the neighbourhoods of interest, for either course. Notably, the similarity of a horizontal spread for both plots of Figure \[fig:Figf\] indicates that the goodness of fit of a model is similar for both courses. The spread is slightly narrower for the inclined course; this might be a result of a lower average speed, $\overline V_{\!\rightarrow}$, which allows for more data points for a given distance and, hence, a higher accuracy of information. Interpretation ============== Model considerations {#sec:MC} -------------------- The misfit minimization of equation (\[eq:model\]), $\min f$, treats $\rm C_{d}A$, $\rm C_{rr}$ and $\lambda$ as adjustable parameters. The values in Table \[table:ModelRates\] are the changes of $\rm C_{d}A$ due to a change in $\rm C_{rr}$ or $\lambda$; in either case, the other quantities are kept constant. Let us examine the first row. For the flat course—in the neighbourhood of $\overline V_{\!\rightarrow}=10.51$ and $\overline P=258.8$, wherein ${\rm C_{d}A}=0.2607$ and ${\rm C_{rr}}=0.00231$—$\partial_{\rm C_{rr}}{\rm C_{d}A}=-16.3745$ and, in accordance with expression (\[eq:Thm\]), its reciprocal is $\partial_{\rm C_{d}A}{\rm C_{rr}}=-0.0610705$. We write the corresponding differentials as $${\rm d(C_{d}A)} = \dfrac{\partial{\rm C_{d}A}}{\partial{\rm C_{rr}}}\, {\rm d(C_{rr})} = {-16.3745}\,{\rm d(C_{rr})}$$ and $${\rm d(C_{rr})} = \dfrac{\partial{\rm C_{rr}}}{\partial{\rm C_{d}A}}\, {\rm d(C_{d}A)} = {-0.0610705}\,{\rm d(C_{d}A)} \,;$$ in other words, an increase of $\rm C_{rr}$ by a unit corresponds to a decrease of $\rm C_{d}A$ by $16.3745$ units, and an increase of $\rm C_{d}A$ by a unit corresponds to a decrease of $\rm C_{rr}$ by ${0.0610705}$ of a unit. Thus, $$\left. \dfrac{\rm d(C_{d}A)}{\rm C_{d}A} \right|_{{\rm C_{d}A}=0.2607} = \dfrac{1}{0.2607} \,,$$ which is an increase of about $384\%$, corresponds to $$\left. \dfrac{\rm d(C_{rr})}{\rm C_{rr}} \right|_{{\rm C_{rr}}=0.00231} = -\dfrac{0.0610705}{0.00231} \,,$$ which is a decrease of about $2644\%$. For the inclined course—in the neighbourhood of $\overline V_{\!\rightarrow}=4.138$ and $\overline P=286.5783$, wherein ${\rm C_{d}A}=0.2702$ and ${\rm C_{rr}}=0.01298$—$\partial_{\rm C_{rr}}{\rm C_{d}A}=-108.754$ and its reciprocal is $\partial_{\rm C_{d}A}{\rm C_{rr}}=-0.0091951$. Following the same method as for the flat course, we see that an increase of $\rm C_{d}A$ by about $1/0.2702=370\%$ corresponds to a decrease of $\rm C_{rr}$ by about $0.0091951/0.01298=71\%$. Remaining within a linear approximation, an increase of $\rm C_{d}A$ by $1\%$ corresponds to a decrease of $\rm C_{rr}$ by $6.89\%$, for the flat course, and a decrease of only $0.19\%$, for the inclined course. This result quantifies that the dependence between $\rm C_{d}A$ and $\rm C_{rr}$, within adjustments of the model, is more pronounced for the flat course than for the inclined course, as expected in view of expression (\[eq:P\]), whose value—for the inclined course—is dominated by the first summand in the numerator, which includes neither $\rm C_{d}A$ nor $\rm C_{rr}$. This result provides a quantitative justification for the observation that the dependance of the accuracy of the estimate of power on the accuracies of $\rm C_{d}A$ and $\rm C_{rr}$ varies depending on the context; it is more pronounced on flat and fast courses. Similar evaluations can be performed using the values of derivatives contained in the second row of Table \[table:ModelRates\]. Therein, an increase in $\lambda$ results in a decrease of $\rm C_{d}A$, with different rates, for the flat and inclined courses. Physical considerations ----------------------- Physical inferences—based on minimization of expression (\[eq:model\])—are accurate in a neighbourhood of $\overline V_{\!\rightarrow}$ and $\overline P$, wherein the set of values for ${\rm C_{d}A}$, ${\rm C_{rr}}$ and $\lambda$ is estimated, since, as discussed in Section \[sec:MC\], these values—in spite of their distinct physical interpretations—are related among each other by the process of optimization of the model. In view of expression (\[eq:P\]), and as illustrated in Figure \[fig:FigPowerSpeed\], power as a function of ground speed is a cubic. The inflection point of the curve corresponds to the speed for which there is no air resistance, since the ground speed is equal to the tailwind, $V_{\!\rightarrow}=-w_{\leftarrow}$. At that point, $P$ is the power to overcome the rolling and drivetrain resistance, only. To the left of that point, the empirical adequacy of expression (\[eq:P\]) is questionable. However, for the results presented in this article, we consider the cases of $w\gg -V_{\!\rightarrow}$, which are well to the right of the inflection point. ![Power as a function of speed [*ceteris paribus*]{}[]{data-label="fig:FigPowerSpeed"}](FigPowerSpeed.pdf) The values in Table \[table:PhysicalRates\] are the changes of ground speed due to a change in power, mass, slope and wind; in each case, the other quantities are kept constant. These values allow us to answer such questions as what increase of speed would result from an increase of power by $1$ watt? To answer this question, let us examine the first row. For the flat course—considering the neighbourhood of $\overline V_{\!\rightarrow}=10.51$ and $\overline P=258.8$—$\partial_{P}V_{\!\rightarrow}=0.0176846$ and, in accordance with expression (\[eq:Thm\]), its reciprocal is $\partial_{V_{\!\rightarrow}}P=56.5462$. We write the corresponding differential as $${\rm d}P = \dfrac{\partial P}{\partial V_{\!\rightarrow}}\, {\rm d}V_{\!\rightarrow} = {56.5462}\,{\rm d}V_{\!\rightarrow} \,;$$ in other words, an increase of $V_{\!\rightarrow}$ by a unit requires an increase of $P$ by $56.5462$ units. This means that an increase of speed of $1$ metre per second requires an increase of power of $56.5462$ watts.[^5] For the inclined course—considering the neighbourhood of $\overline V_{\!\rightarrow}=4.138$ and $\overline P=286.6$—$\partial_{P}V_{\!\rightarrow}=0.0134356$ and, in accordance with expression (\[eq:Thm\]), its reciprocal is $\partial_{V_{\!\rightarrow}}P=74.4293$. Thus, an increase of speed of about $1$ metre per second requires an increase of power of about $74.4293$ watts. Hence, for the flat course, $$\label{eq:dVV} \left. \dfrac{{\rm d}V_{\!\rightarrow}}{V_{\!\rightarrow}} \right|_{\overline V_{\!\rightarrow}=10.51} = \dfrac{1}{10.51} \,,$$ which is a increase in speed of about $9.5\%$, requires $$\left. \dfrac{{\rm d}P}{P} \right|_{\overline P=258.8} = \dfrac{56.5462}{258.8} \,,$$ which is an increase in power of about $22\%$. For the inclined course, ${{\rm d}V/V}=1/{4.138}$ and ${\rm d}P/P={74.4293}/286.6$, which means that a $24\%$ increase in speed requires about $26\%$ increase in power. Remaining within a linear approximation, an increase of speed by $1\%$ requires an increase of power by about $2.3\%$, for the flat course, and an increase of only about $1.1\%$, for the inclined course. This result provides a quantitative justification for a time-trial adage of pushing on the uphills and recovering on the flats, to diminish the overall time. Since, as illustrated in Figure \[fig:FigPowerSpeed\], the slope of the tangent line changes along the curve, the value of expression (\[eq:Thm\]) corresponds to a given neighbourhood of pairs, $\overline V_{\!\rightarrow}$ and $\overline P$. Our interpretation is tantamount to comparing the slopes of two such curves—one corresponding to the model of the flat course and the other of the inclined course—at two distinct locations, $(\overline V_{\!\rightarrow},\overline P)=(10.51,258.8)$ and $(\overline V_{\!\rightarrow},\overline P)=(4.138,286.6)$. Even though the slope of the tangent line changes, it is positive for all values. This means that the function is monotonically increasing, even though it is a third degree polynomial. In other words, the relation of power and speed is a bijection, as illustrated in Figure \[fig:FigPowerSpeed\] and as discussed in Appendix \[sec:One-to-one\]. Let us examine the second row of Table \[table:PhysicalRates\]. For the flat course, $\partial_{m}V_{\!\rightarrow}=-0.00436803$ and its reciprocal is $\partial_{V_{\!\rightarrow}}m=-228.936$. This means that an increase of speed by $1$ metre per second—due only to the loss of mass—requires a decrease of mass of about $229$ kilograms. For the inclined course, $\partial_{m}V_{\!\rightarrow}=-0.0330937$, and its reciprocal is $\partial_{V_{\!\rightarrow}}m=-30.2172$; thus, an increase of speed by $1$ metre per second requires a decrease of mass of about $30$ kilograms. Thus, for the flat course, in accordance with expression (\[eq:dVV\]), an increase in speed by about $9.5\%$ requires ${\rm d}m/m=228.936/111$, which is a decrease of mass of about $206\%$. For the inclined course, ${{\rm d}V/V}=1/4.138$; hence, an increase in speed by about $24\%$ requires ${\rm d}m/m=30.2172/111$, which is a decrease of mass of about $27\%$. Remaining within a linear approximation, for the flat course, an increase of speed by $1\%$ requires a decrease of mass of about $22\%$, and for the inclined course, it requires a decrease of mass of only about $1\%$. This is supportive evidence of an empirical insight into the importance of lightness for climbing; in contrast to flat courses, in the hills, even a small loss of weight results in a noticeable advantage. Also, this result can be used to quantify the importance of the power-to-weight ratio, which plays an important role in climbing, but a lesser one on a flat. Similar evaluations can be performed using the values of derivatives contained in the third and fourth rows of Table \[table:PhysicalRates\]. In both cases, the sign is negative; hence, as expected, the increase of steepness or headwind results in a decrease of speed. These rates of decrease, which are different for the flat and inclined courses, can be quantified in a manner analogous to the one presented in this section. Discussion and conclusions ========================== The results derived herein are sources of information for optimizing the performance in a time trial under a variety of conditions, such as the strategy of the distribution of effort over the hilly and flat portions or headwind and tailwind sections. For instance, examining $\partial_{w_{\leftarrow}}V_{\!\rightarrow}$, for a flat course, we could quantify another time-trial adage of pushing against the headwind and recovering with the tailwind, to diminish the overall time, under a constraint of cyclist’s capacity; such a conclusion is illustrated in Appendix \[sec:Appendix\]. A further insight into this statement is provided by the following example. Let us consider a five-kilometre section against the headwind, $w_{\leftarrow}=5$, and, following a turnaround, the same five-kilometre section with the tailwind, $w_{\leftarrow}=-5$. If we keep a constant power, $P=258.8$, and use equation (\[eq:P\]) to find the corresponding speed, we achieve the total time of $946$ seconds, for ten kilometres, with the upwind speed of $V_{\!\rightarrow}=8.27286$ and the downwind speed of $V_{\!\rightarrow}=14.6269$. If we maintain the same average power, over ten kilometres, but increase the power on the upwind section by $10\%$ and decrease the power on the downwind section by $10\%$, we reduce the total time by about $16$ seconds, with the upwind speed of $V_{\!\rightarrow}=8.64701$ and the downwind speed of $V_{\!\rightarrow}=13.7839$. For reliable results—in view of Figure \[fig:FigPowerSpeed\] and the linear approximation within a neighbourhood of the average speed for which the flat-course model is established—one should not consider excessive increases or decreases of speed or power. To conclude this example, let us consider the case of keeping a constant speed, $V_{\!\rightarrow}=11.4256$, which is the average for the latter scenario, for ten kilometres. Such a strategy requires the power for the upwind section to be $P={531.557}\gg 258.8$. Thus, even though we should push harder against the wind than with the wind, we should not try to keep the same speed for both the upwind and downwind sections. This conclusion is consistent with the partial-derivative values of Table \[table:PhysicalRates\]. This conclusion is—only in part—consistent with a “Rule of Thumb” of @Anton2013. > First, recognize that the equal power outputs recipe, which would have you maintain the same pedal cadence and heart rate in headwind or tailwind, may feel optimal, but it actually isn’t. In fact, it is only barely faster than suffering the punishing swing in power-output that would be required to maintain equal out-and-back speeds. Your overall speed (and your finishing position, of course) will benefit from expending some extra energy when the wind is in your face and conserving some energy when the wind is at your back, but [*not too much*]{}, because going too far slows you down again as you approach the equal-speeds scenario. A quantification of this, and another, rule of thumb of @Anton2013 is presented in Appendix \[sec:Appendix\], where we question their generality. Also, results derived in this paper allow for a quantitative evaluation of the aerodynamic efficiency and—for team time trials—of the efficiency of drafting. Under various conditions, there are different relations between the rates of change of quantities in question. In this paper, as a consequence of the implicit function theorem, relations between the rates of change of all quantities that are included in a model are explicitly stated, and each relation can be evaluated for given conditions. Furthermore, the derived expressions allow us to interpret the obtained measurements in a quantitative manner, since the values of these expressions entail concrete issues to be addressed for a given bicycle course. The reliability of information—which depends on the accuracy of measurements and the empirical adequacy of a model—is quantified by a misfit and by standard deviations of model parameters. Also, using partial derivatives listed in Section \[sub:Partials\], we can write the differential of $P$, and, hence, estimate its error inherited from the errors of other quantities, $${\rm d}P = \partial_{m}P\,{\rm d}m +\,\cdots\,+ \partial_{\lambda}P\,{\rm d}\lambda\,,$$ where, in accordance with equation (\[eq:Thm\]) and in view of $\partial_{P}f=1$, $\partial_{m}P=-\partial_{m}f,\,\ldots\,,\partial_{\lambda}P=-\partial_{\lambda}f$. Time minimization with Lagrange multipliers {#sec:Appendix} =========================================== Preliminary remarks ------------------- Consider a flat course of length $d$, whose one half is covered against the wind, as illustrated in Figure \[fig:FigLagrange\], and the other half with the wind. ![Joseph-Louis (Giuseppe Luigi) Lagrange examining his optimizations based on the fact that, in general, in accordance with expression (\[eq:P\]), headwinds, $w_\leftarrow>0$, increase the air resistance, and tailwinds, $w_\leftarrow<0$, decrease it, provided that $w_{\leftarrow}\nless-V_{\!\rightarrow}$.[]{data-label="fig:FigLagrange"}](FigLagrange.pdf) To minimize the time, $t$, we need to maximize the average speed, $$\label{eq:AveV} \overline V_{\!\rightarrow} = \dfrac{d}{t} = \dfrac{d}{\dfrac{d}{2\,V_{U}}+\dfrac{d}{2\,V_{D}}} = \dfrac{2\,V_{U}V_{D}}{V_{U}+V_{D}} \,,$$ where $V_{U}$ and $V_{D}$ are the speeds on the upwind and downwind sections, respectively. The maximum of this function occurs for all values along $V_{U}=V_{D}$. To get a pair of values that corresponds to a realistic scenario, we invoke the method of Lagrange multipliers and find the maximum of speed , subject to constraints. To do so, we state the problem as a Lagrangian function of two variables with $n$ constraints, $$\label{eq:L_general} L(V_{U},V_{D}) = \overline V_{\!\!\rightarrow}(V_{U},V_{D}) + \Lambda_{1}\,\Gamma_{1}(V_{U},V_{D}) + \dots + \Lambda_{n}\,\Gamma_{n}(V_{U},V_{D}) \,,$$ where $\Lambda_i$, with $i=1,\ldots,n$, is a Lagrange multiplier. The optimization is achieved at the stationary points of function , which we find by solving the system of equations, $$\label{eq:LagPar} \dfrac{\partial L}{\partial V_{U}}=0 \,,\quad \dfrac{\partial L}{\partial V_{D}}=0 \,,\quad \dfrac{\partial L}{\partial\Lambda_{1}}= 0 \,,\quad\ldots\,,\quad \dfrac{\partial L}{\partial\Lambda_{n}}=0 \,,$$ whose solution is the pair, $V_{U},V_{D}$, that extremizes expression (\[eq:AveV\]) and satisfies the constraints, $\Gamma_i$, where $i=1,\ldots,n$, within the physical realm. Constraint of total work {#app:ConstraintWork} ------------------------ Let us impose a constraint in terms of the amount of total work, $W_{0}=W_{U}+W_{D}$, to be done by a cyclist on the upwind and downwind sections, whose proportions of length are stated in expression (\[eq:AveV\]), $$\label{eq:Gamma_{W}} \Gamma_{W} = \overbrace{ \underbrace{ \frac{ {\rm C_{rr}}\,m\,g + \tfrac{1}{2}\,{\rm C_{d}A}\,\overline\rho \left(V_{U}+w_{\leftarrow}\right)^{2} }{ 1-\lambda } }_{F_{\!\leftarrow}} \dfrac{d}{2} }^{W_{U}} + \overbrace{ \underbrace{ \frac{ {\rm C_{rr}}\,m\,g + \tfrac{1}{2}\,{\rm C_{d}A}\,\overline\rho \left(V_{D}-w_{\leftarrow}\right)^{2} }{ 1-\lambda } }_{F_{\!\leftarrow}} \dfrac{d}{2} }^{W_{D}}-W_{0} = 0 \,.$$ Herein, we assume $$W_{0} = \underbrace{ \frac{ {\rm C_{rr}}\,m\,g + \tfrac{1}{2}\,{\rm C_{d}A}\,\overline\rho\,{\overline V_{\!\rightarrow}}^{2} }{ 1-\lambda } }_{F_{\!\leftarrow}} d$$ to be the total amount of energy available to the cyclist, which corresponds to the work done on the same course, with a maximum effort—with no wind, $w_{\leftarrow}=0$—resulting in a given value of $\overline V_{\!\rightarrow}$. We write function  as $$\label{eq:L_{W}} L_{W} = \overline V_{\!\rightarrow} + \Lambda_{W}\,\Gamma_{W} \,.$$ Considering $d=10000$, model parameters stated in Section \[sec:FlatCourseVals\], namely, $m=111$, $g=9.81$, $\overline\rho=1.20406$, ${\rm C_{d}A}=0.2607$, ${\rm C_{rr}}=0.00231$, $\lambda=0.03574$, and letting $\overline V_{\!\rightarrow}=10.51$, we obtain $W_{0}=205878$. To minimize the traveltime with $w_{\leftarrow}=5$, we write system (\[eq:LagPar\]), in terms of function (\[eq:L\_[W]{}\]), $$\label{eq:L_{W}_system} \begin{dcases} \dfrac{\partial L_{W}}{\partial V_{U}} = \dfrac{2\,{V_{D}}^2}{\left(V_{D} + V_{U}\right)^{2}} + \Lambda_{W} \left(1627.66\,V_{U}+8138.32\right) = 0 \,, \\ \dfrac{\partial L_{W}}{\partial V_{D}} = \dfrac{2\,{V_{U}}^2}{\left(V_{D} + V_{U}\right)^{2}} + \Lambda_{W} \left(1627.66\,V_{D}-8138.32\right) = 0 \,, \\ \dfrac{\partial L_{W}}{\partial\Lambda_{W}} = 813.832\left({V_{U}}^2 + {V_{D}}^2\right) + 8138.32\left(V_{U}-V_{D}\right) - 139100 = 0 \,. \end{dcases}$$ Solving system  numerically, we obtain a single physical solution, $$\label{eq:L_{W}_system_sol} V_{U} = 8.27945 \quad{\rm and}\quad V_{D} = 11.6766$$ which is the pair that both maximizes expression (\[eq:AveV\]) and satisfies constraint (\[eq:Gamma\_[W]{}\]). In accordance with expression (\[eq:AveV\]), the average speed is $\overline V_{\!\rightarrow} = 9.68886$, which is lower than the speed under the assumption of $w_{\leftarrow}=0$, namely, $\overline V_{\!\rightarrow}=10.51$. This quantifies an adage that riding with the wind does not compensate for the speed lost by riding against the wind. The loss is due to the dissipation of energy due to the air, rolling and drivetrain resistances, which are present on both the upwind and downwind sections. Constraint of average power {#app:ConstraintPower} --------------------------- Let us impose a constraint in terms of the value of average power, $P_0$, maintained by a cyclist on the upwind and downwind sections. In contrast to work, power is not a cumulative quantity. Hence, the distance does not appear explicitly in a constraint, and we require constraints for both the upwind and downwind sections, $$\begin{aligned} \label{eq:Gamma_{P_{U}}} \Gamma_{P_{U}} &= \overbrace{ \underbrace{ \frac{ {\rm C_{rr}}\,m\,g + \tfrac{1}{2}\,{\rm C_{d}A}\,\overline\rho \left(V_{U}+w_{\leftarrow}\right)^{2} }{ 1-\lambda } }_{F_{\!\leftarrow}} V_{U} }^{P_{U}} - P_{0} \,, \\ \label{eq:Gamma_{P_{D}}} \Gamma_{P_{D}} &= \overbrace{ \underbrace{ \frac{ {\rm C_{rr}}\,m\,g + \tfrac{1}{2}\,{\rm C_{d}A}\,\overline\rho \left(V_{D}-w_{\leftarrow}\right)^{2} }{ 1-\lambda } }_{F_{\!\leftarrow}} V_{D} }^{P_{D}} - P_{0} \,.\end{aligned}$$ Herein, we assume $$P_{0} = \underbrace{ \frac{ {\rm C_{rr}}\,m\,g + \tfrac{1}{2}\,{\rm C_{d}A}\,\overline\rho\,{\overline V_{\rightarrow}}^{2} }{ 1-\lambda } }_{F_{\!\leftarrow}} \overline V_{\rightarrow}$$ to be the average power available to the cyclist, which corresponds to the average power achieved on the same course, with a maximum effort—with no wind, $w_{\leftarrow}=0$—resulting in a given value of $\overline V_{\!\rightarrow}$. Function  is $$\label{eq:L_{P}} L_{P} = \overline V_{\!\rightarrow} + \Lambda_{P_{U}}\,\Gamma_{P_{U}} + \Lambda_{P_{D}}\,\Gamma_{P_{D}} \,.$$ For $m=111$, $g=9.81$, $\overline\rho=1.20406$, ${\rm C_{d}A}=0.2607$, ${\rm C_{rr}}=0.00231$, $\lambda=0.03574$, $\overline V_{\!\rightarrow}=10.51$, we obtain $P_{0} = 216.378$. To minimize the traveltime with $w_{\leftarrow}=5$, we write system (\[eq:LagPar\]), in terms of function (\[eq:L\_[P]{}\]), $$ \begin{dcases} \dfrac{\partial L_{P}}{\partial V_{U}} = \dfrac{2\,{V_{D}}^2}{\left(V_{D} + V_{U}\right)^{2}} + \Lambda_{P_{U}} \left(0.488299\,{V_{U}}^2+3.25533\,V_{U}+6.67778\right) = 0 \,, \\ \dfrac{\partial L_{P}}{\partial V_{D}} = \dfrac{2\,{V_{U}}^2}{\left(V_{D} + V_{U}\right)^{2}} + \Lambda_{P_{D}} \left(0.488299\,{V_{D}}^2-3.25533\,V_{D}+6.67778\right) = 0 \,, \\ \dfrac{\partial L_{P}}{\partial\Lambda_{P_{U}}} = 0.162766\,{V_{U}}^{3} + 1.62766\,{V_{U}}^{2} + 6.67778\,V_{U} - 216.378 = 0 \,, \\ \dfrac{\partial L_{P}}{\partial\Lambda_{P_{D}}} = 0.162766\,{V_{U}}^{3} - 1.62766\,{V_{U}}^{2} + 6.67778\,V_{U} - 216.378 = 0 \,. \end{dcases}$$ The single physical solution is $$\label{eq:L_{P}_system_sol} V_{U} = 7.60272 \quad{\rm and}\quad V_{D} = 13.9163 \,,$$ which both maximizes expression  and satisfies constraints  and . The corresponding average is $\overline V_{\!\!\rightarrow} = 9.83332$, which confirms an adage that riding with the wind does not compensate for the speed lost by riding against the wind. Relation between differences and derivatives -------------------------------------------- To conclude this appendix, let us comment on difference $\Delta V_{\!\rightarrow}/\Delta w_{\leftarrow}$, discussed herein, in the context of $\partial_{w_{\leftarrow}}V_{\!\rightarrow}$, whose value is presented in Table \[table:PhysicalRates\]. Partial derivatives correspond to a tangent to a curve at a point, and the differences to a secant over a segment of the curve. Also, partial derivatives are obtained under the assumption that all other quantities are constant. The latter requirement is satisfied in Appendix \[app:ConstraintPower\], where $$\frac{\Delta V_{\rightarrow}}{\Delta w_{\leftarrow}} = \frac{V_{U}-V_{D}}{w_{\leftarrow}-(-w_{\leftarrow})} = \frac{7.60272-13.9163}{10} = -0.631356\,,$$ which agrees with $\partial_{w_{\leftarrow}}V_{\rightarrow}$, in Table \[table:PhysicalRates\], to two decimal points. For $w_{\leftarrow}=0.05$, we obtain $V_{U} = 10.4782$ and $V_{D} = 10.5418$; hence, $\Delta V_{\rightarrow}/\Delta w_{\leftarrow}=-0.635911$, which agrees with $\partial_{w_{\leftarrow}}V_{\rightarrow}$ to six decimal points. In general, $$\lim_{\Delta w_{\leftarrow}\rightarrow0} \frac{\Delta V_{\rightarrow}}{\Delta w_{\leftarrow}} = \frac{\partial V_{\rightarrow}}{\partial w_{\leftarrow}} \,,$$ as expected, in view of a secant approaching a tangent. The requirement of constant quantities is not satisfied in Appendix \[app:ConstraintWork\], since $P$ is allowed to vary to maintain the imposed value of $W_{0}$. In Appendix \[app:ConstraintPower\], $W$ varies to maintain the imposed value of $P_0$, but $W$ is not a variable in function (\[eq:f\]), used in partial derivatives. As shown in this appendix, properties of partial derivatives need to be considered in examining time-trial strategies. In contrast to common optimization methods, partial derivatives correspond to a change of a single variable, only. Closing remarks --------------- Let us examine the constraints discussed in this appendix in terms of required powers. For the work constraint, the average speed is $\overline V_{\!\rightarrow}= 9.68886$. Following expression (\[eq:P\]), the required powers are $P_{U}=365.537$ and $P_{D}=59.9459$, for the upwind and downwind sections, respectively. For windless conditions, we have $\overline{P} =173.316$. Thus, $P_{U}$ is significantly greater than $\overline P$. For the power constraint, with $P_{0}=216.378$, the average speed is $\overline V_{\!\rightarrow}= 9.83332$. Since the average speed is greater than for the work-constraint optimization and the average power does not exceed the value obtained in windless conditions, this appears to be the preferable strategy. Also, power is provided as an instantaneous quantity by the power meters, which allows the rider to follow a given strategy, whose further refinements are to be considered in future studies. To close, let us consider a rule of thumb of @Anton2013. > Choose a target-speed $v_0$. $\,[\ldots]\,$ Endeavor to ride at $v\cong v_0+w/4$ when the wind is at your back and at $v\cong v_0-w/2$ when the wind is at your face. If we choose $\overline V_{\!\rightarrow}=10.51=:v_0$ to be a target speed, with $w=5$, speeds , which result from the power constraint, are less congruent with this rule than speeds , which result from the work constraint, yet—according to the present analysis—speeds  appear to be preferable. This is an indication of further subtleties that need to be considered in developing a time-trial strategy. One-to-one relation between power and speed {#sec:One-to-one} =========================================== \[prop:one-to-one\] According to model (\[eq:P\]), with $a=0$ and $\eta=1$, the relation between the measured power, $P$, and the bicycle speed, $V_{\!\rightarrow}$, is one-to-one. It suffices to show that $\partial P/\partial V_{\!\rightarrow}>0$, for $V_{\!\rightarrow}\in(0,\infty)$. Since $$\dfrac{\partial P}{\partial V_{\!\rightarrow}} =\frac{({\rm C_{rr}}\cos\theta+\sin\theta)\,g\,m+\frac{3}{2}{\rm C_dA}\,\rho\,V_{\!\rightarrow}^{\,2}}{1-\lambda}\,,$$ where all quantities are positive and $\lambda\ll 1$, it follows that $\partial P/\partial V_{\!\rightarrow}>0$ and, hence, the relation between power and speed is one-to-one. This bijection means that power and speed are related by an invertible function, which is consistent with unique physical solutions obtained in Appendices \[app:ConstraintWork\] and \[app:ConstraintPower\]. A consequence of Proposition \[prop:one-to-one\] is that $\partial^2P/\partial V_{\!\rightarrow}^{\,2}>0$; hence—in contrast to Figure \[fig:FigPowerSpeed\], where $\eta=\pm 1$—the corresponding curve is concave up for $V_{\!\rightarrow}\in(0,\infty)$. Another consequence is that—[*ceteris paribus*]{}—the increase of speed requires increase of power, and an increase of power results in an increase of speed. As illustrated in Figure \[fig:FigPowerSpeed\], this remains true even for $\eta=-1$, and can be viewed as a physical law for bicycling. Acknowledgements {#acknowledgements .unnumbered} ================ We wish to acknowledge Len Bos, for fruitful discussions, David Dalton, for his scientific editing and proofreading, Elena Patarini, for her graphic support, and Roberto Lauciello, for his artistic contribution. Furthermore, we wish to acknowledge Favero Electronics for inspiring this study by their technological advances and for supporting this work by providing us with their latest model of Assioma Duo power meters. Conflict of Interest {#conflict-of-interest .unnumbered} ==================== The authors declare that they have no conflict of interest. [^1]: AGH–University of Science and Technology, Kraków, Poland, `[email protected]` [^2]: Memorial University of Newfoundland, Canada, `[email protected]` [^3]: Memorial University of Newfoundland, Canada, `[email protected]` [^4]: For consistency with power meters, whose measurements are expressed in watts, which are $\rm{kg\,m^2/s^3}$, we use the [*SI*]{} units for all quantities. Mass is given in kilograms, $\rm{kg}$, length in metres, $\rm{m}$, and time in seconds, $\rm{s}$; hence, speed is in metres per second, change of speed in metres per second squared, force in newtons, $\rm{kg\,m/s^2}$, and work and energy in joules, $\rm{kg\,m^{2}/s^{2}}$; angle is in radians. [^5]: $\partial P/\partial V$ can be also found by differentiating expression (\[eq:P\]) with respect to $V_{\!\rightarrow}$. However, the implicit function theorem allows us to obtain relations between quantities, without explicitly expressing one in terms of the other. Also, in accordance with the inverse function theorem, $\partial V/\partial P=1/(\partial P/\partial V)$, which is justified by the fact that, for expression (\[eq:P\]), $\partial P/\partial V_{\!\rightarrow}\neq 0$, in the neighbourhood of interest, as required by the theorem. However, expression (\[eq:Thm\]), which states the implicit function theorem, provides a convenience of examining the relations between the rates of change of any two quantities without invoking the inverse function theorem and requiring an explicit expression for either of them.
{ "pile_set_name": "ArXiv" }
--- author: - 'Mohammadreza Doostmohammadian,  Hamid R. Rabiee,  Houman Zarrabi,  and Usman Khan, [^1]' bibliography: - 'bibliography.bib' --- [Doostmohammadian : Observational Equivalence in System Estimation: Contractions in Complex Networks]{} networks have recently gained considerable attentions in control and estimation theory [@liu_pnas; @ruths2014control; @liu2016tutorial; @das2015distributed; @asilomar14; @Liu_nature; @barabasi2016social]. This interest stems from the challenge to understand and infer the fundamental aspects of system behavior. Such complex networks exist in nature for example in chemical reaction networks and biological networks [@liu_pnas] as in proteomics and gene networks. Other than these natural complex networks, synthetic large-scale networks are recently considered due to emergence of the so-called Internet-of-Things (IoT) and Cyber-Physical-Systems (CPS) [@das2015distributed; @asilomar14]. Interestingly the design of such man-made networks are significantly tied by control and estimation principles as they are genuinely constructed based on these principles [@liu2016tutorial]. Examples range from consensus networks [@scientia] and social networks [@barabasi2016social] to more technological networks including electric power grids, computer networks, the Internet, etc. Indeed, many networks are a formalism to describe phenomena and systems in real life[^2]. In these and other similar applications the research focus is to uncover the tie between the internal system/network dynamics and the controllability and estimation properties. It is known that, the internal states of complex systems are to a great extent dependent on each other, which is due to interaction of different components on each other and therefore these complex systems are represented as networks [@liu_pnas]. This inter-dependence is such that by measuring and tracking certain variables of complex system one can infer sufficient information about the rest of the system for filtering and tracking purposes. This implies that measuring well-selected variables give an *observable* inference of complex system. The term observability is a measure defining whether the internal states of a system can be determined by knowledge of its measurements. The system is said to be observable if one can reconstruct the *complete* state of the complex system from the set of measured states also known as system outputs [@bay]. There are different methods to check for observability of dynamic systems, namely: (i) algebraic method based on Gramian test [@bay]; (ii) the symbolic method also known as Popov-Belevitch-Hautus (PBH) test [@hautus]; and, (iii) the structural observability method introduced by pioneering work of Lin [@lin]. The first two methods are based on numerical values of system parameters while the third method is irrespective of the parameters and only relies on the *structure* of the underlying system. Therefore, the third method has certain benefits over the two other methods as it is computationally efficient and only requires the structural information and sparsity pattern of the system instead of exact numerical values [@woude:03]. In other words, the structural method only relies on the knowledge of complex system as a graph/network and therefore is extensively studied in the literature [@lin; @liu_pnas; @ruths2014control; @liu2016tutorial; @asilomar14; @Liu_nature; @woude:03]. The system graph representation, referred to as *system graph*, is an abstract way of modeling complex systems and has recently applied widely in the literature to reduce the complexity of such systems [@liu_pnas; @asilomar14; @Liu_nature]. In the system graph, every graph node represents a state (a variable or a parameter) and every link (or edge) between a pair of nodes represents derivative functional connection relating the state variables [@liu_pnas; @Liu_nature; @woude:03]. In general, graph representation approach is more conventional in networked systems [@acc13_mesbahi; @jstsp14] where the network structure is embedded into the system structure. The structural observability of networks, or any system graph as a network, is also referred to as *network observability* [@liu_pnas; @Liu_nature] and is the adopted methodology in this paper.[^3] The network observability, as an abstract observability model of the system is closely related to system graph properties.[^4] The main theorem on this topic is originally stated in [@lin] and further developed recently in [@liu_pnas; @Liu_nature; @asilomar11; @commault_recovery; @globalsip14; @boukhobza_recovery; @jstsp14][^5]. Existence of disjoint cycles and output connected paths in the graph is closely tied with its observability. In this direction, recently the concept of matching and dilations in graph [@Liu_nature], and Strongly Connected Components (SCC) [@asilomar11] are introduced to be related to network observability/controllability. Among these, the concept of contraction is the focus of this paper. An introductory description of contraction in the network is the set of nodes contracting (linking) to fewer number of nodes. In system estimation perspective, nodes in the same contraction are observationally equivalent, i.e. in case of losing observability of one (unmatched) node/state in the network/system another node in the same contraction can be measured to recover for the loss of observability. This is applicable in estimation of large scale systems such as power grid [@asilomar14] and internet based autonomous systems. For example when a sensor fails to measure a state –due to excessive noise, disturbance, or even external attacks– some necessary information of the system is lost and system/network cannot be tracked globally. To recover for this loss of information another sensor can be applied to measure equivalent state/node of the system/network. This is why the contractions play an important role in estimation. Indeed, one can apply a new sensor to measure an equivalent state in the same contraction and recover for the observability loss. In this regard, the size of contraction determines the possible number of equivalent options for observability recovery. Larger contractions imply more options among which one can choose the most efficient state measurement in terms of cost [@IJSS2017], reliability, etc. This is the main motivation to analyze the size and distribution of contractions as they play a major role in system observability recovery. *Related Literature:* Structural observability of full-rank systems (having no contraction in system graph) is considered in [@liu_pnas; @asilomar11]. In these works, structural observability is shown to be closely related to network SCC classification. In [@Liu_nature] using cavity method the authors find considerable relation between average network degree and number of unmatched nodes. As one of their main results, they find that denser networks have less number of unmatched nodes and therefore it is less challenging to control and direct the network to the desired state. In [@globalsip14] the authors consider *distributed* estimation and formulate necessary and sufficient conditions for distributed structural observability. This work finds the connection between the structure of complex system and the structure of monitoring sensor network. In [@commault_recovery; @boukhobza_recovery], the authors classify sensors based on their essentiality for observability using combinatorial algorithms with application to sensor failure and diagnosis. Among these and other literature, what missing is on the concept of contraction and the relation between distribution of contractions and properties of the network (or system graph). *Contribution:* In this paper, we study the properties of contractions in undirected networks/system-graphs as a key factor in estimation and observability. Adopting the structural/network observability method, the related question addressed here is that: how to find the equivalent state nodes in the network/system-graph to infer observationally equivalent information of the associated system? and we show that by finding contractions in the system-graph (or network), one can find the system states (or network nodes) equivalent in terms of observability and estimation. In this regard, the size of a contraction determines the potential number of *equivalent* sensing locations in networks as model of complex systems, which is discussed in Section \[sec\_cont\]. Further, a *polynomial order* algorithm is applied to find the contractions in the system graph. This algorithm is a modification of the algorithm for unmatched node detection given in [@murota]. Contractions are of particular interest in recovering sensor failure and loss of observability in tracking/filtering noise-corrupted global state of the system/network. Detailed discussion on application of contractions in system estimation and observability including example of observability recovery in Kalman filtering is provided in Section \[sec\_obsrv\]. Introducing the contraction set, the follow-up question is: how do the properties of these contraction sets change based on different characteristics of the underlying network? We investigate the effect of two factors on the size and distribution of contraction components: degree heterogeneity and clustering coefficient. First result of this paper is that the clustering coefficient as a network characteristic is related to average size and number of contractions. In particular, our results show that for Scale-Free networks, with power-law degree distribution, increase in clustering coefficient results in a decrease in average contraction size in the network. Further, we observe decrease in the number of contractions in high clustering coefficient Scale-Free networks. As the next contribution, we check the effect of degree heterogeneity in Small-World networks on contraction properties. Specifically, our results show that increase in randomness of link connectivity (tuning the $p$ factor) results in decrease in the average contraction size but increase in the number of contractions in the network. These results are addressed in Section \[sec\_rand\]. Further in Section \[sec\_real\], as a practical contribution, the contraction properties including the size distribution and prevalence are discussed for two real world networks: a Power-grid network and a Route-view network. Noting that the degree distribution of many real-world networks show power-law degree distribution, including the two example here, the results for these two practical examples corresponds with contraction properties of scale-free networks. More detailed discussion on these results and concluding remarks are stated in Section \[sec\_conc\]. It should be noted that in this paper the results are particularly stated for *undirected* networks/system graphs. Notions on Graph Theory and Definition of Contraction {#sec_cont} ===================================================== In this section we define the contraction sets in graphs by first introducing the relevant graph theoretic notions. Define a graph as $\mc{G}=(\mc{V},\mc{E})$, where $\mc{V}$ is the node set containing $n$ graph nodes, and $\mc{E}=\{(v_i,v_j)\}$ is the set of edges connecting the nodes. Define a path as a sequence of distinct nodes with every consecutive nodes as an edge in $\mc{E}$. Further, define a cycle as a path starting and ending at the same node. Define $\mc{N}(i)$ as the degree of node $i$. The adjacency matrix of the graph $A_G=\{a_{ij}\}$ is defined as $a_{ij}=1$, if $(v_j,v_i) \in \mc{E}$, otherwise $a_{ij}=0$. We further introduce the following graph-theoretic concepts to define contractions: - *Bipartite graphs:* Define a bipartite graph, $\Gamma=(\mc{V}^+,\mc{V}^-,\mc{E}_\Gamma)$, such that its nodes are partitioned into two disjoint sets: $\mc{V}^+$ and $\mc{V}^-$, and all of its edges $\mc{E}_\Gamma$ start in $\mc{V}^+$ and end in $\mc{V}^-$. We construct a bipartite graph, $\Gamma$, from  $\mc{G}$ with the edge set $\mc{E}_{\Gamma}$, defined as the collection of $(v_j^-,v_i^+)$, if  $(v_j,v_i) \in \mc{E}$.[^6] - *Matching:* A matching, $\underline{\mc{M}}$, on the system graph, $\mc{G}$, is defined as a subset of the edge set, $\mc{E}$, with no common end-nodes. In the bipartite graph, $\Gamma$, it is defined as a subset of edges where no two of them are incident on the same vertex in $\mc{V}^+$, i.e. all the edges in $\mc{M}$ are all disjoint. The number of edges, $|\underline{\mc{M}}|$, is the size of the matching. A matching, $\underline{\mc{M}}$, with maximum size is called maximum matching, denoted by $\mc{M}$, which is non-unique in general. - *Matched/Unmatched nodes:* Let $\mc{M}$ be a maximum matching defined on the bipartite graph, $\Gamma$. Let $\partial \mc{M}^+$ and $\partial \mc{M}^-$ denote the nodes incident to $\mc{M}$ in $\mc{V}^+$ and $\mc{V}^-$ respectively. Denote by $\delta \mc{M}$ the set of unmatched nodes in $\mc{V}^+$ as $\delta \mc{M} = \mc{V}^+ \backslash \partial \mc{M}^+$. Note that maximum matching $\mc{M}$ is not unique in general. - *Auxiliary graph*, denoted by $\Gamma^\mc{M}$, is a graph associated to a maximum matching, $\mc{M}$. It is constructed by reversing all the edges of maximum matching, $\mc{M}$, and keeping the direction of all other edges $\mc{E}_{\Gamma} \backslash \mc{M}$, in the bipartite graph, $\Gamma$. This graph is defined to localize the contractions in the system graph. - *${\mc{M}}$-alternating path:* In the auxiliary graph, define an ${\mc{M}}$-alternating path as a sequence of edges starting from an unmatched node in $\delta \mc{M}$ and every second edge in $\mc{M}$, and denote it by $\mc{Q}_{\mc{M}}$. The name comes from the alternating edges between unmatched part, $\mc{E} \backslash \mc{M}$, and matched part, $\mc{M}$, of the auxiliary graph. - *${\mc{M}}$-augmenting path:* In the auxiliary graph, define an ${\mc{M}}$-augmenting path, denoted by $\mc{P}_{\mc{M}}$, as an ${\mc{M}}$-alternating path with begin node and end node in $\delta \mc{M}$. Having defined these preliminary notions on graph theory, the notion of a contraction set is defined as follows: In the auxiliary graph representation of a network, $\Gamma^\mc{M} _A$, define a contraction set for every unmatched node $v_j \in \delta \mc{M}$, as the set of nodes containing all states in $\mc{V}^+$ reachable by ${\mc{M}}$-alternating paths starting from $v_j$. Denote this set by $\mc{C}_i$ and further define $\mc{C}$ as the set of all contractions, i.e. $\mc{C}=\{\mc{C}_1,...,\mc{C}_m\}$. Intuitively, in graph $\mc{G}$, a contraction set defines nodes that are connected (contracted) to less number of nodes.[^7] ![This figure shows a network contraction in the left, where the three contraction nodes are shown in red color. Figures in the right show three different maximum matching in bipartite representations of the same contraction. The red edges represent maximum matching and the red node represents the unmatched node.[]{data-label="fig_3nodecont"}](fig_3nodecont.pdf){width="3.5in"} ![This figure illustrates the procedure of finding contractions explained in the paper. Graph in the left shows one possible matching and the unmatched node in the bipartite representation of the graph in Fig \[fig\_3nodecont\]. The middle graph shows the auxiliary representation, where all matching edges are reversed. In the right graph an ${\mc{M}}$-alternating path is shown in black. Starting from the unmatched node, this path is used to find the contraction nodes (shown by dashed squares). Later in this paper we name these contraction nodes as observationally equivalent nodes.[]{data-label="fig_3nodecont2"}](fig_3nodecont2.pdf){width="2.7in"} For better illustration of the above definitions a contraction of 3 nodes into 2 nodes is shown in Fig. \[fig\_3nodecont\]. The bipartite representation, the maximum matching, and the unmatched node are illustrated in the figure. We further illustrate the definition of auxiliary graph and ${\mc{M}}$-alternating path in Fig. \[fig\_3nodecont2\]. The algorithm to find the contraction sets in network is given in Algorithm \[alg\_cont\]. \[alg\_cont\] **Given:** System graph $\mc{G}_A$ Construct the bipartite graph $\Gamma=(\mc{V}^+,\mc{V}^-,\mc{E}_\Gamma)$ Find a matching $\underline{\mc{M}}$ as the set of edges with no common end nodes Construct the auxiliary graph $\Gamma^{\underline{\mc{M}}}_A$ by reversing the edges in matching $\underline{\mc{M}}$ Define $\partial {\underline{\mc{M}}}^+$ as the nodes in $\mc{V}^+$ incident to ${\underline{\mc{M}}}$ Define the set of unmatched nodes $\delta{\underline{\mc{M}}}$ as $\delta {\underline{\mc{M}}} = \mc{V}^+ \backslash \partial {\underline{\mc{M}}}^+$ Define the ${\mc{M}}$-alternating path, $\mc{Q}_{\underline{\mc{M}}}$, as a sequence of edges starting from an unmatched node in $\delta {\underline{\mc{M}}}$ and every second edge in ${\underline{\mc{M}}}$ Define an ${\mc{M}}$-augmenting path, $\mc{P}_{\underline{\mc{M}}}$, as an ${\mc{M}}$-alternating path with begin node and end node in $\delta {\underline{\mc{M}}}$ Construct the auxiliary graph $\Gamma^{\mc{M}}_A$ for the maximum matching ${\mc{M}}$ Define  $\partial {\mc{M}}^+$ for the maximum matching ${\mc{M}}$ and define $\delta {\mc{M}} = \mc{V}^+ \backslash \partial \mc{M}^+$ Define the ${\mc{M}}$-alternating path, $\mc{Q}_{\mc{M}}$, for the maximum matching ${\mc{M}}$ **Return** $\mc{C}_i, i = \{1,...,m\}$ In Algorithm \[alg\_cont\], $\oplus$ is the XOR operator in set theory. As a result of this operator, each augmenting path increases the size of the matching till it reaches the maximum matching. The computational complexity of this algorithm is on the order of $\mc{O}(\sqrt{n} |\mc{E}|)$ or $\mc{O}(n^{\frac{5}{2}})$ in worst case. In general, given the system graph $\mc{G}_A$ there are other efficient algorithms to compute the maximum matching, $\mc{M}$, e.g., the maximum flow algorithm [@hopcraft]. The notions $\Gamma,\mc{M}$ can be obtained by the Dulmage-Mendelsohn (DM) decomposition [@dulmage58]. Other than DM decomposition, maximum matchings can be efficiently computed in $\mc{O}(\sqrt{n} |\mc{E}_A|)$ using the approach in [@maxmatching]. In the following, we state two main lemmas relating the maximum matchings and contractions. \[lem\_unmatched\] Any choice of maximum matching, $\mc{M}$, includes one and only one unmatched node in every contraction $\mc{C}_i, i\in \{1,...,m\}$. The detailed proof is provided in [@murota; @berge]. \[lem\_contr\] For two sets of maximum matching, $\mc{M}_1 \neq \mc{M}_2$, any unmatched node $v_i \in \delta \mc{M}_1$ can be reached along an alternating path from a node $v_j \in \delta \mc{M}_2$. This further implies that the set $\mc{C}$ is the same for any choice of maximum matching. The proof is given in [@murota]. Application in Observability and System Estimation {#sec_obsrv} ================================================== In this section, we first discuss the concept of structural observability in networks and then its application to system estimation. To further illustrate the results a network estimation example is provided. Network Observability {#subsec_netobsrv} --------------------- Observability of networks quantifies whether given measurements contain sufficient information to comprehensively reconstruct the states of all nodes in the network. For a network, or a system graph representing a complex system, the necessary and sufficient conditions for (structural) observability is given in the following theorem. \[thm\_obsrv\] A network (or system graph) is structurally observable if and only if: (i) every node can reach to an output/measurement via a path of state nodes, and (ii) there exist a family of disjoint cycles and output-connected paths covering all nodes. The proof is given in [@lin] and in [@rein_book] for the dual case of structural controllability. In Theorem \[thm\_obsrv\], condition (i) is known as *accessibility* and condition (ii) as the *S-rank* condition. Note that for connected undirected networks the accessibility is already satisfied. This is because, in a connected undirected network every node is reachable by every other node and therefore output connectivity of one node implies the reachability of all other nodes to that output. \[thm\_unmatched\] In a connected undirected network with the set of unmatched nodes, $\delta \mc{M}$, observation/measurement of every unmatched node is necessary and sufficient for network observability. The proof is given in [@Liu_nature] for the dual case of network controllability. Following the definition of contraction and results in previous section here we state the theorem on the concept of observational equivalence in contractions. \[thm\_contr\] In a connected undirected network with the set of contractions $\mc{C}=\{\mc{C}_1,...,\mc{C}_m\}$, a measurement/observation of one state node in every contraction $\mc{C}_i, i\in \{1,...,m\}$ is necessary and sufficient for network observability. From Theorem \[thm\_unmatched\] observation of every unmatched node is necessary and sufficient for network observability. Note that based on Lemma \[lem\_unmatched\] for every contraction $\mc{C}_i$, every node $v_j$ is an unmatched node for a choice of maximum matching $\mc{M}$. This implies that observing at least one node in every contraction is necessary for observability. Further, by measuring node $v_j$ in $\mc{C}_i$ from Lemma \[lem\_contr\] all other nodes in $\mc{C}_i \backslash v_j$ are matched for the choice of maximum matching $\mc{M}$ and therefore only one node is sufficient for network observability. \[lem\_rankdeficiency\] Number of contractions in a network $\mc{G}$ equals the (structural) rank deficiency of its associated adjacency matrix, $A_G$. Indeed the rank deficiency of the adjacency matrix, $A_G$, equals the number of unmatched nodes in the network $\mc{G}$. Indeed from Lemma \[lem\_unmatched\] the number of contractions equals the number of unmatched nodes in the network. Note that, by definition, unmatched nodes appear on acyclic part of network while the cyclic part is completely matched. It is known that the rank of the network adjacency matrix is structurally defined by the number of nodes included in a set of disjoint cycles [@godsil]. This implies that the rank deficiency can be structurally defined by the number of unmatched nodes, which are contained in the acyclic part of the network. The concept of contraction is closely related to the concept of *observational equivalence*. Let $C_i$ denote the measurement matrix of node/state $x^i$. Let $\mc{O}(A,C_i)$ represent the observability Grammian of the pair $A$ and $C_i$. The observational equivalence relation among two states/nodes $x^i$ and $x^j$, denoted by $x^i \sim x^j$, is defined as: $$\begin{aligned} \mbox{rank}~\mc{O}(A, C_i) = \mbox{rank}~\mc{O}(A, C_j) = \mbox{rank}~\mc{O}\left(A, \left[ \begin{array}{c} C_i\\ C_j \end{array} \right] \right)\end{aligned}$$ Note that this follows the three properties of the equivalence relation, i.e. transitivity, reflexivity, and symmetry.[^8] \[lem\_equiv\] The algebraic implication of observational equivalence relation among states in each contraction is defined as follows. For any two (or more) measurements of states $x^i$ and $x^j$ in the same contraction, the structural-rank recovery of system matrix $A$ is equal to $1$, i.e., $$\begin{aligned} \label{eq_srank} \mbox{S-rank}\left( \left[ \begin{array}{c} A \\ C_{i} \end{array} \right] \right) &= & \mbox{S-rank}\left( \left[ \begin{array}{c} A \\ C_{j} \end{array} \right] \right) \nonumber \\ &=& \mbox{S-rank}\left( \left[ \begin{array}{c} A \\ C_{i} \\ C_{j} \end{array} \right] \right) \nonumber \\ &=& \mbox{S-rank}(A)+1. \end{aligned}$$ where S-rank implies the structural-rank[^9] of the matrix. The proof directly follows the three properties of observational equivalence relation. One can easily check that the reflexivity, symmetry, and transitivity of equation directly follows. \[cor\_equiv\] Theorem \[thm\_contr\] along with Lemma \[lem\_unmatched\], \[lem\_contr\], and \[lem\_equiv\] imply that all nodes in a contraction are *equivalent* in terms of observability. In other words, measurement of *any* node in each contraction, assuming that all other contractions have one observation, provides network observability. As a result of the equivalent observability relation, nodes in the same contraction recover loss of observability. In other words, in the case of observation failure of a node, say node $v_i$, some information of the system is lost. In this case, observation of another node, say node $v_j$, sharing a contraction with node $v_i$ recovers the observability loss. In this regard, the *size of contraction* defines the possible number of equivalent sensing nodes for recovering observability loss.The implication of equivalent relation is further discussed in next subsection; we show how dynamic systems can be represented structurally as networks, where we can apply the above Theorems and Lemmas to find equivalent states in terms of system observability and estimation. System Estimation ----------------- Consider the system model to be a discrete-time linear dynamic system[^10]: $$\begin{aligned} \label{sys1} \mb{x}_{k+1} &=& A\mb{x}_k + \mb{v}_k,\end{aligned}$$ with  $\mb{x}_k\in\mathbb{R}^n$ $$\begin{aligned} \nonumber \mb{x}_k = \left( \begin{array}{c} x_k^1\\ \vdots\\ x_k^n \end{array} \right)\end{aligned}$$ as the state vector,  $A=\{a_{ij}\}\in\mathbb{R}^{n\times n}$ as the system matrix, and $\mb{v}_k\sim\mathcal{N}(0,V)$ as the system noise. Assume the dynamical system to be monitored by measurement/observation model: $$\begin{aligned} \label{sys2} \mb{y}_k = C\mb{x}_k + \mb{r}_k,\end{aligned}$$ where $$\begin{aligned} \nonumber \mb{y}_k = \left( \begin{array}{c} y_k^1\\ \vdots\\ y_k^m \end{array} \right),~~ C = \left( \begin{array}{c} C_1\\ \vdots\\ C_m \end{array} \right),~~ \mb{r}_k = \left( \begin{array}{c} r_k^1\\ \vdots\\ r_k^m \end{array} \right),\end{aligned}$$ Here, $\mb{r}_k\sim\mathcal{N}(0,R)$ is the observation noise with $R = \mbox{blockdiag}[R_1,\ldots,R_N]$, and $C$ is the measurement matrix. In structured systems theory, the LTI system in Eqs. -, can be modeled as a *system graph*. In this scenario, every node is a system state and every edge represents the interaction of two states based on the system matrix, $A$. Denote the set of system states by $\mc{X}\triangleq\{x^1,\ldots,x^n\}$ and denote the set of system observations/measurements by $\mc{Y}\triangleq\{y^1,\ldots,y^m\}$. Then the system graph is defined by $\mc{G}_A=(\mc{X},\mc{E}_A)$ where the edge set, $\mc{E}_A$, is defined as $\mc{E}_A=\{(x^i,x^j)~|~a_{ji}\neq0\}$, to be interpreted as $x^i\rightarrow x^j$. One should note that, in this graph representation of system the structure of system graph only relies on free parametric entries of matrix $A$. In other words, the graph structure depends on each entry $a_{ij}$ being a free parameter and not on the exact numerical value of $a_{ij}$. Therefore, any Linear Structure Invariant (LSI) system with fixed structure and time-varying parameters can be modeled as a system graph.[^11] The motivation of applying graph representation of system is that one can check its characteristics by using equivalent graphical properties. The system characteristic of interest here is system observability, which plays a crucial role in system estimation and filtering. To illustrate more we consider the role of system graph observability in Kalman estimation as discussed next. Let $\widehat{\mb{x}}_{k|k}$ be the Kalman estimator tracking system state $\mb{x}_k$ at time $k$ given all the measurements, $\mb{y}_k$, up to time $k$. The dynamics of this estimator is defined as follows [@kalman:61]: $$\begin{aligned} \label{eq_kalman1} \widehat{\mb{x}}_{k|k-1} &=& A\widehat{\mb{x}}_{k-1|k-1} \\ \label{eq_kalman2} \widehat{\mb{x}}_{k|k} &=&\widehat{\mb{x}}_{k|k-1} + K_k C^T (\mb{y}_k-C\widehat{\mb{x}}_{k|k-1})\end{aligned}$$ where the $K_k$ is the Kalman gain computed in a recursive procedure as proposed by Kalman [@kalman:61]. It can be shown that the error, $\widehat{\mb{e}}_{k|k} = \mb{x}_k - \widehat{\mb{x}}_{k|k}$, in this estimator is given by, $$\begin{aligned} \label{ge} \widehat{\mb{e}}_{k|k} = (A - K_kC^TCA)\widehat{\mb{e}}_{k-1|k-1} + \eta_k,\end{aligned}$$ where the vector $\eta_k$ collects the remaining terms (noise terms) that are independent of $\widehat{\mb{e}}_{k-1|k-1}$ and $\widehat{\mb{e}}_{k|k}$. It is known that the dynamics of Kalman error, $\widehat{\mb{e}}_{k|k}$ is stable if the measurements defined by matrix $C$ give an observable inference of system defined by $A$ [@kalman:61]. In other words, the Mean Squared Estimation Error (MSEE) reaches bounded stability over time if the pair $(A,C)$ is observable. Following the results of the graph-theoretic method in Section \[subsec\_netobsrv\], we consider two applications in the following. (i) As the first application one can check the observability constraint using results of Theorem \[thm\_contr\]. For a system to be observable, according to Theorem \[thm\_contr\], one state node in each contraction set in the system graph has to be measured. (ii) The other, and more important, application is in case of observability loss. Assume that one (or more) of the sensor measurements fail and therefore the system is not observable anymore. To recover for this loss of observability, one can assign measurements of equivalent states as stated in Corollary \[cor\_equiv\]. The set of equivalent states for observability and estimation can be determined by finding contractions in the system digraph representation using Algorithm \[alg\_cont\]. For example, loosing the measurement of state $x^i$ one can measure another state $x^j$ sharing a contraction with $x^i$ in $\mc{G}_A$ to recover for system observability. These graph-theoretic applications are explained more in the following example. *Illustrative example:* Consider a system with $n=11$ states represented as a system graph in Fig. \[fig\_graph1\]-Top. ![(Top) This figure shows the example system graph with $3$ measurements of states $x^1,x^8,x^9$. The same-colored state nodes in red, green, and orange each represent states in the same contraction component and the blue states are not part of any contraction. (Bottom) The time evolution of the MSEE using the estimator in Eqs. - applied on the same system. The three measurements give an observable inference and therefore the MSEE is bounded steady-state stable.[]{data-label="fig_graph1"}](fig_graph1.pdf "fig:"){width="2.45in"} ![(Top) This figure shows the example system graph with $3$ measurements of states $x^1,x^8,x^9$. The same-colored state nodes in red, green, and orange each represent states in the same contraction component and the blue states are not part of any contraction. (Bottom) The time evolution of the MSEE using the estimator in Eqs. - applied on the same system. The three measurements give an observable inference and therefore the MSEE is bounded steady-state stable.[]{data-label="fig_graph1"}](fig_kalman_MSEE1.pdf "fig:"){width="2.45in"} ![(Top) This figure shows the same example system graph in Fig. \[fig\_graph1\] with $3$ new measurements of states $x^3,x^6,x^{11}$ sharing a contraction with the states measured in Fig. \[fig\_graph1\]. (Bottom) The time evolution of the MSEE of system estimation using the estimator in Eqs. -. As it can be seen, the equivalent measurements also provide an observable estimation with bounded steady-state MSEE.[]{data-label="fig_graph2"}](fig_graph2.pdf "fig:"){width="2.45in"} ![(Top) This figure shows the same example system graph in Fig. \[fig\_graph1\] with $3$ new measurements of states $x^3,x^6,x^{11}$ sharing a contraction with the states measured in Fig. \[fig\_graph1\]. (Bottom) The time evolution of the MSEE of system estimation using the estimator in Eqs. -. As it can be seen, the equivalent measurements also provide an observable estimation with bounded steady-state MSEE.[]{data-label="fig_graph2"}](fig_kalman_MSEE2.pdf "fig:"){width="2.45in"} Assume this graph represents a dynamic system, where each node is system state and each link represents the dynamic interaction of two states (for more details on such representation of systems as networks see [@Liu_nature]). The associated system matrix elements in $A_G$ (i.e. the link weights) are chosen randomly. For sake of illustration and avoiding trivial solutions the elements in $A_G$ are such that the spectral radius of adjacency matrix is greater than $1$, $\rho (A_G)>1$, i.e. the system is unstable. To determine the necessary states for observability we find the set of contractions in the system graph using Algorithm \[alg\_cont\] as $\mc{C}_1=\{x^1,x^3,x^5\}$, $\mc{C}_2=\{x^6,x^8\}$, and $\mc{C}_3=\{x^9,x^{11}\}$.[^12] The system is tracked by $m=3$ measurement of three states each in one contraction set of the system graph. This satisfies the condition in Theorem \[thm\_contr\] for observability and thus leads to stable estimation. These measurements along with system parameters are used in estimator Eqs. - to estimate and track the global state $\mb{x}$ of the system over time iterations $k$. The Mean Squared Estimation Error (MSEE) over time is shown in Fig. \[fig\_graph1\]-Bottom, which is bounded. Note that if we loose the measurement of a state in a contraction, according to Theorem \[thm\_contr\] we loose the system observability. Without observability, we loose the stability of the MSEE and the estimation error goes unbounded. To recover for loss of observability, one can take a measurement of an equivalent state in the same contraction, as shown in Fig. \[fig\_graph2\]-Top. Indeed, measuring any state in the same contraction set is sufficient for observability and yields stable estimation as shown in Fig. \[fig\_graph2\]-Bottom. The key point is that number of possible state to recover for observability directly relates to the size of contraction sets. In this example, there are two options to recover for loss of observation of $x^1$, while there is only one replacement for observation of $x^8$ and replacement for observation of $x^9$. Synthetic and Real Case Studies {#sec_sim} =============================== In this section, we analyze the number and size of contraction sets in both real and random complex networks. Recall that the contraction size is of interest because it determines the number of equivalent nodes for observability recovery, and number of contractions determine the number of node measurements necessary and sufficient for observability. First, random networks as models of complex systems are reviewed and relation between features of these networks and size/number of contractions are discussed. Next, the distribution of contraction sets in some examples of real-world networks are analyzed. Here, we proceed by first reviewing the definitions of relevant network properties. The local clustering coefficient of a node in a graph is defined as the fraction of pair of node neighbors that are linked together. On the other hand, the global clustering coefficient is defined as the fraction of the closed triplets (triangles) to the total number of the triplet paths in the graph [@newman2003structure]. Mathematically the clustering coefficient is defined as: $$\begin{aligned} CC(i) = \frac{2.tr(i)}{\mc{N}(i)(\mc{N}(i)-1)} \end{aligned}$$ where $tr(i)$ is the number of triangles that node $i$ forms with two of its neighbors. The global clustering coefficient of the network is defined as, $$\begin{aligned} CC = \frac{3.tr}{trp} \end{aligned}$$ where the $tr$ is the number of triangles and $trp$ is the number of connected triplets in the network. It is known that the clustering coefficient is a good measure of well-connectivity of the network and presence of strong community-structure in the network [@alstott2016cluster; @newman2003structure]. Degree heterogeneity is an intuitive concept related to the degree distribution of networks. Degree heterogeneity, as opposed to degree homogeneity, determines if the nodes in the network have various degrees (heterogeneous), or have similar degrees (homogeneous) to one another. Various measures of degree heterogeneity resembling the global differences in the node degrees are discussed in [@wu2010heterogeneity; @jacob2017measure]. The most well-known formula for degree heterogeneity is given by the variance of node degrees as follows [@jacob2017measure]: $$\begin{aligned} VAR = \frac{1}{n}\sum_{i=1}^{n}(\mc{N}(i) - \bar{\mc{N}})^2 \end{aligned}$$ where $\bar{\mc{N}}$ is the average node degree, $$\begin{aligned} \bar{\mc{N}} = \frac{1}{n}\sum_{i=1}^{n}\mc{N}(i) \end{aligned}$$ More details on these definitions can be found in [@newman2003structure; @jacob2017measure]. Contraction sets in Random Networks {#sec_rand} ----------------------------------- Random graphs are widely used to model complex systems facilitating analysis of different processes over networks, e.g., spreading processes or cascading failures [@newman2003structure; @barabasi2003scale; @barabasi_albert1999; @Holme2002clusteringScaleFree; @Toivonen2006social; @watts1998smallworld; @newman2002random]. The graphs are called random since the nodes in the graph are randomly connected with each other. We investigate two well-known models for random graphs. We particularly analyze the relation between number and size of contractions with clustering coefficient in Scale Free networks and with degree heterogeneity in Small-World Networks as discussed next. *Scale-Free networks:* Many complex networks are modeled by this type of random network. It is known that degree distribution of such networks follows a power-law distribution [@barabasi2003scale], i.e. the portion of nodes having degree $d$, represented by $f(d)$, follows the following formula: $$\begin{aligned} f(d) = d^{-\sigma},~2<\sigma <3\end{aligned}$$ In log-log scale, the distribution represents a linear function, hence it is named Scale-Free (SF) network. This implies that these graphs have a large number of low-degree nodes and few hubs with high connectivity. A well-known approach to build such networks is proposed by Barabasi and Albert [@barabasi_albert1999]. The Barabasi and Albert (BA) approach considers an initial graph of few number of nodes, called *initial seed* where recursively a new node with $m$ new links is added to the network. The probability that the new node makes a link to old nodes is proportional to the degree of old nodes, implying that the new node *preferably* links to high degree nodes, and is known as *preferential attachment*. In this method hubs with high degree are more likely to connect to the newly added nodes while the low degree nodes are unlikely to gain new links. These types of Scale-Free networks, e.g. BA model, are known to have low clustering coefficient. Therefore recently new random models of networks are proposed in the literature to account for high clustering of real networks [@Holme2002clusteringScaleFree; @Toivonen2006social]. These works propose to modify preferential attachment method such that the resulting networks, beside having power-law distribution, have high clustering. The network growth procedure is similar to the preferential attachment of BA model with some modification. Similar to BA model, they consider an initial seed. Then, recursively a new node connects to $m_r$ nodes in the network based on preferential attachment. But further, in each step the new node also makes $m_s$ links to randomly chosen neighbors of preferentially attached nodes in the network. This additional step is called *triad formation*. This increases the prevalence of triads (triangle cliques) in the network, and therefore results in high clustering coefficient. These random networks are called Clustered Scale Free (CSF) networks. For these networks we analyze the size and number of contractions. It should be noted that for simulation we consider $m=m_r+m_s$, i.e. number of links each new node makes in SF network equals number of links each new node makes in CSF network. This implies that the number of edges in both SF and CSF network are the same. This also implies that the average degree of the network is equal for both SF and CSF networks. This is important as all properties of both SF and CSF networks are similar except their clustering coefficient [@Holme2002clusteringScaleFree; @Toivonen2006social; @assenza2008enhancement]. Simulations are performed over $1000$ different realizations of 1000 node SF and CSF networks and the results are shown in Fig. \[fig\_SF\_CSF\]. ![This figure shows the distribution of contraction size on SF and CSF networks with similar number of preferentially attached nodes. The simulation is performed over $1000$ realizations of $1000$-nodes networks.[]{data-label="fig_SF_CSF"}](fig_SF_CSF.pdf){width="3in"} As it can be seen contractions are more prevalent in SF networks as compared to CSF networks and there are more contractions (and unmatched nodes) in SF networks as compared to CSF networks. Further, the contraction sets are in average larger in SF as compared to CSF networks. The results for $1000$ network realizations are summarized in Table \[tab\_SF\]. Type of network   SF  CSF -------------------------------- ---------- --------- Average Contraction size  $18.89$  $6.72$ Average number of Contractions  $156$  $109$ : Average size and number of contractions in $1000$ realizations of SF and CSF networks. \[tab\_SF\] This implies that by increasing the clustering coefficient as in CSF networks the number and size of contractions decreases. *Small-World networks:* The idea behind this model is to imitate the graphical properties of real-world networks. One of the main structural feature of the real graphs is that they show high level of community structure while keeping small average distance (shortest path), which is known as the small-world phenomena. Such features are not present in typical random models, for example in Erdos-Renyi graphs. Therefore, Watts and Strogatz [@watts1998smallworld] proposed a new semi-random graph, named Small-World model. The Watts and Strogatz (WS) model starts with a $k$-regular network in which every node is connected to its $k$ nearest neighbors (in both sides). Randomly pick links in the $k$-regular network with uniform probability $p$ independent of each other. Then, choose the end node of this link and randomly rewire it to another node. The rewiring must be such that the new link is not a self-link or a link that already exist in the network. By increasing the rewiring probability $p$ one can generate random networks which are more random in terms of their degree heterogeneity and as $p \rightarrow 1$ the model reaches an Erdos-Renyi (ER) graph with Poisson degree distribution [@newman2002random]. On the other hand, small $p$ implies that network conserves its regularity and degrees of most nodes lie around the average degree $2k$. Such networks have adjustable degree heterogeneity by tuning $p$. Indeed, regular networks are the most degree homogeneous and next are small world networks with tunable degree heterogeneity by factor $p$. By increasing the factor $p$ the degree heterogeneity increases up to the point where $p=100\%$ and the graph models the ER network. To relate the contraction size with degree heterogeneity, $1000$ different realizations of 1000 node SW networks with $9$ different $p$ factors are considered and the simulation results are shown in Fig. \[fig\_SW\]. Note that for this simulation only rewiring probability $p$ is changing, therefore graph properties such as number of edges and average node-degree remains unchanged and the only property that changes is degree-heterogeneity [@dwivedi2017optimization]. The average size and number of contractions in $1000$ network realizations are also summarized in Table \[tab\_SW\]. ![This figure shows the distribution of contraction size on different Small-World networks tuning the $p$ factor. The simulation is performed over $1000$ realizations of $1000$-nodes networks.[]{data-label="fig_SW"}](fig_SW.pdf){width="3in"} $p$ factor of SW network $10\%$ $20\%$ $30\%$ $40\%$ $50\%$ -------------------------- --------- -------- -------- -------- --------- avg Contraction size $13.71$ $7.62$ $5.62$ $4.68$ $4.14$ number of Contractions $25$ $47$ $68$ $86$ $101$ $p$ factor of SW network $60\%$ $70\%$ $80\%$ $90\%$ $100\%$ avg Contraction size $3.83$ $3.61$ $3.50$ $3.44$ $3.42$ number of Contractions $113$ $123$ $130$ $132$ $133$ : Average size and number of contractions in $1000$ realizations of SW networks with $10$ different $p$ factors. \[tab\_SW\] As it can be seen from the results by increasing the $p$ factor and degree heterogeneity, in average contractions are decreased in size but increased in number. Contraction sets in Real Networks {#sec_real} --------------------------------- *Power grid network:* As the first example, we consider the power grid network that represents the grid of the Western States of the United States of America. In this network a link is a power supply line and a node represents either a generator, a transformator or a substation. The network is originally addressed in [@watts1998smallworld] but the data is taken from [@UCI_power] where the description of state nodes can be found. It is known that such networks resemble the sparsity of system dynamic matrix where the states represent power flow, voltage, or phase angles [@asilomar14; @poor2012grid], and therefore the network can model a dynamic system type of Eq. , for more details see [@poor2012grid; @khan2013secure; @camsap11]. This network contains $6594$ interaction links connecting $4941$ state nodes. Applying the DM decomposition finds one possible set of unmatched nodes in the network. The network structure is shown in Fig. \[fig\_power\], including $575$ unmatched nodes represented in red color. Recall that From Theorem \[thm\_unmatched\], for observable estimation all the unmatched nodes must be observed by a sensor. These set of observable measurements gives one possible stable estimation of system state nodes over time. ![This figure shows the structure of Western-State Power grid network with $4941$ state nodes and $6594$ links; red nodes in the network represent unmatched nodes monitored by a sensor.[]{data-label="fig_power"}](fig_power.pdf){width="3in"} The distribution of all $575$ contractions in this network are shown in Fig. \[fig\_power\_distribution\]. It should be noted that the clustering coefficient of this network is $10.3\%$ and the average contraction size is $4.98$. ![This figure shows the frequency of size of different contractions in power grid network.[]{data-label="fig_power_distribution"}](fig_power_distribution.pdf){width="2.5in"} Applying the Contraction Detection Algorithm \[alg\_cont\] finds all the contraction sets in the network, where two examples are shown in Fig. \[fig\_power\_cont\]. These examples include contraction sets of size $3$ (green colored contraction) and $52$ (blue colored contraction). Recall that from Corollary \[cor\_equiv\] each contraction set associated with an unmatched node represents all possible options to recover for loss of measurement/observation. In this scenario, losing the observation of any node in the 3-nodes green contraction implies that there are measurements of only $2$ other nodes to recover for possible loss of observability, while for the blue colored contraction there are $51$ possible options to recover for the loss of observability. ![This figure includes two subnetworks of Power-grid network of Fig. \[fig\_power\]. Each subnetwork shows an example of nodes making a contraction, represented as blue and green colored nodes. These colored state nodes are equivalent in terms of network observability.[]{data-label="fig_power_cont"}](fig_power_cont1.pdf "fig:"){width="2in"} ![This figure includes two subnetworks of Power-grid network of Fig. \[fig\_power\]. Each subnetwork shows an example of nodes making a contraction, represented as blue and green colored nodes. These colored state nodes are equivalent in terms of network observability.[]{data-label="fig_power_cont"}](fig_power_cont2.pdf "fig:"){width="1in"} *Route-view network:* This network represents the network of connected autonomous systems of Internet. Every node is an autonomous system and every link represents communication between two systems. The data is taken from [@Konect_Routeviews], but the original description of the network is given in [@leskovec2007graph]. As stated in [@leskovec2007graph] every node is indeed a subgraph of Internet-connected routers that exchanges traffic flow with its peer neighbors. The network contains $6474$ nodes connected with $13895$ links, and is represented in Fig. \[fig\_routeview\_unmatched\]. In this figure regular nodes are represented in black while $3568$ unmatched nodes are shown in red color. ![This figure shows the Route-views network representing internet-connected autonomous systems. The network contains $6474$ nodes connected with $13895$ links. Nodes in red color are unmatched nodes each monitored by a sensor.[]{data-label="fig_routeview_unmatched"}](fig_routeview_unmatched.pdf){width="3.3in"} Applying the Contraction Detection Algorithm \[alg\_cont\] all $3568$ contractions in the network are found. The distribution of contraction sets is as shown in Fig. \[fig\_RouteView\_distribution\]. The average size of contractions in this network is $7.65$ and the clustering coefficient of the network is $0.959 \%$. ![This figure shows the frequency of size of different contractions in Route-view network.[]{data-label="fig_RouteView_distribution"}](fig_RouteView_distribution.pdf){width="2.5in"} Two examples of contraction sets in the Route-view network are shown in Fig. \[fig\_Routeview\_cont\]; one includes set of $2$ contraction nodes (in green color) and the other one includes set of $7$ contraction nodes (in orange color). As mentioned before, each contraction set represents the state nodes giving equivalent information for network observability and estimation. ![This figure shows two subnetworks of Route-view network of Fig. \[fig\_routeview\_unmatched\]. In each subnetwork colored nodes in orange and green represent example of nodes making a contraction. In network observability, these colored nodes represent equivalent states.[]{data-label="fig_Routeview_cont"}](fig_Routeview_Cont2.pdf "fig:"){width="2in"} ![This figure shows two subnetworks of Route-view network of Fig. \[fig\_routeview\_unmatched\]. In each subnetwork colored nodes in orange and green represent example of nodes making a contraction. In network observability, these colored nodes represent equivalent states.[]{data-label="fig_Routeview_cont"}](fig_Routeview_Cont1.pdf "fig:"){width="1in"} The results for these two networks are summarized in Table \[tab\_realnet\]. Name of network   Power grid  Route-view ------------------------ -------------- -------------- avg Contraction size  $4.98$  $7.65$ Contractions/nodes  $575/4941$  $3568/6474$ Clustering Coefficient  $10.3\%$  $0.959\%$ : Characteristics of two examples of real networks including average contraction size, ratio of number of contractions to number of nodes, and clustering coefficient. \[tab\_realnet\] Discussion and Conclusions {#sec_conc} ========================== Comparing the SF and CSF network, we observe a significant raise in average size of contractions in SF network. Noting that SF and CSF networks apply the same preferential attachment model and are similar in terms of most graph statistics including power-law degree distribution and logarithmically increasing average shortest-path length [@Holme2002clusteringScaleFree; @Toivonen2006social; @assenza2008enhancement], therefore, the only difference is low clustering coefficient as the key factor affecting the jump in average size of contractions in SF network. Similar statement holds for the average number of contractions in the network. Note that this number is decreased for clustered version of Scale-Free network while other network characteristics are unchanged. This implies that by increase in the clustering coefficient in average more contractions with larger size appear. This is also the case in real-world network examples stated in Section \[sec\_real\][^13]. For Power grid network with high clustering coefficient the ratio of number of contractions to the total number of nodes is lower than the Route-view network with low clustering coefficient. Similar statement holds for the average contraction size as the size of contractions are in average smaller in Power grid network. For observability and estimation of networks with power-law degree distribution (SF and CSF networks) these results imply that: (i) estimation/tracking of such networks with high clustering coefficient requires (in average) lower number of observations/measurements as there are less number of contractions, but (ii) in case of measurement/sensor failure there are less number of possible equivalent states for observability recovery as the average size of contractions are low. One application of these results is that one can tune the clustering coefficient of (synthetic) networks [@serrano2005tuning] to reduce the challenge for observability recovery and estimation. The other result of this paper is that in Small World networks the average size of contractions is to a great extent related to the degree homogeneity. Increasing the heterogeneity in Small World networks, by increasing the rewiring probability $p$ [@watts1998smallworld], is one key factor on the decrease of average contraction size as mentioned in Table \[tab\_SW\]. Note that by only changing the rewiring probability in SW networks the number of links, average node degree, and the size of graphs are unchanged. On the other hand, by increasing the degree heterogeneity, while the other graph characteristics of in SW networks are unchanged, the average number of contractions is increased. In terms of observability and estimation of SW networks these results imply that: (i) estimation of networks with high level of degree heterogeneity requires more number of measurements/observations which is due to prevalence of contractions, and (ii) in case of sensor/observation failure there are less number of possible options to recover for the loss of observability as the contraction sets in average are smaller in degree heterogeneous SW networks. As an application of these results one may decrease the degree heterogeneity by tuning the $p$ factor in synthetic networks to reduce the number of necessary measurements for observability and further increase the contraction size providing more possible countermeasures for observability recovery. Note that the above mentioned results are applicable for specific networks. In other words, we claim the results regarding the size/distribution of contractions and the clustering coefficient only for power-law degree distribution networks (SF-CSF networks). Further, the results on the relation of degree-heterogeneity and size/distribution of contractions are only stated for networks with Small-World property. For other kind of networks, for example Erdos-Renyi graphs, such results may not apply in general. Note that to make a justified claim about effect of clustering-coefficient/degree-heterogeneity we need to keep other graph properties (e.g. degree distribution, average degree, number of edges) unchanged so we can claim that the only effective property is clustering-coefficient/degree-heterogeneity. We cannot claim this for general graphs as they may differ in terms of, for example, degree distribution. It should be noted that the algorithms to check the matching properties of the network, namely Hopcroft-Karp algorithm [@hopcraft] or the Dulmage-Mendelsohn decomposition [@dulmage58] are of $\mc{O}(n^{2.5})$ complexity. Particularly, the complexity of Algorithm \[alg\_cont\] is in polynomial order $\mc{O}(n^{2.5})$. Note that, polynomial time algorithms are suitable for large-scale system analysis as their running time is upper-bounded by a polynomial expression in system size. The polynomial order complexity of the algorithms motivates application in observability analysis of large-scale networks/systems similar to the real examples given in the previous section. It is worth mentioning that, the results in this paper can be extended to the dual case of large-scale network controllability. As the final comment, it should be noted that this paper considers undirected networks and system graphs. The reason is that for directed networks *root SCCs* play important role in observability [@liu_pnas; @jstsp14]. Therefore, along with contractions, root SCCs are effective in observability recovery. In order to solely consider the role of contractions in observability recovery in this paper we focus on undirected networks. As the direction of future research, we plan to seek whether other graph properties such as network community structure and degree-degree correlation [@posfai2013correlation] are effective on the contraction analysis and observability properties. Acknowledgement {#acknowledgement .unnumbered} =============== The authors would like to thank Professor Glenn Lawyer from Max-Planck-Institute for Informatics for providing us with real network data. [Mohammadreza Doostmohammadian]{} received his B.Sc. and M.Sc. in Mechanical Engineering from Sharif University of Technology, Tehran, Iran, respectively in 2007 and 2010, where he worked on different applications of control systems and robotics. He received his PhD in Electrical and Computer Engineering from Tufts University, Medford, USA in 2015. During his PhD at Signal Processing and Robotic Network (SPARTN) lab he worked on control and signal processing over networks with particular application in social networks. From 2015 to 2017 he was a post-doctoral researcher at ICT Innovation Center for Advanced Information and Communication Technology (AICT), School of Computer Engineering, Sharif University of Technology, where he focused his research on control and estimation over networks and network epidemic. He is currently a researcher at Iran Telecommunication Research Center (ITRC), Tehran, Iran and a lecturer at Semnan University, Semnan, Iran. His general research interests include control and estimation, complex networks, and graph theory. He is a reviewer for IFAC and IEEE journals and conferences. [Hamid R. Rabiee]{} received his B.S. and M.S. degrees (with great distinction) in electrical engineering from California State University, Long Beach, CA, in 1987 and 1989, respectively; the EEE degree in electrical and computer engineering from the University of Southern California, Los Angeles, CA, in 1993; and the Ph.D. degree in electrical and computer engineering from Purdue University, West Lafayette, IN, in 1996. From 1993 to 1996, he was a Member of Technical Staff at AT&T Bell Laboratories. From 1996 to 1999, he worked as a Senior Software Engineer at Intel Corporation. From 1996 to 2000, he was also an Adjunct Professor of electrical and computer engineering at Portland State University, Portland, OR; Oregon Graduate Institute, Beaverton, OR; and Oregon State University, Corvallis, OR. Since September 2000, he has been with the Department of Computer Engineering, Sharif University of Technology, Tehran, Iran, where he is currently a Professor of computer engineering and Director of Sharif University Advanced Information and Communication Technology Research Institute (AICT), Digital Media Laboratory (DML), and Mobile Value Added Services Laboratory (M-VASL). He was also the founder of AICT, Advanced Technologies Incubator (SATI), DML, and M-VASL. He has been the Initiator and Director of national- and international-level projects in the context of United Nation Open Source Network program and Iran National ICT Development Plan, and holds 3 patents. Prof. Rabiee has received numerous awards and honors for his industrial, scientific, and academic contributions. He is a Senior Member of IEEE. [Houman Zarrabi]{} received his PhD from Concordia University in Montreal, Canada in 2011. Since then he has been involved in various industrial and research projects. His main expertise includes IoT, M2M, CPS, big data, embedded systems and VLSI. He is currently the national IoT program director and assistant professor at Iran Telecommunication Research Center (ITRC). [Usman Khan]{} received his B.S. degree (with honors) in Electrical Engineering from University of Engineering and Technology, Lahore-Pakistan, in 2002, M.S. degree in Electrical and Computer Engineering (ECE) from the University of Wisconsin-Madison (UW-Madison) in 2004, and Ph.D. degree in ECE from Carnegie Mellon University in 2009. Currently, he is an Assistant Professor with the ECE Department at Tufts University. He received the NSF Career award in January 2014 and was elevated to IEEE Senior Member grade in March 2014. His research interests include statistical signal processing, networked control and estimation, and distributed linear/nonlinear iterative algorithms for efficient operation and planning of complex infrastructures. Prof. Khan was a post-doctoral researcher in the Electrical and Systems Engineering department at the University of Pennsylvania from September 2009 to December 2010. He worked as a researcher in National Magnetic Resonance Facility at Madison (NMRFAM) from 2003 to 2005, as a research assistant in the Computer Science Dept. at UW-Madison from 2004 to 2005, and as an intern in AKAMAI Technologies in 2007. Dr. Khan is an associate member of the Sensor Array and Multichannel Technical Committee with the IEEE Signal Processing Society. He served on the Technical Program Committees of several IEEE conferences and has organized and chaired several IEEE workshops and sessions. His work was presented as Keynote speech at BiOS SPIE Photonics West-Nanoscale Imaging, Sensing, and Actuation for Biomedical Applications IX. [^1]: Manuscript received April 5, 2017. [^2]: In this paper a network describes the underlying dynamic system or phenomena. Therefore, throughout the paper the network and system are used interchangeably. [^3]: See [@liu_pnas; @Liu_nature; @nonlin] for extension to nonlinear case. [^4]: It should be noted that structural observability and graph theoretic method applied as a *tool* to solve network observability problem. See reference [@liu_pnas; @Liu_nature] for more information. [^5]: Note that many of stated references deal with dual problem of network controllability. The graph properties and notions can be simply redefined for network observability. [^6]: Note that, in general, edges in a bipartite graph have no direction. However in this paper, following the definition in [@murota], it is assumed that the edges have direction from $\mc{V}^+$ to $\mc{V}^-$. This kind of representation is later used in the definition of *Auxiliary graph*. [^7]: It should be mentioned that the concept of contraction is dual of dilation defined in the network controllability problem [@Liu_nature]. In a dilation set, less number of nodes are dilated into more number of nodes. So that we don’t need to continually refer to the dual graph, we define a contraction that is the natural dual of dilation. [^8]: Transitivity implies that if $x^i \sim x^j$ and $x^j \sim x^k$, then $x^i \sim x^k$. Reflexivity implies that every state/node is equivalent to itself, and symmetry implies that $x^i \sim x^j$, then $x^j \sim x^i$. [^9]: Note that, the structural rank (or S-rank) is defined as the maximum rank of the system matrix, $A$, by changing its free parametric entries. In the system graph, $\mc{G}$, S-rank is the size of the *maximum matching*, $\mc{M}$, see [@shields; @van1999generic] for details. [^10]: The results carry forward are also applicable to continuous-time systems. [^11]: This is not a straighforward procedure as the edge weights vary over time while the structure is time-invariant. Note that here we only convey the idea behind LSI dynamics with fixed sparsity pattern on the adjacency matrix. [^12]: Note that contraction $\mc{C}_1$ is similar to the contraction described in Fig. \[fig\_3nodecont\] and \[fig\_3nodecont2\]. [^13]: Note that it is known that most real-world networks including the two examples given in this paper follow a power-law degree distribution [@barabasi_albert1999].
{ "pile_set_name": "ArXiv" }
--- abstract: 'We consider a shape optimization problem for the first mixed Steklov-Dirichlet eigenvalues of domains bounded by two balls in two-point homogeneous space. We give a geometric proof which is motivated by Newton’s shell theorem.' address: 'Department of Mathematical Sciences, KAIST, 291 Daehak-ro Yuseong-gu, Daejeon 34141, Korea' author: - 'Dong-Hwi Seo' bibliography: - 'annot.bib' title: 'A shape optimization problem for the first mixed Steklov-Dirichlet eigenvalue' --- Introduction ============ Let $M^m$ be a Riemannian manifold of dimension $m\ge 2$ and $\Omega\subset M$ a bounded domain with Lipschitz boundary $\partial \Omega$. Let $\partial \Omega = C_1 \cup C_2$ with $C_1 \cap C_2 =\phi$. A mixed Steklov-Dirichlet eigenvalue problem is to find $\sigma \in \mathbb{R}$ for which there exists $u\in C^{\infty}(\Omega)$ satisfying $$\begin{aligned} \label{problem} \left\{ \begin{array}{rcll} \Delta u &=& 0& \text{in } \Omega \\ u &=& 0& \text{on } C_1\\ \frac{\partial u}{\partial \eta} &=& \sigma u &\text{on } C_2 \end{array} \right. ,\end{aligned}$$ where $\eta$ is the outward unit normal vector along $C_2$. When $C_1 = \phi$ and $C_2$ is connected, the problem becomes the Steklov eigenvalue problem introduced by Steklov in 1902 [@stekloff1902problemes]. We will find a domain maximizing the lowest $\sigma$ in a class of subsets in $M$. We call this problem by a shape optimization problem of the first eigenvalue. The shape optimization problem of the first nonzero Steklov eigenvalue in Euclidean space has been studied since the 1950s. In 1954, Weinstock considered the case when $M=\mathbb{R}^2$ [@weinstock1954inequalities]. He showed that the disk is the maximizer among all the simply connected domains with the same boundary lengths. Recently, Bucur, Ferone, Nitsch, and Trombetti studied this perimeter constraint shape optimization problem in any dimension among all the convex sets, and showed that the ball is the maximizer [@bucur2017weinstock]. Without the convexity condition, Fraser and Schoen proved the ball cannot be a maximizer even among all the smooth contractible domains of fixed boundary volume in $\mathbb{R}^m$, $m\ge 3$ [@fraser2019shape]. On the other hand, Brock [@brock2001isoperimetric] proved in 2001 that the ball is the maximizer among all the smooth domains with fixed domain volume in $\mathbb{R}^m$, $m \ge 2$. Note that he does not need any topological restriction. These shape optimization problems have been extended to non-Euclidean spaces as well. The first result in this direction was given by Escobar [@escobar1999isoperimetric] who showed that the first nonzero eigenvalue is maximal for the geodesic disk among all the simply connected domains with fixed domain area in simply connected complete surface $M^2$ with constant Gaussian curvature. In 2014, Binoy and Santhanam extended this result to noncompact rank one symmetric spaces of any dimension [@binoy2014sharp]. Regarding mixed Steklov-Dirichlet eigenvalue problems, it was considered by Hersch and Payne in 1968 [@hersch1968extremal]. They considered the problem (\[problem\]) when $\Omega\subset \mathbb{R}^2$ is a doubly connected region bounded by the inner and the outer boundaries, $C_1$ and $C_2$, respectively. Then among all the conformally equivalent domains with fixed perimeter of $C_2$, the annulus bounded by two concentric circles is the maximizer. Recently, Verma considered connected regions in $\mathbb{R}^m$ with $m \ge 2$ that are bounded by two spheres of given radii and gave the Dirichlet condition only on the inner sphere. Then the maximizer is obtained by the domain bounded by two concentric spheres [@verma2018eigenvalue]. The aim of this paper is to extend Verma’s result [@verma2018eigenvalue] from Euclidean spaces to two-point homogeneous spaces. The main theorem is as follows. We denote the injectivity radius of $M$ and the closure of a set $A\subset M$ by $inj(M)$ and $\text{cl}(A)$, respectively. \[main\] Let $M$ be a two-point homogeneous space. Let $\mathbf{B}_1$ and $\mathbf{B}_2^\prime$ be geodesic balls of radii $R_1, R_2>0$, respectively, such that $\textnormal{cl}(\mathbf{B}_1)\subset \mathbf{B}_2^\prime$ and $R_2 < inj(M)/2$. Then the first mixed Steklov-Dirichlet eigenvalue of the problem $$\begin{aligned} \label{vermaproblem} \left\{ \begin{array}{rcll} \Delta u &=& 0 &\textnormal{in } \mathbf{B}_2^\prime \symbol{92} \textnormal{cl}(\mathbf{B}_1)\\ u &=& 0 &\textnormal{on } \partial \mathbf{B}_1\\ \frac{\partial u}{\partial \eta} &=& \sigma u &\textnormal{on } \partial \mathbf{B}_2^\prime \end{array} \right.\end{aligned}$$ ($\eta$ : the outward unit normal vector along $\partial \mathbf{B}_2^\prime$) attains maximum if and only if $\mathbf{B}_1$ and $\mathbf{B}_2^\prime$ are concentric. Two-point homogeneous space has similar geometric properties with Euclidean space. For example, for two geodesic balls $\mathbf{B}_3$ and $\mathbf{B}_4'$ of radii $R_1$ and $R_2$, respectively, satisfying $\text{cl}(\mathbf{B}_3) \subset \mathbf{B}_4'$, $\mathbf{B}_4' \setminus \text{cl}(\mathbf{B}_3)$ is isometric to $\mathbf{B}_2' \setminus \text{cl}(\mathbf{B}_1)$ if and only if the distance of the centers of $\mathbf{B}_3$ and $\mathbf{B}_4'$ is equal to that of $\mathbf{B}_1$ and $\mathbf{B}_2'$. Furthermore, using additional angles, which are not usual Riemannian angles, there are laws of trigonometry and conditions for triangle conditions (for example, see Proposition \[congruence\]) in two-point homogeneous space. In order to prove the theorem, we estimate the first eigenvalue by substituting an appropriate test function on the Rayleigh quotient (see (\[variational characterization\]) in Section 2.1). We suggest a geometric proof to obtain the lower bound of the denominator of the quotient (see Corollary \[L2sum\]). It is similar to the proof of Newton’s shell theorem (see Remark after Proposition \[gradientofL2sum\]). Newton’s shell theorem is first proved by Newton [@Newton1972principia] (see Propositio LXX Theorema XXX in Sectio XII). It is extended to constant curvature spaces by Kozlov [@Kozlov2000newton] and Izmestiev and Tabachnikov [@Izmestiev2017ivory]. We prove that it is also holds for two-point homogeneous spaces with some restriction (see Corollary \[newton\] and the following Remark). In Section 2, we will briefly review the variational characterization of the mixed Steklov-Dirichlet eigenvalue problem (\[vermaproblem\]) as well as two-point homogeneous spaces and its trigonometry. Section 3 is devoted to the proof of the main theorem. In Section 3.1 we calculate the first mixed Steklov-Dirichlet eigenfunction on the annulus. In Section 3.2, we introduce some crucial lemmas (Section 3.2.1) and prove the main theorem (in Section 3.2.2 (noncompact rank one symmetric space, noted nCROSS) and in Section 3.2.3 (compact rank one symmetric space, noted CROSS)). Especially, in Section 3.2.1, we give a proof of Newton’s shell theorem for a two-point homogeneous space.\ Acknowledgement {#acknowledgement .unnumbered} =============== The author wishes to express his gratitude to Jaigyoung Choe for helpful discussions. This research was partially supported by NRF-2018R1A2B6004262. Background ========== The eigenvalue problem ---------------------- A mixed Steklov-Dirichlet eigenvalue problem (\[problem\]) is equivalent to the eigenvalue problem of the Dirichlet-to-Neumann operator : $$\begin{aligned} L : C^{\infty}(C_2) &\rightarrow C^{\infty}(C_2)\\ u &\mapsto \frac{\partial \hat{u}}{\partial \eta},\end{aligned}$$ where $\hat{u}$ is the harmonic extension of $u$ satisfying the following $$\begin{aligned} \left\{ \begin{array}{rcll} \Delta \hat{u} &=& 0& \text{in } \Omega \\ \hat{u} &=& 0& \text{on } C_1\\ \hat{u} &=& u &\text{on } C_2 \end{array} \right. .\end{aligned}$$ Then $L$ is a positive-definite, self-adjoint operator with discrete spectrum (see for instance [@agranovich2006mixed]), $$\begin{aligned} 0<\sigma_1(\Omega)\le \sigma_2(\Omega) \le \cdots \rightarrow \infty,\end{aligned}$$ provided that $C_1\neq \phi$. We call $\sigma_k(\Omega)$ by the $k$th mixed Steklov-Dirichlet eigenvalue, or simply the $k$th eigenvalue. An eigenfunction of $L$ corresponding to $\sigma_k(\Omega)$ is called the $k$th mixed Steklov-Dirichlet eigenfunction, or the $k$th eigenfunction. Then the first eigenvalue $\sigma_1(\Omega)$ is characterized variationally as follows $$\begin{aligned} \label{variational characterization} \sigma_1(\Omega) = \inf \left\{ \left.\frac{\int_{\Omega}|\nabla v|^2 dV}{\int_{C_2}v^2 ds} \right| v \in H^1(\Omega)\setminus\{0\} \textnormal{ and } v=0 \textnormal{ on } C_1 \right\}.\end{aligned}$$ For convenience we shall call the harmonic extension of the $k$th eigenfunction by the $k$th mixed Steklov-Dirichlet eigenfunction or the $k$th eigenfunction. Two-point homogeneous spaces and triangle congruence conditions --------------------------------------------------------------- Three points in a Euclidean space determine a triangle when three points are not lie on a single line. In classical geometry, there are several congruence conditions on triangles and it is determined by lengths of sides and angles. For example, side-angle-side (SAS) congruence is given by two side lengths and the included angle. In two-point homogeneous spaces, analogous properties also hold with additional angles. These facts are obtained by the laws of trigonometry. In this section, we give some information about two-point homogeneous spaces and its congruence conditions of triangles which will be used later. See [@wolf1972spaces],[@hsiang1989laws],[@brehm1990shape] for more details. A connected Riemannian manifold $M$ is called two-point homogeneous space if $x_i,y_i \in M, i=1,2$ with $dist(x_1, y_1)=dist(x_2,y_2)$, there is an isometry $g$ of $M$ such that $g(x_1)=x_2$ and $g(y_1)=y_2$. In fact, two-point homogeneous spaces are Euclidean spaces or rank one symmetric spaces. We will call the latter spaces by ROSSs. Furthermore, compact ROSS and noncompact ROSS are denoted by CROSS and nCROSS, respectively. Then two-point homogeneous spaces with their isotropy representations are classified as in the Table \[twopointhomogeneous space\] (see [@wolf1972spaces],[@hsiang1989laws]). Here $m\ge 1, n\ge 2$ and $m=\textnormal{dim}_{\mathbb{R}}M = n \cdot \textnormal{dim}_{\mathbb{R}}\mathbb{K}$. CROSS nCROSS Isotropy representation ------------------------- ---------------- ------------------------------- ----------------- --------------------------------------- $\mathbb{K}=\mathbb{R}$ $\mathbb{R}^m$ $\mathbb{S}^m, \mathbb{R}P^n$ $\mathbb{R}H^n$ $(O(m),\mathbb{R}^m)$ $\mathbb{K}=\mathbb{C}$ $\cdot$ $\mathbb{C}P^n$ $\mathbb{C}H^n$ $(U(n),\mathbb{R}^{2n})$ $\mathbb{K}=\mathbb{H}$ $\cdot$ $\mathbb{H}P^n$ $\mathbb{H}H^n$ $(Sp(1)\times Sp(n),\mathbb{R}^{4n})$ $\mathbb{K}=\mathbb{O}$ $\cdot$ $\mathbb{O}P^2$ $\mathbb{O}H^2$ $(Spin(9),\mathbb{R}^{16})$ : Two-point homogeneous spaces, $m\ge 1, n\ge 2$.[]{data-label="twopointhomogeneous space"} An angle is given by two directions at a point $P$. It is classified by its congruence classes which are given by the orbit space of $U_{P}M \times U_{P}M/K$, where $U_PM$ is the unit sphere in the tangent space of $M$ at $P$, and $K$ is the isotropy subgroup of the isometry group $M$ at $P$. The orbit space can be seen by fixing the first component by the action of $K$. More precisely, it is equivalent to an orbit space $U_PM/H$ of an isotropy group $H\subset K$ with respect to a point in $U_PM$. Then it can be checked that for given $\vec{v}_1\in U_PM$, $H$-invariant subspaces are $\mathbb{R}\cdot \vec{v}_1, \mathbb{K}'\cdot \vec{v}_1,$ and the subspace orthogonal to $\mathbb{K}\cdot \vec{v}_1$, where $\mathbb{K}=\mathbb{R}, \mathbb{C}, \mathbb{H}$, and $\mathbb{O}$ and $\mathbb{K}'$ is the set of pure imaginary numbers in $\mathbb{K}$. Then a direction $\vec{v}_2$ is determined up to $H$-action by the following angular invariants (for more details, see [@hsiang1989laws],[@brehm1990shape]): - $\lambda(\vec{v}_1,\vec{v}_2) = \angle(\vec{v}_1,\vec{v}_2)$ ; $0\le\lambda\le\pi$, - $\varphi(\vec{v}_1,\vec{v}_2)=\angle(\vec{v}_1, \mathbb{K}\cdot \vec{v}_2)$ ; $0\le\varphi\le\frac{\pi}{2}$, where $\angle(\vec{v}_1, \vec{v}_2)$ is the usual (Riemannian) angle and $\angle(\vec{v}_1,\mathbb{K}\cdot \vec{v}_2)$ is the angle between $\vec{v}_1$ and the subspace $\mathbb{K}\cdot \vec{v}_2$. Note that when $\mathbb{K}=\mathbb{R}$, $\lambda=\varphi$ or $\lambda=\pi-\varphi$. Then angular invariants satisfy following relations : $$\begin{aligned} \label{180} \lambda(\vec{v}_1,-\vec{v}_2) &= \pi - \lambda(\vec{v}_1,\vec{v}_2), \\ \label{180_2} \varphi(\vec{v}_1,-\vec{v}_2) &= \varphi(\vec{v}_1,\vec{v}_2).\end{aligned}$$ Using the previous $H$-invariant decomposition, we can write the metric of ROSS $M$ explicitly. Let $s(r)$ and $c(r)$ be functions defined as follows : $$\begin{aligned} s(r)= \begin{cases} \sin{r} \text{ with } 0\le r<\pi & \text{if } M=\mathbb{S}^m\\ \sin{r} \text{ with } 0\le r<\frac{\pi}{2} & \text{if } M=\mathbb{R}P^n, \mathbb{C}P^n, \mathbb{Q}P^n, \mathbb{O}P^2\\ \sinh{r} & \text{if } M \text{ is nCROSS} \end{cases}\\\end{aligned}$$ and $$\begin{aligned} c(r)= \begin{cases} \cos{r} \text{ with } 0\le r<\frac{\pi}{2} & \text{if } M=\mathbb{C}P^n, \mathbb{Q}P^n, \mathbb{O}P^2\\ \cosh{r} & \text{if } M \text{ is nCROSS}. \end{cases} \end{aligned}$$ Then the metric $(ds)^2$ is given by $$\begin{aligned} \label{metric} (ds)^2 = (dr)^2+(s(r))^2(c(r))^2g+(s(r))^2h, \end{aligned}$$ where $(dr)^2,g,$ and $h$ are written by $\sigma_1^2$ with the coframe $\sigma_1$ dual to $\vec{v}_1$; $\sigma_2^2+\cdots +\sigma_k^2$ with coframes $\sigma_2,\dots, \sigma_k$ dual to orthonormal basis of $\mathbb{K}'\cdot \vec{v}_1$; $\sigma_{k+1}^2+\cdots +\sigma_m^2$ with coframes $\sigma_{k+1},\dots, \sigma_m$ dual to the complement orthonormal basis of $\mathbb{R}^m$. Since the density function $\omega$ only depends on distance, we may define $\omega$ as a one-variable function $$\omega(r) = (s(r))^{m-1}(c(r))^{k-1}.$$ Then the sectional curvature $K_M$ of $M$: $$\begin{aligned} \label{curvature} \begin{cases} 1\le K_M \le 4 & \text{if } M \text{is CROSS}\\ -4\le K_M \le -1 & \text{if }M \text{is nCROSS}. \end{cases}\end{aligned}$$ In particular, $\mathbb{S}^m$ and $\mathbb{R}P^n$ has sectional curvature 1. Then the condition $0<R_2 < \frac{inj(M)}{2}$ in Theorem 1 implies: $$\begin{aligned} \left\{ \begin{array}{ll} 0<R_2<\frac{\pi}{2} & \text{if } M=\mathbb{S}^m \\ 0<R_2<\frac{\pi}{4} & \text{if } M=\mathbb{R}P^n, \mathbb{C}P^n, \mathbb{H}P^n, \mathbb{O}P^2\\ 0<R_2 & \text{otherwise}. \end{array} \right.\end{aligned}$$ Now consider a triangle $(PQR)$ in $M$ with the metric (\[metric\]), which consists of three distinct points $P,Q,R\in M$ and three connecting geodesics $QR, RP, PQ$. The side lengths will be denoted by $p, q$, and $r$, respectively and the two angular invariants $\lambda, \varphi$ determined by the two tangent vectors of geodesic rays $\vv{PQ}$ and $\vv{PR}$ at $P$ will be denoted by $\lambda(P)$ and $\varphi(P)$, respectively. Furthermore we can denote $\lambda(Q)$, $\varphi(Q)$, $\lambda(R)$, and $\varphi(R)$ in an analogous way. Then it is known that there are congruent conditions of triangles. We introduce some conditions which will be used later. For more conditions, see [@brehm1990shape]. \[congruence\] A triangle (PQR) in ROSS with the metric $\textnormal{(\ref{metric})}$ is uniquely determined up to isometry as follows : 1. \[a\] $p, q,$ and $\lambda(P)$ with $0<p,q,r<\pi$ and $q<p<\frac{\pi}{2}$ if $M$ is $\mathbb{S}^m$. 2. \[b\] $p, q,$ and $\lambda(P)$ with $0<p,q,r<\frac{\pi}{2}$ and $q<p<\frac{\pi}{4}$ if $M$ is $\mathbb{R}P^n$. 3. \[c\] $p, q,$ $\lambda(P)$, and $\varphi(P)$ with $0<p,q,r<\frac{\pi}{2}$ and $(p-q)(\cos{p}-\sin{q}\cos{\varphi(P)})>0$ if $M$ is $\mathbb{C}P^n, \mathbb{H}P^n$ or $\mathbb{O}P^2$. 4. \[d\] $p, q, \lambda(P)$, and $\varphi(P)$ with $0<p,q,r$ and $q<p$ if $M$ is nCROSS. The proof of \[a\] can be found in Section VI in [@todhunter1863spherical]. In fact, the condition $p<\frac{\pi}{2}$ can be replaced by $p+q < \pi$. The proof of \[b\] follows from \[a\]. The proofs of \[c\] and \[d\] can be found in (ix) and (ix’) of Theorem 4 and 4’ in [@brehm1990shape]. Main proof ========== Let $M$ be a ROSS with the metric (\[metric\]). Let $X$ and $C$ be the centers of $\mathbf{B}_1$ and $\mathbf{B}_2^\prime$, respectively. Define $\mathbf{B}_2$ to be the ball of radius $R_2$, centered at $X$. The first eigenfunctions ------------------------ In this section, we derive an explicit formula for the first mixed Steklov-Dirichlet eigenfunctions in $\mathbf{B}_2\setminus \textnormal{cl}(\mathbf{B}_1)$. Using the following standard argument as in [@castillon2016spectral] and [@verma2018eigenvalue], we can show that the first eigenfunction is a function that only depends on the distance from $X$. Using seperation of variables, a mixed Steklov-Dirichlet eigenfunction $u(r,\theta_1,\dots \theta_{m-1})$ in $\mathbf{B}_2\setminus \textnormal{cl}(\mathbf{B}_1)$ is obtained by multiplying a Laplacian eigenfunction $f(\theta_1,\dots, \theta_{m-1})$ on the unit sphere $\mathbb{S}^{m-1}$ by an appropriate radial function $a(r)$. Here, $(r, \theta_1,\dots \theta_{m-1})$ is the polar coordinate in $T_XM$. Since Laplace eigenfunctions on $\mathbb{S}^{m-1}$ are indeed Laplace eigenfunctions on $\partial\mathbf{B}_2$ (see Theorem 3.1 in [@castillon2016spectral], or Corollary 5.5 in [@bergery1982laplacians]) and it consists of a basis of $L^2(\partial\mathbf{B}_2)$, our mixed Steklov-Dirichlet eigenfunctions restrict to $\partial \mathbf{B}_2$ become a basis of $L^2(\partial \mathbf{B}_2)$. It implies the $k$th mixed Steklov-Dirichlet eigenfunction is written by a product of a Laplacian eigenfunction and a radial function. By some computations as in Section 2.1 in [@verma2018eigenvalue], we can conclude that the $k$th mixed Steklov-Dirichlet eigenfunction is corresponding to the $k$th Laplacian eigenfunction. Since the first Laplacian eigenfunctions are constants, we obtain the following. \[firsteigenfunction\] Let $r_X : M\rightarrow [0,\infty)$ be the distance function from $X$. Let $a : [R_1, \infty) \rightarrow \mathbb{R}$ be a function defined by $$\begin{aligned} a(r) = \int_{R_1}^r \frac{1}{\omega(t)}dt. \end{aligned}$$ Then the first mixed Steklov-Dirichlet eigenfunction in $\mathbf{B}_2\setminus \textnormal{cl}(\mathbf{B}_1)$ is $a\circ r_X$ up to constant. By the argument in the paragraph, the first eigenfunction can be written by $$\begin{aligned} a\circ r_X, \end{aligned}$$ where $a:[R_1, \infty) \rightarrow \mathbb{R}$ is a real-valued function. Then, the harmonicity of the eigenfunction implies $$\begin{aligned} 0=\Delta a(r) = a^{\prime\prime}(r)+\frac{\omega^\prime(r)}{\omega(r)}a^\prime(r) = \frac{1}{\omega(r)}(a^\prime(r)\omega(r))^\prime. \end{aligned}$$ Here, we used $r$ instead of $r_X$ for simplicity of notation. With the fact that $a(R_1)=0$ from the boundary condition, we obtain the formula of $a(r)$ up to constant. Crucial lemmas and the proof for nCROSS --------------------------------------- We begin with two definitions. For given $X\in \mathbf{B}_2^\prime$, a vector-valued function $\vec{v}_X: M\setminus{\{X\}}\rightarrow U_XM$ is defined by $P\in M\setminus \{X\}$ and $\vec{v}_X(P)\in U_XM$ such that $\vec{v}_X(P)$ is the unit tangent vector of the geodesic ray $\vv{XP}$ at $X$. For a given parametrization of $M$ around $X$, we can identify $T_XM$ with $\mathbb{R}^m$. Then we can give the following definition. For given $X\in \mathbf{B}_2^\prime$ and a parametrization of $M$ around $X$, a map $\pi_X : \mathbb{S}^{m-1}\cong U_XM \rightarrow \partial \mathbf{B}_2^\prime$ is defined by $\pi_X(v) = \textnormal{exp}_X([0,\infty)\cdot v) \cap \partial \mathbf{B}_2^\prime$, i.e. $\pi_X(v)$ is the point of $\partial \mathbf{B}_2'$ in the geodesic emanating from $X$ in $v$ direction. Note that $\pi_X$ has the inverse map. Thus, for any $P\in \partial \mathbf{B}_2^\prime$, we can find $P_s\in \mathbb{S}^{m-1}$ such that $P=\pi_X(P_s)$. Furthermore, let $C_s\in\mathbb{S}^{m-1}$ such that the geodesic ray $\text{exp}_X([0,\infty)\cdot C_s)$ passes through $C$. Then we can define $-P_s$ and $\bar{P_s}$ in $\mathbb{S}^{m-1}$ such that they are the symmetric points of $P_s$ with respect to $X$ and the line passing through $X$ and $C_s$, respectively. In addition, $-\bar{P_s}$ can be defined as the symmetric point of $\bar{P_s}$ with respect to $X$. Now we denote $\text{exp}_X(-P_s)$, $\text{exp}_X(\bar{P_s})$, and $\text{exp}_X(-\bar{P_s})$ by $-P$, $\bar{P}$, and $-\bar{P}$, respectively. Figure \[points\] explains the situation. \(C) at (0,0); (D) at (2,0); (X) at (-1,0); at (X) [$X$]{}; \(P) at ($(C)!1!60:(D)$); (bP) at ($(C)!1!-60:(D)$); let 1=($(P)-(X)$), 1=[atan2(1,1)]{} in (X)–($(X)+(180+\n1 : 2*2cm)$); let 2=($(bP)-(X)$), 2=[atan2(2,2)]{} in (X)–($(X)+(180+\n2 : 2*2cm)$); (X)–(P); (X)–(bP); \(C) circle(2cm); (X) circle (0.5cm); (-P) node\[ left\][$-P$]{}–(X); (-bP) node\[ left\][$-\bar{P}$]{}–(X); (X)–(bP); (X)–(C); (X)–(P); (X)–(C); (X)–(bP); ; ; ; ; ; ### Properties of angles and distances In this section, we prove the lemmas which are essential in the proof of the main proposition in the next section. We prove a lemma about the “symmetric properties" of angles and distances. In addition, we obtain a lemma which is motivated from the concept of solid angle. As a corollary, we introduce Newton’s shell theorem with an infinitesimally thin “shell" in ROSS. We begin with a lemma, which are useful for the lemmas below. \[rightangle\] A triangle $(PQR)$ in ROSS $M$ with the metric $\textnormal{(\ref{metric})}$ satisfies : 1. If $M=\mathbb{S}^m$, $0<p,q,r<\pi$, and $p\le q <\frac{\pi}{2}$, then $\lambda(P)<\frac{\pi}{2}$. 2. If $M=\mathbb{R}P^n, \mathbb{C}P^n, \mathbb{H}P^n,$ or $\mathbb{O}P^2$, $0<p,q,r<\frac{\pi}{2}$, and $p\le q <\frac{\pi}{4}$, then $\lambda(P)<\frac{\pi}{2}$. 3. If $M$ is nCROSS, $0<p,q,r$, and $p\le q$, then $\lambda(P)<\frac{\pi}{2}$. <!-- --> 1. Suppose $\lambda(P)\ge \frac{\pi}{2}$. Using the law of cosines of spherical triangles (see p. 179 in [@karcher1989comparison]), $$\begin{aligned} \cos{p} = \cos{q}\cos{r}+\sin{q}\sin{r}\cos{P} < \cos{q}\cos{r}. \end{aligned}$$ Combining the previous inequality with $\cos{p},\cos{q}>0$, we obtain $\cos{r}>0$ and $\cos{p}<\cos{q}$. It implies $p>q$, contradiction to our assumption. 2. Suppose $\lambda(P)\ge \frac{\pi}{2}$. Since $M$ has sectional curvature $K_M \le 4$ as in (\[curvature\]), we can apply the triangle comparison theorem (see p. 197 in [@karcher1989comparison]). $$\begin{aligned} \cos{2p} \le \cos{2q}\cos{2r}+\sin{2q}\sin{2r}\cos{P}<\cos{2q}\cos{2r}. \end{aligned}$$ Then by an analogous argument in (a), we obtain a contradiction. 3. Suppose $\lambda(P)\ge \frac{\pi}{2}$. Since $M$ has sectional curvature $K_M \le -1$ as in (\[curvature\]), we can apply the triangle comparison theorem (see p. 197 in [@karcher1989comparison]). $$\begin{aligned} \cosh{p} \ge \cosh{q}\cosh{r}-\sinh{q}\sinh{r}\cos{P}>\cosh{q}. \end{aligned}$$ Thus $p>q$, which contradicts to our assumption. For $P\in \partial \mathbf{B}_2^\prime$, consider a triangle $(PXC)$ in $\textnormal{cl}(\mathbf{B}_2^\prime)$ defined in the beginning of Section 3, which consists of the center $X$ of $\mathbf{B}_1$, the center $C$ of $\mathbf{B}_2^\prime$, $P$, and godesics connecting two of them. Then the next lemma explains relations of distances from X to $P,\bar{P},-P$, and $-\bar{P}$ and relations of angles at those points. ![Illustration of Lemma \[symmetry\]. The circles in (a),(b),(c) represent $\partial \mathbf{B}_2'$.[]{data-label="Lemma explained"}](a.png){width="\linewidth"} ![Illustration of Lemma \[symmetry\]. The circles in (a),(b),(c) represent $\partial \mathbf{B}_2'$.[]{data-label="Lemma explained"}](b.png){width="\linewidth"} ![Illustration of Lemma \[symmetry\]. The circles in (a),(b),(c) represent $\partial \mathbf{B}_2'$.[]{data-label="Lemma explained"}](c.png){width="\linewidth"} \[symmetry\] Let $\lambda_X : \partial \mathbf{B}_2^\prime \rightarrow [0,\pi]$ be an angle function with respect to $X$ that assigns to each $P \in \partial \mathbf{B}_2^\prime$ an angle $\lambda(P)$ of the triangle (PXC). Define $r_X$ as in the Proposition \[firsteigenfunction\]. Then, $\lambda_X$ and $r_X$ satisfy the following. 1. $0 \le \lambda_X(P) <\frac{\pi}{2}$. 2. $\lambda_X(P) = \lambda_X(\bar{P})$, $r_X(P)=r_X(\bar{P})$ for all $P\in \partial \mathbf{B}_2^\prime$. 3. $\lambda_X(P) = \lambda_X(-P)$, $r_X(P)\ge r_X(-P)$ for all $P\in \partial \mathbf{B}_2^\prime$ satisfying\ $\angle(\vec{v}_X(P), \vec{v}_X(C)) \le \frac{\pi}{2}$. The equality holds if and only if $\angle(\vec{v}_X(P), \vec{v}_X(C)) = \frac{\pi}{2}$. We will prove this lemma when $M = \mathbb{C}P^n, \mathbb{H}P^n$ or $\mathbb{O}P^2$. Then we have $R_2 < \frac{inj (M)}{2}=\frac{\pi}{4}$. 1. Note that $R_2 <\frac{\pi}{4}$ and $|CX|<|CP|=R_2$. Then the statement follows from Lemma \[rightangle\]. 2. Consider two triangles $(PXC)$ and $(\bar{P}XC)$. By the constructions of $P$ and $\bar{P}$, $\lambda(X)$ of $(PXC)$ and $(\bar{P}XC)$ are identical. The same holds for $\varphi(X)$. Note that the two triangles have the common edge $XC$ and $|CP|=|C\bar{P}|=R_2$. From the fact that $|CX|<|CP|=R_2<\frac{\pi}{4}$ we have $\sin{|CX|}<\cos{|CP|}$. Therefore by Proposition \[congruence\], $(PXC)$ and $(\bar{P}XC)$ are congruent. Then our statement follows. 3. Using the fact that $\mathbf{B}_2^\prime$ is convex (see p. 148 in [@petersen2006riemannian]), we can define a point $R \in \mathbf{B}_2^\prime$ in the complete geodesic containing $X$ and $P$ such that the geodesic meets $CR$ perpendicularly. Then under the condition on $P$, we claim that $|PR|\le |PX|$. It is equivalent to showing that $\lambda(\vv{XR},\vv{XC})\le \frac{\pi}{2}$. If $X=R$, $\lambda(\vv{XR},\vv{XC})= \frac{\pi}{2}$. Otherwise, we have $|RC|<|XC|<\frac{\pi}{4}$. Then by Lemma \[rightangle\] for $(XCR)$, our claim follows. On the other hand, two triangles $(PRC)$ and $(-PRC)$ are congruent by (\[180\]),(\[180\_2\]), and Proposition \[congruence\] as in the proof of (b). Thus we obtain that $\lambda_X(P)=\lambda_X(-P)$ and $|PR|=|-PR|$, which imply the desired conclusion. A slight change in the proof shows it also holds if $M$ is $\mathbb{S}^m, \mathbb{R}P^n$ or nCROSSs. Now we will give another lemma that explains an “infinitesimal area of $\partial \mathbf{B}_2$ from $X$" can be calculated by $\lambda_X$ and $r_X$. \(C) at (0,0); (X) at (-1,0); (D) at (2,0); (P1) at ($(C)!1!105:(D)$); (P2) at ($(C)!1!40:(D)$); let 2=($(P1)-(X)$), 2=[atan2(2,2)]{} in (X) –($(X)+(180+\n2 :2*2cm)$); let 3=($(P2)-(X)$), 3=[atan2(3,3)]{} in (X) –($(X)+(180+\n3:2*2cm)$); (C) circle(2cm); (X) circle (0.5cm); (P1)–(X); (P2)–(X); let 2=($(P1)-(X)$), 2=[atan2(2,2)]{},3=($(P2)-(X)$), 3=[atan2(3,3)]{} in (X) + (3 : 0.5) arc (3 : 2 : 0.5) – (P1) arc (105 :40 : 2)–cycle; at ($(C)!1!72.5:(D)$) [$\mathcal{A}$]{}; at ($(X)!1!72.5:(-0.5,0)$) [$\pi_X^{-1}(\mathcal{A})$]{}; let 2=($(P1)-(X)$), 2=[atan2(2,2)]{},3=($(P2)-(X)$), 3=[atan2(3,3)]{} in (X) + (3 : 0.5) arc (3 : 2 : 0.5); (P1) arc (105 : 40 : 2); at ($(C)!1!92.5:(1,0)$) [$\mathcal{R}$]{}; \[radon\] Let $\mu$ be the Lebesgue measure on $\mathbb{S}^{m-1}$ and consider the pushforward ${\pi_X}_{\#}\mu$ on $\partial \mathbf{B}_2^\prime$. Then for a measurable set $\mathcal{A} \subset \partial \mathbf{B}_2^\prime$, we have $$\begin{aligned} {\pi_X}_{\#}\mu(\mathcal{A})= \mu(\pi_X^{-1}(\mathcal{A}))= \int_{\mathcal{A}} \frac{\cos{\lambda_X}}{\omega(r_X)}dS_2^\prime, \end{aligned}$$ where $S_2^\prime$ is the induced measure on $\partial \mathbf{B}_2^\prime$ from the metric of $M$. Equivalently, $$\begin{aligned} dS_2^\prime = \frac{\omega(r_X)}{\cos{\lambda_X}}d{\pi_X}_{\#}\mu. \end{aligned}$$ It is clear that $S_2^\prime$ and ${\pi_X}_{\#}\mu$ are $\sigma$-finite and ${\pi_X}_{\#}\mu \ll S_2^\prime$ that is to say $(\pi_X)_{\#}\mu$ is absolutely continuous with respect to $S_2^\prime$. Furthermore, $S_2^\prime \ll {\pi_X}_{\#}\mu$. By the Radon-Nikodym theorem, there are functions $f_1$ and $f_2$ on $\partial \mathbf{B}_2^\prime$ such that $$\begin{aligned} {\pi_X}_{\#} \mu(\mathcal{A})=\int_{\mathcal{A}}f_1 dS_2^\prime \end{aligned}$$ and $$\begin{aligned} dS_2^\prime(\mathcal{A})=\int_{\mathcal{A}}f_2 d{\pi_X}_{\#}\mu. \end{aligned}$$ Consider a vector field $\mathbb{F}$ on $M\setminus \{X\}$ defined by $$\begin{aligned} \mathbb{F}(Y) = \left(\frac{1}{\omega(r_X)}\frac{\partial}{\partial r}\right)(Y), \end{aligned}$$ where $\frac{\partial}{\partial r}(Y)$ is the vector in $T_YM$ obtained by the parallel transport of the unit tangent vector $\vec{v}_X(Y)$ along $XY$. Then $$\begin{aligned} \text{div}(\mathbb{F})= \frac{1}{\omega(r_X)}\frac{\partial}{\partial r}\left(\omega(r_X)\cdot \frac{1}{\omega(r_X)} \right)=0. \end{aligned}$$ Now consider a region $\mathcal{R}$ that is the region of the solid cone from $X$ over a geodesic ball $\mathcal{B}\subset \partial\mathbf{B}_2^\prime$ bounded by $\partial \mathbf{B}_1$ and $\partial \mathbf{B}_2^\prime$. Equivalently, $$\begin{aligned} \mathcal{R} = \{\text{exp}_X (t\cdot \vec{v}_X(Y)) | Y\in \mathcal{B}, R_1 \le t \le r_X(Y)\}. \end{aligned}$$ Let $\mathcal{R}\cap \partial \mathbf{B}_1 = \mathcal{B}_1$. Then applying the divergence theorem to $\mathbb{F}$ on $\mathcal{R}$, we have $$\begin{aligned} 0=\int_{\mathcal{R}}\text{div}\mathbb{F} = \int_{\mathcal{B}} \frac{\cos{\lambda_X}}{\omega(r_X)}dS_2^\prime-\int_{\mathcal{B}_1} \frac{1}{\omega(R_1)}dS_1, \end{aligned}$$ where $S_1$ is the measure on $\partial \mathbf{B}_1$ induced by the metric of $M$. Combining it with the fact that $$\begin{aligned} \int_{\mathcal{B}_1} \frac{1}{\omega(R_1)} dS_1= \mu(\pi_X^{-1}(\mathcal{A})), \end{aligned}$$ the first statement is proved for $\mathcal{B}$. Then by Theorem 4.7 in [@Simon1983measure], the first statement is proved. Since $\cos{\lambda_X}\neq 0$ from Lemma \[symmetry\], the second argument follows. The following corollary is not necessary for the proof of the main theorem. \[newton\] We have $$\begin{aligned} \int_{\partial \mathbf{B}_2^\prime} \frac{\vec{v}_X}{\omega(r_X)}dS_2^\prime =0.\end{aligned}$$ Using the previous lemma, the left hand side is equal to $$\begin{aligned} \label{symmetric identity} \int_{\mathbb{S}^{m-1}} \left(\frac{\vec{v}_X}{\cos{\lambda_X}}\circ \pi_X\right) d\mu. \end{aligned}$$ By Lemma \[symmetry\], we have $$\begin{aligned} \left(\frac{\vec{v}_X}{\cos{\lambda_X}} \right) \circ \pi_X(P_s) + \left(\frac{\vec{v}_X}{\cos{\lambda_X}} \right) \circ \pi_X(-P_s) =0 \end{aligned}$$ for $P_s \in \mathbb{S}^{m-1}$. Then this relation gives the desired result. Note that if $M=\mathbb{R}^{3}$, then $\omega(r)=r^2$. Furthermore $\vec{v}_X(\pi_X(p))$ is the unit vector from $X$ to $P=\pi_X(p)$ at $X$. Thus the equation becomes Newton’s shell theorem, which implies that the net gravitational force of a spherical shell acting on any object inside is zero. ### The proof for nCROSS In this section, we prove the main theorem for nCROSS. We use the fact that the first mixed Steklov-Dirichlet eigenfunction, $a\circ r_X$, of the annulus $\mathbf{B}_2\setminus \mathbf{B}_1$ is a test function in both of the variational characterizations of $\sigma_1(\mathbf{B}_2^\prime \setminus \mathbf{B}_1)$ and $\sigma_1(\mathbf{B}_2\setminus \mathbf{B}_1)$. Substituting the test function into the two Rayleigh quotients, we compare the two denominators and the two numerators in the following two propositions.\ Define a map $$\begin{aligned} \int_{\partial\mathbf{B}_2^\prime}(a\circ r_{(\cdot)})^2 dS_2^\prime : \mathbf{B}_2^\prime &\rightarrow \mathbb{R}\end{aligned}$$ that assigns to $X \in \mathbf{B}_2^\prime$ $$\begin{aligned} \int_{\partial\mathbf{B}_2^\prime}(a\circ r_{X})^2 dS_2^\prime\end{aligned}$$ In the following proposition, we show that the function has a minimum value at $C$ by analyzing the gradient of the function at each $X\in \mathbf{B}_2^\prime$, $$\begin{aligned} \nabla \left(\int_{\partial \mathbf{B}_2^\prime} (a\circ r_{(\cdot)})^2 dS_2^\prime \right)(X)\in T_XM.\end{aligned}$$ \[gradientofL2sum\] We have $$\begin{aligned} \nabla \left(\int_{\partial \mathbf{B}_2^\prime} (a\circ r_{(\cdot)})^2 dS_2^\prime \right)(X) = \begin{cases} -g(X)\cdot \vec{v}_X(C) &\text{if}\enspace X\neq C,\\ 0 &\text{if}\enspace X=C, \end{cases}\end{aligned}$$ where $g: \mathbf{B}_2^\prime \setminus\{C\} \rightarrow \mathbb{R}^+$ is a positive function. Furthermore, $$\begin{aligned} \int_{\partial \mathbf{B}_2^\prime} (a\circ r_C)^2 dS_2^\prime \le \int_{\partial \mathbf{B}_2^\prime} (a\circ r_X)^2 dS_2^\prime,\end{aligned}$$and equality holds if and only if $X=C$. The gradient is calculated at $X\in \mathbf{B}_2^\prime$, so it does not affect on the integration region $\partial \mathbf{B}_2^\prime$. Then for $P\in \partial \mathbf{B}_2^\prime$, $\nabla (a\circ r_{(\cdot)}(P))^2(X) \in T_X M$. Thus $$-\nabla \left( \int_{\partial \mathbf{B}_2^\prime} (a\circ r_{(\cdot)})^2(P) dS_2^\prime(P)\right)(X) =\int_{\partial \mathbf{B}_2^\prime}\frac{2(a\circ r_X)}{\omega(r_X)}\cdot(-\nabla r_{(\cdot)}(P)(X))dS_2^\prime(P).$$ With $-\nabla (r_{(\cdot)}(P))(X)=\vec{v}_X(P)$ and Lemma \[radon\], the previous equation is equal to $$\begin{aligned} \label{symmetric identity2} \int_{\mathbb{S}^{m-1}} \left( 2(a\circ r_X) \cdot \frac{\vec{v}_X}{\cos{\lambda_X}} \right) \circ \pi_X d\mu.\end{aligned}$$ \(C) at (0,0); (D) at (2,0); (X) at (-1,0); ; at (-1.2,0) [$X$]{}; (P) at ($(C)!1!60:(D)$); (bP) at ($(C)!1!-60:(D)$); let 1=($(P)-(X)$), 1=[atan2(1,1)]{} in (X)–($(X)+(180+\n1 : 2*2cm)$); let 2=($(bP)-(X)$), 2=[atan2(2,2)]{} in (X)–($(X)+(180+\n2 : 2*2cm)$); (X)–(P); (X)–(bP); (C) circle(2cm); (-P) node\[ left\][$-P$]{}–(X); (-bP) node\[ left\][$-\bar{P}$]{}–(X); (X)–(bP); (X)–(C); (X)–(P); (X)–(bP); (P\_s) at ($(X)!0.8cm!(P)$); (X) – (P\_s); (bP\_s) at ($(X)!0.8cm!(bP)$); (X) – (bP\_s); (X,P\_s); (X,bP\_s); (-P\_s) at ($(X)!0.4cm!(-P)$); (-bP\_s) at ($(X)!0.4cm!(-bP)$); (X) – (-P\_s); (X) – (-bP\_s); (X,-P\_s); (X,-bP\_s); If $X=C$, the integral has value 0. Otherwise, we consider the integrand at $P_s \in \{v|\langle v, c \rangle \ge 0 \}\subset \mathbb{S}^{m-1}$, $\bar{P_s}, -P_s,$ and $-\bar{P_s}$. Note that the condition for $P_s$ is equivalent to $\angle(\vec{v}_X(P),\vec{v}_X(C))\le \frac{\pi}{2}$. Then using Lemma \[symmetry\], $$\begin{aligned} &\left(\left( 2(a\circ r_X) \cdot \frac{\vec{v}_X}{\cos{\lambda_X}} \right)(P)+\left( 2(a\circ r_X) \cdot \frac{\vec{v}_X}{\cos{\lambda_X}} \right)(\bar{P})\right) \\ +&\left(\left( 2(a\circ r_X) \cdot \frac{\vec{v}_X}{\cos{\lambda_X}} \right)(-P)+\left( 2(a\circ r_X) \cdot \frac{\vec{v}_X}{\cos{\lambda_X}} \right)(-\bar{P})\right) \\ =&2(a\circ r_X)(P)\cdot \frac{2\langle\vec{v}_X(P),\vec{v}_X(C)\rangle }{\cos{\lambda_X}}\cdot \vec{v}_X(C)\\ +&2(a\circ r_X)(-P)\cdot \frac{2\langle\vec{v}_X(-P),\vec{v}_X(C)\rangle }{\cos{\lambda_X}}\cdot \vec{v}_X(C)\\ =&4\left( (a\circ r_X)(P)-(a\circ r_X)(-P)\right)\cdot \frac{\langle\vec{v}_X(P),\vec{v}_X(C)\rangle }{\cos{\lambda_X}}\cdot \vec{v}_X(C).\end{aligned}$$ Furthermore, Lemma \[symmetry\] implies $ (a\circ r_X)(P)-(a\circ r_X)(-P)>0$ unless $\angle(\vec{v}_X(P),\vec{v}_X(C)) =\pi/2$. Thus our integration has a form $g(X)\cdot \vec{v}_X(C)$ for some positive function $g$. Note that we actually proved that the gradient of the function has the opposite direction from $X$ to $C$. It implies our desired inequality. - In the proof, the function $g$ only depends on the distance between $X$ and $C$. - The proof is similar to the proof of Corollary \[newton\] if we compare (\[symmetric identity\]) and (\[symmetric identity2\]). The difference between the two proofs is the fact that $a$ is an increasing function. \[L2sum\]We have $$\int_{\partial \mathbf{B}_2} (a\circ r_X)^2 dS_2 \le \int_{\partial \mathbf{B}_2^\prime} (a\circ r_X)^2 dS_2^\prime,$$ where $S_2$ is the measure on $\partial \mathbf{B}_2$ induced from the metric of $M$. The equality holds if and only if $\mathbf{B}_2^\prime = \mathbf{B}_2$. Note that $\mathbf{B}_2$ is a ball of radius $R_2$, centered at $X$. Therefore we have $$\int_{\partial \mathbf{B}_2} (a\circ r_X)^2 dS_2 = \int_{\partial \mathbf{B}_2^\prime} (a\circ r_C)^2 dS_2^\prime.$$ Then Proposition \[gradientofL2sum\] implies the statement. In the following proposition, $(\nabla (a\circ r_X))(Z)$ for $Z\in M \setminus \{X\}$ is the gradient of $$\begin{aligned} a\circ r_X(\cdot) : M\setminus\{X\} \rightarrow \mathbb{R}\end{aligned}$$ at $Z$. \[gradient\] We have $$\int_{\mathbf{B}_2^\prime \setminus \textnormal{cl}(\mathbf{B}_1)}|\nabla(a\circ r_X)|^2dV^\prime \le \int_{\mathbf{B}_2 \setminus \textnormal{cl}(\mathbf{B}_1)}|\nabla(a\circ r_X)|^2 dV,$$ where $V$ and $V^\prime$ are measures on $\mathbf{B}_2$ and $\mathbf{B}_2^\prime$ induced from the metric of $M$, respectively, and equality holds if and only if $\mathbf{B}_2^\prime = \mathbf{B}_2$. Note that $|\nabla(a\circ r_X(\cdot))|=|\nabla a|\circ r_X(\cdot)$ and it is easy to check that $|\nabla a|(r) = |a^\prime(r)| = \frac{1}{\omega(r)}$ is a decreasing function since we only consider when $M$ is nCROSS. Then $$\begin{aligned} &\int_{\mathbf{B}_2 \setminus \textnormal{cl}(\mathbf{B}_1)} |\nabla (a\circ r_X)|^2 dV- \int_{\mathbf{B}_2^\prime \setminus \textnormal{cl}(\mathbf{B}_1)} |\nabla (a\circ r_X)|^2 dV^\prime \\ =&\int_{\mathbf{B}_2 \setminus \mathbf{B}_2^\prime} |\nabla (a\circ r_X)|^2 dV - \int_{\mathbf{B}_2^\prime \setminus \mathbf{B}_2} |\nabla (a\circ r_X)|^2 dV^\prime \\ \ge &\int_{\mathbf{B}_2 \setminus \mathbf{B}_2^\prime} | \nabla a(R_2)|^2 dV - \int_{\mathbf{B}_2^\prime \setminus \mathbf{B}_2} |\nabla a(R_2)|^2 dV^\prime =0. \end{aligned}$$ To satisfy the equality, $|\mathbf{B}_2^\prime \setminus \mathbf{B}_2|=|\mathbf{B}_2 \setminus \mathbf{B}_2^\prime|=0$, or $\mathbf{B}_2^\prime=\mathbf{B}_2$. We used only the fact that $\omega(r)$ is a concave function in $[0,2R_2)$. Thus the proof also applies when $M$ is CROSS and $R_2 < \frac{inj(M)}{4}$. Now we have the following proof of the main theorem when $M$ is a nCROSS. Note that $u\circ r_X =0$ on $\partial \mathbf{B}_1$. By the variational characterization of $\sigma_1(\mathbf{B}_2^\prime \setminus \textnormal{cl}(\mathbf{B}_1))$, $$\sigma_1(\mathbf{B}_2^\prime \setminus\textnormal{cl}(\mathbf{B}_1)) \le \frac{\int_{\mathbf{B}_2^\prime\setminus \textnormal{cl}(\mathbf{B}_1)} |\nabla (a\circ r_X)|^2 dV^\prime}{\int_{\partial \mathbf{B}_2^\prime}(a\circ r_X)^2 dS_2^\prime}.$$ By Corollary \[L2sum\] and Proposition \[gradient\], we have $$\sigma_1(\mathbf{B}_2^\prime \setminus\textnormal{cl}(\mathbf{B}_1)) \le \frac{\int_{\mathbf{B}_2 \setminus \textnormal{cl}(\mathbf{B}_1)} |\nabla (a\circ r_X)|^2 dV}{\int_{\partial \mathbf{B}_2}(a\circ r_X)^2 dS_2}.$$ Since we have shown that $a\circ r_X$ is the first mixed Steklov-Dirichlet eigenfunction on the annulus $\mathbf{B}_2 \setminus \textnormal{cl}(\mathbf{B}_1)$ in Proposition \[firsteigenfunction\], the right hand side is $\sigma_1(\mathbf{B}_2 \setminus \textnormal{cl}(\mathbf{B}_1))$. It is the desired inequality. In addition, the equality condition is followed from the equality conditions in Corollary \[L2sum\] and Proposition \[gradient\]. The method of the proof carries over to Euclidean space $\mathbb{R}^m$. ### The proof for CROSS In this section we modify the proof of Proposition \[gradient\] to show that the inequality in this proposition also holds when $M$ is CROSS and $R_2<\frac{inj(M)}{2}$. Then using the same argument in Section 3.2.2, we can show that the main theorem holds in this situation. $B_r(X)$ denotes the ball of radius $r$, centered at $X$ and $d:=r_X(C)$ denotes the distance between $X$ and $C$. Then the difference between the two sides of the inequality in Proposition \[gradient\] becomes $$\begin{aligned} &\int_{\mathbf{B}_2\setminus \text{cl}(\mathbf{B}_1)}\left(\frac{1}{\omega(r)}\right)^2dV-\int_{\mathbf{B}_2'\setminus \text{cl}(\mathbf{B}_1)}\left(\frac{1}{\omega(r)}\right)^2dV \\ =&\int_{\mathbf{B}_2\setminus \mathbf{B}_2'}\left(\frac{1}{\omega(r)}\right)^2dV-\int_{\mathbf{B}_2'\setminus \mathbf{B}_2}\left(\frac{1}{\omega(r)}\right)^2dV \\ =&\int_{R_2-d}^{R_2}\int_{\pi_X^{-1}((\mathbf{B}_2\setminus \mathbf{B}_2')\cap \partial B_{r_1}(X))}\frac{1}{\omega(r_1)}d\mu dr_1\\ &-\int_{R_2}^{R_2+d}\int_{\pi_X^{-1}((\mathbf{B}_2'\setminus \mathbf{B}_2)\cap \partial B_{r_2}(X))}\frac{1}{\omega(r_2)}d\mu dr_2\\ =&\int_0^d \left( \int_{\pi_X^{-1}((\mathbf{B}_2\setminus \mathbf{B}_2')\cap \partial B_{R_2-s}(X))}\frac{1}{\omega(R_2-s)} \right. \\ &\left. -\int_{\pi_X^{-1}((\mathbf{B}_2'\setminus \mathbf{B}_2)\cap \partial B_{R_2+s}(X))}\frac{1}{\omega(R_2+s)} \right)d\mu ds.\end{aligned}$$ The last equality is obtained by substituting $r_1$ and $r_2$ by $R_2-s$ and $R_2+s$ for $s<d$, respectively. Then the integral becomes nonnegative provided that the following two lemmas hold. \(C) at (0,0); (D) at (2,0); (X) at (-1,0); ; ; at (-1.2,2) [$\mathbf{B}_2$]{}; at (0.2,2) [$\mathbf{B}_2'$]{}; (C) circle(2cm); (X) circle(2cm); (X) circle (0.5cm); (X) circle (2.4cm); (X) circle (1.6cm); (X)–(C); ; ; let 1=($(first intersect)+(1,0)$), 1=[atan2(1,1)]{},2=($(second intersect)+(1,0)$), 2=[atan2(2,2)]{} in (X) + (2 : 2.4) arc (2 : 1 :2.4); let 3=($(third intersect)+(1,0)$), 3=[atan2(3,3)]{},4=($(fourth intersect)+(1,0)$), 4=[atan2(4,4)]{} in (X) + (4+180:-1.6) arc (4+180 : 3-180 :-1.6); (X)to \[dim above = $R_2+s$\] ($(X)!1!-45:(1.4,0)$); (X)to \[dim below = $R_2-s$\] ($(X)!1!1600:(0.6,0)$); \[measure comparison\] We have $$\begin{aligned} |\pi_X^{-1}((\mathbf{B}_2'\setminus \mathbf{B}_2)\cap \partial B_{R_2+s}(X))|\le |\pi_X^{-1}((\mathbf{B}_2\setminus \mathbf{B}_2')\cap \partial B_{R_2-s}(X))|\end{aligned}$$ for $s<R_2$. Consider $S \in (\mathbf{B}_2\setminus \mathbf{B}_2')\cap \partial B_{R_2-s}(X)$. Then the triangle $(SXC)$ has side lengths $$\begin{aligned} |XC|=d, |XS|=R_2-s, |CS|\ge R_2. \end{aligned}$$ Consider the space form $\mathbb{S}_{\kappa}^m$ of constant curvature $\kappa$, where $\kappa\in \mathbb{R}^+$ is a constant such that a geodesic ball of radius $R_2$ is a hemisphere in $\mathbb{S}_{\kappa}^m$. Then we have $$\begin{aligned} \frac{\pi}{2\sqrt{\kappa}}=R_2, \end{aligned}$$ so $\kappa$ is bigger than the sectional curvature of $M$. Now consider a triangle $(S_{\kappa}X_{\kappa}C_{\kappa})$ with the same side lengths as $(SXC)$ in $\mathbb{S}_{\kappa}^m$. Then by the triangle comparison theorem (see p. 197 in [@karcher1989comparison]), $$\begin{aligned} \angle{SXC} \le \angle{S_{\kappa}X_{\kappa}C_{\kappa}}. \end{aligned}$$ Then it implies the following inequality. $$\begin{aligned} \label{measure1} &|\pi_X^{-1}((\mathbf{B}_2\setminus \mathbf{B}_2')\cap \partial B_{R_2-s}(X))|\nonumber\\ =&|\{\pi_X^{-1}(S)| |XS|=R_2-s, |CS|\ge R_2 \}|\nonumber\\ \ge& |\{S_{\kappa}| |X_{\kappa}S_{\kappa}|=R_2-s, |C_{\kappa}S_{\kappa}|\ge R_2\}|\times \frac{1}{s_{\kappa}(R_2-s)}\nonumber\\ =&|\{S_{\kappa}| S_{\kappa} \in ((\mathbf{B}_2)_{\kappa}\setminus (\mathbf{B}_2')_{\kappa})\cap \partial (B_{R_2-s})_{\kappa}(X_{\kappa})\}|\times \frac{1}{s_{\kappa}(R_2-s)}, \end{aligned}$$ where $(\mathbf{B}_2)_{\kappa}$ and $(\mathbf{B}_2')_{\kappa}$ are geodesic balls of radius $R_2$ in $\mathbb{S}_{\kappa}^m$, centered at $X_{\kappa}$ and $C_{\kappa}$, respectively, and $$\begin{aligned} s_{\kappa}(r)=\frac{\sin{\sqrt{\kappa}r}}{\sqrt{\kappa}}. \end{aligned}$$ By a similar argument, we obtain $$\begin{aligned} \label{measure2} &|\pi_X^{-1}((\mathbf{B}_2'\setminus \mathbf{B}_2)\cap \partial B_{R_2+s}(X))|\nonumber\\ \le& |\{S_{\kappa}'| S_{\kappa}' \in ((\mathbf{B}_2')_{\kappa}\setminus (\mathbf{B}_2)_{\kappa})\cap \partial (B_{R_2+s})_{\kappa}(X_{\kappa})\}|\times \frac{1}{s_{\kappa}(R_2+s)} \end{aligned}$$ Since $$\begin{aligned} s_{\kappa}(R_2-s)=s_{\kappa}(R_2+s), \end{aligned}$$ and the set $$\begin{aligned} \{S_{\kappa}| S_{\kappa} \in ((\mathbf{B}_2)_{\kappa}\setminus (\mathbf{B}_2')_{\kappa})\cap \partial (B_{R_2-s})_{\kappa}(X_{\kappa})\} \end{aligned}$$ is the image of the antipodal map in $\mathbb{S}_{\kappa}^m$ of $$\begin{aligned} \{S_{\kappa}'| S_{\kappa}' \in ((\mathbf{B}_2')_{\kappa}\setminus (\mathbf{B}_2)_{\kappa})\cap \partial (B_{R_2+s})_{\kappa}(X_{\kappa})\}, \end{aligned}$$ the right hand sides of (\[measure1\]) and (\[measure2\]) are equal. Thus our desired inequality is obtained. We have $$\begin{aligned} \omega(R_2-s) < \omega(R_2+s)\end{aligned}$$ for $0<s<R_2$. We begin with $M=\mathbb{R}P^n, \mathbb{C}P^n, \mathbb{H}P^n, \mathbb{O}P^2$, which are CROSS except for $\mathbb{S}^m$. Then $s<R_2<\frac{\pi}{4}$. We have two observations of the density function $\omega(t)=(\sin{t})^{m-1}(\cos{t})^{k-1}$ : $$\begin{aligned} \begin{cases} \omega'(t)>0 &\text{if}\enspace t<\arctan{\sqrt{\frac{m-1}{k-1}}},\\ \omega'(t)<0 &\text{if}\enspace t>\arctan{\sqrt{\frac{m-1}{k-1}}}. \end{cases}\end{aligned}$$ and $$\begin{aligned} \omega(t) \le \omega(\frac{\pi}{2}-t)\end{aligned}$$ for $t<\frac{\pi}{4}$. The second observation follows from $$\begin{aligned} \omega(\frac{\pi}{2}-t)-\omega(t) &=(\cos{t})^{m-1}(\sin{t})^{k-1}-(\sin{t})^{m-1}(\cos{t})^{k-1}\\ &=(\sin{t})^{k-1}(\cos{t})^{k-1}((\cos{t})^{m-k}-(\sin{t})^{m-k})>0. \end{aligned}$$ Therefore if $$\begin{aligned} R_2+s<\arctan{\sqrt{\frac{m-1}{k-1}}}, \end{aligned}$$ the first observation implies $$\begin{aligned} \omega(R_2-s)<\omega(R_2+s). \end{aligned}$$ Otherwise, the two observations give $$\begin{aligned} \omega(R_2-s)<\omega(\frac{\pi}{2}-(R_2-s))< \omega(R_2+s). \end{aligned}$$ Therefore the proof for CROSS follows except for $\mathbb{S}^m$. The same proof also works for $\mathbb{S}^m$ if we replace $\frac{\pi}{4}$ and $\frac{\pi}{2}$ by $\frac{\pi}{2}$ and $\pi$, respectively.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Recent results establish the optimality of interference alignment to approach the Shannon capacity of interference networks at high SNR. However, the extent to which interference can be aligned over a finite number of signalling dimensions remains unknown. Another important concern for interference alignment schemes is the requirement of global channel knowledge. In this work we provide examples of iterative algorithms that utilize the reciprocity of wireless networks to achieve interference alignment with only local channel knowledge at each node. These algorithms also provide numerical insights into the feasibility of interference alignment that are not yet available in theory.' author: - '\' bibliography: - 'biblio.bib' title: Approaching the Capacity of Wireless Networks through Distributed Interference Alignment --- Introduction ============ The recent emergence of the idea of interference alignment for wireless networks has shown that the capacity of wireless networks can be much higher than previously believed [@Cadambe_Jafar_int]. The canonical example of interference alignment is a communication scenario where, regardless of the number of interferers, every user is able to access one half of the spectrum free from interference from other users [@Cadambe_Jafar_int]. For the interference channel with $K$ transmitters and $K$ receivers and random, time varying channel coefficients drawn from a continuous distribution, reference [@Cadambe_Jafar_int] characterizes the network sum capacity as $$C_\Sigma(SNR) = \frac{K}{2} \log(SNR) + o(\log(SNR))\label{eq:dofint}$$ so that the capacity per user is $\frac{1}{2}\log(SNR)+o(\log(SNR))$. Here SNR is defined as the total transmit power of all the transmitters in the network when the local noise power at each node is normalized to unity. The $o(\log(SNR))$ term, by definition, becomes negligible compared to $\log(SNR)$ at high SNR. Therefore the accuracy of the capacity approximation in (\[eq:dofint\]) approaches $100\%$ at high SNR. Since the capacity of a single user in the absence of all interference is $\log(SNR)+o(\log(SNR))$, the main result of [@Cadambe_Jafar_int] may be summarized as: “*At high SNR, every user in a wireless interference network is (simultaneously and almost surely) able to achieve (nearly) one half of the capacity that he could achieve in the absence of all interference.*" The capacity-optimal achievable scheme within $o(\log(\mbox{SNR}))$ is shown to be the interference alignment scheme. Interference alignment on the $K$ user interference channel refers to the idea of constructing signals in such a way that they cast overlapping shadows over one half of the signal space observed by each receiver where they constitute interference, leaving the other half of the signal space free of interference for the desired signal. This approach reveals the sub-optimality of the cake-cutting view of spectrum allocation between co-existing wireless systems because, essentially, everyone gets “half the cake”. Interference alignment schemes are presented in [@Cadambe_Jafar_int] in the form of closed form expressions for the transmit precoding matrices. However, these closed form expressions require global channel knowledge which can be an overwhelming overhead in practice. Moreover, closed form solutions have only been found in certain cases. In general, analytical solutions to interference alignment problem are difficult to obtain and even the feasibility of interference alignment over a limited number of signalling dimensions is an open problem. In this paper we explore distributed interference alignment algorithms to accomplish the following objectives. - Require only local channel knowledge at each node. Specifically, each receiver is assumed to know only the channel to its desired transmitter and the covariance matrix of its effective noise (consisting of the AWGN and the interference from all other users). - Provide numerical insights into the feasibility of alignment. We propose iterative algorithms that take a cognitive approach to interference management and utilize only the local side information available naturally due to the reciprocity of wireless networks. The two key properties can be summarized as follows. - [*Cognitive Principle:*]{} Unlike selfish approaches studied in prior work where each transmitter tries to maximize his own rate by transmitting along those signaling dimensions where his desired receiver sees the least interference, we follow an unselfish approach where each transmitter primarily tries to minimize the interference to unintended receivers. Since avoiding interference to unintended receivers is the defining feature of cognitive radio [@cogtutorial] the unselfish approach is a cognitive approach. The cognitive approach is found to lead to interference alignment, and is thus capable of approaching network capacity at high SNR. - [*Reciprocity:*]{} For a given transmitter, learning how much interference is caused at unintended receivers can require too much side information, and is one of the key challenges for cognitive radio systems. However, this information is naturally available because of the reciprocity of the channel for networks where two-way communication is based on time-division duplex operation with synchronized time-slots. Due to reciprocity, the signalling dimensions along which a receiving node sees the least interference from other users are also the same signalling dimensions along which this node will cause the least interference to other nodes in the reciprocal network where all transmitters and receivers switch roles. The paper is organized as follows. In the next section, we review some optimization approaches for interference networks in existing literature. In Section \[sec:recalign\] we state the open problem of determining the feasibility of interference alignment over a limited number of signaling dimensions. The same section also describes the reciprocity property of interference alignment. In Section \[sec: algo\] we present iterative interference alignment algorithms. In Section \[sec: results\], we show that these algorithm can achieve performance close to the theoretical results and discuss a few applications. We conclude with Section \[sec: concl\]. [*Notation:*]{} We use lower case for scalars, upper case for vectors and bold font to denote matrices. ${\bf A}_{\star d}$ represents the $d^{th}$ column of matrix ${\bf A}$. ${\bf I}_{d}$ represents the $d\times d$ identity matrix. Tr$[{\bf A}]$ denotes the trace of the matrix ${\bf A}$ and ${\bf A}^\dagger$ is the conjugate transpose of matrix ${\bf A}$. Finally, $\mathcal{K}\triangleq\{1,2,\cdots, K\}$ is the index set of $K$ users. Interference Optimization Approaches ==================================== The optimality of interference alignment schemes at high SNR is interesting because these schemes treat all interference as noise and require no multi-user detection. Achievable schemes based on treating interference as noise have been explored extensively over the last decade. Prominent among these are the interference avoidance and iterative waterfilling algorithms where each transmitter acts selfishly to align its transmissions along those directions where its desired receiver sees the least interference [@rose_ia; @rose_interfering; @game_fair; @ia_game; @iterative_ulukus_yates], and network duality approaches [@wc_book_tse; @transmit_beamforming_farrokhi; @network_dual_rao; @babadi-2007] that are based on the reciprocity of the wireless propagation channel. Interference Avoidance and Iterative Waterfilling ------------------------------------------------- Iterative algorithms are commonly used for various resource allocation problems, such as “interference avoidance” and “iterative waterfilling”. However, the philosophy of interference alignment is quite distinct from both iterative waterfilling and interference avoidance. With iterative waterfilling/interference avoidance algorithms [@yu_iterative_waterfilling; @rose_ia; @etkin_tse_2007], each transmitter tries to do what is best for his own receiver, i.e., each transmitter allocates its power in a manner best suited for his desired receiver. With interference alignment each transmitter tries to minimize the interference he causes to other receivers. The interference alignment schemes in [@Jafar_Shamai; @Cadambe_Jafar_int; @Cadambe_Jafar_X; @Cadambe_Jafar_XFB] show that for interference networks, the “do no harm” approach is much more powerful, and is in fact capacity-optimal within $o(\log(\mbox{SNR}))$, than the “help yourself” approach of interference avoidance and iterative waterfilling schemes. From a game-theoretic perspective, interference avoidance and iterative waterfilling algorithms lead to a stable operating point commonly known as the Nash equilibrium. At Nash equilibrium, there is no incentive for any user to unilaterally change his transmit strategy. Often these points are not optimum from a network perspective and indicate inefficiency in wireless network operation [@comp_coll_miso_int; @leshem-2007]. Interestingly, interference alignment is not a Nash equilibrium point if the goal of each user is to maximize his own rate. Fig. \[fig: nash\] shows the interference alignment solution for the three user two antenna case. Notice that interfering signals are co-linear at each receiver while the desired signal may not be exactly orthogonal to the interference - a price paid for interference alignment. It can be easily observed that the interference alignment solution is not a Nash equilibrium point. Fixing the transmit strategy for users 2 and 3, the best strategy for user 1 is to choose ${\bf u}_1$ such that his signal is orthogonal to the interference at receiver 1. Although this strategy is good for user 1, it will destroy interference alignment at receivers 2 and 3. Thus Fig. \[fig: nash\] clearly highlights the difference between the optimal strategy of interference alignment and the selfish strategy of iterative waterfilling or interference avoidance. Iterative schemes have also been used to implicitly achieve interference alignment on the $2$ user $X$ channel in [@MMK; @MMKreport1; @MMKreport2]. However, for the $2$ user $X$ channel interference alignment can be explicitly achieved with roughly the same amount of channel knowledge as required by the iterative schemes, without the need for an iterative process [@Jafar_Shamai]. The iterative schemes of [@MMK] are specialized for the $2$ user $X$ channel and generalizations to $X$ networks and interference networks with more than $2$ users are not straightforward. Network Duality --------------- Another approach taken in prior work is to exploit the duality relationships enabled by the reciprocity of the propagation channel. For example, network duality ensures that the same set of signal to interference and noise ratios (SINRs) can be achieved in the original and the reciprocal network with the same total transmit power [@wc_book_tse; @transmit_beamforming_farrokhi]. Network duality is used in [@transmit_beamforming_farrokhi; @network_dual_rao] to minimize the total transmit power required to support a feasible rate vector. Reciprocity of propagation channels is used in [@babadi-2007] for optimal frequency allocation problem. In this work we provide examples of iterative algorithms to achieve interference alignment on wireless interference channels. These algorithms combine elements of all the above-mentioned approaches, especially [@MMK] and [@babadi-2007]. System Model ============ Consider the $K$-user MIMO interference channel where the $k^{th}$ transmitter and receiver are equipped with $M^{[k]}$ and $N^{[k]}$ antennas respectively. Note that the antennas could represent symbol extensions in time or frequency as well. However, if the antennas correspond to symbol extensions over orthogonal dimensions (time, frequency slots) then the channel matrices will have a diagonal structure. The channel is defined as: $$\begin{aligned} Y^{[k]}(n)=\sum_{l=1}^K {\bf H}^{[kl]}(n)X^{[l]}(n)+Z^{[k]}(n), ~~\forall k\in\mathcal{K}\end{aligned}$$ where, at the $n^{th}$ channel use, $Y^{[k]}(n), Z^{[k]}(n)$ are the $N^{[k]}\times 1$ received signal vector and the zero mean unit variance circularly symmetric additive white Gaussian noise vector (AWGN) at receiver $k$, $X^{[l]}(n)$ is the $M^{[l]}\times 1$ signal vector transmitted by transmitter $l$, and ${\bf H}^{[kl]}(n)$ is the $N^{[k]}\times M^{[l]}$ matrix of channel coefficients between transmitter $l$ and receiver $k$. The transmit power at transmitter $l$ is $\mbox{E}[||X^{[l]}||^2]=P^{[l]}$. For the $K$ user interference channel defined above, we also define a reciprocal channel, where the role of transmitters and receivers are switched. For every variable on the original channel, the corresponding variable on the reciprocal channel is denoted with a left arrow on top. The reciprocal channel is defined as: $$\begin{aligned} \overleftarrow{Y}^{[k]}(n)=\sum_{l=1}^K \overleftarrow{\bf H}^{[kl]}(n)\overleftarrow{X}^{[l]}(n)+\overleftarrow{Z}^{[k]}(n), ~~\forall k\in\mathcal{K}\end{aligned}$$ where, at the $n^{th}$ channel use, $\overleftarrow{Y}^{[k]}(n), \overleftarrow{Z}^{[k]}(n)$ are the $M^{[k]}\times 1$ received signal vector and the zero mean unit variance circularly symmetric additive white Gaussian noise vector (AWGN) at receiver $k$ which is equipped with $M^{[k]}$ antennas, $\overleftarrow{X}^{[l]}(n)$ is the $N^{[l]}\times 1$ signal vector transmitted by transmitter $l$, and $\overleftarrow{\bf H}^{[kl]}(n)={\bf H}^{[lk]\dagger}(n)$ is the $M^{[k]}\times N^{[l]}$ matrix of channel coefficients between transmitter $l$ and receiver $k$. The transmit powers at transmitter $l$ on the reciprocal channel is $\mbox{E}[||\overleftarrow{X}^{[l]}||^2=\overleftarrow{P}^{[l]}$. The channel use index $n$ is henceforth suppressed to avoid cumbersome notation. Interference Alignment over Limited Dimensions - An Open Problem {#sec:recalign} ================================================================ References [@Cadambe_Jafar_int] and [@Cadambe_Jafar_X] present interference alignment schemes that are constructed over symbol extensions of time-varying channels. It is shown that by using *long* symbol extensions the degrees of freedom achieved per dimension approach arbitrarily close to the theoretical outerbound, thereby establishing the degrees of freedom of time-varying interference and $X$ networks. However, the extent to which interference can be aligned over a *limited* number of dimensions remains an open problem. As a consequence, the maximum number of degrees of freedom that can be achieved through alignment of interference signal vectors is not known in general. Note that interference alignment can also be accomplished in terms of signal *levels* rather than signal vectors by using structured codes (multilevel and lattice codes) as shown in [@Bresler_Parekh_Tse; @Cadambe_Jafar_Shamai]. However, in this work our focus is on interference alignment through signal vectors. We first review the interference alignment problem and the reciprocity property of interference alignment. Let $d^{[k]}\leq\min(M^{[k]},N^{[k]}), k\in\mathcal{K}$ denote the degrees of freedom for user $k$’s message. [*Precoding at Transmitter:*]{} Let ${\bf V}^{[k]}$ be an $M^{[k]}\times d^{[k]}$ matrix whose columns are the orthonormal basis of the transmitted signal space of user $k$. Mathematically, the transmitted signal vector of user $k$ is given by: $$\begin{aligned} {X}^{[k]}=\sum_{d=1}^{d^{[k]}}{\bf V}^{[k]}_{[\star d]}\overline{X}^{[k]}_d={\bf V}^{[k]}{\overline{X}}^{[k]}, ~~\overline{X}^{[k]}\sim\mathcal{N}\left(0,\frac{P^{[k]}}{d^{[k]}}{\bf I}_{d^{[k]}}\right)\end{aligned}$$ Each element of the $d^{[k]}\times 1$ vector ${\overline{X}}^{[k]}$ represents an independently encoded Gaussian codebook symbol with power $\frac{P^{[k]}}{d^{[k]}}$ that is beamformed with the corresponding vector of ${\bf V}^{[k]}$. [*Remark:*]{} For interference alignment or the achievability of the degrees of freedom it suffices if the beamforming vectors are linearly independent. However, we assume the beamforming vectors are orthonormal in the formulation above. Note that this does not affect the feasibility of interference alignment, and it naturally leads to the iterative algorithm to be presented in this paper. [*Interference Suppression at Receiver:*]{} Let ${\bf U}^{[k]}$ be an $N^{[k]}\times d^{[k]}$ matrix whose columns are the orthonormal basis of the interference-free desired signal subspace at receiver $k$. The $k^{th}$ receiver filters its received signal to obtain: $$\begin{aligned} \overline{Y}^{[k]}={\bf U}^{[k]\dagger} Y^{[k]}\end{aligned}$$ If interference is aligned into the null space of ${\bf U}^{[k]}$ then the following condition must be satisfied: $$\begin{aligned} {\bf U}^{[k]\dagger}{\bf H}^{[kj]}{\bf V}^{[j]} &=& 0, \forall j\neq k\\ \mbox{rank}\left({\bf U}^{[k]\dagger}{\bf H}^{[kk]}{\bf V}^{[k]}\right)&=&d_k\end{aligned}$$ In other words the desired signals are received through a $d^{[k]}\times d^{[k]}$ full rank channel matrix $$\overline{\bf H}^{[kk]}\triangleq {\bf U}^{[k]\dagger}{\bf H}^{[kk]}{\bf V}^{[k]}$$ while the interference is completely eliminated. The effective channel for user $k$ is then expressed as: $$\begin{aligned} \overline{Y}^{[k]}&=&\overline{\bf H}^{[kk]}{\overline{X}}^{[k]}+\overline{Z}^{[k]}\end{aligned}$$ where $\overline{Z}^{[k]} \sim\mathcal{N}\left(0,I\right)$ is the effective $d^{[k]}\times 1$ AWGN vector at receiver $k$. The rate achieved on this channel is: $$\begin{aligned} R^{[k]}&=&\log\left|{\bf I}_{d^{[k]}}+\frac{P^{[k]}}{d^{[k]}}\overline{\bf H}^{[kk]}\overline{\bf H}^{[kk]\dagger}\right|\\&=&d^{[k]}\log(P^{[k]})+o(\log(P^{[k]}))\end{aligned}$$ Thus, $d^{[k]}$ degrees of freedom are achieved by user $k$. Feasibility of Alignment ------------------------ Given the channel matrices ${\bf H}^{[kj]}, k,j\in\mathcal{K}$, we say that the degrees of freedom allocation $(d^{[1]}, d^{[2]}, \cdots, d^{[K]})$ is feasible if there exist transmit precoding matrices ${\bf V}^{[k]}$ and receive interference suppression matrices ${\bf U}^{[k]}$: $$\begin{aligned} {\bf V}^{[k]}:M^{[k]}\times d^{[k]},&&{\bf V}^{[k]\dagger}{\bf V}^{[k]}={\bf I}_{d^{[k]}}\\ {\bf U}^{[k]}:N^{[k]}\times d^{[k]},&&{\bf U}^{[k]\dagger}{\bf U}^{[k]}={\bf I}_{d^{[k]}}\end{aligned}$$ such that $$\begin{aligned} {\bf U}^{[k]\dagger}{\bf H}^{[kj]}{\bf V}^{[j]} &=& 0, \forall j\neq k\label{eq:cross}\\ \mbox{rank}\left({\bf U}^{[k]\dagger}{\bf H}^{[kk]}{\bf V}^{[k]}\right)&=&d_k, ~\forall k\in\mathcal{K} \label{eq:direct}\end{aligned}$$ [*Remark:*]{} Suppose all the elements of the channel matrices are randomly and independently generated from continuous distributions and ${\bf V}^{[k]},{\bf U}^{[k]}, k\in\mathcal{K}$ can be found to satisfy condition (\[eq:cross\]). Then, condition (\[eq:direct\]) will also be satisfied with probability 1. This is because the direct channel matrices ${\bf H}^{[kk]}$ do not appear in condition (\[eq:cross\]). So the choice of transmit and receive filters ${\bf V}^{[k]},{\bf U}^{[k]}, k\in\mathcal{K}$ to satisfy (\[eq:cross\]) does not depend on the direct channel matrices ${\bf H}^{[kk]}$. Since ${\bf H}^{[kk]}$ is independent of ${\bf V}^{[k]},{\bf U}^{[k]}$ and all its elements are randomly generated from a continuous distribution (i.e. it lacks any special structure), the product matrix ${\bf U}^{[k]T}{\bf H}^{[kk]}{\bf V}^{[k]}$ has full rank with probability $1$. Thus, for random MIMO channels without time-extensions, if (\[eq:cross\]) can be satisfied then (\[eq:direct\]) is automatically satisfied almost surely as well. However, if time-extensions are considered then the channel matrices may have a block diagonal structure and (\[eq:direct\]) cannot be taken for granted. The solution to the feasibility problem is not known in general. In other words, given a set of randomly generated channel matrices and a degree-of-freedom allocation $(d^{[1]}, d^{[2]}, \cdots, d^{[K]})$, it is not known if one can almost surely find transmit and receive filters that will satisfy the feasibility conditions. The distributed interference algorithm developed in this paper will be useful in (numerically) solving this open problem. Reciprocity of Alignment ------------------------ An interesting observation from the problem formulation above is the duality relationship between interference alignment on a given interference channel and its reciprocal channel obtained by switching the direction of communication. Specifically, let $\overleftarrow{\bf V}^{[k]}, \overleftarrow{\bf U}^{[k]}$ denote the transmit precoding filters and the receive interference suppression filters on the reciprocal channel. The feasibility conditions on the reciprocal channel are: $$\begin{aligned} \overleftarrow{\bf V}^{[k]}:N^{[k]}\times d^{[k]},&&\overleftarrow{\bf U}^{[k]\dagger}\overleftarrow{\bf U}^{[k]}={\bf I}_{d^{[k]}}\\ \overleftarrow{\bf U}^{[k]}:M^{[k]}\times d^{[k]},&&\overleftarrow{\bf U}^{[k]\dagger}\overleftarrow{\bf U}^{[k]}={\bf I}_{d^{[k]}}\end{aligned}$$ such that $$\begin{aligned} \overleftarrow{\bf U}^{[j]\dagger}\overleftarrow{\bf H}^{[jk]}\overleftarrow{\bf V}^{[k]} &=& 0, \forall j\neq k\\ \mbox{rank}\left(\overleftarrow{\bf U}^{[k]\dagger}\overleftarrow{\bf H}^{[kk]}\overleftarrow{\bf V}^{[k]}\right)&=&d_k, ~\forall k\in\mathcal{K}\end{aligned}$$ Suppose we set $\overleftarrow{\bf V}^{[k]}={\bf U}^{[k]}, \overleftarrow{\bf U}^{[k]}={\bf V}^{[k]}$. Then the feasibility conditions on the reciprocal channel become identical to the original feasibility conditions. Thus, the following observation can be made: [**Reciprocity of Alignment:**]{} [*Since the feasibility conditions are identical, if the degrees of freedom allocation $(d^{[1]}, d^{[2]}, \cdots, d^{[K]})$ is feasible on the original interference network then it is also feasible on the reciprocal network (and vice versa). Interference alignment on the reciprocal interference network is simply achieved by choosing the transmit filters and the receive filters on the reciprocal channel as the receive filters and the transmit filters (respectively) of the original channel.*]{} Reciprocity of alignment is a key property used for distributed interference alignment algorithms, described in the next section. Distributed algorithm for interference alignment {#sec: algo} ================================================ In this section we construct distributed interference alignment algorithms for the interference channel with multiple-antenna nodes and no symbol extensions. Continuing with the system model of Section \[sec:recalign\], this implies that the relevant interference alignment feasibility condition is (\[eq:cross\]) while (\[eq:direct\]) is automatically satisfied. Basically, (\[eq:cross\]) requires that at each receiver, all interference is suppressed, leaving as many interference-free dimensions as the degrees of freedom allocated to that receiver. Since we are interested in distributed algorithms, we start with arbitrary transmit and receive filters ${\bf V}^{[k]}, {\bf U}^{[k]}$ and iteratively update these filters to approach interference alignment. The quality of alignment is measured by the power in the *leakage interference* at each receiver, i.e. the interference power remaining in the received signal after the receive interference suppression filter is applied. The goal is to achieve interference alignment by progressively reducing the leakage interference. If interference alignment is feasible then eventually leakage interference will be zero. The total interference leakage at receiver $k$ due to all undesired transmitters ($j \neq k$) is given by: $$\begin{aligned} I^{[k\star]}=\mbox{Tr}\left[{\bf U}^{[k]\dagger}{\bf Q}^{[k]}{\bf U}^{[k]}\right]\end{aligned}$$ where $$\begin{aligned} {\bf Q}^{[k]}=\sum_{j=1,j\neq k}^K \frac{P^{[j]}}{d^{[j]}}{\bf H}^{[kj]}{\bf V}^{[j]}{\bf V}^{[j]\dagger}{\bf H}^{[kj]\dagger}\end{aligned}$$ is the interference covariance matrix at receiver $k$. Similarly, in the reciprocal network, the total interference leakage at receiver $j$ due to all undesired transmitters ($k \neq j$) is given by: $$\begin{aligned} \overleftarrow{I}^{[j\star]}=\mbox{Tr}\left[\overleftarrow{\bf U}^{[j]\dagger}\overleftarrow{\bf Q}^{[j]}\overleftarrow{\bf U}^{[j]}\right]\end{aligned}$$ where $$\begin{aligned} \overleftarrow{\bf Q}^{[j]}=\sum_{k=1,k\neq j}^K \frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\overleftarrow{\bf H}^{[jk]}\overleftarrow{\bf V}^{[k]}\overleftarrow{\bf V}^{[k]\dagger}\overleftarrow{\bf H}^{[jk]\dagger}\end{aligned}$$ is the interference covariance matrix at receiver $j$. The iterative algorithm alternates between the original and reciprocal networks. Within each network only the receivers update their interference suppression filters to minimize their total leakage interference. [*Step I:*]{} In the original network, each receiver solves the following optimization problem. $$\begin{aligned} \min_{{\bf U}^{[k]}:N^{[k]}\times d^{[k]}, ~~{\bf U}^{[k]}{\bf U}^{[k]\dagger}={\bf I}_{d^{[k]}}} I^{[k\star]}\end{aligned}$$ In other words, receiver $k$ chooses its interference suppression filter ${\bf U}^{[k]}$ to minimize the leakage interference due to all undesired transmitters. The $d^{[k]}$ dimensional received signal subspace that contains the least interference is the space spanned by the eigenvectors corresponding to the $d^{[k]}$ smallest eigenvalues of the interference covariance matrix ${\bf Q}^{[k]}$. Thus, the $d^{[k]}$ columns of ${\bf U}^{[k]}$ are given by: $${\bf U}^{[k]}_{\star d}={\bf \nu}_{d}[{\bf Q}^{[k]}],~~~~ d=1, \cdots,d^{[k]} \label{eqn: basis_vectors}$$ where ${\bf \nu}_{d}[{\bf A}]$ is the eigenvector corresponding to the $d^{th}$ smallest eigenvalue of $\bf A$. [*Step II:*]{} The second step is identical to the first step, but performed in the reciprocal network. Consider the reciprocal network obtained by reversing the roles of the transmitters and the receivers. The transmit precoding matrices in the reciprocal network, $\overleftarrow{\bf V}^{[k]}$, are the receive interference suppression matrices ${\bf U}^{[k]}$ from the original network that were determined in Step I. Each receiver in the reciprocal network solves the following optimization problem. $$\begin{aligned} \min_{\overleftarrow{\bf U}^{[j]}:M^{[j]}\times d^{[j]}, ~~\overleftarrow{\bf U}^{[j]}\overleftarrow{\bf U}^{[j]\dagger}={\bf I}_{d^{[j]}}} \overleftarrow{I}^{[j\star]}\end{aligned}$$ Similar to Step I, the $d^{[j]}$ columns of $\overleftarrow{\bf U}^{[j]}$ are given by: $$\overleftarrow{\bf U}^{[j]}_{\star d}={\bf \nu}_{d}[\overleftarrow{\bf Q}^{[j]}],~~~~ d=1, \cdots,d^{[j]}$$ The receive interference suppression filters in the reciprocal network are then used as the transmit precoding matrices in the original network, and the algorithm returns to Step I. The iterations continue in this manner until the algorithm converges. The iterative procedure is summarized in Algorithm \[alg:iter\_ia\]. A pictorial representation is shown in Fig. \[fig: algo\]. Start with arbitrary precoding matrices ${\bf V}^{[j]}: M^{[j]}\times d^{[j]}, {\bf V}^{[j]}{\bf V}^{[j]\dagger}={\bf I}_{d^{[j]}}$. Begin iteration Compute interference covariance matrix at the receivers: $${\bf Q}^{[k]}=\sum_{j=1,j\neq k}^K \frac{P^{[j]}}{d^{[j]}}{\bf H}^{[kj]}{\bf V}^{[j]}{\bf V}^{[j]\dagger}{\bf H}^{[kj]\dagger}$$ Compute the interference suppression matrix at each receiver: $${\bf U}^{[k]}_{\star d}={\bf \nu}_{d}[{\bf Q}^{[k]}],~~~~ d=1, \cdots,d^{[k]}\nonumber$$ Reverse the communication direction and set $\overleftarrow{\bf V}^{[k]}={\bf U}^{[k]}$. Compute interference covariance matrix at the new receivers: $$\overleftarrow{\bf Q}^{[j]}=\sum_{k=1,k\neq j}^K \frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\overleftarrow{\bf H}^{[jk]}\overleftarrow{\bf V}^{[k]}\overleftarrow{\bf V}^{[k]\dagger}\overleftarrow{\bf H}^{[jk]\dagger}$$ Compute the interference suppression matrix at each receiver: $$\overleftarrow{\bf U}^{[j]}_{\star d}={\bf \nu}_{d}[\overleftarrow{\bf Q}^{[j]}],~~~~ d=1, \cdots,d^{[k]}\nonumber$$ Reverse the communication direction and set ${\bf V}^{[k]}=\overleftarrow{\bf U}^{[k]}$. Continue till convergence. Proof of Convergence -------------------- We now show that the algorithm must converge. The proof also highlights the intuition behind the algorithm. We define a metric called the weighted leakage interference (WLI) as: $$\begin{aligned} I_{w}&=&\sum_{k=1}^{K}\sum_{j=1, j\neq k}^K\frac{\overleftarrow{P}^{[k]}}{d^{[k]}}I^{[kj]}\\ &=&\sum_{k=1}^K\sum_{j=1, j\neq k}^K\frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\frac{{P}^{[j]}}{d^{[j]}}\mbox{Tr}\left[{\bf U}^{[k]\dagger}{\bf H}^{[kj]}{\bf V}^{[j]}{\bf V}^{[j]\dagger}{\bf H}^{[kj]\dagger}{\bf U}^{[k]}\right]\end{aligned}$$ We show that each step in the algorithm reduces the value of WLI. Since WLI is bounded below by zero, this implies that the algorithm must converge. Note that an interference alignment solution corresponds to WLI$=0$. The WLI associated with receiver $k$ is $$\begin{aligned} I^{[k\star]}_{w}&=&\frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\sum_{j=1, j\neq k}^K\frac{{P}^{[j]}}{d^{[j]}}\mbox{Tr}\left[{\bf U}^{[k]\dagger}{\bf H}^{[kj]}{\bf V}^{[j]}{\bf V}^{[j]\dagger}{\bf H}^{[kj]\dagger}{\bf U}^{[k]}\right]\nonumber \label{eqn:leakkstar}\\ &=&\frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\mbox{Tr}\left[{\bf U}^{[k]\dagger}{\bf Q}^{[k]}{\bf U}^{[k]}\right]=\frac{\overleftarrow{P}^{[k]}}{d^{[k]}}I^{[k\star]}\nonumber \end{aligned}$$ Therefore the value of ${\bf U}^{[k]}$ computed in Step 4 to minimize $I^{[k\star]}$ also minimizes $I^{[k\star]}_{w}$. Since $I_w=\sum_{k=1}^KI^{[k\star]}_w$, we have $$\begin{aligned} \min_{{\bf U}^{[1]},{\bf U}^{[2]}, \cdots, {\bf U}^{[K]}} I_w &=& \min_{{\bf U}^{[1]},{\bf U}^{[2]}, \cdots, {\bf U}^{[K]}} \sum_{k=1}^KI_w^{[k\star]}\\ &=&\sum_{k=1}^K \left[\min_{{\bf U}^{[k]}}I_w^{[k\star]}\right]=\sum_{k=1}^K \frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\left[\min_{{\bf U}^{[k]}}I^{[k\star]}\right]\end{aligned}$$ In other words, given the values of ${\bf V}^{[j]}, j\in\{1,2,\cdots, K\}$, Step 4 minimizes the value of $I_w$ over all possible choices of ${\bf U}^{[k]}, k\in\{1,2,\cdots, K\}$. In particular, Step 4 can only reduce the value of $I_w$. The weighted leakage interference associated with transmitter $j$ is [$$\begin{aligned} I^{[\star j]}_w&=&\frac{{P}^{[j]}}{d^{[j]}}\sum_{k=1}^K\frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\mbox{Tr}\left[{\bf U}^{[k]\dagger}{\bf H}^{[kj]}{\bf V}^{[j]}{\bf V}^{[j]\dagger}{\bf H}^{[kj]\dagger}{\bf U}^{[k]}\right]\\ &=&\frac{{P}^{[j]}}{d^{[j]}}\sum_{k=1}^K\frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\mbox{Tr}\left[\overleftarrow{\bf V}^{[k]\dagger}\overleftarrow{\bf H}^{[jk]\dagger}\overleftarrow{\bf U}^{[j]}\overleftarrow{\bf U}^{[j]\dagger}\overleftarrow{\bf H}^{[jk]}\overleftarrow{\bf V}^{[k]}\right]\\ &=&\frac{{P}^{[j]}}{d^{[j]}}\sum_{k=1}^K\frac{\overleftarrow{P}^{[k]}}{d^{[k]}}\mbox{Tr}\left[\overleftarrow{\bf U}^{[j]\dagger}\overleftarrow{\bf H}^{[jk]}\overleftarrow{\bf V}^{[k]}\overleftarrow{\bf V}^{[k]\dagger}\overleftarrow{\bf H}^{[jk]\dagger}\overleftarrow{\bf U}^{[j]}\right]\\ &=&\frac{{P}^{[j]}}{d^{[j]}}\mbox{Tr}\left[\overleftarrow{\bf U}^{[j]\dagger}\overleftarrow{\bf Q}^{[j]}\overleftarrow{\bf U}^{[j]}\right]\\\end{aligned}$$ ]{} Therefore the value of $\overleftarrow{\bf U}^{[j]}$ computed in Step 7 to minimize $\overleftarrow{I}^{[j\star]}$ also minimizes $I^{[\star j]}_{w}$. Since $I_w=\sum_{k=1}^KI^{[\star j]}_w$, it is easily seen that Step 7 can also only reduce the value of $I_w$. Since the value of $I_w$ is monotonically reduced after every iteration, convergence of the algorithm is guaranteed. [*Remark:*]{} While the algorithm minimizes leakage interference at every iteration and is guaranteed to converge, convergence to global minimum is not guaranteed due to the non-convex nature of the interference optimization problem. Numerical results for the performance of the algorithm are presented in the next section. The following observations summarize the intuition behind the iterative algorithm. 1. Dimensions along which a receiver sees the least interference from other nodes are also the dimensions along which it causes the least interference to other nodes in the reciprocal network where it functions as a transmitter. 2. The weighted leakage interference is unchanged in the original and reciprocal networks if the transmit and receive filters are switched. Max-SINR Algorithm ------------------ The algorithm presented above seeks perfect interference alignment. In particular it seeks to create an interference-free subspace of the required number of dimensions, that is designated as the *desired* signal subspace. However, note that interference alignment makes no attempt to maximize the desired signal power within the desired signal subspace. In fact the algorithm described above does not depend at all on the direct channels ${\bf H}^{[kk]}$ through which the desired signal arrives at the intended receiver. Therefore, while the interference is eliminated within the desired space, no coherent combining gain (array gain) for the desired signal is obtained with interference alignment. While this is optimal as all signal powers approach infinity, it is not optimal in general at intermediate SNR values. Therefore other algorithms may be designed which will perform better than the interference alignment algorithm at intermediate SNR values. In this section we consider one such natural extension of the interference alignment algorithm where the receive filters ${\bf U}^{[k]}$ and $\overleftarrow{\bf U}^{[k]}$ are chosen to maximize SINR at the receivers instead of only minimizing the leakage interference. While there is no loss of generality in assuming orthogonal precoding vectors for the streams sent from the same transmitter as far as interference alignment is concerned, orthogonal precoding vectors are in general suboptimal for SINR optimization. We therefore no longer assume that the columns of ${\bf V}^{[k]}$ (the transmit precoding vectors) are mutually orthogonal. We also identify the columns of ${\bf U}^{[k]}$ to be the specific combining vectors for the corresponding desired data stream, so that they are not necessarily orthogonal either. With these modified definitions, the SINR of the $l^{th}$ stream of the $k^{th}$ receiver is $$\mbox{SINR}_{kl}=\frac{{\bf U}^{[k]\dagger}_{\star l}{\bf H}^{[kk]}{\bf V}^{[k]}_{\star l}{\bf V}^{[k]\dagger}_{\star l}{\bf H}^{[kk]\dagger}{\bf U}^{[k]}_{\star l}}{{\bf U}^{[k]\dagger}_{\star l}{\bf B}^{[kl]}{\bf U}^{[k]}_{\star l}}\frac{P^{[k]}}{d^{[k]}}$$ where $$\begin{aligned} \small {\bf B}^{[kl]}&=& \sum_{j=1}^{K}\frac{P^{[j]}}{d^{[j]}}\sum_{d=1}^{d^{[j]}}{\bf H}^{[kj]}{\bf V}^{[j]}_{\star d}{\bf V}^{[j]\dagger}_{\star d}{\bf H}^{[kj]\dagger}\nonumber\\ && ~~-\frac{P^{[k]}}{d^{[k]}}{\bf H}^{[kk]}{\bf V}^{[k]}_{\star l}{\bf V}^{[k]\dagger}_{\star l}{\bf H}^{[kk]\dagger}+{\bf I}_{N^{[k]}}\label{eq:Bkl}\end{aligned}$$ The unit vector ${\bf U}^{[k]}_{\star l}$ that maximizes $\mbox{SINR}_{kl}$ is given by $${\bf U}^{[k]}_{\star l}= \frac{\left({\bf B}^{[kl]}\right)^{-1}{\bf H}^{[kk]}{\bf V}^{[k]}_{\star l}}{\parallel \left({\bf B}^{[kl]}\right)^{-1}{\bf H}^{[kk]}{\bf V}^{[k]}_{\star l}\parallel}.\label{eq:Ukl}$$ The steps of the iteration are given in Algorithm \[alg:max\_sinr\]. Start with any ${\bf V}^{[k]}:M^{[k]}\times d^{[k]}$, columns of ${\bf V}^{[k]}$ are linearly independent unit vectors. Begin iteration Compute interference plus noise covariance matrix for ${\bf B}^{[kl]}$ for stream $l$ at receiver $k$ according to (\[eq:Bkl\]), $\forall ~k\in\{1,2,\cdots, K\}, l\in\{1,2,\cdots,d^{[k]}\}$. Calculate receive combining vectors ${\bf U}^{[k]}_{\star l}$ at receiver $k$ according to (\[eq:Ukl\]), $\forall ~k\in\{1,2,\cdots, K\}, l\in\{1,2,\cdots,d^{[k]}\}$. Reverse the communication direction and use the receive combining vectors as precoding vectors: $\overleftarrow{\bf V}^{[k]}={\bf U}^{[k]}$, $\forall ~k\in\{1,2,\cdots, K\}$. In the reciprocal network, compute interference plus-noise covariance matrix $\overleftarrow{\bf B}^{[kl]}$ for stream $l$ at receiver $k$, $\forall ~k\in\{1,2,\cdots, K\}, l\in\{1,2,\cdots,d^{[k]}\}$. Calculate receive combining vectors $\overleftarrow{\bf U}^{[k]}_{\star l}$, $\forall ~k\in\{1,2,\cdots, K\}, l\in\{1,2,\cdots,d^{[k]}\}$. Reverse the communication direction and use the receive combining vectors as precoding vectors: ${\bf V}^{[k]}=\overleftarrow{\bf U}^{[k]}$, $\forall ~k\in\{1,2,\cdots, K\}$. Repeat until convergence. Performance Results and Applications {#sec: results} ==================================== Consider the $3$ user interference channel where each node is equipped with $2$ antennas and all channel coefficients are i.i.d. zero mean unit variance circularly symmetric complex Gaussian. As shown in Fig. \[fig: nash\] with interference alignment each user achieves $1$ degree of freedom. In Fig. \[algo\_3u\_2a\] we compare the performance of the iterative interference alignment algorithm for this channel with the theoretical interference alignment which assumes global channel knowledge. It can be seen that the algorithm performs very close to the theoretical case and more importantly it provides significant benefits over the orthogonal case. The sum rate for the orthogonal scheme is calculated assuming equal time sharing for the users, and with power $3P$ per node. The modified interference alignment scheme where the receive-combining vectors maximize the SINR provides a considerable performance gain at low and intermediate $P$. As expected, it achieves the same performance of interference alignment at high $P$. We also plot the performance of interference avoidance algorithm which is a selfish approach. The isotropic transmission case refers to the case where each transmitter sends $2$ streams of equal power without regard to the channel information. In the following, we explore some of the applications of the iterative interference alignment algorithm. Feasibility of Interference Alignment ------------------------------------- While the iterative algorithm is useful for circumventing the need for global channel knowledge, it can also be used to check theoretical feasibility of interference alignment for a given number of streams per user. Let $(d^{[1]},d^{[2]}, \cdots d^{[k]})$ denote the number of transmit streams of the users. For perfect interference alignment $\sum_{j=1}^{d^{[k]}}\lambda_j[{\bf Q}^{[k]}]=0$ at receiver $k$ where $\lambda_j[{\bf A}]$ denotes the $j$th smallest eigenvalue of $\bf A$. Note that $\sum_{j=1}^{d^{[k]}}\lambda_j[{\bf Q}^{[k]}]$ indicates the interference power in the desired signal space. Using the algorithm, we plot in Fig. \[int\_percent\], the percentage of interference in the desired signal space versus the total number of transmit streams in the network. The fraction of interference in the desired signal space of receiver $k$ is defined as $$p_{k}= \frac{\sum_{j=1}^{d^{[k]}}\lambda_j[{\bf Q}^{[k]}]}{\mbox{Tr}[{\bf Q}^{[k]}]}.$$ When interference alignment is feasible the fraction of interference in desired signal space will be zero (within numerical errors). Fig. \[int\_percent\] suggests that interference alignment is feasible on the four user interference channel with 5 antennas at each node when each transmitter sends two streams. With increase in the number of streams the interference in desired signal space increases which is an indication that interference alignment is not possible. Although the upperbound on the degrees of freedom for this network is 10, approaching the upperbound may entail channel extensions. The plot suggests that 8 degrees of freedom ($d^{[1]}=d^{[2]}=d^{[3]}=d^{[4]}=2$) can be achieved without channel extension. Similarly for the 4 antenna case, the plot indicates that interference alignment is possible for up to a total of 6 streams in the $4$ user interference network with only $4$ antennas at each node. Networks with single antenna nodes ---------------------------------- MIMO nodes are not necessary in order to achieve interference alignment in wireless networks. Interference can be aligned even in networks with single antenna nodes through channel extension in frequency or time as long as the channel is varying across frequency or time [@Cadambe_Jafar_int]. However, one caveat with this approach is the need for long symbol extensions. Aligning interference with a large number of beams over a large number of dimensions can be especially challenging for iterative algorithms due to large dimensionality of the optimization space and the inherently non-convex nature of the problem. Moreover, symbol extensions over orthogonal dimensions produce structured (diagonal or block diagonal) matrices for which both conditions (\[eq:direct\]) and (\[eq:cross\]) are non-trivial. Due to these difficulties, it is preferred if interference alignment can be accomplished without symbol extensions or with limited symbol extensions. In this section, we give an example of how long symbol extensions may be avoided by the use of relays. Note that [@Cadambe_Jafar_XFB] has shown that relays cannot increase the degrees of freedom for time-varying wireless networks. However, as we show in this section, relays can be very useful by reducing the size of the signalling space over which interference alignment can be accomplished. The key idea is to employ relays to create a *virtual* MIMO system. Consider an interference relay channel with three sources, three destinations and a half-duplex relay (node $0$) as shown in Fig. \[fig: relay\_int\]. Recall that in the absence of relays this network is shown to approach the upperbound of $3/2$ degrees of freedom per orthogonal dimension in the asymptotic limit of infinitely long symbol extensions [@Cadambe_Jafar_int]. However, we show that with relays only a two time slots are required to achieve the outerbound, i.e. $3/2$ degrees of freedom. Consider the following two slot protocol. In the first slot, the relay is silent and the received signal at destination $j$ is given by $$y^{[j]}(1)=\sum_{i=1}^{3}h^{[ji]}(1)x^{[i]}(1)+z^{[j]}(1), ~~~j=1,2,3$$ The received signal at the relay can be expressed as $$y^{[0]}(1)=\sum_{i=1}^{3}{h}^{[0i]}(1)x^{[i]}(1)+z^{[0]}(1)$$ In the second slot, the relay transmits a scaled version of its received symbol while source $i$ transmits $x^{[i]}(2)$. The received signal at the $j^{th}$ destination node in the second slot ($j=1,2,3$) $$y^{[j]}(2)=\sum_{i=1}^{3}{h}^{[ji]}(2)X^{[i]}(2)+{h}^{[j0]}(2)\beta y^{[0]}(1)+z^{[j]}(2)$$ Let ${Y}^{[j]}=[y^{[j]}(1)~ y^{[j]}(2)]^{T}$ and ${X}^{[i]}=[x^{[i]}(1)~ x^{[i]}(2)]^{T}$. In vector form, the received signal at destination $j$ can be expressed as $${Y}^{[j]}=\sum_{i=1}^{3}{\bf H}^{[ji]}{X}^{[i]}+ {Z}^{[j]}$$ where $$\small{\bf H}^{[ji]}=\left[\begin{array}{cc} h^{[ji]}(1) &0\\ \beta h^{[j0]}(2)h^{[0i]}(1) & h^{[ji]}(2) \end{array}\right]$$ and $${Z}^{[j]}=\left[\begin{array}{c} z^{[j]}(1) \\ \beta h^{j0}(2)z^{[0]}(1) +z^{[j]}(2)\end{array}\right]$$ Thus, over two time slots, the relay network reduces to a three user MIMO interference channel with (non-diagonal) structure on the channel matrix. Since the channel matrix is non-diagonal (unlike symbol extensions in the absence of a relay) it is easy to verify that when the channels are random and independent of each other, a multiplexing gain of $\frac{3}{2}$ is achieved with probability $1$. The advantage of this scheme is that it requires only two time-slots to achieve $3/2$ degrees of freedom per time-slot, whereas in the absence of the relay, infinitely many time-slots are used in [@Cadambe_Jafar_int]. Since fewer dimensions are needed, iterative algorithms work better (faster convergence) in this setting. In Fig. \[int\_relay\] we plot the performance of the interference alignment schemes for the case of time extensions. For the interference alignment scheme with relay, the transmit power per node is $P$. That is, the three transmitters and the relay have a total power of $4P$. The transmit power for the orthogonal scheme where only transmitter is active at a time is $4P$. The interference alignment scheme without relay achieves a multiplexing gain of 4 in 3 time slots [@Cadambe_Jafar_int]. Here transmitter 1 sends two streams while transmitters 2 and 3 send one stream each. The transmit power per stream is $P$ adding up to $4P$ for the scheme. We can also increase the number of time slots for channel extension to improve the multiplexing gain. However it requires very high $P$ to outperform the orthogonal scheme. It can be seen that adding an extra relay helps in achieving the multiplexing gain of $\frac{3}{2}$ in two slots. Further the performance improves at low $P$ as well. It must be stressed that the main idea behind the interference alignment scheme with the relay is to show that there are benefits in employing relays when the nodes do not have multiple antennas, especially for iterative algorithms. The scheme can be further improved by optimally allocating power to the relay and employing Algorithm \[alg:max\_sinr\] to improve its performance at low $P$. Conclusion {#sec: concl} ========== Distributed interference alignment algorithms are investigated. Interference alignment is found to be achievable through iterative algorithms based on network reciprocity and the “minimize interference to others” approach. Numerical comparisons to orthogonal schemes, simultaneous transmission schemes and selfish interference avoidance schemes show that the benefits of distributed interference alignment algorithm are significant and close to the theoretical predictions. As mitigating interference is the fundamental problem of wireless networks, the ’do no harm’ approach based algorithms have enormous applications in wireless networks.
{ "pile_set_name": "ArXiv" }
--- abstract: 'In de Sitter space, scale invariant fluctuations give rise to infrared logarithmic corrections to physical quantities, which eventually spoil perturbation theories. For models without derivative interactions, it has been known that the field equation reduces to a Langevin equation with white noise in the leading logarithm approximation. The stochastic equation allows us to evaluate the infrared effects nonperturbatively. We extend the resummation formula so that it is applicable to models with derivative interactions. We first consider the nonlinear sigma model and next consider a more general model which consists of a noncanonical kinetic term and a potential term. The stochastic equations derived from the infrared resummation in these models can be understood as generalizations of the standard one to curved target spaces.' --- NCTS-TH/1814 [**Infrared resummation for derivative interactions\ in de Sitter space**]{} Hiroyuki <span style="font-variant:small-caps;">Kitamoto</span>[^1], *Physics Division, National Center for Theoretical Sciences*\ *National Tsing-Hua University, Hsinchu 30013, Taiwan*\ July 2019 Introduction ============ In de Sitter (dS) space, the propagator for a massless and minimally coupled scalar field has a dS symmetry breaking term. This term is a direct consequence of the scale invariant spectrum at the superhorizon scale and is expressed as a logarithm of the scale factor of the Universe [@Vilenkin1982; @Linde1982; @Starobinsky1982]. In the presence of this scalar field, physical quantities may acquire time dependences through the propagator. In tribute to their origin, we call them quantum infrared (IR) effects in dS space. By employing the Schwinger-Keldysh formalism [@Schwinger1961; @Keldysh1964], we can study interacting field theories in dS space perturbatively. The IR effects at each loop level manifest as polynomials in the IR logarithm whose degrees increase with the loop level. At a late time, the leading IR effects come from the leading IR logarithms at each loop level. This fact indicates that the perturbative study breaks down after a large enough cosmic expansion. In order to understand such a situation, we need to evaluate the IR effects nonperturbatively. For models without derivative interactions, Starobinsky and Yokoyama proposed that the IR dynamics can be described nonperturbatively by a Langevin equation with white noise [@Starobinsky1986; @Starobinsky1994]. Tsamis and Woodard proved that the stochastic approach is equivalent to the resummation of the leading IR logarithms to all-loop orders [@Woodard2005]. In this paper, we extend the resummation formula of the leading IR logarithms so that it is applicable to models with derivative interactions. As concrete examples, we consider the nonlinear sigma model and a more general model which consists of a noncanonical kinetic term and a potential term. Although there has been an attempt to derive stochastic equations in these models [@Tada2018], the consistency with the resummation of the leading IR logarithms has not been verified completely. We show that even when the kinetic term is noncanonical, the Yang-Feldman equation reduces to a Langevin equation with white noise in the leading logarithm approximation. The resulting stochastic equation is expressed in a covariant form with respect to the field coordinate transformation. As a specific feature of derivative interactions, we need to take into account the contribution from the subhorizon scale in evaluating the leading IR effects. This fact was found in [@Kitamoto2011; @Kitamoto2012] where the energy-momentum tensor of the nonlinear sigma model was studied. This study takes a first step toward the understanding of the nonperturbative IR effects from gravity. The gravitational field theory includes massless and minimally coupled modes [@Woodard1994], and it consists of derivative interactions in a similar way to the general scalar field theory. There are perturbative studies of gravitational IR effects; e.g. the gravitational IR corrections on the slow-roll parameters $\epsilon$ and $\eta$ were studied in single-field inflation models [@Kitamoto2017]. At the one-loop level, $\epsilon$ acquires a secular growth while $\eta$ does not. This result indicates that if $\epsilon$ and $\eta$ are vanishingly small at the beginning due to the pseudo shift symmetry, the quantum mechanism leads to an inflation model with a linear potential. In order to evaluate the whole time evolution of physical quantities, the IR resummation formula for gravity is necessary but it has not been known. The organization of this paper is as follows. In Sec. 2, we review the propagator for a massless and minimally coupled scalar field in dS space, in particular, its IR behavior. In Sec. 3, we derive a Langevin equation for the nonlinear sigma model by applying the leading logarithm approximation to its Yang-Feldman equation. In Sec. 4, we apply the same approach to a general scalar field theory which consists of a noncanonical kinetic term and a potential term. We also mention a relation to the Euclidean field theory on a sphere. We conclude with discussions in Sec. 5. Free scalar field theory ======================== Here we review a free scalar field which is massless and minimally coupled to the dS background. In particular, we focus on its IR dynamics which is a source of time dependent quantum effects. In the Poincaré coordinate, the metric of dS space is given by $$\begin{aligned} ds^2=-dt^2+a^2(t)d\textbf{x}^2,\hspace{1em}a(t)=e^{Ht}, \label{metric}\end{aligned}$$ where the dimension of the spacetime is taken as an arbitrary $D$, and $H$ is the Hubble parameter which is constant in dS space. In the conformally flat coordinate, $$\begin{aligned} g_{\mu\nu}=a^2(\tau)\eta_{\mu\nu},\hspace{1em}a(\tau)=\frac{1}{-H\tau}. \label{metric'}\end{aligned}$$ The conformal time $\tau$ is related to the cosmic time $t$ as $\tau=-\frac{1}{H}e^{-Ht}$. In this paper, $D$-dimensional vectors and tensors are expressed in the conformally flat coordinates. The quadratic action for a massless and minimally coupled scalar field is given by $$\begin{aligned} S=\int \sqrt{-g}d^Dx\ \left[-\frac{1}{2}g^{\mu\nu}\partial_\mu\varphi\partial_\nu\varphi\right]. \end{aligned}$$ The free scalar field $\varphi_0$ can be expanded by the annihilation and the creation operators $a_\textbf{p},\ a_\textbf{p}^\dagger$ as $$\begin{aligned} \varphi_0(x)=\int\frac{d^{D-1}p}{(2\pi)^{D-1}}\ \left[\phi_\textbf{p}(x)a_\textbf{p}+\phi_\textbf{p}^*(x)a_\textbf{p}^\dagger\right], \end{aligned}$$ $$\begin{aligned} \phi_\textbf{p}(x)=\frac{\sqrt{\pi}}{2}H^\frac{D-2}{2}(-\tau)^\frac{D-1}{2}H_\frac{D-1}{2}^{(1)}(-p\tau) e^{i\textbf{p}\cdot\textbf{x}}, \label{WF}\end{aligned}$$ where $p=|\textbf{p}|$ and $H_\frac{D-1}{2}^{(1)}$ is the Hankel function of the first kind. See, e.g., [@ToI] for details of special functions. Throughout this paper, we consider the Bunch-Davies vacuum $a_\textbf{p}|0\rangle=0$. It should be noted that $p$ denotes the comoving momentum, and the physical momentum is given by $P=p/a(\tau)$. In this paper, we mainly consider the contribution from the superhorizon scale where curved spacetime-specific effects become manifest. At the superhorizon scale $P\ll H \Leftrightarrow -p\tau\ll 1$, the wave function behaves as $$\begin{aligned} \phi_\textbf{p}(x)\simeq -i\frac{2^\frac{D-3}{2}\Gamma(\frac{D-1}{2})}{\sqrt{\pi}} \frac{H^\frac{D-2}{2}}{p^\frac{D-1}{2}}e^{i\textbf{p}\cdot\textbf{x}}. \label{WF-IR}\end{aligned}$$ The quantum fluctuation of a massless and minimally coupled scalar field is frozen at the superhorizon scale as the leading IR term of the wave function has no time dependence. The contribution from (\[WF-IR\]) to the propagator at the coincident point is given by $$\begin{aligned} \langle\varphi_0^2(x)\rangle\simeq \int_{p<Ha(\tau)}\frac{d^{D-1}p}{(2\pi)^{D-1}}\ \frac{2^{D-3}\Gamma^2(\frac{D-1}{2})}{\pi}\frac{H^{D-2}}{p^{D-1}}. \label{P-IR'}\end{aligned}$$ The upper bound of the momentum integral is taken on the horizon because the $1/p^{D-1}$ spectrum is dominant at the superhorizon scale. The propagator has an IR logarithmic divergence due to the presence of the scale invariant spectrum. In order to regularize the IR divergence, we introduce an IR cutoff $p_0$ which fixes the minimum value of the comoving momentum. Under the current setting, the IR contribution to the propagator is expressed as the logarithm of the scale factor [@Vilenkin1982; @Linde1982; @Starobinsky1982]: $$\begin{aligned} \langle\varphi_0^2(x)\rangle &\simeq \frac{2H^{D-2}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \int^{Ha(\tau)}_{p_0}\frac{dp}{p} \notag\\ &=\frac{2H^{D-2}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \int^H_{p_0/a(\tau)}\frac{dP}{P} \notag\\ &=\frac{2H^{D-2}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}\log (a(\tau)/a_0). \label{P-IR}\end{aligned}$$ From (\[P-IR’\]) to (\[P-IR\]), we used $$\begin{aligned} \Gamma(2z)=\frac{2^{2z-1}}{\sqrt{\pi}}\Gamma(z)\Gamma(z+\frac{1}{2}). \label{formula}\end{aligned}$$ This formula of the gamma function is used repeatedly in this paper. The scale factor at the initial time $a_0$ is related to the IR cutoff as $a_0=p_0/H$. Physically, $a(\tau)/p_0$ means the size of the Universe, and it expands with the scale factor. Since the Hubble scale $1/H$ is constant, the degrees of freedom at the superhorizon scale increase with time. This increase gives rise to the logarithmic dependence of the scale factor to the propagator through the scale invariant spectrum. Here we write down the explicit form of the propagator for a massless and minimally coupled scalar field [@Woodard2010]: $$\begin{aligned} \langle\varphi_0(x)\varphi_0(x')\rangle = I(y)+\frac{H^{D-2}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}\log (a(\tau)a(\tau')/a_0^2), \label{propagator}\end{aligned}$$ $$\begin{aligned} I(y)&=\frac{H^{D-2}}{(4\pi)^\frac{D}{2}}\left\{\Gamma(\frac{D}{2}-1)\left(\frac{y}{4}\right)^{1-\frac{D}{2}} -\frac{\Gamma(\frac{D}{2}+1)}{2-\frac{D}{2}}\left(\frac{y}{4}\right)^{2-\frac{D}{2}} +\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}\delta\right. \\ &\hspace{5em}\left.+\sum^{\infty}_{n=1} \left[\frac{\Gamma(D-1+n)}{n\Gamma(\frac{D}{2}+n)}\left(\frac{y}{4}\right)^n -\frac{\Gamma(\frac{D}{2}+1+n)}{(2-\frac{D}{2}+n)(n+1)!}\left(\frac{y}{4}\right)^{2-\frac{D}{2}+n} \right]\right\}, \notag\end{aligned}$$ $$\begin{aligned} \delta\equiv-\psi(1-\frac{D}{2})+\psi(\frac{D-1}{2})+\psi(D-1)+\psi(1),\hspace{1em} \psi(z)\equiv \frac{1}{\Gamma(z)}\frac{d\Gamma(z)}{dz}, \end{aligned}$$ where $y$ denotes the square of the physical distance: $$\begin{aligned} y=H^2a(\tau)a(\tau')\Delta x^2,\hspace{1em}\Delta x^2\equiv -(\tau-\tau')^2+(\textbf{x}-\textbf{x}')^2. \end{aligned}$$ The metric of dS space and $y$ are invariant under the rescaling $\tau\to C\tau,\ \textbf{x}\to C\textbf{x}$. The propagator for a massless and minimally coupled scalar field (\[propagator\]) does not respect this dS symmetry. This is due to the IR cutoff dependence of the second term. We evaluate the twice-differentiated propagator at the coincident point $\langle\partial_\mu\varphi_0(x)\partial_\nu\varphi_0(x)\rangle$ as we use it in the subsequent discussion. In the dimensional regularization, the $y^\alpha$ term disappears at the coincident point if $\alpha$ becomes positive for a sufficiently low $D$. Considering this fact, the twice-differentiated propagator at the coincident point is evaluated as $$\begin{aligned} \langle\partial_\mu\varphi_0(x)\partial_\nu\varphi_0(x)\rangle &=\frac{H^{D-2}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D)}{4\Gamma(\frac{D}{2}+1)} \partial_\mu\partial'_\nu y^1|_{x'\to x} \notag\\ &=-\frac{H^D}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D)}{D\Gamma(\frac{D}{2})}g_{\mu\nu}. \label{tdp}\end{aligned}$$ The twice-differentiated propagator respects the dS symmetry. That is because if differential operators act on both $x$ and $x'$, the dS symmetry breaking term in (\[propagator\]) does not have any contribution, $$\begin{aligned} \partial_\mu\partial'_\nu \log (a(\tau)a(\tau')/a_0^2)=0. \label{disappear}\end{aligned}$$ In interacting field theories with massless and minimally coupled scalar fields, physical quantities may acquire IR logarithmic corrections through the propagator. At the late time $\log(a(\tau)/a_0)\gg 1$, the leading IR effects come from the leading IR logarithms at each loop level. As an example, let us consider the $\lambda \varphi^i\varphi^i\varphi^j\varphi^j$ interaction. With each increase in the loop level, quantum corrections from the nonderivative interaction are multiplied by at most a factor of $$\begin{aligned} \lambda H^{D-4} \log^2(a(\tau)/a_0). \end{aligned}$$ As another example, let us consider the $f^2\varphi^i\varphi^ig^{\mu\nu}\partial_\mu\varphi^j\partial_\nu\varphi^j$ interaction. With each increase in the loop level, quantum corrections from the derivative interaction are multiplied by at most a factor of $$\begin{aligned} f^2H^{D-2} \log (a(\tau)/a_0). \end{aligned}$$ These secular growths indicate that even though the dimensionless couplings are small as $\lambda H^{D-4},\ f^2H^{D-2}\ll 1$, the perturbative study breaks down after a long enough time: $$\begin{aligned} \lambda H^{D-4} \log^2(a(\tau)/a_0) \sim 1\hspace{1em}\text{or}\hspace{1em} f^2H^{D-2} \log (a(\tau)/a_0) \sim 1. \end{aligned}$$ A rational approach for evaluating such nonperturbative effects is to resum the leading IR logarithms to all-loop orders. The resummation formula for nonderivative interactions has been derived by Tsamis and Woodard [@Woodard2005]. In subsequent sections, we derive the resummation formula for derivative interactions. Nonlinear sigma model ===================== As a concrete model with derivative interactions, we consider the nonlinear sigma model: $$\begin{aligned} S=\int \sqrt{-g}d^Dx\ \left[-\frac{1}{2}G_{ij}(\varphi)g^{\mu\nu}\partial_\mu\varphi^i\partial_\nu\varphi^j\right], \label{NLsigma}\end{aligned}$$ where $G_{ij}\ (i=1,\cdots,N)$ is the metric of the target space. The action is invariant under the general coordinate transformation where $\varphi^i$ are identified as coordinates. The global symmetry guarantees that this model consists of massless and minimally coupled scalar fields without fine-tuning the quadratic action. For later use, we introduce the vielbein: $$\begin{aligned} G_{ij}(\varphi)=E_i^{\ a}(\varphi)E_j^{\ b}(\varphi)\delta_{ab}, \end{aligned}$$ where $a$ denotes the index of the flat tangent space. The field equation for the nonlinear sigma model is given by $$\begin{aligned} \frac{1}{\sqrt{-g}}\partial_\mu\left\{\sqrt{-g}\ G_{ij}(\varphi)g^{\mu\nu}\partial_\nu\varphi^j\right\} -\frac{1}{2}\frac{\partial}{\partial\varphi^i}G_{jk}(\varphi)g^{\mu\nu}\partial_\mu\varphi^j\partial_\nu\varphi^k=0. \label{FE0}\end{aligned}$$ As is well known, the equation can be rewritten in the following form: $$\begin{aligned} \nabla^\mu\nabla_\mu\varphi^i +\Gamma^i_{\ jk}(\varphi)g^{\mu\nu}\partial_\mu\varphi^j\partial_\nu\varphi^k=0, \label{FE0'}\end{aligned}$$ where $\nabla_\mu$ is the covariant derivative in the spacetime and $\Gamma^i_{\ jk}$ is the Levi-Civita connection in the target space. However, in order to find a covariant Yang-Feldman equation, it is more useful to rewrite (\[FE0\]) as $$\begin{aligned} \frac{1}{\sqrt{-g}}\partial_\mu\big\{\sqrt{-g}\ E_i^{\ a}(\varphi)g^{\mu\nu}\partial_\nu\varphi^i\big\}-D_iE_j^{\ a}(\varphi)g^{\mu\nu}\partial_\mu\varphi^i\partial_\nu\varphi^j=0, \label{FE1}\end{aligned}$$ where $D_i$ is the covariant derivative in the target space. The covariance is manifest in (\[FE1\]) in contrast to (\[FE0\]) and (\[FE0’\]). The Yang-Feldman equation for the nonlinear sigma model is given by $$\begin{aligned} -i\int d^D&x'\ \partial'_\mu\left\{\sqrt{-g(x')}g^{\mu\nu}(x')\partial'_\nu G^R(x,x') E_i^{\ a}(\varphi(x'))\right\} \varphi^i(x') \notag\\ &=\varphi_0^a(x) -i\int\sqrt{-g(x')}d^Dx'\ G^R(x,x')D_iE_j^{\ a}(\varphi(x')) g^{\mu\nu}(x')\partial'_\mu\varphi^i(x')\partial'_\nu\varphi^j(x'), \label{YF1}\end{aligned}$$ where $G^R$ denotes the retarded propagator: $$\begin{aligned} G^R(x,x')=\theta(\tau-\tau')\langle[\varphi_0(x),\varphi_0(x')]\rangle, \label{retarded}\end{aligned}$$ which satisfies $$\begin{aligned} \nabla^\mu\nabla_\mu G^R(x,x')=\frac{i\delta^{(D)}(x-x')}{\sqrt{-g(x)}}. \label{kernel}\end{aligned}$$ We can reproduce (\[FE1\]) by applying the d’Alembert operator to both sides of (\[YF1\]). It should be noted that the homogeneous solution, i.e. the free field, has the index $a$ rather than $i$, and both sides of (\[YF1\]) have the index $a$. The Yang-Feldman equation describes quantum effects from all momentum scales, and it is not exactly solvable in general. We show that in the leading logarithm approximation, the Yang-Feldman equation reduces to a Langevin equation with white noise, which is a more suitable tool for studying the IR dynamics. For the first term in the right-hand side of (\[YF1\]), we extract its leading IR behavior as $$\begin{aligned} \varphi_0^a(x)\simeq\bar{\varphi}_0^a(x) \equiv\int\frac{d^{D-1}p}{(2\pi)^{D-1}}\ \theta(Ha(\tau)-p) \left[-i\frac{2^\frac{D-3}{2}\Gamma(\frac{D-1}{2})}{\sqrt{\pi}}\frac{H^\frac{D-2}{2}}{p^\frac{D-1}{2}} e^{i\textbf{p}\cdot\textbf{x}}a_\textbf{p}^a+\text{(H.c.)}\right], \label{right1}\end{aligned}$$ where $[a_\textbf{p}^a,a_{\textbf{p}'}^{\dagger b}] =(2\pi)^{D-1}\delta^{(D-1)}(\textbf{p}-\textbf{p}')\delta^{ab}$ and the other commutation relations are zero. Since the $1/p^\frac{D-1}{2}$ term is dominant at the superhorizon scale, we introduce the step function which is nonzero only at this IR scale. Substituting (\[right1\]), each Wightman function without differential operators $\langle\varphi_0(x)\varphi_0(x')\rangle$ provides a single IR logarithm. The second term in the right-hand side of (\[YF1\]) includes two differentiated fields $\partial'_\mu\varphi^i(x')\partial'_\nu\varphi^j(x')$. At each order of the expansion in the interaction vertex $$\begin{aligned} \frac{1}{2}\delta G_{ij}(\varphi)g^{\mu\nu}\partial_\mu\varphi^i\partial_\nu\varphi^j,\hspace{1em} \delta G_{ij}(\varphi)=G_{ij}(\varphi)-\delta_{ij}, \label{vertex}\end{aligned}$$ the diagram with the leading IR logarithms includes a loop of twice-differentiated propagators, which starts at $x'$ and ends at $x'$: $$\begin{aligned} \parbox{\boxal}{\usebox{\boxa}}\ ,\hspace{1em} \parbox{\boxbl}{\usebox{\boxb}}\ ,\hspace{1em} \parbox{\boxcl}{\usebox{\boxc}}\ ,\hspace{1em}\cdots\hspace{1em}. \label{diagrams}\end{aligned}$$ Here the horizontal line segment denotes $G^R(x,x')$, and $\partial$ denotes a differential operator. The other diagrams, which are obtained by transferring any of the differential operators from the propagators inside the loop to the nondifferentiated fields outside the loop, have reduced powers of the IR logarithms. In other words, we need to minimize the number of differentiated propagators to obtain the leading IR logarithms.[^2] Using partial integrations, we can rewrite the diagrams as $$\begin{aligned} \parbox{\boxbl}{\usebox{\boxb}}\ \simeq\parbox{\boxbbl}{\usebox{\boxbb}}\ \simeq\parbox{\boxbbbl}{\usebox{\boxbbb}}\ , \end{aligned}$$ $$\begin{aligned} \parbox{\boxcl}{\usebox{\boxc}}\ \simeq\parbox{\boxccl}{\usebox{\boxcc}}\ \simeq\parbox{\boxcccl}{\usebox{\boxccc}}\ , \end{aligned}$$ where we used the fact that the diagrams with differentiated $\delta G_{ij}$ are negligible in the leading logarithm approximation. The presence of the propagator with the d’Alembert operator $\nabla^2$ allows us to evaluate the vertex integral trivially.[^3] Applying the same procedure also at higher orders of the expansion in the interaction vertex, the total contribution of the diagrams (\[diagrams\]) can be summarized as $$\begin{aligned} \parbox{\boxdl}{\usebox{\boxd}}\ , \label{diagram}\end{aligned}$$ where we used $$\begin{aligned} \delta_{ij}-\delta G_{ij}+\delta G_{ik}\delta G_{kj}+\cdots=G^{ij}. \end{aligned}$$ In the second term in the right-hand side of (\[YF1\]), the diagrammatic discussion can be expressed as $$\begin{aligned} \partial'_\mu\varphi^i(x')\partial'_\nu\varphi^j(x') &\simeq G^{ij}(\varphi(x'))\langle\partial'_\mu\varphi_0(x')\partial'_\nu\varphi_0(x')\rangle \notag\\ &=-G^{ij}(\varphi(x'))\frac{H^D}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D)}{D\Gamma(\frac{D}{2})}g_{\mu\nu}(x'). \label{right2}\end{aligned}$$ We substituted the value of the twice-differentiated propagator (\[tdp\]) in the last line. It should be noted that (\[right2\]) is the approximation for the diagrams with a loop of twice-differentiated propagators (\[diagrams\]). This approximation means that we can integrate out differentiated fields as if nondifferentiated fields were constant. In other words, we cannot treat $G_{ij}$ as a constant coefficient of the kinetic term in evaluating correlation functions of nondifferentiated fields. The second term in the right-hand side of (\[YF1\]) also includes the retarded propagator. In order to evaluate it, we need to know the real and the imaginary parts of the wave function. Therefore, the wave function (\[WF\]) should be expanded further than in (\[WF-IR\]): $$\begin{aligned} \phi_\textbf{p}(x)\simeq -i\left\{\frac{2^\frac{D-3}{2}\Gamma(\frac{D-1}{2})}{\sqrt{\pi}}\frac{H^\frac{D-2}{2}}{p^\frac{D-1}{2}} +i\frac{\sqrt{\pi}}{2^\frac{D+1}{2}\Gamma(\frac{D+1}{2})}H^\frac{D-2}{2}p^\frac{D-1}{2}(-\tau)^{D-1}\right\} e^{i\textbf{p}\cdot\textbf{x}}. \label{WF-IR'}\end{aligned}$$ Substituting it in (\[retarded\]), the retarded propagator is evaluated as $$\begin{aligned} G^R(x,x')&\simeq \theta(\tau-\tau')\int\frac{d^{D-1}p}{(2\pi)^{D-1}}\ \frac{-iH^{D-2}}{D-1}\left[(-\tau')^{D-1}-(-\tau)^{D-1}\right]e^{i\textbf{p}\cdot(\textbf{x}-\textbf{x}')} \notag\\ &=\frac{-i}{(D-1)H}\theta(\tau-\tau')\left[a^{-(D-1)}(\tau')-a^{-(D-1)}(\tau)\right] \delta^{(D-1)}(\textbf{x}-\textbf{x}'). \label{right3'}\end{aligned}$$ The retarded propagator itself does not have an IR logarithm, in contrast to the Wightman function, while the vertex integral with it induces a single IR logarithm as $$\begin{aligned} \int\sqrt{-g(x')}d^Dx'\ G^R(x,x') &\simeq -\frac{i}{(D-1)H}\int^\tau a(\tau')d\tau'\ \left[1-\left(a(\tau')/a(\tau)\right)^{D-1}\right] \notag\\ &\simeq-\frac{i}{(D-1)H}\int^tdt', \label{right3}\end{aligned}$$ where we dropped the trivial spatial integral $\int d^{D-1}x'\ \delta^{(D-1)}(\textbf{x}-\textbf{x}')$. In the last line, we neglected the second term because it does not induce an IR logarithm. As $dt'=d(\log a(t'))/H$, the vertex integral provides a single IR logarithm in addition to the contribution from the integrand $D_iE_j^{\ a}(\varphi(x'))g^{\mu\nu}(x')\partial'_\mu\varphi^i(x')\partial'_\nu\varphi^j(x')$. It should be noted that the higher-order terms neglected in (\[right3’\]) reduce the powers of the IR logarithms from the integrand. The left-hand side of (\[YF1\]) can be rewritten as $$\begin{aligned} E_i^{\ a}(\varphi(x))\varphi^i(x) -i\int \sqrt{-g(x')}d^Dx'\ g^{\mu\nu}(x')\partial'_\mu G^R(x,x') \partial'_\nu E_i^{\ a}(\varphi(x'))\varphi^i(x'). \label{YF1'}\end{aligned}$$ We can extract the leading IR logarithms from the first term just by iteratively substituting $\varphi$ in it. In order to evaluate the leading IR effects from the second term, we approximate the differentiated retarded propagator as $$\begin{aligned} \partial'_\mu G^R(x,x') &\simeq \theta(\tau-\tau')\delta_\mu^{\ 0}\int\frac{d^{D-1}p}{(2\pi)^{D-1}}\ iH^{D-2}(-\tau')^{D-2}e^{i\textbf{p}\cdot(\textbf{x}-\textbf{x}')} \notag\\ &= i\theta(\tau-\tau')\delta_\mu^{\ 0} a^{-(D-2)}(\tau')\delta^{(D-1)}(\textbf{x}-\textbf{x}'). \label{left'}\end{aligned}$$ As we did in (\[WF-IR’\]) and (\[right3’\]), we expanded the differentiated retarded propagator in powers of $(-p\tau)$ and $(-p\tau')$ and kept the zeroth-order term.[^4] Substituting (\[left’\]), the vertex integral with the differentiated retarded propagator is given by $$\begin{aligned} \int \sqrt{-g(x')}d^Dx'\ g^{\mu\nu}(x')\partial'_\mu G^R(x,x') &\simeq -i\int^\tau d\tau'\ \delta_0^{\ \nu}\notag\\ &=-i\int^tdt'\ a^{-1}(t')\delta_0^{\ \nu}, \label{left}\end{aligned}$$ where we dropped the trivial spatial integral. For the integrand $\partial'_0 E_i^{\ a}(\varphi(x'))\varphi^i(x')$, the differential operator $\partial'_0=a(t')\partial'_t$ removes a single IR logarithm from $E_i^{\ a}(\varphi(x'))$ and adds one scale factor $a(t')$. The scale factor from the differential operator is canceled by $a^{-1}(t')$ in (\[left\]), and then the vertex integral provides a single IR logarithm as $dt'=d(\log a(t'))/H$. That is how the second term in (\[YF1’\]) induces the leading IR logarithms. Applying the leading logarithm approximation (\[right1\]), (\[right2\]), (\[right3\]) and (\[left\]), the Yang-Feldman equation (\[YF1\]) reduces to $$\begin{aligned} E_i^{\ a}(\varphi(x))\varphi^i&(x) -\int^tdt'\ \partial'_tE_i^{\ a}(\varphi(t',\textbf{x}))\varphi^i(t',\textbf{x}) \notag\\ &=\bar{\varphi}_0^a(x) +\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \int^tdt'\ D_iE^{ia}(\varphi(t',\textbf{x})). \end{aligned}$$ Differentiating both sides with respect to $t$, we obtain the following equation: $$\begin{aligned} E_i^{\ a}(\varphi(x))\dot{\varphi}^i(x)=\dot{\bar{\varphi}}_0^a(x) +\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}D_iE^{ia}(\varphi(x)), \label{Langevin1}\end{aligned}$$ where $\dot{\varphi}^i=\partial_t \varphi^i$ and the correlation function of $\dot{\bar{\varphi}}_0^a$ is given by $$\begin{aligned} \langle\dot{\bar{\varphi}}_0^a(t,\textbf{x})\dot{\bar{\varphi}}_0^b(t',\textbf{x})\rangle &=\delta^{ab}\int \frac{d^{D-1}p}{(2\pi)^{D-1}}\ (H^2a(t))^2\delta(Ha(t)-p)\delta(Ha(t)-Ha(t')) \notag\\ &\hspace{8em}\times \frac{2^{D-3}\Gamma^2(\frac{D-1}{2})}{\pi}\frac{H^{D-2}}{p^{D-1}}\notag\\ &=\frac{2H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}\delta^{ab}\delta(t-t'). \label{white1}\end{aligned}$$ It should be recalled that the time dependence of $\bar{\varphi}_0^a$ appears only through the step function which is nonzero only at the superhorizon scale. As a consequence of this fact, the correlation function of $\dot{\bar{\varphi}}_0^a$ is proportional to the temporal delta function. This type of fluctuation is called a white noise. The equation in (\[Langevin1\]) and (\[white1\]) is known as a Langevin equation with white noise [@Risken]. From the Langevin equation, we can derive the equation satisfied by the probability density $\rho$: $$\begin{aligned} \dot{\rho}(t,\phi) =\ &\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \frac{\partial^2}{\partial\phi^i\partial\phi^j}\left\{G^{ij}(\phi)\rho(t,\phi)\right\} \notag\\ &-\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \frac{\partial}{\partial\phi^i}\left\{\left[\frac{\partial}{\partial \phi^j}E^i_{\ a}(\phi) E^{ja}(\phi) +E^i_{\ a}(\phi)D_jE^{ja}(\phi)\right]\rho(t,\phi)\right\} \notag\\ =\ &\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \frac{\partial^2}{\partial\phi^i\partial\phi^j}\left\{G^{ij}(\phi)\rho(t,\phi)\right\} \notag\\ &+\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \frac{\partial}{\partial\phi^i}\left\{\Gamma^i_{\ jk}(\phi)G^{jk}(\phi)\rho(t,\phi)\right\}, \label{FP1}\end{aligned}$$ which is known as a Fokker-Planck equation. By using its solution, we can evaluate the vacuum expectation value (vev) of any operator made of $\varphi^i$ as $$\begin{aligned} \langle F^{i_1\cdots i_n}(\varphi(x))\rangle=\int d^N\phi\ F^{i_1\cdots i_n}(\phi)\rho(t,\phi), \label{F}\end{aligned}$$ where $F^{i_1\cdots i_n}$ denotes an arbitrary function. It should be noted that $\varphi^i$ are operators while $\phi^i$ are $c$-numbers. For a general case, we review the relation between the Langevin equation and the Fokker-Planck equation in the Appendix. We can rewrite (\[FP1\]) in the following covariant form: $$\begin{aligned} \frac{\dot{\rho}(t,\phi)}{\sqrt{G(\phi)}} =\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} D^iD_i\left\{\frac{\rho(t,\phi)}{\sqrt{G(\phi)}}\right\}. \label{FP1'}\end{aligned}$$ It should be noted that as seen in (\[F\]), $\rho$ itself is not a scalar quantity but $\rho/\sqrt{G}$ is a scalar quantity. The Fokker-Planck equation is exactly solvable for an equilibrium state which is eventually reached: $\rho(t,\phi)\to \rho_\infty(\phi)$ at $t\to\infty$ if it exists. Since $-D^iD_i=D^{\dagger i} D_i$ is non-negative–definite, the solution for the equilibrium state is given by $$\begin{aligned} D_i\left\{\frac{\rho_\infty(t,\phi)}{\sqrt{G(\phi)}}\right\}=0\hspace{1em} \Rightarrow\hspace{1em} \rho_\infty(\phi)=Z^{-1}\sqrt{G(\phi)}, \label{solution1}\end{aligned}$$ where $Z$ is constant. Since the total integral of the probability density is kept at unity, the overall coefficient is fixed as $$\begin{aligned} Z=\int d^N\phi\ \sqrt{G(\phi)}. \label{solution1'}\end{aligned}$$ As seen in (\[solution1\]) and (\[solution1’\]), the convergence of $\sqrt{G(\phi)}$ at $|\phi|\to\infty$ determines whether an equilibrium state is eventually reached or not. If the asymptotic behavior of $\sqrt{G(\phi)}$ is at most $|\phi|^{-\beta}$, $\beta>0$, no equilibrium state is reached. Here we mention the previous study of the stochastic equation for the same model [@Tada2018]. The previous study did not take into account the contribution from the second term in the left-hand side of (\[FE1\]). As discussed in (\[vertex\])–(\[right2\]), this term also induces the leading IR logarithms and leads to the drift term in (\[Langevin1\]). In the previous study, this drift term was introduced by hand, backwards from the Fokker-Planck equation (\[FP1’\]). Let us recall that the diagrams with a loop of twice-differentiated propagators (\[diagrams\]) induce this drift term. As seen in (\[disappear\]), the IR logarithm does not have any contribution to the twice-differentiated propagator. This fact shows that we need to take into account the contribution from the subhorizon scale if derivative interactions are present. We give a concrete example which induces an equilibrium state: $$\begin{aligned} G_{ij}(\varphi)=\exp\left(-\frac{f^2}{2N}\varphi^k\varphi^k\right)\delta_{ij}, \label{compact}\end{aligned}$$ where $f$ is a coupling constant. In this case, the saturation values of $\langle (\varphi^i(x)\varphi^i(x))^n \rangle$ are given by $$\begin{aligned} \rho_\infty(\phi)=Z^{-1}\exp\left(-\frac{f^2}{4}\phi^k\phi^k\right) \hspace{1em}\Rightarrow\hspace{1em} \langle (\varphi^i(x)\varphi^i(x))^n \rangle|_{t\to\infty} =\frac{\Gamma(\frac{N}{2}+n)}{\Gamma(\frac{N}{2})}\left(\frac{2}{f}\right)^{2n}. \label{example}\end{aligned}$$ The $1/f^{2n}$ dependences of (\[example\]) can be understood in the following way. Up to the leading logarithm accuracy, the time evolution of $\langle (\varphi^i(x)\varphi^i(x))^n \rangle$ is given by $$\begin{aligned} \langle (\varphi^i(x)\varphi^i(x))^n \rangle \simeq H^{n(D-2)}\sum_{m=0}^\infty c_m (f^2H^{D-2})^m\log^{n+m}(a(t)/a_0), \label{estimate1}\end{aligned}$$ where $c_m$ denote numerical coefficients. In the case that an equilibrium state is eventually reached, the vev grows until the time when the perturbation theory breaks down: $$\begin{aligned} f^2H^{D-2}\log (a(t_c)/a_0)\sim 1\hspace{1em}\Leftrightarrow\hspace{1em} t_c-t_0\sim \frac{1}{f^2H^{D-2}}\frac{1}{H}, \end{aligned}$$ where $t_0=\frac{1}{H}\log a_0$. Therefore, substituting $\log (a(t)/a_0)\sim 1/(f^2H^{D-2})$, we can estimate the saturation value as $$\begin{aligned} \langle (\varphi^i(x)\varphi^i(x))^n \rangle \sim H^{n(D-2)}\cdot \frac{1}{(f^2H^{D-2})^n}=\frac{1}{f^{2n}}\hspace{1em}\text{at $t-t_0\gg t_c-t_0$}. \label{estimate2}\end{aligned}$$ We also mention the flat space limit of the IR effect. Although the saturation value (\[example\]) itself does not depend on the Hubble constant, this fact does not mean that this calculation result also holds true in flat space. As discussed above, $\langle (\varphi^i(x)\varphi^i(x))^n \rangle$ approaches the saturation value at $t-t_0 \gg t_c-t_0$. This time region disappears in the flat space limit $H\to 0$. Therefore, we should take the flat space limit of (\[estimate1\]) rather than that of (\[estimate2\]). The flat space limit of (\[estimate1\]) shows the expected result that there is no IR effect in flat space. More general model ================== As a more general model, we consider the hybrid model which consists of a noncanonical kinetic term and a potential term: $$\begin{aligned} S=\int \sqrt{-g}d^Dx\ \left[-\frac{1}{2}G_{ij}(\varphi)g^{\mu\nu}\partial_\mu\varphi^i\partial_\nu\varphi^j-V(\varphi)\right]. \label{Hybrid}\end{aligned}$$ We consider the case that each coupling in the potential term, which is made dimensionless by the Hubble constant, is kept small. In other words, the scalar fields can be identified as pseudo Nambu-Goldstone bosons. The field equation for the hybrid model is written in the covariant form $$\begin{aligned} \frac{1}{\sqrt{-g}}\partial_\mu\big\{\sqrt{-g}\ E_i^{\ a}(\varphi)g^{\mu\nu}\partial_\nu\varphi^i\big\}-D_iE_j^{\ a}(\varphi)g^{\mu\nu}\partial_\mu\varphi^i\partial_\nu\varphi^j -E^{ia}(\varphi)D_iV(\varphi)=0, \label{FE2}\end{aligned}$$ and thus the Yang-Feldman equation is given by $$\begin{aligned} -i\int d^D&x'\ \partial'_\mu\left\{\sqrt{-g(x')}g^{\mu\nu}(x')\partial'_\nu G^R(x,x') E_i^{\ a}(\varphi(x'))\right\} \varphi^i(x') \notag\\ &=\varphi_0^a(x) -i\int\sqrt{-g(x')}d^Dx'\ G^R(x,x')\left\{D_iE_j^{\ a}(\varphi(x')) g^{\mu\nu}(x')\partial'_\mu\varphi^i(x')\partial'_\nu\varphi^j(x')\right. \notag\\ &\hspace{17.5em}\left.+E^{ia}(\varphi(x'))D_iV(\varphi(x'))\right\}. \label{YF2}\end{aligned}$$ Applying the leading logarithm approximation (\[right1\]), (\[right2\]), (\[right3\]) and (\[left\]), the Yang-Feldman equation reduces to $$\begin{aligned} E_i^{\ a}(\varphi(x))\varphi^i&(x) -\int^tdt'\ \partial'_tE_i^{\ a}(\varphi(t',\textbf{x}))\varphi^i(t',\textbf{x}) \notag\\ &=\bar{\varphi}_0^a(x) +\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \int^tdt'\ D_iE^{ia}(\varphi(t',\textbf{x})) \notag\\ &\hspace{3.8em}-\frac{1}{(D-1)H}\int^tdt'\ E^{ia}(\varphi(t',\textbf{x}))D_iV(\varphi(t',\textbf{x})). \end{aligned}$$ Differentiating both sides with respect to $t$, we obtain the Langevin equation with white noise: $$\begin{aligned} E_i^{\ a}(\varphi(x))\dot{\varphi}^i(x)=\dot{\bar{\varphi}}_0^a(x) &+\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}D_iE^{ia}(\varphi(x)) \notag\\ &-\frac{1}{(D-1)H}E^{ia}(\varphi(x))D_iV(\varphi(x)), \label{Langevin2}\end{aligned}$$ $$\begin{aligned} \langle\dot{\bar{\varphi}}_0^a(t,\textbf{x})\dot{\bar{\varphi}}_0^b(t',\textbf{x})\rangle =\frac{2H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}\delta^{ab}\delta(t-t'). \label{white2}\end{aligned}$$ Here we explain why the Langevin equation for the hybrid model can be derived in the same way as that for the nonlinear sigma model. Since (\[right1\]), (\[right3\]) and (\[left\]) are the approximations of the free field $\varphi_0$, we can apply them regardless of the type of interactions. Although (\[right2\]) is the approximation of the interacting field $\varphi$, it holds true regardless of whether the potential term is present or not. Since the differential operators in (\[right2\]) reduce the number of IR logarithms from nonderivative interactions, (\[right2\]) does not depend on the potential term up to the leading logarithm accuracy. As discussed in (\[diagrams\]), we can keep up the number of IR logarithms from derivative interactions by constructing a loop of twice-differentiated propagators. From the Langevin equation in (\[Langevin2\]) and (\[white2\]), we can derive the Fokker-Planck equation: $$\begin{aligned} \frac{\dot{\rho}(t,\phi)}{\sqrt{G(\phi)}} &=\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} D^iD_i\left\{\frac{\rho(t,\phi)}{\sqrt{G(\phi)}}\right\} +\frac{1}{(D-1)H}D^i\left\{D_iV(\phi)\frac{\rho(t,\phi)}{\sqrt{G(\phi)}}\right\} \notag\\ &=\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} D^i\left\{\left[D_i+\frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D}D_iV(\phi)\right] \frac{\rho(t,\phi)}{\sqrt{G(\phi)}}\right\}. \label{FP2}\end{aligned}$$ In order to solve the Fokker-Planck equation for an equilibrium state, it is useful to introduce $$\begin{aligned} \Psi(t,\phi)=\exp\left\{\frac{1}{2}\frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D}V(\phi)\right\} \frac{\rho(t,\phi)}{\sqrt{G(\phi)}}. \end{aligned}$$ We can rewrite (\[FP2\]) as the equation for the rescaled probability density: $$\begin{aligned} \dot{\Psi}(t,\phi)=-\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} \mathcal{A}^{\dagger i} \mathcal{A}_i\Psi(t,\phi), \end{aligned}$$ $$\begin{aligned} \mathcal{A}_i&=D_i+\frac{1}{2}\frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D}D_iV(\phi), \notag\\ \mathcal{A}^{\dagger i}&=-D^i+\frac{1}{2}\frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D}D^iV(\phi). \end{aligned}$$ Since $\mathcal{A}^{\dagger i} \mathcal{A}_i$ is non-negative–definite, the solution for the equilibrium sate, $\Psi(t,\phi)\to \Psi_\infty(\phi)$ at $t\to\infty$, is given by $$\begin{aligned} \mathcal{A}_i\Psi_\infty(\phi)=0\hspace{1em}\Rightarrow\hspace{1em} \rho_\infty(\phi)=Z^{-1} \sqrt{G(\phi)}\exp\left\{-\frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D}V(\phi)\right\}, \label{solution2}\end{aligned}$$ where the normalization factor is given by $$\begin{aligned} Z=\int d^N\phi\ \sqrt{G(\phi)}\exp\left\{-\frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D}V(\phi)\right\}. \label{solution2'}\end{aligned}$$ If $V(\phi)$ behaves like $\lambda|\phi|^\alpha,\ \lambda,\ \alpha>0$ at $|\phi|\to \infty$, even though the asymptotic behavior of $\sqrt{G(\phi)}$ is at most $|\phi|^{-\beta}$, $\beta>0$, an equilibrium state is eventually reached. The equilibrium solution in (\[solution2\]) and (\[solution2’\]) indicates a relation to the Euclidean field theory on a $D$-dimensional sphere. The IR effects discussed in this paper are interpreted as nonequilibrium phenomena as they break the dS symmetry and evolve with time. However, if an equilibrium state is eventually reached, it may be described by the Euclidean field theory on a $D$-dimensional sphere. For nonderivative interactions, it has been known that the equilibrium solution of the stochastic equation can be reproduced by considering the zero mode dynamics in the Euclidean field theory on a $D$-dimensional sphere [@Hu1986; @Hu1987; @Rajaraman2010]. As shown below, the discussion holds true even when the kinetic term is noncanonical. The scalar fields can be expanded by the spherical harmonics $Y_{\{l_1,\cdots,l_D\}}$ on a $D$-dimensional sphere as $$\begin{aligned} \varphi^i(x_E)=\sum_{\{l_1,\cdots,l_D\}}\varphi^i_{\{l_1,\cdots,l_D\}}Y_{\{l_1,\cdots,l_D\}}(x_E), \end{aligned}$$ where $x_E$ are the spherical coordinates and $l_1\ge |l_2| \ge \cdots \ge |l_D|\ge 0$ are the angular momentums. In order to evaluate the IR effects, we extract the zero mode as $$\begin{aligned} \varphi^i(x_E)\simeq \varphi^i_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}}. \end{aligned}$$ It should be recalled that the zero mode has no coordinate dependence. In the zero mode approximation, only the potential term contributes to the Euclidean action as $$\begin{aligned} S_E &=\int\sqrt{g_E}d^Dx_E\ \left[\frac{1}{2}G_{ij}(\varphi)g_E^{\mu\nu}\partial_\mu\varphi^i\partial_\nu\varphi^j +V(\varphi)\right] \notag\\ &\simeq V(\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}})\int\sqrt{g_E}d^Dx_E \notag\\ &=V(\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}})\cdot \frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D}, \label{SE}\end{aligned}$$ where the coefficient of the potential term is nothing but the volume of a $D$-dimensional sphere of radius $1/H$. In the path integral formalism with (\[SE\]), the vev of any operator made of $\varphi^i$ is evaluated as $$\begin{aligned} &\langle F^{i_1\cdots i_n}(\varphi(x))\rangle \label{vev'}\\ =\ &\frac{\int \sqrt{\mathcal{G}}\mathcal{D}\varphi\ F^{i_1\cdots i_n}(\varphi(x))e^{-S_E}} {\int \sqrt{\mathcal{G}}\mathcal{D}\varphi\ e^{-S_E}} \notag\\ \simeq\ &\frac{\int\sqrt{G}d^N(\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}})\ F^{i_1\cdots i_n}(\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}}) \exp\left\{-\frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D} V(\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}})\right\}} {\int\sqrt{G}d^N(\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}})\ \exp\left\{-\frac{2\pi^\frac{D+1}{2}}{\Gamma(\frac{D+1}{2})H^D} V(\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}})\right\}}. \notag\end{aligned}$$ The covariant functional integral $\sqrt{\mathcal{G}}\mathcal{D}\varphi$ reduces to $\sqrt{G}d^N(\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}})$ in the zero mode approximation. Identifying $\varphi_{\{0,\cdots,0\}}Y_{\{0,\cdots,0\}}$ as $\phi$, it can be confirmed that (\[vev’\]) is equivalent to the stochastic evaluation with the equilibrium solution in (\[solution2\]) and (\[solution2’\]). Conclusion ========== We extended the resummation formula of the leading IR logarithms so that it is applicable to a general scalar field theory whose kinetic term is noncanonical. In the presence of derivative interactions, we need to take into account not only the contribution from the superhorizon scale but also that from the subhorizon scale. The discussion around (\[vertex\])–(\[right2\]) shows that up to the leading logarithm accuracy, we can integrate out differentiated fields as if nondifferentiated fields were constant. Specifically, the subhorizon dynamics can be treated as in the free field theory with the coefficient of the kinetic term $G_{ij}(\varphi)$. After integrating out the contribution from the subhorizon scale, we treat nondifferentiated fields as dynamical variables. In the general scalar field theory, the leading logarithm approximation of the Yang-Feldman equation leads to a Langevin equation with white noise. As seen in (\[FP2\]), the resulting stochastic equation can be understood as a generalization of the standard one to curved target spaces. The generalized stochastic equation shows that if the target space is sufficiently compact as in (\[compact\]), even though the potential term is absent, an equilibrium state is eventually reached. The equilibrium state can be reproduced by considering the zero mode dynamics in the Euclidean field theory on a sphere. It should be emphasized that if the target space is not so compact, i.e. the asymptotic behavior of $\sqrt{G(\phi)}$ is at most $|\phi|^{-\beta}$, $\beta>0$, and the potential term is absent, no equilibrium state is reached. We derived the IR resummation formula for the general scalar field theory as a first step to obtain that for gravity. The Einstein gravity consists of derivative interactions in a similar way to the general scalar field theory[^5] and includes gauge degrees of freedom and tensor fields in contrast to it. This study shows how to resum the leading IR logarithms from derivative interactions of the scalar field. The resummation formula in the presence of gauge degrees of freedom and tensor fields is a subject for future research. The contribution from gravity should be considered also in evaluating the IR effects in inflation theories. If only the inflaton induced the IR effects, we could evaluate them nonperturbatively by using the current resummation formula. However, there is no reason to neglect the IR effects from gravity. In particular, if the pseudo shift symmetry is respected, the IR effects from the gravitational interaction are more dominant compared with those from the self-interaction of the inflaton [@Kitamoto2017]. Acknowledgment {#acknowledgment .unnumbered} ============== This work was supported in part by the National Center of Theoretical Sciences (NCTS) and Grant-in-Aid for Scientific Research (B) No. 26287044. We thank G. Cho, C. S. Chu, K. Furuuchi, B. L. Hu, C. H. Kim, Y. Kitazawa, S. P. Miao, S. Nojiri, T. Tanaka, and R. P. Woodard for discussions. In particular, we thank G. Cho and C. H. Kim for their participation at an early stage of this work, and K. Furuuchi for reading the manuscript and making comments. Langevin and Fokker-Planck equations ==================================== Here we review the relation between the Langevin equation and the Fokker-Planck equation. For a derivation, see, e.g., [@Risken]. If the Langevin equation is expressed as $$\begin{aligned} \dot{\xi}^i(t)=A^i_{\ a}(t,\xi(t))\eta^a(t)+B^i(t,\xi(t)), \label{Langevin0}\end{aligned}$$ $$\begin{aligned} \langle \eta^a(t)\rangle=0,\hspace{1em} \langle \eta^a(t)\eta^b(t')\rangle=\delta^{ab}\delta(t-t'). \label{white0}\end{aligned}$$ The corresponding Fokker-Planck equation is given by $$\begin{aligned} \dot{\rho}(t,x)= \ &\frac{1}{2}\frac{\partial^2}{\partial x^i\partial x^j}\left\{A^i_{\ a}(t,x)A^{ja}(t,x)\rho(t,x)\right\} \notag\\ &-\frac{\partial}{\partial x^i}\left\{\left[\frac{1}{2}\frac{\partial}{\partial x^j}A^i_{\ a}(t,x) A^{ja}(t,x) +B^i(t,x)\right]\rho(t,x)\right\}. \label{FP0}\end{aligned}$$ Using its solution, the vev of any operator made of $\xi^i$ can be evaluated as $$\begin{aligned} \langle F^{i_1\cdots i_n}(\xi(t))\rangle=\int d^Nx\ F^{i_1\cdots i_n}(x)\rho(t,x), \label{F0}\end{aligned}$$ where $F^{i_1\cdots i_n}$ denotes an arbitrary function. In deriving (\[FP1\]) from (\[Langevin1\]) and (\[white1\]), we made the following replacement: $$\begin{aligned} &\xi^i(t)\to \varphi^i(t,\textbf{x}), \notag\\ &A^i_{\ a}\to \left\{\frac{2H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}\right\}^\frac{1}{2} E^i_{\ a}, \notag\\ &B^i\to \frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}E^i_{\ a}D_jE^{ja}. \end{aligned}$$ In the main text, the probability density for one-point functions is discussed. Here we study the probability density for spatially separated two-point functions. In order to study it, we need to know the correlation function of $\dot{\bar{\varphi}}_0^a$ at spatially separated points: $$\begin{aligned} \langle\dot{\bar{\varphi}}_0^a(t,\textbf{x})\dot{\bar{\varphi}}_0^b(t',\textbf{x}')\rangle &=\frac{2H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})}\delta^{ab}\delta(t-t') \theta(1-Ha(t)r), \label{white1'}\end{aligned}$$ where $r=|\textbf{x}-\textbf{x}'|$. In deriving (\[white1’\]), we approximated the Bessel function of the first kind by the step function as $$\begin{aligned} \frac{2^\frac{D-3}{2}\Gamma(\frac{D-1}{2})}{(Ha(t)r)^\frac{D-3}{2}}J_\frac{D-3}{2}(Ha(t)r) \simeq \theta(1-Ha(t)r). \end{aligned}$$ From (\[Langevin1\]) and (\[white1’\]), the Fokker-Planck equation for spatially separated two-point functions is given by $$\begin{aligned} \frac{\dot{\rho}(t,\phi,\phi')}{\sqrt{G(\phi)G(\phi')}}= \ &\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} D^iD_i\left\{\frac{\rho(t,\phi,\phi')}{\sqrt{G(\phi)G(\phi')}}\right\} \\ &+\frac{H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} D'^iD'_i\left\{\frac{\rho(t,\phi,\phi')}{\sqrt{G(\phi)G(\phi')}}\right\} \notag\\ &+\theta(1-Ha(t)r)\frac{2H^{D-1}}{(4\pi)^\frac{D}{2}}\frac{\Gamma(D-1)}{\Gamma(\frac{D}{2})} D_i\left\{E^i_{\ a}(\phi)D'_j\left[E^{ja}(\phi')\frac{\rho(t,\phi,\phi')}{\sqrt{G(\phi)G(\phi')}}\right]\right\}, \notag\end{aligned}$$ where $\phi$ and $\phi'$ denote $\phi(\textbf{x})$ and $\phi(\textbf{x}')$. Using its solution, any spatially separated two-point function can be evaluated as $$\begin{aligned} \langle P^{i_1\cdots i_m}(\varphi(t,\textbf{x}))Q^{j_1\cdots j_n}(\varphi(t,\textbf{x}'))\rangle =\int d^N\phi d^N\phi'\ P^{i_1\cdots i_m}(\phi)Q^{j_1\cdots j_n}(\phi')\rho(t,\phi,\phi'), \end{aligned}$$ where $P^{i_1\cdots i_m}$ and $Q^{j_1\cdots j_n}$ denote arbitrary functions. [999]{} A. Vilenkin and L. H. Ford, Phys. Rev. D **26**, 1231 (1982). A. D. Linde, Phys. Lett. **116B**, 335 (1982). A. A. Starobinsky, Phys. Lett. **117B**, 175 (1982). J. S. Schwinger, J. Math. Phys. **2**, 407 (1961). L. V. Keldysh, Zh. Eksp. Teor. Fiz. **47**, 1515 (1964) \[Sov. Phys. JETP **20**, 1018 (1965)\]. A. A. Starobinsky, Lect. Notes Phys. **246**, 107 (1986). A. A. Starobinsky and J. Yokoyama, Phys. Rev. D **50**, 6357 (1994) \[astro-ph/9407016\]. N. C. Tsamis and R. P. Woodard, Nucl. Phys. B **724**, 295 (2005) \[gr-qc/0505115\]. L. Pinol, S. Renaux-Petel, and Y. Tada, Class. Quant. Grav.  [**36**]{}, no. 7, 07LT01 (2019) \[arXiv:1806.10126 \[gr-qc\]\]. H. Kitamoto and Y. Kitazawa, Phys. Rev. D **83**, 104043 (2011) \[arXiv:1012.5930 \[hep-th\]\]. H. Kitamoto and Y. Kitazawa, Phys. Rev. D **85**, 044062 (2012) \[arXiv:1109.4892 \[hep-th\]\]. N. C. Tsamis and R. P. Woodard, Commun. Math. Phys. **162**, 217 (1994). H. Kitamoto, Y. Kitazawa, and R. Kojima, Phys. Rev. D **96**, no. 2, 023535 (2017) \[arXiv:1704.08845 \[hep-th\]\]. I. S. Gradshteyn and I. M. Ryzhik, *Table of Integrals, Series, and Products* (Academic Press, New York, 2007). S. P. Miao, N. C. Tsamis, and R. P. Woodard, J. Math. Phys. **51**, 072503 (2010) \[arXiv:1002.4037 \[gr-qc\]\]. H. Risken, *The Fokker-Planck Equation: Methods of Solution and Applications* (Springer, New York, 1989). B. L. Hu and D. J. O’Connor, Phys. Rev. Lett. **56**, 1613 (1986). B. L. Hu and D. J. O’Connor, Phys. Rev. D **36**, 1701 (1987). A. Rajaraman, Phys. Rev. D **82**, 123522 (2010) \[arXiv:1008.1271 \[hep-th\]\]. [^1]: E-mail: [email protected] [^2]: The statement holds true only if the retarded propagator starting at $x'$ is not differentiated. Therefore, it does not apply to the left-hand side of (\[YF1\]) with $\partial'_\nu G^R(x,x')$. [^3]: The calculation technique can also be adopted in evaluating the energy-momentum tensor of the nonlinear sigma model [@Kitamoto2012]. [^4]: Each contribution from the $(-p\tau')^n$ term for $n\ge 1$ induces the leading IR logarithms, while the leading IR logarithms from the higher-order terms are canceled in total. The cancellation holds true in scalar field theories where the Lorentz indices are contracted within differential operators. [^5]: More specifically, neither the Einstein gravity nor the models adopted in this paper include higher derivative interactions.
{ "pile_set_name": "ArXiv" }
25.0 cm 17.5cm 0.0cm 0.0cm 0.0cm -0.7cm -0.7cm 0.35cm P. Grandi$^1$, C. M. Urry$^2$, L. Maraschi$^{3}$, M. Guainazzi$^4$, E. Massaro$^{5}$, G. Matt$^{6}$, L. Bassani$^7$, A. Cimatti$^8$, P. Giommi$^9$, M. Dadina$^{9}$, G. C. Perola$^{6}$, L. Piro$^1$, A. Santangelo$^{10}$. $^1$[ IAS/CNR, Area di Ricerca Tor Vergata, Via Fosso del Cavaliere, I-00133 Roma, Italy ]{}\ $^2$[ STScI, 3700 San Martin Drive, Baltimore, MD 21218, USA]{}\ $^3$[ Osservatorio Astronomico di Brera, Via Brera 28, I-20121 Milano, Italy]{}\ $^4$[ SSD/ESA, ESTEC, Postbus 299, 2200 AG Noordwijk, The Netherlands]{}\ $^{5}$[ Istituto Astronomico di Roma, Via Lancisi 29 Roma, Italy]{}\ $^{6}$[ Università degli Studi “Roma 3”, Via della Vasca Navale 84, I-00146 Roma, Italy]{}\ $^7$[ ITESRE/CNR, Via P. Gobetti 101, I-40129, Bologna, Italy]{}\ $^8$[ Osservatorio Astronomico di Arcetri, Largo E. Fermi 5, I-50125, Firenze, Italy]{}\ $^9$[ BeppoSAX SDC, c/o Nuova Telespazio, Via Corcolle 19, I-90146, Roma, Italy]{}\ $^{10}$[IFCAI/CNR, Via U. La Malfa 153, I-90146, Palermo, Italy]{}\ ABSTRACT {#abstract .unnumbered} ======== We present preliminary results from two observations of the radio galaxy Centaurus A performed by the BeppoSAX satellite on 1997 February 20-21 and on 1998 January 6-7. In the second pointing the source was brighter by a factor 1.3. We did not detect any spectral variation of the nuclear continuum in spite of the long-term flux change between the two observations. At both epochs, the nuclear point-like emission was well fitted with a strongly absorbed ($N_H\sim10^{23}$ cm$^{-2}$) power law with an exponential cutoff at high energies (E$_{cut}> 200$ keV). We also observed a significant flux variation of the iron line between the two observations. The flux of the line and of the continuum changed in the opposite sense. The line is more intense at the first epoch, when the nuclear source was at the lower intensity level. The implied delay between the continuum and line variations strongly suggests that the cold material responsible for the iron line production is not located very near to the primary X-ray source. There is also evidence that the line profile changed between the two epochs, being broader and slightly blueshifted when the source was fainter. It is possible that the emission feature is a blend of cold and ionized iron lines produced in separate regions surrounding the nuclear source. INTRODUCTION ============ The radio galaxy Centaurus A (Cen A) is a well-known radio-loud AGN which has been studied extensively over the whole X-ray and $\gamma$-ray range. Its proximity (z=0.008) means its X-ray emission has been spatially resolved, into at least five different regions: the nucleus, the jet, a middle NE radio lobe between 20$^\prime$-25$^\prime$ from the nucleus, a faint diffuse emission within 6$^\prime$ probably from hot interstellar medium and two ridges of thermal emission along each edge of the dust line (Feigelson et al. 1981, Turner et al. 1997). Above 3 keV the spectral emission is essentially dominated by the nuclear point-like source, while at soft energies the extended thermal components become important because of the large extinction suffered by the nucleus. The X-ray spectrum between 3-10 keV is generally modeled by a power law heavily absorbed by dense and probably stratified matter and by a fluorescence iron line (Morini et al. 1989, Miyazaki et al 1996, Turner et al. 1997, Sugizaki et al. 1998). Hard X-ray and soft $\gamma$-ray observations performed with the GRO satellite show a steepening of the power law at energies larger than 120 keV (Kinzer et al. 1995, Steinle et al. 1997). Here we present preliminary results of two observations of Cen A performed with the BeppoSAX satellite. We concentrate on the nuclear X-ray emission, giving information on the continuum spectral shape and on the iron line feature. RESULTS ======= The Narrow Field Instruments (LECS: 0.1-10 keV; MECS: 1-10 keV; HPGSPC: 4-120 keV; PDS: 15-300 keV) of BeppoSAX (Parmar et al. 1997, Boella et al. 1997, Manzo et al. 1997, Frontera et al. 1997) observed Cen A twice on 1997 February 20-12 and 1998 January 6-7 for 35 and 75 ksec, respectively. The average source flux increased between the two observations, being brighter by about a factor 1.3 in 1998. The brightness level was however relatively low in both observations compared to the historic light curve. The observed flux, $F_{\rm 3-12~keV}\sim 0.2-0.3\times10^{-9}$ photons cm$^{-2}$ sec$^{-1}$, is within the range observed in the last 10 years and lower at least by a factor $\sim 5$ than the outburst in 1974-75 (Turner et al. 1997). Here we present the results of the preliminary analysis of the MECS and PDS instruments, covering the 1.5-150 keV energy band. The MECS spectra were accumulated with 8$^\prime$ extraction radii. Because of a failure in the MECS unit 1 on 1997 May 6 only data from two units were available in the 1998 observation. The extraction region of the MECS was large enough to include extended components, in particular the X-ray emission from the galactic ridges and from the jet. The diffuse thermal ($kT=0.9$ keV) emission produced by interstellar medium is very soft and therefore negligible in the 1.5-150 keV range. A spatially resolved spectroscopic study based on ROSAT observations (Turner et al. 1997) has shown that the galactic ridge emission can be well fitted with a Raymond-Smith model ($kT=5$ keV, metal abundances 0.4 times solar) modified by Galactic absorption ($N_H=7\times10^{20}$ cm$^{-2}$), and the jet emission with an absorbed power law (N$_H=9.7 \times10^{20}$ cm$^{-2}$, $\Gamma=2.29$). We used those parameters to model the extended emission components and used the observed luminosity corrected by absorption ($L_{\rm jet}^{\rm 0.1-2~keV}=2\times10^{39}$ erg sec$^{-1}$; $L_{\rm gal. ridges}^{\rm 0.1-2~keV}=2.4\times10^{39}$ erg sec$^{-1}$) to set the normalization constants. [lccccccccc]{} &&&&&&&&&\ Model & A$^a$ &$N_{H}$ & $\Gamma$ & E$_{\it cutoff}$ & R& E$_{\it Fe}$ & $\sigma$ & F$_{\it Fe}^b$ & $\chi^2(dof)$\ && $10^{22}$ cm$^{-2}$& & (keV) & & keV & keV &&\ &&&&&&&&&\ &&&&&&&&&\ \ &&&&&&&&&\ PL &8.8$\pm0.3$&9.9$\pm0.2$ & 1.77$^{+0.02}_{-0.01}$& ... &... &6.49$^{+0.07}_{-0.08}$& 0.3$\pm0.1$& 5.1$^{+1.1}_{-0.9}$&124(89)\ CPL & 7.7$\pm0.4$ & 9.5$\pm0.2$& 1.68$\pm0.03$& 215$^{+107}_{-55}$& 0.0 & 6.47$^{+0.06}_{-0.07}$ & 0.3$\pm0.1$ & 4.8$^{+1.1}_{-0.8}$& 97(88)\ CPL+ REF & 8.5$^{+0.9}_{-0.6}$ & 9.8$^{+0.2}_{-0.3}$ & 1.76$^{+0.04}_{-0.06}$ & 286$^{+257}_{-94}$& 0.21$^{+0.05}_{-0.14}$&6.48$\pm0.07$ & 0.3$\pm0.1$ & 5.2$^{+1.2}_{-1.0}$&90(87)\ &&&&&&&&&\ &&&&&&&&&\ \ &&&&&&&&&\ PL &11.9$\pm0.2$&9.6$^{+0.1}_{-0.2}$ & 1.78$\pm0.01$&... &... & 6.39$^{+0.07}_{-0.08}$& 0.07$^{+0.22}_{-0.07}$ & 2.5$^{+1.1}_{-0.7}$ & 147(89)\ CPL &10.5$\pm0.4$&9.2$\pm0.2$& 1.70$\pm0.02$&250$^{+81}_{-50}$& 0.0 & 6.38$^{+0.07}_{-0.08}$ & 0.06$^{+0.23}_{-0.06}$ &2.5$^{+0.8}_{-0.7}$ &98(88)\ CPL+REF &11.1$^{+0.9}_{-0.7}$ & 9.2$^{+0.3}_{-0.1}$ & 1.73$^{+0.05}_{-0.04}$& 297$^{+156}_{-79}$ & 0.08$^{+0.09}_{-0.08}$& 6.38$\pm0.08$ & 0.08$^{+0.23}_{-0.08}$ & 2.6$\pm0.8$ & 95(87)\ &&&&&&&&\ \ \ With the contributions of the galactic ridges and the jet fixed, we studied the spectral shape of the nuclear point-like component. In Table 1 are listed the spectral fit results from both observations. An absorbed power law (PL) was not an acceptable model (see Table 1 and Fig. 1) for the continuum in both the observations. More reasonable fits were obtained with a cutoff power law (CPL). The CPL fit to the 1997 data improved significantly ($P_{F_{\it test}}=98.9\%$) if a reflection component was added to the model (R in Table 1 is the ratio between the reflection and power law normalizations). In the 1998 data, characterized by a larger nuclear flux, the inclusion of a reflection component gave a less significant improvement ($P_{F_{\it test}}=94\%$) and R was consistent with zero. It is possible that reprocessed X-ray radiation was present in the first observation and not in the second, as can occur if the X-ray reflecting region is far from the nuclear source. However, since the amount of reflection is strongly dependent on the relative flux normalization adopted for the MECS and PDS instruments, which have intercalibration uncertainties of $\sim 3\%$, and since the amount of reflection ($R=0.2$) is quite small, we can not exclude that the reflection component may be produced by miscalibration effects. Comparing the results in Table 1, we can conclude that the spectral shape of the primary X-ray radiation did not change significantly with the intensity. A Gaussian profile fits the iron line in both observations, but in contrast to the primary X-ray continuum spectrum, the emission feature changed shape. Figure 2 (left panel) shows the ratio between the 1997 and 1998 spectra. It is evident that the ratio is constant across the entire 3-150 keV range apart from the iron line region. The iron line was indeed clearly more intense when the source was fainter. As shown in Table 1, the total number of photons in the line was about twice as large in the first observation, when the nuclear flux was $\sim 25\%$ lower. This result implies that the region responsible for the iron line responds with a significant delay to the continuum variations and therefore the reprocessing material cannot be located very near to the primary X-ray source. The data also indicate that the shape of the Gaussian profile changed. The iron line appears broader and slightly blueshifted in the first observation relative to the second (Fig. 2, right panel). Note that, while the iron line is well resolved in the 1997 observation ($\sigma=0.3\pm0.1$ keV), its width is consistent with zero in 1998. One possibility is that the 1997 feature was a blend of cold and ionized lines. As often observed in Seyfert 2 galaxies, the 6.4 keV line might be produced by a molecular dusty torus and the ionized one by warm scattering material out of the line of sight. Alternatively, the reprocessing material might be stratified and roughly spherical with the ionized region closer to the nuclear source. The presence of the reflection component when the iron line flux was larger, even if marginal, seems however to favour a toroidal geometry for the cold absorber. CONCLUSION ========== The long term variability of Cen A observed with BeppoSAX offered a unique opportunity to study the spectral variations of the nuclear emission in the medium and hard X-ray (3-150 keV) bands as a function of the intensity. We observed that: (1) Independent of the brightness of the nuclear component, the spectral shape of point-like nuclear source did not change significantly. In both observations, the spectrum was a power law with an exponential cutoff at energies $> 200$ keV, heavily absorbed by a large invariable column density $N_H\sim10^{23}$ cm$^{\-2}$. An additional hard (reflection) component was present when the nuclear source was fainter. Since the amount of reflection was small and the MECS and PDS intercalibration uncertainties are of the order of $\sim 3\%$, this result should be considered with caution. (2) The strength of the iron line flux is not directly correlated with the intensity of the nuclear source. We conclude that material responsible for reprocessing the primary X-ray radiation is not very near to the nuclear source. In addition, the change of line profile (intrinsic width and energy peak) with nuclear flux suggests the feature is a blend of cold and ionized line, probably produced by separate regions with different opacities. REFERENCES ========== - Boella G., et al., A$\&$AS,122, 327 (1997) - Feigelson, E. D., et al., ApJ, 251, 32 (1981) - Frontera et al., A$\&$AS, 122, 357 (1977) - Kinzer, R. L., et al., ApJ, 449, 118 (1995) - Manzo G., et al. A$\&$AS, 122, 341 (1977) - Miyazaki, S., et al., PASJ, 48, 801 (1996) - Morini, M., F. Anselmo , D. Molteni, ApJ, 347, 750 (1989) - Parmar A., et al., A$\&$AS, 122, 309 (1977) - Steinle, H., et al., A$\&$A, 330, 97 (1998) - Sugizaki, M., et al., PASJ, 49, 59 (1997) - Turner, T. J., I. M. George, R. F. Mushotzky, K. Nandra, ApJ 475, 118 (1997)
{ "pile_set_name": "ArXiv" }
--- abstract: 'We discuss the present status of the description of the structure of the very neutron rich nuclei, in the framework of modern large scale shell model calculations. Particular attention is paid to the interaction related issues, as well as to the problems of the shell model approach at the neutron drip line. We present detailed results for nuclei around N=20 and, more briefly, we discuss some salient features of the regions close to N=8, 28 and 40. We show that most experimental features can be understood in a shell model context.' address: - 'Institut de Recherches Subatomiques, IN2P3-CNRS-Université Louis Pasteur, F–67037 Strasbourg Cedex 2, France' - 'Laboratoire de Physique Théorique, Université Louis Pasteur, 3-5 rue de l’Université, F–67084 Strasbourg Cedex, France' - | Departamento de Física Teórica C–XI, Universidad Autónoma de Madrid\ 28049 Madrid, Spain author: - 'E. Caurier' - 'F. Nowacki' - 'A. Poves' title: Shell model studies of neutron rich nuclei --- and Neutron rich nuclei. Large scale shell model calculations. 21.60.Cs, 21.60.-n, 21.10.Hw, 21.10.Ky Introduction ============ Our knowledge of the properties of the nuclei lying far from the valley of stability has increased a lot in the last decade, thanks to the work carried out at the isotope separators on line (ISOL) and fragment separators. In some cases, mostly in light nuclei, the neutron and/or proton drip lines have been reached. The situation will be much improved with the advent of the new generation of Radiactive Ion Beam facilities that will be discussed at length in other papers of this volume. These experimental advances have been accompanied by intense developments in nuclear structure theory. In the region of medium-light nuclei, the shell model description in large valence spaces, that gives the most complete and reliable picture, has become available. A novelty in the very neutron rich side, is that the situation at the Fermi surface resembles that of the heavy nuclei at the valley of stability, with the proton and neutron Fermi levels sitting at different major shells. This represents a new challenge for the shell model approach, because two contiguous major shells have to be included in the valence space, in contrast with the usual calculations in a single major shell. This brings in new problem; on one side the size of the calculations can become very large, demanding novel shell model codes or new techniques; on the other, the effective in medium interaction is richer and therefore more difficult to keep under control. Besides, very close to the neutron drip line, where the physical states are only slightly bound and the wave functions may exhibit very long tails, the validity of an approach based in a Fock space representation may appear at first sight dubious, because of the entanglement of configuration space and Fock space degrees of freedom. The predictive power of the shell model descriptions in these new regions is seriously hindered by the lack of an “universal” shell model interaction. Whereas it has been demonstrated that one can obtain a fully reliable multipole hamiltonian from modern G-matrices, the monopole hamiltonian is usually incorrect [@duzu]. The monopole hamiltonian contains all the terms depending on the number of particles and on the isospin. The isospin dependent terms of the spherical mean field play a dominant role in the location of the different configurations far from stability and therefore determine which dynamical aspects will be manifest in each regime. Unfortunately, the experimental knowledge of the spherical mean field close to the valley of stability is, in general, not sufficient to move safely far out. The success of the large scale shell model calculations depends crucially on the correctness of the monopole hamiltonian, that’s why the experimental information on some “simple” (closed shells plus or minus one nucleon) exotic nuclei is invaluable. The first extensive survey of neutron rich nuclei in the shell model context is due to Wildenthal, Curtin and Brown [@wcb83] using their fitted USD interaction [@usd] to compute energy spectra and beta decay properties of all the neutron rich nuclei in the $sd$-shell. However, some years before, similar calculations by Wildenthal and Chung, when compared with the experimental data of $^{31}$Na, had led these authors to entitle his paper “The collapse of the shell model ordering in the very neutron rich isotopes of Na and Mg” [@wch80]. This was a word of warning on the weird behaviour to be expected when exploring the far from stability land. Later, more experimental findings confirmed this premonitory view, and, nowadays, expressions like “new phases of nuclear matter” or “vanishing of shell closures” are part of our current jargon. N=8: Halos ========== The neutron rich side of the $p$-shell has since long provided us with the most dramatic example of intruder state; the ground state 1/2$^+$ in $^{11}$Be. The expected 0$\hbar \omega$ “normal” state lies at 300 keV. The shell model description of such inversion requires obviously two major shells. The mechanism of this inversion is common to other cases that we will study later and can be schematically understood as consisting of two major ingredients: 1\) The monopole hamiltonian that gives the “unperturbed” or spherical Hartree-Fock energy of the different distributions of the valence particles among the valence orbits (configurations). Far from stability, the energy gaps between these configurations may be eroded because of the small binding of the orbits at the top of the well. 2\) The multipole terms (mainly pairing and quadrupole) that mix the components belonging to each configuration, produce different levels of coherence and different energy gains relative to the centroïd for different configurations, depending on the structure of the spherical mean field. They can even invert the energy ordering of the configurations given by the monopole terms. There is a very nice application of this scheme to the $^{11}$Be case by B. A. Brown in ref. [@bro94]. He examines first the $p$-$sd$ gap evolution towards the neutron rich side, concluding that it is reduced, but still 4 MeV wide. Promoting a particle across the gap cost therefore 4 MeV. However, it opens the possibility for neutron pairing correlations (gain $\sim$ 2 MeV) and also allows for the quadrupole coupling of the $^{10}$Be 2$^+$ with the 1d$_{5/2}$ neutron (gain $\sim$ 2 MeV). Summing up all those contributions the intruder wins. Notice the subtle balance between spherical mean field properties and correlations, the latter depending very much on the detailed location of the orbits around the Fermi level. When many particles many holes excitations are at play, the monopole effects are more involved, as discussed by A. Zuker [@zuk98] in the case of the 4p-4h 0$^+$ excited state of $^{16}$O at $\sim$ 6 MeV and clustering (mainly $\alpha$ correlations) effects are surely present in the physical solution. These calculations and a few similar ones [@n8; @bor98], have provided a solid shell model interpretation to the behaviour of the nuclei in the region, in agreement with or confirmed by subsequent experiments. A most prominent member of this region is $^{11}$Li, which sits at the drip line and has a very small ($\sim$ 200 keV) two-neutron separation energy. The last two neutrons have a very large spatial extension, forming what is called a neutron halo [@halo]. In spite of the exotic matter distribution and of the closeness of the continuum, the shell model picture can still cope with many of the structural properties of $^{11}$Li. The calculations predicted a ground state of $^{11}$Li dominated by a configuration with two neutrons in the 2s$_{1/2}$ orbit. Indeed, the most recent experimental information confirms this extreme [@bor99]. Another fingerprint of the dominance of intruder configurations in $^{11}$Li was provided by a classical beta decay experiment, measuring its lifetime and the branching ratio of its decay to the first excited 1/2$^{-}$ state in $^{11}$Be. The results were only compatible, by large, with the assumption of $s$-wave dominance [@bor98]. Very recent experimental work at MSU, has shown that a similar situation happens in $^{12}$Be. The expectedly semi-magic isotope of Berilium (N=8) turns out to be dominated by the intruder configuration with two neutrons in the $sd$-shell [@navin]. With this, it joins its forerunner cousin $^{32}$Mg in the realm of the intruders. N=20: Intruders =============== It was in this region were the massive breaking of a semi-magic closure far from stability was first detected, in what is known as “the island of inversion” around $^{31}$Na. The anomalous experimental data on the mass and the spin of $^{31}$Na [@na31] were attributed to a transition from spherical to prolate shape at N=20 [@campi]. Later, the measures were extended to other Neon, Sodium and Magnesium isotopes [@n20exp] and were interpreted in a shell model context as due to the inversion of the neutron closed shell configuration and a 2p-2h intruder configuration, intrinsically deformed [@pr87]. The intrusion mechanism, that works as we have explained in section 2, was already sketched in ref. [@pr87]. The N=20 quasiparticle gap diminishes when the neutron rich area is approached. For instance, its experimental value for $^{40}$Ca is $g$=7 MeV, for $^{36}$S is $g$= 5.6 MeV and for $^{34}$Si is $g$= 5.1 MeV. $^{34}$Si is a good reference because it is clearly semi-magic, as we shall discuss later, and very neutron rich. Extrapolating smoothly these numbers one should expect $g\sim$ 4.6 MeV for $^{32}$Mg, whereas experimentally $g$= 3.6 MeV. This difference may be actually due to the onset of deformation. A two particle-two hole neutron excitation across the N=20 closure, would accordingly cost $\sim$ 7 MeV, much less than what it takes in $^{40}$Ca. When the calculations are performed, it turns out that the gain in correlation energy of the intruder state overshoots the monopole gap by 1.5 MeV, producing the famous inversion. Most of the intruder’s gain in correlation energy is quadrupole, because of the presence of open shell $sd$ protons and open shell $pf$ neutrons. When we move to $^{34}$Si, the gap becomes larger and the quadrupole correlation is hindered due to the closure of the 1d$_{5/2}$ proton subshell. As a result, the neutron closed shell is now the ground state and the intruder becomes an excited state. Different groups have made calculations in this region. In ref. [@bro.msu] the diagonalizations were supplemented with a weak coupling approximation to delineate the contour of the “island of inversion”. More recently [@fuku; @siis] similar calculations have been undertaken using different truncations of the $sd$-$pf$ space. In ref. [@us98] the full $sd$-shell for neutrons and the full $pf$-shell for neutrons is considered. The calculations allow up to two particle jumps from the $sd$-shell to the $pf$-shell. Besides, it turns out that the cross shell proton excitations are irrelevant in this zone. There is a good level of agreement between the results of these groups, in particular they share the same strong and weak points. One of the strong points is the description of the structure of the isolated intruders as well as its location relative to the normal states for nuclei N=20. Another common strong point concerns the predicted limits in Z of the island of inversion; in all the calculations only Ne, Na and Mg belong to it, while F and Al sit at its very edge. In the weak side are the limits in N of the region. Different choices of the monopole hamiltonian produce small shifts in the borders. As a consequence, N=19 and N=22 are inside or outside depending on the calculation. Also in the weak side is the amount of mixing between normal and intruder states. Very recently, the Quantum Monte Carlo diagonalization method has been also implemented in this region, using as valence space the $sd$-shell plus the two lower $pf$-shell orbits, 1f$_{7/2}$ and 2p$_{3/2}$ [@otsu99]. The effective interaction employed in this reference has been adjusted to produce a 1d$_{3/2}$–$fp$ gap that decreases rapidly between $^{34}$Si ($g$=4.4 MeV) and $^{28}$O ($g$=1.2 MeV), whereas in our case these figures are 4.7 MeV and 3.4 MeV respectively. It is evident that this choice results in an enhancement of the intruder mixing and in an enlarged “island of inversion”. We shall now present some of our latest results. We use the same valence space we had in ref [@us98] and essentially the same effective interaction. A modification has been forced by the recent experimental measure at Isolde [@guy] of the excitation energy of the 3/2$^-$ state in $^{35}$Si (1 MeV). In order to fix the monopole terms of the cross shell interaction in the $sd$-$pf$ valence space this information is vital. In our old interaction we had taken the conservative view and had put the 3/2$^-$ state at 2 MeV (as in $^{41}$Ca). The effect of this change on the results of ref. [@us98] is not dramatic and amounts to enhance moderately the quadrupole correlations, increasing the binding energy and the deformation of the intruders. Besides, this modification binds $^{31}$F, in agreement with the most recent experimental result [@saku]. In what follows we compare the structure of the normal and intruder states in the different nuclei of interest and give our predictions for their relative position. $^{30}$Mg $^{32}$Mg $^{34}$Mg ------------------------------- ------- ----------- ------- ------- ----------- -------- ------- ----------- -------- N I EXP N I EXP N I EXP $\Delta$E(0$^+_{\textrm{I}}$) +3.1 -1.4 +1.1 0$^+$ 0.0 0.0 0.0 0.0 0.0 0.0 2$^+$ 1.69 0.88 1.48 1.69 0.93 0.89 1.09 0.66 (0.67) 4$^+$ 4.01 2.27 2.93 2.33 (2.29) 2.41 1.86 (2.13) 6$^+$ 6.82 3.75 9.98 3.81 3.52 3.50 BE2 2$^+$ $\rightarrow$ 0$^+$ 53 112 59(5) 36 98 90(16) 75 131 4$^+$ $\rightarrow$ 2$^+$ 35 144 17 123 88 175 6$^+$ $\rightarrow$ 4$^+$ 23 140 2 115 76 176 Q$_{spec}$(2$^+$) -12.4 -19.9 -11.4 -18.1 -15.4 -22.7 : Properties of the even magnesium isotopes. N is for normal and I for intruder. Energies in MeV, BE2’s in e$^2$fm$^4$ and Q’s in efm$^2$ in all the tables We start the tour with the even Mg isotopes with N=18, N=20 and N=22. Our results are gathered in table 1. In $^{30}$Mg the configuration with normal filling gives clearly the best reproduction of the (scarce) existing experimental data [@pry]. Notice however, that the 2p-4h intruder is very collective, actually as much as the intruder in $^{32}$Mg. We can therefore conclude that $^{30}$Mg is outside the inversion zone. In $^{32}$Mg the situation is the opposite. Our calculation places the 2p-2h intruder well below the closed shell configuration. The differences between both are manifest, and the excellent agreement between the properties of the calculated intruder and the experimental data (the 2$^+$ excitation energy [@guille], the 0$^+$ $\rightarrow$ 2$^+$ BE2 [@moto] and the 4$^{+}$ excitation energy [@aza]) make it posible to assign this configuration unambiguously. Data and calculations suggest, at low spin, a prolate deformed structure, certainly perturbed, with $\beta \sim$ 0.5. But, what about the mixing between different np-nh configurations? Indeed, the true physical state must be mixed to a larger or smaller extent. However, in view of the excellent agreement of our fixed 2p-2h solution with the experiment, it is clear that any mixing would deteriorate it. The way out of this dilemma –to mix or not to mix– refers to the effective interaction. The $sd$-shell and $pf$-shell parts of our effective interaction are well suited for 0$\hbar \omega$ calculations, and contain implicitly part of the effects of the cross shell mixing. Thus, in a mixed calculation one has to take care of properly unrenormalising the interaction. With this caveat, the mixed results may come back to agree with the experiment. In $^{34}$Mg the normal configuration that contains two $pf$ neutrons, is already quite collective. It can be seen in the table that it resembles very much the intruder configuration in $^{32}$Mg. Its own 4p-2h intruder is even more deformed ($\beta \sim$ 0.6) and a better rotor. Therefore, it is more difficult to make a sharp distinction between them both and their different mixed combinations. In our calculation the normal state is 1 MeV below the intruder. However, it is by no means excluded that they could be much closer or even than the intruder would come below. We have put in parenthesis the very recent and preliminary results from Riken [@mg34] that seem to favour the intruder option. Let’s mention however, that results equivalent to those given by the intruder configuration alone, can be also obtained with a 50% mixed solution, provided the $pf$-shell pairing is reduced. In the QMCD calculations of ref. [@otsu99], the ground state band is dominantly 4p-2h; the 2$^+$ comes at the right place but the 4$^{+}$ is too high, making the solution to over-rotate. Clearly, $^{34}$Mg is at the edge of the “island of inversion”, whether it is more on the inside or the outside is (theoretically) a matter of subtle arrangements that can only be decided by better experimental data. $^{28}$Ne $^{30}$Ne $^{32}$Ne ------------------------------- ------ ----------- -------- ------ ----------- ----- ------- ----------- ----- N I EXP N I EXP N I EXP $\Delta$E(0$^+_{\textrm{I}}$) +2.6 -1.4 +1.5 0$^+$ 0.0 0.0 0.0 0.0 0.0 0.0 2$^+$ 1.81 0.87 (1.32) 1.90 0.85 1.01 0.68 4$^+$ 3.34 2.21 2.87 2.08 2.09 1.82 6$^+$ 6.35 3.90 3.61 3.29 3.42 BE2 2$^+$ $\rightarrow$ 0$^+$ 36 78 54(27) 29 72 72 100 4$^+$ $\rightarrow$ 2$^+$ 31 105 22 97 74 137 6$^+$ $\rightarrow$ 4$^+$ 15 104 89 60 133 Q$_{spec}$(2$^+$) -1.2 -17.8 -1.1 -16.4 -13.7 -20.0 : Properties of the even neon isotopes. N is for normal and I for intruder. In table 2 we have collected the results for the even Neon isotopes to which, [*mutatis mutandis*]{}, most of the arguments advanced in the discussion of the magnesiums apply. Now, the experimental information is even meagrer than before. Let’s just comment on the $^{28}$Ne case because it has been argued in ref. [@otsu99] that it could be substantially more mixed than its neighbour $^{30}$Mg. This claim originates in the comparison between the $sd$ prediction for the 2$^+$ excitation energy (1.81 MeV using the USD interaction) and the experimental result (1.32 MeV [@aza]). However, there could be another explanation; that the discrepancy were due, instead, to a defect of the USD interaction, for the experimental 2$^+$ excitation energy of $^{28}$Ne (1.32 MeV) is only slightly lower than that of $^{30}$Mg, (1.48 MeV). The recent measure of the BE2 [@pry] does not settle the case yet, because its large error bar do not discards a low mixing scenario. In the N=19 isotones $^{29}$Ne and $^{31}$Mg the situation is even more complex, because the normal configurations are almost degenerate with the opposite parity 1p-2h intruders, while the 2p-3h intruders appear a bit above. The competition for the ground state is between 3/2$^+$ and 3/2$^-$ in both cases. The beta decay data from Isolde [@klotz] favour the positive parity for the ground state of $^{31}$Mg. On its side, the calculation explains the occurrence of such a high level density at low excitation energy in the experiment. In the N=21 isotones $^{31}$Ne and $^{33}$Mg the lowest configurations are the positive parity 2p-1h intruders and the negative parity 3p-2h intruders. Like in the N=19 case they are nearly degenerate and the ground state candidates are 3/2$^+$ and 3/2$^-$. In a recent experimental study of the decay of $^{33}$Na [@walter] it is found that the ground state is 3/2$^+$ . This agrees with our calculation that, in addition, predicts a 3/2$^-$ at 300 keV. In table 3. we show the results for $^{31}$Na. In this nucleus the intruder configuration is clearly dominant and very distinct from the normal one. The calculation reproduces the occurrence of 3/2$^+$ as spin of the ground state as well as the excitation energy of the 5/2$^+$ state, recently measured at MSU [@pryt]. In this reference, shell model calculations along the same lines than ours, although in a somewhat smaller valence space, are reported. They agree reasonably well with the present results. [ccccccccc]{} \ & 2J & N & & 2J & I(2p-2h) & & 2J& exp\ &5$^+$ & 0.0 & & 3$^+$ & 0.0 & & 3$^+$ & 0.0\ &3$^+$ & 0.45 & & 5$^+$ & 0.28 & & (5$^+$) & 0.35(2)\ &1$^+$ & 3.18 & & 7$^+$ & 1.06 & & &\ &7$^+$ & 4.42 & & 1$^+$ & 2.28 & & &\ BE2(5$^+ \rightarrow$ &3$^+$) & 62 &BE2(3$^+ \rightarrow$ & 5$^+$) & 216 & & &\ &1$^+$) & 12 & & 7$^+$) & 110 & & &\ &7$^+$) & 40 & & 1$^+$) & 6 & & &\ N exp I(2p-2h) exp I(1p-1h) exp ------- ------ ----- -- ------- ---------- ------- -- ------- ---------- ------ 0$^+$ 0.0 0.0 0$^+$ 1.7 (2.1) 4$^-$ 4.19 4.38 2$^+$ 4.86 5.3 2$^+$ 3.0 3.3 3$^-$ 4.40 4.26 4$^+$ 7.92 4$^+$ 4.7 5$^-$ 4.53 4.97 : $^{34}$Si level scheme. N is for normal and I for intruder. Finally, in table 4. we move outside of the “island of inversion”. In its “normal” ground state $^{34}$Si is closer to doubly-magic than to semi-magic. The different intruders fit very nicely with the experimental data. Notice the not very frequent level scheme, with a 0$^+$ as first excited state. It stems from the calculation that this state is oblate with $\beta \sim$ 0.4. The band built upon it is not very regular, but the first couple of transitions are in accord with the rotor picture and are consistent with the value of the 2$^+$ spectroscopic moment. N=28: Coexistence ================= If we move towards the neutron drip line, the Mg isotopes continue being deformed; even the N=28, $^{40}$Mg, which sits at the drip line. Another candidate to shell closure disappearance! Experimentally N=28 has been reached for Z as low as 14 and there is a lot of debate on the physical interpretation of the data. In ref. [@reta] we studied the region. When the reduction of the $1f_{7/2}-2p_{3/2}$ splitting in $^{35}$Si is plugged in the calculations, its main effect is to erode the N=28 shell closure, bringing the 2$^+$ excitation energy and the BE2(0$^+$ $\rightarrow$ 2$^+$) in $^{44}$S and $^{46}$Ar into full agreement with the data [@glas]. In $^{42}$Si the excitation energy of the first 2$^+$ drops by $\sim$ 1 MeV. We can follow the behaviour of the sulphur isotopes in the same valence space crossing two magic numbers and fifteen units of mass. $^{36}$S is spherical, its N=20 neutron closure is reinforced by the proton 2s$_{1/2}$ closure, resulting in a 2$^+$ at 3.5 MeV. Adding neutrons, the 2s$_{1/2}$ and 1d$_{3/2}$ become degenerate and, at N=26, $^{42}$S is a nearly perfect prolate rotor. At N=28 the spherical and the deformed solutions appear at the same energy and mix at 50%. $^{43}$S could provide a nice example of shape coexistence and isomerism. Our calculation produces a 3/2$^-$ deformed ground state, a low lying 7/2$^-$ spherical isomer and another 7/2$^-$ at higher energy, belonging to the ground state band. All these in full correspondence with the experimental results of ref. [@sara]. One could even attempt N=40, but realism advises to increase Z a few units. We suggested a few years ago that $^{64}$Cr could be another “semi-magic” prolate rotor, because of the strong quadrupole coherence of the intruder configurations with four $pf$ protons and four $gds$ neutrons. Preliminary evidence of such behaviour has been reported in its neighbour $^{66}$Fe [@hanna]. For the moment no signs of weakening of N=50 in $^{78}$Ni have been found [@daug]. [50]{} M Dufour and A. P. Zuker, Phys. Rev. C[**54**]{} (1996) 1641. B. H. Wildenthal, M. S. Curtin and B. A. Brown, Phys. Rev. C[**28**]{} (1983) 1343. B.H. Wildenthal, [*Prog. Part. Nucl. Phys.*]{} [**11**]{} (1984) 5. B. H. Wildenthal and W. Chung, Phys. Rev. C[**22**]{} (1980) 2260. B. A. Brown, Proc. ENAM95 p. 451, M. de Saint Simon and O. Sorlin eds. Ed. Frontieres (France) 1995; T. Aumann [*et al.*]{}, Phys. Rev. Lett. [**84**]{}, (2000) 35. A. P. Zuker, “Contemporary Nuclear Shell Models” Lecture Notes in Physics 482, p. 93, Springer 1997. T. Suzuki and T. Otsuka, Phys. Rev. C[**50**]{}, (1994) R555. M. J. G. Borge, [*et al.*]{}, Phys. Rev. C[**55**]{}, (1997) R8. I. Tanihata, [*et al.*]{}, Phys. Rev. Lett. [**55**]{}, (1985) 380; P. G. Hansen and B. Jonson, Europhys. Lett. [**4**]{} (1987) 409. H. Simon, [*et al.*]{}, Phys. Rev. Lett. [**83**]{}, (1999) 496. A. Navin, [*et al.*]{}, Phys. Rev. Lett. [**85**]{}, (2000) 266. C. Thibault, [*et al.*]{}, Phys. Rev. C[**12**]{} (1975) 193. X. Campi, H. Flocard, A. K. Kerman, S. Koonin, Nucl. Phys. A[**251**]{} (1975) 193. C. Detraz, [*et al.*]{}, Phys. Rev. C[**19**]{} (1978) 171. A. Poves and J. Retamosa, Phys. Lett B[**184**]{} (1987) 311. A. Poves and J. Retamosa, Nucl. Phys A[**571**]{} (1994) 221. E. K. Warburton, J. A. Becker and B. A. Brown, Phys. Rev. C[**41**]{} (1990) 1147. N. Fukunishi, T. Otsuka, and T. Sebe, Phys. Lett. [**B296**]{} (1992) 279. T. Otsuka and N. Fukunishi, Phys. Rep. [ **264**]{} (1996) 297. T. Siiskonen, P. O. Lippas and J. Rikovska, Phys. Rev. C[**60**]{} (1999) 034312. E. Caurier, F. Nowacki, A. Poves, J. Retamosa, Phys. Rev. C[**58**]{} (1998) 2033. Y. Utsuno, T. Otsuka, T. Mizusaki, M. Honma, Phys. Rev. C[**60**]{} (1999) 054315. S. Nummela, [*et al.*]{}, “Experimental Nuclear Physics in Europe” AIP Conf. Proc. [**495**]{} (1999) 55, B. Rubio, M. Lozano and W. Gelletly eds. H. Sakurai, [*et al.*]{}, Phys. Lett. B[**448**]{} (1999) 180. B. V. Pritychenko, [*et al.*]{}, Phys. Lett. B [ **461**]{} (1999) 322. D. Guillemaud, [*et al.*]{}, Nucl Phys A[**246**]{} (1984) 37. T. Motobayashi, [*et al.*]{}, Phys. Lett. B[**346**]{} (1995) 9. F. Azaiez [*et al.*]{}, “Experimental Nuclear Physics in Europe” AIP Conf. Proc. [**495**]{} (1999) 171, B. Rubio, M. Lozano and W. Gelletly eds. K. Yoneda, [*et al.*]{}, Proc. Int. Conf. RIB2000, Divonne (France), in press. G. Klotz [*et al.*]{}, Phys. Rev. C[**47**]{} (1993) 2502. G. Walter, [*et al.*]{}, Proc. Int. Conf. RIB2000, Divonne (France), in press. B. V. Pritychenko, [*et al.*]{}, MSU-NSCL preprint-1156, june 2000. J. Retamosa, E. Caurier, F. Nowacki, A. Poves, Phys. Rev. C[**55**]{} (1997) 1266. T. Glasmacher, Annu. Rev. Nuc. Part. Sci. [**48**]{} (1998) 1. F. Sarazin, [*et al.*]{}, Phys. Rev. Lett. [**84**]{} (2000) 5062; R. W. Ibbotson, [*et al.*]{}, Phys. Rev. C[**59**]{} (1999) 642. M. Hannawald, [*et al.*]{}, Phys. Rev. Lett. [**82**]{} (1999) 1391. J. M. Daugas, [*et al.*]{}, Phys. Lett. B[**476**]{} (2000) 213.
{ "pile_set_name": "ArXiv" }
-0.0in-0.01in 15.5cm **Effects of density-dependent quark mass on phase** **diagram of three-flavor quark matter** Xiao-Bing Zhang and Xue-Qian Li Department of Physics, Nankai University, Tianjin 300071, China [Considering the density dependence of quark mass, we investigate the phase transition between the ( unpaired ) strange quark matter and the color-flavor-locked matter, which are supposed to be two candidates for the ground state of strongly interacting matter. We find that if the current mass of strange quark $m_s$ is small, the strange quark matter remains stable unless the baryon density is very high. If $m_s$ is large, the phase transition from the strange quark matter to the color-flavor-locked matter in particular to its gapless phase is found to be different from the results predicted by previous works. A complicated phase diagram of three-flavor quark matter is presented, in which the color-flavor-locked phase region is suppressed for moderate densities.]{} [**I. INTRODUCTION**]{} The quark matter with three flavors ( $u$, $d$ and $s$ ) has been intensively studied for two decades. When the down quark chemical potential is larger than the strange quark mass, the strange quark matter ( SQM ) might be energetically favored with respect to two-flavor quark matter and even nuclear matter so that it should be the ground state of strongly interacting matter [@witt]. Within the framework of the bag model, Farhi and Jaffe pointed out that SQM with the strange quark mass $m_s <140$MeV ( and with appropriate bag constant ) becomes the stable ground state for low baryon densities [@jaff]. Based on this consideration, it is further speculated that some compact stars are made up not of neutrons but SQM, which are termed as strange quark stars [@ss]. On the other hand, the study of dense quark matter draws much attention due to the recent progress in understanding of color superconductivity. At high densities, the original color and flavor symmetries of three-flavor QCD, namely $SU(3)_{color}\times SU(3)_{L} \times SU(3)_{R}$, is suggested to be broken down to a diagonal subgroup $SU(3)_{color+L+R}$ via the Bardeen-Cooper-Schrieffer ( BCS ) pairing [@alf]. The three-flavor quark matter with the particular symmetry is called the color-flavor locked ( CFL ) matter and it is different from SQM which is the matter without the BCS pairing. As another state of strongly interacting matter, the CFL matter is widely believed to become “ absolutely ” stable for sufficiently high densities [@alf03]. Thus there are two candidates for the ground state of three-flavor quark matter, CFL and SQM, which are stable for high and low densities respectively. The question is, in the moderate density region, which one of them is the ground state. In the other words, one concerns how SQM undergoes a phase transition to CFL with increase of density. Investigation on these issues is important for exploring the physics of strange quark stars and/or the interior structure of compact stars. Ignoring the $u$ and $d$ quark masses, the CFL free energy takes the form $$\begin{aligned} \Omega_{CFL}= - \frac{3\mu^4}{4\pi^2}+ \frac{3m_s^2 \mu^2}{4\pi^2} -\frac{3\Delta^2 \mu^2}{\pi^2} ,\label{pcfl0}\end{aligned}$$ to order of $m_s^2/\mu$, where $\mu$ is the quark chemical potential and $\Delta$ denotes the color superconducting gap. Using Eq.(\[pcfl0\]), Alford *et. al.* concluded that the CFL matter is more stable than the unpaired quark matter ( exactly, SQM ) as long as [@alfd] $${\mu}\geq \frac{m_s^2}{4\Delta}. \label{cflbreak}$$ As the necessary condition for CFL existence [@alfj02], Eq.(\[cflbreak\]) is valid only for high densities. This inequality can not fully answer the question raised above because it does not address the phase structure at the moderate densities. To illustrate this point more clearly, we draw the phase diagram based on Eq.(\[cflbreak\]) in the ( $m_s, \mu$ ) plane ( Fig. 1 ) . When the strange quark mass is small as $m_s <175$MeV, Fig. 1 shows that SQM is excluded completely from the moderate density region $\mu=0.3-1$GeV and the CFL matter including its gapless phase ( gCFL [@alf04] ) dominates all over. However, for small $m_s$, SQM has been predicted to be the stable ground state [@jaff] so that it should be favorable at least for low densities such as $\mu\sim 0.3$GeV. It implies that the relation determined by Eq.(\[cflbreak\]) is problematic more or less. If assuming that CFL emerges in strange quark stars, this contradiction becomes more obvious. Starting at very low density and increasing the matter density by increasing $\mu$, the CFL formation must be preceded by presence of the stable SQM state. From this point of view, SQM remains stable for relatively low densities; otherwise, the self-bound quark stars could not exist and then the CFL formation would be impossible. Therefore, the moderate-density phase diagram shown in Fig. 1 needs to be reexamined, especially in the quark-star environment. In fact, the implicit assumption for Fig. 1 is that only the current mass of strange quark $m_s$ was considered in the descriptions for both CFL and SQM. According to the low-density QCD, the strange quark mass not merely originates from the explicit breaking of chiral symmetry. For low densities where SQM exists as the stable ground state, there is no reason to neglect the dynamical mass induced by the spontaneous chiral breaking. Once the dynamical mass is taken into account, it is found that the SQM stability window, e. g. the allowed region of the current mass $m_s$, is widened [@lug; @pion]. This motivates us to reexamine both SQM and CFL while the dynamical mass is incorporated. In this work, we will introduce the density dependence of quark mass to investigate the phenomenological effects of the dynamical mass on the moderate-density phase diagram. This approach should be closer to reality and obviously helpful to clarify the problem of Fig. 1. In Sec.II, we briefly review the mass-density-dependent model [@lug] and consider the free energy of the CFL matter when the density-dependent quark mass is incorporated. We emphasize that the mechanism for existence of the gCFL phase needs to be reexamined in particular for relatively low densities. In Sec.III, we investigate the phase transitions from SQM to the conventional CFL phase and/or the gCFL phase and present a new phase diagram which is very different from Fig. 1. [**II. THE MODEL** ]{} [**A. SQM and its stability** ]{} Following Ref.[@lug], the density-dependent quark mass is given by $$\begin{aligned} {m}_{D}= {C}/(3\rho),\label{md}\end{aligned}$$ where $\rho$ denotes the matter density and $C$ is a model parameter which is constrained by the SQM stability conditions. If ignoring the current masses of $u$ and $d$ quarks, the masses for the light and strange quarks in this model are $$\begin{aligned} {M}_{u}={M}_{d}= {m}_{D};\;\; {M}_{s}= {m}_{s}+{m}_{D},\label{mq}\end{aligned}$$ respectively. The SQM free energy contributed by the Fermi gas reads [@lug] $$\begin{aligned} {\Omega}(\mu_i,M_i,p^i_F)&=& {{\sum}\atop{i=u,d,s}} \int_0^{p_F^i} \frac{3}{\pi^2} p^2 (\sqrt{p^2+M_i^2}-\mu_i )dp \nonumber \\ &=& -{{\sum}\atop{i=u,d,s}} \frac{1}{4\pi^2}[{\mu_i}{p_F^i}( {\mu_i^2}-{5\over2}{M_i^2})+ {3\over2}{M_i^4}\ln (\frac{\mu_i+p_F^i}{M_i})] .\label{psqm1}\end{aligned}$$ For each flavor the Fermi momentum ${p_F^i}$ is defined by $p_F^i=\sqrt{\mu_i^2-M_i^2}$ where $\mu_u=\mu-2\mu_e/3$ and $\mu_d=\mu_s=\mu+\mu_e/3$ if the electron chemical potential $\mu_e\neq 0$. On the SQM side, the Fermi momenta of $u$, $d$ and $s$ quarks are different and are related to the corresponding densities via $\rho_i={(p_F^i)}^3/{\pi^2}$. Therefore, for SQM, the electrical neutrality is realized by $$\begin{aligned} \frac{2}{3}\rho_u-\frac{1}{3}(\rho_d+\rho_s)=\rho_e=\frac{\mu_e^3}{3\pi^2}, \label{neu}\end{aligned}$$ and the baryon density is $$\begin{aligned} {\rho}&=&\frac{1}{3}(\rho_u+\rho_d+\rho_s).\label{nsqm}\end{aligned}$$ When the contribution from electrons is included, the total free energy for SQM becomes $$\begin{aligned} {\Omega}_{SQM}&=& {\Omega}(\mu_i,M_i,p_F^i) -\frac{\mu_e^4} {12\pi^2}.\label{psqm}\end{aligned}$$ With respect to nuclear matter, SQM becomes energetically stable for low densities as long as its energy per baryon satisfies $${({\cal E}/\rho)}_{SQM} \le 930\textrm{MeV}, \label{stb}$$ at zero pressure, where $930$MeV corresponds to a typical value of the energy per baryon in nuclei. For our purpose, Eq.(\[stb\]) needs to be considered seriously to guarantee that not nuclear matter but SQM undergoes a phase transition to CFL ( if without this constraint the nuclear-CFL transition [@alfd] would be very likely ). In this work, we fix the parameter $C$ by the critical condition of Eq.(\[stb\]) for certainty. For instance, the value of $C$ is adopted to be $110$ and $70$MeV/fm$^3$ as $m_s=0$ and $180$MeV respectively ( see Ref.[@lug] for details ). On the other hand, the energy per baryon for two-flavor quark matter ( 2QM ) is required to satisfy the inequality $${({\cal E}/\rho)}_{2QM} > 930\textrm{MeV} , \label{1'}$$ at zero pressure [@jaff]. By using Eqs.(\[stb\]) and (\[1’\]), the stability window can be obtained in which SQM corresponds to the stable ground state at low densities. But Eq.(\[1’\]) does not apply to the moderate-density case. Due to the appearance of strange flavor, SQM is favored over the regular two-flavor matter as long as $$\begin{aligned} \mu_d=\mu+\mu_e/3 \geq M_s. \label{2s}\end{aligned}$$ Instead of Eq.(\[1’\]), thus, Eq.(\[2s\]) needs to be taken into account in the following calculation. [**B. CFL and gCFL in the model**]{} Different from the unpaired one, the CFL quark matter is an insulator in which no electrons are required for the electrical neutrality [@neu]. The Fermi momenta have the common value [@alfd; @neu] $$\begin{aligned} \nu=2\mu-\sqrt{\mu^2+m_s^2/3},\label{pfcom}\end{aligned}$$ for all three flavors and $\mu_e$ does not influence the CFL free energy directly . Thus the CFL free energy contributed by the Fermi gas is obtained by replacing the variables $\mu_i$ and $p_F^i$ in Eq.(\[psqm1\]) by $\mu$ and $\nu$ respectively. Together with the contribution from the CFL pairing ( the third term in Eq.(\[pcfl0\]) ), the total free energy for CFL takes the form $$\begin{aligned} \Omega_{CFL}= {\Omega}(\mu,M_i,\nu)-\frac{3\Delta^2 \mu^2}{\pi^2} ,\label{pcfl}\end{aligned}$$ when the density dependence of quark mass is considered. At high density, the value of $m_D$ is close to zero so that the difference between Eqs.(\[pcfl0\]) and (\[pcfl\]) becomes negligible. For the concerned density region, Eq.(\[pcfl\]) means that not only $m_s$ but also $m_D$ contribute to the CFL free energy. As a consequence, the previous results of the SQM-CFL transition e.g. Fig. 1 might be no longer valid at low/moderate densities ( see Sec.III for details ). Now we turn to consider the gapless color-flavor-locked ( gCFL ) phase where the Cooper pairs between the blue-down ( $bd$ ) and green-strange ( $gs$ ) quasi quarks becomes unstable. At sufficiently high densities, the strange quark mass is just the current mass $m_s$ and the light quark masses are very small so that the mass matrix reduces to $diag(0,0,m_s)$ approximately. It is well known that the quark mass term can be simplified as $$\begin{aligned} -\frac{m_s^2}{2\mu}(\tilde{\psi_L^s}^+ \tilde{\psi_L^s} + L\rightarrow R), \label{lq}\end{aligned}$$ at the leading order of the CFL effective Lagrangian [@hong; @sch]. Here $\tilde{\psi}$ is not the ordinary quark field, but the quasi-quark field in the vicinity of the Fermi surface. For the $bd$-$gs$ Cooper pairs, $\tilde{\psi}_{L/R}^s$ in Eq.(\[lq\]) denotes the $gs$ modes near the Fermi surface. It is worth being notified that Eq.(\[lq\]) conserves chiral symmetry although the current mass term in the low-density QCD theory does not. As suggested in Ref.[@sch], $m_s^2/(2\mu)$ is regarded as the chemical potential associated with strangeness, i.e. $\mu_S=m_s^2/(2\mu)$. Obviously, the effective chemical potential for the $gs$ modes is influenced by the nonzero $\mu_S$ while that for the $bd$ modes is independent of $\mu_S$. As a result, the effective chemical potentials $\mu_{gs}^{eff}$ and $\mu_{bd}^{eff}$ become different. To account for the relative chemical potential of the paired $bd$ and $gs$ modes, the variation is [@alf04] $$\begin{aligned} \delta \mu=\frac{\mu_{bd}^{eff}-\mu_{gs}^{eff}}{2}= \frac{m_s^2}{2\mu}, \label{dmu0}\end{aligned}$$ where the contribution from the chemical potential associated with the color charge has been included. When the variation is larger than the color superconducting gap, i.e. $$\begin{aligned} \frac{m_s^2}{2\mu}\geq \Delta, \label{gcfl}\end{aligned}$$ gCFL emerges as the more stable phase than CFL. Based on Eq.(\[gcfl\]), gCFL was predicted to emerge at relatively large $m_s$ and/or relatively small $\mu$, as shown in Fig. 1. However, this is not the whole story yet. At densities where $m_D$ becomes nonzero, the variation $\delta\mu$ caused by the strange mass term mainly is influenced by $m_D$ also. This means that the gCFL existence at moderate densities must be reexamined once the density dependence of quark mass is considered. Noticing that $m_D$ is the dynamical mass essentially, it does not enter the CFL effective Lagrangian via a simple replacement $m_s \rightarrow M_s=m_s+m_D $ directly. Therefore, an extrapolation like $\delta \mu \rightarrow \delta \mu'=\frac{M_s^2}{2\mu}=\frac{(m_s+m_D)^2}{2\mu}$ is not feasible in principle. To incorporate the effect of $m_D$ self-consistently, let us consider the dynamical quark mass term in the Lagrangian involving the quasi-quark degrees of freedom. Note that the chiral symmetry pattern exhibits a resemblance to the color-flavor-locking pattern. It is reasonable to assume that $m_D$ enters the CFL effective Lagrangian via the chiral-invariant form $$\begin{aligned} \xi m_D(\tilde{\psi_L^s}^+ \tilde{\psi_L^s} + L\rightarrow R).\label{lq2}\end{aligned}$$ Without losing generality, we introduce an unknown coefficient $\xi$ in Eq.(\[lq2\]). If the density-dependent-mass $m_D$ does account for all the non-perturbative effects of the dynamical chiral breaking, $\xi$ is equal to $1$ which is adopted in the calculations of Sec.III. Eq.(\[lq2\]) is expected to deviate the strangeness chemical potential from the original value $\mu_S=m_s^2/(2\mu)$. Correspondingly, only the effective chemical potential $\mu_{gs}^{eff}$ is influenced by Eq.(\[lq2\]) for the $bd$-$gs$ Cooper pairs. In analogy with the treatment of yielding Eq.(\[dmu0\]), there exists a decrease in the value of $\delta \mu$ so that we redefine the variation as $$\begin{aligned} \delta \mu \rightarrow \delta \mu'=\frac{m_s^2}{2\mu}-\frac{\xi}{2} m_D.\label{dmu}\end{aligned}$$ By inserting Eq.(\[dmu\]) into the dispersion relation for the $bd$-$gs$ pairs [@alfkou], the gapless modes become possible at the momenta $$\begin{aligned} p^{\pm}_{gapless}= \overline{\mu}\pm \sqrt{(\delta \mu ')^2-\Delta^2},\end{aligned}$$ where $\overline{\mu}$ is the average value of the $bd$ and $gs$ effective chemical potentials. Compared with the case without $m_D$, the “ blocking ” region [@liu], i.e. the width between $p^{+}_{gapless}$ and $p^{-}_{gapless}$, is suppressed for not-very-high densities. As a consequence, the free energy contributed by the gapless phenomenon is expected to be influenced by both $m_s$ and $m_D$ since the relative free energy of gCFL and CFL depends on the magnitude of the “ blocking ” region mainly. Based on Eq.(\[dmu\]), the condition Eq.(\[gcfl\]) is replaced by $$\begin{aligned} \frac{m_s^2}{2\mu}-\frac{\xi}{2} m_D \geq \Delta. \label{gcfl2}\end{aligned}$$ If CFL emerges at the densities where $m_D$ becomes negligible, Eq.(\[gcfl2\]) reduces back to Eq.(\[gcfl\]) and the critical condition for the CFL-gCFL transition is still $m_s^2/(2\mu)=\Delta$ approximately. Also, the necessary condition Eq.(\[cflbreak\]) for the existence of the CFL matter needs to be reexamined in the case of $m_D\neq 0$. Once the variation between the $bd$ and $gs$ chemical potentials is too large, the color-flavor-locked pairing might be broken completely. According to Ref.[@liu], the CFL matter including gCFL exists only when the variation is not larger than $2\Delta$ which is the energy cost for breaking a Cooper pair. So the previous condition Eq.(\[cflbreak\]) is extended to $$\begin{aligned} \frac{m_s^2}{2\mu}-\frac{\xi}{2} m_D \leq 2\Delta.\label{cflbreak2}\end{aligned}$$ For the moderate density region, Eq.(\[cflbreak2\]) provides a more realistic boundary for the CFL matter. [**III. NUMERICAL RESULTS AND CONCLUSIONS** ]{} As a strong-coupling effect, the nonzero $m_D$ is expected to affect the phase diagram of three-flavor quark matter for not-very-high densities. Before going to be specific, let us firstly discuss *whether or not* the phase transition between SQM and CFL/gCFL occur in the moderate density region. The answer is not always positive and it is actually linked to the value of $m_s$, as argued in the following. Now both SQM and CFL/gCFL are the deconfined phases, therefore the physics of confinement does not play a role in determining the SQM-CFL/gCFL transitions. So the negative values of the free energies obtained in Sec. II are related to the corresponding pressure directly and then the Gibbs condition for the pressure equilibrium reads $$\begin{aligned} P_{CFL/gCFL}- P_{SQM}={\Omega}_{SQM}-{\Omega}_{CFL/gCFL}=0.\end{aligned}$$ For $m_s=10$, $50$, $100$ and $150$MeV, we show $\delta P=P_{CFL}-P_{SQM}$ as a function of $1/\mu$ in Fig. 2. It is found that $\delta P$ does no longer approach to zero monotonously with increasing $1/\mu$, i.e. decreasing $\mu$. As shown in Fig. 2, there exists a rising tendency of $\delta P$ in the vicinity of $\mu\simeq 0.3$GeV. This leads to the fact that no any pressure equilibrium appears in the moderate density region so that a first-order SQM-CFL phase transition does not occur. Although the CFL pressure is relatively large, the absence of the phase transition means that the CFL matter is impossible to exist at least in our concerned density region. Therefore, SQM with small $m_s$ still remains as a stable state for moderate densities while CFL with small $m_s$ is prohibited unless the density is very large . The above physical picture holds valid until $m_s$ is large. For larger $m_s$, more pressure is paid to maintain the common Fermi momentum so that the pressure of CFL decreases. As long as $m_s$ is large enough, the SQM-CFL transition in moderate density region becomes possible to occur. Our numerical calculation shows that, as $m_s$ is about $150$MeV, the pressure equilibrium comes to appear in the vicinity of $\mu\simeq 0.3$GeV ( see Fig. 2 also ). Noticing that such pressure equilibrium behaves like “ crossover ” of CFL and SQM, $m_s \simeq150$MeV is regarded as the minimal value allowed for the CFL existence at moderate density. Once $m_s$ is larger than $150$MeV, the transition from SQM to CFL/gCFL might occur in moderate density region. As a typical example, the result of $\delta P=P_{CFL}-P_{SQM}$ for $m_s=200$MeV is given by the solid line in Fig. 3. As shown in Fig. 3, the critical chemical potential $\mu_c$ for the SQM-CFL transition is about $0.4$GeV. When the gCFL phase and the SQM-gCFL transition are incorporated, however, the SQM-CFL transition might not occur in nature since gCFL is more energetically favorable than CFL. The dashed line in Fig. 3 gives the value of $P_{gCFL}-P_{SQM}$ for $m_s=200$MeV. As shown in Fig. 3, the SQM-gCFL transition occurs at the critical chemical potential $\mu_c'\simeq 0.35$GeV, which is smaller than the value of $\mu_c$. Therefore, it is not CFL but gCFL to emerge firstly in the quark star environment. Also, the critical point for the gCFL-CFL transition $\mu_c''$ is shown to be about $0.67$GeV in Fig. 3. For $m_s=200$MeV, we conclude that, gCFL exists as a stable state in the region of $\mu_c'<\mu \leq \mu_c''$ and CFL emerges as $\mu>\mu_c''$. Here we would like to emphasize the importance of the nonzero $m_D$ for the properties of gCFL. In Eq.(\[dmu\]) the $m_s^2/(2\mu)$ and $m_D$ terms provide opposite contributions to the gCFL pressure actually. Within the framework where the density dependence of quark mass is considered, if the gapless phenomenon were determined by $m_s^2/(2\mu)$ exclusively the pressure of gCFL would be much larger than that of SQM with increasing $1/\mu$, as shown by the dotted line in Fig. 3. In that case, the SQM-gCFL pressure equilibrium and thus the corresponding transition no longer exists in the moderate density region. Therefore the effect of $m_D$ is relevant to the gCFL presence for moderate densities. Based on the above arguments, a schematic phase diagram is given for the moderate density region in Fig. 4. There are three kinds of different structures in the phase diagram according to the value of $m_s$ : \(i) As $m_s$ is small such as $m_s <150$MeV , the effect of $m_D$ prohibits the CFL formation for not-very-high densities. In this case, it is not CFL but SQM to be the stable phase in the whole moderate density region, as shown in Fig. 4. This conclusion is very different from that obtained by Fig. 1, but agrees with the original prediction that CFL with zero ( or small ) $m_s$ becomes possible only when the density is high enough [@alf]. When $m_s$ is slightly larger than $150$MeV, a first-order transition from SQM to CFL takes place. For instance, let us consider the horizontal line of $m_s=175$MeV in Fig. 4. At the critical chemical potential $\mu_c \simeq 0.35$GeV, the SQM-CFL transition occurs so that SQM remains stable for $\mu<\mu_c$ whereas CFL emerges for $\mu\ge\mu_c$. \(ii) As $m_s$ is large, the gCFL phase becomes more likely than the conventional CFL phase and then the SQM-gCFL transition replaces the SQM-CFL transition to be relevant to the phase diagram. Considering the horizontal line of $m_s=200$MeV, gCFL comes to emerge at $\mu_c'\simeq 0.35$GeV and it undergoes a continuous transition to CFL at $\mu_c''\simeq 0.67$GeV. Interestingly, the SQM-gCFL transition curve in the low density region is qualitatively different from the result of Fig. 1 : the critical value of $m_s$ does increase with decreasing $\mu$, as shown by the solid line in Fig. 4. The reason is that the gapless phase is determined by Eq.(\[dmu\]), in which the $m_s^2/(2\mu)$ and $m_D$ terms offer opposite effects on the gCFL pressure. For lower $\mu$ the latter effect is more important, so that the transition becomes possible only when $m_s$ is relatively large. Also, the continuous gCFL-CFL transition curve is given by Eq.(\[gcfl2\]) as shown by the dashed line in Fig. 4. Together with the SQM-CFL and SQM-gCFL transition curves, there exists an intersection of the three phase transitions in the vicinity of ( $m_s$,$\mu$ )= ( 185MeV,0.4GeV ). \(iii) As $m_s$ is larger, the variation $\delta \mu'$ might be too large to allow existence of the color-flavor-locked pairing in quark matter. Most importantly, the necessary condition Eq.(\[cflbreak2\]) for existence of the CFL matter no longer coincides with the SQM-gCFL transition curve, although Eq.(\[cflbreak\]) does in the case without $m_D$. Thus, the gCFL phase region is surrounded by the CFL boundary curve ( that obtained from Eq.(\[cflbreak2\]) ), the SQM-gCFL transition curve ( that determined by the pressure equilibrium ) and the gCFL-CFL transition curve ( that obtained from Eq.(\[gcfl2\]) ), as shown in Fig. 4. Comparing with Fig. 1, we find that the gCFL phase region is suppressed for low $\mu$. On the other hand, the existence of three-flavor quark matter might be prohibited if $m_s$ is very large. The boundary curve of SQM is given by Eq.(\[2s\]) and is shown in Fig. 4 , which seems to imply that 2QM is irrelevant to the gCFL presence. At this point, it must be stressed that a possibility of the 2QM-gCFL transition could not be ruled out simply if the color superconductivity exists in the two-flavor quark matter. In that case, the boundary of SQM shown in Fig. 4 needs to be modified also. Details involving such 2QM and its transition to gCFL and/or SQM depend on what kind of two-flavor color superconducting phase be taken into account, which is beyond the scope of the present work. In summary, we extend the descriptions of CFL and gCFL from high-density case to the moderate density region when the density dependence of quark mass is considered. Starting at low density and raising the matter density, the physical picture that SQM remains as the stable state at first and then undergoes a first-order phase transition to gCFL and/or CFL is examined in details. As a result, we predict a more complicated phase diagram of three-flavor quark matter, in which the CFL/gCFL phase region is suppressed for low densities. The present phase diagram is helpful to better understand the ground state of three-flavor quark matter in the environment of quark stars. Of course, there are some uncertainties of the color superconducting gap used in this work. When the value of the gap is chosen in other ways, we can give the similar phase diagram as Fig. 4. For instance, if the gap is large such as $\Delta\sim 80$MeV [@kap] we find that the minimum value of $m_s$ allowed for the CFL existence increases so that the CFL phase region for moderate densities might be further suppressed. Even if the density dependence of the gap is included, the change of $\Delta$ in the finite region of $\mu=0.3-1$GeV is not too drastic and the conclusion obtained from Fig. 4 is still qualitatively correct. In the further work one should construct the dynamical quark mass within a more realistic framework such as that beyond the bag model as well as take the contributions from the color-sextet pairing and the gap equation into account. Some of the problems are being investigated. [**Acknowledgements**]{} This work was supported by National Natural Science Foundation of China ( NSFC ) under Contract No.10405012. [99]{} E. Witten, Phys. Rev. D [**30**]{}, 272 (1984); A. Bodmer, Phys. Rev. D [**4**]{}, 1601 (1971); S. Chin and A. K. Kerman, Phys. Rev. Lett. [**43**]{}, 1292 (1979). E. Farhi and R. L. Jaffe, Phys. Rev. D [**30**]{}, 2379 (1984). For reviews, see e. g., C. Alcock and A. Olito, Annu. Rev. Nucl. Part. Sci. [**38**]{}, 161 (1988). M. Alford, K. Rajagopal and F. Wilczek, Phys. Lett. [**B422**]{}, 247 (1998); M. Alford, K. Rajagopal and F. Wilczek, Nucl. Phys. [**B537**]{}, 443 (1999). For reviews, see e. g., M. Alford, Prog. Thero. Phys. Suppl. [**153**]{}, 1 (2004). M. Alford, K. Rajagopal, S. Reddy and F. Wilczek, Phys. Rev. D [**64**]{}, 074017 (2001). M. Alford and K. Rajagopal, J. High Energy Phys. [**0206**]{}, 031 (2002); M. Alford, C. Kouvaris and K. Rajagopal, Phys. Rev. Lett. [**92**]{}, 222001 (2004). O. G. Benvenuto and G. Lugones, Phys. Rev. D [**51**]{}, 1989 (1995); see also, G. N. Fowler, S. Raha and R. W. Weiner, Z. Phys. C [**9**]{}, 271 (1981). V. Soni and D. Bhattacharya, Phys. Rev. D [**69**]{}, 074001 (2004). K. Rajagopal and F. Wilczek, Phys. Rev. Lett [**86**]{}, 3492 (2001). X. B. Zhang and X. Q. Li, Phys. Rev. D [**70**]{}, 054010 (2004). See e. g. D. K. Hong, Phys. Lett. B [**473**]{}, 118 (2000); S. R. Beane, P. F. Bedaque and M. J. Savage, Phys. Lett. B [**483**]{}, 131 (2000); D. T. Son and M. Stephanov, Phys. Rev. D [**61**]{}, 074012 (2000). T. Schäfer, Nucl. Phys. [**A575**]{}, 269 (2000); P. F. Bedaque and T. Schäfer, Nucl. Phys. [**A697**]{}, 802 (2002). M. Alford, C. Kouvaris and K. Rajagopal, Phys. Rev. D [**71**]{}, 054009 (2005). E. Gubankova, W. V. Liu and F. Wilczek, Phys. Rev. Lett. [**91**]{}, 032001 (2003) D. B. Kaplan and S. Reddy, Phys. Rev. D [**65**]{}, 054042 (2002).
{ "pile_set_name": "ArXiv" }
--- abstract: 'Using a variational method, we prove a conjecture of Carlen, Frank and Lieb, which concerns the joint convexity of the the trace function $$\Psi_{p,q,s}(A,B)={\textnormal{Tr}}(B^{\frac{q}{2}}K^*A^{p}KB^{\frac{q}{2}})^s,$$ where $-1\leq q< 0,~1\leq p\leq 2,~(p,q)\ne (1,-1),~s\geq\frac{1}{p+q}$, $A$ and $B$ are $N\times N$ positive semi-definite matrices and $K$ is a fixed $N\times N$ matrix. This admits the Audenaert-Datta conjecture with $s=\frac{1}{p+q}$ as a special case. Together with other known results, we will give full range of $(p,q,s)$ for $\Psi_{p,q,s}$ to be joint convex/concave. As a consequence, we obtain the full range of $(\alpha,z)$ for $\alpha-z$ Rényi relative entropies to be monotone under the completely positive trace preserving maps. We will also use this method to give simple proofs for some known results on joint convexity/concavity of $\Psi_{p,q,s}$.' address: 'Laboratoire de Mathématiques, Université de Bourgogne Franche-Comté, 25030 Besançon, France and Institute of Mathematics, Polish Academy of Sciences, ul. Śniadeckich 8, 00-956 Warszawa, Poland' author: - Haonan Zhang title: 'Carlen-Frank-Lieb conjecture and monotonicity of $\alpha-z$ Rényi relative entropy' --- Introduction {#sect:intro} ============ Given two probability density functions $\rho$ and $\sigma$ on $\mathbb{R}$, the *relative entropy*, or *Kullbach-Liebler divergence* of $\rho$ with respect to $\sigma$ is given by $$\label{classical relative entropy} D(\rho||\sigma):=\int_{\mathbb{R}}\rho(x)(\log\rho(x)-\log\sigma(x)){\text{d}x}.$$ For $\alpha\in(0,1)\cup(1,\infty)$, the *$\alpha$-Rényi relative entropy* of $\rho$ with respect to $\sigma$ [@Renyi61] is defined as $$\label{classcial alpha relative entropy} D_{\alpha}(\rho||\sigma):=\frac{1}{\alpha-1}\log\int_{\mathbb{R}}\rho(x)^{\alpha}\sigma(x)^{1-\alpha}{\text{d}x}.$$ Both relative entropies have been generalized to quantum case, where density functions are replaced by density matrices, i.e., positive semi-definite matrices of trace 1, and integrals are replaced by traces. In the sequel we shall use ${\mathcal{D}}_N$ to denote the set of $N\times N$ density matrices, ${\mathcal{P}}_N$ to denote the set of $N\times N$ positive semi-definite matrices and ${\mathcal{M}}_N$ to denote the set of all $N\times N$ matrices. And ${\textnormal{Tr}}$ always denotes the usual trace on matrices. Now for $\rho,\sigma\in{\mathcal{D}}_N$, a natural quantum analog of , is the so-called *Umegaki relative entropy* [@Umegaki62] $$\label{Umegaki relative entropy} D(\rho||\sigma):={\textnormal{Tr}}\rho(\log\rho-\log\sigma).$$ If $\text{supp}(\rho)\not\subset\text{supp}(\sigma)$, then $D(\rho||\sigma)$ is understood as $\infty$, where $\text{supp}(x)$ denotes the support of $x$. All the relative entropies in the sequel will be understood in a similar way. The quantum analogs of might take various forms. One of the most important generalizations of is *quantum $\alpha$-Rényi relative entropy* $$\label{quantum alpha relative entropy} D_{\alpha}(\rho||\sigma):=\frac{1}{\alpha-1}\log{\textnormal{Tr}}(\rho^{\alpha}\sigma^{1-\alpha}),~~ \alpha\in(0,1)\cup(1,\infty).$$ It admits Umegaki relative entropy $D(\rho||\sigma)$ as a limit case when $\alpha\to 1$. Remark that throughout this paper, whenever we write $X^\beta$ with negative $\beta$, automatically we assume $X$ is invertible (or one can consider $X^\beta$ in the sense of generalized inverse, i.e., $X^\beta=(X|_{\text{supp}X})^\beta$, which won’t bring any trouble for our results). Another generalization of , introduced by Müller-Lennert, Dupuis, Szehr, Fehr, Tomamichel [@MDSFT13-sandwich] and Wilde, Winter, Yang [@WWY14-sandwich], is known as *sandwiched $\alpha$-Rényi entropy*: $$\label{sanwished alpha relative entropy} \widetilde{D}_{\alpha}(\rho||\sigma):=\frac{1}{\alpha-1}\log{\textnormal{Tr}}(\sigma^{\frac{1-\alpha}{2\alpha}}\rho\sigma^{\frac{1-\alpha}{2\alpha}})^{\alpha},~~\alpha\in(0,1)\cup(1,\infty).$$ In recent years, Audenaert and Datta [@AD-alpha-z] introduced a new family of quantum Rényi relative entropies, which unify $\alpha$-Rényi relative entropy $D_{\alpha}$ and sandwiched $\alpha$-Rényi relative entropy $\widetilde{D}_{\alpha}$ together by using two parameters, called *$\alpha-z$ Rényi relative entropies*: $$\label{alpha-z entropy} D_{\alpha,z}(\rho||\sigma):=\frac{1}{\alpha-1}\log{\textnormal{Tr}}(\sigma^{\frac{1-\alpha}{2z}}\rho^{\frac{\alpha}{z}}\sigma^{\frac{1-\alpha}{2z}})^{z},~~\alpha\in(-\infty,1)\cup(1,\infty),~~z>0.$$ Note that by taking $z=1$ and $\alpha=z$, one recovers $D_{\alpha}$ and $\widetilde{D}_{\alpha}$, respectively. We comment here that the $\alpha-z$ Rényi relative entropies have appeared earlier in a paper by Jaksic, Ogata, Pautrat and Pillet [@JOPP12]. We recommend here a very nice paper [@CFL18-conjecture] by Carlen, Frank and Lieb, which tells the story of relative entropies more seamlessly and enables new comers to this field, including the author, to understand the background and get access to recent advances more easily and quickly. Most of notions in this paper come from [@CFL18-conjecture] and one can find all needed information there and the references therein. Now we come back to $\alpha-z$ Rényi relative entropies $D_{\alpha,z}$. It has operational meaning only if it is monotone under the completely positive trace preserving (CPTP) maps. That is, $$\label{DPI} D_{\alpha,z}(\mathcal{E}(\rho)||\mathcal{E}(\sigma))\leq D_{\alpha,z}(\rho||\sigma),$$ for all CPTP $\mathcal{E}$ and $\rho,\sigma\in{\mathcal{D}}_N$. This inequality is known as *Data Processing Inequality* (DPI). Although DPI is throughly studied for $D,~D_{\alpha}$ and $\widetilde{D}_{\alpha}$, it remains open for $D_{\alpha,z}$ to satisfy DPI for some range of $(\alpha,z)$. It is a standard argument that DPI is essentially equivalent to the joint convexity/concavity of the trace functions in the definition of $D_{\alpha,z}$. We will go back on this in the end of this section. Recall here that $f:{\mathcal{M}}_N\times{\mathcal{M}}_N\to\mathbb{R}$ is said to be *jointly convex* if $$f(\lambda A_1+(1-\lambda)A_2,\lambda B_1+(1-\lambda)B_2)\leq \lambda f(A_1,B_1)+(1-\lambda)f(A_2,B_2),$$ for all $\lambda\in(0,1)$ and $A_i,B_i\in{\mathcal{M}}_N,~i=1,2$. The notion of joint concavity follows a similar manner. Then from some known results on joint convexity and joint concavity of certain trace functions, Audenaert and Datta obtained DPI for $D_{\alpha,z}$ for some-but not full-range of $(\alpha,z)$ [@AD-alpha-z Theorem 1]. By saying full we mean under some certain necessary conditions (which we shall see in Proposition \[ness for psi\]) of $(\alpha,z)$. It is then natural to ask whether DPI holds for the remaining range of $(\alpha,z)$. This motived Audenaert and Datta to raise the following conjecture: \[conj of Audenaert-Datta\][@AD-alpha-z Conjecure 1] If $1\leq p\leq 2,~-1\leq q<0$ and $(p,q)\ne(1,-1)$, then for any $K\in {\mathcal{M}}_N$ and any $N$ $${\mathcal{P}}_N\times{\mathcal{P}}_N\ni(A,B)\mapsto {\textnormal{Tr}}(B^{\frac{q}{2}}K^*A^{p}KB^{\frac{q}{2}})^{\frac{1}{p+q}}$$ is jointly convex. We cheat a little bit here, since their original form of conjecture concerns the convexity of ${\mathcal{P}}_N\ni A\mapsto {\textnormal{Tr}}(A^{\frac{q}{2}}K^*A^{p}KA^{\frac{q}{2}})^{\frac{1}{p+q}}$. However, by doubling dimension, a standard argument shows that they are equivalent. See the discussions after [@CFL18-conjecture Conjecture 1] for example. More generally, consider the joint convexity/concavity of trace functions $$\Psi_{p,q,s}(A,B)={\textnormal{Tr}}(B^{\frac{q}{2}}K^*A^{p}KB^{\frac{q}{2}})^s,$$ for $A,B\in {\mathcal{P}}_N$, $K\in M_N$ and $p,q,s\in\mathbb{R}$. Note that $\Psi_{q,p,s}(B,A)=\Psi_{p,q,s}(A,B)$ with $K$ replaced by $K^*$. By an approximation argument we may assume that $K$ is invertible. Then $\Psi_{-p,-q,-s}(A,B)=\Psi_{p,q,s}(A,B)$ with $K$ replaced by $(K^{-1})^*$. So in the sequel, we always assume that $p\geq q$ and $s>0$. The current knowledge of joint convexity/concavity of $\Psi_{p,q,s}$ is summarized in the following proposition: \[known results on psi\] Fix $K\in M_N$. 1. If $0\leq q\leq p\leq1$ and $0<s\leq\frac{1}{p+q}$, then $\Psi_{p,q,s}$ is jointly concave. 2. If $-1\leq q\leq p\leq 0$ and $s>0$, then $\Psi_{p,q,s}$ is jointly convex. 3. If $-1\leq q\leq0,~1\leq p<2,~(p,q)\ne(-1,1)$ and $s\geq\min\{\frac{1}{p-1},\frac{1}{q+1}\}$, then $\Psi_{p,q,s}$ is jointly convex. If $p=2,~-1\leq q\leq0$ and $s\geq\frac{1}{q+1}$, then $\Psi_{p,q,s}$ is jointly convex. The proofs of (1)(2) for full range are due to Hiai [@Hiai16-concavity-II Theorem 2.1]. The proofs of (3) are due to Frank and Lieb [@FL13-DPI-sandwich Proposition 3], and Carlen, Frank and Lieb [@CFL16-some]. For more details of history on these results, see the discussions after [@CFL18-conjecture Theorem 2]. We only comment here that the case for $s=1$, which was firstly studied in the history, are due to Lieb [@Lieb73-WYDconjecture] for $0\leq q\leq p\leq1$ with $p+q\leq 1$, as well as for $-1\leq q\leq0$, and due to Ando [@Ando79] for $-1\leq q\leq0,~1\leq p<2$, with $p+q\geq1$. Their work played an important role in the development of matrix analysis. The following proposition gives necessary conditions for $\Psi_{p,q,s}$ to be jointly convex or joint concave. \[ness for psi\] Let $p\geq q$ and $s>0$. Suppose that $(p,q)\ne(0,0)$. 1. If ${\mathcal{P}}_2\times {\mathcal{P}}_2\ni(A,B)\mapsto\Psi_{p,q,s}$ is jointly concave for $K=I$, then $0\leq q\leq p\leq1$ and $0<s\leq\frac{1}{p+q}$. 2. If ${\mathcal{P}}_4\times {\mathcal{P}}_4\ni(A,B)\mapsto\Psi_{p,q,s}$ is jointly convex for $K=I$, then either $-1\leq q\leq p\leq 0$ and $0<s\leq\frac{1}{p+q}$ or $-1\leq q\leq0,~1\leq p\leq 2,~ (p,q)\ne(-1,1)$ and $s\geq\frac{1}{p+q}$. From the above two propositions, Carlen, Frank and Lieb conjectured [@CFL18-conjecture] that: \[conj of CFL\][@CFL18-conjecture Conjecture 4] If $1\leq p\leq 2,~-1\leq q<0,~(p,q)\ne(1,-1)$ and $s\geq\frac{1}{p+q}$, then for any $K\in {\mathcal{M}}_N$ and any $N$ $${\mathcal{P}}_N\times{\mathcal{P}}_N\ni(A,B)\mapsto {\textnormal{Tr}}(B^{\frac{q}{2}}K^*A^{p}KB^{\frac{q}{2}})^s$$ is jointly convex. Partial results of Conjecture \[conj of CFL\] are known, as pointed out in Proposition \[known results on psi\] (3). The main result of this paper is to prove Conjecture \[conj of CFL\], which, together with Proposition \[known results on psi\] (1)(2) and Proposition \[ness for psi\], will give the full range of $(p,q,s)$ for $\Psi_{p,q,s}$ to be jointly convex or jointly concave: \[main thm\] Fix $K\in M_N$. Suppose that $p\geq q$ and $s>0$. 1. If $0\leq q\leq p\leq1$ and $0<s\leq\frac{1}{p+q}$, then $\Psi_{p,q,s}$ is jointly concave. 2. If $-1\leq q\leq p\leq 0$ and $s>0$, then $\Psi_{p,q,s}$ is jointly convex. 3. If $-1\leq q\leq0,~1\leq p\leq 2,~(p,q)\ne(-1,1)$ and $s\geq\frac{1}{p+q}$, then $\Psi_{p,q,s}$ is jointly convex. The following figure summarizes the joint convexity/concavity of $\Psi_{p,q,s}$ for all $p,q,s$. Note that $(1,-1)$ and $(-1,1)$ don’t belong to the area of convexity. (-4,0)–(7,0) node\[right\][p]{}; (0,-4)–(0,7) node\[above\][q]{}; \(A) at (0,0); (B) at (0,3); (C) at (3,3); (D) at (3,0); (A) – (B) – (C) – (D) – cycle; (A)–(B)–(C)–(D)–cycle; (1.5,50pt) node\[\] [**concave**]{}; (1.5,35pt) node\[\] [for $0\le s \le \frac{1}{p+q} $]{}; \(E) at (6,0); (F) at (6,-3); (G) at (3,-3); (D) – (E) – (F) – (G) – cycle; (D)–(E)–(F)–(G)–cycle; (4.5,-35pt) node\[\] [**convex**]{}; (4.5,-50pt) node\[\] [for $s \ge \frac{1}{p+q} $]{}; \(H) at (0,-3); (I) at (-3,-3); (J) at (-3,0); (A) – (H) – (I) – (J) – cycle; (A)–(H)–(I)–(J)–cycle; (-1.5,-35pt) node\[\] [**convex**]{}; (-1.5,-50pt) node\[\] [for $s \geq 0 $]{}; \(K) at (-3,3); (L) at (-3,6); (M) at (0,6); (B) – (K) – (L) – (M) – cycle; (B)–(K)–(L)–(M)–cycle; (-1.5,135pt) node\[\] [**convex**]{}; (-1.5,120pt) node\[\] [for $s \ge \frac{1}{p+q} $]{}; (-3,0) node\[above\] [-1]{}; (0,-3) node\[right\] [-1]{}; (3.2,0) node\[above\] [1]{}; (6,0) node\[above\] [2]{}; (0,3.2) node\[right\] [1]{}; (0,6) node\[right\] [2]{}; (-2.5,-2.5)–(-3.5,-3.5); (-0.5,-0.5)–(0.5,0.5); (2.5,2.5)–(6.5,6.5) node\[above\] [q=p]{}; We shall prove (3), which is nothing but the Conjecture \[conj of CFL\], via a refinement of a variational method that originates in [@CL08-minkowski-II]. We will also give simple proofs for (1) with $q\ne0$ and (2) with $p\ne0$ using the same method. In other words we will prove joint convexity/concavity of $\Psi_{p,q,s}$ for genuine two variables case. The one variable case, Theorem \[thm:one variable case\], as a known result, is our building block. Since Carlen-Frank-Lieb Conjecture \[conj of CFL\] includes Audenaert-Datta Conjecture \[conj of Audenaert-Datta\] as a special case with $s=\frac{1}{p+q}$, we will also obtain the full range of $(\alpha,z)$ for $D_{\alpha,z}$ to satisfy DPI. Namely, as a corollary of Proposition \[ness for psi\] and Theorem \[main thm\] we have The $\alpha-z$ relative Rényi entropy $D_{\alpha,z}$ is monotone under completely positive trace preserving maps on ${\mathcal{P}}_N$ for all $N$ if and only if one of the following holds 1. $0<\alpha<1$ and $z\geq\max\{\alpha,1-\alpha\}$; 2. $1<\alpha\leq 2$ and $\frac{\alpha}{2}\leq z\leq\alpha$; 3. $2\leq\alpha<\infty$ and $\alpha-1\leq z\leq\alpha$. Now we close this section by explaining why monotonicity of $D_{\alpha,z}$ under CPTP maps is related to the joint convexity/concavity of trace functions $\Psi_{p,q,s}(A,B)$. The following proposition comes from [@CFL18-conjecture], which follows from a well-known argument of Lindblad [@Lindblad74] and Ulhmann [@Uhlmann73]. Let $\alpha,z>0$ and $\alpha\ne 1$. Set $p=\frac{\alpha}{z}$ and $q=\frac{1-\alpha}{z}$. Then $D_{\alpha,z}$ is monotone under completely positive trace preserving maps on ${\mathcal{P}}_N$ for all $N$ if and only if one of the following holds 1. $\alpha<1$ and $\Psi_{p,q,\frac{1}{p+q}}$ with $K=I$ is jointly concave; 2. $\alpha>1$ and $\Psi_{p,q,\frac{1}{p+q}}$ with $K=I$ is jointly convex. We use $\Psi$ to denote $\Psi_{p,q,\frac{1}{p+q}}$ with $K=I$. We only prove here that when $\alpha>1$, $D_{\alpha,z}$ is monotone under CPTP maps on ${\mathcal{P}}_N$ for all $N$ if and only if $\Psi$ is jointly convex, since the proof for $\alpha<1$ is similar. Note that when $\alpha>1$, $D_{\alpha,z}$ is monotone under CPTP maps if and only if $\Psi$ is monotone increasing under CPTP maps. To show the “if” part, for each CPTP map $\mathcal{E}$, we use Steinspring’s Theorem [@Stinespring55] to write $\mathcal{E}$ as $$\mathcal{E}(\gamma)={\textnormal{Tr}}_2U(\gamma\otimes \delta)U^*,$$ where $\delta\in{\mathcal{D}}_{N'}$, $U$ is unitary on $\mathbb{C}^N\otimes\mathbb{C}^{N'}$ and $N'$ is an integer not bigger than $N^2$. ${\textnormal{Tr}}_2$ denotes the usual partial trace over $\mathbb{C}^{N'}$. Let $du$ denote the normalized Haar measure on the group of all unitaries on $\mathbb{C}^{N'}$, then $$\label{equation:Schur's lemma} \mathcal{E}(\gamma)\otimes\frac{1_{\mathbb{C}^{N'}}}{N'}=\int(1\otimes u)U(\gamma\otimes\delta)U^*(1\otimes u^*)du.$$ By the tensor property of $\Psi$, we have $$\Psi(\mathcal{E}(\rho),\mathcal{E}(\sigma)) =\Psi\left(\mathcal{E}(\rho)\otimes\frac{1_{\mathbb{C}^{N'}}}{N'},\mathcal{E}(\sigma)\otimes\frac{1_{\mathbb{C}^{N'}}}{N'}\right).$$ From the joint convexity of $\Psi$ and it follows $$\Psi(\mathcal{E}(\rho),\mathcal{E}(\sigma)) \leq\int\Psi((1\otimes u)U(\rho\otimes\delta)U^*(1\otimes u^*),(1\otimes u)U(\sigma\otimes\delta)U^*(1\otimes u^*))du.$$ By the unitary invariance and tensor property of $\Psi$ we obtain $$\Psi(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq\Psi(\rho,\sigma),$$ as desired. To show the “only if" part, for any $\rho_1,\rho_2,\sigma_1,\sigma_2\in{\mathcal{P}}_N$ and any $0<\lambda<1$, define $$\rho= \begin{pmatrix*} \lambda\rho_1&0\\ 0&(1-\lambda)\rho_2 \end{pmatrix*} \text{ and } \sigma= \begin{pmatrix*} \lambda\sigma_1&0\\ 0&(1-\lambda)\sigma_2 \end{pmatrix*},$$ in ${\mathcal{P}}_{2N}$. Since the map $$\mathcal{E} \begin{pmatrix*} a&b\\ c&d \end{pmatrix*} =a+d,$$ is CPTP, we obtain from the monotonicy of $\Psi$ that $$\Psi(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq\Psi(\rho,\sigma),$$ which is nothing but $$\Psi(\lambda\rho_1+(1-\lambda)\rho_2,\lambda\sigma_1+(1-\lambda)\sigma_2)\leq\lambda\Psi(\rho_1,\sigma_1)+(1-\lambda)\Psi(\rho_2,\sigma_2).$$ This finishes the proof of joint convexity of $\Psi$. The proofs ========== This section is devoted to the proof of Theorem \[main thm\]. Consider a special case of $\Psi_{p,q,s}$ with $q=0$ $$\Upsilon_{p,s}(A):={\textnormal{Tr}}(K^*A^{p}K)^s,$$ for $A\in{\mathcal{P}}_N,~K\in M_N$ and $s>0$. We shall use the convexity/convavity of $\Upsilon_{p,s}$, which is throughly studied, as a building block to achieve joint convexity/convavity of $\Psi_{p,q,s}$. \[thm:one variable case\] Fix $K\in M_N$. 1. If $0< p\leq1$ and $0<s\leq\frac{1}{p}$, then $\Upsilon_{p,s}$ is concave. 2. If $-1\leq p\leq 0$ and $s>0$, then $\Upsilon_{p,s}$ is convex. 3. If $1\leq p\leq 2$ and $s\geq\frac{1}{p}$, then $\Upsilon_{p,s}$ is convex. The proofs of (1) and (2) are due to Hiai [@Hiai13-concavity-I Theorem 4.1]. The proof of (3) is due to Carlen and Lieb [@CL08-minkowski-II Theorem 1.1]. Again, see the discussions after Proposition 5 in [@CFL18-conjecture] for more historical information. We only comment here that the proof of concavity for $0< p\leq 1$ with $s=\frac{1}{p}$ is due to Epstein [@Epstein73]. His method uses complex analysis and is nowadays developed as an important tool to deal with joint convexity/concavity of trace functions, usually known as “analytic method", compared with “variational method". The next proposition, due to Hiai [@Hiai13-concavity-I Propositions 5.1(1) and 5.4(1)], completes Theorem \[thm:one variable case\] with necessary conditions. Let $s>0$ and $p\ne0$. 1. If ${\mathcal{P}}_2\ni A\mapsto \Upsilon_{p,s}(A)$ is concave for any invertible $K$, then $0<p\leq 1$; 2. f ${\mathcal{P}}_4\ni A\mapsto \Upsilon_{p,s}(A)$ is convex for any invertible K, then either $-1\leq p<0$ and $s>0$ or $1\leq p\leq 2$ and $s\geq\frac{1}{p}$. Now let’s state our varitional method, which admits an independent interest. For $r_i>0,i=0,1,2$ such that $\frac{1}{r_0}=\frac{1}{r_1}+\frac{1}{r_2}$, we have for any $X,Y\in{\mathcal{M}}_N$ that $$\label{varitional mathod inf-concave} {\textnormal{Tr}}|XY|^{r_0}=\inf\{\frac{r_0}{r_1}{\textnormal{Tr}}|XZ|^{r_1}+\frac{r_0}{r_2}{\textnormal{Tr}}|Z^{-1}Y|^{r_2}:Z\textnormal{ invertible} \}$$ and $$\label{varitional mathod sup-convex} {\textnormal{Tr}}|XY|^{r_1}=\sup\{\frac{r_1}{r_0}{\textnormal{Tr}}|XZ|^{r_0}-\frac{r_1}{r_2}{\textnormal{Tr}}|Y^{-1}Z|^{r_2}:Z\in{\mathcal{M}}_N\}.$$ Let $\Vert\cdot\Vert_p$ denote the matrix $p$-norm. For any $Z$ invertible, we have by Hölder’s inequality that $${\textnormal{Tr}}|XY|^{r_0}\leq\Vert XZ\Vert^{r_0}_{r_1}\Vert Z^{-1}Y\Vert^{r_0}_{r_2}=[{\textnormal{Tr}}|XZ|^{r_1}]^{\frac{r_0}{r_1}}[{\textnormal{Tr}}|Z^{-1}Y|^{r_2}]^{\frac{r_0}{r_2}}.$$ Then from the Young’s inequality for numbers (or AM-GM inequality): $x^{\alpha}y^{\beta}\leq\alpha x+\beta y$ for positive $x,y$ and positive $\alpha,\beta$ such that $\alpha+\beta=1$, it follows $$\label{ineq of Holder plus Young-inf} {\textnormal{Tr}}|XY|^{r_0}\leq[{\textnormal{Tr}}|XZ|^{r_1}]^{\frac{r_0}{r_1}}[{\textnormal{Tr}}|Z^{-1}Y|^{r_2}]^{\frac{r_0}{r_2}}\leq\frac{r_0}{r_1}{\textnormal{Tr}}|XZ|^{r_1}+\frac{r_0}{r_2}{\textnormal{Tr}}|Z^{-1}Y|^{r_2}.$$\[ineq of Holder plus Young-sup\] Exchanging $Y$ and $Z$, we have $${\textnormal{Tr}}|XY|^{r_1}\geq\frac{r_1}{r_0}{\textnormal{Tr}}|XZ|^{r_0}-\frac{r_1}{r_2}{\textnormal{Tr}}|Y^{-1}Z|^{r_2}.$$ Now to prove assume first that both $X$ and $Y$ are invertible, then $$\label{varitional mathod min-concave} {\textnormal{Tr}}|XY|^{r_0}=\min\{\frac{r_0}{r_1}{\textnormal{Tr}}|XZ|^{r_1}+\frac{r_0}{r_2}{\textnormal{Tr}}|Z^{-1}Y|^{r_2}:Z\textnormal{ invertible} \}$$ To see this, let $Y^*X^*=U|Y^*X^*|$ be the polar decomposition of $Y^*X^*$, then $XYU=|Y^*X^*|$. Set $Z:=YU|Y^*X^*|^{-\frac{r_1}{r_1+r_2}}$, we have $$XZ=XYU|Y^*X^*|^{-\frac{r_1}{r_1+r_2}}=|Y^*X^*|^{\frac{r_2}{r_1+r_2}},~~Z^{-1}Y=|Y^*X^*|^{\frac{r_1}{r_1+r_2}}U^*.$$ Using the facts that $\Vert\cdot \Vert_p$ is unitary invariant and $\Vert A \Vert_p=\Vert A^*\Vert_p$ for all $A$, we have $${\textnormal{Tr}}|XZ|^{r_1}={\textnormal{Tr}}|Y^*X^*|^{\frac{r_1 r_2}{r_1+r_2}}={\textnormal{Tr}}|XY|^{\frac{r_1 r_2}{r_1+r_2}}={\textnormal{Tr}}|XY|^{r_0},$$ and $${\textnormal{Tr}}|Z^{-1}Y|^{r_2}={\textnormal{Tr}}|Y^*X^*|^{\frac{r_1 r_2}{r_1+r_2}}={\textnormal{Tr}}|XY|^{\frac{r_1 r_2}{r_1+r_2}}={\textnormal{Tr}}|XY|^{r_0}.$$ Hence $\frac{r_0}{r_1}{\textnormal{Tr}}|XZ|^{r_1}+\frac{r_0}{r_2}{\textnormal{Tr}}|Z^{-1}Y|^{r_2}={\textnormal{Tr}}|XY|^{r_0}$, which proves . For general $X,Y$, $X_{\epsilon}:=X+\epsilon I$ and $Y_{\epsilon}:=Y+\epsilon I$ are both invertible for small $\epsilon>0$. Then we have shown that there exists invertible $Z_{\epsilon}$ such that $${\textnormal{Tr}}|X_{\epsilon}Y_{\epsilon}|^{r_0}=\frac{r_0}{r_1}{\textnormal{Tr}}|XZ_{\epsilon}|^{r_1}+\frac{r_0}{r_2}{\textnormal{Tr}}|Z_{\epsilon}^{-1}Y|^{r_2}.$$ Thus the proof of is finished, as soon as one observes ${\textnormal{Tr}}|XY|^{r_0}=\lim\limits_{\epsilon\to 0^+}{\textnormal{Tr}}|X_{\epsilon}Y_{\epsilon}|^{r_0}$. Now we prove in a similar way. Assume that $X$ is invertible and we show first that $$\label{varitional mathod max-concave} {\textnormal{Tr}}|XY|^{r_1}=\max\{\frac{r_1}{r_0}{\textnormal{Tr}}|XZ|^{r_0}-\frac{r_1}{r_2}{\textnormal{Tr}}|Y^{-1}Z|^{r_2}:Z\in{\mathcal{M}}_N\}.$$ Let $U$ be as above and choose $Z$ to be $YU|Y^*X^*|^{\frac{r_1}{r_2}}$, then $$XZ=XYU|Y^*X^*|^{\frac{r_1}{r_2}}=|Y^*X^*|^{\frac{r_1+r_2}{r_2}},~~ Y^{-1}Z=U|Y^*X^*|^{\frac{r_1}{r_2}}.$$ It follows that $${\textnormal{Tr}}|XZ|^{r_0}={\textnormal{Tr}}|Y^*X^*|^{\frac{r_1+r_2}{r_0 r_2}}={\textnormal{Tr}}|Y^*X^*|^{r_1}={\textnormal{Tr}}|XY|^{r_1},$$ and $${\textnormal{Tr}}|Y^{-1}Z|^{r_2}={\textnormal{Tr}}|Y^*X^*|^{r_1}={\textnormal{Tr}}|XY|^{r_1}.$$ Hence ${\textnormal{Tr}}|XY|^{r_1}=\frac{r_1}{r_0}{\textnormal{Tr}}|XZ|^{r_0}-\frac{r_1}{r_2}{\textnormal{Tr}}|Y^{-1}Z|^{r_2}$. This proves and then the proof of follows a similar limit argument as above. It is possible generalize this variational method to infinite dimensional case or even more general von Neumann algebras, which is beyond the aim of this paper. It is also possible to apply this variational method to trace functions with $n\geq 3$ variables. Indeed, let $r_j>0,j=0,1,\dots,n$ such that $\frac{1}{r_0}=\sum_{j=1}^{n}\frac{1}{r_j}$. Then we have for $X_1,\dots,X_n\in{\mathcal{M}}_N$ $$\label{n variables-inf} {\textnormal{Tr}}|X_1\cdots X_n|^{r_0}=\inf\{\frac{r_0}{r_1}{\textnormal{Tr}}|X_1Z_1|^{r_1}+\sum_{j=2}^{n-1}\frac{r_0}{r_j}{\textnormal{Tr}}|Z^{-1}_{j-1}X_jZ_j|^{r_j}+\frac{r_0}{r_n}{\textnormal{Tr}}|Z^{-1}_{n-1}X_n|^{r_n}\},$$ and $$\label{n variables-sup} {\textnormal{Tr}}|X_1\cdots X_n|^{r_n} =\sup\{\frac{r_n}{r_0}{\textnormal{Tr}}|Z_{n-1}X_{n}|^{r_0}-\sum_{j=2}^{n-1}\frac{r_n}{r_{j}}{\textnormal{Tr}}|Z_{j}X^{-1}_{j}Z^{-1}_{j-1}|^{r_{n-1}}-\frac{r_n}{r_1}{\textnormal{Tr}}|Z_{1}X^{-1}_{1}|^{r_1}\},$$ where the supremum and infimum run over all invertible $Z_1,\dots,Z_{n-1}$. The proof is similar to two variables case. We only explain here when infimum is achieved for . By a similar limit argument in above theorem it is reduced to show with inf replaced by min for all invertible $X_j$. Let $X^*_n\cdots X^*_1=U|X^*_n\cdots X^*_1|$ be the polar decomposition of $X^*_n\cdots X^*_1$. Then set $$Z_j:=X_{j+1}\cdots X_nU|X^*_n\cdots X^*_1|^{\alpha_j},~~\alpha_j=\sum_{k=1}^{j}\frac{r_0}{r_{k}}-1$$ for $1\leq j\leq n-1$. One can check that $${\textnormal{Tr}}|X_1\cdots X_n|^{r_0}=\frac{r_0}{r_1}{\textnormal{Tr}}|X_1Z_1|^{r_1}+\sum_{j=2}^{n-1}\frac{r_0}{r_j}{\textnormal{Tr}}|Z^{-1}_{j-1}X_jZ_j|^{r_j}+\frac{r_0}{r_n}{\textnormal{Tr}}|Z^{-1}_{n-1}X_n|^{r_n}.$$ Now we are ready to prove Theorem \[main thm\]. Before proceeding the proof note first that $$\Psi_{p,q,s}(A,B)={\textnormal{Tr}}(B^{\frac{q}{2}}K^*A^{p}KB^{\frac{q}{2}})={\textnormal{Tr}}|A^{\frac{p}{2}}KB^{\frac{q}{2}}|^{2s}.$$ We shall use an easy fact that the joint convexity (resp. joint concavity) is stable under taking supremum (resp. infimum).\ (1) If $q=0$ then it is reduced to Theorem \[thm:one variable case\] (1). To show the case $0<q\leq p\leq 1$ and $0\leq s\leq\frac{1}{p+q}$, set $\lambda:=s(p+q)\in(0,1]$ and we apply to $(r_0,r_1,r_2)=(2s,\frac{2\lambda}{p},\frac{2\lambda}{q})$ and $(X,Y)=(A^{\frac{p}{2}}K,B^{\frac{q}{2}})$: $$\label{proof of (1)} \Psi_{p,q,s}(A,B)=\inf\{\frac{p}{p+q}{\textnormal{Tr}}|A^{\frac{p}{2}}KZ|^{\frac{2\lambda}{p}}+\frac{q}{p+q}{\textnormal{Tr}}|Z^{-1}B^{\frac{q}{2}}|^{\frac{2\lambda}{q}}:Z\textnormal{ invertible} \}$$ Since $0<\frac{\lambda}{p}\leq\frac{1}{p}$ and $0<\frac{\lambda}{q}\leq\frac{1}{q}$, from Theorem \[thm:one variable case\] (1) it follows the maps $$A\mapsto\frac{p}{p+q}{\textnormal{Tr}}|A^{\frac{p}{2}}KZ|^{\frac{2\lambda}{p}}=\frac{p}{p+q}{\textnormal{Tr}}(Z^*K^*A^{p}KZ)^{\frac{\lambda}{p}}$$ and $$B\mapsto\frac{q}{p+q}{\textnormal{Tr}}|Z^{-1}B^{\frac{q}{2}}|^{\frac{2\lambda}{q}}=\frac{q}{p+q}{\textnormal{Tr}}(Z^{-1}B^{q}(Z^{-1})^*)^{\frac{\lambda}{q}}$$ are both concave. Hence they are both joint concave in $(A,B)$ and so is $\Psi_{p,q,s}$ by .\ (2) If $p=0$, then it is reduced to Theorem \[thm:one variable case\] (2). Suppose $-1\leq q\leq p<0$ and $s>0$, then we apply to $(r_0,r_1,r_2)=(2t,2s,\frac{2}{-q})$ with $\frac{1}{t}=\frac{1}{s}-q$ and $(X,Y)=(A^{\frac{p}{2}}K,B^{\frac{q}{2}})$: $$\label{proof of (2)} \Psi_{p,q,s}(A,B)=\sup\{\frac{s}{t}{\textnormal{Tr}}|A^{\frac{p}{2}}KZ|^{2t}+sq{\textnormal{Tr}}|B^{-\frac{q}{2}}Z|^{\frac{2}{-q}}:Z\in{\mathcal{M}}_N\}.$$ Note that $t>0$, $sq<0$ and $0<-q\leq 1$. By Theorem \[thm:one variable case\] (1) and (2), the maps $$A\mapsto\frac{s}{t}{\textnormal{Tr}}|A^{\frac{p}{2}}KZ|^{2t}=\frac{s}{t}{\textnormal{Tr}}(Z^*K^*A^{p}KZ)^{t}$$ and $$B\mapsto sq{\textnormal{Tr}}|B^{-\frac{q}{2}}Z|^{\frac{2}{-q}}=sq{\textnormal{Tr}}(Z^*B^{-q}Z)^{\frac{1}{-q}}$$ are both convex. Hence they are both joint convex in $(A,B)$ and so is $\Psi_{p,q,s}$ by .\ (3) If $q=0$, then it is reduced to Theorem \[thm:one variable case\] (3). Suppose $-1\leq q<0,~1\leq p\leq 2,~(p,q)\ne(1,-1)$ and $s\geq\frac{1}{p+q}$, then we apply to $(r_0,r_1,r_2)=(2t,2s,\frac{2}{-q})$ with $\frac{1}{t}=\frac{1}{s}-q$ and $(X,Y)=(A^{\frac{p}{2}}K,B^{\frac{q}{2}})$: $$\label{proof of (3)} \Psi_{p,q,s}(A,B)=\sup\{\frac{s}{t}{\textnormal{Tr}}|A^{\frac{p}{2}}KZ|^{2t}+sq{\textnormal{Tr}}|B^{-\frac{q}{2}}Z|^{\frac{2}{-q}}:Z\in{\mathcal{M}}_N\}.$$ Since $sq<0$, $0<-q\leq 1$ and $t=\frac{1}{s^{-1}-q}\geq\frac{1}{p}$, we have by Theorem \[thm:one variable case\] (1) and (3) that the maps $$A\mapsto\frac{s}{t}{\textnormal{Tr}}|A^{\frac{p}{2}}KZ|^{2t}=\frac{s}{t}{\textnormal{Tr}}(Z^*K^*A^{p}KZ)^{t}$$ and $$B\mapsto sq{\textnormal{Tr}}|B^{-\frac{q}{2}}Z|^{\frac{2}{-q}}=sq{\textnormal{Tr}}(Z^*B^{-q}Z)^{\frac{1}{-q}}$$ are both convex. Hence they are both joint convex in $(A,B)$ and so is $\Psi_{p,q,s}$ by . \[last rem\] Although the variational methods of and admit analogs and of $n(\geq 3)$ variables, the joint convexity/concavity of $${\mathcal{P}}_N\times{\mathcal{P}}_N\times\cdots\times{\mathcal{P}}_N\ni (A_1,A_2,\dots,A_n)\mapsto{\textnormal{Tr}}(A_n^{\frac{p_n}{2}}K^{*}_{n-1}\cdots K^{*}_{1}A_1^{p_1}K_1\cdots K_{n-1}A_n^{\frac{p_n}{2}})^s$$ can not be derived directly from Theorem \[thm:one variable case\] because of the appearance of the term ${\textnormal{Tr}}|Z^{-1}_{j-1}X_jZ_j|^{r_j}$. For example, we have $$\label{variation-three variables} {\textnormal{Tr}}|X_{1}X_{2}X_{3}|^{r_0}=\inf\{\frac{r_0}{r_1}{\textnormal{Tr}}|X_{1}Z_{1}|^{r_1}+\frac{r_0}{r_2}{\textnormal{Tr}}|Z_{1}^{-1}X_{2}Z_{2}|^{r_2}+\frac{r_0}{r_3}{\textnormal{Tr}}|Z^{-1}_{2}X_{3}|^{r_3}\},$$ where the infimum runs over all invertible $Z_1$ and $Z_{2}$. To obtain the joint concavity of $${\mathcal{P}}_N\times{\mathcal{P}}_N\times{\mathcal{P}}_N\ni (A_1,A_2,A_3)\mapsto{\textnormal{Tr}}(A_3^{\frac{p_3}{2}}K^{*}_{2}A_2^{\frac{p_2}{2}}K^{*}_{1}A_1^{p_1}K_1A_2^{\frac{p_2}{2}}K_{2}A_3^{\frac{p_3}{2}})^s,$$ via the variational method , the concavity of the function of the form $${\mathcal{P}}_N\ni A_2\mapsto{\textnormal{Tr}}|Y_{1}A_{2}^{\frac{p_2}{2}}Y_{2}|^{r_2}={\textnormal{Tr}}(Y_2^*A_{2}^{\frac{p_2}{2}}Y_1^*Y_1A_{2}^{\frac{p_2}{2}}Y_2)^{\frac{r_2}{2}}$$ is required. Unfortunately, few is known for general $Y_1^*Y_1$. Indeed, Carlen, Frank and Lieb proved that [@CFL16-some Corollary 3.3] for $p,q,r\in\mathbb{R}\setminus\{0\}$, the function $$(A,B,C)\mapsto{\textnormal{Tr}}C^{\frac{r}{2}}B^{\frac{q}{2}}A^{p}B^{\frac{q}{2}}C^{\frac{r}{2}}$$ is never concave, and it is convex if and only if $q=2,~p,r<0$ and $-1\leq p+r<0$. Acknowledgment {#acknowledgment .unnumbered} -------------- The author would like to thank his advisor Professor Quanhua Xu for holding the Summer Working Seminar on Quantum Information in Harbin, July-August 2018, which brought him into this field. He is also grateful for Quanhua Xu, Adam Skalski and Zhi Yin for some valuable comments on a draft of this paper. The research was partially supported by the NCN (National Centre of Science) grant 2014/14/E/ST1/00525, the French “Investissements d’Avenir" program, project ISITE-BFC (contract ANR-15-IDEX-03) and NSFC No. 11431011. [MLDSF]{} K. M. R. Audenaert and N. Datta. -[$z$]{}-[R]{}ényi relative entropies. , 56(2):022202, 16, 2015. T. Ando. Concavity of certain maps on positive definite matrices and applications to [H]{}adamard products. , 26:203–241, 1979. E. A. Carlen, R. L. Frank, and E. H. Lieb. Some operator and trace function convexity theorems. , 490:174–185, 2016. E. A. Carlen, R. L. Frank, and E. H. Lieb. . , June 2018. E. A. Carlen and E. H. Lieb. A [M]{}inkowski type trace inequality and strong subadditivity of quantum entropy. [II]{}. [C]{}onvexity and concavity. , 83(2):107–126, 2008. H. Epstein. Remarks on two theorems of [E]{}. [L]{}ieb. , 31:317–325, 1973. R. L. Frank and E. H. Lieb. Monotonicity of a relative [R]{}ényi entropy. , 54(12):122201, 5, 2013. F. Hiai. Concavity of certain matrix trace and norm functions. , 439(5):1568–1589, 2013. F. Hiai. Concavity of certain matrix trace and norm functions. [II]{}. , 496:193–220, 2016. V. Jaksic, Y. Ogata, Y. Pautrat, C. Pillet. Entropic Fluctuations in Quantum Statistical Mechanics. An Introduction. Volume 95, August 2010, Oxford University Press, 2012. E. H. Lieb. Convex trace functions and the [W]{}igner-[Y]{}anase-[D]{}yson conjecture. , 11:267–288, 1973. G. Lindblad. Expectations and entropy inequalities for finite quantum systems. , 39:111–119, 1974. M. Müller-Lennert, F. Dupuis, O. Szehr, S. Fehr, and M. Tomamichel. On quantum [R]{}ényi entropies: a new generalization and some properties. , 54(12):122203, 20, 2013. A. Rényi. On measures of entropy and information. In [*Proc. 4th [B]{}erkeley [S]{}ympos. [M]{}ath. [S]{}tatist. and [P]{}rob., [V]{}ol. [I]{}*]{}, pages 547–561. Univ. California Press, Berkeley, Calif., 1961. W. F. Stinespring. Positive functions on [$C^*$]{}-algebras. , 6:211–216, 1955. A. Uhlmann. Endlich-dimensionale [D]{}ichtematrizen. [II]{}. , 22:139–177, 1973. H. Umegaki. Conditional expectation in an operator algebra. [IV]{}. [E]{}ntropy and information. , 14:59–85, 1962. M. M. Wilde, A. Winter, and D. Yang. Strong converse for the classical capacity of entanglement-breaking and [H]{}adamard channels via a sandwiched [R]{}ényi relative entropy. , 331(2):593–622, 2014.
{ "pile_set_name": "ArXiv" }
[**The analog of the Schauder inequality for closed surfaces in Euclidean spaces\ **]{} [**Andrei I. Bodrenko**]{} [^1] [**Abstract**]{}\ [The analog of the Schauder inequality for closed surfaces in Euclidean spaces is obtained in this article. ]{} Introduction {#introduction .unnumbered} ============ Let $E^{n}$ be $n$-dimensional ($n>1$) Euclidean space. Let $D$ be the finite domain in $E^n$, $\partial D$ be the boundary of $D,$ $\bar D$ be the closure of $D.$ Let $(x^{1},...,x^{n})$ be the Cartesian coordinates in $E^{n}.$ \[definition 1\]. We say, that function $f$ on $D$ is of class $C^{k,s}(\bar D)$ , $k\geq 0,$ $s\in (0,1),$ if it has continuous partial derivatives up to $k$-th order inclusively and bounded value $$|f|_{(D) k,s}= \sum_{|i|=0}^{k} \sup_{x\in D} | \partial^{|i|} f(x)| + \sum_{|i|=k} \sup_{x_1, x_2 \in D} \frac{|\partial^{|i|} f(x_{1})- \partial^{|i|}f(x_{2})|} {|x_{1}-x_{2}|^{s}}. \eqno(1)$$ Partial derivatives of function $f(x)$ are denoted by $\partial^{|i|} f(x)\equiv$ $\frac{\partial^{|i|}f(x)} {\partial^{i_{1}}x^{1}...\partial^{i_{n}}x^{n}}$, where $|i|=i_{1}+ ... +i_{n}$ is the order of derivative. $|x|=(\sum_{i=1}^{n}(x^{i})^{2})^{1/2}$, where $(x^{1},...,x^{n})$ are the coordinates of point $x\in E^{n}$. The value $|f|_{(D) k,s}$ we call the norm of function $f$ in the space $C^{k,s}(\bar D)$. It is known (see \[1\]) that the space $C^{k,s}(\bar D)$ with norm denoted by formula $(1)$ is complete normed space. We define the cylinder $C_{R,L}$ in $E^{n}$ by the following formula: $$C_{R,L}=\Biggl\{x: \sum_{i=1}^{n-1}(x^{i})^{2} < R^{2}, -2LR < x^{n} < 2LR \Biggr\},$$ where $ L=const>0, R=const>0, $ and let $x=(0,0,...,0)$ be called its center. \[definition 2\]. Domain $D$ is called strictly Lipschitzian if for every point $x_{0}\in \partial D$ it can be introduced the coordinates $$u^{k}=\sum_{l=1}^{n} c^{k}_{l}(x^{l}-x^{l}_{0}), \qquad k=\overline {1,n},$$ where $||c^{k}_{l}||$ is orthogonal matrix such that the intersection of $\partial D$ and the cylinder $\bar C_{R,L}$ corresponding to the coordinates $\{u^{k}\}$ , is given by the equation $$u^{n}=\omega (u^{,n}), \qquad u^{,n}\equiv (u^{1},...,u^{n-1}),$$ where $\omega (u^{,n})$ is Lipschitzian function for $|u^{,n}|\leq R$ with Lipschitz constant bounded by $L,$ and $$\bar D \cap \bar C_{R,L} =\Biggl\{ u: |u^{,n}|\leq R, \quad -2LR\leq u^{n}\leq\omega (u^{,n}) \Biggr\}.$$ The numbers $R$ and $L$ are fixed for the domain $D$. Arbitrary convex domain is strictly Lipschitzian (see \[1\]). Let $x_{0}=(x^{1}_{0},...,x^{n}_{0})$ be a point of $\partial D$, where surface $\partial D$ has tangent plane. \[definition 3\]. We call $(u^{1},...,u^{n})$ the specific local coordinate system with origin at the point $x_{0}$ if the coordinates $\{u^{k}\}$ and $\{x^{k}\}$ satisfy the following equations: $u^{k}=\sum_{l=1}^{n} c^{k}_{l}(x^{l}-x^{l}_{0}),$ $ k=\overline {1,n},$ and the axis $u^{n}$ is directed to the normal of $\partial D$ at the point $x_{0}$, that is outward for $D$. \[definition 4\]. Domain $D$ is called the domain of class $C^{l,s}$, $l\geq 1$, if it is strictly Lipschitzian and the coordinates $\{u^{k}\}$ that are given in the definition 2 are the specific local coordinates, the function $u^{n}=\omega (u^{,n})$, defining the equation of the surface $\partial D$ is of class $C^{l,s}(|u^{,n}|\leq R)$. \[definition 5\]. We will say that the boundary $\partial D$ of domain $D$ is of class $C^{l,s}$ if for every point $x_{0}\in \partial D$ there can be introduced the the specific local coordinates such that the function $u^{n}=\omega (u^{,n})$ is of class $C^{l,s}(|u^{,n}|\leq R).$ Let the boundary $\partial D$ of the domain $D$ is of class $C^{l_{1},s_{1}},$ $s_{1}\in (0;1)$. Let on $\partial D$ be given the function $\varphi (x), x\in \partial D$. \[definition 6\]. We will say that function $\varphi (x)$ is of class $C^{l,s}(\partial D)$ if it as a function of the specific local coordinates $u^{,n}=(u^{1},...,u^{n-1})$ introduced for every point $x_{0}\in\partial D$ is of class $C^{l,s}(|u^{,n}|\leq R),$ where $|u^{,n}|\leq R$ is the base of cylinder corresponding to the point $x_{0}.$ \[definition 7\]. Norm $|\varphi|_{(\partial D) l,s}$ of function $\varphi$ , given on the surface $\partial D$ is called the greatest of the norms $|\varphi(u^{,n})|_{(|u^{,n}|\leq R) l,s} ,$ calculated for all points $x_{0}\in \partial D.$ Let $F$ be the closed orientable hypersurface in Euclidean space $E^{n+1}.$ Let $(y^{1},...,y^{n+1})$ be the Cartesian coordinates in $E^{n+1}.$ Let $U$ be arbitrary open set on $F.$ \[definition 8\]. Couple $(U,h)$ is called the admissible map of class $C^{k,s}$ if: 1\) $h$ is the homeomorphism $U$ on the open ball $K_{r}$ of radius $r>0$ in $E^{n};$ 2\) the inverse mapping $\bar h^{-1}(x) \equiv (f^{1}(x),...,f^{n+1}(x))$, $x\in K_{r}$, satisfies the condition: $f^{\alpha}\in C^{k,s}(\bar K_{r})$, $\alpha=\overline{1,n+1}$. \[definition 9\]. $F$ is called the surface of class $C^{k,s}$ if the following conditions hold: 1\) $F$ is the surface of class $C^{k}$; 2\) on $F,$ there exists the finite aggregate $\{(U_{i},h_{i})\}_{i=\overline {1,N}}$ of admissible maps of class $C^{k,s},$ where the collection of sets $(U_{i})_{i=\overline {1,N}}$ is open covering of $F$; 3\) if $U_{i}\cap U_{j}\not =\emptyset$ then the mapping $h_{j}\circ h^{-1}_{i}$ is diffeomorphism of class $C^k$ of $h_{i}(U_{i}\cap U_{j})$ on set $h_{j}(U_{i}\cap U_{j})$. \[definition 10\]. The aggregate $\{(U_{i},h_{i})\}_{i=\overline {1,N}}$ of the admissible maps considered in definition 9 is called the admissible atlas of class $C^{k,s}$ of hypersurface $F.$ \[definition 11\]. Function $f$ determined on surface $F,$ is called the function of class $C^{p,s}(F)$, $p<k $, if 1\) on hypersurface $F,$ there exists the admissible atlas $\{(U_{i},h_{i})\}_{i=\overline {1,N}}$ of class $C^{k,s}$ and 2\) $f\circ h_{i}^{-1}\in C^{p,s}(\bar K_{r_{i}})$, $i=\overline {1,N}$. We fix the admissible atlas $\{(U_{i},h_{i})\}_{i=\overline {1,N}}$ of class $C^{k,s}$ of hypersurface $F$. Let be given the function $f\in C^{p,s}(F)$. The norm of function $f$ in space $C^{p,s}(F)$ is determined by the following formula: $$|f|_{(F) p,s}= \max_{i} |f\circ h^{-1}_{i}|_{(K_{r_{i}}) p,s}.$$ We will prove that the obtained normed space is complete. Let $\{f_{m}\}_{m=1}^{\infty}$ be the Cauchy sequence of functions $f_m$ of class $C^{p,s}(F),$ therefore $\forall \varepsilon>0 $ and for every natural number $l$ $$|f_{m+l}-f_{m}|_{(F) p,s}< \varepsilon,$$ for all safficiently large $m.$ Then, from the definition of norm, we obtain $$|f_{m+l}\circ h^{-1}_{i}-f_{m}\circ h^{-1}_{i}|_{(K_{r_{i}})p,s}<\varepsilon.$$ Since the function space $C^{p,s}(\bar K_{r_{i}})$ is complete normed space, then the sequence of functions $\{f_{m}\circ h^{-1}_{i}\}$ on $\bar K_{r_{i}}$ has limit: $f\circ h^{-1}_{i}=\lim_{m\rightarrow\infty}f_{m}\circ h^{-1}_{i}.$ Since $f\circ h^{-1}_{i}\in C^{p,s}(\bar K_{r_{i}}), \forall i=\overline{1,N},$ then $f\in C^{p,s}(F)$. Statement of the result. ======================== Let $F\in C^{3,s},$ where $s\in (0;1).$ Let $\{(U_{i},h_{i})\}_{i=\overline {1,N}}$ be the admissible atlas $F$ of class $C^{3,s}$. Let $(U,h)$ be an arbitrary map from the collection $\{(U_{i},h_{i})\}_{i=\overline {1,N}}$. Then the hypersurface $F$ on map $(U,h)$ is determined by the following equation system: $$y^{\alpha}\equiv h^{-1 \alpha}(x)=f^{\alpha}(x^{1},...,x^{n}), \alpha=\overline{1,n+1}, (x^{1},...,x^{n})\in K_{r} \eqno(2)$$ Consider the differential operator $A$ on $F$ that, on map $(U,h),$ is defined by: $$A=\sum_{k=1}^{n}\sum_{p=1}^{n}a^{kp} \partial_{k}(\partial_{p}) + \sum_{j=1}^{n}b^{j}\partial_{j}+c. \eqno(3)$$ Let $a^{kp}=a^{pk},$ and the operator $A$ is strictly elliptic on $F,$ i. e. $$\sum_{k=1}^{n}\sum_{p=1}^{n} a^{kp}(x) \zeta_{k}\zeta_{p} \geq \nu \sum_{k=1}^{n}(\zeta_{k})^{2}, \quad \nu=const>0, \quad \forall \zeta_{k}, \quad k=\overline {1,n}.$$ Let, on $F,$ be given a function $f$ of class $C^{2,s}(F).$ Then we have: $$A(f\circ h^{-1}(x))=\sum_{k=1}^{n}\sum_{p=1}^{n}a^{kp}(x) \partial_{k}(\partial_{p}(f\circ h^{-1}(x))) + \sum_{j=1}^{n}b^{j}(x)\partial_{j}(f\circ h^{-1}(x)) + c(x) f\circ h^{-1}(x),$$ where $x\in K_{r}$ for every admissible map $(U,h)$ of class $C^{3,s}.$ \[theorem1\]. Let function $f$ be a solution of class $C^{2,s}(F)$ of the problem: $Af=\gamma,$ where $c(x)\neq 0$ on $F,$ $\gamma\in C^{0,s}(F),$ $a^{kp}\in C^{0,s}(F),$ $b^{j}\in C^{0,s}(F),$ $c \in C^{0,s}(F).$ Then the following inequality holds: $$|f|_{(F) 2,s}\leq M|\gamma|_{(F)0,s},$$ where the constant $M$ depends on $s,n,$ the surface $F,$ the coefficients $a^{kp}, b^{j} , c $ $ ( k,p,j=\overline {1,n} )$ and the admissible atlas on $F$ $\{(U_{i},h_{i})\}_{i=\overline {1,N}}$ of class $C^{3,s}$. Auxiliary conjectures. ====================== \[note1\]: Let $f$ be the function of class $C^{2,s}(F),$ $\{U_{i},h_{i}\}$ be the admissible map on $F,$ $h_{i}(U_{i})=K_{r_{i}}$. Let $x_{0}\in \partial K_{r_{i}}$. Consider the specific local coordinates $(x^{1},...,x^{n})$ for the point $x_{0}$ where $ x^{n}=\omega(x^{1},...,x^{n-1}).$ Then the intersection of the cylinder $\bar{C}_{R,L}$ at the point $x_{0}$ and the surface $\partial K_{r_{i}}$ is given by: $ x^{n}=\omega(x^{1},...,x^{n-1}), |(x^{1},...,x^{n-1})|\leq R.$ We assume that the specific local coordinates were introduced in the ball $K_{r_{i}}$. Let $O_{i}(x_{0})$ be an open domain in $K_{r_{i}}$ such that $\bar{O}_{i}(x_{0})\supset (\partial K_{r_{i}}\cap \bar{C}_{R,L})$. Let $B_{\rho_{j}}(x_{0})$ be an open ball â $E^{n}$ of radius $\rho_{j}$ with center at the point $x_{0}.$ We will prove several lemmas before the theorem 1. \[lemma1\]. There exist numbers $R,L,Q$ and set collection $\{O_{i}(x_{0})\}_{i=\overline {1,N}}$ such that for every point $x_{0}\in \partial K_{r_{i}}$ and for all $i=\overline {1,N}$ the following conditions hold: 1\) $\exists j\not =i$ such that $h_{i}^{-1}(O_{i}(x_{0}))\subset U_{j}$. 2\) $\exists \rho_{j}>0$ such that $h_{j}(h_{i}^{-1}(O_{i}(x_{0})))\subset B_{\rho_{j}}(x_{0}) \subset K_{r_{j}}$, where $dist(B_{\rho_{j}}(x_{0}),\partial K_{r_{j}})\geq Q > 0$. Proof of lemma 1 follows from definition 10, compactness of $\partial K_{r_{j}},$ definition 2 and finiteness of covering $\{U_{i}\}_{i=\overline {1,N}},$ for sufficiently small numbers $R$ and $L.$ We fix numbers $R,L,Q,$ point $x_{0}\in \partial K_{r_{i}}$ and set collection $\{O_{i}(x_{0})\}_{i=\overline {1,N}},$ that satisfy lemma 1. \[lemma2\]. Under the conditions of lemma 1, the following inequality holds: $$\sup_{|(x^{1},...,x^{n-1})|\leq R} \left| f\circ h^{-1}_{i} (x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1})) \right| \leq \sup_{u\in B_{\rho_{j}}} \left| f\circ h^{-1}_{j}(u) \right|.$$ [**Proof.**]{} Since there exists a diffeomorphism of class $C^{3}$ of the neighborhood $O_{i}(x_{0})$ into the ball $B_{\rho_{j}}(x_{0}),$ then there exist mappings: $u^{p}=k^{p}(x^{1},...,x^{n}),p=\overline {1,n},$ $\forall x\in O_{i}(x_{0}),$ $x^{p}=g^{p}(u^{1},...,u^{n}), \forall u\in h_{j}(h^{-1}_{i}(O_{i}(x_{0})).$ Hence for every point $x\in O_{i}(x_{0})$ there exists point $u\in B_{\rho_{j}}(x_{0})$ such that $f\circ h^{-1}_{i}(x)=f\circ h^{-1}_{j}(u)$. Lemma is proved. \[lemma3\]. The following inequality holds: $$\sup_{|(x^{1},...,x^{n-1})|\leq R} \left| \frac{\partial}{\partial x^{k}} (f\circ h^{-1}_{i} (x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1}))) \right| \leq$$ $$\leq M \max_{p=\overline{1,n}}\sup_{u\in B_{\rho_{j}}} \left| \frac{\partial}{\partial u^{p}}(f\circ h^{-1}_{j}(u)) \right| , \quad k=\overline{1,n-1},$$ where $M=\mbox{const}>0$. [**Proof.**]{} We have $$\frac{\partial}{\partial x^{k}} \left( f\circ h^{-1}_{i} (x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1})) \right) =$$ $$= \frac{\partial}{\partial x^{k}} \left( f\circ h^{-1}_{j}(u^{1},...,u^{n}) \right) = \frac{\partial}{\partial u^{p}} \left( f\circ h^{-1}_{j}(u^{1},...,u^{n}) \right) \frac{\partial k^{p}}{\partial x^{k}},$$ where the point $(u^{1},...,u^{n})\in B_{\rho_{j}}(x_{0})$, $u^{l}=k^{l}(x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1}))$, $l=\overline{1,n}$. Since the functions $k^{p}\in C^{3}$ then the functions $\frac{\partial k^{p}}{\partial x^{k}}$ are bounded on $|(x^{1},...,x^{n-1})|\leq R$. Lemma 3 is proved. \[lemma4\]. The following inequality holds: $$\sup_{|(x^{1},...,x^{n-1})|\leq R} \left| \frac{\partial^{2}}{\partial x^{k}\partial x^{q}} (f\circ h^{-1}_{i} (x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1}))) \right| \leq$$ $$\leq M_{1} \left( \max_{p=\overline{1,n}}\sup_{u\in B_{\rho_{j}}} \left| \frac{\partial}{\partial u^{p}}(f\circ h^{-1}_{j}(u)) \right| +\max_{p,r=\overline{1,n}}\sup_{u\in B_{\rho_{j}}} \left| \frac{\partial^{2}}{\partial u^{p}\partial u^{r}}(f\circ h^{-1}_{j}(u)) \right| \right) ,$$ $$k, q= \overline{1,n-1},$$ where $M_1=\mbox{const}>0$. [**Proof.**]{} We have $$\frac{\partial^{2}}{\partial x^{k}\partial x^{q}} \left( f\circ h^{-1}_{i} (x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1})) \right) =$$ $$= \frac{\partial^{2}}{\partial u^{p}\partial u^{r}} \left( f\circ h^{-1}_{j}(u^{1},...,u^{n}) \right) \frac{\partial k^{p}}{\partial x^{k}} \frac{\partial k^{r}}{\partial x^{q}}+ \frac{\partial}{\partial u^{p}} \left( f\circ h^{-1}_{j}(u^{1},...,u^{n}) \right) \frac{\partial^{2}k^{p}}{\partial x^{k}\partial x^{q}},$$ where the point $(u^{1},...,u^{n})\in B_{\rho_{j}}(x_{0})$, $u^{l}=k^{l}(x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1}))$, $l=\overline{1,n}$. Since the functions $k^{p}\in C^{3}$ therefore the functions $\frac{\partial k^{p}}{\partial x^{k}}$ and $\frac{\partial^{2}k^{p}}{\partial x^{k}\partial x^{q}}$ are bounded on $|(x^{1},...,x^{n-1})|\leq R$. Therefore we obtain the proof of lemma 4. \[lemma5\]. The following inequality holds: $$\left| f\circ h^{-1}_{i}(x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1})) \right| _{(|(x^{1},...,x^{n-1})|\leq R) 2,s} \leq$$ $$\leq M_{1} \left( |f\circ h^{-1}_{j}(u)|_{B_{\rho_{j}} 2,s}+ |f\circ h^{-1}_{j}(u)|_{B_{\rho_{j}} 1,s} \right),$$ where $M_1=\mbox{const}>0$. Proof follows from lemmas 2, 3 and 4. \[lemma6\]. For any function $f\in C^{2,s}(\bar {B}_{\rho _{j}})$ the following inequality holds: $$\left| f \right|_{(B_{\rho_{j}}) 1,s} \leq M_{2} \left| f \right|_{(B_{\rho_{j}}) 2,s},$$ where $M_2=\mbox{const}>0$. [**Proof.**]{} We have the inequality (see \[1\]): $$\sup_{u_{1},u_{2}\in B_{\rho_{j}}} \frac{|\partial(f(u_{1}))-\partial(f(u_{2}))|}{|u_{1}-u_{2}|^{s}}\leq$$ $$\leq M_{3} \left( \sum_{k=0}^{2}\sup_{B_{\rho_{j}}}|\partial^{k}(f)| \right)^{s} \left( \sum_{k=0}^{1}\sup_{B_{\rho_{j}}}|\partial^{k}(f)| \right)^{1-s} \leq M_{3} \left( \sum_{k=0}^{2}\sup_{B_{\rho_{j}}}|\partial^{k}(f)| \right),$$ where $M_3=\mbox{const}>0$. Therefore we obtain the proof of lemma 6. Proof of theorem 1. =================== By lemmas 5 and 6 we have the inequality: $$|f\circ h^{-1}_{i}(x^{1},...,x^{n-1},\omega(x^{1},...,x^{n-1}))| _{(|(x^{1},...,x^{n-1})|\leq R) 2,s} \leq$$ $$\leq M_{4}|f\circ h^{-1}_{j}(u)|_{B_{\rho_{j}} 2,s},$$ where $M_4=\mbox{const}>0$. Since the ball $\bar {K}_{r_{i}}$ is compact, then for some $j$ we obtain: $$\left| f \right|_{\partial K_{r_{i}}2,s} \leq M_{5} \left| f \right|_{B_{\rho_{j}}2,s} \eqno(4),$$ where $M_5=\mbox{const}>0$. Since function $f$ is a solution of class $C^{2,s}(F)$ of the problem $A f =\gamma$ then we have the inequality (see \[1\]): $$|f|_{B_{\rho_{j}}2,s}\leq M(B_{\rho_{j}}) \left( |\gamma|_{K_{r_{j}}0,s}+\max_{K_{r_{j}}}|f| \right) ,$$ where $\bar{B}_{\rho_{j}}\subset K_{r_{j}}$. We have the inequality (see \[1\]): $$\max_{K_{r_{j}}}|f| \leq \max \left( \max_{\partial K_{r_{j}}}|f|; \max_{K_{r_{j}}}\Bigl|\frac{\gamma}{c}\Bigr| \right).$$ Since $F$ is compact then there exists point $x_{0}\in F$ such that $|f(x_{0})|=\max_{F}|f|$. Therefore there exists neighborhood $U_{l}$ such that $x_{0}\in U_{l}$. Hence, we obtain: $$\max_{K_{r_{j}}}|f|\leq \max_{K_{r_{l}}}|f|\leq \max_{K_{r_{l}}}\Bigl|\frac{\gamma}{c}\Bigr|. \eqno(5)$$ We have the estimate (see \[1\]): $$|f|_{K_{r_{i}}2,s}\leq M_{6} \left( |\gamma|_{K_{r_{i}}0,s}+\max_{K_{r_{i}}}|f|+|f|_{\partial K_{r_{i}}2,s} \right),$$ where $M_6=\mbox{const}>0.$ Using inequalities (4) and (5), we finish the proof of theorem 1. [**References**]{} 1. O.A. Ladyjenskaya, N.N. Uraltseva. Linear and quasilinear equations of elliptic type. M: Nauka, 1973. [^1]: Andrei I. Bodrenko, associate professor, Department of Mathematics, Volgograd State University,\ Universitó Prospekt 100, Volgograd, 400062, RUSSIA. Email: [email protected]
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study preconditioning techniques for discontinuous Galerkin discretizations of isotropic linear elasticity problems in primal (displacement) formulation. We propose subspace correction methods based on a splitting of the vector valued piecewise linear discontinuous finite element space, that are optimal with respect to the mesh size and the Lamé parameters. The pure displacement, the mixed and the traction free problems are discussed in detail. We present a convergence analysis of the proposed preconditioners and include numerical examples that validate the theory and assess the performance of the preconditioners.' address: - 'Centre de Recerca Matemàtica, Campus de Bellaterra, 08193 Bellaterra, (Barcelona), Spain. Email: `[email protected]`' - ' Johann Radon Institute for Computational and Applied Mathematics, Austrian Academy of Sciences Altenberger Str. 69, 4040 Linz, Austria. Email: `[email protected]` ' - 'Johann Radon Institute for Computational and Applied Mathematics, Austrian Academy of Sciences Altenberger Str. 69, 4040 Linz, Austria. Email:`[email protected]` ' - 'Department of Mathematics, The Pennsylvania State University, University Park, PA 16802, USA. Email: `[email protected]` ' author: - Blanca Ayuso de Dios - Ivan Georgiev - Johannes Kraus - Ludmil Zikatanov title: A subspace correction method for ciscontinuous Galerkin discretizations of linear elasticity equations --- Introduction ============ The finite element approximation of the equations of isotropic linear elasticity may be accomplished in various ways. The most straightforward approach is to use the primal formulation and conforming finite elements. It is well known that such a method, in general, does not provide approximation to the displacement field when the material is nearly incompressible (the Poisson ratio is close to $1/2$). This phenomenon is called *volume locking*. To alleviate locking, several approaches exist. Among the possible solutions, we mention the use of mixed methods, reduced integration techniques, stabilization techniques, nonconforming methods, and the use of discontinuous Galerkin methods. We refer to [@FalkR-1991aa; @HansboP_LarsonM-2002aa] for further discussions on such difficulties and their remedies. In this work we focus on the Symmetric Interior Penalty discontinuous Galerkin (SIPG) methods introduced in [@HansboP_LarsonM-2002aa; @HansboP_LarsonM-2003aa; @WihlerT-2004aa; @WihlerT-2006aa] for the approximation of isotropic linear elasticity. We have chosen to work with these DG discretizations, since we have in mind a method that is simple but still applicable to different types of boundary conditions. In fact, unlike classical low order non-conforming methods (see [@FalkR-1991aa]), the Interior Penalty (IP) stabilization methods introduced in [@HansboP_LarsonM-2002aa; @HansboP_LarsonM-2003aa] can be shown to be stable in the case of essential (Dirichlet or pure displacement) boundary conditions, or natural (Neumann type, or traction free) boundary conditions. As a consequence, these IP methods provide a robust approximation to the displacement field and avoid the volume locking regardless the boundary conditions of the problem. For the design of the preconditioners we follow the ideas introduced in [@Ayuso-de-DiosB_ZikatanovL-2009aa] for second order elliptic problems. However, such extensions are not straightforward, since we aim at constructing preconditioners that work well for three different types of boundary conditions: essential, natural and mixed boundary conditions, used in linear elasticity. This complicates the matters quite a bit. We consider a splitting of the vector valued, piecewise linear, discontinuous finite element space, into two subspaces: the vector valued Crouzeix-Raviart space and a space complementary to it which consists of functions whose averages are $L^2$ orthogonal to the constants on every edge/face of the partition. This space decomposition is direct and the spaces are orthogonal with respect to a bilinear form obtained via using “reduced integration” to calculate the contributions of the penalty terms in SIPG. In the pure displacement case (essential boundary conditions), the restriction of the bilinear form based on reduced integration is coercive on the Crouzeix-Raviart space and is spectrally equivalent to the SIPG bilinear form. The space decomposition mentioned above is then orthogonal in this reduced integration bilinear form. Thus, in case of essential boundary conditions we have a natural block diagonal preconditioner for the linear elasticity problem: (1) a solution of a problem arising from discretization by nonconforming Crouzeix-Raviart elements; (2) solution of a well-conditioned problem on the complementary space. For traction free problems or problems with Dirichlet conditions only on part of the boundary, the situation is quite different. On one hand the reduced integration bilinear form when restricted to the Crouzeix-Raviart space has a null space whose dimension depends on the size of the problem (see [@FalkR-1991aa]). On the other hand in the full SIPG bilinear form (without reduced integration) the space splitting discussed above is no longer orthogonal. Our approach in resolving these issues is based on a delicate estimate given in §\[subsect:cbs\] which shows a uniform bound on the angle between the Crouzeix-Raviart and its complementary space in the SIPG bilinear form for all types of boundary conditions. Once such a bound is available we show that a uniform block diagonal preconditioner can be constructed. The rest of the paper is organized as follows. We present the linear elasticity problem, the basic notation and discuss the DG discretizations considered in §\[DG-IP-discretizations\]. Next, in §\[sect:split\] we introduce the splitting of the vector valued piecewise linear DG space and discuss some properties of the related subspaces. In section §\[sect:preconditioning\], we introduce the subspace correction methods, and we prove that they give rise to a uniform preconditioner for the symmetric IP method. The last section §\[sect:numerics\] contains several numerical tests that support the theoretical results. Interior Penalty Discontinuous Galerkin methods for linear elasticity equations {#DG-IP-discretizations} =============================================================================== In this section, we introduce the linear elasticity problem together with the basic notation and the derivation of the Interior Penalty (IP) methods and we discuss the stability of these methods. Linear Elasticity: Problem formulation and notation {#sect:problem-formulation-and-notation} --------------------------------------------------- Let $\Omega\subset {{\rm I\! R}^{d}}$, $d=2,3$, be a polygon or polyhedron (not necessarily convex) and let ${\bm{u}}$ be a vector field in ${{\rm I\! R}^{d}}$, defined on $\Omega$ such that ${\bm{u}}\in [H^1(\Omega)]^d$. The elasticity tensor, which we denote by $\mathcal{C}$, is a linear operator, i.e., $\mathcal{C}: {{\rm I\! R}^{d\times d}}_{\textrm{sym}} \mapsto {{\rm I\! R}^{d\times d}}_{\textrm{sym}}$, acting on a symmetric matrix $A\in {{\rm I\! R}^{d\times d}}_{\textrm{sym}}$, in the following way: $$\mathcal{C}\;A = 2\mu A + \lambda \operatorname{trace}(A) I,$$ where $\mu $ and $\lambda$ are the Lamé parameters and satisfy $0<\mu_1<\mu<\mu_2$ and $0\leq \lambda <\infty$. In terms of the modulus of elasticity (Young’s modulus), $\mathfrak{E}$, and Poisson’s ratio, $\nu$, the Lamè parameters can be rewritten in the case of plane strain as: $\mu=\mathfrak{E}/(2(1+\nu)) $ and $\lambda=\nu\mathfrak{E}/ ((1+\nu)(1-2\nu)$. The material tends to the incompressible limit (becomes incompressible) when the Lamé parameter $\lambda \to \infty$ or equivalently when the Poisson’s ratio $\nu\to 1/2$. One can show that the linear operator $\mathcal{C}$ is selfadjoint and has two eigenvalues: (1) a simple eigenvalue equal to $(2\mu + d\lambda)$ corresponding to the identity matrix; (2) an eigenvalue equal to $2\mu$, corresponding to the $\frac{d(d+1)}{2}-1$ dimensional space of traceless, symmetric, real matrices. Thus for $d=2,3$, we always have that $$\label{eigsC} 2\mu\langle A:A\rangle \le \langle \mathcal{C}A:A\rangle \le (2\mu + d\lambda)\langle A:A\rangle ,$$ where $\langle\cdot : \cdot\rangle$ denotes the Frobenius scalar product of two tensors in ${{\rm I\! R}^{d\times d}}$. We also denote by $\langle\cdot,\cdot\rangle$ the Euclidean scalar product of two vectors in ${{\rm I\! R}^{d}}$, i.e., $$\langle {\bm{v}},{\bm{w}}\rangle = \sum_{k=1}^d v_k w_k, \qquad \langle {\bm{v}}:{\bm{w}}\rangle = \sum_{j=1}^d\sum_{k=1}^d v_{jk} w_{jk}.$$ The corresponding inner products in $[L^2({\Omega})]^d$ and $[L^2({\Omega})]^{d\times d}$ are denoted by $$({\bm{v}},{\bm{w}}) = \int_{{\Omega}}\langle{\bm{v}},{\bm{w}}\rangle, \qquad ({\bm{v}}:{\bm{w}}) = \int_{{\Omega}}\langle{\bm{v}}:{\bm{w}}\rangle .$$ We write $\partial\Omega=\Gamma_N\cup \Gamma_D$ with $\Gamma_N$ and $\Gamma_D$ referring respectively to the subsets of the $\partial {\Omega}$ where Neumann and Dirichlet boundary conditions are imposed. Let ${{\bm{\varepsilon}}({\bm{u}})}=\frac{1}{2} ( {{\boldsymbol \nabla}}{{\bm{u}}}+ ({{\boldsymbol \nabla}}{{\bm{u}}})^{T})$ be the symmetric part of the gradient of a vector valued function ${\bm{u}}$. The elasticity problem in primal formulation then is: Find ${\bm{u}} \in [H^{1+\alpha}_{\Gamma_D}(\Omega)]^d$, $\alpha>0$, which is the unique minimizer of the energy functional $\mathcal{J}({\bm{u}})$, given by $$\label{eq:minimize0} \mathcal{J}({\bm{u}}):= \frac12 ({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}:{{\bm{\varepsilon}}({\bm{u}})})- ({\bm{f}},{\bm{u}}) -({\bm{g}}_N, {\bm{v}})_{\Gamma_N}$$ Here ${\bm{f}}\in [L^2(\Omega)]^d$ is a given volume force and ${\bm{g}}_N \in [H^{3/2}(\Gamma_N)]^d$ is a given surface force acting on $\Gamma_N\subset \partial{\Omega}$. The Euler-Lagrange equations corresponding to the minimization problem  give the following well known system of linear PDEs for the unknown displacement field ${\bm{u}}$: $$\label{eq:PDEform} \begin{aligned} -{\textrm{div}}(\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}) & ={\bm{f}}, \quad &&\mbox{ on } {\Omega}, \\ (\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}){\bm{n}} & ={\bm{g}}_{N}, \quad &&\mbox{ on }\Gamma_N,\\ {\bm{u}}&={\bm{0}}, \quad && \mbox{ on } \Gamma_D. \end{aligned}$$ In the above equations, ${\bm{n}}$ is the outward unit normal vector to $\partial{\Omega}$. The solution ${\bm{u}}$ vanishes on a closed part of the boundary $\Gamma_D$ (Dirichlet boundary) and the normal stresses are prescribed on $\Gamma_N$ (Neumann part of the boundary). In the traction free case ($\Gamma_N=\partial\Omega$), the existence of a unique solution to is guaranteed if the data satisfy the following compatibility condition: $$\int_{\Omega} {\bm{f}}\cdot {\bm{v}} dx+\int_{\partial \Omega} {\bm{g}}_{N}\cdot {\bm{v}} ds =0 \quad \forall\, {\bm{v}} \in \mathbf{RM}({\Omega}),$$ where $ \mathbf{RM}({\Omega})$ is the space of rigid motions, defined by: $$\label{rm00} \mathbf{RM}({\Omega}):=\left\{ {\bm{v}}={\bm{a}}+{\bm{b}}{\bm{x}} \quad :\quad {\bm{a}}\in \mathbb{R}^{d} \quad {\bm{b}}\in \mathfrak{so}(d)\,\,\right\}$$ where ${\bm{x}}$ is the position vector function in $\Omega$ and $ \mathfrak{so}(d)$ is the Lie algebra of skew-symmetric $d\times d$ matrices. In this case, the uniqueness of solution is guaranteed up to a rigid motion (and is unique, if we require that the solution is orthogonal to any element from $\mathbf{RM}(\Omega)$). In the case of $\Gamma_D\neq \emptyset$ and closed with respect to $\partial\Omega$ no extra conditions are required to guarantee uniqueness. By considering the variational formulation of , the issue of solvability and uniqueness of the problem reduces to show coercivity of the associated bilinear form. As it is well known, for linear elasticity, this hinges on the classical Korn’s inequality [@DuvautG_LionsJL-1976aa] which guarantees the existence of a generic positive constant $C_{\Omega}>0$ such that: $$\label{korn:cont} \|\nabla {\bm{v}}\|_{0,\Omega}^{2} \leq C_{\Omega}\left( \|{{\bm{\varepsilon}}({\bm{v}})}\|_{0,\Omega}^{2} +\|{\bm{v}}\|_{0,\Omega}^{2}\right), \qquad \forall\, {\bm{v}}\in [H^{1}(\Omega)]^{d}\;.$$ The second term on the right hand side can be omitted as follows from the Poincaré or Poincaré-Friedrich’s inequality, obtaining thus first Korn’s inequality for ${\bm{v}}\in [H^{1}_{0,\Gamma_D}(\Omega)]^{d}$ and second Korn’s inequality for ${\bm{v}}\in [H^{1}(\Omega)]^{d}/\mathbf{RM}(\Omega)$. Interior penalty methods: Preliminaries and notation ---------------------------------------------------- We now introduce the basic notations and tools needed for the derivation of the DG methods. [**Domain partitioning.**]{} Let ${\mathcal{T}_h}$ be a shape-regular of partition of ${\Omega}$ into $d$-dimensional simplices ${T}$ (triangles if $d=2$ and tetrahedrons if $d=3$). We denote by ${{h}_{{T}}}$ the diameter of ${T}$ and we set ${h}=\max_{{T}\in {\mathcal{T}_h}} {{h}_{{T}}}$. We also assume that ${\mathcal{T}_h}$ is conforming in the sense that it does not contain hanging nodes. A face (shared by two neighboring elements or being part of the boundary) is denoted by $E$. Clearly, such a face is a $(d-1)$ dimensional simplex, that is, a line segment in two dimensions and a triangle in three dimensions. We denote the set of all faces by ${{\mathcal{E}_h}}$, and the collection of all interior faces and boundary faces by ${{\mathcal{E}^{o}_h}}$ and ${{\mathcal{E}^{\partial}_h}}$, respectively. Further, the set of Dirichlet faces is denoted by ${{\mathcal{E}^{D}_h}}$, and the set of Neumann faces by ${{\mathcal{E}^{N}_h}}$. We thus have, $${{\mathcal{E}_h}}={{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{\partial}_h}},\quad {{\mathcal{E}^{D}_h}}={{\mathcal{E}^{\partial}_h}}\cap \Gamma_D,\quad {{\mathcal{E}^{N}_h}}={{\mathcal{E}^{\partial}_h}}\cap \Gamma_N,\quad {{\mathcal{E}^{\partial}_h}}={{\mathcal{E}^{D}_h}}\cup {{\mathcal{E}^{N}_h}}.$$ [**Trace operators (average and jump) on $E\in{{\mathcal{E}_h}}$.**]{} \[subsect:traces\] To define the average and jump trace operators for an interior face $E\in {{\mathcal{E}^{o}_h}}$, and any ${T}\in {\mathcal{T}_h}$, such that $E\in \partial {T}$ we set ${\bm{n}}_{E,T}$ to be the unit outward (with respect to $T$) normal vector to $E$. With every face $E\in {{\mathcal{E}^{o}_h}}$ we also associate a unit vector ${\bm{n}}_E$ which is orthogonal to the $(d-1)$ dimensional affine variety (line in 2D and plane in 3D) containing the face. For the boundary faces, we always set ${\bm{n}}_E={\bm{n}}_{E,T}$, where $T$ is the *unique* element for which we have $E\subset \partial T$. In our setting, for the interior faces, the particular direction of ${\bm{n}}_E$ is not important, although it is important that this direction is fixed. For every face $E\in {{\mathcal{E}_h}}$, we define $T^{+}(E)$ and $T^{-}(E)$ as follows: $$\label{eq:t-plus-minus} \begin{array}{rcl} T^{+}(E)&{{:=}}& \{T\in {\mathcal{T}_h}\ \mbox{such that} \ E\subset \partial T, \ \mbox{and} \ \langle{\bm{n}}_{E}, {\bm{n}}_{E,T}\rangle>0\}, \\ T^{-}(E)&{{:=}}& \{T\in {\mathcal{T}_h}\ \mbox{such that} \ E\subset \partial T, \ \mbox{and} \ \langle{\bm{n}}_{E}, {\bm{n}}_{E,T}\rangle < 0\}. \end{array}$$ It is immediate to see that both sets defined above contain *no more than* one element, that is: for every face we have exactly one $T^{+}(E)$ and for the interior faces we also have exactly one $T^{-}(E)$. For the boundary faces we only have $T^{+}(E)$. In the following, we write $T^\pm$ instead of $T^\pm(E)$, when this does not cause confusion and ambiguity. For a given function ${\bm{w}}\in [L^2({\Omega})]^d$ the average and jump trace operators for a fixed $E\in {{\mathcal{E}^{o}_h}}$ are as follows: $$\label{eq:definition0} {\{\!\!\{{\bm{w}}\}\!\!\}}:=\left(\frac{{\bm{w}}^+ + {\bm{w}}^-}{2}\right),\qquad {\lbrack\!\lbrack {\bm{w}} \rbrack\!\rbrack}:= ({\bm{w}}^+ - {\bm{w}}^-),$$ where ${\bm{w}}^{+}$ and ${\bm{w}}^{-}$ denote respectively, the traces of ${\bm{w}}$ onto $E$ taken from within the interior of ${T}^{+}$ and ${T}^{-}$. On boundary faces $E\in {{\mathcal{E}^{\partial}_h}}$, we set ${\{\!\!\{{\bm{w}}\}\!\!\}}={\bm{w}}$ and ${\lbrack\!\lbrack {\bm{w}} \rbrack\!\rbrack}={\bm{w}}$. We remark that our notation differs from the one used in [@ArnoldD_BrezziF_CockburnB_MariniL-2001aa], [@ArnoldD_BrezziF_MariniL-2005aa], [@ArnoldD_BrezziF_FalkR_MariniL-2007aa] (which is considered a classical one for the IP methods). We have chosen a notation that is consistent with the one used in [@HansboP_LarsonM-2003aa], where the IP method we consider was introduced for the pure displacement problem. In addition, it seems that such a choice leads to a shorter and simpler description of the preconditioners we propose here. [**Finite Element Spaces.**]{} The piecewise linear DG space is defined by $${V^{\textrm{DG}}}:=\{u\in L^{2}({\Omega})~\mbox{such that}~u\big|_{{T}} \in \mathbb{P}^{1}({T}), \quad \forall\, {T}\in {\mathcal{T}_h}\,\},$$ where $\mathbb{P}^{1}({T})$ is the space of linear polynomials on ${T}$. The corresponding space of vector valued functions is defined as $${{\bm{V}}^{\textrm{DG}}}:=[{V^{\textrm{DG}}}]^d.$$ For a given face $E$, we denote by ${\mathcal{P}^0_{E}}:L^{2}(E)\mapsto \mathbb{P}^{0}(E)$ the $L^{2}$-projection onto the constant (vector valued or scalar valued) functions on $E$ defined by $$\begin{aligned} {\mathcal{P}^0_{E}}w = \frac{1}{|E|}\int_E w &\quad\mbox{for all}\quad w\in L^{2}(E), \label{eq:l2edge}\\ {\mathcal{P}^0_{E}}\bm{w} = \frac{1}{|E|}\int_E \bm{w} &\quad\mbox{for all}\quad \bm{w}\in [L^{2}(E)]^d. \label{eq:l2edgev}\end{aligned}$$ Observe that for $\bm{w}\in {{\bm{V}}^{\textrm{DG}}}$ the mid-point integration rule implies that ${\mathcal{P}^0_{E}}\bm{w}=\bm{w}(m_E)$ for all $E\in {{\mathcal{E}_h}}$, with $m_E$ denoting the barycenter of the edge or face $E$. The classical Crouzeix-Raviart finite element space can be defined as a subspace of $V^{DG}$, as follows: $$\label{defCRs} {V^{\textrm{CR}}}=\left\{ v\in {V^{\textrm{DG}}} \, \, \,: \quad {\mathcal{P}^0_{E}}{\lbrack\!\lbrack v \rbrack\!\rbrack}=0, \,\, \forall\, E\in {{\mathcal{E}^{o}_h}}\right\}.$$ The corresponding space of vector valued functions is $$\label{defCR} {{\bm{V}}^{\textrm{CR}}}:=[{V^{\textrm{CR}}}]^d$$ Weighted residual derivation of the IP methods ---------------------------------------------- In [@HansboP_LarsonM-2003aa] the authors introduced a symmetric interior penalty method for the problem of linear elasticity in the pure displacement case (i.e, $\Gamma_D=\partial{\Omega}, \,\,\Gamma_N=\emptyset$). We define the function space $$[H^{2}({\mathcal{T}_h})]^{d}=\left\{ {\bm{u}} \in [L^{2}({\Omega})]^{d}~\mbox{such that}~{\bm{u}}\big|_{{T}} \in [H^{2}({T})]^{d}, \quad \forall\, {T}\in {\mathcal{T}_h}\,\right\}.$$ For any pair of vector fields (or tensors) ${\bm{v}}$ and ${\bm{w}}$, we denote $${({\bm{v}},{\bm{w}})_{\mathcal{T}_h}} = \sum_{{T}\in {\mathcal{T}_h}}\int_{{T}}\langle {\bm{v}},{\bm{w}}\rangle.$$ For scalar and vector valued functions we also use the notation $$\label{eq:def-edge-product} (v,w)_{\mathcal{E}} = \sum_{E\in { \mathcal{E}}}\int_E vw, \quad\mbox{and}\quad ({{\bm{v}}},{{\bm{w}}})_{\mathcal{E}} = \sum_{E\in { \mathcal{E}}}\int_E\langle {\bm{v}},{\bm{w}}\rangle\;.$$ We now derive, using the weighted residual framework [@BrezziF_CockburnB_MariniL_SuliE-2006aa], the IP methods for the more general case of mixed boundary conditions. To present a short derivation of the methods, we assume ${\bm{u}}\in [H^{2}(\Omega)]^{d}$. Such assumption is not required for the methods to work. We present the derivation under such assumption in order to avoid unnecessary details which would shift the focus of our presentation on preconditioners. By assuming that the solution of  is a priori discontinuous, ${\bm{u}}\in [H^{2}({\mathcal{T}_h})]^{d}$, we may rewrite the continuous problem  as follows: Find ${\bm{u}}\in [H^{2}({\mathcal{T}_h})]^{d}$ such that $$\label{wr00} \left\{\begin{aligned} -{\textrm{div}}(\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}) &={\bm{f}} \quad &&\mbox{ on } {T}\in {\mathcal{T}_h}\;, \\ {\lbrack\!\lbrack (\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}){\bm{n}} \rbrack\!\rbrack}_E &={\bm{0}} \quad &&\mbox{ on } E \in {{\mathcal{E}^{o}_h}}\;,\\ {\lbrack\!\lbrack {\bm{u}} \rbrack\!\rbrack}_E &={\bm{0}} \quad &&\mbox{ on } E \in {{\mathcal{E}^{o}_h}}\;,\\ {\lbrack\!\lbrack {\bm{u}} \rbrack\!\rbrack}_E &={\bm{0}} \quad &&\mbox{ on } E \in {{\mathcal{E}^{D}_h}}\;,\\ {\lbrack\!\lbrack (\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}){\bm{n}} -{\bm{g}}_N \rbrack\!\rbrack}_E &={\bm{0}} \quad &&\mbox{ on } E \in{{\mathcal{E}^{N}_h}}\;. \end{aligned}\right.$$ where we recall that $\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})} = 2\mu {{\bm{\varepsilon}}({\bm{u}})} + \lambda \operatorname{trace}({{\bm{\varepsilon}}({\bm{u}})}) I$. Following [@BrezziF_CockburnB_MariniL_SuliE-2006aa], we next introduce a variational formulation of by considering the following five operators $$\begin{aligned} \mathcal{B}_0: [H^{2}({\mathcal{T}_h})]^{d}& {\longrightarrow}[L^{2}({\mathcal{T}_h})]^{d},& \qquad\qquad \\ \mathcal{B}_1: [H^{2}({\mathcal{T}_h})]^{d} &{\longrightarrow}[L^{2}({{\mathcal{E}^{o}_h}})]^{d},& \mathcal{B}^{\partial}_1: [H^{2}({\mathcal{T}_h})]^{d} &{\longrightarrow}[L^{2}({{\mathcal{E}^{D}_h}})]^{d} \\ \mathcal{B}_2: [H^{2}({\mathcal{T}_h})]^{d}& {\longrightarrow}[L^{2}({{\mathcal{E}^{o}_h}})]^{d}, & \mathcal{B}^{\partial}_2: [H^{2}({\mathcal{T}_h})]^{d} &{\longrightarrow}[L^{2}({{\mathcal{E}^{N}_h}})]^{d},\end{aligned}$$ and weighting each equation in appropriately. This then amounts to considering the following problem: Find ${\bm{u}}\in [H^{2}({\mathcal{T}_h})]^{d}$ such that for all ${\bm{v}}\in [H^{2}({\mathcal{T}_h})]^{d}$ $$\label{wr0} \begin{aligned} &(-{\textrm{div}}(\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})})-{\bm{f}}, \mathcal{B}_0({\bm{v}}))_{{\mathcal{T}_h}} + ( {\lbrack\!\lbrack (\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}) {\bm{n}} \rbrack\!\rbrack} , \mathcal{B}_2({\bm{v}}))_{{{\mathcal{E}^{o}_h}}} + ( {\lbrack\!\lbrack {\bm{u}} \rbrack\!\rbrack}, \mathcal{B}_1({\bm{v}}))_{{{\mathcal{E}^{o}_h}}}\qquad\qquad &&\\ &\qquad\qquad \qquad\quad+ ( {\lbrack\!\lbrack {\bm{u}} \rbrack\!\rbrack}, \mathcal{B}^{\partial}_1({\bm{v}}))_{{{\mathcal{E}^{D}_h}}}+ ( {\lbrack\!\lbrack (\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}) {\bm{n}}-{\bm{g}}_N \rbrack\!\rbrack} , \mathcal{B}^{\partial}_2({\bm{v}}))_{ {{\mathcal{E}^{N}_h}}}={\bm{0}}. && \end{aligned}$$ Different choices of the operators ${\mathcal{B}}_0$, ${\mathcal{B}}_1$, ${\mathcal{B}}_2$, ${\mathcal{B}}^{\partial}_1$ and ${\mathcal{B}}^{\partial}_2$ above give rise to different variational formulations and, consequently to different DG methods. We refer to [@BrezziF_CockburnB_MariniL_SuliE-2006aa Theorem 6] for sufficient conditions on the operators ${\mathcal{B}}_0$, ${\mathcal{B}}_1$, ${\mathcal{B}}_2$, ${\mathcal{B}}^{\partial}_1$ and ${\mathcal{B}}^{\partial}_2$ to guarantee[^1] the uniqueness of the solution of . To derive the IP method of interest, we take ${\bm{v}}$ piecewise smooth and we set $\mathcal{B}_0({\bm{v}})={\bm{v}}$, $\mathcal{B}_2({\bm{v}})={\{\!\!\{{\bm{v}}\}\!\!\}}$ and $\mathcal{B}^{\partial}_2({\bm{v}})={\bm{v}}$ in , to obtain that $$\label{eq:ortho0} \begin{array}{l} (-{\textrm{div}}(\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}), {\bm{v}} )_{{\mathcal{T}_h}} + ( {\lbrack\!\lbrack ( \mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}){\bm{n}} \rbrack\!\rbrack} , {\{\!\!\{{\bm{v}}\}\!\!\}})_{{{\mathcal{E}^{o}_h}}\cup {{\mathcal{E}^{D}_h}}}+ ({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack}, \mathcal{B}_1({\bm{v}}))_{{{\mathcal{E}^{o}_h}}\cup {{\mathcal{E}^{D}_h}}}\\ ~~~= ({\bm{f}},{\bm{v}})_{{\mathcal{T}_h}}+ ( {\bm{g}}_N ,{\bm{v}} )_{{{\mathcal{E}^{N}_h}}}\;. \end{array}$$ Defining $$\label{rhs00} \mathcal{F}({\bm{v}})=({\bm{f}},{\bm{v}})_{{\mathcal{T}_h}}+( {\lbrack\!\lbrack{\bm{g}}\rbrack\!\rbrack} , \mathcal{B}^{\partial}_1({\bm{v}}) )_{{{\mathcal{E}^{D}_h}}} +( {\bm{g}}_N ,{\bm{v}} )_{{{\mathcal{E}^{N}_h}}},$$ and integrating by parts the first term on the left side of  then leads to $$\label{eq:a-bilinear-form} {(\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}:{{\bm{\varepsilon}}({\bm{v}})})_{\mathcal{T}_h}} - ( {\{\!\!\{(\mathcal{C}{{\bm{\varepsilon}}({\bm{u}})}){\bm{n}}\}\!\!\}} , {\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack} )_{{{\mathcal{E}^{o}_h}}\cup {{\mathcal{E}^{D}_h}}}+ ({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},\mathcal{B}_1({\bm{v}})) _{{{\mathcal{E}^{o}_h}}\cup {{\mathcal{E}^{D}_h}}} = \mathcal{F}({\bm{v}}).$$ For a fixed edge $E\in{{\mathcal{E}^{o}_h}}\cup {{\mathcal{E}^{D}_h}}$ the operator $ \mathcal{B}_1({\bm{v}})$ is defined by $$\label{defB1} \mathcal{B}_1({\bm{v}}):= -{\{\!\!\{(\mathcal{C}{{\bm{\varepsilon}}({\bm{v}})}) {\bm{n}}\}\!\!\}} + \alpha_0\beta_0 {\mathcal{P}^0_{E}}{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}+\alpha_1\beta_1{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack},$$ where, following [@HansboP_LarsonM-2003aa], the parameters $\beta_0$ and $\beta_1$ are chosen depending on the Lamé constants $\lambda$ and $\mu$: $$\label{defbeta} \beta_0:=d\lambda +2\mu, \qquad \quad \beta_1:=2\mu\;.$$ The remaining two parameters, $\alpha_0$ and $\alpha_1$, are still at our disposal to ensure (later on) stability and to avoid locking of the resulting method. We define $$\label{eq:aj0aj1} \begin{aligned} a_{j,0}({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})&{{:=}}\alpha_0\beta_0 \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \int_E\langle h_E^{-1}{\lbrack\!\lbrack {\bm{u}} \rbrack\!\rbrack}, {\mathcal{P}^0_{E}}{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}\rangle,\\ a_{j,1}({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})&{{:=}}\alpha_1\beta_1 \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \int_E\langle h_E^{-1}{\lbrack\!\lbrack {\bm{u}} \rbrack\!\rbrack},{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\rangle\;, \end{aligned}$$ and set $$a_j({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})=a_{j,0}({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})+ a_{j,1}({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}).$$ Then, the weak formulation of Problem  reads: Find ${\bm{u}}\in [H^{2}({\mathcal{T}_h})]^{d}$ such that $$\label{eq:weakform-ok} { \mathcal{A}}({\bm{u}},{\bm{w}})=\mathcal{F}({\bm{w}}),\qquad \forall \,\,{\bm{w}}\in [H^{2}({\mathcal{T}_h})]^{d}.$$ The bilinear form ${ \mathcal{A}}(\cdot,\cdot)$ is given by $$\label{ipAstep} { \mathcal{A}}({\bm{u}},{\bm{w}})={ \mathcal{A}}_0({\bm{u}},{\bm{w}})+ a_{j,1}({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{w}}\rbrack\!\rbrack}),$$ where $$\label{ipA0} \begin{array}{rcl} { \mathcal{A}}_0({\bm{u}},{\bm{w}})& =& {({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}:{{\bm{\varepsilon}}({\bm{w}})})_{\mathcal{T}_h}}- ({\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}){\bm{n}}\}\!\!\}},{\lbrack\!\lbrack{\bm{w}}\rbrack\!\rbrack})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\\ && - ( {\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{w}})}){\bm{n}}\}\!\!\}})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} + a_{j,0}({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{w}}\rbrack\!\rbrack}). \end{array}$$ It is straightforward to see that $$\label{ipA} \begin{array}{rcl} { \mathcal{A}}({\bm{u}},{\bm{w}}) &=& {({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}:{{\bm{\varepsilon}}({\bm{w}})})_{\mathcal{T}_h}}- ({\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}){\bm{n}}\}\!\!\}},{\lbrack\!\lbrack{\bm{w}}\rbrack\!\rbrack})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\\ && +\theta ({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{w}})}){\bm{n}}\}\!\!\}})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} + a_{j}({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{w}}\rbrack\!\rbrack}). \end{array}$$ To obtain the discrete formulation, we replace the function space $[H^{2}({\mathcal{T}_h})]^{d}$ in   by ${{\bm{V}}^{\textrm{DG}}}$, and we get the [**IP-1 approximation**]{} to the problem: Find ${{\bm{u}}_h}\in {{\bm{V}}^{\textrm{DG}}}$ such that: $$\label{eq:DG-formulation} { \mathcal{A}}({{\bm{u}}_h},{\bm{w}})=\mathcal{F}({\bm{w}}),\qquad \forall \,\,{\bm{w}}\in {{\bm{V}}^{\textrm{DG}}}.$$ We could also consider the approximation given by the [**IP-0 method**]{}: Find ${{\bm{u}}_h}\in {{\bm{V}}^{\textrm{DG}}}$ such that: $$\label{eq:DG-formulation0} { \mathcal{A}}_0({{\bm{u}}_h},{\bm{w}})=\mathcal{F}({\bm{w}}),\qquad \forall \,\,{\bm{w}}\in {{\bm{V}}^{\textrm{DG}}}.$$ As we see next, the [**IP-0**]{} method provides a robust approximation to the problem in the pure displacement problem $\Gamma_D=\partial\Omega$. As we mentioned earlier, for other types of boundary conditions such equivalence in general does not hold. \[theta\] Although we do not consider non-symmetric IP methods in this paper, let us remark that non-symmetric versions can be easily incorporated in the definition of $\mathcal{B}_1({\bm{v}})$. For example, by setting: $$\mathcal{B}_1({\bm{v}}):= \theta{\{\!\!\{(\mathcal{C}{{\bm{\varepsilon}}({\bm{v}})}) {\bm{n}}\}\!\!\}} + \alpha_0\beta_0 {\mathcal{P}^0_{E}}{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}+\alpha_1\beta_1{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack},$$ we obtain a non-symmetric bilinear form for the values $\theta=0$ or $\theta=1$. Such values of $\theta$ correspond to the Incomplete Interior Penalty (IIPG, $\theta=0$) and Non-symmetric Interior Penalty (NIPG, $\theta=1$) discretizations, respectively. Stability Analysis\[sec:2\] --------------------------- We close this section presenting the stability and continuity results pertinent to our work. We start by introducing some norm notation. For ${\bm{v}}\in [H^2({\mathcal{T}_h})]^d$ we define the semi-norms $$\label{semi-norm} \begin{aligned} &\|\nabla{\bm{v}}\|_{0,{\mathcal{T}_h}}^{2}=\sum_{{T}\in {\mathcal{T}_h}}\|\nabla {\bm{v}}\|_{0,{T}}^{2} \qquad && \|\mathcal{C}^{1/2}{{{\bm{\varepsilon}}({\bm{v}})}}\|_{0,{\mathcal{T}_h}}^{2}=\sum_{T\in\mathcal{T}_h}\int_{T}\langle{\mathcal{C}{\bm{\varepsilon}}({\bm{v}})}:{{{\bm{\varepsilon}}({\bm{v}})}}\rangle \\ & | { \mathcal{P}}^{0}_E{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}|^{2}_{\ast}=\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}h_E^{-1} \|{ \mathcal{P}}^{0}_E{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}\|_{0,E}^{2} \qquad &&|{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}|^{2}_{\ast}=\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}h_E^{-1} \|{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}\|_{0,E}^{2}\;, \end{aligned}$$ and norms: $$\label{norm_h} \|{\bm{v}}\|_h^2 = \|\mathcal{C}^{1/2}{{{\bm{\varepsilon}}({\bm{v}})}}\|_{0,{\mathcal{T}_h}}^{2} + \beta_0| { \mathcal{P}}^{0}_E{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}\|_{\ast}^{2} +\beta_1|{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}|^{2}_{\ast}+ \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} h_E\|{\bf \mathcal{C}}^{1/2} {{\bm{\varepsilon}}({\bm{v}})}\cdot {{\bf n}}\|_{0,E}^{2}\;.$$ For ${\bm{v}}\in {{\bm{V}}^{\textrm{DG}}}$ we define the norms $$\label{eq:Me_D_norm} \|{\bm{v}}\|_{DG0}^2= \|\mathcal{C}^{1/2}{{{\bm{\varepsilon}}({\bm{v}})}}\|_{0,{\mathcal{T}_h}}^{2} +\beta_0 | { \mathcal{P}}^{0}_E{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}|^{2}_{\ast}$$ and $$\label{normDG} \|{\bm{v}}\|_{DG}^2 =\|{\bm{v}}\|_{DG0}^2+ \beta_1 |{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}|^{2}_{\ast}\;.$$ Notice that for ${\bm{v}}\in {{\bm{V}}^{\textrm{DG}}}$ the norms and are equivalent. We finally introduce the norm: $$\label{fullnorm} \|{\bm{v}}\|_{H^{1}({\mathcal{T}_h})}^2 =\|\nabla {\bm{v}}\|_{0,{\mathcal{T}_h}}^{2} + \beta_0| { \mathcal{P}}^{0}_E{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}\|_{\ast}^{2}++ \beta_1| {\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}\|_{\ast}^{2}\;.$$ Notice that continuity of the [**IP-1**]{} and [**IP-0**]{} bilinear forms with respect to the norm follows easily from Cauchy-Schwarz inequality together with the bound on the maximum eigenvalue of $\mathcal{C}$, i.e., for all ${\bm{u}} \in [H^2({\mathcal{T}_h})]^d$ and all ${\bm{v}} \in {{\bm{V}}^{\textrm{DG}}}$ we have $$\begin{aligned} ({\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}{\bm{n}})\}\!\!\}},{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}&=& ({\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}{\bm{n}})\}\!\!\}},{ \mathcal{P}}^{0}_E{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\\ & \leq& \frac{1}{\alpha_0\beta_0} h^{1/2}_E\|{\bf \mathcal{C}} {{\bm{\varepsilon}}({\bm{u}})}\cdot {{\bf n}}\|_{0,{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \cdot \frac{ \alpha_0\beta_0}{4} \|h_{E}^{-1/2}{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\|_{0,{{\mathcal{E}^{o}_h}}\cup\Gamma_D} \\ & \leq & \frac{1}{\alpha_0} \|h_{E}^{1/2}\mathcal{C}^{1/2}{{\bm{\varepsilon}}({\bm{u}})}\cdot {{\bf n}}\|_{0,{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \frac{ \alpha_0\beta_0}{4} \|h_{E}^{-1/2}{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\|_{0,{{\mathcal{E}^{o}_h}}\cup\Gamma_D}.\end{aligned}$$ The equivalence of the norms and for any ${\bm{v}}\in {{\bm{V}}^{\textrm{DG}}}$ guarantees therefore the continuity of the [**IP-1**]{} bilinear form with respect to the norm defined in for finite element functions. The solvability of the discrete methods and is guaranteed if and only if, a discrete version of the Korn’s inequality holds on ${{\bm{V}}^{\textrm{DG}}}$. In [@BrennerS-2004aa] the following discrete Korn inequality is shown for $[H^{1}({\mathcal{T}_h})]^{d}$-vector fields: $$\label{korn:disc} \|{\bm{\nabla}} {\bm{v}}\|_{0,{\mathcal{T}_h}}^{2} \leq C \left( \|{{\bm{\varepsilon}}({\bm{v}})}\|_{0,{\mathcal{T}_h}}^{2} + |\pi_1 {\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}|_{\ast}^{2} +\|{\bm{\nabla}} \times {\bm{v}}\|_{0,{\mathcal{T}_h}}^{2} \right)$$ where $\pi_1 : [L^{2}({{\mathcal{E}_h}})]^{d}{\longrightarrow}{\bm{\mathbb{P}}}^{1}({{\mathcal{E}_h}})$ is the $L^{2}$-orthogonal projection onto the space of piecewise linear vector valued functions on ${{\mathcal{E}_h}}$ (or a subset of it). Coercivity of the [**IP-1**]{} bilinear form with respect to the norm can be easily shown by taking ${\bm{u}}={\bm{w}}={\bm{v}}$ in : $$\begin{aligned} { \mathcal{A}}({\bm{v}},{\bm{v}}) &=( {\mathcal{C}{\bm{\varepsilon}}({\bm{{\bm{v}}}})} : {{\bm{\varepsilon}}({\bm{v}})} )_{{\mathcal{T}_h}}+ \alpha_0 \beta_0 \|h_{E}^{-1/2}{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\|_{0,{{\mathcal{E}^{o}_h}}\cup\Gamma_D}^{2} + \alpha_1\beta_1\|h_{E}^{-1/2}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\|_{0,{{\mathcal{E}^{o}_h}}\cup\Gamma_D}^{2} &&\\ &\qquad \quad -2 ({\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{v}})}{\bm{n}})\}\!\!\}}, {\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\;.\end{aligned}$$ Using Cauchy-Schwarz, trace and inverse inequalities together with the arithmetic-geometric inequality and the bound on the maximum eigenvalue of $\mathcal{C}$ it follows that $$\begin{aligned} \label{bella0} ({\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{v}})}{\bm{n}})\}\!\!\}},{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}&=& ({\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{v}})}{\bm{n}})\}\!\!\}},{ \mathcal{P}}^{0}_E{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \nonumber \\ & \leq& \frac{C_{t}(1+C_{inv})}{\alpha_0\beta_0} \|\mathcal{C}{{\bm{\varepsilon}}({\bm{v}})}\|_{0,{\mathcal{T}_h}}^{2} + \frac{ \alpha_0\beta_0}{4} \|h_{E}^{-1/2}{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\|_{0,{{\mathcal{E}^{o}_h}}\cup\Gamma_D}^{2} \nonumber \\ & \leq & \frac{C_{t}(1+C_{inv})}{\alpha_0} \|\mathcal{C}^{1/2}{{\bm{\varepsilon}}({\bm{v}})}\|_{0,{\mathcal{T}_h}}^{2} + \frac{ \alpha_0\beta_0}{4} \|h_{E}^{-1/2}{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\|_{0,{{\mathcal{E}^{o}_h}}\cup\Gamma_D}^{2}.\end{aligned}$$ Hence, we finally have $$\begin{aligned} { \mathcal{A}}({\bm{v}},{\bm{v}})&\geq& (1-\frac{2C_{t}(1+C_{inv})}{\alpha_0} ) \|\mathcal{C}^{1/2} {{\bm{\varepsilon}}({\bm{v}})}\|_{0,{\mathcal{T}_h}}^{2} + \alpha_1\beta_1\|h_{E}^{-1/2}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\|_{0,{{\mathcal{E}^{o}_h}}\cup\Gamma_D}^{2}\\ &&~~~~+ \frac{\alpha_0}{2} \beta_0 \|h_{E}^{-1/2}{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\|_{0,{{\mathcal{E}^{o}_h}}\cup\Gamma_D}^{2}, \quad \forall\, {\bm{v}} \in {{\bm{V}}^{\textrm{DG}}}\;,\end{aligned}$$ and therefore by taking $\alpha_{0}=\max{(1,4C_{t}(1+C_{inv}))}$ (sufficiently large) we ensure the coercivity of ${ \mathcal{A}}(\cdot,\cdot)$ with respect to the $\|\cdot\|_{DG}$-norm with constant independent of $h$, $\mu$, and $\lambda$. Using now (since the norm contains the full jump) we conclude that ${ \mathcal{A}}(\cdot,\cdot)$ is coercive with respect to the $\|\cdot\|_{H^{1}({\mathcal{T}_h})}$-norm . Therefore the [ **IP-1**]{} method defined by provides a robust approximation to and does not lock as $\lambda \rightarrow \infty$. As we mentioned earlier, in the pure displacement case ($\Gamma_D=\partial{\Omega}, \, \Gamma_N=\emptyset$) the bilinear form ${ \mathcal{A}}_0(\cdot\cdot)$ defined in is coercive. Indeed we may use the identity (which holds for $C_0^{\infty}(\Omega)$ functions): $${\textrm{div}}{{\bm{\varepsilon}}({\bm{v}})}=\frac{1}{2}\left( {\textrm{div}}\nabla {\bm{v}} +\nabla {\textrm{div}}{\bm{v}} \right)$$ and rewrite the volume term in (also in ) as follows: $$\begin{aligned} {({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}:{{\bm{\varepsilon}}({\bm{w}})})_{\mathcal{T}_h}}&=& \sum_{T\in\mathcal{T}_h}\int_{T}\langle{\mathcal{C}{\bm{\varepsilon}}({\bm{v}})}:{{{\bm{\varepsilon}}({\bm{v}})}}\rangle\\ &=& \sum_{T\in\mathcal{T}_h}\int_{T} \left( 2\mu \langle \nabla {\bm{u}}:\nabla {\bm{v}}\rangle + (\mu +\lambda) \langle {\textrm{div}}{\bm{u}}, {\textrm{div}}{\bm{v}}\rangle \right).\end{aligned}$$ Then, from the discrete Poincaré inequality [@FengX_KarakashianO-2001aa; @BrennerS-2003aa], the resulting modified bilinear form for ${ \mathcal{A}}_0(\cdot,\cdot)$ is now coercive in ${\bm{V}}^{DG}$ with respect to the $\|\cdot\|_{H^{1}({\mathcal{T}_h})}$ norm, with coercivity constant independent of $h$ and $\lambda$; $$\label{st:A00} { \mathcal{A}}_0({\bm{v}},{\bm{v}}) \geq C \|{\bm{v}}\|_{H^{1}({\mathcal{T}_h})}^{2} \quad \forall\, {\bm{v}}\in {\bm{V}}^{DG}\;.$$ Therefore, the discrete problem is well posed and the [**IP-0**]{} method is stable and robust (locking free in the limit $\lambda \to \infty$). Notice that in we are using the $\|\cdot \|_{H^{1}({\mathcal{T}_h})}$-norm which includes not only the norm $|P^{0}_E{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}|_{\ast}$, but also the norm $|{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}|_{\ast}$. This is a consequence of the vector valued counterpart of [@Ayuso-de-DiosB_ZikatanovL-2009aa Lemma 2.3]. The stability property given in implies that the [ **IP-0**]{} and [**IP-1**]{} methods are spectrally equivalent for the pure displacement problem. These observations are summarized in the next Lemma: \[le:A0A1\] Let $\mathcal{A}(\cdot,\cdot)$ and ${ \mathcal{A}}_0(\cdot,\cdot)$ be the bilinear forms of the [**IP-1**]{} and [**IP-0**]{} methods for the linear elasticity problem, defined in and , respectively. For the pure displacement problem $\Gamma_D=\partial{\Omega}, \,\, \Gamma_N=\emptyset$, there exist a constant $c>0$ that depends only on the geometry of the domain ${\Omega}$ but is independent of the mesh size and the Lamé parameters $\mu$ and $\lambda$ such that $$\label{A0A1:0} { \mathcal{A}}_0({\bm{v}},{\bm{v}})\leq { \mathcal{A}}({\bm{v}},{\bm{v}})\leq c{ \mathcal{A}}_0({\bm{v}},{\bm{v}}) \quad \forall\, {\bm{v}} \in {\bm{V}}^{DG}\;.$$ The above lemma guarantees that for the pure displacement problem, constructing a uniform preconditioner for the [**IP-1**]{} is equivalent to constructing a uniform preconditioner for the [ **IP-0**]{} method (see [@Ayuso-de-DiosB_ZikatanovL-2009aa]). For linear elasticity equations, unlike for scalar equations, this can be done only when $\Gamma_D=\partial\Omega$. For a detailed derivation and error estimates, we refer to [@HansboP_LarsonM-2003aa Theorem 2.5]. Space decomposition {#sect:split} =================== We present now a decomposition of the DG space of piecewise linear vector valued functions that plays a key role in the construction of iterative solvers. This decomposition was introduced in [@Ayuso-de-DiosB_ZikatanovL-2009aa] for scalar functions and also in [@BurmanE_StammB-2008aa] in a different context. Its extension to vector valued functions is more or less straightforward. We omit those proofs which are just an easy modification of the corresponding proofs in the scalar case. However, we review the main ingredients and ideas behind such proofs, since they play an important role in the analysis of the preconditioner given later on. In the last part of the section we give some properties of the spaces entering in the and prove a result that is essential for showing that the proposed preconditioner is uniform. Following [@Ayuso-de-DiosB_ZikatanovL-2009aa] we introduce the space complementary to ${V^{\textrm{CR}}}$ in ${V^{\textrm{DG}}}$, $$\label{defZ} {{ \mathcal{Z}}^{\textrm{}}}=\left\{ z\in \, {V^{\textrm{DG}}}~\mbox{and}~{ \mathcal{P}}_{E}^{0} {\{\!\!\{z\}\!\!\}}=0,~\mbox{for all}~E\in {{\mathcal{E}^{o}_h}}\right\}.$$ The corresponding space of vector valued functions is $$\label{defZ2} {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}=[{{ \mathcal{Z}}^{\textrm{}}}]^d.$$ To describe the basis functions associated with the spaces and , let $\varphi_{E ,{T}}$ denote the scalar basis function on ${T}$, dual to the degree of freedom at the mass center of the face $E$, and extended by zero outside ${T}$. For $E\in\partial{T}$, $E^{\prime}\in \partial {T}$, the function $\varphi_{E,T}$ satisfies $$\varphi_{E,{T}}(m_{E'}) = \left\{ \begin{array}{ll} 1 & \quad \mbox{if }\quad E=E^{\prime},\\ 0 &\quad \mbox{otherwise,} \end{array} \right.$$ and also we have $$\varphi_{E,{T}}\in \mathbb{P}^{1}({T}), \quad \varphi_{E,{T}}(x)=0, \forall\, x\notin {T}.$$ For all ${\bm{u}}\in {{\bm{V}}^{\textrm{DG}}}$ we then have $$\label{eq:represent0} {\bm{u}}(x) = \sum_{{T}\in {\mathcal{T}_h}}\sum_{E\in \partial{T}}{\bm{u}}_{T}(m_E)\varphi_{E,{T}}(x) =\sum_{E\in {{\mathcal{E}_h}}} {\bm{u}}^{+}(m_E)\varphi_{E}^{+}(x)+\sum_{E\in {{\mathcal{E}^{o}_h}}} {\bm{u}}^{-}(m_E)\varphi_{E}^{-}(x),$$ where in the last identity we have just changed the order of summation and used the short hand notation $\varphi_{E}^{\pm}(x):=\varphi_{E,{T}^{\pm}}(x)$ together with $$\begin{aligned} {\bm{u}}^{\pm}(m_{E})&:={\bm{u}}_{{T}^{\pm}}(m_E)=\frac{1}{|E|}\int_{E} {\bm{u}}_{{T}^{\pm}} ds, \quad &\forall\, E\in {{\mathcal{E}^{o}_h}}, \,\,:\,\, E=\partial{T}^{+}\cap \partial {T}^{-}, &&\\ {\bm{u}}(m_{E})&:={\bm{u}}_{{T}}(m_E)=\frac{1}{|E|}\int_{E} {\bm{u}}_{{T}} ds, \quad &\forall\, E\in {{\mathcal{E}^{\partial}_h}},~\mbox{such that}~E=\partial{T}\cap \partial {\Omega}. && \end{aligned}$$ Recalling now the definitions of $T^{+}(E)$ and $T^{-}(E)$ given in  we set $$\label{basis-CR} \begin{aligned} ~& \varphi_{E}^{CR}=\varphi_{E,{T}^{+}(E)}+\varphi_{E,{T}^{-}(E)}, &\quad \forall\, E\in {{\mathcal{E}^{o}_h}}, &&\\ ~& \varphi_{E}^{CR}=\varphi_{E,{T}^{+}(E)}, &\quad \forall\, E\in {{\mathcal{E}^{N}_h}}. && \end{aligned}$$ and $$\label{basis-Z} \begin{aligned} ~& \psi_{E}^{z}=\frac{\varphi_{E,{T}^{+}(E)}-\varphi_{E,{T}^{-}(E)}}{2}, &\quad \forall\, E\in {{\mathcal{E}^{o}_h}}, &&\\ ~& \psi_{E}^{z}=\varphi_{E,{T}^{+}(E)}, &\quad \forall\, E\in {{\mathcal{E}^{D}_h}}. \end{aligned}$$ ![Basis functions associated with the face $E$: $\psi_{E}^{z}$ (left) and $\varphi_{E}^{CR}$ (right). \[fig:ZCR\]](Z "fig:"){height="2in" width="40.00000%"} ![Basis functions associated with the face $E$: $\psi_{E}^{z}$ (left) and $\varphi_{E}^{CR}$ (right). \[fig:ZCR\]](CR "fig:"){height="2in" width="40.00000%"} Some clarification is needed here. Note that from the definition of $\varphi_{E,{T}^{+}(E)}$ and $\varphi_{E,{T}^{-}(E)}$ for an interior edge $E\in {{\mathcal{E}^{o}_h}}$, it does not follow that their sum is even defined on the edge $E$, since it is just a sum of two functions from $L^2({\Omega})$. However, the sum $(\varphi_{E,{T}^{+}(E)}+\varphi_{E,{T}^{-}(E)})$ has a representative, which is continuous across $E$ and this representative is denoted here with ${\bm{ \varphi}}_{E}^{CR}$, see Figure \[fig:ZCR\]. Clearly, $\{\varphi_{E}^{CR}\}_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{N}_h}}}$ are linearly independent, and $\{{\bm{\psi}}^{z}_{E}\}_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}$ are linearly independent. A simple argument then shows that $${{\bm{V}}^{\textrm{CR}}} = \operatorname{span} \left\{ \{\varphi_{E}^{CR}{\bm{e}}_k\}_{k=1}^d\right\}_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{N}_h}}}, \quad {{\bm{{ \mathcal{Z}}}}^{\textrm{}}} = \operatorname{span} \left\{ \{\psi_{E}^{z}{\bm{e}}_k\}_{k=1}^d\right\}_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}.$$ Here ${\bm{e}}_k$, $k=1,\ldots,d$ is the $k$-th canonical basis vector in ${{\rm I\! R}^{d}}$. Hence by performing a change of basis in , we have obtained a “natural” splitting of $${{\bm{V}}^{\textrm{DG}}}={{\bm{V}}^{\textrm{CR}}}\oplus {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$$ and the set $$\label{nat_basis} \{{\bm{\psi}}^{z}_{E}\}_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \cup \{\varphi_{E}^{CR}\}_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{N}_h}}},$$ provides a natural basis for the DG finite element space. This is summarized in the next proposition. \[split:0\] For any ${\bm{u}}\in {{\bm{V}}^{\textrm{DG}}}$ there exist unique ${\bm{v}}\in{{\bm{V}}^{\textrm{CR}}}$ and a unique ${\bm{z}}\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ such that $$\label{eq:decomposition} \begin{array}{ll} {\bm{u}}={\bm{v}}+{\bm{z}}\quad \mbox{and}\quad& \begin{array}{rcl} {\bm{v}}&=&\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{N}_h}}} \left(\frac{1}{|E|}\int_{E}{\{\!\!\{{\bm{u}}\}\!\!\}}ds\right)\varphi^{CR}_{E}(x) \in {{\bm{V}}^{\textrm{CR}}},\\ {\bm{z}}&=&\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \left(\frac{1}{|E|}\int_{E}{\lbrack\!\lbrack {\bm{u}} \rbrack\!\rbrack}ds \right){\bm{\psi}}^{z}_{E}(x) \in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}. \end{array} \end{array}$$ The proof of the above result follows by arguing as for the scalar case in [@Ayuso-de-DiosB_ZikatanovL-2009aa Proposition 3.1], but proceeding componentwise. The next Lemma shows that the splitting we have proposed is orthogonal with respect to the inner product defined by $\mathcal{A}_0(\cdot,\cdot)$. \[le:orto\] The splitting ${{\bm{V}}^{\textrm{DG}}}={{\bm{V}}^{\textrm{CR}}}\oplus {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ is ${ \mathcal{A}}_{0}$-orthogonal. That is $$\label{orto:1} { \mathcal{A}}_0({\bm{v}},{\bm{z}})={ \mathcal{A}}_0({\bm{z}},{\bm{v}})=0 \qquad \forall\, {\bm{v}} \in {{\bm{V}}^{\textrm{CR}}},\quad \forall\, {\bm{z}} \in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}} .$$ The proof follows straightforwardly by using the weighted residual formulation - and the definition of the spaces $ {{\bm{V}}^{\textrm{CR}}}$ and ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$. Some properties of the space ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$\[subsect:cbs\] --------------------------------------------------------------------------------- We now present some properties of the functions in the space ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$. We start with a simple observation. From the definition of the spaces ${{\bm{V}}^{\textrm{CR}}}$ and ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ it is easy to see that $$\sum_{{T}\in {\mathcal{T}_h}}\|\nabla {\bm{z}}\|_{0,{T}}^2=({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack},{\{\!\!\{\nabla {\bm{z}}\}\!\!\}})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}.$$ Applying the Schwarz inequality, one then gets the following estimate $$\sum_{{T}\in {\mathcal{T}_h}}\|\nabla {\bm{z}}\|^{2}_{0,{T}}\leq C\|h^{-1/2}P^{0}_E{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}\|^{2}_{0,{{\mathcal{E}_h}}},$$ which is a straightforward way to see that the restriction of the [**IP-1**]{} and [**IP-0**]{}-bilinear forms (even for $\theta=0,1$ as in Remark \[theta\]) to the space ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ are coercive in the $\|\cdot\|_{H^{1}({\mathcal{T}_h})}$-norm (regardless whether the boundary conditions are Dirichlet, Neumann or mixed type). Therefore the resulting stiffness matrices are positive definite. The next result provides bounds on the eigenvalues of ${ \mathcal{A}}_0(\cdot,\cdot)$ and ${ \mathcal{A}}(\cdot,\cdot)$, when restricted to ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$. \[le:poincZ\] Let ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ be the space defined in . Then for all ${\bm{z}}\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$, the following estimates hold $$\label{poincZ:0} h^{-2}\|{\bm{z}}\|_{0}^{2}\lesssim { \mathcal{A}}_{0}({\bm{z}},{\bm{z}})\lesssim h^{-2}\|{\bm{z}}\|_{0}^{2}\;,$$ and also, $$\label{poincZ:1} [(\alpha_0)\beta_0 +\alpha_1\beta_1] h^{-2}\|{\bm{z}}\|_{0}^{2}\lesssim { \mathcal{A}}({\bm{z}},{\bm{z}})\lesssim [\alpha_0\beta_0 +\alpha_1\beta_1] h^{-2}\|{\bm{z}}\|_{0}^{2}\;,$$ where $\beta_0$ and $\beta_1$ are as defined in . Arguing as in [@Ayuso-de-DiosB_ZikatanovL-2009aa Lemma 5.3] (but now componentwise for vector valued functions) one can show that (due the special structure of the space ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$). $$\label{poinc:00} h^{-2}\|{\bm{z}}\|_{0}^{2}\lesssim \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} h_{E}^{-1}\|{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}\|^{2}_{0,E} \lesssim h^{-2}\|{\bm{z}}\|_{0}^{2}\;.$$ From the coercivity of ${ \mathcal{A}}_0$ it follows then $$\alpha_0\beta_0 h^{-2}\|{\bm{z}}\|_{0}^{2}\lesssim \alpha_0\beta_0\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} h_{E}^{-1}\|{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}\|^{2}_{0,E} \leq { \mathcal{A}}_0({\bm{z}},{\bm{z}})\;.$$ Similarly, the $L^{2}({{\mathcal{E}_h}})$ stability of the projection ${ \mathcal{P}}_E^{0}$ together with the coercivity of ${ \mathcal{A}}$ gives $$\begin{aligned} ( \alpha_0\beta_0 +\alpha_1\beta_1) h^{-2}\|{\bm{z}}\|_{0}^{2}&\lesssim \alpha_0\beta_0\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} h_{E}^{-1}\|{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}\|^{2}_{0,E} + \alpha_1\beta_1\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} h_{E}^{-1}\|{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}\|^{2}_{0,E} &&\\ &\lesssim \alpha_0\beta_0\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} h_{E}^{-1}\|{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}\|^{2}_{0,E} + C\alpha_1\beta_1\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} h_{E}^{-1}\|{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}\|^{2}_{0,E} &&\\ & \leq { \mathcal{A}}({\bm{z}},{\bm{z}}), && \end{aligned}$$ and so, the lower bounds in and follow. We next show the upper bound in , and the upper bound in is obtained in an analogous fashion. Using together with we get $$\begin{aligned} { \mathcal{A}}_0({\bm{z}},{\bm{z}}) &\leq \alpha_0 \beta_0\sum_{E\in {{\mathcal{E}^{o}_h}}\cup \Gamma_D} h_{E}^{-1}\|{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}\|^{2}_{0,E} + \|\mathcal{C}^{1/2}{{\bm{\varepsilon}}({\bm{z}})}\|_{0,{\mathcal{T}_h}}^{2} &&\\ &\leq \beta_0 \left(\alpha_0 \|h_{E}^{-1/2}{ \mathcal{P}}^{0}_{E}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}\|^{2}_{{{\mathcal{E}^{o}_h}}\cup \Gamma_D} +C\|{{\bm{\varepsilon}}({\bm{z}})}\|_{0,{\mathcal{T}_h}}^{2}\right)\;.\end{aligned}$$ Hence, the upper bound in follows in a straightforward fashion using the trace and inverse inequalities together with the obvious inequality $\|{{\bm{\varepsilon}}({\bm{z}})}\|_{0,{\mathcal{T}_h}}\leq \|{{\boldsymbol \nabla}}{\bm{z}}\|_{0,{\mathcal{T}_h}}$. We close this section with establishing a uniform bound on the angle between ${{\bm{V}}^{\textrm{CR}}}$ and ${{\bm{Z}}^{\textrm{}}}$ in the inner product given by the bilinear form ${ \mathcal{A}}(\cdot,\cdot)$. The estimate is given in Proposition \[aux\_aux\_lema\]. It plays a crucial role in bounding the condition number of the preconditioned system. We remind that $E \in{{\mathcal{E}_h}}$ denotes a $(d-1)$-dimensional simplex (a face), which is either the intersection of two $d$-dimensional simplices $T\in {\mathcal{T}_h}$ or an intersection of a $d$-dimensional simplex $T\in{\mathcal{T}_h}$ and the complement of $\Omega$, i.e., $E=T\cap ({{\rm I\! R}^{d}}\setminus\Omega)$. In the former case, the face $E$ is called an interior face and in the latter it is called a boundary face. The proof of Proposition \[aux\_aux\_lema\] requires arguments involving the incidence relations between simplices $T\in{\mathcal{T}_h}$ and faces $E\in {{\mathcal{E}_h}}$, and estimates on the cardinality of these incidence sets. For the readers’ convenience, we provide a list of such estimates below. - We define $\mathcal{N}_0(E)$ to be the set of $d$-dimensional $T\in {\mathcal{T}_h}$ simplices that contain $E$: $$\mathcal{N}_0(E):=\{T\in {\mathcal{T}_h},\quad\mbox{such that}\quad E\in T\}$$ By definition, for the cardinality of this set we have $|\mathcal{N}_0(E)|=2$ for the interior faces and $|\mathcal{N}_0(E)|=1$ for the boundary faces. - We define the set of neighbor (or neighboring) faces $\mathcal{N}_1(E)$ to be the set of faces which share an element with $E$: $$\mathcal{N}_1(E):=\{E'\in {{\mathcal{E}_h}},\quad\mbox{such that}\quad \mathcal{N}_0(E)\cap\mathcal{N}_0(E')\neq \emptyset\}$$ From Proposition \[prop:cardinal\] (see Appendix \[ap0\]) we have that $|\mathcal{N}_1(E)|\le (2d+1)$. - Next, we define $\mathcal{N}_2(E)$ to be the set of faces which share at least one neighboring face with $E$: $$\mathcal{N}_2(E):=\{E'\in {{\mathcal{E}_h}},\quad\mbox{such that}\quad \mathcal{N}_1(E)\cap\mathcal{N}_1(E')\neq \emptyset\}$$ From Proposition \[prop:cardinal\] we have the estimate $|\mathcal{N}_2(E)|\le (2d+1)^2$. - For the basis functions $\{\psi_E^z\}_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}$ we have the following relations: $$\label{jumpZ} \frac{1}{|E|}\int_E {\lbrack\!\lbrack \psi_{E'}^z \rbrack\!\rbrack} = \delta_{EE'},\quad\mbox{and}\quad {\lbrack\!\lbrack \psi_{E}^z \rbrack\!\rbrack}(x) = 1,\quad\mbox{for all}\quad x\in E,$$ $$\label{jumpZ1} |{\lbrack\!\lbrack \psi_{E}^z \rbrack\!\rbrack}(x)| \le 1, \quad \mbox{for all}\quad x\in E', \quad\mbox{and all}\quad E'\in \mathcal{N}_2(E).$$ The above relations all follow from the definition of $\psi^z_E(x)$ and the fact that ${\lbrack\!\lbrack \psi^z_E \rbrack\!\rbrack}$ is linear function on every face in ${{\mathcal{E}_h}}$, and therefore $\int_E{\lbrack\!\lbrack \psi^z_{E'} \rbrack\!\rbrack}=|E|{\lbrack\!\lbrack \psi^z_{E'} \rbrack\!\rbrack}(m_{E})$. - Finally, for $E\in {{\mathcal{E}_h}}$, $E'\in {{\mathcal{E}_h}}$, and $E''\in {{\mathcal{E}_h}}$ it is straightforward to see that we have: $$\label{jumpZ31} \mbox{If}\quad E\notin\mathcal{N}_1(E')\cap \mathcal{N}_1(E'') \quad \mbox{then}\quad\int_E{\lbrack\!\lbrack \psi^z_{E'} \rbrack\!\rbrack}{\lbrack\!\lbrack \psi_{E''}^z \rbrack\!\rbrack}=0.$$ An easy consequence from the definitions then is the following: $$\label{jumpZ3} \mbox{If}\quad E'\notin \mathcal{N}_2(E'') \quad \mbox{then}\quad\int_E{\lbrack\!\lbrack \psi^z_{E'} \rbrack\!\rbrack}{\lbrack\!\lbrack \psi_{E''}^z \rbrack\!\rbrack}=0, \quad \mbox{for all}\quad E\in {{\mathcal{E}_h}}.$$ We finally give Proposition \[aux\_aux\_lema\]. To avoid unnecessary complications with the notation, we state and prove the result for scalar valued functions. The proof for vector valued functions is easy to obtain, and with the same constant, by just applying the scalar valued result component-wise. \[aux\_aux\_lema\] The following inequality holds for $z\in {{ \mathcal{Z}}^{\textrm{}}}$: $$\label{the_estimate} \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\|h_E^{-1/2}({\lbrack\!\lbrack z \rbrack\!\rbrack}-{ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack})\|_{0,E}^2 \le (1-\frac{1}{\rho})\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\|h_E^{-1/2}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}\|_{0,E}^2,$$ with a constant $\rho\ge 1$ which depends on the shape regularity of the mesh. Since ${ \mathcal{P}}_{E}^{0}$ is the $L^{2}$ orthogonal projection on the constants, we have that $$\label{ineq:zPz} \|h_E^{-1/2}({\lbrack\!\lbrack z \rbrack\!\rbrack}-{ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack})\|_{0,E}^2 = \|h_E^{-1/2}{\lbrack\!\lbrack z \rbrack\!\rbrack}\|_{0,E}^2- \|h_E^{-1/2}{ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}\|_{0,E}^2.$$ Let $z \in { \mathcal{Z}}$, i.e., $z=\sum_{E' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} z_{E'} \psi_{E'}^{z}$. From  we have that ${ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack \psi_{E'}^{z} \rbrack\!\rbrack} = \delta_{EE'}$, and hence, we may conclude that $$\begin{aligned} \|h_E^{-1/2}{ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}\|_{0,E}^2 &=& \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\sum_{E'\in{{\mathcal{E}_h}}}\delta_{EE'}\frac{|E|}{h_E} z_Ez_{E'}\\ &=&\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\mathbb{D}_{EE} z_E^2=\langle \mathbb{D}\tilde{z},\tilde{z} \rangle. \end{aligned}$$ Here we have denoted by $\mathbb{D}:{{\rm I\! R}^{|{{\mathcal{E}_h}}|}}\mapsto {{\rm I\! R}^{|{{\mathcal{E}_h}}|}}$ a diagonal matrix with non-zero elements $\mathbb{D}_{EE}:=\frac{|E|}{h_E}$ and by $\tilde{z}\in {{\rm I\! R}^{|{{\mathcal{E}_h}}|}}$ the vector of coefficients $\tilde{z}=\{z_E\}_{E\in {{\mathcal{E}_h}}}$ in the expansion of $z\in { \mathcal{Z}}$ via the basis $\{\psi_{E}^{z}\}_{E\in {{\mathcal{E}_h}}}$. Further we consider the right hand side of  and we have $$\begin{aligned} \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\|h_E^{-1/2}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}\|_{0,E}^2 &=& \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} h_E^{-1} \left\|\sum_{E' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} z_{E'} {\lbrack\!\lbrack \psi_{E'}^{z} \rbrack\!\rbrack}\right\|_{0,E}^2\\ & = & \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \int_E h_E^{-1} \sum_{E' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\sum_{E'' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} z_{E'} z_{E''} {\lbrack\!\lbrack \psi_{E'}^{z} \rbrack\!\rbrack}{\lbrack\!\lbrack \psi_{E''}^{z} \rbrack\!\rbrack}\\ &= & \sum_{E' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\sum_{E'' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} z_{E'} z_{E''} \left(\sum_{E\in {{\mathcal{E}_h}}}\int_E h_E^{-1} {\lbrack\!\lbrack \psi_{E'}^{z} \rbrack\!\rbrack}{\lbrack\!\lbrack \psi_{E''}^{z} \rbrack\!\rbrack}\right)\\ &= & \sum_{E' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\sum_{E'' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} z_{E'} z_{E''}\mathbb{S}_{E'E''} = \langle \mathbb{S}\tilde{z},\tilde{z} \rangle. \end{aligned}$$ Here, $\mathbb{S}:{{\rm I\! R}^{|{{\mathcal{E}_h}}|}}\mapsto {{\rm I\! R}^{|{{\mathcal{E}_h}}|}}$ denotes the symmetric real matrix with elements $$\label{See} \mathbb{S}_{E'E''}= \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \int_Eh_E^{-1}{\lbrack\!\lbrack \psi_{E'}^{z} \rbrack\!\rbrack}{\lbrack\!\lbrack \psi_{E''}^{z} \rbrack\!\rbrack}= \sum_{E\in\mathcal{N}_1(E')\cap \mathcal{N}_1(E'')}\int_Eh_E^{-1}{\lbrack\!\lbrack \psi_{E'}^{z} \rbrack\!\rbrack}{\lbrack\!\lbrack \psi_{E''}^{z} \rbrack\!\rbrack}.$$ In the last identity above, we have used . Note that according to , if $E'\notin\mathcal{N}_2(E'')$ then $\mathbb{S}_{E'E''}=0$. Thus, $$\langle \mathbb{S}\tilde{z},\tilde{z} \rangle= \sum_{E' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\sum_{E''\in \mathcal{N}_2(E)} z_{E'} z_{E''}\mathbb{S}_{E'E''}.$$ From this identity and  and , we obtain that $$|\mathbb{S}_{E'E''}|\le |\mathcal{N}_1(E')\cap\mathcal{N}_1(E'')|\max_{E\in\mathcal{N}_1(E')\cap \mathcal{N}_1(E'')} \frac{|E|}{h_E}\le (2d+1)\max_{E\in\mathcal{N}_1(E')\cap \mathcal{N}_1(E'')}\frac{|E|}{h_E}.$$ Introducing $$\rho=\sup_{\tilde{w}\in {{\rm I\! R}^{|{{\mathcal{E}_h}}|}}} \frac{\langle \mathbb{S} \tilde{w},\tilde{w}\rangle}{\langle \mathbb{D} \tilde{w},\tilde{w}\rangle}, =\sup_{\tilde{w}\in {{\rm I\! R}^{|{{\mathcal{E}_h}}|}}} \frac{\langle \mathbb{D}^{-1/2} \mathbb{S} \mathbb{D}^{-1/2} \tilde{w},\tilde{w}\rangle}{\langle \tilde{w},\tilde{w}\rangle},$$ we obtain that $$\label{eq:ineq-s-d} \langle \mathbb{S}\tilde{z},\tilde{z} \rangle = \langle \mathbb{D}^{-1/2}\mathbb{S}\mathbb{D}^{-1/2}\mathbb{D}^{1/2}\tilde{z}, \mathbb{D}^{1/2}\tilde{z} \rangle \le \rho \langle \mathbb{D}\tilde{z},\tilde{z} \rangle.$$ This inequality can be rewritten as $\frac{1}{\rho}\langle \mathbb{S}\tilde{z},\tilde{z} \rangle \le \langle \mathbb{D}\tilde{z},\tilde{z} \rangle$ and hence $$\langle \mathbb{S}\tilde{z},\tilde{z} \rangle - \langle \mathbb{D}\tilde{z},\tilde{z} \rangle \le \langle \mathbb{S}\tilde{z},\tilde{z} \rangle - \frac{1}{\rho}\langle \mathbb{S}\tilde{z},\tilde{z} \rangle = (1-\frac{1}{\rho}) \langle \mathbb{S}\tilde{z},\tilde{z} \rangle.$$ Note that  implies that $$\label{eq:rho-gt-1} \langle \mathbb{S}\tilde{z},\tilde{z} \rangle = \langle \mathbb{D}\tilde{z},\tilde{z} \rangle +\sum_{E\in {{\mathcal{E}_h}}}\|h_E^{-1/2}({\lbrack\!\lbrack z \rbrack\!\rbrack}-{ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack})\|_{0,E}^2,$$ and thus $\langle \mathbb{S}\tilde{z},\tilde{z} \rangle \ge \langle \mathbb{D}\tilde{z},\tilde{z} \rangle$. This shows that $\rho \ge 1$ in . It remains to show that $\rho$ can be bounded by quantities depending only on the shape regularity of the mesh. Again, by  we have that: if $E'\notin\mathcal{N}_2(E'')$ then $\mathbb{S}_{E'E''}=0$. Hence: $$\begin{aligned} \rho &\le& \|\mathbb{D}^{-1/2}\mathbb{S}\mathbb{D}^{-1/2}\|_{\ell^{\infty}} \le \max_{E'' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\sum_{E'\in \mathcal{N}_2(E'')} \frac{|\mathbb{S}_{E''E'}|}{\sqrt{\mathbb{D}_{E'E'}\mathbb{D}_{E''E''}}}\\ &\le& \max_{E'' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\left[|\mathcal{N}_2(E'')|\max_{E'\in \mathcal{N}_2(E'')} \frac{|\mathbb{S}_{E'E''}|}{\sqrt{\mathbb{D}_{E'E'}\mathbb{D}_{E''E''}}}\right]\\ &\le& (2d+1)^3\max_{E'' \in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}\max_{E'\in \mathcal{N}_2(E'')} \max_{E\in\mathcal{N}_1(E')\cap \mathcal{N}_1(E'')}\frac{|E|}{h_E} \sqrt{\frac{h_{E'}h_{E''}}{|E'||E''|}}.\end{aligned}$$ The quantity on the right side of this estimate only depends on the shape regularity of the mesh and the proof is complete. We remark that the constants in Proposition \[aux\_aux\_lema\] can be sharpened, at the price of further complicating the proof. The result given above is sufficient for our purposes, and we do not further comment on the possible “optimal” value of the constant $\rho$ above. Another relevant observation is that the inequality in Proposition \[aux\_aux\_lema\] holds true, with the same or even smaller $\rho$, if we replace ${{\mathcal{E}^{o}_h}}\cup {{\mathcal{E}^{D}_h}}$ with a subset of edges $\mathcal{E}\subset ({{\mathcal{E}^{o}_h}}\cup {{\mathcal{E}^{D}_h}})$ in . The proof is completely analogous (just ${{\mathcal{E}^{o}_h}}\cup {{\mathcal{E}^{D}_h}}$ is replaced by $\mathcal{E}$). Preconditioning {#sect:preconditioning} =============== In this section, we present the construction and convergence analysis of the preconditioners we propose for the considered IP-methods. To construct the preconditioners, we use the subspace splitting given in Proposition \[split:0\], which suggests a simple change of basis. We have that for any ${\bm{u}},{\bm{w}} \in {{\bm{V}}^{\textrm{DG}}}$, we can write ${\bm{u}}={\bm{z}}+{\bm{v}}$, and ${\bm{w}}={\bm{\zeta}}+{\bm{\varphi}}$, where ${\bm{z}},{\bm{\zeta}} \in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ and ${\bm{v}}, {\bm{\varphi}} \in {{\bm{V}}^{\textrm{CR}}}$. Therefore, by performing this change of basis we can write $\mathcal{A}({\bm{u}},{\bm{w}})=\mathcal{A}(({\bm{z}},{\bm{v}}),({\bm{\zeta}},{\bm{\phi}}))$. The ${ \mathcal{A}}_0$-orthogonality of the subspaces in the splitting gives $${ \mathcal{A}}_0 (({\bm{z}},{\bm{v}}),({\bm{\zeta}},{\bm{\phi}})) = { \mathcal{A}}_0 ({\bm{z}},{\bm{\zeta}}) +{ \mathcal{A}}_0 ({\bm{v}},{\bm{\phi}}).$$ which implies that the resulting stiffness matrix of ${ \mathcal{A}}_0$ in this new basis is block diagonal. For the pure displacement problem ($\Gamma_N=\emptyset$), as discussed in Section \[sec:2\], the spectral equivalence given in Lemma \[le:A0A1\], guarantees that an optimal preconditioner for ${ \mathcal{A}}_0$ is also optimal for ${ \mathcal{A}}$. Therefore it is enough to study how to efficiently solve each of the blocks in the above block diagonal structure of ${ \mathcal{A}}_0$: the subproblem resulting from the restriction of ${ \mathcal{A}}_0$ to ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ and the subproblem on the space ${{\bm{V}}^{\textrm{CR}}}$. For traction free or mixed type of boundary conditions, although a preconditioner for ${ \mathcal{A}}_0$ does not result in an optimal solution method. However, the block structure of ${ \mathcal{A}}_0$ in the new basis already suggests that a reasonable choice for an approximation of ${ \mathcal{A}}(\cdot,\cdot)$ is $$\label{ipB} {\mathcal{B}}(({\bm{z}},{\bm{v}}),({\bm{\zeta}},{\bm{\phi}})) = { \mathcal{A}}({\bm{z}},{\bm{\zeta}}) +{ \mathcal{A}}({\bm{v}},{\bm{\phi}}).$$ The following algorithm describes the application of a preconditioner, which is based on the bilinear form in the equation . \[alg2\] Let ${\bm{r}}\in [L^{2}({\Omega})]^d$ be given. Then the action of the preconditioner on ${\bm{r}}$ is the function ${\bm{u}}\in {{\bm{V}}^{\textrm{DG}}}$ which is obtained from the following three steps. 1. Find $z\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ such that $${ \mathcal{A}}({\bm{z}},{\bm{\zeta}}) = ({\bm{r}},{\bm{\zeta}})_{{\mathcal{T}_h}} \quad \mbox{for all}\quad {\bm{\zeta}}\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}.$$ 2. Find ${\bm{v}}\in {{\bm{V}}^{\textrm{CR}}}$ such that $${ \mathcal{A}}({\bm{v}},{\bm{\varphi}}) = ({\bm{r}},{\bm{\varphi}})_{{\mathcal{T}_h}} \quad \mbox{for all}\quad {\bm{\varphi}}\in {{\bm{V}}^{\textrm{CR}}}.$$ 3. Set ${\bm{u}}={\bm{z}}+{\bm{v}}$. As before, the application of this preconditioner corresponds to solving the subproblem of the restriction of ${ \mathcal{A}}(\cdot,\cdot)$ to ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ and the subproblem of the restriction of ${ \mathcal{A}}(\cdot,\cdot)$ to ${{\bm{V}}^{\textrm{CR}}}$. We now briefly discuss how the two smaller sub-problems can be efficiently solved in both cases: (1) the case of Dirichlet boundary conditions on all of $\partial\Omega$; and (2) the case of Neumann or mixed boundary conditions. [**Solution in the subspace ${\bf { \mathcal{Z}}}$:**]{} Lemma \[le:poincZ\] guarantees that the restriction of ${ \mathcal{A}}(\cdot,\cdot)$ and ${ \mathcal{A}}_0(\cdot,\cdot)$ to ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ is well-conditioned with respect to both, the mesh size and the Lamé constants $\lambda, \mu$. Therefore, the linear system corresponding to the subproblem of the restriction to ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ can be efficiently solved by the method of Conjugate Gradients (CG). A simple consequence of the well known estimate on the convergence of CG (see, e.g., [@SaadY-2003aa; @HestenesM_StiefelE-1952aa]) shows that the number of CG iterations required to achieve a fixed error tolerance is uniformly bounded, independently of the size of the problem and the parameters. [**Solution in**]{} ${{\bm{V}}^{\textrm{CR}}}$: We now briefly discuss how to construct a uniform preconditioner for the corresponding subproblem on the space ${{\bm{V}}^{\textrm{CR}}}$. Rather than developing a completely new method, the idea is to use the optimal preconditioners that have already been studied in literature, and modify them if needed so that they fit in the present framework. For our discussion, we distinguish two cases: the pure displacement problem ($\Gamma_N=\emptyset$) and the case with mixed or traction free boundary conditions ($\Gamma_N\ne \emptyset$). - For the case of Dirichlet boundary conditions on the entire boundary–the so-called pure displacement problem–it is known how to construct optimal order multilevel preconditioners that are robust with respect to the parameter $\lambda$, see e.g. [@BlahetaR_MargenovS_NeytchevaM-2006aa; @KrausJ_MargenovS-2009aa; @GeorgievI_KrausJ_MargenovS-2009aa] and the references therein. - The traction free problem or the case of mixed boundary conditions is more difficult to handle because the (discrete) Korn inequality is not satisfied for the standard discretization by Crouzeix-Raviart elements without additional stabilization, as was shown in [@FalkR-1991aa]. The design of optimal and robust solution methods for stabilized discretizations is still an open problem, however, auxiliary space techniques might bridge this gap soon. Convergence Analysis -------------------- We now prove that the proposed block preconditioners are indeed optimal so that their convergence is uniform with respect to mesh size and the Lamé parameters. This result is given in Theorem \[the-theorem\]. The following Lemma is crucial for this proof, since it gives estimates on the norm of the off-diagonal blocks in the $2\times 2$ block form of the stiffness matrix associated to $\mathcal{A}(\cdot,\cdot)$, corresponding to the space splitting ${{\bm{V}}^{\textrm{DG}}}={{\bm{V}}^{\textrm{CR}}}\oplus {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$. The result provides a measure of the angle between the subspaces ${{\bm{V}}^{\textrm{CR}}}$ and ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$, with respect to the ${ \mathcal{A}}$-norm. The proof of this result uses Proposition \[aux\_aux\_lema\]. \[lem\_gama\] [**Strengthened Cauchy-Schwarz inequality:**]{} The following inequality holds for any ${\bm{z}}\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ and any ${\bm{v}}\in {{\bm{V}}^{\textrm{CR}}}$ $${ \mathcal{A}}({\bm{z}},{\bm{v}})^2 \leq \gamma^2 { \mathcal{A}}({\bm{z}},{\bm{z}}) { \mathcal{A}}({\bm{v}},{\bm{v}})$$ where $\gamma< 1$ and $\gamma$ depends only on $\alpha_0$, $\alpha_1$ and the constant from Proposition \[aux\_aux\_lema\]. We know that we can always choose $\alpha_0$ large enough, such that for all ${\bm{u}} \in {{\bm{V}}^{\textrm{DG}}}$ we have $${ \mathcal{A}}_0({\bm{u}},{\bm{u}})={({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}:{{\bm{\varepsilon}}({\bm{u}})})_{\mathcal{T}_h}} -2({\{\!\!\{({\mathcal{C}{\bm{\varepsilon}}({\bm{u}})}) {\bm{n}}\}\!\!\}},{\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack})_{{{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}}+ \alpha_0 a_{j,0}({\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{u}}\rbrack\!\rbrack}) \ge 0.$$ Then it is sufficient to prove that there exists $\gamma=\gamma(\alpha_1)<1$ such that for all ${\bm{z}}\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ and for all ${\bm{v}} \in {{\bm{V}}^{\textrm{CR}}}$ the inequality $$\left[a_{j,1}({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})\right]^2 \leq \gamma^2 a_{j,1}({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}) a_{j,1}({\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}),$$ holds. By the definition of the spaces ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ and ${{\bm{V}}^{\textrm{CR}}}$, on the boundary edges $E\in {{\mathcal{E}^{\partial}_h}}$ we have either ${ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}=0$ (if $E\in{{\mathcal{E}^{N}_h}}$) or ${ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}=0$ (if $E\in{{\mathcal{E}^{D}_h}}$). Hence, from the symmetry of ${ \mathcal{P}}_{E}^{0}$ we conclude that $$\int_E\langle {\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack},{ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\rangle= \int_E\langle { \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack},{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\rangle=0, \quad \mbox{for all}\quad E\in {{\mathcal{E}^{\partial}_h}}, \quad \mbox{and all}\quad {\bm{z}}\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}, \quad{\bm{v}}\in {{\bm{V}}^{\textrm{CR}}}.$$ Since for the interior edges $E\in {{\mathcal{E}^{o}_h}}$ we also have ${ \mathcal{P}}_{E}^{0}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}=0$, the above relation and the definition of ${\mathcal{P}^0_{E}}$ altogether imply that for all ${\bm{z}}\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$, and ${\bm{v}}\in {{\bm{V}}^{\textrm{CR}}}$ $$\label{eq:zero} \alpha_1\beta_1 \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \int_E\langle h_E^{-1}{\mathcal{P}^0_{E}}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack},{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\rangle= \alpha_1\beta_1 \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \int_E\langle h_E^{-1}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack},{\mathcal{P}^0_{E}}{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\rangle=0.$$ The equation  and the Schwarz inequality then lead to $$\begin{aligned} [a_{j,1}({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})]^2 & = & \left[ \alpha_1\beta_1 \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \int_E\langle h_E^{-1}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack},{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\rangle \right]^2 \\ &=& \left[ \alpha_1\beta_1 \sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \int_E\langle h_E^{-1}({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}-{\mathcal{P}^0_{E}}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack}),{\lbrack\!\lbrack {\bm{v}} \rbrack\!\rbrack}\rangle \right]^2 \\ & \leq & a_{j,1}({\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}) \left[\alpha_1\beta_1\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \|h_E^{-1/2}({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}-{\mathcal{P}^0_{E}}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack})\|_{0,E}^2\right].\end{aligned}$$ Next, the result in Proposition \[aux\_aux\_lema\] (more precisely its vector valued form) implies that $$\alpha_1\beta_1\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \|h_E^{-1/2}({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}-{\mathcal{P}^0_{E}}{\lbrack\!\lbrack {\bm{z}} \rbrack\!\rbrack})\|_{0,E}^2 \le \left(1-\frac{1}{\rho}\right) \alpha_1\beta_1\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \|h_E^{-1/2}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}\|_{0,E}^2$$ Therefore, we have $$\begin{aligned} [a_{j,1}({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack})]^2 & \leq & \left(1-\frac{1}{\rho}\right) a_{j,1}({\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}) \left[ \alpha_1\beta_1\sum_{E\in {{\mathcal{E}^{o}_h}}\cup{{\mathcal{E}^{D}_h}}} \|h_E^{-1/2}{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}\|_{0,E}^2 \right]\\ & \leq & \left(1-\frac{1}{\rho}\right)a_{j,1}({\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{z}}\rbrack\!\rbrack}) a_{j,1}({\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack},{\lbrack\!\lbrack{\bm{v}}\rbrack\!\rbrack}),\end{aligned}$$ which shows the desired inequality. We are now in a position to prove that the preconditioner given by Algorithm \[alg2\] is uniform with respect to the mesh size and the problem parameters. \[the-theorem\] Let $\mathcal{A}(\cdot,\cdot)$ be the symmetric bilinear form defined by (\[ipA\]) where $\theta=-1$ and ${\mathcal{B}}(\cdot,\cdot)$ be the bilinear form defined by (\[ipB\]). Then the following estimates hold for all ${\bm{z}}\in {{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ and for all ${\bm{v}} \in {{\bm{V}}^{\textrm{CR}}}$ $$\label{eq:equiv-A-A_0} \frac{1}{1+\gamma}\mathcal{A}(({\bm{z}},{\bm{v}}),({\bm{z}},{\bm{v}})) \leq {\mathcal{B}}(({\bm{z}},{\bm{v}}),({\bm{z}},{\bm{v}})) \leq \frac{1}{1-\gamma}\mathcal{A}(({\bm{z}},{\bm{v}}),({\bm{z}},{\bm{v}})).$$ The constant $\gamma<1$ is the constant from Lemma \[lem\_gama\]. Using Lemma \[lem\_gama\] we have $$- 2 \gamma \sqrt{{ \mathcal{A}}({\bm{z}},{\bm{z}}) \, { \mathcal{A}}({\bm{v}},{\bm{v}})} \le 2 { \mathcal{A}}({\bm{z}},{\bm{v}}) \le 2 \gamma \sqrt{{ \mathcal{A}}({\bm{z}},{\bm{z}}) \, { \mathcal{A}}({\bm{v}},{\bm{v}})}$$ and since $-a^2-b^2 \le 2 a b \le a^2 + b^2$ for any real numbers $a$ and $b$ we obtain $$(1 - \gamma) \left( { \mathcal{A}}({\bm{z}},{\bm{z}}) + { \mathcal{A}}({\bm{v}},{\bm{v}}) \right) \le { \mathcal{A}}({\bm{z}},{\bm{z}}) + { \mathcal{A}}({\bm{v}},{\bm{v}}) + 2 { \mathcal{A}}({\bm{z}},{\bm{v}}) \le (1 + \gamma) \left( { \mathcal{A}}({\bm{z}},{\bm{z}}) + { \mathcal{A}}({\bm{v}},{\bm{v}}) \right)$$ which is the same as $$(1 - \gamma) {\mathcal{B}}(({\bm{z}},{\bm{v}}),({\bm{z}},{\bm{v}})) \le { \mathcal{A}}(({\bm{z}},{\bm{v}}),({\bm{z}},{\bm{v}})) \le (1 + \gamma) {\mathcal{B}}(({\bm{z}},{\bm{v}}),({\bm{z}},{\bm{v}}))$$ and thus (\[eq:equiv-A-A\_0\]) holds with the same constant $\gamma < 1$ as used in the estimate of Lemma \[lem\_gama\]. \[rem\_gama\] Note that $\gamma\le q < 1$ is uniformly bounded away from $1$ and this bound holds independently of the parameters $h$, $\lambda$, and $\mu$. Numerical experiments {#sect:numerics} ===================== In this section we present a set of numerical tests that illustrate our theoretical results. We consider the SIPG discretization of the model problem (\[eq:PDEform\]) on the unit square in ${{\rm I\! R}^{2}}$ with mixed boundary conditions. For the penalty parameters in (\[eq:aj0aj1\]) we choose the values $\alpha_0=4$ and $\alpha_1=1$. The coarsest mesh (at level $0$) consists of eight triangles and is refined four times. Each refined mesh at level $\ell$, $\ell=1,2,3,4$ is obtained by subdividing every triangle at level $(\ell-1)$ into four congruent triangles. The CBS constants and the spectral condition numbers summarized in the tables below have been computed using MATLAB. In Table \[gama\_tabl\] we list the values of the constant $\gamma^{2}$ in the inequality stated in Lemma \[lem\_gama\] for different levels of refinement. Evidently, $\gamma$ is uniformly bounded with respect to the mesh size (or the number of refinement levels) and also with respect to the material parameters, Young’s modulus $\mathfrak{E}$ and Poisson ratio $\nu$ (see Remark \[rem\_gama\]). $\gamma^{2}$ $\nu=0.25$ $\nu=0.4$ $\nu=0.49$ $\nu=0.499$ $\nu=0.49999$ -------------- ------------ ----------- ------------ ----------------------- ----------------------- $\ell=1$ 0.0664 0.025 0.0024 2.4024$\times10^{-4}$ 2.4015$\times10^{-6}$ $\ell=2$ 0.0678 0.0255 0.0025 2.4567$\times10^{-4}$ 2.4559$\times10^{-6}$ $\ell=3$ 0.0684 0.0258 0.0025 2.4866$\times10^{-4}$ 2.4857$\times10^{-6}$ $\ell=4$ 0.0686 0.0259 0.0025 2.4974$\times10^{-4}$ 2.4966$\times10^{-6}$ : Observed CBS constant $\gamma^{2}$ for $\Omega=(0,1)^2$. \[gama\_tabl\] It can be seen from Table \[gama\_tabl\_jump\] that the two subspaces ${{\bm{V}}^{\textrm{CR}}}$ and ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$ remain nearly ${ \mathcal{A}}$-orthogonal when we introduce a jump in the Poisson ratio (on the coarsest mesh); In our experiment we set $\nu=\nu_1=0.3$ (and $E=E_1=1$) in the subdomain $\Omega_1=[0,0.5]\times[0,0.5]\cup [0.5,1]\times[0.5,1]$, and $\nu=\nu_2$ (and $E_2=1$) in the subdomain $\Omega_2=\Omega \setminus \Omega_2$, respectively. $\gamma^{2}$ $\nu_2=0.3$ $\nu_2=0.4$ $\nu_2=0.49$ $\nu_2=0.499$ $\nu_2=0.49999$ -------------- ------------- ------------- -------------- --------------- ----------------- $\ell=1$ 0.0451 0.0177 0.0442 0.0509 0.0517 $\ell=2$ 0.0460 0.0180 0.0689 0.0803 0.0816 $\ell=3$ 0.0464 0.0182 0.0689 0.0802 0.0816 $\ell=4$ 0.0466 0.0182 0.0689 0.0802 0.0816 : Observed CBS constant $\gamma^{2}$ for $\Omega=(0,1)^2$ and jumps in $\nu$. \[gama\_tabl\_jump\] Next we consider an L-shaped domain $\Omega=[0,1]\times[0,1] \setminus (0.5,1]\times(0.5,1]$ with Neumann boundary conditions on the sides $y=0$ and $y=1$ and Dirichlet boundary conditions on the remaining part of the boundary. The initial triangulation (level 0) consists of 4 similar triangles. The angle is almost the same as for the square domain, see Table \[Lgama\_tabl\]. $\gamma^{2}$ $\nu=0.25$ $\nu=0.4$ $\nu=0.49$ $\nu=0.499$ $\nu=0.49999$ -------------- ------------ ----------- ------------ ----------------------- ----------------------- $\ell=1$ 0.0561 0.0202 0.0019 1.8918$\times10^{-4}$ 1.8906$\times10^{-6}$ $\ell=2$ 0.0631 0.0233 0.0022 2.2118$\times10^{-4}$ 2.2106$\times10^{-6}$ $\ell=3$ 0.0672 0.0252 0.0024 2.4216$\times10^{-4}$ 2.4207$\times10^{-6}$ $\ell=4$ 0.0682 0.0257 0.0025 2.4810$\times10^{-4}$ 2.4801$\times10^{-6}$ : Observed CBS constant $\gamma^{2}$ for L-shaped domain.\[Lgama\_tabl\] Furthermore, we computed the relative condition number of the preconditioner $B$ corresponding to the bilinear form (\[ipB\]) for the model problem on the L-shaped domain. The results of this experiment, which are listed in Table \[Lrel\_cond\_num\], confirm the uniform bound provided by Theorem \[the-theorem\]. $\kappa(B^{-1}A)$ $\nu=0.25$ $\nu=0.4$ $\nu=0.49$ $\nu=0.499$ $\nu=0.49999$ ------------------- ------------ ----------- ------------ ------------- --------------- $\ell=1$ 1.6204 1.3314 1.0912 1.0279 1.0028 $\ell=2$ 1.6713 1.3606 1.0990 1.0302 1.0030 $\ell=3$ 1.6997 1.3774 1.1037 1.0316 1.0031 $\ell=4$ 1.7073 1.3820 1.1050 1.0320 1.0032 : Tabulated values of $\kappa(B^{-1}A)$ for L-shaped domain. \[Lrel\_cond\_num\] Finally, we computed the condition number $\kappa(A_{zz})$ of the matrix $A_{zz}$ related to the restriction of ${ \mathcal{A}}(\cdot,\cdot)$ to the space ${{\bm{{ \mathcal{Z}}}}^{\textrm{}}}$, again for the model problem on the L-shaped domain. In view of Lemma \[le:poincZ\] we already know that $A_{zz}$ is well-conditioned, and this is clearly seen in Table \[Lcond\_Z\_block\] where the values of $\kappa(A_{zz})$ are listed. $\kappa(A_{zz})$ $\nu=0.25$ $\nu=0.4$ $\nu=0.49$ $\nu=0.499$ $\nu=0.49999$ ------------------ ------------ ----------- ------------ ------------- --------------- $\ell=1$ 8.9067 7.1484 6.4788 6.4220 6.4158 $\ell=2$ 9.0875 7.1932 6.4829 6.4229 6.4164 $\ell=3$ 9.1577 7.2080 6.4841 6.4230 6.4164 $\ell=4$ 9.1794 7.2118 6.4844 6.4230 6.4164 : Values of $\kappa(A_{zz})$ for L-shaped domain. \[Lcond\_Z\_block\] Acknowledgments =============== Part of this work was completed while the fourth author was visiting RICAM, Austrian Academy of Sciences in Linz. Thanks go to the RICAM for the kind hospitality and support. The work of the first author was partially supported by the Spanish MEC under projects MTM2008-03541 and HI2008-0173. The work of the second author has been partially supported by the Bulgarian NSF, Grant DO 02-338/08. We also gratefully acknowledge the support by the Austrian Science Fund, Grants P19170-N18 and P22989-N18. The work of the fourth author has been supported in part by the US National Science Foundation, Grants DMS-0810982, and OCI-0749202. Auxiliary results {#ap0} ================= Bounds on the cardinality of $\mathcal{N}_1(E)$ and $\mathcal{N}_2(E)$ ---------------------------------------------------------------------- We first recall the definitions of $\mathcal{N}_0(E)$, $\mathcal{N}_1(E)$ and $\mathcal{N}_2(E)$, already given in §\[subsect:cbs\]: $$\begin{aligned} \mathcal{N}_0(E)&:=&\{T\in {\mathcal{T}_h},\quad\mbox{such that}\quad E\in T\},\\ \mathcal{N}_1(E)&:=&\{E'\in {{\mathcal{E}_h}},\quad\mbox{such that}\quad \mathcal{N}_0(E)\cap\mathcal{N}_0(E')\neq \emptyset\},\\ \mathcal{N}_2(E)&:=&\{E'\in {{\mathcal{E}_h}},\quad\mbox{such that}\quad \mathcal{N}_1(E)\cap\mathcal{N}_1(E')\neq \emptyset\}.\end{aligned}$$ In the proof of the strengthened Cauchy-Schwarz inequality §\[subsect:cbs\] we needed several estimates on the cardinality of these sets and these estimates are given in the proposition below. We remind the reader that we have $|\mathcal{N}_0(E)|\le 2$. \[prop:cardinal\] The following inequalities hold: $$\label{eq:cardinal} |\mathcal{N}_1(E)|\le (2d+1)\quad\mbox{and}\quad |\mathcal{N}_2(E)|\le (2d+1)^2.$$ Let $E\in {{\mathcal{E}_h}}$ be fixed. To prove the bound on $|\mathcal{N}_1(E)|$ we consider the elements $T\in {\mathcal{T}_h}$, such that $E\in T$. In each such element $T$, there are exactly $d$ faces $E'\in T$, $E'\neq E$. Since there are at most two elements $T\in {\mathcal{T}_h}$ containing $E$ we have at most $2d$ faces $E'\in{{\mathcal{E}_h}}$ such that $E'\in \mathcal{N}_1(E)$, and $E'\neq E$. Adding $E$ itself to the total count gives $|\mathcal{N}_1(E)|\le (2d+1)$. The second bound given in  follows from the first and the following inclusion: $$\displaystyle \mathcal{N}_2(E)\subset \bigcup_{E'\in \mathcal{N}_{1}(E)} \mathcal{N}_1(E').$$ To show the above inclusion, we consider an arbitrary $E''\in \mathcal{N}_2(E)$. By the definition of $\mathcal{N}_2(E)$, the intersection of $\mathcal{N}_1(E'')$ and $\mathcal{N}_1(E)$ is not empty. Equivalently, there exists $E'\in{{\mathcal{E}_h}}$ such that $E'\in\mathcal N_1(E'')$ and $E'\in\mathcal{N}_1(E)$. On the other hand, from the definition of $\mathcal{N}_1(E'')$, we have that $E'\in\mathcal{N}_1(E'')$ implies that $E''\in\mathcal{N}_1(E')$, i.e., if $E'$ is a neighbor of $E''$, then $E''$ is a neighbor of $E'$. Putting this together, we conclude that: if $E''\in \mathcal{N}_2(E)$, then there exists $E'\in \mathcal{N}_1(E)$, such that $E''\in\mathcal{N}_1(E')$, and this is exactly the inclusion we wanted to show. To prove the desired bound is then straightforward: $$\begin{aligned} \displaystyle \left|\bigcup_{E'\in \mathcal{N}_{1}(E)} \mathcal{N}_1(E')\right|&\le& \sum_{E'\in \mathcal{N}_{1}(E)} |\mathcal{N}_1(E')| \le \sum_{E'\in \mathcal{N}_{1}(E)} (2d+1) \\ &=&(2d+1)|\mathcal{N}_{1}(E)| \le (2d+1)^2.\end{aligned}$$ A multiplicative relation {#subsect:dumb} ------------------------- This is to prove a basic relation used to derive  as well as . Let $\odot$ be a map $V\times W\mapsto U$, where $U$, $V$, and $W$ are linear vector spaces over the real numbers. We assume that $\odot$ satisfies the following distributive laws: $$a\odot (b+c)= a\odot b + a\odot c, \qquad (a+b)\odot c= a\odot c + b\odot c,$$ and we assume that for all $\xi\in {{\rm I\! R}^{}}$ and all $\eta\in {{\rm I\! R}^{}}$, we have: $$\label{eq:prop0} (\xi a)\odot (\eta b) = (\xi\eta) (a\odot b).$$ We have the following identities, based on the definitions : $$\label{eq:identity0} a^+\odot b^+-a^-\odot b^- = {\lbrack\!\lbrack a \rbrack\!\rbrack}\odot {\{\!\!\{b\}\!\!\}}+{\{\!\!\{a\}\!\!\}}\odot {\lbrack\!\lbrack b \rbrack\!\rbrack}.$$ Proving this relation is indeed trivial. Some examples for which the reader should verify these identities are: (1) For real numbers $a$ and $b$ one may take as $\odot$ the usual multiplication of real numbers; (2) $a$ and $b$ elements of a real Hilbert space and $\odot$ inner product; (3) $a$ and $b$ are linear operators, and $\odot$ is then the multiplication of linear operators. Note that in such case $\odot$ is not necessarily commutative; (4) $a$ is a matrix and $b$ is a vector, or more generally, $a$ is a linear operator and $b$ is an element of a Hilbert space. From , we have that the right side of the identity  is $${\lbrack\!\lbrack a \rbrack\!\rbrack}\odot {\{\!\!\{b\}\!\!\}}+{\{\!\!\{a\}\!\!\}}\odot {\lbrack\!\lbrack b \rbrack\!\rbrack}= (a^+ - a^-)\odot \left(\frac{b^+ + b^-}{2}\right)+ \left(\frac{a^+ + a^-}{2}\right)\odot (b^+ - b^-)$$ Using the distributive law, and  (linearity of $\odot$ with respect to scalar multiplication), we have $$\begin{aligned} \lefteqn{(a^+ - a^-)\odot \left(\frac{b^+ + b^-}{2}\right)+ \left(\frac{a^+ + a^-}{2}\right)\odot (b^+ - b^-)}\\ &=& \frac12(a^+ - a^-)\odot (b^+ + b^-)+ \frac12(a^+ + a^-)\odot (b^+ - b^-)\\ &=& \frac12a^+ \odot (b^+ + b^-)- \frac12a^- \odot (b^+ + b^-)+ \frac12a^+ \odot (b^+ - b^-)+ \frac12a^- \odot(b^+ - b^-)\\ &=& \frac12a^+ \odot b^+ +\frac12a^+ \odot b^- -\frac12a^- \odot b^+ -\frac12a^- \odot b^-\\ && +\frac12a^+ \odot b^+ -\frac12a^+ \odot b^- +\frac12a^- \odot b^+ -\frac12a^- \odot b^-\\ & = & \frac12a^+ \odot b^+ -\frac12a^- \odot b^- +\frac12a^+ \odot b^+ -\frac12a^- \odot b^- = a^+ \odot b^+ - a^- \odot b^-.\end{aligned}$$ [10]{} Douglas N. Arnold, Franco Brezzi, Bernardo Cockburn, and L. Donatella Marini. Unified analysis of discontinuous [G]{}alerkin methods for elliptic problems. , 39(5):1749–1779 (electronic), 2001/02. Douglas N. Arnold, Franco Brezzi, Richard Falk, and L. Donatella Marini. Locking-free reissner-mindlin elements without reduced integration. , 196(37-40):3660–3671, 2007. Douglas N. Arnold, Franco Brezzi, and L. Donatella Marini. A family of discontinuous [G]{}alerkin finite elements for the [R]{}eissner-[M]{}indlin plate. , 22/23:25–45, 2005. Blanca Ayuso de Dios and Ludmil Zikatanov. Uniformly convergent iterative methods for discontinuous [G]{}alerkin discretizations. , 40(1-3):4–36, 2009. Radim Blaheta, Svetozar Margenov, and Maya Neytcheva. Aggregation-based multilevel preconditioning of non-conforming fem elasticity problems. In Jack Dongarra, Kaj Madsen, and Jerzy Wasniewski, editors, [ *Applied Parallel Computing. State of the Art in Scientific Computing*]{}, volume 3732 of [*Lecture Notes in Computer Science*]{}, pages 847–856. Springer Berlin / Heidelberg, 2006. Susanne C. Brenner. Poincaré-[F]{}riedrichs inequalities for piecewise [$H\sp 1$]{} functions. , 41(1):306–324 (electronic), 2003. Susanne C. Brenner. Korn’s inequalities for piecewise [$H^1$]{} vector fields. , 73(247):1067–1087 (electronic), 2004. F. Brezzi, B. Cockburn, L. D. Marini, and E. S[ü]{}li. Stabilization mechanisms in discontinuous [G]{}alerkin finite element methods. , 195(25-28):3293–3310, 2006. Erik Burman and Benjamin Stamm. Low order discontinuous [G]{}alerkin methods for second order elliptic problems. , 47(1):508–533, 2008. G. Duvaut and J.-L. Lions. . Springer-Verlag, Berlin, 1976. Translated from the French by C. W. John, Grundlehren der Mathematischen Wissenschaften, 219. Richard S. Falk. Nonconforming finite element methods for the equations of linear elasticity. , 57(196):529–550, 1991. Xiaobing Feng and Ohannes A. Karakashian. Two-level additive [S]{}chwarz methods for a discontinuous [G]{}alerkin approximation of second order elliptic problems. , 39(4):1343–1365 (electronic), 2001. I. Georgiev, J. K. Kraus, and Margenov S. Multilevel preconditioning of [C]{}rouzeix-[R]{}aviart [3D]{} pure displacement elasticity problems. In I. Lirkov, S. Margenov, and J. Wasniewski, editors, [*Large Scale Scientific Computing*]{}, volume 5910 of [*Lecture Notes in Computer Science (LNCS)*]{}, pages 103–110. Springer, Berlin, Heidelberg, 2010. Peter Hansbo and Mats G. Larson. Discontinuous [G]{}alerkin methods for incompressible and nearly incompressible elasticity by [N]{}itsche’s method. , 191(17-18):1895–1908, 2002. Peter Hansbo and Mats G. Larson. Discontinuous [G]{}alerkin and the [C]{}rouzeix-[R]{}aviart element: application to elasticity. , 37(1):63–72, 2003. Magnus R. Hestenes and Eduard Stiefel. Methods of conjugate gradients for solving linear systems. , 49:409–436 (1953), 1952. Johannes Kraus and Svetozar Margenov. , volume 5 of [*Radon Series on Computational and Applied Mathematics*]{}. Walter de Gruyter GmbH & Co. KG, Berlin, 2009. Yousef Saad. . Society for Industrial and Applied Mathematics, Philadelphia, PA, second edition, 2003. Thomas P. Wihler. Locking-free [DGFEM]{} for elasticity problems in polygons. , 24(1):45–75, 2004. Thomas P. Wihler. Locking-free adaptive discontinuous [G]{}alerkin [FEM]{} for linear elasticity problems. , 75(255):1087–1102 (electronic), 2006. [^1]: We note that in [@BrezziF_CockburnB_MariniL_SuliE-2006aa] the focus is on the scalar Laplace equation. The arguments for the elasticity problem, are basically the same.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We test the ability of equilibrium galactic disk and one-zone interstellar medium models to describe the physical and emission properties of quasar hosts, submillimeter galaxies, and Lyman-$\alpha$ emitters at $z\gtrsim6$. The size, line widths, star formation rates, black hole accretion rates, gas masses and temperatures, and the relationships between these properties are all well-described by our model, and we provide approximate fitting formulae for comparison with future observations. However, comparing our carbon line predictions to observations reveals differences between the ISM at low and high redshifts. Our underestimate of the \[CII\] line emission indicates either higher star formation efficiencies in high-redshift molecular clouds or less depletion of metals into dust at fixed metallicity. Further, our over-prediction of the CO(6–5)/CO(1–0) ratio suggests that molecular clouds in real high-redshift galaxies have a lower turbulent Mach number and more subthermal CO(6–5) emission than expected owing either to sizes smaller than the local Jeans mass or to a pressure support mechanism other than turbulence.' author: - | Joseph A. Mu[ñ]{}oz[^1] and Steven R. Furlanetto\ Department of Physics and Astronomy, University of California Los Angeles; Los Angeles, CA 90095, USA bibliography: - 'ms.bib' title: 'Extreme Galaxies During Reionization: Testing ISM and Disk Models' --- galaxies: high-redshift — galaxies: evolution — galaxies: ISM — radio lines: galaxies — submillimeter: galaxies — quasars: general Introduction ============ Observations continue to probe the universe at ever higher redshifts with new discoveries of quasar hosts [e.g., @Fan00; @Fan01; @Fan03; @Willott07; @Willott09; @Jiang09; @Mortlock09; @Willott10; @Mortlock11], sub-millimeter galaxies [SMGs; e.g., @Vieira13; @Riechers13], Lyman-break galaxies [LBGs; e.g., @Bouwens06; @Bunker10; @McLure10; @Finkelstein10; @Yan10; @Bouwens11a; @Bouwens11b; @Oesch12; @Coe13; @Ellis13; @Oesch13], and Lyman-$\alpha$ emitters [LAEs; e.g., @Hu02a; @Taniguchi05; @Iye06; @Ouchi09a; @Ouchi09b; @Ouchi10] at $z>6$. These systems are different than those observed locally, with higher surface densities and smaller sizes besides being bathed in a significantly hotter Cosmic Microwave Background (CMB). The more common LBGs and LAE, may also have much lower metallicities [e.g., @Finlator11; @MF13a; @Ouchi13]. While observations of atomic and molecular lines have already begun to probe the state of the cool gas in these galaxies [see @CW13 for a recent review], putting them into a complete theoretical framework for galaxy formation across cosmic time requires understanding the implications of these physical differences on the interstellar medium (ISM) at high-redshift [e.g., @Narayanan12; @MF13a]. Such a framework for galaxy formation includes the relationships among cosmic inflows, stellar and quasar outflows, gas transport through the galaxy, the buildup of the stellar population, and the growth of a central supermassive black hole in an active galactic nucleus (AGN). We would also like to understand how the state of the ISM gas, the efficiency with which it forms stars, and the resulting molecular line emission relates to and can be a probe of the larger-scale physics. In @Munoz12, we described the evolution of the galaxy luminosity function as a balance among star formation, star formation-driven outflows, and the gas accretion rate onto galaxies from the buildup of cosmological structure. We then developed a model for galactic disks in @MF12—initially derived by @Thompson05—and embedded it into our cosmic setting to study the relationship between gas transport within galaxies and the growth of central black holes in the faintest high-redshift systems. Finally, in @MF13a, we overlaid a photo-dissociative model for molecular clouds on sub-galactic scales to understand the chemical state of CO and make realistic predictions for ALMA and JVLA observations that self-consistently account for the physical conditions in the $z\sim6$ ISM. Here we found that the CO signal from typical galaxies at these redshifts will be very difficult to observe if they are as metal- and dust-poor as expected, which both reduces the total amount of carbon and allows a higher fraction of the CO to be dissociated. In the present work, we will continue to use this theoretical framework to describe galaxy formation during the reionization epoch. While we previously focused on the CO signal from LBGs, we now turn our attention to CII—a product of CO dissociation—and its associated $158\,{\rm \mu m}$ line emission, which may be easier to detect than CO in low-metallicity environments. We also apply our model to a broader range of $z=6$ galaxy types from LAEs—with relatively low star formation rates ($\lesssim 10\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr}$)—to SMGs and quasar hosts—with star formation rates of tens to thousands of solar masses per year. More data are available for these brighter systems [e.g., @Riechers09; @Wang13; @Willott13; @Riechers13], and the environments are undoubtedly extreme,making comparisons to our models more informative. Our goals will be to ask whether our idealized framework for galaxy formation can describe SMGs and quasar hosts at all, to understand the physical sources of any disparity, and learn how to improve our treatment of sources currently undetected in molecular emission lines. Missing \[CII\] Emission {#sec:intro:cii} ------------------------ One key aim of this work is to illuminate a puzzle in the relationship between molecular gas physics and the observed \[CII\] emission at high redshift. Here we illustrate the difficulty with a simplified calculation (many of the parameter choices made here will be justified later on). Let us assume that the host galaxies of the brightest $z=6$ quasars and SMGs live in $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ halos, that the cold gas in the galactic disk comprises 10% of the halo’s baryons,[^2] i.e., just over $10^{11}\,{\ensuremath{{\rm M_{\odot}}}}$ worth, all of which has a temperature of $50\,{\rm K}$. The carbon in some fraction, $\bar{f}_{\rm CO}$, of this gas is in the form of CO, while the rest is dissociated into CII. The galactic disk, which extends out to 5%/$\sqrt{2}$ of the halo virial radius, is about $3.5\,{\ensuremath{{\rm kpc}}}$ in this halo. The average surface density of the disk is, thus, a few thousand solar masses per square parsec ($4200\,{\rm {\ensuremath{{\rm M_{\odot}}}}/pc^2}$ for a cosmic baryon fraction of $f_{\rm b}=0.16$) and, given the redshift dependence of the halo virial radius, is expected to scale as $(1+z)^2$ for fixed mass. This is well above the typical surface density of $85\,{\rm {\ensuremath{{\rm M_{\odot}}}}/pc^2}$ observed in local clouds, so we assume that clouds in this galaxy have surface densities of order this higher value [@Krumholz09b]. For a clumping factor of $c=5$, the chemical equilibrium calculation of @Krumholz09b predicts that such dense gas should be fully molecular unless the metallicity is below about $0.05\%$ of solar (below the threshold at which the equilibrium approximations in the model holds). Moreover, at solar metallicity, the visual extinction through one of these clouds is about 200, and thus, approximately 99% all of carbon gas is in the form of CO—sufficiently shielded against dissociation—according to the PDR model of @Wolfire10 [Eq. \[eq:fco\]]. 50% of the carbon is dissociated if the metallicity is approximately 1% of solar. The maximum amount of \[CII\] emission is produced if the line is thermalized and optically thin. In this optimistic case, the luminosity can be calculated simply by $$\label{eq:Lcii} {L_{\rm [CII]}}=\frac{\left(1-\bar{f}_{\rm CO}\right)\,{M_{\rm gas}}\,{\rm C/H}}{m_{\rm p}}\,\frac{4\,{\rm e}^{-{\ensuremath{{h_{\rm p}}}}\,\nu_{\rm [CII]}/k_{b}\,\bar{T}}}{2+4\,{\rm e}^{-{\ensuremath{{h_{\rm p}}}}\,\nu_{\rm [CII]}/k_{b}\,\bar{T}}}\,A_{\rm [CII]}\,{\ensuremath{{h_{\rm p}}}}\,\nu_{\rm [CII]},$$ where $A_{\rm [CII]}=2.3\times10^{-6}\,{\rm s^{-1}}$, $\nu_{\rm [CII]}=1900.5\,{\rm GHz}$, $m_{\rm p}$ is the proton mass, and ${\ensuremath{{h_{\rm p}}}}$ is the Planck constant. We assume, moreover, that the carbon abundance is ${\rm C/H}=1.5\times10^{-4}$ at solar metallicity. Then, a CO fraction of $\bar{f}_{\rm CO}=0.99$ gives a luminosity of about $0.5\times10^9\,{\ensuremath{{\rm L_{\odot}}}}$. At high temperature, the ratio of the number of molecules in the excited state to the number in the ground state saturates at $2/3$ giving a luminosity of $1.4\times10^9\,{\ensuremath{{\rm L_{\odot}}}}$. For comparison, the quasars J2310$+$1855 [@Wang13] and J1148$+$5251 [@Riechers09] and the submillimeter galaxy HFLS3 [@Riechers13] have \[CII\] luminosities of $8.8$, $26$, and $16\times10^{9}\,{\ensuremath{{\rm L_{\odot}}}}$, respectively. These observations are an order-of-magnitude higher than the most optimistic cases in this simple calculation. Clearly, one of our seemingly reasonable assumptions does not extrapolate to these systems at $z\sim6$. We plan to use our more sophisticated, physically motivated model of galaxy formation and line emission at high redshift to investigate this issue. This Paper {#sec:intro:outline} ---------- We begin with a brief summary of the relevant details of the galaxy formation model developed in our previous work (§\[sec:model\]). We then describe the observations of quasar hosts, SMGs, and LAEs that we take from the literature (§\[sec:obs\]) and interpret physical properties from these data that we will compare to our model. We consider the suitability of our theoretical galactic disks for describing the observations (§\[sec:comp\]) with respect to their disk structure (§\[sec:comp:struc\]), quasar luminosities (§\[sec:comp:lbh\]), gas masses (§\[sec:comp:mgas\]), and gas temperatures (§\[sec:comp:temp\]). Next, we turn our attention to the emission from carbon lines (§\[sec:c\]). We compute the \[CII\] emission, presenting results using our fiducial model (§\[sec:c:cii\]) and considering additions (§\[sec:c:alt\]) that improve agreement with the \[CII\] data, before turning our attention to the line ratios of CO (§\[sec:c:co\]). Finally, we conclude with a summary and discussion our results and their implications for high-redshift galaxy formation (§\[sec:discussion\]). Throughout this work, we assume a flat, $\Lambda$CDM universe with $H_0=70\,{\rm km/s/Mpc}$, $\Omega_{\rm m}=0.28$, $\Omega_{\rm b}=0.046$, and $\sigma_8=0.82$. Galactic Disk Model {#sec:model} =================== We use the radiation pressure- and supernovae-supported disk models of @Thompson05 and @MF12 combined with the treatments of molecular clouds and carbon emission lines in @MF13a. We refer the reader to these works for details of our model while only briefly summarizing our methods here. We assume that gas accretes onto the outer edge of each disk at the cold flow rate [e.g., @McBride09] and, in our fiducial model, ignore suppression from AGN feedback in very large halo masses. Star formation and winds deplete the gas as it moves toward the disk center. The rate of star formation at each radius is set such that radiation pressure from starlight on dust and mechanical pressure from supernovae maintain marginal Toomre-instability (i.e., $Q=1$) and vertical hydrostatic equilibrium.[^3] Meanwhile, we assume momentum driven super-winds eject gas at a rate $(400\,{\rm km/s})/\sigma$ times higher than the star formation rate, where $\sigma$ is the halo velocity dispersion, matching observations of the UV luminosity function at $z=6$–$8$ [@ML11; @Munoz12] and consistent with the results from numerical simulations [@OD08].[^4] We consider two phenomenological models for the transport rate of gas through the disk: a linear spiral wave (LSW) model in which the inflow velocity is $v_{\rm in}=m\,c_{\rm s}$ and a shocked, nonlinear inflow model where $v_{\rm in}=\sqrt{2}\,\beta\,\sigma$. Here, $c_{\rm s}$ is the local sound speed and the two free-parameters defining these prescriptions are the Mach number, $m$, and the constant $\beta$. We further assume that the gas remaining after depletion by star formation and winds transitions smoothly into the accretion disk of a central black hole and powers an AGN. In LBGs, the black hole growth rate is negligible compared to the total star formation rate, and while the resulting X-rays dominate the contribution from high-mass X-ray binaries in the stellar disk, they are currently undetectable in stacked samples of [*[Chandra]{}*]{} data [@MF12]. However, with high halo masses and for very rapid gas inflow, e.g., $\beta$-models with $\beta \gtrsim 0.01$, the black hole growth rate can be significant enough to power an observable quasar. At each radius in the disk, we then assume that the gas fragments into clouds on the scale of the local Jeans mass. Calculating the fraction of molecular gas turned into stars, ${{\rm SFR_{\rm ff}}}$, per free-fall time, ${t_{\rm ff}}$, as prescribed by @KM05, we set the fraction of cloud gas in molecular form, ${f_{\rm H_{2}}}$, to be such that the required star formation rate is produced. That is, $$\label{eq:fmol} {f_{\rm H_{2}}}=\frac{{\dot{\Sigma}_{\star}}}{{\Sigma_{\rm g}}}\, \left( \frac{{{\rm SFR_{\rm ff}}}}{{t_{\rm ff}}}\right)^{-1},$$ where ${\dot{\Sigma}_{\star}}$ and ${\Sigma_{\rm g}}$ are the surface star formation rate and gas densities as a function of galactocentric radius. Additionally, where ${f_{\rm H_{2}}}$ would be greater than unity, we simply set ${f_{\rm H_{2}}}=1$.[^5] Thus, in general, more efficient star formation results in lower molecular fractions since the dynamically-balanced star formation rate is unchanged. We then use the @Wolfire10 model of cloud photo-dissociated regions (PDRs) to calculate the fraction, ${f_{\rm CO}}$, of cloud mass in which carbon is in the form of CO rather than dissociated CII, where ${f_{\rm CO}}\leq {f_{\rm H_{2}}}$. CO in the clouds are shielded from the external dissociating radiation field, $G_0$, by turbulently-generated inhomogeneities that follow a log-normal distribution where the ratio of the median density to the average is $\sqrt{1+3\,{\mathcal{M}}^2/4}$, with ${\mathcal{M}}$ is the thermal Mach number of the turbulent gas. @Wolfire10 models these inhomogeneities as uniform clumps of density $n_{\rm c}$ embedded within smooth, more diffuse gas. The ratio ${f_{\rm CO}}/{f_{\rm H_{2}}}$ is then given by $$\label{eq:fco} {\ln}\left(\frac{{f_{\rm CO}}}{{f_{\rm H_{2}}}}\right)=\frac{-4.0}{A_{V}}\,\left[0.53-0.045\,{\ln}\left(\frac{G'_0}{n_{\rm c}\,{\rm cm^3}}\right)-0.097\,{\ln}(Z')\right],$$ where $G'_0=G_0/2.7\times10^{-3}\,{\rm erg\,s^{-1}\,cm^{-2}}$ is the external far-UV radiation field bathing the cloud in units of the local Galactic interstellar field found by @Draine78, $Z'$ is the metallicity in solar units, and we set $n_{\rm c}$ to be the median gas density in the cloud. The far UV dissociating radiation field is determined produced primarily by starlight, and we have checked that emission from the AGN in our model galaxies contributes negligibly. --------------------------------- ------ ----------------- ------------- --------- ------------------------------------------- ------ -------- Name $z$ $R_{\rm disk} $ $v_{\rm c}$ sin $i$ SFR Type Refer. kpc km/s ${\rm {\ensuremath{{\rm M_{\odot}}}}/yr}$ (1) (2) (3) (4) (5) (6) (7) (8) SDSS J231038.88$+$185519.7 6.00 3.18$\pm$0.17 549$\pm$46 0.71 2100\* QSO A ULAS J131911.29$+$095051.4 6.13 3.26$\pm$0.39 620$\pm$143 0.84 1400\* QSO A SDSS J205406.49$-$000514.8 6.04 1.98$\pm$0.28 589$\pm$55 0.41 380\* QSO A SDSS J012958.51$-$003539.7 5.78 2.41$\pm$0.35 242$\pm$31 0.80 40\* QSO A SDSS J104433.04$-$012502.2 5.78 3.54$\pm$0.64 160$\pm$60 –– 2800\* QSO A SDSS J114816.64$+$525150.3 6.42 4.93$\pm$0.88 287$\pm$28 –– 3300 QSO B CFHQS J021013$-$045620 6.43 2.85$\pm$1.36 189$\pm$18 –– 48$\pm$14 QSO C CFHQS J232908$-$030158 6.42 –– –– –– $< 40$ QSO C 1HERMES S350 J170647.8$+$584623 6.34 3.4 470$\pm$135 –– 2900 SMG D [*Himiko*]{} 6.60 $\sim 16$ –– –– $\sim 100$ LAE E, F [*IOK$-$1*]{} 6.96 $< 5.35$ –– –– $\sim 16$ LAE F HCM 6A 6.56 –– –– –– $9$ LAE G --------------------------------- ------ ----------------- ------------- --------- ------------------------------------------- ------ -------- [Col. (1): Source name. Col. (2): Redshift. Col. (3): Disk radius. Col. (4): Disk circular velocity. Col. (5): Disk inclination. Col. (6): Star formation rate. Col. (7): Reference: (A) @Wang13; (B) @Riechers09; (C) @Willott13; (D) @Riechers13; (E) @Ouchi13; (F) @Walter12; (G) @Hu02b. \*Predicted values based on Eq. \[eq:fits\] with $\beta=0.1$.]{} \[tab:disk\] The temperature of the gas in the absence of the CMB is given by the disk model. We combine this value with the CMB temperature at high redshift as in @MF13a assuming that all of the CMB radiation is ultimately transferred to the gas. The temperature and density of each cloud is then input into a version of the escape-probability formalism code by @KT07, where the modifications, described in @MF13a, are consistent with those implemented in the new DESPOTIC code [@Krumholz13]. Given the lack of observational constraints at high redshift, this procedure requires fewer arbitrary parameter choices than does the detailed heating and cooling balance performed by DESPOTIC and assumes a regime in which the dust and gas are tightly coupled. However, since our temperatures are consistent with those determined from observations of emission lines in high-z quasar hosts (see §\[sec:comp:temp\]), we conclude that our method is sufficient for our purposes. The escape-probability code includes the full log-normal density distribution to determine the CO and CII level populations and calculates the resulting line emission from each cloud. Throughout the rest of this Paper, we assume a carbon abundance of ${\rm C/H}=1.5\times10^{-4}\,Z'$. In considering only the emission from the PDRs of cold molecular clouds, we have ignored any contribution from HII regions, which are not well-described by our model. These ionized regions would both lower the molecular fraction of the galactic gas and increase its temperature. However, observational and theoretical evidence suggest that this is likely a small effect. As we will see in §\[sec:comp:temp\], the cold gas temperatures in our models without HII regions are comparable to those inferred from both dust and molecular line observations of quasar hosts suggesting that the hotter, ionized gas in HII regions does not contribute significantly. Moreover, in a sample of 60 normal, star-forming galaxies at relatively low redshift, @Malhotra01 attributed only about 50% of the total \[CII\] flux to ionized gas as probed by \[NII\], while @Vasta10 found somewhat less emission in more recent work. We would expect HII regions in molecular clouds to be even less important in starbursts or AGN or at much higher redshifts, at least [*[dynamically]{}*]{}, as external pressure from the ISM becomes significant [@Krumholz06]. Additionally, the assumption of no contribution from HII regions is consistent with the lack of a \[NII\] detection in J1148$+$5251 [@Walter09a]. Despite this evidence, in §\[sec:c:alt\], we will consider the effect of an alternative addition to our model in which the molecular fraction is lower than computed by equation \[eq:fmol\], which is similar to the effect of including a contribution from HII regions, albeit at lower temperatures. Finally, we arrive at the observed signal by summing the emission from all clouds in our model galaxies and subtracting the CMB as an observational background. We note that, while our model accurately determines the intrinsic cloud emission as well as the effect of the CMB background, in some cases, the two are of similar magnitude and the subtracted result is not well-determined [@MF13a]. This is particularly true for the CO(1–0) line of galaxies in low-mass halos but is not an issue for the \[CII\] line in the present work. Observed Systems {#sec:obs} ================ ---------------- ------------------------------------------------- --------------------------------- ------------------------------------------------ -------------------------------- ------ -------- Shortened Name $L_{\rm BH}$ ${L_{\rm [CII]}}$ ${L_{\rm CO(1-0)}}$ ${S\,{\Delta}v_{\rm CO(6-5)}}$ Type Refer. $10^{13}\,{\rm {\ensuremath{{\rm L_{\odot}}}}}$ $10^{9}\,{\rm Jy\,km/s\,Mpc^2}$ $10^{5}\,{\rm {\ensuremath{{\rm L_{\odot}}}}}$ Jy km/s (1) (2) (3) (4) (5) (6) (7) J2310$+$1855 9.3 64$\pm$9.0 32.1 1.52$\pm$0.13 QSO A J1319$+$0950 7.0 28$\pm$6 9.4 0.43$\pm$0.09 QSO A J2054$-$0005 2.8 22$\pm$3 7.3 0.34$\pm$0.07 QSO A J0129$-$0035 0.57 12$\pm$2 7.4 0.37$\pm$0.07 QSO A J1044$-$0125 11.6 10$\pm$3 4.2 0.21$\pm$0.04 QSO A J1148$+$5251 10 165$\pm$13 $< 7100$ 0.67$\pm$0.08 QSO B J0210$-$0456 0.53 1.94$\pm$0.265 –– –– QSO C J2329$-$0301 1 $< 0.71$ –– –– QSO C HFLS3 –– 100.$\pm$21 4800$\pm$1500 2.74$\pm$0.68 SMG D [*Himiko*]{} –– $< 0.33$ –– –– LAE E [*IOK$-$1*]{} –– $< 2.86$ –– –– LAE F [*HCM 6A*]{} –– $< 0.41$ –– –– LAE G ---------------- ------------------------------------------------- --------------------------------- ------------------------------------------------ -------------------------------- ------ -------- [Col. (1): Shortened source name. Col. (2): Bolometric black hole luminosity. Col. (3): \[CII\] line luminosity. Col. (4): CO(1–0) line luminosity. Col. (5): CO(6–5) velocity-integrated line flux. Col. (6): Source type. Col. (7): Reference: (A) @Wang13; (B) @Willott03 and @Riechers09; (C) @Willott13; (D) @Riechers13; (E) @Ouchi13; (F) @Walter12; (G) @Kanekar13. Calculated by @Wang13 from CO(6–5) observations assuming an excitation ladder similar to that of J1148$+$5251 [@Riechers13].]{} \[tab:lum\] In this section, we discuss the observed systems to which we will compare our model. These sources fall into three categories: (1) quasar hosts (QSOs), (2) sub-mm galaxies (SMGs), and (3) Lyman-$\alpha$ emitters (LAEs). Since the redshift of the system is an [*a priori*]{} requirement for observing the \[CII\] line flux (at least given current instruments), Lyman-break galaxies with no discernible line emission are not as good candidates for this type of followup at high-redshift. We plan to extend our analysis to these objects using the aggregate, unresolved intensity on the sky in forthcoming work. Table \[tab:disk\] shows the sources we have compiled from the literature and lists, where available, their derived physical properties: size, circular velocity, inclination, and star formation rate, along with a designation of the type of system. We have interpreted each object as if it were a galactic disk. We determined the disk radius for the quasar hosts and SMGs by taking the semi-major axis of the \[CII\] emission on the sky (where detected) except in the case of J1148$+$5251, whose more extended CO emission clearly traces galactic disk gas. Sizes for LAEs were derived from the extent of the Lyman-$\alpha$ emission, but we do not interpret these as disk radii since the Lyman-$\alpha$ results from scattering of material outside of the galactic disk. We computed the angular diameter distance based on the specific redshift of each object, where an angular size of 0.1" corresponds to about $5.7\,{{\ensuremath{{\rm kpc}}}}$ at $z=6$. We relate the circular velocity, ${V_{\rm c}}$, to the \[CII\] FWHM line width, ${\Delta}v_{\rm [CII]}$, by assuming ${V_{\rm c}}={\Delta}v_{\rm [CII]}$ [@Nagamine06b]. We use this relation in the absence of other inclination information and set ${V_{\rm c}}={\Delta}v_{\rm [CII]}/{\rm sin}\,i$, otherwise. Quoted errors in ${V_{\rm c}}$ combine contributions from both ${\Delta}v_{\rm [CII]}$ and ${\rm sin}\,i$ where available. The star formation rates from our sample were typically derived by assuming a ratio with observed FIR luminosity. However, no such measurements or derived star formation rates are yet available for the @Wang13 systems. Of particular difficulty is the star formation rate of HCM 6A [@Hu02a], which varies wildly when inferred from different observational indicators [see @Kanekar13 for a summary]. For this object, we assume the star formation rate inferred from UV continuum measurements of roughly $9\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr}$ [@Hu02b], which ignores the effect of dust obscuration. Table \[tab:lum\] lists, for these same systems, their luminous properties: bolometric black hole luminosity, \[CII\] luminosity, CO(1–0) luminosity, and CO(6–5) flux. Typically, the bolometric black hole luminosities were derived from measurements in the rest-frame UV assuming a bolometric correction.[^6] Direct information about the CO(1–0) line is not actually available for the sample of quasars from @Wang13, but these authors have assigned a value to each object based on the CO(6–5) data assuming a model excitation ladder similar to that of J1148$+$5251 [@Riechers09]. However, the extrapolation of high-order CO lines to lower orders is quite uncertain [e.g. @CW13]. Particularly pertinent to comparison with our models is the presence (or lack thereof) of an AGN in HFLS3. The absence of AGN activity was determined from the galaxy’s lack of excess radio emission, hot dust component, rest-frame optical continuum emission up to $2.4\,{\rm \mu m}$ beyond that associated with the starburst, and broad rest-frame UV lines [@Riechers13 and private communication with the authors]. However, these criteria may not be sufficient to completely rule out the presence of a weak AGN [@Rush96; @Spinoglio02; @EM12]. If HFLS3 does indeed have such a faint nuclear component, we can ask at what level such a component would remain undetected. Combining the measured 2.2$\,\mu$m and 3.6$\,\mu$m emission with the recommended bolometric correction fits of @Runnoe12, we conservatively estimate the bolometric black hole luminosity to be less than about $3\times10^{12}\,{\ensuremath{{\rm L_{\odot}}}}$. For comparison, J0129$-$0035 was detected in an extremely deep stripe of SDSS data with a rest-frame UV magnitude of 22.28 [@Jiang09]—well-below the typical SDSS-DR9 $z$-band detection limit of 20.5—corresponding to an isotropic bolometric black hole luminosity of $5.7\times10^{12}\,{\ensuremath{{\rm L_{\odot}}}}$. We nominally set a maximum luminosity of $3\times10^{12}\,{\ensuremath{{\rm L_{\odot}}}}$ on an AGN component of HFLS3. Where necessary, we have assumed a maximum black hole luminosity for the LAEs in our sample of $10^{12}\,{\ensuremath{{\rm L_{\odot}}}}$, small enough that the rest-frame UV emission in these systems is still dominated by starlight. Comparison of Model Galactic Disks to Observations {#sec:comp} ================================================== In this section, we compare our model galactic disks to observed systems with respect to their disk structure and dynamics (§\[sec:comp:struc\]), quasar luminosities (§\[sec:comp:lbh\]), gas masses (§\[sec:comp:mgas\]), and gas temperatures (§\[sec:comp:temp\]). Structure and Dynamics {#sec:comp:struc} ---------------------- Here, we compare the structure and dynamics of our model disks to those observed and show that our simple framework describes the data reasonably well. We first consider disk radii by assuming a constant spin parameter of $\lambda=0.05$. The outer radius of the disk is then given by $$\label{eq:rdisk} {R_{\rm d}}=\frac{\lambda\,{R_{\rm vir}}}{\sqrt{2}}=1.65\,{\ensuremath{{\rm kpc}}}\,\left(\frac{{M_{\rm h}}}{10^{12}\,{\ensuremath{{\rm M_{\odot}}}}}\right)^{1/3}\,\left(\frac{1+z}{7}\right)^{-1},$$ where ${R_{\rm vir}}$ is the virial radius of a dark matter halo with mass ${M_{\rm h}}$ at redshift $z$. The circular velocity of a model disk is taken to be $$\label{eq:vcirc} {V_{\rm c}}=\sqrt{2}\,\sigma=303\,{\ensuremath{{\rm km\,s}^{-1}}}\,\left(\frac{{M_{\rm h}}}{10^{12}\,{\ensuremath{{\rm M_{\odot}}}}}\right)^{1/3}\,\left(\frac{1+z}{7}\right)^{1/2},$$ where $\sigma$ is the halo velocity dispersion [@BL01]. Inspection of equations \[eq:rdisk\] and \[eq:vcirc\] indicates that the ratio of ${V_{\rm c}}$ to ${R_{\rm d}}$ will be independent of mass and fixed at a given redshift. Figure \[fig:disk\] shows this relationship at $z=6$. ![\[fig:disk\] Disk circular velocity, ${V_{\rm c}}$, as a function of disk radius, ${R_{\rm d}}$, for quasar hosts and SMGs at $z\sim6$. The solid line indicates our model derived from Eq. \[eq:rdisk\] and \[eq:vcirc\] at $z=6$, while vertical, dotted lines denote disk radii corresponding to $10^{12}$ and $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$. The points show observations from Table \[tab:disk\] where available. Circles correspond to the quasar sample of @Wang13, the pentagon is J1148$+$5251, the hexagon is J0210$-$0456, and the triangle is the SMG HFLS3. Filled points indicate observations corrected for inclination. ](disk.eps){width="\columnwidth"} ![\[fig:rdist\] The star formation rate surface density (top panel) and mass inflow rate (bottom panel) as a function of galactocentric radius, $r$, in our disk model for a host halo mass of $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ at $z=6$. Results are plotted out to the disk radius of approximately $3.55\,{\ensuremath{{\rm kpc}}}$. We show the behavior of a range of angular momentum transport models: short-dashed (green), solid (magenta), and dot-short-dashed (blue) lines denote $\beta$-models with $\beta=1$, 0.1, and 0.01, while long-dashed (red) and dot-long-dashed (brown) curves indicate LSW models with $m=1$ and 0.02, respectively. The thick, solid line in the upper panel references a proportional relationship with $r^{-2}$. ](rdist.eps){width="\columnwidth"} Our model disk sizes and circular velocities appear to describe some of the objects in our sample very well. Systems for which inclinations are available (filled symbols) are particularly well-represented. On the other hand, objects for which we do not have inclinations (open symbols) are also those most off-set from our model relation. Moreover, these objects are all below the relation so that incorporating an inclination would likely improve agreement in each case. Additionally, we note that the errors on the size of J0210$-$0456 (hexagon) are particularly large making a comparison difficult. The vertical, dotted lines in the figure indicate disk radii for halo masses of $10^{12}$ and $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ at $z=6$ according to equation \[eq:rdisk\]. Judged purely on inferred disk radius (to avoid the additional complication of missing inclinations), the objects in our sample all have halo masses roughly in this range. The nature of our models allows us to consider the disk structure in even finer detail. Figure \[fig:rdist\] plots the radial distribution of star formation and gas inflow in our model for a $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ halo at $z=6$ and a range of angular momentum transport models. We specifically consider the range of transport models summarized in Table \[tab:amt\]. Models with faster inflow rates of gas maintain higher star formation rates into the inner regions of the galactic disk [see also @Thompson05; @MF12; @MF13a]. Additionally, for $\beta \gtrsim 0.1$, ${\dot{\Sigma}_{\star}}$ peaks at a radius of $\sim100\,{\rm pc}$ reaching nearly $10^{3}\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr/kpc^2}$. This peak is qualitatively similar to the extreme nuclear star formation rate densities noted in HFLS3—${\dot{\Sigma}_{\star}}\sim600\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr/kpc^2}$ distributed over a $1.3\,{\ensuremath{{\rm kpc}}}$ radius region [@Riechers13]—and J1148$+$5251—$\sim10^3\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr/kpc^{2}}$ over $\sim750\,{\rm pc}$ [@Walter09b]—two of the most massive systems in our sample as determined by galactic radius. This behavior in our models is not a consequence of the opacity gap discussed in @Thompson05 and appears even when the opacity is set to be constant. Rather, it results simply from a mass inflow rate and inflow velocity that are roughly constant with radius. As a result, the disk scale height $h \propto r$ and the star formation surface density required to maintain marginal Toomre-instability and vertical hydrostatic equilibrium is ${\dot{\Sigma}_{\star}}\propto r^{-2}$. The star formation rate turns over at very small radii as thermal pressure from the gas becomes more important than stellar feedback [@Thompson05; @MF12]. As radius decreases toward the center of the disk, once star formation is no longer necessary for the stability of the disk, its value drops, no further gas is depleted by star formation or winds, and the mass inflow rate becomes constant. The behavior of constant inflow rate at small radii is shown in the lower panel of Figure \[fig:rdist\]. Our model assumes that all remaining inflowing gas is accreted onto the central black hole and powers an AGN. ![image](lbh.eps){width="\textwidth"} [lcc]{}\ model name & parameter & possible galaxy types\ $\beta$-model & $\beta=0.01$ & QSOs, SMGs\ $\beta$-model & $\beta=0.1$ & QSOs, SMGs\ $\beta$-model & $\beta=1$ & QSOs, SMGs\ LSW & $m=0.2$ & LBGs, LAEs\ LSW & $m=1$ & LBGs, LAEs\ The consistency between our simple disk model and the structural features and relationships of high-redshift quasars is somewhat surprising given the picture of these systems as the product of rapid mergers and disruptions [e.g., @Li07]. However, this agreement supports our continued use of the model to describe quasar luminosities, star formation rates, and molecular line luminosities in the subsequent sections. Quasar Luminosities {#sec:comp:lbh} ------------------- Next, we consider the luminosity of central black holes in our model galaxies and the ability of this framework to describe the relationships between observed $z\sim6$ quasars and their host properties. Figure \[fig:lbh\] plots the bolometric luminosities of our central black holes, ${L_{\rm BH}}$, versus ${R_{\rm d}}$, ${V_{\rm c}}$, and the total star formation rate, respectively. Solid curves assume that the disk gas inflow is non-linear (i.e. for a $\beta$-model), while dashed curves show results for an LSW disk. Table \[tab:amt\] summarizes the range of angular momentum transport models that we consider in this work. We find that using the optically thick dust model of @Thompson05 produces negligible differences from the dust-free model of @MF12 for rapid angular momentum transport, but the two begin to deviate for models where the inflow rate is slower with the dusty model gives somewhat more black hole growth than in the dust-free case if one assumes an LSW disk. The figure shows that, as expected, the faster transport of angular momentum in the $\beta$-model produced significantly more accretion onto the central black hole than does that for the LSW models for the parameter ranges we considered. Here, we present fits to our model results to aid in future predictions.[^7] The fits are described by the form $$\label{eq:fits} y/y_0=\zeta^{a}\,\left(\frac{x}{x_0}\right)^{b+c\,{\rm log}_{10}\,\zeta},$$ with $y/y_0$, $x/x_0$, $a$, $b$, and $c$ given in Table \[tab:fits\]. For nonlinear $\beta$-models, $\zeta=\beta$, while for LSW models, $\zeta=m$. Inflow model $y/y_0$ $x/x_0$ $a$ $b$ $c$ --------------- ----------------------------------------------------------------- ----------------------------------------------------------- ------ ------ ------- $\beta$-model ${L_{\rm BH}}/10^{14}\,{\ensuremath{{\rm L_{\odot}}}}$ ${R_{\rm d}}/{\ensuremath{{\rm kpc}}}$ 1.5 3.3 0.4 $\beta$-model ${L_{\rm BH}}/1.3\times10^{13}\,{\ensuremath{{\rm L_{\odot}}}}$ ${V_{\rm c}}/100\,{\rm km\,s^{-1}}$ 1.5 3.3 0.4 $\beta$-model ${L_{\rm BH}}/2.4\times10^{13}\,{\ensuremath{{\rm L_{\odot}}}}$ ${\rm SFR}/{\rm {\ensuremath{{\rm M_{\odot}}}}\,yr^{-1}}$ 1.68 0.8 0.086 LSW ${L_{\rm BH}}/4.5\times10^{10}\,{\ensuremath{{\rm L_{\odot}}}}$ ${R_{\rm d}}/{\ensuremath{{\rm kpc}}}$ 1.56 0.9 0.0 LSW ${L_{\rm BH}}/2.3\times10^{10}\,{\ensuremath{{\rm L_{\odot}}}}$ ${V_{\rm c}}/100\,{\rm km\,s^{-1}}$ 1.56 0.98 0.0 LSW ${L_{\rm BH}}/2.0\times10^{10}\,{\ensuremath{{\rm L_{\odot}}}}$ ${\rm SFR}/{\rm {\ensuremath{{\rm M_{\odot}}}}\,yr^{-1}}$ 1.57 0.18 0.0 For a $\beta$-model, we find that ${L_{\rm BH}}$ is nearly proportional to halo mass. Moreover, for $\beta>0.1$, the inflow rate becomes high enough that the amount of gas bypassing star formation to fuel the central black hole becomes significant, and the star formation rate loses its proportionality with halo mass. This effect is shown in panel (c) of Figure \[fig:lbh\], where the squares indicating $10^{12}$ and $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ halos are at lower star formation rates for $\beta=1$ than are those for $\beta=0.1$. While the lower inflow velocities associated with LSW models result in dependences on ${R_{\rm d}}$, ${V_{\rm c}}$, and star formation rate that are not quite as well-fit by power-laws as those for a $\beta$-model, we nevertheless attempt to describe the behavior as such for interpolation purposes. However, we caution against using these fits to extrapolate beyond the range plotted in Figure \[fig:lbh\]. We find that ${L_{\rm BH}}$ is a much shallower function of halo mass ($\propto {M_{\rm h}}^{\sim 1/3}$) than for $\beta$-models, which implies that the BH fueling is not supply-limited. Note that we obtain better descriptions of the results for each independent variable by fitting each directly from the plots in Figure \[fig:lbh\] than by assuming any knowledge of the true relationships between the variables (for example, by using equations \[eq:rdisk\] and \[eq:vcirc\]) owing to the different parameter ranges over which we fit the results. As can be seen in Figure \[fig:lbh\], the relationships between $L_{\rm BH}$ and the disk parameters of ${R_{\rm d}}$, ${V_{\rm c}}$, and star formation rate for observed quasar hosts are well-matched by a $\beta$-model disk with $\beta$ slightly less than 0.1 and halo masses in the range of $10^{12}$–$10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$. These sources produce black hole luminosities well in excess of that expected in any of the LSW models, which appear to be ruled out as descriptions for quasar hosts. This is in contrast to more typical galaxies, such as LBGs and LAEs, which can only be represented by $\beta$-model systems if a significant amount of the resulting black hole emission is obscured [@MF12]. Given the agreement between our model and the data, we use equation \[eq:fits\] and the fit parameters in Table \[tab:fits\] with $\beta=0.1$ to predict star formation rates for the @Wang13 sample of quasars based on their bolometric black hole luminosities and list these predictions in Table \[tab:disk\] for comparison with upcoming data. While Figure \[fig:disk\] showed several disagreements between our sample and our model ${V_{\rm c}}$-${R_{\rm d}}$ relation, particularly for J1044$-$0125 and J1148$+$5251, we note that the relationship between ${L_{\rm BH}}$ and ${R_{\rm d}}$ for these systems is generally consistent with that of the other quasars in Panel (a) of Figure \[fig:lbh\]. Consequently, we expect that the discrepancies in Figure \[fig:disk\]—and consequently in panel (b) of Figure \[fig:lbh\]—indeed owe simply to an under-estimated ${V_{\rm c}}$ resulting from a inclined disks. Interestingly, both CFHQS quasars that we consider—J0210$-$0456 (hexagon) and J2329$-$0301 (upper limit on star formation rate in panel c)—are consistent with our model if $\beta=0.1$, though the radius of J0210$-$0456 may be somewhat overestimated—not completely unexpected given the large errors on the observed size. While @Willott13 suggested that J2329$-$0301, in particular, may be observed in a rare state of quasar-suppressed star formation, we obtain agreement with our model with no such feedback mechanism. J2329$-$0301 appears to have a smaller star formation rate simply because it is hosted in a smaller system Given its apparent lack of quasar activity, HFLS3 (upper limit on ${L_{\rm BH}}$) is inconsistent with a $\beta$-model with $\beta>0.01$. However, assuming the data allow an undetected bolometric luminosity of $3\times10^{12}\,{\ensuremath{{\rm L_{\odot}}}}$, we cannot rule out a $\beta$-model description of HFLS3 with $\beta=0.01$. Of course, a variety of uncertainties in both our models and in the observations may cause shifts in the plotted quantities. For example, our model assumes that all gas not turned into stars or ejected by stellar radiation-pressure driven winds accretes onto the central black hole and that all of the resulting radiation escapes. Thus, our model black hole luminosities may be over-estimates (1) if angular momentum is not shed quite so quickly in the central portions of the disk, resulting in the buildup of a bulge [e.g. @Dekel09b] rather than direct accretion onto the black hole, (2) if a wind from the central AGN—inferred in some observations by a broad component of \[CII\] [@Maiolino12; @Valiante12] and not included in our model—blows out gas from the center that would otherwise be accreted, (3) if AGN feedback results in a reduced gas accretion rate onto the galactic disk [e.g., @Voort11], or (4) if radiation trapping [e.g., @WL11b] or obscuration prevents some of the emission from the accreted gas from escaping. Including a significant contribution from any of these effects would result in lower values of ${L_{\rm BH}}$ in our model and would require a larger value of $\beta$ to reproduce the observations. At the same time, the bolometric corrections used by @Wang13 are subject to their own uncertainties and extrapolations, and the total luminosity is over-estimated if the emission is not isotropic [@Runnoe12]. In this latter case, the observations could be better described by our models with a lower value of $\beta$. Additionally, the relationship between FIR luminosity and star formation rate is only empirically calibrated at low redshift and may not hold at $z\sim6$. Ultimately, however, if any of the above effects are significant, they would all have to conspire fairly precisely to produce the agreement we find between most of the quasars in our sample and a disk model with fixed $\beta$. Gas Masses {#sec:comp:mgas} ---------- The mass in cold gas in our disks is critical for the calculation of CO and \[CII\] lines since more gas will imply more emission. We would naively expect the amount of cold gas to be approximately the same as the mass in stars, which in turn has been estimated via abundance matching to be roughly 1% of the host halo mass [e.g., @Behroozi13]. However, without making any additional assumptions, we can calculate the gas mass from our disk models by simply integrating the density required to ensure $Q=1$ at each radius. Figure \[fig:mgas\] shows the total gas masses at $z=6$ in our models for both LSW and $\beta$-model gas transport and different values of $m$ and $\beta$ as a function of halo mass. The masses have been scaled by halo mass and the cosmic baryon fraction to give the fraction of halo baryons in the cold phase of the disk. ![\[fig:mgas\] Fraction of halo baryons in the cold gas disk as a function of halo mass at $z=6$. As in Fig. \[fig:rdist\], short-dashed (green), solid (magenta), and dot-short-dashed (blue) lines denote results for $\beta$-models with $\beta=1$, 0.1, and 0.01, while long-dashed (red) and dot-long-dashed (brown) curves indicate results for LSW models with $m=1$ and 0.02, respectively. The gas mass inferred for HFLS3 by @Riechers13 is noted by the triangle assuming a host halo mass of $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$. The horizontal, dotted line marks the typical fraction of halo baryons in stars at $z\sim6$ from @Behroozi13. ](mgas.eps){width="\columnwidth"} Our model results are not very sensitive to choice of model or parameter values—less than a factor of a few variation among all models considered—and are consistent with calculations of the mass of the cold neutral medium from the numerical simulations of @Nagamine06b and approximately equal to the stellar masses estimated by @Behroozi13. The size, velocity, and star formation rate of HFLS3 are roughly consistent with its being hosted by a halo of mass just under $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$. @Riechers13 find a total molecular mass of $(1\pm0.09)\times10^{11}\,{\ensuremath{{\rm M_{\odot}}}}$ (assuming $\alpha_{\rm CO}=1$) and $2.0\times10^{10}\,{\ensuremath{{\rm M_{\odot}}}}$ worth of atomic gas in the SMG. Together these two components represent a total gas mass (denoted by the triangle in Fig. \[fig:mgas\]) about 40% larger than that in our $m=1$ LSW model for $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$. This is remarkably good agreement given the simplicity of our models and the uncertainties in $\alpha_{\rm CO}$. Using the universal model for the CO to H$_2$ conversion factor in @Narayanan12, we estimate a value of $\alpha_{\rm CO}\approx0.84$ would be appropriate for HFLS3 given its CO luminosity and assuming solar metallicity gas.[^8] Such a value brings the total inferred gas mass even closer to our estimate. ![image](Tave.eps){width="\textwidth"} Gas Temperatures {#sec:comp:temp} ---------------- Similar to the gas mass, the gas temperature is also an important ingredient in determining the luminosity of emission lines. While the temperature varies as a function of radius throughout our model disks, a single value is often inferred observationally for the emitting gas in an entire galaxy. Therefore, we calculate an equivalent temperature from our models for each halo mass and angular momentum transport model. First, we calculate the \[CII\] luminosity assuming thermal, optically thin gas at the temperature prescribed by our disk model: $$\label{eq:Lcii_th} {L_{\rm [CII]}}=\int_0^{{R_{\rm d}}}\!\! \frac{2\,{\rm \pi}\,r\,\left(1-{f_{\rm CO}}\right)\,{\rm C/H}}{m_{\rm p}}\,\frac{4\,{\rm e}^{-{\ensuremath{{h_{\rm p}}}}\,\nu_{\rm [CII]}/k_{b}\,T}}{2+4\,{\rm e}^{-{\ensuremath{{h_{\rm p}}}}\,\nu_{\rm [CII]}/k_{b}\,T}}\,A_{\rm [CII]}\,{\ensuremath{{h_{\rm p}}}}\,\nu_{\rm [CII]}\,{\Sigma_{\rm g}}\,dr,$$ where $A_{\rm [CII]}=2.3\times10^{-6}\,{\rm s^{-1}}$, $\nu_{\rm [CII]}=1900.5\,{\rm GHz}$, the carbon abundance is ${\rm C/H}=1.5\times10^{-4}\,Z'$, $m_{\rm p}$ is the proton mass and where ${f_{\rm CO}}$, $T$, and ${\Sigma_{\rm g}}$ are each functions of the galactocentric radius, $r$. We then determine the temperature, $\bar{T}$, at which this result is reproduced by equation \[eq:Lcii\], taking the gas mass, ${M_{\rm gas}}$, and the total fraction of gas containing CO, $\bar{f}_{\rm CO}$, from our disk model. We plot this average temperature as a function of total star formation rate (left panel) and bolometric black hole luminosity (right panel) in Figure \[fig:Tave\]. To compare $\bar{T}$ to observations, we note that the dust and gas are implicitly assumed to be strongly coupled at the same temperature in our model, and observationally, the dust temperature traces the gas temperature in most systems [@Malhotra01 but see discussion of HFLS3]. Moreover, if ionized gas is a significant source of \[CII\] emission, then the observed dust temperature should include contain a contribution from these hotter pockets unless HII regions are completely devoid of dust. The estimated temperature of the dust responsible for the strong 250 GHz continuum emission observed in roughly 30% of $z=6$ SDSS quasars by the MAMBO survey [@Bertoldi03a; @Petric03; @Wang11b] is in the range of 40–60${\rm K}$ [see also @Wang13]. We fiducially set an approximate span of star formation rates for these systems from J0210$-$0456 and J1148$+$5251, the quasars with the lowest and highest measured values in our sample of quasars (though the undetermined values for J0129$-$0035 and J2329$-$0301 may turn out to be lower). Similarly, the faintest and brightest bolometric black hole luminosities in our sample range from 0.53–11.6$\times 10^{13}\,{\rm {\ensuremath{{\rm L_{\odot}}}}}$ (from J0210$-$0456 and J1044$+$5251, respectively). In Figure \[fig:Tave\], we compare this set of observed values to our model calculations of $\bar{T}$ and find good agreement with the quasar observations independent of the specifics of angular momentum transport. Because our model does not allow the properties of the central emitting black hole to affect the gas in the outer parts of the galactic disk, this consistency implies that quasar heating of galactic gas—at least of the gas that dominates the \[CII\] emission—is minimal. Note that in the right panel, we do not calculate $\bar{T}$ beyond $10^{14}\,{\rm {\ensuremath{{\rm M_{\odot}}}}}$. The agreement between our model and the observed quasars is in contrast to the discrepancy with regard to HFLS3 (triangle). @Riechers13 infer a much higher kinetic gas temperature—$144^{+59}_{-30}\,{\rm K}$—than predicted by any of our models. However, the fact that the SED-derived dust temperature is so much lower than that of the gas—$56^{+9}_{-12}\,{\rm K}$—suggests that our coupled dust-gas model may be poor description of this system. We further compare our results to the star burst and AGN contributions to the gas temperature assumed by @Obreschkow09b: $T_{\rm SB}=T^{\rm max}_{\rm SB}\,\left[\dot{\Sigma}_{\rm SF}/\left(\dot{\Sigma}_{\rm SF}+\dot{\Sigma}^{\rm c}_{\rm SF}\right)\right]^{1/4}$ and $T_{\rm AGN}=T^{\rm max}_{\rm AGN}\,\left[\dot{M}_{\rm BH}/\left(\dot{M}_{\rm BH}+\dot{M}^{\rm c}_{\rm BH}\right)\right]^{1/4}$, respectively, where $T^{\rm max}_{\rm SB}=60\,{\rm K}$, $\dot{\Sigma}^{\rm c}_{\rm SF}=500\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr/kpc^2}$, $T^{\rm max}_{\rm SB}=150\,{\rm K}$, and $\dot{M}^{\rm c}_{\rm AGN}=10\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr}$. For this comparison, we calculate the average star formation rate surface density of a galaxy as $\dot{\Sigma}_{\rm SF}={{\rm SFR}}/{\rm \pi}\,{R_{\rm d}}^2$ and the black hole accretion rate as $\dot{M}_{\rm BH}={L_{\rm BH}}/(0.08\,c^2)$.[^9] [^10] $T_{\rm AGN}$ clearly overestimates the temperature of the gas and dust in these systems, while $T_{\rm SB}$ under-estimates it. These deficiencies are likely symptomatic of over-extrapolation from fits at lower redshift. However, we note that the kinetic temperature of HFLS3 may be reproduced by the @Obreschkow09b model if this system were, in fact, a quasar with ${L_{\rm BH}}\gtrsim10^{13}\,{\ensuremath{{\rm L_{\odot}}}}$. While our model considers in detail the emission from the PDRs of cold molecular clouds, recent works have attributed the \[CII\] emission at high-redshift to smoother, warmer gas components—$\gtrsim200\,{\rm K}$ [@Gong12; @Vallini13]. While these studies considered the emission from LBGs and LAEs rather than from quasars, their temperatures still exceed those found in the much larger systems. This suggests that PDRs of molecular clouds are indeed the dominant source of \[CII\] emission at high redshift. Carbon Emission Lines {#sec:c} ===================== In the previous section, we showed that our disk models provide reasonable descriptions for the observed systems considered in this work. We are now in a position to calculate the amount of \[CII\] emission in our model galaxies. In §\[sec:c:cii\], we present results from the fiducial model described in §\[sec:model\] and compare them with observations. Then, in §\[sec:c:alt\], we consider changes to our base model suggested by the observations to improve agreement. Finally, we will turn our attention the line luminosities of CO in§\[sec:c:co\]. The Fiducial \[CII\] Model {#sec:c:cii} -------------------------- Combining the gas masses and temperatures described in §\[sec:comp:mgas\] and §\[sec:comp:temp\] with the physical properties of our molecular cloud PDRs, Figure \[fig:cii-mhalo\] shows the resulting model \[CII\] luminosities for $z=6$ quasar hosts as a function of halo mass. Here we assume $\beta=0.1$ and $Z'=1$ to demonstrate the effect of several different assumptions on the calculation. ![\[fig:cii-mhalo\] \[CII\] luminosity as a function of halo mass for our model quasar hosts. The solid (magenta) curve shows results for a fiducial model of $\beta$-disks with $\beta=0.1$ and $Z'=1$, where the emission has been calculated using our modified version of the @KT07 escape probability code. The short-dashed and long-dashed curves explore the effects of cumulatively ([*not*]{} alternative) assuming a thermal populated of states and optically thin gas. Thick, black curves assume an optically thin, thermal population at a fixed temperature of 50 K and including either all disk gas (upper) or only gas containing CII (lower). Finally, these models can be compared to the approximate 10 hour ALMA detection threshold for \[CII\] at $z=6$—extrapolated from the Cycle 0 configuration of @Wang13—denoted by the dotted curve. ](cii_mhalo.eps){width="\columnwidth"} The upper, thick, long-dashed line in the figure shows the simplest method for calculating the \[CII\] luminosity; the emission is derived from equation \[eq:Lcii\] assuming thermal, optically thin gas with $\bar{f}_{\rm CO}=0$ and $\bar{T}=50\,{\rm K}$. The choice of $\bar{f}_{\rm CO}=0$ implies that carbon everywhere in the galaxy is in the form of dissociated CII. However, at the surface densities of halos larger than about $10^{11}\,{\ensuremath{{\rm M_{\odot}}}}$, the gas is expected to be nearly all in molecular form [@Krumholz09b] and, at solar metallicity, nearly all of the carbon is in the form of CO [@Wolfire10 Eq. \[eq:fco\]] with only a small fraction in CII. Using $\bar{f}_{\rm CO}$—calculated properly according to our model via equations \[eq:fmol\] and \[eq:fco\]—results in the lower, thick, long-dashed curve in the figure and represents the largest effect toward accurately determining the amount of \[CII\] emission. The effect is stronger at higher halo masses as the commensurately higher surface densities keep more of the carbon in the form of CO. At low metallicities, there is much less dust in our model, which results in lower values of ${f_{\rm CO}}$, but the net \[CII\] emission ultimately decreases because, while $(1-\bar{f}_{\rm CO})$ asymptotically approaches unity, the total amount of carbon scales as $Z'$. However, at high metallicities, the \[CII\] emission also decreases because $(1-\bar{f}_{\rm CO})$ decreases faster than the carbon abundance increases. Continuing to build up the complexity of the calculation in Figure \[fig:cii-mhalo\], the thin, long-dashed (magenta) curve shows the emission determined using equation \[eq:Lcii\_th\]; the gas is still assumed to be thermal and optically thin, but now the gas temperature is allowed to varying according to our disk model. Note that the approximate agreement between equations \[eq:Lcii\_th\] and \[eq:Lcii\] for $\bar{T}=50\,{\rm K}$ was previously indicated by the flatness of the curves in Figure \[fig:Tave\]. The thin, short-dashed (magenta) curve in Figure \[fig:cii-mhalo\] further improves the calculation of the \[CII\] emission by relaxing the assumption of optically thin gas. This additional complication does not significantly decrease the predicted emission. Finally, the full calculation is denoted by the thin, solid (magenta) line, which shows a very similar prediction when the assumption of thermal emission is removed—less than a factor of two decrease in $10^{11}\,{\ensuremath{{\rm M_{\odot}}}}$ halos. Thus, Figure \[fig:cii-mhalo\] demonstrates that our model luminosity can be approximated—at the large halo masses considered here and for a $\beta=0.1$ model—by assuming thermal, optically thin gas at a fixed temperature of $\sim 50\,{\rm K}$. These assumptions get worse as halo mass decreases, and care should be taken when considering the high-redshift LBGs hosted by smaller systems in future work. ![image](cii1.eps){width="\textwidth"} ![image](cii2.eps){width="\textwidth"} In Figures \[fig:cii1\] and \[fig:cii2\], we show the \[CII\] luminosity assuming solar metallicity ($Z'=1$) as a function of disk radius, total star formation rate, and flux in the CO(6–5) line for a range of angular momentum transport parameter-space: $\beta$-models with $\beta=0.01$, 0.1, and 1 (Fig. \[fig:cii1\]) and LSW models with $m=0.2$ and 1 (Fig. \[fig:cii2\]). As discussed in §\[sec:comp:lbh\], we expect the $\beta$-models of Figure \[fig:cii1\] to be more descriptive of quasar hosts with $\beta$ close to 0.1 (middle row) if we neglect quasar outflows and 1 (top row) if outflows are significant. On the other hand, the LSW models of Figure \[fig:cii2\] are likely more representative of typical galaxies such as LBGs and LAEs, though perhaps at lower metallicities. In both figures, solid curves show our fiducial calculations as described in §\[sec:model\], which we denote “Model 1." We find that Model 1 under-predicts the observed \[CII\] luminosities in our sample by approximately an order-of-magnitude on average at fixed ${R_{\rm d}}$ or star formation rate. At fixed ${S\,{\Delta}v_{\rm CO(6-5)}}$, the agreement is even worse, but we expect that this is due to a corresponding over-production of CO(6–5) emission (see §\[sec:c:co\]). We anticipated this degree of discrepancy between our models and the observations in §\[sec:intro:cii\]. Simply increasing the metallicity beyond $Z_{\odot}$ will produce more total carbon, but the implied increase in dust content will put a higher fraction of the carbon in the form of CO. In §\[sec:c:alt\], we will consider two interesting departures from our fiducial model suggested by the observations to improve agreement; these alternatives are indicated by short- and long-dashed lines in the figures. [lcc]{}\ model name & implementation & possible physical cause\ Model 1 & described in §\[sec:model\] & fiducial case\ Model 2 & multiply ${{\rm SFR_{\rm ff}}}/{t_{\rm ff}}$ & higher star\ & by factor of 10 & formation efficiency\ Model 3 & divide ${f_{\rm CO}}/{f_{\rm H_{2}}}$ & lower dust-to-\ & by factor of 10 & metals ratio\ In the central columns of Figures \[fig:cii1\] and \[fig:cii2\], we compare our sample and Model 1 results to the local scaling relation of \[CII\] luminosity with star formation rate [@deLooze11; @Ouchi13]. The quasar measurements and that of HFLS3 are consistently below the average local relation but still within the observed scatter. However, @Ouchi13 recently placed a 3$\sigma$ upper limit on the \[CII\] luminosity of [*[Himiko]{}*]{}—${L_{\rm [CII]}}<3\times10^{8}\,{\rm Jy km/s Mpc^2}$—and determined an estimated star formation rate of $100\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr}$, which place the LAE significantly below the local relation. These authors suggested that such an offset implies a metallicity for [*[Himiko]{}*]{} below 10% of solar. If Model 1 is an accurate description of LAEs—despite its deficiencies with respect to quasars—then the metallicity of [*[Himiko]{}*]{} could be as high as solar if the inflow rate of gas is quite low $\beta\lesssim0.01$ (consider the bottom, central panel of Fig. \[fig:cii1\]) or $m\lesssim0.2$ (bottom, central panel of Fig. \[fig:cii2\]). However, we showed in @MF13a that a consistent picture of the molecular fraction in chemical equilibrium in typical $z\sim6$ galaxies implies gas-phase metallicities of only a few percent of solar. Thus, the likeliest scenario is that the metallicity of [*[Himiko]{}*]{} is indeed very low and that Model 1 does a poor job of representing the \[CII\] emission across our whole sample. Alternative Molecular Gas Models {#sec:c:alt} -------------------------------- Given that we roughly predict the correct gas masses, star formation rates, and temperatures of these systems, the disagreement between observations and Model 1 must lie in the ratio of \[CII\]-traced to CO-traced gas. Figure \[fig:cii-mhalo\] indicated that we can increase the amount of \[CII\] emission by shifting more of the carbon from CO into dissociated CII by decreasing either ${f_{\rm H_{2}}}$ or ${f_{\rm CO}}/{f_{\rm H_{2}}}$ or both. A decrease in ${f_{\rm H_{2}}}$ is represented in our model by an increase in the efficiency of star formation from molecular gas—thus maintaining the ${\dot{\Sigma}_{\star}}$ required at each radius to ensure $Q=1$—but may physically be a result of the destruction of molecular clouds before the equilibrium value of ${f_{\rm H_{2}}}$ is reached [@MLG12] and star formation occurring in atomic gas [@Krumholz12]. We, thus, define “Model 2" as one in which ${{\rm SFR_{\rm ff}}}/{t_{\rm ff}}$ from equation \[eq:fmol\] increased by a factor of 10 (short-dashed curves in Figs. \[fig:cii1\] and \[fig:cii1\]). We further consider a model in which we instead manipulate the fraction of gas containing CO directly, dividing ${f_{\rm CO}}/{f_{\rm H_{2}}}$ given in equation \[eq:fco\] by a factor of 10, and denote this variation as “Model 3" (long-dashed curves). However, it is worth keeping in mind that both effects could be happening simultaneously to some degree. Table \[tab:mod\] summarizes the three different molecular gas models we consider in this work. As shown in Figures \[fig:cii1\] and \[fig:cii2\], the predicted \[CII\] emission increases substantially in Models 2 and 3. For Model 2, the increase in the \[CII\] luminosity is sufficient to explain the observations only if gas is transported very quickly through the disk with $\beta=1$. On the other hand, Model 3 reproduces the observations more generically; the relationships between \[CII\] emission and either star formation rate or ${S\,{\Delta}v_{\rm CO(6-5)}}$ are more robust to changes in the angular momentum transport mechanism or parameter values. In this case, with respect to star formation rate, we find, $$\label{eq:lcii-sfr} {L_{\rm [CII]}}=5\,\times10^{8}\,{\ensuremath{{\rm L_{\odot}}}}\,\left(\frac{{\rm SFR}}{100\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr}}\right)^{0.9}.$$ As a function of CO flux, all but one of the models give results well-presented by $$\label{eq:lcii-lco6} {L_{\rm [CII]}}=6\,\times10^{9}\,{\ensuremath{{\rm L_{\odot}}}}\,\left(\frac{{S\,{\Delta}v_{\rm CO(6-5)}}}{{\rm Jy\,km/s}}\right)^{0.9}$$ with the values for $\beta=1$ being about 0.3 dex brighter. As discussed in §\[sec:c:cii\], the upper limits on the LAE [*Himiko*]{} are inconsistent with Models 2 and 3 for $Z'=1$. Similarly, HCM 6A (central panels; SFR$=9\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr}$; ${L_{\rm [CII]}}<4.1\times10^{8}\,{\rm Jy km/s Mpc^2}$) may only be in marginal agreement with Models 2 and 3 at solar metallicity (consider the bottom, central panel of Fig. \[fig:cii2\]). A stronger disagreement would result if the star formation rate for this LAE is significantly higher than that estimated from its UV continuum. However, this tension may simply indicate a metallicity for HCM 6A somewhat less than solar. In Figure \[fig:lbh-cii\], we combine our \[CII\] calculations from Figures \[fig:cii1\] and \[fig:cii2\] with our black hole accretion rate results from Figure \[fig:lbh\]. We plot \[CII\] luminosity versus bolometric black hole luminosity for Models 1, 2, and 3. Model 3 with $\beta \approx 0.1$ clearly does the best job of describing the observed quasars for reasonable halo masses ($10^{12}$–$10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$). Here, HFLS3 seems more in tension with $\beta$ even as low as 0.01. If SMGs in halos comparable to those of quasars are actually better-described by a different mode of angular momentum transport (in this case, by LSW disks), it may have implications for the quasar duty cycle and the ability of systems to form supermassive black holes by $z\sim7$ [@Mortlock11]. ![\[fig:lbh-cii\] Bolometric black hole luminosity as a function of \[CII\] luminosity. This figure combines results from Fig. \[fig:lbh\], \[fig:cii1\], and \[fig:cii2\]. As in Fig. \[fig:lbh\], green, magenta, and blue lines correspond to $\beta$-models with $\beta=1$, 0.1, and 0.01, while red and brown curves denote LSW models with $m=1$ and 0.2, respectively. Solid, short-dashed, and long-dashed curves indicate results for Models 1, 2, and 3, respectively, and filled squares denote $10^{12}$ and $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ for each model. Lines end at $10^{14}\,{\ensuremath{{\rm M_{\odot}}}}$, the largest halo mass for which values were calculated. As in Fig. \[fig:disk\], circles represent the observed quasar hosts of @Wang13, the pentagon is J1148$+$5251, the hexagon is J0210$-$0456, the triangle is the SMG HFLS3, and the limits correspond to J2329$-$0301 and the LAEs. Also as before, filled points indicate observations for which ${\rm sin} i$ is available, though no inclination correction is relevant for this figure. ](lbh_cii.eps){width="\columnwidth"} As in §\[sec:comp:lbh\], we again find that the quasar J2329$-$0301 (SFR$<40\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr}$; ${L_{\rm BH}}=10^{13}\,{\ensuremath{{\rm L_{\odot}}}}$; ${L_{\rm [CII]}}<7.9\times10^{8}\,{\rm Jy km/s Mpc^2}$) is consistent with our models with no hint of the uniqueness ascribed to it by @Willott13, though Figure \[fig:lbh-cii\] suggests it may have a value of $\beta$ slightly above 0.1 or non-negligible mass drain from either a quasar wind or a growing bulge. While this consistency may hardly be surprising given that only upper limits exist on both the star formation rate and \[CII\] luminosity for this object, it is worth noting that a star formation rate above the $40\,{\rm {\ensuremath{{\rm M_{\odot}}}}/yr}$ upper limit would be inconsistent with the upper limit on its \[CII\] flux in Models 2 and 3 (central panels in Fig. \[fig:cii1\]), particularly for $\beta>0.1$. CO Line Ratios {#sec:c:co} -------------- ![\[fig:co\] The CO(6–5) velocity-integrated flux versus CO(1–0) luminosity. The gray, shaded region shows the range predicted by our model when both the angular momentum transport mechanism and the choice of Models 1, 2, or 3 is varied. As in Fig. \[fig:disk\], circles represent the observed quasar hosts of @Wang13 with CO(1–0) values estimated from the J1148$+$5251 excitation ladder of @Riechers09—the solid, black line indicates the enforced proportional relationship between the emission from the two emission lines. The pentagon is J1148$+$5251 itself and the triangle is the SMG HFLS3. Also as before, filled points indicate observations for which ${\rm sin}\,i$ is available, though no inclination correction is relevant for this figure. ](co_z6.eps){width="\columnwidth"} In this section, we turn our attention from the \[CII\] line to those of CO. While we considered the line luminosities of CO in typical LBGs during reionization in previous work [@MF13a], we revisit it here for more extreme systems and focus particularly on the ratio of the $J=6\rightarrow5$ and $J=1\rightarrow0$ lines. Figure \[fig:co\] shows the CO(1–0) luminosity as a function of CO(6–5) velocity-integrated flux for our models compared to observations. Circles mark the @Wang13 sample of $z\sim6$ quasars. For these objects, CO(1–0) luminosity values are not directly observed but, rather, extrapolated from the excitation ladder of J1148$+$5251 [@Riechers09 pentagon] by assuming a proportional relationship with CO(6–5) as indicated by the solid line. The detected CO lines from HFLS3 (triangle) are consistent with this extrapolation. For convenience, we denote our models by a shaded region that encompasses the range for a variety of angular momentum transport parameters and for Models 1, 2, and 3. All of our models seem roughly to agree with the expected proportional relationship between emission from the two CO lines, though the shaded region smears out some scatter in the slope of our model relation among different parameter sets. This suggests that the excitation states of the lines do not vary strongly with halo mass but, rather, that larger halo masses produce brighter lines simply because their hosted galaxies contain more gas. However, we fail to reproduce the ratio of these two lines, predicting much more CO(6–5) flux for a given CO(1–0) luminosity. If we assume thermalized and optically thin lines at an excitation temperature of, $T$, the ratio between the luminosities of any two CO lines, represented by upper-state rotational quantum numbers $J$ and $K$, is $$\begin{aligned} \label{eq:Lrat} \frac{L_{J \rightarrow J-1}}{L_{K \rightarrow K-1}}&=&\frac{J}{K}\,\frac{A_{J}}{A_{K}}\,\frac{n_{J}}{n_{K}}\nonumber\\ &=&\frac{J\,(2\,J+1)}{K\,(2\,K+1)}\,\frac{A_{J}}{A_{K}}\,{\rm e}^{-\frac{(J-K)\,(J+K+1)\,{\ensuremath{{h_{\rm p}}}}\,\nu_{\rm CO}}{2\,k_{b}\,T}}.\end{aligned}$$ For CO, $A_{J=1}=7.203\times10^{-8}\,{\rm s^{-1}}$ and $A_{J=6}=2.137\times10^{-5}\,{\rm s^{-1}}$, while $\nu_{\rm CO}=115.3\,{\rm GHz}$. For $J=6$, $K=1$, and $T=50\,{\rm K}$, equation \[eq:Lrat\] gives about 800, roughly a factor of four higher than the @Wang13 value of ${L_{\rm CO(6-5)}}/{L_{\rm CO(1-0)}}\approx 170$. Assuming a temperature of $40\,{\rm K}$ results in a ratio of roughly 500. Of course, the lines are likely not both thermally populated and perhaps not optically thin. Of the two, the $J=6$ state is likely to be the most subthermal since its critical density is on the order of $10^5\,{\rm cm^{-3}}$. Subthermal suppression should be a bigger effect than any optical thickness of the CO(1–0) line and will likely decrease the expected ${L_{\rm CO(6-5)}}/{L_{\rm CO(1-0)}}$ from the value approximated in equation \[eq:Lrat\]. Moreover, subtraction of the CMB background will not affect both lines equally; the CMB is brighter in the CO(6–5) line than in the CO(1–0) line, yet subtraction of the same background will affect optically thick lines more than optically thin ones. Despite achieving roughly correct gas temperatures, our model over-produces the CO(6–5) flux for a given CO(1–0) luminosity, predicting ${L_{\rm CO(6-5)}}/{L_{\rm CO(1-0)}}\sim 10^3$, close to the thermal, optically thin value of equation \[eq:Lrat\]. Additionally, our models produce about the same ratio independent of angular momentum transport mechanism or Model (i.e., any model variation summarized in Tables \[tab:amt\] or \[tab:mod\]).[^11] The discrepancy implies that, inside real galaxies, the CO(6–5) line is more sub-thermally populated than predicted by our models. The likely culprit is the high level of turbulent support (and associated density fluctuations) within our model’s molecular clouds. To see how turbulently-generated density fluctuations affect the thermalization of the CO(6–5) line in our model, consider the simplified treatment of a $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ halo from §\[sec:intro:cii\] hosting a galaxy with a surface gas density of $4200\,{\rm {\ensuremath{{\rm M_{\odot}}}}/pc^2}$. The Jeans mass at the disk radius for this surface density is $3\times10^9\,{\ensuremath{{\rm M_{\odot}}}}$. Clouds of this mass require a turbulent velocity dispersion of  100 km/s to achieve a virial ratio of $\alpha=2$, and the resulting thermal sound speed at 50 K is  0.05 km/s. Thus, the 1D thermal Mach number is  2000, about a factor of 20 larger than the Mach numbers in strong, local starbursts. Since these objects have the same temperatures, the high Mach number results directly from the large cloud masses and high surface densities. As a result, the mass-weighted median density is a factor of 1700 larger than the mean density. Therefore, even though the critical density for thermalization of the CO(6–5) line is nearly a factor of 100 larger than that of CO(1–0), the high median density of the gas ensures that both lines are roughly thermal throughout our model disks. In @MF13a, we showed how including the effect of such turbulent clumps may have important effects for the observability of high excitation CO lines in more typical galaxies at high redshift. In contrast to these theoretical expectations, observers have found subthermally excited CO(6–5) lines in two systems, as indicated by the mean molecular densities found for J1148$+$5251 [@Riechers09] and HFLS3 [@Riechers13]: $n_{\rm H_2}=10^{4.2}\,{\rm cm^{-3}}$ and $n_{\rm H_2}=10^{3.8^{+0.28}_{-0.17}}\,{\rm cm^{-3}}$, respectively. These authors fit the CO excitation using temperature and a single “effective" density as free parameters. However, following the elementary assumptions of our model, the median density of our clouds is much higher than these inferred values, owing to the aforementioned turbulent fragmentation in the disk. Resolving the disagreement with observations will require changing the model of ISM physics to produce lower turbulent Mach numbers and hence lower median densities. Discussion & Conclusions {#sec:discussion} ======================== One of the main goals of this work is to understand whether our galaxy formation framework is an adequate description of observed LAEs, SMGs, and quasar hosts during the epoch of reionization. In §\[sec:comp:struc\], §\[sec:comp:lbh\], §\[sec:comp:mgas\], and §\[sec:comp:temp\], we showed that the relationships among radii, circular velocities, star formation rates, black hole luminosities, gas masses, and temperatures are all reasonably well-captured by our model. That we can come anywhere close to reproducing these properties with our idealized disk models is impressive, particularly with regard to quasar hosts, which are often thought to be built up in a succession of very rapid mergers [e.g., @Li07] since such mergers may provide the gravitational torques necessary to channel gas into the center of the disks. We find that most of the quasar hosts at $z=6$ are well-described by $\beta$-disks with $\beta=0.1$ living inside $\sim10^{12}$–$10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ halos with no direct feedback from the quasar onto star formation in the galactic disk. We point out that the relatively low star formation rates of the CFHQS quasars J0210$-$0456 and J2329$-$0301 are completely consistent with this lack of feedback. Our analysis ignores the effect of quasar outflows, which have been observed in some of these systems [@Maiolino12; @Valiante12]. Such outflows may imply that a higher value of $\beta$ is appropriate for the gas flows in these objects. In this case, the extra gas that would have accreted onto the central black hole is instead expelled. However, since our model reproduces the dependencies of ${L_{\rm BH}}$ on ${R_{\rm d}}$, ${V_{\rm c}}$, and star formation rate while ignoring quasar winds, the inclusion of such outflows and the dependence of the outflow rate on quasar properties is necessarily highly constrained. These constraints may prove an interesting avenue toward theoretically modeling quasar outflows in future work. Further, we use fits to our model results (Eq. \[eq:fits\]) to make predictions for the star formation rates of the @Wang13 quasar hosts in advance of observational determinations from rest-frame UV (see Table \[tab:disk\]). The SMG HFLS3 is an interesting case for our models since it has a large radius, line width, and star formation rate but no apparent AGN activity. If the data constrain the central bolometric black hole luminosity in HFLS3 to be less than $3\times10^{12}\,{\ensuremath{{\rm L_{\odot}}}}$—somewhat lower than could be observed in the deepest SDSS stripe based on its UV continuum component—the object would be only marginally consistent with a $\beta$-disk for $\beta=0.01$. However, within the framework of our model, the large radius, broad line width, high star formation rate, and low AGN activity of HFLS3 could easily be explained by any of our LSW disks inside a halo just under $10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$. However, our fiducial treatment of molecular gas is not completely consistent with observations. For example, we have difficulty reproducing the molecular gas fraction of HFLS3. Our method of calculating ${f_{\rm H_{2}}}$ from equation \[eq:fmol\] in @MF13a predicts an average molecular fraction of more than 99% for this object using parameters consistent with its other observables. Yet, given its inferred atomic gas content, this system has an average molecular fraction of only about 80%. The measurement of atomic gas via the observed \[CII\] emission could be an over-estimate if the signal were dominated by dissociated carbon in otherwise molecular gas, but this creates a new problem since our model also predicts ${f_{\rm CO}}/{f_{\rm H_{2}}}>0.97$, implying that nearly all of the molecular gas contains carbon in the form of CO. We seem to under-estimate the amount of CII in this system, but whether this is because we over-estimate the molecular fraction, ${f_{\rm H_{2}}}$, or the fraction of un-dissociated molecular gas, ${f_{\rm CO}}/{f_{\rm H_{2}}}$, is unclear. Regardless of the reason, this deficiency in our base model (Model 1) is related to our under-prediction of the \[CII\] luminosity compared to all of the detections we considered. While our model does ignore the contribution from ionized gas, we argued that PDRs of molecular clouds are the dominant source of \[CII\] emission at these high redshifts. Figures \[fig:cii1\] and \[fig:cii2\] also showed that our base model over-predicts the CO(6–5) fluxes of these systems if they are indeed associated with halos in the $10^{12}$–$10^{13}\,{\ensuremath{{\rm M_{\odot}}}}$ range. The combination of under-predicted \[CII\] and over-predicted CO again suggests that our models do not properly calculate ${f_{\rm CO}}$ in these systems. Models 2 and 3 are two [*[ad hoc]{}*]{} variations to our base model (Model 1) that try to improve this deficiency. While there are no \[CII\] or CO detections of $z\sim6$ LBGs, if our model similarly over-predicts ${f_{\rm CO}}$ in these systems, the prospects for observing them in CO would be even worse than suggested by @MF13a. In Model 2, we increase the star formation rate per dynamical time, ${{\rm SFR_{\rm ff}}}$, by a factor of ten (see summary in Table \[tab:mod\]). This efficiency is used to calculate the molecular fraction via equation \[eq:fmol\]. As discussed in @MF13a, we also set ${f_{\rm H_{2}}}=1$ when equation \[eq:fmol\] would otherwise predict a value greater than unity. This threshold, therefore, already effectively represents an increase in ${{\rm SFR_{\rm ff}}}$ above the value predicted by @Krumholz09b. The further increase to ${{\rm SFR_{\rm ff}}}$ in Model 2 reduces the molecular fraction required to produce a given star formation rate and reduces the frequency with which the ${f_{\rm H_{2}}}=1$ cutoff comes into effect. Physically, this change in the star formation efficiency may arise from the somewhat different physical properties expected in high-redshift molecular clouds, such as their very high surface densities or short free-fall times. Alternatively, star formation in atomic gas [@Krumholz12] would provide the same ultimate result. A contribution from HII regions may also work in a similar way by reducing the amount of molecular gas, albeit at higher temperatures. Regardless of the physical mechanism, we assume that changing ${{\rm SFR_{\rm ff}}}$ in this alternate model does not affect ${f_{\rm CO}}/{f_{\rm H_{2}}}$—calculated via equation \[eq:fco\]—but ultimately affects ${f_{\rm CO}}$. The magnitude of the increase in ${{\rm SFR_{\rm ff}}}$ that we selected for Model 2 is somewhat arbitrary and was not chosen to produce agreement with the observations but rather to demonstrate the effect of the modification. Nevertheless, we can ask whether this amount of increase in the star formation efficiency is sufficient to reproduce the observations. We find that Model 2 changes the results in low-mass halos more than in high-mass ones. This is expected since high-mass halos are more effected by the ${f_{\rm H_{2}}}=1$ cutoff which essentially nullifies the effect of increasing ${{\rm SFR_{\rm ff}}}$ on ${f_{\rm H_{2}}}$. The specific increase in Model 2 is only sufficient to explain the \[CII\] observations if $\beta=1$. To further match the relationships among ${L_{\rm BH}}$, ${R_{\rm d}}$, ${V_{\rm c}}$, and star formation rate in the quasar hosts, one must additionally invoke quasar outflows to reduce the accretion rate of the central black hole and the resulting luminosity. For HFLS3, however, this model appears to fail—despite predicting a total molecular fraction of 80%—unless ${{\rm SFR_{\rm ff}}}$ is still higher than we set here, since the disk parameters that produce the correct \[CII\] luminosity over-predict that of the central black hole even with winds. In Model 3, on the other hand, we create more CII by dissociating more of the carbon residing in otherwise molecular gas. Here, we decrease ${f_{\rm CO}}/{f_{\rm H_{2}}}$ by a factor of ten while leaving the molecular fraction, ${f_{\rm H_{2}}}$, unchanged. Physically, this scenario may result from a lower dust-to-metals ratio at high redshift so that the same metallicity produces less dust-shielding against CO dissociation. Again, the magnitude of this change is somewhat arbitrary and set primarily for demonstration purposes. Nevertheless, we find that Model 3 produces \[CII\] luminosity as a function of disk radius, star formation rate, or CO(6–5) flux consistent with observations independent of angular momentum transport mechanism. Thus, unlike in Model 2, we can simultaneously reproduce both the \[CII\] luminosity and low central black hole accretion rate of HFLS3. A second major issue revealed by our work is that, while Models 2 and 3 seem to do a better job of producing the observed \[CII\] emission than does Model 1, none of these scenarios adequately describes the CO excitation ladder of $z\sim6$ systems. Our model does yield a roughly constant value of ${L_{\rm CO(6-5)}}/{L_{\rm CO(1-0)}}$ with luminosity suggesting that the fluxes scale with gas mass, but our low value of the ratio compared to observations likely results from an over-estimate of the Mach number in our molecular clouds and a correspondingly more thermalized CO(6–5) population than observed. We can reproduce the correct ratio of CO(6–5)/CO(1–0) in our models if the Mach numbers are about a factor of 12 lower than we estimate. Since our temperatures are consistent with observations, lower Mach numbers require lower turbulent velocity dispersions. While this could be effected by a lower virial ratio, it is unclear how to physically interpret the necessary value of $\alpha\sim0.01\ll1$. Further, this scenario conflicts with the results of @Mashian13. These authors fit the velocity gradient and gas density for the $z\sim5$ SMG HDF 850.1 under different virialization assumptions and concluded that high-redshift molecular clouds are likely un-virialized, i.e., $\alpha>1$, though they did not take into account the effect of the velocity gradient on the clumpiness of the gas in the case that the support is provided by turbulence. Alternatively, a lower velocity dispersion could result from smaller clouds. While we assume $M_{\rm cl}=M_{\rm Jeans}$ as a function of radius, where our estimate of the Jeans mass already takes into account the self-gravity of the gas disk, a large contribution to the self-gravity from stars could lower the Jeans mass further. Clouds might also have smaller masses if they form as a result of turbulent rather than gravitational fragmentation and can achieve a dissociative equilibrium before they can be smoothed out by the Jeans stability. Finally, we note that, while the magnitude of the effect is unclear, fewer turbulent clumps could also result in more dissociation of CO into CII and boost the \[CII\] luminosity. This study reveals two puzzles in relating recent observations to existing ISM theory—namely, reproducing the \[CII\] emission and the ratio CO(6–5)/CO(1–0)—and suggests some schematic ways to improve agreement. Clearly, as samples of $z\gtrsim6$ systems with well-determined atomic and molecular emission continue grow, more theoretical work will be needed to understand the small scale physics of high-redshift molecular clouds and appreciate the differences between those at low-redshift. Acknowledgements ================ We thank Jean Turner, Mark Krumholz, Chris Carilli, Dominik Riechers, Matt Malkan, and Desika Narayanan for helpful discussions and suggestions. This research was partially supported by the David and Lucile Packard Foundation. [^1]: E-mail: [email protected] [^2]: If we assume that another 10% of halo baryons are in stars, this results in a gas fraction—${\rm gas}/({\rm gas}+{\rm stars})$—of 50%. [^3]: As in @Thompson05, we use opacities as functions of temperature and density derived from @BL94 but have checked that the specific fits do not significantly affect our results. [^4]: While @Munoz12 determined this mass-loading factor for LBGs at $z=6$–$8$, we will also apply it here to quasar hosts whose metallicities, stellar populations, and environments may be somewhat different. However, given the large host halo masses of these systems and their correspondingly high velocity dispersions, these stellar winds contribute only minimally to gas depletion so that our assumption does not substantially affect the calculation. [^5]: Imposing the threshold of ${f_{\rm H_{2}}}=1$ effectively increases the star formation efficiency of the gas beyond that prescribed by @KM05. Physically, this may result from gas out of chemical equilibrium in which star formation is produced in atomic gas [e.g., @Krumholz12]. [^6]: @Wang13, for example, uses ${L_{\rm BH}}=4.2\,\nu\,L_{\rm 1450}$, the simplest fit for the isotropic correction from @Runnoe12. [^7]: We define “fits" loosely here and without any statistical rigor. Our goal is merely to describe our results for interpolation purposes rather than to provide physical insight. [^8]: We note, however, that @Narayanan12 ignored the CMB as an observing background, which, while insignificant at $z=2$, is important at $z=6$ [@Obreschkow09b; @MF13a; @daCunha13b]. As a result, the true value of $\alpha_{\rm CO}$ and the molecular mass associated with a given luminosity may be somewhat higher. [^9]: This is not quite equivalent to the procedure in @Obreschkow09b, where the radius used to compute $\dot{\Sigma}_{\rm SF}$ is the half-mass radius of the molecular gas rather than the disk radius. However, the final temperature is not very sensitive to this difference. The star formation rate surface density may rise higher than this average value in the inner regions of galaxy (both real and in our model), but this resolution is not achievable in the @Obreschkow09b semi-analytic prescription. [^10]: A radiative efficiency of 8% is appropriate for Schwarzchild black holes ignoring a general relativistic correction of order unity, but the efficiency may be as high as tens of percent for rotating black holes. However, the final temperature is insensitive to these subtleties. [^11]: Large deviations among model results occur only where our CMB subtraction implies that a line is seen in absorption rather than emission. However, as in @MF13a, we caution against taking these values too seriously since our model more accurately predicts the intrinsic emission and the amount of absorption separately than it does the difference between the two when similarly valued.
{ "pile_set_name": "ArXiv" }
--- abstract: 'In order to analyze the fidelity susceptibility of non-relativistic field theories, which are important in condensed matter systems, we generalize the proposal to obtain the fidelity susceptibility holographically to Lifshitz geometries. It will be argued that this proposal can be used to study the fidelity susceptibility for various condensed matter systems. To demonstrated this, we will explicitly use this proposal to analyze the fidelity susceptibility for a non-relativistic many-body system, and argue that the fidelity susceptibility of this theory can be holographically obtained from a bulk Lifshitz geometry. In fact, using a Einstein-Dilaton-Maxwell-AdS-Lifshitz theory, we explicitly demonstrated that the fidelity susceptibility obtained from this bulk geometry is equal to the fidelity susceptibility of a bosonic many-body system.' author: - | Davood Momeni\ Department of Physics, College of Science, Sultan Qaboos University,\ P.O. Box 36, P.C. 123, Al-Khowd, Muscat, Sultanate of Oman\ \ Mir Faizal\ Irving K. Barber School of Arts and Sciences,\ University of British Columbia - Okanagan, 3333 University Way,\ Kelowna, British Columbia V1V 1V7, Canada\ Department of Physics and Astronomy, University of Lethbridge,\ Lethbridge, Alberta, T1K 3M4, Canada\ \ Aizhan Myrzakul, Ratbay Myrzakulov\ Eurasian International Center for Theoretical Physics\ and Department of General Theoretical Physics,\ Eurasian National University, Astana 010008, Kazakhstan title: Fidelity susceptibility for Lifshitz geometries via Lifshitz Holography --- Introduction ============ It is known that the entropy of a black hole scales with its area. As black holes are maximum entropy objects, this implies that the maximum entropy of that certain region of space scales with the area of its boundary. This observation has led to the development of the holographic principle, which equates the number of degrees of freedom in a region of space to the number of degrees of freedom on the boundary surrounding that region of space [@1; @2]. The AdS/CFT correspondence is a realization of the holographic principle as it is a duality between the string theory/supergravity in AdS spacetime and the field theory on its boundary [@M:1997]. As AdS/CFT correspondence is a duality between two very different theories, it seems from the AdS/CFT correspondence and the holographic principle that laws of physics are fundamentally just information theoretical processes. In fact, various studies done in different fields of science seem to indicate that the laws of physics are informational theoretical processes [@info; @info2]. So, the AdS/CFT can be used to obtain information theoretical information relating to a conformal field theory from the bulk geometry. The entanglement entropy of a field theory is a most important informational theoretical quantity relating to a conformal field theory. It has been demonstrated that the entanglement entropy of a conformal field theory can be holographically obtained from the bulk AdS spacetime, as it is dual a minimal surface in asymptotically AdS spacetime [@6; @6a]. It is also important to know how much information is retained in a system, and holographic entanglement entropy can be used to quantify this as it measures the loss of information in a system. However, it is also important to know the difficulty to obtain this information, and this can be quantified using complexity. As laws of physics can be understood in terms of information theoretical processes [@info; @info2], and complexity is an important informational theoretical quantity, complexity is expected to be an important physical quantity used in the laws of physics. In fact, complexity has been used to understand the behavior of condensed matter systems [@c1; @c2] and molecular physics [@comp1], quantum computing [@comp2]. In fact, it has been argued that the information might not be ideally lost in a black hole, but it would be effectively lost, as it would not be possible obtain this information from a black hole due to its chaotic nature [@hawk]. The complexity of a conformal field theory can also be obtained holographically, as the holographic complexity is dual to a volume in AdS spacetime [@Susskind:2014rva1; @Susskind:2014rva2; @Stanford:2014jda; @Momeni:2016ekm; @Alishahiha:2015rta; @Alishahiha:2017cuk]. It has been demonstrated that the holographic complexity of a field theory can be related to the fidelity susceptibility of the boundary field theory. So, the fidelity susceptibility of a field theory can be holographically calculated using a maximal volume $ V(\gamma_{max})$ in the AdS which ends on the time slice at the AdS boundary [@r5; @r7]. As fidelity susceptibility is important to understand the behavior of condensed matter systems [@f1; @f2; @f4; @f5; @f7], it is important to generalize this proposal to non-relativistic field theories describing condensed matter systems. It may be noted that such non-relativistic condensed matter systems can be holographically analyzed using Lifshitz holography [@cond1; @cond2; @cond4; @cond5; @cond6]. In the Lifshitz holography, the Lifshitz deformed of AdS can be related to the Lifshitz field theories, in which spacetime have different scaling behavior [@Griffin:2012qx]. As Schrödinger invariant quantum system, which describe condensed matter system, the space and time scale differently, we will use Lifshitz holography to analyze such a system. Lifshitz Fidelity Susceptibility ================================ In this section, motivating by [@Griffin:2012qx], in this work we will now generalize the relativistic proposal to obtain the fidelity susceptibility from holographic complexity [@r5; @r7] to Lifshitz geometries. So, if we consider one parameter a Lifshitz field theory, with states denoted by $|\Psi(\lambda)>$, then the inner product of two such states separated by an infinitesimally perturbation $\delta \lambda$, can be expressed as $$<\Psi(\lambda|\Psi(\lambda +\delta\lambda)> = 1-\Xi_F (\delta \lambda)^2.$$ Now using the argument used in [@r5; @r7], it can be argued that $\Xi_F$ can be written as $$\Xi_F = \int dm <O(x), O(x')>,$$ where $dm$ is a suitable integral measure for this system, and $O(x), O(x')$ are suitable local operator in the Lifshitz theory. Now it can again be argued that this quantity will be described holographically. However, as these operators are defined in a Lifshitz theory, and the holographic dual of such a field theory is described by Lifshitz gravity [@Griffin:2012qx], we can use the argument of [@r5; @r7], to argue that $\Xi_F$ can also be obtained holographically from a Lifshitz-AdS geometry. In fact, as this quantity has to reduce to holographic complexity for $z=1$ for usual theory, we can also calculate this quantity in Lifshitz geometries using the volume of maximum surfaces. This can be done by defining $\mathcal{V}(\gamma_{max})$ as a maximal volume in the Lifshitz deformation of the AdS spacetime, which ends on the time slice at the Lifshitz-AdS boundary. We can use this maximal volume in the Lifshitz geometry to define the holographic complexity in such a geometry as $$\begin{aligned} &&F=\frac{V(\gamma_{max})}{8\pi R G}, \end{aligned}$$ It may be noted that for $z=1$ this maximal volume reduces to the usual maximal volume in AdS spacetime, and so this holographic complexity reduces to the usual holographic complexity for $z=1$ [@r5; @r7]. Now the Lifshitz holography reduces to the usual holography for $z =1$, and it is known that there are divergences associated with such volumes for $z=1$ [@Carmi:2016wjl]. So, we need to regularize this volume, before we can define the fidelity susceptibility for Lifshitz geometries. This will be done by subtracting the background Lifshitz-AdS geometry ${V}(\gamma_{max})_{LAdS}$ from the deformed Lifshitz-AdS geometry ${V}(\gamma_{max})_{DLAdS}$. So, we can define a regularized Lifshitz maximal volume $$\begin{aligned} \mathcal{V}(\gamma_{max}) ={V}(\gamma_{max})_{DLAdS}- {V}(\gamma_{max})_{LAdS} \end{aligned}$$ Now using this regularized maximal volume in the Lifshitz geometry, we can define the regularized holographic complexity of a non-relativistic boundary theory as $$\begin{aligned} \Xi_F &=& F_{DLADS}- F_{LAdS}\nonumber \\ &= & \frac{\mathcal{V}(\gamma_{max})}{8\pi R G}, \label{b} \end{aligned}$$ where $R$ is the radius of the curvature of this Lifshitz-AdS geometry. This regularized holographic complexity is equal to fidelity susceptibility of the boundary field theory, and so it fidelity susceptibility of the boundary field theory can be holographically calculated from holographic complexity. It may be noted that for $z=1$ this expression reduces to the usual expression for the regularized fidelity susceptibility [@rf14; @rf17]. Boundary Bosonic System ======================= Now we will use this proposal for holographically analyze a simple system of system of $N$ bosons, without any self-interaction. These bosons will be placed in a background uniform magnetic field $\vec{H}$ in the $z$ direction, with $\vec{H}=H\hat{e}_z$. So, even though the bosons do not interact with a dynamical field, they do interact with a background field. It is possible to holographically analyze such systems in a background field [@b1a0; @b2a0; @b4a0; @b6a0]. We will analyze this proposal for such a simple system, to demonstrate how such a holographic correspondence can work, and so this proposal can be used for holographically analyzing more complicated systems. Now before we analyze the bulk Lifshitz geometry dual to such a system, we will calculate the fidelity susceptibility of this bosonic theory. The Hamiltonian for each of these bosonic particles is $ H_i= {(-i\vec{\nabla}_i-q\vec{A})^2}/{2m} , $ where $\vec{A}=\vec{\nabla}\times \vec{H}$, so the time-dependent Schrödinger equation for these bosonic particles can be written as $$\begin{aligned} \sum\frac{(-i\vec{\nabla}_i-q\vec{A})^2}{2m}\Psi_{tot}(\vec{x}_1,\vec{x}_2,...,\vec{x}_N;t) =i\frac{\partial}{\partial t}\Psi_{tot}(\vec{x}_1,\vec{x}_2,...,\vec{x}_N;t). &&\end{aligned}$$ Now we can write the total wave function as $$\Psi_{tot}(\vec{x}_1,\vec{x}_2,...,\vec{x}_N;t)=\Pi_{i=1}^N\psi_{i}(\vec{x}_i;t).$$ We need to find only ground state wave function $\Psi^{0}_{tot}(\vec{x}_1,\vec{x}_2,...,\vec{x}_N;t)$. Choosing the gauge, $\vec{A}=(0,Hx,0)$, we can write $\vec{A}=\vec{\nabla}\times \vec{H}$. Using this gauge for $\vec{A}$, Schrödinger equation can be expressed as $$\begin{aligned} &&-\nabla_i^2\psi_{i}(\vec{x}_i;t)+2iqHx\frac{\partial \psi_{i} (\vec{x}_i;t)}{\partial y} +q^2H^2x^2\psi_{i}(\vec{x}_i;t)=2mE_i\psi_{i}(\vec{x}_i;t)\label{H2}.\end{aligned}$$ We can express $\psi_i(\vec{x}_i;t)$ as $$\psi_{i}(\vec{x}_i;t)=e^{i(k_i z_i+\beta_i y_i-E_i t)}\phi_{i}(x_i).$$ Now we can write $$\begin{aligned} -\frac{d^2\phi_{i}(x_i)}{dx_i^2}+\Big(q^2H^2x_i^2-2q\beta_iH x_i\Big)\phi_{i}(x_i) =(2mE_i-k_i^2-\beta_i^2)\phi_{i}(x_i)\label{1D-phi}.&&\end{aligned}$$ It may be noted that this looks just like the Schrödinger equation for a simple harmonic oscillator, such that the coordinates have been shifted as $x_i\to \xi+\frac{\beta_i}{qH} $. So, we can express the Schrödinger equation for this system as $$\begin{aligned} -\frac{1}{2m}\frac{d^2\phi_{i}(\xi_i)}{d\xi_i^2}+\frac{1}{2}m\omega_i^2\xi_i^2\phi_{i}(\xi_i)=\Big(E_i-\frac{k_i^2}{2m}\Big)\phi_{i}(\xi_i) \label{1D-psi},\end{aligned}$$ where $\omega_i=\frac{qH}{m}$ denotes frequency (energy) of the system. Exact solution for this equation can be expressed in terms of Hermite functions. Now using the ground state wave function for a bosonic particle, $ \phi_{i,0}(\xi_i)=\sqrt[4]{\frac{qH}{\pi}}e^{-\frac{qH}{2}\xi_i^2},\ \ E_{i,0} =\frac{qH+k_i^2}{2m}. $, the ground state wave function for whole system can be written as $$\begin{aligned} \Psi_{tot,0}(\vec{x}_1,\vec{x}_2,...,\vec{x}_N;t)&=&(\frac{qH}{\pi})^{N/4}\Pi_{i=1}^N \nonumber \\ && \times e^{i\Big(k_i z_i+\beta_i y_i-(\frac{qH+k_i^2}{2m}) t\Big)-\frac{qH}{2}(x_i-\frac{\beta_i}{qH})^2}. \end{aligned}$$ The fidelity susceptibility can be obtained by varying $H\to H+\delta H$, and computing the following inner product, $$F= <{\Psi_{tot,0}(H)|\Psi_{tot,0}(H+\delta H)}>= 1-(\delta H)^2\Xi_F+...$$ So, we expand $F$ in series up to the second order [@b1], $$\begin{aligned} &&<{\Psi_{tot,0}(H)|\Psi_{tot,0}(H+\delta H)}> \nonumber \\ &=& \Pi_{i=1}^N\int d^3x_i \Psi^{\star}_{tot,0}(H+\delta H)\Psi_{tot,0}(H) \\\nonumber &=& \Big(\int d^3x\psi^{\star}(\vec{x};t,H+\delta H)\psi(\vec{x};t,H) \Big)^N. \end{aligned}$$ Thus, we can express the fidelity susceptibility of this system as $$\begin{aligned} &&\Xi_F=\frac{N}{8 q H^3}(qH+4\beta^2)\label{XIF}. \end{aligned}$$ Expression given in (\[XIF\]) is the exact fidelity susceptibility for a system of $N$ charged bosonic particles in a uniform magnetic field. Lifshitz Bulk Dual ================== This boundary theory studied in the previous section is defined by non-relativistic Hamiltonian, so it is expected to be dual to a non-relativistic Lifshitz bulk theory [@a1; @a2; @a4; @a5]. Now we will suppose that the Einstein-Dilaton-Maxwell-AdS-Lifshitz bulk action [@k1; @k2; @l1; @l2] is holographically dual to the quantum non-relativistic bosonic system, and the fidelity susceptibility from the bulk theory matches the fidelity susceptibility of the boundary theory. Thus, we propose the following action for the bulk theory [@l1; @l2], $$\begin{aligned} S_{\mbox{Bulk}}=\int_{\mathcal{M}} d^{4}x\sqrt{-g}\left[\frac{(R-2\Lambda)}{2\kappa^2}-\frac{1}{2} \partial^\mu\phi\partial_\mu\phi +V(\phi) -\xi\mathrm{e}^{\lambda\phi}(F^{\mu\nu}F_{\mu\nu})\right]\,.\label{action}\end{aligned}$$ where the potential is $V(\phi)=V_0\mathrm{e}^{\gamma\phi}$ with parameters $V_0$ and $\gamma$. Here $\phi$ is non-minimally coupled with electromagnetic potential, and the electromagnetic field strength coupled to scalar field as $\xi\mathrm{e}^{\lambda\phi}(F^{\mu\nu}F_{\mu\nu})$, such that $\xi, \lambda$ are suitable constants. The metric for a static, spherically symmetric solution in this Einstein-Dilaton-Maxwell-AdS-Lifshitz can be written as [@l1; @l2] $$ds^2=-e^{2\alpha(r)}B(r)dt^2+\frac{d r^2}{B(r)}+r^2d\sigma_{2,k}^2 \,,\label{metric}$$ where $\alpha(r), B(r)$ are function of $r$. Here $d \sigma_{2,k}^2$ is the metric for a topological two-dimensional surface parametrized by $k= 0, \pm 1$. This two-dimensional manifold is a sphere $S_{2}$ for $k=1$, a torus $T_{2}$ for $k=0$, and a compact hyperbolic manifold $Y_{2}$ for $k=-1$. Now we can choose $k=0$ and write the planar Euclidean coordinates as $d\sigma_{2,k}^2=dx^idx_i,\ \ i=1,2,x^i=\{x,y\}$. It may be noted that due to Lifshitz scaling, $\alpha(r)\propto\log r^{z/2}$, where $z$ is the Lifshitz parameter. So, the general form of the metric with Lifshitz symmetry can be written as $$ds^2=-\left(\frac{r}{r_0}\right)^{z}B(r)dt^2+\frac{d r^2}{B(r)}+r^2 dx_i dx^i\,.\label{exsolution}$$ Here the function $B(r)$ can be written as [@l1; @l2] $$\begin{aligned} B(r)&=&\frac{2}{(2+z)}\left[\frac{\tilde V_0}{2}\right]+ \Big(\frac{r_{+}}{r}\Big)^{1+z/2}\Big(\Big(\frac{2(\Lambda+\tilde Q^2\xi)}{(6+z)}\Big)r_{+}^2 \nonumber \\ && -\frac{2}{(2+z)}\left[\frac{\tilde V_0}{2}\right] \Big) -\Big(\frac{2(\Lambda+\tilde Q^2\xi)}{(6+z)}\Big)r^2. \end{aligned}$$ This bulk theory is dual to the bosonic system, we have analyzed in the previous section. Holographic Complexity ====================== As we have obtained the fidelity susceptibility of the bosonic system in the previous section, we will holographically analyze it in this section. So, we will use generalization the fidelity susceptibility [@r5; @r7] to a Lifshitz geometry given by Eq. (\[b\]). The fidelity susceptibility in such geometries depends on $V(\gamma_{max})$, and we can obtain $V(\gamma_{max})$ using $$\begin{aligned} &&V(\gamma_{max})=\int_{r_{+}}^{r_{\infty}}\frac{r^2dr}{\sqrt{B(r)}}\label{Vmax} \end{aligned}$$ where $r_{+}$ is horizon, and $r_{\infty}$ is an IR cutoff. Now we can use the Poincare coordinate $w=\frac{r}{r_{+}}$ to evaluate integral (\[Vmax\]) as $$\begin{aligned} &&V(\gamma_{max})=r_{+}^3\int_{\epsilon}^{1} \frac{dw}{w^4\sqrt{b(w)}} \label{Vmax-z} \end{aligned}$$ where $\epsilon\to 0$ is an $UV$ cutoff. We also have $$\begin{aligned} && b(w)= b_1w^{1+z/2} -\frac{b_{-2}}{w^2}. \end{aligned}$$ It may be noted as $z=-\frac{4\tilde Q^2\xi}{\Lambda+\tilde Q^2\xi}\geq 3, \Lambda=-\frac{3}{L^2}$, so $z=4$ is an interesting solution. In this case, the coefficients $b_n$ are given as following $$\begin{aligned} && b_1= \Big(\frac{\Lambda+\tilde Q^2\xi}{5}\Big)r_{+}^2-\frac{1}{3}\left[\frac{\tilde V_0}{2} \right],\\&& b_{-2}=\frac{r_{+}^2(\Lambda+\tilde Q^2\xi)}{5}.\end{aligned}$$ Now we obtain $$\begin{aligned} \label{volume22} &&V(\gamma_{max})=r_{+}^3\Big(\frac{A}{3840 (-b_{-2})^{9/2}} \Big)+\frac{r_{+}^3}{2 \epsilon ^2 \sqrt{-b_{-2}}}\end{aligned}$$ here $A = 640 b_{-2}^3 (b_1-3 b_{-2})$. Here the bulk charge $\tilde{Q}$ is dual to the magnetic charge (strength) $ H$ of the boundary theory, and $H$ varying smoothly. So, we can the volume (\[volume22\]) for $\mathcal{O}(\frac{1}{H^3})$, and obtain, $$\begin{aligned} && F_{DAdS}=\frac{\sqrt{5} r_{+}^2\sqrt{-\xi } \left(2 L^2 \xi \tilde Q+3\right)}{48 \pi G L^3 \xi ^2 \tilde Q^3}-\frac{\sqrt{5} r_{+}^2\sqrt{-\xi } \left(2 L^2 \xi \tilde Q^2+3\right)}{32 \pi G L^3 \xi ^2 \tilde Q^3 r^2 \epsilon ^2}. \end{aligned}$$ Now this equation contains both the finite and divergent parts of the holographic fidelity susceptibility. However, we can regularize it by subtracting it from the background AdS geometry, and obtain the regularized finite fidelity susceptibility. $$\begin{aligned} \Xi_F &=& F_{DLADS}- F_{LAdS}\nonumber \\ &=& \frac{\sqrt{5} r_{+}^2\sqrt{-\xi } \left(2 L^2 \xi \tilde Q+3\right)}{48 \pi G L^3 \xi ^2 \tilde Q^3}. \label{fid-holo} \end{aligned}$$ It may be noted that that the regularized fidelity susceptibility calculated holographically in (\[fid-holo\]) is same as the fidelity susceptibility of the boundary theory obtained in (\[XIF\]). So, we can write $$\begin{aligned} &&\frac{\sqrt{5} r_{+}^2\sqrt{-\xi } \left(2 L^2 \xi \tilde Q+3\right)}{48 \pi G L^3 \xi ^2 \tilde Q^3}=\frac{N}{8 q H^3}(qH+4\beta^2) \end{aligned}$$ Now if we holographically identify $\tilde Q=H$, we can also identify many parameters in the bulk to boundary theories. In fact, from this identification, we obtain $$\begin{aligned} &&\frac{2 \sqrt{5}L^2 \xi r_{+}^2 \sqrt{-\xi} }{48 \pi G L^3 \xi ^2 }\equiv\frac{ N}{8 },\\&& \frac{3\sqrt{5} r_{+}^2 \sqrt{-\xi}}{48 \pi G L^3 \xi ^2 }=\frac{N\beta^2}{2 q }.\end{aligned}$$ So, the number of boundary quantum systems $N$ and $\beta^2 q^{-1}$ can be expressed as $$\begin{aligned} &&N^{\mbox{boundary}}=\Big(-\frac{16 \sqrt{5}L^2 r^2_{+} }{48 \pi G L^3 \sqrt{-\xi} }\Big)^{\mbox{Bulk}},\\&& \Big(\frac{\beta^2}{ q }\Big)^{\mbox{boundary}}=\Big(\frac{3}{8L^2 \xi}\Big)^{\mbox{Bulk}}.\end{aligned}$$ It may be noted that $\xi<0$, for this system. So, we have analyzed a system of bosonic particles in a magnetic field, and we obtained the fidelity susceptibility for this system. As this system was a non-relativistic system, it was expected to be dual to a AdS-Lifshitz spacetime. We have demonstrated that this theory is dual to a Einstein-Dilaton-Maxwell-AdS-Lifshitz, and the fidelity susceptibility calculated from the bulk using this theory matches with the fidelity susceptibility of the boundary theory. Conclusion ========== In this paper, we propose that the fidelity susceptibility of a non-relativistic system can be obtained holographically from the holographic complexity of a Lifshitz-AdS theory. We use this proposal to holographically analyze the fidelity susceptibility of non-relativistic field theories, and demonstrated that fidelity susceptibility of the bulk theory is the same as the boundary theory. So, using a Einstein-Dilaton-Maxwell-AdS-Lifshitz theory, we explicitly demonstrated that the fidelity susceptibility obtained from this bulk geometry is equal to the fidelity susceptibility of a bosonic many-body system. It may be noted that as the boundary system considered explicitly in this paper described a simple system, it was possible to calculate the fidelity susceptibility for this system, both in the boundary and in the bulk. However, it is not always possible to calculate the fidelity susceptibility for the boundary and the bulk system. It would be difficult to analyze the strongly coupled field theories, and perform such calculations in the field theory side of the duality. However, it is known that a strongly coupled limit of the field theory can be holographically analyzed using a weakly coupled limit on the gravity side of the duality. Thus, for such non-relativistic systems, where such calculations cannot be performed on the field theory side of the duality, this holographic calculations can be performed using the gravitational side of the duality. As fidelity susceptibility is an very important quantity in condensed matter systems, and many condensed matter systems can be modeled using conformal field theories, it would be possible to use the formalism developed in this paper for analyzing such condensed matter systems. It is possible to describe condensed matter systems like Weyl semi-metal as strongly coupled systems [@wely], it would be possible and interesting to use the results of this paper for holographically analyzing such fidelity susceptibility of such system. As we have generalized the fidelity susceptibility to Lifshitz geometries, and Lifshitz geometries can have important condensed matter applications [@cond1; @cond2; @cond4; @cond5; @cond6], the results of this paper can have interesting condensed matter applications. So, it would be interesting to analyze realistic condensed matter systems, and then use the Lifshitz holography to understand the behavior of fidelity susceptibility for such condensed matter systems. It is also possible to study various interesting time-dependent generalizations of this solution. It is expected that such a time-dependent system on the boundary will be dual to some time-dependent Lifshitz bulk solution. It may be noted that fidelity susceptibility for time-dependent geometries has been studied [@time], and it is expected that this formalism can be generalized to bulk geometries with Lifshitz symmetry. This would be interesting as such time-dependent systems are important in condensed matter physics [@time1]. It would be interesting to analyze such geometries, and use them to understand the boundary fidelity susceptibility. Thus, the proposal developed in this paper can be used to analyze interesting condensed matter systems, and calculate the fidelity susceptibility for such systems. [99]{} G. ’t Hooft, \[arXiv:gr-qc/9310026\] L. Susskind, J. Math. Phys. 36, 6377 (1995) J. M. Maldacena, Adv. Theor. Math. Phys. 2, 231 (1998) K. H. Knuth , AIP Conf. Proc. 1305, 3 (2011) P. Goyal, Information 3, 567 (2012) S. Ryu and T. Takayanagi, Phys. Rev. Lett. 96, 181602 (2006) V. E. Hubeny, M. Rangamani and T. Takayanagi, JHEP 0707, 062 (2007) F. Barahona, J. Phys. A 15, 3241 (1982) M. Troyer and U. J. Wiese. Phys. Rev. Lett 94, 170201 (2005) J. Grunenberg, Phys. Chem. Chem. Phys. 13, 10136 (2011) M. Stanowski, Complicity 2, 78 (2011) S. W. Hawking, M. J. Perry and A. Strominger, Phys. Rev. Lett. 116, 231301 (2016) L. Susskind, Fortsch. Phys.  [ 64]{}, 24 (2016) L. Susskind, Fortsch. Phys.  [ 64]{}, 24 (2016) D. Stanford and L. Susskind, Phys. Rev. D [ 90]{}, 12, 126007 (2014) D. Momeni, S. A. H. Mansoori and R. Myrzakulov, Phys. Lett. B 756, 354 (2016) M. Alishahiha, Phys. Rev. D [ 92]{}, 126009 (2015) M. Alishahiha and A. Faraji Astaneh, Phys. Rev. D [ 96]{}, no. 8, 086004 (2017) M. Miyaji, T. Numasawa, N. Shiba, T. Takayanagi, and K. Watanabe, Phy. Rev. Lett 115, 261602 (2015) N. S. Mazhari, D. Momeni, S. Bahamonde, M. Faizal and R. Myrzakulov, Phys. Lett. B 766, 94 (2017) M. Weber, F. F. Assaad and M. Hohenadler, Phys. Rev. B 94, 245138 (2016) V. Mukherjee, S. Montangero and R. Fazio, Phys. Rev. A 93, 062108 (2016) E. J. Konig, A. Levchenko and N. Sedlmayr, Phys. Rev. B 93, 235160 (2016) B. Damski, Sci. Rep. 5, 15779 (2015) R. Jafari, J. Phys. A: Math. Theor. 49, 185004 (2016) S. A. Hartnoll, Class. Quant. Grav. 26, 224002 (2009) C. J. McGreevy, Adv. High Energy Phys. 2010, 723105 (2010) P. Herzog, J. Phys. A 42, 343001 (2009) J. Gath, J. Hartong, R. Monteiro and N. A. Obers, JHEP 1304, 159 (2013) S. Sachdev, Ann. Rev. Cond. Matt. Phys. 3, 9 (2012) T. Griffin, P. Horava and C. M. Melby-Thompson, Phys. Rev. Lett. [ 110]{}, 081602 (2013) D. Carmi, R. C. Myers and P. Rath, JHEP [ 1703]{}, 118 (2017) D. Momeni, M. Faizal, K. Myrzakulov and R. Myrzakulov, Phys. Lett. B 765, 154 (2017) N. S. Mazhari, D. Momeni, S. Bahamonde, M. Faizal and R. Myrzakulov, Phys. Lett. B [ 766]{}, 94 (2017) D. T. Son, Phys. Rev. D 78, 046003 (2008) K. Balasubramanian and J. McGreevy, Phys. Rev. Lett. 101, 061601 (2008) Y. Nishida and D. T. Son, Phys. Rev. D 76, 086004 (2007) D. Martelli and Y. Tachikawa, JHEP 1005, 091 (2010) S. Sachdev, *Quantum Phase Transitions*, Cambridge University Press, Cambridge, UK (2000) X. Wang, J. Yang, M. Tian, A. Wang, Y. Deng and G. Cleaver, Phys. Rev. D91, 064018 (2015) G. Tallarita, Phys. Rev. D 89, 106005 (2014) T. Andrade, C. Keeler, A. Peach and S. F. Ross, Class. Quant. Grav. 32, 085006 (2015) P. Burda, R. Gregory and S. Ross, JHEP 1411, 073 (2014) M. K. Zangeneh, A. Dehyadegari, M. R. Mehdizadeh, B. Wang and A. Sheykhi, arXiv:1610.06352 \[hep-th\] J. Tarrio and S. Vandoren, JHEP 1109, 017 (2011) D. Momeni, R. Myrzakulov, L. Sebastiani and M. R. Setare, Int. J. Geom. Meth. Mod. Phys. 12, 1550015 (2015) M. H. Dehghani, R. Pourhasan and R. B. Mann, Phys. Rev. D 84, 046002 (2011) T. Bzdusek, A. Ruegg and M. Sigrist, Phys. Rev. B 91, 165105 (2015) D. Momeni, M. Faizal, S. Bahamonde and R. Myrzakulov, Phys. Lett. B 762, 276 (2016) F. Mahfouzi, B. K. Nikolic and N. Kioussis, Phys. Rev. B 93, 115419 (2016) D. T. Son, Phys. Rev. D 78, 046003 (2008)
{ "pile_set_name": "ArXiv" }
--- abstract: 'We use the unintegrated Parton Density Functions of the gluon obtained from a fit to measurements of the structure functions $F_2(x, Q^2)$ and $F_2^c(x,Q^2)$ at HERA to describe the experimental data for $F_2^b(x,Q^2), F_L(x,Q^2)$ and $F_L$ at fixed $W$.' author: - | H. Jung$^1$, A.V. Kotikov$^2$, A.V. Lipatov$^3$ and N.P. Zotov$^3$\ 1- DESY, Hamburg, Germany\ \ 2- BLTHPH - JINR, Dubna, Russia\ \ 3- SINP - MSU, Moscow, Russia\ title: Critical tests of unintegrated gluon distributions --- Introduction ============ The purpose of the present investigation is to study the longitudinal structure function (SF) $F_L (x, Q^2)$ as well as the charm and beauty contributions to the proton SF $F_2(x, Q^2)$ using the $k_T-$factorization approach of QCD [@CCH]. The SF $F_L (x, Q^2)$ is directly connected to the gluon density in the proton. Only in the naive quark-parton-model $F_L (x, Q^2)=0$, and becomes non-zero in pQCD. However the pQCD leads to controversal results still. It was shown recently [@Thorne], that the $F_L$ experimental data from HERA seem to be inconsistent with some of the NLO predictions (in particular the MRST one) at small $x$. BFKL effects significantly improve the description of the low $x$ data when compared to a standard NLO $\bar{MS}$-scheme global fit. The NNLO global fit becomes better when taking into account higher order terms involving powers of $\ln (1/x)$. It means, that we need a resummation procedure. On the other hand it is known, that the BFKL effects are taken into account from the very beginning in the $k_T-$factorization approach [@CCH], which is based on the BFKL [@BFKL] or CCFM [@CCFM] evolution equations summing up the large logarithmic terms proportional to $\ln (1/x)$ or $\ln (1/(1-x))$ in the LLA. Some applications of the $k_T-$factorization approach were shown in Refs. [@Smallx]. In the framework of $k_T$-factorization the study of the longitudinal SF $F_L$ began already ten years ago [@CH], where the small $x$ asymptotics of $F_L$ has been evaluated, using the BFKL results. Since we want to analyze the SF data in a broader range at small $x$ we use a more phenomenological approach in our analyses of $F_2$ and $F_L$ data [@BKS; @KLZ]. Using the $k_T$-factorization approach for the description of different SF at small $x$ we hope to obtain additional information (or restrictions), in particular, about one of the main ingradient of $k_T$-factorization approach - the unintegrated gluon distribution (UGD) In the $k_T$-factorization the SF $F_{2,L}(x,Q^2)$ are driven at small $x$ primarily by gluons and are related in the following way to the UGD $x{\cal A}(x,{\mathbf k}_{T}^2,\mu^2)$ $$\begin{aligned} F_{2,L}(x,Q^2) ~=~\int^1_{x} \frac{dz}{z} \int^{Q^2} dk^2_T \sum_{i=u,d,s,c} e^2_i \hat C^g_{2,L}(x/z,Q^2,m_i^2,k^2_T)x{\cal A}(x,{\mathbf k}_{T}^2,\mu^2).\end{aligned}$$ The functions $\hat C^g_{2,L}(x,Q^2,m_i^2,k^2_T)$ can be regarded as SF of the off-shell gluons with virtuality $k^2_T$ (hereafter we call them [*hard structure functions* ]{}). They are described by the sum of the quark box (and crossed box) diagram contribution to the photon-gluon interaction. To apply Eq.(1) for SF at low $Q^2$ we change the low $Q^2$ asymptotics of the QCD coupling constant within hard structure functions. We have used the so called “freezing” procedure in the “soft” form, when the argument of the strong coupling constant is shifted from $Q^2$ to $Q^2 + M^2$ [@NZ]. Then $\alpha_s = \alpha_s(Q^2 + M^2)$. For massless quarks $M=m_{\rho}$ and for massive ones with mass $m_Q, M=2m_Q$. To calculate the SF $F_2^{c,b}$ and $F_{L}(x,Q^2)$ we used the hard SF $\hat C^g_{2,L}(x,Q^2,m^2,k^2_T)$ from Ref. [@KLZ; @KLZ2][^1] and two UGD ${\cal A}(x,{\mathbf k}_{T}^2,\mu^2)$ obtained in our previous paper [@JKLZ]. These UGD are determined by a convolution of the non-perturbative starting distribution ${\cal{A}}_0(x)$ and CCFM evolution denoted by $\bar{\cal A}(x,{\mathbf k}_{T}^2,\mu^2)$: $$\begin{aligned} x{\cal A}(x,{\mathbf k}_{T}^2,\mu^2)~=~\int dz {\cal A}_0(z) {x\over z} {\bar{\cal A}}({x\over z},{\mathbf k}_{T}^2,\mu^2),\end{aligned}$$ [r]{}[0.5]{} ![image](nikolai_zotov.fig1.eps){width="0.45\columnwidth"} where $$x{\cal{A}}_0(x)~=~Nx^{-B_g}(1-x)^{C_g}(1-D_{g}x).$$ The parameters $N, B_g, C_g, D_g$ of ${\cal A}_0$ were determined in the fits to $F_2$ and $F_2^c$ data [@H1; @H12] independently (see [@JKLZ]) Fig. 1 shows the two different UGD. The small $x$ behaviour of these UGD is very different[^2]. To calculate the SF $F_2^b(x, Q^2)$ and $F_L(x,Q^2)$ we took $m_c=1.4$ GeV and $m_b=4.75$ GeV and used the $m^2 = 0$ limit of the above Eq. 1 to evaluate the corresponding lightquark contributions to the $F_L$. Fig. 2 (left panel) shows the $F_2^b$ as a function of $x$ at fixed $Q^2$. Fig.2 (right panel) shows the $F_L$ as a function of $x$ at fixed $Q^2$. Fig. 3 shows the SF $F_L(Q^2)$ at fixed $W$ compared to the H1 data [@H14]. It is interesting to observe, that the measured $F_2^b$ seems to prefer the UGD obtained from the fit to $F_2$ and is inconsitent with the one obtained from $F_2^c$. Also the measured $F_L$ is better described with the UGD from the $F_2$ fit. ![*The SF $F_2^b$ as a function of $x$ at fixed $Q^2$ compared to the H1 data [@H12](left panel) The solid and dotted lines are from CCFM-evolved UGD obtained from the fits to $F_2(x,Q^2)$ and $F_2^c(x,Q^2)$. The SF $F_L$ as a function of $x$ at fixed $Q^2$ compared to the H1 data [@H1; @H13](right panel) The solid and dotted lines are from CCFM-evolved UGD obtained from the fits to $F_2(x,Q^2)$ and $F_2^c(x,Q^2)$.*[]{data-label="Fig:2"}](nikolai_zotov.fig2.eps){height="8.0cm" width="7.0cm"} ![*The SF $F_2^b$ as a function of $x$ at fixed $Q^2$ compared to the H1 data [@H12](left panel) The solid and dotted lines are from CCFM-evolved UGD obtained from the fits to $F_2(x,Q^2)$ and $F_2^c(x,Q^2)$. The SF $F_L$ as a function of $x$ at fixed $Q^2$ compared to the H1 data [@H1; @H13](right panel) The solid and dotted lines are from CCFM-evolved UGD obtained from the fits to $F_2(x,Q^2)$ and $F_2^c(x,Q^2)$.*[]{data-label="Fig:2"}](nikolai_zotov.fig3.eps){height="8.0cm" width="7.0cm"} ![*The $Q^2$ dependence of SF $F_L(Q^2)$ at fixed $W = 276$ GeV compared to the H1 data [@H14] The solid and dotted lines are from CCFM-evolved UGD obtained from the fits to $F_2(x,Q^2)$ and $F_2^c(x,Q^2)$.*[]{data-label="Fig:3"}](nikolai_zotov.fig4.eps){height="5cm" width="13cm"} In summary the $k_T-$ factorization approach with the CCFM-evolved UGD obtained from the fits to the $F_2(x, Q^2)$ data reproduces the H1 data for SF $F_2^b(x, Q^2)$, $F_L(x, Q^2)$ and $F_L$ at fixed $W$ (see [@JKLZ]). The UGD obtained from the fit to $F_2^c$ seems to overshoot the measured $F_2^b$ and $F_L$ at small $x$. New experimental data for $F_L (x, Q^2)$ but also more precise measurements of the heavy quark structure functions are very important for a precise determination of the UGD. [99]{} S. Catani, M. Ciafaloni and F. Hautmann, Nucl. Phys. [**B366**]{} 135 (1991);\ J.C. Collins and R.K. Ellis, Nucl. Phys. [**B360**]{} 3 (1991);\ E. Levin, M. Ryskin, Yu. Shabelski and A. Shuvaev Sov. J. Nucl. Phys. [**53**]{} 657 (1991). R.S. Thorne, arXiv:hep-ph/0511351;\ C.D. White and R.S. Thorne, Phys. Rev. [**D74**]{} 014002 (2006); [**D75**]{} 034005 (2007). L.N. Lipatov, Sov. J. Nucl. Phys. [**23**]{} 338 (1976);\ E.A. Kuraev, L.N. Lipatov and V.S. Fadin, Sov. Phys. JETP [**44**]{} 443 (1976); [**45**]{} 199 (1977);\ Ya.Ya. Balitzki and L.N. Lipatov, Sov. J. Nucl. Phys. [**28**]{} 822 (1978);\ L.N. Lipatov, Sov. Phys. JETP [**63**]{} 904 (1986). M. Ciafaloni, Nucl. Phys. [**B296**]{} 49 (1988);\ S. Catani, F. Fiorani and G. Marchesini, Nucl. Phys. [**B336**]{} 18 (1995);\ G. Marchesini, Nucl. Phys. [**B445**]{} 49 (1995). Bo Andersson et al. (Small $x$ Collaboration), Eur. Phys. J. [**C25**]{} 77 (2002);\ J. Andersen et al. (Small $x$ Collaboration), Eur. Phys. J. [**C25**]{} 67(2002); [**C35**]{} 67 (2004) S. Catani and F. Hautmann, Nucl. Phys. [**B427**]{} 475 (1994);\ S. Catani, arXiv:hep-ph/9608310. B.Badelek, J.Kwiecinski and A. Stasto, Z. Phys. [**C74**]{} 297 (1997). A.V. Kotikov, A.V. Lipatov and N.P. Zotov, Eur. Phys. J. [**C26**]{} 51 (2002). N. Nikolaev and B.M. Zakharov, Z. Phys. [**C49**]{} 607 (1991); [**C53**]{} 331 (1992). A.V. Kotikov, A.V. Lipatov and N.P. Zotov, Eur. Phys. J. [**C27**]{} 219 (2003). S. Catani, M. Ciafaloni and F. Hautmann, [*Proceedings of the Workshop on Physics at HERA, Hamburg, Germany (1991), v. 2, p. 690*]{}. H. Jung, A.V. Kotikov, A.V. Lipatov and N.P. Zotov, arXiv: hep-ph/0611093. H1 Collab., A. Adloff et al., Eur. Phys. J. [**C21**]{} 33 (2001). H1 Collab., A. Adloff et al., Phys. Lett. [**B528**]{} 199 (2002); A. Aktas et al., Eur. Phys. J. [**C40**]{} 349 (2005); [**C45**]{} 23 (2006). H. Jung, talk in HFS working group on DIS’07. H1 Collab., A. Aid et al., Phys. Lett. [**B393**]{} 452 (1997); N. Gogitidze, J. Phys. [**G28**]{} 751 (2002). E.M. Lobodzinska, [*Proceedings of the DIS 2003, Gatchina, St. Petersburg, Russia, p. 93*]{}. [^1]: There is full agreement of our results with the formulae for the photoproduction of heavy quarks from Ref. [@CCH2]. [^2]: See also Ref. [@HJ].
{ "pile_set_name": "ArXiv" }
--- abstract: 'We introduce a version of Stein’s method for proving concentration and moment inequalities in problems with dependence. Simple illustrative examples from combinatorics, physics, and mathematical statistics are provided.' address: '367 Evans Hall \#3860Department of StatisticsUniversity of CaliforniaBerkeley, CA 94720-3860 ' author: - Sourav Chatterjee title: 'Stein’s method for concentration inequalities' --- \[section\] \[thm\][Lemma]{} \[thm\][Corollary]{} \[thm\][Proposition]{} \[thm\][Definition]{} Introduction and results {#intro} ======================== Stein’s method was introduced by Charles Stein [@stein72] in the context of normal approximation for sums of dependent random variables. Stein’s version of his method, best known as the “method of exchangeable pairs”, attained maturity in his later work [@stein86]. A reasonably large literature has developed around the subject, but it has almost exclusively developed as a method of proving distributional convergence with error bounds. Stein’s attempts at getting large deviations in [@stein86] did not, unfortunately, prove fruitful. Some progress for sums of dependent random variables was made by Raič [@raic04]. A general version of Stein’s method for concentration inequalities was introduced for the first time in the Ph.D. thesis [@chatterjee05] of the present author. The purpose of this paper is to explain the theory developed in [@chatterjee05] via examples. Another application is in [@chatterjee05a]. This section is organized as follows: First, we give three examples, followed by the main abstract theorem; finally, towards the end of the section, we present very condensed overviews of Stein’s method, concentration of measure, and the related literature. Proofs are in section \[proofs\]. A generalized matching problem {#perm} ------------------------------ Let $\{a_{ij}\}$ be an $n\times n$ array of real numbers. Let $\pi$ be chosen uniformly at random from the set of all permutations of $\{1,\ldots,n\}$, and let $X=\sum_{i=1}^n a_{i\pi(i)}$. This class of random variables was first studied by Hoeffding [@hoeffding51], who proved that they are approximately normally distributed under certain conditions. It is easy to see that various well-studied functions of random permutations, like the number of fixed points, the sum of a random sample picked without replacement from a finite population, and the function $\sum_i|i-\pi(i)|$ (known as Spearman’s footrule [@diaconisgraham77]), are all instances of Hoeffding’s statistic. Hoeffding’s statistic has a long history of association with Stein’s method. In fact, in an unpublished work Stein introduced his method to treat the normal approximation problem for this object. Bolthausen [@bolthausen84] used Stein’s method to give a Berry-Esseen bound. Bolthausen and Götze [@bolthausengotze93] gave multivariate central limit theorems under a further generalized setup. However, we have not seen large deviations or concentration bounds using any method. Our version of Stein’s method enables us to easily derive the following nice tail bound. \[hoeff3\] Let $\{a_{ij}\}_{1\le i,j\le n}$ be a collection of numbers from $[0,1]$. Let $X = \sum_{i=1}^n a_{i\pi(i)}$, where $\pi$ is drawn from the uniform distribution over the set of all permutations of $\{1,\ldots,n\}$. Then $${\mathbb{P}}\{|X - {\mathbb{E}}(X)| \ge t\} \le 2\exp\biggl(-\frac{t^2}{4{\mathbb{E}}(X) + 2t}\biggr)$$ for any $t\ge 0$. Note that the bound does not have an explicit dependence on $n$. Note also the automatic transition from Poissonian to gaussian tails as ${\mathbb{E}}(X)$ becomes large (when ${\mathbb{E}}(X)$ is small the bound is like $\exp(-Ct)$, whereas when ${\mathbb{E}}(X)$ is large, it is essentially a gaussian tail with standard deviation $\sqrt{{\mathbb{E}}(X)}$.). These two properties characterize it as a so-called “Bernstein type inequality”, named after the classical Bernstein inequality (see [@shorackwellner86], page 855) for sums of bounded independent random variables. The classical result of Maurey [@maurey79] can only imply the weaker inequality $P(X> {\mathbb{E}}(X) +t) \le e^{-t^2/4n}$. However, it is possible to derive a Bernstein bound similar to Proposition \[hoeff3\] (albeit with a significantly worse constant in the exponent) using Michel Talagrand’s deep theorem about concentration of random permutations (Theorem 5.1 in Section 5 of [@talagrand95]; see also McDiarmid [@mcdiarmid02] and Luczak & McDiarmid [@luczakmcdiarmid03]). For a concrete application, let $X$ be the number of fixed points of a random permutation $\pi$. Then $X =\sum_{i=1}^n a_{i\pi(i)}$, where $a_{ij} = {\mathbb{I}}_{\{i=j\}}$. Since ${\mathbb{E}}(X)=1$, Proposition \[hoeff3\] gives ${\mathbb{P}}\{|X-1|\ge t\} \le 2\exp(-t^2/(4+2t))$. Of course, we do not expect this to be the best possible bound in this very well-understood problem; this is just meant to be an illustration. In fact, the exact distribution of the the number of fixed points is known (see Feller [@feller1], section IV.4), which gives a tail bound like $\exp(-Ct\log t)$. Finally, we also have a “Burkholder-Davis-Gundy” type inequality for Hoeffding’s statistic which does not require a bound on the $a_{ij}$’s. \[hoeff2\] Let $\{a_{ij}\}_{1\le i,j\le n}$ be an arbitrary collection of real numbers. Let $\pi$ be a uniform random permutation, and let $X = \sum_{i=1}^n a_{i\pi(i)}$. Define $$\Delta = \frac{1}{4n}\sum_{i,j} (a_{i\pi(i)} + a_{j\pi(j)}-a_{i\pi(j)}-a_{j\pi(i)})^2.$$ Then for every positive integer $k$, we have ${\mathbb{E}}(X-{\mathbb{E}}(X))^{2k} \le (2k-1)^k {\mathbb{E}}\Delta^k$. For a general exposition about the famous Burkholder-Davis-Gundy martingale inequalities we refer to the article by Burkholder [@burkholder73]. Magnetization in the Curie-Weiss model {#curieweiss} -------------------------------------- Fix any $\beta \ge 0$, $h\in {\mathbb{R}}$, and consider the probability mass function (the Gibbs measure) on $\{-1,1\}^n$ given by $$\label{gibbs} {\mathbb{P}}(\{\sigma\}) := Z^{-1}\exp\biggl(\frac{\beta}{n}\sum_{i<j} \sigma_i\sigma_j +\beta h \sum_i \sigma_i\biggr),$$ where $\sigma = (\sigma_1,\ldots,\sigma_n)$ is a typical element of $\{-1,1\}^n$ and $Z$ is the normalizing constant (depends on $\beta$ and $h$). This is known as the ‘Curie-Weiss model of ferromagnetic interaction’ at inverse temperature $\beta$ and external field $h$. The $\sigma_i$’s stand for the spins of $n$ particles, each having a spin of $+1$ or $-1$. The ferromagnetic interaction between the particles is captured in a very simplistic manner by the first term in the hamiltonian. The [*magnetization*]{} of the system, as a function of the configuration $\sigma$, is defined as $m(\sigma) := \frac{1}{n}\sum_{i=1}^n \sigma_i$. If $n$ is large and $\sigma$ is drawn from the Gibbs measure, then the magnetization satisfies $$\label{meq} m(\sigma) \approx \tanh(\beta m(\sigma) + \beta h).$$ with high probability. The equation has a unique root for small values of $\beta$ and multiple solutions for $\beta$ above a critical value. In the physics parlance, this is described by saying that the Curie-Weiss model exhibits “spontaneous magnetization” at low temperatures. For a formal discussion with rigorous proofs, we refer to Ellis [@ellis85], section IV.4. The following proposition formalizes with finite sample tail bounds. \[curie\] Suppose $\sigma$ is drawn from the Gibbs measure . Then, for any $\beta \ge 0$, $h\in {\mathbb{R}}$, $n\ge 1$, and $t\ge 0$, the magnetization $m := \frac{1}{n}\sum_i \sigma_i$ satisfies $${\mathbb{P}}\biggl\{\bigl|m - \tanh(\beta m + \beta h)\bigr| \ge \frac{\beta}{n} + \frac{t}{\sqrt{n}}\biggr\} \le 2\exp\biggl(-\frac{t^2}{4(1+\beta)}\biggr).$$ Although the Curie-Weiss model is a simple model of ferromagnetic interaction, we haven’t encountered any result in the literature which gives an explicit bound like the above. In particular, the result shows concentration of $m(\sigma)$ around the [*set*]{} of roots of $x=\tanh(\beta x +\beta h)$, and not just its mean. However, concentration inequalities for Gibbs measures without explicit constants under various mixing conditions have been obtained before. For a history of the literature and some significant recent progress, we refer to Chazottes et. al. [@chazottes06]. Least squares estimation in the Ising model {#least} ------------------------------------------- The Ising model is another model of ferromagnetic interaction. Given an undirected graph $G = (V,E)$ on the vertex set $V = \{1,\ldots,n\}$, the Ising model without external field assigns the following probability density on $\{-1,1\}^n$: $$\label{ising} {\mathbb{P}}(\{\sigma\}) = Z(\beta)^{-1} \exp\biggl(\beta \sum_{\{i,j\}\in E } \sigma_i\sigma_j\biggr).$$ Here, as before, $\beta$ is the inverse temperature and $Z(\beta)$ is the normalizing constant. A natural statistical problem in this model is the following: How to make inference about $\beta$ when your data is a single configuration generated from the Gibbs measure? The classical maximum likelihood approach for this problem was first considered by Pickard [@pickard87]. Iterative methods for computing the maximum likelihood estimator (e.g. Geyer & Thompson [@geyerthompson92], Jerrum & Sinclair [@jerrumsinclair93]) are widely used nowadays. The Jerrum-Sinclair algorithm for computing the normalizing constant in the Ising model provably converges in polynomial time. However, it is not so clear whether the MLE is a good estimator at all, particularly at critical temperatures. Here we investigate a method of estimating $\beta$ by minimizing an explicit sum-of-squares. First, let $\sigma$ be drawn from the Gibbs measure on $\{-1,1\}^n$, and for each $i$, let $$m_i := \sum_{j:\{i,j\}\in E} \sigma_j.$$ For each $u \ge 0$, let $$\label{stheta} S(u) := \frac{1}{n}\sum_{i=1}^n \bigl(\sigma_i - \tanh(u m_i))^2.$$ The ‘least-squares estimate’ of $\beta$ is defined to be $$\hat{\beta}_{LS} := {\operatorname{argmin}}_{u\ge 0} S(u).$$ Note that it is practically very easy to compute $\hat{\beta}_{LS}$, because $S$ is a smooth function of a single variable. The least-squares technique is well-known and commonly used in the analysis of gaussian Markov random field (GMRF) models (probably originating from Besag [@besag75]), but rigorous results are scarce. Proposition \[isingprop\] (stated below) shows that the random function $S$ indeed attains an approximate global minimum near $\beta$. In fact, it gives $${\mathbb{E}}|S(\beta) - \min_{u\ge 0} S(u)| = O\biggl(\sqrt{\frac{r\log n}{n}}\biggr),$$ where $r$ is the maximum degree of the dependency graph $G$ (recall that the degree of a vertex is the number of neighbors of that vertex, and the maximum degree of a graph is the maximum vertex degree). \[isingprop\] Let $r$ be the maximum degree of the dependency graph $G$ in the Ising model , and let $S(u)$ be defined as in . Take any $t\ge 0$ and let $$\varepsilon = \sqrt{\frac{r(\log n + t)}{n}}.$$ Then we have $${\mathbb{P}}\{S(\beta) \ge \min_{u\ge 0} S(u) + C\varepsilon\} \le \exp(-Kt^2),$$ where $C$ and $K$ are numerical constants. Although it is unclear whether Proposition \[isingprop\] is useful from a statistical point of view, it seems to be interesting as a mathematical result. For instance, observe that the conclusion is valid at any temperature. This is quite remarkable, since the low temperature phase in the Ising model is notoriously intractable for most graphs. Here we should also mention that the technique can be easily applied to the Ising model with an external field, but we prefer to restrict ourselves to the problem of estimating a single parameter (the temperature) for the sake of clarity. The abstract result {#theory} ------------------- The following theorem encapsulates the concentration and moment inequalities used to work out all the examples in this paper. \[conc\] Let ${\mathcal{X}}$ be a separable metric space and suppose $(X,{X^\prime})$ is an exchangeable pair of ${\mathcal{X}}$-valued random variables. Suppose $f:{\mathcal{X}}{\rightarrow}{\mathbb{R}}$ and $F:{\mathcal{X}}\times {\mathcal{X}}{\rightarrow}{\mathbb{R}}$ are square-integrable functions such that $F$ is antisymmetric (i.e. $F(X,{X^\prime})=-F({X^\prime},X)$ a.s.), and ${\mathbb{E}}(F(X,{X^\prime})\mid X) = f(X)$ a.s. Let $$\Delta(X) := \frac{1}{2}{\mathbb{E}}\bigl(|(f(X)-f({X^\prime}))F(X,{X^\prime})|\,\bigl|\, X\bigr).$$ Then ${\mathbb{E}}(f(X)) = 0$, and the following concentration results hold for $f(X)$: 1. If ${\mathbb{E}}(\Delta(X)) <\infty$, then ${\mathrm{Var}}(f(X)) = \frac{1}{2}{\mathbb{E}}((f(X)-f({X^\prime}))F(X,{X^\prime}))$. 2. Assume that ${\mathbb{E}}(e^{\theta f(X)} |F(X,{X^\prime})|) < \infty$ for all $\theta$. If there exists nonnegative constants $B$ and $C$ such that $\Delta(X) \le B f(X) + C$ almost surely, then for any $t\ge 0$, $${\mathbb{P}}\{f(X) \ge t\}\le \exp\biggl(-\frac{t^2}{2C + 2Bt}\biggr) \ \ \text{and} \ \ {\mathbb{P}}\{f(X) \le -t\}\le \exp\biggl(-\frac{t^2}{2C}\biggr).$$ 3. For any positive integer $k$, we have the following exchangeable pairs version of the Burkholder-Davis-Gundy inequality: $${\mathbb{E}}(f(X)^{2k}) \le (2k-1)^k {\mathbb{E}}(\Delta(X)^k).$$ To see how the exchangeable pairs are constructed and the theorem is applied in our examples, one has to look at the proofs in section \[proofs\]. However, for a quick illustration, we will now work out the inequalities for sums of independent random variables, taking care to spell out details. Simplest example ---------------- Let $X = \sum_{i=1}^n Y_i$, where $Y_i$’s are independent square integrable random variables. Let $\mu_i ={\mathbb{E}}(Y_i)$ and $\sigma_i^2 = {\mathrm{Var}}(Y_i)$. An exchangeable pair is created by choosing a coordinate $I$ uniformly at random from $\{1,\ldots,n\}$, and defining $${X^\prime}= \sum_{j\ne I} Y_j + Y^\prime_I,$$ where $Y^\prime_1,\ldots,Y^\prime_n$ are independent copies of $Y_1,\ldots,Y_n$. Let $$F(x,y) = n(x-y).$$ Then $$\begin{aligned} {\mathbb{E}}(F(X,{X^\prime})\mid Y_1,\ldots,Y_n) &= \frac{1}{n}\sum_{i=1}^n {\mathbb{E}}(n(Y_i - Y^\prime_i)\mid Y_1,\ldots,Y_n) \\ &= \sum_{i=1}^n (Y_i - \mu_i) = X - {\mathbb{E}}(X).\end{aligned}$$ Since the right hand side depends only on $X$, we have $$f(X) = {\mathbb{E}}(F(X,{X^\prime})\mid X) = X - {\mathbb{E}}(X).$$ Thus, from part ($i$) of Theorem \[conc\] we get the elementary identity $${\mathrm{Var}}(X) = \frac{1}{2}\sum_{i=1}^n {\mathbb{E}}(Y_i - Y^\prime_i)^2 = \sum_{i=1}^n \sigma_i^2.$$ Now note that $$\begin{aligned} \Delta(X) &= \frac{n}{2}{\mathbb{E}}((X-{X^\prime})^2 \mid X) \\ &= \frac{1}{2}\sum_{i=1}^n{\mathbb{E}}((Y_i-{Y^\prime}_i)^2\mid X).\end{aligned}$$ If $c_1,\ldots,c_n$ are constants such that $|Y_i-\mu_i|\le c_i$ a.s. for each $i$, then $$\begin{aligned} {\mathbb{E}}((Y_i-{Y^\prime}_i)^2 \mid X) &= {\mathbb{E}}((Y_i-\mu_i)^2\mid X) + {\mathbb{E}}(({Y^\prime}_i-\mu_i)^2)\\ &\le c_i^2 + \sigma_i^2.\end{aligned}$$ Part ($ii$) of Theorem \[conc\] now implies that $${\mathbb{P}}\{|X- {\mathbb{E}}(X)|\ge t\} \le 2 \exp\biggl(-\frac{t^2}{\sum_{i=1}^n (c_i^2 + \sigma_i^2)}\biggr).$$ This is similar to (but not exactly the same as) the classical Hoeffding inequality [@hoeffding63] for sums of bounded random variables. Now suppose that $0\le Y_i \le 1$ a.s. for each $i$. If the $\mu_i$’s are very small, then the Hoeffding bound is wasteful. A more careful analysis gives a better result, as follows. First, note that $$\begin{aligned} \Delta(X) &= \frac{1}{2} \sum_{i=1}^n {\mathbb{E}}((Y_i-{Y^\prime}_i)^2\mid X)\\ &= \frac{1}{2}\sum_{i=1}^n ({\mathbb{E}}Y_i^2 - 2\mu_i {\mathbb{E}}(Y_i\mid X) + {\mathbb{E}}(Y_i^2\mid X)).\end{aligned}$$ Using the assumption that $0\le Y_i\le 1$, we get $$\Delta(X) \le \frac{1}{2}\sum_{i=1}^n ({\mathbb{E}}(Y_i) + {\mathbb{E}}(Y_i\mid X)) = \frac{1}{2}({\mathbb{E}}(X) + X) = \frac{1}{2}f(X) + {\mathbb{E}}(X).$$ Thus, we can take $B = 1/2$ and $C={\mathbb{E}}(X)$ in part ($ii$) of Theorem \[conc\], which gives $${\mathbb{P}}\{|X-{\mathbb{E}}(X)| \ge t\} \le 2\exp\biggl(-\frac{t^2}{2{\mathbb{E}}(X)+t}\biggr).$$ Again, this is a version of the classical Bernstein inequality (see [@shorackwellner86], page 855) for sums of independent random variables. Finally observe that by part ($iii$) of Theorem \[conc\] and an application of Jensen’s inequality, we have for each positive integer $k$, $$\begin{aligned} {\mathbb{E}}(X^{2k}) &\le (2k-1)^k {\mathbb{E}}\biggl(\frac{1}{2}\sum_{i=1}^n {\mathbb{E}}((Y_i-\mu_i)^2 + ({Y^\prime}_i-\mu_i)^2\mid X)\biggr)^k\\ &\le (2k-1)^k {\mathbb{E}}\biggl(\sum_{i=1}^n (Y_i-\mu_i)^2\biggr)^k.\end{aligned}$$ This is exactly what the Burkholder-Davis-Gundy inequality [@burkholder73] would give us for sums of independent random variables (although in this case, it can be derived by easier methods). In the remainder of this section, we give very short overviews of Stein’s method and concentration of measure. Stein’s method -------------- Suppose we want to show that a random variable $X$ taking value in some space ${\mathcal{X}}$ has approximately the same distribution as some other random variable $Z$. The classical version of Stein’s method [@stein72; @stein86] involves four steps: 1. Identify a “characterizing operator” $T$ for $Z$, which has the defining property that for any function $g$ belonging to a fixed large class of functions, ${\mathbb{E}}Tg (Z) = 0$. For instance, if ${\mathcal{X}}= {\mathbb{R}}$ and $Z$ is a standard gaussian random variable, then $Tg(x) := g^\prime(x)-xg(x)$ is a characterizing operator, acting on all locally absolutely continuous $g$ with subexponential growth at infinity. 2. Construct a random variable ${X^\prime}$ such that $(X,{X^\prime})$ is an exchangeable pair. 3. Find an operator $\alpha$ such that for any suitable $h:{\mathcal{X}}{\rightarrow}{\mathbb{R}}$, $\alpha h$ is an antisymmetric function (i.e. $\alpha h(x,y)\equiv -\alpha h(y,x)$) and $$|{\mathbb{E}}(\alpha h(X,{X^\prime})|X = x) - Th(x)| \le \varepsilon_h,$$ where $\varepsilon_h$ is a small error depending only on $h$. 4. Take a function $g$ and find $h$ such that $Th(x) = g(x) - {\mathbb{E}}g(Z)$. By antisymmetry of $\alpha h$ and the exchangeability of $(X,{X^\prime})$, it follows that ${\mathbb{E}}(\alpha h(X,{X^\prime}))=0$. Combining with the previous step, we have the error bound $|{\mathbb{E}}g(X) - {\mathbb{E}}g(Z)|\le \varepsilon_h$. There are other variants of Stein’s method, most notably the generator method of Andrew Barbour [@barbour90], the dependency graph approach introduced by Chen [@chen75] and Baldi and Rinott [@baldirinott89] and popularized by Arratia, Goldstein and Gordon [@agg90], the size-biased coupling method of Barbour, Holst and Janson [@bhj92], and the zero-biased coupling method due to Goldstein and Reinert [@goldsteinreinert97]. The recent applications to algebraic problems by Jason Fulman [@fulman04; @fulman05], and the quest for Berry-Esseen bounds by Rinott and Rotar [@rinottrotar97] and Shao and Su [@shaosu04] are also worthy of note. However, it is not our purpose here to go deeply into the regular versions of Stein’s method. For further references and exposition, we refer to the recent monograph [@stein04]. For applications of the method of exchangeable pairs and other versions of Stein’s method to Poisson approximation, one can look at the survey paper by Chatterjee, Diaconis & Meckes [@cdm05]. Concentration inequalities -------------------------- The theory of concentration inequalities tries to answer the following question: Given a random variable $X$ taking value in some measure space ${\mathcal{X}}$ (which is usually some high dimensional Euclidean space), and a measurable map $f:{\mathcal{X}}{\rightarrow}{\mathbb{R}}$, what is a good explicit bound on ${\mathbb{P}}\{|f(X)-{\mathbb{E}}f(X)| \ge x\}$? Exact evaluation or accurate approximation is, of course, the central purpose of probability theory itself. In situations where this is not possible, concentration inequalities aim to do the next best job by providing rapidly decaying tail bounds. The literature on concentration inequalities is huge — from the pioneering inequalities of Hoeffding [@hoeffding63] to the momentous work of Talagrand [@talagrand95] — but most of it revolves around well-behaved functions of independent random variables. For a nearly complete account of the literature until the year 2001, we redirect the reader to the definitive resource in this subject — the monograph [@ledoux01] by Michel Ledoux. The methods of Kim and Vu [@kimvu04] and Boucheron, Lugosi, and Massart [@blm03] are significant recent developments. The techniques developed in [@chatterjee05] (and partially presented here) have some basic similarities with the concentration results of Schmuckenschläger [@schmuck98], but go much beyond that in terms of applications. Other than that (and log-Sobolev inequalities, which are much harder to obtain anyway) there is very little — even in the vast concentration literature — about the concentration of functions of dependent random variables, particularly in the discrete setting. We hope that our version of Stein’s method will partially fill this void. .2in [**Acknowledgments.**]{} I am grateful to Persi Diaconis and Yuval Peres for many useful comments and suggestions. Thanks are also due to the two anonymous referees for pointing out several omissions and errors. Proofs ====== Before proving Theorem \[conc\], let us see how it is applied to work out the three examples described in section \[intro\]. .1in [**Proof of Proposition \[hoeff3\].**]{} Construct ${X^\prime}$ as follows: Choose $I,J$ uniformly and independently at random from $\{1,\ldots, n\}$. Let $\pi^\prime = \pi\circ (I,J)$, where $(I,J)$ denotes the transposition of $I$ and $J$. It can be easily verified that $(\pi,\pi^\prime)$ is an exchangeable pair. Hence if we let $${X^\prime}:= \sum_{i=1}^n a_{i\pi^\prime(i)},$$ then $(X,{X^\prime})$ is also an exchangeable pair. Now note that $$\begin{aligned} \frac{1}{2}{\mathbb{E}}(n(X - {X^\prime})|\pi) &= \frac{n}{2}{\mathbb{E}}(a_{I\pi(I)} + a_{J\pi(J)} - a_{I\pi(J)} - a_{J\pi(I)}|\pi) \\ &= \frac{1}{n} \sum_{i,j} a_{i\pi(i)} - \frac{1}{n}\sum_{i,j} a_{i\pi(j)} \\ &= X-{\mathbb{E}}(X).\end{aligned}$$ Thus, we can take $f(x)= x-{\mathbb{E}}(X)$ and $F(x,y) = \frac{1}{2}n(x-y)$. Now note that since $0\le a_{ij}\le 1$ for all $i$ and $j$, we have $$\begin{aligned} &\frac{1}{2}{\mathbb{E}}\bigl(|(f(X)-f({X^\prime}))F(X,{X^\prime})| \, \bigl| \,\pi\bigr) = \frac{n}{4}{\mathbb{E}}((X-{X^\prime})^2|\pi) \\ &=\frac{1}{4n}\sum_{i,j} (a_{i\pi(i)} + a_{j\pi(j)}-a_{i\pi(j)}-a_{j\pi(i)})^2 \\ &\le \frac{1}{2n} \sum_{i,j}(a_{i\pi(i)} + a_{j\pi(j)}+a_{i\pi(j)}+a_{j\pi(i)}) \\ &= X + {\mathbb{E}}(X) = f(X) + 2{\mathbb{E}}(X).\end{aligned}$$ Since the last quantity depends only on $X$ it follows that $\Delta(X)= f(X)+2{\mathbb{E}}(X)$. Applying part ($ii$) of Theorem \[conc\] with $B = 1$ and $C= 2{\mathbb{E}}(X)$ completes the proof. $\Box$ .2in [**Proof of Proposition \[hoeff2\].**]{} Follows directly from part $(iii)$ of Theorem \[conc\] and the computations done in the proof of Proposition \[hoeff3\].$\Box$ .2in [**Proof of Proposition \[curie\].**]{} Suppose $\sigma$ is drawn from the Gibbs distribution. We construct ${\sigma^\prime}$ by taking a step in the Gibbs sampler as follows: Choose a coordinate $I$ uniformly at random, and replace the $I^{\mathrm{th}}$ coordinate of $\sigma$ by an element drawn from the conditional distribution of the $I^{\mathrm{th}}$ coordinate given the rest. It is well-known and easy to prove that $(\sigma,{\sigma^\prime})$ is an exchangeable pair. Let $$F(\sigma, {\sigma^\prime}) := \sum_{i=1}^n (\sigma_i - {\sigma^\prime}_i).$$ Now define $$m_i(\sigma) := \frac{1}{n}\sum_{j \le n, j\ne i} \sigma_j, \ \ i=1,\ldots,n.$$ Since the Hamiltonian is a simple explicit function, the conditional distribution of the $i^{\mathrm{th}}$ coordinate given the rest is easy to obtain. An easy computation gives ${\mathbb{E}}(\sigma_i|\{\sigma_j, j\ne i\}) = \tanh(\beta m_i+\beta h)$. Thus, we have $$\begin{aligned} f(\sigma) = {\mathbb{E}}(F(\sigma,{\sigma^\prime})|\sigma) &= \frac{1}{n}\sum_{i=1}^n (\sigma_i -{\mathbb{E}}(\sigma_i|\{\sigma_j, j\ne i\}))\\ &= m - \frac{1}{n}\sum_{i=1}^n \tanh(\beta m_i + \beta h).\end{aligned}$$ Now note that $|F(\sigma,{\sigma^\prime})| \le 2$, because $\sigma$ and ${\sigma^\prime}$ differ at only one coordinate. Also, since the map $x\mapsto \tanh x$ is $1$-Lipschitz, we have $$\begin{aligned} |f(\sigma)-f({\sigma^\prime})| &\le |m(\sigma)-m({\sigma^\prime})| +\frac{\beta}{n}\sum_{i=1}^n |m_i(\sigma) - m_i({\sigma^\prime})| \le \frac{2(1+\beta)}{n}.\end{aligned}$$ Thus, by part $(ii)$ of Theorem \[conc\] we have $${\mathbb{P}}\biggl\{\biggl|m - \frac{1}{n}\sum_{i=1}^n \tanh(\beta m_i + \beta h)\biggr| \ge \frac{t}{\sqrt{n}}\biggr\} \le 2\exp\biggl(-\frac{t^2}{4(1+\beta)}\biggr).$$ Finally note that for each $i$, by the Lipschitz nature of the $\tanh$ function, we get $$\begin{aligned} &\biggl|\frac{1}{n}\sum_{i=1}^n \tanh(\beta m_i + \beta h) - \tanh(\beta m + \beta h) \biggr|\\ &\le \frac{1}{n}\sum_{i=1}^n|\tanh(\beta m_i+\beta h) - \tanh(\beta m+\beta h)| \\ &\le \frac{1}{n}\sum_{i=1}^n\beta|m_i-m| \le \frac{\beta}{n}.\end{aligned}$$ This completes the proof. $\Box$ .2in [**Proof of Proposition \[isingprop\].**]{} As in the proof of Proposition \[curie\], we produce ${\sigma^\prime}$ by taking a step in the Gibbs sampler: A coordinate $I$ is chosen uniformly at random, and $\sigma_I$ is replace by ${\sigma^\prime}_I$ drawn from the conditional distribution of the $I^{\mathrm{th}}$ coordinate given $(\sigma_j)_{j\ne I}$. For each $i$, let $$m_i = m_i(\sigma) := \sum_{j: \{i,j\}\in E} \sigma_j.$$ Now fix $u\ge 0$ and define $$F(\sigma, {\sigma^\prime}) := (\sigma_I- {\sigma^\prime}_I) (\tanh(\beta m_I) - \tanh(u m_I)).$$ Then $F(\sigma,{\sigma^\prime})= -F({\sigma^\prime},\sigma)$ because $m_I(\sigma) = m_I({\sigma^\prime})$ . Now let $$\begin{aligned} &f(\sigma) := {\mathbb{E}}(F(\sigma,{\sigma^\prime})\mid \sigma) \\ &= \frac{1}{n}\sum_{i=1}^n(\sigma_i - \tanh(\beta m_i)) (\tanh(\beta m_i) - \tanh(u m_i)).\end{aligned}$$ Now, if $r$ is the maximum degree of $G$, then at most $r+1$ terms in the sums defining $f(\sigma)$ and $f({\sigma^\prime})$ are unequal, and they all lie in the interval $[-4,4]$. Thus, $|f(\sigma)-f({\sigma^\prime})|\le 8(r+1)/n$. Also, evidently, $|F(\sigma, {\sigma^\prime})|\le 4$. Using all this information in part ($ii$) of Theorem \[conc\], we get $${\mathbb{P}}\{f(\sigma)\le -t\} \le \exp\biggl(-\frac{nt^2}{32(r+1)}\biggr).$$ Now, a direct verification shows that $$\begin{aligned} S(u) - S(\beta) &= \frac{1}{n}\sum_{i=1}^n (\tanh\beta m_i)-\tanh(u m_i))^2 + 2f(\sigma).\end{aligned}$$ Thus, $$\label{tail1} {\mathbb{P}}\{S(\beta) \ge S(u) + t\} \le {\mathbb{P}}\{2f(\sigma) \le -t\} \le \exp\biggl(-\frac{nt^2}{128(r+1)}\biggr).$$ Now note that for any $u,v\ge 0$, we have $$\begin{aligned} &|S(u)-S(v)| \\ &\le \frac{1}{n}\sum_{i=1}^n |(2\sigma_i - \tanh(u m_i) - \tanh(vm_i))(\tanh(vm_i)-\tanh(um_i))|\\ &\le \frac{4}{n}\sum_{i=1}^n |\tanh(vm_i) - \tanh(um_i)| \le 4r|u-v|,\end{aligned}$$ since $|m_i(u-v)|\le r|u-v|$. Let $N = \lfloor\sqrt{nr\log n}\rfloor$, and let $$u_k = k\sqrt{\frac{\log n}{nr}} \ \text{ for } \ k=1,2,\ldots,N.$$ Then, if $u_{k-1} \le u\le u_k$, the above inequality gives $$|S(u)-S(u_k)| \le 4r|u-u_k|\le 4\sqrt{\frac{r\log n}{n}}.$$ Now take any $u \ge u_N$. Since $m_i \in \{0,\pm 1,\ldots, \pm r\}$, therefore $|\tanh(um_i) - \tanh(u_N m_i)|\le 1-\tanh(u_N|m_i|)\le 1-\tanh(u_N)$. Thus, $$\begin{aligned} |S(u)-S(u_N)| &\le \frac{4}{n}\sum_{i=1}^n |\tanh(um_i)-\tanh(u_N m_i)| \\ &\le 4(1-\tanh(u_N)) \le 4e^{-u_N} \le \frac{4e}{n}.\end{aligned}$$ If $n \ge 3$, then $\sqrt{\log n/n} \ge e/n$. Combining the steps, we see that for $n\ge 3$, $$\min_{1\le k\le N} S(u_k) \le \min_{u\ge 0} S(u) + 4\sqrt{\frac{r\log n}{n}}.$$ Finally, combining this with , we get $$\begin{aligned} &{\mathbb{P}}\biggl\{S(\beta) \ge \min_{u\ge 0} S(u) + 4\sqrt{\frac{r\log n}{n}} + t\biggr\}\\ &\le {\mathbb{P}}\bigl\{S(\beta) \ge \min_{1\le k\le N} S(u_k) + t\bigr\}\\ &\le \sum_{k=1}^N {\mathbb{P}}\bigl\{S(\beta) \ge S(u_k) + t\bigr\} \le N \exp\biggl(-\frac{nt^2}{128(r+1)}\biggr).\end{aligned}$$ It is now easy to complete the proof by substituting the value of $N$ and choosing $t > \sqrt{Cr\log n/n}$ for sufficiently large $C$, so that the effect of $N$ washes out. $\Box$ .2in Finally, let us prove our main result. .1in [**Proof of Theorem \[conc\].**]{} Let us begin with a useful general identity. Suppose $h:{\mathcal{X}}{\rightarrow}{\mathbb{R}}$ is any measurable map such that ${\mathbb{E}}|h(X) F(X,{X^\prime})|<\infty$. Then clearly ${\mathbb{E}}(h(X)f(X)) = {\mathbb{E}}(h(X)F(X,{X^\prime}))$. Using the exchangeability of $X$ and ${X^\prime}$, and the antisymmetric nature of $F$, we have $${\mathbb{E}}(h(X)F(X,{X^\prime})) = {\mathbb{E}}(h({X^\prime})F({X^\prime},X)) = -{\mathbb{E}}(h({X^\prime})F(X,{X^\prime})).$$ Thus, we have $$\label{basis} {\mathbb{E}}(h(X) f(X)) = {\mathbb{E}}(h(X)F(X,{X^\prime})) = \frac{1}{2}{\mathbb{E}}((h(X)-h({X^\prime}))F(X,{X^\prime})).$$ The above equation is the basis of all that follows. First, note that by putting $h\equiv 1$, we immediately get ${\mathbb{E}}(f(X))=0$, Similarly, part ($i$) of the Theorem follows by putting $h=f$. Next, let us start proving ($ii$). Let $m(\theta) := {\mathbb{E}}(e^{\theta f(X)})$ be the moment generating function of $f(X)$. We can differentiate $m(\theta)$ and move the derivative inside the expectation because of the assumption that ${\mathbb{E}}(e^{\theta f(X)} |F(X,{X^\prime})|) < \infty$ for all $\theta$. Thus, by equation (\[basis\]), we have $$\begin{aligned} m^\prime(\theta) &= {\mathbb{E}}(e^{\theta f(X)} f(X)) = \frac{1}{2} {\mathbb{E}}((e^{\theta f(X)} - e^{\theta f({X^\prime})})F(X,{X^\prime})).\end{aligned}$$ Now note that for any $x,y\in {\mathbb{R}}$, $$\label{expin} \begin{split} &\biggl|\frac{e^x-e^y}{x-y}\biggr| =\int_0^1 e^{tx + (1-t)y}dt \\ &\le \int_0^1 (te^x+(1-t)e^y)dt = \frac{1}{2}(e^x + e^y). \end{split}$$ Using this inequality, and the exchangeability of $X$ and ${X^\prime}$, we get $$\begin{aligned} |m^\prime(\theta)| &\le \frac{|\theta|}{4}{\mathbb{E}}((e^{\theta f(X)} + e^{\theta f({X^\prime})})|(f(X)-f({X^\prime})) F(X,{X^\prime})|) \\ &= \frac{|\theta|}{2} {\mathbb{E}}(e^{\theta f(X)} \Delta(X) + e^{\theta f({X^\prime})} \Delta({X^\prime}))\\ &= |\theta|{\mathbb{E}}(e^{\theta f(X)} \Delta(X))\\ &\le |\theta|{\mathbb{E}}(e^{\theta f(X)} (Bf(X) + C)) = B|\theta| m^\prime(\theta) + C|\theta| m(\theta).\end{aligned}$$ Since $m$ is a convex function and $m^\prime(0) = {\mathbb{E}}(f(X)) = 0$, therefore $m^\prime(\theta)$ always has the same sign as $\theta$. Thus, for $0\le \theta< 1/B$, the above inequality translates into $$\frac{d}{d\theta} \log m(\theta) \le \frac{C\theta}{1-B\theta}.$$ Using this and recalling that $m(0)=1$, we have $$\begin{aligned} \log m(\theta) &\le \int_0^\theta \frac{Cu}{1-Bu}du \le \frac{C\theta^2}{2(1 - B\theta)}.\end{aligned}$$ Putting $\theta = t/(C+Bt)$, we get $$\begin{aligned} {\mathbb{P}}\{f(X) \ge t\} &\le \exp(-\theta t + \log m(\theta)) \le e^{-t^2/(2C + 2Bt)}.\end{aligned}$$ The lower tail can be done similarly; note that for $\theta \le 0$, we have $m^\prime(\theta)\le 0$, and hence $$|m^\prime(\theta)|\le B|\theta| m^\prime(\theta) + C|\theta| m(\theta)\le C|\theta| m(\theta),$$ and this is the reason why $B$ does not appear in the lower tail bound. This completes the proof of part ($ii$). For the moment inequalities in part ($iii$), first observe that by equation (\[basis\]), we have $${\mathbb{E}}(f(X)^{2k}) = \frac{1}{2} {\mathbb{E}}((f(X)^{2k-1} - f({X^\prime})^{2k-1})F(X,{X^\prime})).$$ By the inequality $$|x^{2k-1} - y^{2k-1}| \le \frac{2k-1}{2}(x^{2k-2} + y^{2k-2})|x-y|$$ which follows easily from a convexity argument very similar to (\[expin\]), we have $$\begin{aligned} {\mathbb{E}}(f(X)^{2k}) &\le (2k-1){\mathbb{E}}(f(X)^{2k-2} \Delta(X))\end{aligned}$$ By Hölder’s inequality, we get $${\mathbb{E}}(f(X)^{2k}) \le (2k-1)({\mathbb{E}}(f(X)^{2k}))^{(k-1)/k}({\mathbb{E}}(\Delta(X)^k))^{1/k}.$$ The proof is completed by transferring ${\mathbb{E}}(f(X)^{2k})^{(k-1)/k}$ to the other side.$\Box$ .2in [99]{} and [Gordon, L. (1989).]{} Two moments suffice for Poisson approximations: the Chen-Stein method. [*Ann. Probab.*]{} [**17**]{} no. 1, 9–25. and [Gordon, L. (1992).]{} Poisson approximation and the Chen-Stein method. [*Statist. Sci.*]{} [**5**]{} no. 4, 403–434. and [Rinott, Y. (1989).]{} On normal approximations of distributions in terms of dependency graphs. [*Ann. Probab.*]{} [**17**]{} no. 4, 1646–1650. Stein’s method for diffusion approximations. [*Probab. Theory Related Fields*]{} [**84**]{} no. 3, 297–322. and [Janson, S. (1992).]{} [*Poisson approximation.*]{} Oxford Studies in Probability, [**2**]{} The Clarendon Press, Oxford University Press, New York. Statistical analysis of non-lattice data. [*The Statistician,*]{} [**24**]{} 179–195. An estimate of the remainder in a combinatorial central limit theorem. [*Z. Wahrsch. Verw. Gebiete*]{} [**66**]{} no. 3, 379–386. and [Götze, F. (1993).]{} The rate of convergence for multivariate sampling statistics. [*Ann. Statist.*]{} [**21**]{} no. 4, 1692–1710. and [Massart, P. (2003).]{} Concentration inequalities using the entropy method. [*Ann. Probab.* ]{} [**31**]{} No. 3, 1583–1614. Distribution function inequalities for martingales. [*Ann. Probab.*]{} [**1**]{} 19–42. Ph.D. thesis. Department of Statistics, Stanford University.\ Available at [http://arxiv.org/math.PR/0507526]{} Concentration of Haar measures, with an application to random matrices. [*Submitted.*]{} Available at [http://arxiv.org/math.PR/0508518]{} and [Meckes, E. (2005).]{} Exchangeable pairs and Poisson approximation. [*Probab. Surv.*]{} [**2**]{} 64–106. and [Redig, F. (2006).]{} Concentration inequalities for random fields via coupling. [*Submitted.*]{} Available at [http://arxiv.org/math.PR/0503483]{} Poisson approximation for dependent trials. [*Ann. Probab.*]{} [**3**]{} No. 3, 534–545. and [Graham, R. L. (1977).]{} Spearman’s footrule as a measure of disarray. [*J. Roy. Statist. Soc. Ser. B*]{} [**39**]{} No. 2, 262–268. and [Holmes, S. (editors) (2004).]{} [*Stein’s method: expository lectures and applications.*]{} IMS Lecture Notes—Monograph Series, [**46**]{}. Grund. der Mathemat. Wissenschaften, [**271**]{}. [*Springer-Verlag, New York.*]{} Third edition. John Wiley & Sons, Inc., New York-London-Sydney. Stein’s method and non-reversible Markov chains. [*Stein’s method: expository lectures and applications,*]{} 69–77, IMS Lecture Notes Monogr. Ser., [**46**]{}, IMS, Beachwood, OH. Stein’s method and Plancherel measure of the symmetric group. [*Trans. Amer. Math. Soc.*]{} [**357**]{} no. 2, 555–570 (electronic). and [Thompson, E. A. (1992).]{} Constrained Monte Carlo maximum likelihood for dependent data. [*J. Roy. Statist. Soc. Ser. B*]{} [**54**]{} no. 3, 657–699. and [Reinert, G. (1997).]{} Stein’s method and the zero bias transformation with application to simple random sampling. [*Ann. Appl. Probab.*]{} [**7**]{} no. 4, 935–952. A combinatorial central limit theorem. [*Ann. Math. Statist.* ]{} [**22** ]{} no. 4, 558–566. Ê Probability inequalities for sums of bounded random variables. Ê[*J. Amer. Stat. Assoc.*]{} [**58** ]{} 13–30. and [Sinclair, A. (1993).]{} Polynomial-time approximation algorithms for the Ising model. [*SIAM J. Comput.*]{} [**22**]{} no. 5, 1087–1116. and [Vu, V. H. (2004).]{} Divide and conquer martingales and the number of triangles in a random graph. [*Random Struct. Algorithms*]{} [**24**]{} no. 2, 166–174. Amer. Math. Soc., Providence, RI. and [McDiarmid, C. (2003).]{} Concentration for locally acting permutations. [*Discrete Math.*]{} [**265**]{} no. 1–3, 159–171. Construction de suites symétriques. [*C. R. Acad. Sci. Paris Sér. A-B*]{} [**288**]{} no. 14, A679–A681. Concentration for independent permutations. [*Combin. Probab. Comput.*]{} [**11**]{} no. 2, 163–178. Inference for discrete Markov fields: the simplest nontrivial case. [*J. Amer. Statist. Assoc.*]{} [**82**]{} no. 397, 90–96. CLT-related large deviation bounds based on Stein’s method. [*Preprint.*]{} and [Rotar, V.]{} (1997). On coupling constructions and rates in the CLT for dependent summands with applications to the antivoter model and weighted $U$-statistics. [*Ann. Appl. Probab.*]{} [**7**]{} no. 4, 1080–1105. and [Su, Z. (2004).]{} The Berry-Esseen bound for character ratios. [*Preprint.*]{} Curvature of nonlocal Markov generators. In [*Convex Geometric Analysis: MSRI Publications.*]{} [**34** ]{} 189–197. and [Wellner, J. A. (1986).]{} [*Empirical processes with applications to statistics.*]{} John Wiley & Sons, Inc., New York. A bound for the error in the normal approximation to the distribution of a sum of dependent random variables. [*Proc. of the Sixth Berkeley Symp. on Math. Statist. and Probab., Vol. II: Probability theory.* ]{} 583–602. IMS Lecture Notes—Monograph Series, [**7**]{}. Concentration of measure and isoperimetric inequalities in product spaces. [*Inst. Hautes Études Sci. Publ. Math.*]{} [**81**]{} 73–205.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present results from 16 snapshots of Aql X-1 with the [*Rossi X-ray Timing Explorer*]{}during the rising phase of its recent outburst. The observations were carried out at a typical rate of once or twice per day. The source shows interesting spectral evolution during this period. Phenomenologically, it bears remarkable similarities to “atoll” sources. Shortly after the onset of the outburst, the source is seen to be in an “island” state, but with little X-ray variability. It then appears to have made a rapid spectral transition (on a time scale less than half a day) to another “island” state, where it evolves slightly and stays for 4 days. In this state, the observed X-ray flux becomes increasingly variable as the source brightens. Quasi-period oscillation (QPO) in the X-ray intensity is detected in the frequency range 670–870 Hz. The QPO frequency increases with the X-ray flux while its fractional rms decreases. The QPO becomes undetectable following a transition to a “banana” state, where the source continues its evolution by moving up and down the “banana” branch in the color-color diagram as the flux (presumably, the mass accretion rate) fluctuates around the peak of the outburst. Throughout the entire period, the power density spectrum is dominated by very-low frequency noises. Little power can be seen above $\sim$1 Hz, which is different from typical “atoll” sources. In the “banana” state, the overall X-ray variability remains low (with fractional rms $\sim$3–4%) but roughly constant. The observed X-ray spectrum is soft with few photons from above $\sim$25 keV, implying the thermal origin of the emission. The evolution of both spectral and temporal X-ray properties is discussed in the context of disk-instability models.' author: - 'Wei Cui, Didier Barret, S. N. Zhang, Wan Chen, Laurence Boirin, and Jean Swank' title: 'Evolution of AQL X-1 During the Rising Phase of its 1998 Outburst' --- Introduction ============ Soft X-ray transients (SXTs) constitute an important subclass of low-mass X-ray binaries. For most of the time, they appear as extremely faint X-ray sources, if detected at all, but occasionally they brighten up by orders of magnitude in X-ray intensity, becoming the brightest X-ray sources in the sky in some cases. There appears to be some consensus that thermal instability causes a sudden surge in the mass accretion rate through the disk and, thus, initiates an X-ray outburst (review by King 1995 and references therein). Archival databases have been established over the years for the study of SXTs in outbursts and have been continually enriched by the on-going missions. Few observations, however, have been made of SXTs during the rising phase of an outburst because of the difficulty in catching such a brief period (typically lasting for a few days; cf., Chen, Shrader, & Livio 1997). The situation has greatly improved with the launch of [*Rossi X-ray Timing Explorer*]{} (RXTE; Bradt, Rothschild, & Swank 1993). [*RXTE*]{} carries an all-sky monitor (ASM) that continuously scans the sky for transient events. Upon detection, the main instruments can be re-programmed to observe the event within hours. To take advantage of these unprecedented capabilities, we have established a comprehensive program to sample the rising phase of an outburst at a high rate, in order to facilitate systematic studies of the phenomenon. Luckily, we have already caught several outbursts from different sources since the beginning of this year. In this Letter, we present results from observations of Aql X-1 during the rising phase of its recent outburst (Swank et al. 1998). Aql X-1 belongs to a minority group of SXTs that have been determined to contain neutron stars (as opposed to black holes) because they display Type I X-ray bursts. It is known to experience frequent outbursts (roughly once every year; see, e.g., Priedhorsky & Terrell 1984; also public ASM/RXTE light curves). The [*RXTE*]{} observations of Aql X-1 during a previous outburst have revealed the presence of quasi-periodic oscillations (QPOs) near kHz range in the X-ray light curves (Zhang et al. 1997), which seem to be common for X-ray binaries with a weakly magnetized neutron star (review by van der Klis 1997). The kHz QPO phenomenon is currently thought to be associated with the accretion disk, perhaps representing processes varying on dynamical time scales in the inner portion of the disk. There is suggestive evidence that the magnetic field might play a crucial role in actually producing the observed X-ray modulation (see, e.g., Zhang, Yu, & Zhang 1998). One of the neutron star SXTs, 4U 1608-52, is further classified as an “atoll” source, whose neutron star is thought to have even weaker magnetic field than a “Z” source (van der Klis 1995). The classification of Aql X-1 in this scheme is not certain. Near the end of February 1998, the ASM detected the onset of an outburst from Aql X-1 (Swank et al. 1998). Fig. 1 shows a portion of the long-term ASM light curve for the source that highlights the event. The outburst peaked at roughly 34 c/s (or $\sim$450 mCrab in the 1.3-12 keV band). The initial rise lasted for about 12 days. The pending Target-of-Opportunity proposals of ours were triggered by the ASM alert, subsequently, a series of pointed [*RXTE*]{} observations were carried out. Observations ============ The data used for this study come from a total of 16 snapshots of Aql X-1 with the [*Proportional Counter Array*]{} aboard [*RXTE*]{}, covering a major portion of the rising phase and the peak of the outburst. The times when these observations were made are indicated in Fig. 1. The first observation took place when the ASM flux was about one quarter of the peak value. Each pointed [*RXTE*]{} observation lasted for one satellite orbit with an effective exposure time in the range 1.2–3.6 ks. The observations were conducted at a typical rate of once or twice per day, with occasional gaps due to scheduling constraints. Besides the two [*Standard*]{} data modes, a combination of [*Burst Trigger*]{} and [*Burst Catcher*]{} modes were used to zoom in onto any Type I bursts detected; but, no X-ray bursts were seen. An [*Event*]{} mode with $\sim 122\mu s$ timing resolution and 64 energy bands was adopted to facilitate high-resolution timing analysis with a moderate energy resolution. Data Analysis and Results ========================= We have carried out preliminary spectral analysis, using the [*Standard 2*]{} data. Throughout the entire period, the observed X-ray spectrum can be modeled quite well by a composite of a blackbody (BB) component (for emission from the neutron star surface), a multi-color disk (MCD) component (for emission from the accretion disk), and a thermal bremsstrahlung component (for emission from hot plasma perhaps present in the vicinity of the neutron star); an emission line at $\sim$6.6 keV is also required. Fig. 2 shows the evolution of the BB and MCD components. The spectrum remains soft, with few photons above $\sim$25 keV, implying the thermal origin of the emission. The fluxes have been computed using the best-fit parameters, with the H I column density fixed at $5\times 10^{21}\mbox{ }cm^{-2}$ (Christian & Swank 1997) during the fit. The discussion of physical models for such a spectrum is beyond the scope of this work; it will be presented in a future paper on detailed spectral modeling. For each observation, we have made light curves (with background subtracted) from the [*Standard 2*]{} data (with 16-second timing resolution) in three energy bands: 2–5.2 keV, 5.2–8.9 keV, and 8.9–19.7 keV, from which two hardness ratios, 5.2–8.9 keV/2–5.2 keV (soft color) and 8.9–19.7 keV/5.2–8.9 keV (hard color), have been calculated. The results are summarized in Fig. 3a, in the form of a color-color diagram for each observation. Fig. 3b shows the overall spectral evolution by putting together all the results. There are two apparent spectral transitions between three disjointed branches. Initially, the source is on the upper of the two left branches. It then seems to have made a rapid spectral transition to the lower one (on a time scale less than half a day; see Fig. 1). As the X-ray flux increases, the source evolves slightly, moving from the upper-right side to the lower-left side, and stays on this branch for 4 days before making another transition to the right branch. Subsequently, the source continues its evolution by moving up and down the branch from left to right as the source flux fluctuates around the peak of the outburst. Using the [*Event*]{} mode data, we have carried out FFT for every 128-second data segment of each observation. Individual power-density spectra (PDSs) are then properly weighted and co-added to obtain the average PDS for the observation. A kHz QPO is detected only when the source is on the left lower branch of the color-color diagram (see Fig. 4 for an example). The QPO profile can be modeled well by a simple Lorentzian function, and the best-fit parameters are shown in Table 1. The errors are estimated based on $\Delta \chi^2 = 1$. From the table, we can see that the QPO frequency varies in the range 670–870 Hz and width (FWHM) in the range 9–20 Hz. As the flux increases, the QPO frequency also increases, and appears to have reached a plateau at $\sim$850 Hz (following Observation 3; see Table 1); the fractional rms decreases. We have also performed FFT for different energy bands to investigate the energy dependence of the QPO. The results show that the fractional rms increases almost linearly with energy, up to $\sim$20% in the 13.1–19.7 keV band. The QPO becomes unmeasurable following the second transition. Except for the first observation, the power-density continuum is dominated by low-frequency “1/f” noises (with the power-law slope varying in the range -0.8– -1.5); very little power is measured above $\sim$1 Hz (see, e.g., Fig. 4). The source shows very little variability when it is on the left upper branch (i.e., Observation 1). It becomes more variable, following the first transition. The measured fractional rms (for the continuum) increases from $\sim$1.5% and to $\sim$6.4%. The variability seems to drop slightly, following the second transition to the right branch, with the fractional rms staying at $\sim$3–4% subsequently. Discussion ========== Is Aql X-1 an “Atoll” Source? ----------------------------- The phenomenology of the observed spectral evolution is very similar to that of “atoll” sources (van der Klis 1995, and references therein). In the color-color diagram (Fig. 3b), the left branches may represent so-called “island” states, while the curved right branch certainly shows the characteristics of a “banana” state. As the observed X-ray flux (presumably, the mass accretion rate) increases, the source moves from one “island” to another, then to the “banana” branch. Fig. 3a shows that the correlation between spectral states and mass accretion rate holds remarkably well in detail. For instance, as soon as the X-ray flux begins to drop in observations 15 and 16, the source reverses its motion along the branch and moves from right to left down the “banana”. Perhaps, Aql X-1 should be classified as an “atoll” source, like its cousin 4U 1608-52. However, Aql X-1 also differs from 4U 1608-52 or other typical “atoll” sources in that the high-frequency noise component (van der Klis 1995) is absent from the measured PDS for both the “island” states and “banana” state. For “atoll” sources, the physical processes that trigger the transition between “island” states themselves or between an “island” state and a “banana” state are still not well understood. In the following, we discuss a physical scenario that can account for the observed spectral and temporal evolution of Aql X-1 during the rising phase of the outburst, based on our current knowledge of SXTs and their outburst mechanisms. Transition between “Island” States ---------------------------------- It is thought that for SXTs in quiescence the mass accretion process is likely to operate in the form of “advection-dominated” accretion flows (ADAFs; Narayan & Yi 1994, 1995; Narayan 1997). In this regime, the compact object is surrounded by a hot ADAF region that is sub-Keplerian and experiences a phase transition to the standard thin disk (Shakura & Sunyaev 1973) thousands of Schwarzschild radii away (Narayan 1997). In the context of disk instability models, when thermal instability sets in, the mass accretion rate surges, thus, an outburst is under way. In the process, the ADAF region is cooled efficiently via inverse Compton scattering processes, because of the increase of “seed” photons from the disk. Consequently, the region shrinks; at the same time, the thin disk in the outer region fills in. Such a transition is likely to continue until the inner edge of the disk reaches the last (marginally) stable orbit (Narayan 1997). Although many details in the ADAF model, such as the location of the transition zone between the ADAF and the thin disk, still need to be worked out, the general evolution sequence has been shown to be followed well (see, e.g., Esin, McClintock, & Narayan 1997). For SXTs that contain a neutron star, the effects of magnetic field must be taken into account. Unfortunately, neither do we know the precise configuration of the magnetic field in the presence of an accretion disk, nor do we fully understand the disk-field interaction (see, e.g., Ghosh & Lamb 1979 and Lovelace, Romanova, & Bisnovatyi-Kogan 1995). Assuming a simple dipole field, the magnetospheric radius is given by (see, e.g., Lamb, Pethick, & Pines 1973 and Cui 1997) $$r_m = 10^7 \xi L^{-2/7}_{x,36} M^{1/7}_{1.4} B^{4/7}_{9} R^{10/7}_6 cm$$ where $\xi$ is a model-dependent and dimensionless number, ranging from 0.52 (Ghosh & Lamb 1979) to $\sim$1 (Arons 1993; Ostriker & Shu 1995; Wang 1996); $L_{x,36}$ is the bolometric X-ray luminosity in units of $10^{36}\mbox{ } ergs/s$; $B_9$ is the dipole field strength at the poles in units of $10^9$ G; $M_{1.4}$ is the mass of the neutron star in units of $1.4 M_{\odot}$, and $R_6$ is its radius in units of 10 km. At the beginning of an outburst, the mass accretion rate is low, so the magnetosphere initially extends much beyond the last stable orbit. As the inner edge of the disk subsequently moves in toward the neutron star, it is bound to encounter the magnetosphere first. We argue that it is the onset of this disk-field engagement that might have caused the transition between the two “island” states. Perhaps, observations 1 and 2 were made just before and after the engagement. For Observation 2, $L_{x,36}=5.7$ (for a source distance of 2.5 kpc, after being corrected for absorption), so $r_m = 61 \xi B_9^{4/7}\mbox{ }km$. For $r_m > 3R_s$ (the radius of the last stable orbit for a slowly rotating neutron star), where $R_s$ is the Schwarzschild radius, $B_9 > 0.062$ (or $0.20$) for $\xi=1$ (or $0.52$). As the mass accretion rate increases, the BB temperatures increases, as shown in Fig. 2. At the same time, the magnetosphere is pushed back further, so the disk extends closer to the neutron star and becomes hotter (also Fig. 2). The intensified emission, both from the neutron star surface and accretion disk (mostly in the 2–5.2 keV band, and also in the 5.2–8.9 keV band), results in the decrease of hardness ratios as the source evolves in the second “island” state. Transition between “Island” and “Banana” States ----------------------------------------------- The evolution continues until the mass accretion rate becomes so large that the magnetosphere is entirely inside the last stable orbit (as indicated by the constancy of the inferred radius of the inner edge of the disk following Observation 6; see Fig. 2). Then, the disk is once again disengaged from the magnetic field. This might have triggered the transition from the lower “island” state to the “banana” state, as indicated by the disappearance of the kHz QPO. Then, an [*upper*]{} limit could be derived for the magnetic field from equation (1). The transition is clearly under way during Observation 6. For this observation, $L_{x,36}=10.7$, so $r_m \lesssim 3R_s$ leads to $B_9 \lesssim 0.085$ (or $0.27$) for $\xi=1$ (or $0.52$). In the “banana” state, the observed spectral evolution implies that the soft fluxes (both from the disk and the neutron star surface) shift more toward the the middle energy band (5.2-8.9 keV), as the mass accretion rate increases, and the hard flux increases even more. Summary ------- We have proposed a simple explanation for the observed spectral and temporal evolution of Aql X-1 during the rising phase of the outburst. It remains to be seen whether or not the model also applies to “atoll” sources in general. Progress can be made by carefully studying “atoll” sources in a similar way (i.e., correlating the occurrence of kHz QPOs with distinct spectral states). The model is almost certainly over-simplified, but seems to be qualitatively valid. It has been speculated that kHz QPOs might originate in the inhomogeneity (or “hot spots”) at the inner edge of the disk and the QPO frequency might simply be the local Keplerian frequency or the beat frequency between the Keplerian motion and the spin of the neutron star (see van der Klis 1997 for a review). If so, the occurrence and disappearance of the kHz QPO for Aql X-1 would strongly suggest that the magnetic field plays an essential role in producing such “hot spots” by interacting with the disk. With limited statistics (Table 1), the kHz QPO frequency does seem to level off as the mass accretion rate continues to increase, perhaps indicating that the inner edge of the disk has reached (or approached close to) the last stable orbit as early as Observation 3. This is not inconsistent with the spectral results (see Fig. 2). It is instructive to compare this outburst with previous ones. A spectral transition was also observed near the end of the decaying phase of the first 1997 outburst (Zhang, Yu, & Zhang 1998). It appears to be similar to the first transition observed here, although it started at a lower flux level. Interestingly enough, the kHz QPO was also detected at lower fluxes (Zhang et al. 1998). Although it is true that the peak intensity of the current outburst is higher than the previous one, it is not clear how this would affect the fundamental properties of the system. It is possible, however, that the disk structure might be different between the rising and decaying phases, which could result in hysteresis. More studies are clearly required to address this issue. This work is supported partially by NASA through Contract NAS5-30612. We would like to thank Alan Harmon and Jean Francois Olive for helpful discussions. Arons, J. 1993, , 408, 160 Bradt, H. V., Rothschild, R. E., & Swank, J. H. 1993, , 97, 355 Chen, W., Shrader, C. R., & Livio, M. 1997, , 491, 312 Christian, D. J., & Swank, J. H. 1997, , 109, 177 Cui, W. 1997, , 482, L163 Esin, A. A., McClintock, J. E., & Narayan, R. 1997, , 489,865 Ghosh, P., & Lamb, F. K. 1979, , 234, 296 King, A. R. 1995, in “X-ray Binaries”, eds. W. H. G. Lewin, J. van Paradijs, & E. P. J. van den Heuvel (Cambridge U. Press, Cambridge) p. 419 Lamb, F. K., Pethick, C. J., and Pines, D. 1973, , 184, 271 Lovelace, R. V. E., Romanova, M. M., & Bisnovatyi-Kogan, G. S. 1995, , 275, 244 Narayan, R., & Yi, I. 1994, , 428, L13 Narayan, R., & Yi, I. 1995, , 444, 231 Narayan, R. 1997, in “Accretion Phenomena and Related Outflows”, eds D. T. Wickramasinghe, G. V. Bicknell, & L. Ferrario, ASP Conf. Series Vol. 121 Ostriker, J., & Shu, F. 1995, , 477, 813 Priedhorsky, W. C., & Terrell, J. 1984, , 280, 661 Shakura, N. I., & Sunyaev, R. A. 1973, , 24, 337 Swank, J. H., Smith, E., Levine, A. M., & Remillard, R. 1998,  6828 van der Klis, M. 1995, in “X-ray Binaries”, eds. W. H. G. Lewin, J. van Paradijs, & E. P. J. van den Heuvel (Cambridge U. Press, Cambridge) p. 252 van der Klis, M. 1997, to appear in Proc. NATO Advanced Institue “The Many Faces of Neutron Stars,” Lipari, Italy, 1996 Wang, Y.-M. 1996, , 465, L111 Zhang, S. N. Yu, W. & Zhang, W. 1998, , 494, L71 Zhang, W., Jahoda, K., Kelley, R. L., Strohmayer, T. E., Swank, J. H., & Zhang, S. N. 1998, , 495, L9 [lccc]{} 2 & $677.0^{+0.4}_{-0.3}$ & $9.1^{+0.7}_{-0.9}$ & $6.3^{+0.2}_{-0.2}$\ 3 & $856.3^{+1.1}_{-1.0}$ & $19.7^{+2.5}_{-2.0}$ & $5.7^{+0.3}_{-0.2}$\ 4 & $839.5^{+0.5}_{-0.4}$ & $11.0^{+1.1}_{-0.9}$ & $5.7^{+0.2}_{-0.2}$\ 5 & $871.4^{+1.0}_{-1.0}$ & $11.8^{+3.0}_{-2.4}$ & $3.3^{+0.3}_{-0.3}$\
{ "pile_set_name": "ArXiv" }
--- abstract: 'Recently, a number of nonlocal integrable equations, such as the -symmetric nonlinear Schrödinger (NLS) equation and -symmetric Davey-Stewartson equations, were proposed and studied. Here we show that many of such nonlocal integrable equations can be converted to local integrable equations through simple variable transformations. Examples include these nonlocal NLS and Davey-Stewartson equations, a nonlocal derivative NLS equation, the reverse space-time complex modified Korteweg-de Vries (CMKdV) equation, and many others. These transformations not only establish immediately the integrability of these nonlocal equations, but also allow us to construct their analytical solutions from solutions of the local equations. These transformations can also be used to derive new nonlocal integrable equations. As applications of these transformations, we use them to derive rogue wave solutions for the partially -symmetric Davey-Stewartson equations and the nonlocal derivative NLS equation. In addition, we use them to derive multi-soliton and quasi-periodic solutions in the reverse space-time CMKdV equation. Furthermore, we use them to construct many new nonlocal integrable equations such as nonlocal short pulse equations, nonlocal nonlinear diffusion equations, and nonlocal Sasa-Satsuma equations.' address: | [*$^1$Shanghai Key Laboratory of Trustworthy Computing, East China Normal University, Shanghai 200062, China*]{}\ [*$^2$Department of Mathematics and Statistics, University of Vermont, Burlington, VT 05401, U.S.A*]{}\ [ *Corresponding author, email address: [email protected]*]{} author: - 'Bo Yang$^{1,2}$ and Jianke Yang$^{2*}$' title: Transformations between nonlocal and local integrable equations --- Introduction ============ The study of integrable nonlinear wave equations has a long history [@Ablowitz1981; @Zakharov1984; @Faddeev1987; @Ablowitz1991; @Yang2010]. Most of those integrable equations are local equations, i.e., the solution’s evolution depends only on the local solution value and its local space and time derivatives. The Korteweg-de Vries equation and the nonlinear Schrödinger (NLS) equation are such examples. Recently, a number of new nonlocal integrable equations were proposed and studied [@AblowitzMussPRL2013; @AblowitzMussPRE2014; @Yan; @Khara2015; @AblowitzMussNonli2016; @Zhu1; @Fokas2016; @Lou; @Lou2; @AblowitzMussSAPM; @Chow; @ZhoudNLS; @ZhouDS; @HePTDS; @HePPTDS; @Zhu2; @Zhu3; @Gerdjikov2017; @Ablowitz_arxiv]. The first such nonlocal equation was the -symmetric NLS equation [@AblowitzMussPRL2013] $$\label{e:PTNLS} iu_t(x,t)+u_{xx}(x,t)+2\sigma u^2(x,t)u^*(-x,t)=0,$$ where $\sigma=\pm 1$ is the sign of nonlinearity (with the plus sign being the focusing case and minus sign the defocusing case), and the asterisk \* represents complex conjugation. Notice that here, the solution’s evolution at location $x$ depends on not only the local solution at $x$, but also the nonlocal solution at the distant position $-x$. That is, solution states at distant locations $x$ and $-x$ are directly coupled, reminiscent of quantum entanglement between pairs of particles. Eq. (\[e:PTNLS\]) was called -symmetric because the nonlinearity-induced potential $V\equiv \sigma u(x,t)u^*(-x, t)$ satisfies the symmetry condition $V^*(x,t)=V(-x,t)$. In addition, this equation is invariant under the action of the operator, i.e., the joint transformations $x\to -x$, $t\to -t$ and complex conjugation \[hence if $u(x,t)$ is a solution, so is $u^*(-x,-t)$)\]. It is noted that symmetric systems, which behave like conservative systems despite gain and loss, attracted a lot of attention in optics and other physical fields in recent years [@Kivsharreview; @Yangreview]. The application of this -symmetric NLS equation for an unconventional system of magnetics was reported in [@PTNLSmagnetics]. Following this nonlocal -symmetric NLS equation, other new nonlocal integrable equations were quickly reported. Examples include the fully -symmetric and partially -symmetric Davey-Stewartson (DS) equations [@Fokas2016; @AblowitzMussSAPM], the nonlocal derivative NLS equation [@ZhoudNLS], the reverse space-time complex modified Korteweg-de Vries (CMKdV) equation [@AblowitzMussNonli2016; @Zhu3], the reverse time NLS equation [@AblowitzMussSAPM], the reverse space-time NLS equation [@Lou; @AblowitzMussSAPM], and many others. These nonlocal equations are distinctly different from local equations for their novel space and/or time coupling, which could induce new physical effects and thus inspire novel physical applications. Indeed, solution properties in some of these nonlocal equations have been analyzed by the inverse scattering transform method, Darboux transformation or the Hirota bilinear method, and interesting behaviors such as finite-time solution blowup [@AblowitzMussPRL2013] and the simultaneous existence of soliton and kink solutions [@Zhu2] have been revealed. In this article, we report that many of these nonlocal integrable equations can be converted to their local integrable counterparts through simple variable transformations. Such nonlocal equations include the -symmetric NLS and DS equations, the nonlocal derivative NLS equation, the reverse space-time CMKdV equation, and many others. This conversion puts these nonlocal equations in a totally different light and opens up a totally new way to study their solution behaviors. First of all, this conversion immediately establishes the integrability of these nonlocal equations. Secondly, it allows us to construct analytical solutions of these nonlocal equations from solutions of their local counterparts. Thirdly, it can be used to derive new nonlocal integrable equations. As applications of these transformations, we use them to derive rogue wave solutions for the partially -symmetric DS equations and the nonlocal derivative NLS equation. In addition, we use them to derive multi-soliton and quasi-periodic solutions in the reverse space-time CMKdV equation. Furthermore, we use them to construct many new nonlocal integrable equations such as nonlocal short pulse equations, nonlocal nonlinear diffusion equations, nonlocal Sasa-Satsuma equations and nonlocal Chen-Lee-Liu equations. Transformations between nonlocal and local integrable equations =============================================================== In this section, we present transformations which convert many nonlocal integrable equations to their local counterparts. Our first example is the -symmetric NLS equation (\[e:PTNLS\]). Under the variable transformations $$\label{transNLS} x=i\hat{x}, \hspace{0.2cm} t=-\hat{t}, \hspace{0.2cm} u(x,t)=\hat{u}(\hat{x},\hat{t}),$$ this nonlocal equation becomes $$\label{e:PTNLSL} i\hat{u}_{\hat{t}}(\hat{x},\hat{t})+\hat{u}_{\hat{x}\hat{x}}(\hat{x},\hat{t})-2\sigma \hat{u}^2(\hat{x},\hat{t})\hat{u}^*(\hat{x},\hat{t})=0,$$ which is the local NLS equation but with the opposite sign of nonlinearity. In other words, the -symmetric focusing NLS equation is converted to the local defocusing NLS equation, and the -symmetric defocusing NLS equation is converted to the local focusing NLS equation. The key reason for this nonlocal to local conversion is that, in the nonlocal equation (\[e:PTNLS\]), $x$ is treated real when taking the complex conjugate $u^*(-x,t)$. But under the $x=i\hat{x}$ transformation with real $\hat{x}$, $x$ becomes imaginary. In this case, when taking the complex conjugate of $u(-x,t)$, the sign of $x$ flips, turning the nonlocal term $u^*(-x,t)$ in (\[e:PTNLS\]) to the local term $\hat{u}^*(\hat{x},\hat{t})$ in (\[e:PTNLSL\]). Following similar ideas, we can transform many more nonlocal integrable equations to their local counterparts. Some examples are listed below. 1. Consider the -symmetric DS equations [@Fokas2016; @AblowitzMussSAPM] $$\begin{aligned} \label{PTDS} && \hspace{-1.4cm} iA_t(x,y,t) = A_{xx}(x,y,t)+\sigma^2 A_{yy}(x,y,t) +\left[\epsilon A(x,y,t)A^*(-x,-y,t)-2Q(x,y,t)\right]A(x,y,t), \nonumber \\ && \hspace{-1.4cm} Q_{xx}(x,y,t)-\sigma^2Q_{yy}(x,y,t)=\epsilon \left[A(x,y,t)A^*(-x,-y,t)\right]_{xx},\end{aligned}$$ where $\sigma^2=\pm 1$ is the equation-type parameter (with $\sigma^2=1$ being DSI and $\sigma^2=-1$ DSII), and $\epsilon=\pm 1$ is the sign of nonlinearity. Under the variable transformations $$x=i\hat{x}, \hspace{0.2cm} y=i\hat{y}, \hspace{0.2cm} t=-\hat{t}, \hspace{0.2cm} A(x,y,t)=\widehat{A}(\hat{x},\hat{y},\hat{t}), \hspace{0.2cm} Q(x,y,t)=-\widehat{Q}(\hat{x},\hat{y},\hat{t}),$$ and dropping the bars, i.e., with $$x\to ix, \hspace{0.2cm} y\to iy, \hspace{0.2cm} t\to -t, \hspace{0.2cm} Q\to -Q,$$ these -symmetric DS equations are converted to the following local (classical) DS equations $$\begin{aligned} && \hspace{-1.4cm} iA_{t}(x,y,t)=A_{xx}(x,y,t)+\sigma^2 A_{yy}(x,y,t) +\left[-\epsilon A(x,y,t)A^*(x,y,t)-2Q(x,y,t)\right] A(x,y,t), \nonumber \\ && \hspace{-1.4cm} Q_{xx}(x,y,t)-\sigma^2Q_{yy}(x,y,t)=-\epsilon \left[A(x,y,t)A^*(x,y,t)\right]_{xx}.\end{aligned}$$ Similar to the -symmetric NLS equation above, the sign of nonlinearity $\epsilon$ has switched after the nonlocal-to-local conversion, but the sign of $\sigma^2$ remains the same. Thus, the -symmetric focusing DSI equations are converted to local defocusing DSI equations, and so on. 2. Consider the partially -symmetric DS equations [@Fokas2016; @AblowitzMussSAPM] $$\begin{aligned} \label{PPTDS} && \hspace{-1.4cm} iA_t(x,y,t)=A_{xx}(x,y,t)+\sigma^2 A_{yy}(x,y,t) +\left[\epsilon A(x,y,t)A^*(-x,y,t)-2Q(x,y,t)\right]A(x,y,t), \nonumber \\ && \hspace{-1.4cm} Q_{xx}(x,y,t)-\sigma^2Q_{yy}(x,y,t)=\epsilon \left[A(x,y,t)A^*(-x,y,t)\right]_{xx},\end{aligned}$$ where $\sigma^2=\pm 1$ and $\epsilon=\pm 1$. Under the variable transformations $$\label{transPPTDS} x\to ix, \hspace{0.2cm} t\to -t, \hspace{0.2cm} Q\to -Q,$$ these nonlocal DS equations reduce to the local DS equations $$\begin{aligned} \label{PPTDSL} && \hspace{-1.4cm} iA_{t}(x,y,t)=A_{xx}(x,y,t)-\sigma^2 A_{yy}(x,y,t) +\left[-\epsilon A(x,y,t)A^*(x,y,t)-2Q(x,y,t)\right] A(x,y,t), \nonumber \\ && \hspace{-1.4cm} Q_{xx}(x,y,t)+\sigma^2Q_{yy}(x,y,t)=-\epsilon \left[A(x,y,t)A^*(x,y,t)\right]_{xx}.\end{aligned}$$ Here, after the nonlocal-to-local conversion, not only the sign of nonlinearity $\epsilon$, but also the equation-type parameter $\sigma^2$, switches. Thus, the partially -symmetric focusing DSI equations are converted to local defocusing DSII equations, and so on. The equations (\[PPTDS\]) are partially -symmetric in $x$. Similar equations partially -symmetric in $y$ can also be converted to their local counterparts through the sole transformation $y\to iy$. Under this conversion, the sign of $\sigma^2$ switches, but not the sign of $\epsilon$. 3. Consider the nonlocal derivative NLS equation [@ZhoudNLS] $$\label{e:PTDNLS} iu_t(x,t)+u_{xx}(x,t)+\sigma \left[u^2(x,t)u^*(-x,t)\right]_x=0,$$ where $\sigma=\pm 1$. Under the variable transformations $$\label{transDNLS} x\to ix, \hspace{0.15cm} t\to -t,$$ this nonlocal equation becomes the local derivative NLS equation [@KaupNewell] $$\label{e:PTDNLSL} iu_t(x,t)+u_{xx}(x,t)+i\sigma \left[u^2(x,t)u^*(x,t)\right]_x=0.$$ 4. Consider the reverse space-time CMKdV equation [@AblowitzMussNonli2016; @Zhu3] $$\label{e:NCMKdV} q_{t}(x,t)+q_{xxx}(x,t)+6\sigma q(x,t)q^{*}(-x,-t)q_{x}(x,t)=0,$$ where $\sigma=\pm 1$. Under the variable transformations $$\label{transCMKdV} x\rightarrow ix,\ t\rightarrow -it,$$ this equation become the following local (classical) CMKdV equation $$\label{e:CMKdV} q_{t}(x,t)+q_{xxx}(x,t)-6\sigma q(x,t)q^{*}(x,t)q_{x}(x,t)=0.$$ Note that the sign of nonlinearity has flipped under this conversion. 5. Consider the multidimensional reverse space-time nonlocal three wave interaction equations [@AblowitzMussSAPM] $$\begin{aligned} && \hspace{-1.2cm} Q_{1,t}(\x,t)+\mathbf{C}_1\cdot \nabla Q_1(\x,t)=\sigma_1 Q_2^*(-\x,-t)Q_3^*(-\x,-t), \nonumber \\ && \hspace{-1.2cm} Q_{2,t}(\x,t)+\mathbf{C}_2\cdot \nabla Q_2(\x,t)=\sigma_2 Q_1^*(-\x,-t)Q_3^*(-\x,-t), \\ && \hspace{-1.2cm} Q_{3,t}(\x,t)+\mathbf{C}_3\cdot \nabla Q_3(\x,t)=\sigma_3 Q_1^*(-\x,-t)Q_2^*(-\x,-t), \nonumber\end{aligned}$$ where $\sigma_j=\pm 1$, $j=1, 2, 3$, $\sigma_1\sigma_2\sigma_3=-1$, $\x$ is a multidimensional spacial variable, and $\mathbf{C}_1$, $\mathbf{C}_2$, $\mathbf{C}_3$ are constant vectors. Under the variable transformations $$\x\to i\hspace{0.04cm} \x, \quad t\to i \hspace{0.04cm} t, \nonumber$$ the above nonlocal three wave interaction equations reduce to the local counterparts [@Ablowitz1981] $$\begin{aligned} && \hspace{-0.8cm} Q_{1,t}(\x,t)+\mathbf{C}_1\cdot \nabla Q_1(\x,t)=i\sigma_1 Q_2^*(\x,t)Q_3^*(\x,t), \nonumber \\ && \hspace{-0.8cm} Q_{2,t}(\x,t)+\mathbf{C}_2\cdot \nabla Q_2(\x,t)=i\sigma_2 Q_1^*(\x,t)Q_3^*(\x,t), \\ && \hspace{-0.8cm} Q_{3,t}(\x,t)+\mathbf{C}_3\cdot \nabla Q_3(\x,t)=i\sigma_3 Q_1^*(\x,t)Q_2^*(\x,t). \nonumber\end{aligned}$$ In addition to the above nonlocal integrable equations, many others, such as the vector or matrix extensions of the -symmetric NLS equations [@Yan; @Zhu1; @AblowitzMussSAPM], can also be converted to local integrable equations through similar transformations. These transformations between nonlocal and local integrable equations offer a totally different way of studying these nonlocal equations, and they can be used for many purposes. First of all, these transformations immediately establish the integrability of the underlying nonlocal equations in view of the integrability of their local counterparts. Secondly, these transformations allow us to obtain the infinite number of conservation laws for the nonlocal equations from those of local equations. For example, from the first two conserved quantities of the NLS equation (\[e:PTNLSL\]), $$I_1=\int_{-\infty}^\infty \hat{u}(\hat{x},\hat{t})\hat{u}^*(\hat{x},\hat{t}) d\hat{x}, \quad I_2=\int_{-\infty}^\infty \hat{u}^*(\hat{x},\hat{t})\hat{u}_{\hat{x}}(\hat{x},\hat{t}) d\hat{x}, \nonumber$$ we immediately obtain through variable transformations (\[transNLS\]) the first two conserved quantities of the -symmetric NLS equation (\[e:PTNLS\]) [@AblowitzMussPRL2013; @AblowitzMussNonli2016] $$I_1=\int_{-\infty}^\infty u(x,t)u^*(-x, t)dx, \quad I_2=\int_{-\infty}^\infty u^*(-x, t)u_x(x,t) dx. \nonumber$$ Thirdly, these transformations allow us to construct analytical solutions of nonlocal equations from those of local ones. Examples of this will be presented in the next three sections. Fourthly, these transformations can be used to derive new nonlocal integrable equations from their local counterparts. This will be demonstrated in Sec. \[sec:newnonlocal\]. It is noted that the solution construction of nonlocal equations through these transformations may not be as trivial as it seems. The reason is that well-behaved solutions of the local equations may become ill-behaved under these transformations. For instance, the soliton solution of the local focusing NLS equation (\[e:PTNLSL\]) (with $\sigma=-1$), $$\hat{u}(\hat{x},\hat{t})=e^{i\hspace{0.04cm} \hat{t}} \mbox{sech}\hspace{0.03cm} \hat{x},$$ under transformations (\[transNLS\]), becomes a singular solution $$u(x,t)=e^{-i\hspace{0.04cm} t} \mbox{sec}\hspace{0.03cm} x$$ of the nonlocal defocusing NLS equation (\[e:PTNLS\]). Thus, in order to derive nonsingular solutions of nonlocal integrable equations through these transformations, one needs to choose solutions of local equations carefully. Rogue waves in partially -symmetric DS equations ================================================ In this section, we derive rogue wave solutions in partially -symmetric DS equations (\[PPTDS\]) using the transformation method. These rogue wave solutions have not been reported before to our best knowledge. Partially -symmetric DSI equations ---------------------------------- Rogue waves are rational solutions. According to the transformations (\[transPPTDS\]), rational solutions in partially -symmetric DSI equations (\[PPTDS\]) (with $\sigma^2=1$) can be obtained from rational solutions in local DSII equations (\[PPTDSL\]). These latter solutions have been reported in [@Satsuma_Ablowitz; @YangDSII]. Utilizing those solutions and the reverse variable transformations of (\[transPPTDS\]), i.e., $x\to -ix$, $t\to -t$ and $Q\to -Q$, and accounting for the sign switching of the nonlinearity parameter $\epsilon$, rational solutions in the partially -symmetric DSI equations (\[PPTDS\]) can be obtained. By imposing parameter conditions on these rational solutions, rogue wave solutions in these nonlocal DSI equations can then be derived. First, we consider fundamental rational solutions in the partially -symmetric DSI equations (\[PPTDS\]). These solutions are deduced from the fundamental rational solutions (11)-(12) of the local DSII equations in Ref. [@YangDSII] under the reverse variable transformations as $$\label{PPTAxyt} A(x,y,t)=\sqrt{2}\left[1-\frac{2i(-ia_2x+b_2y-\omega_2t+\theta_2)+1}{f}\right],$$ $$\label{PPTQxyt} Q(x,y,t)=\epsilon-(2\ln f)_{xx},$$ where $$\label{e:fDSI} f=(-ia_1x+b_1y-\omega_1t+\theta_1)^2+(-ia_2x+b_2y-\omega_2t+\theta_2)^2+\Delta,$$ $$\label{aa1a2} a=a_1+ia_2, \hspace{0.2cm} b=b_1+ib_2, \hspace{0.2cm} \omega=\omega_1+i\omega_2, \hspace{0.2cm} \theta=\theta_1+i\theta_2,$$ $$\label{abomega} a\equiv \frac{p+\epsilon/p}{2}, \quad b\equiv \frac{p-\epsilon/p}{2}i, \quad \omega\equiv \frac{p^2+1/p^2}{i}, \quad \Delta\equiv \frac{\epsilon|p|^2}{(|p|^2+\epsilon)^2},$$ $p$, $\theta$ are free complex parameters, and $a_{1,2}, b_{1,2}, \omega_{1,2}, \theta_{1,2}$ are the real and imaginary parts of complex numbers $a, b, \omega, \theta$. Performing solution analysis analogous to that in [@YangDSII], we find that this rational solution is a rogue wave when $p$ is purely imaginary. In this case, the solutions go to a constant background, $A\to \sqrt{2}$, $Q\to \epsilon$, as $t\to -\infty$. Since $p$ is imaginary, $a$ and $\omega$ are imaginary, and $b$ is real. Thus $a_1=b_2=\omega_1=0$, and the function $f$ in (\[e:fDSI\]) becomes $$f=(b_1y+\theta_1)^2+(-ia_2x-\omega_2 t+\theta_2)^2+\Delta.$$ This function is nonzero as long as $\omega_2t-\theta_2$ is nonzero. Thus, the above rogue wave is nonsingular as long as $\omega_2t-\theta_2\ne 0$. But when $\omega_2t-\theta_2=0$, i.e., at a critical time $t_c=\theta_2/\omega_2$, the function $f$ becomes zero at spatial positions where $$\label{e:hyperbola} (b_1y+\theta_1)^2-a_2^2x^2+\Delta=0.$$ In the generic case where $b_1\ne 0$ and $a_2\ne 0$, i.e., $p^2 \ne -1$, this equation defines a hyperbola on the $(x,y)$ plane. Thus, this rogue wave, which arises from a constant background, develops finite time singularity on the entire hyperbola (\[e:hyperbola\]) at the critical time $t_c=\theta_2/\omega_2$. In addition, this rogue wave exists for both signs of nonlinearity $\epsilon=\pm 1$. In the non-generic case where $p^2=-1$ and $\epsilon=1$, $a=0$, hence the above rogue wave is $x$-independent. Without $x$ dependence, the partially -symmetric DSI equations (\[PPTDS\]) degenerate to the local NLS equation with the spatial variable $y$, and the above rogue wave degenerates to the Peregrine rogue wave of the NLS equation [@Peregrine]. Note that the case of $p^2=-1$ and $\epsilon=-1$ is inadmissible since $\Delta$ is infinite here. To illustrate this fundamental rogue wave, we choose $\epsilon=1$, $p=0.5i$ and $\theta=1+i$. The corresponding rogue wave is plotted in Fig. \[f:fig1\]. In this case, the finite-time singularity occurs at $t_c=4/17\approx 0.2353$, thus we only plotted solutions up to time $t=0.22$, shortly before the blowup. ![(color online) An exploding fundamental rogue wave (\[PPTAxyt\]) in the partially -symmetric DSI equations (\[PPTDS\]) with $\sigma^2=1$, $\epsilon=1$, $p=0.5i$ and $\theta=1+i$. []{data-label="f:fig1"}](fig1){width="70.00000%"} It is interesting to compare this fundamental rogue wave of the nonlocal DSI equation with that of the local DSII equation [@YangDSII]. First of all, the parameter conditions are very different. In the local DSII equation, rogue waves require $|p|=1$; if $|p|\ne 1$, the rational solution would be a two-dimensional lump moving on a constant background. In the nonlocal DSI equation (\[PPTDS\]), rogue waves require $q$ to be purely imaginary. In this case, $|p|\ne 1$ generically, and thus the transformations (\[transPPTDS\]) convert moving-lump solutions of the local DSII equation into rogue waves of the nonlocal DSI equation. Secondly, in the local DSII equation, rogue waves exist only when $\epsilon=-1$; but in the nonlocal DSI equation (\[PPTDS\]), rogue waves exist for both signs of nonlinearity $\epsilon=\pm 1$. Thirdly, in the local DSII equation, fundamental rogue waves are line rogue waves; but in the nonlocal DSI equation, fundamental rogue waves have richer structures. Fourthly, in the local DSII equation, fundamental rogue waves never blow up in finite time; but in the nonlocal DSI equation, fundamental rogue waves generically blow up in finite time. Although some non-generic multi-rogue waves and higher-order rogue waves of the local DSII equation can also blow up in finite time, they only do so at a single spatial point, unlike the fundamental rogue waves of the nonlocal DSI equation where the blowup occurs on an entire hyperbola of the spatial plane. Multi-rogue waves of the nonlocal DSI equations (\[PPTDS\]) can be similarly derived from those of the local DSII equations in [@Satsuma_Ablowitz; @YangDSII] under the reverse variable transformations $x\to -ix$, $t\to -t$, $Q\to -Q$ and the parameter conditions of $p_j \hspace{0.07cm} (1\le j\le n)$ being purely imaginary. These multi-rogue waves describe the nonlinear interaction of several individual fundamental rogue waves. Details are omitted. It is noted that rogue waves in the fully -symmetric DS equations (\[PTDS\]) have been reported in [@HePTDS; @HePPTDS]. Those rogue waves can also be derived using our transformation method in view of the conversion of these nonlocal equations to the local DS equations as discussed in the previous section. Partially -symmetric DSII equations ----------------------------------- Rogue waves in partially -symmetric DSII equations (\[PPTDS\]), with $\sigma^2=-1$, can be obtained from the rational solutions in the local DSI equations (\[PPTDSL\]) under the reverse of transformations (\[transPPTDS\]). These rational solutions in the local DSI equations have been reported in [@Satsuma_Ablowitz; @YangDSI]. Imposing suitable parameter restrictions, rogue waves in the nonlocal DSII equations (\[PPTDS\]) will be obtained. Fundamental rational solutions in the nonlocal DSII equations (\[PPTDS\]) can be obtained from analogous solutions in Ref. [@YangDSI] for the local DSI equations under the reverse variable transformations $x\to -ix$, $t\to -t$, $Q\to -Q$, and accounting for the sign switching of the nonlinearity parameter $\epsilon$. These fundamental rational solutions of the nonlocal DSII equations are given by the same formulae (\[PPTAxyt\])-(\[abomega\]), except that the expressions for parameters $b$ and $\Delta$ are different: $$b\equiv \frac{p-\epsilon/p}{2}, \quad \Delta\equiv \frac{|p|^2}{(p+p^*)^2}.$$ As before, $p$ and $\theta$ are free complex constants. Analysis of these rational solutions shows that they become rogue waves when $\epsilon=-1$ and $|p|=1$, in which case $a, \omega$ are imaginary and $b$ real. These rogue waves approach a constant background as $t\to -\infty$, but develop finite-time singularity at time $t_c=\theta_2/\omega_2$ and on the hyperbola $$(b_1y+\theta_1)^2-a_2^2x^2+\Delta=0.$$ Graphs of these rogue waves are qualitatively similar to those in Fig. 1. Multi-rogue waves in the nonlocal DSII equations (\[PPTDS\]) can be derived from those of the local DSI equations in [@Satsuma_Ablowitz; @YangDSI] under variable transformations and parameter conditions of $\epsilon=-1$, $|p_j|=1, j=1, \dots, n$. Details are omitted. Rogue waves in the nonlocal derivative NLS equation =================================================== Now, we consider rogue waves in the nonlocal derivative NLS equation (\[e:PTDNLS\]), i.e., $$\label{e:PTDNLS2} iu_t(x,t)+u_{xx}(x,t)+\sigma \left[u^2(x,t)u^*(-x,t)\right]_x=0,$$ where $\sigma=\pm 1$. These rogue waves can be obtained from rational solutions of the local derivative NLS equation (\[e:PTDNLSL\]) through the variable transformation (\[transDNLS\]). This local equation is invariant when $\sigma\to -\sigma, x\to -x$, thus we fix $\sigma=1$ without loss of generality. For this $\sigma$ value, the fundamental rational solution in the local derivative NLS equation is given by Eq. (47) in Ref. [@GuoDNLS], and it is a moving soliton on a constant background. Then under the reverse of the transformation (\[transDNLS\]), i.e., $x\to -ix$, $t\to -t$, this moving soliton of the local equation is converted to the following fundamental rational solution of the nonlocal equation, $$\label{e:DNLSf} u(x,t)=\frac{(2ix-6t-i)(2ix-6t+3i)}{(2ix-6t+i)^2}.$$ The graph of this solution is displayed in Fig. 2(a). It is seen that this is a rogue wave, rising from a constant background and then retreating back to the same background, analogous to the Peregrine solution of the NLS equation. However, the present rogue wave blows up to infinity at $x=-1/2$ and finite time $t_c=0$, unlike the Peregrine solution. ![(color online) Rogue waves ($|u|$) in the nonlocal derivative NLS equation (\[e:PTDNLS2\]) with $\sigma=1$. (a) The fundamental rogue wave (\[e:DNLSf\]); (b, c, d) second-order rogue waves (\[e:DNLS2nd\]) with $\alpha=-10, 0$ and $10$ respectively.[]{data-label="f:fig2"}](fig2){width="70.00000%"} Higher-order rogue waves in the nonlocal equation (\[e:PTDNLS2\]) can be obtained from higher-order rational solutions of the local derivative NLS equation. For instance, the second-order rational solution of the local equation is given by Eq. (48) in Ref. [@GuoDNLS]. Then under the transformations $x\to -ix$, $t\to -t$, we get the second-order rational solution of the nonlocal equation (\[e:PTDNLS2\]) as $$\label{e:DNLS2nd} u(x,t)=\frac{(F_1-iF_2)F_3}{(F_1+iF_2)^2},$$ where $$F_1=8\eta^3+18\eta-48t+12\alpha, \quad F_2=12\eta^2+3, \nonumber$$ $$F_3=8\eta^3-30\eta-48t+12\alpha+i(36\eta^2-15), \quad \eta=ix-3t, \nonumber$$ and $\alpha$ is a free real constant. Graphs of these rational solutions are plotted in Fig. 2(b,c,d) for $\alpha=-10, 0$ and $10$ respectively. It is seen that these rational solutions are second-order rogue waves which arise from and retreat back to the same constant background. But they blow up to infinity at three points of the $(x,t)$ plane. Multi-solitons and quasi-periodic solutions in the reverse space-time CMKdV equation ==================================================================================== In this section, we derive analytical solutions for the reverse space-time CMKdV equation (\[e:NCMKdV\]), i.e., $$\label{e:NCMKdV2} q_{t}(x,t)+q_{xxx}(x,t)+6\sigma q(x,t)q^{*}(-x,-t)q_{x}(x,t)=0,$$ where $\sigma=\pm 1$. The case of $\sigma=1$ will be called the focusing case, and that of $\sigma=-1$ the defocusing case. As we have shown, this nonlocal equation, under transformations (\[transCMKdV\]), becomes the local CMKdV equation (\[e:CMKdV\]) with the opposite sign of nonlinearity. Thus, we will derive analytical solutions for the defocusing/focusing nonlocal CMKdV equation from those of the local focusing/defocusing CMKdV equation. Multi-solitons in the nonlocal focusing equation ------------------------------------------------ Eq. (\[e:NCMKdV2\]) in the focusing case has $\sigma=1$. Solitons and multi-solitons in this nonlocal focusing equation can be constructed from singular solutions in the local defocusing equation (\[e:CMKdV\]). The local defocusing equation admits the following singular solutions $$\begin{aligned} q(x,t)=\sqrt{\nu} \exp(i \phi) \sec \hspace{0.04cm} [\sqrt{\nu} \hspace{0.06cm} (x+\nu t+\delta)], \quad \nu>0,\end{aligned}$$ where $\nu$, $\phi$ and $\delta$ are real constants. Then under the reverse of transformations (\[transCMKdV\]), i.e., $x\to -ix$, $t\to it$, this singular solution becomes $$q(x,t)=\sqrt{\nu} \exp(i \phi) \hspace{0.06cm} \mbox{sech} \hspace{0.04cm} [\sqrt{\nu} \hspace{0.06cm} (x- \nu t + i \delta)], \quad \nu>0,$$ which is the fundamental soliton in the nonlocal focusing CMKdV equation (\[e:NCMKdV2\]). Its peak amplitude, which occurs at $x-\nu t=0$, is $\sqrt{\nu}\sec(\sqrt{\nu}\hspace{0.05cm}\delta)$. Thus, for a given $\nu$, this peak amplitude can vary from $\sqrt{\nu}$ to infinity depending on the choice of the $\delta$ values. Second-order singular solutions in the local defocusing equation (\[e:CMKdV\]) can be obtained from Ref. [@CMKdV2soliton] under certain parameter constraints \[specifically, by requiring $c_1, c_2$ negative in Eq. (3.18) of that paper\]. Then, under the above variable transformations, we get two-soliton solutions in the nonlocal focusing CMKdV equation (\[e:NCMKdV2\]) as $$\label{e:CMKdV2soliton} q(x,t)=\frac{{\kappa} \left[ \sqrt{\nu_{1}}\exp(i\phi_{1})\cosh({\theta_{2}})+\sqrt{\nu_{2}}\exp(i\phi_{2})\cosh({\theta_{1}}) \right]} {({\kappa}^2-1)\cos(\phi_{1}-\phi_{2})+{\kappa}^2\cosh({\theta_{1}}-{\theta_{2}})+\cosh({\theta_{1}}+{\theta_{2}})},$$ with $${\kappa}=\frac{\sqrt{\nu_1}+\sqrt{\nu_2}}{\sqrt{\nu_1}-\sqrt{\nu_2}}, \quad {\theta_{k}}=\sqrt{\nu_{k}} \hspace{0.08cm} {\xi_{k}}, \nonumber$$ $${\xi_{k}}=x-\nu_{k}t+i \hspace{0.04cm} \delta_{k}, \quad k=1,2. \nonumber$$ For parameter choices $$\label{CMKdVpara} \nu_{1}=1, \hspace{0.15cm} \nu_{2}=3, \hspace{0.15cm} \phi_{1}=\phi_{2}=0, \hspace{0.15cm} \delta_{1}=\frac{1}{2}, \hspace{0.15cm} \delta_{2}=-\frac{1}{2},$$ this two-soliton solution is displayed in Fig. 3(a). ![(color online) (a) A two-soliton solution (\[e:CMKdV2soliton\]) in the nonlocal focusing CMKdV equation (\[e:NCMKdV2\]) under parameter choices (\[CMKdVpara\]). (b) A quasi-periodic solution (\[e:CMKdVquasi\]) in the nonlocal defocusing CMKdV equation (\[e:NCMKdV2\]) under parameter choices (\[CMKdVpara2\]). []{data-label="f:fig3"}](fig3){width="70.00000%"} Higher-order solitons in the nonlocal focusing CMKdV equation (\[e:NCMKdV2\]) can be obtained similarly. Quasi-periodic solutions in the nonlocal defocusing equation ------------------------------------------------------------ Eq. (\[e:NCMKdV2\]) in the defocusing case has $\sigma=-1$. In this case, the local focusing CMKdV equation (\[e:NCMKdV\]) has a soliton solution $$\begin{aligned} q(x,t)=\sqrt{c} \exp(i \phi) \hspace{0.04cm} \mbox{sech} \hspace{0.04cm} [\sqrt{c} \hspace{0.07cm} (x-c \hspace{0.04cm} t+\delta)], \quad c>0,\end{aligned}$$ where $c, \phi$ and $\delta$ are real constants. Then under the transformations, we obtain the following solution for the nonlocal defocusing equation (\[e:NCMKdV2\]), $$\begin{aligned} q(x,t)=\sqrt{c}\exp(i \phi) \sec \hspace{0.04cm} [\sqrt{c} \hspace{0.07cm} (x+c \hspace{0.04cm} t+i \hspace{0.04cm} \delta)], \quad c>0.\end{aligned}$$ This solution is a traveling wave and is periodic in both space $x$ and time $t$. Second-order extensions of this periodic solution can be obtained from two-soliton solutions of the local focusing CMKdV equation. The latter can be found in Eq. (3.18) of Ref. [@CMKdV2soliton]. Then, under variable transformations $x\to -ix$, $t\to it$, we get the following solution for the defocusing nonlocal CMKdV equation (\[e:NCMKdV2\]) as $$\label{e:CMKdVquasi} q(x,t)=\frac{\kappa \left[\sqrt{c_{1}}\exp(i \hspace{0.04cm} \phi_{1})\cos(\theta_{2})+\sqrt{c_{2}}\exp(i \hspace{0.04cm} \phi_{2})\cos(\theta_{1})\right]} {(\kappa^2-1)\cos(\phi_{1}-\phi_{2})+\kappa^2\cos(\theta_{1}-\theta_{2})+\cos(\theta_{1}+\theta_{2})},$$ where $$\kappa =\frac{\sqrt{c_1}+\sqrt{c_2}}{\sqrt{c_1}-\sqrt{c_2}}, \quad \theta_{j}=\sqrt{c_{j}} \hspace{0.08cm} \xi_{j}, \quad \xi_{j}=x+c_{j} \hspace{0.04cm} t+i \hspace{0.04cm} \delta_{j}. \nonumber$$ This solution contains two frequencies and is generically quasi-periodic in both space and time. Under the parameter choices $$\label{CMKdVpara2} c_{1}=1, \hspace{0.15cm} c_{2}=3, \hspace{0.15cm} \phi_{1}=\phi_{2}=0, \hspace{0.15cm} \delta_{1}=\frac{1}{2}, \hspace{0.15cm} \delta_{2}=-\frac{1}{2},$$ this double-frequency quasi-periodic solution is displayed in Fig. 3(b). Higher-order quasi-periodic solutions in the nonlocal defocusing CMKdV equation (\[e:NCMKdV2\]) can be obtained from higher-order solitons of the local focusing CMKdV equation in a similar way. New nonlocal integrable equations {#sec:newnonlocal} ================================= In this last section, we show how these variable transformations can be used to derive new integrable nonlocal equations from their local counterparts. Nonlocal complex short pulse equations -------------------------------------- As the first example, we consider the integrable local complex short pulse (CSP) equation [@Feng_shortpulse; @Feng3], $$\label{e:CSP} q_{xt} + q+\frac{1}{2}\sigma \left(|q|^2q_x\right)_x=0, \quad \sigma=\pm 1,$$ where $q(x,t)$ is a complex function. Under variable transformations $x \to -ix$, we get an integrable reverse space nonlocal CSP equation, $$\label{e:CSPspace} q_{xt}(x,t)-iq(x,t)+\frac{1}{2}i\sigma [q(x,t)q_x(x,t)q^*(-x,t)]_x=0.$$ Under a different variable transformation $t\to it$, we get an integrable reverse time nonlocal CSP equation, $$\label{e:CSPtime} q_{xt}(x,t)+i q(x,t)+\frac{1}{2}i\sigma[q(x,t)q_x(x,t)q^*(x,-t)]_x=0.$$ Under the combined transformations $x\to -ix, t\to it$, we get an integrable reverse space-time nonlocal CSP equation, $$\label{e:CSPspacetime} q_{xt}(x,t)+ q(x,t)-\frac{1}{2}\sigma[q(x,t)q_x(x,t)q^*(-x,-t)]_x=0.$$ Notice that this last nonlocal equation admits a reduction of $q(x,t)$ being real-valued. Under this reduction, we get an integrable reverse space-time real short-pulse equation $$\label{e:RSPspacetime} q_{xt}(x,t)+ q(x,t)-\frac{1}{2}\sigma[q(x,t)q_x(x,t)q(-x,-t)]_x=0,$$ where $q(x,t)$ is a real function. Infinite numbers of conservation laws for these new nonlocal short-pulse equations can be inferred directly from those of local short-pulse equations through the corresponding variable transformations. For instance, the first two conserved quantities of the local CSP equation (\[e:CSP\]) are $$\begin{aligned} I_1&=&\int_{-\infty}^\infty \sqrt{1+\sigma \hspace{0.05cm} q_x(x,t) \hspace{0.05cm} q^*_x(x,t)} \hspace{0.1cm} dx, \\ I_2&=&\int_{-\infty}^\infty \frac{q_{xx}(x,t)}{q_x(x,t)\sqrt{1+\sigma \hspace{0.05cm} q_x(x,t) \hspace{0.05cm} q^*_x(x,t)}} \hspace{0.1cm} dx.\end{aligned}$$ The former quantity has been reported in [@Feng_shortpulse; @Feng3], and we found the latter quantity by inspiration of conserved quantities for the Wadati-Konno-Ichikawa hierarchy (which contains the real short pulse equation) [@Franca2012]. Then, using these conserved quantities and the transformation $x \to -ix$, we obtain the first two conserved quantities of the reverse space nonlocal CSP equation (\[e:CSPspace\]) as $$\begin{aligned} I_1&=&\int_{-\infty}^\infty \sqrt{1-\sigma \hspace{0.05cm} q_x(x,t) \hspace{0.05cm} q^*_x(-x,t)} \hspace{0.1cm} dx, \\ I_2&=&\int_{-\infty}^\infty \frac{q_{xx}(x,t)}{q_x(x,t)\sqrt{1-\sigma \hspace{0.05cm} q_x(x,t) \hspace{0.05cm} q^*_x(-x,t)}} \hspace{0.1cm} dx.\end{aligned}$$ Under the transformation $t\to it$, we obtain the first two conserved quantities of the reverse time nonlocal CSP equation (\[e:CSPtime\]) as $$\begin{aligned} I_1&=&\int_{-\infty}^\infty \sqrt{1+\sigma \hspace{0.05cm} q_x(x,t) \hspace{0.05cm} q^*_x(x,-t)} \hspace{0.1cm} dx, \\ I_2&=&\int_{-\infty}^\infty \frac{q_{xx}(x,t)}{q_x(x,t)\sqrt{1+\sigma \hspace{0.05cm} q_x(x,t) \hspace{0.05cm} q^*_x(x,-t)}} \hspace{0.1cm} dx.\end{aligned}$$ Under the combined transformations $x\to -ix, t\to it$, we obtain the first two conserved quantities of the reverse space-time nonlocal CSP equation (\[e:CSPspacetime\]) as $$\begin{aligned} I_1&=&\int_{-\infty}^\infty \sqrt{1-\sigma \hspace{0.05cm} q_x(x,t) \hspace{0.05cm} q^*_x(-x,-t)} \hspace{0.1cm} dx, \label{e:CSPST1} \\ I_2&=&\int_{-\infty}^\infty \frac{q_{xx}(x,t)}{q_x(x,t)\sqrt{1-\sigma \hspace{0.05cm} q_x(x,t) \hspace{0.05cm} q^*_x(-x,-t)}} \hspace{0.1cm} dx. \label{e:CSPST2}\end{aligned}$$ The first two conserved quantities of the reverse space-time real short-pulse equation (\[e:RSPspacetime\]) are simply these $I_1, I_2$ in (\[e:CSPST1\])-(\[e:CSPST2\]) with the complex conjugate removed. Higher conserved quantities of these new nonlocal CSP equations can be similarly obtained. Nonlocal nonlinear diffusion equations -------------------------------------- As a second example, we consider the local integrable NLS equation $$\label{e:NLS6} iu_t+u_{xx}+2\sigma |u|^2u=0, \quad \sigma=\pm 1.$$ Under the variable transformation $t\to -it$, we get an integrable reverse time nonlinear diffusion equation $$\label{e:nonlocal_diffision1} u_t(x,t)-u_{xx}(x,t)-2\sigma u^2(x,t)u^*(x,-t)=0.$$ Under the variable transformations $x\to ix, t\to it$, we get an integrable reverse space-time nonlinear diffusion equation $$\label{e:nonlocal_diffision2} u_t(x,t)-u_{xx}(x,t)+2\sigma u^2(x,t)u^*(-x,-t)=0.$$ Notice that both of these nonlocal equations admit the reduction of $u(x,t)$ being real. Under this reduction, we also obtain integrable reverse-time and reverse-space-time real nonlinear diffusion equations $$\label{e:realdiffusion1} u_t(x,t)-u_{xx}(x,t)-2\sigma u^2(x,t)u(x,-t)=0,$$ and $$\label{e:realdiffusion2} u_t(x,t)-u_{xx}(x,t)+2\sigma u^2(x,t)u(-x,-t)=0.$$ Infinite numbers of conservation laws for these new nonlocal diffusion equations can be readily derived from those of the local NLS equation (\[e:NLS6\]) through variable transformations. For instance, using the first four conserved quantities of the local NLS equation [@Yang2010], we immediately obtain the first four conserved quantities of the reverse time nonlinear diffusion equation (\[e:nonlocal\_diffision1\]) as $$\begin{aligned} && \hspace{-1cm} I_1=\int_{-\infty}^\infty u(x,t)u^*(x, -t)dx, \\ && \hspace{-1cm} I_2=\int_{-\infty}^\infty u^*(x, -t)u_x(x,t)dx, \\ && \hspace{-1cm} I_3=\int_{-\infty}^\infty \left[u^*(x, -t)u_{xx}(x,t)+u^2(x,t)u^{*2}(x, -t)\right] dx, \\ && \hspace{-1cm} I_4=\int_{-\infty}^\infty u^*(x, -t)\left\{ u_{xxx}(x,t)+[u^2(x,t)u^*(x, -t)]_x +2u^*(x,-t)u(x,t)u_x(x,t)\right\}dx.\end{aligned}$$ Likewise, the first four conserved quantities of the reverse space-time nonlinear diffusion equation (\[e:nonlocal\_diffision2\]) are found to be $$\begin{aligned} && \hspace{-1cm} I_1=\int_{-\infty}^\infty u(x,t)u^*(-x, -t)dx, \\ && \hspace{-1cm} I_2=\int_{-\infty}^\infty u^*(-x, -t)u_x(x,t)dx, \\ && \hspace{-1cm} I_3=\int_{-\infty}^\infty \left[-u^*(-x, -t)u_{xx}(x,t)+u^2(x,t)u^{*2}(-x, -t)\right] dx, \\ && \hspace{-1cm} I_4=\int_{-\infty}^\infty u^*(-x, -t)\left\{-u_{xxx}(x,t)+[u^2(x,t)u^*(-x, -t)]_x +2u^*(-x,-t)u(x,t)u_x(x,t)\right\}dx.\end{aligned}$$ Conserved quantities for the reverse-time and reverse-space-time real nonlinear diffusion equations (\[e:realdiffusion1\])-(\[e:realdiffusion2\]) are simply those of the complex equations above with the conjugation removed. Analytical solutions to these nonlocal diffusion equations can also be derived from solutions of the local NLS equation through transformations. For example, from the soliton solution of the local focusing NLS equation (\[e:NLS6\]), $$u(x,t)=\eta \hspace{0.05cm} \mbox{sech} \hspace{0.05cm} [\eta (x-c \hspace{0.04cm} t)]\exp\left\{\frac{1}{2}i \hspace{0.04cm} c \hspace{0.04cm} x+i\left(\eta^2-\frac{1}{4}c^2\right)t\right\},$$ with $\eta, c$ being real constants, we obtain the solution to the reverse time nonlocal diffusion equation (\[e:nonlocal\_diffision1\]) with $\sigma=1$ as $$\label{e:diffusion1} u(x,t)=\eta \hspace{0.05cm} \mbox{sech} \hspace{0.05cm} [\eta (x+i\hspace{0.04cm} c \hspace{0.04cm} t)]\exp\left\{\frac{1}{2}i \hspace{0.04cm} c \hspace{0.04cm} x+\left(\eta^2-\frac{1}{4}c^2\right)t\right\}.$$ This solution exponentially grows or decays depending on the sign of $\eta^2-c^2/4$. In addition, it periodically collapses at location $x=0$. For $\eta=c=1$, this solution is illustrated in Fig. 4 (left panel). ![(color online) Left: solution (\[e:diffusion1\]) to the reverse time nonlocal diffusion equation (\[e:nonlocal\_diffision1\]) with $\sigma=\eta=c=1$. Right: solution (\[e:diffusion2\]) to the reverse space-time nonlocal diffusion equation (\[e:nonlocal\_diffision2\]) with $\sigma=1$. []{data-label="f:fig4"}](fig4){width="70.00000%"} As another example, from the Peregrine rogue wave solution of the local focusing NLS equation (\[e:NLS6\]), $$u(x, t)=e^{2it}\left(1-\frac{4(1-4it)}{1+4x^2+16t^2}\right),$$ and utilizing the transformations $x\to ix, t\to it$, we obtain the following solution to the reverse space-time nonlocal diffusion equation (\[e:nonlocal\_diffision2\]) with $\sigma=1$ as $$\label{e:diffusion2} u(x, t)=e^{-2t}\left(1-\frac{4(1+4t)}{1-4x^2-16t^2}\right).$$ This solution is illustrated in Fig. 4 (right panel). It decays exponentially with time, but blows up to infinity on the ellipse $4x^2+16t^2=1$ of the $(x,t)$ plane. Other new nonlocal integrable equations --------------------------------------- In addition to the above new nonlocal integrable equations, we can also obtain many other such equations using transformations. For instance, from the local Sasa-Satsuma equation [@SS] $$u_t+u_{xxx}+6|u|^2u_x+3u\left(|u|^2\right)_x=0,$$ and employing the variable transformations $x\to ix, t\to -it$, we get an integrable reverse space-time Sasa-Satsuma equation, $$\begin{aligned} && \hspace{-1cm} u_t(x,t)+u_{xxx}(x,t)-6u(x,t)u^*(-x,-t)u_x(x,t)-3u(x,t)\left[u(x,t)u^*(-x,-t)\right]_x=0.\end{aligned}$$ From the local Chen-Lee-Liu equation [@ChenLeeLiu] $$iu_t+u_{xx}+i|u|^2u_x=0,$$ and employing the variable transformation $x\to ix, t \to -t$, we obtain an integrable reverse-space Chen-Lee-Liu equation $$iu_t(x,t)+u_{xx}(x,t)-u(x,t)u^*(-x,t)u_x(x,t)=0.$$ A different transformation $x\to ix, t\to it$ yields a different integrable reverse-space-time nonlinear diffusion equation $$u_t(x,t)-u_{xx}(x,t)+u(x,t)u^*(-x,-t)u_x(x,t)=0.$$ From the local modified NLS equation [@WadatiMNLS] $$iu_t+u_{xx}+i\alpha (|u|^2u)_x+\beta |u|^2u=0$$ with real constants $\alpha, \beta$, and under transformations $x\to ix, t \to -t$, we get an integrable reverse-space modified NLS equation $$iu_t(x,t)+u_{xx}(x,t)-\alpha \left[u^2(x,t)u^*(-x,t)\right]_x-\beta u^2(x,t)u^*(-x,t)=0.$$ From an integrable (2+1)-dimensional NLS equation [@2DNLS_lump_rogue] $$iq_t+q_{xy}+2iq(qq_x^*-q^*q_x)=0,$$ and taking the transformations $x\to ix, y\to -iy$, we obtain an integrable reverse-space (2+1)-dimensional NLS equation $$\begin{aligned} && \hspace{-1.3cm} iq_t(x,y,t)+q_{xy}(x,y,t)+2q(x,y,t)\left[q(x,y,t)q_x^*(-x,-y,t) -q^*(-x,-y,t)q_x(x,y,t)\right]=0.\end{aligned}$$ Thus, this transformation technique is a powerful tool to generate a large class of new nonlocal integrable equations. Solution dynamics in these new nonlocal equations can also be studied through this transformation, as we have demonstrated earlier in this article. Summary ======= In summary, we have reported that many recently proposed nonlocal integrable equations can be converted to local integrable equations through simple variable transformations. Examples include -symmetric NLS and Davey-Stewartson equations, a nonlocal derivative NLS equation, the reverse space-time complex modified Korteweg-de Vries equation, reverse space-time three wave interaction equations, and many others. These transformations not only establish the integrability of these nonlocal equations, but also allow us to construct their analytical solutions from solutions of the local equations. These transformations can also be used to derive new nonlocal integrable equations. As applications of these transformations, we have used them to derive rogue wave solutions for the partially -symmetric Davey-Stewartson equations and the nonlocal derivative NLS equation. In addition, we have used them to derive multi-soliton and quasi-periodic solutions in the reverse space-time complex modified KdV equation. Furthermore, we have used them to construct many new nonlocal integrable equations such as nonlocal short pulse equations, nonlocal nonlinear diffusion equations, nonlocal Sasa-Satsuma equations and nonlocal Chen-Lee-Liu equations. These transformations reveal an intimate and deep connection between many nonlocal and local integrable equations. They are expected to provide a new and powerful tool in the study of these nonlocal equations. Acknowledgment {#acknowledgment .unnumbered} ============== This material is based upon work supported by the Air Force Office of Scientific Research under award number FA9550-12-1-0244, and the National Science Foundation under award number DMS-1616122. The work of B.Y. is supported by a visiting-student scholarship from the Chinese Scholarship Council. References {#references .unnumbered} ========== [10]{} M.J. Ablowitz and H. Segur, *Solitons and Inverse Scattering Transform* (SIAM, Philadelphia, 1981). S.P. Novikov, S.V. Manakov, L.P. Pitaevskii and V.E. Zakharov, *Theory of Solitons* (Plenum, New York, 1984) L. Takhtadjan and L. Faddeev, *The Hamiltonian Approach to Soliton Theory* (Springer Verlag, Berlin, 1987). M.J. Ablowitz and P.A. Clarkson, *Solitons, Nonlinear Evolution Equations and Inverse Scattering* (Cambridge University Press, 1991). J. Yang, *Nonlinear Waves in Integrable and Non integrable Systems* (SIAM, Philadelphia, 2010). M.J. Ablowitz and Z.H. Musslimani, “Integrable nonlocal nonlinear Schrödinger equation", Phys. Rev. Lett. 110, 064105 (2013). M.J. Ablowitz and Z.H. Musslimani, “Integrable discrete symmetric model", Phys. Rev. E 90, 032912 (2014). Z. Yan, “Integrable -symmetric local and nonlocal vector nonlinear Schrödinger equations: A unified two-parameter model," Appl. Math. Lett. 47, 61–68 (2015). A. Khara and A. Saxena, “Periodic and hyperbolic soliton solutions of a number of nonlocal nonlinear equations", J. Math. Phys. 56, 032104 (2015). M.J. Ablowitz and Z.H. Musslimani, “Inverse scattering transform for the integrable nonlocal nonlinear Schrödinger equation," Nonlinearity 29, 915–946 (2016). C.Q. Song, D.M. Xiao and Z.N. Zhu, “Solitons and dynamics for a general integrable nonlocal coupled nonlinear Schrödinger equation", Commun. Nonlinear Sci. Numer. Simul. 45, 13–28 (2017). A.S. Fokas, “Integrable multidimensional versions of the nonlocal nonlinear Schrödinger equation", Nonlinearity 29, 319–324 (2016). S.Y. Lou, “Alice-Bob systems, $P_s$-$T_d$-$C$ principles and multi-soliton solutions", https://arxiv.org/abs/1603.03975 (2016). S.Y. Lou and F. Huang, “Alice-Bob physics: coherent solutions of nonlocal KdV systems", Scientific Reports 7, 869 (2017). M.J. Ablowitz and Z.H. Musslimani, “Integrable nonlocal nonlinear equations", Stud. Appl. Math. DOI: 10.1111/sapm.12153 (2016). Z.X. Xu, and K.W. Chow, “Breathers and rogue waves for a third order nonlocal partial differential equation by a bilinear transformation", Appl. Math. Lett. 56, 72–77 (2016). Z.X. Zhou, “Darboux transformations and global solutions for a nonlocal derivative nonlinear Schrödinger equation", arXiv:1612.04892 \[nlin.SI\] (2016). Z.X. Zhou, “Darboux transformations and global explicit solutions for nonlocal Davey-Stewartson I equation", arXiv:1612.05689 \[nlin.SI\] (2016). J.G. Rao, Y.S. Zhang, A.S. Fokas, and J.S. He, “Rogue waves of the nonlocal Davey-Stewartson I equation" (preprint 2016). J.G. Rao, Y. Cheng and J.S. He, “Rational and semi-rational solutions of the nonlocal Davey-Stewartson equations", to appear in Stud. Appl. Math. (2017). J.L. Ji and Z.N. Zhu, “On a nonlocal modified Korteweg-de Vries equation: integrability, Darboux transformation and soliton solutions", Commun. Nonlinear Sci. Numer. Simul. 42 699–708 (2017). L.Y. Ma, S.F. Shen and Z.N. Zhu, “Integrable nonlocal complex mKdV equation: soliton solution and gauge equivalence", arXiv:1612.06723 \[nlin.SI\] (2016). V. S. Gerdjikov and A. Saxena, “Complete integrability of nonlocal nonlinear Schrödinger equation", J. Math. Phys. 58, 013502 (2017). M.J. Ablowitz, X. Luo, and Z.H. Musslimani, “Inverse scattering transform for the nonlocal nonlinear Schrödinger equation with nonzero boundary conditions", arXiv:1612.02726 \[nlin.SI\] (2016). S.V. Suchkov, A.A. Sukhorukov, J. Huang, S.V. Dmitriev, C. Lee and Y.S. Kivshar, “Nonlinear switching and solitons in -symmetric photonic systems", Laser Photonics Rev. 10, 177 (2016). V.V. Konotop, J. Yang and D.A. Zezyulin, “Nonlinear waves in -symmetric systems," Rev. Mod. Phys. 88, 035002 (2016). T.A. Gadzhimuradov and A.M. Agalarov, “Towards a gauge-equivalent magnetic structure of the nonlocal nonlinear Schrödinger equation", Phys. Rev. A 93, 062124 (2016). D.J. Kaup and A.C. Newell, “An exact solution for a derivative nonlinear Schrödinger equation", J. Math. Phys. 19, 798 (1978). J. Satsuma and M.J. Ablowitz, “Two-dimensional lumps in nonlinear dispersive systems", J. Math. Phys. 20, 1496–503 (1979). Y. Ohta and J. Yang, “Dynamics of rogue waves in the Davey-Stewartson II equation", J. Phys. A 46, 105202 (2013). D.H. Peregrine, “Water waves, nonlinear Schrödinger equations and their solutions", J. Aust. Math. Soc. B 25, 16–43 (1983). Y. Ohta and J. Yang, “Rogue waves in the Davey-Stewartson-I equation", Phys. Rev. E 86, 036604 (2012). B. Guo, L. Ling, and Q.P. Liu, “High-order solutions and generalized Darboux transformations of derivative nonlinear Schrödinger equations", Stud. Appl. Math. 130, 317–344 (2013). S. C. Anco, N. T. Ngatat and M. Willoughby, “Interaction properties of complex modified Korteweg–de Vries (mKdV) solitons", Physica D, 240, 1378–1394 (2011). B.F. Feng, “Complex short pulse and coupled complex short pulse equations," Physica D 297, 62–75 (2015). B.F. Feng, K. Maruno and Y. Ohta, “Geometric formulation and multi-dark soliton solution to the defocusing complex short pulse equation", Stud. Appl. Math. 138, 343–367 (2017). G.S. Franca, J.F. Gomes and A.H. Zimerman, “The higher grading structure of the WKI hierarchy and the two-component short pulse equation", J. High Energ. Phys. 2012:120, doi:10.1007/JHEP08(2012)120 (2012). N. Sasa and J. Satsuma, “New type of soliton solutions for a higher-order nonlinear Schrodinger equation", J. Phys. Soc. Jpn. 60, 409 (1991). H.H. Chen, Y.C. Lee and C.S. Liu, “Integrability of nonlinear Hamiltonian system by inverse scattering method", Phys. Scr. 20, 490–492 (1979). M. Wadati, K. Konno, and Y.H. Ichikawa, “A generalization of inverse scattering method", J. Phys. Soc. Jpn. 46, 1965 (1979). A. Kundu, A. Mukherjee and T. Naskar, “Modelling rogue waves through exact dynamical lump soliton controlled by ocean currents", Proc. R. Soc. A 470, 20130576 (2014).
{ "pile_set_name": "ArXiv" }
--- author: - 'Sangmin Choi, Ratindranath Akhoury' bibliography: - 'references.bib' title: 'BMS Supertranslation Symmetry Implies Faddeev-Kulish Amplitudes' --- Introduction ============ Scattering amplitudes in gauge and gravitational theories suffer from infrared divergences, which (upon resummation) have the net effect of rendering all such amplitudes zero. The traditional way of dealing with this problem is to employ the Bloch-Nordsieck method [@BN], where one constructs an inclusive cross section out of all processes that are physically indistinguishable, i.e., including the contributions of undetectable soft bosons (photons or gravitions). While this approach allows one to obtain cross sections that can be used to match with experiments, it has the shortcoming of giving up on the notion of a well-defined S-matrix element. An alternative to this method is to use the asymptotic states of Faddeev and Kulish [@Kulish:1970ut] in place of Fock states. These states can be interpreted as Fock states dressed by an infinite number of soft bosons, which are commonly referred to as the boson clouds. It has been shown that using a set of such states as basis yields well-defined, infrared finite S-matrix elements. [@Chung:1965zza; @Ware:2013zja]. There has been work done in the recent years, for example [@Kapec:2015ena; @He:2014cra; @He:2014laa; @Strominger:2013jfa], that revealed the existence of an infinite number of degenerate vacua due to the spontaneous breaking of the asymptotic symmetries in gauge and gravity theories. It has been argued that the vanishing of all S-matrix elements in the traditional approach is reflecting the fact that scattering processes induce a transition between the degenerate vacua, in a way that conserves the charges of the broken symmetries. This paved the way to understanding the connection of asymptotic symmetries to the formalism of Faddeev and Kulish. The boson clouds of the Faddeev-Kulish states have been shown, in [@Gabai:2016kuf] for QED and [@Choi:2017bna] for gravity, to precisely cancel the vacuum transitions induced by the scattering operator, which explains why Faddeev-Kulish states yield well-defined S-matrix elements. Also, it has been shown in [@Kapec:2017tkm] that a gauge-invariant formulation of the charged particles in QED analogous to [@Dirac; @Bagan:1999jf] yields coherent states that are essentially equivalent to Faddeev-Kulish states, which turn out to be the charge eigenstates of the large gauge symmetry. In this paper, we take one step further and argue that the infrared-finite scattering amplitudes constructed using the Faddeev-Kulish states (henceforth referred to as Faddeev-Kulish amplitudes) naturally arise as a consequence of asymptotic symmetry. Thus, since Faddeev-Kulish amplitudes are infrared finite, so are all BMS supertranslation charge conserving amplitudes. We demonstrate this in the case of perturbative quantum gravity, by constructing eigenstates of BMS supertranslation charge and showing that any scattering amplitude that conserves this charge is equivalent to the Faddeev-Kulish amplitude. In this process, we show that the graviton clouds “weakly commute" with the scattering operator, in the sense that clouds in the incoming state can freely be moved to the outgoing state and vice versa. Our work provides a natural proof of the conjecture made in [@Kapec:2017tkm], which claims that amplitudes conserving charges of asymptotic symmetries are infrared finite. We conclude with an application of our results to the study initiated in [@Carney:2017jut], where information theoretic properties of low energy photons and gravitons are analyzed through the study of the relevant density matrices. Our approach here, in contrast to [@Carney:2017jut], is to derive expressions for the density matrices which satisfy conservation of BMS supertranslation charge at all stages. If the measurements are sensitive only to the momenta of the hard matter particles, then the conclusions of [@Carney:2017jut; @Carney:2017oxp] are unchanged. There is a decoherence of the momentum configurations of these particles. This decoherence and the consequent high degree of correlations between the hard and soft quanta was also noted independently in [@Strominger:2017aeh] using a different approach. The paper is organized as follows. In section \[EIGENSTATE\], we construct the eigenstates of BMS supertranslation charge, and study the implications of the charge conservation on the scattering amplitudes. We establish in section \[MOVECLOUD\] that the Faddeev-Kulish graviton clouds weakly commute with the scattering operator. This result is used in section \[SAME\] to show the equality between amplitudes that conserve the supertranslation charge and the Faddeev-Kulish amplitudes. In section \[DECOHERENCE\], we apply the preceding results to the analysis of [@Carney:2017jut]. We wrap up with a brief discussion in section \[DISCUSSION\]. BMS charge and eigenstates {#EIGENSTATE} ========================== In order to establish notation and to make connections with earlier work, we will begin with a review of BMS symmetry and the conserved charges. As is customary, we will employ the retarded coordinates $(u,r,z,{{\bar{z}}})$, defined in terms of the Cartesian coordinates $(t,x_1,x_2,x_3)$ as $$\begin{aligned} u = t-r,\qquad r^2=x_1^2+x_2^2+x_3^2,\qquad z=\frac{x_1+ix_2}{r+x_3}.\end{aligned}$$ Here $u$ is the retarded time and $z$ is the complex coordinate on the unit 2-sphere with the metric ${{\gamma_{z\bar{z}}}}= \frac{2}{(1+z{{\bar{z}}})^2}$. Then in the Bondi gauge [@Bondi:1962px; @Sachs:1962wk], the asymptotically flat metric has the expansion [@Strominger:2013jfa] $$\begin{aligned} \begin{split} ds^2 =& -du^2 - 2dudr + 2r^2 {{\gamma_{z\bar{z}}}}dz d{{\bar{z}}}\\ & +\frac{2m_B}{r}du^2 + rC_{zz}dz^2 + rC_{{{\bar{z}}}{{\bar{z}}}}d{{\bar{z}}}^2 + D^zC_{zz}dudz + D^{{\bar{z}}}C_{{{\bar{z}}}{{\bar{z}}}}dud{{\bar{z}}}\\ & +\cdots, \end{split}\end{aligned}$$ where $m_B$ is the Bondi mass aspect and $D^z$, $D^{{\bar{z}}}$ are the 2-sphere covariant derivatives. The gravitational radiation is characterized by the Bondi news tensor $N_{zz}={{\partial}}_u C_{zz}$. The BMS supertranslation charge for a 2-sphere function $f=f(w,{{\bar{w}}})$ is then $$\begin{aligned} Q(f) = Q_S(f) + Q_H(f),\end{aligned}$$ where, explicit expressions for the soft part $Q_S$ and the hard part $Q_H$ are given in [@Strominger:2017zoo; @Campiglia:2015kxa]. We are interested in these expressions at the leading terms in the large-$r$ expansion which are known to be gauge-invariant [@Avery:2015gxa]. The action of the hard charge $Q_H$ on a Fock state of $N$ massive particles can be expressed as [@Campiglia:2015kxa] $$\begin{aligned} Q_H\ket{{\mathbf}{p}_1,\ldots,{\mathbf}{p}_N} &= \sum_{i=1}^N \tilde{f}(p_i)\ket{{\mathbf}{p}_1,\ldots,{\mathbf}{p}_N},\end{aligned}$$ where $p_i^\mu = (E_k,{\mathbf}{p}_i)$, and $$\begin{aligned} \tilde{f}(p) = -\frac{1}{2\pi} \int d^2w \, \frac{(\epsilon^+(w,{{\bar{w}}})\cdot p)^2}{p\cdot \hat{x}_w} D^2_{{\bar{w}}}f(w,{{\bar{w}}}).\end{aligned}$$ Here $\hat{x}^\mu_w=(1,\hat{{\mathbf}{x}}_w)$ with the unit vector $\hat{{\mathbf}{x}}_w$ pointing in the direction $(w,{{\bar{w}}})$, and the polarization vectors have components $$\begin{aligned} \epsilon^{-\mu}(z,{{\bar{z}}})=\frac{1}{\sqrt{2}}(z,1,i,-z) \quad\text{and}\quad \epsilon^{+\mu}(z,{{\bar{z}}})=\frac{1}{\sqrt{2}}({{\bar{z}}},1,-i,-{{\bar{z}}}).\end{aligned}$$ The action of the soft charge $Q_S$ on the same state is [@Strominger:2013jfa] $$\begin{aligned} Q_S\ket{{\mathbf}{p}_1,\ldots,{\mathbf}{p}_N} &= -\frac{1}{8\pi G}\int du\, d^2w\, {{\gamma_{w\bar{w}}}}N^{{{\bar{w}}}{{\bar{w}}}}D_{{\bar{w}}}^2 f \ket{{\mathbf}{p}_1,\ldots,{\mathbf}{p}_N}.\end{aligned}$$ Conservation of BMS supertranslation charges imply, $$\begin{aligned} \label{Scons} \braket{{{\text{out}}}|\, [Q(f),{\mathcal{S}}]\,|{{\text{in}}}} = 0,\end{aligned}$$ which should hold for all functions $f(w,{{\bar{w}}})$. In particular, let us choose $$\begin{aligned} f(w,{{\bar{w}}}) = \frac{(1+w{{\bar{w}}})({{\bar{w}}}-{{\bar{z}}})}{(1+z{{\bar{z}}})(w-z)},\end{aligned}$$ such that [@Campiglia:2015kxa] $$\begin{aligned} D_{{\bar{w}}}^2 f(w,{{\bar{w}}}) = 2\pi \delta^2(w-z).\end{aligned}$$ With this choice, the conservation law reads $$\begin{aligned} \label{rawcons} \frac{{{\gamma_{z\bar{z}}}}}{4 G}\int_{-\infty}^\infty du\braket{{{\text{out}}}|(N^{{{\bar{z}}}{{\bar{z}}}}{\mathcal{S}}- {\mathcal{S}}N^{{{\bar{z}}}{{\bar{z}}}})|{{\text{in}}}} = -\sum_i \eta_i \frac{(p_i\cdot \epsilon^+(z,{{\bar{z}}}))^2}{p_i\cdot \hat{x}_z} \braket{{{\text{out}}}|{\mathcal{S}}|{{\text{in}}}},\end{aligned}$$ where the sum on the RHS runs over all external particles and $\eta_i=+1$ ($-1$) if $i$ is an outgoing (incoming) particle. Let us define the operator $$\begin{aligned} N(z,{{\bar{z}}}) \equiv {{\gamma_{z\bar{z}}}}\int^\infty_{-\infty} du\,N^{{{\bar{z}}}{{\bar{z}}}} = {{\gamma^{z\bar{z}}}}\int^\infty_{-\infty}du\,N_{zz}.\end{aligned}$$ Then becomes $$\begin{aligned} \braket{{{\text{out}}}|(N(z,{{\bar{z}}}){\mathcal{S}}- {\mathcal{S}}N(z,{{\bar{z}}}))|{{\text{in}}}} = -\frac{\kappa^2}{8\pi}\sum_i \eta_i \frac{(p_i\cdot \epsilon^+(z,{{\bar{z}}}))^2}{p_i\cdot \hat{x}_z} \braket{{{\text{out}}}|{\mathcal{S}}|{{\text{in}}}},\end{aligned}$$ where $\kappa = \sqrt{32\pi G}$. If the in- and out-states are eigenstates of $N(z,{{\bar{z}}})$ such that $$\begin{aligned} \bra{{{\text{out}}}}N(z,{{\bar{z}}}) &= N_{{\text{out}}}\bra{{{\text{out}}}} \qquad\text{and}\qquad N(z,{{\bar{z}}})\ket{{{\text{in}}}} = N_{{\text{in}}}\ket{{{\text{in}}}},\end{aligned}$$ then we obtain $$\begin{aligned} \label{refined_cons} \left(N_{{\text{out}}}- N_{{\text{in}}}\right)\braket{{{\text{out}}}|{\mathcal{S}}|{{\text{in}}}} = \Omega^\text{soft} \braket{{{\text{out}}}|{\mathcal{S}}|{{\text{in}}}},\end{aligned}$$ with a soft factor that is analogous to that of [@Kapec:2017tkm]: $$\begin{aligned} \Omega^\text{soft} = -\frac{\kappa^2}{8\pi}\sum_i \eta_i \frac{p_i^\mu p_i^\nu}{p_i\cdot \hat{x}_z}\epsilon^+_{{\mu\nu}}.\end{aligned}$$ To see what the eigenstates look like, we first note that $N(z,{{\bar{z}}})$ can be expressed in terms of the graviton creation and annihilation operators as [@He:2014laa] $$\begin{aligned} N(z,{{\bar{z}}}) = -\frac{\kappa}{8\pi}\lim_{\omega\to 0} \left[ \omega a_+(\omega \hat{x}_z) + \omega a_-^\dagger (\omega \hat{x}_z) \right ].\end{aligned}$$ This suggests that the eigenstate should take some form of a coherent graviton state. Next, consider the following state $$\begin{aligned} \label{Nstate} \ket{N}=\exp\left\{ \int{\widetilde{d^3 k}\,} N^{{\mu\nu}}(k) \left[a^\dagger_{{\mu\nu}}(k) - a_{{\mu\nu}}(k)\right] \right\}\ket{0},\end{aligned}$$ where ${\widetilde{d^3 k}\,} = \frac{d^3k}{(2\pi)^3 (2{{\omega_{\mathbf}{k}}})}$ is the Lorentz-invariant measure, $$\begin{aligned} a_{{\mu\nu}}^\dagger(k) = \sum_r \epsilon^r_{{\mu\nu}}(k) a^{r\dagger}(k), \qquad a_{{\mu\nu}}(k) = \sum_r \epsilon^{r*}_{{\mu\nu}}(k) a^{r}(k),\end{aligned}$$ $N^{\mu\nu}$ is an arbitrary symmetric tensor and the sum runs over all polarizations, including the unphysical ones. We will next show that if the symmetric tensor $N^{{\mu\nu}}(k)$ has soft poles, then the above state is an eigenstate of both $\lim \omega a_+$ and $\lim\omega a^\dagger_-$. Indeed, $$\begin{aligned} \lim_{\omega\to 0}\omega a_+(\omega \hat{x}_z)\ket{N} &= \lim_{\omega\to 0}\omega \left[ a_+(\omega \hat{x}_z), \int{\widetilde{d^3 k}\,} N^{{\mu\nu}}(k) \left(a^\dagger_{{\mu\nu}}(k) - a_{{\mu\nu}}(k)\right) \right] \ket{N} \\ &= \lim_{\omega\to 0}\frac{\omega}{2} N_{{\mu\nu}}(\omega \hat{x}_z) I^{{\mu\nu\rho\sigma}}\epsilon^+_{{\rho\sigma}}(z,{{\bar{z}}}) \ket{N} \\ &= \lim_{\omega\to 0}\omega N^{{\mu\nu}}(\omega \hat{x}_z) \epsilon^+_{{\mu\nu}}(z,{{\bar{z}}}) \ket{N}.\end{aligned}$$ Thus we see that the eigenvalue is non-zero only if $N^{{\mu\nu}}$ has poles for soft momenta. Similarly, $$\begin{aligned} \label{convac} \lim_{\omega\to 0}\omega a_-^\dagger(\omega \hat{x}_z)\ket{N} &= \lim_{\omega\to 0}\omega N^{{\mu\nu}}(\omega \hat{x}_z) \epsilon^+_{{\mu\nu}}(z,{{\bar{z}}}) \ket{N}.\end{aligned}$$ It should be noted that in , the term with the creation operator acting on the vacuum vanishes upon taking the soft limit $\omega \rightarrow 0$. From this we can immediately see that $\ket{N}$ is an eigenstate of $N(z,{{\bar{z}}})$, i.e., $$\begin{aligned} N(z,{{\bar{z}}}) \ket{N} = -\frac{\kappa}{4\pi} \left(\, \lim_{\omega\to 0}\omega N^{{\mu\nu}}\epsilon^+_{{\mu\nu}}\right)\ket{N}.\end{aligned}$$ In particular, the Fock vacuum $\ket{0}$ , which corresponds to $N^{{\mu\nu}}=0$, is itself an eigenstate with eigenvalue $0$. Later, when considering S matrix elements, we will for convenience put $N^{{\mu\nu}}=0$ for the incoming state, which amounts to assuming that the incoming state is a Fock state. This does not entail a loss of generality because as can be seen from , it is only the difference $N_{{\text{out}}}^{{\mu\nu}}-N_{{\text{in}}}^{{\mu\nu}}$ that matters. Similarly, the bra state $$\begin{aligned} \bra{N}=\bra{0}\exp\left[ -\int{\widetilde{d^3 k}\,} N^{{\mu\nu}}\left(a^\dagger_{{\mu\nu}}- a_{{\mu\nu}}\right) \right]\end{aligned}$$ is an eigenstate of $N(z,{{\bar{z}}})$: $$\begin{aligned} \bra{N}N(z,{{\bar{z}}}) = -\frac{\kappa}{4\pi}\bra{N} \left(\, \lim_{\omega\to 0}\omega N^{{\mu\nu}}\epsilon^+_{{\mu\nu}}\right).\end{aligned}$$ We want to treat these eigenstates as alternative vacuums, so we will restrict the momentum integrals to run over only the soft momenta. With these choices, $\ket{N}$ will remain an eigenstate with any number of hard particle operators acting on it. The conservation law $N_\text{out}-N_\text{in}=\Omega^\text{soft}$ implied by is then $$\begin{aligned} \lim_{\omega\to 0}\omega \big[N^{{\mu\nu}}_\text{out}(\omega\hat{x}_z) - N^{{\mu\nu}}_\text{in}(\omega\hat{x}_z)\big] \epsilon^+_{{\mu\nu}}(z,{{\bar{z}}}) &= \frac{\kappa}{2}\sum_i \eta_i\frac{p_i^\mu p_i^\nu }{p_i\cdot \hat{x}_z}\epsilon^+_{{\mu\nu}}(z,{{\bar{z}}}).\end{aligned}$$ As shown above, the leading soft terms in $N^{\mu\nu}$ are the only ones contributing to the eigenvalue, which therefore satisfy $$\begin{aligned} \label{eigen_cons} N^{{\mu\nu}}_\text{out}(k) - N^{{\mu\nu}}_\text{in}(k) = \frac{\kappa}{2}\sum_i \eta_i\frac{p_i^\mu p_i^\nu}{p_i\cdot k},\end{aligned}$$ where we have put $k = \omega \hat{x}_z$. We should emphasize that either this conservation law is satisfied or the amplitude ${\braket{\text{out}|{\mathcal{S}}|\text{in}}}$ vanishes. This implies that if the initial state is built on the Fock vacuum $\ket{0}$, i.e. $$\begin{aligned} \ket{\text{in}} = \prod_{i\in\text{in}}b^\dagger(p_i)\ket{0},\end{aligned}$$ where $b^\dagger$ is the creation operator of hard massive particles, then this state does not scatter into any state built on the same vacuum $\ket{0}$, since in that case $N_\text{out}=N_\text{in}=0$, thereby violating the conservation law $N_{{\text{out}}}- N_{{\text{in}}}= \Omega^\text{soft}$. Instead, scattering must take place into states built on the vacuum $\ket{N_{{\text{out}}}}$ with $$\begin{aligned} N^{{\mu\nu}}_\text{out}=\frac{\kappa}{2}\sum_i \eta_i\frac{p_i^\mu p_i^\nu}{p_i\cdot k}.\end{aligned}$$ Such states therefore have the form, (see Eq. ) $$\begin{aligned} \bra{\text{out}}=\bra{0}\left[\prod_{j\in\text{out}}b(p_j)\right] \exp\left[ -\frac{\kappa}{2}\sum_i\eta_i \int{\widetilde{d^3 k}\,} \frac{p_i^\mu p_i^\nu}{p_i\cdot k} (a^\dagger_{{\mu\nu}}- a_{{\mu\nu}}) \right].\end{aligned}$$ The scattering amplitude now can be written in the form: $$\begin{aligned} \label{amplitude} \bra{\text{out}} {\mathcal{S}}\ket{\text{in}} = \bra{{{\Psi_\text{out}}}} \exp\left[ -\frac{\kappa}{2}\sum_i\eta_i\int{\widetilde{d^3 k}\,} \frac{p_i^\mu p_i^\nu}{p_i\cdot k} (a^\dagger_{{\mu\nu}}- a_{{\mu\nu}}) \right] {\mathcal{S}}\ket{{{\Psi_\text{in}}}},\end{aligned}$$ where ${{\Psi_\text{out}}}$, ${{\Psi_\text{in}}}$ denote the usual Fock states for the hard particles. The form of is reminiscent of the Faddeev-Kulish amplitudes. In the following two sections we will spell out this equivalence more precisely. It will turn out that any amplitude that obeys the conservation law , an example being , is equal to the Faddeev-Kulish amplitude and is therefore IR-finite. Relation to Faddeev-Kulish amplitudes ===================================== As a first step in establishing this equality, we will demonstrate a crucial feature of the Faddeev-Kulish amplitudes which, although technical, has important physical consequences. Since a Faddeev-Kulish amplitude is constructed by dressing each external particle with its cloud of soft gravitons, an amplitude with $n$ incoming and $n'$ outgoing particles necessarily has $n$ clouds on the right of the scattering operator ${\mathcal{S}}$, and $n'$ clouds on the left. Although the clouds commute with each other, it was not clear how things change if, for example, one moves a cloud dressing an incoming particle (therefore sitting on the right of ${\mathcal{S}}$) to the left of ${\mathcal{S}}$. In this connection, based on the conservation of supertranslation charge and the crossing symmetry, the authors of [@Kapec:2017tkm] conjectured that such amplitudes exhibit the same cancellation of IR divergences. In this section, we will explicitly show that the clouds “weakly commute" with ${\mathcal{S}}$, in the sense that in an ${\mathcal{S}}$ matrix element, any incoming cloud can be moved to the outgoing state without affecting the amplitude, and vice versa. This result proves the aforementioned conjecture, since it follows that the amplitudes considered in [@Kapec:2017tkm] are equal to the Faddeev-Kulish amplitude. Then in the next section, we will use this to show that any amplitude that conserves supertranslation charge, for example , is equal to the Faddeev-Kulish amplitude with the same external particle configuration. This will establish the notion that Faddeev-Kulish amplitudes naturally arise from the charge conservation of asymptotic symmetries. In order to relate to the Faddeev-Kulish amplitude, let us denote its left hand side by ${\mathcal{M}}$, $$\begin{aligned} {\mathcal{M}}&= \bra{{{\Psi_\text{out}}}} \exp\left[ -\frac{\kappa}{2}\sum_i\eta_i\int{\widetilde{d^3 k}\,} \frac{p_i^\mu p_i^\nu}{p_i\cdot k} (a^\dagger_{{\mu\nu}}- a_{{\mu\nu}}) \right] {\mathcal{S}}\ket{{{\Psi_\text{in}}}}.\end{aligned}$$ Next, consider another amplitude ${\mathcal{M}}_c$, given by $$\begin{aligned} {\mathcal{M}}_c &= \bra{{{\Psi_\text{out}}}} \exp\left\{ -\frac{\kappa}{2}\sum_i\eta_i\int{\widetilde{d^3 k}\,} \left[\frac{p_i^\mu p_i^\nu}{p_i\cdot k} + \frac{c^{{\mu\nu}}(p_i,k)}{{{\omega_{\mathbf}{k}}}}\right] (a^\dagger_{{\mu\nu}}- a_{{\mu\nu}}) \right\} {\mathcal{S}}\ket{{{\Psi_\text{in}}}} \\ &= \label{new_amp} \bra{{{\Psi_\text{out}}}}\exp\left[-\sum_i \eta_i R_f(p_i)\right]{\mathcal{S}}\ket{{{\Psi_\text{in}}}},\end{aligned}$$ where we inserted a term proportional to $c^{{\mu\nu}}/{{\omega_{\mathbf}{k}}}$ to the argument of the exponential. Here $c^{{\mu\nu}}(p,k)$ is the tensor of [@Ware:2013zja] that parametrizes the asymptotic space, and $$\begin{aligned} \label{Rf} R_f(p_i) = \frac{\kappa}{2}\int {\widetilde{d^3 k}\,} \left[\frac{p_i^\mu p_i^\nu}{p_i\cdot k} + \frac{c^{{\mu\nu}}(p_i,k)}{{{\omega_{\mathbf}{k}}}}\right] (a^\dagger_{{\mu\nu}}- a_{{\mu\nu}})\end{aligned}$$ is the anti-Hermitian operator appearing in the construction of Faddeev-Kulish state [@Ware:2013zja] with $\phi = 1$. In contrast to ${\mathcal{M}}$ and ${\mathcal{M}}_c$, the IR-finite Faddeev-Kulish amplitude ${\mathcal{M}}_\text{FK}$ is given by $$\begin{aligned} \label{FKamp} {\mathcal{M}}_\text{FK} = \braket{{{\Psi_\text{out}}}|\exp\left[-\sum_{i\in{{\text{out}}}} R_f(p_i)\right] {\mathcal{S}}\exp\left[\sum_{i\in{{\text{in}}}} R_f(p_i)\right]|{{\Psi_\text{in}}}}.\end{aligned}$$ We aim to establish ${\mathcal{M}}_{FK}={\mathcal{M}}_c={\mathcal{M}}$. Moving the graviton clouds {#MOVECLOUD} -------------------------- Let us start by considering the simplest case, i.e., the Faddeev-Kulish amplitude for single-particle external states to leading order in the interaction. We follow the shorthand notations used in [@Choi:2017bna]: $$\begin{aligned} P_{{\mu\nu}}(p,k) = \frac{\kappa}{2}\left(\frac{p_\mu p_\nu}{p\cdot k}\right), \qquad C_{{\mu\nu}}(p,k) = \frac{\kappa}{2}\frac{c_{{\mu\nu}}(p,k)}{{{\omega_{\mathbf}{k}}}},\end{aligned}$$ and $S_{{\mu\nu}}(p,k) = P_{{\mu\nu}}(p,k)+C_{{\mu\nu}}(p,k)$. These allow us to write, (see [@Choi:2017bna] for details) $$\begin{aligned} {\mathcal{M}}_\text{FK} = \bra{0} b(p_f) e^{ -S_f \cdot(a^\dagger - a) } {\mathcal{S}}e^{ S_i \cdot(a^\dagger - a) } b^\dagger(p_i) \ket{0},\end{aligned}$$ where $S_f^{{\mu\nu}}\equiv S^{{\mu\nu}}(p_f,k)$ and $S_i^{{\mu\nu}}\equiv S^{{\mu\nu}}(p_i,k)$. The subscript FK is written to emphasize that this is a Faddeev-Kulish amplitude. In what follows we will employ the following notation, $$\begin{aligned} S\cdot (a^\dagger-a) \equiv \int{\widetilde{d^3 k}\,}S^{{\mu\nu}}(a^\dagger_{{\mu\nu}}- a_{{\mu\nu}}),\end{aligned}$$ and $$\begin{aligned} S_f\cdot I \cdot S_i \equiv \int{\widetilde{d^3 k}\,} S_f^{{\mu\nu}}I_{{\mu\nu\rho\sigma}}S_i^{{\rho\sigma}},\end{aligned}$$ where $$\begin{aligned} I^{{\mu\nu\rho\sigma}}=\eta^{\mu\rho}\eta^{\nu\sigma}+\eta^{\mu\sigma}\eta^{\nu\rho}-\eta^{{{\mu\nu}}}\eta^{{{\rho\sigma}}}.\end{aligned}$$ Upto the one loop order, this amplitude is $$\begin{aligned} {\mathcal{M}}_\text{FK} = \bra{0} b(p_f) \left( 1+S_f \cdot a - \frac{1}{4}S_f\cdot I\cdot S_f \right) {\mathcal{S}}\left( 1+S_i \cdot a^\dagger - \frac{1}{4}S_i \cdot I\cdot S_i \right) b^\dagger(p_i) \ket{0}.\end{aligned}$$ Working out the infrared divergences (see [@Choi:2017bna] for details), we see that they factor out and cancel as $$\begin{aligned} \Bigg( 1 -\underbrace{\frac{1}{4}P\cdot I \cdot P}_\text{virtual} +\underbrace{\frac{1}{2}S\cdot I \cdot P}_\text{interacting} -\underbrace{\frac{1}{4}S\cdot I \cdot S}_\text{cloud-to-cloud} \Bigg)\braket{p_f|{\mathcal{S}}|p_i} = \braket{p_f|{\mathcal{S}}|p_i},\end{aligned}$$ where $P=P_f-P_i$ and $S=S_f-S_i$. Note that the various infrared divergent contributions are indicated in braces. These are (1) corrections due to virtual graviton exchange, (2) the interacting graviton corrections arising from gravitons connecting the Faddeev-Kulish clouds to external legs, and finally (2) corrections due to cloud-to-cloud graviton exchanges. These have been discussed in detail in appendix B of [@Choi:2017bna]. [0.24]{} ![Diagrams (a)-(d) represent processes with Faddeev-Kulish asymptotic states. Diagrams (e)-(h) represent the same processes with the incoming cloud moved to the outgoing state. Notice the “wrong" sign $+R_f$ compared to a normal outgoing cloud with $-R_f$.[]{data-label="movecloud"}](out_Pf "fig:"){width="\textwidth"} [0.24]{} ![Diagrams (a)-(d) represent processes with Faddeev-Kulish asymptotic states. Diagrams (e)-(h) represent the same processes with the incoming cloud moved to the outgoing state. Notice the “wrong" sign $+R_f$ compared to a normal outgoing cloud with $-R_f$.[]{data-label="movecloud"}](out_pi "fig:"){width="\textwidth"} [0.24]{} ![Diagrams (a)-(d) represent processes with Faddeev-Kulish asymptotic states. Diagrams (e)-(h) represent the same processes with the incoming cloud moved to the outgoing state. Notice the “wrong" sign $+R_f$ compared to a normal outgoing cloud with $-R_f$.[]{data-label="movecloud"}](in_Pf "fig:"){width="\textwidth"} [0.24]{} ![Diagrams (a)-(d) represent processes with Faddeev-Kulish asymptotic states. Diagrams (e)-(h) represent the same processes with the incoming cloud moved to the outgoing state. Notice the “wrong" sign $+R_f$ compared to a normal outgoing cloud with $-R_f$.[]{data-label="movecloud"}](in_pi "fig:"){width="\textwidth"} [0.24]{} ![Diagrams (a)-(d) represent processes with Faddeev-Kulish asymptotic states. Diagrams (e)-(h) represent the same processes with the incoming cloud moved to the outgoing state. Notice the “wrong" sign $+R_f$ compared to a normal outgoing cloud with $-R_f$.[]{data-label="movecloud"}](out_Pf2 "fig:"){width="\textwidth"} [0.24]{} ![Diagrams (a)-(d) represent processes with Faddeev-Kulish asymptotic states. Diagrams (e)-(h) represent the same processes with the incoming cloud moved to the outgoing state. Notice the “wrong" sign $+R_f$ compared to a normal outgoing cloud with $-R_f$.[]{data-label="movecloud"}](out_pi2 "fig:"){width="\textwidth"} [0.24]{} ![Diagrams (a)-(d) represent processes with Faddeev-Kulish asymptotic states. Diagrams (e)-(h) represent the same processes with the incoming cloud moved to the outgoing state. Notice the “wrong" sign $+R_f$ compared to a normal outgoing cloud with $-R_f$.[]{data-label="movecloud"}](in_Pf2 "fig:"){width="\textwidth"} [0.24]{} ![Diagrams (a)-(d) represent processes with Faddeev-Kulish asymptotic states. Diagrams (e)-(h) represent the same processes with the incoming cloud moved to the outgoing state. Notice the “wrong" sign $+R_f$ compared to a normal outgoing cloud with $-R_f$.[]{data-label="movecloud"}](in_pi2 "fig:"){width="\textwidth"} Now let us see what happens if we put all the clouds in the outgoing state. We will denote this amplitude as, $$\begin{aligned} {\mathcal{M}}_c = \bra{0} b(p_f) e^{ -S_f \cdot(a^\dagger - a) } e^{ S_i \cdot(a^\dagger - a) } {\mathcal{S}}b^\dagger(p_i) \ket{0}.\end{aligned}$$ Let us consider the various infrared divergent contributions in this case. The virtual graviton contribution remains unchanged. For the interacting gravitons, it used to be that the graviton contractions with a cloud gives the factor $$\begin{aligned} \frac{1}{2}\int{\widetilde{d^3 k}\,}S^{{\mu\nu}}I_{{\mu\nu\rho\sigma}},\end{aligned}$$ and depending on whether it was an incoming or an outgoing cloud, the contraction became $$\begin{aligned} +\frac{\eta}{2}\int{\widetilde{d^3 k}\,}S^{{\mu\nu}}_f I_{{\mu\nu\rho\sigma}}P^{{\rho\sigma}}\quad&\text{for outgoing cloud (Figures \ref{movecloud}(a),(b)), and}\\ -\frac{\eta}{2}\int{\widetilde{d^3 k}\,}S^{{\mu\nu}}_i I_{{\mu\nu\rho\sigma}}P^{{\rho\sigma}}\quad&\text{for incoming cloud (Figures \ref{movecloud}(c),(d)),}\end{aligned}$$ due to the difference in the sign of soft factor for absorption and emission. Figures \[movecloud\](a) and \[movecloud\](c) have $\eta=+1$, while \[movecloud\](b) and \[movecloud\](d) have $\eta=-1$. But now, we have two clouds that are in the outgoing state, so the graviton contraction gives the factor $$\begin{aligned} +\frac{1}{2}\int{\widetilde{d^3 k}\,}S_f^{{\mu\nu}}I_{{\mu\nu\rho\sigma}}\quad&\text{for the $p_f$ cloud, and}\\ -\frac{1}{2}\int{\widetilde{d^3 k}\,}S_i^{{\mu\nu}}I_{{\mu\nu\rho\sigma}}\quad&\text{for the $p_i$ cloud,}\end{aligned}$$ due to the difference in the signs of $R_f$. Since both are outgoing clouds, we have the same sign for the soft factor, $$\begin{aligned} +\frac{\eta}{2}\int{\widetilde{d^3 k}\,}S^{{\mu\nu}}_f I_{{\mu\nu\rho\sigma}}P^{{\rho\sigma}}\quad&\text{for the $p_f$ cloud (Figures \ref{movecloud}(e),(f)), and}\\ -\frac{\eta}{2}\int{\widetilde{d^3 k}\,}S^{{\mu\nu}}_i I_{{\mu\nu\rho\sigma}}P^{{\rho\sigma}}\quad&\text{for the $p_i$ cloud (Figures \ref{movecloud}(g),(h)),}\end{aligned}$$ where Figures \[movecloud\](e) and \[movecloud\](g) have $\eta=+1$, while \[movecloud\](f) and \[movecloud\](h) have $\eta=-1$. One can see that the results stay the same, meaning that contributions of interacting gravitons are unaltered. It remains to check the cloud-to-cloud contributions, but since these arise from contractions between operators in the clouds, they do not depend on which side of ${\mathcal{S}}$ the cloud is located and therefore are unchanged. We have thus shown that the infrared divergent part of the single-particle, leading order amplitudes ${\mathcal{M}}_c $ and ${\mathcal{M}}_\text{FK}$ remains unchanged upon shifting the cloud around, i.e. between the in and out states. Next, we will generalize this result to the most general case of multiple external particles and all loop orders. Again, we begin by considering the individual contributions, i.e., virtual, interacting, and cloud-to-cloud gravitons. The virtual graviton contribution is unchanged from the one given in [@Choi:2017bna]. For the interacting gravitons, consider the amplitude of a diagram with $N$ ($N'$) absorbed (emitted) interacting gravitons, $$\begin{aligned} \label{multint} \begin{split} (-1)^N& \left[ \prod_{r=1}^{N+N'}\frac{1}{2}\int{\widetilde{d^3 k_r}\,}S_{{\mu\nu}}(p_r,k_r) I^{{{\mu\nu}}\rho_r \sigma_r} \right] {\mathcal{J}}_{\rho_1\sigma_1\rho_2\sigma_2\cdots\rho_{N+N'}\sigma_{N+N'}}, \end{split}\end{aligned}$$ where $p_r$ is the momentum of the external particle that exchanges graviton $r$, and ${\mathcal{J}}$ is a complicated tensor whose detailed form is given in equation (B.58) of [@Choi:2017bna]. Taking the $k$-th incoming cloud and moving it to the outgoing state will have the two following effects (see Figure \[factors\]): ![An example of an incoming cloud being moved to the out-state. Each boson connecting this cloud to an external propagator obtains two factors of $(-1)$, one from the soft factor and the other from the “wrong" sign of $R_f$. These two factors cancel, and thus the overall amplitude is unaffected by such a change.[]{data-label="factors"}](factors){width=".95\textwidth"} 1. The factor $(-1)^N$, which comes from the signs in the soft factors, will become $(-1)^{N-n_k}$, where $n_k$ is the number of interacting gravitons connected to the $k$-th (previously) incoming cloud. This is because these gravitons used to be absorbed but are now emitted. 2. The following factor in , $$\begin{aligned} \left[ \prod_{r=1}^{N+N'}\frac{1}{2}\int{\widetilde{d^3 k_r}\,}S_{{\mu\nu}}(p_r,k_r) I^{{{\mu\nu}}\rho_r\sigma_r} \right], \end{aligned}$$ came from contractions of gravitons with the clouds. For the Faddeev-Kulish amplitude, where all clouds are in the proper locations, each cloud gives the same factor $\frac{1}{2}\int{\widetilde{d^3 k}\,}S^{{\mu\nu}}I_{{\mu\nu\rho\sigma}}$ upon contraction. But now that the $k$-th incoming cloud is sitting in the outgoing state with a wrong sign (incoming and outgoing clouds have different signs $e^{\pm R_f}$), only this cloud gives an additional factor of $-1$. The above factor changes to $$\begin{aligned} \left[ (-1)^{n_k}\prod_{r=1}^{N+N'}\frac{1}{2}\int{\widetilde{d^3 k_r}\,}S_{{\mu\nu}}(p_r,k_r) I^{{{\mu\nu}}\rho_r\sigma_r} \right]. \end{aligned}$$ It follows that we obtain two factors $(-1)^{-n_k}$ and $(-1)^{n_k}$, which cancel each other and the overall contribution remains unchanged. It remains to consider the cloud-to-cloud gravitons. There are three types: out-to-out, in-to-in, and the disconnected. The contributions of $l$ disconnected gravitons factored out as $$\begin{aligned} l!\left[\frac{1}{2} S^\text{out}\cdot I\cdot S^\text{in}\right]^l,\end{aligned}$$ but with the $k$-th incoming cloud moved to the out-state (as an outgoing cloud with the wrong sign), this is adjusted to $$\begin{aligned} l!\left[\frac{1}{2} (S^\text{out}-S^k)\cdot I\cdot (S^\text{in} - S^k) \right]^l,\end{aligned}$$ which eventually exponentiates to $$\begin{aligned} \label{cloud1} \exp\left\{ \frac{1}{2}(S^\text{out}-S^k)\cdot I\cdot (S^\text{in} - S^k) \right\}.\end{aligned}$$ The in-to-in and out-to-out contributions change from $$\begin{aligned} \exp\left\{ -\frac{1}{4}S^\text{out}\cdot I\cdot S^\text{out} -\frac{1}{4}S^\text{in}\cdot I\cdot S^\text{in} \right\}\end{aligned}$$ to $$\begin{aligned} \label{cloud2} \exp\left\{ -\frac{1}{4}(S^\text{out}-S^k)\cdot I\cdot (S^\text{out}-S^k) -\frac{1}{4}(S^\text{in}-S^k)\cdot I\cdot (S^\text{in}-S^k) \right\}.\end{aligned}$$ Putting and together, we obtain $$\begin{aligned} \exp\left\{ -\frac{1}{4}(S^\text{out}-S^\text{in})\cdot I \cdot (S^\text{out}-S^\text{in}) \right\},\end{aligned}$$ which is the same factor that was obtained without moving the cloud, and thus the cloud-to-cloud contribution also remains unaltered. It follows that we can write $$\begin{aligned} &\bra{{{\Psi_\text{out}}}} \left[\prod_{j\in\text{out}}e^{-S_j\cdot (a^\dagger-a)}\right] {\mathcal{S}}\left[\prod_{i\in\text{in}}e^{S_i\cdot (a^\dagger-a)}\right] \ket{{{\Psi_\text{in}}}} \\ &= \bra{{{\Psi_\text{out}}}} \left[\prod_{j\in\text{out}}e^{-S_j\cdot (a^\dagger-a)}\right] \left[\prod_{i\in\text{in}}e^{S_i\cdot (a^\dagger-a)}\right] {\mathcal{S}}\ket{{{\Psi_\text{in}}}} \\ &= \bra{{{\Psi_\text{out}}}} {\mathcal{S}}\left[\prod_{j\in\text{out}}e^{-S_j\cdot (a^\dagger-a)}\right] \left[\prod_{i\in\text{in}}e^{S_i\cdot (a^\dagger-a)}\right] \ket{{{\Psi_\text{in}}}},\end{aligned}$$ and so on. Therefore, we conclude that the Faddeev-Kulish amplitude does not change under a shift of the cloud from one side of the scattering operator to the other. Equality of the amplitudes {#SAME} -------------------------- From and , it is clear that the only difference between ${\mathcal{M}}_c$ and ${\mathcal{M}}_\text{FK}$ is in the location of the clouds; the incoming cloud, which should be dressing the incoming state, is located in the out-state. We have seen in the previous subsection that in an amplitude the clouds can freely be commuted through the scattering operator. This implies that the amplitude ${\mathcal{M}}_c$, which has all the clouds in the outgoing state, is actually equal to the Faddeev-Kulish amplitude, i.e. $$\begin{aligned} {\mathcal{M}}_c = {\mathcal{M}}_\text{FK}.\end{aligned}$$ Now let us consider the original amplitude ${\mathcal{M}}$ of that emerged from the conservation of supertranslation charge. This is a special case of ${\mathcal{M}}_c$, in the sense that putting $c^{{\mu\nu}}= 0$ in ${\mathcal{M}}_c$ recovers ${\mathcal{M}}$. Thus, ${\mathcal{M}}$ is equal to the Faddeev-Kulish amplitude constructed using the $R(t)$ operator of [@Ware:2013zja] instead of $R_f$. Since states constructed with $R(t)$ and $R_f$ are related by a unitary transformation, this implies that ${\mathcal{M}}= {\mathcal{M}}_c$. We can see this more directly by noting that amplitudes constructed with $c^{{\mu\nu}}=0$ are related to those with non-zero $c^{{\mu\nu}}$ by the following relation [@Choi:2017bna] $$\begin{aligned} {\mathcal{M}}_c &= \exp\left[-\frac{\kappa^2}{4}\sum_{n,m}\eta_n\eta_m \int\frac{{\widetilde{d^3 k}\,}}{\omega_{\mathbf}{k}^2} c_{{\mu\nu}}(p_n,k)I^{{\mu\nu\rho\sigma}}c_{{\rho\sigma}}(p_m,k)\right]{\mathcal{M}}= {\mathcal{M}},\end{aligned}$$ where each sum runs over the whole set of external particles. The summand vanishes term by term, due to one of the constraints that $c^{{\mu\nu}}$ has to satisfy. Therefore ${\mathcal{M}}= {\mathcal{M}}_c = {\mathcal{M}}_\text{FK}$, and the amplitude ${\mathcal{M}}$ of is the IR-finite Faddeev-Kulish amplitude. Soft gravitons and decoherence of momentum configurations of hard matter particles {#DECOHERENCE} ================================================================================== In this section we will reconsider the problem of the decoherence of momentum superpositions of hard matter particles due to low energy soft gravitons that was discussed in [@Carney:2017jut]. The same conclusions were reached in [@Strominger:2017aeh] using a different approach. In [@Carney:2017jut] the usual Bloch-Nordsieck mechanism was introduced to cancel the infrared divergences in order to obtain finite density matrices, and the asymptotic symmetries discussed in section \[EIGENSTATE\] do not play any role. The question we address in this section is how a consistent application of the results of the previous sections might change the conclusions of [@Carney:2017jut]. First, we will briefly outline the logic of [@Carney:2017jut]. Consider an “in" Fock state $\ket{\alpha}_{{\text{in}}}$ at time $t=-\infty$, which is related to the “out" Fock state at $t=\infty$ by the S matrix: $$\begin{aligned} \ket{\alpha} \quad\to\quad \ket{\alpha}_{{\text{in}}}&={\mathcal{S}}\ket{\alpha}_{{\text{out}}}\\ &= \left(\sum_{\beta b}\ket{\beta b}\bra{\beta b}\right) {\mathcal{S}}\ket{\alpha}_{{\text{out}}}\\ &= \sum_{\beta b}S_{\beta b,\alpha}\ket{\beta b}_{{\text{out}}},\end{aligned}$$ where, $S_{\beta b,\alpha} \equiv \braket{\beta b|{\mathcal{S}}|\alpha}$, and $\beta$ ($b$) stands for the set of hard (soft) particles. We will drop subscripts on the kets which, unless specified, will always be the asymptotic out-states. Then the authors construct a reduced density matrix by tracing out the external soft bosons $\ket{b}$: $$\begin{aligned} \label{reduced_dm} \rho = \sum_{\beta\beta' b}S_{\beta b,\alpha}S^*_{\beta' b,\alpha}\ket{\beta}\bra{\beta'}.\end{aligned}$$ By factoring out the divergences from the sum, $$\begin{aligned} \label{oldform} \sum_b S_{\beta b,\alpha}S^*_{\beta' b,\alpha} &= S_{\beta,\alpha}S^*_{\beta',\alpha} \underbrace{ \left(\frac{E}{\lambda}\right)^{\tilde{A}_{\beta\beta',\alpha}} \left(\frac{E}{\lambda}\right)^{\tilde{B}_{\beta\beta',\alpha}} f\left(\frac{E}{E_T},\tilde{A}_{\beta\beta',\alpha}\right) f\left(\frac{E}{E_T},\tilde{B}_{\beta\beta',\alpha}\right) }_\text{real soft bosons} \\ &= S^\Lambda_{\beta,\alpha}S^{\Lambda*}_{\beta',\alpha} \underbrace{ \left(\frac{\lambda}{\Lambda}\right)^{A_{\beta,\alpha}/2+A_{\beta',\alpha}/2} \left(\frac{\lambda}{\Lambda}\right)^{B_{\beta,\alpha}/2+B_{\beta',\alpha}/2} }_\text{virtual bosons} \\ &\qquad\times \underbrace{ \left(\frac{E}{\lambda}\right)^{\tilde{A}_{\beta\beta',\alpha}} \left(\frac{E}{\lambda}\right)^{\tilde{B}_{\beta\beta',\alpha}} f\left(\frac{E}{E_T},\tilde{A}_{\beta\beta',\alpha}\right) f\left(\frac{E}{E_T},\tilde{B}_{\beta\beta',\alpha}\right) }_\text{real soft bosons},\end{aligned}$$ and by considering the limit as the IR cut-off $\lambda$ is removed, the authors of [@Carney:2017jut] observed the decoherence of momentum configurations of hard particles or conversely, the strong correlations between the hard and soft particles. We will refer to [@Carney:2017jut] for details of the notations and derivations of this equation. However, note that it is essential in this approach to sum over the soft bosons because otherwise the infrared divergences will not cancel. We now show that this conclusion implicitly assumes that the vacuum is unique and before the cancellation of IR divergences for the inclusive process, one is dealing with S matrix elements which vanish as the cut-off is removed. We have seen that conservation of BMS charge, namely $$\begin{aligned} \left(N_{{\text{out}}}- N_{{\text{in}}}\right)\braket{{{\text{out}}}|{\mathcal{S}}|{{\text{in}}}} = \Omega^\text{soft}\braket{{{\text{out}}}|{\mathcal{S}}|{{\text{in}}}},\end{aligned}$$ dictates that scattering processes starting from a state built on the Fock vacuum $\ket{0}$ evolves only into states that are built on the coherent vacuum $$\begin{aligned} \exp\left[\int_\text{soft}{\widetilde{d^3 k}\,}N_{{\text{out}}}^{{\mu\nu}}(a^\dagger_{{\mu\nu}}- a_{{\mu\nu}})\right] \ket{0},\end{aligned}$$ where, $$\begin{aligned} N_{{\text{out}}}^{{\mu\nu}}= \frac{\kappa}{2}\sum_{i}\eta_i\frac{p_i^\mu p_i^\nu}{p_i\cdot k},\end{aligned}$$ with the sum running over all external particles[^1]. Therefore, if we started with a state $\ket{\alpha}$ built on $\ket{0}$, then the outgoing state cannot be just $\ket{\beta b}$, which is a Fock state built on $\ket{0}$; all S-matrix elements between such states will vanish. Instead, $\ket{\alpha}$ will scatter into states accompanied by a coherent cloud, $$\begin{aligned} \ket{\beta;N_{{\text{out}}}}=\ket{\beta}\exp\left[\int_\text{soft}{\widetilde{d^3 k}\,}N_{{\text{out}}}^{{\mu\nu}}(a^\dagger_{{\mu\nu}}- a_{{\mu\nu}})\right],\end{aligned}$$ with $N_{{\text{out}}}$ dependent on the sets of external hard momenta $\alpha$ and $\beta$. We therefore should consider, $$\begin{aligned} \ket{\alpha}_{{\text{in}}}&= \sum_{\beta} S^\text{FK}_{\beta,\alpha}\ket{\beta;N_{{\text{out}}}},\end{aligned}$$ where we have written, $$\begin{aligned} S_{\beta,\alpha}^\text{FK} \equiv \bra{\beta} \exp\left[-\int_\text{soft}{\widetilde{d^3 k}\,}N^{{\mu\nu}}_{{\text{out}}}(a^\dagger_{{\mu\nu}}-a_{{\mu\nu}})\right] {\mathcal{S}}\ket{\alpha}.\end{aligned}$$ The states $\ket{\alpha}$ and $\ket{\beta}$ are just the conventional Fock states. We have seen earlier that the right hand side is exactly equivalent to the amplitude constructed using the Faddeev-Kulish asymptotic states, i.e., $$\begin{aligned} \bra{\beta}e^{-R_f}{\mathcal{S}}e^{R_f}\ket{\alpha},\end{aligned}$$ and hence the the left hand side has the superscript $\text{FK}$ on the S matrix element. Now the density matrix becomes $$\begin{aligned} \label{newrho} \sum_{\beta \beta'}S^{\text{FK}}_{\beta,\alpha}S^{\text{FK}*}_{\beta',\alpha} \ket{\beta;N_{{\text{out}}}}\bra{\beta';N_{{\text{out}}}'}.\end{aligned}$$ The amplitudes $S^{\text{FK}}_{\beta,\alpha}$ do not have infrared divergences coming from the virtual bosons. In the “virtual bosons" part of , the $\lambda$-dependent part is exactly canceled by interactions involving the clouds, as seen in [@Choi:2017bna]. Thus, in this framework there is no longer the decoherence that was observed in [@Carney:2017jut]. To sum up, due to the conservation of BMS charge, any conventional Fock state $\ket{\alpha}$ evolves not into another Fock state $\ket{\beta b}$, but instead into a coherent state $\ket{\beta;N_{{\text{out}}}}$. If the starting state is a coherent state, then the end state will just be another coherent state, and the BMS charge conservation will guarantee that the amplitudes $S^\text{FK}_{\beta,\alpha}$ coincide with the infrared-finite Faddeev-Kulish amplitudes. We reiterate, that the presence of the coherent boson cloud cancels all the problematic dependence on the infrared cut-off $\lambda$, and therefore one is no longer mathematically forced to sum over the soft particles in order to obtain well-defined density matrix elements. It is noteworthy that although the density matrix elements are now well-defined, depending on what kind of measurement is being carried out, one may still construct a reduced density matrix by summing over the soft particles. Would the decoherence of the momentum configurations of the hard matter particles return in this case? This analysis has recently been carried out in [@Carney:2017oxp]. We will next reanalyze this within the framework introduced in the previous sections of this paper. The $\beta \beta'$-component of the reduced density matrix is $$\begin{aligned} \rho_{\beta\beta'} &= \sum_b S^{\text{FK}}_{\beta,\alpha}S^{\text{FK}*}_{\beta',\alpha} \braket{b|N_{{\text{out}}}}\braket{N_{{\text{out}}}'|b} \\&= S^{\text{FK}}_{\beta,\alpha}S^{\text{FK}*}_{\beta',\alpha} \bra{N_{{\text{out}}}'}\left(\sum_b\ket{b}\bra{b}\right)\ket{N_{{\text{out}}}} \\&= S^{\text{FK}}_{\beta,\alpha}S^{\text{FK}*}_{\beta',\alpha} \braket{N_{{\text{out}}}'|N_{{\text{out}}}}.\end{aligned}$$ By normal-ordering the graviton operators, we obtain $$\begin{aligned} \braket{N_{{\text{out}}}'|N_{{\text{out}}}} &= \bra{0}\exp\left \{ \frac{\kappa}{2}\int_\text{soft}{\widetilde{d^3 k}\,} \left( N_{{\text{out}}}^{{\mu\nu}}- N'^{{\mu\nu}}_{{\text{out}}}\right) (a_{{\mu\nu}}^\dagger - a_{{\mu\nu}}) \right \}\ket{0} \\ \label{vanishing_W2} &= \exp\left \{ -\frac{\kappa^2}{16}\int_\text{soft}{\widetilde{d^3 k}\,} \left( N_{{\text{out}}}^{{\mu\nu}}- N'^{{\mu\nu}}_{{\text{out}}}\right) I_{{\mu\nu\rho\sigma}}\left( N_{{\text{out}}}^{{\rho\sigma}}- N'^{{\rho\sigma}}_{{\text{out}}}\right) \right \},\end{aligned}$$ where we can write $$\begin{aligned} N^{{\mu\nu}}_{{\text{out}}}- N'^{{\mu\nu}}_{{\text{out}}}= \sum_{p\in\beta}\frac{p^\mu p^\nu}{p\cdot k} - \sum_{p\in\beta'}\frac{p^\mu p^\nu}{p\cdot k}.\end{aligned}$$ Therefore, if $\beta \neq \beta'$ then the integral in is infrared-divergent and the expression vanishes. This implies that the off-diagonal elements of the reduced density matrix is zero and the decoherence of momentum configurations of the hard particles reappears. Does this conclusion change if we include external states with soft gravitons? The density matrix with external soft gravitons is $$\begin{aligned} \sum_{\beta\beta'bb'}S^\text{FK}_{\beta b,\alpha}S^{\text{FK}*}_{\beta' b',\alpha} \ket{\beta b;N_{{\text{out}}}}\bra{\beta' b';N'_{{\text{out}}}},\end{aligned}$$ and the reduced density matrix, after tracing out the soft particles, becomes $$\begin{aligned} \rho_{\beta\beta'} &= \sum_{b''}\sum_{b b'} S^\text{FK}_{\beta b,\alpha}S^{\text{FK}*}_{\beta' b',\alpha} \braket{b''|b;N_{{\text{out}}}}\braket{ b';N'_{{\text{out}}}|b''} \\ &= \sum_{b b'} S^\text{FK}_{\beta b,\alpha}S^{\text{FK}*}_{\beta' b',\alpha} \braket{b';N'_{{\text{out}}}| b;N_{{\text{out}}}}.\end{aligned}$$ Let us employ a notation similar to that of [@Carney:2017oxp]: $$\begin{aligned} W(\beta) &= \exp\left\{ \frac{\kappa}{2}\int_\text{soft}{\widetilde{d^3 k}\,} \sum_{p\in\beta}\frac{p^\mu p^\nu}{p\cdot k}(a^\dagger_{{\mu\nu}}-a_{{\mu\nu}}) \right\}, \\ W^\dagger(\beta') &= \exp\left\{ -\frac{\kappa}{2}\int_\text{soft}{\widetilde{d^3 k}\,} \sum_{p\in\beta'}\frac{p^\mu p^\nu}{p\cdot k}(a^\dagger_{{\mu\nu}}-a_{{\mu\nu}}) \right\},\end{aligned}$$ such that $\ket{b;N_{{\text{out}}}}=W(\beta)\ket{b}$. Then, the reduced density matrix element is $$\begin{aligned} \label{rdm_ext_soft} \rho_{\beta\beta'}= \sum_{b b'} S^\text{FK}_{\beta b,\alpha}S^{\text{FK}*}_{\beta' b',\alpha} \bra{b'}{{W^\dagger(\beta')W(\beta)}}\ket{b}.\end{aligned}$$ Let us see what we can say about $\braket{b'|{{W^\dagger(\beta')W(\beta)}}|b}$. Let $m$ and $n$ be the particle number of $b'$ and $b$, respectively. Then, $$\begin{aligned} \braket{b'|{{W^\dagger(\beta')W(\beta)}}|b} = & \bra{0} a_{\ell'_1}(k'_1)\cdots a_{\ell'_m}(k'_m) W^\dagger(\beta')W(\beta) a^\dagger_{\ell_1}(k_1)\cdots a^\dagger_{\ell_n}(k_n) \ket{0},\end{aligned}$$ where $\ell'_i$ and $k'_i$ ($\ell_i$ and $k_i$) are the polarization and momentum of the $i$-th graviton in $b'$ ($b$). Let us use the shorthand $$\begin{aligned} W^2 \equiv {{W^\dagger(\beta')W(\beta)}},\end{aligned}$$ and observe that since $$\begin{aligned} a_\ell(k) &= \epsilon^{{{\mu\nu}}}_\ell(k) a_{{\mu\nu}}(k), \\ a^\dagger_\ell(k) &= \epsilon^{{{\mu\nu}}*}_\ell(k) a^\dagger_{{\mu\nu}}(k),\end{aligned}$$ we have the commutators $$\begin{aligned} \left[W^2,a^\dagger_{\ell}(k)\right] &= -\frac{\kappa}{2}\int_\text{soft}{\widetilde{d^3 k'}\,}\sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k'} \left[a^\dagger_{{\mu\nu}}(k')-a_{{\mu\nu}}(k'),a^\dagger_\ell(k)\right] W^2 \\ &= +\frac{\kappa}{2}\sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k} \epsilon^{\ell*}_{{\mu\nu}}(k) W^2, \\ \left[a_{\ell}(k),W^2\right] &= -\frac{\kappa}{2}\int_\text{soft}{\widetilde{d^3 k'}\,}\sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k'} \left[a_\ell(k),a^\dagger_{{\mu\nu}}(k')-a_{{\mu\nu}}(k')\right] W^2 \\ &= -\frac{\kappa}{2}\sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k} \epsilon^{\ell}_{{\mu\nu}}(k) W^2,\end{aligned}$$ where $\eta_p=+1$ if $p\in\beta'$ and $\eta_p=-1$ if $p\in\beta$. Using this, we can commute the left-most creation operator $a^\dagger_{\ell_1}(k_1)$ to the left side of $W^2$ to obtain $$\begin{aligned} \nonumber \braket{b'|W^2|b} &= \left[\frac{\kappa}{2} \sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k_1} \epsilon^{\ell_1*}_{{\mu\nu}}(k_1) \right] \bra{0} a_{\ell'_1}(k'_1)\cdots a_{\ell'_m}(k'_m) W^2 a^\dagger_{\ell_2}(k_2)\cdots a^\dagger_{\ell_n}(k_n) \ket{0} \\&\qquad +\bra{0} a_{\ell'_1}(k'_1)\cdots a_{\ell'_m}(k'_m) a^\dagger_{\ell_1}(k_1) W^2 a^\dagger_{\ell_2}(k_2)\cdots a^\dagger_{\ell_n}(k_n) \ket{0}.\end{aligned}$$ However, we will next show that the contribution from the second term in the parentheses is vanishingly small. To see this, one may consider commuting $a^\dagger_{\ell_1}(k_1)$ all the way to the left, aiming to act it on the vacuum. This will create one term for each annihilation operator which has a factor of the following form, $$\begin{aligned} \nonumber &\int_\text{soft}{\widetilde{d^3 k'_j}\,}{\widetilde{d^3 k_1}\,} S^\text{FK}_{\beta b,\alpha}S^{\text{FK}*}_{\beta' b',\alpha} \left[a_{\ell'_j}(k'_j),a^\dagger_{\ell_1}(k_1)\right] \\\label{momvol} &= \delta_{\ell'_j,\ell_1} \int d\Omega'_jd\Omega_1 \delta^2(\Omega'_j-\Omega_1) S^\text{FK}_{\beta b,\alpha}S^{\text{FK}*}_{\beta' b',\alpha} \int_\text{soft} \frac{|{\mathbf}{k}'_j|^2 d|{\mathbf}{k}'_j|}{(2\pi)^3|{\mathbf}{k}'_j|},\end{aligned}$$ where we have separated out the radial parts from the momentum integrals. ![Diagrams contributing to an amplitude with external soft boson. The first two diagrams cancel the last two diagrams, and only the diagram in the middle remains, which is of zeroth order in the soft momentum.[]{data-label="singlegrav"}](multiin "fig:"){width=".19\textwidth"} ![Diagrams contributing to an amplitude with external soft boson. The first two diagrams cancel the last two diagrams, and only the diagram in the middle remains, which is of zeroth order in the soft momentum.[]{data-label="singlegrav"}](multiout "fig:"){width=".19\textwidth"} ![Diagrams contributing to an amplitude with external soft boson. The first two diagrams cancel the last two diagrams, and only the diagram in the middle remains, which is of zeroth order in the soft momentum.[]{data-label="singlegrav"}](multibody "fig:"){width=".19\textwidth"} ![Diagrams contributing to an amplitude with external soft boson. The first two diagrams cancel the last two diagrams, and only the diagram in the middle remains, which is of zeroth order in the soft momentum.[]{data-label="singlegrav"}](multiincloud "fig:"){width=".19\textwidth"} ![Diagrams contributing to an amplitude with external soft boson. The first two diagrams cancel the last two diagrams, and only the diagram in the middle remains, which is of zeroth order in the soft momentum.[]{data-label="singlegrav"}](multioutcloud "fig:"){width=".19\textwidth"} The radial integrals can be computed separately, because the Faddeev-Kulish amplitudes $S^\text{FK}_{\beta b,\alpha}$ and $S^{\text{FK}*}_{\beta' b',\alpha}$ are $O(|{\mathbf}{k}|^0)$ in each soft momentum ${\mathbf}{k}$, which can be seen in figure \[singlegrav\] for a single outgoing soft graviton; the first two diagrams cancel the last two diagrams, and only the one in the middle, which is infrared-finite, contribute. If the momentum integral was over the whole momentum space, then the last integral in will diverge. But since it is only over the soft momentum space, it has a vanishingly small value (proportional to some momentum cutoff squared, $\omega_c^2$, where we think of $\omega_c\to 0$) and therefore the expression vanishes. Thus, we have $$\begin{aligned} \braket{b'|W^2|b} = \left[\frac{\kappa}{2} \sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k_1} \epsilon^{\ell_1*}_{{\mu\nu}}(k_1) \right] \bra{0} a_{\ell'_1}(k'_1)\cdots a_{\ell'_m}(k'_m) W^2 a^\dagger_{\ell_2}(k_2)\cdots a^\dagger_{\ell_n}(k_n) \ket{0}.\end{aligned}$$ Each creation operator gives a factor analogous to that in the square brackets, so we may write $$\begin{aligned} \braket{b'|W^2|b} = \prod_{i=1}^n\left[\frac{\kappa}{2} \sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k_i} \epsilon^{\ell_i*}_{{\mu\nu}}(k_i)\right] \bra{0} a_{\ell'_1}(k'_1)\cdots a_{\ell'_m}(k'_m) W^2 \ket{0}.\end{aligned}$$ We can perform a similar process for the annihilation operators, where this time the factors have an additional minus sign, and this yields $$\begin{aligned} \braket{b'|W^2|b} = \prod_{i=1}^n\left[\frac{\kappa}{2} \sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k_i} \epsilon^{\ell_i*}_{{\mu\nu}}(k_i)\right] \prod_{j=1}^m\left[-\frac{\kappa}{2} \sum_{p\in\beta,\beta'}\eta_p\frac{p^\mu p^\nu}{p\cdot k'_j} \epsilon^{\ell_j}_{{\mu\nu}}(k'_j)\right] \bra{0} W^2 \ket{0}.\end{aligned}$$ This explicitly shows that each term in the sum of contains a product of infrared-finite integrals as well as the vacuum expectation value $\braket{0|W^2|0}=\braket{N'_{{\text{out}}}|N_{{\text{out}}}}$, but we have seen that this value vanishes for the off-diagonal elements $\beta\neq\beta'$. Therefore, the reduced density matrix still exhibits a complete decoherence of the hard particle momentum configurations. We will conclude this section with a discussion of the two formulations of the density matrix: the one using the Bloch-Nordsieck mechanism and the one using dressed states. It is straightforward to see that only the off-diagonal elements of the reduced density matrix are different, whereas, the diagonal element which is essentially the Bloch-Nordsieck cross section is the same in the two approaches. Indeed, the cross section of the process $\alpha\to\beta b$ is given (up to a factor) by the absolute square of the amplitude: $$\begin{aligned} \Gamma_{\beta b,\alpha} = S_{\beta b,\alpha}S^*_{\beta b,\alpha}.\end{aligned}$$ These cross sections exhibit two types of infrared divergence, one arising from the real soft bosons and the other from the virtual bosons. The Bloch-Nordsieck method of dealing with these divergences is to sum over all unobservable soft bosons, $$\begin{aligned} \label{BN} \Gamma_{\beta, \alpha} = \sum_b \Gamma_{\beta b,\alpha} = \sum_b S_{\beta b,\alpha}S^*_{\beta b,\alpha},\end{aligned}$$ and performing this sum results in the exponentiation of the soft factors of real bosons, which then cancels the divergence due to virtual bosons. It is clear that every diagonal element of the reduced density matrix in is a Bloch-Nordsieck cross section: $$\begin{aligned} \rho_{\beta\beta} = \braket{\beta|\rho|\beta} = \sum_b S_{\beta b,\alpha}S^*_{\beta b,\alpha}.\end{aligned}$$ Thus, only the off-diagonal elements are affected (see Eq. ). The practical use of the Bloch-Nordsieck mechanism for obtaining IR finite cross sections does not require any modifications. Discussion {#DISCUSSION} ========== We have demonstrated that graviton cloud operators weakly commute with the scattering operator, and used this to show that scattering amplitudes which conserve BMS supertranslation charge are equal to the Faddeev-Kulish amplitudes. Since Faddeev-Kulish amplitudes are free of infrared divergence, this proves the conjecture in [@Kapec:2017tkm], that conservation of asymptotic charge leads to infrared finite scattering amplitudes. Our work ties up some loose ends on the relation between the Faddeev-Kulish formalism, asymptotic symmetry and infrared divergences. In particular, it supports the viewpoint that in QED and perturbative gravity, infrared divergences in the usual Dyson expansion of the S matrix are not real in the sense that they arise only as a result of using states that violate the conservation of the BMS supertranslation charge. Our paper also clarifies a common misconception in the literature that soft clouds “surround” the asymptotic particle. In actual fact, soft photons always go out to null infinity and massive particles to time-like infinity. Thus at large enough retarded times, one has matter particles surrounded by static fields (for example, coulomb fields in QED)[^2]. Indeed, it was conjectured in [@Kapec:2017tkm] and proved in this paper that the soft clouds can be moved from the in state to the out state without loosing infrared finiteness. We have also applied our formalism to the intriguing problem considered in [@Carney:2017jut] where it was found that tracing out soft degrees of freedom leads to the decoherence of hard particle momenta, whether or not one employs the Faddeev-Kulish states [@Carney:2017oxp]. In contrast to their work, we have constructed the corresponding reduced density matrices conserving the BMS supertranslation charge at all stages and arrived at a similar conclusion. It seems puzzling how to reconcile this with the fact that there are entanglement phenomenon observed in nature. Perhaps the large time S matrix approach may not be well suited to this problem. It is worth noting that while we have worked exclusively with gravity, a similar analysis can be applied to QED. This would lead to an analogous conclusion, namely, that large gauge transformation charge conservation implies infrared-finite Faddeev-Kulish amplitudes of QED. Therefore, a natural direction for future study would be the extension of these results to QCD. It will be very interesting and non-trivial to see how the asymptotic symmetry of QCD relates to the infrared-finite Faddeev-Kulish amplitude in that theory. More specifically, in QCD what is the connection between large gauge transformation charge conservation and infrared finiteness of the appropriate S matrix elements? S.C. acknowledges a fellowship from the Samsung Foundation of Culture. We are grateful to Andy Strominger for his insightful comments on the manuscript. We would like to thank Sandeep Pradhan for an initial collaboration and him, Gordon Semenoff and Malcolm Perry for discussions. [^1]: Note that here the in and out labels refer to incoming or outgoing particles. The Fock states are all in the “out" basis. [^2]: We thank Andy Strominger for clarifying this.
{ "pile_set_name": "ArXiv" }
--- abstract: 'An orthogonal representation of a graph is an assignment of nonzero real vectors to its vertices such that distinct non-adjacent vertices are assigned to orthogonal vectors. We prove general lower bounds on the dimension of orthogonal representations of graphs using the Borsuk-Ulam theorem from algebraic topology. Our bounds strengthen the Kneser conjecture, proved by Lovász in 1978, and some of its extensions due to [Bárány]{}, Schrijver, [Donikov]{}, and [Kriz]{}. As applications, we determine the integrality gap of fractional upper bounds on the Shannon capacity of graphs and the quantum one-round communication complexity of certain promise equality problems.' author: - 'Ishay Haviv[^1]' bibliography: - 'kneser.bib' title: '[**Topological Bounds on the Dimension of Orthogonal Representations of Graphs**]{}' --- Introduction ============ A $t$-dimensional [*orthogonal representation*]{} of a graph $G=(V,E)$ over a field ${\mathbb{F}}$ is an assignment of a vector $u_i \in {\mathbb{F}}^t$ to each vertex $i \in V$ such that $\langle u_i,u_i \rangle \neq 0$ for every $i \in V$, and $\langle u_i,u_j \rangle = 0$ for every distinct non-adjacent vertices $i$ and $j$ in $G$. The [*orthogonality dimension*]{} of a graph $G$ over a field ${\mathbb{F}}$, denoted by $\xi_{\mathbb{F}}(G)$, is defined as the smallest integer $t$ for which there exists a $t$-dimensional orthogonal representation of $G$ over ${\mathbb{F}}$. It is easy to verify that the orthogonality dimension of a graph is sandwiched between its independence number and its clique cover number, that is, for every graph $G$ and a field ${\mathbb{F}}$, $\alpha(G) \leq \xi_{\mathbb{F}}(G) \leq \chi(\overline{G})$. The notion of orthogonal representations over the real field was introduced by Lovász [@Lovasz79] in the study of the Shannon capacity of graphs and was later involved in a geometric characterization of connectivity properties of graphs by Lovász, Saks, and Schrijver [@LovaszSS89]. The orthogonality dimension over the complex field was used by de Wolf [@deWolfThesis] in a characterization of the quantum one-round communication complexity of promise equality problems and by Cameron et al. [@CameronMNSW07] in the study of the quantum chromatic number of graphs (see also [@ScarpaS12; @BrietBLPS15; @BrietZ17]). An extension of orthogonal representations, called orthogonal bi-representations, was introduced by Haemers [@Haemers81] (see also [@Peeters96]). Their smallest possible dimension, known as the minrank parameter of graphs, has found further applications in information theory, e.g., [@BBJK06; @mazumdar2014duality; @maleki2014index], and in theoretical computer science, e.g., [@Valiant92; @PudlakRS97; @HavivL13] (see Section \[sec:minrank\]). The present paper provides lower bounds on the orthogonality dimension of graphs over the real and complex fields using topological methods. The use of topological methods in combinatorics was initiated in the study of the chromatic number of the Kneser graph defined as follows. For integers $d \geq 2s$, the Kneser graph $K(d,s)$ is the graph whose vertices are all the $s$-subsets of $[d]=\{1,\ldots,d\}$ where two sets are adjacent if they are disjoint. In 1955, Kneser [@Kneser55] observed that $K(d,s)$ admits a proper coloring with $d-2s+2$ colors, simply by coloring every set $A$ by the smallest integer in $A \cup \{d-2s+2\}$, and conjectured that fewer colors do not suffice. In 1978, Lovász [@LovaszKneser] confirmed the conjecture by a breakthrough application of a tool from algebraic topology, the Borsuk-Ulam theorem [@Borsuk33]. Since then, topological methods have led to additional important results in combinatorics, discrete geometry, and theoretical computer science. For an in-depth background to the topic we refer the interested reader to Matou[š]{}ek’s excellent book [@MatousekBook]. Following Lovász’s proof of the Kneser conjecture, several alternative proofs were given in the literature. The simplest known proof of the conjecture is the one of Greene [@Greene02], inspired by a proof found by [Bárány]{} [@Barany78] soon after Lovász’s. Other proofs were given by [Donikov]{} [@Dolnikov82], Sarkaria [@Sarkaria90], and Matou[š]{}ek [@Matousek04], where Matou[š]{}ek’s proof is the only one derived from a combinatorial argument (but the topological inspiration is still around). Schrijver considered in [@SchrijverKneser78] the graph $S(d,s)$ defined as the subgraph of $K(d,s)$ induced by the collection of all $s$-subsets of $[d]$ that include no consecutive integers modulo $d$ (that is, the $s$-subsets $A \subseteq [d]$ such that if $i \in A$ then $i+1 \notin A$, and if $d \in A$ then $1 \notin A$). It was shown in [@SchrijverKneser78] that $S(d,s)$ is a vertex-critical subgraph of $K(d,s)$, that is, its chromatic number is equal to that of $K(d,s)$ and a removal of any vertex of $S(d,s)$ decreases its chromatic number. The various known proofs of the Kneser conjecture extend far beyond the chromatic number of the Kneser graph. Extensions of these proofs to lower bounds on the chromatic number of general graphs were given by [Donikov]{} [@Dolnikov82], by [Kriz]{} [@Kriz92], and by Matou[š]{}ek and Ziegler [@MatousekZ04] who generalized the proof techniques of Lovász, Sarkaria, and [Bárány]{}. Such extensions are usually stated for a generalized Kneser graph $K(\calF)$, defined as the graph whose vertex set is a set system $\calF$ where two sets are adjacent if they are disjoint. The generalized bounds are tight for the collection $\calF$ of all $s$-subsets of $[d]$ which corresponds to the graph $K(d,s)$, and some of them imply a tight lower bound on the chromatic number of the graph $S(d,s)$ as well. It is not difficult to see that every graph is isomorphic to $K(\calF)$ for some set system $\calF$ (see, e.g., [@MatousekZ04]), hence the bounds hold for all graphs (but for certain graphs they are quite weak). It was shown in [@MatousekZ04] that the extensions of the proofs of Lovász, Sarkaria, [Bárány]{}, [Donikov]{}, and [Kriz]{} can be (almost) linearly ordered by strength, where Lovász’s original proof technique is the strongest. Topological Bounds on the Orthogonality Dimension ------------------------------------------------- We prove two general lower bounds on the orthogonality dimension of graphs over the real and complex fields and on the minrank parameter over the real field (see Definition \[def:minrank\]). For convenience, we state the results for the complements of the generalized Kneser graphs $K(\calF)$. As mentioned before, every graph can be represented in this form. The two bounds are proved using the Borsuk-Ulam theorem from algebraic topology. The statement of our first bound is purely combinatorial. It strengthens the lower bounds on the chromatic number obtained independently by [Donikov]{} [@Dolnikov82] and by [Kriz]{} [@Kriz92]. Our proof is inspired by the proof of the Kneser conjecture by Greene [@Greene02]. To state the bound, we need the following definition (see, e.g., [@MatousekBook Section 3.4]). \[def:cd2\] Let $\calF$ be a set system with ground set $[d]$ such that $\emptyset \notin \calF$. The [*$2$-colorability-defect*]{} of $\calF$, denoted by ${\mathop{\mathrm{cd}}}_2(\calF)$, is the minimum size of a set $X \subseteq [d]$ such that the hypergraph on the vertex set $[d] \setminus X$ with the hyperedge set $\{ A \in \calF \mid A \cap X = \emptyset \}$ is $2$-colorable. Equivalently, ${\mathop{\mathrm{cd}}}_2(\calF)$ is the minimum number of white elements in a coloring of $[d]$ by red, blue and white, such that no set of $\calF$ is completely red or completely blue (but it may be completely white). \[thm:cd\] For every set system $\calF$ such that $\emptyset \notin \calF$, 1. \[thm:cd\_itm:1\] $\xi_\R(\overline{K(\calF)}) \geq {\mathop{\mathrm{cd}}}_2(\calF)$, 2. \[thm:cd\_itm:2\] $\xi_\C(\overline{K(\calF)}) \geq \frac{{{\mathop{\mathrm{cd}}}}_2(\calF)}{2}$, and 3. \[thm:cd\_itm:3\] ${{\mathop{\mathrm{minrk}}}}_\R(\overline{K(\calF)}) \geq \sqrt{\frac{{{\mathop{\mathrm{cd}}}}_2(\calF)}{2}}$. As an easy consequence of Theorem \[thm:cd\], we obtain that the orthogonality dimension of the complement of the Kneser graph $K(d,s)$ over the reals is $d-2s+2$ (see Corollary \[cor:xiKneser\]). Our second bound has a geometric nature. Its proof employs the approach of [Bárány]{} [@Barany78] to the Kneser conjecture and strengthens a general lower bound on the chromatic number that follows from [@Barany78] and is given explicitly in [@MatousekZ04]. In what follows, $\S^t$ stands for the $t$-dimensional unit sphere $\{ x \in \R^{t+1} \mid \|x\|=1 \}$, and an open hemisphere of $\S^t$ is a set of the form $\{ z \in \S^t \mid \langle x,z \rangle >0 \}$ for some $x \in \S^t$. \[thm:geo\] Let $\calF$ be a set system with ground set $[d]$ such that $\emptyset \notin \calF$. Suppose that for an integer $t \geq 2$ there exist points $y_1,\ldots,y_d \in \S^{t-2}$ such that every open hemisphere of $\S^{t-2}$ contains the points of $\{y_i \mid i \in A\}$ for some $A \in \calF$. Then, 1. \[thm:geom\_itm:1\] $\xi_\R(\overline{K(\calF)}) \geq t$, 2. \[thm:geom\_itm:2\] $\xi_\C(\overline{K(\calF)}) \geq \frac{t}{2}$, and 3. \[thm:geom\_itm:3\] ${{\mathop{\mathrm{minrk}}}}_\R(\overline{K(\calF)}) \geq \sqrt{\frac{t}{2}}$. The theorem is used to prove that the orthogonality dimension of the complement of the Schrijver graph $S(d,s)$ over the reals is $d-2s+2$ (see Corollary \[cor:xiSchrijver\]). The proof technique of Theorem \[thm:geo\] is also used to prove a lower bound on the orthogonality dimension over the reals of the complement of the Borsuk graph defined by Erd[ő]{}s and Hajnal [@ErdosH67] (see Section \[sec:Borsuk\]). Applications ------------ We describe below applications of our results to information theory and to quantum communication complexity. ### Shannon Capacity {#sec:Shannon} The strong product $G_1 \cdot G_2$ of two graphs $G_1=(V_1,E_1)$ and $G_2=(V_2,E_2)$ is defined as the graph whose vertex set is $V_1 \times V_2$ where two distinct vertices $(u_1,u_2)$ and $(v_1,v_2)$ are adjacent if for every $i \in \{1,2\}$ the vertices $u_i$ and $v_i$ are either equal or adjacent in $G_i$. The $k$-th power of a graph $G$, denoted by $G^k$, is defined as the product of $k$ copies of $G$. The Shannon capacity of a graph $G$, introduced in 1956 by Shannon [@Shannon56], is the limit $c(G) = \lim_{k \rightarrow \infty}{(\alpha(G^k))^{1/k}}$ whose existence follows from super-multiplicativity and Fekete’s lemma. This graph parameter is motivated by a question in information theory on the capacity of noisy channels. Indeed, if $G$ is the graph whose vertices are the symbols that a channel can transmit, and two symbols are adjacent if they may be confused in the transmission, then $c(G)$ can be intuitively interpreted as the effective alphabet size of the channel. The Shannon capacity parameter of graphs is very far from being well understood. It is not known if the problem of deciding whether the Shannon capacity of an input graph exceeds a given value is decidable, and the exact Shannon capacity is unknown even for small and fixed graphs, e.g., the cycle on $7$ vertices. Several upper bounds on the Shannon capacity of graphs were presented in the literature over the years. It is easy to see that $c(G) \leq \chi(\overline{G})$, and Shannon showed already in [@Shannon56] the stronger bound $c(G) \leq \chi_f(\overline{G})$, where $\chi_f$ stands for the fractional chromatic number. A useful way to obtain an upper bound on the Shannon capacity is to come up with a real-valued non-negative sub-multiplicative function on graphs that forms an upper bound on the independence number, that is, a function $f$ satisfying $f(G_1 \cdot G_2) \leq f(G_1) \cdot f(G_2)$ for every two graphs $G_1,G_2$ and $\alpha(G) \leq f(G)$ for every graph $G$. Indeed, for such an $f$ we have $$c(G) = \lim_{k \rightarrow \infty}{(\alpha(G^k))^{1/k}} \leq \lim_{k \rightarrow \infty}{(f(G^k))^{1/k}} \leq \lim_{k \rightarrow \infty}{(f(G)^k)^{1/k}} = f(G).$$ For example, it is not difficult to verify that the orthogonality dimension $\xi_{\mathbb{F}}$ over any field ${\mathbb{F}}$ is a sub-multiplicative upper bound on the independence number, hence $c(G) \leq \xi_{\mathbb{F}}(G)$ for every graph $G$. Other upper bounds on the Shannon capacity obtained in this way are the $\vartheta$-function due to Lovász [@Lovasz79], the ${{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}$ parameter due to Haemers [@Haemers81], and the minimum dimension of polynomial representations due to Alon [@AlonUnion98]. In a recent work, Hu, Tamo, and Shayevitz [@HuTS17] defined for every function $f$ as above a fractional linear programming variant $f^*$. For a graph $G$ on the vertex set $V$, $f^*(G)$ is the value of the following linear program. $$\begin{aligned} \label{Primal} \begin{split} \textup{maximize } &\sum_{x \in V} w(x) \\ \textup{subject to } & \sum_{x\in S}w(x) \le f(G[S]) \textup{~~~ for each set $S \subseteq V$},\\ & w(x)\ge 0 \textup{~~~ for each vertex $x \in V$}, \end{split}\end{aligned}$$ where $G[S]$ stands for the subgraph of $G$ induced by $S$. It was proved in [@HuTS17] that if $f$ is a sub-multiplicative upper bound on the independence number then so is $f^*$, hence $f^*$ also forms an upper bound on the Shannon capacity. Moreover, the upper bound $f^*$ is at least as strong as $f$, that is, $c(G) \leq f^*(G) \leq f(G)$ for every graph $G$ (see Section \[sec:Shannon4\]). It was shown in [@HuTS17] that the Lovász $\vartheta$-function satisfies $\vartheta(G)=\vartheta^*(G)$ for every graph $G$, whereas for other upper bounds $f$ on the Shannon capacity one can have $f^*(G) < f(G)$. For example, for every odd integer $n \geq 5$ the cycle $C_n$ satisfies ${{\mathop{\mathrm{minrk}}}}^*_\R(C_n) = \frac{n}{2} < \frac{n+1}{2} = {{\mathop{\mathrm{minrk}}}}_\R(C_n)$. In this work, we aim to study the integrality gap of the fractional quantities $f^*$ as a function of the number of vertices. Namely, we would like to estimate the largest possible ratio $f(G)/f^*(G)$ over all $n$-vertex graphs $G$. We start with a general upper bound. A function $f$ on graphs is said to be sub-additive if for every graph $G$ on the vertex set $V$ and every sets $S_1$ and $S_2$ such that $V = S_1 \cup S_2$, $f(G) \leq f(G[S_1]) + f(G[S_2])$. The following theorem shows that for sub-additive functions $f$ the ratio between $f$ and $f^*$ is at most logarithmic in the number of vertices. \[thm:fractional\_upper\] For every sub-additive function $f$ and every $n$-vertex graph $G$, $$f(G) \leq O(\log n) \cdot f^*(G).$$ It is easy to verify that all the aforementioned upper bounds on the Shannon capacity are sub-additive, hence their fractional variants cannot improve the bound on the Shannon capacity by more than a logarithmic multiplicative term. As an application of our results on the Kneser graph, we obtain a matching lower bound on the integrality gap of the fractional orthogonality dimension over the real and complex fields. \[thm:IntroIntegrality\] For every fixed ${{\varepsilon}}>0$, there exists an explicit family of $n$-vertex graphs $G$ such that $$\xi_\C^*(G) \leq \xi_\R^*(G) \leq 2+{{\varepsilon}}\mbox{~~whereas~~} \xi_\R (G)= \Theta(\log n) \mbox{~~and~~} \xi_\C (G)= \Theta(\log n).$$ We also show an unbounded integrality gap for the fractional minrank parameter over various fields (see Theorem \[thm:AllBounds\]). ### Quantum Communication Complexity In the standard model of communication complexity, two parties Alice and Bob get inputs $x,y$ from two sets $\calX,\calY$ respectively, and they have to compute by a communication protocol the value of $g(x,y)$ for a two-variable function $g$. In a promise communication problem, the inputs are guaranteed to be drawn from a subset of $\calX \times \calY$ known to the parties in advance. In a one-round protocol, the communication flows only from Alice to Bob. The classical, respectively quantum, communication complexity of a problem is the minimal number of bits, respectively qubits, that the parties have to exchange on worst-case inputs in a communication protocol for the problem. The orthogonality dimension of graphs over the complex field plays a central role in the study of the quantum communication complexity of promise equality problems.[^2] In such problems, Alice and Bob get either equal or adjacent vertices of a graph $G$ and their goal is to decide whether their inputs are equal. De Wolf [@deWolfThesis Section 8.5] showed that the classical one-round communication complexity of the promise equality problem associated with a graph $G$ is $\lceil \log_2 \chi(G) \rceil$, and that its quantum one-round communication complexity is $\lceil \log_2 \xi_\C(\overline{G}) \rceil$. Bri[ë]{}t et al. [@BrietBLPS15] proved that any classical protocol for such a problem can always be reduced to a classical one-round protocol with no extra communication, while in the quantum setting the one-round and two-round communication complexities of a promise equality problem can have an exponential gap. This separation was obtained using the Lovász $\vartheta$-function and the relation $\xi_\C(G) \geq \vartheta(G)$ (see [@BrietBLPS15 Lemma 2.5]). For a set system $\calF$ with $\emptyset \notin \calF$, consider the promise equality problem in which Alice and Bob get either equal or disjoint sets from $\calF$, and their goal is to decide whether their inputs are equal. Observe that the graph associated with this problem is the generalized Kneser graph $K(\calF)$, hence its quantum one-round communication complexity is precisely $\lceil \log_2 \xi_\C(\overline{K(\calF)}) \rceil$. Our bounds on the orthogonality dimension of such graphs over $\C$ have applications to the quantum one-round communication complexity of promise equality problems, as demonstrated below. For integers $d \geq 2s$, consider the communication complexity problem in which Alice and Bob get two $s$-subsets of $[d]$, their sets are guaranteed to be either equal or disjoint, and their goal is to decide whether the sets are equal. The graph associated with this problem is the Kneser graph $K(d,s)$, so its classical communication complexity is $\lceil \log_2 \chi(K(d,s)) \rceil = \lceil \log_2 (d-2s+2) \rceil$. As an application of Theorem \[thm:cd\], we get that $\lceil \log_2 \xi_\C(\overline{K(d,s)}) \rceil \geq \lceil \log_2 (d-2s+2)-1 \rceil$ (see Corollary \[cor:xiKneser\]), yielding the precise quantum one-round communication complexity of the problem up to an additive $1$. We note that the lower bound on $\xi_\C(\overline{K(d,s)})$ obtained from the Lovász $\vartheta$-function would not suffice here, since $\vartheta(\overline{K(d,s)}) = \frac{d}{s}$ (see [@Lovasz79]). The orthogonality dimension over the complex field was also used by Bri[ë]{}t et al. [@BrietBLPS15] to characterize the quantum one-round communication complexity of a family of problems called list problems, originally studied by Witsenhausen [@Witsenhausen76]. In the list problem that corresponds to the Kneser graph $K(d,s)$, Alice gets an $s$-subset $A$ of $[d]$, Bob gets a list of pairwise disjoint $s$-subsets of $[d]$ that includes the set $A$, and his goal is to discover $A$. It follows from [@BrietBLPS15] that the quantum one-round communication complexity of this problem is equal to the quantity $\lceil \log_2 \xi_\C(\overline{K(d,s)}) \rceil$ determined above. Outline ------- The rest of the paper is organized as follows. In Section \[sec:preliminaries\] we provide some background on the Borsuk-Ulam theorem and on the minrank parameter of graphs. In Section \[sec:bounds\] we prove Theorems \[thm:cd\] and \[thm:geo\] and obtain our results on the Kneser, Schrijver, and Borsuk graphs. Finally, in Section \[sec:Shannon4\] we study the integrality gap of fractional upper bounds on the Shannon capacity and prove Theorems \[thm:fractional\_upper\] and \[thm:IntroIntegrality\]. Preliminaries {#sec:preliminaries} ============= The Borsuk-Ulam Theorem {#sec:BU} ----------------------- For an integer $t \geq 0$, let $\S^{t} = \{ x \in \R^{t+1} \mid \|x\|=1 \}$ denote the $t$-dimensional unit Euclidean sphere. For a point $x \in \S^t$, let $H(x) = \{ z \in \S^t \mid \langle x,z \rangle >0 \}$ denote the open hemisphere of $\S^t$ centered at $x$. We state below the Borsuk-Ulam theorem, proved by Borsuk in 1933 [@Borsuk33]. \[thm:BU\] For every continuous function $f: \S^t \rightarrow \R^t$ there exists $x \in \S^t$ such that $f(x)=f(-x)$. Equivalently, if a continuous function $f: \S^t \rightarrow \R^{t'}$ satisfies $f(x) \neq f(-x)$ for all $x \in \S^t$ then $t' > t$. For several other equivalent versions of Theorem  \[thm:BU\], see [@MatousekBook Section 2.1]. We also need a variant of the Borsuk-Ulam theorem for complex-valued functions. We start with some notations. For a complex number $z \in \C$ we denote by ${\mathsf{Re}}(z)$ and ${\mathsf{Im}}(z)$ the real and imaginary parts of $z$ respectively, hence $z = {\mathsf{Re}}(z)+{\mathsf{Im}}(z) \cdot i$. For an integer $t \geq 1$, let $\phi_t:\C^t \rightarrow \R^{2t}$ be the natural embedding of $\C^t$ in $\R^{2t}$ defined by $$\begin{aligned} \label{eq:phi} \phi_t(x) = ({\mathsf{Re}}(x_1),{\mathsf{Im}}(x_1),\ldots,{\mathsf{Re}}(x_t),{\mathsf{Im}}(x_t)) \in \R^{2t}.\end{aligned}$$ Clearly, $\phi_t$ is a bijection from $\C^t$ to $\R^{2t}$. Notice that we have $\phi_t(-x)=-\phi_t(x)$ for every $x \in \C^t$. Our variant of the Borsuk-Ulam theorem for complex-valued functions is given below. \[thm:BU\_complex\] For every continuous function $f: \S^{2t} \rightarrow \C^t$ there exists $x \in \S^{2t}$ such that $f(x)=f(-x)$. For a continuous function $f: \S^{2t} \rightarrow \C^t$ consider the function $\widetilde{f}: \S^{2t} \rightarrow \R^{2t}$ defined by the composition $\widetilde{f} = \phi_t \circ f$. Applying Theorem \[thm:BU\] to $\widetilde{f}$, we get that there exists $x \in \S^{2t}$ such that $\widetilde{f}(x)=\widetilde{f}(-x)$. By the invertibility of $\phi_t$, this implies that $f(x) = f(-x)$, as required. Minrank {#sec:minrank} ------- The minrank parameter of graphs, introduced by Haemers in [@Haemers81], is defined as follows. \[def:minrank\] Let $G$ be a graph on the vertex set $V = [n]$ and let ${\mathbb{F}}$ be a field. We say that an $n \times n$ matrix $M$ over ${\mathbb{F}}$ [*represents*]{} $G$ if $M_{i,i} \neq 0$ for every $i \in V$, and $M_{i,j}=0$ for every distinct non-adjacent vertices $i,j \in V$. The [*minrank*]{} of $G$ over ${\mathbb{F}}$ is defined as $${{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}(G) = \min\{{{\mathop{\mathrm{rank}}}}_{{\mathbb{F}}}(M)\mid M \mbox{ represents }G\mbox{ over }{\mathbb{F}}\}.$$ The minrank parameter can be equivalently defined in terms of orthogonal bi-representations. A $t$-dimensional [*orthogonal bi-representation*]{} of a graph $G=(V,E)$ over a field ${\mathbb{F}}$ is an assignment of a pair $(u_i,v_i) \in {\mathbb{F}}^t \times {\mathbb{F}}^t$ to each vertex $i \in V$ such that $\langle u_i,v_i \rangle \neq 0$ for every $i \in V$, and $\langle u_i,v_j \rangle = \langle u_j,v_i \rangle = 0$ for every distinct non-adjacent vertices $i$ and $j$ in $G$. It can be verified that ${{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}(G)$ is the smallest integer $t$ for which there exists a $t$-dimensional orthogonal bi-representation of $G$ over ${\mathbb{F}}$ (see, e.g., [@Peeters96; @ChlamtacH14]). Since orthogonal bi-representations generalize orthogonal representations, we clearly have ${{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}(G) \leq \xi_{\mathbb{F}}(G)$ for all graphs $G$ and fields ${\mathbb{F}}$. The minrank parameter is always bounded from above by the clique cover number. The following lemma shows a lower bound on the minrank parameter over finite fields in terms of the clique cover number. Its proof is implicit in [@LangbergS08] and we give here a quick proof for completeness. \[claim:minrankF\] For every graph $G$ and a finite field ${\mathbb{F}}$, ${{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}(G) \geq \log_{|{\mathbb{F}}|}\chi(\overline{G})$. Denote $t = {{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}(G)$. Then there exists an assignment of a pair $(u_i,v_i) \in {\mathbb{F}}^t \times {\mathbb{F}}^t$ to each vertex $i \in V$ that forms an orthogonal bi-representation of $G$ over ${\mathbb{F}}$. Consider the coloring that assigns to every vertex $i \in V$ the color $u_i \in {\mathbb{F}}^t$. We claim that this is a proper coloring of $\overline{G}$. Indeed, for two vertices $i$ and $j$ adjacent in $\overline{G}$ we have $\langle u_i,v_i \rangle \neq 0$ and $\langle u_j,v_i \rangle = 0$, hence $u_i \neq u_j$. Since the number of used colors is at most $|{\mathbb{F}}|^t$, it follows that $\chi(\overline{G}) \leq |{\mathbb{F}}|^t$, as required. Topological Bounds on the Orthogonality Dimension {#sec:bounds} ================================================= In this section we prove Theorems \[thm:cd\] and \[thm:geo\] and obtain our results on the Kneser, Schrijver, and Borsuk graphs. The proofs employ the Borsuk-Ulam theorem given in Section \[sec:BU\]. Proof of Theorem \[thm:cd\] --------------------------- We start with the lower bound on the orthogonality dimension over the real field. Let $\calF$ be a set system with ground set $[d]$ such that $\emptyset \notin \calF$, and put $t = \xi_\R(\overline{K(\calF)})$. Then there exists an assignment of a nonzero vector $u_A \in \R^t$ to every set $A \in \calF$, such that $\langle u_A,u_B \rangle = 0$ for every disjoint sets $A,B \in \calF$. It can be assumed without loss of generality that the first nonzero coordinate in every vector $u_A$ is positive (otherwise replace $u_A$ by $-u_A$). Let $y_1,\ldots,y_d \in \S^t$ be $d$ points in a general position, that is, no $t+1$ of them lie on a $(t-1)$-dimensional sphere. For a set $A \subseteq [d]$ denote $y_A = \{ y_i \mid i \in A \}$. Define a function $f:\S^t \rightarrow \R^t$ by $$f(x) = \sum_{A \in \calF}{u_A \cdot \prod_{j \in A}{\max(\langle x, y_j\rangle,0)}}.$$ Observe that for every $x \in \S^{t}$, $f(x)$ is a linear combination with positive coefficients of the vectors $u_A$ such that $y_A \subseteq H(x)$, where $H(x)$ is the open hemisphere of $\S^t$ centered at $x$. The function $f$ is clearly continuous, hence by Theorem \[thm:BU\] there exists $x \in \S^t$ such that $f(x) = f(-x)$. However, $f(x)$ is a linear combination of the vectors $u_A$ with $y_A \subseteq H(x)$ whereas $f(-x)$ is a linear combination of the vectors $u_A$ with $y_A \subseteq H(-x)$. Since $H(x) \cap H(-x) = \emptyset$, it follows that the sets $A$ involved in the linear combination of $f(x)$ are all disjoint from those involved in the linear combination of $f(-x)$. The fact that $\langle u_A,u_B \rangle = 0$ for every disjoint sets $A,B \in \calF$ yields that the vectors $f(x)$ and $f(-x)$ are orthogonal, and by $f(x) = f(-x)$ they must be equal to the zero vector. We claim now that there is no $A \in \calF$ with $y_A \subseteq H(x)$. To see this, assume in contradiction that $A_1, \ldots, A_m$ are the sets with this property ($m \geq 1$), and let $j \in [t]$ be the least coordinate in which at least one of the vectors $u_{A_1},\ldots,u_{A_m}$ is nonzero. Since $f(x)$ is a linear combination of these vectors with positive coefficients, using the assumption that the first nonzero coordinate of every $u_{A}$ is positive, it follows that the $j$th coordinate of $f(x)$ is positive in contradiction to $f(x)$ being the zero vector. By the same reasoning, there is no $A \in \calF$ with $y_A \subseteq H(-x)$. Finally, let $X$ denote the set of indices $i \in [d]$ for which $y_i$ does not belong to $H(x)$ nor to $H(-x)$. By the assumption of general position, $|X| \leq t$. We color the elements of $[d] \setminus X$ as follows: If $y_i \in H(x)$ then $i$ is colored red, and if $y_i \in H(-x)$ then $i$ is colored blue. Since no set $A \in \calF$ satisfies $y_A \subseteq H(x)$ or $y_A \subseteq H(-x)$, we get a proper $2$-coloring of the hypergraph $([d] \setminus X, \{ A \in \calF \mid A \cap X = \emptyset \})$. This implies that ${\mathop{\mathrm{cd}}}_2(\calF) \leq t$, as required. We next prove our lower bound on the orthogonality dimension over the complex field. Its proof is similar to the proof over the reals but requires the Borsuk-Ulam theorem for complex-valued functions (Theorem \[thm:BU\_complex\]). Recall that $\phi_t$ stands for the natural embedding of $\C^t$ in $\R^{2t}$ given in . Let $\calF$ be a set system with ground set $[d]$ such that $\emptyset \notin \calF$, and put $t = \xi_\C(\overline{K(\calF)})$. Then there exists an assignment of a nonzero vector $u_A \in \C^t$ to every set $A \in \calF$, such that $\langle u_A,u_B \rangle = 0$ for every disjoint sets $A,B \in \calF$. It can be assumed without loss of generality that the first nonzero coordinate in every vector $\phi_t(u_A)$ is positive (otherwise replace $u_A$ by $-u_A$). Let $y_1,\ldots,y_d \in \S^{2t}$ be $d$ points in a general position. As before, for a set $A \subseteq [d]$ denote $y_A = \{ y_i \mid i \in A \}$. Define a function $f:\S^{2t} \rightarrow \C^t$ by $$f(x) = \sum_{A \in \calF}{u_A \cdot \prod_{j \in A}{\max(\langle x, y_j\rangle,0)}}.$$ Observe that for every $x \in \S^{2t}$, $f(x)$ is a linear combination with real positive coefficients of the vectors $u_A$ satisfying $y_A \subseteq H(x)$. The function $f$ is clearly continuous, hence by Theorem \[thm:BU\_complex\] there exists $x \in \S^{2t}$ such that $f(x) = f(-x)$. However, $f(x)$ is a linear combination of the vectors $u_A$ with $y_A \subseteq H(x)$ whereas $f(-x)$ is a linear combination of the vectors $u_A$ with $y_A \subseteq H(-x)$. Since $H(x) \cap H(-x) = \emptyset$, it follows that the sets $A$ involved in the linear combination of $f(x)$ are all disjoint from those involved in the linear combination of $f(-x)$. The fact that $\langle u_A,u_B \rangle = 0$ for every disjoint sets $A,B \in \calF$ yields that the vectors $f(x)$ and $f(-x)$ are orthogonal, and by $f(x) = f(-x)$ they must be equal to the zero vector. We claim now that there is no $A \in \calF$ with $y_A \subseteq H(x)$. To see this, assume in contradiction that $A_1, \ldots, A_m$ are the sets with this property ($m \geq 1$), and let $j \in [2t]$ be the least coordinate in which at least one of the vectors $\phi_{t}(u_{A_1}),\ldots,\phi_{t}(u_{A_m})$ is nonzero. It follows that the $j$th coordinate of $\phi_t(f(x))$ is positive in contradiction to $f(x)$ being the zero vector. By the same reasoning, there is no $A \in \calF$ with $y_A \subseteq H(-x)$. Finally, let $X$ denote the set of indices $i \in [d]$ for which $y_i$ does not belong to $H(x)$ nor to $H(-x)$. By the assumption of general position, $|X| \leq 2t$. We color the elements of $[d] \setminus X$ as follows: If $y_i \in H(x)$ then $i$ is colored red, and if $y_i \in H(-x)$ then $i$ is colored blue. Since no set $A \in \calF$ satisfies $y_A \subseteq H(x)$ or $y_A \subseteq H(-x)$, we get a proper $2$-coloring of the hypergraph $([d] \setminus X, \{ A \in \calF \mid A \cap X = \emptyset \})$. This implies that ${\mathop{\mathrm{cd}}}_2(\calF) \leq 2t$, as required. Finally, we prove our lower bound on the minrank parameter over the real field (recall Definition \[def:minrank\]). We start with the following lemma. Here, a real matrix is said to be non-negative if all of its entries are non-negative. \[lemma:minrank\_nonneg\] Let $\calF$ be a set system such that $\emptyset \notin \calF$, and let $M$ be a real non-negative matrix that represents the graph $\overline{K(\calF)}$ over $\R$. Then, ${{\mathop{\mathrm{rank}}}}_\R(M) \geq \frac{1}{2} \cdot {\mathop{\mathrm{cd}}}_2(\calF)$. Let $\calF$ be a set system of size $n$ with ground set $[d]$ such that $\emptyset \notin \calF$, let $M \in \R^{n \times n}$ be a non-negative matrix that represents the graph $\overline{K(\calF)}$ over $\R$, and put $t = {{\mathop{\mathrm{rank}}}}_\R(M)$. Write $M = M_1^T \cdot M_2$ for matrices $M_1,M_2 \in \R^{t \times n}$. For every set $A \in \calF$, let $u_A$ and $v_A$ be the $t$-dimensional columns associated with $A$ in $M_1$ and $M_2$ respectively, and let $w_A = u_A \circ v_A$ be the $2t$-dimensional concatenation of $u_A$ and $v_A$. Since $M$ represents $\overline{K(\calF)}$, we have $\langle u_A,v_A \rangle \neq 0$ for every $A \in \calF$, and $\langle u_A,v_B \rangle = 0$ for every disjoint sets $A,B \in \calF$. By the assumption that $M$ is non-negative, we also have $\langle u_A,v_B \rangle \geq 0$ for all $A,B \in \calF$. Let $y_1,\ldots,y_d \in \S^{2t}$ be $d$ points in a general position. As before, for a set $A \subseteq [d]$ denote $y_A = \{ y_i \mid i \in A \}$. Define a function $f:\S^{2t} \rightarrow \R^{2t}$ by $$f(x) = \sum_{A \in \calF}{w_A \cdot \prod_{j \in A}{\max(\langle x, y_j\rangle,0)}}.$$ Observe that for every $x \in \S^{2t}$, $f(x)$ is a linear combination with positive coefficients of the vectors $w_A$ such that $y_A \subseteq H(x)$. The function $f$ is clearly continuous, hence by Theorem \[thm:BU\] there exists $x \in \S^{2t}$ such that $f(x) = f(-x)$. For this $x$, denote $f(x) = f(-x) = w = w_1 \circ w_2$ where $w_1,w_2 \in \R^t$. By $f(x)=w$ we get that $w_1$ is a linear combination of the vectors $u_A$ with $y_A \subseteq H(x)$, and by $f(-x)=w$ we get that $w_2$ is a linear combination of the vectors $v_A$ with $y_A \subseteq H(-x)$. However, the sets $A$ with $y_A \subseteq H(x)$ are all disjoint form the sets $A$ with $y_A \subseteq H(-x)$, hence the fact that $\langle u_A,v_B \rangle = 0$ for every disjoint sets $A,B \in \calF$ yields that the vectors $w_1$ and $w_2$ are orthogonal. We claim now that there is no $A \in \calF$ with $y_A \subseteq H(x)$. To see this, assume in contradiction that $A_1, \ldots, A_m$ are the sets with this property ($m \geq 1$). By the definition of $f$ we can write $w = f(x)= \sum_{i=1}^{m}{c_i \cdot w_{A_i}}$ for some positive coefficients $c_i>0$. However, this implies that $$\langle w_1,w_2\rangle = \langle \sum_{i=1}^{m}{c_i \cdot u_{A_i}} ~, \sum_{i=1}^{m}{c_i \cdot v_{A_i}}\rangle = \sum_{1 \leq i,j \leq m}{c_i c_j \cdot \langle u_{A_i},v_{A_j} \rangle} > 0,$$ where the inequality holds since $\langle u_{A_i},v_{A_j} \rangle \geq 0$ for all pairs $i,j$ and $\langle u_{A_i},v_{A_i} \rangle > 0$ for every $i$. This is in contradiction to the fact that the vectors $w_1$ and $w_2$ are orthogonal. By the same reasoning, there is no $A \in \calF$ with $y_A \subseteq H(-x)$. Finally, let $X$ denote the set of indices $i \in [d]$ for which $y_i$ does not belong to $H(x)$ nor to $H(-x)$. By the assumption of general position, $|X| \leq 2t$. We color the elements of $[d] \setminus X$ as follows: If $y_i \in H(x)$ then $i$ is colored red, and if $y_i \in H(-x)$ then $i$ is colored blue. Since no set $A \in \calF$ satisfies $y_A \subseteq H(x)$ or $y_A \subseteq H(-x)$, we get a proper $2$-coloring of the hypergraph $([d] \setminus X, \{ A \in \calF \mid A \cap X = \emptyset \})$. This implies that ${\mathop{\mathrm{cd}}}_2(\calF) \leq 2t$, as required. Equipped with Lemma \[lemma:minrank\_nonneg\], we are ready to complete the proof of Theorem \[thm:cd\]. Let $\calF$ be a set system of size $n$ such that $\emptyset \notin \calF$, and let $M$ be an $n \times n$ matrix that represents the graph $\overline{K(\calF)}$ over $\R$. Consider the $n \times n$ matrix $M'$ defined by $M'_{i,j} = M_{i,j}^2$ for all $i,j$. It is well known and easy to check that $M'$ is a principal sub-matrix of the tensor product $M \otimes M$ of $M$ with itself, hence $${{\mathop{\mathrm{rank}}}}_\R (M') \leq {{\mathop{\mathrm{rank}}}}_\R (M \otimes M) = {{\mathop{\mathrm{rank}}}}_\R (M)^2.$$ The non-negative matrix $M'$ represents $\overline{K(\calF)}$ since it has the same zero pattern as $M$, so we can apply Lemma \[lemma:minrank\_nonneg\] to obtain that $${{\mathop{\mathrm{rank}}}}_\R(M) \geq \sqrt{{{\mathop{\mathrm{rank}}}}_\R(M')} \geq \sqrt{\frac{{{\mathop{\mathrm{cd}}}}_2(\calF)}{2}},$$ completing the proof. ### The Kneser Graph {#sec:Kneser} Recall that for integers $d \geq 2s$, the Kneser graph $K(d,s)$ is the graph whose vertices are all the $s$-subsets of $[d]$, where two sets are adjacent if they are disjoint. We need the following simple claim (see, e.g., [@MatousekBook Section 3.4]). \[claim:cd2\_Kneser\] For integers $d \geq 2s$, let $\calF$ be the collection of all $s$-subsets of $[d]$. Then, ${\mathop{\mathrm{cd}}}_2(\calF) = d-2s+2$. Let $X \subseteq [d]$ be an arbitrary set of size $d-2s+2$, and consider an arbitrary balanced $2$-coloring of the $2s-2$ elements of $[d] \setminus X$. Clearly, no $s$-subset of $[d] \setminus X$ is monochromatic, hence ${\mathop{\mathrm{cd}}}_2(\calF) \leq d-2s+2$. For the other direction, notice that for every $X \subseteq [d]$ of size at most $d-2s+1$ there are at least $2s-1$ elements in $[d] \setminus X$, hence every $2$-coloring of $[d] \setminus X$ includes a monochromatic $s$-subset. This implies that ${\mathop{\mathrm{cd}}}_2(\calF) \geq d-2s+2$ and completes the proof. The following corollary summarizes our bounds for the Kneser graph. \[cor:xiKneser\] For every integers $d \geq 2s$, 1. \[itm:Kneser\] $\xi_\R(\overline{K(d,s)}) = d-2s+2$, 2. $\xi_\C(\overline{K(d,s)}) \geq (d-2s+2)/2$, and 3. ${{\mathop{\mathrm{minrk}}}}_\R(\overline{K(d,s)}) \geq \sqrt{(d-2s+2)/2}$. Notice that $K(d,s)$ is the graph $K(\calF)$ where $\calF$ is the collection of all $s$-subsets of $[d]$. The three lower bounds follow directly by combining Theorem \[thm:cd\] with Claim \[claim:cd2\_Kneser\]. The matching upper bound in Item \[itm:Kneser\] follows by $\xi_\R(\overline{K(d,s)}) \leq \chi(K(d,s)) = d-2s+2$. Proof of Theorem \[thm:geo\] ---------------------------- We prove below Item \[thm:geom\_itm:1\] of Theorem \[thm:geo\]. The other two items follow similarly, using ideas from the proofs of Items \[thm:cd\_itm:2\] and \[thm:cd\_itm:3\] of Theorem \[thm:cd\]. To avoid repetitions, we omit the details. Let $\calF$ be a set system with ground set $[d]$ such that $\emptyset \notin \calF$, and let $y_1,\ldots,y_d \in \S^{t-2}$ be the points given in the theorem. Put $t' = \xi_\R(\overline{K(\calF)})$. Then there exists an assignment of a nonzero vector $u_A \in \R^{t'}$ to every set $A \in \calF$, such that $\langle u_A,u_B \rangle = 0$ for every disjoint sets $A,B \in \calF$. It can be assumed without loss of generality that the first nonzero coordinate in every vector $u_A$ is positive (otherwise replace $u_A$ by $-u_A$). Define a function $f:\S^{t-2} \rightarrow \R^{t'}$ by $$f(x) = \sum_{A \in \calF}{u_A \cdot \prod_{j \in A}{\max(\langle x, y_j\rangle,0)}}.$$ For any $x \in \S^{t-2}$, let $C_x$ be the collection of sets $A \in \calF$ such that $y_A \subseteq H(x)$, where, as before, $y_A = \{ y_i \mid i \in A \}$. By the assumption on the points $y_1,\ldots,y_d$ we have $|C_x| \geq 1$. Observe that $f(x)$ is a linear combination with positive coefficients of the vectors $u_A$ with $A \in C_x$. Letting $j \in [t']$ be the least coordinate in which at least one of the vectors of $\{ u_{A} \mid A \in C_x \}$ is nonzero, it follows that the $j$th coordinate of $f(x)$ is positive, hence $f(x)$ is nonzero. Further, by $H(x) \cap H(-x) = \emptyset$ we get that the sets of $C_x$ are all disjoint from those of $C_{-x}$, hence the fact that $\langle u_A,u_B \rangle = 0$ for every disjoint sets $A,B \in \calF$ yields that the vectors $f(x)$ and $f(-x)$ are orthogonal. Consider the function $g:\S^{t-2} \rightarrow \S^{t'-1}$ defined by $$g(x) = \frac{f(x)}{\|f(x)\|}.$$ Note that $g$ is well defined as $f(x)$ is nonzero for every $x \in \S^{t-2}$. Consider also the function $\widetilde{g} : \S^{t-2} \rightarrow \R^{t'-1}$ that maps every $x \in \S^{t-2}$ to the projection of $g(x)$ to its last $t'-1$ coordinates (i.e., all of its coordinates besides the first one). We claim that there is no $x \in \S^{t-2}$ such that $\widetilde{g}(x)=\widetilde{g}(-x)$. To see this, notice that $g(x)$ is a unit vector whose first entry is non-negative, so the projection of $g(x)$ to its last $t'-1$ coordinates fully determines $g(x)$. This implies that if there exists an $x \in \S^{t-2}$ satisfying $\widetilde{g}(x)=\widetilde{g}(-x)$ then this $x$ also satisfies $g(x)=g(-x)$, in contradiction to the orthogonality of $f(x)$ and $f(-x)$. Since $\widetilde{g}$ is continuous we can apply Theorem \[thm:BU\] to derive that $t'-1 > t-2$ which implies that $t' \geq t$ and completes the proof. ### The Schrijver Graph We say that a set $A \subseteq [d]$ is [*stable*]{} if it does not contain two consecutive elements modulo $d$ (that is, if $i \in A$ then $i+1 \notin A$, and if $d \in A$ then $1 \notin A$). In other words, a stable subset of $[d]$ is an independent set in the cycle $C_d$ with the numbering from $1$ to $d$ along the cycle. Recall that for $d \geq 2s$, the Schrijver graph $S(d,s)$ is the graph whose vertices are all the stable $s$-subsets of $[d]$, where two sets are adjacent if they are disjoint. We need the following strengthening of a lemma of Gale [@Gale56] proved by Schrijver in [@SchrijverKneser78]. See [@MatousekBook Section 3.5] for a nice proof by Ziegler based on the moment curve. \[lemma:Gale\] For every integers $d \geq 2s$, there exist points $y_1,\ldots,y_d \in \S^{d-2s}$ such that every open hemisphere of $\S^{d-2s}$ contains the points of $\{ y_i \mid i \in A \}$ for some stable $s$-subset $A$ of $[d]$. For $d \geq 2s$, consider the collection $\calF$ of all stable $s$-subsets of $[d]$, and notice that $S(d,s)$ is the graph $K(\calF)$. By Lemma \[lemma:Gale\], $\calF$ satisfies the condition of Theorem \[thm:geo\] for $t=d-2s+2$. This directly implies the following corollary which summarizes our bounds for the Schrijver graph. \[cor:xiSchrijver\] For every integers $d \geq 2s$, 1. \[itm:Kneser\] $\xi_\R(\overline{S(d,s)}) = d-2s+2$, 2. $\xi_\C(\overline{S(d,s)}) \geq (d-2s+2)/2$, and 3. ${{\mathop{\mathrm{minrk}}}}_\R(\overline{S(d,s)}) \geq \sqrt{(d-2s+2)/2}$. We note that the bounds that Theorem \[thm:cd\] implies for the Schrijver graph $S(d,s)$ are weaker than the bounds obtained above. Indeed, it is easy to check that the set system $\calF$ that corresponds to the graph $S(d,s)$ satisfies ${\mathop{\mathrm{cd}}}_2(\calF) = d-4s+4$. For a discussion comparing the bounds derived from Theorems \[thm:cd\] and \[thm:geo\], see [@MatousekZ04 Section 6]. ### The Borsuk Graph {#sec:Borsuk} For $0< \alpha <2$ and an integer $d$, the Borsuk graph $B(d,\alpha)$ is defined as the (infinite) graph on the vertex set $\S^{d-1}$ where two points $y,y' \in \S^{d-1}$ are adjacent if $\|y-y'\| \geq \alpha$. It is known that the Borsuk-Ulam theorem implies that $\chi(B(d,\alpha)) \geq d+1$ for every $0< \alpha <2$ and $d$, and that this bound is tight whenever $\alpha \geq \sqrt{2(d+1)/d}$ (see, e.g., [@Lovasz83]). We apply here the proof technique of Theorem \[thm:geo\] to obtain the same bound on the orthogonality dimension over $\R$ of the complement of $B(d,\alpha)$. We first prove the following. \[thm:BorsukGraph\] For $0< {{\varepsilon}}<1$ and an integer $d$, let $V \subseteq \S^{d-1}$ be a finite collection of points in $\S^{d-1}$ such that for every $x \in \S^{d-1}$ there exists $y \in V$ for which $\|x-y\| < {{\varepsilon}}$. Let $G$ be the graph on the vertex set $V$ where two points $y,y' \in V$ are adjacent if $\|y-y'\| \geq 2-2{{\varepsilon}}$. Then, $\xi_\R(\overline{G}) \geq d+1$. For a graph $G$ as in the theorem, put $t = \xi_\R(\overline{G})$. Then there exists an assignment of a nonzero vector $u_y \in \R^t$ to every point $y \in V$, such that $\langle u_y,u_{y'} \rangle = 0$ for every points $y,y' \in V$ satisfying $\|y-y'\| \geq 2-2{{\varepsilon}}$. It can be assumed without loss of generality that the first nonzero coordinate in every vector $u_y$ is positive (otherwise replace $u_y$ by $-u_y$). Define a function $f:\S^{d-1} \rightarrow \R^t$ by $$f(x) = \sum_{y \in V}{u_y \cdot \max ( {{\varepsilon}}- \|x-y\| ,0 )}.$$ For an $x \in \S^{d-1}$, let $C_x \subseteq V$ be the set of all points $y \in V$ such that $\|x-y\| < {{\varepsilon}}$. By the assumption on $V$ we have $|C_x| \geq 1$. Observe that $f(x)$ is a linear combination with positive coefficients of the vectors $u_y$ with $y \in C_x$. Letting $j \in [t]$ be the least coordinate in which at least one of the vectors of $\{ u_{y} \mid y \in C_x \}$ is nonzero, using the assumption that the first nonzero coordinate of every $u_{y}$ is positive, it follows that the $j$th coordinate of $f(x)$ is positive, hence $f(x)$ is nonzero. Further, since the distance between $x$ and $-x$ is $2$, for every $y \in C_x$ and $y' \in C_{-x}$ we have $\|y-y'\| \geq 2-2{{\varepsilon}}$, and thus $\langle u_y, u_{y'} \rangle = 0$. This implies that the vectors $f(x)$ and $f(-x)$ are orthogonal. Consider the function $g:\S^{d-1} \rightarrow \S^{t-1}$ defined by $g(x) = \frac{f(x)}{\|f(x)\|}$. Note that $g$ is well defined as $f(x)$ is nonzero for every $x \in \S^{d-1}$. Consider also the function $\widetilde{g} : \S^{d-1} \rightarrow \R^{t-1}$ that maps every $x \in \S^{d-1}$ to the projection of $g(x)$ to its last $t-1$ coordinates. We claim that there is no $x \in \S^{d-1}$ such that $\widetilde{g}(x)=\widetilde{g}(-x)$. To see this, notice that $g(x)$ is a unit vector whose first entry is non-negative, so the projection of $g(x)$ to its last $t-1$ coordinates fully determines $g(x)$. This implies that if there exists an $x \in \S^{d-1}$ satisfying $\widetilde{g}(x)=\widetilde{g}(-x)$ then this $x$ also satisfies $g(x)=g(-x)$, in contradiction to the orthogonality of $f(x)$ and $f(-x)$. Since $\widetilde{g}$ is continuous we can apply Theorem \[thm:BU\] to derive that $t-1 > d-1$ which implies that $t \geq d+1$ and completes the proof. \[cor:BorsukGraph\] For every $0< \alpha <2$ and an integer $d$, $\xi_\R(\overline{B(d,\alpha)}) \geq d+1$. For $0< \alpha <2$ and an integer $d$, let $V$ be a maximal collection of points in $\S^{d-1}$ with pairwise distances at least ${{\varepsilon}}= 1-\frac{\alpha}{2}$. Observe that $V$ is finite and that for every $x \in \S^{d-1}$ there exists $y \in V$ for which $\|x-y\| < {{\varepsilon}}$. The graph $G$ associated with $V$ and ${{\varepsilon}}$ in Theorem \[thm:BorsukGraph\] is a subgraph of $B(d,\alpha)$, hence $\xi_\R(\overline{B(d,\alpha)}) \geq \xi_\R(\overline{G}) \geq d+1$, and we are done. Fractional Upper Bounds on the Shannon Capacity {#sec:Shannon4} =============================================== Let $f$ be a real-valued non-negative function on graphs. As explained in Section \[sec:Shannon\], if $f$ is a sub-multiplicative upper bound on the independence number then it forms an upper bound on the Shannon capacity. A fractional variant $f^*$ of $f$ was introduced in [@HuTS17], and it was shown there that if $f$ is a sub-multiplicative upper bound on the independence number then so is $f^*$, hence $f^*$ also forms an upper bound on the Shannon capacity. For a graph $G$ on the vertex $V$, the definition of $f^*(G)$ is given in , and by duality it is equal to the value of the following linear program. $$\begin{aligned} \label{Dual} \begin{split} \textup{minimize } &\sum_{S \subseteq V} q(S) \cdot f(G[S]) \\ \textup{subject to } &\sum_{S: x \in S}q(S) \ge 1 \textup{~~~ for each vertex $x \in V$},\\ &q(S)\ge 0 \textup{~~~ for each set $S \subseteq V$}. \end{split}\end{aligned}$$ Note that every graph $G$ satisfies $f^*(G) \leq f(G)$, as follows by taking $q(S)=1$ for $S=V$ and $q(S)=0$ otherwise. We study here the integrality gap of fractional upper bounds $f^*$ on the Shannon capacity, namely, the largest possible ratio $f(G)/f^*(G)$ over all $n$-vertex graphs $G$. Upper Bound ----------- We prove now Theorem \[thm:fractional\_upper\], which claims that for sub-additive functions $f$ on graphs, the integrality gap of $f^*$ is at most logarithmic in the number of vertices. Recall that $f$ is sub-additive if for every graph $G$ on the vertex set $V$ and every sets $S_1$ and $S_2$ such that $V = S_1 \cup S_2$, it satisfies $f(G) \leq f(G[S_1]) + f(G[S_2])$. We use the dual definition of $f^*$ given in . Let $G=(V,E)$ be an $n$-vertex graph, and let $q$ be an optimal solution of . Denote $Q = \sum_{S \subseteq V}{q(S)} \geq 1$, and let $D$ be the distribution over the subsets of $V$ that assigns to every set $S \subseteq V$ the probability $\frac{q(S)}{Q}$. For, say, $t = \lceil Q \cdot \ln (3n) \rceil$, let $S_1,\ldots,S_t$ be $t$ random subsets of $V$ chosen independently from the distribution $D$. We first claim that the probability that the sets $S_1,\ldots,S_t$ do not cover the entire vertex set $V$ is at most $1/3$. Indeed, for every $x \in V$ and $i \in [t]$ the probability that $x \in S_i$ is $$\sum_{S:x \in S}{\frac{q(S)}{Q}} = \frac{1}{Q} \cdot \sum_{S:x \in S}{q(S)} \geq \frac{1}{Q},$$ where the inequality holds since $q$ is a feasible solution of . Hence, the probability that $x$ does not belong to any of the sets $S_1,\ldots,S_t$ is at most $$\Big (1-\frac{1}{Q} \Big )^t \leq e^{-t/Q} \leq \frac{1}{3n}.$$ By the union bound, the probability that there exists $x \in V$ such that $x \notin \cup_{i=1}^{t}{S_i}$ is at most $1/3$. We next consider the expectation of the sum $\sum_{i=1}^{t}{f(G[S_i])}$. For every $i \in [t]$ we have $${\mathop{{\mathbb{E}}}}[f(G[S_i])] = \sum_{S \subseteq V} {\frac{q(S)}{Q} \cdot f(G[S])} = \frac{f^*(G)}{Q},$$ where the second equality holds since $q$ is an optimal solution of . By linearity of expectation, $${\mathop{{\mathbb{E}}}}\Big [\sum_{i=1}^{t}{f(G[S_i])} \Big ] = t \cdot \frac{f^*(G)}{Q} = \lceil Q \cdot \ln (3n) \rceil \cdot \frac{f^*(G)}{Q} \leq 2 \cdot \ln (3n) \cdot f^*(G).$$ By Markov’s inequality, the probability that $\sum_{i=1}^{t}{f(G[S_i])} \geq 6 \cdot \ln (3n) \cdot f^*(G)$ is at most $1/3$. Finally, by the union bound, there are sets $S_1,\ldots,S_t$ that cover the vertex set $V$ and satisfy $\sum_{i=1}^{t}{f(G[S_i])} \leq 6 \cdot \ln (3n) \cdot f^*(G)$. For these sets the sub-additivity of $f$ implies that $$f(G) \leq \sum_{i=1}^{t}{f(G[S_i])} \leq 6 \cdot \ln (3n) \cdot f^*(G),$$ and we are done. Lower Bounds ------------ We turn to show that the ratio between $f$ and $f^*$ is unbounded for several upper bounds $f$ on the Shannon capacity. In fact, we present an explicit family of graphs for which $f^*$ is bounded from above by a constant whereas $f$ is arbitrarily large. We consider here the graph parameters $\xi_{\mathbb{F}}$ and ${{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}$, which, as mentioned before, are sub-additive and sub-multiplicative upper bounds on the independence number for every field ${\mathbb{F}}$. We need the well-studied notion of fractional chromatic number of graphs. For a graph $G$ on the vertex set $V$, let $\calI(G)$ denote the collection of all independent sets of $G$. The [*fractional chromatic number*]{} of $G$, denoted by $\chi_f(G)$, is defined as the value of the following linear program. $$\begin{aligned} \label{fractional_chi} \begin{split} \textup{minimize } &\sum_{I \in \calI(G)} q(I) \\ \textup{subject to } &\sum_{I \in \calI(G): x \in I}q(I) \ge 1 \textup{~~~ for each vertex $x \in V$},\\ &q(I)\ge 0 \textup{~~~ for each set $I \in \calI(G)$}. \end{split}\end{aligned}$$ The following claim, given in [@HuTS17], follows directly from  and . \[claim:f\*chi\_f\] Let $f$ be a function on graphs satisfying $f(G)=1$ whenever $G$ is complete. Then for every graph $G$, $f^*(G) \leq \chi_f(\overline{G})$. In particular, for every graph $G$ and every field ${\mathbb{F}}$, $${{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}^*(G) \leq \xi_{\mathbb{F}}^*(G) \leq \chi_f(\overline{G}).$$ For integers $d \geq 2s$, consider the Kneser graph $K(d,s)$ (see Section \[sec:Kneser\]). The number of vertices in $K(d,s)$ is $\binom{d}{s}$. By the Erd[ő]{}s-Ko-Rado theorem [@ErdosKoRado61] its independence number is $\alpha(K(d,s)) = \binom{d-1}{s-1}$, and as already mentioned, its chromatic number is $\chi(K(d,s)) = d-2s+2$ [@LovaszKneser]. It is known that every vertex-transitive graph $G=(V,E)$ (that is, a graph whose automorphism group is transitive), satisfies $\chi_f(G) = \frac{|V|}{\alpha(G)}$ (see, e.g., [@GraphTheoryBook]). Hence, $$\begin{aligned} \label{eq:fracKneser} \chi_f(K(d,s)) = \frac{\binom{d}{s}}{\binom{d-1}{s-1}} = \frac{d}{s}.\end{aligned}$$ Now we are ready to derive the following theorem, which confirms Theorem \[thm:IntroIntegrality\]. \[thm:AllBounds\] For every fixed ${{\varepsilon}}>0$, there exists an explicit family of $n$-vertex graphs $G$ such that 1. \[itm:1\] ${{\mathop{\mathrm{minrk}}}}^*_{\mathbb{F}}(G) \leq \xi^*_{\mathbb{F}}(G) \leq 2+{{\varepsilon}}$   for every field ${\mathbb{F}}$, 2. \[itm:2\] $\xi_\R (G)= \Theta(\log n)$, 3. \[itm:3\] $\xi_\C (G)= \Theta(\log n)$, 4. \[itm:4\] ${{\mathop{\mathrm{minrk}}}}_\R (G) \geq \Omega( \sqrt{ \log n})$, and 5. \[itm:5\] $\xi_{\mathbb{F}}(G) \geq {{\mathop{\mathrm{minrk}}}}_{\mathbb{F}}(G) \geq \Omega( \log \log n)$   for every fixed finite field ${\mathbb{F}}$. For a fixed ${{\varepsilon}}>0$, let $d$ and $s$ be two integers satisfying $d = (2+{{\varepsilon}}) \cdot s$. Let $G$ be the complement of the Kneser graph $K(d,s)$. We show that $G$ satisfies the assertion of the theorem. By , we have $$\chi_f(\overline{G}) = \chi_f(K(d,s)) = \frac{d}{s} = 2+{{\varepsilon}},$$ hence Item \[itm:1\] of the theorem follows from Claim \[claim:f\*chi\_f\]. The graph $G$ has $n = \binom{d}{s} = 2^{\Theta(s)}$ vertices, so we have $d-2s+2 = {{\varepsilon}}\cdot s +2 = \Theta(\log n)$. For the upper bounds in Items \[itm:2\] and \[itm:3\] notice that for every field ${\mathbb{F}}$, $\xi_{\mathbb{F}}(G) \leq \chi(\overline{G}) = d-2s+2 = \Theta(\log n)$. The lower bounds in Items \[itm:2\], \[itm:3\], and \[itm:4\] follow from Corollary \[cor:xiKneser\]. Finally, Item \[itm:5\] follows from Claim \[claim:minrankF\]. Acknowledgments {#acknowledgments .unnumbered} =============== We would like to thank Itzhak Tamo for useful conversations and for his comments on an early version of the paper. We are also grateful to Florian Frick and Günter M. Ziegler for helpful discussions. [^1]: School of Computer Science, The Academic College of Tel Aviv-Yaffo, Tel Aviv 61083, Israel. [^2]: Note that the orthogonality dimension parameter (also known as orthogonality rank) is sometimes defined in the quantum communication complexity literature as the orthogonality dimension of the complement graph, namely, the definition requires vectors associated with [*adjacent*]{} vertices to be orthogonal. In this paper we have decided to follow the definition commonly used in the information theory literature.
{ "pile_set_name": "ArXiv" }