text
stringlengths 4
2.78M
| meta
dict |
---|---|
---
abstract: 'In an ensemble of laser-driven atoms involving strongly interacting Rydberg states, the excitation probability is usually strongly suppressed. In contrast, here we identify a regime in which the steady-state Rydberg excited fraction is enhanced by the interaction. This effect is associated with the build-up of many-body coherences, induced by coherent multi-photon excitations between collective states. The excitation enhancement should be observable under currently-existing experimental conditions, and may serve as a direct probe for the presence of coherent multi-photon dynamics involving collective quantum states.'
author:
- Martin Gärttner
- Shannon Whitlock
- 'David W. Schönleber'
- Jörg Evers
title: 'Collective excitation of Rydberg-atom ensembles beyond the $\sqrt{N}$ enhancement'
---
The emergence of collective quantum effects in a many-body system is a hallmark of the strongly-correlated regime. A celebrated early example is the phenomenon of Dicke superradiance, in which $N$ two-level atoms coherently interacting with a common optical field acquire enhanced emission properties [@dicke1954]. In cavity quantum electrodynamics, an ensemble of $N$ atoms placed inside an optical cavity experiences a collective enhancement of the couping to a single cavity photon which scales as $\sqrt{N}$ [@colombe2007; @brennecke2007]. Similar collective effects arise in the coupling of superconducting devices to nitrogen vacancy centers in diamond [@kubo2010; @zhu2011; @amsuess2011], and could play important roles in circuit-QED [@wallraff2004] and hybrid optical-micromechanical systems [@vogell2013]. Ensembles of highly-excited Rydberg atoms offer an alternative system with which to study correlations and collective effects due to strong interatomic interactions. Here the competition between the laser excitation process and the interactions ensures only one Rydberg excitation can be accommodated within a critical distance (dipole blockade effect) leading to strong spatial correlations. In this regime a $\sqrt{N}$ enhancement of the atom-light coupling has been demonstrated [@heidemann2007; @gaetan2009; @urban2009; @dudin2012c]. By exploiting these effects it should be possible to realize new quantum technologies such as non-classical light sources [@dudin2012b; @peyronel2012; @maxwell2013] and quantum gates based on collectively enhanced interactions [@lukin2001].
Here, we point out a surprising enhancement of the steady-state Rydberg population for driven three-level atoms which goes beyond the $\sqrt{N}$ enhancement typically associated with the dipole blockade effect. The enhancement occurs for resonant two-step excitation from the ground state to the Rydberg state (under electromagnetically-induced-transparency (EIT) conditions) and for repulsive Rydberg-Rydberg interactions. We show that the effect is related to direct multi-photon transitions between collective states. As a result, its investigation requires exact calculations of the many-body master equation or analytical calculations beyond second order in the driving fields. We identify regimes in which the effect is most pronounced, and find that the Rydberg population of the repulsively interacting system can even be larger than for the completely non-interacting system, highlighting the genuine many-body nature of the system. As this effect should be observable in existing cold atom experiments, it may serve as a direct probe for coherent multi-photon processes between collective states via the observation of a global steady state observable.
![(Color online) (a) Illustration of the setup: Three-level atoms with the Rydberg state ${\left|r\right>}$. The blockade radius exceeds the trap size such that at most one atom can be in the Rydberg state. (b) Steady-state values of the Rydberg fraction $f_r$ for $N=4$ atoms as a function of the probe Rabi frequency ($\Delta=0$). The coherent laser driving is strong compared to the decay rate $\Gamma$ of the intermediate level. Solid black line: exact ME solution, red dashed line: result of a SA model with $\sqrt{N}$-enhanced $\Omega_p$, blue dot-dashed line: value of $f_r$ for non-interacting atoms, black dotted line: analytical solution for $f_r$ in the weak probe limit. The gray shading denotes the region of interaction-induced excitation enhancement beyond the non-interacting value.[]{data-label="fig:MERE_pop"}](fig1){width="\columnwidth"}
First, we consider an ensemble of $N$ three-level atoms all confined to a volume comparable to, or smaller than, a single blockade sphere (Fig. \[fig:MERE\_pop\]a). The atoms are driven by two laser fields which couple the ground and intermediate state ${\left|g\right>}\rightarrow {\left|e\right>}$ and the intermediate and the Rydberg state ${\left|e\right>}\rightarrow {\left|r\right>}$, referred to as the probe and coupling transitions, respectively. The ${\left|e\right>}$ state decays rapidly (decay rate $\Gamma$) via spontaneous emission to the ground state while the other two states are long-lived.
The Hamiltonian of this system reads ($\hbar=1$) $$H=\sum_{i=1}^N H_L^{(i)} - \sum_{i=1}^N \Delta{\left|e_i\right>}{\left<e_i\right|} + \sum_{i<j}\frac{C_6 {\left|r_i r_j\right>}{\left<r_i r_j\right|}}{|\mathbf{x}_i-\mathbf{x}_j|^6}
\label{eq:Hamiltonian}$$ where $H_L^{(i)}=\Omega_p/2 {\left|g_i\right>}{\left<e_i\right|} + \Omega_c/2 {\left|e_i\right>}{\left<r_i\right|} + h.c.$ and $\mathbf{x}_i$ are the atomic positions. We allow for a detuning $\Delta$ from the intermediate state while the two-photon transition is kept resonant. We assume that the Rydberg states interact repulsively via isotropic van der Waals interactions with strength $C_6$. Incoherent processes like spontaneous emission and dephasing are treated using a master equation (ME) including Lindblad terms [@supplement]. The resulting ME reads $\dot{\rho}=-i[H,\rho]+\mathcal{L}[\rho]$. We assume continuous, spatially homogeneous laser driving and focus on the Rydberg fraction $f_r=\mathrm{Tr}[(N^{-1}\sum_i {\left|r_i\right>}{\left<r_i\right|}) \rho]$ in the steady state ($\dot{\rho}=0$) as our main observable. We are able to simulate the dynamics and steady states for up to $N=10$ three-level atoms using the wave function Monte Carlo method [@molmer1993].
Solving the ME for a single atom ($N=1$) and under perfect EIT conditions (zero dephasing and decay of the Rydberg state), up to a normalization factor, yields the steady state ${\left|d\right>}={\left|g\right>}-\Omega_p/\Omega_c{\left|r\right>}$, which is the EIT dark state. Thus the Rydberg excitation probability, equivalent to the steady state Rydberg fraction, becomes $f_0=\Omega_p^2/(\Omega_c^2+\Omega_p^2)$. In the case of a fully blockaded ensemble of $N>1$ particles, the atoms are excited to collective Dicke states, which leads to a $\sqrt{N}$ enhancement of the atom-light coupling [@dicke1954; @lukin2001]. In this spirit one can describe the fully blockaded ensemble as a single three-level “super-atom” (SA) with a $\sqrt{N}$ larger transition dipole moment for the probe transition. The resulting Rydberg excitation probability of a SA is obtained by replacing $\Omega_p$ with $\sqrt{N}\Omega_p$ in $f_0$, i.e., $N f{_{\text{SA}}}=N\Omega_p^2/(\Omega_c^2+N\Omega_p^2)$. Comparing $f{_{\text{SA}}}$ to $f_0$ shows that in the simple SA picture, the SA excitation probability cannot be enhanced more than $N$-fold over the single-atom value, $Nf{_{\text{SA}}}\leq Nf_0$. In the following we will show that the exact solution of the ME can violate this bound and thus an excitation enhancement beyond the bare $\sqrt{N}$ is possible.
Figure \[fig:MERE\_pop\](b) shows the Rydberg fraction $f_r$ for a fully blockaded ensemble as a function of $\Omega_p/\Omega_c$ for $N=4$, $\Omega_c = 4\Gamma$ and $\Delta=0$. For weak driving $\Omega_p\rightarrow 0$ the exact solution agrees with the non-interacting (or single-atom) value $f_0$. For large $\Omega_p$ the Rydberg fraction $f_r$ approaches a constant value of $\approx 1/N$ due to the blockade. In between (for $0.15 \lesssim\Omega_p/\Omega_c\lesssim 0.4$), the Rydberg excited fraction significantly exceeds the non-interacting value. This feature is counter to the usual expectation for the dipole blockade in which repulsive interactions lead to a reduction of the number of Rydberg excitations compared to the non-interacting case. The red dashed line shows the prediction of the SA model, which predicts $f{_{\text{SA}}}\leq f_0$ and is thus insufficient to explain this feature.
![(Color online) (a) Level scheme of symmetrized (Dicke) states. Diagonalizing the horizontal couplings ($\Omega_c$) a dressed state picture is obtained. (b) and (c) Dressed state energies as a function of the detuning from the intermediate state $\Delta$ in the (b) non-interacting and (c) blockaded case. In the blockaded case resonances between the state ${\left|G\right>}$ and the doubly excited dressed states (red) occur at $\Delta\approx \pm\Omega_c/2$ (circles). Gray dotted lines show bare energies (or asymptotes). (d) Rydberg fraction as a function of $\Delta$. The parameters are as in Fig. \[fig:MERE\_pop\]. Solid black line: exact ME solution, red dashed line: result of the SA model, blue dot-dashed line $f_0$, black dotted line: analytical model. The resonances between the dressed Dicke states result in a large excitation enhancement in comparison to the non-interacting Rydberg fraction.[]{data-label="fig:delta_dep"}](fig2){width="\columnwidth"}
To provide a physical picture for the collective excitation enhancement we express the system in the basis of symmetrized Dicke states [@dicke1954; @lukin2001; @carmele2013; @liu2014]. These states are fully symmetric superpositions of all excited states with the same number of $e$ and $r$ excitations, $${\left|E^j R^s\right>}=\mathcal{N}_{j,s}\left(\sum_{i=1}^N{\left|e_i\right>}{\left<g_i\right|}\right)^j\left(\sum_{i=1}^N{\left|r_i\right>}{\left<g_i\right|}\right)^s{\left|G\right>}$$ with the normalization $\mathcal{N}_{j,s}$ [@supplement]. We use the shorthand notation ${\left|E^0 R^0\right>}={\left|G\right>}$, ${\left|E^1 R^0\right>}={\left|E\right>}$ and ${\left|E^0 R^1\right>}={\left|R\right>}$ \[Fig. \[fig:delta\_dep\](a)\]. Note that $s\in\{0,1\}$ for a perfectly blockaded ensemble, so the state ${\left|R^2\right>}$ is not present in this case. The Hamiltonian preserves the symmetry of these states and population of states outside the manifold of symmetric states only arises due to incoherent effects such as spontaneous emission. The coupling laser does not change the total number of excitations $j+s$. Therefore, the $\Omega_c$-part of the Hamiltonian can be diagonalized, resulting in the dressed many body eigenstates shown in Figs. \[fig:delta\_dep\](b) and (c). Due to the exclusion of states with multiple Rydberg excitations the interacting system exhibits new level crossings. In particular we point out the crossings between the states ${\left|G\right>}$ and the dressed states of the $j+s=2$ manifold at $\Delta=\pm\Omega_c/2$ where we expect an excitation enhancement due to direct two-photon processes. In the non-interacting case, a level crossing appears at $\Delta=0$, however, here the dressed Dicke state picture is incomplete since interference effects between different excitation channels are not visualized. Such effects lead to the absence of $e$-excitations in the non-interacting case (EIT) and thus the non-interacting Rydberg fraction is independent of the intermediate state detuning $\Delta$. Figure \[fig:delta\_dep\](d) shows the detuning dependence of the Rydberg fraction. The full calculations show a significant detuning dependence of $f_r$ with two peaks centered around $\Delta\approx \pm\Omega_c/2$ with widths of $\approx \Gamma$. At its maximum the enhancement is approximately a 50% effect. On resonance ($\Delta=0$) the enhancement is still visible, but only on the order of 5% for these parameters. Clearly, here the simple SA picture fails, since it amounts to neglecting all but the states ${\left|G\right>}$, ${\left|E\right>}$, and ${\left|R\right>}$. The presence of the resonances, however, shows that states with $j+s>1$ play a crucial role. In particular, from the dressed state representation we find that coherent two-photon excitation from the ground to the doubly excited collective states is essential for the observed enhancement effect.
![(Color online) (a) Dissipative phase diagram showing the enhancement factor $f_r/f_0$ as a function of Rabi frequencies for $N=4$ atoms and $\Delta=0$. For strong coherent driving and weak probe field, a region of $f_r>f_0$ is encountered. The dashed line shows the transition from suppressed to enhanced excitation ($f_r=f_0$). The arrow marks the critical value of $\Omega_c/\Gamma \approx 1.27$. (b) Maximal value of $f_r/f_0$ as a function of atom number $N$. (c) Few atoms ($N=4$) in a spherical volume of varying radius with random position sampling. The possibility for multiple Rydberg excitations is included in the simulation and realistic values for the laser dephasings and the spontaneous decay rate of the Rydberg state have been used (see text for details). The parameters correspond to the the black cross (lower black curves) and plus sign (upper red curves) in (a). For the higher Rabi frequencies the excitation enhancement is present at all sphere radii (densities). []{data-label="fig:fbl_ME"}](fig3){width="\columnwidth"}
On resonance, this effect appears most surprising as in the non-interacting case the dressed Dicke states show a crossing at $\Delta=0$, while in the blockaded case they do not. Therefore, in the following we focus on the excitation enhancement for the special case of $\Delta=0$. Figure \[fig:fbl\_ME\](a) shows the ratio $f_r/f_0$, obtained from exact ME calculations as a function of the Rabi frequencies for the case of four fully blockaded atoms and perfect EIT conditions. For strong coupling and weak probe fields we observe a range of parameters where the Rydberg fraction exceeds that of a non-interacting ensemble, $f_r>f_0$. The dashed line marks the border of this region ($f_r=f_0$). As $N$ is increased the qualitative features of this dissipative phase diagram stay the same and the structure is compressed vertically, consistent with a $\sqrt{N}$ rescaling of $\Omega_p$, while the onset in $\Omega_c/\Gamma$ is independent of $N$ in the weak probe limit. We also determine the values of $\Omega_p$ for which $f_r/f_0$ is maximized. For $\Omega_c=4\Gamma$ the maximum is found at $\sqrt{N}\Omega_p\approx0.5\Omega_c$. Figure \[fig:fbl\_ME\](b) shows the maximal value of $f_r/f_0$ as a function of the atom number $N$. We find that $f_r/f_0$ appears to saturate but reaches values larger than $1.4$ if extrapolated to larger $N$. Similarly, the enhancement factor saturates as a function of $\Omega_c/\Gamma$.
In the following we derive analytical expressions for the Rydberg fraction for a fully blockaded ensemble using fourth order perturbation theory for small $\Omega_p/\Omega_c$. By dividing the ME by $\Omega_c/2$, we separate off the term associated with probe laser proportional to $\epsilon=\Omega_p/\Omega_c$ which can be treated as a perturbation in the weak probe limit. We then expand the steady state density matrix in orders of the small parameter $\epsilon$ ($\rho=\sum_n \rho_n \epsilon^n$) and solve for the $\rho_n$ recursively [@supplement]. To fourth order we obtain $f_r^{(4)} = \epsilon^2 + c_4\epsilon^4 + \mathcal{O}(\epsilon^6)$ where $$\label{eq:fr4main}
c_4 = 2(N-1)\frac{1+4\mathrm{Im[\beta]}^2-|\beta|^2-|\beta|^4}{|1+\beta^2|^2}-1$$ with $\beta=(\Gamma+2i\Delta)/\Omega_c$. The predictions of Eq. are shown as dotted lines in Figs. \[fig:MERE\_pop\](b) and \[fig:delta\_dep\](d). We can now compare the fourth order terms in the non-interaction case $f_0=\epsilon^2-\epsilon^4+\mathcal{O}(\epsilon^6)$, the SA case $f{_{\text{SA}}}=\epsilon^2-N\epsilon^4+\mathcal{O}(\epsilon^6)$, and the master equation calculation $f_r^{(4)}$. By solving $c_4>-1$ for $\beta$ with $\Delta=0$ we find that an enhancement ($f_r>f_0$) occurs if $\Omega_c/\Gamma>\sqrt{\varphi}$ with $\varphi=(\sqrt{5}-1)/2$, independent of $N$ \[arrow in Fig. \[fig:fbl\_ME\](a)\]. A closer analysis of $c_4$ reveals that in the limit $\Omega_c\gg \Gamma$, Eq. describes two Lorentzian peaks of width $\Gamma$ centered at $\Delta=\pm\Omega_c/2$. The enhancement of the fourth order term beyond the non-interacting value is $c_4+1\approx 2(N-1)$ on resonance ($\Delta=0$), while at the maxima ($\Delta=\pm\Omega_c/2$) one obtains $c_4+1\approx 3\Omega_c^2(N-1)/(2\Gamma^2)$.
We now ask if the excitation enhancement persists under realistic experimental conditions, for example with imperfect blockade and including dephasing and the finite lifetime of the Rydberg state. We consider $^{87}$Rb atoms with states ${\left|g\right>}={\left|5s_{1/2}\right>}$, ${\left|e\right>}={\left|5p_{3/2}\right>}$, and ${\left|r\right>}={\left|55s\right>}$ as in [@hofmann2012]. The atoms are assumed to be randomly distributed inside a sphere of variable radius. The corresponding interaction coefficient is $C_6/2\pi=50\,\textrm{GHz}~\mu\textrm{m}^6$, and the spontaneous decay rates are $\Gamma/2\pi=6.06$MHz and $\Gamma_r/2\pi=2$kHz. An overview of the simulations is shown in Fig. \[fig:fbl\_ME\](c) (solid lines). The lower black lines correspond to the Rabi frequencies $\Omega_p/2\pi=1\,$MHz, $\Omega_c/2\pi=5.1$MHz, and the laser linewidths $\gamma_p/2\pi=0.33\,$MHz and $\gamma_c/2\pi=1.4\,$MHz, typical of recent experiments [@hofmann2012]. The upper red lines correspond to the strong driving case with both Rabi frequencies increased by a factor of $3.75$. In the case of weak driving no excitation enhancement is observed. For strong driving $f_r/f_0$ exceeds unity (enhancement) and increases towards higher densities (smaller sphere radius). Although the laser dephasings are chosen relatively large, the enhancement persists for all densities. This means that the excitation enhancement should be observable under realistic experimental conditions as long as the regime of strong coherent driving ($\Omega_c/\Gamma>\sqrt{\varphi}$) is reached. Through time-dependent simulations we also verified that the steady state is reached in less than $1\,\mu$s for these parameters. One option to observe the excitation enhancement effect could thus be to vary the cloud density at constant atom number, for example by thermal expansion.
The observation of the excitation enhancement has a number of consequences for understanding light-matter interactions in strongly-interacting systems. Firstly, the nonlinear optical response of a Rydberg gas driven under EIT conditions is predicted to obey a universal relation between the optical susceptibility $\chi=\textrm{Im}[\chi_{eg}]$ and the Rydberg fraction [@ates2011; @sevincli2011b; @gaerttner2014b]. This relation states that in the case of perfect EIT $\chi/\chi_{2L} = 1-f_r/f_0$, where $\chi_{2L}$ is the susceptibility in the two-level case, i.e., without coupling laser. Our finding that $f_r>f_0$ would therefore predict a negative susceptibility, hence showing the universal relation must break down in the collectively enhanced regime. Secondly, we show that coherences between $N$-atom collective states and direct two-photon processes are important for determining the steady-state of this system under realistic experimental conditions. This indicates that the classical rate equation (RE) models which neglect these coherences [@ates2011; @petrosyan2011; @sevincli2011b; @heeg2012; @hofmann2012; @gaerttner2013; @gaerttner2014b] are not sufficient to describe the most important details of the experiments. While these models do reproduce the $\sqrt{N}$ collective enhancement, they cannot describe the effect reported here since $f_r<f_0$ is assumed by construction [@gaerttner2014b]. For comparison, the predictions of a RE model [@heeg2012] have been added as dashed lines in Fig. \[fig:fbl\_ME\](c), which qualitatively fail to reproduce the enhancement reported here. The excitation enhancement therefore acts as a clear experimental signature for the breakdown of the RE approach.
Finally, we discuss the influence of decoherence. Typically, it is challenging to distinguish between coherent and incoherent excitation dynamics [@schempp2013; @schoenleber2013; @malossi2013; @weber2014; @schauss2012; @petrosyan2013c; @lesanovsky2014; @urvoy2014]. The excitation enhancement reported here manifests itself in a global observable in steady-state and is thus comparatively easy to access experimentally. Since it is associated with a direct two-photon process, it requires at least partly coherent dynamics. Indeed, we find that if single atom dephasing is added, $f_r/f_0$ decreases monotonically and falls below unity when dephasing dominates.
In summary, we have discovered a counter-intuitive excitation enhancement effect that occurs for strong driving of the upper and weak driving of the lower transition of interacting three-level Rydberg atoms. We have shown that this effect is connected to direct multi-photon transitions between collective Dicke states and have derived analytical expressions for the steady state density for arbitrary atom number $N$, which are capable of reproducing the observed enhancement effect in the weak probe limit. The enhancement allows to detect the presence of coherent multi-photon processes, and to identify parameter regimes in which the RE approach breaks down. As the enhancement involves a global steady state observable it should be observable with existing experimental setups [@hofmann2012; @maxwell2013; @peyronel2012; @schauss2012].
A natural next step would be to ask how the observed excitation enhancement is connected to the build-up of multi-partite entanglement. Furthermore, it is interesting to ask whether similar enhancements could be observed in other strongly interacting driven systems, or if the generated multi-particle coherences will find applications in metrology or in quantum information science.
We thank K. P. Heeg and M. Höning for discussions. This work was supported by University of Heidelberg (Center for Quantum Dynamics, LGFG). SW acknowledges support by the DFG Emmy Noether grant Wh141-1-1.
[35]{}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 [****, ()](\doibase 10.1103/PhysRev.93.99) [****, ()](http://dx.doi.org/10.1038/nature06331) [****, ()](http://dx.doi.org/10.1038/nature06120) [****, ()](\doibase
10.1103/PhysRevLett.105.140502) [****, ()](http://dx.doi.org/10.1038/nature10462) [****, ()](\doibase 10.1103/PhysRevLett.107.060502) [****, ()](http://dx.doi.org/10.1038/nature02851) [****, ()](\doibase 10.1103/PhysRevA.87.023816) [****, ()](\doibase 10.1103/PhysRevLett.99.163601) [****, ()](\doibase 10.1038/nphys1183) [****, ()](\doibase
10.1038/nphys1178) [****, ()](\doibase
10.1038/nphys2413) [****, ()](\doibase 10.1126/science.1217901) [****, ()](\doibase
10.1038/nature11361) [****, ()](\doibase
10.1103/PhysRevLett.110.103001) [****, ()](\doibase 10.1103/PhysRevLett.87.037901) @noop [****, ()](\doibase 10.1364/JOSAB.10.000524) @noop [****, ()](\doibase
10.1103/PhysRevA.89.033839) [****, ()](\doibase
10.1103/PhysRevLett.110.203601) [****, ()](\doibase 10.1103/PhysRevA.83.041802) [****, ()](\doibase doi:10.1088/0953-4075/44/18/184018) [****, ()](\doibase 10.1103/PhysRevA.89.063407) [****, ()](\doibase 10.1103/PhysRevLett.107.213601) [****, ()](\doibase 10.1103/PhysRevA.86.063421) [****, ()](\doibase 10.1103/PhysRevA.88.033417) [****, ()](\doibase
10.1103/PhysRevLett.112.013002) [****, ()](\doibase 10.1103/PhysRevA.89.033421) [****, ()](\doibase
10.1103/PhysRevLett.113.023006) @noop [****, ()](http://dx.doi.org/10.1038/nature11596) [“,” ](http://stacks.iop.org/0953-4075/46/i=14/a=141001) () [****, ()](\doibase 10.1103/PhysRevA.90.011603) @noop
Supplemental material
=====================
Here we provide the details of our analytical calculations of the steady state of a fully blockaded ensemble of atoms in the weak probe limit.
[*Steady-state perturbation:*]{} In order to calculate the steady state density matrix perturbatively, we exploit that in the weak probe limit, the ME can be divided into two parts, one of which is proportional to the small parameter $\epsilon=\Omega_p/\Omega_c$: $$\label{eq:ME_general}
0=\mathcal{L}_0[\rho]+\epsilon \mathcal{L}_1[\rho] \, .$$ We now expand the full steady state in terms of the small parameter $\epsilon$ $$\rho^{(ss)}=\rho_0+\epsilon \rho_1 + \epsilon^2 \rho_2 + \ldots \,.$$ Substituting this into Eq. and comparing terms that are of the same order in $\epsilon$ we can solve for the $\rho_i$ recursively assuming that the zeroth order term $\rho_0$, i.e., the steady state of $\mathcal{L}_0$ is known. Then the higher order terms are obtained by solving $$\label{eq:ss_iterate}
0=\mathcal{L}_0[\rho_{i}]+\mathcal{L}_1[\rho_{i-1}]$$ for $\rho_i$.
In our case we seek to solve $$\label{eq:ME_full}
0= -i\frac{\Omega_p}{2}[X_p,\rho] -i\frac{\Omega_c}{2}[X_c,\rho] + \frac{\Gamma}{2}\mathcal{L}_{se}[\rho] \, ,$$ where $$X_p=\sum_{i=1}^N\left({\left|g_i\right>}{\left<e_i\right|}+{\left|e_i\right>}{\left<g_i\right|}\right) \,,$$ $$X_c=\sum_{i=1}^N\left({\left|e_i\right>}{\left<r_i\right|}+{\left|r_i\right>}{\left<e_i\right|}\right) \,,$$ and $$\label{eq:decay}
\mathcal{L}_{se}[\rho]=\sum_{i=1}^N\left(2{\left|g_i\right>}{\left<e_i\right|}\rho{\left|e_i\right>}{\left<g_i\right|} - \left\{{\left|e_i\right>}{\left<e_i\right|},\rho\right\}\right) \,.$$ The last term accounts for spontaneous decay from the intermediate level.
[*Symmetrized basis:*]{} As the Hamiltonian is invariant under exchange of the particles, it is convenient to use a symmetrized basis, or Dicke state basis [@dicke1954; @lukin2001] (see also [@carmele2013]), ${\left|E^j R^s\right>}$ with $s\in\{0,1\}$, $j\in\{0,\ldots, N-s\}$, ${\left|E^0 R^0\right>}={\left|G\right>}={\left|g_1\ldots g_N\right>}$ and $${\left|E^j R^s\right>}=\mathcal{N}_{j,s}\left(\sum_{i=1}^N{\left|e_i\right>}{\left<g_i\right|}\right)^j\left(\sum_{i=1}^N{\left|r_i\right>}{\left<g_i\right|}\right)^s{\left|G\right>}$$ with the normalization $$\mathcal{N}_{j,s}=\sqrt{\frac{(N-j-s)!}{N^s(N-s)!j!}}.$$ The matrix elements of the Hamiltonian in this basis are $$\begin{split}
{\left<E^j R^s\right|}X_p{\left|E^{j\prime} R^{s\prime}\right>} & = \sqrt{(j+1)(N-j-s)}\delta_{j,j^\prime - 1}\delta_{s,s^\prime} \\
& +\sqrt{j(N-j+1-s)}\delta_{j,j^\prime + 1}\delta_{s,s^\prime}
\end{split}$$ and $${\left<E^j R^s\right|}X_c{\left|E^{j\prime} R^{s\prime}\right>} = \sqrt{j+s}\delta_{j+s,j^\prime + s^\prime}\delta_{1-s,s^\prime}$$ We observe that the couplings between the symmetrized states of small $j+s$ induced by the probe laser scale with $\sqrt{N}$, while those of the coupling laser are independent of $N$. This is the reason why all characteristic features in Fig. 3(a) become approximately $N$-independent in the weak probe limit if the parameter $\Omega_p$ is rescaled by $\sqrt{N}$.
The Lindblad term does not preserve the symmetry, so we introduce asymmetric states for $j+s=1$, which are required for the 4th order perturbative calculation
$$\begin{aligned}
{\left|E_k\right>} &= \frac{1}{\sqrt{N}}\sum_{j=1}^N {\left|e_j\right>}{\left<g_j\right|}{\rm e}^{2\pi i \frac{j k}{N}} {\left|G\right>} \\
{\left|R_k\right>} &= \frac{1}{\sqrt{N}}\sum_{j=1}^N {\left|r_j\right>}{\left<g_j\right|}{\rm e}^{2\pi i \frac{j k}{N}} {\left|G\right>} \end{aligned}$$
where for $k=N$ the symmetric states ${\left|E_N\right>}={\left|E^1R^0\right>}={\left|E^1\right>}$ and ${\left|R_N\right>}={\left|E^0R^1\right>}={\left|R\right>}$ are recovered.
We summarize the action of the Lindblad term on the relevant density matrix elements ${\left<E^j R^s\right|}\rho{\left|E^{j\prime} R^{s\prime}\right>}$. For matrix elements involving a state with no $e$-excitation, we can at most have a diagonal term: $$\mathcal{L}_{se}[{\left|E^j R^s\right>}{\left<E^0R^{s\prime}\right|}] = -j{\left|E^j R^s\right>}{\left<E^0R^{s\prime}\right|}$$ Matrix elements involving ${\left|E^1\right>}$ decay only to symmetric states: $$\begin{split}
\mathcal{L}_{se}[{\left|E^j R^s\right>} &{\left<E^1\right|}] = -(j+1){\left|E^j R^s\right>}{\left<G\right|}\\
& + 2\sqrt{j\left(1-\frac{j-1+s}{N}\right)}{\left|E^{j-1} R^s\right>}{\left<G\right|}
\end{split}$$ The action on elements with $j+s=j^\prime+s^\prime=2$ can be summarized as $$\begin{split}
\mathcal{L}_{se}[{\left|E^jR^s\right>} &{\left<E^{j\prime}R^{s\prime}\right|}] = -(j+j^\prime){\left|E^jR^s\right>}{\left<E^{j\prime}R^{s\prime}\right|} \\
&+ 2\sqrt{jj^\prime}\left(1-\frac{1}{N}\right){\left|E^{j-1}R^s\right>}{\left<E^{j\prime-1}R^{s\prime}\right|} \\
&+ \frac{2\sqrt{jj^\prime}}{N(N-1)}\sum_{k=1}^{N-1}{\left|(E^{j-1}R^s)_k\right>}{\left<(E^{j\prime-1}R^{s\prime})_k\right|}
\end{split}$$ with ${\left|(E^1R^0)_k\right>}={\left|E_k\right>}$ and ${\left|(E^0R^1)_k\right>}={\left|R_k\right>}$.
![(Color online) Structure of the ME in the symmetrized basis, including couplings between density matrix elements ${\left|E^jR^s\right>} {\left<E^{j\prime}R^{s\prime}\right|}$. $H_c$ connects neighboring cells separated by a dashed line, i.e., differing only in $s$ or $s^\prime$. $H_p$ connects states differing by one in $j$ or $j^\prime$. The Hamiltonian cannot induce couplings between states of different symmetry quantum number $k$. The Lindblad term connects matrix elements in which both $j$ and $j^\prime$ differ by one, irrespective of the symmetry of the state. Note, that only a selection of couplings is shown. The gray shading shows the part of the density matrix that is being populated in the different orders of the weak probe expansion of the steady state, starting with the ground state $\rho_0={\left|G\right>}{\left<G\right|}$ and extending to higher and higher excited states as the order of $\epsilon$ increases.[]{data-label="fig:ME_sketch"}](ME_visualized){width="\columnwidth"}
The structure of the ME is illustrated in Fig. \[fig:ME\_sketch\]. The time derivative of a matrix element $\rho_{j,s;j\prime,s\prime}$ represented by a cell of the table depends on all the matrix elements from which arrows point to it.
[*Up to fourth order results in the weak probe limit:*]{} Here we want to perturbatively solve the steady state equation $$\label{eq:ME_weakP}
0= -i\epsilon[X_p,\rho] -i[X_c,\rho] + \beta\mathcal{L}_{se}[\rho] \, ,$$ where $\epsilon=\Omega_p/\Omega_c$ and $\beta=\Gamma/\Omega_c$. We have omitted the intermediate state detuning for the sake of readability. In the above notation we thus have $\mathcal{L}_0[\rho]=-i[X_c,\rho] + \beta\mathcal{L}_{se}[\rho]$ and $\mathcal{L}_1[\rho]=-i[X_p,\rho]$.
The zeroth order equation $\mathcal{L}_{0}[\rho_0]=0$ is trivially solved by $\rho_0={\left|G\right>}{\left<G\right|}$. Solving the recursion up to the fourth order leads to the unnormalized steady state $$\label{eq:full_rhoss}
\begin{split}
\rho^{(4)} &= {\left|G\right>}{\left<G\right|} + N\epsilon^2{\left|R\right>}{\left<R\right|} \\
+ & \frac{\epsilon^4}{x^2} \biggr\{ (N-1)^2{\left|E^1\right>}{\left<E^1\right|} \\
& + [N(N-1)(3-\beta^2 x)-(N-1)y]{\left|R\right>}{\left<R\right|} \\
& + \frac{N(N-1)}{2}{\left|E^2\right>}{\left<E^2\right|} \\
& + N(N-1)\beta^2{\left|E^1R\right>}{\left<E^1R\right|} \\
& + \sum_{k=1}^{N-1} \left[x{\left|E_k\right>}{\left<E_k\right|} + y{\left|R_k\right>}{\left<R_k\right|} \right]\biggr\}\\
& + \text{off-diagonal terms}\,,
\end{split}$$ where we abbreviated $x=1+\beta^2$ and $y=1-\beta^2+\beta^4$. As we are only interested in populations of the Rydberg and intermediate states, so we do not list all the coherences here. From this, one obtains $$\label{eq:fr2}
f_r^{(2)}=\frac{\epsilon^2}{1+ N\epsilon^2}=\epsilon^2 + \mathcal{O}(\epsilon^4)$$ and $$\label{eq:fr4}
\begin{split}
f_r^{(4)} & = \frac{1}{\mathrm{Tr}[\rho^{(4)}]}\left[ \epsilon^2 + (N-1)\frac{3-\beta^4}{(1+\beta^2)^2} \right] \epsilon^4 \\
& = \epsilon^2 + \left[\frac{2(N-1)(1-\beta^2-\beta^4)}{(1+\beta^2)^2}-1\right] \epsilon^4 + \mathcal{O}(\epsilon^6)
\end{split}$$ as well as $$\label{eq:fe4}
f_e^{(4)} = \frac{1}{\mathrm{Tr}[\rho^{(4)}]}\frac{2(N-1)}{1+\beta^2}\epsilon^4 = \frac{2(N-1)}{1+\beta^2}\epsilon^4 + \mathcal{O}(\epsilon^6)\,.$$ The last expression of Eq. is what is plotted as a black dotted line in Fig. 1(b) in the main text. The second order steady state involves at most singly excited states. It is thus equivalent to a three-level atom with $\sqrt{N}$ enhanced probe Rabi frequency and also $f_r^{(2)}$ coincides with the result of a classical rate equation model [@heeg2012].
| {
"pile_set_name": "ArXiv"
} |
---
author:
- |
Mathias Butenschön and\
II. Institut für Theoretische Physik, Universität Hamburg, Luruper Chaussee 149, 22761 Hamburg\
E-mail: ,
title: 'Direct $J/\psi$ photoproduction at next-to-leading-order in nonrelativistic QCD'
---
The factorization formalism of nonrelativistic quantum chromodynamics (NRQCD) [@Bodwin:1994jh] provides a consistent theoretical framework for the description of heavy-quarkonium production and decay. This implies a separation of process-dependent short-distance coefficients, to be calculated perturbatively as expansions in the strong-coupling constant $\alpha_s$, from supposedly universal long-distance matrix elements (LDMEs), to be extracted from experiment. The relative importance of the latter can be estimated by means of velocity scaling rules; [*i.e.*]{}, the LDMEs are predicted to scale with a definite power of the heavy-quark ($Q$) velocity $v$ in the limit $v\ll1$. In this way, the theoretical predictions are organized as double expansions in $\alpha_s$ and $v$. A crucial feature of this formalism is that it takes into account the complete structure of the $Q\overline{Q}$ Fock space, which is spanned by the states $n={}^{2S+1}L_J^{[a]}$ with definite spin $S$, orbital angular momentum $L$, total angular momentum $J$, and color multiplicity $a=1,8$. In particular, this formalism predicts the existence of color-octet (CO) processes in nature. This means that $Q\overline{Q}$ pairs are produced at short distances in CO states and subsequently evolve into physical, color-singlet (CS) quarkonia by the nonperturbative emission of soft gluons. In the limit $v\to0$, the traditional CS model (CSM) is recovered in the case of $S$-wave quarkonia.
Fifteen years after the introduction of the NRQCD factorization formalism [@Bodwin:1994jh], the existence of CO processes and the universality of the LDMEs are still at issue and far from proven, despite an impressive series of experimental and theoretical endeavors. The greatest success of NRQCD was that it was able to explain the $J/\psi$ hadroproduction yield at the Fermilab Tevatron [@Cho:1995vh], while the CSM prediction lies orders of magnitudes below the data, even if the latter is evaluated at next-to-leading order (NLO) or beyond [@Campbell:2007ws]. Also in the case of $J/\psi$ photoproduction at DESY HERA, the CSM cross section significantly falls short of the data, as demonstrated by a recent NLO analysis [@Artoisenet:2009xh] using up-to-date input parameters and standard scale choices, leaving room for CO contributions [@Cacciari:1996dg]. Similarly, the $J/\psi$ yields measured in electroproduction at HERA and in two-photon collisions at CERN LEP2 were shown [@Kniehl:2001tk; @Klasen:2001cu] to favor the presence of CO processes. As for $J/\psi$ polarization in hadroproduction, neither the leading-order (LO) NRQCD prediction [@Braaten:1999qk], nor the NLO CSM one [@Campbell:2007ws] leads to an adequate description of the Tevaton data. The situation is quite similar for the polarization in photoproduction at HERA [@Artoisenet:2009xh].
In order to convincingly establish the CO mechanism and the LDME universality, it is an urgent task to complete the NLO description of $J/\psi$ hadro- [@Campbell:2007ws] and photoproduction [@Artoisenet:2009xh; @Kramer:1994zi], regarding both $J/\psi$ yield and polarization, by including the full CO contributions at NLO. While the NLO contributions due to the $^1\!S_0^{[8]}$ and $^3\!S_1^{[8]}$ CO states may be obtained using standard techniques [@Kramer:1994zi], the NLO treatment of $^3\!P_J^{[8]}$ states in $2\to2$ processes requires a more advanced technology, which has been lacking so far. In fact, the $^3\!P_J^{[8]}$ contributions represent the missing links in all those previous NLO analyses [@Campbell:2007ws; @Artoisenet:2009xh; @Kramer:1994zi], and there is no reason at all to expect them to be insignificant. Specifically, their calculation is far more intricate because the application of the $^3\!P_J^{[8]}$ projection operators to the short-distance scattering amplitudes produce particularly lengthy expressions involving complicated tensor loop integrals and exhibiting an entangled pattern of infrared singularities. This technical bottleneck, which has prevented essential progress in the global test of NRQCD factorization for the past fifteen years, was overcome for the first time in Ref. [@Butenschoen:2009zy], which we review here.
![Sample diagrams contributing at LO (a and d) and to the virtual (b and e) and real (c and f) NLO corrections.[]{data-label="fig:Examples"}](ExampleDiagrams.eps){width="6cm"}
In direct photoproduction, a quasi-real photon $\gamma$ that is radiated off the incoming electron $e$ interacts with a parton $i$ stemming from the incoming proton $p$. Invoking the Weizsäcker-Williams approximation and the factorization theorems of the QCD parton model and NRQCD [@Bodwin:1994jh], the inclusive $J/\psi$ photoproduction cross section is evaluated from $$d\sigma(ep\to J/\psi+X)
=\sum_{i,n} \int dxdy\, f_{\gamma/e}(x)f_{i/p}(y)
\langle{\cal O}^{J/\psi}[n]\rangle
d\sigma(\gamma i\to c\overline{c}[n]+X),
\label{Overview.Cross}$$ where $f_{\gamma/e}(x)$ is the photon flux function, $f_{i/p}(y)$ are the parton distribution functions (PDFs) of the proton, $\langle{\cal O}^{J/\psi}[n]\rangle$ are the LDMEs, and $d\sigma(\gamma i\to c\overline{c}[n]+X)$ are the partonic cross sections. Working in the fixed-flavor-number scheme, $i$ runs over the gluon $g$ and the light quarks $q=u,d,s$ and anti-quarks $\overline q$. The Fock states contributing through the order of our calculation include $n={}^3S_1^{[1]},{}^1S_0^{[8]},{}^3S_1^{[8]},{}^3P_J^{[8]}$. Example Feynman diagrams for partonic LO subprocesses $\gamma i\to c\overline{c}[n]+X$ as well as virtual- and real-correction diagrams are shown in Fig. \[fig:Examples\].
We now describe our theoretical input and the kinematic conditions for our numerical analysis. We set $m_c=m_{J/\psi}/2$, adopt the values of $m_{J/\psi}$, $m_e$, and $\alpha$ from Ref. [@Amsler:2008zzb], and use the one-loop (two-loop) formula for $\alpha_s^{(n_f)}(\mu)$, with $n_f=3$ active quark flavors, at LO (NLO). As for the proton PDFs, we use set CTEQ6L1 (CTEQ6M) [@Pumplin:2002vw] at LO (NLO), which comes with an asymptotic scale parameter of $\Lambda_\mathrm{QCD}^{(4)}=215$ MeV (326 MeV), so that $\Lambda_\mathrm{QCD}^{(3)}=249$ MeV (389 MeV). We evaluate the photon flux function using Eq. (5) of Ref. [@Kniehl:1996we] with the cut-off $Q_\mathrm{max}^2=2$ GeV$^2$ [@Adloff:2002ex; @H1.prelim] on the photon virtuality. Our default choices for the renormalization, factorization, and NRQCD scales are $\mu_r=\mu_f=m_T$ and $\mu_\Lambda=m_c$, respectively, where $m_T=\sqrt{p_T^2+4m_c^2}$ is the $J/\psi$ transverse mass. We adopt the LDMEs from Ref. [@Kniehl:1998qy], which were fitted to Tevatron I data using the CTEQ4 PDFs, because, besides the usual LO set, they also comprise a [*higher-order-improved*]{} set determined by approximately taking into account dominant higher-order effects due to multiple-gluon radiation in inclusive $J/\psi$ hadroproduction, which had been found to be substantial by a Monte Carlo study [@CanoColoma:1997rn]. We disentangle $\langle {\cal O}^{J/\psi}(^1\!S_0^{[8]}) \rangle$ and $\langle {\cal O}^{J/\psi}(^3\!P_0^{[8]}) \rangle$, a linear combination of which is fixed by the fit only, as in Ref. [@Klasen:2004tz]. The LO CO LDMEs are similar to the those obtained in Ref. [@Kniehl:2006qq] by fitting Tevatron II data using the CTEQ6L1 PDFs [@Pumplin:2002vw]. The higher-order-improved CO LDMEs are likely to undershoot the genuine ones, which are presently unknown.
Recently, the H1 Collaboration presented preliminary data on inclusive $J/\psi$ photoproduction taken in collisions of 27.6 GeV electrons or positrons on 920 GeV protons in the HERA II laboratory frame [@H1.prelim]. They nicely agree with their previous measurement at HERA I [@Adloff:2002ex]. These data come as singly differential cross sections in $p_T^2$, $W=\sqrt{(p_\gamma+p_p)^2}$, and $z=(p_{J/\psi}\cdot p_p)/(p_\gamma\cdot p_p)$, in each case with certain acceptance cuts on the other two variables. Here, $p_\gamma$, $p_p$, and $p_{J/\psi}$ are the photon, proton, and $J/\psi$ four-momenta, respectively. In the comparisons below, we impose the same kinematic conditions on our theoretical predictions.
----------------------------------------------- ------------------------------------------- -------------------------------------------
{width="4.7cm"} {width="4.7cm"} {width="4.7cm"}
----------------------------------------------- ------------------------------------------- -------------------------------------------
The H1 measurements [@Adloff:2002ex; @H1.prelim] of the $p_T^2$, $W$, and $z$ distributions of inclusive $J/\psi$ photoproduction are compared with our new NLO predictions in full NRQCD in Fig. \[fig:results\](a)–(c), respectively. The uncertainty due the LDMEs is indicated by shaded (yellow) bands, whose upper margins (solid lines) refer to the LO set. For comparison, also the default predictions at LO (dashed lines) as well as those of the CSM at NLO (dot-dashed lines) and LO (dotted lines) are shown. Notice that the experimental data are contaminated by the feed-down from heavier charmonia, mainly due to $\psi^\prime\to J/\psi+X$, which yields an estimated enhancement by about 15% [@Kramer:1994zi]. Furthermore, our predictions do not include resolved photoproduction, which contributes appreciably only at $z{\,\rlap{\lower 3.5 pt \hbox{$\mathchar \sim$}} \raise 1pt
\hbox {$<$}\,}0.3$ [@Kniehl:1998qy], and diffractive production, which is confined to the quasi-elastic domain at $z\approx1$ and $p_T\approx0$. These contributions are efficiently suppressed by the cut $0.3<z<0.9$ in Figs. \[fig:results\](a) and (b), so that our comparisons are indeed meaningful. We observe that the NLO corrections enhance the NRQCD cross section, by up to 115%, in the kinematic range considered, except for $z{\,\rlap{\lower 3.5 pt \hbox{$\mathchar \sim$}} \raise 1pt
\hbox {$<$}\,}0.45$, where they are negative. As may be seen from Fig. \[fig:results\](c), the familiar growth of the LO NRQCD prediction in the upper endpoint region, leading to a breakdown at $z=1$, is further enhanced at NLO. The solution to this problem clearly lies beyond the fixed-order treatment and may be found in soft collinear effective theory [@Fleming:2006cd]. The experimental data are nicely gathered in the central region of the error bands, except for the two low-$z$ points in Fig. \[fig:results\](c), which overshoot the NLO NRQCD prediction. However, this apparent disagreement is expected to fade away once the NLO-corrected NRQCD contribution due to resolved photoproduction is included. In fact, the above considerations concerning the large size of the NLO corrections to hadroproduction directly carry over to resolved photoproduction, which proceeds through the same partonic subprocesses. On the other hand, the default CSM predictions significantly undershoot the experimental data, by typically a factor of 4, which has already been observed in Ref. [@Artoisenet:2009xh]. Except for $p_T^2{\,\rlap{\lower 3.5 pt \hbox{$\mathchar \sim$}} \raise 1pt
\hbox {$>$}\,}4$ GeV$^2$, the situation is even deteriorated by the inclusion of the NLO corrections.
Despite the caveat concerning our limited knowledge of the CO LDMEs at NLO, we conclude that the H1 data [@Adloff:2002ex; @H1.prelim] show clear evidence of the existence of CO processes in nature, as predicted by NRQCD, supporting the conclusions previously reached for hadroproduction at the Tevatron [@Cho:1995vh] and two-photon collisions at LEP2 [@Klasen:2001cu]. In order to further substantiate this argument, it is indispensable to complete the NLO analysis of inclusive $J/\psi$ hadroproduction in NRQCD, by treating also the $^3\!P_J^{[8]}$ channels at NLO, so as to permit a genuine NLO fit of the relevant CO LDMEs to Tevatron and CERN LHC data. This goal is greatly facilitated by the technical advancement achieved in the present analysis.
This work was supported in part by BMBF Grant No. 05H09GUE, DFG Grant No. KN 365/6–1, and HGF Grant No. HA 101.
[99]{}
G.T. Bodwin, E. Braaten and G.P. Lepage, *Phys. Rev.* D [**51**]{} (1995) 1125; [**55**]{} (1997) 5853(E). P.L. Cho and A.K. Leibovich, *Phys. Rev.* D [**53**]{} (1996) 150; [**53**]{} (1996) 6203.
J. Campbell, F. Maltoni and F. Tramontano, *Phys. Rev. Lett.* [**98**]{} (2007) 252002; P. Artoisenet, J.P. Lansberg and F. Maltoni, *Phys. Lett.* B [**653**]{} (2007) 60; B. Gong and J.-X. Wang, *Phys. Rev. Lett.* [**100**]{} (2008) 232001; B. Gong, X.Q. Li and J.-X. Wang, *Phys. Lett.* B [**673**]{} (2009) 197. P. Artoisenet, J. Campbell, F. Maltoni and F. Tramontano, *Phys. Rev. Lett.* [**102**]{} (2009) 142001; C.-H. Chang, R. Li and J.-X. Wang, *Phys. Rev.* D [**80**]{} (2009) 034020. M. Cacciari and M. Krämer, *Phys. Rev. Lett.* [**76**]{} (1996) 4128; P. Ko, J. Lee and H.S. Song, *Phys. Rev.* D [**54**]{} (1996) 4312; [**60**]{} (1999) 119902(E). B.A. Kniehl and L. Zwirner, *Nucl. Phys.* B [**621**]{} (2002) 337; [**637**]{} (2002) 311; [**678**]{} (2004) 258. M. Klasen, B.A. Kniehl, L.N. Mihaila and M. Steinhauser, *Phys. Rev. Lett.* [**89**]{} (2002) 032001. E. Braaten, B.A. Kniehl and J. Lee, *Phys. Rev.* D [**62**]{} (2000) 094005; B.A. Kniehl and J. Lee, *ibid.* [**62**]{} (2000) 114027. M. Krämer, J. Zunft, J. Steegborn and P.M. Zerwas, *Phys. Lett.* B [**348**]{} (1995) 657; M. Krämer, *Nucl. Phys.* B [**459**]{} (1996) 3. M. Butenschön and B.A. Kniehl, *Phys. Rev. Lett.* [**104**]{} (2010) 072001. Particle Data Group, C. Amsler [*et al.*]{}, *Phys. Lett.* B [**667**]{} (2008) 1.
CTEQ Colaboration, J. Pumplin [*et al.*]{}, *JHEP* [**0207**]{} (2002) 012 . B.A. Kniehl, G. Kramer and M. Spira, *Z. Phys.* C [**76**]{} (1997) 689. H1 Collaboration, C. Adloff [*et al.*]{}, *Eur. Phys. J.* C [**25**]{} (2002) 25. H1 Collaboration, F.D. Aaron [*et al.*]{}, *DESY Report* 09–225 \[[arXiv:1002.0234 \[hep-ex\]]{}\].
B.A. Kniehl and G. Kramer, *Eur. Phys. J.* C [**6**]{} (1999) 493. B. Cano-Coloma and M.A. Sanchis-Lozano, *Nucl. Phys.* B [**508**]{} (1997) 753. M. Klasen, B.A. Kniehl, L.N. Mihaila and M. Steinhauser, *Nucl. Phys.* B [**713**]{} (2005) 487; *Phys. Rev.* D [**71**]{} (2005) 014016. B.A. Kniehl and C.P. Palisoc, *Eur. Phys. J.* C [**48**]{} (2006) 451. S. Fleming, A.K. Leibovich and T. Mehen, *Phys. Rev.* D [**74**]{} (2006) 114004.
| {
"pile_set_name": "ArXiv"
} |
---
author:
- Todd Karin
- Scott Dunham
- 'Kai-Mei Fu'
bibliography:
- 'toddkarin.bib'
- 'fu\_lab\_bib.bib'
date: 'May 20, 2014'
title: |
Alignment of the Diamond Nitrogen Vacancy Center by Strain Engineering:\
Supplemental Material
---
=1
The response of a bulk diamond crystal to strain along $z\propto [111]$ can be found by transforming the rank-four elastic constants tensor in the original coordinate system (\[100\],\[010\],\[001\]), $$c = \left(\begin{array}{cccccc}
c_{11} & c_{12} & c_{12} & 0 & 0&0 \\
c_{12} & c_{11} & c_{12} & 0 & 0&0 \\
c_{12} & c_{12} & c_{11} & 0 & 0&0 \\
0&0&0&c_{44}&0&0 \\
0&0&0&0&c_{44}&0 \\
0&0&0&0&0&c_{44} \\
\end{array}
\right),$$ into the new coordinate system ($[1\bar{1}0],\,[11\bar{2}],\,[111]$).[@Nye] Here we have written the strain tensor in its reduced form, where components 1,2,3 correspond to uniaxial strain along x,y,z and 4,5,6 to shear strains along $yz, xz, xy$. [@Dresselhaus] The generalized Hooke’s law strain energy is $$W = \sum_{ij}\frac{1}{2} \varepsilon_i c_{ij} \varepsilon_j .$$ If a strain $\varepsilon_3$ is applied in the $z$ direction, the diamond will accumulate a strain in $x,y$ in order to minimize the strain energy: $$\varepsilon_1 = \varepsilon_2 = - \overbrace{\frac{c_{11} +2 c_{12} -c_{44}}{2c_{11}+4c_{12}+c_{44}} }^{0.232 \pm 0.002} \varepsilon_3 , \quad \varepsilon_4 =\varepsilon_5 = \varepsilon_6 = 0.$$ Here we have used experimentally determined elastic constants for diamond.[@McSkimin1972b] Alternately, if biaxial compressive strain $\varepsilon_1 = \varepsilon_2$ is applied, then the diamond acquires a strain $$\varepsilon_3 = - \overbrace{\frac{c_{11} +2 c_{12} -c_{44}}{c_{11}+2c_{12}+2c_{44}} }^{0.302 \pm 0.003} (\varepsilon_1+ \varepsilon_2), \quad \varepsilon_4 =\varepsilon_5 = \varepsilon_6 = 0.$$ These proportionality constants were used to deform the supercell given an applied uniaxial $z$ or biaxial $x,y$ strain.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Given a sufficiently large training dataset, it is relatively easy to train a modern convolution neural network (CNN) as a required image classifier. However, for the task of fish classification and/or fish detection, if a CNN was trained to detect or classify particular fish species in particular background habitats, the same CNN exhibits much lower accuracy when applied to new/unseen fish species and/or fish habitats. Therefore, in practice, the CNN needs to be continuously fine-tuned to improve its classification accuracy to handle new project-specific fish species or habitats. In this work we present a labelling-efficient method of training a CNN-based fish-detector (the Xception CNN was used as the base) on relatively small numbers (4,000) of project-domain underwater fish/no-fish images from 20 different habitats. Additionally, 17,000 of known negative (that is, missing fish) general-domain (VOC2012) above-water images were used. Two publicly available fish-domain datasets supplied additional 27,000 of above-water and underwater positive/fish images. By using this multi-domain collection of images, the trained Xception-based binary (fish/not-fish) classifier achieved 0.17% false-positives and 0.61% false-negatives on the project’s 20,000 negative and 16,000 positive holdout test images, respectively. The area under the ROC curve (AUC) was 99.94%.'
author:
-
-
-
-
-
-
title: 'Underwater Fish Detection with Weak Multi-Domain Supervision'
---
fish, detection, convolution neural network, image, video
Materials and methods
=====================
[10]{} \[1\][\#1]{} url@samestyle \[2\][\#2]{} \[2\][[l@\#1=l@\#1\#2]{}]{}
F. Shafait, A. Mian, M. Shortis, B. Ghanem, P. F. Culverhouse, D. Edgington, D. Cline, M. Ravanbakhsh, J. Seager, and E. S. Harvey, “Fish identification from videos captured in uncontrolled underwater environments,” *ICES Journal of Marine Science*, vol. 73, pp. 2737–2746, 2016.
S. A. Siddiqui, A. Salman, M. I. Malik, F. Shafait, A. Mian, M. R. Shortis, and E. S. Harvey, “Automatic fish species classification in underwater videos: exploiting pre-trained deep neural network models to compensate for limited labelled data,” *ICES Journal of Marine Science*, vol. 75, pp. 374–389, 2018.
Y. LeCun, Y. Bengio, and G. Hinton, “Deep learning,” *Nature*, vol. 521, pp. 436–444, 2015.
A. Rivas, P. Chamoso, A. González-Briones, and J. M. Corchado, “Detection of cattle using drones and convolutional neural networks,” *Sensors*, vol. 18, no. 7, 2018.
F. Chollet, “Xception: Deep learning with depthwise separable convolutions,” in *2017 IEEE Conference on Computer Vision and Pattern Recognition (CVPR2017)*, 2017.
A. Joly, H. Go[ë]{}au, H. Glotin, C. Spampinato, P. Bonnet, W.-P. Vellinga, R. Planque, A. Rauber, R. Fisher, and H. M[ü]{}ller, “Lifeclef 2014: Multimedia life species identification challenges,” in *Information Access Evaluation. Multilinguality, Multimodality, and Interaction*, ser. Lecture Notes in Computer Science, E. Kanoulas, M. Lupu, P. Clough, M. Sanderson, M. Hall, A. Hanbury, and E. Toms, Eds., vol. 8685.1em plus 0.5em minus 0.4emCham: Springer International Publishing, 2014, pp. 229–249.
B. J. Boom, J. He, S. Palazzo, P. X. Huang, C. Beyan, H.-M. Chou, F.-P. Lin, C. Spampinato, and R. B. Fisher, “A research tool for long-term and continuous analysis of fish assemblage in coral-reefs using underwater camera footage,” *Ecological Informatics*, vol. 23, pp. 83 – 97, 2014, special Issue on Multimedia in Ecology and Environment.
R. B. Fisher *et al.* \[Online\]. Available: [www.fish4knowledge.eu](www.fish4knowledge.eu)
——. \[Online\]. Available: <https://bit.ly/2Ex7dnZ>
C. Spampinato, S. Palazzo, P. H. Joalland, S. Paris, H. Glotin, K. Blanc, D. Lingrand, and F. Precioso, “Fine-grained object recognition in underwater visual data,” *Multimedia Tools and Applications*, vol. 75, pp. 1701–1720, 2016.
A. Vedaldi and B. Fulkerson, “[VLFeat]{}: An open and portable library of computer vision algorithms,” 2008. \[Online\]. Available: <http://www.vlfeat.org>
——, “[VLfeat]{}: An open and portable library of computer vision algorithms,” in *Proceedings of the 18th ACM International Conference on Multimedia*, ser. MM ’10.1em plus 0.5em minus 0.4emNew York, NY, USA: ACM, 2010, pp. 1469–1472.
A. Salman, A. Jalal, F. Shafait, A. Mian, M. Shortis, J. Seager, and E. Harvey, “Fish species classification in unconstrained underwater environments based on deep learning,” *Limnology and Oceanography: Methods*, vol. 14, pp. 570–585, 2016.
O. Barnich and M. V. Droogenbroeck, “[ViBe]{}: A universal background subtraction algorithm for video sequences,” *IEEE Transactions on Image Processing*, vol. 20, no. 6, pp. 1709–1724, 2011.
K. Blanc, D. Lingrand, and F. Precioso, “Fish species recognition from video using svm classifier,” in *Proceedings of the 3rd ACM International Workshop on Multimedia Analysis for Ecological Data*, ser. MAED ’14. 1em plus 0.5em minus 0.4emNew York, NY, USA: ACM, 2014, pp. 1–6.
A. Krizhevsky, I. Sutskever, and G. E. Hinton, “Imagenet classification with deep convolutional neural networks,” in *Proceedings of the 25th International Conference on Neural Information Processing Systems - Volume 1*, ser. NIPS’12.1em plus 0.5em minus 0.4emUSA: Curran Associates Inc., 2012, pp. 1097–1105.
R. Girshick, “Fast r-cnn,” in *The IEEE International Conference on Computer Vision (ICCV)*, December 2015.
X. Li, M. Shang, H. Qin, and L. Chen, “Fast accurate fish detection and recognition of underwater images with fast r-cnn,” in *OCEANS 2015 - MTS/IEEE Washington*, 2015, pp. 1–5.
Y. Hu, A. S. Mian, and R. Owens, “Face recognition using sparse approximated nearest points between image sets,” *IEEE Transactions on Pattern Analysis and Machine Intelligence*, vol. 34, pp. 1992–2004, 2012.
Y.-H. Hsiao, C.-C. Chen, S.-I. Lin, and F.-P. Lin, “Real-world underwater fish recognition and identification, using sparse representation,” *Ecological Informatics*, vol. 23, pp. 13 – 21, 2014, special Issue on Multimedia in Ecology and Environment.
C. Spampinato, Y.-H. Chen-Burger, G. Nadarajan, and R. B. Fisher, “Detecting, tracking and counting fish in low quality unconstrained underwater videos,” in *Proceedings of 3rd International Conference on Computer Vision Theory and Applications (VISAPP)*, vol. 2, 2008, pp. 514–519.
Z. Zivkovic and F. van der Heijden, “Efficient adaptive density estimation per image pixel for the task of background subtraction,” *Pattern Recognition Letters*, vol. 27, pp. 773 – 780, 2006.
Itseez, “Open source computer vision library,” <https://github.com/itseez/opencv>, 2017.
A. Joly, H. Go[ë]{}au, H. Glotin, C. Spampinato, P. Bonnet, W.-P. Vellinga, R. Planqu[é]{}, A. Rauber, S. Palazzo, B. Fisher, and H. M[ü]{}ller, “Lifeclef 2015: Multimedia life species identification challenges,” in *Experimental IR Meets Multilinguality, Multimodality, and Interaction*, ser. Lecture Notes in Computer Science, J. Mothe, J. Savoy, J. Kamps, K. Pinel-Sauvagnat, G. Jones, E. San Juan, L. Capellato, and N. Ferro, Eds., vol. 9283.1em plus 0.5em minus 0.4emCham: Springer International Publishing, 2015, pp. 462–483.
“Fish species recognition.” \[Online\]. Available: <https://bit.ly/2LomRTp>
E. S. Harvey, M. Cappo, G. A. Kendrick, and D. L. McLean, “Coastal fish assemblages reflect geological and oceanographic gradients within an australian zootone,” *PLOS ONE*, vol. 8, 2013.
K. Simonyan and A. Zisserman, “Very deep convolutional networks for large-scale image recognition,” *CoRR*, vol. abs/1409.1556, 2014.
K. He, X. Zhang, S. Ren, and J. Sun, “Deep residual learning for image recognition,” in *2016 IEEE Conference on Computer Vision and Pattern Recognition (CVPR)*, 2016, pp. 770–778.
J. Deng, W. Dong, R. Socher, L. J. Li, K. Li, and L. Fei-Fei, “Imagenet: A large-scale hierarchical image database,” in *2009 IEEE Conference on Computer Vision and Pattern Recognition*, 2009, pp. 248–255.
A. S. Razavian, H. Azizpour, J. Sullivan, and S. Carlsson, “Cnn features off-the-shelf: An astounding baseline for recognition,” in *2014 IEEE Conference on Computer Vision and Pattern Recognition Workshops*, 2014, pp. 512–519.
M. Oquab, L. Bottou, I. Laptev, and J. Sivic, “Learning and transferring mid-level image representations using convolutional neural networks,” in *2014 IEEE Conference on Computer Vision and Pattern Recognition*. 1em plus 0.5em minus 0.4emIEEE, 2014, pp. 1717–1724.
S. Marini, E. Fanelli, V. Sbragaglia, E. Azzurro, J. Del Rio Fernandez, and J. Aguzzi, “Tracking fish abundance by underwater image recognition,” *Scientific Reports*, vol. 8, p. 13748, 2018.
“The western mediterranean expandable seafloor observatory ([OBSEA]{}).” \[Online\]. Available: <http://www.obsea.es>
L. Corgnati, S. Marini, L. Mazzei, E. Ottaviani, S. Aliani, A. Conversi, and A. Griffa, “Looking inside the ocean: Toward an autonomous imaging system for monitoring gelatinous zooplankton,” *Sensors*, vol. 16, 2016.
K. Zuiderveld, *“Contrast Limited Adaptive Histogram Equalization”*. 1em plus 0.5em minus 0.4emSan Diego: “Academic Press Professional”, 1994, pp. 474–485.
M. Bradley, R. Baker, I. Nagelkerken, and M. Sheaves, “Context is more important than habitat type in determining use by juvenile fish,” *Landscape Ecology*, 2019, in press.
, *Specification ICC.1:2004-10 (Profile version 4.2.0.0) Image technology colour management - Architecture, profile format, and data structure*, 2004. \[Online\]. Available: <http://www.color.org/icc1v42.pdf>
O. Russakovsky, A. L. Bearman, V. Ferrari, and F. Li, “What’s the point: Semantic segmentation with point supervision,” *CoRR*, vol. abs/1506.02106, 2015.
F. Chollet *et al.*, “Keras: The python deep learning library,” 2015. \[Online\]. Available: <https://keras.io/>
M. Abadi *et al.*, “[TensorFlow]{}: Large-scale machine learning on heterogeneous systems,” 2015. \[Online\]. Available: <http://tensorflow.org/>
D. P. Kingma and J. Ba, “Adam: [A]{} method for stochastic optimization,” *CoRR*, vol. abs/1412.6980, 2014.
M. Everingham, S. M. A. Eslami, L. Van Gool, C. K. I. Williams, J. Winn, and A. Zisserman, “The pascal visual object classes challenge: A retrospective,” *International Journal of Computer Vision*, vol. 111, pp. 98–136, 2015.
K. Anantharajah, Z. Ge, C. McCool, S. Denman, C. Fookes, P. Corke, D. Tjondronegoro, and S. Sridharan, “Local inter-session variability modelling for object classification,” in *IEEE Winter Conference on Applications of Computer Vision*, 2014, pp. 309–316.
“[QUT]{} fish dataset.” \[Online\]. Available: <https://bit.ly/2APDvGB>
I. H. Laradji, N. Rostamzadeh, P. O. Pinheiro, D. Vazquez, and M. Schmidt, “Where are the blobs: Counting by localization with point supervision,” in *Computer Vision – ECCV 2018*, V. Ferrari, M. Hebert, C. Sminchisescu, and Y. Weiss, Eds.1em plus 0.5em minus 0.4emCham: Springer International Publishing, 2018, pp. 560–576.
T. Fawcett, “An introduction to [ROC]{} analysis,” *Pattern Recognition Letters*, vol. 27, pp. 861–874, 2006.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The interpolation of couples of separable Hilbert spaces with a function parameter is studied. The main properties of the classic interpolation are proved. Some applications to the interpolation of isotropic Hörmander spaces over a closed manifold are given.'
address:
- 'Institute of Mathematics NAS of Ukraine, Tereshchenkivska str. 3, Kyiv, Ukraine, 01601'
- 'Institute of Mathematics NAS of Ukraine, Tereshchenkivska str. 3, Kyiv, Ukraine, 01601; Chernigiv State Technological University, Shevchenka str. 95, Chernigiv 14027, Ukraine'
author:
- 'Vladimir A. Mikhailets, Alexandr A. Murach'
date: '07/12/2007'
title: |
Interpolation with a function parameter and\
refined scale of spaces
---
Introduction
============
In this paper we study the interpolation of couples of separable Hilbert spaces with a functional parameter. We generalize the classical theorems on interpolation with a power parameter of order $\theta\in(0,1)$ to the maximal class of functions.
As an application, we consider the interpolation of isotropic Hörmander spaces over a closed manifold $$H^{s,\varphi}:=H_{2}^{\langle\cdot\rangle^{s}\,\varphi(\langle\cdot\rangle)},
\quad\langle\xi\rangle:=\bigl(1+|\xi|^{2}\bigr)^{1/2}. \leqno(1.1)$$ Here, $s\in\mathbb{R}$ and $\varphi$ is a functional parameter slowly varying at $+\infty$ in Karamata’s sense. In particular, every standard function $$\varphi(t)=(\log t)^{r_{1}}(\log\log
t)^{r_{2}}\ldots(\log\ldots\log
t)^{r_{n}},\quad\{r_{1},r_{2},\ldots,r_{n}\}\subset\mathbb{R},\;n\in\mathbb{N},$$ is admissible. This scale was introduced and investigated by the authors in [@MM2a; @MM2b]. It contains the Sobolev scale $\{H^{s}\}\equiv\{H^{s,1}\}$ and is attached to it by the number parameter $s$ and being considerably finer.
Spaces of form (1.1) arise naturally in different spectral problems: convergence of spectral expansions of self-adjoint elliptic operators almost everywhere, in the norm of the spaces $L_{p}$ with $p>2$ or $C$ (see survey [@AIN76]); spectral asymptotics of general self-adjoint elliptic operators in a bounded domain, the Weyl formula, a sharp estimate of the remainder in it (see [@Mikh82; @Mikh89]) and others. They may be expected to be useful in other “fine” questions. Due to their interpolation properties, the spaces $H^{s,\varphi}$ occupy a special position among the spaces of a generalized smoothness which are actively investigated and used today (see survey [@KalLiz87], recent articles [@HarMou04; @FarLeo06] and the bibliography given there).
One of the main results of the article is a description of the refined scale by means of regularly varying functions of a positive elliptic pseudodifferential operator.
The related questions were studied in [@Merucci84; @CobFern86] and by the authors in \[11–20\].
An interpolation with a function parameter
==========================================
A definition of the interpolation.
----------------------------------
An ordered couple $[X_{0},X_{1}]$ of complex Hilbert spaces $X_{0}$ and $X_{1}$ is called *admissible* if the spaces $X_{0}$, $X_{1}$ are separable and the continuous imbedding $X_{1}\hookrightarrow X_{0}$ holds.
Let an admissible couple $X=[X_{\,0},X_{1}]$ of Hilbert spaces be given. It is known [@LM71 Ch. 1, Sec. 2.1], [@Kr72 Ch. IV, Sec. 9.1] that for this couple $X$ there exists the isometric isomorphism $J:X_{1}\leftrightarrow X_{\,0}$ such that $J$ is a self-adjoint positive operator on the space $X_{\,0}$ with the domain $X_{1}$. The operator $J$ is called a *generating* one for the couple $X$. This operator is uniquely determined by the couple$X$. Indeed, assume that $J_{1}$ is also a generating operator for the couple $X$. Then the operators $J$ and $J_{1}$ are metrically equal, that is $\|Ju\|_{X_{\,0}}=\|u\|_{X_{1}}=\|J_{1}u\|_{X_{\,0}}$ for any $u\in X_{1}$. Moreover, these operators are positive. Hence, they are equal.
We denote by $\mathcal{B}$ the set of all functions $\psi:(0,+\infty)\rightarrow(0,+\infty)$ such that
a)
: $\psi$ is Borel measurable on the semiaxis $(0,+\infty)$;
b)
: $\psi$ is bounded on each closed interval $[a,b]$, where $0<a<b<+\infty$;
c)
: $1/\psi$ is bounded on each set $[r,+\infty)$, where $r>0$.
Let $\psi\in\mathcal{B}$. Generally, the unbounded operator $\psi(J)$ is defined in the space $X_{0}$ as a function of $J$. We denote by $[X_{0},X_{1}]_{\psi}$ or, simply, by $X_{\psi}$ the domain of the operator $\psi(J)$, equipped with the inner product $(u,v)_{X_{\psi}}:=(\psi(J)u,\psi(J)v)_{X_{0}}$ and the corresponding norm $\|\,u\,\|_{X_{\psi}}=(u,u)_{X_{\psi}}^{1/2}$.
The space $X_{\psi}$ is the Hilbert separable one and, moreover, the continuous dense imbedding $X_{\psi}\hookrightarrow X_{0}$ is fulfilled. Indeed, we have $\mathrm{Spec}\,J\subseteq[r,+\infty)$ and $\psi(t)\geq c$ for $t\geq r$, where $r,c$ are some positive numbers. Hence, $\mathrm{Spec}\,\psi(J)\subseteq[c,+\infty)$, that implies the isometric isomorphism $\psi(J):X_{\psi}\leftrightarrow
X_{0}$. It follows that the space $X_{\psi}$ is complete and separable as well as that the function $\|\cdot\|_{X_{\psi}}$ is positive definite, so this is a norm. Next, since the operator $\psi^{-1}(J)$ is bounded in the space $X_{0}$, a bounded imbedding operator $I=\psi^{-1}(J)\psi(J):X_{\psi}\rightarrow
X_{0}$ exists. The imbedding $X_{\psi}\hookrightarrow X_{0}$ is dense because the domain of the operator $\psi(J)$ is a dense linear manifold in the space $X_{0}$.
Further, it is useful to note the following. Let functions $\varphi,\psi\in\mathcal{B}$ be such that $\varphi\asymp\psi$ in a neighborhood of $+\infty$. Then, by the definition of the set $\mathcal{B}$, we have $\varphi\asymp\psi$ on $\mathrm{Spec}\,J$. Hence, $X_{\varphi}=X_{\psi}$ up to equivalent norms.
A function $\psi\in\mathcal{B}$ is called an *interpolation parameter* if the following condition is satisfied for all admissible couples $X=[X_{0},X_{1}]$, $Y=[Y_{0},Y_{1}]$ of Hilbert spaces and an arbitrary linear mapping $T$ given on $X_{0}$: if the restriction of the mapping $T$ to the space $X_{j}$ is a bounded operator $T:X_{j}\rightarrow Y_{j}$ for each $j=0,\,1$, then the restriction of the mapping $T$ to the space $X_{\psi}$ is also a bounded operator $T:X_{\psi}\rightarrow Y_{\psi}$.
Otherwise speaking, $\psi$ is an interpolation parameter if and only if the mapping $X\mapsto X_{\psi}$ is an interpolation functor given on the category of all admissible couples $X$ of Hilbert spaces [@Tr80 Sec. 1.2.2], [@BL80 Sec. 2.4]. In the case where $\psi$ is an interpolation parameter, we say that the space $X_{\psi}$ is obtained by the *interpolation with the function parameter* $\psi$ of the admissible couple $X$.
Further we investigate the main properties of the mapping $X\mapsto X_{\psi}$.
Imbeddings of spaces.
---------------------
Let $\psi\in\mathcal{B}$ be an interpolation parameter and $X=[X_{0},X_{1}]$ be an admissible couple of Hilbert spaces. Then the continuous dense imbeddings $X_{1}\hookrightarrow X_{\psi}\hookrightarrow X_{0}$ hold.
According to Subsection 2.1 it only remains to prove the continuous dense imbedding $X_{1}\hookrightarrow X_{\psi}$. Consider two bounded imbedding operators $I:X_{1}\rightarrow
X_{0}$ and $I:X_{1}\rightarrow X_{1}$. Since $\psi$ is an interpolation parameter, these operators imply the bounded imbedding operator $I:X_{1}\rightarrow X_{\psi}$. Thus, the continuous imbedding $X_{1}\hookrightarrow X_{\psi}$ is valid. We will prove that it is dense. For an arbitrary $u\in X_{\psi}$, we have $v:=(1+\psi^{2}(J))^{1/2}\,u\in X_{0}$. Since $X_{1}$ is dense in $X_{0}$, the sequence $(v_{k})\subset X_{1}$ such that $v_{k}\rightarrow v$ in $X_{0}$ as $k\rightarrow\infty$ exists. From this and from (1.1) it follows that $$u_{k}:=(1+\psi^{2}(J))^{-1/2}\,v_{k}\rightarrow
u\quad\mbox{â}\quad X_{\psi}\quad\mbox{for}\quad
k\rightarrow\infty.$$ It remains to note that $$u_{k}=(1+\psi^{2}(J))^{-1/2}J^{-1}Jv_{k}=J^{-1}(1+\psi^{2}(J))^{-1/2}Jv_{k}
\in X_{1}.$$ Theorem 2.1 is proved.
Let functions $\psi,\chi\in\mathcal{B}$ be such that the function $\psi/\chi$ is bounded in a neighborhood of $+\infty$. Then, for each admissible couple $X=[X_{0},X_{1}]$ of Hilbert spaces, the continuous and dense imbedding $X_{\chi}\hookrightarrow X_{\psi}$ holds. If the imbedding $X_{1}\hookrightarrow X_{0}$ is compact and $\psi(t)/\chi(t)\rightarrow0$ as $t\rightarrow+\infty$, then the imbedding $X_{\chi}\hookrightarrow X_{\psi}$ is also compact.
Let $J$ be a genarating operator for the couple $X$. Let us note that $\mathrm{Spec}\,J\subseteq[r,+\infty)$ for some number $r>0$. According to the condition of the theorem, we have $\psi(t)/\chi(t)\leq c$ for $t\geq r$. Therefore $$X_{\chi}=\mathrm{Dom}\,\chi(J)\subseteq\mathrm{Dom}\,\psi(J)=X_{\psi},
\quad\|\psi(J)\,u\|_{X_{0}}\leq c\,\|\chi(J)\,u\|_{X_{0}}.$$ From this formulae and from the definition of the spaces $X_{\chi}$, $X_{\psi}$ we obtain the continuous imbedding $X_{\chi}\hookrightarrow X_{\psi}$. Let us prove its density.
We consider the isometric isomorphisms $\psi(J):X_{\psi}\leftrightarrow X_{0}$ and $\chi(J):X_{\chi}\leftrightarrow X_{0}$. For any given $u\in
X_{\psi}$, we have $\psi(J)\,u\in X_{0}$. Since the space $X_{\chi}$ is densely embedded into $X_{0}$, a sequence $(v_{k})\subset X_{\chi}$ such that $v_{k}\rightarrow\psi(J)\,u$ in $X_{0}$ as $k\rightarrow\infty$ exists. Hence, $\psi^{-1}(J)\,v_{k}\rightarrow u$ in $X_{\psi}$ as $k\rightarrow\infty$, where $$\psi^{-1}(J)\,v_{k}=\psi^{-1}(J)\,\chi^{-1}(J)\,
\chi(J)\,v_{k}=\chi^{-1}(J)\,\psi^{-1}(J)\,\chi(J)\,v_{k} \in
X_{\chi}.$$ Thus, we have proved the density of the embedding $X_{\chi}\hookrightarrow X_{\psi}$.
Now let us assume that the imbedding $X_{1}\hookrightarrow X_{0}$ is compact and $\psi(t)/\chi(t)\rightarrow0$ as $t\rightarrow+\infty$. We will prove the compactness of the embedding $X_{\chi}\hookrightarrow X_{\psi}$. Let $(u_{k})$ be an arbitrary bounded sequence belonging to $X_{\chi}$. Since the sequence of elements $w_{k}:=J^{-1}\,\chi(J)\,u_{k}$ is bounded in $X_{1}$, we can select a subsequence of elements $w_{k_{n}}=J^{-1}\,\chi(J)\,u_{k_{n}}$ being the Cauchy sequence in $X_{0}$. We show that $(u_{k_{n}})$ is the Cauchy sequence in $X_{\psi}$.
Let $E_{t}$, $t\geq r$, be a resolution of the unity in $X_{0}$, corresponding to the self-adjoint operator$J$. We can write $$\|u_{k_{n}}-u_{k_{m}}\|_{X_{\psi}}^{2}=
\|\psi(J)\,(u_{k_{n}}-u_{k_{m}})\|_{X_{0}}^{2}
=\|\psi(J)\,\chi^{-1}(J)\,J\,(w_{k_{n}}-w_{k_{m}})\|_{X_{0}}^{2}$$ $$=\int_{r}^{+\infty}\,\psi^{2}(t)\,\chi^{-2}(t)\,t^{2}\:
d\,\|E_{t}(w_{k_{n}}-w_{k_{m}})\|_{X_{0}}^{2}. \leqno(2.1)$$ Let us choose an arbitrary number $\varepsilon>0$. There is a number $\rho=\rho(\varepsilon)>r$ such that $$\psi(t)/\chi(t)\leq(2c_{0})^{-1}\varepsilon\quad\mbox{for}\quad
t\geq\rho\quad\mbox{and}\quad
c_{0}:=\sup\,\{\,\|w_{k}\|_{X_{1}}:\,k\in\mathbb{N}\,\}<\infty.$$ Hence, for all indices $n,m$ we have $$\int_{\rho}^{+\infty}\psi^{2}(t)\,\chi^{-2}(t)\,t^{2}\:
d\,\|E_{t}(w_{k_{n}}-w_{k_{m}})\|_{X_{0}}^{2}\,\leq\,
(2c_{0})^{-2}\,\varepsilon^{2}\,\int_{\rho}^{+\infty}\,t^{2}\:
d\,\|E_{t}(w_{k_{n}}-w_{k_{m}})\|_{X_{0}}^{2}$$ $$\leq\,(2c_{0})^{-2}\,\varepsilon^{2}\,\|J\,(w_{k_{n}}-w_{k_{m}})\|_{X_{0}}^{2}\,
=\,(2c_{0})^{-2}\,\varepsilon^{2}\,\|w_{k_{n}}-w_{k_{m}}\|_{X_{1}}^{2}\,\leq
\,\varepsilon^{2} \leqno(2.2)$$ In addition, by the inequality $\psi(t)/\chi(t)\leq c$ for $t\geq
r$, we can write the following: $$\int_{r}^{\rho}\,\psi^{2}(t)\,\chi^{-2}(t)\,t^{2}\:
d\,\|E_{t}(w_{k_{n}}-w_{k_{m}})\|_{X_{0}}^{2}\,\leq\,c^{2}\rho^{2}\,
\int_{r}^{\rho}\,d\,\|E_{t}(w_{k_{n}}-w_{k_{m}})\|_{X_{0}}^{2}$$ $$\leq
c^{2}\rho^{2}\,\|w_{k_{n}}-w_{k_{m}}\|_{X_{0}}^{2}\rightarrow0
\quad\mbox{as}\quad n,m\rightarrow\infty. \leqno(2.3)$$ Now formulae (2.1) — (2.3) imply the inequality $\|u_{k_{n}}-u_{k_{m}}\|_{X_{\psi}}\leq2\varepsilon$ for sufficiently large $n,m$. Therefore $(u_{k_{n}})$ is the Cauchy sequence in the space $X_{\psi}$ which means the compactness of the imbedding $X_{\chi}\hookrightarrow X_{\psi}$. Theorem 2.2 is proved.
Reiteration.
------------
Let functions $f,g,\psi\in\mathcal{B}$ be given. Suppose that the function $f/g$ is bounded in a neighborhood of $+\infty$. Then $[X_{f},X_{g}]_{\psi}=X_{\omega}$ holds with the equality of norms for each admissible couple $X$ of Hilbert spaces. Here the function $\omega\in\mathcal{B}$ is given by the formula $\omega(t):=f(t)\,\psi(g(t)/f(t))$ for $t>0$. If $f,g,\psi$ are interpolation parameters, so is $\omega$.
Since the function $f/g$ is bounded in a neighborhood of $+\infty$, the couple $[X_{f},X_{g}]$ is admissible by Theorem 2.2 and, in addition, $\omega\in\mathcal{B}$. So, the spaces $[X_{f},X_{g}]_{\psi}$ and $X_{\omega}$ are well defined. We will prove them to be equal.
Let an operator $J$ be generating for the couple $X=[X_{0},X_{1}]$, where $\mathrm{Spec}\,J\subseteq[r,+\infty)$ for some number $r>0$. We have three isometric isomorphisms $$f(J):X_{f}\leftrightarrow X_{0},\quad g(J):X_{g}\leftrightarrow
X_{0},\quad B:=f^{-1}(J)\,g(J):X_{g}\leftrightarrow X_{f}.$$ Let us consider $B$ as a closed operator in the space $X_{f}$, defined on $X_{g}$. The operator $B$ is generating for the couple $[X_{f},X_{g}]$ because $B$ is positive and self-adjoint on $X_{f}$. The positiveness of $B$ follows from the condition $f(t)/g(t)\leq c$ for $t\geq r$ which implies $$(Bu,u)_{X_{f}}=(g(J)\,u,f(J)\,u)_{X_{0}}\geq
c^{-1}\,(f(J)\,u,f(J)\,u)_{X_{0}}= c^{-1}\,\|u\|_{X_{f}}^{2}.$$ The self-adjointness follows from the fact that $0$ is a regular point for the operator $B$.
Using the spectral theorem, we reduce the self-adjoint on $X_{0}$ operator $J$ to the form of multiplication by a function: $J=I^{-1}(\alpha\cdot I)$. Here, $I:X_{0}\leftrightarrow
L_{2}(U,d\mu)$ is an isometric isomorphism, $(U,\mu)$ is a space with a finite measure, $\alpha:U\rightarrow[r,+\infty)$ is a measurable function. The isometric isomorphism $If(J):X_{f}\leftrightarrow L_{2}(U,d\mu)$ reduces the self-adjoint on $X_{f}$ operator $B$ to the form of multiplication by the function $(g/f)\circ\alpha$: $$If(J)\,B\,u=Ig(J)\,u=(g\circ\alpha)\,Iu=(g\circ\alpha)\,If^{-1}(J)f(J)\,u=
((g/f)\circ\alpha)\,If(J)\,u,\quad u\in X_{g}.$$ Therefore, for an arbitrary $u\in X_{\omega}$, we have $$\|\psi(B)\,u\|_{X_{f}}=
\|(\psi\circ(g/f)\circ\alpha)\cdot(If(J)\,u)\|_{L_{2}(U,d\mu)}=
\|(\omega\circ\alpha)\cdot(Iu)\|_{L_{2}(U,d\mu)}=\|\omega(J)\,u\|_{X_{0}}.$$ Note that the function $f/\omega$ is bounded in a neighborhood of $+\infty$. Hence, $X_{\omega}\hookrightarrow X_{f}$ and the expression $f(J)\,u$ is well defined. Thus, the equality $[X_{f},X_{g}]_{\psi}=X_{\omega}$ is proved.
We now assume that $f,g,\psi$ are interpolation parameters. We show that $\omega$ is also an interpolation parameter. Let arbitrary admissible couples $X=[X_{0},X_{1}]$, $Y=[Y_{0},Y_{1}]$ and a linear mapping $T$ be the same as those in Definition 2.2. We have the bounded operators $T:X_{f}\rightarrow Y_{f}$ and $T:X_{g}\rightarrow Y_{g}$ which imply the boundedness of the operator $T:[X_{f},X_{g}]_{\psi}\rightarrow[Y_{f},Y_{g}]_{\psi}$. We have already proved that $[X_{f},X_{g}]_{\psi}=X_{\omega}$ and $[Y_{f},Y_{g}]_{\psi}=Y_{\omega}$. So, a bounded operator $T:X_{\omega}\rightarrow Y_{\omega}$ exists. It means that $\omega$ is an interpolation parameter. Theorem 2.3 is proved.
The interpolation of dual spaces.
---------------------------------
Let $H$ be a Hilbert space. We denote by $H'$ the space dual to $H$. Thus, $H'$ is the Banach space of all linear continuous functionals $l:H\rightarrow\mathbb{C}$. By the Riesz theorem, the mapping $S:v\mapsto(\,\cdot,v)_{H}$, where $v\in H$, establishes the antilinear isometric isomorphism $S:H\leftrightarrow H'$. This implies that $H'$ is the Hilbert space with respect to the inner product $(l,m)_{H'}:=(S^{-1}l,S^{-1}m)_{H}$. We emphasize that we do not identify $H'$ as $H$ by means of the isomorphism $S$.
Let $\psi\in\mathcal{B}$ be such that the function $\psi(t)/t$ is bounded in a neighborhood of $+\infty$. Then, for each admissible couple $[X_{0},X_{1}]$ of Hilbert spaces, the equality of spaces $[X_{1}',X_{0}']_{\psi}=[X_{0},X_{1}]_{\chi}'$ with the equality of norms holds. Here the function $\chi\in\mathcal{B}$ is given by the formula $\chi(t):=t/\psi(t)$ for $t>0$. If $\psi$ is an interpolation parameter, so is $\chi$.
Note that the couple $[X_{1}',X_{0}']$ is admissible, provided that we naturally identify functionals from $X_{0}'$ as their restrictions to the space $X_{1}$. From the condition of the theorem it follows that $\varphi\in\mathcal{B}$. Thus, the spaces $[X_{1}',X_{0}']_{\psi}$ and $[X_{0},X_{1}]_{\chi}'$ are well defined. Let us prove these spaces to be equal.
Let $J:X_{1}\leftrightarrow X_{0}$ be a generating operator for the couple $[X_{0},X_{1}]$. Let us consider the isometric isomorphisms $S_{j}:X_{j}\leftrightarrow X_{j}'$, $j=0,\,1$, which appear in the Riesz theorem. The operator $J'$, being adjoint to $J$, satisfies the equality $J'=S_{1}J^{-1}S_{0}^{-1}$. This results from the following: $$(J'l)u=l(Ju)=(Ju,S_{0}^{-1}l)_{X_{0}}=(u,J^{-1}S_{0}^{-1}l)_{X_{1}}=
(S_{1}J^{-1}S_{0}^{-1}l)u\quad\mbox{for each}\quad l\in X_{0}',\;
u\in X_{1}.$$ Thus, the isometric isomorphism $$J'=S_{1}J^{-1}S_{0}^{-1}:\,X_{0}'\leftrightarrow X_{1}'
\leqno(2.4)$$ exists.
Let us note that the equalities $$(u,JS_{1}^{-1}l)_{X_{0}}=(J^{-1}u,S_{1}^{-1}l)_{X_{1}}=l(J^{-1}u),\quad
(u,J^{-1}S_{0}^{-1}l)_{X_{0}}=(J^{-1}u,S_{0}^{-1}l)_{X_{0}}=l(J^{-1}u),$$ where $l\in X_{0}'\hookrightarrow X_{1}'$, $u\in X_{0}$, imply the property $$JS_{1}^{-1}l=J^{-1}S_{0}^{-1}l\in X_{1}\quad\mbox{for each}\quad
l\in X_{0}'.\leqno(2.5)$$
Let us consider $J'$ as a closed operator in the space $X_{1}'$ with the domain $X_{0}'$. The operator $J'$ is generating for the couple $[X_{1}',X_{0}']$ because $J'$ is positive and self-adjoint on $X_{1}'$. The positiveness of $J'$ results from the positiveness of the operator $J$ on the space $X_{0}$ and from (2.5) in the following way: $$(J'l,l)_{X_{1}'}=(S_{1}J^{-1}S_{0}^{-1}l,l)_{X_{1}'}=
(J^{-1}S_{0}^{-1}l,S_{1}^{-1}l)_{X_{1}}=
(JJ^{-1}S_{0}^{-1}l,JS_{1}^{-1}l)_{X_{0}}$$ $$=(JJS_{1}^{-1}l,JS_{1}^{-1}l)_{X_{0}}\geq
c\,\|JS_{1}^{-1}l\|_{X_{0}}^{2}=c\,\|S_{1}^{-1}l\|_{X_{1}}^{2}=
c\,\|l\|_{X_{1}'}^{2}.$$ Here the number $c>0$ does not depend on $l\in X_{0}'$. The operator $J'$ is self-adjoint because $0$ is its regular point (see (2.4)). Let us reduce the operator $J$ to the form of multiplication by a function: $J=I^{-1}(\alpha\cdot I)$ as it has been done in the proof of Theorem 2.3. The isometric isomorphism $$IJS_{1}^{-1}:\,X_{1}'\leftrightarrow L_{2}(U,d\mu) \leqno(2.6)$$ reduces the operator $J'$ to the form of multiplication by the same function $\alpha$: $$(IJS_{1}^{-1})J'l=IS_{0}^{-1}l=IJJ^{-1}S_{0}^{-1}l=\alpha\cdot
IJ^{-1}S_{0}^{-1}l=\alpha\cdot IJS_{1}^{-1}l\quad\mbox{for
each}\quad l\in X_{0}'.$$ The last equality follows from (2.5).
By Theorem 2.2, two continuous dense imbeddings $X_{0}'\hookrightarrow[X_{1}',X_{0}']_{\psi}$ and $[X_{0},X_{1}]_{\chi}\hookrightarrow X_{0}$ hold. The second imbedding implies the continuous dense imbedding $X_{0}'\hookrightarrow[X_{0},X_{1}]_{\chi}'$. Let us show that the norms in the spaces $[X_{1}',X_{0}']_{\psi}$ and $[X_{0},X_{1}]_{\chi}'$ are equal on the dense subset $X_{0}'$. For each $l\in X_{0}'$, $u\in[X_{0},X_{1}]_{\chi}$, we can write $$l(u)=(u,S_{0}^{-1}l)_{X_{0}}=(\chi(J)u,\chi^{-1}(J)S_{0}^{-1}l)_{X_{0}}=
(v,\chi^{-1}(J)S_{0}^{-1}l)_{X_{0}}$$ with $v:=\chi(J)u\in X_{0}$. It implies the following: $$\|l\|_{\,[X_{0},X_{1}]_{\chi}'}=
\sup\,\{\,|l(u)|\,/\,\|u\|_{\,[X_{0},X_{1}]_{\chi}}: u\in
[X_{0},X_{1}]_{\chi},\,u\neq0\,\}$$ $$=\sup\,\{\,|(v,\chi^{-1}(J)S_{0}^{-1}l)_{X_{0}}|\,/\,\|v\|_{X_{0}}:
\,v\in X_{0},\,v\neq0\,\}$$ $$=\|\chi^{-1}(J)S_{0}^{-1}l\|_{X_{0}}=
\|I\chi^{-1}(J)S_{0}^{-1}l\|_{L_{2}(U,d\mu)}=
\|(\chi^{-1}\circ\alpha)\cdot IS_{0}^{-1}l\|_{L_{2}(U,d\mu)}.$$ On the other hand, using isomorphisms (2.6), (2.4), we have $$\|l\|_{\,[X_{1}',X_{0}']_{\psi}}=\|\psi(J')l\|_{X_{1}'}=
\|\chi^{-1}(J')J'l\|_{X_{1}'}=
\|(IJS_{1}^{-1})\chi^{-1}(J')J'l\|_{L_{2}(U,d\mu)}$$ $$=\|(\chi^{-1}\circ\alpha)\cdot(IJS_{1}^{-1})J'l\|_{L_{2}(U,d\mu)}
=\|(\chi^{-1}\circ\alpha)\cdot IS_{0}^{-1}l\|_{L_{2}(U,d\mu)}.$$ Thus, norms in the spaces $[X_{1}',X_{0}']_{\psi}$ and $[X_{0},X_{1}]_{\chi}'$ are equal on the dense subset $X_{0}'$. So, these spaces coincide.
Now suppose $\psi$ to be an interpolation parameter. We will show that so is $\chi$. Let admissible couples $X=[X_{0},X_{1}]$, $Y=[Y_{0},Y_{1}]$ and a linear mapping $T$ be the same as those in Definition 2.2. Passing to the adjoint operator $T'$, we get the bounded operators $T':Y_{j}'\rightarrow X_{j}'$, $j=0,\,1$. Since $\psi$ is an interpolation parameter, a bounded operator $T':[Y_{1}',Y_{0}']_{\psi}\rightarrow[X_{1}',X_{0}']_{\psi}$ exists. As we have already proved, $[X_{1}',X_{0}']_{\psi}=[X_{0},X_{1}]_{\chi}'$ and $[Y_{1}',Y_{0}']_{\psi}=[Y_{0},Y_{1}]_{\chi}'$ with equalities of norms. Therefore a bounded operator $T':[Y_{0},Y_{1}]_{\chi}'\rightarrow[X_{0},X_{1}]_{\chi}'$ exists. Thus, passing to the second adjoint operator $T''$, we get the bounded operator $T'':[X_{0},X_{1}]_{\chi}''\rightarrow[Y_{0},Y_{1}]_{\chi}''$. It remains to identify the second dual spaces with original spaces which leads us to the bounded operator $T:[X_{0},X_{1}]_{\chi}\rightarrow[Y_{0},Y_{1}]_{\chi}$. This means that $\chi$ is an interpolation parameter. Theorem 2.4 is proved.
The ihterpolation of direct products of spaces.
-----------------------------------------------
Let a finite or countable set of admissible couples of Hilbert spaces $X^{(k)}:=[X_{0}^{(k)},X_{1}^{(k)}]$, $k\in\omega$, be given. Suppose that the set of norms of the imbedding operators $X_{1}^{(k)}\hookrightarrow X_{0}^{(k)}$, $k\in\omega$, is bounded. Then, for an arbitrary function $\psi\in\mathcal{B}$, the equality of spaces $$\left[\:\prod_{k\in\omega}X_{0}^{(k)},\,
\prod_{k\in\omega}X_{1}^{(k)}\right]_{\psi}=
\,\prod_{k\in\omega}\left[X_{0}^{(k)},X_{1}^{(k)}\right]_{\psi}$$ and the equality of norms in them hold.
We assume that $\omega=\mathbb{N}$ (the case of finite set $\omega$ is treated analogously and easier). The spaces $X_{0}:=\prod_{k=1}^{\infty}X_{0}^{(k)}$, $X_{1}:=\prod_{k=1}^{\infty}X_{1}^{(k)}$ are Hilbert and separable ones. The continuous imbedding $X_{1}\hookrightarrow X_{0}$ is evident due to the condition of the theorem. Let $u:=(u_{1},u_{2},\ldots)\in X_{0}$. For all indices $n,k$ an element $v_{n,k}\in X_{1}^{(k)}$ such that $\|u_{k}-v_{n,k}\|_{X_{0}^{(k)}}<1/n$ exists. Let us form a sequence of vectors $v^{(n)}:=(v_{n,1},\ldots,v_{n,n},0,0,\ldots)\in X_{1}$. We have $$\|u-v^{(n)}\|_{X_{0}}^{2}=\sum_{k=1}^{n}\|u_{k}-v_{n,k}\|_{X_{0}}^{2}+
\sum_{k=n+1}^{\infty}\|u_{k}\|_{X_{0}}^{2}\leq \frac{n}{n^{2}}+
\sum_{k=n+1}^{\infty}\|u_{k}\|_{X_{0}}^{2}\rightarrow0\;\;\mbox{as}
\;\;n\rightarrow\infty.$$ Thus, the couple $X:=[X_{0},X_{1}]$ is admissible.
Denote by $J_{k}$ a generating operator for the couple $X^{(k)}$. An operator $J:=(J_{1},J_{2},\ldots)$ is generating for the couple $X$ which may be proved directly. Moreover, it is natural to expect that $\psi(J)=(\psi(J_{1}),\psi(J_{2}),\ldots)$ and $\mathrm{Dom}\,\psi(J)=\prod_{k=1}^{\infty}X_{\psi}^{(k)}$. Now we prove these equalities. Let us reduce the operator $J_{k}$ to the form of multiplication by a function: $I_{k}J_{k}=\alpha_{k}\cdot
I_{k}$. Here $I_{k}:X_{0}^{(k)}\leftrightarrow
L_{2}(V_{k},d\mu_{k})$ is an isometric isomorphism, $V_{k}$ is a space with a finite measure $\mu_{k}$ and $\alpha_{k}:V_{k}\rightarrow(0,+\infty)$ is a measurable function. We may consider the sets $V_{k}$ to be mutually disjoint. Let us set $V:=\bigcup_{k=1}^{\infty}V_{k}$. We call $\Omega\subseteq V$ a measurable set if, for every index $k$, the set $\Omega\cap
V_{k}$ is $\mu_{k}$-measurable. On the $\sigma$-algebra of all measurable sets $\Omega\subseteq V$, we introduce the $\sigma$-finite measure $\mu(\Omega):=\sum_{k=1}^{\infty}\mu_{k}(\Omega\cap V_{k})$. Further, for a vector $u:=(u_{1},u_{2},\ldots)\in X_{0}$, we consider the measurable functions $Iu$ and $\alpha$, defined on the set $V$ by the formulae $(Iu)(\lambda):=(I_{k}u_{k})(\lambda)$ and $\alpha(\lambda):=\alpha_{k}(\lambda)$ with $\lambda\in
V_{k}$. Now we have the isometric isomorphism $I:X_{0}\leftrightarrow L_{2}(V,d\mu)$. It reduces the operator $J$ to the form of multiplication by the function $\alpha$ because $$(IJu)(\lambda)=(I_{k}J_{k}u_{k})(\lambda)=
\alpha_{k}(\lambda)(I_{k}u_{k})(\lambda)=\alpha(\lambda)(Iu)(\lambda)\quad
\mbox{for}\;\;u\in X_{1},\,\lambda\in V_{k}.$$ Hence, we can write down the following: $$X_{\psi}=\mathrm{Dom}\,\psi(J)=\{\,u\in
X_{0}:\,(\psi\circ\alpha)\cdot(Iu)\in L_{2}(V,d\mu)\,\}$$ $$=\{\,u\in
X_{0}:\,\sum_{k=1}^{\infty}\:\|(\psi\circ\alpha_{k})\cdot(I_{k}u_{k})\|
_{L_{2}(V_{k},d\mu_{k})}^{2}<\infty\,\}$$ $$=\{\,u:\,u_{k}\in\mathrm{Dom}\,\psi(J_{k}),\;
\sum_{k=1}^{\infty}\:\|\psi(J_{k})u_{k}\|
_{X_{0}^{(k)}}^{2}<\infty\,\}=\prod_{k=1}^{\infty}X_{\psi}^{(k)}.$$ Furthermore, for each $u\in \mathrm{Dom}\,\psi(J)$, we have $$(I\psi(J)u)(\lambda)=\psi(\alpha(\lambda))\,(Iu)(\lambda)=
\psi(\alpha_{k}(\lambda))\,(I_{k}u_{k})(\lambda)$$ $$=(I_{k}\psi(J_{k})u_{k})(\lambda)=
\bigl(I(\psi(J_{1})u_{1},\psi(J_{2})u_{2},\ldots)\bigr)(\lambda)
\quad\mbox{for}\;\;\lambda\in V_{k}.$$ Therefore $\psi(J)u=(\psi(J_{1})u_{1},\psi(J_{2})u_{2},\ldots)$ which implies
$$\|u\|_{X_{\psi}}^{2}=\|\psi(J)u\|_{X_{0}}^{2}=
\sum_{k=1}^{\infty}\,\|\psi(J_{k})u_{k}\|_{X_{0}^{(k)}}^{2}=
\sum_{k=1}^{\infty}\,\|u_{k}\|_{X_{\psi}^{(k)}}^{2}.$$ Theorem 2.5 is proved.
An operator norm in interpolation spaces.
-----------------------------------------
For given interpolation parameter $\psi\in\mathcal{B}$ and number $m>0$, there exists a number $c=c(\psi,m)>0$ such that $$\|T\|_{X_{\psi}\rightarrow Y_{\psi}}\leq c
\max\,\bigl\{\,\|T\|_{X_{j}\rightarrow Y_{j}}:\,j=0,\,1\,\bigr\}.$$ Here $X=[X_{0},X_{1}]$ and $Y=[Y_{0},Y_{1}]$ are admissible couples of Hilbert spaces for which the norms of the imbedding operators $X_{1}\hookrightarrow X_{0}$ and $Y_{1}\hookrightarrow
Y_{0}$ do not exceed the number $m$, and $T$ is any linear mapping defined on the space $X_{0}$ and establishing the bounded operators $T:X_{j}\rightarrow Y_{j}$ with $j=0,\,1$.
Let us suppose the contrary. Then we can write: $$\|T_{k}\|_{X^{(k)}_{\psi}\rightarrow Y^{(k)}_{\psi}}>k\,m_{k}
\quad\mbox{for each index}\;\;k. \leqno(2.7)$$ Here, $X^{(k)}:=[X_{0}^{(k)},X_{1}^{(k)}]$ and $Y^{(k)}:=[Y_{0}^{(k)},Y_{1}^{(k)}]$ are some admissible couples of Hilbert spaces for which the norms of the imbedding operators $X_{1}^{(k)}\hookrightarrow X_{0}^{(k)}$ and $Y_{1}^{(k)}\hookrightarrow Y_{0}^{(k)}$ do not exceed the number $m$. Furthermore, $T_{k}$ is a certain linear mapping defined on the space $X_{0}^{(k)}$ and establishing the bounded operators $T_{k}:X_{j}^{(k)}\rightarrow Y_{j}^{(k)}$ with $j=0,\,1$. We also use the notation $$m_{k}:=\max\left\{\,\|T_{k}\|_{X^{(k)}_{0}\rightarrow
Y^{(k)}_{0}},\;\|T_{k}\|_{X^{(k)}_{1}\rightarrow
Y^{(k)}_{1}}\,\right\}>0.$$ Now let us consider the bounded operators $$T:u=(u_{1},u_{2},\ldots)\mapsto
(m_{1}^{-1}\,T_{1}u_{1},m_{2}^{-1}\,T_{2}u_{2},\ldots),\quad
T:\,\prod_{k=1}^{\infty}X_{j}^{(k)}\rightarrow\prod_{k=1}^{\infty}Y_{j}^{(k)},
\;j=0,\,1. \leqno(2.8)$$ Their boundedness results from the following inequalities: $$\sum_{k=1}^{\infty}\;\left\|m_{k}^{-1}\,T_{k}u_{k}\right\|_{Y_{j}^{(k)}}^{2}\leq
\sum_{k=1}^{\infty}\;m_{k}^{-2}\,\|T_{k}\|_{X_{j}^{(k)}\rightarrow
Y_{j}^{(k)}}^{2}\,\|u_{k}\|_{X_{j}^{(k)}}^{2}\leq
\sum_{k=1}^{\infty}\;\|u_{k}\|_{X_{j}^{(k)}}^{2}.$$ Since $\psi$ is an interpolation parameter, the boundedness of operators (2.8) implies the existence of the bounded operator $$T:\,\left[\:\prod_{k=1}^{\infty}X_{0}^{(k)},\,
\prod_{k=1}^{\infty}X_{1}^{(k)}\right]_{\psi}\rightarrow\;
\left[\:\prod_{k=1}^{\infty}Y_{0}^{(k)},\,
\prod_{k=1}^{\infty}Y_{1}^{(k)}\right]_{\psi}$$ which by Theorem 2.5 means the boundedness of the operator $$T:\,\prod_{k=1}^{\infty}X_{\psi}^{(k)}\rightarrow\;
\prod_{k=1}^{\infty}Y_{\psi}^{(k)}.$$ Let $c_{0}$ be the norm of the last operator. For every index $k$ we consider a vector $u^{(k)}:=(u_{1},\ldots,u_{k},\ldots)$ such that $u_{k}\in X_{\psi}^{(k)}$ and $u_{j}=0$ for $j\neq k$. We have: $$\|T_{k}u_{k}\|_{Y_{\psi}^{(k)}}=
m_{k}\,\|Tu^{(k)}\|_{\prod_{j=1}^{\infty}Y_{\psi}^{(j)}}\leq
m_{k}\,c_{0}\,\|u^{(k)}\|_{\prod_{j=1}^{\infty}X_{\psi}^{(j)}}=
m_{k}\,c_{0}\,\|u_{k}\|_{X_{\psi}^{(k)}}$$ for each $u_{k}\in X_{\psi}^{(k)}$. Hence, $$\|T_{k}\|_{X^{(k)}_{\psi}\rightarrow Y^{(k)}_{\psi}}\leq
c_{0}\,m_{k} \quad\mbox{for every index}\;\;k,$$ contrary to inequality (2.7). Thus, our supposition is false and the theorem is true.
A criterion for a function to be an interpolation parameter.
------------------------------------------------------------
Using Peetre’s results [@Peet68], [@BL80 Sec. 5.4], we prove the following criterion.
Let a function $\psi:(0,+\infty)\rightarrow(0,+\infty)$ and a number $r\geq0$ be given. The function $\psi$ is called *quasiconcave* (or pseudoconcave) on the semiaxis $(r,+\infty)$ if a concave function $\psi_{1}:(r,+\infty)\rightarrow(0,+\infty)$ such that $\psi(t)\asymp \psi_{1}(t)$ for $t>r$ exists. The function $\psi$ is called quasiconcave in a neighborhood of $+\infty$ if it is quasiconcave on a certain semiaxis $(r,+\infty)$, where $r$ is a sufficiently large number.
A function $\psi\in\mathcal{B}$ is an interpolation parameter if and only if it is quasiconcave in a neighborhood of $+\infty$.
To prove this theorem we need two lemmas.
Let a function $\psi$ belong to the set $\mathcal{B}$ and be quasiconcave in a neighborhood of $+\infty$. Then there exists a concave function $\psi_{0}:(0,+\infty)\rightarrow(0,+\infty)$ such that for every number $\varepsilon>0$ it holds $\psi(t)\asymp
\psi_{0}(t)$ with $t\geq\varepsilon$.
It is evident.
Let a function $\psi\in\mathcal{B}$ and a number $r\geq0$ be given. The function $\psi$ is quasiconcave on the semiaxis $(r,+\infty)$ if and only if there exists a number $c>0$ such that $$\psi(t)/\psi(s)\leq c\,\max\{1,\,t/s\}\quad\mbox{for each}\quad
t,s>r.$$
In the case where $r=0$ this lemma was proved by J. Peetre [@Peet68], [@BL80 Lemma 5.4.3] (the condition $\psi\in\mathcal{B}$ being superfluous). In the case where $r>0$ the sufficiency can be proved analogously. The necessity is be reduced to the case $r=0$ with the help of Lemma 2.1. Indeed, let us put $\varepsilon=r$ in this lemma. Then we have a function $\psi_{0}$ such that $$\psi(t)/\psi(s)\asymp\psi_{0}(t)/\psi_{0}(s)\leq
c_{0}\max\{1,\,t/s\}\quad\mbox{for each}\quad t,s>r.$$ (In fact, $c_{0}=1$ for a concave function $\psi_{0}$ [@Peet68]). Lemma 2.2 is proved.
*Sufficiency.* Let us suppose that a function $\psi\in\mathcal{B}$ is quasiconcave in a neighborhood of $+\infty$. We need to prove that $\psi$ is an interpolation parameter.
Let admissible couples $X=[X_{0},X_{1}]$, $Y=[Y_{0},Y_{1}]$ and a linear mapping $T$ be the same as those in Definition 2.2. In addition, let operators $J_{X}:X_{1}\leftrightarrow X_{0}$ and $J_{Y}:Y_{1}\leftrightarrow Y_{0}$ be generating ones for the couples $X$ and $Y$ respectively. Using the spectral theorem we reduce these operators, self-adjoint in $X_{0}$ and in $Y_{0}$ respectively, to the form of multiplication by a function: $$J_{X}=I_{X}^{-1}\,(\alpha\cdot I_{X})\quad\mbox{and}\quad
J_{Y}=I_{Y}^{-1}\,(\beta\cdot I_{Y}). \leqno(2.9)$$ Here, $I_{X}:X_{0}\leftrightarrow L_{2}(U,d\mu)$ and $I_{Y}:Y_{0}\leftrightarrow L_{2}(V,d\nu)$ are certain isometric isomorphisms, $(U,\mu)$ and $(V,\nu)$ are spaces with finite measures and $\alpha:U\rightarrow(0,+\infty)$ and $\beta:V\rightarrow(0,+\infty)$ are some measurable functions. Since the operators $T:X_{0}\rightarrow Y_{0}$ and $T:X_{1}\rightarrow Y_{1}$ are bounded, so are the operators $$I_{Y}\,T\,I_{X}^{-1}:\,L_{2}(U,d\mu)\rightarrow L_{2}(V,d\nu),
\leqno(2.10)$$ $$I_{Y}\,J_{Y}\,T\,J_{X}^{-1}\,I_{X}^{-1}:\,L_{2}(U,d\mu)\rightarrow
L_{2}(V,d\nu). \leqno(2.11)$$ By virtue of (2.9), we can write $$I_{Y}\,J_{Y}\,T\,J_{X}^{-1}\,I_{X}^{-1}=(\beta\cdot
I_{Y})\,T\,(\alpha^{-1}\cdot I_{X}^{-1}).$$ Hence (2.11) implies the boundedness of the operator $$I_{Y}\,T\,I_{X}^{-1}=\beta^{-1}\cdot(I_{Y}\,J_{Y}\,T\,J_{X}^{-1}\,I_{X}^{-1})
\cdot\alpha:\,L_{2}(U,\alpha^{2}d\mu)\rightarrow
L_{2}(V,\beta^{2}d\nu). \leqno(2.12)$$
Let a concave function $\psi_{0}:(0,+\infty)\rightarrow(0,+\infty)$ be the same as that in Lemma 2.1. Note that $\psi_{0}\in\mathcal{B}$ and (see Subsection 2.1) $$X_{\psi}=X_{\psi_{0}},\quad Y_{\psi}=Y_{\psi_{0}}\quad\mbox{with
equivalence of norms}. \leqno(2.13)$$ J. Peetre [@Peet68], [@BL80 Theorem 5.4.4] proved that a positive function is quasiconcave on $(0,+\infty)$ if and only if it is an interpolation function in the sense of the definition stated in [@BL80 Definition 5.4.2]. Hence, for the function $\psi_{0}$, the boundedness of operators (2.10), (2.12) implies the existence of a bounded operator $$I_{Y}\,T\,I_{X}^{-1}:\,L_{2}(U,(\psi_{0}\circ\alpha^{2})\,d\mu)\rightarrow
L_{2}(V,(\psi_{0}\circ\beta^{2})\,d\nu). \leqno(2.14)$$ Let us pass from (2.14) to the operator $T:X_{\psi_{0}}\rightarrow
Y_{\psi_{0}}$ with the help of the isometric isomorphisms $\psi_{0}(J_{X}):\,X_{\psi_{0}}\leftrightarrow X_{0}$ and $\psi_{0}(J_{Y}):\,Y_{\psi_{0}}\leftrightarrow Y_{0}$. We reduce these isomorphisms (which are self-adjoint operators on $X_{0}$ and $Y_{0}$ respectively) to the form of the multiplication by a function: $$I_{X}\,\psi_{0}(J_{X})=(\psi_{0}\circ\alpha)\cdot I_{X}:\,
X_{\psi_{0}}\leftrightarrow L_{2}(U,d\mu),$$ $$I_{Y}\,\psi_{0}(J_{Y})=(\psi_{0}\circ\beta)\cdot I_{Y}:\,
Y_{\psi_{0}}\leftrightarrow L_{2}(V,d\nu).$$ We get the isometric isomorphisms $$I_{X}=(\psi_{0}^{-1}\circ\alpha)\cdot(I_{X}\,\psi_{0}(J_{X})):\,
X_{\psi_{0}}\leftrightarrow L_{2}(U,(\psi^{2}\circ\alpha)\,d\mu),$$ $$I_{Y}=(\psi_{0}^{-1}\circ\beta)\cdot(I_{Y}\,\psi_{0}(J_{Y})):\,
Y_{\psi_{0}}\leftrightarrow L_{2}(V,(\psi^{2}\circ\beta)\,d\nu).$$ From this and (2.14) the existence of the bounded operator $$T=I_{Y}^{-1}(I_{Y}\,T\,I_{X}^{-1})I_{X}:\,X_{\psi_{0}}\rightarrow
Y_{\psi_{0}}$$ follows.
Thus, due to equations (2.13) we have $$(T:X_{j}\rightarrow
Y_{j},\,j=0,1)\,\Rightarrow\,(T:X_{\psi_{0}}\rightarrow
Y_{\psi_{0}})\,\Rightarrow\,(T:X_{\psi}\rightarrow Y_{\psi}),$$ where the linear operators are bounded. So, by Definition 2.2 the function $\psi$ is an interpolation parameter. Sufficiency is proved.
*Necessity.* Now we suppose that a function $\psi\in\mathcal{B}$ is an interpolation parameter. We need to prove that $\psi$ is quasiconcave in a neighborhood of $+\infty$. The proof is similar to [@Peet68], [@BL80 Sec. 5.4].
Let us consider the space $L_{2}(U,d\mu)$ with $U=\{0,\,1\}$, $\mu(\{0\})=\mu(\{1\})=1$ and define on it the linear mapping $T$ by the formula $(Tu)(0):=0$, $(Tu)(1):=u(0)$, where $u\in
L_{2}(U,d\mu)$. Choose arbitrary numbers $s,t>1$ and put $\omega(0):=s^{2}$, $\omega(1):=t^{2}$. We have the admissible couple $X:=[L_{2}(U,d\mu),L_{2}(U,\omega\,d\mu)]$ and bounded operators $$T:\,L_{2}(U,d\mu)\rightarrow L_{2}(U,d\mu)\quad\mbox{and}\quad
T:\,L_{2}(U,\omega\,d\mu)\rightarrow L_{2}(U,\omega\,d\mu)$$ with norms $1$ and $t/s$ respectively. From this, since $\psi$ is an interpolation parameter, it follows that a bounded operator $T:X_{\psi}\rightarrow X_{\psi}$ exists. By Theorem2.6 with $Y=X$ and $m=1$ we conclude that the norm of this operator satisfies the inequality $$\|T\|_{X_{\psi}\rightarrow X_{\psi}}\leq\,c\,\max\{1,t/s\}.
\leqno(2.15)$$ Here, the number $c>0$ does not depend on $t,s>1$.
It is not difficult to calculate the norm in the space $X_{\psi}$. Indeed, the operator $J$ of multiplication by the function $\omega^{1/2}$ is generating for the couple $X$. Hence, since $\psi(J)$ is the operator of multiplication by the function $\psi\circ\omega^{1/2}$, we can write $$\bigl\|u\bigr\|_{X_{\psi}}^{2}=
\bigl\|(\psi\circ\omega^{1/2})\,u\bigr\|_{L_{2}(U,d\mu)}^{2}=
\psi^{2}(s)\,|u(0)|^{2}+\psi^{2}(t)\,|u(1)|^{2},\quad
\|Tu\|_{X_{\psi}}^{2}=\psi^{2}(t)\,|u(0)|^{2}.$$ It follows that $$\|T\|_{X_{\psi}\rightarrow X_{\psi}}=\psi(t)/\psi(s). \leqno(2.16)$$ Now relations (2.15), (2.16) imply the inequality $$\psi(t)\leq c\,\max\{1,t/s\}\,\psi(s)\quad\mbox{for each}\quad
t,s>1.$$ According to Theorem 2.2, the last statement is equivalent to the quasiconcavity of the function $\psi$ on the semiaxis $(1,+\infty)$. Necessity is proved.
A refined scale of spaces
=========================
Quasiregularly varying functions.
---------------------------------
We recall the following:
A positive function $\psi$ defined on a semiaxis $[b,+\infty)$ is called a function *regularly varying* at $+\infty$ with the index $\theta\in\mathbb{R}$ if $\psi$ is Borel measurable on $[b_{\,0},+\infty)$ for some number $b_{\,0}\geq b$ and $$\lim_{t\rightarrow\,+\infty}\;\psi(\lambda\,t)/\psi(t)=
\lambda^{\theta}\quad\mbox{for each}\quad \lambda>0.$$ A function regularly varying at $+\infty$ with the index $\theta=0$ is called a function *slowly varying* at $+\infty$.
The theory of regularly varying functions was founded by J. Karamata in the 1930s. These functions are closely related to the power functions and have numerous applications, mainly due to their special role in Tauberian-type theorems [@Se85; @Re87; @BiGoTe89; @Ma00]. A standard example of functions regularly varying at $+\infty$ with the index $\theta$ is $$\psi(t)=t^{\theta}\,(\ln t)^{r_{1}}\,(\ln\ln t)^{r_{2}} \ldots
(\ln\ldots\ln t)^{r_{k}}\quad\mbox{for}\quad t\gg1,$$ where $r_{1}, r_{2},\ldots,r_{k}\in\mathbb{R}$. In the case where $\theta=0$ these functions form the *logarithmic multiscale* which has a number of applications in the theory of function spaces.
A positive function $\psi$ defined on a semiaxis $[b,+\infty)$ is called a function *quasiregularly varying* at $+\infty$ with the index $\theta\in\mathbb{R}$ if there exist a number $b_{1}\geq b$ and a function $\psi_{1}:[b_{1},+\infty)\rightarrow
(0,+\infty)$ regularly varying at $+\infty$ with the index $\theta$ such that $\psi(t)\asymp\psi_{1}(t)$ with $t\geq b_{1}$. A function quasiregularly varying at $+\infty$ with the index $\theta=0$ is called a function *quasislowly varying* at $+\infty$.
We denote by $\mathrm{QSV}$ the set of all functions quasislowly varying at $+\infty$. It is evident that $\psi$ is a function quasiregularly varying at $+\infty$ with the index $\theta$ if and only if $\psi(t)=t^{\theta}\varphi(t)$, $t\gg1$, for some function $\varphi\in\mathrm{QSV}$. From the known [@Se85 Theorem 1.2] integral representation of a slowly varying function it immediately results the following description of the set $\mathrm{QSV}$.
Let $\varphi\in\mathrm{QSV}$. Then $$\varphi(t)=
\exp\left(\beta(t)+\int_{r}^{\:t}\frac{\alpha(\tau)}{\tau}\;d\tau\right),
\quad t\geq r, \leqno(3.1)$$ for some number $r>0$, a continuous function $\alpha:
[r,+\infty)\rightarrow\mathbb{R}$ approaching zero at $+\infty$ and a bounded function $\beta: [r,+\infty)\rightarrow\mathbb{R}$. The converse statement is also true: every function of form (3.1) *belongs to the set $\mathrm{QSV}$.*
Following interpolation property of quasiregular varying functions will play a decisive role in further.
Let a function $\psi\in\mathcal{B}$ be quasiregularly varying at $+\infty$ with the index $\theta\in(0,1)$. Then $\psi$ is an interpolation parameter.
We can write $\psi(t)=t^{\theta}\varphi(t)$ for $t>0$ with $\varphi\in\mathrm{QSV}$. According to Theorem 3.1, the function $\varphi$ can be represented in form (3.1). Let us set $\varepsilon:=\min\{\theta,1-\theta\}>0$ and choose a number $r_{\varepsilon}\geq r$ such that $|\alpha(t)|<\varepsilon$ for $t> r_{\varepsilon}$. For each $t,s>r_{\varepsilon}$, we have by virtue of (3.1) the following: $$\frac{\varphi(t)}{\varphi(s)}=\exp\left(\beta(t)-\beta(s)+
\int_{s}^{\:t}\frac{\alpha(\tau)}{\tau}\;d\tau\right) \leq
c\exp\left|\int_{s}^{\:t}\frac{\varepsilon}{\tau}\;d\tau\right|=
c\max\left\{(t/s)^{\varepsilon},(s/t)^{\varepsilon}\right\}.$$ Here the number $c>0$ does not depend on $t$ and $s$ because the function $\beta$ is bounded. From this and from the inequality $0\leq\theta\pm\varepsilon\leq1$ it follows that $$\psi(t)/\psi(s)=(t^{\theta}\varphi(t))/(s^{\theta}\varphi(s))\leq
c\max\left\{(t/s)^{\theta+\varepsilon},(t/s)^{\theta-\varepsilon}\right\}\leq
c\max\{1,t/s\}.$$ Hence, by Theorem 2.2 the function $\psi\in\mathcal{B}$ is quasiconcave in a neighborhood of $+\infty$. According to Theorem 2.7, this is equivalent to the statement that $\psi$ is an interpolation parameter. Theorem 3.2 is proved.
We need the following properties of the set $\mathrm{QSV}$.
Let $\varphi,\chi\in\mathrm{QSV}$. The following assertions are true.
- *There is a positive function $\varphi_{1}\in
C^{\infty}((0;+\infty))$ regularly varying at $+\infty$ such that $\varphi(t)\asymp\varphi_{1}(t)$ with $t\gg1$.*
- *For each $\theta>0$, the limits $t^{-\theta}\varphi(t)\rightarrow0$ and $t^{\theta}\varphi(t)\rightarrow+\infty$ as $t\rightarrow+\infty$ hold.*
- *The functions $\varphi+\chi$, $\varphi\,\chi$, $\varphi/\chi$ and $\varphi^{\sigma}$, where $\sigma\in\mathbb{R}$, belong to the set $\mathrm{QSV}$.*
- *Let $\theta\geq0$ and in the case where $\theta=0$ suppose that $\varphi(t)\rightarrow\infty$ as $t\rightarrow+\infty$. Then the composite function $\chi(t^{\theta}\varphi(t))$ of the argument $t$ belongs to the set $\mathrm{QSV}$.*
For regularly varying functions $\varphi,\chi$ these assertions are known [@Se85 Sec. 1.5] (even with the strong equivalence being in assertion (i)). This implies immediately assertions (i), (ii), (iii) for the functions $\varphi,\chi\in\mathrm{QSV}$.
It remains to prove assertion (iv). Let $\lambda>0$. Since $\varphi\in\mathrm{QSV}$, the functions $\varphi(\lambda
t)/\varphi(t)$ and $\varphi(t)/\varphi(\lambda t)$ are bounded in a neighborhood of $+\infty$. Therefore applying the theorem [@Se85 Sec. 1.2] on uniform convergence to a positive slowly varying function $\chi_{1}$ such that $\chi_{1}(\tau)\asymp\chi(\tau)$ with $\tau\gg1$, we can write $$\chi_{1}\left((\lambda t)^{\theta}\varphi(\lambda t)\right)\big/
\chi_{1}\left(t^{\theta}\varphi(t)\right)=
\chi_{1}\left(\frac{\lambda^{\theta}\varphi(\lambda
t)}{\varphi(t)}\:t^{\theta}\varphi(t)\right)\Big/
\chi_{1}\left(t^{\theta}\varphi(t)\right)\rightarrow1\quad\mbox{as}
\quad t\rightarrow+\infty.$$ Here we use the limit $t^{\theta}\varphi(t)\rightarrow\infty$ as $t\rightarrow+\infty$. Hence, the function $\chi_{1}(t^{\theta}\varphi(t))$ is slowly varying at $+\infty$. In addition, $\chi(t^{\theta}\varphi(t))\asymp\chi_{1}(t^{\theta}\varphi(t))$ with $t\gg1$. Thus, the function $\chi(t^{\theta}\varphi(t))$ belongs to the set $\mathrm{QSV}$. Assertion (iv) is proved.
A refined scale over the Euclidean space.
-----------------------------------------
Let $n\in\mathbb{N}$. As usual, $\mathbb{R}^{n}$ denotes the $n$-dimensional Euclidean space and $\mathcal{S}'(\mathbb{R}^{n})$ denotes the linear topological Schwartz space of tempered distributions on $\mathbb{R}^{n}$. We use also the following notations: $\langle\xi\rangle=(1+\xi_{1}^{2}+\ldots+\xi_{n}^{2})^{1/2}$ denotes the smoothed modulus of a vector $\xi=(\xi_{1},\ldots,\xi_{n})\in \mathbb{R}^{n}$ and $\widehat{u}$ denotes the Fourier transform of the distribution $u\in\mathcal{S}'(\mathbb{R}^{n})$. We write an integral taken on the space $\mathbb{R}^{n}$ without limits.
Let $\mathcal{M}$ denote the set of all functions $\varphi:[1;+\infty)\rightarrow(0;+\infty)$ such that
a
: $\varphi$ is Borel measurable on the set $[1;+\infty)$;
b
: functions $\varphi$ and $1/\varphi$ are bounded on every closed interval $[1;b]$, where $1<b<+\infty$;
c
: $\varphi\in\mathrm{QSV}$.
Let $s\in\mathbb{R}$, $\varphi\in\mathcal{M}$.
We denote by $H^{s,\varphi}(\mathbb{R}^{n})$ the space of all distributions $u\in\mathcal{S}'(\mathbb{R}^{n})$ such that the Fourier transform $\widehat{u}$ is a function locally Lebesgue integrable on $\mathbb{R}^{n}$ which satisfies the inequality $$\int\langle\xi\rangle^{2s}\varphi^{2}(\langle\xi\rangle)\,|\widehat{u}(\xi)|^{2}\,
d\xi<\infty.$$ The inner product in the space $\mathrm{H}^{s,\varphi}(\mathbb{R}^{n})$ is defined by the formula $$(u,v)_{\mathrm{H}^{s,\varphi}(\mathbb{R}^{n})}:=
\int\langle\xi\rangle^{2s}\varphi^{2}(\langle\xi\rangle)
\,\widehat{u}(\xi)\,\overline{\widehat{v}(\xi)}\,d\xi$$ and generates the norm in the usual way.
The space $H^{s,\varphi}(\mathbb{R}^{n})$ is a special isotropic Hilbert case of the spaces introduced by L. Hörmander [@Her65 Sec. 2.2], [@Her86 Sec. 10.1] and L. R. Volevich, B. P. Paneah [@VoPa65 Sec. 2], [@Pa00 Sec. 1.4.2]. Note that this space is actually defined with the help of the function $\varphi_{s}(t)=t^{s}\varphi(t)$ regularly varying at $+\infty$ with the index $s$. However it is more convenient for us to represent the parameter $\varphi_{s}$ as the couple of two parameters $s$ and $\varphi$.
In the particular case where $\varphi\equiv1$ the space $H^{s,\varphi}(\mathbb{R}^{n})$ coincides with the Sobolev space $H^{s}(\mathbb{R}^{n})$. In general, the following inclusions are true: $$\bigcup_{\varepsilon>0}H^{s+\varepsilon}(\mathbb{R}^{n})=:H^{s+}(\mathbb{R}^{n})
\subset H^{s,\varphi}(\mathbb{R}^{n})\subset
H^{s-}(\mathbb{R}^{n}):=\bigcap_{\varepsilon>0}H^{s-\varepsilon}(\mathbb{R}^{n}).
\leqno(3.2)$$ They result from assertion (ii) of Theorem 3.3 and from the definition of the set $\mathcal{M}$, according to which, for each $\varepsilon>0$, there is a number $c_{\varepsilon}\geq1$ such that $c_{\varepsilon}^{-1}t^{-\varepsilon}\leq\varphi(t)\leq
c_{\varepsilon}t^{\varepsilon}$ for $t\geq1$. Inclusions (3.2) mean that, in the collection of the spaces $$\{H^{s,\varphi}(\mathbb{R}^{n}):s\in\mathbb{R},\varphi\in\mathcal{M}\,\},
\leqno(3.3)$$ the function parameter $\varphi$ *refines* the basic (power) $s$-smoothness. Therefore it is natural to give the following definition.
The collection of function spaces (3.3) is called a *refined scale* over $\mathbb{R}^{n}$ (with respect to the Sobolev scale).
Besides the properties inherent to the Hörmander spaces [@Her65 Sec. 2.2 ], [@Her86 Sec. 10.1] and the Volevich-Paneah spaces [@VoPa65 Ch. I, II], [@Pa00 Sec. 1.4], the refined scale over $\mathbb{R}^{n}$ possesses the following fundamental interpolation property:
Let a function $\varphi\in\mathcal{M}$ and positive numbers $\varepsilon,\delta$ be given. Let $\psi(t):=t^{\,\varepsilon/(\varepsilon+\delta)}\,
\varphi(t^{1/(\varepsilon+\delta)})$ for $t\geq1$ and $\psi(t):=\varphi(1)$ for $0<t<1$. Then the following assertions are true:
- The function $\psi$ belongs to the set $\mathcal{B}$ and is an interpolation parameter.
- For an arbitrary $s\in\mathbb{R}$, the equality of spaces $$\left[H^{s-\varepsilon}(\mathbb{R}^{n}),H^{s+\delta}(\mathbb{R}^{n})\right]_{\psi}
=H^{s,\varphi}(\mathbb{R}^{n})$$ and equality of norms in them hold.
*Assertion* (i). By virtue of assertions (ii), (iv) of Theorem 3.3, the function $\psi$ belongs to the set $\mathcal{B}$ and is a function regular varying at $+\infty$ with the index $\theta=\varepsilon/(\varepsilon+\delta)\in(0,\,1)$. Therefore $\psi$ is an interpolation parameter because of Theorem 3.2. Assertion (i) is proved.
*Assertion* (ii). Let $s\in\mathbb{R}$. It follows from the properties of the Sobolev spaces that the couple $[H^{s-\varepsilon}(\mathbb{R}^{n}),H^{s+\delta}(\mathbb{R}^{n})]$ is admissible and the pseudodifferential operator with symbol $\langle\xi\rangle^{\varepsilon+\delta}$ is a generating operator $J$ for this couple. Applying the Fourier transform $\mathcal{F}:H^{s-\varepsilon}(\mathbb{R}^{n})\leftrightarrow
L_{2}(\mathbb{R}^{n},\langle\xi\rangle^{2(s-\varepsilon)}d\xi)$, we reduce the operator $J$ to the form of multiplication by the function $\langle\xi\rangle^{\varepsilon+\delta}$ of $\xi\in\mathbb{R}^{n}$. Hence, the operator $\psi(J)$ is reduced to the form of multiplication by the function $\psi(\langle\xi\rangle^{\varepsilon+\delta})=
\langle\xi\rangle^{\varepsilon}\varphi(\langle\xi\rangle)$. This permits us to write the following in view of (3.2): $$\left[H^{s-\varepsilon}(\mathbb{R}^{n}),H^{s+\delta}(\mathbb{R}^{n})\right]_{\psi}=
\left\{u\in H^{s-\varepsilon}(\mathbb{R}^{n}):
\langle\xi\rangle^{\varepsilon}\varphi(\langle\xi\rangle)\
\widehat{u}(\xi)\in
L_{2}(\mathbb{R}^{n},\langle\xi\rangle^{2(s-\varepsilon)}d\xi)\right\}$$ $$=\left\{u\in H^{s-\varepsilon}(\mathbb{R}^{n}):
\int\langle\xi\rangle^{2s}\varphi^{2}(\langle\xi\rangle)
\left|\widehat{u}(\xi)\right|^{2}d\xi<\infty\right\}=
H^{s-\varepsilon}(\mathbb{R}^{n})\cap
H^{s,\varphi}(\mathbb{R}^{n})=H^{s,\varphi}(\mathbb{R}^{n}).$$ In addition the norm in the space $\left[H^{s-\varepsilon}(\mathbb{R}^{n}),
H^{s+\delta}(\mathbb{R}^{n})\right]_{\psi}$ is equal to $$\|\psi(J)u\|_{H^{s-\varepsilon}(\mathbb{R}^{n})}=
\left(\int|\langle\xi\rangle^{\varepsilon}\varphi(\langle\xi\rangle)\
\widehat{u}(\xi)|^{2}\,\langle\xi\rangle^{2(s-\varepsilon)}\
d\xi\right)^{1/2}=\|u\|_{H^{s,\varphi}(\mathbb{R}^{n})}.$$ Assertion (ii) is proved.
A refined scale over a closed manifold
--------------------------------------
Further let $\Gamma$ be a closed (that is compact and without a boundary) infinitely smooth manifold of dimension $n$. We suppose that a certain $C^{\infty}$-density $dx$ is defined on $\Gamma$. We denote by $\mathcal{D}'(\Gamma)$ the linear topological space of all distributions on $\Gamma$. Thus $\mathcal{D}'(\Gamma)$ is the space antidual to the space $C^{\infty}(\Gamma)$ with respect to the natural extension of the scalar product in $L_{2}(\Gamma,dx)=:L_{2}(\Gamma)$ by continuity. This extension is denoted by $(f,w)_{\Gamma}$ for $f\in\mathcal{D}'(\Gamma)$, $w\in
C^{\infty}(\Gamma)$.
The refined scale over the manifold $\Gamma$ is constructed from scale (3.3) in the following way. We choose a finite atlas from the $C^{\infty}$-structure on $\Gamma$ consisting of the local charts $\alpha_{j}:\mathbb{R}^{n}\leftrightarrow U_{j}$, $j=1,\ldots,r$. Here the open sets $U_{j}$ form the finite covering of the manifold $\Gamma$. Let functions $\chi_{j}\in
C^{\infty}(\Gamma)$, $j=1,\ldots,r$, form a partition of unity on $\Gamma$ satisfying the condition $\mathrm{supp}\,\chi_{j}\subset
U_{j}$. As before, $s\in\mathbb{R}, \varphi\in\mathcal{M}$.
We denote by $H^{s,\varphi}(\Gamma)$ the space of all distributions $f\in\mathcal{D}'(\Gamma)$ such that $(\chi_{j}f)\circ\alpha_{j}\in H^{s,\varphi}(\mathbb{R}^{n})$ for each $j=1,\ldots,r$. Here $(\chi_{j}f)\circ\alpha_{j}$ is the representation of the distribution $\chi_{j}f$ in the local chart $\alpha_{j}$. The inner product in the space $H^{s,\varphi}(\Gamma)$ is defined by the formula $$(f,g)_{H^{s,\varphi}(\Gamma)}:=\sum_{j=1}^{r}\,((\chi_{j}f)\circ\alpha_{j},
(\chi_{j}\,g)\circ\alpha_{j})_{H^{s,\varphi}(\mathbb{R}^{n})}$$ and induces the norm in the usual way.
The collection of function spaces $\{H^{s,\varphi}(\Gamma):s\in\mathbb{R},\varphi\in\mathcal{M}\}$ is called a *refined scale* over the closed manifold $\Gamma$.
In the particular case where $\varphi\equiv1$ the space $H^{s,\varphi}(\Gamma)$ coincides with the Sobolev space $H^{s}(\Gamma)$. Sobolev spaces are known [@Her65 Sec. 2.6], [@Shubin78 Sec. 7.5] to be complete and independent (up to equivalence of norms) of the choice of the atlas and the partition of unity. We will show that every space $H^{s,\varphi}(\Gamma)$ can be obtained by the interpolation of the proper couple of Sobolev’s spaces. It implies that the space $H^{s,\varphi}(\Gamma)$ is Hilbert and independent of this choice.
Let a function $\varphi\in\mathcal{M}$ and positive numbers $\varepsilon,\delta$ be given. Then, for each $s\in\mathbb{R}$, the equality of spaces $$\left[H^{s-\varepsilon}(\Gamma),\
H^{s+\delta}(\Gamma)\right]_{\psi}=H^{s,\varphi}(\Gamma)\quad\mbox{with
equivalence of norms} \leqno(3.4)$$ hold. Here $\psi$ is the interpolation parameter from Theorem $3.4$.
The couple of the Sobolev spaces on the left-hand side of equality (3.4) is admissible [@Shubin78 Sec. 7.5, 7.6]. We deduce this equality from Theorem 3.4 with the help of the well known method of “rectification” and “sewing” of the manifold $\Gamma$. According to Definition 3.5, the linear mapping of “rectification” $$T:\,f\mapsto(\,(\chi_{1}f)\circ\alpha_{1},\ldots,(\chi_{r}f)\circ\alpha_{r}\,),
\quad f\in\mathcal{D}'(\Gamma),$$ defines the isometric operators $$T:\,H^{\sigma}(\Gamma)\rightarrow(H^{\sigma}(\mathbb{R}^{n}))^{r},\quad
\sigma\in\mathbb{R}, \leqno(3.5)$$ $$T:\,H^{s,\varphi}(\Gamma)\rightarrow(H^{s,\varphi}(\mathbb{R}^{n}))^{r}.
\leqno(3.6)$$ Since $\psi$ is the interpolation parameter and operators (3.5) are bounded for $\sigma\in\{s-\varepsilon,s+\delta\}$, the bounded operator $$T:\,\left[H^{s-\varepsilon}(\Gamma),H^{s+\delta}(\Gamma)\right]_{\psi}
\rightarrow\left[\,(H^{s-\varepsilon}(\mathbb{R}^{n}))^{r},\,
(H^{s+\delta}(\mathbb{R}^{n}))^{r}\,\right]_{\psi} \leqno(3.7)$$ exists. By virtue of Theorems 2.5, 3.4, the following equalities of spaces and norms in them are true: $$\left[\,(H^{s-\varepsilon}(\mathbb{R}^{n}))^{r},\,
(H^{s+\delta}(\mathbb{R}^{n}))^{r}\,\right]_{\psi}=
\left(\,\left[H^{s-\varepsilon}(\mathbb{R}^{n}),
H^{s+\delta}(\mathbb{R}^{n})\right]_{\psi}\,\right)^{r}
=(H^{s,\varphi}(\mathbb{R}^{n}))^{r}. \leqno(3.8)$$ Thus, since operator (3.7) is bounded, so is the operator $$T:\,\left[H^{s-\varepsilon}(\Gamma),H^{s+\delta}(\Gamma)\right]_{\psi}
\rightarrow(H^{s,\varphi}(\mathbb{R}^{n}))^{r}. \leqno(3.9)$$
Now we construct for $T$ the left inverse operator $K$ of “sewing” of the manifold $\Gamma$. For each $j=1,\ldots,r$ we take a function $\eta_{j}\in C_{0}^{\infty}(\mathbb{R}^{n})$ such that $\eta_{j}=1$ on the set $\alpha_{j}^{-1}(\mathrm{supp}\,\chi_{j})$. Let us consider the linear mapping $$K:\,(h_{1},\ldots,h_{r})\mapsto\sum_{j=1}^{r}\,
\Theta_{j}\left((\eta_{j}h_{j})\circ\alpha_{j}^{-1}\right),\quad
h_{1},\ldots,h_{r}\in\mathcal{S}'(\mathbb{R}^{n}).$$ Here $(\eta_{j}h_{j})\circ\alpha_{j}^{-1}$ is a distribution in the open set $U_{j}\subseteq\Gamma$ such that its representative in the local map $\alpha_{j}$ has the form $\eta_{j}h_{j}$. In addition, $\Theta_{j}$ denotes the operator of extension by zero from $U_{j}$ to $\Gamma$. This operator is well defined on distributions with support belonging to $U_{j}$. By the choice of the functions $\chi_{j}$, $\eta_{j}$, we have $$KTf=\sum_{j=1}^{r}\,\Theta_{j}\left((\eta_{j}\,
((\chi_{j}f)\circ\alpha_{j}))\circ\alpha_{j}^{-1}\right)=
\sum_{j=1}^{r}\,\Theta_{j}\left(\,
(\chi_{j}f)\circ\alpha_{j}\circ\alpha_{j}^{-1}\right)=
\sum_{j=1}^{r}\,\chi_{j}f=f,$$ that is $$KTf=f\quad\mbox{for each}\quad f\in\mathcal{D}'(\Gamma).
\leqno(3.10)$$
Let us show that the linear mapping $K$ defines the bounded operator $$K:\,(H^{s,\varphi}(\mathbb{R}^{n}))^{r}\rightarrow
H^{s,\varphi}(\Gamma).\leqno(3.11)$$ For an arbitrary vector $h=(h_{1},\ldots,h_{r})$ from the space $(H^{s,\varphi}(\mathbb{R}^{n}))^{r}$, we write $$\bigl\|Kh\bigr\|^{2}_{H^{s,\varphi}(\Gamma)}
=\sum_{l=1}^{r}\;\bigl\|(\chi_{l}\,Kh)
\circ\alpha_{\,l}\bigr\|_{H^{s,\varphi}(\mathbb{R}^{n})}^{2}
=\sum_{l=1}^{r}\,\Bigl\|\Bigl(\chi_{\,l}\,\sum_{j=1}^{r}\
\Theta_{j}\bigl((\eta_{j}h_{j})\circ\alpha_{j}^{-1}\bigr)\Bigr)
\circ\alpha_{\,l}\,\Bigr\|_{H^{\,s,\varphi}(\mathbb{R}^{n})}^{2}$$ $$=\sum_{l=1}^{r}\;\Bigl\|\,\sum_{j=1}^{r}(\eta_{j,l}\,h_{j})
\circ\beta_{\,j,l}\,\Bigr\|_{H^{s,\varphi}(\mathbb{R}^{n})}^{2}
\leq\sum_{l=1}^{r}\;\Bigl(\sum_{j=1}^{r}\,\bigl\|(\eta_{j,l}\,h_{j})\circ
\beta_{j,l}\,\bigr\|_{H^{\,s,\varphi}(\mathbb{R}^{n})}\Bigr)^{2}.\leqno(3.12)$$ Here $\eta_{j,l}:=(\chi_{\,l}\circ\alpha_{j})\,\eta_{j}\in
C_{0}^{\infty}(\mathbb{R}^{n})$ and $\beta_{j,l}:\,\mathbb{R}^{n}\,\leftrightarrow\,\mathbb{R}^{n}$ is a $C^{\infty}$-diffeomorphism such that $\beta_{j,l}\,=\alpha_{j}^{-1}\circ\alpha_{l}$ in a neighborhood of $\mathrm{supp}\,\eta_{j,l}$ and $\beta_{j,l}(x)=x$ for all $x\in\mathbb{R}^{n}$ sufficiently large in modulus. The operator of multiplication by a function of the class $C_{0}^{\infty}(\mathbb{R}^{n})$ and the operator of change of variables $u\mapsto u\circ\beta_{j,l}$ are known [@Her87 Theorems B.1.7, B.1.8] to be bounded on every space $H^{\sigma}(\mathbb{R}^{n})$ with $\sigma\in\mathbb{R}$. Therefore the linear operator $v\mapsto (\eta_{j,l}\,v)\circ\beta_{j,l}$ is bounded on the space $H^{\sigma}(\mathbb{R}^{n})$. Then boundedness of this operator in the space $H^{s,\varphi}(\mathbb{R}^{n})$ follows by Theorem 3.3. Hence relations (3.12) imply the estimate $$\bigl\|Kh\bigr\|_{H^{s,\varphi}(\Gamma)}^{2}\leq
c\,\sum_{j=1}^{r}\
\bigl\|h_{j}\bigr\|_{H^{s,\varphi}(\mathbb{R}^{n})}^{2},$$ where the constant $c>0$ is independent of $h=(h_{1},\ldots,h_{r})$. Thus, operator (3.11) is bounded for each $s\in\mathbb{R}$, $\varphi\in\mathcal{M}$.
In particular, the operators $K:\,(H^{\sigma}(\mathbb{R}^{n}))^{r}\rightarrow
H^{\sigma}(\Gamma)$ with $\sigma\in\mathbb{R}$ are bounded. Let us set $\sigma\in\{s-\varepsilon,\,s+\delta\}$ and use the interpolation with the parameter $\psi$. Due to equality (3.8), we obtain the bounded operator $$K:\,(H^{s,\varphi}(\mathbb{R}^{n}))^{r}\rightarrow\
\left[H^{s-\varepsilon}(\Gamma),\,H^{s+\delta}(\Gamma)\right]_{\psi}.\leqno(3.13)$$ Now formulae (3.6), (3.13) and (3.10) imply that the identity mapping $KT$ establishes the continuous imbedding of the space $H^{s,\varphi}(\Gamma)$ into the interpolation space $[H^{s-\varepsilon}(\Gamma),\,H^{s+\delta}(\Gamma)]_{\psi}$. Moreover, formulae (3.10) and (3.13) imply that the same mapping $KT$ establishes also the inverse continuous imbedding. Theorem 3.5 is proved.
The following properties of the refined scale over the manifold $\Gamma$ can be deduced from Theorem 3.5 and the interpolation properties established in Section 2.
Let $s\in\mathbb{R}$ and $\varphi,\varphi_{1}\in\mathcal{M}$. The following assertions are true:
- The space $H^{s,\varphi}(\Gamma)$ is Hilbert separable and does not depend (up to equivalence of norms) on the choice of an atlas for $\Gamma$ and partition of unity used in Definition $3.5$.
- The set $C^{\infty}(\Gamma)$ is dense in the space $H^{s,\varphi}(\Gamma)$.
- For each $\varepsilon>0$, the compact and dense imbedding $H^{s+\varepsilon,\varphi_{1}}(\Gamma)\hookrightarrow
H^{s,\varphi}(\Gamma)$ holds.
- Suppose that the function $\varphi/\varphi_{1}$ is bounded in a neighborhood of $+\infty$. Then continuous dense imbedding $H^{s+\varepsilon,\varphi_{1}}(\Gamma)\hookrightarrow
H^{s,\varphi}(\Gamma)$ is valid. It is compact if $\varphi(t)/\varphi_{1}(t)\rightarrow0$ as $t\rightarrow+\infty$.
- The spaces $H^{s,\varphi}(\Gamma)$ and $H^{-s,1/\varphi}(\Gamma)$ are mutually dual (up to equivalence of norms) with respect to the extension of the inner product in $L_{2}(\Gamma)$ by continuity.
*Assertion* (i). The space $H^{s,\varphi}(\Gamma)$ is Hilbert and separable because, according to Theorem 3.5, this space is obtained by the interpolation of a certain couple of the Sobolev spaces. Let us consider two couples $\mathcal{A}_{1}$ and $\mathcal{A}_{2}$ each of which consists of an atlas and a partition of unity on $\Gamma$. We denote by $H^{s,\varphi}(\Gamma,\mathcal{A}_{j})$ and $H^{\sigma}(\Gamma,\mathcal{A}_{j})$ respectively the spaces from the refined scale and the Sobolev spaces which correspond to the couple $\mathcal{A}_{j}$, where $j=1,\,2$. For the Sobolev spaces, the identity mapping establishes the topological isomorphism $I:H^{\sigma}(\Gamma,\mathcal{A}_{1})\leftrightarrow
H^{\sigma}(\Gamma,\mathcal{A}_{2})$ for each $\sigma\in\mathbb{R}$. Let us set $\sigma=s\mp1$ and use the interpolation with the parameter $\psi$ from Theorem 3.4. By Theorem 3.5 we arrive at the topological isomorphism $I:H^{s,\varphi}(\Gamma,\mathcal{A}_{1})\leftrightarrow
H^{s,\varphi}(\Gamma,\mathcal{A}_{2})$. It means that the space $H^{s,\varphi}(\Gamma)$ is independent of the choice of an atlas and a unity partition mentioned above. Assertion (i) is proved.
*Assertion* (ii). By virtue of Theorems 2.1 and 3.5, we have the continuous dense imbedding $H^{s+\delta}(\Gamma)\hookrightarrow H^{s,\varphi}(\Gamma)$. Besides, the set $C^{\infty}(\Gamma)$ is dense in the Sobolev space $H^{s+\delta}(\Gamma)$ [@Shubin78 Proposition 7.4]. These two facts imply assertion (ii).
*Assertion* (iii). Assume that $\varepsilon>0$. By Theorem 3.5, there exist interpolation parameters $\chi,\eta\in\mathcal{B}$ such that the following equalities of spaces with equivalence of norms in them is true: $$\left[H^{s+\varepsilon/2}(\Gamma),H^{s+2\varepsilon}(\Gamma)\right]_{\chi}=
H^{s+\varepsilon,\varphi_{1}}(\Gamma)\quad\mbox{and}\quad
\left[H^{s-\varepsilon}(\Gamma),H^{s+\varepsilon/3}(\Gamma)\right]_{\eta}=
H^{s,\varphi}(\Gamma).$$ It implies by Theorem 2.1 the next chain of continuous imbeddings $$H^{s+\varepsilon,\varphi_{1}}(\Gamma)\hookrightarrow
H^{s+\varepsilon/2}(\Gamma)\hookrightarrow
H^{s+\varepsilon/3}(\Gamma)\hookrightarrow H^{s,\varphi}(\Gamma).$$ Here the central imbedding of Sobolev spaces is compact [@Shubin78 Theorem 7.4]. Therefore the imbedding $H^{s+\varepsilon,\varphi_{1}}(\Gamma)\hookrightarrow
H^{s,\varphi}(\Gamma)$ is compact as well. This imbedding is dense because of assertion (ii). Assertion (iii) is proved.
*Assertion* (iv). Let us assume that the function $\varphi/\varphi_{1}$ is bounded in a neighborhood of $+\infty$. By Theorem 3.5, we have the following equalities of spaces with equivalence of norms in them: $$\left[H^{s-1}(\Gamma),H^{s+1}(\Gamma)\right]_{\psi}
=H^{s,\varphi}(\Gamma)\quad\mbox{and}\quad
\left[H^{s-1}(\Gamma),H^{s+1}(\Gamma)\right]_{\psi_{1}}
=H^{s,\varphi_{1}}(\Gamma).$$ Here the interpolation parameters $\psi,\psi_{1}\in\mathcal{B}$ satisfy the condition $$\psi(t)/\psi_{1}(t)=\varphi(t^{1/2})/\varphi_{1}(t^{1/2})\quad
\mbox{for}\quad t\geq1.$$ Hence, the function $\psi/\psi_{1}$ is bounded in a neighborhood of $+\infty$ that, by Theorem 2.2, implies the continuous dense imbedding $H^{s,\varphi_{1}}(\Gamma)\hookrightarrow
H^{s,\varphi}(\Gamma)$. Now suppose that $\varphi(t)/\varphi_{1}(t)\rightarrow0$ as $t\rightarrow+\infty$. It implies the limit $\psi(t)/\psi_{1}(t)\rightarrow0$ as $t\rightarrow+\infty$. In addition, we recall that the imbedding of Sobolev spaces $H^{s+1}(\Gamma)\hookrightarrow H^{s-1}(\Gamma)$ is compact. It follows by Theorem 2.2 that the imbedding $H^{s,\varphi_{1}}(\Gamma)\hookrightarrow H^{s,\varphi}(\Gamma)$ is compact as well. Assertion (iv) is proved.
*Assertion* (v) is known (see e.g. [@Shubin78 Theorem 7.7]) in the case $\varphi\equiv1$. From this the case of an arbitrary $\varphi\in\mathcal{M}$ can be obtained as follows. First we note that $1/\varphi\in\mathcal{M}$ and therefore the space $H^{-s,1/\varphi}(\Gamma)$ is well defined. The Sobolev spaces $H^{s\pm1}(\Gamma)$ and $H^{-s\mp1}(\Gamma)$ are mutually dual with respect to the extension of the inner product in $L_{2}(\Gamma)$ by continuity. This means that the linear mapping $Q:w\mapsto(\,\cdot\,,\overline{w})_{\Gamma}$, $w\in C^{\infty}(\Gamma)$, is extended by continuity to the topological isomorphisms $Q:H^{s\mp1}(\Gamma)\leftrightarrow(H^{-s\pm1}(\Gamma))'$. Let us apply to them the interpolation with the parameter $\psi$ from Theorem 3.5 in the case where $\varepsilon=\delta=1$. We obtain one more topological isomorphism $$Q:\left[H^{s-1}(\Gamma),H^{s+1}(\Gamma)\right]_{\psi}\leftrightarrow
\left[(H^{-s+1}(\Gamma))',(H^{-s-1}(\Gamma))'\right]_{\psi}.
\leqno(3.14)$$ Here the left-hand interpolation space equals to $H^{s,\varphi}(\Gamma)$ and, by Theorem 2.4, the right-hand one can be written as $$\left[(H^{-s+1}(\Gamma))',(H^{-s-1}(\Gamma))'\right]_{\psi}=
\left[H^{-s-1}(\Gamma),H^{-s+1}(\Gamma)\right]_{\chi}'=
(H^{-s,1/\varphi}(\Gamma))'.$$ Let us note that the last equality is valid because $\chi(t):=t/\psi(t)=t^{1/2}/\varphi(t^{1/2})$ for $t\geq1$. Thus, (3.14) implies the topological isomorphism $Q:H^{s,\varphi}(\Gamma)\leftrightarrow(H^{-s,1/\varphi}(\Gamma))'$, which means the mutual duality of the spaces $H^{s,\varphi}(\Gamma)$ and $H^{-s,1/\varphi}(\Gamma)$ in the sense mentioned above. Assertion (v) is proved.
The refined scale is closed with respect to the interpolation with a function parameter regular varying at $+\infty$.
Let $s_{0},s_{1}\in\mathbb{R}$, $s_{0}\leq s_{1}$ and $\varphi_{0},\varphi_{1}\in\mathcal{M}$. In the case where $s_{0}=s_{1}$ we suppose that the function $\varphi_{0}/\varphi_{1}$ is bounded in a neighborhood of $+\infty$. Let $\psi\in\mathcal{B}$ be a function regularly varying at $+\infty$ with the index $\theta$, where $0<\theta<1$. By Theorem $3.2$, $\psi$ is an interpolation parameter. We represent it as $\psi(t)=t^{\theta}\chi(t)$ with $\chi\in\mathrm{QSV}$. Let us set $s:=(1-\theta)s_{0}+\theta
s_{1}$ and $$\varphi(t):=\varphi_{0}^{1-\theta}(t)\,\varphi_{1}^{\theta}(t)\,
\chi\left(t^{s_{1}-s_{0}}\varphi_{1}(t)/\varphi_{0}(t)\right)\quad\mbox{for}\quad
t\geq1.$$ Then $\varphi\in\mathcal{M}$ and $$\left[\,H^{s_{0},\varphi_{0}}(\Gamma),
H^{s_{1},\varphi_{1}}(\Gamma)\,\right]_{\psi}=
H^{s,\varphi}(\Gamma)\quad\mbox{with equivalence of norms}.$$
This theorem is a direct consequence of Theorems 3.5 and 2.3.
Theorem 3.7 is true in the limiting case where $\theta=0$ or $\theta=1$ under additional supposition that the function $\psi$ is quasiconcave in a neighborhood of $+\infty$. Then, by Theorem2.7, $\psi$ is an interpolation parameter. For example, Theorem 3.7 is true for each of the functions $\psi(t):=\ln^{r}t$ and $\psi(t):=t/\ln^{r} t$, where $t\gg1$ and $r>0$.
An alternative definition of the refined scale.
-----------------------------------------------
Let $A$ be an elliptic pseudodifferential operator on $\Gamma$ of order $m>0$. We suppose that the operator $A:C^{\infty}(\Gamma)\rightarrow C^{\infty}(\Gamma)$ is positive on the space $L_{2}(\Gamma)$, that is there exists a number $r>0$ such that $$(Lu,u)_{\Gamma}\geq r\,(u,u)_{\Gamma}\quad\mbox{for each}\quad
u\in C^{\infty}(\Gamma). \leqno(3.15)$$ In the present subsection, $(\,\cdot\,,\,\cdot\,)_{\Gamma}$ is the inner product in $L_{2}(\Gamma)$.
We denote by $A_{0}$ the closure of the operator $A:C^{\infty}(\Gamma)\rightarrow C^{\infty}(\Gamma)$ on the space $L_{2}(\Gamma)$. This closure exists and has the domain $H^{m}(\Gamma)$ because the operator $A$ is elliptic on $\Gamma$ [@Shubin78 Corollary 8.3], [@Agr90 Theorem 2.3.5]. The pseudodifferential operator $A$ is formally self-adjoint due to condition (3.15). Hence [@Shubin78 Theorem 8.3], [@Agr90 Theorem 2.3.7], $A_{0}$ is an unbounded self-adjoint operator in the space $L_{2}(\Gamma)$ with $\mathrm{Spec}\,A_{0}\subseteq[r,+\infty)$. In particular, we have $0\notin\mathrm{Spec}\,A_{0}$, that implies the topological isomorphism $$A:\,H^{s+m,\varphi}(\Gamma)\,\leftrightarrow\,H^{s,\varphi}(\Gamma)
\quad\mbox{for each}\quad s\in\mathbb{R},\;\varphi\in\mathcal{M}.
\leqno(3.16)$$ In the Sobolev case where $\varphi\equiv1$ this result is well known (see e.g. [@Shubin78 Theorem 8.1, Proposition 8.5], [@Her87 Theorem 19.2.1], [@Agr90 Sec. 2.3]). The general case of an arbitrary $\varphi\in\mathcal{M}$ follows immediately from the case $\varphi\equiv1$ by virtue of Theorem 3.5.
Let $s\in\mathbb{R}$ and $\varphi\in\mathcal{M}$. We set $\varphi_{s}(t):=t^{s/m}\varphi(t^{1/m})$ for $t\geq1$ and, moreover, $\varphi_{s}(t):=\varphi(1)$ for $0<t<1$. Since the function $\varphi$ is positive and Borel measurable on the semiaxis $(0,+\infty)$, a self-adjoint operator $\varphi_{s}(A_{0})$ is well-defined in the space $L_{2}(\Gamma)$ as the function $\varphi$ of $A_{0}$.
The following assertions are true.
- The domain of the operator $\varphi_{s}(A_{0})$ contains the set $C^{\infty}(\Gamma)$.
- The mapping $$f\,\mapsto\,\|\varphi_{s}(A_{0})f\|_{L_{2}(\Gamma)},\quad f\in
C^{\infty}(\Gamma), \leqno(3.17)$$ is a norm in the space $C^{\infty}(\Gamma)$.
*Assertion* (i). Let us choose an integer $k$ so that $k>s/m$. Since $\varphi\in\mathcal{M}$, the function $\varphi_{s}$ is bounded on every compact subset of the semiaxis $(0,+\infty)$ and, moreover, $t^{-k}\varphi_{s}(t)\rightarrow0$ as $t\rightarrow+\infty$ because of assertions (ii), (iv) of Theorem 3.3. Hence, there is a number $c>0$ such that $\varphi_{s}(t)\leq
c\,t^{k}$ for $t\geq r$. Let us consider the unbounded operator $A_{0}^{k}$ on the space $L_{2}(\Gamma)$. Since $A:C^{\infty}(\Gamma)\rightarrow C^{\infty}(\Gamma)$, we can write $C^{\infty}(\Gamma)\subset\mathrm{Dom}\,A_{0}^{k}\subset
\mathrm{Dom}\,\varphi_{s}(A_{0})$. Assertion (i) is proved.
*Assertion* (ii). According to assertion (i), mapping (3.17) is well-defined. For this mapping, all norm properties are evident except for the positive definiteness property. Let us prove it. Applying the spectral theorem, we can write for an arbitrary function $f\in C^{\infty}(\Gamma)$: $$\bigl\|\varphi_{s}(A_{0})f\bigr\|_{L_{2}(\Gamma)}^{2}=
\int_{r}^{+\infty}\varphi_{s}^{2}(t)\,d(E_{t}f,f)_{\Gamma}\quad\mbox{and}
\quad\bigl\|f\bigr\|_{L_{2}(\Gamma)}^{2}=
\int_{r}^{+\infty}d(E_{t}f,f)_{\Gamma}. \leqno(3.18)$$ Here $E_{t}$, $t\geq r$, is the resolution of identity in the space $L_{2}(\Gamma)$ which corresponds to the self-adjoint operator $A_{0}$. If $\|\varphi_{s}(A_{0})f\|_{L_{2}(\Gamma)}^{2}=0$, then from the first equality in (3.18) and from the inequality $\varphi_{s}>0$ it follows that the measure $(E(\cdot)f,f)_{\Gamma}$ of the set $[r,+\infty)$ is equal to 0. Now the second equality in (3.18) implies that $f=0$ on $\Gamma$. Assertion (ii) is proved.
The space $H^{s,\varphi}_{A}(\Gamma)$ is a completion of the space $C^{\infty}(\Gamma)$ with respect to norm (3.17).
The space $H^{s,\varphi}_{A}(\Gamma)$ is Hilbert one because norm (3.17) is generated by the inner product $(\varphi_{s}(A_{0})f,\varphi_{s}(A_{0})g)_{\Gamma}$ of functions $f,g\in C^{\infty}(\Gamma)$.
For arbitrary $s\in\mathbb{R}$, $\varphi\in\mathcal{M}$, the norms in the spaces $H^{s,\varphi}_{A}(\Gamma)$ and $H^{s,\varphi}(\Gamma)$ are equivalent on the dense linear manifold $C^{\infty}(\Gamma)$. Thus, $H^{s,\varphi}_{A}(\Gamma)=H^{s,\varphi}(\Gamma)$ up to equivalence of norms.
At first let us suppose that $s>0$. We choose $k\in\mathbb{N}$ so that $k\,m>s$. Since the operator $A_{0}^{k}$ is closed and positive on $L_{2}(\Gamma)$, its domain $\mathrm{Dom}\,A_{0}^{k}$ is Hilbert space with respect to the inner product $(A_{0}^{k}f,A_{0}^{k}g)_{\Gamma}$ of functions $f,g$. Note that the couple of spaces $[L_{2}(\Gamma),\mathrm{Dom}\,A_{0}^{k}]$ is admissible, and the operator $A_{0}^{k}$ is a generating one for it. Moreover, since $A_{0}^{k}$ is a closure of the elliptic pseudodifferential operator $A^{k}$ on $L_{2}(\Gamma)$, the spaces $\mathrm{Dom}\,A_{0}^{k}$ and $H^{km}(\Gamma)$ are equal up to equivalent norms. Let a function $\psi$ be the interpolation parameter from Theorems 3.3, 3.4 with $\varepsilon=s$ and $\delta=k\,m-s$. Then $\psi(t^{k})=\varphi_{s}(t)$ for $t>0$, so by Theorem 3.4 we can write $$\bigl\|f\bigr\|_{H^{s,\varphi}(\Gamma)}\asymp
\bigl\|f\bigr\|_{[H^{0}(\Gamma),\,H^{km}(\Gamma)]_{\psi}}\asymp
\bigl\|f\bigr\|_{[L_{2}(\Gamma),\,\mathrm{Dom}\,A_{0}^{k\,}]_{\psi}}=
\bigl\|\psi(A_{0}^{k})f\bigr\|_{L_{2}(\Gamma)}=
\bigl\|\varphi_{s}(A_{0})f\bigr\|_{L_{2}(\Gamma)},$$ for each $f\in C^{\infty}(\Gamma)$.
Now let the number $s\in\mathbb{R}$ be arbitrary. Choose $k\in\mathbb{N}$ so that $s+k\,m>0$. As has been proved, $$\bigl\|g\bigr\|_{H^{s+km,\varphi}(\Gamma)}\asymp
\bigl\|\varphi_{s+km}(A_{0})\,g\bigr\|_{L_{2}(\Gamma)},\quad g\in
C^{\infty}(\Gamma). \leqno(3.19)$$ The following topological isomorphism holds due to (3.16) : $$A^{k}:\,H^{\sigma+km,\varphi}(\Gamma)\leftrightarrow
H^{\sigma,\varphi}(\Gamma)\quad\mbox{for
each}\quad\sigma\in\mathbb{R}. \leqno(3.20)$$ Denote by $A^{-k}$ the inverse operator to $A^{k}$. For every function $f\in C^{\infty}(\Gamma)$, we have $A^{-k}f\in
C^{\infty}(\Gamma)$ and $A_{0}^{k}A^{-k}f=f$. Hence, by virtue of (3.20), (3.19), we can write $$\bigl\|f\bigr\|_{H^{s,\varphi}(\Gamma)}\asymp
\bigl\|A^{-k}f\bigr\|_{H^{s+km,\varphi}(\Gamma)}\asymp
\bigl\|\varphi_{s+km}(A_{0})A^{-k}f\bigr\|_{L_{2}(\Gamma)}$$ $$=\bigl\|\varphi_{s}(A_{0})A_{0}^{k}A^{-k}f\bigr\|_{L_{2}(\Gamma)}=
\bigl\|\varphi_{s}(A_{0})f\bigr\|_{L_{2}(\Gamma)},\quad f\in
C^{\infty}(\Gamma).$$ Theorem 3.8 is proved.
Let $s\geq0$ and $\varphi\in\mathcal{M}$. In the case where $s=0$ we suppose that the function $1/\varphi$ is bounded in a neighborhood of $+\infty$. Then the space $H^{s,\varphi}(\Gamma)$ coincides with the domain of the operator $\varphi_{s}(A_{0})$ and the norm in the space $H^{s,\varphi}(\Gamma)$ is equivalent to the graph norm of the operator $\varphi_{s}(A_{0})$.
The domain $\mathrm{Dom}\,\varphi_{s}(A_{0})$ of the closed operator $\varphi_{s}(A_{0})$ is Hilbert space with respect to the scalar product of the graph of this operator. Let us prove that the norms in the spaces $\mathrm{Dom}\,\varphi_{s}(A_{0})$ and $H^{s,\varphi}_{A}(\Gamma)$ are equivalent on the dense linear manifold $C^{\infty}(\Gamma)$. By Theorem 3.8, it will imply the present theorem. According to the condition of the present theorem and by virtue of assertion (ii) of Theorem 3.3, there is a number $c>0$ such that $\varphi_{s}(t)\geq c$ for $t>0$. Therefore $$\bigl\|\varphi_{s}(A_{0})f\bigr\|_{L_{2}(\Gamma)}\geq c\,
\bigl\|f\bigr\|_{L_{2}(\Gamma)}\quad\mbox{for each}\quad f\in
C^{\infty}(\Gamma).$$ It yields the equivalence of norms mentioned above. It remains to prove the density of the set $C^{\infty}(\Gamma)$ in the space $\mathrm{Dom}\,\varphi_{s}(A_{0})$.
Let $f\in\mathrm{Dom}\,\varphi_{s}(A_{0})$. Since $\varphi_{s}(A_{0})f\in L_{2}(\Gamma)$, there is a sequence of functions $h_{j}\in C^{\infty}(\Gamma)$ such that $h_{j}\rightarrow\varphi_{s}(A_{0})f$ in $L_{2}(\Gamma)$ as $j\rightarrow\infty$. Note that the operator $\varphi_{s}^{-1}(A_{0})$ is bounded on the space $L_{2}(\Gamma)$ because $1/\varphi_{s}(t)\leq 1/c$ for $t>0$. Hence, $$f_{j}:=\varphi_{s}^{-1}(A_{0})h_{j}\rightarrow
f\quad\mbox{and}\quad
\varphi_{s}(A_{0})f_{j}=h_{j}\rightarrow\varphi_{s}(A_{0})f\quad\mbox{in}\quad
L_{2}(\Gamma)\quad\mbox{as}\quad j\rightarrow\infty.$$ In other words, $f_{j}\rightarrow f$ with respect to the graph norm of the operator $\varphi_{s}(A_{0})$. Moreover, since $h_{j}\in C^{\infty}(\Gamma)$, then $f_{j}=A_{0}^{-k}\varphi_{s}^{-1}(A_{0})A_{0}^{k}\,h_{j}\in
H^{km}(\Gamma)$ for every $k\in\mathbb{N}$. Consequently, $f_{j}\in C^{\infty}(\Gamma)$ and the density of the set $C^{\infty}(\Gamma)$ in the space $\mathrm{Dom}\,\varphi_{s}(A_{0})$ is established. Theorem 3.9 is proved.
A significant example of the operator $A$ investigated above is the operator $1-\triangle_{\Gamma}$, where $\triangle_{\Gamma}$ is the Beltrami-Laplace operator on the Riemannian manifold $\Gamma$ (then $m=2$).
[999999]{} V. A. Mikhailets, A. A. Murach, *Improved scale of spaces and elliptic boundary-value problems. I*, Ukrainian Math. J. **58** (2006), no. 2, 244–262. V. A. Mikhailets, A. A. Murach, *Improved scale of spaces and elliptic boundary-value problems. II*, Ukrainian. Math. J. **58** (2006), no. 3, 398–417. Sh. A. Alimov, V. A. Il’in, E. M. Nikishin, *Convergence problems of multiple trigonometric series and spectral decompositions. I*, Russian Math. Surveys **31** (1976), no. 6, 29–86. V. A. Mikhailets, *Asymtotics of the spectrum of elliptic operators and boundary conditions*, Soviet Math. Dokl. **26** (1982), no. 5, 464–468. V. A. Mikhailets, *A precise estimate of the remainder in the spectral asymptotics of general elliptic boundary problems*, Funct. Anal. Appl. **23** (1989), no. 2, 137–139. G. A. Kalyabin, P. I. Lizorkin, *Spaces of functions of generalized smoothness*, Math. Nachr., **133** (1987), 7–32. D. D. Haroske, S. D. Moura, *Continuity envelopes of spaces of generalised smoothness, entropy and approximation numbers*, J. Approximation Theory, **128** (2004), 151–174. W. Farkas, H.-G. Leopold, *Characterisations of function of generalized smoothness*, Ann. Mat. Pura Appl. **185** (2006), no. 1, 1–62. C. Merucci, *Application of interpolation with a function parameter to Lorentz, Sobolev and Besov spaces*, Proc. Lund Conf. 1983, Lecture Notes in Math. **1070**, 183–201, Springer-Verlag, Berlin etc, 1984. F. Cobos, D. L. Fernandez, *Hardy-Sobolev spaces and Besov spaces with a function parameter*, Proc. Lund Conf. 1986, Lecture Notes in Math. **1302**, 158–170, Springer-Verlag, Berlin etc, 1988. V. A. Mikhailets, A. A. Murach, *The interpolation of spaces with functional parameter and the spaces of differential functions*, Dopov. Nats. Acad. Nauk. Ukr. Mat. Pryr. Tehn. Nauky (2006), no. 6, 13–18. (Russian) V. A. Mikhailets, A. A. Murach, *An elliptic operator in the refined scale of spaces on a closed manifold*, Dopov. Nats. Acad. Nauk. Ukr. Mat. Pryr. Tehn. Nauky (2006), no. 10, 27–33. (Russian) V. A. Mikhailets, A. A. Murach, *Regular elliptic boundary-value problem for a homogeneous equation in a two-sided improved scale of spaces*, Ukrainian Math. J. **58** (2006), no. 11, 1748–1767. V. A. Mikhailets, A. A. Murach, *Elliptic operator with homogeneous regular boundary conditions in two-sided refined scale of spaces*, Ukr. Math. Bull **3** (2006), no. 4, 529–560. V. A. Mikhailets, A. A. Murach, *Refined scale of spaces and elliptic boundary-value problems. III*, Ukra¿n. Mat. Zh. **59** (2007), no. 5, 679–701. (Russian) A. A. Murach, *The systems of differential equations elliptic in the sense of Petrovskii in a refined scale of spaces on a close manifold*, Dopov. Nats. Acad. Nauk. Ukr. Mat. Pryr. Tehn. Nauky (2007), no. 5, 29–35. (Russian) A. A. Murach, *The boundary-value problem for a system of differential equations elliptic in the sensce of Petrovskii in a refined scale of spaces*, Dopov. Nats. Acad. Nauk. Ukr. Mat. Pryr. Tehn. Nauky (2007), no. 6, 24–31. (Russian) A. A. Murach, *Elliptic pseudodifferential operators in refined scale of spaces over closed manifold*, Ukra¿n. Mat. Zh. **59** (2007), no. 6, 798–814. (Russian) V. A. Mikhailets, A. A. Murach, *Regular elliptic boundary-value problem in a two-sided improved scale of spaces*, Ukra¿n. Mat. Zh. **60** (2008), no. 4. (to appear, Russian) V. A. Mikhailets, A. A. Murach, *Elliptic systems of pseudodifferential equations in a refined scale on a closed manifold*, arXiv:0711.2164v1 \[math.AP\] 14 Nov 2007. J.-L. Lions, E. Magenes, *Problèmes aux Limites non Homogènes et Applications, Vol.1*, Dunod, Paris, 1968. *Functional analysis*, Nauka, Moscow, 1972. (Russian) H. Triebel, *Interpolation. Function Spaces. Differential Operators*, North-Holland, Amsterdam, 1978. J. Berg, J. Löfström, *Interpolation Spaces*, Springer-Verlag, Berlin etc, 1976. J. Peetre, *On interpolation functons. II*, Acta Sci. Math. **29** (1968), no. 1–2, 91–92. E. Seneta, *Regularly Varying Functions*, Lect. Notes in Math. 508, Springer-Verlag, Berlin etc, 1976. S. I. Reshnick, *Extreme Values, Regular Variation and Point Processes*, Springer-Verlag, Berlin etc, 1987. N. H. Bingham, C. M. Goldie, J. L. Teugels, *Regular Variation*, Cambridge Univ. Press, Cambridge, 1989. V. Maric, *Regular Variation and Differential Equations*, Springer-Verlag, Berlin etc, 2000. L. Hörmander, *Linear Partial Differential Operators*, Springer-Verlag, Berlin etc, 1963. L. Hörmander, *The Analysis of Linear Partial Differential Operators, Vol. 2*, Springer-Verlag, Berlin etc, 1983. L. R. Volevich, B. P. Paneah, *Certain spaces of generalized functions and imbedding theorems*, Uspehi Mat. Nauk. **20** (1965), no. 1, 3–74. (Russian) B. Paneah, *The Oblique Derivative Problem. The Poincaré Problem*, Wiley – VCH, Berlin etc, 2000. M. A. Shubin, *Pseudodifferential Operators and Spectral Theory*, Springer-Verlag, Berlin etc, 2001. L. Hörmander, *The Analysis of Linear Partial Differential Operators, Vol. 3*, Springer-Verlag, Berlin etc, 1985. M. S. Agranovich, *Partial differential equations. VI. Elliptic operators on closed manifolds*, Encycl. Math. Sci. **63**, Springer-Verlag, Berlin etc, 1994, 1–130.
| {
"pile_set_name": "ArXiv"
} |
---
author:
-
-
-
-
bibliography:
- 'main.bib'
title: 'Improving Web Content Blocking With Event-Loop-Turn Granularity JavaScript Signatures'
---
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The intermediate valent systems TmSe and [SmB$_6$]{} have been investigated up to 16 and 18 GPa by ac microcalorimetry with a pressure ($p$) tuning realized in situ at low temperature. For TmSe, the transition from an antiferromagnetic insulator for $p<3$ GPa to an antiferromagnetic metal at higher pressure has been confirmed. A drastic change in the $p$ variation of the Néel temperature ($T_N$) is observed at 3 GPa. In the metallic phase ($p>3$ GPa) , [$T_{N}$]{} is found to increase linearly with $p$. A similar linear $p$ increase of [$T_{N}$]{} is observed for the quasitrivalent compound TmS which is at ambiant pressure equivalent to TmSe at $p \sim 7$ GPa. In the case of [SmB$_6$]{} long range magnetism has been detected above $p \sim 8$ GPa, i.e. at a pressure slightly higher than the pressure of the insulator to metal transition. However a homogeneous magnetic phase occurs only above 10 GPa. The magnetic and electronic properties are related to the renormalization of the 4f wavefunction either to the divalent or the trivalent configurations. As observed in SmS, long range magnetism in [SmB$_6$]{} occurs already far below the pressure where a trivalent [Sm$^{3+}$]{} state will be reached. It seems possible, to describe roughly the physical properties of the intermediate valence equilibrium by assuming formulas for the Kondo lattice temperature depending on the valence configuration. Comparison is also made with the appearance of long range magnetism in cerium and ytterbium heavy fermion compounds.'
address: 'Département de Recherche Fondamentale sur la Matière Condensée, CEA Grenoble, 17 rue des Martyrs, 38054 Grenoble Cedex 9, France'
author:
- 'J. Derr, G. Knebel, G. Lapertot, B. Salce, M-A. Méasson and J. Flouquet'
title: 'Valence and magnetic ordering in intermediate valence compounds : TmSe versus SmB$_6$'
---
Introduction
============
####
Recently, a major interest was the study of the high pressure phase diagrams of heavy fermion compounds (HFC)[@revuedeJacques]. However, in these systems, the departure from the trivalent configuration is weak ; the occupation number [n$_f$]{} of the $4f^{1}$ is nearly one. Unusual magnetic properties found notably on ytterbium HFC such as YbRh$_2$Si$_2$ push to revisit other situations with magnetic and valence fluctuations occuring between two 4f configurations. The cases of intermediate valence compounds (IVC) as SmS, [SmB$_6$]{} and TmSe, are particularely interesting[@Wachter].
####
To caracterize the intermediate valent state, a key parameter is the occupation number [n$_f$]{} of the trivalent configuration linked to the valence $v$ by $v=2+$ [n$_f$]{} when the valence fluctuation occurs between the divalent and the trivalent state (case of Sm, Tm and Yb) or $v=4-$ [n$_f$]{} when it happens between the trivalent and tetravalent state (case of Ce). The important difference between Sm, Tm or Yb compounds is that [n$_f$]{} can vary from 0 to 1 while in Ce intermetallic compounds : [n$_f$]{} $>0.8$ and at least long range magnetic ordering (M) occurs only for [n$_f$]{} $>0.9$[@Malterre]. TmSe[@Launois; @Haen; @Wertheim; @Brewer] as well as SmS[@Wachter] and SmB$_6$[@Beaurepaire] in their low pressure intermediate valent gold phase have a valence near 2.6-2.7. Their trivalent limit will be reached smoothly only at very high pressure above 10 GPa for TmSe and 20 GPa for SmS and SmB$_6$[@Roeler; @Dallera; @Ogita] . As will be discussed, the striking point is that for these three systems, the change from insulating to metallic conduction at low temperature occurs when [n$_f$]{}$\sim 0.8$.
####
The valence mixing between the divalent (2+) and trivalent (3+) configurations of the rare earth (RE) ions is associated to the release of an itinerant 5d electron according to the relation RE$^{2+} \Longleftrightarrow$ RE$^{3+} + e^{-} 5d$. Experimentally, the effect of pressure is to broaden the bands and move this equilibrium to the right (increasing [n$_f$]{} ). Of course, band structure calculations are necessary to describe the real situation, but the chemical equilibrium is worthwhile to consider. In the divalent black (B) phase, the ground state is a classical insulator. Through a first order transition at $V=V_{B-G}$, a valence transition occurs to an intermediate valence (IV) gold (G) phase which is still insulating. However, under pressure the insulating gap will close for a fixed volume $V_\Delta$. At $V=V_\Delta$, metallic conduction appears for [n$_f$]{}$\sim 0.8$ at a volume quite larger than the volume $V_{3+}$ calculated for a pure trivalent configuration. Figure \[differentcompounds\] represents the location of the different compounds at ambiant pressure.
![Valence state, as a function of the density, i.e. the inverse of the molar volume $V$. For $V<V_{B-G}$, the system jumps from 2+ black (B) phase to IV gold (G) phase. up to $V_{B-G}>V>V_{\Delta}$ the system is still insulating (The dashed lines represent the insulating caracter). Magnetism looks also governed by the 2+ configuration. For $V<V_{\Delta}$, the system is metallic, and magnetism looks governed by the 3+ configuration. The trivalent limit will appear for $V_{3+}<V_{\Delta}$. Long range magnetism in SmS and [SmB$_6$]{} appear for $\frac{1}{V}\sim \frac{1}{V_{\Delta}}+\epsilon < \frac{1}{V_{3+}}$. Of course it occurs always for TmSe whatever is the valence[]{data-label="differentcompounds"}](figure1.eps){width="7cm"}
####
In Sm compounds, the intermediate valent state occurs between a non magnetic $4f^{6}$ configuration of Sm$^{2+}$, with a zero angular momentum J and the Kramer’s configuration (J=$\frac{5}{2}$) of Sm$^{3+}$ ($4f^{5}$). It looks worthwhile to predict that, as in cerium HFC, magnetic ordering will occur when the occupancy n$_f$ of the trivalent configuration approaches one. In this case, following the Doniach model (see [@revuedeJacques]), the Kondo coupling should be small enough, so that the Kondo energy becomes smaller than the RKKY energy. However, recently it was shown by use of a microscopic hyperfine technique as nuclear forward scattering and a macroscopic probe as ac microcalorimetry that magnetic ordering occurs already for a rather large departure from [n$_f$]{}$=1$[@BarlaSmS; @Haga; @BarlaSmB6]. Up to [n$_f$]{}$\leq 0.8$, the 4f wavefunction seems to be renormalized to the 2+ configuration while above [n$_f$]{}$\sim 0.8$, it seems linked to the 3+ configuration. Furthermore, this is related to the conduction properties : insulating below [n$_f$]{}$\sim 0.8$ and metallic above.
####
In Tm chalcogenides, the ground state of the divalent configuration ([n$_f$]{} $=0$, case of TmTe) is insulating, and becomes metallic for the trivalent form. In the IVC (case of TmSe) with low [n$_f$]{} ([n$_f$]{} $\leq 0.8$), the many body effects of the correlation lead to the survival of an insulator. In the specific case of Tm, the novelty is that mixing occurs between two configurations with non zero angular momentum. The divalent one ([Tm$^{2+}$]{} $4f^{13}$) is a Kramer’s configuration with $J=\frac{7}{2}$ which leads to a doublet or a quartet crystal field ground state, while the trivalent one ([Tm$^{3+}$]{} $4f^{12}$) is a non Kramer’s ion with $J=6$ which may lead to a singlet crystal field ground state. The pressure induced collapse of the insulating state is associated with a change in the magnetic structure at $p_{\Delta}\sim 3~$GPa[@Ribault]. Below $p_{\Delta}$, i.e. for [n$_f$]{}$\leq 0.8$, the ground state is insulating, like in the low pressure intermediate valence phase of SmS and SmB$_6$, and antiferromagnetic of type I with properties basically given by a dressing towards a divalent renormalization (insulating conduction, doublet degeneracy of the local magnetic level). Above $p_{\Delta}$, the ground state is metallic (like TmS, or SmS and SmB$_6$ at high pressure), again antiferromagnetic, but of type II, with properties renormalized to the trivalent configuration. A surprising report was that near p$\sim 6$ GPa, TmSe may become insulating again[@Ohashi1; @Ohashi2].
####
In this paper, we present a detailed study of the high pressure phase diagrams of TmSe, TmS and [SmB$_6$]{} . Since for the two first cases, specific heat is already well known for $p=0$, and also interplay occurs between pressure and ligand effects, those compounds allow to verify the faisability and difficulties of high pressure microcalorimetry experiments. As TmSe is already magnetically ordered at ambiant pressure, up to 3 GPa, one may expect a signal in the ac calorimetry equivalent to ambiant pressure. Above 3 GPa, the signal may change as the signal may be normalized to the 3+ configuration like in TmS. Special attention is given on the pressure range around 6 GPa. The evolution of $T_N(p)$ of TmSe above 3 GPa will be compared to the nearly trivalent TmS. In [SmB$_6$]{} , we found evidence for a magnetically ordered ground state for $p>8$ GPa. However, a homogeneous ground state appears only above 10 GPa.
####
The paper is organized as follows. First we will discuss details of the ac calorimetry technique. Then, the experimental results on TmSe, TmS ans [SmB$_6$]{} will be presented and an experimental conclusion will be given. In the last part, the influence of the valence on the appearance of magnetic order will be discussed in detail and a comparison to the well known high pressure phase diagrams of Ce and Yb Kondo lattice will be given.
Experimental
============
####
The TmSe and TmS single crystals were prepared by F. Holtzberg in IBM research center, New York, and samples of the same batch have been intensively studied previously in CNRS Grenoble[@TmSeTmSauCNRS]. [SmB$_6$]{} single crystals were grown in CEA Grenoble, out of an aluminium flux. The samples studied were cleaved to be approximately 200\*100\*50$\mu m^3$ in size. The high pressure experiment were performed in a diamond anvil pressure cell (see figure \[photo\]). Argon is used as a pressure transmitter. The pressure is measured at low temperature by the shift of the ruby fluorescence line. In the ac calorimetry, a laser is used as heater. The beam is modulated using a mechanical chopper which works in the frequency range 50 Hz$<f<$5000 Hz. The temperature oscillations of the sample are measured with a Au/AuFe(0.07%) thermocouple which is spot welded on the sample. In the case of TmSe, it was glued with very diluted General Electric varnish. It is important that the thermocouple is welded in one point to avoid contributions of the thermoelectric power of the sample itself. A lock-in amplifier is used to measure the voltage of the thermocouple.
![Zoom on the high pressure cell. A thermocouple made of Au and AuFe is welded on the sample. Argon is used as pressure medium. The pressure is measured due to the fluorescence shift of ruby. The diamater of the hole is about 350 $\mu m$.[]{data-label="photo"}](figure2.eps){width="5cm"}
The measurements were performed in a $^4$He bath cryostat.
####
This experimental situation can be described by a first order model neglecting all internal time constants between sample, heater and thermometer[@accalorimetry] : $T_{\rm ac} = \frac{P_0}{\kappa + i \omega C}$, where $T_{\rm ac}$ is the amplitude of the temperature oscillation, $P_0$ the average power transmited, $\kappa$ the thermal conductivity to the bath and $C$ the specific heat. Even if the leak $\kappa$ is unknown, the phase measured by the lock-in is supposed to give the possibility to extract the value of the specific heat. $C=\frac{P_0.S}{V.\omega} \sin (\phi -\phi_0)$, where $V$ is the voltage of the thermocouple, $S$ its relative thermopower and $(\phi - \phi_0)$ the phase of the signal. If we want to minimize the importance of the phase correction, the choice of the frequency is crucial, as it balances the importance of the specific heat compared to the leak in the signal measured. From this point of view (without considering noise problems due to a decrease of the signal at high frequency), the frequency should be the highest possible. However, the experiment will show that this model is no longer valid at higher frequencies. If the frequency is too high, the sample decouples from the thermocouple and, the thermocouple can be directly excited by the laser and measure only its own temperature at high frequency[@chrisetmam].
![Schematic view of the thermal system in the pressure cell. The laser gives the power $aP_0$ and $(1-a)P_0$ respectively to the thermocouple and the sample[]{data-label="schema"}](figure3.eps){width="7cm"}
The next step is to include a thermal conductivity $\kappa_S$ between the sample and the thermocouple and to consider that a small proportion $a$ of the power is directly received on the thermocouple. In this situation (see figure \[schema\]), $T_{\rm ac}$ can be reestimated:[@TheseMAM] $T_{\rm ac} = \frac{P_0.(1-\frac{\kappa_{eff}}{\kappa_S})}{\kappa_{eff}}\frac{1 + i \omega a \frac{C}{\kappa_S}}{1 + i \omega \frac{C}{\kappa_{eff}}}$ with $\kappa_{eff}=\frac{\kappa \kappa_{S}}{\kappa+\kappa_S}$ representing the total parallel thermal conductivity of leak. Three different limits can be distinguished :
- At low frequency, if $\omega C \ll \kappa_{eff}$ then $T_{\rm ac}= \frac{P_0.(1-\frac{\kappa_{eff}}{\kappa_S})}{\kappa_{eff}}$. The value of the basic model is recovered : the phase of the signal is nearly zero and the inverse of the module is small.
- For the intermediate regime $\kappa_{eff} \ll \omega C \ll \kappa_S$ we recover also the basic model $T_{\rm ac}= \frac{P_0.(1-\frac{\kappa_{eff}}{\kappa_S})}{\kappa_{eff}+i\omega C}$. In good conditions, if frequency becomes high enough compared to the leak, the phase reaches nearly -$\frac{\Pi}{2}$.
- Finally, at high frequency, for $\kappa_{S} \ll \omega C$, $T_{\rm ac}=(1-\frac{\kappa_{eff}}{\kappa_S})\frac{a P_0}{\kappa_S}$.The phase reaches zero and the module decreases again. Physically, the thermocouple is decoupled from the sample.
To view more clearly the frequency dependence of the system, let us consider the complex number $\frac{1}{T_{\rm ac}}$. The phase measured by the lock in is directly the opposite of the phase of this complex number, and the signal $\frac{1}{V}$ is directly linked to the module of this complex number. Part (a) of the figure \[complex\] explains the different regimes depending on the frequency. From that picture, we can roughly draw the shape of the phase and of the inverse of the module (see part (b) of the figure \[complex\])
![Part (a) shows graphically the inverse of $T_{\rm ac}$ in the complex representation. Leak phenomena is on the real axis whereas capacitive effects are on the imaginary axis. The three limit cases are : (1), capacitive effect is negligible and the power $P_0$ is transmited to the bath with the leak $\kappa$ ; (2), capacitive effect becomes dominant and the component $\omega C$ is added ; (3) The sample is decoupled from the thermocouple. the power received is only the fraction $a P_0$ and the main leak is still $\kappa_S$, towards the sample. In part (b), the schematic shape of the phase and the module of $\frac{1}{T_{\rm ac}}$ is deducted from the evolution drawn in part (a). The vertical dashed lines show the cut off frequencies $\frac{\kappa_{eff}}{C}$ and $\frac{\kappa_{S}}{C}$ which indicate the change of regime corresponding to (1), (2) and (3).[]{data-label="complex"}](figure4a.eps "fig:"){width="7cm"} ![Part (a) shows graphically the inverse of $T_{\rm ac}$ in the complex representation. Leak phenomena is on the real axis whereas capacitive effects are on the imaginary axis. The three limit cases are : (1), capacitive effect is negligible and the power $P_0$ is transmited to the bath with the leak $\kappa$ ; (2), capacitive effect becomes dominant and the component $\omega C$ is added ; (3) The sample is decoupled from the thermocouple. the power received is only the fraction $a P_0$ and the main leak is still $\kappa_S$, towards the sample. In part (b), the schematic shape of the phase and the module of $\frac{1}{T_{\rm ac}}$ is deducted from the evolution drawn in part (a). The vertical dashed lines show the cut off frequencies $\frac{\kappa_{eff}}{C}$ and $\frac{\kappa_{S}}{C}$ which indicate the change of regime corresponding to (1), (2) and (3).[]{data-label="complex"}](figure4b.eps "fig:"){width="7cm"}
Moreover, if we consider the variable change $\omega\leftrightarrow \omega C$, the shape of the dependance in $\omega$ of the argument and module of $\frac{1}{T_{ac}}$(figure \[complex\]b) can be expanded to the dependance in $\omega C$. Then, considering a jump in the specific heat $C$ at the magnetic transition, the phase will be changed differently at high and low frequency. Around $\omega=\frac{\kappa_{eff}}{C}$ the signal in the phase will be a negative peak, but around $\omega=\frac{\kappa_S}{C}$, the signal of the phase can be a positive peak. This will be confirmed later by the experimental results.
Thus, the best frequency for the measurement is between this two cut-off frequencies. Typically, the best frequency was about 90 Hz for TmSe, 800 Hz for TmS and 4500 Hz for [SmB$_6$]{}. Assuming that the specific heat of TmSe is higher than that of TmS which is higher than that of [SmB$_6$]{} (at least at low pressure as indicated in figure \[c0\]), this support the model since in the conditions of measure $C \omega$ stays roughly constant.
####
Nevertheless, even if the behaviour of the phase is understood, the incertitude on the reference phase $\phi_0$ and the complex influence of pressure keep the situation delicate. Therefore, in the following, we will usually estimate the specific heat via the simplest expression : $C=\frac{P_0.S}{V.\omega}$.
####
The main point of the apparatus is the possibility to change the pressure at low temperature and also to use a excellent hydrostatic medium (Ar or He). To improve the faisability of the difficult microcalorimetric measurement under hydrostatic pressure, the choice has been made to minimize the number of electrical leads and thus, to use a laser as heater. The advantage of the technique is to give the pressure variation of the Néel temperature with a great accuracy i.e. a large set of pressure. If it is an excellent method to determine the phase diagram, the difficulty is to extract the specific heat in absolute units.
Results
=======
Preliminary results {#preliminary-results .unnumbered}
-------------------
![Specific heat divided by temperature measured at ambiant pressure for TmSe[@Berton], TmS[@TmSPzero], SmB$_6$(measured on our sample) and CeB$_6$[@CeB6][]{data-label="c0"}](figure5.eps){width="7cm"}
####
Before discussing the specific heat under pressure, we present the specific heat at ambiant pressure for the different systems in figure \[c0\]. The behaviour of $C$ for the two Tm compounds is quite different. For TmSe, the specific heat has a sharp anomaly at [$T_{N}$]{}[@Berton]. TmS is metallic and the crystal field ground state may be a singlet. Here, large fluctuations are already oberved above [$T_{N}$]{}[@TmSPzero]. In the other case, as [SmB$_6$]{} is non magnetic we have reported here the results for CeB$_6$[@CeB6] in order to have an idea of the amplitude of the signal under pressure. The comparison is worthwhile as both ($4f^{1}$) Ce$^{3+}$ and ($4f^{5}$) Sm$^{3+}$ are Kramer’s ions with the same angular momentum $J=\frac{5}{2}$ with a lifting of the degeneracy by the crystal field in a $\Gamma_{7}$ doublet and a $\Gamma_{8}$ quartet. The successive transitions observed for CeB$_6$ are now well understood by a cascade from paramagnetism to quadrupolar ordering at $T_{N_{1}}\sim 2.9$ K and to dipolar ordering at $T_{N_{2}}\sim 2.2$ K, the crystal field ground state being a $\Gamma_{8}$ quartet[@Shiina98].
TmSe {#tmse .unnumbered}
----
####
The temperature dependence of the specific heat of TmSe has been measured in a wide pressure range (1 to 14 GPa). Raw data are plotted in figure \[raw\_TmSe\] for different pressures.
![Raw data measured for TmSe, for several pressures (1.5, 5 and 11.2 GPa.). Module data are normalized at high temperature ; therefore, the three curves looks continuous after the anomaly.[]{data-label="raw_TmSe"}](figure6a.eps "fig:"){width="7cm"} ![Raw data measured for TmSe, for several pressures (1.5, 5 and 11.2 GPa.). Module data are normalized at high temperature ; therefore, the three curves looks continuous after the anomaly.[]{data-label="raw_TmSe"}](figure6b.eps "fig:"){width="7cm"}
The modulus and also the phase show a clear anomaly at the magnetic transition. The measurement has been realised at low frequency, so that the magnetic anomaly is seen in the phase as a negative peak. A second very sharp positive peak is also observed, inside the first negative peak, especially at low pressure. A simulation[@TheseMAM] shows that the huge value of the specific heat jump in TmSe can induce this second positive peak changing from the regime of low $\omega C$ to the one of high $\omega C$. Nevertheless, the strongly negative peak of the phase shows that we are in the low frequency regime where phase correction is supposed to be used. Moreover, as the signal is huge on the modulus, the phase correction is not significant (This is explained because close to $-\frac{\Pi}{2}$ the sinus is not really sensitive). Thus we present here the simplest estimation of the specific heat $C=\frac{P_0.S}{V.\omega}$. Some of the calculated curves are plotted in figure \[anomaly\_TmSe\].
![Evolution of the specific heat anomaly of TmSe under pressure. Data are normalized at high temperature and plotted for 1.5, 5, 7.6, 11.2 and 13.1 GPa.[]{data-label="anomaly_TmSe"}](figure7.eps){width="7cm"}
####
A first observation is that the magnetic anomaly is very well defined, so that, the Néel temperatures can be easily extracted. To define [$T_{N}$]{} we choose the maximum of the anomaly. Furthermore, we found an unexpected broadening of the anomaly as the pressure increases. This appears below 10 GPa, when hydrostaticity is still very good (less than 0.1 GPa of variation in the cell)[@jean]. This broadenning cannot be explained by pressure inhommogeneities as $\frac{dT_N}{dp}$ is small. Figure \[split\] shows $\frac{C}{T}$ for pressures above 10 GPa. The data indicate a splitting of the magnetic anomaly, which can be followed under pressure.
![Observation of the spliting of the specific heat anomaly of TmSe at high pressure.[]{data-label="split"}](figure8.eps){width="6cm"}
The resulting phase diagram is then represented in figure \[phasediagram\_tmse\]
![(p, T) Phase diagram of TmSe. The squares represent the maximum of each specific heat anomaly, and the circles indicate the second maximum observed at high pressures. The dashed line show the phase transition coresponding to the slope change.[]{data-label="phasediagram_tmse"}](figure9.eps){width="7cm"}
####
The phase diagram can be distinguished in three parts. At low pressures, the evolution of T$_N$ is quite flat and a maximum can be seen around 1.3 GPa. Then, a break in the slope around 3 GPa corresponds to the pressure of transition from the insulating AF1 phase to the metallic AF2 phase. The second magnetic structure is caracterized by a linear increase of the Néel temperature with pressure.
####
At low pressure, our data are completely consistent with previous resistivity measurements[@Ribault; @Ohashi1; @Ohashi2]. The important observation is the continuous increase of [$T_{N}$]{} with pressure at high pressure. Contrary to recent resistivity measurements who showed a discontinuity in the Néel temperature around 6 GPa[@Mignot], no anomaly in $T_N(p)$ is seen in our data. Actually, our observation is consistent with a release of the 5d electrons near 3 GPa. Recent neutrons measurements[@Mignot] confirm this idea as no change in the magnetic structure is found at 6 GPa. Finally, an interesting splitting of the magnetic anomaly is observed at high pressure, above 10 GPa. The evolution of the signal shape was detailled in figure \[split\]. The origin of this splitting and the new phase is not clear. This observation pushs us to study TmS, which can be seen as high pressure analog of TmSe. In TmS, evidence has been reported for two different magnetic phases [@OashiTmS] around 5 GPa.
TmS {#tms .unnumbered}
---
####
The specific heat of TmS was measured up to 19 GPa. Raw data are plotted for several pressures in figure \[raw\_tms\].
![Raw data measured for TmS. The module has been drawn for 1.8, 11.3 and 18.7 GPa. The signal has been followed until very high pressure, but at the end, it desapears as shows the curve at 18.7 GPa. Then, the behaviour of the phase has been detailled : the measurement at low frequency ($200$ Hz) and high frequency ($800$ Hz) have been compared and followed with pressure from 1.8 GPa to 2.7 GPa. For the picture the phase have been arbitrary shifted.[]{data-label="raw_tms"}](figure10a.eps "fig:"){width="7cm"} ![Raw data measured for TmS. The module has been drawn for 1.8, 11.3 and 18.7 GPa. The signal has been followed until very high pressure, but at the end, it desapears as shows the curve at 18.7 GPa. Then, the behaviour of the phase has been detailled : the measurement at low frequency ($200$ Hz) and high frequency ($800$ Hz) have been compared and followed with pressure from 1.8 GPa to 2.7 GPa. For the picture the phase have been arbitrary shifted.[]{data-label="raw_tms"}](figure10b.eps "fig:"){width="7cm"}
The behaviour of the phase is detailled for the low pressures. The previous explanation is confirmed : we can observe two different regimes for the phase, depending if the measurement is performed at low or high frequency. This is really reproducible and stable with pressure change. That confirms that the feature occuring on the phase is very useful to detect the magnetic transition. Unfortunately, in the low frequency regime, the feature on the modulus is very small and don’t allow us to extract a good shape of the specific heat. On the other hand, figure \[raw\_tms\] shows that the module measured at high frequency is more clear. Even if the first order model is valid only at low frequency, figure \[complex\]b shows that the evolution of the module is still monotonous even after the first cut off. Thus, in order to avoid a correction with an arbitrary phase $\phi_0$, we prefer to show the estimation at zero order of the specific heat at 800 Hz. Some selected pressures are shown on figure \[anomaly\_tms\].
![Evolution of the magnetic anomaly of TmS under pressure ; specific heat divided by temperature has been normalized at high temperature and plotted for different pressures : 1.8, 4.8, 11.3 and 18.7 GPa[]{data-label="anomaly_tms"}](figure11.eps){width="7cm"}
####
Increasing the pressure, the maximum is shifted to higher temperature, from 6 to 12 K. Until 15 GPa the signal is only slighty broadened, and still very clear, but at higher pressure, the signal decreases. The phase diagram of TmS is shown in figure \[phasediagramm\_TmS\]. Anomalies found in previous resistivity measurements[@OashiTmS] and neutron scattering[@Mignot] have also been plotted. $T_1$ and $T_2$ are kinks observed in the resistivity curve. $T_1$ looks linked to $T_N$ and $T_2$ indicates a new phase which has also been evidenced by neutron scattering.
 and resistivity measurements[@OashiTmS] (empty symbols). []{data-label="phasediagramm_TmS"}](figure12.eps){width="7cm"}
![Combination of the phase diagram of both TmSe and TmS. For the abscisse axis, we have choosen a typical volume linked to the pressure in the TmSe compound. That means that pressure for TmS has been renormalized .[]{data-label="phasediagramm_combine"}](figure13.eps){width="7cm"}
Our study indicates a linear $p$ increase of the Néel temperature. This observation differs from published results obtained by resistivity or neutron scattering experiments. The sensitivity of TmS to defects is well known. At ambiant pressure, the value of [$T_{N}$]{} is sample dependent and varies between 5.2 K and 7.05 K[@TmSdefect; @TmSPzero]. Our sample comes from the same batch than the crystal measured in reference[@TmSPzero] where excellent agreements was found between different methods in the [$T_{N}$]{} determination. The second anomaly below [$T_{N}$]{} observed by neutron scattering in the $p$ range above 5 GPa is due to a “lock-in” transition from an incommensurate to a commensurate structure. Therefore, if entropy is just slighty changed, it might be not detected by our specific heat measurement. Of course, an open question is again here the reproductibility of this second anomaly with the defects’ content
In order to compare these results to TmSe, we have scaled the pressure applied on TmS, into an equivalent pressure applied on TmSe, to obtain the same volume. The pressure range has been shifted of 7 GPa, corresponding to the value where TmSe is more or less trivalent, and then normalized by the ratio of the compressibility of the two compounds (1.5 $10^{-6}$ bar$^{-1}$ for TmS and 3.5 $10^{-6}$ bar$^{-1}$ for TmSe from reference[@compress1; @compress2]). The resulting phase diagram is plotted in figure [\[phasediagramm\_combine\]]{}.
[$T_{N}$]{} of TmS scales very well to TmSe. Of course, the points of TmSe don’t follow completely the same alignement at too high pressure: the TmSe measurements themselfs have to be renormalized at very high pressure as the compressibility of TmSe decreases[@compress1; @compress2].
SmB$_6$ {#smb_6 .unnumbered}
-------
####
Finally, similar experiments have been performed for SmB$_6$. Long range magnetic ordering has been found above 8 GPa[@BarlaSmB6]. The features observed on the raw data are already clear. They have been plotted in figure \[raw\_smb6\_2\]
![Raw data measured for SmB$_6$ for several pressures (4.7, 10.2 and 12.5 GPa.). Module data are normalized at high temperature[]{data-label="raw_smb6_2"}](figure14a.eps "fig:"){width="7cm"} ![Raw data measured for SmB$_6$ for several pressures (4.7, 10.2 and 12.5 GPa.). Module data are normalized at high temperature[]{data-label="raw_smb6_2"}](figure14b.eps "fig:"){width="7cm"}
and the magnetic anomaly shows clearly up in the modulus. The feature in the modulus is so huge, and we never reach the “high frequency regime” with inversion of the phase, even for the highest frequency allowed by the set up. Thermal contact between the sample and the thermocouple was very good. Therefore, the specific heat has been extracted only from the modulus measured at very high frequency, and a selection of the results have been plotted in figure \[anomaly\_smb6\]. With increasing pressure, the anomaly gets more and more pronounced. Contrary to the case of TmSe, the peak gets sharper, even above 10 GPa.
![Growth of the magnetic anomaly of [SmB$_6$]{} under pressure ; specific heat divided by temperature has been normalized at high temperature and plotted for different pressures : 4.7, 7.8, 8.7, 10.2, 10.6 and 12.5 GPa[]{data-label="anomaly_smb6"}](figure15.eps){width="7cm"}
####
The phase diagram of [SmB$_6$]{} is shown in figure \[smb6diagram\]. We choosed as criterium for [$T_{N}$]{} the maximum of the anomaly in $\frac{C}{T}$. In order to look more carefully at the change of the shape of the signal, we have also investigated the broadening of the anomaly which is plotted in figure \[smb6diagram\] too.
![Phase diagram of [SmB$_6$]{}. The Néel temperature (dark square) and the broadening $\Delta$T$_N$ have been ploted in Kelvin (the broadening is the width ot the anomaly peak at half of the height). We have also plotted [$T_{N}$]{} for another cell measured previously (light square). These results are compared to the magnetic fraction measured by NFS [@BarlaSmB6]. The vanishing of the gap is also represented by arrows coresponding to different studies : 1-Sample given by K. Flachbart measured in the laboratory, 2-ref.[@Gabani], 3-ref.[@Moshchakov], 4-ref.[@Cooley], 5-Sample grown by G. Lapertot and measured in the laboratory and 6-ref.[@Beille]. The dashed box shows the wide pressure range corresponding to the collapse of the hybridisation gap observed in different samples[]{data-label="smb6diagram"}](figure16.eps){width="7cm"}
####
The evolution of the broadening, shows that the anomaly peak is first very broad and then sharper. Moreover, a change of regime appears around 10 GPa. This change is significant as we can observe a clear change in the slope of the broadening i.e. roughly at the pressure where 100% of magnetic sites has been detected by NFS[@BarlaSmB6].
Experimental conclusion {#experimental-conclusion .unnumbered}
-----------------------
####
It has been shown that the experimental set up of the cell is critical to obtain correct shapes of the specific heat. Especially the link between the thermocouple and the sample must be very good. The main incertitude concerns the knwoledge of the reference phase $\phi_0$. With that information, it could be possible to correct the variation due to the leak but non monotonous behaviour of the phase before 4 K (certainly due to a $T$ dependence of $\kappa(T)$) has discouraged us to associate $\phi_0$ with the phase measured at low temperature. So that, the extraction of an absolute value of the specific heat remains difficult and as the phase correction is generally small, we have prefered to show here estimation derived only from the module. Nevertheless, the method is very useful to detect the pressure induced phase transitions (here, long range magnetism), and the in situ pressure generation gives a fine pressure tuning. Thus this technique is well adapted to draw phase diagrams.
####
The main experimental problem is to understand the broadening and the loss of the magnetic anomaly under pressure. One could imagine that the thermal contact between the sample and the thermocouple is one of the issue. But, as the systems were well welded, we don’t believe in a loss of the contact. Another consideration is the behaviour of the thermal leak, as it can become huge at high pressure. The first guilty phenomenum is the argon conductivity, if we extrapole some conductivity measurements done at higher temperature[@argon], the conductivity increases with more than a factor 10 between 1 and 10 GPa. At 1 GPa, the two terms $\omega C$ and $\kappa$ can already be estimated of the same order (10$^{-3}$ W K$^{-1}$), so that a factor 10 will be a huge effect for the relative signal measured at 10 GPa. Of course this doesn’t explain the relative sudden character of the effect as $\kappa$ increases roughly linearly. If we consider the big compressibility of the argon[@argonbis] we can expect a reduction of the volume of the pressure chamber of the order of 30% .In this case, if you consider the geometry of the chamber (see figure \[photo\]), a possible contact could occur between the sample and the gasket at high pressure ; this could imply a big thermal contact, and a sudden increase of the thermal leak. In the case of TmSe, the anomaly is lost quickly (before 10 GPa) but for other compounds, the set up allows us to follow correctly the magnetic anomaly until around 15 GPa.
####
For [SmB$_6$]{} , the situation is completely different as the broadening occurs at low pressure. There are two possible explanations for the broadening of the magnetic anomaly. First, if we assume a very sharp transition (as it seems to be, since T$_N$, nearly jumps from zero to its maximum value), the broadening could be the effect of the pressure inhomogeneity, as even a small pressure gradient would imply a large average of the Néel temperatures. Nevertheless, to explain the experiment, one has to assume a quite big inhomogeneity of the order of 1 GPa. Typical deviation is about only 0.1 GPa[@jean]. Therefore, a sound explanation is to consider the observed broadening as the signature of an intrinsic phenomena which may be a mixed state linked to a first order transition. The system becomes homogeneous and reach a full long range magnetic ordering only at high pressure. This idea is consistent with NFS measurements which evidences a coexistence of two phases between 5 and 10 GPa.
Discussion
==========
####
There are different approaches for the description of magnetism of TmSe and SmB$_6$ ; but, in order to make a comparison between Tm and Sm, and even with the case of Ce and Yb, we will assume that each integer valent configuration is associated to a Kondo lattice temperature [$T_{KL}$]{} and that it is the comparison of this characteristic energy with other energy scales like the crystal field splitting or the magnetic intersite interaction which will be led to the renormalization towards a given configuration.
####
The basic idea[@Flouquet] is that, compared to a single impurity, due to the release of an itinerant electron related to the valence mixing, a feedback occurs between the Kondo effect and the number of itinerant electrons. In analogy, to the theoretical results known for the Kondo effect of the cerium ion in the $\frac{1}{N_f}$ expansion[@Hewson], we will assume that for the $3+$ configuration, $T_{K}^{3+}= (1-n_f)N_f\Delta_0$, where [n$_f$]{} is the occupation number of the trivalent state, $N_f$ the degeneracy ($N_f=2J+1$) and $\Delta_0$ the width of the virtual 4f level in the Anderson lattice related to the density of states of the light conduction electron ($N(E_f)$) and to the hybridization mixing potential ($V_{df}$) : $\Delta_0=\pi V_{df}^2 N(E_f)$. Of course $\Delta_0$ must be very sensitive to the spatial extension of the 4f orbits. One can note that the usual Kondo formula of the susceptibility $\chi$ will be recovered for the Cerium case as it will correspond to $\chi^{3+}$ [n$_f$]{} i.e. to [$T_{K}$]{} $=\frac{T_{K}^{3+}}{n_f}$.
####
In the so called [f$^{1}$]{}-[f$^{2}$]{}model (instead of the [f$^{0}$]{}-[f$^{1}$]{}model suitable for the Cerium electron case, and for the Kondo hole analog ytterbium), there are theoretical studies on TmSe[@Flouquet; @Newns; @Read; @Nunes; @Yafet; @Saso], for Tm impurity, with [n$_f$]{} going from zero ([f$^{1}$]{}) to one ([f$^{2}$]{}). Basically, the large $\frac{1}{N_f}$ theory leads to very similar physics than that of the [f$^{0}$]{}-[f$^{1}$]{}model with however, a maxima of the Kondo temperature around [n$_f$]{}$\sim 1.7$. A discussion on the Kondo effect on Sm ions can be found in reference[@Sakai]. For the cerium case, [$T_{K}$]{} will continuously increase as [n$_f$]{} decreases. Yb HFC are often viewed as the hole analog (4f$^{13}$ configuration for Yb$^{3+}$) of the Ce HFC with a decrease of [$T_{K}$]{} under pressure. The Tm compounds are always magnetically ordered as the exchange energy always exceeds $T_K$ , either of [Tm$^{2+}$]{} or of [Tm$^{3+}$]{}.
####
Our physical picture stresses out the role of the valence mixing and the release of the 5d electron. This is the key point concerning the magnetic ground state but also the electronic ground state. That pushes us to extend the Kondo temperature formula to the lattice where the virtual bound width $\Delta_0$ is now directly related to the bare bandwidth $D$ of the 5d light conduction electrons : $\Delta_0=\alpha D$, with $D$ depending on [n$_f$]{} and $\alpha$ typically of the order of $10^{-2}$ in order to recover a narrow virtual bound state for the impurity. The change of the numbers of carrier will give here $D(n_f)=D_0 n_f^{\frac{2}{3}}$.
####
If we apply this rule to TmSe, the Kondo temperature $T_{KL}$ of the trivalent and divalent configuration in a lattice will be :
- [$T_{KL}^{3+}$]{}$=\alpha D_0 (1-$[n$_f$]{}$)$[n$_f$]{}$^{\frac{2}{3}}N_f^{3+}$
- [$T_{KL}^{2+}$]{}$=\alpha D_0 $[n$_f$]{}$^{\frac{5}{3}}N_f^{2+}$
where the degeneracy $N_f^{3+}$ and $N_f^{2+}$ are respectevely 13 and 8.
![Comparison of the Kondo lattice temperature and the crystal field spliting in arbitrary units, in the case of TmSe, for both the 2+ and the 3+ configuration. With the criteria choosen, long range magnetism is allowed if T$_{KL}$ is smaller than [$\Delta_{CF}$]{}.[]{data-label="simultmse"}](figure17.eps){width="7cm"}
Figure \[simultmse\] represents the Kondo temperature for the two configurations. A typical value of the overall crystal field splitting $\Delta_{CF}$ has been added to the plot. Of course, a crucial point has been to choose the ratio between [$\Delta_{CF}$]{} and $D_0$ to compare [$\Delta_{CF}$]{} with $T_{KL}$. In order to have a coherent behaviour, we put $\frac{\alpha D_0}{\Delta_{CF}}\sim 4$ which correspond to a very small effective bandwidth. Anyway, if we assume that [$\Delta_{CF}$]{} $\sim 100$ K[@Berton], $\alpha D_0$ can be nearly the order of magnitude of 400 K. The different energies has been traced versus [n$_f$]{}, varying in the same way as the pressure. If there is an extra effect as a electron gap, a simple way would be to add an extra pressure dependence on $D$ ($D=0$ for $p<p_{\Delta}$).
####
Extrapolating from the numerous studies performed on Ce HFC, the occurence of long range magnetism requires at least the recovery of usual rare earth properties, notably the full reaction to the crystal field splitting i. e. k$_BT_{KL}<\Delta_{CF}$. Of course a main consideration is the relative strength of the intersite exchange interaction and [$T_{KL}$]{} as discussed for the usual Doniach model. Long range magnetism will occur only if the energy scale of the coupling is stronger than the Kondo energy. Nevertheless, in our simple view, we compare only $\Delta_{CF}$ and [$T_{KL}$]{}. Therefore, we only indicate when long range magnetism will be possible. For each configuration, long range magnetism will be possible while $T_{KL}$ is smaller than [$\Delta_{CF}$]{}; that means we assume the coupling is already strong enough. Of course, the position of the intersections are very sensitive to the ratio $\frac{\Delta_{CF}}{\alpha D_0}$. Nevertheless, this basic model explains qualitatively the general shape of the phase diagram. At low pressure, [n$_f$]{} is small, and long range magnetism is due to Tm$^{2+}$ ; then at higher pressure, when [n$_f$]{} increases, this long range magnetism disapears and long range magnetism due to Tm$^{3+}$ appears. The change of regime observed at 3 GPa is well reproduced. At this critical pressure, a critical value of [n$_f$]{} is reached, where the renormalization of the wavefunction changes from the 2+ to the 3+ ground state since the Kondo effect becomes crucial for the Tm$^{2+}$ ions and is not strong enough for Tm$^{3+}$.
![Comparison of the Kondo lattice temperature and the crystal field spliting in arbitrary units, for the 3+ configuration. In the case of [SmB$_6$]{} as only the trivalent state is magnetic, long range magnetism will be allowed if T$_{KL}^{3+}$ become smaller than [$\Delta_{CF}$]{}.[]{data-label="simul_smb6"}](figure18.eps){width="7cm"}
For Sm Kondo lattice, the previous formula of [$T_{KL}^{3+}$]{} is plotted in figure \[simul\_smb6\]. Now, $\frac{\alpha D_0}{\Delta_{CF}}\sim 2$ is choosen. The interpretation is the same : a magnetic ground state is possible when its trivalent Kondo lattice temperature (in charge of the long range magnetism here, since Sm$^{2+}$ is non magnetic) is low enough compared to a crystal field energy. At low value of [n$_f$]{} the long range magnetism will disappear as the exchange energy will drop.
####
The important point is that [$T_{KL}^{3+}$]{} reaches a broad maxima near [n$_f$]{}$=0.4$. This is a critical difference with the Cerium case which correspond to the release of the 4f electron from the 4f shell : Ce$^{3+}\Longleftrightarrow$ Ce$^{4+}+5d$. If no extra electron is considered, one may find that [$T_{KL}^{3+}$]{} goes, in Ce case, as [$T_{KL}^{3+}$]{}$= (1-$[n$_f$]{}$)^{\frac{5}{3}} \alpha D_0 N_f$ (with $N_f=2J+1=6$). By contrast to the previous case, [$T_{KL}^{3+}$]{} never reaches a maximum in the Ce case. Actually, this naive scheme gives the correct result that in Ce HFC, [$T_{K}$]{} decreases continuously with increasing [n$_f$]{}.
####
In those considerations, there is the underlining assumption that the valence fluctuation can be slow enough to follow the motion of the spin dynamics of the trivalent configuration even for n$_f \sim 0.8$ as observed in SmS, [SmB$_6$]{} by NFS or in YbRh$_2$Si$_2$[@Sichelschmidt03]. This suggest that the 4f-5d correlation is a favorable factor to slow down the valence fluctuation. This consideration lead to propose that SmB$_6$, like SmS, can be regarded in the low pressure gold phase ($p<p_{\Delta}$) as an excitonic dielectric semiconductor with the electron promoted from f shells spread over the p orbitals of neighboring boron sites but with the same symmetry as the f electron in the central Sm site[@Kikoin1; @Kikoin2]. An alternative idea is that the electron (5d) created by the mixing of the 4f state and the hole produced in the conduction band screen the 4f hole and form a bound state in a low carrier density medium[@Kasuya]. Up to now, there is no consideration on the pressure dependence of the 5d screening and thus on the disappearance of the reported many body effects. In term of a Kondo approach, one may think that one way to describe the extra many body effect is to consider the possibility of the Kondo effect of the 5d electron itself. A many body treatment will be required, so far its [$T_{K}$]{} (5d) is lower than its crystal field splitting $\Delta_{CF}$ (5d). Of course, [$T_{K}$]{} (5d) will be far greater than [$T_{K}$]{} (4f) but also $\Delta_{CF}$ (5d) $>\Delta_{CF}$(4f). A change will occur under pressure since in all reported cases (Sm$^{3+}$, Yb$^{3+}$, Tm$^{3+}$) their $T_K$(4f) decrease under $p$ while $T_K$(5d) increase with pressure. When $T_K$(5d)$>\Delta_{CF}$(5d) there will be no more reason to consider the extra many body effects of the 5d electron which could be considered then as dissolved in the Fermi sea.
####
In the case of TmSe, entering in the trivalent state, there are two reasons that the physics will be dominated by the formation of a magnetic moment on an initial singlet ground state : the Kondo effect and a probable singlet crystal field level. As pointed out, the two mechanisms leads to rather similar increase of the sublattice magnetization under pressure on increasing the intersite exchange coupling. Thus the difference in the crystal field ground state limits the comparison of TmSe with SmB$_6$ and SmS. However, let’s emphasize the similarity : up to [n$_f$]{}$\sim 0.8$, the physics appear renormalized to the divalent configuration, not only governing the magnetic properties, but also the electronic properties (formation of many body insulating state) ; above [n$_f$]{}$\sim 0.8$ the physics is now governed by the trivalent configuration (metallic conduction and nature of the magnetic order parameter).
####
Microscopic evidence to the 2+ configuration in SmS, [SmB$_6$]{} but even TmSe was given by inelastic neutron squattering experiments and measurement of the magnetic form factor[@MignotphysicaB; @Alekseev2002]. The demonstration of a dressing towards the 3+ configuration for SmS and [SmB$_6$]{} was done by NFS as both the quadrupolar and dipolar magnetic hyperfine structure are caracteristic of a 3+ state even for [n$_f$]{} $\sim 0.8$. Macroscopically, suggestions of the 2+ renormalization of TmSe at $p=0$ comes from the specific heat, and of the 3+ renormalization of TmSe above 3 GPa from its continuity with the quasitrivalent compound TmS.
####
Concerning SmS and [SmB$_6$]{} we have to be careful on the coincidence in the appearance of long range magnetism and closing of the hybridization gap. For SmS, the coincidence has been found. An extrapolation made from inelastic measurement on Sm$_{0.83}$Y$_{0.17}$S suggests strongly that Sm-Sm exchange interactions play a major role even in the low pressure gold phase[@Alekseev2002]. No similar influence is observed for [SmB$_6$]{} certainly due to the isolation of the Sm ion with the B cage. The gap is closed as function of pressure before long range magnetic ordering appears. Typically, the gap is closed between 4 and 6 GPa[@Gabani; @Cooley; @Moshchakov; @Beille], but long range magnetism do not appear before 8 GPa and a homogeneous magnetic phase picture without phase separation may occur only above 10 GPa.
####
Finally, the difference between [SmB$_6$]{} and SmS is not so surprising, as their band structures are completely different due to symmetries which are different. Local spin density approximation (LSDA+U approach) were published for Sm monochalcogenides[@Antonovsms] and [SmB$_6$]{}[@Antonovsmb]. For SmS, NaCl type crystal structure with the space group Fm3m, the occupation number [n$_f$]{} is found equal to 0.55 (valence $v=2.55$) in the low pressure gold phase, a non zero magnetic moment is always obtained. For [SmB$_6$]{} , CaB$_6$ type crystal structure with the space group Pn3m, the calculations produce always an integer valence ground state either divalent or trivalent. A small hybridization energy gap is recovered in [SmB$_6$]{} for samarium in the divalent state. It was emphasized that the magnetism of golden SmS as well as the formation of the IV state in [SmB$_6$]{} requires to go beyond this mean field approximation.
Conclusion
==========
####
Ac calorimetry with in situ $p$ variation at low temperature is a powerful technique to define without ambiguity the magnetic phase diagram under pressure. We hope that our experimental report may help to experimental progresses.
####
The common point in the three investigated systems TmSe, [SmB$_6$]{}, and SmS[@Haga; @BarlaSmS] is the link between the electric conduction and the renormalization to divalent or trivalent configurations at low temperature. Looking more deeply on SmS and [SmB$_6$]{}, a main difference appear between the clear onset of antiferromagnetism at $p_{\Delta}$ in SmS and the large pressure window in [SmB$_6$]{} ($6<p<10$ GPa) where an inhomogeneous behaviour is observed. A homogeneous magnetic phase occurs in [SmB$_6$]{} only above 10 GPa. It is amazing to observe that if $p_{\Delta}=2$ GPa is remarkably reproducible in SmS[@Haga], a large dispersion appears for [SmB$_6$]{} (around 3 GPa). The next step is to understand the role of the disorder in [SmB$_6$]{} and the impact on the collapse of the gap and the appearance of long range magnetism.
####
Finally by comparison to results on Ce intermetallic heavy fermion compounds, in these Sm and Tm systems, a long range magnetism characteristic of the trivalent configuration occurs far below the pressure where the trivalent state will be reached. This phenomena is quite similar to that observed in YbRh$_2$Si$_2$. Physically, the interesting fact is that both slow spin and valence fluctuations must interfer.
Aknowledgment {#aknowledgment .unnumbered}
=============
####
We would like to thank Christophe Marcenat for his precious clarification in the explanation of the phase behaviour of ac microcalorimetry measurements.
[99]{} J. Flouquet, Cond-mat/0501602 (2005) P. Wachter, Handbook of physics and chemistry of rare earths, Edited by K.A. Gschneider [*et al.*]{} North holland Amsterdam (1994). D. Malterre Adv. Phys. [**45**]{} 299 (1996) H. Launois, M. Rawiso, E. Holland-Moritz, R. Pott, and D. Wohlleben, Phys. Rev. Lett. [**44**]{} 1271 (1980).
P. Haen, F. Lapierre, J. M. Mignot, J. P. Kappler, G. Krill and M. F. Ravet, J. Magn. Magn. Mater. [**47-48**]{}, 490 (1985). G. K. Wertheim, W. Eib, E. Kaldis and M. Campagna, Phys. Rev. B [**22**]{}, 6240 (1980). W. D. Brewer, G. Kalkowski, G. Kaindl and F. Holtzberg, Phys. Rev. B [**32**]{}, 3676 (1985). E. Beaurepaire, J. P. Kappler and G. Krill, Phys. Rev. B [**41**]{}, 6768 (1990). J. Röhler et al., in Valence Instabilities, edited by P. Wachter and H. Boppart (North-Holland, Amsterdam, 1982), p. 215. C. Dallera, E. Annese, J-P Rueff, M. Grioni, G. Vanko, L. Braicovich, A. Barla, J-P. Sanchez, R. Gusmeroli, A. Palenzona, L. Degiorgi and G. Lapertot, J. Phys.: Condens. Matter [**17**]{} S849 (2005).
N. Ogita, S. Nagai, M. Udagawa, F. Iga, M. Sera, T. Oguchi, J. Akimitsu and S. Kunii, Physica B [**359-361**]{} 941 (2005).
A. Barla, J. P. Sanchez, Y. Haga, G. Lapertot, B. P. Doyle, O. Leupold, R. Ruffer, M. M. Abd-Elmeguid, R. Lengsdorf and J. Flouquet, Phys. Rev. Lett. [**92**]{} 066401 (2004). Y. Haga, J. Derr, A. Barla, B. Salce, G. Lapertot, I. Sheikin, K. Matsubayashi, N. K. Sato and J. Flouquet, Phys. Rev. B [**70**]{}, 220406 (2004) A. Barla, J. Derr, J. P. Sanchez, B. Salce, G. Lapertot, B. P. Doyle, R. Ruffer, R. Lengsdorf, M. M. Abd-Elmeguid and J. Flouquet, Phys. Rev. Lett. [**94**]{}, 166401 (2005). M. Ribault, J. Flouquet, P. Haen, F. Lapierre, J. M. Mignot and F. Holtzberg, Phys. Rev. Lett. [**45**]{} 1295 (1980). M. Ohashi, N. Takeshita, H. Mitamura, T. Matsumura, T. Suzuki, T. Goto, H. Ishimoto and N. Môri, Physica B [**259-261**]{}, 326-328 (1999). M. Ohashi, N. Takeshita, H. Mitamura, T. Matsumura, T. Suzuki, T. Mori, T. Goto, H. Ishimoto and N. Môri, J. Magn. Magn. Mater. [**226-230**]{}, 158-160 (2001).
F. Holtzberg, J. Flouquet, P. Haen, F. Lapierre, Y. Lassailly and C. Vettier, J. of Appl. Phys. [**57**]{} (1985) 3152-3153. B. Salce, J. Thomasson, A. Demuer, J. J. Blanchard, J. M. Martinod, L. Devoile and A. Guillaume, Rev. sci. Instrum. [**71**]{}, 2461 (2000). A. Demuer, C. Marcenat, J. Thomasson, R. Calemczuk, B. Salce, P. Lejay, D. Braithwaite and J. Flouquet, J. of low Temp Phys, [**120**]{} 245-57 (2000)
Paul F. Sullivan and G. Seidel, Phys. Rev. [**173**]{}, 679-685 (1968). C. Marcenat and M. A. Méasson, private communication. A. Berton, J. Chaussy, B. Cornut, J. Flouquet, J. Odin, J. Peyrard and F. Holtzberg, Phys. Rev. B, [**23**]{} 3504 (1981). A. Berton, J. Chaussy, J. Flouquet, J. Odin, J. Peyrard and F. Holtzberg, Phys. Rev. B, [**31**]{} 4313 (1985).
T. Fujita, M. Suzuki, T. Komatsubara, S. Kunii, T. Kasuya and T. Ohtsuka, Solid-State-Communications. Aug. 1980; 35(7): 569-72 R. Shiina, O. Sakai, H. Shiba and P. Thalmeier, J. Phys. Soc. Jpn. [**67**]{} 941 (1998). M. A. Méasson, Thèse de doctorat, Université Joseph Fourrier, Grenoble (2005).
J. Thomasson [*et al.*]{} Private communication.
J. M. Mignot, I. N. Goncharenko,P. Link,T. Maysumura and T. Suzuki, Hyperfine interactions [**128**]{} (2000) 207-224.
M. Ohashi, N. Takeshita, H. Mitamura, T. Matsumura, Y. Uwatoko, T. Suzuki, T. Goto, H. Ishimoto and N. Môri, Physica B [**281-282**]{}, 264-266 (2000).
G. Chouteau, F. Holtzberg, O. Pena, T. Penney and R. Tournier, J. Physique, colloq 40 (1979) C5-361. Y. Lassailly, C. Vettier ,F. Holtzberg, J. Flouquet, C. M. E. Zeyen and F. Lapierre, Phys. Rev. B 28, 2880-2882 (1983) B. Batlogg, H. R. Ott, E. Kaldis, W. Thöni, and P. Wachter, Phys. Rev. B [**19**]{}, 247-259 (1979).
S. Gabani, E. Bauer, S. Berger, K. Flashbart, Y. Paderno, C. Paul, V. Pavlík and N. Shitsevalova, Phys. Rev. B, [**67**]{} 172406 (2003). V.V. Moshchalkov, I. V. Berman, N.B. Brandt, S. N. Pashkevich, E.V. Bogdanov, E. S. Konovalova and M. V. Semenov, J. Magn. Magn. Mater. [**47-48**]{}, 289-291 (1985). J. C. Cooley, M. C. Aronson, Z. Fisk and P. C. Canfield, Phys. Rev. Lett. [**74**]{} 9 (1995). J. Beille, M. B. Maple, J. Wittig, Z. Fisk and L. E. Delong, Phys. Rev. B, [**28**]{} 12 (1983).
K. V. Tretiakov and S. Scandolo, Journal of Chemical Physics. (2004) 121(22) 11177-82 H. Shimizu, H. Tashiro, T. Kume, S. Sasaki, Phys. Rev. Lett. [**86**]{} 20, 4568-71 (2001)
J. Flouquet, A. Barla, R. Boursier, J. Derr and G. Knebel, J. Phys. Soc. Jpn. [**74**]{} 178-185 (2005).
A. C. Hewson, The kondo problem to heavy fermions (Cambridge University Press, Cambridge, 1993). D.M. Newns and N. Read, Adv in Physcics, [**36(6)**]{} 799-849 (1987) N. Read, K. Dharamvir, J. W. Rasul and D. M. Newns, J. Phys. C [**19**]{} (1986) 1597. A. C. Nunes, J. W. Rasul and G. A. Gehring, J. Phys. C [**19**]{} (1986) 1017. Y. Yafet, C. M. Varma and B. Jones, Phys. Rev. B, [**32**]{} 360 (1985). T. Saso, J. Magn. Magn. Mater. [**76-77**]{}, 176-178 (1988).
O. Sakai, Y. Shimizu and T. Kasuya, Prog. Theor. Physics, [**108**]{} 73 (1992).
J. T. Waber and D. T. Cromer, J. Chem. Phys. [**42**]{} (1965) 4116. H. Harima, private communication. J. Plessel, M. M. Abd-Elmeguid, J. P. Sanchez, G. Knebel, C. Geibel, O. Trovarelli and F. Steglich, Phys. Rev. B, [**67**]{} 180403 (2003). G. Knebel, V. Glaskov, A. Pourret, P. G. Nicklowitz, G. Lapertot, B. Salce and J. Flouquet, Physica B [**359-361**]{} (2005) 20-22.
J. Sichelschmidt, V. A. Ivanshin, J. Ferstl, C. Geibel, and F. Steglich, Phys. Rev. Lett. [**91**]{} 156401 (2003). K. A. Kikoin and A. S. Mishchenko, J. Phys. cond. matt. [**7**]{} 307 (1995). S. Curnoe and K. A. Kikoin, Phys. Rev. B, [**61**]{} 15714 (2000). T. Kasuya, Europhysics letter [**26**]{} 277 (1994).
J. M. Mignot and P. A. Alekseev, Physica B [**215**]{}, 99 (1995).
P. A. Alekseev, J.-M. Mignot, A. Ochiai, E. V. Nefeodova, I. P. Sadikov, E. S. Clementyev, V. N. Lazukov, M. Braden, and K. S. Nemkovski, Phys. Rev. B, [**65**]{} 153201 (2002).
V. N. Antonov, B. N. Harmon and A. N. Yaresko Phys. Rev. B, [**66**]{} 165208 (2002). V. N. Antonov, B. N. Harmon and A. N. Yaresko Phys. Rev. B, [**66**]{} 165209 (2002).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Concept Analysis [@Wi82] provides a principled approach to effective management of wide area information systems, such as the Nebula File System and Interface [@BDBCP94]. This not only offers evidence to support the assertion [@MiDo94] that a digital library is a bounded collection of incommensurate information sources in a logical space, but also sheds light on techniques for collaboration through coordinated access to the shared organization of knowledge [@Fu94].'
author:
- |
[^1]\
University of Arkansas\
[email protected]\
- |
[**C. Mic Bowman**]{}[^2]\
Transarc Corporation\
[email protected]\
bibliography:
- 'dl95.bib'
title: '[**Digital Libraries, Conceptual Knowledge Systems, and the Nebula Interface**]{}'
---
Introduction {#introduction .unnumbered}
============
In their lead-off paper for last year’s [*Digital Library’ 94*]{} conference [@MiDo94], Francis Miksa and Philip Doty ask the question “Why should a digital library be called a ‘library’?” They examine three aspects of the traditional library which may reveal the meaning of a digital library. These three aspects were highlighted in their statement (our emphases) that
----------------------------------------------------------------------------------
“a library is a [*collection*]{} of [*information sources*]{} in a [*place*]{}.”
----------------------------------------------------------------------------------
- A [*collection*]{} consists of objects gathered and assembled together with boundaries based upon pragmatic considerations and purpose.
- [*Information sources*]{} are separate and unique intellectual and artistic entities, which inform the user by highly incommensurate methods, purposes, and intensions.
- A [*place*]{} is an intellectual construct, a logical or intellectual space, with location of place meaning “a rationalized set of relationships” which structure the members of a collection.
In the announcement for this year’s [*Digital Library’ 95*]{} conference [@Fu94], Richard Furuta directs our attention to a further collaborative aspect of digital libraries, which was highlighted in his statement (our emphases) that
---------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------
“scholarly work in the digital library of the future will be mediated through coordinated access to [*shared information spaces*]{}. Patrons will [*organize*]{} their own private digital libraries, [*collaborate*]{} with colleagues through shared digital libraries, and have access to huge amounts of multimedia information in global, public digital libraries”.
---------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------
In this paper we describe a mathematical framework, Concept Analysis and conceptual knowledge systems, and an implementation prototype in the areas of resource discovery and wide area information management, the Nebula File System and Interface, which potentially satisfy all four of these criteria.
- The bounded collection is represented by the notions of [*many-valued context*]{} and [*formal context*]{} in Concept Analysis or [*context*]{} in Nebula.
- The information source is represented by an [*object*]{} in Concept Analysis or a [*file object*]{} in Nebula. The incommensurability of information sources is represented by the [*typing*]{} of synoptic information which has been abstracted from objects, and afterward is associated with them.
- The place, or logical space, is represented by the notion of a [*concept lattice*]{} in Concept Analysis or a [*view taxonomy*]{} in Nebula.
- Collaboration through coordinated access to the organization of shared information spaces, is provided in Nebula by the specification of connectivity between information spaces via scoping, and is represented in Concept Analysis in terms of elaboration of the notion of a [*conceptual knowledge system*]{} as a constrained sum of formal contexts.
Table \[aspects\] summarizes the analogies between these four aspects of Digital Libraries and various notions in Concept Analysis and the Nebula Interface.
Since libraries function as shared repositories of information, their contents must be classified and catalogued for easy access by patrons. Classification is a common technique for organizing information [@Ro87]. A taxonomy or classification scheme is an orderly arrangement of items into classes according to shared characteristics [@WyTa92]. There is a strong and useful analogy displayed in Table \[ls:rd:analogies\] between classification in Library Science and classification in Network Information Discovery and Retrieval (NIDR) systems such as Nebula. The Nebula File System [@BDBCP94] and Interface [@Po94] is a prototype wide area information system [@BoSpSp94], which offers a new model for the organization and coordinated access of information spaces. The Nebula system integrates information management with traditional file system operations. In Nebula, resource meta-information is classified by collocation in views. Views are conceptual classes which organize resource meta-information into dynamically customizable information spaces.
Concept Analysis [@Wi82] provides a principled approach to the effective management of wide area information systems such as Nebula. It is a new approach to the analysis of information, which provides the mathematical foundation for faceted, synthetic classification. This is the appropriate model for the organization of knowledge in dynamic view-oriented NIDR systems. Conceptual scaling [@GaWi89] provides the mathematical foundation for faceted analysis via the user’s view and interpretation of information. A flexible dynamic organizing/browsing mechanism, based on ideas from conceptual scaling, is important as publishing moves from a “push” model, where an editor determines what readers see, to a “pull” model, where users decide for themselves what to read by selecting from a variety of information sources. Conceptual knowledge systems [@Wi92] provide an adequate theory of knowledge consisting of: knowledge representation, knowledge inferencing, knowledge acquisition, and knowledge communication tools. Shared access to information spaces, as specified by Nebula scoping, can be formally represent by conceptual knowledge systems.
The paper is organized as follows. Section \[section:abstract\] discusses the nature and benefits of the data abstraction known as meta-information. It is also concerned with the interpretation of such data by faceted analysis. Interpreted meta-information is modeled by formal contexts in Concept Analysis. Such contexts correspond to the idea that digital libraries are “collections” with boundaries. In Section \[section:organize\] the organization of knowledge through synthetic classification is described in terms of the notion of conceptual knowledge systems from Concept Analysis. An instance of conceptual knowledge systems called view taxonomy is constructed in Nebula by the specification of views as descriptive names within scoped indexes. The conceptual space of such organized knowledge corresponds to the notion of digital libraries as collections of information resources “in a place.” Section \[section:manage\] is concerned with the management of information. It discusses the twin paradigms of searching and browsing, and introduces the idea of concept neighborhood navigation. Section \[section:share\] indicates how collaboration between digital libraries, small local conceptual spaces (private or group) and large global conceptual spaces, can be defined via shared organization. Finally, Section \[section:future\] summarizes what we have accomplished in this paper and briefly describes some future work.
Abstracting Meta-Information {#section:abstract}
============================
In the universe of all knowledge, both verbal knowledge and the non-verbal knowledge in music and art, there is a certain amount of summarization or synoptic knowledge that has been recorded [@WyTa92]. This knowledge is called meta-information, and is referred to as the bibliographic universe in Library Science. This meta-information often consists of a set of tag/value attributes or elements. The Nebula File System abstracts from files various elements of meta-information such as filename, size, date created, owner, etc. Nebula refers to such meta-information as a [*file object*]{}. A file object consists of a collection of attributes which the represented file possesses or has. Nebula file objects are gathered together into contexts. A Nebula context stores not only file objects, but also views and a collection of binary index relations that connect attribute values to file objects. There is one index relation for each attribute tag. Table \[document:mvcxt\] displays a Nebula context of documents (see Figure 2 in [@BDBCP94]).
Nebula contexts are instances of the Concept Analysis notion of a many-valued context. The unanalyzed and uninterpreted data of many application domains can often be conceptualized as a constrained collection of many-valued contexts. A [*many-valued context*]{} [@GaWi89] is a quadruple ${\mbox{$ \langle G,N,D,\phi \rangle $}}$, where: $G$ is a set of objects, which models the set of file objects and views in Nebula; $N$ is a set of sorts, which model attribute-tags in Nebula (and database or entity/relationship attributes); $D = \{ D_a \mid a {\,{\in}\,}N \}$ is an $N$-indexed collection of domains of values, corresponding to the attribute-values in Nebula; and $\phi = \{ G \stackrel{\phi_a}{\rightarrow} D_a \mid a {\,{\in}\,}N \}$ is an $N$-indexed collection of functions, called an [*information function*]{}, whose inverses correspond to the index relations in a Nebula context. The Nebula context of documents in Table \[document:mvcxt\] is a many-valued context. The set of sorts is $N = \{ \mbox{project}, \mbox{format} \}$, the set of objects is $G = \{ \mbox{plan1.ps}, \mbox{plan2.ps}, \mbox{plan2.doc}, \mbox{notes1.txt}, \mbox{notes2.txt} \}$, the domains are $D_{\rm project} = \{ \mbox{plan1}, \mbox{plan2} \}$ and $D_{\rm format} = \{ \mbox{postscript}, \mbox{text} \}$, and the information function $\phi_{\rm project}(\mbox{plan1.ps}) = \mbox{plan1}$, $\phi_{\rm project}(\mbox{notes1.txt}) = \mbox{plan2}$, etc.
The [**has**]{} relationship between Nebula file objects and attributes is an instance of the binary incidence matrix of a formal context. A [*(formal) context*]{} is a triple ${\mbox{$ \langle G,M,I \rangle $}}$ consisting of two sets $G$ and $M$ and a binary [*incidence*]{} relation $I \subseteq { {G} {\times} {M} }$ between $G$ and $M$. Intuitively, the elements of $G$ are thought of as [*entities*]{} or [*objects*]{}, the elements of $M$ are thought of as [*properties*]{}, [*characteristics*]{} or [*attributes*]{} that the entities might have, and $g{I}m$ asserts that “object $g$ [*has*]{} attribute $m$.” Table \[document:cxt\] displays a formal context obtained by nominal conceptual scaling of the attributes from the multi-valued context displayed in Table \[document:mvcxt\].
[c@c@c]{}
[|l@l|]{}\
1 & plan1.ps\
2 & plan2.ps\
3 & plan2.doc\
4 & notes1.txt\
5 & notes2.txt\
&
[|l|c@c@c@c|]{}\
& 1 & 2 & 3 & 4\
1 &$\times$&&$\times$&\
2 &&$\times$&$\times$&\
3 &&$\times$&&\
4 &&$\times$&&$\times$\
5 &&$\times$&&$\times$\
&
[|l@l|]{}\
1 & project=plan1\
2 & project=plan2\
3 & format=postscript\
4 & format=text\
Organizing Conceptual Space {#section:organize}
===========================
Nebula identifies resources through descriptive names [@Po94; @BoSpSp94]. A [*descriptive name*]{} is an expression which selects objects based upon their registered attributes. A basic form of descriptive name is just a conjunction of attributes, such as
`format=text & project=plan2 & name=notes2.txt`.
Descriptive names provide file name resolution through [*associative access*]{}. They are particularly important for finding resources in very large information spaces when the resource location is unknown. A set of functions for resolving descriptive names exists in each Nebula context. These include attribute domain operations such as equality and order, and regular expression matching.
Nebula replaces the directories in traditional file systems with database views. The directory structure of traditional file systems is a static organization, which “reflects the requirements of the system designer, not the varied requirements of a diverse user-base” [@BDBCP94]. Nebula views provide a powerful and flexible mechanism for the logical and dynamic organization of an information space, which allows user customization. In Nebula a [*view*]{} is a query which by using a descriptive name selects file objects from within a scope index. Views relate file objects by containment and scope. Containment defines an abstraction relationship over the file objects contain in a view. The properties shared by those objects are abstracted in the descriptive name in the [containment]{} attribute of the view. The view serves as the conceptual class for the objects it contains. Scope defines a generalization-specialization relationship between the conceptual classes denoted by the views. By using views the Nebula File System provides a classification scheme which is both synthetic and faceted: synthetic, because it generates conceptual categorizations “on the fly”; faceted, because it uses atomic units of information known as facets to accomplish this categorization.
Nebula views are instances of the notion of a formal concept in Concept Analysis, used within the confines of a conceptual knowledge system. A [*formal concept*]{} or [*conceptual class*]{} consists of any group of entities or objects exhibiting one or more common characteristics, traits or attributes. Conceptual classes are logically characterized by their extension and intension. The [*extension*]{} of a class is the aggregate of entities or objects which it includes or denotes. The [*intension*]{} of a class is the sum of its unique characteristics, traits or attributes, which, taken together, imply the concept signified by the conceptual class. The process of subordination of conceptual classes and collocation of objects exhibits a natural order, proceeding top-down from the more general classes with larger extension and smaller intension to the more specialized classes with smaller extension and larger intension. This [**isa**]{} relationship is a partial order called generalization-specialization. Conceptual classes with this generalization-specialization ordering form a class hierarchy for the context, which mathematically is a complete lattice, Figure \[document:lat\] displays the lattice of conceptual classes associated with the formal context of documents displayed in Table \[document:cxt\].
(500,100) (180,0)
(100,100) [(0,0)]{} [(50,100)]{} [(0,75)]{} [(5,80)]{} (0,75)[(2,1)[50]{}]{} [(100,75)]{} [(105,80)]{} [(105,69)]{} (100,75)[(-2,1)[50]{}]{} [(50,50)]{} [(55,44)]{} (50,50)[(-2,1)[50]{}]{} (50,50)[(2,1)[50]{}]{} [(0,25)]{} [(5,30)]{} [(5,19)]{} (0,25)[(0,1)[50]{}]{} [(100,25)]{} [(105,30)]{} [(105,19)]{} [(105,11)]{} (100,25)[(0,1)[50]{}]{} [(50,0)]{} (50,0)[(-2,1)[50]{}]{} (50,0)[(0,1)[50]{}]{} (50,0)[(2,1)[50]{}]{}
According to Concept Analysis, in addition to modeling knowledge representation in file systems, with the notion of a [*conceptual knowledge system*]{} [@Wi92] we will be able to do knowledge inferencing, knowledge acquisition, and knowledge communication. There are 3 basic notions in conceptual knowledge systems: objects, attributes, and concepts. In Nebula, these correspond to: file objects, attributes, and views, respectively. There are 4 basic relationships in conceptual knowledge systems: an object has an attribute, an object belongs to a concept, an attribute abstracts from a concept, and a concept is a subconcept of another concept. In Nebula, these correspond to: has, containment, constructor, and scope relationships, respectively. These notions and relationships partition the frame of a conceptual knowledge system as in Table \[fca:relationships\].
Table \[document:cks:cxt\] represents a conceptual knowledge system within the conceptual universe ${\cal D}$ of all documents in an information system and their properties (see Figure 2 in [@BDBCP94]). The conceptual knowledge system in Table \[document:cks:cxt\] extends the formal context of documents represented by Table \[document:cxt\]. It consists of a set $B$ of five concepts of ${\cal D}$ defined as views, in addition to the set of objects and attributes in Table \[document:cxt\]. For convenience of illustration, we have added an additional object “notes0.txt”. The crosses in Table \[document:cks:cxt\] represent four relations (Boolean matrices): the organization submatrix is the order relation on the 5 classes, and the having submatrix is identical to the incidence matrix in Table \[document:cxt\]. Figure \[document:cks:lat\] represents the lattice of conceptual classes for the conceptual knowledge system displayed in Table \[document:cks:cxt\]. The set of concepts $B = \{ {\rm objs}, {\rm docs}, {\rm postscript}, {\rm plan1}, {\rm plan2} \}$, which were added to the document space to form the conceptual knowledge system, are each represent by a node at the top of the line diagram in Figure \[document:cks:lat\].
[c@c@c]{}
[|r@l|]{}\
1 & Object\
2 & Document\
3 & PostScript\
4 & Plan1\
5 & Plan2\
6 & plan1.ps\
7 & plan2.ps\
8 & plan2.doc\
9 & notes0.txt\
9 & notes1.txt\
10 & notes2.txt\
&
[|r|c@c@c@c@ c@|c@c@c@c|]{}\
& 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9\
1 &$\times$&&&& &&&&\
2 &$\times$&$\times$&&& &&&&\
3 &$\times$&$\times$&$\times$&& &&&$\times$&\
4 &$\times$&$\times$&&$\times$& &$\times$&&&\
5 &$\times$&$\times$&&&$\times$ &&$\times$&&\
6 &$\times$&$\times$&$\times$&$\times$& &$\times$&&$\times$&\
7 &$\times$&$\times$&$\times$&&$\times$ &&$\times$&$\times$&\
8 &$\times$&$\times$&&&$\times$ &&$\times$&&\
9 &$\times$&$\times$&&$\times$&&$\times$&&&$\times$\
10 &$\times$&$\times$&&&$\times$&&$\times$&&$\times$\
11 &$\times$&$\times$&&&$\times$&&$\times$&&$\times$\
&
[|r@l|]{}\
1 & Object\
2 & Document\
3 & PostScript\
4 & Plan1\
5 & Plan2\
6 & project=plan1\
7 & project=plan2\
8 & format=postscript\
9 & format=text\
(300,175) (50,0)
(200,175) [(0,0)]{} [(100,175)]{} [(105,171)]{} [(100,150)]{} [(105,146)]{} (100,150)[(0,1)[25]{}]{} [(25,100)]{} [(0,108)]{} [(30,96)]{} (25,100)[(3,2)[75]{}]{} [(75,100)]{} [(70,108)]{} [(80,96)]{} [(80,85)]{} (75,100)[(1,2)[25]{}]{} [(125,100)]{} [(125,108)]{} [(130,96)]{} (125,100)[(-1,2)[25]{}]{} [(175,100)]{} [(180,108)]{} (175,100)[(-3,2)[75]{}]{} [(0,50)]{} [(5,43)]{} (0,50)[(1,2)[25]{}]{} (0,50)[(3,2)[75]{}]{} [(50,50)]{} [(55,43)]{} (50,50)[(-1,2)[25]{}]{} (50,50)[(3,2)[75]{}]{} [(150,50)]{} [(155,42)]{} [(155,36)]{} (150,50)[(-3,2)[75]{}]{} (150,50)[(1,2)[25]{}]{} [(200,50)]{} [(205,43)]{} (200,50)[(-3,2)[75]{}]{} (200,50)[(-1,2)[25]{}]{} [(100,0)]{} (100,0)[(-2,1)[100]{}]{} (100,0)[(-1,1)[50]{}]{} (100,0)[(1,1)[50]{}]{} (100,0)[(2,1)[100]{}]{}
Managing Information {#section:manage}
====================
In Library Science only the bibliographic universe can be controlled. Such control is performed by means of bibliographic tools. Catalogs are bibliographic tools, which exercise bibliographic control through three basic functions: identification, collocation, and evaluation. A user, who has a citation or bibliographic item in mind, should be able to match or [*identify*]{} or find a bibliographic entry for that item. This is the searching paradigm. A user, based upon various bibliographic data and connecting references, should be able to bring together in one place or [*collocate*]{} bibliographic entries for similar and closely related material. This is part of the browsing paradigm. A user should be able to choose by [*evaluation*]{} from among many bibliographic entries the one that best represents the knowledge, information or specific item desired. This also is part of the browsing paradigm.
Wide area information systems provide access to networked information resources through many different application interfaces: WWW hypertext, Gopher menues, and Archie search. File management systems, such as Archie, Gopher, Prospero, WAIS, and WWW, use an existing file system as a storage repository. Resource discovery systems, such as Harvest and Whois$+$$+$, effectively manipulate the tremendous amount of heterogeneous information in wide area information systems and wide area file systems, by using the uniform, logical interface provided by the resource meta-information in bibliographic records: Harvest defines its Summary Object Interchange Format (SOIF) and Whois$+$$+$ can use the resource description called the Uniform Resource Characteristic (URC) currently being developed by the Internet Engineering Task Force (IETF) working group on uniform resource identifiers. Just as the ISBD in Library Science, this meta-information is typed in order to represent the heterogeneity of networked information resources. Resource discovery systems use two paradigms for managing information: searching and organizing/browsing.
--------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------
[*Searching*]{} is the process of locating resources. The user provides a description or query of the resources being sought. The resource discovery system resolves the query and returns to the user a list of resources which match the description. Archie is the canonical example of an information system based upon the search paradigm. To search for a file the user must possess enough information about the file to formulate a query. Searching is most effective when a user can formulate a precise query from attributes that are indexed for efficient lookup. The main weakness of searching is that formation of good queries can be a difficult task, especially in an information space unfamiliar to the user. [*Organizing*]{} is the human-guided process (on the server side) of deciding how to interrelate information, usually by placing it into some sort of hierarchy (for example, the hierarchy of direcetories in an FTP file system). [*Browsing*]{} is the corresponding human-guided process (on the client side) of exploring the organization and contents of a resource space. The main weakness of organizing/browsing is that (1) it is done by someone else, (2) it is not easy to change (for example, in Library Science the standard classification systems of Dewey and the Library of Congress are fixed and possibly not relevant to the present-day patron), and (3) it is difficult to keep a large amount of data “well organized”. For effective browsing the system does not resolve queries, but instead it organizes the information space so that the user can navigate it easily. The key to effective browsing is a well-organized, flexible, dynamic information space. Classification is a common technique for organizing information. A taxonomy or classification scheme is an orderly arrangement of terms or classes. Such a scheme arranges a set of objects into classes with shared characteristics.
--------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------
Concept Analysis defines a navigation method called [*concept neighborhood navigation*]{}, which moves between local conceptual neighborhoods. This allows for an interactive exploration of information spaces, individual or shared. A subset of facets can be chosen and starting from these, a local environment of related items can be explored. When browsing a very large data repository, it can be desirable to select a set of starting objects and to interactively explore their neighborhood. This process consists of the following steps:
Initialization
: The facets are evaluated with respect to the uninterpreted data and transformed into binary relations (formal contexts). The evaluated facets are composed into a single formal context using the operation of [*apposition*]{} (this operation requires the contexts to share a common object set). A global analysis is performed on the total context, chiefly in terms of the collection of local neighborhood lattices.
Browse Loop
: A new seed is chosen. This may be either an object or an attribute. The local neighborhood of a given seed object is analyzed and previewed. To simplify the visualization data to be presented to the user (and possibly reach an acceptable number of conceptual classes), the local neighborhood is modified using various means: raising the connectivity threshold, rank-ordering the attributes and restricting to the most important ones, restricting to a ball around the seed induced by a similarity metric, etc. The local neighborhood is visualized. At this time the user may want to visualize the union context of the local neighborhoods for the old seed and the new seed — this allows comparison of “distance” moved and classes in common.
Sharing Organization {#section:share}
====================
Scoping in Nebula can provide coordinated access and efficient sharing of the organization of separate, sharable information spaces [@Po94]. Comparison of the private individual information spaces of experts in the context of psychoanalysis has been discussed in Concept Analysis [@SWFL94]. The mathematical foundations, for collaboration by coordinated access to the knowledge organized in logical spaces, is defined in terms of constrained sums of formal contexts [@Ke92] applied to conceptual knowledge systems. The sharing of knowledge organization can be accomplished with the specification of connectivity between logical information spaces. Sharing organization between two logical spaces is visualize in Table \[sharing\] in terms of elaborated conceptual knowledge systems. The first logical space makes use of the organization of the second logical space by specifying the link connectivity $\mbox{sharing}_{1,2}$. These links represent scoping the first logical space conceptual classes on appropriate classes in the second logical space. So with links in $\mbox{sharing}_{1,2}$ the second logical space shares its organization with the first logical space. The linked connectivity in $\mbox{sharing}_{2,1}$ represent the dual situation — the sharing of the first by the second. The cross-linked instantiation in blocks $\mbox{instantiation}_{1,2}$ and $\mbox{instantiation}_{2,1}$ is derived connectivity defined by incidence matrix closure.
incidence $\mbox{classes}_1$ $\mbox{classes}_2$ attributes
-------------------- ------------------------------ ------------------------------ ---------------------------
$\mbox{classes}_1$ $\mbox{organization}_1$ $\mbox{sharing}_{1,2}$ $\mbox{distinguishing}_1$
$\mbox{classes}_2$ $\mbox{sharing}_{2,1}$ $\mbox{organization}_2$ $\mbox{distinguishing}_2$
$\mbox{objects}_1$ $\mbox{instantiation}_1$ $\mbox{instantiation}_{1,2}$ $\mbox{having}_1$
$\mbox{objects}_2$ $\mbox{instantiation}_{2,1}$ $\mbox{instantiation}_2$ $\mbox{having}_2$
: [**Sharing Organization**]{}[]{data-label="sharing"}
As Francis Miksa and Philip Doty have pointed out [@MiDo94], the Internet Gopher space today is not a digital library — it contains a huge variety of useful sources, but these are “not tied together as a single intellectual construct, neither in the sense of structure nor in the sense of access methods.” We contend that such legacy information can be augmented and organized by coordinated access into a collection of collaborating digital libraries. The following example, which is part of the LC Marvel Gopher space at the Library of Congress, describes how this can be done.
The Library of Congress Machine-Assisted Realization of the Virtual Electronic Library (LC Marvel located at `gopher://marvel.loc.gov/`) classifies government publications. The conceptual space defined includes government publications as objects and abstracted content as attributes. Table \[tab:govpub\] shows attributes registered for three documents in the conceptual space defined by Marvel. The conceptual space is structured according to government branch, office, and project. Table \[tab:govclass\] shows several document classes that exist. When implemented in Nebula, each publication is a file object with attributes registered for relevant properties. In practice, these attributes are collected by textual summarizers that process the typical structure of government documents such as press releases, newsletters, and BAA’s. A special “class” attribute is registered for a document that corresponds to the path of Gopher menus traversed to retrieve the document. The “class” attribute provides some additional information about the content of the document.
Nebula represents each class as a view that contains all objects in the corresponding menu or in any submenu. That is, the “Executive” view contains all publications for the White House, the Department of Agriculture, and any other Executive branch department or committee. In addition to the conceptual space of government publications, each user defines a “reader” space that augments the structure of the conceptual space defined by Marvel. The user space represents a profile of information the user finds interesting. For example, a user may consider interesting the class of documents regarding nuclear waste disposal. This class of documents crosses the boundaries of classes defined by Marvel. In fact it might include publications by the Judicial branch, executive orders and press releases from the White House and the Department of Energy, and legislative actions from Congress.
Intuitively, the reason for constructing user classes is that formal, organizational classification of government publications is too general. While the general classification is useful for browsing the collection of documents, it is not sufficient for issue-specific location of publications. Nebula accommodates user classes by layering them on top of the more general structure. The formal structure is available globally. In a separate user space, an issue-specific view is constructed by scoping it on the appropriate organizational classes. In this way, the formal structure provides a reasonable first approximation on which users can construct a customized structure.
Conclusions & Future Work {#section:future}
=========================
In this paper we have demonstrated in a very real sense how the Nebula prototype and Concept Analysis articulate the idea of digital libraries as bounded collections of typed information sources in conceptual spaces with collaboration defined by coordinated access to the shared organization of knowledge.
In particular, we have shown how Concept Analysis provides a principled foundation for the Nebula Interface, by giving it a mathematically rigorous base, which fits well with the intuitions behind Nebula. For example, the notion of a concept lattice provides an explicit mathematical structure for Nebula view taxonomies. In addition, Concept Analysis reveals the composite nature of view specification. The first step is the explicit specification of the lattice join of the superordinate views in the [scope]{} component. The second step is the specialization of the superordinate join view via the conceptually scaled attributes in the [constructor]{} component.
The Concept Analysis foundation should allow us to apply the same extensions to Nebula which have been given to Concept Analysis: the Fuzzy Logic extension of Concept Analysis [@Ke94b] will allow us to extend Nebula to fuzzy taxonomies of views; and the Rough Set extension of Concept Analysis [@Ke93b] will allow us to define rough Nebula views. Concept Analysis strongly suggests the use of faceted analysis via conceptual scaling, along with the current awareness ideas of SDI and continuous queries, to serve as a basis for the descriptive name component in the specification of views. It suggests the importance of “implications” and expert system type rules as an augmented means for analyzing taxonomic structures in Nebula.
A formal basis for describing views in terms of objects and a more expressive environment for analyzing the relationships between views may make it possible to simplify a collection of views, thus providing query/space optimization. In the simple case, two views could be collapsed into one if they always contains the same set of objects. This is a straightforward application of the Concept Analysis optimization technique of “purification”. Also applicable is the optimization technique of “reduction”, which eliminates objects and attributes which are not irreducible.
More generally, a “cluster” of similar (but not identical) views might be abstracted into a more general view that many would find interesting. In Concept Analysis this kind of clustering has been captured by the notion of a [*rough concept*]{} in the Rough Set extension, which is based upon an indiscernibility relation on objects (which could dually be on attributes). It can also be realized by the notion of a generalized metric over conceptual classes (views), which would cluster views by a tolerance setting (a ball around a view of a certain small radius).
Traditional browsing systems, such as Gopher and the World Wide Web, define browsing transitions along the physical links of the knowledge organization, whether these are hierarchical as in Gopher-space or cross-referential as in the Web. Concept Analysis offers a more flexible and logical alternative. By introducing the idea of a concept neighborhood, it defines browsing transitions along conceptual links of the knowledge organization. The knowledge organization here consists of the whole information space, which would normally be only virtually represented. One travels from the local concept neighborhood of one object to the local concept neighborhood of a neighboring object. The transition can be pictured by the union neighborhood.
Future work will include: the incorporation into Nebula of Fuzzy Logic and Rough Set extensions; application of conceptual scaling techniques from Concept Analysis to provide for a rigorous foundation and elaboration of the associative access in Nebula view specification; realization of a testbed for collaborative studies; and development of client software for analysis and browsing by concept neighborhood navigation.
[^1]: This research was funded by a grant from [**Intel**]{} Corporation.
[^2]: This research was funded by the Advanced Research Project Agency, under contract number F19628-93-C-1076.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
In this paper we describe the structure of a group of conjugating automorphisms $C_n$ and prove that this structure is similar to the structure of a braid group $B_n$ with $n\geq 2$ strings. We find the linear representation of group $C_n$. Also we prove that the braid group $B_n(S^2)$ of 2–sphere, mapping class group $M(0,n)$ of the $n$–punctured 2–sphere and the braid group $B_3(P^2)$ of the projective plane are linear. Using result of J. Dyer, E. Formanek, E. Grossman and the faithful linear representation of Lawrence–Krammer of $B_4$ we construct faithful linear representation of the automorphism group ${\rm Aut}(F_2)$.\
[*Mathematics Subject Classification:*]{} 20F28, 20F36, 20G35\
[*Key words and phrases:*]{} free group, conjugating automorphism, automorphism group, braid group, braid group of manifold, linear representations, width of verbal subgroups.
author:
- Valerij Bardakov
title: ' The structure of the group of conjugating automorphisms and the linear representation of the braid groups of some manifolds [^1]'
---
Introduction
============
It is well known \[1,2\] that the automorphism group ${\rm Aut}(F_2)$ of a two generated free group $F_2$ can be constructed from cyclic groups using free and semi direct products. The following problem is open: is it true for ${\rm Aut}(F_n)$ for $n>2$. Denote $F'_n=[F_n, F_n]$ (the commutator subgroup of $F_n$) and $A_n=F_n/F_n'$. Now $A_n$ is a free abelian group and we have ${\rm Aut}(A_n)\simeq {\rm GL}_n(\mathbb{Z}).$ The natural map from ${\rm Aut}(F_n)$ into ${\rm Aut}(A_n)$ is an epimorphism. The kernel of this map is called the group of $IA$–[*automorphisms*]{} and denoted as ${\rm IA}(F_n)$. The group ${\rm IA}(F_2)$ is the group of inner automorpisms ${\rm Inn}(F_2)$ which is isomorphic to $F_2$, but ${\rm Inn}(F_n)$ is a proper subgroup of ${\rm
IA}(F_n)$ if $F_n$ is free of rank greater than 2.
J. Nielsen showed for $n\leq 3$ and W. Magnus showed for all $n$ (see \[3, chapter 1, § 4\]), that ${\rm IA}(F_n)$ is generated by the automorphisms $$\varepsilon_{ijk} : \left\{
\begin{array}{ll}
x_{i} \longmapsto x_{i}[x_j, x_k] & \mbox{for } k\neq i, j, \\
x_{l} \longmapsto x_{l} & \mbox{for } l\neq i,
\end{array} \right. ~~~~~
\varepsilon_{ij} : \left\{
\begin{array}{ll}
x_{i} \longmapsto x_{j}^{-1}x_ix_j & \mbox{for } i\neq j, \\
x_{l} \longmapsto x_{l} & \mbox{for } l\neq i,
\end{array} \right.$$ where $[a, b]=a^{-1}b^{-1}ab$ is a commutator $a$ and $b$. J. Nielsen showed that ${\rm IA}(F_n)$ is the normal closure in ${\rm
Aut}(F_n)$ of $\varepsilon_{12}.$ The defining relation for ${\rm
IA}(F_n)$, $n\geq 3,$ is not known but S. Krstic and J. McCool \[4\] proved that ${\rm IA}(F_3)$ is not finite presented.
Denote subgroup of ${\rm IA}(F_n)$ which is generated by $\varepsilon_{ij}$, $1\leq i\neq j\leq n$, by $Cb_n$. Group $Cb_n$ is called [*group of basis–conjugating automorphisms*]{}. J. McCool \[5\] proved that this group is finitely presented and found defining relations (see below).
Group $Cb_n$ is subgroup of group conjugating automorphisms $C_n$, where automorphism from ${\rm Aut}(F_n)$ is called [*conjugating automorphism*]{} if it maps each generator $x_i$ to $f_i^{-1}x_{\pi (i)}f_i$, where $f_i \in F_n$ and $\pi $ is a permutation from symmetric group $S_n$. It is clear that if $\pi $ is identical permutation then this conjugating automorphism lies in $Cb_n$. Automorphisms from $C_n$ which fix product $x_1x_2...x_n$ form the braid group $B_n$ on $n$ strings. The braid group $B_n$ has a normal subgroup $P_n$ of finite index. Subgroup $P_n$ is called [*pure braid group*]{} and quotient $B_n/P_n$ is the symmetric group $S_n$. It is clear that $P_n$ is subgroup of $Cb_n$. The structure of $P_n$ is well known \[6, 7\]. It is the semi direct product of free groups: $$P_n=U_n\leftthreetimes (U_{n-1}\leftthreetimes (\ldots \leftthreetimes
(U_3\leftthreetimes U_2))\ldots),$$ where $U_i\simeq F_{i-1}, i=2,3,\ldots,n.$ The following natural question arises: is it true that group of basis–conjugating automorphisms $Cb_n$ is a semi direct product of some group? In this paper we give positive answer on this question. It will be proved the following theorem
The group of basis–conjugating automorphisms $Cb_n$, $n\geq 2,$ is a semi direct product $$Cb_n=D_{n-1}\leftthreetimes (D_{n-2}\leftthreetimes (\ldots
\leftthreetimes (D_2\leftthreetimes D_1))\ldots ),$$ where $D_i=<\varepsilon_{i+1,1},\varepsilon_{i+1,2},
\ldots,\varepsilon_{i+1,i},
\varepsilon_{1,i+1},\varepsilon_{2,i+1},\ldots,\varepsilon_{i,i+1}>$, $i=1,2,\ldots,n-1$. Elements $\varepsilon_{i+1,1},\varepsilon_{i+1,2},
\ldots,\varepsilon_{i+1,i}$ generate a free group of rank $i$ and elements $\varepsilon_{1,i+1},\varepsilon_{2,i+1},\ldots,\varepsilon_{i,i+1}$ generate a free abelian group of rank $i$. This decomposition is consistent with the decomposition of $P_n$, i. e., there is inclusion $U_{i+1}\leq D_i,$ $i=1,2,\ldots,n-1.$
Remark that analogous theorem was proved by D. J. Collins and N. D. Gilbert \[8\] for group which is the kernel of homomorphism ${\rm Aut}(G) \longrightarrow
{\rm Aut}(\overline{G})$, where $G = G_1 * G_2 * \ldots * G_n$ is a free product, group $\overline{G} = G_1 \times G_2 \times \ldots \times G_n$ is a direct product and each factor $G_i$ is indecomposible and is not isomorphic to infinite cyclic group.
As a corollary from this theorem we will find the some normal form for words which is presented elements from $C_n$. We will prove some properties of $C_n$ and $Cb_n$. We will prove that for $n\geq 2$ the group $C_n$ has no more than 4 generators and we will find defining relations. For $n\geq 2$ we will prove that $Cb_n$ has not proper verbal subgroups of finite width. This result is closely related to the question 14.15 from “Kourovka Notebook” \[9\].
Also one can ask the following question:
For which $n$ the group generated by $\varepsilon_{ijk}$, $1\leq k\neq i,j\leq n,$ is finitely defined?
One of the most intriguing problems in the theory of braid groups is the question of whether $B_n$ is linear, i. e., whether it admits a faithful representation into a group of matrices over field. This question was formulated by W. Burau in 1936. W. Burau \[10\] found an $n$–dimensional linear representation of $B_n$ which for a long time had been considered as a candidate for faithful representation. However, as it was established by J. A. Moody \[11\] in 1991 this representation is not faithful for $n\geq 9$. This bound was improved for $n\geq 6$ by D. Long and M. Paton \[12\] and to $n\geq 5$ by S. Bigelow \[13\]. To this date, it is unknown whether the Burau representation of $B_4$ is faithful.
In 1990 R. J. Lawrence \[14\] introduced a family of linear representations of $B_n$. Later D. Krammer \[15\] and S. Bigelow \[16\] proved that one of this representation is faithful. Therefore $B_n$ are linear for all $n\geq
2$.
As it was mention above, $B_n$ is subgroup of ${\rm Aut}(F_n).$ But E. Formanek and C. Procesi \[17\] proved that ${\rm Aut}(F_n)$ is not linear for all $n\geq 3$. So the following question arises:
Find the maximal related inclusion subgroup of ${\rm Aut}(F_n)$, $n\geq 3$, which contains $B_n$ and is linear. In particular, is it true that group $C_n$ is linear?
The second part of this problem was formulated in \[9, Problem 15.9\]. The one natural way for answer on this question is to extend the known representations of $B_n$ to a representation of $C_n$.
In this paper we define an extension of the Burau representation on $C_n$ and prove that the linear representation of Lawrence–Krammer extending on $C_3$ and for $n\geq 4$ we define the extension of this representation for some restrictions on parameters of representation.
Braid group $B_n$ is a particular case of general construction braid group $B_n(M)$ of manifold $M$ on $n$ strings.
For which manifold $M$ and natural $n$ the braid group $B_n(M)$ on $n$ strings is linear?
The greatest interest in this problem is connected to manifolds of dimension two. In the braid groups of these manifolds only $B_n(S^2)$ of 2–sphere $S^2$ and $B_n(P^2)$ of projective plane $P^2$ have a torsion.
In this article we will prove that the mapping class group $M(0,n)$ of the $n$–punctured 2–sphere and braid group $B_n(S^2)$ of 2–sphere are linear for all $n\geq 2$. Also we will prove that braid group $B_3(P^2)$ of projective plane is linear (for $n=1, 2$ this group is finite). In all cases we will construct corresponding linear representation. The result that $B_n(S^2)$ is linear was announced in \[18\]. As it was known to the author S. Bigelow and R. D. Budney \[19\] give another proof that groups $B_n(S^2)$ and $M(0,n)$ are linear.
As we noted above group ${\rm Aut}(F_n)$ is not linear for $n\geq
3$. It was proved in \[20\] that ${\rm Aut}(F_2)$ is linear if and only if braid group $B_4$ on 4 strings is linear. In last section we will construct the faithful linear representation of ${\rm Aut}(F_2)$.
I thankful to J. Meier for pointing out on work \[8\] and on some non-corrected statement in the first version of the manuscript.
Definitions and notations
=========================
We remind some known facts which can be found in \[21, 6, 7\].
The group $G$ is a [*semi direct product*]{} of $A$ and $B$ if there exist subgroups $H$ and $K$ in $G$ such that $$G=HK,~~~A\simeq H\unlhd G,~~~B\simeq K,~~~H\cap K=1.$$ Semi direct product is denoted as $G=A\leftthreetimes B$. If $A$ and $B$ have presentations $$A=<a_1, a_2,\ldots,a_k \parallel A_1,
A_2,\ldots,A_p>,~~~
B=<b_1, b_2,\ldots,b_l \parallel B_1, B_2,\ldots,B_q>,$$ then $G=A\leftthreetimes B$ have presentation $$G=<a_1, a_2,\ldots,a_k; b_1, b_2,\ldots,b_l \parallel A_1, A_2,\ldots,A_p;
B_1, B_2,\ldots,B_q; b_i^{-1}a_jb_i=C_{ij},$$ $$~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~1\leq i\leq l,~~~ 1\leq j\leq k>,$$ where $A_1, A_2,\ldots,A_p, C_{ij}$ are words in alphabet $\{ a_1^{\pm 1},a_2^{\pm 1},\ldots,a_k^{\pm 1}\}$, $B_1, B_2,\ldots,B_q$ are words in alphabet $\{ b_1^{\pm 1},b_2^{\pm 1},\ldots,b_l^{\pm 1}\},$ elements $C_{i1}, C_{i2},\ldots,C_{ik}$ generate group $A$ for each $i\in \{ 1,2,\ldots,l \}$ and conjugation by element $b_i$ induces automorphism of group $A$.
Let $M$ be a manifold of dimension $\geq 2$. [*Configuration space*]{} $F_n(M)$, $n\in \mathbb{N}$, for manifold $M$ is a set $$F_n(M)=\{ (z_1, z_2,\ldots,z_n)\in M^n \vert z_i\neq z_j~~
\mbox{for}~~i\neq j \}$$ of ordering collections of $n$ distinct points from $M$. The fundamental group $P_n(M)=\pi_1(F_n(M))$ of the space $F_n(M)$ is the [*pure braid group*]{} with $n$ strings of the manifold $M$. Symmetric group $S_n$ acts on the space $F_n(M)$ permute the coordinats. This action is free and induce the regular covering of orbit space $F_n(M)/S_n$ by $F_n(M)$. The fundamental group $B_n(M)=\pi_1(F_n(M)/S_n)$ is called the [*full braid group*]{} of $M$, or more simply the [*braid group*]{} of $M$. The regular covering projection $F_n(M)\longrightarrow F_n(M)/S_n$ is induce the short exact sequence $$1\longrightarrow P_n(M)\longrightarrow B_n(M)\longrightarrow
S_n\longrightarrow 1.$$ If $M$ is a closed, smooth manifold of dimension $n\geq 2$, then the inclusion map $F_n(M)\longrightarrow M^n$ induces a surjective homomorphism $P_n(M)\longrightarrow \pi_1(M)\times \ldots \times \pi_1(M)$ pure braid group $P_n(M)$ on direct product $n$ copies of fundamental group $\pi_1(M)$. If $\mbox{dim}M>2$ then this homomorphism also injective. The braid groups of manifolds of dimension 2 represent the largest interest.
Among the 2–dimension manifolds special role play the 2–sphere $S^2$ and the projective plane $P^2$ because the braid groups only this manifolds have elements of finite order. If $M$ is a closed surface different from $S^2$ or $P^2$, then in the following sequence of groups $$1\longrightarrow P_n(E^2)\longrightarrow P_n(M)\longrightarrow
\prod^n_{i=1}\pi_1(M)\longrightarrow 1$$ the kernel of each homomorphism is equal to the normal closure of the image of the previous homomorphism in the sequence. In this sequence $E^2$ denotes the Euclidean plane and $\prod^n_{i=1}\pi_1(M)$ denoted the direct product of $n$ copies of $\pi_1(M)$.
The classical braid group of Artin $B_n$ is the braid group $B_n(E^2)$ of Euclidean plane $E^2$. We will call $B_n$ simply braid group.
The braid group $B_n$, $n\geq 2$, with $n$ strings can be defined as the group generated by $n-1$ generators $\sigma_1,\sigma_2,...,\sigma_{n-1}$ with defining relations $$\sigma_i\sigma_{i+1}\sigma_i=\sigma_{i+1}\sigma_i\sigma_{i+1},~~~ i=1,2,...,n-2, \eqno{(1)}$$ $$\sigma_i \sigma_j = \sigma_j \sigma_i,~~~ |i-j|\geq 2. \eqno{(2)}$$ There is a homomorphism $\nu : B_n\longrightarrow S_n$ from the group $B_n$ to the symmetric group $S_n$ on $n$ letters defined by $$\nu(\sigma_i)=(i,i+1),~~~i=1,2,\ldots,n-1.$$ The kernel of homomorphism $\nu $ is the pure braid group $P_n$. The group $P_n$ admits a presentation with generators $$a_{i,i+1}=\sigma_i^2,$$ $$a_{ij}=\sigma_{j-1}\sigma_{j-2}\ldots\sigma_{i+1}\sigma_i^2\sigma_{i+1}^{-1}\ldots
\sigma_{j-2}^{-1}\sigma_{j-1}^{-1},~~~i+1< j \leq n.$$ and defining relations: $$\begin{array}{ll}
a_{ik}^{-\nu }a_{kj}a_{ik}^{\nu }=\left( a_{ij}a_{kj}\right) ^{\nu }a_{kj}
\left( a_{ij}a_{kj}\right) ^{-\nu }, & \\
& \\
a_{km}^{-\nu }a_{kj}a_{km}^{\nu }=\left( a_{kj}a_{mj}\right) ^{\nu }a_{kj}
\left( a_{kj}a_{mj}\right) ^{-\nu }, &~~~m < j,\\
& \\
a_{im}^{-\nu }a_{kj}a_{im}^{\nu }=\left[ a_{ij}^{-\nu }, a_{mj}^{-\nu }\right] ^{\nu }a_{kj}
\left[ a_{ij}^{-\nu }, a_{mj}^{-\nu }\right] ^{-\nu }, & ~~~i < k <
m,\\
& \\
a_{im}^{-\nu }a_{kj}a_{im}^{\nu }=a_{kj}, & ~~~k < i;~~~m < j~~
\mbox{or}~~ m < k,
\end{array}$$ where $\nu=\pm 1$.
The subgroup $U_n$ of $P_n$ which is generated by $a_{1,n},
a_{2,n},\ldots,a_{n-1,n}$ is free and normal in $P_n$. The group $P_n$ is a semi direct product of $U_n$ and $P_{n-1}$. Hence the group $P_n$ is a semi direct product $$P_n=U_n\leftthreetimes (U_{n-1}\leftthreetimes (\ldots \leftthreetimes
(U_3\leftthreetimes U_2))\ldots),$$ where $U_i=<a_{1,i},
a_{2,i},\ldots,a_{i-1,i}>, i=2,3,\ldots,n,$ is a free group of rank $i-1$.
The braid group $B_n$ has a faithful representation as a group of automorphisms of a free group $F_n=<x_1, x_2,\ldots, x_n>$ of rank $n$. The representation is induced by a mapping from $B_n$ to ${\rm Aut}(F_n)$ defined by $$\sigma_{i} : \left\{
\begin{array}{ll}
x_{i} \longmapsto x_{i}x_{i+1}x_i^{-1}, & \\ x_{i+1} \longmapsto
x_{i}, & \\ x_{l} \longmapsto x_{l} & ~~ l\neq i,i+1.
\end{array} \right.$$ The generator $a_{rs}$ of $P_n$ defines the following automorphism $$a_{rs} : \left\{
\begin{array}{ll}
x_{i} \longmapsto x_{i} & \mbox{if }~~ s < i~~ \mbox{or}~~ i < r, \\
& \\
x_{r} \longmapsto x_{r}x_{s}x_rx_s^{-1}x_r^{-1}, & \\
& \\
x_{i} \longmapsto [x_{r}^{-1}, x_s^{-1}]x_i[x_{r}^{-1}, x_s^{-1}]^{-1} & \mbox{if }~~ r < i <s,\\
& \\
x_{s} \longmapsto x_{r}x_sx_r^{-1}. &
\end{array} \right.$$
By the theorem of Artin \[7, Theorem 1.9\] automorphism $\beta \in B_n \subset {\rm Aut}(F_n)$ if and only if $\beta $ satisfies to the two conditions $$1)~~~ \beta(x_i)=f_i^{-1}x_{\pi(i)}f_i,~~~1\leq i\leq n,$$ $$2)~~~ \beta(x_1x_2 \ldots x_n)=x_1x_2 \ldots x_n,$$ where $\pi $ is a permutation of $\{1, 2, \ldots, n \}$ and $f_i=f_i(x_1,x_2,\ldots,x_n)$ is a word in the generators of $F_n$.
The automorphism of $F_n$ calls [*conjugating automorphism*]{} if it satisfies condition 1). Let $C_n$ be a group of conjugating automorphisms. Obvious, that it has the normal subgroup $Cb_n$ with factor group $C_n/Cb_n$ isomorphic to the symmetric group $S_n$. The group $Cb_n$ calls the group of [*basis–conjugating automorphisms*]{}. J. McCool \[5\] proved, that the group $Cb_n$ is generated by automorphisms $$\varepsilon_{ij} : \left\{
\begin{array}{ll}
x_{i} \longmapsto x_{j}^{-1}x_ix_j & i\neq j, \\
x_{l} \longmapsto x_{l} & l\neq i,
\end{array} \right.$$ $i\leq i\neq j \leq n$, with defining relations $$\begin{array}{lr}
\varepsilon_{ij}\varepsilon_{kl}=\varepsilon_{kl}\varepsilon_{ij},
&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ (3) \\
& \\
\varepsilon_{ij}\varepsilon_{kj}=\varepsilon_{kj}\varepsilon_{ij},
&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ (4) \\
& \\
(\varepsilon_{ij}\varepsilon_{kj})\varepsilon_{ik}=\varepsilon_{ik}(\varepsilon_{ij}
\varepsilon_{kj}).
&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ (5) \\
\end{array}$$ where distinct letters denote distinct indexes.
Decomposition of the group of basis–conjugating automorphisms
=============================================================
In this section it will be proved that the structure of $Cb_n$ is similar to the structure of pure braid group $P_n$. The result is the following:
The group of basis–conjugating automorphisms $Cb_n$, $n\geq 2$ is a semi direct product $$Cb_n=D_{n-1}\leftthreetimes (D_{n-2}\leftthreetimes (\ldots
\leftthreetimes (D_2\leftthreetimes D_1))\ldots ),$$ where $D_i=<\varepsilon_{i+1,1},\varepsilon_{i+1,2},
\ldots,\varepsilon_{i+1,i},
\varepsilon_{1,i+1},\varepsilon_{2,i+1},\ldots,\varepsilon_{i,i+1}>$, $i=1,2,\ldots,n-1$. Elements $\varepsilon_{i+1,1},\varepsilon_{i+1,2},
\ldots,\varepsilon_{i+1,i}$ generate a free group of rank $i$ and elements $\varepsilon_{1,i+1},\varepsilon_{2,i+1},\ldots,\varepsilon_{i,i+1}$ generate a free abelian group of rank $i$. This decomposition is consistent with the decomposition of $P_n$, i. e., there is inclusion $U_{i+1}\leq D_i,$ $i=1,2,\ldots,n-1.$
As an immediate consequence of this theorem we obtain the normal form of words in $C_n$.
Every element $\beta \in C_n$ can be written in the form $$\beta=\beta_1\beta_2\ldots \beta_{n-1}\pi_{\beta},$$ where $\pi_{\beta}$ is a permutation automorphism and each $\beta_j$ belongs to the subgroup $D_j$.
For prove of the theorem 3.1 we need the following lemma which follow from defining relations of $Cb_n$
The following conjugating rules are true in $Cb_n$: $$\begin{array}{ll}
1)~~& \varepsilon_{ij}^{-\nu }\varepsilon_{kl}\varepsilon_{ij}^{\nu }=\varepsilon_{kl}, \\
\\
2)~~& \varepsilon_{ij}^{-\nu }\varepsilon_{kj}\varepsilon_{ij}^{\nu }=\varepsilon_{kj}, \\
\\
3)~~& \varepsilon_{ij}^{-\nu }\varepsilon_{ki}\varepsilon_{ij}^{\nu }=
\varepsilon_{kj}^{\nu }\varepsilon_{ki}\varepsilon_{kj}^{-\nu }, \\
\\
4)~~& \varepsilon_{ij}^{-\nu }\varepsilon_{ik}\varepsilon_{ij}^{\nu }=
\varepsilon_{kj}^{\nu }\varepsilon_{ik}\varepsilon_{kj}^{-\nu }, \\
\\
5)~~& \varepsilon_{ij}^{-\nu }\varepsilon_{jk}\varepsilon_{ij}^{\nu }=
[\varepsilon_{kj}^{-\nu }, \varepsilon_{ik}]\varepsilon_{jk}. \\
\end{array}$$
where $\nu =\pm 1$ and distinct letters denote distinct indexes.
Define the following subgroups in $Cb_n$: $$D_i=<\varepsilon_{i+1,1}, \varepsilon_{i+1,2},\ldots,\varepsilon_{i+1,i},
\varepsilon_{1,i+1},
\varepsilon_{2,i+1},\ldots,\varepsilon_{i,i+1}>,~~~i=1,2,\ldots,n-1.$$
Let $k$ and $l$ be the integer numbers such that $2 \leq k < l \leq n$. Then the group $D_{k-1}$ lies in normalizer of $D_{l-1}$.
From definition we have $$D_{k-1}=<\varepsilon_{k,1}, \varepsilon_{k,2},\ldots,\varepsilon_{k,k-1},
\varepsilon_{1,k},
\varepsilon_{2,k},\ldots,\varepsilon_{k-1,k}>,$$ $$D_{l-1}=<\varepsilon_{l,1}, \varepsilon_{l,2},\ldots,\varepsilon_{l,l-1},
\varepsilon_{1,l},
\varepsilon_{2,l},\ldots,\varepsilon_{l-1,l}>.$$ Let $\varepsilon_{k,i}$, $1\leq i \leq k-1$ be some generator of $D_{k-1}$. We should prove that element $\varepsilon^{-\nu }_{k,i}d\varepsilon^{-\nu }_{k,i},$ $\nu =\pm 1$, lies in $D_{l-1}$ for each generator $d$ of $D_{l-1}$. By Lemma 1 we have the following conjugating rules: $$\begin{array}{l}
\varepsilon_{ki}^{-\nu }\varepsilon_{lj}\varepsilon_{ki}^{\nu }=\varepsilon_{lj},~~~
\varepsilon_{ki}^{-\nu }\varepsilon_{jl}\varepsilon_{ki}^{\nu }=\varepsilon_{jl},~~~j\neq i,
k,
\\
\\
\varepsilon_{ki}^{-\nu }\varepsilon_{li}\varepsilon_{ki}^{\nu }=\varepsilon_{li}, \\
\\
\varepsilon_{ki}^{-\nu }\varepsilon_{lk}\varepsilon_{ki}^{\nu }=
\varepsilon_{li}^{\nu }\varepsilon_{lk}\varepsilon_{li}^{-\nu }, \\
\\
\varepsilon_{ki}^{-\nu }\varepsilon_{kl}\varepsilon_{ki}^{\nu }=
\varepsilon_{li}^{\nu }\varepsilon_{kl}\varepsilon_{li}^{-\nu }, \\
\\
\varepsilon_{ki}^{-\nu }\varepsilon_{il}\varepsilon_{ki}^{\nu }=
[\varepsilon_{li}^{-\nu }, \varepsilon_{kl}]\varepsilon_{il}. \\
\end{array}$$ We see that right parts of this rules lie in $D_{l-1}$. Hence in this case lemma is true.
Let $\varepsilon_{ik}$, $1\leq i \leq k-1$ be a generator of $D_{k-1}$. By Lemma 1 we have the following conjugating rules: $$\begin{array}{l}
\varepsilon_{ik}^{-\nu }\varepsilon_{lj}\varepsilon_{ik}^{\nu }=\varepsilon_{lj},~~~
\varepsilon_{ik}^{-\nu }\varepsilon_{jl}\varepsilon_{ik}^{\nu }=\varepsilon_{jl},~~~j\neq i,
k,
\\
\\
\varepsilon_{ik}^{-\nu }\varepsilon_{lk}\varepsilon_{ik}^{\nu }=\varepsilon_{lk}, \\
\\
\varepsilon_{ik}^{-\nu }\varepsilon_{li}\varepsilon_{ik}^{\nu }=
\varepsilon_{lk}^{\nu }\varepsilon_{li}\varepsilon_{lk}^{-\nu }, \\
\\
\varepsilon_{ik}^{-\nu }\varepsilon_{il}\varepsilon_{ik}^{\nu }=
\varepsilon_{lk}^{\nu }\varepsilon_{il}\varepsilon_{lk}^{-\nu }, \\
\\
\varepsilon_{ik}^{-\nu }\varepsilon_{kl}\varepsilon_{ik}^{\nu }=
[\varepsilon_{lk}^{-\nu }, \varepsilon_{il}]\varepsilon_{kl}. \\
\end{array}$$ We see again that right parts of this rules lie in $D_{l-1}$. Hence the lemma is true in all cases.
From this lemma with the help of induction we have
Subgroup $D_{n-1}$ is normal in $Cb_n$.
The following lemma is a part of Theorem 3.1.
If $n\geq 2$ then the group of basis–conjugating automorphisms $Cb_n$ is a semi direct product $$Cb_n=D_{n-1}\leftthreetimes (D_{n-2}\leftthreetimes \ldots (\leftthreetimes (D_{2}\leftthreetimes
D_{1}))\ldots),$$ where $D_i=<\varepsilon_{i+1,1}, \varepsilon_{i+1,2},\ldots,\varepsilon_{i+1,i},
\varepsilon_{1,i+1},
\varepsilon_{2,i+1},\ldots,\varepsilon_{i,i+1}>,$ $
<\varepsilon_{i+1,1}, \varepsilon_{i+1,2},\ldots,\varepsilon_{i+1,i}>\simeq F_i,$ $<\varepsilon_{1,i+1},
\varepsilon_{2,i+1},\ldots,\varepsilon_{i,i+1}>\simeq \mathbb{Z}^i,$ $i=1,2,\ldots,n-1.$
For $n=2$ we have $$Cb_2=D_1=<\varepsilon_{21}, \varepsilon_{12}>.$$ Since the group $Cb_2$ has not relations, then $D_1\simeq F_2=F_1*\mathbb{Z}$ and we have the basis of induction.
We will prove that $Cb_n=D_{n-1}\leftthreetimes Cb_{n-1}$. From Corollary 3.2 subgroup $D_{n-1}$ is normal in $Cb_n$. We divide defining relations (3)–(5) of $Cb_n$ on three parts which is not intersecting. In the first part we pose defining relations which contain only generators from $D_{n-1}$. We see that in every pair of generators of $D_{n-1}$ one index is equal to $n$. Hence in the first part of defining relations only relations from (4) lie, i. e. $$\varepsilon_{in}\varepsilon_{jn}=\varepsilon_{jn}\varepsilon_{in},~~~1 \leq i\neq
j\leq n-1. \eqno{(6)}$$ Therefore $\varepsilon_{1n}, \varepsilon_{2n},\ldots,\varepsilon_{n-1,n}$ generate the abelian subgroup of $D_{n-1}$. Also, it is not hard to check that elements $\varepsilon_{n1}, \varepsilon_{n2},\ldots,\varepsilon_{n,n-1}$ are the free generators of free group $F_{n-1}$. In the second part of relations relations lie only in generators of $Cb_{n-1}$. The third part of relations relations in generators of $D_{n-1}$ and $Cb_{n-1}$ simultaneously lie. We should prove that this relations define an action of $Cb_{n-1}$ on $D_{n-1}$.
Consider relations from (3) which include generators of $D_{n-1}$ and $Cb_{n-1}$ simultaneously. This relations look as $$\varepsilon_{ij}\varepsilon_{kn}=\varepsilon_{kn}\varepsilon_{ij},~~~
\varepsilon_{ij}\varepsilon_{nk}=\varepsilon_{nk}\varepsilon_{ij},~~~1 \leq
i,j,k < n. \eqno{(7)}$$ From these relations we have the following conjugating rules $$\varepsilon_{ij}^{-\nu }\varepsilon_{kn}\varepsilon_{ij}^{\nu }=\varepsilon_{kn},~~~
\varepsilon_{ij}^{-\nu }\varepsilon_{nk}\varepsilon_{ij}^{\nu }=\varepsilon_{nk},~~~
\nu =\pm 1. \eqno{(8)}$$
If defining relation from (4) includes generators of $D_{n-1}$ and $Cb_{n-1}$, then it look as $$\varepsilon_{nj}\varepsilon_{kj}=\varepsilon_{kj}\varepsilon_{nj},~~~1 \leq k,j < n.$$ From this relations we have $$\varepsilon_{kj}^{-\nu }\varepsilon_{nj}\varepsilon_{kj}^{\nu }=\varepsilon_{nj},
~~~1 \leq k,j < n~~~
\nu =\pm 1. \eqno{(9)}$$
If relation from (5) includes generators of $D_{n-1}$ and $Cb_{n-1}$ simultaneously, then this relation has on of the following three types $$\begin{array}{l}
(\varepsilon_{ij}\varepsilon_{nj})\varepsilon_{in}=
\varepsilon_{in}(\varepsilon_{ij}\varepsilon_{nj}), \\
\\
(\varepsilon_{nj}\varepsilon_{ij})\varepsilon_{ni}=
\varepsilon_{ni}(\varepsilon_{nj}\varepsilon_{ij}), \\
\\
(\varepsilon_{in}\varepsilon_{jn})\varepsilon_{ij}=
\varepsilon_{ij}(\varepsilon_{in}\varepsilon_{jn}),~~~1\leq i,j < n. \\
\end{array}$$ From this relations we have the formulas of conjugating: $$\begin{array}{lr}
\varepsilon_{ij}^{-\nu }\varepsilon_{in}\varepsilon_{ij}^{\nu }=
\varepsilon_{nj}^{\nu }\varepsilon_{in}\varepsilon_{nj}^{-\nu },
&~~~~~~~~~~~~~~~~~~~~~~~~~~(10) \\
& \\
\varepsilon_{ij}^{-\nu }\varepsilon_{ni}\varepsilon_{ij}^{\nu }=
\varepsilon_{nj}^{\nu }\varepsilon_{ni}\varepsilon_{nj}^{-\nu },
&~~~~~~~~~~~~~~~~~~~~~~~~~~(11) \\
& \\
\varepsilon_{ij}^{-\nu }\varepsilon_{jn}\varepsilon_{ij}^{\nu }=
[\varepsilon_{nj}^{-\nu },
\varepsilon_{in}]\varepsilon_{jn},~~~1\leq i,j < n,~~~\nu =\pm 1.
&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~(12) \\
\end{array}$$ So, defining relations of $Cb_n$ which include generators of $D_{n-1}$ and $Cb_{n-1}$ simultaneously are equivalent to (8)–(12). Hence, we define action of $Cb_{n-1}$ on $D_{n-1}$. Since $D_{n-1}\bigcap Cb_{n-1}=1$ then $Cb_n=D_{n-1}\leftthreetimes Cb_{n-1}$. From inductive hypothesis we have the needed decomposition.
The group of conjugating automorphisms $C_n$ include the group of basis–conjugating automorphisms $Cb_n$. A. G. Savushkina \[22\] proved that $C_n$ is a semi direct product $C_n=Cb_n\leftthreetimes S_n$ of $Cb_n$ and symmetric group $S_n$, where $S_n$ generating by automorphisms: $$\alpha_{i} : \left\{
\begin{array}{ll}
x_{i} \longmapsto x_{i+1}, & \\
x_{i+1} \longmapsto x_{i}, & \\
x_{l} \longmapsto x_{l}, & l\neq i,i+1,\\
\end{array} \right.$$ $i=1,2,\ldots, n-1,$ and defining relations: $$\alpha_j^2=1,~~~j=1,2,\ldots, n-1, \eqno{(13)}$$ $$\alpha_j\alpha_{j+1}\alpha_j=\alpha_{j+1}\alpha_{j}\alpha_{j+1},~~~j=1,2,\ldots,n-2, \eqno{(14)}$$ $$\alpha_k\alpha_{l}=\alpha_{l}\alpha_{k},~~~1\leq k,l\leq n-1,~~~|k-l|\geq 2. \eqno{(15)}$$ Whole group $C_n$ is generated by $\varepsilon_{ij}$, $1\leq i\neq j\leq n$; $\alpha_{k}$, $1\leq k\leq n-1$ and defined by relations (3)–(5), (13)–(15) and $$\varepsilon_{ij}\alpha_{k}=\alpha_{k}\varepsilon_{ij},~~~k\neq
i-1,i,j-1,j, \eqno{(16)}$$ $$\varepsilon_{ij}\alpha_{i}=\alpha_{i}\varepsilon_{i+1,j},~~~
\varepsilon_{ij}\alpha_{j}=\alpha_{j}\varepsilon_{i,i+1},~~~
\varepsilon_{i,i+1}\alpha_{i}=\alpha_{i}\varepsilon_{i+1,i},~~~
\eqno{(17)}$$ (see \[22, Lemma 1\]).
Evidently, pure braid group $P_n$ is subgroup of $Cb_n$. We can express generators of $P_n$ over generators of $Cb_n$.
Generators of $P_n$ can be express as $$a_{i,i+1}=\varepsilon_{i,i+1}^{-1}\varepsilon_{i+1,i}^{-1},~~~i=1,2,\ldots,n-1, \eqno{(18)}$$ $$a_{ij}=\varepsilon_{j-1,i}\varepsilon_{j-2,i}\ldots\varepsilon_{i+1,i}
(\varepsilon_{ij}^{-1}\varepsilon_{ji}^{-1})\varepsilon_{i+1,i}^{-1}\ldots
\varepsilon_{j-2,i}^{-1}\varepsilon_{j-1,i}^{-1}=$$ $$=\varepsilon_{j-1,j}^{-1}\varepsilon_{j-2,j}^{-1}\ldots\varepsilon_{i+1,j}^{-1}
(\varepsilon_{ij}^{-1}\varepsilon_{ji}^{-1})\varepsilon_{i+1,j}\ldots
\varepsilon_{j-2,j}\varepsilon_{j-1,j},~~~2 \leq i+1 < j \leq n. \eqno{(19)}$$
The formula (18) follows from actions of $P_n$ and $Cb_n$ on free group $F_n$. For proof of the first part of (19) we should verify that automorphism $$a_{ij}\varepsilon_{ji}
(\varepsilon_{j-1,i}\varepsilon_{j-2,i}\ldots\varepsilon_{i+1,i})
\varepsilon_{ij}(\varepsilon_{i+1,i}^{-1}\ldots\varepsilon_{j-2,i}^{-1}\varepsilon_{j-1,i}^{-1})$$ is identical. For proof of the second part of (19) we should verify relation $$\varepsilon_{ki}(\varepsilon_{ij}^{-1}\varepsilon_{ji}^{-1})\varepsilon_{ki}^{-1}=
\varepsilon_{kj}^{-1}(\varepsilon_{ij}^{-1}\varepsilon_{ji}^{-1})\varepsilon_{kj},~~~
i+1\leq k\leq j-1. \eqno{(20)}$$ From Lemma 3.1 we have $$\varepsilon_{ki}\varepsilon_{ij}^{-1}\varepsilon_{ki}^{-1}=\varepsilon_{ij}^{-1}
[\varepsilon_{kj}, \varepsilon_{ji}],~~~
\varepsilon_{ki}\varepsilon_{ji}^{-1}\varepsilon_{ki}^{-1}=\varepsilon_{ji}^{-1}.$$ We use this formulas and rewrite the left–hand side of (20) in the following form $$\varepsilon_{ki}(\varepsilon_{ij}^{-1}\varepsilon_{ji}^{-1})\varepsilon_{ki}^{-1}=
\varepsilon_{ij}^{-1}
[\varepsilon_{kj}, \varepsilon_{ji}]\varepsilon_{ji}^{-1}=
\varepsilon_{ij}^{-1}\varepsilon_{kj}^{-1}\varepsilon_{ji}^{-1}\varepsilon_{kj}.$$ Since $\varepsilon_{ij}^{-1}$ and $\varepsilon_{kj}^{-1}$ are commutative, then we have (20). From this relation we have required formula.
Now we can complete the proof of the Theorem 3.1. We know that free group $U_{i+1}$ generating by $a_{1,i+1}, a_{2,i+1},\ldots,a_{i,i+1}$. By Lemma 3.4 we have $$a_{k,i+1}=
\varepsilon_{i,i+1}^{-1}\varepsilon_{i-1,i+1}^{-1}\ldots\varepsilon_{k+1,i+1}^{-1}
(\varepsilon_{k,i+1}^{-1}\varepsilon_{i+1,k}^{-1})\varepsilon_{k+1,i+1}\ldots
\varepsilon_{i-1,i+1}\varepsilon_{i,i+1},~~~1 \leq k \leq i-1,$$ $$a_{i,i+1}=\varepsilon_{i,i+1}^{-1}\varepsilon_{i+1,i}^{-1}.$$ Hence $U_{i+1}$ is subgroup of $D_i$.
If we use the Tit’s transformation, then not hard to prove that $D_i$ generating by $a_{1,i+1}, a_{2,i+1},\ldots,a_{i,i+1};$ $\varepsilon_{1,i+1},
\varepsilon_{2,i+1},\ldots,\varepsilon_{i,i+1}.$ The proof of the theorem is complete.
This theorem reduces the description of $Cb_n$ to the description of its subgroups $D_{n-1}, \ldots, D_1.$ We can see that this subgroups (with the exception of $D_1$) have the complicated structure. In \[8, p. 167\] the hypothesis was formulated that this subgroups are not finitely define.
By Theorem 3.1 we have $Cb_3 = D_2 \leftthreetimes D_1,$ where $D_2 = < \varepsilon_{31},
\varepsilon_{32}, \varepsilon_{13}, \varepsilon_{23}>$, $D_1 = < \varepsilon_{21},
\varepsilon_{12}> \simeq F_2.$ Elements $\varepsilon_{13}$ and $\varepsilon_{23}$ from $D_2$ are commute. Let us prove that there are many another relations in $D_2$. As it was noted in proof of Theorem 3.1 conjugation by element from $D_1$ induses automorphism of $D_2.$ This implies that if we conjugate relation $\varepsilon_{13} \varepsilon_{23} =
\varepsilon_{23} \varepsilon_{13}$ then we get relation from $D_2.$ In particular, conjugating relation $\varepsilon_{13} \varepsilon_{23} =
\varepsilon_{23} \varepsilon_{13}$ by different powers of $\varepsilon_{12}$ and $\varepsilon_{21}$ we get the following relations $$[\varepsilon_{13} \varepsilon_{23}, \varepsilon_{32}^k
\varepsilon_{13} \varepsilon_{32}^{-k}] = 1,$$ $$[\varepsilon_{13} \varepsilon_{23}, \varepsilon_{31}^k
\varepsilon_{23} \varepsilon_{31}^{-k}] = 1,$$ for each integer $k$.
Some property of group of conjugating automorphisms
===================================================
We find the abelianized groups $C_n/C_n'$ and $Cb_n/Cb_n'$ from the following known fact.
(\[22, Theorem 1\]). Group $C_n$ is generated by automorphisms $\sigma_i$, $\alpha_i$, $i=1,2,\ldots,n-1$ and determined by relations (1)–(2), (13)–(15) and $$\alpha_i\sigma_j=\sigma_j\alpha_i,~~~~~~~|i-j|\geq 2, \eqno{(21)}$$ $$\sigma_i\alpha_{i+1}\alpha_i=\alpha_{i+1}\alpha_i\sigma_{i+1},
~~~\sigma_i\sigma_{i+1}\alpha_i=\alpha_{i+1}\sigma_i\sigma_{i+1}. \eqno{(22)}$$
a\) The abelianized group $C_n/C_n'$, $n\geq 2,$ is isomorphic to $\mathbb{Z}_2\times \mathbb{Z}$.
b\) The abelianized group $Cb_n/Cb_n'$, $n\geq 2,$ is isomorphic to $\mathbb{Z}^{n(n-1)}$.
The group of inner automorphisms ${\rm Inn}(F_n)$ of $F_n$ is subgroup of $Cb_n$.
Intersection $B_n\bigcap {\rm Inn}(F_n)$, $n\geq 2$, is infinite cyclic group which generating by inner automorphisms that is conjugation by $x_1x_2\ldots x_n$.
immediately follows from Artin’s theorem (see § 2) which gives description of $B_n$ as subgroup of ${\rm Aut}(F_n)$. Elements from $B_n$ fixing the product $x_1x_2\ldots x_n$. Inner automorphisms which has this property is only conjugation by $(x_1x_2\ldots x_n)^k$, $k\in \mathbb{Z}$. This completes the proof.
We will show that $C_n$, $n\geq 2$, can be generated at most 4 elements. By the Lemma 4.1 we have $$C_2=<\alpha_1, \sigma_1 \parallel \alpha_1^2=1>\simeq
\mathbb{Z}_2 * \mathbb{Z}.$$ For $n\geq 3$ we have
The group $C_n$, $n\geq 3$, can be generated by $\alpha_1$, $\alpha $, $\sigma_1$, $\sigma $ with defining relations $$\sigma^n=(\sigma \sigma_1)^{n-1},~~~\sigma_1(\sigma^{-j}\sigma_1\sigma^j)=
(\sigma^{-j}\sigma_1\sigma^j)\sigma_1,~~~2\leq j\leq n/2, \eqno{(23)}$$ $$\alpha^2=1,~~~\alpha^n=(\alpha_1\alpha )^{n-1},~~~\alpha_1(\alpha^{-j}\alpha_1\alpha^j)=
(\alpha^{-j}\alpha_1\alpha^j)\alpha_1,~~~2\leq j\leq n/2, \eqno{(24)}$$ $$\alpha^{-i}\alpha_1\alpha^i\sigma^{j}\sigma_1\sigma^{-j}=
\sigma^{j}\sigma_1\sigma^{-j}\alpha^{-i}\alpha_1\alpha^i,~~~1\leq i,j\leq n-2,~~~
|i-j|\geq 2, \eqno{(25)}$$ $$\sigma^{i-1}\sigma_1\sigma^{-(i-1)}\alpha^{-i}\alpha_1\alpha \alpha_1\alpha^{i-1}=
\alpha^{-i}\alpha_1\alpha
\alpha_1\alpha^{i-1}\sigma^{i}\sigma_1\sigma^{-i},~~~2\leq i\leq n-1,
\eqno{(26)}$$ $$\sigma^{i-1}\sigma_1\sigma \sigma_1\sigma^{-i}\alpha^{-(i-1)}\alpha_1\alpha^{i-1}=
\alpha^{-i}\alpha_1\alpha^{i}\sigma^{i-1}\sigma_1\sigma \sigma_1\sigma^{-i},~~~2\leq i\leq
n-1.
\eqno{(27)}$$ New and old generators connecting by formulas $$\sigma = \sigma_1\sigma_2\ldots\sigma_{n-1},~~~\alpha = \alpha_{n-1}\alpha_{n-2}\ldots\alpha_1,
~~~ \eqno{(28)}$$ $$\sigma_{i+1} = \sigma^i\sigma_1\sigma^{-i},~~~\alpha_{i+1} = \alpha^{-i}\alpha_{1}\alpha^i,
~~~ 1\leq i\leq n-2. \eqno{(29)}$$
It is well–known \[23, Chapter 6\] that $B_n$, $n\geq 3$, is generated by $\sigma_1$ and $\sigma $ and defined by relations (23). Analogous, symmetric group $S_n$ is generated by $\alpha_1, \alpha $ and defined by relations (24). New and old generators are connecting by formulas (28) and (29). Hence in presentation of $C_n$ from the Lemma 4.1 we can replace the generators $\alpha_i$, $\sigma_i$, $i=1,2,\ldots,n-1$, by generators $\alpha_1$, $\alpha $, $\sigma_1$, $\sigma $, and replace relations (1)–(2) by relations (23) and replace relations (13)–(15) by relations (24). Substitute in relations (21)–(22) expression $\alpha_i$, $\sigma_j$ through $\alpha $, $\alpha_1$, $\sigma_1$, $\sigma $, we obtain (25)–(27). This completes the proof.
Let $V$ be a set of words in an alphabet of variables ranging over elements of group $G$. The subgroup $V(G)$ of the group $G$ generated by all values of words from $V^{\pm 1}$ is called the [*verbal subgroup*]{} defined by the set $V$. The [*width*]{} wid$(G, V)$ of the verbal subgroup $V(G)$ with respect to $V$ is defined to be the minimal number $m\in
\mathbb{N} \bigcup \{+\infty \}$ such that any element of $V(G)$ can be presented as a product of $\leq m$ values of words from $V^{\pm 1}$.
Subgroup is called the [*proper subgroup*]{} if it is not trivial and it is not all group. The set $V$ is called the [*proper set of words*]{}, if the verbal subgroup $V(F_2)$ is a proper subgroup of $F_2$. If $V$ is not a proper set of words, then it is called [*unproper set of words*]{}. The width of verbal subgroup with respect to unproper set of words is finite \[24, Lemma 1\]. If in group $G$ each finite proper set of words defines the proper verbal subgroup, then we will said that $G$ [*is rich by verbal subgroups*]{}. We will said that $G$ [*has not proper verbal subgroups of finite width*]{}, if for each finite proper set $V$ the width wid$(G, V)$ is infinite.
Some years ago author has proved \[25\] that the braid group $B_n$ for $n\geq 3$ has not proper verbal subgroups of finite width. This result finishes investigations of some authors which started with the G. S. Makanin’s question from “The Kourovka Notebook” (Unsolved problems in group theory) \[9\].
The width of verbal subgroups for HNN–extensions and groups with one relations were investigated in \[26\]. The width of derived groups of some matrix groups over rings were investigated in \[27\]. The connection between the width of derived group, commutator length and equations in free group were studied in \[28\]. Some Artin groups were investigated in \[29\]. In this article were proved
(\[29, Lemma 3\]). If there is an epimorphism from $G$ on group which has not proper verbal subgroups of finite width and which is rich by verbal subgroups, then $G$ also has not proper verbal subgroups of finite width.
We will use this lemma and prove
The group of bases–conjugating automorphisms $Cb_n$ for $n\geq 2$ has not proper verbal subgroups of finite width.
By the Theorem 3.1 there exists an epimorphism of $Cb_{n}$ for $n \geq 2$ on $Cb_{2}$. The group $Cb_2$ is isomorphic to $F_2$. From result of A. H. Rhemtulla \[24\] we have that $F_2$ has not proper verbal subgroups of finite width and it is evident that $F_2$ is rich by verbal subgroups. The result follows from the Lemma 4.2.
Is it true that for $n\geq 2$ the group $C_n$ has not proper verbal subgroups of finite width?
Expansion of known linear representations of braid group on $C_n$
==================================================================
The Burau representation $\psi : B_n\longrightarrow {\rm GL}(W_n)$ maps $B_n$ in group of automorphisms ${\rm GL}(W_n)$ of free $n$ dimension module $W_n$ over ring $\mathbb{Z}[q^{\pm 1}]$ of Laurent polynomials with an integer coefficients. If $w_1, w_2,\ldots , w_n$ is a free basis of $W_n$, then $\psi(\sigma_i)$, $i=1,2,\ldots,n-1$ acts by the rule (we will write $\sigma_i(w_j)$ or $w_j^{\sigma_i}$ instead $\psi(\sigma_i)(w_j)$ or $w_j^{\psi(\sigma_i)}$ accordingly): $$\left.
\begin{array}{ll}
\sigma_i(w_i)=(1-q)w_i+qw_{i+1}, & \\
\sigma_i(w_{i+1})=w_{i}, & \\
\sigma_i(w_l)=w_l,~~~ \mbox{if}~~~l\neq i,i+1.&
\end{array}
\right.$$ We will consider right action of $B_n$ on $W_n$.
A. G. Savushkina \[22\] proved that group of conjugating automorphisms $C_n$ generating by automorphisms $\sigma_{1}, \sigma_{2},\ldots, \sigma_{n-1},$ $\alpha_{1}, \alpha_{2}, \ldots, \alpha_{n-1}$ of free group $F_n$. The set $\sigma_{1}, \sigma_{2},\ldots, \sigma_{n-1}$ generates braid group $B_n$ and satisfy relations (1)–(2). The set of automorphisms $\alpha_{1}, \alpha_{2}, \ldots, \alpha_{n-1}$ generates the symmetric group $S_n$ which defines by relations (13)–(15). Group $C_n$ is defined by relations of $B_n$, $S_n$ and relations $$\alpha_i \sigma_j = \sigma_j \alpha_i~~~\mbox{if}~~~ |i-j|\geq 2,
\eqno{(30)}$$ $$\sigma_i\alpha_{i+1}\alpha_i=\alpha_{i+1}\alpha_i\sigma_{i+1}~~~\mbox{if}~~~
i=1,2,...,n-2, \eqno{(31)}$$ $$\sigma_{i+1}\sigma_{i}\alpha_{i+1}=\alpha_{i}\sigma_{i+1}\sigma_{i}~~~\mbox{if}~~~
i=1,2,...,n-2. \eqno{(32)}$$
There is a linear representations $\widetilde{\psi} : C_n\longrightarrow {\rm GL}(W_n)$ of $C_n$ which is an expansion of Burau representation $\psi : B_n\longrightarrow {\rm GL}(W_n)$.
For the generators of braid group $B_n$ we define $\widetilde{\psi}(\sigma_i)=\psi(\sigma_i)$. For the generators $\alpha_{1}, \alpha_{2}, \ldots, \alpha_{n-1}$ define the action on basis of $W_n$ by equalities $$\left.
\begin{array}{l}
\alpha_i(w_i)=w_{i+1},\\
\alpha_i(w_{i+1})=w_{i},\\
\alpha_i(w_l)=w_l,~~~ \mbox{if}~~~l\neq i,i+1.\\
\end{array}
\right.$$ Every elements of $C_n$ is a word in alphabet $\sigma^{\pm 1}_{1}, \sigma^{\pm 1}_{2},\ldots, \sigma^{\pm 1}_{n-1}$, $\alpha_{1}, \alpha_{2}, \ldots, \alpha_{n-1}$. Hence we defined the map $\widetilde{\psi} : C_n\longrightarrow {\rm GL}(W_n)$. We should prove that this map is homomorphism. For this we should prove that all relations of $C_n$ are carry out in $\widetilde{\psi}(C_n)$. It is a simple exercise. The proposition is proved.
We remind (see \[14, 15, 16\]) definition of faithful linear representation of Lawrence–Krammer braid group $B_n$. Let $V_n$ be a free module dimension $m=n(n-1)/2$ with basis $v_{ij}$, $1\leq i < j\leq n$, over ring $\mathbb{Z}[t^{\pm 1}, q^{\pm 1}]$ of Laurent polynomials on two variables. Then representation $\rho
: B_n\longrightarrow {\rm GL}(V_n)$ was defined by action by $\sigma_i$, $i=1,2,\ldots,n-1$, on module $V_n$ by equalities $$\left.
\begin{array}{l}
\sigma_i(v_{k,i})=(1-q)v_{k,i}+qv_{k,i+1}+q(q-1)v_{i,i+1},\\
\\
\sigma_i(v_{k,i+1})=v_{k,i}~~~ \mbox{if }~~~k < i,\\
\\
\sigma_i(v_{i,i+1})=tq^2v_{i,i+1},\\
\\
\sigma_i(v_{i,l})=tq(q-1)v_{i,i+1}+(1-q)v_{i,l}+qv_{i+1,l}~~~ \mbox{if}~~~i+1 <
l,\\
\\
\sigma_i(v_{i+1,l})=v_{i,l},\\
\\
\sigma_i(v_{k,l})=v_{k,l}~~~ \mbox{if }~~~\{ k, l\} \bigcap \{ i,
i+1\}=\emptyset.\\
\end{array}
\right.$$ From this equalities it is not difficult to find the action by $\sigma_i^{-1}$: $$\left.
\begin{array}{l}
\sigma_i^{-1}(v_{k,i})=v_{k,i+1},\\
\\
\sigma_i^{-1}(v_{k,i+1})=q^{-1}v_{k,i}+q^{-1}(q-1)v_{k,i+1}-t^{-1}q^{-2}(q-1)v_{i,i+1}~~~
\mbox{if }~~~k < i,\\
\\
\sigma_i^{-1}(v_{i,i+1})=t^{-1}q^{-2}v_{i,i+1},\\
\\
\sigma_i^{-1}(v_{i,l})=v_{i+1,l}~~~ \mbox{if }~~~i+1 < l,\\
\\
\sigma_i^{-1}(v_{i+1,l})=-q^{-2}(q-1)v_{i,i+1}+q^{-1}v_{i,l}+q^{-1}(q-1)v_{i+1,l},\\
\\
\sigma_i^{-1}(v_{k,l})=v_{k,l}~~~ \mbox{if }~~~\{ k, l\} \bigcap \{ i,
i+1\}=\emptyset.\\
\end{array}
\right.$$ and the actions by $a_{i,i+1}=\sigma_i^2$ $(i=1,2,\ldots,n-1)$: $$\left.
\begin{array}{l}
a_{i,i+1}(v_{k,i})=(q^2-q+1)v_{k,i}+q(1-q)v_{k,i+1}+q(q-1)(tq^2-q+1)v_{i,i+1},\\
\\
a_{i,i+1}(v_{k,i+1})=(1-q)v_{k,i}+qv_{k,i+1}+q(q-1)v_{i,i+1}~~~ \mbox{if }~~~k <
i,\\
\\
a_{i,i+1}(v_{i,i+1})=t^2q^4v_{i,i+1},\\
\\
a_{i,i+1}(v_{i,l})=tq(q-1)(tq^2-q+1)v_{i,i+1}+(q^2-q+1)v_{i,l}+q(1-q)v_{i+1,l}~~~ \mbox{if }~~~i+1 <
l,\\
\\
a_{i,i+1}(v_{i+1,l})=tq(q-1)v_{i,i+1}+(1-q)v_{i,l}+qv_{i+1,l},\\
\\
a_{i,i+1}(v_{k,l})=v_{k,l}~~~ \mbox{if }~~~\{ k, l\} \bigcap \{ i,
i+1\}=\emptyset.\\
\end{array}
\right.$$
We want to expansion the representation of Lawrence–Krammer on $C_n$. For this we should define the action of $S_n$ on module $V_n$. Let $S_n$ act on the set $\{1,2,\ldots,n \}$ by permutation. If $\pi $ is a permutation from $S_n$, then let $\pi
(v_{ij})=v_{\pi(i),\pi(j)}$. Thus automorphism $\alpha_i$ act on basis of $V_n$ by the rules: $$\left.
\begin{array}{l}
\alpha_i(v_{k,i})=v_{k,i+1},\\
\\
\alpha_i(v_{k,i+1})=v_{k,i}~~~ \mbox{if}~~~k < i,\\
\\
\alpha_i(v_{i,i+1})=v_{i,i+1},\\
\\
\alpha_i(v_{i,l})=v_{i+1,l}~~~ \mbox{if}~~~i+1 < l,\\
\\
\alpha_i(v_{i+1,l})=v_{i,l},\\
\\
\alpha_i(v_{k,l})=v_{k,l}~~~ \mbox{if}~~~\{ k, l\} \bigcap \{ i,
i+1\}=\emptyset.\\
\end{array}
\right.$$
It is not difficult to see that this action define a homomorphism from $S_n$ to ${\rm GL}(V_n)$. For this we note that equalities for action of $\alpha_i$ are obtained from equalities for action of $\sigma_i$ if we take $q=t=1$. Hence relations (14)–(15) are valid. The relation $\alpha_i^2=1$ also are valid that is following from the action of $a_{i,i+1}=\sigma_i^2$ on $V_n$. Hence we construct the linear representation of $S_n$. As well–known if $n\geq 5$ then $S_n$ has a simple subgroup $A_n$ of index 2 and our representation are faithful. We have proved
The linear representation $S_n\longrightarrow {\rm GL}(V_n)$ constructed above are faithful linear representation of $S_n$.
We have the map of $C_n$ in ${\rm GL}(V_n)$ to be defined by action on generators $\alpha_1, \alpha_2,\ldots,\alpha_{n-1}$, $\sigma_1, \sigma_2,\ldots,\sigma_{n-1}$. This map is keep all relations which include only generators $\alpha_i$ and keep all relations which include only generators $\sigma_i$. We should consider remaining relations. It is not difficult to show that all relations $$\alpha_i\sigma_j=\sigma_j\alpha_i,~~ \mbox{if} ~~|i-j|\geq 2,$$ $$\sigma_{i+1}\sigma_{i}\alpha_{i+1}=\alpha_i\sigma_{i+1}\sigma_i,$$ are keep.
Consider relation $\sigma_i\alpha_{i+1}\alpha_i=\alpha_{i+1}\alpha_i\sigma_{i+1}$. Acting by the left–hand side of this relations on basis of $V_n$ we have $$\left.
\begin{array}{l}
v_{k,i}^{\sigma_i\alpha_{i+1}\alpha_i}=(1-q)v_{k,i+1}+qv_{k,i+2}+q(q-1)v_{i+1,i+2},\\
\\
v_{k,i+1}^{\sigma_i\alpha_{i+1}\alpha_i}=v_{k,i+1}~~~ \mbox{if}~~~k <
i,\\
\\
v_{i,i+1}^{\sigma_i\alpha_{i+1}\alpha_i}=tq^2v_{i+1,i+2},\\
\\
v_{i,l}^{\sigma_i\alpha_{i+1}\alpha_i}=tq(q-1)v_{i+1,i+2}+(1-q)v_{i+1,l}+qv_{i+2,l}~~~ \mbox{if}~~~i+2 <
l,\\
\\
v_{i,i+2}^{\sigma_i\alpha_{i+1}\alpha_i}=tq(q-1)v_{i+1,i+2}+(1-q)v_{i,i+1}+qv_{i,i+2},\\
\\
v_{i+1,l}^{\sigma_i\alpha_{i+1}\alpha_i}=v_{i+1,l}~~~ \mbox{if}~~~l >
i+2,\\
\\
v_{i+1,i+2}^{\sigma_i\alpha_{i+1}\alpha_i}=v_{i,i+1},\\
\\
v_{k,l}^{\sigma_i\alpha_{i+1}\alpha_i}=v_{k,l}~~~ \mbox{if}~~~\{ k, l\} \bigcap \{ i,
i+1, i+2\}=\emptyset,\\
\\
v_{k,i+2}^{\sigma_i\alpha_{i+1}\alpha_i}=v_{k,i}~~~ \mbox{if}~~~k <
i,\\
\\
v_{i+2,l}^{\sigma_i\alpha_{i+1}\alpha_i}=v_{i,l}.\\
\end{array}
\right.$$ Acting by the right–hand side of this relations, we have $$\left.
\begin{array}{l}
v_{k,i}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=(1-q)v_{k,i+1}+qv_{k,i+2}+q(q-1)v_{i+1,i+2},\\
\\
v_{k,i+1}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=v_{k,i+1}~~~ \mbox{if}~~~k <
i,\\
\\
v_{i,i+1}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=tq^2v_{i+1,i+2},\\
\\
v_{i,l}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=tq(q-1)v_{i+1,i+2}+(1-q)v_{i+1,l}+qv_{i+2,l}~~~ \mbox{if}~~~i+2 <
l,\\
\\
v_{i,i+2}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=(1-q)v_{i,i+1}+qv_{i,i+2}+q(q-1)v_{i+1,i+2},\\
\\
v_{i+1,l}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=v_{i+1,l}~~~ \mbox{if}~~~l >
i+2,\\
\\
v_{i+1,i+2}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=v_{i,i+1},\\
\\
v_{k,l}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=v_{k,l}~~~ \mbox{if}~~~\{ k, l\} \bigcap \{ i,
i+1, i+2\}=\emptyset,\\
\\
v_{k,i+2}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=v_{k,i}~~~ \mbox{if}~~~k <
i,\\
\\
v_{i+2,l}^{\alpha_{i+1}\alpha_i\sigma_{i+1}}=v_{i,l}.\\
\end{array}
\right.$$
Comparing the actions of automorphisms $\sigma_i\alpha_{i+1}\alpha_i$ and $\alpha_{i+1}\alpha_i\sigma_{i+1}$ on $v_{i,i+2}$ we see that they are equal if and only if $t=1$. Therefore the following proposition is true
The map $\rho : C_n\longrightarrow {\rm GL}(V_n)$ constructed above is a linear representation if and only if $t=1$.
For $t=1$ $B_n$–module $V_n$ is isomorphic to the symmetric square of $B_n$–module $W$ and in this case representation $\rho : C_n\longrightarrow {\rm GL}(V_n)$ is unfaithful for $n\geq 5$. The question of the faithfulity of this representation for $n=3,4$ remained open.
We can define the action of $\alpha_i$ on $V_n$ by another way. Not difficult to prove
For every nonzero number $x\in \mathbb{C}$ the map $\rho_1 : S_n\longrightarrow {\rm GL}(V_n)$ defining by action of $\alpha_i$ on the $V_n$ by equalities $$\left.
\begin{array}{l}
\alpha_i(v_{k,i})=xv_{k,i+1},\\
\\
\alpha_i(v_{k,i+1})=x^{-1}v_{k,i}~~~ \mbox{if}~~~k < i,\\
\\
\alpha_i(v_{i,i+1})=v_{i,i+1},\\
\\
\alpha_i(v_{i,l})=xv_{i+1,l}~~~ \mbox{if}~~~i+1 < l,\\
\\
\alpha_i(v_{i+1,l})=x^{-1}v_{i,l},\\
\\
\alpha_i(v_{k,l})=v_{k,l}~~~ \mbox{if}~~~\{ k, l\} \bigcap \{ i,
i+1\}=\emptyset.\\
\end{array}
\right.$$ is the faithful linear representation.
If we will expand the map $\rho_1$ on all $C_n$ defining $\rho_1(\sigma_i)=\rho(\sigma_i)$, then we will have the map $\rho_1 : C_n\longrightarrow {\rm GL}(V_n)$. We see that this map keeps relations $$\sigma_i\alpha_{i+1}\alpha_i=\alpha_{i+1}\alpha_i\sigma_{i+1},$$ $$\sigma_{i+1}\sigma_{i}\alpha_{i+1}=\alpha_i\sigma_{i+1}\sigma_i$$ if we take $x=t^{-1/3}$. but in this case the relation $\alpha_i\sigma_j=\sigma_j\alpha_i$ is not keep since $$v_{i,j}^{\alpha_i\sigma_j}=x(1-q)v_{i+1,j}+xqv_{i+1,j+1}+xq(q-1)v_{j.j+1},$$ $$v_{i,j}^{\sigma_j\alpha_i}=x(1-q)v_{i+1,j}+xqv_{i+1,j+1}+q(q-1)v_{j.j+1},$$ and $v_{i,j}^{\alpha_i\sigma_j}=v_{i,j}^{\sigma_j\alpha_i}$ if and only if $x=1$. Since there not exist relations $\alpha_i\sigma_j=\sigma_j\alpha_i$ in $C_3$ then we have
The linear representation $\rho_1 : C_3\longrightarrow {\rm GL}(V_3)$ is expansion of Lawrence–Krammer representation on $C_3$.
Linear representations of braid groups of some manifolds
========================================================
As it was proved (see \[7, Theorem 1.11; 32\]) that braid group $B_n(S^2)$ of the 2–sphere $S^2$ admits a presentation with generators $\delta_1, \delta_2,\ldots,\delta_{n-1}$ and defining relations: $$\delta_i\delta_{i+1}\delta_{i}=\delta_{i+1}\delta_{i}\delta_{i+1}~~~\mbox{if}~~~
i=1,2,...,n-2,$$ $$\delta_i\delta_{j}=\delta_{j}\delta_{i}~~~\mbox{if}~~~ |i-j|~~~ \geq
2,$$ $$\delta_1\delta_{2}\ldots \delta_{n-2}\delta_{n-1}^2\delta_{n-2}\ldots
\delta_{2}\delta_1=1.$$ From this relations we see that $B_n(S^2)$ is a homomorphic image of $B_n$. Since $B_2(S^2)$ is a cyclic group of order 2 and $B_3(S^2)$ is a metacyclic group of order 12, then we will consider $n>3$.
R. Gillette and J. Van Buskirk \[30\] have studied the structure of $B_n(S^2)$. We recall some of their results. Let $$a_{i,i}=1,~~~a_{i,j}=\delta_i^{-1}\delta_{i+1}^{-1}\ldots \delta_{j-2}^{-1}\delta_{j-1}^2\delta_{j-2}\ldots
\delta_{i+1}\delta_i,~~~1\leq i< j\leq n. \eqno{(33)}$$ These elements generate the pure braid group $P_n(S^2)$ which is a kernel of homomorphism $\nu : B_n(S^2)\longrightarrow S_n$ and $\nu $ send $\delta_i$ in transposition $(i,i+1)$, $1\leq i\leq n-1$, from $S_n$. Let define the subgroup $A_{n-i+1}=<a_{i,i+1}, a_{i,i+2},\ldots,a_{i,n}>$ for each $i=1,2,\ldots,n-1$. Subgroup $A_n$ is normal in $P_n(S^2)$ and we have the short exact sequence $$1\longrightarrow A_n\longrightarrow P_n(S^2)\longrightarrow
P_{n-1}(S^2)\longrightarrow 1$$ and $P_n(S^2)$ is a semi direct product: $P_n(S^2)=A_n\leftthreetimes
P_{n-1}(S^2)$. The generators of $A_n$ are connected by the relation $$a_{1,2}a_{1,3}\ldots a_{1,n}=1.$$ Since the following relations are hold in $P_n(S^2)$ $$a_{i,i+1}a_{i,i+2}\ldots a_{i,i+n-1}=1,~~~i=1,2,\ldots,n-1,$$ where $a_{j,i}=a_{i,j}$ if $j > i$ and indices are taken mod$n$, then $$a_{i,n}=a_{i,n-1}^{-1}\ldots a_{i,i+1}^{-1}a_{i-1,i}^{-1}\ldots
a_{1,i}^{-1},~~~i=1,2,\ldots,n-1.$$ Using this formulas we can exclude $a_{1,n}, a_{2,n},\ldots, a_{n-1,n}$ from the set of generators of $P_n(S^2)$. Hence $A_{n-i+1}$ is freely generated by $a_{i,i+1}, a_{i,i+2},\ldots, a_{i,n-1}$.
The center of $B_n(S^2)$ is the cyclic group of order 2 generated by the Dirac braid $$\Delta_n=(\delta_1\delta_2 \ldots \delta_{n-2})^{n-1}=
(a_{1,2}a_{1,3}\ldots a_{1,n-1})(a_{2,3}a_{2,4}\ldots a_{2,n-1})\ldots
(a_{n-2,n-1}).$$
The group $P_n(S^2)$ was split in the semi direct product $$P_n(S^2)=A_n\leftthreetimes (A_{n-1}\leftthreetimes (\ldots
\leftthreetimes A_3)\ldots).$$ Let $L_n=A_n\leftthreetimes (A_{n-1}\leftthreetimes (\ldots
\leftthreetimes A_4)\ldots)$. Since $A_3$ is a cyclic group generating by $a_{n-2,n-1}$, then $P_n(S^2)=L_n\times <\Delta_n>\simeq L_n\times \mathbb{Z}_2$. Group $L_n$ is isomorphic to subgroup $U_n\leftthreetimes (U_{n-1}\leftthreetimes (\ldots
\leftthreetimes U_4)\ldots)\leq P_n$.
The braid group of sphere closed relates with the group $M(0,n)$ of mapping classes of the $n$–punctured 2–sphere. Presentation of the $M(0,n)$ derived from that of $B_n(S^2)$ by adding the relation: $\Delta_n=(\delta_1\delta_2 \ldots \delta_{n-2})^{n-1}=1$. There is an epimorphism from $B_n(S^2)$ to $M(0,n)$. The kernel of this epimorphism is equal to the center of $B_n(S^2)$. A mapping class group of the $n$–punctured 2–sphere can be visualized as an $n$–string braid between concentric 2–spheres, where the inner sphere is free to execute full revolutions.
For $M(0,n)$ there is a short exact sequence $$1\longrightarrow L_n\longrightarrow M(0,n)\longrightarrow
S_{n}\longrightarrow 1,$$ where $L_n$ is a kernel of epimorphism $\nu : M(0,n)\longrightarrow S_n$ that map $\delta_i$ to the transposition $(i,i+1)$, $i=1,2,\ldots, n-2$.
A. I. Malcev \[31, Lemma 1; 21\] have proved that if $H$ is subgroup of finite index in group $G$ and $H$ is linear, then $G$ is linear too. Recall his construction. Let $\vert G : H \vert =m$ and $\psi
: H\longrightarrow {\rm GL}_{l}(F)$ is a faithful linear representation of $H$ by matrices of order $l$ over field $F$. In this case $G$ is a union of right cosets $$G=He \bigcup Hg_2 \bigcup \ldots \bigcup Hg_m.$$ For each $g\in G$ we can write the product $g_ig$ uniquely in the form $h_ig_{n_i}$, $h_i \in H$. Therefore for each $g$ we have the sequence $h_1, h_2, \ldots, h_m$ from $H$ and the sequence numbers $n_1, n_2, \ldots, n_m$. For the sequence $\{ n_i \}$ we construct the matrix $D(n_i)$ with integer coefficients by the rule: $$D(n_i)=\parallel d_{j,k} \parallel \in
M_{lm}(\mathbb{Z}),~~~d_{j,n_j}=E_l,~~~d_{j,k}=0~~~
\mbox{if}~~~k\neq n_j,$$ where $E_l$ is identical matrix of order $l$. Then we can map $g\in G$ to the matrix $${\rm diag}(\psi(h_1), \psi(h_2),\ldots, \psi(h_m))D(n_i).$$ Hence we constructed the linear representation of $G$ in ${\rm GL}_{lm}(F)$. This linear representation is faithful. Using this construction we will prove
Let $R=\mathbb{Z}[t^{\pm 1}, q^{\pm 1}]$ be a ring of Laurent polynomials on two variables. Then the groups $M(0,n)$ and $B_n(S^2)$ are linear for each $n\geq 2$. For each $n\geq 4$ there are inclusion maps $\varphi : M(0,n)\longrightarrow {\rm GL}_m(R)$ and $\varphi_1 : B_n(S^2)\longrightarrow {\rm GL}_{m_1}(R)$, where $m=(n-1)(n-2)n!/2$, $m_1=2m$.
As we noted above, if $n=2,3$ groups $M(0,n)$, $B_n(S^2)$ are finite and hence are linear. Since $L_n$ is isomorphic to subgroup of $B_{n}$, then there are faithful linear representation $\rho : L_n\longrightarrow {\rm GL}_l(R)$ for $l=(n-1)(n-2)/2$. This representation is induced by Lawrence–Krammer representation of $B_n$.
Let $m_1, m_2, \ldots, m_{n!}$ be coset representatives of $M(0,n)$ by subgroup $L_n$. Since $M(0,n)$ is generated by $\delta_1, \delta_2, \ldots, \delta_{n-1}$ then we can define $\varphi $ only on these elements. Each generator $\delta_k$ acts on set of coset representatives by permutation. We find $$m_i\delta_k =
h_i^km_{\pi_k(i)},~~~i=1,2,\ldots,n!,~~~k=1,2,\ldots,n-1,$$ where symbol $k$ in the upper part is an index but not exponent (this rule will be true to the end of this proof). To compare with $\delta_k$ the matrix $$\varphi(\delta_k) = {\rm diag}(\rho (h_1^k), \rho (h_2^k),\ldots, \rho (h_{n!}^k))
\pi(\delta_k)\in {\rm GL}_m(R),$$ where ${\rm diag}(\rho (h_1^k), \rho (h_2^k),\ldots, \rho (h_{n!}^k))$ is a block–diagonal matrix; $\pi(\delta_k)$ is a block–monomial matrix in which the block on position $(j,\pi(j))$ is an identical matrix of order $l$, but the block on position $(j,s)$ for $s\neq \pi(j)$ is a null matrix of order $l$. This linear representation $\varphi $ is a faithful representation of $M(0,n)$.
Consider the group $B_n(S^2)$. This group contains the linear subgroup $L_n$ index $2n!$. As a set of right coset representatives of $B_n(S^2)$ by subgroup $L_n$ we take elements $m_i\Delta_n^{\epsilon }$, $i=1,2,\ldots,n!$, $\epsilon =0,1,$ where $\Delta_n$ is the generator of center of $B_n(S^2)$. Since $B_n(S^2)$ is generated by $\delta_1, \delta_2, \ldots, \delta_{n-1}$ then it is enough to define representation $\varphi_1$ on these elements. Let us ordere the coset representatives $m_i\Delta_n^{\epsilon }$ and denote them by $n_1, n_2,\ldots, n_{2n!}$. Each generator $\delta_k$ of $B_n(S^2)$ acts on these representatives by the rules $$n_j\delta_k =
g_j^k n_{\pi_k(j)},~~~j=1,2,\ldots,2n!,~~~k=1,2,\ldots,n-1.$$ To compare with $\delta_k$ the matrix $$\varphi_1(\delta_k) = {\rm diag}(\rho (g_1^k), \rho (g_2^k),\ldots, \rho (g_{2n!}^k))
\pi(\delta_k)\in {\rm GL}_{2m}(R),$$ where ${\rm diag}(\rho (g_1^k), \rho (g_2^k),\ldots, \rho (g_{2n!}^k))$ is a block–diagonal matrix and $\pi(\delta_k)$ is a block–monomial matrix in which the block on position $(j,\pi(j))$ is an identical matrix of order $l$ but the block on position $(j,s)$ for $s\neq \pi(j)$ is a null matrix of order $l$. This linear representation $\varphi_1$ is a faithful representation of $B_n(S^2)$ over ring $R$. Since $R$ is included in the field of complex numbers $\mathbb{C}$ (it is enough to take numbers $t$ and $q$ which are non–zero transcendental over $\mathbb{Q}$) then we obtain the required assertions.
Let $P^2$ be a projective plane. The braid group $B_n(P^2)$, $n\geq 1,$ admits a presentation \[32\] with generators $\delta_1, \delta_2,\ldots,\delta_{n-1},$ $\rho_1, \rho_2,\ldots,\rho_{n}$ and defining relations: $$\delta_i\delta_{i+1}\delta_{i}=\delta_{i+1}\delta_{i}\delta_{i+1}~~~\mbox{if}~~~ i=1,2,...,n-2,$$ $$\delta_i\delta_{j}=\delta_{j}\delta_{i}~~~\mbox{if}~~~ |i-j| \geq 2,$$ $$\delta_i\rho_{j}=\rho_{j}\delta_{i},~~~j\neq i,i+1,$$ $$\rho_i = \delta_i\rho_{i+1}\delta_{i}~~~\mbox{if}~~~ i=1,2,...,n-1,$$ $$\rho_{i+1}^{-1}\rho_{i}^{-1}\rho_{i+1}\rho_{i}=\delta_{i}^2~~~\mbox{if}~~~ i=1,2,...,n-1,$$ $$\delta_1\delta_{2}\ldots \delta_{n-2}\delta_{n-1}^2\delta_{n-2}\ldots
\delta_{2}\delta_1=\rho_1^2.$$
It is clear that the map taking $\delta_i$ onto the transposition $(i,i+1)$, $i=1, 2, \ldots, n-1,$ and each $\rho_j$ onto the identity is homomorphism of $B_n(P^2)$ onto the symmetric group $S_n$, with kernel $P_n(P^2)$.
The group $B_1(P^2)$ is a cyclic group of order 2, $B_2(P^2)$ is a finite group of order 16. For $n\geq 3$ the group $B_n(P^2)$ is infinite. Consider the case $n=3$. The group $P_3(P^2)$ contains a free subgroup $A_3(P^2)=<\rho_1, a_2, a_3 \parallel a_2a_3=\rho_1^2>$, where $a_2=\delta_1^2$, $a_3=\delta_1^{-1}\delta_2^2\delta_1$. The factor group $P_3(P^2)/A_3(P^2)\simeq P_2(P^2)$ is a quaternion group of order 8. The set $$\{ 1, \delta_1,
\delta_2, \delta_2\delta_1, \delta_1\delta_2, \delta_1\delta_2\delta_1 \}$$ is a right coset representatives of $B_3(P^2)$ by $P_3(P^2)$. Note that $P_2(P^2)$ has a presentation $$P_2(P^2) = <\rho_2, \rho_3 \parallel
\rho_2^2=\rho_3^2=(\rho_3\rho_2)^2>.$$ In $P_3(P^2)$ the following relations are true: $$\rho_3^{-\epsilon }a_2\rho_3^{\epsilon
}=a_2,~~~\epsilon=\pm 1,~~~\rho_j^{-1}\rho_1\rho_j=\rho_1a_j^{-1},~~j=2,3,$$ $$\rho_2^{-1}a_3\rho_2=a_2a_3a_2^{-1},~~~\rho_2^{-1}a_2\rho_2=\rho_1a_2^{-1}\rho_1^{-1}.$$ From defining relations of $B_3(P^2)$ we get $\delta_1$, $\delta_2$.
In $B_3(P^2)$ the following rules of conjugating are true $$a_2^{\delta_1}=a_2,~~~\rho_1^{\delta_1}=\rho_2a_2,~~~\rho_2^{\delta_1}=a_2^{-1}\rho_1,
~~~\rho_3^{\delta_1}=\rho_3,$$ $$a_2^{\delta_2}=\rho_1^2a_2^{-1},~~~\rho_1^{\delta_2}=\rho_1,
~~~\rho_2^{\delta_2}=\rho_3\delta_2^2,~~~\rho_3^{\delta_2}=\delta_2^{-2}\rho_2.$$
Since $A_3(P^2)$ is a free group with free generators $\rho_1$, $a_2$ then using Sanov’s representation \[21\] we can include $A_3(P^2)$ in ${\rm SL}_2(\mathbb{Z})$ assuming $$\overline{\rho}_1=
\left(
\begin{array}{cc}
1 & 2 \\
0 & 1 \\
\end{array}
\right),~~~
\overline{a}_2=
\left(
\begin{array}{cc}
1 & 0 \\
2 & 1 \\
\end{array}
\right)~~~$$ as images of $\rho_1$ and $a_2$ accordingly.
The group $P_3(P^2)$ is an extension of $A_3(P^2)$ by the group of quaternion of order 8. Let as chose the set $$\{ e, \rho_2, \rho^2_2, \rho^3_2, \rho_3, \rho_3\rho_2, \rho_3\rho^2_2, \rho_3\rho^3_2
\}$$ as the set of coset representatives of $P_3(P^2)$ by $A_3(P^2)$. Thus the free group $A_3(P^2)$ has index 48 in $B_3(P^2)$. Using Malcev’s construction we get
Braid group $B_3(P^2)$ of projective plane $P^2$ is included in ${\rm SL}_{96}(\mathbb{Z})$.
The faithful linear representation of ${\rm Aut}(F_2)$
======================================================
It is known \[17\] that ${\rm Aut}(F_n)$ is not linear for $n\geq 3$. If $n=2$ then in \[20\] was proved that ${\rm Aut}(F_2)$ is linear if and only if $B_4$ is. Since $B_4$ is linear then we can construct the faithful linear representation of ${\rm Aut}(F_2)$. Recall some results from \[20\].
Let $B_4^{*}=B_4/Z(B_4)$ where $Z(B_4)$ is a center of $B_4$. As it was noted above $Z(B_4)$ is an infinite cyclic group generating by $\Delta_4=(\sigma_1\sigma_2\sigma_3)^4$. The group $B_4^{*}$ is isomorphic to ${\rm Aut}^+(F_2)$ which is preimage of ${\rm SL}_2(\mathbb{Z})$ by homomorphism $\xi : {\rm Aut}(F_2)\longrightarrow {\rm GL}_2(\mathbb{Z})$ sending automorphism of $F_2$ to automorphism of $\mathbb{Z}^2$. The group ${\rm
Aut}^+(F_2)$ is subgroup of index 2 in ${\rm Aut}(F_2)$.
Let $\rho : B_4\longrightarrow {\rm GL}_6(\mathbb{C})$ be a faithful linear representation of Lawrense–Krammer. M. Zino \[33\] have proved that this representation is irreducible.
Denote by $F$ subgroup of $B_4$ generating by $x=\sigma_1\sigma^{-1}_3$, $y=\sigma_2\sigma_1\sigma^{-1}_3\sigma^{-1}_2$. This group is free with free generators $x$, $y$ and is normal in $B_4$. Inner automorphisms of $B_4$ induced automorphisms of $F$, i. e., there is epimorphism $h : B_4\longrightarrow {\rm Aut}^+(F_2)$. The kernel of this epimorphism is equal to the center of $B_4$ and images of generators $\sigma_1$, $\sigma_2$, $\sigma_3$ defining automorphisms $$h(\sigma_{1})=\alpha_1 : \left\{
\begin{array}{ll}
x \longmapsto x, & \\
y \longmapsto yx^{-1}, & \\
\end{array} \right.~~~
h(\sigma_{2})=\alpha_2 : \left\{
\begin{array}{ll}
x \longmapsto y, & \\
y \longmapsto yx^{-1}y, & \\
\end{array} \right.~~~
h(\sigma_{3})=\alpha_3 : \left\{
\begin{array}{ll}
x \longmapsto x, & \\
y \longmapsto x^{-1}y. & \\
\end{array} \right.~~~$$
The whole group ${\rm Aut}(F_2)$ is generated by automorphisms $$P : \left\{
\begin{array}{ll}
x \longmapsto y, & \\
y \longmapsto x, & \\
\end{array} \right.~~~
\omega : \left\{
\begin{array}{ll}
x \longmapsto x^{-1}, & \\
y \longmapsto y, & \\
\end{array} \right.~~~
U : \left\{
\begin{array}{ll}
x \longmapsto xy, & \\
y \longmapsto y, & \\
\end{array} \right.~~~$$ and defined by relations $$P^2=\omega^2=(\omega P)^4=(P\omega PU)^2=(UP\omega
)^3=[\omega, \omega U\omega ]=1.$$ The following formulas are hold $$\alpha_1=PU^{-1}P,~~~\alpha_2=PU\omega U^{-1},~~~\alpha_3=P\omega U\omega P.$$
Using the Lawrense–Krammer representation $\rho : B_4\longrightarrow {\rm GL}_6(\mathbb{C})$ we find the images of generators of $B_4$: $$\rho_1=\rho(\sigma_1)=
\left(
\begin{array}{cccccc}
tq^2 & 0 & 0 & 0 & 0 & 0 \\
tq(q-1) & 1-q & 0 & q & 0 & 0 \\
tq(q-1) & 0 & 1-q & 0 & q & 0 \\
0 & 1 & 0 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 & 1 \\
\end{array}
\right),$$ $$\rho_2=\rho(\sigma_2)=
\left(
\begin{array}{cccccc}
1-q & q & 0 & q(q-1) & 0 & 0 \\
1 & 0 & 0 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 & 0 & 0 \\
0 & 0 & 0 & tq^2 & 0 & 0 \\
0 & 0 & 0 & tq(q-1) & 1-q & q \\
0 & 0 & 0 & 0 & 1 & 0 \\
\end{array}
\right),$$ $$\rho_3=\rho(\sigma_3)=
\left(
\begin{array}{cccccc}
1 & 0 & 0 & 0 & 0 & 0 \\
0 & 1-q & q & 0 & 0 & q(q-1) \\
0 & 1 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 1-q & q & q(q-1) \\
0 & 0 & 1 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 & tq^2 \\
\end{array}
\right).$$ It is not difficult to check that $(\rho_1\rho_2\rho_3)^4=t^2q^8E_6$ is the image of center of $B_4$. Let $\mu=1/\sqrt[12]{t^2q^8}$ and define $\overline{\rho}(\sigma_i)=\mu \rho(\sigma_i)=\mu \rho_i$. Then $\overline{\rho}$ induce the faithful linear representation of $B_4^{*}\simeq {\rm Aut}^+F_2$.
Since ${\rm Aut}^+F_2$ is subgroup of index 2 in ${\rm Aut}F_2$ then $e$ and $\omega $ is a right coset representatieves. It is not difficult to check
In ${\rm Aut}F_2$ the following equalities are hold $$\begin{array}{ll}
1) & \omega \alpha_1=\alpha^{-1}_1\omega,\\
2) & \omega
\alpha_2=\alpha_3\alpha_1\alpha^{-1}_2\alpha^{-1}_1\alpha^{-1}_3\omega,\\
3) & \omega \alpha_3=\alpha^{-1}_3\omega.
\end{array}$$
Since ${\rm Aut}F_2$ is generated by automorphisms $\alpha_1, $ $\alpha_2,$ $\alpha_3$, $\omega,$ then acting on coset representatives of ${\rm Aut}F_2$ by ${\rm Aut}^+F_2$ we find the correspondence diagonal and monomial matrix. Define $$\psi(\alpha_1)={\rm diag}(\overline{\rho }_1, (\overline{\rho
}_1)^{-1}),~~~\psi(\alpha_2)={\rm diag}(\overline{\rho }_2,
\overline{\rho }_3\overline{\rho }_1(\overline{\rho }_2)^{-1}(\overline{\rho
}_1)^{-1}(\overline{\rho}_3)^{-1}),$$ $$\psi(\alpha_3)={\rm diag}(\overline{\rho }_3, (\overline{\rho
}_3)^{-1}),~~~
\psi(\omega )=
\left(
\begin{array}{cc}
0 & E_6 \\
E_6 & 0 \\
\end{array}
\right),$$ where $E_6$ is the identical matrix of order 6 and $\overline{\rho }_i$ defined above. If we conjugate this representation by matrix $c={\rm diag}(E_6, \overline{\rho }_3\overline{\rho}_1)$ we receive the new representation: $$\overline{\psi }(\alpha_i)=c^{-1}\psi(\alpha_i)c=
\left(
\begin{array}{cc}
\overline{\rho }_i & 0 \\
0 & \overline{\rho }_i^{-1} \\
\end{array}
\right),~~~i=1,2,3;~~~
\overline{\psi }(\omega )=c^{-1}\psi(\omega )c=
\left(
\begin{array}{cc}
0 & \overline{\rho }_3 \overline{\rho }_1 \\
(\overline{\rho }_3 \overline{\rho }_1)^{-1} & 0 \\
\end{array}
\right).
\eqno{(34)}$$
The following theorem is true
The map $\overline{\psi } : {\rm Aut}F_2\longrightarrow {\rm GL}_{12}(\mathbb{Z}[t^{\pm 1}, q^{\pm 1}])$, defining on generators in (34) is a faithful linear representation of ${\rm Aut}F_2$.
REFERENCES
1. D. Z. Djokovic, The structure of the automorphism group of a free group on two generators, Proc. Amer. Math. Soc., 88, N 2 (1983), 218–220.
2. G. T. Kozlov, The structure of the automorphism group ${\rm Aut}(F_2)$, Algebra, Logika i prilozhenija, Irkutsk, IGU, 1994, 28–32 (in Russian).
3. R. C. Lyndon and P. E. Schupp, Combinatorial group theory, Springer–Verlag, 1977.
4. S. Krstic, J. McCool, The non-finite presentability of ${\rm IA}(F_3)$ and ${\rm GL}_2(\mathbb{Z}[t,t^{-1}])$, Inven. Math., 129, N 3 (1997), 595–606.
5. J. McCool, On basis–conjugating automorphisms of free groups, Can. J. Math., 38, N 6 (1986), 1525–1529.
6. A. A. Markov, Foundations of the algebraic theory of braids, Trudy Math. Inst. Steklov, 1945, 16, 1–54 (in Russian).
7. J. S. Birman, Braids, links and mapping class group, Princeton–Tokyo: Univ. press, 1974.
8. D. J. Collins, N. D. Gilbert, Structure and torsion in automorphism groups of free products, Quart. J. Math. Oxford, 41, N 162 (1990), 155–178.
9. The Kourovka Notebook (Unsolved problems in group theory), 15th edn., Institute of Mathematics SO RAN, Novosibirsk 2002.
10. W. Burau, Uber Zopfgruppen und gleichsinnig verdrillte Verkettungen, Abh. Math. Semin. Hamburg Univ, 11, 1936, 179–186.
11. J. A. Moody, The Burau representation of the braid group $B_n$ is unfaithful for large $n$, Bull. Amer. Math. Soc. 25, N 2 (1991), 379-384.
12. D. D. Long, M. Paton, The Burau representation is not faithful for $n\geq 6$, Topology, 32, N 2 (1993), 439-447.
13. S. Bigelow, The Burau representation of $B_5$ is not faithful, Geom. Topology, 3, (1999), 397-404.
14. R. J. Lawrence, Homological representation of the Hecke Algebra, Commun. Math. Phys., 135, N 1 (1990), 141-191.
15. D. Krammer, Braid groups are linear, Annals of Math., 155, N 1 (2002), 131-156.
16. S. Bigelow, Braid groups are linear, J. Amer. Math. Soc., 14, (2001), 471-486.
17. E. Formanek, C. Procesi, The automorphism groups of a free group is not linear, J. Algebra, 149, N 2 (1992), 494–499.
18. V. G. Bardakov, M. V. Nechshadim, Some property of braid groups of compact oriented 2–manifolds, IV International Algebraic Conference dedicated to the 60th anniversary of Yu. I. Merzljakov, Novosibirsk, Institute of Mathematics, 2000, 9–13 (in Russian).
19. S. Bigelow, R. D. Budney, The mapping class group of genus two surface is linear, to appear in “Algebraic and Geometric Topology”.
20. J. L. Dyer, E. Formanek, E. K. Grossman, On the linearity of automorphism groups of free groups, Arch. Math., 38, N 5 (1982), 404–409.
21. M. I. Kargapolov, Yu. I. Merzljakov, Fundamentals of the theory of groups, New York: Springer, 1979.
22. A. G. Savushkina, On group of conjugating automorphisms of free group, Matem. Zametki, 60, N 1 (1996), 92–108 (in Russian).
23. H. S. M. Coxeter, H. W. O. J. Mozer, Generators and relations for discrete groups, Berlin; Heidelberg; New York: Springer, 1957.
24. A. H. Rhemtulla, A problem of bounded expressibility in free products, Proc. Cambridge Phil. Soc., 64, N 3 (1969), 573–584.
25. V. G. Bardakov, On the theory of braid groups, Russ. Acad. Sci., Sb. Math., 76, N 1 (1993), 123–153 (Zbl 0798.20029).
26. V. G. Bardakov, On the width of verbal subgroups of certain free constructions, Algebra Logic, 36, N 5 (1997), 288–301 (Zbl 0941.20017).
27. V. G. Bardakov, On the decomposition of automorphisms of free modules into simple factors, Izv. Ross. Acad. Nauk, Ser. Mat., 59, N 2 (1995), 109–128 (Zbl 0896.20031).
28. V. G. Bardakov, Computing the commutator length in free groups, Algebra Log., 39, N 4 (2000), 395–440 (Zbl 096.20019).
29. V. G. Bardakov, The width of verbal subgroups of some Artin groups, Lavrent’ev (ed.), Group and metric properties of mappings. Novosibirsk: Novosibirskij Gosudarstvennyj Universitet, 8–18 (1995), (in Russian) (Zbl 0943.20034).
30. R. Gillette, J. Van Buskirk, The word problem and consequences for the braid groups and mapping class groups of the 2–sphere, Trans. Amer. Math. Soc., 131, N 2 (1968), 277–296.
31. A. I. Malcev, On isomorphic representation of infinite groups by matrix, Matem. Sb., 8, N 3 (1940), 405–422 (in Russian).
32. J. Van Buskirk, Braid groups of compact 2–manifolds with elements of finite order, Trans. Amer. Math. Soc., 122, N 1 (1966), 81–97.
33. M. G. Zinno, On Krammer’s representation of the braid group, Math. Ann., 321, N 1 (2000), 197–211.
Author address:
Valerij Bardakov\
Sobolev Institute of Mathematics,\
Novosibirsk, 630090, Russia\
[[email protected]]{}
[^1]: Partially supported by the Russian Foundation for Basic Research (grant number 02–01–01118).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The performance of all-optical dual-hop relayed free-space optical communication systems is analytically studied and evaluated. We consider the case when the total received signal undergoes turbulence-induced channel fading, modeled by the versatile mixture-Gamma distribution. Also, the misalignment-induced fading due to the presence of pointing errors is jointly considered in the enclosed analysis. The performance of both amplify-and-forward and decode-and-forward relaying transmission is studied, when heterodyne detection is applied. New closed-form expressions are derived regarding some key performance metrics of the considered system; namely, the system outage probability and average bit-error rate.'
author:
- 'Nikolaos I. Miridakis'
- 'Dimitrios D. Vergados'
- Angelos Michalas
bibliography:
- 'References.bib'
date: 'Received: date / Accepted: date'
title: 'All-Optical FSO Relaying Under Mixture-Gamma Fading Channels and Pointing Errors'
---
[example.eps]{} gsave newpath 20 20 moveto 20 220 lineto 220 220 lineto 220 20 lineto closepath 2 setlinewidth gsave .4 setgray fill grestore stroke grestore
Introduction {#intro}
============
One of the main challenges in free-space optical (FSO) communications is the channel fading due to turbulence-induced scintillation and misalignment. The former is mainly caused by the random fluctuation of the refractive index, whereas the latter by time-varying pointing errors due to thermal expansion, dynamic wind loads, and internal vibrations [@j:FaridHranilovic2012]. To mitigate channel fading, an all-optical cooperative transmission via relaying can be used, which brings impressive performance improvements in FSO systems [@j:KashaniUysal2013; @j:ChatzidiamantisMichalopoulos2013; @j:YangGaoAlouini2014].
To efficiently address the composite turbulence-induced and misalignment fading, various distribution models have been proposed. Among them, the most popular ones (due to their accuracy and generality) are the Gamma-Gamma and Málaga distribution models. These models, combined with the pointing error effect, result to highly complicated probability density function (PDF)/cumulative distribution function (CDF) because the rather cumbersome high-order Meijer’s $G$-function is included therein [@j:AnsariYilmaz2016]. Due to this reason, the performance of relayed transmission has been only merely studied so far in FSO communication systems. Specifically, only approximate analytical results have been presented to date, regarding the asymptotically high signal-to-noise ratio (SNR) regime (e.g., see [@j:AnsariYilmaz2016; @j:YangGaoAlouini2014] and relevant references therein). Nevertheless, analytical results for the performance of all-optical relayed FSO systems in the entire SNR regime (low-to-high) are not available so far.
Capitalizing on the aforementioned observations, we study the performance of all optical dual-hop relayed FSO systems over composite fading channels and pointing errors in arbitrary SNR regions. Three popular relaying transmission protocols are adopted; namely, the channel state information (CSI)-assisted amplify-and-forward (AF), the AF with a fixed (statistical) gain, and the decode-and-forward (DF) relaying transmission schemes. The mixture Gamma (MG) distribution model is used to capture the statistical properties of turbulence-induced channel fading. It was recently indicated in [@j:SandalidisMG2016] that the MG distribution sharply coincides to both the Gamma-Gamma and Málaga generalized distribution models. Thus, it serves as an effective model to accurately approach channel fading in weak-to-strong turbulence conditions. Further, the misalignment-induced fading is also considered, by assuming the commonly adopted model of zero-boresight pointing errors.
The analysis of current work focuses on the coherent heterodyne detection type. Although it represents a more complicated detection method than other alternatives, it has the ability to better overcome the thermal noise effects [@j:YangGaoAlouini2014; @j:LiuYao2010]. New closed-form expressions are derived for key system performance metrics, i.e., the system outage probability and ABER. The derived results can serve as an efficient means of evaluating the performance of the considered relaying schemes, whereas they are much more time-effective than the available methods so far (i.e., time-consuming manifold numerical integrations and/or exhaustive simulations).
*Notation*: $f_{X}(\cdot)$, $F_{X}(\cdot)$ and $\bar{F}_{X}(\cdot)$ represent PDF, CDF and complementary CDF of the random variable (RV) $X$, respectively. Moreover, $\Gamma(\cdot)$ denotes the Gamma function [@tables Eq. (8.310.1)], $\Gamma(\cdot,\cdot)$ is the upper incomplete Gamma function [@tables Eq. (8.350.2)], and ${\rm erf(\cdot)}$ represents the error function [@tables Eq. (8.253.1)]. Finally, ${}_2F_1(\cdot,\cdot;\cdot;\cdot)$ denotes the Gauss hypergeometric function [@tables Eq. (9.100)], while $W_{\mu,\nu}(\cdot)$ is the Whittaker’s-$W$ hypergeometric function [@tables Eq. (9.220.4)].
System Model {#System Model}
============
Consider an all-optical dual-hop relayed FSO system, which consists of a source, relay and destination node; all equipped with a single aperture. Two transmission phases (e.g., two consecutive time slots) are required to complete the end-to-end information transmission. In the first transmission phase, the source converts the RF signal to optical signal and send it to the relay node. Upon reception, the relay process the latter signal according to the underlying relaying transmission protocol. Then, in the second transmission phase, the relay retransmits the signal to the destination. It is assumed that the direct source-to-destination link does not exist, due to strong propagation attenuation and/or severe channel fading conditions. We retain our focus on three popular relaying protocols; namely, CSI-assisted AF relaying, AF relaying using a fixed (statistical) gain, and DF relaying. Correspondingly, the end-to-end SNR is given by [@j:YangGaoAlouini2014] $$\begin{aligned}
\nonumber
\left\{
\begin{array}{c l}
\gamma_{\rm CSI}&\triangleq \frac{\gamma_{1}\gamma_{2}}{\gamma_{1}+\gamma_{2}+q}, \\
\gamma_{\rm Fixed}&\triangleq \frac{\gamma_{1}\gamma_{2}}{\gamma_{2}+U},\\
\gamma_{\rm DF}&\triangleq \min\{\gamma_{1},\gamma_{2}\},
\end{array}\right.\end{aligned}$$ where $\gamma_{1}$ and $\gamma_{2}$ denote the received SNR at the two consecutive transmission phases, while $U$ is a fixed constant related to the fixed AF gain. Also, the parameter $q\in\{0,1\}$ corresponds to the approximate CSI-assisted AF relayed case (when $q=0$) or the exact case (when $q=1$).
The received signal at the two transmission phases undergoes joint channel fading (irradiance) and misalignment due to the possible pointing errors, whereas it is subject to an additive white zero-mean Gaussian noise with power $N_{0}$. The recently proposed MG distribution is used to model the turbulence-induced fading. Doing so, it turns out that the PDF of SNR at the $j^{\rm th}$ transmission hop, with $j\in \{1,2\}$, reads as [@j:SandalidisMG2016] $$\begin{aligned}
\nonumber
f_{\gamma_{j}}(x)=&\sum^{L}_{i_{j}=1}\frac{a_{i_{j}}c^{\xi^{2}_{j}-b_{i_{j}}}_{i_{j}}\xi^{2}_{j}x^{\xi^{2}_{j}-1}}{(A_{0_{j}}\bar{\gamma}_{j})^{\xi^{2}_{j}}}\\
&\times \Gamma\left(b_{i_{j}}-\xi^{2}_{j},\frac{c_{i_{j}}}{A_{0_{j}}\bar{\gamma}_{j}}x\right),\ \ x>0,
\label{snrpdf}\end{aligned}$$ where $\bar{\gamma}_{1}\triangleq P_{S}\eta_{1}\bar{I}_{1}/N_{0}$ and $\bar{\gamma}_{2}\triangleq P_{R}\eta_{2}\bar{I}_{2}/N_{0}$ denote the average received SNRs at the relay and destination nodes, respectively, while $\eta_{j}$ is the electrical-to-optical conversion coefficient of the $j^{\rm th}$ hop.[^1] Also, $P_{S}$ and $P_{R}$ represent the transmit power of the source and relay, respectively, while $\bar{I}_{j}$ denotes the average received irradiance at the $j^{\rm th}$ hop, which is expressed as [@j:SandalidisMG2016] $$\begin{aligned}
\bar{I}_{j}=\left(\frac{\xi^{2}_{j}A_{0_{j}}}{1+\xi^{2}_{j}}\right)\sum^{L}_{i_{j}=1}a_{i_{j}}\Gamma(1+b_{i_{j}})c^{-(1+b_{i_{j}})}_{i_{j}}.
\label{meanI}\end{aligned}$$ Moreover, $a_{i_{j}}$, $b_{i_{j}}$ and $c_{i_{j}}$ denote the parameters of MG distribution. As an illustrative example, to model Gamma-Gamma channel fading, these parameters are defined in [@j:SandalidisMG2016 Eq. (5)]. A similar matching can be easily obtained when considering another generalized fading model, namely, the Málaga distribution [@j:SandalidisMG2016 Eq. (9)]. At this point, it is worth noting that $b_{i_{j}}$ reflects the minimum between the small- and large scale channel fading parameters at the $j^{\rm th}$ hop [@j:MiridakisTsiftsisletter2017; @j:MiridakisVergadosMichalllllas2017]. The total number of sum terms in determines the accuracy level of MG distribution. It was shown in [@j:SandalidisMG2016 Figs. 2 and 3] that a condition of $L\leq 10$ satisfies an acceptable accuracy level for most practical applications, resulting to a rapidly converging series of . This reveals the great success of the MG distribution model. Further, $A_{0}$ is a constant term that defines the pointing loss given by $A_{0}=[{\rm erf}(\sqrt{\pi}r/(\sqrt{2}w_{z}))]^{2}$, while $r$ is the radius of the detection aperture, and $w_{z}$ is the beam waist. In addition, $\xi$ denotes the ratio between the equivalent beam radius at the receiver and the pointing error displacement standard deviation (jitter) at the receiver (i.e., the condition when $\xi\rightarrow +\infty$ reflects the non-pointing error case).
Statistical Analysis {#Statistical Analysis}
====================
We commence by extracting some key statistical results, namely, the CDF of the end-to-end received SNR for various relaying transmission modes. These results are quite useful for the overall performance evaluation of the considered system.
Under the condition $b_{i_{j}}-\xi^{2}_{j} \in \mathbb{N}^{+}$ and using [@tables Eq. (8.352.4)], reduces to $$\begin{aligned}
f_{\gamma_{j}}(x)=\sum^{L}_{i_{j}=1}\sum^{b_{i_{j}}-\xi^{2}_{j}-1}_{k_{j}=0}\Xi(i_{j})x^{\xi^{2}_{j}+k_{j}-1}\exp\left(-\frac{c_{i_{j}}}{A_{0_{j}}\bar{\gamma}_{j}}x\right),
\label{snrpdf1}\end{aligned}$$ where $$\begin{aligned}
\Xi(i_{j})\triangleq \frac{a_{i_{j}}c^{\xi^{2}_{j}-b_{i_{j}}+k_{j}}_{i_{j}}\xi^{2}_{j}(b_{i_{j}}-\xi^{2}_{j}-1)!}{k_{j}!(A_{0_{j}}\bar{\gamma}_{j})^{\xi^{2}_{j}+k_{j}}}.
\label{Xi}\end{aligned}$$
Notice that the restriction $b_{i_{j}}-\xi^{2}_{j} \in \mathbb{N}^{+}$ applies for integer channel fading parameters (or to be more precise, $b_{i_{j}}-\xi^{2}_{j}$ should be an integer value). In the case when at least one of the involved parameters is non-integer, can serve as a close-approximate of and/or as a performance benchmark, since it bounds the SNR performance between the nearest upper and lower integer values of the corresponding fading parameter.
Moreover, in the case when $b_{i_{j}}-\xi^{2}_{j}\leq 0$ (i.e., the misalignment-induced parameter due to the pointing error effect is equal or greater than the turbulence-induced fading parameter), then the following inequality is used $$\begin{aligned}
\nonumber
\Gamma(-j,x)&\triangleq \int^{\infty}_{x}t^{-j-1}\exp(-t)dt\\
\nonumber
&=\exp(-x)\int^{\infty}_{0}(t+x)^{-j-1}\exp(-t)dt\\
&\leq \exp(-x)x^{-j-1},\ j\geq 0,
\label{inequality}\end{aligned}$$ which gets more tight for low-to-moderate SNR regions.[^2] According to the latter inequality, it stems that $$\begin{aligned}
f_{\gamma_{j}}(x)\leq f^{(B)}_{\gamma_{j}}(x),\ x>0,
\label{snrpdfbound}\end{aligned}$$ where $$\begin{aligned}
f^{(B)}_{\gamma_{j}}(x)=\sum^{L}_{i_{j}=1}\frac{a_{i_{j}}c^{-1}_{i_{j}}\xi^{2}_{j}}{(A_{0_{j}}\bar{\gamma}_{j})^{b_{i_{j}}-1}}x^{b_{i_{j}}-2}\exp\left(-\frac{c_{i_{j}}}{A_{0_{j}}\bar{\gamma}_{j}}x\right),\end{aligned}$$ stands for the upper bound of the true PDF $f_{\gamma_{j}}(\cdot)$ and can be directly obtained from by simply setting $k_{j}=b_{i_{j}}-\xi^{2}_{j}-1$. In what follows, we use the general form of and we revisit the tightness of the bound (in the case when $b_{i_{j}}-\xi^{2}_{j}\leq 0$) in the numerical results.
The CCDF of SNR at the $j^{\rm th}$ transmission hop is given by $$\begin{aligned}
\nonumber
\bar{F}_{\gamma_{j}}(x)&=1-F_{\gamma_{j}}(x)\\
\nonumber
&=\sum^{L}_{i_{j}=1}\sum^{b_{i_{j}}-\xi^{2}_{j}-1}_{k_{j}=0}\sum^{\xi^{2}_{j}+k_{j}-1}_{r_{j}=0}\frac{\Xi(i_{j})(\xi^{2}_{j}+k_{j}-1)!}{r_{j}!\left(\frac{c_{i_{j}}}{A_{0_{j}}\bar{\gamma}_{j}}\right)^{\xi^{2}_{j}+k_{j}-r_{j}}}\\
&\ \ \ \ \times x^{r_{j}} \exp\left(-\frac{c_{i_{j}}}{A_{0_{j}}\bar{\gamma}_{j}}x\right),
\label{ccdfsnr}\end{aligned}$$
Based on the above statistics, the following lemmas provide the CDF for end-to-end SNR for different types of relayed transmission mode.
The CDF of the end-to-end SNR for a dual-hop CSI-assisted relayed transmission is derived by in a closed-form expression.
$$\begin{aligned}
\nonumber
F_{\gamma_{\rm CSI}}(x)&=1-\sum^{L}_{i_{1}=1}\sum^{b_{i_{1}}-\xi^{2}_{1}-1}_{k_{1}=0}\sum^{\xi^{2}_{1}+k_{1}-1}_{r_{1}=0}\sum^{L}_{i_{2}=1}\sum^{b_{i_{2}}-\xi^{2}_{2}-1}_{k_{2}=0}\sum^{\xi^{2}_{2}+k_{2}-1}_{p=0}\sum^{r_{1}}_{s=0}\binom{r_{1}}{s}\binom{\xi^{2}_{2}+k_{2}-1}{p}\\
\nonumber
&\times \frac{2\Xi(i_{1})\Xi(i_{2})(\xi^{2}_{1}+k_{1}-1)!\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}\right)^{\frac{p-s+1}{2}+r_{1}-\xi^{2}_{1}-k_{1}}(x^{2}+q x)^{\frac{p+s+1}{2}}}{r_{1}!\left(\frac{c_{i_{2}}}{A_{0_{2}}\bar{\gamma}_{2}}\right)^{\frac{p-s+1}{2}}\exp\left(\left[\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}+\frac{c_{i_{2}}}{A_{0_{2}}\bar{\gamma}_{2}}\right]x\right)}\\
&\times x^{\xi^{2}_{2}+k_{2}+r_{1}-p-s-1} K_{p-s+1}\left(2\sqrt{\frac{c_{i_{1}}c_{i_{2}}}{A_{0_{1}}A_{0_{2}}\bar{\gamma}_{1}\bar{\gamma}_{2}}(x^{2}+q x)}\right)
\label{cdfafcsi}\end{aligned}$$
It holds that [@tsiftsis2006nonregenerative] $$\begin{aligned}
F_{\gamma_{\rm CSI}}(x)\triangleq 1-\int^{\infty}_{0}\bar{F}_{\gamma_{1}}\left(x+\frac{x^{2}+q x}{y}\right)f_{\gamma_{2}}(x+y)dy.
\label{cdfcsiii}\end{aligned}$$ Hence, inserting and into , utilizing [@tables Eqs. (1.111) and (3.471.9)], and after performing some straightforward manipulations, (\[cdfafcsi\]) is obtained.
The CDF of the end-to-end SNR for a dual-hop relayed transmission with a fixed gain is provided as $$\begin{aligned}
\nonumber
F_{\gamma_{\rm Fixed}}(x)&=1-\sum^{L}_{i_{1}=1}\sum^{b_{i_{1}}-\xi^{2}_{1}-1}_{k_{1}=0}\sum^{\xi^{2}_{1}+k_{1}-1}_{r_{1}=0}\sum^{L}_{i_{2}=1}\sum^{b_{i_{2}}-\xi^{2}_{2}-1}_{k_{2}=0}\sum^{r_{1}}_{s=0}\binom{r_{1}}{s}\Xi(i_{1})\Xi(i_{2})\\
\nonumber
&\times \frac{2(\xi^{2}_{1}+k_{1}-1)!\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}\right)^{\frac{\xi^{2}_{2}+k_{2}-s}{2}+r_{1}-\xi^{2}_{1}-k_{1}}U^{\frac{\xi^{2}_{2}+k_{2}+s}{2}}x^{\frac{\xi^{2}_{2}+k_{2}-s}{2}+r_{1}}}{r_{1}!\left(\frac{c_{i_{2}}}{A_{0_{2}}\bar{\gamma}_{2}}\right)^{\frac{\xi^{2}_{2}+k_{2}-s}{2}}\exp\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}x\right)}\\
&\times K_{\xi^{2}_{2}+k_{2}-s}\left(2\sqrt{\frac{c_{i_{1}}c_{i_{2}}}{A_{0_{1}}A_{0_{2}}\bar{\gamma}_{1}\bar{\gamma}_{2}}U x}\right),
\label{cdfaffixed}\end{aligned}$$ where $$\begin{aligned}
\nonumber
U\triangleq \left(\mathbb{E}\left[\frac{1}{\gamma_{1}+1}\right]\right)^{-1}=\Bigg[&\sum^{L}_{i_{1}=1}\sum^{b_{i_{1}}-\xi^{2}_{1}-1}_{k_{1}=0}\frac{\Xi(i_{1})\Gamma(\xi^{2}_{1}+k_{1})}{\exp\left(-\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}\right)}\\
&\times \Gamma\left(1-\xi^{2}_{1}-k_{1},\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}\right)\Bigg]^{-1}.
\label{ugain}\end{aligned}$$
It holds that [@tsiftsis2006nonregenerative] $$\begin{aligned}
F_{\gamma_{\rm Fixed}}(x)\triangleq \int^{\infty}_{0}F_{\gamma_{1}}\left(x+\frac{U x}{y}\right)f_{\gamma_{2}}(y)dy,
\label{cdffixed}\end{aligned}$$ Hence, following the same steps as for deriving (\[cdfafcsi\]), yields (\[cdfaffixed\]). Also, regarding the derivation of , the fixed AF gain is defined as [@j:HasnaAlouini2004] $U\triangleq (\int^{\infty}_{0}(x+1)^{-1}f_{\gamma_{1}}(x)dx)^{-1}$. Hence, using [@tables Eq. (3.383.10)], is directly extracted.
The CDF of the end-to-end SNR for a DF relayed transmission is expressed as $$\begin{aligned}
F_{\gamma_{\rm DF}}(x)\triangleq F_{\min\{\gamma_{1},\gamma_{2}\}}(x)=1-\prod^{2}_{j=1}\bar{F}_{\gamma_{j}}(x).
\label{cdfdf}\end{aligned}$$
System Performance {#System Performance}
==================
Capitalizing on the previously derived statistics, some useful metrics that define the overall system performance are analytically evaluated.
Outage Probability
------------------
The outage probability is defined as the probability that the SNR falls below a certain threshold value, $\gamma_{\text{th}}$, such that $P_{\rm out}(\gamma_{\rm th})=F_{\gamma}(\gamma_{\rm th})$. With the aid of the closed-form expressions of the previously derived lemmas, the system outage performance can be easily computed for each relayed transmission mode, correspondingly.
Average Bit-Error Rate
----------------------
The ABER is defined as [@j:AnsariYilmaz2016] $$\begin{aligned}
\overline{P}_{b}\triangleq \frac{Q^{P}}{2 \Gamma(P)}\int^{\infty}_{0}z^{P-1}\exp\left(-Q z\right)F_{\gamma}(z)dz,
\label{aberdef}\end{aligned}$$ where $P$ and $Q$ are fixed modulation-specific parameters. For the CSI-assisted AF relaying scenario, the exact ABER is analytically intractable. Yet, the case when $q=0$ can serve as a sharp approximation in moderate-to-high SNR.
The ABER of the end-to-end SNR for a dual-hop CSI-assisted relayed transmission, when $q=0$, is presented as $$\begin{aligned}
\nonumber
&\bar{P_{b}}_{\gamma_{\rm CSI}}=\scriptstyle \frac{1}{2}-\sum^{L}_{i_{1}=1}\sum^{b_{i_{1}}-\xi^{2}_{1}-1}_{k_{1}=0}\sum^{\xi^{2}_{1}+k_{1}-1}_{r_{1}=0}\sum^{L}_{i_{2}=1}\sum^{b_{i_{2}}-\xi^{2}_{2}-1}_{k_{2}=0}\sum^{\xi^{2}_{2}+k_{2}-1}_{p=0}\sum^{r_{1}}_{s=0}\binom{r_{1}}{s}\binom{\xi^{2}_{2}+k_{2}-1}{p}\\
\nonumber
&\times \scriptstyle \frac{\Xi(i_{1})\Xi(i_{2})(\xi^{2}_{1}+k_{1}-1)!\sqrt{\pi}4^{p-s+1}\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}\right)^{p-s+r_{1}-\xi^{2}_{1}-k_{1}+1}\Gamma(\xi^{2}_{2}+k_{2}+r_{1}+p-s+P+1)\Gamma(\xi^{2}_{2}+k_{2}+r_{1}-p+s+P-1)}{r_{1}!\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}+\frac{c_{i_{2}}}{A_{0_{2}}\bar{\gamma}_{2}}+Q+2\sqrt{\frac{c_{i_{1}}c_{i_{2}}}{A_{0_{1}}A_{0_{2}}\bar{\gamma}_{1}\bar{\gamma}_{2}}}\right)^{\xi^{2}_{2}+k_{2}+r_{1}+p-s+P+1}\Gamma\left(\xi^{2}_{2}+k_{2}+r_{1}+P+\frac{1}{2}\right)}\\
&\times \scriptstyle {}_2F_1\left(\xi^{2}_{2}+k_{2}+r_{1}+p-s+P+1,p-s+\frac{3}{2};\xi^{2}_{2}+k_{2}+r_{1}+P+\frac{1}{2};\frac{\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}+\frac{c_{i_{2}}}{A_{0_{2}}\bar{\gamma}_{2}}+Q-2\sqrt{\frac{c_{i_{1}}c_{i_{2}}}{A_{0_{1}}A_{0_{2}}\bar{\gamma}_{1}\bar{\gamma}_{2}}}}{\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}+\frac{c_{i_{2}}}{A_{0_{2}}\bar{\gamma}_{2}}+Q+2\sqrt{\frac{c_{i_{1}}c_{i_{2}}}{A_{0_{1}}A_{0_{2}}\bar{\gamma}_{1}\bar{\gamma}_{2}}}}\right)
\label{aberafcsi}\end{aligned}$$
The desired expression directly arises by setting $q=0$ in , using , and utilizing [@tables Eq. (6.621.3)].
The ABER of the end-to-end SNR for a dual-hop relayed transmission with a fixed gain is derived in a closed-form expression as $$\begin{aligned}
\nonumber
&\bar{P_{b}}_{\gamma_{\rm Fixed}}=\scriptstyle \frac{1}{2}-\sum^{L}_{i_{1}=1}\sum^{b_{i_{1}}-\xi^{2}_{1}-1}_{k_{1}=0}\sum^{\xi^{2}_{1}+k_{1}-1}_{r_{1}=0}\sum^{L}_{i_{2}=1}\sum^{b_{i_{2}}-\xi^{2}_{2}-1}_{k_{2}=0}\sum^{r_{1}}_{s=0}\binom{r_{1}}{s}\Xi(i_{1})\Xi(i_{2})\\
\nonumber
&\times \scriptstyle \frac{(\xi^{2}_{1}+k_{1}-1)!\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}\right)^{\frac{\xi^{2}_{2}+k_{2}-s-1}{2}+r_{1}-\xi^{2}_{1}-k_{1}}U^{\frac{\xi^{2}_{2}+k_{2}+s}{2}}Q^{P}\Gamma(\xi^{2}_{2}+k_{2}-s+r_{1}+P)\Gamma(r_{1}+P)}{2 r_{1}!\Gamma(P)\left(\frac{c_{i_{2}}}{A_{0_{2}}\bar{\gamma}_{2}}\right)^{\frac{\xi^{2}_{2}+k_{2}-s+1}{2}}\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}+Q\right)^{\frac{\xi^{2}_{2}+k_{2}-s+2r_{1}+2P-1}{2}}\exp\left(-\frac{c_{i_{1}}c_{i_{2}}}{2 A_{0_{1}}A_{0_{2}}\bar{\gamma}_{1}\bar{\gamma}_{2}\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}+Q\right)}\right)}\\
&\times \scriptstyle W_{-\frac{(\xi^{2}_{2}+k_{2}-s+2 r_{1}+2 P-1)}{2},\frac{\xi^{2}_{2}+k_{2}-s}{2}}\left(\frac{c_{i_{1}}c_{i_{2}}}{A_{0_{1}}A_{0_{2}}\bar{\gamma}_{1}\bar{\gamma}_{2}\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}+Q\right)}\right)
\label{aberaffixed}\end{aligned}$$
The closed-form ABER directly arises by inserting into , implementing a change of variables $\sqrt{x}\rightarrow x$ in the resultant expression and utilizing [@tables Eq. (6.631.3)].
The ABER of the end-to-end SNR for a dual-hop DF relayed transmission is presented as $$\begin{aligned}
\nonumber
\bar{P_{b}}_{\gamma_{\rm DF}}=&\frac{1}{2}-\sum^{L}_{i_{1}=1}\sum^{b_{i_{1}}-\xi^{2}_{1}-1}_{k_{1}=0}\sum^{\xi^{2}_{1}+k_{1}-1}_{r_{1}=0}\sum^{L}_{i_{2}=1}\sum^{b_{i_{2}}-\xi^{2}_{2}-1}_{k_{2}=0}\sum^{\xi^{2}_{2}+k_{2}-1}_{r_{2}=0}\prod^{2}_{j=1}\left[\frac{\Xi(i_{j})(\xi^{2}_{j}+k_{j}-1)!}{r_{j}!\left(\frac{c_{i_{j}}}{A_{0_{j}}\bar{\gamma}_{j}}\right)^{\xi^{2}_{j}+k_{j}-r_{j}}}\right] \\
&\times \frac{Q^{P}\Gamma(r_{1}+r_{2}+P)}{2 \Gamma(P)\left(\frac{c_{i_{1}}}{A_{0_{1}}\bar{\gamma}_{1}}+\frac{c_{i_{2}}}{A_{0_{2}}\bar{\gamma}_{2}}+Q\right)^{r_{1}+r_{2}+P}}
\label{aberdf}\end{aligned}$$
The desired result trivially follows by evaluating the integral that arises by inserting into .
Numerical Results {#Numerical Results}
=================
In this Section, the accuracy of the proposed approach is numerically verified. For the sake of clarity and without loss of generality, the analytical results are cross-compared with numerical results modeled by simulating Gamma-Gamma faded channels. Hence, $\alpha_{j}$ and $\beta_{j}$ denote the small- and large-scale channel fading parameters of the $j^{\rm th}$ transmission hop, respectively. Recall that $b_{i_{j}}=\min\{\alpha_{j},\beta_{j}\}$. Also, $L=10$, $r/w_{z}=0.1$, and $[P,Q]=[0.5,1]$ (i.e., the coherent BPSK modulation scheme is considered). The considered outage threshold is $\gamma_{\rm th}=0$dB (i.e., ensuring a minimum target data rate of 1 bps/Hz). Curve-lines and circle-marks stand for the analytical and simulation results, respectively.
In Figs. \[fig1\] and \[fig2\], the outage and ABER performance is, respectively, illustrated for the considered dual-hop relay schemes. The CSI-assisted AF relaying outperforms the AF scheme with a fixed gain, while DF outperforms both AF schemes, as expected. Insightfully, the impact of the turbulence-induced fading severity is more critical to the system performance than the pointing error effect. Furthermore, when $\min\{\alpha_{j},\beta_{j}\}> \xi^{2}_{j}$, the analytical results perfectly match the corresponding simulation points in the entire SNR regime. In the case when $\min\{\alpha_{j},\beta_{j}\}\leq \xi^{2}_{j}$, the analytical results closely bound the simulations, while the tightness of this bound is more emphatic for low-to-moderate SNR regions. Thus, the remark of is verified. It is also noteworthy that the diversity order is the same for all the relay schemes and is related to the minimum of the involved channel fading parameters [@j:FaridHranilovic2012; @j:YangGaoAlouini2014].
Conclusion {#Conclusion}
==========
The performance of all-optical dual-hop relayed FSO systems was analytically studied and evaluated in the entire SNR regime. Three popular relaying transmission protocols were adopted, i.e., the CSI-assisted AF, AF with a fixed gain and DF relaying schemes. The joint effect of turbulence- and misalignment-induced channel fading was considered. Important system performance metrics were derived in new closed formulations, i.e., the system outage probability and ABER. Finally, a numerical verification of the analytical results was implemented, manifesting the efficiency of the proposed approach.
[^1]: In the AF relaying case, where there is no optical-to-electrical conversion at the relay node, $\eta_{2}=1$ [@j:YangGaoAlouini2014].
[^2]: The sharpness of this bound is manifested when $x\rightarrow +\infty$, such that $\Gamma(-j,x)\rightarrow \exp(-x)x^{-j-1}$, which reflects low-to-moderate SNR regions according to . It is noteworthy that the latter SNR regions are of particular interest in current study, since analytical results only for the asymptotically high SNR region exist so far.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We introduce geometric measures of entanglement for indistinguishable particles, which apply to mixed states, multipartite systems, and arbitrary dimensions. They are based on generalized (i.e., not necessarily finite) norms on the set of quantum states and lead to the first necessary and sufficient computational separability criterion in this general setting. The coherent approach developed in the paper allows us to compare, in particular, entanglement for fermionic and distinguishable particles: The entanglement measure for fermionic particles coincides with the corresponding entanglement measure for distinguishable particles up to a factor of $k!$ where $k$ is the number of particles involved. By this result the amount of entanglement emerging from fermi statistics alone is clearly separated from the overall amount of entanglement. Finally, our techniques are applied to entanglement related to Schmidt and Slater numbers.'
author:
- |
Florian Sokoli, Burkhard Kümmerer\
Fachbereich Mathematik, Technische Universität Darmstadt
title: Entanglement of Indistinguishable Particles and its Quantification
---
Introduction
============
Entanglement of indistinguishable particles has gained an increasing amount of interest during the last decade. Due to the different nature of fermionic and bosonic quantum systems in comparison to distinguishable particles the conventional approach to entanglement needs to be modified considerably. In particular, entanglement of fermionic particles has been be considered as being ”unphysical” and ”useless” for some time. Meanwhile, the conceptional problem of defining entanglement of indistinguishable particles in an appropriate way has widely been solved [@Benatti; @Eckert; @Ghi02; @Grabowski; @Li; @Schliemann; @Tichy].\
However, the possibilities for detecting and quantifying this kind of entanglement are still rather limited and many questions are open.\
Schliemann et al. presented the first characterization of entanglement for quantum states of indistinguishable particles. They studied entanglement of mixed states of bipartite quantum systems using a *correlation measure*, also called *Schliemann concurrence* [@Eckert; @Schliemann]. A more general and widely used approach is to consider the reduced one-particle-state of a given multipartite state of indistinguishable particles. In this context several entanglement measures have been proposed such as the *purity* or the *von Neumann entropy* of the reduced state [@Balachandran; @Ghi04; @Ghi05; @Iemini2; @Paskauskas; @Plastino; @Zander]. It can shown that these techniques are sufficient in order to characterize entanglement of pure fermionic states [@Plastino]. Moreover, measures based on a *shifted negativity* [@Iemini2] and *Rényi entropies* [@Zander] have been used in this context as well. Lévay and Vrana [@Levay] have introduced an algebraic measure for tripartite fermionic entanglement as well as a separability criterion for pure multipartite states based on methods from algebraic geometry. Finally, Iemini et al. [@Iemini1] have considered *entanglement witnesses* in order to obtain an entanglement measure called *generalized robustness*. However, up to now there is no necessary and sufficient separability criterion for indistinguishable particles which works for arbitrary mixed and multipartite states in arbitrary dimensions. The present paper contributes to this problem by introducing measures for fermionic and bosonic entanglement in this general setting which lead to necessary and sufficient separability criteria.\
\
Comparing the various definitions of entanglement for distinguishable, bosonic and fermionic particles one realizes a great deal of common mathematical structure which indicates the possibility of a more abstract and unified treatment. In a groundbreaking work [@Arveson] Arveson provided such a theory of generalized entanglement which allows a coherent approach to the detection and quantification of all different kinds of quantum entanglement at once. The general idea is to use *Minkowski functionals* with respect to convex balanced sets as entanglement measures. It turns out that they can be interpreted as *generalized norms* in the sense that all properties of a norm are satisfied except that they might be infinite for infinite dimensional Hilbert spaces. Although these concepts are already very general they still need to be adapted in order to control entanglement of quantum states restricted to linear subspaces. We therefore develop a slight generalization of Arveson’s concepts which, in particular, leads to a simper and easier-to-handle theory.\
A widely used treatment of entanglement of fermionic particles is to consider them as being distinguishable and to study their entanglement in the ”classical” way as for distinguishable particles. Our treatment of entanglement allows us to provide a mathematical justification of this approach. We show that the entanglement measure for fermions coincides with the corresponding entanglement measure for distinguishable particles up to a state-independent factor of $k!$ where $k$ equals the number of particles involved. Moreover, the amount of ”trivial” entanglement, i.e. entanglement due to antisymmetrization is clearly separated from ”genuine” fermionic entanglement.\
\
This paper is organized as follows. In Section \[SN\] we describe our notational and terminological conventions. Section \[SA\] gives an introduction to relevant aspects of Arveson’s concepts with a focus on its geometrical background. Afterwards we describe our generalization of his methods which is used in the subsequent proof of the main result of Section \[SA\] on controling entanglement of states that are restricted to linear subspaces. In Section \[SB\] we use these abstract results in order to define measures for bosonic and fermionic entanglement. Moreover, we establish the aforementioned relation between the entanglement measures for fermionic and distinguishable particles. We extend these results in Section \[SC\] by introducing entanglement measures related to Schmidt and Slater numbers. For bipartite systems we derive similar quantitative relations between the measures for distinguishable and fermionic particles as in Section \[SB\]. Moreover, we generalize a result of Johnston [@Johnston] concerning the quantification of Schmidt number entanglement which was formulated for bipartite finite dimensional quantum systems.
Notation and Terminology {#SN}
========================
Throughout this paper $\mathcal{H}$ will always denote a (possibly infinite dimensional) complex Hilbert space. The scalar-product on $\mathcal{H}$ is denoted by $\langle\cdot,\cdot\rangle$ and assumed to be linear in the first and anti-linear in the second component. General vectors of $\mathcal{H}$ are typically denoted by $\xi,\eta$, and $\zeta$ whereas for vectors of orthogonal systems we use $e$ and $f$. The algebra of bounded linear operators on $\mathcal{H}$ is denoted by $\mathcal{B}(\mathcal{H})$. A continuous linear functional $\varphi$ on $\mathcal{B}(\mathcal{H})$ is called *normal* if there is a trace-class operator $\Phi$ on $\mathcal{H}$ such that $$\varphi(x)=\mbox{tr}(\Phi x)\quad\mbox{for all }x\in\mathcal{B}(\mathcal{H})\,.$$ The set of normal linear functionals on $\mathcal{B}(\mathcal{H})$ is denoted by $\mathcal{B}(\mathcal{H})_{\ast}$. If $\dim\,\mathcal{H}=n<\infty$ every linear functional on $\mathcal{B}(\mathcal{H})$ is normal and the algebra $\mathcal{B}(\mathcal{H})$ can be identified with the set of complex $n\times n$-matrices which we denote by $M_n({\mathbbmss{C}})$. A *state* on $\mathcal{B}(\mathcal{H})$ is a positive (hence continuous) linear functional of norm one. For a normal state $\varphi$ on $\mathcal{B}(\mathcal{H})$ its associated trace-class operator $\rho$ is referred to as *density operator*. In this paper we use linear functionals instead of density operators since it leads to simpler notations.\
For vectors $\xi,\eta\in\mathcal{H}$ the *rank-one-operator* $t_{\xi,\eta}$ is the linear operator on $\mathcal{H}$ defined by $$t_{\xi,\eta}\zeta:=\langle\zeta,\eta\rangle\xi\quad\mbox{for }\zeta\in\mathcal{H}$$ and the corresponding (normal) *vector functional* $\omega_{\xi,\eta}:\mathcal{B}(\mathcal{H})\rightarrow{\mathbbmss{C}}$ induced by $\xi,\eta$ is given by $$\omega_{\xi,\eta}(x):=\langle x\xi,\eta\rangle=\mbox{tr}(t_{\xi,\eta}x)\quad\mbox{for }x\in\mathcal{B}(\mathcal{H})\,.$$ Occasionally we write $t_{\xi}$ and $\omega_{\xi}$ for $t_{\xi,\xi}$ and $\omega_{\xi,\xi}$ respectively.\
The unit ball and the unit sphere of a complex normed space $X$ are denoted by $B_{1,X}$ and $S_{1,X}$ respectively. A set $M\subset X$ is called *balanced* if for all $x\in M$ and all numbers $\lambda\in{\mathbbmss{C}}$ with $|\lambda|=1$ we have $\lambda x\in M$. The convex hull of $M$ is denoted by $\mbox{co}(M)$ and we write $\overline{\mbox{co}}^{{\left\|{\cdot}\right\|}}(M)$ for its norm closure. More generally, for an arbitrary topology $\mathcal{T}$ on $X$ the $\mathcal{T}$-closure of $\mbox{co}(M)$ is denoted by $\overline{\mbox{co}}^{\mathcal{T}}(M)$.\
We say that a set of continuous linear functionals $N$ on $X$ is *separating* or *separates points of $X$* if for every $x\in X$ the relation $$\varphi(x)=0\quad\mbox{for all }\varphi\in N$$ already implies $x=0$. In particular, a subset $N\subset \mathcal{H}$ is separating if for every $\xi\in\mathcal{H}$ the condition $$\langle\xi,\eta\rangle=0\quad\mbox{for all }\eta\in N$$ implies $\xi=0.$
Abstract Theory {#SA}
===============
In this section we give a brief introduction to Arveson’s entanglement theory and generalize it according to our needs. Theorem \[TA\], the main result of this section, prepares grounds for comparing entanglement of distinguishable and fermionic particles in the sections \[SB\] and \[SC\].
\[D4\] Let $V\subset \mathcal{H}$ be a balanced set of unit vectors that separates points of $\mathcal{H}$ and $$K:=\overline{{\mathrm{co}}}^{{\left\|{\cdot}\right\|}}\{\omega_{\xi,\eta}:\,\xi,\eta\in V\}\subset\mathcal{B}(\mathcal{H})_{\ast}\,.$$ Moreover, we define functions on $\mathcal{B}(\mathcal{H})$ and $\mathcal{B}(\mathcal{H})_{\ast}$ by $${\left\|{x}\right\|}_{K}:=\sup{\left\{{|\psi(x)|:\,\psi\in K}\right\}}\quad \mbox{for }x\in\mathcal{B}(\mathcal{H})$$ and $$q_K(\psi):=\inf{\left\{{\alpha\geq 0:\,\psi\in\alpha K}\right\}}\quad\mbox{for }\psi\in\mathcal{B}(\mathcal{H})_{\ast}$$ respectively. For the latter we use the convention that $q_K(\psi)=\infty$ whenever the set ${\left\{{\alpha\geq 0:\,\psi\in\alpha K}\right\}}$ is empty.
The function $q_K$ is called *Minkowski functional* associated to the set $K$ and quantifies the distance of a point $\psi\in\mathcal{B}(\mathcal{H})_{\ast}$ to $K$ along rays passing through the origin. In the case that $\mathcal{H}$ is a tensor product of Hilbert spaces appropriate choices of $V$ will allow to quantify multipartite entanglement in many different scenarios.
\[PA\] With $K$ and its associated functions ${\left\|{\cdot}\right\|}_K$ and $q_K$ as defined above, we have the following properties:
1. \[PA1\] The set $K$ is a norm-closed convex and balanced subset of the unit ball of $\mathcal{B}(\mathcal{H})_{\ast}$ that separates points of $\mathcal{B}(\mathcal{H})$. In particular, we have $0\in K$ and $${\left\|{\psi}\right\|}\leq q_K(\psi)\quad\mbox{for all }\psi\in\mathcal{B}(\mathcal{H})_{\ast}\,.$$ Moreover, $K$ equals the ”unit ball” of $q_K$, i.e., $$K={\left\{{\psi\in\mathcal{B}(\mathcal{H})_{\ast}:\,q_K(\psi)\leq 1}\right\}}\,.$$
2. \[PA2\] The function ${\left\|{\cdot}\right\|}_K$ is a norm on $\mathcal{B}(\mathcal{H})$ such that ${\left\|{x}\right\|}_K\leq{\left\|{x}\right\|}$ for all $x\in\mathcal{B}(\mathcal{H})$ and we have $${\left\|{x}\right\|}_K=\sup{\left\{{|\omega_{\xi,\eta}(x)|:\,\xi,\eta\in V}\right\}}=\sup{\left\{{|\langle x\xi,\eta\rangle|:\,\xi,\eta\in V}\right\}}\,.$$
3. \[PA3\] The Minkowski functional $q_K$ has the alternative representation $$q_K(\psi)=\sup{\left\{{|\psi(x)|:\,x\in\mathcal{B}(\mathcal{H}),\,{\left\|{x}\right\|}_K\leq 1}\right\}}\quad\mbox{for all }\psi\in\mathcal{B}(\mathcal{H})_{\ast}\,.$$
$\,$
1. The set $K$ is convex and norm-closed by definition. In order to show that $K$ separates points of $\mathcal{B}(\mathcal{H})$ it is sufficient to show that the set $${\left\{{\omega_{\xi,\eta}:\,\xi,\eta\in V}\right\}}\subset K$$ has this property. Therefore, suppose that $\omega_{\xi,\eta}(x)=0$ for all $\xi,\eta\in V$. Then $$\langle x\xi,\eta\rangle=0\quad\mbox{for all }\xi,\eta\in V$$ and therefore $x\xi=0$ for all $\xi\in V$ since $V$ separates points of $\mathcal{H}$. For every $\zeta\in\mathcal{H}$ and $\xi\in V$ it follows $$0=\langle x\xi,\zeta\rangle=\langle\xi,x^{\ast}\zeta\rangle\,.$$ Hence, $x^{\ast}\zeta=0$ for all $\zeta\in\mathcal{H}$ which implies $$x=x^{\ast\ast}=0^{\ast}=0\,.$$ Since $${\left\|{\omega_{\xi,\eta}}\right\|}={\left\|{\xi}\right\|}\cdot{\left\|{\eta}\right\|}=1\mbox{ for all }\xi,\eta\in V$$ we have $$K=\overline{\mbox{co}}^{{\left\|{\cdot}\right\|}}{\left\{{\omega_{\xi,\eta}:\,\xi,\eta\in V}\right\}}\subset B_{1,\mathcal{B}(\mathcal{H})_{\ast}}\,.$$ It is a well-known fact (compare for example Section I.3 of [@Defant]) that for every $\varepsilon>0$ we have $$K\subset {\left\{{\varphi\in\mathcal{B}(\mathcal{H})_{\ast}:\,q_K(\varphi)\leq 1}\right\}}\subset (1+\varepsilon)\cdot K$$ and using that $K$ is norm-closed it easily follows that $K$ equals the unit ball of $q_K$.
2. For all $x\in\mathcal{B}(\mathcal{H})$ the mapping $$\mathcal{B}(\mathcal{H})_{\ast}\ni\psi\mapsto |\psi(x)|\in{\mathbbmss{C}}$$ is convex and norm-continuous so that $$\begin{aligned}
{\left\|{x}\right\|}_K&=&\sup{\left\{{|\psi(x)|:\,\psi\in\overline{\mbox{co}}^{{\left\|{\cdot}\right\|}}{\left\{{\omega_{\xi,\eta}:\,\xi,\eta\in V}\right\}}}\right\}}\\
&=&\sup{\left\{{|\omega_{\xi,\eta}(x)|:\,\xi,\eta\in V}\right\}}\,.\end{aligned}$$ The remaining properties of ${\left\|{\cdot}\right\|}_K$ are obvious.
3. Using the concept of the *polar* $K^{\circ}\subset\mathcal{B}(\mathcal{H})$ of $K$ with respect to the dual pair $(\mathcal{B}(\mathcal{H})_{\ast},\mathcal{B}(\mathcal{H}))$ (compare for example Section V.1 of [@Convay]) we have $$\begin{aligned}
K^{\circ}&=&{\left\{{x\in\mathcal{B}(\mathcal{H}):\,|\varphi(x)|\leq 1\mbox{ for all }\varphi\in K}\right\}}\\
&=&{\left\{{x\in\mathcal{B}(\mathcal{H}):\,\sup{\left\{{|\varphi(x)|:\,\varphi\in K}\right\}}\leq 1}\right\}}\\
&=&{\left\{{x\in\mathcal{B}(\mathcal{H}):\,{\left\|{x}\right\|}_K\leq 1}\right\}}\\
&=&B_{1,{\left\|{\cdot}\right\|}_K}\,.\end{aligned}$$ The *bipolar* $K^{\circ\circ}\subset\mathcal{B}(\mathcal{H})_{\ast}$ of $K$ is then given by $$\begin{aligned}
K^{\circ\circ}&=&{\left\{{\varphi\in\mathcal{B}(\mathcal{H})_{\ast}:\,|\varphi(x)|\leq 1\mbox{ for all }x\in K^{\circ}}\right\}}\\
&=&{\left\{{\varphi\in\mathcal{B}(\mathcal{H})_{\ast}:\,\sup{\left\{{|\varphi(x)|:\,x\in K^{\circ}}\right\}}\leq 1}\right\}}\\
&=&{\left\{{\varphi\in\mathcal{B}(\mathcal{H})_{\ast}:\,\sup{\left\{{|\varphi(x)|:\,x\in \mathcal{B}(\mathcal{H}),\,{\left\|{x}\right\|}_K\leq 1}\right\}}\leq 1}\right\}}\,.\end{aligned}$$ Using the fact that the $\sigma(\mathcal{B}(\mathcal{H})_{\ast},\mathcal{B}(\mathcal{H}))$-closure of a convex subset of $\mathcal{B}(\mathcal{H})_{\ast}$ equals its norm closure the Bipolar Theorem and $(i)$ imply $$K\stackrel{(i)}{=}\overline{\mbox{co}}^{{\left\|{\cdot}\right\|}}(K\cup{\left\{{0}\right\}})=\overline{\mbox{co}}^{\sigma(\mathcal{B}(\mathcal{H})_{\ast},\mathcal{B}(\mathcal{H}))}(K\cup{\left\{{0}\right\}})=K^{\circ\circ}\,.$$ Hence $K$ equals the unit ball of the function $$f(\varphi):=\sup{\left\{{|\varphi(x)|:\,x\in \mathcal{B}(\mathcal{H}),\,{\left\|{x}\right\|}_K\leq 1}\right\}}\quad\mbox{for }\varphi\in\mathcal{B}(\mathcal{\varphi})_{\ast}\,.$$ Therefore, since the functions $q_K$ and $f$ possess the same unit ball and are homogeneous for positive scalars it follows $q_K=f$.
Note that from the representation of $q_K$ according to Proposition \[PA\].(\[PA3\]) it easily follows that $q_K$ satisfies the triangle inequality. Hence, $q_K$ has all characteristic properties of a norm except that it may attain the value ”$\infty$”. We may therefore speak of a *generalized norm*.\
The above construction slightly differs from Arveson’s approach. In his publication [@Arveson] the unital $\mbox{C}^{\ast}$-algebra $$\mathcal{A}:=\mathcal{K}(\mathcal{H})+{\mathbbmss{C}}\mathbbmss{1}$$ is used in order to define a function $E:\mathcal{B}(\mathcal{H})_{\ast}\rightarrow[0,\infty]$ given by $$E(\psi):=\sup{\left\{{|\psi(x)|:\,x\in\mathcal{A},\,{\left\|{x}\right\|}_K\leq 1}\right\}}\quad\mbox{for }\psi\in\mathcal{B}(\mathcal{H})_{\ast}\,.$$ Hence, comparing $q_K$ with $E$ we see that for $q_K$ the supremum according to Proposition \[PA\].(\[PA3\]) runs over $\mathcal{B}(\mathcal{H})$ whereas in the case of $E$ the supremum is restricted to the $\mbox{C}^{\ast}$-algebra $\mathcal{A}$. As a matter of fact both functions $E$ and $q_K$ can detect separability equally well.
\[TB\] Let $\varphi$ be a normal state on $\mathcal{B}(\mathcal{H})$. Then the following conditions are equivalent:
1. $\varphi\in\overline{\mathrm{co}}^{{\left\|{\cdot}\right\|}}\{\omega_{\xi}:\,\xi\in V\}$.
2. $E(\varphi)=1$.
3. $q_K(\varphi)=1$.
The equivalence of (i) and (ii) has been proven by Arveson (compare Theorem 6.2 in [@Arveson]). A careful inspection of his proof reveals that the only relevant property of the algebra $\mathcal{A}$ is the fact that it containes the unit operator (which guarantees the set of states on $\mathcal{A}$ to be compact in its $\sigma^{\ast}$-topology) as well as the set of compact operators on $\mathcal{H}$. Both conditions are met by the $\mbox{C}^{\ast}$-algebra $\mathcal{B}(\mathcal{H})$ as well so that the equivalence can be proven literary.
Note that according to the proof of Theorem \[TB\] *any* $\mbox{C}^{\ast}$-subalgebra of $\mathcal{B}(\mathcal{H})$ containing the set of compact operators and the unit operator induces an (abstract) entanglement measure with essentially the same structure as $E$ and $q_K$. Therefore, it is justified to speak of a generalization of Arveson’s result.\
The equivalence of (i) and (ii) in the above theorem is the basis of Arveson’s entanglement quantification. However, due to the fact that these statements are equivalent to (iii) as well both functions $E$ and $q_K$ may serve as entanglement measures. We will use the latter since our main result Theorem \[TA\] needs to be formulated in terms of $q_K$.\
Now suppose we are given a nonzero orthogonal projection $p\in\mathcal{B}(\mathcal{H})$ and a normal state $\varphi$ on $\mathcal{B}(\mathcal{H})$ such that $$\varphi(x)=\varphi(pxp)\quad\mbox{for all }x\in\mathcal{B}(\mathcal{H})\,.$$ Since $\varphi(p)=\varphi(p\mathbbmss{1}p)=\varphi(\mathbbmss{1})=1$ we may identify $\varphi$ with a normal state on the algebra $\mathcal{B}(p\mathcal{H})$. On the other hand, projecting the set $V$ to the subspace $p\mathcal{H}$ (along with some further operations) gives rise to a convex and balanced subset $K_p\subset\mathcal{B}(p\mathcal{H})_{\ast}$ according to Definition \[D4\]. Therefore, it is a natural question to ask as to how the values $q_K(\varphi)$ and $q_{K_p}(\varphi)$ are related to each other. Theorem \[TA\] gives a very general answer to that question. In the remainder of this section we will discuss it in detail and to this end we introduce the following notation.
\[D5\] Let $0\neq p\in\mathcal{B}(\mathcal{H})$ be an orthogonal projection. For a number $\lambda\geq 1$ we define subsets of $p\mathcal{H}$ and $\mathcal{B}(p\mathcal{H})_{\ast}$ by $$V_{p,\lambda}:=\lambda pV\cap S_{1,\mathcal{H}}$$ and $$K_{p,\lambda}:=\overline{\mathrm{co}}^{{\left\|{\cdot}\right\|}}\{\omega_{\xi,\eta}:\,\xi,\eta\in V_{p,\lambda}\}$$ respectively. If $V_{p,\lambda}$ satisfies the condition that for all $\xi\in V$ with $p\xi\neq 0$ the relation $$\frac{p\xi}{{\left\|{p\xi}\right\|}}\in V_{p,\lambda}\,,$$ is fulfilled we say that the projection $p$ is *compatible* with the pair $(V,\lambda)$.
Note that the set $V_{p,\lambda}$ as defined above may be empty. In particular, it is not necessarily a separating subset of $p\mathcal{H}$. However, it is always balanced since this property is inherited from $V$.\
We will see in the sections \[SB\] and \[SC\] that the conditions formulated in the following theorem are met by important physical scenarios although this may seem very unlikely on the first glance.
\[TA\] For a nonzero orthogonal projection $p\in\mathcal{B}(\mathcal{H})$ and a number $\lambda\geq 1$ let $V_{p,\lambda}$ and $K_{p,\lambda}$ be the sets introduced in Definition \[D5\].
1. \[TA1\] Suppose that $V_{p,\lambda}$ separates points of $p\mathcal{H}$ and that the mapping $$\mathcal{B}(\mathcal{H})\ni y\mapsto pyp\in\mathcal{B}(\mathcal{H})$$ is a ${\left\|{\cdot}\right\|}_{K}$-contraction. Identifying $\mathcal{B}(p\mathcal{H})$ with the subalgebra $p\mathcal{B}(\mathcal{H})p\subset\mathcal{B}(\mathcal{H})$ we have the following inequalities:
1. ${\left\|{x}\right\|}_{K_{p,\lambda}}\leq\lambda^2\cdot{\left\|{x}\right\|}_{K}\mbox{ for all }x\in\mathcal{B}(p\mathcal{H})\,.$
2. \[TA12\] $q_{K_{p,\lambda}}(\psi)\geq\frac{1}{\lambda^2}\cdot q_{K}(\psi)\mbox{ for all }\psi\in\mathcal{B}(p\mathcal{H})_{\ast}\,.$
2. \[TA2\] If $p$ is compatible with the pair $(V,\lambda)$ then $V_{p,\lambda}$ separates points of $p\mathcal{H}$. Moreover, defining $$\mu:=\sup{\left\{{r\geq 1:\,rpV\in B_{1,\mathcal{H}}}\right\}}$$ we have:
1. ${\left\|{x}\right\|}_{K_{p,\lambda}}\geq\mu^2\cdot{\left\|{x}\right\|}_{K}\mbox{ for all }x\in\mathcal{B}(p\mathcal{H})$.
2. \[TA22\] $q_{K_{p,\lambda}}(\psi)\leq\frac{1}{\mu^2}\cdot q_{K}(\psi)\mbox{ for all }\psi\in\mathcal{B}(\mathcal{H})_{\ast}\,.$
$\,$
1. By the construction of $V_{p,\lambda}$ we obtain $$\begin{aligned}
{\left\|{x}\right\|}_{K_{p,\lambda}}&=&\sup{\left\{{\left|\left\langle x\xi,\eta\right\rangle\right|:\,{\xi,\eta\in V_{p,\lambda}}}\right\}}\\
&=&\lambda^2\cdot\sup{\left\{{\left|\left\langle \underbrace{pxp}_{=x}\xi,\eta\right\rangle\right|:\,\xi,\eta\in V,\,{\left\|{\lambda p\xi}\right\|}={\left\|{\lambda p\eta}\right\|}=1}\right\}}\\
&\leq&\lambda^2\cdot\sup{\left\{{\left|\left\langle x\xi,\eta\right\rangle\right|:\,{\xi,\eta\in V}}\right\}}=\lambda^2\cdot{\left\|{x}\right\|}_{K}\end{aligned}$$ which proves inequality (a) (note that we did not use the contraction condition imposed on ${\left\|{\cdot}\right\|}_K$).\
In order to prove (b) we first consider the case $q_{K}(\psi)=\infty$. Then by Proposition \[PA\].(\[PA3\]) for every $C>0$ there is an element $x\in \mathcal{B}(\mathcal{H})$ with ${\left\|{x}\right\|}_{K}\leq 1$ such that $|\psi(x)|\geq C$. Inequality (a) and the assumption about ${\left\|{\cdot}\right\|}_{K}$ then imply $${\left\|{pxp}\right\|}_{K_{p,\lambda}}\leq\lambda^2\cdot{\left\|{pxp}\right\|}_{K}\leq \lambda^2\cdot{\left\|{x}\right\|}_{K}\leq\lambda^2\,.$$ Using $\psi(x)=\psi(pxp)$ for all $x\in\mathcal{B}(\mathcal{H})$ a further application of Proposition \[PA\].(\[PA3\]) to $q_{K_{p,\lambda}}$ gives $$q_{K_{p,\lambda}}(\psi)\geq\left|\psi\left(\frac{pxp}{\lambda^2}\right)\right|=\frac{1}{\lambda^2}|\psi(x)|\geq\frac{C}{\lambda^2}\,,$$ that is $q_{K_{p,\lambda}}(\psi)=\infty$.\
On the other hand, if $q_{K}(\psi)<\infty$, Proposition \[PA\].(\[PA3\]) allows to choose for every $\varepsilon>0$ an element $x\in \mathcal{B}(\mathcal{H})$ with ${\left\|{x}\right\|}_{K}\leq 1$ such that $$q_{K}(\psi)-\varepsilon<|\psi(x)|\,.$$ By the same computation as above we have ${\left\|{pxp}\right\|}_{K_{p,\lambda}}\leq\lambda^2$ and it follows $$|\psi(x)|=\lambda^2\cdot\left|\psi\left(\frac{pxp}{\lambda^2}\right)\right|\leq\lambda^2\cdot\sup{\left\{{|\psi(y)|:\,y\in \mathcal{B}(p\mathcal{H}),\,{\left\|{y}\right\|}_{K_{p}}\leq 1}\right\}}=\lambda^2\cdot q_{K_{p,\lambda}}({\psi})$$ and therefore $$\frac{1}{\lambda^2}\cdot q_{K}(\psi)<\frac{\varepsilon}{\lambda^2}+q_{K_{p,\lambda}}(\psi)\,.$$ Hence, $\frac{1}{\lambda^2}\cdot q_{K}(\psi)\leq q_{K_{p,\lambda}}(\psi)$.
2. Let $\xi\in p\mathcal{H}$ such that $\langle\xi,\eta\rangle=0$ for all $\eta\in V_{p,\lambda}$. For $\zeta\in V$ with $p\zeta\neq 0$ it follows from our assumption $$\langle\xi,\zeta\rangle=\langle p\xi,\zeta\rangle=\langle\xi,p\zeta\rangle={\left\|{p\zeta}\right\|}\cdot\left\langle\xi,\frac{p\zeta}{{\left\|{p\zeta}\right\|}}\right\rangle=0$$ and for $p\zeta=0$ we find analogously that $\langle\xi,\zeta\rangle=0$. Since $V$ separates points of $\mathcal{H}$ this implies $\xi=0$. Hence, $V_{p,\lambda}$ separates points of $p\mathcal{H}$.\
In order to prove (a) we may assume that $x\neq 0$. Using Proposition \[PA\].(\[PA2\]) for ${\left\|{x}\right\|}_{K}>\varepsilon>0$ we can choose $\xi,\eta\in V$ such that $${\left\|{x}\right\|}_{K}-\varepsilon<\left|\left\langle x\xi,\eta\right\rangle\right|\,.$$ Then we must have $$0\neq|\langle x\xi,\eta\rangle|=|\langle xp\xi,p\eta\rangle|$$ and, in particular, $p\xi,p\eta\neq 0$. By the choice of $\mu$ and the condition imposed on $V_{p,\lambda}$ we obtain $$\begin{aligned}
{\left\|{x}\right\|}_K-\varepsilon&<&\left|\left\langle x\xi,\eta\right\rangle\right|
=\frac{1}{\mu^2}\cdot\underbrace{{\left\|{\mu p\xi}\right\|}}_{\leq 1}\cdot\underbrace{{\left\|{\mu p\eta}\right\|}}_{\leq 1}\cdot\left|\left\langle x\frac{p\xi}{{\left\|{p\xi}\right\|}},\frac{p\eta}{{\left\|{p\eta}\right\|}}\right\rangle\right|\\
&\leq&\frac{1}{\mu^2}\cdot\sup{\left\{{|\langle x\xi',\eta'\rangle|:\,\xi',\eta'\in V_{p,\lambda}}\right\}}=\frac{1}{\mu^2}{\left\|{x}\right\|}_{K_{p,\lambda}}\,.\end{aligned}$$ It follows $$\mu^2\cdot{\left\|{x}\right\|}_{K}<\mu^2\cdot\varepsilon+{\left\|{x}\right\|}_{K_{p,\lambda}}\,;$$ that is $\mu^2\cdot{\left\|{x}\right\|}_{K}\leq{\left\|{x}\right\|}_{K_{p,\lambda}}$. By Proposition \[PA\].(\[PA3\]) the inequality (b) now follows from (a) via $$\begin{aligned}
q_{K_{p,\lambda}}(\psi)&=&\sup{\left\{{|\psi(x)|:\,x\in \mathcal{B}(p\mathcal{H}),\,{\left\|{x}\right\|}_{K_{p,\lambda}}\leq 1}\right\}}\\
&\stackrel{\mbox{\footnotesize{(a)}}}{\leq}&\frac{1}{\mu^2}\cdot\sup{\left\{{|\psi(x)|:\,{x\in \mathcal{B}(p\mathcal{H})\,,{\left\|{x}\right\|}_{K}\leq 1}}\right\}}\\
&\leq&\frac{1}{\mu^2}\cdot\sup{\left\{{|\psi(x)|:\,{x\in \mathcal{B}(\mathcal{H})\,,{\left\|{x}\right\|}_{K}\leq 1}}\right\}}\\
&=&\frac{1}{\mu^2}\cdot q_{K}(\psi)\,.\end{aligned}$$
Note that if the conditions of Theorem \[TA\].(\[TA1\]) and (\[TA2\]) are both satisfied we have the chains of inequalities $$\mu^2\cdot{\left\|{x}\right\|}_K\leq{\left\|{x}\right\|}_{K_{p,\lambda}}\leq\lambda^2\cdot{\left\|{x}\right\|}_K\quad\mbox{for all }x\in\mathcal{B}(p\mathcal{H})$$ and $$\frac{1}{\lambda^2}\cdot q_K(\psi)\leq q_{K_{p,\lambda}}(\psi)\leq\frac{1}{\mu^2}q_K(\psi)\quad\mbox{for all }\psi\in\mathcal{B}(p\mathcal{H})_{\ast}\,.$$
Quantifying Multipartite Entanglement of Indistinguishable Particles {#SB}
====================================================================
For a number $k\in\mathbbmss{N}$ we denote the $k$-fold Hilbert space tensor product of $\mathcal{H}$ with itself by $\mathcal{H}^{\otimes k}$. The following notion is the starting point for the quantification of entanglement of distinguishable particles.
\[D1\] We denote the set of unit product vectors of $\mathcal{H}^{\otimes k}$ by $V^{\otimes k}$, i.e. $$V^{{\otimes k}}:={\left\{{\eta_1\otimes...\otimes\eta_k:\,\eta_i\in\mathcal{H},\,{\left\|{\eta_i}\right\|}=1}\right\}}\,.$$ A normal state $\varphi$ on $\mathcal{B}(\mathcal{H}^{\otimes k})$ is called *separable* if $$\varphi\in\overline{\mathrm{co}}^{{\left\|{\cdot}\right\|}}{\left\{{\omega_{\xi}:\,\xi\in V^{\otimes k}}\right\}}\,.$$ Otherwise, it is called *entangled*.
Noting that the set $V^{{\otimes k}}$ is a balanced subset of the unit sphere that separates points of $\mathcal{H}^{\otimes k}$ we obtain a necessary and sufficient separability criterion for normal states.
\[CA\] Let $K^{\otimes k}$ denote the norm-closed convex hull of the set $$\left\{\omega_{\xi,\eta}:\,\xi,\eta\in V^{\otimes k}\right\}\,.$$ Then every normal state $\varphi$ on $\mathcal{B}(\mathcal{H}^{\otimes k})$ satisfies $q_{K^{\otimes k}}(\varphi)\geq 1$ and $q_{K^{\otimes k}}(\varphi)=1$ if and only if $\varphi$ is separable.
The claim immediately follows from Proposition \[PA\].(\[PA1\]) and Theorem \[TB\].
It is shown in [@Arveson] that for finite dimensional Hilbert spaces Corollary \[CA\] is the *greatest cross-norm criterion* introduced by Rudolph [@Rudolph0; @Rudolph; @Rudolph2]. We will encounter such situations in the examples \[E1\]-\[E2\] below.\
In order to define separability for states of indistinguishable particles some conceptual and notational preparation is required.\
Let $\pi\in S_k$ be a permutation of $k$ points. We denote the associated permutation operator on $\mathcal{H}^{\otimes k}$ by $U_{\pi}$. The orthogonal projections onto the bosonic and fermionic subspaces are given by $$\begin{aligned}
P_{+}:=\frac{1}{k!}\sum_{\pi\in S_k}{U_{\pi}}\quad\mbox{and}\quad P_{-}:=\frac{1}{k!}\sum_{\pi\in S_k}{\mathrm{sign}(\pi)U_{\pi}}\\\end{aligned}$$ and we write $$\mathcal{H}^{\vee k}:=P_+\mathcal{H}^{\otimes k}\quad\mbox{and}\quad\mathcal{H}^{\wedge k}:=P_-\mathcal{H}^{\otimes k}\,.$$ Furthermore, for $\eta_1,...,\eta_k\in\mathcal{H}$ we define $$\bigvee_{i=1}^k{\eta_i}:=\eta_1\vee...\vee\eta_k:=P_+\eta_1\otimes...\otimes\eta_k$$ and $$\bigwedge_{i=1}^k{\eta_i}:=\eta_1\wedge...\wedge\eta_k:=\sqrt{k!}P_-\eta_1\otimes...\otimes\eta_k\,.$$
It is a well-known fact that for all $\eta_i\in\mathcal{H}$ and $\pi\in S_k$ we have $$U_{\pi}\left(\eta_1\vee...\vee\eta_k\right)=\eta_1\vee...\vee\eta_k\quad\mbox{and}\quad U_{\pi}\left(\eta_1\wedge...\wedge\eta_k\right)=\mbox{sign}(\pi)\eta_1\wedge...\wedge\eta_k\,.$$ Using the latter relation it is easy to prove that $$\langle\xi_1\wedge...\wedge\xi_k,\eta_1\wedge...\wedge\eta_k\rangle=\det\left(\begin{array}{ccc}\langle\xi_1,\eta_1\rangle&\cdots&\langle\xi_k,\eta_1\rangle\\\vdots&&\vdots\\\langle\xi_1,\eta_k\rangle&\cdots&\langle\xi_k,\eta_k\rangle\end{array}\right)$$ for all $\xi_i,\eta_i\in\mathcal{H}$. In particular, this implies $$\eta_1\wedge...\wedge\eta_k=0$$ whenever the vectors $\eta_1,...,\eta_k$ are linearly dependent.
\[LA\] Let $\eta_1,...,\eta_k\in\mathcal{H}$.
1. \[LA1\] There exists an orthogonal system $\{\eta_1',...,\eta_k'\}\subset\mathcal{H}$ such that $$\bigwedge_{i=1}^k{\eta_i}=\bigwedge_{i=1}^k{\eta_i'}$$ and it can be chosen in such a way that ${\left\|{\eta_i'}\right\|}\leq{\left\|{\eta_i}\right\|}$ for all $1\leq i\leq k$.
2. \[LA2\] We have the estimation $${\left\|{\bigwedge_{i=1}^k{\eta_i}}\right\|}\leq\prod_{i=1}^k{{\left\|{\eta_i}\right\|}}\,.$$ and equality is given if and only if $\{\eta_1,...,\eta_k\}$ is an orthogonal system.
$\,$
1. For every $i\in\{2,...,k\}$ there are elements $\eta_i^{(1)},\eta_i^{(2)}\in\mathcal{H}$ with $\eta_i^{(1)}\in\mathbbmss{C}\cdot\eta_1$ and $\eta_i^{(2)}\in(\mathbbmss{C}\cdot\eta_1)^{\bot}$ such that $\eta_i=\eta_i^{(1)}+\eta_i^{(2)}$. It follows $$\bigwedge_{i=1}^k{\eta_i}=\eta_1\wedge\bigwedge_{i=2}^k{\left(\eta_i^{(1)}+\eta_i^{(2)}\right)}=\sum_{j_2=1}^2{...\sum_{j_k=1}^2{\underbrace{\eta_1\wedge\bigwedge_{i=2}^k{\eta_i^{(j_i)}}}_{=0\mbox{\footnotesize{ if }}j_i=1}}}=\eta_1\wedge\bigwedge_{i=2}^k{\eta_i^{(2)}}$$ and the vectors $\eta_i^{(2)}$ are orthogonal to $\eta_1$. If we set $\eta_1':=\eta_1$ and apply the same procedure to $\bigwedge_{i=2}^{k}{\eta_i^{(2)}}$ we iteratively obtain the vectors $\eta_1',...,\eta_k'$ as claimed. Due to $${\left\|{\eta_2'}\right\|}^2={\left\|{\eta_2^{(2)}}\right\|}^2={\left\|{\eta_2}\right\|}^2-{\left\|{\eta_2^{(1)}}\right\|}^2\leq{\left\|{\eta_2}\right\|}^2$$ one also iteratively validates the inequalities ${\left\|{\eta_i'}\right\|}\leq{\left\|{\eta_i}\right\|}$.
2. By (i) there is an orthogonal system $\left\{\eta_1',...,\eta_k'\right\}\subset\mathcal{H}$ with $\bigwedge_{i=1}^k{\eta_i}=\bigwedge_{i=1}^k{\eta_i'}$ and ${\left\|{\eta_i'}\right\|}\leq{\left\|{\eta_i}\right\|}$. It follows $$\begin{aligned}
{\left\|{\eta_1\wedge...\wedge\eta_k}\right\|}^2&=&{\left\|{\eta_1'\wedge...\wedge\eta_k'}\right\|}^2=\det\left(\begin{array}{ccc}\langle\eta_1',\eta_1'\rangle&\cdots&\langle\eta_k',\eta_1'\rangle\\\vdots&&\vdots\\\langle\eta_1',\eta_k'\rangle&\cdots&\langle\eta_k',\eta_k'\rangle\end{array}\right)\\
&=&{\left\|{\eta_1'}\right\|}^2\cdot...\cdot{\left\|{\eta_k'}\right\|}^2\leq{\left\|{\eta_1}\right\|}^2\cdot...\cdot{\left\|{\eta_k}\right\|}^2\end{aligned}$$ which proves the inequality. Moreover, we see that $${\left\|{\eta_1\wedge...\wedge\eta_k}\right\|}={\left\|{\eta_1}\right\|}\cdot...\cdot{\left\|{\eta_k}\right\|}$$ if $\{\eta_1,...,\eta_k\}$ happens to be an orthogonal system. Conversely, if this is not the case then the procedure described in (i) allows to choose the orthogonal system $\{\eta_1',...,\eta_k'\}$ in such a way that at least one of the inequalities ${\left\|{\eta_i'}\right\|}\leq{\left\|{\eta_i}\right\|}$ is strict, so that the above inequality is strict as well.
In particular, Lemma \[LA\] implies that the space $\mathcal{H}^{\wedge k}$ does not contain any nonzero product vectors. Therefore, the ”classical” notion of entanglement of fermionic particles does not apply to this case.\
Whereas for fermionic particles an accepted definition of entanglement is found [@Eckert; @Grabowski; @Levay; @Plastino; @Schliemann; @Zander] there is a still ongoing debate as to how entanglement of bosonic particles should be defined [@Eckert; @Ghi02; @Ghi04; @Ghi05; @Grabowski; @Paskauskas; @Li]. In the following we will concentrate on the definition of bosonic entanglement to which our techniques apply.
\[D2\]$\,$
1. \[D21\] The set of symmetric product vectors of $\mathcal{H}^{\vee k}$ is denoted by $$V^{\vee k}:={\left\{{\eta\otimes...\otimes\eta:\,\eta\in\mathcal{H},\,{\left\|{\eta}\right\|}=1}\right\}}\,.$$ We say that a normal state $\varphi$ on $\mathcal{B}\left(\mathcal{H}^{\vee k}\right)$ is *bosonic separable* if $$\varphi\in\overline{\mathrm{co}}^{{\left\|{\cdot}\right\|}}{\left\{{\omega_{\xi}:\,\xi\in V^{\vee k}}\right\}}\,.$$
2. \[D22\] Similarly, we define $$V^{\wedge k}:={\left\{{\eta_1\wedge...\wedge\eta_k:\,\eta_i\in\mathcal{H},\,{\left\|{\eta_1\wedge...\wedge\eta_k}\right\|}=1}\right\}}$$ and call a normal state $\varphi$ on $\mathcal{B}\left(\mathcal{H}^{\wedge k}\right)$ *fermionic separable* if $$\varphi\in\overline{\mathrm{co}}^{{\left\|{\cdot}\right\|}}{\left\{{\omega_{\xi}:\,\xi\in V^{\wedge k}}\right\}}\,.$$
Comparing the definitions \[D1\], \[D2\].(\[D21\]) and \[D2\].(\[D22\]) one recognize that these are just special cases of the more general concept discussed in Section \[SA\]. In addition, the sets $V^{\otimes k}$, $V^{\vee k}$ and $V^{\wedge k}$ fulfill the following geometric relations which are of significant importance for entanglement quantification.
\[PC\] According to Theorem \[TA\].(\[TA2\]) let us define the numbers $$\mu_{\pm}:=\sup\left\{r\geq 1:\,rP_{\pm}V^{\otimes k}\subset B_{1,\mathcal{H}^{\otimes k}}\right\}\,.$$
1. \[PC1\] We have $\mu_+=1$ and $\mu_-=\sqrt{k!}$ as well as $$V^{\vee k}=\left(P_{+}V^{\otimes k}\right)\cap S_{1,\mathcal{H}^{\otimes k}}\quad\mbox{and}\quad V^{\wedge k}=\left(\sqrt{k!}P_{-}V^{\otimes k}\right)\cap S_{1,\mathcal{H}^{\otimes k}}\,.$$
2. \[PC2\] The projection $P_-$ is compatible with the pair $(V^{\otimes k},\mu_-)$ in the sense of Definition \[D5\] whereas $P_+$ is not compatible with $(V^{\otimes k},\mu_+)$.
3. \[PC3\] The mappings $$\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)\ni x\mapsto P_{\pm}xP_{\pm}\in\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)$$ are ${\left\|{\cdot}\right\|}_{K^{\otimes k}}$-contractions (compare Definition \[D4\] and Corollary \[CA\]).
$\,$
1. We begin by proving the statements about $\mu_+$ and $V^{\vee k}$. Clearly, we have $V^{\vee k}\subset V^{\otimes k}$ and every vector $\eta\in V^{\vee k}$ is a fixed-point of $P_+$. Hence, $\mu_+=1$ and $V^{\vee k}\subset (P_+V^{\otimes k})\cap S_{1,\mathcal{H}^{\otimes k}}$. Conversely, for $\eta=\eta_1\otimes...\otimes\eta_k\in V^{\otimes k}$ we have $$P_+\eta=\frac{1}{k!}\sum_{\pi\in S_k}U_{\pi}\eta=\sum_{\pi\in S_k}{\frac{1}{k!}\cdot\eta_{\pi(1)}\otimes...\otimes\eta_{\pi(k)}}\,;$$ that is $P_+\eta$ is a convex combination of unit vectors. Since the unit ball of a Hilbert space is strictly convex it follows that $P_+\eta$ is a unit vector if and only if $\eta=U_{\pi}\eta$ for all $\pi\in S_k$ which implies $\eta\in V^{\vee k}$.\
Let us now consider the fermionic analogue. Using Lemma \[LA\].(\[LA2\]) we find for $\eta_1\otimes...\otimes\eta_k\in V^{\otimes k}$ that $${\left\|{\sqrt{k!}P_-\eta_1\otimes...\otimes\eta_k}\right\|}={\left\|{\eta_1\wedge...\wedge\eta_k}\right\|}\leq{\left\|{\eta_1}\right\|}\cdot...\cdot{\left\|{\eta_k}\right\|}=1$$ and we have equality if and only if $\{\eta_1,...,\eta_k\}$ is an orthogonal system. This proves $\mu_-=\sqrt{k!}$. On the other hand, for a given unit vector $\eta_1\wedge...\wedge\eta_k\in V^{\wedge k}$ Lemma \[LA\].(\[LA1\]) allows to choose an orthogonal system $\{\eta_1',...,\eta_k'\}\subset\mathcal{H}$ with $$\eta_1\wedge...\wedge\eta_k=\eta_1'\wedge...\wedge\eta_k'\,.$$ Due to $$1={\left\|{\eta_1\wedge...\wedge\eta_k}\right\|}={\left\|{\eta_1'\wedge...\wedge\eta_k'}\right\|}={\left\|{\eta_1'}\right\|}\cdot...\cdot{\left\|{\eta_k'}\right\|}$$ and using that $\left\{\frac{\eta_1'}{{\left\|{\eta_1'}\right\|}},...,\frac{\eta_k'}{{\left\|{\eta_k'}\right\|}}\right\}$ is an orthonormal system it follows $$\eta_1\wedge...\wedge\eta_k=\frac{\eta_1'}{{\left\|{\eta_1'}\right\|}}\wedge...\wedge\frac{\eta_k'}{{\left\|{\eta_k'}\right\|}}=\sqrt{k!}P_-\frac{\eta_1'}{{\left\|{\eta_1'}\right\|}}\otimes..\otimes\frac{\eta_k'}{{\left\|{\eta_k'}\right\|}}\in\sqrt{k!}P_-V^{\otimes k}$$ so that $V^{\wedge k}\subset\left(\sqrt{k!}P_-V^{\otimes k}\right)\cap S_{1,\mathcal{H}}$. The converse inclusion is trivial.
2. If $\eta=\eta_1\otimes...\otimes\eta_k\in V^{\otimes k}$ with $P_-\eta\neq 0$, then $$\frac{P_-\eta}{{\left\|{P_-\eta}\right\|}}=\frac{\eta_1\wedge...\wedge\eta_k}{{\left\|{\eta_1\wedge...\wedge\eta_k}\right\|}}\in V^{\wedge k}$$ by definition of $V^{\wedge k}$ so that $P_-$ is compatible with $(V^{\otimes k},\mu_-)$.\
In order to see that $P_+$ is not compatible with $(V^{\otimes k},\mu_+)$ it is sufficient to consider the case $k=2$. For example, choosing two linearly independent unit vectors $\xi,\eta\in\mathcal{H}$ it is clear that $P_+(\xi\otimes \eta)=\frac{1}{2}\left(\xi\otimes \eta+\eta\otimes \xi\right)$ is nonzero but no product vector. In particular, if rescaled to unit length it cannot belong to $V^{\vee 2}$.
3. For every element $\pi\in S_k$ the permutation operator $U_{\pi}$ is a bijection of $V^{\otimes k}$ onto itself. Thus, for all $\pi,\sigma\in S_k$ the mapping $$\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)\ni x\mapsto U_{\pi}xU_{\sigma}\in\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)$$ is ${\left\|{\cdot}\right\|}_{K^{\otimes k}}$-isometric since for every $x\in\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)$ we have $$\begin{aligned}
{\left\|{U_{\pi}xU_{\sigma}}\right\|}_{K^{\otimes k}}&=&\sup\left\{\left|\left\langle U_{\pi}xU_{\sigma}\xi,\eta\right\rangle\right|:\,\xi,\eta\in V^{\otimes k}\right\}\\
&=&\sup\left\{\left|\left\langle xU_{\sigma}\xi,U_{\pi}^{\ast}\eta\right\rangle\right|:\,\xi,\eta\in V^{\otimes k}\right\}\\
&=&\sup\left\{\left|\left\langle x\xi,\eta\right\rangle\right|:\,\xi,\eta\in V^{\otimes k}\right\}\\
&=&{\left\|{x}\right\|}_{K^{\otimes k}}\,.\end{aligned}$$ It follows $$\begin{aligned}
{\left\|{P_{\pm}xP_{\pm}}\right\|}_{K^{\otimes k}}\leq\frac{1}{(k!)^2}\sum_{\pi,\sigma\in S_k}{\underbrace{{\left\|{U_{\pi}xU_{\sigma}}\right\|}_{K^{\otimes k}}}_{={\left\|{x}\right\|}_{K^{\otimes k}}}}={\left\|{x}\right\|}_{K^{\otimes k}}\,.\end{aligned}$$
Using Proposition \[PC\] and by analogy with Corollary \[CA\] we arrive at the following result concerning entanglement quantification of indistinguishable particles.
\[PD\] Let $K^{\vee k}$ and $K^{\wedge k}$ denote the norm-closed convex hulls of the sets $$\left\{\omega_{\xi}:\,\xi\in V^{\vee k}\right\}\quad\mbox{and}\quad \left\{\omega_{\xi}:\,\xi\in V^{\wedge k}\right\}$$ respectively. Then for normal states $\varphi_+\in\mathcal{B}\left(\mathcal{H}^{\vee k}\right)_{\ast}$ and $\varphi_-\in\mathcal{B}\left(\mathcal{H}^{\wedge k}\right)_{\ast}$ we have the following equivalences:
1. \[PD1\] We have $q_{K^{\vee k}}(\varphi_+)\geq 1$ and $q_{K^{\vee k}}(\varphi_+)= 1$ if and only if $\varphi_+$ is bosonic separable.
2. \[PD2\] We have $q_{K^{\wedge k}}(\varphi_-)\geq 1$ and $q_{K^{\wedge k}}(\varphi_-)= 1$ if and only if $\varphi_-$ is fermionic separable.
Furthermore, considering $\varphi_-$ as a normal state on $\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)$ via $$\mathcal{B}\left(\mathcal{H}^{\wedge k}\right)\cong P_-\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)P_-\subset\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)$$ we have $$q_{K^{\wedge k}}(\varphi_-)=\frac{1}{k!}\cdot q_{K^{\otimes k}}(\varphi_-)\,.$$ In particular, $q_{K^{\otimes k}}(\varphi_-)\geq k!$ and $q_{K^{\otimes k}}(\varphi_-)=k!$ if and only if $\varphi_-$ is fermionic separable.
Both sets $V^{\wedge k}$ and $V^{\vee k}$ are balanced subsets of the corresponding unit spheres. Moreover, since $P_-$ is compatible with $(V^{\otimes k},\mu_-)$ by Proposition \[PC\].(\[PC2\]) the set $V^{\wedge k}$ separates points of $\mathcal{H}^{\wedge k}$ according to Theorem \[TA\].(\[TA2\]). It is shown in [@OpenQuantumSystems] (compare Lemma 1.1 of Chapter 8) that the linear hull of $V^{\vee k}$ is a dense subspace of $\mathcal{H}^{\vee k}$. Hence, $V^{\vee k}$ separates points of $\mathcal{H}^{\vee k}$. The statements (\[PD1\]) and (\[PD2\]) then follow from Theorem \[TB\]. Moreover, in the fermionic case by Proposition \[PC\].(\[PC1\]) we have $\mu_-=\sqrt{k!}$ and be applying Theorem \[TA\].\[TA12\] and \[TA\].\[TA22\] with $\lambda=\sqrt{k!}$ we obtain $$\frac{1}{k!}\cdot q_{K^{\otimes k}}(\varphi_-)\leq q_{K^{\wedge k}}(\varphi_-)\leq\frac{1}{k!}q_{K^{\otimes k}}(\varphi_-)\,.$$
Let us consider some simple but instructive examples for finite dimensional situations; that is for $n:=\dim\,\mathcal{H}<\infty$ and therefore $\mathcal{B}(\mathcal{H})\cong M_n({\mathbbmss{C}})$. The quantification of entanglement of distinguishable particles in terms of the Minkowski functional $q_{K^{\otimes k}}$ then proceeds as follows.\
With respect to the trace norm ${\left\|{\cdot}\right\|}_1$ on $M_n({\mathbbmss{C}})$ we can construct the *greatest cross norm* or *projective tensor norm* ${\left\|{\cdot}\right\|}_{\pi}$ on the space $\left(M_n({\mathbbmss{C}})\right)^{\otimes k}$ (compare for example [@Ryan]). As was demonstrated by Arveson (compare Theorem 9.1 in [@Arveson]) if $\varphi$ is a (normal) state on $\left(M_n({\mathbbmss{C}})\right)^{\otimes k}$ with density operator $\rho\in\left(M_n({\mathbbmss{C}})\right)^{\otimes k}$ then $q_{K^{\otimes k}}(\varphi)={\left\|{\rho}\right\|}_{\pi}$.
\[E1\]
Consider the Hilbert space $\mathbbmss{C}^2\otimes\mathbbmss{C}^2$ and the vector state $\omega_{\xi}$ on $M_2(\mathbbmss{C})\otimes M_2({\mathbbmss{C}})$ where $$\xi:=\frac{1}{\sqrt{2}}(e_1\otimes e_2-e_2\otimes e_1)=e_1\wedge e_2$$ for an orthonormal basis $\{e_1,e_2\}$ of $\mathbbmss{C}^2$ (a so called *singlet state*). Because of $\dim\,{\mathbbmss{C}}^2\wedge{\mathbbmss{C}}^2=1$ and since $\xi\neq 0$ it is clear that ${\mathbbmss{C}}^2\wedge{\mathbbmss{C}}^2$ is generated by $\xi$. In particular, there are no fermionic entangled states on $\mathcal{B}({\mathbbmss{C}}^2\wedge{\mathbbmss{C}}^2)$. Indeed, using the known fact that $q_{K^{\otimes 2}}(\omega_{\xi})=2$ (compare Theorem 14.1 in [@Arveson]) and Proposition \[PD\] we have $$q_{K^{\wedge 2}}(\omega_{\xi})=\frac{1}{2!}\cdot q_{K^{\otimes 2}}(\omega_{\xi})=1\,.$$ so that $\omega_{\xi}$ is indeed fermionic separable.
We have seen in the above example that for dimensional reasons there are no fermionic entangled states on $\mathcal{B}({\mathbbmss{C}}^2\wedge{\mathbbmss{C}}^2)$. However, using elementary algebra one can prove that every vector $\xi\in{\mathbbmss{C}}^3\wedge{\mathbbmss{C}}^3$ is of the form $\eta\wedge\zeta$ for appropriate vectors $\eta,\zeta\in{\mathbbmss{C}}^3$ as well. This means that there are no fermionic entangled states on $\mathcal{B}({\mathbbmss{C}}^3\wedge{\mathbbmss{C}}^3)$ too. It may be instructive to reproduce this fact by using the methods we developed so far.\
To this end let $\xi\in\mathbbmss{C}^3\wedge\mathbbmss{C}^3$ be a unit vector and let $\{e_1,e_2,e_3\}$ be an orthonormal basis of $\mathbbmss{C}^3$. Then there are numbers $a,b,c\in\mathbbmss{C}$ such that $$\begin{aligned}
\xi&=&a\cdot e_1\wedge e_2+b\cdot e_1\wedge e_3+c\cdot e_2\wedge e_3\\
&=&e_1\otimes\frac{1}{\sqrt{2}}(ae_2+be_3)+e_2\otimes\frac{1}{\sqrt{2}}(-ae_1+ce_3)+e_3\otimes\frac{1}{\sqrt{2}}(-be_1-ce_2)\,.\end{aligned}$$ Defining the matrix $$A:=\left(\begin{array}{ccc}0&a&b\\-a&0&c\\-b&-c&0\end{array}\right)$$ we obtain $$A^{\ast}A=\mathbbmss{1}_3-\left(\begin{array}{c}c\\-b\\a\end{array}\right)\left(\overline{c},-\overline{b},\overline{a}\right)\,.$$ Since $\left({c},-{b},{a}\right)^T$ is a unit vector the matrix $A^{\ast}A$ is the orthogonal projection of rank 2 onto the orthogonal complement of $\left({c},-{b},{a}\right)^T$. Therefore, the theorems 7.2 and 8.2 in [@Arveson] and Proposition 7 in [@Rudolph] imply that $$\begin{aligned}
q_{K^{\otimes 2}}(\omega_{\xi})={\left\|{t_{\xi,\xi}}\right\|}_{\pi}={{\left\|{\frac{1}{\sqrt{2}}A}\right\|}_1}^2=\frac{1}{{2}}\left(\mbox{tr}\sqrt{A^{\ast}A}\right)^2=\frac{2^2}{2}=2\,.\end{aligned}$$ Like in the above example we obtain $$q_{K^{\wedge 2}}(\omega_{\xi})=\frac{1}{2}\cdot q_{K^{\otimes 2}}(\omega_{\xi})=1\,.$$ Thus, $\omega_{\xi}$ is fermionic separable, i.e. there are vectors $\eta,\zeta\in{\mathbbmss{C}}^3$ with $\xi=\eta\wedge\zeta$.
\[E2\]
Let $\mathcal{H}=\mathbbmss{C}^n$ and $\varphi_{\mbox{\footnotesize{tr}}}$ be the tracial (or chaotic) state on $\mathcal{B}\left(\left({\mathbbmss{C}}^n\right)^{\wedge k}\right)$. In particular $k\geq n$. For an arbitrary orthonormal basis $\{e_1,...,e_n\}$ of $\mathbbmss{C}^n$ we have $$\varphi_{\mbox{\footnotesize{tr}}}=\left(\begin{array}{c}n\\k\end{array}\right)^{-1}\sum_{1\leq i_1\leq...\leq i_k\leq n}{\omega_{e_{i_1}\wedge...\wedge e_{i_k}}}\in\mbox{co}\left\{\omega_{\xi}:\,\xi\in V^{\wedge k}\right\}\subset K^{\wedge k}\,.$$ Hence, $\varphi_{\mbox{\footnotesize{tr}}}$ is fermionic separable and $q_{K^{\otimes k}}(\varphi_{\mbox{\footnotesize{tr}}})$ can be computed according to $$q_{K^{\otimes k}}(\varphi_{\mbox{\footnotesize{tr}}})=k!\cdot q_{K^{\wedge k}}(\varphi_{\mbox{\footnotesize{tr}}})=k!$$ by Proposition \[PD\]. This result has been obtained earlier by Maassen [@Maassen] using different methods.
Entanglement related to Schmidt and Slater numbers {#SC}
==================================================
In addition to the concepts discussed in the previous section several other notions of entanglement are in use which aim at a somewhat finer entanglement classification. We concentrate on fermionic systems since our techniques work most efficiently in this case. Nevertheless, several aspects can partly be transferred to bosonic particles.\
In this section we will assume that $\mathcal{H}$ is infinite dimensional in order to avoid unnecessary complications when discussing the various ranks of vectors $\xi\in\mathcal{H}$ below. This is not really a restriction since every finite dimensional Hilbert space can be considered as a subspace of $\mathcal{H}$.
\[D3\] Let $\xi\in\mathcal{H}^{\otimes k}$, $\xi_-\in\mathcal{H}^{\wedge k}$ be unit vectors and $\varphi\in\mathcal{B}\left(\mathcal{H}^{\otimes k}\right)_{\ast}$, $\varphi_-\in\mathcal{B}\left(\mathcal{H}^{\wedge k}\right)_{\ast}$ normal states.
1. \[D31\] If $\xi$ can be expressed as a linear combination vectors of $V^{\otimes k}$ we define the *Schmidt rank* of $\xi$ by $$\mathrm{SR}(\xi):=\min\left\{l\in{\mathbbmss{N}}:\,\xi=\sum_{i=1}^l{\lambda_i\eta_i}:\,\eta_i\in V^{\otimes k},\,\lambda_i\in{\mathbbmss{C}}\right\}\,.$$ Otherwise, we put $\mathrm{SR}(\xi):=\infty$. The set of unit vectors having Schmidt rank smaller or equal to a number $l\in{\mathbbmss{N}}$ is denoted by $V^{\otimes k}_{l}$.
2. If $\varphi\in\bigcup_{l\in{\mathbbmss{N}}}\overline{\mathrm{co}}^{{\left\|{\cdot}\right\|}}{\left\{{\omega_{\xi}:\,\xi\in V^{\otimes k}_l}\right\}}$ the *Schmidt number* of $\varphi$ is defined by $$\mathrm{SN}(\varphi):=\min\left\{l\in{\mathbbmss{N}}:\,\varphi\in\overline{\mathrm{co}}^{{\left\|{\cdot}\right\|}}{\left\{{\omega_{\xi}:\,\xi\in V^{\otimes k}_l}\right\}}\right\}$$ and $\mathrm{SN}(\varphi):=\infty$ if this is not the case.
3. Analogously, the *Slater rank* $\mathrm{SLR}(\xi_-)$ and the *Slater number* $\mathrm{SLN}(\varphi_-)$ are defined by replacing the set $V^{\otimes k}$ in (i) and (ii) by $V^{\wedge k}$.
Note that by the definitions \[D1\], \[D2\].(\[D22\]) and \[D3\].(\[D31\]) we have $V^{\otimes k}=V_1^{\otimes k}$ and $V^{\wedge k}=V_1^{\wedge k}$.\
Schmidt and Slater numbers of normal states in the various contexts are rough measure of how entangled a given normal state is (compare for example [@Eckert; @Johnston; @Schliemann; @Terhal]). For bipartite finite dimensional quantum systems Johnston [@Johnston] has constructed a cross norm based measure for the detection of states having Schmidt number smaller or equal to a number $l\in{\mathbbmss{N}}$ by using the concept of $l$-block positivity. A comparison of this construction to Proposition \[PA\].(\[PA2\]) and (\[PA3\]) reveals that this is a special case of the following result.
\[PF\] For $l\in{\mathbbmss{N}}$ let $K^{\otimes k}_l$ and $K^{\wedge k}_l$ denote the norm-closed convex hulls of the sets $$\left\{\omega_{\xi,\eta}:\,\xi,\eta\in V^{\otimes k}_l\right\}\quad\mbox{and}\quad\left\{\omega_{\xi,\eta}:\,\xi,\eta\in V^{\wedge k}_l\right\}$$ respectively.
1. For a normal state $\varphi\in\mathcal{B}(\mathcal{H}^{\otimes k})_{\ast}$ we have $q_{K^{\otimes k}_l}(\varphi)\geq 1$ and $q_{K^{\otimes k}_l}(\varphi)=1$ if and only if $\mathrm{SN}(\varphi)\leq l$.
2. For a normal state $\varphi_-\in\mathcal{B}(\mathcal{H}^{\wedge k})_{\ast}$ we have $q_{K^{\wedge k}_l}(\varphi_-)\geq 1$ and $q_{K^{\wedge k}_l}(\varphi_-)=1$ if and only if $\mathrm{SLN}(\varphi_-)\leq l$.
Both sets $V^{\otimes k}_l$ and $V^{\wedge k}_l$ are balanced and separate the points of $\mathcal{H}^{\otimes k}$ and $\mathcal{H}^{\wedge k}$ respectively since this is the case for $V^{\otimes k}$ and $V^{\wedge k}$. The claim then follows from Theorem \[TB\].
We have seen in Proposition \[PD\] that the measures for entanglement of distinguishable and fermionic particles are related to each other by a factor of $k!$. In the sequel we generalize this result to the measures $q_{K_l^{\otimes k}}$ and $q_{K_l^{\wedge k}}$ for $l\geq 2$ and $k=2$. The restriction to the bipartite case is necessary since the proofs strongly rely on the following canonical forms for state vectors.
Let $\mathcal{H}$ be a Hilbert space and $\xi\in\mathcal{H}\otimes\mathcal{H}$. Then there are a sequence of numbers $\lambda_n\geq 0$ with $(\lambda_n)_{n\in {\mathbbmss{N}}}\in\ell^2({\mathbbmss{N}})$ and orthonormal systems $(e_n)_{n\in{\mathbbmss{N}}}, (f_n)_{n\in{\mathbbmss{N}}}\subset\mathcal{H}$ such that the following holds:
1. The vector $\xi$ can be written in the form $$\xi=\sum_{i=1}^{\infty}\lambda_n e_n\otimes f_n\,.$$ This representation is called a *Schmidt decomposition* of $\xi$ and the numbers $\lambda_n$ are called *Schmidt coefficients*.
2. If $\xi\in\mathcal{H}\wedge\mathcal{H}$ the numbers $\lambda_n$ and the vectors $e_n,f_n$ can be chosen such that the orthonormal systems $(e_n)_{n\in{\mathbbmss{N}}}$ and $(f_n)_{n\in{\mathbbmss{N}}}$ are orthogonal to each other and $$\xi=\sum_{i=1}^{\infty}\lambda_n e_n\wedge f_n\,.$$ This representation is called a *Slater decomposition* of $\xi$ with *Slater coefficients* $\lambda_n$.
In both cases the sequence $\lambda_n$ is unique up to ordering. Moreover, the number of non-vanishing coefficients $\lambda_n$ in 1. and 2. equals the Schmidt rank and the Slater rank of $\xi$ respectively.
The first statement is a simple consequence of the normal form for compact operators on a Hilbert space whereas (ii) has been proven by Schliemann et al. [@Schliemann] for finite dimensional Hilbert spaces. However, using a block diagonal argument for blocks of constant Schmidt coefficients the general case can be easily reduced to situations involving only finite dimensional Hilbert spaces:\
For a fixed Schmidt coefficient $\lambda\neq 0$ let $$N_{\lambda}:=\{n\in\mathbbmss{N}:\,\lambda_n=\lambda\}\,.$$ Since the sequence of Schmidt coefficients converges to zero the set $N_{\lambda}$ is finite and the space $$\mathcal{K}_{\lambda}:=\mbox{lin}\{e_n:\,n\in N_{\lambda}\}$$ has finite dimension. Moreover, since $\xi\in\mathcal{H}\wedge\mathcal{H}\subset\mathcal{H}\otimes\mathcal{H}$ we have $$\sum_{n=1}^{\infty}\lambda_ne_n\otimes f_n=\xi=\sum_{n=1}^{\infty}{\lambda_n(-f_n)\otimes e_n}$$ and using the fact that both sides of the above equation constitute a Schmidt decomposition of $\xi$ with the same Schmidt coefficients it follows $$\mathcal{K}_{\lambda}=\mbox{lin}\{f_n:\,n\in N_{\lambda}\}\,.$$ Furthermore, since $\mathcal{K}_{\lambda}\bot \mathcal{K}_{\mu}$ for every other Schmidt coefficient $\mu\neq \lambda$ we have $$\sum_{n\in N_{\lambda}}\lambda e_n\otimes f_n=-\sum_{n\in N_{\lambda}}\lambda f_n\otimes e_n\,.$$ This shows that $$\sum_{n\in N_{\lambda}}\lambda e_n\otimes f_n\in\mathcal{K}_{\lambda}\wedge\mathcal{K}_{\lambda}$$ and the claim now follows from the finite dimensional case.
We have the following analogue of Proposition \[PC\] concerning the geometrical properties of the sets $V^{\otimes 2}_l$ and $V^{\wedge 2}_l$.
\[PE\] For $l\in{\mathbbmss{N}}$, $l \geq 2$ let us define $$\mu_{-,l}:=\sup\left\{r\geq 1:\,rP_{-}V^{\otimes 2}_l\subset B_{1,\mathcal{H}^{\otimes 2}}\right\}\,.$$
1. \[PE1\] We have $\mu_{-,l}=1$ and $$V^{\wedge 2}_l=\left(\sqrt{2}P_-V^{\otimes 2}_l\right)\cap S_{1,\mathcal{H}^{\otimes 2}}\,.$$
2. The projection $P_-$ is compatible with $\left(V_l^{\otimes 2},\sqrt{2}\right)$.
3. The mapping $$\mathcal{B}(\mathcal{H}^{\otimes 2})\ni x\mapsto P_{\pm}xP_{\pm}\in\mathcal{B}(\mathcal{H}^{\otimes 2})$$ is a ${\left\|{\cdot}\right\|}_{K^{\otimes 2}_{l}}$-contraction.
$\,$
1. For $l\geq 2$ the projection $P_-$ has fixed-points in $V_l^{\otimes 2}$, for example vectors of the type $$\frac{1}{\sqrt{2}}(e_1\otimes e_2-e_2\otimes e_1)$$ for an orthonormal system $\{e_1,e_2\}\subset\mathcal{H}$.\
Clearly, projecting a vector $\xi\in V_l^{\otimes 2}$ onto $\mathcal{H}^{\wedge 2}$ by $P_-$ results in a vector having Slater rank less than or equal to $l$; that is $\left(\sqrt{2}P_-V_l^{\otimes 2}\right)\cap S_{1,\mathcal{H}^{\otimes 2}}\subset V_l^{\wedge 2}$. On the other hand, every vector $\xi\in V_l^{\wedge 2}$ with Slater decomposition $\xi=\sum_{i=1}^{l}{\lambda_i e_i\wedge f_i}$ can be written in the form $$\xi=\sum_{i=1}^{l}{\frac{\lambda_i}{\sqrt{2}}(e_i\otimes f_i-f_i\otimes e_i)}=\sqrt{2}P_-\sum_{i=1}^{l}{\lambda_ie_i\otimes f_i}\in \sqrt{2}P_-V_l^{\otimes 2}\,.$$
2. As mentioned in (i) we have $\mbox{SLR}(P_-\xi)\leq\mbox{SR}(\xi)$ for all $\xi\in\mathcal{H}\otimes\mathcal{H}$. Therefore, it is clear that $$\frac{P_-\xi}{{\left\|{P_-\xi}\right\|}}\in V_l^{\wedge 2}\quad\mbox{for all }\xi\in V^{\otimes 2}_l\mbox{ with }P_-\xi\neq 0$$ so that $P_-$ is compatible with $\left(V^{\otimes 2}_l,\sqrt{2}\right)$.
3. Using that for every permutation $\pi\in S_k$ the corresponding permutation operator induces a bijection from $V^{\otimes 2}_l$ onto itself the proof of Proposition \[PC\].(\[PC3\]) can be mimicked literally.
Taking a look at Proposition \[PE\].(\[PE1\]) we see that the constant $\mu_{-,l}$ no longer equals the corresponding constant $\mu_-=\sqrt{2!}=\sqrt{2}$ from Proposition \[PC\].(\[PC1\]). As a consequence, for $l\geq 2$ the Minkowski functionals $q_{K^{\otimes 2}_l}$ and $q_{K^{\wedge 2}_{l}}$ are no longer multiples of each other as it is the case for $l=1$ according to Proposition \[PD\]. Nevertheless, they are still ”equivalent” in the following sense.
\[PG\] For $l\in{\mathbbmss{N}}$, $l\geq 2$ and a normal state $\varphi_-\in\mathcal{B}(\mathcal{H}^{\wedge 2})_{\ast}$ we have the chain of inequalities $$\frac{1}{2}\cdot q_{K^{\otimes 2}_l}(\varphi_-)\leq q_{K^{\wedge 2}_l}(\varphi_-)\leq q_{K^{\otimes 2}_l}(\varphi_-)$$ and on both sides equality can be achieved for an appropriate choice of $\varphi_-$. In particular, these inequalities can in general not be sharpened any further.
Both inequalities follow from Proposition \[PE\] and Theorem \[TA\].\
In order to achieve equality we choose mutually orthogonal orthonormal systems $\{e_1,...,e_l\}, \{f_1,...,f_l\}\subset\mathcal{H}$ and consider the family of vectors $$\xi_k:=\frac{1}{\sqrt{2k}}\sum_{i=1}^k{(e_i\otimes f_i-f_i\otimes e_i)}=\frac{1}{\sqrt{k}}\sum_{i=1}^k{e_i\wedge f_i}\in\mathcal{H}\wedge\mathcal{H}\quad\mbox{for }1\leq k\leq l\,.$$ Then we have $\mbox{SR}(\xi_k)=2k$ and $\mbox{SLR}(\xi_k)=k$. Hence the corresponding normal states $\omega_{\xi_k}$ satisfy $\mbox{SN}(\omega_{\xi_k})\leq 2k$ and $\mbox{SLN}(\omega_{\xi_k})\leq k$. Therefore, equality on the right-hand-side can be achieved as follows.\
For $1\leq k \leq\left\lfloor\frac{l}{2}\right\rfloor$ (note that this condition can always be satisfied by $k$ since $l\geq 2$) we have $\mbox{SN}(\omega_{\xi_k})\leq l$ and $\mbox{SLN}(\omega_{\xi_k})\leq l$. It follows $$q_{K_l^{\otimes 2}}(\omega_{\xi_k})=1=q_{K_l^{\wedge 2}}(\omega_{\xi_k})$$ by Proposition \[PF\].\
In order to achieve equality on the left-hand-side we consider the vector $\sqrt{2}\xi_l$ (with Schmidt rank $2l$). As a first step we show that the corresponding rank-one-operator $t_{\sqrt{2}\xi_l}$ satisfies $${\left\|{t_{\sqrt{2}\xi_l}}\right\|}_{K^{\otimes 2}_l}\leq 1\,.$$ Indeed, choosing an arbitrary unit vector $\eta\in V^{\otimes 2}_l$ with Schmidt decomposition $$\eta=\sum_{j=1}^l{\lambda_je_j'\otimes f_j'}$$ a two-fold application of the Cauchy-Schwarz inequality followed by Bessel’s inequality with respect to the orthonormal system ${\left\{{e_1,...,e_l,f_1,...,f_l}\right\}}$ yields $$\begin{aligned}
\left|\left\langle\sqrt{2}\xi_l,\eta\right\rangle\right|^2&=&\left|\left\langle\frac{1}{\sqrt{l}}\sum_{i=1}^l{(e_i\otimes f_i-f_i\otimes e_i)},\sum_{j=1}^l{\lambda_je_j'\otimes f_j'}\right\rangle\right|^2\\
&=&\frac{1}{l}\left|\sum_{i=1}^l\sum_{j=1}^l\lambda_j\langle e_i,e_j'\rangle\langle f_i,f_j'\rangle-\lambda_j\langle f_i,e_j'\rangle\langle e_i,f_j'\rangle\right|^2\\
&=&\frac{1}{l}\left|\sum_{j=1}^l\lambda_j\cdot\left\langle\sum_{i=1}^l\langle e_i,e_j'\rangle f_i-\langle f_i,e_j'\rangle e_i,f_j'\right\rangle\right|^2\\
&\leq&\frac{1}{l}\underbrace{\left(\sum_{j=1}^l\lambda_j^2\right)}_{=1}\left(\sum_{j=1}^l\left|\left\langle\sum_{i=1}^l\langle e_i,e_j'\rangle f_i-\langle f_i,e_j'\rangle e_i,f_j'\right\rangle\right|^2\right)\\
&=&\frac{1}{l}\sum_{j=1}^l\left|\left\langle\sum_{i=1}^l\langle e_i,e_j'\rangle f_i-\langle f_i,e_j'\rangle e_i,f_j'\right\rangle\right|^2\\
&\leq&\frac{1}{l}\sum_{j=1}^l{\left\|{\sum_{i=1}^l\langle e_i,e_j'\rangle f_i-\langle f_i,e_j'\rangle e_i}\right\|}^2\cdot{\left\|{f_j'}\right\|}^2\\
&=&\frac{1}{l}\sum_{j=1}^l\sum_{i=1}^l\left(|\langle e_i,e_j'\rangle|^2+|\langle f_i,e_j'\rangle|^2\right)\\
&\leq&\frac{1}{l}\sum_{j=1}^l{{\left\|{e_j'}\right\|}^2}=1\,.\end{aligned}$$ It follows $$\begin{aligned}
{\left\|{t_{\sqrt{2}\xi_l}}\right\|}_{K^{\otimes 2}_l}&=&\sup\left\{\left|\left\langle t_{\sqrt{2}\xi_l}\eta,\zeta\right\rangle\right|:\,\eta,\zeta\in V^{\otimes 2}_l\right\}\\
&=&\sup_{\eta\in V_l^{\otimes 2}}\sup_{\zeta\in V_l^{\otimes 2}}\left|\left\langle\sqrt{2}\xi_l,\zeta\right\rangle\left\langle\eta,\sqrt{2}\xi_l\right\rangle\right|\\
&=&\sup_{\eta\in V_l^{\otimes 2}}\left|\left\langle\sqrt{2}\xi_l,\eta\right\rangle\right|\cdot\sup_{\zeta\in V_l^{\otimes 2}}\left|\left\langle\sqrt{2}\xi_l,\zeta\right\rangle\right|\\
&=&\sup\left\{\left|\left\langle \sqrt{2}\xi_l,\eta\right\rangle\right|^2:\,\eta\in V^{\otimes 2}_l\right\}\leq 1\\\end{aligned}$$ as claimed. By Proposition \[PA\].(\[PA3\]) $q_{K^{\otimes 2}_l}(\omega_{\xi_l})$ can therefore be estimated according to $$\begin{aligned}
q_{K^{\otimes 2}_l}(\omega_{\xi_l})&=&\sup\left\{|\omega_{\xi_l}(x)|:\,x\in\mathcal{B}(\mathcal{H}),\,{\left\|{x}\right\|}_{K^{\otimes 2}_l}\leq 1\right\}\\
&\geq&\left|\omega_{\xi_l}\left(t_{\sqrt{2}\xi_l}\right)\right|=2|\langle t_{\xi_l}\xi_l,\xi_l\rangle|=2\,.\end{aligned}$$ Since $\mbox{SLN}(\omega_{\xi_l})\leq l$, it follows $$1\leq\frac{1}{2}\cdot q_{K_l^{\otimes}}(\omega_{\xi_l})\leq q_{K_l^{\wedge 2}}(\omega_{\xi_l})=1\,.$$
According to Proposition \[PG\] for $l\geq 2$ the measures $q_{K_l^{\otimes 2}}$ and $q_{K_l^{\wedge 2}}$ are not as strongly coupled as it is the case for $l=1$ (compare Proposition \[PD\]). Nevertheless, Proposition \[PG\] can be interpreted as an equivalence of the generalized norms $q_{K_l^{\otimes 2}}$ and $q_{K_l^{\wedge 2}}$. In particular, $q_{K_l^{\otimes 2}}$ is finite if and only if $q_{K_l^{\wedge 2}}$ is finite.
Summary and Conclusions
=======================
In this paper we have demonstrated how the detection and quantification of many different kinds of quantum entanglement can be dealt with by a generalization of Arveson’s unifying approach [@Arveson]. We have constructed new and general entanglement measures for fermionic and bosonic particles which may be interpreted as generalized norms and work for mixed states and multipartite systems regardless of the dimensions. These entanglement measures also constitute necessary and sufficient computational separability criteria.\
We have seen that the measures for distinguishable and fermionic particles are multiples of each other by a factor of $k!$ where $k$ is the number of particles. When it comes to the quantification of fermionic entanglement this fact allows to treat the particles as if they were distinguishable. Moreover, the amount of entanglement for distinguishable particles emerging from fermi statistics alone is at least $k!$.\
Furthermore, we have introduced analog entanglement measures for more sophisticated notions of entanglement related to Schmidt and Slater numbers. Thereby, a result of Johnston [@Johnston] on the detection and quantification of Schmidt number entanglement which was formulated for bipartite finite dimensional systems has been generalized to arbitrary quantum systems. In the bipartite case we have demonstrated that the measures for Schmidt and Slater number entanglement satisfy a chain of inequalities which may be interpreted as an equivalence of generalized norms. Furthermore, it has been shown that these inequalities can in general not be sharpened any further.\
It is an important open question for further research whether the relations according to Proposition \[PG\] are still valid for multipartite quantum systems where the possibility of performing Schmidt and Slater decompositions is no longer given. On the other hand, using the Schmidt decomposition it is possible to find explicit formulas for the amount of entanglement of pure bipartite states for distinguishable particles [@Rudolph2]. According to the proof of Proposition \[PG\] it seems likely that this is possible for the measures $q_{K^{\otimes 2}_l}$ and $q_{K^{\wedge 2}_l}$ as well. However, many aspects concerning the subtle geometric properties of these measures are not yet fully understood. This will be another subject of further studies as well.
Acknowledgments {#acknowledgments .unnumbered}
===============
We are grateful to Hans Maassen for numerous stimulating and fruitful discussions especially concerning fermionic entanglement.
[1]{}
<span style="font-variant:small-caps;">W. Arveson</span>: *Maximal vectors in Hilbert spaces and quantum entanglement* Journal of Functional Analysis **256**, 1476-1510 (2009)
<span style="font-variant:small-caps;">A.P. Balachandran, T.R. Govindarajan, A.R. de Queiroz and A.F. Reyes-Lega</span>: *Algebraic approach to entanglement and entropy* Phys. Rev. A **88** (2013)
<span style="font-variant:small-caps;">F. Benatti, R. Floreanini, K. Titimbo</span>: *Entanglement of Identical Particles* Open Systems and Information Dynamics **21** (2014)
Springer-Verlag, New York (1990)
Academic Press, New York (1976)
North-Holland (1993)
: *Quantum Correlations in Systems of Indistinguishable Particles* Annals of Physics **299** 88-127 (2002)
<span style="font-variant:small-caps;">G. Ghiradi, L. Marinatto and T. Weber</span> *Entanglement and properties of composite quantum systems: a conceptional and mathematical analysis* J. Stat. Physics **108** (2002)
<span style="font-variant:small-caps;">G. Ghiradi, L. Marinatto</span> *General criterion for the entanglement of two indistinguishable particles* Phys. Rev. A **70** (2004)
<span style="font-variant:small-caps;">G. Ghiradi, L. Marinatto</span> *Identical particles and entanglement* Opt. Spectrosc. **99** (2005)
: *Entanglement for multipartite systems of indistinguishable particles* Journal of Physics A, **44** (2011)
: *Quantifying quantum correlations in fermionic systems* Quantum Inf. Process. **12** 733 (2013)
: *Computable measures for the entanglement of indistinguishable particles* Phys. Rev. A **87** (2013)
<span style="font-variant:small-caps;">P. Levay, P. Vrana</span>: *Three fermions with six single-particle states can be entangled in two inequivalent ways* Phys. Rev. A **78** (2008)
<span style="font-variant:small-caps;">H. Maassen</span>: *Greatest cross norm progress report* Private communication (2010).
<span style="font-variant:small-caps;">R. Paskauskas and L. You</span> *Quantum correlations in two boson wave functions* Phys. Rev. A **64** (2001)
<span style="font-variant:small-caps;">A.R. Plastino, D. Manzano and J.S. Dehesa</span>: *Separability criteria and entanglement measures for pure states of $N$ identical fermions* EPL **86** (2009)
<span style="font-variant:small-caps;">N. Johnston</span>: *Norm duality and the cross norm criteria for quantum entanglement* Linear and Multilinear Algebra (2013)
<span style="font-variant:small-caps;">Y.S. Li, B. Zeng, X.S. Liu ans G.L. Long</span>: *Entanglement in a two-identical-particle system.* Phys. Rev. A **64** (2001)
<span style="font-variant:small-caps;">O. Rudolph</span>: *A separability criterion for density operators* Journal of Physics A: Math. Gen. **33** 3951 (2000)
<span style="font-variant:small-caps;">O. Rudolph</span>: *A new class of entanglement measures* J. Math. Phys. **42** 5306-5314 (2001)
<span style="font-variant:small-caps;">O. Rudolph</span>: *Further results on the cross norm criterion for separability.* Quantum Inf. Process. **4**, 219 (2005)
<span style="font-variant:small-caps;">R.A. Ryan</span>: Introduction to Tensor Products of Banach Spaces, Springer Verlag (2002)
<span style="font-variant:small-caps;">J. Schliemann, J.I. Cirac, M. Kus, M. Lewenstein and D. Loss</span>: *Quantum correlations in two-fermion systems* Phys. Rev. A **64** (2001)
<span style="font-variant:small-caps;">B. Terhal and P. Horodecki</span>: *Schmidt number for density matrices* Phys. Rev. A **61** (2000)
<span style="font-variant:small-caps;">M.C. Tichy, F. de Melo, M. Kus, F. Mintert and A. Buchleitner</span>: *Entanglement of identical particles and the detection process*
: *Entropic entanglement criteria for Fermion systems* Eur. Phys. J. D (2012) 66: 14
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'It is currently believed that the “standard” accretion disk theory under-predicts the observed X-ray luminosity from Soft X-ray Transients (SXT) in quiescence by as much as 4 to 6 orders of magnitude. This failure of the standard model is considered to be an important argument for the existence of the alternative mode of accretion – Advection Dominated Accretion Flows (ADAF) in astrophysics, since these flows allow a much higher level of X-ray emission in quiescence, in agreement with the observations. Here we point out that, in stark contrast to steady-state standard disks, such disks in quiescence (being non-steady) produce most of the X-ray emission very far from the last stable orbit. Taking this into account, these disks [*can*]{} accommodate the observed X-ray luminosities of SXTs rather naturally. Our theory predicts that [Fe K$\alpha$ ]{}lines from standard accretion disks in quiescence should be narrow even though the cold disk goes all the way down to the last stable orbit.'
author:
- Sergei Nayakshin and Roland Svensson
title: 'On the “failure” of the standard model for Soft X-ray Transients in quiescence.'
---
Introduction {#sect:intro}
============
Soft X-ray transients are mass transfer binaries which periodically undergo outbursts in which their luminosity increase by several orders of magnitude (e.g., see reviews by Tanaka & Lewin 1995; van Paradijs & McClintok 1995; Wheeler 1996; Cannizzo 1998). The compact object may be either a neutron star (NS) or a black hole (BH). SXTs are particularly interesting given the ongoing debate about the structure of the inner accretion disk. In quiescence, the luminosity of SXTs seems to be much higher (e.g., Verbunt 1995; Lasota 1996, Yi et al. 1996; Lasota, Narayan & Yi 1996a; Narayan, McClintock & Yi 1996) than allowed by the standard model (Shakura & Sunyaev 1973; expected to be modified for SXTs due to hydrogen-ionization instability – e.g., Mineshige & Wheeler 1989 \[MW89\]). Lasota, Narayan & co-workers argued that because the “standard” theory predicts that the inner accretion disk is essentially empty of mass in quiescence and accretes at a very small accretion rate, $\dot{M}_{\rm in}$, it cannot reproduce the observed X-ray luminosity $L_{\rm x, obs}$: typical numbers are $0.1
\dot{M}_{\rm in} c^2\sim (10^{-4} - 10^{-6})\,\times L_{\rm x,
obs}$. They conclude that SXTs in quiescence clearly invalidate the standard model.
However, in this Letter, we argue that this argument is applicable only to “bare” standard accretion disks – those without an overlying hot corona. We find that quiescent standard disks [*with a corona*]{} should produce most of their X-ray emission at very large radii precisely due to the fact that the inner disk is almost empty of mass. At large radii, the standard theory of hydrogen-ionization instability allows much larger accretion rates, and the observed X-ray luminosities can be easily reproduced if the corona reprocesses $\sim
10$% of the disk energy budget. We also make a testable prediction – [Fe K$\alpha$ ]{}lines from quiescent accretion disks should be narrow (even though these disks extend down to the last stable orbit or the surface of the compact star).
Hydrogen ionization instability: basic facts {#sect:inst}
============================================
Because the argument of Lasota (1996) against the standard model in SXTs is qualitative, we will address it with a qualitative model as well. The hydrogen ionization instability of accretion disks has been studied for more than two decades for a range of central objects – White Dwarfs (WD), NS, and galactic and extra galactic BHs (e.g., Hoshi 1979; Smak 1982; 1984; MW89; Mineshige & Shields 1990 \[MS90\]; Cannizzo 1993; Siemiginowska, Czerny & Kostyunin 1996 \[SCK96\]). Cannizzo (1998, §2 & 3) gives a short overview of the development of the theory and an easily accessible and enlightening discussion of the background physics. Here we will only sketch the basic principles of the instability to the extent needed to explain our main point. The behavior of the unstable disk depends on the so-called “S-curve” that relates the local disk effective temperature and the column density (see Figures in Smak 1982, MS90, SCK96; and Fig. 31 in Frank, King & Raine 1992). The curves are different for different radii, but luckily only slightly. The steady state Shakura-Sunyaev solution predicts that the disk effective temperature, ${T_{\rm eff}}(R)$, changes as $R^{-3/4}$ with radius. In general, this predicts that for small $R$, hydrogen is completely ionized, then there should be a range of radii where hydrogen is partially ionized, and finally, at large $R$, hydrogen should recombine completely.
However, it turns out that the partially ionized solution is thermally unstable. In terms of the S-curve, this region is between points C and E (see Fig. 31 in Frank et al. 1992). Numerical simulations show that the unstable disks oscillate between two quasi-stable states. The accretion rate is not constant with time and radius, but the time-averaged accretion rate equals that at which the matter is supplied at the outer edge of the disk, $\dot{M}_0$. One of these states – the region between points E and D on the S-curve – is the outburst state in which hydrogen is completely ionized. The other thermally stable state – the one between points B and C on the S-curve in Fig. 31 of Frank et al. 1992 – is the quiescent state when hydrogen is neutral. To remain in this state, the mid-plane gas temperature should be kept below $\sim 10^4$ Kelvin, which in turn requires ${T_{\rm eff}}(R) < T_C(R) \simlt {\rm few} \times 10^3 $ (depending on the black hole mass, $M$, and viscosity prescription).
Time-dependent (full disk) calculations of many authors show that when the unstable region of the disk returns to quiescence, it settles around point B on the S-curve, which is a little below the maximum quiescent temperature and surface density (point C). The accretion rate corresponding to point B, $\dot{M}_B(R)$, is lower than $\dot{M}_0$, so that the matter starts accumulating, most quickly at the outermost part of the unstable region. At a future time, this piling up of the mass will lead to $\dot{M}(R)> \dot{M}_C$ at some radius (equivalently, ${T_{\rm eff}}(R)> T_C(R)$), and the outburst will be triggered at that radius (see Cannizzo 1998, §3).
Fortunately, for the present analysis, these details are mostly irrelevant. What is important for us here is that both local disk calculations and full disk time-dependent ones show that the gas effective temperature in quiescence is between $T_B(R)$ and $T_C(R)$; that $T_B(R)$ is typically smaller than $T_C(R)$ by a factor of 2 or less, and that both of these functions are extremely weakly dependent on $R$. In particular, the scaling law given by Smak (1984), MS90 and SCK96 is $T_C(R)\propto R^{-q}$, where $q$ is between 0.08 and 0.12. Full disk calculations give temperature profiles slowly increasing and then decreasing with $R$ (e.g., Fig. 13 and 16 in MS90; Fig. 10 in SCK96), which is probably due to the fact that the outer disk fills up quicker and hence lies closer to point C than the inner disk does. There are also observations of the disk effective temperature in quiescence for eclipsing dwarf novae, systems to which the hydrogen-ionization instability was applied first. These observations show that ${T_{\rm eff}}(R)$ is either a constant or a slowly decreasing function of $R$ (Fig. 11 in Wood et al. 1986; Fig. 14 in Wood et al. 1989; and Figs. 5 & 6 in Bobinger et al. 1997).
While the exact run of ${T_{\rm eff}}(R)$ with radius depends on details of the model and the object under consideration, [*all*]{} of the above observational and theoretical papers show that ${T_{\rm eff}}(R)\sim $ const in quiescence within a factor no larger than $2$. This is quite remarkable given that the radii of the unstable region can span some two or more orders of magnitude: the steady-state Shakura-Sunyaev solution with ${T_{\rm eff}}(R)\propto R^{-3/4}$ would predict a change in ${T_{\rm eff}}$ by about a factor of $30$. Hence, below we will assume that ${T_{\rm eff}}(R)\sim T_U \equiv$ const $\simlt T_C\sim $ few $\times 10^3$ K in quiescence. Physically, the near isothermality of the disk in quiescence is due to the fact that the ionization state of hydrogen is a very sensitive function of temperature, and large changes in radius cause only a slight difference in $T$. For example, the opacity is an extremely strong function of midplane disk temperature $T_d$: in the temperature range $6000 \simlt T_d \simlt 8000$ K, the opacity $\propto T_d^b$ where $b$ is as large as 10–20 (see Cannizzo & Wheeler 1984).
Of course, the effective temperature is approximately constant only inside the unstable disk region. Let us call the corresponding radius $R_U$. For larger radii, the disk is always non-ionized and so it follows the steady state Shakura-Sunyaev solution, i.e., ${T_{\rm eff}}(R)\propto R^{-3/4}$. Summarizing, $${T_{\rm eff}}(R) \simeq \cases{ T_U, & if\, $R \, < R_U$ \cr T_{\rm SS}(R) &
otherwise \, ,\cr}
\label{tu}$$ where $T_{\rm SS}(R)$ is the Shakura-Sunyaev effective temperature of the disk corresponding to the given radius, accretion rate $\dot{M}_0$ and the black hole mass. The thermal disk flux is simply given by ${F_{\rm d}}(R) = \sigma_B {T_{\rm eff}}^4(R)$.
What are the implications of this for the accretion rate through the disk? At $R\gg R_s\equiv 2GM/c^2$, ${F_{\rm d}}(R) = (3/8\pi)
GM\dot{M}(R)/R^3$, which can be solved for $\dot{M}(R)$: $$\dot{M}(R) \simeq \cases{ \dot{M}_0 (R/R_U)^3, & for $R < R_U$ \cr
\dot{M}_0 & for $R > R_U$\cr}
\label{mofr}$$ That is, the accretion rate scales as $\dot{M}(R)\propto R^3$ within the unstable radius. Physically, this comes about from the fact that it is most difficult to keep the hydrogen recombined for smaller radii because the gravitational and viscous energy release scale as $R^{-3}$, and hence the disk has to be most strongly depleted for the smallest $R$ to remain quiescent (see also Cannizzo 1998).
X-rays do not always come from the inner disk {#sect:cr}
=============================================
A frequent implicit or explicit assumption in the common logic about the X-ray emitting region(s) is the following. Since X-rays are the “hottest” part of the overall disk spectrum, and since the Shakura & Sunyaev (1973) model predicts that the disk is hottest in its innermost region, then that has to be the region where the X-rays are produced. Seemingly innocent, this suggestion is actually appropriate for the ADAF-type models (i.e., the inner-ADAF plus outer cold disk models), but is actually entirely foreign for standard disks with coronae in quiescence.
To show this, let us first assume without any justification that the corona above the disk exists at all radii, and that the fraction of power transferred into the corona, $f$, is independent of radius (e.g., Svensson & Zdziarski 1994). That is, ${F_{\rm x}}(R) = f {F_{\rm d}}(R)$. Next, we define the X-ray luminosity per decade of radius, as $L_x(R)\equiv R^2 {F_{\rm x}}(R)$, which obviously scales according to $$L_x(R) = \cases{ f L_U (R/R_U)^2, & for \, $R < R_U$ \cr
f L_U (R/R_U)^{-1} & for $R > R_U$ \cr}\, ,
\label{lofr}$$ where $L_U = {F_{\rm d}}(R_U) R_U^2$. Therefore, [*most*]{} of the viscous energy liberation [*and X-ray production*]{} for a quiescent disk occurs at $R\sim R_U \gg 3 R_S$, if $f=$ const. The inner disk in the standard model is simply very dim and for many purposes may be unobservable [*at any wavelength*]{} in quiescence.
Is it appropriate to assume $f\simeq $ const? To date, detailed MHD simulations of accretion disks have not yet addressed this question, because global MHD disk simulations are forbiddenly numerically extensive, and hence calculations are limited to a small region in radius (see, e.g., Miller and Stone 2000). In the absence of these, we will make a simple analytical argument. If one defines the $\alpha$-parameter through $\nu = \alpha c_s H$, where $c_s$ and $H$ are the sound speed and the pressure scale height, respectively, then the total energy generated by the disk per unit surface area is $2
{F_{\rm d}}(R) = (9/8)\, (GM/R^3)\, \nu \Sigma = (9/8)\, \alpha c_s {P_{\rm tot}}$, where $\Sigma$ is the vertical mass column density, and ${P_{\rm tot}}$ is the total disk mid-plane pressure. The flux of magnetic energy, ${F_{\rm mag}}$, delivered by buoyancy to the disk surface and then presumably transformed into X-rays by reconnection is $ {F_{\rm mag}}\sim v_b
\langle{P_{\rm mag}}\rangle$, where $v_b\simlt c_s$ is the average buoyant velocity, and $\langle{P_{\rm mag}}\rangle$ is the average magnetic field pressure ($\sim$ energy density). The viscosity parameter $\alpha$ was defined by Shakura & Sunyaev (1973) to be $\alpha \simeq \alpha_t +
{\langle{P_{\rm mag}}\rangle\over P_{\rm tot}}$, where the first term, $\alpha_t$ is the turbulent viscosity coefficient. It is now believed that the magnetic viscosity is significantly more important than the turbulent one (e.g., Hawley, Gammie, & Balbus 1995; 1996), and that $\alpha \sim \langle{P_{\rm mag}}\rangle/ P_{\rm tot}$, and hence $${F_{\rm mag}}\sim v_b \alpha {P_{\rm tot}}\simlt \alpha c_s {P_{\rm tot}}\sim {F_{\rm d}}\; .
\label{fm2}$$ In other words, this equation shows that $f$ is indeed expected to be a constant with respect to $R$.
Further, if corona is powered by magnetic flares (e.g., Galeev, Rosner & Vaiana 1979; Haardt, Maraschi & Ghisselini 1994; Beloborodov 1999; Miller & Stone 2000), then it is clear that the resulting X-ray spectra have little to do with the temperature of the underlying disk. The spectra depend primarily on the Thomson depth of the hot plasma and [*its*]{} temperature. The temperature within a flare is controlled by the ratio of the X-ray flux to the disk flux, ${F_{\rm x}}/{F_{\rm d}}$, (e.g., Stern et al. 1995; Poutanen & Svensson 1996), plus the exact (unknown) geometry of the reconnection site. Therefore, it is entirely possible that a magnetic flare that occurs at $R\sim
1000\, R_S$ will have an X-ray spectrum similar to, or even harder than that of a flare at $R \sim 10 R_S$, especially for the quiescent disks.
To summarize this discussion, there is nothing special about the innermost region of the disk for quiescent accretion disks. If there is a corona above the disk, then most of the emission at any wavelength will be produced at $R\sim R_U$ \[Note that in a steady Shakura-Sunyaev disk, most of the viscous energy release happens in the inner disk, and this is why it is safe to assume that X-rays come from very small radii [*for such disks*]{}\]. Since there are several parameters that will determine the resulting X-ray spectra (e.g., reconnection rate, height and geometry of the X-ray source, Thomson depth of the X-ray emitting plasma, etc.), it is difficult to calculate the spectra model-independently. Nevertheless, we note that variations of these parameters are sure to yield a broad range of spectral indices (see, e.g., Stern et al. 1995; Poutanen & Svensson 1996), and hence it may be entirely possible to reproduce the observed spectra of SXTs.
X-ray “Problem” for standard disks in quiescent SXTs {#sect:sxts}
====================================================
Currently, it is believed that there exists a clear-cut “evidence” for a major failure of the standard disk instability model for the quiescent state accretion disks around dwarf novae, neutron stars and galactic black holes (SXTs). The popularity of the ADAF model itself in part rests on this fact because it is claimed to be the only model that can explain self-consistently the level of X-rays in quiescence for the systems mentioned above. Let us for a moment accept the suggestion that X-rays do emanate from the innermost region and follow the arguments of Lasota (1996), as well as many of other authors (e.g., Lasota et al. 1996a; Yi et al. 1996). The requirement for the disk material to be below the instability point C on the S-curve limits the accretion rate to be (Lasota 1996): $$\dot{M}_{\rm m}(R) \simeq 2.76 \times 10^3 t_9^{-1}\, r_7^{3.11}
M_1^{-0.37}\, \alpha^{-0.79}\; {\rm g\ s}^{-1}\;,
\label{lasota1}$$ where $M_1$ is the mass of the black hole in units of $10 {{\,M_\odot}}$, $t_9$ is the recurrence time in units of $10^9$ s, and $r_7 \equiv
R/10^7$ cm. Numerical models (e.g., MW89) satisfy this requirement. This condition (eq. 5) is similar to the simpler one given by equation (\[mofr\]) with a scaling $\dot{M}_{\rm
m}(R)\propto R^{3.11}$ instead of $R^3$. Using a $10$ % efficiency for conversion of mass accretion rate into X-rays, Lasota (1996) noted that two SXTs, A0620-00 and V404 Cyg, have been detected at the accretion rates $\sim 10^{10}$ and $\sim 3\times 10^{12}$ g s$^{-1}$ (see Verbunt 1995). This estimate is 4-6 orders of magnitude larger than that given by equation (\[lasota1\]). However, as we argued in §\[sect:cr\], most of X-ray emission comes from $R\sim R_U \gg 3 R_S$ for standard disks in quiescence. Approximately, we have $$L_x \simlt f\, \frac{GM\dot{M}(R_U)}{R_U} = f \frac{R_S}{2 R_U}
\dot{M}(R_U)\, c^2\;,
\label{lasota2}$$ where $\dot{M}(R_U)\sim \dot{M}_0$. Since in quiescence $\dot{M}(R_U)\simlt \dot{M}_{\rm m}(R_U)$, we estimate $$L_x \simlt 3.7 \times 10^{23} \, f\, t_9^{-1}\, r_7^{2.11}
M_1^{0.63}\, \alpha^{-0.79}\; {\rm erg\ s}^{-1}\;.
\label{lasota3}$$ Clearly, by allowing $r_7 = R_U/10^7 {\rm cm}$ to be much greater than unity, we can obtain much higher $L_x$ than the estimate based on 10% efficiency and $r\sim $few.
Let us now use theoretical and observational constraints on the parameters in equation (\[lasota3\]) to determine $L_x$ more accurately for A0620-00. From observations, we know that $L_x\sim 0.1 L_{\rm bb}$, where $L_{\rm bb}\simeq 10^{32}
{\rm erg\ s}^{-1}$ is the blackbody optical luminosity of the disk as inferred by McClintock et al. (1995). This means that we need to choose $f\simeq 0.1$. Further, analyses of H$\alpha$ and $H\beta$ spectroscopic data (Marsh et al. 1994) indicate that the radius where the Balmer line luminosity peaks is $\sim 5\times 10^{10}$. This should roughly equal $R_U$ since that is where most of the luminosity is produced in our model. Therefore, $$L_x \sim 9.4 \times 10^{30} \left(\frac{f}{0.1\, t_9}\right)\,
\left(\frac{M}{5{{\,M_\odot}}}\right)^{0.63}\,
\alpha_1^{-0.79}\; \frac{\rm erg}{\rm s}\;,
\label{me}$$ where $\alpha_1\equiv \alpha/0.1$. This value of $L_x$ is in a very good agreement with the observed $L_x\sim 10^{31} {\rm erg\ s}^{-1}$. A further independent check of the self-consistency of this picture can be made by computing the temperature $T_U$, from $\sigma_B T_U^4 = (1-f)
(3/8\pi) GM\dot{M}_0/R_U^3$. For $\dot{M}_0 \simeq 10^{16}$ g s$^{-1}$ (see McClintock et al. 1995), we find $$T_U = 3.2 \times 10^3\, \left(\frac{M}{5{{\,M_\odot}}}\right)^{1/4}\;.
\label{tucheck}$$ This temperature estimate agrees quite well with the expected temperature at the lower kink of the S-curve (see Figs. in Smak 1982, MS90, SCK96).
Discussion {#sect:conclusion}
==========
In this paper we re-examined the argument given by Lasota (1996), Lasota et al. (1996a) and Yi et al. (1996) with which they rule out the standard accretion disks extending down to the last stable orbit for quiescent states of SXTs. The argument is based on the assumption that all X-rays are emitted very close to the last stable orbit. Because the accretion rate through this region is so low in quiescence, the standard theory appears to under-predict the observed X-ray luminosities by orders of magnitude. However, we have shown that this argument is only applicable to bare standard accretion disks, i.e., those that have no hot X-ray producing corona. Such disks are ruled out already due to the fact that they do not produce hard X-rays at all, and hence there is no reason to invoke them as a model for SXTs in the first place (and not only for SXTs but for all X-ray emitting accreting sources). We have shown that if a corona exists at all radii, then most of X-ray emission will come from large radii in quiescence. At these radii, the maximum allowed accretion rate – the maximum one at which hydrogen is neutral – is much larger. Accordingly, much larger X-ray luminosities may be sustained by such “non-bare” disks, in qualitative agreement with the X-ray observations of quiescent SXTs.
Further progress in testing this and the ADAF models for SXTs can only be made by a detailed comparison of theoretical and observed spectra. While there is a large body of literature on the spectra of the ADAF model for SXTs (e.g., see the recent review by Lasota 2001) and none for the model proposed here, both of these face serious theoretical challenges and hence neither model can be preferred (in our opinion). In particular, standard disks with magnetic flares is perhaps the most viable model for [*luminous*]{} accreting black holes, both AGN and GBHCs (e.g., Galeev et al. 1979; Beloborodov 1999; Nayakshin & Dove 2001; Done & Nayakshin 2001). However, it is not clear at all how spectra of the high luminosity magnetic flares (e.g., Poutanen & Svensson 1996) should be modified when applied to SXTs. Ironically, the uncertainties in the theory may be great enough to preclude both positive or negative statements about the “standard” model for SXTs, if these require detailed spectra.
The ADAF model, on the other hand, has fundamental difficulties in the modeling of the cold-disk ADAF transition, which is therefore often done in an ad-hoc way. Another major uncertainty of the theory is the accretion rate through the disk as a function of radius ($\dot{M}(R)$). The earlier papers assumed $\dot{M}(R)=$ const. However, this would only be the case if the ADAF solution joined the cold disk at $R_t \simgt R_U$, where $\dot{M}(R)$ is indeed constant (see eq. \[mofr\]). Yet $R_t$ is treated as a free parameter of the model with $\dot{M}$ being independent of it (e.g., Lasota et al. 1996b; Yi et al. 1996). From our analysis above, we believe that if one chooses $R_t < R_U$, the accretion rate through the ADAF region should be smaller by the factor $(R_t/R_U)^3$ (see also Wheeler 1996, on a similar point). Since the ADAF luminosity scales as $\dot{M}^2$ for small accretion rates, this factor translates into the sixth power of $R_t/R_U$ in $L_x$. In addition, the issue is further complicated by the more recent discovery of the importance of convection and winds in the ADAF solution (e.g., Blandford & Begelman 1999), neglected in the earlier papers. Taken together, this implies that at the present, the X-ray spectra of the ADAF model are no more well determined than the spectra of magnetic flares.
One obvious consequence of our model is that because most of the X-ray power is radiated away at $R\sim R_U$, [Fe K$\alpha$ ]{}lines of quiescent SXTs will have to be narrow even though the disk extends down to the last stable orbit. While narrow-ish lines are also predicted by the ADAF model, we believe that a detailed modeling and observations of these lines are still the key to distinguish the two models. The ADAF model produces the lines mostly due to collisional processes, and the line is dominated by H- and He-like components (e.g., Narayan & Raymond 1999) with the line centroid energy of $\sim$ 6.9 and 6.7 keV, respectively. In the case of a cold disk plus corona model, depending on the parameters of X-ray spectra we can expect the line centroid energy to be either at $\sim 6.4$ or $6.7$ keV or be a combination of these two (e.g., Nayakshin & Kallman 2001). Highly ionized collisionally excited [Fe K$\alpha$ ]{}lines may also be expected based on the analogy with [Fe K$\alpha$ ]{}lines from solar flares (e.g., see Pike et al. 1996). The ADAF model is however unlikely to produce a neutral-like [Fe K$\alpha$ ]{}line, because most of X-ray emission in the ADAF is expected at small radii which yields little illuminating flux at large radii where the cold disk is located.
Finally, we would like to stress the following general “theorem”. The absence of emission from the inner disk (whether this emission is the [Fe K$\alpha$ ]{}line emission, the X-ray continuum or the thermal disk emission), is [*not*]{} an evidence for the absence of the standard disk there. The particular case of SXTs discussed here proves this point. Because hydrogen is neutral in quiescence, the inner disk is expected to be very dim at all wavelengths and so it may be difficult to discern in the data. This point is quite general as it may also be relevant for other than hydrogen ionization disk instabilities. For example, Nayakshin, Rappaport & Melia (2000) have devised a specific model for the radiation-pressure driven instabilities in the galactic micro-quasar GRS 1915+105. In this model, the inner disk is also very much dimmer during the “quiescence” than expected in the steady-state Shakura-Sunyaev solution (see Fig. 13 in Nayakshin et al. 2000). Therefore, one must not use the steady-state Shakura-Sunyaev theory for unsteady disks – e.g., low luminosity systems, such as SXTs and Low Luminosity AGN (since one may suspect that the reason these objects are dim is that they are in a quiescent state), and those verifiably unsteady such as GRS 1915+105.
The authors acknowledge many useful discussions with John Cannizzo, Chris Done and Demos Kazanas.
Beloborodov, A. 1999, , 510, L123
Blandford, R. D., & Begelman, M. C. 1999, , 303, L1
Bobinger, A., Horne, K., Mantel, K.-H., Wolf, S. 1997, , 327, 1023
Cannizzo, J. K. 1993, in Accretion Disks in Compact Stellar Systems, ed. J. C. Wheeler (Singapore: World Scientific), 6
Cannizzo, J. K. 1998, in ASP Conf. Ser. 137: Wild Stars in The Old West, ed. S. Howel, E. Kuulkers, & C. Woodward (San Francisco: ASP), 308
Cannizzo, J. K., & Wheeler, J. C. 1984, , 55, 367
Done, C., & Nayakshin, S. 2001, , 546, 419
Frank, J., King, A., & Raine, D. 1992, Accretion Power in Astrophysics (Cambridge, UK: Cambridge University Press)
Galeev, A. A., Rosner, R., & Vaiana, G. S., 1979, , 229, 318
Haardt F., Maraschi, L., & Ghisellini, G. 1994, , 432, L95
Hawley, J. F., Gammie, C. F., & Balbus, S. A. 1995, , 440, 742
Hawley, J. F., Gammie, C. F., & Balbus, S. A. 1996, , 464, 690
Hoshi, R., 1979, Progr. Theor. Phys., 61, 1307
Lasota, J.-P. 1996, in IAU Symp. 165: Compact Stars in Binaries, ed. J. van Paradijs, E. P. J. van den Heuvel & E. Kuulkers (Dordrecht: Kluwer), 43
Lasota, J.-P. 2001, New Astronomy Reviews, in press.
Lasota, J.-P., Narayan, R., & Yi, I. 1996a, , 314, 813
Lasota, J.-P., Abramowicz, M.A., Chen, X., Krolik, J., Narayan, R., & Yi, I. 1996b, , 462, 142
Marsh, T. R., Robinson, E. L., & Wood, J. H., 1994, , 266, 137
McClintock, J. E., Horne, K., & Remillard, R. A. 1995, , 442, 358
Miller, K. A., & Stone, J. M. 2000, , 534, 398
Mineshige, S., & Shields, G.A. 1990, , 351, 47 (MS90)
Mineshige, S., & Wheeler, J.C. 1989, , 343, 241 (MW89)
Narayan, R., & Raymond, J. 1999, , 515, L69
Narayan, R., McClintock, J. E., & Yi, I. 1996, , 457, 821
Nayakshin, S., & Dove, J. 2001, , in press
Nayakshin, S., & Kallman, T. 2001, , 546, 406
Nayakshin, S., Rappaport, S., & Melia, F. 2000, , 535, 798
Pike, C.D. et al. 1996, ApJ, 464, 487
Poutanen, J. & Svensson, R. 1996, , 470, 249
Shakura, N. I., & Sunyaev, R. A. 1973, , 24, 337
Siemiginowska, A., Czerny, B., & Kostyunin, V. 1996, , 458, 491 (SCK96)
Smak, J. 1982, Acta Astr., 32, 199
Smak, J. 1984, Acta Astr., 34, 161
Stern, B. E., Poutanen, J., Svensson, R., Sikora, M., & Begelman, M. C. 1995, , 449, L13
Svensson, R. 1996, , 120C, 475
Svensson, R. & Zdziarski, A. A. 1994, , 436, 599
Tanaka, Y., & Lewin, W. H. G. 1995, in X-Ray Binaries, ed. W. H. G. Lewin, J. van Paradijs, & E. P. J. van den Heuvel (Cambridge, UK: Cambridge Univ. Press), 126
van Paradijs, J., & McClintock, J. E. 1995, in X-Ray Binaries, ed. J. van Paradijs, & W. H. G. Lewin, E. P. J. van den Heuvel (Cambridge, UK: Cambridge Univ. Press), 58
Verbunt, F. 1995, in IAU Symp. 165: Compact Stars in Binaries, ed. J. van Paradijs, E. P. J. van den Heuvel, & E. Kuulkers (Dordrecht: Kluwer), 333
Wheeler, J. C. 1996, in Relativistic Astrophysics: A Conference in Honor of Igor Novikov’s 60th Birthday", ed. B. Jones & D. Marković (Cambridge, UK: Cambridge Univ. Press), 211
Wood, J, Horne, K., Berriman, G., Wade, R., O’Donoghue, D., & Warner, B. 1986, , 219, 629
Wood, J, Horne, K., Berriman, G., & Wade, R. 1989, , 341, 974
Yi, I., Narayan, R., Barret, D., & McClintock, J. E. 1996, , 120C, 187
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
The thermally assisted detachment of a self-avoiding polymer chain from an adhesive surface by an external force applied to one of the chain ends is investigated. We perform our study in the “fixed height” statistical ensemble where one measures the fluctuating force, exerted by the chain on the last monomer when a chain end is kept fixed at height $h$ over the solid plane at different adsorption strength $\epsilon$. The phase diagram in the $h -
\epsilon$ plane is derived both analytically and by Monte Carlo simulations. We demonstrate that in the vicinity of the polymer desorption transition a number of properties like fluctuations and probability distribution of various quantities behave differently, if $h$ rather than $f$ is used as an independent control parameter.
author:
- 'S. Bhattacharya$^1$, A. Milchev$^{1,2}$, V.G. Rostiashvili$^1$ and T.A. Vilgis$^1$'
title: Pulling an adsorbed polymer chain off a solid surface
---
Introduction {#Intro}
============
The properties of single polymer chains at surfaces have received considerable attention in recent years. Much of this has been spurred by new experimental techniques such as atomic force microscopy (AFM) and optical/magnetic tweezers [@Smith] which allow one to manipulate single polymer chains. Study of single polymer molecules at surfaces, such as mica or self-assembled monolayers, by Atomic Force Microscopy (AFM) method provides a great scope for experimentation [@Senden; @Hugel; @Seitz; @Rohrbach; @Gaub_2; @Friedsam; @Gaub_1]. Applications range from sequential unfolding of collapsed biopolymers over stretching of coiled synthetic polymers to breaking individual covalent bonds[@Rief; @Ortiz; @Grandbois].
In these experiments it is customary to anchor a polymer molecule with one end to the substrate whereas the other end is fixed on the AFM cantilever. The polymer molecule can be adsorbed on the substrate while the cantilever recedes from the substrate. In so doing one can prescribe the acting force in AFM experiment whereas the distance between the tip and the surface is measured. Conversely, it is also possible to fix the distance and measure the corresponding force, a method which is actually more typical in AFM-experiments. From the standpoint of statistical mechanics these two cases could be qualified as $f$-ensemble (force is fixed while fluctuating the height of the chain end is measured) and as $h$-ensemble ($h$ is fixed while one measures the fluctuating $f$). Recently these two ways of descriptions as well as their interrelation were discussed for the case of a phantom polymer chain by Skvortsov et al. [@Skvortsov]. In recent papers [@PRER; @Bhattacharya] we studied extensively the desorption of a single tethered [*self-avoiding*]{} polymer from a flat solid substrate with an external force applied to a free chain’s end in the $f$-ensemble.
In the present paper we consider the detachment process of a single self-avoiding polymer chain, keeping the distance $h$ between the free chain’s end and the substrate as the control parameter. We derive analytical results for the main observables which characterize the detachment process. The mean value as well as the probability distribution function (PDF) of the order parameter are presented in close analytical expressions using the Grand Canonical Ensemble (GCE) method[@Bhattacharya]. The basic force-height relationship which describes the process of polymer detachments by pulling, that is, the relevant equation of state for this system is also derived both for extendible as well as for rigid bonds and shown to comply well with our Monte Carlo simulation results. We demonstrate also that a number of properties behave differently in the vicinity of the phase transition, regarding which of the two equivalent ensembles is used as a basis for the study of systems’s behavior.
Single chain adsorption: using distance as a control parameter {#Theory}
==============================================================
Deformation of a tethered chain
-------------------------------
Before considering the adsorption-desorption behavior of a polymer in terms of the chain end distance $h$, we first examine how a chain tethered to a solid surface responds to stretching. This problem amounts to finding the chain free end probability distribution function (PDF) $P_{N}(h)$ where $N$ is the chain length, i.e., the number of beads. The partition function of such a chain at fixed distance $h$ of the chain end from the anchoring plane is given by $$\begin{aligned}
\Xi_{\rm tail}(N, h) = \frac{\mu_{3}^{N}}{N^{\beta}} \: l_0 P_{N} (h)
\label{Partition_Tail}\end{aligned}$$ where $\beta = 1 - \gamma_1$ and the exponent $\gamma_1 =
0.680$[@Vanderzande]. Here $\mu_3$ is a model dependent connective constant (see e.g. ref.[@Vanderzande]). In Eq. (\[Partition\_Tail\]) $l_0$ denotes a short-range characteristic length which depends on the chain model. Below we discuss the chain deformation within two models: the bead-spring (BS) model for elastic bonds and the freely jointed bond vectors (FJBV) model where the bonds between adjacent beads are considered rigid.
### Bead-spring model
The form of $P_N(h)$ has been discussed earlier [@Binder; @Eisenriegler] and later used in studies of the monomer density in polymer brushes [@Kreer]. Here we outline this in a way which is appropriate for our purposes. The average end-to-end chain distance scales as $R_N = l_0 N^{\nu}$, where $l_0$ is the mean distance between two successive beads on a chain and $\nu \approx
0.588$ is the Flory exponent [@Grosberg]. The short distance behavior, $h
\ll R_N$, is given by $$\begin{aligned}
P_N(h) \propto \left(\frac{h}{R_N}\right)^{\zeta}
\label{Short}\end{aligned}$$ where the exponent $\zeta \approx 0.8$. For the long distance behavior, $h/R_N \gg 1$, we assume, following Ref.[@Kreer], that the PDF of the end-to-end vector ${\bf r}$ is given by des Cloizeaux’s expression [@Cloiseaux] for a chain in the bulk : $P_N({\bf r})=(1/R_N) \: F({\bf r}/R_N)$ where the scaling function $F(x)
\propto x^t \: \exp [-D x^\delta]$, and the exponents $t=(\beta-d/2+\nu
d)/(1-\nu)$ , $\delta=1/(1-\nu)$. Here and below $d$ denotes the space dimensionality. One should emphasize that the presence of a surface is manifested only by the replacement of the universal exponent $\gamma$ with another universal exponent $\gamma_1$ (as compared to the pure bulk case!). By integration of $P_N({\bf r})$ over the $x$ and $y$ coordinates while $h$ is measured along the $z-$coordinate, one obtains $P_N(h) \propto
(h/R_N)^{2+t-\delta} \:\: \exp [-D(h/R_N)^{\delta}]$. As the long distance behavior is dominated mainly by the exponential function while the short distance regime is described by Eq. (\[Short\]), we can approximate the overall behavior as $$\begin{aligned}
P_N(h) = \frac{A}{R_N} \: \left(\frac{h}{R_N}\right)^{\zeta} \: \exp \left[-D
\left(\frac{h}{R_N}\right)^{\delta}\right]
\label{Overall}\end{aligned}$$
![Probability distribution $P_N(h)$ of chain end positions $h$ above the grafting plane for a polymer with $N = 128$ monomers at zero strength of the adsorption potential $\epsilon = 0.0$. In the inset the MC data for $P_N(h)$ (solid black line) is compared to the theoretic result, Eq. (\[Overall\]). Dashed line denotes the expected slope of $\zeta \approx 0.78$ of the probability distribution for small heights.[]{data-label="pdf_ends"}](P_h.eps)
A comparison of the distribution, Eq. (\[Overall\]), with our simulation data is shown in Fig. \[pdf\_ends\]. The constants $A$ and $D$ in Eq. (\[Overall\]) can be found from the conditions: $\int P_N (h) d h =1$ and $\int h^2 P_N (h) d h =R_N^2$. This leads to: $$\begin{aligned}
A = \delta
\left[\Gamma\left(\frac{1+\zeta}{\delta}\right)\right]^{-(1+\zeta)/2}
\left[\Gamma\left(\frac{3+\zeta}{\delta}\right)\right]^{-(1-\zeta)/2}
\label{A}\end{aligned}$$ and $$\begin{aligned}
D = \left[\Gamma\left(\frac{3+\zeta}{\delta}\right)\right]^{\delta/2}
\left[\Gamma\left(\frac{1+\zeta}{\delta}\right)\right]^{-\delta/2}
\label{D}\end{aligned}$$ where $\delta \approx 2.43$ and $\zeta \approx 0.8$. One gets thus the estimates $A\approx 2.029$ and $D \approx 0.670$.
The free energy of the tethered chain with a fixed distance $h$ takes on the form $F_{\rm tail}(N, h) = - k_BT \ln \Xi_{\rm tail} (N, h)$ where $k_B$ denotes the Boltzmann constant. By making use of Eqs. (\[Partition\_Tail\]) and (\[Overall\]) the expression for the force $f_N$, acting on the end-monomer when kept at distance $h$ is given by $$\begin{aligned}
f_N = \frac{\partial}{\partial h} \: F_{\rm tail}(N, h)
= \frac{k_BT}{R_N} \:\left[ \delta D \left( \frac{h}{R_N}\right)^{\delta-1} -
\zeta \left( \frac{R_N}{h}\right)\right]
\label{Deformation}\end{aligned}$$ One should note that at $h/R_N \gg 1$ we have $h \propto R_N (R_N
f_N/k_BT)^{1/(\delta-1)}$ which, after taking into account that $\delta^{-1} =
1-\nu$, leads to the well known Pincus deformation law: $h \propto l_0 N (l_0
f_N/k_BT)^{1/\nu -1}$ [@de; @Gennes]. Within the framework of this approximation the (dimensionless) elastic energy reads $U_{\rm el}/k_BT = - N
(l_0 f_{N}/k_BT)^{1/\nu}$. In result the corresponding free energy of the chain tail is given by $$\begin{aligned}
\frac{F_{\rm tail}}{k_BT} = - N \left(\frac{l_0 f_N}{k_BT}\right)^{1/\nu} - N
\ln \mu_3
\label{Free_Energy_BS}\end{aligned}$$
Eq. (\[Deformation\]) indicates that there exists a height $h_0 = (\zeta/\delta D)^{1/\delta} R_N$ over the surface where the force $f_N$ changes sign and becomes negative (that is, an entropic repulsion dominates). According to Eq. (\[Deformation\]) the force diverges as $f_N \propto - k_BT/h$ upon further decrease of the distance $h$.
### Freely jointed chain
It is well known [@Grosberg] that the Pincus law, Eq. (\[Deformation\]), describes the deformation of a linear chain at intermediate force strength, $1/N^{\nu} \ll l_0 f_N/k_BT \leq 1$. Direct Monte Carlo simulation results indicate that, depending on the model, deviations from Pincus law emerge at $h/R_N \ge 3$ (bead-spring off-lattice model) [@Lai], or $h/R_N \ge 6$ (Bond Fluctuation Model [@Wittkop]). In such “overstretched” regime (when the chain is stretched close to its contour length) one should take into account that the chain bonds cannot expand indefinitely. This case could be treated, therefore, within the simple freely jointed bond vectors (FJBV) model [@Lai; @Schurr] where the bond length $l_0$ is fixed. In this model the force - deformation relationship is given by $$\begin{aligned}
f_N = \frac{k_BT}{l_0} \: {\cal L}^{-1} \left( \frac{h}{l_0 N}\right)
\label{L}\end{aligned}$$ where ${\cal L}^{-1}$ denotes the inverse Langevin function ${\cal L}(x) =
\coth(x) - 1/x$ and $l_0$ is the fixed bond vector length. We discuss the main results pertaining to the FJBV model in Appendix A. The elastic deformation energy reads $U_{\rm el}/k_BT = - (l_0 f_N/k_BT) \sum_{i=1}^{N} <\cos \theta_i>
= - N (l_0 f_N/k_BT) {\cal L}(l_0 f_N/k_BT)$, where $\theta_i$ is the average polar angle of the $i$-th bond vector (see Appendix A). Thus the corresponding free energy of the chain tail for the FJBV model reads $$\begin{aligned}
\frac{F_{\rm tail}}{k_BT} = - N {\cal G}\left(\frac{l_0 f_N}{k_BT} \right) -
N \ln \mu_3
\label{Free_Energy_FJBV}\end{aligned}$$ where we have used the notation ${\cal G}(x) = x {\cal L}(x) = x \coth (x) - 1$. Now we are in a position to discuss the pulling of the adsorbed chain controlled by the chain height $h$.
Pulling controlled by the chain end position
--------------------------------------------
Consider now an adsorbed chain when the adsorption energy per monomer is sufficiently large, $\varepsilon \geq \varepsilon_c$, where $\varepsilon_c$ denotes a corresponding [*critical*]{} energy of adsorption. Below we will also use the notation $\epsilon = \varepsilon/k_B T$ for the dimensionless adsorption energy. The problem of force-induced polymer desorption could be posed as follows: how is the process of polymer detachment governed by the chain end position $h$? Figure \[Tearing\]a gives a schematic representation of such a system, and the situation in a computer experiment, as shown in the snapshot Fig. \[Tearing\]b, is very similar.
![ (a) Schematic graph of an adsorbed polymer chain, partially detached from the plane by an external force which keeps the last monomer at height $h$. The total chain is built up from a tail of length $M$ and an adsorbed part of length $N - M$. The force $f_M$ acting on the chain end is conjugated to $h$, i.e., $f_M = \partial F_{\rm tail}/\partial h$. (b) A snapshot from the MC simulation: $N = 128,\; h = 25.0,\; \epsilon = 4.0$ and $\langle f \rangle =
6.126$. []{data-label="Tearing"}](ads_desorp.eps "fig:") ![ (a) Schematic graph of an adsorbed polymer chain, partially detached from the plane by an external force which keeps the last monomer at height $h$. The total chain is built up from a tail of length $M$ and an adsorbed part of length $N - M$. The force $f_M$ acting on the chain end is conjugated to $h$, i.e., $f_M = \partial F_{\rm tail}/\partial h$. (b) A snapshot from the MC simulation: $N = 128,\; h = 25.0,\; \epsilon = 4.0$ and $\langle f \rangle =
6.126$. []{data-label="Tearing"}](pull2.eps "fig:")
As is evident from Fig. \[Tearing\]a, the system is built up from a tail of length $M$ and an adsorbed portion of length $N-M$. The adsorbed part can be treated within the GCE approach [@Bhattacharya]. In our earlier treatment [@Bhattacharya] it was shown that the free energy of the adsorbed portion is $F_{\rm ads} = k_BT (N - M)\ln z^{*}(\epsilon)$, where the fugacity per adsorbed monomer $z^{*}(\epsilon)$ depends on $\epsilon$ and can be found from the basic equation $$\begin{aligned}
\Phi(\alpha, \mu_3 z^{*}) \: \Phi (\lambda, \mu_2 w z^{*})= 1
\label{Basic}\end{aligned}$$ The so called [*polylog function*]{} in Eq. (\[Basic\]) is defined as $\Phi
(\alpha, z) = \sum_{n=1}^{\infty} z^n/n^{\alpha}$ and the connective constants $\mu_3$, $\mu_2$ in three and two dimensional space have values which are model dependent [@Vanderzande]. The exponents $\alpha = 1+ \phi$ and $\lambda =
1-\gamma_{d=2}$ where $\phi \approx 0.5$ is the [*crossover exponent*]{} which governs the polymer adsorption at criticality, and in particular, the fraction of adsorbed monomers at the critical adsorption point (CAP) $\epsilon =
\epsilon_{c}$. The constant $\gamma_{d=2} = 1.343$ [@Vanderzande]. Finally $w = \exp(\epsilon)$ is the additional statistical weight gained by each adsorbed segment.
In equilibrium, the force conjugated to $h$, that is, $f_M = \partial F_{\rm
tail}/\partial h$, should be equal to the [*chain resistance force to pulling*]{} $f_{\rm p} = (k_BT/l_0) {\cal F}(\epsilon)$ (where ${\cal F}(\epsilon)$ is a scaling function depending only on $\epsilon$), i.e., $$\begin{aligned}
f_M = \begin{cases}\frac{k_BT}{R_M} \:\left[ \delta D \left(
\frac{h}{R_M}\right)^{\delta-1} -
\zeta \left( \frac{R_M}{h}\right)\right] = f_{\rm p} &\mbox{, for BS-model}\\
\\
\frac{k_BT}{l_0} \: {\cal L}^{-1}\left(\frac{h}{l_0 M}\right) = f_{\rm p}
&\mbox{,
for FJBV-model}
\end{cases}
\label{Mech_Euqilibrium}\end{aligned}$$ The resisting force $f_{\rm p}$ holds the last adsorbed monomer on the adhesive plane (see again Fig.\[Tearing\]a whereby this monomer is shown to experience a force $f_{M}$). One should emphasize that the force $f_{\rm p}$ [*stays constant*]{} in the course of the pulling process as long as one monomer, at least, is adsorbed on the surface. Thus $f_{\rm p}$ corresponds to a [*plateau*]{} on the deformation curve (force $f$ vs. chain end position $h$). The adsorbed monomer (see Fig. \[Tearing\]) has a chemical potential, $\mu_{\rm ads}=\ln
z^{*}$, which in equilibrium should be equal to the chemical potential of a desorbed monomer in the tail, $\mu_{\rm des}= \partial (F_{\rm
tail}/k_BT)/\partial N$. The expression for $F_{\rm tail}$ depends on the model and is given either by Eq.(\[Free\_Energy\_BS\]) for the BS-model or by Eq.(\[Free\_Energy\_FJBV\]) in the case of FJBV-model. Taking this into account the condition $\mu_{\rm ads}=\mu_{\rm des}$ leads to the following “plateau law” relationship $$\begin{aligned}
\frac{l_0 \: f_{\rm p}}{k_BT} = \begin{cases}\left|\ln [\mu_3
z^{*}(\epsilon)]\right|^{\nu} &\mbox{, for BS-model}\\
\\
{\cal G}^{-1}\left(\left|\ln [\mu_3 z^{*} (\varepsilon)]\right|\right) &\mbox{,
for
FJVB-model}
\end{cases}
\label{Local_Equilibrium}\end{aligned}$$ where ${\cal G}^{-1}(x)$ stands for the inversion of the function ${\cal G}(x) =
x \coth (x) - 1$. One should note that Eq.(\[Local\_Equilibrium\]) coincides with Eq.(3.16) in Ref. [@Bhattacharya] which determines the detachment line in the pulling process controlled by the applied force. Close to the critical point $\epsilon_c$, the plateau force $f_{\rm p}$ goes to zero. Indeed, since in the vicinity of the critical point $\ln [\mu_3 z^{*}(\epsilon)] \propto -
(\epsilon - \epsilon_c)^{1/\phi}$ (see ref.[@Bhattacharya]), and ${\cal
G}^{-1} (x)\approx (3 x)^{1/2}$, one may conclude that $f_{\rm p} \propto
(\epsilon - \epsilon_c)^{\nu/\phi}$ for the BS-model and $f_{\rm p} \propto
(\epsilon - \epsilon_c)^{1/2 \phi}$ for the FJVB-model.
One can solve Eq.(\[Mech\_Euqilibrium\]) with respect to $M$ (taking into account that $h \gg R_M$), and arrive at an expression for the tail length $$\begin{aligned}
M(h, \epsilon) = \begin{cases} \frac{h}{l_0} \: \left(\frac{k_BT}{l_0
f_{\rm p}}\right)^{1/\nu - 1} &\mbox{, for BS-model}\\
\\
\frac{h}{l_0} \:\left[{\cal L}\left(\frac{l_0
f_{\rm p}}{k_BT}\right)\right]^{-1}&\mbox{, for
FJVB-model}
\end{cases}
\label{M_vs_h}\end{aligned}$$ where the force at the plateau, $f_{\rm p}$, is described by Eq. (\[Local\_Equilibrium\]). If for the degree of adsorption one uses as an order parameter the fraction of chain contacts with the plane, $n = N_s/N$, where $N_s$ is the number of monomers on the surface, one can write [@Bhattacharya] $$\begin{aligned}
n = - \frac{1}{k_BT N} \frac{\partial}{\partial \epsilon} \left(F_{\rm ads} +
F_{\rm tail}\right)
\label{Fraction}\end{aligned}$$ where $F_{\rm ads}$ and $F_{\rm tail}$ are free energies of the adsorbed and desorbed portions of the chain respectively. The free energy $F_{\rm ads} =
k_BT [ N - M(h, \epsilon)] \ln z^{*}(\epsilon)$ whereas $F_{\rm tail} = k_BT
\mu_{\rm des} M(h, \epsilon)$ (recall that $\mu_{\rm des}$ is the chemical potential of a desorbed monomer). After substitution of these expressions in eq. (\[Fraction\]) and taking into account that in equilibrium $\mu_{\rm ads} =
\mu_{\rm des}$ (the sequence of operations is important: taking the derivative with respect to $\epsilon$ is to be followed by the condition $\mu_{\rm ads}
=\mu_{\rm des}$) so one gets $$\begin{aligned}
n = - \left[1 - \frac{M(h, \epsilon)}{N}\right] \frac{\partial \ln
z^{*}(\epsilon)}{\partial
\epsilon}
\label{Order_Parameter}\end{aligned}$$ i.e. the order parameter $n$ is defined by the product of monomer fraction in the adsorbed portion, $1 - M/N$, and the fraction of surface contacts in this portion, $ - \partial \ln z^{*}/\partial \epsilon$. The expressions for the order parameter can be recast in the form $$\begin{aligned}
n = \left|\frac{\partial \ln
z^{*}(\epsilon)}{\partial
\epsilon}\right| \times \begin{cases}
1 - \frac{h}{c_1 l_0 N} \left(\frac{k_BT}{l_0 f_{\rm
p}}\right)^{1/\nu - 1} &\mbox{, BS-model}\\
\\
1 - \frac{h}{c_2 l_0 N} \left[{\cal L} \left(\frac{l_0 f_{\rm
p}}{k_BT}\right)\right]^{-1}
&\mbox{, FJVB-model}
\end{cases}
\label{Order_Parameter1}\end{aligned}$$ Here $c_1$ and $c_2$ are some constants of the order of unity.
As one can see from Eq. (\[Order\_Parameter1\]), the order parameter decreases linearly and steadily with $h/N$. This behavior is qualitatively different from the abrupt jump of $n$ when the pulling force $f$ is changed as a control parameter. In Section \[MC\_results\] we will show that this predictions is in a good agreement with our MC - findings. The transition point on the $n$ vs. $h$ curve corresponds to total detachment, $n = 0$. The corresponding distance $h$ will be termed “detachment height” $h_D$. The dependence of $h_D$ on the adsorption energy $\epsilon$ can be obtained from Eq.(\[Order\_Parameter1\]) where $n$ is set to zero, i.e. $$\begin{aligned}
\frac{h_D}{l_0 N} = \begin{cases}
\left( \frac{l_0 f_{\rm p}}{k_BT}\right)^{1/\nu -1} &\mbox{, BS-model}\\
\\
{\cal L}\left( \frac{l_0 f_{\rm p}}{k_BT}\right) &\mbox{, FJVB-model}
\end{cases}
\label{Detachment_Line}\end{aligned}$$ where again $f_{\rm p}$ as a function of $\epsilon$ is given by Eq. (\[Local\_Equilibrium\]). The line given by Eq. (\[Detachment\_Line\]), is named “detachment line”. It corresponds to an adsorption - desorption polymer transition which appears as of [*second order*]{} since this order parameter $n$ goes to zero continuously as $h$ increases. One should emphasize, however, that this “detachment” transition has the same nature as the force-induced desorption transition [@Bhattacharya] in the $f$-ensemble where the pulling force $f$, rather than the distance $h$, is fixed and used as a control parameter. This phase transformation is known to be of [*first order*]{}.
It is easy to understand (cf. with Eq.(\[M\_vs\_h\]) ) that the condition $M(h,
\varepsilon) = N$ corresponds to the detachment line as well as to a terminal point of the force plateau. It can be seen in the MC - simulation results in Section\[MC\_results\], Fig. \[force\].
Probability distribution $P(K)$ of the number of adsorbed monomers {#Theory2}
==================================================================
The grand canonical ensemble (GCE) method, which has been used in our recent paper [@Bhattacharya], is a good starting point to calculate the probability distribution function $P(K)$ of the adsorbed monomers number $K$. According to this approach, the GCE-partition function of an adsorbed chain has the form $$\Xi (z, w) = \sum_{N=1}^{\infty} \: \sum_{K=0}^{\infty} \Xi_{N, K} \: z^N \:
w^K = \dfrac{V_0 (wz) \: Q (z)}{1 - V(w z)\: U(z)}
\label{GCE}$$ where $z$ and $w = \exp(\epsilon)$ are the fugacities conjugated to chain length $N$ and to the number of adsorbed monomers $K$, respectively. In Eq.(\[GCE\]) $U(z)$, $V(wz)$ and $Q(z)$ denote the GCE partition functions for loops, trains and tails, respectively. The building block adjacent to the tethered chain end corresponds to $V_0 (wz) = 1 + V (wz)$. It has been shown [@Bhattacharya] that the functions $U(z)$, $V(wz)$ and $Q(z)$ can be expressed in terms of polylog functions, defined in the paragraph after Eq. (\[Basic\]), as $U (z) =
\Phi (\alpha, \mu_3 z)$, $V (wz) = \Phi (1 - \gamma_{d=2}, \mu_2 w z)$ and $Q
(z) = 1 + \Phi (1 - \gamma_{1}, \mu_3 z)$, where $\mu_3$ and $\mu_2$ are $3d$- and $2d$ - connective constants respectively. By making use of the inverse Laplace transformation of $\Xi (z, w)$ with respect to $z$ (see, e.g. [@Rudnick]) the (canonical with respect to the chain length $N$) partition function is obtained as $$\Xi_{N}(w) = \sum_{K=0}^{\infty} \: \Xi_{N,K} \: {w}^{K} =\exp[-N
\ln z^{*} (w)]
\label{GC}$$ where $z^{*}(w)$ is a simple pole of $\Xi (z, w)$ in complex $z$-plane given by equation $V(w z^*) U(z^*) = 1$, i.e. by Eq.(\[Basic\]).
The (non-normalized) probability for the chain to have $K$ adsorbed monomers is $P(K) \propto \Xi_{N,K} \: {\rm e}^{\epsilon K} = \exp[- {\cal F}(K)/k_BT]$, where ${\cal F}(K)$ is the free energy at given $K$. It is convenient to redefine the fugacity $w$ as $w \rightarrow \xi w$, (as well as $w_c
\rightarrow \xi w_c$) where $\xi$ is an arbitrary complex variable. Then the probability $P(K)$ can be found as the coefficient of $\xi^K$ in the function $\Xi_{N} (\xi w)$ expansion in powers of $\xi$. Therefore $$P(K) = \exp[- {\cal F}(K)/k_BT] = \frac{1}{2\pi i} \: \displaystyle \oint
\dfrac{\exp[- N \ln z^{*} (\xi \: w)]}{\xi^{K+1}} \: d \xi
\label{Integral}$$ where the contour of integration is a closed path in the complex $\xi$ plane around $\xi =0$. (see e.g. [@Rudnick]). To estimate the integral in Eq.(\[Integral\]) we use the [*steepest descent method*]{} [@Rudnick].
For large $N$ the main contribution to the integral in Eq. (\[Integral\]) is given by the saddle point $\xi = \xi_0$ of the integrand which is defined by the extremum of the function $g(\xi) = - \ln z^{*}(\xi \: w) -
[(K+1)/N] \ln \xi$, i.e., by the condition $$\dfrac{K+1}{\xi_0} = - \left. N \: \dfrac{\partial \ln z^{*}}{\partial
\xi}\right|_{\xi = \xi_0}
\label{SP}$$ The integral is dominated by the term $\exp [- N \ln z^{*}(\xi_0 w) - (K+1)
\ln \xi_0]$. Another contribution comes from the integration along the steepest descent line. As a result one obtains $$\begin{aligned}
P(K) \propto \dfrac{\exp [- N \ln z^{*}(\xi_0 \: w) - (K+1) \ln
\xi_0]}{\sqrt{N\left[\left(\dfrac{N}{K}\right)
\left[\dfrac{\partial\ln z^{*}}{\partial \xi}\right]_{\xi = \xi_0}^2 -
\left.\dfrac{\partial^2 \ln z^{*}}{\partial \xi^2}\right|_{\xi = \xi_0}\right]}}
\label{Method}\end{aligned}$$ The validity of the steepest descent method is ensured by the condition of the second derivative $N g''(\xi_0)$ being large, which yields $$\begin{aligned}
N g''(\xi_0) = N\left[\left(\dfrac{N}{K}\right)
\left[\dfrac{\partial\ln z^{*}}{\partial \xi}\right]_{\xi = \xi_0}^2 -
\left.\dfrac{\partial^2 \ln z^{*}}{\partial \xi^2}\right|_{\xi = \xi_0}\right]
\gg 1
\label{Valid}\end{aligned}$$ A more explicit calculation whithin this method can be performed in the vicinity of the critical point $\epsilon = \epsilon_c$.
PDF of the number of adsorbed monomers close to the critical point of adsorption
--------------------------------------------------------------------------------
In this case the explicit form of $\ln z^{*}$ is known [@Bhattacharya] and after the redefinition of the fugacity, $w \rightarrow \xi w$, it reads $$\begin{aligned}
\ln z^{*} (\xi w) = - a_1 (w - w_c)^{1/\phi} \xi^{1/\phi} - \ln \mu_3
\label{Z_Star}\end{aligned}$$ where $a_1$ is a constant of the order of unity. The critical adsorption fugacity $w_c = {\exp (\epsilon_c)}$ is defined by the equation $$\begin{aligned}
\zeta (\alpha)\: \Phi (1 - \gamma_{d=2}, \mu_2 w_c/\mu_3) = 1
\label{Critical}\end{aligned}$$ with $\zeta (x)$ denoting the Riemann zeta-function.
By using Eq.(\[Z\_Star\]) in Eq.(\[SP\]) one arrives at the expression for the saddle point $$\begin{aligned}
\xi_0 = \left( \dfrac{K}{N}\right)^{\phi} \dfrac{a_2}{w - w_c}
\label{SP_Solution}\end{aligned}$$ where $a_2=(\phi/a_1)^{\phi}$. Using Eq. (\[Z\_Star\]) and Eq. (\[SP\_Solution\]) in Eq. (\[Method\]) then yields the expression for PDF $$\begin{aligned}
P(K) \propto [(w - w_c) {\rm e}^{a_2}]^{K} \left(\dfrac{K}{N}\right)^{- \phi K
- 1/2} \: \dfrac{\mu_{3}^{N}}{N^{1/2}}
\label{Distr1}\end{aligned}$$ For reasonably large $K$ and after normalization one arrives at the final expression for the PDF $$\begin{aligned}
P(K) = \dfrac{\eta^K}{{\cal C}(\eta) K^{\phi K} }
\label{Distr3}\end{aligned}$$ where we have introduced the usual adsorption scaling variable $\eta = b_1 (w -
w_c)N^\phi$ ($b_1$ is a constant of the order of unity; see e.g. [@Hsu]) as well as the normalization constant ${\cal C}$: $$\begin{aligned}
{\cal C}(\eta) = \sum_{K=1}^{N} \: \dfrac{\eta^K}{K^{\phi K}}
\label{Normalization}\end{aligned}$$ One can readily see that the width of the distribution increases with $w$ or with $\epsilon$. To this end one may directly calculate the fluctuation variance as follows: $$\begin{aligned}
\overline{(K - \overline{K})^2} = - N \dfrac{\partial^2 \ln z^{*}(w)}{\partial
(\ln w )^2} \propto N w (w/\phi - w_c) (w - w_c)^{1/\phi - 2}
\label{Variance_1}\end{aligned}$$ where the expression for $\ln z^{*}(w)$ given by Eq.(\[Z\_Star\]) (where also $\xi=1$ ) has been used. Taking into account that $\phi \approx
0.5$, it becomes clear that the variance really grows with $w$.
The validity of the steepest descent method is ensured by the condition Eq.(\[Valid\]). Using Eq.(\[Z\_Star\]) and Eq. (\[SP\_Solution\]), one can verify that this criterion holds when $N (w - w_c)^2 (K/N)^{1 - 2\phi} \gg 1$. We recall that $\phi \approx 0.5$, so that $(w - w_c) N^{1/2} \gg 1$. In result the criterion becomes $$\begin{aligned}
N^{-1/2} \ll (w - w_c) \ll 1
\label{Criterion}\end{aligned}$$ In the deep adsorption regime this condition might be violated. Nevertheless, the steepest descent method could still be used there, provided that the appropriate solution for $z^{*}(w)$ (see Eq.(\[Basic\])) is chosen.
The regime of deep adsorption
-----------------------------
In the deep adsorption regime one should use the the solution for $z^{*} (w)$ which was also discussed in ref. [@Bhattacharya]. Namely, in this case $$\begin{aligned}
z^{*} (w) \approx \dfrac{1}{\mu_2 w} \left[1 - \left( \dfrac{\mu_3}{\mu_2
w}\right)^{1/(1-\lambda)} \right]
\label{Deep_Ads}\end{aligned}$$ With Eq. (\[Deep\_Ads\]) the mean value $\overline{K} = - N \partial \ln
z^{*}(w)/\partial \ln w$ can be written as $$\begin{aligned}
\dfrac{\overline{K}}{N} = 1 - \dfrac{1}{1 - \lambda} \left(
\dfrac{\mu_3}{\mu_2 w}\right)^{1/(1-\lambda)}
\label{K}
\end{aligned}$$ thus $\overline{K}$ tends to $N$ with $w$ growing as it should be. The variance of the fluctuations within the GCE then becomes $$\begin{aligned}
\overline{(K - \overline{K})^2} = - N \dfrac{\partial^2 \ln
z^{*}(w)}{\partial (\ln w )^2} = \dfrac{N}{(1 - \lambda)^2}
\left(\dfrac{\mu_3}{\mu_2 w}\right)^{1/(1-\lambda)},
\label{Variance_2}
\end{aligned}$$ i.e., the fluctuations decrease when the adsorption energy grows. Comparison of this result with the result given by Eq. (\[Variance\_1\]) leads to the important conclusion that the fluctuations of the number of adsorbed monomers first grow with $\epsilon$, attain a maximum, and finally decrease with increasing surface adhesion $\epsilon$. The position of the maximum reflects the presence of finite-size effects, and, as the chain length $N\to \infty$, this maximum occurs at the CAP.
Consider now the steepest descent treatment for the deep adsorption regime. With Eq.(\[Deep\_Ads\]) (after the rescaling $w \rightarrow \xi w$) in Eq.(\[SP\]), the saddle point becomes $$\begin{aligned}
\xi_0 = b_2 \left(\dfrac{\mu_3}{\mu_2 w} \right) \left(1 -
\dfrac{K}{N}\right)^{-(1 - \lambda)}
\label{SP_Deep}
\end{aligned}$$ where $b_2 = (1 - \lambda)^{-(1 - \lambda)}$ (recall that $1-\lambda =
\gamma_{d=2} = 1.343$). The main contribution comes from the exponential term in Eq.(\[Method\]) which is given by $$\begin{aligned}
\exp\left[-N \ln z^{*} (w \xi_0) - (K+1) \ln \xi_0\right] = (\mu_2 w)^{N}
\dfrac{\left(b_2 \mu_3 {\rm e}^{1-\lambda} N^{1-\lambda}/\mu_2
w\right)^{N-K}}{\left(N - K\right)^{(1-\lambda)(N-K)}}
\label{Exponential}\end{aligned}$$ For $N-K \gg 1$ the expression for the non-normalized PDF takes on the form $$\begin{aligned}
P(K) \propto \dfrac{\left(b_3 \mu_3 N^{1-\lambda}/\mu_2
w\right)^{N-K} }{(N-K)^{(1-\lambda) (N - K)} }\end{aligned}$$ where $b_3$ is a constant of the order of unity. Normalization of this distribution yields $$\begin{aligned}
P(K) = \dfrac{\chi^{N-K}
}{{\cal R}(\chi) (N-K)^{(1-\lambda)(N - K)} }
\label{PDF_Deep}\end{aligned}$$ where the parameter $\chi = b_3 (\mu_3/\mu_2 w) N^{1-\lambda}$ and the normalization constant $$\begin{aligned}
{\cal R}(\chi) = \sum_{K=1}^{N-1} \dfrac{\chi^{N-K}
}{(N-K)^{(1-\lambda)(N - K)} }.
\label{R}\end{aligned}$$ The validity condition, Eq.(\[Valid\]), in the deep adsorption regime (after substitution of Eq. (\[Deep\_Ads\]) into Eq.(\[Valid\])) requires $$\begin{aligned}
(N - K)^{3/2 - \lambda} \gg (\mu_3/\mu_2 w) N^{1 - \lambda},
\label{Valid_Deep}\end{aligned}$$ i.e., $K$ should not be very close to $N$. In Fig. \[pdf\_theory\]a we show the PDF of the number of chain contacts, $P(K)$, for a free chain without pulling and several adsorption strengths of the substrate. One can readily verify that visually the shape of $P(K)$ resembles very much a Gaussian distribution at moderate values of $\epsilon_c\approx 1.7 < \epsilon < 6.0 $. The PDF variance goes through a sharp maximum at $\epsilon \gtrsim \epsilon_c$ and then declines, as expected from Eq. (\[Variance\_2\]).
One should note that in the $f$-ensemble (where the force $f$ and not the distance $h$ acts as a controll parameter [@Bhattacharya]) the order parameter $n = \overline{K}/N$ undergoes a jump at the detachment adsorption energy $\epsilon_D$. This means that $N (\partial^2 \ln z^{*}/\partial
\epsilon^2)_{\epsilon_D} = - (\partial \overline{K}/\partial
\epsilon)_{\epsilon_D} \rightarrow - \infty$. Thus at the detachment point the variance of the fluctuations $\overline{(K - \overline{K})^2} = - N (\partial^2
\ln z^{*}/\partial \epsilon^2)_{\epsilon_D} \rightarrow \infty$, which practically means that for chains of a finite length the distribution at $\epsilon = \epsilon_D$ becomes very broad, in sharp contrast to Eq. (\[Variance\_2\]). This has indeed been observed in our MC-simulation results (see Fig.12 in ref. [@Bhattacharya]).
$P(K)$ distribution in the subcritical regime $w < w_c$ of underadsorption
--------------------------------------------------------------------------
In the subcritical regime, $w < w_c$, the fraction of adsorbed points (order parameter) $n = \overline{K}/N = 0$, in the thermodynamic limit. Nevertheless, $\overline{K} \neq 0$ and one can examine the form of the PDF $P(K)$. At $w <
w_c$ the solution for $z^* (w)$ (the simple pole of $\Xi (z, w)$ in the complex $z$-plane) does not exist because $V(w z) U(z) < 1$ (see Eq.(\[GCE\])). However, the tail GCE-partition function $Q(z) = 1 + \Phi (1- \gamma_1, \mu_3
z) \propto \Gamma (\gamma_1)/(1 - \mu_3 z)^{\gamma_1}$ has a [*branch point*]{} at $z = 1/\mu_3$ (see Eq. (A 11) in ref. [@Bhattacharya]) which governs the coefficient at $z^N$, i.e., the partition function $\Xi_{N}(w)$. The calculation (following Section 2.4.3 in ref. [@Rudnick]) yields $$\begin{aligned}
\Xi_{N}(w) = \dfrac{1 + \Phi (\lambda, \mu_2 w/\mu_3)}{1 - \zeta
(\alpha) \: \Phi (\lambda, \mu_2 w/\mu_3)} \: \mu_3^N \: N^{\gamma_{1}
- 1}
\label{Subcritical_GC}\end{aligned}$$ where $\lambda = 1 - \gamma_{d=2}$ and we have also used that at $z = 1/\mu_3$ the loop and train GCE-partition functions are $U(1/\mu_3)= \Phi (\alpha, 1) = \zeta
(\alpha)$ and $V(w/\mu_3) = \Phi (\lambda, \mu_2 w/\mu_3)$ respectively. This expression has a pole at $w = w_c$ (cf. Eq. (\[Critical\])) which yields the coefficient of $w^K$ , i.e. $\Xi_{N,K}$. Recall that $\Xi_N(w) =
\sum_{K=0}^{\infty} \Xi_{N,K} w^K$, so that $P(K) \propto \Xi_{N,K} w^K$. Expansion of the denominator in Eq. (\[Subcritical\_GC\]) around $w = w_c$ reveals the simple pole as follows $$\begin{aligned}
\Xi_{N}(w) = \dfrac{[1 + 1/\zeta(\alpha)] \: \mu_3^N N^{\gamma_1 -
1}}{\zeta(\alpha) \: \Phi (-\gamma_{d=2}, \mu_2 w_c/\mu_3)} \left(
\dfrac{w_c}{w_c - w}\right).
\label{Pole}\end{aligned}$$ In (\[Pole\]) we have used the relationship $x (d/d x) \Phi (1-\gamma_{d=2},
x) = \Phi(- \gamma_{d=2}, x)$. The coefficient of $w^K$, i.e., $\Xi_{N,K}$, is proportional to $w_c^{- K}$. Therefore $P(K) \propto (w/w_c)^{K} = \exp
[-(\epsilon_c - \epsilon)K]$. Taking the normalization condition $\sum_{K=0}^{N} P(K) = 1$ into account, the final expression for $P(K)$ can be recast in the form $$\begin{aligned}
P(K) = \dfrac{1 - {\exp}[-(\epsilon_c - \epsilon)]}{1 - {\exp}[- N
(\epsilon_c - \epsilon)]} \: {\exp}[- K(\epsilon_c - \epsilon)],
\label{Exp}\end{aligned}$$ i.e., $P(K)$ has a simple exponential form. The calculation of the average $\overline{K} = \sum_{K=0}^{N} K P(K)$ leads to the simple result $\overline{K}
= [\exp (\epsilon_c - \epsilon) - 1]^{-1} \approx 1/(\epsilon_c - \epsilon)$, i.e., $\overline{K} \rightarrow \infty$ at $\epsilon \rightarrow \epsilon_c$. On the other hand we know that $\overline{K} = N^{\phi}$ at $\epsilon =
\epsilon_c$. In order to prevent a divergency at $\epsilon = \epsilon_c$, one should incorporate an appropriate cutoff in the PDF given by Eq.(\[Exp\]). With this the distribution is given by $$\begin{aligned}
P(K) = \dfrac{1 - {\exp}[-(\epsilon_c - \epsilon + 1/N^{\phi})]}{1 - {\exp}[- N
(\epsilon_c - \epsilon + 1/N^{\phi})]} \: {\exp}[- K(\epsilon_c - \epsilon +
1/N^{\phi})].
\label{Exp_Cutoff}\end{aligned}$$ Thus the expression for the average number of adsorbed monomer has the correct limit behavior, i.e., $$\begin{aligned}
\overline{K} = \dfrac{1}{\epsilon_c - \epsilon + 1/N^{\phi}}
\label{Average_Cutoff}\end{aligned}$$
Probability distribution function $P(K)$ in the $h$-ensemble
------------------------------------------------------------
Eventually we examine how the fixed chain-end hight $h$ affects the PDF of the number of contacts $K$. To this end we refer again to Fig. \[Tearing\]a where an adsorbed chain with a fixed height $h$ of the last monomer is depicted. The adsorbed chain consists of a tail of length $M$ and of an adsorbed part with $N
- M$ beads. One should bear in mind that $M$ is function of the control parameters $h$ and of $w = \exp (\epsilon)$ given by eqs.(\[Local\_Equilibrium\]) and (\[M\_vs\_h\]). The partition function of the adsorbed part is then given by $$\begin{aligned}
\Xi_{N} (w) = \sum_{K=0}^{N} \: \Xi_{N, K} \: w^K = \exp\left\lbrace - [N -
M(h, w)] \ln z^{*} (w)\right\rbrace
\label{Adsorbed_Part}\end{aligned}$$ where we took into account that the free energy of the adsorbed portion is given as $F_{\rm ads} = k_B T [N - M(h, w)] \ln z^{*}(w)$ (see Sec. II B).
![\[pdf\_theory\] (a) Probability Distribution Function of the number of chain contacts with the grafting plane, $K$, for different degrees of adhesion $\epsilon$ and no pulling force $f = 0$ as follows from eqs. (\[Distr3\]) and (\[PDF\_Deep\]). The variance of $P(K)$ is shown in the inset for different $\epsilon$. (b) The same as in (a) for $\epsilon = 1.75$ and different heights of the chain end $h$. The change of the variance $\langle K^2\rangle - \langle K\rangle^2$ with $h$ is shown in the inset.](PDF_theor.eps "fig:") ![\[pdf\_theory\] (a) Probability Distribution Function of the number of chain contacts with the grafting plane, $K$, for different degrees of adhesion $\epsilon$ and no pulling force $f = 0$ as follows from eqs. (\[Distr3\]) and (\[PDF\_Deep\]). The variance of $P(K)$ is shown in the inset for different $\epsilon$. (b) The same as in (a) for $\epsilon = 1.75$ and different heights of the chain end $h$. The change of the variance $\langle K^2\rangle - \langle K\rangle^2$ with $h$ is shown in the inset.](PDF_pull.eps "fig:")
As mentioned above, the PDF $P(K) \propto \Xi_{N, K} \: w^K $, so that by means of rescaling $w \rightarrow \xi w$ and $w_c \rightarrow \xi w_c$ the PDF can be found as the coefficient of $\xi^K$, i.e., $$P(K) = \frac{1}{2\pi i} \: \displaystyle \oint
\dfrac{\exp\left\lbrace - [N -
M(h, \xi w)] \ln z^{*} (\xi w)\right\rbrace}{\xi^{K+1}} \: d
\xi
\label{Coefficient}$$ As before, the steepest descent method can be used to calculate the integral in Eq. (\[Coefficient\]). However, in this case the calculations are more complicated and we have relegated most of them to Appendix B. As may be seen there, the saddle point equation can not be solved analytically in the general case but could be treated iteratively for relatively small heights $h$.
One can readily see that for $h = 0$ Eq. (\[PDF\_Pull\]) reduces to Eq.(\[Distr3\]). The PDF, following from Eq. (\[PDF\_Pull\]) is shown in Fig. \[pdf\_theory\]b for $\epsilon = 1.75$ and several values of the height $h$. It can be seen that the curve for $h = 0$ coincides with the curve for $\epsilon = 1.75$ in Fig. \[pdf\_theory\]a as it should be. Evidently, both the mean value $\overline{K}$ and $\overline{K^2} - \overline{K}^2$ decline with growing $h$.
Monte Carlo Simulation Model {#MC_model}
============================
We use a coarse grained off-lattice bead-spring model [@MC_Milchev] which has proved rather efficient in a number of polymers studies so far. The system consists of a single polymer chain tethered at one end to a flat impenetrable structureless surface. The surface interaction is described by a square well potential, $$\label{ads_pot}
U_w(z) = \begin{cases}
\epsilon, & z < r_c \\
0, & z \ge r_c
\end{cases}$$ The strength $\epsilon$ is varied from $2.0$ to $5.0$ while the interaction range $r_c = 0.125$. The effective bonded interaction is described by the FENE (finitely extensible nonlinear elastic) potential: $$U_{FENE}= -K(1-l_0)^2ln\left[1-\left(\frac{l-l_0}{l_{max}-l_0} \right)^2 \right]
\label{fene}$$ with $K=20$, and the fully stretched-, mean-, and minimum bond lengths $l_{max}=1, l_0 =0.7, l_{min} =0.4$. The nonbonded interactions between monomers are described by the Morse potential: $$\label{Morse}
\frac{U_M(r)}{\epsilon_M} =\exp(-2\alpha(r-r_{min}))-2\exp(-\alpha(r-r_{min}))$$ with $\alpha =24,\; r_{min}=0.8,\; \epsilon_M/k_BT=1$. In few cases, needed to clarify the nature of the polymer chain resistance to stretching, we have taken the nonbonded interactions between monomers as purely repulsive by shifting the Morse potential upward by $\epsilon_M$ and removing its attractive branch for $r \ge r_{min}$.
We employ periodic boundary conditions in the $x-y$ directions and impenetrable walls in the $z$ direction. The lengths of the studied polymer chains are typically $64$, and $128$. The size of the simulation box was chosen appropriately to the chain length, so for example, for a chain length of $128$, the box size was $256 \times 256 \times 256$ . All simulations were carried out for constant position of the last monomer $z$-coordinate, that is, in the fixed height ensemble. The the fluctuating force $f$, exerted on the last bead by the rest of the chain was measured and average over about $2000$ measurements. The standard Metropolis algorithm was employed to govern the moves with self avoidance automatically incorporated in the potentials. In each Monte Carlo update, a monomer was chosen at random and a random displacement attempted with $\Delta x,\;\Delta y,\;\Delta z$ chosen uniformly from the interval $-0.5\le
\Delta x,\Delta y,\Delta z\le 0.5$. If the last monomer was displaced in $z$ direction, there was an energy cost of $-f\Delta z$ due to the pulling force. The transition probability for the attempted move was calculated from the change $\Delta U$ of the potential energies before and after the move was performed as $W=exp(-\Delta U/k_BT)$. As in a standard Metropolis algorithm, the attempted move was accepted, if $W$ exceeds a random number uniformly distributed in the interval $[0,1)$.
As a rule, the polymer chains have been originally equilibrated in the MC method for a period of about $ 5 \times 10^5$ MCS after which typically $500$ measurement runs were performed, each of length $2 \times 10^6$ MCS. The equilibration period and the length of the run were chosen according to the chain length and the values provided here are for the longest chain length.
Monte Carlo simulation results {#MC_results}
==============================
In order to verify the theoretical predictions, outlined in Section \[Theory\], we carried out extensive Monte Carlo simulations with the off-lattice model, defined in Section \[MC\_model\]. In these simulations we fix the end monomer of the polymer chain at height $h$ above the adsorbing surface, and measure the (fluctuating) force, needed to keep the last bead at distance $h$, as well as the corresponding fraction of adsorbed monomers $n$. These computer experiments are performed at different strengths $\epsilon$ of the adsorption potential, Eq. (\[ads\_pot\]).
![\[OP\] (a) Order parameter (fraction of adsorbed monomers) $n$ variation with changing height $h/l_0N$ of the fixed chain-end for polymers of length $N = 128$ and different adsorption strength $\varepsilon/k_BT$. (b) Variation of $n$ with $\varepsilon/k_BT$ for different fixed positions of the chain-end $h/l_0N$ as it is seen from MC-data. Insets show the resulting $n-
\epsilon$ relationship at several fixed heights $h$.](op_theor.eps "fig:") ![\[OP\] (a) Order parameter (fraction of adsorbed monomers) $n$ variation with changing height $h/l_0N$ of the fixed chain-end for polymers of length $N = 128$ and different adsorption strength $\varepsilon/k_BT$. (b) Variation of $n$ with $\varepsilon/k_BT$ for different fixed positions of the chain-end $h/l_0N$ as it is seen from MC-data. Insets show the resulting $n-
\epsilon$ relationship at several fixed heights $h$.](op_MC.eps "fig:")
In Figs. \[OP\]a,b we compare the predicted dependence of the order parameter $n$ on the (dimensionless) height $h/l_0N$ at several values of $2.0 \le
\epsilon \le 5.0$ with the results from MC simulations. Note, that the critical point of adsorption $\epsilon_c \approx 1.69$ so we take our measurements above the region of critical adsorption. Typically, both in the analytic results, Fig. \[OP\]a, and in the MC-data, Fig. \[OP\]b, for $N=128$, one recovers the predicted linear decrease of $n$ with growing $h$. Finite-size effects lead to some rounding of the simulation data (in Fig. \[OP\]b these effects are seen to be larger for $N=64$ than for $N=128$) when $n\to 0$ so that the height of detachment $h_D$ is determined from the intersection of the tangent to $n(h)$ and the $x-$axis where
![Phase diagram showing the dependence of the critical height of polymer detachment from the substrate, $h_D/l_0 N$, with the relative strength of adsorption $(\varepsilon - \varepsilon_{c})/k_BT $, where $\varepsilon_{c}/k_BT$ is the critical point of adsorption at zero force. The theoretical curves follow Eq. (\[Detachment\_Line\]).[]{data-label="phase_diag"}](phase.eps)
$n=0$. Evidently, with growing adsorption strength, $\epsilon$, larger height $h_D$ is needed to detach the polymer from the substrate. Thus, one may construct a phase diagram for the desorption transition, which we show in Fig. \[phase\_diag\]. The theoretical prediction is given by eq. (\[Detachment\_Line\]).
In the insets of Fig. \[OP\]a,b we also show the variation of the fraction of adsorbed segments $n$ with adsorption strengths $\epsilon$ for several heights $20 \le h \le 50$ of a chain with $N = 128$. It is evident that, apart from the rounding of the MC data at $n \to 0$, one finds again good agreement between the behavior, predicted by Eq. (\[Order\_Parameter\]), and the simulation results.
The gradual change of $n$ in the whole interval of possible variation of $h$ suggests a pseudo-continuous phase transition, as pointed out in the end of Section \[Theory\], Eq. (\[Order\_Parameter1\]). Of course, if $h$ is itself expressed in terms of the measured pulling force, one would again find that $n$ changes abruptly with varying $f$ at some threshold value $f_D$, indicating a first order transformation from an adsorbed into a desorbed state of the polymer chain.
![PDF of the order parameter (fraction of contacts with the plane) for different adsorption strength $\epsilon$ at zero force. The critical adsorption point (CAP) $\epsilon_{c} \approx 1.67$. The change of the variance $\langle n^2\rangle -
\langle n\rangle^2$ with varying $\epsilon$ is displayed in the inset. []{data-label="pdf_MC"}](op_raw.eps)
It has been pointed out earlier by Skvortsov et al.[@Skvortsov] that, while both the fixed-force and the fixed-height ensembles are equivalent as far as the mean values of observables such as the fraction of adsorbed monomers and other related quantities are concerned, this does not apply to some more detailed properties like those involving fluctuations. Therefore, it is interesting to examine the fluctuations of the order parameter, $n$, for different values of our control parameter $h$, and compare them to theoretical predictions for $P(K)$ from Section \[Theory\]. First we compare the order parameter distribution $P(n)$ for zero force, Fig. \[pdf\_MC\], obtained from our computer experiment, to that, predicted by Eqs. (\[Distr3\]), (\[PDF\_Deep\]), (\[Exp\_Cutoff\]), and displayed in Fig. \[pdf\_theory\]a. It is evident from Fig. \[pdf\_MC\] that for free chains at different strengths of adhesion there is a perfect agreement between analytical and simulational results. For rather weak adsorption $\epsilon = 1.3\pm 1.6 < \epsilon_c=1.67$ in the subcritical regime, one can verify from Fig. \[pdf\_MC\] that $P(n)$ gradually transforms from nearly Gaussian into exponential distribution, as expected from Eq. (\[Exp\_Cutoff\]). For $\epsilon > \epsilon_c$ the distribution width $\langle (n - \langle n \rangle)^2\rangle$ grows and goes through a sharp maximum in the vicinity of $\epsilon_c$, and then drops as $\epsilon$ increases further - compare insets in Fig. \[pdf\_MC\] and Fig. \[pdf\_theory\]a.
![\[PDF\_OP\] (a) Probability distribution $P(n)$ of the order parameter $n$ (i.e., the fraction of adsorbed monomers) for $N = 128$ and $\epsilon = 3.0$ at different heights of the chain-end $h$ over the grafting plane. In the inset we show $P(n)$ at the detachment line $h_D = 54.3$. (b) Variation of the second- and third central moments of $P(n)$ with $h$. The maximum of $\langle (n - \langle n \rangle)^3\rangle$ is reached at $h = h_D$.](ophistogram.eps "fig:") ![\[PDF\_OP\] (a) Probability distribution $P(n)$ of the order parameter $n$ (i.e., the fraction of adsorbed monomers) for $N = 128$ and $\epsilon = 3.0$ at different heights of the chain-end $h$ over the grafting plane. In the inset we show $P(n)$ at the detachment line $h_D = 54.3$. (b) Variation of the second- and third central moments of $P(n)$ with $h$. The maximum of $\langle (n - \langle n \rangle)^3\rangle$ is reached at $h = h_D$.](moments.eps "fig:")
Let us consider now PDF in the presence of pulling. In Fig. \[PDF\_OP\]a we display the distribution $P(n)$ measured in the MC simulations for different heights $h$ and constant adsorption energy $\epsilon =
3.0$. One can readily verify from our results that far enough from the detachment line, $h < h_D$, the shape of $P(n)$ looks like Gaussian and that the second moment, $\langle (\Delta n)^2\rangle$, remains unchanged with varying height $h$. Of course, when $h \to h_D$, the maximum of $P(n)$ shifts to lower values of $n$. Only in the immediate vicinity of $h_D$, where $n \to 0$ and the fluctuations strongly decrease, one observes a significant deviation from the Gaussian shape - cf. the inset in Fig. \[PDF\_OP\]a. The latter is illustrated in more detail in Fig. \[PDF\_OP\]b where we show the measured variation of the second moment, $\langle (\Delta n)^2\rangle$, and that of the third moment, $\langle (n - \langle n \rangle)^3\rangle$ with increasing height $h$. The deviation from Guassinity in $P(n)$, measured by the deviation of the third moment from zero, is localized in the vicinity of the detachment height $h_D$. The corresponding theoretical prediction for the relatively weak adsorption strength is depicted in Fig. \[pdf\_theory\]b. It can be seen that with increasing $h$ the almost Gaussian distribution tends to Poisson-like one. Also the fluctuations decrease with $h$ in accordance with MC-findings.
The force $f$, exerted by the chain on the end-monomer, when the latter is kept at height $h$ above the surface, is one of the main properties which can be measured in experiments carried out within the fixed-height ensemble. Note that $f$ has the same magnitude and opposite sign, regarding the force, applied by the experimentalist. The variation of the force $f$ with increasing height $h$ is shown in Fig. \[force\]a for several values of the adsorption potential $2.0 \le \epsilon \le 5.0$. In Fig. \[force\]a we distinguish between two contributions to the total force $f$, acting on the end bead. The first stems from the quasi-elastic forces of the bonded interaction (FENE) whereas the second contribution is due to the short-range (attractive) interactions between non-bonded monomers (in our model - the Morse potential). A typical feature of the $f - h$ relationship, namely, the existence of a broad interval of heights $h$ where the force remains constant (a plateau in the force) is readily seen in Fig. \[force\]a. With growing strength of adsorption $\epsilon$ the length of this plateau as well as the magnitude of the plateau force increase. Note, that for $\epsilon = 0$ no plateau whatsoever is found. Upon further extension (by increasing $h$) of the chain, the plateau ends and the measured force starts to grow rapidly in magnitude - an effect, caused by a change of the chain conformation itself in the entirely desorbed state.
A closer inspection of Fig. \[force\]a reveals that the non-bonded contribution to $f$, which is generally much weaker than the bonded one, behaves differently, depending on whether the forces between non-nearest neighbors along the backbone of the chain are purely repulsive, or contain an attractive branch. While for strong adsorption, $\epsilon \ge 3.0$, a plateau is observed even for attractive non-bonded interactions, for weak adsorption, $\epsilon \le 2.0$, an increase of the non-bonded contribution at $h/l_0N
\approx 0.35$, (seen as a [*minimum*]{} in Fig. \[force\]a) is observed. This effect is entirely missing in the case of purely repulsive nonbonded
![\[force\] (a) Variation of the two components to the total force, exerted by the chain on the end-monomer which is fixed at (dimensionless) height $h/l_0N$ for different adsorption potentials $2.0 \le \epsilon/k_BT \le 5.0$: bonding interactions (full symbols) and non-bonding Morse interactions (empty symbols). In the inset the same is shown for a neutral plane $\epsilon = 0.0$ for purely repulsive monomers (triangles) and for such with a weak Morse attraction (circles). (b) Variation of the total force (plateau hight) exerted by the AFM tip on the chain-end for chain length $N=128$ with the relative strength of adsorption $(\varepsilon - \varepsilon_c)/k_BT$.](f_h_plateau.eps "fig:") ![\[force\] (a) Variation of the two components to the total force, exerted by the chain on the end-monomer which is fixed at (dimensionless) height $h/l_0N$ for different adsorption potentials $2.0 \le \epsilon/k_BT \le 5.0$: bonding interactions (full symbols) and non-bonding Morse interactions (empty symbols). In the inset the same is shown for a neutral plane $\epsilon = 0.0$ for purely repulsive monomers (triangles) and for such with a weak Morse attraction (circles). (b) Variation of the total force (plateau hight) exerted by the AFM tip on the chain-end for chain length $N=128$ with the relative strength of adsorption $(\varepsilon - \varepsilon_c)/k_BT$.](forceb_b.eps "fig:")
interactions - see the inset in Fig. \[force\]a where the contributions from bonded and non-bonded interactions are shown for a neutral surface $\epsilon = 0$. If one plots the magnitude of the measured force at the plateau against the corresponding value of the the adsorption potential, $\epsilon$, one may check the theoretical result, Eq. (\[Local\_Equilibrium\]) - Fig. \[force\]b. Evidently, the theoretical predictions about $f_{\rm p}$ agree well with the observed variation of the detachment force in the MC simulation, both within the BS- or FJBV models, as long as only the excluded volume interactions in the MC data are taken into account. If the total contribution to $f_{\rm p}$, including also attractive non-bonded interactions in the chain, is depicted - black triangles in Fig. \[force\]b - then the agreement with the theoretical curves deteriorates since the latter do not take into account the possible presence of attractive non-bonded interactions.
![\[f-h\] (a) Variation of the total applied force $f$ with growing height of the end monomer in terms of Pincus reduced variables, $f l_0
N^{\nu}/k_B T$ versus $h/l_0N^\nu$, for a polymer with purely repulsive nonbonded forces for $N=64,\;128$. (b) The same as in (a) but in terms of reduced units $f l_0/k_B T$ versus $h/l_0N$ for purely repulsive (empty symbols) as well as for usual Morse potential (full symbols) of nonbonded interactions between monomers. The FJBV-model results, Eq. (\[L\]), is shown by a solid line. Arrows indicate the unperturbed gyration radius positions $R_g/N$ for $N=64,\;128$](h_f_Pincus.eps "fig:") ![\[f-h\] (a) Variation of the total applied force $f$ with growing height of the end monomer in terms of Pincus reduced variables, $f l_0
N^{\nu}/k_B T$ versus $h/l_0N^\nu$, for a polymer with purely repulsive nonbonded forces for $N=64,\;128$. (b) The same as in (a) but in terms of reduced units $f l_0/k_B T$ versus $h/l_0N$ for purely repulsive (empty symbols) as well as for usual Morse potential (full symbols) of nonbonded interactions between monomers. The FJBV-model results, Eq. (\[L\]), is shown by a solid line. Arrows indicate the unperturbed gyration radius positions $R_g/N$ for $N=64,\;128$](force_B.eps "fig:")
The $f - h$ relationship, which gives the equation of state of the stretched polymer, may be derived within one of the different theoretical models, e.g., that of BS-, Eq. (\[Deformation\]), or FJBV-model, Eq. (\[L\]), as mentioned in Section \[Theory\]. Which of these theoretical descriptions is the more adequate can be decided by comparison with experiment. In Fig. \[f-h\]a,b, we present such comparison by plotting our simulation data using different normalization for the height $h$. From Fig. \[f-h\]a it becomes evident that the data from our computer experiment for $N=64$ and $N=128$ collapse on a single curve, albeit this collapse only holds as long as $h/l_0N^\nu \le 3.0$ for the BS-model while it fails for stronger stretching. In contrast, this collapse works well for all values of $h$, provided the height is scaled with the contour length of the chain $N$, rather than with $N^\nu$, as in the FJBV model - Fig. \[f-h\]b - regardless of whether a purely repulsive, or the full Morse potential (which includes also an attractive part) of interactions is involved. The analytical expression, Eq. (\[L\]), is found to provide perfect agreement with the simulation data for strong stretching, $h/l_0N \ge 0.4$. From the simulation data on Fig. \[f-h\] one may even verify that the force $f$ goes through zero at some height $h > 0$ and then turns negative, provided one keeps the chain end very close to the grafting surface (cf. eq.(\[Deformation\])).
Summary
=======
In the present work we have treated the force-induced desorption of a self-avoiding polymer chain from a flat structureless substrate both theoretically and by means of Monte Carlo simulation within the constant-height ensemble. The motivation for this investigation has been the necessity to distinguish between results obtained in this ensemble and results, derived in the constant-force ensemble, considered recently[@Bhattacharya], as far as both ensembles could in principle be used by experimentalists. We demonstrate that the observed behavior of the main quantity of interest, namely, the fraction of adsorbed beads $n$ (i.e., the order parameter of the phase transition) with changing height $h$ differs qualitatively from the variation of the order parameter when the pulling force is varied. In the constant-height ensemble one observes a steady variation of $n$ with changing $h$ whereas in the constant-force ensemble one sees an abrupt jump of $n$ at a particular value of $f_D$, termed a detachment force. However, this should not cast doubts on the genuine first-order nature of the phase transition which can be recovered within the constant-height ensemble too, provided one expresses the control parameter $h$ in terms of the average force $f$. This equivalence has been studied extensively for Gaussian chains by Skvortsov et al. [@Skvortsov] who noted that ensemble equivalence does not apply to fluctuations of the pertinent quantities too.
Indeed, in our earlier study[@PRER] we found diverging variance of the PDF $P(n)$ at $f_D$ whereas in our present study the fluctuations of the order parameter are observed to stay finite at the transition height $h_D$. These findings confirm theoretical predictions based on analytic results which we derive within the GCE-approach. Within this approach we have explored two different theoretical models for the basic force - extension relationship, namely, the bead-spring (BS) model as well as that of a Freely-Jointed Bond-Vectors (FJBV) model. Our simulation results indicate a good agreement between theory and computer experiment.
Acknowledgments {#acknowledgments .unnumbered}
===============
We are indebted to A. Skvortsov for useful discussions during the preparation of this work. A. Milchev thanks the Max-Planck Institute for Polymer Research in Mainz, Germany, for hospitality during his visit at the institute. A. Milchev and V. Rostiashvili acknowledge support from the Deutsche Forschungsgemeinschaft (DFG), grant No. SFB 625/B4.
Freely jointed bond vectors model
=================================
The deformation law in the overstretched regime (when the chain deformation is close to its saturation) could be treated better within the FJBV model. Consider a tethered chain of length $N$ with one end anchored at the origin of the coordinates and an external force $f_N$ acting on the free end of the chain. The corresponding deformation energy reads $$\begin{aligned}
U_{\rm ext } = - f_N \: r_{N}^{\perp} = - f_N \: \sum_{i=1}^{N} \: b_i
\cos \theta_i\end{aligned}$$ where $r_{N}^{\perp}$ is the $z$-coordinate (directed perpendicular to the surf ace) of the chain end, $b_i$ and $\theta_i$ are the length and the polar angle of the $i$-th bond vector respectively. The corresponding partition function of the FJBV model is given by $$\begin{aligned}
Z_{N} (f_N) &=& \int \prod_{i=1}^{N} d \phi_i \sin \theta_i d \theta_i
\exp\left(\frac{f_N}{k_B T} \sum_{i=1}^{N} b_i \cos \theta_i\right)\nonumber\\
&=& (4\pi)^{N} \prod_{i=1}^{N} \left(\frac{k_B T}{b_i f_N} \right) \cosh
\left(\frac{b_i f_N}{k_B T} \right)
\label{Partition_Func}\end{aligned}$$ The average orientation of the $i$-th bond vector can be calculated as $$\begin{aligned}
<\cos \theta_i> = \left(\frac{k_B T}{f_N}\right) \: \frac{\partial}{\partial
b_i}
\ln Z_N (f_N) = \coth \left(\frac{b_i f_N}{k_B T}\right) - \left(\frac{k_B
T}{b_i f_N
}\right)
\label{Cos}\end{aligned}$$ From Eq.(\[Cos\]) the chain end mean distance from the surface, $h$, is given by $$\begin{aligned}
h = \sum_{i=1}^{N} b_i <\cos \theta_i> = b N {\cal L} \left(\frac{b f_N }{k_B T}
\right)
\label{H_vs_Force}\end{aligned}$$ where we have taken into account that the lengths of all bond vectors are of equal length, $b_i = b$, and ${\cal L} (x) = \coth (x) - 1/x$ is the Langevin function. This leads to the force - distance relationship $$\begin{aligned}
\frac{b f_N}{k_B T} = {\cal L}^{-1}\left(\frac{h}{b N}\right) = \begin{cases}
\frac{1}{1-h/b
N} & \text{, at} \:\: h/b N \leq 1\\
\frac{2 h}{b N} & \text{, at} \:\: h/b N \ll 1
\end{cases}
\label{Inversion}\end{aligned}$$ which we use in Sec.II. The notation ${\cal L}^{-1}(x)$ stands for the inverse Langevin function.
Calculation of PDF in the $h$-ensemble
======================================
Using Eq.(\[Z\_Star\]) for $\ln z^{*} (w)$ as well as Eqs. (\[Local\_Equilibrium\]) and (\[M\_vs\_h\]) (for the BS-model), one obtains an expression for the tail length $$M(h, \xi \: w) = \dfrac{h/l_0}{\left[a_1^\phi (w - w_c)
\: \xi \right]^{(1-\nu)/\phi} }
\label{Tail_Length}$$ The saddle point (SP) equation in this case reads (cf. Eq.(\[SP\])) $$\dfrac{K+1}{\xi_0} = - \left. \left[N - M(h, \xi_0 w) \right] \:
\dfrac{\partial \ln z^{*}}{\partial
\xi}\right|_{\xi = \xi_0} + \left. \dfrac{\partial M}{\partial \xi} \right|_{\xi
= \xi_0} \: \ln z^{*}(\xi_0 w)
\label{SP_General}$$ Taking into account Eqs. (\[Z\_Star\]) and (\[Tail\_Length\]), after introducing the notation $y = (w - w_c) \xi_0$, the SP-equation can be recast into $$\left( \dfrac{K}{N}\right) = y^{1/\phi} - B_1 \left( \dfrac{h}{l_0 N}\right)
y^{\nu/\phi} - B_2 \left( \dfrac{h}{l_0 N}\right) \dfrac{\ln
\mu_3}{y^{(1-\nu)/\phi}}\;,
\label{SP_Recast}$$ where $B_1$ and $B_2$ are constants of the order of unity. In the particular case $h = 0$ Eq.(\[SP\_Recast\]) goes back, as expected, to Eq.(\[SP\_Solution\]). Eq. (\[SP\_Recast\]) can be solved iteratively for $h
\ll l_0 N$ as $$y = \left[ \left( \dfrac{K}{N}\right) + B_1 \left( \dfrac{K}{N}\right)^\nu
\left( \dfrac{h}{l_0 N}\right) + B_2 \ln \mu_3 \left(
\dfrac{N}{K}\right)^{1-\nu} \left( \dfrac{h}{l_0 N}\right) \right]^{\phi}
\label{Iteration}$$ As before, the main contribution in the integral given by Eq. (\[Coefficient\]) reads $$\begin{aligned}
P(K) &\propto& \exp \left\lbrace- \left[N - M(h, \xi_0 w)\right] \ln
z^{*}(\xi_0 w) - (K+1) \ln \xi_0 \right\rbrace \nonumber\\
&=& \left\lbrace (w -
w_c) N^{\phi} \exp \left[1 - \dfrac{(h/l_0 N)}{(K/N)^{1-\nu} \left[\rho
(K/N, h/l_0 N)\right]^{1-\nu}}\right] \rho (K/N, h/l_0 N)\right\rbrace^K \:
\left[K \rho (K/N, h/l_0 N) \right]^{-\phi K}
\end{aligned}$$ where one introduces the notation $$\begin{aligned}
\rho \left(K/N, h/l_0 N \right) \equiv 1 + B_1 \dfrac{h/l_0
N}{(K/N)^{1-\nu}} + B_2 \ln \mu_3 \: \dfrac{(h/l_0 N)}{(K/N)^{2-\nu}}
\label{Notation}\end{aligned}$$ After normalization, the final expression for the PDF reads $$\begin{aligned}
P(K) = \dfrac{1}{{\cal W}(\eta, h/l_0 N)} \: \left\lbrace \eta \: \exp
\left[1 - \dfrac{(h/l_0 N)}{(K/N)^{1-\nu} \left[\rho
(K/N, h/l_0 N)\right]^{1-\nu}}\right] \rho (K/N, h/l_0 N)\right\rbrace^K \:
\left[K \rho (K/N, h/l_0 N) \right]^{-\phi K}
\label{PDF_Pull}\end{aligned}$$ where $\eta = (w - w_c) N^\phi$ as before and the normalization constant reads $$\begin{aligned}
{\cal W}(\eta, h/l_0 N) = \sum_{K=1}^{N} \: \left\lbrace \eta \: \exp
\left[1 - \dfrac{(h/l_0 N)}{(K/N)^{1-\nu} \left[\rho
(K/N, h/l_0 N)\right]^{1-\nu}}\right] \rho (K/N, h/l_0 N)\right\rbrace^K \:
\left[K \rho (K/N, h/l_0 N) \right]^{-\phi K}
\label{Normalization_Pull}\end{aligned}$$ Again one can readily see that for $h = 0$ Eq.(\[PDF\_Pull\]) reduces to Eq.(\[Distr3\]). The PDF, following from Eq. (\[PDF\_Pull\]) is shown in Fig. \[pdf\_theory\]b for several values of the height $h$. Evidently, both the mean value $\overline{K}$ and the variance $\overline{K^2}-\overline{K}^2$ decline with growing $h$.
[99]{} S. B. Smith, Y. Cui, and C. Bustamante, Science [**271**]{}, 795(1996). T. J. Senden, J.-M. di Meglio, and P. Auroy, Europ. Phys. J. B[**3**]{}, 211(1998). T. Hugel, M. Grosholz, H. Clausen-Schaumann, A. Pfau, H. E. Gaub, and M. Seitz, Macromolecules, [**34**]{}, 1039(2001) H. Hugel, M. Seitz, Macromol. Rapid. Commun. [**22**]{}, 989 (2001). A. Rohrbach, E.H.K. Stelzer, J. Appl. Phys. [**91**]{}, 5474 (2002). L. Sonnenberg, Y. Luo, H. Schaad, M. Seitz, H. Gölfen, H.E. Gaub, J. Am. Chem. Soc. [**129**]{}, 15364 (2007). C. Friedsam, A. Del Campo Becares, M. Seitz, and H. E. Gaub, New J. Phys. [**6**]{}, 9(2004). S.K. Kufer, E.M. Puchner, H. Gumpp, T. Liedl, H.E. Gaub, Science, [**319**]{}, 594 (2008). M. Rief, F. Oesterhelt, B. Heymann, and H. E. Gaub, Science [**275**]{}, 1295(1997). C. Ortiz and G. Hadziioannou, Macromolecules [**32**]{}, 780(1990). M. Grandbois, M. Beyer, M. Rief, H. Clausen-Schaumann, and H. E. Gaub, Science [**283**]{}, 1727(1999). A.M. Skvortsov, L.I. Klushin, T.M. Birshtein, Polymer Sci. (Moscow) Ser. A, [**51**]{}, 1 (2009). S. Bhattacharya, V.G. Rostiashvili, A. Milchev, T.A. Vilgis, Phys. Rev. E [**79**]{}, 030802(R) (2009). S. Bhattacharya, V.G. Rostiashvili, A. Milchev, T.A. Vilgis, Macromolecules, [**42**]{}, 2236 (2009). C. Vanderzande, [*Lattice Model of Polymers*]{}, Cambridge University Press, Cambridge, 1998. E. Eisenriegler, K. Kremer, K. Binder, J. Chem. Phys. [**77**]{}, 6296 (1982). E. Eisenriegler, [*Polymers Near Surfaces*]{}, World Scientific, 1993. T. Kreer, S. Metzger, M. Müller, K. Binder, J. Chem. Phys. [**120**]{}, 4012 (2004). A.Yu. Grosberg, A.R. Khokhlov,[*Statistical Physics of Macromolecules*]{} (AIP, New York, 1994). J. des Cloiseaus, G. Jannink, [Polymers in Solution]{}, Clarendon Press, Oxford, 1990. P.-G. de Gennes, [*Scaling Consept in Polymer Physics*]{}, Cornell Univercity Press, 1979. Y.-J. Sheng, P.-Y. Lai, Phys. Rev. E [**56**]{}, 1900 (1997). M. Wittkop, J.-U. Sommer, S. Kreitmeier, D. Göritz, Phys. Rev. E [**49**]{}, 5472 (1994). J.U. Schurr, S.B. Smith, Biopolymers, [**29**]{}, 1161 (1990). J. Rudnick, G. Gaspari, [*Elements of Random Walk*]{}, Cambridge Univercity Press, Cambridge, 2004. S. Bhattacharya, H.-P. Hsu, V.G. Rostiashvili, A. Milchev, T.A. Vilgis, Macromolecules, [**41**]{}, 2920 (2008). K. Binder and A. Milchev, J. Computer-Aided Material Design, [**9**]{}, 33(2002)
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
The inferential model (IM) framework provides valid prior-free probabilistic inference by focusing on predicting unobserved auxiliary variables. But, efficient IM-based inference can be challenging when the auxiliary variable is of higher dimension than the parameter. Here we show that features of the auxiliary variable are often fully observed and, in such cases, a simultaneous dimension reduction and information aggregation can be achieved by conditioning. This proposed conditioning strategy leads to efficient IM inference, and casts new light on Fisher’s notions of sufficiency, conditioning, and also Bayesian inference. A differential equation-driven selection of a conditional association is developed, and validity of the conditional IM is proved under some conditions. For problems that do not admit a valid conditional IM of the standard form, we propose a more flexible class of conditional IMs based on localization. Examples of local conditional IMs in a bivariate normal model and a normal variance components model are also given.
*Keywords and phrases:* Ancillary; auxiliary variable; Bayes; belief function; differential equation; sufficiency; predictive random set; validity.
author:
- |
Ryan Martin\
Department of Mathematics, Statistics, and Computer Science\
University of Illinois at Chicago\
[[email protected]]([email protected])\
\
Chuanhai Liu\
Department of Statistics\
Purdue University\
[[email protected]]([email protected])
bibliography:
- '/Users/rgmartin/Dropbox/Research/mybib.bib'
title: 'Conditional inferential models: combining information for prior-free probabilistic inference'
---
Introduction {#S:intro}
============
Fisher’s brand of statistical inference [@fisher1973] is often viewed as a middle-ground between the Bayesian and frequentist approaches. Two important examples are his fiducial argument and his ideas on conditional inference. Perhaps influenced by Fisher’s ideas, a current focus in foundational research is on achieving some kind of compromise between the Bayesian and frequentist ideals. See, for example, recent work on fiducial inference [@hannig2009; @hannig2012; @hannig.lee.2009], confidence distributions [@xie.singh.strawderman.2011; @xie.singh.2012], Dempster–Shafer theory [@dempster2008; @shafer2011], and objective Bayes with default, reference, and/or data-dependent priors [@berger2006; @bergerbernardosun2009; @fraser2011; @fraser.reid.marras.yi.2010]. Recently @imbasics have laid out the details of a promising new *inferential model* (IM) approach. IMs take the usual input—sampling model and observed data—and produce prior-free, probabilistic measures of certainty about any assertion/hypothesis of interest, with an almost automatic calibration property. The fundamental idea is that uncertainty about the parameter of interest $\theta$, given observed data $X=x$, is fully characterized by the of an unobservable auxiliary variable $U$. So, the problem of inference about $\theta$ can be translated into one of predicting this unobserved $U$ with a random set. In Section \[S:ims\] we briefly review the construction and basic properties of IMs.
The discussion in @imbasics focuses on the case where $\theta$ and $U$ are of the same dimension. But there are many problems, e.g., iid data from scalar parameter models, where the dimension of the auxiliary variable is greater than that of the parameter. In such cases, efficiency can be gained by first reducing the dimension of the auxiliary variable to be predicted, though it is not obvious how this should be done in general. Here we focus our attention on an auxiliary variable dimension reduction step based on conditioning. The key observation is that, typically, certain functions of the auxiliary variables are fully observed. By conditioning on those observed characteristics of the auxiliary variable, we can effectively reduce the dimension of the unobserved characteristics to be predicted. A motivating example, demonstrating the efficiency gain from dimension reduction, along with the detailed developments are presented in Section \[S:conditional\]. The proposed dimension-reduction approach, based on conditioning, can be viewed as a general tool for combining information about $\theta$ across samples—a counterpart to Bayes’ theorem and sufficiency. With the lower-dimensional auxiliary variable, we proceed to construct what is called a *conditional IM*. We prove a validity theorem that establishes a desirable calibration property of the conditional IM, and facilitates a common interpretation across users and experiments.
Finding the dimension-reduced representation is sometimes a familiar task. For example, when the minimal sufficient statistic has dimension matching that of the parameter, the conditional IM is exactly that obtained by working directly with said statistic. In other cases, finding the lower-dimensional representation is not so simple. For this, in Section \[S:finding\], we propose a new differential equation-driven technique for identifying observed characteristics of the auxiliary variable. Two classical conditional inference problems are worked out in Section \[S:more-examples\], one showing how the proposed differential equation technique leads to an additional dimension reduction beyond what ordinary sufficiency provides. So, besides the development of conditional IMs, the proposed framework also casts new light on the familiar notion of sufficiency, as well as Fisher’s attractive but elusive ideas on ancillary statistics and conditional inference.
In some cases, however, it may not be possible to produce a valid conditional IM with these somewhat standard techniques. For this, we propose an extension of the conditional IM framework, in Section \[S:gcim\], which allows the lower-dimensional auxiliary variable representation to depend on $\theta$ in a certain sense. We refer to these as *local conditional IMs*, and we describe their construction and prove a validity theorem. An important example of such a problem is the bivariate normal model with known means and variances but unknown correlation. For this example, we construct a local conditional IM based on a modification of our differential equations technique, and provide the results of a simulation study that shows that our conditional plausibility intervals outperform the classical $r^\star$-driven asymptotically approximate confidence intervals [@bn1986; @fraser1990] in both small and large samples. A local conditional IM analysis of the variance-components model, another benchmark problem, is also given.
Review of IMs {#S:ims}
=============
Notation and construction {#SS:notation}
-------------------------
To fix notation, let $X$ be the observable data, taking values in a space ${\mathbb{X}}$, and let $\theta$ be the parameter of interest, taking values in the parameter space $\Theta$. The starting point of the IM framework is similar to that of fiducial, in the sense that an auxiliary variable, denoted by $U$ and taking values in a space ${\mathbb{U}}$ with probability measure ${\mathsf{P}}_U$, is associated with $X$ and $\theta$. It is this association, together with the distribution $U \sim {\mathsf{P}}_U$, that characterizes the sampling distribution $X \sim {\mathsf{P}}_{X|\theta}$. In particular, if we write this association as $$\label{eq:assoc}
X = a(\theta,U), \quad U \sim {\mathsf{P}}_U.$$ Throughout, subscripts on ${\mathsf{P}}$ indicate which quantity is random.
Fiducial inference employs the sampling distribution ${\mathsf{P}}_U$ *after* $X=x$ is observed. The IM approach is different in that it treats the unobserved value of $U$, which is tied to the observed data $X=x$ and the *true value* of $\theta$, as the fundamental quantity. Then the goal is to predict this unobserved value with a random set before conditioning on $X=x$ and inverting . Let $({\mathbb{U}},\mathscr{U},{\mathsf{P}}_U)$ be a probability space, where $\mathscr{U}$ is rich enough to contain all closed subsets of ${\mathbb{U}}$. Take a collection ${\mathbb{S}}$ of closed (hence ${\mathsf{P}}_U$-measurable) subsets of ${\mathbb{U}}$, assumed to contain $\varnothing$ and ${\mathbb{U}}$. This collection will serve as the support of the predictive random set. We shall also assume that the collection ${\mathbb{S}}$ is nested, i.e., either $S \subseteq S'$ or $S' \subseteq S$ for all $S,S' \in {\mathbb{S}}$. Now define the predictive random set ${\mathcal{S}}\sim {\mathsf{P}}_{\mathcal{S}}$, supported on ${\mathbb{S}}$, with distribution ${\mathsf{P}}_{\mathcal{S}}$ satisfying $${\mathsf{P}}_{\mathcal{S}}\{{\mathcal{S}}\subseteq K\} = \sup_{S \in {\mathbb{S}}: S \subseteq K} {\mathsf{P}}_U(S), \quad K \subseteq {\mathbb{U}}.$$ Predictive random sets with these properties are called *admissible*. @imbasics give a sort of complete-class theorem for admissible predictive random sets. In scalar $\theta$ problems, ${\mathsf{P}}_U $ is often ${{\sf Unif}}(0,1)$, so an important example of an admissible predictive random set is $$\label{eq:default.prs}
{\mathcal{S}}= \{u: |u-\tfrac12| \leq |U-\tfrac12|\}, \quad U \sim {{\sf Unif}}(0,1).$$ @imbasics [Corollary 1] show that this ${\mathcal{S}}$—called the “default” predictive random set—has a variety of good properties, and these good properties often carry over to the corresponding IM. It should be mentioned that admissible predictive random sets are not unique. In this paper we focus primarily on simplicity, but the ideas on optimal predictive random sets in @imbasics can also be applied here.
The following three steps—association, predict, and combine—define an IM.
Associate $X$, $\theta$, and $U \sim {\mathsf{P}}_U$, consistent with the sampling distribution $X \sim {\mathsf{P}}_{X|\theta}$, such that, for all $(x,u)$, there is a unique subset $\Theta_x(u) = \{\theta: x=a(\theta,u)\} \subseteq \Theta$, possibly empty, containing all possible candidate values of $\theta$ given $(x,u)$.
Predict the unobserved value $u^\star$ of $U$ associated with the observed data by an admissible predictive random set ${\mathcal{S}}$.
Combine ${\mathcal{S}}$ and the association $\Theta_x(u)$ specified in the A-step to obtain $$\label{eq:new.focal}
\Theta_x({\mathcal{S}}) = \bigcup_{u \in {\mathcal{S}}} \Theta_x(u).$$ Then compute the *belief function* $$\label{eq:belief}
{\mathsf{bel}}_x(A; {\mathcal{S}}) = {\mathsf{P}}_{\mathcal{S}}\{\Theta_x({\mathcal{S}}) \subseteq A \mid \Theta_x({\mathcal{S}}) \neq \varnothing\},$$ where $A \subseteq \Theta$ is the assertion/hypothesis about $\theta$ of interest. @leafliu2012 give an alternative to conditioning on the event “$\Theta_x({\mathcal{S}}) \neq \varnothing$,” based on stretching, that tends to be more efficient.
The belief function is just one part of the inferential output. Since the belief function is sub-additive, i.e., ${\mathsf{bel}}_x(A;{\mathcal{S}}) + {\mathsf{bel}}_x(A^c;{\mathcal{S}}) \leq 1$, one actually needs both ${\mathsf{bel}}_x(A;{\mathcal{S}})$ and ${\mathsf{bel}}_x(A^c;{\mathcal{S}})$ to summarize the information in $x$ concerning the truthfulness of assertion $A$. In some cases, it is more convenient to report the *plausibility function* $$\label{eq:plausibility}
{\mathsf{pl}}_x(A;{\mathcal{S}}) = 1-{\mathsf{bel}}_x(A^c; {\mathcal{S}}).$$ Then the pair $({\mathsf{bel}}_x,{\mathsf{pl}}_x)(A;{\mathcal{S}})$ characterize the IM output. The IM and fiducial approaches both start with a representation of the sampling model using auxiliary variables, but, beyond that, there are some important differences. First, by taking the predictive random set ${\mathcal{S}}= \{U\}$, with $U \sim {\mathsf{P}}_U$, a random singleton, the corresponding IM is exactly the fiducial distribution for $\theta$. However, this singleton random set is not nested, hence, not admissible, so the desirable validity properties in Section \[SS:validity\] are not guaranteed. Second, a subtle point, the interpretation of probability changes as one proceeds along the fiducial argument. One starts with a non-subjective probability ${\mathsf{P}}_U$ for $U$ before $X=x$ is observed. Then, after $X=x$ is observed, the conditional distribution of $U$, which is concentrated on the set $\{u: x=a(\theta,u)\}$, where $\theta$ is the fixed true parameter value, is replaced by the original ${\mathsf{P}}_U$, i.e., one “continues to regard” $U$ as a sample from ${\mathsf{P}}_U$ after $X=x$ is observed [@dempster1963]. Therefore, despite starting with a non-subjective probability ${\mathsf{P}}_U$, the choice to replace the conditional distribution of $U$, given $X=x$, with ${\mathsf{P}}_U$ ultimately makes the fiducial probabilities subjective. This explains why fiducial inference is not valid for some assertions (and finite $n$). IM probabilities, on the other hand, are based on ${\mathsf{P}}_{\mathcal{S}}$, and the fact that there are no direct links between data and ${\mathcal{S}}$ means that the ${\mathsf{P}}_{\mathcal{S}}$-probabilities are the same—in terms of both computation and interpretation—before and after $X=x$ is observed. This explains why IM-based inference is valid whenever the predictive random set is suitably calibrated to ${\mathsf{P}}_U$; see Section \[SS:validity\] below.
Finally, without practical loss of generality, assume that $\{{\mathsf{P}}_{X|\theta}: \theta \in \Theta\}$ has a common dominating measure, say $\mu$. Then we require that ${\mathsf{bel}}_x(A;{\mathcal{S}})$ be a $\mu$-measurable function in $x$. This is easy to check in examples, but general sufficient conditions are more elusive. To keep the presentation simple, we shall mostly ignore these technicalities.
Validity of IMs {#SS:validity}
---------------
Given ${\mathcal{S}}$, the corresponding IM is *valid for* $A$ if the belief function satisfies $$\label{eq:bel.valid}
\sup_{\theta \not\in A} {\mathsf{P}}_\theta\bigl\{ {\mathsf{bel}}_X(A;{\mathcal{S}}) \geq 1-\alpha \bigr\} \leq \alpha, \quad \forall\;\alpha \in (0,1).$$ The IM is simply called *valid* if it is valid for all $A$. In other words, the IM is valid for $A$ if ${\mathsf{bel}}_X(A;{\mathcal{S}})$ is stochastically no larger than ${{\sf Unif}}(0,1)$ when $X \sim {\mathsf{P}}_{X|\theta}$ with $\theta \not\in A$. That is, if $A$ is false, then the amount of support in data $X$ for $A$ will be large only for a relatively small proportion of $X$ values. @imbasics [@imbasics.c] show that this validity property holds whenever the predictive random set ${\mathcal{S}}$ is admissible. If the IM is valid for all $A$, then can be equivalently stated in terms of the plausibility function: $$\label{eq:pl.valid}
\sup_{\theta \in A} {\mathsf{P}}_{X|\theta}\{{\mathsf{pl}}_X(A; {\mathcal{S}}) \leq \alpha\} \leq \alpha, \quad \forall \; \alpha \in (0,1).$$ This formulation is occasionally more convenient than .
There are two important consequences of the validity theorem. First, it helps determine an objective scale on which the belief probabilities can be interpreted. Unlike valid IMs, the output from default-prior Bayesian, fiducial, and Dempster–Shafer inference does not have a specified scale for interpretation. For example, is a Bayesian or fiducial posterior probability of 0.9 a large value? It is common to think on the usual frequency scale, i.e., betting on an event with 0.9 probability wins 90% of the time, but there is no justification for this without some notion of validity as in or . Second, validity justifies the use of IM output to construct frequentist decision procedures with control on error rates. For example, a $100(1-\alpha)$% plausibility region for $\theta$ is given by $$\label{eq:plaus.region}
\{\theta: {\mathsf{pl}}_x(\theta; {\mathcal{S}}) > \alpha\}.$$ It follows easily from that this plausibility region has nominal $1-\alpha$ coverage probability. But we should emphasize here that, although plausibility functions can be used to construct frequentist procedures, the interpretation is quite different. For example, the plausibility region is understood as the collection of $\theta$’s such that each is *individually* sufficiently plausible, given $X=x$. Confidence/credible regions, on the other hand, say nothing about the plausibility of any particular $\theta$ they contain.
Conditional IMs {#S:conditional}
===============
Motivation {#SS:normal.example}
----------
In the case of a scalar auxiliary variable, construction of efficient predictive random sets is relatively easy. However, rarely does the model directly admit a scalar auxiliary variable representation. To see this, suppose $X_1,\ldots,X_n$ are independent ${{\sf N}}(\theta,1)$ with unknown mean $\theta$. In vector notation, an association is $X = \theta 1_n + U$, where $1_n$ is an $n$-vector of unity, and $U \sim {\sf N}_n(0,I)$. At first look, it seems that one must predict an $n$-dimensional auxiliary variable $U$. But efficient prediction of $U$ would be challenging, even for moderate $n$, so reducing its dimension—ideally to one—would be a desirable first step. After reducing the dimension to one, choosing efficient predictive random sets is as easy as in the scalar auxiliary variable case considered in @imbasics.
The basic point is that one pays a price, in terms of efficiency, for predicting higher-dimensional auxiliary variables. To see this better, we shall take a closer look at the normal mean problem above, with $n=2$. That is, we have a baseline association $$X_1 = \theta + U_1 \quad \text{and} \quad X_2 = \theta + U_2,$$ where $U_1,U_2$ are independent ${{\sf N}}(0,1)$. To make things simple, consider the following change of variables: $Y_1 = X_1+X_2$ and $Y_2 = X_1-X_2$. In the new variables, we have $$Y_1 = 2\theta + V_1 \quad \text{and} \quad Y_2 = V_2,$$ where $V_1,V_2$ are independent ${{\sf N}}(0,2)$. This completes the A-step. Following the basic procedure described in Section \[S:ims\], for the P-step, we should predict the pair $(V_1,V_2)$ with a predictive random set ${\mathcal{S}}$. A simple $L_\infty$ generalization of the default predictive random set to the case of a two-dimensional auxiliary variable is a random square: $${\mathcal{S}}= \{(v_1,v_2): \max(|v_1|,|v_2|) \leq \max(|V_1|,|V_2|)\}, \quad V_1,V_2 {\overset{\text{\tiny iid}}{\,\sim\,}}{{\sf N}}(0,2).$$ For a singleton assertion $\{\theta\}$, the C-step gives plausibility function $${\mathsf{pl}}_y(\theta) = \frac{1 - G( 2^{-1/2}\max\{|y_1-2\theta|,|y_2|\} )^2}{1 - G( 2^{-1/2}|y_2| )^2},$$ where $G(z)=1-2(1-\Phi(z))$ is the $|{{\sf N}}(0,1)|$ distribution function. The unusual form here is due to the conditioning in to remove conflict cases where $\Theta_y({\mathcal{S}}) = \varnothing$.
As an alternative approach, note that the value of $V_2$ is known once $Y_2$ is observed. So rather than trying to predict this component, as in the approach just described, we might condition on this observed value, to sharpen our uncertainty for predicting $V_1$. Since $V_1$ and $V_2$ are actually independent, it suffices to work with the marginal distribution, $V_1 \sim {{\sf N}}(0,2)$. For the A-step, we get $Y_1 = 2\theta + V_1$ and, for the P-step, we use a default predictive random set ${\mathcal{S}}=\{v_1: |v_1| \leq V_1\}$, where $V_1 \sim {{\sf N}}(0,2)$. For the same singleton assertion, the C-step this time gives plausibility function $${\mathsf{pl}}_y(\theta) = 1 - |2\Phi\bigl(2^{-1/2}(y_1-2\theta)\bigr) - 1|.$$
The claim is that inference based on the latter IM formulation is more efficient than that based on the former. To check this, we consider the sampling distribution of ${\mathsf{pl}}_Y(0)$ in the case where $Y=(Y_1,Y_2)$ is an independent ${{\sf N}}(0,2)$ random vector. Figure \[fig:plaus.qq\] shows a quantile plot of the two simulated samples. By the validity theorem, the plausibilities are both stochastically no smaller than ${{\sf Unif}}(0,1)$. However, we see that plausibilities for the reduced, one-dimensional predictive random set are exactly ${{\sf Unif}}(0,1)$ distributed, while those based on the two-dimensional predictive random set tend to be considerably larger. The larger plausibility means less efficiency, e.g., wider plausibility intervals, so the IM based on the reduced one-dimensional predictive random set is preferred. This difference in efficiency is explained by the fact that the two-dimensional predictive random set for $(V_1,V_2)$ corresponds to a larger-than-necessary predictive random set for $V_1$; the conflict cases in the two-dimensional case have little to no effect on efficiency.
In the remainder of this section, we will give a general prescription for increasing efficiency by reducing the dimension. The key is that, in general, some functions of the original auxiliary variable are fully observed, like $V_2=U_1-U_2$ in this simple example. Then the strategy is to condition on what is fully observed to sharpen prediction of what is not observed. Since this “conditioning to sharpen inference” strategy is commonly used in statistics, similar considerations are natural in the IM framework.
Dimension reduction via conditioning {#SS:cim}
------------------------------------
Here we propose a conditioning strategy, whereby a simultaneous information aggregation and dimension reduction is achieved, that results in an overall gain in efficiency. The intuition is that some functions of the unobserved ${u^\star}$ are actually observed, so these characteristics do not need to be predicted. Focusing only on the unobserved characteristics of ${u^\star}$ leads directly to a dimension reduction. However, knowledge about the observed characteristics helps to better predict those unobserved characteristics, so information is accumulated and prediction is sharpened. The general strategy is as follows:
- Identify an observed characteristic, $\eta(U)$, of the auxiliary variable $U$ whose distribution is free (or at least mostly free) of $\theta$, and
- define a conditional association that relates an unobserved characteristic, $\tau(U)$, of the auxiliary variable $U$ to $\theta$ and some function $T(X)$ of the data $X$.
The second step is familiar, as it relates to working with, say, a minimal sufficient statistic. The first step, however, is less familiar and generally more difficult; see Section \[S:finding\].
To make this formal, suppose that $x \mapsto (T(x),H(x))$ and $u \mapsto (\tau(u),\eta(u))$ are one-to-one functions. Suppose that the relationship $x=a(u,\theta)$ in the baseline association can be decomposed as
\[eq:decomposition\] $$\begin{aligned}
H(x) & = \eta(u), \label{eq:condition} \\
T(x) & = b(\tau(u), \theta). \label{eq:condition0} \end{aligned}$$
This decomposition immediately suggests an alternative association. Let $(V_T,V_H) \in {\mathbb{V}}_T \times {\mathbb{V}}_H$ be the image of $U$ under $(\tau,\eta)$, and let ${\mathsf{P}}_{V_T|h}$ be the conditional distribution of $V_T$, given $V_H=h$, where $h \in H({\mathbb{X}})$. Since $H(x)$ provides no information about $\theta$, we can take a new association $$\label{eq:cond-aeqn}
T(X) = b(V_T,\theta), \quad V_T \sim {\mathsf{P}}_{V_T | H(x)}.$$ We shall refer to this as a *conditional association*. This alternative association can be understood via a certain hierarchical representation of the sampling model; see Remark \[re:hierarchy\]. The important point is that $\tau$ can often be chosen so that $V_T$ is of lower dimension than $U$. In fact, $V_T$ will often have dimension the same as that of $\theta$. In addition to providing a sort of summary of the data, like in the classical context, this auxiliary variable dimension reduction has a unique advantage in the IM context: efficient predictive random sets for the lower-dimensional $V_T$ are easier to construct. Furthermore, the conditioning aspect sharpens our predictive ability, improving efficiency even more. We witnessed, empirically, these gains in efficiency in the simple example in Section \[SS:normal.example\]. Some further remarks on this conditional association, and its connections to Fisher’s sufficiency and Bayes theorem, are collected in Section \[SS:remarks\].
Once a decomposition is available, construction of the corresponding IM follows exactly as in Section \[S:intro\]. To simplify the presentation later on, here we restate the three-step construction of a *conditional IM*.
Associate $T(x)$ and $\theta$ with the new auxiliary variable $v_T = \tau(u)$ to get the collection of sets $\Theta_{T(x)}(v_T) = \{\theta: T(x) = b(v_T,\theta)\}$, $v_T \in {\mathbb{V}}_T$, based on .
Fix $h=H(x)$. Predict the unobserved value ${v^\star}_T$ of $V_T$ with a *conditionally admissible* predictive random set ${\mathcal{S}}\sim {\mathsf{P}}_{{\mathcal{S}}|h}$ (see Section \[SS:condcred\]).
Combine results of the A- and P-steps to get $$\label{eq:condfocal}
\Theta_{T(x)}({\mathcal{S}}) = \bigcup_{v_T \in {\mathcal{S}}} \Theta_{T(x)}(v_T) \subseteq \Theta.$$ Then the corresponding conditional belief and plausibility functions are given by $$\label{eq:condbelief}
\begin{split}
{\mathsf{cbel}}_{T(x)|h}(A;{\mathcal{S}}) & = {\mathsf{P}}_{{\mathcal{S}}|h}\{\Theta_{T(x)}({\mathcal{S}}) \subseteq A \mid \Theta_{T(x)}({\mathcal{S}}) \neq \varnothing\} \\
{\mathsf{cpl}}_{T(x)|h}(A;{\mathcal{S}}) & = 1-{\mathsf{cbel}}_{T(x)|h}(A^c; {\mathcal{S}}).
\end{split}$$ These functions can be used for inference on $\theta$ just like those in Section \[S:ims\]. The notation ${\mathsf{cbel}}$ and ${\mathsf{cpl}}$ is meant to indicate that these are belief and plausibility functions, depending on $(T,H)(x)$, based on predicting the lower-dimensional auxiliary variable $V_T = \tau(U)$ in the conditional association .
When a decomposition is available, the conditional association and the corresponding conditional IM analysis is intuitively reasonable. One could ask, however, if there is any loss in using IMs built from instead of . The following proposition establishes that there is no loss.
\[prop:conditioning\] Suppose the baseline association admits a decomposition of the form . Let ${\mathcal{S}}$ be a predictive random set for $U$ in the baseline association with the property that ${\mathsf{P}}_{\mathcal{S}}\{\Theta_x({\mathcal{S}}) \neq \varnothing\} > 0$ for all $x$. Then there exists a predictive random set ${\mathcal{S}}_{H(x)}$ for $\tau(U)$, depending on ${\mathcal{S}}$ and $H(x)$, in the conditional association such that ${\mathsf{bel}}_x(A;{\mathcal{S}}) = {\mathsf{cbel}}_{T(x)}(A; {\mathcal{S}}_{H(x)})$ for all $x$ and all assertions $A \subseteq \Theta$.
See Appendix \[S:proofs\] for the proof. This says that the conditional association is as good a starting point as the baseline association in the sense that any belief function that obtains from the latter can be matched by one that obtains from the former. Therefore, the best conditional IM can be no worse than the best baseline IM. However, as in Section \[SS:normal.example\], by working with predictive random sets for the lower-dimensional auxiliary variable in the conditional association, efficiency can be improved. The point of this paper, in fact, is that the best conditional IM is more efficient than the best baseline IM.
A shortcoming of Proposition \[prop:conditioning\] is that the predictive random set ${\mathcal{S}}_{H(x)}$ constructed in the proof may not be valid for prediction of $\tau(U)$, which prevents us from making a direct efficiency comparison of conditional versus baseline IMs. However, in the special case where $\tau(U)$ and $\eta(U)$ are independent, validity of that predictive random set obtains.
\[co:indep.cond\] In the setup of Proposition \[prop:conditioning\], suppose that $\tau(U)$ and $\eta(U)$ are independent. Then the predictive random set ${\mathcal{S}}'=\tau({\mathcal{S}})$ for $\tau(U)$ is valid and ${\mathsf{bel}}_x(A;{\mathcal{S}}) = {\mathsf{cbel}}_{T(x)}(A; {\mathcal{S}}')$ for all $x$ and all assertions $A \subseteq \Theta$.
Since the predictive random set for $\tau(U)$ is valid, a requirement for IMs, in this case we can conclude that the conditional IM is at least as efficient as the baseline IM. The independence condition holds in many examples; see Section \[S:more-examples\].
Remarks {#SS:remarks}
-------
\[re:separable\] More general decompositions in are possible. That is, one may replace “$H(x) = \eta(u)$” in with “$c(x,u) = 0$” for a function $c$. However, this more general “non-separable” case may not fit into the context of the conditional validity theorem; see Theorem \[thm:cond.valid\]. We will have more to say about this in Sections \[SS:condcred\] and \[S:gcim\].
\[re:hierarchy\] The decomposition boils down to a particular hierarchical representation of the sampling model for $X$. Indeed, for functions $H$ and $T$ as in , with $V_H=\eta(U)$, and $V_T = \tau(U)$, data $X \sim {\mathsf{P}}_{X|\theta}$ can be simulated as follows.
1. Get $(V_T,V_H)$ by sampling $V_H \sim {\mathsf{P}}_{V_H}$ and $V_T \mid V_H \sim {\mathsf{P}}_{V_T|V_H}$;
2. Obtain $X$ by solving the system $H(X)=V_H$ and $T(X)=b(V_T,\theta)$.
This hierarchical model representation also provides the following insight: when $X=x$ is observed, so too is the value of $V_H$, and this knowledge can be used to update the auxiliary variable distribution, analogous to Bayes’ theorem.
\[re:sufficiency\] There are close connections between the conditional IM and Fisher’s notion of sufficiency. At a high level, both theories provide a sort of dimension reduction. The key difference between the two is that sufficiency focuses on reducing the dimension of the data, while our approach focuses on reducing the dimension of the auxiliary variable. Although the conditional IM can, in some cases, correspond to a sufficient statistic-type of reduction, this is not necessary; see the remarks at the end of Section \[SS:student\]. Proper conditioning appears to be more important than the use of sufficient statistics. In fact, in some cases, it is possible, within the IM framework, to reduce the dimension further than that which is provided by sufficiency; see Section \[S:gcim\].
\[re:bayes\] As we mentioned previously, conditional IMs have some connections to Bayes’ theorem, in particular, in how information is combined or aggregated across samples. In fact, it can be shown that, in a certain sense, the Bayes solution is a special case of conditional IMs. To see this, consider a simple but generic example. The Bayes model, cast in terms of associations, is of the following form: $$\theta = U_0, \quad U_0 \sim {\mathsf{P}}_{U_0} \quad \text{and} \quad X=a(U_0,U_1), \quad U_1 \sim {\mathsf{P}}_{U_1},$$ where ${\mathsf{P}}_U$ for $U=(U_0,U_1)$ is such that $U_1$ is conditionally independent given $U_0$. Here ${\mathsf{P}}_{U_0}$ is like the prior, and the distribution induced by $u_1 \mapsto a(\theta,u_1)$ given $U_0=\theta$ determines the likelihood. It is clear that the function $a(U_0,U_1)$ is fully observed, so the conditional IM strategy would employ the conditional distribution of $U_0$ given the observed value $x$ of $a(U_0,U_1)$. It is not hard to see that the belief function based on the “naive” predictive random set ${\mathcal{S}}=\{U_0\}$ is exactly the Bayesian posterior distribution function. So in any problem with a known prior distribution, the Bayes solution can be obtained as a special case of the conditional IM. No non-naive predictive random set is needed here because the naive IM itself is valid; this is consistent with the simple corresponding fact for posterior probabilities under a Bayes model with known prior.
\[re:more.bayes\] As a follow-up to Remark \[re:bayes\], since a full prior is not required to construct a conditional IM, it is possible to develop an inferential framework based on conditional IMs and “partial prior information.” For example, valid prior information may be available for some but not all components of $\theta$. Incorporating the prior information where it is available while remaining prior-free where it is not can be obtained by slight extension of the argument in the previous remark. This important application of conditional IMs deserves further investigation. See, also, [@xie.singh.2012].
Validity of conditional IMs {#SS:condcred}
---------------------------
Here we extend the validity results in @imbasics to the conditional IM context. The main obstacle is that the distribution function ${\mathsf{P}}_{\mathcal{S}}$, determined by the conditional distribution ${\mathsf{P}}_{V_T|H(x)}$ in , depends on data through the value $H(x)$. This is handled in Theorem \[thm:cond.valid\] below by conditioning on the observed value of $H(X)$.
Fix $h \in H({\mathbb{X}})$, and let ${\mathbb{S}}_h$ be a collection of closed ${\mathsf{P}}_{V_T|h}$-measurable subsets of ${\mathbb{V}}_T$ that contains both $\varnothing$ and ${\mathbb{V}}_T$. Like before, we also assume that ${\mathbb{S}}_h$ is nested in the sense that either $S \subseteq S'$ or $S' \subseteq S$ for all $S,S' \in {\mathbb{S}}_h$. Then ${\mathcal{S}}$ is a *conditionally admissible* predictive random set, given $h$, if its distribution ${\mathsf{P}}_{{\mathcal{S}}|h}$ satisfies $$\label{eq:cond.natural.measure}
{\mathsf{P}}_{{\mathcal{S}}|h}\{{\mathcal{S}}\subseteq K\} = \sup_{S \in {\mathbb{S}}_h: S \subseteq K} {\mathsf{P}}_{V_T|h}\{S\}, \quad K \subseteq {\mathbb{V}}_T.$$ In this case, the distribution of ${\mathcal{S}}$ depends on the particular $h$. We now have the following extension of the validity theorem to the case of conditional IMs.
\[thm:cond.valid\] For any $h$, suppose that ${\mathcal{S}}$ is conditionally admissible, given $h$, with distribution ${\mathsf{P}}_{{\mathcal{S}}|h}$ as in . If $\Theta_{T(x)}({\mathcal{S}}) \neq \varnothing$ with ${\mathsf{P}}_{{\mathcal{S}}|h}$-probability 1 for all $x$ such that $H(x)=h$, then the conditional IM is conditionally valid, i.e., for any $A \subseteq \Theta$, $$\label{eq:cond.valid.bel}
\sup_{\theta \not\in A} {\mathsf{P}}_{X|\theta}\{{\mathsf{cbel}}_{T(X)|h}(A;{\mathcal{S}}) \geq 1-\alpha \mid H(X) = h\} \leq \alpha, \quad \forall \; \alpha \in (0,1).$$
Now is a good time to recall Remark \[re:separable\]. More general decompositions of the baseline association are allowed in the discussion in Section \[SS:cim\], but only for the “separable” version is it possible to prove a conditional validity theorem. The point is that a condition like $c(X,U)=0$ does not identify a fixed subset of the sample space on which probability calculations can be restricted—the subspace would depend on $U$.
Since the calibration property in Theorem \[thm:cond.valid\] holds for all assertions $A$, we may translate to a statement in terms of the corresponding plausibility function: $$\label{eq:cond.valid.pl}
\sup_{\theta \in A} {\mathsf{P}}_{X|\theta}\{{\mathsf{cpl}}_{T(X)|h}(A;{\mathcal{S}}) \leq \alpha \mid H(X)=h\} \leq \alpha, \quad \forall\;\alpha \in (0,1).$$ So, in addition to providing an objective scale for interpreting the conditional belief and plausibility function values, provides desirable properties of conditional IM-based frequentist procedures. For example, if $h=H(x)$ is observed, the conditional $100(1-\alpha)\%$ plausibility region for $\theta$ is $\{\theta: {\mathsf{cpl}}_{T(x)|h}(\theta; {\mathcal{S}}) > \alpha\}$. Then, by , the conditional coverage probability is ${\mathsf{P}}_{X|\theta}\{{\mathsf{cpl}}_{T(X)|h}(\theta;{\mathcal{S}}) > \alpha \mid H(X)=h\} \geq 1-\alpha$. In Fisher’s mind, this is a more meaningful coverage probability since it is conditioned on a particular aspect of the observed data, namely, $H(x)=h$. In other words, the probability calculation focuses on a relevant subset $\{x: H(x)=h\}$ of the sample space. In some cases, though, conditional validity is the same as ordinary validity.
\[co:cond.valid\] Suppose that the predictive random set ${\mathcal{S}}$ does not depend on the observed $H(x)=h$, so that ${\mathsf{P}}_{{\mathcal{S}}|h} \equiv {\mathsf{P}}_{\mathcal{S}}$ and ${\mathsf{cbel}}_{T(x)|h} \equiv {\mathsf{cbel}}_{T(x)}$. Then under the conditions of Theorem \[thm:cond.valid\], the conditional IM is unconditionally valid, i.e., for any $A \subseteq \Theta$, $$\sup_{\theta \not\in A} {\mathsf{P}}_{X|\theta}\{{\mathsf{cbel}}_{T(X)}(A;{\mathcal{S}}) \geq 1-\alpha\} \leq \alpha, \quad \forall \; \alpha \in (0,1).$$
Two possible ways the condition of Corollary \[co:cond.valid\] may hold are as follows. First, in the P-step, the user may specify ${\mathcal{S}}$ directly without dependence on the observed $H(x)=h$; see Section \[SS:student\]. Second, it could happen that $V_T$ and $V_H$ are statistically independent, in which case the distribution ${\mathsf{P}}_{\mathcal{S}}$ for ${\mathcal{S}}$ is determined by the marginal distribution of $V_T$, which does not depend on $h$.
Finding conditional associations {#S:finding}
================================
Familiar things: likelihood and symmetry
----------------------------------------
In many problems, finding a decomposition and the corresponding conditional association is easy to do. In general, the definition of sufficiency implies that we can define a conditional association via, say, the marginal distribution of the minimal sufficient statistic; see Section \[SS:gamma\]. In standard problems, such as full-rank exponential families, minimal sufficient statistics are easily obtained, so this is probably the simplest approach. This, of course, includes both discrete and continuous problems. Similarly, if the problem has a group structure, invariance considerations can be used to find a decomposition; see Section \[SS:student\]. But one can consider other conditional associations if desirable. For example, when the minimal sufficient statistic has dimension larger than that of the parameter, like in curved exponential families, then some special conditioning can potentially further reduce the dimension; see Section \[SS:nile\].
A new differential equations-based technique {#SS:diffeq}
--------------------------------------------
Here we describe a novel technique for finding conditional associations, based on differential equations. The method can be used for going directly from the baseline association to something lower-dimensional. In fact, in those nice problems mentioned above, it is easy to check that this differential equation-based technique reproduces the solutions based on minimal sufficiency, group invariance, etc. However, in our experience, this new approach is especially powerful in cases where the familiar things fail to give a fully satisfactory reduction. In such cases, the differential equation-based technique can provide a further dimension reduction, beyond what sufficiency alone can give.
For concreteness, suppose $\Theta \subseteq {\mathbb{R}}$; the multi-parameter case can be handled similarly, as in Section \[SS:variance.components\]. The intuition is that $\tau$ should map ${\mathbb{U}}\subseteq {\mathbb{R}}^n$ to $\Theta$, so that $V_T=\tau(U)$ is one-dimensional, like $\theta$. Moreover, $\eta$ should map ${\mathbb{U}}$ into a $(n-1)$-dimensional manifold in ${\mathbb{R}}^n$, and be insensitive to changes in $\theta$ in the following sense. For baseline association $x=a(\theta,u)$, suppose that $u_{x,\theta}$ is the unique solution for $u$. Then for fixed $x$, we require that $\eta(u_{x,\theta})$ be constant in $\theta$. In other words, we require that ${\partial}u_{x,\theta}/{\partial}\theta$ exists and $$\label{eq:diffeq}
\underset{n \times 1}{0} = \frac{{\partial}\eta(u_{x,\theta})}{{\partial}\theta} = \underset{\text{$n \times n$, rank $n-1$}}{\frac{{\partial}\eta(u)}{{\partial}u} \Bigr|_{u=u_{x,\theta}}} \cdot \underset{n \times 1}{\frac{{\partial}u_{x,\theta}}{{\partial}\theta}}.$$ It is clear from the construction that, if a solution $\eta$ of this partial differential equation exists, then the value of $\eta(U)$ is fully observed, i.e., there is a corresponding function $H$, not depending on $\theta$, such that $H(X)=\eta(U)$. So, with appropriate choice of $\tau$, the solution $\eta$ of determines the decomposition . A different but related use of $\theta$-derivatives of the association is presented in @fraser.fraser.staicu.2010.
Formal theory on existence of solutions and on solving the differential equation system is available. For example, the method of characteristics described in @polyanin2002 is powerful tool for solving such systems. However, such formalities here will take us too far off track. Examples of this method in action are given in Section \[SS:nile\], \[SS:bvn2\], and \[SS:variance.components\]. In all three cases, this differential equations method is applied after an initial step based on sufficiency provides an unsatisfactory dimension reduction.
Three detailed examples {#S:more-examples}
=======================
A Student-t location problem {#SS:student}
----------------------------
Suppose $X_1,\ldots,X_n$ is an independent sample from a Student-t distribution ${{\sf t}}_\nu(\theta)$, where the degrees of freedom $\nu$ is known but the location $\theta$ is unknown. This is a somewhat peculiar problem because there is no satisfactory reduction via sufficiency. For the IM approach, start with a baseline association $X = \theta 1_n + U$, with $U=(U_1,\ldots,U_n)^\top$ and $U_i \sim {{\sf t}}_\nu$, independent, for $i=1,\ldots,n$. For this location parameter problem, invariance considerations suggest the following decomposition: $$X-T(X)1_n = U-T(U)1_n \quad \text{and} \quad T(X) = \theta + T(U),$$ where $T(\cdot)$ is the maximum likelihood estimator. Let $V_T=T(U)$ and $V_H = H(U)=U-T(U)1_n$. If $h$ is the observed $H(X)$, then it follows from the result of @bn1983 that the conditional distribution of $V_T$, given $V_H=h$, has a density $$f_{\nu,h}(v_T) = c(\nu,h) \prod_{i=1}^n \bigl\{ \nu + (v_T+h_i)^2 \bigr\}^{-(\nu+1)/2},$$ where $c(\nu,h)$ is a normalizing constant that depends only on $\nu$ and $h$. If we write $F_{\nu,h}$ for the distribution function corresponding to the density $f_{\nu,h}$ above, then a conditional IM for $\theta$ can be built based on the following association: $$T(X) = \theta + F_{\nu,h}^{-1}(W), \quad W \sim {{\sf Unif}}(0,1).$$ With this conditional association, we are ready for the P- and C-steps. For simplicity, in the P-step we elect to take the predictive random set ${\mathcal{S}}$ as in ; this also has some theoretical justification since $f_{\nu,h}$ should be approximately symmetric about $v_T=0$ [@imbasics Sec. 4.3.2]. For the C-step, the random set $\Theta_{T(x)}({\mathcal{S}})$ is $$\bigl[ T(x) - F_{\nu,h}^{-1}\bigl(\tfrac12+|W-\tfrac12|\bigr), \, T(x) - F_{\nu,h}^{-1}\bigl(\tfrac12-|W-\tfrac12|\bigr) \bigl], \quad W \sim {{\sf Unif}}(0,1).$$ From this point, numerical methods can be used to compute the conditional belief and plausibility functions. For example, if $A=\{\theta\}$ is a singleton assertion, then $${\mathsf{cpl}}_{T(x)|h}(\theta; {\mathcal{S}}) = 1-\bigl|1-2F_{\nu,h}(\theta-T(x)) \bigr|,$$ and the corresponding $100(1-\alpha)$% plausibility interval for $\theta$ is $$\{\theta: {\mathsf{cpl}}_{T(x)|h}(\theta; {\mathcal{S}}) > \alpha\} = \bigl( T(x) + F_{\nu,h}^{-1}(\alpha/2), \, T(x) + F_{\nu,h}^{-1}(1-\alpha/2) \bigr).$$
For illustration, we present the results of a simple simulation study. In particular, for several pairs $(n,\nu)$, 5000 Monte Carlo samples of size $n$ are obtained from a Student-t distribution with $\nu$ degrees of freedom and center $\theta = 0$. For each sample, the 95% plausibility interval for $\theta$ based on the conditional IM above is obtained. For comparison, we also compute the 95% confidence interval based on the asymptotic normality of the maximum likelihood estimate, and a 95% flat-prior Bayesian credible interval. The results of this simulation are summarized in Table \[table:t\]. We find that the results here are almost indistinguishable, so favor must go to the plausibility intervals, since these have guaranteed coverage for all $n$, while the other two are only asymptotically correct.
We also did the conditional IM calculations with an alternative decomposition, which took $V_T=U_1$ and $V_H=(0,U_2-U_1,\ldots,U_n-U_1)$. We were surprised to see that the results obtained with this “naive” decomposition were indistinguishable from those shown here based on the arguably more reasonable maximum likelihood-driven decomposition. This suggests that the particular choice of decomposition may not be so important; instead, it is the conditioning part that seems to matter most; see @fraser2004.
Fisher’s problem of the Nile {#SS:nile}
----------------------------
Suppose two independent exponential samples, namely $X_1 = (X_{11},\ldots,X_{1n})$ and $X_2 = (X_{21},\ldots,X_{2n})$, are available, the first with mean $\theta^{-1}$ and the second with mean $\theta$. The goal is to make inference on $\theta > 0$. The name comes from an application [@fisher1973] to fertility of land in the Nile river valley. In this example, the maximum likelihood estimate is not sufficient, so conditioning on an ancillary statistic is recommended.
Sufficiency considerations suggest the following initial dimension reduction step: $$S(X_1) = \theta^{-1} U_1 \quad \text{and} \quad S(X_2) = \theta U_2, \quad U_1,U_2 \sim {{\sf Gam}}(n,1),$$ where $S(X_i) = \sum_{j=1}^n X_{ij}$. But efficiency can be gained by considering a further reduction to a scalar auxiliary variable. Here we employ the differential equation technique in Section \[SS:diffeq\]. Start with $u_{x,\theta} = (\theta S(x_1), \theta^{-1}S(x_2))^\top$. Differentiating with respect to $\theta$ reveals that our (real valued) conditioning function $\eta$ must satisfy $$\frac{{\partial}\eta(u)}{{\partial}u} \Bigr|_{u=u_{x,\theta}} \binom{S(x_1)}{-\theta^{-2}S(x_2)} = 0.$$ If we take $\eta(u) = \{u_1u_2\}^{1/2}$, then $$\frac{{\partial}\eta(u)}{{\partial}u} \Bigr|_{u=u_{x,\theta}} = \frac{1}{2\{S(x_1)S(x_2)\}^{1/2}} \bigl(\theta^{-1} S(x_2), \,\theta S(x_1)\bigr)$$ and, clearly, this satisfies the differential equation above. Therefore, for , we take $$\label{eq:nile3}
H(X) = V_H \quad \text{and} \quad T(X) = \theta V_T,$$ where $T(X) = \{S(X_1)/S(X_2) \}^{1/2}$, $H(X) = \{S(X_1) S(X_2)\}^{1/2}$, $V_T = \{U_1/U_2\}^{1/2}$, and $V_H = \{U_1U_2\}^{1/2}$. These quantities are familiar from the classical approach: $T(X)$ is the maximum likelihood estimate of $\theta$, $H(X)$ is an ancillary statistic, and the pair $(T,H)(X)$ is a jointly minimal sufficient statistic [@ghoshreidfraser2010].
By and our general discussion in Section \[SS:cim\], we can focus on a conditional association based on $T(X)=\theta V_T$. The conditional distribution of $V_T$ given $V_H=h$ is a generalized inverse Gaussian distribution [@bn1977] with density function $$\label{eq:gig}
f_h(v_T) = \frac{1}{2 v_T K_0(2h)} \exp\{-h (v_T^{-1}+v_T) \},$$ where $K_0$ is the modified Bessel function of the second kind. As a final simplifying step, write the conditional association as $$\label{eq:nile4}
T(X) = \theta F_h^{-1}(W), \quad W \sim {\sf Unif}(0,1),$$ where $F_h$ is the distribution function corresponding to the density $f_h$ in . This completes the A-step. If we take ${\mathcal{S}}$ as in for the P-step, then the C-step gives $$\Theta_{T(x)}({\mathcal{S}}) = \Bigl[ \frac{T(x)}{F_h^{-1}(\tfrac12 + |W-\tfrac12|)}, \, \frac{T(x)}{F_h^{-1}(\tfrac12-|W-\tfrac12|)} \Bigr], \quad W \sim {{\sf Unif}}(0,1).$$ From this, the conditional belief/plausibility functions are readily evaluated.
For illustration, we display plausibility functions ${\mathsf{cpl}}_t(\theta;{\mathcal{S}})$ for two conditional IMs. The first is based on that derived above; the second is based on a similar derivation, but we ignore $V_H$ and simply work with the marginal distribution of $V_T$ in . Figure \[fig:nile\] shows plausibility functions for $T(x)=0.90$, with $n=20$ and true $\theta=1$, sampled from its conditional distribution given $h$, for two different values of $h$. In this case, if $h$ is large (i.e., $h > n$), then the bona fide conditional IM has narrower level sets than the naive conditional IM. The opposite is true when $h$ is small (i.e., $h < n$). This is due to the fact that the conditional Fisher information in $T$ is an increasing function in $h$; see @ghoshreidfraser2010 [Example 1]. Therefore, $T$ has more variability when $h$ is small, and this adjustment should be reflected in the plausibility function. The bona fide conditional IM catches this phenomenon while the naive one does not.
A two-parameter gamma problem {#SS:gamma}
-----------------------------
Let $X=(X_1,\ldots,X_n)$ be an independent sample from ${{\sf Gam}}(\theta_1,\theta_2)$, where $\theta_1 > 0$ and $\theta_2 > 0$ are the shape and scale parameters, respectively, both unknown. In this case, we may construct a conditional association based on the marginal distribution of the two-dimensional complete sufficient statistic, which we choose to represent as $T_1 = \sum_{i=1}^n X_i$ and $T_2 = n^{-1}\sum_{i=1}^n \log X_i - \log(T_1/n)$. Then we have a conditional association $$T_1 = \theta_2 F_{n\theta_1}^{-1}(U_1) \quad \text{and} \quad T_2 = G_{\theta_1}^{-1}(U_2),$$ where $U_1,U_2$ are independent ${{\sf Unif}}(0,1)$, $F_a$ is the distribution function of ${{\sf Gam}}(a,1)$, and $G_b$ is some distribution function without a familiar form. For the P-step, consider an analogue of the default predictive random set , given by the random square: $${\mathcal{S}}= \{(u_1,u_2): \max(|u_1-\tfrac12|,|u_2-\tfrac12|) \leq \max(|U_1-\tfrac12|,|U_2-\tfrac12|)\},$$ with $U_1,U_2 {\overset{\text{\tiny iid}}{\,\sim\,}}{{\sf Unif}}(0,1)$. In this case, with observed $t=(t_1,t_2)$, the C-step gives $$\Theta_t({\mathcal{S}}) = \{(\theta_1,\theta_2): \max(|F_{n\theta_1}(t_1/\theta_2) - \tfrac12|,|G_{\theta_1}(t_2)-\tfrac12|) \leq \max(|U_1-\tfrac12|,|U_2-\tfrac12|)\}.$$ From here, we can write down the plausibility function for a singleton assertion: $${\mathsf{cpl}}_t(\{\theta_1,\theta_2\}; {\mathcal{S}}) = 1- \max\bigl\{ |2F_{n\theta_1}(t_1/\theta_2) - 1|, \, |2G_{\theta_1}(t_2) - 1| \bigr\}^2.$$ Evaluating $G_{\theta_1}(\cdot)$ requires Monte Carlo but, since $T_2$ is $\theta_2$-ancillary, the same Monte Carlo samples can be used for all candidate $\theta_2$ values.
For illustration, we simulated a single sample of size $n=25$ from a gamma distribution with shape $\theta_1=7$ and scale $\theta_2=3$. Figure \[fig:gamma.contour\] displays a sample of size 5000 from a Bayesian posterior distribution for $(\theta_1,\theta_2)$ based on Jeffreys’ prior. Also displayed are the 90% confidence ellipse based on the asymptotic normality of the maximum likelihood estimator, and the 90% conditional IM plausibility region $$\{(\theta_1,\theta_2): {\mathsf{cpl}}_t(\{\theta_1,\theta_2\}; {\mathcal{S}}) > 0.1\}.$$ Besides having guaranteed coverage, the plausibility region captures the non-elliptical shape of the posterior. For larger $n$, all three regions will have a similar shape.
Local conditional IMs {#S:gcim}
=====================
Motivation: bivariate normal model {#SS:bvn1}
----------------------------------
So far we have seen that the conditional IM approach is successful in problems where the baseline association admits a decomposition of the form . However, as alluded to above, there are interesting and important problems where apparently no such decomposition exists. Next is one such problem, which may be considered as a “benchmark example” for conditional inference [@ghoshreidfraser2010 Example 5].
Suppose $(X_{11},X_{21}),\ldots,(X_{1n},X_{2n})$ is an independent sample from a standard bivariate normal distribution with zero means, unit variances, but unknown correlation coefficient $\theta \in (-1,1)$. A natural first step towards inference on $\theta$ is to take advantage of the fact that $X_1+X_2$ and $X_1-X_2$ are independent. In particular, by defining $$X_1 \gets \frac12 \sum_{i=1}^n (X_{1i}+X_{2i})^2 \quad \text{and} \quad X_2 \gets \frac12 \sum_{i=1}^n (X_{1i}-X_{2i})^2,$$ we may rewrite the baseline association as $$\label{eq:bvn.amodel.1}
X_1 = (1+\theta)U_1 \quad \text{and} \quad X_2 = (1-\theta)U_2, \quad U_1,U_2 \sim {{\sf ChiSq}}(n).$$ Sufficiency justifies this first reduction. Equation is equivalent to $$\label{eq:bvn.amodel.2}
\frac{X_1}{U_1} + \frac{X_2}{U_2} = 2 \quad \text{and} \quad \frac{X_1}{X_2} = \frac{1+\theta}{1-\theta} \, \frac{U_1}{U_2} .$$ The first equation depends on data and auxiliary variable—free of $\theta$—while the second depends also on $\theta$. But note that the first expression in is not of the form specified in . Actually, this first expression is of the more general “non-separable” form $c(X,U) = 0$ described in Remark \[re:separable\]. So, although provides a suitable decomposition of the baseline association, the requirements of Theorem \[thm:cond.valid\] are not met, so the resulting conditional IM may not be valid. This warrants an alternative approach.
To elaborate on this last point, observe that the distribution for $\theta$ obtained via the distribution of $(U_1,U_2)$, given $X_1/U_1 + X_2/U_2 = 2$, is exactly a type of fiducial distribution. As we pointed out in Section \[SS:notation\], conditioning on the full data, $(X_1,X_2)$ in this case, for fixed $\theta$, makes the distribution of $(U_1,U_2)$ degenerate. Therefore, “continuing to regard” $(U_1,U_2)$ as independent chi-squares, given data, may be difficult to justify.
Relaxing via localization {#SS:localization}
-------------------------
As describe above, the separability in can be too strict, but extending the conditional validity theorem to allow non-separablility appears difficult. The idea here is to relax in a different direction. Specifically, we propose to allow the pair of function $(H,\eta)$ in to depend, locally, on the parameter. This generalization allows us some additional flexibility in finding an auxiliary variable dimension reduction.
Start by fixing an arbitrary $\theta_0 \in \Theta$. As in Section \[SS:cim\], consider a pair of functions $(T,H_{\theta_0})$, depending on $\theta_0$, such that $x \mapsto (T(x),H_{\theta_0}(x))$ is one-to-one. Now take the corresponding functions $u \mapsto (\tau(u), \eta_{\theta_0}(u))$, one-to-one, such that the baseline association, at $\theta=\theta_0$, can be decomposed as $$\label{eq:gcim.decomp}
H_{\theta_0}(X) = \eta_{\theta_0}(U) \quad \text{and} \quad T(X) = b(\tau(U), \theta_0).$$ That is, , with $U \sim {\mathsf{P}}_U$, describes the sampling distribution $X \sim {\mathsf{P}}_{X|\theta_0}$. Suppose $H_{\theta_0}(X)=h_0$ is observed. We can compute the conditional distribution ${\mathsf{P}}_{V_T|h_0,\theta_0}$ of $V_T = \tau(U)$ given $\eta_{\theta_0}(U) = h_0$, which is then used to construct predictive random sets.
From this point, we may proceed exactly as before. That is, for the A-step, we get sets $\Theta_{T(x)}(v_T) = \{\theta: T(x)=b(v_T,\theta)\}$ just like before. For the P-step, we pick a conditionally admissible predictive random set ${\mathcal{S}}\sim {\mathsf{P}}_{S|h_0,\theta_0}$. Finally, the C-step produces conditional the plausibility function $${\mathsf{cpl}}_{T(x)|h_0,\theta_0}(A;{\mathcal{S}}) = 1-{\mathsf{P}}_{{\mathcal{S}}|h_0,\theta_0}\{\Theta_{T(x)}({\mathcal{S}}) \subseteq A^c\}, \quad A \subseteq \Theta.$$ We shall refer to the corresponding conditional IM as a *local* conditional IM at $\theta=\theta_0$. The adjective “local” is meant to indicate the dependence of the construction on the particular point $\theta_0$. As we see below, the validity properties of this local conditional IM are, in a certain sense, also local.
Validity of local conditional IMs {#SS:local.valid}
---------------------------------
The following theorem shows that for each $\theta_0$ value, the local conditional IM at $\theta_0$ is valid for some important assertions depending on the particular $\theta_0$. The proof is exactly like that of Theorem \[thm:cond.valid\] and, hence, omitted.
\[thm:local.valid\] For any $\theta_0$, take $h_0 \in H_{\theta_0}({\mathbb{X}})$. Suppose that ${\mathcal{S}}\sim {\mathsf{P}}_{{\mathcal{S}}|h_0,\theta_0}$ is conditionally admissible. If $\Theta_{T(x)}({\mathcal{S}}) \neq \varnothing$ with ${\mathsf{P}}_{{\mathcal{S}}|h_0,\theta_0}$-probability 1 for all $x$ such that $H_{\theta_0}(x)=h_0$, then the local conditional IM at $\theta_0$ is conditionally valid for $A=\{\theta_0\}$, i.e., $${\mathsf{P}}_{X|\theta_0}\{{\mathsf{cpl}}_{T(X)|h_0,\theta_0}(\theta_0;{\mathcal{S}}) \leq \alpha \mid H_{\theta_0}(X) = h_0\} \leq \alpha, \quad \forall \; \alpha \in (0,1).$$
The validity result here is not as strong as in Theorem \[thm:cond.valid\], a consequence of the localization. It does, however, imply that the local conditional plausibility region, $$\label{eq:cond.pl.region}
\{\theta: {\mathsf{cpl}}_{T(x)|H_\theta(x)}(\theta;{\mathcal{S}}) > \alpha\},$$ has the nominal (conditional) $1-\alpha$ coverage probability. This theoretical result is confirmed by the simulation experiment in Section \[SS:bvn2\] below. Observe that, in the definition of conditional plausibility region , the plausibility function depends on $\theta$ in two places—in the argument (the assertion) and in the local conditional IM itself. The latter structural dependence of the IM on the particular assertion is consistent with the optimality developments described in @imbasics.
Bivariate normal model, revisited {#SS:bvn2}
---------------------------------
Here we demonstrate that the localization technique can be successfully used to solve the bivariate normal problem described above. Start with the relation in . Fix $\theta_0$. To construct the functions $(H, \eta_{\theta_0})$, depending on $\theta_0$, and the corresponding local conditional IM at $\theta_0$, we shall modify the differential equation approach in Section \[SS:diffeq\].
In this case, if we let $u_{x,\theta} = (x_1/(1+\theta), x_2/(1-\theta))^\top$, then we have $$\frac{{\partial}u_{x,\theta}}{{\partial}\theta} = \Bigl(-\frac{x_1}{(1+\theta)^2}, \frac{x_2}{(1-\theta)^2}\Bigr)^\top.$$ For a local conditional IM at $\theta_0$, we propose to choose a real-valued $\eta_{\theta_0}(u)$ such that ${\partial}\eta_{\theta_0}(u_{x,\theta})$ vanishes at $\theta=\theta_0$. If we take $$\label{eq:phi}
\eta_{\theta_0}(u) = (1+\theta_0)\log u_1 + (1-\theta_0)\log u_2,$$ then $$\frac{{\partial}\eta_{\theta_0}(u)}{{\partial}u} = \Bigl(\frac{1+\theta_0}{u_1}, \,\frac{1-\theta_0}{u_2}\Bigr),$$ so the derivative of $\psi_{H_0}(u_{x,\theta})$ with respect to $\theta$ is $$\begin{aligned}
\frac{{\partial}\eta_{\theta_0}(u_{x,\theta})}{{\partial}\theta} & = \frac{{\partial}\eta_{\theta_0}(u)}{{\partial}u} \Bigr|_{u=u_{x,\theta}} \cdot \frac{{\partial}u_{x,\theta}}{{\partial}\theta} \\
& = -\frac{(1+\theta_0)^2}{x_1} \cdot \frac{x_1}{(1+\theta)^2} + \frac{(1-\theta_0)^2}{x_2} \cdot \frac{x_2}{(1-\theta)^2} \\
& = -\frac{(1+\theta_0)^2}{(1+\theta)^2} + \frac{(1-\theta_0)^2}{(1-\theta)^2}. \end{aligned}$$ The latter expression clearly evaluates to zero at $\theta=\theta_0$, so $\eta_{\theta_0}$ satisfies the desired differential equation. The corresponding function $H(x)=H_{\theta_0}(x)$ is given by $$H_{\theta_0}(x) = (1+\theta_0) \log\{x_1/(1+\theta_0)\} + (1-\theta_0)\log\{x_2 / (1-\theta_0)\}.$$ For the local conditional association—the second expression in —we take $$T(X) = z(\theta) + V_T,$$ where $T(x) = \log(x_1/x_2)$, $z(\theta) = \log\{(1+\theta)/(1-\theta)\}$, and $V_T=T(U)$. Then ${\mathsf{P}}_{V_T|\theta_0,h_0}$ is the conditional distribution of $V_T$, given $(\theta_0,h_0)$, where $h_0$ is the observed $H_{\theta_0}(X)=H_{\theta_0}(x)$. This conditional distribution has a density, given by $$f_{h_0,\theta_0}(v_T) \propto \exp\bigl\{ -n\theta_0 v_T/2 - \cosh(v_T/2) e^{(h_0-\theta_0 v_T)/2} \bigr\}.$$ If we let $F_{h_0,\theta_0}$ denote the corresponding distribution function, then we can describe this conditional association model by $$T(X) = z(\theta) + F_{h_0,\theta_0}^{-1}(W), \quad W \sim {\sf Unif}(0,1).$$ If, for the P-step, we use the predictive random set ${\mathcal{S}}$ in , then the local conditional plausibility function is $${\mathsf{cpl}}_{T(x)|h_0,\theta_0}(\theta_0; {\mathcal{S}}) = 1-\bigl| 1 - 2 F_{h_0,\theta_0}(T(x)-z(\theta_0)) \bigr|.$$ A local conditional $100(1-\alpha)$% plausibility interval for $\theta$ can be found just as before, by thresholding the plausibility function at $\alpha$. It follows from Theorem \[thm:local.valid\] that these intervals will have the nominal coverage probabilities.
For illustration, we consider a simple simulation example. We compute the local conditional 90% plausibility interval for $\theta$ in for 5000 Monte Carlo samples where, in each case, the true $\theta$ is sampled from $\{0.0, 0.3, 0.6, 0.9\}$. For several values of $n$, the estimated coverage probabilities and expected lengths are compared, in Table \[table:bvn\], to those of the conditional frequentist interval based on the so-called “$r^\star$” approximation due to @bn1986 and @fraser1990, summarized nicely in @reid1995 [@reid2003], and a Bayesian credible interval based on Jeffreys prior. The general message is that, compared to the other methods, the local conditional IM intervals have exact coverage for all $n$, though the intervals appear to be slightly longer on average when $n$ is small. But when $n$ is moderate or large, there is no apparent difference in the performance. Since one cannot hope to do much better than the Jeffreys’ prior Bayes intervals for large $n$, we see that the local conditional IM results are at least asymptotically efficient, along with being valid for all $n$.
A normal variance components model {#SS:variance.components}
----------------------------------
Consider the following standard two variance components model, $$Y^{(g)} = (Y_{g1},\ldots,Y_{gn_g})^\top \sim {{\sf N}}_{n_g}(\mu 1_{n_g}, \theta_{\varepsilon}I_{n_g} + \theta_\alpha J_{n_g}),$$ independent across $g=1,\ldots,G$. Here $G$ is the number of treatments, and $n_g$ is the number of replications under treatment $g$. Not all $n_g$ are equal, so this is an unbalanced design. The parameter of interest is $\theta=(\theta_\alpha,\theta_{\varepsilon})$, the variance components. This model corresponds to the marginal distribution of the response in a simple one-way random effects model, $$Y = \mu 1_n + Z\alpha + {\varepsilon},$$ where $n=\sum_{g=1}^G n_g$ is the total sample size, $Y$ is a $n$-vector obtained by stacking the $Y^{(g)}$s, $Z$ is a $n \times G$ binary matrix such that ${\mathsf{E}}(Y^{(g)} \mid \alpha_g) = (\mu + \alpha_g)1_{n_g}$, the random effects $\alpha_1,\ldots,\alpha_G$ are iid ${{\sf N}}(0,\theta_\alpha)$, and ${\varepsilon}$ is a $n$-vector of independent ${{\sf N}}(0,\theta_{\varepsilon})$ noise. These models are very useful in problems where variability comes from two sources. The goal is to make inference on these two sources of variation.
The common mean $\mu$ is a nuisance parameter, which we will eliminate with a transformation; the more general mixed-effects model case where $\mu$ is a linear function of some fixed covariates can be handled similarly. This marginalization can be justified within the IM framework; see @immarg. Our setup here is like that in @e.hannig.iyer.2008; the more general case, with more than two variance components, as in @cisewski.hannig.2012, shall be considered elsewhere.
Following @olsen.seely.birkes.1976, let $K$ be a $n \times (n-1)$ matrix such that $K^\top K = I_{n-1}$ and $KK^\top = I_n - n^{-1} 1_n1_n^\top$. Find the matrix $M = K^\top Z Z^\top K$ and let $\lambda_1 > \cdots > \lambda_L \geq 0$ be the distinct eigenvalues of $M$; let $r_\ell$ be the multiplicity of $\lambda_\ell$, $\ell=1,\ldots,L$. Take $P=[P_1,\ldots,P_L]$ a $(n-1) \times (n-1)$ orthogonal matrix, such that $P^\top M P$ is a diagonal matrix with the eigenvalues, in their multiplicities, fall on the diagonal. Here $P_\ell$, which corresponds to $\lambda_\ell$, is a $(n-1) \times r_\ell$ matrix. Define $$X_\ell = Y^\top K P_\ell P_\ell^\top K^\top Y, \quad \ell=1,\ldots,L.$$ Then $(X_1,\ldots,X_L)$ are minimal sufficient for $\theta=(\theta_\alpha,\theta_{\varepsilon})$, and they satisfy the distributional equations $$X_\ell = (\lambda_\ell \theta_\alpha + \theta_{\varepsilon}) U_\ell, \quad \ell=1,\ldots,L,$$ where $U_1,\ldots,U_L$ are independent, with $U_\ell \sim {{\sf ChiSq}}(r_\ell)$. In our case of an unbalanced one-way random effects model, we know that $L$ is 1 plus the number of distinct group sample sizes $n_g$. Thus, $L > 2$, and since the parameter of interest $\theta$ is two-dimensional, there is room to reduce the auxiliary variable down further from $L$ to 2. To accomplish this, we shall employ the differential equation-based technique proposed above. To make some connection to the original $Y$ sample, note that $\lambda_L=0$ and $X_L = \sum_{g=1}^G \sum_{j=1}^{n_g} (Y_{gj}-\bar Y_{g\cdot})^2$ is the usual error sum of squares. The other $X_\ell$’s, for $\ell=1,\ldots,L-1$, are also sums of squares, but these are less familiar than $X_L$.
To start, for a given $X=x$ and $\theta$, we can solve for $u$ in the above association: $$u_{x,\theta,\ell} = \frac{x_\ell}{\lambda_\ell \theta_\alpha + \theta_{\varepsilon}}, \quad \ell=1,\ldots,L-1, \quad u_{x,\theta,L} = \frac{x_L}{\theta_{\varepsilon}}.$$ Differentiating this expression with respect to both components of $\theta$ gives an $L \times 2$ matrix ${\partial}u_{x,\theta} / {\partial}\theta = {\mathrm{diag}}\{u_{x,\theta}\}W(\theta)$, where the rows of $W(\theta)$ are given by $$w_\ell(\theta) = \Bigl( -\frac{\lambda_\ell}{\lambda_\ell \theta_\alpha + \theta_{\varepsilon}}, \, -\frac{1}{\lambda_\ell \theta_\alpha + \theta_{\varepsilon}} \Bigr), \quad \ell=1,\ldots,L-1, \quad w_L(\theta) = (0, -\theta_{\varepsilon}^{-1}).$$ Choose an arbitrary localization point $\theta_0 = (\theta_{0\alpha}, \theta_{0{\varepsilon}})$. The goal is to find a function $\eta_{\theta_0}(u)$ that satisfies $$\label{eq:vc.diffeq}
\underset{(L-2) \times L}{\frac{{\partial}\eta_{\theta_0}(u)}{{\partial}u} \Bigr|_{u=u_{x,\theta}}} \cdot \underset{L \times L}{{\mathrm{diag}}\{u_{x,\theta}\}} \cdot \underset{L \times 2}{W(\theta)} = \underset{(L-2) \times 2}{0} \qquad \text{at $\theta=\theta_0$}.$$ The method of characteristics [@polyanin2002] suggests the logarithmic function $$\eta_{\theta_0}(u)^\top = \bigl( \log u_1 , \cdots , \log u_L \bigr) \Pi(\theta_0)^\top,$$ where $\Pi(\theta_0)$ is a $(L-2) \times L$ matrix with rows orthogonal to the columns of $W(\theta_0)$. Since $\Pi(\theta_0)W(\theta_0)$ vanishes, it is easy to check that holds for this $\eta_{\theta_0}$. Then the corresponding $H_{\theta_0}(x)$ satisfies $$H_{\theta_0}(x)^\top = \Bigl( \log\frac{x_1}{\lambda_1 \theta_{0\alpha} + \theta_{0{\varepsilon}}} , \cdots, \log\frac{x_{L-1}}{\lambda_{L-1} \theta_{0\alpha} + \theta_{0{\varepsilon}}}, \log\frac{x_L}{\theta_{0{\varepsilon}}} \Bigr) \Pi(\theta_0)^\top.$$ Take two orthogonal $L$-vectors which are not orthogonal to the columns of $W(\theta_0)$. One of these vectors should be $(0,\ldots,0,1)$, so that one component of the conditional association will be free of $\theta_\alpha$. The other vector can be, say, $(1,\ldots,1,0)$. Then define the mapping $\tau(u)$, taking values in ${\mathbb{R}}^2$, via the equation $$\begin{pmatrix} \eta_{\theta_0}(u) \\ \tau(u) \end{pmatrix} = \begin{pmatrix} \Pi(\theta_0) \\ 1 \; \cdots \; 1 \; 0 \\ 0 \; \cdots \; 0 \; 1 \end{pmatrix} \begin{pmatrix} \log u_1 \\ \vdots \\ \log u_L \end{pmatrix}.$$ This, in turn, defines the two-dimensional $(T_1,T_2)(x)$ to be used in the conditional association. In particular, the conditional association is given by $$\sum_{\ell=1}^{L-1} \log X_\ell = \sum_{\ell=1}^{L-1} \log(\lambda_\ell \theta_\alpha + \theta_{\varepsilon}) + \sum_{\ell=1}^{L-1} \log U_\ell, \qquad \log X_L = \log\theta_{\varepsilon}+ \log U_L,$$ and we set $T_1=\sum_{\ell=1}^{L-1} \log X_\ell$, $T_2=\log X_L$, $\tau(U)_1 = \sum_{\ell=1}^{L-1} \log U_\ell$, and $\tau(U)_2=\log U_L$. Furthermore, since this representation is linear on the log scale, and $U_1,\ldots,U_L$ are independent chi-square, the conditional distribution of $\tau(U)$, given $\eta_{\theta_0}(U) = H_{\theta_0}(x)$, can be readily found numerically. Samples from this conditional distribution, and of the corresponding predictive random set, can be obtained via Markov chain Monte Carlo.
Details on efficient implementation of the IM-based solution to this important problem, along with comparisons with other methods, will be presented elsewhere. Here we only give a brief illustration. We simulate data from the one-way random effects model above, with $\theta=(1,1)$, $\mu=0$, $G=5$, and group sample sizes$(4,4,4,8,48)$; this configuration is one considered in @e.hannig.iyer.2008 [Section 4], having moderate degree of unbalance. A box plot of the data, in Figure \[fig:vc\](a), shows evidence that suggests $\theta_\alpha > 0$. Figure \[fig:vc\](b) shows the 90% local conditional IM plausibility region for $(\log\theta_\alpha,\log\theta_{\varepsilon})$ which, in this case, contains the true parameter value $(0,0)$. This region is computed by simulating from the conditional distribution of $\tau(U)$, given $\eta_{\theta_0}(U)=H_{\theta_0}(x)$, for each $\theta_0$, with a random walk Metropolis–Hastings procedure. The predictive random set used here is an ellipse in the $\tau(U)$-space, a $L_2$ generalization of the default predictive random set , with an elasticity feature to avoid conflict [@leafliu2012]. For comparison, we also display the contours of the fiducial density for $(\log\theta_\alpha,\log\theta_{\varepsilon})$ as given in @e.hannig.iyer.2008. This indicates that the plausibility region shape is roughly consistent with the fiducial distribution, though efficiency comparisons remain to be worked out.
Discussion {#S:discuss}
==========
This paper extends the basic IM framework laid out in @imbasics by developing an auxiliary variable dimension reduction strategy. This reduction simultaneously accomplishes two goals. First, it provides a suitable combination of information across samples, and we argue in Remarks \[re:sufficiency\] and \[re:bayes\] in Section \[SS:remarks\] that Fisher’s concept of sufficiency and Bayes’ theorem can both be viewed as special cases of this combination of information via conditioning. Second, this reduction makes construction of efficient predictive random sets considerably simpler. A new differential equation technique is proposed by which an auxiliary variable dimension reduction can be found even in cases where sufficiency fails to give a satisfactory reduction. In addition, as our simulation results in Sections \[SS:student\] and \[SS:bvn2\] demonstrate, even with a default choice of predictive random set, the conditional IMs are as good or better than those standard likelihood and Bayes methods. This suggests that our proposed method of combining information is efficient. We expect that the conditional IM approach, paired with the optimal predictive random sets, will have even better performance. However, more work is needed on efficient computation of these optimal predictive random sets.
The local conditional IMs considered in Section \[S:gcim\] are an important contribution. Indeed, these tools provide a means to reduce the effective dimension even in cases where the minimal sufficient statistic has dimension greater than that of the parameter. For example, in the variance-components problem in Section \[SS:variance.components\], we identified a one-dimensional auxiliary variable to predict, even though there is no dimension reduction that can be achieved via sufficiency. The idea of focusing on validity locally at a single $\theta=\theta_0$ itself seems to provide an improvement, this is, in fact, a special case of a more general idea. One could measure locality by a general assertion $A$, not necessarily a singleton $A=\{\theta_0\}$. In this way, one can develop a conditional IM that focuses on validity at a particular assertion $A$, thus extending the range of application of local conditional IMs. Even this latter extension is a special case of a more general idea, where associations are based on generic functions of $(X,\theta,U)$, not necessarily exact formulations of the sampling model. This new idea will be explored elsewhere.
The examples in this paper focused on continuous distributions. Efficient inference in discrete problems is challenging in any framework, and IMs are no different. For nice discrete problems, e.g., regular exponential families, the IM analysis described herein can be carried out without difficulty. However, when sufficiency considerations alone provide inadequate auxiliary variable dimension reduction, new tools are needed. Here, the goal was to combine information about a single quantity coming from different sources, and conditioning is shown to be the right tool. There are other cases, however, where dimension reduction is needed because the real quantity of interest is some lower-dimensional characteristic of the full unknown parameter. For these nuisance parameter problems, a different sort of dimension reduction is needed. The companion paper [@immarg] deals with marginalization from an IM point of view.
Acknowledgments {#acknowledgments .unnumbered}
===============
This work is partially supported by the U.S. National Science Foundation, grants DMS-1007678, DMS-1208833, and DMS-1208841. The authors thank Dr. Jing-Shiang Hwang for comments on an earlier draft, as well as the helpful suggestions given by the Editor and the anonymous Associate Editor and referees.
Proofs {#S:proofs}
======
For the given predictive random set ${\mathcal{S}}$ for $U$ in the baseline association, the corresponding random set $\Theta_x({\mathcal{S}})$ in the C-step can be written as $$\begin{aligned}
\Theta_x({\mathcal{S}}) & = \bigcup_{u \in {\mathcal{S}}} \{\theta: T(x) = b(\tau(u), \theta), \, H(x)=\eta(u)\} \\
& = \bigcup_{u \in {\mathcal{S}}} \bigl[ \{\theta: T(x) = b(\tau(u),\theta)\} \cap \{\theta: H(x)=\eta(u)\} \bigr] \\
& = \bigcup_{u \in {\cal R}_{H(x)}} \{\theta: T(x) = b(\tau(u), \theta)\} \\
& = \Theta_{T(x)}( \tau({\cal R}_{H(x)}) ),\end{aligned}$$ where ${\cal R}_{H(x)} = {\mathcal{S}}\cap \{u: \eta(u)=H(x)\}$. Next, set ${\mathcal{S}}_{H(x)}=\tau({\cal R}_{H(x)})$ and let ${\mathcal{S}}_{H(x)}$ have the distribution it inherits from ${\mathcal{S}}$ through the mapping just described. It is important to note that the distribution of ${\mathcal{S}}_{H(x)}$ is not a conditional distribution of, e.g., ${\mathcal{S}}$ given that ${\mathcal{S}}\cap \{u: \eta(u)=H(x)\} \neq \varnothing$, etc; it is a well-defined distribution obtained via the mapping ${\mathcal{S}}\mapsto \tau({\mathcal{S}}\cap \{u: \eta(u)=H(x)\})$. We have shown that $\Theta_x({\mathcal{S}}) = \Theta_{T(x)}({\mathcal{S}}_{H(x)})$ with ${\mathsf{P}}_{\mathcal{S}}$-probability 1 for all $x$, so $$\Theta_x({\mathcal{S}}) \neq \varnothing \iff \Theta_{T(x)}({\mathcal{S}}_{H(x)}) \neq \varnothing \quad \text{and} \quad \Theta_x({\mathcal{S}}) \subseteq A \iff \Theta_{T(x)}({\mathcal{S}}_{H(x)}) \subseteq A.$$ Since ${\mathsf{P}}_{\mathcal{S}}\{\Theta_x({\mathcal{S}}) \neq \varnothing\} > 0$ for all $x$, the two belief functions $$\begin{aligned}
{\mathsf{bel}}_x(A;{\mathcal{S}}) & = {\mathsf{P}}_{\mathcal{S}}\{\Theta_x({\mathcal{S}}) \subseteq A \mid \Theta_x({\mathcal{S}}) \neq \varnothing\}, \\
{\mathsf{cbel}}_{T(x)}(A; {\mathcal{S}}_{H(x)}) & = {\mathsf{P}}_{\mathcal{S}}\{\Theta_{T(x)}({\mathcal{S}}_{H(x)}) \subseteq A \mid \Theta_{T(x)}({\mathcal{S}}_{H(x)}) \neq \varnothing\}, \end{aligned}$$ are well-defined conditional probabilities, i.e., no Borel paradox issues, and, moreover, must be equal for all $x$ and all assertions $A$, as the proposition claimed.
\[lem:prs.valid\] Fix $h \in H({\mathbb{X}})$ and take conditionally admissible ${\mathcal{S}}\sim {\mathsf{P}}_{{\mathcal{S}}|h}$ as in Section \[SS:condcred\]. Write $Q_{{\mathcal{S}}|h}(v_T) = {\mathsf{P}}_{{\mathcal{S}}|h}\{{\mathcal{S}}\not\ni v_T\}$. Then $Q_{{\mathcal{S}}|h}(V_T)$ is stochastically no larger than ${{\sf Unif}}(0,1)$ for $V_T \sim {\mathsf{P}}_{V_T|h}$.
Just like that of Theorem 1$'$ in @imbasics.c.
Take any $\theta \not\in A$ as the true value of the parameter. Next, note that $T(X) = b(V_T,\theta)$, with $V_T \sim {\mathsf{P}}_{V_T|h}$, characterizes the conditional distribution of $X$, given $H(X)=h$. Since $A \subset \{\theta\}^c$, monotonicity of the belief function gives $${\mathsf{cbel}}_{T(X)|h}(A;{\mathcal{S}}) \leq {\mathsf{cbel}}_{T(X)|h}(\{\theta\}^c;{\mathcal{S}}) = {\mathsf{P}}_{{\mathcal{S}}|h}\{\Theta_{T(X)}({\mathcal{S}}) \not\ni \theta\} = Q_{{\mathcal{S}}|h}(V_T).$$ By Lemma \[lem:prs.valid\], the right-hand side is stochastically no larger than ${\sf Unif}(0,1)$. This, in turn, implies the same of the left-hand side ${\mathsf{cbel}}_{T(X)|h}(A;{\mathcal{S}})$, as a function of $X \sim {\mathsf{P}}_{X|\theta}$, given $H(X)=h$. Therefore, $${\mathsf{P}}_{X|\theta}\{{\mathsf{cbel}}_{T(X)|h}(A;{\mathcal{S}}) \geq 1-\alpha \mid H(X) = h\} \leq {\mathsf{P}}\{ {\sf Unif}(0,1) \geq 1-\alpha\} = \alpha.$$ Taking supremum over $\theta \not\in A$ proves .
Since the distribution of ${\mathcal{S}}$ is free of $h$ in this case, the belief function ${\mathsf{cbel}}_{T(X)|h} \equiv {\mathsf{cbel}}_{T(X)}$ is also free of $h$. Therefore, before taking supremum in the last line of the proof of Theorem \[thm:cond.valid\], we can take expectation over $h$ to remove the conditioning, so that the validity property holds unconditionally, like in .
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We give a new, intrinsic, mass definition for spacetimes asymptotic to the Randall-Sundrum braneworld models, RS1 and RS2. For this mass, we prove a first law for static black holes, including variations of the bulk cosmological constant, brane tensions, and RS1 interbrane distance. Our first law defines a thermodynamic volume and a gravitational tension that are braneworld analogs of the corresponding quantities in asymptotically AdS black hole spacetimes and asymptotically flat compactifications, respectively. We also prove the following related variational principle for asymptotically RS black holes: instantaneously static initial data that extremizes the mass yields a static black hole, for variations at fixed apparent horizon area, AdS curvature length, cosmological constant, brane tensions, and RS brane warp factors. This variational principle is valid with either two branes (RS1) or one brane (RS2), and is applicable to variational trial solutions.'
author:
- Scott Fraser
- 'Douglas M. Eardley'
bibliography:
- 'first-law-vp.bib'
title: |
First law and a variational principle for\
static asymptotically Randall-Sundrum black holes
---
Introduction
=============
Static and stationary black holes should obey four classical laws [@Bardeen-four-laws]. The first law expresses conservation of energy, and has been proven in spacetimes with four dimensions [@Bardeen-four-laws; @sudarsky-wald] and higher dimensions [@Gibbons-first-law; @bh-enthalpy; @Gibbons-enthalpy], including a compact extra dimension [@harmark-obers-def-tension; @*harmark-obers-first-law; @*kol-sorkin-piran], but not in spacetimes asymptotic to the Randall-Sundrum (RS) braneworld models [@RS1; @RS2]. In this paper, we close this gap by proving a general first law for static asymptotically RS black holes. The first law relates the variations of mass and other physical quantities. For this purpose, we provide a new, intrinsic, mass definition. Our first law defines a thermodynamic volume and a gravitational tension that are braneworld analogs of thermodynamic volume in asymptotically AdS black hole spacetimes [@bh-enthalpy; @Gibbons-enthalpy] and gravitational tension in asymptotically flat compactifications [@harmark-obers-def-tension; @*harmark-obers-first-law; @*kol-sorkin-piran].
The RS models are phenomenologically interesting, and have holographic interpretations [@RS-AdS-CFT] in the AdS/CFT correspondence. In the RS models, our observed universe is a brane surrounded by an AdS bulk. The bulk is warped by a negative cosmological constant. The RS1 model [@RS1] has two branes of opposite tension, with our universe on the negative-tension brane. Tuning the interbrane distance appropriately predicts the production of small black holes at TeV-scale collider energies [@Banks-Fischler; @*Giddings], and LHC experiments [@LHC-RS-extra-dimensions; @*LHC-RS-black-holes] continue to test this hypothesis. In the RS2 model [@RS2], our universe resides on the positive-tension brane, with the negative-tension brane removed to infinite distance. Perturbations of RS2 reproduce Newtonian gravity at large distance on the brane, while in RS1 this requires a mechanism to stabilize the interbrane distance [@Tanaka]. In RS2, solutions for static black holes on the brane have been found numerically, for both small black holes [@Kudoh-smallBH-1; @*Kudoh-smallBH-2; @*Kudoh-smallBH-6D] and large black holes [@Figueras-Wiseman; @*Abdolrahimi], compared to the AdS curvature length. The only known exact analytic black hole solutions are the static and stationary solutions [@ehm; @*ehm-2] in a lower-dimensional version of RS2. An exact solution for a large black hole on the brane was also found in [@nonempty-bulk] in a generalized RS2 setup with matter in the bulk.
In general, the first law for a static or stationary black hole takes the form $\delta M = (\kappa/8\pi G) \delta A + \sum_i p_i \,\delta Q_i$. This relates the variations of mass $M$, horizon area $A$, and other physical quantities $Q_i$. Thus, if a black hole is static or stationary, it extremizes $M$ under variations that hold constant the remaining variables ($A$, $Q_i$). The converse of this statement motivates a variational principle: [*If a black hole’s exterior spatial geometry is initially static (or initially stationary) and extremizes the mass with other physical variables held fixed, then the black hole is static (or stationary).*]{} In this variational principle, the specific variables to hold fixed depend on the form of the first law. The appropriate area to hold fixed is that of the black hole apparent horizon, which is determined by the spatial geometry alone (unlike the event horizon, which is a global spacetime property). The apparent horizon generally lies inside the event horizon, and coincides with it for a static or stationary black hole spacetime.
For asymptotically flat black holes in four spacetime dimensions, a version of the above variational principle was proved by Hawking for stationary black holes [@hawking], and was extended to Einstein-Yang-Mills theory by Sudarsky and Wald [@sudarsky-wald; @chrusciel-wald]. In this paper, we prove a version of the above variational principle for static asymptotically RS black holes. The quantities held fixed in our variational principle are the AdS curvature length, cosmological constant, brane tensions, and RS values (at spatial infinity) of warp factors on each brane. The variations of these quantities appear in the general first law that we prove in this paper.
This paper is organized as follows. After reviewing the RS spacetimes in section \[RS-review\], we define the mass for an asymptotically RS spacetime in section \[sec:mass\], and evaluate the mass for a static asymptotic solution in section \[sec:static-soln\]. We prove the first law for static black holes in section \[sec:first-law\]. We prove the variational principle in section \[vp-proof\], including an explicit application using a trial solution. We conclude in section \[sec:conclusion\].
Throughout this paper, we use two branes, so our results apply to either RS1 or RS2 in the appropriate limit. We work on the orbifold region (between the branes) and use $D$ spacetime dimensions. A timelike surface has metric $\gamma_{ab}$, extrinsic curvature $K_{ab}=\gamma_a{}^c\nabla_c n_b$, and outward unit normal $n_b$. A spatial hypersurface $\Sigma$ has unit normal $u_a$, metric $h_{ab}$, and covariant derivative $D_a$. Each boundary $B$ of $\Sigma$ has metric $\sigma_{ab}$, extrinsic curvature $k_{ab}=h_a{}^c D_c n_b$, and outward unit normal $n_b$. The boundaries $B$ of $\Sigma$ are illustrated in Fig. \[initial-data-figure\]. These boundaries are: spatial infinity $B_\infty$, the branes ($B_1$, $B_2$), and the black hole apparent horizon $B_{H}$. If the black hole is static, then $B_{H}$ coincides with the black hole event horizon in $\Sigma$.
![\[initial-data-figure\] Illustration of a spatial hypersurface $\Sigma$, for a black hole with apparent horizon $B_H$ not intersecting the branes, $B_1$ and $B_2$. Spatial infinity $B_\infty$ is a single boundary, transverse to both branes. Each boundary $B$ has outward normal $n$. ](fig1-eps-converted-to.pdf)
The Randall-Sundrum spacetimes \[RS-review\]
============================================
The RS spacetimes [@RS1; @RS2] are portions of an anti-de Sitter (AdS) spacetime, with metric $$\label{g_RS}
ds^2_{\rm RS}
=
\Omega(Z)^2 \left(- dt^2 + d\rho^2 + \rho^2d\omega_{D-3}^2 + dZ^2 \right) \ .$$ Here $d\omega_{D-3}^2$ denotes the unit $(D-3)$-sphere. The warp factor is $\Omega(Z)=\ell/Z$, with values $\Omega_i$ on each brane. Here $\ell$ is the AdS curvature length, related to the bulk cosmological constant $\Lambda<0$ given below. The RS1 model [@RS1] contains two branes, which are the surfaces $Z=Z_i$ with brane tensions $\lambda_i$, where $i=1,2$. The brane tensions $\lambda_i$ and bulk cosmological constant $\Lambda$ are $$\label{RS-tensions}
\lambda_1 = -\lambda_2 =
\frac{2(D-2)}{8\pi G_D\ell}
\ , \quad
\Lambda = -\,\frac{(D-1)(D-2)}{2\ell^2} \ .$$ The dimension $Z$ is compactified on the orbifold $S^1/\mathbb{Z}_2$ and the branes have orbifold mirror symmetry: in the covering space, symmetric points across a brane are identified. There is a discontinuity in the extrinsic curvature $K_{ab}$ across each brane given by the Israel condition [@israel]. Using orbifold symmetry, the Israel condition requires the extrinsic curvature at each brane to satisfy $$\label{RS-K}
2 K_{ab}
= \left(\frac{8\pi G_D\lambda}{D-2}\right) \gamma_{ab}
\ , \quad
2k_{ab} = \left(\frac{8\pi G_D\lambda}{D-2}\right) \sigma_{ab} \ .$$ Using (\[RS-tensions\]), this can also be written as $$2 K_{ab}
= \frac{\varepsilon}{\ell}\gamma_{ab}
\ , \quad
2k_{ab} = \frac{\varepsilon}{\ell} \sigma_{ab} \ ,$$ where $\varepsilon=\pm 1$ is the sign of each brane tension. The RS2 spacetime [@RS2] is obtained from RS1 by removing the negative-tension brane (now a regulator) to infinite distance ($Z_2 \rightarrow \infty$) and the orbifold region has $Z \ge Z_1$.
Mass definition \[sec:mass\]
============================
For an asymptotically RS spacetime, we will define the mass $M$ using a counterterm method [@balasubramanian-kraus; @*kraus-larsen-siebelink]. This is an intrinsic approach, which is well suited to the variations we will perform in section \[sec:first-law\] to prove the first law, and is also useful for spacetimes with nontrivial topology. By comparison, other definitions, such as the Brown-York mass [@brown-york], use an auxiliary reference spacetime, which may be awkward when the topology is nontrivial. A reference spacetime approach is also implicit when the asymptotic geometry serves this purpose, as in the Abbott-Deser mass [@abbott-deser] and its counterpart with an asymptotically flat compact dimension, the Deser-Soldate mass [@Deser-Soldate]. For an asymptotically RS spacetime, it is straightforward to verify that evaluating our mass definition, as in (\[mass\]) below, reduces to the same result as the Abbott-Deser mass formula [@abbott-deser].
In the counterterm approach [@balasubramanian-kraus; @*kraus-larsen-siebelink], for a spacetime with metric $g_{ab}$, one first evaluates the bare action $\widetilde S$. If this diverges, one constructs an action counterterm $S_{ct}$ to render the sum $S=\widetilde S+S_{ct}$ finite, as follows. Let the metric $g_{ab}$ asymptote to $g_{ab}^{(0)}$ whose bare action $\widetilde S_0$ also diverges. We express $\widetilde S_0$ in terms of its intrinsic boundary invariants, and define the action counterterm $S_{ct}=-\widetilde S$ where $\widetilde S$ is the same functional of its boundary geometry that $\widetilde S_0$ is of its boundary geometry. This gives $S_0=0$ for $g_{ab}^{(0)}$ and a finite action $S$ for $g_{ab}$.
To define the mass $M$, one proceeds from the action to the Hamiltonian (defined on an arbitrary initial value spatial hypersurface $\Sigma$), which is given by a bulk term involving initial value constraints, and surface terms. For a solution to the constraints, the bulk term vanishes and the bare mass at spatial infinity is [@brown-york] $$\label{M-bare-formula}
\widetilde M = - \, \frac{1}{8\pi G_D}\int_{B_\infty} d^{D-2}x\,N
\sqrt{\sigma} \, k \ .$$ Here $k$ is the extrinsic curvature of the boundary $B_\infty$ and for a static spacetime, the lapse function is $N=\sqrt{-g_{tt}}$. If $\widetilde M$ diverges, one constructs a mass counterterm $$\label{M-ct-formula}
M_{ct} = \int_{B_\infty} d^{D-2} x \, N\sqrt{\sigma} \, u_a u_b \left[
\frac{2}{\sqrt{-\gamma}} \frac{\delta
S_{ct}}{\delta\gamma_{ab}} \right]$$ such that the mass $M$ is finite, defined by $$\label{finite M}
M = \widetilde M + M_{ct} \ .$$ Following the above procedure, we begin with the RS spacetime. The bare action $\widetilde S_{\rm RS}$ consists of a bulk term $S_\Sigma$ and a Gibbons-Hawking term at each boundary, $$\label{S_RS}
\widetilde S_{\rm RS} = S_{\Sigma} + S_1 + S_2 + S_\infty \ .$$ Here $$\begin{aligned}
\nonumber
S_{\Sigma} &=& \frac{1}{16\pi G_D} \int dt \int_{\Sigma}
d^{D-1}x\,\sqrt{-g}\left( R -2\Lambda\right) \ ,
\\
\nonumber
S_i &=& \frac{1}{8\pi G_D}\int dt \int_{B_i} d^{D-2}x\,\sqrt{-\gamma}
\left(K-\frac{8\pi G_D\lambda_i}{2}\right) \ ,
\\
S_\infty &=& \frac{1}{8\pi G_D} \int dt \int_{B_\infty} d^{D-2}x\,\sqrt{-\gamma}\,
K \ .\end{aligned}$$ For the RS solution (\[g\_RS\]), the Ricci scalar is $R=2\Lambda D/(D-2)$ and after integrating $ S_{\Sigma}$ in $Z$, we find $$S_{\Sigma} + S_1 + S_2 = 0 \ .$$ We now specialize to the case $D=5$, for which $$\label{S-RS-1} \widetilde S_{\rm RS}
= S_\infty
= \frac{1}{8\pi G_5} \int d^{4}x\,\sqrt{-\gamma}\, \frac{2}{\Omega \rho}
\ .$$ This diverges as $\rho \rightarrow \infty$. The metric $\gamma_{ab}$ on this boundary is (\[g\_RS\]) with $\rho=$ constant. Let $\hat\gamma_{ab}$ be the submetric with $Z=$ constant, and let $\hat\sigma_{ab}$ be the submetric on a 2-sphere (constant $\rho$, $Z$, $t$). Their Ricci scalars are $$\hat{\cal R}(\hat\gamma) =\hat{\cal R}(\hat\sigma)
= \frac{2}{(\Omega \rho)^2 }\ .$$ If we express $\widetilde S_{\rm RS}$ in terms of $\hat{\cal R}$, then an asymptotically RS spacetime has action counterterm $
S_{ct} = -\widetilde S
$ where $\widetilde S$ is the same functional of its boundary geometry that $\widetilde S_{\rm RS}$ is of its geometry. Thus $$\label{S-ct-hybrid}
S_{ct} = - \, \frac{\sqrt{2}}{8\pi G_5} \int d^4x\,
\sqrt{-\gamma}\,\sqrt{\hat{\cal R}}
\ .$$ All quantities in (\[S-ct-hybrid\]) refer to the boundary geometry of a general metric $g_{ab}$, not the RS metric (\[g\_RS\]). As shown in Appendix \[appendix-M-ct\], the mass counterterm (\[M-ct-formula\]) then yields our mass definition for an asymptotically RS spacetime, $$\label{mass-geometric}
M =
\frac{1}{8\pi G_5}
\int_{B_\infty} d^{3} x\, N \sqrt{\sigma} \left(- k + \sqrt{2\hat{\cal R}}\right)
\ .$$
Static asymptotic solution \[sec:static-soln\]
==============================================
Here we evaluate our mass definition (\[mass-geometric\]) for a static asymptotic solution, which we will use in section \[sec:first-law\] below. We define the branes as the surfaces $Z=Z_1$ and $Z=Z_2$. We consider a static asymptotically RS metric with functions $F_\nu$ that fall off at large $\rho$ as follows, $$\begin{aligned}
\nonumber
ds^2 &=&
\Omega^2 \left(-e^{2F_t}dt^2 +
e^{2F_\rho}d\rho^2 + e^{2F_\omega}{\rho^2d\omega_{2}^2} + e^{2F_Z}dZ^2 \right)
\\
\label{ansatz-1}
F_\nu &=& \frac{a_\nu(Z)}{\rho} +
\frac{b_\nu(Z)}{\rho^2} + \frac{c_\nu(Z)}{\rho^3}
+ O(1/\rho^4)\ .\end{aligned}$$ For these asymptotics, the mass (\[mass-geometric\]) evaluates to $$M =
\label{mass}
\frac{1}{G_5} \int_{Z_1}^{Z_2} dZ\, \Omega^3
\left( a_\rho + \frac{a_Z}{2}
\right) \ .$$ The value of $M$ also appears in the solution to (\[ansatz-1\]). To see this, we find it is necessary to solve the Einstein equations through third order. The first order solutions $a_\nu$ are
\[Order1-3\] $$\begin{aligned}
\label{a_t}
a_t(Z) &=& a_\rho(Z) + \mu_0 \ ,
\\
a_\omega(Z) &=& a_\rho(Z) + \mu_1 \ ,
\\
a_Z(Z) &=& -\,Z \,a_\rho^\prime \ .
\end{aligned}$$
Here $\mu_0$ and $\mu_1$ are integration constants and $^\prime=d/dZ$. The constant $\mu_1$ can be removed by a gauge transformation $\rho \rightarrow \rho + \mu_1/2$. Two identities we will need are
$$\begin{aligned}
\label{identity1}
\int_{Z_1}^{Z_2} dZ\,\Omega^3\left(a_t + a_\rho + a_Z \right) &= 0 \ ,
\\
\label{int-field} \int_{Z_1}^{Z_2} dZ\,\Omega^{n+1} \left(n \,a_\rho +
a_Z \right)
&=
- \ell
\left(\Omega^{n} a_\rho\right)\Big|_{Z_1}^{Z_2} \ .\end{aligned}$$
The identity (\[int-field\]) also holds with $a_\rho$ replaced by $a_t$ or $a_\omega$. Using (\[int-field\]), the mass can be written as $$\label{M-local}
M = -\,\frac{\ell}{2G_5}\left(\Omega^2 a_\rho\right)\Big|^{Z_2}_{Z_1}
\ .$$ The third order solutions $c_\nu$ involve an integration constant $q_0$. The Israel condition provide one equation at each brane, which can be solved for $\mu_0$ and $q_0$ as $$\begin{aligned}
\label{mu_0}
\mu_0\left(\Omega_2{}^2-\Omega_1{}^2\right)
&=&
-2\left(\Omega^2 a_\rho\right)\Big|^{Z_2}_{Z_1} \ ,
\\
\label{q_0}
q_0 (\Omega_2^{-2} -\Omega_1^{-2}) &=&
2 a_\rho\Big|_{Z_1}^{Z_2}
\ .\end{aligned}$$ Here $\Omega_i=\ell/Z_i$ the warp factor at each brane. We see from (\[int-field\]) and (\[mu\_0\]) that $M$ is proportional to $\mu_0$, $$\label{M and mu0}
M = \frac{\ell\mu_0}{4G_5} \left(\Omega_2{}^2-\Omega_1{}^2\right)
\ .$$ We see from (\[q\_0\]) that $q_0$ is proportional to a quantity $\mathcal{Q}$ parametrizing the interbrane distance near $\rho \rightarrow \infty$, $${\cal L}_{\rm branes}
= L + \frac{\mathcal{Q}}{\rho} + O(1/\rho^2) \ ,$$ where the distance $L$ at infinity and the constant $\cal Q$ are $$\label{L asymp}
L
= \ell\ln \left(\frac{\Omega_1}{\Omega_2}\right)
\quad , \quad
\mathcal{Q} = -\ell a_\rho\Big|^{Z_2}_{Z_1} \ .$$ We will refer to $L$ and $\cal Q$ in the next section. We will also use the fact that $M$ and $\cal Q$ appear in the values of $a_\rho$ at each brane, which we find by solving (\[M-local\]) and (\[L asymp\]), $$\label{a-M-Q}
a_\rho(Z_i)=\frac{2G_5M -\Omega_j{}^2 {\cal Q}}{\ell(\Omega_1{}^2-\Omega_2{}^2)}
\quad , \quad j \neq i \ .$$ In the RS1 case, $M$ and $\cal Q$ have lower-dimensional interpretations on each brane. This follows since there is an effective Brans-Dicke gravity on each brane [@garriga-tanaka], and Brans-Dicke gravity contains two asymptotic quantities, a tensor mass and a scalar mass [@Lee-BD-masses]. On each brane, one can verify that $M$ and $\cal Q$ are proportional to the effective tensor mass and scalar mass, respectively.
First law for static black holes \[sec:first-law\]
==================================================
Preliminary form
----------------
For a static or stationary black hole, the first law relates the variations of mass, black hole horizon area, and other physical quantities. We will include variations of the bulk cosmological constant $\Lambda$ and brane tensions $\lambda_i$ that preserve the RS conditions (\[RS-tensions\]), $$\label{d-ell}
-\, \frac{\delta\ell}{\ell}
= \frac{\delta\lambda_1}{\lambda_1}
= \frac{\delta\lambda_2}{\lambda_2}
= \frac{\delta\Lambda}{2\Lambda} \ .$$ We will also include the variation $\delta L$ of the interbrane separation. From (\[L asymp\]), $$\label{dL}
\delta L = \delta\ell\,\ln\left(\frac{\Omega_1}{\Omega_2}\right)
+ \ell \, \frac{\delta\Omega_1}{\Omega_1}
-\ell \, \frac{\delta\Omega_2}{\Omega_2}
\ .$$
Our setup is general: it applies to a black hole localized on a brane, or isolated in the bulk (away from either brane), and also applies to the asymptotically RS black string [@black-string]. Our method is based on the Hamiltonian approach of [@sudarsky-wald], with the additional variations (\[d-ell\])(\[dL\]). The full Hamiltonian contains a bulk term and boundary terms. The bulk term $H_{\Sigma}$ is defined on a spatial hypersurface $\Sigma$, $$\label{H_sigma}
H_{\Sigma} =\int_{\Sigma} d^{D-1} x\, \left( N
{\cal C}_0 + N^a {\cal C}_{a} \right) \ .$$ Here $N$ and $N^a$ are the lapse and shift in the standard decomposition of the spacetime metric. Our focus is the initial data ($h_{ab}$, $p^{ab}$) on $\Sigma$, where $h_{ab}$ is the spatial metric and $p^{ab}$ is its canonically conjugate momentum, $$16\pi G_D \, p^{ab}=\sqrt{h} {\cal K}^{ab} - {\cal K}\, h^{ab}
\quad , \quad
{\cal K}_{ab} = h_a{}^c \nabla_c u_b
\ .$$ Initial data must satisfy constraints, ${\cal C}_0 = 0$ and ${\cal C}_a= 0$, which we henceforth assume, where $$\begin{aligned}
\nonumber {\cal C}_0 &=&
\frac{\sqrt{h}}{16\pi G_D}\left(2\Lambda - {\cal R}\right) +
\frac{16\pi G_D}{\sqrt{h}} \left( p^{ab} p_{ab} - \frac{ p^2}{D-2}
\right) \ ,
\\
{\cal C}_a &=& -2 D_b p_a{}^b \ .
\label{constraints}\end{aligned}$$ Here $\cal R$ and $D_a$ are the Ricci scalar and covariant derivative associated with $h_{ab}$. We now consider the change $\delta H_{\Sigma}$ under variations ($\delta h_{ab}$, $\delta p^{ab}$). One finds $\delta{\cal C}_0$ and $\delta{\cal C}_a$ involve derivatives ($D_c\delta h_{ab}$, $D_c\delta p^{ab}$). Integrating by parts to remove these derivatives yields surface terms $I_B$, $$\begin{aligned}
\nonumber
\delta H_{\Sigma} &=& \int_{\Sigma} d^{D-1}x
\left[
{\cal P}^{ab}\delta h_{ab} + {\cal H}_{ab}\, \delta p^{ab}
\right]
\\
& &
+ \
\frac{\delta\Lambda}{8\pi G_D}
\int_{\Sigma} d^{D-1}x \,
N \sqrt{h}
\ + \
\sum_B I_B
\label{delta_H_sigma-1}
\ .\end{aligned}$$ The sum in (\[delta\_H\_sigma-1\]) is over all of the boundaries $B$ illustrated in Fig. \[initial-data-figure\]. The quantities ${\cal P}_{ab}$ and ${\cal H}_{ab}$ appear in the time evolution equations, $$\label{evolution}
\dot h_{ab} = {\cal H}_{ab} \ , \quad
\dot p^{ab}=-{\cal P}^{ab} \ ,$$ where the overdot denotes the Lie derivative along the time evolution vector field $t^a= N u^a + N^a$ with $u^a$ the unit normal to $\Sigma$. We will not need the most general forms of ${\cal P}_{ab}$, ${\cal H}_{ab}$, and $I_B$. We will give their simplified forms below, after implementing some of our key assumptions.
We now assume our variations take one solution of the constraints to another solution of the constraints, so we take $\delta{\cal C}_0=0$ and $\delta{\cal C}_a=0$. Then the variation of (\[H\_sigma\]) immediately gives $$\label{delta_H_sigma-2}
\delta H_{\Sigma}=0 \ .$$ We henceforth assume the initial data is instantaneously static, for which $p^{ab}=\delta p^{ab}=0$ and we take $N^a=0$. One then explicitly finds ${\cal H}_{ab}=0$, so (\[delta\_H\_sigma-1\]) and (\[delta\_H\_sigma-2\]) give $$0 = \int_{\Sigma} d^{D-1}x
\left[
{\cal P}^{ab}\delta h_{ab}
+
\frac{\delta\Lambda}{8\pi G_D}
N \sqrt{h}
\right]
\ + \
\sum_B I_B
\label{delta_H_sigma-1b}
\ .$$ This result will be the primary equation for proving our variational principle in section \[vp-proof\]. Here ${\cal P}^{ab}$ is given by $${\cal P}^{ab} = \frac{\sqrt{h}}{16\pi G_D}
\left(
{\cal R}^{ab}+ h^{ab}D_c D^c - D^a D^b
\right)N \ .$$ We now assume a static black hole with timelike Killing field $\xi^a$ and choose $t^a=\xi^a$. For a static solution, ${\cal P}^{ab}=0$ by (\[evolution\]). Then (\[delta\_H\_sigma-1b\]) gives $$\label{first-law-1}
0 = \frac{\delta\Lambda}{8\pi G_D}\int_{\Sigma} d^{D-1} x\, N
\sqrt{h} \ + \ \sum_B I_B \ .$$ This equation is our preliminary form of the first law. It simply remains to express (\[first-law-1\]) in terms of physical quantities. Each surface term $I_B$ can be written [@brown-lau-york] $$I_B =
\int_{B} d^{D-2}x\, N
\left[
\frac{\delta\left(\sqrt{\sigma} k\right)}{8\pi G_D}
- \frac{\sqrt{\sigma}}{2} s^{ab}\delta\sigma_{ab}
\right]
\label{surface-terms-2}
- J_B \ ,$$ where $$\begin{aligned}
\label{stress}
8\pi G_D\, s^{ab} &= - k^{ab} + \left[k + n^c
(D_cN)/N\right]\sigma^{ab} \ ,
\\
16\pi G_D J_B &=
\sum_{B^\prime \not= B}\int_{B \cap B^\prime}
d^{D-3}x \, N\sqrt{\hat\sigma}\, n^\prime_a\, \sigma^a{}_b\,
\delta n^b \ .\end{aligned}$$ Here $\hat\sigma$ denotes the metric on $B \cap B^\prime$. In what follows, we will have $J_B=0$. This is due to $N=0$ on $B_H$, and due to orthogonality ($n^\prime_a n^a=0$) at the other boundaries $B$.
We now evaluate the boundary terms (\[surface-terms-2\]) for $D=5$. The results at the horizon $B_H$ and the branes $B_i$ are $$\label{da}
I_{B_{H}}
=\frac{\kappa}{8\pi G_5}\, \delta A
\quad , \quad
I_{B_i} =
\frac{\delta\lambda_i}{2} \int_{B_i} d^{3}x\,N \sqrt{\sigma}
\ ,$$ with $A$ the black hole horizon area. These results are straightforward to derive. At the black hole horizon $B_H$, we have $N=0$ and $D_aN = -\kappa\, n_a$ where $\kappa$ is the constant surface gravity [@sudarsky-wald]. This gives $8\pi G_5 N s^{ab} = -\kappa\,\sigma^{ab}$ and the result in (\[da\]) follows. At each brane $B_i$, we use $n^c (D_cN)/N =K-k$, which is a general result [@brown-york] valid when $u^a n_a=0$. Using (\[RS-K\]) gives $s^{ab} =(\lambda_i/2)\sigma^{ab}$ which yields the result in (\[da\]). In Appendix \[appendix-first-law\], we show the boundary term $I_{B_\infty}$ is $$\label{text-I-infty-3}
I_{B_\infty} = - \delta M
+ {\cal F}_\infty \, \delta \ell
+ {\cal U}_1 \frac{\delta\Omega_1}{\Omega_1}
- {\cal U}_2 \frac{\delta\Omega_2}{\Omega_2}
\ ,$$ where the boundary quantity ${\cal F}_\infty$ at infinity is $$\label{F-infty}
{\cal F}_\infty =
- \frac{1}{2G_5\ell}\int_{Z_1}^{Z_2} dZ\, \Omega^3\left( a_t + 2a_Z \right) \ ,$$ and the coefficients $ {\cal U}_i$ are $$\label{U-def}
{\cal U}_i =
M\left( \frac{\Omega_i{}^2 }{\Omega_1{}^2-\Omega_2{}^2}\right)
-
\frac{3 {\cal Q}}{2G_5}
\left(
\frac{\Omega_1{}^2\Omega_2{}^2}{\Omega_1{}^{2}-\Omega_2{}^{2}}
\right) \ .$$ Here $\cal Q$ parametrizes the asymptotic interbrane separation (\[L asymp\]). We also define $\cal F$ by the following sum, $$\label{F-def}
{\cal F} \,\delta \ell =
\frac{\delta\Lambda}{8\pi G_5}\int_{\Sigma} d^{4}x\, N
\sqrt{h}
+ I_{B_1} +\ I_{B_2} + {\cal F}_\infty \delta\ell \ .$$ Here the two brane terms $I_{B_i}$ render the volume integral finite, as one can verify. The term ${\cal F}_\infty$ renders ${\cal F}$ gauge invariant, as shown in Appendix \[gauge invariance\]. We also define ${\cal V}$ by $$\label{V-def}
{\cal V}
= \left(\frac{\ell}{2P}\right) {\cal F}
\quad ,
\quad
-\,{\cal V} \, \delta P = {\cal F} \,\delta \ell \ ,$$ where $P = -\Lambda/(8\pi G_5)$ is the pressure due to the cosmological constant. We have now evaluated all the terms needed to rewrite the preliminary first law (\[first-law-1\]).
The first law
-------------
We will give four versions of the first law, corresponding to different choices of variations. Substituting (\[da\]), (\[text-I-infty-3\]), and (\[V-def\]) into (\[first-law-1\]) gives the first law in the form $$\label{first-law-v1}
\delta M = \frac{\kappa \,\delta A}{8\pi G_5}
- {\cal V}\, \delta P
+ {\cal U}_1 \frac{\delta\Omega_1}{\Omega_1}
- {\cal U}_2 \frac{\delta\Omega_2}{\Omega_2}
\ .$$ The area term is standard. The last two terms are changes in mass due to changes in the branes’ gravitational field, since $\delta\Omega_i$ are variations of gravitational redshift factors on each brane. The last term is absent in RS2, which removes the negative-tension brane to $Z_2 \rightarrow \infty$, for which $\Omega_2 \rightarrow 0$ and ${\cal U}_2/\Omega_2 \rightarrow 0$ by (\[U-def\]).
For discussion purposes, we will take ${\cal V} >0$. This is easily verified for the static asymptotically RS black string [@black-string], which is the only known exact solution for an asymptotically RS black object in 5-dimensional spacetime. We will also see in (\[tensions-def\]) below that ${\cal V} >0$ if and only if a gravitational tension ${\cal T}_0$ is positive.
The coefficient of $\delta P$ defines a thermodynamic volume $V_{\rm eff}$ in a black hole first law [@KNAdS; @bh-enthalpy; @Gibbons-enthalpy]. For a static asymptotically AdS black hole, it was found in [@bh-enthalpy] that $V_{\rm eff}>0$ is the volume *removed* by the black hole (the volume of pure AdS space minus the volume outside the black hole). In our first law, $V_{\rm eff}=-{\cal V}<0$ suggests that net volume is *added* outside the black hole (compared to the case with no black hole). Added volume makes sense physically: in RS2 the black hole repels the positive-tension brane, and in RS1 we would expect a version of the black hole Archimedes effect [@Archimedes-1; @*Archimedes-2; @Archimedes-3-tension], where the black hole increases the size of the compact dimension (here the interbrane distance). We also note that $V_{\rm eff}<0$ occurs, with a natural interpretation as an added volume, in AdS-Taub-NUT-AdS spacetime [@negative-volume].
In RS1, there are three ways the variation $\delta L$ of the interbrane distance can be introduced into the first law, using (\[dL\]). In each case, the coefficient of $\delta L$ defines a gravitational tension ${\cal T}$ that depends on which quantities are held fixed. The three gravitational tensions we refer to below are $$\label{tensions-def}
{\cal T}_0 = \frac{2P\,\cal V}{L}
\quad , \quad
{\cal T}_1 = \frac{{\cal U}_1}{\ell}
\quad , \quad
{\cal T}_2 = \frac{{\cal U}_2}{\ell} \ .$$ Using (\[dL\]) in (\[first-law-v1\]) to change variables from $\delta\ell$ to $\delta L$ gives $$\label{first-law-v2}
\delta M = \frac{\kappa \,\delta A}{8\pi G_5}
+ {\cal T}_0 \,\delta L
+ \sum_{i=1,2} \pm \Big({\cal U}_i - {\cal T}_0\,\ell \Big) \frac{\delta\Omega_i}{\Omega_i} \ .$$ Here $\pm$ is the sign of each brane tension $\lambda_i$ and ${\cal T}_0$ is a gravitational tension at fixed values of ($A$, $\Omega_1$, $\Omega_2$). Using (\[dL\]) in (\[first-law-v1\]) to eliminate $\delta \Omega_1$ or $\delta \Omega_2$ gives the first law as $$\label{first-law-v3}
\delta M = \frac{\kappa \,\delta A}{8\pi G_5}
+ {\cal T}_1\, \delta L
+ \left(-{\cal V} + \frac{{\cal T}_1 \, L}{2P} \right) \delta P
+ M \frac{\delta\Omega_2}{\Omega_2}$$ and $$\label{first-law-v4}
\delta M = \frac{\kappa \,\delta A}{8\pi G_5}
+ {\cal T}_2\, \delta L
+ \left(-{\cal V} + \frac{{\cal T}_2 \, L}{2P} \right) \delta P
+ M \frac{\delta\Omega_1}{\Omega_1} \ .$$ Each term ${\cal T} \delta L$ in (\[first-law-v2\])$-$(\[first-law-v4\]) is the work needed to vary the RS1 interbrane distance (with different quantities held fixed), analogous to the work terms in the first law in the case of a compact dimension without branes [@harmark-obers-def-tension; @harmark-obers-first-law; @kol-sorkin-piran]. Our gravitational tensions (\[tensions-def\]) are easily verified to be positive for the asymptotically RS1 black string [@black-string]. We would also expect our gravitational tensions to be positive due the black hole’s attraction to its images in the covering space, which has been shown [@Archimedes-3-tension; @tension-interpretation; @positive-tension] in the case of a compact dimension without branes.
Since each version of the first law reparametrizes the geometry, in (\[first-law-v3\]) and (\[first-law-v4\]) each thermodynamic volume $V_{\rm eff}= -{\cal V}+{\cal T}_i L/P$ differs from $-\cal V$. For $-{\cal V}<0$, this indicates that positive gravitational tension ${\cal T}_i$ opposes the black hole Archimedes effect, and the sign of $V_{\rm eff}$ depends on their relative strengths.
Reparametrizing the geometry also transforms the brane terms in the first law, but with the interesting property that the coefficients of $\delta\Omega_i/\Omega_i$ always add to $M$, in each version of the first law. A brane term in (\[first-law-v1\]) or (\[first-law-v2\]) shifts into both $V_{\rm eff}\, \delta P$ and ${\cal T}_i \,\delta L$ in (\[first-law-v3\]) and (\[first-law-v4\]), and this shift incorporates the brane’s orbifold symmetry into the gravitational tension, since ${\cal T}_i$ is due to the black hole’s attraction to its orbifold mirror images.
Variational principle \[vp-proof\]
==================================
The first law we proved in section \[sec:first-law\] includes the variations of the AdS curvature length, cosmological constant, brane tensions, and RS brane warp factors. This motivates the following variational principle, which we prove in this section.
[**Variational principle for asymptotically RS black holes:**]{} [ *Instantaneously static initial data that extremizes the mass $M$ is initial data for a static black hole, for variations that leave fixed the apparent horizon area $A$, the AdS curvature length $\ell$, cosmological constant $\Lambda$, brane tensions $\lambda_i$, and RS values (at spatial infinity) of the warp factors $\Omega_i$ on each brane.*]{}
Main proof \[main step\]
------------------------
Our proof of the variational principle proceeds in two steps. Here we perform the main step, which reduces the proof to two auxiliary boundary value problems. These boundary value problems are the topics of section \[BVPs\]. Our setup is general: it applies to a black hole localized on a brane, or isolated in the bulk (away from either brane), and also applies to the asymptotically RS black string [@black-string].
Our key assumptions will be the following. We assume our initial data $h_{ab}$ is instantaneously static. We also assume the variations $\delta h_{ab}$ extremize the mass ($\delta M=0$), while holding fixed the apparent horizon area $A$ and the remaining quantities ($\ell$, $\Lambda$, $\lambda_i$, $\Omega_i$).
Our proof closely follows the proof of the first law in section \[sec:first-law\]. The initial steps are the same, as we indicated after the result (\[delta\_H\_sigma-1b\]). Thus, for an instantaneously static geometry $h_{ab}$, we proceed exactly as in section \[sec:first-law\] through (\[delta\_H\_sigma-1b\]). Now taking $\delta\Lambda=0$ in (\[delta\_H\_sigma-1b\]) gives $$\label{vp-delta_H_sigma}
\int_{\Sigma} d^{D-1}x \,{\cal P}^{ab}\delta h_{ab}
+
\sum_B I_B
=0 \ .$$ In what follows, (\[vp-delta\_H\_sigma\]) will be our primary equation, where $$\label{vp-Pab-initial}
{\cal P}^{ab} = \frac{\sqrt{h}}{16\pi G_D}
\left(
{\cal R}^{ab}+ h^{ab}D_c D^c - D^a D^b
\right)N \ .$$ For instantaneously static initial data, the constraint ${\cal C}_0=0$ in (\[constraints\]) simplifies to ${\cal R}=2\Lambda$ and $\delta {\cal C}_a$ vanishes identically. The remaining linearized constraint ($\delta{\cal C}_0=0$) simplifies to $$\label{vp-linearized-constraint-1-red}
\left({\cal R}^{ab} + h^{ab}D^c D_c -D^a D^b\right)
\delta h_{ab} = 0 \ .$$
We now evaluate the boundary terms $I_B$ in (\[vp-delta\_H\_sigma\]). The boundaries $B$ are illustrated in Fig. \[initial-data-figure\]. In section \[sec:first-law\], we evaluated $I_{B_i}$ (at each brane) and $I_{B_\infty}$ (at spatial infinity), including the variations of quantities ($\ell$, $\Lambda$, $\lambda_i$, $\Omega_i$) held constant here by assumption. In this case, (\[da\]) and (\[text-I-infty-3\]) reduce to $$\label{surface-terms-values}
I_{B_i} = 0 \ , \quad
I_{B_\infty}= -\delta M \ .$$ Additionally, we have $\delta M=0$, by our assumption of a mass extremum, so $I_{B_\infty}=0$. At the apparent horizon $B_H$, we use an alternate form to that given in (\[surface-terms-2\]), $$\label{surface-terms}
I_{B_H} = \int d^{D-2}x \sqrt{\sigma}
{\cal A}^{bcd} \left[(D_bN)\delta h_{cd}
- ND_b\delta h_{cd} \right] \ ,$$ where $$16\pi G_D \, {\cal A}^{bcd}= n_a(h^{ac}h^{bd} -h^{ab}h^{cd}) \ .$$ The boundary condition on the lapse is $N=0$, whence $D_aN = -f n_a$, where $f^2 = (D^bN)(D_bN)$. Then (\[surface-terms\]) is $$I_{B_H}=
\frac{1}{8\pi G_D} \int d^{D-2}x \, f\,
\delta\sqrt{\sigma} \ ,$$ using $\sigma_{ab}=h_{ab}- n_a n_b$ and $
\delta \sqrt{\sigma}
= \sqrt{\sigma}\sigma^{ab}\delta \sigma_{ab}/2
$. For convenience, we now choose to set $I_{B_H}=0$ using the following gauge transformation, $$\label{horizon gauge}
\delta\sigma_{ab} \rightarrow \delta\sigma_{ab} + 2{\cal D}_{(a}\xi_{b)} \ ,
\quad
\sigma^{ab}\delta \sigma_{ab}
\rightarrow
0 \ ,$$ where ${\cal D}_a$ is the covariant derivative associated with $\sigma_{ab}$. If we let $\xi_a={\cal D}_aF$, then $
\sigma^{ab}\delta \sigma_{ab}
\rightarrow
0
$ requires $$\label{vp-throat-gauge}
-\sqrt{\sigma}\,{\cal D}^a{\cal D}_a F = \delta \sqrt{\sigma} \ .$$ Note the apparent horizon is a closed surface (this is most clearly seen in the covering space, if the apparent horizon intersects a brane). A solution $F$ to (\[vp-throat-gauge\]) on a closed surface is well known to exist if and only if the surface integral of the right-hand side of (\[vp-throat-gauge\]) vanishes. This integral is simply $\delta A$, which indeed vanishes since we hold $A$ constant. Thus a solution $\xi_a$ exists to achieve (\[horizon gauge\]), and we henceforth set $I_{B_H}=0$. Since $I_{B_i}=I_{B_\infty}=0$ and we set $I_{B_H}=0$, our primary equation (\[vp-delta\_H\_sigma\]) simplifies to $$\label{vp-no-go-converse-1}
\int_{\Sigma} d^{D-1}x \, {\cal P}^{ab}\delta h_{ab} =0
\ .$$
Our goal is to conclude that the initial geometry $h_{ab}$ evolves to a static spacetime. The well known condition for this is that ${\cal P}^{ab}=0$ on $\Sigma$. We cannot, however, immediately conclude that ${\cal P}^{ab}=0$ from (\[vp-no-go-converse-1\]), because not all of the variations $\delta h_{ab}$ are arbitrary: the linearized constraint (\[vp-linearized-constraint-1-red\]) removes one degree of freedom, which can be taken as $h^{ab}\delta h_{ab}$ or as the variation $\delta h$ of the determinant $h$. These are related by $
h^{ab} \delta h_{ab}
=
\delta h/h$. As an identity, we may decompose $\delta h_{ab}$ into a trace-free ($TF$) part and a part proportional to $\delta h$: $$\label{variations-decomp}
\delta h_{ab} =
(\delta h_{ab})^{TF}
+
\frac{1}{D-1}\left(\frac{\delta h}{h}\right)h_{ab} \ .$$ Using (\[variations-decomp\]), our primary equation (\[vp-no-go-converse-1\]) then becomes $$\label{vp-no-go-converse-2}
\int_{\Sigma} d^{D-1}x
\left[ ({\cal P}^{ab})^{TF}(\delta h_{ab})^{TF}
+
\frac{{\cal P}\, \delta h}{(D-1)h}
\right]
=0 \ ,$$ where $$\begin{aligned}
\label{P-decomp}
{\cal P}_{ab} &=&
( {\cal P}_{ab})^{TF}
+
\frac{\cal P}{D-1} h_{ab} \ ,
\\
{\cal P} &=& h_{ab}{\cal P}^{ab} \ .\end{aligned}$$ The arbitrary variations are $(\delta h_{ab})^{TF}$, subject to smoothness at the apparent horizon, $(\delta h_{ab})^{TF} \rightarrow 0$ at $B_\infty$, and boundary conditions at the branes that we will specify in the next section. As a completeness check, the arbitrary variations $(\delta h_{ab})^{TF}$ alone should determine the dependent quantity $\delta h$, which we verify below by showing the linearized constraint (\[vp-linearized-constraint-1-red\]) is a well posed boundary value problem for $\delta h$.
Our proof then reduces to showing ${\cal P}=0$, which allows us to conclude from (\[vp-no-go-converse-2\]) that $({\cal P}^{ab})^{TF}=0$, since $(\delta h_{ab})^{TF}$ are arbitrary variations. It then follows from (\[P-decomp\]) that ${\cal P}^{ab}=0$, which is the desired result. The statement ${\cal P}=0$ is a boundary value problem for $N$ that we demonstrate is solvable in the following section, which completes our proof of the variational principle.
Auxiliary boundary value problems \[BVPs\]
------------------------------------------
The boundary value problem for $N$ is
\[N-BVP\] $$\begin{aligned}
\label{vp-N-eom}
D^a D_a N - \frac{(D-1)}{\ell^2} N
&=& 0 \ ,
\\*
\label{vp-N-bc-brane}
\left(n^a D_a N - \frac{\varepsilon}{\ell}\, N \right)\Big|_{B_i}
&=& 0 \ ,
\\*
\label{vp-N-bc-throat}
N \, \Big|_{B_H} &=& 0 \ ,
\\*
\label{vp-N-bc-infty}
N\, \Big|_{B_\infty} &\rightarrow& \Omega \ .\end{aligned}$$
Here, $\Omega$ is the warp factor of the asymptotic RS solution (\[g\_RS\]) and $\varepsilon=\pm 1$ is the sign of each brane tension. The result (\[vp-N-eom\]) follows from setting ${\cal P}= h_{ab}{\cal P}^{ab}=0$, using (\[vp-Pab-initial\]) and ${\cal R}=2\Lambda$. The boundary conditions (\[vp-N-bc-throat\]) and (\[vp-N-bc-infty\]) are straightforward. Our main concern is the brane boundary condition (\[vp-N-bc-brane\]), which results from using $$2K_{ab} = n^c \partial_c \gamma_{ab} + \gamma_{ac} \partial_b n^c + \gamma_{bc} \partial_a n^c \ .$$ Now $\gamma_{tt}=-N^2$ and $\gamma_{ta}=0$ gives $
2K_{tt} = n^c \partial_c (-N^2)
$, and the Israel condition $K_{tt} = (\varepsilon/\ell)\gamma_{tt}$ then gives (\[vp-N-bc-brane\]).
As shown in [@inverse-positivity], the following approach can put a Robin boundary condition (\[vp-N-bc-brane\]) in a standard form while keeping its associated elliptic equation (\[vp-N-eom\]) in a divergence form. Let $w_a$ be any vector field and define ${\cal W}_aN = (D_a - w_a)N$. Then (\[vp-N-eom\]) and (\[vp-N-bc-brane\]) become $$\label{vp-N-eom-2}
D^a {\cal W}_a N
+ w^a D_a N
+ \left[ D_a w^a - \frac{(D-1)}{\ell^2}\right] N
=0$$ and $$\label{vp-N-bc-brane-2}
\left[n^a {\cal W}_a N+ \left(n^a w_a - \frac{\varepsilon}{\ell}\right)N \right]\Big|_{B_i}
= 0 \ .$$ As in [@inverse-positivity], we now choose $w_a$ so $(n^aw_a - \varepsilon/\ell) \ge 0$ at $B_i$, which is the usual prerequisite for applying an existence theorem to a boundary value problem of the form (\[vp-N-eom-2\])(\[vp-N-bc-brane-2\]). For example, we choose $w_a = - \widetilde n_a/\ell$, where $\widetilde n_a$ is any vector field, pointing from $B_1$ to $B_2$, that interpolates from the inward unit normal ($-n_a$) of $B_1$ to the outward unit normal $n_a$ of $B_2$. Then $n^a\widetilde n_a = -\varepsilon$ at each brane $B_i$, and $w_a = - \widetilde n_a/\ell$ gives $(n^aw_a - \varepsilon/\ell) = 0$ in (\[vp-N-bc-brane-2\]). With the brane boundary conditions in standard form, and the remaining standard (Dirichlet) boundary conditions, (\[vp-N-bc-throat\]) and (\[vp-N-bc-infty\]), we then readily infer that the boundary value problem (\[N-BVP\]) for $N$ is solvable.
We now turn to the boundary value problem for $\delta h$, which we will state in terms of a scalar quantity $(\delta h/h)$,
\[dh BVP\] $$\begin{aligned}
\label{vp-dh-eom}
D^a D_a (\delta h/h) - \frac{(D-1)}{\ell^2} \, (\delta h/h)
&=& f_{\Sigma} \ ,
\\*
\label{vp-dh-bc-brane}
\left[n^a D_a (\delta h/h) - \frac{\varepsilon}{\ell} \, (\delta h/h)
\right]\Big|_{B_i}
&=& f_{i} \ ,
\\*
\label{vp-dh-bc-throat}
n^a D_a (\delta h/h) \Big|_{B_H} &=& f_H \ ,
\\*
\label{vp-dh-bc-infty}
(\delta h/h)\ \Big|_{B_\infty} &\rightarrow& 0 \ .\end{aligned}$$
As above, $\varepsilon =\pm 1$ is the sign of each brane tension. The result (\[vp-dh-eom\]) follows from substituting (\[variations-decomp\]) into the linearized constraint (\[vp-linearized-constraint-1-red\]) with ${\cal R}=2\Lambda$. We will give the source terms and derive the boundary conditions below.
The key point is that (\[dh BVP\]) is a well posed boundary value problem. The elliptic equation (\[vp-dh-eom\]) and the boundary conditions (\[vp-dh-bc-brane\]) are similar in form to (\[vp-N-eom\]) and (\[vp-N-bc-brane\]) in the previous boundary value problem (\[N-BVP\]). The remaining boundary conditions, (\[vp-dh-bc-throat\]) and (\[vp-dh-bc-infty\]), are well known types: Neumann and Dirichlet, respectively.
In the remainder of this section, we provide the details of the source terms and the boundary conditions in (\[dh BVP\]). The source terms in (\[dh BVP\]) are $$\begin{aligned}
f_{\Sigma} &=&
\frac{D-1}{D-2} \left[D^a D^b(\delta h_{ab})^{TF}
-
{\cal R}^{ab} (\delta h_{ab})^{TF}
\right]\ ,
\\*
-f_{i} &=& \frac{D-1}{D-2}
\left[
\sigma^{ab} n^c D_c (\delta
h_{ab})^{TF} + \frac{\varepsilon D}{\ell}n^a n^b (\delta h_{ab})^{TF}
\right]
\ ,
\\*
f_H &=&
\frac{1}{D-2}
\left[
2k^{ab} (\delta h_{ab})^{TF} - \sigma^{ab} n^c D_c (\delta h_{ab})^{TF}
\right] \ .\end{aligned}$$ The boundary conditions on $\delta h$ are given by varying those on $h_{ab}$, which at the apparent horizon and the branes involve the extrinsic curvature $k_{ab}=\sigma_a{}^c D_c n_b$, $$\label{k-boundary-conditions}
k \Big|_{B_H} = 0
\ , \quad
k_{ab} \Big|_{B_i} = \frac{\varepsilon}{\ell} \, \sigma_{ab} \ .$$ By varying these, we obtain $$\label{unconstrained-var}
\delta k \Big|_{B_H} = 0
\ , \quad
\delta k \Big|_{B_i} = 0
\ , \quad
\delta k_{ab} \Big|_{B_i} = \frac{\varepsilon}{\ell} \, \delta \sigma_{ab} \ .$$ To evaluate these, we use the general results
$$\begin{aligned}
\label{unconstrained-var-dk_ab}
2\delta k_{ab} &=&
\left(n^c n^d\delta h_{cd}\right)k_{ab}
-
\sigma_a{}^c \sigma_b{}^d n^f
J_{cdf} \ ,
\\
\label{unconstrained-var-dk}
-2\delta k &=& 2k^{ab} \delta \sigma_{ab} - k\,n^a n^b \delta h_{ab}
+ \sigma^{ab} n^c J_{abc} \ ,
\\
J_{abc} &=& D_{a}\delta h_{bc} + D_{b}\delta h_{ac} - D_c\delta h_{ab} \ .\end{aligned}$$
The boundary conditions at the branes (\[vp-dh-bc-brane\]) and the apparent horizon (\[vp-dh-bc-throat\]) result from evaluating $\delta k=0$ using (\[variations-decomp\]), (\[k-boundary-conditions\]), and (\[unconstrained-var-dk\]). The last relation in (\[unconstrained-var\]) expresses brane boundary conditions for $(\delta h_{ab})^{TF}$, since it reduces to a form independent of $\delta h$ after substituting (\[RS-K\]), (\[variations-decomp\]), (\[vp-dh-bc-brane\]), and (\[unconstrained-var-dk\_ab\]).
An application of the variational principle \[black string application\]
------------------------------------------------------------------------
Here we demonstrate the utility of the variational principle, by applying it to a trial solution and reproducing the static asymptotically RS black string [@black-string], which is the only known exact solution for an asymptotically RS black object in 5-dimensional spacetime. We first specify a trial geometry for an initially static a black string. After evaluating the apparent horizon area $A$ and mass $M$, we then apply the variational principle.
A black string is a set of lower dimensional black holes stacked in an extra dimension $Z$, which is how we will construct the trial geometry. We take $$\label{ansatz}
ds^2 =
\Omega(Z)^2
\left[
\Psi^4({\bf x},Z) \, d{\bf x}^2+ dZ^2
\right] \ ,$$ where $\Omega=\ell/Z$ and the branes are the surfaces $Z=Z_1$ and $Z=Z_2$. We will take $$\label{Psi-1}
\Psi = 1 + \frac{\rho_0}{2}\left(
\frac{1}{|{\bf x}+{\bf x}_0|} + \frac{1}{|{\bf x}-{\bf x}_0|}
\right) \ .$$ Here $\rho_0 >0 $ is a constant and ${\bf x}_0= (0,0, d)$, using Cartesian coordinates ${\bf x} =(x, y, z)$ with origin at ${\bf x}=0$. We choose a function $d(Z)$ as follows. The constraint, ${\cal R}= -12/\ell^2$, after linearizing in $d$ and its derivatives, has the solution $d(Z) =d_0 + c_0 Z^4$, where $c_0$ and $d_0$ are constants. For the case $c_0=0$, (\[Psi-1\]) is an exact solution and the RS1 limit ($\Omega_2\rightarrow 0$) is easily taken. We will work in RS2 and take $c_0$ as a small nonzero parameter.
On each slice $Z$=constant, we now transform to spherical coordinates centered at ${\bf x}=0$, with $z=\rho \cos\theta$, and expand $\Psi$ in Legendre polynomials $P_{k}(\cos\theta)$ for $\rho > d$, $$\label{Psi-2}
\Psi =1 + \frac{\rho_0}{\rho} +
\frac{\rho_0}{\rho} \sum_{j=1}^\infty \left(\frac{d}{\rho}\right)^{2j} P_{2j}(\cos\theta) \ .$$ On each slice $Z$=constant, we will take $\rho_0 \gg d$, which describes a 3-dimensional black hole [@Brill-Lindquist], and $d$ parametrizes the 2-dimensional apparent horizon’s distortion from the sphere $\rho=\rho_0$. The full 3-dimensional apparent horizon (the union of the 2-dimensional apparent horizons) therefore describes a distorted black string.
On each slice $Z$=constant, as in [@Brill-Lindquist], the surface $\rho(\theta)$ of the 2-dimensional apparent horizon can be found as the sum of Legendre polynomials that minimizes the area $A_2$, $$A_2 = 2\pi
\int_0^{\pi} d\theta\, \Psi^4 \rho \sqrt{\rho^2 + \left( \frac{d\rho}{d\theta}\right)^2} \ .$$ To lowest order in $(d/\rho_0) \ll 1$, we find the apparent horizon on each slice $Z$=constant is $$\label{bs-AH}
\rho(\theta) = \rho_0
\left[ 1 + \frac{5}{7}\left(\frac{d}{\rho_0} \right)^2 P_2(\cos\theta) \right] \ .$$ This agrees with the numerical results of [@Brill-Lindquist], and gives, to lowest order in $(d/\rho_0$), $$\label{bs-area-2D}
A_2(Z) =
64\pi \rho_0^2
\left[
1 - \frac{5 }{7}\left(\frac{d}{\rho_0}\right)^4
\right] \ .$$ Integrating in $Z$ gives the 3-dimensional area of the full apparent horizon in the black string geometry, $$\label{bs-area}
A = \int_{Z_1}^{Z_2} dZ\, \Omega^3
A_2
=
32\pi w
\left(
\rho_0^2 - \frac{5Q}{7\rho_0^2}
\right) \ ,$$ where, with $\Omega_i$ the warp factor at each brane, $$w = \ell \left(\Omega_1^2 - \Omega_2^2\right)
\ ,
\quad
Q = d_0^4
+
\frac{8 \, c_0 \, d_0^3 \, \ell^4}{\Omega_1\Omega_2(\Omega_1+\Omega_2)} \ .$$ We defined the mass $M$ in (\[mass-geometric\]), which for (\[Psi-2\]) gives $$\label{bs-mass}
G_5 M =w \rho_0 \ .$$ To lowest order in $Q$, combining (\[bs-area\])(\[bs-mass\]) gives $$\label{bs mass function}
\left(\frac{G_5 M}{w}\right)^2 = \left(\frac{A}{32\pi w } \right)+ \frac{5}{7} \left(\frac{32\pi w }{A} \right) Q \ .$$ We now apply our variational principle: we extremize $M$ in (\[bs mass function\]) at fixed $A$, $\ell$, and $\Omega_i$. This yields the conditions $c_0=0$ and $d_0=0$, which we conclude describes a static black string. We can verify this directly, since $d=0$ in (\[Psi-2\]) gives $\Psi = 1+\rho_0/\rho$ and the apparent horizon (\[bs-AH\]) is located at $\rho=\rho_0$. This is indeed the initial geometry of the static black string [@black-string] in isotropic coordinates.
We can also deduce that the evolution of the distorted black string, with $d \neq 0$, will not be static, since each slice $Z$=constant is the initial geometry for an attracting two-body problem [@Brill-Lindquist], and the 2-dimensional apparent horizon considered above describes a black hole formed by, and surrounding, two closely separated smaller black holes (with a small minimal surface surrounding each point ${\bf x}=\pm{\bf x}_0$). From the perspective in each slice $Z$=constant, as in [@Brill-Lindquist], the two small interior black holes will coalesce as the initial data evolves, due to mutual gravitational attraction. This results in a time-dependent geometry on each slice $Z$=constant, and results in a time-dependent black string geometry in the bulk perspective.
Conclusion \[sec:conclusion\]
=============================
We have derived the first law for a static asymptotically RS black hole, whose mass $M$ is defined in (\[mass-geometric\]) and (\[mass\]). Four versions of this law are given in (\[first-law-v1\])(\[first-law-v4\]) for different choices of variations. In both RS1 and RS2, the general first law contains brane terms and a thermodynamic volume. In RS1, we can define both a thermodynamic volume and a gravitational tension, due to the presence of both a cosmological constant and a compact interbrane distance. This differs from the first law in previously studied spacetimes (with either a cosmological constant or a compact dimension), where the analogs of our thermodynamic volume and gravitational tension are isolated from each other, appearing in the separate first laws of separate spacetimes.
The variational principle we developed in this paper states that for an asymptotically RS black hole initially at rest, initial data that extremizes the mass yields a static black hole, for variations at fixed values of the apparent horizon area and the remaining physical variables in the first law ($L$, $\Omega_i$, $\ell$, $\Lambda$, $\lambda_i$). It would be interesting to investigate the consequences of holding fewer variables fixed. An example of this in four-dimensional spacetime is Hawking’s proof [@hawking] that the static (Schwarzschild) black hole is an extremum of mass at fixed apparent horizon area but arbitrary angular momentum.
Our example application of the variational principle to a trial solution serves as a prelude to the approach we will take in a sequel paper [@paper-3-extrema]. In [@paper-3-extrema], we will conclude that solutions exist for small static black holes in RS2, both on and off the brane, as special members of a general family of initially static black holes. This family of black hole initial data will also indicate that a small black hole on an orbifold-symmetric brane in RS2 is stable against leaving the brane, which generalizes to other models with an orbifold-symmetric brane. If we inhabit such a brane, then small black holes, if produced in high energy collider experiments on the brane, could be studied directly (instead of leaving behind a signature of missing energy), which is an important result for future experiments at the LHC.
We thank D. Kastor and J. Traschen for useful discussions at KITP. This research was supported in part by the National Science Foundation under Grant No. NSF PHY11-25915.
Mass counterterm $M_{ct}$ {#appendix-M-ct}
=========================
In this appendix, we derive the mass counterterm $M_{ct}$ from (\[M-ct-formula\]) and (\[S-ct-hybrid\]), and thereby prove the mass formula (\[mass-geometric\]). We begin with the variation of (\[S-ct-hybrid\]), $$\label{d S_ct}
\delta S_{ct}
= - \int
d^4x
\frac{\sqrt{-\gamma} }{ 8\sqrt{2}\pi G_5}
\left(\sqrt{\hat{\cal R}}
\gamma^{ab} \delta\gamma_{ab}
+ \frac{\delta \hat{\cal R}}{\sqrt{\hat{\cal R}}} \right)
\ .$$ The standard variation of the Ricci scalar is $$\delta{\hat{\cal R}} = - {\hat{\cal R}}^{ab}\delta\hat\sigma_{ab} + \hat{d}_a v^a$$ where $v^a = 2 \hat\sigma^{b[a}\hat{d}^{c]}\delta\hat\sigma_{bc}$ and $\hat{d}_a$ is the covariant derivative associated with $\hat\sigma_{ab}$. This gives $$\label{A3}
\delta S_{ct} = \int d^4x\, {\cal S}^{ab} \delta\gamma_{ab} - (8\sqrt{2}\pi G_5)I \ ,$$ where ${\cal S}^{ab}$ is given below and, with $J=\sqrt{-\gamma_{tt}\gamma_{ZZ}/\hat{\cal R}}$, $$I
=
\int dt\,dZ\,
\int d^2x \, \sqrt{\hat\sigma}\,
\left[
\hat {d}_a (J v^a) -v^a \hat {d}_a J
\right]
\ .$$ We conclude that $I=0$, as follows. The first term is a total divergence, but the 2-sphere has no boundary. Also, $\hat {d}_a J=0$ since we take $J$ independent of the angular coordinates. Then (\[A3\]) gives $${\cal S}^{ab} = \frac{\delta S_{ct}}{\delta\gamma_{ab}} \ .$$ In (\[d S\_ct\]) we now use $$\gamma^{ab}\delta\gamma_{ab} = \gamma^{tt}\delta\gamma_{tt}
+ \hat\sigma^{ab}\delta\hat\sigma_{ab}
+ \sigma^{ZZ}\delta\sigma_{ZZ}$$ which gives
$$\begin{aligned}
\label{fderiv-S-ct-1}
\frac{\delta S_{ct}}{\delta\gamma_{tt}} &=&
- \, \frac{1}{8\sqrt{2}\pi G_5}\, \sqrt{-\gamma\hat{\cal R}}
\, \gamma^{tt} \ , \
\\
\label{fderiv-S-ct-2} \frac{\delta S_{ct}}{\delta\hat\sigma_{ab}}
&=&
- \, \frac{1}{8\sqrt{2}\pi G_5}\,
\sqrt{-\gamma \hat{\cal R}} \left(
\hat\sigma^{ab} -
\frac{\hat{\cal R}^{ab}}{\hat{\cal R}}
\right) \ , \qquad
\\
\label{fderiv-S-ct-3} \frac{\delta S_{ct}}{\delta\sigma_{ZZ}} &=&
- \, \frac{1}{8\sqrt{2}\pi G_5}\,
\sqrt{-\gamma \hat{\cal R}} \, \sigma^{ZZ} \ .\end{aligned}$$
The mass counterterm from (\[M-ct-formula\]) and (\[fderiv-S-ct-1\]) is then $$\label{M-ct}
M_{ct}
=
\frac{\sqrt{2}}{8\pi G_5}
\int d^3 x \, \sqrt{-\gamma \hat{\cal R}} \ .$$ Combining this with (\[M-bare-formula\]) now gives the mass formula (\[mass-geometric\]).
Boundary term $I_B$ at infinity {#appendix-first-law}
===============================
Here we evaluate the term $I_{B_\infty}$ given by (\[surface-terms-2\]) at the boundary $\rho \rightarrow \infty$. Throughout this appendix, $\simeq$ denotes evaluating at leading order and neglecting terms of higher order in $1/\rho$. We will relate $I_{B_\infty}$ to the mass variation $\delta M$ and additional terms. The mass is a sum of two terms, $M=\widetilde M+M_{ct}$, whose individual variations are $$\label{dM-tilde}
\delta \widetilde M
= -\, \frac{1}{8\pi G_5}\int_{B_\infty} d^{3}x\,\left[N
\delta\left(\sqrt{\sigma} \,k\right) + \sqrt{\sigma}\, k\, \delta
N \right]$$ and $$\label{dM-ct-3}
\delta M_{ct} =
\int_{B_\infty} d^3 x \, \left[\frac{\delta M_{ct}}{\delta\sigma_{ab}} \, \delta\sigma_{ab}
+ \frac{ \sqrt{2\sigma \hat{\cal R}}}{8\pi G_5}
\, \delta N \right] \ .$$ Note $\delta M_{ct}/\delta \sigma_{ab} =
-\delta S_{ct}/\delta \sigma_{ab}$ since $\int dt\,M_{ct}=-S_{ct}$ by (\[M-ct\]) and (\[S-ct-hybrid\]). From (\[fderiv-S-ct-2\]) and (\[fderiv-S-ct-3\]), we find $$\label{dM-ct-X}
\frac{\delta M_{ct}}{\delta\sigma_{ab}} -
\frac{N}{2} \sqrt{\sigma} s^{ab} = X^{ab}$$ at large $\rho$, where the quantities $X^{ab}$ are given below. Using (\[dM-tilde\])(\[dM-ct-X\]), we rewrite the boundary term (\[surface-terms-2\]) as $$\begin{aligned}
\nonumber I_{B_\infty} &=& - \delta M
+ \int_{B_\infty} d^{3}x \, X^{ab}\delta\sigma_{ab}
\\
& &
- \frac{1}{8\pi G_5} \int_{B_\infty} d^{3}x \, \sqrt{\sigma}
\left(k -\sqrt{2\hat{\cal R}}\right)\delta N \ . \qquad
\label{I-infty-2}\end{aligned}$$ For the metric asymptotics (\[ansatz-1\]), the quantities $X^{ab}$ are
$$\begin{aligned}
16\pi G_5\, X^{\chi\chi} &\simeq& \frac{\Omega \sin\chi}{\rho^2}
\left(a_t+a_\rho+a_Z\right) \ ,
\\
16\pi G_5\, X^{\phi\phi} &\simeq& \frac{\Omega }{\rho^2\sin\chi}
\left(a_t+a_\rho+a_Z\right) \ ,
\\
\label{X_ab}
16\pi G_5\, X^{ZZ} &\simeq& \Omega\sin\chi
\left( a_t +2a_\rho\right) \ .\end{aligned}$$
Here $\chi$ is the polar angle on the 2-sphere with radius $\rho$. We now proceed to evaluate (\[I-infty-2\]). We begin with three convenient variables ($\ell$, $Z_1$, $Z_2$) and then express results in terms of three physical variables ($\ell$, $\Omega_1$, $\Omega_2$). We first consider the variation $\delta\ell$ at fixed ($Z_1$, $Z_2$). At large $\rho$, we have $\delta g_{ab} \simeq 2(\delta\ell/\ell) g_{ab}$. Then $$\int_{B_\infty} d^{3}x\, X^{ab}\delta\sigma_{ab} =
{\cal F}_\infty \delta \ell$$ where $$\label{F-infty-appendix}
{\cal F}_\infty =
\frac{1}{2G_5\ell}\int_{Z_1}^{Z_2} dZ\, \Omega^3\left( a_t + 2a_\rho \right) \ .$$ This is entirely due to $\delta \sigma_{ZZ}$ since the integral contributions from $\delta\sigma_{\phi\phi}$ and $\delta\sigma_{\chi\chi}$ vanish by the identity (\[identity1\]). This identity can also be used to rewrite ${\cal F}_\infty$ in the form given in (\[F-infty\]). The last line in (\[I-infty-2\]) yields $M \delta \ell/\ell$. Hence (\[I-infty-2\]) yields, at fixed ($Z_1, Z_2)$, $$\label{I-fixed Z}
\left(I_{B_\infty}\right)_{Z_1 Z_2} = - \delta M +
\left({\cal F}_\infty +\frac{M}{\ell}\right) \delta \ell
\ .$$ We now consider variations ($\delta Z_1$, $\delta Z_2$) at fixed $\ell$. We then perform a coordinate transformation $$Z \rightarrow \widetilde Z = (1-\epsilon)Z - \zeta$$ such that the branes again reside at $Z=Z_i$. The required transformation is $$\epsilon = \frac{\delta Z_2-\delta Z_1}{Z_2-Z_1}
\quad , \quad
\zeta = \frac{Z_2\,\delta Z_1- Z_1\,\delta Z_2}{Z_2-Z_1} \ .$$ At large $\rho$, the resulting metric perturbation is $$\delta g_{ab}
\simeq
2\Omega^2\left[
\epsilon\left(\delta^Z_a \delta^Z_b - \eta_{ab}\right)
-\frac{\zeta}{\ell} \,\Omega\eta_{ab}\right]
\ ,$$ where $\eta_{ab}$ is the 5-dimensional Minkowski metric. Then (\[I-infty-2\]) becomes, at fixed $\ell$, $$\label{I-fixed-ell}
\left(I_{B_\infty}\right)_\ell = - \delta M
- \epsilon M - \zeta\,{\cal I}
\ ,$$ where the integral $\cal I$ is $${\cal I} =
\frac{3}{2G_5\ell}
\int_{Z_1}^{Z_2} dZ\, \Omega^4 \left(a_t + 2a_\rho + a_Z\right)
\ .$$ For the case when all three quantities ($\ell$, $Z_1$, $Z_2$) are varied, we combine (\[I-fixed Z\]) and (\[I-fixed-ell\]) to obtain $$\label{I-infty-3}
I_{B_\infty} = - \delta M +
\left({\cal F}_\infty +\frac{M}{\ell}\right) \delta \ell
- \epsilon M - \zeta\,{\cal I} \ .$$ We can evaluate the integral $\cal I$ using (\[a\_t\]) and the identity (\[int-field\]). We then express the result in terms of $M$ and $\cal Q$ using (\[M and mu0\]) and (\[a-M-Q\]). This gives $$\label{integral I}
{\cal I} \ell =
M\left( \frac{\Omega_1{}^3 -\Omega_1{}^3 }{\Omega_1{}^2-\Omega_2{}^2}\right)
-
\frac{3 {\cal Q}}{2G_5}
\left(
\frac{\Omega_1{}^2\Omega_2{}^2}{\Omega_1{}^{2}-\Omega_2{}^{2}}
\right) \ .$$ We also express $\epsilon$ and $\zeta$ in terms of three physical variables ($\ell$, $\Omega_1$, $\Omega_2$) using $$\label{dZ}
\delta Z_i = \frac{1}{\Omega_i}
\left( \delta\ell
-
\frac{\ell}{\Omega_i} \, \delta\Omega_i
\right) \ .$$ Using (\[integral I\]) and (\[dZ\]) in (\[I-infty-3\]) then yields the result for $ I_{B_\infty}$ given in (\[text-I-infty-3\]).
Gauge invariance {#gauge invariance}
================
It is important to confirm that our quantities ($M$, $\cal Q$, $\cal V$, $L$, ${\cal T}_0$, ${\cal T}_i$, ${\cal U}_i$) are gauge invariant at infinity. As one can verify, these quantities are invariant under the following metric transformation that leaves the branes fixed, $$a_\nu \rightarrow a_\nu - \frac{\Omega^\prime}{\Omega}w - \delta^Z_\nu w^\prime
\ \; , \ \;
w(Z_1)=w(Z_2)=0
\ ,$$ with $^\prime=d/dZ$. This is generated by the coordinate transformation $x^a \rightarrow x^a + \varepsilon^a$, where to leading order in $1/\rho$, $$\varepsilon^Z = \frac{w}{\rho}
\quad , \ \quad
\varepsilon^\rho = \frac{W}{\rho^2}
\quad , \ \quad
w=W^\prime \ .$$ In particular, we consider the quantity ${\cal F}$, and write $${\cal F} = {\cal F}_\Sigma + {\cal F}_\infty$$ where $ {\cal F}_\Sigma$ is the sum of bulk and brane terms in (\[F-def\]), $${\cal F}_\Sigma \,\delta\ell \ \equiv \
\frac{\delta\Lambda}{8\pi G_5}\int_{\Sigma} d^{4}x\, N
\sqrt{h}
\ + \ I_{B_1} \ +\ I_{B_2} \ .$$ We note that $\cal F$ is gauge invariant, but neither ${\cal F}_\Sigma$ nor ${\cal F}_\infty$ is separately invariant, since they transform as $${\cal F}_\Sigma \rightarrow {\cal F}_\Sigma - \varphi
\quad , \quad
{\cal F}_\infty \rightarrow {\cal F}_\infty + \varphi \ ,$$ where $$\varphi =
\frac{3}{2G_5\ell^2}\int_{Z_1}^{Z_2} dZ \,\Omega^4 w \ .$$
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Lower bounds for the average probability of error of estimating a hidden variable $X$ given an observation of a correlated random variable $Y$, and Fano’s inequality in particular, play a central role in information theory. In this paper, we present a lower bound for the average estimation error based on the marginal distribution of $X$ and the principal inertias of the joint distribution matrix of $X$ and $Y$. Furthermore, we discuss an information measure based on the sum of the largest principal inertias, called $k$-correlation, which generalizes maximal correlation. We show that $k$-correlation satisfies the Data Processing Inequality and is convex in the conditional distribution of $Y$ given $X$. Finally, we investigate how to answer a fundamental question in inference and privacy: given an observation $Y$, can we estimate a function $f(X)$ of the hidden random variable $X$ with an average error below a certain threshold? We provide a general method for answering this question using an approach based on rate-distortion theory.'
author:
- |
Flávio P. Calmon, Mayank Varia, Muriel Médard,\
Mark M. Christiansen, Ken R. Duffy, Stefano Tessaro [^1]
bibliography:
- 'references.bib'
title: Bounds on inference
---
Concluding remarks {#sec:conclusion}
==================
We illustrated in this paper how the principal inertia-decomposition of the joint distribution matrix can be applied to derive useful bounds for the average estimation error. The principal inertias are a more refined metric of the correlation between $X$ and $Y$ than, say, mutual information. Furthermore, the principal inertia components can be used in metrics, such as $k$-correlation, that share several properties with mutual information (e.g. convexity).
Furthermore, we also introduced a general method for bounding the average estimation error of functions of a hidden random variable. This method depends on the Schur-concavity of a lower bound for the error-rate function. We proved that the $e_{\mathcal{I}}({p_X},\theta)$ itself is Schur-concave whenever the measure of information is concave in ${p_X}$. It remains to be shown if $e_{\mathcal{I}}({p_X},\theta)$ is Schur-concave for more general measures of information (such as $k$-correlation), and finding the necessary and sufficient conditions for Schur-concavity would be of both theoretical and practical interest.
Finally, the creation of bounds for $P_e$ and $P_{e,M}$ given constraints on different metrics of information is a promising avenue of research. Most information-theoretic lower bounds for the average estimation error are based on mutual information. However, in statistics, a wide range of metrics are used to estimate the information between an observed and a hidden variable. Relating such metrics with the fundamental limits of inference is relevant for practical applications in both security and machine learning.
Acknowledgement {#acknowledgement .unnumbered}
===============
The authors would like to thank Prof. Shafi Goldwasser and Prof. Yury Polyanskiy for the insightful discussions and suggestions throughout the course of this work.
[^1]: This work is sponsored by the Intelligence Advanced Research Projects Activity under Air Force Contract FA8721-05-C-0002. Opinions, interpretations, conclusions and recommendations are those of the authors and are not necessarily endorsed by the United States Government. M.C. and K.D. are supported by Science Foundation Ireland Grant No. 11/PI/1177.
F. P. Calmon and M. Médard are with the Research Laboratory of Electronics at the Massachusetts Institute of Technology, Cambridge, MA (e-mail: [email protected]; [email protected]). M. Varia is with MIT Lincoln Laboratory, Lexington, MA (e-mail: [email protected]). M. M. Christiansen and K. R. Duffy are with the Hamilton Institute at the National University of Ireland, Maynooth, Ireland (e-mail: [email protected]; [email protected]). S. Tessaro is with the Computer Science and Artificial Intelligence Laboratory at the Massachusetts Institute of Technology, Cambridge, MA (e-mail: [email protected]).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We deduce a nonlinear and inhomogeneous Fokker-Planck equation within a generalized Stratonovich, or stochastic $\alpha$-, prescription ($\alpha=0$, $1/2$ and $1$ respectively correspond to the Itô, Stratonovich and anti-Itô prescriptions). We obtain its stationary state $p_{st}(x)$ for a class of constitutive relations between drift and diffusion and show that it has a $q$-exponential form, $p_{st}(x) = N_q[1 - (1-q)\beta V(x)]^{1/(1-q)}$, with an index $q$ which does [*not*]{} depend on $\alpha$ in the presence of any nonvanishing nonlinearity. This is in contrast with the linear case, for which the index $q$ is $\alpha$-dependent.'
author:
- Zochil González Arenas
- 'Daniel G. Barci'
- Constantino Tsallis
title: 'Nonlinear inhomogeneous Fokker-Planck equation within a generalized Stratonovich prescription'
---
Introduction
============
The nonlinear Fokker-Planck (FP) equation has been largely used to study a wide class of physical systems which exhibit anomalous diffusion [@Plastino95; @Tsallis96; @Frankbook; @Tsallis2002; @CuradoNobreetal]. A particular feature of this equation is that its stationary solutions are probability distributions obeying nonextensive statistical mechanics [@Tsallis88; @books].
From a mesoscopic point of view, nonlinear FP equations are related with a class of Langevin equations with multiplicative noise [@Lisa1]. In these processes, the inhomogeneity of the diffusion function is proportional to a function of the probability density itself. Therefore, the computation of stochastic trajectories turn out to be very cumbersome. Indeed, one needs, for each noise realization, to know the complete time evolution of the probability density, making the problem a self-consistent one, very hard to deal with. Moreover, it is well known that, to correctly define the stochastic multiplicative process it is necessary to fix a prescription to perform the Wiener integrals. The stochastic evolution depends on this prescription and the final stationary state, if it exists, might also be prescription-dependent. The most popular conventions are the Itô and Stratonovich ones. However, it is possible to work within a more general scheme usually referred to as *generalized Stratonovich prescription* [@Hanggi1978] or *$\alpha$-prescription* [@Janssen-RG]. This convention is parametrized by a parameter $\alpha$ ($0 \le \alpha \le 1$) and recovers the Itô and Stratonovich prescriptions as the $\alpha=0$ and $\alpha=1/2$ particular cases. The concept of equilibrium in these systems should be carefully defined, since the forward and backward stochastic evolutions are generally performed with different dual prescriptions and, as a consequence, the usual detailed balance relations should be properly generalized [@ArenasBarci3].
Recently, a class of inhomogeneous and nonlinear FP equations [@MarizTsallis] was considered within the Itô prescription. It was shown that stationary solutions for the probability density are of the $q$-exponential form, namely $$\begin{aligned}
p_{\rm st}(x) = P_q(V) &\equiv& N_q \,e_q^{-\beta V(x)} \nonumber \\
&\equiv& N_q \, [1 - (1-q)\beta V(x)]^{\frac{1}{1-q}},
\label{q-exp}\end{aligned}$$ where $V(x)$ is any confining potential, $N_q$ a normalization constant, $\beta$ and $q$ are two real numbers characterizing the distribution, and $e_1^z=e^z$; $\beta$ is an inverse effective “temperature” [@effectivetemperature]. In particular, $\lim_{q\to 1} P_q(V)= N_1\, e^{-\beta V(x)}$, the usual Boltzmann-Gibbs (BG) distribution.
As mentioned above, in [@MarizTsallis], the FP equation was deduced within the Itô convention. In the present paper we analyze how the stationary distributions depend on the particular prescription used to define the stochastic process. To do this, we first deduce the inhomogeneous and nonlinear FP equation within the generalized Stratonovich convention, parametrized by $\alpha$, and then we look for stationary solutions of a family of processes, defined by a particular class of constitutive relations between drift and dissipation. As we shall see, the model is parametrized with two real numbers, namely $\eta$ and $\theta$, defined hereafter. The first one measures the nonlinearity of the system, while the second one is related with inhomogeneity; the point $(\eta,\theta)=(0,0)$ corresponds to the linear homogeneous particular case and represents a normal diffusion process. The stationary-state solutions depend on the values of these parameters. In particular, it will become clear that the solutions are, in the linear limit $\eta \to 0$, nonanalytic in the space $(\eta, \theta)$. We have found that, for the general case in which $(\eta,\theta)\neq (0,0)$, the stationary probability distributions are $q$-exponentials with an index $q$ which is [*independent of the stochastic prescription*]{}. The different conventions, characterized by $\alpha$, do modify the temperature parameter $\beta$, but [*not*]{} $q$.
The paper is organized as follows. In section \[NLFP\] we present the nonlinear inhomogeneous FP equation and define our model. In section \[eta0\] we study the linear limit ($\eta = 0$, $\forall \theta$), and in section \[etaneq0\] we address the general case. Finally, we discuss our results in section \[conclusions\].
The nonlinear Fokker-Planck equation for multiplicative Markov processes {#NLFP}
========================================================================
Consider a Markovian multiplicative stochastic process described by the Langevin equation $$\frac{dx}{dt} = F(x,t) + [\phi(x,t)]^{1/2}\xi(t),
\label{Langevin}$$ where $\langle\xi(t)\rangle=0$ and $\langle\xi(t)\xi(t')\rangle=\delta(t-t')$. $F(x,t)$ is the drift force, and $\phi(x,t)$ is in principle an arbitrary function that models the state-dependent diffusion process. As it is well known, this equation should be complemented with a prescription to integrate the Wiener integral. In this paper, we use the *generalized Stratonovich prescription* [@Hanggi1978] or *$\alpha$-prescription* [@Janssen-RG]. Briefly speaking, it is necessary to give sense to the ill-defined product $[\phi(x(t),t)]^{1/2}\xi(t)$, since $\xi(t)$ is delta-correlated. By definition, the Riemann-Stieltjes integral of a Wiener process $W(t)$ with $\xi(t)=dW(t)/dt$ is $$\begin{aligned}
\lefteqn{
\int [\phi(x(t),t)]^{1/2}\; dW(t)=} \nonumber \\
&& \lim _{n\to\infty}
\sum_{j=1}^n
[\phi(x(\tau_j),\tau_j)]^{1/2}(W(t_{j+1})-W(t_j))
\label{eq.Wiener}\end{aligned}$$ where $\tau_j$ is taken in the interval $[t_j,t_{j+1}]$ and the limit is taken in the sense of [*mean-square limit*]{} [@Gardiner]. For a smooth measure $W(t)$, the limit converges to a unique value, regardless the value of $\tau_j$. However, $W(t)$ is not smooth, in fact, it is nowhere integrable. In any interval, white noise fluctuates an infinite number of times with infinite variance. Therefore, the value of the integral depends on the prescription for the choice of $\tau_j$. In the “generalized Stratonovich prescription” we choose $$x(\tau_j)=(1-\alpha)x(t_j)+\alpha x(t_{j+1})\mbox{~~ with~~} 0\le \alpha
\le 1.
\label{eq.prescription}$$ In this way, $\alpha=0$ corresponds with the pre-point Itô interpretation and $\alpha=1/2$ coincides with the (mid-point) Stratonovich one. Moreover, the post-point prescription, $\alpha=1$, is also known as the kinetic or anti-Itô interpretation. In principle, each particular choice of $\alpha$ fixes a different stochastic evolution.
In many physical applications, a weakly colored Gaussian-Markov noise with a finite variance [@Hanggi-shot] is considered. In this case, there is no problem with the interpretation of equation (\[Langevin\]) and the limit of infinite variance can be taken at the end of the calculations. This regularization procedure is equivalent to the Stratonovich interpretation, $\alpha=1/2$ [@vanKampen; @Zinn-Justin]. However, in other applications, like chemical Langevin equations [@vanKampen] or econometric problems [@Mantegna; @Bouchaud], the noise can be considered principally white, since it could be a reduction of jump-like or Poisson-like processes. In such cases, the Itô interpretation ($\alpha=0$) should be more suitable. Hence, the interpretation of equation (\[Langevin\]) depends on the physics behind a particular application. Once the interpretation is fixed, the stochastic dynamics is unambiguously defined.
From the stochastic equation (\[Langevin\]) it is possible to derive a Fokker-Planck equation, given by [@Celia; @Coutinho; @ArenasBarci2; @ArenasBarci3] $$\begin{aligned}
\frac{\partial p(x,t)}{\partial t} &=& - \frac{\partial}{\partial x} \left\{\left[ F(x)
+ \frac{\alpha}{2} \frac{\partial \phi(x,t)}{\partial x} \right] p(x,t)\right\} \nonumber \\
&& + \frac{1}{2} \frac{\partial^2}{\partial x^2}\left\{\phi(x,t)p(x,t)\right\},
\label{FP} \end{aligned}$$ where $p(x,t)$ is the time-dependent probability distribution and $\alpha \in
[0,1]$ parametrize the stochastic prescription.
If the function $\phi(x,t)$ is an “external” fixed function, modeling a simple state diffusion process, then, Eq. (\[FP\]) is linear. However, as discussed in Ref. [@Lisa1], the diffusion function could depend on the probability distribution itself, for instance, $$\phi (x,t) = D \left[ g(x)\right]^{\theta} \left[ p(x,t)\right]^{\eta},
\label{phi}$$ where $D$ is a constant diffusion coefficient, $g(x)$ is an arbitrary well-behaved function and $p(x,t)$ is a solution of the FP equation. With this choice, Eq. (\[FP\]) is a nonlinear equation describing a state-dependent diffusion process with non-trivial particle-bath couplings [@Lisa1]. The real constants $\theta$ and $\eta$ control the relative strength of these effects. For instance, the point $\eta=\theta=0$ represents a normal diffusion process driven by a stochastic additive Langevin equation. On the other hand, the line $\eta=0, \theta\neq 0$, represents a usual state-dependent diffusion process, described by a multiplicative Langevin equation. Moreover, the general case $\eta\neq 0$, is a multiplicative process, whose diffusion functions should be self-consistently computed by solving the related nonlinear FP equation.
Eq. can be written as a continuity equation, $$\frac{\partial p(x,t)}{\partial t} = \frac{\partial J(x,t)}{\partial x}$$ where the current of probability is given by $$\begin{aligned}
J(x,t) &=& \left[-F(x) + (1-\alpha) \frac{1}{2} D \frac{\partial \phi(x,p)}{\partial x}+ \right. \nonumber \\
&&+ \left. \frac{1}{2}D\phi(x,p) \frac{\partial}{\partial x} \right]p(x,t).
\label{J} \end{aligned}$$ Here we have indicated that $\phi(x,p(x,t))$ could be a function of $p(x,t)$ given by Eq. (\[phi\]).
The equilibrium distribution is defined as the stationary solution with zero current of probability, [*i.e.*]{}, $$P_{eq}(x)=\lim_{t\to\infty} p(x,t)$$ supplemented with $\lim_{x \to \pm \infty}J(x,t)=0$. In the following sections we will find the equilibrium probability distribution in the whole parameter range $\{\eta, \theta\}$.
The linear case, $\eta=0$ {#eta0}
=========================
Let us begin by analyzing stationary states of the simpler case $\eta=0$. For this case, $$\phi(x)=D[g(x)]^\theta$$ and the Fokker-Planck equation (Eq. (\[FP\])) is linear. The stationary states have been studied in Refs. [@Celia; @Coutinho; @ArenasBarci3] for different particular cases. In this section we summarize the main results and procedures in order to present them in a unified scheme and to compare them with the nonlinear case.
The equilibrium solution takes the form [@ArenasBarci3] $$P_{eq}(x)=N\; e^{-U_{eq}(x)},
\label{Peq}$$ where $N$ is a normalization constant and the “effective potential” is given by $$U_{\rm eq}(x)=-2 \int^x \frac{F(x')}{D g(x')^\theta} dx'+ (1-\alpha) \theta\ln
g(x).
\label{Ueq}$$ Thus, as already mentioned, the equilibrium distribution depends on the particular stochastic prescription used to define the Langevin equation. For general functions $F(x)$ and $g(x)$, the probability density distribution is given by Eqs. (\[Peq\]) and (\[Ueq\]). The only constraint is a condition of integrability in order to compute the normalization factor $N$. To go further, we need to impose constitutive relations between drift and dissipation. For instance, suppose that the system is submitted to a conservative force, with energy potential $V(x)$. For $\theta=0$, the resulting process is additive and $$U_{\rm eq}(x)= \left(\frac{2}{D}\right)\; V(x)
\label{Ueq0}$$ up to an unimportant constant term that is absorbed in the normalization. Then, Einstein relation imposes for the inverse temperature $\beta=2/D$, leading to the Boltzmann distribution. On the other hand, we could impose, for $\theta\neq 0$, a local generalization of Einstein relation $$F(x)=-\left(\frac{\beta}{2D g^{\theta}(x)}\right) \; V'(x)$$ ending with the solution [@ArenasBarci2; @ArenasBarci3] $$U_{\rm eq}(x)=\beta V(x)+ (1-\alpha)\theta\ln g(x).
\label{Ueq1}$$ We see that, for multiplicative noise, the final distribution is generally not of the Boltzmann type, even for the usual prescriptions of Itô ($\alpha=0$) or Stratonovich ($\alpha=1/2$). The exception is the anti-Itô interpretation ($\alpha=1$) which, together with the local Einstein relation, leads to the usual thermodynamical equilibrium distribution. For this reason, this convention is also called the kinetic prescription. An interesting particular case is to consider a “free” particle in an inhomogeneous dissipative medium, where $V=0$ and the probability distribution is a power law of the form $$P_{eq}= N \frac{1}{g(x)^{\theta(1-\alpha)}}\,,$$ assuming it is normalizable.
Moreover, we could impose constitutive relations different from the local Einstein relation, such as the one used in Ref. [@Celia]. We can choose, for instance, $F(x)=- V'(x)$, $D g(x)= A + B V(x)$ and $\theta=1$; for simplicity we shall assume $A>0$ and $B>0$. Substituting these expressions in Eq. (\[Ueq\]) we immediately find a $q$-exponential form (Eq. (\[q-exp\])) with $$q =\frac{2(B+1)-\alpha B}{B +2 -\alpha B} \mbox{~~~~and~~~~} \beta = \frac{B(1-\alpha)+2}{A}.
\label{qalpha}$$
Therefore, we have shown that, using the general solution Eqs. (\[Peq\]) and (\[Ueq\]), in the linear case $\eta=0$, we can find different types of equilibrium distributions, such as the Boltzmann or the $q$-exponential distribution, depending on the constitutive relation imposed between drift and dissipation and on the particular stochastic prescription used to derived the FP equation. Let us also notice that, whenever $A$ and $B$ have the same sign, the inverse temperature $\beta$ is positive, as normally expected; if both are positive (negative), then $q>1$ ($q<1$), which corresponds to long-tailed distributions (compact support distributions).
The nonlinear case, $\eta\neq 0$ {#etaneq0}
================================
The solution of the nonlinear and inhomogeneous FP equation (\[FP\]) with (\[phi\]) for general values of $g(x)$ and $F(x)$ is quite involved. We will look for solutions imposing the constitutive relations [@MarizTsallis] $$F(x)=-V'(x)\mbox{~~~~and~~~~}g(x)=A+BV(x),
\label{constraints}$$ where, as already mentioned, $A, B$ are real positive constants.
Looking for stationary solutions $\partial p(x,t) / \partial t = 0$ and assuming appropriate boundary conditions which guarantee a null net flux, we have $$\begin{aligned}
\lefteqn{
\frac{\partial F(x)p(x,\infty)}{\partial x} = } \nonumber \\
&& \frac{D}{2}\frac{\partial}{\partial x}\left[ (1-\alpha)\frac{\partial \phi(x,p)}{\partial x} + \phi(x,p) \frac{\partial}{\partial x} \right]p(x,\infty).
\label{StationaryFP}\end{aligned}$$ The choices made in Eq. (\[constraints\]) allow us to write the differential equation Eq. (\[StationaryFP\]) in terms of the variable $V$, obtaining $$(1-\alpha)\frac{\partial [g(V)^{\theta}p(V)^{\eta}]}{\partial V}p(V) + g(V)^{\theta}p(V)^{\eta} \frac{\partial p(V)}{\partial V} = - \frac{2p(V)}{D},$$ where $p(V)\equiv p(V(x),\infty)$. This equation can be re-written in the form of a *Bernoulli equation* [@Bernoulli] (see also [@bemski]), $$\begin{aligned}
\frac{d p(V)}{d V}& +&
\frac{(1-\alpha)\theta B g(V)^{-1}}{[(1-\alpha)\eta +1]}p(V) =
\nonumber \\
&&- \frac{2g(V)^{-\theta}}{D[(1-\alpha)\eta +1]}[p(V)]^{1-\eta},
\label{Bernoulli}\end{aligned}$$ a class of nonlinear differential equations that can be linearized by a suitable change of variables.
For $\eta=0$, we recover the linear equation we have treated in the last subsection. For $\eta\neq 0$ we can perform the nonlinear change of variables, $$Z(V)=C_N\; p(V)^\eta\; , \label{Z}$$ where $C_N$ is a normalization constant. With this, Eq. (\[Bernoulli\]) becomes a first order linear ordinary differential equation $$\frac{d Z}{dV} + \frac{(1-\alpha)\eta \theta B g(V)^{-1}}{[(1-\alpha)\eta +1]}Z = - \frac{2g(V)^{-\theta}C_N \eta}{D[(1-\alpha)\eta +1]},
\label{ODE}$$ with the general solution $$\begin{aligned}
Z&=& (A+BV)^{-\frac{(1-\alpha)\eta \theta}{(1-\alpha)\eta +1}}
\left\{C_I - \frac{2 C_N \eta}{BD[(1-\alpha)\eta + 1- \theta]}
\times \right. \nonumber \\
&\times& \left. \left[(A+BV)^{\frac{(1-\alpha)\eta +1 -\theta}{(1-\alpha)\eta +1}} - A^{\frac{(1-\alpha)\eta +1 -\theta}{(1-\alpha)\eta +1}} \right] \right\},
\label{Bernoulli_sol}\end{aligned}$$ where $C_I$ is an integration constant which depends on the “initial” condition. Thus, Eqs. (\[Z\]) and (\[Bernoulli\_sol\]) provides a family of explicit solutions of the nonlinear FP equation in terms of two constants, $C_N$ and $C_I$, which should be adjusted by means of an initial condition and the probability distribution normalization.
$\theta=0$
----------
In the particular case $\theta= 0$, the inhomogeneity of the dissipation function $\phi(x)$ comes only from the probability density. The stationary solution can be read from Eq. (\[Bernoulli\_sol\]) $$Z=C_N p(V)^\eta=C_I\left(1- \frac{2C_N\eta}{D C_I[1+\eta-\alpha\eta]}\; V \right),$$ which can be rewritten in terms of a $q$-exponential (Eq. (\[q-exp\])) with $N_q=(C_I/C_N)^{1/\eta}$, $$q=1-\eta \mbox{~~~~and~~~~}N_q^{1-q}\beta=\frac{2}{D(1+\eta-\alpha\eta)} \,.
\label{theta0}$$ Interestingly enough, the index $q$ is $\alpha$-independent. We will show that this is a general feature of nonlinearity ($\eta\neq 0$). In contrast, the inverse temperature $\beta$ depends on the prescription that has been used. In particular, Eq. (\[theta0\]), in the Itô prescription, i.e. $\alpha=0$, coincides with the result for the homogeneous nonlinear model obtained in Refs. [@Plastino95; @Tsallis96].
$\theta\neq0$
-------------
For the inhomogeneous and nonlinear case, i.e. $\theta \neq 0$, we can straightforwardly verify $$C_I = -\frac{2C_N \eta A^{\frac{(1-\alpha)\eta +1 -\theta}{(1-\alpha)\eta +1}}}{BD[(1-\alpha)\eta +1 -\theta]} \,.$$ Consequently, using Eq. (\[Bernoulli\_sol\]), we obtain $$Z = - \frac{2C_N \eta}{BD[(1-\alpha)\eta +1 -\theta]}(A+BV)^{1-\theta} \,,$$ hence, from Eq. (\[Z\]), $$p(V) = \left\{ -\frac{2 \eta}{BD[(1-\alpha)\eta +1 -\theta]} \right\}^{\frac{1}{\eta}}\!\!(A+BV)^{\frac{1-\theta}{\eta}} \,.
\label{p_sol}$$ This expression can be written in the form of a $q$-exponential (Eq. (\[q-exp\])) with $$q = 1-\frac{\eta}{1-\theta},
\label{q}$$ and the parameter $\beta$ $$\begin{aligned}
\beta &=& \frac{B}{A}\frac{1-\theta}{\eta} \\
&=& \frac{C_N}{C_I}\frac{2 (1-\theta)}{D[(1-\alpha)\eta + 1 -\theta]}A^{-\theta / (1-\alpha)\eta +1}.
\label{beta}\end{aligned}$$
We see that the $q$-exponential distribution is a solution of the nonlinear and inhomogeneous FP equation [*for any value of the stochastic prescription $\alpha$*]{}. Moreover, the value of $q$ itself is [*universal*]{} in the sense that it does not depend on the prescription $\alpha$, as can be seen in Eq. (\[q\]). The index $q$ only depends on the parameters $\eta$ and $\theta$ that measure the relative importance of nonlinearity and inhomogeneity. Of course, this value of $q$ coincides with the one computed in Ref. [@MarizTsallis] using the Itô prescription. Different stochastic prescriptions only affect the temperature parameter $\beta$.
From Eq. (\[q\]), it could be wrongly concluded that in the linear limit $\eta\to 0$, the probability distribution is of the Boltzmann type, since $q\to 1$. However, as we showed in the previous section, in the linear case, we find power-law solutions with non-universal $q$ ($\alpha$-dependent). In other words, the solutions are [*not*]{} analytic in the linear limit $\eta\to 0$. In variance with this fact, the limit $\theta\to 0$ is perfectly well defined as can be seen from Eqs. (\[q\]) and (\[theta0\]).
Conclusions
===========
We have presented an inhomogeneous and nonlinear Fokker-Planck equation describing a generalized Markov stochastic process in the “generalized Stratonovich prescription”. This prescription is parametrized by a real parameter $0\leq \alpha \leq 1$, and contains the usual Stratonovich ($\alpha=1/2$), Itô ($\alpha=0$) and kinetic ($\alpha=1$) prescriptions as particular cases. We have also parametrized nonlinearity and inhomogeneity by means of two parameters, $\eta$ and $\theta$, in such a way that the point $\eta=\theta=0$ represents a normal difusion process. In this way, we have unified and generalized different results already obtained in the literature [@Plastino95; @Tsallis96; @ArenasBarci3; @MarizTsallis; @Celia; @Coutinho].
--------------------- -- --------------------------------------------- -- ------------------------------------
Fokker-Planck Linear Nonlinear
equation $(\eta = 0)$ $(\eta \neq 0)$
\[1ex\] Homogeneous Additive noise Multiplicative noise
$(\theta = 0)$ $q =1$ $q = 1- \eta$
Inhomogeneous Multiplicative noise Multiplicative noise
$(\theta \neq 0)$ $q =\frac{2(B+1)-\alpha B}{B +2 -\alpha B}$ $q = 1-\frac{\eta}{1-\theta}$
\[1ex\]
--------------------- -- --------------------------------------------- -- ------------------------------------
: Stationary solutions of nonlinear inhomogeneous FP equation. The $q$-exponential distribution has been obtained for the family of constitutive relations given by . Notice that, in the nonlinear case, the exponent $q$ does not depend on the stochastic prescription, while it is not the case for the linear inhomogenous FP equation (the $q$ value for this case recovers the results in [@Celia] for the particular instances $\alpha=0$ and $\alpha=1/2$ by doing $B \to 2M/\tau C$; it also recovers, for $\alpha=0$, the results in [@MarizTsallis] by doing $B \to BD$). Let us emphasize that $q$ cannot be arbitrarily large for a given $V(x)$, otherwise the normalizability property will be lost. For example, if we are dealing with $q$-Gaussians, then it must be $q<3$.[]{data-label="Summary"}
We have solved the stationary FP equation for all values of the parameters $\eta, \theta, \alpha$. For the linear case, $\eta=0$, we have found a general solution depending on the drift, the dissipation and the prescription $\alpha$. The Boltzmann distribution is obtained when a generalization of the Einstein relation and the kinetic prescription, $\alpha=1$, are imposed. In all other cases, the solution is more involved.
There exist a link between the stationary solutions of the linear or nonlinear FP equation and the distribution obtained by extremizing a particular entropy under simple specific constraints. Indeed, the connection between the linear FP equation and Boltzmann-Gibbs entropy is well known since long. Analogously, for nonlinear FP equations yielding specific classes of anomalous diffusion, nonadditive entropic functionals have been analyzed in detail [@Plastino95; @Tsallis96]. In addition, this remarkable link has also been found for even more general nonlinear FP equations and entropic forms [@CuradoNobreetal; @RibeiroTsallis].
We analyzed a family of constitutive relations between drift and dissipation that results in a $q$-exponential distribution, thus exhibiting a possible mechanism compatible with nonextensive statistical mechanics. In the nonlinear case, $\eta\neq 0$, the value of the exponent $q$ is, remarkably enough, $\alpha$-independent. The different prescriptions that define the stochastic process [*only*]{} affect the inverse temperature $\beta$. In the linear case, in contrast, the exponent of the power-law does depend on the stochastic prescription $\alpha$. This clearly shows that the linear limit of the solutions is not analytic, namely, $ \lim_{\eta\to 0} p_{eq}[\eta,\theta]\neq p_{eq}[0,\theta] $.
Table \[Summary\] summarizes the values of the entropic index $q$ for the $q$-exponential distributions obtained as the stationary-state distributions for nonlinear inhomogeneous Fokker-Planck equation when using the particular relations given by . A variety of possible physical applications of the present results can be found in [@books] and references therein.
The Brazilian agencies CNPq and FAPERJ are acknowledged for partial financial support. ZGA is a CNPq Postdoctoral Fellow.
[99]{}
A.R. Plastino, A. Plastino, Physica A [**222**]{}, 347 (1995) .
C. Tsallis, D.J. Bukman, Phys. Rev. E [**54**]{}, R2197 (1996).
T.D. Frank, [*Nonlinear Fokker-Planck Equations: Fundamentals and Applications*]{}, Springer, Germany (2005).
C. Tsallis, E.K. Lenzi, Chem. Phys. [**284**]{}, 341 (2002).
F.D. Nobre, E.M.F. Curado and G.A. Rowlands, Physica A 334 (2004), p. 109; V. Schwammle, E.M.F. Curado and F.D. Nobre, Eur. Phys. J. B 70 (2009), p. 107; M.S. Ribeiro, F.D. Nobre and E.M.F. Curado, Entropy 13 (2011), p. 1928 \[[*Tsallis Entropy*]{}, Special Issue\].
C. Tsallis, J. Stat. Phys. [**52**]{}, 479 (1988).
M. Gell-Mann and C. Tsallis Eds., [*Nonextensive Entropy – Interdisciplinary Applications*]{}, Oxford University Press, New York (2004); C. Tsallis, [*Introduction to Nonextensive Statistical Mechanics: approaching a complex world*]{}, Springer, New York (2009).
L. Borland, Phys. Rev. E [**57**]{}, 6634 (1998)
P. Hänggi, Helv. Phys. Acta [**51**]{}, 183 (1978).
H.K. Janssen, [*From phase transitions to chaos*]{}, Topics in Modern Statistical Physics, ed. G. Györgyi, I. Kondor, L. Sasvári and T. Tél, World Scientific, Singapore (1992).
Z. González Arenas, D. Barci, J. Stat. Mech. P12005 (2012).
A.M. Mariz, C. Tsallis, Phys. Lett. A [**376**]{}, 3088 (2012).
L.A. Rios, R.M.O. Galvão, L. Cirto, Phys. Plasmas [**19**]{}, 034701 (2012); L.J.L. Cirto, V.R.V. Assis and C. Tsallis, Physica A, 393 (2014), p. 286; W. Niedenzu, T. Grießer and H. Ritsch Europhys. Lett., 96 (2011), p. 43001.
C. W. Gardiner, [*Handbook of stochastic methods for physics, chemistry and the natural sciences*]{}, Heidelberg: Springer-Verlag, Berlin (1996).
P. Hänggi, [*Z. Physik B* ]{} [**75**]{}, 275 (1989).
N. G. van Kampen, [*Stochastic Processes in Physics and Chemistry*]{}, UK: Elsevier, London, (2007).
J. Zinn-Justin, [*Quantum field theory and critical phenomena*]{}, USA: Oxford University Press), (2002).
R. N. Mantegna and H. E. Stanley, [*An introduction to econophysics: correlations and complexity in finance*]{}, Cambridge, UK: Cambridge University Press, (2000).
J. P. Bouchaud and M. Potters, [*Theory of financial risk and derivative pricing: from statistical physics to risk management*]{}, Cambridge University Press, (2003).
Z. González Arenas, D. Barci, Phys. Rev. E [**85**]{}, 041122 (2012).
C. Anteneodo, C. Tsallis, J. Math. Phys. [**44**]{}, 5194 (2003).
B. Coutinho dos Santos and C. Tsallis, Phys. Rev. E, [**82**]{}, 061119 (2010).
L. Elsgoltz, [*Ecuaciones diferenciales y cálculo variacional*]{}, Edit. MIR, Moscow (1969).
C. Tsallis, G. Bemski and R.S. Mendes, Phys. Lett. A [**257**]{}, 93 (1999).
M.S. Ribeiro, C. Tsallis and F.D. Nobre, Phys. Rev. E [**88**]{}, 052107 (2013); M.S. Ribeiro, F.D. Nobre and C. Tsallis, Phys. Rev. E [**89**]{}, 052135 (2014).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'An Adversarial System to attack and an Authorship Attribution System (AAS) to defend itself against the attacks are analyzed. Defending a system against attacks from an adversarial machine learner can be done by randomly switching between models for the system, by detecting and reacting to changes in the distribution of normal inputs, or by using other methods. Adversarial machine learning is used to identify a system that is being used to map system inputs to outputs. Three types of machine learners are using for the model that is being attacked. The machine learners that are used to model the system being attacked are a Radial Basis Function Support Vector Machine, a Linear Support Vector Machine, and a Feedforward Neural Network. The feature masks are evolved using accuracy as the fitness measure. The system defends itself against adversarial machine learning attacks by identifying inputs that do not match the probability distribution of normal inputs. The system also defends itself against adversarial attacks by randomly switching between the feature masks being used to map system inputs to outputs.'
author:
-
title: Defending Against Adversarial Machine Learning
---
adversarial machine learning, accuracy, probability, feature mask, genetic algorithm, authorship attribution system
Introduction
============
A system can defend itself against adversarial machine learners by randomly switching the feature mask it is using, by detecting unusual distributions of system inputs and reacting defensively to them, or by using other methods. Authorship attribution can be discovered using adversarial machine learning to learn the patterns between the character unigrams and authors. An adversarial machine learner can learn the mappings between system inputs and outputs and learn the Neural Network (NN) or Support Vector Machine (SVM).
A defensive system can detect the system inputs from an adversarial machine learner by comparing the normal probability distribution of the inputs versus the current probability distribution of the inputs and detecting a difference between them. If a difference is detected between the normal inputs to the system and the current inputs to the system based on the probability, then it can be assumed that the current inputs may come from an adversarial machine learner which has learned the input-output mapping to model the NN. Also, if it is discovered that the inputs are along a certain decision boundary that the system has, then the defending system can become wary of the inputs and suspect them of being from an adversarial machine learning attacker system. Since the adversarial machine learner can develop a NN or SVM to model the system, given the input-output mappings, the adversarial machine learning system can then use knowledge of the decision boundaries in the model to inform creating inputs that will “trick" the system that the adversarial machine learning system is attacking.
Generative Adversarial Networks (GANs) are a type of adversarial artificial intelligence. The discriminator operates based on a neural network that has decision boundaries and makes decisions based on inputs. The generative network, the adversary, learns the distribution in order to “fake out" the discriminator. Generative Adversarial Networks learn the distribution, create synthetic data, and can then trick the discriminator system and its decision boundaries. Examples of GANs tricking systems are “deep fakes", where an attacker creates an input to trick the decision boundaries of the original system. The analysis of variance, such as an F-test, ANOVA test, or Student t-Test, can be done. Then, the P-value and T-values can be analyzed to make sure that the guessed value is correct.
The seven phases in the Machine Learning Model Kill Chain are the:
1. Reconnaissance (Recon) Phase
2. Weaponization Phase
3. Delivery Phase
4. Exploitation
5. Installation
6. Command and Control
7. Action
During the Reconnaissance (Recon) Phase, the Machine Learning (ML) models are determined. The ML models are used to protect the defending system from the type of attacks which are to be launched by the attacking system. Then, during the Weaponization Phase, the results of probes are used in an effort to develop an attack on the defending system by learning the defending system’s decision boundary. During the Decision Phase, the defending system’s decision boundary is attacked by the attacking adversarial machine learning system. During the Exploitation Phase, the adversarial machine learning system gathers deeper information about the defending system’s underlying model. The attacking system learns how the defending system’s model will be tuned, how fast new rules can be formed, and how threats are ranked. During the Installation Phase, new rules, or features, that will allow future attacks to happen, are set up. During the Command and Control Phase, a hidden command and control channel is set up to allow for expansion of the attack. Finally, during the Action Phase, the attackers act on their main objective[@Nyugen; @2017].
30
feature masks are developed, and the system switches between the $30$ feature masks to get roughly the same results. If an attacker knows the coding language that the system is written in (Python, in this case), the random generator, and the seed, then the attacker can figure out the $30$ feature masks.
If the number of features is reduced, the accuracy will increase. Feature reduction is done $30$ times to develop the $30$ feature masks. The cycle of attack by the Adversarial Author (Attacking System) and defense by the Authorship Attribution System (AAS) (Defending System) is shown in Figure \[figure\_AdvAuthor\_AAS\].
![Cycle of Adversarial Author (Attacking System) vs. Authorship Attribution System (Defending System)[]{data-label="figure_AdvAuthor_AAS"}](AdvAuthor_AAS.png){width="40mm"}
Acceptable ranges of drops in accuracy by the AAS whilst under attack are less than $11$ percent. The baseline accuracy of the AAS is between $60$ to $80$ percent, depending on which algorithm is being used. Therefore, an $11$ percent drop in accuracy on $60$ to $80$ percent accuracy would result in $49$ to $69$ percent accuracy.
Term Frequency-Inverse Document Frequency (TFIDF) is a numerical statistic that is intended to reflect how important a letter is in a character unigram. TFIDF is used as a weighting factor in searches, and the value increases proportionally to the number of times a letter appears in the character unigram and is offset by the number of character unigrams in the set that contain the letter.
When the simple (unweighted) Euclidean distance is used, normalization provides an equal weighting to all features, whereas some features would have more importance than others when normalization is not used. Normalization of the training data is done by subtracting the mean and dividing the mean and standard deviation of the training data.
Standardization provides a way to scale the test data with the mean and standard deviation of the training data.
It is important for a defensive system to establish a sense of normalcy for inputs. A measure of normalcy is created by the algorithm knowing the distribution of normal inputs. If the distribution of the inputs has changed, a notification is made by the defensive algorithm to alert the system that there is something that has changed and the machine learning algorithm is receiving abnormal inputs. This violates the stationarity assumption, since the inputs are not normal and expected inputs. The inputs with abnormal distributions can trick the decision boundary and create bad results from the machine learning algorithm.
In machine learning, Naïve Bayes classifiers are a family of simple “probabilistic classifiers” based on applying Bayes’ theorem with strong (naïve) independence assumptions between the features. They are among the simplest Bayesian network models.
Naïve Bayes classifiers are highly scalable, requiring a number of parameters linear in the number of variables (features/predictors) in a learning problem. Maximum-Likelihood training can be done by evaluating a closed-form expression, which takes linear time, rather than by expensive iterative approximation as used for many other types of classifiers.
In the statistics and computer science literature, Naïve Bayes models are known under a variety of names, including simple Bayes and independent Bayes. All these names reference the use of Bayes’ theorem in the classifier’s decision rule, but Naïve Bayes is not (necessarily) a Bayesian method.
The discussion so far has derived the independent feature model, that is, the Naïve Bayes probability model. The Naïve Bayes classifier combines this model with a decision rule. One common rule is to pick the hypothesis that is most probable; this is known as the Maximum A-Posteriori (MAP) decision rule.
When dealing with continuous data, a typical assumption is that the continuous values associated with each class are distributed according to a normal (or Gaussian) distribution. For example, suppose the training data contains a continuous attribute, $x$. The data is first segmented by the class, and then compute the mean and variance of $x$ in each class.
The algorithms are sorted into equivalence classes based on whether there are statistically significant differences between the algorithms. The ANOVA and Student t-Tests are used to determine statistical significance and to compare the algorithms.
Methodology
===========
An Authorship Attribution System (AAS) that is resistant to Adversarial Authorship Attacks is developed. After developing the AAS, a method for testing it is developed. The AAS reads from a file that contains the adversarial texts, called “AdversarialTests.txt". This file contains the list of adversarial texts that will be located in the same directory as the AAS. The “AdversarialTests" will contain the file names of a number of adversarial texts (advText00.txt, advText01.txt, ..., advText24.txt). The AAS reads in each adversarial text and classifies it by placing the results in a file, called “AdversarialTestResults.txt". The “AdversarialTestResults.txt" file will have a classification associated with each adversarial text of the form:
- advText00.txt $\rightarrow 1000$
- advText01.txt $\rightarrow 1022$
- ...
- advText24.txt $\rightarrow 1005$
Defending System
----------------
Principle Component Analysis (PCA) may be used to reduce the number of features need for the Feature Mask to correctly identify the author. Principal Component Analysis consists of keeping the best and most effective (or highest magnitude) features and then either not using or getting rid of the other features. np.count\_nonzero() is a function that can be used to count the number of features that are being used in the feature mask. The reduction of features results in an increase in accuracy.
Switching between Feature Masks may be used as a way to keep the attacking system from learning the decision boundaries of the defending system. Using this method, the defending system randomly switches between equivalent feature masks.
Use a vector of attack criteria to decide whether or not the defending system thinks that it is being attacked. If multiple factors give positive values, then the system will determine that it must currently be under attack. Then, the defending system will respond accordingly to the perceived attack.
Attacking System
----------------
The defending system may be attacked during any of the $5$ Stages of Machine Learning. The $5$ Stages of Machine Learning are:
1. Measuring
2. Feature Selection
3. Learning Model - Lazy Learner General Regression Neural Network (GRNN) No Model - Support Vector Machine (SVM) - Feedforward Neural Network (FFNN))
4. Training
5. Prediction
There is a conditional probability, or an association, between the inputs the defending system should expect to receive and the inputs that it is actually receiving. If the system begins to receive inputs that do not match the distribution that it is expecting, then the system can recognize that the input distribution is different and react accordingly. This makes the attacking system have to then change its method of attack, in order to continue attacking the defending system in an effective manner.
The effect of feedback control on measurement noise is that the system handles measurement noise well, if the noise is Gaussian white noise. Therefore, the system may give misleading outputs. One possibility to avoid mistakes is to compute the correlation of the equation error and check whether it is white noise. The decision boundaries are “tricked" or “fooled" by modifying the input based on knowing how the system will respond to changes in the input. This is a technique used by adversarial machine learning attackers to trick a system’s decision boundaries. The defending system must find a way either to mitigate the effect of, or to recognize, the attack.
The Steady-State Genetic Algorithm (SSGA) is used to evolve the feature masks in a way that increases the accuracy. The SSGA replaces the worst fit individual in the population with the child. A population size of $30$ is used. A mutation rate of $0.5$ is used. There are $95$ features in each feature vector, and there are $30$ feature masks. To clarify, each feature mask is one member of the population, so there are $30$ feature masks, or population members. Binary Tournament Selection is used for selecting parents from the feature masks. Uniform Crossover is used to produce the child’s features from the parents’ features.
Radial Basis Function Support Vector Machine
--------------------------------------------
The Radial Basis Function Support Vector Machine (RBFSVM) is implemented using scikit-learn with its default kernel being a radial basis function and this is a method whose value depends on the distance from the origin or from some point. In this paper, the output received from the processing model is then evaluated on the RBFSVM model and the accuracy is obtained.
Linear Support Vector Machine
-----------------------------
The Linear Support Vector Machine (LSVM) is implemented using scikit-learn. The LSVM is the simplest form of a support vector machine. Along with the RBFSVM, the output of the processing model is passed through the LSVM and its accuracy is also obtained.
Feedforward Neural Network
--------------------------
The advantages of Multi-Layer Perception (MLP) are its capability to learn nonlinear models and its capability to learn models in real-time (online learning). There are three main disadvantages of MLP. The first disadvantage is that MLP with hidden layers has a non-convex loss function where there exists more than one local minimum. Therefore, different random weight initializations can lead to different validation accuracy. Secondly, MLP requires tuning a number of hyper-parameters such as the number of hidden neurons, layers, and iterations. Lastly, MLP is sensitive to feature scaling.
NNs may be very large, so it is impractical to write down the gradient formula by hand for all parameters. Back-propagation is the recursive application of the Chain Rule along a computational graph to compute the gradients of all inputs/parameters/intermediates. The implementations of back-propagation maintain a graph structure, where the nodes implement forward()/backward() API. The forward pass computes the result of an operation and saves any intermediates needed for gradient computation in memory. The backward pass applies the Chain Rule to compute the gradient of the loss function with respect to the inputs.
Experiment
==========
Each author uses $4$ writing samples, and there are $25$ authors. There are thus $100$ training sets. There are $7$ letters in the alphabet which are used. A Genetic Algorithm (GA) with uniform crossover is used. The algorithm is run $30$ times, and the best feature mask is found out of all $30$ of the runs. The best fit individual is extracted from the character unigrams, or the number of features. The feature consistency is the consistency of a feature being present over $30$ runs.The feature mask consists of $1$’s and $0$’s. The feature mask multiplied by the learning set vector results in the derived learning set vector gets rid of the components that are not to be considered.
For feature selection, a population is randomly generated. The population size that is used is $30$. The mutation rate is $0.5$. There are $100$ feature vectors. The feature extraction is used to come up with the dataset, or feature set. The fitness value is found for each feature. In each algorithm, accuracy is used for the fitness value. In other words, the accuracy for the derived training set is used for determining the fitness for each individual in the population.
The evolutionary process is continued until a stopping rule is reached. Stopping rules that can be used are number of function evaluations, reaching a certain fitness value, or another stopping criteria may be used. The algorithm used in the test had a stopping rule of the limiting number of function evaluations being $1000$ and the fitness value, or accuracy, reaching $0.99754$. This means that the algorithm must run until it reaches either the limit of $1000$ function evaluations or $99.754\%$ accuracy, depending on whichever stopping limit is reached first.
**Algorithm** **Training** **Evaluation** **Validation** **Test**
--------------- -------------- ---------------- ---------------- ----------
***GRNN*** 70% 10% 10% 10%
***RBFNN*** 80% 0% 10% 10%
***SVM*** 90% 0% 0% 10%
***FFNN*** 80% 0% 10% 10%
: Algorithm Dataset Allocation
\[table:Algorithm\_Dataset Allocation\_Summary\]
Radial Basis Function Support Vector Machine
--------------------------------------------
For the RBFSVM model, the test set is evaluated for the default values of the RBFSVM function available built into the function in scikit-learn. The function is passed with train data and train labels as parameter for training purpose. It is evaluated with the evaluation set and evaluation labels as parameters which are then used to obtain the accuracy.
Linear Support Vector Machine
-----------------------------
For the LSVM model, the linear kernel is used to process the evaluation set. The function is passed with train data and train labels as parameter for training purpose. It is then evaluated with evaluation set and evaluation labels as parameters which are used to obtain the accuracy.
Feedforward Neural Network
--------------------------
The MultiLayer Perceptron (MLP) function is used to train the model with training data and train labels for the FFNN and then the evaluation data and evaluation labels are used to test the model for accuracy.
Results
=======
The ANOVA Tests gives a P-value ${< 0.5}$ which suggests that there is no statistical significant difference and has weak acceptance of hypothesis. This means that all the three models are in different equivalence classes. The Student t-Test values for all combinations of RBFSVMs, LSVMs and FFNNs have t-Stat value ${< -0.5}$ which suggests that the null hypothesis is accepted for each combination. The results of the statistical analysis of a comparison of all the algorithms is shown in Table \[table\_comparison\_results\].
--------------- -------------- ------------ ------------
**** **** **** ****
**Run** ***RBFSVM*** ***LSVM*** ***FFNN***
1 0.68 0.72 0.68
2 0.72 0.64 0.68
3 0.6 0.72 0.76
4 0.68 0.72 0.68
5 0.68 0.72 0.72
6 0.6 0.68 0.68
7 0.68 0.72 0.72
8 0.56 0.68 0.68
9 0.6 0.68 0.72
10 0.56 0.72 0.76
11 0.68 0.72 0.84
12 0.6 0.8 0.72
13 0.6 0.72 0.76
14 0.6 0.68 0.8
15 0.6 0.76 0.76
16 0.64 0.68 0.76
17 0.64 0.68 0.68
18 0.64 0.68 0.72
19 0.6 0.72 0.72
20 0.6 0.84 0.72
21 0.6 0.68 0.68
22 0.6 0.72 0.68
23 0.68 0.64 0.76
24 0.68 0.8 0.76
25 0.64 0.76 0.76
26 0.6 0.68 0.64
27 0.72 0.76 0.76
28 0.6 0.68 0.76
29 0.64 0.8 0.68
30 0.68 0.76 0.72
***Average*** 0.633 0.719 0.725
--------------- -------------- ------------ ------------
: Comparison of Algorithms
\[table\_comparison\_results\]
The results place the three algorithms in the same equivalence class using both the ANOVA and Student t-Tests. When the ANOVA test and the Student t-Tests are performed, the ANOVA test of the three algorithms yields a p-value of 3.35E-12, so the F-Test is then performed to determine which two-tailed two-sample Student t-Test to use. In each comparison between algorithms, the Student t-Test results in a t-Stat value that is smaller than the t Critical value. Therefore, the null hypothesis is accepted.
The results from each algorithm are shown in Table \[table\_comparison\_results\], while the ANOVA test results are shown in Tables \[table:ANOVA\_Single\_Factor\_Summary\] and \[table:ANOVA\_Single\_Factor\_Var\_Summary\].
-------------- ----------- --------- ------------- --------------
**** **** **** **** ****
**Groups** **Count** **Sum** **Average** **Variance**
***RBFSVM*** 30 19.08 0.636 0.00180
***LSVM*** 30 21.56 0.719 0.00237
***FFNN*** 30 21.76 0.725 0.00196
-------------- ----------- --------- ------------- --------------
: ANOVA Test Summary
\[table:ANOVA\_Single\_Factor\_Summary\]
--------------- -------- -------- -------- ------- ------------- ------------
**** **** **** **** **** **** ****
**SS** **df** **MS** **F** **P-value** **F crit**
***Between*** 0.149 2 0.074 36.35 3.35E-12 3.10
***Within*** 0.178 87 0.002
***Total*** 0.326 89
--------------- -------- -------- -------- ------- ------------- ------------
: ANOVA Test Variation Summary
\[table:ANOVA\_Single\_Factor\_Var\_Summary\]
Representative Student t-Tests are shown in Tables \[table:tTest\_RS\_SS\] and \[table:tTest\_SS\_SA\].
---------------------------------- ------------ ----------
**** **** ****
**** **RBFSVM** **LSVM**
**Mean** 0.636 0.718
**Variance** 0.00180 0.00237
**Observations** 30 30
**Pooled Variance** 0.00209
**Hypothesized Mean Difference** 0
**df** 58
**t-Stat** -7.008
**P(T${<=}$t) one-tail** 1.42E-09
**t Critical one-tail** 1.671
**P(T${<=}$t) two-tail** 2.84E-09
**t Critical two-tail** 2.002
---------------------------------- ------------ ----------
: Student t-Test: Two-Sample Assuming Equal Variances
\[table:tTest\_RS\_SS\]
---------------------------------- ---------- ----------
**** **** ****
**** **LSVM** **FFNN**
**Mean** 0.719 0.725
**Variance** 0.00237 0.00196
**Observations** 30 30
**Hypothesized Mean Difference** 0
**df** 57
**t-Stat** -0.555
**P(T${<=}$t) one-tail** 0.290
**t Critical one-tail** 1.672
**P(T${<=}$t) two-tail** 0.581
**t Critical two-tail** 2.002
---------------------------------- ---------- ----------
: Student t-Test: Two-Sample Assuming Unequal Variances
\[table:tTest\_SS\_SA\]
There are $2$ equivalence classes. The LSVM with Normalization, Standardization, and TFIDF and the FFNN with Normalization, Standardization, and TFIDF are in the same equivalence class, while the RBSVM with Normalization, Standardization, and TFIDF is in a different equivalence class.
The resulting accuracies are shown in Table \[table\_variations\_comparison\_results\].
**Algorithm** ***RBFSVM*** ***LSVM*** ***FFNN***
--------------------------------------- -------------- ------------ ------------
None 0.54 0.32 0.42
Normalization, Standardization, TFIDF 0.7 0.62 0.682
Standardization, TFIDF 0.51 0.52 0.61
Normalization, TFIDF 0.46 0.48 0.601
Normalization, Standardization 0.7 062 0.674
Normalization 0.59 0.29 0.44
Standardization 0.51 0.52 0.65
TFIDF 0.42 0.47 0.6
: Effects of Normalization, Standardization, and TFIDF on Algorithm Accuracy
\[table\_variations\_comparison\_results\]
The results of the statistical analysis of a comparison of all the algorithms is shown in Table \[table\_comparison\_results\].
[|c|c|c|c|c|]{} ****&\
**Algorithm** & ***Accuracy*** & ***Precision*** & ***Recall*** & ***F1***\
GRNN & 0.4315 & 0.3693 & 0.2457 & 0.2571\
EGRNN & 0.4328 & 0.4163 & 0.3682 & 0.3369\
RBFNN$^{\mathrm{1}}$& 0.2367 & 0.3454 & 0.2335 & 0.2342\
RBFNN$^{\mathrm{2}}$& 0.2735 & 0.4245 & 0.3253 & 0.3726\
SVM$^{\mathrm{3}}$ & 0.4342 & 0.4137 & 0.4577 & 0.4643\
SVM$^{\mathrm{4}}$ & 0.3172 & 0.2845 & 0.3435 & 0.2281\
FFNN & 0.1293 & 0.1346 & 0.3946 & 0.2986\
\
\
\
\
\[table\_comparison\_results\]
Radial Basis Function Support Vector Machine
--------------------------------------------
The effect of not using TFIDF is minimal. The accuracies produced are, on average, exactly the same as the accuracies using the TFIDF. The effect of not using Standardization is an average drop in accuracy of about 10%. If a dataset is not standardized before it is used and does not naturally have a normal distribution, then the algorithm is going to perform badly. The effect of not using Normalization is not a large change in accuracy values. The average for 5 runs is approximately the same with and without normalization. Since the accuracies cannot be negative values, the effect of normalization on all three of the algorithms is minimal. If the dataset produced negative values, the effect would be more noticeable.
Linear Support Vector Machine
-----------------------------
The effect of not using TFIDF is basically no change to the accuracy values. The accuracies are within the same range when TFIDF is used or is not used. The effect of not using Standardization is also an average drop in accuracy of about 10%. The StandardScalar in scikit-learn removes the mean and scales the data to unit variance. This still allows for outliers to influence the computation, and therefore does not guarantee balanced feature scales in the cases of having outliers. The effect of not using Normalization is a drop in accuracy of about 10%. The LSVM is the most affected by not using normalized data.
Feedforward Neural Network
--------------------------
The effect of not using TFIDF also results in no change to the accuracy values. The effect of not using Standardization is a convergence warning. The FFNN has a maximum iteration value of 2000, and when the data is not standardized, the optimization cannot converge in the 2000 iterations. The effect of not using Normalization is minimal. It results in a drop of about 5%, at the most.
Conclusions
===========
The RBFSVM, LSVM, and FFNN are all affected in similar ways when TFIDF, Standardization, and Normalization are not used. It is evident from the ANOVA test and Student t-Test that the LSVM and FFNN are in the same equivalence class. The FFNN had the highest average accuracy for 30 function evaluations, with an average accuracy of 72.5%. The ANOVA Test gives a P-value ${< 0.05}$, which both suggests that there is no statistical significant difference and has weak acceptance of the hypothesis. This means that all the three models are in different equivalence classes. The Student t-Test values for all combinations of RBFSVMs, LSVMs and FFNNs have t-Stat value ${< -0.05}$ which suggests that the null hypothesis is accepted for each combination.
Particle Swarm Optimization (PSO) is an option for an algorithm that may be used instead a GA and may be explored in future work.
Breakdown of the Work
=====================
Alison Jenkins - Research, Code, Analysis, [LaTeX]{} Report, and Extra Credit.
[00]{} Casey, Kenan, et. al. *An Evolutionary Approach for Achieving Scalability with General Regression Neural Networks*. Auburn University, 2019. Dozier, J. *Computational Intelligence and Adversarial Machine Learning: Genetic Algorithms*. Powerpoint Presentation, COMP6970 - Computational Intelligence and Adversarial Machine Learning class, Auburn University, 2019. Dozier, J. *Computational Intelligence and Adversarial Machine Learning: GRNNs, RBFNs, SVMs, NNs*. Powerpoint Presentation, COMP6970 - Computational Intelligence and Adversarial Machine Learning class, Auburn University, 2019. Dozier, J. *Computational Intelligence and Adversarial Machine Learning: Instance-Based Machine Learning*. Powerpoint Presentation, COMP6970 - Computational Intelligence and Adversarial Machine Learning class, Auburn University, 2019. Dozier, G. *Computational Intelligence and Adversarial Machine Learning: Particle Swarm Optimization*. Powerpoint Presentation, COMP6970 - Computational Intelligence and Adversarial Machine Learning class, Auburn University, 2019. Engelbrecht, Andries P. *Computational Intelligence: An Introduction*. John Wiley & Sons, 2007. Haruna Matsushita, Yoshifumi Nishio and Chi K. Tse: *Network-Structured Particle Swarm Optimizer That Considers Neighborhood Distances and Behaviors*. Journal of Signal Processing, Vol.18, No.6, pp 291-292, 2014. Joseph, Anthony D., et al. *Adversarial Machine Learning*. Cambridge University Press, 2018. Nguyen, Tam N. *Attacking Machine Learning Models as Part of a Cyber Kill Chain*. 2017. Sarkar, Dipanjan. *Text Analytics with Python*. Apress, 2016.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We consider one-dimensional random boolean cellular automata, i.e., the cells are identified with the integers from 1 to $N$. The behavior of the automaton is mainly determined by the support of the random variable that selects one of the sixteen possible Boolean rules, independently for each cell. A cell is said to stabilize if it will not change its state anymore after some time. We classify the one-dimensional random boolean automata according to the positivity of their probability of stabilization. Here is an example of a consequence of our results: if the support contains at least 5 rules, then asymptotically as $N\rightarrow \infty$ the probability of stabilization is positive, whereas there exist random boolean cellular automata with 4 rules in their support for which this probability tends to 0.'
address:
- |
Delft Institute of Applied Mathematics, Technical University of Delft, Mekelweg 4, 2628 CD Delft, The Netherlands\
email: [email protected]
- |
Eurandom, Eindhoven, The Netherlands\
email: [email protected]
author:
- 'F. Michel Dekking, Leonard van Driel and Anne Fey'
bibliography:
- 'RBCA.bib'
nocite: '[@*]'
title: Stability in random Boolean cellular automata on the integer lattice
---
Introduction
============
We study the dynamics of cellular automata as amply discussed in Stephen Wolfram’s book “A New Kind of Science" ([@NKS]). The cells, which are identified with the integers from 1 to $N$, can be in two possible states 0 or 1, and the state at the next time instant is a function of the states of the two neighbouring cells. For this function, also called (evolution) rule, there are 16 different possibilities, see also Table \[tab-bca-phi\]. Although there is a chapter in [@NKS] on *random* initial conditions for CA’s, there is hardly anything on *random* rules, or even on the possibility to assign different rules to different cells (so called *inhomogeneous* CA’s). In this paper we will give a classification of one-dimensional Random Boolean CA’s.
Our results are also of interest in the context of Random Boolean Networks introduced by Stuart Kauffman in 1969 ([@Kauffman1]). Here not only the rules and the initial condition are chosen randomly, but also the neighbours of a cell, i.e., the $N$ cells obtain a random graph structure. For each cell $i$ independently a mother $M(i)$ and a father $F(i)$ are chosen uniformly from the remaining pairs of cells. This gives a graph with vertices $i$ and directed edges $(M(i),i)$, $(F(i),i)$ for $i=1,\dots,N$. Next, each cell $i$ is assigned independently a Boolean rule $\Phi_i$ uniformly from the set of 16 Boolean rules (i.e., maps from $\{0,1\}^2$ to $\{0,1\}$, see also Table \[tab-bca-phi\]). To start, all cells obtain independently an initial value $X_i(0)$ from the set $\{0,1\}$ with equal probability. The state of a RBN at time $t$ will be denoted by ${X}(t)$. Then the dynamics is given by the map ${\Psi_{\!\Phi}}$: $${X}(t+1)={\Psi_{\!\Phi}}({X}(t))$$ defined by $$X_i(t+1) = \Phi_i(X_{M(i)}(t),X_{F(i)}(t))$$ for all cells $i$.
A $1$-dimensional random Boolean cellular automaton can be considered as an RBN with $$M(i)\equiv i-1\quad\mathrm{ and } \quad F(i)\equiv i+1,$$ for $i\!=\!2,\dots,N\!-\!1$. We put $M(1)\!=\!N, F(1)\!=\!2$, and $M(N)\!=\!N-1, F(N)\!=\!1$. In Figure \[fig-intro\] we display a realization of the evolution of a RBCA in the usual way: zero’s and ones are coded by white and black squares, space is in the horizontal direction, time in the vertical direction.
\[h\]
\[fig-intro\]
We say a cell *stabilizes* if its value is not changed anymore after some time. In Figure \[fig-intro\] the stable cells yield the black or white stripes; in this RBCA example about 56% of the cells stabilize.
Kauffman ([@MR762660]) found in his simulations that about 60-80% of the cells in a two-connected RBN stabilize. It was later shown rigorously by Luczak and Cohen ([@luczak:1991]) that in the limit as the number of cells $N$ tends to infinity 100% of the cells will stabilize—size $N=10\,000$ in Kauffmans simulations is much too small to unearth this behavior.\
For the RBN’s there is no natural model for the infinite size ($N\rightarrow\infty$) network. For a one-dimensional CA there is: the graph is simply the set of all integers with edges between nearest neighbors. It therefore makes sense to speak of the probability $\sigma_*$ of stabilization of the infinite automaton. The result of Luczak and Cohen leads to the question whether this $\sigma_*=1$. In this paper we will show that this is rarely the case, and moreover that there are many cases where actually $\sigma_*=0$. By different cases we mean different distributions on the set of 16 Boolean rules. We remark that the results in [@MR762660] and [@luczak:1991] are restricted to the uniform distribution on the set of rules—see [@Drossel-2007; @Lynch-1993; @Lynch-1993-2; @Lynch-PhD] for some results for RBN’s with other rule distributions.\
The second problem we will discuss is whether $\sigma_N\rightarrow\sigma_*$ when $N\rightarrow\infty$. We will show that this will often be the case, but may also fail to hold.\
For some previous work with rather partial results on 2–dimensional versions of RBCA’s see e.g., [@fogelman]. We will give an application of our results at the end of Section \[imp\] to a recent paper on RBN’s of Greil and Drossel ([@Drossel-2007]).
Finally we mention that our interest in RBCA was also motivated by certain CA models of the subsurface in the geosciences (see [@geo]).
Random Boolean cellular automata {#HD}
================================
Before we define RBCA, we describe their realizations which are inhomogeneous cellular automata. The dynamics in such a CA is defined by attaching a Boolean rule to each cell. The rule is given by specifying what its value is on the four different values $(1,1)$, $(1,0)$, $(0,1)$ and $(0,0)$ of the neighbors of a cell. In Table \[tab-bca-phi\], all 16 different rules are defined.
${\mathtt{j}}$ $\!{\varphi_{\mathtt{j}}}(0,0)\!$ ${\varphi_{\mathtt{j}}}(0,1)$ ${\varphi_{\mathtt{j}}}(1,0)$ ${\varphi_{\mathtt{j}}}(1,1)$ affine
---------------- ----------------------------------- ------------------------------- ------------------------------- ------------------------------- ---------
0 0 0 0 0 $0$
1 1 0 0 0 no
2 0 1 0 0 no
3 1 1 0 0 $x+1$
4 0 0 1 0 no
5 1 0 1 0 $y+1$
6 0 1 1 0 $x+y$
7 1 1 1 0 no
8 0 0 0 1 no
9 1 0 0 1 $x+y+1$
10 0 1 0 1 $y$
11 1 1 0 1 no
12 0 0 1 1 $ x$
13 1 0 1 1 no
14 0 1 1 1 no
15 1 1 1 1 $1$
: Definition of the 16 Boolean rules. The last column gives a formula for ${\varphi_{\mathtt{j}}}(x,y)$ when ${\varphi_{\mathtt{j}}}$ is affine.
\[tab-bca-phi\]
For ease of notation these are indexed as ${\varphi_{\mathtt{j}}}$ where ${\mathtt{j}}$ is the number in binary notation obtained from the four bits in the row describing ${\varphi_{\mathtt{j}}}$. Anytime we refer to one of these 16 rules we use index ${\mathtt{j}}$ (note the different font); when we use index $i$ we refer abusively to the rule assigned to cell $i$.
An inhomogeneous CA’s mapping ${\Psi_{\varphi}}: \{0,1\}^N\to \{0,1\}^N$ is determined by a vector of rules, ${\varphi}=(\varphi_1, \ldots,
\varphi_N)$. Here for all time steps, $\varphi_i$ is the rule of cell $i$. Hence given a state ${x}= (x_1 \ldots x_N)$ we get $({\Psi_{\varphi}}({x}))_i = \varphi_i(x_{i-1},x_{i+1})$ for $i=2,\dots, N-1$. To deal with the problem that occurs in the end points we consider the automata as defined on a circle; a convenient way to express this is to define $x_{i+kN}=x_i$ for $i=1,\dots, N$, and all integers $k$. To address ${\Psi_{\varphi}}$ we will also write the vector of lower indices of the $\varphi_i$ $(1 \leq i \leq N)$ mentioned. E.g., mapping (0,3,3,15) describes ${\Psi_{\varphi}}:\{0,1\}^4\to\{0,1\}^4$ with rule vector $({\varphi_{\mathtt{0}}},{\varphi_{\mathtt{3}}},{\varphi_{\mathtt{3}}},{\varphi_{\mathtt{15}}})$.\
In a RBCA the rules $\varphi_i$ are chosen according to a distribution $\Phi$ on the set of rules, independently for each $i$. The initial state is an i.i.d. sequence $(X_i(0))$ from a Bernoulli(1/2) distribution, independent from the $(\Phi_i)$. Then the value of the $i^\textrm{th}$ cell at time $t+1$ is given by $$X_i(t+1)=\Phi_i(X_{i-1}(t),X_{i+1}(t)).$$ We denote by ${{\mathbb P\!}_N}$ the probability measure generated by the ($X_i(0)$) and the ($\Phi_i$) on the space of infinite periodic sequences with period $N$. It is natural to consider also the probability measure ${{\mathbb P\!}_*}$ on $\{0,1\}^{\mathbb{Z}}$ generated by two infinite independent i.i.d. sequences ($X_i(0)$) and ($\Phi_i$).
Stability {#stab}
=========
We say cell $i$ *stabilizes at time* $T$ if $$X_i(t)=X_i(T) \quad \text{for all } t\ge T,$$ and $T$ is the smallest integer with this property. Let $\sigma_N$ be the probability that a cell in a random cellular automaton of size $N$ with rule distribution $\Phi$ stabilizes. Because of stationarity, $$\sigma_N={{\mathbb P\!}_N\!\left({\text{cell } 0 \text{ stabilizes}}\right)}.$$
We first give an example were $\sigma_N$ does not converge as $N\rightarrow\infty$.\
Consider the RBCA’s for $N=1,2\dots$ with rule distribution given by $${{\mathbb P\!}_N\!\left({\Phi={\varphi_{\mathtt{6}}}}\right)}=1.$$ Note that randomness is only involved because of the randomness in the initial state ${X}(0)=(X_i(0))$, and that ${\varphi_{\mathtt{6}}}$ is the well known linear rule which can be written as ${\varphi_{\mathtt{6}}}(x,y)=x+y$. This addition, as all additions with 0’s and 1’s, is modulo 2.
\[prop:discontin\] For the ${\varphi_{\mathtt{6}}}$-RBCA one has $$\liminf_{N\rightarrow\infty} \sigma_N=0,
\qquad \limsup_{N\rightarrow\infty} \sigma_N =1.$$
*Proof:* It is well known (see e.g. [@algebra]) and easy to prove that the ${\varphi_{\mathtt{6}}}$-automaton satisfies $$X_i(t)=X_{i-t}(0)+X_{i+t}(0) \qquad \text{ if } t=2^p \text{ for
some } p.$$ It follows that $\sigma_N=1$ for each $N=2^p$ which is a power of 2: $$X_i(N)=X_{-N+i}(0)+X_{N+i}(0)=2X_i(0)=0 \qquad \text{for all } i,$$ which implies that $X_i(t)=0$ for all $t\ge N$. This proves that $\limsup \sigma_N =1$. To obtain the statement on $\liminf$, consider the subsequence of $N$ of the form $N=2^p+1$. Then we find for all $i$ $$X_i(N-1)=X_i(2^p)=X_{-2^p+i}(0)+X_{2^p+i}(0)=X_{i+1}(0)+X_{i-1}(0).$$ From this we directly obtain that $$X_0(N)=X_{-1}(N-1)+X_{1}(N-1)=X_{-2}(0)+X_{0}(0)+X_{0}(0)+X_{2}(0),$$ and so $$\label{even}
X_0(N)=X_{-2}(0)+X_{2}(0).$$ Thus $${{\mathbb P\!}_N\!\left({X_0(N-1)\ne X_0(N)}\right)}={{\mathbb P\!}_N\!\left({X_{-1}(0)+X_{1}(0)\ne X_{-2}(0)+X_{2}(0)}\right)}=\frac12.$$ Note that this computation does not only hold for $t=N-1$, but for any $t$ which satisfies $t=-1 \mod N$. An infinite sequence of such $t$’s is given by $t=2^{(2k+1)p}, k=0,1,2,\dots$. For $N$ of the form $2^p+1$ we therefore obtain that $$1-\sigma_N={{\mathbb P\!}_N\!\left({X_0(t)\ne X_0(t+1)\quad \text{for infinitely many } t}\right)}\ge\frac12.$$ This inequality is a warm up for a more general result: we can generalize Equation (\[even\]) to for all $m$ $$X_0(N+2^m-2)=X_{-2^m}(0)+X_{2^m}(0).$$ Since for $m=0,1,\dots,\ell$ the $X_{-2^m}(0)+X_{2^m}(0)$ are independent Bernoulli variables with success probability $1/2$, it follows that $${{\mathbb P\!}_N\!\left({\exists m, 0\le m < \ell: X_0(N+2^m-2)\ne X_0(N-1)}\right)}=1-\frac1{2^\ell},$$ and employing the same infinite sequence of $t$’s as for the $\ell=1$ case, we can deduce that $ \liminf_{N\rightarrow\infty} \sigma_N=0$.$\Box$\
It is convenient to define random variables $Z_i$ by $Z_i=1$ if cell $i$ stabilizes, and 0 otherwise. We then have (independently of $i$): $$\sigma_N={{\mathbb P\!}_N\!\left({Z_i=1}\right)}, \quad \text{and} \quad \sigma_*={{\mathbb P\!}_*\!\left({Z_i=1}\right)}.$$ The *uniform* RBCA is given by the rule distribution ${{\mathbb P\!}_N\!\left({\Phi={\varphi_{\mathtt{j}}}}\right)}=1/16$ for $j=0,1,\ldots,15$.
\[unif\] Let $\sigma_N$ be the probability of stabilization of a size $N$ uniform RBCA, $\sigma_*$ that of an infinite size uniform RBCA. Then $$\lim_{N\rightarrow\infty}\sigma_N=\sigma_*\quad \textrm{and}\quad \sigma_*>0.$$
*Proof:* The two rules ${\varphi_{\mathtt{0}}}$ and ${\varphi_{\mathtt{15}}}$ are called *walls*. We denote $W=\{{\varphi_{\mathtt{0}}}$,${\varphi_{\mathtt{15}}}\}$. Cells that have obtained a rule from $W$ are already stable at time 0 or 1. We define for positive integers $k$ and $l$ the *chambers* $C_{k,l}$ by $$C_{k,l}=\{\Phi_{-k}\in W, \Phi_{-k+1}\notin W, \Phi_{-k+2}\notin W,\ldots,\Phi_{\l-1}\notin W,\Phi_l\in W\}.$$ In addition we put $C_{0,0}=\{\Phi_0\in W\}$. Since the chambers are disjoint events whose union has probability 1 we can write $$\begin{aligned}
\sigma_*={{\mathbb P\!}_*\!\left({Z_0=1}\right)}={\ensuremath{{\mathbb P\!}_*\!}\left( Z_0=1 \, |\, C_{0,0}\right)}{{\mathbb P\!}_*\!\left({C_{0,0}}\right)}+
\sum_{k,l=1}^{\infty}{\ensuremath{{\mathbb P\!}_*\!}\left( Z_0=1 \, |\, C_{k,l}\right)}{{\mathbb P\!}_*\!\left({C_{k,l}}\right)}.\end{aligned}$$ We will make a similar splitting for finite RBCA of size $N$. In case $N=2M+1$ we consider the indices of the chambers from the set $\{-M,\dots,0,\dots,M\}$, in case $N=2M$ from $\{-M\!+\!1,\dots,0,\dots,M\}$. In the sequel we will assume $N$ is odd, as our arguments can be transferred trivially to the case $N$ even. Let $D$ be the event $$D=\Big(C_{0,0}\cup\bigcup_{k,l=1}^M C_{k,l}\Big)^c.$$ Then $D$ occurs iff there is no wall among $\Phi_{-M}, \dots, \Phi_0$ or among $\Phi_{0}, \dots, \Phi_M$. Since ${{\mathbb P\!}_N\!\left({\Phi_i\notin W}\right)}={{\mathbb P\!}_*\!\left({\Phi_i\notin W}\right)}=7/8$, we obtain that $${{\mathbb P\!}_*\!\left({D}\right)}={{\mathbb P\!}_N\!\left({D}\right)}=2\Big(\frac78\Big)^{M+1}\!\!-\Big(\frac78\Big)^{2M+1}
\le2\Big(\frac78\Big)^{N/2}.$$ The crux of the proof is that for all $1\le k,l\le M$ and for $k=l=0$ we have $${\ensuremath{{\mathbb P\!}_N\!}\left( Z_0=1 \, |\, C_{k,l}\right)}={\ensuremath{{\mathbb P\!}_*\!}\left( Z_0=1 \, |\, C_{k,l}\right)},$$ simply because both the finite automaton and the infinite automaton evolve between the walls independently of the evolution outside the walls. Thus $$\begin{aligned}
|\sigma_N-\sigma_*|&\le & {\ensuremath{{\mathbb P\!}_*\!}\left( Z_0=1 \, |\, D\right)}{{\mathbb P\!}_*\!\left({D}\right)}+{\ensuremath{{\mathbb P\!}_N\!}\left( Z_0=1 \, |\, D\right)}{{\mathbb P\!}_N\!\left({D}\right)}\\
&\le & 4\Big(\frac78\Big)^{N/2}.\end{aligned}$$ It follows that $\sigma_N \rightarrow \sigma_*$ exponentially fast. Note that $$\sigma_N\ge{{\mathbb P\!}_N\!\left({\Phi\in W}\right)}=\frac18,$$ and so $\sigma_*\ge 1/8>0$. $\Box$\
Actually we can obtain a good estimate of $\sigma_*$ using the exponential convergence of $\sigma_N$, by sampling uniform RBCA’s for some large $N$ and computing a 95% confidence interval. We found that $\sigma_{100}=0.678058\pm 0.000029$. Interestingly, an exhaustive enumeration for $N=6$ yields already a value $\sigma_6$ close to 0.68.
We end this section with the remark that the ‘wall’-property directly implies that the sequence $(Z_i)$ is strongly mixing under ${{\mathbb P\!}_*}$: $${{\mathbb P\!}_*\!\left({Z_0={\varepsilon_{0}}, Z_i={\varepsilon_{1}}}\right)}\rightarrow {{\mathbb P\!}_*\!\left({Z_0={\varepsilon_{0}}}\right)}{{\mathbb P\!}_*\!\left({Z_0={\varepsilon_{1}}}\right)}\quad \mathrm{as} \quad i\rightarrow\infty.$$ Therefore the ergodic theorem applies, yielding that we also have pathwise about 68% of stable cells.
Stability by means of impermeable blocks {#imp}
========================================
Note that the proof of Proposition \[unif\] will still be valid if $\Phi$ is not uniformly distributed, but has an arbitrary distribution with ${{\mathbb P\!}_N\!\left({\Phi={\varphi_{\mathtt{j}}}}\right)}>0$ for all ${\mathtt{j}}$ (of course $7/8$ has to be replaced by ${{\mathbb P\!}_N\!\left({\Phi={\varphi_{\mathtt{0}}}}\right)}+{{\mathbb P\!}_N\!\left({\Phi={\varphi_{\mathtt{15}}}}\right)}$). The only thing that matters is the *support* of $\Phi$ defined by $$\text{Supp}(\Phi)=\{{\mathtt{j}}: {{\mathbb P\!}_N\!\left({\Phi={\varphi_{\mathtt{j}}}}\right)}>0\}=\{{\mathtt{j}}: {{\mathbb P\!}_*\!\left({\Phi={\varphi_{\mathtt{j}}}}\right)}>0\}.$$ In this section we will generalize the concept of a wall. An easy adaptation of its proof will permit to draw the conclusions of Proposition \[unif\] for a large majority of the $2^{16}$ different supports.
Consider an inhomogeneous BCA. We call adjacent cells $\{i+1,
\ldots,i+p\}$ an *impermeable block* of size $p$, if there exist $b=(b_1,\ldots,b_p)\in\{0,1\}^p$, and $({\mathtt{j_1}},\ldots,{\mathtt{{\mathtt{j_p}}}})\in\{0,\ldots,15\}^p$ such that if $$(x_{i+1}(0),\ldots,x_{i+p}(0))=(b_1,\ldots,b_p)\quad \text{and}
\quad
(\varphi_{i+1},\ldots,\varphi_{i+p})=(\varphi_{{\mathtt{j_1}}},\ldots\varphi_{{\mathtt{j_p}}}),$$ then for all $t$ $$(x_{i+1}(t),\ldots,x_{i+p}(t)),$$ is unchanged, whatever values are chosen for $x_i(s)$ and $x_{i+p+1}(s)$, for $s=0,1\dots,t-1$. Here we assume that $N$ is larger than $p+2$.\
As an example take $p=2$, $b=(0,0)$ and $({\mathtt{j_1}},{\mathtt{j_2}})=(2,4)$.\
If a particular $\Phi$ has ${\varphi_{\mathtt{j}}}$’s in its support that yield an impermeable block, then any $\Phi$ having this set of ${\varphi_{\mathtt{j}}}$’s in its support will also have an impermeable block. We therefore look for minimal sets of ${\mathtt{j}}$’s such that the ${\varphi_{\mathtt{j}}}$’s give rise to impermeable blocks. Here is a listing of these sets: $$\begin{aligned}
\mathcal{G}=\big\{\{0\}, \{1\}, \{7\}, \{8\}, \{14\}, \{15\}, \{2,
4\}, \{2, 5\},
\{2, 9\}, \{2, 12\}, \{2, 13\},\quad\\
\{3, 4\},\{3, 5\}, \{3,10\}, \{3, 13\}, \{4, 9\}, \{4, 10\}, \{4, 11\}, \{5,11\},\quad\\
\{5,12\},\{6, 11\},\{6, 13\}, \{10, 12\}, \{10, 13\}, \{11,12\}, \{11, 13\},\quad\\
\{2, 6, 10\}, \{4, 6, 12\}, \{9, 10, 11\},\{9, 12, 13\}\big\}.\end{aligned}$$ Because of the symmetries of the collection of sets of CA rules (cf. Section \[symm\]), of these 30 subsets there is a subcollection $\mathcal{\widetilde{G}}$ of only 12 where the dynamics of the associated ${\Psi_{\!\Phi}}$ is essentially different. Table \[imperm\] gives these supports with their $\varphi$-blocks $(\varphi_{{\mathtt{j_1}}},\ldots\varphi_{{\mathtt{j_p}}})$, and their $b$-blocks $(b_1,\ldots,b_p)$.
------------- ----------------- ----------- -------------- ----------------- ---------------
[\ $\varphi$-block $b$-block Support $\varphi$-block $b$-block
]{} Support
[\ (0) (0) $\{2,12\}$ (2,12) (0,0)
\
]{} $\{0\}$
$\{1\}$ (1,1,1) (0,1,0) $\{2, 13\}$ (13, 2) (1,0)
$\{8\}$ (8,8) (0,0) $\{3, 5\}$ (5,3) (0,0)
$\{2,4\}$ (2,4) (0,0) $\{3,10\}$ (10,3) (0,0)
$\{2,5\}$ (5,2) (1,0) $\{10,12\}$ (10,12) (0,0)
$\{2,9\}$ (2,9,9,2) (0,0,1,0) $\{2,6,10\}$ (10,6,2) (1,1,0) [\
]{}\[imperm\]
------------- ----------------- ----------- -------------- ----------------- ---------------
: Impermeable blocks.
Often there is more than one $b$-block that will work, in the table we have chosen $b$-blocks with minimal length. It can be quickly verified that the entries in Table \[imperm\] yield impermeable blocks. See Appendix \[imp-ver\] for two examples of the type of verification one has to perform.
We can describe the supports which do *not* give rise to impermeable blocks by the list $$\begin{aligned}
\mathcal{B}=\big\{ \{2, 3, 6\}, \{2, 3, 11\}, \{2, 10, 11\},\{3,
9, 11\}, \{4, 5, 6\}, \{4, 5, 13\}, \quad\\ \hspace*{5cm} \{4,
12, 13\}, \{5, 9, 13\}, \{3,6, 9, 12\},
\{5,6, 9, 10\} \big\}.\end{aligned}$$ Any support which does not admit an impermeable block is a subset of a set in $\mathcal{B}$. For an example of a proof that with these supports no impermeable blocks occur, see Appendix \[no-imp\].
One can check that any subset of $\{0,1,\ldots,15\}$ is either a subset of a set from $\mathcal{B}$, or contains a set from $\mathcal{G}$, and so we have completely described the supports with regard to the existence of impermeable blocks.
[**Example**]{}. In [@Drossel-2007] the two-connected RBN with solely rule ${\varphi_{\mathtt{7}}}$ is studied, and it is claimed that “every node oscillates with period 2". Since ${\varphi_{\mathtt{7}}}$ is mirror equivalent to ${\varphi_{\mathtt{1}}}$, we see that actually the RBCA with support only consisting of ${\varphi_{\mathtt{7}}}$, has absorbing block (1,0,1). Let $i$ be a node in the ${\varphi_{\mathtt{7}}}$-RBN with parents $j$ and $k$. Then, if $i$ happens to be one of the parents of both $j$ and $k$ (figure 8 configuration), and the state of $i$ is 0, and that of $j$ and $k$ is 1, then these three nodes keep that state forever. This gives that the expected number of stable nodes in the RBN is at least $$N\cdot {{\mathbb P\!}_N\!\left({\mathrm{figure\; 8\; configuration}}\right)}\cdot{{\mathbb P\!}_N\!\left({\mathrm{initial\; values\; 1,0,1}}\right)}=N\left(\frac2{N-1}\right)^2.\frac18.$$
Classifying RBCA’s {#class}
==================
In the previous section we have established that all the RBCA’s whose support contains a set in $\mathcal{G}$ are well behaved in the sense that they are stability continuous ($\sigma_N\rightarrow\sigma_*$), and do have a positive probability to stabilize ($\sigma_*>0$). As there are only 53 supports described by $\mathcal{B}$ this means that this regular behaviour manifests itself for at least 99,9% of the RBCA’s. The following result gives a complete classification.
\[main\] An infinite one-dimensional random Boolean cellular automaton has no stable cells, i.e., $\sigma_*=0$, if and only if the support ${\mathrm{Supp}{(\Phi)}}$ of the rule distribution $\Phi$ is a subset of $\{{\varphi_{\mathtt{2}}}, {\varphi_{\mathtt{10}}}\}, \{{\varphi_{\mathtt{10}}}, {\varphi_{\mathtt{11}}}\}, \{{\varphi_{\mathtt{4}}}, {\varphi_{\mathtt{12}}}\}, \{{\varphi_{\mathtt{12}}}, {\varphi_{\mathtt{13}}}\},
\{ {\varphi_{\mathtt{3}}}, {\varphi_{\mathtt{6}}}, {\varphi_{\mathtt{9}}}, {\varphi_{\mathtt{12}}}\}$ or $\{ {\varphi_{\mathtt{5}}}, {\varphi_{\mathtt{6}}}, {\varphi_{\mathtt{9}}}, {\varphi_{\mathtt{10}}}\}$.
*Proof:* From the results of the previous section it is clear that any CA with $\sigma_*=0$ must have its support in (the subsets of ) collection $\mathcal{B}$. Because of symmetries it is enough to consider the 18 supports which occur as subsets of the sets in the collection $\mathcal{\widetilde{B}}$ given by $$\begin{aligned}
\mathcal{\widetilde{B}}=\big\{\{2, 3, 6\}, \{2, 3, 11\}, \{2, 10,
11\}, \{3,6, 9, 12\}\big\}.\end{aligned}$$ In Section \[linbadguys\] we will prove that subsets of rules in $\{3,6,9, 12\}$ all give supports that belong to CA with $\sigma_*=0$.
It occurs that the remaining supports can be split into two parts. The first part consists of the supports which occur as subsets of $\mathcal{\widetilde{B}}$ given by $$\begin{aligned}
\mathcal{\widetilde{S}}=\big\{\{2, 3\}, \{2,6\},\{2, 11\}, \{2, 3,6\},\{2,
3,11\},\{2,10,11\}\big\}.\end{aligned}$$ The $\widetilde{}$ here indicates that we removed those supports that are symmetry equivalent to supports that are already listed, as, for example, $\{3,11\}$. The supports in $\mathcal{\widetilde{S}}$ will be shown in Section \[absorb\] to be of the same regular type as found in Section \[imp\]. The second part consists of the two supports $$\{2\} \textrm{ and} \; \{2,10\}.$$ Here the automata behave very similarly.
\[fig-2-10\]
First note that ${\varphi_{\mathtt{10}}}$ is simply the left shift. The crucial fact is that $${\varphi_{\mathtt{2}}}(x,y)={\varphi_{\mathtt{10}}}(x,y) \quad\mathrm{if}\quad (x,y)\ne (1,1).$$ First we consider the case ${\mathrm{Supp}{(\Phi)}}=\{{\varphi_{\mathtt{2}}}\}$. Then ${\varphi_{\mathtt{2}}}$ acts as the left shift for all times $t\ge 1$, since after one application of ${\varphi_{\mathtt{2}}}$ there will be no blocks $1 \ast 1$ (i.e., 101 or 111) left anymore, since in the initial sequence $0\ast 1$ cannot overlap with itself with lag 2. The map ${\varphi_{\mathtt{2}}}$ acts then as the shift on a subshift in which blocks of more than two ones can not occur. It follows that for each $i$ the process $(X_i(t)_{t\ge 1})$ is the same subshift, which clearly can not stabilize. (Actually it can be shown that this process is a hidden Markov chain on 8 equiprobable states.)
In case ${\mathrm{Supp}{(\Phi)}}=\{{\varphi_{\mathtt{2}}}, {\varphi_{\mathtt{10}}}\}$, the CA runs as a left shift with occasional transformations of 1’s into 0’s. This is a form of self-organization. In fact any block of 1’s that occurs at time 0 to the right of a cell $i$ with $\Phi_i={\varphi_{\mathtt{2}}}$ will be cut into blocks of 1 and 11 separated by 0’s. Since any cell $i$ has a cell $i'$ to the right of it with $\Phi_{i'}={\varphi_{\mathtt{2}}}$ with probability 1, there exists a (random) $t_0$, such that the process $(X_i(t))_{t\ge t_0}$ will only contain 1 and 11, and cannot stabilize. $\Box$
Stability by means of absorbing blocks {#absorb}
======================================
Here we show that we may define a weaker notion than that of an impermeable block which still permits us to conclude to continuity of stabilization and a positive probability to stabilize.\
Consider an inhomogeneous CA. We call adjacent cells $\{i+1,
\ldots,i+p\}$ an *absorbing block* of size $p$, if there exist $b$ in $\{0,1\}^p$, and $({\mathtt{j_1}},\ldots,{\mathtt{j_p}})$ in $\{0,\ldots,15\}^p$ such that if $$(x_{i+1}(0),\ldots,x_{i+p}(0))=(b_1,\ldots,b_p)\quad \text{and}
\quad
(\varphi_{i+1},\ldots,\varphi_{i+p})=(\varphi_{{\mathtt{j_1}}},\ldots\varphi_{{\mathtt{j_p}}}),$$ then for some $c$ with $1<c<p$ the value $x_{i+c}(t)$ is unchanged for all $t$, whatever values are chosen for $x_i(s)$ and $x_{i+p+1}(s),s=0,1\dots,t-1$. (Here we assume that $N$ is larger than $p+2$). The idea is that the evolution of one or both adjacent cells $i$ and $i+p+1$ may ‘penetrate’ the block, but will never influence the ‘central’ cell $i+c$.\
We give an example for an inhomogeneous automaton with rules ${\varphi_{\mathtt{2}}}$ and ${\varphi_{\mathtt{11}}}$. Let $p=4$, $b=(0,1,1,0)$, and $({\mathtt{j_1}},\ldots,{\mathtt{j_4}})=(2,2,11,2)$. Then this block is absorbing with $c=2$:
---------- -------------------------- -------------------------- --------------------------- -------------------------- ------------
cell $i$ cell $i+1$ cell $i+2$ cell $i+3$ cell $i+4$ cell $i+5$
${\varphi_{\mathtt{2}}}$ ${\varphi_{\mathtt{2}}}$ ${\varphi_{\mathtt{11}}}$ ${\varphi_{\mathtt{2}}}$
[\ 0 1 1 0 $y$
\
]{} $x$
$x'$ $1-x$ 1 0 0 $y'$
$x''$ $1-x'$ 0 0 $y'$ $y''$
$x'''$ 0 0 1 $y''$ $y'''$
0 1 1 0
---------- -------------------------- -------------------------- --------------------------- -------------------------- ------------
Note that at time $t=4$ the block $b=(0,1,1,0)$ reappears, so cell $i+2$ has a periodic evolution with period 4, independently of the evolution of cell $i$ and cell $i+5$. This establishes that $(2,2,11,2)$ is an absorbing block. Similarly it can be verified that $(2,2,3)$ is an absorbing block for the automata with rules ${\varphi_{\mathtt{2}}}$ and ${\varphi_{\mathtt{3}}}$.
For the support $\{2,6\}$, we present an absorbing block *family* with stable cells: take $\varphi$-block $$(2,2,2,6,6,6,2,2,2,6,6,6,2,2,2,2)$$ and $b$-block family $$(0,0,1,1,0,1,0,1,0,1,1,0,x,y,z,w),$$ where $x,y,z$ and $w$ can have any value as long as both $xz$ and $yw$ are not $11$. As a consequence of the ${\varphi_{\mathtt{2}}}$ rule, neither $xz$ nor $yw$ can be $11$ at any later time. One can then check, in a case by case analysis for all remaining possibilities for $xz$ and $yw$ (three possibilities each), that the $b$-block family keeps reappearing, and that the fourth and seventh cell are stable. We find that if $yw = 01$ or $yw = 10$, it takes 8 time steps for another member of the $b$-block family to reappear, for all other possibilities, this takes 4 time steps.
It follows from these observations that the rules with supports in the collection $\mathcal{\widetilde{S}}$ considered in the proof of Theorem \[main\] do not have stable cells.
We can not yet conclude that $\sigma_*>0$ for supports $\{{\varphi_{\mathtt{2}}},{\varphi_{\mathtt{3}}}\}$ and $\{{\varphi_{\mathtt{2}}},{\varphi_{\mathtt{11}}}\}$ , since in both cases the central cell has a period 4 dynamics. However, we can find larger absorbing blocks with a stable central cell $c$: for $\{2,3\}$ take $p=9$, $c=6$, $\varphi$-block: $(2,2,3,2,3,2,2,2,3)$ and $b$-block: $(0,0,1,1,0,0,0,0,1)$. For $\{2,11\}$ take $p=10$, $c=6$, $\varphi$-block: $(2,2,11,2,11,2,2,2,11,2)$ and $b$-block: $(0,1,1,0,0,0,0,1,1,0)$. The existence of such absorbing blocks with a stable cell implies that $\sigma_*>0$.
Affine chaos {#linbadguys}
============
Here we will analyze the RBCA’s whose support is a subset of $\{3,6,9,12\}$.
\[h\]
\[fig-36912\]
The crucial observation is that not only these rules are affine, but they can all be written as $${\varphi_{\mathtt{j}}}(x,y)=x+f(y,{\varphi_{\mathtt{j}}}),$$ where $f(y,{\varphi_{\mathtt{j}}})=1,\,y,\,y+\!1,\,0$ for ${\mathtt{j}}=3,6,9,12$ respectively. It follows from this that we can write for some function $f_t$ with $4t-2$ arguments $$\label{xprop}
x_i(t)=x_{i-t}(0)+f_t(x_{i-t+2}(0),\dots,x_{i+t}(0),\varphi_{i-t+1},\dots,\varphi_{i+t-1}).$$ Indeed, for $t=1$ we have $$x_i(1)=x_{i-1}(0)+f_1(x_{i+1}(0),\varphi_i),$$ where $f_1(y,\varphi_i)=f(y,\varphi_i )$ as above. For $t=2$ we obtain, applying this equation or its shift three times $$\begin{aligned}
x_i(2)&=&x_{i-1}(1)+f_1(x_{i+1}(1),\varphi_i)\\
&=&x_{i-2}(0)+f_1(x_{i}(0),\varphi_{i-1})+
f_1(x_{i}(0)+f_1(x_{i+2}(0),\varphi_{i+1}),\varphi_i).\end{aligned}$$ Defining $$\begin{aligned}
f_2(x_{i}(0),\dots,x_{i+2}(0),\varphi_{i-1},\dots,\varphi_{i+1})\hspace*{5cm}\\
\hfill = f_1(x_{i}(0),\varphi_{i-1})+f_1(x_{i}(0)+f_1(x_{i+2}(0),\varphi_{i+1}),\varphi_i),\end{aligned}$$ we obtain Equation (\[xprop\]) for $t=2$. Continuing in this fashion we obtain Equation (\[xprop\]) for all $t$.
The following proposition will immediately imply that $\sigma_*=0$ for the RBCA’s from this section.
Consider an RBCA with $\mathrm{Supp}(\Phi) \subset
\{{\varphi_{\mathtt{3}}},{\varphi_{\mathtt{6}}},{\varphi_{\mathtt{9}}},{\varphi_{\mathtt{12}}}\}$, and let $i$ be a fixed cell. Then the random variables $(X_i(t))_{t=0}^{t=\infty}$ are independent [*Ber$\mspace{2mu}(\frac12)$*]{} random variables under ${{\mathbb P\!}_*}$.
*Proof:* We prove this by induction w.r.t. the length $t$ of the cylinders. We will show that for each cell $i$ and for all $t=1,2,\dots$ $$\label{ber}
{{\mathbb P\!}_*\!\left({X_{i}(0)={\varepsilon_{0}},X_i(1)={\varepsilon_{1}},\dots,X_i(t-1)={\varepsilon_{t-1}}}\right)}=2^{-t}$$ for all $({\varepsilon_{0}},{\varepsilon_{1}},\dots,{\varepsilon_{t-1}})$ from$\{0,1\}^{t}$.
\[h\]
![\[fig-prop2\] Additive propagation.](Fig4_propagation.eps){height="8cm"}
Although in general the $(X_i(t))$ will not be stationary (any cell which stabilizes but not at $t=0$ has a transient evolution!), a simple computation shows that Equation (\[ber\]) implies that shifted cylinders of length $t$ also have probability $2^{-t}$ to occur.
We will use the following abbreviations: $$\begin{aligned}
Y_{i,t} &=& f_t(X_{i-t+2}(0),\dots,X_{i+t}(0),\Phi_{i-t+1},\dots,\Phi_{i+t-1})\\
E_{i,t,\varepsilon} &=& \{X_{i}(0)={\varepsilon_{0}},X_i(1)={\varepsilon_{1}},\dots,X_i(t-1)={\varepsilon_{t-1}}\}.\end{aligned}$$ First note that Equation (\[ber\]) is true for $t=1$ by definition of the RBCA. Figure \[fig-prop2\] is useful in the remainder of the proof.
Suppose that Equation (\[ber\]), i.e., ${{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon}}\right)}=2^{-t}$ has been proved for cylinders of length $t$. Then for all ${\varepsilon_{0}},\dots,{\varepsilon_{t-1}}$ from$\{0,1\}^{t}$ $$\begin{aligned}
&&{{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon},X_i(t)=1}\right)}=\\
&&{{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon},X_{i-t}(0)=0,Y_{i,t}=1}\right)}+
{{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon},X_{i-t}(0)=1,Y_{i,t}=0}\right)}=\\
&&{{\mathbb P\!}_*\!\left({X_{i-t}(0)=0}\right)}{{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon},Y_{i,t}=1}\right)}+
{{\mathbb P\!}_*\!\left({X_{i-t}(0)=1}\right)}{{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon},Y_{i,t}=0}\right)}=\\
&&\tfrac12\big({{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon},Y_{i,t}=1}\right)}+
{{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon},Y_{i,t}=0}\right)}\big)=\\
&&\tfrac12{{\mathbb P\!}_*\!\left({E_{i,t,\varepsilon}}\right)}=2^{-(t+1)},\end{aligned}$$ where we used that $X_{i-t}(0)$ is independent of $E_{i,t,\varepsilon}$ and of $Y_{i,t}$. This equality proves Equation (\[ber\]) for length $t+1$ cylinders.\
What remains for the automata in this section is the question whether $\sigma_N \rightarrow \sigma_{*}$. For some automata this is easy to answer. We have already seen that for support $\{6\}$ there is *no* stability continuity (see Proposition \[prop:discontin\]). By symmetry, the same is true for support $\{9\}$ and with a little more work this can also be shown for support $\{6, 9\}$. On the other hand it is easy to show that $\sigma_{2N+1}=0$ and $\sigma_{2N}=2^{-N}$ for $\mathrm{Supp}(\Phi)=\{3\}$ and that $\sigma_N=2^{-N+1}$ for $\mathrm{Supp}(\Phi)=\{12\}$, hence we do have stability continuity in these cases. Also for the combined support $\{3, 12\}$ it is easily shown that $\sigma_N\le
2^{-N+1}$, hence here too there is stability continuity. However, for the automata with supports $\{3,6\}$, $\{6,12\}$, $\{3,6,9\}$, $\{3,6,12\}$, $\{6,9,12\}$, and $\{3,6,9,12\}$, the computation of $\sigma_N$ becomes quite involved. We conjecture that there is stability continuity for these cases.
Conclusion
==========
We have studied the behavior of one-dimensional random Boolean cellular automata with two inputs. Although this behavior can be quite diverse, we have shown that it does not depend on the probabilities with which the Boolean rules $\varphi_{{\mathtt{0}}},\dots,\varphi_{{\mathtt{15}}}$ are attached to the integers, but only on the support of this random variable $\Phi$, i.e. on the probabilities ${\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{j}}} \right)}}$ being 0 or positive. This contrasts with the behavior of random Boolean networks as determined by Lynch ([@Lynch-PhD; @Lynch-2006]), see also [@Drossel-2007]. His result is that there is ordered behavior as long as $${\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{0}}} \right)}}+{\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{15}}} \right)}}\ge
{\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{6}}} \right)}}+{\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{9}}} \right)}},$$ and chaotic behavior when the opposite inequality holds. Nevertheless, there is an interesting parallel: for the one-dimensional random cellular automata ${\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{0}}} \right)}}+{\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{15}}} \right)}}>0$ implies regular behavior in the sense that $\sigma^*>0$ and exponential convergence of $\sigma_N$ to $\sigma^*$, while ${\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{6}}} \right)}}+{\ensuremath{{\rm P}\!\left( \Phi=\varphi_{{\mathtt{9}}} \right)}}=1$ implies that $\sigma^*=0$, and *no* convergence of $\sigma_N$ to $\sigma^*$.
Appendix: CA symmetries {#symm}
=======================
There are two symmetry operations on the collection of sets of CA rules. The first one is an extension of the mirror map $\mathcal{M}$ defined by $$\mathcal{M}(0)=1,\quad \mathcal{M}(1)=0.$$ It is given by $$(\mathcal{M}\varphi)(x,y)=\mathcal{M}(\varphi(\mathcal{M}(x),\mathcal{M}(y))).$$ The second one is space reversal $\mathcal{R}$, defined by $$(\mathcal{R}\varphi)(x,y)=\varphi(y,x).$$ Both operations are involutions. The effect of $\mathcal{M}$, written as a permutation is: $$(0\; 15) (1 \;7) (2 \;11) (4 \;13) (6\; 9) (8\; 14).$$ The effect of $\mathcal{R}$ is: $$(2 \;4) (3\; 5) (10\; 12) (11\; 13).$$
Appendix: impermeable blocks {#imp-ver}
=============================
Here we give two examples of the proofs that the blocks obtained from Table \[imperm\] yield impermeable blocks.
To prove that the pair $(b_1,\ldots, b_4)=(0,0,1,0),
({\mathtt{j_1}},\ldots,{\mathtt{j_4}})=(2,9,9,2)$ yields an impermeable block, we have to consider a configuration as
--------- -------------------------- -------------------------- -------------------------- -------------------------- -----
${\varphi_{\mathtt{2}}}$ ${\varphi_{\mathtt{9}}}$ ${\varphi_{\mathtt{9}}}$ ${\varphi_{\mathtt{2}}}$
[\ 0 0 1 0 $y$
\
]{} $x$
--------- -------------------------- -------------------------- -------------------------- -------------------------- -----
Filling in ${\varphi_{\mathtt{2}}}(x,0)=0,\, {\varphi_{\mathtt{9}}}(0,1)=0,\,{\varphi_{\mathtt{9}}}(0,0)=1$ and ${\varphi_{\mathtt{2}}}(1,y)=0$, which is true for arbitrary $x$ and $y$, we obtain the next line (time $t=1$):
--------- -------------------------- -------------------------- -------------------------- -------------------------- ------
${\varphi_{\mathtt{2}}}$ ${\varphi_{\mathtt{9}}}$ ${\varphi_{\mathtt{9}}}$ ${\varphi_{\mathtt{2}}}$
[\ 0 0 1 0 $y$
\
]{} $x$
$x'$ 0 0 1 0 $y'$
--------- -------------------------- -------------------------- -------------------------- -------------------------- ------
Since this is the same as the state at time 0, it follows that we do indeed have an impermeable block, consisting of four cells that always stabilize at time 0.
There is only one exception of an impermeable block which does *not* consist of stable cells:
--------- -------------------------- -------------------------- --------
${\varphi_{\mathtt{5}}}$ ${\varphi_{\mathtt{3}}}$
[\ 0 0 $y$
\
]{} $x$
$x'$ 1 1 $y'$
$x''$ 0 0 $y''$
$x'''$ 1 1 $y'''$
--------- -------------------------- -------------------------- --------
This time we obtain a period 2 impermeable block, but actually a period 1 impermeable block is also possible by taking $(b_1,b_2)=(0,1)$.
Appendix: no impermeable blocks {#no-imp}
================================
We will show that the RBCA’s with their support in $\{{\varphi_{\mathtt{2}}},{\varphi_{\mathtt{3}}},{\varphi_{\mathtt{11}}}\}$ do not admit any impermeable blocks. We will do this by showing that diagrams as in the previous appendix can not exist. In the following we will use frequently that a block is impermeable if and only if its mirror image, respectively space reversal is impermeable.
Consider the first element $\varphi_{{\mathtt{j_1}}}$ of the $\varphi$-block. This can not be ${\varphi_{\mathtt{3}}}$, since ${\varphi_{\mathtt{3}}}(x,y)=x$, which is not compatible with impermeability from the left. So it has to be ${\varphi_{\mathtt{2}}}$ or ${\varphi_{\mathtt{11}}}$. Since $\mathcal{M}({\varphi_{\mathtt{2}}})={\varphi_{\mathtt{11}}}$, and $\mathcal{M}({\varphi_{\mathtt{3}}})={\varphi_{\mathtt{3}}}$, we can assume without loss of generality that it is ${\varphi_{\mathtt{11}}}$. Since ${\varphi_{\mathtt{11}}}(x,y)=0$ if and only if $(x,y)=(1,0)$ the first element of the iterates of the $b$-block must be equal to 1:
--------- --------------------------- ------------------- ------------------- ------------------- ---
${\varphi_{\mathtt{11}}}$ $\varphi_{\cdot}$ $\varphi_{\cdot}$ $\varphi_{\cdot}$ …
[\ $\cdot$ $\cdot$ $\cdot$ $\cdot$ …
\
]{} $x$
$x'$ $1$ $\cdot$ $\cdot$ $\cdot$ …
$x''$ $1$ $\cdot$ $\cdot$ $\cdot$ …
--------- --------------------------- ------------------- ------------------- ------------------- ---
We next consider all three possibilities for $\varphi_{{\mathtt{j_2}}}$, filling in, if possible, the values of the second cell given by the rules. For ${\mathtt{j_2}}=2$ we obtain:
--------- --------------------------- -------------------------- ------------------- ------------------- ----------
${\varphi_{\mathtt{11}}}$ ${\varphi_{\mathtt{2}}}$ $\varphi_{\cdot}$ $\varphi_{\cdot}$ …
[\ $\cdot$ $\cdot$ $\cdot$ $\cdot$ $\ldots$
\
]{} $x$
$x'$ $1$ $\cdot$ $\cdot$ $\cdot$ $\ldots$
$x''$ $1$ $0$ $\cdot$ $\cdot$ $\ldots$
--------- --------------------------- -------------------------- ------------------- ------------------- ----------
However this gives a contradiction at $t=3$, since ${\varphi_{\mathtt{11}}}(x'',0)=1-x''$ depends on $x''$. Exactly the same contradiction occurs when ${\mathtt{j_2}}=3$. Consequently we must have ${\mathtt{j_2}}=11$. The whole second column of the $b$-block must consist of 1’s, again because of ${\varphi_{\mathtt{11}}}(x,0)=1-x$. But then, since ${\varphi_{\mathtt{11}}}(1,0)=0$ also the 3$^{rd}$ column must be filled with 1’s starting from $t=1$. It is then quickly checked in the following diagram that ${\mathtt{j_3}}=2$ and ${\mathtt{j_3}}=3$ are impossible:
--------- --------------------------- --------------------------- ---------------------------- ------------------- ----------
${\varphi_{\mathtt{11}}}$ ${\varphi_{\mathtt{11}}}$ $\varphi_{{\mathtt{j_3}}}$ $\varphi_{\cdot}$ …
[\ $\cdot$ $1$ $\cdot$ $\cdot$ $\ldots$
\
]{} $x$
$x'$ $1$ $1$ $1$ $\cdot$ $\ldots$
$x''$ $1$ $1$ $1$ $\cdot$ $\ldots$
--------- --------------------------- --------------------------- ---------------------------- ------------------- ----------
Conclusion: also ${\mathtt{j_3}}=11$. Continuing in this fashion we find that the (string of indices of the) $\varphi$-block has the form $({\mathtt{j_1}},\ldots,{\mathtt{j_p}})=(11,\ldots,11)$. But then we have a problem in the $p^{\rm th}$ column since ${\varphi_{\mathtt{11}}}(1,y)=y$ depends on $y$. Hence for any $p$ this possibility is ruled out.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
Autapses are synapses that connect a neuron to itself in the nervous system. Previously, both experimental and theoretical studies have demonstrated that autaptic connections in the nervous system have a significant physiological function. Autapses in nature provide self-delayed feedback, thus introducing an additional time scale to neuronal activities and causing many dynamic behaviors in neurons. Recently, theoretical studies have revealed that an autapse provides a control option for adjusting the response of a neuron: e.g., an autaptic connection can cause the electrical activities of the Hindmarsh-Rose neuron to switch between quiescent, periodic, and chaotic firing patterns; an autapse can enhance or suppress the mode-locking status of a neuron injected with sinusoidal current; and the firing frequency and interspike interval distributions of the response spike train can also be modified by the autapse. In this paper, we review recent studies that showed how an autapse affects the response of a single neuron.
**Keywords:** Autapse, Self-delay feedback, single neuron, firing pattern\
**PACS:** 87.19.ll, 87.19.ls, 87.18.Sn, 87.19.lg
author:
- Hengtong Wang
- Yong Chen
title: Firing dynamics of an autaptic neuron
---
[UTF8]{}[gbsn]{}
Introduction
============
The nervous system is a complex network of neurons, synapses, and other specialized cells, and through this system, neurons receive and transmit information between different parts of the body [@kandel]. The responses of neurons and the function of synapses have received considerable attention because of their status as building blocks of the nervous system [@Bartos; @Bennett; @Connors]. Nearly a century ago, neuroscientists found a special synaptic structure, the autapse, which is a synapse between different parts of the same neuron [@vander; @Bekkers1998; @yamaguichi]. Because of its odd structure, the self-synaptic connection has remained poorly understood. However, many recent studies found that autapses are much more widespread in the nervous system than previously thought. Autapses have been reported in various brain areas, such as the neocortex, cerebellum, hippocampus, striatum, and substantia nigra [@lubke; @Bekkers1998; @flight; @branco; @Kimura; @tamas]. Interestingly, about 80% of cortical pyramidal neurons have autaptic connections [@lubke]. Many studies have revealed that autapses are not merely curiosities, but play an authentic physiological role in the nervous system. Autapses can maintain persistent activity in the nervous system by mediating positive feedback [@saada].
Recent experiments also demonstrated that autapses are very important for the processing function of the brain [@ikeda]. Bacci et al. recorded the activity of fast spiking interneurons in acute brain slices of juvenile rats and found that autaptic transmission increased the spike-timing precision [@bacci]. A somatic spike can evoke excitatory postsynaptic currents with various amplitudes through an autapse [@Kimura]. However, the function of autapses and their contribution to information processing are still unclear [@ikeda; @bekkers09]. Thus, understanding the effect of autaptic activities on the responses of a neuron is a fundamental step to elucidating the process of information transfer in the nervous system [@salinas1; @ychen2008; @chenyl; @quiroga].
In fact, natural autapses are self-feedback connections in the nervous system and may allow a unique type of self-control, similar to the feedback in other systems [@ikeda]. Actually, feedback creates a circuit, or loop, that connects a system to itself and commonly occurs in many systems, e.g., gene regulatory networks [@davidson; @zhanghui], population dynamics [@cabrera], and climate systems [@bonan]. Feedback is also used extensively to control the state of a nonlinear system, such as to stabilize periodic orbits or to control coherence resonance [@post; @sethia]. Feedback is also very commonly used in the design of electric circuit elements [@waik]. In these systems, information about the past or the present influences the same phenomena in the present or future, respectively. Moreover, feedback has a marked effect on the dynamics of nonlinear systems. The frequency at the onset of the oscillation of a nonlinear system can be modified by the feedback loop [@Gaudreault]. Furthermore, self-delayed feedback with a fixed delay time can suppress chaos and also control steady states [@Ahlborn; @Balanov]. In a simulation study, Popovych et al. showed that time-delayed feedback has the ability to desynchronize groups of model neurons [@popovych].
A neuron with an autaptic connection can provide a distinctive physiological self-feedback model with many dynamic properties. Rusin et al. performed a time-delayed feedback stimulation of a group of cultured neurons and experimentally demonstrated that time-delayed feedback could cause synchronization of the action potentials of a group of neurons [@rusin]. Prager et al. found that time-delayed feedback can facilitate the noise-induced oscillation of a neuronal system [@prager]. Autaptic delayed feedback was found to modify the bursting in the distribution of interspike intervals of a stochastic Hodgkin-Huxley (HH) neuron [@liyy]. Such autaptic delayed feedback can also reduce the spontaneous spiking activity of a stochastic neuron at the characteristic frequencies. The effect of an autapse on the spike rate of a single neuronal system shows dependence on the duration of the autapse activity [@hashemi]. Modulation of autaptic delay feedback can cause the dynamic behaviors of a Hindmarsh-Rose (HR) neuron to switch between quiescent, periodic, and chaotic firing patterns [@wht2014a].
In this article, we review the current stat of knowledge of the effect of an autapse on single neurons.
Autaptic connection
===================
Scientists have long investigated the functions and properties of neurons and synapses. Neurons vary widely with different sizes and forms, and neuron responses are also involved in many dynamic behaviors. A synapse is a structure in the nervous system that permits a neuron (or nerve cell) to transmit an electrical or chemical signal to another cell (neural or otherwise) [@schacter]. The self-synapse, or the autapse, was first described as a synapse between the axon of a pyramidal cell and its own dendrites by Van der Loos and Glaser in 1972 [@vander]. Before the report by Van der loos et al., other terms, such as ’self-excitation’ and ’self-sensing’, were used to describe the self-synaptic structure. For a long time, autapses in the nervous system seemed like anatomical curiosities with questionable functional significance. However, recent experiments began to reveal how autapses might play an important role in brain function. Some reports also suggested that autapses are connected to some neural diseases [@jang]. Neurobiology experiments have always focused on excitatory (glutamate-releasing) and inhibitory (GABA-releasing) autapses. To date, autaptic structures have been found in almost all parts of the nervous system [@Bekkers1998]. Fig. \[autapse2\] shows photos of the many autapses of a single rat hippocampal neuron.
![A single rat hippocampal neuron (arrowed, left) grown in culture on a ‘microdot’ of glia (flat gray cells), showing abundant autapses labeled with an antibody (small dark spots, right) [@ikeda].[]{data-label="autapse2"}](fig1_autapse.eps){width="50.00000%"}
![Schematic drawing of a neuron with autaptic feedback. The structure in the circle represents an autaptic connection between dendritic and axonal terminals. []{data-label="autapse3"}](fig2_autapse.eps){width="45.00000%"}
In nature, autaptic connections enable self-feedback of a neuron. Thus, in almost all current autapse studies, the model of a neuron with an autaptic connection contains a single compartment that exhibits a self-delayed feedback mechanism, as shown in Fig. \[autapse3\]. The delay time represents the elapsed time associated with the axonal propagation prior to the reintroduction of the signal into the neuron. The delay time is also believed to be one of the important properties of autaptic connections. Mathematical models of an autapse in previous studies of the effect of self-synapses on neurons fall into two main categories: one has linear self-coupling and describes a gap junction, and the other has nonlinear self-coupling and describes a chemical synapse, which may be an excitatory or inhibitory synapse. The specific mathematical form of the autapse depends on the chosen neuronal model.
The autaptic connection of a simplified neuron model (such as the HR neuron, FitzHugh-Nagumo neuron, Izhikevich neuron and others) can often be described by one of the following formulas [@hashemi; @wht2014a; @majun1]:
- Linear coupling gives as electrical diffusive-type: $$I_{aut}=g_{aut}[u(t-\tau)-u(t)].
\label{eq1}$$ $g_{aut}$ is the autaptic conductivity. $\tau$ is the delay time. This form of autaptic current is proportional to the difference between the membrane potential at $t$ and that at an earlier time point $t-\tau$.
- The chemical synapse function is modeled using the so-called fast threshold modulation (FTM) scheme [@Belykh; @buric]: $$\begin{aligned}
I_{aut}(t)&=&-g_{aut}(u(t)-V_{syn})S(t-\tau), \\
S(t-\tau)&=&1/\{1+\exp[-k(u(t-\tau)-\theta)]\},\end{aligned}$$ where $g_{aut}$ is the autaptic intensity, and $V_{syn}$ is the synaptic reversal potential. $V_{syn}$=2 and $V_{syn}$=-2 correspond to excitatory and inhibitory autapses, respectively.
For a conductance-based neuron model (such as the HH neuron, Morris-Lecar neuron, Connor-Stevens neuron and others), the mathematic model of the autapse is often given as follows [@liyy; @hashemi; @wht_jtb; @wht_chaos]
- Linear coupling is also always set as electrical diffusive-type and has the same form as Eq. \[eq1\].
- Conductance-based chemical autapses can be described using a bi-exponential function, which has been reported to fit well with experiment results [@william], $$I_{aut}(t)=G(t-t_{delay})(V-E_{syn}),
\label{chemsyn}$$ where $G(t-t_{delay})$ is the autaptic conductance, and $E_{syn}$ is the autaptic reversal potential. In this equation, we used $E_{syn}=0\;\mathrm{mV}$ for excitatory neurons and -80 $\mathrm{mV}$ for inhibitory neurons. The autaptic conductance is defined as $$\begin{aligned}
G(t)&=&g_{aut}\sigma(t-t_{fire})\\
\sigma(t)&=&\frac{\exp(-t/t_d)-\exp(-t/t_r)}{t_d-t_r}\end{aligned}$$ with presynaptic spikes occurring at $t_{fire}$. The parameters $t_d$ and $t_r$ represent the decay and rise times of the function, respectively, and these parameters determine the duration of the response. The autaptic rise and decay times were set to $t_r=0.1\;\mathrm{ms}$ and $t_d=3\;\mathrm{ms}$, respectively.
In the following sections, we will review the two main forms of an autapse. Studies have shown that autapses that exhibit delayed feedback provide a control option for adjusting neuron responses.
Firing pattern transition in a bursting neuron
==============================================
Experimental observations have indicated that action potentials can occur with different firing patterns, and the primary firing patterns are spiking and bursting [@cocatre; @mainen; @izikevich2000; @kepecs2002; @guhuaguang; @yuht]. Bursting is an extremely diverse general phenomenon in firing patterns exhibited by neurons in the central nervous system and spinal cord [@wagenaar; @wangj]. Previous studies on a bursting neuron with an autapse showed the novel dynamics and the transition of firing patterns induced by an autapse [@wht2014a].
The simplified model of Hindmarsh and Rose has turned out to capture the features of experimentally measured electrical data quite accurately, particularly for studies of the spiking-bursting behavior of neuron membrane potentials.
Without an autapse, HR neurons exhibit many dynamic behaviors, including quiescent, regular spiking, periodic, and chaotic bursting firing patterns (see Fig. 1 in Ref. [@wht2014a] and Ref. [@tang]).
The presence of an autapse completely changes the firing patterns of the original HR neuron. The firing pattern can be adjusted from a periodic or chaotic pattern to another periodic pattern or to a chaotic bursting pattern as the autaptic parameters change, independently of the original firing pattern. Fig. \[fig3\_hr\_e1\] shows the time courses of the membrane potentials of an HR neuron with an electrical autapse as an example. Importantly, the maximum action potential is increased by the electrical autapse.
![Time courses of the membrane potentials in response to different $I_{ext}$ with $g_{aut}=0.5$ [@wht2014a]. The blue curves represent the time course of the HR neuron with an autapse, and the red curves indicate that without autapse.[]{data-label="fig3_hr_e1"}](fig3_electrical_all_v.eps){width="50.00000%"}
![Bifurcation diagram of the interspike interval of an HR neuron with an autapse versus the autaptic delay time [@wht2014a]. From left to right, the panels show the results of the electrical autapse, excitatory chemical autapse, and inhibitory chemical autapse.[]{data-label="hr_e2"}](fig4_bifurcation.eps){width="45.00000%"}
The ISI plot bifurcation diagrams of the HR neuron clearly show the transition between periodic and aperiodic firing. For an electrical autapse, the periodic state transits to the chaotic state exhibiting an alternating behavior as the time delay increased. With higher autaptic intensity, this alternating behavior is more noticeable and occurs more frequently. For shorter delay times, the firing pattern of the HR neuron showed periodic spiking independent of the external DC input. The neuron with an excitatory chemical autapse exhibits chaotic firing patterns in a larger area of $g_{aut}$-$\tau$ space than that of the neuron with an electrical autapse. For strong external stimuli, the chaotic region of the combinational parameters in the $g_{aut}$-$\tau$ space is enlarged. The excitatory autapse plays a positive role in generating and enhancing chaos as a whole. As neurons with an inhibitory autaptic connection, the chaotic spiking of the HR neuron can be decreased and suppressed. With the proper inhibitory autaptic parameters, the HR neuron could be driven to a resting state. Fig. \[hr\_e2\] shows the bifurcation of the ISIs of an HR neuron for the three types of autapses.
![An example of the two ways of transition (the discontinuous transition \[upper\] and the continuous transition \[bottom\]) into chaos from periodic firing [@wht2014a]. External stimuli is set as $I_{ext}$=2.67.[]{data-label="cnsns_returnmap1"}](fig5_ways_trasition_chaos.eps){width="50.00000%"}
Without an autapse, the firing pattern of an HR neuron switches from silent to periodic bursting and then to chaotic firing with period-doubling bifurcation as the DC current increased [@wangxj; @gnzalez; @innocenti; @innocenti2009]. For neurons with an autapse, the firing pattern can switch into any other firing pattern, regardless of the original pattern. As a whole, there are two main approaches to transition from a periodic to a chaotic firing pattern: discontinuous and continuous. The interspike interval return map is a useful approach to characterizing the transition to chaos. Fig. \[cnsns\_returnmap1\] shows an example of the two approaches for the transition of a neuron to chaos. In the discontinuous transition, the neuron has a periodic bursting firing pattern because the delay time is short. When the delay time increases, the system suddenly enters a chaotic firing state. In the continuous transition, the neuron first displays a periodic firing pattern. As the autaptic parameters change, the number of spikes in the single burst as well as the period of the periodic bursting increase almost continuously, and the neuron transitions into the chaotic state.
Mode-locking behavior of a regular spiking neuron
=================================================
When excited by a periodic stimulus, neurons respond with various mode-locking firing patterns and quasi-periodic states [@leesg; @chey; @wanght; @fellous; @wht_jtb]. When an autapse is present, the mode-locking firing of a neuron can be switched to another state, depending on the chosen autapse parameters. An autapse provides self-feedback and contributes an additional time scale to the dynamic neuronal system. Thus, the neuron-autapse system may exhibit very complex dynamics, due to the interplay between autaptic delayed feedback, an external periodic stimulus, and the intrinsic activity of the neuron.
In previous studies, the model neuron-autapse system often contained an HH neuron and an autapse. The HH model is a conductance-based model that describes how action potentials in neurons are initiated and propagated [@hodgkin]. In this section, we review the mode-locking firing pattern of an HH neuron with an electrical autapse that has the same mathematical form as Eq. (\[eq1\]).
![Mode-locking of responses of an HH autaptic neuron with different delay times [@wht_jtb]. The lines with different markers denote the firing rate of the neuron excited by different input frequencies. []{data-label="jtb_fig1"}](fig6a.eps "fig:"){width="23.00000%"} ![Mode-locking of responses of an HH autaptic neuron with different delay times [@wht_jtb]. The lines with different markers denote the firing rate of the neuron excited by different input frequencies. []{data-label="jtb_fig1"}](fig6b.eps "fig:"){width="23.00000%"}
![Mode-locking of responses of an HH autaptic neuron with different delay times [@wht_jtb]. The lines with different markers denote the firing rate of the neuron excited by different input frequencies. []{data-label="jtb_fig1"}](fig6c.eps "fig:"){width="23.00000%"} ![Mode-locking of responses of an HH autaptic neuron with different delay times [@wht_jtb]. The lines with different markers denote the firing rate of the neuron excited by different input frequencies. []{data-label="jtb_fig1"}](fig6d.eps "fig:"){width="23.00000%"}
Without an autapse, the mode-locking behaviors depend on the values of the stimulus frequencies and amplitudes. As shown in Fig. 1 in Ref [@wht_jtb], the Arnold tongues in the frequency and amplitude space show the overall features of various phase-locked states and provide the different $p:q$ (denoting output action potentials per input spikes) mode-locking regions. Ref [@leesg] also shows the bifurcation mechanisms that create the boundaries of the complex mode-locking structure.
The presence of an autapse substantially modifies the mode-locking patterns of the neuron. For the same sinusoidal stimulus, the neuron with an autapse can fire more or fewer action potentials than that of a neuron without an autapse. The activities of the neuron can also be driven into sub-threshold oscillation. The electrical autapse displays a regulating function that controls the mode-locking firing of the neuron. The time courses of the membrane potentials of the neuron with and without an autapse are shown in Fig. 2 of Ref. [@wht_jtb]. With an autapse, the modification of the mode-locking firing pattern depends on the autaptic conductivity and the delay time. When the synaptic conductivity is small, the mode-locking $p:q$ is similar to that without an autaptic connection. When the synaptic conductivity is large, the neuron displays very complex mode-locking firing. The $p:q$ value of mode-locked firing increases from zero to a very large value when the frequency of the sinusoidal current increases from zero to the “threshold” frequency (the minimal input frequency for which the neuron fires an action potential given a fixed input amplitude). For further increases in the frequency of the sinusoidal current, the $p:q$ value of mode-locking firing decreases. In the case of extremely strong autaptic conductivity \[see Fig. 3 (d) in Ref. [@wht_jtb]\], the autaptic activities completely disrupt the original mode-locking firing of the neuron without an autapse.
For continuous changes in the delay time, the model-locked firing of the neuron is also quite interesting. Fig. \[jtb\_fig1\] shows the dependence of the mode-locking state on the delay time for different input frequencies and autaptic conductivities. For weak autaptic conductivities \[Fig. \[jtb\_fig1\](a)\], the $p:q$ values of mode-locking are similar to those without self-feedback for the short delay time. As the delay time increases, the values of $p:q$ mode-locking begin to fluctuate. The smaller the input frequency is, the larger the resulting fluctuation. For strong autaptic intensities \[shown in Fig. \[jtb\_fig1\](b-d)\], the $p:q$ values of mode-locking decrease with increasing delay time in a stepwise manner. With further increases in the delay time, the $p:q$ values of mode-locking suddenly jump to a very large value (larger than that without autaptic self-feedback) and then decrease smoothly. As the delay time increases, the $p:q$ value of mode-locking of a neuron with a high autaptic intensity exhibits periodic behavior.
With autaptic self-feedback, the responses of an HH neuron to a sinusoidal stimulus can have higher or lower $p:q$ mode-locking responses than that of a neuron without an autapse. These dynamical behaviors depend on the autaptic intensity and the delay time. Moreover, the presence of an autapse increases the range of values for which the HH neuron spiking is locked with the sinusoidal current. When the autaptic intensity is weak, the mode-locking behaviors show no marked changes. That is, an autapse with a weak intensity will have a negligible effect on the response of the neuron. For stronger autaptic intensities, the $p:q$ value of the mode-locking increases and then decreases as the delay time increases, exhibiting nearly periodic behaviors. Thus, for sufficiently strong autaptic intensity, changing the delay time provides better regulation of mode-locking than does changing the autaptic intensity. The autaptic connection may also provide a control option for adjusting mode-locked firing in a neural information process.
Responses of a regular spiking neuron
=====================================
Response to DC currents
-----------------------
![(Color online) Firing frequencies ( (a),(c)) and ISI distribution ((b), (d)) of a neuron with a chemical autapse (upper panels: excitatory autapse; lower panels: inhibitory autapse) excited by a DC current [@wht_chaos]. Dotted lines gives the odd multiples of half the intrinsic period of a single HH neuron excited by the corresponding DC current.[]{data-label="dc5"}](fig7a.eps "fig:"){width="23.60000%"} ![(Color online) Firing frequencies ( (a),(c)) and ISI distribution ((b), (d)) of a neuron with a chemical autapse (upper panels: excitatory autapse; lower panels: inhibitory autapse) excited by a DC current [@wht_chaos]. Dotted lines gives the odd multiples of half the intrinsic period of a single HH neuron excited by the corresponding DC current.[]{data-label="dc5"}](fig7b.eps "fig:"){width="23.60000%"}
![(Color online) Firing frequencies ( (a),(c)) and ISI distribution ((b), (d)) of a neuron with a chemical autapse (upper panels: excitatory autapse; lower panels: inhibitory autapse) excited by a DC current [@wht_chaos]. Dotted lines gives the odd multiples of half the intrinsic period of a single HH neuron excited by the corresponding DC current.[]{data-label="dc5"}](fig7c.eps "fig:"){width="23.60000%"} ![(Color online) Firing frequencies ( (a),(c)) and ISI distribution ((b), (d)) of a neuron with a chemical autapse (upper panels: excitatory autapse; lower panels: inhibitory autapse) excited by a DC current [@wht_chaos]. Dotted lines gives the odd multiples of half the intrinsic period of a single HH neuron excited by the corresponding DC current.[]{data-label="dc5"}](fig7d.eps "fig:"){width="23.60000%"}
When injected with a DC current, an HH neuron without an autapse displays regular spiking when the current intensity is larger than a critical current value. As for most neuronal models, increases in the current will increase the firing rate of the neuron, and the HH neuron undergoes a Hopf bifurcation. In the presence of an autapse, the output frequency can be higher or lower than the intrinsic frequency (the firing frequency of an HH neuron without an autapse) when the autaptic conductance and delay time are changed. Interestingly, the output frequency, as well as the output ISI distribution, shows periodic behaviors as the autaptic delay time increases. This periodicity of the changes in the output frequency (ISI distributions) in response to changes in the delay time is also similar to the intrinsic period of the neuron. When an HH neuron is connected to an electrical or excitatory chemical autapse, the firing rate and the ISI distribution of the response spike trains can both be either amplified or depressed. When the delay time approaches the odd multiples of the intrinsic half period of the firing of an HH neuron with an electrical or excitatory chemical autapse, the ISIs of the spike train are very different from the intrinsic ISIs. When the delay time approaches a multiple of the intrinsic period, the ISIs of the spike train are similar to the intrinsic ISIs. However, the inhibitory chemical autapse can only suppress the firing responses of the neuron. The ISIs of the response spike train of a neuron with an inhibitory autapse are always greater than the intrinsic period, whereas the corresponding response frequency is not less than the intrinsic response frequency. Current neurobiological experiments also reveal that autapses in the nervous system can be either self-inhibitory or self-excitatory, depending on their location on the neuron [@Bekkers1998; @bekkers2003; @gulledge1; @gulledge2].
Because autaptic pulses perturb neural spiking through a process that is similar to that of an oscillating system with a fixed-delay feedback, the phase-response curve (PRC) theory can provide useful insight into the phenomena of the neuron-autapse system [@glass1; @glass2; @glass3; @glass4]. For an oscillating system, the PRC describes the phase shift that occurs in response to a brief external stimulus [@Carmen1; @Carmen2]. Fig. \[rep\_fig2\] gives the phase-response curves of the three types of autapses. The PRCs of a neuron with an electrical autapse or an excitatory chemical autapse can be negative or positive, depending on the stimulus phase (delay time). Thus, an electrical autapse or excitatory chemical autapse could delay or advance the next spike. Therefore, the firing rate of a neuron with an electrical autapse or an excitatory chemical autapse can be higher or smaller than that of the neuron without an autapse. For the inhibitory chemical autapse, however, the PRC is always positive, because the delay time increased. That is, an inhibitory autapse always postpones the next spike. Thus, the firing rate of an HH neuron with an inhibitory autapse is not larger than that of the neuron without an autapse.
![(Color online) The phase-response curve of three types of autapses with different autaptic conductivity $g_{aut}$, (a) electrical autapse, (b) excitatory chemical autapse, and (c) inhibitory chemical autapse. The external DC stimuli is set as 26 $\mu$A/cm$^2$.[]{data-label="rep_fig2"}](fig8_prc_electrical_all.eps "fig:"){width="22.40000%"} ![(Color online) The phase-response curve of three types of autapses with different autaptic conductivity $g_{aut}$, (a) electrical autapse, (b) excitatory chemical autapse, and (c) inhibitory chemical autapse. The external DC stimuli is set as 26 $\mu$A/cm$^2$.[]{data-label="rep_fig2"}](fig8_prc_excitatory_all.eps "fig:"){width="22.40000%"} ![(Color online) The phase-response curve of three types of autapses with different autaptic conductivity $g_{aut}$, (a) electrical autapse, (b) excitatory chemical autapse, and (c) inhibitory chemical autapse. The external DC stimuli is set as 26 $\mu$A/cm$^2$.[]{data-label="rep_fig2"}](fig8_prc_inhibitory_autapse_all.eps "fig:"){width="22.40000%"}
When the DC current increases, the HH neuron without an autapse undergoes a Hopf bifurcation from a quiescent state to a periodic spiking state. Although the limit cycle and the oscillation period of the neuron have been disturbed by the autaptic current, the neuron-autapse system still undergoes the Hopf bifurcation, generating a stable periodic orbit. Interestingly, an HH neuron with some special autaptic parameters does not fire regular action potentials and is attracted to the quiescent state. This result reveals the phenomenon of spiking death that is induced by an autaptic connection, which is more clearly shown in the membrane potential and autaptic current traces in Fig. \[rfig4\_v\]. With an upper-threshold DC current, the neuron fires action potentials at first. After the delay time, the subsequent autaptic pulse reaches the neuron and drives the neuron to a fixed point. Moreover, the spiking death induced by the autapse is independent of the initial conditions of the neuron. As shown in the inset of Fig. \[rfig4\_v\] (a), the neuron-autapse system can return to the fixed point even when the input is a brief external perturbation. The figure also shows that the spiking death induced by the autapse is stable. The Ref. [@sjw] also shows a similar phenomenon: an HH neuron transitions from a limit cycle to a fixed point when the neuron is perturbed by an excitatory chemical synaptic pulse.
![(Color online) (a) Traces of the membrane potential (dashed lines) and autaptic current (solid lines) [@wht_chaos]. (b) The phase portrait corresponds to (a). (c) and (d) are the enlarged plots of (a) and (b), respectively. The different colors indicate different integral initial conditions (The specific set of parameters can be found in Ref. [@wht_chaos]). The inset in (a): The membrane potential of HH neuron-autapse system perturbed by a brief step pause (black solid line).[]{data-label="rfig4_v"}](fig9_bifurcation_stable_point.eps){width="50.00000%"}
Response to a random synaptic pulse-like input
----------------------------------------------
The assumed $\alpha$ form of the postsynaptic current model that is used in many works is perfect for generating pulse-like currents that are similar to the synaptic pulses observed experimentally and also enables easy modification of the ISI of the input current to investigate the input-output properties conveniently [@wanght]. Injecting this type of synaptic pulse-like signal with different ISIs into the HH neuron, Hideo et al. have reported that the mean firing frequency decreases as the mean input ISI increases [@Hideo; @Borkowski]. When a neuron is injected with a synaptic-like pulsed current with random ISIs, the neuron displays interesting autapse-induced response behaviors [@wht_chaos]. As the delay time increases, the frequency of the neuron with a sufficient electrical or excitatory chemical autapse shows nearly periodic behaviors. Such periodicity is not changed when the mean ISIs of the input synaptic current changes. When given the input synaptic pulses with a large mean value of ISIs, however, the response frequency of a neuron with an inhibitory autapse does not show periodic behavior when the autaptic delay time increases.
Autaptic activities also influence the detailed response of the ISI distribution. For a short autaptic delay time, the output ISIs are distributed in an area \[blue bars in Fig. \[returnmap\]\] that is smaller than the input area (red lines in Fig. \[returnmap\]). When the delay time increases, the region in which the output ISIs are distributed does not change much compared with that for the neuron without an autapse (green lines in Fig. \[returnmap\]). For further increases in the delay time, the distribution of the output ISIs suddenly decreases almost to a single point ($t_{delay}=8.5$ ms in Fig. \[returnmap\]). For these autaptic parameters, the responding spike train is strongly regulated. Then, the size of the distribution of ISIs increases slowly as the delay time is increased further.
When the size of the distribution of the output ISIs increases to its maximum (this maximum distribution area is smaller than the area for the low delay time condition $t_{delay}<5.5 \mathrm{ms}$), the distribution suddenly shrinks nearly to a point again (see $t_{delay}=20.5$ ms in Fig. \[returnmap\]). For the entire range of delay times, the distribution of the output ISIs periodically exhibits the above behaviors when the delay time increases. Thus, the delayed feedback activities of an autapse can act as a regulator that adjusts the ISIs of the output spike train. For some specific autaptic parameters, the resulting spike train of a neuron can be modified to obtain an almost regular spiking, even for a very random ISI input.
![(Color online) Histograms of output (blue bars) ISIs of a neuron with an electrical autapse [@wht_chaos]. The green lines gives the output ISI of a neuron without an autapse and the red lines give the ISI of the random input pulse train.[]{data-label="returnmap"}](fig10.eps){width="45.00000%"}
Without an autapse, the neuron will filter the spikes with a short ISI and thus show low-pass filtering behavior when the neuron is injected with a random synaptic pulse-like current [@Hideo]. For a neuron with an autapse, long ISI pulses can be filtered in addition to the short ISI spikes. Thus, the neuron-autapse system displays complex filtering behaviors, including low-pass filtering and band-filtering behaviors.
When the delay time is short, an HH neuron with an electrical or excitatory chemical autapse acts as a low-pass filter similar to a neuron without an autapse and removes the spikes with short ISIs. When the autaptic delay time is long enough, the neuron displays band-pass filtering behavior and removes both the short- and long-ISI spikes. The cut-off value for the ISIs (either the minimum or the maximum ISIs that will be retained in the output spike train) can be changed by changing the autaptic parameters. For some specific autaptic parameters, the neuron can filter most of the input pulse, and the output spike train can be altered into a nearly regular spike train, even for an input with highly random ISIs. More interestingly, a neuron with an inhibitory chemical autapse can only act as a low-pass filter. Thus, the inhibitory chemical autapse does not have a significant effect on the frequency of the output spike train.
It can be conjectured that the filtering properties depend on the intrinsic properties of both the neuron and the autapse. For electrical or excitatory chemical autapses, the delay time will act as a border or as a cut-off value for the ISI filtering of synaptic pulses with a long ISI. The intrinsic filtering property of neurons removes the short input ISI pulses, and the time delay of the electrical or excitatory chemical autapse removes the long input ISI pulses. However, an inhibitory autapse delays the spikes of a neuron and thus filters the spikes that have short ISI pulses. Thus, the firing frequency of a neuron with an inhibitory autapse is not higher than that of a neuron without an autapse.
Effect of noise on the firing dynamics of a neuron
==================================================
Neuronal noise is random electrical fluctuations within neuronal networks and affects the patterns of neural activity in a determinant way [@white; @Jacobson; @wanghuiqiao; @Destexhe]. In the presence of noise, autapses can also substantially affect the response of a neuron. In this section, we review previous studies on the interplay of autapses and noise on the firing of an autaptic neuron.
Considering the subthreshold dynamics of a neuron with interaction between autaptic-delayed feedback and noise, Masoller et al. investigated the firing patterns of an HH model of a thermoreceptor neuron with an electrical autaptic feedback in the presence of a Gaussian white noise [@masoller]. In their studies, the neuron displays only subthreshold oscillations in the absence of feedback and noise. Their results show that the interaction among weak autaptic feedback, noise, and the subthreshold intrinsic activity is nontrivial. The subthreshold oscillation amplitude can be enhanced by the autapse in the presence of external noise, and this enhancement is more pronounced for certain delay values. For negative autaptic delay feedback, the firing rate can be lower than that of the noise-free situation, depending on the delay. This is because noise inhibits feedback-induced spikes by driving the neuronal oscillations away from the firing threshold. For positive autaptic delay feedback, there are regions of delay values where the noise-induced spikes are inhibited by the feedback; in this case, the autaptic feedback drives the neuronal oscillations away from the threshold.
Li and his co-workers analyzed the effects of electrical autapses on the spiking dynamics of a stochastic HH neuron [@liyy], considering the stochastic gating of ion channels or the so-called intrinsic channel noise is considered. They found that the delayed feedback manifests itself in the occurrence of bursting and a rich multimodal interspike interval distribution, exhibiting a delay-induced reduction in the spontaneous spiking activity at characteristic frequencies. For small numbers of ion channels, the channel noise is sizable and the excitatory dynamics remain practically unaffected by the delay. However, smaller noise levels and stronger autaptic intensities induce different synchronization phenomena between the delay time and the intrinsic time scales. The delay time and the intrinsic time scales determine the number of spikes that will be induced and become subsequently locked during one delay epoch.
Recently, Yilmaz and Ozer showed that the electrical autaptic delayed feedback either enhances or suppresses the weak signal detection, depending on the parameters of autapse and channel noise [@ozer]. When the delay time is close to integer multiples of the period of the intrinsic oscillations, the autapse enhances the weak periodic signal detection for the optimal values of the intrinsic noise and the autaptic intensity. The system response also exhibits stochastic resonance behavior, depending on the autaptic intensity. Moreover, the weak signal detection capability of the HH neuron is strongly dependent on the cell size and autaptic strength.
Conclusion and outlook
======================
Self-delayed feedback significantly affects nonlinear dynamic systems because such self-feedback loops introduce a new time scale into the dynamics of the system [@soriano; @park; @soriano]. The system of a neuron with an autapse contains two time scales: the intrinsic time scale of the neuron and the time scale of the autapse. However, the situation is more complex in the nervous system, because there are many more time scales, including the time scales of the neurons, the synapse, the environment, and the autaptic connections. This complexity raises the question of how the activity time of an autapse influences the activities of coupled neurons in the nervous system.
Autapses provide self-feedback circuits that are common in the nervous system. Since the naming of autapses, many experimental studies have revealed that autapses play important roles in brain function [@Bekkers1998; @ikeda; @bacci; @Kimura; @salinas1; @ychen2008; @quiroga]. The current theoretical studies on the topic of autaptic connections also reported the importance of autapses and the many novel dynamic behaviors induced by the autapse [@rusin; @liyy; @hashemi; @wht2014a; @wht_jtb; @wht_chaos]. Autapses provide a control option that can sufficiently adjust the firing behaviors of a neuron for any form of input stimulus, regardless of the neuron type.
Autapses offer a new mechanism for switching between quiescent, periodic and chaotic firing patterns in bursting neurons [@wht2014a]. Additionally, the nervous system responds rapidly to an external stimulus based on the transition between neuron firing patterns [@Belykh; @erichsen]. Thus, the results of the firing pattern transition induced by an autapse also indicate that an autapse could act as an efficient tool for controlling the transition among different relevant neuronal activities in the nervous system.
For an HH neuron with an autapse, the firing frequency and interspike interval distributions of the output spike train show periodic behavior when the delay time is increased [@wht_chaos]. When specific autaptic parameters are chosen, the response spike trains are nearly regular, and the ISI distribution covers a small area, even with a highly random input. This phenomenon emphassises the autapse-induced filtering behaviors of the neuron. These results about the autapse are useful for studying the control of a nonlinear system.
In neuronal systems, an autapse can take the form of recurrent excitation, i.e., the discharge of a neuron, possibly after passing through the axon, can subsequently induce an excitatory response in the same neuron. This self-excitation mechanism is important for maintaining persistent activity, particularly the feeding behavior in sea slugs. Leonel et al. also reported that such processes play functional roles in amplifying activity in neuronal assemblies, causing reverberating activity, inducing some form of memory, or generating rhythmic patterns, such as those in central pattern generators [@gomez]. The previous experimental data also indicate that the brain tissue expresses a novel form of self inhibition, namely autaptic inhibitory transmission in Fast-Spike (FS) cells slow self activities in interneurons [@Pawelzik]. Ma et al. also investigated the activities of a two-dimensional neural network containing autapses with different time delays and found that the autapses induced many novel collective behaviors [@majun2]. The effect of autapse on the synchronization of neural network have also been conducted in a group of HH neurons with small world network structure [@wu]. It was found that the neurons exhibit synchronization transitions as autaptic delay feedback is varied, and fine synchronized network activities occur when an optimal autaptic strength is chosen. Alberto Bacci and his colleagues studied the autaptic self-inhibition of basket cells and the role of the autaptic connections of disinhibition within cortical circuits and argued that autaptic feedback could have a dual function in temporally coordinating parvalbumin basket cells during cortical network activity [@deleuze]. Another study also indicated that fast spiking neurons with autapses showed the strongest asynchronous release in brain slices obtained from patients with intractable epilepsy, and that such discharges may be involved in generating and regulating network activities, including epileptic activity [@jang].
Although there are many studies on the topic of the autapse, the precise function of autapses and their contribution to information processing are remain unclear. There are also many open questions both in the experimental and theoretical areas: 1) What is the role of autapses in coordinating network activities? i.e., FS cell autapses in adjusting fast network synchrony. 2) What are the molecular mechanisms underlying autaptic asynchronous release and what is its relevance during physiological and pathological network activities? 3) What are the functions of autapses in the information propagation in the neural circuit and neural network? Addressing these questions will also facilitate our understanding of the fundamental mechanisms governing several core functions of cortical activities.
[999]{} Kandel E R, Schwartz J H, Jessell T M, Siegelbaum S A and Hudspeth A J 2000 *Principles of Neural Science* (New York: McGraw-Hill Medical)
Bartos M, Vida I and Jonas P 2007 *Nat. Rev. Neurosci.* **8** 45
Bennett M V L and Zukin R S 2004 *Neuron* **41** 495
Connors B W and Long M A 2004 *Annu. Rev. Neurosci.* **27** 393
Van der Loos H and Glaser E M 1972 *Brain Res.* **48** 355
Bekkers J M 1998 *Curr. Biol.* **8** R52
Yamaguchi K 2008 *Autapse Encyclopedia of Neuroscience* (M. D. Binder ed.), N Hirokawa and U Windhorst (Springer Berlin Heidelberg) pp 229-32
Lübke J, Markram H, Frotscher M and Sakmann B 1996 *J. Neurosci.* **16** 3209
Flight M H 2009 *Nat. Rev. Neurosci.* **10** 316
Branco T and Staras K 2009 *Nat. Rev. Neurosci.* **10** 373
Kimura F, Otsu Y and Tsumoto T 1997 *J. Neurophysiol.* **77** 2805
Tamás G, Buhl E H and Somogyi P 1997 *J. Neurosci.* **17** 6352
Saada R, Miller N, Hurwitz I and Susswein A J 2009 *Curr. Biol.* **19** 479
Ikeda K and Bekkers J M 2006 *Curr. Biol.* **16** R308
Bacci A and Huguenard J R 2006 *Neuron* **49** 119
Bekkers J M 2009 *Curr. Biol.* **19**, R296
Salinas E and Sejnowski T J 2001 *Nat. Rev. Neurosci.* **2** 539
Chen Y, Yu L and Qin S M 2008 *Phys. Rev. E* **78** 051909
Chen Y, Zhang H, Wang H, Yu L and Chen Y 2013 *PLoS ONE* **8** e56822
Quiroga R Q and Panzeri S 2009 *Nat. Rev. Neurosci.* **10** 173
Davidson E. and Levin M. 2005 *Proc. Natl. Acad. Sci. U.S.A.* **102** 4935
Zhang H, Chen Y and Chen Y 2012 *PLoS ONE* **7** e51840
Cabrera J L and Milton J G 2002 *Phys. Rev. Lett.* **89** 158702
Bonan G B 2008 *Science* **320** 1444
Postlethwaite C M and Silber M 2007 *Phys. Rev. E* **76** 056214
Sethia G C, Kurths J, and Sen A 2007 *Phys. Lett. A* **364** 227
Chen W-K 2005 *Circuit Analysis and Feedback Amplifier Theory* (Bocan Raton, FL: CRC Press)
Gaudreault M, Drolet F and Viñals J 2012 *Phys. Rev. E* **85** 056214
Ahlborn A and Parlitz U 2004 *Phys. Rev. Lett.* **93** 264101
Balanov A G, Janson N B and Schöll E 2005 *Phys. Rev. E* **71** 016222
Popovych O V, Hauptmann C and Tass P A 2006 *Biol. Cybern.* **95** 69
Rusin C G, Johnson S E, Kapur J and Hudson J L 2011 *Phys. Rev. E* **84** 066202
Prager T, Lerch H P, Schimansky-Geier L and Schöll E 2007 *J. Phys. A* **40** 11045
Li Y, Schmid G, Hänggi P and Schimansky-Geier L 2010 *Phys. Rev. E* **82** 061907
Hashemi M, Valizadeh A and Azizi Y 2012 *Phys. Rev. E* **85** 021917
Wang H, Ma J, Chen Y and Chen Y 2014 *Commun. Nonlinear Sci. Numer. Simul.* **19** 3242
Schacter D L, Gilbert D T, Wegner D M and Nock M K 2014 *Psychology* (New York, NY: Worth Publishers)
Jiang M, Zhu J, Liu Y, Yang M, Tian C, Jiang S, Wang Y, Guo H, Wang K and Shu Y 2012 *PLoS Biol.* **10** e1001324
Jun Ma H Q 2015 *Int. J. of Mod. Phys. B* **29** 1450239
Belykh I, de Lange E and Hasler M 2005 *Phys. Rev. Lett.* **94** 188101
Burić N, Todorović K and Vasović N 2008 *Phys. Rev. E* **78** 036211
Wang H, Sun Y, Li Y and Chen Y 2014 *J. Theor. Biol.* **358** 25
Wang H, Wang L, Chen Y and Chen Y 2014 *Chaos* **24** 033122
Connelly W M and Lees G 2010 *J. Physiol.* **588**, 2047
Cocatre-Zilgien JH, Delcomyn F 1992 *J. Neurosci. Meth.* **41** 19
Mainen Z F and Sejnowski T J 1996 *Nature* **382** 363
Izhikevich E M 2000 *Int. J. Bifurcat. Chaos* **10** 1171
Kepecs A, Wang X-J and Lisman J 2002 *J. Neurosci.* **22** 9053
Gu Hua-Guang,Zhu Zhou,Jia Bing. 2011 *Acta Physica Sinica* **60** 100505
Yu H-T, Wang J, Deng B and Wei X-L 2013 *Chin. Phys. B* **22** 018701
Wagenaar D A, Pine J and Potter S M 2006 *BMC Neurosci.* **7** 11
Yu H, Wang J, Deng B, Wei X, Wong Y K, Chan W L, Tsang K M, Ziqi Y. et al. 2011 *Chaos* **21** 013127
Tang G, Xu K and Jiang L 2011 *Phys. Rev. E* **84** 046207
Wang X-J 1993 *Physica D* **62** 263
Jm G-M 2003 *Chaos* **13** 845
Innocenti G, Morelli A, Genesio R and Torcini A 2007 *Chaos* **17** 043128
Innocenti G and Genesio R 2009 *Chaos* **19** 023124
Lee S-G and Kim S 2006 *Phys. Rev. E* **73** 041924
Che Y-Q, Wang J, Si W-J and Fei X-Y 2009 *Chaos, Solitons* & *Fractals* **39** 454
Wang H, Wang L, Yu L and Chen Y 2011 *Phys. Rev. E.* **83** 021915
Fellous J-M, Houweling, A R, Modi, R H, Rao, R P N, Tiesinga P H E and Sejnowski T J2001 *J Neurophysiol.* **85** 1782
Hodgkin A L and Huxley A F 1952 *The J. of Physiol.* **117** 500
Bekkers J M 2003 *Curr. Biol.* **13**, R433
Gulledge A T and Stuart G J 2003 *Neuron* **37**, 299
Saada-Madar R, Miller N and Susswein A J 2012 *J. Mol. Histol.* **43** 431
Lewis J, Bachoo M, Glass L, and Polosa C 1987 *Phys. Lett. A* **125**, 119
Glass L and Zeng W-Z 1990 *Ann.NY Acad.Sci.* **591** 316
Lewis J E, Glass L, Bachoo M, and Polosa C 1992 *J. Theor. Biol.* **159** 491
Kunysz A M, Shrier A, and Glass L 1997 *Am. J. Physiol.* **273** C331
Smeal R M, Ermentrout G B, and White J A 2010 *Trans. R. Soc. B* **365** 2407
Krogh-Madsen T, Butera R, Ermentrout G B, and Glass L, in Phase Response Curves in Neuroscience, edited by N.W. Schultheiss, Prinz A A., and Butera R J (Springer New York, 2012), pp. 33
Wang S-J, Xu X-J, Wu Z-X, Huang Z-G, and Wang Y-H 2008 *Phys. Rev. E* **78** 061906
Hasegawa H 2000 *Phys. Rev. E* **61** 718 Borkowski L S *Phys. Rev. E* **80** 051914
White J A, Rubinstein J T and Kay A R 2000 *Trends in Neurosci.* **23** 131
Jacobson G A, et al. 2005 *J. Physiol.* **564** 145
Wang H, Yu L and Chen Y 2009 *Acta Phys. Sin.* **58** 5070
Destexhe A and Rudolph-Lilith M 2012 Neuronal Noise (New York: Springer)
Masoller C, Torrent M C and Garcia-Ojalvo J 2008 *Phys. Rev. E* **78** 041907
Yilmaz E and Ozer M 2015 *Physica A* **421** 455
Soriano M C, Garcia-Ojalvo J, Mirasso C R and Fischer I 2013 *Rev. Mod. Phys.* **85** 421
Park J-H, Huh S-H, Kim S-H, Seo S-J and Park G-T 2005 *IEEE Trans. Neural. Netw.* **16** 414
Erichsen R J, Brunnet L G. 2008 *Phys. Rev. E* **78** 061917
Gómez L, Budelli R and Pakdaman K 2001 *Phys. Rev. E* **64** 061910
Pawelzik H, Hughes D I and Thomson A M 2003 *J Physiol* **546** 701
Qin H, Wu Y, Wang C and Ma J 2015 *Commun. Nonlinear. Sci. Numer. Simulat.* **23** 164
Wu Y, Gong Y and Wang Q 2015 *Chaos* **25** 043113
Deleuze C, Pazienti A and Bacci A 2014 *Curr. Opin. Neurobiol.* **26** 64
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'This paper presents a simple and elementary proof of Butcher’s theorem on the order conditions of Runge-Kutta methods. It is based on a recursive definition of rooted trees and avoids combinatorial tools such as labelings and Faà di Bruno’s formula. This strictly recursive approach can easily and elegantly be implemented using modern computer algebra systems like [*Mathematica*]{} for automatically generating the order conditions. The full, but short source code is presented and applied to some instructive examples.'
address: 'Munich University of Technology, 80290 Munich, Germany'
author:
- Folkmar Bornemann
date: 'November 4, 2002'
title: 'Runge-Kutta Methods, Trees, and [*Mathematica*]{}'
---
Introduction
============
A first step towards the construction of Runge-Kutta methods is the calculation of the order conditions that the coefficients have to obey. In the old days they were obtained by expanding the error term in a Taylor series by hand, a procedure which for higher orders sooner or later runs into difficulties because of the largely increasing combinatorial complexity. It was a major break-through when Butcher [@Butc63] published 1963 his result of systematically describing order conditions by rooted trees. The proof of this result has evolved very much in meantime, mainly under the influence of Butcher’s later work [@Butc87] and the contributions of Hairer and Wanner [@HaWa74; @HNWa87]. In this paper we will present a simple and elementary proof of Butcher’s theorem by using very consequently the recursive structure of rooted trees. This way we avoid lengthy calculations of combinatorial coefficients, the use of tree-labelings, or Faà di Bruno’s formula as in [@Butc87; @HNWa87]. Our proof is very similar in spirit to the presentation of $B$-series by Hairer in Chapter 2 of his lecture notes [@Hair99].
As early as 1976 Jenks [@Jenk76] posed the problem of automatically generating order conditions for Runge-Kutta methods using computer algebra systems, but no replies were received. In 1988 Keiper of Wolfram Research concluded that the method of automatically calculating Taylor expansions by brute force was bound to be very inefficient. Naturally he turned to the elegant results of Butcher’s. Utilizing them, he wrote the [*Mathematica*]{} package [Butcher.m]{}, which has been available as part of the standard distribution of [*Mathematica*]{} since then. This package was later considerably improved by Sofroniou [@Sofr94] and offers a lot of sophisticated tools.
While teaching the simple proof of Butcher’s result in a first course on numerical ODEs, the author realized that the underlying recursive structure could also be exploited for a simple and elegant computer implementation. This approach differs from the work of Sofroniou in various respects. We will present the full source code in [*Mathematica*]{} and some applications.
Runge-Kutta methods
===================
The Runge-Kutta methods are one-step discretizations of initial-value problems for systems of $d$ ordinary differential equations, $$x' = f(t,x),\qquad x(t_0) = x_0,$$ where the right-hand side $f: [t_0,T]\times \Omega \subset {{\mathbb R}}\times {{\mathbb R}}^d
\to{{\mathbb R}}^d$ is assumed to be sufficiently smooth. The [*continuous evolution*]{} $x(t) = \Phi^{t,t_0} x_0$ of the initial-value problem is approximated in steps of length $\tau$ by $\Psi^{t+\tau,t} x \approx
\Phi^{t+\tau,t}x$. This [*discrete evolution*]{} $\Psi^{t+\tau,t}x$ is defined as an approximation of the integral-equation representation $$\Phi^{t+\tau,t} x = x + \int_t^{t+\tau} f(\sigma,\Phi^{\sigma,t}
x) \,d\sigma$$ by appropriate quadrature formulas: $$\label{eq:rk}
\begin{array}{rcl}
\Psi^{t+\tau,t}x = x + \tau \displaystyle\sum_{i=1}^s b_i
k_i,\qquad
k_i = f\left(t+c_i\tau, x + \tau\displaystyle\sum_{j=1}^s
a_{ij} k_j\right).
\end{array}$$ The vectors $k_i \in {{\mathbb R}}^d$, $i=1,\ldots,s$, are called [*stages*]{}, $s$ is the stage number. Following the standard notation, we collect the coefficients of the method into a matrix and two vectors $${{\mathcal{A}}}= (a_{ij})_{ij} \in {{\mathbb R}}^{s \times s},\; b = (b_1,\ldots,b_s)^T \in
{{\mathbb R}}^s,\;
c = (c_1,\ldots,c_s)^T \in {{\mathbb R}}^s.$$ The method is [*explicit*]{}, if ${{\mathcal{A}}}$ is strictly lower triangular. The method has [*order*]{} $p \in {{\mathbb N}}$, if the error term expands to $$\Phi^{t+\tau,t}x - \Psi^{t+\tau,t}x = O(\tau^{p+1}).$$ In terms of the Taylor expansion of the error $\Phi - \Psi$, the vanishing of all lower order terms in $\tau$ just defines the conditions which have to be satisfied by the coefficients ${{\mathcal{A}}}$, $b$ and $c$ of a Runge-Kutta method.
If we choose $c_i = \sum_{j=1}^s a_{ij}$, it can be shown [@HNWa87], that there is no loss of generality in considering [*autonomous*]{} systems only, i.e., those with no dependence of $f$ on $t$. Doing so, the expressions $\Phi^{t+\tau,t}x$ and $\Psi^{t+\tau,t} x$ are likewise independent of $t$. We will write $\Phi^\tau x$ and $\Psi^\tau x$ for short, calling them the [*flow*]{} and the [*discrete flow*]{} of the continuous resp. the discrete system.
Elementary differentials and rooted trees
=========================================
The Taylor expansions of both the phase flow $\Phi^\tau x$ and the discrete flow $\Psi^\tau x$ are linear combinations of [*elementary*]{} differentials like $$f'''(f'f,f'f,f) = \sum_{ijklm} \frac{\partial^3 f}{\partial x_i
\partial x_j\partial x_k} \cdot \frac{\partial f_i}{\partial x_l
} f_l \cdot \frac{\partial f_j}{\partial x_m } f_m \cdot f_k .$$ We will use the short multilinear notation of the left hand side for the rest of the paper.
An elementary differential can be expressed uniquely by the structure of how the subterms enter the multilinear maps. For instance, looking at the expression $f'''(f'f,f'f,f)$ we observe that $f'''$ must be a third derivative, since [*three*]{} arguments make it [*three*]{}-linear. This structure can be expressed in general by [*rooted trees*]{}, e.g., $$\begin{picture}(5.5,2)
\put(0.75,0.5){\line(0,1){0.6}} \put(0.75,1.1){\line(1,1){0.5}}
\put(0.75,1.1){\line(-1,1){0.5}} \put(0.75,0.5){\circle{0.2}}
\put(0.75,0.5){\circle*{0.1}} \put(0.75,1.1){\circle*{0.1}}
\put(0.25,1.6){\circle*{0.1}} \put(1.25,1.6){\circle*{0.1}}
\put(3.25,1){\makebox(0,0){ expressing $\; f'f''(f,f),$}}
\end{picture}
\begin{picture}(5.5,2)
\put(0.75,0.5){\line(1,1){0.5}} \put(0.75,0.5){\line(-1,1){0.5}}
\put(0.25,1.0){\line(0,1){0.6}} \put(0.75,0.5){\circle{0.2}}
\put(0.75,0.5){\circle*{0.1}} \put(0.25,1){\circle*{0.1}}
\put(1.25,1){\circle*{0.1}} \put(0.25,1.6){\circle*{0.1}}
\put(3.5,1){\makebox(0,0){ expressing $\; f''(f'f,f).$}}
\end{picture}$$ Every node with $n$ children denotes a $n$th derivative of $f$, which is applied as a multilinear map to further elementary differentials, according to the structure of the tree. We start reading off this structure by looking at the root. This defines a [*recursive*]{} procedure, if we observe the following: Having removed the root and its edges, a rooted tree $\beta$ decomposes into rooted subtrees $\beta_1,\ldots,\beta_n$ with strictly less nodes. The roots of the subtrees $\beta_1,\ldots,\beta_n$ are exactly the $n$ children of $\beta$’s root. This way a rooted tree $\beta$ can be defined as the [*unordered*]{} list of its successors $$\label{eq:treedef}
\beta=[\beta_1,\ldots,\beta_n],\qquad \#\beta = 1 + \#\beta_1 +
\ldots + \#\beta_n.$$ Here, we denote by $\#\beta$ the [*order*]{} of a rooted tree $\beta$, i.e., the number of its nodes. The root itself can be identified with the [*empty*]{} list, $\odot = [\,]$.
An application of this procedure shows for the examples above that $f'(f''(f,f))$ is expressed by $[[\odot,\odot]]$ and $f''(f'(f),f)$ is expressed by $[[\odot],\odot]$. The reader will observe the perfect matching of parentheses and commas. In general the relation between a rooted tree $\beta=[\beta_1,\ldots,\beta_n]$ and its corresponding elementary differential $f^{(\beta)}(x)$ is recursively defined by $$\label{eq:diffdef}
f^{(\beta)}(x)=f^{(n)}(x) \cdot
\left(f^{(\beta_1)}(x),\ldots,f^{(\beta_n)}(x)\right).$$ The dot of multiplication denotes the multilinear application of the derivative to the $n$ given arguments. Due to the symmetry of the $n$-linear map $f^{(n)}$, the order of the subtrees $\beta_1,\ldots,\beta_n$ does not matter, which means, that $f^{(\beta)}$ depends in a well-defined way on $\beta$ as an unordered list only.
From $\odot = [\,]$ we deduce $f^{(\odot)} = f$. Analogously, each of the recursive definitions in the following will have a well-defined meaning if applied to the single root $\odot =
[\,]$, mostly by using the reasonable convention that empty products evaluate to one and empty sums to zero—a convention that is also observed by most computer algebra systems.
A simple proof of Butcher’s theorem
===================================
We are now in a position to calculate and denote the Taylor expansion of the continuous flow $\Phi^\tau$ in a clear and compact fashion.
Given $f\in C^p(\Omega,{{\mathbb R}}^d)$ the flow $\Phi^\tau x$ expands to $$\label{eq:phitay}
\Phi^\tau x = x + \sum_{\#\beta\leq p}
\frac{\tau^{\#\beta}}{\beta!} \alpha_\beta\, f^{(\beta)}(x) +
O(\tau^{p+1}).$$ The coefficients $\beta!$ and $\alpha_\beta$ belonging to a rooted tree $\beta=[\beta_1,\ldots,\beta_n]$ are recursively defined by $$\label{eq:factorialdef}
\beta ! = (\#\beta)\, {\beta_1}!\cdot\ldots\cdot{\beta_n}!, \qquad
\alpha_\beta = \frac{\delta_\beta}{n !}\,
\alpha_{\beta_1}\cdot\ldots\cdot \alpha_{\beta_n}.$$ By $\delta_\beta$ we denote the number of different ordered tuples $(\beta_1,\ldots,\beta_n)$ which correspond to the same unordered list $\beta =[\beta_1,\ldots,\beta_n]$.
The assertion is obviously true for $p=0$. We proceed by induction on $p$. Using the assertion for $p$, the multivariate Taylor formula and the multilinearity of the derivatives we obtain $$\begin{split}
f(\Phi^\tau x) &= f\left(x + \sum_{\#\beta\leq p}
\frac{\tau^{\#\beta}}{\beta!} \alpha_\beta f^{(\beta)} +
O(\tau^{p+1})\right)\\ &= \sum_{n = 0}^p \frac{1}{n!}
f^{(n)} \!\cdot\!\! \left( \sum_{\#\beta_1\leq p}
\frac{\tau^{\#\beta_1}}{\beta_1!} \alpha_{\beta_1}
f^{(\beta_1)},\ldots,\!\sum_{\#\beta_n\leq p}
\frac{\tau^{\#\beta_n}}{\beta_n!}
\alpha_{\beta_n}
f^{(\beta_n)} \right) + O(\tau^{p+1})\\
\end{split}$$ $$\begin{split}
\phantom{f(\Phi^\tau x)}&=\sum_{n=0}^p
\frac{1}{n!}\sum_{\#\beta_1+\ldots+\#\beta_n\leq p}
\frac{\tau^{\#\beta_1+\ldots+\#\beta_n}}{\beta_1! \cdot
\ldots\cdot \beta_n!} \cdot
\alpha_{\beta_1}\cdot\ldots\cdot\alpha_{\beta_n}\cdot
\\ &\qquad f^{(n)} \cdot \left( f^{(\beta_1)},\ldots,f^{(\beta_n)}\right) +
O(\tau^{p+1})\\
&= \sum_{n=0}^p \sum_{\substack{
\beta =[\beta_1,\ldots,\beta_n]\\
\#\beta \leq p+1
}}
\frac{\#\beta\cdot
\tau^{\#\beta-1}}{\beta!}\! \cdot\!
\underbrace{\frac{\delta_\beta}{n!}\,
\alpha_{\beta_1}\cdot\ldots\cdot\alpha_{\beta_n}}_{=\alpha_\beta}
\,f^{(\beta)} + O(\tau^{p+1})\\ &=
\sum_{\#\beta \leq p+1} \frac{\#\beta\cdot
\tau^{\#\beta-1}}{\beta!} \alpha_\beta\,f^{(\beta)}
+O(\tau^{p+1}).
\end{split}$$ Plugging this into the integral form of the initial value problem we obtain $$\Phi^\tau x = x + \int_0^\tau f(\Phi^\sigma x) \,d\sigma = x + \sum_{\#\beta \leq p+1}
\frac{\tau^{\#\beta}}{\beta!} \alpha_\beta\,f^{(\beta)}
+O(\tau^{p+2}),$$ which proves the assertion for $p+1$.
A likewise clear and compact expression can be calculated for the Taylor expansion of the discrete flow.
Given $f\in C^p(\Omega,{{\mathbb R}}^d)$ the discrete flow $\Psi^\tau x$ expands to $$\label{eq:psitay}
\Psi^\tau x = x + \sum_{\#\beta\leq p} \tau^{\#\beta}
\alpha_\beta\cdot b^T{{\mathcal{A}}}^{(\beta)}\, f^{(\beta)}(x) +
O(\tau^{p+1}).$$ The vector ${{\mathcal{A}}}^{(\beta)} \in
{{\mathbb R}}^s$, $\beta=[\beta_1, \ldots,\beta_n]$, is recursively defined by $$\label{eq:abeta}
{{\mathcal{A}}}^{(\beta)}_i = \left({{\mathcal{A}}}\cdot {{\mathcal{A}}}^{(\beta_1)}\right)_i
\cdot\ldots\cdot \left({{\mathcal{A}}}\cdot{{\mathcal{A}}}^{(\beta_n)}\right)_i,\qquad
i=1,\ldots,s.$$
Because of the definition (\[eq:rk\]) of the discrete flow we have to prove that the stages $k_i$ expand to $$k_i = \sum_{\#\beta\leq p} \tau^{\#\beta-1}
\alpha_\beta\,{{\mathcal{A}}}^{(\beta)}_i f^{(\beta)} + O(\tau^{p}).$$ This is obviously the case for $p=0$. We proceed by induction on $p$. Using the assertion for $p$, the definition of the stages $k_i$, the multivariate Taylor formula and the multilinearity of the derivatives we obtain $$\begin{split}
k_i &=
f\left(x + \tau \left(\sum_{\#\beta\leq p} \tau^{\#\beta-1}
\alpha_\beta\, \left({{\mathcal{A}}}\cdot {{\mathcal{A}}}^{(\beta)}\right)_i\,
f^{(\beta)} + O(\tau^{p}) \right)\right)\\ &= \sum_{n = 0}^p
\frac{1}{n!} f^{(n)} \cdot \left( \sum_{\#\beta_1\leq p}
\tau^{\#\beta_1} \alpha_{\beta_1}\,\left({{\mathcal{A}}}\cdot
{{\mathcal{A}}}^{(\beta_1)}\right)_i\, f^{(\beta_1)},\ldots\right.\\
&\qquad\qquad\ldots\left., \sum_{\#\beta_n\leq p}
\tau^{\#\beta_n} \alpha_{\beta_n}\,\left({{\mathcal{A}}}\cdot
{{\mathcal{A}}}^{(\beta_n)}\right)_i \, f^{(\beta_n)} \right) \;+
O(\tau^{p+1})\\
&=\sum_{n=0}^p
\frac{1}{n!}\sum_{\#\beta_1+\ldots+\#\beta_n\leq p}
\tau^{\#\beta_1+\ldots+\#\beta_n} \cdot
\alpha_{\beta_1}\cdot\ldots\cdot\alpha_{\beta_n}\cdot\left({{\mathcal{A}}}\cdot {{\mathcal{A}}}^{(\beta_1)}\right)_i \cdot \\ &\;
\ldots\cdot
\left({{\mathcal{A}}}\cdot{{\mathcal{A}}}^{(\beta_n)}\right)_i \, f^{(n)} \cdot \left(
f^{(\beta_1)},\ldots,f^{(\beta_n)}\right) + O(\tau^{p+1})
\end{split}$$ $$\begin{split}
&= \sum_{n=0}^p \;\sum_{\substack{
\beta =[\beta_1,\ldots,\beta_n]\\
\#\beta \leq p+1
}}
\tau^{\#\beta-1} \cdot \underbrace{\frac{\delta_\beta}{n!}\,
\alpha_{\beta_1}\cdot\ldots\cdot\alpha_{\beta_n}}_{=\alpha_\beta}
\cdot\,{{\mathcal{A}}}^{(\beta)}_i \,f^{(\beta)} + O(\tau^{p+1})\\
&= \sum_{\#\beta \leq p+1} \tau^{\#\beta-1}
\alpha_\beta\cdot{{\mathcal{A}}}^{(\beta)}_i\,f^{(\beta)} +O(\tau^{p+1}),
\end{split}$$ which proves the assertion for $p+1$.
Comparing the coefficients of the elementary differentials in the expansions of both the phase flow and the discrete flow, we immediately obtain Butcher’s theorem [@Butc63].
\[satz:butcher\] A Runge-Kutta-method $(b,{{\mathcal{A}}})$ is of order $p \in {{\mathbb N}}$ for all $f \in C^p(\Omega,{{\mathbb R}}^d)$, if the order conditions $$\label{eq:cond}
b^T {{\mathcal{A}}}^{(\beta)} = 1/\beta !$$ are satisfied for all rooted trees $\beta$ of order $\#\beta\leq p$.
Generating order conditions with [*Mathematica*]{}
==================================================
The recursive constructions underlying the proof of Butcher’s theorem can easily be realized using modern computer algebra systems like [*Mathematica*]{}.[^1] We assume that the reader is familiar with this particular package.
We begin by defining the recursive data-structure of a rooted tree as an unordered list—together with two simple routines for input and output:
[Attributes\[Tree\]={Orderless,Listable};\
ToTree\[f\_\]:=Tree@@ToTree/@f; ToTree\[f\_Symbol\]:=Tree\[\];\
Format\[$\boldsymbol\beta$\_Tree\]:=StringReplace\[ToString\[List@@$\boldsymbol\beta$/.{}->“$\boldsymbol\odot$”\],\
{“{”->“\[”,“}”->“\]”}\] ]{}
Now, here is an example for the input of the tree representing the elementary differential $f''(f''(f,f'(f)),f) =
f''(f,f''(f'(f),f))$:
[$\boldsymbol\beta_{\tt 1}$=ToTree\[f${^{\prime\prime}}$\[f$^{\prime\prime}$\[f,f$^{\prime}$\[f\]\],f\]\]; $\boldsymbol\beta_{\tt
2}$=ToTree\[f$^{\prime\prime}$\[f,f$^{\prime\prime}$\[f$^{\prime}$\[f\],f\]\]\]; If\[$\boldsymbol\beta_{\tt 1}$==$\boldsymbol\beta_{\tt
2}$,$\boldsymbol\beta_{\tt 1}$,,\]]{}
$$[\odot ,[\odot ,[\odot ]]]$$ The definition (\[eq:treedef\]) of the order $\# \beta$ is simply expressed by the recursive procedure:
[TreeOrder\[$\boldsymbol\beta\_\,$\]:=1+Plus@@TreeOrder/@$\boldsymbol\beta$]{}
As an example, we take the elementary differential $f''(f'''(f'(f),f'(f),f),f)$:
[$\boldsymbol\beta$ = ToTree\[f$^{\prime\prime}$\[f$^{\prime\prime\prime}$\[f$^{\prime}$\[f\],f$^{\prime}$\[f\],f\],f\]\]; TreeOrder\[$\boldsymbol\beta$\]]{}
$${8}$$
The definition (\[eq:factorialdef\]) of $\beta !$ is analogously expressed by:
[ TreeFactorial\[$\boldsymbol\beta$\_\]:=TreeOrder\[$\boldsymbol\beta$\]Times@@TreeFactorial/@$\boldsymbol\beta$ ]{}
The above used tree $\beta$ of order 8 gives: [TreeFactorial\[$\boldsymbol\beta$\] ]{} $${192}$$
For the sake of completeness we also express the recursive definition (\[eq:factorialdef\]) of $\alpha_\beta$ using [*Mathematica*]{}:
[TreeAlpha\[$\boldsymbol\beta$\_\]:=Length\[Permutations\[$\boldsymbol\beta$\]\]/Length\[$\boldsymbol\beta$\]!Times@@TreeAlpha/@$\boldsymbol\beta$]{}
An application to the above example yields: [TreeAlpha\[$\boldsymbol\beta$\]]{} $${\frac{1}{2}}$$
We are now ready to map the recursive definition (\[eq:abeta\]) of ${{\mathcal{A}}}^{(\beta)}$ and of the order condition $b^T {{\mathcal{A}}}^{(\beta)} = 1/\beta!$ to [*Mathematica*]{}:
[TreeA\[$\boldsymbol\beta$\_,n\_:1\]:=(vars={i,j,k,l,m,p,q,r,u,v,w};\
Times@@(Sum\[a$_{\tt{vars[[n]],vars[[n+1]]}}$TreeA\[\#,n+1\]//Evaluate,\
{vars\[\[n+1\]\],s}//Evaluate\]&/@$\boldsymbol\beta$))\
TreeOrderCondition\[$\boldsymbol\beta$\_\]:=Sum\[b$_{\tt i}$TreeA\[$\boldsymbol\beta$\]//Evaluate,{i,s}\]\
==1/TreeFactorial\[$\boldsymbol\beta$\] ]{}
For convenience, the coordinate index of the vector ${{\mathcal{A}}}^{(\beta)}$ can be chosen from the list [{i,j,k,l,m,p,q,r,u,v,w}]{} and is passed by number as the second argument to [TreeA]{}. The order condition belonging to the above example is obtained as follows:
[TreeOrderCondition\[$\boldsymbol\beta$\] ]{}
$${\sum _{i=1}^{s}{b_i} \bigg(\sum _{j=1}^{s}{a_{i,j}}\bigg) \sum _{j=1}^{s}{a_{i,j}} \bigg(\sum _{k=1}^{s}{a_{j,k}}\bigg)
{{\Bigg(\sum _{k=1}^{s}{a_{j,k}} \sum
_{l=1}^{s}{a_{k,l}}\Bigg)}^2}==\frac{1}{192}}$$
Even the typesetting of this formula was done completely automatically, using [*Mathematica*]{}’s ability to generate TeX-sources.
To generate all the order conditions for a given order $p$, we need a device that constructs the set of all trees $\beta$ with $\# \beta
\leq p$. There are, in principle, two different recursive approaches:
- [*root-oriented*]{}: generate all trees $\beta$ of order $\#
\beta = p$ by first, listing all integer partitions $p-1 = p_1 +
\ldots + p_n$, $n=1,\ldots,p-1$, and next, setting $\beta = [\beta_1,\ldots,\beta_n]$ for all trees $\beta_1,\ldots,\beta_n$ of order $\# \beta_1 =
p_1 < p ,\ldots,\#\beta=p_n < p$. These trees have already been generated by the recursion.
- [*leaf-oriented*]{}: Add a leaf to each node of the trees $\hat\beta = [\beta_1,\ldots,\beta_n]$ of order $\# \hat\beta = p-1$, increasing thereby the order exactly by one. This can be done recursively by adding a leaf to every node of the subtrees $\beta_1,\ldots,\beta_n$.
The root-oriented approach was chosen by Sofroniou [@Sofr94] in his [*Mathematica*]{} package [Butcher.m]{}. It requires an efficient integer partition package and the handling of cartesian products. The leaf-oriented approach is as least as efficient as the other one, but much easier to code:
[AddLeave\[$\boldsymbol\beta$\_\]:=AddLeave\[$\boldsymbol\beta$\]=Fold\[Union,{Prepend\[$\boldsymbol\beta$,Tree\[\]\]},\
ReplacePart\[$\boldsymbol\beta$,AddLeave\[$\boldsymbol\beta$\[\[\#\]\]\],\#\]&/@Range\[Length\[$\boldsymbol\beta$\]\]\]\
Trees\[order\_\]:=NestList\[Union@@AddLeave/@\#&,{Tree\[\]},order-1\] ]{}
Given an order $p$ this procedure generates a list of the sets of trees for each order $q \leq p$, e.g., [Trees\[4\] ]{} $$\{\{\odot \},\{[\odot ]\},\{[[\odot ]],[\odot , \odot ]\},
\{[[[\odot ]]],[[\odot ,\odot ]],[\odot ,[\odot ]],[\odot ,\odot , \odot ]\}\}$$ For instance, the number of trees for each order $p \leq 10$ is given by the entries of the following list: [Length/@Trees\[10\] ]{} $${\{1,1,2,4,9,20,48,115,286,719\}}$$ The number of order conditions for $p=10$ can thus be obtained by: [Plus@@% ]{} $$1205$$
Finally, for concrete calculations one has to specify the number $s$ of stages. The following procedure then generates the specific set of equations for [*explicit*]{} Runge-Kutta methods:
[ explicit={a$_{\tt i\_\,,j\_}$:>0/;i<=j,c$_{\tt 1}$->0};\
OrderConditions\[order\_,stages\_\]:=(\
autonomous=Table\[Sum\[a$_{\tt i,j}$,{j,stages}\]==c$_{\tt i}$,{i,stages}\];\
{(TreeOrderCondition/@Union@@Trees\[order\])/.s->stages/.\
ToRules\[And@@autonomous\],autonomous}/.explicit) ]{}
This way, we can automatically generate and typeset the order conditions for the classical explicit 4-stage Runge-Kutta methods of order 4:
[First\[OrderConditions\[4,4\]\] ]{}
$$\begin{split}
\big\{& {b_1}+{b_2}+{b_3}+{b_4}==1,\;{b_2}{\,}{c_2}+{b_3}
{c_3}+{b_4}{\,}{c_4}==\frac{1}{2}, \\*[2mm] & {b_3}{\,}{c_2}{\,}{a_{3,2}}+{b_4}{\,}({c_2}{\,}{a_{4,2}}+{c_3}{\,}{a_{4,3}})==\frac{1}{6},\;{b_4}{\,}{c_2}{\,}{a_{3,2}}{\,}{a_{4,3}}==\frac{1}{24}, \\*[2mm] &
{b_3}{\,}c_{2}^{2}{\,}{a_{3,2}}+{b_4}{\,}(c_{2}^{2}{\,}{a_{4,2}}+c_{3}^{2}{\,}{a_{4,3}})==\frac{1}{12},\;{b_2}{\,}c_{2}^{2}+{b_3}{\,}c_{3}^{2}+{b_4}{\,}c_{4}^{2}==\frac{1}{3}, \\*[2mm] & {b_3}{\,}{c_2}{\,}{c_3}{\,}{a_{3,2}}+{b_4}{\,}{c_4}{\,}({c_2}{\,}{a_{4,2}}+{c_3}{\,}{a_{4,3}})==\frac{1}{8},\;{b_2}{\,}c_{2}^{3}+{b_3}{\,}c_{3}^{3}+{b_4}{\,}c_{4}^{3}==\frac{1}{4}\big\}
\end{split}$$
[Even in the more recent literature one can find examples like [@GaGu99], where order conditions for Runge-Kutta methods are generated by using a computer algebra system to calculate the Taylor expansions of the flow and the discrete flow directly. This approach is typically bound to [*scalar*]{} non-autonomous equations, i.e., $d=1$. Besides being inefficient for higher orders, it is well-known [@Butc63b] that for $p \geq 5$ additional order conditions for general systems make an appearance, which do not show up in the scalar case.]{}
Examples of usage
=================
The following simple procedure tempts to solve the order conditions for a given order $p$ and stage number $s$ by using brute force, i.e., [*Mathematica*]{}’s [Solve]{}-command. To simplify the task, the user is allowed to supply a set [pre]{} of a priori chosen additional equations and assignments that he thinks to be helpful.
[RungeKuttaMethod\[p\_,s\_,pre\_\]:=(\
rkTemplate={Table\[a$_{\tt i, j}$,{i,s},{j,s}\],Table\[b$_{\tt
i}$,{j,1},{i,s}\],\
Table\[c$_{\tt i}$,{j,1},{i,s}\]}/.explicit;\
conditions=pre$\,\boldsymbol\cup\,$Union@@OrderConditions\[p,s\];\
solveVars=Complement\[Flatten\[rkTemplate\],{0}\];\
sol=Solve\[conditions,solveVars\];\
Thread\[{A,b,c}==MatrixForm/@rkTemplate/.\#\]&/@sol) ]{}
Since [*Mathematica*]{}’s [Solve]{}-command uses a Gröbner-basis approach for solving systems of polynomial equations, we can show this way that the classical explicit 4-stage Runge-Kutta method of order $p=4$ is uniquely given by the additional constraints $b_2=b_3$ and $c_2=c_3$: [RungeKuttaMethod\[4,4,{b$_{\tt
2}$==b$_{\tt 3}$,c$_{\tt
2}$==c$_{\tt 3}$}\] ]{} $$\Big\{\Big\{A==\left(
\begin{matrix}
0&0&0&0 \\
\frac{1}{2}&0&0&0 \\
0&\frac{1}{2}&0&0 \\
0&0&1&0
\end{matrix} \right),b==\left(\begin{matrix}
\frac{1}{6}&\frac{1}{3}&\frac{1}{3}&\frac{1}{6}
\end{matrix}\right),c==\left(\begin{matrix}
0&\frac{1}{2}&\frac{1}{2}&1
\end{matrix} \right)\Big\}\Big\}$$
The next example is more demanding. In his book [@Butc87 p. 199], Butcher describes an algorithm for the construction of explicit 6-stage methods of order $p=5$. The choices $c_6=1$ and $b_2=0$ together with the free parameters $c_2,c_3,c_4,c_5$ and $a_{43}$ yield a unique method. Butcher provides a two-parameter example by choosing $c_2 = u,c_3=1/4,c_4=1/2,c_5=3/4,a_{43}=v$. By just passing this additional information to [*Mathematica*]{}’s [Solve]{}-command we obtain the following solution:
[ pre = {c$_{\tt 2}$==u,c$_{\tt 3}$==1/4,c$_{\tt 4}$==1/2,c$_{\tt 5}$==3/4,c$_{\tt 6}$==1,b$_{\tt
2}$==0,a$_{\tt 4,3}$==v};\
First\[RungeKuttaMethod\[5,6,pre\]\] ]{}
$$\begin{split}
\Big\{& A==\left(\begin{matrix}
0&0&0&0&0&0 \\*[2mm]
u&0&0&0&0&0 \\*[2mm]
\frac{-1+8{\,}u}{32{\,}u}&\frac{1}{32{\,}u}&0&0&0&0 \\*[2mm]
\frac{-1+4{\,}u+2{\,}v-8{\,}u{\,}v}{8{\,}u}&
\frac{1-2{\,}v}{8{\,}u}&v&0&0&0 \\*[2mm]
\frac{3{\,}(1-3{\,}u-v+4{\,}u{\,}v)}{16{\,}u}&
\frac{3{\,}(-1+v)}{16{\,}u}&
-\frac{3}{4}{\,}(-1+v)&\frac{9}{16}&0&0\\*[2mm]
\frac{-7+22{\,}u+6{\,}v-24{\,}u{\,}v}{14{\,}u}&
\frac{7-6{\,}v}{14{\,}u}&\frac{12{\,}v}{7}&-\frac{12}{7}&\frac{8}{7}&0
\end{matrix} \right)\\*[2mm]
& b==\left(\begin{matrix}
\frac{7}{90} &
0 &
\frac{16}{45} &
\frac{2}{15} &
\frac{16}{45} &
\frac{7}{90}
\end{matrix} \right),\quad c==\left(\begin{matrix}
0 &
u &
\frac{1}{4} &
\frac{1}{2} &
\frac{3}{4} &
1
\end{matrix} \right)\quad \Big\}
\end{split}$$
This result shows that the coefficients $a_{51}$ and $a_{52}$ of Butcher’s solution [@Butc87 p. 199] are in error, a fact that was already observed by Sofroniou [@Sofr94] using the [*Mathematica*]{} package [Butcher.m]{}.
[1]{}
J. C. Butcher, *Coefficients for the study of [Runge-Kutta]{} integration processes*, J. Austral. Math. Soc. **3** (1963), 185–201.
[to3em]{}, *On the integration processes of [A. Huta]{}*, J. Austral. Math. Soc. **3** (1963), 202–206.
[to3em]{}, *The numerical analysis of ordinary differential equations. [Runge-Kutta]{} methods and general linear methods*, [John Wiley & Sons]{}, Chichester, 1987.
W. Gander and D. Gruntz, *Derivation of numerical methods using computer algebra*, SIAM Review **41** (1999), 577–593.
E. Hairer, *Numerical geometric integration. [Lecture notes of a course given in 1998/99]{}*, [http://www.unige.ch/math/folks/hairer/polycop.html]{}, 1999.
E. Hairer, S. P. [Nørsett]{}, and G. Wanner, *Solving ordinary differential equations i. nonstiff problems*, 2. ed., [Springer-Verlag]{}, Berlin, Heidelberg, New York, 1993.
E. Hairer and G. Wanner, *On the [Butcher]{} group and general multi-value methods*, Computing **13** (1974), 1–15.
R. J. Jenks, *Problem \# 11: Generation of [Runge-Kutta]{} equations*, SIGSAM Bulletin **10** (1976), 6.
M. Sofroniou, *Sysmbolic derivation of [Runge-Kutta]{} methods*, J. Symb. Comp. **18** (1994), 265–296.
[^1]: A [*Mathematica*]{} notebook with all the programs and examples of this e-print is included with the source files: [http://arxiv.org/e-print/math.NA/0211049.tar.gz]{} .
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We prove the necessity of the UMD condition, with a quantitative estimate of the UMD constant, for any inequality in a family of $L^p$ bounds between different partial derivatives $\partial^\beta u$ of $u\in C^\infty_c({{\mathbb{R}}}^n,X)$. In particular, we show that the estimate $\|u_{xy}\|_p\leq K(\|u_{xx}\|_p+\|u_{yy}\|_p)$ characterizes the UMD property, and the best constant $K$ is equal to one half of the UMD constant. This precise value of $K$ seems to be new even for scalar-valued functions.'
address:
- ' A.J. Castro, Departamento de Análisis Matemático, Universidad de la Laguna, Campus de Anchieta, Avda. Astrofísico Francisco Sánchez, s/n, 38271, La Laguna (Sta. Cruz de Tenerife), Spain'
- ' T.P. Hytönen, Department of Mathematics and Statistics, University of Helsinki, Gustaf Hällströmin katu 2b, FI-00014 Helsinki, Finland'
author:
- 'Alejandro J. Castro'
- 'Tuomas P. Hytönen'
title: |
Bounds for partial derivatives:\
necessity of UMD and sharp constants
---
Introduction {#sec:intro}
============
A priori estimates between different partial derivatives such as $$\label{eq:estim}
\|\partial^\beta u\|_{L^{q}({{\mathbb{R}}}^n,X)}
\leq K \sum_{j=1}^N \|\partial^{\alpha^j} u\|_{L^{p_j}({{\mathbb{R}}}^n,X)}, \quad u \in C_c^\infty({{\mathbb{R}}}^n,X),$$ play an important role in Analysis. Here $\partial^\beta=\partial_1^{\beta_1} \cdots \partial_n^{\beta_n}$ is the usual derivative associated to the multi-index $\beta=(\beta_1, \dots, \beta_n)$. In the case of scalar-valued functions, $X={\mathbb{C}}$ or $X={{\mathbb{R}}}$, and exponents in the range $q,p_j\in(1,\infty)$, a complete characterization of admissible sets of $q,p_j$ and $\beta,\alpha^j$ is a classical result of O. Besov, V. Il$'$in and S. Nikol$'$ski[ĭ]{} [@BNS]. The second author showed in [@Hy4] that the same characterization remains valid for vector-valued functions, provided that the target space $X$ has the UMD property.
A Banach space $X$ has this property if, for some (equivalently, for all) $1<p<\infty$, there exists a constant $C_p<\infty$, such that $$\label{eq:defUMD}
\Big\| \sum_{\ell=1}^r \sigma_\ell d_\ell \Big\|_{L^p(\Omega,X)}
\leq C_p \Big\| \sum_{\ell=1}^r d_\ell \Big\|_{L^p(\Omega,X)},$$ for each $r \in {\mathbb{N}}$, every sequence of signs $(\sigma_\ell)_{\ell \in {\mathbb{N}}}$ in $\{-1,1\}$ and any martingale difference sequence $(d_\ell)_{\ell \in {\mathbb{N}}}$ in $L^p(\Omega,X)$. Here $L^p(\Omega,X)$ represents the space of all functions that take values in the Banach space $X$ and their $X$ norm is $p$-integrable on $\Omega$. We denote by $\beta_p(X)$ the smallest admissible constant $C_p$. Main properties of UMD spaces can be found in the survey of J. L. Rubio de Francia [@Rub].
Our present goals are two-fold. On the one hand, we show that the result of [@Hy4] is optimal in the sense that it establishes the most general class of Banach spaces where the classical Besov–Il’in–Nikol’ski[ĭ]{} result [@BNS] can be generalized; namely, we exhibit instances of , for which UMD is not only sufficient (as shown in [@Hy4]) but also necessary. This continues a tradition of diverse characterizations of the UMD property (see for instance [@Bou3; @Bu2; @CL; @GMS; @Gue; @HTV; @Hy3; @Hy; @KaWe; @KuWe; @Xu], amongst others). On the other hand, as a byproduct, we obtain information on the size of the constant $K$ in , which seems to be new already in the scalar case. Recall that the list of analytic inequalities, where the sharp constant can be determined, is somewhat restricted, whereas the UMD inequality for $X={\mathbb{C}}$ or $X={{\mathbb{R}}}$ is one of the prominent positive examples with $\beta_p({\mathbb{C}})=\beta_p({{\mathbb{R}}})=(p^*-1)$; here $p^*=\!
max\{p,p'\}$. This is a celebrated result of Burkholder; see [@Bu5], [@Bu4 p. 12] and [@Bu6 Theorem 14]. Thus it is useful to relate other estimates to .
A particular case of that we have in mind is the basic bound $$\label{eq:mainExample}
\|\partial_1\partial_2 u\|_{L^p({{\mathbb{R}}}^2,X)}
\leq K(\|\partial_1^2 u\|_{L^p({{\mathbb{R}}}^2,X)}+\|\partial_2^2 u\|_{L^p({{\mathbb{R}}}^2,X)}).$$ Since $\partial_1\partial_2 u=R_1R_2\triangle u$, where $R_i$ is the Riesz transform in the $i$th direction, it follows at once that $$\label{eq:R12comput}
\begin{split}
\|\partial_1\partial_2 u\|_{L^p({{\mathbb{R}}}^2,X)}
&\leq\|R_1R_2\|_{\mathcal{L}(L^p({{\mathbb{R}}}^2,X))}\|\partial_1^2 u+\partial_2^2 u\|_{L^p({{\mathbb{R}}}^2,X)} \\
&\leq\|R_1R_2\|_{\mathcal{L}(L^p({{\mathbb{R}}}^2,X))}(\|\partial_1^2 u\|_{L^p({{\mathbb{R}}}^2,X)}+\|\partial_2^2 u\|_{L^p({{\mathbb{R}}}^2,X)}),
\end{split}$$ and $\|R_1R_2\|_{\mathcal{L}(L^p({{\mathbb{R}}}^2,X))}=\beta_p(X)/2$ by [@GMS], so that $K\leq \beta_p(X)/2$ in . We prove that this constant, optimal for the first line in , is also sharp for the seemingly weaker inequality .
It is interesting that the best constant in such a simple-looking classical inequality appears to be previously unknown. This constant is somewhat connected to the famous open problem of determining the $L^p$ norm of the Beurling transform, whose imaginary part is $2R_1R_2$; see [@GMS] for more details. It has been long conjectured [@Iwa] that the norm of the Beurling operator on $L^p({{\mathbb{R}}}^2)$ is $\beta_p({\mathbb{C}})$, and a surprising contribution of [@GMS] was to show that this conjectured norm is already achieved by the imaginary part alone. Our result further amplifies this phenomenon.
We then turn to the precise formulation of our main results. We are only concerned with the case that $p_j\equiv q=:p$ in , and all multi-indices $\beta,\alpha^j$ have equal length, which we denote by $|\beta|=\beta_1 + \dots +\beta_n$, when $\beta=(\beta_1, \dots, \beta_n) \in {\mathbb{N}}^n$. We also need to assume that the components of $\beta$ have sufficiently different parity from those of each $\alpha^j$.
\[Th:main\] Let $1<p<\infty$, $N \in {\mathbb{N}}$ and $\alpha^j$, $\beta \in {\mathbb{N}}^n$. Let $|\beta|=|\alpha^j|$ for all $j=1, \ldots, N$, and suppose that there exists a set $F\subset\{1,\ldots,n\}$ such that $\sum_{\ell\in F}\alpha^j_\ell$ has the same parity (even or odd) for each $j=1,\ldots,N$, which is different from the parity of $\sum_{\ell\in F}\beta_{\ell}$. If the following estimate holds: $$\label{eq:estim2}
\|\partial^\beta u\|_{L^p({{\mathbb{R}}}^n,X)}
\leq K \sum_{j=1}^N \|\partial^{\alpha^j} u\|_{L^p({{\mathbb{R}}}^n,X)}, \quad u \in C_c^\infty({{\mathbb{R}}}^n,X),$$ then $X$ is UMD and $\beta_p(X) \leq KN$.
It is known (see e.g. [@Hy4]) that a necessary condition for to hold in any Banach space is that $\beta$ is a convex combination of the multi-indices $\alpha^j$. Although we will not explicitly use this fact, it is worth observing that the assumption is only meaningful for such sets of multi-indices.
The following corollaries demonstrate the somewhat technical conditions on the multi-indices assumed in the theorem:
Let $X$ be a Banach space and $1<p<\infty$. Let $\beta\in{\mathbb{N}}^n$ be a multi-index such that $|\beta|$ is even, but $\beta_s$ is odd for at least one index $s\in\{1,\ldots,n\}$. Then the estimate $$\|\partial^\beta u\|_{L^p({{\mathbb{R}}}^n,X)}\leq K\sum_{j=1}^n\|\partial_j^{|\beta|}u\|_{L^p({{\mathbb{R}}}^n,X)},\qquad u\in C^\infty_c({{\mathbb{R}}}^n,X),$$ holds if and only if $X$ is a UMD space, and in this case $\beta_p(X)\leq Kn$.
The “if” part is well known; for instance, it is a special case of the results of [@Hy4]. For the “only if” part, we apply Theorem \[Th:main\] with $\alpha^j=|\beta|e_j$ and $F=\{s\}$. Then $\sum_{\ell\in F}\beta_\ell=\beta_s$ is odd, whereas $\sum_{\ell\in F}\alpha^j_\ell=|\beta|\delta_{js}$ is even for all $j$, as both $0$ and $|\beta|$ are even. Thus the Theorem implies that $X$ is UMD, with the asserted estimate for the constant.
\[Cor:bestconst1\] Let $X$ be a Banach space and $1<p<\infty$. Then holds if and only if $X$ is a UMD space, and the best constant $K$ satisfies $K=\beta_p(X)/2$.
We already explained that $K\leq\beta_p(X)/2$ in ; this part of the result is not new. The converse estimate is a special case of the previous Corollary with $n=2$ and $\beta=(1,1)$.
There are many characterizations of the UMD property by estimates between $L^p$ norms of two different objects. A novel qualitative feature of the above characterizing conditions is that they involve a sum of several $L^p$ norms on the right hand side.
The proof of Theorem \[Th:main\] proceeds via the theory of Fourier multipliers. We denote by $\widehat{f}$ the Fourier transform of the function $f$, defined as usual by $$\widehat{f}(\xi)
= \int_{{{\mathbb{R}}}^n} e^{-2 \pi i \xi \cdot x} f(x) dx, \quad \xi \in {{\mathbb{R}}}^n,$$ and by $\widecheck{f}$ its inverse, given as above but without the plus sign in the complex exponential. The Fourier multiplier of a function $m$ is defined by $T_m f = (m \widehat{f})^{\vee}$.
We write $$\label{eq:mmj}
m(\xi)= \frac{\xi^\beta}{|\xi|^{|\beta|}}
\quad \text{ and } \quad
m_j(\xi)= \frac{\xi^{\alpha^j}}{|\xi|^{|\alpha^j|}}, \qquad \beta, \alpha^j \in {\mathbb{N}}^n, \ j=1, \dots, N,$$ where $\xi^\beta=\xi_1^{\beta_1} \cdots \xi_n^{\beta_n}$. By well-known properties of the Fourier transform we have that $$\partial^\beta u = T_m (\Delta^{|\beta|/2}u),$$ where $\Delta=-\partial_1^2 - \dots - \partial_n^2$ is the positive Laplacian in ${{\mathbb{R}}}^n$ and $\Delta^{|\beta|/2}u=[(2\pi i|\xi|)^{|\beta|} \widehat{u}]^{\vee}$. Hence, with $|\alpha^j|=|\beta|$, $j=1, \dots N$, implies that $$\label{eq:estim3}
\|T_m v\|_{L^p({{\mathbb{R}}}^n,X)}
\leq K \sum_{j=1}^N \| T_{m_j}v\|_{L^p({{\mathbb{R}}}^n,X)}, \quad v \in \mathcal{S}_0({{\mathbb{R}}}^n,X),$$ where $\mathcal{S}_0({{\mathbb{R}}}^n,X)$ is the subspace of the Schwartz class constituted by functions having zero mean value. Moreover, in order to prove that $X$ is UMD it is more convenient to have an estimate of type , but in terms of discrete Fourier multipliers (see Propositions \[Prop:RnTn\] and \[Prop:TnTrn\] below).
This paper is organized as follows. In Section \[sec:transf\] we collect some transference results about Fourier multipliers and in Section \[sec:proofTh\] we present the proof of Theorem \[Th:main\].
Transference results for Fourier multipliers {#sec:transf}
============================================
Throughout this section we assume that $m$ is a measurable bounded function verifying that $$\label{eq:mk}
m(k)=\lim_{r \to 0} \frac{1}{|B(k,r)|} \int_{B(k,r)}m(\xi)d\xi,$$ exists for every $k \in {\mathbb{Z}}^n$. The discrete Fourier multiplier $T_{\widetilde{m}}f$, of a function $f$ defined in the $n$-dimensional torus ${{\mathbb{T}}}^n=[-1/2,1/2)^n$, is given by $$\label{eq:Tmdiscret}
T_{\widetilde{m}}f (t)
= \sum_{k \in {\mathbb{Z}}^n} m(k) \widehat{f}(k) e_k(t), \quad t \in {{\mathbb{T}}}^n,$$ where $e_k(t)=e^{2\pi i t \cdot k}$ and $\widehat{f}(k)=\int_{{{\mathbb{T}}}^n} e_k(-t) f(t)dt$.
Next, we present some lemmas which will be very useful in the proof of Proposition \[Prop:RnTn\]. We say that $f$ is a trigonometric polynomial in $X$ if $$f(t)
= \sum_{k \in {\mathbb{Z}}^n} a_k e_k(t), \quad t \in {{\mathbb{T}}}^n,$$ with finitely many nonzero coefficients $a_k \in X$. Note that $a_k=\widehat{f}(k)$.
\[Lem:11.20\] Let $\phi$, $\psi \in \mathcal{S}({{\mathbb{R}}}^n)$ be radial functions satisfying $\int_{{{\mathbb{R}}}^n} \phi \psi d\xi=1 $. Then, $$\int_{{{\mathbb{T}}}^n}\langle g(t), T_{\widetilde{m}}f(t) \rangle dt
= \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^n \int_{{{\mathbb{R}}}^n}\langle \psi({{\varepsilon}}x)g(x) , T_m(\phi({{\varepsilon}}\cdot )f)(x)\rangle dx,$$ for every trigonometric polynomials $f:{{\mathbb{T}}}^n \longrightarrow X$, $g:{{\mathbb{T}}}^n \longrightarrow X^*$. Here $\langle \cdot , \cdot \rangle$ denotes the pairing between $X$ and its dual $X^*$.
The key step in this proof is the identity $$\label{eq:11.21}
\lim_{{{\varepsilon}}\to 0} \frac{1}{{{\varepsilon}}^n}
\int_{{{\mathbb{R}}}^n} m(\xi) \widehat{\phi}\Big(\frac{\xi - k}{{{\varepsilon}}} \Big) \widecheck{\psi}\Big(\frac{\xi - \ell}{{{\varepsilon}}} \Big) d\xi
= \chi_{\{k=\ell\}} m(k), \quad k, \ell \in {\mathbb{Z}}^n,$$
Assume first that $k=\ell$. Since the Fourier transform (and its inverse) of a radial function, is also radial (see for example [@Ste2 p. 430]), we consider $\rho(|\xi|)=\widehat{\phi}(\xi) \widecheck{\psi}(\xi)$, which satisfies $\int_{{{\mathbb{R}}}^n}\rho(|\xi|)d\xi=\int_{{{\mathbb{R}}}^n}\widehat\phi\widecheck\psi\,d\xi=\int_{{{\mathbb{R}}}^n}\phi\psi\,dx=1$. We have that $$\begin{aligned}
& \frac{1}{{{\varepsilon}}^n} \int_{{{\mathbb{R}}}^n} m(\xi) \widehat{\phi}\Big(\frac{\xi - k}{{{\varepsilon}}} \Big) \widecheck{\psi}\Big(\frac{\xi - k}{{{\varepsilon}}} \Big) d\xi
= \int_{{{\mathbb{R}}}^n} m({{\varepsilon}}\eta + k) \rho(|\eta|) d\eta
= - \int_{{{\mathbb{R}}}^n} m({{\varepsilon}}\eta + k) \int_{|\eta|}^\infty \rho'(r) dr d\eta \\
& \qquad = - \int_0^\infty \rho'(r) \int_{B(0,r)} m({{\varepsilon}}\eta + k) d\eta dr
= - \int_0^\infty \rho'(r) \frac{|B(0,r)|}{|B(k,{{\varepsilon}}r)|}\int_{B(k,{{\varepsilon}}r)} m(\xi) d\xi dr.
\end{aligned}$$ By taking into account $\rho \in \mathcal{S}({{\mathbb{R}}})$, $m \in L^\infty(R^n)$ and , we can apply dominated convergence and then integrate by parts to obtain $$\begin{aligned}
& - \lim_{{{\varepsilon}}\to 0} \int_0^\infty \rho'(r) \frac{|B(0,r)|}{|B(k,{{\varepsilon}}r)|}\int_{B(k,{{\varepsilon}}r)} m(\xi) d\xi dr
= - m(k) |B(0,1)| \int_0^\infty r^n \rho'(r) dr \\
& \qquad = m(k) |B(0,1)| \int_0^\infty n r^{n-1} \rho(r) dr
= m(k) \int_0^\infty \int_{\partial B(0,r)} \rho(r) d\sigma dr
= m(k) \int_{{{\mathbb{R}}}^n} \rho(|\xi|) d\xi
= m(k),
\end{aligned}$$ because $|\partial B(0,r)|=|B(0,1)| n r^{n-1}$.
Suppose now that $k \neq \ell$. This time, $$\frac{1}{{{\varepsilon}}^n} \int_{{{\mathbb{R}}}^n} m(\xi) \widehat{\phi}\Big(\frac{\xi - k}{{{\varepsilon}}} \Big) \widecheck{\psi}\Big(\frac{\xi - \ell}{{{\varepsilon}}} \Big) d\xi
= \int_{{{\mathbb{R}}}^n} m({{\varepsilon}}\eta + k) \widehat{\phi}(\eta) \widecheck{\psi}\Big(\eta + \frac{k - \ell}{{{\varepsilon}}} \Big) d\eta,$$ where the integrand is bounded by $\|m\|_\infty \|\widecheck\psi\|_\infty |\widehat\phi(\eta)| \in L^1(d\eta)$, and converges pointwise to zero as ${{\varepsilon}}\to 0$, since $\widecheck{\psi}(\eta + (k - \ell)/{{\varepsilon}})\to 0$. Applying again dominated convergence we readily see that the integral converges to zero as ${{\varepsilon}}\to 0$.
Take the trigonometric polynomials $f=\sum_{k \in {\mathbb{Z}}^n} a_k e_k$ and $g=\sum_{\ell \in {\mathbb{Z}}^n} b_\ell e_\ell$ with $a_k \in X$ and $b_\ell \in X^*$. An application of and making use of the usual properties of the Fourier transform, we deduce $$\begin{aligned}
& \int_{{{\mathbb{T}}}^n}\langle g(t), T_{\widetilde{m}}f(t) \rangle dt
= \sum_{\ell, k \in {\mathbb{Z}}^n} \langle b_\ell , a_k \rangle m(k) \int_{{{\mathbb{T}}}^n} e_\ell(t) e_k(t) dt
= \sum_{\ell, k \in {\mathbb{Z}}^n} \langle b_{-\ell} , a_k \rangle \chi_{\{k=\ell\}} m(k) \\
& \qquad = \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^n \int_{{{\mathbb{R}}}^n}
\Big \langle \sum_{\ell\in {\mathbb{Z}}^n} b_{-\ell} \frac{1}{{{\varepsilon}}^n} \widecheck{\psi}\Big(\frac{\xi-\ell}{{{\varepsilon}}}\Big),
m(\xi) \sum_{k \in {\mathbb{Z}}^n} a_k \frac{1}{{{\varepsilon}}^n} \widehat{\phi}\Big(\frac{\xi-k}{{{\varepsilon}}}\Big) \Big \rangle d\xi\\
& \qquad = \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^n \int_{{{\mathbb{R}}}^n}
\Big \langle \Big[ \psi({{\varepsilon}}\cdot) g \Big]^{\vee}(\xi),
m(\xi) \Big[ \phi({{\varepsilon}}\cdot) f \Big]^{\wedge}(\xi) \Big \rangle d\xi
= \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^n \int_{{{\mathbb{R}}}^n}\langle \psi({{\varepsilon}}x)g(x) , T_m(\phi({{\varepsilon}}\cdot )f)(x)\rangle dx.
\end{aligned}$$
\[Lem:11.21\] Let $1<p<\infty$. Then, for every $\phi \in \mathcal{S}({{\mathbb{R}}}^n)$ and $f \in L^p({{\mathbb{T}}}^n,X)$, we have that $$\lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p} \| \phi({{\varepsilon}}\cdot) f \|_{L^p({{\mathbb{R}}}^n,X)}
= \| \phi \|_{L^p({{\mathbb{R}}}^n)} \| f \|_{L^p({{\mathbb{T}}}^n,X)}.$$
The periodicity of the function $f$ allow us to write $$\begin{aligned}
{{\varepsilon}}^n \int_{{{\mathbb{R}}}^n} \| \phi({{\varepsilon}}x) f(x)\|_X^p dx
& = {{\varepsilon}}^n \sum_{k \in {\mathbb{Z}}^n} \int_{{{\mathbb{T}}}^n+k} |\phi({{\varepsilon}}x)|^p \|f(x)\|_X^p dx \\
& = \int_{{{\mathbb{T}}}^n} \Big({{\varepsilon}}^n \sum_{k \in {\mathbb{Z}}^n} |\phi({{\varepsilon}}(x-k))|^p \Big) \|f(x)\|_X^p dx,
\end{aligned}$$ where the quantity in parentheses is a Riemann sum of the function $|\phi(\cdot)|^p$, which is uniformly bounded in $x$ and ${{\varepsilon}}$. Now, this lemma is a simple consequence of the dominated convergence theorem.
\[Lem:11.23\] Let $1<p<\infty$ and $p'=p/(p-1)$. Then, for each $f \in \mathcal{S}({{\mathbb{R}}}^n,X)$, we have that $$\lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p'} \Big\| \sum_{k \in {\mathbb{Z}}^n} \widehat{f}({{\varepsilon}}k) e_k \Big\|_{L^p({{\mathbb{T}}}^n,X)}
= \| f \|_{L^p({{\mathbb{R}}}^n,X)}.$$
Recall that we identify ${{\mathbb{T}}}^n=[-1/2,1/2)^n$; whence the notation ${{\varepsilon}}^{-1}{{\mathbb{T}}}^n=[-1/(2{{\varepsilon}}),1/(2{{\varepsilon}}))^n$ is meaningful below. Define $f_{{\varepsilon}}={{\varepsilon}}^{-n}f(\cdot/{{\varepsilon}})$. The Poisson summation formula (see [@Graf Theorem 3.1.17]), tells us that $$\sum_{k \in {\mathbb{Z}}^n} \widehat{f}({{\varepsilon}}k) e_k(t)
= \sum_{k \in {\mathbb{Z}}^n} \widehat{f_{{\varepsilon}}}(k) e_k(t)
= \sum_{k \in {\mathbb{Z}}^n} f_{{\varepsilon}}(t+k)
= \frac{1}{{{\varepsilon}}^n}\sum_{k \in {\mathbb{Z}}^n} f\Big(\frac{t}{{{\varepsilon}}}+\frac{k}{{{\varepsilon}}}\Big), \quad t \in {{\mathbb{T}}}^n.$$ Then, making the change of variables $u=t/{{\varepsilon}}$, $$\begin{aligned}
{{\varepsilon}}^{n/p'} \Big\| \sum_{k \in {\mathbb{Z}}^n} \widehat{f}({{\varepsilon}}k) e_k \Big\|_{L^p({{\mathbb{T}}}^n,X)}
& = {{\varepsilon}}^{-n/p} \Big\| \sum_{k \in {\mathbb{Z}}^n} f\Big(\frac{\cdot}{{{\varepsilon}}}+\frac{k}{{{\varepsilon}}}\Big) \Big\|_{L^p({{\mathbb{T}}}^n,X)}
= \Big\| \sum_{k \in {\mathbb{Z}}^n} f\Big(\cdot +\frac{k}{{{\varepsilon}}}\Big) \Big\|_{L^p({{\varepsilon}}^{-1}{{\mathbb{T}}}^n,X)} \\
& =\|f\|_{L^p({{\varepsilon}}^{-1}{{\mathbb{T}}}^n,X)}
+ \mathcal{O}\Big( \sum_{k \in {\mathbb{Z}}^n \setminus \{0\}} \Big\| f\Big(\cdot +\frac{k}{{{\varepsilon}}}\Big) \Big\|_{L^p({{\varepsilon}}^{-1}{{\mathbb{T}}}^n,X)} \Big).
\end{aligned}$$ It is obvious that $$\|f\|_{L^p({{\varepsilon}}^{-1}{{\mathbb{T}}}^n,X)} \longrightarrow \| f \|_{L^p({{\mathbb{R}}}^n,X)}, \quad \text{as } {{\varepsilon}}\to 0.$$ We analyze the second term. Since $f \in \mathcal{S}({{\mathbb{R}}}^n,X)$, for every $N \in {\mathbb{N}}$, we have that $$\Big\| f\Big(u +\frac{k}{{{\varepsilon}}} \Big)\Big\|_X
\leq C \Big(\frac{{{\varepsilon}}}{|k|}\Big)^{N}, \quad u \in {{\varepsilon}}^{-1}{{\mathbb{T}}}^n.$$ Hence, it is enough to take $N>n$ to see that $$\sum_{k \in {\mathbb{Z}}^n \setminus \{0\}} \Big\| f\Big(\cdot +\frac{k}{{{\varepsilon}}}\Big) \Big\|_{L^p({{\varepsilon}}^{-1}{{\mathbb{T}}}^n,X)}
\leq C {{\varepsilon}}^{N-n/p} \sum_{k \in {\mathbb{Z}}^n \setminus \{0\}} \frac{1}{|k|^N}
\leq C {{\varepsilon}}^{N-n/p} \longrightarrow 0, \quad \text{as } {{\varepsilon}}\to 0.\qedhere$$
We now establish our transference result between ${{\mathbb{R}}}^n$ and ${{\mathbb{T}}}^n$. We say that $m$ is a homogeneous function (of order zero) when $m(\lambda \xi)=m(\xi)$, $\xi \in {{\mathbb{R}}}^n$, $\lambda>0$. We denote $$L^p_0({{\mathbb{T}}}^n,X)=\Big\{f \in L^p({{\mathbb{T}}}^n,X) : \int_{{{\mathbb{T}}}^n}f dt = 0\Big\}.$$
\[Prop:RnTn\] Let $1<p<\infty$, $N \in {\mathbb{N}}$ and $m$, $m_j \in C^\infty({{\mathbb{R}}}^n \setminus \{0\})$, $j=1, \dots, N$, be homogeneous functions. Then, the following assertions are equivalent:
- $\displaystyle \|T_m f\|_{L^p({{\mathbb{R}}}^n,X)}
\leq K \sum_{j=1}^N \| T_{m_j}f\|_{L^p({{\mathbb{R}}}^n,X)}, \quad f \in L^p({{\mathbb{R}}}^n,X),$
- $\displaystyle \|T_{\widetilde{m}} f\|_{L^p({{\mathbb{T}}}^n,X)}
\leq K \sum_{j=1}^N \| T_{\widetilde{m}_j}f\|_{L^p({{\mathbb{T}}}^n,X)}, \quad f \in L^p_0({{\mathbb{T}}}^n,X).$
Let $f \in \mathcal{S}({{\mathbb{R}}}^n,X)$ be such that $\operatorname{supp}\widehat{f} \subset {{\mathbb{R}}}^n \setminus \{0\}$ is compact. Observe that this class of functions is dense in $L^p({{\mathbb{R}}}^n,X)$. By the smoothness of the multiplier $m$, we also have that $T_m f \in \mathcal{S}({{\mathbb{R}}}^n,X)$. Hence, we can apply Lemma \[Lem:11.23\] and get $$\begin{aligned}
\| T_m f \|_{L^p({{\mathbb{R}}}^n,X)}
= & \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p'} \Big\| \sum_{k \in {\mathbb{Z}}^n} \widehat{T_m f}({{\varepsilon}}k) e_k \Big\|_{L^p({{\mathbb{T}}}^n,X)}
= \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p'} \Big\| \sum_{k \in {\mathbb{Z}}^n} m(k) \widehat{f}({{\varepsilon}}k) e_k \Big\|_{L^p({{\mathbb{T}}}^n,X)} \\
= & \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p'} \Big\| T_{\widetilde{m}}\Big(\sum_{k \in {\mathbb{Z}}^n} \widehat{f}({{\varepsilon}}k) e_k \Big) \Big\|_{L^p({{\mathbb{T}}}^n,X)}.
\end{aligned}$$ Notice that in the second equality we have applied that $m$ is a homogeneous function. Now, since the function between parentheses is in $L^p_0({{\mathbb{T}}}^n,X)$, we use hypothesis $(ii)$ and again Lemma \[Lem:11.23\] to conclude that $$\begin{aligned}
\| T_m f \|_{L^p({{\mathbb{R}}}^n,X)}
\leq & K \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p'} \sum_{j=1}^N\Big\| T_{\widetilde{m}_j}\Big(\sum_{k \in {\mathbb{Z}}^n} \widehat{f}({{\varepsilon}}k) e_k \Big) \Big\|_{L^p({{\mathbb{T}}}^n,X)} \\
= & K \lim_{{{\varepsilon}}\to 0} \sum_{j=1}^N{{\varepsilon}}^{n/p'} \Big\| \sum_{k \in {\mathbb{Z}}^n} \widehat{T_{m_j}f}({{\varepsilon}}k) e_k \Big\|_{L^p({{\mathbb{T}}}^n,X)}
= K \sum_{j=1}^N \| T_{m_j}f \|_{L^p({{\mathbb{R}}}^n,X)}.\qedhere
\end{aligned}$$
Let $f$ be a trigonometric polynomial with $\widehat{f}(0)=0$. This family of functions is dense in $L^p_0({{\mathbb{T}}}^n,X)$. We can write $f= \sum_{k \in {\mathbb{Z}}^n} \widehat{f}(k)e_k=\sum_{k\in I}\widehat{f}(k)e_k$, where $I=\{ k \in {\mathbb{Z}}^n : \widehat{f}(k) \neq 0\}$ is a finite set such that $0 \notin I$.
We choose the auxiliary functions $\phi(x)=e^{-\pi|x|^2/p}$ and $\psi(x)=e^{-\pi|x|^2/p'}$, which clearly satisfy (see, for instance, [@Foll Proposition 8.24]) $$1
=\|\phi\|_{L^p({{\mathbb{R}}}^n)}
=\|\psi\|_{L^{p'}({{\mathbb{R}}}^n)}
=\int_{{{\mathbb{R}}}^n} \phi(x) \psi(x) dx
=\int_{{{\mathbb{R}}}^n} \widehat{\phi}(\xi) \widecheck{\psi}(\xi) d\xi.$$ Then, by Lemma \[Lem:11.20\], Hölder’s inequality, Lemma \[Lem:11.21\] and hypothesis $(i)$ we arrive at $$\begin{aligned}
\|T_{\widetilde{m}} f\|_{L^p({{\mathbb{T}}}^n,X)}
= & \sup_{g} \int_{{{\mathbb{T}}}^n}\langle g(t), T_{\widetilde{m}}f(t) \rangle dt
= \sup_{g} \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^n \int_{{{\mathbb{R}}}^n}\langle \psi({{\varepsilon}}x)g(x) , T_m(\phi({{\varepsilon}}\cdot )f)(x)\rangle dx \\
\leq & \sup_{g} \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p'} \|\psi({{\varepsilon}}\cdot)g \|_{L^{p'}({{\mathbb{R}}}^n,X^*)} {{\varepsilon}}^{n/p} \|T_m(\phi({{\varepsilon}}\cdot )f)\|_{L^{p}({{\mathbb{R}}}^n,X)} \\
\leq & K\lim_{{{\varepsilon}}\to 0} \sum_{j=1}^N {{\varepsilon}}^{n/p} \|T_{m_j}(\phi({{\varepsilon}}\cdot )f)\|_{L^{p}({{\mathbb{R}}}^n,X)},
\end{aligned}$$ where the supremum is over all trigonometric polynomials $g : {{\mathbb{T}}}^n \to X^*$ such that $\|g\|_{L^{p'}({{\mathbb{T}}}^n,X^*)} \leq 1$.
To finish the reasoning we are going to prove that $$\label{eq:goal1}
\lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p} \|T_{m_j}(\phi({{\varepsilon}}\cdot )f)\|_{L^{p}({{\mathbb{R}}}^n,X)}
\leq \|T_{\widetilde{m}_j}f\|_{L^{p}({{\mathbb{T}}}^n,X)}, \quad j=1, \dots, N.$$ For simplicity, from now on, we just write $m$ instead of $m_j$, $j=1, \dots, N$.
Take $\theta_0 \in C^\infty_c({{\mathbb{R}}}^n)$ such that $\operatorname{supp}\theta_0 \subset B(0,2)$, $\theta_0 \equiv 1$ in $B(0,1)$ and consider $$\theta_\ell (\xi)
= \theta_0 \Big( \frac{\xi}{2^\ell} \Big) - \theta_0 \Big( \frac{\xi}{2^{\ell-1}} \Big), \quad \xi \in {{\mathbb{R}}}^n, \ \ell \geq 1.$$ Note that the function $\theta_\ell$ is supported in the annulus $B(0,2^{\ell+1}) \setminus B(0,2^{\ell-1})$. Then, we have the partition of the unity $$1= \sum_{\ell \in {\mathbb{N}}} \theta_\ell (\xi), \quad \xi \in {{\mathbb{R}}}^n,$$ where for each $\xi$ there exists at most two nonzero terms in the above sum (see [@Ste2 p. 242] for details). Let also $\Theta_\ell =\theta_{\ell-1}+\theta_{\ell}+\theta_{\ell+1}$ ($\theta_{-1}=0$), so that $\Theta_\ell$ is supported in $B(0,2^{\ell+2})$ for $\ell=0,1$, and in $B(0,2^{\ell+2})\setminus B(0,2^{\ell-2})$ for $\ell\geq 2$; and $\Theta_\ell=1$ on the support of $\theta_\ell$.
For every $\ell \in {\mathbb{N}}$, we define $\phi_\ell \in \mathcal{S}({{\mathbb{R}}}^n)$ to be the function such that $\widehat{\phi}_\ell = \widehat{\phi} \ \theta_\ell$. Now, we can write $$\begin{aligned}
T_m(\phi({{\varepsilon}}\cdot )f)
& = \sum_{k \in {\mathbb{Z}}^n} \widehat{f}(k) T_m(\phi({{\varepsilon}}\cdot )e_k)
= \sum_{k \in {\mathbb{Z}}^n} \widehat{f}(k) \Big[ m(\xi) \frac{1}{{{\varepsilon}}^n} \widehat{\phi} \Big( \frac{\xi-k}{{{\varepsilon}}} \Big) \Big]^{\vee} \\
& = \sum_{\ell \in {\mathbb{N}}} \sum_{k \in {\mathbb{Z}}^n} \widehat{f}(k) \Big[ m(\xi) \frac{1}{{{\varepsilon}}^n} \widehat{\phi}_\ell \Big( \frac{\xi-k}{{{\varepsilon}}} \Big) \Big]^{\vee}
= \sum_{\ell \in {\mathbb{N}}} T_m(\phi_\ell({{\varepsilon}}\cdot )f).
\end{aligned}$$ Since $\widehat{\phi}_\ell((\cdot - k)/{{\varepsilon}})$ is supported in $B(k,2^{\ell+1}{{\varepsilon}})$, it suggests that we may replace $m(\xi)$ by $m(k)$ with the advantage that $$\sum_{\ell \in {\mathbb{N}}} \sum_{k \in {\mathbb{Z}}^n} \widehat{f}(k) \Big[ m(k) \widehat{\phi_\ell({{\varepsilon}}\cdot)e_k}(\xi) \Big]^{\vee}
= \sum_{\ell \in {\mathbb{N}}} \phi_\ell({{\varepsilon}}\cdot) T_{\widetilde{m}} f
= \phi({{\varepsilon}}\cdot) T_{\widetilde{m}} f.$$ With this motivation in mind, we write $$\begin{aligned}
\lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p} \|T_{m}(\phi({{\varepsilon}}\cdot )f)\|_{L^{p}({{\mathbb{R}}}^n,X)}
\leq & \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p} \|T_{m}(\phi({{\varepsilon}}\cdot )f) - \phi({{\varepsilon}}\cdot) T_{\widetilde{m}} f\|_{L^{p}({{\mathbb{R}}}^n,X)} \\
& + \lim_{{{\varepsilon}}\to 0} {{\varepsilon}}^{n/p} \|\phi({{\varepsilon}}\cdot) T_{\widetilde{m}} f\|_{L^{p}({{\mathbb{R}}}^n,X)}.
\end{aligned}$$ Thus, by Lemma \[Lem:11.21\], to establish we only need to see that the first term converges to zero. Even more, it is sufficient to show that $$\label{eq:goal2}
\lim_{{{\varepsilon}}\to 0} \sum_{\ell \in {\mathbb{N}}} {{\varepsilon}}^{n/p} \|T_{m(\cdot)-m(k)}(\phi_\ell({{\varepsilon}}\cdot )e_k)\|_{L^{p}({{\mathbb{R}}}^n)} = 0, \quad k \in I.$$ Notice that is a scalar property, so the Banach space $X$ does not play any role hereinafter.
We first observe an easy estimate: $$\begin{split}
{{\varepsilon}}^{n/p} \|T_{m(\cdot)-m(k)}(\phi_\ell({{\varepsilon}}\cdot )e_k)\|_{L^{p}({{\mathbb{R}}}^n)}
&\leq {{\varepsilon}}^{n/p} \Big(\|T_m\|_{L^p({{\mathbb{R}}}^n)\longrightarrow L^p({{\mathbb{R}}}^n)}+\|m\|_{L^\infty({{\mathbb{R}}}^n)}\Big)\|\phi_\ell({{\varepsilon}}\cdot )e_k\|_{L^{p}({{\mathbb{R}}}^n)} \\
&\leq C\|\phi_\ell\|_{L^{p}({{\mathbb{R}}}^n)}, \\
\end{split}$$ where moreover we claim that $$\label{eq:claim2}
\|\phi_\ell\|_{L^p({{\mathbb{R}}}^n)}
\leq C 2^{-\rho \ell}, \quad \rho \in {\mathbb{N}}.$$ Indeed, fixing some $\nu>n/(2p)$ and any (large) $\mu\in{\mathbb{N}}$, we have $$\begin{aligned}
\|\phi_\ell\|_{L^p({{\mathbb{R}}}^n)}
& \leq \|(1+|x|^2)^\nu \phi_\ell\|_{L^\infty({{\mathbb{R}}}^n)} \| (1+|x|^2)^{-\nu} \|_{L^p({{\mathbb{R}}}^n)} \\
& \leq C \Big\|\Big[(1+ \Delta)^\nu \widehat{\phi_\ell}\Big]^{\vee}\Big\|_{L^\infty({{\mathbb{R}}}^n)}
\leq C \| (1+ \Delta)^\nu \widehat{\phi_\ell} \|_{L^1({{\mathbb{R}}}^n)}
\leq C \sum_{|\gamma| \leq 2 \nu } \| \partial^\gamma \widehat{\phi_\ell} \|_{L^1({{\mathbb{R}}}^n)}\\
&\leq C \sum_{|\gamma| \leq 2 \nu }\sum_{\delta\leq\gamma}\binom{\gamma}{\delta}
\| \partial^{\delta} \widehat{\phi}\cdot \partial^{\gamma-\delta}\theta_{\ell} \|_{L^1({{\mathbb{R}}}^n)}
\leq C \sum_{|\delta| \leq 2 \nu }
\| \partial^{\delta} \widehat{\phi}\|_{L^1(B(0,2^{\ell+1})\setminus B(0,2^{\ell-1}))} \\
& \leq C 2^{(n-\mu) \ell} \sum_{|\delta| \leq 2 \nu } \| |\xi|^\mu \partial^\delta \widehat{\phi}(\xi) \|_{L^\infty({{\mathbb{R}}}^n)}
\leq C 2^{(n-\mu) \ell},
\end{aligned}$$ since $\phi$ and thus $\widehat\phi$ belong to $\mathcal{S}({{\mathbb{R}}}^n)$.
By dominated convergence, it suffices to show that each term in , for a fixed $\ell\in{\mathbb{N}}$, tends to zero as ${{\varepsilon}}\to 0$.
Fix $k \in I$, $\ell \in {\mathbb{N}}$ and $0<{{\varepsilon}}<2^{-\ell-3}$. If we define $M_\ell^{{{\varepsilon}},k} = (m(\cdot)-m(k))\Theta_\ell((\cdot-k)/{{\varepsilon}})$, it is clear that $$T_{m(\cdot)-m(k)}(\phi_\ell({{\varepsilon}}\cdot )e_k)
= T_{M_\ell^{{{\varepsilon}},k}}(\phi_\ell({{\varepsilon}}\cdot )e_k),$$ and $$\begin{aligned}
\label{eq:phiTM}
{{\varepsilon}}^{n/p} \|T_{M_\ell^{{{\varepsilon}},k}}(\phi_\ell({{\varepsilon}}\cdot )e_k)\|_{L^{p}({{\mathbb{R}}}^n)}
\leq & \| \phi_\ell \|_{L^{p}({{\mathbb{R}}}^n)} \|T_{M_\ell^{{{\varepsilon}},k}}\|_{L^{p}({{\mathbb{R}}}^n) \longrightarrow L^{p}({{\mathbb{R}}}^n)}.
\end{aligned}$$ We already estimate $\| \phi_\ell \|_{L^{p}({{\mathbb{R}}}^n)}$ above, and we now turn to the multiplier norm. First of all, it is more convenient to consider the new multiplier $$\widetilde{M_\ell^{{{\varepsilon}},k}}(\eta)
= M_\ell^{{{\varepsilon}},k}(2^\ell {{\varepsilon}}\eta + k) = [m(2^\ell {{\varepsilon}}\eta + k) - m(k)] \Theta_\ell (2^\ell \eta), \quad \eta \in {{\mathbb{R}}}^n,$$ because the $L^p$-norm of Fourier multipliers is invariant under dilations and translations in the multiplier function (see for example [@Graf (2.5.14) and (2.5.15)]). Note that $\operatorname{supp}\Theta_\ell (2^\ell \cdot) \subset B(0,4)$, so that $|k+2^\ell{{\varepsilon}}\eta|\geq|k|- |\eta| \cdot 2^\ell{{\varepsilon}}\geq 1- 4 \cdot 1/8=1/2$, and thus $m$ is only evaluated in $B(0,1/2)^c$, where it is $C^\infty$. Also note that $\Theta_\ell(2^\ell\eta)=\Theta_2(2^2\eta)$ for all $\ell\geq 2$.
We first estimate $\widetilde{M_\ell^{{{\varepsilon}},k}}$ pointwise. By applying the mean value theorem: $$\begin{aligned}
|\widetilde{M_\ell^{{{\varepsilon}},k}}(\eta)|
&\leq \max_{z\in[k,k+2^\ell{{\varepsilon}}\eta]} |\nabla m(z)|\cdot 2^\ell{{\varepsilon}}|\eta|\cdot |\Theta_\ell(2^\ell\eta)| \\
& \leq 2^{\ell+2} {{\varepsilon}}\sup_{|z|\geq 1/2} |\nabla m(z)|
\leq C 2^{\ell} {{\varepsilon}}, \quad \eta \in {{\mathbb{R}}}^n,
\end{aligned}$$ where the last bound follows from the homogeneity of $m$ (see for example [@Graf p. 366]). (Here $[x,y]$ stands for the segment connecting points $x$, $y \in {{\mathbb{R}}}^n$ and $C$ does not depend on $\eta$, $\ell$ or ${{\varepsilon}}$.)
Next, we estimate the size of the derivatives of $\widetilde{M_\ell^{{{\varepsilon}},k}}$. From the Leibniz rule, we get for every $\gamma \in {\mathbb{N}}^n \setminus \{0\}$, such that $|\gamma| \leq n+1$, $$\begin{split}
|\partial^\gamma \widetilde{M_\ell^{{{\varepsilon}},k}}(\eta)|
& \leq \Big\{ |m(2^{\ell}{{\varepsilon}}\eta + k) - m(k)|\cdot |(\partial^\gamma) (\Theta_\ell)(2^{\ell}\cdot))(\eta)| \\
& \qquad\qquad+ \sum_{0 \neq \tau \leq \gamma} \binom{\gamma}{\tau} (2^\ell{{\varepsilon}})^{|\tau|} |(\partial^\tau m)(2^{\ell}{{\varepsilon}}\eta + k)|
\cdot |(\partial^{\gamma - \tau})( \Theta_\ell) (2^{\ell}\cdot))(\eta)|\Big\} \\
& \leq C\Big\{ 2^{\ell} {{\varepsilon}}+\sum_{0\neq\tau\leq\gamma}(2^\ell{{\varepsilon}})^{|\tau|}\Big\}\leq C 2^\ell{{\varepsilon}}, \quad \eta \in {{\mathbb{R}}}^n,
\end{split}$$ for certain constant $C$ independent of $\eta$, $\ell$ and ${{\varepsilon}}$. The same bound holds for $|\eta|^{|\gamma|}|\partial^\gamma \widetilde{M_\ell^{{{\varepsilon}},k}}(\eta)|$, since $|\eta|\leq 4$ on the support of $ \widetilde{M_\ell^{{{\varepsilon}},k}}$. Hence, Mihlin’s multiplier theorem (see for instance [@Graf Theorem 5.2.7]) implies that $$\label{eq:Mihlin}
\| T_{\widetilde{M_\ell^{{{\varepsilon}},k}}} \|_{L^{p}({{\mathbb{R}}}^n) \longrightarrow L^{p}({{\mathbb{R}}}^n)}
\leq C 2^{\ell} {{\varepsilon}}.$$
Putting together , and we conclude .
To finish this section we present a transference result from ${{\mathbb{T}}}^n$ to ${{\mathbb{T}}}^{rn}$. Its idea when $n=1$ goes back to Bourgain [@Bou]. Given an operator $S$ acting on $L^p_0({{\mathbb{T}}}^n,X)$ we define its tensor extension to $L^p_0({{\mathbb{T}}}^n,L^p({{\mathbb{T}}}^{(\ell-1)n},X)) \simeq L^p({{\mathbb{T}}}^{(\ell-1)n},L^p_0({{\mathbb{T}}}^{n},X))$, $\ell \in {\mathbb{N}}$, by $$S_\ell = I_{L^p({{\mathbb{T}}}^{(\ell-1)n})} \otimes S.$$
\[Prop:TnTrn\] Let $1<p<\infty$, $N,r \in {\mathbb{N}}$ and $m$, $m_j \in C^\infty({{\mathbb{R}}}^n \setminus \{0\})$, $j=1, \dots, N$, be homogeneous functions. If the following estimate holds $$\displaystyle \|T_{\widetilde{m}} f\|_{L^p({{\mathbb{T}}}^n,X)}
\leq K \sum_{j=1}^N \| T_{\widetilde{m}_j}f\|_{L^p({{\mathbb{T}}}^n,X)}, \quad f \in L^p_0({{\mathbb{T}}}^n,X),$$ then, $$\displaystyle \Big \| \sum_{\ell=1}^r (T_{\widetilde{m}})_{\ell} f_\ell \Big\|_{L^p({{\mathbb{T}}}^{rn},X)}
\leq K \sum_{j=1}^N \Big\| \sum_{\ell=1}^r (T_{\widetilde{m}_j})_\ell f_\ell \Big \|_{L^p({{\mathbb{T}}}^{rn},X)},
\quad f_\ell \in L^p_0({{\mathbb{T}}}^n,L^p({{\mathbb{T}}}^{(\ell-1)n},X)).$$
By density arguments it is enough to consider trigonometric polynomials $(f_\ell)_{\ell=1}^r$. For each $\ell =1, \dots, r$, we write $$\label{eq:expansion}
f_\ell(\bar{t}_{\ell-1},t_\ell)
= \sum_{\substack{k \in {\mathbb{Z}}^n \setminus \{0\} \\ |k| \leq B}}
\sum_{\substack{s \in {\mathbb{Z}}^{(\ell-1)n} \\ |s| \leq B}}
a_{s,k}^{(\ell)} e_s(\bar{t}_{\ell-1}) e_k(t_\ell),$$ where $\bar{t}_{\ell-1}=(t_1, \dots, t_{\ell-1}) \in {{\mathbb{T}}}^{(\ell-1)n}$, $t_\ell \in {{\mathbb{T}}}^n$ and $a_{s,k}^{(\ell)} \in X$. Notice that it is posible to choose the same $B \in {\mathbb{N}}$ for all $f_\ell$. Then, $$\begin{aligned}
(T_{\widetilde{m}})_{\ell} f_\ell(\bar{t}_{\ell-1},t_\ell)
& = \sum_{\substack{k \in {\mathbb{Z}}^n \setminus \{0\} \\ |k| \leq B}}
\sum_{\substack{s \in {\mathbb{Z}}^{(\ell-1)n} \\ |s| \leq B}}
a_{s,k}^{(\ell)} e_s(\bar{t}_{\ell-1}) T_{\widetilde{m}} \Big(e_k(t_\ell)\Big) \\
& = \sum_{\substack{k \in {\mathbb{Z}}^n \setminus \{0\} \\ |k| \leq B}}
\sum_{\substack{s \in {\mathbb{Z}}^{(\ell-1)n} \\ |s| \leq B}}
a_{s,k}^{(\ell)} e_s(\bar{t}_{\ell-1}) m(k) e_k(t_\ell), \quad (\bar{t}_{\ell-1},t_\ell) \in {{\mathbb{T}}}^{\ell n}.
\end{aligned}$$
Fix some $\bar{t}_\ell=(\bar{t}_{\ell-1},t_\ell)=(t_1, \dots, t_\ell) \in {{\mathbb{T}}}^{\ell n}$ and take $\bar{M}_\ell = (\bar{M}_{\ell-1},M_\ell)=(M_1, \dots, M_\ell) \in ({\mathbb{N}}\setminus\{0\})^\ell$ to be chosen below. We introduce some operations between vectors of different lengths, $$\bar{M}_\ell \otimes t
= (\bar{M}_{\ell-1} \otimes t, M_\ell t)= (M_1 t , \dots, M_\ell t) \in ({{\mathbb{T}}}^{n})^\ell,
\quad t \in {{\mathbb{T}}}^n,$$ $$\bar{M}_{\ell-1} \odot s
= M_1 s_1 + \dots + M_{\ell-1}s_{\ell-1} \in {\mathbb{Z}}^n, \quad s=(s_1, \dots, s_{\ell-1}) \in ({\mathbb{Z}}^n)^{\ell-1}.$$ It is verified that $s \cdot (\bar{M}_{\ell-1} \otimes t) = (\bar{M}_{\ell-1} \odot s) \cdot t$, and hence $e_s(\bar{M}_{\ell-1} \otimes t)=e_{\bar{M}_{\ell-1} \odot s}(t)$.
Now, we consider Bourgain’s transformation of $f_\ell$, which is given by $$\begin{aligned}
\label{eq:11.22}
\widetilde{f}_\ell(t)
= f_\ell(\bar{t}_{\ell} + \bar{M}_\ell \otimes t)
= \sum_{\substack{k \in {\mathbb{Z}}^n \setminus \{0\} \\ |k| \leq B}}
\sum_{\substack{s \in {\mathbb{Z}}^{(\ell-1)n} \\ |s| \leq B}}
a_{s,k}^{(\ell)} e_s(\bar{t}_{\ell-1}) e_k(t_\ell)e_{\bar{M}_{\ell-1} \odot s + M_\ell k}(t),
\quad t \in {{\mathbb{T}}}^n.
\end{aligned}$$ The function $\widetilde{(T_{\widetilde{m}})_\ell f_\ell}$ is defined analogously.
We want to compare $\widetilde{(T_{\widetilde{m}})_\ell f_\ell}$ with $T_{\widetilde{m}} \widetilde{f}_\ell$. Observe that both of them are multipliers transforms of $\widetilde{f}_\ell$, where in the first case the exponential factor $e_{\bar{M}_{\ell-1} \odot s + M_\ell k}$ is multiplied by $m(k)$; and in the second one by $m(\bar{M}_{\ell-1} \odot s + M_\ell k)$. Using the homogeneity of $m$ and the mean value theorem, we arrive at $$\begin{aligned}
|m(k) - m(\bar{M}_{\ell-1} \odot s + M_\ell k)|
& = \Big| m\Big( \frac{k}{|k|} \Big) - m\Big( \frac{\bar{M}_{\ell-1} \odot s}{M_\ell|k|} + \frac{k}{|k|}\Big) \Big| \\
& \leq \Big| \frac{\bar{M}_{\ell-1} \odot s}{M_\ell|k|} \Big| \sup_z |\nabla m(z)|
\leq \frac{|\bar{M}_{\ell-1}|}{|M_\ell|} \sup_z |\nabla m(z)|,
\end{aligned}$$ where the supremum is taken over certain compact set away from the origin. This supremum is finite because $m \in C^\infty({{\mathbb{R}}}^n \setminus \{0\})$. If we choose the sequence $M_1 < M_2 < \dots$ to be sufficiently rapidly increasing, we can make above difference smaller than any preassigned ${{\varepsilon}}>0$.
In conclusion, denoting by $\|g\|_A$ the sum of the $X$-norms of the Fourier coefficients of a trigonometric polynomial $g$ (on a torus of any dimension), we have deduced that $$\Big\|\widetilde{(T_{\widetilde{m}})_\ell f_\ell} - T_{\widetilde{m}} \widetilde{f}_\ell \Big\|_{L^p({{\mathbb{T}}}^n,X)}
\leq \Big\|\widetilde{(T_{\widetilde{m}})_\ell f_\ell} - T_{\widetilde{m}} \widetilde{f}_\ell \Big\|_{A}
\leq {{\varepsilon}}\|f_\ell\|_A, \quad \ell=1, \dots r,$$ provided that $M_1 < M_2 < \dots$ is sufficiently rapidly increasing.
Now, we apply the hypothesis in order to get $$\begin{aligned}
\Big\| \sum_{\ell=1}^r \widetilde{(T_{\widetilde{m}})_\ell f_\ell} \Big\|_{L^p({{\mathbb{T}}}^n,X)}
& \leq \Big\| T_{\widetilde{m}} \Big( \sum_{\ell=1}^r \widetilde{f}_\ell \Big) \Big\|_{L^p({{\mathbb{T}}}^n,X)}
+ {{\varepsilon}}\sum_{\ell=1}^r \|f_\ell\|_A \\
& \leq K \sum_{j=1}^N \Big \| T_{\widetilde{m}_j} \Big( \sum_{\ell=1}^r \widetilde{f}_\ell \Big) \Big\|_{L^p({{\mathbb{T}}}^n,X)}
+ {{\varepsilon}}\sum_{\ell=1}^r \|f_\ell\|_A.
\end{aligned}$$ So far, we have taken $L^p$-norms with respect to the variable $t \in {{\mathbb{T}}}^n$. It is time to take $L^p$-norms with respect to the fixed variables $\bar{t}_r=(t_1, \dots, t_r) \in {{\mathbb{T}}}^{rn}$, $$\begin{aligned}
& \Big( \int_{{{\mathbb{T}}}^{rn}} \int_{{{\mathbb{T}}}^n} \Big\| \sum_{\ell=1}^r
(T_{\widetilde{m}})_\ell f_\ell(\bar{t}_\ell + \bar{M}_\ell \otimes t) \Big\|^p_X dt d\bar{t_r} \Big)^{1/p} \\
& \qquad \leq K \sum_{j=1}^N \Big( \int_{{{\mathbb{T}}}^{rn}} \int_{{{\mathbb{T}}}^n} \Big\| \sum_{\ell=1}^r
T_{\widetilde{m}_j} \widetilde{f}_\ell (\bar{t}_\ell + \bar{M}_\ell \otimes t) \Big\|^p_X dt d\bar{t_r} \Big)^{1/p}
+ {{\varepsilon}}\sum_{\ell=1}^r \|f_\ell\|_A.
\end{aligned}$$ Finally, exchanging the order of integration we notice that the dependence in $t$ and $\bar{M}_\ell$ disappears, resulting in $$\Big \| \sum_{\ell=1}^r (T_{\widetilde{m}})_{\ell} f_\ell \Big\|_{L^p({{\mathbb{T}}}^{rn},X)}
\leq K \sum_{j=1}^N \Big\| \sum_{\ell=1}^r (T_{\widetilde{m}_j})_\ell f_\ell \Big \|_{L^p({{\mathbb{T}}}^{rn},X)}
+ {{\varepsilon}}\sum_{\ell=1}^r \|f_\ell\|_A,$$ which finishes the proof of this lemma, once we take ${{\varepsilon}}\to 0$.
Proof of Theorem \[Th:main\] {#sec:proofTh}
============================
We start this section with an alternative, but equivalent, definition of a UMD Banach space. Let $(\Omega,d\mu)$ be a probability space. We call $({{\varepsilon}}_\ell d_\ell({{\varepsilon}}_1, \dots, {{\varepsilon}}_{\ell-1}))_{\ell \in {\mathbb{N}}}$ a Paley–Walsh martingale difference sequence when $d_\ell : {{\mathbb{R}}}^{\ell-1} \longrightarrow X$, $\ell \geq 2$; $d_1$ is a constant map in $X$ and $({{\varepsilon}}_\ell)_{\ell \in {\mathbb{N}}}$ are independent Bernoulli random variables, that is, $\mu(\{\omega \in \Omega : {{\varepsilon}}_\ell(\omega)=\pm 1\})=1/2$, $\ell \in {\mathbb{N}}$.
The Banach space $X$ has UMD if for some (equivalently, for every) $1<p<\infty$, there exists a constant $c_p$ such that $$\label{eq:defUMD2}
\Big\| \sum_{\ell=1}^r \sigma_\ell {{\varepsilon}}_\ell d_\ell({{\varepsilon}}_1, \dots, {{\varepsilon}}_{\ell-1}) \Big\|_{L^p(\Omega,X)}
\leq c_p \Big\| \sum_{\ell=1}^r {{\varepsilon}}_\ell d_\ell({{\varepsilon}}_1, \dots, {{\varepsilon}}_{\ell-1}) \Big\|_{L^p(\Omega,X)},$$ for each $r \in {\mathbb{N}}$, all Walsh-Paley martingale difference sequence $({{\varepsilon}}_\ell d_\ell)_{\ell=1}^r$ and every $(\sigma_\ell)_{\ell =1}^r \in \{-1,1\}^r$. It is well-known that $\beta_p(X)= \inf c_p$ (see [@Bu4 p. 12] and [@Mau]).
The next lemma reduces checking the condition to the following inequality, more convenient for our purposes.
\[Lem:reduction\] Let $1<p<\infty$. Suppose that there exists $C_p>0$ such that the following holds: For every $r\in{\mathbb{N}}$ and a sequence $(\sigma_\ell)_{\ell=1}^r$ in $\{-1,+1\}$, there is a sequence $(b_\ell)_{\ell=1}^r$ in $\{-1,+1\}^n$ such that $$\label{eq:reduction}
\Big\| \sum_{\ell=1}^r \sigma_\ell a(b_\ell \cdot t_\ell) \Phi_\ell(t_1, \dots, t_{\ell-1}) \Big\|_{L^p({{\mathbb{T}}}^{rn},X)}
\leq C_p \Big\| \sum_{\ell=1}^r a(b_\ell \cdot t_\ell) \Phi_\ell(t_1, \dots, t_{\ell-1}) \Big\|_{L^p({{\mathbb{T}}}^{rn},X)},$$ for every $1$-dimensional trigonometric polynomial $a$ with zero mean value and every trigonometric polynomial $\Phi_\ell:{{\mathbb{T}}}^{(\ell-1)n}\longrightarrow X$. Then $\beta_p(X) \leq C_p$.
Let $r \in {\mathbb{N}}$, let $({{\varepsilon}}_\ell d_\ell)_{\ell=1}^r$ be a given Paley–Walsh martingale difference sequence and $(\sigma_\ell)_{\ell =1}^r \in \{-1,1\}^r$. Take also $b_\ell \in \{-1,1\}^n$, $\ell=1, \dots, r$, satisfying .
It is obvious that the function $\operatorname{sgn}(\theta)$, $\theta \in {{\mathbb{T}}}$, is a Bernoulli random variable. Moreover, for every $b \in \{-1,1\}^n$, the function $\operatorname{sgn}( b \cdot t)$, $t \in {{\mathbb{T}}}^n$, is also a Bernoulli random variable. Since $({{\varepsilon}}_1(\omega), \dots, {{\varepsilon}}_r(\omega))$ and $(\operatorname{sgn}(b_1 \cdot t_1), \dots, \operatorname{sgn}(b_r \cdot t_r))$ have the same distribution, we can write $$\label{eq:equid}
\begin{split}
&\Big\| \sum_{\ell=1}^r \sigma_\ell' {{\varepsilon}}_\ell d_\ell\Big({{\varepsilon}}_1, \dots, {{\varepsilon}}_{\ell-1})\Big) \Big\|_{L^p(\Omega,X)} \\
&= \Big\| \sum_{\ell=1}^r \sigma_\ell' \operatorname{sgn}(b_\ell \cdot t_\ell) d_\ell\Big(\operatorname{sgn}(b_1 \cdot t_1), \dots, \operatorname{sgn}(b_{\ell-1} \cdot t_{\ell-1})\Big) \Big\|_{L^p({{\mathbb{T}}}^{rn},X)},
\end{split}$$ for both $\sigma_\ell'=\sigma_\ell$ and $\sigma_\ell'=1$.
For a fixed $\delta>0$, we choose a $1$-dimensional trigonometric polynomial $a$ with zero mean value, verifying that $$\|\operatorname{sgn}(\cdot) - a \|_{L^p({{\mathbb{T}}})} < \delta,$$ where $\operatorname{sgn}$ is the $1$-periodic extension of the sign on $[-1/2,1/2)$. Then it easily follows that, for every $\ell=1, \dots, r$, $$\|\operatorname{sgn}(b_\ell \cdot t) - a(b_\ell \cdot t) \|_{L^p({{\mathbb{T}}}^n)} < \delta,$$ and so in particular $$\|a(b_\ell \cdot t) \|_{L^p({{\mathbb{T}}}^n)} \leq \|\operatorname{sgn}(b_\ell \cdot t)\|_{L^p({{\mathbb{T}}}^n)} +\delta=1+\delta.$$
We also choose trigonometric polynomials $\Phi_\ell:{{\mathbb{T}}}^{(\ell-1)n}\longrightarrow X$ such that $$\Big\|\Phi_\ell(t_1,\ldots,t_{\ell-1})-d_\ell\Big(\operatorname{sgn}(b_1 \cdot t_1), \dots, \operatorname{sgn}(b_{\ell-1} \cdot t_{\ell-1})\Big) \Big\|_{L^p({{\mathbb{T}}}^{(\ell-1)n},X)}<\delta,
\quad \ell=1, \dots, r.$$ It follows that $$\label{eq:approx}
\begin{split}
&\Big\|a(b_\ell\cdot t_\ell)\Phi_\ell(t_1,\ldots,t_{\ell-1})
-\operatorname{sgn}(b_\ell\cdot t_{\ell})d_\ell\Big(\operatorname{sgn}(b_1 \cdot t_1), \dots, \operatorname{sgn}(b_{\ell-1} \cdot t_{\ell-1})\Big) \Big\|_{L^p({{\mathbb{T}}}^{\ell n},X)} \\
&\leq\|a(b_\ell\cdot t_{\ell})\|_{L^p({{\mathbb{T}}}^n)}\Big\|\Phi_\ell(t_1,\ldots,t_{\ell-1})
-d_\ell\Big(\operatorname{sgn}(b_1 \cdot t_1), \dots, \operatorname{sgn}(b_{\ell-1} \cdot t_{\ell-1})\Big) \Big\|_{L^p({{\mathbb{T}}}^{(\ell-1) n},X)} \\
&\quad+\|a(b_\ell\cdot t_\ell)-\operatorname{sgn}(b_\ell\cdot t_\ell)\|_{L^p({{\mathbb{T}}}^n)}
\Big\|d_\ell\Big(\operatorname{sgn}(b_1 \cdot t_1), \dots, \operatorname{sgn}(b_{\ell-1} \cdot t_{\ell-1})\Big) \Big\|_{L^p({{\mathbb{T}}}^{(\ell-1) n},X)} \\
&\leq (1+\delta)\cdot\delta+\delta\cdot C \leq C\delta, \quad \ell=1, \dots, r,
\end{split}$$ where the first $C$ is the maximum of the $L^p$ norms of the given functions $d_\ell({{\varepsilon}}_1,\ldots,{{\varepsilon}}_{\ell-1})$.
A combination of and with the assumption now shows that $$\begin{split}
&\Big\| \sum_{\ell=1}^r \sigma_\ell {{\varepsilon}}_\ell d_\ell({{\varepsilon}}_1, \dots, {{\varepsilon}}_{\ell-1}) \Big\|_{L^p(\Omega,X)}
\leq \Big\| \sum_{\ell=1}^r \sigma_\ell a(b_\ell \cdot t_\ell) \Phi_\ell(t_1,\ldots,t_{\ell-1}) \Big\|_{L^p({{\mathbb{T}}}^{rn},X)} +Cr\delta \\
&\leq C_p \Big\| \sum_{\ell=1}^r a(b_\ell \cdot t_\ell) \Phi_\ell(t_1,\ldots,t_{\ell-1}) \Big\|_{L^p({{\mathbb{T}}}^{rn},X)} +Cr\delta \\
&\leq C_p\Big\| \sum_{\ell=1}^r {{\varepsilon}}_\ell d_\ell({{\varepsilon}}_1,\ldots,{{\varepsilon}}_{\ell-1}) \Big\|_{L^p(\Omega,X)}+(C_p+1)Cr\delta.
\end{split}$$ Since $\delta>0$ is arbitrarily small, we conclude that holds and $\beta_p(X) \leq C_p$.
Let us first make an additional useful reduction. Note that the set $F$ can be neither $\varnothing$ or $\{1,\ldots,n\}$, because for these sets we have $\sum_{\ell\in F}\beta_\ell=\sum_{\ell\in F}\alpha^j_\ell$ (using $|\beta|=|\alpha^j|$ if $F=\{1,\ldots,n\}$), contradicting the fact that these have different parity. Thus, there exists an index $s\in F$ as well as $t\in\{1,\ldots,n\}\setminus F$. By applying the assumption to $\partial_k u$ in place of $u$, we see that implies a similar estimate with $\beta+e_k$ in place of $\beta$ and $\alpha^j+e_k$ in place of $\alpha^j$. Thus, if $\sum_{\ell\in F}\beta_\ell$ is not already odd, we can make it odd by adding $e_s$ to both $\beta$ and $\alpha^j$; note that this preserves the other assumptions. After, if the new $|\beta|$ is not even, we can make it even by adding $e_t$ to both $\beta$ and $\alpha^j$, which again preserves the other assumptions, including the value of $\sum_{\ell\in F}\b!
eta_\ell$, since $t\notin F$. Thus, if the assumptions of Theorem \[Th:main\] hold for some $\beta,\alpha^j$, they also hold for some (possibly different) $\beta,\alpha^j$ which satisfy in addition that $|\beta|=|\alpha^j|$ is even and $\sum_{\ell\in F}\beta_{\ell}$ is odd (so that $\sum_{\ell\in F}\alpha^j_\ell$ is again even). We henceforth assume this, and proceed to prove the claim that $X$ is UMD with $\beta_p(X)\leq KN$.
So, let $\alpha^j$, $\beta$ and $F$ be as in the statement of the theorem, with the additional assumptions just made, and $m$, $m_j$ be the multipliers defined in . The evenness of $|\beta|=|\alpha^j|$ gives the advantage that the multipliers $m(\xi)=\xi^\beta/|\xi|^{|\beta|}$ and $m_j(\xi)=\xi^{\alpha^j}/|\xi|^{|\alpha^j|}$ are not only homogeneous, but also even, so that $m(\lambda\xi)=m(\xi)$ holds for all $\lambda\in{{\mathbb{R}}}\setminus\{0\}$, not just $\lambda>0$.
Fix a 1-dimensional trigonometric polynomial $$a(\theta)
= \sum_{\ell \in {\mathbb{Z}}} \widehat{a}(\ell) e_\ell(\theta), \quad \theta \in {{\mathbb{T}}},$$ with $\widehat{a}(0)=0$, and consider the function $a_b(t)=a(b \cdot t)$, $t \in {{\mathbb{T}}}^n$, for certain $b \in \{-1,1\}^n$. Also, if $J \subset \{1, \dots, n\}$ we define $b_J \in \{-1,1\}^n$ to be the vector whose $\ell$-th component is equal to $-1$, if $\ell \in J$; or $1$ otherwise. With this notation, take $a^+=a_{(1, \dots, 1)}$ and $a^-=a_{b_{F}}$.
Since $m(\ell b)=m(b)$ for all $\ell\in{\mathbb{Z}}\setminus\{0\}$, we find that $$\label{eq:claim}
T_{\widetilde{m}} a_b(t)
=T_{\widetilde{m}}\sum_{\ell\in{\mathbb{Z}}}\widehat{a}(\ell)e_{\ell b}(t)
=\sum_{\ell\in{\mathbb{Z}}}\widehat{a}(\ell)m(\ell b)e_{\ell b}(t)
= m(b)a_b(t), \quad b \in \{-1,1\}^n,$$ where $T_{\widetilde{m}}$ the discrete Fourier multiplier defined in . Now, taking in mind the assumptions imposed over $F$, we deduce as a direct consequence of that $$\label{eq:pm}
T_{\widetilde{m}} a^{\pm} = \pm n^{-|\beta|/2} a^{\pm}
\quad \text{and} \quad
T_{\widetilde{m}_j} a^{\pm} = n^{-|\beta|/2} a^{\pm}, \ j=1, \dots, N.$$
Let $r \in {\mathbb{N}}$, $\sigma_\ell \in \{-1,1\}$ and $\Phi_\ell:{{\mathbb{T}}}^{(\ell-1)n}\longrightarrow X$ be trigonometric polynomials for $\ell=1, \dots, r$. Moreover, we define $\zeta_\ell = a^+$, if $\sigma_\ell=1$; and $\zeta_\ell = a^-$, when $\sigma_\ell=-1$. As it was commented in the introduction, implies . Then, Propositions \[Prop:RnTn\] and \[Prop:TnTrn\] together with , allows us to write $$\begin{aligned}
\Big\| \sum_{\ell=1}^r \sigma_\ell \zeta_\ell \Phi_\ell \Big\|_{L^p({{\mathbb{T}}}^{rn},X)}
& = n^{|\beta|/2} \Big\| \sum_{\ell=1}^r T_{\widetilde{m}}\zeta_\ell \Phi_\ell \Big\|_{L^p({{\mathbb{T}}}^{rn},X)}
\leq K n^{|\beta|/2} \sum_{j=1}^N \Big\| \sum_{\ell=1}^r T_{\widetilde{m}_j}\zeta_\ell \Phi_\ell \Big\|_{L^p({{\mathbb{T}}}^{rn},X)} \\
& = K N \Big\| \sum_{\ell=1}^r \zeta_\ell \Phi_\ell \Big\|_{L^p({{\mathbb{T}}}^{rn},X)}.\end{aligned}$$ Finally, an application of Lemma \[Lem:reduction\] implies that $\beta_p(X) \leq KN$, and hence the proof of this theorem is completed.
[10]{}
, [ *Integral representations of functions and imbedding theorems. [V]{}ol. [I]{}*]{}, V. H. Winston & Sons, Washington, D.C., 1978. Translated from the Russian, Scripta Series in Mathematics, Edited by Mitchell H. Taibleson.
, [*Some remarks on [B]{}anach spaces in which martingale difference sequences are unconditional*]{}, Ark. Mat., 21 (1983), pp. 163–168.
height 2pt depth -1.6pt width 23pt, [*Vector-valued singular integrals and the [$H^1$]{}-[BMO]{} duality*]{}, in Probability theory and harmonic analysis ([C]{}leveland, [O]{}hio, 1983), vol. 98 of Monogr. Textbooks Pure Appl. Math., Dekker, New York, 1986, pp. 1–19.
, [*A geometric condition that implies the existence of certain singular integrals of [B]{}anach-space-valued functions*]{}, in Conference on harmonic analysis in honor of [A]{}ntoni [Z]{}ygmund, [V]{}ol. [I]{}, [II]{} ([C]{}hicago, [I]{}ll., 1981), Wadsworth Math. Ser., Wadsworth, Belmont, CA, 1983, pp. 270–286.
height 2pt depth -1.6pt width 23pt, [*Boundary value problems and sharp inequalities for martingale transforms*]{}, Ann. Probab., 12 (1984), pp. 647–702.
height 2pt depth -1.6pt width 23pt, [*Explorations in martingale theory and its applications*]{}, in École d’Été de [P]{}robabilités de [S]{}aint-[F]{}lour [XIX]{}—1989, vol. 1464 of Lecture Notes in Math., Springer, Berlin, 1991, pp. 1–66.
height 2pt depth -1.6pt width 23pt, [*Martingales and singular integrals in [B]{}anach spaces*]{}, in Handbook of the geometry of [B]{}anach spaces, [V]{}ol. [I]{}, North-Holland, Amsterdam, 2001, pp. 233–269.
, [*Régularité [$L^p$]{} pour les équations d’évolution*]{}, in Séminaire d’[A]{}nalyse [F]{}onctionelle 1984/1985, vol. 26 of Publ. Math. Univ. Paris VII, Paris, 1986, pp. 155–165.
, [*Real analysis. [M]{}odern techniques and their applications*]{}, Pure and Applied Mathematics (New York), John Wiley & Sons Inc., New York, second ed., 1999.
, [*On singular integral and martingale transforms*]{}, Trans. Amer. Math. Soc., 362 (2010), pp. 553–575.
, [*Classical [F]{}ourier analysis*]{}, vol. 249 of Graduate Texts in Mathematics, Springer, New York, second ed., 2008.
, [*Some remarks on complex powers of [$(-\Delta)$]{} and [UMD]{} spaces*]{}, Illinois J. Math., 35 (1991), pp. 401–407.
, [*Vector-valued extensions of operators related to the [O]{}rnstein-[U]{}hlenbeck semigroup*]{}, J. Anal. Math., 91 (2003), pp. 1–29.
, [*Aspects of probabilistic [L]{}ittlewood-[P]{}aley theory in [B]{}anach spaces*]{}, in Banach spaces and their applications in analysis, Walter de Gruyter, Berlin, 2007, pp. 343–355.
height 2pt depth -1.6pt width 23pt, [*Estimates for partial derivatives of vector-valued functions*]{}, Illinois J. Math., 51 (2007), pp. 731–742.
height 2pt depth -1.6pt width 23pt, [ *Littlewood-[P]{}aley-[S]{}tein theory for semigroups in [UMD]{} spaces*]{}, Rev. Mat. Iberoam., 23 (2007), pp. 973–1009.
, [*Extremal inequalities in Sobolev spaces and quasiconformal mappings*]{}, Z. Anal. Anwendungen 1 (1982), pp. 1–16.
, [*Wavelet transform for functions with values in [UMD]{} spaces*]{}, Studia Math., 186 (2008), pp. 101–126.
, [*Maximal [$L_p$]{}-regularity for parabolic equations, [F]{}ourier multiplier theorems and [$H^\infty$]{}-functional calculus*]{}, in Functional analytic methods for evolution equations, vol. 1855 of Lecture Notes in Math., Springer, Berlin, 2004, pp. 65–311.
, [*Système de [H]{}aar*]{}, in Séminaire [M]{}aurey-[S]{}chwartz 1974–1975: [E]{}spaces [L[$\sup{p}$]{}]{}, applications radonifiantes et géométrie des espaces de [B]{}anach, [E]{}xp. [N]{}os. [I]{} et [II]{}, Centre Math., École Polytech., Paris, 1975, pp. 26 pp. (erratum, p. 1).
, [*Martingale and integral transforms of [B]{}anach space valued functions*]{}, in Probability and [B]{}anach spaces ([Z]{}aragoza, 1985), vol. 1221 of Lecture Notes in Math., Springer, Berlin, 1986, pp. 195–222.
, [*Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals*]{}, vol. 43 of Princeton Mathematical Series, Princeton University Press, Princeton, NJ, 1993.
, [*Littlewood-[P]{}aley theory for functions with values in uniformly convex spaces*]{}, J. Reine Angew. Math., 504 (1998), pp. 195–226.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We calculate the $D^*D\pi$ and $B^*B\pi$ couplings using QCD sum rules on the light-cone. In this approach, the large-distance dynamics is incorporated in a set of pion wave functions. We take into account two-particle and three-particle wave functions of twist 2, 3 and 4. The resulting values of the coupling constants are $g_{D^*D\pi}= 12.5\pm 1$ and $g_{B^*B\pi}= 29\pm 3 $. From this we predict the partial width $ \Gamma (D^{*+} {\rightarrow}D^0 \pi^+ )=32 \pm 5~ keV $. We also discuss the soft-pion limit of the sum rules which is equivalent to the external axial field approach employed in earlier calculations. Furthermore, using $g_{B^*B\pi}$ and $g_{D^*D\pi}$ the pole dominance model for the $ B {\rightarrow}\pi$ and $D{\rightarrow}\pi$ semileptonic form factors is compared with the direct calculation of these form factors in the same framework of light-cone sum rules.'
---
MPI-PhT/94-62\
CEBAF-TH-94-22\
LMU 15/94\
September 1994
[**$D^*D\pi$ and $B^*B\pi$ couplings in QCD $^+$**]{}\
\
[$^1$ CEBAF, Newport News, Virginia 23606, USA]{}\
[$^2$ Max-Planck-Institut für Physik, Werner-Heisenberg-Institut, D-80805 München, Germany ]{}\
[$^3$ Sektion Physik der Universität München, D-80333 München , Germany ]{}\
[$^a$ on leave from Institute of Theoretical and Experimental Physics, 117259 Moscow, Russia ]{}\
[$^b$ on leave from St.Petersburg Nuclear Physics Institute, 188350 Gatchina, Russia; now at DESY, D–22603 Hamburg, Germany]{}\
[$^c$ on leave from Yerevan Physics Institute, 375036 Yerevan, Armenia ]{}\
[$^{*}$*Alexander von Humboldt Fellow* ]{}\
$^+$ Work supported by the German Federal Ministry for Research and Technology under contract No. 05 6MU93P\
Introduction
============
The extraction of fundamental parameters from data on heavy flavoured hadrons inevitably requires some information about the physics at large distances. Numerous theoretical studies have been devoted to making this extraction as reliable as possible. While the inclusive $B$ and $D$ decays appear to be the most clean reactions theoretically, exclusive decays are often much easier to measure experimentally. However, for their interpretation one needs accurate estimates of decay form factors and other hadronic matrix elements. In the exceptional case $B {\rightarrow}D e \nu $, the form factor at zero recoil can be calculated in the heavy quark limit [@IW89; @N91]. In most other important cases, one has to rely on less rigorous nonperturbative approaches. Among those, QCD sum rules [@SVZ] have proved to be particularly powerful.
In this paper we employ sum rule methods in order to calculate the $D^*D\pi$ and $B^*B\pi$ coupling constants. These couplings are interesting for several reasons. In particular, they determine the normalization of the heavy-to-light form factors $D {\rightarrow}\pi $ and $B {\rightarrow}\pi $ near zero pionic recoil, where the $D^*$- and $B^*$- poles are believed to dominate. For further discussion one may consult refs. [@IW90; @BLN]. Recently, it has been argued that in the combined heavy quark and chiral limit vector meson dominance becomes even exact [@GM]. As noted in refs. [@BBD1; @BKR], $B^*$ dominance is also compatible with the dependence of the $B {\rightarrow}\pi$ form factor on the momentum transfer $p$ predicted by QCD sum rules at low values of $p^2$. A similar conclusion is reached in ref. [@Ball] concerning $D^*$ dominance in the $D {\rightarrow}\pi$ form factor. In addition, the $D^*D\pi$ coupling can be directly measured in the decay $D^* {\rightarrow}D\pi$ and thus provides one more independent test of the sum rule approach.
Calculations of couplings of heavy mesons to a pion have already been undertaken several times in the framework of QCD sum rules. Unfortunately, the sum rules obtained in refs. \[10-13\] differ in nonleading terms and, to some extent, also in numerical results. Here, we suggest an alternative method known as QCD sum rules on the light-cone. In this approach, the ideas of duality and matching between parton and hadron descriptions, intrinsic to the QCD sum rules, are combined with the specific operator product expansion (OPE) techniques used to study hard exclusive processes in QCD [@exclusive; @BLreport]. In contrast to the conventional sum rules based on the Wilson OPE of the T-product of currents at small distances, one considers expansions near the light-cone in terms of nonlocal operators, the matrix elements of which define hadron wave functions of increasing twist. As one advantage, this formulation allows to incorporate additional information about the Euclidean asymptotics of correlation functions in QCD for arbitrary external momenta. These new features are related to the (approximate) conformal invariance of QCD and are coded in the hadron wave functions. Many of the theoretical results obtained in the context of exclusive processes (see e.g. ref. [@CZreport]) are very useful in the present context as well. In turn, we will see that heavy-flavour decays can provide valuable constraints on the wave functions.
Previous applications of light-cone sum rules include calculations of the amplitude of the radiative decay $\Sigma\rightarrow p\gamma$ [@BBK], the nucleon magnetic moments [@BF1], the strong couplings $g_{\pi N N}$ and $g_{\rho\omega\pi}$ [@BF1], form factors of semileptonic and radiative $B$- and $D$-meson decays \[8,19-21\], the pion form factor at intermediate momentum transfers [@BH], and the $ \pi A \gamma^* $ form factors [@belyaev94]. In all these cases the results are encouraging.
The light-cone sum rule for the coupling of heavy mesons to a pion is the principal result of the present paper which is organized as follows. In Sect. 2 we discuss possible strategies in constructing sum rules for coupling constants and explain the concept of light-cone sum rules. The derivation of the sum rule for the $B^*B\pi$ and $D^*D\pi$ couplings is then completed in Sect. 3, taking into account the pion two- and three-particle wave functions up to twist 4. Sect. 4 is devoted to a detailed numerical analysis. In Sect. 5 we show that in a simplified case, putting the external momenta in the correlation function equal to each other and performing a Borel transformation in one momentum instead of two, we obtain the sum rule proposed previously in refs. \[10-13\]. We demonstrate that despite the slightly different terminology of these papers the sum rules must coincide with each other, and elaborate on possible subtleties in these earlier calculations. Furthermore, in Sect. 6 using our results on the $B^*B\pi$ and $D^*D\pi$ coupling constants, we confront the pole model for the heavy-to-light form factors $B \rightarrow \pi$ and $D \rightarrow \pi$ with a direct calculation of these form factors in the same framework of light-cone sum rules following ref. [@BKR]. A comprehensive comparison of our results on the $ D^*D\pi$ and $B^*B\pi$ couplings with other estimates and our conclusions are presented in Sect. 7.
Technical details are collected in two Appendices. Appendix A summarizes the relevant features of the pion wave functions and specifies the input in our numerical calculations. In Appendix B we derive a simple rule how to subtract the contribution from excited resonances and continuum states in the sum rule.
Light-cone versus short-distance expansion
===========================================
For definiteness, we focus on the $D^{*}D \pi$ coupling defined by the on-mass-shell matrix element D\^[\*+]{}(p)\^-(q)D\^0(p+q)&=& -g\_[D\^\*D]{}q\_\^, \[1\] where the momentum assignment is specified in brackets and $\epsilon_\mu$ is the polarization vector of the $D^*$. The couplings for the different charge states are related by isospin symmetry: g\_[D\^\*D]{} g\_[D\^[\*+]{}D\^0\^+]{}= -g\_[D\^[\*+]{}D\^[+]{}\^0]{}=g\_[D\^[\*0]{}D\^0\^0]{} =-g\_[D\^[\*0]{}D\^+\^-]{} . \[gd\] Most of what is said below applies equally to $B^*B\pi$ couplings. The corresponding relations are obtained by the obvious replacements $ c {\rightarrow}b $, $ D^* {\rightarrow}\bar B^*$ and $D {\rightarrow}\bar B $.
Following the general strategy of QCD sum rules, we want to obtain quantitative estimates for $g_{D^*D\pi}$ by matching the representations of a suitable correlation function in terms of hadronic and quark-gluon degrees of freedom. For this purpose, we choose F\_(p,q)= id\^4xe\^[ipx]{} \^-(q)|T{ |[d]{}(x)\_c(x), |[c]{}(0)i\_5 u(0)} |0 . \[19\] To the best of our knowledge, the study of correlation functions with the T-product of currents sandwiched between the vacuum and one-pion state was first suggested in ref. [@CS83].
With the pion being on mass-shell, $q^2=m_\pi^2 $, the correlation function (\[19\]) depends on two invariants, $p^2$ and $(p+q)^2$. Throughout the paper we set $m_\pi=0$. The contribution of interest is the one having poles in $p^2$ and $(p+q)^2$ : F\_(p,q)& = & (q\_+12(1-)p\_) . \[22\] Obviously, this term stems from the ground states in the $(\bar{d}c)$ and $(\bar c u)$ channels. To derive eq. (\[22\]) we have made use of eq. (\[1\]) and the decay constants $f_D$ and $f_{D^*}$ defined by the matrix elements D |[c]{}i\_5u 0&=& \[fD1\] and 0 |[d]{}\_cD\^\*&=&m\_[D\^\*]{}f\_[D\^\*]{}\_ , \[fD\*\] respectively.
The main theoretical task is the calculation of the correlation function (\[19\]) in QCD. This problem can be solved in the Euclidean region where both virtualities $p^2$ and $(p+q)^2$ are negative and large, so that the charm quark is sufficiently far off-shell. Substituting, as a first approximation, the free $c$-quark propagator 0|T{c(x)|[c]{}(0)}|0= i\_c\^0(x) = e\^[-ikx]{} \[prop\] into eq. (\[19\]) one readily obtains F\_(p,q)&=&i e\^[i(p-k)x]{}(m\_c (q)||[d]{}(x)\_\_5u(0)|0.\
&&+. k\^(q) ||[d]{}(x)\_\_\_5u(0)|0) . \[23\] Diagramatically, this contribution is depicted in Fig. 1a. Applying the short-distance expansion (SDE) in terms of local operators to the first matrix element of eq. (\[23\]), |[d]{}(x)\_\_5u(0)=\_n |[d]{}(0)(x)\^n\_\_5 u(0) , \[expan\] one has after integration over $x$ and $k$ : F\_(p,q)&=&i \_[n=0]{}\^M\_n q\_ , \[expans\] where $$\langle\pi(q) |\bar{d}\stackrel{\leftarrow}{D}_{\alpha_1}
\stackrel{\leftarrow}{D}_{\alpha_2}...
\stackrel{\leftarrow}{D}_{\alpha_n}{\gamma}_\mu{\gamma}_5 u |0\rangle
=(i)^n q_\mu q_{\alpha_1} q_{\alpha_2}...q_{\alpha_n}M_n + ...~,$$ $D$ being the covariant derivative, has been used. One immediately encounters the following problem. If the ratio =2(p q)/(m\_c\^2-p\^2)= ((p+q)\^2-p\^2)/(m\_c\^2-p\^2) \[ksi\] is finite one must keep an [*infinite*]{} series of local operators in eq. (\[expans\]). All these operators give contributions of the same order in the heavy quark propagator $1/(m_c^2-p^2)$, differing only by powers of the dimensionless parameter $ \tilde{\xi}$ [^1]. Therefore, SDE of eq. (\[23\]) is useful [*only*]{} if $ \tilde{\xi} \rightarrow 0$, i.e. if $p^2 \simeq (p+q)^2$ or, equivalently, $q \simeq 0$. Under this condition, the series in eq. (\[expans\]) can be truncated after a few terms involving only a small number of unknown matrix elements $M_n$. However, for general momenta with $p^2 \neq (p+q)^2$ one has to sum up the infinite series of matrix elements of local operators in some way.
This formidable task is solved by using the techniques developed for hard exclusive processes in QCD [@BLreport; @CZreport]. We illustrate the solution for the correlation function id\^4xe\^[ipx]{}\^0(q)|T{ |[q]{}(x)\_q(x), |[q]{}(0)\_ q(0)} |0= \_ p\^q\^F\^\*(p\^2,(p+q)\^2) , \[photonff\] which is similar to eq. (\[19\]) and defines the form factor of the coupling of a pion to a pair of virtual photons [@CZreport; @gorsky]. In eq. (\[photonff\]), ${\cal Q}$ is a matrix of electromagnetic charges, and $q$ is a row vector composed of the up and down quark flavours. As well-known [@exclusive], for sufficiently virtual photons, $p^2 {\rightarrow}-\infty$ and $(p+q)^2 {\rightarrow}-\infty$, this form factor can be calculated in perturbative QCD. The principal result reads F\^\*(p\^2,(p+q)\^2) = F\_0 \_0\^1 , F\_0= , \[photon\] with calculable radiative corrections, and with power corrections suppressed by the photon virtualities. Here, ${\varphi}_\pi(u)$ is the pion wave function of leading twist, defined by the following matrix element of a nonlocal operator on the light-cone $x^2=0$ : (q)||[d]{}(x)\_\_5u(0)|0= -iq\_f\_\_0\^1due\^[iuqx]{}\_(u) . \[pionwf\] Physically, ${\varphi}_\pi$ represents the distribution in the fraction of the light-cone momentum $q_0 +q_3 $ of the pion carried by a constituent quark. Note the normalization of ${\varphi}_\pi$ to unity following from eq. (\[pionwf\]) for $x=0$ .
Let us first concentrate on the form factor (\[photon\]) at (almost) equal photon virtualities, i.e. at $\xi= (2p\cdot q)/(-p^2) \ll 1$. Expanding the denominator in eq. (\[photon\]) around $\xi=0$ one obtains a sum over moments of the pion wave function: F\^\*(p\^2,(p+q)\^2)= \_n \^n\_0\^1duu\^n\_(u) . \[summom\] From the definition (\[pionwf\]) it is easy to see that these moments are given by vacuum-to-pion transition matrix elements involving increasing powers of the covariant derivative. For $ p^2=(p+q)^2 $, i.e. $q=0$, only the lowest moment $n=0$ contributes in eq. (\[summom\]), and the form factor reduces to $F_0/p^2$ which is the classical result. In contrast, if the photon virtualities differ strongly from each other, then many moments contribute to eq. (\[summom\]). In this case, the calculation of the form factor requires the knowledge of the shape of the pion wave function.
Returning to the correlation function (\[19\]) one realizes that the same technique may be used to obtain a representation analogous to eq. (\[photon\]). The only new element in the correlation function (\[19\]) is the virtual heavy quark propagating between the points $x$ and $0$ instead of the light quarks present in eq. (\[photonff\]). This gives rise to important differences which however do not change the formalism substantially. For the present discussion it is sufficient to stick to the approximation (\[23\]) and confine ourselves to the first term proportional to $m_c$. The complete analysis of this expression and the calculation of further corrections will be carried out in the next section. Furthermore, writing $F_\mu$ in terms of invariant amplitudes: F\_(p,q)=F(p\^2,(p+q)\^2)q\_+ (p\^2,(p+q)\^2)p\_ , \[220\] we focus on the function $F$. Using the definition eq. (\[pionwf\]) of the leading twist wave function and integrating over $x$ and $k$ one finds F(p\^2,(p+q)\^2)=m\_cf\_\_0\^1 . \[Fzeroth\] Thus, the infinite series of matrix elements of local operators encountered before in eq. (\[expans\]) is effectively replaced by an unknown wave function. The expression (\[Fzeroth\]) is rather similar to the one quoted in eq. (\[photon\]) for the $ \pi^0 {\gamma}^* {\gamma}^* $ form factor. Most noteworthy is the fact that the large-distance dynamics is described by one and the same pion wave function. This universal property is essential for the whole approach.
Next we indicate how the relation (\[Fzeroth\]) can be turned into a sum rule for the coupling constant $g_{D^*D\pi}$. The key idea is to write a hadronic representation of $F$ by means of a double dispersion integral: F(p\^2,(p+q)\^2)=\
+\
+ + . \[221\] The first term arises from the ground state contribution already indicated in eq. (\[22\]), while the spectral function $\rho^h(s_1,s_2)$ is supposed to take into account higher resonances and continuum states in the $D^*$ and $D$ channels. The additional single dispersion integrals originate in subtractions which are generally necessary to make the double dispersion integral finite. Then, considering $p^2$ and $(p+q)^2$ as independent variables one can perform the usual Borel improvement in both channels. Applying the Borel operator \_[M\^2]{}f(Q\^2)=lim\_[Q\^2,n , Q\^2/n=M\^2 ]{} ( -)\^n f(Q\^2) f(M\^2) \[B\] to eq. (\[221\]) with respect to $p^2$ and $(p+q)^2$, we obtain F(M\_1\^2,M\_2\^2)&& [B]{}\_[M\_1\^2]{}[B]{}\_[M\_2\^2]{}F(p\^2,(p+q)\^2)= e\^[--]{}\
&+&e\^[--]{} \^h(s\_1,s\_2)ds\_1ds\_2 , \[222\] where $M_1^2$ and $M_2^2$ are the Borel parameters associated with $p^2$ and $(p+q)^2$, respectively. Note that contributions from heavier states are now exponentially suppressed by factors $\exp\{-\frac{s_{1,2}^2-m_{D^*,D}^2}{M_{1,2}^2} \}$ as desired, while the subtraction terms depending only on one of the variables, $p^2$ or $(p+q)^2$, vanish.
The same transformation has to be applied to the expression (\[Fzeroth\]). To this end we rewrite $(p+uq)^2 = (1-u) p^2 + u (p+q)^2$, and use $${\cal B}_{M_1^2}{\cal B}_{M_2^2}
\frac{(l-1)!}{[m_c^2 -(1-u) p^2 - u (p+q)^2]^l }
= (M^2)^{2-l} e^{-m_c^2/M^2 }\delta(u-u_0) ~,
\label{dBorel}$$ where the Borel parameters $M_1^2$ and $M^2_2$ have been replaced by u\_0=, M\^2= . \[37\] Finally, equating the quark-gluon and the hadronic representations of $F(M_1^2,M_2^2) $ and discarding for a moment contributions of higher states, we end up with the sum rule $$\frac{m_D^2m_{D^*}f_Df_{D^*}}{m_c} \cdot g_{D^*D\pi}
= m_c f_\pi~{\varphi}_\pi(u_0)M^2
\exp \left[ \frac{m^2_{D^\ast}-m_c^2}{M_1^2} +
\frac{m^2_{D}-m_c^2}{M_2^2}\right] +\ldots
\label{SR0}$$ The ellipses refer to higher-twist contributions which we discuss in detail later. Since $M_1^2$ and $M_2^2$ are expected to be quite similar in magnitude, the coupling constant $g_{D^*D\pi}$ is determined by the value of the pion wave function at $u\simeq 1/2$, that is by the probability for the quark and the antiquark to carry equal momentum fractions in the pion [@BBK]. This interesting feature is shared by the sum rules for many other important hadronic couplings involving the pion.
The quantity ${\varphi}_\pi(1/2)$ is considered to be a nonperturbative parameter, similar to quark and gluon condensates in the standard approach. It may be determined from suitable sum rules in which the phenomenological part is known experimentally. We use the value $${\varphi}_\pi(1/2) = 1.2\pm 0.2
\label{WF12}$$ obtained in ref. [@BF1].
The dependence on the pion wave function disappears in the kinematical limit $q\rightarrow 0$ as can be seen from eq. (\[Fzeroth\]). This is just the limit where the correlation function (\[19\]) can be treated in SDE. The condition $q \simeq 0$ is implicitly assumed in refs. [@EK85; @GY94] where the correlation function (\[19\]) is calculated using the external field method. This technique is equivalent to the soft-pion approximation used in refs. [@ovch; @Cetc94] as will become clear later. For comparison, we present the sum rule following from eqs. (\[Fzeroth\]) and (\[221\]) by putting $q=0$, or equivalently $(p+q)^2=p^2$. Since $p^2$ is the only variable left, one now can only perform a single Borel transformation and, hence, the subtraction terms in the double dispersion relation (\[221\]) are no more eliminated. Moreover, the contributions to eq. (\[221\]) of transitions from excited states to ground states are not suppressed after Borel transformation [@ioffe]. This point will be explained in more detail later in Sect. 5 . In the approximation considered in eq. (\[SR0\]) one obtains $$\frac{m_D^2m_{D^*}f_Df_{D^*}}{m_c} \cdot g_{D^*D\pi} + M^2 A
= m_c f_\pi M^2
\exp \left[ \frac{m^2_{D^\ast}-m_c^2}{2M^2} +
\frac{m^2_{D}-m_c^2}{2M^2}\right] +\ldots ~,
\label{SR1}$$ where $A$ is an unknown constant corresponding to the contributions of unwanted transitions and subtraction terms.
From eqs. (\[SR0\]) and (\[SR1\]) one can clearly see the advantages and disadvantages of the two approaches. In the light-cone sum rule (\[SR0\]) the hadronic input is simple, whereas the theoretical expression involves a new [*universal*]{} nonperturbative parameter, namely ${\varphi}_\pi(1/2)$. Just the opposite is the case for the sum rule (\[SR1\]) at $q =0$. Here the QCD part is straightforward, while the hadronic representation now involves an additional unknown quantity, which is [*non-universal*]{} and specific for this particular sum rule. A comparison of the results obtained in these two approaches should allow one to check the reliability and improve the accuracy of the predictions.
Light-cone sum rule for $g_{D^*D\pi}$ and $g_{B^*B\pi}$
========================================================
In this section we systematically derive the light-cone sum rule for the $D^*D\pi$ and $B^*B\pi$ couplings taking into account the two- and three-particle pion wave functions up to twist 4. First, we complete the calculation of the diagram Fig. 1a which represents the contribution from quark-antiquark wave functions. To this end we return to the expression (\[23\]). In the first matrix element we include the twist 4 corrections in addition to the leading twist term already given in eq. (\[pionwf\]): $$\begin{aligned}
\langle\pi(q)|\bar{d}(x){\gamma}_\mu{\gamma}_5u(0)|0\rangle &=&
-iq_\mu f_\pi\int_0^1du\,e^{iuqx}
\left({\varphi}_\pi (u)+x^2g_1(u)+O(x^4)\right)
\nonumber \\
&&{}+
f_\pi\left( x_\mu -\frac{x^2q_\mu}{qx}\right)\int_0^1
du\,e^{iuqx}g_2(u) ~.
\label{25}\end{aligned}$$ On the r.h.s. of this relation one sees the first few terms of the light-cone expansion in $x^2$ of the matrix element on the l.h.s.. While ${\varphi}_\pi$ parametrizes the leading twist 2 contribution, $g_1$ and $g_2$ are associated with twist 4 operators. In the second matrix element of eq. (\[23\]) we substitute \_\_=-i\_+g\_ \[gammaid\] and express the result in terms of the twist 3 wave functions ${\varphi}_p$ and ${\varphi}_\sigma $ defined by the matrix elements (q)||[d]{}(x)i\_5u(0)|0= \_0\^1due\^[iuqx]{}\_[p]{}(u) \[26\] and (q)||[d]{}(x)\_\_5u(0)|0= i(q\_x\_-q\_x\_) \_0\^1due\^[iuqx]{}\_(u) \[27\] . It should be noted that in eqs. (\[25\],\[26\],\[27\]) the path-ordered gauge factors { ig\_s \^1\_0 d x\_A\^(x) } , \[path\] appearing in between the quark fields and assuring gauge invariance, are not shown for brevity since they formally disappear in the light-cone gauge $x_\mu A^\mu =0 $ assumed throughout this paper. More details on these wave functions can be found in refs. [@BF1; @BB; @BF2] and in Appendix A.
Collecting all terms , we obtain the following result for the invariant function $F$ as defined in eq. (\[220\]): F\^[(a)]{}(p\^2,(p+q)\^2)&=&\_0\^1{ m\_cf\_\_(u)+.\
&& +. m\_cf\_} , \[28\] where $$G_2(u)=-\int_0^ug_2(v)dv ~.$$ The suffix $(a)$ refers to the diagram in Fig. 1a which represents the leading twist term in the light-cone expansion of the c-quark propagator given in eq. (\[prop\]).
In addition, to the accuracy of eq. (\[28\]) we must also take into account higher twist terms in the propagator up to twist 4 which are numerous, in general. The complete expansion is given in ref. [@BB]. One has contributions from $\bar q G q$, $\bar q GG q $ and $\bar q q\bar q q$ nonlocal operators, $G$ denoting the gluon field strength. Here we only consider operators with one gluon field, corresponding to quark-antiquark-gluon components in the pion, and neglect components with two extra gluons, or with an additional $\bar q q$ pair. This is consistent with the approximation of the twist 4 two-particle wave functions derived in ref. [@BF2] and used here. Taking into account higher Fock-space components would demand corresponding modifications in the two-particle functions via the equations of motion. Formally, the neglect of the $\bar q GG q $ and $\bar q q\bar q q$ terms can be justified on the basis of an expansion in conformal spin [@BF2]. In this approximation the $c-$quark propagator reads $$\begin{aligned}
\langle 0 |T\{c(x)\bar{c}(0)\}|0\rangle &=& i\hat{S}_c^0(x)
-ig_s\int\frac{d^4k}{(2\pi )^4}e^{-ikx}
\int_0^1dv\left[ \frac12\frac{\not\!k+m_c}{(m^2_c-k^2)^2}
G^{\mu\nu}(vx)\sigma_{\mu\nu}\right.
\nonumber\\
&& \mbox{}
+\left.\frac1{m_c^2-k^2}vx_\mu G^{\mu\nu}(vx){\gamma}_\nu
\right]~,
\label{32}\end{aligned}$$ where $G_{\mu\nu}=G_{\mu\nu}^c\frac{\lambda^c}2$ with $\mbox{\rm tr}(\lambda^a\lambda^b)=2\delta^{ab}$, and $g_s$ is the strong coupling constant.
Substituting eq. (\[32\]) into eq. (\[19\]) and using eq. (\[220\]) one obtains the contribution to the invariant function $F$ represented by the diagram in Fig. 1b: F\_\^[(b)]{} (p\^2, (p+q)\^2)&=&i ||[d]{}(x)\_\_5u(0)|0 . \[33\] With eq. (\[gammaid\]) and the identities \_\_&=& i(g\_\_-g\_\_) +\_\^\_5 \[ggg\] and \_\_\_&=& (\_g\_-\_g\_)+ i(g\_g\_-g\_g\_)\
&&- \_\_5-i\_ g\^ \_\_5 \[34\] one is led to the three-particle pion wave functions [@gorsky; @BF2] defined by\
&=&if\_[3]{}\[(q\_q\_g\_-q\_q\_g\_) -(q\_q\_g\_-q\_q\_g\_)\] \_i\_[3]{}(\_i)e\^[iqx(\_1+v\_3)]{}, \[29\]\
&=&f\_\_i\_(\_i)e\^[iqx(\_1+v\_3)]{}\
&& +f\_(q\_x\_-q\_x\_) \_i\_(\_i)e\^[iqx(\_1+v\_3)]{}, \[30\]\
&=&if\_\_i\_(\_i)e\^[iqx(\_1+v\_3)]{}\
&&+if\_(q\_x\_-q\_x\_) \_i\_(\_i)e\^[iqx(\_1+v\_3)]{}, \[31\] where $\tilde{G}_{{\alpha}{\beta}}= \frac12
\epsilon _{{\alpha}{\beta}\sigma\tau}G^{\sigma\tau}$ and ${\cal D}{\alpha}_i= d{\alpha}_1 d{\alpha}_2 d{\alpha}_3 \delta(1-{\alpha}_1-{\alpha}_2-{\alpha}_3)$. While ${\varphi}_{3\pi}({\alpha}_i)$ is a twist 3 wave function, the remaining functions ${\varphi}_\perp$, ${\varphi}_\parallel$, $\tilde{{\varphi}}_\perp$ and $\tilde{{\varphi}}_\parallel$ are all of twist 4. Substitution of these expressions in eq. (\[33\]), finally, yields F\^[(b)]{}(p\^2,(p+q)\^2)&=& \_0\^1 dv\_i{ .\
&&+ m\_cf\_.} . \[35\]
In addition to Fig. 1b there are further gluonic diagrams such as the ones depicted in Figs. 1c and 1d. Note, however, that it is not necessary to take the diagram in Fig. 1c into account separately, since its contribution (to twist 4 accuracy) is already included in the two particle wave functions. In contrast, the two-loop perturbative corrections exemplified in Fig. 1d should be included in a systematic way, but their calculation lies beyond the scope of this paper.
Putting together eqs. (\[28\]) and (\[35\]) and applying the double Borel transformation (\[dBorel\]) with respect to $p^2$ and $(p+q)^2$, we end up with the following expression for the invariant amplitude $F$ :\
&=& e\^[-]{}M\^2 { m\_cf\_\_(u\_0) + ( u\_0 \_p(u\_0) +13\_(u\_0)+16u\_0 (u\_0) +\_(u\_0) )\
&& +u\_0g\_2(u\_0)- (g\_1(u\_0) +G\_2(u\_0)) + 2f\_[3]{}I\^G\_3(u\_0)+ m\_cf\_ } . \[F\] Here, $I^G_{3}$ and $I^G_{4}$ involve the three-particle wave functions of twist 3 and 4, respectively : I\^G\_3(u\_0)= \_0\^[u\_0]{}d\_1 , \[fg3\] I\^G\_4(u\_0)= \_0\^[u\_0]{}d\_1\_[u\_0-\_1]{}\^[1-\_1]{} \[2\_(\_i)-\_(\_i)+ 2\_(\_i)-\_(\_i)\] . \[fg4\] The Borel parameters $M^2$ and $u_0 $ are given in eq. (\[37\]). The above is the desired quark-gluon representation of the invariant amplitude $F$ in the correlation function (\[19\]).
The remaining task now is to match eq. (\[F\]) with the corresponding hadronic representation (\[222\]) and to extract the coupling $g_{D^*D\pi}$. As usual, invoking duality, we assume that above certain thresholds in $s_1$ and $s_2$ the double spectral density $\rho^h(s_1,s_2)$ associated with higher resonances and continuum states coincides with the spectral density derived from the diagrams in Fig. 1. The procedure is explained in detail in Appendix B. For $M_1^2=M_2^2=2M^2$ and $u_0=\frac12$, and for standard polynomial wave functions, the effect of the continuum subtraction is remarkably simple [@BBK; @BF1]. It amounts to the following replacement of the exponential factor multiplying the twist 2 and 3 terms proportional to $M^2$ in eq. (\[F\]): e\^[-]{}( e\^[-]{}- e\^[-]{}) , \[39\] $s_0$ being the threshold parameter defined in eq. (\[boundary\]). The higher twist terms which are suppressed in eq. (\[F\]) by inverse powers of $M^2$ with respect to leading one remain unaffected. With eqs. (\[222\]), (\[F\]) and (\[39\]) it is then easy to derive the following QCD sum rule for the $DD^*\pi$ coupling: $$f_Df_{D^*}g_{D^*D\pi}=\frac{m_c^2f_\pi}{m_D^2m_D^*}
e^{\frac{m_{D}^2+m_{D*}^2}{2M^2}}
\Bigg\{M^2 [e^{-\frac{m_c^2}{M^2}} - e^{-\frac{s_0}{M^2}}]
\Big[{\varphi}_\pi(u_0)
$$ $$+ \frac{\mu_\pi}{m_c} \left( u_0{\varphi}_p(u_0)
+\frac13{\varphi}_\sigma (u_0)+\frac16u_0\frac{d{\varphi}_\sigma}{du}(u_0)\right)
+ \frac{2f_{3\pi}}{m_cf_\pi}I^G_3(u_0)\Big]
$$ +e\^[-]{} }\_[u\_0=1/2]{} , \[fin\] where $$\mu_\pi=\frac{m_\pi^2}{m_u+m_d}
= \frac{-2\langle\bar q q\rangle}{f_\pi^2} ~.
\label{qqbar}$$ In eq. (\[qqbar\]) we have used the familiar PCAC relation between $m_\pi$, $f_\pi$ , the quark masses and the quark condensate density $\langle\bar q q\rangle$. Note that the twist 2 and 3 and the twist 4 wave functions have different dimensions (see Appendix A). $G$-parity implies $g_2(1/2)=d{\varphi}_\sigma/du(1/2)=0$, so that these terms vanish in the sum rule (\[fin\]).
For completeness, we also repeat the standard two-point sum rules for the decay constants $f_D$ and $f_{D^\ast}$ : = \_[m\_c\^2]{} \^[s\_0]{}ds e\^s -m\_c |q q e\^ , \[fD\] and f\_[D\^\*]{}\^2m\_[D\*]{}\^2= \_[m\_c\^2]{} \^[s\_0]{} ds e\^s (2+) -m\_c |q q e\^ (1-) , \[fDs\] where $m_0^2= \langle\bar q \sigma_{\alpha\beta}
G^{\alpha\beta}q \rangle/ \langle \bar q q \rangle $ is a conventional parametrization for the quark-gluon condensate. In the above, we have omitted numerically insignificant contributions of the gluon and four-quark condensates. For consistency, we do not take into account perturbative $O(\alpha_s)$ corrections to these sum rules, since they are also not included in the sum rule (\[fin\]).
Numerical analysis
==================
The principal nonperturbative input in the sum rule (\[fin\]) are pion wave functions on the light-cone. In ref. [@exclusive] a theoretical framework has been developed to study these functions. In particular, it has been shown that the wave functions can be expanded in terms of matrix elements of conformal operators which in leading logarithmic approximation do not mix under renormalization. For example, for the leading twist pion wave function one finds an expansion in Gegenbauer polynomials, $${\varphi}_\pi(u,\mu) = 6 u(1-u)\Big[1+a_2(\mu)C^{3/2}_2(2u-1)+
a_4(\mu)C^{3/2}_4(2u-1)+\ldots\Big]\,,
\label{expansion}$$ where all the nonperturbative information is included in the set of multiplicatively renormalizable coefficients $a_n$, $n=2,4,\ldots$. The corresponding anomalous dimensions are such that the coefficients $a_n$ vanish for $\mu \rightarrow \infty $, and the wave function is uniquely determined by the first term in the expansion. Therefore, this term is called asymptotic wave function. Similar expansions also exist for the wave functions of nonleading twist [@BF2].
For practical applications it is important that the expansion in conformal spin converges sufficiently fast. How fast the wave functions approach their well-known asymptotic form is still under debate. However, there are indications [@radnew] that the nonasymptotic deviations have been overestimated previously. Corrections to the asymptotic expressions in next-to-leading (and in some cases also next-to-next-to-leading order) in conformal spin are known for all wave functions which appear in the sum rule (\[fin\]). For details, we must refer the reader to the original literature [@CZreport; @BF2].
In our numerical analysis we use the set of nonleading twist wave functions proposed in ref. [@BF2]. The explicit expressions and the values of the various parameters are collected in Appendix A. Furthermore, we take $f_\pi =132$ MeV, $\mu_\pi(1 GeV)=1.65$, corresponding to $ \langle \bar q q \rangle = -(243$ MeV$)^3 $, $m_0^2=0.8$ GeV$^2$ and, in the charmed meson channels, $m_c=1.3$ GeV, $s_0=6$ GeV$^2$, $m_D=1.87$ GeV and $m_{D^\ast}=2.01$ GeV. The same parameters are also used to determine the decay constants $f_D$ and $f_{D^{\ast}}$ from the sum rules (\[fD\]) and (\[fDs\]). One obtains f\_D = 170 10 MeV, f\_[D\^\*]{} = 240 20 MeV . \[fDDstar\] The uncertainty quoted characterizes the variation with the Borel parameter $M^2$ in the interval 1 GeV$^2< M^2 < $ 2 GeV$^2$.
Having fixed the input parameters, one must find the range of values of $M^2$ for which the sum rule (\[fin\]) is reliable. The lowest possible value of $M^2$ is usually determined by the requirement that the terms proportional to the highest inverse power of the Borel parameter stay reasonably small. The upper limit is determined by demanding that the continuum contribution does not get too large. In the $D^*D\pi$ sum rule we take the interval 2 GeV$^2< M^2<$ 4 GeV$^2$. In this interval the twist 4 term proportional to $M^{-2}$ does not exceed $5\%$. Simultaneously, higher states contribute less than $30\%$. The dependence of the r.h.s. of eq. (\[fin\]) on the Borel parameter is shown in Fig. 2a. As can be seen, in the fiducial range of $M^2$ given above, the sum rule is quite stable. From Fig. 2a one can directly read off the prediction f\_Df\_[D\^[\*]{}]{}g\_[D\^\*D]{} = 0.51 0.05 GeV\^2 . \[combinD\] Dividing this product of couplings by the decay constants (\[fDDstar\]) finally yields for the $D^*D\pi$ coupling constant: g\_[D\^\*D]{}= 12.51.0 , \[constD\*Dpi\] where the error is understood to indicate the range of values corresponding to the correlated variation of the results (\[fDDstar\]) and (\[combinD\]) within the fiducial intervals. While the combination of couplings (\[combinD\]) is not affected by uncertainties in the decay constants, it is more sensitive to the charm quark mass and threshold $s_0$ than the coupling $g_{D^*D\pi}$ itself.
A few comments are in order. The twist 3 terms contribute to the sum rule at the level of (50 $\div$ 60) $\%$ and are therefore as important as the twist 2 contribution. On the other hand, the impact of twist 4 is small amounting to about 5%. Two sources of uncertainties not included in eq. (\[combinD\]) are the nonasymptotic corrections to the leading twist wave function ${\varphi}_\pi$ and to the three-particle (twist 3) wave function ${\varphi}_{3\pi}$. The latter in turn induce modifications in the two-particle (twist 3) wave functions ${\varphi}_p$ and ${\varphi}_\sigma$, apart from the corrections generated by the asymptotic ${\varphi}_{3\pi}$ wave function. In order to estimate the sensitivity of our results to the nonasymptotic effects in ${\varphi}_\pi$ and ${\varphi}_{3\pi}$ we drop these corrections altogether and recalculate the product $f_Df_{D^{*}}g_{D^*D\pi }$. Remarkably, the result changes by only 5$\%$. One can thus be confident that the total uncertainty in eq. (\[combinD\]) does not exceed 20$\%$.
We should emphasize that the above prediction can be directly tested experimentally in the decay $D^* \rightarrow D \pi $. With the value of $g_{D^*D \pi }$ given in eq. (\[constD\*Dpi\]) one predicts the decay width ( D\^[\*+]{} D\^0 \^+) = | q |\^3 = 32 5 . \[Gamma\] Predictions for other charge combinations are easily obtained from eq. (\[Gamma\]) taking into account the isospin relations (\[gd\]) as well as small differences in the phase space volumes: ( D\^[\*+]{} D\^0 \^+) = 21.1(D\^[\*+]{} D\^+ \^0)= 20.72(D\^[\*0]{} D\^[0]{} \^0) . \[widths\] The current experimental upper limit ( D\^[\*+]{} D\^0 \^+) < 89 \[exp\] is obtained by combining the limit $ \Gamma_{tot}(D^{*+}) < 131 ~ \mbox{\rm keV}$ [@ACCMOR] with the branching ratio\
$BR(D^{*+} \rightarrow D^0 \pi^+ ) = (68.1 \pm 1.0 \pm 1.3) \% $ [@CLEO]. Our prediction is well below this upper limit.
The sum rule for $g_{D^*D\pi}$ given in eq. (\[fin\]) is easily converted into a sum rule for the coupling $g_{B^*B\pi}= g_{\bar B^{*0}B^-\pi^+}$ by replacing $c$ with $b$, $D$ with $\bar B$, and $D^*$ with $\bar B^*$. The corresponding parameters are $m_B=5.279$ GeV, $m_{B^*}=5.325$ GeV, $m_b=4.7$ GeV, and $s_0=35$ GeV$^2$. In addition, one has to evolute the wave function parameters to a higher scale $\mu_b$ (see Appendix A). With these changes the two-point sum rules (\[fD\]) and (\[fDs\]) yield $$f_B=140\,\mbox{\rm MeV}\,,~~~~f_{B^*}=160\,\mbox{\rm MeV}
\label{fB}$$ with negligible uncertainties due to variation of $M^2$. Using the same criteria as for the $g_{D^*D\pi }$ sum rule the fiducial range in $M^2$ turns out to be 6 GeV$^2 < M^2 <~ $12 GeV$^2$. The stability of the sum rule for the product of couplings $f_Bf_{B^*}g_{B^*B\pi }$ is illustrated in Fig. 2b. We obtain f\_Bf\_[B\^[\*]{}]{}g\_[B\^\*B]{}= 0.64 0.06\^2 , \[combinB\] and with eq. (\[fB\]) g\_[B\^\*B]{}=293 . \[41\] The hierarchy of various twists as well as the uncertainty due to the nonasymptotic corrections are found to be similar as in the case of $g_{D^*D\pi}$ . Contrary to the latter, the coupling constant $g_{B^*B\pi}$ cannot be measured directly, since the corresponding decay $B^* \rightarrow B \pi $ is kinematically forbidden. However, the $B^*B\pi$ on-shell vertex is of a great importance for understanding of the behaviour of heavy-to-light form factors as will be discussed in Sect. 6.
Sum rules from the short-distance expansion
============================================
With the results of Sect. 4 at hand we are now also in a position to study in more detail the soft-pion limit (\[SR1\]) of the sum rule (\[fin\]) which is obtained from the correlation function (\[19\]) at $p^2=(p+q)^2$ or, equivalently, for $ q{\rightarrow}0$. As discussed in Sect. 2, in this limit one can apply a short-distance expansion in terms of local operators with increasing dimensions in contrast to the light-cone expansion involving nonlocal operators with increasing twist. In technical terms, at $q \rightarrow 0 $ only the lowest moments of the wave functions contribute. Thus the integrals reduce to overall normalization factors. The explicit expression for the invariant function $F$ in this limit is directly obtained from eqs. (\[28\]) and (\[35\]): F(p\^2)= , \[46\] where the parameter $\delta^2$ is specified in Appendix A. After Borel transformation in $p^2$ one has $$\begin{aligned}
F(M^2)&= &m_cf_\pi e^{-\frac{m_c^2}{M^2}}\left[
1+\frac{2\mu_\pi}{3m_c}
+\frac1{M^2}\left( \frac{\mu_\pi m_c}3+\frac{10}9\delta^2
\right)-\frac{5m_c^2\delta^2}{9M^4}\right] ~.
\label{47}\end{aligned}$$
As indicated in eq. (\[SR1\]), the price for simplifying the QCD representation of the correlation function is a more complicated hadronic representation. This in turn makes it more difficult to extract the ground state contribution containing the $D^*D\pi$ coupling. For illustration, we consider the contribution in the dispersion representation of the correlation function (\[19\]) from the transition of a given excited state in the $D^*$-channel with mass $m_* > m_{D^*}$ to the ground state $D$-meson at $q \rightarrow 0$. This contribution is proportional to , \[mmm\] and, after Borel transformation, to ( e\^[-]{}- e\^[-]{}) . \[unsuppr\] Similar expressions hold for the ground state transition $D^*\rightarrow D$ with $m_* = m_{D^*}$. In the limit $m_D = m_{D^*}$, one has a double pole instead of eq. (\[mmm\]) and e\^[-]{} \[e2\] instead of eq. (\[unsuppr\]). Clearly, the contribution (\[unsuppr\]) is not exponentially suppressed relative to the ground state contribution (\[e2\]), and can therefore not be subtracted assuming duality. On the other hand, transitions involving excited states in both the initial and the final states are suppressed by Borel transformation with respect to the ground state transitions and cause no problems. Schematically, the complete hadronic part of the invariant amplitude $F(M^2)$ can be written as follows: F(M\^2) 1[M\^2]{}{ g\_[D\^\*D]{} +AM\^2} e\^[-]{} + Ce\^[-]{} , \[44\] where the constant $A$ incorporates all unsuppressed contributions of the type (\[unsuppr\]), while the term proportional to $C$ contains all exponentially suppressed contributions.
To get rid of the contaminating term $ AM^2$ we follow ref. [@ioffe] and apply the operator ( 1-M\^2 )M\^2e\^ \[oper\] to both representations (\[47\]) and (\[44\]). Now it is possible to subtract the continuum contribution by the same substitution (\[39\]) which we have already employed to get the light-cone sum rule (\[fin\]). In this way we obtain a new sum rule for $g_{D^*D\pi}$: f\_Df\_[D\^\*]{}g\_[D\^\*D]{}&=&( 1-M\^2)e\^\
&& . \[48\] For the same input parameters and the range of $M^2$ leading to the prediction (\[constD\*Dpi\]) this sum rule yields g\_[D\^\*D]{}=112 . \[49\] Similarly, replacing in eq. (\[48\]) the charmed meson parameters by the corresponding quantities in the beauty channel, one finds g\_[B\^\*B]{}=286 . \[50\] As compared with the predictions (\[constD\*Dpi\]) and (\[41\]) the uncertainties are larger by a factor of two due to the worse stability of the sum rule (\[48\]) against variation of $M^2$. The agreement of the results indicates selfconsistency of the sum rule approach and gives support to the approximations used in the pion wave functions.
Furthermore, one can show that the sum rule (\[48\]) is actually equivalent to the sum rule obtained in ref. [@EK85] using external field techniques. Indeed, applying the usual reduction formalism to the pion, one can rewrite the correlation function (\[19\]) in the following form: F\_(p,q)=(q\^2-m\_\^2)d\^4x d\^4y e\^[i(px-qy)]{} 0T{|[d]{}(x)\_c(x), \_(y),|[c]{}(0)\_5u(0)}0 , \[q1\] where $\phi_\pi(y)$ is the interpolating pion field. According to PCAC \_(y)= , \[q2\] $ j_\mu^5(y)=\bar{u}(y){\gamma}_\mu{\gamma}_5d(y)$ being the axial current. Subsituting eq. (\[q2\]) in eq. (\[q1\]) and integrating by parts, one gets [^2] F\_(p,q) &=&iT\_(p,q)q\^= -T\_(p,q)q\^ , \[Fmu\] with T\_(p,q)&=&d\^4x d\^4y e\^[i(px-qy)]{} 0|T{ |[d]{}(x)\_c(x), j\_\^5(y), |[c]{}(0)\_5u(0)}|0 . \[q3\]
Instead of dealing with the three-point correlation function $T_{\mu\rho}(p,q)$ directly, it is more convenient to consider the following two-point correlation function in the constant external axial field $A_\mu^5$: $$T^A_{\mu}(p,q)=\int d^4x~e^{ipx}
\langle0 \mid T\{ \bar{d}(x){\gamma}_\mu c(x),\bar{c}(0)
{\gamma}_5u(0) \}\mid 0\rangle_A ~.
\label{q4}$$ It is assumed that a term $A^{5\mu}j_\mu^5$ corresponding to the interaction of the external field with the light quarks is added to the QCD Lagrangian. To first order in the external field, this correlation function is given by $$T^A_{\mu}(p,q) = T_{\mu\rho}(p,q)A^{5\rho}\, ,$$ with $T_{\mu\rho}(p,q)$ as defined in eq. (\[q3\]). In the above sense, the two-point correlation function (\[q4\]) in the constant axial field, is equivalent to the three-point function (\[q3\]) and, via the PCAC relation (\[q2\]), also to the two-point correlation function (\[19\]) at $q {\rightarrow}0$ [^3]. Consequently, the sum rules obtained in refs. [@EK85] and [@Cetc94] should coincide with each other and with the sum rule (\[48\]) derived in this paper.
In particular, the expression (\[47\]) for $F(M^2)$ can be compared with the result given in eq. (19) of ref. [@EK85] and in eq. (2.15) of ref. [@Cetc94], after normalization and kinematical structures are adjusted properly. In refs. [@EK85; @Cetc94] the correlation function (\[19\]) is separated as follows: F\_= Aq\_+ B(2p\_+q\_) , \[choice\] and the sum rule is obtained by evaluating of the invariant function $A$ for $q {\rightarrow}0$. In terms of the invariant amplitudes defined in eq. (\[220\]) one has A= F-2, B= 2 . \[our\] Hence, for comparison we need also the second invariant function $\tilde F$, which we have not discussed so far, but which can be calculated along the same lines. Adding this contribution to eq. (\[47\]), we have checked that our result for the invariant amplitude $A$ coincides with the one presented in ref. [@Cetc94], apart from terms proportional to $m_0^2$ which, being associated with twist 5 contributions in the light-cone sum rules, are neglected in our approximation. Numerically, this terms are not important. On the other hand, we disagree with ref. [@EK85] in the non-leading terms proportional to $\delta^2$.
Although it is legitimate to use different Lorentz decompositions of the correlation function (\[19\]) in order to derive the desired sum rule, we think that the choice adopted in the present paper is more adequate for the following reason. Since the vector current $\bar q {\gamma}_\mu c $ is not conserved, it not only couples to $J^P=1^-$ vector mesons, but also to $J^P=0^+$ scalar mesons ($D_0$). The corresponding transition matrix element is proportional to the momentum $p_\mu$: 0 |q \_c D\_0= f\_[D\_0]{}m\_[D\_0]{}p\_ . \[scalar\] The mass of the ground state $D_0$ meson is expected to be in the vicinity of 2400 MeV which is not far from the mass of the $D^*$ and below the accepted continuum threshold in the $D^\ast$ channel. For this reason, the $D_0$ contribution should be added to the sum rule. Unfortunately, this introduces additional uncertainties in the hadronic representation as is the case, for example, in ref. [@Cetc94]. In contrast, our sum rules based on the invariant function $F$ defined in eq. (\[220\]) are not affected by scalar meson contributions, which is a clear advantage.
A calculation rather similar to ref. [@Cetc94], but with particular emphasis on the heavy quark limit, has been carried out earlier in ref. [@ovch]. Very recently, another calculation in the heavy quark limit appeared in ref. [@GY94] using the external field technique. Unfortunately, in this paper a wrong expression for the induced quark condensate in the external field is used, as can be seen by consulting refs. [@EK85; @ga]. The error can be traced back to the equation of motion for the quark field in presence of the external field which is modified from $i\!\not\!\!\!D q = 0 $ to $i\!\not\!\!\!D q = \not\!\!\!A^5 q $. By this modification the axial current insertions into the vacuum quark legs are properly taken into account. The numerical comparison of these different calculations is left for the concluding section.
Pole model for $D{\rightarrow}\pi$ and $B{\rightarrow}\pi$ form factors and QCD sum rules
=========================================================================================
The couplings $g_{D^*D\pi}$ and $g_{B^*B\pi}$ fix the normalization of the form factors of the heavy-to-light transitions $D{\rightarrow}\pi$ and $B{\rightarrow}\pi$, respectively, in the pole model description \[5,6\]. This model is based on the vector dominance idea suggesting a momentum dependence dominated by the $D^*$ and $B^*$ poles, respectively. More definitely, the form factor $f^+_D(p^2)$ defined by the matrix element (q) |d\_c D(p+q) = 2f\^+\_D(p\^2)q\_+ ( f\^+\_D(p\^2) -f\^-\_D(p\^2))p\_\[formfdef\] is predicted to be given by f\^+\_D(p\^2)= . \[onepole\] An analogous expression holds for the form factor $f^+_B(p^2)$.
It is difficult to justify the pole model from first principles. Generally, it is believed that the vector dominance approximation is valid at zero recoil, that is at $p^2{\rightarrow}m_D^2$. Arguments based on heavy quark symmetry suggest a somewhat larger region of validity characterized by $(m_D^2-p^2)/m_c \sim O(1 $GeV). However, there are no convincing arguments in favour of this model to be valid also at small values of $p^2$ which are most interesting from a practical point of view. Therefore, the finding [@BBD1; @Ball] that the pole behaviour is consistent with the $p^2$ dependence at $p^2{\rightarrow}0$ predicted by sum rules, is very remarkable. Meanwhile, this claim has been confirmed by independent calculations within the framework of the light-cone sum rules [@BKR].
In this section we want to demonstrate that not only the shape but also the absolute normalization of the above form factors appears to be comparable with the pole model description. This assertion is non-trivial, since contributions of several low-lying resonances in the $D^*$ or $B^*$ channel could still mimic the $p^2$ dependence of a single pole, but the relation to the coupling $g_{D^*D\pi}$ or $g_{B^*B\pi}$ should then be lost [@IW90]. However, despite of the overall agreement in the mass range of $D$ and $B$ mesons, there is a clear disagreement on the asymptotic dependence of the form factors on the heavy mass. The QCD sum rules on the light-cone provide a unique framework to examine these issues, since both the form factors $f^+_{D,B}(p^2)$ at $m_{c,b}^2-p^2\geq O(1$ GeV$^2$) and the couplings $g_{D^*D\pi}$ and $g_{B^*B\pi}$ can be calculated from the same correlation function (\[19\]) using the same technique. In addition, contrary to conventional sum rules [@BBD1], this approach leads to consistent results in the heavy quark limit [@ABS].
The detailed derivation of the light-cone sum rules for the $D{\rightarrow}\pi$ and $B {\rightarrow}\pi $ form factors is discussed in ref. [@BKR] (see also refs. \[19-21\]). Here we just mention that the sum rule for $f^+_D(p^2)$ is obtained by matching the expressions (\[28\]) and (\[35\]) for the invariant amplitude $F(p^2, (p+q)^2)$ in terms of the pion wave functions with the hadronic representation F(p\^2, (p+q)\^2) = + \_[s\_0]{}\^ . \[physpart\] In the above, the pole term is due to the ground state in the heavy channel, while the excited and continuum states are taken into account by the dispersion integral above the threshold $s_0$. Invoking duality, the latter contributions are cancelled against the corresponding pieces in eqs. (\[28\]) and (\[35\]). After Borel transformation in the variable $(p+q)^2$, the resulting sum rule takes the form\
&& -\_0\^1u du \_3(u,M\^2,p\^2) } , \[formSR\] where $$\begin{aligned}
\Phi_2& = &{\varphi}_\pi(u) +
\frac{\mu_\pi}{m_c}\Bigg[u {\varphi}_{p}(u)
+ \frac16 {\varphi}_{\sigma }(u)
\left(2 + \frac{m_c^2+p^2}{uM^2}\right)\Bigg]
\nonumber\\
&&{}
- \frac{4m_c^2g_1(u)}{u^2M^4} - \frac{2G_2(u)}{uM^2} \left(1+
\frac{m^2_c+p^2}{uM^2} \right) ~,
\label{XXX}
\\
\Phi_3&=& \frac{2f_{3\pi}}{f_{\pi}m_c}
\varphi_{3\pi}(\alpha_i)
\left[1-\frac{ m^2_c -p^2 }{(\alpha_1+u\alpha_3)M^2}\right]
\nonumber\\&&{}
-\frac1{uM^2} \Bigg[2{\varphi}_\perp ({\alpha}_i)-{\varphi}_\parallel ({\alpha}_i)+
2\tilde{{\varphi}}_\perp ({\alpha}_i)-\tilde{{\varphi}}_\parallel ({\alpha}_i)\Bigg]~,
\label{xxz}\end{aligned}$$ and $\Delta = (m_c^2-p^2)/(s_0-p^2) $. The notation is as in eq. (\[fin\]). Improving the approximation given in ref. [@BKR] we have added the contributions of three-particle wave functions of twist 4. The analogous sum rule for the $B {\rightarrow}\pi$ form factor follows from the above by replacing $c {\rightarrow}b $ and $D {\rightarrow}\bar B $, and by rescaling $\mu_\pi$ and the wave function parameters from $\mu_c$ to $\mu_b$.
The maximum momentum transfer $p^2$ at which these sum rules are applicable is estimated to be about 15 GeV$^2$ for B mesons, and 1 GeV$^2$ for D mesons. For numerical evaluation we use the approximations of the wave functions given in Appendix A. We emphasize that the input here is exactly the same as in the calculation of the couplings $g_{D^*D\pi}$ and $g_{B^*B\pi}$ . The form factor $f^+_D( p^2)$ resulting from the sum rule (\[formSR\]) is plotted in Fig. 3a, together with the corresponding prediction (\[onepole\]) of the pole model. We see that in the region of overlap both calculations approximately agree with each other. To a lesser extent, this also applies to the form factor $f^+_B( p^2)$ illustrated in Fig. 3b [^4]. Quantitatively, at $p^2=0$ we find \[fD0\] f\^+\_D(0)\_[SR]{}= 0.66, f\^+\_D(0)\_[PM]{}=0.75 , and \[fB0\] f\^+\_B(0)\_[SR]{}= 0.29, f\^+\_B(0)\_[PM]{}= 0.44 .
Thus, in the regions $m_Q^2-p^2 > O(1$ GeV$^2$) with $Q=c$ and $b$, respectively, the numerical agreement between the light-cone sum rule and the pole model is better than 15% for $f^+_D$, but only within 50% for $f^+_B$.
At this point, we must add a word of caution concerning the applicability of the pole model too far away from the zero recoil point, in particular at $p^2=0$. The two descriptions differ markedly in the asymptotic dependence of the form factors on the heavy mass. Focusing on $B$ mesons and using the familiar scaling laws f\_B= \_B, f\_[B\^\*]{} = \_[B\^\*]{} , \[consthq\] and \[gscale\] g\_[B\^\*B]{} = g , which are expected to be valid at $m_b\to\infty$ modulo logarithmic corrections, one obtains f\^+\_B(0)\_[PM]{} \~1/ , whereas the light-cone sum rule (79) yields [@CZ90] \[SRlim\] f\^+\_B(0)\_[SR]{} \~1/[m\_B\^[3/2]{}]{} . This result rests on the QCD prediction [@exclusive] of the behaviour of the leading twist pion wave function near the end point, that is ${\varphi}_\pi(u)\sim 1-u$ at $u {\rightarrow}1$. It should be noted that the contribution estimated by the sum rules corresponds to the so-called Feynman mechanism. In the case of heavy-to-light transitions it leads to the same asymptotic behaviour as the hard rescattering mechanism [@CZ90; @BurDon]. Recently it has been shown [@Akhoury] that the power behaviour (\[SRlim\]) of hard rescattering is not modified by the Sudakov type double logarithmic corrections. We believe that the light-cone sum rules reproduce the true asymptotic behaviour, although a rigorous proof in QCD is still lacking. On the other hand, we see no theoretical justification for extrapolating the pole model to the region $p^2=0$. The solution suggested by Fig. 3 is to match the two descriptions in the region of intermediate momentum transfer $ p^2\simeq m_Q^2-O(1$GeV$^2)$.
Referring for a detailed discussion to refs. [@ABS] and [@X] we want to emphasize that the light-cone sum rules seem to be generally consistent with the heavy quark expansion. In particular, the light-cone sum rule (\[fin\]) correctly reproduces the heavy quark mass dependence of the coupling $g_{B^*B\pi}$. Fitting our predictions for $g_{B^*B\pi}$ and $g_{D^*D\pi}$ to the form \[1/m\] g\_[B\^\*B]{} = g and the analogous expression for $g_{D^*D\pi}$, we find for the coupling $\hat g$ and the strength $\Delta$ of the $1/m_Q$ correction: g =0.320.02 , =(0.7 0.1) GeV . \[fit\]
Furthermore, we are able to make a numerical prediction for the theoretically interesting ratio 0.92 . \[ratio3\] This ratio is unity in the heavy quark limit and is shown to be subject to $1/m_Q$ corrections only in the next-to-leading order [@BLN]. Our result (\[ratio3\]) is nicely consistent with this expectation. The deviation from unity also agrees in magnitude with the estimate in ref. [@Cetc94], but has a different sign. This is due to a sizeable difference in the ratio $f_{B^*}/f_{D^*}$. While the values of the decay constants given in eqs. (\[fDs\]) and (\[fB\]) yield =1.12 \[rrr\] in agreement with the expectation quoted in ref. [@BLN], the latter ratio turns out to be larger by 30% if calculated from $f_{B^*}$ and $f_{D^*}$ as assumed in ref. [@Cetc94].
Summary and conclusions
=======================
We have presented a comprehensive analysis of the pion couplings to heavy mesons in the framework of QCD sum rules. The main new result of this paper is the light-cone sum rule (\[fin\]) providing the numerical estimates for $g_{D^*D\pi}$ and $g_{B^*B\pi}$ given in eqs. (\[constD\*Dpi\]) and (\[41\]), respectively. The decay width $ \Gamma(D^* {\rightarrow}D\pi)$ predicted in eq. (\[Gamma\]) turns out to be three times smaller than the present experimental upper limit. We have compared our results to earlier QCD sum rule calculations \[10-13\], and resolved the existing discrepancies.
A rather complete compilation of estimates [^5] on the pion couplings to heavy mesons is given in Table 1. In the first row we show predictions on the reduced coupling $\hat g$ defined in eq. (\[gscale\]). As one can see, the values obtained by combining the nonrelativistic constituent quark model with PCAC [@IW90; @NW; @Yan] are roughly two times larger than the values favoured by our sum rule. However, more recent analyses [@CG; @Ametal] combining chiral HQET with experimental constraints on $D^*$ decays tend to give somewhat smaller values of $\hat g$. Moreover, another recent calculation [@BardeenHill] based on the extended Nambu-Jona-Lasinio model and chiral HQET is in perfect agreement with our estimate.
The next two rows list the estimates of the couplings $g_{B^*B\pi}$ and $g_{D^*D\pi}$. These predictions are even wider spread. Quark models [@Eichtetal; @DoXu] seem to give the strongest couplings, whereas SU(4) symmetry [@Kam] and the reggeon quark-gluon string model [@KN] predict a relatively small coupling. Two comments are in order concerning the analysis of ref. [@Cetc94]. Firstly, these predictions are consistently lower than ours. There are several reasons for that: the different Lorentz decompositions (\[220\]) and (\[choice\]) of the correlation function (\[19\]), the differences between the sum rule (\[fin\]) and the soft-pion limit (\[48\]) of it, the different regions of the Borel parameter $M^2$, and the different values used for the decay constants $f_{D^{(*)}}$ and $f_{B^{(*)}}$. In fact, as can be seen in Fig. 2, the couplings shrink with $M^2$. However, given the reliability criteria, generally accepted for sum rules, we see no possibility to shift $M^2$ to larger values beyond the regions considered in this paper, in contrast to ref. [@Cetc94]. Secondly, we find it inconsistent to include the perturbative gluon correction in the estimates of $f_{D,D^*}$ and $f_{B,B^*}$, since they are not included in the sum rule for the combination of couplings $f_Df_{D^*}g_{D^*D\pi}$ and $f_Bf_{B^*}g_{B^*B\pi}$. At least, we see no convincing argument in favour of such a procedure. For these two reasons we believe that the couplings are underestimated in ref. [@Cetc94].
For convenience and direct comparison with future measurements the decay width\
$\Gamma(D^{*+}{\rightarrow}D^0 \pi^+)$ as calculated from $g_{D^*D\pi}$ or $\hat g$ is quoted in the last row of Table 1. The widths in the channels $D^{*+} \rightarrow D^+ \pi^0$ and $D^{*0} \rightarrow D^{0} \pi^0$ are related to the above by coefficients which can be read off from eq. (\[widths\]). Note that in contrast to the evaluation of $\Gamma(D^*{\rightarrow}D\pi)$ from $g_{D^*D\pi}$ in this paper and in ref. [@Cetc94] the estimates in refs. [@Yan; @CG] using the reduced coupling $\hat g$ do not include $1/m$ corrections. However, the latter are important as can be seen from eq. (\[fit\]).
In addition, we have examined the pole model for the $B\to\pi$ and $D\to\pi$ form factors. Using our results on the $g_{B^*B\pi}$ and $g_{D^*D\pi}$ coupling constants, we have found approximate numerical agreement between the pole model description and the direct sum rule calculation. However, the dependence on the heavy quark mass is found to be completely different in the region of small momentum transfers. We have argued in favour of the sum rule approach. Moreover, writing a heavy quark expansion for the couplings $g_{B^*B\pi}$ and $g_{D^*D\pi}$ we have determined the expansion coefficients, in particular, the leading $1/m$ correction.
Last but not least, we have discussed in some detail the theoretical foundations and advantages of the light-cone sum rules, complementing the work of refs. \[14-19\]. We believe that this approach is especially suitable for the study of heavy-to-light decay form factors, and coupling constants of the type considered in this paper. Further obvious applications include the radiative decays $D^*\to D\gamma$ and $B^*\to B\gamma$. Since the photon wave functions are expected to deviate less from their asymptotic forms than the pion wave functions [@BBK], these decays should provide a rather conclusive consistency check of the light-cone approach.
[**Acknowledgements**]{}
V.M. Belyaev is grateful to DAAD for financial support during his visit at the University of Munich. This work is also partially supported by the EC grant INTAS-83-283.
Appendix A {#appendix-a .unnumbered}
==========
For convenience, we collect here the explicit expressions for the pion wave functions used in our numerical calculations and specify the values of the parameters involved.
For the leading twist two wave function we take [@BF1] $${\varphi}_\pi(u,\mu) = 6 u(1-u)\Big[1+a_2(\mu)C^{3/2}_2(2u-1)+
a_4(\mu)C^{3/2}_4(2u-1)\Big]
\label{expansion1}$$ with the Gegenbauer polynomials C\_2\^[3/2]{}(2u-1)=2\[5(2u-1)\^2-1\] ,\
C\_4\^[3/2]{}(2u-1)=8\[21(2u-1)\^4-14(2u-1)\^2+1\] , \[G2\] and the coefficients $a_2=\frac23$, $a_4=0.43$ corresponding to the normalization point $\mu=0.5$ GeV. In the present applications the appropriate scales are set by the typical virtuality of the heavy quark. We choose \_c = 1.3, \_b = 2.4 . \[scales\] Renormalization group evolution of the coefficients $a_2$ and $a_4$ to these higher scales yields a\_2(\_c) = 0.41 , a\_4(\_c) = 0.23 , \[a24c\]\
a\_2(\_b) = 0.35 , a\_4(\_b) = 0.18 . \[a24b\] We stress that the value of ${\varphi}_\pi$ at $u=1/2 $ varies by only 2% when the scale $\mu$ is increased from 0.5 GeV to 2.4 GeV. Obviously, one can neglect this effect given the 15% uncertainty in the value of ${\varphi}_\pi(u=1/2,\mu=0.5$ GeV$)$ quoted in eq. (\[WF12\]).
According to the analysis in refs. [@BF1; @BF2] the set of wave functions of twist three is uniquely specified by the choice of the three-particle wave function ${\varphi}_{3\pi}$. Taking into account the contributions to ${\varphi}_{3\pi}$ up to next-to-next-to-leading order in conformal spin, one has \_[3]{}(\_i)&=&360 \_1\_2\_3\^2 (1+\_[1,0]{}12(7\_3-3)\
&&+\_[2,0]{}(2-4\_1\_2-8\_3+8\_3\^2) +\_[1,1]{}(3\_1\_2-2\_3+3\_3\^2)\] . \[3pi\] This implies for the two-particle wave functions of twist three [@BF2]: \_p(u)&=&1+B\_212(3(u-|[u]{})\^2-1)+B\_418(35(u-|[u]{})\^4 -30(u-|[u]{})\^2+3) \[bbb\] and \_(u)&=&6u|[u]{} , \[tw3\] where $$\begin{aligned}
\label{wf3}
B_2&=&30R\,,
\nonumber\\
B_4&=&\frac32R(4\omega_{2,0}-\omega_{1,1}-2\omega_{1,0})\,,
\nonumber\\
C_2&=&R(5-\frac12\omega_{1,0})\,,
\nonumber\\
C_4&=&\frac1{10}R(4\omega_{2,0}-\omega_{1,1})\,,
\label{ccc}\end{aligned}$$ with $$\begin{aligned}
R&=&\frac{f_{3\pi}}{\mu_\pi f_\pi}\,.
\label{RR}\end{aligned}$$ The coefficients $f_{3\pi}$ and $\omega_{i,k}$ have been determined at the normalization point $\mu = 1$ GeV from QCD sum rules [@CZreport]: f\_[3]{} = 0.0035 \^2, \_[1,0]{}= -2.88, \_[2,0]{}= 10.5, \_[1,1]{}= 0 . After renormalization [@BB] to the relevant scales (\[scales\]), we get f\_[3]{}(\_c) = 0.0032 \^2, \_[1,0]{}(\_c)= -2.63 , \_[2,0]{}(\_c)= 9.62 , \_[1,1]{}(\_c)= -1.05 ,\
f\_[3]{}(\_b) = 0.0026 \^2, \_[1,0]{}(\_b)= -2.18, \_[2,0]{}(\_b)= 8.12, \_[1,1]{}(\_b)= -2.59 . The corresponding numerical values of the coefficients (\[ccc\]) are as follows: B\_2(\_c)=0.41, B\_4(\_c)=0.925, C\_2(\_c)=0.087, C\_4(\_c)=0.054,\
B\_2(\_b)=0.29, B\_4(\_b)=0.58, C\_2(\_b)=0.059, C\_4(\_b)=0.034. In addition, the running of light-quark masses induces a scale dependence of the parameter $\mu_\pi= m_\pi^2/(m_u+m_d)$: \_(1 ) = 1.65, \_(\_c) = 1.76, \_(\_b) = 2.02.
The wave functions of twist four are more numerous. The complete set given in ref. [@BF2] (see also ref. [@gorsky]) includes four three-particle wave functions. However, in leading and next-to-leading order in conformal spin, these are specified by only two parameters: \_(\_i)&=&30\^2 (\_1-\_2)\_3\^2\[13+2 (1-2\_3)\] ,\
\_(\_i)&=&120\^2(\_1-\_2)\_1\_2\_3 ,\
\_(\_i)&=&30\^2\_3\^2(1-\_3)\[13+2 (1-2\_3)\] ,\
\_(\_i)&=&-120\^2\_1\_2\_3\[13+ (1-3\_3)\] . \[tw4gluon\] The two-particle twist 4 wave functions are related to these by equations of motion. To the above order in conformal spin they involve no new parameters and are given by $$\begin{aligned}
g_1(u)&=&\frac{5}2\delta^2\bar{u}^2u^2+\frac{1}{2}\varepsilon\delta^2[
\bar{u}u(2+13\bar{u}u)+10u^3\ln u(2-3u+\frac65u^2)
\nonumber
\\
&&{}+10\bar{u}^3\ln \bar{u}(2-3\bar{u}+\frac65\bar{u}^2)]\,,
\nonumber
\\
g_2(u)&=&\frac{10}3\delta^2\bar{u}u(u-\bar{u})\,,
\nonumber\\
G_2(u)&=& \frac53\delta^2 \bar u ^2 u^2 ~.
\label{tw4}\end{aligned}$$ One of the parameters is defined by the matrix element |g\_s|[d]{}\_\^u|0 = i\^2f\_q\_ . \[delta\] The QCD sum rule estimate of ref. [@nov] yields $\delta^2=0.2$ GeV$^2$ at $\mu=1$ GeV. The remaining parameter is associated with the deviation of twist four wave functions from their asymptotic form. At $\mu=1$ GeV it takes the value $\varepsilon =0.5$ [@BF2]. Renormalization to the relevant scales (\[scales\]) gives \^2(\_c)=0.19, (\_c) =0.45,\
\^2(\_b)=0.17, (\_b) =0.36. This completes the description of the pion wave functions, as far as it is needed for the applications in this paper.
Appendix B {#appendix-b .unnumbered}
==========
Here we derive the substitution (\[39\]) used in the sum rule (\[fin\]) in order to subtract the continuum contribution. For this purpose we have to write the invariant amplitude $F$ given by eqs. (\[28\]) and (\[35\]) in the form of a double dispersion integral: F( p\^2, (p+q)\^2)=\^\_[m\_c\^2]{} \^\_[m\_c\^2]{} \^[QCD]{}(s\_1,s\_2) . \[repr\] Focusing first on the leading contribution (\[Fzeroth\]): F(p\^2, (p+q)\^2)= m\_cf\_\^1\_0 = m\_cf\_\^1\_0 , \[integr\] and changing variable from $u$ to $(m_c^2-p^2)/(s-p^2) $, one obtains F( p\^2, (p+q)\^2)=m\_cf\_\^\_[m\_c\^2]{} . \[integrs\] In general, the wave function ${\varphi}_\pi(u)$ can be expressed as a power series in $(1-u)$: \_(u) = \_k a\_k(1-u)\^k = \_ka\_k( )\^k . \[sum\] Substituting this representation into eq. (\[integrs\]) and introducing formally two variables $s_1$ and $s_2$ instead of $s$, it is easy to rewrite this expression in the form (\[repr\]) with the double spectral density \^[QCD]{}(s\_1,s\_2)=m\_cf\_\_k (s\_1-m\_c\^2)\^k\^[(k)]{}(s\_1-s\_2) . \[dens\] The validity of eq. (\[dens\]) can easily be checked by direct calculation. The above derivation may seem tricky. However, there is a convenient general method [@rad2] to find the appropriate double spectral densities. One takes the Borel transformed invariant amplitude $F( M_1^2, M_2^2)$ and performs two more Borel transformations in the variables $\tau_1=1/M^2_1$ and $\tau_2=1/M^2_2$, to get \_[\_1]{}[B]{}\_[\_2]{}F(1/\_1,1/\_2) = \^[QCD]{}(1/\_1,1/\_2) . Details and useful examples can be found in ref. [@BB94].
To proceed, we apply a double Borel transformation to the dispersion integral (\[repr\]) with $\rho^{QCD}$ given by eq. (\[dens\]): \_[M\_1\^2]{}[B]{}\_[M\_2\^2]{}F=m\_cf\_\_k \^\_[m\_c\^2]{}ds\_1 \^\_[m\_c\^2]{} ds\_2 (s\_1-m\_c\^2)\^k\^[(k)]{}(s\_1-s\_2) e\^[-s\_1/M\_1\^2]{}e\^[-s\_2/M\_2\^2]{} . \[check\] Introducing again new variables $s=s_1+s_2$ and $v=s_1/s$ we can use the $\delta$-function to evaluate the integral over $v$. The result is F(M\_1\^2, M\_2\^2)= m\_cf\_\_k \_[2m\_c\^2]{}\^ ds ( )\^k\_[v=1/2]{} . \[form1\] At $M_1^2=M_2^2=2M^2$ the $v$-dependence of the exponent disappears and the differentiation of the bracket gives a factor $k!$. We then get F(M\^2)= m\_cf\_\_k \_[2m\_c\^2]{}\^ds e\^[-]{}= m\_cf\_\_(1/2)M\^2e\^[-]{} , \[form\] which is the leading contribution in the sum rule (\[fin\]). For arbitrary values of $M_1^2$ and $M_2^2$ a similar expression is obtained, with the argument of the wave function and the Borel parameter in eq. (\[form\]) generalized to $u_0$ and $M^2$ , respectively, as defined in eq. (\[37\]).
We now turn to the problem of subtracting the contributions from excited and continuum states in sum rules. In the usual approximation based on duality, one identifies the spectral functions $\rho^{QCD}$ and $\rho^h$ beyond a given boundary in the $(s_1,s_2)$-plane. Then, the subtraction effectively amounts to restricting the dispersion integrals in eq. (\[repr\]) to the region below this boundary. Ideally, the result should not depend on the precise shape of this region. To be specific, one may take s\_1\^a +s\_2\^a s\_0\^a , \[boundary\] where $s_0$ plays the role of an effective threshold. Popular choices of the duality region are triangles in the $(s_1,s_2)-$ plane corresponding to $a=1$, and squares corresponding to $a\rightarrow \infty$. Since the spectral density (\[dens\]) vanishes everywhere except at $s_1=s_2$, it is actually irrelevant which form of the boundary we adopt [^6].
Using duality as outlined above we have to evaluate the integral in eq. (\[check\]) with the integration region restricted by eq. (\[boundary\]). Changing variables and integrating over $v$ one obtains an expression similar to eq. (\[form1\]), but with the upper limit of integration in $s$ lowered to $2s_0$ and with the addition of surface terms. The latter disappear for $M_1^2=M_2^2$. Hence, one is again led to eq. (\[form\]) with a simple modification of the integration limit: F(M\^2)= m\_cf\_\_k () \_[2m\_c\^2]{}\^[2s\_0]{} ds e\^[-]{}= m\_cf\_\_(1/2)M\^2 . \[formh\] This proves the substitution rule quoted in eq. (\[39\]). It is important to note that the proportionality of the Borel transform $F(M_1^2, M_2^2)$ given in eq. (\[form1\]) to the wave function ${\varphi}_\pi$ at the point $u_0=M_1^2/(M_1^2+M_2^2)$ is generally destroyed by the continuum subtraction. It is only retained in the symmetric point $M_1^2=M_2^2$ implying $u_0=1/2$.
The above procedure is not possible for higher twist contributions which are proportional to zero or negative powers of the Borel parameters. The reason is that the corresponding spectral densities are not concentrated near the diagonal $s_1=s_2$. In fact, the continuum subtraction is rather complicated in these cases. For further discussion we refer the reader to the second paper of ref. [@BB94]. Here, we neglect the continuum subtraction in higher twist terms altogether. This is justified to a good approximation since the corresponding spectral densities decrease fast with $s_1$ and $s_2$ as a consequence of ultraviolet convergence and, hence, the continuum contribution is expected to be small anyway.
[99]{}
N. Isgur and M.B. Wise, Phys. Lett. [**B232**]{} (1989) 113; [**B237**]{} (1990) 527.
M. Neubert, Phys. Lett. [**B264**]{} (1991) 455.
M.A. Shifman, A.I. Vainshtein and V.I. Zakharov, Nucl. Phys. [**B147**]{} (1979) 385, 448.
N. Isgur and M.B. Wise, Phys. Rev. [**D41**]{} (1990) 151.
G. Burdman, Z. Ligeti , M. Neubert and Y. Nir, Phys. Rev. [**D49**]{} (1994) 2331.
B. Grinstein and P.F. Mende, Nucl. Phys. [**B425**]{} (1994) 451 .
P. Ball, V.M. Braun and H.G. Dosch, Phys. Lett. [**B273**]{} (1991) 316.
V.M. Belyaev, A. Khodjamirian and R. Rückl, Z. Phys. [**[C60]{}**]{} (1993) 349.
P. Ball, Phys. Rev. [**D48**]{} 3190 (1993).
V.L. Eletsky and Ya.I. Kogan, Z. Phys. [**C28**]{} (1985) 155.
A.A. Ovchinnikov, Sov. J. Nucl. Phys. [**50**]{} (1989) 519.
A.G. Grozin and O.I. Yakovlev, preprint BUDKERINP–94–3, hep-ph/9401267.
P. Colangelo et al., preprint UGVA–DPT 1994/06–856, hep-ph/9406295.
V.L. Chernyak and A.R. Zhitnitsky, JETP Lett. [**[25]{}**]{} (1977) 510; Yad. Fiz. [**[31]{}**]{} (1980) 1053. A.V. Efremov and A.V. Radyushkin, Phys. Lett. [**[B94]{}**]{} (1980) 245; Teor. Mat. Fiz. [**[42]{}**]{} (1980) 147. G.P. Lepage and S.J. Brodsky, Phys. Lett. [**[B87]{}**]{} (1979) 359; Phys. Rev. [**[D22]{}**]{} (1980) 2157.
S.J. Brodsky and G.P. Lepage, in: Perturbative Quantum Chromodynamics, ed. A.H. Mueller (World Scientific, Singapore, 1989) pp. 93-240.
V.L. Chernyak and A.R. Zhitnitsky, Phys. Rep. [**[112]{}**]{} (1984) 173.
I.I. Balitsky, V.M. Braun and A.V. Kolesnichenko, Sov. J. Nucl. Phys. [**44** ]{} (1986) 1028; Nucl. Phys. [**B312**]{} (1989) 509.
V.M. Braun and I.B. Filyanov, Z. Phys. [**C 44**]{} (1989) 157.
V.L. Chernyak and I.R. Zhitnitsky, Nucl. Phys. [**B345**]{} (1990) 137.
P. Ball, V.M. Braun and H.G. Dosch, Phys. Rev. [**D44**]{} (1991) 3567.
A. Ali, V.M. Braun and H. Simma, Z. Phys. [**[C63]{}**]{} (1994) 437.
V.M. Braun and I. Halperin, Phys.Lett. [**B328**]{} (1994) 457.
V.M. Belyaev, preprint CEBAF-TH-94-09, hep-ph/9404279.
N.S. Craigie and J. Stern, Nucl. Phys. [**B216**]{} (1983) 209.
A.S. Gorsky, Sov. J. Nucl. Phys. [**41**]{} (1985) 1008; ibid. [**45**]{} (1987) 512; ibid. [**49**]{} (1990).
B.L. Ioffe and A.V. Smilga, Pisma v ZhETF [**37**]{} (1983) 250; Nucl. Phys. [**B232**]{} (1984) 109.
I.I. Balitsky and V.M. Braun, Nucl. Phys. [**B311**]{} (1988) 541.
V.M. Braun and I.B. Filyanov, Z. Phys. [**C 48**]{} (1990) 239.
A.V. Radyushkin, preprint CEBAF-TH-93-12, CEBAF-TH-94-13, hep-ph/9406237.
ACCMOR Collab., S. Barlag et al., Phys. Lett. [**B278**]{} (1992) 480.
CLEO Collab., F. Butler et al., Phys. Rev. Lett. [**69** ]{} (1992) 2041.
V.M. Belyaev and Ya.I. Kogan, Pisma v ZhETF [**37**]{} (1983) 611.
G. Burdman and J.F. Donoghue, Phys. Lett. [**B270**]{} (1991) 55.
R. Akhoury and G. Sterman, Phys. Rev. [**D50**]{} (1994) 358.
A. Khodjamirian and R. Rückl, [*in preparation* ]{}.
V.L. Eletsky, [*private communication* ]{}.
S. Nussinov and W. Wetzel, Phys. Rev. [**D36**]{} (1987) 130.
T.-M. Yan et al., Phys. Rev. [**D46**]{} (1992) 1148.
P. Cho and H. Georgi, Phys. Lett. [**B296**]{} (1992) 408.
J.F. Amundson et al., Phys. Lett. [**B296**]{} (1992) 415.
W.A. Bardeen and C.T. Hill, Phys. Rev. [**D49**]{} (1994) 409.
E. Eichten et al., Phys. Rev. [**D21**]{} (1988) 203.
P.J. O’Donnell and Q.P. Xu, Phys. Lett. [**B336**]{} (1994) 113.
R.L. Thews and A.N. Kamal, Phys. Rev. [**D32**]{} (1985) 810.
A.B. Kaidalov and A.V. Nogteva, Sov. J. Nucl. Phys. [**47**]{} (1988) 321.
V.A. Novikov, M.A. Shifman, A.I. Vainshtein and V.I. Zakharov, Nucl.Phys. [**B237**]{} (1984) 525.
V.A. Nesterenko and A.V. Radyushkin, Phys. Lett. [**B115**]{} (1982) 410.
V.A. Beilin and A.V. Raduyshkin, Nucl. Phys. [**B260**]{} (1985) 61;\
P. Ball and V.M. Braun, Phys. Rev. [**D49**]{} (1994) 2472.
[lllll]{}\
Reference & $\hat{g}$ & $g_{B^*B\pi}$ & $g_{D^*D\pi}$ & $\Gamma(D^{*+}{\rightarrow}D^0\pi^+)$ (keV)\
\
\
This paper & 0.32 $\pm$ 0.02 & 29 $\pm$ 3 & 12.5 $\pm$ 1.0 & 32 $\pm$ 5\
\
This paper$^a$ & – & 28 $\pm$ 6& 11 $\pm$ 2 & –\
\
[@ovch]$^a$ &–& 32 $\pm$ 6& – & –\
\
[@Cetc94]$^a$ &0.39 $\pm$ 0.16& 20 $\pm$ 4&9 $\pm$ 1& –\
\
[@Cetc94]$^{a *}$ &0.21 $\pm$ 0.06& 15 $\pm$ 4 &7 $\pm$ 1& 10 $\pm$ 3\
\
[@NW]$^b$ & 0.7&–& –& –\
\
[@IW90]$^{b}$ &–& 64&–&–\
\
[@Yan]$^b$ & 0.75 $\div$ 1.0&–& –& 100 $\div$ 180\
\
[@CG]$^c$ & 0.6 $\div $ 0.7 & – & – & 61 $\div$ 78\
\
[@Ametal]$^c$ & 0.4 $\div$ 0.7 & – & – & –\
\
[@BardeenHill]$^d$ & 0.3 & – & – & –\
\
[@Eichtetal]$^e$ & –& – & 16.2&53.4\
\
[@DoXu]$^f$ & –& – &19.5 $\pm$ 1.0&76 $\pm$ 7\
\
[@Kam]$^g$ & –& – &8.9&16\
\
[@KN]$^h$ & –& – &8.2& 13.8\
\
Experiment$^i$&–& – &$<$ 21 & $<$ 89\
\
\
$^a$ QCD sum rules in external axial field or soft pion limit.\
$^*$ including perturbative correction to the heavy meson decay constants.\
$^b$ Quark model + chiral HQET.\
$^c$ Chiral HQET with experimental constraints on $D^*$ decays.\
$^d$ Extended NJL model + chiral HQET .\
$^e$ Quark Model + scaling relation.\
$^f$ Relativistic quark model.\
$^g$ SU(4) symmetry.\
$^h$ Reggeon quark-gluon string model.\
$^i$ Combination of ACCMOR \[33\] and CLEO \[34\] measurements.\
[^1]: This feature is also observed in deep inelastic scattering, with the variables $\{ Q^2, ~\nu, ~x \}$ playing the role of $ \{-p^2,~ p\cdot q ,~ \tilde{\xi} \}$ . As well known, there one applies an expansion near the light-cone in terms of operators of increasing twist, rather than of increasing dimension.
[^2]: In addition, one obtains two-point correlation functions because of contact terms. These do not lead to double poles in $p^2$ in the relevant dispersion relation and are therefore eliminated by applying the differentiation operator (\[oper\]).
[^3]: As a side remark, the light-cone approach leading to eq. (\[fin\]) corresponds to a calculation in the background of a [*variable*]{} external axial field [@BBK; @BF1].
[^4]: The dependence of eq. (\[formSR\]) on the Borel parameter is weak [@BKR]. For definiteness, we take here $M^2=4$ GeV$^2$ for the $D{\rightarrow}\pi$ form factor and $M^2=10$ GeV$^2$ for the $B{\rightarrow}\pi$ form factor.
[^5]: We have not included the results of ref. [@EK85] since to our knowledge this analysis is being reconsidered [@PC:Eletsky]. The result of ref. [@GY94] is omitted for reasons explained in Sect. 5.
[^6]: This is literally true only if the power series defining the wave function (\[sum\]) is truncated at some finite order, or if it converges rapidly. However, this condition is always met at a sufficiently high normalization point where the wave function deviates little from the asymptotic form.
| {
"pile_set_name": "ArXiv"
} |
---
author:
- Mukund Rangamani
- ', Moshe Rozali'
- ', Anson Wong'
title: Driven Holographic CFTs
---
Introduction and Conclusions {#sec:intro}
============================
The dynamics of quantum field theories driven far from equilibrium is a fascinating topic, owing to the complex interplay of quantum and statistical behaviours in the system. While a quantitative understanding of how field theories respond to non-linear external sources remains in general an open problem, in recent years one has gained some insight into such phenomena.
On the one hand progress in this direction has been driven by experimental developments which allow for a detailed study. For instance the ability to simulate many-body dynamics in cold-atom systems has led to the opening of a new frontier in dynamical simulations, cf., [@Polkovnikov:2010yn] for a recent review. On the other hand, theoretical horizons have been broadened with the gauge/gravity duality providing an excellent arena to explore the dynamics of strongly interacting many-body systems using (classical) gravitational dynamics in a suitable limit (cf., [@Hubeny:2010ry] for a not so recent review). Coupled with the development of excellent numerical algorithms for studying dynamical problems in AdS gravity [@Chesler:2008hg; @Chesler:2013lia], the confluence of ideas and techniques provides an excellent opportunity to further our understanding of out-of-equilibrium dynamics.
A much studied protocol in this context is the quantum quench dynamics, wherein one takes a system initially in equilibrium, typically in the ground state, and subjects it to external sources which change the subsequent dynamics by modifying the Hamiltonian. The rate at which sources act on the system controls the features of the subsequent relaxation, assuming that the sources are non-vanishing for a finite amount of time. The analysis of such a quench protocol has benefited both from theoretical understanding using standard quantum field theory technology in low dimensions [@Calabrese:2005in; @Calabrese:2006rx; @Calabrese:2007rg] and from a wide array of examples that have been studied holographically in the recent past [@Bhattacharyya:2009uu; @Das:2010yw; @Basu:2011ft; @Buchel:2012gw; @Bhaseen:2012gg; @Basu:2012gg; @Nozaki:2013wia; @Buchel:2013lla; @Hartman:2013qma; @Basu:2013vva; @Li:2013fhw; @Buchel:2013gba; @Auzzi:2013pca; @Buchel:2014gta]. In most cases the interest is in the approach to equilibrium at late times and the rate at which various observables thermalize [@Danielsson:1999fa; @Hubeny:2007xt; @AbajoArrastia:2010yt; @Albash:2010mv; @Balasubramanian:2010ce; @Balasubramanian:2011ur; @Aparicio:2011zy; @Balasubramanian:2011at; @Keranen:2011xs; @Galante:2012pv; @Caceres:2012em; @Hubeny:2013hz; @Hartman:2013qma; @Liu:2013iza; @Balasubramanian:2013oga; @Liu:2013qca; @Abajo-Arrastia:2014fma]. Note that since we inject energy in the process of the quench, even an initially pure state will appear to be well approximated by a thermal ensemble asymptotically (assuming that the field theory dynamics are sufficiently ergodic).
A slightly different but related scenario is one where we subject a system, again initially in an equilibrium configuration, to an external driving source which keeps doing work on it throughout the entire time period under study. More specifically, we will be interested in examining the behaviour when the initial state is chosen to be a thermal density matrix, so that one can simultaneously explore the response of a quantum dissipative system. For non-linear dynamical systems the response under such external driving can provide insight into the dynamics via the coherent build-up of the response.
Classical analogs of what we have in mind are situations where we drive a (damped) pendulum steadily or subject a viscous fluid to external forcing. The latter is particularly apposite, for the problem we study can be thought of as a hot deconfined plasma of a planar gauge theory disturbed by an external source, as studied in the hydrodynamic context in [@Bhattacharyya:2008ji]. Rather than letting the driving grow without bound, we will subject our plasma to a periodic driving by turning on the source for a relevant operator. One therefore has two relevant dimensionful parameters characterizing the situation: (a) The amplitude of the external force whose scaling dimension is set by the conformal weight of the operator we exploit and (b) The frequency of the external driving. The third scale which is the temperature of the initial equilibrium state can be factored out, if we are interested in describing the dynamics for a conformally invariant system, which is most natural in the gauge/gravity context. This scenario was explored in [@Auzzi:2013pca], who carried out a perturbative analysis for small amplitudes of the driving source. A related analysis of periodically driving a quantum system near a critical point was undertaken in [@Li:2013fhw].
Gravitationally the problem we study is the following: we have a Schwarzschild- black hole modeling our initial thermal density matrix of a three-dimensional CFT. At some instant of time on the boundary we turn on a periodic source for a relevant scalar operator, which we specifically choose to be of dimension $2$ for simplicity.[^1] The physics of the system is captured by examining the behaviour or various observables as we vary the amplitude $\amp$ and the period $\per$ of the driving (measured e.g. in units of the initial temperature). We will in particular extend the perturbative analysis of [@Auzzi:2013pca] valid for $\amp \ll 1$ to the non-perturbative regime $\amp \gg 1$ for a wide range of driving frequencies. We find that the system naturally exhibits at least four different phases which are depicted in phase diagram Fig. \[fig:PP\_qualitative\]; two of these (labeled IIb and III) are non-perturbative in $\amp$.
![The “phase diagram” of the driven holographic plasma characterized by the period ($\per$) and amplitude ($\amp$) of the driving force, measured in units of the initial thermal scale $T_0$. There are four distinct regimes marked on the diagram which are explained in the main text. $\sigma_\text{in}$ refers here to the imaginary (or in-phase) part of the conductivity defined in Eq. . As we move from southwest to northeast in the figure, the system is driven into a more non-linear regime; the crossing of the grey-dashed boundary is the turn on of the in-phase part of the conductivity $\sigma_{\text{in}}$ in regime II, and the crossing of the blue-dashed boundary signifies the entrance into the resonance phase of regime III i.e., $|\phi^{\text{max}}_{1}| \rightarrow \infty$. The character of the different regimes is further illustrated by displaying the phase portrait of the scalar operator (expectation value against source) used to drive the plasma.[]{data-label="fig:PP_qualitative"}](figures/phase_diagram_qualitative.pdf){width="85.00000%"}
(0.3,0.4)(0,0) (0.15,0.56)[(0,0)[ $\per\, T_0$]{}]{} (-13.04,10.5)[(0,0)[ $\amp/T_0$]{}]{}
Before we describe the different phases, let us examine for a moment the physics of the gravitational system qualitatively. Initially we have a planar black hole in . When we turn on the scalar source, we are injecting energy into the bulk. This energy does work on the system and simultaneously heats it up. The latter is seen by the fact that some of the energy falls behind the horizon, which grows[^2] – this is the gravitational response to the disturbance of the plasma. However, in this process we also induce an expectation value for the operator whose source we tweak. When we disturb the system ‘slowly enough’, the operative parameter measuring this being the product of the amplitude and the period, the system has time to catch-up. This is the dissipation dominated regime indicated by I in Fig. \[fig:PP\_qualitative\]. In this regime the injected energy falls behind the horizon with little fanfare.
As we ramp up the disturbance, the plasma is driven more and more non-linear, with a dynamical cross-over visible as we move into phases IIa or IIb of Fig. \[fig:PP\_qualitative\]. Note that the entire non-linear dynamics in the system is induced by the non-linearities of gravity, for we model the system simply by a free (massive) scalar field. In this phase the response gets more in-phase with the source. It is amusing to contrast this with non-linear scalar dynamics; we find that in this phase we can model the scalar 1PI effective action induced from the gravitational interactions to be well mimicked by a polynomial potential (see [@Basu:2013vva] for previous studies of self-interacting scalars in AdS). In this regime there is less dissipation; the entropy production by the growth of the horizon area is slowed down relative to region I. The primary distinction between the two phases IIa and IIb themselves is the lag in the response seen as the period is increased (hence the tilt in the phase portrait).
For even larger disturbances, we enter region III, where the system response gets highly resonant and there is a steep growth in the response. As one might suspect this is the domain where the gravitational non-linearities are strongest and indeed one can check that such behaviour is not visible for a polynomially (self-) interacting scalar. In the course of our investigation we explore not just the phase portrait, but various other physical quantities of interest, such as the growth of entropy and dissipation in the system, the rate at which entanglement is produced, etc.. For instance, region IIb is characterized by enormous fluctuations in the energy of the system over a single period and continuous but non-differentiable behaviour in the entanglement entropy of a sub-system.
Let us contrast our results with the analysis in the perturbative regime of small amplitudes undertaken in [@Auzzi:2013pca].[^3] As one can see from phase diagram Fig. \[fig:PP\_qualitative\] for small amplitudes, $A \ll 1$, one is largely in the dissipation dominated linear response regime. This is indeed consistent with the analysis of [@Auzzi:2013pca], who explore the dependence of observables on both the period of the driving as well as the dimension of the perturbing operator $\Delta$. As for us the latter remains frozen and we are unable to check the detailed scaling relations they find, but in the common domain of overlap we do indeed have agreement. In particular, for perturbing operators of dimension $\Delta =2$ in CFT$_3$ we expect to see that the energy dissipation as a function of the period scales as $E_\text{diss} \sim \per^{-1}$ (for $\amp \ll 1$), independent of the initial temperature. Furthermore, we also expect that the work done in each cycle, measured by the entropy produced, to scale with the increased energy density. We find that in the slow driving regime this scaling closely tracks the prediction from local thermal equilibrium, but starts to grow more steeply as we transit into more interesting non-linear regimes.
While the various response functions provide us with a useful diagnostic of the phase structure of the dynamical evolution, we also attempt to gain insight into the non-equilibrium dynamics using entanglement entropy for small sub-systems. This non-local probe exhibits distinct characteristic features in the various regimes: for weak driving, the growth of entanglement is gradual (and appears to track the growth of thermal entropy), while for strong driving there are steep oscillations and glitches in its evolution. We should caution the reader that we have only examined entanglement entropy for relatively small sub-systems, owing to technical complications with numerical stability. Nevertheless these results suggest a rather rich structure in the temporal growth of entanglement with driving, which deserves further detailed exploration [@Rangamani:2015ys].
The outline of this paper is as follows. We begin in §\[sec:setup\] by giving a quick overview of the basic set-up and the numerical solutions. Following this in §\[sec:obs\], we set out the various observables we use to explore the behaviour of the system. In particular, we justify the rationale behind phase diagram Fig. \[fig:PP\_qualitative\] and how we should physically think of the different regimes. §\[sec:ee\] is devoted to the study of entanglement entropy in this system where we focus on the region of an infinite strip and exploit the underlying homogeneity of the set-up. We conclude with a discussion in §\[sec:discuss\]. Some technical results about holographic renormalization required for computing various observables is collected in the Appendices; Appendix \[sec:holren\] collects some useful information about holographic renormalization in our models while Appendix \[sec:eeapp\] provides details relevant for computing entanglement entropy.
Driven CFTs and their Holographic Duals {#sec:setup}
========================================
We first take the opportunity to set up the basic problem of a field theory driven out of equilibrium by turning on a source for a relevant operator. We then go on to describe how to model this in the holographic set-up and present the basic methodology and results from the numerical simulations.
Driving CFTs by Relevant Operators {#sec:cftdriving}
----------------------------------
We are interested in the dynamics of strongly coupled plasmas that are driven by an external source. The initial plasma is in equilibrium in some homogeneous thermal state at a temperature $T_0 $ for $t <0$. At $t=0$ we introduce external sources with some specified spatial-temporal profile that we control. We focus exclusively on situations where the external sources are spatially homogeneous, but otherwise arbitrary and tunable at will.
To wit, the system under consideration can be modeled by an equilibrium density matrix, evolved under a time-dependent Hamiltonian, i.e., we take $$S_{CFT} = S_{{\cal J}=0} + \int d^d x \sqrt{-\gamma}\; {\cal J}(x)\, {\cal O}(x)
\label{}$$ where ${\cal O}(x)$ is a single trace (gauge-invariant) relevant operator of conformal dimension $\Delta<d$. The source ${\cal J}(x)$ is chosen to have no spatial dependence and be temporally periodic and thus can be represented as $${\cal J}(x) = \amp \, \cos(\omega t)\, \Theta(t) \,.
\label{}$$ Here $\Theta(t)$ is the Heaviside step function for turning on the periodic perturbation of amplitude $\amp$ and driving frequency $\omega = 2 \pi / \per$ at $t=0$; later in actual (numerical) implementations we will choose a suitable ramp factor to smoothly turn the perturbation on.
In the presence of the source, the Ward identities following from diffeomorphism and Weyl invariance get modified. A simple analysis shows that the boundary conservation equation now has an explicit source term $$\nabla_\mu T^{\mu}_{\ \alpha} = {\cal O}\, \nabla_\alpha {\cal J} \,.
\label{eq:cward}$$ indicative of the work done by the driving source on the CFT. Likewise the one-point function of the trace of the energy-momentum tensor no longer vanishes but satisfies $$T^\mu_\mu = \left(d-\Delta \right)\, {\cal J}(x) {\cal O}(x)
\label{eq:tward}$$ Since the boundary theory is conformal, it does not have any intrinsic time scale. The time scales in the problem come from only the driving force, namely its amplitude and period. The situation of interest is thus characterized by three scales:
- $T_0$: the initial thermal scale for the homogeneous plasma.
- $\amp$: the amplitude of the source whose scaling dimension is $d-\Delta$.
- $\omega$: the driving frequency or the time-scale set by the period $\per = 2\pi/\omega$.
Holographic Driving {#sec:gdual}
-------------------
The gravity dual to this set-up is modeled by the dynamics of a scalar field $\phi$ with mass $m_\phi^2=-2$, dual to a relevant perturbation of the boundary theory. $$S_{\text{bulk}} = \frac{1}{16\pi G_N}\; \int d^{d+1} x\; \sqrt{-g} \, \left( R + d(d-1) - \frac{\alpha_g}{2} \left[ \, (\partial \phi)^2 + m^2 \phi^2 \right] \right)
\label{eq:bulkS}$$ In our holographic implementation of this set-up we will work in $d=3$ and consider a scalar operator with conformal dimension $\Delta =2$. While this is rather specific, we will explore the phase structure of the driven system as a function of the ratio of scales outlined above. The qualitative features we believe are independent of these actual choices.[^4] We have included a dimensionless gravity-scalar coupling $\alpha_g$ which we can use to tune the amount of backreaction on the geometry; for the most part we will focus on $\alpha_g = 0$ or $\alpha_g =1$, to model probe and interacting scalar fields respectively.
We want to study gravitational dynamics driven by a scalar field whose non-normalizable mode is turned out as dictated by the source ${\cal J}(x)$, i.e., take $\phi_0(t) = \amp\,\cos(\omega t)$ and study the behaviour of the theory with varying amplitude $\amp$ and frequency $\omega$. The gravitational background is an asymptotically spacetime, which we write in ingoing Eddington-Finkelstein coordinates (sometimes called the Bondi-Sachs form) as: $$ds^2 =-2 \, \gtt (t,r)\, e^{2 \br (t,r)}\,dt^2+2 \,e^{2 \br (t,r)}\,dt \,dr + \gxx (t,r)^2\, (dx^2+dy^2)
\label{eq:bulkcy}$$ The coordinate dependences of the metric functions $\gtt$, $\br$, $\gxx$ are explicitly indicated with homogeneity ensuring that $\partial_x$ and $\partial_y$ are Killing vector fields.
Our initial state is a planar Schwarzschild- black hole with temperature $T_0=3/\pi$, corresponding to horizon size $r_+ =1$. This bulk solution is given by $\gtt = r^2(1-\frac{1}{r^3})$, $\br = 0$, $\gxx = r$ with metric $$ds^2_{t \le 0} = - r^2 \left(1-\frac{1}{r^3}\right)\, dt^2 +2\, dt\, dr + r^2\, \left(dx^2 + dy^2\right).
\label{}$$ For our choices of $m_\phi^2 = -2$ in $d=3$, the amplitude $\amp$ has mass dimension $1$. Thus we have two interesting time scales associated with the external driving force: the period $\per$ and the inverse amplitude $\amp^{-1}$. To capture universal physics, we look at relatively late times of the non-thermalized system compared to both of these scales. Note also that in those late times the initial value of the temperature, $T_0$, becomes irrelevant.
There has been much interest recently in holographic [*quenches*]{}, in which the system is initially driven to an excited state, and then is allowed to return to equilibrium, a process which exhibits some degree of universality. In contrast, we are interested in the dynamics of the steady state system while it is being driven. Hence, in our solutions we do not turn off the driving force at late times, and seek universal features associated with the driven steady state system. We will see that such dynamical features exist, and they strongly depend on the parameter $$\xi (\per,\amp) \equiv \per\,\amp \,,
\label{eq:xidef}$$ the unique dimensionless parameter formed from the two time scales associated with the driving force. Below we refer to the regime $\xi \ll 1$ as the weak driving regime, and $\xi \gg 1$ as the strong driving regime (which is further divided into two separate dynamical regimes). We also measure time in units of the period $\per$, thus we vary and discuss the dependence of observables on the two dimensionless parameters: the strength of the drive and time.
Bulk solutions {#sec:bulksol}
--------------
We solve the equations of motion resulting from the scalar-gravity Lagrangian by direct numerical integration. The boundary conditions on the scalar are prescribed by the source and the metric is required to be asymptotically . The AdS boundary is attained as $r\to \infty$ and the asymptotic behaviour of the fields is $$\begin{aligned}
\phi(t,r) &= \frac{\phi_0(t)}{r} + \frac{\phi_1(t)}{r^2} + {\cal O}(r^{-3})
\nonumber \\
\gxx (t,r) &= r + \lambda(t) - \frac{\alpha_g}{4}\, \frac{\phi_0(t)^2}{r} + {\cal O}(r^{-2})
\nonumber \\
\gtt (t,r) &= \frac{1}{2} ( r + \lambda(t))^2 - \lambda'(t) - \frac{\alpha_g}{4}\, \phi_0(t)^2
+ {\cal O}(r^{-1})
\nonumber \\
\chi(t,r) &= {\cal O}(r^{-4})
\label{}\end{aligned}$$
[0.45]{} ![A sample solution displaying the scalar field $\phi(t,u)$ and the temporal component of the metric function $\gtt (t,u)$ for $\xi(\per=1, \amp=1)=1$. Time is measured in units of $\per$ and the radial component is compactified as $u = 1/r$. []{data-label="fig:sample_solution"}](figures/phi_amp=1_per=1_tf=10.pdf "fig:"){width="\textwidth"} (-47,8)[(0,0)[$\frac{t}{\per}$]{}]{} (-160,13)[(0,0)[$u$]{}]{} (-202,95)[(0,0)[$\phi$]{}]{}
[0.45]{} ![A sample solution displaying the scalar field $\phi(t,u)$ and the temporal component of the metric function $\gtt (t,u)$ for $\xi(\per=1, \amp=1)=1$. Time is measured in units of $\per$ and the radial component is compactified as $u = 1/r$. []{data-label="fig:sample_solution"}](figures/gtt_amp=1_per=1_tf=10.pdf "fig:"){width="\textwidth"} (-47,8)[(0,0)[$\frac{t}{\per}$]{}]{} (-160,13)[(0,0)[$u$]{}]{} (-202,95)[(0,0)[$\gtt$]{}]{}
More specifically, we use the characteristic formulation of the resulting partial differential equations as explained in detail in [@Chesler:2013lia] to numerically integrate for the solution. The advantage of the method is that it allows us to use constrained evolution: at each time step we solve a nested set of ODEs to determine the time derivatives of all dynamical quantities, and then we use one of the standard time evolution schemes to march forward in time. While we follow the general logic of [@Chesler:2013lia], in our implementation we found that some of elements described in [@Balasubramanian:2013yqa] enabled for a more robust evolution.
To solve the radial ODEs we discretize the equations using a Chebyshev basis in the radial direction, typically taking a grid of 60 points. For time evolution we use an explicit Runge-Kutta method of order 4, with an adaptive step size. We filter at each time step by throwing out the top third of the Fourier modes for each dynamical variable to avoid artificial and unphysical growth in amplitudes of short wavelength modes associated with the UV cutoff.
In the regime of strong driving, we found it necessary to turn on the perturbation gradually from zero. Therefore we include a ramp-up time of $2\,\per$, after which the amplitude reaches its intended value. Thus, the first few periods of each solution show behaviour sensitive to details of the ramp-up protocol. We look at observables only after this ramp-up time of $2\,\per$.
In Fig. \[fig:sample\_solution\] we show one example of evolved bulk fields for a specific solution. As we perturb the system by a relevant operator, the scalar field grows towards the horizon. All fields are (at least approximately) modulated with the period of the source.
At this point it is worthwhile mentioning one important consistency check on the numerical scheme, which relies on the existence of a smooth horizon in the spacetime. Given a metric and a Cauchy slice in the bulk spacetime, one can find the outermost trapped surface on this slice. If we have a set of Cauchy slices that foliate the spacetime, then the future outermost trapping horizon, which we simply refer to as the apparent horizon by a common abuse of terminology, is typically defined by taking the union of the outermost trapped surface on all the slices. The apparent horizon thus defined is subject to an area law which was originally discussed by [@Hayward:1993wb] – we refer the reader to [@Booth:2005qc] for a concise modern summary and proof of the statement. It is however important to note that the statement relies on the existence of a sensible foliation of the spacetime by Cauchy slices. Indeed, it is possible as discussed in [@Wald:1991zz] to find exotic symmetry breaking foliations (which are however incomplete) in which even the Schwarzschild black hole solutions fails to have a trapped surface.
We mention this in passing, as [@Auzzi:2013pca] quotes the result of [@Wald:1991zz] to argue that apparent horizon areas need not be monotone generically. They however do not encounter such behaviour, for with the choice of ingoing coordinates in , there is a canonical choice of bulk Cauchy slices respecting the homogeneity of the disturbance. In this foliation the result quoted in [@Booth:2005qc] does apply and in fact simply follows from properties of null congruences using Raychaudhuri’s equation.[^5] Our results are indeed consistent with this expectation and we have checked that the area of the apparent horizon does grow monotonically in $t$ (which labels the leaves of the foliation chosen), as we shall extensively see in the sequel. While initial results of [@Buchel:2014gta] appeared to suggest otherwise, upon closer scrutiny, one finds that in numerical analyses so far the area of the apparent horizon does respect the second law as derived by [@Hayward:1993wb].[^6]
Driving Diagnostics {#sec:obs}
===================
Having constructed the holographic duals we now turn to lessons that can be extracted from the geometry for the dynamics of strongly coupled field theories. A-priori there are a number of observables which are useful probes of the out-of-equilibrium situation and we will focus on those that offer most clear insight into the dynamics. Our primary goal is to quantify the behaviour of the system as a function of $\{\per,\amp\}$ and construct a phase diagram demarcating the various regimes in this phase space. Let us quickly enumerate the observables we will use and proceed to explain why they give us some insight into the dynamics:
- The phase portrait of response $\phi_1(t)$ as a function of the source $\phi_0(t)$. Alternatively, this relation can be codified in a conductivity $\sigma(t)$, as defined below in . We find 4 underlying phases regions that the system can fall into.
- The $\phi_1$-$\phi_0$ phase portrait features for polynomial and non-polynomial potentials with the gravity-scalar coupling $\alpha_g$ switched on and off.
- The cycle-averaged thermodynamics quantified by the energy density $\epsilon_\text{avg}(t)$ and entropy density $s_{\text{avg}}(t)$, and the scaling relation $s_{\text{avg}} \sim \epsilon^{\gamma}_\text{avg}$ between them.
- The work done in each cycle, measured as the difference in average energy between two successive cycles, $\epsilon_\text{cycle} = \epsilon_\text{avg}^{(n+1)} - \epsilon_\text{avg}^{(n)}$. We typically take $n$ to correspond to the penultimate cycle of our simulation.
- Fluctuations $\epsilon_\text{fluc}(t)$ in the energy density around $\epsilon_\text{avg}(t)$ and the maximal response $|\phi^{\text{max}}_1(t)|$.
- Entanglement entropy and extremal surface evolution for fixed spatial strips ${\cal A}$ on the boundary.
When the system is driven by an external source, the most basic quantity is the response, which is characterized by the scalar one-point function in the presence of the source. In linear response theory, this can be obtained from the retarded Green’s function of the operator ${\cal O}(x)$ evaluated in equilibrium. We are not just interested in the linear response regime, which would correspond in our set-up to $\amp \ll T_0$, but in the full non-linear response. To visualize the response of the strongly coupled plasma, especially in the non-linear regime, where its phase relative to the source is important, we will find it instructive to exhibit the phase portrait, the trajectory traced by the system in the $\phi_0$-$\phi_1$ plane. We also codify the relation between scalar source and response by a complex conductivity, defined below.
In addition to the one-point function of the operator deforming the CFT, we are interested in the boundary energy-momentum tensor. This can be decomposed in to an energy density $\epsilon(t)$ and a pressure. In the holographic set-up one has $$\vev{{\cal O}(t)} = \phi_1(t)\,, \quad \epsilon(t) = \vev{T^t_{\ t} (t)} \,, \quad p(t) = \vev{T^i_{\ i} (t)}
\label{}$$ The scale Ward identity implies that pressure is not an independent observable since it can be obtained from knowledge of $\epsilon(t)$ and $\phi_1(t)$, so we will not discuss the pressure separately. Additionally, to probe the local thermodynamics we will monitor the local entropy density $s(t)$, obtained by computing the area of the apparent horizon at time $t$.[^7]
The dynamics of the bulk gravitational fields encode the heat production resulting from supplying external energy to the system. We monitor the explicit time dependence of the energy density $\epsilon(t)$ and the entropy density $s(t)$ along with their values averaged over each driving cycle period $\per$, and find for the most part that the averaged values are increasing with time.[^8] These provide a useful diagnostic of the departure from equilibrium, as one can monitor the scaling relation to infer the local thermodynamic equation of state. We define the thermodynamic scaling exponent $\gamma$ when the system is in a steady state $t>t_{\text{s}}$ via $$s_{\text{avg}} \sim \epsilon_{\text{avg}}^{\gamma}
\label{eq:gammadef}$$ Note that in thermal equilibrium, conformal invariance predicts $\gamma_0 = \tfrac{2}{3}$. We will encounter this and other scaling regimes in our driven system when conformal invariance is broken.
Note that one natural set of non-local observables we could use are the multi-point correlation function for gauge invariant local operators, perhaps for ${\cal O}$ itself. However, realistically this computation involves solving the wave-equation for the linearized scalar fluctuations on top of the background we have constructed, together with the imposition of suitable boundary conditions on the future horizon, to obtain sensible time-ordered correlation functions. These boundary conditions are somewhat tricky to implement (see however [@CaronHuot:2011dr; @Chesler:2011ds]) – we will therefore postpone a discussion of correlators to the future.[^9]
Below we describe the behaviour of the observables mentioned above in three distinct dynamical regimes, and comment on the bulk interpretation of those regimes. Once we have gained sufficient intuition from this exercise, we will then examine the entanglement entropy for a specified boundary region.
Dissipation Dominated Regime {#sec:dissdom}
----------------------------
The simplest situation occurs in the regime of weak driving $\xi \ll 1$, which is best described as the *dissipation-dominated regime* (phase I). This includes the regime of small amplitudes, studied perturbatively in [@Auzzi:2013pca]. In this weak driving regime, the behaviour of all observables is dominated by dissipation, which we now demonstrate by looking at some specific observables.
As we drive the system by the scalar non-normalizable mode $\phi_0$ it is instructive to divide the scalar response $\phi_1$ to the part in-phase with the driving force, and the part completely out-of-phase with the perturbation. In analogy with an electromagnetic perturbation in linear response, we can complexify the time dependence of the scalar field[^10] and define a complex conductivity $$\sigma(t) \equiv \frac{1}{i \omega} \frac{\phi_1(t)}{\phi_0(t)} = \sigmaOut(t) + i \, \sigmaIn(t) .
\label{eq:sigdef}$$ With this notation the out-of-phase and in-phase parts of the response correspond to the real and imaginary parts of the complex conductivity, $\sigmaOut(t)$ and $\sigmaIn(t)$, respectively. This is the usual convention for the more familiar conductivity, related to electromagnetic perturbations.
(0.3,0.4)(0,0) (-108,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-203,92)[(0,0)[$\tilde{\phi}_1$]{}]{}
As shown in Fig. \[fig:PP\_boring\] in the low driving regime the scalar response is precisely out of phase with the scalar source, $\sigmaIn=0$, meaning all the energy is dissipated and none of it used to excite the internal energy associated with the scalar field i.e., no work is being done on the system. This is the quench limit and it matches with what we expect from the behaviour of the perturbation in linear response. The complex conductivity $\sigma=\sigmaOut$ is purely real and has constant amplitude as a function of time at high frequencies.[^11] This is manifested in the final steady state being reached almost immediately and consisting of closed untilted trajectories in phase space. As we shall see below, tilting of the trajectories in phase space is indicative of non-trivial response and work done onto the system. Fig. \[fig:conductivity\_phasediagram\] shows what fraction of the complex conductivity $\sigma_{\text{out}}$ is present on each point on the $(\per,\amp)$ phase diagram, and for what we are concerned with currently, the system has the response being completely out-of-phase with the source when the period is low.
(-60,13)[(0,0)[$\log_{10} \per$]{}]{} (-235,23)[(0,0)[$\amp$]{}]{} (-265,225)[(0,0)[$\left| \frac{\sigma_{\text{in}}}{\sigma} \right|$]{}]{}
Both the energy and entropy density, averaged over each cycle, grow linearly with time in the dissipation-dominated regime . As the black hole grows, its entropy growth tracks its energy growth at a slightly higher rate than the equilibrium relation $s_{\text{avg}} \sim \epsilon_{\text{avg}}^{2/3}$, i.e., $\gamma \gtrsim 2/3$. This entropy-energy scaling is shown in Fig. \[fig:S\_vs\_E\_boring\] along with their own evolution with time. Note that the expansion of the black hole horizon is not necessarily adiabatic (as measured e.g., by the rate of entropy increase $\frac{1}{T}\frac{\dot{S}}{S}$).
[0.43]{} ![The fitted average entropy $s_{\text{avg}}$ versus the average energy $\epsilon_{\text{avg}}$ (left) and their individual values as a function of time (right) for $\xi(\per = 0.01, \amp = 1) = 0.01$. Fitting for $s_{\text{avg}} \sim \epsilon_{\text{avg}}^{\gamma}$, we find a fitted value of $\gamma=0.6682 \pm 0.0023 \gtrsim \tfrac{2}{3}$ with 95% confidence.[]{data-label="fig:S_vs_E_boring"}](figures/SvE_amp=1_per=0.01_tf=0.1.pdf "fig:"){width="\textwidth"} (-82,-8.0)[(0,0)[$\tilde{\epsilon}_{\text{avg}}$]{}]{} (-194,90)[(0,0)[$\tilde{s}_{\text{avg}}$]{}]{}
[0.465]{} ![The fitted average entropy $s_{\text{avg}}$ versus the average energy $\epsilon_{\text{avg}}$ (left) and their individual values as a function of time (right) for $\xi(\per = 0.01, \amp = 1) = 0.01$. Fitting for $s_{\text{avg}} \sim \epsilon_{\text{avg}}^{\gamma}$, we find a fitted value of $\gamma=0.6682 \pm 0.0023 \gtrsim \tfrac{2}{3}$ with 95% confidence.[]{data-label="fig:S_vs_E_boring"}](figures/SEvt_amp=1_per=0.01_tf=0.1.pdf "fig:"){width="\textwidth"} (-200,122)[(0,0)[$\tilde{\epsilon}_{\text{avg}}$]{}]{} (-200,49)[(0,0)[$\tilde{s}_{\text{avg}}$]{}]{} (-90,-8.0)[(0,0)[$\frac{t}{\per}$]{}]{}
In the low amplitude regime, one can also estimate in perturbation theory the amount of energy dissipated per cycle $\epsilon_\text{cycle}$ which we define as the difference of the average energy $\epsilon_\text{avg}$ between two successive cycles; for simplicity we take the result for the last two cycles of our evolution in quoting the results below. One expects the relation to take a scaling form $\epsilon_\text{cycle} \sim \omega^\alpha$. The scaling exponent $\alpha$ should be a non-trivial function of frequency itself; for low frequencies it is independent of the driving operator, but the high frequency limit cares about the spectral properties about the operator in question. Specifically, one finds that [@Auzzi:2013pca]: $\epsilon_\text{cycle} \sim \omega$ for small frequencies and $\epsilon_\text{cycle} \sim \omega^{2\,\Delta - d}$ for high frequencies. Since we are not scanning over different choices of the driving operator, we have a single shot at determining this result. As depicted in Fig. \[fig:alpha\_omega\_low\_A\] we indeed find that the energy dissipated is linear both at low and high frequencies: $\alpha(\omega) \to 1$ both for $\omega \gg 1$ and for $\omega \ll 1$ (a coincidence owing to our choice $\Delta =2$ and $d=3$). Interestingly there is some non-trivial intermediate frequency behaviour which appears to amplify the energy dissipated in a single cycle.
(0.3,0.4)(0,0) (-91,-8.0)[(0,0)[$\text{log}_{10} \ \omega$]{}]{} (-207,98)[(0,0)[$\alpha$]{}]{}
The bulk picture of the process is also very simple: as we send energy pulses, which are either weak or infrequent, they interact very rarely before falling into the black hole horizon. All injected energy from the boundary goes towards steadily increasing the black hole mass and the scalar field remains unexcited. The more diverse behaviour observed below can be attributed to gravitational interactions of those energy pulses before they fall into the black hole.
Dynamical Crossover Tilted Regime {#sec:dc}
---------------------------------
We now discuss the qualitative changes in the system as we begin to move from the weak driving $\xi \ll 1$ to the strong driving regime $\xi \gg 1$ (from regime I to regime II through the grey-dashed line in phase diagram Fig. \[fig:PP\_qualitative\]). Fig. \[fig:PP\_wobbly\] depicts a typical phase portrait of the system as we cross into the new dynamical regime. We see that this regime is characterized by an onset of excitations of the scalar field and breaking of discrete time translation symmetry.
[0.45]{} ![The dimensionless phase portrait of the response $\tilde{\phi}_1$ versus the source $\tilde{\phi}_0$ for $\xi(\per = 0.1, \amp = 20)=2$ (left) and $\xi( \per = 10, \amp = 1) = 10$ (right). The conventions are as in Fig. \[fig:PP\_boring\]. The left panel shows the behaviour in phase IIb while the right panel pertains to phase IIa.[]{data-label="fig:PP_wobbly"}](figures/PP_amp=20_per=0.1_tf=1.pdf "fig:"){width="\textwidth"} (-104,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-199,95)[(0,0)[$\tilde{\phi}_1$]{}]{}
[0.45]{} ![The dimensionless phase portrait of the response $\tilde{\phi}_1$ versus the source $\tilde{\phi}_0$ for $\xi(\per = 0.1, \amp = 20)=2$ (left) and $\xi( \per = 10, \amp = 1) = 10$ (right). The conventions are as in Fig. \[fig:PP\_boring\]. The left panel shows the behaviour in phase IIb while the right panel pertains to phase IIa.[]{data-label="fig:PP_wobbly"}](figures/PP_amp=1_per=10_tf=100.pdf "fig:"){width="\textwidth"} (-104,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-199,95)[(0,0)[$\tilde{\phi}_1$]{}]{}
The left panel of Fig. \[fig:PP\_wobbly\] shows the transition from $\xi \ll 1 \rightarrow \xi \gg 1$ at high amplitudes: the trajectories are no longer closed, rather they precess as a function of time and are slightly tilted. The breaking of discrete time-translation invariance is an interesting effect of the gravitational interactions of the scalar field.
In the right panel of Fig. \[fig:PP\_wobbly\] we see the effect of moving into the new dynamical regime at low amplitudes: there is a clear tilt in the phase portrait from the one in Fig. \[fig:PP\_boring\] with $\xi \gg 1$ which indicates that the response is no longer completely out of phase with the source. The tilting of the trajectories at lower frequencies corresponds to the emergence of a finite in-phase contribution $\sigmaIn > 0$ in the conductivity; this sets the system somewhere between one with a purely out-of-phase conductivity (closed circular trajectories) and one with a purely in-phase conductivity (straight diagonal line trajectories). In other words not all of the injected energy is dissipated as was the case in regime I, but rather, work is actually being done on the system.
As a result of having less dissipation in this regime, the energy and entropy of the black hole grow more slowly with time. Moreover, we find the scaling behaviour between the average energy and entropy, with a thermodynamic scaling exponent $\gamma > \tfrac{2}{3}$, for all values of $(\per,\amp)$, as shown in Fig. \[fig:gamma\_phasediagram\]. In other words, while the work done in the system slows down the energy increase of the black hole, the entropy production is affected less.
(-60,13)[(0,0)[$\log_{10} \per$]{}]{} (-235,23)[(0,0)[$\amp$]{}]{} (-265,222)[(0,0)[$\log_{10}(\gamma-\gamma_0)$]{}]{}
To understand this regime further, it is instructive to reproduce this type of phase portrait for a system without gravity. To that effect, we can study the special case of scalar field evolution in a fixed black hole background, with no backreaction on the geometry (i.e., $\alpha_g =0$). To include non-linearity into the problem, we add self-coupling to the scalar field, to mimic the effect of the non-linearities due to gravitational interactions (see also [@Basu:2013vva]). Fig. \[fig:PP\_phi2\] depicts the phase portrait of a self-coupled scalar field with two types of [*polynomial*]{} potentials, which we took to be our original form (free massive scalar) and also one with quartic self-interactions: $$V_{\text{poly},4}(\phi)=-2 \phi^2-\frac{1}{2}\phi^4.
\label{eqn:V4}$$
[0.45]{} ![The phase portrait of the dimensionless response $\tilde{\phi}_1 \equiv \tfrac{\per}{\amp} \,\phi_1$ versus the dimensionless source $\tilde{\phi}_0 \equiv \tfrac{1}{\amp}\, \phi_0$ for $\xi(\per = 10, \amp = 1)=10$ with $\alpha_g=0$ and different polynomial potentials $V(\phi)$. The conventions are as described in Fig. \[fig:PP\_boring\].[]{data-label="fig:PP_phi2"}](figures/PPp1_amp=1_per=10_tf=100.pdf "fig:"){width="\textwidth"} (-104,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-198,93)[(0,0)[$\tilde{\phi}_1$]{}]{}
[0.46]{} ![The phase portrait of the dimensionless response $\tilde{\phi}_1 \equiv \tfrac{\per}{\amp} \,\phi_1$ versus the dimensionless source $\tilde{\phi}_0 \equiv \tfrac{1}{\amp}\, \phi_0$ for $\xi(\per = 10, \amp = 1)=10$ with $\alpha_g=0$ and different polynomial potentials $V(\phi)$. The conventions are as described in Fig. \[fig:PP\_boring\].[]{data-label="fig:PP_phi2"}](figures/PPp2_amp=1_per=10_tf=100.pdf "fig:"){width="\textwidth"} (-104,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-197,92)[(0,0)[$\tilde{\phi}_1$]{}]{}
We can see that without non-linearity as in Fig. \[subfig:PP\_phi2\_xi10\], the phase portrait is tilted, but sharp features of the phase portrait are lost compared to the case with the same driving but also gravitational backreaction, depicted in Fig. \[subfig:PP\_boring\_low\_freq\]. Adding a polynomial non-linearity, as done in Fig. \[subfig:PP\_phi4\_xi10\], gives a phase portrait that starts to form slightly sharper features along with some amplification of the response. Thus, the simple system of self-interacting scalar field allows us sufficiently separate the two effects in regime II: we see that the tilt in the phase diagram is associated with decreased frequency, whereas the breaking of time-translation invariance is associated with increased amplitude. We note also that for this simple system, the third dynamical regime of unbounded amplification discussed in the next subsection seems to be absent.
Thus, the bulk interpretation of this dynamical regime becomes clear: the pulses of energy injected at the boundary interact gravitationally before falling into the black hole. This results in additional physics to that of simple dissipation, modeled here by infalling the black hole. The gravitational interaction is due to perturbative exchange of gravitons, and can be mimicked by a polynomial self-interaction of the scalar field. In the next subsection we will see the effect of the gravitational interactions becoming strong when both $\amp$ and $\per$ are large.
Unbounded Amplification Regime {#sec:unamp}
------------------------------
As we increase the driving strength further in both $\amp$ and $\per$ directions (from regime II to regime III through the blue-dashed line in phase diagram Fig. \[fig:PP\_qualitative\]), we enter a dynamical regime no longer reproducible by polynomial self-interactions of the scalar field. We see the phase portrait of the scalar field in Fig. \[fig:PP\_interesting\] for two instances of parameters in this regime. Moreover, we find this dynamical regime to be characterized by unbounded response and restoration of time translation symmetry.
[0.45]{} ![The phase portrait of the response $\tilde{\phi}_1$ versus the source $\tilde{\phi}_0$ for $\xi=20$ (left) and $\xi= 200$ (right) in the non-perturbative dynamical regime (regime III). The conventions are as in Fig. \[fig:PP\_boring\].[]{data-label="fig:PP_interesting"}](figures/PP_amp=20_per=1_tf=10.pdf "fig:"){width="\textwidth"} (-102,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-195,90)[(0,0)[$\tilde{\phi}_1$]{}]{}
[0.46]{} ![The phase portrait of the response $\tilde{\phi}_1$ versus the source $\tilde{\phi}_0$ for $\xi=20$ (left) and $\xi= 200$ (right) in the non-perturbative dynamical regime (regime III). The conventions are as in Fig. \[fig:PP\_boring\].[]{data-label="fig:PP_interesting"}](figures/PP_amp=20_per=10_tf=100.pdf "fig:"){width="\textwidth"} (-102,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-194,90)[(0,0)[$\tilde{\phi}_1$]{}]{}
As we increase the strength of the driving force $\xi$, the phase portrait becomes sharper and tilted, corresponding to an increased response and, again, less lag with the source as seen in Fig. \[fig:conductivity\_phasediagram\]. The ‘slowness’ of the energy injection from the boundary allows the scalar field to heat up as if the entire process were adiabatic, consequently allowing the scalar response to respond relatively quicker to the source. Note that although Fig. \[fig:conductivity\_phasediagram\] shows $\left| \sigma_{\text{in}} / \sigma \right| \approx 1$ in this regime, the absolute value $\left| \sigma \right|$ is actually very large in this unbounded amplification regime so that a small $\left| \sigmaOut / \sigma \right|$ is still strong enough to keep the black hole perpetually growing in size.
The maximal response $|\phi^{\text{max}}_1|$ over our ten cycles of driving is plotted in Fig. \[fig:maxresponse\_phasediagram\] throughout the phase diagram. It is seen to increase rapidly with $\xi$ past the dissipation-dominated regime. This seems to indicate the presence of a non-linear resonance, which allows the scalar response to grow without bound. An interesting feature of Fig. \[fig:maxresponse\_phasediagram\] is that the maximal response does not grow in the high frequency regime regardless of how large $\xi$ is by increasing $\amp$. It seems unlikely that unbounded behaviour is attainable even for amplitudes drastically higher than the bounds of numerical explorations reported in Fig. \[fig:maxresponse\_phasediagram\]. Physically, this means that a rapid pulsing of small packets of energies can barely amplify the response of the system; the frequency of driving has to be below a certain bound for resonance to be possible – or in other words, a certain slowness in the sourcing is required. We conjecture that one should would see unbounded amplification only in the combined large $\per$, large $\amp$ regime which is slightly different from the traditional definition of resonance that depends only on frequency. An interesting curiousity is a slight dip in the response for moderate values of $\xi$ preceding the rapid growth. This trough appears to demarcate the domains of bounded (regime II) and unbounded responses (regime III) empirically. It would be interesting to come up with a explanation for this phenomenon.
(-60,13)[(0,0)[$\log_{10} \per$]{}]{} (-235,23)[(0,0)[$\amp$]{}]{} (-270,224)[(0,0)[$\log_{10} \left| \tilde{\phi}^{\text{max}}_1 \right| $]{}]{}
Finally, it is amusing to model the non-linear effects of gravity in terms of an effective scalar potential to see what is necessary to attain regime III. We find that while a scalar field with polynomial self-interaction does not seem to posses this regime, one can reproduce similar features by [*non-polynomial*]{} potentials. For example, we can discuss a self-interacting scalar field probe, with $$V_{\text{non-poly}}(\phi)=-2 \sinh^2\phi+\frac{1}{6}\sinh^4\phi\,.
\label{eqn:V_nonpoly}$$ This choice of scalar self-interaction is chosen to agree with our previous example in the small field regime, but of course behaves differently for large field values. In Fig. \[fig:PP\_nonpoly\] we see that indeed similar features of the phase diagram are reproduced: narrow closed trajectories and resonant response. We conclude therefore that the features of this dynamical regime are due to strong, non-perturbative gravitational effects occurring outside the black hole horizon. The fact that the non-linearities induced by gravity can be extremely strong, should perhaps be borne in mind while attempting to come up with simplified models of gravitational dynamics in AdS spacetime.
[0.45]{} ![The phase portrait of the response $\tilde{\phi}_1$ versus the source $\tilde{\phi}_0$ for $\xi= 15$ (left) and $\xi= 20$ (right) for the non-polynomial potential Eq. , in the conventions of Fig. \[fig:PP\_boring\]. []{data-label="fig:PP_nonpoly"}](figures/PPnp1_amp=1.5_per=10_tf=100.pdf "fig:"){width="\textwidth"} (-102,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-195,91)[(0,0)[$\tilde{\phi}_1$]{}]{}
[0.45]{} ![The phase portrait of the response $\tilde{\phi}_1$ versus the source $\tilde{\phi}_0$ for $\xi= 15$ (left) and $\xi= 20$ (right) for the non-polynomial potential Eq. , in the conventions of Fig. \[fig:PP\_boring\]. []{data-label="fig:PP_nonpoly"}](figures/PPnp2_amp=2_per=10_tf=100.pdf "fig:"){width="\textwidth"} (-102,-8.0)[(0,0)[$\tilde{\phi}_0$]{}]{} (-195,91)[(0,0)[$\tilde{\phi}_1$]{}]{}
Energy Fluctuations
-------------------
Another observable we monitor is the behaviour of energy fluctuations. More precisely, we consider the deviations from the average energy in a each cycle, $\epsilon_\text{fluc}(t) = |\epsilon(t) - \epsilon_\text{avg}(t)|$. These cycle fluctuations are a crude proxy for genuine fluctuation information that can be extracted, for instance, by considering symmetrized two-point functions of the boundary energy momentum tensor. Such ensemble-averaged fluctuations are known to exhibit phase transitions in periodically driven systems [@nature]. Some indication those transitions are possible in holographic systems is given in [@Auzzi:2013pca].
The results for our simulations in various regimes are plotted in Fig. \[fig:energyfluc\_phasediagram\]. We observe a qualitative change in these cycle fluctuations between different regimes. While in the dissipation-dominated phase we do not see a lot of deviation from the mean, there is a steep growth in fluctuations as we enter the non-linear phases. The fluctuations are maximal in the unbounded amplification regime (regime III). We note that in contrast to the maximal scalar response, which also grows dramatically in that phase, the fluctuations do track the driving frequency, with there being more deviations in the large period limit.
It would be useful to confirm this behaviour directly with the computation of correlation functions, a task we leave for future investigation.
(-60,13)[(0,0)[$\log_{10} \per$]{}]{} (-235,23)[(0,0)[$\amp$]{}]{} (-260,217)[(0,0)[$\tilde{\epsilon}_{\text{fluc}}$]{}]{}
Entanglement entropy {#sec:ee}
====================
Thus far we have discussed various local observables (response functions and thermodynamic data) which have served to help us chart the phase diagram of the driven system in Fig. \[fig:PP\_qualitative\]. We now turn to other non-local field theory observables that are sensitive to the non-equilibrium dynamics. Since we are not going to examine the behaviour of higher point correlation functions, we will dive right into the dynamical behaviour of entanglement entropy.
In the boundary we have a density matrix $\rho(t)$ which is time-evolving with respect to the perturbed Hamiltonian. At any given instant of (boundary) time, we pick a spatial region ${\cal A}$ and construct the matrix elements of the reduced density matrix $\rho_{\cal A} (t) = {\rm Tr}_{{\cal A}^c}\left(\rho(t)\right)$ by tracing out the degrees of freedom in the complement (on the chosen Cauchy slice). The entanglement entropy is given by the von Neumann entropy of $\rho_{\cal A}$, i.e., $S_{\cal A}(t) = -{\rm Tr}_{\cal A} \left(\rho_{\cal A}\, \log \rho_{\cal A}\right)$ which we can monitor as a function of time.
Holographically computing the entanglement entropy for boundary regions in time dependent situations involves finding bulk codimension-2 extremal surfaces $\cal E_{\cal A}$ anchored on the said boundary region ${\cal A}$ [@Hubeny:2007xt]. We study the evolution of entanglement entropy focusing in particular on translationally invariant strip regions: $${\cal A} = \{ t = t_{\cal A}, -a \leq x \leq a , y \in {\mathbb R} \}\,.
\label{eqn:spatial_strip_defn}$$ The bulk codimension-2 surface ends at $x=\pm a$ at some chosen instance of boundary time $t_{\cal A}$ and is obtained by solving effectively a set of geodesic-like equations with our interpolated metric functions $\Sigma$, $f$, and $\chi$ (see Appendix \[sec:extrdet\] for details). The covariant holographic entanglement entropy prescription [@Hubeny:2007xt] generalizing [@Ryu:2006bv; @Ryu:2006ef] states that $$S_{\cal A} = \frac{ \text{Area}({\cal E}_{\cal A} ) }{ 4 \, \GN }\,.
\label{eqn:RT_prescription}$$ Should there be multiple extremal surfaces, we choose the one with minimal area (homologous to ${\cal A}$). The proper area of these surfaces diverges owing to the locality of the underlying QFT. In our case we encounter potential divergences not only from the surface reaching out to the asymptotic boundary, but also from the presence of the sources driving the system. The physical result we are after is the finite universal contribution $S^{\text{fin}}_{\cal A}$, which will measure the entanglement created/destroyed as we drive the system away from thermal equilibrium. Fortuitously, for our choice of scalar operator, there are no contributions due to the source, and hence we can simply regulate by background subtraction.[^12] As a result we will consider as our entanglement diagnostic, the following finite quantity $$\sreg(t)= \frac{4 \, \GN}{L_y} \, \big[ S_{\cal A}(t) - S_{\cal A}(t=0) \big]
\label{eqn:S_reg}$$ where $L_y$ is the IR regulator in the non-compact translationally invariant direction. Since we drive the system away from thermal equilibrium, $S_{\cal A}(t=0)$ is the corresponding value of the entanglement entropy computed in the Schwarzschild- geometry. In what follows we will simply quote the results of our numerical simulations both for the behaviour of the extremal surfaces themselves and $\sreg(t)$.
Extremal surfaces in the driven geometries {#sec:extrsufaces}
------------------------------------------
The extent to which the extremal surfaces penetrate into the bulk can for the most part be determined from the location of the cap-off point which we parameterize as $(t^*, u^*=1/r^*,x=0)$.[^13] For very small regions we are reasonably close to the AdS boundary whence, the curves are approximately semi-circles $u^2 + x^2 \approx a^2$. As we increase to larger strip widths the extremal surfaces start to probe the interesting regions of the driven geometry and thus allows us to see qualitative differences between the four phases.
Generically we see that the following statements hold irrespective of the phases we consider:
1. The radial depth and the temporal extent spanned by the surface evolves non-trivially as a function of $t_{\cal A}$. One consequence of working with ingoing coordinates is that the surfaces naturally dip back in time (see [@Hubeny:2013hz; @Hubeny:2013dea]).
2. The oscillatory driving of the system imprints itself in the profile of the extremal surfaces, with the scale of these oscillations set by the the driving parameters $\amp$ and $\per$. The periodic movement of the surface can be seen in pulsations of the turnaround point of the surface: $u^*$ and $t^*$ have oscillations of the same period superposed over some enveloping function.
3. On average, the extremal surfaces reach further into the bulk with time; $u^*(t_{\cal A})$ is monotonically increasing for the range of parameters explored. To understand this note, we gauge fixed the bulk coordinate chart such that the horizon is at $u_+=1$. In these coordinates the proper size of the region ${\cal A}$ increases (due to $\gxx(t,r)$) which means that the surfaces want to get closer to the horizon to extremize the area functional. The rate at which this happens depends on both the amplitude and the frequency of the driving. We also note that surfaces dip less temporally, i.e., $t^* - t_{\cal A}$ is increasing.
4. We also note that the location of the extremal surface appears to be consistent with causality of entanglement entropy [@Headrick:2014cta]. While we have not explicitly checked that the surface lies in the casual shadow of the boundary region ${\cal A}$, one simple consistency check visible from our results for $t^*$ is that $t^*< t_{\cal A}-a$. We remind the reader that in lines of constant $t$ and $x$ are radially ingoing null geodesics. Causality at the very least requires that the cap-off point of the extremal surface lies below the ingoing null geodesic from the domain of dependence. Since for the strip region the boundary domain of dependence is a diamond anchored at $(t_{\cal A}\pm a,0)$ and $(0,\pm a)$, we note that the ingoing light ray from the bottom tip of this diamond cannot signal to the cap-off point.
In the following discussion we will illustrate the behaviour of the extremal surfaces more explicitly in each of our phases. We have been reasonably conservative in our analysis and have chosen to work only with surfaces that do not get too close to the horizon (in fact $u^* <0.2$). This is to avoid both numerical issues as well as to avoid complications from the existence of multiple extremal surfaces. We follow a single branch of solutions as described at the end of Appendix \[sec:extrdet\]. The primary results of the extremal surfaces are shown in the plots Figs. \[fig:ES\_evolution\_per01\_amp1\], \[fig:ES\_evolution\_per10\_amp1\], \[fig:ES\_evolution\_per10\_amp20\], and \[fig:ES\_evolution\_per01\_amp20\], where we show the evolution of the extremal surface as well as $u^*(t_{\cal A})$ and $t^*(t_{\cal A})$.
![Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=0.1,\amp=1)=0.1$ (phase I; dissipation-dominated). We pick a UV cutoff $u_{\cal A}=10^{-3}$ and have defined $\tilde{t}^* \equiv (t^*-t_{\cal A})/\per$ to measure the cap-off $t^*$ point relative to the boundary.[]{data-label="fig:ES_evolution_per01_amp1"}](figures/ES_evo_a=0.05_amp=1_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-11,19)[(0,0)[$\frac{x}{a}$]{}]{} (-140,-10)[(0,0)[$u$]{}]{} (-290,100)[(0,0)[$\frac{t}{\per}$]{}]{}
![Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=0.1,\amp=1)=0.1$ (phase I; dissipation-dominated). We pick a UV cutoff $u_{\cal A}=10^{-3}$ and have defined $\tilde{t}^* \equiv (t^*-t_{\cal A})/\per$ to measure the cap-off $t^*$ point relative to the boundary.[]{data-label="fig:ES_evolution_per01_amp1"}](figures/ustar_evo_a=0.05_amp=1_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-105,105)[(0,0)[$u^*$]{}]{} ![Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=0.1,\amp=1)=0.1$ (phase I; dissipation-dominated). We pick a UV cutoff $u_{\cal A}=10^{-3}$ and have defined $\tilde{t}^* \equiv (t^*-t_{\cal A})/\per$ to measure the cap-off $t^*$ point relative to the boundary.[]{data-label="fig:ES_evolution_per01_amp1"}](figures/tstar_evo_a=0.05_amp=1_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-103,99)[(0,0)[$\tilde{t}^*$]{}]{} (-56,-8.0)[(0,0)[$t_{\cal A}/\per$]{}]{}
![Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=10,\amp=1)=10$ (phase II; tilted). Conventions described in Fig. \[fig:ES\_evolution\_per01\_amp1\] apply.[]{data-label="fig:ES_evolution_per10_amp1"}](figures/ES_evo_a=0.05_amp=1_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-11,19)[(0,0)[$\frac{x}{a}$]{}]{} (-140,-10)[(0,0)[$u$]{}]{} (-290,100)[(0,0)[$\frac{t}{\per}$]{}]{}
![Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=10,\amp=1)=10$ (phase II; tilted). Conventions described in Fig. \[fig:ES\_evolution\_per01\_amp1\] apply.[]{data-label="fig:ES_evolution_per10_amp1"}](figures/ustar_evo_a=0.05_amp=1_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-105,105)[(0,0)[$u^*$]{}]{} ![Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=10,\amp=1)=10$ (phase II; tilted). Conventions described in Fig. \[fig:ES\_evolution\_per01\_amp1\] apply.[]{data-label="fig:ES_evolution_per10_amp1"}](figures/tstar_evo_a=0.05_amp=1_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-103,99)[(0,0)[$\tilde{t}^*$]{}]{} (-116,114)[(0,0)[$\times 10^{-3}$]{}]{} (-56,-8.0)[(0,0)[$t_{\cal A}/\per$]{}]{}
#### Linear regime (small $\amp$):
Although all phases display extremal surfaces that sink into the bulk with each driving cycle, the growth of $u^*$ in the linear regime of small amplitudes is most steady. We focus here on phases I (high frequency; dissipation-dominated) illustrated in Fig. \[fig:ES\_evolution\_per01\_amp1\] and IIa (low frequency; tilted) illustrated in Fig. \[fig:ES\_evolution\_per10\_amp1\], which fall under this characterization. As the frequency is lowered and we pass from the dissipation-dominated phase to the tilted phase, there is drastic reduction in the growth of $u^*$ per cycle.
The evolution of $t^*$ in the two phases is also interesting; $t^*-t_{\cal A}$ is gradually increasing on average with time (recall that in the stationary geometry $t^*-t_{\cal A}$ would be constant). It turns out to be useful to look at a dimensionless parameter $\tilde{t}^* \equiv (t^*-t_{\cal A})/\per$ which measures the cap-off time relative to the boundary. In this context, there is more time-lag in phase I i.e., $\tilde{t}^*_{\text{\tiny{I}}} \ll \tilde{t}^*_{\text{\tiny{IIa}}} \lesssim 0$, which hints at the cause for why the surfaces do not penetrate as far deep in the bulk in phase IIa as opposed to phase I.[^14] In addition we see strong oscillatory patterns in phase II in spite of having only a steady increase in $u^*$; such a feature is absent in phase I.
![ Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=10,\amp=20)=200$ (phase III; unbounded amplification). Conventions described in Fig. \[fig:ES\_evolution\_per01\_amp1\] apply.[]{data-label="fig:ES_evolution_per10_amp20"}](figures/ES_evo_a=0.05_amp=20_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-11,19)[(0,0)[$\frac{x}{a}$]{}]{} (-140,-10)[(0,0)[$u$]{}]{} (-290,100)[(0,0)[$\frac{t}{\per}$]{}]{}
![ Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=10,\amp=20)=200$ (phase III; unbounded amplification). Conventions described in Fig. \[fig:ES\_evolution\_per01\_amp1\] apply.[]{data-label="fig:ES_evolution_per10_amp20"}](figures/ustar_evo_a=0.05_amp=20_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-105,105)[(0,0)[$u^*$]{}]{} ![ Evolution of the extremal surfaces for a strip of width $a=0.05$ with driving parameters $\xi(\per=10,\amp=20)=200$ (phase III; unbounded amplification). Conventions described in Fig. \[fig:ES\_evolution\_per01\_amp1\] apply.[]{data-label="fig:ES_evolution_per10_amp20"}](figures/tstar_evo_a=0.05_amp=20_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-103,99)[(0,0)[$\tilde{t}^*$]{}]{} (-110,114)[(0,0)[$\times 10^{-3}$]{}]{} (-56,-8.0)[(0,0)[$t_{\cal A}/\per$]{}]{}
![Evolution of the extremal surfaces for a strip of width $a=0.01$ with driving parameters $\xi(\per=0.1,\amp=20)=2$ (phase IIb; dynamical crossover). Conventions are as described in Fig. \[fig:ES\_evolution\_per01\_amp1\].[]{data-label="fig:ES_evolution_per01_amp20"}](figures/ES_evo_a=0.01_amp=20_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-11,19)[(0,0)[$\frac{x}{a}$]{}]{} (-140,-10)[(0,0)[$u$]{}]{} (-290,100)[(0,0)[$\frac{t}{\per}$]{}]{}
![Evolution of the extremal surfaces for a strip of width $a=0.01$ with driving parameters $\xi(\per=0.1,\amp=20)=2$ (phase IIb; dynamical crossover). Conventions are as described in Fig. \[fig:ES\_evolution\_per01\_amp1\].[]{data-label="fig:ES_evolution_per01_amp20"}](figures/ustar_evo_a=0.01_amp=20_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-105,105)[(0,0)[$u^*$]{}]{} ![Evolution of the extremal surfaces for a strip of width $a=0.01$ with driving parameters $\xi(\per=0.1,\amp=20)=2$ (phase IIb; dynamical crossover). Conventions are as described in Fig. \[fig:ES\_evolution\_per01\_amp1\].[]{data-label="fig:ES_evolution_per01_amp20"}](figures/tstar_evo_a=0.01_amp=20_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-103,99)[(0,0)[$\tilde{t}^*$]{}]{} (-56,-8.0)[(0,0)[$t_{\cal A}/\per$]{}]{}
#### Non-linear regime (large $\amp$):
We now turn to the phases III (low frequency; unbounded amplification) illustrated in Fig. \[fig:ES\_evolution\_per10\_amp20\] and IIb (high frequency; wobbly) illustrated in Fig. \[fig:ES\_evolution\_per01\_amp20\] in the non-linear regime of high amplitude. Some of the features seen in the linear regime continue to pertain: we see more pronounced oscillations in $\tilde{t}^*$ and a decreased tendency for the surfaces to lag behind in time at lower frequencies.
In the unbounded amplification regime (phase III), we see significant bursts of growth of the extremal surfaces. The oscillatory driving is felt rather acutely by the surfaces and the evolution is considerably violent. On average however, $u^*$ appears to advance more serenely despite having large amplitude oscillations per cycle.
In the dynamical crossover wobbly regime (phase IIb), there is a considerable amount of instability. We chose here to work with smaller strip widths $a=0.01$ (instead of $a=0.05$) to avoid complications of phase transitions between multiple competing extremal surfaces. The early part of the evolution is in line with what happens in the dissipation-dominated regime (phase I), but shortly after, there are discontinuities in the $\tilde{t}^*$ parameter with no noticeable effect in $u^*$. Around $t_{\cal A}/P \approx 4.0 - 4.2$ and $t_{\cal A}/P \approx 4.6$, we see an exchange of dominance in the extremal surface, which starts out at a higher value of $\tilde{t}^*$.
All in all, the extremal surfaces in the non-linear regime definitely has elements of intrigue owing to the large pulses of energy that affect the bulk geometry significantly. Although we do not delve into extremal surfaces that are positioned deeper into the bulk, we notice in the course of our analysis that the surfaces tend towards the horizon as expected. More curiously, we also find that for larger regions we cannot find extremal surfaces that stay outside the apparent horizon. This is not surprising since we expect based on earlier results that there will be surfaces that penetrate the apparent horizon of the black hole (cf., [@AbajoArrastia:2010yt]). However, one of the disadvantages of our numerical scheme is that we are unable to explore this interesting regime due to the fact that the spacetime inside the apparent horizon has been excised. As explained in [@Chesler:2013lia], this was to avoid complications with having caustics in the coordinate chart. Analysis of entanglement entropy however does require us to have the complete bulk geometry.
The evolution of entanglement {#sec:eegrow}
-----------------------------
We now turn to the evolution of the entanglement entropy; the results are presented in Fig \[fig:EE\_unreg\] for the regulated quantity $\sreg$ as introduced in .
In the dissipation-dominated regime (phase I), the entanglement entropy gradually increases, though in each cycle of forcing there is a time period for which the growth is negligible. We expect this feature is simply a consequence of the entanglement entropy tracking the thermal entropy. Even though we are not quite probing the full thermal contribution with the relatively small regions ${\cal A}$, it bears to reason that the variation of the geometry is more or less equitable on all radial scales. This appears consistent with other probes of this phase. As we discussed in §\[sec:dissdom\] the weak driving allows the system to efficiently dissipate the energy induced by the source and the conductivity $\sigma(t)$ was purely imaginary. Basically the dominant effect here is the growth of the black hole horizon due to the driving and this in turn imprints itself into the growth of $\sreg$ seen in Fig. \[subfig:EE\_reg\_phase1\].
[0.42]{} ![The evolution of the regularized entanglement entropy, $\sreg$ defined in Eq. , for the four phases for a radial cutoff of $u_{\cal A} = 10^{-3}$. The strip widths are $a=0.05$ for panels (a), (b), (c), and $a=0.01$ for panel (d).[]{data-label="fig:EE_unreg"}](figures/EE_reg_evo_a=0.05_amp=1_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-82,-7)[(0,0)[$t_{\cal A}/\per$]{}]{} (-191,92)[(0,0)]{}
[0.438]{} ![The evolution of the regularized entanglement entropy, $\sreg$ defined in Eq. , for the four phases for a radial cutoff of $u_{\cal A} = 10^{-3}$. The strip widths are $a=0.05$ for panels (a), (b), (c), and $a=0.01$ for panel (d).[]{data-label="fig:EE_unreg"}](figures/EE_reg_evo_a=0.05_amp=1_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-82,-7)[(0,0)[$t_{\cal A}/\per$]{}]{} (-198,92)[(0,0)]{}
[0.42]{} ![The evolution of the regularized entanglement entropy, $\sreg$ defined in Eq. , for the four phases for a radial cutoff of $u_{\cal A} = 10^{-3}$. The strip widths are $a=0.05$ for panels (a), (b), (c), and $a=0.01$ for panel (d).[]{data-label="fig:EE_unreg"}](figures/EE_reg_evo_a=0.05_amp=20_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-82,-7)[(0,0)[$t_{\cal A}/\per$]{}]{} (-191,92)[(0,0)]{}
[0.42]{} ![The evolution of the regularized entanglement entropy, $\sreg$ defined in Eq. , for the four phases for a radial cutoff of $u_{\cal A} = 10^{-3}$. The strip widths are $a=0.05$ for panels (a), (b), (c), and $a=0.01$ for panel (d).[]{data-label="fig:EE_unreg"}](figures/EE_reg_evo_a=0.01_amp=20_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-82,-7)[(0,0)[$t_{\cal A}/\per$]{}]{} (-191,92)[(0,0)]{}
[0.42]{} ![The evolution of the regularized entanglement entropy, $\sreg$ defined in Eq. , against the normalized entropy of the black hole, $s/s_0 = s/s(t=0)$, for the four phases for a radial cutoff of $u_{\cal A} = 10^{-3}$. The strip widths are $a=0.05$ for panels (a), (b), (c), and $a=0.01$ for panel (d). We include the Spearman and Pearson rank coefficients, $-1 \leq \rho_s \leq 1$ and $-1\leq \rho_p \leq 1$ respectively, for each plot to demonstrate the linearity of the correlation between the entanglement entropy and the thermal entropy (see text for explanation).[]{data-label="fig:EE_reg_S"}](figures/EE_reg_S_evo_a=0.05_amp=1_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-82,-7)[(0,0)[$s/s_0$]{}]{} (-191,92)[(0,0)]{}
[0.438]{} ![The evolution of the regularized entanglement entropy, $\sreg$ defined in Eq. , against the normalized entropy of the black hole, $s/s_0 = s/s(t=0)$, for the four phases for a radial cutoff of $u_{\cal A} = 10^{-3}$. The strip widths are $a=0.05$ for panels (a), (b), (c), and $a=0.01$ for panel (d). We include the Spearman and Pearson rank coefficients, $-1 \leq \rho_s \leq 1$ and $-1\leq \rho_p \leq 1$ respectively, for each plot to demonstrate the linearity of the correlation between the entanglement entropy and the thermal entropy (see text for explanation).[]{data-label="fig:EE_reg_S"}](figures/EE_reg_S_evo_a=0.05_amp=1_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-82,-7)[(0,0)[$s/s_0$]{}]{} (-198,92)[(0,0)]{}
[0.42]{} ![The evolution of the regularized entanglement entropy, $\sreg$ defined in Eq. , against the normalized entropy of the black hole, $s/s_0 = s/s(t=0)$, for the four phases for a radial cutoff of $u_{\cal A} = 10^{-3}$. The strip widths are $a=0.05$ for panels (a), (b), (c), and $a=0.01$ for panel (d). We include the Spearman and Pearson rank coefficients, $-1 \leq \rho_s \leq 1$ and $-1\leq \rho_p \leq 1$ respectively, for each plot to demonstrate the linearity of the correlation between the entanglement entropy and the thermal entropy (see text for explanation).[]{data-label="fig:EE_reg_S"}](figures/EE_reg_S_evo_a=0.05_amp=20_per=10_tfinal=100_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-82,-7)[(0,0)[$s/s_0$]{}]{} (-191,92)[(0,0)]{}
[0.42]{} ![The evolution of the regularized entanglement entropy, $\sreg$ defined in Eq. , against the normalized entropy of the black hole, $s/s_0 = s/s(t=0)$, for the four phases for a radial cutoff of $u_{\cal A} = 10^{-3}$. The strip widths are $a=0.05$ for panels (a), (b), (c), and $a=0.01$ for panel (d). We include the Spearman and Pearson rank coefficients, $-1 \leq \rho_s \leq 1$ and $-1\leq \rho_p \leq 1$ respectively, for each plot to demonstrate the linearity of the correlation between the entanglement entropy and the thermal entropy (see text for explanation).[]{data-label="fig:EE_reg_S"}](figures/EE_reg_S_evo_a=0.01_amp=20_per=0.1_tfinal=1_cutoff=0.001.pdf "fig:"){width="\textwidth"} (-82,-7)[(0,0)[$s/s_0$]{}]{} (-191,92)[(0,0)]{}
On the other hand when we reach phase IIa (tilted regime) by way of small amplitudes, we start to see definite oscillatory evolution of $\sreg$. In each oscillatory period we see a local reduction in $\sreg$. On the other hand the temporal radial depth attained by the extremal surface as measured by $u^*$ is almost similar to that in phase I by juxtaposing the behaviour in Fig. \[fig:ES\_evolution\_per01\_amp1\] and Fig. \[fig:ES\_evolution\_per10\_amp1\]. In phase IIa however, our extremal surfaces are closer to the boundary in contrast to phase I. We conjecture that the origin of the reduction in the $\sreg$ is associated with the sharp oscillations in $t^*$ or equivalently $\tilde{t}^*$. These imprint themselves into the actual value of the area despite the surface not getting too far into the bulk (which is possible since even the asymptotics of the geometry is sensitive to the driving, cf., ). The onset of non-monotone growth of $\sreg$ in Fig. \[subfig:EE\_reg\_phase2\] characterizes the departure from the linear regime to the non-linear domain in line with the behavior of the phase portrait which in turns modifies the conductivity (which picks up a real part $\sigmaIn > 0$ in phase IIa).
The temporal change of $\sreg$ is much more pronounced in the non-linear regime. In the unbounded amplification phase III (see Fig. \[subfig:EE\_reg\_phase3\]) and the dynamical crossover wobbly phase IIb (see Fig. \[subfig:EE\_reg\_phase4\]), the $\sreg$ appears to track the time-coordinate of the cap-off point $\tilde{t}^*$ quite efficiently. Indeed here we expect the non-linearities of the system to be the dominant effect. We know that the black hole grows quite rapidly in response to the energy injected into the system at the boundary from our discussion in §\[sec:dc\] and §\[sec:unamp\]. The behaviour in phase III is smooth with large amplitude oscillations, which qualitatively track quite well the behaviour of $\tilde{t}^*$. The dynamical crossover wobbly phase (phase IIb) exhibits a lot more drastic behaviour. We encounter for the first time a jumps in the family of extremal surface that minimize the area (satisfying the boundary conditions and the homology constraint). These jumps translate into continuous but non-differentiable kinks in $\sreg$ visible in Fig. \[subfig:EE\_reg\_phase4\]. We again note that the radial position of the cap-off point of the extremal surface behaves much more smoothly and the glitches appear in $\tilde{t}^*$. Furthermore, the growth of the entanglement itself is rather steep as we see about an order of magnitude difference in $\sreg$ between the low amplitude and high amplitude regimes.
It is interesting to contrast the change of entanglement entropy with the change in the thermal entropy to see how the two are correlated. As we have argued above, the fact that we have an ever increasing thermal entropy (the bulk black hole is constantly growing) implies that even for small sub-systems we will quickly see overwhelming thermal contribution. We display in Fig. \[fig:EE\_reg\_S\] the functional dependence of $\sreg$ on the (normalized) instantaneous thermal entropy $s(t)/s(t=0)$. It is immediately apparently by eyeballing the plots that there appears to be near-perfect correlation in three phases with Fig. \[subfig:EE\_regS\_phase3\] corresponding to phase III being the only outlier. To get a quantitative feeling for the correlation we have also indicated the Pearson correlation coefficient $\rho_p$ as well as the Spearman rank coefficient $\rho_s$. These are statistical markers for measuring correlations between two sets of data and are defined to take values in the interval $[-1,1]$. The values $\rho_s, \rho_p = 0, \pm 1$ signify zero, perfect positive and perfect negative correlation respectively. While the Spearman coefficient indicates that the observables in question are monotonically related, the Pearson coefficient provides an accurate measure of linear correlation. Indeed from the results quoted in Fig. \[fig:EE\_reg\_S\] we see that $\sreg(s)$ is a linear function to a very good approximation in phases I, IIa and IIb. It is curious that the linearity is respected even in the presence of the glitches in the growth of entanglement entropy (we do not see any drastic behaviour in the area of the apparent horizon). The unbounded amplification phase III clearly demonstrates the effects of non-linearities by decorrelating $\sreg$ and $s(t)$.
Discussion {#sec:discuss}
==========
The non-equilibrium dynamics of strongly coupled field theories is amenable to detailed quantitative exploration using the AdS/CFT correspondence. We have exploited this set-up to study the behaviour when a homogeneous thermal plasma is driven away from equilibrium by a periodically sourcing a relevant (composite) scalar operator. The resulting dynamics exhibits a rather rich phase structure illustrated in Fig. \[fig:PP\_qualitative\].
We identified four distinct phases, characterizing them in terms of the frequency and amplitude of the external driving force. Of these the dissipation dominated phase I is perhaps most intuitive for here the weakness of the driving, allows the system to to catch up with the driving. This is clearly visible in the various observables we studied; the complex conductivity of the response is purely real owing to the phase lag between the source and response and the evolution of entanglement is pretty quiescent.
There is more structure when we ramp up either the period of driving, or the amplitude, for now the system departs quite rapidly away from equilibrium. The response therefore is more pronounced; we see more in phase response and greater temporal oscillations. In phases IIa to IIb there emerges a non-vanishing imaginary part to the conductivity, which in fact appears to capture the entire response for high values of the period and amplitude. We also notice that there are significant fluctuations in the energy density and the entanglement entropy and furthermore, the entropy density grows rather rapidly in this regime. Perhaps most intriguing is the unbounded amplification of phase III, where we see sharp fluctuations and a highly non-linear response. We argue that this response appears to be not captured by polynomial self-interactions of the composite operator; the intricate dynamics of gravity in AdS appears to induce effective non-polynomial couplings in the effective action for the operator ${\cal O}$ we use to perturb the system away from equilibrium. We believe this fact is significant and should be taken into account when attempting to construct effective models distilling the effects of gravitational interactions for strongly coupled systems .
While our focus has been on computing the simplest set of observables, essentially one-point functions and entanglement entropy for small sub-systems, the power of holography is that we can do much more. In time independent equilibrium scenarios it is straightforward to use the holographic map to compute correlation functions (at least two point functions). In the genuine non-equilibrium scenarios as those we have focused on the technology for computing such observables, whilst present [@CaronHuot:2011dr; @CaronHuot:2011dr] is still a bit cumbersome to work with (at least numerically). It would be interesting to develop these techniques further perhaps taking inspiration from the analytical models of [@Ebrahim:2010ra; @Keranen:2014lna]. This would allow us with a direct probe of fluctuations in the plasma, which can be contrasted with the dissipation in the system, the latter being measured by the entropy production through the growth of the horizon.
Likewise our exploration of the behaviour of entanglement entropy has been restricted to analysis of small sub-systems for pragmatic reasons. While the sub-system under consideration was chosen to have fixed size, the fact that we are continuously driving the system leads to an ever increasing thermal contribution to the entanglement. Geometrically this is easy to understand since the horizon for our bulk solution is ever growing (as we have indicated that both the event and apparent horizons are required to be monotonic in our set-up) and reaches out towards the boundary in the course of the evolution. As a result, the local thermal scale can overwhelm the relative smallness of the sub-region we choose. To have precise mapping of the entanglement structure we need to be able to ascertain the true minimum of the area functional in such scenarios bearing in mind that the extremal surface can (and often does) penetrate various horizons. A significant obstacle in ascertaining this is the fact that the characteristic method for solving Einstein’s equations developed in [@Chesler:2013lia] excises the region of the spacetime behind the apparent horizon. While this is a technical obstacle, overcoming it would not only enable us to probe the interior of a highly non-equilibrium black hole using holographic entanglement, but it could also allow us to explore other interesting scenarios such as the effect of perturbing the ground state of the system by external sources.
We would like to thank Veronika Hubeny and Henry Maxfield for useful discussions. M. Rangamani and M. Rozali would like to acknowledge the hospitality of Yukawa Institute for Theoretical Physics, Kyoto during the course of the project. In addition M. Rangamani would also like to acknowledge the hospitality of IAS, Princeton, University of Amsterdam and Aspen Center for Physics.
M. Rangamani acknowledges support from the Ambrose Monell foundation, by the National Science Foundation under Grant 1066293, by the FQXi under grant “Measures of Holographic Information" (FQXi-RFP3-1334), by the STFC Consolidated Grant ST/L000407/1, and the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013), ERC Consolidator Grant Agreement ERC-2013-CoG-615443: SPiN (Symmetry Principles in Nature). M. Rozali and A. Wong are supported by NSERC.
Holographic Renormalization {#sec:holren}
===========================
We collect here some salient results for the computation of physical field theory quantities using standard holographic techniques.
Scalar deformations {#sec:}
-------------------
The bulk action should be supplemented by boundary counter-terms to ensure that (a) the bulk equations of motion follow from a consistent variational principle and (b) the on-shell action evaluated on the solutions is finite.
In standard Poincaré- $$ds^2 = r^2\, \eta_{\mu\nu}\, dx^\mu\,dx^\nu + \frac{dr^2}{r^2} \equiv \frac{\eta_{\mu\nu}\, dx^\mu \, dx^\nu + dz^2}{z^2}
\label{}$$ the scalar field behaves asymptotically as $$\begin{aligned}
\phi(r,x) &\to \frac{1}{r^{d-\Delta}} \, \phi_0 + \frac{1}{r^\Delta}\, \phi_1
\nonumber \\
\phi(z,x) &\to z^{d-\Delta} \, \phi_0 + z^\Delta \, \phi_1
\label{}\end{aligned}$$ We will work with standard quantization (Dirichlet boundary conditions) for the scalar field, which involves treating the mode that fall-off as $r^{\Delta-d}$ as the source for the scalar field.
In the presence of the source we let the metric to take the FG form, $$ds^2 = \frac{dz^2}{z^2} + \frac{g_{\mu\nu}(x,z) \,dx^\mu\, dx^\nu}{z^2}
\label{}$$ where $g_{\mu\nu}(z,x) = \gamma_{\mu\nu} + {\cal O}(z)$. If necessary we will denote by $\gamma_{\epsilon}$ the induced metric on the surface $z = z_\epsilon$ which differs from the boundary metric by a conformal transformation by $z_\epsilon^2$. We will ignore this issue for most part and write the counter-terms in terms of $\gamma_{\mu\nu}$ below for simplicity.
With these conventions we find the following boundary counter-terms: $$\begin{aligned}
S_{bdy} &= \frac{1}{16\pi\, G_N}\; \int d^d x \; \sqrt{-\gamma} \, \left( 2 K - 2\, (d-1) -\frac{1}{d-2}\, ^\gamma R
\right.
\nonumber \\
& \left. \qquad \qquad \qquad
-\frac{1}{2}\, \Delta_- \, \phi^2 + \frac{1}{2\,(2\,\Delta-d-2)}\;\left[ (\partial \phi)^2 + c_1 \; ^\gamma R\; \phi^2 \right]\right)
\label{}\end{aligned}$$ We are using conventional AdS/CFT definitions: $$\Delta_\pm = \frac{d}{2} \pm \sqrt{\frac{d^2}{4} + m^2} = \frac{d}{2} \pm \nu
\label{}$$ Our interest concerns conformally coupled scalar field which has a mass in AdS units given by $$m_c^2 = -\frac{d^2-1}{4} \qquad \Longrightarrow \qquad \Delta_\pm = \frac{d\pm1}{2}
\label{}$$ To compute the boundary energy momentum tensor we vary $$T^{\mu\nu} = \frac{2}{\sqrt{-\gamma}} \, \frac{(\delta S_{bulk} + S_{bdy})}{\delta\gamma_{\mu\nu}}
\label{}$$ where we should take care to include the appropriate radial dependence in the definition of $\gamma_{\mu\nu}$.
Lets split the contribution from the graviton and the scalar and write $$T^{\mu\nu} = T^{\mu\nu}_g + T^{\mu\nu}_\phi
\label{}$$ where the split is determined by the requirement that $T^{\mu\nu}_\phi \propto \phi$. Then the two pieces can be computed efficiently as follows:
$$\begin{aligned}
T^{\mu\nu}_g = \frac{1}{16\pi \,G_N}\; \frac{2}{\sqrt{-\gamma}} \;\frac{\delta}{\delta \gamma_{\mu\nu}}
& \left[
\int d^{d+1} x\; \sqrt{-g} \, (R+d(d-1))
\right. \nonumber \\
& \left.
+\int d^dx\;\sqrt{-\gamma} \left(2 \, K - 2\, (d-1) - \frac{1}{d-2} \; {}^\gamma R \right)
\right]
\label{tmng1}\end{aligned}$$
which one can show evaluates to a nice covariant expression: $$T^{\mu\nu}_g = \frac{2}{16\pi \,G_N} \; \left(K^{\mu\nu} - K\, \gamma^{\mu\nu} + (d-1)\, \gamma^{\mu\nu} -\frac{1}{d-2}\, \left( {}^\gamma R^{\mu\nu} - \frac{1}{2} {}^\gamma R \, \gamma^{\mu\nu}\right) \right)
\label{tmng}$$ where $z_\epsilon$ is the location of the cut-off surface.
The scalar contribution can be evaluated by using the fact that we are interested in the boundary variations to obtain: $$\begin{aligned}
T^{\mu\nu}_{\phi} = \frac{1}{16\pi \,G_N}\; \frac{2}{\sqrt{-\gamma}} \;\frac{\delta}{\delta \gamma_{\mu\nu}}
\left[
\int d^dx \sqrt{-\gamma}\; \left(\frac{1}{2\, z_\epsilon^{d-1}} \phi\, \partial_z \phi - \frac{1}{2\,z_\epsilon^d} \,\Delta_- \phi^2 + \cdots \right)
\right]
\label{tmnphi1}\end{aligned}$$ where $\cdots$ indicate the contribution from the higher order counter-terms and we have put back the powers of $z_\epsilon$ now. The details now depend on the asymptotic expansion of $\phi$. For general $\Delta$ we have to worry about the fact that the Taylor series solution in the neighbourhood of $z \simeq 0$ looks like $$\phi(z,x) = \phi_0 \, z^{\Delta_-} + a_1(\phi_0) \, z^{\Delta_- -2} + \cdots + \phi_1 \, z^{\Delta_+} + \cdots$$ and we need to know the various intermediate pieces to complete the analysis. The case we are interested in is rather special, where there are no powers of $z$ in the Taylor expansion between the source and the vev, so let us simply record the result for this case for now leaving a more general analysis for later.
Before proceeding though, let us note that we can express covariantly as follows ($ r=z^{-1}$): $$\begin{aligned}
T^{\mu\nu}_{\phi} = \frac{1}{16\pi \,G_N}\; \frac{2}{\sqrt{-\gamma}} \;\frac{\delta}{\delta \gamma_{\mu\nu}}
\left[
\int d^dx \sqrt{-\gamma_\epsilon}\; \left(\frac{1}{2\, r_\epsilon} \; \phi \, n^A \nabla_A\phi- \frac{1}{2} \,\Delta_- \phi^2 + \cdots \right)
\right]
\label{tmnphi2}\end{aligned}$$ where $n^A$ is the unit normal perpendicular to the cut-off surface.
Specializing to $\Delta_- - \Delta_+ < 2$ ($\Delta_+ >d-2$) {#sec:}
-----------------------------------------------------------
In this case the asymptotic expansion belongs to the special kind where $$\phi(z,x) = \phi_0 \, z^{\Delta_-} + \phi_1 \, z^{\Delta_+} + \cdots$$ where we are allowed to use the fact that $\phi_0\, z^{\Delta_-}$ is the beginning of an independent Taylor series where the powers of $z$ change by $2$ units (use the fact that the Lagrangian has $\phi \to -\phi$ symmetry). This corresponds to the case we are interested where $\Delta_+ = 2$, $\Delta_- =1$ in $d= 3$.
In this circumstance we can simply use the terms written explicitly in to obtain $$\begin{aligned}
T^{\mu\nu}_{\phi} = \frac{1}{16\pi \,G_N}\; \frac{1}{2} (2\,\Delta_+ -d) \, \phi_0\, \phi_1\, \gamma^{\mu\nu}\end{aligned}$$ Then we find $$T_{\mu\nu} = \frac{1}{16\pi\,G_N} \, \left(K_{\mu\nu} - K\, \gamma_{\mu\nu} + (d-1)\, \gamma_{\mu\nu} + \frac{1}{2}\, (2\Delta_+ -d)\, \phi_0 \, \phi_1 \, \gamma_{\mu\nu}\right)$$
$m^2 =-2$ in $d=3$ {#sec:}
------------------
Now, we can get the final answer for the case of interest either by working with the Fefferman-Graham expansion in which case we need to know that $$\begin{aligned}
g_{\mu\nu}(z,x) dx^\mu\, dx^\nu&= - \left(1 -\frac{1}{4}\, \phi_0^2\, z^2+ \frac{4}{3}\, a_3 \, z^3 + \cdots \right) dt^2
\nonumber \\
& \qquad + \left(1- \frac{1}{4}\, \phi_0^2 \, z^2- \frac{2}{3}\, (a_3 + \phi_0\, \phi_1) \, z^3 + \cdots \right) \, (dx^2 + dy^2)
\label{eq:metfg}
\end{aligned}$$ The metric fall-offs allow us to compute the pieces in the boundary stress tensor directly since the $z^3$ term above is the correct answer.
Using this or directly computing from the CY-ansatz we claim to obtain (rescaled the result by a factor of $3/2$). $$T^\mu_\nu = \text{diag} \bigg\{2\, a_3 + \, \phi_0\, \phi_1 , -a_3, - a_3\bigg\}$$ We can check that this satisfies the Ward identities: $$T^\mu_\mu = \phi_0\, \phi_1 = {\cal J}\, {\cal O}_2 \,, \qquad \nabla_\mu T^{\mu0} = -2\, \dot{a_3} - \phi_1\, \dot{\phi}_0 - \phi_0\,\dot{\phi}_1 = -\phi_1\, \dot{\phi}_0 = {\cal O}\, \nabla^\nu\, {\cal J}$$ where we used the boundary conservation law derived from the solution $\dot{a}_3 = -\frac{1}{2}\, \phi_0\, \dot{\phi}_1$.
Extremal surfaces and entanglement for strips {#sec:eeapp}
=============================================
In this appendix we describe our methodology for finding extremal surfaces relevant for the computation of entanglement entropy. For simplicity we will focus on regions which exploit the symmetry of our set-up and consider ${\cal A}$ to be a strip extended along one of the translationally invariant directions, say $y$ without any loss of generality, as in Eq. . We need a bulk codimension-2 surface that ends on the boundary of this region i.e., at $x=\pm a$ (at the chosen instant of boundary time $t_{\cal A} $). We describe our strategy for finding this surface and computing its (regulated) area below.
Determining extremal surfaces {#sec:extrdet}
-----------------------------
To find the extremal surface, we start by gauge fixing the reparameterization invariance on the surface. We take $y$ to be one of the coordinates. Dimensionally reducing in this direction, we construct an effective action for a curve in the remaining directions and pick a proper-length parameter $\lambda$ as the second coordinate. Thus, the extremal surface ${\cal E}_{\cal A}$ is embedded in the bulk as $$X^{\mu} = \left( t(\lambda),r(\lambda),x(\lambda) ,y\right) .
\label{eqn:embedding_coordinates}$$ We choose the proper-length parameter to ensure that $\sqrt{ \text{det} \gamma_{ab} } = 1$, which implies that the unregulated area of the extremal surface is given as $$\text{Area}({\cal E}_{\cal A}) = L_y\, \int_{{\cal E}_{\cal A}} d\lambda \sqrt{ \text{det} \gamma_{ab} } = \lambda_{{\cal E}_{\cal A}} \, L_y \,,
\label{eqn:area_proper_length_param}$$ in terms of parameter distance $\lambda_{{\cal E}_{\cal A}}$ spanned by the curve and the IR regulator $L_y$.
In practical terms we work with the effective Lagrangian $$\mathcal{L}= \gxx^{2}\, \left[ 2 \, t'\, e^{2 \br} \left( r' - t'\, \gtt \right)+\gxx^2 \, x'^2\right]
\label{eqn:detgamma}$$ where the metric functions $\rho$, $\gtt$, $\br$ are obtained by interpolation of our numerical solutions. This is a geodesic problem, with some non-minimal coupling from the dimensional reduction along the translationally invariant direction of the strip. Instead if using the geodesic equations, we found it convenient to pass to a set of six *first-order* Hamilton-like equations by introducing $P_{t} = t'$, $P_{x} = x'$, and $P_{+} = r' - \gxx \,t'$ which are related to the conjugate momenta. The equations we solve are the above three and $$\begin{split}
P'_{x} &= -\frac{4 P_x}{\gxx} \left( (P_t \gtt+P_{+} )\partial_r \gxx + P_t \partial_t \gxx \right) =0 \\
P'_{t} &= 2 P_x^2 \gxx e^{-2 \chi} \partial_r \gxx-P_t^2 \left(\partial_r \gtt+2 \gtt \partial_r \chi +2 \partial_t \chi \right)-\frac{2 P_t^2}{\gxx} \left(\gtt \partial_r \gxx + \partial_t \gxx \right) =0 \\
P'_{+} &= 2 P_x^2 \gxx e^{-2 \chi} \left(\gtt \partial_r \gxx + \partial_t \gxx \right) + P_t P_+ \partial_r \gtt- \frac{2 P_+^2}{\gxx} \left(\partial_r \gxx+ \gxx \partial_r \chi \right) =0
\end{split}
\label{eqn:P_eom}$$ We start from $x=0$ in the bulk at some smooth cap-off point $(x=0,t^*,r^*)$ where $t'=r'=0$.[^15] and propagate out to the boundary. We evolve until a with a fixed UV cut-off at $r_{\eebdy}$ and regulate the final answer for the entanglement entropy by background subtraction (see below).
In the main text we illustrate the temporal dependence of the extremal surfaces and $S_{\cal A}^\text{reg}$ for each of the four phases (I-IV) of Fig. \[fig:PP\_qualitative\] for fixed strip width $a$. Since we numerically control the data of the cap-off point we work iteratively: we start by fixing a suitable strip width $a$ by tuning $r^*$ and $t^*$, then we evolve the extremal surfaces by increasing $t^*$ and re-adjusting $r^*$ such that the strip width remains as $a$. We note that we assume that there are no discontinuities or multi-valuedness in the map from $(r^*,t^*) \to (a,r_{\eebdy})$, which we believe makes sense for small strip widths.[^16] Finally, to work in a compact domain we choose $u = 1/r \in [0,1]$ which we will use to explain the properties of the extremal surfaces.
Regulated entanglement entropies {#sec:regent}
--------------------------------
Since the extremal surfaces reach out all the way to the boundary, the proper area is divergent with the coefficient of the leading divergent term fixed by the area of the entangling surface $\partial {\cal A}$. For a state of the CFT with vanishing sources for operators it is well known [@Ryu:2006ef] that the entanglement entropy behaves as $$S_{\cal A}= \frac{ \text{Area}(\partial {\cal A}) }{u_{\cal A}} + S_{\cal A}^\text{fin} + {\cal O}(u_{\cal A}) \,.
\label{eq:sdiv}$$ where $S_{\cal A}^\text{fin}$ is finite in the limit $u_{\cal A} \to 0$. In normalizable states of the field theory $S_{\cal A}^\text{fin}$ is the universal contribution to entanglement which should be independent of the cutoff value $u_{\cal A}$.[^17] One natural way for us to extract this quantity is to measure the entanglement relative to the $t=0$ thermal Schwarzschild state $\Delta S_{\cal A}(t) = S_{\cal A}(t) - S_{\cal A}(t=0)$, which can be extracted simply by vacuum subtraction.
Usually, when we turn on sources for relevant operators, these can contribute additional divergences to the entanglement entropy [@Hung:2011ta]. In general in the presence of additional relevant scales one naively expects there to be logarithmically divergent terms polluting and rendering vacuum subtraction meaningless. Fortuitously, this does no happen for the problem at hand. This can be extracted by examining the detailed discussion of [@Hung:2011ta], which we paraphrase below.
There is however a quick argument for the absence of logarithmic terms which we now describe. For scalar operators in CFT$_d$ with operator dimension $\frac{d}{2}< \Delta < \frac{d}{2}+1$, as we have considered, it is well known in AdS/CFT that the corresponding bulk field has mass in the window where both asymptotic fall-offs are normalizable, i.e., $m^2 \in \left(m_{BF}^2, m_{BF}^2 +1\right)$ with the Breitenlohner-Freedman bound mass $m_{BF}^2 = -\frac{d^2}{4}$ as usual.[^18] In this window note that $\Delta_- - \Delta_+ <2$ and we have the Legendre transformed theory with an operator of dimension $\Delta_-$ by switching to alternate quantization [@Klebanov:1999tb].
Turning on a source for the faster-fall off mode $\Delta_+$ is equivalent, insofar as the leading back-reaction on the metric, to considering instead a state in the Legendre transformed theory where the alternate quantized operator with dimension $\Delta_-$ acquires a vacuum expectation value. However, since the divergence structure of entanglement is the same in all states of the field theory, and the conformal vacua of the two theories (standard and alternate quantization) coincide, it follows that the divergence structure of $S_{\cal A}$ should be unchanged from , even with ${\cal J}(x) \neq 0$. Our story is of course a special case with $\Delta_+ = 2, \Delta_- = 1$ in $d=3$. This observation is consistent with the results of [@Hung:2011ta] and the counter-terms used in [@Auzzi:2013pca].
To explicitly analyze the structure of the divergences in the entanglement entropy, let us consider the metric given in . Since the details of the divergences are blind to the boundary spatio-temporal behaviour of the sources we will examine the somewhat simplified setting where $\phi_0 = \text{const}$ to glean the relevant information.
With the time-translational symmetry restored by this choice, the Lagrangian for the extremal surface (which now is minimal) is simpler: $${\cal L} = \frac{\sqrt{g_{ii}(z)}}{z^2}\, \sqrt{g_{ii} (z)+ z'(x)^2} \,$$ where $g_{ii}(z)$ is the spatial component of the metric in . This system has a conserved Hamiltonian, which we can exploit to write down an expression for the area directly. Introducing, $z_*$ which captures depth to which the minimal surface penetrates into the bulk, we have for the on-shell value of the area $$\text{Area }({\cal E}_{\cal A}) \propto \int_\epsilon^{z_*} \, dz\; \frac{\sqrt{g_{ii}}}{z^2}
\left(1- \frac{g_{ii}(z_*)^2\, z^4}{ g_{ii}(z)^2\, z_*^4} \right)^{-\frac{1}{2}}$$ Using the explicit form of $g_{ii}$, the second term is at least $z^4$ near the boundary so we can forget about it. The first term is all that matters, so lets look at $$\frac{\sqrt{g_{ii}}}{z^2} = \frac{1}{z^2}- \frac{1}{4}\, \phi_0^2 - \frac{2}{3}\, (a_3 + \phi_0\, \phi_1)
\, z + \cdots$$ which has the $z^{-1}$ divergence expected upon integration, but no further contribution of relevance in $z \to 0$ limit. From the $\phi_0^2$ term we get a contribution to the finite part of the entanglement, and this is indeed the physically relevant answer. It should be clear from this discussion is not specific to the choice $m^2 = -2$ in $d=3$, but should hold for $\frac{d}{2}< \Delta_+ < \frac{d}{2}+1$ as we argued abstractly above.
[10]{}
A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalattore, [ *[Nonequilibrium dynamics of closed interacting quantum systems]{}*]{}, [ *Rev.Mod.Phys.*]{} [**83**]{} (2011) 863, \[[[arXiv:1007.5331]{}](http://xxx.lanl.gov/abs/1007.5331)\].
V. E. Hubeny and M. Rangamani, [*[A Holographic view on physics out of equilibrium]{}*]{}, [*Adv.High Energy Phys.*]{} [**2010**]{} (2010) 297916, \[[[arXiv:1006.3675]{}](http://xxx.lanl.gov/abs/1006.3675)\].
P. M. Chesler and L. G. Yaffe, [*[Horizon formation and far-from-equilibrium isotropization in supersymmetric Yang-Mills plasma]{}*]{}, [*Phys.Rev.Lett.*]{} [**102**]{} (2009) 211601, \[[[ arXiv:0812.2053]{}](http://xxx.lanl.gov/abs/0812.2053)\].
P. M. Chesler and L. G. Yaffe, [*[Numerical solution of gravitational dynamics in asymptotically anti-de Sitter spacetimes]{}*]{}, [*JHEP*]{} [ **1407**]{} (2014) 086, \[[[ arXiv:1309.1439]{}](http://xxx.lanl.gov/abs/1309.1439)\].
P. Calabrese and J. L. Cardy, [*[Evolution of entanglement entropy in one-dimensional systems]{}*]{}, [*J.Stat.Mech.*]{} [**0504**]{} (2005) P04010, \[[[cond-mat/0503393]{}](http://xxx.lanl.gov/abs/cond-mat/0503393)\].
P. Calabrese and J. L. Cardy, [*[Time-dependence of correlation functions following a quantum quench]{}*]{}, [*Phys.Rev.Lett.*]{} [**96**]{} (2006) 136801, \[[[cond-mat/0601225]{}](http://xxx.lanl.gov/abs/cond-mat/0601225)\].
P. Calabrese and J. Cardy, [*[Quantum Quenches in Extended Systems]{}*]{}, [ *J.Stat.Mech.*]{} [**0706**]{} (2007) P06008, \[[[arXiv:0704.1880]{}](http://xxx.lanl.gov/abs/0704.1880)\].
S. Bhattacharyya and S. Minwalla, [*[Weak Field Black Hole Formation in Asymptotically AdS Spacetimes]{}*]{}, [*JHEP*]{} [**0909**]{} (2009) 034, \[[[arXiv:0904.0464]{}](http://xxx.lanl.gov/abs/0904.0464)\].
S. R. Das, T. Nishioka, and T. Takayanagi, [*[Probe Branes, Time-dependent Couplings and Thermalization in AdS/CFT]{}*]{}, [*JHEP*]{} [**1007**]{} (2010) 071, \[[[arXiv:1005.3348]{}](http://xxx.lanl.gov/abs/1005.3348)\].
P. Basu and S. R. Das, [*[Quantum Quench across a Holographic Critical Point]{}*]{}, [*JHEP*]{} [**1201**]{} (2012) 103, \[[[arXiv:1109.3909]{}](http://xxx.lanl.gov/abs/1109.3909)\].
A. Buchel, L. Lehner, and R. C. Myers, [*[Thermal quenches in N=2\* plasmas]{}*]{}, [*JHEP*]{} [**1208**]{} (2012) 049, \[[[arXiv:1206.6785]{}](http://xxx.lanl.gov/abs/1206.6785)\].
M. Bhaseen, J. P. Gauntlett, B. Simons, J. Sonner, and T. Wiseman, [ *[Holographic Superfluids and the Dynamics of Symmetry Breaking]{}*]{}, [ *Phys.Rev.Lett.*]{} [**110**]{} (2013) 015301, \[[[arXiv:1207.4194]{}](http://xxx.lanl.gov/abs/1207.4194)\].
P. Basu, D. Das, S. R. Das, and T. Nishioka, [*[Quantum Quench Across a Zero Temperature Holographic Superfluid Transition]{}*]{}, [*JHEP*]{} [**1303**]{} (2013) 146, \[[[ arXiv:1211.7076]{}](http://xxx.lanl.gov/abs/1211.7076)\].
M. Nozaki, T. Numasawa, and T. Takayanagi, [*[Holographic Local Quenches and Entanglement Density]{}*]{}, [*JHEP*]{} [**1305**]{} (2013) 080, \[[[arXiv:1302.5703]{}](http://xxx.lanl.gov/abs/1302.5703)\].
A. Buchel, L. Lehner, R. C. Myers, and A. van Niekerk, [*[Quantum quenches of holographic plasmas]{}*]{}, [*JHEP*]{} [**1305**]{} (2013) 067, \[[[arXiv:1302.2924]{}](http://xxx.lanl.gov/abs/1302.2924)\].
T. Hartman and J. Maldacena, [*[Time Evolution of Entanglement Entropy from Black Hole Interiors]{}*]{}, [*JHEP*]{} [**1305**]{} (2013) 014, \[[[arXiv:1303.1080]{}](http://xxx.lanl.gov/abs/1303.1080)\].
P. Basu and A. Ghosh, [*[Dissipative Nonlinear Dynamics in Holography]{}*]{}, [*Phys.Rev.*]{} [**D89**]{} (2014) 046004, \[[[arXiv:1304.6349]{}](http://xxx.lanl.gov/abs/1304.6349)\].
W.-J. Li, Y. Tian, and H.-b. Zhang, [*[Periodically Driven Holographic Superconductor]{}*]{}, [*JHEP*]{} [**1307**]{} (2013) 030, \[[[arXiv:1305.1600]{}](http://xxx.lanl.gov/abs/1305.1600)\].
A. Buchel, R. C. Myers, and A. van Niekerk, [*[Universality of Abrupt Holographic Quenches]{}*]{}, [*Phys.Rev.Lett.*]{} [**111**]{} (2013) 201602, \[[[arXiv:1307.4740]{}](http://xxx.lanl.gov/abs/1307.4740)\].
R. Auzzi, S. Elitzur, S. B. Gudnason, and E. Rabinovici, [*[On periodically driven AdS/CFT]{}*]{}, [*JHEP*]{} [**1311**]{} (2013) 016, \[[[arXiv:1308.2132]{}](http://xxx.lanl.gov/abs/1308.2132)\].
A. Buchel, R. C. Myers, and A. van Niekerk, [*[Nonlocal probes of thermalization in holographic quenches with spectral methods]{}*]{}, [[arXiv:1410.6201]{}](http://xxx.lanl.gov/abs/1410.6201).
U. H. Danielsson, E. Keski-Vakkuri, and M. Kruczenski, [*[Black hole formation in AdS and thermalization on the boundary]{}*]{}, [*JHEP*]{} [**0002**]{} (2000) 039, \[[[ hep-th/9912209]{}](http://xxx.lanl.gov/abs/hep-th/9912209)\].
V. E. Hubeny, M. Rangamani, and T. Takayanagi, [*[A Covariant holographic entanglement entropy proposal]{}*]{}, [*JHEP*]{} [**0707**]{} (2007) 062, \[[[arXiv:0705.0016]{}](http://xxx.lanl.gov/abs/0705.0016)\].
J. Abajo-Arrastia, J. Aparicio, and E. Lopez, [*[Holographic Evolution of Entanglement Entropy]{}*]{}, [*JHEP*]{} [**1011**]{} (2010) 149, \[[[arXiv:1006.4090]{}](http://xxx.lanl.gov/abs/1006.4090)\].
T. Albash and C. V. Johnson, [*[Evolution of Holographic Entanglement Entropy after Thermal and Electromagnetic Quenches]{}*]{}, [*New J.Phys.*]{} [ **13**]{} (2011) 045017, \[[[ arXiv:1008.3027]{}](http://xxx.lanl.gov/abs/1008.3027)\].
V. Balasubramanian, A. Bernamonti, J. de Boer, N. Copland, B. Craps, [ *et. al.*]{}, [*[Thermalization of Strongly Coupled Field Theories]{}*]{}, [ *Phys.Rev.Lett.*]{} [**106**]{} (2011) 191601, \[[[arXiv:1012.4753]{}](http://xxx.lanl.gov/abs/1012.4753)\].
V. Balasubramanian, A. Bernamonti, J. de Boer, N. Copland, B. Craps, [ *et. al.*]{}, [*[Holographic Thermalization]{}*]{}, [*Phys.Rev.*]{} [**D84**]{} (2011) 026010, \[[[ arXiv:1103.2683]{}](http://xxx.lanl.gov/abs/1103.2683)\].
J. Aparicio and E. Lopez, [*[Evolution of Two-Point Functions from Holography]{}*]{}, [*JHEP*]{} [**1112**]{} (2011) 082, \[[[arXiv:1109.3571]{}](http://xxx.lanl.gov/abs/1109.3571)\].
V. Balasubramanian, A. Bernamonti, N. Copland, B. Craps, and F. Galli, [ *[Thermalization of mutual and tripartite information in strongly coupled two dimensional conformal field theories]{}*]{}, [*Phys.Rev.*]{} [**D84**]{} (2011) 105017, \[[[arXiv:1110.0488]{}](http://xxx.lanl.gov/abs/1110.0488)\].
V. Keranen, E. Keski-Vakkuri, and L. Thorlacius, [*[Thermalization and entanglement following a non-relativistic holographic quench]{}*]{}, [ *Phys.Rev.*]{} [**D85**]{} (2012) 026005, \[[[arXiv:1110.5035]{}](http://xxx.lanl.gov/abs/1110.5035)\].
D. Galante and M. Schvellinger, [*[Thermalization with a chemical potential from AdS spaces]{}*]{}, [*JHEP*]{} [**1207**]{} (2012) 096, \[[[arXiv:1205.1548]{}](http://xxx.lanl.gov/abs/1205.1548)\].
E. Caceres and A. Kundu, [*[Holographic Thermalization with Chemical Potential]{}*]{}, [*JHEP*]{} [**1209**]{} (2012) 055, \[[[arXiv:1205.2354]{}](http://xxx.lanl.gov/abs/1205.2354)\].
V. E. Hubeny, M. Rangamani, and E. Tonni, [*[Thermalization of Causal Holographic Information]{}*]{}, [*JHEP*]{} [**1305**]{} (2013) 136, \[[[arXiv:1302.0853]{}](http://xxx.lanl.gov/abs/1302.0853)\].
H. Liu and S. J. Suh, [*[Entanglement Tsunami: Universal Scaling in Holographic Thermalization]{}*]{}, [*Phys.Rev.Lett.*]{} [**112**]{} (2014) 011601, \[[[arXiv:1305.7244]{}](http://xxx.lanl.gov/abs/1305.7244)\].
V. Balasubramanian, A. Bernamonti, J. de Boer, B. Craps, L. Franti, [ *et. al.*]{}, [*[Inhomogeneous holographic thermalization]{}*]{}, [*JHEP*]{} [ **1310**]{} (2013) 082, \[[[ arXiv:1307.7086]{}](http://xxx.lanl.gov/abs/1307.7086)\].
H. Liu and S. J. Suh, [*[Entanglement growth during thermalization in holographic systems]{}*]{}, [*Phys.Rev.*]{} [**D89**]{} (2014) 066012, \[[[arXiv:1311.1200]{}](http://xxx.lanl.gov/abs/1311.1200)\].
J. Abajo-Arrastia, E. da Silva, E. Lopez, J. Mas, and A. Serantes, [ *[Holographic Relaxation of Finite Size Isolated Quantum Systems]{}*]{}, [ *JHEP*]{} [**1405**]{} (2014) 126, \[[[ arXiv:1403.2632]{}](http://xxx.lanl.gov/abs/1403.2632)\].
S. Bhattacharyya, R. Loganayagam, S. Minwalla, S. Nampuri, S. P. Trivedi, [ *et. al.*]{}, [*[Forced Fluid Dynamics from Gravity]{}*]{}, [*JHEP*]{} [**0902**]{} (2009) 018, \[[[ arXiv:0806.0006]{}](http://xxx.lanl.gov/abs/0806.0006)\].
S. Hayward, [*[General laws of black hole dynamics]{}*]{}, [*Phys.Rev.*]{} [ **D49**]{} (1994) 6467–6474.
I. Booth, [*[Black hole boundaries]{}*]{}, [*Can.J.Phys.*]{} [**83**]{} (2005) 1073–1099, \[[[ gr-qc/0508107]{}](http://xxx.lanl.gov/abs/gr-qc/0508107)\].
M. Rangamani, M. Rozali, and A. Wong, [*[work in progress]{}*]{}.
K. Balasubramanian and C. P. Herzog, [*[Losing Forward Momentum Holographically]{}*]{}, [*Class.Quant.Grav.*]{} [**31**]{} (2014) 125010, \[[[arXiv:1312.4953]{}](http://xxx.lanl.gov/abs/1312.4953)\].
R. M. Wald and V. Iyer, [*[Trapped surfaces in the Schwarzschild geometry and cosmic censorship]{}*]{}, [*Phys.Rev.*]{} [**D44**]{} (1991) 3719–3722.
P. Figueras, V. E. Hubeny, M. Rangamani, and S. F. Ross, [*[Dynamical black holes and expanding plasmas]{}*]{}, [*JHEP*]{} [**0904**]{} (2009) 137, \[[[arXiv:0902.4696]{}](http://xxx.lanl.gov/abs/0902.4696)\].
S. Caron-Huot, P. M. Chesler, and D. Teaney, [*[Fluctuation, dissipation, and thermalization in non-equilibrium AdS$_5$ black hole geometries]{}*]{}, [ *Phys.Rev.*]{} [**D84**]{} (2011) 026012, \[[[arXiv:1102.1073]{}](http://xxx.lanl.gov/abs/1102.1073)\].
P. M. Chesler and D. Teaney, [*[Dynamical Hawking Radiation and Holographic Thermalization]{}*]{}, [[ arXiv:1112.6196]{}](http://xxx.lanl.gov/abs/1112.6196).
V. Balasubramanian and S. F. Ross, [*[Holographic particle detection]{}*]{}, [*Phys.Rev.*]{} [**D61**]{} (2000) 044007, \[[[hep-th/9906226]{}](http://xxx.lanl.gov/abs/hep-th/9906226)\].
J. Louko, D. Marolf, and S. F. Ross, [*[On geodesic propagators and black hole holography]{}*]{}, [*Phys.Rev.*]{} [**D62**]{} (2000) 044041, \[[[hep-th/0002111]{}](http://xxx.lanl.gov/abs/hep-th/0002111)\].
M. Headrick, V. E. Hubeny, A. Lawrence, and M. Rangamani, [*[Causality & holographic entanglement entropy]{}*]{}, [*JHEP*]{} [**1412**]{} (2014) 162, \[[[arXiv:1408.6300]{}](http://xxx.lanl.gov/abs/1408.6300)\].
G. [Bunin]{}, L. [D’Alessio]{}, Y. [Kafri]{}, and A. [Polkovnikov]{}, [*[Universal energy fluctuations in thermally isolated driven systems]{}*]{}, [*Nature Physics*]{} [**7**]{} (Nov., 2011) 913–917, \[[[arXiv:1102.1735]{}](http://xxx.lanl.gov/abs/1102.1735)\].
S. Ryu and T. Takayanagi, [*[Holographic derivation of entanglement entropy from AdS/CFT]{}*]{}, [*Phys.Rev.Lett.*]{} [**96**]{} (2006) 181602, \[[[hep-th/0603001]{}](http://xxx.lanl.gov/abs/hep-th/0603001)\].
S. Ryu and T. Takayanagi, [*[Aspects of Holographic Entanglement Entropy]{}*]{}, [*JHEP*]{} [**0608**]{} (2006) 045, \[[[hep-th/0605073]{}](http://xxx.lanl.gov/abs/hep-th/0605073)\].
V. E. Hubeny and H. Maxfield, [*[Holographic probes of collapsing black holes]{}*]{}, [*JHEP*]{} [**1403**]{} (2014) 097, \[[[arXiv:1312.6887]{}](http://xxx.lanl.gov/abs/1312.6887)\].
H. Ebrahim and M. Headrick, [*[Instantaneous Thermalization in Holographic Plasmas]{}*]{}, [[arXiv:1010.5443]{}](http://xxx.lanl.gov/abs/1010.5443).
V. Keranen and P. Kleinert, [*[Non-equilibrium scalar two point functions in AdS/CFT]{}*]{}, [[arXiv:1412.2806]{}](http://xxx.lanl.gov/abs/1412.2806).
R. C. Myers and A. Sinha, [*[Seeing a c-theorem with holography]{}*]{}, [ *Phys.Rev.*]{} [**D82**]{} (2010) 046006, \[[[arXiv:1006.1263]{}](http://xxx.lanl.gov/abs/1006.1263)\].
D. L. Jafferis, I. R. Klebanov, S. S. Pufu, and B. R. Safdi, [*[Towards the F-Theorem: N=2 Field Theories on the Three-Sphere]{}*]{}, [*JHEP*]{} [**1106**]{} (2011) 102, \[[[ arXiv:1103.1181]{}](http://xxx.lanl.gov/abs/1103.1181)\].
H. Liu and M. Mezei, [*[A Refinement of entanglement entropy and the number of degrees of freedom]{}*]{}, [*JHEP*]{} [**1304**]{} (2013) 162, \[[[arXiv:1202.2070]{}](http://xxx.lanl.gov/abs/1202.2070)\].
H. Casini and M. Huerta, [*[On the RG running of the entanglement entropy of a circle]{}*]{}, [*Phys.Rev.*]{} [**D85**]{} (2012) 125016, \[[[arXiv:1202.5650]{}](http://xxx.lanl.gov/abs/1202.5650)\].
L.-Y. Hung, R. C. Myers, and M. Smolkin, [*[Some Calculable Contributions to Holographic Entanglement Entropy]{}*]{}, [*JHEP*]{} [**1108**]{} (2011) 039, \[[[arXiv:1105.6055]{}](http://xxx.lanl.gov/abs/1105.6055)\].
I. R. Klebanov and E. Witten, [*[AdS / CFT correspondence and symmetry breaking]{}*]{}, [*Nucl.Phys.*]{} [**B556**]{} (1999) 89–114, \[[[hep-th/9905104]{}](http://xxx.lanl.gov/abs/hep-th/9905104)\].
[^1]: This choice turns out to have several advantages as the dual scalar being conformally coupled to gravity in the bulk allows a certain level of technical simplification in various holographic renormalizations we need to do to extract physical data.
[^2]: As we will be describing the dynamics of Einstein-scalar system with the scalar field satisfying the null energy condition, the areas of the event and apparent horizon (in the canonical foliation) have to grow monotonically – a consequence of the area theorem [@Hayward:1993wb] (see [@Booth:2005qc] for an excellent overview). We will elaborate on this point in §\[sec:bulksol\].
[^3]: We note that [@Li:2013fhw] study the influence of a periodic electric fields on the phase transition between a normal and superconducting phase using holography. It is clear in this case that a driving the system will make it exit the low temperature superconducting phase as the energy expended heats up the system past the critical point, as their analysis confirms.
[^4]: We have also set $\ell_\text{AdS} =1$ for simplicity.
[^5]: To be sure the statement of the area increase theorem does rely on the null energy condition, which we happily assume, for it is always satisfied by scalar fields with sensible kinetic terms.
[^6]: We thank Alex Buchel for checking this and confirming the monotone growth of the apparent horizon area.
[^7]: Using the area of the apparent horizon (defined as the outermost trapped surface in the foliation respecting spatial homogeneity) results a causal boundary observable. One maps points on the apparent horizon to boundary points by Lie transport along radially ingoing null geodesics, which in the ansatz are simply lines of constant $\{t,x,y\}$. On the other hand the teleological nature of the event horizon implies that its area would not provide a good measure for the boundary entropy density, cf., [@Chesler:2008hg; @Figueras:2009iu] for a discussion of this point.
[^8]: Note that the averaging makes $\epsilon_\text{avg}(t)$ and $s_\text{avg}(t)$ discrete in time.
[^9]: We could following standard practice attempt to compute two-point correlation functions using the geodesic approximation [@Balasubramanian:1999zv]. However, as discussed in [@Louko:2000tp] and more recently in [@Headrick:2014cta], this prescription doesn’t generically reproduce correct time-ordered correlation functions (we really want in-in correlation functions in our set-up). As a result we will also refrain from computing geodesics in the numerical background.
[^10]: That is, regard $\cos \omega t$ and $\sin \omega t$ as the real and imaginary parts of $e^{i \omega t}$.
[^11]: This is similar to the behaviour of the conductivity for electromagnetic perturbations in asymptotically AdS space.
[^12]: Details of the divergent structure and the counter-terms necessary to compute the area functional in our set up can be found in Appendix \[sec:regent\].
[^13]: The coordinate $u =1/r$ is chosen such that the horizon remains at $u=1$ during the entire course of the evolution (the boundary is at $u=0$).
[^14]: Note that in absolute terms however, $t^*$ in both regimes is comparable in magnitude.
[^15]: This cap-off point is not necessarily the deepest point in the bulk; for the examples shown in this paper it however does turn out to coincide.
[^16]: Such behaviour was noticed in extremal surface computation in global Vaidya-AdS by [@Hubeny:2013dea].
[^17]: For the vacuum state of a CFT$_3$ with ${\cal A}$ being a circular disc $S_{\cal A}^\text{fin}$ would give the F-function [@Myers:2010xs; @Jafferis:2011zi] (the latter defined as the logarithm of the partition function of the theory a three-sphere). In fact, this can be used to define a UV finite quantity without recourse to background subtraction: following [@Liu:2012eea; @Casini:2012ei] we can just as well consider $\left(R\,\frac{d}{dR} -1\right) S_{\cal A}$, with $R$ being the disc radius, as the measure of entanglement growth.
[^18]: Implicit in this statement is the fact that we are quantizing the scalar field with standard (Dirichlet) boundary conditions, so that the dimension of the dual operator is $\Delta = \Delta_+$.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Evaluation of quality of experience (QoE) based on electroencephalography (EEG) has received great attention due to its capability of real-time QoE monitoring of users. However, it still suffers from rather low recognition accuracy. In this paper, we propose a novel method using deep neural networks toward improved modeling of EEG and thereby improved recognition accuracy. In particular, we aim to model spatio-temporal characteristics relevant for QoE analysis within learning models. The results demonstrate the effectiveness of the proposed method.'
author:
-
bibliography:
- 'refs.bib'
title: Evaluation of preference of multimedia content using deep neural networks for electroencephalography
---
Introduction {#sec:intro}
============
With the noticeable growing of the demand for multimedia content, user-adaptive content delivery has become a key to success of many multimedia services. Consequently, it is crucial to understand how users perceive the multimedia, which is the concept of quality of experience (QoE) defined as “the degree of delight or annoyance of the user of an application or service” [@Qualinet13]. Traditionally, QoE has been measured explicitly, i.e., subjects are asked about their experience with the given multimedia content via an interview or a questionnaire. However, it is difficult for this explicit approach to capture the user responses in real-time because the evaluation is typically implemented after the presentation of content.
On the other hand, QoE also can be monitored through the implicit cues obtained from the physiological or behavioral signals of users, which enables real-time monitoring of QoE. Particularly, the brain signals such as electroencephalography (EEG) are expected to provide deeper insight into the perceptual experience of multimedia because they contain the whole information of the multimedia perception, whereas the explicit approach can measure only predefined final outputs of the perception.
Many studies have employed the EEG signals to capture the degradation of QoE [@Antons12; @Scholler12; @Mustafa12] and the overall QoE [@Antons13; @Arndt14; @Koelstra12; @Moon15; @Perrin15; @Kroupi14; @Kroupi16]. They showed the potential of EEG to automatically monitor QoE of users, which can be used for many applications such as QoE-aware video scaling for content delivery [@Lee12] and personalized multimedia recommendation [@Moon17survey].
A limitation of the existing EEG-based implicit QoE assessment systems is that their performance still remains at insufficient levels for real-world applications where high reliability is critical. For instance, Table \[table:previous\] summarizes the results reported in representative studies on binary classification of content preference for the DEAP database [@Koelstra12]. It can be seen that even the recent deep learning approaches show accuracies lower than 90%.
\[table:previous\]
[>m[0.5cm]{} >m[2cm]{} >m[2.5cm]{} >m[1.5cm]{}]{} **Ref.** & **Classifier** & **Classification scheme** & **Classification accuracy**
[@Koelstra12] & Gaussian naive Bayes classifier & leave-one-video-out for each subject & 0.502 (F1-score) [@Gupta16] & Relevance vector machine & leave-one-video-out for each subject & 0.65 (F1-score) [@Amjadzadeh17] & Ensemble classifier\* & leave-one-trial-out & 0.647 [@Zhuang14] & Support vector machine & leave-one-video-out for each subject & 0.705 [@Xu16] & Deep belief network & five-fold cross-validation for each subject & 0.867 (F1-score) [@Alhagry17] & Recurrent neural network & four-fold cross-validation & 0.880
\
\*Ensemble of support vector machine, nearest mean, 1-nearest neighbor, k-nearest neighbor, and linear discriminant analysis
We notice that the spatial relationship of EEG signals has not been significantly considered in the previous EEG-based QoE recognition studies although it possibly includes useful information of neural activities. In the resting state of the brain, the neural activities of different brain regions show a certain relationship that comprises the functional resting-state network [@Beckmann05]. However, if any stimulus is given, the spatial relationship is altered because the neural activity of interest appears.
In this paper, we propose a novel approach to improve the accuracy of EEG-based preference recognition. Particularly, convolutional neural networks (CNNs) are employed, which has the capability to analyze the spatial information of EEG signals. The contributions of this work can be summarized as follows:
- We achieve high recognition accuracy of preference based on EEG by adopting deep CNNs that enable spatial analysis of EEG signals. This demonstrates the feasibility of the real-world applications using EEG such as real-time QoE monitoring, automatic feedback generation, QoE-aware multimedia compression, and so on.
- We compare various types of EEG features, input shapes for CNNs, and CNN structures with different complexity, which contributes to further related studies by providing guidelines for system design.
EEG signal features {#sec:feature}
===================
This section describes the EEG features that are employed as inputs of the CNNs. They can be categorized depending on whether the feature considers the activation of a single region or multiple regions. In this paper, one feature indicating the activation level of a single brain region and three features that consider the activation of multiple brain regions are employed. Details of the features are explained below.
Power spectral density (PSD) {#subsec:psd}
----------------------------
PSD represents the activation level of a single electrode. It is calculated using the Welch’s method, which is a non-parametric spectral estimation method based on the Fourier transform. For the $i$-th window of EEG signals $x_i, (i=1, 2, ..., M)$, the periodogram at frequency $f$ is calculated as: $$P_i(f) = \frac{1}{NW}\left| \sum_{n=1}^{N} w(n)x_i(n)e^{-j2\pi fn} \right| ^2,$$ where $N$ is the number of data points in the window, $w(n)$ is a window function, and $W$ is a normalization constant given as $W=\frac{1}{N}\sum_{n=1}^{N} \left|w(n)\right|^2$. PSD is obtained by averaging the periodogram over the windows: $$PSD(f) = \frac{1}{M}\sum_{i=1}^{M} P_i(f).$$
Furthermore, the PSD values of the baseline signals (five seconds before presentation of stimuli) are subtracted from those of the corresponding trial signals to eliminate irrelevant neural activities.
Pearson correlation coefficient (PCC) {#subsec:pcc}
-------------------------------------
PCC is a measure of the linear relationship between two signals, which ranges from -1 to 1. A PCC value of -1 (1) corresponds to the perfect negative (positive) linear relationship, and a PCC value of zero indicates that there is no linear relationship between the two signals. It is calculated as follows: $$PCC = \frac{cov(X,Y)}{\sigma_X\sigma_Y},$$ where $\sigma_X$ and $\sigma_Y$ indicate the standard deviations of the given two signals $X=\{x_i\}$ and $Y=\{y_i\}$, respectively, and $cov(X,Y)$ is the covariance between them.
Phase locking value (PLV) {#subsec:plv}
-------------------------
PLV [@Lachaux99] describes the phase synchronization between two signals, which is calculated as an absolute value of the average phase differences over temporal windows. This can be presented as: $$PLV = \frac{1}{M} \left| \sum_{i=1}^{M} e^{j\Delta\phi_i} \right|,$$ where $\Delta\phi_i$ indicates the phase difference of the $i$-th window. PLV ranges between 0 to 1, which correspond to independence and perfect synchronization of two signals, respectively.
Transfer entropy (TE) {#subsec:te}
---------------------
TE [@Schreiber00] measures information flow between two time series, assuming that the two time series can be approximated by Markov chains. It is defined as: $$TE_{Y\to X} = \sum p(x_{i+1},x_i,y_i)\log \frac{p(x_{i+1}|x_i,y_i)}{p(x_{i+1}|x_i)}.
\label{eq:te}$$ The result of (\[eq:te\]) is the directional information that indicates the ability of time series $Y$ to improve the prediction of time series $X$. We use the Java Information Dynamics Toolkit [@Lizier14] to obtain TE features.
System design {#sec:system}
=============
Input {#subsec:input}
-----
PSD indicates the activation level of a single regions of the brain. Therefore, the PSD values can be represented as a topography, i.e., they are allocated to the locations of the corresponding electrodes and the rest of the scalp is filled by interpolation [@Bashivan16]. Examples of the topography are shown in Figure \[fig:topoexp\]. The outside of the head is filled with zeros.
In contrast to PSD, it is difficult to describe the other features as a topographic figure because they measure the relationship between two regions of the brain, which is called the brain connectivity [@Friston11]. We transform the features into matrices used as CNN inputs, whose ($i, j$)-th element is the feature value obtained by using the data of the $i$-th and $j$-th electrodes.
Here, the order of electrodes in the input matrix becomes important because the filters of a CNN learn localized patterns of the matrix. We consider two different ordering methods, namely, ‘distance’ and ‘random’. The first method arranges the EEG electrodes according to the distance between two electrodes so that physically neighboring electrodes are adjacent in the matrix. At the same time, it considers the hemispheric structure of the brain as shown in Figure \[fig:dist1\]. That is, the ordering starts from the left frontal electrode and proceeds to the nearest electrode in the depth direction of the head within the left hemisphere; after finishing the left side of the head, it shifts to the occipital area of the right hemisphere, repeats the same process for the right side, and ends at the center. The second method simply randomizes the order of electrodes.
![Ordering method of the EEG electrodes based on the distance between electrodes and hemispheric structure of the head.[]{data-label="fig:dist1"}](ord_01.eps){width="0.6\columnwidth"}
CNN structure {#subsec:cnn}
-------------
![Example CNN structure.[]{data-label="fig:cnn"}](cnn-2.eps){width="\columnwidth"}
Three CNN structures with different complexity are adopted for the EEG-based preference recognition. The simplest structure includes one convolutional layer and one max-pooling layer as illustrated in Figure \[fig:cnn\]. The second structure has one convolutional layer, one max-pooling layer, two convolutional layers, one max-pooling layer, and finally a fully connected layer. The third structure consists of five convolutional layers and five max-pooling layers, one after the other, before the fully connected layer. The three CNN structures are denoted as CNN1, CNN2, and CNN3 in the following.
The first convolutional layer of each CNN structure has 32 filters, and the number of filters of the following convolutional layer becomes twice that of the previous convolutional layer. The size of the filters is fixed as 3$\times$3 for all convolutional layers. The rectified linear unit (ReLU) is employed as the activation function. The max-pooling is conducted for 2$\times$2 patches, and the batch normalization is implemented after every max-pooling.
The CNNs are implemented in Theano. The Adam algorithm is used for training by minimizing the loss defined by the cross-entropy function. The training is conducted with a Tesla K80 GPU, where the batch size is set to 256.
Database {#subsec:data}
--------
We employ the DEAP database [@Koelstra12] that contains one-minute-long 32-channel EEG signals recorded while 32 subjects were watching videos and the corresponding preference scores that indicate how much the subjects like the videos. It has been popularly used for analyzing multimedia experience based on EEG. As the number of electrodes is 32, the sizes of the feature matrices for PCC, PLV, and TE become 32$\times$32, and the topographies for PSD are also rendered into 32$\times$32 pixels to have the same input size.
The EEG signals are divided into three-second-long segments with an overlap of 2.5 seconds. Thus, the total number of data is 147,200 (32 subjects$\times$40 videos$\times$115 segments). These data are divided into five clusters randomly, which are used for a five-fold leave-one-cluster-out cross-validation scheme. The features are calculated for delta (0-3 Hz), theta (4-7 Hz), low alpha (8-9.5 Hz), high alpha (10.5-12 Hz), alpha (8-12 Hz), low beta (13-16 Hz), mid beta (17-20 Hz), high beta (21-29 Hz), beta (13-29 Hz), and gamma (30-50 Hz) frequency bands of EEG signals. Consequently, the sizes of CNN inputs become 32$\times$32$\times$10.
We examine two scenarios of preference prediction. First, a binary classification of preference (liking vs. disliking) is considered. As the original preference score in the database lies in a 9-point scoring scale, we define the videos received preference scores between 1 and 5 as one class, and the rest as the other. As a result, 33.52% of the entire data are labeled as the ‘disliking’ class, and 66.48% of the data are assigned as the ‘liking’ class. Note that the sizes of the two classes are highly imbalanced. Therefore, the F1-score is used for evaluation of prediction results. For the random ordering, the final result is obtained by averaging the F1-score with three differently randomized orders. Second, the subjective preference score is estimated, which is a regression task. The regression performance is assessed in terms of the root-mean-square error (RMSE) between the ground truth and predicted preference scores.
Results {#sec:res}
=======
\[table:clsresult\]
[>m[1.5cm]{} c c c c c]{} & &\
& & PCC & PLV & TE & PSD\
& distance & 0.932 & 0.967 & 0.945 &\
& random & 0.936 & 0.966 & 0.942 &\
& distance & 0.938 & 0.969 & 0.921 &\
& random & 0.937 & 0.959 & 0.911 &\
& distance & 0.927 & 0.907 & 0.811 &\
& random & 0.914 & 0.895 & 0.808 &\
Binary classification {#subsec:res-cls}
---------------------
The results of binary classification are shown in Table \[table:clsresult\]. Overall, the proposed system results in much higher F1-scores than the previous works shown in Table \[table:previous\]. The best performance (F1-score = 0.969) is obtained with the combination of CNN2, the distance-based ordering method, and PLV. There are also several other cases showing comparable performance to the best case.
When the complexity of CNNs is examined, the obtained results indicate that a more complex structure does not necessarily produces better classification performance. The most complex network, i.e., CNN3, show rather degraded classification performance compared with the simpler networks for PCC, PLV, and TE, and CNN2 also results in lower F1-scores for TE and PLV (random order). Only the performance of PSD is improved by adopting more complex CNN architectures.
The features concerning the relationship between different brain regions demonstrate better classification results, i.e., F1-scores with PCC, PLV, and TE significantly exceed those with PSD, except for the case with TE and CNN3. In particular, PLV shows the best performance among the features with relatively shallow CNN structures (CNN1 and CNN2), but PCC is better than the other features with the most complex CNN structure (CNN3).
The prediction performance also varies depending on the ordering method. Overall, the physical distance-based ordering yields better accuracy than the random ordering, which indicates the strategy of distance-based ordering, highlighting the information of interest by allocating a single receptive field to feature values that are possibly similar to each other, works better for recognition.
\[table:regresult\]
[>m[1.5cm]{} c c c c c]{} & &\
& & PCC & PLV & TE & PSD\
& distance & 1.538 & 1.429 & 1.557 &\
& random & 1.536 & 1.417 & 1.546 &\
& distance & 1.340 & 1.320 & 1.517 &\
& random & 1.386 & 1.309 & 1.549 &\
& distance & 1.238 & 1.228 & 1.643 &\
& random & 1.252 & 1.237 & 1.644 &\
Score regression {#subsec:res-reg}
----------------
Table \[table:regresult\] shows the results of the preference score regression. The best result (RMSE = 1.228) is achieved when PLV matrices formed using the distance-based method are employed for CNN3. This demonstrates that it is feasible to specify the level of preference in a finer scale than the binary classification.
As in the binary classification results, the features containing relational information outperform PSD. The RMSE of PSD is always larger than those of PCC, PLV, and TE, and PLV shows the best performance for all CNN structures.
The influence of the complexity of CNN structures is different from that for the classification. The results of PCC, PLV, and PSD are improved by using more complex networks; for TE, the best performance is obtained with CNN2.
Furthermore, although the distance-based ordering consistently provides better performance for CNN3, the result of the random ordering is better, in particular, for CNN1.
Discussion {#subsec:res-dis}
----------
From the results of the binary classification and score regression, we consistently observed the superiority of the features that measure the relationship between different brain regions. That is, such relationship includes useful information for prediction of preference.
In particular, we compared three features that reflect different aspects of the relationship. PLV that measures the phase synchronization between brain regions achieves high prediction accuracy with relatively shallow structures in the binary classification. In the score regression, PLV consistently provides the best performance regardless of CNN structures.
PCC and PLV are rather traditional, simple approaches to analyze the brain connectivity without consideration of directionality, but show better performance than TE measuring directional information flow between different regions. This is somewhat surprising, and would require further investigation in the future.
The influence of the network complexity was also examined from the results. In the binary classification, the CNN structure with medium complexity provides the best recognition result, and the most complex network shows the worse result. However, a more complex network shows a better result overall for the score regression. This is probably because the score regression is more difficult to solve than the binary classification, i.e., the mapping function of the score regression is more complex than that of the binary classification.
It is observed that feature matrices aligned according to the distance-based order performs better than those using the random order when the deep CNN structure is used, whereas such superiority is not prominent (binary classification) or not observed (score regression) in CNN1. This is probably because the structure of CNN1 is not complex enough to take the advantage of distance-based ordering.
We further analyze the failure cases of the binary classification to verify whether such cases are influenced by the specific subject or video. Histograms of the number of misclassifications for the best case in Table \[table:clsresult\] are shown in Figure \[fig:fail\]. The indexes of subjects and videos are in a descending order of the number of failures.
It can be noticed that the recognition performance notably varies depending on the subject. The first three subjects take 17% of the failure cases, whereas the last three subjects occupy only 3%. As no noticeable rating tendency is found for those subjects, this variance of recognition performance indicates that there is significant individual difference in neural activities related to QoE.
It is also observed that the performance significantly differs depending on the video. The number of false classifications of the first five videos occupies 25% of the entire failure cases. In order to analyze the relationship between the classification performance and video characteristics, Figure \[fig:emotion-mis\] plots the preference, valence, and arousal scores with respect to the number of misclassifications. The five videos with larger numbers of failure received lower preference scores (2.933 on average) compared with the others (5.518 on average)[^1]. Those videos also have an affective characteristic in common, i.e., low valence and high arousal. While the average valence and arousal scores are 5.254 and 5.157 for all videos, respectively, the videos showing high misclassification rates received the valence and arousal scores of 3.507 and 5.665 on average, respectively. In summary, it seems that the neural activities for videos inducing low preference, low valence, and high arousal are relatively difficult to classify. One possibility that can explain this tendency is the negative emotion (low valence) of stimuli influences on the classification performance. From previous psychological researches, it was revealed that the negative emotion tend to induce more intensive responses than the positive emotion [@Peeters90], and the bias to the negative emotion probably becomes strong with the high arousal in this experiment. Therefore, neural activities related to the negative emotion may overwhelm the QoE-related neural activities so that the classification performance is degraded.
Moreover, we conducted the same analysis for the score regression for the best case in Table \[table:regresult\]. We found that the regression for the five videos with high misclassification ratios also shows relatively large errors. These videos are included in the bottom 20% in terms of the sum of absolute differences between the ground truths and predicted preference scores.
Conclusion {#sec:conclu}
==========
We have proposed a novel preference prediction approach using EEG based on CNN. We demonstrated significantly improved performance of the proposed method in comparison to previous works for preference prediction. Moreover, we examined various combination of network complexity, feature type, and arrangement of the input matrix.
In the future, other types of brain connectivities will be examined; in particular, an improved version of the phase synchronization may be helpful to enhance the prediction performance. Furthermore, it will be also interesting to consider the characteristics of subjects and videos to obtain the robust performance of QoE recognition.
Acknowledgment {#acknowledgment .unnumbered}
==============
This work was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Korea government (MSIT) (NRF-2016R1E1A1A01943283).
[^1]: This result is not because of different performance for the two classes. The classification accuracies for the low preference and high preference classes are almost the same, i.e., 0.974 and 0.972, respectively.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
Recently, we developed a method for calculating the *lifetime* of the particle in the special situation where there is no potential barrier, as a first step in our efforts to understand the quantum-mechanics of magnetic traps. The toy model that was used in this study was *physically* *unrealistic* because the magnetic field did not obey Laplace’s equation. Here, we study, both classically and quantum-mechanically, the problem of a neutral particle with spin $S$, mass $m$ and magnetic moment $\mu$, moving in two-dimensions in an inhomogeneous physically *realistic* magnetic field given by $$\mathbf{B}=B_{\perp}^{\prime}(x\mathbf{\hat{x}}-y\mathbf{\hat{y}})+B_{0}\mathbf{\hat{z}.}$$ We identify $$K\equiv\sqrt{\dfrac{S^{2}\left( B_{\bot}^{\prime}\right) ^{2}}{\mu
mB_{0}^{3}}},$$ which is the ratio between the precessional frequency of the particle and its vibration frequency, as the relevant parameter of the problem.
Classically, we find that when $\mu$ is antiparallel to $\mathbf{B}$, the particle is trapped provided that $K<\sqrt{4/27}$. We also find that viscous friction, be it translational or precessional, destabilizes the system.
Quantum-mechanically, we study the problem of spin $S=\hbar/2$ particle in the same field. Treating $K$ as a small parameter for the perturbation from the adiabatic Hamiltonian, we find that the lifetime $T_{esc}$ of the particle in its trapped ground-state is $$T_{esc}=\dfrac{T_{vib}}{128\pi^{2}}\exp\left[ \dfrac{2}{K}\right] \text{ ,}$$ where $T_{vib}=2\pi\sqrt{mB_{0}/\mu\left( B_{\bot}^{\prime}\right) ^{2}}$ is the classical period of the particle when placed in the adiabatic potential $V=\mu\left| \mathbf{B}\right| $.
author:
- |
S. Gov[^1] and S. Shtrikman[^2]\
The Department of Electronics,\
Weizmann Institute of Science,\
Rehovot 76100, Israel
- |
H. Thomas\
The Department of Physics and Astronomy,\
University of Basel,\
CH-4056 Basel, Switzerland
title: 'Magnetic Trapping of Neutral Particles: A Study of a Physically Realistic Model. '
---
Introduction.\[intro\]
======================
Magnetic traps for neutral particles.\[traps\]
----------------------------------------------
Recently there has been rapid progress in techniques for trapping samples of neutral atoms at elevated densities and extremely low temperatures. The development of magnetic and optical traps for atoms has proceeded in parallel in recent years, in order to attain higher densities and lower temperatures [@t1; @t2; @t3; @t4; @t6]. We should note here that neutral traps have been around much longer than their realizations for neutral atoms might suggest, and the seminal papers for neutral trapping as applied to neutrons and plasmas date from the sixties and seventies. Many of these papers are referenced by the authors of Refs.[@t1; @t2; @t3]. In this paper we concentrate on the study of *magnetic* traps. Such traps exploit the interaction of the magnetic moment of the atom with the inhomogeneous magnetic field to provide spatial confinement.
Microscopic particles are not the only candidates for magnetic traps. In fact, a vivid demonstration of trapping large scale objects is the hovering magnetic top[@levitron; @ucas; @harrigan; @patent]. This ingenues magnetic device, which hovers in mid-air for about 2 minutes, has been studied recently by several authors [@edge; @Berry; @bounds; @simon; @dynamic].
Qualitative description.\[desc\]
--------------------------------
The physical mechanism underlying the operation of magnetic traps is the adiabatic principle. The common way to describe their operation is in terms of *classical* mechanics: As the particle is released into the trap, its magnetic moment points antiparallel to the direction of the magnetic field. While inside the trap, the particle experiences lateral oscillations $\omega_{vib}$ which are slow compared to its precession $\omega_{prec}$. Under this condition the spin of the particle may be considered as experiencing a *slowly* rotating magnetic field. Thus, the spin precesses around the *local* direction of the magnetic field $\mathbf{B}$ (adiabatic approximation) and, on the average, its magnetic moment $\mathbf{\mu}$ points *antiparallel* to the local magnetic field lines. Hence, the magnetic energy, which is normally given by $-\mathbf{\mu}\cdot\mathbf{B}$, is now given (for small precession angle) by $\mu\left|
\mathbf{B}\right| $. Thus, the overall effective potential seen by the particle is $$V_{eff}\simeq\mu\left| \mathbf{B}\right| . \label{energy}$$ In the adiabatic approximation, the spin degree of freedom is rigidly coupled to the translational degrees of freedom, and is already incorporated in Eq.(\[energy\]). Thus, under the adiabatic approximation, the particle may be considered as having only translational degrees of freedom. When the strength of the magnetic field possesses a* minimum*, the effective potential becomes attractive near that minimum and the whole apparatus acts as a trap.
As mentioned above, the adiabatic approximation holds whenever $\omega
_{prec}\gg\omega_{vib}$. As $\omega_{prec}$ is inversely proportional to the spin, this inequality can be satisfied provided that the spin of the particle is small enough. If, on the other hand, the spin of the particle is too large, it cannot respond fast enough to the changes of the direction of the magnetic field. In this limit $\omega_{prec}\ll\omega_{vib}$, the spin has to be considered as fixed in space and, according to Earnshaw’s theorem[@earnshaw], becomes unstable against *translations*. Note also that $\omega_{prec}$ is proportional to the field $\left| \mathbf{B}\right| $. To prevent $\omega_{prec}$ of becoming too small, resulting in spin-flips (Majorana transitions), most magneto-static traps include a *bias* field, so that the effective potential $V_{eff}$ possesses a *nonvanishing* minimum.
The purpose and structure of this paper.\[purp\]
------------------------------------------------
The discussion of magnetic traps in the literature is, almost entirely, done in terms of *classical* mechanics. In microscopic systems, however, quantum effects become dominant, and in these cases *quantum mechanics* is suited for the description of the trap [@quant]. An even more interesting issue is the understanding of how the classical and quantum descriptions of a *given* system are related.
As a first step in our efforts to understand the quantum-mechanics of magnetic traps, we recently developed a method for calculating the *lifetime* of the particle in the special situation where there is no potential barrier[@life1d]. The toy model that was used in this study consisted of a particle with spin, having only a single translational degree of freedom, in the presence of a 1D inhomogeneous magnetic field. We found that the trapped state of the particle decays with a lifetime given by $\sim1/\left( \sqrt
{K}\omega_{vib}\right) \exp\left( 2/K\right) $ where $K=\omega_{vib}/\omega_{prec}$. The field that was used in this model was not divergenceless, and in this sense, the model is *unrealistic*. The next step, presented in this paper, is to study, both classically and quantum-mechanically, the case of a particle with spin, having *two* translational degree of freedom, in the presence of a *physically* *realistic* (i.e. divergenceless) inhomogeneous magnetic field. This model is reminiscent of a Ioffe-Pritchard trap[@t2; @ife], but without the axial translational degree of freedom. We neglect the effect of interactions between the particles in the trap and so we analyze the dynamics of a *single* particle inside the trap.
The structure of this paper is as follows: In Sec.(\[def\]) we start by defining the system we study, together with useful parameters that will be used throughout this paper. Next, we carry out a classical analysis of the problem in Sec.(\[class\]). Here, we find two stationary solutions for the particle inside the trap. One of them corresponds to a state whose spin is *parallel* to the direction of the magnetic field whereas the other one corresponds to a state whose spin is *antiparallel* to that direction. When considering the dynamical stability of these solutions, we find that only the *antiparallel* stationary solution is stable. We also study the same problem but with viscous friction added, and arrive at the result that friction *destabilizes* the system. In Sec.(\[quant\]) we reconsider the problem, from a quantum-mechanical point of view. Here, we also find states that refer to *parallel* and *antiparallel* orientations of the spin, one of them being bound while the other one unbounded. In this case, however, these two states are *coupled* due to the inhomogeneity of the field and we move on to calculate the *lifetime* of the bound state. Finally, in Sec.(\[dis\]) we compare the results of the classical analysis with these of the quantum analysis and comment on their implications to practical magnetic traps.
Description of the problem.\[def\]
==================================
We consider a particle of mass $m$, magnetic moment $\mu$ and intrinsic spin $S$ (aligned with $\mu$) moving in an inhomogeneous magnetic field $\mathbf{B}$ given by $$\mathbf{B=}B_{0}\mathbf{\hat{z}+}B_{\bot}^{\prime}\left( x\mathbf{\hat{x}-}y\mathbf{\hat{y}}\right) \text{.} \label{d0}$$ This field possesses a nonzero minimum of amplitude at the origin, which is the essential part of the trap. The Hamiltonian for this system is$$H=\dfrac{p^{2}}{2m}-\mathbf{\mu\cdot B} \label{d0.1}$$ where $p$ is the momentum of the particle.
The Hamiltonian is invariant under a group of operations consisting of a rotation of position space about the $z$-axis by an arbitrary angle $\gamma$ combined with a rotation of spin space about the $S_{z}$-axis by the *opposite* angle $-\gamma$. Since the generators of these two rotations are the $z$-components of orbital angular momentum $L_{z}=xp_{y}-yp_{x}$ and of spin angular momentum $S_{z}$, respectively, this symmetry gives rise to a constant of motion, $$\Lambda=L_{z}-S_{z}=\mathrm{const}. \label{d0.2}$$ Since the magnetic field $\mathbf{B}$ does not depend on $z$, the motion along the $z$-direction is trivial. Therefore, we restrict ourselves to studying the motion in the $(x,y)$-plane.
We define $\omega_{prec}$ as the precessional frequency of the particle when it is at the origin $(x=0,y=0)$. Since at that point the magnetic field is $\mathbf{B=}B_{0}\mathbf{\hat{z}}$ we find that $$\omega_{prec}\equiv\dfrac{\mu B_{0}}{S}\text{.} \label{d1}$$ Next, we define $\omega_{vib}$ as the small-amplitude vibrational frequency of the particle when it is placed in the adiabatic potential field given by $$V(x)=\mu\left| \mathbf{B}(x)\right| =\mu B_{0}\left( 1+\dfrac{1}{2}\left(
\dfrac{B_{\bot}^{\prime}}{B_{0}}\right) ^{2}\left( x^{2}+y^{2}\right)
\right) +\mathcal{O}\left( x^{4},x^{2}y^{2},y^{4}\right) .$$ For this potential we have $$k_{x}=k_{y}=\left. \dfrac{\partial^{2}V}{\partial x^{2}}\right| _{@x=0}=\mu\dfrac{\left( B_{\bot}^{\prime}\right) ^{2}}{B_{0}}\text{,}$$ and therefore $$\omega_{vib}\equiv\sqrt{\dfrac{k_{x}}{m}}=\sqrt{\dfrac{\left( B_{\bot
}^{\prime}\right) ^{2}\mu}{mB_{0}}}\text{.} \label{d2}$$ We also define the ratio between $\omega_{vib}$ and $\omega_{prec}$, $$K\equiv\dfrac{\omega_{vib}}{\omega_{prec}}=\sqrt{\dfrac{S^{2}(B_{\bot}^{\prime})^{2}}{\mu mB_{0}^{3}}}\text{.} \label{d3}$$ This will be our ‘measure of adiabaticity’. It is clear that as $K$ becomes smaller and smaller, the adiabatic approximation becomes more and more accurate. Note that when the bias field $B_{0}$ vanishes, $K$ becomes infinite, and the adiabatic approximation fails. We will later show that, under this condition, the system become *unstable* against spin flips, which is in agreement with our discussion at the beginning. This shows that the introduction of the bias field $B_{0}$, is *essential* to the operation of the trap with regard to spin-flips. Note also that $K$ is the only possibility to form a non-dimensional quantity (up to an arbitrary power) out of the parameters of the system. The value of $K$ therefore, completely determines the behavior of the system.
Classical analysis.\[class\]
============================
The stationary solutions.\[stat\]
---------------------------------
We denote by $\mathbf{\hat{n}}$ a unit vector in the direction of the spin (and the magnetic moment). Thus, the equations of motion for the center of mass of the particle are $$\begin{aligned}
m\dfrac{d^{2}x}{dt^{2}} & =\mu\dfrac{\partial}{\partial x}\left(
\mathbf{\hat{n}\cdot B}\right) \label{c1}\\
m\dfrac{d^{2}y}{dt^{2}} & =\mu\dfrac{\partial}{\partial y}\left(
\mathbf{\hat{n}\cdot B}\right) ,\nonumber\end{aligned}$$ and the evolution of its spin is determined by $$S\dfrac{d\mathbf{\hat{n}}}{dt}=\mu\mathbf{\hat{n}\times B}\text{.} \label{c2}$$ It is straightforward to check that the quantity $\Lambda=L_{z}-S_{z}$ is indeed conserved.
The two equilibria solutions to Eqs.(\[c1\]) and (\[c2\]) are $$\mathbf{\hat{n}}(t)=\mp\mathbf{\hat{z}} \label{c3}$$ with$$\begin{aligned}
x(t) & =0\\
y(t) & =0\end{aligned}$$ representing a motionless particle at the origin with its magnetic moment (and spin) pointing *antiparallel* ($\mathbf{\hat{n}}(t)=-\mathbf{\hat{z}}$) to the direction of the field at that point and a similar solution but with the magnetic moment pointing *parallel* to the direction of the field ($\mathbf{\hat{n}}(t)=+\mathbf{\hat{z}}$).
Stability of the solutions.
---------------------------
To check the stability of these solutions we now add first-order perturbations. We set $$\begin{aligned}
\mathbf{\hat{n}(}t\mathbf{)} & =\mathbf{\mp\hat{z}+}\epsilon_{x}(t)\mathbf{\hat{x}+}\epsilon_{y}(t)\mathbf{\hat{y}}\label{c4}\\
x(t) & =0+\delta x(t)\nonumber\\
y(t) & =0+\delta y(t),\nonumber\end{aligned}$$ (note that, to first order, the perturbation $\delta\mathbf{\hat{n}}=\epsilon_{x}(t)\mathbf{\hat{x}+}\epsilon_{y}(t)\mathbf{\hat{y}}$ is taken to be *orthogonal* to the value of $\mathbf{\hat{n}}$ for the stationary solution $\mathbf{\hat{n}}_{0}=\mathbf{\mp\hat{z}}$, since $\mathbf{\hat{n}}$ is a unit vector) substitute these in Eqs.(\[c1\]) and (\[c2\]), and retain only first-order terms. We find that the resulting equations for $\delta x(t)$, $\delta y(t)$, $\epsilon_{x}(t)$ and $\epsilon_{y}(t)$ are $$\begin{aligned}
\dfrac{d^{2}\delta x}{dt^{2}} & =\dfrac{\mu B_{\bot}^{\prime}}{m}\epsilon_{x}\label{c5}\\
\dfrac{d^{2}\delta y}{dt^{2}} & =-\dfrac{\mu B_{\bot}^{\prime}}{m}\epsilon_{y}\nonumber\\
\dfrac{d\epsilon_{x}}{dt} & =\dfrac{\mu}{S}\left( \mp B_{\bot}^{\prime
}\delta y+B_{0}\epsilon_{y}\right) \nonumber\\
\dfrac{d\epsilon_{y}}{dt} & =\dfrac{\mu}{S}\left( \mp B_{\bot}^{\prime
}\delta x-B_{0}\epsilon_{x}\right) \text{.}\nonumber\end{aligned}$$ The normal modes of the system transform as the irreducible representations of the symmetry group. The 4-dimensional linear space spanned by the deviations $(\delta x,\delta y,\epsilon_{x},\epsilon_{y})$ from the stationary state carries the irreducible representations $\Gamma_{+}$ with characters $e^{-i\gamma}$ and $\Gamma_{-}$ with characters $e^{+i\gamma}$, and may thus be decomposed into the two 2-dimensional invariant subspaces transforming as $\Gamma_{+}$ and $\Gamma_{-}$, respectively. These subspaces are spanned by the circular position coordinates and precessional spin coordinates$$\begin{aligned}
\Gamma_{+}:\; & (\rho_{+}=\delta x+i\delta y,\epsilon_{-}=\epsilon
_{x}-i\epsilon_{y});\label{c5.1}\\
\Gamma_{-}:\; & (\rho_{-}=\delta x-i\delta y,\epsilon_{+}=\epsilon
_{x}+i\epsilon_{y}). \label{c5.2}$$ Thus, the normal modes consist of a circular motion in the $(x,y)$-plane coupled to a precession of the spin vector in the *opposite* sense.
Indeed, after introducing the $(\rho_{\pm},\epsilon_{\mp})$-coordinates into Eqs.(\[c5\]), this set of four equations decomposes into one pair of equations for $(\rho_{+},\epsilon_{-})$ and another pair for $(\rho
_{-},\epsilon_{+})$. We now look for oscillatory (stable) solutions of these equations and set $$\rho_{\pm}=\rho_{\pm,0}e^{-i\omega t},\quad\epsilon_{\pm}=\epsilon_{\pm
,0}e^{-i\omega t}. \label{c7}$$ This yields the algebraic equations $$\Gamma_{+}:\;\left(
\begin{array}
[c]{cc}\omega^{2} & \omega_{vib}^{2}B_{0}/B_{\bot}^{\prime}\\
\pm i\omega_{prec}B_{\bot}^{\prime}/B_{0} & i(\omega+\omega_{prec})
\end{array}
\right) \cdot\left(
\begin{array}
[c]{l}\rho_{+,0}\\
\epsilon_{-,0}\end{array}
\right) =\left(
\begin{array}
[c]{l}0\\
0
\end{array}
\right) , \label{c8.1}$$ $\allowbreak$$$\Gamma_{-}:\;\left(
\begin{array}
[c]{cc}\omega^{2} & \omega_{vib}^{2}B_{0}/B_{\bot}^{\prime}\\
\mp i\omega_{prec}B_{\bot}^{\prime}/B_{0} & i(\omega-\omega_{prec})
\end{array}
\right) \cdot\left(
\begin{array}
[c]{l}\rho_{-,0}\\
\epsilon_{+,0}\end{array}
\right) =\left(
\begin{array}
[c]{l}0\\
0
\end{array}
\right) . \label{c8.2}$$ These equations have non-trivial solutions whenever the determinant of either of the two matrices vanishes. This yields the secular equations $$\begin{aligned}
\Gamma_{+} & :\;K\left( \frac{\omega}{\omega_{vib}}\right) ^{3}+\left(
\frac{\omega}{\omega_{vib}}\right) ^{2}\mp1=0,\label{c11.1}\\
\Gamma_{-} & :\;K\left( \frac{\omega}{\omega_{vib}}\right) ^{3}-\left(
\frac{\omega}{\omega_{vib}}\right) ^{2}\pm1=0, \label{c11.2}$$ which determine the eigenfrequencies $\omega$ of the various modes. Since the system has three degrees of freedom, we expect to have three normal modes. Indeed, when $\omega$ is a solution of the first equation, then $-\omega$ is a solution of the second equation. We define the mode frequencies in Eq.(\[c11.1\]) to be positive (or, in the case of complex $\omega$, to have positive real part); the negative $\omega$-values are needed to construct real solutions. Then, the $\Gamma_{+}$-modes describe vibrational motions turning counter-clockwise coupled to spin precessions turning clockwise, i.e., opposite to the natural spin precession, and the $\Gamma_{-}$-modes describe vibrational motions turning clockwise coupled to spin precessions turning counter-clockwise, i.e., in the same sense as the natural spin precession.
When the *lower* sign is taken in Eqs.(\[c4\]), corresponding to a spin *parallel* to the magnetic field, $\mathbf{\hat{n}}=+\mathbf{\hat{z}}$, we find two $\Gamma_{+}$-modes with complex-conjugate mode frequencies. Thus, one of the mode frequencies possesses a positive imaginary part, indicating that the spin-up state is unstable for *any* value of $K$. We therefore concentrate on the stationary state with the spin *antiparallel* to the magnetic field, $\mathbf{\hat{n}}=-\mathbf{\hat{z}}$, corresponding to the *upper* sign in Eqs.(\[c4\]). For the spin-down state we find one $\Gamma_{+}$-mode for any value of $K$, and for $K$ smaller than a critical value $K_{c}=\sqrt{4/27}$ two $\Gamma_{-}$-modes, with real frequencies, as is shown in Fig.(\[fig1\]).
\[ptb\]
[graph1.eps]{}
For $K\rightarrow0$, the $\Gamma_{+}$-mode, which goes like $\omega
\simeq\omega_{vib}\left( 1-K/2\right) $, and the slower of the $\Gamma_{-}$-modes, which behaves as $\omega\simeq\omega_{vib}\left( 1+K/2\right) $, become degenerate with frequency $\omega=\omega_{vib}$, corresponding to two linear independent purely vibrational modes. With increasing $K$, the coupling between translational motion and spin motion lifts the degeneracy and gives rise to an increasing spin component, which leads to a decrease of the $\Gamma_{+}$ mode frequency and an increase of the slow $\Gamma_{-}$ mode frequency. The fastest of the $\Gamma_{-}$-modes, on the other hand, which for $K\rightarrow0$ is a pure spin precession mode with frequency $\omega
=\omega_{vib}/K=\omega_{prec}$, acquires with increasing $K$ an increasing vibrational component, leading to a decrease of the mode frequency. At $K=K_{c}$, it becomes degenerate with the slower $\Gamma_{-}$-mode, with mode frequency $\omega=\sqrt{3}\omega_{vib}$. For $K>K_{c}$, the two $\Gamma_{-}$-modes have complex-conjugate frequencies. Fig.(\[fig1\]) shows the real and imaginary parts of the frequencies $\omega$ of the three modes as a function of $K$.
The dependence of the modes on the parameter $K$ is shown more explicitly by the form of the eigenvectors. The general form of these is given by$$\left(
\begin{array}
[c]{c}\rho_{\pm,0}\\
\epsilon_{\mp,0}\end{array}
\right) =\left(
\begin{array}
[c]{c}\dfrac{B_{0}}{B_{\perp}^{\prime}}\dfrac{\omega_{vib}}{\omega}\\
-\dfrac{\omega}{\omega_{vib}}\end{array}
\right) A_{\pm}\text{ , } \label{c12.1}$$ where $A_{\pm}$ are dimensionless amplitude parameters.
The excitation energy of the modes.
-----------------------------------
The excitation energy of a given mode $\xi$ is defined as the difference between the energy of the mode and the energy of the stationary state,$$\xi=-\mu\mathbf{\hat{n}}\cdot\mathbf{B+}\dfrac{1}{2}m\left[ \left(
\dfrac{d\delta x}{dt}\right) ^{2}+\left( \dfrac{d\delta y}{dt}\right)
^{2}\right] -\mu B_{0}. \label{ec1.1}$$ Note that the energy contains bilinear terms in the coordinates and hence, one cannot neglect the $\hat{z}$-component of the spin. Instead, one must set$$\mathbf{\hat{n}\cdot\hat{z}=-}\sqrt{1-\left( \epsilon_{x}^{2}+\epsilon
_{y}^{2}\right) }\mathbf{\simeq}-\left( 1-\dfrac{1}{2}\left( \epsilon
_{x}^{2}+\epsilon_{y}^{2}\right) \right) .$$ Thus, the correct expression of the energy for small amplitudes is$$\xi\simeq-\mu\left[ \dfrac{1}{2}\left( \epsilon_{x}^{2}+\epsilon_{y}^{2}\right) B_{0}+B_{\perp}^{\prime}\left( \delta x\epsilon_{x}-\delta
y\epsilon_{y}\right) \right] \mathbf{+}\dfrac{1}{2}m\left[ \left(
\dfrac{d\delta x}{dt}\right) ^{2}+\left( \dfrac{d\delta y}{dt}\right)
^{2}\right] . \label{ec1.2}$$ The deviations $\delta x(t),\delta y(t),\epsilon_{x}(t),\epsilon_{y}(t)$ from the stationary state belonging to the normal modes are given in real form by $$\delta x(t)=\frac{1}{2}\rho_{\pm,0}e^{-i\omega t}+\mathrm{c.c.},\quad\delta
y(t)=\pm\frac{1}{2i}\rho_{\pm,0}e^{-i\omega t}+\mathrm{c.c.}, \label{ec2}$$$$\epsilon_{x}(t)=\frac{1}{2}\epsilon_{\mp,0}e^{-i\omega t}+\mathrm{c.c.},\quad\epsilon_{y}(t)=\mp\frac{1}{2i}\epsilon_{\mp,0}e^{-i\omega
t}+\mathrm{c.c.}. \label{ec3}$$ With the help of Eq.(\[c12.1\]), one obtains $$\xi=\mu B_{0}\frac{3\omega_{vib}^{2}-\omega^{2}}{\omega_{vib}^{2}}|A_{\pm
}|^{2}. \label{ec4}$$ From this result we conclude that for $0<K<\sqrt{4/27}$, the excitation energy of the vibrational modes, for which $\omega^{2}<3\omega_{vib}^{2}$, is *positive* while the excitation energy of the precessional mode, satisfying $\omega^{2}>3\omega_{vib}^{2}$, is always *negative*. At the point $K=\sqrt{4/27}$, where the clockwise vibrational mode and the precessional mode coalesce, the excitation energy vanishes. We will further refer to these observations in the following section.
The effect of viscous friction.\[fric\]
---------------------------------------
When friction is introduced into the system, the equations of motion become $$\begin{aligned}
m\dfrac{d^{2}x}{dt^{2}} & =\mu\dfrac{\partial}{\partial x}\left(
\mathbf{\hat{n}\cdot B}\right) -r_{t}\dfrac{dx}{dt}\label{fric1}\\
m\dfrac{d^{2}y}{dt^{2}} & =\mu\dfrac{\partial}{\partial y}\left(
\mathbf{\hat{n}\cdot B}\right) -r_{t}\dfrac{dy}{dt}\nonumber\end{aligned}$$ and $$S\dfrac{d\mathbf{\hat{n}}}{dt}=\mu\mathbf{\hat{n}\times B-}r_{p}\mathbf{\hat{n}\times}\dfrac{d\mathbf{\hat{n}}}{dt}\text{,} \label{fric2}$$ where $r_{t}$ and $r_{p}$ are translational and precessional friction coefficients, respectively. The second term in the right-hand side of Eq.(\[fric2\]) is the spin-damping contributed by the *change* in the direction of the spin from $\mathbf{\hat{n}}$ to $\mathbf{\hat{n}+}d\mathbf{\hat{n}}$. Since, by definition, $\mathbf{\hat{n}}$ is a unit vector, it must point perpendicular to $d\mathbf{\hat{n}}$. Thus, $\Omega_{\perp}=\left| d\mathbf{\hat{n}/}dt\right| $ is the angular velocity associated with the change of $\mathbf{\hat{n}}$. Since the direction of $\Omega_{\perp}$ must be perpendicular to both $d\mathbf{\hat{n}}$ and $\mathbf{\hat{n}}$ we form the cross product $\mathbf{\Omega}_{\bot
}=\mathbf{\hat{n}\times}\left( d\mathbf{\hat{n}/}dt\right) $ which incorporates both the correct value and the right direction. Multiplying $\mathbf{\Omega}_{\bot}$ by $r_{p}$ yields the spin-damping term.
To first order in $r_{r}$ and $r_{t}$ the secular equation in this case is given by$\allowbreak$ $$\begin{aligned}
0 & =-K^{2}\omega_{n}^{6}+\omega_{n}^{4}-2\omega_{n}^{2}+1+2iK\omega_{n}^{5}\dfrac{r_{p}}{S}-2iK^{3}\omega_{n}^{5}\dfrac{r_{t}}{S}\left( \dfrac
{B_{0}}{B_{\bot}^{\prime}}\right) ^{2}\label{fric3.1}\\
& +2iK\omega_{n}^{3}\dfrac{r_{t}}{S}\left( \dfrac{B_{0}}{B_{\bot}^{\prime}}\right) ^{2}-2iK\omega_{n}^{3}\dfrac{r_{p}}{S}-2iK\omega_{n}\dfrac{r_{t}}{S}\left( \dfrac{B_{0}}{B_{\bot}^{\prime}}\right) ^{2},\nonumber\end{aligned}$$
where we defined$$\omega_{n}\equiv\dfrac{\omega}{\omega_{vib}}$$ to make the expression simple. Let $\omega_{n,0}$ be the eigenfrequencies $\omega_{n}$ of the frictionless problem, given by Eq.(\[c11.2\]). When adding small friction to the problem, the eigenfrequencies will change by a small amount $\delta\omega_{n}$. We find an approximate expression for $\delta\omega_{n}$ by expanding Eq.(\[fric3.1\]) around $\omega_{n,0}$ to first order in $\delta\omega_{n}$ and making use of Eq.(\[c11.2\]). This gives $$\delta\omega_{n}=\dfrac{iK}{S}\left( \dfrac{r_{p}\omega_{n,0}^{4}+r_{t}\left( B_{0}/B_{\perp}^{\prime}\right) ^{2}}{\omega_{n,0}^{2}-3}\right) +\mathcal{O}\left( r_{t}^{2},r_{t}r_{p},r_{p}^{2}\right) .
\label{fric4}$$
Eq.(\[fric4\]) has an interesting consequence: The numerator in Eq.(\[fric4\]) is positive for all three modes while the denominator is negative for the two vibrational modes and positive for the precessional mode. We therefore conclude that friction, either translational or precessional, stabilizes the vibrational modes and, simultaneously, destabilizes the precessional mode. The system all together becomes of course, *unstable*.
The fact that spin damping leads to an exponential growth of the precessional mode is no surprise in view of its negative excitation energy. Also, the exponential decay of the vibrational modes due to translational friction is to be expected on account of their positive excitation energy. What *is* important is the fact that due to the coupling between translation and precession, *translational* friction causes an exponential growth of the *precessional* mode, with a growth time which, compared to the effect of spin damping, is smaller by a factor of $r_{t}K^{2}S^{2}/\mu mr_{p}B_{0}^{2}$ in the limit of small $K$.
Quantum-mechanical analysis.\[quant\]
=====================================
The Hamiltonian and its diagonalized form.\[ham\]
-------------------------------------------------
In this section we consider the problem of a neutral particle with spin half ( $S=\hbar/2)$ in a 2D inhomogeneous magnetic field from a quantum-mechanical point of view. Unlike the classical analysis, in which the derivation was valid for any value of the adiabaticity parameter $K$, we concentrate here on the behavior of the system when $K$ is *small*. We choose to analyze the case of a spin half particle because this case already shows the essentials of the quantum-mechanical problem. Note also that, quantum mechanically, the magnetic moment $\mu$ and the spin $S$ of a particle are related by$$\mu=\gamma S,$$ where $\gamma$ is the gyromagnetic ratio of the particle. Setting $\mu=\gamma
S$ and $S=\hbar/2$ in Eq.(\[d3\]) gives$$K=\sqrt{\dfrac{\hbar(B_{\bot}^{\prime})^{2}}{2\gamma mB_{0}^{3}}}.$$
Now, it is convenient to express the spatial dependence of the magnetic field in polar coordinates $r=\sqrt{x^{2}+y^{2}}$ and $\phi=\arctan\left(
y/x\right) $. We also denote by $B$ the amplitude of $\mathbf{B}$, by $\theta$-its direction with respect to the $\mathbf{\hat{z}}$ axis and by $\varphi$ the angle between the projection of $\mathbf{B}$ onto the $x$-$y$ plane and the $\hat{x}$-axis. Thus, Eq.(\[d0\]) is rewritten as $$\mathbf{B}=B\left[ \sin\theta\cos\varphi\mathbf{\hat{x}+}\sin\theta
\sin\varphi\mathbf{\hat{y}}+\cos\theta\mathbf{\hat{z}}\right] \label{h1}$$ where $$\begin{aligned}
B & =B_{0}\sqrt{1+\left( \dfrac{B_{\bot}^{\prime}}{B_{0}}\right) ^{2}r^{2}}\text{ ,}\label{h1.1}\\
\theta & =\arctan\left( \dfrac{B_{\bot}^{\prime}r}{B_{0}}\right)
,\nonumber\\
\varphi & =\arctan\left( \dfrac{B_{y}}{B_{x}}\right) =-\arctan\left(
\dfrac{y}{x}\right) =-\phi.\nonumber\end{aligned}$$ Thus, $B$ and $\theta$ depends only on $r$ whereas $\varphi$ depends (linearly) only on $\phi$.
The time-independent Schrödinger equation for this system is $$\left[ -\frac{\hbar^{2}}{2m}\nabla^{2}-\mu B\left( \sin\theta\cos\varphi
\hat{\sigma}_{x}\mathbf{+}\sin\theta\sin\varphi\hat{\sigma}_{y}+\cos\theta
\hat{\sigma}_{z}\right) \right] \Psi^{\prime}\left( r,\phi\right)
=E\Psi^{\prime}\left( r,\phi\right) \label{h5}$$ where $\hat{\sigma}_{x}$, $\hat{\sigma}_{y}$ and $\hat{\sigma}_{z}$ are the Pauli matrices given by $$\begin{array}
[c]{ccc}\hat{\sigma}_{x}=\left(
\begin{array}
[c]{cc}0 & 1\\
1 & 0
\end{array}
\right) & \hat{\sigma}_{y}=\left(
\begin{array}
[c]{cc}0 & -i\\
i & 0
\end{array}
\right) & \hat{\sigma}_{z}=\left(
\begin{array}
[c]{cc}1 & 0\\
0 & -1
\end{array}
\right) ,
\end{array}$$ $E$ is the eigenenergy and $\Psi^{\prime}$ is the two-components spinor $$\Psi^{\prime}=\left(
\begin{array}
[c]{c}\psi_{\uparrow}^{\prime}\left( r,\phi\right) \\
\psi_{\downarrow}^{\prime}\left( r,\phi\right)
\end{array}
\right) . \label{h5.1}$$ In matrix form Eq.(\[h5\]) becomes $$\left( H_{K}+H_{M}\right) \left(
\begin{array}
[c]{c}\psi_{\uparrow}^{\prime}\left( r,\phi\right) \\
\psi_{\downarrow}^{\prime}\left( r,\phi\right)
\end{array}
\right) =E\left(
\begin{array}
[c]{c}\psi_{\uparrow}^{\prime}\left( r,\phi\right) \\
\psi_{\downarrow}^{\prime}\left( r,\phi\right)
\end{array}
\right) \label{h6.0}$$ where $H_{K}$ and $H_{M}$, given by $$\begin{aligned}
H_{K} & \equiv-\dfrac{\hbar^{2}}{2m}\nabla^{2}\label{h6.1}\\
H_{M} & \equiv-\mu B\left(
\begin{array}
[c]{cc}\cos\theta & \sin\theta e^{-i\varphi}\\
\sin\theta e^{i\varphi} & -\cos\theta
\end{array}
\right) ,\nonumber\end{aligned}$$ are the kinetic part and the magnetic part of the Hamiltonian $H$, respectively.
In order to diagonalize the magnetic part of the Hamiltonian, we make a local *passive* transformation of coordinates on the wavefunction such that the spinor is expressed in a new coordinate system whose $\mathbf{\hat{z}}$ axis coincides with the direction of the magnetic field at the point $\left(
r,\phi\right) $. We denote by $R\left( r,\phi\right) $ the required transformation and set $\Psi=R\Psi^{\prime}$. Thus, $\Psi$ represent *the same* direction of the spin as before the transformation but using the *new* coordinate system. The Hamiltonian in this newly defined system is clearly given by $RHR^{-1}$. In the case of the magnetic field given in Eqs.(\[h1\]) and (\[h1.1\]) the required transformation is accomplished by using the three Euler angles: First, we perform a rotation through an angle $\varphi$ around the $\hat{z}$ axis. Second, we make a rotation through an angle $\theta$ around the *new* position of the $\hat{y}$ axis. At the end of this process the new $\hat{z}$ axis coincide with the direction of the magnetic field. Now the value of the last Euler angle, which is a rotation around the new $\hat{z}$ axis, has no effect on this axis. For simplicity we choose this angle to be $0$. Thus, the representation of the complete transformation for spin half particle is given by [@rot] $$R=\exp\left[ i\dfrac{\theta}{2}\hat{\sigma}_{y}\right] \exp\left[
i\dfrac{\varphi}{2}\hat{\sigma}_{z}\right] ,$$ while its inverse is given by $$R^{-1}=\exp\left[ -i\dfrac{\varphi}{2}\hat{\sigma}_{z}\right] \exp\left[
-i\dfrac{\theta}{2}\hat{\sigma}_{y}\right] .$$ It is easily verified that the transformation indeed diagonalizes the magnetic part of the Hamiltonian as $$RH_{M}R^{-1}=-\mu B\hat{\sigma}_{z}.$$ For the kinetic part we find, after some algebra, that $$RH_{K}R^{-1}=-\dfrac{\hbar^{2}}{2m}\left[
\begin{array}
[c]{c}\mathbf{\nabla}^{2}-\dfrac{1}{4}\left( \dfrac{d\theta}{dr}\right)
^{2}-\dfrac{1}{4r^{2}}\\
+\dfrac{i}{r^{2}}\hat{\sigma}_{z}\cos\theta\dfrac{\partial}{\partial\phi}\\
-i\left[ \left( \dfrac{d\theta}{dr}\right) \dfrac{\partial}{\partial
r}+\dfrac{1}{2}\dfrac{d^{2}\theta}{dr^{2}}+\dfrac{1}{2r}\dfrac{d\theta}{dr}\right] \hat{\sigma}_{y}\\
-\dfrac{i}{r^{2}}\hat{\sigma}_{x}\sin\theta\dfrac{\partial}{\partial\phi}\end{array}
\right]$$
Thus, the Hamiltonian of the system in the rotated frame may be written as $$H=H_{diag}+H_{int} \label{h7}$$ where $$\begin{aligned}
H_{diag} & =-\dfrac{\hbar^{2}}{2m}\left[ \mathbf{\nabla}^{2}-\dfrac{1}{4}\left( \dfrac{d\theta}{dr}\right) ^{2}-\dfrac{1}{4r^{2}}+\dfrac{i}{r^{2}}\hat{\sigma}_{z}\cos\theta\dfrac{\partial}{\partial\phi}\right] -\mu
B\hat{\sigma}_{z}\label{h7.01}\\
H_{int} & =-\dfrac{\hbar^{2}}{2m}\left\{ -i\hat{\sigma}_{y}\left[ \left(
\dfrac{d\theta}{dr}\right) \dfrac{\partial}{\partial r}+\dfrac{1}{2}\dfrac{d^{2}\theta}{dr^{2}}+\dfrac{1}{2r}\dfrac{d\theta}{dr}\right]
-\dfrac{i}{r^{2}}\hat{\sigma}_{x}\sin\theta\dfrac{\partial}{\partial\phi
}\right\} .\nonumber\end{aligned}$$ The first part of the Hamiltonian $H_{diag}$ is diagonal. It contains the kinetic part $\sim\mathbf{\nabla}^{2}$, a term whose form is $\mp$ $\mu B$ which is to be identified as the adiabatic effective potential and the terms $\sim1/r^{2},ir^{-2}\hat{\sigma}_{z}\partial/\partial\phi$ which appear due to the rotation. The second part of the Hamiltonian $H_{int}$ contains only non-diagonal components. These will be shown to be of order $\mathcal{O}\left( K\right) $ and hence may be regarded as a small perturbation. We proceed to find the eigenstates of $H_{diag}$.
Stationary states of $H_{diag}$.\[diag\]
----------------------------------------
Since $H_{diag}$ is diagonal, the two spin states of the wavefunction are decoupled. We then seek a solution of the form $$\Psi_{\downarrow}=\left(
\begin{array}
[c]{c}0\\
\psi_{\downarrow}(r,\phi)
\end{array}
\right) \text{ ; }E=E_{\downarrow}, \label{h8.01}$$ referred to as the *spin-down* state, and another solution $$\Psi_{\uparrow}=\left(
\begin{array}
[c]{c}\psi_{\uparrow}(r,\phi)\\
0
\end{array}
\right) \text{ ; }E=E_{\uparrow}, \label{h8.02}$$ which we call the *spin-up* state.
The equation for the non-vanishing component of the spin-down state is given by $$\left\{ -\dfrac{\hbar^{2}}{2m}\left[ \mathbf{\nabla}^{2}-\dfrac{1}{4}\left(
\dfrac{d\theta}{dr}\right) ^{2}-\dfrac{1}{4r^{2}}-\dfrac{i}{r^{2}}\cos
\theta\dfrac{\partial}{\partial\phi}\right] +\mu B\right\} \psi_{\downarrow
}=E_{\downarrow}\psi_{\downarrow}, \label{h8.1}$$ whereas the equation for the non-vanishing component of the spin-up state is $$\left\{ -\dfrac{\hbar^{2}}{2m}\left[ \mathbf{\nabla}^{2}-\dfrac{1}{4}\left(
\dfrac{d\theta}{dr}\right) ^{2}-\dfrac{1}{4r^{2}}+\dfrac{i}{r^{2}}\cos
\theta\dfrac{\partial}{\partial\phi}\right] -\mu B\right\} \psi_{\uparrow
}=E_{\uparrow}\psi_{\uparrow}. \label{h8.2}$$
We now show that in the limit of small $K$ we can neglect the term $\sim\left( d\theta/dr\right) ^{2}$ in both Eq.(\[h8.1\]) and Eq.(\[h8.2\]): We compare the order of magnitude of the term $\mu B$ to that of the term $\hbar^{2}\left( d\theta/dr\right) ^{2}/8m$. Using Eq.(\[h1.1\]) it can be easily shown that the maximum value of $d\theta/dr$ is $B_{\bot}^{\prime}/B_{0}$ whereas the minimum value of $\mu B$ is $\mu
B_{0}$. Thus, $$\dfrac{\left. \mu B\right| _{\text{min}}}{\left( \dfrac{\hbar^{2}}{8m}\left( \dfrac{d\theta}{dr}\right) _{\text{max}}^{2}\right) }=\dfrac{8\mu mB_{0}^{3}}{\left( B_{\bot}^{\prime}\right) ^{2}\hbar^{2}}=\frac{2}{K^{2}},$$ and hence we can neglect the term $\sim(d\theta/dr)^{2}$ when $K$ is small. Furthermore, as we are interested in the solutions near the origin we replace the $\cos\theta$ term by its zeroth-order approximation around $r=0$. We will justify this approximation later. Under these approximations, Eqs.(\[h8.1\]) and (\[h8.2\]) simplify to $$\left\{ -\dfrac{\hbar^{2}}{2m}\left[ \mathbf{\nabla}^{2}-\dfrac{1}{4r^{2}}-\dfrac{i}{r^{2}}\dfrac{\partial}{\partial\phi}\right] +\mu B\right\}
\psi_{\downarrow}=E_{\downarrow}\psi_{\downarrow} \label{h8.3}$$ and $$\left\{ -\dfrac{\hbar^{2}}{2m}\left[ \mathbf{\nabla}^{2}-\dfrac{1}{4r^{2}}+\dfrac{i}{r^{2}}\dfrac{\partial}{\partial\phi}\right] -\mu B\right\}
\psi_{\uparrow}=E_{\uparrow}\psi_{\uparrow}. \label{h8.4}$$
The approximate solutions of these equations is outlined in the next two subsections.
### Stationary spin-down states.\[down\]
Eq.(\[h8.3\]) represents a particle in a symmetric *attractive* potential. If the extent of the wave function is small enough we can expand $B$ in Eq.(\[h1.1\]) to second order in $r$$$B\simeq B_{0}\left[ 1+\dfrac{1}{2}\left( \dfrac{B_{\bot}^{\prime}r}{B_{0}}\right) ^{2}\right] +\mathcal{O}\left( r^{4}\right) , \label{down0}$$ and apply the well-known solution of the harmonic oscillator[@sho] in two dimensions. Under this approximation, Eq.(\[h8.3\]) becomes$$\left[ -\dfrac{\hbar^{2}}{2m}\left( \dfrac{1}{r}\dfrac{\partial}{\partial
r}+\dfrac{\partial^{2}}{\partial r^{2}}-\dfrac{1}{r^{2}}\left( i\dfrac
{\partial}{\partial\phi}+\dfrac{1}{2}\right) ^{2}\right) +\dfrac{\mu\left(
B_{\bot}^{\prime}\right) ^{2}r^{2}}{2B_{0}}\right] \psi_{\downarrow}=\left(
E_{\downarrow}-\mu B_{0}\right) \psi_{\downarrow}. \label{down0.1}$$ We seek a solution whose form is$$\psi_{\downarrow}(r,\phi)=f(r)e^{i\nu\phi} \label{down0.2}$$ and then the equation satisfied by $f\left( r\right) $ is$$-\dfrac{\hbar^{2}}{2m}\left[ \dfrac{1}{r}\dfrac{df}{dr}+\dfrac{d^{2}f}{dr^{2}}-\dfrac{f}{r^{2}}\left( \nu-\dfrac{1}{2}\right) ^{2}\right]
+\dfrac{\mu\left( B_{\perp}^{\prime}\right) ^{2}r^{2}f}{2B_{0}}=\left(
E_{\downarrow}-\mu B_{0}\right) f, \label{down0.3}$$ which is an eigenvalue problem for $f$. The smallest eigenvalue for this problem is obtained by setting$$\nu=\dfrac{1}{2},$$ for which the eigenfunction $f$ is$$f\left( r\right) =D\exp\left[ -\sqrt{\dfrac{\mu m\left( B_{\perp}^{\prime
}\right) ^{2}}{4\hbar^{2}B}}r^{2}\right] =D\exp\left[ -\dfrac{1}{4K}\left(
\dfrac{B_{\perp}^{\prime}r}{B_{0}}\right) ^{2}\right] .$$ Thus, under the harmonic oscillator approximation the down-part of the spin-down state is $$\psi_{\downarrow}=\frac{B_{\perp}^{\prime}}{B_{0}\sqrt{2\pi K}}\exp\left[
-\dfrac{1}{4K}\left( \dfrac{B_{\perp}^{\prime}r}{B_{0}}\right) ^{2}\right]
e^{i\phi/2}\text{ }. \label{down1}$$ where the normalization constant $D$ has been calculated by demanding that$${\displaystyle\int\limits_{0}^{\infty}}
rdr{\displaystyle\int\limits_{0}^{2\pi}}
d\phi\left| \psi_{\downarrow}\right| ^{2}=1,$$ using the definite integral$${\displaystyle\int\limits_{0}^{\infty}}
re^{-ar^{2}}dr=\dfrac{1}{2a}.$$ Note that the extent of this wave function over which it changes appreciably is given by $$\Delta r_{\downarrow}\sim\sqrt{K}\dfrac{B_{0}}{B_{\bot}^{\prime}},
\label{down2}$$ whereas the extent over which $\mu B$ changes significantly (see Eq.(\[h1.1\])) is $$\Delta r_{\mu B}\sim\dfrac{B_{0}}{B_{\bot}^{\prime}}. \label{down3}$$ Thus, the ratio between these two length scales is $$\dfrac{\Delta r_{\downarrow}}{\Delta r_{\mu B}}\sim\sqrt{K}. \label{down4}$$ We therefore conclude that when $K$ is small enough, the harmonic approximation is justified. Note also that $\Delta r_{\mu B}$ is also the typical length of $\cos\theta$. This shows that the substitution of $\cos\theta$ in Eq.(\[h8.1\]) by $1$ is also justified.
The wave function $\psi_{\downarrow}$, given by Eq.(\[down1\]), then represents the lowest possible bound state for this system. This state corresponds to a *trapped* particle. The energy of this state is clearly $$E_{\downarrow}=\mu B_{0}+2\left( \frac{\hbar}{2}\omega_{vib}\right) =\mu
B_{0}\left( 1+2K\right) \simeq\mu B_{0}, \label{down5}$$ while its full spinor representation is $$\Psi_{\downarrow}=\left(
\begin{array}
[c]{c}0\\
\dfrac{B_{\perp}^{\prime}}{B_{0}\sqrt{2\pi K}}\exp\left[ -\dfrac{1}{4K}\left( \dfrac{B_{\perp}^{\prime}r}{B_{0}}\right) ^{2}\right]
e^{i\phi/2}\end{array}
\right) . \label{down6}$$
### Stationary spin-up states.\[up\]
Eq.(\[h8.4\]) describes a particle in a *repulsive* potential. It corresponds to an unbounded state representing an *untrapped* particle. In this case there is a continuum of states, each with its own energy. As we are interested in non-radiative decay, we focus on finding a solution with an energy which is *equal* to the energy found for the trapped state, that is $$E_{\uparrow}=E_{\downarrow}\simeq\mu B_{0}. \label{up0}$$ When evaluating the lifetime in the next section, we compute the matrix element of $H_{int}$ between the states $\psi_{\uparrow}$ and $\psi
_{\downarrow}$. Thus, most of the contribution to this integral comes from the region in $r$ where $\psi_{\downarrow}$ is substantial. According to Eq.(\[down4\]), $\mu B$ changes very little in this range and, as a first approximation, we may take $\cos\theta\simeq1$ and the potential in this region as *uniform*,$$\mu B\simeq\mu B_{0} \label{up0.1}$$ in Eq.(\[h8.4\]). We now set a solution whose form is$$\psi_{\uparrow}(r,\phi)=g(r)e^{i\gamma\phi}.$$ Substituting this, together with Eqs.(\[up0\]) and (\[up0.1\]) into Eq.(\[h8.4\]) gives$$-\dfrac{\hbar^{2}}{2m}\left[ \dfrac{1}{r}\dfrac{dg}{dr}+\dfrac{d^{2}g}{dr^{2}}-\dfrac{g}{r^{2}}\left( \gamma+\dfrac{1}{2}\right) ^{2}\right]
=2\mu B_{0}g,$$ whose non-singular solution is$$g\left( r\right) =J_{\gamma+1/2}\left( \dfrac{B_{\perp}^{\prime}r}{B_{0}K}\right) ,$$ where $J_{\alpha}\left( x\right) $ is the Bessel function of the first kind of order $\alpha$.
We note that all the four terms of $H_{int}$ does not operate on the $\phi$ coordinate. This is a consequence of the fact that $L_{z}-S_{z}$ (where $L_{z}$ is the $z$-component of the orbital angular momentum and $S_{z}=\hbar\sigma_{z}/2$ is the $z$-component of the spin) is conserved. Hence, in order to have a non-vanishing matrix element between the up-state and the down-state, they must have the *same* $\phi$-dependence. Thus, $\gamma=\nu=1/2$ and as a result $$\psi_{\uparrow}=CJ_{1}\left( \dfrac{B_{\perp}^{\prime}r}{B_{0}K}\right)
e^{i\phi/2} \label{up0.2}$$ with$$\Psi_{\uparrow}=\left(
\begin{array}
[c]{c}CJ_{1}\left( \dfrac{B_{\perp}^{\prime}r}{B_{0}K}\right) e^{i\phi/2}\\
0
\end{array}
\right) . \label{up0.3}$$ where $C$ is the normalization constant which is chosen to be real.
The wave function given in Eq.(\[up0.2\]) is oscillatory. It has a period of about $$\Delta r_{\uparrow}\sim K\dfrac{B_{0}}{B_{\perp}^{\prime}} \label{up5}$$ near the origin. Comparing it to $\Delta r_{\downarrow}$ given in Eq.(\[down2\]), we find that $$\dfrac{\Delta r_{\uparrow}}{\Delta r_{\downarrow}}\sim\sqrt{K}, \label{up6}$$ which shows that, for $K\ll1$, the wavefunction $\psi_{\uparrow}$ executes many oscillations in the region where $\psi_{\downarrow}$ is appreciable.
The lifetime.\[time\]
---------------------
To evaluate the lifetime $T_{esc}$ of the particle in its trapped state, which is the average time it takes for the particle to escape, we calculate the transition rate from the bound state given by Eq.(\[down6\]), to the unbounded state Eq.(\[up0.3\]), according to Fermi’s golden rule[@fermi]. Thus, $$\dfrac{1}{T_{esc}}=\dfrac{2\pi}{\hbar}\left| H_{\downarrow,\uparrow}\right|
^{2}g(E_{\uparrow}) \label{t1}$$ where $$\begin{aligned}
H_{\downarrow,\uparrow} & ={\displaystyle\int\limits_{0}^{\infty}}
rdr{\displaystyle\int\limits_{0}^{2\pi}}
d\phi\Psi_{\uparrow}^{\dagger}H_{int}\Psi_{\downarrow}\label{t2}\\
& =-\dfrac{\hbar^{2}}{2m}{\displaystyle\int\limits_{0}^{\infty}}
rdr{\displaystyle\int\limits_{0}^{2\pi}}
d\phi\psi_{\uparrow}^{\ast}\left\{ -\left( \dfrac{d\theta}{dr}\right)
\dfrac{\partial}{\partial r}-\dfrac{1}{2}\dfrac{d^{2}\theta}{dr^{2}}-\dfrac
{1}{2r}\dfrac{d\theta}{dr}-\dfrac{i}{r^{2}}\sin\theta\dfrac{\partial}{\partial\phi}\right\} \psi_{\downarrow}\nonumber\end{aligned}$$ is the matrix element of $H_{int}$ Eq.(\[h7.01\]) between $\Psi_{\downarrow
}$ and $\Psi_{\uparrow}$, and $g(E_{\uparrow})$ is the density of the final states at energy $E_{\uparrow}$.
The integrand in Eq.(\[t2\]) consists of a product of three elements: The function $\psi_{\downarrow}^{\ast}$ whose ‘width’ is about $\Delta
r_{\downarrow}$ (given in Eq.(\[down2\])) around the origin, An operator consisting of four $\theta$-dependent terms whose extent around the origin $\Delta r_{\mu B}$ is roughly $\sqrt{1/K}$ larger than $\Delta r_{\downarrow}$ and the function $\psi_{\uparrow}$ which is an *oscillatory* function with a characteristic period near the origin $\Delta r_{\uparrow}$ which is $\sqrt{K}$ *smaller* than $\Delta r_{\downarrow}$. This suggests that we can approximate the integral in Eq.(\[t2\]) by substituting $\sin\theta$, $d\theta/dr$ and $d^{2}\theta/dr^{2}$ by their value at $r=0$, $$\begin{aligned}
\sin\theta & \simeq\dfrac{B_{\bot}^{\prime}}{B_{0}}r\label{t5}\\
\dfrac{d\theta}{dr} & \simeq\dfrac{B_{\bot}^{\prime}}{B_{0}}\nonumber\\
\dfrac{d^{2}\theta}{dr^{2}} & \simeq0\nonumber\end{aligned}$$
Substituting Eqs.(\[t5\]), (\[down1\]) and (\[up0.2\]) into Eq.(\[t2\]) gives $$H_{\downarrow,\uparrow}\simeq-\sqrt{\pi}\hbar^{2}\frac{B_{\bot}^{\prime}}{mB_{0}}C\sqrt{\dfrac{2}{K}}\exp\left[ -\dfrac{1}{K}\right] , \label{t8}$$ where we have used the definite integral $${\displaystyle\int\limits_{0}^{+\infty}}
r^{2}J_{1}\left( br\right) e^{-ar^{2}}dr=\dfrac{b}{4a^{2}}\exp\left[
-\dfrac{b^{2}}{4a}\right] . \label{t8.1}$$
When Eq.(\[t8\]) is substituted into Eq.(\[t1\]) the term $C^{2}g(E_{\uparrow})$ appears. This term can be calculated by temporarily introducing suitable boundary conditions: Assume that the system is bounded by an infinite potential wall at $r=R$, the radius $R$ being large compared to $\Delta r_{\downarrow}$ yet small when compared to $\Delta r_{\mu B}$. In this case the uniform potential approximation holds for all $r<R$, and the wave function has the form$$\psi_{\uparrow}\left( r,\phi\right) =Cg\left( r\right) e^{i\phi/2}\text{,}$$ where the radial part $g(r)$ satisfies the Schrödinger equation $$-\frac{\hbar^{2}}{2m}\left[ \frac{d^{2}g}{dr^{2}}+\frac{1}{r}\frac{dg}{dr}-\frac{g}{r^{2}}\right] =(\mu B_{0}+E)g.$$ This equation has as non-singular solutions the Bessel functions of order 1, $$g(r)=J_{1}(kr)\quad\mathrm{where}\quad k^{2}=\frac{2m}{\hbar^{2}}(E+\mu
B_{0}),$$ with eigenvalues $k=k_{n}$ determined by the boundary condition. From $R\gg
r_{\downarrow}$ it follows that $kR\gg1$ such that $J_{1}(kR)$ may be approximated by the first term of its asymptotic expansion. Therefore, the boundary condition reads $$J_{1}(kR)=\sqrt{\frac{2}{\pi kR}}\,\cos\left( kR-\frac{3\pi}{4}\right) =0.$$ This yields the eigenvalues $k_{n}=(n+1/4)\pi/R$. The density of states on the $k$-axis is thus given by $dN/dk=R/\pi$, from which one obtains the density of states on the energy axis at $E=\mu B_{0}$ $$\rho(E=\mu B_{0})=\frac{dN}{dE}=\frac{mR}{\pi\hbar^{2}k}=\frac{1}{2\pi}\sqrt{\frac{m}{\hbar^{2}\mu B_{0}}}\,R. \label{t8.2}$$ The constant $C$ is determined by the normalization condition $$\int|\psi_{\uparrow}|^{2}rdrd\phi=2\pi C^{2}\int_{0}^{R}[J_{1}(kr)]^{2}rdr=1,$$ which gives[@grad] $$\int_{0}^{R}[J_{1}(kr)]^{2}rdr=\frac{1}{2}R^{2}[J_{2}(kR)]^{2}.$$ In the asymptotic region $kR\gg1$, the function $J_{2}(kR)$ takes the values $\pm\sqrt{2/(\pi kR)}$ at the zeros of $J_{1}(kR)$. This gives $$C=\frac{k}{2R} \label{t8.3}$$ and therefore $$C^{2}\rho(E=\mu B_{0})=\frac{m}{2\pi\hbar^{2}}. \label{t13}$$
Finally, using Eqs.(\[t13\]) and (\[t8\]) inside Eq.(\[t1\]) gives $$T_{esc}=\dfrac{1}{64\pi\omega_{vib}}\exp\left[ \dfrac{2}{K}\right]
=\dfrac{T_{vib}}{128\pi^{2}}\exp\left[ \dfrac{2}{K}\right] ,$$ where $T_{vib}=2\pi/\omega_{vib}$ is the period of classical oscillations inside the trap.
Discussion.\[dis\]
==================
Summarizing all we have found we conclude that the problem we have studied has three important time scales: The shortest time scale is $T_{prec}$, which is the time required for *one* precession of the spin around the axis of the local magnetic field. The intermediate time scale is $T_{vib}=T_{prec}/K$, which is the time required to complete one cycle of the center of mass around the center of the trap. These two time scales appear both in the classical and the quantum-mechanical analysis. The longest time scale (provided $K$ is small) $T_{esc}$, which is not present at the classical problem, is the time it takes for the particle to escape from the trap.
Whereas the classical analysis yields an upper bound of $K=\sqrt{4/27}$ for trapping to occur, no such sharp bound exists in the quantum-mechanical analysis. This is related to the fact that one cannot associate an effective potential well with a finite barrier with the system.
As an example, we apply our results to the case of a neutron and an atom trapped with a field $B_{0}=100$ Oe and $B_{0}/B_{\bot}^{\prime}=10$cm. These parameters correspond to typical traps used in Bose-Einstein condensation experiments[@bec; @bec2; @bec3; @bec4]. The results, being correct to within an order of magnitude, are outlined in the following table. $$\begin{tabular}
[c]{|c||c|c|}\hline
& \multicolumn{2}{||c|}{$\begin{array}
[c]{c}B_{0}=100\text{ Oe}\\
B_{0}/B_{\bot}^{\prime}=10\text{cm}\end{array}
$}\\\cline{2-3}& Neutron & Atom\\\hline
$m$ gr & $\sim10^{-25}$ & $\sim10^{-22}$\\\hline
$\mu$ emu & $\sim10^{-23}$ & $\sim10^{-20}$\\\hline
$K$ & $\sim10^{-5}$ & $\sim10^{-8}$\\\hline
$T_{prec}$ sec & $\sim10^{-6}$ & $\sim10^{-9}$\\\hline
$T_{vib}$ sec & $\sim10^{-1}$ & $\sim10^{-1}$\\\hline
$T_{esc}$ sec & $\sim10^{\left( 10^{5}\right) }$ & $\sim10^{\left(
10^{8}\right) }$\\\hline
\end{tabular}$$
We note that in both cases $K$ is much smaller than $1$. Also, the calculated lifetime of the particle in the trap is extremely large, suggesting that the particle (either neutron or atom) is tightly trapped in this field.
The problem studied in this paper deals with a spin $1/2$ particle. Though this fact has little influence on the solution of the *classical* problem, the extension to higher spin values complicates the analysis of the *quantum-mechanical* problem. In this case one has to deal with a ($2S+1$)-component spinor, and the interaction Hamiltonian does no longer connect the ($-S$)-state to the ($+S$)-state, but only to the ($-S+1$) and ($-S+2$) states which for $S\geq5/2$ will *still* be trapped.
[99]{} A. L. Migdall, J. V. Prodan, W. D. Phillips, T. H. Bergeman and H. J. Metcalf, “First observation of magnetically trapped neutral atoms”, *Phys. Rev. Lett.*, **54** (24), 2596-2599 (1985).
T. Bergeman, G. Erez and H. J. Metcalf, “Magnetostatic trapping fields for neutral atoms”, *Phys. Rev. A.,* **35** (4), 1535-1546 (1987).
V. S. Bagnato, G. P. Lafyatis, A. G. Martin, E. L. Raab, R. N. Ahmad-Bitar and D. E. Pritchard, “Continous stopping and trapping of neutral atoms”, *Phys. Rev. Lett.*, **58** (21), 2194-2197 (1987).
W. Petrich, M. H. Anderson, J. R. Ensher and E. A. Cornell, “Stable, tightly confining magnetic trap for evaporative cooling of neutral atoms”, *Phys. Rev. Lett.*, **74** (17), 3352-3355 (1995).
M. O. Mewes, M. R. Andrews, N. J. Van-Druten, D. M. Kurn, D. S. Durfee, W. Ketterle, “Bose-Einstein condensation in a tightly confining DC magnetic trap”, *Phys. Rev. Lett.*, **77**(3), 416-419 (1996).
The Levitron is available from ‘Fascinations’, 18964 Des Moines Way South, Seattle, WA 98148.
The U-CAS is available from Masudaya International Inc., 6-4, Kuramae, 2-Chome, Taito-Ku, Tokyo, 111 Japan.
R. Harrigan, U.S. Patent Number: 4,382,245, Date of Patent: May 3, 1983.
Hones et al., U.S. Patent Number: 5,404,062, Date of Patent: Apr. 4, 1995.
R. Edge, “Levitation using only permanent magnets”, *Phys. Teach.* **33**, 252-253 (1995) and “Corrections to the levitation paper”, *ibid*. **34**, 329 (1996).
M. V. Berry, *Proc. R. Soc. Lond.* A **452,** 1207-1220 (1996).
S. Gov and S. Shtrikman, *Proc. of the 19*$^{\text{th}}$* IEEE Conv. in Israel*, 184-187 (1996).
M. D. Simon, L. O. Heflinger and S. L. Ridgway, *Am. J. Phys.* **65** (4), 286-292 (1997).
S. Gov, S. Shtrikman and H. Thomas, *Los-Alamos E-Print Archive*, http://xxx.lanl.gov/physics/9803020 (1998).
S. Earnshaw, *Trans. Cambridge Philos. Soc.* **7**, 97-112 (1842).
The quantized motion of atoms in a quadrupole magnetic trap has been studied *numerically* by T. H. Bergeman, P. McNicholl, J. Kycia, H. Metcalf and N. L. Balazs, “Quantized motion of atoms in a quadrupole magnetostatic trap”, *J. Opt. Soc. Am. B*, **6** (11), 2249 (1989). Here, we use an *analytic* method to find the lifetime of the particle in such a trap.
S. Gov, S. Shtrikman and H. Thomas, *Los-Alamos E-Print Archive*, http://xxx.lanl.gov/physics/9808007, Submitted to *Am. J. Phys.*
J. D. Weinstein and K. G. Libbrecht, *Phys. Rev. A.,* **52** (5), 4004-4008 (1995).
L. D. Landau and E. M. Lifshitz, Quantum Mechanics (Pergamon Press), 3$^{\text{rd}}$ ed., pp. 213-214.
‘Quantum Mechanics’ by E. Merzbacher, *John Wiley & Sons.*, 2$^{\text{nd}}$ Ed., Ch. 5, Sec. 3, 57-61.
‘Quantum Mechanics’ by E. Merzbacher, *John Wiley & Sons.*, 2$^{\text{nd}}$ Ed., Ch. 18, Sec. 8, 475-481.
This integral may be found in ‘Table of Integrals, Series, and Products’ by I. S. Gradshteyn and I. M. Ryzhik, *Academic Press*, 5$^{\text{th}}$ Ed., 6.521.1, pp. 697. Note that this integral make explicit use of the fact that $J_{1}\left( kR\right) =0$.
M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman and E. A. Cornell, “Observation of Bose-Einstein condensation in a dilute atomic vapor”, *Science* **269**, 198 (1995).
K. B. Davis, M. O. Mewes, M. R. Andrews, N. J. van Druten, D. S. Durfee, D. M. Kurm and W. Ketterle, “Bose-Einstein condensation in a gas of Sodium atoms”, *Phys. Rev. Lett.* **75**, 3969 (1995).
C. C. Bradley, C. A. Sackett, J. J. Tollett and R. G. Hulet, “Evidence of Bose-Einstein condensation in an atomic gas with attractive interactions”, *Phys. Rev. Lett.* **75**, 1687 (1995); *ibid* **79**, 1170 (1997).
E. A. Cornell and C. E. Wiemann, “The Bose-Einstein condensate”, *Sci. Am.* **278**, 26-31 (1998).
[^1]: Also with the Center for Technological Education Holon, 52 Golomb St., P.O.B 305, Holon 58102, Israel.
[^2]: Also with the Department of Physics, University of California, San Diego, La Jolla, 92093 CA, USA.
| {
"pile_set_name": "ArXiv"
} |
Introduction
============
Elemental abundance trends of Galactic bulge stars are crucial for understanding bulge formation. Galactic chemical evolution is best traced by dwarf stars because their spectra are straightforward to analyze (Edvardsson 1993). However, observations of bulge dwarfs are challenging due to their faintness ($V$=19–20; Feltzing and Gilmore 2000), impeding spectroscopic observation under normal circumstances. Consequently, many studies have focused on giant stars, despite the difficulty in analyzing their spectra. This difficulty has led to shifts in the mean \[Fe/H\] and the \[Fe/H\] distribution function of the bulge as the analysis techniques are refined (Rich 1988, McWilliam and Rich 1994, Fulbright 2007, Zoccali 2008, Hill 2011). Fortunately, gravitational microlensing offers a unique opportunity to observe bulge dwarfs. When a bulge dwarf is lensed by a foreground object, its brightness can increase by $>$5 magnitudes, enabling spectroscopic observations of sufficiently high resolution and signal-to-noise (S/N) for an abundance analysis (Minniti 1998, Johnson 2007, Bensby 2011 and references therein).
The abundances of different elements are correlated, reflecting their origin in a common nucleosynthetic process, such as Type II or Type Ia supernovae (SNe). However, these correlations are not perfect because the same element can be made in multiple nucleosynthetic processes, whose relative contributions to a star are best distinguished if abundances are measured for many elements. Here we analyze a sample of 35 bulge dwarfs[^1], all of which have at least seven measured elemental abundances and the majority (24/35) of which have 11 or 12 measured elemental abundances, with median errors $\sigma
< 0.25$ dex. Bensby (2011) find that the bulge dwarf \[Fe/H\] distribution function is bimodal, peaked at \[Fe/H\] $\approx -0.6 \; {\rm and} \; +0.3$. They further show that the $\alpha$-element abundances in the bulge dwarfs vary systematically with \[Fe/H\], following the trends found for thin and thick disk dwarfs (Bensby 2003, 2005, Reddy 2003, 2006). Here we revisit this data set with a different analysis technique based on principal component decomposition of the elemental abundance patterns, showing that the bimodality seen in \[Fe/H\] also appears in the [*relative*]{} elemental abundance patterns.
Principal component analysis (PCA) is a natural tool for characterizing correlations in a high-dimensional space, reducing the overall dimensionality of the data set while allowing the data themselves to reveal the strongest patterns of correlations. While PCA has a long history in astronomy, its application to elemental abundance analysis is relatively new. The only such application we know of is the study of Ting (2012), who used this technique to investigate the distributions in elemental abundance space (hereafter, [$\mathcal{C}$-space]{}; Freeman and Bland-Hawthorn 2002) defined by various samples of stars from the disk, halo, clusters, and satellite galaxies. They found that disk stars occupied about 6 dimensions within the 17-dimensional [$\mathcal{C}$-space]{} of the data, but the nucleosynthetic processes likely responsible for the lowest order dimensions changed as a function of \[Fe/H\]. Their work demonstrated the potential usefulness of principal component abundance analysis (PCAA) as a way to identify groups of stars with distinct enrichment histories. Here we apply this approach to bulge dwarfs to shed further light on the formation of the Galactic bulge.
Method
======
PCAA defines a new set of orthogonal basis vectors in [$\mathcal{C}$-space]{} whose components are chosen to align with the maximum variation within the data not already attributed to lower order components. We use standard PCA (see, Jolliffe 1986) with the data matrix {$d_{i,j}$} representing the logarithmic abundance[^2] relative to iron of element $j$ for star $i$: $d_{i,j}$ = \[$X_j$/Fe\] with $X_j$ = O, Na, Mg, Al, Si, Ca, Ti, Cr, Ni, Zn, Y, and Ba for $j$ = 1, 2, ..., 12. PCA identifies orthogonal eigenvectors ${\bf e}_{k} = \{e_{k,j}\}$ such that the abundance of a given element in a given star can be represented as a sum $$\label{eqn:pcbasis}
d_{i,j} = \bar{d_j} + \displaystyle\sum\limits_{k=1}^{N_{\rm PC}} c_{i,k}
e_{k,j},$$ where $\bar{d_{j}} = \frac{1}{N_{\rm star}}
\displaystyle\sum\limits_{i=1}^{N_{\rm star}} d_{i,j}$ is the mean value of \[$X_j$/Fe\] in the full data set and $c_{i,k}$ is the coefficient for the $k^{\rm
th}$ PC of star $i$. The first PC describes the direction in elemental abundance space along which the sample stars exhibit the greatest variation, the second PC describes the direction of the second largest variation, etc. If the number of principal components in the sum is equal to the number of elements measured, then the data can be represented exactly. However, if elemental abundances are correlated so that stars are restricted to a lower dimensional subspace, then the elemental abundances can be represented to good accuracy by a smaller number of PCs.
Our data set comes from the homogeneous elemental abundance and error reanalysis of microlensed bulge dwarfs and subgiants by Bensby et al. (in prep.); abundances for 26 of our 35 sample stars have been previously published (Epstein 2010, Bensby 2010, 2011).
Results
=======
The dimensionality of the subspace occupied by stars within the full [$\mathcal{C}$-space]{}can be expressed as the number of PCs required to explain the intrinsic variation in the data (the variation not attributable to observational errors). Ting (2012) used Monte Carlo simulations spanning a range of intrinsic dimensionality and variance to show that the true dimensionality was recovered when the cumulative variation of the first $k$ PCs was about 85%. For our sample of microlensed bulge dwarfs, the first 1, 2, and 3 PCs describe 64%, 77%, and 84% of the cumulative variation within the data. Although the threshold for completely describing the data likely depends on many factors including sample size, number of observed abundances, and abundance uncertainties, it is fair to say that the bulge dwarfs occupy approximately three dimensions of the 12-dimensional [$\mathcal{C}$-space]{} investigated here.
Fig. 1 shows the first two PCs derived from the observed bulge dwarf elemental abundances, with the uncertainties determined from bootstrap resampling (see below). PC1 is dominated by the abundances of oxygen, other $\alpha$-elements (Mg, Si, Ca, and Ti), and Al, with small uncertainties on the relative contributions of each abundance. SNe II are the primary source of $\alpha$-elements and Al, but they are also a significant producer of Fe, especially at early times. By contrast, SNe Ia create large amounts of Fe and other Fe-peak elements, leading to sub-solar \[$\alpha$/Fe\] yields; once enough time has elapsed for a substantial number of SNe Ia to occur, they become the dominant source of Fe. The interplay between these two nucleosynthetic sources is thought to underpin the dichotomy in \[$\alpha$/Fe\] observed in Galactic disk stars, with high \[$\alpha$/Fe\] for “thick disk” stars reflecting rapid formation (Fuhrmann 1998) and roughly solar \[$\alpha$/Fe\] for “thin disk” stars, which have predominantly higher \[Fe/H\] (Gilmore and Wyse 1985). Although PCA is a “blind” statistical technique with no [*a priori*]{} theoretical input, in this data set (and others we have explored) the first principal component picks up this expected distinction between the two dominant supernova enrichment mechanisms.
PC2 is primarily governed by Na with secondary contributions from Ni (correlated with Na) and Ba (anticorrelated with Na). A similar Na–Ni correlation, attributed to metallicity-dependent SN II yields, has been found previously in halo stars (Nissen and Schuster 1997, 2010); however, this study and the companion paper of Bensby et al. (in prep.) are the first to identify a Na–Ni correlation amongst bulge stars (see Bensby et al. in prep. for more details). Na is primarily produced by hydrostatic carbon burning in the massive stars that explode as SNe II, but proton capture at the same temperatures depletes the pre-explosion Na abundance. As metallicity increases, the neutron excess increases, making Na less susceptible to proton capture and consequently increasing the overall Na yield (Clayton 2003). Similarly, the yield of $^{58}$Ni (the most common Ni isotope) from SNe II is sensitive to the neutron excess and the abundance of neutron-rich nuclei, like $^{23}$Na, in the progenitor star (Woosley 1973). Additional significant $^{58}$Ni production occurs in SNe Ia, whose $^{58}$Ni yield increases with metallicity (Timmes 2003). However, the Na–Ba anticorrelation is not readily explained by a single nucleosynthetic process. One star with distinctive abundances, MOA-2009-BLG-259S (see Fig. 3c), has a large impact on the contributions of Zn, Y, and Ba to PC2 and drives up the uncertainties for these abundances. The crosses/dashed line in Fig. 1b show the effect of omitting this star when defining principal components; PC1 hardly changes, but the contributions of Zn and Y to PC2 switch from being correlated with Na to anticorrelated. The \[Y/Fe\] for MOA-2009-BLG-259S has a large uncertainty, though the elevated ${\rm [Zn/Fe]} = 0.44 \pm 0.17$ appears to be well-established (Fig. 3c). Regardless of MOA-2009-BLG-259S, PC2 is dominated by Na and shows a Na–Ni correlation and a Na–Ba anticorrelation.
For comparison, we have found the principal components of a sample of 702 solar neighborhood thin and thick disk dwarfs from Bensby (2003, 2005, and in prep.) with the same elements measured. The disk PC1 and PC2 are shown as triangles/dotted lines in Fig. 1. The clear similarity between the bulge and disk PC1s suggests that the relative enrichment from SNe II vs. SNe Ia is the main driver of diversity among stars in both samples. On the other hand, the disk and bulge PC2s do not resemble each other: in contrast to the bulge PC2 discussed above, the disk PC2 is dominated entirely by correlated Y and Ba, both neutron-capture elements produced mainly by the $s$-process in disk stars (Sneden 2008). The short vs. long lifetimes of the nucleosynthetic sources driving the bulge and disk PC2s, respectively, implies that the bulge formed too rapidly ($\lesssim$Gyr) for asymptotic giant branch stars to dominate PC2. Johnson (2012) independently reached a similar conclusion based on the relative abundances of $r$- and $s$-process elements.
{width="12cm"}
Fig. 2 shows the distribution of the 35 bulge dwarfs in (PC1, PC2)-space, with the upper panel showing the histograms of the bulge dwarfs in PC1 and of kinematically selected subsets of thin and thick disk dwarfs[^3] projected onto the bulge PC1. It is evident from visual inspection that the bulge dwarfs divide into two distinct groups along the PC1 axis, centered at PC1 values of $\pm0.3$. (Because the sum in Eq. 1 includes the mean elemental abundances of the sample, a value of ${\rm PC1} = -0.3$ corresponds to $\sim$\[$\alpha$/Fe\]$_\odot$.) Thus, this analysis of [*relative*]{} elemental abundances, with no direct input from \[Fe/H\], recovers the bimodality that Bensby (2011) found in the \[Fe/H\] distribution without reference to relative abundances. Bensby (2011) did find elevated \[$\alpha$/Fe\] ratios for the metal-poor bulge dwarfs, and we recover the same correlation in this “reverse” analysis: every star with ${\rm PC1} < -0.1$ has ${\rm [Fe/H]} > -0.02$, and all but two of the stars with ${\rm PC1} > -0.1$ have ${\rm [Fe/H]} < -0.18$. Of these two stars, one (MOA-2009-BLG-259S; square) is a clear PC2 outlier, while the other (MOA-2010-BLG-523S; triangle) is a moderate PC3 outlier that is undistinguished in (PC1, PC2)-space.
Fig. 2a shows that the metal-rich and metal-poor bulge dwarfs track the thin and thick disk dwarfs, respectively, in PC1. PCAA highlights the scarcity of stars with intermediate \[$\alpha$/Fe\] (PC1 $\sim$ 0) in the bulge dwarfs and in the thin and thick disk dwarfs. The fact that the two bulge populations track the two disk populations in \[Fe/H\] and PC1 suggests that the stars had $\alpha$-enrichment histories that produced the same abundance patterns. This similarity provides tentative evidence that the bulge formed through secular processes, such as disk instabilities, that heated inner thin and thick disk stars to form the bulge (Kormendy and Kennicutt 2004). Alternatively, the bulge and disk could have formed by distinct mechanisms but experienced “convergent” enrichment histories. For example, Bournaud (2007) propose that the large cosmological accretion rates of high redshift galaxies enable the rapid ($\sim$0.5–1 Gyr), simultaneous formation of the bulge and inner disk, which could account for the $\alpha$-enhanced subpopulation of each component.
To test the statistical significance of PC1 and PC2 and the robustness of the bimodality in Fig. 2, we created 100 bootstrap resamplings of the data set, redefining principal components each time. PC1 is always recovered, and the general form of PC2 (high Na value, Na–Ni correlation, and Na–Ba anticorrelation) is recovered in 99/100 resamplings. Thus, PC2 is statistically significant even though the contributions of Zn, Y, and Ba to PC2 fluctuate because of its sensitivity to MOA-2009-BLG-259S. The histogram of PC1 values is always multimodal, showing two distinct groupings (as in Fig. 2a) about 90% of the time; the remaining resamplings show three apparent groupings, but the small sample size makes the distribution of values [*within*]{} the ${\rm PC1} > 0$ group difficult to characterize reliably. The significance of a formal test for bimodality will depend on the form of the adopted null hypothesis, but the likelihood ratio of a 2-Gaussian fit to the PC1 distribution to a 1-Gaussian fit is $\mathscr{L}(2{\rm G})/\mathscr{L}(1{\rm G}) = 2 \times 10^5$, a large improvement for the addition of two free parameters.
{width="12cm"}
Fig. 3 shows the decompositions of the elemental abundance patterns of four sample stars into the sum of the sample mean and the first one, two, or three PCs. The best fit coefficients $c_{i, k}$ of Eq. (1) are found for each star by [$\chi^2$]{}-minimization, treating all errors as independent. Panels (a) and (b) show typical examples of metal-poor and metal-rich stars, respectively, each of them fit almost perfectly with a single PC. Panel (c) shows MOA-2009-BLG-259S, which is poorly fit by PC1 alone but well fit when PC2 is also included. Panel (d) shows MOA-2010-BLG-523S, which has one of the largest PC3 coefficients in our sample.
Clearly these PCA fits have low values of [$\chi^2_{\rm red}$]{} = [$\chi^2$]{}/(d.o.f.), once a sufficient number of PCs (one, two, or three) is included. The number of degrees of freedom is the number of elemental abundances measured for the star minus the number of PCs in the fit. Panel (e) shows the histogram of [$\chi^2_{\rm red}$]{}values for single-PC and 2-PC decompositions; median values are [$\chi^2_{\rm red}$]{} = 0.31 and 0.21, respectively. The low [$\chi^2_{\rm red}$]{} values indicate that the observational errors on the elemental abundances are typically overestimated, at least in the sense that they do not represent the variance in estimated elemental abundance that would arise from observing the same star many times. We believe that these low [$\chi^2_{\rm red}$]{} values arise because the quoted errors effectively incorporate a component of systematic calibration error in addition to random error. For example, multiple lines of the same element may give discrepant abundance estimates because of uncertainties in the assumed oscillator strengths; the internal dispersion of these estimates is one useful indication of the absolute elemental abundance uncertainty, but the range of elemental abundance measurements from multiple high-S/N spectra will be smaller than this dispersion because the same oscillator strengths are assumed each time. The treatment of random and systematic errors and error correlations is a significant issue for PCAA and other model-fitting approaches, but we defer it to future investigations.
Conclusions
===========
Our results confirm and extend the findings of Bensby (2011), who identified bimodality in the \[Fe/H\] distribution of microlensed bulge dwarfs, with metal-poor stars showing enhanced $\alpha$-element abundances like those of solar neighborhood thick disk stars. Our principal component analysis demonstrates that the bimodality found by Bensby (2011) is present even in the [*relative*]{} elemental abundances (\[$X$/Fe\]) of the bulge dwarfs alone. The first PC is dominated by $\alpha$-elements, reflecting the dichotomy between SN II and SN Ia enrichment. The second PC captures the Na–Ni correlation caused by the metallicity dependence of SN II yields. Intriguingly, the two metal-rich stars that exhibit $\alpha$-enhancement (MOA-2009-BLG-259S and MOA-2010-BLG-523S) are also outliers from the main locus of stars in (PC1, PC2, PC3)-space, suggesting that they do indeed have unusual enrichment histories. If we project the thin and thick disk dwarfs onto the bulge dwarf PC1, they occupy the locations of the metal-rich and metal-poor bulge dwarfs, respectively. Analyzing local disk dwarfs, we find that the first principal component is nearly identical to the bulge PC1. However, the disk PC2 is governed by Y and Ba, products of long-lived asymptotic giant branch stars, whereas enrichment from short-lived SNe II drives the bulge PC2. Qualitatively, these results support a scenario in which the bulge grows by secular evolution from the inner disk, which itself has the elemental abundance dichotomy seen in local populations, but there may be other bulge formation models that can produce similar results.
Our results, and those of Ting (2012), illustrate the potential of PCAA as a tool for characterizing the distribution of stars in high-dimensional [$\mathcal{C}$-space]{}. One application, highlighted here, is to identify subpopulations in a sample, drawing on the information present in all measured elemental abundances simultaneously. With large multi-element samples, this approach could be used to isolate “interloper” stars accreted from a dissolved satellite, and perhaps to identify cohorts of stars associated with common birth clusters (Freeman and Bland-Hawthorn 2002). A second application, illustrated by the examples of MOA-2009-BLG-259S and MOA-2010-BLG-523S, is to identify outlier stars, either through their unusual locations in PC-space or because they are poorly fit by combinations of PCs that fit most stars well. Such outliers may reveal rare but physically informative enrichment pathways. A third application, highlighted by Ting (2012), is to characterize the dimensionality of the stellar distribution in [$\mathcal{C}$-space]{}, which is a basic test for models of Galactic assembly and enrichment. We will explore this technique in future work using theoretical models.
By allowing the data themselves to define the directions of strongest variation, PCAA complements the usual approach of testing predictive models that adopt nucleosynthetic yields from theoretical calculations. The PCs clearly do have physical content, but the connection of PCs to enrichment mechanisms is not one-to-one (see Ting 2012), and interpreting them will require comparisons to models that vary both enrichment and mixing histories and the nucleosynthetic yields themselves. Our analysis identifies several practical complications of PCAA, including the potential sensitivity to outliers in small samples, the treatment of missing elemental abundance measurements, the impact of heteroscedastic errors and correlated errors, and the mix of random and systematic contributions to the errors quoted in observational analyses. We will investigate these issues in future work. The enormous samples of high-resolution spectra anticipated from the SDSS-III APOGEE survey (Majewski in prep., Eisenstein 2011), the Gaia-ESO survey (Gilmore 2012), and the HERMES survey (Barden 2010) will map the elemental abundance distribution over wide swaths of the Galaxy, and PCAA will be a valuable tool for connecting these measurements to a comprehensive theory of the formation of the Milky Way.
[^1]: Although our sample includes some subgiant stars, we will describe it as “bulge dwarfs” for brevity.
[^2]: Elemental abundances are defined as \[$X/Y$\] $\equiv$ ${\rm
log}(N_X / N_Y) - {\rm log}(N_X / N_Y)_\odot$, with missing data replaced by the average value of that elemental abundance for the other stars.
[^3]: The kinematic designation is based on the Bensby (2003) selection criteria; however, we adopt a more stringent cut on the relative thick/thin disk membership probability: stars with $P$(thick disk)/$P$(thin disk) $>$ 100 and $P$(thick disk)/$P$(thin disk) $<$ 0.01 are designated thick and thin disk stars, respectively.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'An analytical expression for square-well fluid direct correlation function (DCF) obtained recently by Tang (Y.Tang, J. Chem. Phys. **127**, 164504 (2007)) in the first-order mean spherical approximation is extended for wider well widths ($2<\lambda < 3$). Theoretically obtained direct correlation functions and radial distribution functions for square-well fluid with $\lambda=2.1$ and $\lambda=2.5$ are compared with corresponding results of Monte-Carlo simulation.'
author:
- 'S. P. Hlushak, A. Trokhymchuk'
- 'S. Sokołowski'
title: 'Direct correlation function of square well fluid with wide well: First order mean spherical approximation'
---
Introduction
============
Recently, the systems of particles interacting with discrete potentials gained much attention from the scientific community. Such an increased interest to this class of model systems is associated primarely with their ability to mimic the bulk properties of variety of complex fluid systems, like associating fluids, colloids, cluster particles etc. The discrete hard-core potentials composed of an attractive square well and a repulsive square barrier were shown to induce the liquid-liquid phase transition in a single-component model fluid [glaser2007,franzese2001gmg,rzysko2008jcp,cervantes2007jcp]{}. Similar phase transitions take place in the real substances like water, carbon and phosphorus. Moreover, the effective potentials with square-well plus square-barrier components may be used to mimic colloidal particles interactions, which according to the DLVO theory are characterized by a long-range repulsive barrier and one or possibly two attractive wells. Finally, inhomogeneous fluids with descrete repulsive and attractive interaction potentials represent another intriquing area of research because they often serve as a benchmark to study a variety of interesting problems such as interfacial phenomena, surface adsorption, wetting, capillary condensation, etc.
The most suitable method to handle these systems seems to be the classical density-functional theoriy (DFT). There are several versions of the DFT derived usualy from either weighted density approximation or perturbation expansions. Recently, Tang and Wu [@tang2003a1; @tang2004a2; @tang2004a3] have proposed a new DFT in which fundamental measure theory is combined with the so-called first-order mean-spherical approximation (FMSA). Its implementation requires knowledge of the thermodynamics and the direct correlation function of the model bulk fluid. In comparison with alternative approaches, DFT/FMSA possesses a number of advantages due to the both accuracy and simplicity of the FMSA solution. In combination with the DFT, the FMSA was already applied to study Lennard-Jones [@tang2003a1], Yukawa [@tang2003a2], Sutherlad [@Mi2008a4] fluid systems. Regarding the systems with descrete potentials, quite recently [@tang2008a5] the DFT/FMSA approach has been applied to study the structural properties of the square-well (SW) fluid model defined by the pair interaction potential,
$$u\left( r\right) =\left\{
\begin{array}{ll}
\infty , & \quad r<1, \\
-\varepsilon , & \quad 1\leq r<\lambda , \\
0, & \quad r\geq \lambda .%
\end{array}%
\right. \label{SW}$$
with a unit hard-core diameter, the attractive strength $\varepsilon $ and the well width parameter $1<\lambda <2$; the FMSA solution for the direct correlation function (DCF) of this model system has been obtained in a separate study [@tang2007dcf]. It has been shown that DFT/FMSA gives good performance for entire considered range of attraction parameter $1<\lambda <2$.
The fact that FMSA is linear with respect to the interaction potential allows one to use linear combinations of the different potential functions for which FMSA solution is known. Regarding the case of the SW model this means that one can employ the linear combinations of the square wells ($%
\varepsilon >0$) and/or square barriers ($\varepsilon <0$) in order to form all variety of the discrete potential models. Unfortunately, majority of potential applications of the discrete SW-based potential models (e.g., see Refs. [@glaser2007; @franzese2001gmg; @rzysko2008jcp; @cervantes2007jcp]) $\ \ $have the range of interaction that exceeds one particle diameter, that requires to consider $\lambda >2$ in Eq.(1). This means that existing FMSA solution obtained by Tang [@tang2007dcf] is not sufficient to have FMSA be applied in the studies of the complex discrete potential models. The aim of present work is to extend the recent FMSA solution[@tang2007dcf] for the SW model fluid with $1<\lambda <2$ to the case of the SW fluids with $%
2<\lambda <3$. Obtaining of a such FMSA solution for the attractive SW model is crucial for the future application of both the FMSA as well as the DFT/FMSA to the systems with a combined (attractive SW plus repulsive SB) discrete potential. This will be shown in a forthcoming paper \[\]. The actual paper is organized as follows. In the next Section II we outline the necessary details of the FMSA solution for SW potential with $%
1<\lambda <3.$ The FMSA results for the direct and pair correlation function will be presented and compared against computer simulation data in Section III. We conclude the paper in Section IV.
FMSA for square-well model fluids
=================================
The general FMSA formalism, developed by Tang and Lu, is presented in their paper [tang1993nso]{}. These authors solved the Ornstein-Zernike (OZ) equation, $$\tilde{h}\left( k\right) =\tilde{c}\left( k\right) +\rho \tilde{h}\left(
k\right) \tilde{c}\left( k\right)$$for a one-component system of particles/moleculs of a number density $\rho $ by employing the perturbative expansion for the total and direct correlation functions, $h$ and $c,$ respectively, over the pair interaction energy parameter $\varepsilon ,$ $$\begin{aligned}
\tilde{h}\left( k\right) &=&\tilde{h}_{0}\left( k\right) +\varepsilon \tilde{%
h}_{1}\left( k\right) +\dots \label{h_pert} \\
\tilde{c}\left( k\right) &=&\tilde{c}_{0}\left( k\right) +\varepsilon \tilde{%
c}_{1}\left( k\right) +\dots , \label{c_pert}\end{aligned}$$where the subscript $0$ denotes the contribution of a hard-sphere (HS) reference system while the subscript $1$ stands for the first-order perturbation term. Here and in what follows, in accordance with the previous notation[@tang1993nso] all symbols with the tilde denote the three-dimensional Fourier transforms, $$\begin{aligned}
\tilde{h}\left( k\right) &=&\frac{4\pi }{k}\int_{0}^{\infty }sin\left(
kr\right) rh\left( r\right) dr, \\
\tilde{c}\left( k\right) &=&\frac{4\pi }{k}\int_{0}^{\infty }sin\left(
kr\right) rc\left( r\right) dr, \label{fourier3D}\end{aligned}$$while all symbols with the hats denote the one-dimensional Fourier transforms or the Laplace transforms.
By employing the Hilbert transform, the general solution of the first order OZ equation can be obtained. The Fourier transform of the first order contribution to the total correlation function $h$ reads, $$\hat{h}_{1}\left( k\right) =\frac{P\left( ik\right) }{{\hat{Q}_{0}^{2}\left(
ik\right) }}, \label{P_def}$$where
$$P\left( ik\right) =\frac{U_{1}\left( k\right) }{2\hat{Q}_{0}^{2}\left(
-ik\right) }-\frac{e^{-ik{}}}{2i\pi }\int_{-\infty }^{\infty }\frac{%
U_{1}\left( y\right) e^{iy{}}}{\left( y-k\right) \hat{Q}_{0}^{2}\left(
-iy\right) }dy, \label{P_k}$$
and $\hat{Q_{0}}\left( ik\right) $ is the Baxter hard-sphere factorization function with a Laplace transform given by $$\hat{Q}_{0}\left( s\right) =\frac{S\left( s\right) +12\eta L\left( s\right)
e^{-s{}}}{\left( 1-\eta \right) ^{2}s^{3}}, \label{Q_0}$$with $\eta =\frac{1}{6}\pi \rho $, and $$\begin{aligned}
S\left( t\right) &=&\left( 1-\eta \right) ^{2}t^{3}+6\eta \left( 1-\eta
\right) t^{2}+18\eta ^{2}t-12\eta \left( 1+2\eta \right) , \label{S_def} \\
L\left( t\right) &=&\left( 1+\frac{\eta }{2}\right) t+1+2\eta .
\label{L_def}\end{aligned}$$The function $U_{1}\left( k\right) $ in Eq. (\[P\_k\]) is defined as
$$U_{1}\left( k\right) =\int_{1}^{\infty } rc_{1}\left( r\right) e^{-ikr}dr,
\label{U_1_def}$$
where, according to the FMSA closure, $$c_{1}\left( r\right) =-\beta u\left( r\right) ,\quad \mathrm{for}\quad r>1,
\label{c_1_def}$$with $\beta =1/k_{B}T$ and $u\left( r\right) $ being the pair potential. For smooth pair potentials $u\left( r\right) $, for which $U_{1}\left(
k\right) \sim e^{-ik{}}$ (e.g., for the Yukava potential), the integration contour in the right-hand side of Eq. (\[P\_k\]) can be closed in the upper complex half-plane and evaluation of the integral does not require to calculate the residues at zeroes of the function $\hat{Q_{0}}\left(
-ik\right) $. This important simplification does not hold for the pair potentials that vanish for $r>\lambda {}$ ($\lambda >1$), e.g. for the SW potential. For such a potential Tang and Lu expanded ${1}/{\hat{Q_{0}}\left( -ik\right) ^{2}}$ in a following way, $$\frac{1}{\hat{Q_{0}}\left( -ik\right) ^{2}}=\frac{\left( 1-\eta \right)
^{4}\left( -ik\right) ^{6}}{S^{2}\left( -ik\right) }\left[ 1-2\left( 12\eta
\frac{L\left( -ik\right) }{S\left( -ik\right) }e^{ik}\right) +3\left( 12\eta
\frac{L\left( -ik\right) }{S\left( -ik\right) }e^{ik}\right) ^{2}-\cdots %
\right] , \label{Q_expansion}$$and performed calculations according to the scheme described in the Appendix of their work[@tang1993nso]. Since only the first term of the expansion (\[Q\_expansion\]) survives for $\lambda \leq 2$, therefore $$\hat{h}_{1}\left( k\right) =-\frac{\left( 1-\eta \right) ^{4}e^{-ik}}{\hat{Q}%
_{0}^{2}\left( ik\right) }Res\left\{ \frac{W_{\lambda }\left( y\right)
\left( -iy\right) ^{6}e^{iy}}{\left( y-k\right) S^{2}\left( -iy\right) }%
\right\} ,\qquad \lambda \leq 2, \label{h_1_Ll1}$$where $$W_{\lambda }\left( k\right) =\int_{\lambda }^{\infty }r{c}_{1}\left(
r\right) e^{-ikr}dr \label{W_lambda_def}$$should be calculated for ${c}_{1}\left( r\right) $ smoothly extended from $%
r\in \left[ 1,\lambda \right) $ to $r\in \left[ \lambda ,\infty \right) $. For the SW potential we get[@tang1994aas] $$W_{\lambda }\left( k\right) =\beta \varepsilon \frac{1+ik\lambda }{\left(
ik\right) ^{2}}e^{-ik\lambda }. \label{W_lambda_SW_def}$$
Expression (\[h\_1\_Ll1\]) is valid only for the pair potentials $u\left(
r\right) $ with the attractive well width that does not exceed one particle diameter, i.e., for $\lambda \leq 2$. In order to extend the actual FMSA scheme for larger values of parameter $\lambda $, the contribution due to the second term in the expansion (\[Q\_expansion\]) should be taken into account. After doing that we obtain the following result, $$\hat{h}_{1}\left( k\right) =-\frac{\left( 1-\eta \right) ^{4}e^{-ik}}{\hat{Q}%
_{0}^{2}\left( ik\right) }Res\left\{ \frac{W_{\lambda }\left( y\right)
\left( -iy\right) ^{6}e^{iy}}{\left( y-k\right) S^{2}\left( -iy\right) }%
\left[ 1-24\eta \frac{L\left( -iy\right) }{S\left( -iy\right) }e^{iy}\right]
\right\} ,\quad 2<\lambda \leq 3, \label{h_1_Ll2}$$which may be viewed as Eq. (\[h\_1\_Ll1\]) supplemented with an additional term, that contributes to $r>2$ region in the $r$-space.
The therrmodynamic and structural properties of the square-well fluid for $%
\lambda <2$ within the FMSA approximation have been reported in Refs.[@tang1994aas; @tang2007dcf; @tang1994for]. Here, we would like to pay attention to the caase with $\lambda >2,$ i.e., to the development of an additional term that already appears in Eq. (\[h\_1\_Ll2\]). Let us denote it as $$\hat{h}_{1}^{\left( 2\right) }\left( k\right) =\frac{24\eta \left( 1-\eta
\right) ^{4}e^{-ik}}{\hat{Q}_{0}^{2}\left( ik\right) }Res\left\{ \frac{%
W_{\lambda }\left( y\right) \left( -iy\right) ^{6}L\left( -iy\right) e^{2iy}%
}{\left( y-k\right) S^{3}\left( -iy\right) }\right\} . \label{h_1_L2}$$After calculating the sum of residues in above expression, we obtain the following Laplace transform $$\begin{aligned}
\hat{h}_{1}^{\left( 2\right) }\left( s\right) &=&\frac{24\eta \left( 1-\eta
\right) ^{4}e^{-s}}{\hat{Q}_{0}^{2}\left( s\right) }\left\{ -\frac{C\left(
-s\right) L\left( -s\right) e^{-s\left( \lambda -2\right) }}{S^{3}\left(
-s\right) }\right. \notag \\
&+&\left. \frac{1}{2}\sum_{i=0}^{2}\left( \frac{G_{2}\left( t_{i}\right) }{%
s+t_{i}}+\frac{G_{1}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{2}}+\frac{%
G_{0}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{3}}\right) \right\} ,
\label{h_1_sL2}\end{aligned}$$where $C\left( s\right) =\lambda s^{5}-s^{4}$, while $t_{i}$ correspond to the roots of $S\left( t\right) =0$, and functions $G_{i}\left( s\right) $ are, $$\begin{aligned}
&&\!\!\!\!\!\!\!\!\!\!\!\!G_{0}\left( s\right) =-2\frac{C\left( s\right)
L\left( s\right) e^{s\left( \lambda -2\right) }}{S_{1}^{3}\left( s\right) },
\\
&&\!\!\!\!\!\!\!\!\!\!\!\!G_{1}\left( s\right) =2\frac{\left\{ \frac{{}}{{}}%
\frac{d\left[ C\left( s\right) L\left( s\right) \right] }{ds}+\left( \lambda
-2\right) C\left( s\right) L\left( s\right) \right\} e^{s\left( \lambda
-2\right) }}{S_{1}^{3}\left( s\right) }-3\frac{C\left( s\right) L\left(
s\right) S_{2}\left( s\right) e^{s\left( \lambda -2\right) }}{%
S_{1}^{4}\left( s\right) }, \\
&&\!\!\!\!\!\!\!\!\!\!\!\!G_{2}\left( s\right) =-\frac{\left\{ \frac{{}}{{}}%
\frac{d^{2}\left[ C\left( s\right) L\left( s\right) \right] }{ds^{2}}%
+2\left( \lambda -2\right) \frac{d\left[ C\left( s\right) L\left( s\right) %
\right] }{ds}+\left( \lambda -2\right) ^{2}C\left( s\right) L\left( s\right)
\right\} e^{s\left( \lambda -2\right) }}{S_{1}^{3}\left( s\right) } \\
&&+3\frac{\left\{ \frac{{}}{{}}\frac{d\left[ C\left( s\right) L\left(
s\right) \right] }{ds}+\left( \lambda -2\right) C\left( s\right) L\left(
s\right) \right\} S_{2}\left( s\right) e^{s\left( \lambda -2\right) }}{%
S_{1}^{4}\left( s\right) }+\frac{C\left( s\right) L\left( s\right)
S_{3}\left( s\right) e^{s\left( \lambda -2\right) }}{S_{1}^{4}\left(
s\right) } \notag \\
&&-3\frac{C\left( s\right) L\left( s\right) S_{2}^{2}\left( s\right)
e^{s\left( \lambda -2\right) }}{S_{1}^{5}\left( s\right) }, \notag
\label{G2}\end{aligned}$$with $S_{n}\left( s\right) =d^{n}S\left( s\right) /ds^{n}$.
Then the Laplace transform of the first order contribution to the total correlation function $h$ in the case of $2<\lambda \leq 3$ can be written as, $$\hat{h}_{1}\left( s\right) =\frac{\beta \varepsilon \left( 1-\eta \right)
^{2}}{\hat{Q}_{0}^{2}\left( s\right) }\left[ {\hat{p}^{\left( 1\right)
}\left( s\right) }{}+24\eta {\hat{p}^{\left( 2\right) }\left( s\right) }{}%
\right] .$$In the above we introduced two functions $$\hat{p}^{\left( 1\right) }\left( s\right) =e^{-s}Res\left[ \frac{C\left(
t\right) e^{\left( \lambda -1\right) t}}{\left( s+t\right) S^{2}\left(
t\right) }\right] , \label{p1s}$$and $$\hat{p}^{\left( 2\right) }\left( s\right) =e^{-s}Res\left[ \frac{C\left(
t\right) L\left( s\right) e^{\left( \lambda -2\right) t}}{\left( s+t\right)
S^{3}\left( t\right) }\right] . \label{p2s}$$The first one was examined in more details in Ref.[@tang2007dcf], whereas the second function, in agreement with Eq. (\[h\_1\_sL2\]), reads $$\begin{aligned}
\hat{p}^{\left( 2\right) }\left( s\right) &=&e^{-s}\left\{ -\frac{C\left(
-s\right) L\left( -s\right) e^{-s\left( \lambda -2\right) }}{S^{3}\left(
-s\right) }\right. \notag \\
&&+\left. \frac{1}{2}\sum_{i=0}^{2}\left( \frac{G_{2}\left( t_{i}\right) }{%
s+t_{i}}+\frac{G_{1}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{2}}+\frac{%
G_{0}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{3}}\right) \right\} .
\label{p2s_2}\end{aligned}$$
Following by Tang and Lu[@tang2007dcf], the Laplace transform of the DCF can be obtained from the relation $$\hat{c}\left( s\right) =\beta \varepsilon \left( 1-\eta \right) ^{4}\left\{ %
\left[ \hat{Q_{p}}\left( s\right) \right] ^{\left[ 0,\infty \right] }-\left[
\hat{Q_{p}}\left( -s\right) \right] ^{\left[ 0,\infty \right] }\right\} ,
\label{c1s}$$where superscript $\left[ 0,\infty \right] $ denotes the part that is nonzero only in $\left[ 0,\infty \right] $ region of the $r$ space, and $$\hat{Q_{p}}\left( s\right) =\hat{Q_{0}^{2}}\left( -s\right) \left[ \hat{p}%
^{\left( 1\right) }\left( s\right) +24\eta \hat{p}^{\left( 2\right) }\left(
s\right) \right] . \label{Qps}$$In order to invert $\hat{c}\left( s\right) $ to the $r-$space, we should ignore all $\left[ 0,\infty \right] $ superscripts in relation (\[c1s\]) and invert it to $r-$space. Since we are interested only in $r>0$ region, we assume that ${c}\left( r\right) =0$ for $r<0$. Note, that the contribution from the first term $\hat{p}^{\left(
1\right) }\left( s\right) $ in Eq. (\[Qps\]) was already considered in Ref.[@tang2007dcf]. Let us examine now the second term of Eq. (\[Qps\]), $$\hat{Q_{p}}^{(2)}\left( s\right) =24\eta \hat{Q_{0}^{2}}\left( -s\right)
\hat{p}^{\left( 2\right) }\left( s\right) . \label{Qp2s}$$Taking into account that $\hat{Q_{0}}\left( s\right) $ is analytical function in the whole complex plane, we may simplify the expression for $%
\hat{p}^{\left( 2\right) }\left( s\right) $ by subtracting some analytical function. Indeed, the subtracted analytical function being multiplied by $%
\hat{Q_{0}^{2}}\left( -s\right) $ will result in a new analytical function, that has no impact on the resulting expressions for $c_{1}\left( r\right) $ in the $r-$space. In such a way we replace the first term $-{C\left(
-s\right) L\left( -s\right) e^{-s\left( \lambda -2\right) }}/{S^{3}\left(
-s\right) }$ in Eq. (\[p2s\_2\]) by expression $\sum\limits_{i=0}^{2}\left(
\frac{C_{3}\left( t_{i}\right) }{s+t_{i}}+\frac{C_{2}\left( t_{i}\right) }{%
\left( s+t_{i}\right) ^{2}}+\frac{C_{1}\left( t_{i}\right) }{\left(
s+t_{i}\right) ^{3}}\right) e^{-s\left( \lambda -2\right) }$, where the coefficients $C_{1}\left( t_{i}\right) $, $C_{2}\left( t_{i}\right) $ and $%
C_{3}\left( t_{i}\right) $ are found from the equality of residues at all poles $t_{i}$ of the expression $$\underset{\{t_{i}\}}{Res}\left[ \left( \frac{C_{3}\left( t_{i}\right) }{%
s+t_{i}}+\frac{C_{2}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{2}}+\frac{%
C_{1}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{3}}\right) e^{-s\left(
\lambda -2\right) }\right] =\underset{\{t_{i}\}}{Res}\left[ -\frac{C\left(
-s\right) L\left( -s\right) e^{-s\left( \lambda -2\right) }}{S^{3}\left(
-s\right) }\right] . \label{res_equality}$$The new simplified function $\hat{p}^{\prime \left( 2\right) }\left(
s\right) $ reads $$\begin{aligned}
\hat{p}^{\prime \left( 2\right) }\left( s\right)
&=&\sum\limits_{i=0}^{2}\left( \frac{C_{3}\left( t_{i}\right) }{s+t_{i}}+%
\frac{C_{2}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{2}}+\frac{%
C_{1}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{3}}\right) e^{-s\left(
\lambda -1\right) } \notag \\
&&+\frac{1}{2}\sum_{i=0}^{2}\left( \frac{G_{2}\left( t_{i}\right) }{s+t_{i}}+%
\frac{G_{1}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{2}}+\frac{%
G_{0}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{3}}\right) e^{-s},
\label{ps_prime}\end{aligned}$$where $$\begin{aligned}
C_{1}\left( t_{i}\right) &=&\frac{C\left( t_{i}\right) L\left( t_{i}\right)
}{S_{1}^{3}\left( t_{i}\right) }, \notag \\
C_{2}\left( t_{i}\right) &=&-\frac{\frac{d}{ds}\left[ C\left( s\right)
L\left( s\right) \right] _{s=t_{i}}}{S_{1}^{3}\left( t_{i}\right) }+\frac{3}{%
2}\frac{C\left( t_{i}\right) L\left( t_{i}\right) S_{2}\left( t_{i}\right) }{%
S_{1}^{4}\left( t_{i}\right) }, \notag \\
C_{3}\left( t_{i}\right) &=&\frac{\frac{d^{2}}{ds^{2}}\left[ C\left(
s\right) L\left( s\right) \right] _{s=t_{i}}}{2S_{1}^{3}\left( t_{i}\right) }%
-\frac{3}{2}\frac{\frac{d}{ds}\left[ C\left( s\right) L\left( s\right) %
\right] _{s=t_{i}}S_{2}\left( t_{i}\right) }{S_{1}^{4}\left( t_{i}\right) }
\notag \\
&&-\frac{C\left( t_{i}\right) L\left( t_{i}\right) S_{3}\left( t_{i}\right)
}{2S_{1}^{4}\left( t_{i}\right) }+\frac{3}{2}\frac{C\left( t_{i}\right)
L\left( t_{i}\right) S_{2}^{2}\left( t_{i}\right) }{S_{1}^{5}\left(
t_{i}\right) }. \notag \label{C_i}\end{aligned}$$
It is easy to find that the functions $G_{i}\left( s\right) $ from Eqs. (21)-(23) can be expressed in terms of $C_{i}\left( s\right) $ in the following way $$\begin{aligned}
G_{0}\left( s\right) &=&-2C_{1}\left( s\right) e^{\left( \lambda -2\right)
s}, \notag \\
G_{1}\left( s\right) &=&-2\left[ C_{2}\left( s\right) -\left( \lambda
-2\right) C_{1}\left( s\right) \right] e^{\left( \lambda -2\right) s},
\notag \\
G_{2}\left( s\right) &=&-2\left[ C_{3}\left( s\right) -\left( \lambda
-2\right) C_{2}\left( s\right) +\frac{\left( \lambda -2\right) ^{2}}{2}%
C_{1}\left( s\right) \right] e^{\left( \lambda -2\right) s}. \notag
\label{GviaC}\end{aligned}$$Then the inverse Laplace transform of Eq. (\[ps\_prime\]) reads $$\begin{aligned}
{p}^{\prime \left( 2\right) }\left( r\right) &=&\frac{1}{2}%
\sum_{i=0}^{2}\left( \underset{\frac{{}}{{}}}{}G_{2}\left( t_{i}\right)
+\left( r-1\right) G_{1}\left( t_{i}\right) \right. \notag \\
&&+\left. \frac{\left( r-1\right) ^{2}}{2}G_{0}\left( t_{i}\right) \right)
e^{-t_{i}\left( r-1\right) }\left[ H\left( r-1\right) -H\left( r-\lambda
+1\right) \right] , \label{p2r_prime}\end{aligned}$$which is nonzero inside $\left[ 1,\lambda -1\right] $, and $H\left( r\right)$ is the Heaviside step function. Substituting (\[ps\_prime\]) and (\[Q\_0\]) into (\[Qp2s\]), we obtain $$\begin{aligned}
&&\frac{\hat{Q_{p}}^{(2)}\left( s\right) }{24\eta }=\hat{p}^{\prime \left(
2\right) }\left( s\right) +\left[ \sum_{i=1}^{6}\frac{e_{0,i}}{\left(
-s\right) ^{i}}+\sum_{i=2}^{6}\frac{e_{1,i}}{\left( -s\right) ^{i}}%
e^{s}+\sum_{i=4}^{6}\frac{e_{2,i}}{\left( -s\right) ^{i}}e^{2s}\right]
\sum_{i=0}^{2}\left\{ \frac{C_{3}\left( t_{i}\right) }{s+t_{i}}\right.
\notag \\
&&+\left. \frac{C_{2}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{2}}+\frac{%
C_{1}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{3}}\right\} e^{-\left(
\lambda -1\right) s}+\frac{1}{2}\left[ \sum_{i=1}^{6}\frac{e_{0,i}}{\left(
-s\right) ^{i}}+\sum_{i=2}^{6}\frac{e_{1,i}}{\left( -s\right) ^{i}}%
e^{s}\right. \notag \\
&&+\left. \sum_{i=4}^{6}\frac{e_{2,i}}{\left( -s\right) ^{i}}e^{2s}\right]
\sum_{i=0}^{2}\left\{ \frac{G_{2}\left( t_{i}\right) }{s+t_{i}}+\frac{%
G_{1}\left( t_{i}\right) }{\left( s+t_{i}\right) ^{2}}+\frac{G_{0}\left(
t_{i}\right) }{\left( s+t_{i}\right) ^{3}}\right\} e^{-s},
\label{Qp2s_prime}\end{aligned}$$and $$\begin{aligned}
&&\frac{\hat{Q_{p}}^{(2)}\left( -s\right) }{24\eta }=\hat{p}^{\prime \left(
2\right) }\left( -s\right) -\left[ \sum_{i=1}^{6}\frac{e_{0,i}}{s^{i}}%
+\sum_{i=2}^{6}\frac{e_{1,i}}{s^{i}}e^{-s}+\sum_{i=4}^{6}\frac{e_{2,i}}{s^{i}%
}e^{-2s}\right] \sum_{i=0}^{2}\left\{ \frac{C_{3}\left( t_{i}\right) }{%
s-t_{i}}\right. \notag \\
&&-\left. \frac{C_{2}\left( t_{i}\right) }{\left( s-t_{i}\right) ^{2}}+\frac{%
C_{1}\left( t_{i}\right) }{\left( s-t_{i}\right) ^{3}}\right\} e^{\left(
\lambda -1\right) s}-\frac{1}{2}\left[ \sum_{i=1}^{6}\frac{e_{0,i}}{s^{i}}%
+\sum_{i=2}^{6}\frac{e_{1,i}}{s^{i}}e^{-s}\right. \notag \\
&&+\left. \sum_{i=4}^{6}\frac{e_{2,i}}{s^{i}}e^{-2s}\right]
\sum_{i=0}^{2}\left\{ \frac{G_{2}\left( t_{i}\right) }{s-t_{i}}-\frac{%
G_{1}\left( t_{i}\right) }{\left( s-t_{i}\right) ^{2}}+\frac{G_{0}\left(
t_{i}\right) }{\left( s-t_{i}\right) ^{3}}\right\} e^{s},
\label{Qp2ms_prime}\end{aligned}$$where the coefficients $e_{n,i}$ of the expansion of the Baxter function $%
\hat{Q_{0}}^{2}\left( s\right) $ are given in the Appendix.
In order to obtain the DCF in $r-$space we will use the method that is similar to that outlined in Ref.[@tang2007dcf]. Namely, from Eqs. (\[Qp2s\_prime\]) and ([Qp2ms\_prime]{}) it is evident, that the key expressions to be inverted are the functions of following form $$\hat{w}\left( s,z,n,m\right) =\frac{1}{s^{n}\left( s+z\right) ^{m}}=\left\{
\begin{array}{ll}
\sum\limits_{k=1}^{n}\frac{a_{k}^{nm}}{s^{k}}+\sum\limits_{k=1}^{m}\frac{%
b_{k}^{nm}}{\left( s+z\right) ^{k}}, & z\neq 0, \\
\frac{1}{s^{m+n}}, & z=0,%
\end{array}%
\right. \label{wsznm}$$with coefficients $$\begin{aligned}
a_{k}^{nm} &=&\left( -1\right) ^{n-m}\frac{\left( m+n-k-1\right) !}{\left(
n-k\right) !\left( m-1\right) !z^{m+n-k}}, \\
b_{k}^{nm} &=&\left( -1\right) ^{n}\frac{\left( m+n-k-1\right) !}{\left(
m-k\right) !\left( n-1\right) !z^{m+n-k}}. \label{abnmk}\end{aligned}$$The inverse Laplace transform of these functions reads $$\hat{w}\left( r,z,n,m\right) =\left\{
\begin{array}{ll}
\left[ \sum\limits_{k=1}^{n}a_{k}^{nm}\frac{r^{k-1}}{\left( k-1\right) !}%
+\sum\limits_{k=1}^{m}b_{k}^{nm}\frac{r^{k-1}}{\left( k-1\right) !}e^{-zr}%
\right] H\left( r\right) , & z\neq 0, \\
\frac{r^{m+n-1}}{\left( m+n-1\right) !}H\left( r\right) , & z=0,%
\end{array}%
\right. \label{wrznm}$$where $H\left( r\right) $ is the Heaviside step function. With this in hands, we may now write down in terms of $\hat{w}\left( r,z,n,m\right) $the expressions for the inverse Laplace transforms of Eqs. (\[Qp2s\_prime\]) and (\[Qp2ms\_prime\]) for $r\geq 0,$ $$\begin{aligned}
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}\frac{Q_{p}^{\left( 2\right) }\left(
r\right) }{24\eta }=p^{\left( 2\right) }\left( r\right)
+\sum_{i=1}^{6}\sum_{j=0}^{2}\frac{e_{0,i}}{\left( -1\right) ^{i}}\left[
\frac{{}}{{}}C_{3}\left( t_{j}\right) w\left( r-\lambda +1,t_{j},i,1\right)
\right. \notag \\
&&\left. +C_{2}\left( t_{j}\right) w\left( r-\lambda +1,t_{j},i,2\right)
+C_{1}\left( t_{j}\right) w\left( r-\lambda +1,t_{j},i,3\right) \frac{{}}{{}}%
\right] H\left( r-\lambda +1\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}+\sum_{i=2}^{6}\sum_{j=0}^{2}\frac{e_{1,i}}{%
\left( -1\right) ^{i}}\left[ \frac{{}}{{}}C_{3}\left( t_{j}\right) w\left(
r-\lambda +2,t_{j},i,1\right) +C_{2}\left( t_{j}\right) w\left( r-\lambda
+2,t_{j},i,2\right) \right. \notag \\
&&\left. +C_{1}\left( t_{j}\right) w\left( r-\lambda +2,t_{j},i,3\right)
\frac{{}}{{}}\right] H\left( r-\lambda +2\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}+\sum_{i=4}^{6}\sum_{j=0}^{2}\frac{e_{2,i}}{%
\left( -1\right) ^{i}}\left[ \frac{{}}{{}}C_{3}\left( t_{j}\right) w\left(
r-\lambda +3,t_{j},i,1\right) \right. \notag \\
&&\left. +C_{2}\left( t_{j}\right) w\left( r-\lambda +3,t_{j},i,2\right)
+C_{1}\left( t_{j}\right) w\left( r-\lambda +3,t_{j},i,3\right) \frac{{}}{{}}%
\right] H\left( r-\lambda +3\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}+\frac{1}{2}\sum_{i=1}^{6}\sum_{j=0}^{2}%
\frac{e_{0,i}}{\left( -1\right) ^{i}}\left[ \frac{{}}{{}}G_{2}\left(
t_{j}\right) w\left( r-1,t_{j},i,1\right) \right. \notag \\
&&\left. +G_{1}\left( t_{j}\right) w\left( r-1,t_{j},i,2\right) +G_{0}\left(
t_{j}\right) w\left( r-1,t_{j},i,3\right) \frac{{}}{{}}\right] H\left(
r-1\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}+\frac{1}{2}\sum_{i=2}^{6}\sum_{j=0}^{2}%
\frac{e_{1,i}}{\left( -1\right) ^{i}}\left[ \frac{{}}{{}}G_{2}\left(
t_{j}\right) w\left( r,t_{j},i,1\right) \right. \notag \\
&&\left. +G_{1}\left( t_{j}\right) w\left( r,t_{j},i,2\right) +G_{0}\left(
t_{j}\right) w\left( r,t_{j},i,3\right) \frac{{}}{{}}\right] H\left( r\right)
\notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}+\frac{1}{2}\sum_{i=4}^{6}\sum_{j=0}^{2}%
\frac{e_{2,i}}{\left( -1\right) ^{i}}\left[ \frac{{}}{{}}G_{2}\left(
t_{j}\right) w\left( r+1,t_{j},i,1\right) \right. \notag \\
&&\left. +G_{1}\left( t_{j}\right) w\left( r+1,t_{j},i,2\right) +G_{0}\left(
t_{j}\right) w\left( r+1,t_{j},i,3\right) \frac{{}}{{}}\right] H\left(
r+1\right) , \label{Q2pr}\end{aligned}$$$$\begin{aligned}
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}\frac{Q_{p}^{\left( 2\right) }\left(
-r\right) }{24\eta }=p^{\left( 2\right) }\left( -r\right)
-\sum_{i=1}^{6}\sum_{j=0}^{2}e_{0,i}\left[ \frac{{}}{{}}C_{3}\left(
t_{j}\right) w\left( r+\lambda -1,-t_{j},i,1\right) \right. \notag \\
&&\left. -C_{2}\left( t_{j}\right) w\left( r+\lambda -1,-t_{j},i,2\right)
+C_{1}\left( t_{j}\right) w\left( r+\lambda -1,-t_{j},i,3\right) \frac{{}}{{}%
}\right] H\left( r+\lambda -1\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}-\sum_{i=2}^{6}\sum_{j=0}^{2}e_{1,i}\left[
\frac{{}}{{}}C_{3}\left( t_{j}\right) w\left( r+\lambda -2,-t_{j},i,1\right)
\right. \notag \\
&&\left. -C_{2}\left( t_{j}\right) w\left( r+\lambda -2,-t_{j},i,2\right)
+C_{1}\left( t_{j}\right) w\left( r+\lambda -2,-t_{j},i,3\right) \frac{{}}{{}%
}\right] H\left( r+\lambda -2\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}-\sum_{i=4}^{6}\sum_{j=0}^{2}e_{2,i}\left[
\frac{{}}{{}}C_{3}\left( t_{j}\right) w\left( r+\lambda -3,-t_{j},i,1\right)
\right. \notag \\
&&\left. -C_{2}\left( t_{j}\right) w\left( r+\lambda -3,-t_{j},i,2\right)
+C_{1}\left( t_{j}\right) w\left( r+\lambda -3,-t_{j},i,3\right) \frac{{}}{{}%
}\right] H\left( r+\lambda -3\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}-\frac{1}{2}\sum_{i=1}^{6}%
\sum_{j=0}^{2}e_{0,i}\left[ \frac{{}}{{}}G_{2}\left( t_{j}\right) w\left(
r+1,-t_{j},i,1\right) \right. \notag \\
&&\left. -G_{1}\left( t_{j}\right) w\left( r+1,-t_{j},i,2\right)
+G_{0}\left( t_{j}\right) w\left( r+1,-t_{j},i,3\right) \frac{{}}{{}}\right]
H\left( r+1\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}-\frac{1}{2}\sum_{i=2}^{6}%
\sum_{j=0}^{2}e_{1,i}\left[ \frac{{}}{{}}G_{2}\left( t_{j}\right) w\left(
r,-t_{j},i,1\right) \right. \notag \\
&&\left. -G_{1}\left( t_{j}\right) w\left( r,-t_{j},i,2\right) +G_{0}\left(
t_{j}\right) w\left( r,-t_{j},i,3\right) \frac{{}}{{}}\right] H\left(
r\right) \notag \\
&&{\!\!\!\!\!\!\!\!\!\!\!\!\!\!}-\frac{1}{2}\sum_{i=4}^{6}%
\sum_{j=0}^{2}e_{2,i}\left[ \frac{{}}{{}}G_{2}\left( t_{j}\right) w\left(
r-1,-t_{j},i,1\right) \right. \notag \\
&&\left. -G_{1}\left( t_{j}\right) w\left( r-1,-t_{j},i,2\right)
+G_{0}\left( t_{j}\right) w\left( r-1,-t_{j},i,3\right) \frac{{}}{{}}\right]
H\left( r-1\right) . \label{Q2pmr}\end{aligned}$$The above expressions are then substituted into inverse of Eq. (\[Qps\]), $$Q_{p}\left( r\right) =Q_{p}^{\left( 1\right) }\left( r\right) +Q_{p}^{\left(
2\right) }\left( r\right) , \label{Qpr}$$where $Q_{p}^{\left( 1\right) }\left( r\right) $ is the inverese Laplace transform of the function $\hat{Q_{p}}^{(1)}\left( s\right) =24\eta \hat{%
Q_{0}^{2}}\left( -s\right) \hat{p}^{\left( 1\right) }\left( s\right) $ and describes the contribution of the first term of the expansion (\[h\_1\_Ll2\]); the expression for this contribution was given in Ref.[@tang2007dcf]. Then the first-order perturbation contribution to the DCF of the SW model fluid with the radius of interaction that exceeds two particle diameters reads, $$rc_{1}\left( r\right) =\beta \varepsilon \left( 1-\eta \right) ^{4}\left[
Q_{p}\left( r\right) -Q_{p}\left( -r\right) \right] .$$The equations (\[Q2pr\]) and (\[Q2pmr\]) have been presented here in a similar manner as it has been done in the work of Tang and Lu[tang2007dcf]{} for the case of the SW model fluid with the range of interaction that does not exceed two particle diameters, i.e. for $\ \lambda
\leq 2.$ The resulting expression for $c_{1}\left( r\right) $ in the case of $\lambda
>2$ is consequently rather long, but also quite straightforward. It may be summarized as follows, $$rc_{1}\left( r\right) =\left\{
\begin{array}{ll}
0, & r>\lambda , \\
\beta \varepsilon r, & 1<r\leq \lambda , \\
\beta \varepsilon \left( 1-\eta \right) ^{4}\left[ Q_{p}\left( r\right)
-Q_{p}\left( -r\right) \right] , & r\leq 1,%
\end{array}%
\right. \label{rc1f}$$where $Q_{p}\left( r\right) $ is given by Eq. (\[Qpr\]). The direct correlation function is discontinuous at $r=1$ and $r=\lambda $. The continuity of the DCF inside the core, $r\in \left[ 0,1\right] $, is evident from Eqs. ([Qp2s\_prime]{}) and (\[Qp2ms\_prime\]), where the terms, that might contribute to the discontinuities at $r=\lambda -2$ and $r=3-\lambda $, are at least of the second order in ${1}/{s}$.
Results and discussions
=======================
To examine the evolution experienced by the DCF of the SW fluid upon increase of the attractive well width, in Fig. \[fig:DCF\_SWr8\] we show the first-order DCF $c_{1}\left( r\right) $ evaluated for nine different values of width parameter $\lambda $ in the range from $\lambda =1.5$ to $\lambda =3.0.$ All calculations are performed for fixed density $\rho =0.75$ and temperature $\beta \varepsilon
=0.5$. Note, that the dependence of $c_{1}\left( r\right) $ on the attractive well width $\lambda $ in the core region $r<1$ is not trivial. It is very well illustrated by the behavior of the value of $c_{1}\left(
r=0\right).$ One can see that value of $c_{1}\left( r=0\right)$ is first decreasing upon $\lambda $ increases from $\lambda =1.5$ to $\lambda \approx 1.7$ and after this starts to grow with $\lambda $ increases. However, after reaching the value of $\lambda \approx
2.2$, $c_{1}\left( r=0\right)$ starts to drop till the value of $\lambda \approx 2.6$ after which it begins to grow again.
In Figures \[fig:dcf\_rdf\_l25\] and \[fig:dcf\_rdf\_l21\] we show the total DCF $c\left( r\right) $ and total radial distribution function (RDF) $%
g\left( r\right) =h_{1}\left( r\right) +1$ of the SW fluid with attractive well width $\lambda =2.1$ and $2.5$ at the density $\rho =0.75.$ Both functions are obtained within the FMSA and from Monte-Carlo simulations. The FMSA results for the first-order correction term $h_{1}\left( r\right) $ to the RDF were obtained numerically. The correspondin DCF and RDF of the hard-sphere reference system, $c_{0}\left( r\right) $ and $g_{0}\left(
r\right) $, respectively, were obtained from the first-order GMSA theory of Tang and Lu [@tang1995ier]. The results for the intermediate and low densities are not shown, since at such conditions the considered SW fluid seems to be unstable. Good agreement between Monte Carlo simulation data and FMSA theory results could be observed for the SW fluid with $\lambda =2.5$ (see Fig. \[fig:dcf\_rdf\_l25\]). The same doesn’t hold for the SW fluid with $\lambda =2.1$. The discrepancies between FMSA and simulation predictions are clearly visible for both DCF and RDF at the particle contact $r=1$ and at the point of discontinuity at $r=\lambda =2.1$. These shortcomings in the case of RDF can be corrected by employing exponential (EXP) [@chandler1972oce] $$g^{EXP}\left( r\right) =g_{0}\left( r\right) e^{h_{1}\left( r\right) }
\label{EXP_appr}$$or linearized exponential (LEXP) [@verlet1974ptp] $$g^{LEXP}\left( r\right) =g_{0}\left( r\right) \left( 1+h_{1}\left( r\right)
\right) \label{LEXP_appr}$$approximations. Indeed, the LEXP approximation significantly improves the RDF of the SW fluid, giving better contact value at $r=1$ and improving RDF values on both sides of the discontinuity at $r=\lambda =2.1$ (see Fig. [fig:dcf\_rdf\_l21]{}, right panel).
In Fig. \[fig:Compress\] we present the inverse reduced isothermal compressibilities of the SW fluids that was obtained from
$$\frac{1}{\chi _{T}}=\beta \left( \frac{\partial P}{\partial \rho }\right)
_{T}=1-4\pi \rho \int c\left( r\right) r^{2}dr, \label{fluct_eq}$$
The DCFs of the reference system of hard spheres were obtained from the first-order GMSA theory of Tang and Lu [@tang1995ier]. From Fig. [fig:Compress]{} we note that at the high densities ($\eta\approx 0.5$) the compressibility of the SW fluid with $\lambda=2.5$ increases rapidly and almost reaches the level of the fluid $\lambda=2.0$. Mathematically, this can be explained by positive contribution into compressibility of the first-order DCF in the $r<1$ region, which compensates the negative contribution due to the potential well. On Fig.\[fig:spinodal\] we present spinodal curves and critical points for few long-range SW fluids obtained from condition $\left( \frac{\partial P}{\partial \rho}\right)_{T}=0 $ and (\[fluct\_eq\]).
Conclusions
===========
We extended the FMSA theory to deal with the square-well fluids of large wells width ($1<\lambda <3$) than those studied previously and obtained analytical expressions for the DCF. We obtained reasonable agreement of the DCF and RDFs with MC simulation data and proposed exponential (EXP) and linearized exponential (LEXP) approximations to correct RDF values close to the contact.
Acknowledgments {#acknowledgments .unnumbered}
===============
AT thanks MCSU for the hospitality while visiting the Department for the Modelling of Physico-Chemical Processes when this project has been initiated.
Expansions of the Laplace transform of the Baxter factorization function
========================================================================
Laplace transforms of the Baxter hard-sphere factorization function and square of the transform are $$\hat{Q_0}\left(s\right)=1+\sum_{i=1}^3\frac{a_i}{s_i} +\sum_{i=2}^3\frac{b_i%
}{s_i}e^{-s}, \label{Q_0app}$$ and $$\hat{Q_0}^2\left(s\right)=1+\sum_{i=1}^6\frac{e_{0,i}}{s^i} +\sum_{i=2}^6%
\frac{e_{1,i}}{s^i}e^{-s} +\sum_{i=4}^6\frac{e_{2,i}}{s^i}e^{-2s},
\label{Q0s2app}$$ where $$\begin{aligned}
&& b_2=\frac{12\eta\left( 1+\frac{\eta}{2} \right)}{\left( 1-\eta \right)^2}%
, \quad b_3=\frac{12\eta\left(1+{2\eta}\right)}{\left( 1-\eta \right)^2},
\notag \\
&& a_1=b_2-\frac{b_3}{2},\quad a_2=b_3-b_2,\quad a_3=-b_3, \label{ai_bi_app}\end{aligned}$$ and $$\begin{aligned}
&& e_{0,1}=2a_1,\quad e_{0,2}=a_1^2+2a_2, \quad e_{0,3}=2a_3+2a_1a_2, \notag
\\
&& e_{0,4}=2a_1a_3+a_2^2,\quad e_{0,5}=2a_2a_3,\quad e_{0,6}=a_3^2, \notag
\\
&& e_{1,2}=2b_2,\quad e_{1,3}=2b_3+2a_1b_2,\quad e_{1,4}=2a_1b_3+2a_2b_2,
\notag \\
&& e_{1,5}=2a_2b_3+2a_3b_2,\quad e_{1,6}=2a_3b_3, \notag \\
&& e_{2,4}=b_2^2,\quad e_{2,5}=2b_2b_3,\quad e_{2,6}=b_3^2. \notag
\label{e_ni_app}\end{aligned}$$
[99]{} Glaser, M.A., et al, *EPL*, **78**, 46004 (2007)
Franzese, G.; Malescio, G.; Skibinsky, A.; Buldyrev, S. and Stanley, H., * Nature*, **409**, 6821 (2001) 692–695.
Sinanogu, O. and Pitzer, K., *J. Chem. Phys.*, **32**, 1279–1288 (1960).
W. Rżysko, O. Pizio, A. Patrykiejew, S. Sokołowski, J. Chem. Phys. **129**, 124502 (2008).
L. A. Cervantes, A. L. Benavides, and F. del Rio, J. Chem. Phys. **126**, 084507 (2007).
Y. Tang and J. Wu, J. Chem. Phys. **119**, 7388 (2003).
Y. Tang and J. Wu, Phys. Rev. E **70**, 011201 (2004).
Y. Tang, J.Chem. Phys. **121**, 10605 (2004).
J. Mi, Y.Tang, and C. Zhong, J. Chem. Phys. **128**, 0545032 (2008).
Y. Tang, Mol. Phys. **106**, 2431 (2008).
Mej[i]{}a-Rosales, S.; Gil-Villegas, A.; Ivlev, B. and Ruiz-Garcia, J., *J. Phys.: Condens. Matter*, **14**, 4795–4804 (2002).
Benavides, A.; del Pino, L.; Gil-Villegas, A. and Sastre, F., *J. Chem. Phys.*, **125**, 204715 (2006).
Lee, B.; Kim, S.; Lee, C. and Seong, B., *Journal of the Korean Physical Society*, **36**(1), 18–22 (2000).
Henderson, D., *Fundamentals of Inhomogeneous Fluids*, CRC Press, 1992.
Oxtoby, D. W., *Annu. Rev. Mater. Res.*, **32**, 39–52 (2002).
Tang, Y. and Lu B.C.Y., *J. Chem. Phys.*, ** 99**, 9828 (1993).
Tang, Y. and Lu B.C.Y, *J. Chem. Phys.*, ** 100**, 6665 (1994).
Tang, Y., *J. Chem. Phys.*, **127**, 164504 (2007).
Tang, Y. and Lu B.C.Y, *J. Chem. Phys.*, ** 100**, 3079 (1994).
Tang, Y. and Lu B.C.Y, *J. Chem. Phys.*, ** 103**, 7463 (1995).
Guillen-Escamilla, I.; Chavez-Paez, M. and Castaneda-Priego, R., *J. Phys. Cond. Matt.*, **19**(8), 086224 (2007).
Zerah, G. and Hansen, J., *J. Chem. Phys.*, ** 84**, 2336 (1986).
Chandler, D. and Andersen, H.C., *J. Chem. Phys.*, **57**, 1930 (1972)
Verlet, L. and Weis, J.J., *Molecular Physics*, **28**, 665 (1974).
Barker, JA and Henderson, D., *Reviews of Modern Physics*, **48**, 587 (1976)
FIGURE CAPTIONS
**Fig. \[fig:DCF\_SWr8\] First-order DCF for several $\lambda$ values with $\varepsilon\beta=0.5$ and $\rho=0.75$.**
**Fig. \[fig:dcf\_rdf\_l25\] Full DCF (left panel) and full RDF (right panel) for SW fluid with $\lambda=2.5$, $\varepsilon\beta=0.5$ and $\rho=0.75$. Lines denote FMSA results, symbols denote simulation results.**
**Fig. \[fig:dcf\_rdf\_l21\] Full DCF (left panel) and full RDF (right panel) for SW fluid with $\lambda=2.1$, $\varepsilon\beta=0.5$ and $\rho=0.75$. Lines denote FMSA results, symbols denote simulation results. Dotted line denotes linearized exponential approximation (LEXP).**
**Fig. \[fig:Compress\] Inverse reduced isothermal compressibility $\chi_T^{-1}=\beta\left( \frac{\partial P}{\partial\rho}
\right)_{T}$ of the SW fluids at $\varepsilon\beta=0.5$ and $%
\varepsilon\beta=0.25$ .**
**Fig. \[fig:spinodal\] Spinodal curves for SW fluids with $\lambda=2.0, 2.2, 2.4, 2.5, 2.6, 2.8, 3.0$.**
![[]{data-label="fig:DCF_SWr8"}](fig1.eps){width="12cm"}
![[]{data-label="fig:dcf_rdf_l25"}](fig2__3.eps){width="12cm"}
![[]{data-label="fig:dcf_rdf_l21"}](fig3__3.eps){width="12cm"}
![[]{data-label="fig:Compress"}](fig10.eps){width="12cm"}
![[]{data-label="fig:spinodal"}](fig14.eps){width="12cm"}
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We analyze the theoretical description of radiative decays in charmonium. We use an elementary emission decay model to build the most general electromagnetic transition operator. We show that accurate results for the widths can be obtained from a simple quark potential model reasonably fitting the spectroscopy if the complete form of the operator is used instead of its standard p/m approximation and the experimental masses are properly implemented in the calculation.'
author:
- 'R. Bruschini'
- 'P. González'
bibliography:
- 'emcharmbib.bib'
title: 'Radiative decays in charmonium beyond the p/m approximation'
---
Introduction\[SI\]
==================
Electromagnetic decays in heavy quarkonium (bottomonium or charmonium) may play a key role in the understanding of its structure. The current impossibility to directly solve QCD, the theory of strong interactions, for the description of hadrons, forces us to rely, for the knowledge of such structure, on models and/or effective theories, incorporating at the greater extent the properties of QCD (see for instance [@Bra04; @Bra11] and references therein). Among these approximations the most successful one regarding the number of described heavy quarkonium states below the open flavor meson-meson thresholds is undoubtedly the constituent quark model, see [@Bra04] and references therein, where heavy quarkonium is described as a quark-antiquark bound system. Then, as the electromagnetic transition operator is known (with no free parameter) the comparison of the measured widths with their calculated values from different spectroscopic quark models may be an ideal test to discriminate between these models and to advance in the understanding of the heavy quarkonium structure.
In a recent paper [@Br3-19], we have shown that this discrimination may be rather difficult in bottomonium. By using the standard expansion of the electromagnetic transition operator up to $\vb*{p}_{b}/M_{b}$ order, where $\vb*{p}_{b}$ $(M_{b})$ stands for three-momentum (mass) of the $b$ quark, we have shown that accurate results for the widths can be obtained from different quark potential models reasonably fitting the spectroscopy once the experimental masses of the bottomonium states instead of the calculated ones are properly implemented in the calculation. The argument justifying this substitution (and giving meaning to the qualificative “reasonably fitting the spectroscopy” employed) is that the experimental masses can be hopefully obtained from the wavefunctions of the calculated states by applying first order perturbation theory. Therefore, the implementation of the experimental masses allows for a direct test of the quark model wavefunctions.
In this article we analyze electromagnetic decays in charmonium, focusing on the quantitatively most relevant electromagnetic transitions, ${{^{3}\!S_1} \leftrightarrow {^{3}\!P_J}}$, for which there are data available. We first show that the $\vb*{p}_{c}/M_{c}$ order approximation, where the subindex $c$ stands for the $c$ quark, does not give rise to such an accurate description of the decay widths as in bottomonium. This could be somehow expected since the expectation value of $\abs{\vb*{p}_{c}} /M_{c}$, representing the speed of the quark, can be about half the speed of light for the low lying charmonium states, what makes the use of the transition operator up to the $\vb*{p}_{c}/M_{c}$ order debatable. We proceed then to build the complete transition operator and to apply it to the calculation of the decay widths. This allows us to to discriminate between different quark models according to their accuracy in the description of radiative decays.
These contents are organized as follows. In Section \[SII\] we detail the Cornell potential models we use to calculate the masses and wavefunctions of the low lying charmonium states. We expect that the experimental masses can be obtained from the calculated values via corrections to the Hamiltonian evaluated at first order in perturbation theory. In Section \[SIII\] the elementary emission model for electromagnetic transitions is developed in some detail. From it we recover the usual $\vb*{p}_{c}/M_{c}$ approximation in Section \[SIV\]. The comparison of the calculated decay widths with data points out the need to go beyond this approximation. Then, in Section \[SV\] the complete transition operator, to all $\vb*{p}_{c}/M_{c}$ orders, is built. The transition amplitude and the explicit form of its electric and magnetic contributions are detailed in Section \[SV\] and applied in Section \[SVI\] to the calculation of radiative decay widths which are compared to data. Finally, in Section \[SVII\] our main results and conclusions are summarized.
Spectroscopic Quark Models\[SII\]
=================================
For the description of charmonium we shall use a nonrelativistic quark potential model framework defined by the Hamiltonian $$H=\frac{\vb*{p}^{2}}{M_{c}}+V(r),
\label{ham}$$ with a Cornell potential energy $$V(r) =\sigma r-\frac{\zeta}{r}+\beta
\label{pot}$$ where $\vb*{p}$ is the relative momentum operator, $r=\abs{\vb*{r}}$ is the $c-\overline{c}$ distance operator ($\vb*{r}$ is the relative position operator), the parameters $\sigma$ and $\zeta$ stand for the string tension and the chromoelectric coulomb strength respectively, and $\beta$ is a constant to fix the origin of the potential. It is important to emphasize that
i) this potential form arises from spin independent quenched lattice QCD calculations in the Born-Oppenheimer approximation [@Bal01],
ii) in the spirit of the nonrelativistic quark model calculations $\sigma$, $\zeta$, $\beta$ and the quark mass $M_{c}$ should be considered as effective parameters through which spin dependent and/or spin independent corrections may be implicitly incorporated.
Henceforth we shall make use of two different quark models with the same Hamiltonian form , Model I and Model II, that have been used for the analysis of radiative decays in bottomonium [@Br3-19]. As we are dealing with a radial potential we shall denote the spectroscopic states by $n\,{^{2s+1}\!L_{J}}$ where $s,$ $L$ and $J$ stand for the spin, orbital angular momentum and total angular momentum quantum numbers respectively.
Model I, providing a good description of spin triplet state masses in bottomonium [@Gon14] is defined in charmonium [@Gon15] by the set of parameter values $$\begin{split}
\sigma_{I}&=850\,{\text{MeV}}/{\text{fm}},\\
\zeta_{I} &=100 \, {\text{MeV}}\cdot {\text{fm}}, \\
(M_{c})_{I}&=1348.6 \, {\text{MeV}},
\end{split}
\label{par1}$$ giving account of the mass differences between some of the low lying (spin triplet) charmonium states whose electromagnetic transitions are measured, more precisely between $2{^{3}\!S_{1}}$ and $1{^{3}\!P_{1}}$, and between $1{^{3}\!P_{1}}$ and $1{^{3}\!S_{1}}$. Hence, the model also describes accurately, through a convenient choice of the additive constant $(\beta_{c}) _{I}$, the masses of the $1{^{3}\!S_{1}}$, $2{^{3}\!S_{1}}$ and $1{^{3}\!P_{1}}$ states. Furthermore, inasmuch as the mass splittings between $P$ states can be obtained via first order perturbation theory the $1{^{3}\!P_{0,2}}$ states would be described by the same wavefunction as the $1{^{3}\!P_{1}}$ one. In Table \[tabmassdif1\] we list the ratios of the calculated mass differences to the experimental ones.
$i-f$ $\frac{(M_{i}-M_{f})_I^{\text{Theor}}}{(M_{i}-M_{f})^{\text{Expt}}}$
--------------------------------- ----------------------------------------------------------------------
$2{^{3}\!S_{1}}-1{^{3}\!P_{1}}$ $1.00$
$1{^{3}\!P_{1}}-1{^{3}\!S_{1}}$ $0.99$
$2{^{3}\!S_{1}}-1{^{3}\!S_{1}}$ $0.99$
$2{^{3}\!S_{1}}-1{^{3}\!P_{2}}$ $1.09$
$2{^{3}\!S_{1}}-1{^{3}\!P_{0}}$ $0.65$
$1{^{3}\!P_{0}}-1{^{3}\!S_{1}}$ $1.29$
$1{^{3}\!P_{2}}-1{^{3}\!S_{1}}$ $0.89$
: Ratios of the calculated mass differences $(M_{i}-M_{f})_I^{\text{Theor}}$ between $1{^{3}\!S_{1}},$ $2{^{3}\!S_{1}}$, $1{^{3}\!P_{1}}$ charmonium states from Model I as compared to the experimental ones $( M_{i}-M_{f})^{\text{Expt}}$ taken from [@PDG18].[]{data-label="tabmassdif1"}
Model II, defined in reference [@Eic94] by the set of parameter values $$\begin{split}
\sigma_{II}&=925.5\,{\text{MeV}}/{\text{fm}}, \\
\zeta_{II}&=102.6 \, {\text{MeV}}\vdot {\text{fm}}, \\
(M_{c})_{II}&=1840 \, {\text{MeV}},
\end{split}
\label{par}$$ is based on the assumption that mass corrections to the charmonium states calculated from may have to do mainly with non considered spin dependent terms in the potential, so that the quark model should fit the centers of gravity of spin triplet and spin singlet states.
Let us realize that the chosen value for the string tension, $\sqrt{\sigma_{II}} = 427.4$ MeV, agrees with the one derived from the analysis of Regge trajectories in light mesons [@Bal01], and that the Coulomb strength $\zeta_{II}=102.6 {\text{MeV}}{\text{fm}}$ corresponds to a strong quark-gluon coupling $\alpha_{s}=\frac{3\zeta}{4\hbar}\simeq 0.39$ quite in agreement with the value derived in QCD from the fine structure splitting of $1P$ states in charmonium [@Bad99]. As for $M_{c}$ the chosen value gives mass differences between any two of the centers of gravity of the $1S$, $2S$ and $1P$ states in accord with data within a $10\%$ of accuracy. In Table \[tabmassdif\] we list the ratios of the calculated center of gravity mass differences to the experimental ones. One could think of refining the values of the parameters to make all these ratios closer to $1$. This can only be done at the price of loosing the close connection of $\sigma_{II}$ and/or $\zeta_{II}$ with their expected phenomenological values. Instead, we prefer to maintain such connection and the modest discrepancy reflected in Table \[tabmassdif\]. Moreover, the chosen values of $\sigma_{II}$ and $\zeta_{II}$ have been used in bottomonium for an accurate description of electromagnetic decay widths [@Br3-19] what adds interest to the possibility of getting such an accurate description in charmonium as well.
$i-f$ $\frac{(\overline{M}_{i}-\overline{M}_{f})_{II}^{\text{Theor}}}{(\overline{M}_{i}-\overline{M}_{f})^{\text{Expt}}}$
--------- ---------------------------------------------------------------------------------------------------------------------
$2S-1S$ $0.98$
$2S-1P$ $1.09$
$1P-1S$ $0.94$
: Ratios of the calculated center of gravity mass differences $(\overline{M}_{i}-\overline{M}_{f})_{II}^{\text{Theor}}$ between $1S$, $2S$, $1P$ charmonium states from Model II as compared to the experimental ones $(\overline{M}_{i}-\overline{M}_{f})^{\text{Expt}}$ taken from [@PDG18].[]{data-label="tabmassdif"}
It is worth to point out that the quark model wavefunctions are the same for any value of the additive constant $\beta$. Indeed, we could change $(\beta_{c})_{II}$ to obtain from Model II an approximate mass description of the $1{^{3}\!S_{1}},$ $2{^{3}\!S_{1}}$ and $1{^{3}\!P_{1}}$ states, or we could change $(\beta_{c})_{I}$ to obtain from Model I the centers of gravity of the $1S,$ $2S$ and $1P$ states in accord with data within a $5\%$ of accuracy. In this sense, Model I and Model II are quite equivalent with respect to the mass description of the low lying charmonium states except for $2{^{3}\!P_J}$, as shown in Table \[tabIII\]. Hence, any significant difference in their predictions for other observables involving these states can put severe constraints on the possible values of the parameters, in particular on the charm mass whose value in Model I is very different to that in Model II.
$n\,{^{2s+1}\!L_{J}}$ $(M_{n\,{^{2s+1}\!L_{J}}})_{I}$ $(M_{n\,{^{2s+1}\!L_{J}}})_{\text{Expt}}$ $(M_{n\,{^{2s+1}\!L_{J}}})_{II}$
----------------------- --------------------------------- ------------------------------------------- ----------------------------------
$1{^{3}\!S_{1}}$ $3099$ $3096.916\pm0.011$ $3088$
$2{^{3}\!S_{1}}$ $3685$ $3686.09\pm0.04$ $3678$
$1{^{3}\!P_{0}}$ $3509$ $3414.75\pm0.31$ $3516$
$1{^{3}\!P_{1}}$ $3509$ $3510.66\pm0.07$ $3516$
$1{^{3}\!P_{2}}$ $3509$ $3556.20\pm0.09$ $3516$
$2{^{3}\!P_{0}}$ $3911$ $3862_{-32-13}^{+26+40}$ $3959$
$2{^{3}\!P_{1}}$ $3911$ $3959$
$2{^{3}\!P_{2}}$ $3911$ $3927.2\pm2.6$ $3959$
: Calculated masses, $M_{n\,{^{2s+1}\!L_{J}}}$ for the low lying spin triplet states. The subscripts $I$ and $II$ refer to Model I with $(\beta_{c}) _{I}=53 \, {\text{MeV}}$ and Model II with $(\beta_{c})_{II}=-850 \, {\text{MeV}}$ respectively. Experimental masses from [@PDG18]. We have not quoted any mass for the experimental $2{^{3}\!P_{1}}$ state since the $X(3872)$ can not be considered a pure Cornell state.[]{data-label="tabIII"}
It should be also remarked that simplicity is not the only argument to use a Cornell potential model for the study of radiative decays. The absence of any momentum dependence in the potential allows for a complete factorization of the heavy quarkonium mass dependence in the calculation of the electromagnetic decay widths up to $\vb*{p}_{c}/M_{c}$ order. Furthermore, such a complete factorization can be also pursued to higher orders as will be shown later on. Then, if the experimental masses instead of the calculated ones are implemented in the calculation the comparison of the calculated decay widths to data becomes a powerful tool to test the model wavefunctions. As a counterpart, the use of the Cornell potential in charmonium restricts the study to the low lying states where meson-meson threshold effects can be neglected.
Electromagnetic Decay Model\[SIII\]
===================================
Let us consider the decay $I\rightarrow\gamma F$ where $I$ and $F$ are the initial and final charmonium states respectively. In the rest frame of the decaying meson $I$ the total width is given by (we follow the PDG conventions, see [@PDG18 p. 567]) $$\Gamma_{I\rightarrow\gamma F}=\frac{k_{0}}{8\pi M_{I}^{2}}\frac{1}{(2J_{I}+1)}\sum_{\lambda=\pm1}\sum_{m_{I},m_{F}}\abs{\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda}}^{2},
\label{width}$$ where $k_{0}$ is the energy of the photon and $M_{I}$, $J_{I}$ and $m_{I}$ stand for the mass of $I$, its total angular momentum and its third projection respectively. The polarization of the photon is represented by $\lambda$ (as usual we choose the three-momentum of the photon in the $Z$ direction) and the transition amplitude by $\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda}$. This amplitude can be obtained from the interaction Hamiltonian $\mathcal{H}^{int}$ as $$\begin{gathered}
(2\pi)^{3}\delta^{(3)}(\vb*{P}_{I}-\vb*{k}-\vb*{P}_{F})\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda}=\\
\sqrt{2M_{I}}\sqrt{2E_{F}}\sqrt{2k_{0}}\mel{F\gamma}{\mathcal{H}^{int}}{I},
\label{deltas}\end{gathered}$$ where $P_{I}=(E_{I},\vb*{P}_{I}) =(M_{I},\vb*{0})$, $P_{F}=( E_{F},\vb*{P}_{F})$ and $(k_{0},\vb*{k})$ are the meson and photon four-momenta.
In the Elementary Emission Decay Model the radiative transition $I\rightarrow\gamma F$ takes place through the emission of the photon by the quark or the antiquark in the initial state. In QED the interaction Hamiltonian at the quark level is given by $$\mathcal{H}^{int}=\int \dd{\vb*{x}} j^{\mu}(\vb*{x}) A_{\mu}(\vb*{x}),
\label{Hint}$$ where $A_{\mu}$ is the photon field and $j^{\mu}$ the electromagnetic current given by $$j^{\mu}(\vb*{x})= \overline{q}(\vb*{x}) \mathcal{Q} \gamma^{\mu}q(\vb*{x})
\label{emcurr}$$ where $q(\vb*{x})$ is the quark field and $\mathcal{Q}$ the quark charge matrix. Notice that the time dependence has been obviated since in the calculation of the transition amplitude it only gives rise to a Dirac delta accounting for energy conservation.
In quantum field theory the quark field operator with no time dependence is written as (see for example [@Pes95 p. 58], but note that we use Dirac spinors instead of the Weyl representation adopted there) $$\begin{gathered}
q(\vb*{x}) =\int\frac{\dd{\vb{p}}}{(2\pi)^{3}}\frac{1}{\sqrt{2\mathrm{E}(\vb{p})}}\\
\sum_{m_{s}}\bigl( u^{m_{s}}(\vb{p}) b_{1}^{m_{s}}( \vb{p})e^{i\vb{p}\vdot\vb*{x}} + v^{m_{s}}(\vb{p}) b_{2}^{m_{s}\dagger}(\vb{p})e^{-i\vb{p}\vdot\vb*{x}}\bigr),
\label{qfield}\end{gathered}$$ where $\mathrm{E}(\vb{p})$ is the relativistic energy of a quark $Q$ or antiquark $\bar{Q}$, $\mathrm{E}(\vb{p}) =\sqrt{M_{Q}^{2}+\vb{p}^{2}}$, $u^{m_{s}}(\vb{p})$ and $v^{m_{s}}(\vb{p})$ are the Dirac spinors $$u^{m_{s}}(\vb{p}) =\sqrt{M_{Q}+\mathrm{E}(\vb{p})}\pmqty{\chi^{m_{s}}\\\frac{\vb{p}\vdot\vb*{\sigma}}{M_{Q}+\mathrm{E}(\vb{p})}\chi^{m_{s}}}
\label{w1}$$ $$v^{m_{s}}(\vb{p}) =\sqrt{M_{\bar{Q}}+\mathrm{E}(\vb{p})}\pmqty{\frac{\vb{p}\vdot\vb*{\sigma}}{M_{\bar{Q}}+\mathrm{E}(\vb{p})}\chi^{m_{s}} \\ \chi^{m_{s}}} \label{w2}$$ with $\chi^{m_{s}}$ being the Pauli spinor, and $b_{1}^{m_{s}}(\vb{p})$ $( b_{2}^{m_{s}\dagger}( \vb{p}))$ the annihilation (creation) operator of a quark (antiquark) with spin projection $m_{s}$ and three-momentum $\vb{p}$.
Then, by defining $$q_{\alpha}(\vb*{x}) =\int\frac{\dd{\vb{p}_\alpha}}{(2\pi)^{3}}\sqrt{\frac{M_{\alpha}+\mathrm{E}_\alpha(\vb{p}_\alpha)}{2\mathrm{E}_\alpha( \vb{p}_\alpha) }}\sum_{m_{s}}e^{i\vb{p}_\alpha\vdot\vb*{x}} \chi_{\alpha}^{m_{s}}b_{\alpha}^{m_{s}}(\vb{p}_\alpha),$$ where $\alpha=1,2$ refers to quark and antiquark respectively, the electromagnetic current $j^{\mu}(\vb*{x})$ reads
$$j^0( \vb*{x}) = \sum_{\alpha=1,2} e_\alpha \Biggr[q_\alpha^\dagger(\vb*{x})q_\alpha(\vb*{x})+ \pqty{\frac{\grad q_\alpha(\vb*{x})^\dagger}{M_\alpha + {E}_\alpha}\vdot\frac{\grad q_\alpha(\vb*{x})}{M_\alpha + {E}_\alpha}+ i \vb*{\sigma}\!_{\alpha}\vdot \frac{\grad q_\alpha(\vb*{x})^\dagger}{M_\alpha + {E}_\alpha} \cp \frac{\grad q_\alpha(\vb*{x})}{M_\alpha + {E}_\alpha}}\Biggr]$$
and $$\begin{gathered}
\vb*{j}(\vb*{x}) = \sum_{\alpha=1,2} e_\alpha \left\{ -i \bqty{q_\alpha^\dagger(\vb*{x}) \pqty{\frac{\grad{q_\alpha(\vb*{x})}}{M_\alpha + {E}_\alpha}}-\pqty{\frac{\grad{q_\alpha^\dagger(\vb*{x})}}{M_\alpha +{E}_\alpha}} q_\alpha(\vb*{x})}+\right.\\
\left.\bqty{\pqty{\frac{\grad{q_\alpha^\dagger(\vb*{x})}}{M_\alpha + {E}_\alpha}} \cp \vb*{\sigma}\!_{\alpha} \, q_\alpha(\vb*{x})-q_\alpha^\dagger(\vb*{x}) \vb*{\sigma}\!_{\alpha} \cp \pqty{\frac{\grad{q_\alpha(\vb*{x})}}{M_\alpha + {E}_\alpha}}}\right\},\end{gathered}$$ where $e_1 = e_{c}=\frac{2}{3}e$ (with $e=\sqrt{4\pi\alpha_{em}}$ being $\alpha_{em}$ the fine structure constant), $e_2 = e_{\overline{c}} = - e_c$, and $E_{\alpha}=\sqrt{M_{\alpha}^{2}-\vb*{\nabla}_\alpha^{2}}$.
It is important to highlight that in the derivation of the current operator we have kept only terms that conserve separately the number of quarks and antiquarks, since we are only interested in radiative processes where there is no quark-antiquark photoproduction or annihilation.
In order to calculate the matrix element $\mel{F\gamma}{\mathcal{H}^{int}}{I}$ from the spectroscopic quark model wavefunctions one needs the “first quantized” form of the interaction. Following the procedure explained in [@LeY88] and detailed in Appendix \[firstop\], we obtain the first quantized form of the current (in Appendix \[gaugeinv\] it is checked that this current is conserved, as required by gauge invariance) $$\begin{gathered}
j^0_\text{1st}(\vb*{x}) = \sum_{\alpha=1,2} e_\alpha \sqrt{\frac{M_\alpha + {{E}}_\alpha}{2 {{E}}_\alpha}} \Biggr[\delta^{(3)}(\vb*{x}-\vb*{r}_\alpha)+ \\
\left(\frac{{\vb*{p}}_\alpha}{M_\alpha + {{E}}_\alpha}\vdot \delta^{(3)}(\vb*{x}-\vb*{r}_\alpha) \frac{{\vb*{p}}_\alpha}{M_\alpha + {{E}}_\alpha}+ i \vb*{\sigma}\!_{\alpha}\vdot \frac{{\vb*{p}}_\alpha}{M_\alpha + {{E}}_\alpha} \cp \delta^{(3)}(\vb*{x}-\vb*{r}_\alpha) \frac{{\vb*{p}}_\alpha}{M_\alpha + {{E}}_\alpha}\right)\Biggr] \sqrt{\frac{M_\alpha + {{E}}_\alpha}{2 {{E}}_\alpha}},
\label{cur0}\end{gathered}$$ $$\begin{gathered}
\vb*{j}_{\text{1st}}(\vb*{x}) = \sum_{\alpha=1,2} e_\alpha \sqrt{\frac{M_\alpha + {{E}}_\alpha}{2 {{E}}_\alpha}} \left[\pqty{\delta^{(3)}(\vb*{x}-\vb*{r}_\alpha) \frac{{\vb*{p}}_\alpha}{M_\alpha + {{E}}_\alpha} + \frac{{\vb*{p}}_\alpha}{M_\alpha + {{E}}_\alpha} \delta^{(3)}(\vb*{x}-\vb*{r}_\alpha)}+\right.\\
\left.-i \vb*{\sigma}\!_{\alpha} \cp \pqty{\delta^{(3)}(\vb*{x}-\vb*{r}_\alpha) \frac{{\vb*{p}}_\alpha}{M_\alpha + {{E}}_\alpha} - \frac{{\vb*{p}}_\alpha}{M_\alpha + {{E}}_\alpha} \delta^{(3)}(\vb*{x}-\vb*{r}_\alpha)}\right]\sqrt{\frac{M_\alpha + {{E}}_\alpha}{2 {{E}}_\alpha}}.
\label{curvec}\end{gathered}$$
Then the operator to be sandwiched between the meson states reads $$\begin{gathered}
\mel{\vb*{k},\lambda}{\mathcal{H}^{int}_\text{1st}}{0} = -\frac{1}{\sqrt{2 k_0}} \sum_{\alpha=1,2} e_\alpha \sqrt{\frac{M_\alpha + \widehat{E}_\alpha}{2 \widehat{E}_\alpha}} \\
\left[e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_\alpha}\frac{\widehat{\vb*{p}}_\alpha}{M_\alpha+\widehat{E}_\alpha} + \frac{\widehat{\vb*{p}}_\alpha}{M_\alpha+\widehat{E}_\alpha}e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_\alpha} -i \vb*{\sigma}\!_{\alpha} \cp \pqty{e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_\alpha} \frac{\widehat{\vb*{p}}_\alpha}{M_\alpha+\widehat{E}_\alpha} - \frac{\widehat{\vb*{p}}_\alpha}{M_\alpha+\widehat{E}_\alpha} e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_\alpha}}\right] \\
\vdot ( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*} \sqrt{\frac{M_\alpha + \widehat{E}_\alpha}{2 \widehat{E}_\alpha}},
\label{comop}\end{gathered}$$
where we have put a hat above $E$ and $\vb*{p}$ and $\vb*{r}$ to make more clear that they are operators, in contrast to $\vb*{k}$ which is a vector number.
It is a common practice in the analysis of radiative transitions in heavy quarkonium to proceed to a nonrelativistic reduction of this operator up to $\frac{\vb*{p}_{c}}{M_{c}}$ order. This has been justified for the use of the nonrelativistic Schrödinger equation for the calculation of the states, see for example [@LeY88]. However, as the effectiveness of the quark model parameters and the implementation of the experimental masses in the calculation of radiative decays may incorporate in an effective manner relativistic effects the restriction of the operator to the $\frac{\vb*{p}_{c}}{M_{c}}$ order may be questionable, at least for charmonium where the speed of the quarks in the low lying states is not much smaller than the speed of light in vacuum. Therefore we proceed next to the evaluation of the transition amplitude $\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda}$ , first in the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation and then to all $\frac{\vb*{p}_{c}}{M_{c}}$ orders. As we shall see, the consideration of the higher $\frac{\vb*{p}_{c}}{M_{c}}$ orders becomes essential for an accurate description of the decay widths.
The p/m approximation\[SIV\]
============================
The $\frac{\vb*{p}_{c}}{M_{c}}$ approximation is defined from in the limit $\widehat{E}_{\alpha}=M_{\alpha}$. A thorough analysis of this approximation has been carried out in [@Br3-19]. Here we only give the final expressions for the calculation of the amplitude: $$\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda}=\sqrt{2M_{I}}\sqrt{2E_{F}}\sum_{\alpha=1,2}\frac{e_{\alpha}}{2M_{c}}\mel{\Psi_{F}}{\mathcal{O}_{\alpha}}{\Psi_{I}},
\label{p/mamp}$$ where $$\ket{\Psi} \equiv \ket{J,m,nL,s}
\label{wf}$$ stands for the $c\overline{c}$ spectroscopic $n\,{^{2s+1}\!L_{J}}$ state and $\mathcal{O}_{\alpha}$ for an operator that we detail next.
Thus, for $^{3}\!S_{1}\rightarrow\gamma\,{^{3}\!P_{J}}$ transitions we have (here we write only the electric part of the operator; the form for the magnetic part can be found in the appendices of [@Br3-19]) $$\begin{split}
&\langle \mathcal{O}_{\alpha}^{el}\rangle _{\,{^{3}\!S_{1}}\rightarrow\gamma\,{^{3}\!P_{J}}} \equiv \mel{\Psi_{F( {^{3}\!P_{J}})}}{\mathcal{O}_{\alpha}^{el}}{\Psi_{I( \,{^{3}\!S_{1}}) }} \\
&= \mel{\Psi_{F( {^{3}\!P_{J}})}}{e^{i(-1) ^{\alpha}( \frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2})}(-1)^{\alpha}2\widehat
{\vb*{p}}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^*}{\Psi_{I( \,{^{3}\!S_{1}}) }}
\end{split}
\label{Osp0}$$ where $\widehat{\vb*{r}}=\frac{\widehat{\vb*{r}}_{1}-\widehat{\vb*{r}}_{2}}{2}$ is the relative position operator and $\widehat{\vb*{p}}=\frac{\widehat{\vb*{p}}_{1}-\widehat{\vb*{p}}_{2}}{2}$ is the relative three-momentum operator.
By using the equality $$\widehat{\vb*{p}}=-i\frac{M_{c}}{2} \comm{\widehat{\vb*{r}}}{H}
\label{conm}$$ where $H$ is the Cornell spectroscopic Hamiltonian, and introducing a Parseval identity in terms of a complete set of intermediate eigenstates $\Bqty{ \ket{\Psi_{int}}} $ of $H$, the mass dependence in the matrix element can be explicitly extracted: $$\begin{gathered}
\langle \mathcal{O}_{\alpha}^{el}\rangle _{\,{^{3}\!S_{1}}\rightarrow\gamma\,{^{3}\!P_{J}}}=\\
=-i M_{c}\sum_{int}\mel{\Psi_{F( {^{3}\!P_{J}})}}{e^{i(-1) ^{\alpha}(\frac{\vb*{k} \vdot\widehat{\vb*{r}}}{2})}}{\Psi_{int}} (M_{I}-M_{int}) \\
\mel{\Psi_{int}}{(-1)^{\alpha}\widehat{\vb*{r}}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}}{\Psi_{I( \,{^{3}\!S_{1}}) }}.
\label{Osp}\end{gathered}$$ This permits the implementation of the experimental mass differences $(M_{I}-M_{int})$ instead of the calculated ones so that the quark model wavefunctions can be directly tested.
As for $^{3}\!P_{J}\rightarrow\gamma\,{^{3}\!S_{1}}$ transitions, using $\comm{\widehat{\vb*{p}}}{e^{-i\vb*{k}\vdot\widehat{\vb*{r}}}}
=-\vb*{k}e^{-i\vb*{k}\vdot\widehat{\vb*{r}}}$ we get in a similar manner $$\begin{gathered}
\langle \mathcal{O}_{\alpha}^{el}\rangle _{{^{3}\!P_{J}}\rightarrow\gamma\,{^{3}\!S_{1}}}\equiv \mel{\Psi_{F( \,{^{3}\!S_{1}})}}{ \mathcal{O}_{\alpha}^{el}}{\Psi_{I({^{3}\!P_{J}})}} \\
= \mel{\Psi_{F( \,{^{3}\!S_{1}})}}{(-1) ^{\alpha}2\widehat{\vb*{p}}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}e^{i(-1)^{\alpha}(\frac{\vb*{k} \vdot\widehat{\vb*{r}}}{2})}}{\Psi_{_{I( {^{3}\!P_{J}}) }}}\\
=-i M_{c}\sum_{int}( M_{int}-M_{F}) \mel{\Psi_{F(\,{^{3}\!S_{1}})}}{(-1)^{\alpha}\widehat{\vb*{r}}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}}{\Psi_{int}}\\
\mel{\Psi_{int}}{e^{i(-1)^{\alpha}(\frac{\vb*{k} \vdot\widehat{\vb*{r}}}{2})}}{\Psi_{I({^{3}\!P_{J}})}}.
\label{Ops}\end{gathered}$$
These expressions get further simplified in the so called Long Wavelength (LWL) limit, corresponding to take $e^{i(-1)^{\alpha}(\frac{\vb*{k} \vdot\widehat{\vb*{r}}}{2})}=1$: $$\begin{gathered}
\langle ( \mathcal{O}_{\alpha}^{el}) _{LWL}\rangle_{\,{^{3}\!S_{1}}\rightarrow\gamma\,{^{3}\!P_{J}}}=\\
-i M_{c}(M_{I}-M_{F}) \mel{\Psi_{F( {^{3}\!P_{J}}) }}{(\vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}\vdot(-1)^{\alpha}\widehat{\vb*{r}}}{\Psi_{I( \,{^{3}\!S_{1}})}}
\label{LWL1},\end{gathered}$$ $$\begin{gathered}
\langle ( \mathcal{O}_{\alpha}^{el}) _{LWL}\rangle_{{^{3}\!P_{J}}\rightarrow\gamma\,{^{3}\!S_{1}}}=\\
-i M_{c}( M_{I}-M_{F})\mel{\Psi_{F( \,{^{3}\!S_{1}})}}{(\vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}\vdot(-1)^{\alpha}\widehat{\vb*{r}}}{\Psi_{I( {^{3}\!P_{J}})}}
\label{LWL2}.\end{gathered}$$ Actually this is the limit commonly used in the literature [@Eic08] for the calculation of radiative decays despite the fact that it can only be taken for granted when the condition $$\abs{\vb*{k}} ( 2 \langle \widehat{\vb*{r}}^{2}\rangle^{\frac{1}{2}}) _{F}<1
\label{criterium}$$ is satisfied [@Br3-19]. Indeed, regarding the considered transitions in charmonium, the only LWL processes are ${2{^{3}\!S_{1}}\rightarrow\gamma^{3}1P_{1,2}}$ as can be checked from Table \[tabLWL\] where the root mean square radii have been obtained from Model II (the same conclusion comes out from Model I).
$^{3}\!S_{1}\rightarrow\gamma\,{^{3}\!P_{J}}$ $\abs{\vb*{k}}_{\text{Expt}}\text{(MeV)}$ $\abs{\vb*{k}}_{\text{Expt}}( 2\left\langle r^{2}\right\rangle ^{\frac{1}{2}})_{{^{3}\!P_{J}}}$
--------------------------------------------------- ------------------------------------------- --------------------------------------------------------------------------------------------------
$\psi(2S) \rightarrow\gamma\chi_{c_{0}}( 1p) $ $261$ $1.6$
$\psi(2S) \rightarrow\gamma\chi_{c_{1}}( 1p) $ $171$ $1.0$
$\psi(2S) \rightarrow\gamma\chi_{c_{2}}( 1p) $ $128$ $0.8$
$\chi_{c_{0}}(1p) \rightarrow\gamma J/\psi$ $303$ $1.2$
$\chi_{c_{1}}(1p) \rightarrow\gamma J/\psi$ $389$ $1.5$
$\chi_{c_{2}}(1p) \rightarrow\gamma J/\psi$ $430$ $1.6$
: Experimental values of the photon energy $\abs{\vb*{k}}_{\text{Expt}}$ and calculated values of $\abs{\vb*{k}}_{\text{Expt}}( 2\left\langle r^{2}\right\rangle ^{\frac{1}{2}}) _{{^{3}\!P_{J}}}$ from Model II for $^{3}\!S_{1}\rightarrow\gamma\,{^{3}\!P_{J}}$ and $^{3}\!P_{J}\rightarrow\gamma\,{^{3}\!S_{1}}$ radiative transitions.[]{data-label="tabLWL"}
In Table \[tabp/m\] we list the calculated ${{^{3}\!S_1} \leftrightarrow {^{3}\!P_J}}$ transition widths as compared to data for charmonium. In all cases the experimental masses for the initial, final and (when known) intermediate charmonium states instead of the calculated ones from the spectroscopic Hamiltonian have been used. It turns out that the consideration of intermediate $n{^{3}\!P_{J}}$ states with $n\leq2$ is enough in the sense that the inclusion of higher Cornell states hardly changes ($2\%$ at most) the results. Regarding the intermediate $2{^{3}\!P_{J}}$ states we have used the experimental masses for $2{^{3}\!P_{0}}$, corresponding to $\chi_{c0}(3860)$ under the assumption that this resonance is a pure Cornell state, and $2{^{3}\!P_{2}}$, corresponding to $\chi_{c2}( 3930)$ that may be reasonably taken as a pure Cornell state since it lies quite below the first $S$-wave $2^{++}$ meson-meson threshold [@Gon15]. As for $2{^{3}\!P_{1}}$ we have used the calculated mass from Model I, see Table \[tabIII\], since it lies in between those of $2{^{3}\!P_{0}}$ and $2{^{3}\!P_{2}}$ as should be expected when no threshold effects are taken into account (notice that the $X(3872)$ can not be taken as a pure Cornell state).
----------------------------------------------------- --------------------------------------------------------- --------------------------------------------------------- ------------------------------ -------------------------------------------------------- ---------------------------------------------------------
Radiative Decay $(\Gamma_{LWL}^{( {\text{Theor}}-{\text{Expt}})})_{I}$ $(\Gamma_{p/M}^{( {\text{Theor}}-{\text{Expt}})})_{I}$ $\Gamma_{\text{Expt}}^{PDG}$ $(\Gamma_{p/M}^{({\text{Theor}}-{\text{Expt}})})_{II}$ $(\Gamma_{LWL}^{({\text{Theor}}-{\text{Expt}})})_{II} $
KeV KeV KeV KeV KeV
$\psi( 2S) \rightarrow\gamma\chi_{c_{0}}( 1p) $ $61$ $77$ $28.8\pm1.4$ $57$ $47$
$\psi( 2S) \rightarrow\gamma\chi_{c_{1}}( 1p) $ $53$ $52$ $28.7\pm1.5$ $41$ $41$
$\psi( 2S) \rightarrow\gamma\chi_{c_{2}}( 1p) $ $37$ $37$ $28.0\pm1.3$ $29$ $29$
$\chi_{c_{0}}( 1p) \rightarrow\gamma J/\psi$ $186$ $160$ $151\pm14$ $118$ $128$
$\chi_{c_{1}}( 1p) \rightarrow\gamma J/\psi$ $386$ $464$ $288\pm22$ $315$ $266$
$\chi_{c_{2}}( 1p) \rightarrow\gamma J/\psi$ $513$ $616$ $374\pm27$ $419$ $353$
----------------------------------------------------- --------------------------------------------------------- --------------------------------------------------------- ------------------------------ -------------------------------------------------------- ---------------------------------------------------------
A look at Table \[tabp/m\] confirms, through a comparison of the calculated LWL results (second column for Model I and sixth column for Model II) with the $\frac{\vb*{p}_{c}}{M_{c}}$ ones (third column for Model I and fifth column for Model II), the validity of the LWL limit for $2{^{3}\!S_{1}}\rightarrow\gamma^{3}1P_{1,2}$.
It makes also clear, through the comparison of the calculated $\frac{\vb*{p}_{c}}{M_{c}}$ decay widths with data, that the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation does not give an accurate overall description of these decays. More precisely, except for $1{^{3}\!P_{0}}\rightarrow\gamma\,1{^{3}\!S_{1}}$ from Model I and $2{^{3}\!S_{1}}\rightarrow\gamma 1{^{3}\!P_{2}}$ from Model II, all the calculated $\frac{\vb*{p}_{c}}{M_{c}}$ widths are out of the experimental intervals, in some cases with big differences respect to data. This is in contrast with the situation in bottomonium [@Br3-19] where most of the calculated widths were within or pretty close to the measured intervals.
By realizing that this may have to do at least in part with the poor convergence of the expansion of the complete transition operator in powers of $\frac{\vb*{p}_{c}}{M_{c}}$ (let us recall that the expectation value of $\frac{\left\vert \vb*{p}_{c}\right\vert }{M_{c}}$ in the low lying charmonium states can be as big as $0.5$) we develop in what follows the formalism for the application of the complete operator to the calculation of the decay widths.
Beyond the p/m approximation\[SV\]
==================================
As it was mentioned before, the extraction of the mass dependence in the matrix elements involved in the calculation of the radiative decay widths, allowing for the substitution of the calculated masses by the measured ones, is a crucial step to test the spectroscopic quark model wavefunctions. Moreover, it is a condition *sine qua non* to get an accurate description of radiative decay widths in bottomonium [@Br3-19].
When the complete transition operator is considered, the presence of the momentum dependent operator $\widehat{E}_{\alpha}$ (instead of the constant $M_{\alpha}$ in the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation) complicates this mass extraction. To facilitate it, we first rearrange expression to group all the energy dependent operators to the right so that they act on the initial state. This is convenient on the one hand because we can extract the mass dependence in the terms multiplying the energy dependent operators in exactly the same manner as done in the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation (see below), and on the other hand because the action of $\widehat{E}_{\alpha}$ simplifies when acting on a state of zero total three-momentum. In fact, as the initial state is in its rest frame the action of $\widehat{\vb*{p}}_{\alpha}=\frac{\widehat{\vb*{P}}}{2}-(-1)^{\alpha}\widehat{\vb*{p}}$ becomes equivalent to that of $\widehat{\vb*{p}}$: $$\begin{gathered}
\widehat{E}_{\alpha}\ket{\vb*{P=0,}J,m,nL,s} =\\
\sqrt{M_{c}^{2}+\widehat{\vb*{p}}^{2}}\ket{\vb*{P=0},J,m,nL,s}.
\label{P=0}\end{gathered}$$
To relocate the energy dependent operators we use that $e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}$ represents a three-momentum translation, so that $$e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}\ket{\vb*{p}_{\alpha}} =\ket{\vb*{p}_{\alpha}-\vb*{k}}.
\label{trans1}$$ If we expand for convenience the meson states in terms of the quark and antiquark momentum eigenstates, $\ket{\vb*{p}_1,\vb*{p}_2}$ where we obviate the spin quantum number since it does not play any role in this argument, then gives account of the transition from a quark (antiquark) state with three-momentum $\vb*{p}_{\alpha}$ to a quark (antiquark) state with three-momentum $(\vb*{p}_{\alpha}-\vb*{k})$ through the emission of a photon with three-momentum $\vb*{k}$. Then energy conservation tells us that $$\widehat{E}_{\alpha}(e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}} \ket{\vb*{p}_{\alpha}}) = \widehat{E}_{\alpha} \ket{\vb*{p}_{\alpha}-\vb*{k}} =(E_{\alpha}-k_{0}) \ket{\vb*{p}_{\alpha}-\vb*{k}},
\label{trans2}$$ where $E_{\alpha}$ and $k_{0}$ are numbers, not operators, so that we can write $$\begin{split}
\widehat{E}_{\alpha}(e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}\ket{\vb*{p}_{\alpha}}) &= ( E_{\alpha}-k_{0}) \pqty*{e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}} \ket{\vb*{p}_{\alpha}}} \\
&= e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}( E_{\alpha}-k_{0}) \ket{\vb*{p}_{\alpha}} \\
&=e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}( \widehat{E}_{\alpha}-k_{0}) \ket{ \vb*{p}_{\alpha}}.
\end{split}
\label{trans3}$$ Since this equality holds for the complete set of momentum eigenstates $\Bqty{\ket{\vb*{p}_\alpha}}$ we can rewrite it as an identity between operators: $$\widehat{E}_{\alpha}e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}=e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}( \widehat{E}_{\alpha}-k_{0}).
\label{trans4}$$ Then, using and the well known commutator $$\comm{\widehat{\vb*{p}}_{\alpha}}{e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}} =-\vb*{k}e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}
\label{conmpk}$$ we rearrange as
$$\begin{gathered}
\mel{\vb*{k},\lambda }{ \mathcal{H}_\text{1st}^{int}}{0} =-\frac{1}{\sqrt{2k_{0}}}\sum_{\alpha=1,2}\frac{e_{\alpha}}{2M_{\alpha}}\{e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}2\widehat{\vb*{p}}_{\alpha}\;\mathcal{P}_{+}(\widehat{E}_{\alpha})-e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}\vb*{k}\;\mathcal{K}(\widehat{E}_{\alpha})+\\
-i\vb*{\sigma}_{\alpha}\times\lbrack e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}2\widehat{\vb*{p}}_{\alpha}\;\mathcal{P}_{-}(\widehat{E}_{\alpha})+e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}\vb*{k}\;\mathcal{K}(\widehat{E}_{\alpha})]\}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}
\label{2.4}\end{gathered}$$
where $\mathcal{P}_{\pm}(\widehat{E}_{\alpha})$ and $\mathcal{K}(\widehat{E}_{\alpha})$ stand for the energy dependent scalar operators $$\mathcal{P}_{\pm}(\widehat{E}_{\alpha}) \equiv \pqty{\frac{M_{\alpha}}{M_{\alpha}+\widehat{E}_{\alpha}}\pm\frac{M_{\alpha}}{M_{\alpha}+\widehat{E}_{\alpha}-k_{0}}} \sqrt{\frac{( M_{\alpha}+\widehat{E}_{\alpha})( M_{\alpha}+\widehat{E}_{\alpha}-k_{0})}{4\widehat{E}_{\alpha}( \widehat{E}_{\alpha}-k_{0})}},
\label{pen}$$ $$\mathcal{K}(\widehat{E}_{\alpha}) \equiv\pqty{\frac{2M_{\alpha}}{M_{\alpha}+\widehat{E}_{\alpha}-k_{0}}} \sqrt{\frac{( M_{\alpha}+\widehat{E}_{\alpha})(M_{\alpha}+\widehat{E}_{\alpha}-k_{0}) }{4\widehat{E}_{\alpha}( \widehat{E}_{\alpha}-k_{0}) }}.
\label{ken}$$
Equivalently, using we can also write $$\begin{gathered}
\mel{\vb*{k},\lambda}{\mathcal{H}^{int}_\text{1st}}{0} =-\frac{1}{\sqrt{2k_{0}}}\sum_{\alpha=1,2}\frac{e_{\alpha}}{2M_{\alpha}}\{2\widehat{\vb*{p}}_{\alpha}e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}\mathcal{P}_{+}(\widehat{E}_{\alpha})+e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}\vb*{k}\;\mathcal{K}'(\widehat{E}_{\alpha})+\\
-i\vb*{\sigma}_{\alpha}\times\lbrack2\widehat{\vb*{p}}_{\alpha}e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}\mathcal{P}_{-}(\widehat{E}_{\alpha})+e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_{\alpha}}\vb*{k}\;\mathcal{K}'(\widehat{E}_{\alpha})]\}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*},
\label{2.5}\end{gathered}$$
with $$\mathcal{K}'(\widehat{E}_{\alpha}) \equiv\pqty{\frac{2M_{\alpha}}{M_{\alpha}+\widehat{E}_{\alpha}}} \sqrt{\frac{( M_{\alpha}+\widehat{E}_{\alpha})(M_{\alpha}+\widehat{E}_{\alpha}-k_{0}) }{4\widehat{E}_{\alpha}( \widehat{E}_{\alpha}-k_{0}) }}.
\label{kprimen}$$ For ${{^{3}\!S_1} \leftrightarrow {^{3}\!P_J}}$ transitions, the use of makes easier the calculation when the initial state is an $S$-wave ${(L_{I}=0)}$, whereas is more convenient when the initial state is a $P$-wave $(L_{I}=1)$, as it was the case in the $\frac{\vb*{p}}{M}$ approximation [@Br3-19].
From or the transition amplitude $\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda}$ can be straightforwardly derived following the step by step procedure explained in [@Br3-19]. Thus, using $$\begin{aligned}
\ket{I} &= \ket{\vb*{P}_{I},J_{I},m_{I},n_{I}L_{I},s_{I}},
\label{IN} \\
\ket{F} &= \ket{\vb*{P}_{F},J_{F},m_{F},n_{F}L_{F},s_{F}},
\label{FIN}\end{aligned}$$ introducing center of mass $$\widehat{\vb*{R}}=\frac{\widehat{\vb*{r}}_{1}+\widehat{\vb*{r}}_{2}}{2} \qquad \widehat{\vb*{P}}=\widehat{\vb*{p}}_{1}+\widehat{\vb*{p}}_{2}
\label{cm}$$ and relative $$\widehat{\vb*{r}}=\widehat{\vb*{r}}_{1}-\widehat{\vb*{r}}_{2} \qquad \widehat{\vb*{p}}=\frac{\widehat{\vb*{p}}_{1}-\widehat
{\vb*{p}}_{2}}{2}
\label{rel}$$ operators, integrating over the center of mass spatial degrees of freedom, taking into account that ${(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}\vdot\vb*{k}=0}$ and that in the rest frame of the decaying meson one has $\vb*{P}_{I}=\vb*{0}$, $\vb*{P}_{F}=-\vb*{k}$, the transition amplitude becomes $$\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda}=\sqrt{2M_{I}}\sqrt{2E_{F}}\sum_{\alpha=1,2}\frac{e_{\alpha}}{2M_{\alpha}}\mel{\Psi_{F}}{\widetilde{\mathcal{O}}_{\alpha}}{\Psi_{I}}
\label{tmec}$$ where the matrix element $\mel{\Psi_{F}}{\widetilde{\mathcal{O}}_{\alpha}}{\Psi_{I}} \equiv \langle \widetilde{\mathcal{O}}_{\alpha}\rangle _{FI}$ can be conveniently expressed as a sum of electric and magnetic contributions $$\begin{split}
\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI} &=\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{el}+\langle\widetilde{\mathcal{O}}_{\alpha}\rangle _{FI}^{( mag)_{\vb*{\sigma}\cp\vb*{k}}}+\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{( mag) _{\vb*{\sigma}\cp\widehat{\vb*{p}}}}\\
&=\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{el}+\langle\widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle _{FI}^{(mag) _{\vb*{\sigma}\cp\vb*{k}}}+\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{( mag) _{\vb*{\sigma}\cp\widehat{\vb*{p}}}},
\end{split}
\label{reesp}$$ the first decomposition is technically convenient when the initial state is an $S$-wave, the second one when the initial state is a $P$-wave.
More explicitly, in the first decomposition the electric part is given by $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{el}=\\
\mel{\Psi_{F}}{e^{i(-1)^{\alpha}(\frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2})}(-1)^{\alpha}2\widehat{\vb*{p}}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}},
\label{Oel}\end{gathered}$$ the first magnetic term by $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{(mag)_{\vb*{\sigma}\cp\vb*{k}}}=\\
\mel{\Psi_{F}}{e^{i(-1) ^{\alpha}( \frac{\vb*{k}\vdot\widehat{\vb*{x}}}{2}) }i\vb*{\sigma}_{\alpha}\times\vb*{k}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}\mathcal{K}(\widehat{E}_{\alpha})}{\Psi_{I}}.
\label{Omag}\end{gathered}$$ and the second magnetic term by $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{(mag)_{\vb*{\sigma}\cp\widehat{\vb*{p}}}}=\\
-\mel{\Psi_{F}}{e^{i(-1)^{\alpha}(\frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2})}i\vb*{\sigma}_{\alpha}\times(-1)^{\alpha}2\widehat{\vb*{p}}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}\mathcal{P}_{-}(\widehat{E}_{\alpha})}{\Psi_{I}}
\label{Omagp}\end{gathered}$$ As can be easily checked these expressions reduce to the corresponding ones obtained in the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation by taking $\widehat{E}_{\alpha}=M_{\alpha}$ and $k_{0}\ll M_{\alpha}$.
The second decomposition reads $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{el} = \\
\mel{\Psi_{F}}{(-1)^{\alpha}2\widehat{\vb*{p}} \vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}e^{i(-1)^{\alpha}(\frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2}) }\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}},
\label{OpLi1}\end{gathered}$$ $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{( mag)_{\vb*{\sigma}\cp\vb*{k}}} = \\
\mel{\Psi_{F}}{i\vb*{\sigma}_{\alpha}\times\vb*{k}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}e^{i(-1)^{\alpha}(\frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2})}\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}}.
\label{OpLi1mag}\end{gathered}$$ $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{( mag)_{\vb*{\sigma}\cp\widehat{\vb*{p}}}} = \\
-\mel{\Psi_{F}}{i\vb*{\sigma}_{\alpha}\times(-1)^{\alpha}2\widehat{\vb*{p}}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}e^{i(-1)^{\alpha}(\frac{\vb*{k} \vdot\widehat{\vb*{r}}}{2})}\mathcal{P}_{-}(\widehat{E}_{\alpha})}{\Psi_{I}},
\label{Omagprimep}\end{gathered}$$
In order to extract the mass dependence from the $\widehat{\vb*{p}}$ operator we use and introduce two Parseval identities in terms of eigenstates of the Cornell Hamiltonian $$I = \sum_{A}\ketbra{\Psi_{A}}{\Psi_{A}}
\label{Parsid}$$ where $A$ is a shorthand notation for all the quantum numbers labeling the eigenstates ($J_A,m_A,n_A, L_A,s_A$), so that the above expressions become $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{el} = \\
-iM_{c}\sum_{A,B}\mel{\Psi_{F}}{e^{i(-1)^{\alpha}(\frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2})}}{\Psi_{A}}(M_{B}-M_{A}) \\
\mel{\Psi_{A}}{(-1) ^{\alpha}\widehat{\vb*{r}}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}}{\Psi_{B}} \mel{\Psi_{B}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}},
\label{OelFI}\end{gathered}$$ $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{(mag)_{\vb*{\sigma}\cp\vb*{k}}}=\\
\sum_{B}\mel{\Psi_{F}}{e^{i(-1) ^{\alpha}(\frac{\vb*{k} \vdot\widehat{\vb*{r}}}{2}) }i\vb*{\sigma}_{\alpha}\times \vb*{k} \vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}}{\Psi_{B}}\\
\mel{\Psi_{B}}{\mathcal{K}(\widehat{E}_{\alpha})}{\Psi_{I}},
\label{OmagkFI}\end{gathered}$$ $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{(mag)_{\vb*{\sigma}\cp\widehat{\vb*{p}}}}=\\
iM_{c}\sum_{A,B}\mel{\Psi_{F}}{e^{i(-1) ^{\alpha}(\frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2}) }}{\Psi_{A}} (M_{B}-M_{A}) \\
\mel{\Psi_{A}}{(-1)^{\alpha}i\vb*{\sigma}_{\alpha}\times\widehat{\vb*{r}}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}}{\Psi_{B}} \mel{\Psi_{B}}{\mathcal{P}_{-}(\widehat{E}_{\alpha})}{\Psi_{I}},
\label{OmagpFI}\end{gathered}$$ and $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{el} = \\
-iM_{c}\sum_{A,B}\mel{\Psi_{F}}{(-1) ^{\alpha}\widehat{\vb*{r}}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}}{\Psi_{A}}(M_{A}-M_{F}) \\
\mel{\Psi_{A}}{e^{i(-1)^{\alpha}(\frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2})}}{\Psi_{B}} \mel{\Psi_{B}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}},
\label{OprimelFI}\end{gathered}$$ $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{(mag)_{\vb*{\sigma}\cp\vb*{k}}}=\\
\sum_{B}\mel{\Psi_{F}}{e^{i(-1) ^{\alpha}(\frac{\vb*{k} \vdot\widehat{\vb*{r}}}{2}) }i\vb*{\sigma}_{\alpha}\times \vb*{k} \vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}}{\Psi_{B}}\\
\mel{\Psi_{B}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}}.
\label{OprimemagkFI}\end{gathered}$$ $$\begin{gathered}
\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{(mag)_{\vb*{\sigma}\cp\widehat{\vb*{p}}}}=\\
iM_{c}\sum_{A,B}\mel{\Psi_{F}}{(-1)^{\alpha}i\vb*{\sigma}_{\alpha}\times\widehat{\vb*{r}}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}}{\Psi_{A}} (M_{A}-M_{F}) \\
\mel{\Psi_{A}}{e^{i(-1) ^{\alpha}(\frac{\vb*{k}\vdot\widehat{\vb*{r}}}{2}) }}{\Psi_{B}} \mel{\Psi_{B}}{\mathcal{P}_{-}(\widehat{E}_{\alpha})}{\Psi_{I}},
\label{OprimemagpFI}\end{gathered}$$
To proceed further we should extract the state mass dependence, if present, from the matrix elements involving the energy dependent operators. Although in principle one could extract it through the expansion of the energy operators in powers of $\frac{\widehat{\vb*{p}}^{2}}{M_{c}^{2}}$ and the introduction of as many additional complete sets of intermediate states as needed, this is not practicable. Instead we can infer the possible mass dependence by realizing that
i) $k_{0}$ is quite smaller than $M_{c}$, $\frac{k_{0}}{M_{c}}\leq0.3$ for Model I and $\frac{k_{0}}{M_{c}}\leq0.23$ for Model II, as can be verified from Table \[tabLWL\]. Then, up to $( \frac{k_{0}}{M_{c}}) ^{0}$ order the energy dependent operators reduce to (let us recall that as $\widehat{E}_{\alpha}$ is acting on the initial state it can be substituted by $\widehat{E}=\sqrt{M_{c}^{2}+\widehat{\vb*{p}}^{2}}$) $$\mathcal{P}_{+}(\widehat{E})\simeq\frac{M_{c}}{\widehat{E}}\simeq\mathcal{K}(\widehat{E})
\label{ME}$$ $$\mathcal{P}_{-}(\widehat{E})\simeq {0} \label{P-}$$ where $$\frac{M_{c}}{\widehat{E}}=\frac{1}{\sqrt{1+\frac{\widehat{\vb*{p}}^{2}}{M_{c}^{2}}}}=1-\frac{\widehat{\vb*{p}}^{2}}{2M_{c}^{2}}+\frac{3}{2} \pqty{\frac{\widehat{\vb*{p}}^{2}}{2M_{c}^{2}}}^{2}+\dots
\label{M/E}$$ and $$\frac{\widehat{\vb*{p}}^{2}}{2M_{c}^{2}}=\frac{H-V( r)}{2M_{c}} .
\label{HV}$$
ii) as the energy dependent operators are scalars, the only contributions to the matrix elements, for example $\mel{\Psi_{B}}{ \mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}}$, come from $\ket{\Psi_{B}}$ being equal to $\ket{\Psi_{I}}$ or to a radial excitation of $\ket{\Psi_{I}}$. The dominant contribution is $\mel{\Psi_{I}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}}$ involving the matrix element $\mel{\Psi_{I}}{H-V( r)}{\Psi_{I}} =(M_{I}-2M_{c}) -\mel{\Psi_{I}}{V( r)}{\Psi_{I}}$.
iii) according to our assumption that the difference between the calculated mass and the experimental one is due to potential corrections evaluated at first order in perturbation theory it is obvious that $(M_{I})_\text{calculated}-\mel{\Psi_{I}}{V( r)}{\Psi_{I}} =( M_{I}) _{\text{Expt}}-\mel{\Psi_{I}}{V_\text{corrected}( r)}{\Psi_{I}}$ for any of our quark models. Hence, we can appropriately calculate $\mel{\Psi_{I}}{H-V(r)}{\Psi_{I}}$ with Model I or II without the need of extracting the mass dependence to implement the measured values.
Therefore, as the consideration of higher $\frac{k_{0}}{M_{c}}$ orders do not introduce any additional dependence on the mass of the states, we can calculate the matrix elements of the energy dependent operators from Model I or II. One should keep in mind though that the presence of the energy dependent operators, depending on $\frac{\widehat{\vb*{p}}^{2}}{2M_{c}^{2}}=\frac{H-V( r)}{2M_{c}}$, introduces an additional quark mass dependence in the amplitude with respect to the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation. This makes the calculated widths to be more model dependent so that their comparison to data is not only testing the quark model wavefunctions but also the quark mass parameter. This means that this parameter looses part of its effectiveness becoming a more phenomenological one. In fact, we show next that the analysis of radiative transitions may severely constraint its range of values in correlation with the phenomenological string tension.
For the sake of completeness we write also the former matrix elements in the LWL limit: $$\begin{gathered}
( \langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{el}) _{LWL}= (\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{el})_{LWL} =\\
-iM_{c}\sum_{B}\mel{\Psi_{F}}{( M_{B}-M_{F}) (-1) ^{\alpha}\widehat{\vb*{r}}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}}{\Psi_{B}} \\
\mel{\Psi_{B}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}};
\label{finlwlel}\end{gathered}$$ $$\begin{gathered}
( \langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{( mag) _{\vb*{\sigma}\cp\vb*{k}}}) _{LWL}=(\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{(mag)_{\vb*{\sigma}\cp\vb*{k}}})_{LWL}=\\
\sum_{B}\mel{\Psi_{F}}{ ( i\vb*{\sigma}_{\alpha}\times\vb*{k})\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}}{\Psi_{B}}\\
\mel{\Psi_{B}}{ \mathcal{K}(\widehat{E}_{\alpha})}{ \Psi_{I}}.
\label{finlwlmagk}\end{gathered}$$ $$\begin{gathered}
(\langle \widetilde{\mathcal{O}}_{\alpha}\rangle_{FI}^{( mag) _{\vb*{\sigma}\cp\widehat{\vb*{p}}}}) _{LWL}= (\langle \widetilde{\mathcal{O}}_{\alpha}^{\prime}\rangle_{FI}^{(mag)_{\vb*{\sigma}\cp\widehat{\vb*{p}}}})_{LWL} =\\
iM_{c}\sum_{B}\mel{\Psi_{F}}{i\vb*{\sigma}_{\alpha}\times( M_{B}-M_{F}) (-1) ^{\alpha}\widehat{\vb*{r}}\vdot( \vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}}{ \Psi_{B}} \\
\mel{\Psi_{B}}{\mathcal{P}_{-}(\widehat{E}_{\alpha})}{ \Psi_{I}};
\label{finlwlmagp}\end{gathered}$$
Transitions $^{3}\!S_{1}\rightarrow\gamma\,{^{3}\!P_{J}}$
---------------------------------------------------------
In order to evaluate the $^{3}\!S_{1}\rightarrow\gamma\,{^{3}\!P_{J}}$ amplitude we use $\widehat{\vb*{r}}\vdot(\vb*{\epsilon}_{\vb*{k}}^{\lambda}) ^{*}=\sqrt{\frac{4\pi}{3}}r( Y_{1}^{\lambda}(\hat{r}) ) ^{*}$, the well known expansion $$e^{i( \frac{\vb*{k}\vdot\vb*{r}}{2}) }=\sum_{l=0}^{\infty} i^{l}\sqrt{4\pi}\sqrt{2l+1}j_{l}\pqty{\frac{kr}{2}} Y_{l}^{0}(\hat{r}) \label{eikr}$$ where $k$ is in the $Z$ direction, and the fact that the energy dependent operators are scalars, so that for example $$\mel{\Psi_{B}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}} \propto\delta_{J_{B},J_{I}}\delta
_{m_{B},m_{I}}\delta_{L_{B},L_{I}}\delta_{s_{B},s_{I}} .
\label{deltaBI}$$ Thus, from and the electric part of the amplitude reads
$$\begin{gathered}
( \mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{\text{\text{\text{(electric)}}}} }) ^{\,{^{3}\!S_{1}}\rightarrow\gamma\,{^{3}\!P_{J}}}=i\sqrt{2M_{I}}\sqrt{2E_{F}}\delta_{S_{I},S_{F}}e_{c} \sum_{l=0}^{\infty}\sum_{J_{A},L_{A}} \\
(-1) ^{l+L_{F}+L_{I}}(4l+1) ( 2L_{A}+1) C_{2l,\;J_{A},\;J_{F}}^{0,\;m_{F},\;m_{F}}C_{1,\;J_{A},\;J_{I}}^{\lambda,\;m_{F},\;m_{I}}
\pmqty{L_{A} & 2l & L_{F} \\ 0 & 0 & 0} \pmqty{L_{A} & 1 & L_{I} \\ 0 & 0 & 0}
\bmqty{J_{F} & 2l & J_{A} \\ L_{A} & S_{F} & L_{F}} \bmqty{J_{I} & 1 & J_{A} \\ L_{A} & S_{I} & L_{I}}\\
\sum_{n_A,n_B}( M_{B}-M_{A}) \pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{F}L_{F}})^{*}j_{2l}\pqty{\frac{kr}{2}} R_{n_{A}L_{A}}}
\pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{A}L_{A}})^{*}r R_{n_{B}L_{I}}} \mel{J_{I},m_{I},n_{B}L_{I},s_{I}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}} ,
\label{esp}\end{gathered}$$
where $M_{B}$ is the mass of the state $\ket{J_{I},m_{I},n_{B}L_{I},s_{I}}$, $M_{A}$ is the mass of $\ket{J_{A},m_{F},n_{A}L_{A},s_{F}}$, $C$ is the Clebsch-Gordan coefficient $$C_{j_1,\;j_2,\;j_3}^{m_1,m_2,m_3}\equiv(-1)^{j_2-j_1-m_3}\sqrt{2 j_3+1}\pmqty{j_1 & j_2 & j_3 \\ m_1 & m_2 & -m_3}
\label{cg}$$ with $(\;) $ standing for the $3j$ symbol, and $$\begin{bmatrix}
j_1 & j_2 & j_{12} \\
j_3 & j & j_{23}
\end{bmatrix} \equiv
(-1)^{j_1+j_2+j_3+j}\sqrt{(2j_{12}+1)(2j_{23}+1)}
\begin{Bmatrix}
j_1 & j_2 & j_{12} \\
j_3 & j & j_{23}
\end{Bmatrix}
\label{symbol}$$ with $\{\;\} $ standing for the $6j$ symbol.
As for the first magnetic term, we get from and $$\begin{gathered}
( \mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{(magnetic)} _{\vb*{\sigma}\cp\vb*{k}}}) ^{\,{^{3}\!S_{1}}\rightarrow\gamma\,{^{3}\!P_{J}}} = i \sqrt{2E_I}\sqrt{2E_F} e_c \lambda \frac{k}{m} \sum_{l=1}^\infty \sum_{\tilde J = \max(1,\abs{l-2})}^l \\
i^{l} \pqty{(-1)^l + (-1)^{S_F-S_I}} (-1)^{L_F+l-1} \frac{2l - 1}{2} \sqrt{3(2L_F+1)} C_{l-1,1,\tilde J}^{\;0,\;\lambda,\;\lambda}\,C_{\tilde J,\;J_F,\;J_I}^{\lambda,\; m_F,\; m_I} \pmqty{L_F & l-1 & L_I \\ 0 & 0 & 0} \\
\bmqty{S_I & 1 & S_F \\ \frac{1}{2} & \frac{1}{2} & \frac{1}{2}} \bmqty{L_F & l-1 & L_I \\ S_F & 1 & S_I \\ J_F & \tilde J & J_I} \sum_{n_B} \pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{F}L_{F}})^{*}j_{l-1}\pqty{\frac{kr}{2}} R_{n_{B}L_{I}}} \mel{J_{I},m_{I},n_{B}L_{I},s_{I}}{\mathcal{K}(\widehat{E}_{\alpha})}{\Psi_{I}}.
\label{mspk}\end{gathered}$$
Finally, for the second magnetic term we obtain from and $$\begin{gathered}
( \mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{(magnetic)} _{\vb*{\sigma}\cp\widehat{\vb*{p}}}}) ^{\,{^{3}\!S_{1}}\rightarrow
\gamma\,{^{3}\!P_{J}}} = i \sqrt{2E_I}\sqrt{2E_F} e_c \sqrt{\frac{3}{2}} \sum_{l=0}^\infty \sum_{J_A, L_A} i^l (-1)^{L_F+L_I}\pqty{(-1)^l+(-1)^{S_F-S_I}} (2 l +1) (2 L_A + 1) \\
C_{l,\;J_A,\; J_F}^{0,\;m_F,\;m_F} C_{1,\;J_A,\; J_I}^{\lambda,\;m_F,\;m_I} \pmqty{L_A & l & L_F \\ 0 & 0 & 0} \pmqty{L_A & 1 & L_I \\ 0 & 0 & 0} \bmqty{J_F & l & J_A \\ L_A & S_F & L_F} \bmqty{S_I & 1 & S_F \\ \frac{1}{2} & \frac{1}{2} & \frac{1}{2}} \bmqty{L_A & 1 & L_I \\ S_F & 1 & S_I \\ J_A & 1 & J_I} \\
\sum_{n_A,n_B}(M_B - M_A)\pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{F}L_{F}})^{*}j_{l}\pqty{\frac{kr}{2}} R_{n_{A}L_{A}}} \pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{A}L_{A}})^{*}r R_{n_{B}L_{I}}} \mel{J_{I},m_{I},n_{B}L_{I},s_{I}}{\mathcal{P}_{-}(\widehat{E}_{\alpha})}{\Psi_{I}}.
\label{mspp}\end{gathered}$$ For simplicity the matrix elements of the energy dependent operators have been evaluated in momentum space.
Transitions $^{3}\!P_{J}\rightarrow\gamma\,{^{3}\!S_{1}}$
---------------------------------------------------------
We proceed in the same manner for $^{3}\!P_{J}\rightarrow\gamma\,{^{3}\!S_{1}}$ transitions. Thus, from and the electric part of the amplitude reads $$\begin{gathered}
( \mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{\text{\text{(electric)}}} }) ^{{^{3}\!P_{J}}\rightarrow\gamma\,{^{3}\!S_{1}}}= i\sqrt{2M_{I}}\sqrt{2E_{F}}\delta_{S_{I},S_{F}}e_{c} \sum_{l=0}^{\infty}\sum_{J_{A},L_{A}} \\
(-1) ^{l}(4l+1) \sqrt{(2 L_I + 1)(2 L_F + 1)} C_{2l,\;J_{I},\;J_{A}}^{0,\;m_{I},\;m_{I}}C_{1,\;J_{F},\;J_{A}}^{\lambda,\;m_{F},\;m_{I}}
\pmqty{L_{I} & 2l & L_{A} \\ 0 & 0 & 0} \pmqty{L_{F} & 1 & L_{A} \\ 0 & 0 & 0}
\bmqty{J_{A} & 2l & J_{I} \\ L_{I} & S_{I} & L_{A}} \bmqty{J_{A} & 1 & J_{F} \\ L_{F} & S_{F} & L_{A}}\\
\sum_{n_A,n_B}( M_{A}-M_{F}) \pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{F}L_{F}})^{*}r R_{n_{A}L_{A}}} \pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{A}L_{A}})^{*}j_{2l}\pqty{\frac{kr}{2}} R_{n_{B}L_{I}}}
\mel{J_{I},m_{I},n_{B}L_{I},s_{I}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}}.
\label{eps}\end{gathered}$$
where $M_{B}$ is the mass of the state $\ket{J_{I},m_{I},n_{B}L_{I},s_{I}}$ and $M_{A}$ is the mass of the intermediate state $\ket{J_{A},m_{I},n_{A}L_{A},s_{I}}$.
As for the first magnetic term, we get from and $$\begin{gathered}
( \mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{(magnetic)} _{\vb*{\sigma}\cp\vb*{k}}}) ^{{^{3}\!P_{J}}\rightarrow\gamma\,{^{3}\!S_{1}}} = i \sqrt{2E_I}\sqrt{2E_F} e_c \lambda \frac{k}{m} \sum_{l=1}^\infty \sum_{\tilde J = \max(1,\abs{l-2})}^l \\
i^{l} \pqty{(-1)^l + (-1)^{S_F-S_I}} (-1)^{L_F+l-1} \frac{2l - 1}{2} \sqrt{3(2L_F+1)} C_{l-1,1,\tilde J}^{\;0,\;\lambda,\;\lambda}\,C_{\tilde J,\;J_F,\;J_I}^{\lambda,\; m_F,\; m_I} \pmqty{L_F & l-1 & L_I \\ 0 & 0 & 0} \\
\bmqty{S_I & 1 & S_F \\ \frac{1}{2} & \frac{1}{2} & \frac{1}{2}} \bmqty{L_F & l-1 & L_I \\ S_F & 1 & S_I \\ J_F & \tilde J & J_I} \sum_{n_B} \pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{F}L_{F}})^{*}j_{l-1}\pqty{\frac{kr}{2}} R_{n_{B}L_{I}}} \mel{J_{I},m_{I},n_{B}L_{I},s_{I}}{\mathcal{P}_{+}(\widehat{E}_{\alpha})}{\Psi_{I}}.
\label{mpsk}\end{gathered}$$
Finally, for the second magnetic term we obtain from and $$\begin{gathered}
( \mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{(magnetic)} _{\vb*{\sigma}\cp\widehat{\vb*{p}}}}) ^{{^{3}\!P_{J}}\rightarrow
\gamma\,{^{3}\!S_{1}}} = i \sqrt{2E_I}\sqrt{2E_F} e_c \sqrt{\frac{3}{2}} \sum_{l=0}^\infty \sum_{J_A, L_A} i^l \pqty{(-1)^l+(-1)^{S_F-S_I}} (2 l +1) \sqrt{(2L_I+1)(2L_F+1)} \\
C_{l,\;J_I,\; J_A}^{0,\;m_I,\;m_I} C_{1,\;J_F,\; J_A}^{\lambda,\;m_F,\;m_I} \pmqty{L_I & l & L_A \\ 0 & 0 & 0} \pmqty{L_F & 1 & L_A \\ 0 & 0 & 0} \bmqty{J_A & l & J_I \\ L_I & S_I & L_A} \bmqty{S_I & 1 & S_F \\ \frac{1}{2} & \frac{1}{2} & \frac{1}{2}} \bmqty{L_F & 1 & L_A \\ S_F & 1 & S_I \\ J_F & 1 & J_A} \\
\sum_{n_A,n_B}(M_A - M_F)\pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{F}L_{F}})^{*}r R_{n_{A}L_{A}}}\pqty{\int\limits_{0}^{\infty}\dd{r}r^{2}( R_{n_{A}L_{A}})^{*}j_{l}\pqty{\frac{kr}{2}} R_{n_{B}L_{I}}} \mel{J_{I},m_{I},n_{B}L_{I},s_{I}}{\mathcal{P}_{-}(\widehat{E}_{\alpha})}{\Psi_{I}}.
\label{mpsp}\end{gathered}$$
For simplicity the matrix elements of the energy dependent operators have been evaluated in momentum space.
Results and conclusions\[SVI\]
==============================
From the electric and magnetic amplitudes detailed in the preceding section, the total amplitude $$\begin{gathered}
\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda}=\\
\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{(electric)} }+\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{(magnetic)}_{\vb*{\sigma}\cp\vb*{k}}}+\mathcal{M}_{J_{F},m_{F},J_{I},m_{I}}^{\lambda\;\text{(magnetic)} _{\vb*{\sigma}\cp\widehat{\vb*{p}}}}
\label{tme}\end{gathered}$$ and the width for the $I\rightarrow\gamma F$ decay, given by , can be straightforwardly evaluated.
Notice that the intermediate $B$ states are $( n_{B}) \,{^{3}\!S_{1}}$ for $^{3}\!S_{1}\rightarrow\gamma\,{^{3}\!P_{J}}$ and $(n_{B}){^{3}\!P_{J}}$ for $^{3}\!P_{J}\rightarrow\gamma\,{^{3}\!S_{1}}$ transitions. The intermediate $A$ states are $( n_{A}) ^{3}P_{J_{A}}$ in both cases.
We have checked numerically that for the calculated $^{3}\!S_{1}\rightarrow\gamma\,{^{3}\!P_{J}}$ and ${^{3}\!P_{J}}\rightarrow\gamma\,{^{3}\!S_{1}}$ widths it is enough to take $n_{B}\leq2$ and $n_{A}\leq2$ to assure convergence at the level of $1\%$. As we did in the case of the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation we have used the experimental masses for $2{^{3}\!P_{0}}$ (corresponding to $\chi_{c0}( 3860)$) and $2{^{3}\!P_{2}}$ (corresponding to $\chi_{c2}( 3930)$) and the Cornell calculated mass from Model I for $2{^{3}\!P_{1}}$. The results obtained are shown in Table \[tabcomplet\].
----------------------------------------------------- ------------------------------------------------------------------------- ------------------------------ --------------------------------------------------------------------------
Radiative Decay $( \Gamma_\text{complete}^{( {\text{Theor}}-{\text{Expt}}) }) _{I}$ $\Gamma_{\text{Expt}}^{PDG}$ $( \Gamma_\text{complete}^{( {\text{Theor}}-{\text{Expt}}) }) _{II}$
KeV KeV KeV
$\psi( 2S) \rightarrow\gamma\chi_{c_{0}}( 1p) $ $54$ $28.8\pm1.4$ $43$
$\psi( 2S) \rightarrow\gamma\chi_{c_{1}}( 1p) $ $35$ $28.7\pm1.5$ $30$
$\psi( 2S) \rightarrow\gamma\chi_{c_{2}}( 1p) $ $20$ $28.0\pm1.3$ $18$
$\chi_{c_{0}}( 1p) \rightarrow\gamma J/\psi$ $187$ $151\pm14$ $130$
$\chi_{c_{1}}( 1p) \rightarrow\gamma J/\psi$ $415$ $288\pm22$ $284$
$\chi_{c_{2}}( 1p) \rightarrow\gamma J/\psi$ $566$ $374\pm27$ $385$
----------------------------------------------------- ------------------------------------------------------------------------- ------------------------------ --------------------------------------------------------------------------
A first general feature of these results is that the values of the widths from Model I are systematically bigger than from Model II. As this was also the case in the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation, see Table \[tabp/m\], it can be mostly attributed to the different wavefunctions from both models (let us recall that in the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation the quark mass dependence of the calculated width is reduced with respect to the complete calculation).
A second general feature of the results in Table \[tabcomplet\] is that the calculated widths from Model II are much closer to data than the ones from Model I (the only exception being $2{^{3}\!S_{1}}\rightarrow\gamma\,1{^{3}\!P_{2}}$, where they are almost equal). More precisely, half (three) of the calculated widths from Model II are within the experimental intervals; another one, the $1{^{3}\!P_{0}}\rightarrow\gamma\,1{^{3}\!S_{1}}$ calculated width is very close to the experimental value differing from it a $6\%,$ whereas the remaining two calculated widths, $2{^{3}\!S_{1}}\rightarrow\gamma\,1{^{3}\!P_{0,2}}$, differ from data about a $40\%$. In contrast, all the calculated widths from Model I are out of the experimental intervals, in half of the cases with big differences (bigger than $40\%$) with respect to data.
Centering on Model II, the calculated widths show better agreement with data for ${2{^{3}\!S_{1}}\rightarrow \gamma\,1{^{3}\!P_{1}}}$ than for ${2{^{3}\!S_{1}}\rightarrow\gamma\,1{^{3}\!P_{0,2}}}$, and the same tendency, although attenuated, is observed for ${1{^{3}\!P_{1}}\rightarrow\gamma\,1{^{3}\!S_{1}}}$ as compared to $1{^{3}\!P_{0,2}}\rightarrow\gamma\,1{^{3}\!S_{1}}$. This points out to a good wavefunction for $1{^{3}\!P_{1}}$ and to some deficiency in the $1{^{3}\!P_{0,2}}$ wavefunctions, what can be correlated to the mass description of the charmonium states, see Table \[tabIII\]: quite good for $1{^{3}\!S_{1}}$, $2{^{3}\!S_{1}}$ and $1{^{3}\!P_{1}}$ and deficient for $1{^{3}\!P_{0,2}}.$ Then, it is quite possible that nonperturbative or second order perturbative corrections to the $1{^{3}\!P_{0,2}}$ wavefunctions play some role. If we assume that this is the case, so that that the wavefunction corrections are responsible for the observed discrepancies with data, then we may tentatively conclude that an accurate description of radiative decays in charmonium is feasible.
We may also conclude that the analysis of radiative decay processes serves as a stringent test of the spectroscopic quark model. In this regard, let us recall that the difference between Model I and II comes essentially from the different values of the parameters, $\sigma$ and $M_{c}$. Actually, $\sigma$ and $M_{c}$ are correlated in the sense that an increase in $\sigma$ has to be compensated with an increase in $M_{c}$ to maintain the spectral mass fit, see Table \[tabIII\]. The fact that Model II works much better for the description of the studied radiative decays indicates that the value of the effective parameter $\sigma$ has to be quite close to its phenomenological value from the analysis of Regge trajectories in light mesons and, as a consequence, that the quark mass effective parameter, at least for the description of the low lying charmonium states, is constrained to be around $1.84$ GeV.
Certainly, the prediction of the decay widths for radiative transitions not yet measured could provide, if confirmed by future experiments, strong support to our conclusions. In practice, this comparison program presents some difficulties. On the one hand we may expect that the $3{^{3}\!S_{1}}\rightarrow\gamma\,1{^{3}\!P_{0,2}}$ widths suffer at least from the same uncertainty as the $2{^{3}\!S_{1}}\rightarrow\gamma\,1{^{3}\!P_{0,2}}$. On the other hand the experimental resonance $\psi(4040)$ contains presumably a significant mixing of $3{^{3}\!S_{1}}$ and $2^{3}D_{1}$ states and the calculation of $2^{3}D_{1}\rightarrow\gamma\,1{^{3}\!P_{0,1,2}}$ beyond the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation is quite uncertain due to the current dearth of $^{3}D_{J}$ state mass data to be implemented. The same arguments apply to $3{^{3}\!S_{1}}\rightarrow\gamma\,2{^{3}\!P_{0,2}}.$ On the other hand the observed decays $X( 3872) \rightarrow \gamma\,( 1,2) {^{3}\!S_{1}}$ can not be described from Model II since the $X(3872)$ description requires the incorporation of meson-meson threshold effects. Furthermore, the same kind of effects may be present in $\chi_{c0}(3860)$ that we have assumed to be the pure $2{^{3}\!P_{0}}$ Cornell state in spite of being above its first $I(J^{PC})=0(0^{++}) $ meson-meson threshold, and in many other high lying charmonium sates.
This reduces the efficiency of our comparison program to the prediction of the $2{^{3}\!P_{2}}\rightarrow( 1,2) \,{^{3}\!S_{1}}$, the $2{^{3}\!P_{0}}\rightarrow( 1,2) \,{^{3}\!S_{1}}$ and the $3{^{3}\!S_{1}}\rightarrow 1{^{3}\!P_{1}}$ decay widths from Model II.
The results are shown in Table \[tabprediction\]. As said above, the experimental confirmation of these results, at least for $\chi_{c2}(3930) \rightarrow \gamma\,(J/\psi,\psi(2S))$, would be a good test of our decay model. We encourage an experimental effort along this line.
----------------------------------------------- --------------------------------------------------------------------------
Radiative Decay $( \Gamma_\text{complete}^{( {\text{Theor}}-{\text{Expt}}) }) _{II}$
KeV
$\chi_{c2}( 3930) \rightarrow\gamma J/\psi$ $54$
$\chi_{c2}( 3930) \rightarrow\psi( 2S) $ $128$
$\chi_{c0}( 3860) \rightarrow\gamma J/\psi$ $36$
$\chi_{c0}( 3860) \rightarrow\psi( 2S) $ $45$
$3{^{3}\!S_{1}}\rightarrow\chi_{c_{1}}(1p) $ $1.6$
----------------------------------------------- --------------------------------------------------------------------------
: Predicted widths from Model II. Same notation as in Table \[tabcomplet\]. $\Gamma_\text{complete}^{( {\text{Theor}}-{\text{Expt}}) }$: calculated width implemented with the experimental masses and photon energy. For the calculation of the $3{^{3}\!S_{1}}\rightarrow\chi_{c_{1}}(1p)$ width we have used in Eqs. , and intermediate states with $n_A \le 2$ and $n_B \le 3$.[]{data-label="tabprediction"}
Summary\[SVII\]
===============
In this article we have analyzed ${{^{3}\!S_1} \leftrightarrow {^{3}\!P_J}}$ electromagnetic transitions in charmonium. From two constituent quark models reasonably fitting the masses of the low lying charmonium states and from an elementary emission model for the description of the radiative processes we have shown that
i) neither the standard $\frac{\vb*{p}_{c}}{M_{c}}$ approximation to the electromagnetic transition operator nor its commonly used long wave length limit should be taken for granted for they can not give accurate account of data. Although this inaccuracy has been only quantitatively proved for two Cornell potential models it may be reasonably assumed to be valid in general since the Cornell potential form contains the more relevant terms for an appropriate description of the low lying charmonium states. As a consequence, the $\frac{\vb*{p}_{c}}{M_{c}}$ approximation or its long wave length limit should also not be used to discriminate between different spectroscopic quark models.
ii) the use of the complete electromagnetic operator may allow for an accurate description of ${{^{3}\!S_1} \leftrightarrow {^{3}\!P_J}}$ radiative transitions between low lying charmonium states from a Cornell potential model, when the string tension parameter is close to the phenomenological value derived from the analysis of Regge trajectories in light mesons. This restricts drastically the possible values of the charm mass parameter to keep an appropriate charmonium mass description. Therefore, electromagnetic decay processes in charmonium can be used as a powerful tool to constrain the value of the charm mass parameter and to discriminate among different quark models.
iii) the complete electromagnetic operator formalism developed for the calculation of the radiative widths can be straightforwardly applied to bottomonium $b\bar{b}$, and trivially extended to treat radiative decays of bottom, charmed mesons $b\bar{c}$ and $c\bar{b}$.
This work has been supported by of Spain (MINECO) and EU Feder Grant No. FPA2016-77177-C2-1-P, by SEV-2014-0398, and by EU Horizon 2020 research and innovation program under Grant No. 824093 (STRONG-2020). R.B. acknowledges a FPI fellowship from the under Grant No. BES-2017-079860.
First quantized transition operator\[firstop\]
==============================================
Let us consider for instance the quark part of the vector component of the current (following the main text, we denote quark components by subscript $1$) given by
$$\begin{gathered}
\pqty{\vb*{j}(\vb*{x})}_1 = e_1 \left\{ -i \bqty{q_1^\dagger(\vb*{x}) \pqty{\frac{\grad{q_1(\vb*{x})}}{M_1 + {E}_1}}-\pqty{\frac{\grad{q_1^\dagger(\vb*{x})}}{M_1 +{E}_1}} q_1(\vb*{x})}+\right.\\
\left.\bqty{\pqty{\frac{\grad{q_1^\dagger(\vb*{x})}}{M_1 + {E}_1}} \cp \vb*{\sigma}\!_{1} \, q_1(\vb*{x})-q_1^\dagger(\vb*{x}) \vb*{\sigma}\!_{1} \cp \pqty{\frac{\grad{q_1(\vb*{x})}}{M_1 + {E}_1}}}\right\},\end{gathered}$$
and a quark Fock state given by $$\left\vert \eta\right\rangle =\frac{1}{( 2\pi) ^{\frac{3}{2}}}\int \dd{\vb{p}_1}\sum_{m_s}\phi_{\eta}^{m_s}( \vb{p}_1) b_{1}^{m_s\dagger}( \vb{p}_1) \ket{0} =\frac{1}{( 2\pi) ^{\frac{3}{2}}}\int\dd{\vb{p}_1}\frac{1}{( 2\pi) ^{\frac{3}{2}}}\int\dd{\vb{r}_1}e^{-i\vb{p}_1 \vdot\vb{r}_1} \sum_{m_s}\psi_{\eta}^{m_s}( \vb{r}_1) b_{1}^{m_s\dagger}( \vb{p}_1)\ket{0}$$ where $\phi( \vb{p}_1) $ and $\psi( \vb{r}_1) $ stand for the wavefunction in momentum and configuration space respectively.
We define the first quantized quark part of the vector component of the current $( \vb*{j}_\text{1st}( \vb*{x}))_1$ by requiring that $$\mel{\epsilon}{\pqty{\vb*{j}( \vb*{x})}_1}{\eta} =\int \dd{\vb{r}_1}\sum_{m_s,m_s^\prime}\bigl(\psi_{\epsilon}^{m_s^\prime}( \vb{r}_1)\bigr)^\ast (\vb*{j}_\text{1st}( \vb*{x}))_1^{m_s^\prime, m_s} (\vb{r}_1) \psi_{\eta}^{m_s}( \vb{r}_1) .$$ It is then easy to check that $$\begin{gathered}
\pqty{\vb*{j}_\text{1st}(\vb*{x})}_1 = e_1 \sqrt{\frac{M_1 + {E}_1}{2 {E}_1}} \left[\delta^{(3)}(\vb*{x}-\vb*{r}_1) \frac{\vb*{p}_1}{M_1 + {E}_1} + \frac{\vb*{p}_1}{M_1 + {E}_1} \delta^{(3)}(\vb*{x}-\vb*{r}_1)+\right.\\
\left.-i \vb*{\sigma}_1 \cp \pqty{\delta^{(3)}(\vb*{x}-\vb*{r}_1) \frac{\vb*{p}_1}{M_1 + {E}_1} - \frac{\vb*{p}_1}{M_1 + {E}_1} \delta^{(3)}(\vb*{x}-\vb*{r}_1)}\right]\sqrt{\frac{M_1 + {E}_1}{2 {E}_1}} .
\label{jvecQ}\end{gathered}$$ By using the radiation gauge so that the time component of the electromagnetic field vanishes $( A^{0}( \vb*{x})=0) $ we get $$\begin{gathered}
\mathcal{H}^{int} = \int \dd{\vb*{x}}j^{\mu}_\text{1st}(\vb*{x})A_{\mu}( \vb*{x}) = - \int \dd{\vb*{x}}\vb*{j}_\text{1st}(\vb*{x})\vdot \vb*{A}( \vb*{x})\\
=-\sum_{\alpha=1,2} e_\alpha \sqrt{\frac{M_\alpha +{E}_\alpha}{2 {E}_\alpha}} \left[\pqty{\vb*{A}(\vb*{r}_\alpha) \vdot \frac{\vb*{p}_\alpha}{M_\alpha + {E}_\alpha} + \frac{\vb*{p}_\alpha}{M_\alpha + {E}_\alpha} \vdot \vb*{A}(\vb*{r}_\alpha)}+\right.\\
\left.-i \pqty{ \vb*{\sigma}_{\alpha} \cp \vb*{A}(\vb*{r}_\alpha) \vdot \frac{\vb*{p}_\alpha}{M_\alpha + {E}_\alpha} - \frac{\vb*{p}_\alpha}{M_\alpha + {E}_\alpha} \vdot \vb*{\sigma}_{\alpha} \cp \vb*{A}(\vb*{r}_\alpha) }\right]\sqrt{\frac{M_\alpha + {E}_\alpha}{2 {E}_\alpha}} .\end{gathered}$$ To obtain the desired form of the transition operator we expand $\vb*{A}(\vb*{x})$, the electromagnetic field at $t=0$, in terms of creation and annihilation operators (see for example [@Pes95 p. 123]): $$\vb*{A}(\vb*{x}) = \int\frac{\dd{\vb{k}}}{(2\pi)^3}\frac{1}{\sqrt{2 \mathrm{k}_0}} \sum_{\lambda=\pm1} \pqty{\vb*{\epsilon}_{\vb{k}}^{\lambda} \, a_{\vb{k}}^\lambda e^{i \vb{k}\vdot\vb*{x}}+( \vb*{\epsilon}_{\vb{k}}^{\lambda})^{*} \, a_{\vb{k}}^{\lambda\dagger} e^{-i \vb{k}\vdot\vb*{x}}}
\label{elmanop}$$ where $a_{\vb{k}}^{\lambda\dagger}$ is the operator that creates a photon with three-momentum $\vb{k}$ and polarization $\lambda$ $$\ket{\vb{k},\lambda} = a_{\vb{k}}^{\lambda\dagger} \ket{0} .$$ Then, taking the matrix element of between the vacuum and the one photon state yields $$\mel{\vb*{k},\lambda}{\vb*{A}( \vb*{x})}{0} =\frac{1}{\sqrt{2k_{0}}}e^{-i\vb*{k}\vdot\vb*{x}}( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*}
\label{kame}$$ where $\vb*{\epsilon}_{\vb*{k}}^{\lambda}$ stands for the photon polarization vector, and the first quantized transition operator to be sandwiched between the meson states reads $$\begin{gathered}
\mel{\vb*{k},\lambda}{\mathcal{H}^{int}_\text{1st}}{0} = -\frac{1}{\sqrt{2 k_0}} \sum_{\alpha=1,2} e_\alpha \sqrt{\frac{M_\alpha + \widehat{E}_\alpha}{2 \widehat{E}_\alpha}} \\
\left[e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_\alpha}\frac{\widehat{\vb*{p}}_\alpha}{M_\alpha+\widehat{E}_\alpha} + \frac{\widehat{\vb*{p}}_\alpha}{M_\alpha+\widehat{E}_\alpha}e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_\alpha} -i \vb*{\sigma}\!_{\alpha} \cp \pqty{e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_\alpha} \frac{\widehat{\vb*{p}}_\alpha}{M_\alpha+\widehat{E}_\alpha} - \frac{\widehat{\vb*{p}}_\alpha}{M_\alpha+\widehat{E}_\alpha} e^{-i\vb*{k}\vdot\widehat{\vb*{r}}_\alpha}}\right] \\
\vdot ( \vb*{\epsilon}_{\vb*{k}}^{\lambda})^{*} \sqrt{\frac{M_\alpha + \widehat{E}_\alpha}{2 \widehat{E}_\alpha}} .\end{gathered}$$
Gauge invariance\[gaugeinv\]
============================
Let us show that the electromagnetic current satisfies the condition $$k_{\mu}j^{\mu}_\text{1st}(\vb*{x})=0,$$ meaning that the current is conserved as required by gauge invariance. For this purpose, we have to verify that the matrix element of $k_{\mu}j^{\mu}_\text{1st}(\vb*{x})$ between any initial and final meson state is zero. Equivalently, expanding the meson states in terms of quark and antiquark momentum eigenstates, we can verify that $$\begin{split}
\mel{\vb*{p}_{f},m'_{s}}{k_{\mu}j^{\mu}_\text{1st}(\vb*{x})}{\vb*{p}_{i},m_s} &= \mel{\vb*{p}_{f},m'_{s}}{k_{0}j^{0}_\text{1st}(\vb*{x})-\vb*{k}\vdot \vb*{j}_\text{1st}(\vb*{x})}{\vb*{p}_{i},m_s} \\
&=k_{0}\mel{\vb*{p}_{f},m'_{s}}{ j^{0}_\text{1st}(\vb*{x})}{\vb*{p}_{i},m_s} -\vb*{k}\vdot\mel{\vb*{p}_{f},m'_{s}}{\vb*{j}_\text{1st}(\vb*{x})}{\vb*{p}_{i},m_s} =0
\end{split}$$ where $\vb*{p}_{i}$ is the three-momentum of the initial quark state going to a final state with a photon and a quark with three-momentum $\vb*{p}_{f}$.
Let us prove it for the quark contribution. The time component of the current is given by $$\begin{gathered}
\pqty{j_\text{1st}^{0}( \vb*{x})}_1 = e_1 \sqrt{\frac{M_1 +{E}_1}{2 {E}_1}} \Biggr[\delta^{(3)}(\vb*{x}-\vb*{r}_1)+ \\
\left(\frac{\vb*{p}_1}{M_1 +{E}_1}\vdot \delta^{(3)}(\vb*{x}-\vb*{r}_1) \frac{\vb*{p}_1}{M_1 + {E}_1}+ i \vb*{\sigma}_1\vdot \frac{\vb*{p}_1}{M_1 + {E}_1} \cp \delta^{(3)}(\vb*{x}-\vb*{r}_1) \frac{\vb*{p}_1}{M_1 +{E}_1}\right)\Biggr]\\
\sqrt{\frac{M_1 + {E}_1}{2 {E}_1}}
\label{j0Q}\end{gathered}$$ as can be easily checked following the procedure detailed in Appendix \[firstop\], whilst the vector component is given by .
To calculate the matrix elements we take into account the action of the energy and momentum operators: $$\begin{aligned}
\vb*{p}_1\ket{\vb*{p}_{i}, m_s} &= \vb*{p}_{i} \ket{\vb*{p}_{i}, m_s},\\
\bra{\vb*{p}_{f}, m'_s}\vb*{p}_1 &= \bra{\vb*{p}_{f}, m'_s}\vb*{p}_{f},\end{aligned}$$ and $$\begin{aligned}
{E}_1 \ket{\vb*{p}_{i}, m_s} &= \sqrt{M_1^2+\vb*{p}_{i}^2} \ket{\vb*{p}_{i},m_s} = E_{i} \ket{\vb*{p}_{i},m_s},\\
\bra{\vb*{p}_{f}, m'_s} {E}_1 &= \bra{\vb*{p}_{f}, m'_s} \sqrt{M_1^2+\vb*{p}_{f}^2} = \bra{\vb*{p}_{f}, m'_s} E_{f}.\end{aligned}$$ The matrix element of the time component reads $$\begin{gathered}
\mel{\vb*{p}_f,m'_s}{\pqty{j_\text{1st}^0(\vb*{x})}_1}{\vb*{p}_i,m_s} = e_1 \Biggr[\delta^{m'_s,m_s}\pqty{1+ \frac{\vb*{p}_i}{M_1 + E_i} \cdot \frac{\vb*{p}_f}{M_1 + E_f}} - i \vb*{\sigma}_1^{m'_s,m_s}\vdot \frac{\vb*{p}_i}{M_1 + E_i} \cp \frac{\vb*{p}_f}{M_1 + E_f}\Biggr] \\
\sqrt{\frac{M_1 + E_f}{2 E_f}} \sqrt{\frac{M_1 +E_i}{2E_i}} e^{i (\vb*{p}_i - \vb*{p}_f)\vdot \vb*{x}}\end{gathered}$$ where we have used $$\mel{\vb*{p}_{f},m'_{s}}{\delta( \vb*{x}-\vb{r}_1)}{\vb*{p}_{i},m_s} = e^{i (\vb*{p}_i - \vb*{p}_f)\vdot \vb*{x}} \delta^{m'_s,m_s}.$$ Note that for momentum eigenstates we have adopted the nonrelativistic normalization $$\braket{\vb*{p}^\prime,m'_s}{\vb*{p},m_s}=(2\pi)^3 \delta^{(3)}(\vb*{p}-\vb*{p}^\prime)\delta^{m'_s,m_s},$$ consistently with [@PDG18 p. 567]. In the same way, for the matrix element of the vector component we obtain $$\begin{gathered}
\mel{\vb*{p}_f,m'_s}{\pqty{\vb*{j}_\text{1st}(\vb*{x})}_1}{\vb*{p}_i,m_s} = e_1 \left[\delta^{m'_s,m_s}\pqty{\frac{\vb*{p}_i}{M_1 + E_i} + \frac{\vb*{p}_f}{M_1 +E_f}} -i \vb*{\sigma}_1^{m'_s,m_s} \cp \pqty{\frac{\vb*{p}_i}{M_1 + E_i} - \frac{\vb*{p}_f}{M_1 + E_f} }\right] \\
\sqrt{\frac{M_1 + E_f}{2 E_f}} \sqrt{\frac{M_1 +E_i}{2E_i}} e^{i (\vb*{p}_i - \vb*{p}_f)\vdot \vb*{x}} .\end{gathered}$$
Now, using $$\vb*{p}_{i,f} \vdot \vb*{p}_{i,f} = \vb*{p}_{i,f}^2 = E_{i,f}^2 - M_1^2$$ it is easy to show that $$(\vb*{p}_{i}-\vb*{p}_{f}) \vdot \pqty{\frac{\vb*{p}_i}{M_1 + E_i} + \frac{\vb*{p}_f}{M_1 +E_f}} = (E_i - E_f) \pqty{1+\frac{\vb*{p}_i}{M_1 + E_i} \vdot \frac{\vb*{p}_f}{M_1 + E_f}}$$ and $$(\vb*{p}_{i}-\vb*{p}_{f}) \vdot \vb*{\sigma}_1^{m'_s,m_s} \cp \pqty{\frac{\vb*{p}_i}{M_1 + E_i} - \frac{\vb*{p}_f}{M_1 + E_f} } = (E_i - E_f) \vb*{\sigma}_1^{m'_s,m_s} \vdot \frac{\vb*{p}_i}{M_1 + E_i} \cp \frac{\vb*{p}_f}{M_1 + E_f} ,$$ so that $$(\vb*{p}_{i}-\vb*{p}_{f}) \vdot \mel{\vb*{p}_f,m'_s}{\pqty{\vb*{j}_\text{1st}(\vb*{x})}_1}{\vb*{p}_i,m_s} = (E_{i}-E_{f}) \mel{\vb*{p}_f,m'_s}{\pqty{j^0_\text{1st}(\vb*{x})}_1}{\vb*{p}_i,m_s}.
\label{easyeq}$$
Then, applying four-momentum conservation in the emission of the photon by the quark $$\begin{aligned}
k_{0}&=E_{i}-E_{f} , \\
\vb*{k} &= \vb*{p}_{i}-\vb*{p}_{f} ,\end{aligned}$$ becomes $$\vb*{k} \vdot \mel{\vb*{p}_f,m'_s}{\pqty{\vb*{j}_\text{1st}(\vb*{x})}_1}{\vb*{p}_i,m_s} = k_0 \mel{\vb*{p}_f,m'_s}{\pqty{j^0_\text{1st}(\vb*{x})}_1}{\vb*{p}_i,m_s} ,$$ showing that the quark contribution to the current is conserved.
It is clear that the substitutions $e_1\rightarrow e_2$ and $M_1\rightarrow M_2$ do not affect this result, so that the the antiquark contribution is conserved as well. Then, the conservation of both the quark and antiquark components separately implies the conservation of the total current defined by and .
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'It is shown that the fluctuations of the jamming coverage upon Random Sequential Adsorption ($\sigma_{\theta_J}$), decay with the lattice size according to the power-law $\sigma_{\theta_J} \propto L^{-1/\nu_{J}}$, with $\nu_{J} = \frac{2}{2D - d_{\rm f}}$, where $D$ is the dimension of the substrate and $d_{\rm f}$ is the fractal dimension of the set of sites belonging to the substrate where the RSA process actually takes place. This result is in excellent agreement with the figure recently reported by Vandewalle [*et al*]{} ([*Eur. Phys. J.*]{} B. [**14**]{}, 407 (2000)), namely $\nu_{J} = 1.0 \pm 0.1$ for the RSA of needles with $D = 2$ and $d_{\rm f} = 2$, that gives $\nu_{J} = 1$. Furthermore, our prediction is in excellent agreement with different previous numerical results. The derived relationships are also confirmed by means of extensive numerical simulations applied to the RSA of dimers on both stochastic and deterministic fractal substrates.'
author:
- |
Ernesto S. Loscar, Rodolfo A. Borzi and Ezequiel V. Albano$^{a}$\
$^a$[*Instituto de Investigaciones Fisicoquímicas Teóricas y Aplicadas*]{}\
[*(INIFTA), UNLP, CONICET, Suc.4, CC16,*]{}\
[*1900 La Plata, Argentina*]{}
title: 'Scaling behavior of jamming fluctuations upon random sequential adsorption.'
---
.5cm .5cm 21truecm 16truecm
The irreversible deposition of particles on a surface involves two characteristic time scales: the time between depositions, and the diffusion time of the particles on the surface. For very strong interaction between particles and the substrate (chemical adsorption), diffusion becomes irrelevant and the Random Sequential Adsorption (RSA) model provides an excellent description of the underlying processes (for a review on RSA models see [@Evans]). Under these conditions the system evolves rapidly toward far-from equilibrium conditions and the dynamics becomes essentially dominated by geometrical exclusion effects between particles. This kind of effects has been observed in numerous experiments [@9].
The RSA of needles (or linear segments) on homogeneous, two-dimensional samples, has very recently attracted considerable interest [@Galam; @Pekalski]. Particular attention has been drawn to the interplay between the jamming coverage and percolation [@Galam; @Pekalski; @frede]. The percolation problem has also attracted considerable attention in the field of statistical physics due to their relevance for the understanding of processes and phenomena in many other areas such as those occurring in disordered media, porous materials, systems of biological and ecological interest, etc. [@hav1; @hav2; @stau]. Therefore, a great progress in the field of the statistical physics of far-from equilibrium processes could be achieved by establishing links between RSA and percolation [@Galam; @Pekalski; @frede].
The percolation transition is related to the probability of occurrence of an infinite connectivity between randomly deposited objects, as a function of the fraction $p$ of the substrate occupied by the objects. Close to the percolation threshold $p_{c}$, the probability $P$ to find a percolating cluster, on a finite sample of side $L$, is given by an error function [@stau] $$P = \frac{1}{\sqrt{2\pi}\sigma} \int_{-\infty}^{p}
exp\Big{[}-\frac{1}{2}
{\Big{(}\frac{p' - p_{c}}{\sigma}\Big{)}}^2
\Big{]}dp' ,
\label{error}$$ where $\sigma$ is the width of the transition region. It is well known that $\sigma$ vanishes in the thermodynamic limit according to [@stau] $$\sigma \propto L^{-\frac{1}{\nu}},
\label{delta}$$ where $\nu$ is the exponent that governs the divergence of the correlation length as $\xi \propto |p - p_{c}|^{-\nu}$.
Very recently it has been suggested that the jamming probability and the fluctuations of the jamming coverage may obey relationships similar to equations (\[error\]) and (\[delta\]) [@Galam], respectively. The aim of this note is to provide a qualitative derivation of equation (\[delta\]) for the case of RSA on both homogeneous and deterministic fractal substrates. The predictions of the obtained equation will be compared with previously published data and further numerical tests will be performed. To accomplish these goals, the RSA of dimers on deterministic and stochastic fractals such as a Sierpinski Carpets (SC) [@hav2] and the diffusion front [@hav1; @hav2], has been studied.
Let us first establish a link between the fluctuations of the number of deposited particles (at the jamming state) on a subsystem of side $L_0$ with those of a system of side $L$, with $L_0 < L$. Considering a fractal subsystem of side $L_0$, that has itself $Q_0$ minimal pattern blocks, and increasing the size of the subsystem $n$ steps $\lambda$ times until reaching the size $L$, such as $L(n) = \lambda^n L_0$, then $Q_0$ will change to $Q(n) = s^{n} Q_0$. Therefore, eliminating $n$, it follows that $$Q={Q_0}\times \left({\frac{L}{L_0}} \right)^{d_{\rm f}}
\label{fract}$$
where $d_{\rm f}=log(s)/log(\lambda)$ is the fractal dimension.
Let $N_0$ be the number of adsorbed particles in the starting subsystem of side $L_{0}$. For the system of side $L$ the number of adsorbed particles $N(L)$ is given by the sum
$$N(L) = \sum_{\i=1}^{s^n} N_i ,
\label{sum1}$$
where $N_i$ are the number of adsorbed particles on each subsystem of side $L_0$ that form the system of side $L$. Let $\sigma_{N(L_0)}$ be the fluctuations, in the starting $L_0$-subsystem, of $N_0$. If the correlation length associated to the random sequential adsorption ($\xi_{Rsa}$) is short compared with $L_0$ ($\xi_{Rsa} << L_0$), the random variables $N_i$ will be statistically independent an so from Eq.(\[sum1\]) it follows
$$\sigma^2_{N(L)}=
\sum_{\i=1}^{s^n} {\sigma_{i}}^2.
\label{sum2}$$
Furthermore, in the thermodynamic limit, the $L_0$-subsystems should have identical statistical properties. Thus, their respective fluctuations will be the same. So, from the fact that $s^{n} = \frac{Q(n)}{Q_0}$ and Eqs.(\[fract\]) and (\[sum2\]) one has
$$\sigma^2_{N(L)}=\frac{\sigma^2_{N_0}}{L_0^{d_{\rm f}}} \times L^{d_{\rm f}},
\label{sum3}$$
then the fluctuation of the density ($\theta$) in the system of size $L$ can be obtained from Eq.(\[sum3\]) dividing by $L^{2D}$, so that
$$\sigma_{\theta} \propto L^{-\frac{1}{\nu_{J}}} ,
\label{g141}$$
where
$$\quad \nu_{J}= {\frac{2}{2D-d_{\rm f}}}.
\label{g14}$$
It should be stressed that Eqs. (\[g141\]) and (\[g14\]) are quite general relationships valid for substrate systems that are both homogeneous and deterministic fractal. Furthermore, the same relationships hold for the case of substrates globally-invariant under translations, such as random fractals, as it has been demonstrated elsewhere [@nosot]. Also, the condition that the correlation length of the RSA process should be smaller than the system size is usually valid for jammed states, where the correlation length is very short. It is also very interesting to notice that, using these relationships it may be possible to evaluate $d_{\rm f}$ performing both RSA numerical simulations and actual experiments. Furthermore, existing numerical simulations performed in $D = 2$ dimensions with $d_{\rm f} = 2$ are in excellent agreement with equations (\[g141\]) and (\[g14\]) (notice that for these conditions it follows straightforwardly from equation (\[g14\]) that $\nu_{J} = 1$ exactly). In fact, for the jamming upon RSA of needles in two dimensions the value $\nu_{J} = 1.0 \pm 0.1$ has been reported [@Galam] and this figure is independent of the aspect ratio of the needles. Furthermore, early numerical results of Nakamura for the RSA of square blocks are also consistent with $\nu_{J} \simeq 1$ [@Naka], while Kondrat [*et al.*]{} [@Pekalski] have reported $\nu_{J} = 1.00 \pm 0.05$ for the RSA of segments on the square lattice. Since the obtained values for the exponent are independent (within error bars) of: i) the length of the segments (for all a = 1,2,....,45) [@Pekalski], ii) the aspect ratio of the needles [@Galam] and iii) the size of the square blocks [@Naka], it has been suggested that $\nu_{J}$ is a good candidate for an universal quantity of the jamming process [@Pekalski]. Within this context, our finding shows that $\nu_J$ depends on the dimensionality of the substrate and the set where the RSA processes actually takes place.
On homogeneous samples the jamming coverage $(\theta$) and its fluctuations ($\sigma_{\theta}$) can straightforwardly be obtained, since one has to deal with a single stochastic process. However, RSA on nonhomogeneous random substrates requires a careful treatment because two correlated stochastic processes are now involved [@nosot]. One can assume that the fluctuations due to the RSA process are given by an average over $M$ independent samples:
$$\sigma_{\theta}=
\sum_{i=1}^{M}
\frac { {\sigma^{i}_{\theta}}}
{M} ,
\label{sigmarsa}$$
where $\sigma^{i}_{\theta}$ are the fluctuations measured using a [**single**]{} substrate sample but taken averages over independent RSA trials. It has been shown [@nosot] that measuring $\sigma_{\theta}$ with the aid of Eq. (\[sigmarsa\]) one captures the physical behavior of the RSA process. In contrast, measuring the fluctuations of the average jamming coverage of different samples the physical behavior reflects the properties of the substrate [@nosot].
In order to perform additional tests to the obtained analytical results, the RSA of dimers on both stochastic and deterministic fractals has been studied numerically.
As example of an stochastic fractal, we have used a diffusion front. In order to generate the diffusion front, we considered the diffusion of particles at random, but with hard-core interactions, on a 2D square lattice of size $L \times L$. There is a source of particles at the first row of the lattice $y = 1, 1 \leq x \leq L$ kept at concentration $p(1,t) \equiv 1$. Also, at row $y = L+1, 1 \leq x \leq L$ there is a well, $p(L+1,t) \equiv 0$. So, there is a concentration gradient along the source-well direction, while along the perpendicular [*x*]{}-direction periodic boundary conditions are imposed. In the steady state the concentration gradient is constant, so one has
$$\nabla p(y)=L^{-1}.
\label{grad}$$
It is well known that the properties of the diffusion front [@SRG; @MAR1] are closely related to those of the incipient percolation cluster [@hav1; @hav2; @stau]. As the concentration $p(y)$ of particles depends on the position, decreasing from the source to the well, one actually has a [**gradient percolation**]{} system. The structure of the diffusion front is identical to the structure of the hull of the incipient percolation cluster [@SRG]. Furthermore, the concentration of particles at the mean front position $y_f$ is the same as the percolation threshold $p_c$, so that $p(y_f)=p_c$ [@SRG]. The diffusion front is conveniently described by its average width $\sigma_f$ and the total number $N_f$ of particles that constitute it. Using heuristic arguments it has been suggested that $$\begin{aligned}
\frac {N_f}{L}\sim \left| \nabla p(y_f) \right|^{-\alpha_N} \qquad \qquad \textrm{where}\
\quad \alpha_{N} = \frac{1}{\nu+1}
\label{alfa} \end{aligned}$$ being $\nu$ the critical exponent of the correlation length in the percolation problem [@hav1; @hav2]; $\nu = 4/3$ in $2D$, which gives $\alpha_N=3/7$. So, from Eqs.(\[grad\]) and (\[alfa\]) one has
$$N_f \sim L^{d_{\rm f}^{DF}} ,
\label{frente}$$
with $d_{\rm f}^{DF} = \alpha_{N} + 1 = 10/7 \approx 1.4286$, and the diffusion front is a stochastic self-similar fractal [@note].
RSA of dimers on diffusion fronts has been simulated using two rules: according to [**Rule I**]{} only adsorption events of dimers taking place on two nearest-neighbor (NN) sites, such us one of then belongs to the diffusion front and the remaining one is outside it, are considered. On the other hand, using [**Rule II**]{} one only allows the adsorption on NN sites of the diffusion front, disregarding adsorption trials on already occupied sites of the front and sites outside the fractal.
RSA of dimers on deterministic fractals (Sierpinski Carpets [@hav1; @hav2]) is also studied. The SC in $D = 2$ dimensions is generated by dividing a full square into $\lambda^D$ smaller squares of the same size. Out of these squares, $k$ of them are chosen and removed. In the next iteration, the procedure is repeated by dividing each of the small squares left into $\lambda^D$ smaller squares removing those $k$ squares that are located at the same positions as in the first iteration. The resultant fractal dimensions are $$d_{\rm f}(s,\lambda) = log(s)/log(\lambda)$$ where $s=\lambda^D - k$. In principle, this procedure has to be repeated again and again, however for the practical implementation in a computer only a finite number of iterations are actually performed [@hav2; @nazareno]. In a square lattice the smaller subdivision is actually a single site and the length is measured in site units. Furthermore there is a minimal pattern of $\lambda^{d_{\rm f}}$ sites. In the present work various generations of SC’s of different size $L$, with periodical boundary conditions, have been employed. In all cases dimers are allow to adsorb only on NN empty sites belonging to the fractal. For SC’s with $\lambda = 3$ and $k = 1,2,3$, as used in the simulations, the fractal dimensions are $ d_{\rm f_{I}}=log(8)/log(3)\approx 1.8928$, $ d_{\rm f_{II}}=log(7)/log(3)\approx 1.7712$ and $ d_{\rm f_{III}}=log(6)/log(3)\approx 1.6309$, respectively.
Figures \[randomfractals\] and \[sierpi\] show log-log plots of $\sigma_{\theta}$ versus $L$ obtained upon RSA of dimers on diffusion fronts and Sierpinski Carpets, respectively. The obtained results, for these kind of fractals, are in excellent agreement with the prediction of Eq.(\[g14\]) as follows from the comparison of evaluated and theoretical exponents listed in Table I. Further support to the theoretical prediction follows from additional results obtained using homogeneous samples, which are also listed in Table I.
=9.0 cm
=9.0 cm
Substrate $1/\nu_{J}$ $d_{\rm f}^*$ $d_{\rm f}$
------------- ------------- --------------- ----------------------------
SC (a) $1.051(4)$ $1.898(8)$ $ln(8)/ln(3) \simeq 1.893$
SC (b) $1.052(4)$ $1.896(8)$ $ln(8)/ln(3) \simeq 1.893$
SC (c) $1.115(2)$ $1.770(4)$ $ln(7)/ln(3) \simeq 1.771$
SC (d) $1.110(7)$ $1.780(15)$ $ln(7)/ln(3) \simeq 1.771$
SC (e) $1.16(2)$ $1.68(4)$ $ln(6)/ln(3) \simeq 1.631$
DF $1.30(2)$ $1.40(4)$ $10/7 \simeq 1.429$
HS2 (D = 2) $1$ - 2
: Examples of the application of Eq.(\[g14\]) to different fractals as listed in the first column: SC $\equiv$ Sierpinski Carpet, DF $\equiv$ Diffusion front, HS2 Homogeneous Substrate in $D = 2$ dimensions. The 2nd column shows the exponents obtained fitting Eq.(\[g14\]) to the simulation results while the 3rd one shows the estimations of $d_{\rm f}$ obtained using $\frac{1}{\nu_{J}}= {\frac{2D-d_{\rm f}}{2}}$. The 4th column is a list of the exact values of $d_{\rm f}$. Notice that for SC the labels a)-e) allows to identify the generating patterns, as shown in figure \[sierpi\].
Summing up, it is shown that the exponent $\nu_{J}$ can be obtained as a function of the dimensionality $D$ of the space and the fractal dimension $d_{\rm f}$ of the subset site where the RSA process actually takes place. Our main result $\nu_{J}= {\frac{2}{2D-d_{\rm f}}}$, provides a solid ground to previous numerical data [@Galam; @Pekalski; @Naka]. Furthermore, in this work, the validity of the proposed relationship is verified by means of extensive numerical simulations, using both homogeneous substrates as well as different fractals.
[**Acknowledgments**]{}: This work was supported by CONICET, UNLP and ANPCyT (Argentina).
[99]{}
J. W. Evans, Rev. Mod. Phys., [**65**]{}, 1281 (1993).
J. J. Ramsden, J. Stat. Phys., **73**, 853 (1993).
N. Vandewalle, S. Galam and M. Kramer, Eur. Phys. J. B, [**14**]{}, 407 (2000).
G. Kondrat and A. Pekalski, Phys. Rev. E, [**63**]{}, 051108 (2001).
F. Rampf and E. V. Albano, Phys. Rev. E., [**66**]{}, 061106 (2002).
, Eds. A. Bunde and S. Havlin. Springer-Verlag. Heildelberg, (1991).
M. S. Nazarro, A. J. Ramirez Pastor, J. L. Riccardo and V. Pereyra, [**30**]{}, 1925 (1997).
, Eds. A. Bunde and S. Havlin. Springer-Verlag. Heildelberg, (1994).
D. Stauffer and A. Aharoni, [*Introduction to the Percolation Theory*]{} (Taylor & Francis, London, 1994), 2nd edn.
M. Nakamura, J. Phys. A, [**19**]{}, 2345 (1986).
E. S. Loscar, A. Borzi and E. V. Albano, to be published.
B. Sapoval, M. Rosso and J. F. Gouyet. J. Physique. Lett. [**46**]{}, L49 (1985); M. Rosso, J. F. Gouyet and B. Sapoval. Phys. Rev. B. [**32**]{}, 6053 (1985); M. Rosso, J. F. Gouyet and B. Sapoval. Phys. Rev. Lett. [**57**]{}, 3195 (1986).
A. Memsouk, Y. Boughaleb, R. Nassif and H. Ennamiri. Eur. Phys. J. B. [**17**]{}, 137 (2000); A. Hader, A. Memsouk and Y. Boughaleb. Eur. Phys. J. B. [**17**]{}, 137 (2000).
Notice that scaling $N_{f}$ as a function of the diffusion width, instead of the lattice size $L$ as in Eq.(\[frente\]), one can define a different fractal dimension given by $d_{f}^{*DF} = \nu +1 / \nu = 7/4$.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'After introducing the Szekeres and Lemaître–Tolman cosmological models, the real-time cosmology program is briefly mentioned. Then, a few widespread misconceptions about the cosmological models are pointed out and corrected. Investigation of null geodesic equations in the Szekeres models shows that observers in favourable positions would see galaxies drift across the sky at a rate of up to $10^{-6}$ arc seconds per year. Such a drift would be possible to measure using devices that are under construction; the required time of monitoring would be $\approx10$ years. This effect is zero in the FLRW models, so it provides a measure of inhomogeneity of the Universe. In the Szekeres models, the condition for zero drift is zero shear. But in the shearfree normal models, the condition for zero drift is that, in the comoving coordinates, the time dependence of the metric completely factors out.'
author:
- 'Andrzej Krasiński$^*$'
- 'Krzysztof Bolejko$^*$'
title: Exact inhomogeneous models and the drift of light rays induced by nonsymmetric flow of the cosmic medium
---
Inhomogeneous models in astrophysics$^1$
========================================
\[introduction\]
Just as was the case with our earlier reviews [@Kras1997; @BKHC2010; @BCKr2011], we define inhomogeneous cosmological models as those exact solutions of Einstein’s equations that contain at least a subclass of nonvacuum and nonstatic Friedmann – Lemaître – Robertson – Walker (FLRW) solutions as a limit. The reason for this choice is that such FLRW models are generally considered to be a good first approximation to a description of our real Universe, so it makes sense to consider only those other models that have a chance to be a still better approximation. Models that do not include an FLRW limit would not easily fulfil this condition.
This is a live topic, with new contributions appearing frequently, so any review intended to be complete would become obsolete rather soon. Ref. 1 is complete until 1994, Ref. 2 is a selective update until 2009 and Ref. 3 is a more selective update until the end of 2010. The criterion of the selection in the updates was the usefulness of the chosen papers for solving problems of observational cosmology.
In the very brief overview given here we put emphasis on pointing out and correcting the erroneous results that exist in the literature and are being taken as proven truths. Some of them have evolved to become research paradigms, with many followers, some others proceed in this direction. We hope to stop this process, which disturbs and slows down the recognition of the inhomogeneous models as useful devices for understanding the observed Universe.
We first present the two classes of inhomogeneous models that, so far, proved most fruitful in their astrophysical application: the Szekeres model [@Szek1975], and its spherically symmetric limit, the Lemaître [@Lema1933] – Tolman [@Tolm1934] (L–T) model. Then we briefly mention the real-time cosmology program, and we give an overview of the erroneous ideas. Finally, we present an effect newly calculated in a few classes of models that is possible to observe and can become a test of homogeneity of the Universe: the drift of light rays induced by nonsymmetric flow of the cosmic medium
The Szekeres solution {#Szek}
=====================
The (quasi-spherical) Szekeres solution [@Szek1975; @PlKr2006] is, in comoving coordinates $${\rm d} s^2 = {\rm d} t^2 - \frac {{\cal E}^2 {(\Phi / {\cal E}),_r}^2} {1 +
2E(r)} {\rm d} r^2 - \frac {\Phi^2} {{\cal E}^2} \left({\rm d} x^2 + {\rm d}
y^2\right),$$ $${\cal E} {\ {\overset {\rm def} =}\ }\frac {(x - {P})^2} {2{S}} + \frac {(y - {Q})^2} {2{S}} + \frac
{{S}} 2,\label{2.1}$$ where $E(r)$, $M(r)$, ${P}{(r)}$, ${Q}{(r)}$ and ${S}{(r)}$ are arbitrary functions and $\Phi(t,r)$ obeys $$\label{2.2}
{\Phi,_t}^2 = 2E(r) + \frac {2 {M}{(r)}} {\Phi} + \frac 1 3 \Lambda \Phi^2.$$ The source in the Einstein equations is dust, whose mass density in energy units is $$\label{2.3}
\kappa \rho = \frac {2 \left(M / {\cal E}^3\right),_r} {(\Phi / {\cal E})^2
\left(\Phi / {\cal E}\right),_r}.$$ Eq. (\[2.2\]) implies that the bang time is in general position-dependent: $$\label{2.4}
\int\limits_0^{\Phi}\frac{{\rm d} \widetilde{\Phi}}{\sqrt{2E + 2M /
\widetilde{\Phi} + \frac 1 3 \Lambda \widetilde{\Phi}^2}} = t - {t_B(r)}.$$
The general Szekeres metric has no symmetry. It contains the spherically symmetric Lemaître [@Lema1933] – Tolman [@Tolm1934] (L–T) model as the limit of $(P, Q, S)$ being all constant. The latter is usually used with a different parametrisation of the spheres of constant $(t, r)$, namely $$\label{2.5}
{\rm d}s^2 = {\rm d}t^2 - \frac{R,_r^2}{1 + 2E} {\rm d}r^2 - R^2(t,r) \left({\rm
d}\vartheta^2 + \sin^2 \vartheta {\rm d}\varphi^2 \right),$$ where $R \equiv \Phi$, still obeying (\[2.2\]), and (\[2.3\]) simplifies to $$\label{2.6}
\kappa \rho = \frac {2 M,_r} {R^2 R,_r}.$$
The Friedmann limit follows when, in addition, $\Phi (t,r) = r S(t)$, $2E = - k
r^2$ where $k =$ const is the FLRW curvature index, and $t_B$ is constant.
Real-time cosmology
===================
There are ways in which the expansion of the Universe might be directly observed. The authors of Refs. [@QQAm2009; @QABC2012] composed them into a paradigm termed [***real-time cosmology***]{}. One of them is the [***redshift drift***]{}: the change of redshift with time for a fixed light source, induced by the expansion of the Universe.
Consider, as an example, an L–T model with $\Lambda$, given by (\[2.5\]) with $R$ obeying (\[2.2\]). Along a single radial null geodesic, directed toward the observer, $t = T(t_o,r)$ (where $t_o$ is the instant of observation) the redshift is [@Bond1947] $$\label{3.1}
1 + z(t_o,r) = {\rm exp} \left[\int_{r_{\rm em}}^{r_{\rm obs}} \frac
{R,_{tr}(T(t_o,r), r)} {\sqrt{1 + 2E(r)}} {\rm d} r\right].$$
For any fixed source (i.e. constant $r$), eq. (\[2.2\]) defines a different expansion velocity $R,_t$ for $\Lambda = 0$ and for $\Lambda \neq 0$, and thus allows us to calculate the contribution of $\Lambda$ to $z$ via (\[3.1\]). With the evolution type known, the expansion velocity depends on $r$, and so allows us to infer the distribution of mass along the past light cone.
According to the authors of [@QQAm2009; @QABC2012], the “European Extremely Large Telescope” (E-ELT)[^1] could detect the redshift drift during less than 10 years of monitoring a given light source. The Gaia observatory[^2] could achieve this during about 30 years.
For more on redshift drift see the contribution by P. Mishra, M.-N. Célérier and T. Singh in these Proceedings.
The drift of light rays described in Sec. \[Szredshift\] and following is another real-time cosmology effect.
Erroneous ideas and paradigms {#errideas}
=============================
The L–T model was noticed in the astrophysics community – but:
1\. Many astrophysicists treat it as an enemy to kill rather than as a useful new device. (Citation from Ref. [@QABC2012]: The Gaia or E-ELT projects could distinguish FLRW from L–T “possibly eliminating an [*exotic alternative explanation to dark energy*]{}”).
2\. Some astrophysicists practise a loose approach to mathematics. An extreme example is to take for granted every equation found in any paper, without attention being paid to the assumptions under which it was derived.
Papers written in such a style planted errors in the literature, which then came to be taken as established facts. In this section a few characteristic errors are presented (marked by [**[$\bullet$]{}**]{}) together with their explanations (marked by [**[$*$]{}**]{}).
[**[$\bullet$]{}**]{} The accelerating expansion of the Universe is an observationally established fact (many refs., the Nobel Committee among them).
[**[$*$]{}**]{} The established fact is the [*smaller than expected observed luminosity of the SNIa supernovae*]{} (but even this is obtained assuming that FLRW is the right cosmological model). [***The accelerating expansion is an element of theoretical explanation of this observation.***]{} When the SNIa observations are interpreted against the background of a suitably adjusted L–T model, they can be explained by matter inhomogeneities along the line of sight, with decelerating expansion [@KHBC2010; @INNa2002].
[**[$\bullet$]{}**]{} Positive sign of the redshift drift is a direct confirmation of accelerated expansion of the space.
[**[$*$]{}**]{} The sign of redshift drift is only related to acceleration when homogeneity is assumed[@YKN2011] (see also Mishra, Célérier, and Singh in these Proceedings).
[**[$\bullet$]{}**]{} $H_0$, supernovae and the cosmic microwave background radiation (i.e. the size of the sound horizon and the location of the acoustic peaks) are sufficient to rule out inhomogeneous L–T models.
[**[$*$]{}**]{} These observations only depend on $D(z)$ and $\rho(z)$, and thus can be accommodated by the L–T model, which is specified by 2 arbitrary functions. Examples of such constructions are given in [@CBKr2010; @INNa2002]. In fact, as follows from the Sachs equations, in the approximation of small null shear (which for most L–T models works quite well[@BoFe2012]), there is a relation between $D(z)$, $\rho(z)$ and $H(z)$, meaning that these 3 observables are not independent[@BoFe2012], and thus allow the L–T model to accommodate more data on the past null cone.
[**[$\bullet$]{}**]{} The gravitational potential of a typical structure in the Universe is of the order of $10^{-5}$ and thus, by writing the metric in the conformal Newtonian gauge, one immediately shows that inhomogeneities can only introduce minute ($\approx 10^{-5}$) deviations from the RW geometry. This also means that the evolution of the Universe must be Friedmannian (common argument among cosmologists).
[**[$*$]{}**]{} Even if the gravitational potential remains small, its spatial derivatives do not, and thus the model has completely different optical properties than the Friedmann models. This was shown by rewriting one of the L–T Gpc-scale inhomogeneous models in the conformal Newtonian coordinates [@EMR2009]. The gravitational potential of this model remains small yet the distance–redshift relation deviates strongly from that for the background model.
The question remains whether small-scale fluctuations (of the order of tens of Mpc) could also modify optical properties and evolution of the Universe. The problem is complicated as it requires solving the Einstein equations and null geodesics for a general matter distribution, which, with current technology, is not possible to do numerically. Therefore, the problem has been addressed in a number of approximations and toy models. Recent studies showed that the optical properties along a single line of sight can be significantly different than in the Friedmann model. Yet, if averaged over all directions, the average distance–redshift relation closely follows that of the model, which describes the evolution of the average density and expansion rate[@BoFe2012]. Thus, the problem reduces to the following one: do the average density and expansion rate follow the evolution of the homogeneous model, i.e. is the evolution of the background affected by small-scale inhomogeneities and does it deviate from the Friedmannian evolution? Some authors claim that matter inhomogeneities cannot affect the background and that the Universe must have Friedmannian properties[@IsWa2006; @GrWa2011], while others argue for strong deviation from the Friedmannian evolution[@Wilt2011; @RBCO2011]. Studies of this problem within the exact models, like L–T, proved that under certain conditions the back-reaction can be large, while under others it remains quite small [@Suss2011], leaving the problem unsolved. For an informative description of the problem and techniques used to address it see Ref. [@BuRa2012; @CELU2011].
[**[$\bullet$]{}**]{} Fitting an L–T model to number counts or the $D_L(z)$ relation results in predicting a huge void, several hundred Mpc in radius, around the centre (too many papers to be cited, literature still growing). Measurements of the dipole component of the CMB radiation then imply that our Galaxy should be very close to the center of this void, which contradicts the “cosmological principle”.
[**[$*$]{}**]{} , for example constant $t_B$. When the model is employed at full generality, the giant void is not implied [@CBKr2010].
[**[$*$]{}**]{} , it cannot say which model is “right” and which is “wrong”.
[**[$\bullet$]{}**]{} The bang time function must be constant, otherwise the decaying mode of density perturbation is nonzero, which implies large inhomogeneities in the early universe (see, for example, Ref. [@ZMSc2008]).
[**[$*$]{}**]{} The bang time function describes the differences in the age between different regions of the Universe. In the L–T and Szekeres models it is also related to the amplitude of the decaying mode. However, these models describe the evolution of dust and therefore cannot be extended to times before the recombination, when the Universe was in a turbulent state: rotation, plasma, pressure gradients all did affect the proper time of an observer (${\rm d} \tau = {\rm d} t~ \sqrt{g_{00} (t,x^i)}$). Eventually, even if the Universe started with a simultaneous big bang, by the time of recombination, due to the standard physical processes, the age of the Universe would have been different at different spatial positions, giving rise to non-constant $t_B(r)$ of the dust L–T or Szekeres model that takes over there.
Moreover, the relation between the nonsimultaneous big bang and the decaying mode was established only for the L–T [@Silk1977; @PlKr2006] and Szekeres [@GoWa1982] models. For more general models, not yet explicitly known as solutions of Einstein’s equations, like the ones mentioned above, the connection may be more complicated and indirect. Thus, citing this relation for such a general situation is an illegitimate stretching of a theorem beyond the domain of its assumptions (see also the next entry below).
[**[$\bullet$]{}**]{} The L–T models used to explain away dark energy must have their bang-time function constant, or else they “can be ruled out on the basis of the expected cosmic microwave background spectral distortion” [@Zibi2011].
[**[$*$]{}**]{} . This sets them on the wrong track from the beginning.
[**[$*$]{}**]{} , [***one must apply it at every step of analysis***]{} of the observational data. To do so, would require a re-analysis of a huge pool of data. C. Hellaby with coworkers [@McHe2008] is working on such a program applied to the L–T model, but the work is far from being completed.
Lacking any better chance, we currently use observations interpreted in the FLRW framework to infer about the $M(r)$ and $t_B(r)$ functions in the L–T model. This is justified as long as we intend to point out possibilities, under the tacit assumption that these results will be verified in the future within a complete revision of the observational material on the basis of the L–T model. However, putting “precise” bounds on the L–T model functions using the self-inconsistent mixture of FLRW/L–T data available today is a self-delusion. An example: the spatial distribution of galaxies and voids is inferred from the luminosity distance vs. redshift relation that applies [*only*]{} in the FLRW models. Without assuming the FLRW background, we know nothing about this distribution until we reconstruct it using the L–T model from the beginning.
The L–T and Szekeres models cannot be treated as exact models of the Universe, to be taken literally in all their aspects. They are [*exact as solutions of Einstein’s equations*]{}, but when applied in cosmology, they are merely [*the next step of approximation after FLRW*]{}. If the FLRW approximation is good for some purposes, then a more detailed model, [*when applied in a situation, in which its assumptions are fulfilled*]{}, can only be better.
The redshift equations in the Szekeres models {#Szredshift}
=============================================
Consider two light rays, the second one following the first after a short time-interval $\tau$, both emitted by the same source and arriving at the same observer. The trajectory of the first ray is given by $$\label{5.1}
(t, x, y) = (T(r), X(r), Y(r)),$$ the corresponding equation for the second ray is $$\label{5.2}
(t, x, y) = (T(r) + \tau(r), X(r) + \zeta(r), Y(r) + \psi(r)).$$ This means that while the first ray intersects a hypersurface $r = r_0$ at $(t,
x, y) = (T, X, Y)$, the second ray intersects the same hypersurface not only later, but, in general, at a different comoving location. [***$\Longrightarrow$ In general the two rays will intersect different sequences of intermediate matter worldlines***]{}.
The same is true for nonradial rays in the L–T model.
Consequently, the second ray is emitted in a different direction and is received from a different direction by the observer. Thus, a typical observer in a Szekeres spacetime should see each light source slowly [***drift across the sky***]{}. How slowly will be estimated further on. As will be seen from the following, [***the absence of this drift is a property of exceptionally simple geometries***]{} (or exceptional directions in more general geometries).
We assume that $(\zeta, \psi)$ and $({{{\rm d} {}} / {{\rm d} {r}}}) (\tau, \zeta, \psi)$ are small of the same order as $\tau$, so we neglect all terms nonlinear in any of them and terms involving their products.
For any function $f(t, r, x, y)$ the symbol $\Delta f$ will denote $$\label{5.3}
f(t + \tau, r, x + \zeta, y + \psi) - f(t, r, x, y)$$ [***linearized in $(\tau, \zeta, \psi)$.***]{} Note: [***the difference is taken at the same value of $r$***]{}. Applying $\Delta$ to the null geodesic equations parametrised by $r$ we obtain the equations of propagation of $(\tau,
\zeta, \psi)$ and $(\xi, \eta) {\ {\overset {\rm def} =}\ }({{{\rm d} {}} / {{\rm d} {r}}})$ $(\zeta, \psi)$ along a null geodesic – see both sets of equations fully displayed in Ref. [@KrBo2011].
Repeatable light paths {#repeat}
======================
There will be no drift when, for a given source–observer pair, each light ray will proceed through the same intermediate sequence of matter world lines. Rays having this property will be called [***repeatable light paths (RLP)***]{}. For a RLP we have $$\label{6.1}
\zeta = \psi = \xi = \eta = 0$$ all along the ray. The equations of propagation of $(\tau, \zeta, \psi, \xi,
\eta)$ become then overdetermined (3 equations to determine the propagation of $\tau$ along a null geodesic), and imply limitations on the metric components. They can be used in 2 ways:
1\. As the condition (on the metric) for [***all***]{} null geodesics to be RLPs.
2\. As the conditions under which special null geodesics are RLPs in subcases of the Szekeres spacetime.
In the first interpretation, the equations of propagation should be identities in the components of ${{{\rm d} {x^{\alpha}}} / {{\rm d} {r}}}$, and this happens when $$\label{6.2}
\Psi {\ {\overset {\rm def} =}\ }\Phi,_{tr} - \Phi,_t \Phi,_r/\Phi = 0.$$ This means zero shear, i.e. the Friedmann limit. Thus, we have the following
[**Corollary:**]{}
[***The only spacetimes in the Szekeres family in which [all]{} null geodesics have repeatable paths are the Friedmann models.***]{}
In fact, we proved something stronger. Since the drift vanishes in the Friedmann models, we have:
[**Corollary 2:**]{}
[***The presence of the drift would be an observational evidence for the Universe to be inhomogeneous on large scales.***]{}
In the second interpretation, there are only 2 nontrivial (i.e. non-Friedmannian) cases:
A. When the Szekeres spacetime is axially symmetric ($P$ and $Q$ are constant). In this case, the RLPs are those null geodesics that stay on the axis of symmetry in each 3-space of constant $t$.
B. When the Szekeres spacetime is spherically symmetric ($P, Q, S$ are all constant) – then it reduces to the L–T model. In this case, the radial null geodesics are [*the only*]{} RLPs that exist. A formal proof of this statement is highly complicated [@KrBo2011].
For non-radial rays in the L–T model, the non-RLP phenomenon was predicted, by a different method (and under the name of [***cosmic parallax***]{}), by Quercellini [*et al.*]{} [@QQAm2009; @QABC2012]. Their review contains a broad presentation of the real-time cosmology paradigm, oriented toward observational possibilities.

Examples of non-RLPs in the L–T model {#numex}
=====================================
The examples will show the non-RLP effect for nonradial null geodesics in two configurations of the L–T model, shown in Fig. \[densprof\], for different positions of the observers with respect to the center of symmetry. In Example 1 (see Fig. \[example1\]) we use Profile 1, the observer and the light source at 3.5 Gpc from the center of the void, the directions to them at the angle 1.8 rad, and [@BoWy2008]
$t_B =0$ – simultaneous Big Bang,
$\rho(t_0,r) = \rho_0 \left[ 1 + \delta - \delta \exp \left( - {r^2}/{\sigma^2}
\right) \right]$ – the density profile at the current instant,
$r {\ {\overset {\rm def} =}\ }R(t_0,r)$ – the radial coordinate,
$\rho_0 {\ {\overset {\rm def} =}\ }\rho(t_0,0) = 0.3 \times (3H_0^2)/(8 \pi G) \equiv 0.3 \rho_{\rm
critical}$ – the present density at the center of the void,
$H_0 = 72$ km s$^{-1}$ Mpc$^{-1}$ – the present value of the Hubble parameter,
$\delta = 4.05$, $\sigma = 2.96$ Gpc.
The source in Fig. \[example1\] sends three light rays to the same observer, the first of which was received by the observer $5 \times 10^9$ years ago, the second is being received right now, and the third one will be received $5 \times
10^9$ years in the future. Fig. \[fig1\] shows these rays projected on the space $t =$ now along the flow lines of the L–T dust. The source is at the upper right corner of the graph and the observer is at the upper left corner.
![The configuration of the light source and the observer in example 1.[]{data-label="example1"}](example1.ps)

The [***time-averaged***]{} rate of change of the position of the source in the sky, seen by the observer is $$\begin{aligned}
\label{7.1}
&& \dot{\gamma} = \frac {\rm angle\ between\ the\ earlier\ and\ the\ later\ ray}
{\rm time\ interval = 5 \times 10^9\ years} \nonumber \\
&& \sim 10^{-7}\ \frac {\rm arc sec} {\rm year}.\end{aligned}$$ The rate of drift in the next figures is calculated in the same way.
In Examples 2, 3 and 4 (Fig. \[fig2\]), the observer O is at $R_0$ from the center; the angle between the direction toward the galaxy ([$_*$]{}) and toward the origin is $\gamma$. For each $\gamma$ we calculated the rate of change $\dot \gamma$ by eq. (\[7.1\]), and the graphs in Fig. \[fig3\] show $\dot{\gamma}$ as a function of $\gamma$. All the examples have $d = 1$ Gly $\approx 306.6$ Mpc. Example 2 (solid line) has $R_0 = 3$ Gpc and Profile 1; Example 3 (dashed line) has $R_0 = 1$ Gpc and Profile 1; Example 4 (dotted line) has $R_0 = 1$ Gpc and Profile 2 (for which $\delta = 10.0$ instead of $\delta =
4.05$; this is a deeper void in a higher-density background). The amplitude is $\sim 10^{-7}$ for (3) and $\sim 10^{-6}$ for (2) and (4). With the Gaia accuracy of $5-20 \times 10^{-6}$ arcsec, we would need a few years to detect this effect.
![The configuration for Examples 2, 3, 4.[]{data-label="fig2"}](fig2.eps)

RLPs in shearfree normal models {#BarnesRLP}
===============================
In the Szekeres models, the condition for all null geodesics to be RLPs was the vanishing of shear. This suggests that the cause of the non-RLP phenomenon might be shear in the cosmic flow. To test this supposition, the existence of RLPs was investigated in those cosmological models in which shear is zero [@Kras2011] – the shearfree normal models found by Barnes [@Barn1973]. They obey the Einstein equations with a perfect fluid source and contain, as the acceleration-free limit, the whole FLRW family.
There are four classes of them: the Petrov type D metrics that are spherically, plane and hyperbolically symmetric, and the conformally flat metric found earlier by Stephani [@Step1967]. In the Petrov type D case, the metric in comoving coordinates is $${\rm d} s^2 = \left(\frac {F V,_t} V\right)^2 {\rm d} t^2 - \frac 1 {V^2}
\left({\rm d} x^2 + {\rm d} y^2 + {\rm d} z^2\right), \label{8.1}$$ where $F(t)$ is an arbitrary function, related to the expansion scalar $\theta$ by $\theta = 3 / F$. The Einstein equations reduce to the single equation: $$w,_{uu} /w^{2} = f(u), \label{8.2}$$ where $f(u)$ is an arbitrary function, while $u$ and $w$ are related to $(x, y,
z)$ and to $V(t,x,y,z)$ differently in each subfamily. We have $$(u, w) = \left\{ \begin{array}{ll} (r^{2}, V)
& \mbox{with spherical symmetry}, \\
\ & \ r^2 {\ {\overset {\rm def} =}\ }x^2 + y^2 + z^2;\\
(z, V) & \mbox{with plane symmetry};\\
(x/y, V/y) & \mbox{with hyperbolic symmetry}.\end{array} \right. \label{8.3}$$ The FLRW limit follows when $f = 0$ and $V = R(t) g(x,y,z)$.
The conformally flat Stephani solution [@Step1967; @Kras1997] has the metric given by (\[8.1\]), the coordinates are comoving, and $V(t, x, y, z)$ is given by $$\begin{aligned}
&&V = \frac 1 R \left\{1 + \frac 1 4 k(t) \left[\left(x - x_{0}(t)\right)^{2} +
\left(y - y_{0}(t)\right)^{2} \right.\right. \nonumber \\
&&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ + \left.\left.\left(z -
z_{0}(t)\right)^{2}\right]\right\},
\label{8.4}\end{aligned}$$ where $(R, k, x_0, y_0, z_0)$ are arbitrary functions of $t$. This a generalisation of the whole FLRW class, which results when $(k, x_0, y_0, z_0)$ are all constant. In general, (\[8.4\]) has no symmetry.
In these models, in the most general cases, generic null geodesics are not RLPs. Consequently, [***it is not shear that causes the non-RLP property***]{}.[^3]
In the general type D shearfree normal models, the only RLPs are radial null geodesics in the spherical case and their analogues in the other two cases. In the most general Stephani spacetime, RLPs do not exist. In the axially symmetric subcase of the Stephani solution the RLPs are those geodesics that intersect the axis of symmetry in every space of constant time. In the spherically-, plane- and hyperbolically symmetric subcases, the RLPs are the radial geodesics.
The completely drift-free subcases are conformally flat, but more general than FLRW. Their defining property is that their time-dependence in the comoving coordinates can be factored out, and the cofactor metric is static. The FLRW models have the same property. For example, in the drift-free spherically symmetric type D case: $$\begin{aligned}
{\rm d} s^2 = \frac 1 {V^2} && \hspace{-3mm} \left\{\left[\left(A_1 + A_2
r^2\right) \left(F S,_t {\rm d} t\right)\right]^2 - {\rm d} r^2\right.
\nonumber \\
&&- \left.r^2 \left({\rm d} \vartheta^2 + r^2
\sin^2 \vartheta {\rm d} \varphi^2\right)\right\}, \label{8.5}\end{aligned}$$ the whole non-staticity is contained in $V$: $$\label{8.6}
V = B_1 + B_2 r^2 + \left(A_1 + A_2 r^2\right) S(t).$$ The $(A_1, A_2, B_1, B_2)$ are arbitrary constants and $S(t)$ is an arbitrary function. This model is more general than FLRW because the pressure in it is spatially inhomogeneous. The FLRW limit follows when $A_1 \neq 0$ and $B_2 =
(A_2/A_1) B_1$.
Dependence of RLPs on the observer congruence
=============================================
The RLPs are defined relative to the congruence of worldlines of the observers and light sources. So far, we have considered observers and light sources attached to the particles of the cosmic medium, whose velocity field is defined by the spacetime geometry via the Einstein equations. But we could as well consider other timelike congruences, or spacetimes in which no preferred timelike congruence exists, for example Minkowski. It turns out that even in the Minkowski spacetime one can devise a timelike congruence that will display the non-RLP property [@Kras2012].
Take the Minkowski metric in the spherical coordinates $$\label{9.1}
{\rm d} s^2 = {\rm d} {t'}^2 - {\rm d} {r'}^2 - {r'}^2 \left({\rm d} \vartheta^2
+ \sin^2 \vartheta {\rm d} \varphi^2\right),$$ and carry out the following transformation on it: $$\label{9.2}
t' = (r - t)^2 + 1 / (r + t)^2, \qquad r' = (r - t)^2 - 1 / (r + t)^2.$$ The result is the metric $$\begin{aligned}
&& {\rm d} s^2 = \frac 1 {(r + t)^4}\left\{16 u \left({\rm d} t^2 - {\rm d}
r^2\right)\right. \\
&& \ \ \left.- \left(u^2 - 1\right)^2 \left({\rm d} \vartheta^2 + \sin^2
\vartheta {\rm d} \varphi^2\right)\right\}, \qquad u {\ {\overset {\rm def} =}\ }r^2 - t^2. \nonumber
\label{9.3}\end{aligned}$$ Now we assume that the curves with the unit tangent vector field $u^{\alpha} =
\left[(r + t)^2 / \left(4 \sqrt{u}\right)\right] {\delta^{\alpha}}_0$ are world lines of test observers and test light sources.
Proceeding as before we conclude that, with respect to this congruence, generic null geodesics in the Minkowski spacetime have non-repeatable paths. (The exception are those rays that are radial in the coordinates of (\[9.3\])). This is because the time-dependence of (\[9.3\]) cannot be factored out.
Acknowledgments {#acknowledgments .unnumbered}
===============
We are grateful to Marie Noëlle Célérier for several helpful comments.
[99]{} A. Krasiński, [*Inhomogeneous cosmological models*]{}. Cambridge University Press 1997, 317 pp, ISBN 0 521 48180 5.
K. Bolejko, A. Krasiński, C. Hellaby and M.-N. Célérier, Structures in the Universe by exact methods – formation, evolution, interactions. Cambridge University Press 2010, 242 pp, ISBN 978-0-521-76914-3.
K. Bolejko, M.-N. Célérier and A. Krasiński, [ *Class. Quant. Grav.*]{} [**28**]{}, 164002 (2011).
P. Szekeres, [*Commun. Math. Phys*]{}. [**41**]{}, 55 (1975).
G. Lemaître, [*Ann. Soc. Sci. Bruxelles*]{} [**A53**]{}, 51 (1933); English translation: [*Gen. Rel. Grav.*]{} [**29**]{}, 637 (1997).
R. C. Tolman, [*Proc. Nat. Acad. Sci. USA*]{} [**20**]{}, 169 (1934); Reprinted: [*Gen. Rel. Grav.*]{} [**29**]{}, 931 (1997).
J. Plebański, A. Krasiński, [*An introduction to general relativity and cosmology*]{}. Cambridge University Press (2006).
C. Quercellini, M. Quartin and L. Amendola, [ *Phys.Rev.Lett.*]{} [**102**]{}, 151302 (2009).
C. Quercellini, L. Amendola, A. Balbi, P. Cabella, M. Quartin, [*Phys. Reports*]{}. [**521**]{}, 95 – 134 (2012).
H. Bondi, [*Mon. Not. Roy. Astr. Soc.*]{} [**107**]{}, 410 (1947); Reprinted: [*Gen. Rel. Grav.*]{} [**31**]{}, 1777 (1999).
A. Krasiński, C. Hellaby, K. Bolejko and M.-N. Célérier, [*Gen. Rel. Grav.*]{} [**42**]{} 2453–75 (2010).
H. Iguchi, T. Nakamura and K. Nakao, [*Progr. Theor. Phys*]{}. [**108**]{}, 809 (2002).
C-M. Yoo, T. Kai, K-I. Nakao, [*Phys. Rev.*]{} [**D83**]{}, 043527 (2011).
M.-N. Célérier, K. Bolejko and A. Krasiński, [ *Astron. Astrophys.*]{} [**518**]{}, A21 (2010).
K. Bolejko, P.G. Ferreira, [*J. Cosmol. Astropart. Phys*]{}. [**05(2012)**]{}, 003 (2012).
K. Enqvist, M. Mattsson, G. Rigopoulos, [*J. Cosmol. Astropart. Phys*]{}. [**09(2009)**]{}, 022 (2009).
A. Ishibashi, R. M. Wald, [*Class. Q. Grav.*]{} [**23**]{} 235 (2006).
S. R. Green, R. M. Wald, [*Phys. Rev. D*]{} [**83**]{}, 084020 (2011).
D. L. Wiltshire, [*Class. Q. Grav.*]{} [**28**]{}, 164006 (2011).
X. Roy, T. Buchert, S. Carloni, N. Obadia [*Class. Q. Grav.*]{} [**28**]{}, 165004 (2011).
R. A. Sussman, [*Class. Q. Grav.*]{} [**28**]{}, 235002 (2011)
T. Buchert, S. Räsänen, [*Ann. Rev. Nuc. Part. Sci.*]{} [**62**]{}, 57 (2012).
C. Clarkson, G. Ellis, J. Larena, O. Umeh, [*Rept. Prog. Phys.*]{} [**74**]{}, 112901 (2011).
J. P. Zibin, A. Moss and D.Scott, [*Phys. Rev. Lett.*]{} [ **101**]{}, 251303 (2008).
J. Silk, [*Astron. Astrophys.*]{} [**59**]{}, 53 (1977).
S. W. Goode and J. Wainwright, [*Phys. Rev.*]{} [**D26**]{}, 3315 (1982).
J. Zibin, [*Phys. Rev.*]{} [**D84**]{}, 123508 (2011).
M. L. McClure, C. Hellaby, [*Phys. Rev.*]{} [**D78**]{}, 044005 (2008).
A. Krasiński and K. Bolejko, [*Phys. Rev.*]{} [**D83**]{}, 083503 (2011).
K. Bolejko and J. S. B. Wyithe [*J. Cosmol. Astropart. Phys*]{}. [**02(2009)**]{}, 020 (2009).
A. Krasiński, [*Phys. Rev.*]{} [**D84**]{}, 023510 (2011).
A. Barnes, [*Gen. Rel. Grav.*]{} [**4**]{}, 105 (1973).
H. Stephani, [*Commun. Math. Phys*]{}. [**4**]{}, 137 (1967).
A. Krasiński, [*Phys. Rev.*]{} [**D86**]{}, 064001 (2012).
[^1]: Now in the planning, to be built in Chile.
[^2]: http://sci.esa.int/science-e/www/area/index.cfm?fareaid=26
[^3]: Contrary to what the authors of Ref. [@QABC2012] claim throughout their paper.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Using data taken as part of the Bluedisk project we study the connection between neutral hydrogen ([[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}) in the environment of spiral galaxies and that in the galaxies themselves. We measure the total [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass present in the environment in a statistical way by studying the distribution of noise peaks in the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}data cubes obtained for 40 galaxies observed with WSRT. We find that galaxies whose [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass fraction is high relative to standard scaling relations have an excess [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass in the surrounding environment as well. Gas in the environment consists of gas clumps which are individually below the detection limit of our [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}data. These clumps may be hosted by small satellite galaxies and$\slash$or be the high-density peaks of a more diffuse gas distribution in the inter-galactic medium. We interpret this result as an indication for a picture in which the ${\rm H}{\textsc i}$-rich central galaxies accrete gas from an extended gas reservoir present in their environment.'
author:
- |
Jing Wang$^{1}$[^1], Paolo Serra$^{1}$, Gyula I. G. Józsa$^{2,3,4}$, Bärbel Koribalski$^{1}$,\
Thijs van der Hulst$^{5}$, Peter Kamphuis$^{1}$, Cheng Li$^{6}$ , Jian Fu$^{6}$, Ting Xiao$^{6}$,\
Roderik Overzier$^{7}$, Mark Wieringa$^{1}$, Enci Wang$^{6}$\
$^1$Australia Telescope National Facility, CSIRO Astronomy and Space Science, PO box 76, Epping, NSW 1710, Australia\
$^2$SKA South Africa Radio Astronomy Research Group, 3rd Floor, The Park, Park Road, Pinelands, 7405, South Africa\
$^3$Rhodes University, Department of Physics and Electronics, Rhodes Centre for Radio Astronomy Techniques & Technologies,\
PO Box 94, Grahamstown, 6140, South Africa\
$^4$Argelander-Institut für Astronomie, Auf dem Hügel 71, D-53121 Bonn, Germany\
$^5$University of Groningen, Kapteyn Astronomical Institute, Landleven 12, 9747 AD, Groningen, The Netherlands\
$^6$Key Laboratory for Research in Galaxies and Cosmology, Shanghai Astronomical Observatory, Chinese Astronomical Society,\
80 Nandan Rd, Shanghai 200030, China\
$^7$Observatório Nacional, Ministry of Science, Technology, and Innovation, Rio de Janeiro, Brazil\
date: 'Accepted 2014 ???? ?? Received 2014 ???? ??; in original form 2014 January'
title: 'An HI View of Galaxy Conformity: HI-rich Environment around HI-excess Galaxies'
---
galaxies:spiral; intergalactic medium
Introduction
============
Gas accretion is an important process in the evolution of galaxies. Galaxies like the Milky Way would run out of their gas in less than 1 Gyr given their current star formation rate unless their gas reservoir is continuously replenished (Kennicutt 1983). The source of this fresh gas is unknown but the majority is unlikely to be contained in gas-rich satellite galaxies because they can contribute at most 10 percent of the required gas mass (Kauffmann et al. 2010; Di Teodoro & Fraternali 2014, ). Likewise, high-velocity [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}clouds contribute 3-5 times less gas than required to sustain the Milky Way star formation (Richter 2012). A potentially more substantial source of fresh material is the cooling flow of low-metallicity, ionised gas in a galaxy’s halo (Shull et al. 2009). This may be related to the “fountain” gas which is driven to a height of a few kpc from the disc by star formation, and may trigger a gas inflow rate very close to the star formation rate (Fraternali et al. 2013). Recent theoretical studies suggest that the fountain driven gas naturally produces an outside-in shrinking galactic disk (Elmegreen et al. 2014), which is consistent with what is observed for nearby dwarf galaxies (Zhang et al. 2012). However, it does not explain the observed inside-out growth of more massive disk galaxies (J. Wang et al. 2011), or their chemical evolution (Spitoni et al. 2013), which have been both explained in terms of a cosmological radial gas accretion (Kauffmann 1996, Spitoni et al. 2013).
In the framework of the standard $\Lambda$CDM cosmology, two modes of (radial) gas accretion occur. Gas entering the potential well of a galaxy gets shocked to the virial temperature of the dark matter halo at an early epoch of infall, and then gradually cools and falls onto the central galaxy (Rees & Ostriker 1977; Silk 1977; Binney 1977, White &Rees 1978). Gas that has never been shock heated close to virial temperature falls onto the galaxy from outside the virial radius along filamentary structures (e.g. Kere[š]{} et al. 2005). This latter mode is termed “cold mode” accretion, and is believed to be the dominant way of gas accretion for galaxies in low mass dark matter halos (Kere[š]{} et al. 2005, Dekel & Birnboim et al. 2006). Numerical studies characterise the filaments of cold mode accretion as cool and clumpy clouds surrounded by ionised gas (Kere[š]{} & Hernquist 2009).
There have been many observational attempts to trace the cold mode accreting gas around galaxies. Most of the evidence comes from observing ionised gas at intermediate to high redshift. QSO absorption lines in Mg II with an inflowing velocity feature have been found to be prevalent at a distance of tens of kpc around star forming galaxies (e.g. Giavalisco 2011, Kacprzak et al. 2012). At least one of them is observed to have a filamentary structure (Rauch et al. 2011). Furthermore, metal-poor Lyman limit systems are observed to have characteristics expected for cold accretion (Lehner et al. 2013).
Tracing the cold-mode accreting gas in the neutral phase in galaxies at low redshift has been difficult. [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}clouds that form a filamentary structure have been found around some nearby galaxies (e.g., Shostak & Skillman 1989; Oosterloo et al. 2007; de Blok et al. 2014, for a review see Sancisi et al. 2008) and in the Milky Way (HVCs, Putman et al. 2002) but are more likely to trace interaction with companions and star formation feedback. None of them has been confidently demonstrated to be tracing cold-mode gas accretion. Although cold-mode accretion is believed to produce the highly clumpy galactic disks observed at high redshift (Agertz et al. 2009), we do not find an especially clumpy morphology for the ${\rm H}{\textsc i}$-rich galaxies when compared to control galaxies at low redshift (J. Wang et al. 2013, 2014). This may be because at low redshift the accreting gas gets slowed down and blends better with the hot halo gas around the galaxies (Putman et al. 2012). Hence, it may be better to search for the cold-mode accreting neutral gas well outside the galactic disks, by searching for signal from a large volume around the central galaxies. Here we attempt to perform exactly this search by analysing the data taken as part of the Bluedisk project (J. Wang et al. 2012) in a special way.
Previous analysis of the same Bluedisk dataset already shows that the gas richness of a galaxy is linked to the gas richness of its environment. In particular, the satellites of abnormally ${\rm H}{\textsc i}$-rich spirals are themselves abnormally [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}rich (E. Wang et al. 2015). This result may be seen as a new manifestation of the “conformity phenomenon”, whereby the colours of satellites correlate with the colour of central galaxies (Weinmann et al. 2006, Kauffmann et al. 2010, W. Wang et al. 2012). The work of E. Wang et al. (2015) was based on individual [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}detections in the Bluedisk data cubes and, therefore, is limited to galaxies with $\mHI>10^8~\ms$. As previously pointed out by Kauffmann et al. (2010), because the galactic conformity usually extends to scales far beyond dark matter halos, it implies that the satellite galaxies are probably tracing the underlying smooth gas reservoir that is available for fuelling the central galaxies. If this hypothesis is true, the conformity behaviour should still hold when it is examined at low luminosity levels (i.e. by tracing the less bright satellites or clouds in the general environment that are potentially far more numerous than the brighter satellites).
Motivated by these results we perform a new analysis of the Bluedisk data. Our aim is to trace the faint gas distribution that may surround galaxies below the detection limit of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}data. Our method consists of adding up the flux of (both positive and negative) noise peaks in the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}data cubes out to a radius of 16 arc min (typically 500 kpc), trying to detect a statistical [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}signal (i.e. an excess of positive detections) and studying whether its properties correlate with the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}content of the central galaxy.
The paper is organised as follows. In section 2, we describe the Bluedisk data used in this analysis. In section 3, we introduce our new method to measure a total [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass in faint systems around primary galaxies through examining the noise peaks. In section 4, we discuss the origin of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass measured from an analysis of the noise peaks. We find that the excess [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass fraction in the primary galaxies is correlated with an excess [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass in the surrounding environment. In section 5, we discuss the implication of our results for cold mode gas accretion in galaxies. A $\Lambda$CDM cosmology with $\Omega_{m}=0.3$, $\Omega_{\lambda}=0.7$ and $h=0.7$ is assumed throughout the paper.
The sample
==========
The Bluedisk dataset
--------------------
The Bluedisk project aims at searching for clues about gas accretion and galactic disk formation in the local universe. Details about the sample, data processing and analysis can be found in Paper I and II, and here we only review the most relevant information.
The sample consists of 50 massive galaxies (M$_*>10^{10} \ms$) with a redshift between 0.018-0.03 (corresponding to a luminosity distance of 70-130 Mpc). From the atlas in Paper I, most of them are blue spiral galaxies. The target fields are covered by the Sloan Digital Sky Survey (SDSS, Abazajian et al. 2009), the Galaxy Evolution Explorer (GALEX) imaging survey (Martin et al. 2005), and the Wide-Field Infrared Survey Explorer (WISE, Wright et al. 2010), allowing a multi-wavelength analysis. Radio 21 cm synthesis data were obtained with the Westerbork Synthesis Radio Telescope (WSRT), with the main goal of observing the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}emission line. In this paper, we make use of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}data cubes optimised for high sensitivity at lower resolution, which are produced with Robust 0 weighting and 30$"$ tapering. The data cubes have a typical rms of 0.37 mJy beam$^{-1}$, a typical beam size of 38$'' \times 36''$ and a velocity resolution of 12.3 km/s (after Hanning smoothing), which provides sufficient sensitivity to detect point sources down to an [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass of a few $10^8\ms$. The primary beam has a FWHM (full-width-at-half-maximum) of $\sim0.5\degr$, corresponding to a $\sim1$ Mpc sky region around the target galaxies, hence the data is very suitable for studying the relatively large-scale [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}environment around the central sources. The details of the primary beam correction are discussed in E.Wang et al.(2015).
Following Paper I, we use the 42 isolated target galaxies for this study. Hence, the local environment is different from groups or clusters, where the dominant process is gas stripping rather than accretion. We further exclude two galaxies (galaxy 14 and 27) with the lowest redshift, so that we have a narrow redshift range of 0.023–0.03 (a luminosity distance of 100-130 Mpc). We call these 40 target galaxies “primary galaxies”.
These galaxies have a broad distribution in [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass fraction $\fHI=\log \HIs$, ranging from lower than -2 to higher than 1 (Figure \[fig:sample\]). Compared to other samples featuring ${\rm H}{\textsc i}$-rich galaxies (like the samples from Lemonias et al. 2014 and the HIghMass project, High [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}Mass, Huang et al. 2014), the Bluedisk sample has the advantage of being able to compare the most [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}rich galaxies with those of normal [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}content. As we will show later in Section \[sec:conformity\], this is crucial for producing our main results.
As mentioned before, the aim of the paper is to investigate the connection between the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}in the (central) primary galaxies and the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}in the environment. In addition to $\fHI$, we also use a few other parameters to quantify the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}richness of primary galaxies. Using the GALEX Arecibo SDSS Survey (GASS, Catinella et al. 2010), Catinella et al. (2013, [**C13 for short hereafter**]{}) calibrated a photometric estimator of $\fHI$ in galaxies with NUV$-r<4.5$ (which is also the colour range of the Bluedisk sample). The estimator is a combination of the NUV$-r$ colour and the mass surface density, reflecting the connection between star formation rate and cold gas content. The difference between the observed and the estimated $\fHI$, which we denote as $\Delta_{C13}\fHI$, measures the excess gas, and can be used to indicate [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}richness. Xiao et al. (in prep) further add the optical concentration parameter and stellar mass to the estimator, giving more control on the internal structure of galaxies. We also use the difference from this 4-parameter (4p) estimator ( $\Delta_{4p}\fHI$) as a measure of [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}richness.
![ [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass fraction of the Bluedisk sample (dark grey dots), the HIghMass sample (triangles), and the sample from Lemonias et al. (2014) (L14, diamonds). The dashed line shows the median relation between $\fHI$ and stellar mass for general galaxies, taken from Catinella et al. (2010). The dotted line is the dashed line 0.6 dex upward, showing the original division between the Bluedisk high $\fHI$ and control galaxies.[]{data-label="fig:sample"}](plot/sample.ps){width="8cm"}
Method
======
As discussed in the introduction section, the conformity behaviour of central and bright satellite galaxies can be interpreted in terms of an underlying extended reservoir of cold gas. If this hypothesis is correct, we expect to detect conformity also at lower [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass levels than probed by E. Wang et al. (2015), in a regime dominated by very faint satellite galaxies and$\slash$or gas clouds in the inter-galactic medium. In this section we introduce our technique to detect this faint [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}gas, in the environment of the Bluedisk galaxies.
The method entails an analysis of the noise peaks (both positive and negative) in [**the cleaned data cubes**]{} in order to obtain a statistical detection. Because the noise is distributed symmetrically around 0 (both the median and mean background of a cube are typically within $\pm$10$^{-5}$ mJy beam$^{-1}$, while the rms of a cube is $\sim$0.4 mJy beam$^{-1}$), a positive noise peak will statistically be cancelled out by a negative noise peak of the same level. If in addition to noise there is a collection of very faint [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}objects, then these will add a slight positive signal to the noise, that will elevate the positive noise peaks or diminish negatiive noise peaks. This bias is detectable by a careful analysis of the distribution of the noise values throughout the cube. We do this by selecting all (positive and negative) noise peaks above 4 $\sigma$ in the area of interest, after removing the signal that has been reliably detected in the source finding procedure ( see section \[sec:scfinding\]).
Selecting noise peaks above a certain threshold has an important advantage compared to integrating over all voxels within a large area. Bright HI emission as well as diffuse faint HI emission in the data cubes is cleaned by means of clean masks (Paper I). The clean masks do not include the faint noise peaks used for our statistical analysis here, and these are therefore not cleaned. Because the integral over the [**dirty beam**]{} over a large area goes to zero it becomes very difficult to determine fluxes over extended areas without proper deconvolution. This is inherent to the sidelobe structure of the PSF. Any blind measurement of the total flux over a large area is therefore fraught with problems. However, by pre-selecting noise peaks, we analyse only the central part of the PSF and exclude any sidelobes. We are therefore unable to detect the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass of a very extended, smoothly distributed medium, (this is a consequence of using an interferometer with limited short baselines), but we are able to detect small scale faint emission using the method outlined here.
In the following sections, we describe our procedure in details. We improve the data quality by flattening the spectral baselines (section \[sec:flattening\]). We use the source-finding application SoFiA (Serra et al. 2015) to extract all signal exceeding 4 $\sigma$. After removing all bright, reliable detections (section \[sec:scfinding\]), we are left with “noise” peaks which we analyse in order to detect a statistical signal. We perform a series of careful tests to examine the robustness of the accumulative signal against various data reduction artefacts (section \[sec:testing\]).
Flattening the spectral baselines {#sec:flattening}
---------------------------------
In order to improve the quality of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}data cubes we fit second-order polynomial functions in moving windows along each line of sight. This allows us to obtain global functions describing the shape of the spectral baselines without specifying their functional form. In practice, we convolve each spectrum with the Savitzky-Golay filter (S-G, Savitzky & Golay 1964), an analytical solution for the local fitting process. We adopt a S-G filter with a half width of 70 channels (840 km/s), wider than the velocity width of all the galaxies in our data cubes. This ensures that the smoothed spectrum describes the large-scale shape of the baseline only. In Figure 2 we show the shape of the filter and demonstrate the convolution effect on simple functions, including the presence of edge effects. Within our analysis range of $\pm$500 km/s around the spectrum centre (channels 34 to 114), the S-G filter conserves the shape of the functions with a maximum deviation of less than 15%.
Before convolving the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}spectra with the S-G filter we mask channels that belong to one of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}detections presented in Paper I (see also Sec. 3.2 below). We refer to these channels as the initial flagging set, and replace them with values interpolated from a first order polynomial fitting to the un-flagged part of spectrum. We calculate $\sigma_m$, the median absolute deviation of the original spectrum from the convolved spectrum, and update the mask by including all channels that deviate by more than $\pm$3 $\sigma_m$ from the convolved spectrum. A flagged channel is replaced by the value interpolated from a first order polynomial fit to the un-flagged part of the whole spectrum, if it is in the initial flagging set, or in a flagged region that is wider than 10 channels and less than 40 channels from one end of the spectrum (potential bright sources that lie near the edge of the spectrum and cause incorrect continuum over-subtraction during the former data reduction). Otherwise the flagged channel (which is not likely to belong to a bright source) gets replaced by a first order polynomial interpolation from nearby channels. We repeat the above steps of flagging and convolution until $\sigma_m$ (with typical value of a few times 0.1 mJy beam$^{-1}$) varies by less than 10$^{-5}$ mJy beam$^{-1}$, or after 10 iterations. Figure \[fig:subcont\] shows that this procedure models a variety of spectral baseline shapes successfully.
The final step of this procedure consists of combining all spectral baseline models into individual cubes. We smooth these cubes with a gaussian kernel with a FWHM of 2 pixels in the Ra-Dec direction in order to reduce the fluctuation between adjacent pixels (this does not significantly modify the PSF (which has a typical FWHM of 3.7 pixels). Finally, we subtract the resulting spectral baseline model cube from the original cube. After this subtraction, the median rms of each data cube drops very slightly from 0.367 to 0.363 mJy beam$^{-1}$ (and from 0.372 to 0.362 mJy beam$^{-1}$ for the inner channel range 34 to 114). Hence while flattening the baselines, the method does not significantly change the global noise properties of the cube, In Figure \[fig:projection\], we compare two data cubes, using the same position-velocity cut, one as obtained before the additional continuum subtraction, and one as obtained after the additional continuum subtraction. While the noise features look similar, the continuum subtraction is obviously improved.
![The S-G filter (top row) and examples of the convolution effect on simple functions. The area under the S-G filter in the top row is normalised. From the second to the bottom row, the dark grey curves are linear, second order polynomial and third order polynomial functions and a flat line with a spike on one side. The magenta curves are the result of the grey curves convolved with the S-G filter. The horizontal dashed lines mark the y axis position of 0, and the vertical lines mark the channels 34 and 114.[]{data-label="fig:sgfilter"}](plot/sgfilter.ps){width="8cm"}
![Examples of spectra before (black lines) and after applying the S-G filter (magenta lines). The dashed lines mark the zero level.[]{data-label="fig:subcont"}](plot/fitcont.ps){width="8cm"}
![ An example of an x-z projection from an original [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}cube (top) and that after applying the improved continuum subtraction discussed in section \[sec:flattening\] (bottom). The displayed grey scale ranges from -3 to 3 $\sigma$ around the mean of the cube. []{data-label="fig:projection"}](plot/projection40-159.ps){width="7.5cm"}
Source finding and statistical detection of HI in the environment {#sec:scfinding}
-----------------------------------------------------------------
We extract sources (both real detections as well as noise peaks to be used for our statistical analysis) from the improved [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}data cubes using the source detection software SoFiA. We refer to Serra et al. (2015) for a detailed description of the software. Here, we only outline the most relevant steps and describe the key parameter choices.
SoFiA first convolves the data with various smoothing kernels and selects voxels with absolute values above a certain threshold. We use smoothing kernels with a width of 0 and 3 pixels in the projected sky direction (roughly the size of the synthesised beam), and a width of 0, 3 and 5 pixels (corresponding to 0, 36 and 60 km$/$s) in the velocity direction (smoothing with a kernel with width of 0 is equivalent to no smoothing). Each channel of the cube is weighted according to the noise, so that noise variations along the velocity axis are removed. This step is very useful to suppress the detection of weak RFI (radio frequency interference) residuals. We adopt a detection threshold equal to 4 $\sigma$. We note that the clipping is performed on the absolute values of the voxels in the filtered data cubes, meaning that our initial source list contains sources with negative total flux. These parameter settings for SoFiA are chosen after several tests and provide a good balance between suppressing the noise and gathering sufficient statistics for the following analysis..
SoFiA calculates a reliability index $R$ for each source with positive total flux (Serra et al.2012, 2015). We chose 0.01 Jy beam$^{-1}$ as a minimum total flux and R$>$0.99 to classify a positive detection as a reliable source. We will refer to the remaining positive and negative detections (the noise peaks) as the “candidates” hereafter. While the division between reliable sources and candidates is not strict, we will see in section \[sec:matchopt\] that the candidates close to the threshold of reliable sources do not dominate the total signal in candidates.
Using these selection criteria, besides the 40 primary galaxies, SoFiA detects in total 59,757 sources from the 40 cubes, of which, using the scheme as described, 137 are classified as reliable. 42 of the reliable sources lie within a projected distance ($D_{proj}$) of 500 kpc and a systematic velocity difference ($\Delta V_{sys}$) of 500 km$/$s around the primary galaxies. 95$\%$ of them have $\mHI>1.62\times10^8\ms$. All of them can be associated with an optical galaxy down to a $r$ band magnitude of 22 by requiring the centre to be offset by less than the size of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}beam. The connection between primary galaxies and these reliable sources (satellite galaxies) is studied in detail in E. Wang et al. (2015). One of the key findings is that the satellites of ${\rm H}{\textsc i}$-excess galaxies ($\Delta_{C13}\fHI>0$; see Sec. 2) have higher [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}excess than the satellites of ${\rm H}{\textsc i}$-normal galaxies ($\Delta_{C13}\fHI<0$). Here we aim to test whether such a connection between the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}content of galaxies and that of their environment holds also when studying the fainter, candidate [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}sources.
We find 6,369 candidates that lie within a $D_{proj}$ of 100 to 500 kpc and $V_{sys}$ distance of 500 km$/$s from the primary galaxies, and these will serve as the main sample for analysis in what follows. We exclude the inner 100 kpc region because some of our [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}rich galaxies have large [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}disks extending to $\sim$ 100 kpc. The radial range is always confined within the full-width-half-power of the primary beam. Most of the candidates have [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}masses (absolute values for the negative candidates) within the range of $4.0\times10^6\ms$ and $1.4\times10^8\ms$ (5 and 95 percentiles of the distribution).
The ensemble of all positive and negative candidate detections results in a net positive [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}signal in the environment around Bluedisk galaxies. We show this by calculating for each cube the accumulative [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass ($\mHIcand$) by summing up all negative and positive candidates in the volume defined above. To further improve the signal-to-noise ratio, we calculate the average $\mHIcand$ of the 40 cubes. We calculate the error bars as a combination of the variation over the data cubes (var1) and the mean variation within a data cube (var2). The variation over the data cubes (var1) is calculated through a standard bootstrapping procedure. Taking the whole sample of 40 cubes for example, we randomly select 40 cubes from the sample (repetition allowed) and calculate the average $\mHIcand$. We repeat this step for 1000 times and obtain a distribution consisting of 1000 values of average $\mHIcand$. var1 is calculated as the variation of this distribution. The final error is calculated as $\sqrt{var1+\overline{var2}}$. We derive an average $\mHIcand$ of $\sim1.1\times10^8\ms$ for the 40 cubes within the whole analysis volume.
The question is, however, whether and how significantly $\mHIcand$ is affected by remaining artefacts in the data. Asymmetrically distributed noise can come from weak RFI residuals that are not completely removed during data reduction, and can also be produced in the continuum subtraction and cleaning processes. Hence in the following sub-sections we undertake a series of tests to identify and quantify these effects, before moving on to the scientific interpretation in section \[sec:result\].
Influence of synthesis data reduction on the mass in candidates {#sec:testing}
---------------------------------------------------------------
We begin this series of tests by verifying that the white noise in the data does not accumulate to a significant net signal. We test this by applying the same source finding and $\mHIcand$ measurements presented in section \[sec:scfinding\] to the Stokes Q cubes. Since no polarised signal in the line data is expected, the result is expected to be 0 unless indeed a positive accumulative noise signal is present. Figure \[fig:mhidistr\_qq\] confirms that $\mHIcand$ of the Stokes Q cubes is indeed close to 0.
In addition, we visually inspect the data cubes carefully in various projections. The well trained eye is able to spot and identify systematics in the data. Because the WSRT is a linear E-W array, bright RFI residuals show patterns of parallel stripes and bright cleaning residuals show patterns of elliptical rings around bright sources on the Ra-Dec projections. Improper continuum subtraction results in negative regions on the Ra-velocity or Dec-velocity projections (see Figure \[fig:projection\]). We find no obvious cleaning or continuum subtraction residuals. We find some very weak underlying RFI patterns in a few channels, but none of our [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}candidates are associated with any of those. This is partly because the source finder takes into account noise variations within cubes (including those caused by residual RFI; section \[sec:scfinding\])
We continue our search for possible sources of systematic error in our measurement of $\mHIcand$ by investigating the effect of incorrect cleaning and continuum subtraction. Any systematics related to clean residuals [**(i.e. the presence of residual side lobe flux in the HI cubes)**]{} should correlate with a higher cleaned flux or a larger cleaned region. This opens the possibility for the following tests.
We calculate the sum of all positive voxels belonging to candidates and the absolute sum of all the negative voxels belonging to candidates, and we refer to them as $f+_{cand}$ and $f-_{cand}$. We calculate the total flux in the reliable sources and the total flux in the candidates, and refer to these as $f_{rel}$ and $f_{cand}$ respectively. We calculate $f+_{cand}$, $f-_{cand}$, $f_{cand}$ and $f_{rel}$ for each channel map within 500 km$/$s from the systemic velocity and 500 kpc from the position of the primary galaxy. These are the velocity range and sky region within which we perform our analysis. The top row of Figure \[fig:error\_cleaning\] shows that there is no correlation between $f-_{cand}$ and $f_{rel}$, suggesting that clean residuals do not significantly affect the noise (negative voxels) of the data cube. There is a very weak relation between $f+_{cand}$ and $f_{rel}$, resulting in a systematic linear relation of $f_{cand}$ increasing slightly as a function of $f_{rel}$. We argue that part of the relation reflects galaxy-environment connections rather than artefacts, as we will show later. If we attribute the relation fully to clean artefacts around reliable sources, on average an [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass of 0.11$\times10^8\ms$ related to clean artefacts is present in each data cube, which is an order of magnitude lower than the average [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass contained in the candidates ($\mHIcand\sim1.1\times10^8~\ms$, section \[sec:scfinding\]). In the bottom row of Figure \[fig:error\_cleaning\], we perform a similar analysis as in the top row but replace $f_{rel}$ with $N_{rel}$, the number of voxels in the reliable sources. We get similar trends, and the linear relation between $f_{cand}$ and $N_{rel}$ suggests that on average a maximum positive bias in [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass of 0.07$\times10^8\ms$ is associated with the cleaning of reliable sources in a data cube. These tests demonstrate that the cleaning process may indeed leave a small amount of positive residual flux, but it does not significantly affect $\mHIcand$ of the candidates.
Are there faint systematic residuals from continuum subtraction that can not be directly caught by eye? We extract continuum sources from the continuum maps, with a 3 $\sigma$ detection threshold. These sources have a 5 percentile of 0.94 mJy in the flux distribution. We find that 99$\%$ of the flux in candidates comes from sight lines without detectable continuum sources. The fraction is higher than 95% for both positive and negative candidates, if we count them separately. Hence continuum subtraction residuals are not likely to significantly affect $\mHIcand$.
We also find that the flux in candidates is not correlated with radio flux in continuum sources (with a Pearson correlation coefficient of 0.06) or the number of continuum sources (with a Pearson correlation coefficient of -0.27). We divide the galaxy sample equally into two subsets by the total flux in continuum sources, maximum flux in continuum sources and number of continuum sources, and compare their $\mHIcand$ (see Figure \[fig:HIprof0\]). We find no differences for data cubes having different values for the total or maximum radio flux present in the form of continuum sources. There is a trend that data cubes with fewer radio continuum sources have higher $\mHIcand$. However, the number of background continuum sources does not correlate with any of the galaxy properties discussed in section \[sec:conformity\], and we conclude that this correlation does not affect our main results.
Finally, we note that cumulatively there is a weak gradient of $\mHIcand$ on the low and high V$_{sys}$ sides of the primary galaxies, which is unlikely to be physical but does not significantly affect the main result of this paper. We refer the readers to Appendix A for a detailed investigation of this issue.
{width="14.cm"}
{width="16.cm"} {width="16.cm"}
{width="4.cm"} {width="4.cm"} {width="4.cm"}
A summary of the method
-----------------------
To summarise, we used SoFiA to extract candidates with typical mass of 10$^7\ms$, which is lower than that of reliably detected sources but higher than most of the data reduction artefacts. We have shown that the total [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass in candidates, $\mHIcand$, is not sensitive to artefacts produced in the continuum subtraction and cleaning processes of our data reduction, and potentially traces a reservoir of low surface density [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}gas in data cubes. In the following section we study the nature of these candidates and investigate whether $\mHIcand$, i.e., the gas mass contained in the environment of Bluedisk galaxies, correlates with properties of the primary galaxies.
Results {#sec:result}
=======
The nature of the HI mass in candidates {#sec:matchopt}
---------------------------------------
In this section,we attempt to understand better the nature of the signal in the cumulative $\mHIcand$ around the Bluedisk primary galaxies. From the top row of Figure \[fig:mhidistr\], we can see that the average profile of $\mHIcand$ around the 40 primary galaxies rises steadily to $\sim1.1\times10^8\ms$ from a projected distance of 100 kpc out to 500 kpc, and from a systematic velocity distance of 0 out to $\sim\pm300$ km$/$s. In the bottom-left panel, the accumulative $\mHIcand$ is plotted as a function of absolute [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass of individual candidates, and we can see that $\mHIcand$ is not dominated by the most massive candidates. The candidates with $\mHI>10^8\ms$ contributes only $\sim1/3$ of the net signal. We extract the spectra for all the candidates, shift them to have the same central systematic velocity of 7700 km$/$s, and stack them. The stacked spectrum, as displayed in the bottom-right panel of Figure \[fig:mhidistr\], is a narrow emission line with a gaussian fit $\sigma$ of 18.9$\pm$5.7km$/$s. Removing the resolution effect, the line has a width of $\sigma\sim14.3$ km$/$s. If we assume rotational systems and apply this velocity width to the baryonic Tully-Fisher relation (McGaugh et al. 2000), it corresponds to a baryonic mass of 4.4$\pm$2.5$\times10^7\ms$, similar to the typical $\mHI$ of a positive candidate. The candidates could be low mass satellite galaxies, [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}clouds, or [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}clumps in low surface density [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}disks.
If part of $\mHIcand$ probes [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}in galaxies, we expect the positive candidates to trace galaxies more closely than the negative candidates. As a test, we select galaxies with $r$ band flux brighter than 20 mag from the SDSS DR7 photometric catalog. For each candidate we search for the optical galaxy with the smallest projected distance. We compare the distribution of matching distances for the positive candidates, the negative candidates and simulated random positions. In the left panel of Figure \[fig:match\], the negative candidates closely follow the curve for random positions (with a K-S test probability of 0.95, meaning that at a 5$\%$ significance the null hypothesis of the two distributions being drawn from the same parent distribution can be rejected). We find that a slightly higher fraction of positive candidates can be matched to optical galaxies than the negative candidates at small matching distance ($<50$ arsec, $\sim$1.5 times the FWHM of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}PSF), although a K-S test probability of 0.28 suggests the difference is not strong. If we further take only the positive candidates with $\mHI>1.25\times10^7\ms$ (70 percentile in the mass distribution of positive candidates; based on the bottom-left panel of Figure \[fig:mhidistr\], these are the candidates that contribute nearly all the net $\mHIcand$ signal) into account, we find a significant difference in the distribution of matching distances from the negative candidates (K-S test probability $\sim$ 0.04) and from the simulated random positions (K-S test probability $\sim$0.01). We further check if this small excess of optical matches for the bright positive candidates might be due to residual from the continuum subtraction around continuum sources associated with the optical sources. In the middle panel of Figure \[fig:match\], we match the projected position of negative, positive and bright positive candidates to continuum sources. We find the three types of candidates to be indistinguishable in the distribution of matching distances when the distance is smaller than 50 arcsec. The K-S test probability is 0.98 for the comparison between negative and positive candidates. The K-S test probability is 0.14 for the comparison between negative and bright positive candidates, suggesting a weak difference, but the curve shows that the bright positive candidates are less likely to be found along sight lines of continuum sources than the negative candidates. Hence we can conclude that at least the brightest $\sim1/3$ of the positive candidate sample appears to be connected to galaxies, but not necessarily radio bright galaxies.
We also expect the positive candidates to have a broader line width than the negative candidates if they are more likely to be associated with galaxies. This expectation is confirmed in the right panel of Figure \[fig:match\]. The K-S test probability is 7$\times10^ {-6}$ for comparison between positive and negative candidates, and is 0.04 for comparison between the positive and negative candidates with absolute $\mHI>1.25\times10^7 \ms$.
In summary, by measuring $\mHIcand$, we add up the signal from unresolved, low [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass objects distributed in a large volume ($\sim$500 kpc) around the primary galaxies and other reliably detected objects. It is very likely that at least a fraction of these candidates are small galaxies. The remainder might be galaxies too (e.g., objects below the 20 mag cut in r band used here), or small [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}clouds in the circumgalactic medium.
{width="14cm"}
{width="5cm"} {width="5cm"} {width="5cm"}
The HI mass in candidates and the properties of primary galaxies {#sec:conformity}
----------------------------------------------------------------
In this section, we compare $\mHIcand$ between galaxies with different properties, especially between galaxies that are rich in [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}and normal galaxies. We assume that the ${\rm H}{\textsc i}$-rich (either with high $\fHI$ or [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}excess) galaxies have been accreting gas, and hope to find clues about gas accretion by studying $\mHIcand$ around the galaxies.
In Figure \[fig:HIprof\], we divide the sample of primary galaxies evenly into two sub-samples, focusing on different properties, and investigate their difference in averaged accumulative $\mHIcand$ as a function of $D_{proj}$ between 100 and 500 kpc. The dividing galactic properties include the stellar mass (M$_*$), the stellar mass surface density ($\mu_*$), optical concentration (r$_{90}/$r$_{50}$), NUV$-$r colour, the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass ($\mHI$), [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass fraction ($\fHI$), and [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}excess ($\Delta_{C13}\fHI$ and $\Delta_{4p}\fHI$) [^2].
The most prominent trend is the sub-sample with lower M$_*$ having higher $\mHIcand$ than the comparison sub-sample, with a higher than 90% significance that the two distributions are different at a $D_{proj}$ of 500 kpc. The sub-samples with higher $\fHI$ and $\Delta_{4p}\fHI$ also have considerable higher $\mHIcand$ than the comparison sub-samples; especially the latter has a close to 90% significance for the difference in distribution from the comparison sample. There are also weak trends that the sub-sample with lower $\mu_*$ and bluer NUV$-$r colour have higher $\mHIcand$ (considering the difference in average values and the close-to 0.2 K-S test probabilities). There is no significant correlation between $\mHIcand$ and the optical concentration, the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass or $\Delta_{C13}\fHI$ of primary galaxies.
We further control stellar mass to reveal a more intrinsic correlation between $\mHIcand$ and [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}richness of primary galaxies. For each of the sample dividing parameters, we derive the linear relation with stellar mass, and calculate vertical distance to the relation. We use this distance to divide the primary galaxies evenly into two samples, and study their difference in averaged accumulative $\mHIcand$. In Figure \[fig:HIprof2\], we only consider the dividing parameters which have already shown an indication for a correlation with $\mHIcand$ in Figure \[fig:HIprof\]. We can see the trend of galaxies with higher $\fHI$ having on average higher $\mHIcand$ becomes more prominent as suggested by the K-S test probabilities. The trend as a function of $\Delta_{4p}\fHI$ remains almost unchanged, suggesting the advantage of $\Delta_{4p}\fHI$ as an unbiased measure of [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}richness. Bluer primary galaxies also have on average higher $\mHIcand$. $\mHIcand$ is no longer dependent on $\mu_*$ of primary galaxies. These trends are consistent with each other, for at the same stellar mass, ${\rm H}{\textsc i}$-rich galaxies are also more star forming galaxies. Finally, we note that, when the sample is divided by $\fHI$, $\Delta_{4p}\fHI$ and NUV$-$r of the primary galaxies, the difference in the $\mHIcand$ distribution of the sub-samples is tentative, with a K-S test probability of slightly above 0.1 at a $D_{proj}$ of 300-500 kpc.
{width="3.8cm"} {width="3.8cm"} {width="3.8cm"} {width="3.8cm"}
{width="3.8cm"} {width="3.8cm"} {width="3.8cm"} {width="3.8cm"}
{width="17.cm"}
{width="3.8cm"} {width="3.8cm"} {width="3.8cm"} {width="3.8cm"}
Discussion: evidence of gas conformity {#sec:discussion}
======================================
We have found evidence for the presence of an excess of faint gas clouds in a volume out to 500 kpc in projected distance and 500 km$/$s in velocity around [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}excess galaxies, i.e., galaxies that have more [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass than other galaxies with similar optical properties. A parallel study of individual satellite galaxies around the Bluedisk primary galaxies revealed that satellites around [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}excess galaxies have higher [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}excess than satellites around galaxies with normal [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}contents (E. Wang et al. 2015). Although these two studies are based on different signal-to-noise levels of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}data, reflecting different reservoirs of cold gas, they consistently tell the same story: the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}reservoir in the environment of (massive) galaxies appears to reflect the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}content of those galaxies. As discussed in K10, this environmental [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}may trace the dense gas tunnelled into halos through filaments, the so-called “cold mode” accretion (Kere[š]{} et al. 2005).
Recent numerical studies have shown that cold mode accretion is not affected by galactic winds (Powell et al. 2011), hence it can easily leave its footprint on the low-mass satellite galaxies. Beside the gas locked into satellites, our [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}candidates may also include the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}clouds that condense from the ionised cold mode gas (Kere[š]{} et al. 2009a, 2009b). It is interesting to note that the lower limit in the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass of candidates detected in our data ($\sim10^6\ms$, imposed by our 4-$\sigma$ detection threshold) is roughly the lower mass limit required by an [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}cloud to survive in a hot gas halo around galaxies (Murray &Lin 2004). Due to the lack of baseline spacings short-ward of 36 m, we are unable to detect the very extended, smoothly distributed gas. However, the detected candidates that trace the peaks in the IGM or small companion galaxies, indicate that an underlying reservoir of gas from cold mode accretion might be present.
In the picture of cold mode accretion, the outer region of the galactic disks is the favoured location of the accretion taking place (as in the classical model of hot accretion) (Pichon et al. 2011). The accretion rate is bursty because the inflow filament is clumpy (Brooks et al. 2009). These two factors may produce the excess [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}in ${\rm H}{\textsc i}$-excess galaxies. The Bluedisk primary galaxies have stellar masses around that of the MW, for which around 50$\%$ of the gas accretion occurs in the cold mode (van de Voort & Schaye 2012). The most prominent trend found in this paper is that the primary galaxies with lower stellar masses have higher [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass in candidates around them, which may suggest a transition between cold- and hot-mode accretion, as predicted by the simulations (Dekel & Birnboim et al. 2006). Recently Kauffmann (2015) found a tendency for the satellites to align along the major axis of late-type ${\rm H}{\textsc i}$-rich central galaxies with M$_*<10^{10.5} \ms$; they found no such phenomenon for more massive central galaxies. Their result also suggests such a transition. The median M$_*$ of our primary galaxies ($10^{10.56} \ms$ ) is consistent with their suggested transition point.
An alternative picture is that the accumulative signal from [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}candidates traces a rather smooth accretion of dark galaxies. Genel et al. (2010) studied the growth of dark matter halos and found that at least $40\%$ of the baryons in the halos are accreted from very low mass halos, containing smooth cold $T\sim10^4$K gas and no stars. Although no dark galaxies have been found so far from nearby surveys (Zwaan& Briggs 2000, Zwaan 2001, Koribalski et al. 2004, Haynes 2007), this can be explained if dark galaxies are preferentially found around gas accreting massive galaxies: galaxies with very high $\fHI$ and as massive as the Bluedisk sample are rare in the nearby universe. This picture may still be consistent with the cold accretion scenario, since recent simulations show that halos in filaments may have more sub haloes than those in other environment (Guo et al. 2014).
To this date no gas accreting filaments or clouds have directly been observed to exist within a few tens of kpc around galaxies in the local universe, even when the observational depth reaches 10$^{19}$ atom cm$^{-2}$, as in the case of the WSRT Hydrogen Accretion in LOcal GAlaxies Survey (HALOGAS, Heald et al. 2011). However, numerical simulations demonstrate that the accreting gas may just get slowed down and ionised through interaction with the hot gas halo when it gets close to the galactic disks (Putman et al. 2012). In our study here, the candidates are beyond 100 kpc from the galactic centre, far above the disk-halo interface.
Finally, we address the question if the merging of satellites contributes significantly to the cold gas in primary galaxies. We select the 20 primary galaxies with higher than median $\fHI$, and only 8 of them have reliably detected satellites within a $D_{proj}$ of 500 kpc and $\Delta V_{sys}$ of 500 km$/$s. We calculate the timescale for a pair of galaxies to merge following the empirical formula from Kitzbichler & White (2008), $$T=1.6\frac{r}{25h^{-1}kpc} (\frac{M_*}{3\times10^{10}h^{-1}\ms})^{-0.3} Gyr$$ where M$_*$ is the total stellar mass of the pair. We replace M$_*$ in the formula with $\mHI$ plus M$_*$ to minimize the timescale. We calculate the gas merging rate, and compare it with the SFR for these 8 primary galaxies with satellites. The highest ratio is 0.12, hence the merging rate is too low to support the SFR, and can not be a major source for replenishing the gas storage in the Bluedisk ${\rm H}{\textsc i}$-rich galaxies. It is consistent with the findings of many previous studies like K10 and Di Teodoro &Fraternali (2014).
The gas-rich satellites studied by K10 and E. Wang et al. (2015) might be the tip of the iceberg of the missing gas accretion. Here with the detection of an excess in the low-mass low-column density [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}environment, we might have detected a substantial part of the iceberg below the sea-level.
Summary and future prospects
============================
We have developed a new technique to extract information about the large-scale distribution of [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass from 21 cm synthesis data cubes through a technique similar to stacking, which works in an [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass regime which is lower than achievable for reliable direct detections, but still high enough not to be dominated by data reduction artefacts.. We used our technique to investigate the mass contained in candidate [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}sources, possibly associated with individual satellite galaxies, around the Bluedisk primary galaxies. We found a connection between [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}excess ($\Delta_{4p}\fHI$, or high values of $\fHI$ at a fixed stellar mass) in primary galaxies and excess of [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}in low mass candidates in the surroundings. These [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}masses may be the tip of the iceberg of an underlying extended reservoir of gas that fuels the primary galaxies. It is a direct detection of the galactic gas conformity phenomenon down to very low [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}masses ($\sim10^7 \ms$). The result is consistent with the cold mode accretion in cosmological simulations. In such a picture, primary galaxies and satellites are both fuelled by the cosmic web and exhibit the conformity phenomenon of gas richness. The unusually blue outer discs in Bluedisk galaxies with high [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass fractions, indicating inside-out formation, can also be explained with such cosmological gas accretion (Kauffmann 1996).
At this moment, we are unable to establish general relations between the mass, structure and morphology of the stellar disks and the surrounding [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}masses because we are limited by selection effects in our sample of primary galaxies. Our sample misses the most massive galaxies (with M$_*>10^{11}\ms$) in the universe. In more massive halos, cold gas around primary galaxies may behave in a different way from what we find here, for those galaxies are more strongly affected by the halo (Catinella et al. 2014). Limited by primary beam effects, we are also unable to investigate how far away from the primary galaxies the conformity phenomenon extends. Limited by the small sample size, we are unable to statistically measure the total [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass in all the satellite galaxies. Because the accumulative signal ($\mHIcand$) we obtain is only $\sim$3 $\sigma$ above 0, we are unable to calculate spatial gradients, to quantify spatial distribution, or to estimate inflow rate. It also remains unclear how the accreting gas arrives at the ${\rm H}{\textsc i}$-excess galaxies, and how much of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass in candidates is not locked into optical galaxies (i.e. in the form of clumps in the more diffuse regions in and between cosmic filaments, or in dark galaxies).
We note that the emphasis of this paper is the exploration of a new technique. We demonstrate the promising potential to investigate [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}properties in the radio synthesis cubes below 5-6 $\sigma$, which is generally considered as a threshold for detection reliability. We have tried to control the systematics as best as possible and reach the tantalising result that we may have detected evidence for the presence of excess [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}in the intergalactic medium around [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}rich galaxies. We look forward to applying the technique and analysis presented in this paper to a much larger, more uniform and complete dataset, especially from the ASKAP [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}All-Sky Survey, known as WALLABY, and the proposed WSRT Northern Sky [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}Survey (WNSHS), both described in Koribalski (2012).
Acknowledgements {#acknowledgements .unnumbered}
================
We thank the anonymous referee for constructive comments. We gratefully thank T. Oosterloo, V. Kilborn, B. Sault, L. Staveley-Smith and S. Huang for useful discussions. J.M. van der Hulst acknowledges support from the European Research Council under the European Union’s Seventh Framework Programme (FP/2007-2013) / ERC Grant Agreement nr. 291531. J. Fu acknowledges the support from the National Science Foundation of China No. 11173044 and the Shanghai Committee of Science and Technology grant No. 12ZR1452700. T. Xiao acknowledges the support from NSFC under Grant No. 11203056.
GALEX (Galaxy Evolution Explorer) is a NASA Small Explorer, launched in April 2003, developed in cooperation with the Centre National d’Études Spatiales of France and the Korean Ministry of Science and Technology.
Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/.
This publication makes use of data products from the Wide-field Infrared Survey Explorer, which is a joint project of the University of California, Los Angeles, and the Jet Propulsion Laboratory/California Institute of Technology, funded by the National Aeronautics and Space Administration.
Agertz O., Teyssier R., Moore B., 2009, MNRAS, 397, L64
Abazajian K. N., et al., 2009, ApJS, 182, 543
Becker R. H., White R. L., Helfand D. J., 1995, ApJ, 450, 559
Binney J., 1977, ApJ, 215, 492
Brooks A. M., Governato F., Quinn T., Brook C. B., Wadsley J., 2009, ApJ, 694, 396
Catinella B., et al., 2013, MNRAS, 436, 34
Catinella B., et al., 2010, MNRAS, 403, 683
Condon J. J., Cotton W. D., Greisen E. W., Yin Q. F., Perley R. A., Taylor G. B., Broderick J. J., 1998, AJ, 115, 1693
Conselice C. J., Mortlock A., Bluck A. F. L., Gr[ü]{}tzbauch R., Duncan K., 2013, MNRAS, 430, 1051
de Blok W. J. G., et al., 2014, A&A, 569, AA68
Dekel A., Birnboim Y., 2006, MNRAS, 368, 2
Di Teodoro E. M., Fraternali F., 2014, A&A, 567, AA68
Elmegreen B. G., Struck C., Hunter D. A., 2014, ApJ, 796, 110
Fabello S., Catinella B., Giovanelli R., Kauffmann G., Haynes M. P., Heckman T. M., Schiminovich D., 2011, MNRAS, 411, 993
Fraternali F., Marasco A., Marinacci F., Binney J., 2013, ApJ, 764, LL21
Genel S., Bouch[é]{} N., Naab T., Sternberg A., Genzel R., 2010, ApJ, 719, 229
Giavalisco M., et al., 2011, ApJ, 743, 95
Giovanelli R., et al., 2005, AJ, 130, 2598
Guo Q., White S., Li C., Boylan-Kolchin M., 2010, MNRAS, 404, 1111
Guo Q., Tempel E., Libeskind N. I., 2014, arXiv, arXiv:1403.5563
Haynes M. P., 2007, NCimB, 122, 1109
Heald G., et al., 2011, A&A, 526, AA118
Heavens A., Panter B., Jimenez R., Dunlop J., 2004, Natur, 428, 625
Hopkins A. M., McClure-Griffiths N. M., Gaensler B. M., 2008, ApJ, 682, L13
Huang M.-L., Kauffmann G., Chen Y.-M., Moran S. M., Heckman T. M., Dav[é]{} R., Johansson J., 2013, MNRAS, 431, 2622
Huang S., et al., 2014, ApJ, 793, 40
Kacprzak G. G., Churchill C. W., Nielsen N. M., 2012, ApJ, 760, LL7
Kauffmann G., 1996, MNRAS, 281, 475
Kauffmann G., Heckman T. M., De Lucia G., Brinchmann J., Charlot S., Tremonti C., White S. D. M., Brinkmann J., 2006, MNRAS, 367, 1394
Kauffmann G., Li C., Heckman T. M., 2010, MNRAS, 409, 491
Kauffmann G., Li C., Zhang W., Weinmann S., 2013, MNRAS, 430, 1447
Kauffmann G., 2015, MNRAS, 450, 618
Kennicutt R. C., Jr., 1983, ApJ, 272, 54
Kere[š]{} D., Hernquist L., 2009, ApJ, 700, L1
Kere[š]{} D., Katz N., Dav[é]{} R., Fardal M., Weinberg D. H., 2009, MNRAS, 396, 2332
Kere[š]{} D., Katz N., Fardal M., Dav[é]{} R., Weinberg D. H., 2009, MNRAS, 395, 160
Kere[š]{} D., Katz N., Weinberg D. H., Dav[é]{} R., 2005, MNRAS, 363, 2
Kitzbichler M. G., White S. D. M., 2008, MNRAS, 391, 1489
Koribalski B. S., 2012, PASA, 29, 359
Koribalski B. S., et al., 2004, AJ, 128, 16
Larson R. B., 1972, Natur, 236, 21
Lehner N., et al., 2013, ApJ, 770, 138
Lemonias J. J., Schiminovich D., Catinella B., Heckman T. M., Moran S. M., 2014, ApJ, 790, 27
Mannucci F., Cresci G., Maiolino R., Marconi A., Gnerucci A., 2010, MNRAS, 408, 2115
Martin D. C., et al., 2005, ApJ, 619, L1
McGaugh, S. S., Schombert, J. M., Bothun, G. D., & de Blok, W. J. G. 2000, ApJL, 533, L99
Moran S. M., et al., 2012, ApJ, 745, 66
Murray S. D., Lin D. N. C., 2004, ApJ, 615, 586
Oosterloo T., Fraternali F., Sancisi R., 2007, AJ, 134, 1019
Pichon C., Pogosyan D., Kimm T., Slyz A., Devriendt J., Dubois Y., 2011, MNRAS, 418, 2493
Powell L. C., Slyz A., Devriendt J., 2011, MNRAS, 414, 3671
Putman M. E., et al., 2002, AJ, 123, 873
Putman, M. E., Peek, J. E. G., & Joung, M. R. 2012, ARA&A, 50, 491
Rauch M., Becker G. D., Haehnelt M. G., Gauthier J.-R., Ravindranath S., Sargent W. L. W., 2011, MNRAS, 418, 1115
Richter P., 2012, ApJ, 750, 165
Rees M. J., Ostriker J. P., 1977, MNRAS, 179, 541
Sancisi R., Fraternali F., Oosterloo T., van der Hulst T., 2008, A&ARv, 15, 189
Sch[ö]{}nrich R., Binney J., 2009, MNRAS, 399, 1145
Schaye J., et al., 2010, MNRAS, 402, 1536
Serra P., et al., 2015, MNRAS, 448, 1922
Serra P., et al., 2012, MNRAS, 422, 1835
Shostak G. S., Skillman E. D., 1989, A&A, 214, 33
Shull J. M., Jones J. R., Danforth C. W., Collins J. A., 2009, ApJ, 699, 754
Silk J., 1977, ApJ, 211, 638
Spitoni E., Matteucci F., Marcon-Uchida M. M., 2013, A&A, 551, AA123
van de Voort F., Schaye J., 2012, MNRAS, 423, 2991
Wang E., Wang J., Kauffmann G., J[ó]{}zsa G. I. G., Li C., 2015, MNRAS, 449, 2010
Wang J., et al., 2014, MNRAS, 441, 2159
Wang J., et al., 2013, MNRAS, 433, 270
Wang J., et al., 2011, MNRAS, 412, 1081
Wang W., White S. D. M., 2012, MNRAS, 424, 2574
Weinmann S. M., van den Bosch F. C., Yang X., Mo H. J., 2006, MNRAS, 366, 2
White S. D. M., Rees M. J., 1978, MNRAS, 183, 341
Wright E. L., et al., 2010, AJ, 140, 1868
Yates R. M., Kauffmann G., Guo Q., 2012, MNRAS, 422, 215
Zhang H.-X., Hunter D. A., Elmegreen B. G., Gao Y., Schruba A., 2012, AJ, 143, 47
Zhang W., Li C., Kauffmann G., Zou H., Catinella B., Shen S., Guo Q., Chang R., 2009, MNRAS, 397, 1243
Zwaan M. A., 2001, MNRAS, 325, 1142
Zwaan M. A., Briggs F. H., 2000, ApJ, 530, L61
$\mHIcand$ on the low and high V$_{sys}$ sides of the primary galaxies
======================================================================
In the main part of this paper, we report the detection of [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}mass ($\mHIcand$) by adding the candidates outside the galaxies detected in [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}cubes (Sect. 3.3). We find $\mHIcand$ to be higher around galaxies with high [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}excess. We have performed tests to show that $\mHIcand$ is unlikely to be from CLEANing or continuum subtraction residual artefacts (Sect. 3.4). However, as shown in the top-left panel of Figure \[fig:app1\], there is a small gradient in $\mHIcand$ between the blue and red shifted sides of the [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}cube around the primary galaxies. Consequently, part of the difference in $\mHIcand$ found between sub-samples defined by properties of primary galaxies may be related to this gradient. Indeed, the gradient is slightly stronger for the cubes with low $\Delta_{4p}\fHI$ primary galaxies than the cubes with high $\Delta_{4p}\fHI$ primary galaxies (the second and third panels in the top row of Figure \[fig:app1\]); and the difference of $\mHIcand$ around primary galaxies with low and high $\Delta_{4p}\fHI$ is less significant for the low V$_{sys}$ side of the cubes than for the high V$_{sys}$ side of the cubes.
From inspection of the position of candidates in [[H]{}[<span style="font-variant:small-caps;">i</span> ]{}]{}cubes, we find that this gradient is not clearly associated with any failures of continuum subtraction, CLEANing or RFI pattern removal. We do not find such gradients in the test with stokes Q cubes (Sect 3.4). However, we find this gradient is weakly correlated with the date of observation (Figure \[fig:app2\]) with a Pearson correlation coefficient of 0.3, mostly driven by the data points for the year 2011, where the distribution of gradients is strongly biased toward negative values. If we exclude the 12 data cubes observed in 2011, the gradients of $\mHIcand$ is much weaker for the whole sample, and for the primary galaxies with low and high $\Delta_{4p}\fHI$ as well (the first 3 columns in the bottom row of Figure \[fig:app1\]). Meanwhile, the difference in $\mHIcand$ between the two sub-samples with low and high $\Delta_{4p}\fHI$ becomes more significant for the low V$_{sys}$ side of the cubes, and is similar for the high V$_{sys}$ side of the cubes (the fourth and fifth columns in the bottom row of Figure \[fig:app1\]).
Hence we argue that while the cause for the gradient of $\mHIcand$ is not found, the effect is not likely to significantly affect our main results, Because it is unclear whether the gradient effect is an artefact or a statistic coincidence, we still use the full sample of 40 data cubes for analysis in the paper. Nevertheless, we warn the readers of this uncertainty in our experiment and intend to investigate it with larger samples in the future.
{width="3.cm"} {width="3.cm"} {width="3.cm"} {width="3cm"} {width="3.cm"}
{width="3.cm"} {width="3.cm"} {width="3.cm"} {width="3.cm"} {width="3.cm"}
![ Relation between $\mHIcand$ gradients and Modified Julian Date (MJD). The vertical dotted line marks the division of the year 2011 and 2012. The horizontal dotted line marks the position of 0. []{data-label="fig:app2"}](plot/grad_date.ps){width="7cm"}
[^1]: Email: [email protected]
[^2]: We point out the pairs of subsamples investigated here all have similar ranges of redshift (e.g. the difference in median luminosity distance is always less than 6%)
| {
"pile_set_name": "ArXiv"
} |
\[2016/12/29\]
@normal[m]{}
@bold[b]{}
\[counter\][Theorem]{} \[counter\][Lemma]{} \[counter\][Proposition]{} \[counter\][Corollary]{} \[counter\][Conjecture]{}
\[counter\][Definition]{} \[counter\][Example]{} \[counter\][Notation]{} \[counter\][Problem]{} \[counter\][System]{} \[counter\][Algorithm]{} \[counter\][Assumption]{}
\[counter\][Remark]{}
\[subcounter1\][Algorithm]{} \[subcounter2\][Theorem]{}
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Up to now, the effects of having heterogeneous networks of contacts have been studied mostly for diseases which are not persistent in time, i.e., for diseases where the infectious period can be considered very small compared to the lifetime of an individual. Moreover, all these previous results have been obtained for closed populations, where the number of individuals does not change during the whole duration of the epidemics. Here, we go one step further and analyze, both analytically and numerically, a radically different kind of diseases: those that are persistent and can last for an individual’s lifetime. To be more specific, we particularize to the case of Tuberculosis’ (TB) infection dynamics, where the infection remains latent for a period of time before showing up and spreading to other individuals. We introduce an epidemiological model for TB-like persistent infections taking into account the heterogeneity inherent to the population structure. This sort of dynamics introduces new analytical and numerical challenges that we are able to sort out. Our results show that also for persistent diseases the epidemic threshold depends on the ratio of the first two moments of the degree distribution so that it goes to zero in a class of scale-free networks when the system approaches the thermodynamic limit.'
author:
- 'J. Sanz'
- 'L. M. Floría'
- 'Y. Moreno'
title: Spreading of Persistent Infections in Heterogeneous Populations
---
Introduction
============
Disease spreading has been the subject of intense research since long time ago [@MayBook; @Cambridge; @Murray]. Our current knowledge comprises mathematical models that have allowed to better understand how an epidemic spreads and to design more efficient immunization and vaccination policies [@MayBook; @Cambridge; @Murray]. These models have gained in complexity in recent years capitalizing on data collections which have provided information on the local and global patterns of relationships in the population [@geisel; @guimera; @colizza]. In particular, with the advent of modern computational resources and tracking systems, it is now feasible to contact-trace the way the epidemic spreads or at least to predict the paths that a given pathogen might follow. In this way, some of the assumptions at the basis of the theoretical models that were difficult to test -the backbone through which the diseases are transmitted- are now more accurately incorporated into epidemiological models [@egamstw04; @cpv07; @glmp08; @cv08; @mam09].
Strikingly, the systems on top of which diseases spread show common nontrivial topological and statistical properties [@siam; @PhysRep]. A large number of networks of contacts in real-world social, biological and technological systems have been found to be best described by the so-called scale-free (SF) networks. In SF networks, the number of contacts or connections of a node with other nodes in the system, the degree (or connectivity) $k$, follows a power law distribution, $P(k) \sim k^{-\gamma}$. Recent studies have shown that the SF topology has a great impact on the dynamics and function of the system under study [@cnsw00; @ceah01; @siam; @PhysRep]. The reason is that, at variance with homogeneous or regular networks, SF architectures are a limiting case of heterogeneity where the connectivity fluctuations diverges if $2<\gamma\leq 3$ as the system size tends to infinity (the thermodynamic limit). This means that there are nodes in the network which has an eventually unbounded number of connections compared to the average degree. Examples of such networks include the Internet [@falou99; @calda00], the world-wide-web (WWW) [@www99], food-webs, and metabolic or protein networks [@strog01; @PhysRep].
In the context of disease spreading, SF networks lead to a vanishing epidemic threshold in the limit of infinite population when $\gamma \leq 3$ [@Vespignani; @MaySci; @mpv02; @newman02]. This is because the ratio $\langle k\rangle / \langle k^2\rangle$ determines the epidemic threshold above which the outbreak occurs. When $2 < \gamma \leq 3$, $\langle k\rangle$ is finite while $\langle k^2\rangle$ goes to infinity, that is, the transmission probability required for the infection to spread goes to zero. Moreover, the previous result holds both for the Susceptible-Infected-Susceptible (SIS) and Susceptible-Infected-Removed (SIR) epidemiological models [@Vespignani; @MaySci; @mpv02].
In this paper, we will deal with a different kind of diseases -those that are persistent in time and shows a latent period that can be as large as an individual’s lifetime. Our first aim is to enlarge the epidemiological framework for complex networks reported previously for the SIS model [@Vespignani] and proposed as well for the SIR model [@MaySci; @mpv02] by integrating the spreading dynamics of persistent diseases within it. With this purpose, we consider a variation of the Susceptible-Exposed-Infected-Removed model [@MB] on complex heterogeneous networks. As we will see, this kind of dynamics introduces new challenges essentially different to other successfully treated before. In particular, as the latent period is high enough, we shall work with an open system in which new individuals are born and others dead for causes not directly related to the spreading of the disease. We present novel analytical and numerical methods that allow us to obtain the epidemic threshold for this kind of disease in heterogeneous populations. Moreover, we introduce a numerical method that is well-suited to deal with the kind of problems we face. Our results point out that also for persistent infections the virtually unbounded fluctuations of the degree distribution have an important enhancing effect on the epidemic spreading.
The model
=========
To be more precise and without loss of generality, we will particularize our model to one case of persistent infection -the most threatening one which is tuberculosis (TB). TB is an old disease whose world-wide prevalence had been diminishing even before vaccination and prophylaxis strategies were firstly accomplished [@Wilson; @Styblo; @Daniel]. Its recent return in developing countries, mainly in Southeast Asia, have attracted renewed interest in it. The current world estimate of prevalence is about 33% while the number of deaths per year that it is causing reaches more than 3 million people [@Bleed]. Depending of the kind and the intensity of immune response that the host immune system performs after initial infection with M. tuberculosis bacillus, the individual can suffer latent infection, (in which the bacteria are under a growth-arrest regime and the individual neither suffer any clinical symptom nor becomes infectious) or active infection, where the host suffers clinical symptoms and can transmit the pathogen by air [@Comstock; @Styblodos]. Latently infected individuals can, generally after an immune-depression episode, reach the active phase. Estimating the probability of developing direct active infection after a contact, or alternatively, the lifetime’s risk for a latent infected individual to evolve into the active phase, are not easy tasks. However, it is generally accepted that only 5-10% of the infections directly produce active TB [@Comstock; @Styblodos], while the ranges concerning the estimation of typical “half-life” of latent state rounds about 500 years [@MB].
The spreading dynamics of TB like diseases has been studied in recent years. However, to the best of our knowledge, these works assume the homogeneous mixing hypothesis, that is, a perfectly homogeneous system in which all individuals are dynamically equivalent. As mentioned above, many of the systems on top of which diseases spread, are better described by scale-free connectivity patterns. Therefore, in what follows, our main objective will be to assemble a basic model fitted for tuberculosis spreading that firstly takes into account the heterogeneity in the distribution of the networks of contacts. We note that the increasingly alarming situation about TB epidemiology evidences the need to increase the effort in TB research in a global way. In the context of the study of its epidemiology, new models must be developed in order to gain predictive skills, incorporating the recent theoretical advances referring to disease spreading on complex heterogeneous substrates as well as meta-population approaches and new computational tools for numerical analysis and simulation. In this sense, ours is a first contribution that addresses one of the most important parameters in epidemiological description: the epidemic threshold.
Let us then introduce our model. We consider that individuals in the population are compartmentalized into three groups: healthy -$U(t)$-, infected but not infectious -*or latently infected* $L(t)$- and sick individuals $T(t)$ which are infected and are infectious as well. The transition between these subpopulations proceeds in such a way that a healthy individual acquires the bacteria through a contact with an infectious subject with probability $\lambda$. In its turn, this newly infected individual may develop the disease directly with probability $p$. However, the most common case is the establishment of a dynamic equilibrium between the bacillus and the host’s immune system, which allows the survival of the former inside the latter. When this happens, newly infected individuals become latently infected, because despite harboring the bacteria in blood, neither becomes sick nor is able to infect others.
On the other hand, after a certain period of time (which may be several years) and usually following an episode of immunosuppression, the balance between the bacterium and its host can be broken. In this case, the bacteria grow and the individual falls ill beginning to develop the clinical symptoms of the disease. In addition, if the infection attacks the lungs (pulmonary TB), the bacillus is present in the sputum, making the guest infectious.
The dynamics of the disease, in a well mixed population, is then described by the following system of nonlinear differential equations: $$\frac{dU(t)}{dt}=bN(t)-\lambda \beta U(t)\frac{T(t)}{N(t)}-\mu U(t),\nonumber$$ $$\frac{dL(t)}{dt}=(1-p)\lambda \beta U(t)\frac{T(t)}{N(t)}-(\mu+r) L(t),
\label{a}$$ $$\frac{dT(t)}{dt}=p\lambda \beta U(t)\frac{T(t)}{N(t)}+rL(t)-(\mu+\mu_{tb}) T(t),\nonumber$$ in which:
$N(t)=U(t)+L(t)+T(t)$
: represents the total population at time $t$,
$\beta$
: is the number of contacts per time unit,
$\lambda$
: is the probability that the bacteria is transmitted to a new host after a contact with an infectious subject,
$b$
: is the birth rate per capita and per unit time,
$\mu$
: is the natural death rate per capita and unit time,
$\mu_{tb}$
: is the rate of disease-related deaths per capita and unit time,
$r$
: is the transition frequency of latent infection (i.e., the probability that a latently infected individual becomes infectious),
with the closure relationship: $$\frac{dN(t)}{dt}=(b-\mu) N(t) -\mu_{tb}T(t).$$ The model above is a variation of the archetypal SIR model to which a fourth class has been added (latency class L). This kind of model has been largely treated in the literature in its well-mixed version, and it is frequently referred as SEIR model. As a first step, in this work, we will identify the removed individuals mostly with dead ones, and therefore we do not consider the possibility of natural or medical recovery (this simplification is in part justified by the large latency period of infected individuals and the constant flow of newborns into the system). A more refined model would consist of introducing such eventual recovery fluxes in the model, as well as the possibility of further relapses (the so called endogenous reactivation). These phenomenologies might be important mainly for diseases (like TB) for which the only feasible treatment in many areas consists of supplying large series of antibiotics. Thinking on the tuberculosis case, further refinements, like the inclusion of varieties of less infectious extra-pulmonary diseases [@natmed95], could also have consequences on the disease’s dynamics.
Structured Populations
======================
Dynamics
--------
The previous system of differential equations describes the dynamics of the epidemics in the well-mixed case. However, as argued above, the number of contacts of a given individual in a population can vary, which is reflected in an heterogeneous distribution of the number of contacts in the system. To account for this fact, we next consider a structured population described by a connectivity distribution $P(k)$. The system of Eqs (\[a\]) has to be modified accordingly. Assuming that all individuals with the same number of contacts, i.e., belonging to the same connectivity class $k$, are dynamically equivalent, the new system of differential equations are formulated for each degree class. Therefore, for a structured population, we have that: $$N_k(t)=P(k)N(t),$$ with: $$U_k(t)+L_k(t)+T_k(t)=N_k(t).$$ Moreover, it is convenient to express the previous equations in terms of densities, also defined within each connectivity class: $$u_k(t)=\frac{U_k(t)}{N_k(t)},\nonumber$$ $$l_k(t)=\frac{L_k(t)}{N_k(t)},$$ $$t_k(t)=\frac{T_k(t)}{N_k(t)},\nonumber$$ so that the following closure relation for any value of $k$ is verified: $$u_k(t)+l_k(t)+t_k(t)=1\quad\quad\forall (k,t).$$ On the other hand, the probability $\Theta$ that any given link points to an infectious individual is given by:
$$\Theta(t)=\frac{\sum_kkT_k(t)}{\sum_kkN_k(t)}=\frac{\sum_kkP(k)t_k(t)}{\langle k \rangle},$$
which leads to the following set of equations that describes the dynamics within each connectivity class: $$\frac{dU_k(t)}{dt}=bP(k)N-\lambda k\Theta(t) U_k(t)-\mu U_k(t),\nonumber$$ $$\frac{dL_k(t)}{dt}=(1-p) \lambda k\Theta(t) U_k(t)-(\mu+r) L_k(t),
\label{h}$$ $$\frac{dT_k(t)}{dt}=p \lambda k\Theta(t)U_k(t)+rL_k(t)-(\mu+\mu_{tb}) T_k(t).\nonumber$$ Finally, the number of individuals with connectivity $k$ evolves according to: $$\frac{dN_k(t)}{dt}=(b-\mu) N_k(t) -\mu_{tb}T_k(t)=(b-\mu-\mu_{tb}t_k)N_k(t).
\label{d}$$ At this point, and building on the previous equation, it is important to point out a feature of the model: the influence of the infection dynamics on the connectivity distribution $P(k)$. First, if we add the above equation for all $k$, we obtain that the total population evolves as: $$\frac{dN(t)}{dt}=(b-\mu) N(t) -\mu_{tb}\sum_kT_k(t)=\left(b-\mu-\mu_{tb}\sum_k{P(k)t_k}\right)N(t).
\label{e}$$ However, if we substitute $N_k(t)=P(k)N(t)$ directly into Eq. (\[d\]) and assume $P(k)$ to be constant, we would arrive to: $$P(k)\frac{dN(t)}{dt}=P(k)\left(b-\mu-\mu_{tb}t_k\right)N(t).\nonumber$$
The last expression is only compatible with Eq. (\[e\]) under the unrealistic assumption that all connectivity classes have the same proportion of sick individuals. We must therefore assume that the distribution of connectivity is also a function of time: $P(k,t)$, and therefore: $$\frac{dN_k(t)}{dt}=\frac{d[P(k,t)N(t)]}{dt}=N(t)\frac{dP(k,t)}{dt}+P(k,t)\frac{dN(t)}{dt},
\label{f}$$ so, if we substitute Eq. (\[f\]) into Eq. (\[d\]) we get: $$N(t)\frac{dP(k,t)}{dt}+P(k,t)\frac{dN(t)}{dt}=P(k)\left(b-\mu-\mu_{tb}t_k\right)N(t),\nonumber$$ expression from which, if we replace $dN(t)/dt$ from Eq. (\[e\]), we get the temporal evolution of $P(k, t)$ as: $$\frac{dP(k,t)}{dt}=-P(k,t)\mu_{tb}\left[t_k(t)-\langle t_k\rangle (t)\right],$$ where $$\langle t_k\rangle (t)=\sum_k{P(k,t)t_k(t)}.$$ Reformulating the equations in terms of densities using the definitions of the densities given above, Eqs. (\[h\]) become $$\frac{du_k(t)}{dt}=b-u_k(t)(b+\lambda k\Theta(t) -\mu_{tb}t_k(t)),\nonumber$$ $$\frac{dl_k(t)}{dt}=(1-p) \lambda k\Theta(t) u_k(t)-(b+r) l_k(t)+\mu_{tb}l_k(t)t_k(t),
\label{g}$$ $$\frac{dt_k(t)}{dt}=p \lambda k\Theta(t)u_k(t)+rl_k(t)-(b+\mu_{tb})t_k(t)+\mu_{tb}t_k(t)^2.\nonumber$$
Evolution of the degree distribution
------------------------------------
At this point it is appropriate to point out one aspect that will hinder any numerical characterization of the epidemic threshold. Our main goal will be to calculate both numerically and analytically the critical value $\lambda_c$ beyond which the population presents an endemic proportion of sicks individuals. However, we expect $\lambda_c$ to be dependent on the ratio $\frac{\langle k \rangle}{\langle k^2 \rangle}$, which is in turn a function of the connectivity distribution $P(k)$. The degree distribution, as previously shown, changes in time as the dynamics of infection progresses. As we should see, we can handle this time dependence analytically, but we should be forced to design a simulation method to account for the rate of births and deaths and the effects of these two processes on the degree distribution.
The aforementioned features might lead to a situation in which the infection dynamics would modify the underlying structure of the network through which the disease is being spread. Therefore, $\lambda_c$ could also vary as one expects it to be intrinsically related to the first two moments of a seemingly time-dependent degree distribution. The reason why we consider the distribution of contacts per unit time as heterogeneous, even for the current airborne-transmitted disease is based on the observation that the number of contacts a person can have per unit of time is subjected to two sources of heterogeneity. Firstly, what we can call *geo-demographic, macroscopic* heterogeneity, in which the number of contacts depends on the population density in the region in which an individual inhabits. Secondly, at a more *individual, microscopic* level, the heterogeneity arises because the number of contacts depends, in a region of constant population density (i.e., a town or neighborhood in a city), on the daily activity pattern of the individual within that region. These two factors define, ultimately, the function $P(k)$. Having that said, the assumption implicitly incorporated in the first equation of system (\[h\]) does not hold. Note that this equation implies that the connectivity of individuals is hereditary and therefore that the number of births within each $k$ class equals the birth rate times the number of individuals within each $k$ class, $N_k=P(k,t)N$.
The above situation would be equivalent to assume that the dynamics of the disease being studied is the only one that influences the demographic structure of a population, which is not true since it is clear that there are countless cultural, economic and social factors that ultimately define the above two levels of heterogeneity. We therefore assume in what follows that the newborns of each generation are distributed among the $k$ classes according to an invariant distribution function, which we further assume to be the initial degree distribution of the original network: $P(k,t_o)$. As we shall see, this assumption, besides being more plausible, has the advantage that makes the connectivity distribution to be roughly stable, and so will be the critical value $\lambda_c$.
So, we have the following reformulation of the system of differential equations (\[h\]): $$\frac{dU_k(t)}{dt}=bP(k,t_o)N-\lambda k\Theta(t) U_k(t)-\mu U_k(t),\nonumber$$ $$\frac{dL_k(t)}{dt}=(1-p) \lambda k\Theta(t) U_k(t)-(\mu+r) L_k(t),$$ $$\frac{dT_k(t)}{dt}=p \lambda k\Theta(t)U_k(t)+rL_k(t)-(\mu+\mu_{tb}) T_k(t),\nonumber$$ with the definition of the number of individuals in each class of connectivity: $$N_k(t)=N(t)P(k,t),$$ and inside each class: $$U_k(t)=N_k(t)u_k(t),\nonumber$$ $$L_k(t)=N_k(t)l_k(t),$$ $$T_k(t)=N_k(t)t_k(t).\nonumber$$ Now the total population within each connectivity class verifies: $$\frac{dN_k(t)}{dt}=bNP(k,t_o)-\mu NP(k,t) -\mu_{tb}T_k(t),$$ so that, if we add in $k$, the last modification has no effect on the variation of the total volume of the population. The temporal evolution of the degree distribution is now given as: $$\frac{dP(k,t)}{dt}=b\left[P(k,t_o)-P(k,t)\right]-P(k,t)\mu_{tb}\left[t_k(t)-\langle t_k\rangle (t)\right].
\label{i}$$ Finally, writing the equations in terms of the densities we get $$\frac{du_k(t)}{dt}=b\frac{P(k,t_o)}{P(k,t)}\left(1-u_k(t)\right)-u_k(t)\left(\lambda k\Theta(t) -\mu_{tb}t_k(t)\right),\nonumber$$ $$\frac{dl_k(t)}{dt}=(1-p) \lambda k\Theta(t) u_k(t)-\left(b\frac{P(k,t_o)}{P(k,t)}+r\right) l_k(t)+\nonumber$$ $$+\mu_{tb}l_k(t)t_k(t),
\label{j}$$ $$\frac{dt_k(t)}{dt}=p \lambda k\Theta(t)u_k(t)+rl_k(t)-\left(b\frac{P(k,t_o)}{P(k,t)}+\mu_{tb}\right)t_k(t)+\mu_{tb}t_k(t)^2.\nonumber$$
Characterization of the equilibrium points.
-------------------------------------------
The previous set of differential equations tells us how the different densities of interest evolves within each connectivity class. Their corresponding macroscopic quantities are defined as $$\begin{aligned}
\langle u\rangle(t)=\sum{k}{P(k,t)u_k(t)},\nonumber\\
\langle l\rangle(t)=\sum{k}{P(k,t)l_k(t)},\label{k}\\
\langle t\rangle(t)=\sum{k}{P(k,t)t_k(t)},\nonumber\end{aligned}$$ where $\langle u\rangle(t), \langle l\rangle(t)$ and $\langle t\rangle(t)$ are the mean densities of healthy, latent and sick individuals, respectively.
Let us now go one step further and characterize the equilibrium points. The magnitudes of interest are the average densities, so that an equilibrium point $(\langle u\rangle^*,\langle l\rangle^*,\langle t\rangle^*)$ must verify by definition: $$\left(\frac{d\langle u\rangle^*}{dt},\frac{d\langle l\rangle^*}{dt},\frac{d\langle t\rangle^*}{dt}\right)=(0,0,0).\nonumber$$ We also impose a further constraint which is that the degree distribution of the network is stationary, that is: $$\frac{dP(k)^*}{dt}=0\qquad\qquad\forall k.\nonumber$$ At this point one must ask whether macroscopic stability also implies stability within each connectivity class. The answer is yes, if we also demand stability of the degree distribution. Admittedly, if we equate expression (\[i\]) to zero and solve for the stationary $P(k,t)^*$ we get: $$P(k,t)^*=\frac{bP(k,t_o)}{b+\mu_{tb}(t_k^*-\langle t\rangle^*)},\nonumber$$ which shows that this value depends on the microscopic scale $t_k$. Therefore, the stability of the degree distribution imposes a stationary condition on $t_k$ for all $k$, which in its turns extends to the other densities $l_k^*$ and $u_k^*$. Hence, we have: $$\left(\frac{du_k^*}{dt},\frac{dl_k^*}{dt},\frac{dt_k^*}{dt}\right)=(0,0,0)\quad\quad \forall k.\nonumber$$ The above condition is trivially satisfied for the solution $(u_k^*,l_k^*,t_k^*)=(1,0,0)$ $\forall k$, which leads to a degree distribution exactly as the initial distribution. We next analyze the stability of this solution, which shall allow us to characterize the epidemic threshold.
Epidemic threshold
------------------
As stated before, in this section we will study the stability of the solution $(u_k^*,l_k^*,t_k^*)=(1,0,0)$ $\forall k$. At this point, as no latent or infected individuals are introduced in the network, the degree distribution does not change in time; so that $P(k)^*=P(k,t)=P(k,t_o)$. This situation allows to work with the system of differential equations given by (\[g\]) instead of working with the more general case given by the system (\[j\]).
### Case $p=1$
For simplicity and to gain some preliminary insight into the problem, we first study the case $p=1$, which means that there is no latent phase (i.e., the latent subpopulation disappears, $l_k=0$ $\forall k$). Using $u_k+t_k=1$ we get, $$\frac{du_k}{dt}=b-u_k(b+\lambda k \Theta -\mu_{tb})-\mu_{tb}t_k^2,\nonumber$$ where we have omitted temporal dependences, as we will do from now on. Looking for the stationary solution, we have that the condition $\frac{du_k}{dt}=0$ implies: $$u_k=-\left(\frac{1}{2\mu_{tb}}\right)\left(b+\lambda k \Theta -\mu_{tb}\pm\sqrt{(b+\lambda k \Theta -\mu_{tb})^2+4b\mu_{tb}}\right),\nonumber$$ from which the meaningful solution is the one with the negative sign. The previous expression is consistent with the meaning of $u^*$ since we recover the expected result $u^*=1$ when $\Theta=0$. Moreover, if we calculate the derivative with respect to $\Theta$ we get: $$\frac{du_k^*(\Theta)}{d\Theta}=\frac{\lambda k}{2\mu_{tb}}\left(-1+\frac{b+\lambda k \Theta -\mu_{tb}}{\sqrt{(b+\lambda k \Theta -\mu_{tb})^2+4b\mu_{tb}}}\right)<0,\nonumber$$ which guarantees that $u^*$ will always be less than unity and therefore is a real, valid solution. The study of the value of $\Theta$ in the steady state help us to identify the epidemic threshold. We write: $$\Theta^*=\frac{1}{\langle k\rangle }\sum_kkP(k)t_k^*=1-\frac{1}{\langle k\rangle}\sum_kkP(k)u_k^*,\nonumber$$ which, after substituting $u^*_k$ for its value, leads to:
$$\Theta^*=f(\Theta)=\frac{1}{2}-\frac{b}{2\mu_tb}+\frac{\lambda\langle k^2 \rangle}{2\mu_{tb}\langle k \rangle}\Theta-\nonumber$$
$$-\frac{1}{2\mu_{tb}\langle k \rangle}\sum_kkP(k)\sqrt{(b+\lambda k \Theta -\mu_{tb})^2+4b\mu_{tb}}.\nonumber$$
The graphical interpretation of the above equation indicates that the existence of an equilibrium point in which $\Theta^*>0$ is equivalent to the existence of a point at which $f(\Theta)$ crosses the bisector of the first quadrant. Evaluating the second derivative of $f(\Theta)$ one gets: $$\frac{d^2f(\Theta)}{d\Theta^2}=\frac{-2b\lambda^2}{\langle k \rangle}\sum_k\frac{P(k)k^3}{\left[(b+\lambda k \Theta -\mu_{tb})^2+4b\mu_{tb} \right]}<0,\nonumber$$ which ensures that the condition for the existence of such intersection is reduced to: $$\left(\frac{df(\Theta)}{d\Theta}\right)_{\Theta=0}=\frac{\lambda\langle k^2 \rangle}{\langle k \rangle}\frac{1}{b+\mu_{tb}}=1,\nonumber$$ from which the epidemic threshold is derived as: $$\lambda_c=\frac{(b+\mu_{tb})\langle k \rangle}{\langle k^2 \rangle}.\nonumber$$ Note that apart from the factor $(b+\mu_{tb})$, the previous result, formally coincides with the epidemic threshold of the SIR model.
### Case $p\neq1$
This is a somewhat more involved case. For structured populations, the resolution of the system of differential equations (\[j\]) cannot be done explicitly. We next find the epidemic threshold for the case $p\neq1$ using two approaches. On one hand, we study the time derivative of $\Theta$. On the other hand, we will also make use of the singularity of the Jacobian at the point $(u_k,l_k,t_k)=(1,0,0)$ to argue that the expression for the critical threshold is given by: $$(\lambda)_c=\frac{\langle k \rangle}{\langle k^2 \rangle}\frac{(r+b)(\mu_{tb}+b)}{pb+r}.$$
#### Time evolution of $\Theta$ in populations with a low number of sicks.
A first approach to characterize the epidemic threshold in heterogeneous networks when $p\neq1$ is to study the sign of the derivative of $\Theta$ at the onset of an epidemic outbreak. We consider an initially healthy population in which a small proportion of infectious individuals is introduced so that $t_k<<1$ $\forall k$. The derivative of $\Theta$ is: $$\left(\frac{d\Theta}{dt}\right)_{\Theta\sim0}=\frac{\sum_k{P(k)k\frac{dt_k}{dt}}}{\langle k \rangle}+\frac{\sum_k{t_kk\frac{dP(k)}{dt}}}{\langle k \rangle}-\nonumber$$ $$-\left[\frac{\sum_k{P(k)kt_k}}{\langle k \rangle}\right]\left[\frac{\sum_k{k\frac{dP(k)}{dt}}}{\langle k \rangle}\right]\nonumber$$ which, after substitution of the values of the derivatives of $P(k,t)$ and $t_k(t)$ leads to: $$\left(\frac{d\Theta}{dt}\right)_{\Theta\sim0}=\frac{\sum_k{P(k)kl_k}}{\langle k \rangle}+p\lambda\Theta\frac{\sum_k{P(k)k^2u_k}}{\langle k \rangle}-\left(b+\mu_{tb}\right)\Theta+\mu_{tb}\Theta^2.\nonumber$$ At this point we make two simplifications. The first and most easily justifiable is to neglect the term $\Theta^2$. The second is related to the presence of $l_k$ in the above equation, that we have to transform in a dependency with respect to $t_k$. Specifically, we assume to be sufficiently close to the stationary point $(u_k,l_k,t_k)=(1,0,0)$ as to be able to assume that the three derivatives vanish. In other words, and focusing our attention on latent and sick classes, we assume that: $$\left(\frac{dl_k}{dt}\right)_{\Theta\sim0}=(1-p) \lambda k\Theta u_k-(b+r) l_k+\mu_{tb}l_kt_k\simeq0,\nonumber$$ $$\left(\frac{dt_k}{dt}\right)_{\Theta\sim0}=p \lambda k\Theta u_k+rl_k-(b+\mu_{tb})t_k+\mu_{tb}t_k^2\simeq0,\nonumber$$ from which: $$l_k=\frac{\left(1-p)(b+\mu_{tb}\right)t_k-(1-p)\mu_{tb}t_k^2}{r+pb-\mu_{tb}pt_k}=\frac{\left(1-p)(b+\mu_{tb}\right)}{r+pb}t_k + O(t_k^2),\nonumber$$ which allows to express the derivative of $\Theta$ as: $$\left(\frac{d\Theta}{dt}\right)_{\Theta\sim0}=\Theta\left[\frac{r\left(1-p)(b+\mu_{tb}\right)}{r+pb}-(b+\mu_{tb})+\right .\nonumber$$ $$\left .+p\lambda\frac{\sum_k{P(k)k^2u_k}}{\langle k \rangle}\right].\nonumber$$ In the limit $u_k\simeq1$ $\forall k$ the third term within brackets is the ratio $\langle k^2 \rangle/\langle k \rangle$, from which the epidemic threshold condition may be derived as: $$\left(\frac{d\Theta}{dt}\right)_{\Theta\sim0}=\Theta\left[\frac{r\left(1-p)(b+\mu_{tb}\right)}{r+pb}-(b+\mu_{tb})+p\lambda_{c}\frac{\langle k^2 \rangle}{\langle k \rangle}\right]=0,\nonumber$$ finally leading to the expected expression for the threshold: $$\lambda_c=\frac{\langle k \rangle}{\langle k^2 \rangle}\frac{(r+b)(\mu_{tb}+b)}{pb+r}.
\label{threshold}$$
#### Analysis of the Jacobian
While for well-mixed populations the condition of singularity of the Jacobian allows to get the epidemic threshold in a straightforward way, for heterogeneous populations the analysis of the Jacobian is a difficult task because $\Theta$ is a function of each and every one of the $t_k$’s. This translates into the need of calculating a determinant whose order is three times the number of connectivity classes. What we can reasonably do is to verify is the threshold condition is verified for systems in which there are two or three different connectivity classes. In the first case in which only two different classes of connectivity exist, the Jacobian is just a quite distasteful 6x6 determinant that, after some cumbersome and lengthy algebra, can be reduced to the expression: $$J=b^2(b+r)(b+\mu_{tb})\left[(b+r)(b+\mu_{tb})+\frac{\lambda\langle k^2\rangle}{\langle k\rangle}(pb+r)\right],\nonumber$$ which equated to zero leads again to the previously obtained expression for the epidemic threshold. If we instead consider a population with three degree classes, the algebraic complexity of the problem largely increases as we now have to solve a determinant of size 9x9. However, we can proceed as before getting the following expression for the Jacobian: $$J=b^3(b+r)^2(b+\mu_{tb})^2\left[(b+r)(b+\mu_{tb})+\frac{\lambda\langle k^2\rangle}{\langle k\rangle}(pb+r)\right].\nonumber$$ This leads us to the sensible conjecture that increasing the number of connectivity classes does not add new roots to the Jacobian, but it only would increase the degeneracy of the non-interesting solutions $b=0$, $b=-r$ and $b=-\mu_{tb}$.
Numerical simulations
---------------------
When designing numerical simulations to inspect the dynamics of the system under study, we have two difficulties not previously addressed in the literature. These numerical issues with which we have to deal come from the fact that we have a system that is simultaneously open and structured. As a result of dealing with an open system, new individuals are being added to the population at a rate given by the birth rate. Additionally, these new individuals must enter the network of contacts with a predefined connectivity. While deciding how many nodes our new individuals connect to is not a problem, it certainly is to decide what are those nodes the newcomers will be linked to, as this will impact the degree distribution in a nontrivial way. This is an unavoidable numerical complication that we should face relentlessly if the analytical calculations are to be compared with Monte Carlo simulations.
To this end, we have adapted a simulation method based on transition probabilities first proposed in [@PRE] for SIR models in complex networks. The numerical approach considers all transitions between states that take place during the dynamical evolution of the subpopulations, defined by the system of differential equations (\[j\]). When dealing with structured populations, these transition rates depend, in general, on the connectivity class within which they occur. Moreover, within each $k$-class seven transitions are possible (see Fig. \[figure1\]):
-
: Birth of healthy individuals,
-
: Natural death of healthy individuals,
-
: Natural death of latently infected individuals,
-
: Natural or disease-related death of sick individuals,
-
: Transition from a healthy to the latent state,
-
: Transition from a healthy to the sick (infectious) state,
-
: Transition from a latent to the sick (infectious) state.
Each of these transitions is characterized by a characteristic transition rate $\omega_{i,k}$ that can be directly derived from the system of equations that characterizes the rate at which they occur within the class $k$ as:
-----------------------------------------------
$\omega_{1,k}=bNP(k,t_o)$
$\omega_{2,k}=\mu N P(k,t)u_k$
$\omega_{3,k}=\mu NP(k,t)l_k$
$\omega_{4,k}=(\mu+\mu_{tb})NP(k,t)t_k$
$\omega_{5,k}=(1-p)\lambda kNP(k,t)u_k\Theta$
$\omega_{6,k}=p\lambda kNP(k,t)u_k\Theta$
$\omega_{7,k}=r NP(k,t)l_k$
-----------------------------------------------
Similarly, we define the sum of all these transition rates as the average rate at which *one* transition (of any kind) occurs: $$\Omega=\sum_{i,k}{\omega_{i,k}}.$$ This average transition rate in its turn defines the characteristic or average time $\tau$ elapsed between any two consecutive transitions, the latter being defined as the inverse of $\Omega$: $$\tau=\frac{1}{\Omega}.$$ Given the previous definitions, the Monte Carlo algorithm is implemented in such a way that at each MC step (of duration $\tau$) one single transition takes place. Finally, the probability $\Pi_{i,k}$ that a given transition actually happens, is calculated as: $$\Pi_{i,k}=\frac{\omega_{i,k}}{\Omega}=\tau\omega_{i,k},$$ that determines which of all possible transitions is realized at each time step $\tau$.
We have made extensive numerical simulations of the model starting from an initial population made up of $N\simeq10^6$ individuals, whose network of contacts follows an initial degree distribution $P(k)\simeq k^{-3}$. Moreover, every newborn joins the system with a degree that verifies the same connectivity distribution. As for the values of the parameters of the dynamics, and thinking of typical values for persistent diseases, we have set the following values: $\mu=0.009 \quad \text{years}^{{-1}}$, $b=0.010 \quad \text{years}^{{-1}}$, $\mu_{Tb}=0.200 \quad \text{years}^{{-1}}$, $r=0.002 \quad \text{years}^{{-1}}$, and $p=0.070$. Demographical parameters $b$ and $\mu$ are roughly those of a country like Spain, while the parameters $p$ and $r$ are in the range of typical values for the case of tuberculosis. $\mu_{tb}$ has been chosen attending to numerical convenience (tuberculosis reaches a disease-related mortality rate that can be as large as 0.8). On the other hand, we note that a numerical criterion to define stationarity should also be adopted. In our simulations, we first let the system evolve for $4500$ years and later take averages in a window of $10^6$ Monte Carlo steps (which corresponds, roughly, with a temporal lapse of 100 years), for the mean densities of healthy, latent and sick individuals defined in (\[k\]). This is a long enough time and ensures that all of the outputs to the left of the threshold are stabilized to the state $(1,0,0)$ (this is achieved almost surely already for $t\leq4000$ years).
With the values for the parameters as specified above, the epidemic threshold is $\lambda_c=0.305$. In Fig. \[figure2\]), we have plotted the stationary proportion of sick individuals for values of the probability of transmission $\lambda$ in the interval $[0,0.5]$. As can be seen from the figure, the numerical and analytical results are in a reasonable agreement, despite the numerical challenges of simulating an open system in which the dynamics evolves on top of a complex topology. This indicates that the numerical method is accurate enough as to be used in situations where analytical predictions are not at hand.
--------- ---------------- ----------------
$N_{o}$ Analytical Numerical
$\lambda_{cA}$ $\lambda_{cN}$
1000 0.52903 0.46968
10000 0.42559 0.37595
50000 0.37620 0.33088
100000 0.35854 0.31926
--------- ---------------- ----------------
: Dependency of the epidemic thresholds on the initial size of the network. Analytical values are obtained from Eq. (\[threshold\]) using the moments of the distribution generated numerically.
\[default\]
However, it is possible to carry out numerical simulations using a variation of the algorithm above in order to improve the accuracy in the determination of the epidemic threshold $\lambda_{c}$. In this variation, instead of using a criterion for stationarity, we focus our attention in the vicinity of the critical value. More specifically, we first evaluate analytically the value for $\lambda_{c}$ and start the simulation there (obviously, if we don’t have an analytical hint, the simulation can be started at any value of $\lambda$). At each realization, we expect 6000 years for an eventual arrival of the system to the state $(\langle u \rangle,\langle l \rangle, \langle t \rangle)=(1,0,0)$. In the case that this state is not reached in that time, we assume that we are to the right of the critical point and so, we move to the left in $\lambda$ just a little quantity $\delta\lambda=0.01$. If, on the contrary, a minimum number of realizations (we used 10 in our simulations) stabilize at state $(\langle u \rangle,\langle l \rangle, \langle t \rangle)=(1,0,0)$, we assume to be to the left of the critical point and, consequently, we perform a $\delta\lambda$ switch to the right. Each time that such kind of flip-flop algorithm (a sort of bisection method) changes direction, we divide by two the value of $\delta\lambda$ until the desired precision in $\lambda_{c}$ is obtained. Using this numerical approach, we have numerically calculated the values of $\lambda_c$ for different system sizes. The results are reported in Table \[default\]. As expected from finite-size effects, the larger the size of the population, the smaller the absolute error between numerical and analytical thresholds is. Moreover, the larger the system size the smaller the epidemic threshold, which eventually should vanish in the thermodynamic limit.
Conclusions.
============
We have discussed a model for the spreading of persistent infections in complex heterogeneous populations. The framework extends the epidemiological picture proposed in previous works. Our approach is particularly suited for diseases like Tuberculosis, which shows large latency periods. The latent period results from the dynamical equilibrium that is established between the bacterium and the host’s immune system, so that the host might not become infectious during its lifetime. These particular features makes it compulsory to work with an open system where newborns are continuously introduced in the population and individuals might die due to causes different from the disease itself. By assembling a model with all these ingredients, we have shown analytically that the epidemic threshold is proportional to the ratio between the first and second moments of the degree distribution. Therefore, our results point in the same direction that those obtained for the SIS and SIR model on top of the same topologies -the virtually unbounded connectivity fluctuations play a key role in the infection dynamics enhancing the epidemic incidence and lowering the epidemic threshold.
From another point of view, we have developed a method well suited to numerically explore the dynamics of the system under study. In particular, we have been able to deal with the new challenge of having a system where the number of individuals in the population is not constant and, moreover, are connected following a given degree distribution (the well-mixed case is not challenging). Although we have applied the numerical approach to our particular system, it is worth stressing that it is general and can be applied to any problem for which transition rates between different classes and states are known. The results obtained agree well with the analytical estimates with the additional advantage that the whole phase diagram can be explored.
Finally, it is also worth mentioning that the model discussed here is probably the simplest one may devise for the spreading of persistent infections in structured populations. However, despite the recent progresses in modeling disease contagion dynamics and pandemic outbreaks, the kind of spreading phenomena analyzed here is one of the issues that have remained less explored. Our aim is to take a first step towards more realistic modeling of persistent diseases. We have left for future investigation possible extensions of the current model that take into account the influence on the dynamics of vaccination, prophylaxis and recovery rates, as well as the effects of genetic heterogeneity in the pathogen but also in the host [@MB].
We acknowledge financial support of the Government of Aragón through Grant PI038/08. This work has also been partially supported by MICINN through Grants FIS2008-01240 and FIS2009-13364-C02-01.
[99]{}
R. M. Anderson, R. M. May, & B. Anderson, *Infectious diseases of humans: Dynamics and Control* (Oxford University Press, UK, Oxford, 1992).
D. J. Daley, & J. Gani, *Epidemic Modelling* (Cambridge University Press, UK, Cambridge, 1999).
J. D. Murray, *Mathematical Biology* (Springer-Verlag, Germany, Berlin, 2002).
L. Hufnagel, D. Brockmann, & T. Geisel, *Proc. Nat. Acad. Sci. USA*, **101**, 15124-15129 (2004).
R. Guimera, S. Mossa, A. Turtschi, & L. A. N. Amaral, *Proc. Nat. Acad. Sci. USA*, **102**, 7794-7799 (2005).
V. Colizza, A. Barrat, M. Barthélemy, & A. Vespignani, *Proc. Nat. Acad. Sci. USA*, **103**, 2015-2020 (2006).
S. Eubank, H. Guclu, V. S. Anil-Kumar, M. V. Marathe, A. Srinivasan, Z. Toroczkai, & N. Wang, [*Nature*]{} [**429**]{}, 180 (2004).
V. Colizza, R. Pastor-Satorras, & A. Vespignani, [*Nat. Phys.*]{} [**3**]{}, 276 (2007).
J. G. Gardeñes, V. Latora, Y. Moreno, & E. Profumo, [*Proc. Nat. Acad. Sci. USA*]{} [**105**]{}, 1399 (2008).
V. Colizza, & A. Vespignani, [*J. Theor. Biol.*]{} [**251**]{}, 450 (2008).
S. Meloni, A. Arenas, & Y. Moreno, [*Proc. Natl. Acad. Sci. USA*]{} [**106**]{}, 16897 (2009).
M. E. J. Newman, *SIAM Rev.*, [**45**]{}, 167-256 (2003).
S. Boccaletti, V. Latora, Y. Moreno, M. Chavez, & D.-U. Hwang, *Phys. Rep.*, **424**, 175-308 (2006).
D. S. Callaway, M. E. J. Newman, S. H. Strogatz, & D. J. Watts, *Phys. Rev. Lett.*, [**85**]{}, 5468-5471 (2000).
R. Cohen, K. Erez, D. ben Avraham, & S. Havlin, *Phys. Rev. Lett.*, [**86**]{}, 3682-3685 (2001).
M. Faloutsos, P. Faloutsos, & C. Faloutsos, ACM SIGCOMM ’99, Comput. Commun. Rev. [**29**]{}, 251 (1999).
G. Caldarelli, R. Marchetti, & L. Pietronero, *Europhys. Lett.*, [**52**]{}, 386 (2000).
R. Albert, H. Jeong, & A. L. Barabási, *Nature*, [**401**]{}, 130 (1999).
S. H. Strogatz, *Nature*, [**410**]{}, 268 (2001).
R. Pastor-Satorras, & A. Vespignani, *Phys. Rev. Lett.*, **86**, 3200-3203 (2001).
A. L. LLoyd, & R. M. May, *Science*, **292**, 1316-1317 (2001).
Y. Moreno, R. Pastor-Satorras, & A. Vespignani, *Eur. Phys. J. B*, [**26**]{}, 521–529 (2002).
M. E. J. Newman, *Phys. Rev. E.*, [**66**]{}, 016128 (2002).
B. M. Murphy, B.H. Singer, S. Anderson, & S. Kirschner, *Mathematical Biosciences* **180**, 161 (2002).
L. Wilson, *J. Hist. Med. Allied Sci.* [**45**]{} 366 (1990).
K. Styblo, J. Meijer, I. Sutherland, *Bull. Int. Union Tuberc.* [**24**]{}(1), 137 (1969).
T. Daniel, J. Bates, K. Downes, History of Tuberculosis, in *Tuberculosis: Pathogenesis, protection and control*, Edited by B. R. Bloom (American Society for Microbiology, pp.13-24, USA, Washington, 1994).
D. Bleed, C. Watt, C. Dye *World Health Report 2001: Global Tuberculosis Control, Technical Report*, World Health Organization, WHO/CDS/TB/2001.287, 2001. Available from [http://www.who.int/gtb/publications/globrep01/index.html]{}.
D. Bleed, C. Watt, C. Dye, *Am. Rev. Respirat. Dis.* [**125**]{}, 8 (1982).
K. Styblo, Tuberculosis control and surveillance, in *Recent Advances in Respiratory Medicine*, Edited by D. Flenley, T. Petty (Vol. 4, pp. 77-108, UK, London, 1986).
Blower, S. M., McLean, A.R., Porco, T. C., Small, P. M., Hopewell, P. C., S‡nchez, M. A., Moss A. R., *Nature Medicine*, **1(8)**, 815-821 (1995).
Y. Moreno, J. B. Gómez, A. F. Pacheco, *Phys. Rev. E* [**68**]{} 035103(R) (2003).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We compare topological insulator materials and Rashba coupled surfaces as candidates for engineering p+ip superconductivity. Specifically, in each type of material we examine 1) the limitations to inducing superconductivity by proximity to an ordinary s-wave superconductor, and 2) the robustness of the resulting superconductivity against disorder. We find that topological insulators have strong advantages in both regards: there are no fundamental barriers to inducing superconductivity, and the induced superconductivity is immune to disorder. In contrast, for Rashba coupled quantum wires or surface states, the the achievable gap from induced superconductivity is limited unless the Rashba coupling is large. Furthermore, for small Rashba coupling the induced superconductivity is strongly susceptible to disorder. These features pose serious difficulties for realizing p+ip superconductors in semiconductor materials due to their weak spin-orbit coupling, and suggest the need to seek alternatives. Some candidate materials are discussed.'
author:
- 'Andrew C. Potter and Patrick A. Lee'
title: 'Engineering a p+ip Superconductor: Comparison of Topological Insulator and Rashba Spin-Orbit Coupled Materials'
---
Introduction
============
Superconductors with $p+ip$ pairing symmetry have long been expected to posses zero-energy Majorana bound states in vortex cores[@Ivanov] or at the ends of one-dimensional structures [@Kitaev]. These Majorana fermion bound states are expected to exhibit non-Abelian exchange statistics[@Ivanov; @Read/Green], and have been proposed as a basis for topological quantum computers which would be protected from decoherence[@Kitaev; @Nayak; @Alicea1DWires]. Consequently, there is a growing interest in realizing a robust $p+ip$ superconductors in the laboratory. Such $p+ip$ superconductors are thought to naturally occur in triplet paired fermionic superfluids (such as $^3\text{He}$ A or $\text{Sr}_2\text{RuO}_4$)[@He3Ref; @Sr2RuO4Ref], and in the Pfaffian quantum Hall state[@Moore/Read] at $\nu=5/2$. However, these systems are all experimentally delicate, and despite extensive experimental work, direct evidence of Majorana fermions remains elusive.
Recently, the possibility of engineering effective $p+ip$ superconductors in more conventional materials has arisen[@Fujimoto; @Sau; @AliceaSingle; @Potter1; @Potter2; @PALee; @Fu/Kane; @QuantumWires; @HalfMetals]. A common thread in these proposals is the use of spin-orbit coupling to convert conventional superconductivity into $p+ip$ superconductivity, typically by inducing s-wave superconductivity in a 2D material with spin-orbit coupling by proximity to an ordinary bulk superconductor. Among these proposals, two dominant classes of candidate materials have emerged: 1) surface states of topological insulator (TI) materials[@Fu/Kane] and 2) semiconducting quantum-wires or two-dimensional electron gases (2DEGs) with Rashba spin-orbit coupling $U_R$ and induced magnetization $V_z$[@Fujimoto; @Sau; @AliceaSingle]. In this work, we provide a detailed comparison of induced superconductivity in these two classes of materials and discuss the comparative advantages and disadvantages of using each of these materials to construct a $p+ip$ superconductor.
We first examine the prospects for inducing superconductivity in TI surface states and Rashba materials by the proximity effect. For TI surface states, the induced s-wave pairing is always converted into $p+ip$ pairing due to the topologically protected winding of the TI surface Bloch wave-functions. Consequently, for a sufficiently good interface between the TI surface and a bulk superconductor, it is possible to induce the full bulk pairing gap $\Delta_0$ on the TI surface[@TIInducedSC]. The situation is more complicated for Rashba materials, where a more delicate balance of spin-orbit and magnetization is required to achieve p+ip superconductivity. For these materials, the size of the induced superconducting gap is limited not only by the transparency of the interface to the bulk superconductor, but also by the magnetization and Rashba energy scales which we denote by $V_z$ and $U_R$ respectively (see Section \[sec:Rashba2DEG\] for detailed definitions). In particular for small Rashba coupling ($U_R\ll V_z,\Delta_0$), we find that the induced superconducting gap is limited to $\frac{1}{2}\sqrt{U_R\Delta_0}\ll\Delta_0$.
We then analyze the effects of disorder on the induced superconductivity. Since superconductivity in the 2D surface layer is induced by proximity rather than spontaneously developed by phonon interaction, and since the induced superconductivity has s-wave symmetry one might expect that the disorder cannot reduce the induced pairing gap. On the other hand, disorder is pair-breaking for p-wave superconductors, and the induced superconductivity is effectively converted into p+ip superconductivity. Therefore it is not a priori clear what the effect of disorder will be. By computing the disorder averaged density of states, we find that, for TI materials, the induced superconductivity is immune to disorder, and argue on general grounds that this immunity is a direct consequence of time-reversal invariance. In the Rashba 2DEG’s however, the induced magnetization required to realize a single helicity $p+ip$ superconductor inherently breaks time-reversal symmetry leaving the induced superconductivity vulnerable to disorder. We find that, while the induced superconducting gap $\Delta$ never fully closes from disorder, it can be sharply reduced from its clean value. The degree of vulnerability to disorder depends again on the size of the Rashba coupling $U_R$. For small $U_R$, $\Delta$ is strongly suppressed even for very weak disorder for which the superconducting coherence length $\xi_0$ is only a few percent of the mean-free path $\ell$. This sharp decrease is more drastic than the case of magnetic impurities in a conventional superconductor, for which superconductivity is destroyed only when $\xi_0/\ell \sim 1$.
In both regards, the TI materials offer advantages, allowing robust induced superconductivity that is immune to disorder. This suggests that TI materials may therefore be the most promising route to realizing topological superconductivity and Majorana fermions. However, implementing a $p+ip$ superconductor using a TI surface state requires many further developements in material growth and interface engineering, and therefore it may still be desirable to work with more conventional materials with strong Rashba splitting.
So far, the theoretical and preliminary experimental work on building a $p+ip$ superconductor from Rashba 2DEGs has largely focused on semiconductor materials and in particular on semiconductor nanowires[@QuantumWires]. However, in light of our analysis, the very low Rashba energy scales in semiconductors raise serious challenges for inducing superconductivity. Namely, small $U_R$ greatly limits the size of the induced superconducting gap and furthermore, renders the resulting superconductor extremely sensitive to disorder. These drawbacks suggest that an alternative class of materials with stronger spin-orbit coupling should be sought.
Extremely large Rashba splittings on the order of $1$eV have been observed in surface alloys of metals and heavy elements such as Bi on Ag(111)[@SurfaceAlloy]. In an earlier paper[@Potter1], we had proposed this surface alloy as a promising candidate material. However, this extreme Rashba strength can also create new problems. Namely, large Rashba coupling leads to large carrier density, making it difficult to adjust the chemical potential by gating. This is problematic because one must be able to fine-tune the chemical potential to achieve topological superconductivity and to manipulate Majorana end-states. It is therefore desireable to find materials with strong enough Rashba couplings to avoid problems with induced superconductivity and disorder, but not so strong that gating becomes impossible.
One particularly promising candidate is the (110) surface of Au, which first-principles calculations predict will exhibit surface bands with sizeable Rashba splitting[@AuSurface]. These surface bands naturally lie within $\lesssim 50$meV of the bulk Fermi level, indicating that it should be possible to move the chemical potential into the topological regime using gating. Furthermore, the crystal symmetry of the (110) surface allows for a combination of Rashba and Dresselhaus type spin-orbit couplings. If both types of spin orbit coupling are present, magnetization could be induced by applying an in-plane field rather than by proximity to a ferromagnetic insulator[@AliceaSingle]. This would greatly simplify the proposed setup for realizing Majoranas.
This paper is organized as follows: we begin with a review of the proposed route to engineering a $p+ip$ superconductor from TI and Rashba 2DEG materials. We then introduce a simple model of the proximity effect in these materials, and show how to choose system parameters in order optimize the induced superconducting gap. Subsequently, we turn to the issue of disorder, and derive the disorder averaged Green’s functions and density of states for weak to moderate disorder ($k_F\ell\gg 1$). Finally, we close with a discussion of the relative strengths and weaknesses of each class of materials for realizing a $p+ip$ superconductor.
Overview of Proposed Routes to Topological Superconductivity
============================================================
In this section we briefly review the proposed routes to creating a topological superconductor from the surface state of a bulk topological insulator (TI) or from a 2DEG with strong Rashba spin-orbit coupling. In both classes of materials, spin-orbit coupling creates surface bands in which electron spin is locked with respect to the direction of propagation. This helical locking of electron spin direction to propagation directions causes the electron spin to wind as one traverses a loop around the Brillouin zone. Consequently, if s-wave superconductivity is induced by proximity to a bulk superconductor, the helical winding of the surface Bloch wave-functions effectively converts the induced superonductivity into $p+ip$ superconductivity. Specifically, when re-expressed in the basis of the surface bands, the induced s-wave pairing term takes the form of a p-wave pairing term[@Fu/Kane; @AliceaSingle].
Throughout the paper, we work in the basis of time-reversed pairs: $$\begin{aligned}
\Psi_k &=& \begin{pmatrix}\psi_k\\ \mathcal{T}\psi_k \end{pmatrix} = \begin{pmatrix}\psi_k\\ -i\sigma_y\mathcal{K}\psi_k \end{pmatrix}
= \begin{pmatrix}\begin{pmatrix} c_{k,{\uparrow}} \\ c_{k,{\downarrow}} \end{pmatrix} \\ \begin{pmatrix}c^\dagger_{-k,{\downarrow}} \\ -c^\dagger_{-k,{\uparrow}}\end{pmatrix} \end{pmatrix} \label{eq:TRBasis}\end{aligned}$$ which is convenient for discussing superconductivity. Here we take the usual representation $\mathcal{T} = -i\sigma_y\mathcal{K}$ of the time-reversal operator, where $\{\sigma_{x,y,z}\}$ are Pauli matrices in the spin basis, and $\mathcal{K}$ denotes complex conjugation.
We consider inducing the pairing term: $$\begin{aligned}
H_\Delta &=& \Delta\sum_k c_{k,{\uparrow}}^\dagger c_{-k,{\downarrow}}^\dagger +h.c. \nonumber \\
&=&\sum_k \Delta\psi_k^\dagger {\left(}\mathcal{T}\psi_k{\right)}^\dagger +h.c. = \sum_k \Psi_k^\dagger \Delta\tau_1 \Psi_k \label{eq:PairingTerm}\end{aligned}$$ by proximity to an ordinary superconductor, where $\{\tau_{1,2,3}\}$ are Pauli matrices in the particle-hole basis. Here we have chosen a gauge in which the pairing order parameter $\Delta$ is purely real (this is justified, as we are not presently concerned with situations where the superconducting phase is inhomogenous or fluctuating).
In proposals to realize Majorana fermions, there are two relevant energy scales protecting the coherence of information stored among the Majorana fermions. The first is the (p-wave component of the) induced superconducting pairing gap, $\Delta$, which sets the energy scale for single particle excitations that can change the fermionic parity of a pair of Majoranas. The second is the so-called mini-gap to localized excitations near each Majorana. Since localized excitations cannot change non-local fermion parity, the mini-gap is important only when two Majoranas are brought close to each other for measurement purposes[@Akhmerov]. For Majoranas realized as end-states, the mini-gap also scales linearly with the induced p-wave pairing gap[@Potter2]. Therefore, in order to perform quantum coherent manipulations of Majorana fermions, it is important to achieve a robust pairing gap, and to work at temperatures much lower than this gap.
Topological Insulators
----------------------
The low-energy continuum Hamiltonian, $H_{\text{TI}} = \sum_k\Psi_k^\dagger \mathcal{H}_{\text{TI}}(k)\Psi_k$ of a TI surface is[@Fu/Kane]: $$\mathcal{H}_{\text{TI}} = {\left(}v\hat{z}\cdot{\left(}\sigma\times\mathbf{k}{\right)}-\mu{\right)}\tau_3 \label{eq:HTI}$$ where $v$ is the Dirac cone velocity. The surface eigenstates with energies $\pm v|k|$ form the upper and lower branches of a single Dirac cone with definite spin-helicity: $$c_{\pm} = {\left(}\cfrac{c_{k,{\uparrow}}\pm e^{-i\phi_k}c_{k,{\downarrow}}}{\sqrt{2}}{\right)}\label{eq:HelicalBasis}$$ where $\phi_k = \tan^{-1}{\left(}k_x/k_y{\right)}$.
When expressed in the $c_{\pm}$ basis, the pairing term (Eq. \[eq:PairingTerm\]) takes the form of an ideal spinless $p+ip$ superconductor[@Fu/Kane] $$H_\Delta = \sum_k \Delta{\left(}e^{\displaystyle i\phi_k}c^\dagger_{k,+}c^\dagger_{-k,+}+e^{\displaystyle-i\phi_k}c^\dagger_{k,-}c^\dagger_{-k,-}+h.c.{\right)}$$
Rashba 2DEG \[sec:Rashba2DEG\]
------------------------------
The low-energy continuum Hamiltonian, $H_{\text{R}} = \sum_k\Psi_k^\dagger \mathcal{H}_{\text{R}}(k)\Psi_k$ of a Rashba coupled 2DEG with induced magnetization is: $$\displaystyle \mathcal{H}_{\text{R}}(\mathbf{k}) = {\left[}\xi_k+\alpha_R \hat{z}\cdot{\left(}\sigma\times\mathbf{k}{\right)}{\right]}\tau_3 + V_z\sigma_z \label{eq:HR}$$ where $\xi_k = \frac{k^2}{2m}-\mu$ is the spin-independent dispersion, $\mu$ is the chemical potential, $U_R = 2m\alpha_R^2$ is the Rashba coupling strength, and $V_z$ is the induced Zeeman splitting responsible for the surface magnetization.
![(Color online) Band-structure with Rashba coupling, magnetization, and induced superconductivity. Three relevant energy scales are labelled: $2\Delta_{\text{BG}}$ is the energy gap between the ${\varepsilon}_+$ and ${\varepsilon}_-$ bands at the Fermi surface, $\Delta_{\text{FS}}$ is the induced p-wave pairing gap at the Fermi surface, and $2V_Z$ is the induced magnetization gap.[]{data-label="fig:RashbaBS"}](RashbaBS.eps){width="3in"}
In contrast to the TI case, without breaking time-reversal symmetry, there are two helicities present at each energy. Consequently, in order to construct a single species $p+ip$ superconductor, it is necessary to explicitly break time-reversal symmetry, in this case by introducing magnetization, $V_z$, to remove one of these helicities.
The Rashba coupling $\alpha_R$ creates two helical bands with energies ${\varepsilon}_{\pm}^{(R)} = \xi_k\pm \alpha_R|k|$ and spin-wavefunctions $c_{\pm}$ (as for the TI case). $V_z$ cants the helical bands by angle $\theta_M(k)$ out of the xy-plane modifying the surface eigenstates and corresponding dispersions: $$\begin{aligned}
&{\varepsilon}_{\pm}^{(\text{R/FM})} = \xi_k\pm\sqrt{V_z^2+U_R^2\frac{k^2}{2m}} \nonumber \\&c_{\pm}^{(\text{R/FM})}= e^{\displaystyle{-i\phi_k\sigma_z/2}}e^{\displaystyle -i\theta_{M}\sigma_x/2}{\left(}\cfrac{c_{k,{\uparrow}}\pm c_{k,{\downarrow}}}{\sqrt{2}}{\right)}\nonumber \\
&\theta_M(k)=\tan^{-1}{\left(}V_z/\sqrt{V_z^2+U_R^2\frac{k^2}{2m}}{\right)}\end{aligned}$$
Re-expressing $H_\Delta$ in the eigenbasis of both Rashba and Zeeman couplings, one finds that, in addition to $p\pm ip$ pairing $\Delta_{\text{p}}(k) \hat{{\mathbf{k}}}^\pm\sim {\langle}c_{k,\pm}c_{-k,\pm}{\rangle}$ between fermions both in band ${\varepsilon}_{\pm}$, the canting $\theta_M$ introduces an s-wave pairing component $\Delta_{\text{s}}(k) \sim {\langle}c_{k,+}c_{-k,-}{\rangle}$ between fermions $c_+$ and $c_-$ in bands ${\varepsilon}_{+}$ and ${\varepsilon}_{-}$ respectively where: $$\begin{pmatrix} \Delta_{\text{s}}({\mathbf{k}}) \\ \Delta_{\text{p}}(k) \end{pmatrix} = \frac{1}{2\sqrt{V_z^2+U_R^2\frac{k^2}{2m}}}\begin{pmatrix} V_z \\ -\sqrt{U_R^2\frac{k^2}{2m}} \end{pmatrix} \Delta \label{eq:DeltaSP}$$ and $\hat{{\mathbf{k}}}^\pm = {\left(}k_y\pm ik_x{\right)}/k$. As discussed in [@AliceaSingle], one has a topological superconductor with potential Majorana bound states so long as $V_z>\Delta$, and so long as $\mu$ lies within the Zeeman gap ($|\mu|<V_z$). It is most advantageous to set $\mu=0$, placing the chemical potential in the middle of the Zeeman gap (which can be done either by electrostatic gating or chemical doping), and we will take $\mu=0$ throughout the remainder of this paper.
In this system, there are two excitation gap energy-scales: the first is the pairing gap at the Fermi surface ($k=k_F$) given by: $$\begin{aligned}
&\displaystyle \Delta_{FS} = 2\Delta_p=\sqrt{\cfrac{U_R}{\Delta_{BG}}}\Delta \label{eq:DeltaFS} \\
& \Delta_{BG} = \cfrac{{\varepsilon}_+^{(\text{R/FM})}-{\varepsilon}_-^{(\text{R/FM})}}{2} = \sqrt{V_z^2+U_R\cfrac{k_F^2}{2m}} \end{aligned}$$ The second is the Zeeman gap at $k=0$, given (for $\mu=0$) by $|V_z-\Delta|$. The smaller of these two energy scales sets the bulk gap to single-particle excitations which would destroy the non-local information stored among Majorana bound-states. We note that the relative strength of the Rashba spin-orbit coupling $U_R$ and the Zeeman splitting $V_z$ determines the size of the pairing gap at $k_F$. For $V_z\gg U_R$, the pairing gap is only a small fraction of originally induced $\Delta$. If the Zeeman gap closes (i.e. if $V_z\leq \Delta$), then both helicities are present and the system is topologically trivial.
Proximity Induced Superconductivity \[sec:InducedSC\]
=====================================================
Simple Model of Proximity Effect
--------------------------------
In this section, we consider the interface between a bulk s-wave superconductor and either a topological insulator surface or a Rashba coupled surface state with induced magnetization $V_z$. As a simple model of this interface, we consider a bulk superconductor described by the BCS Hamiltonian: $$H_{B} = \sum_{k,\sigma}{\left[}{\varepsilon}_{B,k}b_{k,\sigma}^\dagger b_{k,\sigma} + {\left(}\Delta_0 b_{k,{\uparrow}}^\dagger b_{-k,{\downarrow}}^\dagger+h.c.{\right)}{\right]}$$ coupled to the surface through a clean planar interface described by the bulk–surface tunneling term: $$H_{\text{B--S}} = \displaystyle \sum_{\mathbf{k}_\parallel,k_\perp,\sigma}\Gamma b_{(\mathbf{k}_\parallel,k_\perp),\sigma}^\dagger c_{\mathbf{k}_\parallel,\sigma} +h.c.$$ which conserves momentum $\mathbf{k}_\parallel$ parallel to the interface, and is independent of the transverse momentum $k_\perp$ perpendicular to the interface. Here $b_{k,\sigma}^\dagger$ and $c_{k,\sigma}^\dagger$ are the electron creation operators (with momentum $k$ and spin $\sigma$) for the bulk superconductor and surface respectively, ${\varepsilon}_B$ is the non-superconducting bulk dispersion which we will linearize about the chemical potential $\mu$, and $\Delta_0$ is the bulk s-wave pairing amplitude.
Since surface–bulk tunneling conserves in-plane momentum, the bulk tunneling density of states (in the absence of superconductivity) is given by the one-dimensional expression $N_B({\varepsilon}_B(k))= {\left(}\partial {\varepsilon}_B(k)/\partial k_z{\right)}^{-1}$. Assuming that $N_B$ varies slowly with energy, $H_{\text{B--S}}$ induces the following self-energy correction to the surface Green’s function: $$\Sigma_\Gamma(i\omega) = \frac{\pi\gamma}{\sqrt{\Delta_0^2+\omega^2}}{\left(}-i\omega+\Delta_0\tau_1{\right)}$$ where $\gamma = N_B(0)|\Gamma|^2$ is convenient measure of the strength of surface-bulk coupling corresponding to the width of the surface resonance that would result from $H_{\text{B--S}}$ without bulk-superconductivity ($\Delta_0 = 0$).
Incorporating $\Sigma_\Gamma$ into the surface Green’s function gives: $$\mathcal{G_S}(i\omega) = \frac{Z_\Gamma}{i\omega-Z_\Gamma\mathcal{H}_{\text{TI/R}}-(1-Z_\Gamma)\Delta_0\tau_1} \label{eq:SurfaceGrnsFn}$$ where $Z_\Gamma$ is the reduced in quasi-particle weight due to the bulk–surface hybridization: $$Z_\Gamma(i\omega) = {\left(}1+\frac{\pi\gamma}{\sqrt{\Delta_0^2+\omega^2}}{\right)}^{-1} \label{eq:SurfaceZ}$$ The quasi-particle weight can be interpreted as the fraction of time that a propagating electron resides in the surface, as opposed to the bulk. The surface-bulk tunneling induces a pairing term $\tilde\Delta\tau_1$ in the surface where: $$\tilde{\Delta} = (1-Z_\Gamma)\Delta_0$$ For strong surface-bulk coupling ($\gamma\gg\Delta_0$ or equivalently $Z_\Gamma\ll 1$) a sizeable fraction of the bulk pairing is induced on the surface.
However, this is not the only effect of the interface. From Eq. \[eq:SurfaceGrnsFn\] we see that the surface-bulk coupling renormalizes the surface Hamiltonian, effectively rescaling the coefficients by a factor of $Z_\Gamma$: $$\mathcal{H}_{\text{TI/R}}\rightarrow \tilde{\mathcal{H}}_{\text{TI/R}} = Z_\Gamma\mathcal{H}_{\text{TI/R}}$$ The effects of this renormalization are markedly different for topological insulators and Rashba 2DEG’s, and we will consider each case in turn.
### Proximity Effect for TI Surface
Renormalization of the topological insulator surface due to coupling to a bulk superconductor simply results in rescaling the Fermi velocity $v_F\rightarrow \tilde v_F=Z_\Gamma v_F$ and chemical potential $\mu\rightarrow\tilde{\mu}= Z_\Gamma\mu$. However, the nature of the induced pairing is independent of $v_F$ and $|\mu|$, rather, it depends only on the helical spin-winding of the surface Bloch-wavefunctions as one traverses a loop around the single Dirac cone in the surface Brillouin zone. Since this helical winding is unchanged by the bulk-surface coupling, the induced pairing symmetry will be preserved regardless of $\gamma$.
Therefore, for TI materials there are no fundamental restrictions to pursuing arbitrarily strong coupling between the TI surface and the nearby bulk-superconductor, and consequently it is in principle possible to induce the full bulk-gap $\Delta_0$ on the TI surface. This advantageous feature of TI surfaces was first pointed out in Ref. . In the subsequent section, we will see things are not so simple for the inducing superconductivity in a Rashba coupled 2DEG.
### Proximity Effect in a Rashba Coupled 2DEG
Inducing superconductivity in Rashba coupled surface states is a more delicate matter than in TI surfaces. Whereas the helical TI surface states are topologically guaranteed to convert induced s-wave pairing into effective $p+ip$ pairing, constructing effective $p+ip$ pairing in a Rashba coupled 2DEG requires a careful balance of spin-orbit coupling to create helical winding and magnetization $V_z$ to remove one of the helicities. In particular, the Zeeman gap at $k=0$ is given by $|\tilde{V}_z-\tilde{\Delta}|$, and must not close. Consequently, in order to engineer a topological superconductor, one must arrange for the surface magnetization to exceed the induced pairing amplitude: $\tilde V_Z>\tilde\Delta$. Since $\tilde V_z = Z_\Gamma V_z$, and $\tilde\Delta = (1-Z_\Gamma)\Delta_0$, one immediately sees that a balance must be struck to ensure that the bulk–surface interface is sufficiently strong to induce pairing, but not so strong that it destroys the magnetization gap at $k=0$.
a)![Optimal parameters for producing a large pairing gap in a Rashba coupled surface via the proximity effect as a function of Rashba coupling strenght $U_R$. Panels a.–d. respectively show the optimal excitation gap $\Delta_{\text{FS}}^{(\text{opt})}$, Zeeman splitting $V_z^{(\text{opt})}$, quasi-particle residue $Z^{(\text{opt})}$, and surface-superconductor coupling energy $\gamma^{(\text{opt})}$. All energies are measured with respect to the pairing amplitude $\Delta_0$ of the bulk superconductor.[]{data-label="fig:OptimalInducedGap"}](OptimalGapvsU.eps "fig:"){width="1.5in"} b)![Optimal parameters for producing a large pairing gap in a Rashba coupled surface via the proximity effect as a function of Rashba coupling strenght $U_R$. Panels a.–d. respectively show the optimal excitation gap $\Delta_{\text{FS}}^{(\text{opt})}$, Zeeman splitting $V_z^{(\text{opt})}$, quasi-particle residue $Z^{(\text{opt})}$, and surface-superconductor coupling energy $\gamma^{(\text{opt})}$. All energies are measured with respect to the pairing amplitude $\Delta_0$ of the bulk superconductor.[]{data-label="fig:OptimalInducedGap"}](OptimalVzvsU.eps "fig:"){width="1.5in"} c)![Optimal parameters for producing a large pairing gap in a Rashba coupled surface via the proximity effect as a function of Rashba coupling strenght $U_R$. Panels a.–d. respectively show the optimal excitation gap $\Delta_{\text{FS}}^{(\text{opt})}$, Zeeman splitting $V_z^{(\text{opt})}$, quasi-particle residue $Z^{(\text{opt})}$, and surface-superconductor coupling energy $\gamma^{(\text{opt})}$. All energies are measured with respect to the pairing amplitude $\Delta_0$ of the bulk superconductor.[]{data-label="fig:OptimalInducedGap"}](OptimalZvsU.eps "fig:"){width="1.6in"} d)![Optimal parameters for producing a large pairing gap in a Rashba coupled surface via the proximity effect as a function of Rashba coupling strenght $U_R$. Panels a.–d. respectively show the optimal excitation gap $\Delta_{\text{FS}}^{(\text{opt})}$, Zeeman splitting $V_z^{(\text{opt})}$, quasi-particle residue $Z^{(\text{opt})}$, and surface-superconductor coupling energy $\gamma^{(\text{opt})}$. All energies are measured with respect to the pairing amplitude $\Delta_0$ of the bulk superconductor.[]{data-label="fig:OptimalInducedGap"}](OptimalGammavsU.eps "fig:"){width="1.5in"}
We now turn to a quantitative analysis of the effect of proximity induced pairing in a Rashba 2DEG. The goal will be to find the optimum set of parameters in order to achieve a large p-wave superconducting gap on the Rashba coupled surface. There are two excitation gaps, the magnetization gap, $|\tilde{V}_z-\tilde\Delta| = |Z_\Gamma V_z-(1-Z_\Gamma)\Delta_0|$, at $k=0$ and the p-wave pairing gap $\tilde\Delta_{FS} = (1-Z_\Gamma)\Delta_{FS}$ at the Fermi-surface. The smaller of these two energy scales sets the minimum excitation energy gap. The former decreases with $Z_\Gamma$, whereas the latter increases with $Z_\Gamma$, indicating that the optimal $Z_\Gamma$ is: $$Z^{(\text{opt})} = \frac{\Delta_{FS}+\Delta_0}{V_z+\Delta_{FS}+\Delta_0}$$ which corresponds to an optimal excitation gap of: $$\Delta_{\text{FS}}^{(\text{opt})} = \frac{V_z}{V_z+\Delta_{FS}+\Delta_0}\Delta_{FS}$$ where $\Delta_{FS}$ is a function of $U_R$, $V_z$, and $\Delta_0$ given by Eq. \[eq:DeltaFS\].
Figure \[fig:OptimalInducedGap\]a. shows the optimum achievable value of $E_{\text{gap}}$ as a function of Rashba coupling strength $U_R$, and Figure \[fig:OptimalInducedGap\]b–d. show the corresponding optimal values of $V_z$, $Z_\Gamma$, and $\gamma$. In practice, it will likely not be possible to fine-tune the interface transparency, $\gamma$, between the Rashba surface and the adjacent superconducting layer, or the induced magnetization $V_z$. Rather, these parameters will be determined by the detailed structure of the surface-superconductor and surface-magnetic insulator interfaces. Therefore, one should view Figure \[fig:OptimalInducedGap\]a. as an upper bound on the practically achievable induced pairing gap. Even so, the general trend is clear: *for small Rashba splitting, $U_R$, only a small fraction of the bulk pairing gap $\Delta_0$ is induced on the surface, whereas for large $U_R$ a substantial fraction of $\Delta_0$ is achievable*.
This analysis highlights one of the potentially serious drawbacks of using materials with weak spin-orbit coupling. In the limiting case of small $U_R$, it is advantageous to arrange $Z_\Gamma\simeq 1/2$ , large $\Delta_0\gg U_R$, and $V_z = \Delta_0$, in which case: $$\lim_{\displaystyle U_R\rightarrow 0}E_{\text{gap}} \lesssim\frac{\sqrt{U_R\Delta_0}}{2}$$ In particular for semiconductor materials in which typical Rashba couplings are typically of the order of $0.2-0.8K$[@AliceaSingle]), even after carefully optimizing $V_z$, $\gamma$, and $\Delta_0$ only a p-wave pairing gap on the order of $0.1-0.4K$ is achievable. Such small excitation gaps would require operating at temperatures much smaller than $0.1K$ in order to avoid thermal excitations, which could pose difficulties for experiments. Furthermore, as will be shown in more detail below, small $U_R$ puts stringent restrictions on sample purity, as even small amounts of disorder will further suppress induced pairing. In contrast, the situation is much more hopefull for materials with strong spin orbit couplings. For strong Rashba couplings, one induce nearly the full superconducting gap, $\Delta_0$, given sufficiently transparent superconductor-surface interfaces.
Surface Resonances
------------------
So far, we have been implicitly considering an artificial interface between a 2D material (either a TI surface or Rashba 2DEG) and a different superconducting material. A potentially simpler alternative for realizing $p+ip$ superconductivity, is to use the naturally occuring interface between a bulk-superconductor and its surface. This approach would eliminate the need to find compatible materials to engineer an appropriately transparent interface.
The formalism developed above applies equally well in this case. Namely, if electronic states on the surface of a bulk metal occur at the same energy and momentum as bulk states, then the surface states decay into the bulk leaving behind broadened resonances. If the bulk becomes a superconductor, the surface-bulk coupling induces superconductivity on the surface. Denoting the width of the surface-resonance (in the absence of bulk-superconductivity) by $\gamma$, the induced superconductivity is again described by Eqs. \[eq:SurfaceGrnsFn\] and \[eq:SurfaceZ\]. It is also possible that surface states coexist at the same energy as bulk bands, but reside in regions of the Brillouin zone for which there are no bulk-states. In this case there is no direct tunneling from the surface into the bulk, and the surface state would remain sharp state rather than broadening into a resonance. Consequently to obtain superconductivity on the surface, one would need to rely on some scattering process (e.g. phonon, electron-electron, or disorder scattering) to transfer electrons between surface and bulk states.
For natural superconducting metals with strong spin-orbit coupling (such as Pb), the electrostatic potential created by the material’s surface interupts the bulk inversion symmetry, giving rise to a surface Rashba coupling. If the surface Hamiltonian has appropriate combinations of Rashba spin-orbit coupling and magnetization (as described above), then the induced surface superconductivity will again have effective $p+ip$ pairing symmetry.
A related approach is possible for topological insulator materials, where it has been demonstrated[@TISC] that doping can produce superconductivity with transition temperatures $T_C\sim 0.15-5.5K$. Furthermore, it is common[@Bergman] that samples of materials such as $\text{Bi}_2\text{Se}_3$ that are expected to be bulk-insulators, are actually metallic. In these “topological metals", the topologically protected surface states that would appear for a bulk insulator appear instead as resonances[@Bergman]. In fact a large amount of experimental effort is currently focused on finding materials with genuinely insulating bulks in order to investigate surface electron transport. However, for the purpose of engineering a $p+ip$ superconductor, this surface-bulk coexistence is actually advantageous, and the combination of bulk superconductivity and surface-bulk coupling will result in an effective $p+ip$ superconductor at the surface of a superconducting topological metal.
To examine whether $p+ip$ superconductors built from surface resonances also exhibit Majorana bound states, for example in vortex cores or at the ends of one-dimensional magnetic domains, one can write down the $T$–matrix for scattering from a vortex or domain wall and look for poles at zero-energy. For a static vortex or domain wall configuration, the $T$–matrix at zero-energy is constructed from various products of surface Green’s functions (see Eq.\[eq:SurfaceGrnsFn\]) also at zero-energy. Since $\Sigma_\Gamma(\omega=0) = Z(\omega=0)\Delta_0\tau_1$, the surface Green’s function is identical to that of an ideal $p+ip$ superconductor with gap $\Delta = Z\Delta_0$. *Therefore surface-resonance $p+ip$ superconductors will exhibit zero-energy Majorana bound-states under exactly the same conditions as the effective $p+ip$ superconductor discussed previously*. These Majorana states are localized to the surface layer and are protected against decaying into bulk states because of the bulk superconducting gap.
Disorder
========
In this section we show that disorder effects the induced superconductivity very differently in the TI scheme as opposed to the Rashba scheme. To model disorder, we consider a random on-site potential $$H_{Dis}=\sum_{r,\sigma} V(r)c_{r,\sigma}^\dagger c_{r,\sigma}$$ that has only short-range correlations $$\overline{V(r)V(r')} = W^2\delta(r-r')$$ where $W$ is the disorder strength and $\overline{{\left(}\cdots{\right)}}$ indicates an average over disorder configurations. It is useful to parameterize disorder either by the scattering time $\tau \equiv 1/N(0)W^2$ or the mean free path $\ell = v_F\tau$ where $v_F$ is the Fermi velocity of the surface layer and $N(0)$ is the surface–density of states. Furthermore, we consider moderate disorder that is too weak to induce localization, specifically that $k_F\ell\gg 1$, but make no other assumptions on disorder strength. Since non-planar disorder scattering diagrams are sub-leading in $(k_F\ell)^{-1}$, the regime $k_F\ell\gg 1$ allows for a controlled expansion for the disorder self-energy.
The disorder averaged self-energy and Green’s function are related by the following set of coupled equations: $$\begin{aligned}
\overline{\mathcal{G}}(i\omega,k)&=& {\left[}\mathcal{G}_0(i\omega,k)^{-1}-\Sigma(i\omega){\right]}^{-1} \nonumber \\
\Sigma(i\omega) &=& W^2\tau_3\sum_k \overline{\mathcal{G}}(i\omega,k) \tau_3 \label{eq:SelfConsEqns}
\end{aligned}$$ where the bare (non-disordered Green’s function) $\mathcal{G}_0$ is given by Eq. \[eq:SurfaceGrnsFn\], and incorporates the proximity induced superconductivity.
![Panel A shows a diagrammatic representation of Eq. \[eq:SelfConsEqns\] for the disorder averaged Green’s function and self-energy respectively. Disorder scattering is represented by dashed line originating from an $\times$. For delta-function-correlated impurities only multiple scatterings from the same impurity contribute. Panel B shows an example of a crossed diagram (right) that is sub-leading in $(k_F\ell)^{-1}$ compared to the non-crossed diagram with the same number of disorder scatterings (left).[]{data-label="fig:Diagrams"}](DisorderDiagrams.eps){width="3.2in"}
Time-Reversal Symmetry
----------------------
In the subsequent discussion of disorder, the presence or absence of time-reversal (TR) symmetry plays a key role. We will presently show that when (TR) symmetry is present, induced superconductivity is immune to the presence of disorder. The proof of this principle is most conveniently conducted in the basis of time reversed pairs[@AndersonThm] (see Eq. \[eq:TRBasis\]), in which the Hamiltonian for the disordered system with time-reversal symmetric s-wave pairing induced by proximity effect can be written as the following block-matrix: $$H = \begin{pmatrix} H_0+V&\Delta\mathbb{I} \\ \Delta\mathbb{I}&-(H_0+V) \end{pmatrix} \label{TRHBdG}$$ Here $H_0$ is the Hamiltonian of the (clean) surface, $V$ is random on-site disorder, $\Delta$ is the induced pairing amplitude, $\mathbb{I}$ is the $N\times N$ identity matrix where $N$ is the number of degrees of freedom in the system. Since $\Delta\mathbb{I}$ commutes with $H_0$ and $V$, the eigenvalue problem $\det{\left(}H-{\varepsilon}{\right)}=0$ can be simplified: $$0=\det{\left(}H-{\varepsilon}{\right)}= \det{\left(}{\varepsilon}^2\mathbb{I}-(H_0+V)^2-\Delta^2{\right)}$$ Denoting the eigenvalues of $H_0+V$ by $\{\tilde{{\varepsilon}}_n\}$, the eigenvalues of $H$ are $\pm \tilde{E}_n$ where: $$\tilde{E}_n = \sqrt{\tilde{{\varepsilon}}_n^2+\Delta^2}$$ which is bounded below by $\Delta$, independent of the particular disorder configuration.
These manipulations show that, so long as the 2D surface Hamiltonian is TR invariant, disorder cannot reduce the superconducting gap. As an aside, it is useful to note that the above considerations do not depend on $V$ being spin-independent so long as it preserves TRI. In particular strong spin-orbit impurity scattering will also not reduce the superconducting gap.
Disordered Topological Insulators
---------------------------------
Since the TI Hamiltonian (Eq. \[eq:HTI\]) is TR invariant, from the previous discussion we know that the pairing gap cannot be diminished by disorder. As a simple demonstration of this general principle, one can explicitly calculate the disorder averaged Green’s function and self-energy (see Eq. \[eq:SelfConsEqns\]).
In typical experimental situations, the TI surface states are intrinsically doped away from the surface Dirac point leaving an appreciable density of states at the Fermi surface[@Bergman]. In this case, $\mu\gg\Delta$, and one finds: $$\Sigma(i\omega) = \cfrac{\tau^{-1}}{\sqrt{\Delta^2+\omega^2}}{\left(}i\omega-\Delta\tau_1{\right)}$$ where $\tau^{-1} = \pi N(0)W^2$ is a measure of the disorder strength. Incorporating this disorder self-energy into the Green’s function results in a disorder averaged Green’s function of the same form as the bare Green’s function $\mathcal{G}_0$, except with a reduced quasi-particle weight $Z$ due to disorder scattering: $$\begin{aligned}
\overline{\mathcal{G}}(i\omega) &=& \frac{Z}{i\omega-Z\mathcal{H}_{TI}-\Delta\tau_1} \nonumber \\
Z(i\omega) &=& {\left[}1+\cfrac{\tau^{-1}}{\sqrt{\Delta^2+\omega^2}}{\right]}^{-1} \hspace{.15in} (\mu\gg \Delta) \label{eq:TIDisorderGrnFunction} \end{aligned}$$ For chemical potential tuned to the Dirac point, the results are similar except that the quasi-particle weight takes a different form: $$Z(i\omega) = {\left[}1+\cfrac{W^2}{2\pi v^2}\ln{\left(}\cfrac{\Lambda^2}{\Delta^2+\omega^2}{\right)}{\right]}^{-1} \hspace{.15in} (\mu=0)$$ In either case, inspection of Eq. \[eq:TIDisorderGrnFunction\] reveals that the minimum excitation gap, $\Delta$, is unchanged by disorder scattering, in agreement with the general principles outlined above for TR invariant systems.
Disordered Rashba 2DEGs
-----------------------
In contrast to the TI case, creating a topological superconductor from a Rashba 2DEG requires explicitly breaking TR symmetry by inducing surface magnetization $V_z$. Without TR symmetry, the general arguments outlined above do not apply, and the induced pairing is vulnerable to disorder scattering. The analysis below demonstrates that the pair-breaking effects of disorder scattering are dramatically enhanced by the singular density of states at the superconducting gap edge, and furthermore that these effects are especially pronounced in systems with weak Rashba coupling. Since the pairing on the surface is induced by the bulk, it never vanishes. However, we shall see that pairing can be greatly reduced by even a small amount of disorder unless $U_R\gg V_z$. In fact, when $U_R\ll V_z$, the suppression due to disorder is more severe than for conventional superconductors with magnetic impurities for which the superconducting gap typically closes for $\xi_0/\ell \sim 1$. We will see, for $U_R\ll V_z$, that disorder strongly suppresses the induced superconductivity even for very weak disorder.
For superconductivity induced by proximity effect, the surface state wave-functions are localized to the surface, but extend into the superconductor with characteristic lengthscale $\xi_L$. Therefore electrons residing in the Rashba material are scattered not only by impurities in the Rashba material and interface roughness, but also by impurities in the superconductor. For example, even if one starts with a pristine semiconductor structure (such as a self-assembled nanowire), if superconductivity is induced by proximity to superconductor with some impurities then the mean-free path will be set by the superconductor rather than the semiconductor.
To better understand the distinction between impurities residing in the Rashba coupled surface and those in the superconductor, we consider each separately. To treat either case we find that it is sufficient to replace the bare disorder scattering time $\tau^{-1} = \pi N(0)W^2$ (where $N(0)$ is the surface density of states) by an effective disorder scattering time $${\left(}\tau^{-1}{\right)}_{\text{eff}} = \left\{\begin{array}{ll} Z_\Gamma^2\tau^{-1} &; \text{ Surface Disorder} \\ (1-Z_\Gamma)^2\tau^{-1} &; \text{ Bulk Disorder} \end{array} \right. \label{eq:EffectiveScatteringTime}$$ An explicit derivation of these expressions is given in Appendix A, but the effective scattering time can be understood more simply as follows. The fraction of the surface–resonance wave–function which lies on the surface is $Z_\Gamma$ whereas the fraction residing in the bulk superconductor is $(1-Z_\Gamma)$. Therefore, the disorder scattering matrix elements should be weighted by either $Z_\Gamma$ or $(1-Z_\Gamma)$ for surface and bulk disorder respectively. Eq. \[eq:EffectiveScatteringTime\] can then be simply understood by noting that the scattering time is proportional to the square of the disorder matrix element. By inspection, we see that effects of surface disorder are suppressed in the limit of strong surface–bulk tunneling, whereas the effects of bulk disorder are suppressed in the limit of weak tunneling. Using this effective scattering time, we now turn to the problem of solving the self-consistency relations given in Eq. \[eq:SelfConsEqns\] using the surface Green’s function in Eq. \[eq:SurfaceGrnsFn\] which includes the effects of proximity to the bulk superconductor.
For ordinary disordered superconductors, the strength of disorder is conveniently parametrized by the ratio of the coherence length $\xi_0 = \pi v_F/\Delta$ to the mean free path $\ell=v_F\tau$. For the proximity induced superconductivity these parameters are renormalized by surface–bulk coupling and also depend on the type of disorder (surface or bulk). We will see that the effects of disorder depends on disorder strength only through the ratio $\tilde{\xi_0}/\ell_{\text{eff}}$, where $\tilde{\xi_0} = \pi \tilde{v}_F/\tilde{\Delta}$ is the surface coherence length and $\ell_{\text{eff}}=\tilde{v}_F \tau_\text{eff}$ is the effective mean-free path for disorder electrons. This effective ratio can be written in terms of the intrinsic ratio of the intrinsic mean-free path, $\ell$ (un-renormalized by proximity induced superconductivity) and coherence length of the bulk superconductor, $\xi_0$, as follows: $$\cfrac{\tilde{\xi_0}}{\ell_{\text{eff}}} = \left\{\begin{array}{ll} \displaystyle \frac{Z_\Gamma^2}{1-Z_\Gamma}\frac{\xi_0}{\ell} &; \text{ Surface Disorder} \\ \displaystyle (1-Z_\Gamma)\frac{\xi_0}{\ell} &; \text{ Bulk Disorder} \end{array} \right. \label{eq:EffectiveDisorderRatio}$$ By working in terms of this effective ratio, it is possible to treat both the cases of surface–impurities and bulk–impurties on equal footing.
### Analytic Expressions for Weak Disorder
For weak disorder ($\tilde{\xi_0}/\ell_{\text{eff}}\ll 1$), it is sufficient to evaluate the self-energy to lowest order in disorder scattering strength, corresponding to the first diagram for the self-energy shown in the top line of Fig. \[fig:Diagrams\]: $$\Sigma^{(1)}(i\omega) = W^2\tau_3\sum_k \tilde{\mathcal{G}}_0(i\omega,k) \tau_3$$ Here we emphasize that the value of $W^2$ should be appropriately renormalized according to Eq. \[eq:EffectiveScatteringTime\] depending on whether the scattering considered occurs in the surface or in the bulk superconductor. For $\tilde \Delta_{FS}\ll \tilde V_z$, the dominant contributions to the $k$-integral come from near the Fermi-surface. Linearizing the Bugoliuobov dispersion about the Fermi-surface, and performing the integration yields: $$\Im m\Sigma^{(1)}(i\omega) \simeq -x\omega{\left[}\frac{1}{2}+2\cfrac{\tilde V_z}{\tilde \Delta_{BG}^2}\sigma_z{\left(}\tilde \Delta\tau_1-\tilde U_R\tau_3{\right)}{\right]}$$ $$\begin{aligned}
\Re e\Sigma^{(1)}(i\omega) &\simeq& \frac{x}{\tilde \Delta_{BG}^2}\big{[}{\left(}\omega^2-\tilde \Delta^2{\right)}\tilde U_R\tau_3
\\
&+&\tilde V_z{\left(}\tilde \Delta^2+\omega^2{\right)}\sigma_z+2\tilde U_R^2\tilde \Delta\tau_1\big{]} \end{aligned}$$ $$x = \cfrac{\pi N(0)W^2}{\sqrt{\tilde \Delta_{FS}^2+\omega^2}}\equiv \cfrac{\tau_\text{eff}^{-1}}{\sqrt{\tilde \Delta_{FS}^2+\omega^2}}$$ where $\tau_\text{eff}^{-1}$ is a measure of the disorder strength, given by Eq. \[eq:EffectiveScatteringTime\].
This self-energy alters the spectrum of the disorder averaged BdG Hamiltonian. For weak disorder, we expect the gap at the Fermi surface to change only slightly. To find the correction to $\Delta_{FS}$ due to disorder, one needs to analytically continue the self-energy to real frequency, and then look for a pole in the disorder averaged Green’s function at $\omega = \Delta_{FS}-\delta\omega$, i.e. to solve $$0 = \det{\left[}\tilde \Delta_{FS}-\tilde \delta\omega - \mathcal{H}(k_F)-\Sigma^{(1)}{\left(}\omega=\tilde \Delta_{FS}-\delta\omega{\right)}{\right]}$$ to leading order in $\delta\omega$ one finds: $$\delta\omega = \Psi_0^\dagger \Sigma^{(1)}{\left(}\omega=\tilde \Delta_{FS}-\delta\omega{\right)}\Psi_0$$ where $\Psi_0 = \begin{pmatrix}u_{\uparrow}&u_{\downarrow}&v_{\downarrow}&-v_{\uparrow}\end{pmatrix}^T$ is the eigenvector of $\mathcal{H}(k_F)$ with energy $\tilde \Delta_{FS}$.
In the limiting case where $\tilde V_z\gg \tilde U_R$, $\Psi_0\simeq \frac{1}{\sqrt{2}}\begin{pmatrix}0&1&0&1\end{pmatrix}^T$, and one finds $$\delta\omega \simeq \frac{\tilde \Delta^2}{4V_z}x \simeq \cfrac{\tilde \Delta^2 \tau_\text{eff}^{-1}}{4\tilde V_z\sqrt{2\tilde \Delta_{FS}\delta\omega}}$$ Using $\Delta_{FS}\simeq \Delta\sqrt{\cfrac{U_R}{V_z}}$, and solving for $\delta\omega$ gives the following expression for the disorder renormalized pairing gap at the Fermi-surface: $$\tilde{\Delta}_{FS}(\tau_\text{eff}^{-1}) \simeq \tilde \Delta\sqrt{\cfrac{\tilde U_R}{\tilde V_z}}{\left[}1-{\left(}\cfrac{\tilde V_z}{4\sqrt{2}\tilde U_R }\cfrac{\tilde{\xi_0}}{\ell_{\text{eff}}}{\right)}^{2/3}{\right]}\hspace{.1in}(\tilde V_z\gg \tilde U_R) \label{eq:RashbaDisVz}$$ The unusual non-analytic dependence on disorder strength stems from the singular behavior of $x$ as $\omega\rightarrow \tilde \Delta_{FS}$, which in turn reflects the Van-Hove singularity in the superconducting density of states at the gap edge. This Van-Hove singularity enhances the effective disorder strength $x$, and in particular leads to an infinite slope of $\tilde{\Delta}_{FS}(\tau_\text{eff}^{-1})$ as $\tau_\text{eff}^{-1}\rightarrow 0$.
In the opposite limit, where $\tilde U_R\gg \tilde V_z$, $\Psi_0\simeq \frac{1}{2}\begin{pmatrix}1&-1&1&-1\end{pmatrix}^T$ and consequently the weak disorder correction to the gap energy vanishes to leading order. Including sub-leading contributions in $\tilde U_R/\tilde V_z$ results in: $$\tilde \Delta_{FS}(\tau_\text{eff}^{-1})\simeq\tilde \Delta{\left[}1-{\left(}\cfrac{18}{\sqrt{2}}\cfrac{\tilde V_z^2}{\tilde U_R^2}\cfrac{\tilde{\xi_0}}{\ell_{\text{eff}}}{\right)}^{2/3}{\right]}\hspace{.15in}(\tilde U_R\gg \tilde Vz) \label{eq:RashbaDisUr}$$
ttt\]
![The excitation gap $E_{\text{gap}}$ as a function of coherence length $\tilde{\xi}_0=\pi v_F/\tilde \Delta$ to the effective mean free path $\ell_{\text{eff}} = \tilde{v}_F\tau_{\text{eff}}^{-1}$. $E_{\text{gap}}$ is obtained from numerically solving Eq. \[eq:SelfConsEqns\] for a Rashba 2DEG with induced magnetization $\tilde V_z$ and superconductivity $\tilde \Delta$. Here $\widetilde{{\left(}\cdots{\right)}}$ indicates renormalization due to the proximity effect. The effective disorder strength has a different form depending on whether disorder scattering occurs predominantly in the surface-layer or in the bulk superconductor. Both cases can be treated by choosing the appropriate expression for $\tilde{\xi}_0/\ell_{\text{eff}}$ from Eq. \[eq:EffectiveDisorderRatio\]. The parameters used in this simulation were $t=1$, $\tilde V_z = 0.1$, $\tilde \Delta = 0.01$, and various values of $\tilde U_R$. The top panel shows curves for $\tilde V_z\gg \tilde U_R$, the regime appropriate for semiconductor materials, whereas the bottom panel shows curves in the $\tilde U_R\gtrsim \tilde V_z$ regime which could be achieved by using metallic thin films with stronger spin-orbit coupling. The magnetization $\tilde V_z$ breaks time reversal symmetry rendering the induced pairing susceptible to disorder. For $\tilde V_z\gg \tilde U_R$ the gap is already strongly supressed when $\tilde{\xi}_0$ is only a few percent of $\ell_{\text{eff}}$.\[fig:NumericalDisorder\]](DisorderVz.eps){width="3.2in"}
![The excitation gap $E_{\text{gap}}$ as a function of coherence length $\tilde{\xi}_0=\pi v_F/\tilde \Delta$ to the effective mean free path $\ell_{\text{eff}} = \tilde{v}_F\tau_{\text{eff}}^{-1}$. $E_{\text{gap}}$ is obtained from numerically solving Eq. \[eq:SelfConsEqns\] for a Rashba 2DEG with induced magnetization $\tilde V_z$ and superconductivity $\tilde \Delta$. Here $\widetilde{{\left(}\cdots{\right)}}$ indicates renormalization due to the proximity effect. The effective disorder strength has a different form depending on whether disorder scattering occurs predominantly in the surface-layer or in the bulk superconductor. Both cases can be treated by choosing the appropriate expression for $\tilde{\xi}_0/\ell_{\text{eff}}$ from Eq. \[eq:EffectiveDisorderRatio\]. The parameters used in this simulation were $t=1$, $\tilde V_z = 0.1$, $\tilde \Delta = 0.01$, and various values of $\tilde U_R$. The top panel shows curves for $\tilde V_z\gg \tilde U_R$, the regime appropriate for semiconductor materials, whereas the bottom panel shows curves in the $\tilde U_R\gtrsim \tilde V_z$ regime which could be achieved by using metallic thin films with stronger spin-orbit coupling. The magnetization $\tilde V_z$ breaks time reversal symmetry rendering the induced pairing susceptible to disorder. For $\tilde V_z\gg \tilde U_R$ the gap is already strongly supressed when $\tilde{\xi}_0$ is only a few percent of $\ell_{\text{eff}}$.\[fig:NumericalDisorder\]](DisorderUr.eps){width="3.2in"}
### Numerical Solution for Moderate Disorder
For stronger disorder, Eq. \[eq:SelfConsEqns\] must be solved self-consistently, which can be done numerically. In order to regulate the numerical integrals in the UV we replace the continuum dispersion with a periodic one of the form $\xi_k = -2t\cos(k)$ which naturally introduces a finite band-width. The top and bottom panels of Figure \[fig:NumericalDisorder\] show the dependence of the induced superconducting gap on disorder strength for $\tilde U_R\gtrsim \tilde V_z$ and $\tilde V_z\gg \tilde U_R$ respectively.
For very weak disorder, $\tilde{\xi}_0\ll \ell_{\text{eff}}$, the excitation gap exhibits non-analytic infinite initial slope predicted by Equations \[eq:RashbaDisVz\] and \[eq:RashbaDisUr\]. Stronger disorder never fully closes the superconducting gap, however for $\tilde V_z\gg \tilde U_R$, the gap is largely suppressed even when $\tilde{\xi}_0$ is only a few percent of $\ell_{\text{eff}}$. In most cases, $E_{\text{gap}}$ is suppressed smoothly with increasing disorder strength, however, the $E_{\text{gap}}/\tilde \Delta$ curves for $\tilde U_R\simeq \tilde V_z$ have a knee-shaped kink at $\tilde{\xi}_0/\ell_{\text{eff}}\simeq 0.07$ after which $E_{\text{gap}}$ drops abruptly. This knee occurs when disorder reduces of the magnetization gap at $k=0$ below the pairing gap $\tilde \Delta_{FS}$ at the Fermi-surface.
These results can be readily applied both to the case where disorder scattering is due to the adjacent superconductor, and when disorder scattering occurs in the surface material, by choosing the form for $\tilde{\xi}_0/\ell_{\text{eff}}$ from Eq. \[eq:EffectiveDisorderRatio\]. It is possible to reduce the sensitivity to bulk–disorder by reducing the surface–bulk tunneling strength, $\Gamma$. However, reducing $\Gamma$ will also cause a smaller proximity induced pairing gap $\tilde{\Delta}$. Similarly, it is possible to reduce the sensitivity to surface disorder by increasing the surface–bulk tunneling rate, though, doing so will be detrimental if the surface layer is cleaner than the bulk superconductor.
Here we see a second drawback of using materials with low Rashba coupling: in addition to limiting the size of the induced pairing gap in the absence of disorder, small Rashba coupling renders the topological superconductor susceptible even to small amounts of disorder ($\tilde{\xi}_0/\ell_{\text{eff}} \ll 1$). While bulk semiconductors are typically cleaner than metallic thin films, their extreme sensitivity to disorder will likely be problematic. In particular, great care would need to be taken to limit interfacial roughness between the semiconductor and adjacent bulk superconductor and magnetic insulating film.
Before concluding, we remark on two possible extensions of this analysis. Firstly, the effects of disorder were treated for fully two-dimensional structures, whereas Majorana fermions emerge in one-dimensional (or quasi-one-dimensional) geometries. The effects of disorder in quasi-one-dimensional Rashba coupled structures were analyzed numerically in Ref. , and give similar results to those given above for two-dimensions. Finally, while this analysis has been carried out for the case of Rashba-type spin-orbit coupling, we expect similar results for systems in which both Rashba and Dresselhaus-type spin-orbit couplings are present. The relevant factor in either case is the presence of magnetization $V_z$ which breaks time-reversal symmetry and renders the induced superconductivity susceptible to disorder regardless of spin-orbit type.
Discussion and Conclusion
=========================
In conclusion, we have compared the prospects for constructing an effective $p+ip$ superconductor from TI and Rashba 2DEG based materials. We have focused on technical limitations to inducing superconductivity in these materials, and examined the effects of disorder on the induced superconductivity. In both regards, the TI materials offer natural advantages. In particular, the effective $p+ip$ nature of induced superconductivity in a TI surface is guaranteed by the intrinsic spin-helicity of the bare TI surface states. As a consequence, there are no fundamental limitations for inducing superconductivity by the proximity effect. Furthermore, since time-reversal symmetry remains intact in the TI surface, the induced superconductivity is guaranteed, based on general principles, to be immune to disorder.
While TI surface states offer certain advantages, TI materials are relatively new and many materials challenges remain. It may therefore be desireable to construct a $p+ip$ superconductor from more conventional materials with strong Rashba spin-orbit coupling. Here one needs to induce superconductivity by proximity to a conventional superconductor, and to induce magnetization for example by proximity to a ferromagnetic insulator. In this case one must strike a comparatively delicate balance of spin-orbit coupling and induced magnetization to ensure that the resulting superconductor is effectively $p+ip$[@AliceaSingle].
In this regard, materials with small Rashba coupling strengths face serious difficulties: 1) the size of the induced superconducting gap is limited by the size of the Rashba coupling and 2) for weak Rashba coupling the induced superconductivity is quite fragile and is strongly suppressed even by small amounts of disorder. Our analysis indicates that an alternative class of materials with stronger spin-orbit coupling should be sought. For example, metallic thin films with heavy atomic elements can give orders of magnitude larger spin-orbit couplings than the semiconductor materials that have so far dominated the theoretical discussion.
Constructing a $p+ip$ superconductor from a Rashba 2DEG in this way requires engineering a complicated set of material interfaces between the Rashba 2DEG, superconductor, ferromagnetic insulator, and gate electrodes. However, a fortuitous choice of material could obviate the need for the superconducting and ferromagnetic interfaces. For example, one could try to find bulk metals with strong bulk spin-orbit coupling that naturally superconduct and possess surface resonances. In such a scenario, the coupling between surface-and bulk will then automatically induce superconductivity on the surface, eliminating the need to build an artificial interface for this purpose. A further simplification is possible the metallic surface which has the appropriate symmetry such that both Rashba and Dresselhaus type spin-orbit coupling is present[@AliceaSingle]. In this case, one could induce magnetization by applying an external field rather than by depositing a ferromagnetic insulator[@AliceaSingle]. If a naturally superconducting material can be found with the appropriate spin-orbit coupled surface resonances, this approach might offer the simplest route to an artificial $p+ip$ superconductor and Majorana fermions.
*Acknowledgements -* We thank T. Oguchi for helpful discussion and for sharing data from first-principle calculations of the Au(110) surface band-structure. This work was supported by DOE Grant No. DE–FG02–03ER46076 (PAL) and NSF IGERT Grant No. DGE-0801525 (ACP). Shortly after this work was submitted, another paper appeared that also analyzes the effects of disorder on superconductivity induced in one-dimensional Rashba coupled nanowires[@OtherDisorderPaper].
Surface vs. Bulk Disorder
=========================
In the main text, we argued on conceptual grounds, that the above analysis for disorder is easily modified to separately treat the two distinct cases where disorder scattering occurs predominantly in the surface layer or the adjacent bulk superconductor, by replacing the bare disorder scattering time by an effective scattering time weighted by $Z_\Gamma^2$ or $(1-Z_\Gamma)^2$ respectively. Here we consider each case separately and provide an explicit demonstration of this claim.
Surface Disorder
----------------
Consider first the case of a pristine superconductor so that impurity scattering occurs only in the surface layer. In this case the disorder scattering matrix elements are proportional to the fraction of the electron wave-functions that resides on the surface. More precisely, we can calculate self-energy from disorder scattering in Eq. \[eq:SelfConsEqns\] using the surface Green’s function in Eq. \[eq:SurfaceGrnsFn\]. Equivalently, in Fig. \[fig:Diagrams\], one should take single solid lines to be the surface Green’s function in Eq. \[eq:SurfaceGrnsFn\]. By inspection, we see that the effective disorder matrix elements are reduced by a factor of $Z_\Gamma<1$. Therefore, the disorder self-energy for surface-disorder is suppressed by a factor of $Z_\Gamma^2$.
The analysis in the main text can be easily modified to treat the case where disorder occurs predominantly in the surface layer by replacing $\tau_1$ for the surface (without surface–bulk tunneling) by the effective ratio: $$\tau_{\text{eff}}^{-1} \equiv Z_\Gamma^2\tau^{-1}$$ and keeping the same ratio $U_R/V_z$ (since this ratio is unaffected by the surface–bulk coupling).
Bulk Disorder
-------------
When the dominant source of disorder is in the adjacent bulk–superconductor, the surface electrons must first tunnel into the bulk in order to scatter from the disorder potential. This leads to an renormalized effective disorder strength which is different than the bulk value. To demonstrate this, we focus on the limit where surface–bulk tunneling is much stronger than bulk disorder ($\gamma\ll W$), in which case there is typically no more than one disorder scattering event per surface–bulk tunneling event.
In this limit, it suffices to compute Eq. \[eq:SelfConsEqns\] using the surface-Green’s function in Eq. \[eq:SurfaceGrnsFn\], with an effective disorder vertex is given by the diagram shown in Fig. \[fig:DressedDisorderVertex\]. In this figure the circled $\Gamma$ indicates surface–bulk tunneling and the dashed line indicates scattering from disorder. Written in terms of the bulk Green’s functions the effective disorder vertex for surface states is: $$\begin{aligned}
&&\tilde{V}{\left(}i\omega,\mathbf{k}_\parallel,\mathbf{Q}{\right)}\tau_3 \nonumber \\ && = V|\Gamma|^2\sum_{ k_z} \cfrac{1}{i\omega-\xi_{(k_z,\mathbf{k}_\parallel)}\tau_3-\Delta\tau_1}\tau_3\cfrac{1}{i\omega-\xi_{\mathbf{k}+\mathbf{Q}}\tau_3-\Delta\tau_1}\nonumber\\ \end{aligned}$$ We are interested primarily in low-frequency behavior for which the external momenta $\{\mathbf{k}_\parallel, \mathbf{k}_\parallel + \mathbf{Q}_\parallel\}$ lie near the surface Fermi-level. For a given $\mathbf{Q}_\parallel$ connecting two points on the surface Fermi-surface, there are two values of $Q_z$ for which $\mathbf{k}$ and $\mathbf{k}+\mathbf{Q}$ also lie on the surface. Since the dominant contributions are for momenta lying near the Fermi-surface, we linearize the bulk-dispersion in the z-direction, and denote the tunneling density of states by $N_B(0)\equiv {\left(}\cfrac{\partial {\varepsilon}_B(\mathbf{k})}{\partial{k_z}}{\right)}^{-1}$. With these approximations, the effective disorder potential for surface electrons is independent of momentum transfer $\mathbf{Q}_\parallel$: $$\tilde{V}(i\omega)\simeq \cfrac{\pi N_B(0)|\Gamma|^2}{\sqrt{\Delta^2+\omega^2}} \label{eq:EffectiveDisorderVertex}$$
The disorder self-energy given by the diagram in Fig. \[fig:Diagrams\]a. can then be computed using the surface Green’s function from Eq.\[eq:SelfConsEqns\] and the effective disorder strength in Eq. \[eq:EffectiveDisorderVertex\]. When this self-energy is incorporated into the surface Green’s function, it comes with a factor of ${\left(}Z_\Gamma{\right)}^2$. Since we are interested in the behavior for $\omega\ll \Delta$, we may neglect the frequency dependence of various quantities, and we find that the effective disorder scattering time for surface-electrons is weighted (compared to the bulk quantity) by a factor: $$\begin{aligned}
{\left(}\tau^{-1}{\right)}_{\text{eff}} &=& {\left(}\cfrac{\pi N_B(0)|\Gamma|^2}{\Delta+\pi N_B(0)|\Gamma|^2}{\right)}^2 {\left(}\tau^{-1}{\right)}_{\text{bulk}} \nonumber \\
&=&(1-Z_\Gamma)^2{\left(}\tau^{-1}{\right)}_{\text{bulk}} \end{aligned}$$ For a transparent surface–bulk interface ($\pi N_B(0)|\Gamma|^2\gg \Delta$), this factor approaches unity, justifying the claim made in the main text.
![Effective disorder vertex for surface–electrons in the case where disorder scattering occurs predominantly in the adjacent bulk–superconductor. The solid lines are bulk Green’s functions, circled $\Gamma$s indicate surface–bulk tunneling events, and the dashed line ending in $\times$ indicates disorder scattering that transfers momentum $\mathbf{Q}$[]{data-label="fig:DressedDisorderVertex"}](DressedBulkDisorderVertex.eps){width="1.5in"}
N. Read and D. Green, Phys. Rev. B [**61**]{}, 10267 (2000). D.A. Ivanov, Phys. Rev. Lett. [**86**]{}, 268 (2001). C. Nayak, S.H. Simon, A. Stern, M. Freedman, and S. Das Sarma, Rev. Mod. Phys. [**80**]{}, 1083 (2008). A. Kitaev, arXiv:cond-mat/0010440 (2000). J. Alicea, Y. Oreg, G. Refael, F. von Oppen, M.P.A. Fisher Nature Physics, doi:10.1038/nphys1915 (2011) Lee, D. M. Rev. Mod. Phys. 69, 645–665 (1997). G. M. Luke et al. Nature [**394**]{}, 558-561 (1998); K. Ishida et al. Nature [**396**]{}, 658-660 (1998) G. Moore and N. Read, Nuc. Phys. B [**360**]{}, 362 (1991). S. Fujimoto, Phys. Rev. B [**77**]{}, 220501 (2008). J.D. Sau, R.M. Lutchyn, S. Tewari and S. Das Sarma, Phys. Rev. Lett. [**104**]{}, 040502 (2010). R. M. Lutchyn, J. D. Sau, and S. Das Sarma Phys. Rev. Lett. 105, 077001 (2010); Y. Oreg, G. Refael, and F. von Oppen Phys. Rev. Lett. 105, 177002 (2010) J. Alicea, Phys. Rev. B [**81**]{}, 125318 (2010). A.C. Potter and P.A. Lee, Phys. Rev. Lett. 105, 227003 (2010) A.C. Potter and P.A. Lee, Phys. Rev. B 83, 094525 (2011) P.A. Lee, arXiv: 0907.2681. L. Fu and C.L. Kane, Phys. Rev. Lett. [**100**]{}, 096407 (2008). M. Duckheim and P.W. Brouwer, Phys. Rev. B 83, 054513 (2011); S. B. Chung, H.-J. Zhang, X.-L. Qi, and S.-C. Zhang, arXiv:1011.6422v2 J.D. Sau, R.M Lutchyn, S. Tewari, and S. Das Sarma, Phys. Rev. B 82, 094522 (2010) C. R. Ast et al., Phys. Rev. B 75, 201401(R) (2007). Nagano M, Kodama A, Shishidou T and Oguchi T, J. Phys.: Condens. Matter 21, 064239 (2009) A.R. Akhmerov, Phys. Rev. B 82, 020509(R) (2010) Y.S. Hor, A. J. Williams, J.G. Checkelsky, P. Roushan, J. Seo, Q. Xu, H.W. Zandbergen, A. Yazdani, N.P. Ong, and R. J. Cava, Phys. Rev. Lett. 104, 057001 (2010); Y. S. Hor, J. G. Checkelsky, D. Qu, N. P. Ong, R. J. Cava, arXiv:1006.0317v1 (2010); Y. Chen, Z. Liu, J. G. Analytis, J.H. Chu, H. Zhang, S.K. Mo, R.G. Moore, D. Lu, I. Fisher, S.C. Zhang, Z. Hussain, Z.X. Shen arXiv:1006.3843v1 (2010). D.L. Bergman and G. Refael Phys. Rev. B [**82**]{}, 195417 (2010) P. W. Anderson, J. Phys. Chem. Solids [**11**]{}, 26 (1959) P. W. Brouwer, M. Duckheim, A. Romito, and F. von Oppen, arXiv:1103.2746v1 (2011)
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
We generalize the classical Post correspondence problem ($\PCP_n$) and its non-homogeneous variation ($\GPCP_n$) to non-commutative groups and study the computational complexity of these new problems. We observe that $\PCP_n$ is closely related to the equalizer problem in groups, while $\GPCP_n$ is connected to the double twisted conjugacy problem for endomorphisms. Furthermore, it is shown that one of the strongest forms of the word problem in a group $G$ (we call it the [*hereditary word problem*]{}) can be reduced to $\GPCP_n$ in $G$ in polynomial time.
The main results are that $\PCP_n$ is decidable in a finitely generated nilpotent group in polynomial time, while $\GPCP_n$ is undecidable in any group containing free non-abelian subgroup (though the argument is very different from the classical case of free semigroups). We show that the double endomorphism twisted conjugacy problem is undecidable in free groups of sufficiently large finite rank. We also consider the bounded $\PCP$ and observe that it is in $\NP$ for any group with $\P$-time decidable word problem, meanwhile it is $\NP$-hard in any group containing free non-abelian subgroup. In particular, the bounded $\PCP$ is $\NP$-complete in non-elementary hyperbolic groups and non-abelian right angle Artin groups.
[**Keywords.**]{} Post correspondence problem, nilpotent groups, solvable groups, hyperbolic groups, linear groups, right angle Artin groups, twisted conjugacy problem.
[**2010 Mathematics Subject Classification.**]{} 03D15, 20F65, 20F10.
address:
- 'Department of Mathematics, Stevens Institute of Technology, Hoboken, NJ, 07030 USA'
- 'Department of Mathematics, Stevens Institute of Technology, Hoboken, NJ, 07030 USA'
- 'Department of Mathematics, Stevens Institute of Technology, Hoboken, NJ, 07030 USA'
author:
- Alexei Myasnikov
- Andrey Nikolaev
- Alexander Ushakov
bibliography:
- 'main\_bibliography.bib'
title: The Post Correspondence Problem in groups
---
Introduction
============
Motivation
----------
In this paper, following [@MNU1] we continue our research on non-commutative discrete (combinatorial) optimization. Namely, we define the Post correspondence problem ($\PCP$) for an arbitrary algebraic structure and then study this problem together with its variations for an arbitrary group $G$. The purpose of this research is threefold. Firstly, we approach $\PCP$ in a very different context, facilitating a deeper understanding of the nature of $\PCP$ problems in general. Secondly, we try to tackle several interesting algorithmic problems in group theory that are related to $\PCP$, whose time complexity is unknown. Thirdly, we hope to unify several algebraic techniques through the framework of $\PCP$ problems. We refer to [@MNU1] for the initial motivation, the set-up of the problems, and initial facts on non-commutative discrete optimization.
We would like to thank E. Ventura for his valuable remarks.
The Post correspondence problem in algebra
------------------------------------------
Let $\CA$ be an arbitrary algebraic structure in a language $L$ (for example, a semigroup, a group, or a ring). The Post correspondence problem for $\CA$ (abbreviated as $\PCP(\CA)$) asks to decide when given two tuples of equal length $u = (u_1, \ldots,u_n)$ and $v = (v_1, \ldots,v_n)$ of elements of $\CA$ if there is a term $t(x_1, \ldots,x_n)$ in the language $L$ such that $t(u_1, \ldots,u_n) = t(v_1, \ldots,v_n)$ in $\CA$. In 1946 Post introduced this problem in the case of free monoids (free semigroups) and proved that it is undecidable [@Post:1946]. Since then $\PCP$ took its prominent place in the theory of algorithms and theoretical computer science.
There are some interesting variations of this problem especially in the case of semigroups and groups, which we discuss in detail in Section \[se:PCP\]. Here we mention only one, designed specifically for (semi)groups, to which we refer as a general or a non-homogeneous Post correspondence problem ($\GPCP$). In this case the terms $t$ are just words in a fixed alphabet $X$ (or $X \cup X^{-1}$ in the case of groups), and the problem is to decide when given two tuples $u$ and $v$ of elements in a (semi)group $S$ as above and two extra elements $a, b \in S$ if there is a term $t(x_1, \ldots,x_n)$ such that $at(u_1, \ldots,u_n) = b t(v_1, \ldots,v_n)$ in $S$.
Above we described a [*decision*]{} version of $\PCP$ and $\GPCP$ in a semigroup (or a group) which requires to check if there exists a term $w$, called a [*solution*]{}, for a given instance of the problem. The [*search*]{} variation of the problem asks to find a solution (if it exists) for a given instance. Even more interesting problem is to describe all solutions to the given instance of the problem. We will have more to say about this in due course.
Algebraic meaning of $\PCP$ and $\GPCP$ in groups
-------------------------------------------------
Some connections between Post correspondence problems and classical questions in groups are known. We mention some of them here and refer to Sections \[se:connections\] and \[se:GPCP\] for details.
The standard (homogeneous) $\PCP$ in groups is closely related to the problem of finding the equalizer $E(\phi,\psi)$ of two group homomorphisms $\phi,\psi : H \to G$. This equalizer is defined as $E(\phi,\psi) = \{w \in H \mid \phi(w) = \psi(w)\}$. In particular, $\PCP$ in a group $G$ is the same as to decide if the equalizer of a given pair of homomorphisms $\phi, \psi \in Hom(H,G)$, where $H$ is a free group of finite rank in the variety ${\mathop{\mathrm{Var}}}(G)$ generated by $G$, is trivial or not (see Section \[se:connections\] for details). Indeed, in this case every tuple $u = (u_1, \ldots,u_n)$ of elements of $G$ gives rise to a homomorphism $\phi_u$ from a free group $H$ with basis $x_1, \ldots,x_n$ in the variety ${\mathop{\mathrm{Var}}}(G)$ such that $\phi_u(x_1) = u_1, \ldots, \phi_u(x_n) = u_n$, and vice versa. The equalizer $E(\phi_u,\phi_v)$ describes all solutions $w$ for the instance $u,v$.
It seems that the general Post correspondence problem $\GPCP$ for groups is even more interesting than the standard $\PCP$. Indeed, first of all $\GPCP$ is right in the midst of the endomorphic double twisted conjugacy problem in groups, which is one of the more difficult and less studied group theoretic conjugacy-type problems. In fact, it is shown in Section \[sec:twisted\] (Proposition \[pr:reduction-GPCP-2\]) that the double endomorphism twisted conjugacy problem in a relatively free group in ${\mathop{\mathrm{Var}}}(G)$ is equivalent to $\GPCP(G)$, and, in general, the double endomorphism twisted conjugacy problem in $G$ $\P$-time reduces to $\GPCP(G)$. Furthermore, we prove in Section \[se:GPCP\] that $\GPCP$ in a given group $G$ is intimately related to the word problem in $G$. Namely we show that the hereditary word problem ($\HWP$) in $G$ can be reduced in polynomial time to $\GPCP$ in $G$. Here $\HWP$ in $G$ asks to decide when given an element $w \in G$ and a finite subset $R \subseteq G$ if $w = 1$ in the quotient $H = G/\langle R\rangle_G$ of the group $G$ by the normal subgroup $\langle R\rangle_G$ generated by $R$. Therefore, if $\GPCP$ is decidable in $G$ then there is a [*uniform*]{} algorithm to decide the word problem in every finitely presented (relative to $G$) quotient of $G$. Further, since decidability of $\GPCP$ in $G$ is inherited by all subgroups of $G$ it implies the uniform decidability of $\HWP$ in every subgroup of $G$ (even every section of $G$). Thus, a decision algorithm for $\GPCP$ in $G$ is very powerful and it gives a lot of information about the group $G$. Notice that finitely generated abelian and nilpotent groups have decidable $\HWP$.
Results
-------
In Section \[se:GPCP\] we show that $\GPCP$ is undecidable in every non-abelian free group, as well as in every group containing free non-abelian subgroups. In particular, $\GPCP$ is undecidable in the following groups: non-elementary hyperbolic, non-abelian right angled Artin, braid groups $B_n$ ($n \geq 3$), non-solvable defined by a single relator (thus all one-relator groups with more than two generators), etc. A similar argument shows that the bounded Post correspondence problem in all groups mentioned above is $\NP$-complete. Here in the bounded version of $\GPCP$ one is looking only for solutions (the words $t(x_1, \ldots,x_n)$) whose length is bounded by a given number. We emphasize that the argument used to prove the undecidability results here has nothing to do with the original argument of undecidability of $\PCP$ or $\GPCP$ in free non-commutative semigroups, even though all the groups mentioned above contain such semigroups. In fact, it is still unclear if the $\PCP$ in a free semigroup can be reduced to $\PCP$ in a free non-abelian group. Furthermore, it is still one of the most intriguing open problems whether $\PCP$ in a free non-abelian group is decidable or not.
As a corollary of the undecidability of $\GPCP$ in free non-abelian groups we show that the double endomorphism twisted conjugacy problem in free groups $F_n$ of rank $n \geq 32$ is undecidable. Whether the double endomorphism twisted conjugacy problem is decidable or not in free non-abelian groups of smaller rank remains to be seen.
We also show that free solvable groups $S_{m,n}$ of class $m\geq 3$ and sufficiently high rank $n$ have undecidable double endomorphism twisted conjugacy problem, as well as $\GPCP$. This result is based on examples of finitely presented solvable groups with undecidable word problem constructed by Kharlampovich in [@Kharlampovich:1981].
In the opposite direction we show in Section \[se:nilpotent\] that $\PCP$ is decidable in polynomial time in every finitely generated nilpotent group $G$. This is the best known positive result up to date on $\PCP$ in groups.
Post correspondence problems {#se:PCP}
============================
The classical Post correspondence problem {#se:semigroups}
-----------------------------------------
Let $A$ be a finite alphabet with $|A| \geq 2$. Denote by $A^\ast$ the free monoid with basis $A$ viewed as the set of all words in $A$ with concatenation as the multiplication. Let $X$ be an infinite countable set of variables and $X^\ast$ the corresponding free monoid.
[**The classical Post correspondence problem ($\PCP$) in $A^*$:**]{} Given a finite set of pairs $(g_1,h_1), \ldots , (g_n,h_n)$ of elements of $A^\ast$ determine if there is a non-empty word $w(x_1, \ldots,x_n) \in X^\ast$ such that $w(g_1, \ldots,g_n) = w(h_1, \ldots,h_n) $ in $A^\ast$.
Post showed in [@Post:1946] that the problem is undecidable (see [@Sipser:2005] for a simpler proof).
Nowadays there are several variations of $\PCP$ in $A^*$, the following [*restricted*]{} version is the most typical.
[**$\PCP_n$ in $A^*$:**]{} Let $n$ be a fixed positive integer. Given a finite sequence of pairs $(g_1,h_1), \ldots , (g_m,h_m)$ of $G$, where $m \leq n,$ determine if there is a non-empty word $w(x_1, \ldots,x_m) \in X^*$ such that $w(g_1, \ldots,g_m) = w(h_1, \ldots,h_m)$ in $A^*$.
Breaking $\PCP$ into a collection of the restricted problems $\PCP_n$ makes the boundary between decidable and undecidable more clear: $\PCP_n$ in $A^*$ is decidable for $n \leq 3$, and undecidable for $n\geq 7$, see [@EKR; @Halava00binary(generalized); @Matiyasevich:1996].
Another version of interest is the general $\GPCP$ in the free monoid $A^*$, in which case an input to $\PCP$ contains a sequence of pairs $(g_1,h_1), \ldots , (g_n,h_n)$ as above and also four elements $a,b,c,d \in A^*$, while the task is to find a word $w(x_1, \ldots,x_n) \in X^*$ such that $aw(g_1, \ldots,g_n)b = cw(h_1, \ldots,h_n)d$ in $A^*$. This problem is also undecidable in $A^*$.
There are [*marked*]{} variations of the $\PCP_n$ in $A^*$, in which case for each pair $(g_i,h_i)$ in the instance the initial letters in $g_i$ and $h_i$ are not equal. These problems are known to be decidable [@Halava-Hirvensalo-deWolf:1999]. We refer to a paper [@Halava-Harju:2001] for some recent developments on the Post correspondence problem in free semigroups.
Finishing our short survey of known results we would like to mention that $\PCP$ is undecidable in a free non-abelian semigroup as well (the same argument as for free monoids). Hence semigroup version of $\PCP$ is also undecidable in semigroups that contain free non-abelian subsemigroups, in particular, in groups containing free non-abelian subgroups, or solvable not virtually nilpotent groups (they contain free non-abelian subsemigroups).
In what follows we focus only on the group theoretic versions of the Post corresponding problems $\PCP$ and $\GPCP$ in groups, which is different from the original semigroup version since one has to take inversion of elements into account.
The Post correspondence problem in groups
-----------------------------------------
Throughout the whole paper we use the following notation: $G$ is an arbitrary fixed group generated by a finite set $A$, $F(X)$ a free group with basis $X = \{x_1, \ldots, x_n\}$. We view elements of $F(X)$ as reduced words in $X \cup X^{-1}$. Sometimes we denote $F(X)$ as $F(x_1, \ldots,x_n)$, or simply as $F_n$.
As we mentioned earlier, the group theoretic version of the Post corresponding problem involves terms (words) with inversion.
[**The Post correspondence problem ($\PCP$) in a group $G$:**]{} Given a finite set of pairs $(g_1,h_1), \ldots , (g_n,h_n)$ of elements of $G$ determine if there is a word $w(x_1, \ldots,x_n) \in F(x_1, \ldots,x_n)$, which is not an identity of $G$, such that $w(g_1, \ldots,g_n) = w(h_1, \ldots,h_n)$ in $G$.
Several comments are in order here. Recall that an identity on $G$ is a word $w(x_1, \ldots,x_n)$ such that $w(g_1, \ldots,g_n) = 1$ in $G$ for any $g_1, \ldots, g_n \in G$. If the group $G$ does not have non-trivial identities then the requirement that $w$ is not an identity becomes the same as in the original Post formulation that $w$ is non-empty. Meanwhile, any non-trivial identity $w(x_1, \ldots,x_n)$ in $G$ gives a solution to any instance of $\PCP$ in $G$, which is not very interesting. Sometimes we refer to words $w$ which are identities in $G$ as to [*trivial*]{} solutions of $\PCP$ in $G$, while the solutions which are not identities in $G$ are termed [*non-trivial*]{}. In this regard $\PCP(G)$ asks to find a non-trivial solution to $\PCP$ in $G$.
In the sequel by $\PCP$ for a group $G$ we always, if not said otherwise, understand the group theoretic (not the semigroup one) version of $\PCP$ stated above. By definition $\PCP(G)$ depends on the given generating set of $G$, however it is easy to see that $\PCP(G)$ for different finite generating sets are polynomial time equivalent to each other, i.e., each one reduces to the other in polynomial time. Since in all our considerations the generating sets are finite we omit them from notation and write $\PCP(G)$.
Similar to the classical case one can define the restricted version $\PCP_n$ for a group $G$, in which case the number of pars in each instance of $\PCP_n$ is bounded by $n$, and the general one $\GPCP$ (or $\GPCP_n$), where there are some constants involved. Since the general version is of crucial interest for us we state it precisely.
[**The general Post correspondence problem ($\GPCP$) in a group $G$:**]{} given a finite sequence of pairs $(g_1,h_1), \ldots , (g_n,h_n)$ and two pairs $(a_1,b_1)$ and $(a_2,b_2)$ of elements of $G$ (called the [*constants*]{} of the instance) determine if there is a word $w(x_1, \ldots,x_n) \in F(x_1, \ldots,x_n)$ such that $a_1w(g_1, \ldots,g_n)b_1 = a_2w(h_1, \ldots,h_n)b_2$ in $G$.
Two lemmas are due here.
For any group $G$ $\GPCP(G)$ is linear time equivalent to the restriction of $\GPCP(G)$ where the constants $b_1, b_2,a_2$ are all equal to 1.
Indeed, in the notation above notice that $a_1w(g_1, \ldots,g_n)b_1 = a_2w(h_1, \ldots,h_n)b_2$ in $G$ if and only if $$a_2^{-1}a_1w(g_1, \ldots,g_n)b_1b_2^{-1} = w(h_1, \ldots,h_n),$$ so $\GPCP$ in $G$ is equivalent to $\GPCP$ with $a_2 = 1, b_2 = 1$. Moreover, $$aw(g_1, \ldots,g_n)b = w(h_1, \ldots,h_n)$$ in $G$ if and only if $$abb^{-1}w(g_1, \ldots,g_n)b = w(h_1, \ldots,h_n),$$ i.e., $$abw(g_1^b, \ldots,g_n^b) = w(h_1, \ldots,h_n).$$ Hence $\GPCP(G)$ is linear time equivalent to $\GPCP(G)$ with $b_1= a_2 = b_2 = 1$, as claimed.
From now on we often assume that in $\GPCP$ each instance has the constants $b_1, b_2,a_2$ are all equal to 1, in which case we denote $a_1$ by $a$ and term it the [*constant*]{} of the instance.
\[le:reduction-GPCP\] For any group $G$ and for any instance $(g_1,h_1), \ldots , (g_n,h_n), a$ of $\GPCP(G)$ all solutions $w$ to this instance can be described as $w = w_0u$, where $w_0$ is a particular fixed solution to this instance and $u$ is an arbitrary (perhaps, trivial) solution to $\PCP(G)$ for the instance $(g_1,h_1), \ldots , (g_n,h_n)$.
Suppose $w_0$ is a particular fixed solution to $\GPCP(G)$ for the instance $(g_1,h_1), \ldots , (g_n,h_n), a$, so $aw_0(g_1, \ldots,g_n) = w_0(h_1, \ldots,h_n)$. If $w$ is an arbitrary solution to the same instance in $G$ then $aw(g_1, \ldots,g_n) = w(h_1, \ldots,h_n)$, so $$w_0^{-1}(g_1, \ldots,g_n)w(g_1, \ldots,g_n) = w_0^{-1}(h_1, \ldots,h_n)w(h_1, \ldots,h_n),$$ hence $u = w_0^{-1}w$ solves $\PCP(G)$ for the instance $(g_1,h_1), \ldots , (g_n,h_n)$. Therefore, $w = w_0u$ as claimed.
Lemma \[le:reduction-GPCP\] shows that to get all solutions of $\GPCP$ in $G$ for a given instance one needs only to find a particular solution of $\GPCP(G)$ and all solutions of $\PCP(G)$ for the same instance. In view of this we sometimes refer to $\GPCP$ as the [*non-homogeneous*]{} $\PCP$, and to $\PCP$ as to the [*homogeneous*]{} one.
As usual in discrete optimization there are several other standard variations of $\PCP$ problems: [*bounded, search*]{}, and [*optimal*]{}. We mention them briefly now and refer to [@MNU1] for a thorough discussion of these types of problems in groups. The [*bounded*]{} version of $\PCP$ (or $\GPCP$) requires that the word $w$ in question should be of length bounded from above by a given number $M$. We denote these versions by $\BPCP(G)$ or $\BGPCP(G)$. The [*search*]{} variation of $\PCP$ (or $\GPCP$) asks to find a word $w$ that gives a non-trivial solution to a given instance of the problem (if such a solution exists). The [*optimization*]{} version of $\PCP$ (or $\GPCP$) is a variation of the search problem, when one is asked to find a solution that satisfies some “optimal” conditions. In our case, if not said otherwise, the optimal condition is to find a shortest possible word $w$ which is a solution to the given instance of the problem.
Connections to group theory {#se:connections}
===========================
$\PCP_n$ and the equalizer problem
----------------------------------
Let as above $G$ be a fixed arbitrary group with a finite generating set $A$, $F_n = F(x_1, \ldots,x_n)$ a free group with basis $X= \{x_1, \ldots,x_n\}$.
An $n$-tuple of elements $g= (g_1, \ldots,g_n) \in G^n$ gives a homomorphism $\phi_g:F_n \to G$ where $\phi_g(x_1) = g_1, \ldots, \phi_g(x_n) = g_n$. And vice versa, every homomorphism $F_n \to G$ gives a tuple as above. In this sense each instance $(u_1,v_1), \ldots, (u_n,v_n)$ of $\PCP(G)$ can be uniquely described by a pair of homomorphisms $\phi_u,\phi_v:F_n\to G$, where $u = (u_1, \ldots,u_n), v = (v_1, \ldots, v_n)$. In this case we refer to such a pair of homomorphisms as an instance of $\PCP$ in $G$.
Now given groups $H,G$ and two homomorphism $\phi, \psi \in \Hom(H,G)$ one can define the equalizer $E(\phi,\psi)$ of $\phi,\psi$ as $$\label{eq:equalizer}
E(\phi,\psi) = \{w \in H \mid w^\phi = w^\psi\},$$ which is obviously a subgroup of $H$. If $G$ does not have non-trivial identities then all non-trivial words from $E(\phi,\psi)$ give all solutions to $\PCP$ in $G$ for a given instance $(\phi, \psi) \in \Hom(F_n,G)$. However, if $G$ has non-trivial identities then some words from $E(\phi,\psi)$ are identities which are not solutions to $\PCP(G)$. To accommodate all the cases at once we suggest to replace the free group $F_n$ above by the free group $F_{G,n}$ in the variety ${\mathop{\mathrm{Var}}}(G)$ of rank $n$ with basis $\{x_1, \ldots, x_n\}$. Then similar to the above every tuple $u \in G^n$ gives rise to a homomorphism $\phi_u:F_{G,n} \to G$, where $\phi(x_1) = u_1, \ldots, \phi(x_n) = u_n$, and non-trivial elements of the equalizer $E(\phi_u,\phi_v)$ describe all solutions of $\PCP(G)$ for the instance $u,v \in G^n$. This connects $\PCP_n$ in $G$ with the equalizers of homomorphisms from $\Hom(F_{G,n},G)$.
There are two general algorithmic problems in groups concerning equalizers.
[**The triviality of the equalizer problem $ (\TEP(H,G))$ for groups $H,G$:**]{} Given two homomorphisms $\phi, \psi \in \Hom(H,G)$ decide if the subgroup $E(\phi,\psi)$ in $H$ is trivial or not.
[**The equalizer problem $ (\EP(H,G))$ for groups $H,G$:**]{} Given two homomorphisms $\phi, \psi \in \Hom(H,G)$ find the equalizer $\EP(H,G)$. In particular, if $\EP(H,G)$ is finitely generated then find a finite generating set of $E(\phi,\psi)$.
The formulation above needs some explanation on how we mean “to find” a subgroup in a group. If the subgroup is finitely generated then “to find” usually means to list a finite set of generators. It might happen that the subgroup is not finitely generated, but allows a finite set of generators as a normal subgroup, or as a module under some action. In this case to solve $ \EP(H,G)$ one has to list a finite set of these generators of $ \EP(H,G)$. In this paper we consider equalizers of homomorphisms of finitely generated nilpotent groups, so in this event they are finitely generated and the problem of describing equalizers becomes well-stated.
Equalizers $E(\phi,\psi)$ were studied before, but mostly in the case when $H = G$ and $\phi, \psi$ are automorphisms of $G$. There are few results on equalizers of endomorphisms in groups. Goldstein and Turner have proved in [@Goldstein-Turner:1986] that the equalizer of two endomorphisms of $F_n$ is a finitely generated subgroup in the case one of the two maps is injective. However, is it not known whether there is an algorithm to decide if the equalizer of two endomorphisms in a free group $F_n$ is trivial or not. Ciobanu, Martino and Ventura showed that generically equalizers of endomorphisms in free groups are trivial [@Ciobanu08thegeneric], so on most inputs in a free non-abelian group $F$ $\PCP(F)$ does not have a solution, in this sense $\PCP(F)$ is generically decidable.
We summarize the discussion above in the following easy lemma.
Let $G$ be a group. Then the following holds for any natural $n>0$:
- $\PCP_n(G)$ is equivalent (being just a reformulation) to $\TEP$ for homomorphisms from $Hom(F_{G,n},G)$.
- Finding all solutions for a given instance of $\PCP_n(G)$ is equivalent (being just a reformulation) to $\EP(F_{G,n},G)$ for the same instance.
$\GPCP$ and the double twisted conjugacy {#sec:twisted}
----------------------------------------
Let $\phi, \psi$ be two fixed automorphisms of a group $G$. Two elements $u, v \in G$ are termed [*$(\phi,\psi)$-double-twisted conjugate*]{} if there is an element $w \in G$ such that $uw^\phi = w^\psi v$. In particular, when $\psi = 1$ then $u$ and $v$ are called [*$\phi$-twisted conjugate*]{}, while in the case $\phi = \psi = 1$ $u$ and $v$ are just usual conjugates of each other. The twisted (or double twisted) conjugacy problem in $G$ is to decide whether or not two given elements $u, v \in G$ are twisted (double twisted) conjugate in $G$ for a fixed pair of automorphisms $\phi, \psi \in \mathop{\mathrm{Aut}}(G)$. Observe, that since $\psi$ has the inverse the $(\phi,\psi)$-double-twisted conjugacy problem reduces to $\phi\psi^{-1}$-twisted conjugacy problem, so in the case of automorphisms it is sufficient to consider only twisted conjugacy problem. This problem is much studied in groups, we refer to [@Ventura-Romankov:2009; @Romankov:2010; @Romankov:2011; @BMMV; @Troitsky; @Fel'shtyn; @Fel'shtyn-Leonov-Troitsky] for some recent results.
Much stronger versions of the problems above appear when one replaces automorphisms by arbitrary endomorphisms $\phi, \psi \in End(G)$. Not much is known about double twisted conjugacy problem in groups with respect to endomorphisms.
The next statement (which follows from the discussion above) relates the double-twisted conjugacy problem for endomorphisms to the general Post correspondence problem.
\[pr:reduction-GPCP-2\] Let $G$ be a group generated by a finite set $A = \{a_1, \ldots,a_n\}$. Then the following holds:
- The double-twisted conjugacy problem for endomorphisms in $G$ is linear time reducible to $\GPCP_n(G)$.
- If $G$ is relatively free with basis $A$ then the double-twisted conjugacy problem for endomorphisms in $G$ is linear time equivalent to $\GPCP_n(G)$.
The hereditary word problem and $\GPCP$ {#se:GPCP}
=======================================
It is easy to see that decidability of $\PCP_n$ or $\GPCP_n$ in a group $G$ has some implications for the word problem in $G$. Indeed, an element $g$ is equal to 1 in $G$ if and only if $\GPCP_1$ is decidable in $G$ for the instance consisting of a single pair $(1,1)$ and the constant $g$. Similarly, if $G$ is torsion-free then $g = 1$ in $G$ if and only if $\PCP$ is decidable in $G$ for the instance pair $(g,1)$. In this section we show that the whole lot of word problems in the quotients of $G$ is reducible to $\GPCP$ in $G$.
Let $G$ be a group generated by a finite set $A$. For a subset $R\subseteq G$ by $\langle R\rangle_G$ we denote the normal closure of $R$ in $G$.
[**The hereditary word problem ($\HWP(G)$) in $G$:**]{} Given a finite set $R\cup\{w\}$ of words in the alphabet $A\cup A^{-1}$, decide whether or not $w$ is trivial in the quotient $G/\langle R\rangle_G$.
Note that this problem can also be stated as the uniform membership problem to normal finitely generated subgroups of $G$. Observe also that $\HWP(G)$ requires a [*uniform*]{} algorithm for the word problems in the quotients $G/\langle R\rangle_G$.
It seems that groups with decidable $\HWP$ are rare. Notice that the hereditary word problem is decidable in finitely generated abelian or nilpotent groups.
\[prop:WPtoPCP\] Let $G$ be a finitely generated group. Then the hereditary word problem in $G$ $\P$-time reduces to $\GPCP(G)$.
Let $A$ be a finite generating set of $G$. Suppose $R$ is a finite set of elements of $G$, represented by words in $A\cup A^{-1}$. Denote $H=G/\langle R\rangle_G$. Put $$D_R=\{(a,a^{-1})\mid a\in A\}\cup\{(a,a^{-1})\mid a\in A\}\cup \{(r,1)\mid r \in R\}\cup \{(r^{-1},1)\mid r \in R\}.$$
[**Claim 1.**]{} Let $w$ be a word $w\in (A\cup A^{-1})^\ast$. Then $w=_H1$ if and only if there is a finite sequence of pairs $(u_1,v_1),\ldots,(u_k,v_k)\in D_R$ such that $$\label{eq:u-v}
v_n(\cdots (v_2(v_1 wu_1)u_2)\cdots )u_n=_G 1.$$ Indeed, if (\[eq:u-v\]) holds then $$w=_G v_1^{-1} \ldots v_{n-1}^{-1}(v_n^{-1}u_n^{-1})u_{n-1} ^{-1}\ldots u_1^{-1} =_H 1$$ since for every pair $(u,v) \in D_R$ one has $uv = 1$ in $H$.
To show the converse, suppose $w =_H 1$, i.e., $w \in \langle R\rangle_G$. In this case $$\label{eq:wp}
w=_Gw_1r_1w_2\ldots w_{m}r_mw_{m+1}$$ with $r_i\in R, w_i\in A^\ast$ and $w_1w_2\ldots w_{m+1}=_G1$. Rewriting (\[eq:wp\]) one gets $$\label{eq:wp2}
r_1^{-1} \cdot w_1^{-1} \cdot w\cdot w_1 \cdot 1 =_G w_2r_2w_3\ldots w_{m}r_mw_{m+1}w_1.$$ Notice that the product on the left is in the form required in (\[eq:u-v\]), and the product on the right is in the form required in (\[eq:wp\]). Now the result follows by induction on $m$. This proves the claim.
[**Claim 2.**]{} Let $R\subseteq (A\cup A^{-1})^*$ be a finite set and $w \in (A\cup A^{-1})^*$. Then $\GPCP(G)$ has a solution for the instance $\hat{D}_R = \{(u,v^{-1})\mid(u,v)\in D_R\}$ with the constant $w$ if and only if $w = 1$ in $H$.
Indeed, a sequence $$\label{eq:pcp_witness}
(u_1,v_1^{-1}),\ldots ,(u_M,v_M^{-1}) \in \hat{D}_R$$ gives a solution to $\GPCP(G)$ for the instance $\hat{D}_R$ with the constant $w$ if and only if $$wu_1u_2\cdots u_M=_Gv^{-1}_1v^{-1}_2\cdots v^{-1}_M\iff v_M(\cdots (v_2(v_1 wu_1)u_2)\cdots )u_M=_G 1,$$ which, by the claim above, is equivalent to $w=_H 1$.
This proves Claim 2 together with the proposition.
\[co:PCP\_free\] Let $F$ be a free non-abelian group of finite rank. Then $\GPCP(F)$ is undecidable.
It is known [@Miller] that for any natural number $n \geq 2$ there are finitely presented groups with $n$ generators and undecidable word problem. Therefore, $\HWP(F)$ is undecidable. By Proposition \[prop:WPtoPCP\] $\GPCP(F)$ is also undecidable.
For a finite group presentation $P=\langle a_1, \ldots,a_k \mid r_1, \ldots, r_\ell\rangle$ denote by $N(P) = k+\ell$ the total sum of the number of generators and relators in $P$. Let $N$ be the least number $N(P)$ among all finite presentations $P$ with undecidable word problem. In [@Borisov:1969] Borisov constructed a finitely presented group with $4$ generators and $12$ relations which has undecidable word problem, so $N \leq 16$.
\[co:twisted-free\] Let $F_n$ be a free group of rank $n \geq 32$. Then the endomorphism double twisted conjugacy problem in $F_n$ (as well as $\GPCP_n(F_n)$) is undecidable.
Let $P^\prime = \langle a_1, \ldots,a_4 \mid r_1, \ldots, r_{12}\rangle $ be the Borisov’s presentation and $F_n = \langle a_1, \ldots,a_n \rangle$ a free group of rank $n\geq 32$. Claim 2 in the proof of Proposition \[prop:WPtoPCP\] shows that the word problem in the group $H$ defined by the presentation $P^\prime$ is polynomial time reducible to $\GPCP_n(F_n)$, hence the latter one is undecidable. Now the part 2 in Proposition \[pr:reduction-GPCP-2\] shows that the endomorphism double twisted conjugacy problem in $F_n$ is also undecidable, as claimed.
Note that the twisted conjugacy problem is decidable in free groups [@BMMV]. Together with Corollary \[co:twisted-free\], this gives the following result.
\[co:venturas\_question\] Free groups of rank at least $32$ have decidable twisted conjugacy problem but undecidable endomorphism double twisted conjugacy problem.
Note that for a given group, decidability of the endomorphism double twisted conjugacy problem implies decidability of the twisted conjugacy problem, which in turn implies decidability of the conjugacy problem. It was shown in [@BMV] that the converse to the latter implication is in general false. The above result \[co:venturas\_question\] answers E. Ventura’s question whether the converse to the former implication is true.
Similar results hold for free solvable groups. Let $N_{sol}$ be the least number $N(P)$ among all finite presentations $P$ which define a solvable group with undecidable word problem. In [@Kharlampovich:1981] Kharlampovich constructed a finitely presented solvable group with undecidable word problem, so such number $N_{sol}$ exists.
\[co:PCP\_free\_sol\] Let $S_{m,n}$ be a free solvable non-abelian group of class $m\ge 3$ and rank $n \geq N_{sol}$. Then the endomorphism double twisted conjugacy problem in $S_{m,n}$ (as well as $\GPCP_n(S_{m,n})$) is undecidable.
Similar to the argument in Corollary \[co:twisted-free\].
Observe that it immediately follows from definitions that decidability of $\PCP$ or $\GPCP$ in a finitely generated group is inherited by all finitely generated subgroup of $G$. Therefore, the results above give a host of groups with undecidable $\GPCP$ (as well as $\GPCP_n$).
If a group $G$ contains a free non-abelian subgroup $F_2$ then $\GPCP(G)$ is undecidable.
Therefore $\GPCP$ is undecidable, for example, in non-elementary hyperbolic groups, non-abelian right angled Artin groups, groups with non-trivial splittings into free products with amalgamation or HNN extensions, braid groups $B_n$, non-virtually solvable linear groups, etc.
Another corollary of the results above concerns with complexity of the bounded $\GPCP$ in groups.
\[co:NP\_PCP\_free\] Let $F$ be a non-abelian free group of finite rank. Then the bounded $\GPCP(F)$ is $\NP$-complete.
Let $F = F(A)$ be a free non-abelian group with a finite basis $A$. It is showed in [@Sapir-Birget-Rips:2002 Corollary 1.1] that there exists a finitely presented group $H=\langle B\mid R\rangle$ with $\NP$-complete word problem and polynomial Dehn function $\delta_H(n)$. Passing to a subgroup of $F(A)$, we may assume that $A=B$. One can see that in the case of a free group $G=F(A)$, $M$ in (\[eq:pcp\_witness\]) is bounded by a polynomial (in fact, linear) function of $|w|$ and the number $m$ of relators in (\[eq:wp\]) (see [@Olshanskii_Sapir:2001 Lemma 1] for details). Note that there exists $m$ as above bounded by $\delta_H(|w|)$, so $M$ is bounded by some polynomial $q(|w|)$. Therefore, the map $$w\to (w, D_R, M=q(|w|))$$ is a $\P$-time reduction of the word problem in $H$ to the bounded $\GPCP(F(A))$. It follows that the latter is $\NP$-hard and therefore $\NP$-complete (since the word problem in $F(A)$ is $\P$-time decidable).
\[co:NP\_PCP\_free\_sub\] If a group $G$ contains a free non-abelian subgroup $F_2$ then the bounded $\GPCP(G)$ is $\NP$-hard.
$\PCP$ in nilpotent groups {#se:nilpotent}
==========================
In this section we study complexity of Post correspondence problems in nilpotent groups.
\[prop:abelian\] There is a polynomial time algorithm that given finite presentations of groups $A$, $B$ in the class of abelian groups and a homomorphism $\phi:A \to B$ computes a finite set of generators of the kernel of $\phi$.
Results of [@Kannon-Bachem] provide a polynomial time algorithm to bring an integer matrix to its canonical diagonal (Smith) normal form. Since computing the canonical presentation of a finitely presented abelian group reduces by a standard argument to finding Smith form of an integer matrix (determined by relators in a given presentation), we may find in polynomial time the canonical presentation of $B$, i.e. a direct decomposition $B = \mathbb{Z}^l \times K$, where $K$ is a finite abelian group. Once $B$ is in its canonical form, computing kernel of $\phi$ reduces to solving a system of linear equations in $Z^l$ and $K$, which can be done in polynomial time by the same results [@Kannon-Bachem].
\[co:PCP\_abelian\] There is a polynomial time algorithm that given finite presentations of groups $A$, $B$ in the class of abelian groups and homomorphisms $\phi, \psi \in \Hom(A,B)$ computes a finite set of generators of the equalizer $E(\phi,\psi)$.
Observe that a map $\xi:A \to B$ defined by $\xi(g) = \phi(g)\psi(g)^{-1}$ is a homomorphism from $A$ to $B$ and $E(\phi,\psi) = \ker\xi$. Now the result follows from Proposition \[prop:abelian\].
One can slightly strengthen the corollaries above.
\[co:PCP\_abelian\_2\] Let $c$ be a fixed positive integer.
- There is a polynomial time algorithm that given a finite presentation of a group $A$ (perhaps in the class of nilpotent groups of class $c$), and a finite presentation of a group $B$ in the class of abelian groups, and a homomorphism $\phi \in \Hom(A,B)$ computes a finite set of generators of the kernel $\ker \phi$ modulo the commutant $[A,A]$.
- There is a polynomial time algorithm that given a finite presentation of a group $A$ (perhaps in the class of nilpotent groups of class $c$), and a finite presentation of a group $B$ in the class of abelian groups, and a homomorphism $\phi, \psi \in \Hom(A,B)$ computes a finite set of generators of the equalizer $E(\phi,\psi)$ modulo the commutant $[A,A]$.
Follows immediately from Proposition \[prop:abelian\] and Corollary \[co:PCP\_abelian\].
By $\gamma_{c}(G)$ we denote the $c$’s term of the lower central series of $G$. Recall that the iterated commutator of elements $g_1,\ldots, g_c$ is $[g_1,g_2,\ldots,g_c]=[\ldots [[g_1,g_2],g_3],\ldots ]$. The following lemma is well known (for example, see [@Kargapolov-Merzlyakov Lemma 17.2.1]).
\[le:nilpotent\_basis\] Let $G$ be a group generated by elements $x_1,\ldots, x_n\in G$. Then $\gamma_c(G)$ is generated as a subgroup by $\gamma_{c+1}(G)$ and iterated commutators $[x_{i_1},\ldots, x_{i_{c}}]$.
\[le:commutant\] Let $c_0$ be a fixed positive integer. There is a polynomial time algorithm that given a finite group presentation of a group $G$ in the class of nilpotent groups of class $\le c_0$, finds subgroup generators of $[G,G]$.
Follows from Lemma \[le:nilpotent\_basis\] by an inductive construction since there are at most $n^{c_0+1}$ iterated commutators $[x_{i_1},\ldots, x_{i_{c}}]$, $c\le c_0$, in a group generated by $n\ge 2$ elements $x_1,\ldots,x_n$ (the case $n=1$ is obvious).
\[th:equalizer\] Let $c_0$ be a fixed positive integer. Then there is a polynomial time algorithm that given positive integers $c_H, c_G\le c_0$, finite presentations of groups $H,G$ in the classes of nilpotent groups of class $c_H$ and $c_G$, respectively, and homomorphisms $\phi, \psi \in Hom(H,G)$ computes the generating set of the equalizer $E(H,\phi,\psi)$ as a subgroup of $H$.
Let $Y$ and $Z$ be finite generating sets of $H$ and $G$, respectively. We use induction on the nilpotency class $c=c_G$ of $G$. If $c = 1$ then $G$ is abelian and the result follows from Corollary \[co:PCP\_abelian\_2\] and \[co:PCP\_abelian\].
Suppose now that $c >1$ and we are given $\phi, \psi \in \Hom(H,G)$. Consider the quotient group $\bar G = G/\gamma_c(G)$, which is a nilpotent group of class $c-1$. The homomorphisms $\phi, \psi$ induce some homomorphisms $\phi', \psi' \in \Hom(H, \bar G)$. Observe that the size of $\phi', \psi'$ (the total length of the images $\phi(y), \psi(y), y \in Y$ as words in $Z$) is the same as of $\phi, \psi$. Also observe that $\bar G$ is described in the class of nilpotent groups of class $c-1$ by the same presentation that describes $G$ in the class of nilpotent groups of class $c$. By induction we can compute in polynomial time a finite generating set, say $L = \{h_1, \ldots, h_k\}$, of $E'=E(H,\phi',\psi')$ as a subgroup of $H$. By construction, for $g\in E'$ one has $\phi(g) = \psi(g) \mod \gamma_c(G)$, hence a map $\xi(g) = \phi(g)\psi(g)^{-1}$ defines a homomorphism $\xi:E' \to \gamma_c(G)$. Obviously, $E(\phi,\psi) = \ker \xi$. Further, note that the size of $L$ is polynomial in terms of size the input, and the size of a generating set for $\gamma_c(G)$ is polynomial (of degree that depends on $c$) in terms of size of a generating set for $G$ by Lemma \[le:nilpotent\_basis\]. Now the result follows from Corollary \[co:PCP\_abelian\_2\], item 1), since $\gamma_c(G)$ is abelian, and Lemma \[le:commutant\]
\[th:nilp\_P\] Let $c$ be a fixed positive integer.
- Let $G$ be a finitely generated nilpotent group of class $c$. Then for any $\phi, \psi \in \Hom(F_n,G)$ the subgroup $E(\phi,\psi) \leq F_n$ contains $\gamma_{c+1}(F_n)$ and is finitely generated modulo $\gamma_{c+1}(F_n)$.
- There is a polynomial time algorithm that given a positive integer $n$, a presentation of a group $G$ in the class of nilpotent groups of class $c$ and homomorphisms $\phi, \psi\in \Hom(F_n,G)$ computes a finite set of generators of $E(\phi,\psi)$ in $F_n$ modulo the subgroup $\gamma_{c+1}(F_n)$.
Let $F_n = F_n(X)$, where $X = \{x_1, \ldots,x_n\}$. Fix two homomorphisms $\phi, \psi \in \Hom(F_n,G)$. Since $G$ is nilpotent of class $c$ one has $\gamma_{c+1}(G) = 1$, so $E(\phi,\psi) \geq \gamma_{c+1}(F_n)$. The quotient $N_{n,c} = F_n/\gamma_{c+1}(F_n)$ is a finitely generated free nilpotent group of rank $n$ and class $c$, hence every its subgroup, in particular the image ${\bar E}$ of $E(\phi,\psi)$, is finitely generated. It follows that the group $E(\phi,\psi)$ is finitely generated modulo $\gamma_{c+1}(F_n)$. This proves 1). Notice, that the argument above allows one to reduce everything to the case of nilpotent groups, i.e., to consider the induced homomorphisms $\bar \phi, \bar \psi \in \Hom(N_{n,c}, G)$, instead of $\phi, \psi$, and the subgroup $\bar E$ instead of $E(\phi,\psi)$. Now the result follows from Theorem \[th:equalizer\].
\[th:nilpotent\] Let $G$ be a finitely generated nilpotent group. Then $\PCP_n(G) \in \P$ for every $n \in \mathbb{N}$.
Indeed, by Theorem \[th:nilp\_P\] one can compute in $\P$-time a finite set of elements $h_1, \ldots,h_m \in F_n$ such that $E(\phi,\psi) = \langle h_1, \ldots,h_m, \gamma_{c+1}(F_n) \rangle$. Now the instance of $\PCP_n$ defined by $(\phi, \psi)$ has a non-trivial solution in $G$ if and only if there is $i$ such that $\phi(h_i) \neq 1$ in $G$. Indeed, in this case $\phi(h_i) = \psi(h_i) \neq 1$ in $G$. Otherwise, $\phi(E(\phi,\psi)) = 1$ in $G$ so there is no a non-trivial solution in $G$ to the instance of $\PCP_n$ determined by $\phi$ and $\psi$. This proves the theorem.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The Fridman invariant, which is a biholomorphic invariant on Kobayashi hyperbolic manifolds, can be seen as the dual of the much studied squeezing function. We compare this pair of invariants by showing that they are both equally capable of determining the boundary geometry of a bounded domain if their boundary behaviour is apriori known.'
address:
- 'PM: Department of Mathematics, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India'
- 'KV: Department of Mathematics, Indian Institute of Science, Bangalore 560 012, India'
author:
- 'Prachi Mahajan, Kaushal Verma'
title: A comparison of two biholomorphic invariants
---
Introduction
============
Recall that the squeezing function associated to a bounded domain is a measure of the largest euclidean ball contained in all possible holomorphic embeddings of the given domain into the unit ball in ${\mathbb}C^n$. More precisely, for a bounded domain $D \subset {\mathbb}C^n$ and $p \in D$, let ${\mathcal}F$ be the family of all injective holomorphic maps $ f $ from $D$ to the unit ball ${\mathbb}B^n \subset {\mathbb}C^n$ that map $p$ to the origin. Let $S_D(p, f)$ be the supremum of those $r > 0$ for which the image $f(D)$ contains $B^n(0, r)$, the euclidean ball of radius $r$ around the origin in $ \mathbb{C}^n $. The squeezing function $s_D : D \mapsto (0, 1]$ is defined as $$s_D(p) = \sup \{S_D(p, f) : f \in {\mathcal}F \}.$$ That this is a biholomorphic invariant follows from its definition and when $D = {\mathbb}B^n$, it can be checked that $s_D \equiv 1$. Various aspects of $s_D$ have been studied of late but among those that are directly relevant to this note are its boundary behaviour on some classes of domains (for example [@FW], [@KZ], [@N] and [@NA]) and conversely, its efficacy in determining some geometric properties of the boundary of the domain if its boundary behaviour is a priori known – for example, [@SK] and [@Z2].
It is interesting to note that another biholomorphic invariant, that is dual to the squeezing function in much the same way as the Carathéodory and Kobayashi metrics are, was defined by Fridman in [@Fr1], [@Fr2]. Let us recall its construction: for $X$ a Kobayashi hyperbolic complex manifold of dimension $n$, let $B_X(p, r)$ be the Kobayashi ball around $p \in X$ of radius $r > 0$. Let ${\mathcal}R$ be the set of all $r > 0$ such that there is an injective holomorphic map $f : {\mathbb}B^n {\rightarrow}X$ with $B_X(p, r) \subset f({\mathbb}B^n)$. Note that ${\mathcal}R$ is non-empty. Indeed, the hyperbolicity of $X$ implies that the intrinsic topology on it is equivalent to that induced by the Kobayashi metric. Hence, for small $r >0$, the ball $B_X(p, r)$ is contained in a coordinate chart and this shows that there is an injective holomorphic map from the ball into $X$ whose image contains $B_X(p, r)$. The Fridman invariant is $$h_X(p) = \inf_{r \in {\mathcal}R} \frac{1}{r}$$ which is a non-negative real-valued function on $X$. This is a biholomorphic invariant since the Kobayashi metric is itself preserved by such maps and among other things, Fridman showed that (i) $h_X$ is continuous and that (ii) if $h_X(p) = 0$ for some $p \in X$, then $X$ is biholomorphic to ${\mathbb}B^n$ in [@Fr1]. Other aspects of this invariant such as its boundary behaviour were studied in [@MV].
The purpose of this note is (i) to show that much like the squeezing function, the Fridman invariant can also determine the nature of the boundary of a given domain if its boundary behaviour is a priori known and (ii) to localize and provide a different proof of some of the results in [@N] and [@SK]. While both (i) and (ii) use the methods of scaling, they rely on an observation made in [@MV] namely, the convergence of the integrated Kobayashi distance on each scaled domain to that in the limiting domain. More specifically, refer to Lemma 5.2 and 5.7 of [@MV].
Let $D \subset {\mathbb}C^n$ be a bounded convex domain with $C^{\infty}$-smooth boundary. Then $\partial D$ is strongly pseudoconvex if $h_D(z) {\rightarrow}0$ as $z {\rightarrow}{\partial}D$.
The corresponding statement for the squeezing function $s_D(z)$ is already known – see [@Z2] for instance. Note that the boundary ${\partial}D$ can apriori be of infinite type near $p^0$ in this theorem.
A similar result holds for $h$-extendible boundary points. Before proceeding further, recall that $ p^0 \in \partial D $ is said to be an $h$-extendible boundary point if $ \partial D $ is smooth pseudoconvex finite type near $ p^0 $ and the Catlin and the D’Angelo multitypes at $ p^0 $ coincide. The class of $h$-extendible points includes smooth pseudoconvex finite type boundary points in $ \mathbb{C}^2 $, convex finite type points in $ \mathbb{C}^n $ and pseudoconvex finite type boundary points in $ \mathbb{C}^n $ with Levi-form having at most one degenerate eigenvalue.
Let $D \subset {\mathbb}C^n$ be a bounded domain with $p^0 \in \partial D$. Assume that ${\partial}D$ is $C^{\infty}$-smooth and $h$-extendible near $p^0$. Then $\partial D$ is strongly pseudoconvex near $p^0$ if either $h_D(z) {\rightarrow}0$ or $s_D(z) {\rightarrow}1$ as $z {\rightarrow}p^0$.
Here, it turns out that $\partial D$ is strongly pseudoconvex near $p^0$ if either $h_D(p^j) {\rightarrow}0$ or $s_D(p^j) {\rightarrow}1$ for a sequence $ p^j $ in $ D $ converging to $ p^0 $ only along the inner normal to $ \partial D $ at $ p^0 $. This will be evident from the proof of this theorem.
For a domain $ D $ in $ \mathbb{C}^n $, $ F_D $ denotes its Kobayashi-Royden infinitesimal metric and $ d_D $ its integrated Kobayashi distance.
Before proving these theorems, we begin with:
Two Examples
============
The Fridman invariant for the unit polydisc $ \Delta^n $ is given by $$h_{\Delta^n}(z) = 2 \left(\log \left( \frac{\sqrt{n} + 1}{\sqrt{n} - 1} \right) \right)^{-1}$$ for every $z \in \Delta^n$.
Since $ \Delta^n $ is homogeneous and the function $ h_{\Delta^n} \left( \cdot \right) $ is a biholomorphic invariant, it is enough to compute the explicit formula for $ h_{\Delta^n } $ at the origin. If $ f: \mathbb{B}^n \rightarrow \Delta^n $ is a holomorphic imedding such that $ f(0) = 0 $ and $$\Delta^n \left( 0, \frac{e^{2r} -1}{e^{2r} + 1}\right) = B_{\Delta^n} (0,r) \subset f(\mathbb{B}^n),$$ then it follows from [@Alexander] that $$\frac{e^{2r} -1}{e^{2r} + 1} \leq \frac{1}{\sqrt{n}},$$ or equivalently that, $$\frac{1}{r} \geq 2 \left( \log \left( \frac{ \sqrt{n} + 1}{ \sqrt{n} -1} \right)\right)^{-1},$$ which implies that $$\label{k14}
h_{\Delta^n} \left(0 \right) \geq 2 \left( \log \left( \frac{ \sqrt{n} + 1}{ \sqrt{n} -1} \right)\right)^{-1}.$$ On the other hand, consider $ \psi_0 = i \circ \psi $, where $ \psi $ is an automorphism of $ \mathbb{B}^n $ that preserves the origin and $ i : \mathbb{B}^n \rightarrow \Delta^n $ is the inclusion map. Then $ \psi_0$ is an imbedding of $ \mathbb{B}^n $ into $ \Delta^n $ satisfying $ \psi_0(0) = 0 $ and $$\Delta^n \left( 0, \frac{1}{\sqrt{n}} \right) = B_{\Delta^n} \left(0, \frac{1}{2} \log \left( \frac{ \sqrt{n} + 1}{ \sqrt{n} -1} \right) \right)
\subset \psi_0(\mathbb{B}^n),$$ and hence $$\label{k15}
h_{\Delta^n} \left(0 \right) \leq 2 \left( \log \left( \frac{ \sqrt{n} + 1}{ \sqrt{n} -1} \right)\right)^{-1}.$$ Combining the inequalities (\[k14\]) and (\[k15\]) yields the desired expression for $ h_{\Delta^n} $.
Let $\{p^j\}$ be a sequence in the punctured disc $\Delta {\setminus}\{0\}$ that converges to the origin. Then $$\label{E9}
L_j \le h_{\Delta {\setminus}\{0\}}(p^j) \le U_j$$ where $$L_j^{-1} = \log \left( 2 \left( -\frac{\pi}{\log \vert p^j \vert} \right)^2 + 1 + \frac{2 \pi}{\log \vert p^j \vert} \sqrt{\left( -\frac{\pi}{\log \vert p^j \vert} \right)^2 + 1} \right)$$ and $$U_j^{-1} = \log \left( \left(-\frac{\pi}{\log \vert p^j \vert} \right) + \sqrt{\left(-\frac{\pi}{\log \vert p^j \vert}\right)^2 + 1 } \right).$$
Since $ h_{\Delta {\setminus}\{0\}} (\cdot) $ is a biholomorphic invariant, after composing with an appropriate automorphism of $ \Delta \setminus \{0\} $, we may assume that each $ p^j $ lies in the open interval $ (0,1) $. Consider the slit disc $ \Delta \setminus (-1, 0] $, which is a simply connected domain. Choose a conformal map $ f^j $ from the unit disc $ \Delta $ onto the slit domain $ \Delta \setminus (-1,0] $ such that $ f^j (0) = p^j $. Then $$B_{ \Delta \setminus \{0\} } \left( p^j, r (p^j) \right) \subset \Delta \setminus (-1,0],$$ where $$\label{E8}
r(p^j) = \log \left( - \frac{\pi}{\log p^j} + \sqrt{ \left( - \frac{\pi}{\log p^j}\right)^2 + 1} \right).$$ To establish this claim, it suffices to show that $$\label{E7}
d_{\Delta \setminus \{0\} } \left( p^j, (-1,0) \right) : = \inf_{q \in (-1,0)} d_{\Delta \setminus \{0\} } \left( p^j, q \right) = r(p^j).$$ To verify this, first recall that the upper half-plane $ \mathbb{H} $ is the universal covering space of the punctured disc $ \Delta \setminus \{0\} $, the projection being given by the map $$\mathbb{H} \ni z \mapsto \exp( \iota z ) \in \Delta \setminus \{0\}.$$ Hence, for each $ q $ in $ (-1,0) $, $$d_{\Delta \setminus \{0\} } \left( p^j, q \right) = \inf_{\tilde{q}} d_{\mathbb{H}} \left( - \iota \log p^j, \tilde{q} \right),$$ where the infimum is taken over all preimages $ \tilde{q} $ of $ q $ under the covering map. Furthermore, the preimage of the interval $ (-1,0) $ is the vertical line $ \Re z = \pi $ which is a geodesic in $ \mathbb{H} $. It follows that $$d_{\Delta \setminus \{0\} } \left( p^j, (-1,0) \right) = d_{\mathbb{H}} \left( - \iota \log p^j, \{ z \in \mathbb{H} : \Re z = \pi \} \right).$$ To calculate the right-hand side, observe that there is a unique geodesic (namely, the half-circle centred at $ \pi $ and radius $ | - \iota \log p^j - \pi | $) through $ - \iota \log p^j $ and orthogonal to the line $ \{ \Re z = \pi \} $. Moreover, $ d_{\mathbb{H}} \left( - \iota \log p^j, \{ z: \Re z = \pi \} \right) $ is the distance from $ - \iota \log p^j $ to $ \{ \Re z = \pi \} $ measured along this half-circle. I.e., $$d_{\mathbb{H}} \left( - \iota \log p^j, \{ z \in \mathbb{H} : \Re z = \pi \} \right) = d_{\mathbb{H}} \left( - \iota \log p^j, \pi + \iota | \iota \log p^j + \pi | \right).$$ The Kobayashi distance between two points $ z , w $ on the upper half-plane is given by $$\label{E14}
d_{\mathbb{H}} (z,w) = \log \left( \frac{ |z - \bar{w}| + |z-w| }{|z- \bar{w}| - |z-w| }\right).$$ Using the above formulation of the Kobayashi distance on $ \mathbb{H} $, it can be seen that $$d_{\mathbb{H}} \left( - \iota \log p^j, \pi + \iota | \iota \log p^j + \pi | \right) = \log \left( - \frac{\pi}{\log p^j} + \sqrt{ \left( - \frac{\pi}{\log p^j}\right)^2 + 1} \right),$$ and consequently that, $$d_{\Delta \setminus \{0\} } \left(p^j, (-1,0) \right) = \log \left(- \frac{\pi}{\log p^j} + \sqrt{ \left( - \frac{\pi}{\log p^j}\right)^2 + 1} \right),$$ thereby verifying the equation (\[E7\]). To summarize, there is a biholomorphic imbedding $ f^j : \Delta \rightarrow \Delta \setminus \{0\} $ with $ f^j(0) = p^j $ and $$B_{ \Delta \setminus \{0\} } \left( p^j, r (p^j) \right) \subset f^j(\Delta) = \Delta \setminus (-1,0],$$ where $ r(p^j) $ is as defined by equation (\[E8\]). It follows that $$h_{\Delta \setminus \{0\} } (p^j) \leq 1/r(p^j),$$ which gives the upper estimate (\[E9\]).
For the lower estimate, the following observations will be needed. Firstly, the punctured disc is complete hyperbolic and hence taut. Moreover, for each $j $, $ h_{\Delta \setminus \{0\} } (p^j) > 0 $, and hence there exist a biholomorphic imbedding $ f^j : \Delta \rightarrow \Delta \setminus \{0\} $ with $ f^j(0) = p^j $ and $$\label{E10}
B_{ \Delta \setminus \{0\} } \left( p^j, \frac{1}{h_{\Delta \setminus \{0\} } (p^j) } \right) \subset f^j(\Delta) \subset \Delta \setminus \{0\}.$$ Secondly, consider the circle centred at the origin and radius $ p^j$, $$C^j = \{ w \in \mathbb{C} : |w| = p^j\},$$ and compute $$\sup_{ q \in C^j} d_{\Delta \setminus \{0\} } \left(p^j, q \right).$$ It turns out that $$\label{E12}
s(p^j) := \sup_{ q \in C^j} d_{\Delta \setminus \{0\} } \left(p^j, q \right) =
\log \left( 2 \left( - \frac{\pi}{\log p^j } \right)^2 + 1+ \frac{2 \pi}{\log p^j } \sqrt{ \left( - \frac{\pi}{\log p^j}\right)^2 + 1} \right)$$ Grant this for now. It follows that the circle $ C^j $ is contained in the closure of the Kobayashi ball $ B_{ \Delta \setminus \{0\} } \left( p^j, s(p^j) \right) $. This forces that $$\label{E13}
\frac{1}{h_{\Delta \setminus \{0\} } (p^j)} \leq s(p^j).$$ Indeed, assume on the contrary that the above inequality does not hold, i.e., there is an $ \epsilon_0 > 0 $ such that $$s(p^j) + \epsilon_0 < \frac{1}{h_{\Delta \setminus \{0\} } (p^j) }.$$ Then it is immediate that $$\label{E11}
C^j \subset B_{ \Delta \setminus \{0\} } \left( p^j, s(p^j) + \epsilon_0 \right) \subset
B_{ \Delta \setminus \{0\} } \left( p^j, \frac{1}{h_{\Delta \setminus \{0\} } (p^j) } \right).$$ Combining (\[E10\]) and (\[E11\]) gives $$C^j \subset f^j(\Delta) \subset \Delta \setminus \{0\}.$$ But $ f^j(\Delta) $ is a simply connected sub-domain of the punctured disc and hence it cannot contain any circle centered at the origin. Hence we arrive at a contradiction, thereby proving the inequality (\[E13\]).
The final step is to establish equation (\[E12\]). It follows from the definition that $$s(p^j) := \sup_{ q \in C^j} d_{\Delta \setminus \{0\} } \left(p^j, q \right) =
\sup_{ q \in C^j} \inf_{\tilde{q}} d_{\mathbb{H} } \left(- \iota \log p^j, \tilde{q} \right),$$ where the infimum is taken over all preimages $ \tilde{q} $ of $ q $ under the covering projection $ z \mapsto \exp (\iota z) $. Write $ q = p^j \exp ( \iota \theta) $ for $ \theta \in [0, 2\pi) $, so that the right hand side above equals $$\sup_{\theta \in [0, 2\pi)} \inf_{ k \in \mathbb{Z}} d_{\mathbb{H}} \left( - \iota \log p^j, 2 \pi k + \theta - \iota \log p^j \right)$$ A direct computation using the explicit expression (\[E14\]) for $ d_{\mathbb{H}} (\cdot, \cdot) $ shows that $$\inf_{ k \in \mathbb{Z}} d_{\mathbb{H}} \left( - \iota \log p^j, 2 \pi k + \theta - \iota \log p^j \right) =
\log \left( \frac{ \theta^2 + 2 (\log p^j)^2 + \theta \sqrt{\theta^2 + 4 (\log p^j)^2} }{2 (\log p^j)^2}\right),$$ so that $$\begin{aligned}
{3}
\sup_{\theta \in [0, 2\pi)} \inf_{ k \in \mathbb{Z}} d_{\mathbb{H}} \left( - \iota \log p^j, 2 \pi k + \theta - \iota \log p^j \right)
= & \log \left( \frac{ 4 \pi^2 + 2 (\log p^j)^2 + 2 \pi \sqrt{4 \pi^2 + 4 (\log p^j)^2} }{2 (\log p^j)^2}\right) \\
= & \log \left( 2 \left( - \frac{\pi}{\log p^j } \right)^2 + 1+ \frac{2 \pi}{\log p^j } \sqrt{ \left( - \frac{\pi}{\log p^j}\right)^2 + 1} \right),
\end{aligned}$$ thereby verifying (\[E12\]).
Note that both $L_j, U_j {\rightarrow}+\infty$ as $p^j {\rightarrow}0$ which is expected. Thus $h_{\Delta {\setminus}\{0\}}$ blows up near the origin. On the other hand, that $h_{\Delta {\setminus}\{0\}} (p^j) {\rightarrow}0$ if $\vert p^j \vert {\rightarrow}1$ can be seen from the following:
Let $D \subset {\mathbb}C$ be a bounded domain with $p^0 \in {\partial}D$. Assume that ${\partial}D$ is $C^2$-smooth near $p^0$. Then $h_D(z) {\rightarrow}0$ as $z {\rightarrow}p^0$.
Let $ \rho $ be a $ C^2$-smooth local defining function for $ \partial D $ near $ p^0 $ and $ \{ p^j \} $ be a sequence of points in $ D $ converging to $ p^0 $. Consider the dilations $$T^j(z) = \frac{z-p^j}{- \rho(p^j)}$$ and note that the scaled domains $ D^j = T^j (D)$ are given by $$\{ z \in \mathbb{C}: - 1 + 2 \Re \left( \partial \rho(p^j) z \right) - \psi(p^j) O(1) < 0 \}.$$ near $ T^j(p^0) $. It follows that the sequence of domains $ D^j $ converge in the Hausdorff sense to the half-plane $$D_{\infty} = \{ z \in \mathbb{C}: 2 \Re \left( \partial \rho(p^0) z \right) - 1 < 0 \}.$$ Note that $D$ supports a local holomorphic peak function at $p_0$ since the boundary ${\partial}D$ is $C^2$-smooth near it and hence the proof of Theorem 1.1 of [@MV] can be adapted to show that $$h_{D} (p^j) \rightarrow h_{D_{\infty}} (0).$$ But $ h_{D_{\infty}} (\cdot) \equiv 0 $ as $ D_{\infty} $ is biholomorphically equivalent to $ \mathbb{B}^n $. Hence $ h_{D} (p^j) \rightarrow 0 $ as $ j \rightarrow \infty $.
Proof of Theorem 1.1
====================
Let $p^0 \in {\partial}D$. We will study the behaviour of $h_D(z)$ as $z {\rightarrow}p^0$. The proof of Theorem 1.1 divides into two parts:
1. $ \partial D $ is of finite type near $ p^0 $, or
2. $ \partial D $ is of infinite type near $ p^0 $.
It turns out that $ p^0 \in \partial D $ cannot be of infinite type, thereby, ruling out case(ii).
[*Case (i):*]{} let $ p^j $ be a sequence of points in $ D $ converging to $ p^0 $ along the inner normal to $ \partial D $ at $ p^0 $. Since $ \lim_{j \rightarrow \infty} h_{D} (p^j) = 0 $ by assumption, there exists a sequence of positive real numbers $ R_j \rightarrow \infty $ and a sequence of biholomorphic imbeddings $ F^j: \mathbb{B}^n \rightarrow D $ satisfying $ F^j (0)= p^j $ and $ B_D(p^j, R_j) \subset F^j( \mathbb{B}^n) $.
Before going further, let us briefly recall the scaling technique from [@Mcneal-1994]. Here and in the sequel, we write $ z = (z_1, z_2, \ldots, z_{n-1}, z_n ) = ('z,z_n) \in \mathbb{C}^n $ for brevity. By [@Yu-1992] there exists a local coordinate system $ \Phi $ in a neighbourhood of $ p^0 $ such that $ \Phi(p^0) = ('0, 0) $, $
\Phi (p^j) = ('0, - \| p^0 - p^j \|) $ for each $ j$ and the domain $ \Phi(D) $ near origin can be written as $$\{ ('z,z_n) \in \mathbb{C}^n : 2 \Re z_n + P_0('z) + R(z) < 0 \},$$ where $P_0$ is a nondegenerate weighted homogeneous polynomial of degree $1$ with respect to the weights $ \mathcal{M}(\partial D,
0) $, the multitype of $ \partial D $ near the origin, and $ R $ denotes terms of degree at least two. Define a dilation of coordinates by $$T^j( z_1, z_2, \ldots, z_{n-1}, z_n ) = \left( \delta_j^{-1} z_1, \delta_j^{-1} z_2,
\ldots, \delta_j^{-1} z_{n-1}, \delta_j^{-1} z_n \right),$$ where $ \delta_j = \|p^0 - p^j \| $ for each $j$. Note that $ T^j \big( ('0, - \delta_j) \big) = ('0, -1) $ for all $j$ and the scaled domains $ D^j = T^j \circ \Phi(D) $ converge in the Hausdorff sense to $$D_{ \infty} = \big\{ z \in \mathbb C^n : 2 \Re z_n + P_{0}('z) < 0 \big\}.$$ Furthermore, it follows from Theorem 1.1 of [@Mcneal-1992] that $ D_{\infty} $ is complete hyperbolic and hence $ D_{\infty} $ is taut.
Consider the dilated maps $$\psi^j := T^j \circ \phi \circ F^j : \mathbb{B}^n \rightarrow D^j.$$ Note that $ T^j \circ \phi \circ F^j \big( ('0,0) \big) = ('0,-1)$ for each $ j $. Since the domains $ D^j $ are contained in the intersections of certain half spaces (see [@Gaussier-1997] for details), it follows that the sequence $ \{ T^j \circ \phi \circ F^j \} $ admits a subsequence, that will still be denoted by the same indices, that converges uniformly on compact sets of $ \mathbb{B}^n $ to a holomorphic mapping $ \psi : \mathbb{B}^n \rightarrow D_{\infty} $.
Then $ \psi $ is a biholomorphism from $ \mathbb{B}^n $ onto $ D_{\infty} $. To establish this, first note that for each $ \epsilon > 0 $, $$B_{D_{\infty}} \left( ('0,-1), R - \epsilon \right) \subset B_{D^j} \left( ('0,-1), R \right)
\label{k9}$$ for all $ R > 0 $ and all $j$ large and this will follow from $$\displaystyle\limsup_{j \rightarrow \infty} d_{D^j} \left( ('0,-1), \cdot \right)
\leq d_{D_{\infty}} \left( ('0,-1), \cdot \right).$$ To verify the above inequality, fix $ q \in D_{\infty}$ and let $ \gamma : [0,1]
\rightarrow D_{\infty} $ be a piecewise $C^1$-smooth path in $
D_{\infty} $ such that $ \gamma(0) = ('0,-1), \gamma(1) = q$ and $$\int_0^1 { F_{D_{\infty}} \big( \gamma(t), \dot{\gamma}(t) \big)
dt} \leq d_{D_{\infty}} \left( ('0,-1), q \right) + \epsilon/2.$$ Since the trace of $ \gamma $ is relatively compact in $
D_{\infty}$, it follows that the trace of $\gamma $ is contained uniformly relatively compactly in $ D^j
$ for all large $j$. It follows from Lemma 6.2 of [@MV] that $$\int_0^1 {F_{D^j} \big( \gamma(t), \dot{\gamma}(t) \big) dt}
\leq \int_0^1 { F_{D_{\infty}} \big( \gamma(t), \dot{\gamma}(t)
\big) dt} + \epsilon/2 \leq d_{D_{\infty}} \left( ('0,-1), q \right) + \epsilon.$$ Consequently, $$d_{D^j} \left( ('0,-1), q \right) \leq \int_0^1 {F_{D^j} \big( \gamma(t),
\dot{\gamma}(t) \big) dt} \leq d_{D_{\infty}} \left( ('0,-1), q \right) +
\epsilon$$ which implies that $$\displaystyle\limsup_{j \rightarrow \infty} d_{D^j} \left( ('0,-1), \cdot \right)
\leq d_{D_{\infty}} \left( ('0,-1), \cdot \right).$$
Note that $ B_{ D } ( p^j, R_j) \subset F^j(
\mathbb{B}^n)$. Since $ T^j \circ \phi $ are biholomorphisms and hence Kobayashi isometries, it follows that $$\begin{aligned}
B_{ D^j} \left( ('0,-1), R_j \right) \subset T^j \circ \phi \circ F^j (\mathbb{B}^n). \label{4.0}\end{aligned}$$ Since $ ( D_{\infty}, d_{D_{\infty}}) $ is complete, it is possible to write $$\begin{aligned}
D_{\infty} = \displaystyle \bigcup_{ \nu = 1} ^{\infty}
B_{D_{\infty}} \big( ('0,-1), \nu \big) \label{k10}\end{aligned}$$ which is an exhaustion of $ D_{\infty}$ by an increasing union of relatively compact domains. Now, consider $$\theta^j := \big( T^j \circ \phi \circ F^j \big)^{-1}: T^j \circ \phi \circ F^j (\mathbb{B}^n) \rightarrow
\mathbb{B}^n$$ These mappings are evidently defined on an arbitrary compact subset of $ D_{\infty} $ for large $j$ (cf. (\[k9\]), (\[4.0\]) and (\[k10\])) and hence some subsequence of $ \{ \theta^j \}$ converges to $ \theta: D_{\infty} \rightarrow
\overline{\mathbb{B}}^n$. Moreover, $ \theta \big( ('0,-1) \big) = ('0,0) $ together with the maximum principle shows that $ \theta :
D_{\infty} \rightarrow \mathbb{B}^n $. Finally observe that for $w$ in a fixed compact set in $ D_{\infty} $, $$\begin{aligned}
| \psi \circ \theta (w) - w | & = & | \psi \circ \theta (w) -
\psi^j \circ \theta^j(w) | \\
& = & | \psi \circ \theta (w) - \psi \circ \theta^j (w) | + |
\psi \circ \theta^j(w) - \psi^j \circ \theta^j (w)| \\
& \rightarrow & 0 \ \mbox{as } j \rightarrow \infty\end{aligned}$$ This shows that $ \psi \circ \theta = id $. Similarly, it can be proved that $ \theta \circ \psi = id$. This shows that $
D_{\infty} $ is biholomorphically equivalent to $ \mathbb{B}^n$.
By composing with a suitable Cayley transform, if necessary, we may assume that there is a biholomorphism $ \tilde{\theta} $ from $ D_{\infty}$ onto the unbounded realization of the ball, namely to $$\Sigma = \big\{ z \in \mathbb{C}^n : 2 \Re z_n + {\left\vertz_1\right\vert}^2 + {\left\vertz_2\right\vert}^2 + \ldots +
{\left\vertz_{n-1}\right\vert}^2 < 0 \big\}$$ with the property that the cluster set of $ \tilde{\theta} $ at some point $ ('0, \iota a) \in \partial D_{\infty} $ (for $a \in \mathbb{R} $) contains a point of $ \partial \Sigma $ different from the point at infinity on $ \partial \Sigma $. Then Theorem 2.1 of [@CP] ensures that $ \tilde{\theta} $ extends holomorphically past the boundary of $ D_{\infty} $ to a neighbourhood of $ ('0, \iota a) $. Furthermore, $ \tilde{\theta} $ extends biholomorphically across some point $ ('0, \iota a^0 ) \in \partial D_{ \infty} $. To prove this claim, it suffices to show that the Jacobian of $ \tilde{\theta} $ does not vanish identically on the complex plane $$L = \big \{ ('0, \iota a) : a \in \mathbb{R} \big \} \subset \partial D_{\infty}.$$ If the claim were false, then the Jacobian of $ \tilde{\theta} $ vanishes on the entire $ z_n $-axis, which intersects the domain $ D_{\infty} $. However, $ \tilde{\theta} $ is injective on $ D_{\infty} $, and consequently, has nowhere vanishing Jacobian determinant on $ D_{\infty} $. This contradiction proves the claim.
Next, note that the translations in the imaginary $ z_n $-direction leave $ D_{\infty} $ invariant. Therefore, we may assume that $ ('0, \iota a^0) $ is the origin and that $ \tilde{\theta} $ preserves the origin. Now recall that the Levi form is preserved under local biholomorphisms around a boundary point, thereby yielding the strong pseudoconvexity of $ \partial D_{\infty} $ near the origin. Equivalently, the strong pseudoconvexity of $ p^0 \in \partial D $ follows. Hence the result.
[*Case (ii):*]{} If $ p^0 $ were $ C^{\infty} $-convex of infinite type, by [@Z2], there exists a sequence $ p^j $ in $ D $ converging to $ p^0 \in \partial D $ and affine isomorphisms $ A^j $ of $ \mathbb{C}^n $ so that the domains $ A^j(D) = D^j $ converge in the local Hausdorff topology to a convex domain $ D_{\infty} $ and $ A^j(p^j) \rightarrow
0 \in D_{\infty}$. Moreover, it follows from Proposition 6.1 of [@Z1] that the limit domain $ D_{\infty} $ contains no complex affine lines and hence, it is Kobayashi complete.
Since $ \lim_{j \rightarrow \infty} h_{D} (p^j) = 0 $ by assumption, there exists a sequence of positive real numbers $ R_j \rightarrow \infty $ and a sequence of biholomorphic imbeddings $ F^j: \mathbb{B}^n \rightarrow D $ satisfying $ F^j (0)= p^j $ and $ B_D(p^j, R_j) \subset F^j( \mathbb{B}^n) $. Consider the maps $$\psi^j := A^j \circ F^j : \mathbb{B}^n \rightarrow D^j.$$ Note that $ A^j \circ F^j (0) \rightarrow 0 \in D_{\infty} $. By Proposition 4.2 of [@Z1], some subsequence of the sequence $ \{ A^j \circ F^j \} $ converges uniformly on compact sets of $ \mathbb{B}^n $ to a holomorphic mapping $ \psi : \mathbb{B}^n \rightarrow \overline{ D}_{\infty} $. Since $ \psi(0)= 0 $, it follows that $ \psi ( \mathbb{B}^n ) \subset D_{\infty} $.
Then $ \psi $ is a biholomorphism from $ \mathbb{B}^n $ onto $ D_{\infty} $. This will be done in several steps. The first of these records the stability of the infinitesimal Kobayashi metric, i.e., $$\label{E1}
F_{D^j} ( \cdot, \cdot) \rightarrow F_{D_{\infty}} (\cdot, \cdot)$$ uniformly on compact sets of $ D_{\infty} \times \mathbb{C}^n $. The key step in proving the above assertion is to understand limits of holomorphic mappings $ f^j : \Delta \rightarrow D^j $ that almost realize $ F_{D^j} (\cdot, \cdot) $. Using Proposition 4.2 of [@Z1], it is possible to pass to a subsequence of $ \{ f^j \} $ that converges to a holomorphic mapping $ f : \Delta \rightarrow D_{\infty} $ uniformly on compact sets of $ \Delta $. It follows that the limit map $ f $ provides a candidate in the definition of $ F_{D_{\infty}}( \cdot, \cdot) $.
The second step is to establish that $$\label{E21}
\displaystyle\limsup_{j \rightarrow \infty} d_{D^j} \left( A^j(p^j), \cdot \right)
\leq d_{D_{\infty}} \left( 0, \cdot \right),$$ which would imply that $$\label{k11}
B_{D_{\infty}} \left( 0, R - \epsilon \right) \subset B_{D^j} \left( A^j(p^j), R \right),$$ for all $ R > 0 $ and all $j$ large and for each $ \epsilon > 0 $. To verify (\[E21\]), fix $ q \in D_{\infty}$ as before and let $ \gamma : [0,1]
\rightarrow D_{\infty} $ be a piecewise $C^1$-smooth path in $
D_{\infty} $ such that $ \gamma(0) = 0, \gamma(1) = q$ and $$\int_0^1 { F_{D_{\infty}} \big( \gamma(t), \dot{\gamma}(t) \big)
dt} \leq d_{D_{\infty}} \left( 0, q \right) + \epsilon/2.$$ Define $ \gamma^j: [0,1] \rightarrow \mathbb{C}^n $ by $$\gamma^j(t) = \gamma(t) + A^j(p^j)(1-t).$$ Since the trace of $ \gamma $ is relatively compact in $
D_{\infty}$ and $ A^j (p^j) \rightarrow 0 $, it follows that the trace of $\gamma $ is contained uniformly relatively compactly in $ D^j
$ for all large $j$. Note that $ \gamma^j(0) = A^j(p^j) $ and $ \gamma^j(1) = q $. In addition, $ \gamma^j
\rightarrow \gamma $ and $ \dot{\gamma}^j \rightarrow \dot{\gamma}
$ uniformly on $ [0,1]$. Appealing to (\[E1\]) yields $$\int_0^1 {F_{D^j} \big( \gamma^j(t), \dot{\gamma}^j(t) \big) dt}
\leq \int_0^1 { F_{D_{\infty}} \big( \gamma(t), \dot{\gamma}(t)
\big) dt} + \epsilon/2 \leq d_{D_{\infty}} ( 0, q) + \epsilon.$$ Therefore, $$d_{D^j} \left( A^j(p^j), q \right) \leq \int_0^1 {F_{D^j} \big( \gamma(t),
\dot{\gamma}(t) \big) dt} \leq d_{D_{\infty}} \left( 0, q \right) +
\epsilon$$ which implies that $$\displaystyle\limsup_{j \rightarrow \infty} d_{D^j} \left( A^j(p^j), \cdot \right)
\leq d_{D_{\infty}} \left( 0, \cdot \right),$$ thereby, establishing (\[E21\]).
Next, recall that $ B_{ D } ( p^j, R_j) \subset F^j(\mathbb{B}^n) $. Since $ A^j $ are biholomorphisms and hence Kobayashi isometries, it follows that $$\begin{aligned}
B_{ D^j} \left( A^j(p^j) , R_j \right) \subset A^j \circ F^j (\mathbb{B}^n). \label{5.0}\end{aligned}$$ Since $ ( D_{\infty}, d_{D_{\infty}}) $ is complete, it is possible to write $$\begin{aligned}
D_{\infty} = \displaystyle \bigcup_{ \nu = 1} ^{\infty}
B_{D_{\infty}} \big( 0, \nu \big). \label{k13}\end{aligned}$$ Now, consider the mappings $$\theta^j := \big( A^j \circ F^j \big)^{-1}: A^j \circ F^j (\mathbb{B}^n) \rightarrow
\mathbb{B}^n.$$ It follows from (\[k13\]), (\[k11\]) and (\[5.0\]) that the mappings $ \theta^j$ are defined on any arbitrary compact subset of $ D_{\infty} $ for large $j$. In particular, $ \{\theta^j \} $ is normal. Let $ \theta: D_{\infty} \rightarrow
\overline{\mathbb{B}}^n$ be a holomorphic limit of some subsequence of $ \{ \theta^j \}$. Since $ \theta \big( 0 \big) = 0 $, it follows that $ \theta :
D_{\infty} \rightarrow \mathbb{B}^n $. The final step is to note that $ \psi \circ \theta = id $ and $ \theta \circ \psi = id$ as before. A consequence of all of this is that $
D_{\infty} $ is biholomorphically equivalent to $ \mathbb{B}^n$.
Now we seek an contradiction. If $ p^0 \in \partial D $ were $ C^{\infty} $-convex of infinite type, then the limit domain $ \partial D_{\infty} $ contains a non-trivial complex affine disc (cf. Proposition 6.1, [@Z1]) and hence, it follows from Theorem 3.1 of [@Z1] that $ \left(D_{\infty}, d_{D_{\infty}} \right) $ is not Gromov hyperbolic. Since $ \left( \mathbb{B}^n, d_{\mathbb{B}^n} \right) $ is Gromov hyperbolic and Gromov hyperbolicity is an isometric invariant, it follows that $ \left(D_{\infty}, d_{D_{\infty}} \right) $ and $ \left( \mathbb{B}^n, d_{\mathbb{B}^n} \right) $ cannot be isometric. This is a contradiction since $ D_{\infty} $ is biholomorphic to $ \mathbb{B}^n $. Hence, the boundary point $ p^0 $ has to be of finite type and we are in Case (i).
Proof of Theorem 1.2
====================
Proving Theorem 1.2 involves verifying the following two results:
\[B1\] Let $D \subset {\mathbb}C^n$ be a bounded domain with $p^0 \in \partial D$. Assume that ${\partial}D$ is $C^{\infty}$-smooth and $h$-extendible near $p^0$. Then $\partial D$ is strongly pseudoconvex near $p^0$ if $h_D(z) {\rightarrow}0$ as $z {\rightarrow}p^0$.
\[B2\] Let $D \subset {\mathbb}C^n$ be a bounded domain with $p^0 \in \partial D$. Assume that ${\partial}D$ is $C^{\infty}$-smooth and $h$-extendible near $p^0$. Then $\partial D$ is strongly pseudoconvex near $p^0$ if $s_D(z) {\rightarrow}1$ as $z {\rightarrow}p^0$.
*Proof of Proposition \[B1\].* Let $ p^j $ be a sequence of points in $ D $ converging to $ p^0 $ along the inner normal to $ \partial D $ at $ p^0 $. The boundary point $ p^0 $ is a local peak point by [@Yu-1994] and hence the Fridman’s invariant function can be localized near $ p^0 $ (refer Proposition 3.4 of [@MV]). It follows that $ \lim_{j \rightarrow \infty} h_{U \cap D} (p^j, \mathbb{B}^n ) = 0 $ for any neighbourhood $ U $ of $ z^0 $. As a consequence, there exists a sequence of positive real numbers $ R_j \rightarrow \infty $ and a sequence of biholomorphic imbeddings $ F^j: \mathbb{B}^n \rightarrow U \cap D $ satisfying $ F^j (0)= p^j $ and $ B_{U \cap D}(p^j, R_j) \subset F^j( \mathbb{B}^n) $.
Let us quickly recall the local geometry of $h$-extendible domains. Firstly, if $ \mathcal{M}(\partial D, p^0) = (1, m_2, \ldots, m_n) $ denotes the Catlin’s multitype of $ \partial D $ near $ p^0 $, then there is an automorphism $ \Phi $ of $ \mathbb{C}^n $ such that such that $ \Phi(p^0) = ('0, 0) $, $ \Phi (p^j) = ('0, - \| p^0 - p^j \|) $ for each $ j$ and the defining function for the domain $ \Phi(D) $ can be expanded near the origin as $$\begin{aligned}
{3}
2 \Re z_n + P('z, ' \overline{z}) + R(z), \end{aligned}$$ where $ P('z, ' \overline{z}) $ is a $ (1/m_n , 1/m_{n-1} , \ldots , 1/m_2 ) $ homogeneous polynomial of weight one which is plurisubharmonic and does not contain pluriharmonic terms and $ R $ satisfies $$|R(z)| \lesssim \left( |z_1 |^{m_n} + |z_2 |^{m_{n-1}} + \ldots + |z_n | \right)^{\gamma}$$ for some $ \gamma > 1 $. Let $ T^j $ be the dilation defined by $$T^j( z_1, z_2, \ldots, z_{n-1}, z_n ) = \left( \delta_j^{-1/{m_n}} z_1, \delta_j^{-1/{m_{n-1}}} z_2, \ldots, \delta_j^{-1/m_2} z_{n-1}, \delta_j^{-1} z_n \right),$$ where $ \delta_j = \|p^0 - p^j \| $ for each $j$. Note that $ T^j \big( ('0, - \delta_j) \big) = ('0, -1) $ for all $j$ while the domains $ D^j := T^j \circ \Phi(U \cap D) $ converge in the Hausdorff sense to $$D_{ \infty} = \big\{ z \in \mathbb C^n : 2 \Re z_n + P('z, ' \overline{z}) < 0 \big\}.$$ On the other hand, by Theorem 4.7 of [@Yu-1995], there is an $h$-extendible model $$\Omega_0 = \big\{ z \in \mathbb{C}^n : 2 \Re z_n + Q('z, ' \overline{z}) < 0 \big\},$$ where $ Q('z, ' \overline{z}) $ is a $ (1/m_n , 1/m_{n-1} , \ldots , 1/m_2 ) $ homogeneous polynomial (of weight one which is plurisubharmonic but not pluriharmonic), such that if $ U $ is a small neighbourhood of $ p^0 $, then $$\Phi(U \cap D) \subset \Omega_0.$$ Consequently, the scaled domains $ D^j $ satisfy $$\begin{aligned}
D^j = T^j \circ \Phi(U \cap D) \subset T^j(\Omega_0)\end{aligned}$$ for all $j $ large. The homogenity of $ Q('z, ' \overline{z}) $ with weight one as described above implies that $$Q \left( \delta_j^{1/{m_n}} z_1, \delta_j^{1/{m_{n-1}}} z_2,
\ldots, \delta_j^{1/m_2} z_{n-1} \right) = \delta_j Q ('z, ' \overline{z} )$$ for each $j $. In other words, the mappings $ T^j $ leave $ \Omega_0 $ invariant and hence $$\begin{aligned}
\label{E15}
D^j = T^j \circ \Phi(U \cap D) \subset \Omega_0.\end{aligned}$$ Furthermore, observe that $ D_{\infty} $ and $ \Omega_0 $ are complete hyperbolic (cf. [@Yu-1994]) and hence both $ D_{\infty} $ and $ \Omega_0 $ are taut.
Let us consider the holomorphic mappings $$\psi^j := T^j \circ \phi \circ F^j : \mathbb{B}^n \rightarrow D^j.$$ Note that $ T^j \circ \phi \circ F^j \big( ('0,0) \big) = ('0,-1)$ for each $ j $. As the scaled domains $ D^j $ are contained in a taut domain $ \Omega_0 $, the sequence $ \{ T^j \circ \phi \circ F^j \} $ admits a subsequence, that will still be denoted by the same indices, that converges uniformly on compact sets of $ \mathbb{B}^n $ to a holomorphic mapping $ \psi : \mathbb{B}^n \rightarrow D_{\infty} $.
Here, too, it turns out that $ \psi $ is a biholomorphism from $ \mathbb{B}^n $ onto $ D_{\infty} $. To establish this, first note that $$\begin{aligned}
\label{E4}
F_{D^j} ( \cdot, \cdot) \rightarrow F_{D_{\infty}} ( \cdot, \cdot) \end{aligned}$$ uniformly on compact sets of $ D_{\infty} \times \mathbb{C}^n $. The above assertion can be proved using the fact that the scaled domains $ D^j $ are all contained in $ \Omega_0 $ for large $ j $ (cf. (\[E15\])). Once the stability of the infinitesimal metric is understood (i.e., (\[E4\])), by similar arguments as in the proof of Theorem 1.1(i), it follows that for each $ \epsilon > 0 $, $$B_{D_{\infty}} \left( ('0,-1), R - \epsilon \right) \subset B_{D^j} \left( ('0,-1), R \right)
\label{E2}$$ for all $ R > 0 $ and all $j$ large.
Recall that $ B_{U \cap D } ( p^j, R_j) \subset F^j(
\mathbb{B}^n)$. Since $ T^j \circ \phi $ are biholomorphisms and hence Kobayashi isometries, it follows that $$\begin{aligned}
B_{ D^j} \left( ('0,-1), R_j \right) \subset T^j \circ \phi \circ F^j (\mathbb{B}^n). \label{E5}\end{aligned}$$ Exploiting the Kobayashi completeness of the limit domain $ D_{\infty}$, it is possible to write $$\begin{aligned}
D_{\infty} = \displaystyle \bigcup_{ \nu = 1} ^{\infty}
B_{D_{\infty}} \big( ('0,-1), \nu \big). \label{E6}\end{aligned}$$ Now, consider the mappings $$\theta^j := \big( T^j \circ \phi \circ F^j \big)^{-1}: T^j \circ \phi \circ F^j (\mathbb{B}^n) \rightarrow
\mathbb{B}^n$$ defined on an arbitrary compact subset of $ D_{\infty} $ for large $j$ (cf. (\[E2\]), (\[E5\]) and (\[E6\])) and hence some subsequence of $ \{ \theta^j \}$ converges to $ \theta: D_{\infty} \rightarrow
\overline{\mathbb{B}}^n$. Moreover, $ \theta \big( ('0,-1) \big) = ('0,0) $ together with the maximum principle shows that $ \theta :
D_{\infty} \rightarrow \mathbb{B}^n $. Then it can be checked that $ \psi \circ \theta = id $ and $ \theta \circ \psi = id$ and hence $
D_{\infty} $ is biholomorphically equivalent to $ \mathbb{B}^n$.
Finally, we are ready to prove the theorem. Using similar arguments as in the proof of Theorem 1.1(i), it is possible to show that there is a biholomorphism $ \tilde{\theta} $ from $ D_{\infty}$ onto the unbounded realization of the ball, namely to $$\Sigma = \big\{ z \in \mathbb{C}^n : 2 \Re z_n + {\left\vertz_1\right\vert}^2 + {\left\vertz_2\right\vert}^2 + \ldots +
{\left\vertz_{n-1}\right\vert}^2 < 0 \big\},$$ with the property that $ \tilde{\theta} $ extends biholomorphically past the boundary of $ D_{\infty} $ to a neighbourhood of the origin. Also, $ \tilde{\theta} \big( ('0,0) \big) =
\big( ('0,0) \big) $. Since the Levi form is preserved under local biholomorphisms around a boundary point, it follows that $ \partial D_{\infty} $ must be strongly pseudoconvex near the origin. In particular, $ m_2= \ldots = m_n =2 $ and $ P('z, ' \overline{z}) = |z_1|^2 + \ldots + |z_n|^2 $ which gives the strong pseudoconvexity of $ p^0 \in \partial D $. Hence the result.
*Proof of Proposition \[B2\].* Let $ p^j $ be a sequence of points in $ D $ converging to $ p^0 $ along the inner normal to $ \partial D $ at $ p^0 $. Since $ \lim_{j \rightarrow \infty} s_{D} (p^j) = 1 $ by assumption, there exists a sequence of positive real numbers $ R_j \rightarrow 1 $ and a sequence of biholomorphic imbeddings $ F^j: D \rightarrow \mathbb{B}^n $ satisfying $ F^j(p^j)=0 $ and $ {B}^n(0, R_j) \subset F^j( D) $.
Let us adapt here the method of uniform scaling. Let $ U $ be a neighbourhood of $ p^0 \in \partial D $ and the mappings $ T^j \circ \Phi $ be as described in the the proof of Proposition \[B1\]. So that the domains $ D^j = T^j \circ \Phi (U \cap D) $ converge in the Hausdorff sense to $$D_{ \infty} = \big\{ z \in \mathbb C^n : 2 \Re z_n + P('z, ' \overline{z}) < 0 \big\}.$$ Here, $ P('z, ' \overline{z}) $ is a $ (1/m_n , 1/m_{n-1} , \ldots , 1/m_2 ) $ homogeneous polynomial of weight one that coincides with the polynomial of same degree in the homogeneous Taylor expansion of the defining function for $ \Phi(D) $ near the origin and $ (1, m_2, \ldots, m_n) $ is the Catlin’s multitype of $ \partial D $ near $ p^0 $.
Consider the maps $$\psi^j := F^j \circ \left( T^j \circ \phi \right)^{-1} : D^j \rightarrow F^j(D) \subset \mathbb{B}^n.$$ Note that $ F^j \circ \left( T^j \circ \phi \right)^{-1} \big( ('0,-1) \big) = ('0,0)$ for each $ j $. These mappings are defined on an arbitrary compact subset of $ D_{\infty} $ for large $j$ and hence some subsequence of $ \{ \psi^j \}$ converges to $ \psi: D_{\infty} \rightarrow
\overline{\mathbb{B}}^n$. Since $ \psi \big( ('0,-1) \big) = ('0,0) $, it follows that $ \psi (D_{\infty}) \subset \mathbb{B}^n $.
The claim is that $ \psi $ is a biholomorphism from $ D_{\infty} $ onto $ \mathbb{B}^n $. To begin with, let $ K $ be an arbitrary compact set of $ \mathbb{B}^n $ containing the origin. Since $ {B}^n (0, R_j) \subset F^j(D) $ and $ R_j \rightarrow 1 $, the mappings $ (F^j)^{-1} $ are defined on $ K $ for $ j $ large. Since $ (F^j)^{-1} \big(('0,0) \big) = p^j \rightarrow p^0 \in \partial D $ and the boundary point $ p^0 $ is a local peak point, it follows from the attraction property (see for instance Lemma 2.1.1 of [@Gaussier-1999]) that $$(F^j)^{-1} (K) \subset U \cap D$$ for all $ j $ large. Therefore $$\left( T^j \circ \phi \right) \circ \left(F^j \right)^{-1} (K) \subset \left( T^j \circ \phi \right) (U \cap D) = D^j$$ and hence the scaling sequence $$\theta^j := \left( T^j \circ \phi \right) \circ \left(F^j \right)^{-1}$$ maps $ K $ injectively into $ D^j $ for all $ j $ large. On the other hand, by (\[E15\]), there is a taut domain $ \Omega_0 $ such that $ D^j \subset \Omega_0 $ for all $j $ large. It follows that $ \{ T^j \circ \phi \circ \left(F^j \right)^{-1}|_K \} $ forms a normal family. Let $ \theta : K \rightarrow
\overline{D}_\infty $ be a holomorphic limit of some subsequence of $ \{ T^j \circ \phi \circ \left(F^j \right)^{-1} \} $. Since $ K $ is an arbitrary compact subset of $ \mathbb{B}^n $, $ \theta $ is defined on whole of $ \mathbb{B}^n $. As noted earlier, $ T^j \circ \phi \circ \left(F^j \right)^{-1}
\big( ('0,0) \big) = ('0,-1)$ for each $ j $ and so $ \theta \big( ('0,0) \big) = ('0, -1) $. But $ ('0, -1) \in D_{\infty} $ and $ D_{\infty} $ is open, so $ \theta (\mathbb{B}^n ) \subset D_{\infty} $.
Now an argument similar to the one employed in Theorem 1.1 shows that $ \psi \circ \theta = id $ and $ \theta \circ \psi = id$, which in turn implies that $ D_{\infty} $ is biholomorphically equivalent to $ \mathbb{B}^n$. As in the proof of Proposition \[B1\], the strong pseudoconvexity of $ p^0 \in \partial D $ follows.
[*Concluding Remarks:*]{} While the construction of $h_D$ and $s_D$ are completely dual to each other, the exact relation between them is unclear from their definitions. It would be interesting and useful to clarify this.
The ratio of the Carathéodory–Eisenmann and the Kobayashi volume forms is another biholomorphic invariant that has been studied in the past. It is known that it is at most $1$ everywhere and that if it equals $1$ at an interior point, then the domain is biholomorphic to ${\mathbb}B^n$. Is there an analog of this result for the ratio $s_D/h_D$?
[*Acknowledgements:*]{} The authors would like to thank A. Zimmer for pointing out a gap in an earlier version of this article.
[XYZ]{}
Alexander H.: *Extremal holomorphic imbeddings between the ball and polydisc*. Proc. Amer. Math. Soc. **68** (1978), no. 2, 200–202.
Coupet B., Pinchuk S.: *Holomorphic equivalence problem for weighted homogeneous rigid domains in $ \mathbb{C}^{n+1} $*. (Russian) Complex Analysis in Modern Mathematics, FAZIS, Moscow (2001), 57–70.
Fornaess, J. E.; Wold, E. F.: *A non-strictly pseudoconvex domain for which the squeezing function tends to one towards the boundary*, available at [https://arxiv.org/pdf/1611.04464.pdf]{}, to appear in the Pacific Journal of Mathematics.
Fridman, B. L.: *Biholomorphic invariants of a hyperbolic manifold and some applications*. Trans. Amer. Math. Soc. [**276**]{} (1983), no. 2, 685–698.
Fridman, B. L.: *Imbedding of a strictly pseudoconvex domain in a polyhedron*. (Russian) Dokl. Akad. Nauk SSSR, [**249**]{} (1979), no. 1, 63–67.
Gaussier H.: *Characterization of convex domains with noncompact automorphism group*. Michigan Math. J. **44**(1997), 375–388.
Gaussier H.: *Tautness and complete hyperbolicity of domains in $ \mathbb{C}^n $*. Proc. Am. Math. Soc. **127**(1999), 105–116.
Kim, K.T., Zhang, L.: *On the uniform squeezing property of bounded convex domains in ${\mathbb}C^n$*. Pacific J. Math. 282 (2016), no. 2, 341–358.
Mahajan, P.; Verma, K.: *Some aspects of the Kobayashi and Carathéodory metrics on pseudoconvex domains*. J. Geom. Anal. [**22**]{} (2012), no. 2, 491–560.
McNeal J. D.: *Convex domains of finite type*. J. Funct. Anal. **108**(1992), 361–373.
McNeal J. D.: *Estimates on the Bergman kernels of convex domains*. Adv. Math. **109**(1994), 108–139.
Nikolov, N.: *Behavior of the squeezing function near h-extendible boundary points*. Proc. Amer. Math. Soc. [**146**]{} (2018), no. 8, 3455–3457.
Nikolov, N.; Andreev, L.: *Boundary behavior of the squeezing functions of ${{\mathbb}C}$-convex domains and plane domains*. Internat. J. Math. [**28**]{} (2017), no. 5, https://doi.org/10.1142/S0129167X17500318, 5 pp.
Joo S., Kim K.T.: *On boundary points at which the squeezing function tends to one*. J. Geom. Anal.(2017), https://doi.org/10.1007/s12220-017-9910-4
Yu J.: *Multitypes of convex domains*. Indiana Univ. Math. J. **41**(1992), 837–849.
Yu J.: *Peak functions on weakly pseudoconvex domains*. Indiana Univ. Math. J. **43**(1994), 1271–1295.
Yu J.: *Weighted boundary limits of the generalized Kobayashi-Royden metrics on weakly pseudoconvex domains*. Trans. Amer. Math. Soc. **347**(1995), 587–614.
Zimmer, A.: *Gromov hyperbolicity and the Kobayashi metric on convex domains of finite type*. Math. Ann. [**365**]{} (2016), no. 3–4, 1425–1498.
Zimmer, A.: *A gap theorem for the complex geometry of convex domains*. Trans. Amer. Math. Soc. (2018), https://doi.org/10.1090/tran/7284.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The states of asymptotic relaxation of $2$-dimensional fluids and plasma present a high degree of regularity and obey to the $\sinh $-Poisson equation. We find that embedding the classical fluid description into a field-theoretical framework, the same equation appears as a manifestation of the self-duality.'
---
[**Self-duality of the asymptotic relaxation states of fluid and plasmas**]{}\
\
[*Association Euratom-MER Romania\
National Institute of Laser, Plasma and Radiation Physics, MG-36 Magurele,\
Bucharest, Romania*]{}\
(*e-mail: [email protected], [email protected]*)
The states generated by externally driving (stirring) an ideal fluid can have very irregular form. It is however known from experiments and numerical simulations that after suppressing the drive the system evolves to states with high degree of order, essentially consisting of few large vortices, with very regular geometry. These states are attained after long time evolution and are not due to the residual dissipation. The process consists of vortex merging, which is an essentially topological event where the weak dissipation only allows the reconnection of the field lines but does not produce significant energy loss from the fluid motion. Inferring from results of numerical simulations, Montgomery *et al.* [@Montg1], [@Montgomery] have proved that the scalar stream function $\psi $ describing the motion in two-dimensional space obeys in the far asymtotic regime (where the regular structures are dominant) the *sinh*-Poisson equation $$\Delta \psi +\gamma \sinh \left( \beta \psi \right) =0 \label{sinhP}$$ where $\gamma $ and $\beta $ are *positive* constants. The relations of $\psi $ to the velocity and vorticity are $\mathbf{v=\nabla }\psi \times
\widehat{\mathbf{e}}_{z}$ , $\mathbf{\omega }=\mathbf{\nabla }\times \mathbf{v=-}\nabla ^{2}\psi \widehat{\mathbf{e}}_{z}$ where $\widehat{\mathbf{e}}_{z} $ is the unitary vector perpendicular to the plane. With these variables, the Euler equations for the two dimensional ideal incompressible fluid are $$\begin{aligned}
\frac{\partial \mathbf{\omega }}{\partial t}+\left( \mathbf{v\cdot \nabla }\right) \mathbf{\omega } &=&\mathbf{0} \label{Euler} \\
\mathbf{\nabla \cdot v} &=&0 \notag\end{aligned}$$
We will try to develop a field-theoretical model of the stationary asymptotic relaxed states, *i.e.* we look for a model that could provide a derivation of Eq.(\[sinhP\]). This will be done progressively, examining models ellaborated for closely related problems and collecting the relevant suggestions that could allow us to write a Lagrangian density.
In the study of the two-dimensional Euler fluids, and in particular in explaining the origin of Eq.(\[sinhP\]), an important model consists of a system of $N$ discrete vorticity filaments perpendicular on plane, having circular transversal section of radius $a$ and carrying the vorticity $\omega _{i}$, $i=1,N$. A very comprehensive account of this system is given by Kraichnan and Montgomery [@KraichnanMontgomery] where the correspondence between the continuous and discrete representations of vorticity is discussed in detail. The motion in plane of the $k$-th filament of coordinates $\mathbf{r}_{k}\equiv \left( r_{k}^{1},r_{k}^{2}\right)
\equiv \left( x_{k},y_{k}\right) $ is given by $$\frac{dr_{k}^{i}}{dt}=\varepsilon ^{ij}\frac{\partial }{\partial r_{k}^{j}}\sum_{n=1,n\neq k}^{N}\omega _{n}G\left( \mathbf{r}_{k}-\mathbf{r}_{n}\right) \;,\;i,j=1,2\;,\;k=1,N \label{motion}$$ where the summation is over all the other filaments’ positions $\mathbf{r}_{n}$, $n\neq k$, and $\varepsilon ^{ij}$ is the antisymmetric tensor in two dimensions. As shown in Ref.[@KraichnanMontgomery] $G\left( \mathbf{r}_{k}-\mathbf{r}_{n}\right) $ can be approximated for $a$ small compared to the space extension of the fluid, $L$, $a\ll L$, as the Green function of the Laplacian $$G\left( \mathbf{r},\mathbf{r}^{\prime }\right) \approx -\frac{1}{2\pi }\ln
\left( \frac{\left| \mathbf{r}-\mathbf{r}^{\prime }\right| }{L}\right)
\label{Green}$$
The statistical properties of the system of discrete vortices have been examined using the Liouville theorem and the conservation of energy and momentum . The model consists of an equal number of positive and negative vortices with equal absolute magnitudes $\left| \omega _{i}\right| =\left|
\omega \right| $, in contact with a thermal bath of temperature $T$. The possibility of the generation of two supervortices of opposite signs arises when $T$ is [*n*egative]{} and sufficiently large. When the *most probable state* is attained for a stationary configuration the stream function $\psi $ is shown to verify the $\sinh $-Poisson equation (\[sinhP\]). These statistical considerations remain the reference explanation for the appearence of this equation in this context.
Highly ordered states have been obtained by numerical simulations for other systems in two spatial dimensions: systems of filaments of electric current in $2$-dimensional MHD, guiding centre particles. The asymptotic states are found to be described by scalar functions obeying Liouville equation or $\sinh $-Poisson equation. The theoretical approach which permitted to explain this result was based on the extremum of the entropy over a statistical ensemble of states of a set of discrete objects under the constraints which express the conservation of the invariants of motion and with a negative temperature.
In the equations of motion (\[motion\]) the right hand side contains the *curl* of the Laplacian Green’s function (\[Green\]) (we take $L=1$) $$-\varepsilon ^{ij}\partial _{j}G\left( \mathbf{r,r}^{\prime }\right)
=\varepsilon ^{ij}\partial _{j}\frac{1}{2\pi }\ln r=\frac{1}{2\pi }\varepsilon ^{ij}\frac{r^{j}}{r^{2}} \label{avect}$$ $$\mathbf{\nabla }^{2}\frac{1}{2\pi }\ln r=\delta ^{2}\left( r\right)$$ The term in right hand side of (\[motion\]) can be seen as a vector potential $\mathbf{a}\left( \mathbf{r},t\right) $. The “magnetic” field $\mathbf{\nabla \times a}$ is a sum of Dirac $\delta $ functions at the locations of the vortices. If we take equal strength $\omega $ for all vortices this potential appears as the “statistical potential” and has a topological interpretation [@0010246]. From (\[avect\]) it can be rewritten $$\frac{1}{2\pi }\varepsilon ^{ij}\frac{r^{j}}{r^{2}}=-\frac{1}{2\pi }\partial
_{i}\arctan \frac{y}{x}=-\frac{1}{2\pi }\partial _{i}\theta \label{statist}$$ where $\mathbf{r=}\left( x,y\right) =\left( r\cos \theta ,r\sin \theta
\right) $. The “magnetic” flux through a surface limitted by a large circle is proportional with the number of vortices. The topological nature of this potential suggests that it can be naturally derived in a topological framework, *i.e.* from a Lagrangian density of the Chern-Simons type, $$\mathcal{L}=\frac{1}{4}\varepsilon ^{\mu \alpha \beta }A_{\nu }F_{\alpha
\beta } \label{CSdens}$$ where $F_{\alpha \beta }=\partial _{\alpha }A_{\beta }-\partial _{\beta
}A_{\alpha }$. In the analysis for the two-dimensional Euler equation (as in the investigation of similar systems), the vorticity-type functions $\mathbf{\omega }$ are expressed as the Lapacian operator applied on a scalar function and this naturally invokes the Green function of the Laplacian. Then this approach suggests that the intrinsically determined motion of the fluid can be projected onto two distinct parts of a new model: objects (vortices) and interaction between them (the potential obtained from the Green function). The fact that the interaction potential can be derived from a Chern-Simons topological action suggests that, in order to embedd the original Euler fluid system into a larger field theoretical context, we need (a) a “matter” part in the Lagrangian, which should provide the free dynamics of the vortices; (b) the Chern-Simons term, to describe the free dynamics of the field; (c) the interaction term of the (Chern-Simons-) field and the matter. The main requirement to this field-theoretical extension of the original Euler fluid model is to reproduce the discrete vortices and their equation of motion.
Jackiw and Pi [@JackiwPi], [@JakPi90] have examined a model of $N$ interacting particles (of charges $e_{s}$) moving in plane described by the Lagrangian $$L=\sum_{s=1}^{N}\frac{1}{2}m_{s}\mathbf{v}_{s}^{2}+\frac{1}{2}\int
d^{2}r\varepsilon ^{\alpha \beta \gamma }\left( \partial _{\alpha }A_{\beta
}\right) A_{\gamma }-\int d^{2}rA_{\mu }j^{\mu } \label{LagJP}$$ where $m_{s}\mathbf{v}_{s}=\mathbf{p}_{s}-e_{s}\mathbf{A}\left( \mathbf{r}_{s}|\mathbf{r}_{1},\mathbf{r}_{2},...,\mathbf{r}_{N}\right) $, $\mathbf{A}\left( \mathbf{r}_{s}|\mathbf{r}_{1},\mathbf{r}_{2},...,\mathbf{r}_{N}\right) \equiv (a_{s}^{i}\left( \mathbf{r}_{1},\mathbf{r}_{2},...,\mathbf{r}_{N}\right) _{i=1,2}$ and $a_{s}^{i}\left( \mathbf{r}_{1},\mathbf{r}_{2},...,\mathbf{r}_{N}\right) =\frac{1}{2\pi }\varepsilon ^{ij}\sum_{q\neq
s}^{N}e_{q}\frac{r_{s}^{j}-r_{q}^{j}}{\left| \mathbf{r}_{s}-\mathbf{r}_{q}\right| ^{2}}$. The matter current is defined as $j^{\mu }\equiv \left(
\rho ,\mathbf{j}\right) =\sum_{s=1}^{N}e_{s}v_{s}\delta \left( \mathbf{r-r}_{s}\right) $, $v_{s}^{\mu }=(1,\mathbf{v}_{s})$ with the metric $\left(
1,-1,-1\right) $. By varying the action we obtain the following equations $$\frac{1}{2}\varepsilon ^{\alpha \beta \gamma }F_{\alpha \beta }=\varepsilon
^{\gamma \alpha \beta }\partial _{\alpha }A_{\beta }=j^{\gamma }$$ $$B=-\rho \label{beqrho}$$ $$E^{i}=\varepsilon ^{ij}j^{j} \label{ej}$$ Here $B=\varepsilon ^{ij}\partial _{j}A_{i}$. In these equations the potential appears as being generated by the matter, *i.e.* by the “current” of particles. We note that the Eq.(\[beqrho\]) connects the matter density with the curl of the potential. This is important since in order to derive it in the context of the quantum version of their model, Jackiw and Pi have shown that one needs to include a nonlinear self-interaction of the wave function representing the matter field. The self-interaction of the scalar matter field is of the type $\varphi ^{4}$. The canonical momentum expression becomes the covariant derivative in the field theory version. Even at this point where we only have some hints, it is suggestive to look for possible identifications between the fluid variables and the field theoretical variables of the model of Jackiw and Pi. The fluid variables are: the scalar potential $\psi $, the velocity $v^{i}=\varepsilon ^{ij}\partial _{j}\psi $ and the vorticity $\omega
=\varepsilon ^{ij}\partial _{i}v_{j}$. In order to check the possibility that the fluid model can be embedded into the field theory just described we identify $$\omega \equiv \rho =\Psi ^{\ast }\Psi \label{omegarho}$$ $$v^{i}\equiv A^{i} \label{va}$$ Consider the relation between the vector potential $\mathbf{A}$ and the density $\rho $, obtained from Eq.(\[beqrho\]) $$\partial ^{i}\partial _{i}A^{j}=-\varepsilon ^{jk}\partial _{k}\rho$$ or, writting symbolically the Green function of the Laplace operator for argument $\mathbf{r-r}^{\prime }$ as the inverse of the operator at the left, we have $$A^{j}\left( \mathbf{r},t\right) =\varepsilon ^{jk}\partial _{k}\int
d^{2}r^{\prime }\left( -\partial ^{i}\partial _{i}\right) _{rr^{\prime
}}^{-1}\rho \left( \mathbf{r}^{\prime },t\right) \label{Arho}$$ Since $\omega =-\nabla ^{2}\psi $ (where $\nabla ^{2}\equiv \partial
^{i}\partial _{i}$) we have formally $\psi =-\left( \nabla ^{2}\right) ^{-1}$ $\omega $ and use this formula to calculate the right side term of (\[Arho\]) taking into account the identification (\[omegarho\]) $$\varepsilon ^{jk}\partial _{k}\int d^{2}r^{\prime }\left( -\partial
^{i}\partial _{i}\right) _{rr^{\prime }}^{-1}\omega \left( \mathbf{r}^{\prime }\right) =\varepsilon ^{jk}\partial _{k}\psi =v^{j}$$ which confirms (\[va\]). It results that if we assume the identification (\[omegarho\]) the equation (\[beqrho\]) obtained form the Lagrangian is precisely the *definition* of the vorticity vector.
One important hint from the model of Jackiw and Pi is the idea to represent a classical quantity as the modulus of a fictitious complex scalar field, as in Eq.(\[omegarho\]) and derive dynamical equations from a Lagrangian density expressed in terms of this field. The substitution of the dynamics expressed in terms of usual mechanical quantities by the richer dynamics of the amplitude and phase of the complex scalar field is an example of embedding of one theory into a larger framework and relies on the example of quantum mechanics.
In Ref.([@Nardelli]) the $\sinh $-Poisson equation is derived in a field theoretical model where instead of the topological coupling two complex scalar fields are considered. In that model the dynamics of the two scalar fields is independent except that the self-interaction depends on both. Combining this suggestion with the one exposed in the previous paragraph, we will take the density $\rho $ (which we identify with the local value of the vorticity $\omega $ as in Eq.(\[omegarho\]) ) as a two-component complex field $\Psi $ $$\rho \sim \left[ \Psi ^{\dagger },\Psi \right]$$ where $\Psi =\left(
\begin{array}{c}
\phi _{1} \\
\phi _{2}
\end{array}
\right) $. Keeping the same structure of the Lagrangian density as in the previous example but extending to a non-Abelian $SU\left( 2\right) $ gauge field we have [@Dunne], [@JakPi90] $$\mathcal{L}=-\varepsilon ^{\mu \nu \rho }tr\left( \partial _{\mu }A_{\nu
}A_{\rho }+\frac{2}{3}A_{\mu }A_{\nu }A_{\rho }\right) +itr\left( \Psi
^{\dagger }D_{0}\Psi \right) -\frac{1}{2}tr\left( \left( D_{i}\Psi \right)
^{\dagger }D_{i}\Psi \right) +\frac{1}{4}tr\left( \left[ \Psi ^{\dagger
},\Psi \right] \right) ^{2} \label{LagSU2}$$ where the potantial $A$ takes values in the algebra of the group $SU\left(
2\right) $. This form incorporates and adapt all the suggestions from the previous models: (a) the first term is the general non-Abelian expression of the Chern-Simons term; (b) it uses the covariant derivatives ($\mu =0,1,2$, $i=1,2$) for the minimal coupling $$D_{\mu }\Psi =\partial _{\mu }\Psi +\left[ A_{\mu },\Psi \right]$$ (c) Finally, it includes a scalar self-interaction of $\varphi ^{4}$ type. Analogous to the case treated by Jackiw and Pi for the $U\left( 1\right) $ gauge field (\[LagJP\]), in Ref.[@Dunne] it is shown that the Hamiltonian density corresponding to the Lagrangian density (\[LagSU2\]) is $$\mathcal{H}=\frac{1}{2}tr\left( \left( D_{i}\Psi \right) ^{\dagger }\left(
D_{i}\Psi \right) \right) -\frac{1}{4}tr\left( \left[ \Psi ^{\dagger },\Psi
\right] ^{2}\right) \label{HamiltSU2}$$ since the Chern-Simons term does not contribute to the energy density (being first order in the time derivatives). The equations of motion are $$iD_{0}\Psi =-\frac{1}{2}\mathbf{D}^{2}\Psi -\frac{1}{2}\left[ \left[ \Psi
,\Psi ^{\dagger }\right] ,\Psi \right] \label{ec1}$$ $$F_{\mu \nu }=-\frac{i}{2}\varepsilon _{\mu \nu \rho }J^{\rho } \label{ec2}$$ Using the notation $D_{\pm }\equiv D_{1}\pm iD_{2}$ the first term in (\[HamiltSU2\]) can be written $$tr\left( \left( D_{i}\Psi \right) ^{\dagger }\left( D_{i}\Psi \right)
\right) =tr\left( \left( D_{-}\Psi \right) ^{\dagger }\left( D_{-}\Psi
\right) \right) +\frac{1}{2}tr\left( \Psi ^{\dagger }\left[ \left[ \Psi
,\Psi ^{\dagger }\right] ,\Psi \right] \right)$$ The last term comes from Eq.(\[ec2\]) and from the definition $J^{0}=\left[
\Psi ,\Psi ^{\dagger }\right] $. Then the energy density is $$\mathcal{H}=\frac{1}{2}tr\left( \left( D_{-}\Psi \right) ^{\dagger }\left(
D_{-}\Psi \right) \right) \geq 0 \label{Bogo}$$ and the Bogomol’nyi inequality is saturated at *self-duality* $$D_{-}\Psi =0 \label{sd1}$$ $$\partial _{+}A_{-}-\partial _{-}A_{+}+\left[ A_{+},A_{-}\right] =\left[ \Psi
,\Psi ^{\dagger }\right] \label{sd2}$$ The first equation results from the minimum in Eq.(\[Bogo\]) and the second is actually Eq.(\[ec2\]) with the definitions $$\begin{aligned}
J^{0} &=&\left[ \Psi ^{\dagger },\Psi \right] \\
J^{i} &=&-\frac{i}{2}\left( \left[ \Psi ^{\dagger },D_{i}\Psi \right] -\left[
\left( D_{i}\Psi \right) ^{\dagger },\Psi \right] \right)\end{aligned}$$ The *static* solutions of the *self-duality* equations (\[sd1\], \[sd2\]) are derived in [@Dunne], using the algebraic *ansatz*: $$A_{i}=\sum_{a=1}^{r}A_{i}^{a}H_{a}$$ $$\Psi =\sum_{a=1}^{r}\psi ^{a}E_{a}+\psi ^{M}E_{-M}$$ where $H_{a}$ are the Cartan subalgebra generators for the gauge Lie algebra, $E_{a}$ are the simple-root step operators and $E_{-M}$ is the step operator corresponding to minus the maximal root [@Sattinger]. The rank of the algebra is noted $r$, and $r=1$ for $SU\left( 2\right) $. Then $$\left[ \Psi ^{\dagger },\Psi \right] =\sum_{a=1}^{r}\left| \psi ^{a}\right|
^{2}H_{a}+\left| \psi ^{M}\right| ^{2}H_{-M} \label{psidagpsi}$$ The equations (\[sd1\], \[sd2\]) lead to the affine Toda equations $$\nabla ^{2}\ln \rho _{a}+\sum_{b=1}^{r+1}\widetilde{C}_{ab}\rho _{b}=0
\label{Toda}$$ for $a=1,r$, plus the index for $M$, *i.e.* $a=1,2$. $\widetilde{C}_{ab} $ is the extended Cartan matrix $$\widetilde{C}_{ab}=\frac{2\mathbf{\alpha }^{\left( a\right) }\cdot \mathbf{\alpha }^{\left( b\right) }}{\left| \mathbf{\alpha }^{\left( b\right)
}\right| ^{2}},\;a,b=1,2$$ where $\mathbf{\alpha }^{\left( a\right) }$ are the simple root vectors of the algebra $su\left( 2\right) $, and in addition the minus maximal root. The equations (\[Toda\]) can be written in detail for $\rho _{1}\equiv
\left| \psi ^{1}\right| ^{2}$, $\rho _{2}\equiv \left| \psi ^{-M}\right|
^{2} $ $$\begin{aligned}
\Delta \ln \rho _{1}+2(\rho _{1}-\rho _{2}) &=&0 \label{deltarho12} \\
\Delta \ln \rho _{2}+2(-\rho _{1}+\rho _{2}) &=&0 \notag\end{aligned}$$ and this gives the relation $$\Delta \ln \left( \rho _{1}\rho _{2}\right) =0$$ or $$\rho _{2}=\text{const}\;\rho _{1}^{-1} \label{rho12}$$ since the exponential of a linear term is excluded by the conditions on a circle at infinity. Using Eq.(\[psidagpsi\]) we have to identify $$\omega =\rho _{1}-\rho _{2} \label{omrho1min2}$$ whose equation is obtained from equations in (\[deltarho12\]) using (\[rho12\]) const$=1$ (see bellow) $$\Delta \ln \rho _{1}+2(\rho _{1}-\rho _{1}^{-1})=0 \label{ecrho1}$$ The substitution $\left| \beta \right| \psi \equiv \ln \rho _{1}$ leads both equations (\[omrho1min2\]) and (\[ecrho1\]) to the form $$\Delta \psi +\frac{4}{\left| \beta \right| }\sinh \left( \left| \beta
\right| \psi \right) =0 \label{happy}$$
Choosing a constant different of unity in Eq.(\[rho12\]) imposes a linear substitution of the stream function : $\psi \rightarrow \psi ^{\prime
}\equiv \gamma +\beta \psi $ and modifies the factor multiplying the $\sinh $ function in Eq.(\[happy\]).
We have constructed the Lagrangian density (\[LagSU2\]) with a standard structure of matter-gauge field interaction and incorporating suggestions from models relevant for our objective. The aim was to reproduce the model of fluid vortex filaments supposed to be a faithfull representation of the original Euler fluid dynamics. The fundamental step in the derivation of the $\sinh $-Poisson equation is the assumption that the energy density (\[Bogo\]) is minimum (which is equivalent to minimizing the *action* for these stationary solution) *i.e.* in the condition of *self-duality*.
We conclude that the Euler fluid equations can be embedded into a field theoretical framework via the discrete model of vortex filaments. This framework is able to provide the equation obeyed by the scalar stream function at stationarity and this is precisely the equation inferred from numerical experiments. The particularity of the field theoretical framework is that it shows that the asymptotic states consisting of regular vortices correspond to *self-dual* states in field theory. This might have deep significance and deserves further investigation.
One must also note that the statistical-theoretical model able to explain the $\sinh $-Poisson equation [@Montg2], [@Montg3], [@Joyce], [@Smith] appears to have strong and deep connections with the self-duality. It should be carefully considered as a possible approach in the study of models where self-duality is known to exist, like $O\left( n\right) $ or Self-Dual Yang-Mills theory.
**Aknowledgement**. This work has been presented at the *Forum of Theory* organized by the Department de Recherche sur la Fusion Controlee, CEA-Cadarache, in Aix-en-Provence, France (July 2001). The authors are grateful to the organizers for inviting them to this meeting.
[99]{} D. Montgomery, W.H. Mathaeus, W.T. Stribling, D. Martinez and S. Oughton, Phys. Fluids **A4** (1992) 3
D. Fyfe, D. Montgomery and G. Joyce, J. Plasma Phys. **17**, 369 (1976).
R. H. Kraichnan and D. Montgomery, Rep. Prog. Phys. **43**, 547 (1980)
C.P. Burgess, B.P. Dolan, *Particle vortex duality and the modular group: application to the quantum Hall effect and other 2-D systems*, hep-th/0010246.
R. Jackiw and So-Young Pi, Phys. Rev. D**42**, 3500 (1990).
R. Jackiw and So-Young Pi, Phys. Rev. Lett. **64**, 2969 (1990).
G. Nardelli, Phys. Rev. D**52**, 5944 (1995)
G. Dunne, *Self-dual Chern-Simmons theories*, hep-th/9410065.
T. Hollowood, *Quantum solitons in affine Toda field theories*, hep-th/9110010.
D.H. Sattinger and O.L. Weaver, *Lie Groups and Algebras with Applications to Physics, Geometry and Mechanics*, Applied Mathematical Sciences Vol.61, Springer- Verlag New York, 1986.
D. Montgomery and G. Joyce, Phys. Fluids [**17**]{}, 1139 (1974)
D. Montgomery, L. Turner and G. Vahala, J. Plasma Phys. [**21**]{}, 239 (1979)
G. Joyce and D. Montgomery, J. Plasma Phys. [**10**]{}, 107 (1973)
R.A. Smith, Phys. Rev. A[**43**]{}, 1126 (1991)
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
Using data collected with the BESII detector at $e^{+}e^{-}$ storage ring Beijing Electron Positron Collider, the measurements of relative branching fractions for seven Cabibbo suppressed hadronic weak decays $D^0 \rightarrow K^- K^+$, $\pi^+ \pi^-$, $K^- K^+ \pi^+ \pi^-$ and $\pi^+ \pi^+ \pi^- \pi^-$, $D^+
\rightarrow \overline{K^0} K^+$, $K^- K^+ \pi^+$ and $\pi^- \pi^+
\pi^+$ are presented.
[*PACS:*]{} 13.25.Ft, 14.40.Lb
title: '**Measurements of Cabibbo Suppressed Hadronic Decay Fractions of Charmed $D^0$ and $D^+$ Mesons**'
---
M. Ablikim$^{1}$, J. Z. Bai$^{1}$, Y. Ban$^{11}$, J. G. Bian$^{1}$, X. Cai$^{1}$, J. F. Chang$^{1}$, H. F. Chen$^{16}$, H. S. Chen$^{1}$, H. X. Chen$^{1}$, J. C. Chen$^{1}$, Jin Chen$^{1}$, Jun Chen$^{7}$, M. L. Chen$^{1}$, Y. B. Chen$^{1}$, S. P. Chi$^{2}$, Y. P. Chu$^{1}$, X. Z. Cui$^{1}$, H. L. Dai$^{1}$, Y. S. Dai$^{18}$, Z. Y. Deng$^{1}$, L. Y. Dong$^{1}$$^a$, Q. F. Dong$^{15}$, S. X. Du$^{1}$, Z. Z. Du$^{1}$, J. Fang$^{1}$, S. S. Fang$^{2}$, C. D. Fu$^{1}$, H. Y. Fu$^{1}$, C. S. Gao$^{1}$, Y. N. Gao$^{15}$, M. Y. Gong$^{1}$, W. X. Gong$^{1}$, S. D. Gu$^{1}$, Y. N. Guo$^{1}$, Y. Q. Guo$^{1}$, K. L. He$^{1}$, M. He$^{12}$, X. He$^{1}$, Y. K. Heng$^{1}$, H. M. Hu$^{1}$, T. Hu$^{1}$, X. P. Huang$^{1}$, X. T. Huang$^{12}$, X. B. Ji$^{1}$, C. H. Jiang$^{1}$, X. S. Jiang$^{1}$, D. P. Jin$^{1}$, S. Jin$^{1}$, Y. Jin$^{1}$, Yi Jin$^{1}$, Y. F. Lai$^{1}$, F. Li$^{1}$, G. Li$^{2}$, H. H. Li$^{1}$, J. Li$^{1}$, J. C. Li$^{1}$, Q. J. Li$^{1}$, R. Y. Li$^{1}$, S. M. Li$^{1}$, W. D. Li$^{1}$, W. G. Li$^{1}$, X. L. Li$^{8}$, X. Q. Li$^{10}$, Y. L. Li$^{4}$, Y. F. Liang$^{14}$, H. B. Liao$^{6}$, C. X. Liu$^{1}$, F. Liu$^{6}$, Fang Liu$^{16}$, H. H. Liu$^{1}$, H. M. Liu$^{1}$, J. Liu$^{11}$, J. B. Liu$^{1}$, J. P. Liu$^{17}$, R. G. Liu$^{1}$, Z. A. Liu$^{1}$, Z. X. Liu$^{1}$, F. Lu$^{1}$, G. R. Lu$^{5}$, H. J. Lu$^{16}$, J. G. Lu$^{1}$, C. L. Luo$^{9}$, L. X. Luo$^{4}$, X. L. Luo$^{1}$, F. C. Ma$^{8}$, H. L. Ma$^{1}$, J. M. Ma$^{1}$, L. L. Ma$^{1}$, Q. M. Ma$^{1}$, X. B. Ma$^{5}$, X. Y. Ma$^{1}$, Z. P. Mao$^{1}$, X. H. Mo$^{1}$, J. Nie$^{1}$, Z. D. Nie$^{1}$, H. P. Peng$^{16}$, N. D. Qi$^{1}$, C. D. Qian$^{13}$, H. Qin$^{9}$, J. F. Qiu$^{1}$, Z. Y. Ren$^{1}$, G. Rong$^{1}$, L. Y. Shan$^{1}$, L. Shang$^{1}$, D. L. Shen$^{1}$, X. Y. Shen$^{1}$, H. Y. Sheng$^{1}$, F. Shi$^{1}$, X. Shi$^{11}$$^c$, H. S. Sun$^{1}$, J. F. Sun$^{1}$, S. S. Sun$^{1}$, Y. Z. Sun$^{1}$, Z. J. Sun$^{1}$, X. Tang$^{1}$, N. Tao$^{16}$, Y. R. Tian$^{15}$, G. L. Tong$^{1}$, D. Y. Wang$^{1}$, J. Z. Wang$^{1}$, K. Wang$^{16}$, L. Wang$^{1}$, L. S. Wang$^{1}$, M. Wang$^{1}$, P. Wang$^{1}$, P. L. Wang$^{1}$, S. Z. Wang$^{1}$, W. F. Wang$^{1}$$^d$, Y. F. Wang$^{1}$, Z. Wang$^{1}$, Z. Y. Wang$^{1}$, Zhe Wang$^{1}$, Zheng Wang$^{2}$, C. L. Wei$^{1}$, D. H. Wei$^{1}$, N. Wu$^{1}$, Y. M. Wu$^{1}$, X. M. Xia$^{1}$, X. X. Xie$^{1}$, B. Xin$^{8}$$^b$, G. F. Xu$^{1}$, H. Xu$^{1}$, S. T. Xue$^{1}$, M. L. Yan$^{16}$, F. Yang$^{10}$, H. X. Yang$^{1}$, J. Yang$^{16}$, Y. X. Yang$^{3}$, M. Ye$^{1}$, M. H. Ye$^{2}$, Y. X. Ye$^{16}$, L. H. Yi$^{7}$, Z. Y. Yi$^{1}$, C. S. Yu$^{1}$, G. W. Yu$^{1}$, C. Z. Yuan$^{1}$, J. M. Yuan$^{1}$, Y. Yuan$^{1}$, S. L. Zang$^{1}$, Y. Zeng$^{7}$, Yu Zeng$^{1}$, B. X. Zhang$^{1}$, B. Y. Zhang$^{1}$, C. C. Zhang$^{1}$, D. H. Zhang$^{1}$, H. Y. Zhang$^{1}$, J. Zhang$^{1}$, J. W. Zhang$^{1}$, J. Y. Zhang$^{1}$, Q. J. Zhang$^{1}$, S. Q. Zhang$^{1}$, X. M. Zhang$^{1}$, X. Y. Zhang$^{12}$, Y. Y. Zhang$^{1}$, Yiyun Zhang$^{14}$, Z. P. Zhang$^{16}$, Z. Q. Zhang$^{5}$, D. X. Zhao$^{1}$, J. B. Zhao$^{1}$, J. W. Zhao$^{1}$, M. G. Zhao$^{10}$, P. P. Zhao$^{1}$, W. R. Zhao$^{1}$, X. J. Zhao$^{1}$, Y. B. Zhao$^{1}$, H. Q. Zheng$^{11}$, J. P. Zheng$^{1}$, L. S. Zheng$^{1}$, Z. P. Zheng$^{1}$, X. C. Zhong$^{1}$, B. Q. Zhou$^{1}$, G. M. Zhou$^{1}$, L. Zhou$^{1}$, N. F. Zhou$^{1}$, K. J. Zhu$^{1}$, Q. M. Zhu$^{1}$, Y. C. Zhu$^{1}$, Y. S. Zhu$^{1}$, Yingchun Zhu$^{1}$$^e$, Z. A. Zhu$^{1}$, B. A. Zhuang$^{1}$, X. A. Zhuang$^{1}$, B. S. Zou$^{1}$.\
(BES Collaboration)\
\[att\] $^{1}$ Institute of High Energy Physics, Beijing 100049, People’s Republic of China\
$^{2}$ China Center for Advanced Science and Technology (CCAST), Beijing 100080, People’s Republic of China\
$^{3}$ Guangxi Normal University, Guilin 541004, People’s Republic of China\
$^{4}$ Guangxi University, Nanning 530004, People’s Republic of China\
$^{5}$ Henan Normal University, Xinxiang 453002, People’s Republic of China\
$^{6}$ Huazhong Normal University, Wuhan 430079, People’s Republic of China\
$^{7}$ Hunan University, Changsha 410082, People’s Republic of China\
$^{8}$ Liaoning University, Shenyang 110036, People’s Republic of China\
$^{9}$ Nanjing Normal University, Nanjing 210097, People’s Republic of China\
$^{10}$ Nankai University, Tianjin 300071, People’s Republic of China\
$^{11}$ Peking University, Beijing 100871, People’s Republic of China\
$^{12}$ Shandong University, Jinan 250100, People’s Republic of China\
$^{13}$ Shanghai Jiaotong University, Shanghai 200030, People’s Republic of China\
$^{14}$ Sichuan University, Chengdu 610064, People’s Republic of China\
$^{15}$ Tsinghua University, Beijing 100084, People’s Republic of China\
$^{16}$ University of Science and Technology of China, Hefei 230026, People’s Republic of China\
$^{17}$ Wuhan University, Wuhan 430072, People’s Republic of China\
$^{18}$ Zhejiang University, Hangzhou 310028, People’s Republic of China\
$^{a}$ Current address: Iowa State University, Ames, IA 50011-3160, USA.\
$^{b}$ Current address: Purdue University, West Lafayette, IN 47907, USA.\
$^{c}$ Current address: Cornell University, Ithaca, NY 14853, USA.\
$^{d}$ Current address: Laboratoire de l’Acc[é]{}l[é]{}ratearLin[é]{}aire, F-91898 Orsay, France.\
$^{e}$ Current address: DESY, D-22607, Hamburg, Germany.\
INTRODUCTION {#secintro}
============
Hadronic decays of charmed mesons have been extensively studied. Measurements of relative lifetimes and semileptonic branching fractions for charmed mesons $D^+$ and $D^0$ suggest the presence of nonleptonic processes which enhance the $D^0$ and suppress the $D^+$ width, and lead to the conclusion that the simple spectator model of charmed-meson decay is inadequate. As shown in Figure \[feynme\], all weak decays of heavy mesons may be described by six quark-diagrams: the external W-emission diagrams (a), the internal W-emission diagrams (b), the W-exchange diagram (c), the W-annihilation diagram (d), the horizontal W-loop diagram (e), and the vertical W-loop diagram (f) [@sixfey]. Thus, a further understanding of the $D$ decay mechanism, such as the contributions of other quark-diagrams and final states interactions, requires a systematic study of the hadronic decays.
![ Typical Feynman diagrams for Cabibbo suppressed decays of charm mesons.[]{data-label="feynme"}](a.eps){height="2.0cm" width="3.0cm"}
![ Typical Feynman diagrams for Cabibbo suppressed decays of charm mesons.[]{data-label="feynme"}](b.eps){height="2.0cm" width="3.0cm"}
![ Typical Feynman diagrams for Cabibbo suppressed decays of charm mesons.[]{data-label="feynme"}](c.eps){height="2.0cm" width="3.0cm"}
![ Typical Feynman diagrams for Cabibbo suppressed decays of charm mesons.[]{data-label="feynme"}](d.eps){height="2.0cm" width="3.0cm"}
![ Typical Feynman diagrams for Cabibbo suppressed decays of charm mesons.[]{data-label="feynme"}](e.eps){height="2.0cm" width="3.0cm"}
![ Typical Feynman diagrams for Cabibbo suppressed decays of charm mesons.[]{data-label="feynme"}](f.eps){height="2.0cm" width="3.0cm"}
At present, many experiments, such as, MARKII [@mk2], MARKIII [@mk3], E691 [@e691], E687 [@e687], E791 [@e791], FOCUS [@focus] and CLEO [@cleo], have reported their measurements of Cabibbo suppressed hadronic decay fractions of $D$ mesons. Our present measurements are based on a data sample of integrated luminosity of $\sim 17.3 \mathrm{pb^{-1}}$ at $\psi(3770)$ peak ($\sqrt{s}=3.773$ GeV) and $\sim
16.5\mathrm{pb^{-1}}$ for $\psi(3770)$ peak scan collected by Beijing Spectrometer (BESII) detector at $e^{+}e^{-}$ storage ring Beijing Electron Positron Collider (BEPC) [@bepc]. This paper reports our measurements of Cabibbo suppressed relative branching fractions for several hadronic decay modes of charmed $D$ mesons $D^0 \rightarrow K^- K^+$, $\pi^+ \pi^-$, $K^- K^+ \pi^+ \pi^-$, $\pi^+ \pi^+ \pi^- \pi^-$, $D^+ \rightarrow \overline{K^0} K^+$, $K^- K^+ \pi^+$ and $\pi^- \pi^+ \pi^+$ (through out this paper the charge conjugate states are implicitly included).
BESII DETECTOR {#secbes}
==============
The Beijing Spectrometer (BESII) is a conventional cylindrical magnetic detector that is described in detail in Ref. [@bes2]. A 12-layer Vertex Chamber (VC) surrounds the beryllium beam pipe and provides trigger information, as well as coordinate information. A forty-layer main drift chamber (MDC) located just outside the VC yields precise measurements of charged particle trajectories with a solid angle coverage over $85\%$ of $4\pi$; it also provides ionization energy loss ($dE/dx$) measurements which are used for particle identification. Momentum resolution of $1.7\%\sqrt{1+p^2}$ ($p$ in GeV/$c$) and $dE/dx$ resolution for hadron tracks of $\sim8\%$ are obtained. An array of 48 scintillation counters surrounding the MDC measures the time of flight (TOF) of charged particles with a resolution of about 200 ps for hadrons. Outside the TOF counters, a 12 radiation length, lead-gas barrel shower counter (BSC), operating in limited streamer mode, measures the energies of electrons and photons over $80\%$ of the total solid angle with an energy resolution of $\sigma_{E}/E=0.22/\sqrt{E}$ ($E$ in GeV). Outside the solenoidal coin, which provides a 0.4 T magnetic field over the tracking volume, is an iron flux return that is instrumented with three double-layer muon counters that identify muons with momentum greater than 500 MeV$/c$.
In this analysis, a GEANT3 based Monte Carlo package (SIMBES [@simbes]) with detailed consideration of the detector performance (such as dead electronic channels) is used. The consistency between data and Monte Carlo has been carefully checked in many high purity physics channels, and the agreement is reasonable.
EVENT SELECTION {#secevt}
===============
Charged tracks are required to satisfy $\left| \cos \theta \right|
< 0.8$, where $\theta$ is the polar angle in the MDC, and have good helix fit. The tracks that are not associated with $K^0_S$ reconstruction are required to be originate from interaction point. Pions and kaons are identified by requiring the confidence level of desired hypothesis using combined measurements of time-of-flight [@tof] and energy loss in drift chamber to be greater than 0.1%. In addition, kaon and pion candidates are further identified by requiring the normalized weights, which is defined as ${CL}_{\alpha}/({CL}_{\pi}+{CL}_{K})$, where $\alpha$ denotes desired particle, exceeding $50\%$.
$K^0_S$ candidates are detected through the decay of $K^0_S
\rightarrow \pi^+ \pi^-$. Each oppositely charged track pair is assumed to be $\pi^+$ and $\pi^-$. The decay vertex of $K^0_S$ is required to be 5mm far away from the beam axis. The $\pi^+ \pi^-$ invariant mass is required to be within 0.020 GeV/$c^2$ of the $K^{0}_{S}$ nominal mass.
Cabibbo suppressed hadronic decay modes are expected at a lower rate ($\sim
\tan^2\theta_{C}$, where $\theta_{C}$ is “Cabibbo angle”) compared to relevant Cabibbo favoured modes, for which $\pi/K$ misidentification becomes significant. The unique energy of $D$ meson at the $\psi(3770)$ can be exploited to reduce explicitly background due to incorrect particle assignment. A single particle misidentification results in a reflection peak separated from the beam energy.
The distributions of energy difference($\Delta
E$) between measured energy of $D$ candidates ($E_{\mathrm{tag}}$) and beam energy ($E_{\mathrm{b}}$) are shown in Figure \[dele\] for 4 Cabibbo allowed decay modes with correct $\pi / K$ assignments and the reflection $\Delta E$ distribution with a single particle misidentification. $\Delta E$ of $D$ candidate is required to be less than that of similar topological decay mode to reject particle misidentification combination. $\Delta E$ is further required to be within 50-100 MeV for different decay channels.
![ $\Delta E$ distribution and reflection for similar Cabibbo suppressed modes by using the sample of Cabibbo favoured mode (a) $D^0\rightarrow K^- \pi^+$ (b) $D^0\rightarrow
\overline{K^0} \pi^+$ (c) $D^+\rightarrow K^- \pi^+ \pi^+$ (d) $D^0 \rightarrow K^{-} \pi^{+}\pi^{+}\pi^{-}$. The distributions show the well-separated $\Delta E$ peaks.[]{data-label="dele"}](deltae.eps){height="10.0cm" width="8.0cm"}
The pair production of $D\overline{D}$ at $\psi(3770)$ provides a variable, which is defined as the beam-constrained mass $$M_{\mathrm{bc}}=\sqrt{E^2_{\mathrm{beam}} - (\sum_{\mathrm{i}} p_{\mathrm{i}})^2}$$ exploiting the fact that the total energy of all decay products must sum to the beam energy. As the uncertainty in the beam energy is much smaller than the uncertainty in the total reconstructed energy of the decay tracks, this approach yields much improved mass resolution compared to the invariant mass technique.
The QED processes, $\tau^{+}\tau^{-}$ pair productions and cosmic backgrounds may contribute to $D$ tags. Both of them have a lower multiplicity than that of $D\overline{D}$ decay, the requirement of $\displaystyle N_{\mathrm{ch}}+N_{\mathrm{neu}}/2>3$ will eliminate most of these backgrounds, where $N_{\mathrm{ch}}$ and $N_{\mathrm{neu}}$ represent the total number of charged tracks and neutral tracks respectively. At $\psi(3770)$, $D
\overline{D}$ are produced with the angular distribution $\sin^2
\theta_D$, where $\theta_{D}$ is the production angle of $\psi(3770)\rightarrow D \overline{D}$, $\left|\cos
\theta_{D}\right| < 0.8$ is imposed to each $D$ tag to enhance signal to background ratio.
One event could be counted more than once as a tag candidate. In order to calculate the actual number of tagged events in an unbiased manner, the following criterion is applied to select only one tag combination per event: if more than one combination of tracks form the desired tag, the combination is chosen when the lowest momentum track of this combination has the largest momentum of all other combinations. The result mass plots are shown in Figure \[2body\], \[3body\].
DETECTOR ACCEPTANCE {#secmc}
===================
The detection efficiency is determined by a detailed Monte Carlo simulation of $D \overline{D}$ production, decay and detector response. The decay branching ratios of neutral and charged $D$ meson are taken from the world average values [@PDG], some unseen decay modes are set according to the rules of isospin conservation. Simulated events are processed through the event reconstruction, selection and analysis program.
There are several sub-resonant decay modes in 3-body and 4-body Cabibbo suppressed channels. The detection efficiency is not uniform among these decay modes. The relative decay branching fractions and their errors in PDG are quoted. For $D^{+}\rightarrow\pi^{-}\pi^{+}\pi^{+}$ channel, $\rho^{0}\pi^{+}$ and $\pi^{-}\pi^{+}\pi^{+}$ modes are considered; for $D^{+}\rightarrow K^{-} K^{+}\pi^{+}$ channel, $\phi\pi^{+}$, $\overline{K^{*0}}K^{+}$ and $K^{-}K^{+}\pi^{+}$ are considered; for $D^{0}\rightarrow K^{-} K^{+}\pi^{+}\pi^{-}$ channel, $\phi\pi^{+}\pi^{-}$, $\phi\rho^{0}$, $\overline{K^{*0}}K^{*0}$, $K^{-}K^{+}\rho^{0}$ and $K^{-}K^{+}\pi^{+}\pi^{-}$ are considered; for $D^{0}\rightarrow\pi^{-}\pi^{+}\pi^{+}\pi^{-}$ channel, the uncertainty to Monte Carlo efficiency is estimated to be less than 2$\%$.
RESULTS {#secbr}
=======
The observed number of each Cabibbo suppressed decay channel is determined by fitting the distributions to a function of the form: $$\begin{array}{ccc}
F(m)&=&a_{\mathrm{1}}\left[m\sqrt{1-\left(\frac{m}{E_{\mathrm{b}}}\right)^{2}}
\exp\left(a_{\mathrm{2}}\left[1-\left(\frac{m}{E_{\mathrm{b}}}\right)^2\right]\right)\right]\\
&+
&a_{\mathrm{3}}\exp\left(-\frac{(m-M_{\mathrm{D}})^{2}}{2\sigma^{2}}\right)
+ a_{\mathrm{4}}
\end{array}$$ where the first term parameterizes the background; the Gaussian term accounts for the signal. Just above the $D$ mass, there is no more phase space available for a decay to a pair of $D\overline{D}$ mesons. The background term is ARGUS form [@argusbg], where $a_{\mathrm{1}}$ is the normalization factor; $a_{\mathrm{2}}$, a scale factor for the exponential term; $E_{\mathrm{b}}$, the beam energy and fixed at 1.8865GeV while fitting the mass plot. For $\psi(3770)$ scan data, the beam energy is not a constant, but the background shape is similar, a constant background term $a_{\mathrm{4}}$ is applied to evaluate the varying beam energy points. The mass resolution of each Cabbibo suppressed mode is determined by fitting the $M_{\mathrm{bc}}$ plot of similar Cabbibo allowed modes. The event number for each mode is summarized in Table \[results\].
The pions that are part of reconstructed $K^{0}_{S}$’s, are not used in modes $\pi\pi$, $~\pi\pi\pi$, $~\pi\pi\pi\pi$ and $KK\pi$, $KK\pi\pi$. However, there remain some $D$ decays into $K^{0}_{S}\pi^{+}$, $K^{0}_{S}\pi^{+}\pi^{-}$ and $K^{-}K^{+}K^{0}_{S}$, where the $K^{0}_{S}$ is not identified as a separated vertex, which will be the major background source for decay mode $D^{+}\rightarrow\pi^{-}\pi^{+}\pi^{+}$ and $D^{0}\rightarrow\pi^{-}\pi^{+}\pi^{+}\pi^{-},
K^{-}K^{+}\pi^{+}\pi^{-}$. To reduce the feed-down $K^{0}_{S}\rightarrow\pi^{+}\pi^{-}$ background, a cut of $\left|M_{\pi^{+}\pi^{-}}-M_{K^{0}}\right|>0.040$ GeV$/c^{2}$ is imposed on the invariant mass of each pion pairs for $\pi^{-}\pi^{+}\pi^{+}$ and $K^{-}K^{+}\pi^{+}\pi^{-}$ modes.
For the decay $D^{0}\rightarrow\pi^{-}\pi^{+}\pi^{+}\pi^{-}$, plots of the invariant mass of all $\pi^{+}\pi^{-}$ combinations in $D^0$ candidates within the signal and sideband regions are shown in Figure \[4piks\](a) and (b) respectively. There is a clear $K^{0}_{S}$ peak within $D^0$ signal region with a fitted events number of $112.5 \pm 19.0$ events and no clear $K^0_S$ peak within the sideband region. The $K^0_S$ number is consistent with the expected background (Monte Carlo study gives this number as 98.8) due to $D^{0}\rightarrow\overline{K^{0}}\pi^{+}\pi^{-}$. These are thus subtracted from the $4\pi$ signal.
![Beam energy constrain mass distributions for decays (a) $D^0 \rightarrow K^- K^+$, (b) $D^0 \rightarrow \pi^- \pi^+$, (c) $D^0 \rightarrow K^- \pi^+$ and (d) $D^+ \rightarrow \overline{K^0} K^+ $, (e) $D^+ \rightarrow \overline{K^0} \pi^+$.[]{data-label="2body"}](d0_2trk.eps "fig:"){width="6.5cm"} ![Beam energy constrain mass distributions for decays (a) $D^0 \rightarrow K^- K^+$, (b) $D^0 \rightarrow \pi^- \pi^+$, (c) $D^0 \rightarrow K^- \pi^+$ and (d) $D^+ \rightarrow \overline{K^0} K^+ $, (e) $D^+ \rightarrow \overline{K^0} \pi^+$.[]{data-label="2body"}](dp_k0trk.eps "fig:"){width="6.5cm"}
![Beam energy constrain mass distribution for decay (a) $D^+ \rightarrow K^- K^+ \pi^+$, (b) $D^+ \rightarrow \pi^- \pi^+ \pi^+$, (c) $D^+ \rightarrow K^- \pi^+ \pi^+$ and (d) $D^0 \rightarrow K^- K^+ \pi^+ \pi^-$, (e) $D^0 \rightarrow \pi^- \pi^+ \pi^+ \pi^-$, (f) $D^0 \rightarrow K^- \pi^+ \pi^+ \pi^-$.[]{data-label="3body"}](dp_3trk.eps "fig:"){width="6.5cm"} ![Beam energy constrain mass distribution for decay (a) $D^+ \rightarrow K^- K^+ \pi^+$, (b) $D^+ \rightarrow \pi^- \pi^+ \pi^+$, (c) $D^+ \rightarrow K^- \pi^+ \pi^+$ and (d) $D^0 \rightarrow K^- K^+ \pi^+ \pi^-$, (e) $D^0 \rightarrow \pi^- \pi^+ \pi^+ \pi^-$, (f) $D^0 \rightarrow K^- \pi^+ \pi^+ \pi^-$.[]{data-label="3body"}](d0_4trk.eps "fig:"){width="6.5cm"}
![ $M_{\pi^{+}\pi^{-}}$ for $\pi^- \pi^+ \pi^+ \pi^-$ candidates within $D^0$ signal region (a) and sideband region (b).[]{data-label="4piks"}](mpp.eps){width="6.0cm"}
SYSTEMATIC UNCERTAINTIES {#secse}
========================
In this analysis, we normalize the relative ratios of Cabibbo suppressed decay to similar Cabibbo favoured modes, which permits the cancellation of many common systematic errors. The systematic uncertainties on energy cut and particle identification can not be canceled completely.
Systematic uncertainties on particle identification and $\Delta E$ cuts are about 1-3 % and 1-6% contributing to the relative fractions respectively. The systematic uncertainties on background subtraction due to the $K^{0}_{S}$ contamination are estimated to be $\sim2\%$ [@ks].
The detection efficiencies are not uniform among the different sub-resonant decay modes in the 3 and 4 body decays, $0.5-2.0\%$ are estimated for different modes $\pi^{-}\pi^{+}\pi^{+}$, $K^{-}K^{+}\pi^{+}$ and $\pi^{-}\pi^{+}\pi^{+}\pi^{-}$. Large uncertainty in $K^{-}K^{+}\pi^{+}\pi^{-}$ mode is found due to the decays of low momentum kaons.
CONCLUSIONS {#secconcl}
===========
[cccc]{}
Decay mode & Yield & Relative efficiency & Branching ratio\
[$\frac{K^- K^+}{K^- \pi^+}$]{} & [$\frac{242.2 \pm 20.1}{1934 \pm 49}$]{} & $1.029\pm0.017$ & $0.122 \pm 0.011 \pm 0.004$\
[$\frac{\pi^- \pi^+}{K^- \pi^+}$]{} & [$\frac{75.9 \pm 14.7}{1934 \pm 49}$]{} & $1.146\pm0.030$ & $0.034 \pm 0.007 \pm 0.001$\
[$\frac{\overline{K^0} K^+}{\overline{K^0} \pi^+}$]{} & [$\frac{63.2\pm9.8}{287 \pm 18}$]{} & $0.991\pm0.041$ & $0.222 \pm 0.037 \pm 0.013$\
[$\frac{K^- K^+ \pi^+}{K^- \pi^+ \pi^+}$]{} & [$\frac{181.2\pm20.2}{2324 \pm 53}$]{} & $0.669 \pm 0.015 \pm 0.010$ & $0.117 \pm 0.013 \pm 0.007 $\
[$\frac{\pi^- \pi^+ \pi^+}{K^- \pi^+ \pi^+}$]{} & [$\frac{84.9\pm22.4}{2324 \pm 53}$]{} & $0.888 \pm 0.029 \pm 0.004$ & $0.041 \pm 0.011 \pm 0.003 $\
[$\frac{K^- K^+ \pi^+ \pi^-}{K^- \pi^+ \pi^+ \pi^-}$]{} & [$\frac{19.3\pm8.0}{1540 \pm 51}$]{} & $0.286 \pm 0.021 \pm
0.017$ & $0.044 \pm 0.018 \pm 0.005$\
[$\frac{\pi^- \pi^+ \pi^+ \pi^-}{K^- \pi^+ \pi^+ \pi^-}$]{} & [$\frac{(274.4\pm31.8)-(112.5 \pm 19.0)}{1540 \pm 51}$]{} & $1.336 \pm 0.028 \pm 0.027$ & $0.079 \pm 0.018 \pm 0.005 $\
1.5cm
The relative fractions of seven Cabibbo suppressed decay modes are tabulated in Table.\[results\]. For each mode, fitted event number, background number, relative efficiency and relative branching ratio are listed. The errors of relative efficiency are Monte Carlo statistical error and systematic error due to sub-resonant modes respectively. The first error of branching ratio is statistical, the second one is systematic. The estimated systematic uncertainty is 3-6% for all modes, and the statistical uncertainties of the measurements are about 10% or greater. Results from this measurement are consistent with the world average values and we have improved the previous measurements of the $D^+ \rightarrow \overline{K^0} K^+$ relative branching ratio. The measurements of Cabibbo suppressed branching ratios presented here provide new insights into the mechanism of nonleptoinc $D$ decays. Exact SU(3) symmetry predicts the equality of $\Gamma (D^0
\rightarrow \pi^- \pi^+)$ and $\Gamma (D^0
\rightarrow K^- K^+)$. But the above results show they are not equal. Several distinct effects could contribute to this inequality. Final states interactions breaking SU(3) symmetry could however account for the difference.
ACKNOWLEDGMENTS {#secack}
===============
The BES collaboration thanks the staff of BEPC for their hard efforts. This work is supported in part by the National Natural Science Foundation of China under contracts Nos. 19991480, 10225524, 10225525, the Chinese Academy of Sciences under contract No. KJ 95T-03, the 100 Talents Program of CAS under Contract Nos. U-11, U-24, U-25, and the Knowledge Innovation Project of CAS under Contract Nos. U-602, U-34 (IHEP); and by the National Natural Science Foundation of China under Contract No. 10175060 (USTC), and No. 10225522 (Tsinghua University).
[99]{} L.L. Chau and H.Y. Cheng, Phys. Rev. [**D36**]{}, 137(1987) G.S. Abrams, , Phys. Rev. Lett. [**43**]{}, 481(1979) R.M. Baltrusaitis, , Phys. Rev. Lett. [**55**]{}, 150(1985) E.M. Aitala, , [FNAL E791 Collab.]{} Phys. Lett. [**B421**]{}, 405(1998); E.M. Aitala, , [FNAL E791 Collab.]{} Phys. Lett. [**B423**]{}, 185(1998); E.M. Aitala, , [FNAL E791 Collab.]{} Phys. Rev. Lett. [**86**]{}, 770(2001) J.C. Anjos, , (FNAL E691 Collab.) Phys. Rev. Lett. [**62**]{}, 125(1989); J.C. Anjos, , (FNAL E691 Collab.) Phys. Rev. [**D41**]{}, 2705(1990); J.C. Anjos, , (FNAL E691 Collab.) Phys. Rev. [**D43**]{}, 635(1991); J.C. Anjos, , (FNAL E691 Collab.) Phys. Rev. [**D44**]{}, 3371(1991) P.L. Frabetti, , (FNAL E687 Collab.) Phys. Lett. [**B281**]{}, 167(1992); P.L. Frabetti, , (FNAL E687 Collab.) Phys. Lett. [**B321**]{}, 295(1994); P.L. Frabetti, , (FNAL E687 Collab.) Phys. Lett. [**B354**]{}, 486(1995); P.L. Frabetti, , (FNAL E687 Collab.) Phys. Lett. [**B346**]{}, 199(1995); P.L. Frabetti, , (FNAL E687 Collab.) Phys. Lett. [**B351**]{}, 591(1995); P.L. Frabetti, , (FNAL E687 Collab.) Phys. Lett. [**B407**]{}, 79(1997) R. Ammar , (CLEO Collab.) Phys. Rev. [**D44**]{}, 3383(1991); M. Bishai , (CLEO Collab.) Phys. Rev. Lett. [**78**]{}, 3261(1997); S.E. Csorma , (CLEO Collab.) Phys. Rev. [**D65**]{}, 092001(2002) J.M. Link , (FNAL FOCUS Collab.) Phys. Rev. Lett. [**88**]{} 041602(2002); J.M. Link , (FNAL FOCUS Collab.) Phys. Lett. [**B555**]{}, 167(2003); J.M. Link , (FNAL FOCUS Collab.) hep-ex/0411031 C. Zhang, , in: J. Rossback(Ed.). HEACC’92 Hamburg, XVth Int. Conf. on High Energy Accelerators, Hamburg, Germany, July 20-24, 1992, P.409. J.Z. Bai [*et al.*]{}, (BES Collab.) Nucl. Instr. and Meth. [**A458**]{} (2001) 627. J. C. Chen [*et al.*]{}, BES internal report. H. Albrecht [*et al.*]{}, Phys. Lett. [**B241**]{} (1990) 278. Particle Data Group, Phys. Lett. [**B592**]{} (2004) S.S. Sun, K.L. He , Systematic Study of Time of Flight Correction of BESII (Submitted to HEP &NP). Z. Wang , HEP &NP, 27, 1(2003).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We show that small ball estimates together with Hölder continuity assumption allow to obtain new representation results in models with long memory. In order to apply these results, we establish small ball probability estimates for Gaussian processes whose incremental variance admits two-sided estimates and the incremental covariance preserves sign. As a result, we obtain small ball estimates for integral transforms of Wiener processes and of fractional Brownian motion with Volterra kernels.'
address: 'National Taras Shevchenko University of Kyiv, Mechanics and Mathematics Faculty, 64 Volodymyrska, 01601 Kyiv, Ukraine'
author:
- Yuliya Mishura
- Georgiy Shevchenko
bibliography:
- 'represent-2.bib'
nocite: '[@*]'
title: Small ball properties and representation results
---
integral representation,generalized Lebesgue–Stieltjes integral ,small ball estimate,quasi-helix, fractional Brownian motion 60H05 ,60G15 ,60G22
Introduction
============
One of the most important questions for financial modeling is the question of replication, which loosely can be formulated as follows. Suppose that a continuous time financial market model is driven by a stochastic process $X= {\left\{X_t,t\in[0,1]\right\}}$ given on some stochastic basis $(\Omega, \mathcal{F}, \mathbb{F}= (\mathcal{F}_t)_{t \in [0,1]}, \mathsf{P})$ satisfying usual assumptions. A contingent claim, modeled by an $\mathcal{F}_1$-measurable random variable $\xi$, is hedgeable, if it admits the representation $$\label{equat1}
\xi=\int_0^1\psi_tdX_t,$$ with some $\mathbb{F}$-adapted (replicating) process $\psi$. In the case where $X$ is a Wiener process there are two main representation results. The famous Itô representation theorem establishes for centered square integrable random variables $\xi$. Less known is a result of Dudley [@dudley], who proved that every random variable $\xi$ has representation . There are also a lot of results for martingales or semimartingales, we will not cite them, as this is not our main concern here.
The case where $X$ is not a semimartingale is less studied. The pioneering results were established in [@mish-shev-valk] for fractional Brownian motion (fBm) $B^H$ with Hurst index $H>1/2$. The construction used in [@mish-shev-valk] relies on the Hölder continuity and a small ball estimates for $B^H$. This fact was later used in [@vita-shev2; @vita] to extend the results of [@mish-shev-valk] to a larger class of integrands. In [@vita-shev2], it is also shown that in the case where $X=W+B^H$ is a sum of a Wiener process and an fBm with $H>1/2$, any random variable has representation . It is worth to mention also the article [@vita-shev], where the existence of a continuous integrand $\psi$ is shown in the fBm case.
The main problem with the specific small ball property assumed in the papers [@vita-shev2; @vita] is that it is hard to verify. As it was mentioned in [@li-shao], an upper bound in small ball probability gives lower estimates for metric entropy, which are usually hard to obtain. On the other hand, the assumptions of [@vita-shev2; @vita] are not optimal for establishing representation results.
The goal of this paper is twofold. First, we investigate precise conditions needed to obtain the representation results and compare them to the small ball estimates. Second, we analyze carefully how to get an upper bound for small ball probability for Gaussian processes with variation distance ${\mathbf{E}}|X_t-X_s|^2$ satisfying two-sided power bounds, possibly, with different powers. These two steps allow us to establish the representation results for a wide class of processes. This class includes some Gaussian processes $X$ having non-stationary increments, e.g. processes that can be represented as the integrals of smooth Volterra kernels w.r.t. a Wiener process or fBm.
The paper is organized as follows. In Section \[sec3\], we prove representation theorems for Hölder continuous processes satisfying small ball property. In Section \[sec2\], we establish the small ball estimates for Gaussian processes whose incremental variance satisfies two-sided power estimates and incremental covariance preserves sign. In Section \[sec4\], we prove representation results for the Gaussian processes considered in Section \[sec2\], and give examples of processes, for which the representation results are in place. The examples include subfractional Brownian motion, bifractional Brownian motion, and integral transforms of Wiener process and fractional Brownian motion with Volterra kernels.
Representation theorems for Hölder continuous processes satisfying small ball estimates {#sec3}
=======================================================================================
This section is concerned with the representation results of the form . Here we establish general results for processes satisfying Hölder continuity and small ball assumptions.
Consider an adapted process $X$ satisfying the following assumptions, where $C^\theta[0,1]$ denotes the class of Hölder continuous functions of order $\theta$.
- Hölder continuity: $X\in C^\theta[0,1]$ a.s. for some $\theta>1/2$.
- Small ball estimate: there exist positive constants $\lambda,\mu,K_1,K_2$ such that for all ${\varepsilon}>0$, $\Delta>0$, $s\in[0,1-\Delta]$ $$\label{smallballassumption}
\mathbf{P}\left\{\sup_{s\leq t\leq s+\Delta}|X_t-X_s|\leq K_1 \varepsilon\right\}\leq\exp\left\{-K_2\varepsilon^{-\lambda}\Delta^{\mu}\right\}.$$
In [@vita], the author establishes existence of representation for a centered Gaussian process. The assumptions of [@vita] are close to be a particular case of $(H)$ and $(S)$. Namely, the author assumes $(S)$ with $\lambda=1/H$, $\mu=1$. Instead of $(H)$, the incremental variance is assumed to satisfy ${\mathbf{E}}(X_t-X_s)^2\le C|t-s|^{2H}$, which in the Gaussian case implies that $X\in C^\theta[0,1]$ for any $\theta\in(0,H)$.
\[rem:exponents\] It is clear that the exponents $\lambda,\mu,\theta$ must satisfy $\theta\le \mu/\lambda$. Indeed, assume on the contrary that $\theta>\mu/\lambda$ and take arbitrary $\delta\in(\mu/\lambda,\theta)$. Then for each $n\ge 1$ $$\mathbf{P}\left\{\sup_{0\le t\le n^{-1}}|X_t-X_0|\le n^{-\delta}\right\}\leq\exp\left\{-K_2n^{(\lambda\delta-\mu)}\right\},$$ whence by the Borel-Cantelli lemma, $\sup_{0\le t\le n^{-1}}|X_t-X_0|> n^{-\delta}$ for all $n$ large enough, which contradicts $(H)$.
Further we give basic facts on fractional integration; for more detail, see [@samko; @zahle]. Consider functions $f,g\colon[0,1]\rightarrow \mathbb{R}$, and let $[a,b]\subset [0,1]$. For $\alpha\in (0,1)$ define fractional derivatives $$\begin{gathered}
\big(D_{a+}^{\alpha}f\big)(x)=\frac{1}{\Gamma(1-\alpha)}\bigg(\frac{f(x)}{(x-a)^\alpha}+\alpha
\int_{a}^x\frac{f(x)-f(u)}{(x-u)^{1+\alpha}}du\bigg)1_{(a,b)}(x),\\
\big(D_{b-}^{1-\alpha}g\big)(x)=\frac{e^{i\pi
\alpha}}{\Gamma(\alpha)}\bigg(\frac{g(x)}{(b-x)^{1-\alpha}}+(1-\alpha)
\int_{x}^b\frac{g(x)-g(u)}{(x-u)^{2-\alpha}}du\bigg)1_{(a,b)}(x).\end{gathered}$$ Assuming that $D_{a+}^{\alpha}f\in L_1[a,b]$, $D_{b-}^{1-\alpha}g_{b-}\in
L_\infty[a,b]$, where $g_{b-}(x) = g(x) - g(b)$, the generalized Lebesgue–Stieltjes integral is defined as $$\int_a^bf(x)dg(x)=e^{-i\pi\alpha}\int_a^b\big(D_{a+}^{\alpha}f\big)(x)
\big(D_{b-}^{1-\alpha}g_{b-}\big)(x)dx.$$ It is well known that for $f\in C^\beta[a,b]$, $g\in C^\gamma[a,b]$ with $\beta+\gamma>1$, the generalized Lebesgue–Stieltjes integral $\int_a^bf(x)dg(x)$ exists and equals the limit of Riemann sums.
In order to integrate w.r.t. $X$, fix some $\alpha \in(1-\theta,1/2)$ and introduce the following norm: $$\begin{gathered}
{\left\lVertf\right\rVert}_{\alpha,[a,b]} = \int_a^b \left(\frac{|{f(s)}|}{(s-a)^\alpha} + \int_a^s \frac{|{f(s)-f(z)}|}{(s-z)^{1+\alpha}}dz\right)ds.\end{gathered}$$ For simplicity we will abbreviate ${\left\lVert\cdot\right\rVert}_{\alpha,t} = {\left\lVert\cdot\right\rVert}_{\alpha,[0,t]}$. Denote $\Lambda_\alpha:= \sup_{0\le s<t\le 1} |{D_{t-}^{1-\alpha}X_{t-}}(s)|$. In view of $(H)$, $\Lambda_\alpha<\infty$.
Then for any $t\in(0,1]$ and any $f$ such that ${\left\lVertf\right\rVert}_{\alpha,t}<\infty$, the integral $\int_0^t f(s) dX_s$ is well defined as a generalized Lebesgue–Stieltjes integral, and the following estimate is clear: $$\begin{gathered}
\Big|{\int_0^t f(s)dX_s}\Big|\le \Lambda_\alpha {\left\lVertf\right\rVert}_{\alpha,t}.\end{gathered}$$
\[lemma12\] \[lemma:aux\_general\] Let an adapted process $X$ satisfy conditions $(H)$, $(S)$. Then there exists an adapted process $\phi$ such that ${\left\lVert\phi\right\rVert}_{\alpha,t}<\infty$ for every $t<1$ and $$\label{rep:aux-lemma}
\lim_{t\to 1-} \int_0^t \phi_s{d}X_s = +\infty$$ almost surely.
A slight modification of the argument allows to construct an integrand $\phi$ which is additionally continuous on $[0,1)$. As this is not our primary concern here, we refer to [@vita-shev] for an idea how the modification is carried out.
Choose some $\gamma\in(1,1/\theta)$ and define $\Delta_k = K k^{-\gamma}$, $k\ge 1$, where $K = \left(\sum_{k=1}^\infty k^{-\gamma}\right)^{-1}$. Set $t_0=0$, $t_n = \sum_{k=1}^{n}\Delta_k$, $n\ge 1$. Then $t_n\uparrow 1$, $n\to\infty$.
Further define the sequence of continuously differentiable functions $g_n(x) = \sqrt{x^2 + 4^{-n}} - 2^{-n}$, $n\ge 1$. Obviously, $|x|\ge g_n(x)\ge (|x|-2^{-n})\vee 0$. Finally, fix some $\beta > \mu/\lambda$, introduce a sequence of stopping times $$\tau_n = \min\left\{t\geq t_{n-1} : |X(t) - X({t_{n-1}})| \geq n^{-\beta}\right\} \wedge t_n,$$ and set $$\phi_s = \sum_{k=1}^{\infty} k^{\beta-1}g'_{k}(X_s - X_{t_{k-1}})\textbf{1}_{[t_{k-1},\tau_k)}(s).$$
We will check that $\phi$ is as required. The finiteness of the norm ${\left\lVert\phi\right\rVert}_{\alpha,t}$ is shown exactly as in [@mish-shev-valk] and therefore will be omitted. Thanks to the change of variable integration formula for the generalized Lebesgue–Stieltjes integral (see e.g. [@Mish]), for $t\in[t_{k-1},t_{k})$ $$\begin{gathered}
\int_{t_{k-1}}^t \phi_s dX_s = k^{\beta-1} g_k(X_{t\wedge \tau_k}-X_{t_{k-1}}).\end{gathered}$$ Then for any $n\ge 1$ and $t\in[t_{n-1},t_{n})$ $$\begin{gathered}
\int_0^t\phi_s dX_s = \sum_{k=1}^{n-1} \int_{t_{k-1}}^{t_k} \phi_s dX_s + \int_{t_{n-1}}^t \phi_s dX_s \\
= \sum_{k=1}^{n-1} k^{\beta-1} g_k(X_{\tau_k}-X_{t_{k-1}}) + n^{\beta-1}g_n(X_{t\wedge \tau_n}-X_{t_{n-1}})\\
\ge \sum_{k=1}^{n-1} k^{\beta-1}\big(|X_{\tau_k}-X_{t_{k-1}}|- 2^{-k}\big) \ge \sum_{k=1}^{n-1} k^{\beta-1}|X_{\tau_k}-X_{t_{k-1}}|- \sum_{k=1}^\infty k^{\beta-1} 2^{-k}.\end{gathered}$$ In order to prove the claim, we need to show that the series $\sum_{k=1}^{n-1} k^{\beta-1}|X_{\tau_k}-X_{t_{k-1}}|$ diverges. To this end it suffices to show that, almost surely, $\tau_k <t_k$ for all $k$ large enough so that $k^{\beta-1}|X_{ \tau_k}-X_{t_{k-1}}| = k^{-1}$ eventually.
Clearly, $\tau_k= t_k$ implies $\sup_{t\in[t_{k-1},t_k]}|X_{t}-X_{t_{k-1}}|\le k^{-\beta}$, so by $(S)$, $$\mathbf{P}(\tau_k=t_k)\le K_1\exp\left\{ - K_2 k^{\beta \lambda -\gamma \mu }\right\}.$$ If the exponent near $k$ is positive, then we are done. The positivity is easily seen to be equivalent to $\gamma < \beta \lambda/\mu$. On the other hand, we must have $\gamma>1$. Due to the choice of $\beta$, we can choose $\gamma$ satisfying both requirements, thus finishing the proof.
We are ready to state the main result of this section. While its proof heavily borrows from [@mish-shev-valk; @vita-shev; @vita-shev2], we decided nevertheless to give it for two reasons. Firstly, we aimed to keep the article self-consistent. Secondly, we desired to stress all the key points of the proof in order to make sure that the assumptions are optimal.
\[thm:general\] Let an adapted process $X$ satisfy conditions $(H)$, $(S)$, and a random variable $\xi$ be such that $\xi=Z_1$ for some adapted process $Z$ such that $Z\in C^\rho[0,1]$ with $\rho > \mu/(\lambda\theta)-1$ a.s. Then there exists an adapted process $\psi$ such that ${\left\lVert\psi\right\rVert}_{\alpha,1}<\infty$ for some $\alpha\in (1-\theta,1/2)$ and $$\label{eq:represent}
\int_0^1 \psi_s{d}X_s = \xi$$ almost surely.
One might hope to get this representation result for Hölder continuous process of any order provided that $\mu/(\lambda\theta)\le 1$ (so that the restriction on $\rho$ is void). This, however, is possible only if $\theta = \mu/\lambda$, as it was explained in Remark \[rem:exponents\].
Let ${\left\{t_n,n\ge 1\right\}}\in(0,1)$ be some sequence of points such that $t_n\uparrow 1$, $n\to\infty$. We will construct an adapted process $\psi$ such that
- For all $n$ large enough $\int_0^{t_n}\psi_s dX_s = Z_{t_{n-1}}$.
- ${\left\lVert\psi\right\rVert}_{\alpha,[t_n,1]}\to 0$, $n\to\infty$.
Since $Z_{t_n}\to Z_1$, $n\to\infty$, by continuity, these properties imply .
Denote for $n\ge 1$ $\xi_n = Z_{t_{n}}$, $\Delta_n = t_{n+1}-t_n$, $\delta_n = |\xi_{n}-\xi_{n-1}|$.
We construct the process $\psi$ inductively on $[t_n,t_{n+1}]$. To this end, we take some positive sequences ${\left\{\sigma_n,n\ge 1\right\}}$ and ${\left\{\nu_n,n\ge 1\right\}}$ such that $\sigma_n\to\infty$, $n\to\infty$.
We start the construction setting $\psi_t = 0$ for $t\in[0,t_1]$. Further, assume that $\psi$ is constructed on $[0,t_n)$ and denote $V_t = \int_0^t \psi_s dX_s$. The construction will depend on whether some event $A_n\in \mathcal F_{t_n}$, which will be specified later, or its complement $B_n=\Omega\setminus A_n$ holds.
Case 1: $\omega \in A_n$. Thanks to Lemma \[lemma:aux\_general\], there exists a process ${\left\{\phi_t,t\in [t_n,t_{n+1}]\right\}}$ such that $\int_{t_n}^t \phi_s dX_s \to +\infty$, $t\to t_{n+1}-$. Define $v_n = V_{t_n}-\xi_{n}$, $$\tau_n = \inf{\left\{t\ge t_n: \int_{t_n}^t \phi_s dX_s \ge |v_n|\right\}}$$ and set $$\psi_s = \phi_s \operatorname{sign}v_n\,{\mathbbm{1}_{[t_n,\tau_n]}}(t),\ t\in[t_n,t_{n+1}].$$ It is clear that $\int_{t_n}^{t_{n+1}} \phi_s dX_s = v_n$, hence, $V_{t_{n+1}} = V_{t_n} + v_n = \xi_n$.
Case 2: $\omega\in B_n$. Similarly to the proof of Lemma \[lemma:aux\_general\], define $g_n(x) = \sqrt{x^2 + \nu_n^2} - \nu_n$ so that $g_n\in C^\infty(\mathbb{R})$, $|x|\ge g_n(x)\ge (|x|-\nu_{n})\vee 0$. Introduce the stopping time $$\tau_n = \inf{\left\{t\ge t_n: \sigma_n g_n(X_{t}-X_{t_n})\ge \delta_n\right\}}\wedge t_{n+1}$$ and set $$\psi_s = \sigma_n g'_n(X_t-X_{t_n}) \operatorname{sign}(\xi_n-\xi_{n-1}){\mathbbm{1}_{[t_n,\tau_n]}}(t), t\in[t_n,t_{n+1}].$$
By the change of variable formula for the generalized Lebesgue–Stieltjes integral, $$V_{t_{n+1}}-V_{t_n} = \int_{t_n}^{t_{n+1}} \psi_s dX_s = \sigma_n g_n(X_{t_{n+1}\wedge \tau_n} - X_{t_n}).$$ Therefore, $V_{t_{n+1}}-V_{t_n} = \xi_n -\xi_{n-1}$ provided that $\tau\le \tau_n$. In turn, thanks to the properties for $g_n$, the latter holds if $\sigma_n\sup_{t\in[t_{n-1},t_n]}|X_{t}-X_{t_{n-1}}|\le \delta_n + \sigma_n\nu_n$. In view of this, define $$A_{n+1} = B_n\cap {\left\{\sup_{t\in[t_{n-1},t_n]}|X_{t}-X_{t_{n-1}}|\le \delta_n\sigma_n^{-1} + \nu_n\right\}}, n\ge 1,$$ and $A_1 = \Omega$.
Now we identify conditions under which this construction works. First note that for $(\Phi1)$ it is suffices to ensure that $$\label{eq:borelcantelli}
{ \mathbf{P}}\left(\limsup_{n\to\infty} A_n\right)=0.$$ Indeed, let $N = N(\omega) = \max{\left\{n\ge 1: \omega\in A_n\right\}}$. Then by construction, $V_{t_{N+1}}=\xi_N$. Moreover, since $\omega\in B_n$ for all $n\ge N+1$, we have $V_{t_{n+1}}-V_{t_n}=\xi_n-\xi_{n-1}$ for all $n\ge N+1$, whence $(\Phi1)$ follows.
Now turn to $(\Phi2)$. Write for $n\ge N$ $${\left\lVert\psi\right\rVert}_{\alpha, [t_n,1]} = I_1 + I_2,$$ where $$I_1 = \int_{t_n}^1\frac{{\left\lvert\psi_t\right\rvert}}{(t-t_n)^\alpha} {d}s,\quad I_2 = \int_{t_n}^1\int_{t_n}^{t}\frac{{\left\lvert\psi_t-\psi_s\right\rvert}}{(t-s)^{\alpha+1}}{d}s\,{d}t.$$ Estimate, taking into account that ${\left\lvertg'_k\right\rvert}\le 1$, $$\begin{aligned}
I_1 & = \sum_{k=n}^{\infty} \int_{t_k}^{t_{k+1}}\frac{{\left\lvert\psi_t\right\rvert}}{(t-t_n)^{\alpha}}{d}t \le \sum_{k=n}^{\infty} \int_{t_k}^{t_{k+1}} \frac{\sigma_k}{(t - t_{k})^{\alpha}} \le C\sum_{k=n}^{\infty} \sigma_k\Delta_k^{1-\alpha}.\end{aligned}$$ Further, denoting $\psi(t,s) = {\left\lvert\psi_t-\psi_s\right\rvert}(t-s)^{-\alpha-1}$ and taking into account that $\psi_t = 0$ for $t\in(\tau_k,t_{k+1}]$, write $$\begin{aligned}
I_2 & = \sum_{k=n}^\infty \int_{t_k}^{\tau_{k}} \int_{t_n}^{t} \psi(t,s){d}s \,{d}t
+ \sum_{k=n}^\infty \int_{t_{k}}^{t_{k+1}} \int_{t_n}^{t} \psi(t,s){d}s \,{d}t\\
& = \sum_{k=n}^\infty \int_{t_k}^{\tau_{k}} \int_{t_n}^{t_k} \psi(t,s){d}s \,{d}t
+ \sum_{k=n}^\infty \int_{\tau_{k}}^{t_{k+1}} \int_{t_n}^{\tau_{k}} \psi(t,s){d}s \,{d}t\\
& + \sum_{k=n}^\infty \int_{t_k}^{\tau_{k}} \int_{t_k}^{t} \psi(t,s){d}s \,{d}t
=: J_1 + J_2 + J_3.\end{aligned}$$ Estimate the terms separately: $$\begin{aligned}
&J_1\le 2\sum_{k=n}^{\infty} \sigma_k\int_{t_k}^{\tau_{k}} \int_{t_n}^{t_k}\frac{{d}s\, {d}t}{(t-s)^{\alpha + 1}} \le C\sum_{k=n}^{\infty} \sigma_k\int_{t_k}^{t_{k+1}}\frac{{d}t}{(t-t_n)^{\alpha}}
\le C \sum_{k=n}^{\infty}\sigma_k {\Delta}_k^{1-\alpha}.\end{aligned}$$ Similarly, $J_2 \le C \sum_{k=n}^{\infty}\sigma_k {\Delta}_k^{1-\alpha}$. To estimate $J_3$, denote $r_k = \nu_k^{1/\theta}$ and decompose $$J_3 = \sum_{k=n}^\infty\int_{t_k}^{\tau_{k}}\left(\int_{t_k}^{t-r_k}+\int_{t-r_k}^{t}\right)\psi(t,s){d}s \,{d}t=:J_{31}+J_{32}.$$ Then $$\begin{aligned}
J_{31} &\le 2\sum_{k=n}^\infty \sigma_k\int_{t_k}^{\tau_{k+1}}\int_{t_k}^{t-r_k}(t-s)^{-\alpha-1}{d}s\, {d}t\\&\le
C\sum_{k=n}^\infty \sigma_k\int_{t_k}^{\tau_{k}}r_k^{-\alpha}{d}t = C\sum_{k=n}^{\infty} \sigma_k \Delta^{\vphantom{H}}_k \nu_k^{-\alpha/\theta}.\end{aligned}$$ To estimate $J_{32}$, note that ${\left\lvertg_k''(x)\right\rvert}\le \nu_k^{-1}$. Therefore, $$\begin{aligned}
&J_{32}\le \sum_{k=n}^{\infty}
\sigma_k \nu_k^{-1}\int_{t_k}^{\tau_{k}} \int_{t-r_k}^{t}\frac{{\left\lvertX_t-X_s\right\rvert}}{(t-s)^{\alpha+1}}{d}s \,{d}t\\
& \le C\sum_{k=n}^{\infty}\sigma_k \nu_k^{-1}\int_{t_k}^{\tau_{k}} \int_{t-r_k}^{t}(t-s)^{\theta-\alpha-1}{d}s \,{d}t
\\&\le C \sum_{k=n}^{\infty} \sigma_k \nu_k^{-1} \Delta^{\vphantom{H}}_k r_k^{\theta-\alpha} = C\sum_{k=n}^{\infty} \sigma_k^{\vphantom{H}} \Delta^{\vphantom{H}}_k \nu_k^{-\alpha/\theta}.\end{aligned}$$ Note that the estimation of the summand should be modified for $r_k>\Delta_k$, i.e. $\nu_k>\Delta_k^{\theta}$. In this case the summand is bounded by $\sigma_k \nu_k^{-1}\Delta_k^{\theta-\alpha + 1}<\sigma_k^{\vphantom{H}} \Delta_k^{\vphantom{H}} \nu_k^{-\alpha/\theta}$, which leads to the same estimate.
Summing up, we get that ${\left\lVert\psi\right\rVert}_{\alpha, [t_n,1]}\to 0$, $n\to\infty$, iff $$\label{eq:series}
\sum_{n=1}^{\infty} \sigma_n^{\vphantom{H}}\Delta_n^{1-\alpha}<\infty\ \text{ and }\ \sum_{n=1}^{\infty} \sigma_n^{\vphantom{H}} \Delta^{\vphantom{H}}_n \nu_n^{-\alpha/\theta}<\infty.$$
Let us discuss the choice of parameters. First we need to ensure . Suppose that $\kappa\in (0,\rho)$. Then the Hölder assumption $Z\in C^\rho$ implies that $\delta_n = o(\Delta_n^\kappa)$, $n\to\infty$ a.s. Setting $\nu_n = \Delta_n^\kappa\sigma_n^{-1}$, we get that $$\limsup_{n\to\infty} A_n\subset \limsup_{n\to\infty} C_n,$$ where $$C_{n} = {\left\{\sup_{t\in[t_{n-1},t_n]}|X_{t}-X_{t_{n-1}}|\le 2\Delta_n^\kappa\sigma_n^{-1}\right\}},\ n\ge 1.$$ Then, but virtue of the Borel–Cantelli lemma, it suffices to ensure that $\sum_{n=1}^{\infty} { \mathbf{P}}(C_n)<\infty$. In view of the small ball estimate $(S)$, $${ \mathbf{P}}(C_n) \le K_1\exp{\left\{-2^{-\mu}K_2 \Delta_n^{\mu-\kappa\lambda}\sigma_n^{\lambda}\right\}}.$$ Taking $\sigma_n = n^{\epsilon}\Delta_n^{\kappa-\mu/\lambda}$ with $\epsilon>0$, we get $\sum_{n=1}^{\infty} { \mathbf{P}}(C_n)<\infty$, as required.
Now turn to . With the above choice of $\sigma_n$ and $\nu_n$, they transform to $$\sum_{n=1}^{\infty} n^{\epsilon}\Delta_n^{1+ \kappa - \mu/\lambda -\alpha}<\infty$$ and $$\sum_{n=1}^{\infty} n^{\tau} \Delta^{1+\kappa - \mu(1+\alpha/\theta)/\lambda}_n<\infty,$$ where $\tau = \epsilon(1+\alpha/\theta)$. Taking $\Delta_n=2^{-n}$, it is enough to make the both exponents near $\Delta_n$ positive. Since $\mu/(\lambda\theta)\ge 1$, the second exponent is smaller, so we end up with the requirement that $$1+\kappa - \mu(1+\alpha/\theta)/\lambda>0.$$ The other restrictions we have are $\kappa<\rho$ and $\alpha>1-\theta$. So the choice of $\kappa$ and $\alpha$ is possible iff $$1+\rho - \mu(1+(1-\theta)/\theta)/\lambda>0,$$ which is easily seen to be equivalent to $\rho> \mu/(\lambda\theta)-1$. The proof is now complete.
\[rem:equivalence\] With our approach, the assumption that $Z$ is Hölder continuous is unavoidable. Indeed, in order for the argument to work, one must have $\delta_n \sigma_n^{-1} = O(\Delta_n^\theta)$, $n\to\infty$, as in the opposite case we would have a contradiction with $(H)$. On the other hand, the series $\sum_{n=1}^{\infty} \sigma_n \Delta_n^{1-\alpha}$ must converge. Consequently, $\sum_{n=1}^{\infty} \delta_n \Delta_n^{1-\theta-\alpha}<\infty$, whence $\delta_n = o(\Delta_n^{\theta+\alpha-1})$, $n\to\infty$, as claimed.
Assumption $(S)$ is not optimal for establishing Theorem \[thm:general\]. It is easy to see that the conclusion is of “probability zero” spirit, in particular, it does not change with switching to an equivalent measure. Assumption $(S)$ is more delicate and in general will not hold for an equivalent measure. It is possible to formulate a relevant “almost sure” assumption, for example:
- There exists a number $a>0$ such that for any sequence of points ${\left\{t_n,n\ge1\right\}}\in(0,1)$ such that $t_{n}\uparrow 1$, $n\to\infty$, $t_{n+1} - t_n\sim K n^{-\gamma}$ with some $K>0$, $\gamma>1$, it holds $$\lim_{n\to\infty} n^{\gamma a}\sup_{t\in[t_{n-1},t_n]}|X_{t}-X_{t_{n-1}}|=+\infty.$$
It is easy to check that if one assumes $(S')$ instead of $(S)$, then Theorem \[thm:general\] holds with $\rho_0 = a/\theta-1$. Also, similarly to the proof of , $(S')$ implies $(S)$ with any $a>\mu/\lambda$.
However, $(S')$ is not easy to check. Alternatively, one can assume that the distribution of $X$ is equivalent to that of a process satisfying $(S)$, which seems more natural. However, one needs to ensure that the adaptedness is preserved with the change of measure, e.g. by using some version of the Girsanov theorem.
Small ball probability estimates and representation results for Gaussian processes {#sec2}
==================================================================================
In this section we first establish the small ball property for Gaussian processes satisfying two-sided estimates on the incremental variance and preserving the sign of incremental covariance. We remark that similar assumptions on the incremental variance were imposed in [@azmoo-vita], however, the assumptions on the covariance differ significantly, so the findings are different. Using the small ball estimates, we derive representation results for such Gaussian processes. Finally we give the examples of the processes satisfying these conditions, including integral transforms with Volterra kernels of a Wiener process and of an fBm.
Small ball property for Gaussian properties with variance distance satisfying two-sided estimates
-------------------------------------------------------------------------------------------------
Let $X=\{X_t, t\in [0,1]\}$ be a centered Gaussian process on a finite interval $[0,1]$, whose variance distance ${\mathbf{E}}|X_t-X_s|^2$ satisfies the following two-sided power bounds:
- There exist $H_1\in(0,1]$ and $C_1>0$ such that for any $s,t\in[0,1]$ $${\mathbf{E}}(X_t-X_s)^2\ge C_1|t-s|^{2H_1}.$$
- There exist $H_2\in(0,1]$ and $C_2>0$ such that for any $s,t\in[0,1]$ $${\mathbf{E}}(X_t-X_s)^2\le C_2|t-s|^{2H_2}.$$
Clearly, $H_1\ge H_2$. In the particular case where $H_1=H_2$, such process is called a quasi-helix, see [@kahane; @kahanebook].
Furthermore, assume that the increments of $X$ are either positively or negatively correlated. More precisely, we assume one of the following conditions:
- For any $s_1,t_1,s_2,t_2\in[0,1]$, $ s_1\leq t_1\leq s_2\leq t_2$ $$\pm {\mathbf{E}}(X_{t_1}-X_{s_1})(X_{t_2}-X_{s_2})\geq 0.$$
\[remplusminus\] It is worth to mention that $(A1)$ and $(B^+)$ imply that $H_1\ge 1/2$. Indeed, write for any $n\ge 1$ $$\begin{gathered}
C_2\ge {\mathbf{E}}(X_1-X_0)^2 = \sum_{k=1}^n {\mathbf{E}}(X_{k/n}-X_{(k-1)/n})^2\\ + \sum_{i\neq j} {\mathbf{E}}[(X_{i/n}-X_{(i-1)/n})(X_{j/n}-X_{(j-1)/n})]
\ge C_1 n^{1-2H_1},\end{gathered}$$ whence the claim follows by letting $n\to \infty$. Similarly, $(A2)$ and $(B^-)$ imply that $H_2\le 1/2$.
Further, introduce some notations for different constants. More precisely, denote $$\begin{gathered}
C_0=\frac{64}{C_1},\
C_3=\frac{C_0^{H_1/2}}{4^{H_1}},\ C_4=\left(16 C^2_2 C_0^{({H_2+1})/{2H_1}}\right)^{-1},\
C_5 = \left(32 C^2_2 C_0^{({4H_2+1})/{2H_1}}\right)^{-1}.
$$
The following theorem establishes an upper bound for small deviations of the process $X$.
\[lemma1\] Let $X=\{X_t, t\in [0,1]\}$ be a Gaussian process satisfying $(A1)$ and $(A2)$.
- If $(B^+)$ holds, then for any $\varepsilon\in(0,C_3]$ $$\label{equa}
\mathbf{P}\left\{\sup_{0\leq s\leq t\leq 1}|X_t-X_s|\leq \varepsilon\right\}\leq\exp\left\{-C_4\varepsilon^{4-(2H_2+2)/H_1}\right\}.$$
- If $(B^-)$ holds, then for any $\varepsilon\in(0,C_3]$ $$\mathbf{P}\left\{\sup_{0\leq s\leq t\leq 1}|X_t-X_s|\leq \varepsilon\right\}\leq\exp\left\{-C_5\varepsilon^{4-({4H_2+1})/{H_1}}\right\}.$$
\[remark 2.2\] The small ball estimates of Theorem \[lemma1\] are useful only whenever the exponents of $\varepsilon$ are negative. It is easy to see that in case (1) this happens if $H_2> 2H_1-1$, in case (2), if $H_2> H_1-1/4$. Recall also that in case (1), $H_1\ge 1/2$, in case (2), $H_2\le 1/2$; in both cases $H_2\le H_1$.
We use Theorem 4.4. from [@li-shao]. According to this result, for any any $a\in(0,1/2]$ and any $\varepsilon>0$ the inequality $$\label{equat4} \mathbf{P}\left\{\sup_{0\leq s \leq t\leq 1}|X_t-X_s|\leq \varepsilon\right\}\leq\exp\left\{-\frac{\varepsilon^{4}}{16a^2\sum_{2\leq i,j\leq a^{-1}}({\mathbf{E}}(\xi_i\xi_j))^2}\right\}$$ holds provided that $$\label{32eps2}
a\sum_{2\leq i\leq a^{-1}}{\mathbf{E}}\xi_i^2\geq 32\varepsilon^2,$$ where $\xi_i=X_{ia}-X_{(i-1)a}$. Thanks to $(A1)$ and $(A2)$, $$\label{equat5}
C_1 a^{2H_1}\leq {\mathbf{E}}\xi^2_i\leq C_2 a^{2H_2}.$$ Therefore, $a\sum_{2\leq i\leq a^{-1}}{\mathbf{E}}\xi_i^2\geq C_1 a\left(\left[a^{-1}\right]-1\right) a^{2H_1}\geq C_1 \left(1-2a\right)a^{2H_1}$. If $a\le\frac14$, inequality holds whenever $a^{2H_1}\geq 64{C_1}^{-1}\varepsilon^2=C_0\varepsilon^2.$ Thus, we get the estimate by setting $a = C_0^{1/2H_1}\varepsilon^{1/H_1}$ (the assumption $\varepsilon\le C_3$ entails that $a\le 1/4$).
\(1) If $(B^+)$ holds, then $$\begin{gathered}\label{equat2}
\sum_{2\leq i,j\leq a^{-1}}({\mathbf{E}}(\xi_i\xi_j))^2\leq \max_{2\le i,j\le a^{-1}}|{\mathbf{E}}(\xi_i\xi_j)| \sum_{2\leq i,j\leq a^{-1}} {\mathbf{E}}\xi_i \xi_j\\ \leq C_2 \max_{2\le i\le a^{-1}}{\mathbf{E}}\xi_i^2 \ {\mathbf{E}}\bigg(\sum_{2\le i\le a^{-1}}\xi_i\bigg)^2 = C_2 a^{2H_2} {\mathbf{E}}(X_{a[a^{-1}]}-X_{a})^2\le C^2_2 a^{2H_2}.
\end{gathered}$$ Substituting into , we arrive at the desired estimate.
\(2) If $(B^-)$ holds, then $$\begin{gathered}\label{equat6}\sum_{2\leq i,j\leq a^{-1}}\big({\mathbf{E}}(\xi_i\xi_j)\big)^2
\leq \max_{2\leq i,j\leq a^{-1}}|{\mathbf{E}}(\xi_i\xi_j)| \left(\sum_{2\leq i\leq a^{-1}} {\mathbf{E}}\xi_i^2 -2\sum_{2\leq i<j\leq a^{-1}} {\mathbf{E}}\xi_i \xi_j\right)\\
\le C_2 a^{2H_2} \left(2\sum_{2\leq i\leq a^{-1}} {\mathbf{E}}\xi_i^2-\sum_{2\leq i,j\leq a^{-1}} {\mathbf{E}}\xi_i \xi_j\right)\\
=C_2 a^{2H_2}\left(2\sum_{2\leq i\leq a^{-1}} {\mathbf{E}}\xi_i^2-{\mathbf{E}}(X_{a[a^{-1}]} -X_{a})^2\right)
\le 2C_2^2 a^{4H_2-1}.
\end{gathered}$$ Plugging into and recalling that $a = C_0^{1/2H_1}\varepsilon^{1/H_1}$, we get the desired estimate.
As a corollary, we establish a small ball property on any interval.
\[smallballprop\] Let $X=\{X_t, t\in [t_0,t_0+\Delta]\}$ be a Gaussian process satisfying $(A1)$ and $(A2)$.
- If $(B^+)$ holds, then for any $\varepsilon\in(0,C_3\Delta^{H_1}]$ $$\mathbf{P}\left\{\sup_{t_0\leq s\leq t\leq t_0+\Delta}|X_t-X_s|\leq \varepsilon\right\}\leq\exp\left\{-C_4\varepsilon^{4-(2H_2+2)/H_1}\Delta^{2-2H_2}\right\}.$$
- If $(B^-)$ holds, then for any $\varepsilon\in(0,C_3\Delta^{H_1}]$ $$\mathbf{P}\left\{\sup_{t_0\leq s\leq t\leq t_0+\Delta}|X_t-X_s|\leq \varepsilon\right\}\leq\exp\left\{-C_5\varepsilon^{4-({4H_2+1})/{H_1}}\Delta\right\}.$$
Define $X'_t = \Delta^{-H_1} (X_{t_0+\Delta t}-X_{t_0})$, $t\in[0,1]$. Then $X'$ satisfies $(A1)$ on $[0,1]$. It also satisfies (A2), but with a different constant, namely, $C_2' = C_2 \Delta^{2(H_2-H_1)}$. Setting ${\varepsilon}' = {\varepsilon}\Delta^{-H_1}$ and applying Theorem \[lemma1\], we arrive at the required statement.
Representation results for Gaussian processes {#sec4}
=============================================
Now we apply the obtained representation results to processes from Section \[equat1\]. We could omit the following auxiliary result and derive the required results directly from Theorem \[thm:general\]. However, we give it not only for the sake of completeness, but also to identify relation between assumptions $(H)$, $(S)$ and those from Section \[sec2\].
Assume that an adapted Gaussian process $X$ satisfies conditions $(A1)$, $(A2)$, $(B^+)$ with $0<2H_1-1< H_2\le H_1$. Then there exists an adapted process $\phi$ such that ${\left\lVert\phi\right\rVert}_{\alpha,t}<\infty$ for every $t<1$ and $$\lim_{t\to 1-} \int_0^t \phi_s{d}X_s = +\infty$$ almost surely.
It is not possible to state similar results for processes satisfying $(A1)$, $(A2)$ and $(B^{-})$, since $(H)$ with $\theta>1/2$ requires that $H_2>1/2$.
From $(A2)$ it follows that $X$ is Hölder continuous of any order less than $H_2$, so $(H)$ holds. Proposition \[smallballprop\] (1) yields $$\label{eq:smallballrem}
\mathbf{P}\left\{\sup_{s\leq t\leq s+\Delta}|X_t-X_s|\leq \varepsilon\right\}\leq\exp\left\{-C_4\varepsilon^{-\lambda}\Delta^{\mu}\right\}.$$ with $\lambda=(2H_2+2)/H_1 - 4 = 2(H_2+1-2H_1)>0$, $\mu = 2-2H_2>0$ for all ${\varepsilon}\in (0,C_3 \Delta^{H_1}]$. Setting $K_1 = \exp{\left\{C_4 C_3^{-\lambda}\right\}}$, $K_2 = C_4$, we obtain : for ${\varepsilon}\in (0,C_3 \Delta^{\mu/\lambda}]$ use and the observation that $H_1 \le \mu/\lambda$, for ${\varepsilon}>0$, the left-hand side exceeds $1$. Thus, the statement follows from Lemma \[lemma:aux\_general\].
The following result is a consequence of Theorem \[equat1\].
\[cor:main\] Assume that an adapted Gaussian process $X$ satisfies conditions $(A1)$, $(A2)$, $(B^+)$ with $0<2H_1-1< H_2\le H_1$. Let also a random variable $\xi$ be such that $\xi=Z_1$ for some adapted process $Z\in C^\rho[0,1]$ with $\rho > \rho_0$, where $$\label{rho0}
\rho_0 = \frac{(1+H_2)(H_1-H_2)}{H_2+1-2H_1}.$$ Then there exists an adapted process $\psi$ such that ${\left\lVert\psi\right\rVert}_{\alpha,1}<\infty$ for some $\alpha\in (1-H_2,1/2)$ and $$\int_0^1 \psi_s{d}X_s = \xi$$ almost surely.
Consider the case where $H_1=H_2$. The processes $X$ satisfying $(A1)$ and $(A2)$ are called quasi-helices, see [@kahane; @kahanebook]. For such processes $\rho_0=0$, so all values of $\rho$ are possible, which agrees with the results of [@mish-shev-valk; @vita].
It is natural to study the representation question in the case where $\xi=Z_1$, and the process $Z$ has the same regularity as $X$. In the general case this translates to the requirement that $\theta(1-\theta)>\mu/\lambda$. In the particular case of Corollary \[cor:main\] this translates to the inequality $$H_1 < \frac{2H_2(H_2+1)}{1+3H_2}.$$ A simpler sufficient condition for this is that $H_1\le 4H_2/5 - 1/5$.
Examples
--------
Here we present some examples of processes which satisfy the assumptions $(A1)$, $(A2)$, $(B^+)$ so that the representation result of Theorem \[cor:main\] is true. We remark that it is enough to require the properties to hold on a subinterval $[t_0,1]$.
### Subfractional Brownian motion
Recall the definition of subfractional Brownian motion: this is a centered Gaussian process $G^H = {\left\{G^H_t,t\ge 0\right\}}$ with the covariance function $${\mathbf{E}}G^H_t G^H_u = t^{2H} + s^{2H} -\frac{1}{2}\left((t+s)^{2H} + {\left\lvertt-s\right\rvert}^{2H}\right),$$ where $H\in(0,1)$ is the self-similarity parameter of $G^H$, a counterpart of the Hurst parameter of fractional Brownian motion. It is easy to check that the increments of $G^H$ are not stationary.
It is proved in [@bojdecki] that $G^H$ satisfies properties $(A1)$, $(A2)$ with $H_1=H_2=H$, so $G^H$ is a quasi-helix. It also satisfies $(B^+)$ for $H\in (1/2,1)$, and $(B^-)$ for $H\in(0,1/2)$. Consequently, the conclusion of Theorem \[cor:main\] holds for $G^H$ with any $\rho>0$.
### Bifractional Brownian motion
The bifractional Brownian motion is a centered Gaussian process $B^{H,K} = {\left\{B^{H,K}_t,t\ge 0\right\}}$ with the covariance function $$R(t,s) = {\mathbf{E}}B^{H,K}_t B^{H,K}_u = \frac{1}{2^K}\left( \left(t^{2H} + s^{2H}\right)^K - {\left\lvertt-s\right\rvert}^{2HK}\right),$$ where $H\in(0,1)$, $K\in(0,1]$. This is an $HK$-self-similar process with non-stationary increments, which is also a quasi-helix, that is, it satisfies $(A1)$ and $(A2)$ with $H_1=H_2=HK$ (see [@houdre]).
Concerning $(B^+)$, assume that $HK>1/2$ and write $$\begin{gathered}
\frac{\partial R(t,s)}{\partial t\partial s} = \frac{2HK}{2^K}\left(2H(K-1)t^{2H-1}s^{2H-1}\left(t^{2H}+s^{2H}\right)^{K-1} + (2HK-1){\left\lvertt-s\right\rvert}^{2HK-2} \right) \\
= C_1 s^{2HK-2}\left( |u-1|^{2HK-2} - C_2 u^{2H-1}(u^{2H}+1)\right) \sim C_1 s^{2HK-2} |u-1|^{2HK-2}, u\to 1.\end{gathered}$$ where $u=t/s$ and $C_1,C_2$ are some positive constants. Hence, $(B^+)$ holds on some interval $[t_0,1]$. As a result, we have Theorem \[cor:main\] for $B^{H,K}$ with any $\rho>0$.
### Volterra integral transform of Wiener process
Let $W=\{W(t), t\geq 0\}$ be a standard Wiener process. Consider the processes of the form $X(t) = \int_0^t K(t,s) d W(s)$ with non-random kernels $K=\{K(t,s):[0,1]^2\rightarrow \mathbb{R}\}$ such that $K(t,\cdot)\in L^2{[0,t]}$ for any $t\in [0,1]$. Our goal is to establish the conditions on the kernel $K$ that supply $(A1)$, $(A2)$, and $(B^+)$, so that small ball property of Proposition \[smallballprop\] is in place but only on the intervals separated from $0$. In what follows the constant $r\in[0,1/2)$ is fixed.
\[Th4.1\] Let the kernel $K=\{K(t,s), t,s \in[0,1]\}$ satisfy conditions
- The kernel $K$ is non-negative on $[0,1]^2$ and for any $s\in [0,1]$ $K(\cdot,s)$ is non-decreasing in the first argument.
- There exist constants $D_i>0, i=2,3$ and $1/2<H_2<1$ such that $$|K(t_2,s) - K(t_1,s)| \leq D_2 |t_2-t_1|^{H_2 }s^{-r},\;s,\; t_1,\;t_2 \in [0,1]$$ and $$\ K(t,s)\leq D_3(t-s)^{H_2-1/2}s^{-r},$$
and at least one of the following conditions
- There exist constants $D_1>0$ and $H_1\geq H_2 $ such that $$D_1|t_2-t_1|^{H_1}s^{-r}\leq|K(t_2,s) - K(t_1,s)|,\;s,\; t_1,\;t_2 \in [0,1];$$
- There exist constants $D_1>0$ and $H_1\geq H_2 $ such that $$K(t,s)\geq D_1(t-s)^{H_1-1/2}s^{-r},\;s,\; t_1,\;t_2 \in [0,1].$$
Then the Gaussian process $X(t) = \int_0^t K(t,s) dW(s)$, satisfies conditions $(A1)$, $(A2)$, and $(B^+)$ on any subinterval $[1-\delta, 1]$ for $0<\delta<1$ with powers $H_1, H_2$.
Note that the process $X$ is correctly defined due to condition $K(t,\cdot)\in L^2{[0,t]}$ for any $t\in [0,1]$, and it has a covariance function of the form $${\mathbf{E}}X({t_1}) X({t_2}) = \int_0^{t_1\wedge t_2 } K(t_1,z) K(t_2,z) dz.$$ Further, for any $ 0 < t_1 < t_2 < 1$ $$X_{t_2} - X_{t_1} = \int_0^{t_1}(K(t_2,z)- K(t_1,z))dW(z) + \int_{t_1}^{t_2} K(t_2,z) dW(z).$$ Therefore, for any $s_1,t_1,s_2,t_2\in[0,1]$, $ s_1\leq t_1\leq s_2\leq t_2$ $$\begin{gathered}\label{eqaut11}
{\mathbf{E}}(X_{t_1}-X_{s_1})(X_{t_2}-X_{s_2}) = \int_0^{s_1} (K(t_1,z)- K(s_1,z)) (K(t_2,z)- K(s_2,z)) dz\\+\int_{s_1}^{t_1} K(t_1,z)(K(t_2,z)- K(s_2,z)) dz.
\end{gathered}$$ If the kernel is non-negative and non-decreasing in the 1st argument, the right-hand side of is non-negative so, condition $(B^+)$ holds. Furthermore, for any $0\leq s<t\leq 1$ $$\begin{gathered}\label{eqaut12}
{\mathbf{E}}(X(t)-X(s))^2 = \int_0^{s} (K(t,z)- K(s,z))^2dz+\int_s^t K^2(t,z) dz.
\end{gathered}$$ Let $\delta\in [0,1]$ be fixed. Suppose that $s>1-\delta$ and estimate the right-hand side of from above: $$\begin{gathered}\label{eqaut15}
\int_0^{s} (K(t,z)- K(s,z))^2dz+\int_s^t K^2(t,z) dz\leq
D_2^2|t-s|^{2H_2}\int_0^sz^{-2r}dz\\+D_3^2 \int_s^tz^{-2r}(t-z)^{2H_2-1}dz \leq \frac{D_2^2}{1-2r}|t-s|^{2H_2}+D_3^2(1-\delta)^{-2r}|t-s|^{2H_2}=D_4|t-s|^{2H_2},
\end{gathered}$$ where $D_4=\frac{D_2^2}{1-2r}+D_3^2(1-\delta)^{-2r}$. The estimate under condition $(B3,a)$ is evident: $$\begin{gathered}\label{eqaut16}
\int_0^{s} (K(t,z)- K(s,z))^2dz+\int_s^t K^2(t,z) dz\geq D_1^2|t-s|^{2H_1}\int_0^{1-\delta}z^{-2r}dz\\ \geq (1-\delta)^{1-2r}\frac{D_1^2}{1-2r}|t-s|^{2H_1}.
\end{gathered}$$ The estimate under condition $(B3,b)$ is also simple: $$\begin{gathered}\label{eqaut17}
\int_0^{s} (K(t,z)- K(s,z))^2dz+\int_s^t K^2(t,z) dz\geq D_1^2 \int_s^t (t-z)^{2H_1- 1}z^{-2r}dz\\ \geq (1-\delta)^{1-2r}\frac{D_1^2}{1-2r}|t-s|^{2H_1}.
\end{gathered}$$ Proof follows immediately from –.
Let $H\in(1/2,1)$ and consider the kernel $$K(t,s)=C_Hs^{1/2-H}\varphi(s)\int_s^tu^{H-1/2}(u-s)^{H-\frac32}du,$$ where $\varphi=\{\varphi(s), s\in[0,1]\}$ is nonnegative measurable function satisfying assumption $0<k<\varphi(s)<K$, $C_H>0$ is some constant. Then the kernel is nonnegative and increasing in the first variable, so, condition $(B^+)$ holds and we can state with evidence that for any $0< z <s<t\leq 1$ $$|K(t,z)-K(s,z)|\leq C_H K s^{1/2-H}t^{H-1/2}(t-s)^{H-1/2}\leq C_H K s^{1/2-H}(t-s)^{H-1/2},$$ so, we can put $r=H-1/2$ and $H_2=H-1/2$. Moreover, the bound from below has the form $$K(t,s)\geq C_Hk(t-s)^{H-1/2}\geq C_Hks^{1/2-H}(t-s)^{H-1/2},$$ and condition $H_2> 2H_1-1$ is satisfied since $H_1=H_2=H$. Therefore process $X(t) = \int_0^t K(t,s) d W(s)$ has the properties supplying small ball bound. In the case when $\varphi(s)=1$ and the constant $C_H$ is defined correspondingly, process $\int_0^tK(t,s)dW(s)$ is a fractional Brownian motion. Now we can see that the small ball property of fBm is not a consequence of its stationary increments as one can deduce from previous results ([@li-shao], for example), but of its property to be a quasi-helix.
### Volterra integral transform of fractional Brownian motion
Consider now fractional Brownian motion $B_H=\{B_H(t), t\in[0,1]\}$, the kernel $K=\{K(t,s):[0,1]^2\rightarrow \mathbb{R}\}$ and create Volterra fractional process of the form $$X(t) = \int_0^t K(t,s) dB_H(s)$$ that exists under sufficient condition $$\int_0^t\int_0^t |K(t,u) K(t,v)| |u-v|^{2H-2}du dv<\infty, t\in[0,1].$$ One of the simplest examples of Volterra fractional processes is a fractional Ornstein-Uhlenbeck process $X=(X_t, t \in [0,T])$ satisfying the equation $$X_t = X_0 + a \int_0^t X_s ds + B_t^H
\label{OUeq}$$ with $a\in \mathbb{R}$. For $H>\frac{1}{2}$ the unique solution of equation \[OUeq\] admits the representation $$X_t = X_0 e^{at}+ e^{at}\int_0^t e^{-as}dB_s^H.$$ Considering the particular case $X_0=0 $, we get the process $X_t = \int_0^t e^{a(t-s)} dB_s^H$ with kernel $K(t,s) = e^{a(t-s)}.$
\[Th4\] Let $H>\frac{1}{2}$ and let the kernel $K=\{K(t,s), s,t\in [0,1]\}$ satisfy $(B1)$ and the following conditions:
- The kernel $K$ is bounded, i.e. there are constants $k$ and $K$ such that $0 < k \le K(t,s) \leq K$ for any $s,t \in [0,T]$;
- There exist constants $C>0$ and $H_3 \geq 1/2$ for which $|K(t_2,s) - K(t_1,s)| \leq C |t_2-t_1|^{H_3}$, $s$, $t_1$, $t_2 \in [0,1].$
Then the Gaussian process $X$ with kernel $K$, $X_t = \int_0^t K(t,s) dB_s^H$, satisfies conditions $(A1)$, $(A2)$ and $(B^+)$ with exponents $H_1 = H_3\wedge H$, $H_2 = H_3$.
Firstly, we note that the process $X$ is correctly defined due to condition $(B1)$ and has a covariance function of the form ( see, e.g., [@Mish], [@NVV]) $${\mathbf{E}}X_{t_1} X_{t_2} = H(2H-1) \int_0^{t_1} \int_0^{t_2} K(t_1,s) K(t_2,v) |s-v|^{2H-2}ds dv.$$ Further, for any $ 0 < t_1 < t_2 < 1$ $$X_{t_2} - X_{t_1} = \int_0^{t_1}(K(t_2,s)- K(t_1,s))dB_s^H + \int_{t_1}^{t_2} K(t_2,s) dB_s^H.$$ Therefore, for any $s_1,t_1,s_2,t_2\in[0,1]$, $ s_1\leq t_1\leq s_2\leq t_2$ we have that $$\begin{gathered}
{\mathbf{E}}(X_{t_1}-X_{s_1})(X_{t_2}-X_{s_2})\\
= {\mathbf{E}}\left( \int_0^{s_1} \left(K(t_1,v) -K(s_1,v) \right)dB_v^H + \int_{s_1}^{t_1} K(t_1,v) dB_v^H\right)
\\ \times\left( \int_0^{s_2} \left(K(t_2,v) -K(s_2,v) \right)dB_v^H + \int_{s_2}^{t_2} K(t_2,v) dB_v^H\right)\\\nonumber
= H (2H-1) \Big( \int_0^{s_1} \int_{0}^{s_2} \left( K(t_1,v) -K(s_1,v)\right)\left( K(t_2,z) -K(s_2,z)\right)|v-z|^{2H-2}dvdz \label{l1}\\ +\nonumber
\int_0^{s_1} \int_{s_2}^{t_2} K(t_2,v)\left(K(t_1,z) -K(s_1,z)\right)|v-z|^{2H-2}dvdz
\\\nonumber
+ \int_{s_1}^{t_1} \int_0^{s_2} \left( K(t_2,v) -K(s_2,v)\right)K(t_1,z)|v-z|^{2H-2}dvdz\\\nonumber
+ \int_{s_1}^{t_2} \int_{s_2}^{t_2} K(t_1,v) K(t_2,z) |v-z|^{2H-2}dvdz\Big)\geq 0,
\end{gathered}$$ therefore, property $(B^+)$ holds.
Furthermore, $$\begin{gathered}
{\mathbf{E}}|X_{t}- X_s|^2
={\mathbf{E}}\Big|\int_0^{s } ( K(t,v)- K(s ,v) ) dB_v^H + \int_{s}^{t} K(t,v) dB_v^H \Big|^2\\
\leq 2 {\mathbf{E}}\left|\int_0^{s } ( K( t ,v)- K(s ,v) ) dB_v^H\right|^2 + 2{\mathbf{E}}\Big|\int_{s }^{t} K(t,v) dB_v^H \Big|^2.
\end{gathered}$$
According to [@MMV], there exists a constant $C_H$ depending only on $H$ such that $$\begin{gathered}
{\mathbf{E}}\left|\int_0^{s } ( K(t,v)- K(s ,v) ) dB_v^H \right|^2 \\ \leq C_H \Big(\int_0^{s } \left( K(t,v)- K(s,v)\right)^{\frac{1}{H}} dv\Big)^{2H} \leq C_H \, K^2 \,|t-s|^{2H_3}
\end{gathered}$$
and $$\begin{aligned}
{\mathbf{E}}\left|\int_{s}^{t} K(t,v) dB_v^H \right|^2 \leq C_H \left( \int_{s }^{t} \left(K(t,v)\right)^{\frac{1}{H}} dv \right)^{2H} \leq C_H K^2 |t-s|^{2H},\end{aligned}$$ and condition $(A1) $ holds with $H_2=H_3\wedge H$. Finally, $$\begin{gathered}
{\mathbf{E}}|X_{t}- X_s|^2
\geq C_H\int_s^{t} \int_s^{t} K(t,v)K(t,z)|v-z|^{2H}dvdz\geq k^2|t-s|^{2H},
\end{gathered}$$ whence condition $(A2)$ follows.
Evidently, a fractional Ornstein-Uhlenbeck process with positive drift $a$ and zero initial value satisfies the assumptions of the above theorem.
But the representation result of Theorem \[cor:main\] is also valid for a fractional Ornstein–Uhlenbeck process with a negative drift. Indeed, by the fractional Girsanov theorem (see e.g. [@Mish]), its law is equivalent to that of the fractional Brownian motion driving it. Moreover, the fractional Brownian motion generates the same filtration as the fractional Brownian motion. Therefore, the representation theorem for fractional Brownian motion can be transfered to the fractional Ornstein–Uhlenbeck process via the Girsanov transform. Naturally, the same may be done for a wide class of processes, e.g. solutions of stochastic differential equations with fractional Brownian motion. (See also the discussion in Remark \[rem:equivalence\] above.)
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We study deformations of the discrete Heisenberg group acting properly discontinuously on the Heisenberg group from the left and right and obtain a complete description of the deformation space.'
address: |
Instituto de Matemática, Estatística e Computação Científica, Universidade Estadual de Campinas\
Rua Sérgio Buarque de Holanda, 651, Cidade Universitária “Zeferino Vaz”, Campinas, SP, Brazil
author:
- 'Severin [Barmeier]{}'
title: Deformations of the discrete Heisenberg group
---
Introduction and statement of main result.
==========================================
We will be interested in deformations of the discrete Heisenberg group as a group acting properly discontinuously and cocompactly on a space $X$. The following defines our notion of [*deformation*]{}.
Let $G$ be a Lie group acting continuously on a locally compact space $X$ and let $\Gamma\subset G$ be a discrete subgroup. Define the [*parameter space*]{} of deformations of $\Gamma$ within $G$, acting properly discontinuously on the space $X$ as $$R(\Gamma,G;X)=\left\{\phi\colon\Gamma\to G\middle|
\begin{array}{l}
\phi\text{ is injective},\\
\phi(\Gamma)\text{ acts properly}\\\text{discontinuously}\\
\text{and freely on $X$}
\end{array}\right\}$$ and the [*deformation space*]{} as $$\mathcal T(\Gamma,G;X)=R(\Gamma,G;X)/G,$$ where $G$ acts on $R(\Gamma,G;X)$ by conjugation, so that $\mathcal T(\Gamma,G;X)$ is the space of non-trivial deformations.
There is a natural topology on the parameter space $R(\Gamma,G;X)$ as a subset of ${\mathop{\mathrm{Hom}}\nolimits}(\Gamma,G)$ endowed with the compact open topology. Then we consider the quotient topology on the deformation space $\mathcal T(\Gamma,G;X)$ ([@kobayashi93; @kobayashi01]).
If $X$ is an irreducible Riemannian symmetric space $G/K$, Selberg–Weil rigidity ([@weil]) states that $\mathcal T=\mathcal T(\Gamma,G;G/K)$ is discrete if and only if $G$ is not locally isomorphic to $\operatorname{SL}_2{\mathbb R}$. An example of the failure of rigidity is when $G=\operatorname{PSL}_2{\mathbb R}$, $\Gamma$ is the fundamental group of a Riemann surface of genus $g\geqslant 2$ and $X=\operatorname{SL}_2{\mathbb R}/\operatorname{SO}_2$ is the Poincaré disk. Then $\mathcal T$ is the Teichmüller space, which has dimension $6g-6$.
The study of deformations of discontinuous groups for non-Riemannian homogeneous spaces and the failure of rigidity was initiated by Kobayashi [@kobayashi93]; Kobayashi [@kobayashi98] treats the case when $G$ is semi-simple. A complete description of the parameter and deformation spaces was first given for $\Gamma={\mathbb Z}^{k}$ acting on $X={\mathbb R}^{k+1}$ via some nilpotent group of transformations $G$ in [@kobayashinasrin06] and these results were extended to the case where $G$ is the Heisenberg group, $H$ is any connected Lie subgroup and $\Gamma$ is a subgroup acting properly discontinuously and freely on $X=G/H$, in [@bakloutikedimyoshino08].
In this paper, we give a concrete description of the space $R(\Gamma,G\times G;G)$, where $G$ is the Heisenberg group, $\Gamma=G\cap\operatorname{GL}_3{\mathbb Z}$ is the discrete Heisenberg group and the direct product group $G\times G$ acts on the group manifold $G$ from the left and right. Our main result is the following.
For the deformation space $\mathcal T(\Gamma,G\times G;G)$ of the discrete Heisenberg group acting properly discontinuously on the group manifold $G$ from the left and right, we have the homeomorphism $$\mathcal T(\Gamma,G\times G;G)\cong\operatorname{GL}_2{\mathbb R}\times{\mathbb R}^\times\times{\mathbb R}^3.$$
Notation. {#notation}
=========
Let $G$ denote the Heisenberg group and $\Gamma=G\cap\operatorname{GL}_3{\mathbb Z}$ denote the discrete Heisenberg group. We will replace the matrix notation by defining $$\begin{bmatrix}
a\\b\\c
\end{bmatrix}:=
\begin{pmatrix}
1&a&c\\0&1&b\\0&0&1
\end{pmatrix}.$$
We will fix a presentation $\Gamma=\langle\gamma_1,\gamma_2\rangle$, where $$\begin{aligned}
\label{generators}
\gamma_1=
\begin{bmatrix}
1\\0\\0
\end{bmatrix}
\quad\text{and}\quad
\gamma_2=
\begin{bmatrix}
0\\1\\0
\end{bmatrix}.\end{aligned}$$
As a subgroup $\Gamma$ always acts properly discontinuously and freely on $G$ from the left and the quotient space $\Gamma\backslash G$ is a manifold. Similarly $\Gamma$ always acts properly discontinuously from the right with compact quotient $G/\Gamma$.
To let $\Gamma$ act both from the left and from the right, we rewrite $G$ as the homogeneous space $G\times G/\Delta G$, where $\Delta\colon G\to G\times G$ is the diagonal embedding. Then $\Gamma$ acts on $G\times G/\Delta G$ via homomorphisms $\Gamma\to G\times G$. We note here that ${\mathop{\mathrm{Hom}}\nolimits}(\Gamma,G\times G)\cong(G\times G)\times(G\times G)$ as sets, because each generator $\gamma_1,\gamma_2$ can be assigned any element in $G\times G$, as any relations $\gamma_1$ and $\gamma_2$ satisfy as elements of $G$ are also satisfied by any two arbitrary elements in $G\times G$. Via the topology on $G$, then, ${\mathop{\mathrm{Hom}}\nolimits}(\Gamma,G\times G)$ can be regarded a topological space. In particular, for $G$ being the Heisenberg group we have that $G\cong{\mathbb R}^3$, whence ${\mathop{\mathrm{Hom}}\nolimits}(\Gamma,G\times G)\cong{\mathbb R}^{12}$.
Any homomorphism $\Gamma\to G\times G$ can be written as a pair of homomorphisms $\rho,\rho'\colon\Gamma\to G$. Now write $\Gamma_{\rho,\rho'}=\{(\rho(\gamma),\rho'(\gamma))\mid \gamma\in\Gamma\}$ for the image of the pair $(\rho,\rho')\colon\Gamma\to G\times G$. Then $\Gamma$ acts on $G\times G/\Delta G$ via $\Gamma_{\rho,\rho'}$ and the action of $\Gamma$ on $G$ as subgroup (on the left) is recovered as the action of $\Gamma_{\mathrm{id},\mathbf 1}$ on $G\times G/\Delta G$, where $\mathrm{id}$ is the inclusion and $\mathbf 1$ is the trivial homomorphism. However, for general $\rho,\rho'$ this action is not necessarily properly discontinuous.
Rewriting $G$ as $G\times G/\Delta G$ for $G=\widetilde{\operatorname{SL}_2{\mathbb R}}$ allowed Goldman [@goldman] to construct non-standard Lorentz space forms. Goldman’s conjecture concerning the existence of an open neighbourhood of the embedding ${\rm id}\times{\mathbf 1}$, throughout which the group action remains properly discontinuous was resolved affirmatively for reductive Lie groups by Kobayashi [@kobayashi98]. An analogous result holds if $G$ is a simply connected Lie group and $\Gamma$ is a cocompact discrete group by an unpublished result of T. Yoshino. Our results below show this feature explicitly for $G$ being the Heisenberg group.
Property (CI) and proper actions.
=================================
To check for proper discontinuity of the action of $\Gamma_{\rho,\rho'}$, we will use a criterion by Nasrin [@nasrin01] for 2-step nilpotent groups, which relates properness to the [*property*]{} (CI).
We say the triplet $(L,H,G)$ has the property [(CI)]{} if $L\cap gHg^{-1}$ is compact for any $g\in G$.
(See [@lipsman] for the relationship between the property (CI) and proper actions in the more general context of locally compact topological groups acting on locally compact topological spaces.)
\[thmnasrin\] Let $G$ be a simply connected 2-step nilpotent Lie group, and let $H$ and $L$ be connected subgroups. Then the following conditions are equivalent.
1. $L$ acts properly on $G/H$,
2. the triplet $(L,H,G)$ has the property [(CI)]{},
3. $L\cap gHg^{-1}=\{e\}$ for any $g\in G$.
We will apply this theorem to the triple $(L_{\rho,\rho'},\Delta G,G\times G)$, where $G$ is again the Heisenberg group and $L_{\rho,\rho'}$ is the [*extension*]{} of $\Gamma_{\rho,\rho'}$ defined as follows.
\[extension\] Let $\Gamma$ be a discrete subgroup in a Lie group $G$. A connected subgroup $L\subset G$ is said to be the [*extension*]{} of $\Gamma$ if $L$ contains $\Gamma$ cocompactly.
The following lemma will allow us to use Thm. \[thmnasrin\] to determine the conditions under which $\Gamma_{\rho,\rho'}$ acts properly discontinuously.
\[properproperlydiscontinuous\] Let $L$ be a Lie group acting continuously on a locally compact space $X$. Let $\Gamma\subset L$ be a discrete subgroup such that $\Gamma\backslash L$ is compact. Then the following conditions are equivalent.
1. $\Gamma$ acts properly discontinuously on $X$,
2. $L$ acts properly on $X$.
Main results.
=============
To find the extension of $\Gamma_{\rho,\rho'}$, we use the (global) diffeomorphism $\exp\colon\mathfrak g\to G$, whose inverse we denote by $\log$. Let $\rho,\rho'\colon\Gamma\to G$ be any two homomorphisms. Then $\rho$ and $\rho'$ are determined by their values on the generators, which (in the notation of §\[notation\]) we will set to be $$\begin{aligned}
\label{rho}
\rho(\gamma_i)=
\begin{bmatrix}
a_i\\b_i\\c_i
\end{bmatrix}
\quad\text{and}\quad
\rho'(\gamma_i)=
\begin{bmatrix}
a'_i\\b'_i\\c'_i
\end{bmatrix},\end{aligned}$$ for $i=1,2$. Now, let $\rho_0\colon\mathfrak g\to\mathfrak g$ be a Lie algebra homomorphism defined on the generators by $\rho_0(\log\gamma_i)=\log\rho(\gamma_i)$, for $i=1,2$, and $\rho_0([\log\gamma_1,\log\gamma_2])=\log\rho([\gamma_1,\gamma_2])$, and extended linearly; let $\overline\rho\colon G\to G$ be defined by $\overline\rho=\exp\circ\rho_0\circ\log$. Then $\overline\rho\vert_\Gamma=\rho$, so that $\overline\rho$ extends $\rho$ in the sense that $\overline\rho$ is defined on all of $G$. If we write $\overline\rho'$ for the extension of $\rho'$ to all of $G$, then $L_{\rho,\rho'}=\{(\overline\rho(g),\overline\rho'(g))\mid g\in G\}$ is the extension of $\Gamma_{\rho,\rho'}$ in the sense of Def. \[extension\].
Next, we will check condition (c) of Thm. \[thmnasrin\] for $(L_{\rho,\rho'},\Delta G,G\times G)$. We have that $$\begin{aligned}
&L_{\rho,\rho'}\cap(g_1,g_2)\Delta G(g_1,g_2)^{-1}=\{e\}\notag\\
\Leftrightarrow\;&\overline\rho(g)=g_1^{-1}g_2\overline\rho'(g)(g_1^{-1}g_2)^{-1}\text{ only if $g=e$}\notag\\
\Leftrightarrow\;&\rho_0(\log g)=\mathrm{Ad}_{g_1^{-1}g_2}\rho'_0(\log g)\text{ only if $\log g=0$}\label{condition}\end{aligned}$$ for all $(g_1,g_2)\in G\times G$. Now write $$g=\begin{pmatrix}1&a&c\\0&1&b\\0&0&1\end{pmatrix}
\quad\text{and}\quad\log g=\begin{pmatrix}0&a&c-\frac12ab\\0&0&b\\0&0&0\end{pmatrix}.$$
Calculating the LHS and RHS of (\[condition\]) explicitly, it follows that (\[condition\]) is equivalent to $$\begin{aligned}
&\begin{pmatrix}
a_1 & a_2 & 0\\
b_1 & b_2 & 0\\
\ast & \ast & a_1b_2-a_2b_1
\end{pmatrix}
\begin{pmatrix}
a\\b\\c
\end{pmatrix}\\=
&\begin{pmatrix}
a'_1 & a'_2 & 0\\
b'_1 & b'_2 & 0\\
\ast & \ast & a'_1b'_2-a'_2b'_1
\end{pmatrix}
\begin{pmatrix}
a\\b\\c
\end{pmatrix}\Rightarrow a=b=c=0.\end{aligned}$$ Writing $$\begin{aligned}
\label{a}
A=\begin{pmatrix}
a_1 & a_2\\b_1 & b_2
\end{pmatrix}
\quad\text{and}\quad
A'=\begin{pmatrix}
a'_1 & a'_2\\b'_1 & b'_2
\end{pmatrix}\end{aligned}$$ we can rewrite condition (\[condition\]) as $$\det\begin{pmatrix}
A-A' & 0\\ \ast & \det A-\det A'
\end{pmatrix}\neq 0,$$ and we obtain the following proposition.
\[mainresult\] The group $\Gamma_{\rho,\rho'}$ acts properly discontinuously and cocompactly on $G\times G/\Delta G$ if and only if the following two conditions hold.
1. $\det(A-A')\neq0$, and
2. $\det A-\det A' \neq0$,
where $A,A'$ are determined by $\rho,\rho'$ via (\[rho\]) and (\[a\]).
Proper discontinuity is contained in the above argument. For cocompactness we make use of the following lemma.
\[rhoinjective\] Let $\rho$ be as in (\[rho\]) and $A$ be defined by (\[a\]). Then $\det A\neq0\Leftrightarrow\rho$ is injective.
$\det A$ is precisely the (1,3) entry of the commutator $[\rho(\gamma_1),\rho(\gamma_2)]$ and $\det A\neq0$ if and only if the image $\rho(\Gamma)$ is non-commutative. We show that $\rho(\Gamma)$ being non-commutative is equivalent to $\rho$ being injective.
If $\rho$ is injective, $\rho(\Gamma)\cong\Gamma$ is non-commutative. Conversely, write $N=\ker\rho$ and assume that $\rho(\Gamma)$ is non-commutative. We have the commutative diagram $$\begin{aligned}
\begin{tikzpicture}[baseline=-2.3pt,description/.style={fill=white,inner sep=2pt}]
\matrix (m) [matrix of math nodes, row sep=2.85em,
column sep=1em, inner sep=4pt, text height=1.5ex, text depth=0.25ex, ampersand replacement=\&]
{
\& 0 \& 0 \& 0 \& \\
0 \& N\cap{\mathbb Z}\& N \& N/N\cap{\mathbb Z}\& 0 \\
0 \& {\mathbb Z}\& \Gamma \& {\mathbb Z}^2 \& 0 \\
0 \& {\mathbb Z}/N\cap{\mathbb Z}\& \Gamma/N \& \Gamma/{\mathbb Z}N \& 0 \\
\& 0 \& 0 \& 0 \& \\
};
\draw[-stealth,font=\scriptsize]
(m-1-2) edge node[auto] {} (m-2-2)
(m-1-3) edge node[auto] {} (m-2-3)
(m-1-4) edge node[auto] {} (m-2-4)
(m-2-1) edge node[auto] {} (m-2-2)
(m-2-2) edge node[auto] {} (m-2-3)
(m-2-3) edge node[auto] {} (m-2-4)
(m-2-4) edge node[auto] {} (m-2-5)
(m-2-2) edge node[auto] {} (m-3-2)
(m-2-3) edge node[auto] {} (m-3-3)
(m-2-4) edge node[auto] {} (m-3-4)
(m-3-1) edge node[auto] {} (m-3-2)
(m-3-2) edge node[auto] {} (m-3-3)
(m-3-3) edge node[auto] {} (m-3-4)
(m-3-4) edge node[auto] {} (m-3-5)
(m-3-2) edge node[auto] {} (m-4-2)
(m-3-3) edge node[auto] {} (m-4-3)
(m-3-4) edge node[auto] {} (m-4-4)
(m-4-1) edge node[auto] {} (m-4-2)
(m-4-2) edge node[auto] {} (m-4-3)
(m-4-3) edge node[auto] {} (m-4-4)
(m-4-4) edge node[auto] {} (m-4-5)
(m-4-2) edge node[auto] {} (m-5-2)
(m-4-3) edge node[auto] {} (m-5-3)
(m-4-4) edge node[auto] {} (m-5-4);
\end{tikzpicture}\end{aligned}$$ whose rows and columns are exact by the nine lemma. Turning our attention to the first column, the top left entry $N\cap{\mathbb Z}$ can be considered as a subgroup of ${\mathbb Z}$, and is thus equal to (*i*) $0$, (*ii*) ${\mathbb Z}$, or (*iii*) $m{\mathbb Z}$, for some $m\geq2$.
[**Case [(*ii*)]{}.**]{} If $N\cap{\mathbb Z}={\mathbb Z}$, $N$ contains the commutator ${\mathbb Z}=[\Gamma,\Gamma]$, contradicting the fact that $\rho(\Gamma)\cong\Gamma/N$ was assumed non-commutative.
[**Case [(*iii*)]{}.**]{} If $N\cap{\mathbb Z}=m{\mathbb Z}$, for $m\geq2$, then ${\mathbb Z}/N\cap{\mathbb Z}={\mathbb Z}_m$ in the bottom left entry. However, ${\mathbb Z}_m$ is finite and contains torsion elements and injects into $\Gamma/N$. By the first isomorphism theorem for groups, the induced map $\rho_\ast\colon\Gamma/N\to G$ is injective. But $G$ is torsion-free, whence $\Gamma/N$ is torsion-free also and we obtain a contradiction.
We conclude that $N\cap{\mathbb Z}=0$ (case [(*i*)]{}).
Now, write $\pi\colon\Gamma\to{\mathbb Z}^2$ for the projection and $\pi^\ast\colon N\to N/N\cap{\mathbb Z}$ for the restriction of $\pi$ to $N$. Let $\gamma$ be any element in $\Gamma$ and $n\in N$. Since $N$ is normal, $\gamma n\gamma^{-1}\in N$. Then $$\pi^\ast(\gamma n\gamma^{-1})=\pi^\ast(\gamma)\pi^\ast(n)\pi^\ast(\gamma^{-1})=\pi^\ast(n),$$ where the last equality follows from the fact that $\mathrm{im}\,\pi^\ast$ injects into ${\mathbb Z}^2$ and is therefore commutative. Since $N\cap{\mathbb Z}=0$, $\pi^\ast$ is an isomorphism and we conclude that $\gamma n=n\gamma$, i.e. $N$ is contained in the centraliser ${\mathbb Z}$. Then $N\cap{\mathbb Z}=0$ shows that $N$ is trivial, whence $\rho$ is injective.
Using Thm. \[thmnasrin\], we have shown that $L_{\rho,\rho'}$ acts properly on $G\times G/\Delta G$ if and only if conditions (a) and (b) hold. Applying Lem. \[properproperlydiscontinuous\], $L_{\rho,\rho'}$ acts properly on $G\times G/\Delta G$ if and only if $\Gamma_{\rho,\rho'}$ acts properly discontinuously on $G\times G/\Delta G$.
By Lem. \[rhoinjective\], condition (b) shows that at least one of $\rho,\rho'$ must be injective, whence the cohomological dimension $\operatorname{cd}\Gamma_{\rho,\rho'}=3$. It is a fact, based on a standard argument invoking Poincaré duality, that if a group $\Gamma$ acts (faithfully) on a contractible manifold $X$ and $\operatorname{cd}\Gamma=\dim X$, then $\Gamma\backslash X$ is compact (cf. [@kobayashi89], Cor. 5.5). Since $G\times G/\Delta G\cong{\mathbb R}^3$ is indeed contractible and $\dim G\times G/\Delta G=\operatorname{cd}\Gamma_{\rho,\rho'}=3$, the double quotient $\Gamma_{\rho,\rho'}\backslash G\times G/\Delta G$ is compact.
Prop. \[mainresult\] can be turned into a method for determining pairs of homomorphisms for which $\Gamma_{\rho,\rho'}$ acts properly discontinuously and cocompactly on $G$ from the left and right as follows.
Let $$\begin{aligned}
S=\begin{pmatrix}s_0 & s_1\\s_2 & s_3\end{pmatrix}&\in\operatorname{GL}_2{\mathbb R}\\
\text{and }(S,t_0,t_1,t_2,t_3,c_1,c_2,c_1',c_2')&\in\operatorname{GL}_2{\mathbb R}\times{\mathbb R}^\times\times{\mathbb R}^7.\end{aligned}$$ Define a map $$\begin{aligned}
\label{alpha}
\alpha\colon\operatorname{GL}_2{\mathbb R}\times{\mathbb R}^\times\times{\mathbb R}^7&\to R(\Gamma,G\times G;G)\\
(S,t_0,t_1,t_2,t_3,c_1,c_2,c_1',c_2')&\mapsto\phi,\notag\end{aligned}$$ where $\phi=(\rho,\rho')$ is defined by $$\begin{aligned}
\rho(\gamma_1)&=
\begin{bmatrix}
\tfrac12(s_0(t_0+t_3)+s_0+s_1t_2)\\
\tfrac12(s_2(t_0+t_3)+s_2+s_3t_2)\\
c_1
\end{bmatrix}\\
\rho(\gamma_2)&=
\begin{bmatrix}
\tfrac12(s_1(t_0-t_3)+s_1+s_0t_1)\\
\tfrac12(s_3(t_0-t_3)+s_3+s_2t_1)\\
c_2
\end{bmatrix}\\
\rho'(\gamma_1)&=
\begin{bmatrix}
\tfrac12(s_0(t_0+t_3)-s_0+s_1t_2)\\
\tfrac12(s_2(t_0+t_3)-s_2+s_3t_2)\\
c'_1
\end{bmatrix}\\
\rho'(\gamma_2)&=
\begin{bmatrix}
\tfrac12(s_1(t_0-t_3)-s_1+s_0t_1)\\
\tfrac12(s_3(t_0-t_3)-s_3+s_2t_1)\\
c'_2
\end{bmatrix}.\end{aligned}$$ Determining $A,A'$ via (\[a\]), one checks that $A-A'=S$ and $\det A-\det A'=t_0\cdot\det S\neq0$, as $t_0\in{\mathbb R}^\times$. Thus, conditions (a) and (b) from Prop. \[mainresult\] are satisfied and $\Gamma_{\rho,\rho'}$ acts properly discontinuously and cocompactly on $G$ from the left and right. Moreover, we have the following
The map $\alpha$ (see (\[alpha\])) induces a homeomorphism from $\operatorname{GL}_2{\mathbb R}\times{\mathbb R}^\times\times{\mathbb R}^7$ onto the parameter space $R(\Gamma,G\times G;G)$ of deformations of $\Gamma$ acting properly discontinuously on the group manifold $G$ from the left and right. Furthermore, the deformation space $\mathcal T(\Gamma,G\times G;G)$ is homeomorphic to $\operatorname{GL}_2{\mathbb R}\times{\mathbb R}^\times\times {\mathbb R}^3$.
The idea of the proof and the origin of the map $\alpha$ is the following.
The space of pairs of matrices satisfying (a) and (b) of Prop. \[mainresult\] can be determined as follows. Suppose $A,A'$ satisfy (a) and (b). Consider the map $$\begin{aligned}
\omega\colon(A,A')\mapsto(U,V)=(A-A',(A-A')^{-1}(A+A')),\end{aligned}$$ which is well-defined, since $U=A-A'$ is invertible. We can find an inverse mapping $$\alpha_0\colon(U,V)\mapsto(\tfrac12(UV+U),\tfrac12(UV-U))$$ and one checks that $\alpha_0\circ\omega=\mathrm{id}$ and $\omega\circ\alpha_0=\mathrm{id}$.
For $U$ and $V$, condition (a) is equivalent to the condition that $U\in\operatorname{GL}_2{\mathbb R}$; condition (b) translates into the condition $$\begin{aligned}
\det\tfrac12(UV+U)&\neq\det\tfrac12(UV-U)\\
\Leftrightarrow\hspace{30pt}\det (V+I)&\neq\det(V-I),\label{determinanttrace}\\
\Leftrightarrow\hspace{23.7pt}\det V+\operatorname{tr}V&\neq\det V-\operatorname{tr}V\\
\Leftrightarrow\hspace{59.6pt}\operatorname{tr}V&\neq0,\end{aligned}$$ where $I$ denotes the $2\times2$ identity matrix. Then, writing $M=\{V\in M_2({\mathbb R})\mid\operatorname{tr}V\neq 0\}\cong{\mathbb R}^\times\times{\mathbb R}^3$, the map $$\alpha_0\colon\operatorname{GL}_2{\mathbb R}\times M\to\{(A,A')\mid A,A'\text{ satify (a) \& (b)}\}$$ is a homeomorphism. Writing $\mathrm{id}$ for the identity on ${\mathbb R}^4=\{(c_1,c_2,c_1',c_2')|c_1,c_2,c_1',c_2'\in{\mathbb R}\}$, $\alpha_0\times\mathrm{id}=\alpha$ is the homeomorphism $$\alpha\colon\operatorname{GL}_2{\mathbb R}\times{\mathbb R}^\times\times{\mathbb R}^7\to R(\Gamma,G\times G;G)$$ up to the identification $M\cong{\mathbb R}^\times\times{\mathbb R}^3$.
The conjugation action of $G\times G$ on $\Gamma_{\rho,\rho'}$ leaves the superdiagonal entries of each factor unchanged and is transitive on the $(1,3)$ entries, so that $\mathcal T(\Gamma,G\times G;G)=R(\Gamma,G\times G;G)/(G\times G)$ is homeomorphic to $$\begin{aligned}
\operatorname{GL}_2{\mathbb R}\times{\mathbb R}^\times\times{\mathbb R}^3.\end{aligned}$$
Geometric interpretation of main result.
========================================
Geometrically speaking, we have the central extensions $$\begin{aligned}
\begin{tikzpicture}[baseline=-2.3pt,description/.style={fill=white,inner sep=2pt}]
\matrix (m) [matrix of math nodes, row sep=0.2em,
column sep=3ex, inner sep=2pt, text height=1.5ex, text depth=0.25ex, ampersand replacement=\&]
{0 \& {\mathbb R}\& G \& {\mathbb R}^2 \& 0\\
0 \& {\mathbb Z}\& \Gamma \& {\mathbb Z}^2 \& 0\\};
\path[-stealth,font=\scriptsize]
(m-1-1) edge (m-1-2)
(m-1-2) edge (m-1-3)
(m-1-3) edge (m-1-4)
(m-1-4) edge (m-1-5)
(m-2-1) edge (m-2-2)
(m-2-2) edge (m-2-3)
(m-2-3) edge (m-2-4)
(m-2-4) edge (m-2-5);
\end{tikzpicture}\end{aligned}$$ and by quotienting $\Gamma\backslash G$ can be viewed as a circle bundle over the torus. The two conditions of Prop. \[mainresult\] can then be interpreted as follows. The matrix $A-A'$ determines a Riemannian structure on this torus and $\det A-\det A'$ determines the structure on (i.e. length of) the circle. In particular, the number of connected components (which equals four) of the deformation space $\mathcal T(\Gamma,G\times G;G)$ corresponds to the number of possible combinations of orientations on the torus and the circle.
\[notgraph\] Let $$\rho(\gamma_1)=
\begin{bmatrix}
2\\
c\\
0
\end{bmatrix},\qquad
\rho(\gamma_2)=
\begin{bmatrix}
1\\
2\\
0
\end{bmatrix}$$ and $$\rho'(\gamma_1)=
\begin{bmatrix}
1\\
c\\
0
\end{bmatrix},\qquad
\rho'(\gamma_2)=
\begin{bmatrix}
0\\
1\\
0
\end{bmatrix}.$$ Letting $c$ vary from $0$ to $1$, we obtain a family of groups $\Gamma_{\rho,\rho'}$ (which lies in the component of both base space and fibre orientations being positive), which by Prop. \[mainresult\] act cocompactly and properly discontinuously on $G\times G/\Delta G$, where the length of the fibre varies from $3$ to $2$ and the structure on the torus remains unchanged and is given by the matrix $\left(\begin{smallmatrix}
1 & 1\\0 & 1
\end{smallmatrix}\right)$.
Similarly, it is possible to find families of groups, which only change the structure on the base space, leaving the length of the fibre unchanged; or families, for which both the structure on the base space and the length of the fibre are fixed, but the connection form is deformed.
General examples, like the one above, stand in contrast to the case when $G$ is semisimple of real rank $1$—e.g. $G=\operatorname{SL}_2{\mathbb R}$, $\operatorname{SO}(n,1)$, $\mathrm{SU}(n,1)$, $\mathrm{Sp}(n,1)$—for which any properly discontinuous group for $G\times G/\Delta G$ is a graph up to a finite-index subgroup ([@kobayashi93], Thm. 2 and Rmk. 1).
Acknowledgements. {#acknowledgements. .unnumbered}
=================
The author would like to thank Prof. Taro Yoshino for detailed comments on an earlier version of this paper and Prof. Toshiyuki Kobayashi for his comments, guidance and invaluable advice.
[50]{} Baklouti, A., Kédim, I. & Yoshino, T. “On the deformation space of Clifford–Klein forms of Heisenberg groups”, [*Int. Math. Res. Notices*]{}, rnn066: 26 pp., 2008.
Goldman, W.M. “Nonstandard Lorentz space forms”, [*J. Diff. Geom.*]{}, [**21**]{}:301–308, 1985.
Kobayashi, T. “Proper action on a homogeneous space of reductive type”, [*Math. Ann.*]{}, [**285**]{}:249–263, 1989.
Kobayashi, T. “Discontinuous groups acting on homogeneous spaces of reductive type”, in T. Kawazoe, T. Oshima & S. Sano (eds.) [*Proc. of Fuji–Kawaguchiko Conf. on Representation Th. of Lie Groups and Lie Algebras*]{}, pp. 59–75, River Edge, NJ: World Scientific, 1992.
Kobayashi, T. “On discontinuous groups acting on homogeneous spaces with noncompact isotropy subgroups”, [*J. of Geom. and Phys.*]{}, [**12**]{}:133–144, 1993.
Kobayashi, T. “Deformation of compact Clifford–Klein forms of indefinite-Riemannian homogeneous manifolds”, [*Math. Ann.*]{}, [**310**]{}:395–409, 1998.
Kobayashi, T. “Discontinuous groups for non-Riemannian homogeneous spaces”, in B. Engquist & W. Schmid (eds.) [*Mathematics Unlimited—2001 and Beyond*]{}, pp. 723–747, New York, NY: Springer-Verlag, 2001.
Kobayashi, T. & Nasrin, S. “Deformation of properly discontinuous actions of ${\mathbb Z}^{k}$ on ${\mathbb R}^{k+1}$”, [*Int. J. of Math.*]{}, [**17**]{}(10):1175–1193, 2006.
Lipsman, R.L. “Proper action and a compactness condition”, [*J. of Lie Theory*]{}, [**5**]{}:25–39 1995.
Nasrin, S. “Criterion of proper actions for 2-step nilpotent Lie groups”, [*Tokyo J. Math.*]{}, [**24**]{}(2):535–543, 2001.
Weil, A. “Remarks on the cohomology of groups”, [*Ann. Math.*]{}, [**80**]{}:149–157, 1964.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
In this Paper we report on radio (VLA and ATCA) and X–ray ([*RXTE*]{}, and ) observations of the outburst decay of the transient black hole candidate in early 2008. We find that the X–ray light curve followed an exponential decay, levelling off towards its quiescent level. The exponential decay timescale is $\approx$4 days and the quiescent flux corresponds to a luminosity of 3$\times 10^{32}(\frac{{\rm d}}{7.5
{\rm kpc}})^2$. This together with the relation between quiescent X–ray luminosity and orbital period reported in the literature suggests that has an orbital period longer than $\approx$10 hours. Both the radio and X–ray light curve show evidence for flares. The radio – X–ray correlation can be well described by a power–law with index $\approx$0.18. This is much lower than the index of $\approx$0.6–0.7 found for the decay of several black hole transients before. The radio spectral index measured during one of the radio flares while the source is in the low–hard state, is -0.5$\pm$0.15, which indicates that the radio emission is optically thin. This is unlike what has been found before in black hole sources in the low–hard state. We attribute the radio flares and the low index for the radio – X–ray correlation to the presence of shocks downstream the jet flow, triggered by ejection events earlier in the outburst. We find no evidence for a change in X–ray power law spectral index during the decay, although the relatively high extinction of ${\rm N_H \approx
2.3 \times 10^{22} cm^{-2}}$ limits the detected number of soft photons and thus the accuracy of the spectral fits.
author:
- |
P.G. Jonker$^{1,2}$[^1], J. Miller–Jones$^3$, J. Homan$^{4}$, E. Gallo$^{4}$, M. Rupen$^{5}$, J. Tomsick$^{6}$,R.P. Fender$^{7}$, P. Kaaret$^{8}$, D.T.H. Steeghs$^{9,2}$, M.A.P. Torres$^{2}$, R. Wijnands$^{10}$, S. Markoff$^{10}$, W.H.G. Lewin$^{4}$\
$^1$SRON, Netherlands Institute for Space Research, Sorbonnelaan 2, 3584 CA, Utrecht, The Netherlands\
$^2$Harvard–Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, U.S.A.\
$^3$NRAO Headquarters, 520 Edgemont Road, Charlottesville, VA 22903, USA\
$^4$MIT, Kavli Institute for Astrophysics and Space Research, 70 Vassar Street, Cambridge, MA 02139, USA\
$^5$NRAO, Array Operations Center, 1003 Lopezville Road, Socorro, NM 87801, USA\
$^6$Space Sciences Laboratory, University of California, Berkeley, USA\
$^7$School of Physics and Astronomy, University of Southampton, Southampton SO17 1BJ\
$^8$Department of Physics and Astronomy, University of Iowa, Van Allen Hall, Iowa City, IA 52242, USA\
$^9$Department of Physics, University of Warwick, Coventry CV4 7AL\
$^{10}$Astronomical Institute ‘Anton Pannekoek’, University of Amsterdam, Postbus 94249, 1090 GE Amsterdam, the Netherlands\
title: 'Following the 2008 outburst decay of the black hole candidate in X–ray and radio'
---
stars: individual () — accretion: accretion discs — stars: binaries — X-rays: binaries
Introduction
============
A wide variety of astrophysical objects are thought to be powered by accretion. In the case of accretion onto a black hole, the properties of these flows mainly depend on the mass and spin of the black hole and on the mass accretion rate ([$\dot{m}$]{}). Variations in [$\dot{m}$]{} are probably responsible for the spectral and variability states of black hole X-ray binaries (BHXBs; @2006csxs.book..157M).
Transient BHXBs spend long periods at very low X-ray luminosities, referred to as ‘quiescence’, but during occasional outbursts the luminosity increases by as much as 7 to 8 orders of magnitude, typically reaching values of tens of per cent of the Eddington luminosity ([$L_{Edd}$]{}). has shown that below around a few percent of [$L_{Edd}$]{} BHXBs typically return to the so–called hard state, in which the spectrum is dominated by a power law component with an index $\sim$1.5 (e.g. @2005ApJ...622..508K). BHXBs are still in that state when they have decayed down to $10^{-4}$[$L_{Edd}$]{}.
With [*Chandra*]{} and [*XMM-Newton*]{} it is possible to explore the spectral properties of BHXBs all the way to quiescence, for which log([$L_X$]{}) is typically $\sim$30.5–33.0 ([email protected]; ). This range corresponds to $\sim$3$\times 10^{-9}-$9$\times 10^{-7}$[$L_{Edd}$]{}for a 8.5$M_\odot$ black hole (which is the average black-hole mass from @2006csxs.book..157M). About 15 BHXBs have been observed in quiescence with [*Chandra*]{} and [*XMM-Newton*]{}, but only in a handful of cases could the spectral properties be well constrained. Nevertheless, these few cases suggest an interesting property of quiescent spectra: when fitted with a power-law the indices tend to be considerably softer than the index of $\sim1.5$ found in the hard state (e.g. @2006ApJ...636..971C). Although the errors for individual sources are still large, the overall trend seems to indicate that, at least spectrally, the quiescent state is different from the hard state that is observed three to five decades higher in [$L_X$]{}.
Corbel et al. (2003) and @2003MNRAS.344...60G have demonstrated the existence of a relation between observed X-ray and radio emission from hard state black holes, of the form $L_{\rm radio} \propto L_{\rm
X}^{0.7}$; this relation is nearly the same for V404 Cyg and GX 339–4. However, , @2007ApJ...659..549C, @2007ApJ...655L..97R, @2007ApJ...655..434S and @2008MNRAS.389.1697C found that the relation might not be as universal as previously thought (see the compilation in @2007AIPC..924..715G). On the other hand @2005ApJ...624..295H and @2006MNRAS.371.1334R found further relations between the X–ray and near–infrared fluxes in black hole transients: $L_{IR} \propto L_X^{\,\sim0.6}$. Both relations extend over more than three decades in luminosity down to $\sim 10^{-4}$[$L_{Edd}$]{} and suggest that the X–ray, near–infrared and radio emission in the hard state are intimately connected. Correlations between optical and X–ray light in the black hole source GX 339–4 point in the same direction (@2008MNRAS.390L..29G). [*Chandra*]{} and VLA observations of A 0620–00 in quiescence by @2006MNRAS.370.1351G showed that the radio and X–ray flux lie on the extension of the $L_{\rm radio} \propto L_{\rm X}^{0.7}$ correlation, suggesting that it holds all the way down to quiescence. Some models predict that the relation should break down (i.e. significantly steepen) around $10^{-5}$[$L_{Edd}$]{}(@2005ApJ...629..408Y). The result of Gallo et al. (2006) seems to rule this out, but given the contradicting results on the decay rate and normalisation of the radio – X–ray correlations, this needs to be investigated further.
ADAF and jet–models both predict that the X–ray luminosity scales as $\dot{m}^2$ below $10^{-4}$[$L_{Edd}$]{} (Narayan et al. 1997; @2003MNRAS.343L..99F). Observational evidence for this scaling was deduced from X-ray/radio correlations (; @2006MNRAS.369.1451K). In ADAF models a gradual softening of the power–law photon index is expected until an index of $\sim$2.2 in quiescence (@esmcna1997; @1997ApJ...482..448N). However, in ADAF models the disk should recede in the hard state, something that is recently debated by @2006ApJ...652L.113M [@2006ApJ...653..525M] (but see @2008MNRAS.388..753G; @2009MNRAS.394.2080H).
was discovered in August, 1977 by the Ariel–5 all–sky monitor. The position was more accurately determined by HEAO–I ruling out an association with 4U 1755–388 (@1977IAUC.3099S...1K; @1977IAUC.3106....4K). The source has shown repeated outbursts since 2003 after being rediscovered by the International Gamma–ray Astrophysics Laboratory (INTEGRAL; IGR J17464-3213 @2003ATel..132....1R). @2003ATel..133....1M suggested that the INTEGRAL–found source is the same as . Recently, the source showed several outbursts; one in 2007/2008 (the data presented in this paper is from this outburst; @2008ATel.1348....1K), one in late 2008 () and one in 2009 (@2009ATel.2058....1K).
@2006csxs.book..157M list as a strong black hole candidate. Phase resolved optical or near–infrared observations to determine the mass function of the black hole have not yet been reported, even though the source was found in outburst in near–infrared and optical bands (@2003IAUC.8112....2B; @2003ATel..146....1S). Near–infrared observations of the source in quiescence show an unrelated star within 1 and found the counterpart $K$–band magnitude to be 17.1 (@2009ApJ...698.1398M). is one of the few sources where X–ray jets have been imaged (@2005ApJ...632..504C). The other sources where an X–ray jet has been found are the black hole sources 4U 1755–33 and XTE J1550–564 (@2003ApJ...586L..71A; @2002Sci...298..196C), the neutron star source Cir X–1 (@2007ApJ...663L..93H; @2009MNRAS.397L...1S) and the peculiar source SS 433 (@2002Sci...297.1673M). In addition to X–ray jet outflows @2006ApJ...646..394M found a strongly variable disc wind from . The distance to this source is so far not well constrained, however, @2005ApJ...632..504C find from the fact that the jet speed is limited to maximally the speed of light and the observed X–ray jet proper motion that the upper limit to the distance is 10.4$\pm$2.9 kpc, consistent with the often assumed Galactic Center distance of 7.5 or 8 kpc. This assumption is based on the Galactic coordinates of the source ($l=$357.255 and $b=$-1.83); it lies towards the Galactic bulge. In this paper we assume a distance of 7.5 kpc.
Here, we report on contemporary [*RXTE*]{}, and X–ray and Very Large Array (VLA) radio observations of aimed at following the X–ray and radio light curves and establishing the X–ray – radio correlation while the source decays to quiescence. The observations have been obtained in the last stages of the outburst ending early 2008. For this outburst @2008ATel.1378....1K reported that the source returned to the low–hard state between Jan, 24, 2008 and Feb., 1, 2008. To determine the quiescent spectral properties we further analysed three observations of obtained when it was in quiescence (@2006ApJ...636..971C).
Observations, analysis and results
==================================
[*Chandra*]{} X–ray observations
--------------------------------
We observed with the satellite using the back–illuminated S3 CCD–chip of the Advanced CCD Imaging Spectrometer (ACIS) detector (@1997AAS...190.3404G) on several occasions during the final parts of the decay to quiescence. During the first two observations in 2008 (Obs IDs 8985 and 8986) we employed the High Energy Transmission Grating (HETG) to mitigate effects of photon pile–up. In the subsequent observations we windowed the ACIS–S CCD, providing a frame time of 0.4104 s. We have reprocessed and analysed the data using the [*CIAO 4.0.2*]{} software developed by the Chandra X–ray Center. In our analysis we have selected events only if their energy falls in the 0.3–7 keV range. All data has been used, as background flaring is very weak or absent in all data.
-------------- -------------- ------------- ------------- --------------------------------
Obs ID Observing MJD Time on Count rate
date (UTC) source (ks) 0.3-7 keV (cnt s$^{-1}$)
4565$^a$ 2004 Feb. 13 53047.84959 23.0 $(1.5\pm0.3)\times10^{-3}$
4566$^a$ 2004 Mar. 25 53088.74413 28.4 $(0.7\pm0.2)\times10^{-3}$
4567$^a$ 2004 Mar. 27 53091.25914 40.0 $(1.8\pm0.2)\times10^{-3}$
8985$^{b,d}$ 2008 Feb. 19 54515.14003 6.4 $0.167\pm0.005$$^c$
8986$^{b,d}$ 2008 Feb. 25 54521.04074 7.6 $(8.4\pm0.3)\times10^{-2}$$^c$
8987$^d$ 2008 Mar. 02 54527.13111 6.5 0.101$\pm$0.004
8988$^d$ 2008 Mar. 08 54533.69200 13.7 $(1.9\pm0.1)\times10^{-2}$
8989$^d$ 2008 Mar. 16 54541.23027 20.5 $(4.0\pm0.4)\times10^{-3}$
9833 2008 Mar. 17 54542.08033 11.0 $(2.5\pm0.5)\times10^{-3}$
9838$^d$ 2008 Mar. 21 54546.42945 23.8 $(1.8\pm0.3)\times10^{-3}$
8990 2008 Mar. 22 54547.32918 21.2 $(1.6\pm0.3)\times10^{-3}$
9839 2008 Mar. 23 54548.30002 28.7 $(1.4\pm0.2)\times10^{-3}$
9837 2008 Mar. 24 54549.22579 20.6 $(1.7\pm0.3)\times10^{-3}$
-------------- -------------- ------------- ------------- --------------------------------
[$^a$ Observations reanalysed from @2006ApJ...636..971C.\
$^b$ High energy transmision grating inserted\
$^c$ Zeroth order count rate.\
]{} [$^d$ Data point used in Figure \[rxcorr\].]{}\
Using [*wavdetect*]{} we detect in each of the observations. We have selected a circular region of 10 pixel ($\approx$5) radius centered on the accurately known source position (Steeghs et al. 2003) to extract the source counts. Similarly, we have used a circular annulus with inner and outer radius of 20 and 40 pixels centered on the source position to extract background counts. The source and background region for observation 4566 and 4567 excludes the X–ray jets found by @2005ApJ...632..504C. We have made redistribution and auxilliary response matrices for the source region of each of the observations separately.
The net, background subtracted, source count rate for each observation is given in Table \[log\]. Using [*xspec*]{} version 11.3.2p (@ar1996) we have fitted the spectra of using $\chi^2$ statistics requiring at least 10 counts per spectral bin for the observations with IDs 8985, 8986, 8987, and 8988 and Cash statistics (@1979ApJ...228..939C) modified to account for the subtraction of background counts, the so called W–statistics[^2] for the other observations. We have used an absorbed power–law model ([ *pegpwrlw*]{} in [*xspec*]{}) to describe the data. For the observations with IDs 8985, 8986, 8987, and 8988 we allow all three parameters (neutral hydrogen column density ${\rm N_H}$, the power–law index and normalisation) to float freely. Due to the relatively low number of counts we fix the interstellar extinction during the fits to the rest of the observations to 2.3$\times 10^{22}$ cm$^{-2}$, consistent with the ${\rm N_H}$ found by @2006ApJ...646..394M and for the observations with IDs 8985, 8986, 8987, and 8988. The power law index and normalisation were allowed to float.
We list the results of our spectral analysis in Table \[spec\]. Given the similar fluxes and contemporaneous observations we have also analysed the spectral behaviour of for the observations 9838, 8990, 9839, 9837 together. Furthermore, in order to determine the best–fitting power law index for the source in quiescence we have fitted the spectral model to observations 4565, 4566, 4567, 9838, 8990, 9839, 9837 together (see Table \[spec\]).
------------- --------------------- -------------------------------------------------------------- --------------- --------------------------------------- ---------------------------------
Obs ID N$_H~\times10^{22}$ PL flux (0.3-7 keV ) PL index Unabs. 0.5–10 keV flux Goodness / ${\rm \chi^2_{red}}$
cm$^{-2}$ (0.5-10 keV ) $\times10^{-12}$ erg$^{-1}$ cm$^{-2}$ s$^{-1}$ (erg cm$^{-2}$ s$^{-1}$) per cent / d.o.f.
8985$^d$ 1.8$\pm$0.4 11.8$\pm$1.0 0.87$\pm$0.18 $(1.8 \pm 0.2)\times 10^{-11}$ 1.12/89
8986$^d$ 1.7$\pm$0.6 6.3$\pm$1.0 1.03$\pm$0.27 $(9^{+2}_{-0.3})\times 10^{-12}$ 0.81/55
8987 2.2$\pm$0.3 2.8$\pm$0.5 1.68$\pm$0.21 $(3^{+0.6}_{-0.1})\times 10^{-12}$ 0.86/56
8988 2.2$\pm$0.7 0.55$^{+0.34}_{-0.15}$ 1.75$\pm$0.47 $(5^{+3}_{-0.5})\times 10^{-13}$ 0.7/20
8989 2.3$^a$ $(1.6\pm0.3)\times 10^{-1}$ 2.00$\pm$0.25 $(1.5\pm0.2)\times 10^{-13}$ 52%
9833 2.3$^a$ $(7.6^{+2.7}_{-1.7})\times 10^{-2}$ 1.7$\pm$0.5 $(8.4\pm1.7)\times 10^{-14}$ 54%
9838 2.3$^a$ $(6.1\pm1.5)\times 10^{-2}$ 1.8$\pm$0.4 $(6.2\pm1.0)\times 10^{-14}$ 58%
8990 2.3$^a$ $(6.2^{+2.9}_{-1.6})\times 10^{-2}$ 2.1$\pm$0.4 $(5.6\pm1.0)\times 10^{-14}$ 76%
9839 2.3$^a$ $(3.9\pm0.7)\times 10^{-2}$ 1.3$\pm$0.4 $(4.9\pm0.7)\times 10^{-14}$ 42%
9837 2.3$^a$ $(9\pm4)\times 10^{-2}$ 2.5$\pm$0.4 $(7\pm2)\times 10^{-14}$ 30%
4565 2.3$^a$ $(4.9\pm1.3)\times 10^{-2}$ 1.8$\pm$0.4 $(5\pm1)\times 10^{-14}$ 65%
4566 2.3$^a$ $(2.5^{+1.3}_{-0.7})\times 10^{-2}$ 1.8$\pm$0.6 $(2.6\pm0.6)\times 10^{-14}$ 71%
4567 2.3$^a$ $(5.9\pm1.1)\times 10^{-2}$ 1.8$\pm$0.3 $(6.2\pm0.7)\times 10^{-14}$ 95%
Comb$^b$ 2.3$^a$ $(5.1\pm0.5)\times 10^{-2}$ $1.84\pm0.14$ $(5.2\pm0.3)\times 10^{-14}$ 85%
Comb$^c$ 2.3$^a$ $(5.6\pm0.8)\times 10^{-2}$ $1.90\pm0.19$ $(5.6\pm0.4)\times 10^{-14}$ 57%
00031121001 2.3$^a$ 293$\pm$6 1.74$\pm$0.05 $(2.93\pm0.06)\times 10^{-10}$ 0.97/270
00031121002 2.3$^a$ 106$\pm$3 1.59$\pm$0.08 $(1.06\pm0.03)\times 10^{-10}$ 1.01/166
00031121003 2.3$^a$ 45$\pm$3 1.7$\pm$0.08 $(3.8\pm0.1)\times 10^{-11}$ 0.67/91
00031121004 2.3$^a$ 24$\pm$1 1.6$\pm$0.1 $(2.1\pm0.1)\times 10^{-11}$ 0.95/73
00031121005 2.3$^a$ 10$\pm$1 1.9$\pm$0.2 $(7.6\pm0.6)\times 10^{-12}$ 0.63/19
00031121001 1.85$\pm$0.13 260$\pm$10 1.51$\pm$0.08 $(2.6^{+0.2}_{-0.03})\times 10^{-10}$ 0.94/269
00031121002 1.26$\pm$0.17 90$\pm$3 1.09$\pm$0.12 $(8.9^{+0.7}_{-0.01})\times 10^{-11}$ 0.88/165
00031121003 2.41$\pm$0.28 46$\pm$5 1.77$\pm$0.15 $(3.9^{+0.4}_{-0.01})\times 10^{-11}$ 0.68/90
00031121004 2.06$\pm$0.35 23$\pm$2 1.47$\pm$0.18 $(2.1^{+0.2}_{-0.02})\times 10^{-11}$ 0.95/72
00031121005 1.7$\pm$0.7 9$\pm$2 1.6$\pm$0.4 $(7.1^{+2.0}_{-0.1})\times 10^{-12}$ 0.63/18
------------- --------------------- -------------------------------------------------------------- --------------- --------------------------------------- ---------------------------------
[$^a$ ${\rm N_H}$ has been fixed to 2.3$\times10^{22}$ cm$^{-2}$.]{}\
[$^b$ Fit to Obs ID 4565, 4566, 4567, 9838, 8990, 9839, 9837 combined.]{}\
[$^c$ Fit to Obs ID 9838, 8990, 9839, 9837 combined.]{}\
[$^d$ Fit parameters affected by pile–up. ]{}\
[*Swift*]{} X–ray observations
------------------------------
The [*Swift*]{} satellite observed the field of during the decay of its 2008 outburst. We here report on results obtained using the X–ray telescope (XRT). A log of the observations is presented in Table \[swiftlog\]. The [*ftools*]{} software package tool [ *xselect*]{} has been used to extract source and background photons from regions centered on the known source position or a source free region on the CCD, respectively. Square boxes were used as extraction regions in the first two observations (Obs IDs 00031121001 and 00031121002) that were obtained in windowed timing mode. In that mode ten rows are read–out compressed into one, and the central 200 columns of the CCD are read. Hence, there is only one-dimensional spatial information allowing for faster CCD read–out and reducing effects of photon pile–up. The square boxes have a size of 40 pixels ($\approx
94\arcsec$) along the one available spatial dimension. Events were selected if their event–grade is 0–2[^3]. For the four remaining observations two–dimensional information is available as the CCD is read–out in frame–transfer mode. These observations are in what is called photon–counting mode. Circular extraction regions that have a radius of 50 arcseconds were used. Events were selected if their event–grade is 0–4.
We have determined spectral parameters in the same way as we did for the [*Chandra*]{} observations restricting the fits to photon energies in the range from 0.5–10 keV. We have used two approaches. In the first we keep the neutral hydrogen column density fixed to 2.3$\times
10^{22}$ cm$^{-2}$. All the spectral fits to the [*Swift*]{} data were done using a $\chi^2$ minimisation technique with 10 or more counts per spectral bin, except the spectral fits to the last observation (Obs. ID 00031121006). There, we employed the Cash statistics for estimating the best–fitting parameters, however, due to the low number of counts we could not extract meaningful spectral parameters. In the second approach we leave the neutral hydrogen column density free. The results of the spectral fits using both approaches are presented in Table \[spec\].
--------------------- -------------- ------------- -------------
Obs ID Observing MJD Time on
date (UTC) source (ks)
00031121001$^a$ 2008 Feb. 12 54508.06499 1.74
00031121002$^a$ 2008 Feb. 15 54511.61268 2.26
00031121003$^{b,c}$ 2008 Feb .19 54515.02349 3.35
00031121004$^b$ 2008 Feb. 22 54518.16801 4.62
00031121005$^{b,c}$ 2008 Feb. 26 54522.05481 3.38
00031121006$^b$ 2008 Mar. 04 54529.55583 3.53
--------------------- -------------- ------------- -------------
: A journal of the [*Swift*]{} XRT observations.[]{data-label="swiftlog"}
[$^a$ Windowed timing observing mode.]{}\
[$^b$ Photon counting observing mode.]{}\
[$^c$ Data point used in Figure \[rxcorr\].]{}
[*RXTE*]{} X–ray observations
------------------------------
In addition to the and observations we have used several archival RXTE observations in order to follow the flux evolution of the source after the source entered the low–hard state on the way to quiescence. We do not provide a full description of the source spectrum or power spectral properties since those have been described in depth for the 2003 outburst in @2009ApJ...698.1398M and, for this outburst, in @2008ATel.1378....1K. Instead, we use a simple power–law model to describe the X–ray spectrum in order to derive unabsorbed fluxes in the 0.5–10 keV band.
To this end we have extracted spectra from the [*RXTE*]{}’s Proportional Counter Array data using [*ftools*]{} version 6.6.2. We use only data from the Proportional Counter Unit 2 since that was always operational for these observations. We used the background model appropriate for bright sources and corrected the spectra for dead–time, even though the source count rate is such that dead–time effects are small. We added a systematic error of 0.6 per cent to the count rate in each spectral bin. Finally, we grouped the spectra such that there are at least 10 counts per spectral bin. We estimate the error on the unabsorbed flux that we thus determined to be 10 per cent.
-------------------- -------------- ------------- -------------------
Obs ID Observing Start time Flux (0.5-10 keV)
date MJD (UTC)
93427-01-03-03 2008 Jan. 29 54494.19092 1.5E-08
93427-01-03-04 2008 Jan. 31 54496.74373 6.1E-09
93427-01-04-01 2008 Feb. 1 54497.85618 3.6E-09
93427-01-04-00$^a$ 2008 Feb. 2 54498.83744 8.2E-09
93427-01-04-02$^a$ 2008 Feb. 4 54500.80062 3.8E-09
93427-01-04-03$^a$ 2008 Feb. 6 54502.82833 1.0E-09
93427-01-05-00$^a$ 2008 Feb. 8 54504.92185 9.0E-10
-------------------- -------------- ------------- -------------------
: A journal of the [*RXTE*]{} PCA observations.[]{data-label="rxtelog"}
[$^a$ Data point used in Figure \[rxcorr\].]{}\
Results: X–ray decay
--------------------
In Figure \[xlc\] we have plotted the X–ray decay light curve together with a best–fitting exponential towards the baseline quiescent flux. The best–fitting parameters are an exponential decay timescale of $\approx$4 days and a constant of 4.3$\times10^{-14}$. As can be seen from the figure the fit only globally describes the observations with the $\chi^2$ being very large (728 for 12 degrees of freedom). Obviously, the X–ray decay light curve is subject to additional variations beyond that described by the model (flares and other variability etc). During some of the observations we find strong evidence for such variability. In contrast, during the observations there is no evidence for such variability. Note that we have not used the [ *RXTE*]{} observations in the fit and we have used the and flux values determined fixing the ${\rm N_H}$ to 2.3$\times
10^{22}$ cm$^{-1}$ in the figure.
There is a difference in the spectral parameters obtained from the spectral fits to the observations 8985 and 8986 and the contemporaneous [*Swift*]{} observations 00031121003 and 00031121005. We attribute this to effects of pile–up in the observations. The very flat power law and the somewhat low ${\rm N_H}$ are signs of pile–up effects. Furthermore, at the observed count rates pile–up is indeed expected. The parameters for the spectral fits to the second observation suggest a low ${\rm N_H}$ and a very flat power law index as well. However, the observed count rate of $<$1 count per second is well below the threshold of a few hundred counts per second above which the XRT in windowed timing mode should be suffering from pile–up (@2006NCimB.121.1521M). Close investigation of the spectrum reveals two potential reasons for the low ${\rm N_H}$ and the flat power law index. First, there is emission at low energies that is not present in subsequent observations that are obtained in photon counting mode. It is unclear whether this emission is real (cf. @2006ApJ...653..525M) or due to calibration uncertainties of the windowed timing mode. Nevertheless, even when discarding all data below 1 keV the best–fitting power law still has a low index and the ${\rm N_H}$ is low. Second, there seems to be excess emission around 6.5 keV that can be well described by broad line emission, however, we lack the signal–to–noise to investigate the properties of this line in detail. Given that such (relativistically smeared) emission lines are found in several black hole candidate sources (cf. @2006ApJ...653..525M) we attribute the discordant spectral parameters found when the spectrum is fit with an absorbed power law only, to unresolved line emission affecting the fit.
VLA radio observations and radio decay
--------------------------------------
H1743-322 was observed with the Very Large Array (VLA) under project codes S9208 and AR642. The VLA was in its CnB and C configurations (for the last two observations) implying a fairly large beam size ($>$1.5). Observations were carried out in standard continuum mode with a 50 MHz bandwidth in each of two intermediate frequency (IF) pairs. A maximum of 14 retrofitted EVLA antennas were included in the array. The primary calibrator was 3C 286, which was used to set the flux scale according to the coefficients derived at the VLA by NRAO staff in 1999 as implemented in the 31Dec09 version of AIPS. The secondary calibrator was J1744$-$3116 (1 degree from the target source). Observations were carried out in fast switching mode to reduce target–calibrator slew time, using a 3.3 s integration time. Data calibration and imaging were carried out using standard procedures within AIPS. The source flux density was measured by fitting an elliptical Gaussian to the source in the image plane using the AIPS task JMFIT.
For a journal of the observations see Table \[vla-log\]. In Figure \[rlc\] we plot the observed radio 8.46 GHz decay light curve. Besides the VLA observations we use one observation obtained with the Austrialian Telescope Compact Array (ATCA) on Jan. 28, 2008 (MJD 54493) at 8.64 GHz. During this observation the source was not detected down to a 3 $\sigma$ rms upper limit of 0.15 mJy (@2008ATel.1378....1K; at 8.64 GHz). This stringent ATCA upper limit shows that the radio emission reactivated close in time to the transition to the low–hard state.
The solid line in Figure \[rlc\] is the best–fitting exponential decay to the blue open circles, i.e. the VLA 8.46 GHz detections when the source was in the low–hard state. The exponential decay timescale is 19$\pm$2 days. The last radio observation yields a stringent upper limit of 0.04 mJy, which is below the extrapolation of the best–fitting exponential. This might indicate that the radio decay accelerated. The increase in source flux on Feb. 24, 2008 (MJD 54520) compared with the preceding observations on Feb. 20 and 23, 2008 (MJD 54516, 54519) is atypical since the X–ray flux decayed by a factor of 15 in the same period. Such behaviour could have been caused by a radio flare. Note that the radio flare was apparently not accompanied by an X–ray flare. Alternatively, the X–ray flare has been missed as the radio observation on Feb. 24, 2008 occurred 1.3 days before a observation and 2.5 days after the observation nearest in time.
-------------- ------------ ----------- ---------------- --------------- --------------------------
Calender Frequency On source time Flux density Time between start X-ray
date MJD (days) GHz (minutes) (mJy/beam) and radio obs. (days)
2008 Jan. 21 54486.740 1.4 3.4 2.4$\pm$0.26 XXX
2008 Feb. 20 54516.544 1.4 19 0.6$^b$ XXX
2008 Feb. 23 54519.680 1.4 18 0.6$^b$ XXX
2008 Mar. 01 54526.579 1.4 18 0.5$^b$ XXX
2008 Jan. 19 54484.663 4.86 3.4 1.54$\pm$0.09 XXX
2008 Jan. 21 54486.748 4.86 6 1.10$\pm$0.07 XXX
2008 Feb. 03 54499.737 4.86 17 0.63$\pm$0.09 XXX
2008 Feb. 06 54502.551 4.86 2.4 0.54$\pm$0.12 XXX
2008 Feb. 07 54503.673 4.86 3.4 0.59$\pm$0.14 XXX
2008 Feb. 09 54505.668 4.86 93 0.52$\pm$0.06 XXX
2008 Jan. 10 54475.719 8.46 13 6.16$\pm$0.05 N.A.
2008 Jan. 19 54484.669 8.46 6 0.98$\pm$0.07 N.A.
2008 Jan. 21 54486.760 8.46 10 0.76$\pm$0.06 N.A.
2008 Feb. 03 54499.742 8.46 12 0.52$\pm$0.06 0.9 [*RXTE*]{}
2008 Feb. 05 54501.643 8.46 6 0.48$\pm$0.07 0.8 [*RXTE*]{}
2008 Feb. 06 54502.557 8.46 6 0.45$\pm$0.08 0.3 [*RXTE*]{}
2008 Feb. 07 54503.679 8.46 6 0.46$\pm$0.08 0.8 [*RXTE*]{}
2008 Feb. 09 54505.674 8.46 17 0.56$\pm$0.05 0.8 [*RXTE*]{}
2008 Feb. 19 54515.625 8.46 17 0.23$\pm$0.12 0.6
2008 Feb. 20 54516.552 8.46 18 0.21$\pm$0.05 1.5
2008 Feb. 23 54519.689 8.46 18 0.23$\pm$0.06 1.3
2008 Feb. 24 54520.694 8.46 16 0.31$\pm$0.05 1.3
2008 Mar. 01 54526.587 8.46 18 0.17$\pm$0.04 0.4
2008 Mar. 02 54527.529 8.46 51 0.13$\pm$0.03 0.4
2008 Mar. 08 54533.556 8.46 93 0.07$\pm$0.02 0.1
2008 Mar. 16 54541.550 8.46 124 0.06$^b$ 0.3
2008 Mar. 20 54545.432 8.46 355 0.04$^b$ 1.0
2008 Jan. 10 54475.720 22.46 7.5 3.01$\pm$0.21 XXX
-------------- ------------ ----------- ---------------- --------------- --------------------------
[$^a$ VLA 8.46 GHz observation reported in @2008ATel.1384....1R.]{}\
[$^b$ Three $\sigma$ rms limit.]{}
Radio spectral index
--------------------
On several nights near–simultaneous radio data has been obtained at different frequencies. We use this data to assess the radio spectral index. Initially, the spectral index is negative, implying optically thin radio emission. E.g. the 1.4, 4.86 to 8.46 GHz spectral index on MJD 54486 is -0.64$\pm$0.07.
After the transition to the low–hard state the spectral index is consistent with being 0. To obtain a more accurate measurement of the radio spectral index, we have averaged three short 4.86 GHz and 5 8.46 GHz observations close in time (in the range MJD 54499–54505) where the radio flux was consistent with being constant. The 4.86–8.46 GHz radio spectral index for these averages is 0.03$\pm$0.18. This is consistent with optically thick emission as has been observed in low–hard radio spectra.
However, the radio spectral index does not stay close to 0 during the decay to quiescence. We have averaged the 1.4 GHz and 8.46 GHz data in the MJD range 54516–54526. This data coincides with the radio flare observable at 8.46 GHz. Combining the three 1.4 GHz observations does provide a detection at 0.431$\pm$0.097 mJy. At 8.46 GHz the average flux is 0.175$\pm$0.028 mJy. The 1.4–8.46 GHz radio spectral index is -0.50$\pm$0.15. This implies optically thin radio emission late in the low–hard state.
Radio – X–ray correlation
-------------------------
In Figure \[rxcorr\] we plot the observed correlation between the X–ray and 8.46 GHz radio fluxes for using the X–ray observations closest in time to the radio observations (see footnotes to the journals of the X–ray observations and Table \[vla-log\]). The best–fitting power law with index 0.18$\pm$0.01 (1 $\sigma$) is overplotted. The index is less steep than the index of $\alpha=\approx$0.7 as found for several sources before (S$_\nu
\propto \nu^\alpha; [email protected]; ).
There are several effects that could distort the picture. First, the radio and X–ray data are not strictly simultaneous (see Table \[vla-log\]). There is typically less than 1 day between the start of the radio and X–ray observations, however, in one case the difference in start times is as much as 1.5 days. This together with the flares apparent in the radio light curve as well as in the [*RXTE*]{} and X–ray observations could result in a higher flux in either radio or X–ray which might confuse the apparent correlation. Short duration radio flares have been found in e.g. V404 Cyg (@2008MNRAS.388.1751M). Last, since the spatial resolution of our VLA observations is several arcseconds, the observed radio flux might include emission from a ballistic jet–ejection event earlier in the outburst. Although the ATCA non–detection shows that the initial radio emission probably related to jet–ejection events occurring before the transition to the low–hard state faded away, shocks further down the flow can lead to rebrightenings in radio (and X–ray) at a later stage. The fact that the radio spectral index during the epoch related to the radio flare at MJD 54520 is showing that the radio emission is optically thin strongly argues in favour of this scenario.
Given the low spatial resolution of our radio observations this emission can significantly contribute to the radio flux density during the subsequent measurements. XTE J1550–564 dramatically demonstrates this, with radio (and X–ray) emission from the state transition ejecta still brightening years after the outburst (@2002Sci...298..196C). The part of the radio decay of that our observations sample with detections lasts about 33 days. In the most extreme case of a jet/flow speed of c, in 33 days a distance of 0.028 pc can be travelled. At a source distance of 7.5 kpc this corresponds to 078 for a jet in the plane of the sky. Such an angular extent would still be unresolved by the VLA in the CnB and certainly C configuration. This indicates that similar to the case of XTE J1550–564, shocks in the jet flow could indeed be responsible for the enhanced radio emission. Since such an angular extend could potentially be detected by we averaged the four observations obtained after March 20, 2008. We registered the four observations to a common astrometric frame using the position of the brightest X–ray source (besides ) detected in the observations. After merging the registered frames, we determined the position of the X–ray source associated with in the same way as described above. We find that the position of in the average of the last 4 observations in 2008 is consistent with the best known radio position found by @2004ATel..314....1R, therefore, the shocked radio emission is either not responsible for X–ray emission or it has travelled significantly less than the 078 calculated above.
In order for radio emission from shocks downstream to be the cause of the flattening of the index of the radio – X–ray correlation, the radio emission should not be accompanied by a similar amount of X–ray emission as during the processes responsible for the index of $\approx$0.7 during the normal low–hard state decay.
In order to compare the normalisation of this correlation with that found in other sources we need to know the distance of (see @2004MNRAS.inpress). As mentioned in the introduction we use a distance of 7.5 kpc for . Comparing the normalisation of the radio – X–ray correlation for a distance of 7.5 kpc with that of other sources we find that lies above the area traced out by GX 339–4 and V 404 Cyg (@2003MNRAS.344...60G). Even lowering the distance by a factor 2 one still finds the source above GX 339–4 and V 404 Cyg. Finally, our fluxes and luminosities are given in the 0.5-10 keV band whereas the correlation was plotted in the 2–11 keV band by @2003MNRAS.344...60G. The influence of the different energy bands on the slope of the radio – X–ray correlation is minimal since the X–ray spectrum does not change significantly during the decay. The normalisation is influenced by the difference in energy bands, though. For comparison we also computed the 2–11 keV X–ray fluxes of our X–ray observations and we find that the 2–11 keV values are 80$\pm$5 per cent of the values at 0.5–10 keV. Thus, our 0.5–10 keV luminosities are slightly higher than the 2–11 keV luminosities bringing artificially closer to the area traced by GX 339–4 and V 404 Cyg. The difference in normalisation would be larger if we had used the exact same energy range.
Discussion
==========
Using , and [*RXTE*]{} X–ray observations with (near–)simultaneous radio VLA observations we have observed the decay towards quiescence of the black hole candidate X–ray binary while the source was in the low–hard state. The overall shape of the X–ray decay light curve can be described with an exponential with an e–folding timescale of $\approx$4 days that levels off towards quiescence. X–ray flares are superposed on the exponential decay light curve. During several of the observations flares were observed as well. The unabsorbed 0.5–10 keV quiescent flux is 4.3$\times 10^{-14}$. For an assumed source distance of 7.5 kpc this implies a quiescent 0.5-10 keV X–ray luminosity of $\approx
3\times 10^{32}$. The source luminosity in quiescence after the outburst ending early 2008 is consistent with being the same as that after the 2004 outburst (@2005ApJ...632..504C).
The quiescent X–ray luminosity of $3\times10^{32}$ of implies an orbital period longward of $\approx$10 hours assuming follows the general trend between orbital period and quiescent X–ray luminosity reported in @2001ApJ...553L..47G. Such an orbital period suggests that the mass donor star in has evolved off the main sequence. Alternatively, the mass of the black hole in could be substantially larger than that in the sources defining the relation between orbital period and quiescent X–ray luminosity.
The low–hard state radio light curve shows evidence for radio flares superposed on an exponential decay with an e–folding timescale of $\approx 19$ days. When plotted against the X–ray light curve the radio – X–ray correlation index is lower ($\alpha=0.18\pm0.01$) than found before ($\alpha\approx 0.7$; ; @2003MNRAS.344...60G). The radio flares could be responsible for extended jet emission that due to the relatively large VLA beamsize in our observations using C and CnB configurations was not resolved. The 1.4–8.46 GHz radio spectral index during the MJD 54520 flare is -0.50$\pm$0.15 indicating that the radio emission is probably optically thin.
Relativistic arcsecond scale jets have not been found before for sources in the low–hard state (cf. @2006csxs.book..381F; @2009MNRAS.396.1370F). Therefore, it seems more likely that jet ejections earlier in the outburst of caused shocks when travelling along the jet. We note that similar events could be behind the apparent scatter and variability in the observed radio – X–ray correlation in other sources (cf. @2008MNRAS.389.1697C). The resulting index then depends on the amount and energy in the shocks caused by earlier ejection events. These will change from source to source and from outburst to outburst. For instance, plotting the radio and X–ray detections in the low–hard state during the decay @2009ApJ...698.1398M we find that two of the three 2003 data points fall approximately a factor 2 to 3 below the curve in Figure \[rxcorr\].
Increases in the radio flux density in the low–hard state by factors of 3–10 have been observed in the short–timescale radio flares in the quiescent state of V 404 Cyg (@2008MNRAS.388.1751M). Given that the radio and X–ray data that we presented here are not exactly simultaneous, any hour–long radio flare would have no corresponding flare in the X–ray band. Furthermore, even if the observations had been strictly simultaneous, the short integration times of our initial radio observations implies that an X–ray flare could have been missed since such an X–ray flare would take tens of minutes or longer to propagate downstream to the point in the compact jet where the optical depth is $\approx$1 at 8.46 GHz. Nevertheless, in order to produce a slope of $\approx 0.2$ the radio observations must have been close to the peak of flares, which seems unlikely. We conclude that optically thin radio emission from shocks caused by jet ejections earlier in the outburst influence the late time low–hard state radio emission.
Previous studies of black hole (candidate) sources decaying via the hard state to quiescence suggests that the power–law index of the X–ray spectrum in the quiescent state is different from that in the low–hard state. @1994PASJ...46..375E observed GS 1124–68 with [*Ginga*]{} and found a power-law index that remained essentially constant at $\sim$1.6 during the time the source was decaying in the hard state, except at the lowest observed luminosity of $\sim5.5\times
10^{-5}$[$L_{Edd}$]{} (for $M_{BH}$=7$M_\odot$) when it was found to be 1.84$\pm$0.04. have reported a power-law spectral index of $\sim$2.2 around $5.3\times 10^{-6} (d/6\,{\rm
kpc})^2$[$L_{Edd}$]{} for GX 339–4 (for $M_{BH}$=5.8$M_\odot$). @2004ApJ...601..439T followed XTE J1650–500 down to $\sim1.4\times 10^{-5} (d/6\,{\rm kpc})^2$[$L_{Edd}$]{} (for $M_{BH}$=8.5$M_\odot$) with [*Chandra*]{}, and also found evidence for spectral softening. Finally, @2008MNRAS.389.1697C show that the power–law spectral index is softer in V404 Cyg in quiescence compared to that in the brighter low–hard state.
In Figure \[plindex\] we have plotted the power law index from the and spectral modelling on during the decay towards quiescence and in quiescence. The data point at the lowest X–ray flux is determined by combining the quiescent data obtained right after the outburst of early 2008 as well as the quiescence observations used by @2005ApJ...632..504C. There is no evidence for a softening of the power–law index towards and in quiescence, although for this source the relatively high neutral hydrogen column density limits the accuracy of the power–law determination. A fit of a constant power–law spectral index gives a best–fitting index of 1.704$\pm$0.008 with a $\chi^2=7$ for 9 degrees of freedom.
We have compared the X–ray flux decay rate of with that observed during the last part of the outburst decay from the black hole candidate XTE J1908+094 and we find that the fit function describing the decay of is a good approximation to the last phase of approximately three weeks presented in Jonker et al. (2004) for XTE J1908+094. Note however, that due to a reduced sampling that source was only observed twice during those three weeks. Nevertheless, it is interesting to see that the sources have a similar decay rate. If other sources also follow the same decay rate in X–rays this could provide constraints on the accretion disc model, such as the accretion disc to ADAF evaporation.
Acknowledgments {#acknowledgments .unnumbered}
===============
PGJ acknowledges support from a VIDI grant from the Netherlands Organisation for Scientific Research. MAPT acknowledges support from NASA grant GO8-9042A. DS acknowledges an STFC Advanced FellowshipThe National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.
[59]{} natexlab\#1[\#1]{}
L., [White]{} N. E., 2003, , 586, L71
K. A., 1996, in ASP Conf. Ser. 101: Astronomical Data Analysis Software and Systems V, Vol. 5, p. 17
D., [Nagata]{} T., [Iwata]{} I., [Kato]{} T., [Yamaoka]{} H., 2003, , 8112, 2
M., [Rib[ó]{}]{} M., [Rodriguez]{} J., [Chaty]{} S., [Corbel]{} S., [Goldwurm]{} A., [Frontera]{} F., [Farinelli]{} R., [D’Avanzo]{} P., [Tarana]{} A., [Ubertini]{} P., [Laurent]{} P., [Goldoni]{} P., [Mirabel]{} I. F., 2007, , 659, 549
W., 1979, , 228, 939
S., [Fender]{} R. P., [Tzioumis]{} A. K., [Tomsick]{} J. A., [Orosz]{} J. A., [Miller]{} J. M., [Wijnands]{} R., [Kaaret]{} P., 2002, Science, 298, 196
S., [Kaaret]{} P., [Fender]{} R. P., [Tzioumis]{} A. K., [Tomsick]{} J. A., [Orosz]{} J. A., 2005, , 632, 504
S., [Koerding]{} E., [Kaaret]{} P., 2008, , 389, 1697
S., [Nowak]{} M. A., [Fender]{} R. P., [Tzioumis]{} A. K., [Markoff]{} S., 2003, , 400, 1007
S., [Tomsick]{} J. A., [Kaaret]{} P., 2006, , 636, 971
K., [Ogawa]{} M., [Aoki]{} T., [Dotani]{} T., [Takizawa]{} M., [Tanaka]{} Y., [Yoshida]{} K., [Miyamoto]{} S., [Iga]{} S., [Hayashida]{} K., [Kitamoto]{} S., [Terada]{} K., 1994, , 46, 375
A. A., [McClintock]{} J. E., [Narayan]{} R., 1997, , 489, 865
R., 2006, [Jets from X-ray binaries]{}, [Lewin]{} W. H. G., [van der Klis]{} M., eds., pp. 381–419
R. P., [Gallo]{} E., [Jonker]{} P. G., 2003, , 343, L99
R. P., [Homan]{} J., [Belloni]{} T. M., 2009, , 396, 1370
E., 2007, in American Institute of Physics Conference Series, Vol. 924, The Multicolored Landscape of Compact Objects and Their Explosive Origins, [di Salvo]{} T., [Israel]{} G. L., [Piersant]{} L., [Burderi]{} L., [Matt]{} G., [Tornambe]{} A., [Menna]{} M. T., eds., pp. 715–722
E., [Fender]{} R. P., [Miller-Jones]{} J. C. A., [Merloni]{} A., [Jonker]{} P. G., [Heinz]{} S., [Maccarone]{} T. J., [van der Klis]{} M., 2006, , 370, 1351
E., [Fender]{} R. P., [Pooley]{} G. G., 2003, , 344, 60
P., [Makishima]{} K., [Durant]{} M., [Fabian]{} A. C., [Dhillon]{} V. S., [Marsh]{} T. R., [Miller]{} J. M., [Shahbaz]{} T., [Spruit]{} H. C., 2008, , 390, L29
M. R., [McClintock]{} J. E., [Narayan]{} R., [Callanan]{} P., [Barret]{} D., [Murray]{} S. S., 2001, , 553, L47
G. P., 1997, in Bulletin of the American Astronomical Society, Vol. 29, Bulletin of the American Astronomical Society, pp. 823–+
M., [Done]{} C., [Page]{} K., 2008, , 388, 753
J.-M., [Barret]{} D., [Lasota]{} J.-P., [McClintock]{} J. E., [Menou]{} K., [Motch]{} C., [Olive]{} J.-F., [Webb]{} N., 2003, , 399, 631
S., [Schulz]{} N. S., [Brandt]{} W. N., [Galloway]{} D. K., 2007, , 663, L93
B., [Soleri]{} P., [M[é]{}ndez]{} M., [Belloni]{} T., [Mostafa]{} R., [Wijnands]{} R., 2009, , 394, 2080
J., [Buxton]{} M., [Markoff]{} S., [Bailyn]{} C. D., [Nespoli]{} E., [Belloni]{} T., 2005, , 624, 295
P. G., [Gallo]{} E., [Dhawan]{} V., [Rupen]{} M., [Fender]{} R. P., [Dubus]{} G., 2004, , 351, 1359
, [E]{}, [Tomsick]{}, [A.]{} J., [Yamaoka]{}, [K.]{}, [Ueda]{}, [Y.]{}, 2008, The Astronomer’s Telegram, 1348, 1
E., [Tomsick]{} J. A., [Buxton]{} M. M., [Rothschild]{} R. E., [Pottschmidt]{} K., [Corbel]{} S., [Brocksopp]{} C., [Kaaret]{} P., 2005, , 622, 508
E., [Tomsick]{} J. A., [Corbel]{} S., [Tzioumis]{} T., 2008, The Astronomer’s Telegram, 1378, 1
L. J., [Holt]{} S. S., 1977, , 3099, 1
—, 1977, , 3106, 4
A. K. H., [McClintock]{} J. E., [Garcia]{} M. R., [Murray]{} S. S., [Barret]{} D., 2002, , 570, 277
E. G., [Fender]{} R. P., [Migliari]{} S., 2006, , 369, 1451
H. A., [Barthelmy]{} S. D., [Baumgartner]{} W., [Cummings]{} J., [Fenimore]{} E., [Gehrels]{} N., [Markwardt]{} C. B., [Palmer]{} D., [Parsons]{} A., [Sakamoto]{} T., [Skinner]{} G., [Tueller]{} J., [Ukwatta]{} T., 2009, The Astronomer’s Telegram, 2058, 1
T. J., 2003, , 409, 697
S., [Nowak]{} M., [Corbel]{} S., [Fender]{} R., [Falcke]{} H., 2003, , 397, 645
C. B., [Swank]{} J. H., 2003, The Astronomer’s Telegram, 133, 1
J. E., [Remillard]{} R. A., 2006, [Black hole binaries]{}, [Lewin]{} W. H. G., [van der Klis]{} M., eds., pp. 157–213
J. E., [Remillard]{} R. A., [Rupen]{} M. P., [Torres]{} M. A. P., [Steeghs]{} D., [Levine]{} A. M., [Orosz]{} J. A., 2009, , 698, 1398
S., [Fender]{} R., [M[é]{}ndez]{} M., 2002, Science, 297, 1673
J. M., [Homan]{} J., [Miniutti]{} G., 2006, , 652, L113
J. M., [Homan]{} J., [Steeghs]{} D., [Rupen]{} M., [Hunstead]{} R. W., [Wijnands]{} R., [Charles]{} P. A., [Fabian]{} A. C., 2006, , 653, 525
J. M., [Raymond]{} J., [Homan]{} J., [Fabian]{} A. C., [Steeghs]{} D., [Wijnands]{} R., [Rupen]{} M., [Charles]{} P., [van der Klis]{} M., [Lewin]{} W. H. G., 2006, , 646, 394
J. C. A., [Gallo]{} E., [Rupen]{} M. P., [Mioduszewski]{} A. J., [Brisken]{} W., [Fender]{} R. P., [Jonker]{} P. G., [Maccarone]{} T. J., 2008, , 388, 1751
T., [Romano]{} P., [Mangano]{} V., [Moretti]{} A., [Cusumano]{} G., [La Parola]{} V., [Troja]{} E., [Campana]{} S., [Chincarini]{} G., [Tagliaferri]{} G., [Capalbi]{} M., [Perri]{} M., [Giommi]{} P., [Burrows]{} D., 2006, Nuovo Cimento B Serie, 121, 1521
R., [Barret]{} D., [McClintock]{} J. E., 1997, , 482, 448
L., [Rodriguez]{} J., [Cadolle Bel]{} M., [Kuulkers]{} E., [Hanke]{} M., [Tomsick]{} J., [Corbel]{} S., [Coriat]{} M., [Wilms]{} J., [Goldwurm]{} A., 2009, , 494, L21
M., [Chernyakova]{} M., [Capitanio]{} F., [Westergaard]{} N. J., [Shoenfelder]{} V., [Gehrels]{} N., [Winkler]{} C., 2003, The Astronomer’s Telegram, 132, 1
J., [Bel]{} M. C., [Tomsick]{} J. A., [Corbel]{} S., [Brocksopp]{} C., [Paizis]{} A., [Shaw]{} S. E., [Bodaghee]{} A., 2007, , 655, L97
M. P., [Dhawan]{} V., [Mioduszewski]{} A. J., 2008, The Astronomer’s Telegram, 1384, 1
M. P., [Mioduszewski]{} A. J., [Dhawan]{} V., 2004, The Astronomer’s Telegram, 314, 1
D. M., [Fender]{} R. P., [Hynes]{} R. I., [Brocksopp]{} C., [Homan]{} J., [Jonker]{} P. G., [Buxton]{} M. M., 2006, , 371, 1334
N., [Swank]{} J., [Shrader]{} C. R., [Rupen]{} M., [Beckmann]{} V., [Markwardt]{} C. B., [Smith]{} D. A., 2007, , 655, 434
P., [Heinz]{} S., [Fender]{} R., [Wijnands]{} R., [Tudose]{} V., [Altamirano]{} D., [Jonker]{} P. G., [van der Klis]{} M., [Kuiper]{} L., [Kaiser]{} C., [Casella]{} P., 2009, , 397, L1
D., [Miller]{} J. M., [Kaplan]{} D., [Rupen]{} M., 2003, The Astronomer’s Telegram, 146, 1
J. A., [Kalemci]{} E., [Kaaret]{} P., 2004, , 601, 439
Y. Q., [Cui]{} W., 2007, , 466, 1053
F., [Cui]{} W., 2005, , 629, 408
[^1]: email : [email protected]
[^2]: see http://heasarc.gsfc.nasa.gov/docs/xanadu/xspec/manual/
[^3]: http://swift.gsfc.nasa.gov/docs/swift/analysis/xrt\_swguide\_v1\_2.pdf
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We present a sample of $i_{775}$-dropout candidates identified in five Hubble Advanced Camera for Surveys fields centered on Sloan Digital Sky Survey QSOs at redshift $z\sim 6$. Our fields are as deep as the Great Observatory Origins Deep Survey (GOODS) ACS images which are used as a reference field sample. We find them to be overdense in two fields, underdense in two fields, and as dense as the average density of GOODS in one field. The two excess fields show significantly different color distributions from that of GOODS at the 99% confidence level, strengthening the idea that the excess objects are indeed associated with the QSO. The distribution of $i_{775}$-dropout counts in the five fields is broader than that derived from GOODS at the 80% to 96% confidence level, depending on which selection criteria were adopted to identify $i_{775}$-dropouts; its width cannot be explained by cosmic variance alone. Thus, QSOs seem to affect their environments in complex ways. We suggest the picture where the highest redshift QSOs are located in very massive overdensities and are therefore surrounded by an overdensity of lower mass halos. Radiative feedback by the QSO can in some cases prevent halos from becoming galaxies, thereby generating in extreme cases an underdensity of galaxies. The presence of both enhancement and suppression is compatible with the expected differences between lines of sight at the end of reionization as the presence of residual diffuse neutral hydrogen would provide young galaxies with shielding from the radiative effects of the QSO.'
author:
- Soyoung Kim
- Massimo Stiavelli
- 'M. Trenti'
- 'C.M. Pavlovsky'
- 'S.G. Djorgovski'
- 'C. Scarlata'
- 'D. Stern'
- 'A. Mahabal'
- 'D. Thompson'
- 'M. Dickinson'
- 'N. Panagia'
- 'G. Meylan'
title: The Environments of High Redshift QSOs
---
INTRODUCTION
============
Observational astronomy has finally reached the point of beginning to probe the era of reionization of hydrogen. The long search for Gunn-Peterson [@Gunn65] troughs in the spectra of increasingly higher redshift QSOs has finally become fruitful with the Sloan Digital Sky Survey (SDSS). A dramatic increase in the intergalactic hydrogen absorption at $z\simeq6$ was detected in the spectra of high-redshift SDSS QSOs [e.g., @Becker01; @Djorgovski01; @White03]. This was followed by the possible detection of a Gunn-Peterson trough in the spectrum of QSO SDSS J1030+0524 at $z=6.28$ [@Fan01]. The case for termination of the reionization epoch at $z\sim6$ is now relatively solid [e.g., @Fan06], even if not universally agreed upon [e.g., @Lidz06; @Bolton07]. At the same time, the Compton optical depth $\tau=0.084\pm0.016$ from the five year WMAP data [@Komatsu08] is compatible with a somewhat extended reionization process terminating at $z\simeq6$ [e.g., @Shull07].
Despite the growing consensus that reionization may have terminated at $z\simeq6$, it is extremely unlikely that it occurred in a universally synchronized fashion. Fluctuations from line of sight to line of sight are generally expected due to clumpiness of the IGM, and the gradual development and clumpy distribution of the first ionizing sources, either proto-galaxies or early AGN [e.g., @Miralda00]. Thus, reionization is expected to occur gradually as the UV emissivity increases [cf. @McDonald01], with the lowest density regions becoming fully reionized first. This is also suggested by modern numerical simulations [e.g., @Ciardi03; @Gnedin97; @Gnedin04] which predict an extended period of reionization, starting at $z \sim 15$ or even higher and ending at $z\sim 6$ [see also @Cen03; @Haiman03; @Somerville03; @Wyithe03].
If reionization is completed at $z\simeq6$, it is reasonable to attempt to identify the galaxies responsible for it. The combined Great Observatory Origins Deep Survey [GOODS; @Giavalisco04] and Hubble Ultra Deep Field [HUDF; @Beckwith06] have provided a large sample of $i_{775}$-dropout galaxies. Unfortunately, their estimated ionizing flux is insufficient to reionize the universe under standard assumptions [@Bunker04; @Dickinson04; @Bouwens07]; one would have to assume top heavy, very metal-poor stellar populations [@Stiavelli04], or rely on a burgeoning population of dwarf galaxies brought about by a steep faint end slope of the luminosity function [@Yan04]. A last alternative is that reionization was very gradual [e.g., @Bouwens07]. Unfortunately, testing these ideas is observationally challenging. At the same time, given the predominance of the HUDF data on the derivation of the faint end luminosity function of $i_{775}$-dropouts, one is led to wonder how much these results are affected by cosmic variance given the small volume probed by the HUDF. On this issue, conflicting claims regarding the density of HUDF $i_{775}$-dropouts can be found in the literature, with @Bouwens07 arguing in favor of an underdensity [see also @Oesch07] while @Malhotra05 argued in favor of an overdensity [however, see @TS08].
In general, one would expect very high redshift galaxies to be highly clustered, especially if purely gravitational clustering effects were amplified by positive feedback. Thus, in order to address the importance and sign of feedback in the environments where they should be easiest to detect, we were led to focus on fields centered on $z{{\>\rlap{\raise2pt\hbox{$>$}}\lower3pt\hbox{$\sim$}\>}}6$ QSOs as they should be the most clustered environments at these very high redshifts and the strongest cases of feedback available for study.
Indeed, a generic expectation in most models of galaxy formation is that the most massive density peaks in the early universe are likely to be strongly clustered [@Kaiser84; @Efstathiou88]. The evidence for such bias is already seen with large samples of Lyman-break galaxies at $z\sim3 - 3.5$ [@Steidel03], and in Lyman $\alpha$ selected galaxy samples [e.g., @Venemans03; @Ouchi05], and it should be even stronger at higher redshifts. An excess in the number of galaxies and in the density of star formation was also discovered in a systematic Keck survey of fields centered on known $z>4$ quasars [e.g., @Djorgo99; @Djorgovski99; @Djorgovski03]. The high metallicity associated with QSOs [@Barth03] – even at $z {{\>\rlap{\raise2pt\hbox{$>$}}\lower3pt\hbox{$\sim$}\>}}6$ – is often interpreted as evidence that they are located at the center of massive (proto–)galaxies, thereby corroborating the overall picture. These arguments justify the expectation that QSOs at $z
\simeq 6$ most likely highlight some of the first perturbations that become non–linear in the density distribution of matter [see e.g., @TS07].
However, QSOs are not “quiet neighbors”. The intense emission of ionizing radiation associated with QSOs ionizes the surrounding IGM and may even photo-evaporate gas in neighboring dark halos before this has an opportunity to cool and form stars [@Shapiro01]. In this context, QSOs would suppress galaxy formation in their vicinities. One would then observe a paucity of galaxies near a QSO despite the underlying excess of dark halos. Moreover, near the reionization epoch the fraction of neutral hydrogen in the IGM may change rapidly, possibly shifting the balance of the two effects. It would be exciting to see a change from source enhancement to suppression around reionization by observing a sample of $z=6$ QSOs.
It is with this goal in mind that we started a study of the environment of the five then known QSOs at $z \gtrsim 6$ using the Advanced Camera for Surveys (ACS) on board the Hubble Space Telescope (HST) to obtain images in the F775W ($i_{775}$) and in the F850LP ($z_{850}$) filters so as to identify candidate objects at $z=6$ as $i_{775}$-dropout galaxies. All five fields were observed to the same depth as GOODS in the $i_{775}$ and $z_{850}$ bands so that GOODS can be used as a reference field sample.
In a previous paper [@Stiavelli05], we analyzed the number of $i_{775}$-dropout galaxies identified in a HST/ACS field centered on the SDSS QSO J1030+0524 at $z=6.28$. In this field we found a very significant excess of sources compared to the density of $i_{775}$-dropouts seen in GOODS, thus suggesting that clustering wins over negative feedback. @Zheng06 also observed a radio-loud QSO at $z\sim6$, SDSS J0836+0054, using ACS and detected a significant overdensity of i-dropout galaxies in its vicinity. In this paper, we analyze four additional QSO fields in order to test and expand this result.
Section 2 is a description of the observations and data analysis. Section 3 describes our $i_{775}$-dropout objects and their properties. Section 4 contains discussion of our results and Section 5 summarizes our conclusions. In this paper we use AB magnitudes and assume the cosmological parameters, $H_{0}= 70\, {\rm km}\, {\rm s}^{-1}\, {\rm Mpc}^{-1}$, $\Omega_{m}= 0.26$, and $\Omega_{\Lambda}= 0.74$.
DATA REDUCTION AND ANALYSIS
===========================
We observed five fields centered on five SDSS QSOs at redshift $z{{\>\rlap{\raise2pt\hbox{$>$}}\lower3pt\hbox{$\sim$}\>}}6$ with the ACS/WFC on board HST. The QSOs were the most distant quasars known at the time of our original Cycle 12 proposal. All are radio-quiet. Our targets were SDSS J1148+5251 at $z=6.40\pm0.01$ [@Barth03], SDSS J1030+0524 at $z=6.28\pm0.03$, SDSS J1306+0356 at $z=5.99\pm 0.03$, SDSS J1048+4637 at $z=6.23\pm0.03$, and SDSS J1630+4012 at $z=6.05\pm0.03$ [@Fan01; @Fan03].
Table 1 summarizes the observations. Our observations in the F775W ($i_{775}$) and the F850LP ($z_{850}$) filters were designed to have similar exposure times to those used for the original (version 1.0) GOODS data products. The data were processed by the ACS pipeline CALACS that carries out bias and dark current removal and flat-fielding. The individual calibrated images ([*flt*]{} files) were combined into a single image for each filter using Multidrizzle, a pyraf application based upon the drizzle algorithm [@Fruchter02]. Drizzle also requires weight maps which we computed following the same procedure as was used for the GOODS data reduction: $$Variance = \frac{\left[ (Dt+fB) + \sigma_{read} ^{2}
\right]}{(ft)^{2}}$$ $$Weight = \frac{1}{(Variance)}$$ where $D$ is the dark current (electron/sec/pixel), $f$ is the pixel value of the reference flat field, $B$ is the background (electron/pixel) measured in flat-fielded images, $t$ is the exposure time (second), and $\sigma_{read}$ is the read-out noise (electron/pixel). We ran MultiDrizzle (Koekemoer et al. 2002) with parameters [*pixfrac*]{}= 1.0, [*final\_scale*]{}= 0.03 and [*final\_wht\_type*]{}= [*ivm*]{} (individual weight map). The area of the final images is approximately 11.3 arcmin$^{2}$. We measured the actual background noise in the drizzled ACS images, measuring and correcting for the correlation between pixels introduced by the drizzling and resampling process, and compared this to the variance predicted by the noise model used to generate the weight maps (equations 1 and 2). This correction was also verified by block averaging the images and measuring the resulting noise directly on scales larger than the inter-pixel correlation lengths. The variance maps were adjusted using this correction, and converted to rms maps which were provided to SExtractor [@Bertin96] to modulate the source detection thresholds and to compute photometric uncertainties.
The catalogs were obtained using SExtractor, run on the drizzled science images and with the same input parameters as those for the GOODS catalogs (for both the HDFN and the CDFS). We applied the same procedures to all five fields. The $z_{850}$ band images were used as the detection images when running SExtractor in dual-image mode. We required objects to be detected at a signal-to-noise (S/N)$>5$ in the $z_{850}$ band. For the total magnitude of a source, we adopted SExtractor’s MAG\_AUTO values. The adopted magnitude zero points were 25.6405 and 24.8432 in $i_{775}$ and $z_{850}$, respectively. We computed $i_{775}-z_{850}$ colors using the MAG\_ISO values to compare the same isophotes in the two bands. For $i_{775}$ band sources detected at less than the two sigma level in isophotal apertures, we computed lower limits for the colors using the 2$\sigma$ upper limit to the $i_{775}$ band isophotal flux. The Galactic extinction estimate of E(B-V) was obtained from @Schlegel98 for GOODS and each QSO field. We determined the corrections for the $i_{775}$ and $z_{850}$ magnitudes using SYNPHOT. The actual corrections in the two bands were as follows: 0.024 and 0.018 for HDFN, 0.016 and 0.012 for CDFS, 0.048 and 0.036 for J1030+0524; 0.022 and 0.016 for J1630+4012; 0.036 and 0.027 for J1048+4637; 0.044 and 0.033 for J1148+5251; and 0.060 and 0.045 for J1306+0356. The limiting magnitudes and completeness levels were comparable to those of GOODS catalogs.
CANDIDATE OBJECTS
=================
The selection criteria are based on the $i_{775}-z_{850}$ color, a magnitude limit $z\leq26.5$, limits on S/N ratios, and the SExtractor extraction $flag=0$ which identifies non-saturated and isolated sources outside the masked zones. We have considered two different values of S/N$=5$ and $8,$ and the color limits of $i_{775}-z_{850}=
1.3$ and $1.5$. Objects selected with S/N$>5$ and $i_{775}-z_{850} > 1.3$ will constitute our least restrictive sample S1; objects with S/N$>5$ and $i_{775}-z_{850} > 1.5$ are our sample S2; and those with S/N$>8$ and $i_{775}-z_{850} > 1.3$ are our sample S3. We eliminate objects that reside near the edges and on the star diffraction spikes, as well as objects that appear to be artifacts during visual inspection. GOODS candidates were selected by the same selection criteria using the GOODS catalogs (version 1.1), including visual inspection. However, as the QSO fields only have ACS imaging in two bands, we do not require non-detections ($<2 \sigma$) in the $B_{435}$ and $V_{606}$ as was implemented in the @Dickinson04 selection of $i_{775}$-dropouts in the GOODS fields. Therefore, our GOODS $i_{775}$-dropout sample is different from the one used in @Dickinson04. Table 2 shows the number of $i_{775}$-dropouts selected in QSO fields and GOODS for different S/N ratios and color limits. In Table 2, the number of $i_{775}$-dropouts in GOODS is normalized to the area of a single ACS/WFC field ($\sim$ 11.3 arcmin$^{2}$). The measurements of all quasar field candidates with $i_{775}-z_{850}
> 1.3$ and S/N$>5$ are listed in Table 3.
Contamination by stars is a potential concern. We estimated a priori the possible contamination from stars by using as a proxy the number density of stars brighter than visual magnitude $m_{v}=21$ at the Galactic latitude of the five QSO fields. All fields have lower star density than the mean star density at the galactic latitude of each QSO [@Zombeck90] at the galactic latitude of each QSO. In particular, the J1030+0524 field has a lower star density than GOODS, while the other overdense field, J1630+4012, has a star density 4.8 times higher than GOODS. This suggests a degree of caution is necessary in excluding stars. We have identified stars using the SExtractor star-galaxy index, S/G, half-light radius, $r_{hl}$ and $z_{850}$ mag. The criteria for stars were S/G$\geq0.85$, $r_{hl}\leq0.1$ arcsecond, and $z_{850}<25.5$ applied to the S1 samples. We found no stellar $i_{775}$-dropout candidates in our five fields but found 16 stellar $i_{775}$-dropout candidates (0.55 stars per ACS field) in GOODS.
Our target QSOs are not all flagged as stars because of the long wavelength point source halo effect seen with ACS. The point spread function in the F850LP filter is characterized by a long wavelength halo which is due to light traveling through the CCD, bouncing off the front side at a large angle, going once again through the CCD and being detected. This effect is very wavelength-dependent (and thus, for high-redshift QSOs, redshift-dependent). Well-exposed images of a QSO will show this extended halo and the QSO will fail to be identified as a star. The same would be true for very red stars. However, if we artificially dim the QSOs to have similar apparent magnitudes as the other $i_{775}$-dropouts, the halos drop below the noise level and the fainter versions of our QSOs are identified as stars.
We also estimated the possible contamination by stars fainter than 25.5 by considering the candidates with S/G$\geq0.85$, and half light radius $r_{hl}\leq0.1$ arcsecond. In Table 3, we have two objects (A8 and B2) in J1030+0524 and J1630+4012 that satisfy this relaxed criteria. When applied to GOODS, we found 11 (very red) objects (0.38 objects per ACS field) out of 235 objects selected using the S1 criteria.
For S/N$> 5$ and $i_{775}-z_{850}>1.3$ (our selection S1) we see that two fields, J1030+0524 and J1630+4012, show an overdensity; J1048+4637 has approximately the same number density of $i_{775}$-dropouts as GOODS; and the J1148+5251 and J1306+0356 fields appear underdense compared to GOODS.
We have verified whether the variations in the number of candidates could be due to field-to-field background noise variations. We find these variations to be generally small and that the background noise is highest in the field of J1030+0524, i.e., the one with the largest excess. Thus, we conclude that background noise variations are not affecting our results.
Figures 1 through 5 show for each field the number counts as a function of the $z_{850}$ magnitude (panel a) and as a function of $i_{775}-z_{850}$ color (panel b). Panel c shows the count distribution as a function of magnitude for objects redder than $i_{775}-z_{850}=1.5$ and panel d shows the number of objects redder than a given $i_{775}-z_{850}$ color in 0.1-magnitude bins for $i_{775}-z_{850}>0.9$. The solid line shows the data for galaxies in the QSO fields. The dotted line shows the distributions for the GOODS fields.
Figure 1-(d) shows the color distribution of galaxies in the J1030+0524 field (excluding the QSO) and GOODS. Their distributions appear to be different, especially around $i_{775}-z_{850}\sim 2$. In Figure 2-(d), the color distributions of J1630+4012 and GOODS appear to be different for $i_{775}-z_{850}>1.7$. We applied the Chi-square ($\chi^2$) test on the binned color distributions to determine the significance of the differences between the color distributions of the QSO fields compared to GOODS. We focused on sources with S/N $>5$ that fall in the color interval $1.3<i_{775}-z_{850}< 2.6$. For J1030+0524, the chi-square test yielded a $\chi^2$ statistic of 30 and a probability of $P=0.3$% where P is the one-tailed probability that obtains a value of $\chi^2$ or greater — e.g., there is less than a 0.3% chance that both the GOODS and J1030+0524 $i_{775}$-dropout samples were drawn from the same distribution over the color range considered. For the other overdense field, J1630+4012, we found $\chi^2=52$ and $P<0.1\%$. For the other three fields, $\chi^2= 11$ and $P=41$% for J1048+4637, $\chi^2=7$ and $P=83$% for J1148+5251, and $\chi^2=7$ and $P=83$% for J1306+0356. For two overdense fields, the probability is not more than 0.3% regardless of the specific criterion we use (S2 and S3 samples). Thus, our candidates in both overdense fields have significantly different color distributions compared to GOODS.
Figure 6 shows substantial spatial clustering of the $i_{775}$-dropout candidates in the J1030+0524 field: when the field is divided in half across the diagonal, almost all of the sources are in the south-west half of the field. This makes the excess in J1030+0524 even more significant. The color magnitude diagram of candidates listed in Table 3 is presented in Figure 7, showing that the overdense fields have fainter $i_{775}$-dropouts than GOODS. It is notable that @Willott05 in their less sensitive survey for $i_{775}$-dropouts around high-redshift SDSS QSOs, including J1030+0524, found no overdensities. The upper panel of Figure 8 shows half-light radius versus $z_{850}$ for the $i_{775}$-dropout candidates from GOODS and the QSO fields. There is an upper envelope to the size-magnitude relation, and the bottom panel of Figure 8 shows a histogram comparing the size distribution of GOODS and QSO field $i_{775}$-dropout half-light radii. It appears that the candidates in the overdense fields are more compact than those in GOODS, but this is not statistically significant.
DISCUSSION
==========
Despite a complete reanalysis and a change in the type of SExtractor magnitudes used to compute the $i_{775}-z_{850}$ color for dropout selection (from AUTO to ISO mags), we confirm the overdensity in the J1030+0524 field reported in @Stiavelli05. The overdensity is significant not only in the counts by themselves but also in the color distribution. Indeed, the departure of the color distribution of J1030+0524 and J1630+4012 is in the sense of having an excess of red dropouts with precisely the colors that one would expect from objects at the redshift of the two QSOs. This makes the excess even more convincing.
One uncertain component of the comparison with GOODS is the possible contamination by low redshift and Galactic interlopers. Figure 9 shows the fraction of GOODS $i_{775}$-dropout objects selected by us but rejected when using the full GOODS $i_{775}$-dropout criteria including the $V_{606}$ data [@Beckwith06] to the number of GOODS $i_{775}$-dropouts selected by our criteria vs. the $i_{775}-z_{850}$ color. At $i_{775}-z_{850} > 1.7$, where the excess of $i_{775}$-dropouts is large in the J1030+0524 and J1630+4012 fields, there is less than 15% contamination from potential foreground objects. Statistically, the full GOODS criteria would remove more objects from GOODS than from the J1030+0524 or J1630+4012 field because the latter have a redder color distribution. Thus, we do not think that the detected excess is due to interloper contamination.
In order to understand how unusual it is to identify this distribution of over- and underdensities, we consider the number of $i_{775}$-dropouts identified in 30 distinct and non-overlapping ACS fields in GOODS. Figure 10 presents the resulting histogram of the number of $i_{775}$-dropouts identified per unique GOODS ACS field using the S1 and the S2 selection criteria. These distributions are reasonably well fit by Poisson distributions with a mean of 6.5 (3.13) $i_{775}$-dropouts per ACS field for the S1 (S2) selection criteria. Using these distributions from GOODS, we create 10,000 Monte Carlo (MC) quintuplets, where each MC quintuplet is generated by randomly selecting five independent numbers of $i_{775}$-dropouts, each corresponding to a single ACS field. We then test how many MC quintuplets have the counts we have observed. For the S1, we find that only $0.06 \pm 0.02$% of the MC quintuplets have exactly two overdense and two underdense fields. For the S2, this probability is only $0.03 \pm 0.09$%. For the S3, six $i_{775}$-dropouts in one ACS field is the maximum number among the 30 ACS fields in GOODS so any MC quintuplets cannot be generated to have more than six $i_{775}$-dropouts. However, since one QSO field has 10 $i_{775}$-dropouts, we have zero probability for S3. The error bars on these probabilities are calculated by considering variations between 10 independent subsets of 1,000 MC quintuplets. This comparison to GOODS empirical dropout statistics suggests that the QSOs are indeed affecting their environments.
Estimating the likelihood of the counts observed in our fields on the basis of the $i_{775}$-dropout count distribution in GOODS is not entirely appropriate as even GOODS is affected by cosmic variance because within both the CDFS and the HDFN, the ACS fields are all adjacent. We can use the conservative model of cosmic variance of @TS08 to estimate the likelihood of our detected counts. This model is based on extended Press-Schechter theory as well as synthetic catalogs extracted from N-body simulations of structure formation. In this case, we establish the probability with $10^{6}$ MC quintuplets. We find that the likelihood of a MC quintuplet matching our observed distribution of over- and underdense fields using the S1 criteria is $0.9\pm0.08\%$. S2 has a likelihood of $0.3\pm0.05\%$, and S3 has a likelihood of $0.8\pm0.09\%$. This result is less significant than that derived from the GOODS distribution, but it is comforting that the significance does not decrease when using samples with more stringent color or S/N selections. Thus, while we cannot claim for our overall sample a very significant detection of a discrepancy from a distribution dominated by cosmic variance alone, our distribution remains unlikely at the 99% level.
A criticism to this type of analysis is that these are not a-priori probabilities as we knew the outcome of the experiment before carrying out the statistical tests. This is only partly correct because the main idea of the HST proposal was indeed to look for overdensities or underdensities compared to the field even though the statistical test was not specified. Moreover, it is possible to design an experiment that does not depend as much on the observed counts, namely to evaluate the probability that out of the five fields only one is within one (Poissonian) $\sigma$ of the mean, i.e. within $8.08\pm2.84$ for selection S1, within $3.95\pm1.99$ for selection S2, or within $2.96\pm1.72$ for selection S3. Here the formal Poisson $\sigma$ is used only to define an inner interval and has no attached probability significance. Probabilities are estimated by comparing how our observed object count distribution compares to that expected from cosmic variance. We find that the probability of finding no more than one out of five fields in the inner interval is of 20% for S1, 4% for S2, and 5.8% for S3. The same a priori test based on the observed counts distribution in GOODS would give a probability of finding no more than one object in the inner interval of 1.5% for S1, 0.4% for S2, and 1.5% for S3. This reinforces the view that the QSO fields have a distribution of $i_{775}$-dropout counts broader than what is expected by cosmic variance alone.
CONCLUSIONS
===========
Summarizing our results, we find two fields where the numbers of $i_{775}$-dropout galaxies and their $i_{775}-z_{850}$ color distributions are significantly different (at 99% confidence) than the averages for galaxies selected in the same way from GOODS fields. When we look at the distribution of all five fields, we find that it is likely (at $80 - 96$% confidence level, depending on selection and specific statistical test) that the distribution of counts in the QSO fields is broader than that of GOODS and cannot be explained by cosmic variance alone.
We now discuss the possible implications of our results assuming that the departure from the expected distribution of field $i_{775}$-dropouts is indeed real. The fact that we observe both overdensities and underdensities is somewhat puzzling. We know that QSOs at $z=6$ are very rare objects and are most likely associated with overdensities on large scales. Tracing a pencil beam with the area of an ACS field through a cold dark matter (CDM) simulation box with the method of Trenti & Stiavelli (2008), we do not find correlations over $\Delta z \geq 0.3$. This is not surprising as $\Delta z=0.3$ corresponds to about 90 Mpc $h^{-1}$ at $z\sim6$ and on those scale the CDM power spectrum predicts a value of the mass fluctuation $\sigma_M$ many orders of magnitude lower than the value that can be associated with the QSO itself. From this point of view, the redshift range probed by $i_{775}$-dropouts spans at least three uncorrelated volumes.
A QSO at $z \sim 6$ is expected to live in the most massive halos within $\approx$ Gpc$^3$ comoving volumes, with masses of the order of $\approx 4 \times 10^{12} h^{-1} M_{\sun}$ [e.g., @Springel05]. Thus the dark matter halo mass function in the vicinity of the QSO halo will be biased by the presence of a rare overdensity [e.g., @Barkana04]. To quantify the impact of the QSO on the expected number counts in its immediate neighborhood we use the model of Munoz & Loeb (2007). From their Fig. 4 we derive that around the QSO there should be between 6 and 7 $i_{775}$-dropouts living in dark halos of mass $>5 \times 10^{10} h^{-1} M_{\sun}$ taking into account an assumed duty cycle of 0.25 for LBGs. The duty cycle is used to establish a halo mass scale for the observed galaxies by requiring that the number of halos of the required mass be equal to the number of objects divided by the duty cycle. Adopting a duty cycle allows us to determine a mass scale from the number of objects and to avoid using the ill-measured M/L of galaxies at z ${{\>\rlap{\raise2pt\hbox{$>$}}\lower3pt\hbox{$\sim$}\>}}6$. However, the results do not depend critically on the choice of duty cycle for range between 1 and 0.1. Our fields do not probably reach a depth that allows us to probe these halo masses with high completeness, but still we would expect to detect 2-3 of such LBGs or more if the “duty cycle” were higher.
In this light, deficits in the number of $i_{775}$-dropout candidates are surprising. Indeed 2/3 of the expected objects are in uncorrelated volumes and should not be affected by the presence of the QSO. The one third affected by the QSO now becomes a very small number and detecting a deficit in any single field is generally going to be statistically insignificant. It is interesting to note that at the time this project was planned the expected number of $i_{775}$-dropouts in GOODS was thought to be higher [e.g., @Dickinson04] so that a deficit would have been better quantifiable. Despite these considerations, the fact remains that we do seem to detect fields that have a deficit of $i_{775}$-dropout counts compared to the field. If we really had physical overdensities and physical underdensities near the QSO, what would be the origin of this effect? One possible explanation is that two physical mechanisms are simultaneously at play: the density of halos near the QSO is indeed higher but feedback by the QSO prevents many of these halos from becoming galaxies. The regions generated by luminous quasars can affect the formation and clustering of galaxies. @Wyithe05 derived size from displacement of quasar host galaxy redshift and the Gunn-Peterson trough redshift. The region size of the QSO J1030+0524 is the largest of the five quasars but the second overdense field, J1630+4012 and the most underdense field, J1306+0356, have very similar region sizes. The field with density comparable to GOODS, J1048+4637, has the smallest region size.
Thus, we find no evident correlation between density of $i_{775}$-dropouts and region size. This may or may not be significant as region sizes are roughly correlated with the luminosity of the quasars and their lifetimes; the latter measurements are not very accurate. We see a weak trend between counts and QSO luminosity as the two faintest QSOs are the two overdense ones and the most luminous QSO is one of the underdense ones. However, the most underdense QSO field(J1306+0356) is the third luminous QSO and within 0.04 mag from that of the most overdense (J1030+0524). On the basis of these considerations we conjecture that the suppression of galaxy formation which we may be witnessing could be the result of percolation of ionized Hydrogen bubbles. This would make it dependent, but not uniquely driven, by the QSO properties. Clearly it would be desirable to study these effects with better statistics.
Interestingly, @Maselli08, with an entirely different method, find conclusion similar to ours: namely, that J1630+4012 is overdense while J1148+5251 and J1306+0356 are underdense. They also find an overdensity around SDSS J0836+0054, also found to be overdense by @Zheng06. @Maselli08 predict that ionizing radiation from clustered galaxies for J1630+4012 exceeds the one from the quasar by a factor of five. We estimate the ionizing flux of our candidates. The total UV flux observed by summing the $z_{850}$ photometry from all of our $i_{775}$-dropout candidates is 7.0% and 8.5% of the quasar flux in $z_{850}$ for J1030+0524 and J1630+4012, respectively. For any reasonable spectral energy distribution, the excess is too small to affect the ionizing contributions. In order to have an influence at this level of overdensity, the excess should span a much larger area than that provided here. Clearly to further clarify these findings, we would need a larger sample as well as more extended data over overdense fields.
We thank the referee for careful reading and valuable comments. This work was partially supported by HST GO grant of 01087 and 01168. SGD and AAM acknowledge a partial support from the NSF grant AST-0407448, and the Ajax Foundation. The work of DS was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with NASA.
Barkana, R., & Loeb, A. 2004, , 609, 474
Barth, A. J., Martini, P., Nelson, C. H., & Ho, L. C. 2003, , 594, L95
Becker, R. H., et al.2001, , 122, 2850
Beckwith, S. V. W., et al. 2006, , 132, 1729
Bertin, E., & Arnouts, S. 1996, , 117, 393
Bolton, J. S., & Haehnelt, M. G. 2007, , 374, 493
Bouwens, R. J., Illingworth, G. D., Franx, M., & Ford, H. 2007, , 670, 928
Bunker, A. J., Stanway, E. R., Ellis, R. S., & McMahon, R. G. 2004, , 355, 374
Cen, R. 2003, , 591, 12
Ciardi, B., Ferrara, A., & White, S. D. M. 2003, , 344, L7
Dickinson, M., et al. 2004, , 600, L99
Djorgovski, S. G. 1999, The Hy-Redshift Universe: Galaxy Formation and Evolution at High Redshift, 193, 397
Djorgovski, S. G., Castro, S., Stern, D., & Mahabal, A. A. 2001, , 560, L5
Djorgovski, S. G., Odewahn, S. C., Gal, R. R., Brunner, R. J., & de Carvalho, R. R. 1999, Photometric Redshifts and the Detection of High Redshift Galaxies, 191, 179
Djorgovski, S. G., Stern, D., Mahabal, A. A., & Brunner, R. 2003, , 596, 67
Efstathiou, G., & Rees, M. J. 1988, , 230, 5P
Fan, X., et al. 2001, , 122, 2833
Fan, X., et al. 2003, , 125, 1649
Fan, X., et al. 2006, , 132, 117
Fruchter, A. S., & Hook, R. N. 2002, , 114, 144
Giavalisco, M., et al. 2004, , 600, L93
Gnedin, N. Y. 2004, , 610, 9
Gnedin, N. Y., & Ostriker, J. P. 1997, , 486, 581
Gunn, J. E., & Peterson, B. A. 1965, , 142, 1633
Haiman, Z., & Holder, G. P. 2003, , 595, 1
Kaiser, N. 1984, , 284, L9
Koekemoer, A. M., Fruchter, A. S., Hook, R. N., & Hack, W. 2002, The 2002 HST Calibration Workshop : Hubble after the Installation of the ACS and the NICMOS Cooling System, Proceedings of a Workshop held at the Space Telescope Science Institute, Baltimore, Maryland, October 17 and 18, 2002. Edited by Santiago Arribas, Anton Koekemoer, and Brad Whitmore. Baltimore, MD: Space Telescope Science Institute, 2002., p.337, 337
Komatsu, E., et al.2008, ArXiv e-prints, 803, arXiv:0803.0547
Lidz, A., Oh, S. P., & Furlanetto, S. R. 2006, , 639, L47
Malhotra, S., et al.2005, , 626, 666
Maselli, A., et al.2008, , submitted
McDonald, P., & Miralda-Escud[é]{}, J. 2001, , 549, L11
Miralda-Escud[é]{}, J., Haehnelt, M., & Rees, M. J. 2000, , 530, 1
Mu[ñ]{}oz, J. A., & Loeb, A. 2008, , 385, 2175
Oesch, P. A., et al.2007, , 671, 1212
Ouchi, M., et al. 2005, , 620, L1
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, , 500, 525
Shapiro, P. R., & Raga, A. C. 2001, Revista Mexicana de Astronomia y Astrofisica Conference Series, 10, 109
Shull, M., & Venkatesan, A. 2007, ArXiv Astrophysics e-prints, arXiv:astro-ph/0702323
Somerville, R. S., Bullock, J. S., & Livio, M. 2003, , 593, 616
Springel, V., et al. 2005, , 435, 629
Steidel, C. C., Adelberger, K. L., Shapley, A. E., Pettini, M., Dickinson, M., & Giavalisco, M. 2003, , 592, 728
Stiavelli, M., et al. 2005, , 622, L1
Stiavelli, M., Fall, S. M., & Panagia, N. 2004, , 610, L1
Trenti, M., & Stiavelli, M. 2007, , 667, 38
Trenti, M., & Stiavelli, M. 2008, , 676, 767
Venemans, B. P., Kurk, J. D., Miley, G. K., & R[ö]{}ttgering, H. J. A. 2003, New Astronomy Review, 47, 353
White, R. L., Becker, R. H., Fan, X., & Strauss, M. A. 2003, , 126, 1
Willott, C.J., et al., 2005, in “Growing Black Holes”, ESO (Garching), Merloni et al. eds, (also astro-ph/0410306)
Wyithe, J. S. B., & Loeb, A. 2003, , 588, L69
Wyithe, J. S. B., Loeb, A., & Carilli, C. 2005, , 628, 575
Yan, H., & Windhorst, R. A. 2004, , 600, L1
Zheng, W., et al. 2006, , 640, 574
Zombeck, M.V. 1990, Handbook of Space Astronomy and Astrophysics (Cambridge University Press: Cambridge), p.77
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'In 1976 Procesi and Schacher developed an Artin–Schreier type theory for central simple algebras with involution and conjectured that in such an algebra a totally positive element is always a sum of hermitian squares. In this paper elementary counterexamples to this conjecture are constructed and cases are studied where the conjecture does hold. Also, a Positivstellensatz is established for noncommutative polynomials, positive semidefinite on all tuples of matrices of a fixed size.'
address:
- 'University of California San Diego, Department of Mathematics, 9500 Gilman Drive, La Jolla, California 92093-0112, USA'
- 'School of Mathematical Sciences, University College Dublin, Belfield, Dublin 4, Ireland'
author:
- Igor Klep
- Thomas Unger
date: 18 October 2008
title: 'The Procesi–Schacher conjecture and Hilbert’s 17th problem for algebras with involution'
---
*Dedicated to David W. Lewis on the occasion of his 65th birthday.*
Introduction
============
Artin’s 1927 affirmative solution of Hilbert’s 17th problem (*Is every nonnegative real polynomial a sum of squares of rational functions?*) arguably sparked the beginning of the field of real algebra and consequently real algebraic geometry (cf. [@BCR; @PD]).
Starting with Helton’s seminal paper [@Hel], in which he proved that every positive semidefinite real or complex noncommutative polynomial is a sum of hermitian squares of *polynomials*, variants of Hilbert’s 17th problem in a *noncommutative* setting have become a topic of current interest with wide-ranging applications (e.g. in control theory, optimization, engineering, mathematical physics, etc.); see [@dOHMP] for a nice survey. Most of these results have a functional analytic flavour and are what Helton et al. call *dimensionfree*, that is, they deal with evaluations of noncommutative polynomials in matrix algebras of arbitrarily large size.
Procesi and Schacher in their 1976 Annals of Mathematics paper [@PS] introduce a notion of orderings on central simple algebras with involution, prove a real Nullstellensatz, and a weak noncommutative version of Hilbert’s 17th problem. A strengthening of the latter is proposed as a conjecture [@PS p. 404]: *In a central simple algebra with involution, a totally positive element is always a sum of hermitian squares.*
We explain in Section \[sec5\] how these results can be applied to study non-dimensionfree positivity of noncommutative polynomials. Roughly speaking, a noncommutative polynomial all of whose evaluations in $n\times n$ matrices (for *fixed* $n$) are positive semidefinite, is a sum of hermitian squares with denominators and weights.
A brief outline of the rest of the paper is as follows: in Section \[sec:PSc\] we fix terminology and summarize some of the Procesi–Schacher results in a modern language. Then in Section \[sec:counter\] we present counterexamples to the Procesi–Schacher conjecture, while Section \[para.pos.rel\] contains a study of examples (mainly in the split case) where the conjecture is true.
For general background on central simple algebras with involution we refer the reader to [@BOI] and for the theory of quadratic forms over fields we refer to [@Lam].
The Procesi–Schacher Conjecture {#sec:PSc}
===============================
Let $F$ be a formally real field and let $A$ be a central simple algebra with involution ${\sigma}$ and centre $K$. Assume that $F$ is the fixed field of ${\sigma}$ (i.e., ${\sigma}|_F=\mathrm{id}_F$). The involution ${\sigma}$ is of the *first kind* if $K=F$, and of the *second kind* (also called *unitary*) otherwise. In this case $[K:F]=2$ and ${\sigma}|_K$ is the non-trivial element in $\mathrm{Gal}(K/F)$.
Let $D$ be a division algebra over $K$ with involution $\tau$ and fixed field $F$. Let $h$ be an $n$-dimensional hermitian or skew-hermitian form over $(D,\tau)$. Then $h$ gives rise to an involution on $M_n(D)$, the *adjoint involution* $\operatorname{ad}_h$, defined by $$\operatorname{ad}_h(X)= H\cdot \tau(X)^t\cdot H^{-1},$$ for all $X\in M_n(D)$, where $H$ is the Gram matrix of $h$, $t$ denotes the transpose map on $M_n(D)$ and $\tau(X)$ signifies applying $\tau$ to the entries of $X$. It is well-known that every central simple algebra with involution $(A,{\sigma})$ is of the form $(M_n(D),\operatorname{ad}_h)$, where $n$ is unique, $D$ is unique up to isomorphism and $h$ is unique up to multiplicative equivalence (see [@BOI 4.A]).
If ${\sigma}$ is of the first kind, then ${\sigma}$ is called *orthogonal* or *symplectic* if ${\sigma}$ becomes adjoint to a quadratic or alternating form, respectively, after scalar extension to a *splitting field* of $A$ (i.e., an extension field $L$ of $K$ such that $A{\otimes}_K L\cong M_n(L)$). We denote the subspace of ${\sigma}$-symmetric elements of $A$ by $\operatorname{Sym}(A,{\sigma})$.
Let $\leq$ be an ordering on $F$. We identify $\leq$ with its *positive cone* $P=\{x\in F\,|\, 0\leq x\}$ via $$x\leq y {\Leftrightarrow}y-x\in P$$ for all $x,y\in F$. In this case we also write $\leq_P$ instead of $\leq$.
Procesi and Schacher [@PS §1] consider central simple algebras $A$, equipped with a *positive involution* ${\sigma}$, i.e., an involution whose *involution trace form* $T_{\sigma}$ is positive semidefinite with respect to the ordering $\leq_P$ on $F$, $$T_{\sigma}(x):=\operatorname{Trd}(\sigma(x)x)\geq_P 0 \quad\text{for all } x\in A.$$ Here $\operatorname{Trd}:A\to F$ (the *trace*) denotes the reduced trace $\operatorname{Trd}_{A/F}$ if ${\sigma}$ is of the first kind and the composition $\operatorname{Trd}_{K/F}\circ \operatorname{Trd}_{A/K}$ if ${\sigma}$ is of the second kind. The form $T_{\sigma}$ is a nonsingular quadratic form over $F$, cf. [@BOI §11]. If $\dim_K A=n$, then $\dim T_{\sigma}=n$ if ${\sigma}$ is of the first kind and $\dim T_{\sigma}=2n$ if ${\sigma}$ is of the second kind.
The notion of positive involution seems to have been considered first by Weil in his groundbreaking paper [@Wei].
Procesi and Schacher also define a notion of *positive elements* in $(A,{\sigma})$, cf. [@PS §V]. For greater clarity we have adapted their definitions as follows:
1. An ordering $\leq_P$ of $F$ is called a [*$\sigma$-ordering*]{} if it makes the involution $\sigma$ positive, i.e., if $$0\leq_P \operatorname{Trd}(\sigma(x)x) \quad \text{for all }x\in A.$$
2. Suppose $\leq_P$ is a $\sigma$-ordering on $F$. An element $a\in \operatorname{Sym}(A,{\sigma})$ is called ${\sigma}$-*positive with respect to $\leq_P$* if the quadratic form $ \operatorname{Trd}(\sigma(x)ax)$ is positive semidefinite with respect to $\leq_P$. That is, if $$0\leq_P\operatorname{Trd}(\sigma(x)ax)\quad \text{for all } x\in A.$$
3. An element $a\in \operatorname{Sym}(A,{\sigma})$ is called *totally ${\sigma}$-positive* if it is positive with respect to [all]{} $\sigma$-orderings on $F$.
Elements of the form ${\sigma}(x)x$ with $x\in A$ are called *hermitian squares*. The set of hermitian squares of $A$ is clearly a subset of $\operatorname{Sym}(A,{\sigma})$. It is also clear that the hermitian squares of $K$ are all in $F$.
\[ex2.2\] Sums of hermitian squares and sums of traces of hermitian squares are examples of totally ${\sigma}$-positive elements, as easy verifications will show.
One of the main results in [@PS] explains that these are essentially the only examples. It can be considered as a noncommutative analogue of Artin’s solution to Hilbert’s 17th problem:
[[@PS Theorem 5.4]]{} \[thm:NCh17\] Let $A$ be a central simple algebra with involution ${\sigma}$, centre $K$ and fixed field $F$. Let $\alpha_1,\ldots,\alpha_m\in F$ be elements appearing in a diagonalisation of the quadratic form $\operatorname{Trd}(\sigma(x)x)$. Then for $a\in\operatorname{Sym}(A,\sigma)$ the following statements are equivalent:
1. $a$ is totally ${\sigma}$-positive;
2. there exist $x_{i,{\varepsilon}}\in A$ with $$a=\sum_{{\varepsilon}\in \{0,1\}^m} \alpha^{\varepsilon}\sum_{i} \sigma(x_{i,{\varepsilon}})
x_{i,{\varepsilon}}.$$ $($As usual, $\alpha^{\varepsilon}$ denotes $\alpha_1^{{\varepsilon}_1}\cdots \alpha_m^{{\varepsilon}_m}$.$)$
In the case $n=\deg A=2$, the weights $\alpha_j$ are superfluous (we will come back to this later). Procesi and Schacher [@PS p. 404] conjecture that this is also the case for $n>2$:
\[open:PS\] In a central simple algebra $A$ with involution ${\sigma}$, every totally ${\sigma}$-positive element is a sum of hermitian squares. Equivalently: the trace of a hermitian square is always a sum of hermitian squares.
\[rem2.4\] The two statements in the PS Conjecture are indeed equivalent: the necessary direction follows from the fact that traces of hermitian squares are totally ${\sigma}$-positive, as observed in Example \[ex2.2\].
For the sufficient direction, assume that the trace of a hermitian square is always a sum of hermitian squares. Let $a\in \operatorname{Sym}(A,{\sigma})$ be totally ${\sigma}$-positive. Then $a$ can be expressed in terms of the entries in a diagonalization of the form $\operatorname{Trd}({\sigma}(x)x)$ as in Theorem \[thm:NCh17\](ii). Let $\beta$ be such an entry. Thus, $\beta=\operatorname{Trd}({\sigma}(y)y)$ for some $y\in A$. By the assumption there are $x_1,\ldots, x_\ell\in A$ such that $\beta=\sum_i {\sigma}(x_i)x_i$. Since $\beta\in F$, the expression in Theorem \[thm:NCh17\](ii) can now be rewritten as a sum of hermitian squares.
As mentioned a few lines earlier, Procesi and Schacher provide supporting evidence for their conjecture for the case $\deg A=2$. Another case where the PS Conjecture is true has been well-known since the 1970s:
\[ex:psd\] Let $A$ be the full matrix ring $M_n(F)$ over a formally real field $F$ endowed with the transpose involution ${\sigma}=t$. Since $\operatorname{Trd}=\operatorname{tr}$, every ordering of $F$ is a $\sigma$-ordering. We claim that $a\in\operatorname{Sym}(A,\sigma)$ is totally ${\sigma}$-positive if and only if $a$ is a positive semidefinite matrix in $A\otimes_F R
=M_n(R)$ for any real closed field $R$ containing $F$ (equivalently: for any real closure of $F$).
Indeed, if $a$ is totally ${\sigma}$-positive, then for all $x\in A$, $\operatorname{tr}(x^{t}ax)$ is positive with respect to every ($\sigma$-)ordering of $F$, i.e., $\operatorname{tr}(x^{t}ax)\in{\ensuremath{\sum{F}^{2}}}$. A diagonalisation of the quadratic form $\operatorname{tr}(x^{t}ax)$ will contain only sums of squares in $F$ (as it would otherwise violate the total ${\sigma}$-positivity). Hence this quadratic form remains positive semidefinite under every ordered field extension of $F$.
The converse implication is also easy: if $a$ is positive semidefinite over $M_n(R)$ for every real closed field $R\supseteq F$, then the trace of $x^{t}ax$ for $x\in A$ is nonnegative under the ordering of $R$ and hence under all orderings of $F$. By definition, this means that $a$ is totally ${\sigma}$-positive.
Moreover, every totally ${\sigma}$-positive element of $(A,\sigma)$ is a sum of hermitian squares. Essentially, this goes back to Gondard and Ribenboim [@GR] and has been reproved several times [@Djo; @FRS; @HN; @KS]. It also follows easily from Theorem \[thm:NCh17\] for it suffices to show that the trace of a hermitian square is a sum of hermitian squares. But this is clear: if $a=
\big[
a_{ij}
\big]
_{1\leq i,j\leq n}\in A$, then $$\operatorname{Trd}(\sigma(a)a)=\sum_{i,j=1}^n a_{ij}^2$$ is obviously a sum of (hermitian) squares in $F$.
The reader will have no problems extending this example to the case $K=F(\sqrt{-1})$ and $A=M_n(K)$ endowed with the conjugate transpose involution ${{\overline}{\phantom{x}}}t$.
The Counterexamples {#sec:counter}
===================
When the transpose involution in the previous example is replaced by an arbitrary orthogonal involution ${\sigma}$ on $M_n(F)$ (i.e., an involution which is adjoint to a quadratic form over $F$), the equivalence between totally ${\sigma}$-positive elements and sums of hermitian squares is in general no longer true, as we proceed to show in this section. We assume throughout that $F_0$ is a formally real field.
\[lem3.1\] Let $F=F_0(\!(X)\!)(\!(Y)\!)$, the iterated Laurent series field in two commuting variables $X$ and $Y$. The quadratic form $$q={\langle}X,Y,XY{\rangle}$$ does not weakly represent $1$ over $F$. In fact this is already true over the rational function field $F_0(X,Y)$.
Assume for the sake of contradiction that $m{\times}q$ represents $1$ for some positive integer $m$. Then the form $${\varphi}:={\langle}1{\rangle}\perp m{\times}{\langle}-X,-Y,-XY{\rangle}$$ is isotropic over $F$. This leads to a contradiction by repeated application of Springer’s theorem on fields which are complete with respect to a discrete valuation, cf. [@Lam Chapter VI, §1]. Since $F_0(X,Y)$ embeds into $F$ the proof is finished.
\[thm3.2\] Let $F=F_0(X, Y)$. Let $A=M_3(F)$ and ${\sigma}=\operatorname{ad}_q$, where $$q={\langle}X,Y,XY{\rangle}.$$ The ${\sigma}$-symmetric element $XY$ is totally ${\sigma}$-positive, but is not a sum of hermitian squares in $(A,{\sigma})$.
It is clear that $XY\in\operatorname{Sym}(A,{\sigma})$ since $XY\in F$.
We first show that $XY$ is totally ${\sigma}$-positive. Since $T_{\sigma}\simeq q{\otimes}q$ (see [@Lew p. 227] or [@BOI 11.4]) we have $$\operatorname{sign}_P T_{\sigma}=(\operatorname{sign}_P q)^2 \in \{1,9\}$$ for any ordering $P\in X_F$. (Here $\operatorname{sign}_P T_{\sigma}$ denotes the signature of the quadratic form $T_{\sigma}$ with respect to the ordering $P$.) Hence, the set of ${\sigma}$-orderings on $F$ is not empty. It is exactly the set of $P\in X_F$ with $\operatorname{sign}_P T_{\sigma}=9$. (Note that $F$ has orderings for which both $X$ and $Y$, and thus $XY$, are positive so that the value $\operatorname{sign}_P T_{\sigma}=9$ can indeed be attained.)
Let $P$ be any ${\sigma}$-ordering on $F$. Then we have for any $a\in A$, $$\operatorname{Trd}({\sigma}(a)a)\geq_P 0$$ (by definition) and so for any $a\in A$, $$\operatorname{Trd}\bigl({\sigma}(a) XY a\bigr)=XY \operatorname{Trd}({\sigma}(a)a)\geq_P 0,$$ since $XY\geq_P 0$ (for otherwise $\operatorname{sign}_P T_{\sigma}=1$ and $P$ would not be a ${\sigma}$-ordering on $F$). Hence, $XY$ is totally ${\sigma}$-positive. An alternative argument showing that $XY$ is totally ${\sigma}$-positive can be given by observing that $XY= \operatorname{Trd}({\sigma}(b) b)$ for $$b=\left[\begin{smallmatrix}
0&X&0\\
0&0&0\\
0&0&0
\end{smallmatrix}\right].$$
Next we show that $XY$ is not a sum of hermitian squares in $(A,{\sigma})=(M_3(F), \operatorname{ad}_q)$. We identify $XY$ with $XYI_3$ in $M_3(F)$, where $I_3$ denotes the $3{\times}3$ identity matrix. Assume for the sake of contradiction that $XYI_3$ is a sum of elements of the form ${\sigma}(a)a$ with $a=[a_{ij}]_{1\leq i,j\leq 3} \in M_3(F)$. Recall that $${\sigma}(a)a =\operatorname{ad}_q(a)a=
\left[\begin{smallmatrix}
X & & \\
& Y & \\
& & XY
\end{smallmatrix}\right] \cdot a^t\cdot
\left[\begin{smallmatrix}
X & & \\
& Y & \\
& & XY
\end{smallmatrix}\right]^{-1} \cdot a.$$ The $(3,3)$-entry of ${\sigma}(a)a$ is equal to $$Ya_{13}^2+X a_{23}^2 + a_{33}^2.$$ By our assumption there are $s_1, s_2, s_3 \in \sum F^{{\times}2}$ such that $$XY=Y s_1+ Xs_2 + s_3,$$ which is equivalent with $$1= X^{-1}s_1 + Y^{-1} s_2 + X^{-1} Y^{-1} s_3.$$ Thus, $1$ is weakly represented by the quadratic form $${\langle}X^{-1}, Y^{-1}, X^{-1} Y^{-1}{\rangle}\simeq {\langle}X, Y, XY{\rangle}=q,$$ which is impossible by Lemma \[lem3.1\]. This finishes the proof.
The previous theorem gives us a counterexample to the PS Conjecture. It shows that the conjecture is in general not true for full matrix algebras equipped with an orthogonal involution. In contrast, when we equip a full matrix algebra with a *symplectic* involution, we will show in Theorem \[thm4.7\] below that the conjecture does hold.
Thus, we could ask if the PS Conjecture also holds for non-split central simple algebras with symplectic involution. The answer is “no”:
\[thm3.3\] Let $F=F_0(X, Y)$. Let $A=M_3(F){\otimes}_F H\cong M_3(H)$, where $H=(-1,-1)_F$ is Hamilton’s quaternion division algebra over $F$. Equip $A$ with the involution ${\sigma}=\operatorname{ad}_q{\otimes}\gamma$, where $\gamma$ is quaternion conjugation and ${\sigma}=\operatorname{ad}_q$ for $$q={\langle}X,Y,XY{\rangle}.$$ The algebra $A$ is central simple over $F$ of degree $6$ and the involution ${\sigma}$ is symplectic. The ${\sigma}$-symmetric element $XY$ is totally ${\sigma}$-positive, but is not a sum of hermitian squares in $(A,{\sigma})$.
The assertion about $(A,{\sigma})$ is clear, as is the fact that $XY\in\operatorname{Sym}(A,{\sigma})$ since $XY\in F$.
It is easy to verify that the involution trace form of $\gamma$, $T_\gamma$, is isometric to ${\langle}2{\rangle}{\otimes}N_H$, where $N_H={\langle}1,1,1,1{\rangle}$ is the norm form of $H$. Here $N_H(x):=\operatorname{Nrd}_H(x)$ for all $x\in H$, where $\operatorname{Nrd}_H$ denotes the reduced norm on $H$. Since $T_{\sigma}=T_{\operatorname{ad}_q{\otimes}\gamma}\simeq T_{\operatorname{ad}_q}{\otimes}T_\gamma$, we have $$\operatorname{sign}_P T_{\sigma}= (\operatorname{sign}_P T_{\operatorname{ad}_q})(\operatorname{sign}_P T_\gamma) = 4\operatorname{sign}_P T_{\operatorname{ad}_q}\in \{4, 36\}$$ for any ordering $P\in X_F$. Hence, the set of ${\sigma}$-orderings on $F$ is not empty. It is exactly the set of $P\in X_F$ with $\operatorname{sign}_P T_{\sigma}=36$. (Note again that this value can indeed be attained since there are orderings on $F$ for which both $X$ and $Y$, and thus $XY$, are positive.) Arguing similarly as in the proof of Theorem \[thm3.2\] we can verify that $XY$ is totally ${\sigma}$-positive.
Before proceeding, note that the involution $\gamma$ is adjoint to the hermitian form ${\langle}1{\rangle}_\gamma$ over $(H,\gamma)$. Hence, ${\sigma}$ is adjoint to the hermitian form $h=q{\otimes}{\langle}1{\rangle}_\gamma ={\langle}X,Y,XY{\rangle}_\gamma$ over $(H,\gamma)$. Thus $$h(x,y)= \gamma(x_1) X y_1 +\gamma(x_2) Y y_2 + \gamma(x_3)XY y_3$$ for vectors $x=(x_1,x_2,x_3)$ and $y=(y_1,y_2,y_3)$ in the right $H$-vector space $H^3$.
Next we show that $XY$ is not a sum of hermitian squares in $(A,{\sigma})=(M_3(H), \operatorname{ad}_h)$. We identify $XY$ with $XYI_3$ in $M_3(H)$, where $I_3$ denotes the $3{\times}3$ identity matrix. Assume for the sake of contradiction that $XYI_3$ is a sum of elements of the form ${\sigma}(a)a$ with $a=[a_{ij}]_{1\leq i,j\leq 3} \in M_3(H)$. Recall that $${\sigma}(a)a =\operatorname{ad}_h(a)a=
\left[\begin{smallmatrix}
X & & \\
& Y & \\
& & XY
\end{smallmatrix}\right] \cdot \gamma(a)^t\cdot
\left[\begin{smallmatrix}
X & & \\
& Y & \\
& & XY
\end{smallmatrix}\right]^{-1}\cdot a,$$ where $\gamma(a)=\big[\gamma(a_{ij})\big]_{1\leq i,j\leq 3}$. The $(3,3)$-entry of ${\sigma}(a)a$ is equal to $$\gamma(a_{13})Ya_{13}+\gamma(a_{23})X a_{23} +\gamma(a_{33}) a_{33}=YN_H(a_{13})+XN_H(a_{23})+N_H(a_{33}).$$ Since $N_H={\langle}1,1,1,1{\rangle}$, each of $N_H(a_{13})$, $N_H(a_{23})$, $N_H(a_{33})$ is a sum of four squares in $F$. Thus, by our assumption there are $s_1, s_2, s_3 \in \sum F^{{\times}2}$ such that $$XY=Y s_1+ Xs_2 + s_3.$$ We can now finish the proof with an appeal to Lemma \[lem3.1\], as in the proof of Theorem \[thm3.2\].
By tensoring $(M_3(F), \operatorname{ad}_q)$ with Hamilton’s quaternion division algebra, equipped with a *unitary* involution one obtains a counterexample in the non-split unitary case. We leave the details, which are similar to those in the proof of Theorem \[thm3.3\], to the diligent reader.
From a real algebra perspective it is clear that these counterexamples to the PS Conjecture can easily be seen to work over any formally real field $F$ that admits a proper semiordering (see [@PD §5] for details and unexplained terminology). Given such a field $F$, endowed with a proper semiordering, take negative $a,b\in F$ such that $ab$ is negative as well. Then $q=\langle a,b,ab\rangle$ does not weakly represent $1$ (the quadratic module generated by $\{-a,-b,-ab\}$ is proper) and thus in $M_3(F)$, endowed with the involution $\sigma=\operatorname{ad}_q$, the element $ab$ is totally $\sigma$-positive, but not a sum of hermitian squares (as the proof of Theorem \[thm3.2\] shows).
Positive Results {#para.pos.rel}
================
Procesi and Schacher [@PS p. 404 and 405] prove their conjecture for central simple algebras $A$ of degree two, i.e., quaternion algebras, with arbitrary involution ${\sigma}$ by appealing to matrices and the Cayley–Hamilton theorem. We start this section by giving an alternative argument motivating some of the generalizations that follow.
Throughout this section we assume that the base field $F$ is formally real.
\[prop4.1\] Let $A$ be a quaternion algebra not necessarily division with centre $K$, equipped with an arbitrary involution ${\sigma}$. Let $F$ be the fixed field of $(A,{\sigma})$. Each entry occurring in a diagonalisation of $T_{\sigma}$ is a sum of hermitian squares.
\(i) We first consider involutions of the first kind on $A$. Let $A$ be the quaternion algebra $(a,b)_F$ with $F$-basis $\{1,i,j,k\}$ where $i$, $j$ and $k$ anti-commute, $ij=k$, $i^2=a$ and $j^2=b$.
If ${\sigma}$ is symplectic, then ${\sigma}$ is the unique quaternion conjugation involution $\gamma$ on $A$. An easy computation gives $T_{\sigma}=T_\gamma\simeq {\langle}2{\rangle}{\otimes}{\langle}1,-a,-b,ab{\rangle}$. We have $$1=\gamma(1)1,\ -a=\gamma(i)i,\ -b=\gamma(j)j,\ ab=\gamma(k)k.$$
If ${\sigma}$ is orthogonal, then ${\sigma}=\operatorname{Int}(u)\circ\gamma$, where $u\in A$ satisfies $\gamma(u)=-u$. From [@BOI 11.6] we know that $$T_{\sigma}\simeq {\langle}2{\rangle}{\otimes}{\langle}1, \operatorname{Nrd}_A(u), -\operatorname{Nrd}_A(s), -\operatorname{Nrd}_A(su){\rangle}$$ for some $s\in A$ with ${\sigma}(s)=s=-\gamma(s)$. Now, $$\begin{aligned}
\operatorname{Nrd}_A(u)&=u\gamma(u) =u\gamma(u) u^{-1}u={\sigma}(u)u;\\
-\operatorname{Nrd}_A(s)&=-\gamma(s)s={\sigma}(s)s;\\
-\operatorname{Nrd}_A(su)&=-\operatorname{Nrd}_A(s)\operatorname{Nrd}_A(u)=-\gamma(s)s\operatorname{Nrd}_A(u)={\sigma}(s){\sigma}(u)us={\sigma}(us)us.\end{aligned}$$
\(ii) Finally, let $K=F(\sqrt{\delta})$ and let $A$ be a quaternion algebra over $K$ with unitary involution ${\sigma}$ whose restriction to $K$ is $\tau$, where $\tau$ is determined by $\tau(\sqrt{\delta})=-\sqrt{\delta}$. By a well-known result of Albert [@BOI 2.22] there exists a unique quaternion $F$-subalgebra $A_0\subseteq A$ such that $$A=A_0{\otimes}_F K \text{ and } \sigma= \gamma_0{\otimes}\tau,$$ where $\gamma_0$ is quaternion conjugation on $A_0$. Then $T_{\sigma}\simeq T_{\gamma_0}{\otimes}T_\tau\simeq T_{\gamma_0}{\otimes}{\langle}1,-\delta{\rangle}$. Since $\tau(\sqrt{\delta})\sqrt{\delta}=-\delta$, we are finished by the symplectic part of the proof.
This shows in particular that the PS Conjecture is true for full matrix algebras of degree two over a formally real field $F$ since these are just split quaternion algebras.
Part (ii) of the proof of Proposition \[prop4.1\] motivates the following more general result:
\[thm4.2\] Let $A$ and $B$ be central simple algebras with centre $K$, equipped with arbitrary involutions ${\sigma}$ and $\tau$, respectively. Assume that $(A,{\sigma})$ and $(B,\tau)$ have the same fixed field $F$. If the PS Conjecture holds for $(A,{\sigma})$ and $(B,\tau)$, it also holds for the tensor product $(A{\otimes}_K B, {\sigma}{\otimes}\tau)$.
This is a simple computation, using the fact that $T_{{\sigma}{\otimes}\tau}\simeq T_{\sigma}{\otimes}T_\tau$ and that elements of $A$ commute with elements of $B$ in the tensor product $A{\otimes}_K B$.
\[cor4.3\] Let $(Q_1,{\sigma}_1),\ldots, (Q_\ell,{\sigma}_\ell)$ be quaternion algebras with arbitrary involution over $K$ and with common fixed field $F$. The PS Conjecture holds for the tensor product $\bigotimes_{i=1}^\ell (Q_i,{\sigma}_i)$.
This is an immediate consequence of Proposition \[prop4.1\] and Theorem \[thm4.2\].
Let $A=M_n(F)$ be a split algebra of $2$-power degree $n=2^\ell$, equipped with an orthogonal involution ${\sigma}$ which is adjoint to an $n$-fold Pfister form over $F$. The PS Conjecture holds for $(A,{\sigma})$.
By Becher’s proof of the Pfister Factor Conjecture [@Bec], $(A,{\sigma})$ decomposes as $$(A,{\sigma})\cong \bigotimes_{i=1}^\ell (Q_i,{\sigma}_i),$$ where $(Q_1,{\sigma}_1),\ldots, (Q_\ell,{\sigma}_\ell)$ are quaternion algebras with involution. An appeal to Corollary \[cor4.3\] finishes the proof.
\[cor4.4\] Let $A=M_n(K)$ be a split algebra of $2$-power degree $n=2^\ell$, equipped with a hyperbolic involution ${\sigma}$ of any kind. Let $F$ be the fixed field of $(A,{\sigma})$. The PS Conjecture holds for $(A,{\sigma})$.
Recall from [@BST Theorem 2.1] that the involution ${\sigma}$ is hyperbolic if there exists an idempotent $e\in A$ such that ${\sigma}(e)=1-e$ or, equivalently, if the adjoint (quadratic, alternating or hermitian) form of ${\sigma}$ is hyperbolic.
If $\ell=1$ this is just the split version of Proposition \[prop4.1\]. Assume now that $\ell\geq 2$. By [@BST Theorem 2.2], $(A,{\sigma})$ decomposes as $$(A,{\sigma})\cong \bigotimes_{i=1}^\ell (Q,{\sigma}_i),$$ where $Q=M_2(K)$ and ${\sigma}_1,\ldots, {\sigma}_\ell$ are involutions on $Q$. An appeal to Corollary \[cor4.3\] finishes the proof.
Let $A=M_n(F)$ be a split algebra of $2$-power degree $n=2^\ell$, equipped with a symplectic involution ${\sigma}$. The PS Conjecture holds for $(A,{\sigma})$.
If ${\sigma}$ is a symplectic involution, it is hyperbolic (since it is adjoint to an alternating form over $F$ which is automatically hyperbolic) and we are finished by Corollary \[cor4.4\].
In fact, the PS Conjecture is true for *any* split algebra with symplectic involution. Such an algebra is always of even degree.
\[thm4.7\] Let $A=M_n(F)$ be a split algebra of even degree $n=2m$, equipped with a symplectic involution ${\sigma}$. The PS Conjecture holds for $(A,{\sigma})$.
Since ${\sigma}$ is symplectic, the quadratic form $T_{\sigma}$ is hyperbolic (see [@Lew p. 227] or [@BOI Proof of 11.7]). Thus $T_{\sigma}\simeq m{\times}{\langle}1,-1{\rangle}$ and it suffices to show that $-1$ is a sum of hermitian squares in $A$. We identify $-1$ with $-I_n$, where $I_n$ denotes the $n{\times}n$ identity matrix in $A=M_n(F)$.
Since ${\sigma}$ is symplectic, we have ${\sigma}=\operatorname{Int}(S)\circ t$, where $t$ denotes transposition and $S\in {\mathrm{GL}}_n(F)$ satisfies $S^t=-S$. Since $S$ is skew-symmetric, there exists a matrix $P\in {\mathrm{GL}}_n(F)$ such that $P^tSP=B$, where $B$ is the block diagonal matrix with $m$ blocks $\left[\begin{smallmatrix}
0&1\\
-1 & 0
\end{smallmatrix}\right]$ on the diagonal.
Let $X$ be the block diagonal matrix with $m$ blocks $\left[\begin{smallmatrix}
0&1\\
1 & 0
\end{smallmatrix}\right]$ on the diagonal. Then $X^tBX=B^{-1}$. Hence with $Y=PXP^t$, we have $Y^tSY=S^{-1}$. Thus $${\sigma}(SY)SY=S (SY)^t S^{-1}SY= SY^tS^tY=SY^t(-S)Y=-SS^{-1}=-I_n.
\qedhere$$
Positive noncommutative polynomials {#sec5}
===================================
Algebras of generic matrices with involution {#sec:generic}
--------------------------------------------
After studying the PS Conjecture in the setting of central simple algebras with involution, we proceed to interpret these results as well as Theorem \[thm:NCh17\] for non-dimensionfree positivity of noncommutative (NC) polynomials.
Motivated by problems in optimization and control theory, Helton [@Hel] proved that a symmetric real or complex NC polynomial, all of whose images under algebra $*$-homomorphisms into $M_n({\mathbb{R}})$, $n\in{\mathbb{N}}$, are positive semidefinite (i.e., a dimensionfree positive NC polynomial), is a sum of hermitian squares. What we are interested in, is positivity under evaluations in $M_n({\mathbb{R}})$ for a *fixed* $n$.
To tackle this problem we introduce the language of generic matrices [@Row]. Verifying a condition on evaluations of an NC polynomial in the algebra of $n\times n$ matrices is often conveniently done in the algebra of generic matrices. In this subsection we recall the definition of generic matrices with involution, while our main result on positive NC polynomials (i.e., a Positivstellensatz) is presented in the next subsection.
As in the classical construction of the algebra of generic matrices (see e.g. [@Row §1.3]), it is possible to construct the algebra of generic matrices *with involution* [@PS §II]. To each type of involution (orthogonal, symplectic and unitary) an algebra of generic matrices with involution can be associated, as we now explain. We assume from now on that $K$ is a field of characteristic $0$ with involution $*$ and fixed field $F$.
Let $K{\langle{\overline}X,{\overline}X^*\rangle}$ be the free algebra with involution over $(K,*)$, i.e., the algebra with involution, freely generated by the noncommuting variables ${\overline}X:=
(X_1,X_2,\ldots )$. Its elements (called *NC polynomials*) are (finite) linear combinations of words in (the infinitely many) letters ${\overline}X, {\overline}X^*$.
Fix a type J $\in\{$orthogonal, symplectic, unitary$\}$. Let ${\mathfrak{a}}_{\operatorname{J}_n}\subseteq K{\langle{\overline}X,{\overline}X^*\rangle}$ denote the ideal of all identities satisfied by degree $n$ central simple $K$-algebras with type $\operatorname{J}$ involution. That is, $f=f(X_1,\ldots,X_k,X_1^*,\ldots
,X_k^*)\in K{\langle{\overline}X,{\overline}X^*\rangle}$ is an element of ${\mathfrak{a}}_{\operatorname{J}_n}$ if and only if for every central simple algebra $A$ of degree $n$ with type $\operatorname{J}$ involution $\sigma$ and every $a_1,\ldots,a_k\in A$, $$f(a_1,\ldots,a_k,\sigma(a_1),\ldots,\sigma(a_k))=0.$$ Then $\operatorname{GM}_n(K,{\operatorname{J}}):=
K{\langle{\overline}X,{\overline}X^*\rangle}/{\mathfrak{a}}_{\operatorname{J}_n}$ is the *algebra of generic $n\times n$ matrices with type $\operatorname{J}$ involution*.
An alternative description of the algebra of generic matrices with involution can be obtained as follows. Let $\zeta:=(\zeta_{ij}^{(\ell)}\mid1\leq i, j\leq n,\,\ell\in{\mathbb{N}})$ denote commuting variables and form the polynomial algebra $K[\zeta]$ endowed with the involution extending $*$ and fixing $\zeta_{ij}^{(\ell)}$ pointwise. Consider the $n\times n$ matrices $Y_{\ell}:=\big[
\zeta_{ij}^{(\ell)}\big]_{1\leq i,j\leq n}\in M_n(K[\zeta])$, $\ell\in{\mathbb{N}}$. Each $Y_{\ell}$ is called a *generic matrix*.
1. If $\operatorname{J}\in\{$orthogonal, unitary$\}$, then the (unital) $K$-subalgebra of $M_n(K[\zeta])$ generated by the $Y_\ell$ and their transposes is (canonically) isomorphic to $\operatorname{GM}_n(K,{\rm J})$.
2. If $\operatorname{J}=$ symplectic, then $n$ is even, say $n=2m$. Consider the usual symplectic involution $$\begin{bmatrix}
x&y\\
z&w
\end{bmatrix}
\mapsto
\begin{bmatrix}
w^t&-y^t\\
-z^t&x^t
\end{bmatrix}$$ on $M_{2m}(K[\zeta])$. Then the (unital) $K$-subalgebra of $M_n(K[\zeta])$ generated by the $Y_\ell$ and their images under this involution is (canonically) isomorphic to $\operatorname{GM}_n(K,{\operatorname{J}})$.
If $n=1$, then $\operatorname{J}\in\{$orthogonal, unitary$\}$ and $\operatorname{GM}_1(K,\operatorname{J})$ is isomorphic to $K[\zeta]$ endowed with the involution introduced above. Hence in the sequel we will always assume $n\geq 2$.
Let $\operatorname{J}\in\{$orthogonal, symplectic, unitary$\}$. For $n\geq 2$, $\operatorname{GM}_n(K,\operatorname{J})$ is a PI algebra and a domain (cf. [@PS §II]). Hence its central localization is a division algebra $\operatorname{UD}_n(K,\operatorname{J})$ with involution, which we call the *generic division algebra* with type $\operatorname{J}$ involution of degree $n$. As we will only consider the canonical involution on $\operatorname{GM}_n(K,\operatorname{J})$ and $\operatorname{UD}_n(K,\operatorname{J})$ we use $*$ to denote it.
A Positivstellensatz {#sec6}
--------------------
Let ${\mathbb{K}}\in\{{\mathbb{R}},{\mathbb{C}}\}$ be endowed with the complex conjugation involution ${{\overline}{\phantom{x}}}$. Our aim in this subsection is to deduce a non-dimensionfree version of Helton’s sum of hermitian squares theorem. We will describe symmetric NC polynomials $f$ all of whose evaluations in $M_n({\mathbb{K}})$ are positive semidefinite, see Theorem \[thm:positiv\].
We start with a lemma characterizing total $*$-positivity in the algebra of generic matrices $\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$. The proof of the following proposition uses some elementary model theory, e.g. Tarski’s transfer principle for real closed fields. All the necessary background can be found in [@PD §1 and §2] or, alternatively, [@BCR §1].
\[lem:uf\] Let $n\in {\mathbb{N}}$. If ${\mathbb{K}}={\mathbb{R}}$, let $\operatorname{J}=$ orthogonal and if ${\mathbb{K}}={\mathbb{C}}$, let $\operatorname{J}=$ unitary. If $a=a^*\in\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$ is totally $\sigma$-positive under each $*$-homomorphism from $\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$ to $M_n({\mathbb{K}})$ endowed with a positive type $\operatorname{J}$ involution $\sigma$, then $a$ is totally $*$-positive $($in $\operatorname{UD}_n({\mathbb{K}},\operatorname{J}))$.
Suppose $a\in \operatorname{GM}_n({\mathbb{K}},\operatorname{J})$ is not totally $*$-positive. Then there is a $*$-ordering $\leq$ of the fixed field $Z$ of the centre of $\operatorname{UD}_n({\mathbb{K}},\operatorname{J}))$ under which $\operatorname{Trd}(x^*ax)$ is not positive semidefinite. Let $\langle \alpha_1,\ldots, \alpha_{m}\rangle$ be the diagonalisation of $\operatorname{Trd}(x^*x)$ with $\alpha_i=\alpha_i^* \in Z$. (Here $m=n^2$ if the involution is of the first kind and $m=2n^2$ otherwise.) Given that $Z$ is the field of fractions of the symmetric centre $Z_0$ of $\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$, we may even assume $\alpha_i\in
Z_0$. We also diagonalise $\operatorname{Trd}(x^*ax)$ as $\langle \beta_1,\ldots, \beta_{m}\rangle$ with $\beta_i\in Z_0$. Clearly, $\alpha_i>0$ and one of the $\beta_i$, say $\beta_1$, is negative with respect to the given $*$-ordering $\leq$. Let ${\overline}Z^{ {\rm rc} }$ denote the real closure of $Z$ with respect to this ordering and form $A:=\operatorname{UD}_n({\mathbb{K}},\operatorname{J})\otimes_Z {\overline}Z^{ {\rm rc} }$ endowed with the involution $\sigma=*\otimes {\rm id}$. Then $A$ is a central simple algebra over a real closed (if $\operatorname{J}=$ orthogonal) or algebraically closed field (if $\operatorname{J}=$ unitary). Moreover, its involution $\sigma$ is positive. Hence by the classification result [@PS Theorem 1.2] of Procesi and Schacher, $A$ is either $M_n({\overline}Z^{ {\rm rc} })$ endowed with the transpose (if $\operatorname{J}=$ orthogonal) or $M_n({\overline}Z)$ endowed with the complex conjugate transpose involution (if $\operatorname{J}=$ unitary). Here ${\overline}Z$ is the algebraic closure ${\overline}Z^{ {\rm rc} }(\sqrt{-1})$ of ${\overline}Z^{ {\rm rc} }$ and the complex conjugate maps $r+t \sqrt{-1}
\mapsto r- t \sqrt{-1}$ for $r,t\in{\overline}Z^{ {\rm rc} }$.
For $b\in\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$ let $\hat b\in {\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}$ denote a preimage of $b$ under the canonical map ${\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}\to\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$. Every $*$-homomorphism $\operatorname{GM}_n({\mathbb{K}},\operatorname{J})\to M_n(L)$ for a $*$-field extension $L$ of ${\mathbb{K}}$, where $M_n(L)$ is given a type $\operatorname{J}$ involution, yields a $*$-homomorphism ${\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}\to M_n(L)$, so is essentially given by a point $s\in M_n(L)^{\mathbb{N}}$ describing the images of the $X_i$ under this induces map.
By construction, the image $\beta_1\otimes 1$ of $\beta_1$ under the embedding of algebras with involution $\operatorname{GM}_n({\mathbb{K}},\operatorname{J})\to A$ is not $\sigma$-positive. Let $s$ denote the corresponding evaluation point. By Example \[ex:psd\], this means that $
\hat \beta_1(s,{\overline}{s}^t)=\beta_1\otimes 1$ is not positive semidefinite. Consider the following elementary statement: $$\label{eq:1st}
\begin{split}
\exists \,n\times n \text{ matrices }x=(x_1,\ldots, x_N) : \; &
\hat\alpha_i(x,{\overline}{x}^t) \text{ is positive semidefinite } \land \\
& \hat\beta_1(x,{\overline}{x}^t) \text{ is not positive semidefinite}.
\end{split}$$ ($N$ is the maximal number of variables appearing in one of the $\hat\alpha_i, \hat\beta_1$.)
Obviously such $n\times n$ matrices $x_i$ can be found over ${\overline}Z^{ {\rm rc} }$ or ${\overline}Z$; just take $x_i=s_i$. By Tarski’s transfer principle, the above elementary statement can be satisfied in ${\mathbb{K}}$. This yields a $*$-homomorphism ${\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}\to M_n({\mathbb{K}})$ endowed with the (positive) involution ${{\overline}{\phantom{x}}}t$ and in turn (by universality) a $*$-homomorphism $\operatorname{GM}_n({\mathbb{K}},\operatorname{J})\to (M_n({\mathbb{K}}),{{\overline}{\phantom{x}}}t)$. By the construction, the image of $a$ under this mapping will not be positive semidefinite. This finishes the proof.
In order to state the Positivstellensatz, we need to recall the notion of [*central polynomials*]{} for $n\times n$ matrices. These are $f\in K{\langle{\overline}X,{\overline}X^*\rangle}$ whose image in $\operatorname{GM}_n(K,\operatorname{J})$ is central. Equivalently, the image of $f$ under a $*$-homomorphism from $K{\langle{\overline}X,{\overline}X^*\rangle}$ to $M_n(K)$ endowed with a type $\operatorname{J}$ involution, is always a scalar matrix. If it is nonzero, we call $f$ [*nonvanishing*]{}. The existence of nonvanishing central polynomials is nontrivial; we refer to [@Row §1; Appendix A] for details.
\[thm:positiv\] Suppose ${\mathbb{K}}\in\{{\mathbb{R}},{\mathbb{C}}\}$ is endowed with the complex conjugate involution ${{\overline}{\phantom{x}}}$. Let $g=g^*\in {\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}$, $n\in {\mathbb{N}}$ and fix a type $\operatorname{J}\in\{$orthogonal, unitary$\}$ according to the type of involution on ${\mathbb{K}}$. Choose $\alpha_1,\ldots,\alpha_m\in {\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}$ whose images in $\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$ form a diagonalisation of the quadratic form $\operatorname{Trd}(x^*x)$ on $\operatorname{UD}_n({\mathbb{K}},\operatorname{J})$. Then the following are equivalent:
1. for any $s\in M_n({\mathbb{K}})^{\mathbb{N}}$, $g(s,{\overline}{s}^t)$ is positive semidefinite;
2. there exists a nonvanishing central polynomial $h\in {\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}$ for $n\times n$ matrices and $p_{i,{\varepsilon}}\in {\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}$ with $$h^* g h \equiv
\sum_{{\varepsilon}\in \{0,1\}^m} \alpha^{\varepsilon}\sum_{i} p_{i,{\varepsilon}}^*
p_{i,{\varepsilon}} \pmod {{\mathfrak{a}}_{\operatorname{J}_n}}.$$
Given a congruence as in (ii), it is clear that (i) holds whenever $h(s,{\overline}{s}^t)\neq 0$. As the set of all such $s$ is Zariski dense, (i) holds for all $s\in M_n({\mathbb{K}})^{\mathbb{N}}$.
For the converse implication note that by Lemma \[lem:uf\], $g+{\mathfrak{a}}_{\operatorname{J}_n}$ is totally $*$-positive in $\operatorname{UD}_n({\mathbb{K}},\operatorname{J})$. Hence by Theorem \[thm:NCh17\] we obtain a positivity certificate $$g+{\mathfrak{a}}_{\operatorname{J}_n} =\sum_{{\varepsilon}\in \{0,1\}^m} (\alpha+{\mathfrak{a}}_{\operatorname{J}_n})^{\varepsilon}\sum_{i} (x'_{i,{\varepsilon}})^*
x'_{i,{\varepsilon}}$$ for some $x'_{i,{\varepsilon}}\in\operatorname{UD}_n({\mathbb{K}},\operatorname{J})$. Clearing denominators, there are $x_{i,{\varepsilon}}\in\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$ and a nonzero central $r\in\operatorname{GM}_n({\mathbb{K}},\operatorname{J})$ with $$r^*(g+{\mathfrak{a}}_{\operatorname{J}_n})r= \sum_{{\varepsilon}\in \{0,1\}^m} (\alpha+{\mathfrak{a}}_{\operatorname{J}_n})^{\varepsilon}\sum_{i} x_{i,{\varepsilon}}^*
x_{i,{\varepsilon}}.$$ Lifting this equality to the free algebra yields the desired conclusion.
When $n=2$, the weights $\alpha$ are redundant (cf. §\[para.pos.rel\] or [@PS p. 405]) and we obtain the following strengthening:
\[cor:positiv2\] Suppose ${\mathbb{K}}\in\{{\mathbb{R}},{\mathbb{C}}\}$ is endowed with the complex conjugate involution ${{\overline}{\phantom{x}}}$. Let $g=g^*\in {\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}$, $n\in {\mathbb{N}}$ and fix a type $\operatorname{J}\in\{$orthogonal, unitary$\}$ according to the type of involution on ${\mathbb{K}}$. Then the following are equivalent:
1. for any $s\in M_2({\mathbb{K}})^{\mathbb{N}}$, $g(s,{\overline}{s}^t)$ is positive semidefinite;
2. there exists a nonvanishing central polynomial $h\in {\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}$ for $2\times 2$ matrices and $p_{i}\in {\mathbb{K}}{\langle{\overline}X,{\overline}X^*\rangle}$ with $$h^* g h \equiv
\sum_{i} p_{i}^*
p_{i} \pmod {{\mathfrak{a}}_{\operatorname{J}_2}}.$$
By Tarski’s transfer principle, Theorem \[thm:positiv\] and Corollary \[cor:positiv2\] hold with ${\mathbb{K}}$ replaced by any real closed or algebraically closed field of characteristic $0$.
We conclude the paper with an open problem: can Theorem \[thm:positiv\] be used to give a proof of Helton’s sum of hermitian squares theorem?
Acknowledgments {#acknowledgments .unnumbered}
===============
Both authors wish to thank David Lewis for some very useful observations. The second author wishes to thank Jaka Cimprič for arranging a very nice stay at the University of Ljubljana in May 2008. The first author was supported by the Slovenian Research Agency (project no. Z1-9570-0101-06). The second author was supported by the Science Foundation Ireland Research Frontiers Programme (project no. 07/RFP/MATF191).
[dOHMP]{}
E. Bayer-Fluckiger, D.B. Shapiro, J.-P. Tignol, Hyperbolic involutions, *Math. Z.* **214** (1993), no. 3, 461–476.
K.J. Becher, A proof of the Pfister factor conjecture, *Invent. Math.* **173** (2008), no. 1, 1–6.
J. Bochnak, M. Coste, M.-F. Roy, *Real algebraic geometry*, Springer-Verlag, Berlin, 1998.
M.C. de Olivera, J.W. Helton, S.A. McCullough, M. Putinar, Engineering Systems and Free Semi-Algebraic Geometry, *Emerging Applications of Algebraic Geometry*, 17–62, IMA Vol. Math. Appl. **149**, Springer, 2008.
D.Ž. Djoković, Positive semi-definite matrices as sums of squares, *Linear Algebra and Appl.* [**14**]{} (1976), no. 1, 37–40.
J.F. Fernando, J.M. Ruiz, C. Scheiderer, Sums of squares of linear forms, *Math. Res. Lett.* [**13**]{} (2006), no. 5-6, 947–956.
D. Gondard, P. Ribenboim, Le 17e problème de Hilbert pour les matrices, *Bull. Sci. Math.* (2) [**98**]{} (1974), no. 1, 49–56.
J.W. Helton, “Positive” noncommutative polynomials are sums of squares, *Ann. of Math.* (2) [**156**]{} (2002), no. 2, 675–694.
C.J. Hillar, J. Nie, An elementary and constructive solution to Hilbert’s 17th problem for matrices, *Proc. Amer. Math. Soc.* [**136**]{} (2008), no. 1, 73–76.
I. Klep, M. Schweighofer, Positivity certificates for matrix polynomials, in preparation.
M.-A. Knus, A.S. Merkurjev, M. Rost, J.-P. Tignol, *The Book of Involutions*, Coll. Pub. [**44**]{}, Amer. Math. Soc., Providence, RI (1998).
T.Y. Lam, *Introduction to quadratic forms over fields*, Graduate Studies in Mathematics [**67**]{}, American Mathematical Society, Providence, RI (2005).
D.W. Lewis, Trace forms, Kronecker sums, and the shuffle matrix, *Linear and Multilinear Algebra* **40** (1996), no. 3, 221–227.
A. Prestel, C.N. Delzell, *Positive polynomials. From Hilbert’s 17th problem to real algebra*, Springer-Verlag, Berlin, 2001.
C. Procesi, M. Schacher, A non-commutative real Nullstellensatz and Hilbert’s 17th problem, *Ann. of Math.* (2) [**104**]{} (1976), no. 3, 395–406.
L.H. Rowen, *Polynomial identities in ring theory*, Pure and Applied Mathematics, [**84**]{}, Academic Press, Inc., New York-London (1980).
A. Weil, Algebras with involutions and the classical groups, *J. Indian Math. Soc. (N.S.)* **24** 1960 589–623 (1961).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'It is shown that fully-parallel encoding and decoding schemes with asymptotic block error probability that scales as $O\left(f\left(n\right)\right)$ have Thompson energy that scales as $\Omega\left(\sqrt{\ln f\left(n\right)}n\right)$. As well, it is shown that the number of clock cycles (denoted $T\left(n\right)$) required for any encoding or decoding scheme that reaches this bound must scale as $T\left(n\right)\ge\sqrt{\ln f\left(n\right)}$. Similar scaling results are extended to serialized computation. The Grover information-friction energy model is generalized to three dimensions and the optimal energy of encoding or decoding schemes with probability of block error $P_\mathrm{e}$ is shown to be at least $\Omega\left(n\left(\ln P_{\mathrm{e}}\left(n\right)\right)^{\frac{1}{3}}\right)$.'
author:
- |
Christopher G. Blake and Frank R. Kschischang\
Department of Electrical & Computer Engineering\
University of Toronto\
`[email protected]` `[email protected]`
bibliography:
- 'bibtextDoc.bib'
title: 'Energy, Latency, and Reliability Tradeoffs in Coding Circuits[^1]'
---
Introduction\[sec:Introduction\]
================================
on work started in [@ElGamal] and more recently advanced in [@groverFundamental; @GroverInfoFriction; @blakeConstantPePaper], we borrow a computational complexity model introduced in [@Thompson] that allows us to model the energy and number of clock cycles of a computation. We consider fundamental tradeoffs between the asymptotic energy, number of clock cycles, and block error probability for sequences of good encoders and decoders.
An $f\left(n\right)$*-coding scheme* is a sequence of codes of increasing block length $n$, together with a sequence of encoders and decoders, in which the block error probability associated with the code of block length $n$ is less than $f\left(n\right)$ for sufficiently large $n$.
We show, in terms of $T\left(n\right)$ (the number of clock cycles of the encoder or decoder for the code with block length $n$) that an $f(n)$-coding scheme that is fully parallel has encoding and decoding energy ($E$) that scales as $E\ge\Omega\left(\frac{n\ln f\left(n\right)}{T(n)}\right)$. We show that the energy optimal number of clock cycles for encoders and decoder ($T\left(n\right)$) for an $f\left(n\right)$-coding scheme scales as $O\left(\sqrt{\ln f\left(n\right)}\right)$, giving a universal energy lower bound of $\Omega\left(\sqrt{\ln f\left(n\right)}n\right)$. A special case of our result is that exponentially low probability of error coding schemes thus have encoding and decoding energy that scales at least as $\Omega\left(n^{\frac{3}{2}}\right)$ with energy-optimal number of clock cycles that scales as $\Omega\left(n^{\frac{1}{2}}\right)$. This approach is generalized to serial implementations.
Recent work on the energy complexity of good decoding has focused largely on planar circuits. However, circuits implemented in three-dimensions exist [@3DCircuitBook], and so we generalize the recent information friction (or bit-meters) model introduced by Grover in [@GroverInfoFriction] to circuits implemented in three-dimensions and extend the technique of Grover to show that, in terms of block length $n$, a bit-meters coding scheme in which block error probability is given by $P_{\mathrm{e}}(n)$ has encoding/decoding energy that scales as $\Omega\left(n\left(\ln P_{\mathrm{e}}\left(n\right)\right)^{\frac{1}{3}}\right)$. We show how this approach can be generalized to an arbitrary number of dimensions.
In Section \[sec:Computational-Complexity-Lower\] we discuss prior work, and in particular we discuss existing results on complexity lower bounds for different models of computation for different notions of “good” encoders and decoders. The main technical results of this work are in Section \[sec:ThompsonModelSection\], where we study the Thompson energy model, and in Section \[sec:Information-Friction-Section\], where we study a multi-dimensional generalization of the Grover bit-meters model. In these sections we present lower bounds for decoders, as the derivation for encoding lower bounds is almost exactly the same. We provide an outline of the technique for encoder lower bounds in Section \[sec:encoderLowerBounds\]. In Section \[sec:Limitations-of-Result\] we discuss limitations and weaknesses in the model used. In Section \[sec:Other-Energy-Models\], we discuss other energy models of computation. In Section \[sec:Future-Work\] we discuss possible future work, and conjecture that similar tradeoffs may extend to circuits that perform inference.
*Notation:* We use standard Bachmann-Landeau notation in this paper. The statement $f(x)=O(g(x))$ means that for sufficiently large $x$, $f(x)\le cg(x)$ for some positive constant $c$. The statement $f(x)=\Omega(g(x))$ means that for sufficiently large $x$, $f(x)\ge cg(x)$ again for some constant $c$. The statement $f(x)=\Theta(g(x))$ means that there are two positive constants $b$ and $c$ such that $b\le c$ and for sufficiently large $x$, $bg(x)\le f(x)\le cg(x)$.
Prior Related Work: Computational Complexity Lower Bounds for Good Decoders and Encoders\[sec:Computational-Complexity-Lower\]
==============================================================================================================================
The earliest work on computational complexity lower bounds for good decoding comes from Savage in [@savage1969transactionsArticle] and [@savagePart2], which considered bounds on the memory requirements and number of logical operations needed to compute decoding functions. However, wiring area is a fundamental cost of good decoding and the authors do not consider this. More recently, in [@ElGamal], the authors use a model similar to our model, except the notion of “area” the authors use is the size of the smallest rectangle that completely encloses the circuit under consideration.
In [@groverFundamental], Grover *et al.* consider the same model that we do, and find Thompson energy lower bounds as a function of probability of block error probability for good encoders and decoders. Our analysis of the Thompson model differs from the approach of Grover *et al.* in a number of ways. Firstly, central to the work of Grover *et al.* is a bound on block error probability if inter-subcircuit bits communicated is low (presented in Lemma 2 in the Grover *et al.* paper), which is analogous to our result in (\[eq:theMainArgument\]) of the proof of Theorem \[thm:TheoremRelatingfofntoEnergy\]. Our result simplifies this relationship using simple probability arguments. Secondly, the Grover *et al.* paper does not present what energy-optimal number of clock cycles are in terms of asymptotic probability of block error, nor do they present the fundamental tradeoff between number of clock cycles, energy, and reliability within the Thompson model that we present in this paper. Moreover, the technique of [@groverFundamental] does not extend to serial implementations.
In [@blakeConstantPePaper] we considered the corner case of decoding schemes in which block error probability asymptotically was less than $\frac{1}{2}$ for serial and parallel decoding schemes. We did not, however, analyze schemes in terms of the rate at which block error probability approaches $0$, nor did we compute energy-optimal number of clock cycles as we do herein.
There has also been some work on complexity scaling rules for encoding and decoding of specific types of codes. Low density parity check coding VLSI scaling rules have been studied in [@BlakeLDPCAlmostSure; @ganesanPaper] and polar coding scaling rules have been studied in [@blakeKschischangPolarScaling]. The scaling rules presented in this paper are general and apply to any code.
Another computational model that has proven more tractable than the Turing Time complexity model is the constant depth circuit model (see [@AroraBarakComputationalComplexityBook] for a detailed description of this model). Super-polynomial lower bounds on the size of constant depth circuits that compute certain notions of “good encoding functions” (though not decoding) were derived in [@lovettViolaCannotSampleGoodCodes]. In this case, the notion of “good” considered was the ability to correct at least $\Omega\left(n\right)$ errors at rates asymptotically above $0$. Similar related work exists in [@rychkov] which discovered lower bounds on the formula-size of functions that perform good error control coding; similar bounds were later discovered in [@kojevnikov].
Thompson Model\[sec:ThompsonModelSection\]
==========================================
Circuit Model\[sub:Circuit-Model\]
----------------------------------
The model we will consider derives from Thompson [@Thompson]. The specific model we consider has been studied in [@blakeConstantPePaper; @groverFundamental; @BlakeLDPCAlmostSure; @ganesanPaper]. The reader should refer to [@blakeConstantPePaper] for details of the model. The important parameters to be extracted from the model are $A$, the circuit area, and $T$, the number of clock cycles in a computation. Since in this paper we are only concerned with scaling rules, we assume that both the technology constant and the wire width considered in [@groverFundamental; @blakeConstantPePaper] are equal to $1$. The energy of a computation is thus defined as $E=AT$.
Note that a circuit can be associated with a graph in the natural way, in which a wire corresponds to an edge of the graph and a node corresponds to a vertex. An edge connects two vertices if their associated nodes are connected by wires. A diagram of a small circuit next to its associated graph is given in Fig. \[fig:circuitIsAGraphDiagram\].
Lemma \[lem:ThompsonLemma\] presented below is derived in [@Thompson] and it relates the area of a circuit to its graph’s minumum bisection width, and is a key component of our Thompson model circuit lower bounds.
Definitions and Lemmas
----------------------
To present the main results of this paper we shall present a sequence of definitions and lemmas similar to [@blakeConstantPePaper; @groverFundamental].
\[lem:GuessingLemma\] [@blakeConstantPePaper] Suppose that $X$, $Y$, and $\hat{X}$ are random variables that form a Markov chain $X\rightarrow Y\rightarrow\hat{X}$ and $X$ takes on values from a finite alphabet $\mathcal{X}$ with a uniform distribution, (i.e., $P\left(X=x\right)=\frac{1}{\left|\mathcal{X}\right|}$ for all $x\in\mathcal{X}$), $Y$ takes on values from a finite set $\mathcal{Y}$, and $\hat{X}$ from a set $\mathcal{\hat{X}}$. Suppose as well that $\mathcal{\hat{X}}\in\mathcal{X}$. Then:
$$P\left(\hat{X}=X\right)\le\frac{\left|\mathcal{Y}\right|}{\left|\mathcal{X}\right|}.$$
We will interpret $X$ as the set of symbols a particular subcircuit will need to estimate, $\hat{X}$ as that subcircuit’s estimate of those symbols, and $Y$ as the bits injected into the subcircuit during the computation. Note that this result mirrors the result of Lemma 4 in [@GroverInfoFriction]. In this lemma, the author proves that if a circuit has $\frac{r}{3}$ bits to make an estimate $\hat{X}$ of a random variable $X$ that is uniformly distributed over all binary strings of length $r$, then that circuit makes an error with probability at least $\frac{1}{9}$. Our lemma presented here includes this lemma as a special case by setting $\left|\mathcal{Y}\right|=2^{\frac{r}{3}}$ and $\left|\mathcal{X}\right|=2^{r}$. In this case we can infer: $P\left(\hat{X}\ne X\right)\ge1-\frac{2^{\frac{r}{3}}}{2^{r}}\ge1-2^{-\frac{2}{3}r}>\frac{1}{9}$, where the last inequality is implied by $r\ge1$.
(of Lemma \[lem:GuessingLemma\]) See [@blakeConstantPePaper]. This flows from a simple application of the law of total probability and the definition of a Markov chain.
A *bisection* of a graph $G=\left(V, E\right)$ of a set of vertices $V'\in V$ is a set of edges $E'\in E$ that, once removed from the graph, results in two disconnected subgraphs with vertices $V_{1}$ and $V_{2}$ in which $\left|\left|V'\cap V_{1}\right|-\left|V'\cap V_{2}\right|\right|\le1$. That is, it is the set of edges that, once removed, divides the vertices of $V'$ roughly in half. The *minimum bisection width* of a set of vertices $V'$ is the size of a smallest bisection.
Note that since a circuit is associated with a graph, we can discuss such a circuit’s minimum bisection width, that is the minimum bisection width of the graph with which it is associated. Herein we will consider bisecting the output nodes of a circuit.
\[lem:ThompsonLemma\]All circuits whose associated graphs have minimum bisection width $\omega$ have circuit area $A\ge\frac{\omega^{2}}{4}$.
See Thompson [@Thompson].
We now discuss the notion of nested minimum bisection, a concept introduced by Grover *et al.* in [@groverFundamental] and also used in [@blakeConstantPePaper] which we again present here so the paper is self contained.
Suppose that a circuit has $k$ output nodes. If the output nodes of such a circuit are minimum bisected, this results in two disconnected subcircuits each with, roughly, $\frac{k}{2}$ output nodes. These two subcircuits can each have their output nodes minimum bisected again, resulting in four disconnected subcircuits, now each with roughly $\frac{k}{4}$ output nodes.
This process of nested minimum bisections on a circuit, when repeated $r$ times, is called *performing $r$-stages of nested minimum bisections.* In the case of this paper, the set of nodes to be minimum bisected will be the output nodes. We may also refer to this process as *performing nested bisections,* and a circuit under consideration in which nested bisections have been performed as a *nested bisected circuit*. Note that we will omit the term “minimum” in discussions of such objects, as this is implicit.
Note that associated with an $r$-stage nested bisected circuit are $2^{r}$ subcircuits. Note as well that once a subcircuit has only one node, it does not make sense to bisect that subcircuit again. Suppose we are nested-bisecting the $k$ output nodes of a circuit. In this case, one cannot meaningfully nested-bisect the output nodes of a circuit $r$ times if $2^{r}>k$.
Note that each of the $2^{r}$ subcircuits induced by the $r$-stage nested bisection may have some internal wires, and also wires that were deleted and connect to nodes in other subcircuits. We can index the $2^{r}$ subcircuits with the symbol $i$.
Let the number of wires attached to nodes in subcircuit $i$ that were deleted in the nested bisections be $f_{i}$. This quantity is the *fan-out of subcircuit $i$.*
We shall also consider the bits communicated to a given subcircuit.
\[def:biBitsCommunicatedToiThSubcircuit\]Let $b_{i}=f_{i}T$, where we recall that $T$ is the number of clock cycles used in the running of the circuit under consideration. This quantity is called the *bits communicated to the $i$th subcircuit.*
We can now define an important quantity.
The quantity $B_{r}=\sum_{i=1}^{2^{r}}b_{i}$ is the *inter-subcircuit bits communicated.*
Note that each subcircuit induced by the nested bisections will each have close to $\frac{k}{2^{r}}$ output nodes within them (a consequence of choosing to bisect the output nodes at each stage), however, each may have a different number of input nodes.
This quantity is called the *number of input nodes in the $i$th subcircuit* and we denote it $n_{i}$.
Note that $\sum_{i=1}^{2^{r}}n_{i}=n$ for all valid choices of $r$. That is, the sum over the number of input nodes in each subcircuit is the total number of input nodes in the original circuit.
This now allows us to present an important lemma.
\[lem:GroverATsquaredLemma\]All fully-parallel circuits with inter-subcircuit bits communicated $B_{r}$ have product $AT^{2}$ bounded by:
$$AT^{2}\ge\frac{\left(\sqrt{2}-1\right)^{2}}{32}\frac{B_{r}^{2}}{2^{r}}=c_{1}\frac{B_{r}^{2}}{2^{r}}\label{eq:ATsquaredBound}$$
where we define $c_{1}=\frac{\left(\sqrt{2}-1\right)^{2}}{32}$.
This result, from Grover *et al.* [@groverFundamental] flows from applying Lemma \[lem:ThompsonLemma\] recursively on the nested-bisected structure and optimizing.
\[lem:energyBrLemma\]All fully-parallel circuits with inter-subcircuit bits communicated $B_{r}$ and number of input nodes $n$ have product $AT$ bounded by: $$AT\ge c_{2}\sqrt{\frac{n}{2^{r}}}B_{r}$$ where we define $c_{2}=\frac{\sqrt{2}-1}{4\sqrt{2}}$.
See [@groverFundamental]. This result flows from the observation that $A\ge n$ for a fully parallel circuit and then combining this inequality with (\[eq:ATsquaredBound\]).
An *$(n,k)$-decoder* is a circuit that computes a decoding function $f:\left\{ 0,1\right\} ^{n}\rightarrow\left\{ 0,1\right\} ^{k}$. It is associated with a codebook, (and therefore, naturally, an encoding function, which computes a function $g:\left\{ 0,1\right\} ^{k}\rightarrow\left\{ 0,1\right\} ^{n}$), a channel statistic, $P\left(y^{n}|x^{n}\right)$ (which we will assume herein to be the statistic induced by $n$ channel uses of a binary erasure channel), and a statistic from which the source is drawn $p\left(x^{k}\right)$ (which we will assume to be the statistic generated by $k$ independent fair binary coin flips). The quantity $n$ is the block length of the code, and the quantity $k$ is the the number of bits decoded.
The *block error probability* of a decoder, denoted $P_{\mathrm{e}}$, is the probability that the decoder’s estimate of the original source is incorrect. Note that this probability depends on the source distribution, the channel, and the function that the decoder computes.
A *decoding scheme* is an infinite sequence of circuits $D_{1},D_{2,}\ldots$ each of which computes a decoding function, with block lengths $n_{1}<n_{2}<\ldots$ and bits decoded $k\left(n_{1}\right),k\left(n_{2}\right),\ldots$. They are associated with a sequence of codebooks $C_{1},C_{2},\ldots$ and a channel statistic.
We assume throughout this paper that the channel statistic associated with each decoder is the statistic induced by $n$ uses of a binary erasure channel. Our lower bound results also apply to any channel that is a degraded erasure channel, including the binary symmetric channel. Our results in terms of binary erasure probability $\epsilon$ can be applied to decoding schemes for the binary symmetric channel with crossover probability $p$ by substituting $p=2\epsilon$.
We let $P_{e}\left(n\right)$ denote the *block error probability* for the decoder with input size $n$. We let $R\left(n\right)=\frac{k\left(n\right)}{n}$ be the *rate* of the decoder with input size $n$.
We also classify decoding schemes in terms of how their probability of error scales in the definition below.
An $f\left(n\right)$*-decoding scheme* is a decoding scheme in which for sufficiently large $n$ the block error probability $P_{\mathrm{e}}(n)< f(n)$.
The *asymptotic-rate,* or more compactly, the *rate* of a decoding scheme is $\lim_{n\rightarrow\infty}R\left(n\right)$, if this limit exists, which we denote $R$.
Note that the rate of a decoding scheme may not be the rate of any particular codebook in the decoding scheme.
An *exponentially-low-error decoding scheme* is an $e^{-cn}$-decoding scheme for some $c>0$ with asymptotic rate $R$ greater than $0$.
We will also consider another class of decoding schemes, one which can be considered less reliable.
A *polynomially-low-error decoding scheme* is a $\frac{1}{n^{t}}$-decoding scheme for some $t>0$ with asymptotic rate $R>0$.
We will also need to define a sublinear function, which will be used to deal with a technicality in Theorem \[thm:TheoremRelatingfofntoEnergy\].
A sublinear function $f\left(n\right)$ is a function in which $\lim_{n\rightarrow\infty}\frac{f\left(n\right)}{n}=0$.
Main Lower Bound Results
------------------------
We can now state the main theorem of this paper.
\[thm:TheoremRelatingfofntoEnergy\]All $f\left(n\right)$-decoding schemes associated with a binary erasure channel with erasure probability $\epsilon$ in which $f\left(n\right)$ monotonically decreases to $0$ and in which $-\ln\left(f\left(n\right)\right)$ is a sublinear function have energy that scales as $$E\ge c_{3}\sqrt{\frac{\ln\left(f\left(n\right)\right)}{\ln(\epsilon)}}k\label{eq:energyLowerBound}$$ where $c_{3}=\frac{\sqrt{\ln2}(\sqrt{2}-1)}{16\sqrt{2}}$ and $AT^{2}$ complexity that scales as:
$$AT^{2}\ge c_{4}\frac{k^{2}\ln\left(f\left(n\right)\right)}{n\ln(\epsilon)}\label{eq:ATsquaredTheoremAsFunctionOfFn}$$
for another positive constant $c_{4}=\frac{\ln(2)\left(\sqrt{2}-1\right)^{2}}{512}$.
Associated with each decoder is its $B_{r}$, the inter-subcircuit bits communicated. We can choose $r$ to be any function of $n$ so long as $2^{r}<nR\left(n\right)=k\left(n\right)$. From here on, we will suppress the dependence of $r\left(n\right)$, $k\left(n\right)$, and $R\left(n\right)$ on $n$. For ease of notation, let $N=2^{r}$ be the number of subcircuits induced by the $r$-stages of nested bisections. Consider any specific sufficiently large circuit in our decoding scheme, and suppose that $B_{r}<\frac{k}{2}$. Then there exists at least $\frac{N}{2}$ subcircuits in which $b_{i}<\frac{k}{N}$ (where we recall $b_{i}$ is the bits communicated to the $i$th subcircuit from Definition \[def:biBitsCommunicatedToiThSubcircuit\]). Suppose not, *i.e.*, that there are $\ge\frac{N}{2}$ subcircuits with $b_{i}\ge\frac{k}{N}$. Then, $B_{r}\ge\frac{k}{N}\frac{N}{2}=\frac{k}{2}$, violating the assumption that $B_{r}<\frac{k}{2}$. Call the set of at least $\frac{N}{2}$ subcircuits with bits communicated to them less than $\frac{k}{N}$ $Q$. Using a similar averaging argument, we claim that within $Q$ there must be one subcircuit in which $n_{i}\le\frac{2n}{N}$. If not, if all $\frac{N}{2}$ subcircuits in $Q$ have greater than $\frac{2n}{N}$ input bits injected into them, then the total number of inputs nodes in the entire circuit is greater than $\frac{2n}{N}\frac{N}{2}=n$, but there are only $n$ input nodes in the entire circuit. Thus, there is at least one subcircuit in $Q$ in which $b_{i}<\frac{k}{N}$ and $n_{i}\le\frac{2n}{N}$.
Suppose that all the input bits injected into this special subcircuit are erased. Then, that subcircuit makes an error with probability at least $\frac{1}{2}$ by Lemma \[lem:GuessingLemma\], since it will have to form an estimate of $\frac{k}{N}$ bits by only having injected into it fewer than $\frac{k}{N}$ bits. Thus, if $B_{r}\le\frac{k}{2}$ then: $$\begin{aligned}
P_{e} & \ge & P\left(\mathrm{error|\mathrm{all\ n_{i}\ bits\ erased}}\right)P\left(\mathrm{all\ n_{i}\ bits\ erased}\right)\\
& \ge & \frac{1}{2}\epsilon^{n_{i}}\end{aligned}$$ where this first inequality flows from summing one term in a law of total probability expansion of the probability of block error, and the second from lower bounds on these probabilities.
Combining this observation with the fact the $n_{i}\le\frac{2n}{N}$ gives us the following observation: $$\mathrm{if\ }B_{r}\le\frac{k}{2}\mathrm{\ then\ }P_{\mathrm{e}}\ge\frac{1}{2}\epsilon^{n_{i}}\ge\frac{1}{2}\epsilon^{\frac{2n}{N}}\label{eq:theMainArgument}$$ This is true for any valid choice of $r$.
Now suppose that our decoding scheme is an $f\left(n\right)$-decoding scheme. We choose $r$ to be $$r=\left\lfloor \log_{2}\frac{2n\ln(\epsilon)}{\ln(2)\ln\left(f\left(n\right)\right)}\right\rfloor$$ so that $$N=2^{r}\approx\frac{2n\ln(\epsilon)}{\ln (2)\ln\left(f\left(n\right)\right)}.\label{eq:divisionIntoN}$$ This is a valid choice of $r$ because $N$ cannot grow faster than $O\left(n\right)$ because we assumed $P_{e}\left(n\right)$ was monotonically decreasing (easily checked by inspection). Note as well that $N$ increases with $n$ because of the sub-linearity assumption of $-\ln\left(f(n)\right)$. Then, if $B_{r}\le\frac{k}{2}$, by directly substituting into (\[eq:theMainArgument\]), $$\begin{aligned}
P_{e} & \ge & \frac{1}{2}\exp\left(\frac{\ln(\epsilon)2\ln(2) n\ln\left(f(n)\right)}{2n\ln(\epsilon)}\right)\\
& = & \frac{1}{2}\exp\left(\ln(2)\ln\left(f(n)\right)\right)=f\left(n\right).\end{aligned}$$ In other words, if $B_{r}\le\frac{k}{2}$ then our decoding scheme is not an $f(n)$-decoding scheme. Thus, for this choice of $r$, $B_{r}>\frac{k}{2}$.
Thus, by Lemma \[lem:energyBrLemma\], $$\begin{aligned}
E & \ge & c_{2}\sqrt{\frac{n}{2^{\left\lfloor \log_{2}\frac{2n\ln(\epsilon)}{\ln(2)\ln\left(f(n)\right)}\right\rfloor }}}\frac{k}{2}\\
& \ge & c_{2}\sqrt{\frac{n}{2^{\log_{2}\frac{2n\ln(\epsilon)}{\ln(2)\ln\left(f(n)\right)}+1}}}\frac{k}{2}\\
& \ge & c_{2}\sqrt{\frac{n}{2\left(\frac{2n\ln(\epsilon)}{\ln(2)\ln\left(f(n)\right)}\right)}}\frac{k}{2}\\
& \ge & c_{3}\sqrt{\frac{\ln\left(f(n)\right)}{\ln(\epsilon)}}k\\\end{aligned}$$ where we substituted the value for $N$ in the first line, used the fact that $\left\lfloor x\right\rfloor \le x+1$ in the second, and simplified the lines that followed, proving inequality (\[eq:energyLowerBound\]) of the theorem. As well, by Lemma \[eq:ATsquaredBound\], using $B_{r}>\frac{k}{2}$ for this choice of $r$, following a similar substitution as in the previous paragraph: $$\begin{aligned}
AT^{2} & \ge & c_{1}\frac{B_{r}^{2}}{2^{r}}.\\
& \ge & c_{1}\frac{k^{2}}{4\left(2^{\left\lfloor \log_{2}\frac{2n\ln(\epsilon)}{\ln(2)\ln\left(f(n)\right)}\right\rfloor }\right)}\\
& \ge & c_{1}\frac{k^{2}}{4\left(2^{\log_{2}\frac{2n\ln(\epsilon)}{\ln(2)\ln\left(f(n)\right)}+1}\right)}\\
& = & c_{1}\frac{k^{2}}{8\left(\frac{2n\ln(\epsilon)}{\ln(2)\ln\left(f(n)\right)}\right)}\\
& = & \frac{c_{1}\ln(2)}{16}\frac{k^{2}\ln\left(f(n)\right)}{\ln(\epsilon)}\end{aligned}$$ and the inequality in (\[eq:ATsquaredTheoremAsFunctionOfFn\]) flows from substituting the appropriate value for $c_{1}$ as defined in Lemma \[eq:ATsquaredBound\].
\[cor:exponentialCorollary\]All exponentially low error decoding schemes have energy that scales as $$E\ge\Omega\left(\frac{n^{\frac{3}{2}}}{p\left(n\right)}\right)$$ for all functions $p\left(n\right)$ that increase without bound. In other words, all exponential probability of error decoding schemes have energy at least that scales very close to $\Omega\left(n^{\frac{3}{2}}\right)$. Moreover, any such scheme that has energy that grows optimally, i.e. as $AT=O\left(n^{\frac{3}{2}}\right)$, must have $T\left(n\right)\ge\Omega\left(n^{0.5}\right)$.
Note that an exponentially low error decoding scheme has $P_{e}\le e^{-cn}$. Thus, such a scheme is also an $e^{-c\frac{n}{p\left(n\right)}}$-decoding scheme, for any increasing $p\left(n\right)$. The result then directly flows by substituting $f\left(n\right)=e^{-c\frac{n}{p\left(n\right)}}$ into (\[eq:energyLowerBound\]) of Theorem \[thm:TheoremRelatingfofntoEnergy\].
For the second part of the corollary, suppose that for some constant $c$, a decoding scheme has $$AT=\Theta (n^{\frac{3}{2}} ).\label{eq:ATBoundforExponential}$$ We have as well from (\[eq:ATsquaredTheoremAsFunctionOfFn\]) and substituting $f(n)=e^{-c\frac{n}{p\left(n\right)}}$
$$AT^{2}\ge \Omega \left( \frac{n^{2}}{p\left(n\right)} \right) \label{eq:ATsquaredForExponentialBound}$$
where we use the fact that $k=Rn$ (since by definition exponentially-low error decoding schemes have asymptotic rate greater than $0$).
Suppose that $$T< O\left( \frac{n^{\frac{1}{2}}}{g\left(n\right)} \right) \label{eq:Tbound}$$ for a $g\left(n\right)$ that grows with $n$, *i.e.,* that $T$ asymptotically grows slower than $O\left(n^{\frac{1}{2}}\right)$. Then, to satisfy (\[eq:ATsquaredForExponentialBound\]) we need
$$AT^{2}\ge \Omega \left( \frac{n^{2}}{p\left(n\right)} \right)\label{eq:toBeUnsatisfied}$$
for all increasing $p\left(n\right)$, implying $$A\ge \Omega \left( \frac{ng\left(n\right)^{2}}{p\left(n\right)} \right).$$ To see this precisely, suppose otherwise and then it is easy to see that, combined with (\[eq:Tbound\]) the inequality in (\[eq:toBeUnsatisfied\]) will be unsatisfied. If this is true, however, then the product $$AT\ge \Omega \left( \frac{ng\left(n\right)^{2}}{p\left(n\right)}\frac{n^{\frac{1}{2}}}{g\left(n\right)} \right) = \Omega \left( \frac{n^{\frac{3}{2}}g\left(n\right)}{p\left(n\right)} \right).$$
Since this is true for all increasing $p\left(n\right)$, it is true for, say, $p\left(n\right)=\ln g\left(n\right)$, implying that the product $AT$ grows strictly faster than $\Omega\left(n^{\frac{3}{2}}\right)$, contradicting the assumption of (\[eq:ATBoundforExponential\]).
We generalize Corollary \[cor:exponentialCorollary\] to decoding schemes with different asymptotic block error probabilities below:
\[thm:generalTheoremForoptimalT\] All $f(n)$-decoding schemes with asymptotic rate greater than $0$ in which $f(n)$ is sub-exponential with energy that scales as $E=\Theta\left(\sqrt{\ln f(n)} n\right)$ (that is, their energy matches the lower bound of (\[eq:energyLowerBound\]) of Theorem \[thm:TheoremRelatingfofntoEnergy\]) must have $T\left(n\right)=\Omega\left(\sqrt{\ln f(n)}\right)$. Moreover, for all decoding schemes in which $T\left(n\right)$ is faster than this optimal, $E\ge\Omega\left(\frac{n\ln f(n)}{T\left(n\right)}\right)$.
Suppose that $$AT=\Theta\left(\sqrt{\ln f(n)}n\right)\label{eq:ATAssumptionForGeneralTheorem}$$ Note that from (\[eq:ATsquaredTheoremAsFunctionOfFn\]), $$AT^{2}\ge\Omega\left(n\ln f(n)\right).\label{eq:ATsquaredBoundForThisProof}$$
As well, suppose $T\left(n\right)\le O\left(\frac{\sqrt{\ln f(n)}}{g\left(n\right)}\right)$ for some increasing $g\left(n\right)$. Then, from the bound (\[eq:ATsquaredBoundForThisProof\]) $A\ge\Omega\left(n\sqrt{\ln f(n)}g^{2}\left(n\right)\right)$ (to prove this, suppose otherwise and derive a contradiction). This implies then that $AT\ge\Omega\left(\sqrt{\ln f(n)}ng\left(n\right)\right)$, contradicting (\[eq:ATAssumptionForGeneralTheorem\]).
Moreover, for all $T\left(n\right)$ growing slower than that required for optimal energy, this implies that $A\ge\Omega\left(\frac{n\ln\left(f(n)\right)}{T^{2}\left(n\right)}\right)$, which implies $E\ge\Omega\left(\frac{n\ln f(n)}{T\left(n\right)}\right).$
\[cor:All-polynomially-low-error\]All polynomially-low error decoding schemes have energy that scales at least as $$E\ge\Omega\left(n\sqrt{\ln n}\right).\label{eq:EnergyOfPolynomiallyLower}$$ If this optimal is reached, then $T\left(n\right)\ge \Omega (\sqrt{\ln n})$.
This energy lower bound flows from letting $f(n)=\frac{1}{n^{k}}$ and then substituting this value into (\[eq:energyLowerBound\]). The time lower bound flows from directly applying Theorem \[thm:generalTheoremForoptimalT\].
Serial Decoding Scheme Scaling Rules
------------------------------------
Let the number of output nodes in a particular decoder be denoted $j$ (in a decoding scheme this will be a function of $n$).
A *serial* decoding scheme is one in which $j$ is constant.
In [@blakeConstantPePaper] we considered the case of allowing the number of output nodes $j$ to increase with increasing block length. We required an assumption that such a scheme be output regular, which we define below.
[@blakeConstantPePaper] An *output regular* circuit ** is one in which each output node of the circuit outputs exactly one bit of the computation at specified clock cycles. This definition excludes circuits where some output nodes output a bit during some clock cycle and other output nodes do not during this clock cycle. An *output regular decoding scheme* is one in which each decoder in the scheme is an output regular circuit.
\[thm:All-serial–decoding\]All serial $f(n)$-decoding schemes have energy that scales as $\Omega\left(n\ln f(n)\right)$.
The $\Omega\left(n\ln f(n)\right)$ lower bound flows from following the arguments of the proof of Theorem 2 in [@blakeConstantPePaper], by showing that any decoding scheme in which the area scales less than $O(\ln f(n))$ cannot be an $f\left(n\right)$-decoding scheme.
\[thm:All-output-regular\]All output regular increasing-output node $f\left(n\right)$-decoding schemes have energy that scales as $\Omega\left(n\left(\ln f(n)\right)^{\frac{1}{5}}\right)$.
From the derivations preceding equation (13) in [@blakeConstantPePaper], following a similar argument as in this paper, we divide the circuit into $M=\Theta\left(\frac{n}{A}\right)$ epochs as before, and divide the subcircuits into $N=\Theta\left(\frac{A}{\ln f\left(n\right)}\right)$ subcircuits through nested bisections. With this choice, we can follow the same arguments used in Theorem 3 in [@blakeConstantPePaper], and derive that all $f(n)$-decoding schemes must have $$AT\ge\Omega\left(n\left(\ln\left(f(n)\right)\right)^{\frac{1}{5}}\right).$$
\[sec:Information-Friction-Section\]Information Friction in Three-Dimensional Circuits
======================================================================================
The “information friction” computational energy model was introduced by Grover in [@GroverInfoFriction] and further studied by Vyavahare *et al.* in [@VyavahareBitMeters] and Li *et al.* in [@LiBakshiGroverCompressedSensing]. We generalize (and slightly modify) this model to three dimensions and use a similar approach to Grover to obtain some non-trivial lower bounds on the energy complexity of three dimensional bit-meters decoder circuits, in terms of block length and probability of error. We will discuss how this approach can be generalized to models in arbitrary numbers of dimensions. We present the model below and then prove our main complexity result.
- A circuit is a grid of computational nodes at locations in the set $\mathbb{Z}^{3}$, where $\mathbb{Z}$ is the set of integers. Some nodes are *inputs nodes*, some are *output nodes*, and some are *helper nodes*. Note that Grover [@GroverInfoFriction] considers this model in terms of a parameter characterizing the distance between the nodes, but since we are concerned with scaling rules, we will assume that they are placed at integer locations, allowing us to avoid unnecessary notation. The Grover paper considered scaling rules in which nodes are placed on a plane, in which the number of dimensions $d=2$. In our results we will discuss the case of $d=3$ and afterwards discuss how the approach can be generalized to an arbitrary number of spatial dimensions.
- A circuit is to compute a function of $n$ binary inputs and $k$ binary outputs.
- At the beginning of a computation, the $n$ inputs to the computation are injected into the input nodes. At the end of the computation the $k$ outputs should appear at an output node. A node can be both input and output.
- A node can communicate messages along its links to any other node, and can receive bits communicated to them from any other node.
- Each node has constant memory, and can compute any computable function of all the inputs it has received throughout the computation that is stored in their memory, to produce a message that it can send to any other node.
- We associate a computation with a directed multi-graph, that is, a set of edges linking the nodes. For every computation, there is one edge per bit communicated along a link in the computation’s associated multi-graph. The “cost” of an edge in such a multi-graph is the Euclidean distance between the two nodes that it connects. Note that if a node communicates $m$ bits to another node in a computation, then that computation’s associated multi-graph must have $m$ edges connecting the two nodes. This multi-graph is called a computation’s *communication multi-graph.*
- The *energy*, or the *bit-meters*, denoted $\beta$ of a computation is the sum of the costs of all the edges in the computation’s associated multi-graph (that is, the sum of the Euclidean distances of all the edges).
We consider a grid of three-dimensional cubes, with “inner cubes” nested within them. This object is a generalization of the “stencil” object defined by [@GroverInfoFriction].
An *$\left(L,\lambda\right)-$nested cube grid* is an infinite grid of cubes, with side length $L$ and inner cube side length $L\left(1-2\lambda\right)$. Note that the inner cubes are centered within the outer cubes. Fig. \[fig:nestedCubeGrid\] shows a diagram of one cube in a *$\left(L,\lambda\right)-$nested cube grid*, to which the reader can refer to visualize this nested cube structure. A set of nested cube grid parameters is *valid* if $L>0$ and $0<\lambda<\frac{1}{2}$.
![A diagram of one nested cube in an $(L, \lambda)$-nested cube grid, with the edge lengths labeled. A nested cube grid is an infinite grid of such nested cubes. The outer cubes each have side length $L$ and the inner cubes each have side length $L(1-2\lambda)$ at a distance $L\lambda$ from the faces of the outer cube.[]{data-label="fig:nestedCubeGrid"}](stencil){width="3"}
Note that a nested cube grid can be placed conceptually on top of a bit meters circuit. We will consider placing a nested cube grid in parallel with the Cartesian $3$-space that defines our circuit. We can specify the position of a nested cube grid that is parallel to a set of Cartesian coordinates by calling one of the corners of an outer cube the *origin,* and then specify the location of its origin. A particular set of parameters for a nested cube grid and a location for its origin (called its *orientation*) induces a set of subcircuits, defined below.
A *subcircuit*, associated with a particular orientation of a nested cube grid, is the part of a bit-meters circuit within a particular outer cube.
Nodes in any subcircuit can thus be considered to be either inside an inner cube or outside an inner cube. For any circuit with finite number of nodes there will thus be some cubes that contain computational nodes, and some that do not. We can label the subcircuits that contain nodes with the index $i$. The number of input nodes in cube $i$ we denote $n_{i}$. The number of output nodes in subcircuit $i$ we denote $k_{i}$. Furthermore, we denote the number of input nodes within the inner cube of subcircuit $i$ as $k_{\mathrm{in},i}$.
We define $k_{\mathrm{in}}=\sum k_{\mathrm{in},i}$, which is the *the number of output nodes within inner cubes*, which we will often simply refer to with the symbol $k_{\mathrm{in}}$.
We will show in Lemma \[lem:3DbitmetersLemma\] that there exists a nested cube grid orientation in which $k_{\mathrm{in}}$ is high.
The *internal bit meters* of a subcircuit $i$ is the length of all the communication multigraph edges completely within subcircuit $i$, plus the length of the parts of the edges within subcircuit $i$. This quantity is denoted with the symbol $\beta_{i}$. Note that $\beta=\sum_{\mathrm{all\ subcircuits\ j}}\beta_{j}$ (where we may have to sum over some subcircuits that do not contain any nodes).
Since a computation has associated with it its communication multi-graph, for a given subcircuit we can consider the subgraph formed by all the paths that start outside of the cube and end inside the inner cube. We can group all the vertices of this graph that start outside the outer cube and call this the source, and group all vertices inside an inner cube and call it the sink. For this graph we can consider its min-cut, the minimum set of edges that, once removed, disconnects the source from the sink.
The *number of bits communicated from outside a cube to within an inner cube,* or, *bits communicated,* is the size of this minimum cut. For a particular subcircuit $i$ we refer to this quantity with the symbol $b_{i}$.
This quantity is analogous (but not the same) as the quantity $b_{i}$ for the Thompson circuit model from Definition \[def:biBitsCommunicatedToiThSubcircuit\], and thus we use the same symbol. The reader should not confuse these symbols; the Thompson model definition applies to discussions in Section \[sec:ThompsonModelSection\], and the bit-meters model definition applies in this section, Section \[sec:Information-Friction-Section\].
If the $n_{i}$ internal bits of a subcircuit are fixed, then the subcircuit inside an inner cube will compute a function of the messages passed from outside the outer cube. Clearly, the size of the set of possible messages injected into this internal cube is $2^{b_{i}}$ (since $b_{i}$ is the min cut of the paths leading from outside to inside.)
\[lem:bitMetersProportionalToCommunicatedFromInsideToOut\] All subcircuits with bits communicated $b_{i}$ have internal bit meters at least $b_{i}\lambda L$.
This result flows from Menger’s Theorem [@mengersTheorem; @GoringshortMengersProof], which states that any network with min-cut $b_{i}$ has at least $b_{i}$ disjoint paths from source to sink. Each of these paths must have length at least $\lambda L$ from the triangle inequality.
This lemma makes rigorous the idea that to communicate $b_{i}$ bits from outside a subcircuit to within its inner square, the bit-meters this takes is proportional to the distance from outside an outer square to within an inner square ($\lambda L$) and the number of bits communicated.
In the lemma below we show that there exists an orientation of any nested cube grid such that $k_{\mathrm{in}}$ is high.
\[lem:3DbitmetersLemma\]For all three dimensional bit-meters circuits with $k$ output nodes, all valid nested cube grid parameters $L$ and $\lambda$, there exists an orientation of an $\left(L,\lambda\right)$-nested cube grid in which the number output nodes within inner cubes ($k_{\mathrm{in}}$) is bounded by: $$k_{\mathrm{in}}\ge\left(1-2\lambda\right)^{3}k$$
Note that the relative volume of the inner cubes is $\left(1-2\lambda\right)^{3}.$ This lemma says there exists an orientation of any nested cube grid in which the fraction of output nodes within inner cubes is at least this fraction, so this result is not surprising.
This is a natural generalization of the Grover result (See Lemma 2 of [@GroverInfoFriction]), which uses the probabilistic method. We consider placing the origin of an $\left(L,\lambda\right)$-nested cube grid uniformly randomly within a cube of side length $L$ centered at the origin in the Cartesian $3$-space. We index the $k$ output nodes by $i$. Let $1_{\mathrm{in,}i}$ be the indicator random variable that is equal to $1$ if output node $i$ is within an inner cube. Then, given the uniform measure on the position of the cube, the quantity $k_{\mathrm{in}}$ is a random variable. We observe: $$\begin{aligned}
k_{\mathrm{in}} & = & \sum_{i=1}^{k}1_{\mathrm{in,}i}\mathrm{,\ thus}\nonumber \\
E\left(k_{\mathrm{in}}\right) & = & E\left(\sum_{i=1}^{k}1_{\mathrm{in,}i}\right)\nonumber \\
& = & \sum_{i=1}^{k}E\left(1_{\mathrm{in,}i}\right)\nonumber \\
& = & \sum_{i=1}^{k}\left(1-2\lambda\right)^{3}\label{eq:averageNumberInside}\\
& = & k\left(1-2\lambda\right)^{3}\nonumber \end{aligned}$$ where in (\[eq:averageNumberInside\]) we use the observation that, for each output node, the probability that it is in an inner square is proportional to the relative area of the inner square. Thus, the expected value of $k_{\mathrm{in}}$ is $k\left(1-2\lambda\right)^{3}$ and so there must be at least one nested cube grid orientation in which $k_{\mathrm{in}}$ is greater than or equal to that value.
\[lem:maxNumberOfniLemma\]For all valid nested cube parameters $L$ and $\lambda$, $n_{i}\le\left(L+1\right)^{3}$ and thus for sufficiently large $L$ $n_{i}\le2L^{3}$.
Intuitively, there cannot be more than on the order of $L^{3}$ inner nodes in a cube of volume $L^{3}$. The $\left(L+1\right)^{3}$ bound comes from considering the corner case of a cube whose sides exactly touch output nodes.
We can now state the main results of this section.
\[3DTheorem\] All 3D-bit-meters decoders for a binary erasure channel with erasure probability $\epsilon$ of sufficiently large block length with block error probability $P_{e}$ have bit-meters $\beta$ bounded by:
$$\beta>\frac{27}{512}\left(\frac{\ln\left(4P_{\mathrm{e}}\right)}{2\ln(\epsilon)}\right)^{\frac{1}{3}}k.$$
We consider the number of bits communicated from outside a subcircuit $i$ to within the inner cube of subcircuit $i$ (**$b_{i}$**). It must at least be $k_{\mathrm{in},i}$ to overcome the case that all the input nodes in the entire cube are erased. If this does not happen, then one of the output nodes must guess at least one bit, making an error with probability at least $\frac{1}{2}$, formally justified by Lemma \[lem:GuessingLemma\]. This allows us to argue that: $$\begin{aligned}
P_{\mathrm{e}} & \ge & P\left(\mathrm{error}|\mathrm{all\ }n_{i}\mathrm{\ output\ bits\ are\ erased}\right)\nonumber \\
& & P\left(\mathrm{all\ }n_{i}\mathrm{\ output\ bits\ are\ erased}\right)\nonumber \\
& \ge & \frac{1}{2}\epsilon^{n_{i}}.\label{eq:probaOfErrorBound}\end{aligned}$$ If $\beta<\lambda Lk_{\mathrm{in}}$ then there exists a subcircuit indexed by $i$ in which $b_{i}<k_{\mathrm{in},i}$. Suppose otherwise, i.e. that $b_{i}\ge k_{\mathrm{in},i}$ for all $i$, then: $$\beta\ge\sum_{\mathrm{all\ subcircuits\ }i}\lambda Lb_{i}=\lambda L\sum b_{i}\ge\lambda L\sum k_{\mathrm{in},i}=\lambda Lk_{\mathrm{in}}$$ where we apply Lemma \[lem:bitMetersProportionalToCommunicatedFromInsideToOut\] after the first inequality, and for convenience suppress the subscript on the summation sign after the first instance. This contradicts our assumption that $\beta<\lambda Lk_{\mathrm{in}}$.
We choose the parameter $L$ in terms of probability of error in order to derive a contradiction if a circuit does not have high enough bit-meters. Specifically, we choose $$L=\left(\frac{\ln\left(4P_{\mathrm{e}}\right)}{2\ln(\epsilon)}\right)^{\frac{1}{3}}.\label{eq:LintermsofPe}$$
Consider the nested cube structure that has $k_{\mathrm{in}}\ge\left(1-2\lambda\right)^{3}k$ that must exist by Lemma \[lem:3DbitmetersLemma\]. If $\beta\le\lambda Lk_{\mathrm{in}}$ then there must exist a subcircuit $i$ that has less than $k_{\mathrm{in},i}$ bits injected into it from outside the subcircuit to within its inner cube. Thus: $$\mathrm{if\ }\beta\le\lambda Lk_{\mathrm{in}}\mathrm{\ then\ }P_{\mathrm{e}}\overset{(a)}{\ge}\frac{1}{2}\epsilon^{n_{i}}\overset{(b)}{\ge}\frac{1}{2}\epsilon^{2L^{3}}\overset{(c)}{\ge}2P_{\mathrm{e}}$$ where (a) flows from (\[eq:probaOfErrorBound\]), (b) from Lemma \[lem:maxNumberOfniLemma\], and (c) from the evaluation of this expression by substituting (\[eq:LintermsofPe\]). This is a contradiction. Thus, all bit meters decoders must have $$\begin{aligned}
\beta & > & \lambda Lk_{\mathrm{in}}\\
\beta & > & \lambda\left(1-2\lambda\right)^{3}Lk\\
& \ge & \lambda\left(1-2\lambda\right)^{3}\left(\frac{\ln\left(4P_{\mathrm{e}}\right)}{2\ln(\epsilon)}\right)^{\frac{1}{3}}k.\end{aligned}$$ The second inequality flows from the fact that we are considering the nested cube structure in which $k_{\mathrm{in}}\ge\left(1-2\lambda\right)^{3}k$ that must exist by Lemma \[lem:3DbitmetersLemma\]. We may choose any valid $\lambda$ to maximize this bound, and letting $\lambda=\frac{1}{8}$ gives us: $$\beta>\frac{27}{512}\left(\frac{\ln\left(4P_{\mathrm{e}}\right)}{2\ln(\epsilon)}\right)^{\frac{1}{3}}k.$$
Note that this argument naturally generalizes to $d$-dimensional space, in which all $d$-dimensional bit-meters decoders have energy that scales as $\beta\ge\Omega\left(\left(\ln\left(P_{\mathrm{e}}\right)\right)^{\frac{1}{d}}k\right)$. The key step in the proof to be altered is in a modification of Lemma \[lem:maxNumberOfniLemma\] and a choice of $L=c\left(\frac{\ln\left(4P_{\mathrm{e}}\right)}{\ln(\epsilon)}\right)^{\frac{1}{d}}$in line \[eq:LintermsofPe\] of the proof for some constant $c$ that may vary depending on the dimension. This implies, among other things, that exponentially low probability of error decoding schemes implemented in $d$-dimensions have bit-meters energy that scales as $\Omega\left(n^{1+\frac{1}{d}}\right)$. Obviously, the most engineering-relevant number of dimensions $d$ for this type of analysis are $d=2$ and $d=3$.
Encoder Lower Bounds\[sec:encoderLowerBounds\]
==============================================
In terms of scaling rules, all the decoder lower bounds presented herein can be extended to encoder lower bounds. The main structure of the decoder lower bounds (inspired by [@groverFundamental; @GroverInfoFriction]) involves dividing the circuit into a certain number of subcircuits. Then, we argue that if the bits communicated within the circuit is lower, then there must be one subcircuit where the bits communicated to it are less than the bits it is responsible for decoding. If all the inputs bits in that circuit are erased, the decoder must make an error with probability at least $1/2$.
In the encoder case, we also take inspiration from [@groverFundamental; @GroverInfoFriction]. In this case, the $n$ outputs of the encoder circuit can be divided into a certain number of subcircuits. Then we consider the bits communicated *out* of each subcircuit. This quantity must be proportional to the number of output bits in each subcircuit. Otherwise, there will be at least one subcircuit where the number of bits communicated out is less than the number of output nodes in the subcircuit. Call these bits that were not fully communicated out of this subcircuit *$Q$.* Suppose that once the output bits of the encoder are injected into the channel, all the bits in $Q$ are erased. Now, the decoder must use the other bits of the code to decode. But, the subcircuit containing $Q$ in the encoder communicated less than $\left|Q\right|$ bits to the other outputs of the encoder. By directly applying Lemma \[lem:GuessingLemma\], we see that no matter what function the decoder computes, it must make an error with probability at least $1/2$. An argument of this structure and following exactly the structure of Theorems \[thm:TheoremRelatingfofntoEnergy\], \[thm:generalTheoremForoptimalT\], \[thm:All-serial–decoding\], \[thm:All-output-regular\], and \[3DTheorem\] for the decoders gives us the following theorems, whose proofs are omitted.
All fully-parallel $f(n)$-encoding schemes with number of clock cycles $T(n)$ have energy $$E(n)\ge\Omega\left(\frac{n\log(f(n)}{T(n)}\right)$$ with optimal lower bound of $E\ge\Omega\left(n\sqrt{\log f(n)}\right)$ when $T(n)\ge\sqrt{\log(f(n))}$.
All serial, $f(n)$-encoding schemes have energy that scales as $$E(n)\ge\Omega\left(n\log f(n)\right).$$ All increasing output node, output-regular $f(n)$-encoding schemes have energy that scales as $$E(n)\ge\Omega\left(n\log^{1/5}\left(f(n)\right)\right).$$ Finally, all three-dimensional, bit-meters encoding schemes associated with block error probability $P_{\mathrm{e}}$ have energy that scales
$$E(n)\ge\Omega(n\ln P_{\mathrm{e}}).$$
Limitations of Results\[sec:Limitations-of-Result\]
===================================================
There are a number of weaknesses in the models we have used. Firstly, our results are asymptotic. For some set block error probability and rate, there may be a specific circuit that reaches this block error probability using a circuit design methodology that does not generalize to scale in a way as predicted by our theorems.
Note that our quantity $T$ refers to number of clock cycles, which reflects one of the main “time costs” in a circuit computation. In real circuits, the “time cost” of a computation involves two parameters: the number of clock cycles required, and the time it takes to do each clock cycle. In our model, we do not consider the time per clock cycle. In real circuits, this quantity often varies with wire lengths. We do not consider this in our model.
A particular weakness of the Thompson model we use is that it does not consider a quantity called *switching activity factor*. In circuit design, this quantity is the fraction of the circuit that “switches” during the course of the computation. And yet, our model assumes a switching activity factor of $1$. Thus, in terms of scaling rules, the Thompson model should be considered applicable only to computational schemes in which the switching activity factor does not change with increasing input sizes. On the other hand, the information-friction model accounts for the possibility of schemes in which switching activity factor changes with increasing block length, so, combined with the results of Grover, [@GroverInfoFriction], the asymptotic energy lower bounds we derive apply.
\[sec:Other-Energy-Models\]Other Energy Models of Computation
=============================================================
There has been some work on energy models of computation different from the Thompson energy models and Grover information friction models, and herein we provide a short review.
In [@binghamGreenstreet2012], Bingham *et al.* classify the tradeoffs between the “energy” complexity of parallel algorithms and “time” complexity for the problem of sorting, addition, and multiplication using a model similar to, but not the same as the model we use. In the grid model used by these authors, a circuit is composed of processing elements laid out on a grid, in which each element can perform an operation. In this model the circuit designer has choice over the speed of each operation, but this comes at an energy cost. Real circuits run at higher voltages can result in lower delay for each processing element but higher energy [@Honeisenmead]. The model used by the authors in [@binghamGreenstreet2012] captures some of this fundamental tradeoff. Note that our model assumes constant voltage. Non-trivial results that show how real energy gains can occur by lowering voltages in decoder circuits have been studied in [@Primeau], but we do not study this here.
Another energy model of computation was presented by Jain *et al.* in [@jainMolnarModel]. This model introduced an augmented Turing machine, a generalization of the traditional Turing machine [@turing1936]. The authors introduce a transition function, mapping the current instruction being read, the current state, the next state and the next instruction to the “energy” required to make this transition. This model (once the transition function is clearly defined for a specific processor architecture) would be good for the algorithm designer at the software level. However, we do not believe this model informs the specialized circuit designer. The Thompson model which we analyze, on the other hand, can include, as a special case, the energy complexity of algorithms implemented on a processor, as our model allows for a composition of logic gates to form a processor.
Landauer [@Landauer1961] derives that the energy required to erase one bit of information is at least $kT\ln 2$, where $k$ is Boltzmann’s constant, and $T$ is the temperature. Thus, a fundamental limit of computation comes from having to erase information. Of course, it may be possible to do reversible computation in which no information is erased that can use arbitrarily small amounts of energy, but such circuits must be run arbitrarily slowly. This suggests a fundamental time-energy tradeoff different from the tradeoff discussed herein. Landauer [@LandauerReview], Bennett [@bennetReview1982] and Lloyd [@sethLloydNature] provide detailed discussions and bibliographies on this line of work. Demaine *et al.* [@DemaineLandauerEnergy2016] extract a mathematical model from this line of work and analyze the energy complexity of various algorithms within this model. Note that the Thompson model we use is one informed by how modern VLSI circuits are created, even though they operate at energies far above ultimate physical limits.
Future Work\[sec:Future-Work\]
==============================
Currently, our work on lower bounds has not be extended to other channels, like the additive white Gaussian noise channel. Perhaps more interesting, however, is the question, do there exist polynomially low probability of error decoding schemes with energy that closely matches (\[eq:EnergyOfPolynomiallyLower\]) of Corollary \[cor:All-polynomially-low-error\], i.e., one with energy that scales as $\Omega\left(n\sqrt{\ln n}\right)$? This may have significantly lower energy than an exponentially-low error decoding scheme, and may provide sufficient error control performance. We do not know whether such a decoding scheme exists and this remains an important open question. It may be that decoding strategies with energy that scales like this are already invented but have simply not been analyzed in terms of their energy complexity.
The decoding problem for communication systems is a special case of the more general problem of inference. Well known algorithms used for inference, for example the Sum-Product Algorithm [@KschischangFactorGraphs] and variational methods [@wainwrightVariationalInference], include Gallager’s low-density parity-check decoding algorithms as a special case [@GallagerLDPC]. Thus, we conjecture that there may be similar tradeoffs between energy, latency, and reliability in circuits that perform inference.
[^1]: Part of this work was submitted for presentation at the 2016 International Symposium on Information Theory.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
We study the dimensional asymptotics of the effective actions, or functional determinants, for the Dirac operator $D$ and Laplacians $\Delta +\beta R$ on round $S^n$. For Laplacians the behavior depends on “the coupling strength” $\beta$, and one cannot in general expect a finite limit of $\zeta'(0)$, and for the ordinary Laplacian, $\beta=0$, we prove it to be $+\infty$, for odd dimensions. For the Dirac operator, Bär and Schopka conjectured a limit of unity for the determinant ([@BS]), i.e. $$\lim_{n\to\infty}\det(D, S^n_{\mathrm{can}})=1.$$
We prove their conjecture rigorously, giving asymptotics, as well as a pattern of inequalities satisfied by the determinants. The limiting value of unity is a virtue of having “enough scalar curvature” and no kernel. Thus for the important (conformally covariant) Yamabe operator, $\beta=(n-2)/(4(n-1))$, the determinant tends to unity.
For the ordinary Laplacian it is natural to rescale spheres to unit volume, since $$\lim_{k\to\infty}\det(\Delta, S_\mathrm{rescaled}^{2k+1})=\frac{1}{2\pi e}.$$
author:
- 'Niels Martin M[ø]{}ller'
date: 1 September 2007
title: '[Dimensional asymptotics of effective actions on $S^n$, and proof of Bär-Schopka’s conjecture]{}'
---
Introduction {#sec:Introduction}
============
Ever since spectral zeta functions of natural geometric elliptic differential operators on manifolds started appearing in mathematics ([@RS1], [@RS2]) and in the regularization of path integrals ([@Ha]), there has been interest in calculating values of the associated zeta determinants. Explicit formulae on various spaces have been derived ([@BrFuncDet], [@BS], [@DK], [@Dow1], [@Dow2], [@Ki], [@Va], [@We1], [@We2]), and some authors have displayed numerical features of the behavior of such determinants, culminating in the formulation of Conjecture 1 in [@BS], concerning the limit of the determinants of the Dirac operator on standard $n$-dimensional spheres, as the dimension $n$ grows large. Thus it seems an opportune time for investigating rigorously the asymptotics, using the previously established formulae for determinants. Contrary to what one might think, the rather complicated looking series, which involve Barnes zeta functions and generalized Bernoulli and Stirling numbers, can often be understood via quite elementary methods.
In Section \[sec:Dirac\] we find asymptotics of the effective action for the Dirac operator, and of the phase factor sometimes included. As a corollary this resolves the questions raised in the form of Conjecture 1 in [@BS]. The key point in the proof is the use of explicit formulas by Thomas Branson, Bär-Schopka and J. S. Dowker, and application of recursion to control the indirectly defined special functions and polynomial coefficients that appear, along with the functional equation for the Riemann zeta function $\zeta_R$ and some elementary estimates of $\zeta_R$ ([@BrFuncDet], [@BS], [@Dow1], [@Dow2]. See also the book by K. Kirsten [@Ki]).
In Section \[sec:Laplace\] for the Laplace-Beltrami operator on functions, more refined methods are needed. The reason seems to be that the ordinary Laplacian is not a “natural object” to consider. While an approach using polynomial coefficients such as (generalized) Stirling and Bernouilli numbers works well for the more natural conformally covariant Yamabe and Dirac operators, where in fact the zeta function and derivative at zero tend rapidly to zero, then in the cases when it tends to $\infty$, formidable cancellations amongst large terms should occur in such sum formulas. Namely for Dirac and Yamabe the convergence holds with absolute signs inside the Bernoulli/Stirling expansion sums, while for the ordinary Laplacian they do certainly not. As shown below, the zeta derivative at zero of the ordinary Laplacian goes to $+\infty$ as the logarithm of the dimension, while the same quantity for Dirac and Yamabe operators converges exponentially to zero.
The proof in the ordinary Laplacian case instead relies on estimating a broad range of Barnes zeta functions through convenient contour integral representations. Namely, by exploiting geometry of the situation it is possible to deform the basic Hankel contour to make use of a hybrid of Laplace’s method and the method of stationary phase, the idea of which is exactly that of cancellation of increasingly rapid oscillations. We note that the standard formulations of the method of stationary phase and Laplace’s method for contour integrals (see for example [@Ol]) do not seem feasible for the estimates needed here. What is needed is a mix of these standard tools, due to phase factors which are neither real nor purely imaginary. In this respect the present paper may also turn out useful for treating similar problems on contour integrals.
Dirac operator sphere determinants {#sec:Dirac}
==================================
As in [@BS], we define the Dirac operator determinant including a phase as follows.
$$\det(D):=\exp\Big(i\frac{\pi}{2}\big(\zeta_{D^2}(0)-\eta_D(0)\big)\Big)\exp\Big(-\frac{1}{2}\zeta^{\,\prime}_{D^2}(0)\Big).$$
To state the theorem, we fix the notation that $\varphi_D=\frac{\pi}{2}\zeta_{D^2}(0)-\eta_D(0)$. The theorem in particular proves Conjecture 1 of [@BS].
For the standard round spheres we have: $$\lim_{n\to\infty}\det(D, S^n)=1.$$ In fact we have the following asymptotics $$\begin{split}
|\det(D, S^n)|&=\exp\Big(O\big(({\begin{matrix}\frac{3}{4}\end{matrix}})^n\big)\Big),\\
\varphi_D&=O\big(({\begin{matrix}\frac{3}{4}\end{matrix}})^n\big).
\end{split}$$ Furthermore we have the inequalities $$\begin{split}
&|\det(D, S^n)|< 1,\quad\varphi_D>0,\quad\textrm{if}\quad n\equiv0\:\:\textrm{(mod 4)},\\
&|\det(D, S^n)|> 1,\quad\varphi_D=0,\quad\textrm{if}\quad n\equiv1\:\:\textrm{(mod 4)},\\
&|\det(D, S^n)|> 1,\quad\varphi_D<0,\quad\textrm{if}\quad n\equiv2\:\:\textrm{(mod 4)},\\
&|\det(D, S^n)|< 1,\quad\varphi_D=0,\quad\textrm{if}\quad n\equiv3\:\:\textrm{(mod 4)}.
\end{split}$$
This pattern is also visible in the numerics in [@BS] and is interesting in comparison to the present author’s results on local extremals of determinants, where that pattern is found to be (max, max, min, min), again depending on the respective dimensions mod $4$ (see [@Moeller]). This is yields another example of mod $4$ dependencies generically showing up for zeta regularised quantities.
We need to show that, writing from now on $S^n$ for $S^n_{\mathrm{can}}$, $$\begin{split}
&\lim_{n\to\infty}\zeta_{(D^2, S^n)}(0)=0,\\
&\lim_{n\to\infty}\zeta_{(D^2, S^n)}^{\,\prime}(0)=0.
\end{split}$$
We firstly review the expressions for the zeta functions of Dirac squared, by Tom Branson [@BrFuncDet]. Depending on the parity of $n$ we have as follows, in perfect agreement with [@BS]. $$\label{DiracReview}
\begin{split}
&\zeta_{(D^2,S^{n})}(s)=\frac{2^{k+1}}{(2k-1)!}\sum_{\alpha=0}^{k-1}d_{\alpha,k}\zeta_R(2s-2\alpha-1),\quad\textrm{if}\quad n=2k,\\
&\zeta_{(D^2,S^{n})}(s)=\frac{2^{k}}{(2k-2)!}\sum_{\alpha=0}^{k-1}e_{\alpha,k}\big(2^{2s-2\alpha}-1\big)\zeta_R(2s-2\alpha),\quad\textrm{if}\quad n=2k-1,
\end{split}$$ where the $d_{\alpha,k}$ and $e_{\alpha,k}$ are integers defined indirectly through the following polynomial expressions in $x$. $$\label{deDef}
\prod_{p=1}^{k-1}\big(x-p^2\big)=\sum_{\alpha=0}^{k-1}d_{\alpha,k}x^\alpha,\quad\textrm{and}\quad
\prod_{p=1}^{k-1}\big(x-(p-{\begin{matrix}\frac{1}{2}\end{matrix}})^2\big)=\sum_{\alpha=0}^{k-1}e_{\alpha,k}x^\alpha.$$ Also, to go from Branson’s formulas, we applied the functional equation $$\zeta_{1/2}(s)=\big(2^s-1\big)\zeta_R(s),$$ since it turns out convenient to express everything in terms of the Riemann zeta function $\zeta_R$.
We shall repeatedly rely on the controllability of $\zeta_R$ along the positive real axis. In fact we shall apply only the following two quite elementary facts. $$\begin{aligned}
\label{ZetaEstimates}
&0\leq\zeta_R(n)\leq C_{R},\quad n\in{\mathbb{N}}_+,\\ \label{ZetaPrimeEstimates}
&0\leq\zeta_R^{\,\prime}(n)\leq C_{R}',\quad n\in{\mathbb{N}}_0,\end{aligned}$$ for suitably chosen constants.
If now $n$ is odd, we have as always $\zeta_{(D^2, S^n)}(0)=0$, since the kernel is trivial. Thus we let $n$ be even and apply (\[DiracReview\]). Recalling the functional equation for the Riemann zeta function, in the form most convenient here, $$\label{FuncEq}
\zeta_R(s)=2^s\pi^{s-1}\sin\Big(\frac{\pi s}{2}\Big)\Gamma(1-s)\zeta_R(1-s),$$ we may now rewrite in a form where the asymptotics are more readily seen. $$\zeta_{(D^2,S^{2k})}(0)=\frac{2^{k+2}}{(2k-1)!}\sum_{\alpha=0}^{k-1}d_{\alpha,k}(-1)^{\alpha+1}(2\pi)^{-2\alpha-2}(2\alpha+1)!\zeta_R(2\alpha+2).$$ Note that for fixed $k$ this is a sum in which all the terms have the same sign. Namely $$\operatorname{sign}\big[d_{\alpha,k}\big]=(-1)^{k-1-\alpha},$$ while $\zeta_R$ is positive, since $\zeta_R(0)$ ($=-\frac{1}{2}$) is excluded. One thing we get from this is $$\operatorname{sign}\big[\zeta_{(D^2,S^{2k})}(0)\big]=(-1)^{k}.$$ Very importantly the constant sign (for fixed $k$) allows us to estimate term by term and we write $$\big\vert\zeta_{(D^2,S^{2k})}(0)\big\vert\leq C_RA(k),$$ where the positive numbers $A(k)$ are defined as $$\label{ADef}
A(k):=\frac{2^{k+2}}{(2k-1)!}(-1)^{k+1}\sum_{\alpha=0}^{k-1}d_{\alpha,k}(-1)^\alpha(2\pi)^{-2\alpha-2}(2\alpha+1)!.$$ The point is now that we can control the indirectly defined polynomial coefficients $d_{\alpha,k}$ through a simple recursion. From (\[deDef\]) we get $$d_{\alpha,k+1}=d_{\alpha-1,k}-k^2d_{\alpha,k}.$$ Note that it is of course implicitly understood that $d_{\alpha,k}=0$, if $k$ is not in the range $\{0,1,\ldots, k-1\}$. By a change of index for the first term and using $2\alpha(2\alpha+1)\leq 2k(2k+1)$, this shows the following estimate. $$\begin{split}
A(k+1)&\leq\frac{2}{2k(2k+1)}\bigg\{k^2+\frac{2k(2k+1)}{(2\pi)^2}\bigg\}A(k)\\
&=\bigg\{\frac{k}{2k+1}+\frac{1}{2\pi^2}\bigg\}A(k)\\
&\leq\frac{5}{9}A(k).
\end{split}$$ Thus for instance with $\delta^2=\frac{5}{9}<1$, we get $$\zeta_{(D^2,S^{n})}(0)=O\big(\delta^n\big),$$ proving in particular the claim that $$\label{zetazero}
\lim_{n\to\infty}\zeta_{(D^2, S^n)}(0)=0,$$ meaning that the phase converges to zero. Furthermore note that with $\delta^2=0.55\ldots$ this convergence is indeed rapid, as also witnessed by the numerics of Bär-Schopka.
To deal with the derivative we differentiate (\[DiracReview\]) at $s=0$ and use $\zeta(-2\alpha)=0,\: \alpha\in{\mathbb{N}}_{+}$. $$\begin{aligned}
&\zeta_{(D^2,S^{2k})}^{\,\prime}(0)=\frac{2^{k+2}}{(2k-1)!}\sum_{\alpha=0}^{k-1}d_{\alpha,k}\zeta_R^{\,\prime}(-2\alpha-1),\\ \label{OddDeriv}
&\zeta_{(D^2,S^{2k-1})}^{\,\prime}(0)=\frac{2^{k+1}}{(2k-2)!}\bigg\{-e_{0,k}\log 2+\sum_{\alpha=0}^{k-1}e_{\alpha,k}\big(2^{-2\alpha}-1\big)\zeta_R^{\,\prime}(-2\alpha)\bigg\}.\end{aligned}$$ Again we use the functional equation (\[FuncEq\]) which by differentiation gives $$\begin{split}\label{evenderiv}
&\zeta_R^{\,\prime}(-2\alpha-1)=2(-1)^\alpha(2\pi)^{-2\alpha-2}(2\alpha+1)!\zeta_R^{\,\prime}(2\alpha+2),\quad\alpha\in{\mathbb{N}}_0,\\
&\quad\quad\quad\quad\quad\quad\quad+\big[\log(2\pi)+\gamma-H_{2\alpha+1}\big]\zeta_R(-2\alpha-1),
\end{split}$$ $$\zeta_R^{\,\prime}(-2\alpha)=\pi(-1)^\alpha(2\pi)^{-2\alpha-1}(2\alpha)!\zeta_R(2\alpha+1),\quad\alpha\in{\mathbb{N}}_+.$$ Here $\gamma$ is Euler’s constant, and we have applied $$\label{GammaDiff}
\frac{\Gamma^{\,\prime}(n)}{\Gamma(n)}=H_{n-1}-\gamma,\quad H_{n-1}=\sum_{j=1}^{n-1}\frac{1}{j},\quad n\in{\mathbb{N}}.$$
Dealing first with the even dimensional case $n=2k$, we note that the second term in (\[evenderiv\]) gives a contribution that converges to zero. This follows since in absolute value, after summing over $\alpha$, it is altogether dominated by $$\Big\vert\big(\log(2\pi)+\gamma+H_{2k+1}\big)\zeta_{(D^2, S^{2k})}(0)\Big\vert=O\big(\delta^{2k}\big),$$ where again $\delta^2=\frac{5}{9}<1$ for example is admissible, since $\{\frac{9}{10}(1+1/\pi^2)\}^k H_{2k+1}$ is bounded.
To control now the first term in (\[evenderiv\]), note with $A(k)$ from (\[ADef\]) we have similar estimates here $$\bigg\vert\frac{2^{k+2}}{(2k-1)!}\sum_{\alpha=0}^{k-1}d_{\alpha,k}\zeta_R^{\,\prime}(-2\alpha-1)\bigg\vert\leq C_R'A(k),$$ thus we have the desired convergence $$\lim_{k\to\infty}\zeta_{(D^2, S^{2k})}^{\,\prime}(0)=0,$$ and in fact $\zeta_{(D^2, S^{2k})}^{\,\prime}(0)=O\big(\delta^{2k}\big)$ with, for example, $\delta^2=\frac{5}{9}$ again.
In odd dimensions the situation is changed only slightly. The polynomial coefficients are now $e_{\alpha,k}$ from (\[deDef\]) and as before we have $$\operatorname{sign}[e_{\alpha,k}]=(-1)^{k-1-\alpha}.$$ Writing now (\[OddDeriv\]) as $$\zeta_{(D^2,S^{2k-1})}^{\,\prime}(0)=-\frac{2^{k+1}}{(2k-2)!}\sum_{\alpha=0}^{k-1}e_{\alpha,k}Z(-2\alpha),$$ with $$Z(-2\alpha)= \begin{cases}
\big(1-2^{-2\alpha}\big)\zeta_R^{\,\prime}(-2\alpha) & \alpha\in{\mathbb{N}}_+\\
\log 2 & \alpha=0.
\end{cases}$$ Note that once again the signs match up $$\operatorname{sign}[Z(-2\alpha)]=(-1)^{k-1-\alpha},$$ and for a constant $\tilde{C}_R:=\max\big(\log 2,{\begin{matrix}\frac{C_R}{2}\end{matrix}}\big)$ the estimates $$\big\vert Z(-2\alpha)\big\vert\leq C(2\pi)^{-2\alpha}(2\alpha)!$$ hold, and we define a new sequence of positive numbers $B(k)$ by $$\big\vert\zeta_{(D^2,S^{2k-1})}^{\,\prime}(0)\big\vert\leq B(k):=\frac{\tilde{C}_R2^{k+1}}{(2k-2)!}(-1)^{k-1}\sum_{\alpha=0}^{k-1}e_{\alpha,k}(-1)^\alpha(2\pi)^{-2\alpha}(2\alpha)!.$$ Again we find recursion relations for the relevant polynomials $$e_{\alpha,k+1}=-(k-{\begin{matrix}\frac{1}{2}\end{matrix}})^2e_{\alpha,k}+e_{\alpha-1,k},$$ which gives recursive estimates on the $B(k)$ $$B(k+1)\leq\bigg\{\frac{1}{2}+\frac{1}{2\pi^2}\bigg\}B(k)\leq\frac{5}{9}B(k),$$ and this concludes the proof of the theorem.
Laplace operator sphere determinants {#sec:Laplace}
====================================
Following [@Dow1] we investigate Laplacians of the form $$L_{\alpha}=\Delta +\frac{n-1}{4n}R-\alpha_n^2,$$ where $R$ is the scalar curvature (of $S^n$) and the $\alpha_n\in{\mathbb{R}}$ are constants, but may depend on the dimension $n$. We restrict here to the case $0\leq \alpha_n\leq \frac{n-1}{2}$. Recalling now that $$R(S^n_{\mathrm{can}})=n(n-1),$$ we see that $\alpha_n=\frac{n-1}{2}$ gives the ordinary Laplacian, while $\alpha_n=\frac{1}{2}$ corresponds to the Yamabe operator in each dimension. Note that there is no kernel of the operator $L$ in dimension $n$ if $\alpha_n<\frac{n-1}{2}$, and that for $\alpha_n=\frac{n-1}{2}$ the kernel is the constant functions on $S^n$.
The theorem we will prove in this section is the following.
\[LaplaceTheorem\] On the standard (radius of unity) spheres we have the limits $$\lim_{k\to\infty}\det(\Delta, S^{2k+1}_{\mathrm{can}})=0.$$ and $$\lim_{n\to\infty}\det(Y, S^n_{\mathrm{can}})=1$$ In fact, we can display the asymptotics as (for $n=2k+1$) $$\zeta'_{\Delta, S_\mathrm{can}^{n}}(0)=\log n+O\Big(\frac{\log\log n}{\log n}\Big),$$ and (for any $n$) $$\zeta'_{Y, S^{n}_{\mathrm{can}}}(0)=O\big(2^{-n}\big).$$
The proof of the claimed limit for the conformal Laplacian is easily carried out analogously to the Dirac case. However we give a different type of proof, which works for the ordinary Laplacian as well, and is more suitable for exposing the significance of the coupling constant and kernel of the operator.
From [@Dow1] and [@Dow2] we get the corresponding zeta functions and their first derivatives at zero $$\label{DowkerFormula}
\begin{split}
\zeta_L'(0)=&\zeta_n'(0,a_--\alpha_n)+\zeta_n'(0,a_+-\alpha_n)+\zeta_n'(0,a_-+\alpha_n)+\zeta_n'(0,a_++\alpha_n)\\
&[+\ln(n-1)]\:-\sum_{j=1}^{\lfloor\frac{n}{2}\rfloor}\frac{\alpha_n^{2j}}{j}N_{2j}(n)\sum_{i=0}^{j-1}\frac{1}{2i+1}.
\end{split}$$ In the notation here, the numbers $a_\pm:=(n\pm 1)/2$ originate in the contributions from Dirichlet ($a_+$) and Neumann ($a_-$) boundary conditions on hemispheres. For the ordinary Laplacian, the four functions $a_{\pm}\pm\alpha_n$ are, by order of appearance in the above equation, $(0,1,n-1,n)$. For the Yamabe operator it reads $({\begin{matrix}\frac{n}{2}\end{matrix}}-1,{\begin{matrix}\frac{n}{2}\end{matrix}},{\begin{matrix}\frac{n}{2}\end{matrix}},{\begin{matrix}\frac{n}{2}\end{matrix}}+1)$.
The notation “$[+\ln(n-1)]$” means that this term is only present if $L_{\alpha}$ has non-trivial kernel. The numbers $N_{2j}(n)$ are defined using the following polynomials, namely let $$\prod_{p=1}^{n-2}\Big(x+\frac{n-1}{2}-p\Big)=\sum_{r=0}^{n-2}x^r\tilde{N}_r(n).$$ Then $$N_{2j}(n):=\frac{2}{(n-1)!}\tilde{N}_{2j-2}(n).$$ Finally the zeta functions appearing here are (special cases of) Barnes zeta functions, where for $a\in{\mathbb{R}}$ $$\label{Barnes}
\zeta_n(s,a)=\sum_{m=0}^{\infty}\binom{m+n-1}{n-1}\frac{1}{(a+m)^s},\quad{\mathrm{Re}}{s}>n,$$ Note that here, if $a=0$, it is implicit that the term $m=0$ is omitted.
The meromorphic continuation is carried out using the contour integrals $$\label{BarnesIntegral}
\zeta_n(s,a)=\frac{i\Gamma(1-s)}{2\pi}\int_H\frac{e^{az}}{(1-e^{z})^n}z^{s-1}dz,$$ where $H$ is a left Hankel contour, and $z^{s-1}$ is defined using the negative real axis branch cut of the logarithm enclosed by $H$. The contours will be chosen conveniently for getting estimates of the Barnes functions. Deformations of the curves must of course respect the branch cut and the possible poles at $2\pi i k,k\in{\mathbb{Z}}$.
We note a few general features of the Barnes zeta functions. Each contour will be taken as the positively oriented boundary of some box $(-\infty,r_n]\times[-{\begin{matrix}\frac{i\pi}{2}\end{matrix}},{\begin{matrix}\frac{i\pi}{2}\end{matrix}}]$, where $r_n$ is a real-valued function of the dimension $n$. We denote such a contour by $\gamma_n$ and decompose in the obvious way into one vertical and two horizontal components $h_n^+\cup v_n\cup h_n^-$. Now, since $$\label{NormofDenom}
\big\vert 1-e^{z}\big\vert^n=\big(1+e^{2x}-2e^{x}\cos(y)\big)^{\frac{n}{2}}$$ is the norm of the denominator at $z=x+iy$, we always have $$\label{NormOnPiHalves}
\big\vert 1-e^{z}\big\vert^n=\big(1+e^{2x}\big)^{\frac{n}{2}}$$ on the horizontal parts of the contours. From (\[BarnesIntegral\]) we see, given that $a\leq n$, $$\label{HorizontalParts}
n^k\Bigg\vert\frac{e^{az}}{(1-e^z)^n}\Bigg\vert\leq n^k\big(1+e^{-2x}\big)^{-{\begin{matrix}\frac{n}{2}\end{matrix}}}\to 0,\quad\text{for}\quad n\to\infty,$$ for fixed $z$ and $k$, i.e. pointwise convergence to zero of the integrands. Lebesgue dominated convergence theorem can therefore in most case be used to prove that the contribution from a horizontal piece is $O(n^{-\infty})$ as $n\to\infty$, meaning by definition $O(n^{-k})$ as $n\to\infty$, for any $k$.
When $a>0$, the derivatives at $s=0$ are, using (\[GammaDiff\]) $$\label{Derivatives}
\zeta_n'(0,a_n)=\frac{i}{2\pi}\int_{\gamma_n}\frac{e^{a_nz}}{(1-e^{z})^n}\frac{\log z+\gamma}{z}dz.$$
If $a_n=0,1,{\begin{matrix}\frac{n-1}{2}\end{matrix}}$ or ${\begin{matrix}\frac{n}{2}\end{matrix}}$, then $$\zeta_n'(0,a_n)\to 0,\quad\textrm{as}\quad n\to\infty,$$ in fact it is $O(n^{-\infty})$ as $n\to\infty$.
For dealing with these cases we use the fixed contour consisting of the positively oriented boundary of the box $(-\infty,\log
4]\times[-{\begin{matrix}\frac{i\pi}{2}\end{matrix}},{\begin{matrix}\frac{i\pi}{2}\end{matrix}}]$. From the general remarks following (\[HorizontalParts\]), the horizontal parts of the contours all contribute $O(n^{-\infty})$ for $a_n>0$, by Lebesgue dominated convergence.
In the case $a_n=0$ however, note that the continuation of the integral doesn’t work directly around $s=0$. From (\[Barnes\]) we deduce, by straightforward resummation, a family of functional equations $$\label{BarnesFuncEq}
\zeta_n(s,0)=\zeta_n(s,1)+(n-1)\zeta_n(s+1,1),$$ thus allowing again for meromorphic continuation to ${\mathbb{C}}$ by integral representations. Taking into account the front factor of $\Gamma(1-s)$ in (\[BarnesIntegral\]) and using $$\Gamma(1-s)=\frac{1}{1-s}-\gamma+\ldots$$ near $s=1$, shows $$\label{DerivaZero}
\zeta'_n(s,0)=\zeta'_n(0,1)+\frac{n-1}{2\pi i}\int_H\frac{e^z}{(1-e^z)^n}\big(\gamma\log z+{\begin{matrix}\frac{1}{2}\end{matrix}}\log^2z\big)dz.$$ On the vertical line segments $v_n$ we use $\vert1-e^z\vert\geq3$ and estimate $$\begin{split}
&(n-1)\bigg\vert\int_{v_n}\frac{e^z}{(1-e^z)^n}P(z)\;dz\bigg\vert\leq C_P(n-1)\:3^{-n}\to 0,\quad\mathrm{as}\quad n\to\infty,
\end{split}$$ where $P$ is any polynomial in $\frac{1}{z}$ and $\log z$, as in the expressions appearing in (\[Derivatives\]) and (\[DerivaZero\]).
In the remaining cases $a_n={\begin{matrix}\frac{n-1}{2}\end{matrix}}, {\begin{matrix}\frac{n}{2}\end{matrix}}$ we focus for the sake of argument on the latter, i.e. $$\int_H\frac{1}{\big(e^{z/2}-e^{-z/2}\big)^n}\frac{\log z+\gamma}{z}dz,$$ Then using $\big\vert e^{z/2}-e^{-z/2}\big\vert^2=e^x+e^{-x}-2\cos(y)$ we see that indeed, by Lebesgue dominated convergence, there is rapid convergence to zero, namely $$\zeta_{(Y,S^n)}'(0)= O\big(2^{-n}\big),$$ concluding the proof.
The next proposition deals with the most complicated of the terms, with parameters $a_n=n-1$ and $n$. Here there is no way to deform the Hankel contour in the complex plane, respecting the poles and branch cut of $\log z$, so as to obtain pointwise convergence to zero of the integrand as $n\to 0$ everywhere along the curve. This follows from (\[NormofDenom\]), since it would imply $x\leq 0$ at some point of the curve, which would thus intersect the logarithmic branch cut.
The following lemma essentially gives a suitable hybrid of Laplace’s method and the method of stationary phase, applicable to the cases needed here.
\[stationaryphaselemma\] Assume $\varphi:[a,b)\to{\mathbb{R}}_{\geq 0}$ for $0\leq a< b<\infty$ satisfies
- $\varphi$ is continuous and bounded,
- $x\mapsto e^x\cdot\varphi(x)$ is increasing on $[a,b)$. Then, writing $s_b:=\tan^{-1}(e^{-b})$, $$\Bigg\vert\int_a^b\big(ie^{-x}-1\big)^{-n}\varphi(x)dx\Bigg\vert\leq \frac{2\pi}{n\tan s_b}\max_{s_b\leq x\leq s_b+{\begin{matrix}\frac{\pi}{n}\end{matrix}}}\varphi\big(\log\big({\begin{matrix}\frac{1}{\tan x}\end{matrix}}\big)\big)+O(n^{-\infty}),$$ where the error term is uniform in $b$, but not generally in $a$.
<!-- -->
- Note that in applying this lemma, we will let $b$ depend on $n$, while $a$ will be fixed.
- If $\varphi$ is $C^1$, the second condition in the lemma is equivalent to $\varphi'\geq-\varphi$.
\[nProposition\] If $a_n=n-1$ or $n$, then $$\zeta_n'(0,a_n)\to 0,\quad\text{for}\quad n\to\infty.$$
For proving this, we assume $n\geq 4$ and shift for each $n$ the contour to the positively oriented boundary of the box $(-\infty,\log n]\times[-{\begin{matrix}\frac{i\pi}{2}\end{matrix}},{\begin{matrix}\frac{i\pi}{2}\end{matrix}}]$, which still encloses only the singularity at $z=0$.
The contributions from $h^{\pm}_n$ with, say, ${\mathrm{Re}}z\leq \log 4$, are again $O(n^{-\infty})$ by (\[HorizontalParts\]), while on the moving right hand edges $v_n$, we have the estimates $$\big\vert e^{-z}-1\big\vert^{-n}=(1+{\begin{matrix}\frac{1}{n^2}\end{matrix}}-{\begin{matrix}\frac{2}{n}\end{matrix}}\cos(y))^{-{\begin{matrix}\frac{n}{2}\end{matrix}}}\leq\big(1-{\begin{matrix}\frac{1}{n}\end{matrix}}\big)^{-n},$$ yielding, for some constant $C>0$ $$\bigg\vert\int_{v_n}\frac{1}{(e^{-z}-1)^n}\frac{\log z+\gamma}{z}dz\bigg\vert\leq C\frac{\log\log n}{\log n}.$$ Thus the contribution from this part tends to zero as $n\to\infty$, and similarly for $a_n=n-1$, again with or without the $\log z$ present, as needed for the terms in (\[Derivatives\]).
For the contributions from the right half-plane part of the horizontals, denoted by $h_{R,n}^{\pm}$, in the cases $a_n=n-1,n$, any finite piece gives an $O(n^{-\infty})$ contribution as $n\to\infty$.
To deal with the case $a_n=n$, we calculate explicitly, $$\begin{split}
&\frac{i}{2\pi}\int_{h_{R,n}^{+}\cup h_{R,n}^{-}}\big(e^{-z}-1\big)^{-n}\frac{\log z+\gamma}{z}dz=\\
&-\frac{1}{\pi}\:{\mathrm{Im}}\int_{0}^{\log n}\big(ie^{-x}-1\big)^{-n}\frac{\big(x-i\pi/2\big)\Big(\log\sqrt{x^2+(\frac{\pi}{2})^2}+i\tan^{-1}({\begin{matrix}\frac{\pi}{2x}\end{matrix}})+\gamma\Big)}{x^2+({\begin{matrix}\frac{\pi}{2}\end{matrix}})^2}dx.
\end{split}$$ To take the oscillation into account, we repeatedly apply Lemma \[stationaryphaselemma\] to the expression, with $s_b={\begin{matrix}\frac{1}{n}\end{matrix}}$, considering the product terms separately. For each term the relevant positive function $\varphi$ is decreasing for $x$ large (i.e. for $n$ large, since finite pieces contribute $O(n^{-\infty})$), so the maximum is evaluation at $s_b+\frac{\pi}{n}$. For example one term is analyzed as follows $$\varphi_1(x):=\frac{x\log\sqrt{x^2+(\frac{\pi}{2})^2}}{x^2+(\frac{\pi}{2})^2},\quad x\in[1,\log n),$$ $$\begin{split}
&\Bigg\vert\int_{0}^{\log n}\big(ie^{-x}-1\big)^{-n}\varphi_1(x)dx\Bigg\vert\leq2\pi\frac{{\begin{matrix}\frac{1}{n}\end{matrix}}}{\tan{\begin{matrix}\frac{1}{n}\end{matrix}}}\varphi_1\Big(\log\big({\begin{matrix}\frac{1}{\tan(\frac{1+\pi}{n})}\end{matrix}}\big)\Big)+O(n^{-\infty}),
\end{split}$$ so that by Lemma \[stationaryphaselemma\], the righthand side converges to zero, as the remaining terms can similarly be shown to do.
In the case $a_n=n-1$, Lemma \[stationaryphaselemma\] is not needed, since after a partial integration it is easily seen that $$\int_{{\begin{matrix}\frac{i\pi}{2}\end{matrix}}+[0,\log n]}\big(e^{-z}-1\big)^{-n}e^{-z}\frac{\log
z+\gamma}{z}dz=O(n^{-1})\quad\text{as}\quad n\to\infty,$$ which ends the proof of Proposition \[nProposition\].
Note that the use of Lemma \[stationaryphaselemma\] in the proof was essential, since without the oscillating factor, amounting approximately to $\sin(ne^{-x})$, we would have $$\begin{split}
\frac{1}{2}\big(\log(\log n)\big)^2\geq \int_{\log\sqrt{n}}^{\log n}\big(1+e^{-2x}\big)^{-\frac{n+1}{2}}\frac{\log x}{x}dx\geq\frac{\log 2}{4}\Big\{\log(\log n)-\frac{1}{2}\log 2\Big\},
\end{split}$$ and thus convergence to $+\infty$, at a rate between $\log\log n$ and $\log^2\log n$.
We need to deal with the last term in (\[DowkerFormula\]), which is controlled as follows.
If $n$ is odd, then $$\sum_{j=1}^{\lfloor\frac{n}{2}\rfloor}\frac{\alpha_n^{2j}}{j}N_{2j}(n)\sum_{i=0}^{j-1}\frac{1}{2i+1}=0.$$ Setting $\alpha_n={\begin{matrix}\frac{1}{2}\end{matrix}}$, then for $n$ of any parity $$\lim_{n\to\infty}\sum_{j=1}^{\lfloor\frac{n}{2}\rfloor}\frac{\alpha_n^{2j}}{j}N_{2j}(n)\sum_{i=0}^{j-1}\frac{1}{2i+1}=0,$$ being $O(2^{-n})$ as $n\to\infty$.
Interestingly enough, this term is exactly the “correction term” from a product formula of certain determinants discussed in [@Dow1], and this makes it particularly interesting to see that this tends to zero. The convergence is exponential and this means that the anomaly, in this situation quickly evaporates as $n\to\infty$.
If $n=2k+1$ is odd we exploit symmetry and rewrite the left-hand-side in the defining equation for $\tilde{N}_{r}(n)$ $$x\prod_{p=1}^{k-1}\Big(x^2-p^2\Big)=\sum_{r=0}^{2k-1}\tilde{N}_{r}(n)x^{r},$$ showing the vanishing of all terms with even $r$.
If $n=2k$ is even we again use symmetry to write $$\prod_{p=1}^{k-1}\Big(x^2-(p-{\begin{matrix}\frac{1}{2}\end{matrix}})^2\Big)=\sum_{r=0}^{k-1}\tilde{N}_{2r}(2k)x^{2r}.$$ This shows that $\operatorname{sign}[\tilde{N}_{2r-2}(2k)]=(-1)^{k+r}$ and gives as usual a recurrence relation $$\tilde{N}_{2r-2}(2k+2)=\tilde{N}_{2r-4}(2k)-(k-{\begin{matrix}\frac{1}{2}\end{matrix}})^2\tilde{N}_{2r-2}(2k).$$ We estimate $$\bigg\vert\sum_{j=1}^{k}\frac{\alpha_{2k}^{2j}}{j}N_{2j}(2k)\sum_{i=0}^{j-1}\frac{1}{2i+1}\bigg\vert\leq
2D(k),\quad D(k):=\sum_{j=1}^{k}\alpha_{2k}^{2j}(-1)^{k+j}\frac{\tilde{N}_{2j-2}(2k)}{(2k-1)!}.$$ Letting $\alpha_n={\begin{matrix}\frac{1}{2}\end{matrix}}$ we find the recursive estimates $$\begin{split}
D(k+1)&=\sum_{j=1}^{k+1}\big({\begin{matrix}\frac{1}{2}\end{matrix}}\big)^{2j}(-1)^{k+j}\frac{\tilde{N}_{2j-2}(2k+2)}{(2k+1)!}\\
&=\frac{1}{2k(2k+1)}\bigg\{\frac{1}{4}+(k+{\begin{matrix}\frac{1}{2}\end{matrix}})^2\bigg\}D(k)\leq \frac{1}{2}D(k),
\end{split}$$ for sufficiently large $k$, proving exponential convergence to zero and ending the proof.
Determinants on rescaled spheres {#sec:rescaled}
================================
It may be argued that taking the spheres with the radius unity standard metrics is not as natural for the problem of determinants, and that rescaling to unit volume is more appropriate. For the conformally covariant Yamabe and Dirac operators, this is of course produces no change, and focus remains on the ordinary Laplacian. Here the rescaling indeed cancels the leading term $\log n$ in the zeta derivative at zero.
If we rescale the metric $g$ by the constant $\lambda>0$ to $\tilde{g}:=\lambda^2g$ the determinant of the ordinary Laplacian changes as $$\label{DetChange}
\zeta'_\lambda(0)=\zeta'(0)+2\log\lambda\cdot\zeta(0)=\zeta'(0)-2\log\lambda,$$ where the last equality is due to $\zeta(0)=-1$. To rescale the spheres we use $$\lambda(n)=\big(\operatorname{vol}(S^n)\big)^{-\frac{1}{n}}=\Bigg(\frac{2\pi^{\frac{n+1}{2}}}{\Gamma(\frac{n+1}{2})}\Bigg)^{-\frac{1}{n}},$$ which with Stirling’s formula $$\sqrt{2\pi}n^{n+{\begin{matrix}\frac{1}{2}\end{matrix}}}e^{-n+{\begin{matrix}\frac{1}{12n+1}\end{matrix}}}<n!<\sqrt{2\pi}n^{n+{\begin{matrix}\frac{1}{2}\end{matrix}}}e^{-n+{\begin{matrix}\frac{1}{12n}\end{matrix}}}$$ gives $$\begin{split}
&\log\lambda(n)\\
&\quad=-\frac{1}{n}\Big\{\frac{n+1}{2}\log\pi-{\begin{matrix}\frac{1}{2}\end{matrix}}\log{\begin{matrix}\frac{\pi}{2}\end{matrix}}-\frac{n}{2}\log\frac{n-1}{2}+\frac{n-1}{2}+\frac{1}{6(n-1)}\Big\}+O(n^{-3})\\
&\quad=\frac{1}{2}\log n-\frac{1}{2}(1+\log(2\pi))+O(n^{-1}).
\end{split}$$ Inserting in (\[DetChange\]) gives $$\zeta'_{\mathrm{rescaled}}(0)=\zeta'(0)-\log n+1+\log(2\pi)+O(n^{-1}),$$ thus cancelling the leading order term $\log n$ from the case of radius unity spheres. By the above along with Theorem \[LaplaceTheorem\], we find the new limit and asymptotics.
On the standard spheres rescaled to unit volume, we have the limit $$\lim_{k\to\infty}\det(\Delta, S_\mathrm{rescaled}^{2k+1})=\frac{1}{2\pi e}.$$ In fact the asymptotics can be displayed as (for $n=2k+1$) $$\zeta'_{\Delta, S^{n}}(0)=1+\log(2\pi)+O\Big(\frac{\log\log n}{\log n}\Big).$$
Proof of a stationary phase lemma {#sec:proof}
=================================
Let $s_a:=\tan^{-1}(e^{-a}), s_b:=\tan^{-1}(e^{-b})$ and compute $$\label{LemmaCompute}
\begin{split}
&\int_a^b\sin\big(n\tan^{-1}(e^{-x})\big)\big(1+e^{-2x}\big)^{-{\begin{matrix}\frac{n}{2}\end{matrix}}}\varphi(x)dx\\
&=\int_{s_b}^{s_a}\sin\big(nx\big)\big(1+\tan^{2}x\big)^{-{\begin{matrix}\frac{n-2}{2}\end{matrix}}}\frac{\varphi\big(\log\big(\frac{1}{\tan x}\big)\big)}{\tan x}dx\\
&=\sum_{k=1}^{n_{ab}}\int_{s_b+{\begin{matrix}\frac{(k-1)\pi}{n}\end{matrix}}}^{s_b+{\begin{matrix}\frac{k\pi}{n}\end{matrix}}}\sin\big(nx\big)\big(1+\tan^{2}x\big)^{-{\begin{matrix}\frac{n-2}{2}\end{matrix}}}\frac{\varphi\big(\log\big(\frac{1}{\tan x}\big)\big)}{\tan x}dx\\
&+\int_{s_b+\frac{n_{ab}\pi}{n}}^{s_a}\sin\big(nx\big)\big(1+\tan^{2}x\big)^{-{\begin{matrix}\frac{n-2}{2}\end{matrix}}}\frac{\varphi\big(\log\big(\frac{1}{\tan x}\big)\big)}{\tan x}dx,
\end{split}$$ where $n_{ab}:=\Big\lfloor \frac{n}{\pi}(s_a-s_b)\Big\rfloor$. Thus the last term above can be estimated by $$\begin{split}
&\int_{s_b+\frac{n_{ab}\pi}{n}}^{s_a}\Bigg\vert\big(1+\tan^{2}x\big)^{-{\begin{matrix}\frac{n-2}{2}\end{matrix}}}\frac{\varphi\big(\log\big(\frac{1}{\tan x}\big)\big)}{\tan x}\Bigg\vert dx\leq C_{\varphi,a}\Big(1+\tan^2\big(s_a-\frac{\pi}{n}\big)\Big)^{-{\begin{matrix}\frac{n-2}{2}\end{matrix}}},
\end{split}$$ when $n\geq{\begin{matrix}\frac{\pi}{s_a}\end{matrix}}$, and is $O(n^{-\infty})$ as $n\to\infty$, for any $k$. Note as claimed, that the error term is only uniform in the parameter $b$. The first term in (\[LemmaCompute\]) is written as a sum over the half-periods of the sine function. Fixing $n$, the alternating behavior gives the estimate $$\begin{split}
&\Bigg\vert\sum_{k=1}^{n_{ab}}\int_{s_b+{\begin{matrix}\frac{(k-1)\pi}{n}\end{matrix}}}^{s_b+{\begin{matrix}\frac{k\pi}{n}\end{matrix}}}\sin\big(nx\big)\big(1+\tan^{2}x\big)^{-{\begin{matrix}\frac{n-1}{2}\end{matrix}}}\frac{\varphi\big(\log\big(\frac{1}{\tan x}\big)\big)}{\tan x}dx\Bigg\vert\\
&\leq\int_{s_b}^{s_b+{\begin{matrix}\frac{\pi}{n}\end{matrix}}}\big\vert\sin\big(nx\big)\big\vert\big(1+\tan^{2}x\big)^{-{\begin{matrix}\frac{n-1}{2}\end{matrix}}}\frac{\varphi\big(\log\big(\frac{1}{\tan x}\big)\big)}{\tan x}dx\\
&\leq\frac{{\begin{matrix}\frac{\pi}{n}\end{matrix}}}{\tan s_b}\max_{s_b\leq x\leq s_b+{\begin{matrix}\frac{\pi}{n}\end{matrix}}}\varphi\big(\log\big({\begin{matrix}\frac{1}{\tan x}\end{matrix}}\big)\big).
\end{split}$$ Similar computations apply for the contribution from the real part of the integral.
The author would like to thank C. Bär and S. Schopka for raising the question of limits in their paper [@BS]. Thanks goes to Kate Okikiolu, for useful criticism of the manuscript, and to Department of Mathematics, University of Pennsylvania, Philadelphia for having the author as visiting graduate student during the creation of this paper.
The work is partly supported by my Elite Research Scholarship 2006, from The Danish Ministry of Science, Technology and Innovation.
[99]{}
C. Bär and S. Schopka, *The Dirac determinant of spherical space forms*, Geo. Anal. and Nonlinear PDEs, 39–67, Springer, Berlin, 2003.
T. P. Branson, *The functional determinant*, Lecture Notes Series **4**, Seoul National University, Global Analysis Research Center, Seoul, 1993.
J. S. Dowker, *Effective action in spherical domains*, Commun. Math. Phys. **162** (1994), 633–647.
J. S. Dowker, *Functional determinants on spheres and sectors*, J. Math. Phys. **35** (1994), 4989–4999.
J.S. Dowker, K. Kirsten, *The Barnes $\zeta$-function, sphere determinants and Glaisher-Kinkelin-Bendersky constants*, Anal. Appl. (Singap.) **3** (2005), no. 1, 45–68.
S. W. Hawking, *Zeta function regularization of path integrals in curved space time*, Commun. Math. Phys. **55** (1977), 133–148.
K. Kirsten, *Spectral functions in mathematics and physics*, Chapman & Hall/CRC 2002.
N. M. Møller, *Extremals of spectral determinants for squared dirac operators, and Branson’s conjecture*, arXiv: 0705.3857.
F. W. J. Olver, *Asymptotics and special functions*, Academic Press, New York-London, 1974.
D. B. Ray, I. M. Singer, *$R$-torsion and the Laplacian on Riemannian manifolds*, Advances in Math. **7** (1971), 145–210.
D. B. Ray, I. M. Singer, *Analytic torsion for complex manifolds*, Ann. of Math. (2) **98** (1973), 154–177.
I. Vardi, *Determinants of Laplacians and multiple gamma functions*, SIAM J. Math. Anal. **19** (1988), no. 2, 493–507.
W. I. Weisberger, *Conformal invariants for determinants of Laplacians on Riemann surfaces*, Comm. Math. Phys. **112** (1987), no. 4, 633–638.
W. I. Weisberger, *Normalization of the path integral measure and the coupling constants for bosonic strings*, Nuclear Phys. B **284** (1987), no. 1, 171–200.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Even professional baseball players occasionally find it difficult to gracefully approach seemingly routine pop-ups. This paper describes a set of towering pop-ups with trajectories that exhibit cusps and loops near the apex. For a normal fly ball, the horizontal velocity is continuously decreasing due to drag caused by air resistance. But for pop-ups, the Magnus force (the force due to the ball spinning in a moving airflow) is larger than the drag force. In these cases the horizontal velocity decreases in the beginning, like a normal fly ball, but after the apex, the Magnus force accelerates the horizontal motion. We refer to this class of pop-ups as paradoxical because they appear to misinform the typically robust optical control strategies used by fielders and lead to systematic vacillation in running paths, especially when a trajectory terminates near the fielder. In short, some of the dancing around when infielders pursue pop-ups can be well explained as a combination of bizarre trajectories and misguidance by the normally reliable optical control strategy, rather than apparent fielder error. Former major league infielders confirm that our model agrees with their experiences.'
author:
- 'Michael K. McBeath'
- 'Alan M. Nathan'
- 'A. Terry Bahill'
- 'David G. Baldwin'
title: 'Paradoxical pop-ups: Why are they hard to catch?'
---
Introduction {#sec:intro}
=============
Baseball has a rich tradition of misjudged pop-ups. For example, in April, 1961, Roy Sievers of the Chicago White Sox hit a towering pop-up above Kansas City Athletics’ third baseman Andy Carey who fell backward in trying to make the catch. The ball landed several feet from third base, far out of the reach of Carey. It rolled into the outfield, and Sievers wound up on second with a double.
A few other well-known misplays of pop-ups include: New York Giants’ first baseman Fred Merkle’s failure to catch a foul pop-up in the final game of the 1912 World Series, costing the Giants the series against the Boston Red Sox; St. Louis first baseman Jack Clark’s botched foul pop-up in the sixth game of the 1985 World Series against Kansas City; and White Sox third baseman Bill Melton’s broken nose suffered in an attempt to catch a “routine” pop-up in 1970.
As seen by these examples, even experienced major league baseball players can find it difficult to position themselves to catch pop-ups hit very high over the infield. Players describe these batted balls as “tricky” or “deceptive,” and at times they will be seen lunging for the ball in the last instant of the ball’s descent. “Pop-ups look easy to anyone who hasn’t tried to catch one - like a routine fly ball that you don’t have to run for,” Clete Boyer said, “but they are difficult to judge and can really make you look like an idiot.” Boyer, a veteran of sixteen years in the major leagues, was considered one of the best defensive infielders in baseball.
Several factors can exacerbate the infielder’s problem of positioning himself for a pop-up. Wind currents high above the infield can change the trajectory of the pop-up radically. Also, during day games the sky might provide little contrast as a background for the ball–a condition called a “high sky” by players. Then, there are obstacles on the field–bases, the pitcher’s mound, and teammates–that can hinder the infielder trying to make a catch. But even on a calm night with no obstacles nearby, players might stagger in their efforts to get to the ball.
The frequency of pop-ups in the major leagues–an average of nearly five pop-ups per game–is great enough that teams provide considerable pop-up practice for infielders and catchers. Yet, this practice appears to be severely limited in increasing the skill of these players. Infielders seem unable to reach the level of competency in catching “sky-high” pop-ups that outfielders attain in catching high fly balls, for example. This suggests that the technique commonly used to catch pop-ups might be the factor limiting improvement.
Almost all baseball players learn to catch low, “humpback” pop-ups and fly balls before they have any experience in catching lofty pop-ups. In youth leagues nearly all pop-ups have low velocities and few exceed a height of fifty feet; therefore, they have trajectories that are nearly parabolic. Fly balls, too, have near-parabolic trajectories. Young players develop techniques for tracking low pop-ups and fly balls. If 120-foot pop-ups do not follow similar trajectories, however, major league infielders might find pop-ups are hard to catch because the tracking and navigation method they have learned in their early years is unreliable for high, major league pop-ups.
In the consideration of this hypothesis, we first describe trajectories of a set of prototypical batted balls, using models of the bat-ball collision and ball flight aerodynamics. We then develop models of three specific kinds of typical non-parabolic pop-up trajectories. These “paradoxical” trajectories exhibit unexpected behavior around their apices, including cusps and loops. Several of these paradoxical trajectories are fitted with an optical control model that has been used successfully to describe how players track and navigate to fly balls. For each fit, a prediction of the behavior of infielders attempting to position themselves to catch high pop-ups is compared with the observed behavior of players during games.
Simulations of batted-ball trajectories {#sec:simulations}
=======================================
Forces on a spinning baseball in flight
---------------------------------------
As every student in an introductory physics course learns, the trajectory of a fly ball in a vacuum is a smooth symmetric parabola since the only force acting on it is the downward pull of gravity. However, in the atmosphere the ball is subject to additional forces, shown schematically in Fig. \[fig:forces\]: the retarding force of drag ($F_D$) and the Magnus force ($F_M$). The Magnus force was first mentioned in the scientific literature by none other than a young Isaac Newton in his treatise on the theory of light,[@newton1671] where he included a brief description on the curved trajectory of a spinning tennis ball. Whereas the drag force always acts opposite to the instantaneous direction of motion, the Magnus force is normal to both the velocity and spin vectors. For a typical fly ball to the outfield, the drag force causes the trajectory to be somewhat asymmetric, with the falling angle steeper than the rising angle,[@adair02] although the trajectory is still smooth. If the ball has backspin, as expected for such fly balls, the Magnus force is primarily in the upward direction, resulting in a higher–but still quite smooth–trajectory. However, as we will show the situation is qualitatively very different for a pop-up, since a ball-bat collision resulting in a pop-up will have a considerable backspin, resulting in a significantly larger Magnus force than for a fly ball. Moreover, the direction of the force is primarily horizontal with a sign that is opposite on the upward and downward paths. These conditions will result in unusual trajectories–sometimes with cusps, sometimes with loops–that we label as “paradoxical.”
With this brief introduction, we next discuss our simulations of baseball trajectories in which a model for the ball-bat collision (Sec. \[sec:collision\]) is combined with a model for the drag and Magnus forces (Sec. \[sec:aero\]) to produce the batted-ball trajectories. We discuss the paradoxical nature of these trajectories in Sec \[sec:trajs\] in light of the interplay among the various forces acting on the ball.
Ball-bat collision model {#sec:collision}
------------------------
The collision model is identical to that used both by Sawicki[@sawicki03] and by Cross and Nathan.[@cross06] The geometry of the collision is shown in Fig. \[fig:geom\]. A standard baseball ($r_{ball}$=1.43 inch, mass=5.1 oz) approaches the bat with an initial speed $v_{ball}$=85 mph, initial backspin $\omega_i$=126 rad/s (1200 rpm), and at a downward angle of 8.6$^\circ$ (not shown in the figure). The bat has an initial velocity $v_{bat}$=55 mph at the point of impact and an initial upward angle of 8.6$^\circ$, identical to the downward angle of the ball. The bat was a 34-inch long, 32-oz wood bat with an R161 profile, with radius $r_{bat}$=1.26 inch at the impact point. If lines passing through the center of the ball and bat are drawn parallel to the initial velocity vectors, then those lines are offset by the distance $D$. Simply stated, $D$ is the amount by which the bat undercuts ($D>0$) or overcuts ($D<0$) the ball. In the absence of initial spin on the baseball, a head-on collision ($D=0$) results in the ball leaving the bat at an upward angle of 8.6$^\circ$ and with no spin; undercutting the ball produces backspin and a larger upward angle; overcutting the ball produces topspin and a smaller upward or even a downward angle. The ball-bat collision is characterized by two constants, the normal and tangential coefficients of restitution–$e_N$ and $e_T$, respectively–with the additional assumption that angular momentum is conserved about the initial contact point between the ball and bat.[@cross06] For $e_N$, we use the parameterization e\_N = 0.54 - (v\_N-60)/895 , where $v_N=(v_{ball}+v_{bat})\cos\theta$ is the normal component of the relative ball-bat velocity in units of mph.[@sawicki03] We further assume $e_T=0$, which is equivalent to assuming that the tangential component of the relative ball-bat surface velocity, initially equal to $(v_{ball}+v_{bat})\sin\theta+r_{ball}\omega_i$, is identically zero as the ball leaves the bat, implying that the ball leaves the bat in a rolling motion. The loss of tangential velocity occurs as a result of sliding friction, and it was verified by direct calculation that the assumed coefficient of friction of 0.55[@cross06] is sufficient to bring the tangential motion to a halt prior to the end of the collision for all values of $D<1.7$ inches. Given the initial velocities and our assumptions about $e_N$ and $e_T$, the outgoing velocity $v$, angle $\theta$ , and backspin of the baseball can be calculated as a function of the offset $D$. These parameters, which are shown in Fig. \[fig:d\], along with the initial height of 3 ft., serve as input into the calculation of the batted-ball trajectory. Note particularly that both $\omega$ and $\theta$ are strong functions of $D$, whereas $v$ only weakly depends on $D$.
Baseball aerodynamics model {#sec:aero}
---------------------------
The trajectory of the batted baseball is calculated by numerically solving the differential equations of motion using a fourth-order Runge-Kutta technique, given the initial conditions and the forces. Conventionally, drag and Magnus forces are written as &=&-C\_DAv\^2\
&=&C\_LAv\^2 () , \[eq:drag\] where $\rho$ is the air density (0.077 lb/ft$^3$), $A$ is the cross sectional area of the ball (6.45 inch$^2$), $v$ is the velocity, $\omega$ is the angular velocity, and $ C_D$ and $C_L$ are phenomenological drag and lift coefficients, respectively. Note that the direction of the drag is opposite to the direction of motion whereas the direction of the Magnus force is determined by a right-hand rule. We utilize the parametrizations of Sawicki et al.[@sawicki03] in which $C_D$ is a function of the speed $v$ and $C_L$ is a bilinear function of spin parameter $S=r_{ball}/v$, implying that $F_M$ is proportional to $\omega v$. Since the velocity of the ball does not remain constant during the trajectory, it is necessary to recompute $C_D$ and $C_L$ at each point in the numerical integration. The resulting trajectories are shown in Fig. \[fig:trajs\] for values of $D$ in the range 0-1.7 inches, where an initial height of 3 ft was assumed.
Discussion of trajectories {#sec:trajs}
--------------------------
The striking feature of Fig. \[fig:trajs\] is the qualitatively different character of the trajectories as a function of $D$, or equivalently as a function of the takeoff angle $\theta$. These trajectories range from line drives at small $\theta$, to fly balls at intermediate $\theta$, to pop-ups at large $\theta$. Particularly noteworthy is the rich and complex behavior of the pop-ups, including cusps and loops. The goal of this section is to understand these trajectories in the context of the interplay among the forces acting on the ball. To our knowledge, there has been no previous discussion of such unusual trajectories in the literature. We focus on two particular characteristics that may have implications for the algorithm used by a fielder to catch the ball: the symmetry/asymmetry about the apex and the curvature. Before proceeding, however, we remark that the general features of the trajectories shown in Fig. \[fig:trajs\] are universal and do not depend on the particular model used for either the ball-bat collision or for the drag and lift. For example, using collision and aerodynamics models significantly different from those used here, Adair finds similar trajectories with both cusp-like and loop-like behavior,[@adair02] which we verify with our own calculations using his model. Models based on equations in Watts and Bahill[@watts00] result in similar trajectories.
We first examine the symmetry, or lack thereof, of the trajectory about the apex. Without the drag and Magnus forces, all trajectories would be symmetric parabolas; the actual situation is more complicated. As seen in Fig. \[fig:trajs\], baseballs hit at low and intermediate $\theta$ (line drives and fly balls) have an asymmetric trajectory, with the ball covering less horizontal distance on the way down than it did on the way up. This feature is known intuitively to experienced outfielders. For larger $\theta$ the asymmetry is smaller, and pop-ups hit at a very steep angle are nearly symmetric. How do the forces conspire to produce these results?
We address this question by referring to Figs. \[fig:forcet1\] and \[fig:forcet2\], in which the time dependence of the horizontal components of the velocity and the forces are plotted for a fly ball ($D=0.75$, $\theta=33^\circ$) and a pop-up ($D=1.6$, $\theta=68^\circ$). The initial decrease of the drag force for early times is due to the particular model used for the drag coefficient, which experiences a sharp drop near 75 mph. The asymmetry of the trajectory depends on the interplay between the horizontal components of drag and Magnus, $F_{Dx}$ and $F_{Mx}$, respectively. For forward-going trajectories ($v_x>0$), $F_{Dx}$ always acts in the -x direction, whereas $F_{Mx}$ acts in the -x or +x direction on the rising or falling part of the trajectory, respectively. The relative magnitudes of $F_{Dx}$ and $F_{Mx}$ depend strongly on both $\theta$ and $\omega$. For fly balls, $\theta$ and $\omega$ are small enough (see Fig. \[fig:d\]) that the magnitude of $F_{Dx}$ is generally larger than the magnitude of $F_{Mx}$, as shown in Fig. \[fig:forcet1\]. Therefore $F_x$ is negative throughout the trajectory. Under such conditions, there is a smooth continuous decrease in $v_x$, leading to an asymmetric trajectory, since the horizontal distance covered prior to the apex is greater than that covered after the apex. The situation is qualitatively and quantitatively different for pop-ups, since both $\theta$ and $\omega$ are significantly larger than for a fly ball. As a result, the magnitude of $F_{Mx}$ is much greater than the magnitude of $F_{Dx}$. Indeed, Fig. \[fig:forcet2\] shows that $F_x\approx F_{Mx}$, so that $F_x$ acts in the -x direction before the apex and in the +x direction after the apex. Therefore, the loss of $v_x$ while rising is largely compensated by a gain in $v_x$ while falling, resulting in near symmetry about the apex. Moreover, for this particular trajectory the impulse provided by $F_x$ while rising is nearly sufficient to bring $v_x$ to zero at the apex, resulting in the cusp-like behavior. For even larger values of $\theta$, $F_x$ is so large that $v_x$ changes sign prior to the apex, then reverses sign again on the way down, resulting in the loop-the-loop pattern.
We next address the curvature of the trajectory, $C\equiv
d^2y/dx^2$, which is determined principally by the interplay between the Magnus force $F_M$ and the component of gravity normal to the trajectory $F_{GN}=F_G\cos\theta$. It is straightforward to show that $C$ is directly proportional to the instantaneous value of $(F_M-F_{GN})/(v_x^2\cos\theta)$ and in particular that the sign of $C$ is identical to the sign of $F_M-F_{GN}$. In the absence of a Magnus force, the curvature is always negative, even if drag is present. An excellent example is provided by the inverted parabolic trajectories expected in the absence of aerodynamic forces. The trajectories shown in Fig. \[fig:trajs\] fall into distinct categories, depending on the initial angle $\theta$. For small enough $\theta$, $C$ is negative throughout the trajectory. Indeed, if C is initially negative, then it is always negative, since $F_M$ is never larger and $F_{GN}$ is never smaller than it is at t=0. For our particular collision and aerodynamic model, the initial curvature is negative for $\theta$ less than about $45^\circ$. For intermediate $\theta$, $C$ is positive at the start and end of the trajectory but experiences two sign changes, one before and one after the apex. The separation between the two sign changes decreases as $\theta$ increases, until the two values coalesce at the apex, producing a cusp. For larger values of $\theta$, $C$ is positive throughout the trajectory, resulting in loop-like behavior such as the $D=1.7$ trajectory, where the sign of $v_x$ is initially positive, then changes to negative before the apex, and finally changes to back positive after the apex.
Does the spin remain constant? {#sec:spin}
------------------------------
All the simulations reported thus far assume that the spin remains constant throughout the trajectory. Since the spin plays such a major role in determining the character of the trajectory, it is essential to examine the validity of that assumption. To our knowledge, there have been no experimental studies on the spin decay of baseballs, but there have been two such studies for golf, one by Smits and Smith[@smits] and one by Tavares et al.[@tavares] Tavares et al. propose a theoretical model for the spin decay of a golf ball in which the torque responsible for the decay is expressed as $R\rho AC_Mv^2$, where $R$ is the radius of the ball and $C_M$ is the “coefficient of moment" which is given by $C_M = \beta R\omega/v$. By equating the torque to $Id\omega/dt$, where $I=0.4MR^2$ is the moment of inertia, the spin decay constant $\tau$ can be expressed as = . \[eq:tau\]Using their measurements of $\tau$, Tavares et al. determine $\beta\approx 0.012$, corresponding to $\tau=20$ sec for $v$=100 mph. The measurements of Smits and Smith can be similarly interpreted with $\beta$=0.009, corresponding to $\tau=25$ sec at 100 mph. To estimate the spin decay constant for a baseball, we assume Eq. \[eq:tau\] applies, with $M/R^2$ scaled appropriately for a baseball and with all other factors the same. Using $M/R^2$ = 2.31 and 2.49 oz/inch$^2$ for a golf ball and baseball, respectively, the decay time for a baseball is about 8% longer than for a golf ball, or 22-27 sec at 100 mph and longer for smaller $v$. A similar time constant for baseball was estimated by Sawicki et al.,[@sawicki05] quite possibly using the same arguments as we use here. Since the trajectories examined herein are in the air 7 sec or less, we conclude that our results are not affected by the spin decay. Adair has suggested a much smaller decay time, of order 5 sec,[@adair02] which does not seem to be based on any experimental data. A direct check of our calculations shows that the qualitative effects depicted in Fig. \[fig:trajs\] persist even with a decay time as short as 5 sec.
Optical control model for tracking and navigating baseballs
===========================================================
Overview {#sec:models}
--------
In a seminal article Seville Chapman[@chapman68] proposed an optical control model for catching fly balls, today known as Optical Acceleration Cancellation (OAC). Chapman examined the geometry of catching from the perspective of a moving fielder observing an approaching ballistic target that is traveling along a parabola. He showed that in this case, the fielder can be guided to the destination simply by selecting a running path that keeps the image of the ball rising at a constant rate in a vertical image plane. Mathematically, the tangent of the vertical optical angle to the ball increases at a constant rate. When balls are headed to the side, other optical control strategies become available.[@mcbeath95; @sugar06b] However, in the current paper we examine cases of balls hit directly toward the fielder, so we will emphasize predictions of the OAC control mechanism.
Chapman assumed parabolic trajectories because of his (incorrect) belief that the drag and Magnus forces have a negligible effect on the trajectory. Of course we now know that the effects of these forces can be considerable, as discussed in Sec. \[sec:trajs\]. Yet despite this initial oversight, numerous perception-action catching studies confirm that fielders actually do appear to utilize Chapman’s type of optical control mechanism to guide them to interception, and in particular OAC is the only mechanism that has been supported for balls headed in the sagittal plane directly toward fielders.[@babler93; @mcbeath95; @mcleod96; @sugar06a] Further support for OAC has been found with dogs catching Frisbees as well as functioning mobile robots.[@shafer04; @sugar06b]
Extensive research on the navigational behavior of baseball players supports that perceptual judgment mechanisms used during fly ball catching can generally be divided into two phases.[@mcbeath95; @shaffer05] During the first phase, while the ball is still relatively distant, ball location information is largely limited to the optical trajectory (i.e. the observed trajectory path of the image of the ball). During the second or final phase, other cues such as the increase in optical size of the ball, and the stereo angle between the two eyes also become available and provide additional information for final corrections in fielder positioning and timing. The control parameters in models like OAC are optical angles from the fielder’s perspective, which help direct fielder position relative to the ongoing ball position. Considerable work exploring and examining the final phase of catching has been done by perception scientists[@mazyn07; @salve93] and some recent speculation has been done by physicists.[@adair07] Researchers generally agree that the majority of fielder movement while catching balls takes place during the first phase in which fielders approach the destination region where the ball is headed. In the current work, we focus on control models like OAC that guide fielder position during the initial phase of catching. Thus for example, we would consider the famous play in which Jose Canseco allowed a ball to bounce off of his head for a home run to be a catch, in that he was guided to the correct location to intercept the ball.
An example of how a fielder utilizes the OAC control strategy to intercept a routine fly ball to the outfield is given in Fig. \[fig:oac\]. This figure illustrates the side view of a moving fielder using OAC control strategy to intercept two realistic outfield trajectories determined by our aerodynamics model described in Sec. \[sec:trajs\]. As specified by OAC, the fielder simply runs up or back as needed to keep the tangent of the vertical optical angle to the ball increasing at a constant rate. Since the trajectory deviates from a parabola, the fielder compensates by altering running speed somewhat. Geometrically the OAC solution can be described as the fielder keeping the image of the ball rising at a constant rate along a vertical projection plane that moves forward or backwards to remain equidistant to the fielder. For fly balls of this length, the geometric solution is roughly equivalent to the fielder moving in space to keep the image of the ball aligned with an imaginary elevator that starts at home plate and is tilted forward or backward by the amount corresponding to the distance that the fielder runs. As can be seen in the figure, these outfield trajectories are notably asymmetric, principally due to air resistance shortening, yet OAC still guides the fielder along a smooth, monotonic running path to the desired destination. This simple, relatively direct navigational behavior has been observed in virtually all previous perception-action catching studies with humans and animals.[@shafer04]
Application to examples of paradoxical trajectories {#sec:stategies}
---------------------------------------------------
Most previous models of interceptive perception-action assume that real-world fly ball trajectories remain similar enough to parabolic for robust optical control strategies like OAC to generally produce simple, monotonic running path solutions. Supporting tests have confirmed simple behavior consistent with OAC in relatively extreme interception conditions including catching curving Frisbees, towering outfield blasts and short infield pop-ups.[@mcbeath95; @mcleod96; @sugar06a; @shafer04; @sugar06b] The apparent robustness of these optical control mechanisms implies the commonly observed vacillating and lurching of fielders pursuing high pop-ups must be due to some inexplicable cause. It appears that the infielder is an unfortunate victim of odd wind conditions, if not perhaps a bit too much chew tobacco or a nip of something the inning before. In the current work, we have provided evidence that there is a class of high infield pop-ups that we refer to as paradoxical. Next we show that these deviate from normal parabolic shape in ways dramatic enough to lead fielders using OAC to systematically head off in the wrong direction or bob forward and back. Below we illustrate how a fielder guided by OAC will behave with each of the three paradoxical pop fly trajectories that we determined in Sec. \[sec:simulations\] of this paper.
We first examine perhaps the most extreme paradoxical trajectory of the group, the case of $D=1.7$, shown in Fig. \[fig:pop17\]. This trajectory actually does a full loop-the-loop between the catcher and pitcher, finally curving back out on its descent and landing about 30 feet from home plate. Given the extreme directional changes of this trajectory, we might expect an infielder beginning 100 feet from home plate to experience difficulty achieving graceful interception. Yet, as can be seen in the figure, this case actually results in a relatively smooth running path solution. When the fielder maintains OAC throughout his approach, he initially runs quickly forward, then slightly overshoots the destination, and finally lurches back. In practice, near the interception point, the fielder is so close to the approaching ball that it seems likely the eventual availability of other depth cues like stereo disparity and rate of change in optical size of the ball will mitigate any final lurch, and result in a fairly smooth overall running path to the destination.
Second we examine the case of a pop fly resulting from a bat-ball offset $D=1.6$ in Fig. \[fig:pop16\]. Here the horizontal velocity decreases in the beginning and approaches zero velocity near the apex. Then after the apex, the Magnus force increases the horizontal velocity. Yet, of greater impact to the fielder is that this trajectory’s destination is near where the fielder begins. Thus from the fielder’s perspective, before the discontinuity takes place the trajectory slows in the depth direction such as to guide the fielder to run up too far and then later to reverse course and backtrack to where the ball is now accelerating forward. Here the normally reliable OAC strategy leads the fielder to systematically run up too far and in the final second lurch backwards.
Third, we examine the case of a pop fly that lands just beyond the fielder, the $D=1.5$ condition, in Fig. \[fig:pop15\]. In this case OAC leads the fielder to initially head back to very near where the ball is headed, but then soon after change direction and run forward, only to have to run back again at the end. Certainly, when a fielder vacillates or “dances around" this much, it does not appear that he is being guided well to the ball destination. Yet, this seemingly misguided movement is precisely specified by the OAC control mechanism. Thus, the assumption that fielders use OAC leads to the bold prediction that even experienced, professional infielders are likely to vacillate and make a final lurch backward when navigating to catch some high, hard-hit pop-ups, and indeed this is a commonly witnessed phenomenon. Former major league infielders have affirmed to us that pop-ups landing at the edge of the outfield grass (100 to 130 ft. from home plate) usually are the most difficult to catch.
It is notable that in each of the cases depicted in Figs. \[fig:pop17\]-\[fig:pop15\], the final movement by the fielder prior to catching the ball is backwards. This feature can be directly attributed to the curvature of the trajectory, as discussed in Sec. \[sec:trajs\]. For a typical fly ball, the curvature is small and negative, so the ball breaks slightly towards home plate as it nears the end of its trajectory. For pop-ups, the curvature is large and positive, so the ball breaks away from home plate, forcing the fielder to move backward just prior to catching the ball.
Summary and conclusions {#sec:concl}
=======================
Why are very high pop-ups so hard to catch? Using models of the bat-ball collision and ball flight aerodynamics, we have shown that the trajectories of these pop-ups have unexpected features, such as loops and cusps. We then examined the running paths that occur with these dramatically non-parabolic trajectories when a fielder utilizes OAC, a control strategy that has been shown effective for tracking near-parabolic trajectories. The predicted behavior is very similar to observed behavior of infielders attempting to catch high pop-ups. They often vacillate forward and backward in trying to position themselves properly to make the catch, and frequently these changes in direction can lead to confusion and positioning error. Former major league infielders confirm that our model agrees with their experiences.
Acknowledgments {#acknowledgments .unnumbered}
===============
We are grateful to former major league players Clete Boyer, Jim French, Norm Gigon, Bill Heath, Dave Hirtz, and Wayne Terwilliger for their valuable comments and advice. Also, we thank David W. Smith and Stephen D. Boren for information they provided about pop-ups in the major leagues. Finally, we thank Bob Adair for sharing his own unpublished work with us on judging fly balls and for the insight regarding the final backward movement.
[99]{}
I. Newton, “New theory about light and colors,” Phil. Trans. Royal Soc. [**6**]{}, 3078 (1671).
R. K. Adair [*The Physics of Baseball*]{} (HarperCollins, New York, 2002), 3rd ed.
G. S. Sawicki, M. Hubbard and W.J. Stronge, “How to hit home runs: Optimum baseball swing parameters for maximum range trajectories," Am. J. Phys. [**71**]{}, 1152–1162 (2003).
R. Cross and A. M. Nathan, “Scattering of a baseball by a bat," Am. J. Phys. [**74**]{}, 896–904 (2006).
A. J. Smits and D. R. Smith, “A new aerodynamic model of a golf ball in flight,” Science and Golf II, Proceedings of the 1994 World Scientific Congress on Golf, edited by A. J. Cochran and M. R. Farraly(E&FN Spon., London, 1994), pp. 340-347.
G. Tavares, K. Shannon, and T. Melvin, “Golf ball spin decay model based on radar measurements,” Science and Golf III, Proceedings of the 1998 World Scientific Congress on Golf, edited by M. R. Farraly and A. J. Cochran(Human Kinetics, Champaign IL, 1999), pp. 464-472.
G. S. Sawicki, M. Hubbard, and W. Stronge, “Reply to Comment on How to hit home runs: Optimum baseball bat swing parameters for maximum range trajectories,” Am. J. Phys. [**73**]{}, 185-189 (2005).
R. G. Watts and A. T. Bahill [*Keep Your Eye on the Ball: Curveballs, Knuckleballs and Fallacies of Baseball*]{} (W. H. Freeman, New York, 2000).
S. Chapman, “Catching a baseball,” Am. J. Phys. [**53**]{}, 849–855 (1968).
M. K. McBeath, D. M. Shaffer, and M. K. Kaiser, “How baseball outfielders determine where to run to catch fly balls,” Science [**268**]{}, 569–573 (1995).
T. G. Sugar, et al., ”Mobile robot interception using human navigational principles: Comparison of active versus passive tracking algorithms,” Autonomous Robots, [**21**]{}, 43–54 (2006).
T. G. Babler, T. G. and J. L. Dannemiller, “Role of image acceleration in judging landing location of free-falling projectiles,” J. Expt. Psychology: Human Perception and Performance [**19**]{}, 15–31 (1993)
P. McLeod and Z. Dienes, “Do fielders know where to go to catch the ball or only how to get there?” J. Expt. Psychology: Human Perception and Performance [**22**]{}, 531–543 (1996)
T. G. Sugar, M. K. McBeath, and Z. Wang, “A unified fielder theory for interception of moving objects either above or below the horizon,” Psychonomic Bulletin and Review, [**13**]{}, 908–917 (2006).
D. M. Shafer, et al., “How dogs navigate to catch Frisbees,” Psychological Science, [**15**]{}, 437–441 (2004).
D. M. Shaffer and M. K. McBeath, “Naive beliefs in baseball: Systematic distortion in perceived time of apex for fly balls,” Journal of Experimental Psychology: Learning Memory and Cognition, [**31**]{} 1492–1501 (2005).
L. I. N. Mazyn, G. J. P. Savelsbergh, G. Montagne, and M. Lenoir, “Planned and on-line control of catching as a function of perceptual-motor constraints,” Acta Psychologoica, [**126**]{}, 59–78 (2007).
G. J. P. Savelsbergh, H. T. A. Whiting, J. R. Pijpers, and A. A. M. Vansantvoord 1993), “The visual guidance of catching,” Experimental Brain Research, [**93**]{}, 148–156 (1993).
R. K. Adair, private communication.
[.1in]{}[.1in]{}
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
Although many causal processes have spatial and temporal dimensions, the classical causal inference framework is not directly applicable when the treatment and outcome variables are generated by spatio-temporal point processes. The methodological difficulty primarily arises from the existence of an infinite number of possible treatment and outcome event locations at each point in time. In this paper, we consider a setting where the spatial coordinates of the treatment and outcome events are observed at discrete time periods. We extend the potential outcomes framework by formulating the treatment point process as a stochastic intervention strategy. Our causal estimands include the expected number of outcome events that would occur in an area of interest under a particular stochastic treatment assignment strategy. We develop an estimation technique by applying the inverse probability of treatment weighting method to the spatially-smoothed outcome surfaces. We show that under a set of assumptions, the proposed estimator is consistent and asymptotically normal as the number of time periods goes to infinity. Our motivating application is the evaluation of the effects of American airstrikes on insurgent violence in Iraq from February 2007 to July 2008. We consider interventions that alter the intensity and target areas of airstrikes. We find that increasing the average number of airstrikes from 1 to 6 per day for seven consecutive days increases all types of insurgent violence.
[**Keywords:**]{} inverse probability of treatment weighting, point process, spatial propensity score, stochastic intervention, time-varying confounders
author:
- 'Georgia Papadogeorgou[^1]'
- 'Kosuke Imai[^2]'
- 'Jason Lyall[^3]'
- 'Fan Li[^4]'
bibliography:
- 'Iraq.bib'
- 'intro.bib'
title: 'Causal Inference with Spatio-temporal Data: Estimating the Effects of Airstrikes on Insurgent Violence in Iraq[^5]'
---
Introduction
============
Many causal processes involve both spatial and temporal dimensions. Examples include the environmental impact of newly constructed factories, the economic and social effects of refugee influx, and the various consequences of disease outbreaks. When the treatment and outcome variables are generated by spatio-temporal point processes, the primary methodological challenge arises from the fact that there exists an infinite number of possible treatment and outcome event locations at each point in time.
Unfortunately, the classical causal inference framework that dates back to @neym:23 and @fish:35 is not directly applicable to such settings. Indeed, the standard causal inference approaches assume that the number of units which can receive the treatment is finite [e.g., @rubi:74a; @robi:97]. Although some studies develop a continuous time causal inference framework, they do not incorporate a spatial dimension [e.g., @gill:robi:01; @zhan:joff:small:11]. In addition, a small number of researches have proposed causal inference methods for analyzing fMRI experiments, which have both spatial and temporal dimensions. For example, [@luo2012inference] apply randomized inference, while @sobel2014causal employ structural modeling.
In this paper, we consider a setting where the treatment and outcome events are assumed to be generated by spatio-temporal point processes (Section \[sec:estimands\]). The proposed methodology is based on a single time series of spatial patterns of treatment and outcome variables, and builds upon three strands of the causal inference literature: interference, stochastic interventions, and time series. First, we address the possibility that treatments might affect outcomes at a future time period and at different locations in arbitrary ways. Although some scholars have considered unstructured interference, they assume non-spatial and cross-sectional settings [see @bass:airo:18; @savj:aron:hudg:19 and references therein]. @aronow2019design study spatial randomized experiments in a cross-section setting under the assumption that the number of potential intervention locations is finite and their spatial coordinates are known. In contrast, the proposed spatio-temporal causal inference framework allows for [*temporally and spatially unstructured interference*]{} over an infinite number of locations and across a finite number of time periods.
Second, instead of separately estimating the causal effects of treatment received at each location, we consider the impacts of different [*treatment assignment strategies*]{}, defined formally as the intervention distributions over treatment point patterns. Stochastic interventions have been used to deal with challenging causal inference problems [@Diaz2012population], including the violation of positivity assumption [@kennedy2019nonparametric], mediation analysis [@lok:16], and multiple treatments [@imai:jian:19]. We show that this approach is also useful for dealing with spatio-temporal treatments and outcomes.
Lastly, we do not impose any restriction on the spatial patterns of interference. As such, our estimation does not assume the separation of units into minimally interacting sets [e.g., @tchetgen2017auto]. Instead, we view the region of interest as a single unit where outcome events are observed over time. We then develop estimation and inferential methods by building on the time-series approach in the causal inference literature [@boji:shep:19]. We provide a sufficient set of assumptions that enable consistent estimation based on a spatially-smoothed inverse probability weighting estimator (Section \[sec:estimation\]). We conduct simulation studies to assess the finite sample performance of the proposed estimator (Section \[sec:simulations\]).
Our motivating application is the evaluation of the effects of American airstrikes on insurgent violence in Iraq from February 2007 to July 2008 (Section \[sec:iraq\]). We consider all airstrikes occurring within a day anywhere in Iraq as a [*treatment pattern*]{}, with bombed locations referred to as the *treatment-active locations*. Instead of focusing on the causal effects of each airstrike, we consider the impact of different [*airstrike strategies*]{}, defined formally as the distributions of airstrikes over the entire Iraq. The proposed methodology offers a data-driven approach to evaluate the effectiveness of airstrike strategies on the complex spatial structure of insurgent violence across Iraq.
By using various specifications of stochastic interventions, we examine how different airstrike strategies affect subsequent insurgent attacks (Section \[sec:application\]). First, we show that a higher frequency of airstrikes, without modifying their spatial distribution, increases the number of insurgent attacks, especially near Baghdad, Mosul, and the roads between them. We also find that changing the focal point of airstrikes to Baghdad without modifying the overall frequency shifts insurgent attacks from Baghdad to Mosul and its environs. Finally, our analysis suggests that increasing the number of airstrikes for a few days may initially reduce attacks but ultimately increases them over the long run.
Motivating Application: Airstrikes and Insurgent Activities in Iraq {#sec:iraq}
===================================================================
Background
----------
Airstrikes have emerged as a principal tool for defeating insurgent and terrorist organizations in civil wars around the globe. In the past decade alone, the United States has conducted sustained air campaigns in at least six different countries, including Afghanistan, Iraq, and Syria. Although it has been shown that civilians have all-too-often borne the brunt of these airstrikes [@Lyall:19b], we have few rigorous studies of airstrikes and their effects on subsequent insurgent violence. Even these few studies have largely reached opposite conclusions, with some claiming that airpower reduces insurgent attacks while others arguing they spark escalatory spirals of increased violence [e.g., @Lyall:19a; @Mir:19; @Dell:18; @Kocher:11]. Regardless of conclusions, all of the existing studies discretize a continuous space into arbitrary geographical units and make simplifying assumptions about the patterns of spatial and temporal interference.
We enter this debate by examining the American air campaign in Iraq. We use declassified US Air Force data on airstrikes and shows of force (simulated airstrikes where no weapons are released) for the February 2007 to July 2008 period. The period in question coincides with the “surge” of American forces and airpower designed to destroy multiple Sunni and Shia insurgent organizations in a bid to turn the war’s tide.
Aircraft were assigned to bomb targets via two channels. First, airstrikes were authorized in response to American forces coming under insurgent attack. These close air support (CAS) missions represented the vast majority of airstrikes in 2007-08. Second, a small percentage (about 5%) of airstrikes were pre-planned against high-value targets, typically insurgent commanders, whose presence had been detected from intercepted communications or human intelligence. In each case, airstrikes were driven by insurgent attacks that were either ongoing or had occurred in the recent past in a given location. As a result, the models used later in this paper adjust for prior patterns of insurgent violence in a given location for several short-term windows.
We also account for prior air operations, including shows of force, by American and allied aircraft. Insurgent violence in Iraq is also driven by settlement patterns and transportation networks. Our models therefore include population size and location of Iraqi villages and cities as well as proximity to road networks, where the majority of insurgent attacks were conducted against American convoys. Finally, prior reconstruction spending might also condition the location of airstrikes. Aid is often provided in tandem with airstrikes to drive out insurgents, while these same insurgents often attack aid sites to derail American hearts-and-minds strategies. Taken together, these four factors — recent insurgent attacks, the presence of American forces, settlement patterns, and prior aid spending — drove decisions about the location and severity of airstrikes.
Data
----
\
Figure \[fig:observed\_intensities\] summarizes the spatial and temporal distributions of airstrikes (treatment variable) and insurgent violence (outcome variable). Figure \[fig:air\_time\] presents the temporal distribution of airstrikes recorded by the US Air Force each month. There were a total of 3,254 airstrikes during this period. Figure \[fig:air\_space\] plots the spatial density of these airstrikes across Iraq, with spatial clustering observed around Baghdad and the neighboring “Sunni Triangle,” a hotspot of insurgency. Figure \[fig:ins\_time\] plots the monthly distribution of insurgent attacks by type: Improvised Explosive Devices (IEDs), small arms fire (SAF), and other attacks. A total of 140,977 insurgent attacks recorded by the US Army’s CIDNE database during this time period. Finally, Figure \[fig:ins\_space\] plots the locations of insurgent attacks across Iraq. Baghdad, the Sunni Triangle, and the highway leading north to Mosul are all starkly illustrated.
Causal Inference Framework for Spatio-temporal Data {#sec:estimands}
===================================================
In this section, we propose a causal inference framework for spatio-temporal point processes. We describe the setup, and define causal estimands using stochastic interventions.
The Setup
---------
We represent the locations of airstrikes for each time period (e.g., day) as a spatial point pattern measured at time $t \in {\mathcal{T}}= \{1, 2, \dots, T\}$ where $T$ is the total number of the discrete time periods. Let $W_t(s)$ denote the binary treatment variable at location $s$ for time period $t$, indicating whether or not the site receives the treatment during the time period. We use $W_t$ as a shorthand for $W_t(\Omega)$, which evaluates the binary treatment variable $W_t(s)$ for each element $s$ of the set of all locations $\Omega$ that may receive the treatment. In addition, ${\mathcal{W}}$ represents the set of all possible point patterns at each time period where, for simplicity, we assume that this set does not vary across time periods, i.e., $W_t \in {\mathcal{W}}$ for each $t$. Finally, the set of [*treatment-active locations*]{}, i.e., the sites that receive the treatment, at time $t$ is denoted by $\sparseset[W] = \{s \in \Omega : W_t(s) = 1\}$. We assume that the number of treatment-active locations is finite for each time period, i.e., $|\sparseset[W]| < \infty$ for any $t$. In our study, the treatment-active locations correspond to the set of coordinates of airstrikes.
We use $w_t$ to denote a realization of $W_t$ and $\whist = (w_1, w_2, \dots, w_t)$ to denote the history of treatment point pattern realizations from time 1 through time $t$. Let $Y_t(\whist)$ represent the potential outcome at time $t \in {\mathcal{T}}$ for any given treatment sequence $\whist \in {\mathcal{W}}^t = {\mathcal{W}}\times \cdots \times {\mathcal{W}}$, depending on [*all*]{} previous treatments. Similar to the treatment, $Y_t(\whist)$ represents a point pattern with locations $\sparseset[Y]$, which are referred to as the *outcome-active locations*. In our study, $\sparseset[Y]$ represents the locations of insurgent attacks if the patterns of airstrikes had been $\whist[t]$. Let $\anyhist[t'][{\mathcal{Y}}] = \{Y_{t}(\whist): \whist \in {\mathcal{W}}^{t}, t \leq t' \}$ denote the collection of potential outcomes up to time period $t'$ for all treatment sequences.
Among all of these potential outcomes for time $t$, we only observe the one corresponding to the observed treatment sequence $\whist^{obs}$, denoted by $Y_t^{obs}=Y_t(\whist^{obs})$. We use $\anyhist[t][Y]^{obs} = \{Y_{1}^{obs}, \ldots, Y_{t}^{obs}\}$ to represent the collection of observed outcomes up to and including time period $t$. In addition, let ${{\bm {X}}}_{t-1}$ be the set of possibly time-varying confounders that are realized prior to $W_t$ but after $W_{t-1}$ for $t \geq 1$, and $\anyhist[t'][{\mathcal{{{\bm {X}}}}}] = \{{{\bm {X}}}_t(\whist): \whist \in {\mathcal{W}}^t, t \leq t'\}$ be the set of potential values of ${{\bm {X}}}$ under any possible treatment history $\whist$. Also let $\history = \{\anyhist[t-1][W], \anyhist[t-1][Y]^{obs}, \anyhist[t-1][{{\bm {X}}}]^{obs}\}$ denote all observed history up to time $t-1$.
Causal Estimands under Stochastic Interventions
-----------------------------------------------
A notion central to our proposed causal inference framework is *stochastic interventions*. Instead of setting a treatment variable to a fixed value, a stochastic intervention specifies the probability distribution that generates the treatment. Although our framework accommodates a variety of intervention distributions, we consider intervention distributions based on Poisson point processes, which are fully characterized by an intensity function $\trtintensity: \Omega \rightarrow [0, \infty)$. For example, a homogeneous Poisson point process with $\trtintensity(s) = \trtintensity$ for all $s \in \Omega$, implies that the number of treatment-active locations follows a $\textsf{Poisson}(\trtintensity |\Omega|)$ distribution, with locations distributed independently and uniformly over $\Omega$. In general, the specification of the stochastic intervention should be motivated by policy or scientific objectives. Such examples in the context of our study are given in Section \[subsec:study\_interventions\].
Our causal estimands are the expected number of (potential) outcome-active locations under a specific stochastic intervention of interest, and the comparison of such quantities under different intervention distributions. We begin by defining the causal estimands for a stochastic intervention taking place over a single time period. Let $\intervdist[F]$ denote the distribution of a spatial point process with intensity $\trtintensity$. Also, let $N_B(\cdot)$ denote a counting measure on a region $B \subset \Omega$. Then, we can define the expected number of outcome-active locations for a region $B$ at time $t$ as $$\begin{aligned}
\avgout[F][][1] & \ = \ \int_{{\mathcal{W}}} N_B\left( Y_t(\whist[t-1]^{obs}, w_t)\right) \ \mathrm{d} \intervdist(w_t)
\ = \ \int_{{\mathcal{W}}} \left| S_{Y_t(\whist[t-1]^{obs}, w_t)} \cap B \right| \ \mathrm{d} \intervdist(w_t).
\end{aligned}
\label{eq:avgout_k1}$$ In our application, this quantity represents the expected number of insurgent attacks within a region of Iraq $B$ if the airstrikes at time $t$ were to follow the point process specified by $\intervdist$, given the observed history of airstrikes up to time $t - 1$.
![Graphical Illustration of Stochastic Intervention over Multiple Time Periods for Time Period $t$ and $t + 1$. Under intervention $\interv[T]$, treatments during time periods $t - {M}+ 1, \dots, t-1, t$ are assigned according to distributions $\intervdist[F][{M}], \dots, \intervdist[F][2], \intervdist[F][1]$.[]{data-label="fig:trt_intervention"}](fig_vis_intervention2.png){width="\textwidth"}
We now extend the above estimand to an intervention taking place over ${M}$ consecutive time periods. Consider an intervention, denoted by $\intervdist[T] = \intervdist[F][1] \times \dots \times \intervdist[F][{M}]$, under which the treatment at time $t$ is assigned according to $\intervdist[F][1]$, at time $t - 1$ according to $\intervdist[F][2]$, continuing until time period $t - {M}+ 1$ for which treatment is assigned according to $\intervdist[F][{M}]$. A treatment path based on this intervention is displayed in \[fig:trt\_intervention\](a). Then, we define a general estimand as $$\begin{aligned}
\avgout[T]
& \ = \ \int_{{\mathcal{W}}^{M}} N_B\left( Y_t(\whist[t-{M}]^{obs}, w_{t-{M}+1}, \dots, w_t)\right) \ \mathrm{d} \intervdist[F][{M}](w_{t-{M}+1})\dots \mathrm{d} \intervdist[F][1](w_t) \\
& \ = \ \int_{{\mathcal{W}}^{M}} \left| S_{Y_t(\whist[t-{M}]^{obs}, w_{t-{M}+1}, \dots, w_t)} \cap B \right| \ \mathrm{d} \intervdist[F][{M}](w_{t-{M}+1})\dots \mathrm{d} \intervdist[F][1](w_t).
\end{aligned}
\label{eq:avgout_k}$$ This quantity represents the expected number of outcome events within region $B$ and at time $t$ if the treatment point pattern during the previous ${M}$ time periods were to follow the stochastic intervention with distribution $\intervdist[T]$. Treatments during the initial $t-{M}$ time periods were the same as observed. A special case of $\interv[T]$ assumes that treatments during the ${M}$ time periods are independent and identically distributed draws from the same distribution $\intervdist$, which we denote by $\intervdist[T] = \intervdist[F][][{M}]$.
Given the above setup, we define the average treatment effect of stochastic intervention $\intervdist[T][']$ versus $\intervdist[T]['']$ for a region $B$ at time $t$ as $$\effect[F][T] \ = \ \avgout[T][''] - \avgout[T]['],
\label{eq:effect_k}$$ where ${\mathbf{\trtintensity}}' = (\trtintensity_1', \trtintensity_2', \dots, \trtintensity_{M}')$ represents a collection of treatment intensities over ${M}$ consecutive time periods (similarly for $\mathbf \trtintensity''$).
We further define the average, over time periods $t={M}, {M}+1, \ldots, T$, of the expected potential outcome for region $B$ at each time period if treatments during the ${M}$ proceeding time periods arose from $\interv[T]$ as $$\tempavgout[T] = \frac1{T - {M}+ 1} \sum_{t = {M}}^T \avgout[T].
\label{eq:tempavgout_k}$$ \[fig:trt\_intervention\] shows two of the terms averaged in \[eq:tempavgout\_k\], $\avgout[T]$ and $\avgout[T][][][(t+1)]$. For $\avgout[T]$, treatments up to $t-{M}$ are set to their observed values, and treatments at time periods $t-{M}+1, \dots, t$ are drawn from $\interv[T]$. The same definition applies to $\avgout[T][][][(t +1)]$, but intervention time periods are shifted by 1: treatments up to $t - {M}+ 1$ are set to their observed values, and treatments during time periods $t - {M}+ 2, \dots, t + 1$ are drawn from $\interv[T]$. In \[eq:tempavgout\_k\], the summation starts at $t = {M}$ since the quantity $\avgout[T]$ assumes that there exist ${M}$ prior time periods during which treatments are intervened on.
Based on $\tempavgout[T]$, we define the causal effect of intervention $\interv[T][']$ versus $\interv[T]['']$ as $$\begin{aligned}
\tempeffect[F][T][']['']
& \ = \ \tempavgout[T][''] - \tempavgout[T]['] = \frac1{T - {M}+ 1} \sum_{t = {M}}^T \effect[F][T].
\label{eq:temp_effect_k}
\end{aligned}$$ This estimand represents the average, over time periods $t={M}, {M}+ 1, \ldots, T$, of the expected change in the number of points at each time period when the observed treatment path $\whist[T]^{obs}$ was followed until $t - {M}$ with subsequent treatments $w_{(t - {M}+1):t}$ arising according to $\interv[T]$. The effect size of a point pattern treatment would depend on ${M}$, and a greater value of ${M}$ would be more useful for studying slow-responding outcome processes. Finally, specifying $\interv[T][']$ and $\interv[T]['']$ such that they are identical except at ${M}$ time periods ago yields the lagged effect of a treatment change, as we will discuss in more detail in Section \[subsec:study\_interventions\].
Estimation {#sec:estimation}
==========
In this section, we first describe a set of assumptions sufficient for the identification of the causal quantities presented in the previous section. We then propose an estimator based on the inverse probability of treatment weighting and derive its asymptotic properties. All proofs are given in \[app\_sec:proofs\].
The Assumptions
---------------
Similar to the classic causal inference settings, variants of the unconfoundedness and overlap assumptions are required for the current settings based on stochastic interventions. For simplicity, we focus on stochastic interventions with identical and independent distribution over ${M}$ periods, $\interv[T] = \interv[F][][{M}]$, and intensity $\trtintensity$. Our theoretical results, however, extend straightforwardly to stochastic interventions with non-i.i.d. treatment patterns.
The treatment assignment at time $t$ is independent of all, past or future, potential outcomes and potential confounders conditional on the observed history of treatments, confounders and outcomes up to time $t-1$: $W_t \ {\perp\!\!\!\perp}\ \left\{ \anyhist[T][{\mathcal{Y}}], \anyhist[T][{\mathcal{{{\bm {X}}}}}] \right\} \mid \history.$ \[ass:unmeasured\_conf\]
\[ass:unmeasured\_conf\] resembles the sequential ignorability assumption in longitudinal treatment settings [@robins1999association; @robins2000marginal]. The difference is that \[ass:unmeasured\_conf\] requires the treatment assignment to be conditionally independent of all potential values of the time-varying confounders as well as those of the outcome variable. Unlike the typical longitudinal settings, under which a large number of units are assumed to be independent of one another, the current setting has a [*single*]{} time series of maps. Our assumption is similar to the non-anticipating treatment assumption of [@boji:shep:19], who consider a setting with a single unit being exposed to a binary treatment over time. We extend their assumption by explicitly incorporating time-varying confounders.
The next assumption weakens the standard overlap assumption (also known as positivity). We define the probability *density* of treatment realization $w$ at time $t$ given the history, $\propscore[] = f(W_t = w \mid \history)$, as the propensity score at time period $t$. Also, let $\intervdistf[][F]$ denote the probability density function of the stochastic intervention $\interv[F]$.
There exists a constant $\bound[W] > 0$ such that $\propscore[] > \bound[W] \cdot \intervdistf[][F](w)$ for all $w \in {\mathcal{W}}$. \[ass:positivity\]
Assumption \[ass:positivity\] ensures that all the treatment patterns which are possible under the stochastic intervention of interest can also be observed. If the density $\intervdistf[][F]$ of the stochastic intervention is bounded, \[ass:positivity\] is automatically satisfied so long as the standard overlap assumption holds, i.e., there exists $\bound[W] > 0$ such that $\propscore[] > \bound[W]$ for all $w \in {\mathcal{W}}$.
The Propensity Score for Point Process Treatments {#sec:balance}
-------------------------------------------------
As shown below, the propensity score plays an important role in our estimation. Here, we show that the propensity score for point process treatments has two properties analogous to those of the standard propensity score [@rosenbaum1983central]. That is, the propensity score is a balancing score, and under \[ass:unmeasured\_conf\] the treatment assignment is unconfounded conditional on the propensity score.
The propensity score $\propscore[]$ is a balancing score in that it satisfies: $f(W_t = w \mid \propscore[], \history) \ = \ f(W_t = w \mid \propscore[])$, for all $t$. \[theorem:balancing\_score\]
\[theorem:balancing\_score\] allows us to empirically assess the propensity score model specification by checking the predictive power of covariates in $\history$ in a model for the treatment $W_t$, with and without adjusting for the propensity score. For example, if a covariate significantly improves prediction in a point process model for $W_t$ that adjusts for the estimated propensity score, then the covariate is not balanced and propensity score misspecification is likely.
Under \[ass:unmeasured\_conf\], the treatment assignment at time $t$ is unconfounded given the propensity score at time $t$, that is, given $W_t \ {\perp\!\!\!\perp}\ \left\{ \anyhist[T][{\mathcal{Y}}], \anyhist[T][{\mathcal{{{\bm {X}}}}}] \right\} \mid \history$, we have $W_t {\perp\!\!\!\perp}\left\{ \anyhist[T][{\mathcal{Y}}], \anyhist[T][{\mathcal{{{\bm {X}}}}}] \right\} \mid \propscore[][t][t][W]$. \[theorem:ps\_unconfoundedness\]
\[theorem:ps\_unconfoundedness\] shows that the potentially high-dimensional set $\history$ can be reduced to one dimensional propensity score function $\propscore[]$, which is a sufficient conditioning set for identifying the causal effect of $W_t$.
The Estimators
--------------
We propose a class of estimators for the causal estimands defined in Section \[sec:estimands\], which combines the inverse probability of treatment weighting (IPW) with the kernel smoothing of spatial point patterns. The estimator proceeds in two steps. First, at each time period $t$, the surface of outcome-active locations is spatially smoothed according to a chosen kernel. Then, this surface is weighted by the relative density of the observed treatment pattern under the stochastic intervention of interest and under the actual data generating process.
Formally, consider a univariate kernel $K: [0,\infty) \rightarrow [0, \infty)$ satisfying $\int K(u) \mathrm{d}u = 1$, and bandwidth parameter $b$. Let $K_b$ denote the scaled kernel defined as $K_b(u) = b^{-1} K(u / b)$. We define $\estimatort[F][][{M}]: \Omega \rightarrow \mathbb{R}^{+}$ as $$\begin{aligned}
\estimatort[F][][{M}][T] =
\prod_{j = t - {M}+ 1}^t
\frac{\intervdistf[][T][j]}{\propscore[obs][j][j]}
\Bigg[ \sum_{s \in \sparseset[Y^{obs}][t]} K_b(\|\omega - s\|) \Bigg],
\end{aligned}
\label{eq:estimatort}$$ where $\|\cdot\|$ denotes the Euclidean norm. The summation on the right hand side represents the spatially-smoothed version of the outcome point pattern at time period $t$. The product of ratios represents a weight similar to the weights in the marginal structural models, but in accordance with the stochastic intervention $\interv[F][][{M}]$: each of the ${M}$ terms represents the likelihood ratio of observing the treatment $w_j^{obs}$ in the counterfactual world of the intervention $\interv$ versus the actual world with the observed data at a specific time period.
Assuming that the kernel $K$ is continuous, the estimator in \[eq:estimatort\] defines a continuous surface over $\Omega$. The continuity of $\estimator[F][][{M}]$ allows us to conceptualize causal quantities by viewing $\estimatort[F][][{M}]$ itself as an intensity function. This leads to the following estimator for the expected number of outcome-active locations in any region $B$ at time $t$, defined in \[eq:avgout\_k\], $$\estimatorNt[F][][{M}] \ = \ \int_B \estimatort[F][][{M}][T] \mathrm{d} \omega.
\label{eq:estimatorNt}$$ We use this estimator for the temporally-expected average potential outcome in \[eq:tempavgout\_k\] as $$\estimatorN[F][][{M}] \ = \
\frac1{T-{M}+1} \sum_{t = {M}}^T \estimatorNt[F][][{M}],
\label{eq:estimatorN}$$ and for the causal contrast between two interventions $\interv[F][1][{M}]$ and $\interv[F][2][{M}]$ defined in \[eq:temp\_effect\_k\] as $$\tempeffect[T][F][1][2][{M}] \ = \ \estimatorN[F][2][{M}] - \estimatorN[F][1][{M}].
\label{eq:estimator_effect}$$ For a specific intervention $\interv[F][][{M}]$, the estimator in \[eq:estimatort\] can be calculated once and it can be used to estimate the temporally-expected effects defined in Section \[sec:estimands\] for any $B \subset \Omega$.
Asymptotic Properties of the Proposed Estimators
------------------------------------------------
Below, we establish the asymptotic properties of the proposed estimators. Our results differ from most asymptotic normality results in the causal inference literature in two ways. First, they are based on a single observation (namely, one time series of point patterns that are potentially interconnected over both space and time). Second, we employ a kernel-smoothed version of the outcome. We introduce a set of mild regularity conditions.
\[ass:regularity\_conditions\] The following three conditions hold.
1. \[ass:finite\_points\] There exists $\bound[Y] > 0$ such that $|\sparseset[Y]| < \bound[Y]$ for all $t \in {\mathcal{T}}$ and $\whist \in {\mathcal{W}}^T$.
2. \[ass:convergent\_variance\] Let $\history[t]^* = \{\anyhist[t][W], \anyhist[T][{\mathcal{Y}}], \anyhist[T][{\mathcal{{{\bm {X}}}}}] \} \supset \history[t]$, and ${v}_t = {\text{Var}}\big[ \estimatorNt[F][][{M}] \big| \history[t-{M}]^* \big]$ for $t \geq {M}$. Then, there exists ${v}\in \mathbb{R}^+$ such that $(T-{M}+1)^{-1} \sum_{t = {M}}^T {v}_t \overset{p}{\rightarrow} {v}$ as $T \rightarrow \infty$.
3. \[ass:neighborhood\_boundary\_informal\] There exists a neighborhood of set $B$’s boundary over which outcome active locations are observed during at most $T^{1-Q^*}$ time periods, for some $Q^* \in (1/2, 1)$ and as $T \rightarrow \infty$.
\[ass:regularity\_conditions\]\[ass:finite\_points\] states that there is an upper limit on the number of outcome-active locations at any time period and under any treatment path. In our application, it is reasonable to assume that the number of insurgent attacks occurring during any day is bounded. In \[ass:regularity\_conditions\]\[ass:convergent\_variance\], $\history[t]^*$ represents the expanded history preceding $W_{t+1}$, including previous treatments, all potential outcomes, and all potential confounders. Given the assumptions of bounded relative positivity and bounded number of outcome-active locations, $\estimatorNt[F][][{M}]$ is also bounded, so \[ass:regularity\_conditions\]\[ass:convergent\_variance\] is a weak addition. Lastly, a more rigorous statement of \[ass:regularity\_conditions\]\[ass:neighborhood\_boundary\_informal\] requires additional notation, which is presented in \[app\_sec:proofs\]. Since the size of the boundary’s neighborhood can be arbitrarily small, this assumption is also reasonable. We now present the main theoretical results.
If Assumptions \[ass:unmeasured\_conf\]–\[ass:regularity\_conditions\] hold and the bandwidth $b_T \rightarrow 0$, we have, $$\sqrt{T}(\estimatorN - \tempavgout) \overset{d}{\rightarrow} \mathcal{N}(0, {v}),$$ as $T \rightarrow \infty$, where ${v}$ is the quantity defined in \[ass:regularity\_conditions\]\[ass:convergent\_variance\]. \[theorem:normality\]
The key idea is to separate the estimation error arising due to the treatment assignment, $[W_t \mid \history[t-1]^*]$, from the error due to spatial smoothing. Using martingale theory similar to [@boji:shep:19], we show the former to be $\sqrt{T}$-asymptotically normal. The latter is shown to converge to zero at a rate faster than $1/\sqrt{T}$.
Knowing ${v}$ would allow us to estimate the temporally-expected potential outcome and make inference based on the estimator’s asymptotic distribution, since $ \displaystyle \estimatorN[F][][{M}] \overset{d}{\approx} \mathcal{N}( \tempavgout[F][][{M}], {v}/ T) $ for large $T$. The variance ${v}$ is the converging point of $(T-{M}+1)^{-1}\sum_{t = {M}}^M {v}_t$ in \[ass:regularity\_conditions\]\[ass:convergent\_variance\], i.e., the average over the time $t$ estimator’s variance. Because for each time period $t$ we only observe one treatment path, one cannot directly estimate ${v}_t$ and thus ${v}$ without additional assumptions. We circumvent this problem by using an upper bound of $v$, a quantity, which we can consistently estimate. Specifically, let ${v}_t^* = \text{E}\left\{ [\estimatorNt[F][][{M}]]^2 \mid \history[t-{M}]^* \right\}$. Then, it is clear that ${v}\leq (T-{M}+1)^{-1}\sum_{t = {M}}^T {v}_t^* \overset{p}{\rightarrow} {v}^*$, and an $\alpha$-level confidence interval for $\tempavgout$ which uses the asymptotic variance bound ${v}^*/T$ will achieve the nominal coverage. Even though ${v}^*$ cannot be directly calculated either, there exists a consistent estimator of the asymptotic variance’s upper bound, as stated in the following lemma:
It holds that $ \displaystyle \frac{1}{T-{M}+1} \sum_{t = {M}}^T \left[ \estimatorNt[F][][{M}]^2 - {v}_t^* \right] \overset{p}{\rightarrow} 0. $ \[lemma:consistent\_variance\]
The results presented above for the estimator $\estimatorN$ can be easily extended to the estimator $\tempeffect[T][F][1][2]$:
If Assumptions \[ass:unmeasured\_conf\]–\[ass:regularity\_conditions\] hold, $$\frac{1}{T - {M}+ 1} \sum_{t = {M}}^T {\text{Var}}\left[\effect[T][F][1][2] \mid \history[t - {M}]^*\right] \overset{p}{\rightarrow} \eta,$$ for some $\eta > 0$, and the bandwidth $b_T \rightarrow 0$, we have $$\sqrt{T}(\tempeffect[T][F][1][2] - \tempeffect[F][F][1][2]) \overset{d}{\rightarrow} N(0, \eta),$$ as $T \rightarrow \infty.$ Furthermore, an upper bound of the asymptotic variance $v$ can be consistently estimated by $$\frac{1}{T - {M}+ 1} \sum_{t = {M}}^T \left[\effect[T][F][1][2]\right]^2 \overset{p}{\rightarrow} v^\ast \ge v.$$ \[theorem:normality\_tau\]
All of the theoretical results presented above have been established for the correct and known propensity score $\propscore[]$. However, in practice, the propensity score is unknown and must be estimated. When the propensity score is estimated from a correctly-specified model, it has been shown in various contexts that the resulting estimator maintains the desirable properties described above while having a lower asymptotic variance than the estimator utilizing the true propensity score [e.g., @Hirano2003efficient]. Therefore, even for the estimated propensity score, we make inference based on the upper bound of the asymptotic variance derived above.
Simulation Studies {#sec:simulations}
==================
We conduct simulation studies to evaluate
the performance of the proposed estimator as the number of time periods increases,
the accuracy of the asymptotic approximation,
the difference between the theoretical variance bound and the actual variance,
the performance of the inferential approach based on the estimated asymptotic variance bound, and
the balancing properties of the estimated propensity score.
We use the `spatstat` R package [@Baddeley2015spatial] to generate point patterns from Poisson processes and fitting Poisson process models to the simulated data.
The Simulation Design
---------------------
We consider a time series of point patterns of length $T \in \{200, 400, 500\}$ on the unit square, $\Omega = [0,1]\times[0,1]$. For each time series length $T$, 200 data sets are generated. We now describe our simulation design in detail.
### Time-varying and time-invariant confounders. {#time-varying-and-time-invariant-confounders. .unnumbered}
Our simulation study includes two time-invariant confounders. For the first time-invariant confounder, we construct a hypothetical road network on $\Omega$ using lines and arcs, which is highlighted by bright white lines in \[fig:sims\_roads\]. Then, we define ${X}^1=1.2\exp{(-2D_1)}$ where $D_1$ is the distance to the closest line. The second time-invariant covariate is constructed similarly, as ${X}^2=\exp{(-3D_2)}$ where $D_2$ is the distance to the closest arc. In addition, we generate two time-varying confounders, ${X}_t^3$ and ${X}_t^4$, based on the exponential decay of distance to the closest point; these points are generated according to non-homogeneous Poisson point processes with the following intensity function $$\lambda_t^{X^j}(\omega) = \exp \big\{ \rho_0^j + \rho_1^j X^1(\omega) \big\}, \ j =3,4,$$ where $\rho_1^3 = 1$, and $\rho_1^4 = 1.5$. \[fig:sims\_Xt\_one\] shows one realization of ${X}_t^3$.
### Spatio-temporal point processes for treatment and outcome variables. {#spatio-temporal-point-processes-for-treatment-and-outcome-variables. .unnumbered}
For each time period $t \in \{1, 2, \dots, T\}$, we generate $W_t$ from a non-homogeneous Poisson process that depends on all confounders ${{\bm {X}}}_t(\omega)^T = ({X}^1, {X}^2, {X}_t^3, {X}_t^4)(\omega)$, and the previous treatment and outcome realizations, $W_{t - 1}$ and $Y_{t - 1}$. The intensity of this process is given by $$\lambda_t^W (\omega) \ = \ \exp\big\{ \alpha_0 + {{\bm {X}}}_t(\omega)^T \alpha_{{X}} + \alpha_W W_{t-1}^*(\omega) + \alpha_Y Y_{t - 1}^*(\omega) \big\},
\label{eq:sims_treatment_intensity}$$ where $W_{t - 1}^*(\omega)=\exp(-2 D_W(\omega))$ and $Y_{t - 1}^*(\omega)=\exp(-2 D_Y(\omega))$ with $D_W(\omega)$ and $D_Y(\omega)$ being the minimum distance of $\omega$ to the points in $\sparseset[W][t-1]$ and $\sparseset[Y{}][t-1]$, respectively.
Similarly, we generate $Y_t$ from a non-homogeneous Poisson process with intensity $$\lambda_t^Y(\omega) \ = \ \exp\left\{ \gamma_0 + {{\bm {X}}}_t(\omega)^T \gamma_{{X}} +\gamma_2 {X}_{t-1}^2(\omega) + \gamma_W W_{(t-3):t}^*(\omega) + \gamma_Y Y_{t - 1}^*(\omega) \right\},
\label{eq:outcome_intensity}$$ where $W_{(t-3):t}^*(\omega)=\exp(-2 D_W^\ast(\omega))$ is the minimum distance of $\omega$ to the points in $\bigcup_{j = t-3}^t \sparseset[W][j]$. This specification imposes a lag-four dependence of the outcome on the lagged treatment process. The model gives rise to the average of 5 observed treatment-active locations and 21 observed outcome-active locations within each time period.
### Stochastic interventions. {#stochastic-interventions. .unnumbered}
We consider interventions of the form $\interv[F][][{M}]$ based on a homogeneous Poisson process with intensity $h$ that is constant over $\Omega$ and ranges from 3 to 7. We consider various lengths of each intervention by setting $M \in \{1, 3, 7, 30\}$. The second intervention we consider is defined over the three time periods, i.e., $\interv[T]=\intervdist[F][3] \times \intervdist[F][2] \times \intervdist[F][1]$ with ${M}= 3$. The intervention for the first time period $\intervdist[F][3]$ is a homogeneous Poisson process with intensity $h_3$ ranging from 3 to 7, whereas $\intervdist[F][2] = \intervdist[F][1]$ is a homogeneous Poisson process with intensity equal to 5. For each stochastic intervention, we consider the region of interest, denoted by set $B$, of three different sizes: $B = \Omega=[0, 1] \times [0,1]$, $B = [0, 0.5]\times[0,0.5]$, and $B = [0.75, 1]\times[0.75, 1]$.
### Approximating the true values of the estimands. {#approximating-the-true-values-of-the-estimands. .unnumbered}
From \[eq:outcome\_intensity\], it is clear that the potential outcomes depend on the realized treatments during the last four time points as well as the realized outcomes from the previous time period. This implies that the estimands for all interventions, even for $M > 4$, depend on the observed treatment and outcome paths and are therefore not constant across simulated data sets. Therefore, we approximate the true values of the estimands in the following manner. For each time period $t$, and each $r=1,\ldots,R$ repetition, we generate realizations $w_{t - {M}+ 1}{^{(r)}}, \dots, w_{t - 1}{^{(r)}}, w_t{^{(r)}}$ from the intervention distribution $\intervdist[T]$. Based on the treatment path $(\whist[t - {M}]^{obs}, w_{t - {M}+ 1}{^{(r)}}, \dots, w_t{^{(r)}})$, we generate outcomes $Y_{t-{M}+ 1}{^{(r)}}, \dots, Y_t{^{(r)}}$ using \[eq:outcome\_intensity\]. This yields $\sparseset[Y{^{(r)}}][t]$, which contains the outcome-active locations based on one realization from the stochastic intervention. Repeating this process $R$ times and calculating the average number of points that lie within $B$ provides a Monte Carlo approximation of $\avgout[T]$, and further averaging these over time gives an approximation of $\tempavgout[T]$.
### Estimation. {#estimation. .unnumbered}
We estimate the expected number of points $\tempavgout[T]$ and the effect of a change in the intervention on this quantity $\tempeffect[F][T]$ using the following estimators:
the proposed estimators defined in \[eq:estimatorN\] and \[eq:estimator\_effect\] with the true propensity scores;
the same proposed estimators but with the estimated propensity scores based on the correctly-specified model;
the above two estimators but with the Hájek-type standardization where the weight of time $t$ is divided by the mean of weights across all time periods (see \[app\_sec:hajek\]); and
the unadjusted estimator based on a propensity score model which is specified as a homogeneous Poisson process without confounders.
All estimators utilize the smoothed outcome point pattern. Spatial smoothing is performed using Gaussian kernels with standard deviation equal to $10 T^{-2/3}$, which is decreasing in $T$. We choose this bandwidth such that for $T = 500$ (the longest time series in our simulation scenario) the bandwidth is approximately equal to 0.16, smaller than the size of the smallest $B$ (which is equal to $[0.75,1]^2$). A discussion on choosing the bandwidth is given in Section \[subsec:application\_bandwidth\].
### Theoretical variance and its upper bound. {#theoretical-variance-and-its-upper-bound. .unnumbered}
Theorems \[theorem:normality\] and \[theorem:normality\_tau\] provide the expressions for the asymptotic variances of the proposed estimators. We compute Monte Carlo approximations to these theoretical variances and their upper bounds. Specifically, for each time period $t$ and each replication $r$, the computation proceeds as follows:
We generate treatment and outcome paths $w_{t - {M}+ 1}{^{(r)}}, y_{t - {M}+ 1}{^{(r)}}, \dots, w_t{^{(r)}}, y_t{^{(r)}}$ using the distributions specified in \[eq:sims\_treatment\_intensity\] and \[eq:outcome\_intensity\],
Using the data $(w_{t - {M}+ 1}{^{(r)}}, \dots, w_t{^{(r)}})$ and the outcome $y_t{^{(r)}}$, we compute the estimator according to \[eq:estimatort\] and \[eq:estimatorNt\], and finally
we calculate the variance and the second moment of these estimates over $R$ replications, which can be used to compute the asymptotic variance and variance bound of interest.
The averages of these estimates over time gives the desired Monte Carlo approximations. We use a similar procedure to approximate the theoretical variance and variance bound of $\tempeffect[T][T]$.
### Estimating the variance bound and the resulting inference. {#estimating-the-variance-bound-and-the-resulting-inference. .unnumbered}
We use \[lemma:consistent\_variance\] to estimate the variance bound. This estimated variance bound is then used to compute the confidence intervals and conduct a statistical test of whether the causal effect is zero. Inference based on the Hájek-type adjustment is discussed in \[app\_sec:hajek\].
### Balancing property of the propensity score. {#balancing-property-of-the-propensity-score. .unnumbered}
Using the correctly specified model, we estimate the propensity score at each time period $t$. The inverse of the estimated propensity score is then used as the weight in the weighted Poisson process model for $W_t$ with the intensity specified in \[eq:sims\_treatment\_intensity\]. We compare the statistical significance of the predictors between the weighted and unweighted model fits. Large $p$-values under the weighted model would suggest that the propensity score adequately balances the confounding variables.
Simulation Results
------------------
\[fig:4res1\_estimates\] presents the results for the two simulation scenarios. The top panel shows how the (true and estimated) average potential outcomes in the whole region ($B = \Omega$) change as the intensity varies under the single time period interventions. The bottom panel shows how the true and estimated average potential outcomes in the sub-region $[0.75, 1]^2$ change under the three time period interventions when the intensity at three time periods ago ranges from 3 to 7. For both simulation scenarios, we vary the length of the time series from 200 (left plots) to 500 (right plots).
![Simulation Results for the True and Estimated Average Potential Outcomes. In the top panel, we present the true and estimated average potential outcomes in the entire region $B = \Omega$ under single-time interventions with the varying intensity (horizontal axis). In the bottom panel, we consider the average potential outcome in the sub-region $B = [0.75,1]^2$ for the intervention $\interv[T]$, with $M = 3$, the varying intensity of $\interv[F][3]$ (horizontal axis), and $\interv[F][1], \interv[F][2]$ intensity set to 4. The black lines with solid circles represent the truths, while the other dotted or dashed lines represent the estimates; the estimator based on the true propensity score (purple triangles), the unadjusted estimator (green crosses), the estimator based on the estimated propensity score (blue x’s), the Hájek estimator based on the estimated propensity score (orange rhombuses).[]{data-label="fig:4res1_estimates"}](fig_sims_y_int1_B1.pdf "fig:"){width="98.00000%"}\
![Simulation Results for the True and Estimated Average Potential Outcomes. In the top panel, we present the true and estimated average potential outcomes in the entire region $B = \Omega$ under single-time interventions with the varying intensity (horizontal axis). In the bottom panel, we consider the average potential outcome in the sub-region $B = [0.75,1]^2$ for the intervention $\interv[T]$, with $M = 3$, the varying intensity of $\interv[F][3]$ (horizontal axis), and $\interv[F][1], \interv[F][2]$ intensity set to 4. The black lines with solid circles represent the truths, while the other dotted or dashed lines represent the estimates; the estimator based on the true propensity score (purple triangles), the unadjusted estimator (green crosses), the estimator based on the estimated propensity score (blue x’s), the Hájek estimator based on the estimated propensity score (orange rhombuses).[]{data-label="fig:4res1_estimates"}](fig_sims_y_int4_B3.pdf "fig:"){width="98.00000%"}\
![Simulation Results for the True and Estimated Average Potential Outcomes. In the top panel, we present the true and estimated average potential outcomes in the entire region $B = \Omega$ under single-time interventions with the varying intensity (horizontal axis). In the bottom panel, we consider the average potential outcome in the sub-region $B = [0.75,1]^2$ for the intervention $\interv[T]$, with $M = 3$, the varying intensity of $\interv[F][3]$ (horizontal axis), and $\interv[F][1], \interv[F][2]$ intensity set to 4. The black lines with solid circles represent the truths, while the other dotted or dashed lines represent the estimates; the estimator based on the true propensity score (purple triangles), the unadjusted estimator (green crosses), the estimator based on the estimated propensity score (blue x’s), the Hájek estimator based on the estimated propensity score (orange rhombuses).[]{data-label="fig:4res1_estimates"}](fig_sims_legend.pdf "fig:"){width="80.00000%"}
As expected, the unadjusted estimates (green crosses) are far away from the true average potential outcome (black solid circles) across all simulation scenarios. In contrast, and consistent with the results of Theorems \[theorem:normality\] and \[theorem:normality\_tau\], the accuracy of the proposed estimator (purple triangles based on the true propensity score, blue x’s based on the estimated propensity score) improves as the number of time periods increases. We note that the convergence is slower when $M=3$ than $M=1$. This is expected because the uncertainty of the treatment assignment is greater for a stochastic intervention with a longer time period.
We find that the Hájek estimator performs well across all simulation scenarios even when $T$ is small and $M$ is large, whereas the IPW estimator tends to suffer from extreme weights because the weights are multiplied over the intervention time periods as shown in \[eq:estimatort\]. \[app\_subsec:add\_sims\_estimates\] presents additional simulation results for the IPW and Hájek estimators under the interventions $\interv[F][][{M}]$ with $M = 3, 7, 30$, and the intensity $\trtintensity$ varying from 3 to 7. These results indicate a deteriorating performance of the IPW estimator as the value of $M$ increases. In contrast, the standardization of weights used in the Hájek estimator appears to partially alleviate this issue, with its effect estimates much closer to their true values.
Next, we compare the true theoretical variance, ${v}/ T$, with the variance bound ${v}^* / T$ and its consistent estimate (see \[lemma:consistent\_variance\]). Since part of the goal here is to assess the conservativeness of the theoretical variance bound to the theoretical variance, we focus on the proposed estimators with the true propensity score. \[fig:var\_y\_int2\_B2\] shows the results of an intervention $\interv[F][][M]$ for ${M}=3$, and for region $B = [0, 0.5]^2$. First, we focus on the theoretical variance and variance bound (blue line with open circles, and orange dotted lines with open triangles, respectively). As expected, the true variance decreases as the total number of time periods increases. Moreover, the proposed theoretical variance bound tightly follows the theoretical variance, reflected by the fact that the two lines are visually indistinguishable. These results indicate that the theoretical variance bound is not overly conservative compared to the true theoretical variance. Indeed, the variance bound is substantially larger than the true variance only in the low-variance scenarios (the interventions over a single time period and resembling the actual data generating mechanism, as shown in \[app\_subsec:add\_sims\_variance\]).
Second, we compare the theoretical variance bound with the estimated variance bound (green dashed lines with open rhombuses). As the length of time series increases, the estimated variance bound more closely approximates its theoretical value (consistent with \[lemma:consistent\_variance\]). Furthermore, the estimated variance bound is close to its theoretical value under low uncertainty scenarios and when the intervention intensity closely resembles that of the actual data generating process. However, we find that the estimated variance bound underestimates the true variance bound in high uncertainty scenarios, and convergence to its true value is slower for larger values of ${M}$ (see \[app\_subsec:add\_sims\_variance\]).
The results on the asymptotic variance and variance bound in \[fig:var\_y\_int2\_B2\] lead to similar conclusions with respect to the coverage of 95% confidence intervals of the IPW estimator. In \[app\_subsec:add\_sims\_coverage\], we find that except when $M=30$, the confidence interval for the IPW estimator based on either the true asymptotic variance or the true variance bound has a coverage of about 80% or higher. This implies that the asymptotic normality established in \[theorem:normality\] provides an adequate approximation to the estimator’s sampling distribution for small or moderate values of ${M}$. However, for $M=30$, the confidence interval for the IPW estimator is anti-conservative due to the fact that the weights, which equal the product of ratios across many time periods, become extremely small. In addition, the underestimation of the variance bound in high uncertainty scenarios found in \[fig:var\_y\_int2\_B2\] leads to the under-coverage of the confidence intervals based on the IPW estimator and using the estimated variance bound, especially when the interventions take place over long time periods.
In our simulations, we find that the confidence interval for the Hájek estimator discussed in \[app\_sec:hajek\] has a better coverage probability even for the interventions over long time periods (see \[app\_subsec:add\_sims\_coverage\]). \[app\_subsec:add\_sims\_uncertainty\] also shows that the estimated standard deviation for the Hájek estimator outperforms that for the IPW estimator under many simulation scenarios. Partly based on these findings, we use the Hájek estimator and its associated confidence interval in our application (see Section \[sec:application\]).
Finally, we evaluate the performance of the propensity score as a balancing score, as established in \[theorem:balancing\_score\]. In \[app\_subsec:add\_sims\_balance\], we show that the p-values of the previous outcome-active locations variable ($Y_{t - 1}^*$ in \[eq:sims\_treatment\_intensity\]) are substantially greater in the weighted propensity score model than in the unweighted model, where the weights are given by the inverse of the estimated propensity score. We also find the estimated coefficients in the weighted model are centered around zero. These findings are consistent with the balancing property of the propensity score.
Empirical Analyses {#sec:application}
==================
In this section, we present our empirical analyses of the data sets introduced in Section \[sec:iraq\]. We first describe the airstrike strategies of interest and then discuss the causal effect estimates obtained under those strategies.
Airstrike Strategies and Causal Effects of Interest {#subsec:study_interventions}
---------------------------------------------------
For the purpose of this application, we consider hypothesized stochastic interventions that generate airstrike locations based on a simple non-homogeneous Poisson point process with finite and non-atomic intensity $\trtintensity: \Omega \rightarrow [0, \infty)$. We first specify a baseline probability density $\phi_0$ over $\Omega$. To make this baseline density realistic and increase the credibility of the overlap assumption, we use the airstrike data during January 1 – September 24, 2006 to estimate the baseline distribution $\phi_0$ for our stochastic interventions. This subset of the data is not used in the subsequent analysis. The left plot of \[fig:air\_intens\_pre\_blackout\] shows the estimated baseline density using kernel-smoothing of airstrikes, which is more diffused than the estimated spatial distribution of the entire study period (shown in the right plot of the figure).
We consider the following three questions: (1) How does an increase in the number of airstrikes affect insurgent violence? (2) How does the shift in the prioritization of certain locations for airstrikes change the spatial pattern of insurgent attacks? (3) How long does it take for the effects of change in these airstrike strategies to be realized? The last question examines how quickly the insurgents respond to the change in airstrike strategy.
We address the first question by considering stochastic interventions that have the same spatial distribution but vary in the expected number of airstrikes. We represent such strategies using intensities $\trtintensity(\omega) = c\phi_0(\omega)$ with different values of $c > 0$. Since $\int_\Omega \trtintensity(\omega) \mathrm{d}\omega$ represents the expected number of points from a Poisson point process, these interventions have the same spatial distribution $\phi_0$, but the number of airstrikes is monotonically increasing as a function of $c$. In our analysis, we consider $\{1, 2, \dots, 6\}$ as the range of $c$.
For the second question, we fix the intensity but vary the focal locations of airstrikes. To do this, we specify a distribution over $\Omega$ with power-density $d_\alpha(\omega) = d(\omega)^\alpha / \left(\int_\Omega d(\omega)^\alpha \right)$ and modes located at $s_f \in \Omega$. Based on $d_\alpha$, we specify $\trtintensity_\alpha(\omega) = c_\alpha \phi_0(\omega) d_\alpha(\omega)$ where $c_\alpha$ satisfies the constraint $\int_{\Omega} \trtintensity_\alpha(\omega) \text{d}\omega = c$, so that the overall expected number of airstrikes remains constant. Locations in $s_f$ are prioritized under $\trtintensity_\alpha$ as an increasing function of $\alpha$. For our analysis, we choose the center of Baghdad to be the focal point $s_f$ and $d_\alpha$ to be the normal distribution centered at $s_f$ with precision $\alpha$. We set the expected number of airstrikes per day $c$ to be 3, and vary the precision parameter $\alpha$ from 0 to 3. The visualization of the spatial distributions in $\trtintensity_\alpha$ for the different values of $\alpha$ is shown in \[fig:type2\_interv\].
As discussed in Section \[sec:estimands\], for both of these questions, we can specify airstrike strategies of interest taking place over a number of time periods, ${M}$, by specifying the stochastic interventions as $\interv[T][] = \interv[F][][{M}]$. We assume that insurgent attacks at day $t$ do not affect airstrikes on the same day, and airstrikes at day $t$ can only affect attacks during subsequent time periods. Thus, causal quantities for interventions taking place over ${M}$ time periods refer to insurgent attacks occurring on the following day.
In addition, we may also be interested in the lagged effects of airstrike strategies as illustrated by the third question mentioned above. We specify lagged intervention to be the one which differs only for the ${M}$ time periods ago, i.e., $\intervdist[T][] = \intervdist[F][0][{M}-1] \times \intervdist[F][1]$, where $h_0 = \phi_0$ represents the baseline intensity (with $c = 1$), and $h_1 = c\phi_0$ is the increased intensity with different values of $c$ ranging from 1 to 6. For our analysis, we consider ${M}$ to be equal 1 day, 3 days, 1 week, and 1 month.
Finally, in \[subsec:local\_intervention\], we also introduce local interventions, which change the airstrike strategy in one region while leaving that for the rest of the country unchanged.
The Specification and Diagnostics of the Propensity Score Model {#subsec:application_assumptions}
---------------------------------------------------------------
Our propensity score model is a non-homogeneous Poisson point process with intensity $\lambda_t(\omega), \omega \in \Omega$, where $ \log\lambda_t(\omega) = \bm \beta^T \bm {{\bm {X}}}_t(\omega) $, and $\bm {{\bm {X}}}$ includes an intercept and all the covariates. The two main drivers of military decisions over airstrikes are the prior number and locations of observed insurgent attacks and airstrikes, conditional on which unconfoundedness in Assumption \[ass:unmeasured\_conf\] is expected to approximately hold. Therefore, our model includes the observed airstrikes and attacks during the last day, week, and month. For example, the airstrike history of time $t$ during the previous week is specified as $ \anyhist[t-1][W]^*(\omega) = \sum_{j = 1}^7 \sum_{s \in \sparseset[w^{obs}][t-j]} \exp\{ - \mathrm{dist}(s, \omega) \}$, which represents a continuous surface on $\Omega$ with locations closer to the airstrikes in the previous week having greater positive values than more distant locations.
Our propensity score model also includes additional important covariates that might affect both airstrikes and insurgent attacks. For example, we adjust for shows-of-force (i.e., simulated bombing raids designed to deter insurgents) that occurred one day, one week, and one month before each airstrike. Patterns of US aid spending might also affect the location and intensity of insurgent violence and airstrikes, as we discussed in Section \[sec:iraq\]. We therefore include the amount of aid spent (in \$US dollars) in each Iraqi district in the past month as a time-varying covariate. Finally, we also incorporate temporal splines and several time-invariant spatial covariates, including the airstrike’s distance from major cities, road networks, rivers, local settlements, and the population (logged, measured in 2003) of the governorate in which the airstrike took place.
We evaluate the covariate balance by comparing the $p$-values of estimated coefficients in the propensity score model to the $p$-values in the weighted version of the same model, where each time period is inversely weighted by its truncated propensity score estimate (truncated above at the 90$^{th}$ quantile). Although 13 out of 35 estimated coefficients had $p$-values smaller than 0.05 in the fitted propensity score model, all the $p$-values in the weighted propensity score model are close to 1, suggesting that the estimated propensity score adequately balances these confounders.
The Choice of the Bandwidth Parameter for the Spatial Kernel Smoother {#subsec:application_bandwidth}
---------------------------------------------------------------------
The kernel smoothing part of our estimator is not necessary for estimating the number of points within any set $B \subset \Omega$ since we can simply use an IPW estimator based on the observed number of points within $B$. However, kernel smoothing is useful for visualizing the estimated intensities of insurgent attacks under an intervention of interest over the entire country. One can also use it to acquire estimates of the expected number of insurgent attacks under the intervention for any region of Iraq by considering the intensity’s integral over the region. \[theorem:normality\] shows that, for any set $B\subset \Omega$, kernel smoothing does not affect the estimator’s asymptotic normality as long as the bandwidth converges to zero. In practice, the choice of the bandwidth should be partly driven by the size of the sets $B$.
In our analysis, we estimate the causal quantities for the entire country and the Baghdad administrative unit (see \[fig:local\_interv\]). We choose an adaptive bandwidth separately for each outcome using the `spatstat` package in R. We consider all observed outcome event locations during our study period, and use Scott’s criterion for choosing an optimal, constant bandwidth parameter for an isotropic kernel estimation [@scott1992multivariate]. Using the estimated density as the pilot density, we calculate the optimal adaptive bandwidth surface according to Abramson’s inverse-square-root rule [@abramson1982bandwidth]. This procedure yields a value of the bandwidth used for kernel smoothing at each outcome event location.
Findings
--------
\[fig:res\_IED\_SAF\] illustrates changes in the estimated intensity surfaces for insurgent attacks (measured using IEDs and SAFs) when increasing the intensity of airstrikes (the first two rows) and when shifting the focal point of airstrikes to Baghdad (the bottom two rows), with the varying duration of interventions, $M=1,3,7,30$ days (columns). These surfaces can be used to estimate the causal effect of a change in the intervention over any region. Dark blue areas represent areas where the change in the military strategy would reduce insurgent attacks, whereas red areas correspond to those with an increase in insurgent attacks.
The figure reveals a number of interesting findings. First, there is no substantial change in insurgent attacks if these interventions last only for one day. When increasing airstrikes for a longer duration, however, a greater number of insurgent attacks are expected to occur, with the largest number of attacks per day occurring with the 7 day change in the intervention. These changes are concentrated in the Baghdad area and the roads that connect Baghdad and the northern city of Mosul. These estimated effects, measured in terms of attacks per day, appear to decline when the intervention change lasts for 30 days. These patterns apply to both IEDs and SAFs with slightly greater effects estimated for SAFs.
\
\
\
In contrast, when shifting the focal point of airstrikes to Baghdad for 7 days, the total number of insurgent attacks over the entire country decreases. However, when the intervention lasts for 30 days, we find that insurgents shift their attacks to the areas around Mosul while reducing the number of attacks in Baghdad. This displacement pattern is particularly pronounced for SAFs. For IEDs, insurgents appear to move their attacks to the Mosul area even with the intervention of 7 days and yet the effect size is smaller. Statistical significance of these results is shown in \[app\_tab:res\_tau\].
\[fig:res\_tau\_IED\_SAF\_Type-1\] shows the changes in the estimated average number of insurgent attacks [*in Baghdad*]{} as the airstrike intensity increases from 1 to $2, 3, \dots, 6$ airstrikes per day in the entire country (horizontal axis). We also vary the duration of intervention from $M=1$ day to ${M}= 30$ days (columns). Both the point estimate (solid lines) and 95% CIs (grey bands) are shown. Consistent with \[fig:res\_IED\_SAF\], we find that increasing the number of airstrikes leads to a greater number of attacks when the duration of intervention is 7 days. These effects appear to decrease when the intervention is much shorter or lasts for a month. The patterns are similar for both IEDs and SAFs.
\[fig:res\_tau\_IED\_SAF\_Lagged\] shows the change in the estimated number of IEDs and SAFs attacks in Baghdad when increasing the number of airstrikes ${M}$ [*days before*]{}, while the expected number of airstrikes during the following ${M}-1$ days equals one per day. We find that for ${M}= 3$ all estimated lagged effects are negative, whereas the lagged effects for ${M}= 7$ are estimated to be positive. This suggests that increasing the number of airstrikes may reduce insurgent violence in a short term while leading to an increase in a longer term. \[app\_tab:res\_tau\] presents the effect estimates and 95% CIs for various interventions and outcomes.
Concluding Remarks {#sec:discussion}
==================
In this paper, we provide a framework for causal inference with spatio-temporal point process treatments and outcomes. We demonstrate the flexibility of this proposed methodology by applying it to the estimation of airstrike effects on insurgent violence in Iraq. Our central idea is to use a stochastic intervention, one that represents a distribution of treatments, instead of the standard causal inference approach that estimates the average potential outcomes under some fixed treatment values. A key advantage of our approach is its flexibility: it permits unstructured patterns of both spatial spillover and temporal carryover effects. This flexibility is crucial since for many spatio-temporal causal inference problems, including our own application, we know little about how the treatments in one region affect the outcomes in other regions across different time periods.
The proposed methodology combines spatio-temporal point process modeling with the inverse probability of treatment assignment weighting approach, thereby avoiding the direct modeling of the outcome. Instead, we use kernel smoothing to estimate the spatial surface of causal quantities, enabling effective visualization of estimated causal effects using maps. Our approach enables the specification of various military strategies across space and time. We illustrate this advantage by varying the intensity and focal locations of airstrikes and their effects over multiple temporal windows. In addition, we also evaluate how long it takes for these effects to manifest, offering new insights into how quickly insurgents respond to airstrikes and how far their violence is displaced.
The proposed framework can also be applied to other high-dimensional, and possibly unstructured, treatments. The standard approach to causal inference, which bases causal effects on fixed treatment values, cannot perform well in such settings. Indeed, the sparsity of observed treatment patterns alone makes it difficult to satisfy the required overlap assumption [@imai:jian:19]. In short, the stochastic intervention approach proposed here offers an effective solution to a broad class of causal inference problems. Future research should further develop the methodology for stochastic intervention. In particular, it is important to develop a sensitivity analysis for the potential violation of the unconfoundedness assumption, an improved weighting method for balancing covariates, and an inferential approach with better finite sample performance.
[Supplementary Appendix for “Causal Inference with Spatio-Temporal Data”]{}
Notation {#app_sec:notation}
========
[ll L[10.5cm]{}]{}
\
Paths & $\whist$ & Realization of the treatment assignments for time periods $1,\ldots, t$\
& $\anyhist[t][{\mathcal{Y}}]$ & Collection of all potential outcomes for time periods $1,\ldots, t$\
& $\anyhist[t][Y]^{obs}$ & Observed outcomes for time periods $1, \ldots, t$\
Intervention & ${M}$ & The number of time periods over which we intervene\
& $\trtintensity$ & The Poisson point process intensity defining the stochastic intervention\
Estimands & $\numpoints[t][over]$, $\numpoints[][over]$ & Expected number of outcome-active locations for time period $t$ for an intervention over ${M}$ time periods, and their average over time\
& $\tau_t^{M}$, $\tau^{M}$ & Expected change in the number of outcome-active locations comparing two interventions for time period $t$ and their average over time\
Estimators & $\widehat{Y}_t^{M}$ & Estimated continuous surface the integral of which is used for calculating $\numpoints[t][hat]$\
& $\numpoints[t][hat]$, $\numpoints[][hat]$ & Estimated expected number of points during time period $t$ for an intervention taking place over the preceding ${M}$ time periods, and their average over time\
& $\widehat{\tau}_t^{M}$, $\widehat{\tau}^{M}$ & Estimated expected change in the number of outcome-active locations for time period $t$ comparing two interventions, and their average over time\
Arguments & $B$ & The set over which the number of outcome-active locations are counted\
Theoretical Proofs {#app_sec:proofs}
==================
Note that the collection of variables temporally precedent to treatment at time period $t$ is the expanded history $\history^*$, defined in Assumption \[ass:regularity\_conditions\]. The expanded history $\history^*$ is a filtration generated by the collection of potential confounders $\anyhist[T][{\mathcal{{{\bm {X}}}}}]$, the collection of potential outcomes $\anyhist[T][{\mathcal{Y}}]$, and the previous treatments, and satisfies $\history^* \subset \history[t]^*$.
Let $\error = \estimatorNt - \avgout$ be the estimation error for time period $t$ and lag ${M}$. We will decompose $\error$ in two components, one corresponding to error due to the treatment assignment ($A_{1t}$), and one corresponding to error due to spatial smoothing ($A_{2t}$). Since the bandwidth parameter of the kernel depends on $T$, we write $K_{b_T}$ instead of $K_b$. Specifically, [$$\begin{aligned}
\error & \ = \ \Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[][F](W_j)}{\propscore[][j][j][W]} \Bigg]
\int_B \sum_{s \in \sparseset[{}Y]} K_{b_T}(\omega, s) \mathrm{d}\omega
- \avgout[F][][{M}] \\
& \ = \ \underbrace{\Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[][F](W_j)}{\propscore[][j][j][W]} \Bigg] N_B(Y_t(\whist[t-{M}], W_{t - {M}+ 1}, \dots, W_t)) - \avgout[F][][{M}]}_{\quant[1]} + \\
& \hspace{20pt}
\underbrace{\Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[][F](W_j)}{\propscore[][j][j][W]} \Bigg] \Bigg[
\int_B \sum_{s \in \sparseset[{}Y]} K_{b_T}(\omega, s) \mathrm{d}\omega -
N_B(Y_t(\whist[t-{M}], W_{t - {M}+ 1}, \dots, W_t)) \Bigg]}_{\quant[2]}\end{aligned}$$]{} We show that
1. $\sqrt{T}\big(\frac1{T-{M}+1} \sum_{t = {M}}^T \quant[1] \big)$ is asymptotically normal, and
2. $\sqrt{T}\big(\frac1{T-{M}+1} \sum_{t = {M}}^T \quant[2] \big)$ converges to zero in probability.
#### Showing asymptotic normality of the first error.
$\ $\
We use the central limit theorem for martingale difference series (Theorem 4.16 of [@VanDerVaart2010timeseries]) to show asymptotic normality of $(T-{M}+1)^{-1}\sum_{t = {M}}^T \quant[1]$.
$\quant$ is a martingale difference series with respect to the filtration $\filtration[t] = \history[t-{M}+1]^*$.
To prove this, we show that $E(|\quant|) < \infty$ and $E(\quant \mid \filtration) = E(\quant[1][t] \mid \history[t - {M}]^*) = 0$. For the first part, Assumptions \[ass:positivity\] and \[ass:regularity\_conditions\]\[ass:finite\_points\] imply that $\quant$ is bounded and hence $E[|\quant|] < \infty$: $$\begin{aligned}
|\quant| \ &\leq \ \Bigg| \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[][F](W_j)}{\propscore[][j][j][W]} N_B(Y_t(\whist[t-{M}], W_{t - {M}+ 1}, \dots, W_t)) \Bigg| + \Big| \avgout[F][][{M}] \ \Big| \\
& \leq \ \bound[W]^M \bound[Y] + \bound[Y]
\end{aligned}
\label{proof_eq:bounded_A1t}$$ For the second part, it suffices to show that $$E\Bigg\{
\Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[][F](W_j)}{\propscore[][j][j][W]} \Bigg]
N_B(Y_t(\whist[t-{M}], W_{t - {M}+ 1}, \dots, W_t))
\ \Big| \ \history[t - {M}]^* \Bigg \} \ = \ \avgout[F][][{M}],$$ where the expectation is taken with respect to the assignment of treatments $W_{(t - {M}+ 1):t}$. $$\begin{aligned}
& \quad E\Bigg\{ \Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[][F](W_j)}{\propscore[][j][j][W]}\Bigg]
N_B(Y_t(\whist[t-{M}], W_{t - {M}+ 1}, \dots, W_t)) \ \Big | \ \history[t - {M}]^* \Bigg \} \\
&= \int \Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[][F](w_j)}{\propscore[][j][j]}\Bigg]
N_B(Y_t(\whist[t-{M}], \underbrace{w_{t - {M}+ 1}, \dots, w_t}_{{\bm{w}}_{(t-M+1):t}})) \times \\
& \hspace{0.4in}
f(w_{t - {M}+ 1} \mid \history[t-{M}]^*)
f(w_{t - {M}+ 2} \mid \history[t-{M}]^*, W_{t - {M}+ 1}) \cdots \times \\
& \hspace{0.4in}
f(w_t \mid \history[t-{M}]^*, W_{(t - {M}+ 1):(t - 1)}) \ \mathrm{d}{\bm{w}}_{(t-M+1):t} \\
&= \int \Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[][F](w_j)}{\propscore[][j][j]}\Bigg]
N_B(Y_t(\whist[t-{M}], w_{t - {M}+ 1}, \dots, w_t) \times \\
& \hspace{0.4in}
f(w_{t - {M}+ 1} \mid \history[t-{M}]^*)
f(w_{t - {M}+ 2} \mid \history[t-{M}+1]^*) \cdots
f(w_t \mid \history[t-1]^*) \ \mathrm{d}{\bm{w}}_{(t-M+1):t}
\tag{because $\history[t' +1]^* = \history[t']^* \cup \{W_{t' + 1}\}$} \\
&= \int N_B(Y_t(\whist[t-{M}], w_{t - {M}+ 1}, \dots, w_t))
\left[ \prod_{j = t - {M}+ 1}^t \intervdistf[][F](w_j) \right] \ \mathrm{d}{\bm{w}}_{(t-M+1):t}
\tag{By \cref{ass:unmeasured_conf}} \\
&= \avgout[F][][{M}].\end{aligned}$$ This establishes the claim that $\quant$ is a martingale difference series with respect to filtration $\filtration$.
$(T - {M}+ 1)^{-1} \sum_{t = {M}}^T E\{\quant^2 I(|\quant| > \epsilon \sqrt{T - {M}+ 1}) \mid \filtration\} \overset{p}{\rightarrow} 0$ for every $\epsilon > 0$.
Let $\epsilon > 0$. Note that $\quant$ is bounded by $\bound[Y](\bound[W]^{M}+1)$ (see \[proof\_eq:bounded\_A1t\]). Choose $T_0$ as $$\begin{aligned}
T_0 & \ = \ \underset{t \in \mathbb{N}^+}{\text{argmin}} \{ \epsilon\sqrt{t - {M}+ 1} > \bound[Y](\bound[W]^{M}+ 1) \} \\
& \ = \ \underset{t \in \mathbb{N}^+}{\text{argmin}} \Big\{ t > {M}- 1 + \Big[ \frac{\bound[Y](\bound[W]^{M}+ 1)}{\epsilon} \Big]^2 \Big\} \\
& \ = \ \ceil[\Bigg]{{M}- 1 + \Big[ \frac{\bound[Y](\bound[W]^{M}+ 1)}{\epsilon} \Big]^2}\end{aligned}$$ Then, for $T > T_0$, $\epsilon \sqrt{T - {M}+ 1} > \epsilon \sqrt{T_0 - {M}+ 1} > \bound[Y](\bound[W]^{M}+1)$ which leads to $I(|\quant| > \epsilon \sqrt{T+{M}+1}) = 0$ and $ E(\quant^2 I(|\quant| > \epsilon \sqrt{T - {M}+ 1}) \mid \filtration) = 0. $ This proves the claim.\
*Combining the previous results to show asymptotic normality of the first error*: Since $\quant$ has mean zero, $E(\quant^2 \mid \filtration) = {\text{Var}}(\quant \mid \filtration)$, and since $\avgout$ is fixed, ${\text{Var}}(\quant \mid \filtration) = {\text{Var}}(\estimatorNt \mid \filtration) = {\text{Var}}(\estimatorNt \mid \history[t-{M}]^*)$. This gives us that $$\frac1{T-{M}+1} \sum_{t = {M}}^T E(\quant^2 \mid \filtration) =
\frac1{T-{M}+1} \sum_{t = {M}}^T {\text{Var}}(\estimatorNt[F][][{M}] \mid \history[t-{M}]^*) \overset{p}{\rightarrow} {v},$$ from Assumption \[ass:regularity\_conditions\]\[ass:convergent\_variance\]. Combining these results, using that $\sqrt{T} / \sqrt{T - {M}+ 1} \rightarrow 1$ and Theorem 4.16 of [@VanDerVaart2010timeseries], $$\sqrt{T} \left( \frac{1}{T - {M}+ 1} \sum_{t = {M}}^T \quant \right) \overset{d}{\rightarrow} N(0, v).$$
#### Showing convergence to zero of the second error.
$\ $\
The second error compares the integral of the kernel-smoothed outcome surface over the region of interest $B$ with the actual number of points within the set $B$. We show that as $T$ goes to infinity, and since the bandwidth of the kernel converges to 0, the error due to kernel smoothing also goes to zero. Specifically, we will show that $$\sqrt{T} \left(\frac{1}{T-{M}+1}\sum_{t = {M}}^T \quant[2] \right) \overset{p}{\rightarrow} 0.$$
We start by introducing some notation and making an assumption about the set $B$. For $\epsilon > 0$, we use $\mathcal{N}_\epsilon(A)$ to denote the $\epsilon-$neighborhood of a set $A$: $\mathcal{N}_\epsilon(A) = \{\omega \in \Omega: \text{there exists } a \in A \text{ with } \text{dist}(\omega, a) < \epsilon\}$. Also, we use $\boundary$ to denote the boundary of $B$ (its closure excluding the interior points), $\boundary = \overline{B} \backslash B^o$.
There exists $\bound[B] > 0$ and $Q^* \in (1/2, 1)$ such that $$P \left( \sum_{t = {M}}^T I\Big( \exists s \in \sparseset[{}Y] \cap \mathcal{N}_{\bound[B]}(\boundary) \Big) > T^{1-Q^*} \right) \rightarrow 0, \text{ as } T \rightarrow \infty.$$ \[ass:neighborhood\_boundary\]
\[ass:neighborhood\_boundary\] states that the probability that we observe more than $T^{1-Q^*}$ time periods with outcome-active locations within a $\bound[B]-$neighborhood of $B$’s boundary goes to zero as the number of observed time periods increases.
Let $c_t = \prod_{j = t - {M}+ 1}^t \intervdistf[][F](W_j) / \propscore[][j][j][W]$, and write $$\begin{aligned}
\left| \frac1{T-{M}+1} \sum_{t = {M}}^T \quant[2] \right|
&= \left| \frac1{T-{M}+1} \sum_{t = {M}}^T c_t
\left[ \int_B \sum_{s \in \sparseset[{}Y]} K_{b_T}(\omega;s) \mathrm{d}\omega - N_B(Y_t) \right] \right|.\end{aligned}$$
Then: $$\begin{aligned}
& \int_B \sum_{s \in \sparseset[{}Y]} K_{b_T}(\omega;s) \mathrm{d}\omega - N_B(Y_t) \\
= \ & \sum_{s \in \sparseset[{}Y]\cap B} \int_B K_{b_T}(\omega;s) \mathrm{d}\omega +
\sum_{s \in \sparseset[{}Y] \cap B\complement} \int_B K_{b_T}(\omega;s) \mathrm{d}\omega -
N_B(Y_t) \\
= \ & \sum_{s \in \sparseset[{}Y]\cap B} \Big[1 - \int_{B\complement} K_{b_T}(\omega;s) \mathrm{d}\omega \Big] +
\sum_{s \in \sparseset[{}Y] \cap B\complement} \int_B K_{b_T}(\omega;s) \mathrm{d}\omega -
N_B(Y_t) \\
=\ & \sum_{s \in \sparseset[{}Y] \cap B\complement} \int_B K_{b_T}(\omega;s) \mathrm{d}\omega -
\sum_{s \in \sparseset[{}Y]\cap B} \int_{B\complement} K_{b_T}(\omega;s) \mathrm{d}\omega.\end{aligned}$$ This shows that the error from smoothing the outcome surface at time $t$ comes from (1) the kernel weight from points outside of $B$ that falls within $B$, and (2) the kernel weight from points inside $B$ that falls outside $B$. Using this, we write: $$\begin{aligned}
& \left| \frac1{T-{M}+1} \sum_{t = {M}}^T \quant[2] \right| \ = \\
& \hspace{20pt} \left| \frac1{T-{M}+1} \sum_{t = {M}}^T c_t \Bigg[ \sum_{s \in \sparseset[{}Y] \cap B\complement} \int_B K_{b_T}(\omega;s) \mathrm{d}\omega -
\sum_{s \in \sparseset[{}Y]\cap B} \int_{B\complement} K_{b_T}(\omega;s) \mathrm{d}\omega \Bigg] \right|.\end{aligned}$$
Take $\epsilon > 0$, and $Q \in (1/2, Q^*)$ where $Q^*$ is the one in Assumption \[ass:neighborhood\_boundary\]. Then, we will show that $P ( T^Q \{ | \frac1{T-{M}+1} \sum_{t = {M}}^T \quant[2] |
\} > \epsilon) \rightarrow 0$ as $T \rightarrow \infty$, which implies that the second error converges to zero faster than $\sqrt{T}$ (since $Q > 1/2$). [$$\begin{aligned}
& P \Bigg( T^Q \Bigg\{ \Bigg| \frac1{T-{M}+1} \sum_{t = {M}}^T \quant[2] \Bigg|
\Bigg\} > \epsilon \Bigg) \\
= \ & P \Bigg( \Bigg| \frac1{T-{M}+1} \sum_{t = {M}}^T c_t \Bigg[ \sum_{s \in \sparseset[{}Y] \cap B\complement} \int_B K_{b_T}(\omega;s) \mathrm{d}\omega -
\sum_{s \in \sparseset[{}Y]\cap B} \int_{B\complement} K_{b_T}(\omega;s) \mathrm{d}\omega \Bigg] \Bigg| > \frac{\epsilon}{T^Q} \Bigg) \\
\leq \ & P \Bigg( \frac1{T-{M}+1} \sum_{t = {M}}^T c_t \sum_{s \in \sparseset[{}Y] \cap B\complement} \int_B K_{b_T}(\omega;s) \mathrm{d}\omega > \frac{\epsilon}{2 T^Q} \Bigg) + \\
& \hspace{40pt}
P \Bigg( \frac1{T-{M}+1} \sum_{t = {M}}^T c_t
\sum_{s \in \sparseset[{}Y]\cap B} \int_{B\complement} K_{b_T}(\omega;s) \mathrm{d}\omega > \frac{\epsilon}{2T^Q} \Bigg),\end{aligned}$$ where the last equation holds because $|A - B| > \epsilon$ implies that at least one of $|A|, |B| > \epsilon / 2$. Also, since all quantities are positive, we can drop the absolute value. Then, since $c_t \leq \bound[W]^{M}$ from Assumption \[ass:positivity\], $$\begin{aligned}
& P \Bigg( T^Q \Bigg\{ \Bigg| \frac1{T-{M}+1} \sum_{t = {M}}^T \quant[2] \Bigg|
\Bigg\} > \epsilon \Bigg) \\
\leq \ & P \Bigg( \frac1{T-{M}+1} \sum_{t = {M}}^T \sum_{s \in \sparseset[{}Y] \cap B\complement} \int_B K_{b_T}(\omega;s) \mathrm{d}\omega > \frac{\epsilon}{2 T^Q \bound[W]^{M}} \Bigg) + \\
& \hspace{60pt}
P \Bigg(\frac1{T-{M}+1} \sum_{t = {M}}^T
\sum_{s \in \sparseset[{}Y]\cap B} \int_{B\complement} K_{b_T}(\omega;s) \mathrm{d}\omega > \frac{\epsilon}{2 T^Q \bound[W]^{M}} \Bigg).\end{aligned}$$]{}
Use $\spoint[out]$ to denote the point in $\sparseset[{}Y]$ that lies outside B and is the closest to $B$: $\displaystyle \spoint[out] = \{s \in \sparseset[{}Y] \cap B\complement: \text{dist}(s, B) = \min_{s' \in \sparseset[{}Y] \cap B\complement} \text{dist}(s', B) \} $. Similarly, $\spoint$ is the point in $\sparseset[{}Y] \cap B$ that is closest to $B\complement$. These points are shown graphically in \[app\_fig:proof\_outline\].
![Kernel-smoothed outcome surface, and points $\spoint[in], \spoint[out]$ as the points closest to the boundary of $B$ that lie within and outside $B$ respectively. The amount of kernel weight falling within $B$ from points outside of $B$ is necessarily less or equal to the kernel weight from $\spoint[out]$ (shaded), and similarly for $\spoint[in]$.[]{data-label="app_fig:proof_outline"}](fig_vis_proof_outline.png){width="40.00000%"}
Because there are at most $\bound[Y]$ outcome-active locations, from the definition of $\spoint[in], \spoint[out]$, and because kernels are defined to be decreasing in distance, we have that $$\begin{aligned}
P \Bigg( T^Q \Bigg\{ \Bigg| & \frac1{T-{M}+1} \sum_{t = {M}}^T \quant[2] \Bigg|
\Bigg\} > \epsilon \Bigg) \\
\leq \ & P \Bigg( \frac1{T-{M}+1} \sum_{t = {M}}^T \int_B K_{b_T}(\omega; \spoint[out]) \mathrm{d}\omega > \frac{\epsilon}{2 T^Q \bound[W]^{M}\bound[Y]} \Bigg) \\
& \hspace{40pt} +
P \Bigg( \frac1{T-{M}+1} \sum_{t = {M}}^T
\int_{B\complement} K_{b_T}(\omega; \spoint[in]) \mathrm{d}\omega > \frac{\epsilon}{2 T^Q \bound[W]^{M}\bound[Y]} \Bigg) \\
= \ & \underbrace{P \Bigg( \sum_{t = {M}}^T \int_B K_{b_T}(\omega; \spoint[out]) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} \Bigg)}_{\quantt} \\
& \hspace{40pt} +
\underbrace{P \Bigg( \sum_{t = {M}}^T
\int_{B\complement} K_{b_T}(\omega; \spoint[in]) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} \Bigg)}_{\quantt[2]}.\end{aligned}$$ We show that $\quantt, \quantt[2]$ converge to zero separately. Take $\quantt$: [$$\begin{aligned}
\quantt \ & = \ P \Bigg( \sum_{t = {M}}^T \int_B K_{b_T}(\omega; \spoint[out]) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} \ \Bigg\vert \ \sum_{t={M}}^T I(\spoint[out] \in \mathcal{N}_{\bound[B]}(\boundary) ) > T^{1-Q^*} \Bigg) \\
& \hspace{6cm}
\times \ P \Bigg( \sum_{t={M}}^T I(\spoint[out] \in \mathcal{N}_{\bound[B]}(\boundary)) > T^{1-Q^*} \Bigg) \\
& \hspace{0.5cm} + P \Bigg( \sum_{t = {M}}^T \int_B K_{b_T}(\omega; \spoint[out]) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} \ \Bigg\vert \ \sum_{t={M}}^T I(\spoint[out] \in \mathcal{N}_{\bound[B]}(\boundary)) \leq T^{1-Q^*} \Bigg) \\
& \hspace{6cm}
\times \ P \Bigg( \sum_{t={M}}^T I(\spoint[out] \in \mathcal{N}_{\bound[B]}(\boundary)) \leq T^{1-Q^*} \Bigg)\end{aligned}$$]{} From Assumption \[ass:neighborhood\_boundary\] we have that $$\begin{aligned}
P \Bigg( \sum_{t={M}}^T & I(\spoint[out] \in \mathcal{N}_{\bound[B]}(\boundary)) > T^{1-Q^*} \Bigg)
\\
& \leq \ P \left( \sum_{t = {M}}^T I\Big( \exists s \in \sparseset[{}Y] \cap \mathcal{N}_{\bound[B]}(\boundary) \Big) > T^{1-Q^*} \right)
\rightarrow 0,\end{aligned}$$ and $\lim_{T \rightarrow \infty} \quantt$ is equal to $$\begin{aligned}
\lim_{T \rightarrow \infty}
P \Bigg( \sum_{t = {M}}^T \int_B K_{b_T}(\omega; \spoint[out]) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} \ \Bigg\vert \ \sum_{t={M}}^T I(\spoint[out] \in \mathcal{N}_{\bound[B]}(\boundary)) \leq T^{1-Q^*} \Bigg).\end{aligned}$$ Studying the latter quantity, we have that $$\begin{aligned}
& \quad P \Bigg( \sum_{t = {M}}^T \int_B K_{b_T}(\omega; \spoint[out]) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} \Bigg\vert \sum_{t={M}}^T I(\spoint[out] \in \mathcal{N}_{\bound[B]}(\boundary)) \leq T^{1-Q^*} \Bigg) \\
& \leq P \Bigg( \hspace{20pt} \sum_{\mathclap{\substack{t = {M}\\ \spoint[out] \not\in \mathcal{N}_{\bound[B]}(\boundary)}}}^T \hspace{20pt}
\int_B K_{b_T}(\omega; \spoint[out]) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} - T^{1-Q^*} \Bigg) \\
&
\leq P \Bigg( \hspace{20pt} \sum_{\mathclap{\substack{t = {M}\\ \spoint[out] \not\in \mathcal{N}_{\bound[B]}(\boundary)}}}^T \hspace{20pt}
\int_{\omega: \|\omega\| > \bound[B]} K_{b_T}(\omega; \bm 0) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} - T^{1-Q^*} \Bigg) \\
& \leq P \Bigg( (T-{M}+1)
\int_{\omega: \|\omega\| > \bound[B]} K_{b_T}(\omega; \bm 0) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} - T^{1-Q^*} \Bigg)\\
&= I \Bigg( (T-{M}+1)
\int_{\omega: \|\omega\| > \bound[B]} K_{b_T}(\omega; \bm 0) \mathrm{d}\omega > \frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} - T^{1-Q^*} \Bigg)
{\addtocounter{equation}{1}\tag{{A.\arabic{equation}}}}\label{proof_eq:indicator}\end{aligned}$$ where the first inequality follows from the fact that at most $T^{1-Q^*}$ time periods had $\spoint[out]$ within $\bound[B]$ of set’s $B$ boundary, and $\int_B K_{b_T}(\omega; \spoint[out]) \leq 1$ for those time periods. The second inequality follows from the fact that during the remaining time periods $\spoint[out]$ was further than $\bound[B]$ from $B$ and $\int_B K_{b_T}(\omega;\spoint[out]) \leq \int_{\omega: \|\omega - \spoint[out]\| > \bound[B]} K_{b_T}(\omega; \spoint[out]) = \int_{\omega: \|\omega\| > \bound[B]} K_{b_T}(\omega; \bm 0)$. The third inequality follows from not excluding the time periods with $\spoint[out] \in \mathcal{N}_{\bound[B]}(\boundary)$. Finally, the last equality holds because there is no uncertainty in the statement so the probability turns to an indicator.
Since the bandwidth $b_T \rightarrow 0$ as $T \rightarrow 0$, there exists $T_1 \in \mathbb{N}$ such that $b_T < \bound[B]$ and $\int_{\omega: \|\omega\| > \bound[B]} K_{b_T}(\omega; \bm 0) \mathrm{d}\omega = 0$ for all $T \geq T_1$. Also, since $\frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} - T^{1-Q^*} \rightarrow \infty$, there exists $T_2 \in \mathbb{N}$ such that $\frac{\epsilon (T-{M}+1)}{2 T^Q \bound[W]^{M}\bound[Y]} - T^{1-Q^*} > 1$ for all $T \geq T_2$. Then, for all $T \geq T_0 = \max\{T_1, T_2\}$ we have that the quantity in \[proof\_eq:indicator\] is equal to 0, showing that $\lim_{T\rightarrow \infty} B_1 = 0$. Similarly, we can show that $\lim_{T\rightarrow \infty} B_2 = 0$.
Combining all of these results we have that $$P \Bigg( T^Q \Bigg\{ \Bigg| \frac1{T-{M}+1} \sum_{t = {M}}^T \quant[2] \Bigg|
\Bigg\} > \epsilon \Bigg) \rightarrow 0,$$ as $T \rightarrow \infty$, establishing that the second error converges to zero faster than $1/\sqrt{T}$.
Define $\Psi_t = \big[ \estimatorNt[F][][{M}] \big]^2 - {v}_t^*$. Then, $\Psi_t$ is a martingale difference series with respect to $\filtration[t] = \history[t - {M}+ 1]$:
$E(|\Psi_t|) < \infty$ since $\Psi_t$ is bounded, and
$ \text{E}(\Psi_t \mid \filtration) = \text{E} \Big\{ \big[ \estimatorNt[F][][{M}] \big]^2 \mid \history[t-{M}]^* \Big\} - {v}_t^* = 0. $
Also, since $\estimatorNt[F][][{M}]$ is bounded we have that $\sum_{t= {M}}^\infty t^{-2} \text{E}\big(\Psi_t^2\big) < \infty.$ From Theorem 1 in [@Csorgo1969] we have that $$\frac1{T-{M}+1} \sum_{t = {M}}^T \Psi_t = \frac1{T-{M}+1} \sum_{t = {M}}^T \big[ \estimatorNt[F][][{M}] \big]^2 - \frac1{T-{M}+1} \sum_{t = {M}}^T {v}_t^* \overset{p}{\rightarrow} 0.$$
In order to prove asymptotic normality of $\tempeffect[T][F][1][2]$ we will rely on results in the proof of \[theorem:normality\] above. Take [$$\begin{aligned}
& \effect[T][F][1][2] - \effect[F][F][1][2] = \\
& \Bigg\{ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[2][F](W_j)}{\propscore[][j][j][W]} -
\prod_{j = t - {M}+ 1}^t \frac{\intervdistf[1][F](W_j)}{\propscore[][j][j][W]} \Bigg\}
\int_B \sum_{s \in \sparseset[{}Y]} K_{b_T}(\omega, s) \mathrm{d}\omega
- \effect[F][F][1][2] = \\
& \underbrace{
\Bigg\{ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[2][F](W_j)}{\propscore[][j][j][W]} -
\prod_{j = t - {M}+ 1}^t \frac{\intervdistf[1][F](W_j)}{\propscore[][j][j][W]} \Bigg\}
N_B(Y_t(\whist[t-{M}], W_{t - {M}+ 1}, \dots, W_t)) - \effect[F][F][1][2]}_{\quanttt[1][]} + \\
& \hspace{20pt}
\underbrace{\Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[2][F](W_j)}{\propscore[][j][j][W]} \Bigg] \Bigg[ \int_B \sum_{s \in \sparseset[{}Y]} K_{b_T}(\omega, s) \mathrm{d}\omega -
N_B(Y_t(\whist[t-{M}], W_{t - {M}+ 1}, \dots, W_t)) \Bigg]}_{\quanttt[2][2]} + \\
& \hspace{20pt}
\underbrace{\Bigg[ \prod_{j = t - {M}+ 1}^t \frac{\intervdistf[1][F](W_j)}{\propscore[][j][j][W]} \Bigg] \Bigg[\int_B \sum_{s \in \sparseset[{}Y]} K_{b_T}(\omega, s) \mathrm{d}\omega -
N_B(Y_t(\whist[t-{M}], W_{t - {M}+ 1}, \dots, W_t)) \Bigg]}_{\quanttt[2][1]}\end{aligned}$$]{} Following steps identical to showing $\sqrt{T} \Big[(T-{M}+1)^{-1}\sum_{t = {M}}^T \quant[2] \Big] \overset{p}{\rightarrow} 0$ in the proof of \[theorem:normality\], we can equivalently show that $ \sqrt{T} \Big[(T-{M}+1)^{-1}\sum_{t = {M}}^T \quanttt[2][1] \Big] \overset{p}{\rightarrow} 0$ and $ \sqrt{T} \Big[(T-{M}+1)^{-1}\sum_{t = {M}}^T \quanttt[2][2] \Big] \overset{p}{\rightarrow} 0$.
Therefore, all we need to show is that $ \sqrt{T} \Big[(T-{M}+1)^{-1}\sum_{t = {M}}^T \quanttt[1][] \Big] \overset{d}{\rightarrow} N(0, \eta).$ We will do so by showing again that $\quanttt[1][]$ is a martingale difference series with respect to the filtration $\filtration$:
1. Since $E(|\quant|) < \infty$, from the triangular inequality we straightforwardly have that $E(|\quanttt[1][]|) < \infty$.
2. Since $E(\quant | \filtration[t - 1]) = 0$, we also have that $E(\quanttt[1][] | \filtration[t - 1]) = 0$, from linearity of expectation.
Then, using the triangular inequality and \[proof\_eq:bounded\_A1t\], we have that $\quanttt[1][]$ is bounded by $2 \bound[Y](\bound[W]^{M}+1)$. Then, for $\epsilon > 0$, choosing $T_0 = \underset{t \in \mathbb{N}^+}{\text{argmin}} \{ \epsilon\sqrt{t - {M}+ 1} > 2 \bound[Y](\bound[W]^{M}+ 1) \}$ satisfies that, for $T > T_0$, $ E(\quanttt[1][2] I(|\quanttt[1][]| > \epsilon \sqrt{T - {M}+ 1}) | \filtration) = 0. $ Combining these results, we have that $\sqrt{T}\big[\effect[T][F][1][2] - \effect[F][F][1][2] \big] \rightarrow N(0, \eta)$.
To show $(T - {M}+ 1)^{-1} \sum_{t = {M}}^T \Big\{ \big[\effect[T][F][1][2] \big] ^2 - E \big\{ \big[\effect[F][F][1][2] \big]^2 | \history[t - {M}]^*\big\} \Big\} \overset{p}{\rightarrow} 0$, the proof follows exactly the same way as the proof of \[lemma:consistent\_variance\] and is omitted here.
Note that $f(W_t = w \mid \propscore[], \history) = f(W_t = w \mid \history) = \propscore[]$ since $\propscore[]$ is a function of $\history$. Therefore, it suffices to show that $f(W_t = w \mid \propscore[]) = \propscore[]$: $$\begin{aligned}
f(W_t = w \mid \propscore[]) &= E[f(W_t = w \mid \history) \mid \propscore[]] = E[\propscore[] \mid \propscore[]] = \propscore[].
\label{app_eq:in_proof_balance}\end{aligned}$$
We show that $f \big(W_t = w \mid \propscore[], \anyhist[T][{\mathcal{Y}}], \anyhist[T][{\mathcal{{{\bm {X}}}}}] \big) = f(W_t = w \mid \propscore[])$. $$\begin{aligned}
& f \big(W_t = w \mid \propscore[], \anyhist[T][{\mathcal{Y}}], \anyhist[T][{\mathcal{{{\bm {X}}}}}] \big) \\
&=
E \Big[ f \big(W_t = w \mid \history, \propscore[], \anyhist[T][{\mathcal{Y}}] , \anyhist[T][{\mathcal{{{\bm {X}}}}}] \big) \mid \propscore[], \anyhist[T][{\mathcal{Y}}] , \anyhist[T][{\mathcal{{{\bm {X}}}}}] \Big] \\
&=
E \Big[ f \big(W_t = w \mid \history, \anyhist[T][{\mathcal{Y}}] , \anyhist[T][{\mathcal{{{\bm {X}}}}}] \big) \mid \propscore[], \anyhist[T][{\mathcal{Y}}] , \anyhist[T][{\mathcal{{{\bm {X}}}}}] \Big] \tag{Because $\propscore[]$ is a function of $\history$} \\
&=
E \Big[ f \big(W_t = w \mid \history \big) \mid \propscore[], \anyhist[T][{\mathcal{Y}}] , \anyhist[T][{\mathcal{{{\bm {X}}}}}] \Big] \tag{From \cref{ass:unmeasured_conf}} \\
&=
E \Big[ \propscore[] \mid \propscore[], \anyhist[T][{\mathcal{Y}}] , \anyhist[T][{\mathcal{{{\bm {X}}}}}] \Big] \\
&= \propscore[] \\
&= f(W_t = w \mid \propscore[]) \tag{From \cref{app_eq:in_proof_balance}}\end{aligned}$$
The Hájek Estimator {#app_sec:hajek}
===================
The standardization of weights used in the Hájek estimator is known to be effective in the settings where the weights are extreme. Its sample boundedness property gurantees that the resulting estimate is always within the range of the observed outcome. In our case, the Hájek estimator replaces the division by $T - {M}+ 1$ with that by $\sum_{t = {M}}^T v_t$ where $v_t$ is the product of fractions in \[eq:estimatort\]. For example, $$\begin{aligned}
\estimatorN_{\text{H\'ajek}} & \ = \ \frac{1}{\sum_{t = {M}}^T v_t} \sum_{t = {M}}^T \estimatorNt[F][][{M}] $$ Unfortunately, our asymptotic results (i.e., Theorems \[theorem:normality\] and \[theorem:normality\_tau\]) do not easily extend to the Hájek estimator. The reason is that the estimator includes $\sum_{t = {M}}^T v_t$, a quantity that depends on all time periods, making it difficult to apply the martingale theory used in our proofs. Therefore, we use a heuristic approach to estimating the variance bound of the Hájek estimator. Since the Hájek estimator simply rescales the corresponding IPW estimator by $(T-{M}+1)/\sum_{t = {M}}^T v_t$, we scale the variance bound derived for the estimator by $[(T-{M}+1)/(\sum_{t = {M}}^T v_t)]^2$.
Additional Simulation Results {#app_sec:add_sims}
=============================
Point Estimates for the Interventions over Many Time Periods {#app_subsec:add_sims_estimates}
------------------------------------------------------------
\[app\_fig:res\_sims\_largeM\] shows the performance of the estimators for the interventions over many time periods. The plots show the estimated change in the number of outcome-active locations over the sub-region $B = [0.75, 1]$ for a change in the stochastic intervention from 3 per time period to the value on the horizontal axis. The rows correspond to the interventions over $M = 3 , 7$, and $30$ time periods, respectively, whereas the columns represent the different lengths of time series, i.e., $T = 200, 400$ and $500$. The results are shown for the IPW estimators based on the true propensity score (purple lines with open triangles) and the estimated propensity score (blue lines with x’s) as well as the Hájek estimator based on the estimated propensity score (orange lines with open rhombuses). Only the Hájek estimates are shown for $M = 30$ as the extremely small weights arising from a large number of time periods make the estimates from the other estimators essentially equal to zero. The lines and points in the plot show the median estimate and the rectangles show the interquartile range of estimates across 200 simulated data sets.
\
\
\
![Simulation Results for the Interventions of Increasing Time Lengths. Rows correspond to the interventions taking place over $M=3, 7$, and $30$ time periods. Columns correspond to the increasing length of the time series from 200 (left plots) to 500 (right plots). The vertical axis shows the change in the expected number of the outcome active locations over $[0.75, 1]^2$ for a change in the intervention intensity from 3 under $\trtintensity_1$ to the value shown in the horizontal axis under $\trtintensity_2$, for $M$ time periods. The points in the plot show the median estimate over 200 data sets, and the rectangles show the interquartile range of estimates. Only the Hájek estimates are shown for $M = 30$ as the extremely small weights arising from a large number of time periods make the estimates from the other estimators close to zero.[]{data-label="app_fig:res_sims_largeM"}](fig_sims_legend2.pdf "fig:"){width="85.00000%"}
The results indicate that the IPW estimator’s performance deteriorates as the length of the intervention period $M$ increases. This is expected since the weight at each time point is the product of $M$ ratios as shown in \[eq:estimatort\]. The data provide little information about what would happen under a hypothetical intervention taking place over a long time period unless the specified intervention resembles the actual data generating mechanism for the observed data. However, \[app\_fig:res\_sims\_largeM\] shows that the standardization of weights performed in the Hájek estimator can alleviate some of these issues, yielding effect estimates close to the true values.
Asymptotic Variance and Bound, and Estimated Variance Bound {#app_subsec:add_sims_variance}
-----------------------------------------------------------
\[fig:var\_y\_int2\_B2\] shows the average (over 200 simulated data sets) of the true asymptotic standard deviation and true bound as well as the estimated standard deviation bound of the IPW estimator for the average potential outcome using the true propensity score, for interventions taking place over $M = 3$ time periods. \[app\_fig:var\_y\_other\_B2\] is a similar plot for the interventions taking place over $M = 1, 3$, and $7$ (rows) time periods, and observed time series of length $T = 200, 400, 500$ (columns). These plots show the median and interquartile range of the asymptotic standard deviation, true bound, and estimated bound over 200 simulated data sets.
\
\
We begin by focusing on low uncertainty scenarios, corresponding to the interventions taking place over $M = 1$ or $3$ time periods with the distribution resembling the actual data generating mechanism. We think that the intervention distribution resembles the data generating mechanism in scenarios where the intervention intensity is close to 5, which is the average number of treatment-active locations for the data generating process. In these scenarios, the asymptotic variance bound is distinctly higher than the true asymptotic variance, indicating that the inference based on the true asymptotic bound would be conservative. We find that in these low uncertainty scenarios, the estimated bound is close to the true bound. For that reason, we would expect the confidence intervals for the IPW estimator based on the estimated bound to have a higher coverage probability than its nominal coverage (see \[app\_subsec:add\_sims\_coverage\] for the coverage results).
In contrast, under high uncertainty scenarios such as the interventions over longer time periods, e.g., $M=7$, the asymptotic standard deviation and theoretical bound are essentially indistinguishable. However, under these scenarios, the estimate of the theoretical bound tends to be biased downwards, suggesting that the confidence intervals for the IPW estimator based on the estimated bound would be anti-conservative. Furthermore, we expect it to take a longer time series in order for the estimated bound to converge to its theoretical value when the intervention takes place over a longer time period.
Coverage of the Confidence Intervals for the IPW and Hájek Estimators {#app_subsec:add_sims_coverage}
---------------------------------------------------------------------
#### IPW estimator.
The results in \[app\_fig:var\_y\_other\_B2\] indicate that the coverage of confidence intervals based on the asymptotic variance bound should be similar to those based on the true variance under high uncertainty scenarios, while they should be slightly higher under low uncertainty scenarios. Furthermore, confidence intervals based on the estimated variance bound should yield coverage probability close to (lower than) the nominal coverage under low (high) uncertainty scenarios.
\
These expectations are indeed reflected in the coverage results shown in \[app\_fig:ipw\_cover\]. Furthermore, for $M \in \{1, 3, 7\}$ and $T = 500$, the confidence intervals using the true variance achieve a coverage that is close to the nominal level. In contrast, for $M = 30$, the confidence intervals based on the true asymptotic variance have a coverage of 50% or less, indicating that for interventions taking place over longer time periods, more data are needed to make use of the asymptotic approximation.
#### Hájek estimator.
Motivated by the good performance of the Hájek estimator shown in \[app\_fig:res\_sims\_largeM\], we also investigate the coverage probability of the 95% confidence interval as described in Appendix \[app\_sec:hajek\]. The rows of \[app\_fig:hajek\_cover\] show the coverage results for increasingly small regions, i.e., $B_1=[0,1]^2, B_2=[0, 0.5]^2$, and $B_3=[0.75, 1]^2$, whereas the columns show the results for increasingly long observed time series ($T = 200, 400, 500$). Different colors correspond to the coverage results under interventions taking place over $M = 1$ (black), $3$ (green), $7$ (red), and $30$ (blue) time periods. We find that the coverage is above 85% for all cases, even when an intervention takes place over 30 time periods. As expected, the coverage is higher for smaller values of $M$, since these correspond to lower-uncertainty situations. We also find that the coverage is lower for smaller regions, a phenomenon we do not observe for the IPW estimator.
Uncertainty Estimates {#app_subsec:add_sims_uncertainty}
---------------------
\
We also compute the standard deviation of the estimated average potential outcome across simulated data sets and compare it with the mean of the standard deviations, each of which is used to create the confidence intervals. The similarity of these two quantities implies the accuracy of our uncertainty estimates. \[app\_fig:MCsd\_meansd\] presents the results as the ratio of these two quantities. A value below (above) 1 indicates that the true variability in our point estimates is smaller (greater) than our uncertainty estimate.
While the ratios are always below 1 for the Hájek estimator (bottom panel), they are almost always above 1 for the IPW estimator (top panel). This is consistent with the above results, showing that we tend to overestimate (underestimate) the uncertainty for the Hájek (IPW) estimator. We find that the confidence interval for the Hájek estimator tends to be most conservative when $M$ is small and the region of interest is large. For the IPW estimator, the degree of underestimation decreases as the length of time series $T$ increases but increases as the length of intervention $M$ increases. In fact, when $M=30$, some of the ratios are as large as 20 (hence they are not included in the figure). The results suggest that in practice the Hájek estimator should be preferred over the IPW estimator especially for stochastic interventions over a long time period.
Covariate Balance {#app_subsec:add_sims_balance}
-----------------
We evaluate the balance of covariates based on the estimated propensity score by comparing their p-values in the propensity score model, and in a model with functional form as in the propensity score model but weighted by the inverse of the estimated propensity score. The left plot of \[app\_fig:sims\_balance\] shows the p-value for the previous outcome-active locations, which are one of the time-varying confounders, across 200 simulated data sets. Evidently, the p-values in the unweighted model are close to 0, indicating that previous outcome-active locations form an important predictor of the treatment assignment. However, in the weighted model, the p-values of the same confounder are more evenly distributed across the $(0, 1)$ range, indicating that this confounder is better balanced in the weighted time series. The right plot of \[app\_fig:sims\_balance\] shows the distribution (over 200 simulated data sets) of the estimated coefficient for this confounder in the weighted model. This distribution is centered around zero, indicating that balance is achieved on average across data sets.
Additional Empirical Results {#app_sec:study}
============================
Visualization
-------------
![Visualization of Intensity under Stochastic Interventions whose Focal Point is the Center of Baghdad. Across plots, we vary the degree to which the airstrikes are concentrated around the focal point using the precision parameter, while the expected number of airstrikes is held constant at 3 per day.[]{data-label="fig:type2_interv"}](fig_app_Type2_interv.pdf){width="95.00000%"}
As discussed in Section \[subsec:study\_interventions\], we consider a stochastic intervention whose focal point is the center of Baghdad. The degree of concentration is controlled by the precision parameter $\alpha$ whose greater value, implying that more airstrikes are occurring near the focal point. We vary the value of $\alpha$ from 0 to 3, while keeping the expected number of airstrikes constant at 3 per day. \[fig:type2\_interv\] illustrates intensities for the different values of $\alpha$. The first plot in the figure does not focus on Baghdad at all, representing the baseline spatial distribution $\phi_0$. As the value of $\alpha$ increases, the spatial distribution of airstrikes becomes concentrated more towards the center of Baghdad.
Empirical Results {#app_tab:res_tau}
-----------------
Type ($\interv[T]['], \interv[T]['']$) ${M}$ Outcome Iraq Baghdad Outside Baghdad
---------------------------------------- ------- -------------- -------------------------- --------------------------- -------------------------
3 IED 1.5 (-2.5, 5.5) 0.1 (-1, 1.3) 1.4 (-1.5, 4.3)
SAF 3.6 (-0.4, 7.7) 1.4 (-0.1, 2.8) 2.3 (-0.4, 4.9)
Other Attack 5.1 (-3.9, 14) 1.8 (-1, 4.7) 3.2 (-3.1, 9.5)
Increasing the 7 IED 4.9 (-0.1, 10) 1 (-0.4, 2.4) [**4 (0.2, 7.7)**]{}
intensity SAF [**7 (1.9, 12.1)**]{} [**2 (0.4, 3.5)**]{} [**5 (1.4, 8.6)**]{}
(1, 3) Other Attack [**12.3 (0.4, 24.1)**]{} 3.5 (0, 7.1) [**8.7 (0.3, 17.1)**]{}
30 IED -2.5 (-5.9, 1) [**-1 (-1.8, -0.2)**]{} -1.4 (-4.1, 1.2)
SAF -0.6 (-4.7, 3.5) [**-1.3 (-2.5, -0.2)**]{} 0.8 (-2.3, 3.9)
Other Attack -6.6 (-14.8, 1.7) [**-4 (-6.3, -1.6)**]{} -2.6 (-8.7, 3.5)
3 IED -1 (-9.5, 7.5) -0.4 (-2.6, 1.7) -0.6 (-7, 5.9)
SAF -1.2 (-9.9, 7.5) 0.3 (-2.6, 3.3) -1.5 (-7.5, 4.4)
Other Attack -4.9 (-23.9, 14) -2 (-8, 3.9) -2.9 (-16.1, 10.3)
Changing the 7 IED -3.6 (-14.3, 7) -1.9 (-4.6, 0.8) -1.7 (-9.9, 6.4)
focal points SAF -5.6 (-16.6, 5.4) -3.1 (-6.6, 0.5) -2.5 (-10.2, 5.2)
(0, 3) Other Attack -18.4 (-44, 7.3) -7.1 (-14.9, 0.7) -11.3 (-29.4, 6.9)
30 IED -3.2 (-18.3, 12) -1.5 (-5.3, 2.3) -1.7 (-13.2, 9.8)
SAF 2.4 (-15.1, 19.9) -3.9 (-9.5, 1.6) 6.4 (-6.1, 18.8)
Other Attack -7.8 (-44.3, 28.7) -8.1 (-19.6, 3.4) 0.3 (-25.3, 25.9)
3 IED -2.3 (-10.8, 6.2) -0.6 (-2.9, 1.6) -1.7 (-8, 4.7)
SAF -2.8 (-11.7, 6.2) -1.8 (-4.7, 1.2) -1 (-7.2, 5.1)
Other Attack -7.9 (-28.7, 13) -3.6 (-10.3, 3) -4.2 (-18.6, 10.2)
Lagged 7 IED 6.7 (-0.6, 14) 1.7 (-0.3, 3.6) 5 (-0.4, 10.5)
effects SAF [**7 (0.7, 13.4)**]{} [**2.9 (0.6, 5.2)**]{} 4.1 (-0.1, 8.4)
(1, 5) Other Attack 16.8 (0, 33.5) [**6.8 (0.9, 12.8)**]{} 10 (-1.1, 21)
30 IED 1.8 (-3.9, 7.6) 0.4 (-1, 1.9) 1.4 (-2.9, 5.7)
SAF 1.9 (-3.1, 7) 0.7 (-1.1, 2.5) 1.2 (-2, 4.5)
Other Attack 5 (-9.1, 19.1) 2.5 (-3.2, 8.1) 2.6 (-5.9, 11)
: Causal Effect Estimates and 95% Confidence Intervals for Various Stochastic Interventions. We present the results for three interventions discussed in the main text: increasing the expected number of airstrikes from 1 to 3 per day for $M$ days, changing the focal points of airstrikes from $\alpha = 0$ to $\alpha = 3$ for $M$ days, and the lagged effects of increasing the expected number of airstrikes from 1 to 5 per day $M$ days ago. The range of $M$ we consider is $\{3, 7, 30\}$. The regions of interest are Iraq, Baghdad, and the area outside Baghdad. The results in bold represent statistically significant estimates.[]{data-label="tab:effectestimates"}
\[tab:effectestimates\] presents the numerical effect estimates and 95% confidence intervals for various interventions, including those shown in the main text. We also show the effect estimates for the whole Iraq, Baghdad only, and the area outside Baghdad. Statistically significant estimates include the estimated effects of increasing the expected number of airstrikes from 1 to 3 per day for a period of a week on all types of insurgent attacks, with the increase located mostly outside Baghdad. In addition, continuing this intervention for a period of thirty days leads to a decrease in the number of insurgent attacks in Baghdad. For ${M}= 3$, all estimated lagged effects are negative, whereas all estimated lagged effects for $M= 7$ are positive, indicating that increasing the number of airstrikes might have a short term effect in reducing insurgent violence, but leading to a long-term increase.
Local Interventions {#subsec:local_intervention}
-------------------
In this section, we present the results of [*local interventions*]{}, which change the airstrike strategy over a region of the country, $\intervset \subset \Omega$. Consider intensities $\trtintensity_1, \trtintensity_2$ which are equal to each other outside this region, i.e., $\trtintensity_1(\omega) = \trtintensity_2(\omega) = c_{\notintervset} \phi_0(\omega),$ for $\omega \in \notintervset = \Omega \backslash \intervset$. Then, comparing average potential outcomes under this strategies would represent the change in the expected number of insurgent attacks if the airstrike strategy over the area $\notintervset$ were identical, and the airstrike locations within $\intervset$ were determined according to $\trtintensity_1$ versus $\trtintensity_2$.
In our analysis, for the sake of illustration, we set $\intervset$ to the administrative region of Baghdad and consider the airstrike strategy that sets the expected number of airstrikes outside of Baghdad to 3 per day, and that within the Baghdad administrative unit to $\{1, 2, 3, 4\}$. This strategy is shown in \[fig:local\_interv\].
![Visualization of the Local Intervention Strategy. Under this strategy, we set the expected number of airstrikes outside the Baghdad administrative unit (polygon in the map) to 3 with constant spatial distribution, while changing the intensity within Baghdad from 1 per day to 4 per day.[]{data-label="fig:local_interv"}](fig_app_Local_interv.pdf){width="\textwidth"}
\[app\_tab:res\_tau\_local\] shows the estimated causal effects for this local intervention. We find that the intervention does not lead to any statistically significant change in insurgent attacks of any type.
${M}$ Outcome Iraq Baghdad Outside Baghdad
------- -------------- -------------------- ------------------ --------------------
3 IED 3.9 (-6.5, 14.4) 1.8 (-1.1, 4.6) 2.1 (-5.6, 9.9)
SAF -0.1 (-10.8, 10.5) 0.3 (-3.3, 3.9) -0.4 (-7.7, 6.8)
Other Attack 8.1 (-17.3, 33.4) 3.7 (-4.8, 12.3) 4.3 (-12.9, 21.5)
7 IED 3.8 (-7.3, 14.9) 1 (-1.9, 3.8) 2.8 (-5.6, 11.3)
SAF 1.1 (-9.9, 12.1) 0.8 (-2.9, 4.6) 0.3 (-7.1, 7.7)
Other Attack 8.9 (-18.1, 35.9) 3.7 (-5.1, 12.6) 5.2 (-13.3, 23.7)
30 IED -3.3 (-16.6, 10) -1.7 (-5.2, 1.7) -1.6 (-11.5, 8.4)
SAF 0.1 (-14.4, 14.6) -4 (-8.5, 0.5) 4.1 (-6.3, 14.5)
Other Attack -10.8 (-43, 21.5) -7.6 (-18, 2.9) -3.2 (-25.5, 19.1)
: Causal Effect Estimates for Local Intervention Strategy. The duration of the strategy is set to $M \in \{3, 7, 30\}$. The strategy increases the expected number of airstrikes within the Baghdad administrative unit from 1 to 4 per day while keeping the strategy for the rest of the country unchanged at 3 per day. The 95% confidence intervals are in parentheses.[]{data-label="app_tab:res_tau_local"}
[^1]: Postdoctoral Associate, Department of Statistical Science, Duke University, Durham NC 27708. Email: <[email protected]>, URL: <https://gpapadogeorgou.netlify.com>
[^2]: Professor, Department of Government and Department of Statistics, Harvard University. 1737 Cambridge Street, Institute for Quantitative Social Science, Cambridge MA, 02138. Email: <[email protected]>, URL: <https://imai.fas.harvard.edu>
[^3]: James Wright Chair in Transnational Studies and Associate Professor, Department of Government, Dartmouth College, Hanover, NH 03755. Email: <[email protected]>, URL: [www.jasonlyall.com](www.jasonlyall.com)
[^4]: Associate Professor, Department of Statistical Science, Duke University, Durham, NC 27708. Email: <[email protected]>, URL: <http://www2.stat.duke.edu/~fl35>
[^5]: Lyall gratefully acknowledges financial support from the Air Force Office of Scientific Research (Grant $\#$FA9550-14-1-0072). The findings and conclusions reached here do not reflect the official views or policy of the United States Government or Air Force.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'In this paper, by using the trace map of Frobenius, we consider problems on extending sections for positive characteristic threefolds.'
address: 'Department of Mathematics, Faculty of Science, Kyoto University, Kyoto 606-8502 Japan'
author:
- Hiromu Tanaka
title: The trace map of Frobenius and extending sections for threefolds
---
Introduction
============
In characteristic zero, by the Kodaira vanishing theorem and its generalizations, we can establish some results on adjoint divisors, such as the Kawamata–Shokurov basepoint free theorem (see, for example, [@KM Theorem 3.3]) and the Hacon–M^c^Kernan extension theorem ([@HM Theorem 5.4.21]). These theorems claim, under suitable conditions, that an adjoint divisor $m(K_X+\Delta+A)$ has good properties where $m\in\mathbb Z_{>0}$, $(X, \Delta)$ is a pair and $A$ is an ample divisor. In this paper, we only consider the following very simple situation: $X$ is a smooth projective variety, $\Delta=S$ is a smooth prime divisor and $A$ is an ample Cartier divisor. The following fact immediately follows from the Kodaira vanishing theorem.
\[0char-zero\] Let $k$ be an algebraically closed field of characteristic zero. Let $X$ be a smooth projective variety over $k$. Let $S$ be a smooth prime divisor on $X$ and let $A$ be an ample Cartier divisor on $X$ such that $K_X+S+A$ is nef. Fix $m\in\mathbb Z_{>0}.$ Then, by the Kodaira vanishing theorem, we obtain $$H^1(X, K_X+A+(m-1)(K_X+S+A))=0.$$ Thus, the natural restriction map $$H^0(X, m(K_X+S+A))\to H^0(S, m(K_S+A))$$ is surjective.
It is natural to consider whether the above fact also holds in positive characteristic. Unfortunately, however, there exists the following example.
Let $k$ be an algebraically closed field of positive characteristic. Then, there exist a smooth projective surface $X$ over $k$, a smooth prime divisor $C$ on $X$ and an ample Cartier divisor $A$ on $X$ such that $K_X+C+A$ is nef and that the natural restriction map $$H^0(X, K_X+C+A)\to H^0(C, K_C+A)$$ is not surjective.
Thus, we would like to find a suitable analogy of Fact \[0char-zero\] in positive characteristic. In this paper, we prove the following two theorems.
\[0surface-special-extension\] Let $k$ be an algebraically closed field of positive characteristic. Let $X$ be a smooth projective surface over $k$. Let $C$ be a smooth prime divisor on $X$ and let $A$ be an ample Cartier divisor on $X$. If $H^0(C, K_C+A)\neq 0$, then the natural restriction map $$H^0(X, K_X+C+A)\to H^0(C, K_C+A)$$ is a non-zero map.
\[0threefold-extension\] Let $k$ be an algebraically closed field of positive characteristic. Let $X$ be a smooth projective threefold over $k$. Let $S$ be a smooth prime divisor on $X$ and let $A$ be an ample Cartier divisor on $X$. Assume the following two conditions.
1. [$K_X+S+A$ is nef.]{}
2. [$\kappa(S, K_S+A)\neq 0$.]{}
Then, there exists $m_0\in \mathbb Z_{>0}$ such that, for every $m\geq m_0$, the natural restriction map $$H^0(X, m(K_X+S+A))\to H^0(S, m(K_S+A))$$ is surjective.
To show the above two theorems, we use the trace map of Frobenius. This strategy is essentially the same as [@Schwede2 Proposition 5.3] and its proof. Let us see the idea of the proofs. Let $X$ be a smooth projective variety. Let $S$ be a smooth prime divisor on $X$ and let $A$ be an ample Cartier divisor on $X$. Then, for every $e\in\mathbb Z_{>0}$, we obtain the following commutative diagram by using the trace map of Frobenius: [ $$\begin{CD}
H^0(X, K_X+S+p^eA)@>>>H^0(S, K_S+p^eA)@>>>H^1(X, K_X+p^eA)\\
@VVV @VV{{\operatorname{Tr}}}_S^e(A)V\\
H^0(X, K_X+S+A)@>\rho >> H^0(S, K_S+A)
\end{CD}$$]{} where the lower horizontal arrow $\rho$ is the natural restriction map and the upper horizontal sequence is exact. By the Serre vanishing theorem, for large $e\gg 0$, we obtain the vanishing $H^1(X, K_X+p^eA)=0$. Thus, to prove that the restriction map $\rho$ is surjective (resp. a non-zero map), it is sufficient to show that the trace map ${{\operatorname{Tr}}}_S^e(A)$ is surjective (resp. a non-zero map). Therefore, to prove the above two theorems, we establish the following results on the trace map of Frobenius.
\[0trace-curve\] Let $k$ be an algebraically closed field of positive characteristic. Let $C$ be a smooth projective curve over $k$. Let $A$ be an ample Cartier divisor on $C$. If $H^0(C, \omega_C(A))\neq 0$, then the trace map $${{\operatorname{Tr}}}_C^e(A):H^0(C, \omega_C(p^eA))\to H^0(C, \omega_C(A))$$ is a non-zero map for every $e\in\mathbb Z_{>0}$.
\[0trace-surface\] Let $k$ be an algebraically closed field of positive characteristic. Let $S$ be a smooth projective surface over $k$. Let $A$ be an ample Cartier divisor on $S$. Assume the following two conditions.
1. [$K_S+A$ is nef.]{}
2. [$\kappa(S, K_S+A)\neq 0$.]{}
Then, there exists $m_1\in\mathbb Z_{>0}$ such that the trace map ${{\operatorname{Tr}}}_S^e(A+m(K_S+A))$ $$H^0(S, \omega_S(p^e(A+m(K_S+A))))\to H^0(S, \omega_S(A+m(K_S+A)))$$ is surjective for every $m\geq m_1$ and for every $e\in\mathbb Z_{>0}$.
We would like to consider whether Theorem \[0threefold-extension\] and Theorem \[0trace-surface\] hold for the case where $\kappa(S, K_S+A)=0$. Let us compare Theorem \[0threefold-extension\] with the following basepoint free conjecture (cf. [@KM Theorem 3.3]).
\[0bpf\] Let $k$ be an algebraically closed field of positive characteristic. Let $X$ be a smooth projective threefold over $k$. Let $S$ be a smooth prime divisor on $X$ and let $A$ be an ample Cartier divisor on $X$ such that $K_X+S+A$ is nef. Then, $|b(K_X+S+A)|$ is basepoint free for every $b\gg 0$.
If Conjecture \[0bpf\] holds, then Theorem \[0threefold-extension\] also hold for the case where $\kappa(S, K_S+A)=0$. Then, does Theorem \[0trace-surface\] also hold for the case where $\kappa(S, K_S+A)=0$? Unfortunately, the answer is NO. We can construct the following example in characteristic two.
\[0trace-surface-cex\] Let $k$ be an algebraically closed field of characteristic two. Then, there exists a smooth projective surface $S$ over $k$ such that
1. [$-K_S=:A$ is ample. Note that $\kappa(S, K_S+A)=0$. ]{}
2. [For every $e\in\mathbb Z_{>0}$, the trace map $${{\operatorname{Tr}}}_S^e(A):H^0(S, \omega_S(2^e A))\to H^0(S, \omega_S(A))$$ is a zero map. ]{}
Moreover, Theorem \[0trace-surface-cex\] also shows that we can not generalize Theorem \[0trace-curve\] to dimension two.
In the appendix of this paper (Section 9), we establish the following analogy of the Hacon–M^c^Kernan extension theorem for surfaces.
\[0surface-ext\] Let $k$ be an algebraically closed field of positive characteristic. Let $X$ be a smooth projective surface over $k$ and let $C$ be a smooth prime divisor on $X$. Let $\Delta:=C+B$ where $B$ is an effective [$\mathbb{Q}$]{}-divisor on $X$ which satisfies the following properties.
1. [$C\not\subset{{\operatorname{Supp}}}B$, $\llcorner B\lrcorner=0$ and $(X, \Delta)$ is plt. ]{}
2. [$B\sim_{\mathbb Q}A+F$ where $A$ is an ample [$\mathbb{Q}$]{}-divisor and $F$ is an effective [$\mathbb{Q}$]{}-divisor such that $C\not\subset {{\operatorname{Supp}}}F.$]{}
3. [No prime component of $\Delta$ is contained in the stable base locus of $K_X+\Delta.$]{}
Then, there exists an integer $m_0>0$ such that, for every integer $m>0$, the restriction map $$H^0(X, mm_0(K_X+\Delta))\to H^0(C, mm_0(K_X+\Delta)|_C)$$ is surjective.
But, the proof for Theorem \[0surface-ext\] does not use the trace map of Frobenius. We use results on the minimal model theory established in [@T2] and [@T3].
In Section 1, we summarize the notations. In Section 2, we see the definition and some basic properties of the trace map of Frobenius. The trace map of Frobenius is obtained by applying the functor $\mathcal Hom_{\mathcal O_X}(-, \omega_X)$ to the Frobenius map $\mathcal O_X\to F_*\mathcal O_X.$ In Section 3, we see some known facts on Cartier operator. We can consider the trace map of Frobenius as Cartier operator. The Cartier operator is defined by the de Rham complex. We use Cartier operator to consider the relation between the trace map of Frobenius and étale base changes. In Section 4, we show Theorem \[0surface-special-extension\] and Theorem \[0trace-curve\]. In Section 5, we prove Theorem \[0trace-surface\] for the case where $\kappa=1$. In Section 6, we show Theorem \[0trace-surface\] for the case where $\kappa=2$. In Section 7, by using Theorem \[0trace-surface\], we show Theorem \[0threefold-extension\]. In Section 8, we prove Theorem \[0trace-surface-cex\]. In Section 9, we show Theorem \[0surface-ext\].
We summarize literature related to this paper with respect to the basepoint free theorem, the extension theorem, the trace map of Frobenius and the minimal model theory in positive characteristic.
(Basepoint free theorem and extension theorem) The motivation of this paper is the basepoint free theorem and the extension theorem in characteristic zero. Thus, let us summarize some known results on them. Kawamata and Shokurov established the basepoint fee theorem for klt pairs (cf. [@KMM], [@KM]). [@Ambro] generalizes this result (cf. [@Fujino]). The extension theorem is established by Hacon and M^c^Kernan ([@HM Theorem 5.4.21]). This theorem is a key to show the existence of the flip ([@BCHM]). For related topics, see [@DHP] and [@FG].
(The trace map of Frobenius) The heart of this paper is the trace map of Frobenius. The trace map of Frobenius plays a crucial role in the theory of $F$-singularities (cf. [@BST], [@Schwede1], [@Schwede2]). Moreover, [@CHMS] and [@Mustata] establish results related to birational geometry by using the trace map of Frobenius and the theory of $F$-singularities. For related topics, see [@BSTZ] and [@TW].
(Minimal model theory in positive characteristic) Let us summarize literature on the minimal model theory in positive characteristic. For results on surfaces, see [@Fujita3], [@KK], [@T2] and [@T3]. For results on threefolds, see [@Kawamata], [@Keel], [@Kollar] and [@HX].
The author would like to thank Professor Osamu Fujino for many comments and discussions. He would like to thank Professor Karl Schwede for numerous helpful comments. He thanks Professor Atsushi Moriwaki for warm encouragement. The author is partially supported by JSPS Fellowships for Young Scientists.
Notations
=========
We will freely use the notation and terminology in [@KM].
We will not distinguish the notations invertible sheaves and divisors. For example, we will write $L+M$ for invertible sheaves $L$ and $M$.
For a coherent sheaf $F$ and a Cartier divisor $L$, we define $F(L):=F\otimes \mathcal O_X(L)$.
Throughout this paper, we work over an algebraically closed field $k$ of positive characteristic and let ${\rm char}\,k=:p>0.$
In this paper, [*a variety*]{} means an integral scheme which is separated and of finite type over $k$.
The trace map of Frobenius
==========================
In this section, we define the trace map of the Frobenius and we see some fundamental properties. We only use the smooth case. For the singular case, see [@Schwede2 Section 2].
\[tracedef\] Let $X$ be a smooth projective variety. Let $E$ be an effective [$\mathbb{Z}$]{}-divisor and let $D$ be a [$\mathbb{Z}$]{}-divisor. Then, for every positive integer $e$, there exists a natural $\mathcal O_X$-module homomorphism $${{\operatorname{Tr}}}_{X,E}^e(D):F^e_*(\omega_X(E+p^eD))\to
\omega_X(E+D).$$ We call this a trace map.
Consider the Frobenius map: $$\mathcal O_X\to F^e_*\mathcal O_X,$$ that is, $p^e$-th power map $a\mapsto a^{p^e}$. Since $E$ is effective, we obtain $$\mathcal O_X(-E)\to F^e_*(\mathcal O_X(-p^eE))\hookrightarrow F^e_*(\mathcal O_X(-E)).$$ Tensoring $\mathcal O_X(-D)$, we obtain $$\begin{aligned}
\mathcal O_X(-E-D)
&\to& F^e_*(\mathcal O_X(-E))\otimes \mathcal O_X(-D)\\
&\simeq& F^e_*(\mathcal O_X(-E-p^eD)).\end{aligned}$$ By the duality theorem for finite morphisms, we obtain $$\mathcal Hom_{\mathcal O_X}(F^e_*(\mathcal O_X(-E-p^eD)), \omega_X)\simeq F^e_*(\omega_X(E+p^eD)).$$ Then, apply the functor $\mathcal Hom_{\mathcal O_X}(-, \omega_X)$ and we obtain $$F^e_*(\omega_X(E+p^eD))\to \omega_X(E+D).$$ This is the trace map ${{\operatorname{Tr}}}_{X,E}^e(D)$.
\[pE-vs-E\] By the above construction, ${{\operatorname{Tr}}}_{X,E}^e(D)$ factors through ${{\operatorname{Tr}}}_X^e(E+D):={{\operatorname{Tr}}}_{X, 0}^e(E+D)$: [$${{\operatorname{Tr}}}_{X,E}^e(D):F^e_*(\omega_X(E+p^eD))\hookrightarrow F^e_*(\omega_X(p^eE+p^eD))
\xrightarrow{{{\operatorname{Tr}}}_X^e(E+D)} \omega_X(E+D).$$]{}
\[trace-local\] Let $X$ be a smooth projective variety. Let ${{\operatorname{Spec}}}R\subset X$ be an affine open subset such that $R$ has a $p$-basis $\{x_1, \cdots, x_n\}.$ Then, we obtain $$\Gamma({{\operatorname{Spec}}}R, \omega_X)=Rdx_1\wedge\cdots\wedge dx_n$$ and $$\begin{aligned}
\Gamma({{\operatorname{Spec}}}R, F_*^e\omega_X)
&=&\bigoplus_{0\leq i_j<p^e}R\cdot x_1^{i_1}\cdots x_n^{i_n}dx_1\wedge\cdots\wedge dx_n\\
&=&\bigoplus_{0\leq i_j<p^e}R^{p^e}x_1^{i_1}\cdots x_n^{i_n}dx_1\wedge\cdots\wedge dx_n.\end{aligned}$$ The trace map $${{\operatorname{Tr}}}_X^e:\Gamma({{\operatorname{Spec}}}R, F_*^e\omega_X)\to \Gamma({{\operatorname{Spec}}}R, \omega_X)$$ is described as follows:
1. [${{\operatorname{Tr}}}_X^e(x_1^{p^e-1}\cdots x_n^{p^e-1}dx_1\wedge\cdots\wedge dx_n)=dx_1\wedge\cdots\wedge dx_n.$]{}
2. [${{\operatorname{Tr}}}_X^e(x_1^{i_1}\cdots x_n^{i_n}dx_1\wedge\cdots\wedge dx_n)=0$ if $0\leq i_j<p^e-1$ for some $1\leq j\leq n$.]{}
The following two lemmas are the fundamental properties.
\[D-linearequ\] Let $X$ be a smooth projective variety and let $E$ be an effective [$\mathbb{Z}$]{}-divisor. If $D_1$ and $D_2$ are linearly equivalent [$\mathbb{Z}$]{}-divisors, then the two trace maps ${{\operatorname{Tr}}}_X^e(D_1)$ and ${{\operatorname{Tr}}}_X^e(D_2)$ are the same for every positive integer $e$, that is, there exists a following commutative diagram: $$\begin{CD}
F^e_*(\omega_X(E+p^eD_2))@>{{\operatorname{Tr}}}_X^e(D_2)>> \omega_X(E+D_2)\\
@VV\simeq V @VV\simeq V\\
F^e_*(\omega_X(E+p^eD_1))@>{{\operatorname{Tr}}}_X^e(D_1)>> \omega_X(E+D_1).
\end{CD}$$
Consider the Frobenius map: $$\mathcal O_X(-E)\to F^e_*(\mathcal O_X(-E)).$$ Tensoring $\mathcal O_X(-D_i)$, we obtain $$\begin{aligned}
\mathcal O_X(-E-D_i)
\to F^e_*(\mathcal O_X(-E-p^eD_i)).\end{aligned}$$ Then, we obtain the following commutative diagram: $$\begin{CD}
F^e_*(\mathcal O_X(-E-p^eD_2))@<<< \mathcal O_X(-E-D_2)\\
@AA\times g^{p^e} A @AA\times g A\\
F^e_*(\mathcal O_X(-E-p^eD_1))@<<< \mathcal O_X(-E-D_1)
\end{CD}$$ where $g\in K(X)^*$ is defined by $\mathcal O_X(-D_1)\overset{\times g}\simeq\mathcal O_X(-D_2).$ In particular, the vertical arrows are isomorphisms. Apply the functor $\mathcal Hom_{\mathcal O_X}(-, \omega_X)$ and we obtain the required commutative diagram.
\[e-vs-e+1\] Let $X$ be a smooth projective variety and let $E$ be an effective [$\mathbb{Z}$]{}-divisor. Let $D$ be a [$\mathbb{Z}$]{}-divisor. Then, for every positive integer $e$, $${{\operatorname{Tr}}}_{X, E}^{e+1}(D)={{\operatorname{Tr}}}_{X, E}^e(D)\circ F_*^e({{\operatorname{Tr}}}_{X, E}^1(p^e D)),$$ that is, $$\begin{aligned}
{{\operatorname{Tr}}}_{X, E}^{e+1}(D):
F^{e+1}_*(\omega_X(E+p^{e+1}D))&\xrightarrow{F_*^e({{\operatorname{Tr}}}_{X,E}^1(p^eD))}&F^e_*(\omega_X(E+p^eD))\\
&\xrightarrow{{{\operatorname{Tr}}}_{X,E}^{e}(D)}&\omega_X(E+D).\end{aligned}$$
Consider the Frobenius maps: $$\mathcal O_X(-E)\to F^e_*(\mathcal O_X(-E))\to F^{e+1}_*(\mathcal O_X(-E)).$$ Tensoring $\mathcal O_X(-D)$, we obtain $$\begin{aligned}
\mathcal O_X(-E-D)
\to F^e_*(\mathcal O_X(-E-p^eD))\to
F^{e+1}_*(\mathcal O_X(-E-p^{e+1}D)).\end{aligned}$$ Apply the functor $\mathcal Hom_X(-, \omega_X)$ and we obtain the assertion.
In this paper, we often use the following two commutative diagrams.
\[trace-diagram\] Let $X$ be a smooth projective variety and let $S$ be a smooth prime divisor. Then, there exist the following commutative diagrams.
1. [$$\begin{CD}
0@>>> F_*^e\omega_X@>>> F_*^e(\omega_X(S))@>>> F_*^e\omega_S@>>> 0\\
@. @VV{{\operatorname{Tr}}}_X^e V @VV{{\operatorname{Tr}}}_{X,S}^e V @VV{{\operatorname{Tr}}}_S^e V\\
0@>>> \omega_X@>>> \omega_X(S)@>>> \omega_S@>>> 0.
\end{CD}$$]{}
2. [$$\begin{CD}
0@>>> F_*^e\omega_X@>>> F_*^e(\omega_X(p^eS))@>>> F_*^e(\omega_{p^eS})@>>> 0\\
@. @VV{{\operatorname{Tr}}}_X^e V @VV{{\operatorname{Tr}}}_{X}^e(S) V @VV\Psi V\\
0@>>> \omega_X@>>> \omega_X(S)@>>> \omega_S@>>> 0.
\end{CD}$$ Moreover, ${{\operatorname{Tr}}}_S^e$ factors through $\Psi$ $:$ $${{\operatorname{Tr}}}_S^e:F_*^e\omega_S\to F_*^e(\omega_{p^eS})\overset{\Psi}\to \omega_S.$$ ]{}
\(1) Consider the following commutative diagram: $$\begin{CD}
0@>>> F_*^e(\mathcal O_X(-S))@>>> F_*^e\mathcal O_X@>>> F_*^e\mathcal O_S@>>> 0\\
@. @AAA @AA F_X^e A @AAF_S^e A\\
0@>>> \mathcal O_X(-S)@>>> \mathcal O_X@>>> \mathcal O_S@>>> 0.
\end{CD}$$ Apply the functor $\mathcal Hom_X(-, \omega_X)$ and we obtain the assertion.\
(2) Consider the following commutative diagram: $$\begin{CD}
0@>>> F_*^e(\mathcal O_X(-p^eS))@>>> F_*^e\mathcal O_X@>>> F_*^e\mathcal O_{p^eS}@>>> 0\\
@. @AAA @AA F_X^e A @AAA\\
0@>>> \mathcal O_X(-S)@>>> \mathcal O_X@>>> \mathcal O_S@>>> 0.
\end{CD}$$ Apply the functor $\mathcal Hom_X(-, \omega_X)$ and we obtain the required commutative diagram. Since $$F_S^e:\mathcal O_S\to F_*^e\mathcal O_S$$ factors through $F_*^e\mathcal O_{p^eS}$, we see the assertion.
We would like to know whether the trace map ${{\operatorname{Tr}}}_X^e(D)$ has a good property. It is natural to consider the following question.
\[Tango-cex\] Let $X$ be a smooth projective variety and let $A$ be an ample [$\mathbb{Z}$]{}-divisor on $X$. Then, for every $e\in\mathbb Z_{>0}$, is the trace map $${{\operatorname{Tr}}}_X^e(A):H^0(X, \omega_X(p^eA))\to H^0(X, \omega_X(A))$$ surjective?
The answer to the above question is NO. Indeed, [@Tango] constructs a smooth projective curve $X$ and an ample [$\mathbb{Z}$]{}-divisor $A$ on $X$ such that the trace map ${{\operatorname{Tr}}}_X^e(A)$ is not surjective.
On the other hand, we obtain the affirmative answer for the following two cases: ablelian varieties and $F$-split varieties.
Let $X$ be an abelian variety and let $A$ be an ample [$\mathbb{Z}$]{}-divisor. Then, the trace map $${{\operatorname{Tr}}}_{X}^e(A):H^0(X, \omega_X(p^eA))\to H^0(X, \omega_X(A)).$$ is surjective.
For $n\in\mathbb Z$, let $n_X$ be the $n$-multiplication map of the abelian variety $X$. Let $n\in\mathbb Z_{>0}$ which is not divisible by $p$. Then, $$n_X:X\to X$$ is a finite morphism whose degree is not divisible by $p$. Thus, $$\mathcal O_X\to (n_X)_*\mathcal O_X$$ is split as an $\mathcal O_X$-module homomorphism (cf. [@KM Proposition 5.7(2)]). We obtain the following commutative diagram $$\begin{CD}
F_*^e(n_X)_*(\mathcal O_X(p^e(n_X)^*A))@<<< (n_X)_*(\mathcal O_X((n_X)^*A))\\
@AAA @AA n_X A\\
F_*^e(\mathcal O_X(p^eA))@<<<\mathcal O_X(A)
\end{CD}$$ Apply the functor $\mathcal Hom_{\mathcal O_X}(-, \omega_X)$ and take the cohomology: $$\begin{CD}
H^0(X, \omega_X(p^e(n_X)^*A))@>{{\operatorname{Tr}}}_{X}^e((n_X)^*A))>> H^0(X, \omega_X((n_X)^*A))\\
@VVV @VV n_X^* V\\
H^0(X, \omega_X(p^eA))@>{{\operatorname{Tr}}}_{X}^e(A)>> H^0(X, \omega_X(A)).
\end{CD}$$ Here, $n_X^*$ is surjective by the splitting of $n_X$. Therefore, it is sufficient to show that ${{\operatorname{Tr}}}_{X}^e((n_X)^*A)$ is surjective. By [@Mumford Corollary 3 in Section 6], we obtain $$n_X^*A= \frac{n^2+n}{2}A+\frac{n^2-n}{2}(-1)_X^*A.$$ Note that, since $(-1)_X$ is an automorphism, $(-1)_X^*A$ is ample. Therefore, by the Fujita vanishing theorem ([@Fujita1 Theorem (1)], [@Fujita2 Section 5]), we can find $n\in\mathbb Z_{>0}$ such that ${{\operatorname{Tr}}}_{X}^e((n_X)^*A)$ is surjective.
Let $X$ be a smooth projective variety. We say $X$ is [*F-split*]{} if the Frobenius map $$\mathcal O_X\to F_*\mathcal O_X$$ is split as an $\mathcal O_X$-module homomorphism.
\[F-split-surje\] Let $X$ be an $F$-split smooth projective variety and let $D$ be a [$\mathbb{Z}$]{}-divisor. Then, the trace map $${{\operatorname{Tr}}}_{X}^e(D):H^0(X, \omega_X(p^eD))\to H^0(X, \omega_X(D)).$$ is surjective.
By the definition of $F$-splitting, we see that the Frobenius map $$\mathcal O_X(-D)\to F_*^e(\mathcal O_X(-p^eD))$$ is split. Apply the functor $\mathcal Hom_{\mathcal O_X}(-, \omega_X)$ and we obtain the assertion.
Facts on Cartier operator
=========================
In this section, we summarize the facts on Cartier operator. By Remark \[cartier-remark\], we consider the trace map of the Frobenius as Cartier operator.
\[de-rham\] Let $X$ be a smooth variety. Consider the de Rham complex of $X$ $$\Omega_X^{\bullet}:\mathcal O_X\overset{d_0}\to\Omega^1_X\overset{d_1}\to\Omega^2_X
\overset{d_2}\to\cdots$$ where $\Omega^i_X:=\Omega^i_{X/k}.$ Apply $F_*$ and we obtain a complex $$F_*\Omega_X^{\bullet}:F_*\mathcal O_X\overset{F_*d_0}\to F_*\Omega^1_X\overset{F_*d_1}\to
F_*\Omega^2_X\overset{F_*d_2}\to\cdots.$$ Then, it is easy to see that $F_*d_i$ is an $\mathcal O_X$-module homomorphism. We define $$\begin{aligned}
B_X^i&:=&{{\operatorname{Image}}}(F_*d_{i-1}:F_*\Omega_X^{i-1}\to F_*\Omega_X^{i})\\
Z_X^i&:=&{{\operatorname{Ker}}}(F_*d_i:F_*\Omega_X^i\to F_*\Omega_X^{i+1}).\end{aligned}$$ Note that $B_X^i$ and $Z_X^i$ are coherent sheaves.
\[cartier-def\] Let $X$ be a smooth variety. For every $i\in \mathbb Z$ such that $1\leq i\leq \dim X$, consider the map $$\begin{aligned}
C_X^{-1}:\Omega_X^i&\to& Z_X^i/B_X^i\end{aligned}$$ locally defined by $$\begin{aligned}
C_X^{-1}|_{{{\operatorname{Spec}}}R}:\Gamma({{\operatorname{Spec}}}R, \Omega_X^i)&\to& \Gamma({{\operatorname{Spec}}}R, Z_X^i/B_X^i)\\
da_1\wedge\cdots \wedge da_i&\mapsto&
a_1^{p-1}\cdots a_i^{p-1}da_1\wedge\cdots\wedge da_i\end{aligned}$$ where ${{\operatorname{Spec}}}R$ is an open affine subset of $X$ and $a_1,\cdots, a_i\in R.$ This map $C_X^{-1}$ is a well-defined $\mathcal O_X$-module isomorphism. We call $C_X:=(C_X^{-1})^{-1}$ Cartier operator.
See, for example, [@EV Theorem 9.14].
\[cartier-remark\] Let $X$ be an $n$-dimensional smooth variety. We obtain the following exact sequences.
1. [$0\to \mathcal O_X\to F_*\mathcal O_X\to B_X^1\to 0$.]{}
2. [$0\to Z_X^i\to F_*\Omega_X^i\to B_X^{i+1}\to 0$ for $1\leq i\leq n$.]{}
3. [$0\to B_X^i\to Z_X^i\overset{C_X^i}\to \Omega_X^{i}\to 0$ for $1\leq i \leq n$.]{}
By (2) for $i=n$, we obtain $Z_X^n\simeq F_*\omega_X$. By (3) for $i=n$, we obtain $$0\to B_X^n\to F_*\omega_X\overset{C_X^n}\to \omega_X\to 0.$$
\[cartier-trace\] Let $X$ be an $n$-dimensional smooth projective variety. By Remark \[trace-local\], Cartier operator and the trace map of Frobenius are the same: $C_X^n={{\operatorname{Tr}}}^1_X.$
\[cartier-etale\] Let $\gamma:X\to Y$ be a finite étale morphism between smooth varieties. Then, $$\gamma^*B_Y^i\simeq B_{X}^i\,\,\,{\rm and}\,\,\,\gamma^*Z_Y^i\simeq Z_{X}^i.$$
We may assume $X={{\operatorname{Spec}}}B$ and $Y={{\operatorname{Spec}}}A$. Let $$\varphi: A\to B$$ be the natural homomorphism induced by $\gamma$. Let $$F_A:A\to A\,\,\,{\rm and}\,\,\,F_B:B\to B$$ be the $p$-th power maps respectively. Since $\varphi$ is flat, we see that the natural $B$-module homomorphism $$\begin{aligned}
\theta^i:((F_A)_*\Omega_A^i)\otimes_A B&\to& (F_B)_*(\Omega_A^i\otimes_A B)\\
(\sum_J a_Jdx_J)\otimes_A b&\mapsto& (\sum_J a_Jdx_J)\otimes_A b^p\end{aligned}$$ is an isomorphism where $a_J\in A$ and $dx_J:=dx_{j_1}\wedge\cdots\wedge dx_{j_i}$ for some $x_{j_l}\in A$. Since $\varphi$ is étale, the natural $B$-module homomorphism $$\begin{aligned}
\rho^i:\Omega_A^i\otimes_A B&\to&\Omega_B^i\\
(\sum_J a_Jdx_J)\otimes_A b&\mapsto& \sum \varphi(a_J)bd(\varphi(x_J))\end{aligned}$$ is an isomorphism. It is sufficient to check the commutativity of the following diagram: $$\begin{CD}
(F_B)_*\Omega_B^i@>d>> (F_B)_*\Omega_B^{i+1}\\
@AAF_*\rho^iA @AAF_*\rho^{i+1}A\\
(F_B)_*(\Omega_A^i\otimes_A B)@>d>> (F_B)_*(\Omega_A^{i+1}\otimes_A B)\\
@AA\theta^iA @AA\theta^{i+1}A\\
((F_A)_*\Omega_A^i)\otimes_A B@>d>> ((F_A)_*\Omega_A^{i+1})\otimes_A B.
\end{CD}$$ This follows from $$\begin{aligned}
d\circ F_*\rho^i\circ \theta^i((\sum_J a_Jdx_J)\otimes_A b)
&=&d\circ F_*\rho^i((\sum_J a_Jdx_J)\otimes_A b^p)\\
&=&d(\sum_J \varphi(a_J)b^pd(\varphi(x_J)))\\
&=&\sum_J b^pd(\varphi(a_J))\wedge d(\varphi(x_J))\end{aligned}$$ and $$\begin{aligned}
&&F_*\rho^{i+1}\circ \theta^{i+1}\circ d((\sum_J a_Jdx_J)\otimes_A b)\\
&=&F_*\rho^{i+1}\circ \theta^{i+1}((\sum_J da_J\wedge dx_J)\otimes_A b)\\
&=&F_*\rho^{i+1}((\sum_J da_J\wedge dx_J)\otimes_A b^p)\\
&=&\sum_J b^pd(\varphi(a_J))\wedge d(\varphi(x_J)).\end{aligned}$$
We state the following vanishing result of $F$-split varieties for a later use.
\[F-split-B-vanish\] Let $X$ be an $n$-dimensional $F$-split projective variety and let $A$ be an ample [$\mathbb{Z}$]{}-divisor. Then, $$H^1(X, B_X^n(A))=0.$$
Consider the exact sequence $$0\to B_X^n\to F_*\omega_X\overset{C_X^n}\to \omega_X\to 0.$$ Then, by Proposition \[F-split-surje\], the trace map $$C_X^n={{\operatorname{Tr}}}_{X}^1(A):H^0(X, \omega_X(pA))\to H^0(X, \omega_X(A))$$ is surjective. Therefore, we obtain the exact sequence $$0\to H^1(X, B_X^n(A))\to H^1(X, \omega_X(pA)).$$ Since $F$-split varieties satisfy the Kodaira vanishing theorem ([@MR Proposition 2]), we obtain the vanishing $H^1(X, B_X^n(A))=0$.
The trace map of Frobenius for curves
=====================================
In this section, we calculate the trace map $${{\operatorname{Tr}}}_{X}^e(A):H^0(X, \omega_X(p^eA))\to H^0(X, \omega_X(A))$$ for the case where $X$ is a curve. By Question \[Tango-cex\] and its answer, ${{\operatorname{Tr}}}_X^e(A)$ is not surjective in general. But, for almost all the case, ${{\operatorname{Tr}}}_X^e(A)$ is a non-zero map.
\[curve-trace-nonzero\] Let $X$ be a smooth projective curve whose genus $g(X)$ is not zero. Let $A$ be an ample [$\mathbb{Z}$]{}-divisor. Then, for every $e\in\mathbb Z_{>0}$, the trace map $${{\operatorname{Tr}}}_{X}^e(A):H^0(X, \omega_X(p^eA))\to H^0(X, \omega_X(A))$$ is a non-zero map.
Fix $e\in\mathbb Z_{>0}$. Since $A$ is ample, we see $\deg A\geq 1.$ We consider the two cases: $\deg A>1$ and $\deg A=1$.
In this step, we assume $\deg A>1$ and we prove the assertion. The following argument is the same as the proof of [@Schwede2 Theorem 3.3].
Fix a point $Q\in X$. By Lemma \[trace-diagram\], we obtain the following commutative diagram:
[$$\begin{CD}
H^0(X, K_X+p^eA)@>>> H^0(p^eQ, K_{p^eQ}+p^e(A-Q))@.\to H^1(X, K_X+p^e(A-Q))\\
@VV{{\operatorname{Tr}}}_{X}^e(A) V @VV \Psi V\\
H^0(X, K_X+A)@>\rho >> H^0(Q, K_Q+A-Q).
\end{CD}$$]{} By Serre duality, we obtain the vanishing $$H^1(X, K_X+p^e(A-Q))=0.$$ On the other hand, $\Psi$ is surjective because $\Psi$ satisfies $$\begin{aligned}
{{\operatorname{Tr}}}_Q^e:H^0(Q, K_Q+p^e(A-Q))&\to &H^0(p^eQ, K_{p^eQ}+p^e(A-Q))\\
&\overset{\Psi}\to& H^0(Q, K_Q+A-Q).\end{aligned}$$ Therefore, the composition map $\rho \circ {{\operatorname{Tr}}}_X^e(A)$ is surjective. Then, we see that ${{\operatorname{Tr}}}_{X}^e(A)$ is a non-zero map by $H^0(Q, K_Q+A-Q)\neq 0$.
\[good-1-form\] In this step, we prove that, if $\deg A=1$, then there exists a point $Q\in X$ such that the natural injective map $$H^0(X, \omega_X(p^eA-p^eQ))\to H^0(X, \omega_X(p^eA-(p^e-1)Q))$$ is not surjective.
Since $H^0(Q, L)\simeq k$ for every invertible sheaf $L$, we obtain the following exact sequence $$\begin{aligned}
0&\to& H^0(X, \omega_X(p^eA-p^eQ))\to H^0(X, \omega_X(p^eA-(p^e-1)Q))\to k\\
&\to& H^1(X, \omega_X(p^eA-p^eQ)).\end{aligned}$$ Therefore, it is sufficient to show $$h^1(X, \omega_X(p^eA-p^eQ))=h^0(X, -(p^eA-p^eQ))=0$$ for some point $Q\in X$. Note that the first equality follows from Serre duality. Assume the contrary, that is, assume that $p^eA\sim p^eQ$ for every point $Q\in X$. Since the genus $g(X)$ is not zero, there exists a non-zero $l$-torsion $D$ for a prime number $l\neq p$. Note that $D$ is not a $p^e$-torsion. Take the prime decomposition $$D=\sum m_iQ_i-\sum n_jR_j.$$ Since $\deg D=\sum m_i-\sum n_j=0$, we obtain the following contradiction $$\begin{aligned}
p^eD&=&\sum m_ip^eQ_i-\sum n_jp^eR_j\\
&\sim&\sum m_ip^eA-\sum n_jp^eA\\
&=&(\sum m_i-\sum n_j)p^eA\\
&=&0.\end{aligned}$$
In this step, we assume $\deg A=1$ and we prove the assertion.
We fix a point $Q\in X$ as in Step \[good-1-form\]. If $A\sim A'$, then the corresponding trace maps are the same by Lemma \[D-linearequ\]. Therefore, we may assume that $Q\not\in {{\operatorname{Supp}}}A$. By Step \[good-1-form\], there exists an element $$\eta\in H^0(X, \omega_X(p^eA-(p^e-1)Q))\setminus H^0(X, \omega_X(p^eA-p^eQ)).$$ Take the local ring $(R, \mathfrak m)$ corresponding to the point $Q$. Note that $F^e_*R$ is a free $R$-module. Let $\{x\}$ be the $p$-basis. Then, we obtain $$\omega_R=\bigoplus_{0\leq i<p^e}R^{p^e}x^idx.$$ Thus, we can write $$\eta|_{{{\operatorname{Spec}}}R}=\sum_{0\leq i<p^e}f_i^{p^e}x^idx.$$ $\eta\not\in H^0(X, \omega_X(p^eA-p^eQ))$ means $f_i\not\in \mathfrak m$ for some $0\leq i<p^e$. $\eta\in H^0(X, \omega_X(p^eA-(p^e-1)Q))$ means $f_i\in \mathfrak m$ for every $0\leq i<p^e-1$. Therefore, we obtain $f_{p^e-1}\not\in \mathfrak m.$ Then, we can find $c\in k^{\times}$ and $\mu\in \mathfrak m$ such that $$f_{p^e-1}=c+\mu.$$ By Remark \[trace-local\], we see ${{\operatorname{Tr}}}_{X}^e(A)(\eta)|_{{{\operatorname{Spec}}}R}\neq 0.$
\[curve-trace-cor\] Let $X$ be a smooth projective curve. Let $A$ be an ample [$\mathbb{Z}$]{}-divisor. If $H^0(X, \omega_X(A))\neq 0$, then, for every $e\in\mathbb Z_{>0}$, the trace map $${{\operatorname{Tr}}}_X^e(A):H^0(X, \omega_X(p^eA))\to H^0(X, \omega_X(A))$$ is a non-zero map.
If $g(X)\geq 1$ where $g(X)$ is the genus of $X$, then the assertion follows from Theorem \[curve-trace-nonzero\]. Thus, we may assume that $X\simeq \mathbb P^1$. Since $\mathbb P^1$ is $F$-split, the trace map is surjective.
In characteristic zero, we obtain the following result by the Kodaira vanishing theorem. In positive characteristic, we obtain the following result by the trace map of Frobenius (Corollary \[curve-trace-cor\]).
\[surface-restriction\] Let $X$ be a smooth projective surface and let $C$ be a smooth prime divisor. Let $A$ be an ample [$\mathbb{Z}$]{}-divisor on $X$. If $H^0(C, K_C+A)\neq0$, then the natural restriction map $$H^0(X, K_X+C+A)\to H^0(C, K_C+A)$$ is a non-zero map.
By Lemma \[trace-diagram\], we obtain the following commutative diagram [$$\begin{CD}
H^0(X, K_X+C+p^eA)@>>> H^0(C, K_{C}+p^eA)@>>> H^1(X, K_X+p^eA)\\
@VV{{\operatorname{Tr}}}_{X, C}^e(A) V @VV {{\operatorname{Tr}}}_{C}^e(A) V\\
H^0(X, K_X+C+A)@>>> H^0(C, K_C+A).
\end{CD}$$]{} Then, the assertion follows from Corollary \[curve-trace-cor\].
In characteristic zero, in the above situation, the restriction map is surjective by the Kodaira vanishing theorem. But, in positive characteristic, the restriction map is not surjective in general.
\[cex-surjective\] There exists a smooth projective surface $X$, a smooth prime divisor $H$ on $X$ and an ample [$\mathbb{Z}$]{}-divisor $A$ such that
1. [$|K_X+H+A|$ is basepoint free. ]{}
2. [The natural restriction map $$H^0(X, K_X+H+A)\to H^0(H, K_H+A)$$ is not surjective. ]{}
Let $X$ be a smooth projective surface and let $A$ be an ample [$\mathbb{Z}$]{}-divisor on $X$ such that $$H^1(X, K_X+A)\neq 0.$$ We can find such a surface by [@Raynaud]. Take a smooth hyperplane section $H$ of $X$ such that $|K_X+H+A|$ is basepoint free and that $$H^1(X, K_X+H+A)=0.$$ Consider the exact sequence $$0\to \mathcal O_X(K_X+A)\to\mathcal O_X(K_X+H+A)\to\mathcal O_H(K_H+A)\to0.$$ Then, we obtain the following exact sequence $$H^0(X, K_X+H+A)\to H^0(H, K_H+A)\to H^1(X, K_X+A)\to 0.$$ Since $H^1(X, K_X+A)\neq 0$, the restriction map is not surjective.
We use the following corollary in Section 8.
\[nonzero-criterion\] Let $X$ be a smooth projective surface. Let $L$ be a [$\mathbb{Z}$]{}-divisor on $X$ such that $$L=C+M$$ where $C$ is a smooth prime divisor and $M$ is a nef and big [$\mathbb{Z}$]{}-divisor such that $M|_C$ is ample. If $H^0(C, K_C+M)\neq 0$, then, the trace map $${{\operatorname{Tr}}}_X^e(L):H^0(X, K_X+p^eL)\to H^0(X, K_X+L)$$ is a non-zero map.
By Lemma \[trace-diagram\], we obtain the following commutative diagram [$$\begin{CD}
H^0(X, K_X+p^eL)@>>> H^0(p^eC, K_{p^eC}+p^eM)@>>> H^1(X, K_X+p^eM)\\
@VV{{\operatorname{Tr}}}_{X}^e(L) V @VV \Psi V\\
H^0(X, K_X+L)@>>> H^0(C, K_C+M).
\end{CD}$$]{} By [@T1 Theorem 2.6], we have $H^1(X, K_X+p^eM)=0$ for $e\gg 0$. By Corollary \[curve-trace-cor\], $\Psi$ is a non-zero map because $\Psi$ satisfies $$\begin{aligned}
{{\operatorname{Tr}}}_C^e(M):H^0(C, K_{C}+p^eM)&\to &H^0(p^eC, K_{p^eC}+p^eM)\\
&\overset{\Psi}\to& H^0(C, K_C+M). \end{aligned}$$ Then, the trace map ${{\operatorname{Tr}}}_{X}^e(L)$ is also a non-zero map.
Surjectivity of the trace maps for surfaces ($\kappa=1$)
========================================================
In this section, we show the surjectivity of the trace map $$H^0(X, \omega_X(p^e(A+m(K_X+A))))\to H^0(X, \omega_X(A+m(K_X+A)))$$ for the case where $X$ is a surface and $\kappa(X, K_X+A)=1$. For this, we establish the following vanishing result.
\[ruled-R1\] Let $C$ be a smooth curve. Let $Y:=\mathbb P^1\times C$ and let $\pi:Y\to C$ be the projection. Let $f:X\to Y$ be the blowup at one point and let $$\theta:X\overset{f}\to Y\overset{\pi}\to C.$$ Let $A_X$ be a $\theta$-ample [$\mathbb{Z}$]{}-divisor on $X$. Then, $$R^1\theta_*(B_X^2(A_X))=0.$$
\[step-rational\] In this step, we assume $C$ is rational and we prove the assertion.
Since the assertion is local on $C$, we may assume $C\simeq \mathbb P^1.$ For an arbitrary ample [$\mathbb{Z}$]{}-divisor $A_C$ on $C$, by the Leray spectral sequence, we obtain the following exact sequence $$\begin{aligned}
0&\to & H^1(C, \theta_*(B_X^2(A_X))\otimes \mathcal O_C(A_C))\\
&\to & H^1(X, B_X^2(A_X+\theta^*A_C))\\
&\to & H^0(X, R^1\theta_*(B_X^2(A_X))\otimes \mathcal O_C(A_C))\to 0\\\end{aligned}$$ Let $A_C$ be an ample [$\mathbb{Z}$]{}-divisor on $C$ such that
1. [$A_X+\theta^*A_C$ is ample.]{}
2. [$R^1\theta_*(B_X^2(A_X))\otimes \mathcal O_C(A_C)$ is generated by global sections. ]{}
Then, it is sufficient to show $$H^1(X, B_X^2(A_X+\theta^*A_C))=0.$$ Since $X$ is a toric variety, this follows from Proposition \[F-split-B-vanish\].
\[step-H1-R1\] In this step, we prove that the following assertions are equivalent.
1. [$R^1\theta_*(B_X^2(A_X))=0$.]{}
2. [$H^1(X_c, B_X^2(A_X)|_{X_c})=0$ for every $c\in C$ where $X_c$ is the fiber.]{}
By [@Hartshorne Theorem 12.11], there exists an isomorphism $$R^1\theta_*(B_X^2(A_X))\otimes k(c)\simeq H^1(X_c, B_X^2(A_X)|_{X_c}).$$ By Nakayama’s lemma, if $$R^1\theta_*(B_X^2(A_X))\otimes k(c)=0,$$ then $R^1\theta_*(B_X^2(A_X))|_{U}=0$ for some open neighborhood of $b\in U\subset C$.
We can replace $C$ by a neighborhood of the point corresponding to the singular fiber. Then, we can find the following commutative diagram $$\begin{CD}
X'@<\gamma_X<< X \\
@VVf'V @VVf V\\
Y'@<\gamma_Y<< Y\\
@VV\pi'V @VV\pi V\\
C'@<\gamma_C<< C
\end{CD}$$ where $C'\simeq \mathbb A^1$, each square is a fiber product, $\gamma_C, \gamma_Y$ and $\gamma_X$ are finite étale morphisms. Let $\theta':=\pi'\circ f'.$ Note that $\gamma_X^*B_{X'}^2\simeq B_X^2$ by Lemma \[cartier-etale\].
Let $\tilde c\in C$ be the point corresponding to the singular fiber of $\theta$. Let $\tilde c':=\gamma_C(\tilde c).$ Then, $$\gamma:=\gamma_X|_{X_c}:X_c\to (X')_{c'}$$ is an isomorphism. We can find a $\theta'$-ample [$\mathbb{Z}$]{}-divisor $A_{X'}$ on $X'$ such that $\gamma^*(\mathcal O_{X'}(A_{X'})|_{X_{c'}})\simeq\mathcal O_X(A_X)|_{X_c}$. Then, we obtain $$\begin{aligned}
H^1(X_c, B_X^2(A_X))&=& H^1(X_c, \gamma^*B_{X'}^2(\gamma^*A_{X'}))\\
&=& H^1((X')_{c'}, B_{X'}^2(A_{X'}))\\
&=&0. \end{aligned}$$ The first equation follows from Lemma \[cartier-etale\]. The last equation follows from Step \[step-rational\] and Step \[step-H1-R1\]. Then, the assertion holds by Step \[step-H1-R1\].
Let us prove the main theorem in this section.
\[trace-surface-1\] Let $X$ be a smooth projective surface and let $A$ be an ample [$\mathbb{Z}$]{}-divisor such that $\kappa(X, K_X+A)=1$ and that $K_X+A$ is nef. Then, there exists $m_1\in\mathbb Z_{>0}$ such that the trace map ${{\operatorname{Tr}}}_X^e(A+m(K_X+A))$ $$H^0(X, \omega_X(p^e(A+m(K_X+A))))\to H^0(X, \omega_X(A+m(K_X+A)))$$ is surjective for every $m\geq m_1$ and for every $e\in\mathbb Z_{>0}$.
We see that $K_X+A$ is semi-ample by [@Fujita3].
In this step, we prove that, for some $n_0\in\mathbb Z_{>0}$, the complete linear system $$\Phi_{|n_0(K_X+A)|}=:\theta:X\to C$$ gives a ruled surface structure, that is, $\theta$ is a projective morphism to a smooth projective curve such that $\theta_*\mathcal O_X=\mathcal O_C$ and that a general fiber is $\mathbb P^1$.
We can find $n_0\in\mathbb Z_{>0}$ such that $$\Phi_{|n_0(K_X+A)|}=:\theta:X\to C$$ is a projective morphism to a smooth projective curve such that $\theta_*\mathcal O_X=\mathcal O_C$. Then, by [@Badescu Corollary 7.3], general fibers are integral. Since a general fiber $F$ satisfies $$0=(K_X+A)\cdot F>(K_X+F)\cdot F,$$ we see $F\simeq \mathbb P^1.$ Thus, $\theta$ gives a ruled surface structure.
\[step-R1\] In this step, we prove that it is sufficient to show $$R^1\theta_*(B_X^2(A'))=0$$ for every ample [$\mathbb{Z}$]{}-divisor $A'$.
By Remark \[cartier-remark\] and Remark \[cartier-trace\], we obtain the following exact sequence $$0\to B_X^2\to F_*\omega_X\overset{{{\operatorname{Tr}}}_X^1}\to \omega_X\to 0.$$ By Lemma \[e-vs-e+1\], it is sufficient to show that $$H^1(X, B_X^2(p^d(A+m(K_X+A))))=0$$ for every $0\leq d\leq e$. We can write $K_X+A=\theta^*H$ where $H$ is an ample [$\mathbb{Z}$]{}-divisor on $C$. By the Leray spectral sequence, we obtain $$\begin{aligned}
0&\to&H^1(C, \theta_*(B_X^2(p^d(A+m(K_X+A)))))=0\\
&\to&H^1(X, B_X^2(p^d(A+m(K_X+A))))\\
&\to&H^0(C, R^1\theta_*(B_X^2(p^d(A+m(K_X+A))))\to 0.\end{aligned}$$ The first term vanishes by the Serre vanishing theorem. Thus, it is sufficient to show that $R^1\theta_*(B_X^2(A'))=0$ for every ample [$\mathbb{Z}$]{}-divisor $A'$.
In this step, we prove the assertion. By Step \[step-R1\], it is sufficient to show $$R^1\theta_*(B_X^2(A'))=0$$ for an ample [$\mathbb{Z}$]{}-divisor $A'$. Let $$\theta:X\overset{f}\to Y\overset{\pi}\to C$$ where $\pi$ is a $\mathbb P^1$-bundle structure. Since the problem is local on $C$, we may assume that $\theta$ has only one singular fiber and that $Y=\mathbb P^1\times C.$ Let $F_s$ be the singular fiber. Then, we see $$0=(K_X+A)\cdot F_s=-2+A\cdot F_s.$$ Thus, $F_s$ has at most two irreducible components. This implies that $f$ is the blowup at one point. Then, the assertion follows from Proposition \[ruled-R1\].
Surjectivity of the trace maps for surfaces ($\kappa=2$)
========================================================
In this section, we show the surjectivity of the trace map $$H^0(X, \omega_X(p^e(A+m(K_X+A))))\to H^0(X, \omega_X(A+m(K_X+A)))$$ for the case where $X$ is a surface and $\kappa(X, K_X+A)=2$. Let us recall a lemma on the global generation.
\[uniform-generation\] Let $X$ be a smooth projective variety. Let $A$ be an ample [$\mathbb{Z}$]{}-divisor and let $G$ be a coherent sheaf. Then, there exists $n_0\in\mathbb Z_{>0}$, depending only on $A$ and $G$, such that $$G(n_0A+N)$$ is generated by global sections for every nef [$\mathbb{Z}$]{}-divisor $N$.
The assertion immediately follows from the Castelnuovo–Mumford regularity ([@Lazarsfeld Theorem 1.8.5]) and the Fujita vanishing theorem ([@Fujita1 Theorem (1)], [@Fujita2 Section 5]).
To prove the surjectivity, we establish the following vanishing result.
\[birat-B-vanish\] Let $h:X\to Z$ be a birational morphism between smooth projective surfaces. Let $A_X$ be an ample [$\mathbb{Z}$]{}-divisor and let $A_Z$ be an ample [$\mathbb{Z}$]{}-divisor on $Z$. Then, there exists $m_0\in \mathbb Z_{>0}$ such that $$H^1(X, B_X^2(A_X+h^*(m_0A_Z+N_Z)))=0$$ for every nef [$\mathbb{Z}$]{}-divisor $N_Z$ on $Z$.
The birational morphism $h$ is an $n$-times blowups. We show the assertion by the induction on $n$.
\[step-n=0\] If $n=0$, then the assertion follows from the Fujita vanishing theorem ([@Fujita1 Theorem (1)], [@Fujita2 Section 5]). Thus, we may assume that $n>0$ and the assertion holds for $n-1$.
\[step-one-point\] In this step, we prove that we may assume that $h({{\operatorname{Ex}}}(h))$ is one point.
Let us consider the Leray spectral sequence $$\begin{aligned}
0&\to& H^1(Z, h_*(B_X^2(A_X+h^*(m_0A_Z+N_Z)))=0\\
&\to& H^1(X, B_X^2(A_X+h^*(m_0A_Z+N_Z)))\\
&\to& H^0(Z, R^1h_*(B_X^2(A_X+h^*(m_0A_Z+N_Z))))\to 0\\\end{aligned}$$ where the vanishing $H^1(Z, h_*(B_X^2(A_X+h^*(m_0A_Z+N_Z)))=0$ follows from the Fujita vanishing theorem ([@Fujita1 Theorem (1)], [@Fujita2 Section 5]). By Lemma \[uniform-generation\], the assertion is equivalent to the following vanishing: $$R^1h_*(B_X^2(A_X))=0.$$ Since this problem is local on $Z$, we may assume that $h({{\operatorname{Ex}}}(h))$ is one point.
From now on, we assume that $h({{\operatorname{Ex}}}(h))$ is one point.
\[step-s=0\] Let $$h:X\overset{f}\to Y\overset{g}\to Z$$ where $g$ is the blowup of $Z$ at the point. Let $E_Y$ be the $g$-exceptional curve. Note that $E_Y^2=-1$. We see $$g^*A_Z-\epsilon E_Y$$ is an ample [$\mathbb{Q}$]{}-divisor for every rational number $0<\epsilon\ll 1.$ Thus, by replacing $A_Z$ with its multiple, we may assume that $$A_Y:=g^*A_Z-lE_Y$$ is an ample [$\mathbb{Z}$]{}-divisor for some $l\in\mathbb Z_{>0}$. In particular, we obtain $$h^*A_Z=f^*A_Y+lf^*E_Y.$$ By the induction hypothesis, there exists $m_1\in\mathbb Z_{>0}$ such that $$H^1(X, B_X^2(A_X+f^*(m_1A_Y+N_Y)))=0$$ for every nef [$\mathbb{Z}$]{}-divisor $N_Y$ on $Y$. We have $$m_1h^*A_Z=m_1f^*A_Y+m_1lf^*E_Y.$$
\[step-construct-E\] Let $E_1,\cdots, E_n$ be the $h$-exceptional curves where $E_1$ be the proper transform of $E_Y$. In this step, we construct a sequence of [$\mathbb{Z}$]{}-divisors $$0=:E(0)\leq E(1)\leq E(2)\leq\cdots\leq E(R-1)\leq E(R):=f^*E_Y$$ such that
1. [For every $0\leq r\leq R-1$, $E(r+1)-E(r)=E_j$ for some $1\leq j\leq n$.]{}
2. [$E(r)\cdot E_i\geq -1$ for every $0\leq r\leq R$ and for every $1\leq i\leq n$. ]{}
We consider a decomposition into one point blowups: $$f:X=:X_n\overset{f_n}\to\cdots\overset{f_3}\to X_2\overset{f_2}\to X_1:=Y.$$ We may assume that, for every $2\leq j\leq n$, $E_j\subset X$ is the proper transform of the $f_j$-exceptional curve. For $1\leq j\leq i\leq n$, let $E_j^{(i)}\subset X_i$ be the image of $E_j$. Let $f_{i+1}({{\operatorname{Ex}}}(f_{i+1}))=:P_{i}\in X_{i}$ and let $$g_i:X_i\to Z.$$ Note that $P_i\in {{\operatorname{Ex}}}(g_i)$. Since ${{\operatorname{Supp}}}({{\operatorname{Ex}}}(g_i))$ is simple normal crossing, there are two cases:
1. [$P_i\in E_j^{(i)}$ and $P_i\not\in E_{j'}^{(i)}$ for every $j'\neq j$.]{}
2. [$P_i\in E_j^{(i)}\cap E_{j'}^{(i)}$ for some $j\neq j'$ and $P_i\not\in E_{j''}^{(i)}$ for every $j''\neq j, j'$.]{}
For $1\leq i\leq n$, we construct a finite sequence $({\rm Seq})_i$ of prime divisors on $X$ inductively as follows. Every member of $({\rm Seq})_i$ is $E_j$ for some $j$. Let $$\begin{aligned}
({\rm Seq})_1:=(E_1).\end{aligned}$$ Assume we obtain $({\rm Seq})_{i}$. We construct $({\rm Seq})_{i+1}$ as follows. There are two cases (1) and (2) as above. Assume (1), that is, $P_i\in E_j^{(i)}$ and $P_i\not\in E_{j'}^{(i)}$ for every $j'\neq j$. If $$({\rm Seq})_{i}=(\cdots, E_j,\cdots, E_{j'},\cdots),$$ then we define $({\rm Seq})_{i+1}$ by $$({\rm Seq})_{i+1}:=(\cdots, E_{i+1}, E_j,\cdots, E_{j'},\cdots).$$ In other words, we add $E_{i+1}$ only in front of $E_j$. Assume (2), that is, $P_i\in E_j^{(i)}\cap E_{j'}^{(i)}$ for some $j\neq j'$ and $P_i\not\in E_{j''}^{(i)}$ for every $j''\neq j, j'$. If $$({\rm Seq})_{i}=(\cdots, E_j,\cdots, E_{j'},\cdots, E_{j''},\cdots),$$ then we define $({\rm Seq})_{i+1}$ by $$({\rm Seq})_{i+1}:=(\cdots, E_{i+1}, E_j,\cdots, E_{i+1}, E_{j'},\cdots, E_{j''},\cdots).$$ In other words, we add $E_{i+1}$ only in front of $E_j$ and $E_{j'}$. We obtain a finite sequence $({\rm Seq})_i$ for $1\leq i\leq n.$ Let $$({\rm Seq})_n=(E_{a(1)}, E_{a(2)}, E_{a(3)},\cdots, E_{a(R)})$$ where $a(l)\in\{1, \cdots, n\}.$ We define a finite sequence $({\rm SEQ})$ by $$\begin{aligned}
({\rm SEQ})&=&
(E_{a(1)}, E_{a(1)}+E_{a(2)}, E_{a(1)}+E_{a(2)}+E_{a(3)},\cdots)\\
&=:&(E(1), E(2), E(3),\cdots, E(R)).\end{aligned}$$ By an inductive argument, we see $E(r)\cdot E_j\geq -1$ for every $j$ and $E(R)=f^*E_Y$. This implies the assertion.
\[step-construct-D\] In this step, we construct a sequence of [$\mathbb{Z}$]{}-divisors $$0=:D(0)\leq D(1)\leq D(2)\leq\cdots\leq D(S-1)\leq D(S):=m_1lf^*E_Y$$ such that
1. [For every $0\leq s\leq S-1$, $D(s+1)-D(s)=E_j$ for some $j$. ]{}
2. [$D(s)+A_X+m_1f^*A_Y$ is nef for every $0\leq s\leq S$. ]{}
We define the sequence $\{D(s)\}_{s=0}^S$ by $$\begin{aligned}
&&E(0),E(1),E(2),\cdots, E(R), \\
&&E(R)+E(1),E(R)+E(2),\cdots, 2E(R),\\
&&2E(R)+E(1),2E(R)+E(2),\cdots, 3E(R),\\
&&\cdots\\
&&(m_1l-1)E(R)+E(1),(m_1l-1)E(R)+E(2),\cdots, m_1lE(R).\end{aligned}$$ Then, the sequence $\{D(s)\}_{s=0}^S$ satisfies (a). Thus, we show (b). For every $0\leq s\leq S$, we can write $$D(s)+A_X+m_1f^*A_Y=E(r)+tf^*E_Y+A_X+m_1f^*A_Y.$$ for some $0\leq r\leq R$ and some $0\leq t\leq m_1l-1.$ To show this divisor is nef, it is sufficient to show $$(E(r)+tf^*E_Y+A_X+m_1f^*A_Y)\cdot E_j\geq 0$$ for every $1\leq j\leq n.$ By Step \[step-construct-E\], for every $2\leq j\leq n$, we obtain $$(E(r)+tf^*E_Y+A_X+m_1f^*A_Y)\cdot E_j=(E(r)+A_X)\cdot E_j\geq 0.$$ On the other hand, for the case where $j=1$, we see $$\begin{aligned}
(E(r)+tf^*E_Y+A_X+m_1f^*A_Y)\cdot E_1
&\geq & (tf^*E_Y+m_1f^*A_Y)\cdot E_1\\
&=&(tE_Y+m_1A_Y)\cdot E_Y\\
&\geq &(m_1lE_Y+m_1A_Y)\cdot E_Y\\
&=&0.\end{aligned}$$
\[step-trace\] For a [$\mathbb{Z}$]{}-divisor $D$ on $X$ and for a curve $E\simeq\mathbb P^1$ in $X$, by Lemma \[trace-diagram\], we obtain the following diagram: [$$\begin{CD}
0@>>> H^0(X, \omega_X(pD))@>>> H^0(X, \omega_X(E+pD))@>>> H^0(E, \omega_E(pD))@>>>
H^1(X, \omega_X(pD))\\
@. @VV\alpha:={{\operatorname{Tr}}}_X(D) V @VV\beta:={{\operatorname{Tr}}}_{X, E}(D) V @VV\gamma:={{\operatorname{Tr}}}_E(D) V\\
0@>>>H^0(X, \omega_X(D))@>>> H^0(X, \omega_X(E+D))@>>> H^0(E, \omega_E(D))
\end{CD}$$]{} where the horizontal sequences are exact and the vertical arrows are the trace maps. Then, the following assertions hold.
1. [$\gamma$ is surjective.]{}
2. [If $H^1(X, \omega_X(pD))=0$ and $\alpha$ is surjective, then $\beta$ is also surjective. ]{}
3. [Assume $\beta$ is surjective. Then, the trace map $${{\operatorname{Tr}}}_X(E+D):H^0(X, \omega_X(p(E+D)))\to H^0(X, \omega_X(E+D))$$ is also surjective.]{}
\(1) holds because $E\simeq \mathbb P^1$ is $F$-split (Proposition \[F-split-surje\]). (2) follows from the snake lemma. (3) follows from Remark \[pE-vs-E\].
\[step-final\] Let $m_2\in \mathbb Z_{>0}$ such that $$H^1(X, \omega_X(m_2h^*A_Z+N_X))=0$$ for every nef [$\mathbb{Z}$]{}-divisor $N_X$ on $X$. Note that, since $h^*A_Z$ is nef and big, we can find such an integer $m_2$ by [@T1 Theorem 2.6]. Let $m_0:=m_1+m_2$ and fix a nef [$\mathbb{Z}$]{}-divisor $N_Z$ on $Z$.
We would like to apply the diagram in Step \[step-trace\] for $$D=D(s)+A_X+m_1f^*A_Y+m_2h^*A_Z+h^*N_Z,\,\,\,\,E=D(s+1)-D(s)$$ where $0\leq s\leq S-1$. Note that, by Step \[step-construct-D\], this divisor $D$ is nef. By Step \[step-s=0\], $\alpha$ in Step \[step-trace\] is surjective for $$D=D(0)+A_X+m_1f^*A_Y+m_2h^*A_Z+h^*N_Z.$$ We see $$H^1(X, \omega_X(p(D(s)+A_X+m_1f^*A_Y+m_2h^*A_Z+h^*N_Z)))=0$$ by the choice of $m_2$. Therefore, by Step \[step-construct-D\] and Step \[step-trace\], we obtain the surjection $${{\operatorname{Tr}}}_X(D):H^0(X, \omega_X(pD))\to H^0(X, \omega_X(D))$$ for $$\begin{aligned}
D&=&D(S)+A_X+m_1f^*A_Y+m_2h^*A_Z+h^*N_Z\\
&=&m_1lf^*E_Y+A_X+m_1f^*A_Y+m_2h^*A_Z+h^*N_Z\\
&=&A_X+m_1h^*A_Z+m_2h^*A_Z+h^*N_Z\\
&=&A_X+(m_1+m_2)h^*A_Z+h^*N_Z\\
&=&A_X+h^*(m_0A_Z+N_Z).\end{aligned}$$ Thus, the assertion follows from $$H^1(X, \omega_X(p(A_X+h^*(m_0A_Z+N_Z))))=0.$$
\[birat-surje\] Let $h:X\to Z$ be a birational morphism between smooth projective surfaces. Let $A_X$ be an ample [$\mathbb{Z}$]{}-divisor and let $A_Z$ be an ample [$\mathbb{Z}$]{}-divisor on $Z$. Then, there exists $m_1\in \mathbb Z_{>0}$ such that the trace map ${{\operatorname{Tr}}}^e_X(A_X+h^*(m_1A_Z+N_Z))$ [ $$H^0(X, \omega_X(p^e(A_X+h^*(m_1A_Z+N_Z))))\to H^0(X, \omega_X(A_X+h^*(m_1A_Z+N_Z)))$$]{} is surjective for every $e\in\mathbb Z_{>0}$ and for every nef [$\mathbb{Z}$]{}-divisor $N_Z$ on $Z$.
By Lemma \[e-vs-e+1\], we obtain $${{\operatorname{Tr}}}_{X}^{d+1}(D)={{\operatorname{Tr}}}_{X}^d(D)\circ F_*^d({{\operatorname{Tr}}}_{X}(p^d D)).$$ Thus, the assertion follows from Proposition \[birat-B-vanish\].
Let us prove the main theorem in this section.
\[trace-surface-2\] Let $X$ be a smooth projective surface. Let $A$ be an ample [$\mathbb{Z}$]{}-divisor on $X$ such that $K_X+A$ is nef and big. Then, there exists $m_1\in\mathbb Z_{>0}$ such that the trace map ${{\operatorname{Tr}}}_X^e(A+m(K_X+A))$ $$H^0(X, \omega_X(p^e(A+m(K_X+A))))\to H^0(X, \omega_X(A+m(K_X+A)))$$ is surjective for every $m\geq m_1$ and for every $e\in\mathbb Z_{>0}$.
By Proposition \[birat-surje\], it is sufficient to prove that there exists a birational morphism $$h:X\to Z$$ to a smooth projective surface $Z$ such that $K_X+A$ is the pull-back of an ample [$\mathbb{Z}$]{}-divisor. If $K_X+A$ is ample, then there is nothing to show. We may assume that $K_X+A$ is not ample. Then, by the Nakai–Moishezon criterion, we can find a curve $E$ such that $(K_X+A)\cdot E=0$. This means $K_X\cdot E<0$. Moreover, since $K_X+A$ is big, the equation $(K_X+A)\cdot E=0$ implies $E^2<0$. Therefore, $E$ is a $(-1)$-curve. Contract $E$ and we can repeat this procedure. Then, we obtain $h:X\to Z$.
Main theorem for threefolds
===========================
In this section, we prove the main theorem for threefolds. Let us summarize results on the trace map obtained in previous sections.
\[trace-surface\] Let $X$ be a smooth projective surface. Let $A$ be an ample [$\mathbb{Z}$]{}-divisor on $X$ such that $K_X+A$ is nef and $\kappa(X, K_X+A)\neq 0$. Then, there exists $m_1\in\mathbb Z_{>0}$ such that the trace map ${{\operatorname{Tr}}}_X^e(A+m(K_X+A))$ $$H^0(X, \omega_X(p^e(A+m(K_X+A))))\to H^0(X, \omega_X(A+m(K_X+A)))$$ is surjective for every $m\geq m_1$ and for every $e\in\mathbb Z_{>0}$.
If $\kappa(X, K_X+A)=-\infty$, then there is nothing to show. Thus, we may assume $\kappa(X, K_X+A)\geq 1$. Then, the assertion follows from Theorem \[trace-surface-1\] and Theorem \[trace-surface-2\]
In the above situation, we can show $\kappa(X, K_X+A)\neq -\infty$ by the abundance theorem obtained in [@Fujita3]. Indeed, by Bertini’s theorem, we can find an effective [$\mathbb{Q}$]{}-divisor $D$ such that $\llcorner D\lrcorner=0$ and that $A\sim_{\mathbb Q}D.$
Let us prove the main theorem.
\[threefold-extension\] Let $X$ be a smooth projective threefold. Let $S$ be a smooth prime divisor on $X$ and let $A$ be an ample [$\mathbb{Z}$]{}-divisor on $X$ such that
1. [$K_X+S+A$ is nef. ]{}
2. [$\kappa(S, K_S+A)\neq 0.$]{}
Then, there exists $m_0\in \mathbb Z_{>0}$ such that, for every $m\geq m_0$, the natural restriction map $$H^0(X, m(K_X+S+A))\to H^0(S, m(K_S+A))$$ is surjective.
Let $L:=K_X+S+A$. By Lemma \[trace-diagram\], we obtain the following commutative diagram [ $$\begin{CD}
H^0(X, \omega_X(S+p^eA+p^emL))@>>> H^0(S, \omega_S(p^eA+mp^eL)))@>>>
H^1(X, \omega_X(S+p^eA+p^emL))\\
@VV{{\operatorname{Tr}}}_{X,S}^e(A+mL) V @VV{{\operatorname{Tr}}}_S^e(A+mL) V\\
H^0(X, \omega_X(S+A+mL))@>>> H^0(S, \omega_S(A+mL)).
\end{CD}$$]{} By (2) and Theorem \[trace-surface\], the trace map ${{\operatorname{Tr}}}_S^e(A+mL)$ $$H^0(S, K_S+p^eA+p^emL)\to H^0(S, K_S+A+mL)$$ is surjective for $m\gg 0.$ By the Serre vanishing thoerem, we have $$H^1(X, \omega_X(S+p^eA+p^emL))=0.$$ Therefore, the natural restriction map $$H^0(X, m(K_X+S+A))\to H^0(S, m(K_S+A))$$ is surjective.
Calculation for the case where $\kappa=0$
=========================================
In this section, we consider whether Theorem \[trace-surface\] holds for $\kappa(X, K_X+A)=0$. Let $X$ be a smooth projective surface and let $A$ be an ample [$\mathbb{Z}$]{}-divisor on $X$. Assume that $K_X+A$ is nef and that $\kappa(X, K_X+A)=0$. By the abundance theorem ([@Fujita3]), we see $K_X+A\sim_{\mathbb Q} 0$. Then, $-K_X$ is ample. In particular, $X$ is a rational surface. Thus, we see $K_X+A\sim 0$. We would like to consider the following question.
\[ques-del-Pezzo\] Let $X$ be a smooth projective surface such that $-K_X$ is ample. Is the trace map $${{\operatorname{Tr}}}_X^e(-K_X):H^0(X, \omega_X(-p^eK_X)\to H^0(X, \omega_X(-K_X))$$ surjective?
If $K_X^2\geq 4$, then we obtain an affirmative answer.
Let $X$ be a smooth projective surface such that $-K_X$ is ample. If $K_X^2\geq 4$, then the trace map $${{\operatorname{Tr}}}_X^e(-K_X):H^0(X, \omega_X(-p^eK_X))\to H^0(X, \omega_X(-K_X))$$ is surjective.
Since $h^0(X, \omega_X(-K_X))=1$, it is sufficient to show that the trace map $${{\operatorname{Tr}}}_X^e(-K_X):H^0(X, \omega_X(-p^eK_X))\to H^0(X, \omega_X(-K_X))$$ is a non-zero map. Since $K_X^2\geq 4$, $X$ is obtained by blowing up $\mathbb P^2$ at most $5$ points. Therefore, we can find a smooth conic $C_0$ passing through these points. Let $L_0$ be a line which does not passes through these points. Let $C$ and $L$ be the proper transforms. We see that $L|_C$ is ample, $H^0(C, \omega_C(L|_C))\neq 0$ and $L$ is nef and big. Then, since $C+L\in |-K_X|$, we can apply Corollary \[nonzero-criterion\].
If $X$ is $F$-split, then, by Proposition \[F-split-surje\], the above trace map ${{\operatorname{Tr}}}_X^e(-K_X)$ is surjective. Note that, by [@Hara Example 5.5] and [@Smith Proposition 4.10], if $K_X^2\geq 4$, then $X$ is $F$-split. But, since [@Hara] has no explicit proof, we give the above proof. Moreover, [@Hara Example 5.5] and [@Smith Proposition 4.10] shows that, if $K_X^2=3$ and $X$ is not $F$-split, then $X$ is a Fermat type cubic surface in characteristic two. Indeed, this example gives a negative answer to Question \[ques-del-Pezzo\] as follows.
\[trace-cex\] Let ${\rm char}\,k=p=2$. Consider $\mathbb P^3$ and let $[x:y:z:w]$ be the homogeneous coordinate. Let $$X:=\{[x:y:z:w]\in \mathbb P^3\,|\,x^3+y^3+z^3+w^3=0\}.$$ Then, the trace map $${{\operatorname{Tr}}}_X^e(-K_X):H^0(X, \omega_X(-2^eK_X)\to H^0(X, \omega_X(-K_X))$$ is a zero map.
By Lemma \[e-vs-e+1\], we may assume $e=1$. By Lemma \[trace-diagram\], we obtain the following commutative diagram $$\begin{CD}
H^0(\mathbb P^3, \omega_{\mathbb P^3}(X-2K_{\mathbb P^3}-2X))@>\beta>> H^0(X, \omega_{X}(-2K_{X}))\\
@VV{{\operatorname{Tr}}}_{\mathbb P^3, X}(-K_{\mathbb P^3}-X)V @VV{{\operatorname{Tr}}}_X(-K_X)V\\
H^0(\mathbb P^3, \omega_{\mathbb P^3}(X-K_{\mathbb P^3}-X))@>\alpha>> H^0(X, \omega_{X}(-K_{X})).
\end{CD}$$ We prove that $\alpha$ and $\beta$ are isomorphisms. Since $H^1(\mathbb P^3, L)=0$ for an arbitrary invertible sheaf $L$, it is sufficient to show $$\begin{aligned}
H^0(\mathbb P^3, \omega_{\mathbb P^3}(-K_{\mathbb P^3}-X))
=H^0(\mathbb P^3, \omega_{\mathbb P^3}(-2K_{\mathbb P^3}-2X))
=0.\end{aligned}$$ This follows from $$H^0(\mathbb P^3, \omega_{\mathbb P^3}(-K_{\mathbb P^3}-X))
=H^0(\mathbb P^3, \mathcal O_{\mathbb P^3}(-3))=0$$ and $$H^0(\mathbb P^3, \omega_{\mathbb P^3}(-2K_{\mathbb P^3}-2X))
=H^0(\mathbb P^3, \mathcal O_{\mathbb P^3}(-4+8-6))=0.$$ Therefore, it is sufficient to prove that the trace map ${{\operatorname{Tr}}}_{\mathbb P^3, X}(-K_{\mathbb P^3}-X)$ is a zero map. By Lemma \[D-linearequ\], we obtain $${{\operatorname{Tr}}}_{\mathbb P^3, X}(-K_{\mathbb P^3}-X)={{\operatorname{Tr}}}_{\mathbb P^3, X}(H)$$ where $H$ is defined by $$H:=\{[x:y:z:w]\in \mathbb P^3\,|\,w=0\}.$$ Thus, we show that $${{\operatorname{Tr}}}:={{\operatorname{Tr}}}_{\mathbb P^3, X}(H):H^0(\mathbb P^3, \omega_{\mathbb P^3}(X+2H))\to
H^0(\mathbb P^3, \omega_{\mathbb P^3}(X+H))$$ is a zero map. Let us take a $k$-linear basis of $H^0(\mathbb P^3, \omega_{\mathbb P^3}(X+2H))$. It is easy to see $$h^0(\mathbb P^3, \omega_{\mathbb P^3}(X+2H))=4.$$ Let ${{\operatorname{Spec}}}\,k[X, Y, Z]\subset \mathbb P^3$ be the affine open subset defined by $w\neq 0$. Consider the following four $3$-forms $$\begin{aligned}
\eta_1&:=&\frac{1}{X^3+Y^3+Z^3+1}dX\wedge dY\wedge dZ\\
\eta_X&:=&\frac{X}{X^3+Y^3+Z^3+1}dX\wedge dY\wedge dZ\\
\eta_Y&:=&\frac{Y}{X^3+Y^3+Z^3+1}dX\wedge dY\wedge dZ\\
\eta_Z&:=&\frac{Z}{X^3+Y^3+Z^3+1}dX\wedge dY\wedge dZ.\\\end{aligned}$$ These are elements of $\omega_{k(X, Y, Z)}=(\omega_{\mathbb P^3})_{\xi}$ where $\xi$ is the generic point of $\mathbb P^3$. By a direct calculation, these four elements are linearly independent and $\eta_1, \eta_X,\eta_Y, \eta_Z\in H^0(\mathbb P^3, \omega_{\mathbb P^3}(X+2H)).$ In particular, these four elements form a $k$-linear basis of $H^0(\mathbb P^3, \omega_{\mathbb P^3}(X+2H))$. The trace map is a $p^{-1}$-linear map, that is, for $a, b, c, d\in k$, $$\begin{aligned}
&&{{\operatorname{Tr}}}(a\eta_1+b\eta_X+c\eta_Y+d\eta_Z)\\
&=&a^{\frac{1}{p}}{{\operatorname{Tr}}}(\eta_1)+b^{\frac{1}{p}}{{\operatorname{Tr}}}(\eta_X)+c^{\frac{1}{p}}{{\operatorname{Tr}}}(\eta_Y)+d^{\frac{1}{p}}{{\operatorname{Tr}}}(\eta_Z).\end{aligned}$$ Thus, it is sufficient to show $${{\operatorname{Tr}}}(\eta_1)={{\operatorname{Tr}}}(\eta_X)={{\operatorname{Tr}}}(\eta_Y)={{\operatorname{Tr}}}(\eta_Z)=0.$$ Let us only prove ${{\operatorname{Tr}}}(\eta_X)=0$. This follows from $$\begin{aligned}
&&({{\operatorname{Tr}}}(\eta_X))|_{{{\operatorname{Spec}}}\,k[X, Y, Z]}\\
&=&{{\operatorname{Tr}}}(\frac{X}{X^3+Y^3+Z^3+1}dX\wedge dY\wedge dZ)\\
&=&{{\operatorname{Tr}}}(\frac{X(X^3+Y^3+Z^3+1)}{(X^3+Y^3+Z^3+1)^2}dX\wedge dY\wedge dZ)\\
&=&\frac{1}{X^3+Y^3+Z^3+1}{{\operatorname{Tr}}}((X^4+XY^3+XZ^3+X)dX\wedge dY\wedge dZ)\\
&=&0.\end{aligned}$$ The last equality follows from Remark \[trace-local\].
The following example shows that we cannot prove Theorem \[threefold-extension\] for the case where $\kappa=0$ by the same proof.
Let ${\rm char}\,k=p=2$. Then, there exist smooth projective threefold $X$ over $k$, a smooth prime divisor $S_0$ on $X$ and an ample [$\mathbb{Z}$]{}-divisor $A$ on $X$ which satisfy the following properties.
1. [$|K_X+S_0+A|$ is basepoint free. ]{}
2. [The natural restriction map $$H^0(X, m(K_X+S_0+A))\to H^0(S_0, m(K_{S_0}+A))$$ is surjective for every $m\in\mathbb Z_{>0}$.]{}
3. [ The trace map ${{\operatorname{Tr}}}_{S_0}^e(A+m(K_{S_0}+A))$ [$$H^0(S_0, \omega_{S_0}(2^e(A+m(K_{S_0}+A))))\to
H^0(S_0, \omega_{S_0}(A+m(K_{S_0}+A)))$$]{} is a zero map for every $m\in \mathbb Z_{>0}$ and for every $e\in\mathbb Z_{>0}.$]{}
Let $S$ be the surface in Theorem \[trace-cex\] and let $A_S:=-K_S$. Let $C$ be a smooth projective curve and fix an arbitrary ample [$\mathbb{Z}$]{}-divisor $A_C$ on $C$. Let $X:=S\times C$ and let $\pi_S$ and $\pi_C$ be their projection respectively. Fix a point $c_0\in C$ and let $S_0:=S\times \{c_0\}$. Let $$A:=\pi_S^*A_S+\pi_C^*A_C.$$ Note that $A|_{S_0}=-K_{S_0}$. Thus, (3) follows from Theorem \[trace-cex\]. (1) follows from $$\begin{aligned}
K_X+S_0+A&=&\pi_S^*(K_S+A_S)+\pi_C^*(K_C+c_0+A_C)\\
&=&\pi_C^*(K_C+c_0+A_C).\end{aligned}$$ It is sufficient to show (2). This follows from [ $$\begin{aligned}
&&H^1(X, K_X+A+(m-1)(K_X+S_0+A))\\
&=&H^1(X, \pi_C^*(K_C+A_C+(m-1)(K_C+c_0+A_C)))\\
&\simeq&H^1(S, \mathcal O_S)\otimes_k H^0(C, K_C+A_C+(m-1)(K_C+c_0+A_C))\\
&\oplus &H^0(S, \mathcal O_S)\otimes_k H^1(C, K_C+A_C+(m-1)(K_C+c_0+A_C))\\
&=&0.\end{aligned}$$]{} The last equality holds by $H^1(S, \mathcal O_S)=0$ and $$\deg_C(A_C+(m-1)(K_C+c_0+A_C))>0.$$
Appendix: Extension theorem for surfaces
========================================
For the surface case, we can freely use the minimal model theory (cf. [@Fujita3], [@KK], [@T2]). By using results obtained in [@Fujita3], [@T2] and [@T3], we can establish an analogy of [@HM Theorem 5.4.21] as follows.
\[surface-ext\] Let $X$ be a smooth projective surface and let $C$ be a smooth prime divisor on $X$. Let $\Delta:=C+B$ where $B$ is an effective [$\mathbb{Q}$]{}-divisor which satisfies the following properties.
1. [$C\not\subset{{\operatorname{Supp}}}B$, $\llcorner B\lrcorner=0$ and $(X, \Delta)$ is plt. ]{}
2. [$B\sim_{\mathbb Q}A+F$ where $A$ is an ample [$\mathbb{Q}$]{}-divisor and $F$ is an effective [$\mathbb{Q}$]{}-divisor such that $C\not\subset {{\operatorname{Supp}}}F.$]{}
3. [No prime component of $\Delta$ is contained in the stable base locus of $K_X+\Delta.$]{}
Then, there exists an integer $m_0>0$ such that, for every integer $m>0$, the restriction map $$H^0(X, mm_0(K_X+\Delta))\to H^0(C, mm_0(K_X+\Delta)|_C)$$ is surjective.
In this step we prove that, if $E$ is a curve in $X$ such that $E^2<0$ and $(K_X+\Delta)\cdot E<0$, then the following three assertions hold:
1. [$K_X\cdot E=E^2=-1$.]{}
2. [$E$ is not a prime component of $\Delta$.]{}
3. [$E\cdot C=0$.]{}
Since $(K_X+E)\cdot E\leq (K_X+\Delta)\cdot E<0$, there exists a birational morphism $f:X\to Y$ to a normal [$\mathbb{Q}$]{}-factorial surface $Y$ such that ${{\operatorname{Ex}}}(f)=E$ (cf. [@T2 Theorem 6.2]). Let $\Delta_Y:=f_*\Delta$ and we define $d\in\mathbb Q$ by $$K_X+\Delta=f^*(K_Y+\Delta_Y)+dE.$$ The inequality $(K_X+\Delta)\cdot E<0$ means $d>0.$ We can find a integer $l>0$ such that $l(K_Y+\Delta_Y)$ is Cartier. Then, $E$ is a fixed component of $$l(K_X+\Delta)=f^*(l(K_Y+\Delta_Y))+dlE.$$ Then, the assumption (3) implies (b). This means $E\cdot \Delta\geq 0.$ Then, the assertion (a) follows from $$K_X\cdot E\leq (K_X+\Delta)\cdot E<0.$$ Let us show (c). Assume $E\cdot C>0.$ Then, $E\cdot C\geq 1.$ This means the following contradiction $$0>(K_X+\Delta)\cdot E=K_X\cdot E+C\cdot E+B\cdot E\geq -1+1+0=0.$$
In this step, we prove that we may assume that $K_X+\Delta$ is nef.
Assume that $K_X+\Delta$ is not nef. Then, there exists a curve $E$ such that $(K_X+\Delta)\cdot E<0$. By (3), we can find $l>0$ such that $|l(K_X+\Delta)|\neq\emptyset.$ This means $E^2<0$. We see that $E$ is a $(-1)$-curve by Step 1. Let $f:X\to Y$ be the contraction of $E$. Let $$\Delta_Y:=f_*\Delta\,,\,\,C_Y:=f_*C\,,\,\,B_Y:=f_*B\,,\,\,A_Y:=f_*A\,,\,\,F_Y:=f_*F.$$ Then, $Y$ and these divisors also satisfies the conditions (1), (2) and (3). Let $m_1>0$ be an integer such that $m_1\Delta$ is a [$\mathbb{Z}$]{}-divisor. Then, we have $$m_1(K_X+\Delta)=f^*(m_1(K_Y+\Delta_Y))+eE$$ for some $e\in\mathbb Z_{>0}.$ Let $n$ be an arbitrary positive integer. By $f_*\mathcal O_X=\mathcal O_Y$, we have [$$f_*(\mathcal O_X(nm_1(K_X+\Delta)))\simeq f_*(\mathcal O_X(nm_1f^*(K_Y+\Delta_Y)))\simeq
\mathcal O_Y(nm_1(K_Y+\Delta_Y)).$$]{} By $C\cap E=\emptyset$, we see [$$f_*(\mathcal O_C(nm_1(K_X+\Delta)))\simeq f_*(\mathcal O_C(nm_1f^*(K_Y+\Delta_Y)))\simeq
\mathcal O_{C_Y}(nm_1(K_Y+\Delta_Y)).$$]{} These implies the following commutative diagram $$\begin{CD}
H^0(X, nm_1(K_X+\Delta))@>>> H^0(C, nm_1(K_X+\Delta)|_C)\\
@| @|\\
H^0(Y, nm_1(K_Y+\Delta_Y))@>>> H^0(C_Y, nm_1(K_Y+\Delta_Y)|_{C_Y})
\end{CD}$$ where the horizontal arrows are the natural restriction maps. Thus, we can reduce the problem on $X$ to the problem on $Y$. We can repeat the same argument. Note that $|l(K_X+\Delta)|\neq\emptyset$ and this means the assertion.
By the abundance theorem ([@Fujita3]), $K_X+\Delta$ is semi-ample. Let $$f:=\varphi_{|m_2(K_X+\Delta)|}:X\to R$$ for some $m_2\in\mathbb Z_{>0}$ such that $f_*\mathcal O_X=\mathcal O_R.$ In this step, we prove $f_*\mathcal O_C=\mathcal O_{f(C)}$.
Assume $f_*\mathcal O_C\neq \mathcal O_{f(C)}$. We run the $(K_X+\{\Delta\})$-MMP. By [@T3], there exists a morphisms $$X\overset{g}\to V\to R$$ where $V$ is a smooth projective curve such that a general fiber $G$ of $g$ satisfies $G\simeq\mathbb P^1$ and $\llcorner\Delta\lrcorner\cdot G=2.$ Note that $B=\{\Delta\}$ and $C=\llcorner\Delta\lrcorner$. Since $G$ is a fiber and $B$ is big, we see $G\cdot B>0$. Then, we obtain the following contradiction $$0=(K_X+\Delta)\cdot G\geq(K_X+B)\cdot G+2>(K_X+G)\cdot G+2=0.$$
In this step, we prove the assertion. Let $$f:=\varphi_{|m_2(K_X+\Delta)|}:X\to R$$ such that $f_*\mathcal O_X=\mathcal O_R$ and let $f(C)=:D.$ Then, by Step 3, we see $f_*\mathcal O_C=\mathcal O_D.$ Let $H$ be an ample Cartier divisor on $R$ such that $m_2(K_X+\Delta)=f^*H.$ By the Serre vanishing theorem, we can find $m_3\in\mathbb Z_{>0}$ such that $$H^0(R, mm_3H)\to H^0(D, mm_3 H)$$ is surjective for every $m\in \mathbb Z_{>0}.$ Then, by $f_*\mathcal O_X=\mathcal O_R$ and $f_*\mathcal O_C=\mathcal O_D$, we have the following commutative diagram $$\begin{CD}
H^0(X, mm_2m_3(K_X+\Delta))@>>> H^0(C, mm_2m_3(K_X+\Delta)|_C)\\
@| @|\\
H^0(R, mm_3H)@>>> H^0(D, mm_3H|_{D}).
\end{CD}$$ This implies the assertion.
[Kollár-Mori]{} , [Quasi-log varieties]{}, [Tr. Mat. Inst. Steklova [**240**]{} (2003), Biratsion. Geom. Linein. Sist. Konechno Porozhdennye Algebry, 220–239; translation in Proc. Steklov Inst. Math. 2003, no. 1 (240), 214–233]{}.
, [*Algebraic Surfaces*]{}, [Universitext, Springer-Verlag, New York, 2001]{}.
, [Existence of minimal models for varieties of log general type]{}, [J. Amer. Math. Soc. [**23**]{} (2010), no. 2, 405–468]{}.
, [Discreteness and rationality of $F$-jumping numbers on singular varieties]{}, [Math. Ann. [**347**]{} (2010), 917-949.]{}
, [$F$-singulaties via alterations]{}, [preprint (2012)]{}.
, On the numerical dimension of pseudo-effective divisors in positive characteristic, [preprint (2012).]{}
, Extension theorems, Non-vanishing and the existence of good minimal models, [preprint (2012).]{}
, [*Lectures on Vanishing Theorems*]{}, [Birkhäuser, Basel, 1992.]{}
, Fundamental theorems for the log minimal model program, [ Publ. Res. Inst. Math. Sci. 47 (2011), no. 3, 727–789.]{}
, Log pluricanonical representations and abundance conjecture, [preprint (2011). ]{}
, Vanishing theorems for semipositive line bundles, [Algebraic geometry (Tokyo/Kyoto, 1982), 519–528, Lecture Notes in Math., [**1016**]{}, Springer, Berlin, 1983]{}.
, Semipositive line bundles, [J. Fac. Sci. Univ. Tokyo Sect. IA Math. [**30**]{} (1984), no 3 353–378]{}.
, Fractionally logarithmic canonical rings of algebraic surfaces, [J. Fac. Sci. Univ. Tokyo Sect. IA Math. [**30**]{} (1984), no 3 685–696]{}.
, [Extension theorem and the existence of flips]{}, [in [*Flips for $3$-folds and $4$-folds*]{}, Oxford University Press, 2007, 79–100]{}.
, [On the three dimensional minimal model program in positive characteristic, preprint (2013)]{}.
, [A characterization of rational singularities in terms of injectivity of Frobenius maps]{}, [Amer. J. Math., [**120**]{}, (1998), 981–996]{}.
, [*Algebraic Geometry.*]{}, [Grad. Texts in Math., no [**52**]{}, Springer-Verlag, NewYork, 1977]{}.
, Semistable minimal models of threefolds in positive or mixed characteristic, [J. Alg. Geom. [**3**]{} (1994), 463–491]{}.
, Introduction to the Minimal Model Program, [volume [**10**]{} of adv. Stud. Pure Math., 283–360. Kinokuniya–North–Holland, 1987]{}.
, Basepoint freeness for nef and big linebundles in positive characteristic, [Ann. Math, [**149**]{} (1999), 253–286]{}.
, Extremal rays on smooth threefolds, [Ann. Sci. Ec. Norm. Sup., [**24**]{} (1991), 339–361]{}.
, Birational geometry of log surfaces, [preprint]{}.
, [*Birational geometry of algebraic varieties*]{}, [Cambrigde Tracts in Mathematics, Vol. [**134**]{}, 1998]{}.
, [*Positivity in Algebraic Geometry I*]{}, [A series of Modern Surveys in Mathematics, Vol. [**48**]{}, Springer, 2004]{}.
, [Frobenius splitting and cohomology vanishing for Schubert varieties]{}, [Ann. of Math. (2), [**122**]{} (1985), 27–40]{}.
, [*Abelian varieties*]{}, [Grad. Texts in Math., no [**52**]{}, Springer-Verlag, NewYork, 1977]{}.
, The non-nef locus in positive characteristic , [preprint (2012).]{}
, Contre-exemple au “vanishing theorem” en caractéristique $p>0$, [C. P. Ramanujam — A tribute, Studies in Math., [**8**]{} (1978), 273–278]{}.
, $F$-adjunction, [Algebra and Number Theory, (2009) Vol. 3, No. 8, 907–950]{}.
, A canonical linear system associated to adjoint divisors in characteristic $p > 0$ , [to appear in Journal für die reine und angewandte Mathematik (2012)]{}.
, [Vanishing, singularities and effective bounds via prime characteristic local algebra]{}, [in [*Algebraic geometry–Santa Cruz 1995,*]{} Proc. Sympos. Pure Math., [**62**]{}, Part 1, Amer. Math. Soc., Procidence, RI, 1997, 289–325]{}.
, On $F$-pure threshold, [J. Algebra 282 (2004), no.1, 278-297.]{}
, The X-method for klt surfaces in positive characteristic, [to appear in J. Alg. Geom. (2012)]{}.
, Minimal models and abundance for positive characteristic log surfaces, preprint (2012).
, Abundance theorem for semi log canonical surfaces in positive characteristic, [preprint (2013)]{}.
, On the behavior of extensions of vector bundles under the Frobenius map, [Nagoya Math. J., [**48**]{} (1972), 73–89]{}.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'In this work we prove the nonlinear instability of inhomogeneous steady states solutions to the Hamiltonian Mean Field (HMF) model. We first study the linear instability of this model under a simple criterion by adapting the techniques developed in [@LIN1]. In a second part, we extend to the inhomogeneous case some techniques developed in [@grenier; @Han-Hauray; @Han-Nguyen] and prove a nonlinear instability result under the same criterion.'
author:
- 'M. Lemou'
- 'A. M. Luz'
- 'F. Méhats'
title: Nonlinear instability of inhomogeneous steady states solutions to the HMF Model
---
Introduction
============
In this paper, we are interested in the nonlinear instability of inhomogeneous steady states of the Hamiltonian Mean Field (HMF) system. The HMF system is a kinetic model describing particles moving on a unit circle interacting via an infinite range attractive cosine potential. This 1D model holds many qualitative properties of more realistic long-range interacting systems as the Vlasov-Poisson model. The HMF model has been the subject of many works in the physical community, for the study of non equilibrium phase transitions [@chavanis1; @chavanis2; @antoniazzi; @ogawa2], of travelling clusters [@yamaguchi1; @yamaguchi2] or of relaxation processes [@barre1; @barre2; @chavanis4]. The long-time validity of the N-particle approximation for the HMF model has been investigated in [@rousset1; @rousset2] and the Landau-damping phenomenon near a spatially homogeneous state has been studied recently in [@faou-rousset]. The formal linear stability of inhomogeneous steady states has been studied in [@chavanis3; @OGW; @yamaguchi3]. In particular, a simple criterion of linear stability has been derived in [@OGW]. In [@llmehats], the authors of the present paper have proved that, under the same criterion $\kappa_0<1$ (see below for a precise formulation), the inhomogeneous steady states of HMF that are nonincreasing functions of the microscopic energy are nonlinearly stable. The aim of the present paper is to show, in a certain sense, that this criterion is sharp: we show that if $\kappa_0>1$, the HMF model can develop instabilities, from both the linear and the nonlinear points of view.
In [@LIN2], Guo and Lin have derived a sufficient criterion for linear instability to 3D Vlasov-Poisson by extending an approach developped in [@LIN1] for BGK waves. Let us also mention that both works have adapted some techniques presented in [@guo-strauss] to prove the nonlinear instability of the 3D Vlasov-Poisson system. In the first part of this article we adapt these techniques and prove the linear instability of nonhomogeneous steady states to the HMF system. In [@Han-Hauray], a nonlinear instability result for 1D Vlasov-Poisson equation was obtained for an initial data close to stationary homogeneous profiles that satisfy a Penrose instability criterion by using an approach developed in [@grenier]. In [@Han-Nguyen], starting with the N- particles version of the HMF model, a nonlinear instability result is obtained for the corresponding Vlasov approximation by also considering a Penrose instability condition for stationary homogeneous profiles. In the second part of this article our aim is to prove the nonlinear instability of non-homogeneous steady states of HMF by adapting the techniques developed in [@grenier; @Han-Hauray; @Han-Nguyen].
In the HMF model, the distribution function of particles $f(t,\theta,v)$ solves the initial-valued problem $$\begin{aligned}
\label{sis1}
&\partial_t f+v\partial_{\theta} f-\partial_{\theta} \phi_f\partial_v f=0,\qquad(t,\theta,v)\in{{\mathbb{R}}}_{+}\times\TT\times{{\mathbb{R}}},\\
&f(0,\theta,v)=f_{\rm init}(\theta,v)\geq 0,\nonumber\end{aligned}$$ where $\TT$ is the flat torus $[0,2\pi]$ and where the self-consistent potential $\phi_f$ associated with a distribution function $f$ is defined by $$\label{phi}
\phi_f(\theta)=-\int^{2 \pi}_{0}\rho_f(\theta')\cos(\theta-\theta')d\theta', \qquad \rho_f(\theta)=\int_{{{\mathbb{R}}}}f(\theta,v)dv.$$
Introducing the so-called magnetization vector defined by $$\label{mag2}
M_f=\int^{2 \pi}_{0}\rho_f(\theta)u(\theta) d\theta, \qquad \mbox{with}\quad u(\theta)=(\cos \theta,\sin \theta)^T$$ we have $$\label{phimag}
\phi_f(\theta)=-M_f\cdot u(\theta).$$
In this work will consider steady states of of the form $$\label{Qexp}
f_0(\theta,v)=F(e_0(\theta,v)),$$ where F is a given nonnegative function and where the microscopic energy $e_0(\theta,v)$ is given by $$\label{eqe0}
e_0(\theta,v)=\frac{v^2}{2}+\phi_0(\theta) \quad \text{with}\,\,\phi_0=\phi_{f_0}.$$ Without loss of generality, we assume that $\phi_0(\theta)=-m_0\cos \theta$ with $m_0>0$. Here $m_0$ is the magnetization of the stationary state $f_0$ defined by $m_0=\int \rho_{f_0}\cos \theta d\theta$.
It is shown in [@llmehats] that (essentially) if $F$ is decreasing then $f_0$ is nonlinearly stable by the HMF flow provided that the criterion $\kappa_0<1$ is satisfied, where $\kappa_0$ is given by $$\label{kappa0}
\kappa_0=-\int^{2 \pi}_{0}\!\!\int^{+\infty}_{-\infty} F'\!\left(e_0(\theta,v)\right)\left(\frac{\ds \int_{\mathcal D_{e_0(\theta,v)}}(\cos \theta-\cos \theta')(e_0(\theta,v)+m_0\cos \theta')^{-1/2}d\theta'}{\ds\int_{\mathcal D_{e_0(\theta,v)}}(e_0(\theta,v)+m_0\cos \theta')^{-1/2}d\theta'} \right)^2d\theta dv,$$ with $${\mathcal D_e}=\left\{\theta'\in \TT\,:\,\,m_0\cos \theta'>-e\right\}.$$
In this paper, we explore situations where this criterion is not satisfied, i.e when $\kappa_0>1$. Let us now state our two main results. The first one concerns the linearized HMF equation given by $$\label{sislin}
\partial_t f=Lf,$$ where $$\label{defL}
Lf:=-v\partial_{\theta} f+\partial_{\theta} \phi_0\partial_v f+\partial_{\theta} \phi_f\partial_{v} f_0.$$
\[mainthm1\] Let $f_0\in L^1(\TT,\RR)$ be a stationary solution of of the form , where $F$ is a nonnegative $\mathcal C^1$ function on $\RR$ such that $F'(e_0(\theta,v))$ belongs to $L^1(\TT,\RR)$. Assume that $\kappa_0>1$, where $\kappa_0$ is given by . Then there exists ${\lambda}>0$ and a non-zero $f \in L^1(\TT\times\RR)$ such that $e^{\lambda t}f$ is a nontrivial growing mode weak solution to the linearized HMF equation .
Our second result is the following nonlinear instability theorem.
\[mainthm2\] Let $f_0$ be a stationary solution of of the form , where $F$ is a $\mathcal {C} ^{\infty}$ function on $\RR$, such that $F(e)>0$ for $e<e_*$, $F(e)=0$ for $e\geq e_*$, with $e_*<m_0$ and $|F'(e)|\leq C|e_*-e|^{-\alpha}F(e)$ in the neighborhood of $e_*$, for some $\alpha\geq 1$. Assume that $\kappa_0>1$, where $\kappa_0$ is given by . Then $f_0$ is nonlinearly unstable in $L^1(\TT\times{{\mathbb{R}}})$, namely, there exists $\delta_0>0$ such that for any $\delta>0$ there exists a nonnegative solution $f(t)$ of satisfying $\| f(0)-f_0\|_{L^1}\leq \delta$ and $$\| f(t_{\delta})-f_0\|_{L^1}\geq \delta_0,$$ with $t_{\delta}=O(|\log \delta|)$ as $\delta \rightarrow 0$.
Note that in these two theorems we do not assume that the profile $F$ is a decreasing function. Besides, the set of steady states satisfying the assumptions of these theorems is not empty, as proved in the Appendix (see Lemma \[lemapp\]). Note also that the instability of Theorem \[mainthm2\] is not due to the usual orbital instability. Indeed the functional space of the pertubation can be restricted to the space of even functions in $(\theta,v)$.
The outline of the paper is as follows: Sections \[section2\] and \[section3\] are respectively devoted to the proofs of Theorem \[mainthm1\] and Theorem \[mainthm2\].
A linear instability result: proof of Theorem \[mainthm1\] {#section2}
==========================================================
The aim of this section is to prove Theorem \[mainthm1\]. This proof will be done following the framework used by Lin for the study of periodic BGK waves in [@LIN1], which was generalized to the analysis of instabilities for the 3D Vlasov-Poisson system by Guo and Lin in [@LIN2]. We divide this proof into the three Lemmas \[lem1\], \[lim1G\] and \[lim2G\], respectively proved in Subsections \[fam\], Subsection \[lim0\] and Subsection \[liminfty\].
A growing mode of is a solution of the form $e^{\lambda t}f$, where $f\in L^1(\TT,\RR)$ is an unstable eigenfunction of $L$, i.e. a nonzero function satisfying $Lf=\lambda f$ in the sense of distributions, with $\lambda \in \RR_+^*$ and with $L$ defined by . Note that the equation $Lf-\lambda f=0$ is invariant by translation: if $f(\theta,v)$ is an eigenfunction, then for all $\theta_0$, $f(\theta+\theta_0,v)$ is also an eigenfunction. Since for all $f\in L^1$ we can find a $\theta_0\in \TT$ such that $\int \rho_f(\theta+\theta_0)\sin \theta d\theta=0$, we can assume that our eigenfunction of $L$ always satisfy $\int \rho_f\sin \theta d\theta=0$, i.e. $\phi_f=-m\cos \theta$ with $m=\int\rho_f\cos \theta d\theta$.
Let us first define $\left(\Theta(s,\theta,v),V(s,\theta,v)\right)$ as the solution of the characteristics problem $$\label{siscarac}
\left\{ \begin{array}{l}
\ds\frac{d\Theta(s,\theta,v)}{ds}=V(s,\theta,v)\\[2mm]
\ds\frac{dV(s,\theta,v)}{ds}=-\partial_{\theta}\phi_0\left(\Theta(s,\theta,v)\right)
\end{array} \right.$$ with initial data $\Theta(0,\theta,v)=\theta$, $V(0,\theta,v)=v$. When there is no ambiguity, we denote simply $\Theta(s)=\Theta(s,\theta,v)$ and $V(s)=V(s,\theta,v)$. Since $\phi_0(\theta)=-m_0\cos \theta$, the solution $(\Theta, V)$ is globally defined and belongs to $\mathcal C^\infty(\RR\times\TT\times\RR)$. Note that the energy $e_0\left(\Theta(s),V(s)\right)=\frac{V(s)^2}{2}+\phi_0(\Theta(s))$ does not depend on $s$.
We shall reduce the existence of a growing mode of to the existence of a zero of the following function, defined for all $\lambda\in \RR_+^*$: $$\begin{aligned}
G(\lambda) = &1+\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'\left(\frac{v^2}{2}-m_0\cos\theta\right) \cos^2\theta d\theta dv\nonumber\\
&\quad -\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'\left(\frac{v^2}{2}-m_0\cos\theta\right)\left(\int_{-\infty}^{0}\lambda e^{\lambda s}\cos{\Theta}(s,\theta,v)ds\right) \cos \theta d\theta dv.\label{glambda}\end{aligned}$$
\[lem1\] Let $f_0\in L^1(\TT,\RR)$ be a stationary solution of of the form , where $F$ is a $\mathcal C^1$ function on $\RR$ such that $F'(e_0(\theta,v))$ belongs to $L^1(\TT,\RR)$. Then the function $G$ defined by is well-defined and continuous on $\RR_+^*$. Moreover, there exists a growing mode $e^{\lambda t}f$ solution to associated with the eigenvalue $\lambda>0$ if and only if $G({\lambda})=0$. An unstable eigenfunction $f$ of $L$ is defined by $$\label{flem2}
f(\theta,v)=-F'(e_0(\theta,v))\cos\theta+F'(e_0(\theta,v))\int_{-\infty}^{0}\lambda e^{\lambda s}\cos\left({\Theta}(s,\theta,v)\right)ds.$$
\[lim1G\] Under the assumptions of Lemma \[lem1\], the function $G$ defined by satisfies $$\label{limG}
\lim \limits_{\,\,\lambda\to 0^+}G(\lambda)=1-\kappa_0,$$ where $\kappa_0$ is defined by .
\[lim2G\] Under the assumptions of Lemma \[lem1\], the function $G$ defined by satisfies $$\label{limGinfty}
\lim \limits_{\,\,\lambda\to +\infty}G(\lambda)=1.$$
From these three lemmas, it is clear that if $\kappa_0>1$, we have $\lim \limits_{\,\,\lambda\to 0^+}G(\lambda)<0$ and $\lim \limits_{\,\,\lambda\to +\infty}G(\lambda)>0$ so by continuity of $G$, there exists $\lambda>0$ such that $G(\lambda)=0$. This means that there exists a growing mode to and this proves Theorem \[mainthm1\].
First properties of the function $G(\lambda)$: proof of Lemma \[lem1\] {#fam}
----------------------------------------------------------------------
In this subsection, we prove Lemma \[lem1\]. Let $\lambda\in\RR_+^*$. Since, by assumption, the function $F'(e_0(\theta,v))$ belongs to $L^1(\TT\times\RR)$, and since $$\forall (\theta,v)\qquad \left|\int_{-\infty}^{0}\lambda e^{\lambda s}\cos({\Theta}(s))ds\right|\leq \int_{-\infty}^{0}\lambda e^{\lambda s}ds=1,$$ both functions $$F'(e_0(\theta,v))\cos\theta\quad\mbox{and}\quad F'(e_0(\theta,v))\int_{-\infty}^{0}\lambda e^{\lambda s}\cos\left({\Theta}(s,\theta,v)\right)ds$$ belong to $L^1(\TT\times \RR)$, so the function $f$ defined by also belongs to $L^1(\TT\times \RR)$. Hence, by integrating with respect to $\cos \theta d\theta dv$, we deduce that $G(\lambda)$ is well-defined by . The continuity of $G$ on $\RR_+^*$ stems from dominated convergence.
Consider now a (nonzero) growing mode $e^{\lambda t}f$ of associated to an eigenvalue $\lambda>0$. Let us prove that $G(\lambda)=0$. From $Lf=\lambda f$ and , we get, in the sense of distributions, $$\begin{aligned}
\frac{d}{ds}\left(e^{\lambda s}f\left(\Theta(s),V(s)\right)\right)&=e^{\lambda s}\phi'_f\left(\Theta(s)\right)V(s)F'(e_0(\Theta,V)).\end{aligned}$$ Integrating this equation from $-R$ to $ 0$, we get, for almost all $(\theta,v)$ and all $R$, $$f(\theta,v)=e^{-\lambda R}f\left(\Theta(-R),V(-R)\right)+F'(e_0)\int_{-R}^0 e^{\lambda s} \phi'_f\left(\Theta(s)\right)V(s)ds,$$ where we recall that $e_0(\Theta,V)=e_0(\theta,v)$. We multiply by a test function $\psi(\theta,v) \in C_0^{\infty}(\TT\times\RR)$ and integrate with respect to $(\theta,v)$, $$\begin{aligned}
\int_{0}^{2\pi}\int_{\RR} f(\theta,v)\psi(\theta,v)d\theta dv&=&e^{-\lambda R}\int_{0}^{2\pi}\int_{\RR} f\left(\theta,v\right)\psi\left(\Theta(R),V(R)\right)d\theta dv\\
&&+\int_{-R}^0\int_{0}^{2\pi}\int_{\RR} e^{\lambda s}F'(e_0) \phi'_f\left(\Theta(s)\right)V(s)\psi\left(\theta,v\right)dsd\theta dv. \end{aligned}$$ In the first integral of the right-hand side, we have performed the change of variable $(\theta,v)=\left (\Theta(R,\theta',v'),V(R,\theta',v')\right)$. In the second integral, we remark that $|\phi'_f\left(\Theta(s)\right)|\leq \|f\|_{L^1}$ and, the support of $\psi$ being compact, $v$ is bounded. Hence, by $\frac{v^2}{2}-m_0\cos\theta=\frac{V^2}{2}-m_0\cos\Theta$, $V(s)$ is bounded. Therefore, by dominated convergence (using that $F'(e_0)\in L^1$), as $R\rightarrow\infty$, we get $$\int_{0}^{2\pi}\int_{\RR} f(\theta,v)\psi(\theta,v)d\theta dv=\int_{-\infty}^{0}\int_{0}^{2\pi}\int_{\RR} e^{\lambda s}F'(e_0) \phi'_f\left(\Theta(s)\right)V(s)\psi\left(\theta,v\right)ds d\theta dv$$ i.e. $$\begin{aligned}
f(\theta,v)&=F'(e_0)\int_{-\infty}^0 e^{\lambda s} \phi'_f\left(\Theta(s)\right)V(s)ds=F'(e_0)\int_{-\infty}^0 e^{\lambda s}\frac{d}{ds}\left( \phi_f\left(\Theta(s)\right)\right)ds\\
&=F'(e_0)\phi_f\left(\theta\right)-F'(e_0)\int_{-\infty}^{0}\lambda e^{\lambda s} \phi_f({\Theta}(s))ds\end{aligned}$$ almost everywhere. Recall that $\phi_f(\theta)=-m\cos \theta$, with $m=\int^{2 \pi}_{0}\rho_f(\theta)\cos\theta d\theta$. Then we can rewrite this expression of $f$ as $$f(\theta,v)=-mF'(e_0)\cos\theta-mF'(e_0)\int_{-\infty}^{0}\lambda e^{\lambda s}\cos {\Theta}(s)ds.$$ Integrating both sides of this equation with respect to $\cos \theta d\theta dv$ we get $$\begin{aligned}
m=\int_{0}^{2 \pi}\rho_f(\theta)\cos\theta d\theta =&-m\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0) \cos^2\theta d\theta dv\\
&+m\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0)\left(\int_{-\infty}^{0}\lambda e^{\lambda s} \cos {\Theta}(s)ds\right)\cos\theta d\theta dv,\end{aligned}$$ i.e. $mG(\lambda)=0$. It is clear that $m\neq 0$, otherwise $f=0$ a.e.. Finally, we get $G(\lambda)=0$.
Reciprocally, assume that $G(\lambda)=0$ for some $\lambda>0$. Let $f$ be given by . We have proved above that this function belongs to $L^1(\TT\times\RR)$. Moreover, since $\Theta(s,-\theta,-v)=\Theta(s,\theta,v)$, we have $\int \rho_f\sin \theta d\theta=0$. Multiplying by $\cos \theta$ and integrating with respect to $\theta$ and $v$, and using that $G(\lambda)=0$, we get $\int \rho_f\cos \theta d\theta=1$, so $f$ is not the zero function and we have $\phi_f(\theta)=-\cos \theta$.
We now check that the function $f$ given by is an eigenfunction of $L$ associated with $\lambda$. From , we get $$f(\Theta(t),V(t))=-F'(e_0)\phi_f\left(\Theta(t)\right)+F'(e_0)\int_{-\infty}^{0}\lambda e^{\lambda s}\phi_f\left(\Theta(s,\Theta(t),V(t))\right)ds.$$ Note that $\Theta(s,\Theta(t),V(t))=\Theta\left(s+t,\theta,v\right)$. Therefore $$\begin{aligned}
f(\Theta(t),V(t))&=-F'(e_0)\phi_f\left(\Theta(t)\right)+ F'(e_0)\int_{-\infty}^{0}\lambda e^{\lambda s}\phi_f\left(\Theta(s+t)\right)ds\\
&=-F'(e_0)\phi_f\left(\Theta(t)\right) + F'(e_0)e^{-\lambda t}\int_{-\infty}^{t}\lambda e^{\lambda s} \phi_f\left(\Theta(s)\right)ds\\
&=F'(e_0)e^{-\lambda t}\int_{-\infty}^{t} e^{\lambda s}\phi_f'\left(\Theta(s)\right)V(s)ds,\end{aligned}$$ where we integrated by parts. Then $$\begin{aligned}
e^{\lambda t}f(\Theta(t),V(t))&=F'(e_0)\int_{-\infty}^{t} e^{\lambda s}\phi_f'\left(\Theta(s)\right)V(s)ds\\
&=\int_{-\infty}^{t} e^{\lambda s}\phi_f'\left(\Theta(s)\right)\pa_vf_0(\Theta(s),V(s))ds.\end{aligned}$$ Differentiating both sides with respect to $t$, we obtain in the sense of distributions that for all $t\in \RR$, $$\begin{aligned}
&&e^{\lambda t}\left(\lambda f(\Theta(t),V(t))+V(t)\pa_\theta f(\Theta(t),V(t))-\phi_0'(\Theta(t))\pa_vf(\Theta(t),V(t))\right)\qquad \\
&&=e^{\lambda t}\phi_f'\left(\Theta(t)\right)\pa_vf_0(\Theta(t),V(t)).\end{aligned}$$ By writing this equation at $t=0$, we get $Lf=\lambda f$: $f$ is an unstable eigenfunction of $L$. This ends the proof of Lemma \[lem1\].
Limiting behavior of $G(\lambda)$ near $\lambda=0$: proof of Lemma \[lim1G\] {#lim0}
----------------------------------------------------------------------------
To study $\lim \limits_{\,\,\lambda\to 0^+}G(\lambda)$ we need to analyze the limit of the function $$\label{littleg}
g_\lambda(\theta,v)=\int_{-\infty}^{0}\lambda e^{\lambda s}\cos{\Theta}(s,\theta,v)ds$$ as $\lambda\to 0$. We provide this result in the next lemma, where we also recall some well-known facts on the solution of the characteristics equations , which are nothing but the pendulum equations.
\[lemperiod\] Let $(\theta,v)\in \TT\times \RR$ and $e=\frac{v^2}{2}-m_0\cos\theta$. Consider the solution $\left(\Theta(s,\theta,v),V(s,\theta,v)\right)$ to the characteristics equations . Then the following holds true.
- If $e>m_0$ then, for all $s\in\RR$, we have $$\Theta(s+T_e)=\Theta(s)+2\pi,\qquad V(s+T_e)=V(s),
\label{qperiod}$$ with $$\label{T1}
T_e=\int_{0}^{2\pi}\frac{d\theta^{'}}{\sqrt{2\left(e+m_0\cos\theta^{'}\right)}}>0.$$
- If $-m_0<e<m_0$ then $\Theta$ and $V$ are periodic with period given by $$\label{T2}
{T_e}=4\int_{0}^{\theta_{m_0}}\frac{d\theta^{'}}{\sqrt{2\left(e+m_0\cos\theta^{'}\right)}}>0,$$ where $\theta_{m_0}=\arccos(-\frac{e}{m_0})$.
- We have $$\label{deflim}
\lim \limits_{\,\,\lambda\to 0^+}g_\lambda(\theta,v)=\left\{ \begin{array}{l}
\ds\frac{1}{T_e}\int_{0}^{2 \pi}\!\!\frac{\cos\theta^{'}}{\sqrt{2(e+m_0 \cos\theta^{'})}}d\theta^{'}\qquad \text{if}\; e>m_0,\\[5mm]
\ds\frac{4}{T_e}\int_{0}^{\theta_{m_0}}\!\!\frac{\cos\theta'}{\sqrt{2(e+m_0 \cos\theta')}}d\theta'\qquad \text{if}\; -m_0\!<\!e\!<\!m_0.
\end{array} \right.$$
*(i)* Let $e>m_0$. Without loss of generality, since $\Theta(s,-\theta,-v)=\Theta(s,\theta,v)$ and $V(s,-\theta,-v)=V(s,\theta,v)$, we can only treat the case $v>0$. As we have $$\frac{V(s)^2}{2}=e+m_0\cos \Theta(s)\geq e-m_0>0,$$ $V(s)$ does not vanish and remains positive. Hence $\Theta(s)$ is the solution of the following autonomous equation $$\label{eqautonome}\dot\Theta(s)=V(s)=\sqrt{2(e+m_0\cos \Theta(s))}$$ and is strictly increasing with $\Theta(s)\to +\infty$ as $s\to+\infty$. Let $T_e$ be the unique time such that $\Theta(T_e)=\theta+2\pi$. By Cauchy-Lipschitz’s theorem, we have $\Theta(s+T_e)-2\pi=\Theta(s)$ and holds. Defining $$P(\tau) =\int_0^\tau \frac{d\theta'}{\sqrt{2(e+m_0\cos\theta')}},$$ the solution of satisfies $P(\Theta(s))-P(\theta)=s$. Therefore, we have $T_e=P(\theta+2\pi)-P(\theta)$, from which we get .
*(ii)* Let $-m_0<e<m_0$. In this case, $\Theta(s)$ will oscillate between the two values $\theta_{m_0}=\arccos(-\frac{e}{m_0})$ and $-\theta_{m_0}$ with a period $T_e$ given by . On the half-periods where $\Theta$ is increasing, we also have . We skip the details of the proof, which is classical.
*(iii)* We remark that $\cos \Theta(s+kT_e)=\cos \Theta(s)$ for all $s\in\RR$ and $k\in \ZZ$. Indeed, by [*(i)*]{}, for $e>m_0$ we have $\Theta(s+kT_e)=\Theta(s)+2\pi k$ and, by [*(ii)*]{}, for $-m_0<e<m_0$ we have $\Theta(s+kT_e)=\Theta(s)$. Hence, we compute from $$\begin{aligned}
g_\lambda(\theta,v)&=\sum\limits_{k=0}^{+\infty}\int_{kT_e}^{(k+1)T_e}\lambda e^{-\lambda s}\cos\Theta(-s)ds \\
&=\sum\limits_{k=0}^{+\infty}\int_{0}^{T_e}\lambda e^{-\lambda s-k\lambda T_e}\cos\Theta(-s)ds \\
&=\left(\sum\limits_{k=0}^{+\infty}e^{-k\lambda T_e}\right)\int_{0}^{T_e}\lambda e^{-\lambda s}\cos\Theta(-s)ds\\
&=\frac{\lambda}{1-e^{-\lambda T_e}}\int_{0}^{T_e}e^{-\lambda s}\cos\Theta(-s)ds.\end{aligned}$$ Therefore, clearly, $$\lim \limits_{\,\,\lambda\to 0^+}g_\lambda(\theta,v)
=\frac{1}{T_e}\int_{0}^{T_e}\cos \Theta(-s)ds
=\frac{1}{T_e}\int_{0}^{T_e}\cos\Theta (s)ds.$$ If $e>m_0$, we perform the change of variable $\theta'=\Theta(s)$ which is strictly increasing from $[0,T_e]$ to $[\theta,\theta+2\pi]$. Using , we obtain $$\lim \limits_{\,\,\lambda\to 0^+}g_\lambda(\theta,v)
=\frac{1}{T_e}\int_{\theta}^{\theta+2\pi}\frac{\cos \theta'}{\sqrt{2(e+m_0\cos \theta')}}d\theta'=\frac{1}{T_e}\int_{0}^{2\pi}\frac{\cos \theta'}{\sqrt{2(e+m_0\cos \theta')}}d\theta'.$$ If $-m_0<e<m_0$, we can always choose a time $t_0$ such that $\Theta(t_0)=-\theta_{m_0}$, $\Theta(t_0+T_e/2)=\theta_{m_0}$, $\Theta(s)$ is strictly increasing on $[t_0,t_0+T_e/2]$ and such that $\Theta(s)=\Theta(2t_0+T_e-s)$ for $s\in [t_0+T_e/2,t_0+T_e]$. We have $$\begin{aligned}
\lim \limits_{\,\,\lambda\to 0^+}g_\lambda(\theta,v)&=\frac{1}{T_e}\int_{t_0}^{t_0+T_e/2}\cos\Theta (s)ds+\frac{1}{T_e}\int_{t_0+T_e/2}^{t_0+T_e}\cos\Theta (s)ds\\
&=\frac{1}{T_e}\int_{t_0}^{t_0+T_e/2}\cos\Theta (s)ds+\frac{1}{T_e}\int_{t_0+T_e/2}^{t_0+T_e}\cos\Theta (2t_0+T_e-s)ds\\
&=\frac{2}{T_e}\int_{t_0}^{t_0+T_e/2}\cos\Theta (s)ds\\
&=\frac{2}{T_e}\int_{-\theta_{m_0}}^{\theta_{m_0}}\frac{\cos \theta'}{\sqrt{2(e+m_0\cos \theta')}}d\theta'=\frac{4}{T_e}\int_{0}^{\theta_{m_0}}\frac{\cos \theta'}{\sqrt{2(e+m_0\cos \theta')}}d\theta',\end{aligned}$$ where, on the time interval $[t_0,t_0+T_e/2]$, we performed the change of variable $\theta'=\Theta(s)$.
Now we come back to the definition of $G_\lambda$, which reads $$\begin{aligned}
G(\lambda) = &1+\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'\left(\frac{v^2}{2}-m_0\cos\theta\right) \cos^2\theta d\theta dv\nonumber\\
&\quad -\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'\left(\frac{v^2}{2}-m_0\cos\theta\right)g_\lambda(\theta,v) \cos \theta d\theta dv.\label{Glambda2}\end{aligned}$$ We remark that $|g_\lambda(\theta,v)|\leq 1$ and recall that the function $F'\left(\frac{v^2}{2}-m_0\cos\theta\right)$ belongs to $L^1(\TT\times\RR)$. Therefore, we can pass to the limit in the second integral by dominated convergence and deduce from Lemma \[lemperiod\] [*(iii)*]{} (note that the set $\{(\theta,v):\,e_0(\theta,v)\leq -m_0\}$ is of measure zero) that
$$\begin{aligned}
\lim \limits_{\,\,\lambda\to 0^+}G(\lambda) &= 1+\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0) \cos^2\theta d\theta dv\\
&\quad \quad -\iint_{e_0(\theta,v)>m_0}F'(e_0)\left(\frac{1}{T_e}\int_{0}^{2 \pi}\!\!\frac{\cos\theta^{'}}{\sqrt{2(e_0+m_0 \cos\theta^{'})}}d\theta^{'}\right)\cos \theta d\theta dv\\
&\quad \quad -\iint_{-m_0<e_0(\theta,v)<m_0}F'(e_0)\left(\frac{4}{T_e}\int_{0}^{\theta_{m_0}}\!\!\frac{\cos\theta^{'}}{\sqrt{2(e_0+m_0 \cos\theta^{'})}}d\theta^{'}\right)\cos \theta d\theta dv\\
= &1+\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0)\cos^2\theta d\theta dv\\
&\quad -\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}} F'(e_0)\left(\frac{\ds \int_{\mathcal D_{e_0}}\cos \theta'(e_0+m_0\cos \theta')^{-1/2}d\theta'}{\ds\int_{\mathcal D_{e_0}}(e_0+m_0\cos \theta')^{-1/2}d\theta'} \right)\cos \theta d\theta dv\\
=& 1+\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0) \cos(\theta)^2dv d\theta-\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0) \left(\Pi_{m_0}\left(\cos \theta\right)\right)^2 d\theta dv,\end{aligned}$$
with $$\label{projec}
(\Pi_{m_0} h)(e)=\frac{\ds\int_{\mathcal D_e}(e+m_0\cos \theta)^{-1/2}h(\theta)d\theta}{\ds\int_{\mathcal D_e}(e+m_0\cos \theta)^{-1/2}d\theta},$$ for all function $h(\theta)$ and $${\mathcal D_e}=\left\{\theta'\in \TT\,:\,\,m_0\cos \theta'>-e\right\}.$$ Here the operator $\Pi_{m_0}$ is a variant of the operator $\Pi$ given by (3.8) in [@ORB], this operator should be understood as the “projector”onto the functions which depend only on the microscopic energy $e_0(\theta,v)$. A projector of this type is also mentioned in the work by Guo and Lin [@LIN2].
Now we remark that straightforward calculations give
$$\begin{aligned}
1-\kappa_0&=&1+\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0)\left(\cos^2 \theta -2\cos\theta\Pi_{m_0}\left(\cos \theta\right)+\left(\Pi_{m_0}\left(\cos \theta\right)\right)^2\right)\,d\theta dv,\nonumber\\
&=&1+\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0)\cos^2\theta d\theta dv-\iint F'(e_0) \left(\Pi_{m_0}\left(\cos \theta\right)\right)^2\,d\theta dv,\nonumber\end{aligned}$$
where $$\begin{aligned}
-\kappa_0:&=\int_{0}^{2 \pi}\int_{{{\mathbb{R}}}}F'(e_0)\left(\frac{\ds \int_{\mathcal D_{e}}(\cos \theta-\cos \theta')(e_0+m_0\cos \theta')^{-1/2}d\theta'}{\ds\int_{\mathcal D_{e}}(e_0+m_0\cos \theta')^{-1/2}d\theta'} \right)^2d\theta dv.
$$ This calculation uses that $\Pi_{m_0}$ is a projector. We finally get and the proof of Lemma \[lim1G\] is complete.
Limiting behavior of $G(\lambda)$ as $\lambda\to\infty$: proof of Lemma \[lim2G\] {#liminfty}
---------------------------------------------------------------------------------
In this subsection, we prove Lemma \[lim2G\]. An integration by parts in yields $$g_\lambda(\theta,v)=\cos\theta+\int_{-\infty}^{0}e^{\lambda s}V(s,\theta,v)\sin{\Theta}(s,\theta,v)ds.$$ The velocity can be bounded independently of $s$ thanks to the conservation of the energy, $$|V(s)|=\left(v^2+2m_0\cos \theta -2m_0\cos \Theta(s)\right)^{1/2}\leq \left(v^2+4m_0\right)^{1/2}.$$ Thus $$\left|\int_{-\infty}^{0}e^{\lambda s}\sin{\Theta}(s,\theta,v)V(s,\theta,v)ds\right|\leq \left(v^2+4m_0\right)^{1/2}\int_{-\infty}^{0}e^{\lambda s}ds= \frac{\left(v^2+4m_0\right)^{1/2}}{\lambda}$$ and, for all $(\theta,v)$, $$\lim_{\lambda\to +\infty}g_\lambda(\theta,v)=\cos \theta.$$ Using again that $|g_\lambda(\theta,v)|\leq 1$ and that $F'(e_0(\theta,v))$ belongs to $L^1$, we deduce directly from and from dominated convergence. The proof of Lemma \[lim2G\] is complete.
A nonlinear instability result: proof of Theorem \[mainthm2\] {#section3}
=============================================================
We start by an analysis of the linearized HMF operator $L$ around the inhomogeneous equilibrium state $f_0$, where $L$ is given by . We write $$\label{defL2}
L=L_0 +K,$$ where $$\label{defL0}
L_0f= -v\partial_{\theta} f-E_{f_0}\partial_v f, \qquad Kf=-E_f\partial_{v} f_0, \qquad E_f = -\pa_\theta \phi_f.$$
Estimates on the semigroup $e^{tL}$ {#subsecteL}
-----------------------------------
Let us state some useful properties of the operator $L_0$ given by . Since $\phi_0(\theta)=-m_0 \cos\theta$ is smooth, the characteristics equations admit a unique solution $\Theta(s,\theta, v),V(s, \theta, v)$, which is globally defined and $\mathcal C^\infty$ in the variables $(s,\theta,v)$. Moreover, this solution has bounded derivatives with respect to $\theta$ and $v$, locally in time. Let $k\in \NN$. For any $f$ in the Sobolev space $W^{k,1}(\TT,\RR)$, the function $$\label{solutionL0}
e^{tL_0} f (t,\theta,v):= f\left( \Theta(-t,\theta,v), V(-t,\theta,v) \right), \quad \forall t\geq 0,$$ belongs to $\mathcal C^0(\RR, W^{k,1}(\TT,\RR))$ and is clearly a solution to $\pa_tg= L_0g$ with initial data $f$. This means that the semigroup $e^{tL_0}$ generated by the operator $L_0$ is strongly continuous on $W^{k,1}(\TT,\RR)$.
Our aim is to apply the abstract results in [@shizuta] concerning perturbation theory of linear operators. Hence, we need to prove the following estimate. For all $\beta>0$, there exists a positive constant $M_\beta$ such that $$\label{Lestimate}\| e^{tL_0} f\|_{W^{k,1}} \leq M_\beta e^{t \beta }\| f\|_{W^{k,1}}\qquad \forall f\in W^{k,1}(\TT,\RR), \quad \forall t\geq 0,$$ where $M_{\beta}$ depends on $k$. In fact a stronger estimate will be proved but for a subclass of functions $f$. From the assumptions of Theorem \[mainthm2\] on $F$, there exists $e_*<m_0$ such that support of $F$ is $(-\infty,e_*]$. This means that the support of $f_0$ is contained in $\overline{\Omega_0}$, where $\Omega_0$ is the smooth open set $$\Omega_0= \left\{(\theta, v): \ \frac{v^2}{2} -m_0\cos\theta < e_*\right\}.$$ We then introduce the functional space $$\mathscr E_k = \left\{ f \in W^{k,1}(\TT,\RR): \ \ \mbox{Supp}(f) \subset \overline{\Omega_0} \right\},$$ and claim that for all $f\in \mathscr E_k$, we have $e^{tL_0} f\in \mathscr E_k$ and $$\label{Lestimatebis} \| e^{tL_0} f\|_{W^{k,1}} \leq M_k \| f\|_{W^{k,1}}, \qquad \forall f\in \mathscr E_k,\quad \forall t\geq 0,$$ $M_k$ being a positive constant.
Assume for the moment that estimate holds true. From the assumptions of Theorem \[mainthm2\], one deduces that $\pa_vf_0 \in \mathscr E_k $. It is then easy to check that $e^{tL_0} K$ is a compact operator on $\mathscr E_k$, for all $t\in \RR$, and the map $t\mapsto e^{tL_0} K \in \mathscr{L}(\mathscr E_k)$ is continuous on $\RR$. Hence, $K$ is $L_0$-smoothing in the sense of [@shizuta] (page 707). Assumptions of Theorem 1.1 in [@shizuta] are therefore satisfied, which implies that $L$ generates a strongly continuous semigroup $e^{tL}$. Now, from Theorem 1.2 in [@shizuta], for all $\beta >0$, any point of the spectrum $\sigma(L)$ lying in the half plane $\operatorname{Re}z >\beta$ is an isolated eigenvalue with finite algebraic multiplicity. Furthermore, the set $\sigma (L) \cap \{\operatorname{Re}z >\beta\}$ is finite.
The assumptions of Theorem \[mainthm2\] clearly imply those of Theorem \[mainthm1\]. Hence $L$ admits at least one eigenvalue $\lambda\in \RR_+^*$ associated with an eigenfunction $\widetilde f\in L^1(\TT\times\RR)$. We claim that, in fact, $\widetilde f\in \mathscr E_k$ which will be proved below. This means that the set of eigenvalues of $L$ on $\mathscr E_k$ with positive real part is not empty, and we therefore can choose an eigenvalue $\gamma$ with positive maximal real part. Finally, we apply Theorem 1.3 in [@shizuta] and get that, for all $\beta>\operatorname{Re}\gamma$, there exists a positive constant $M_{\beta,k}$ such that $$\label{Lestimate}\| e^{tL} f\|_{W^{k,1}} \leq M_{\beta,k} \ e^{t \beta }\| f\|_{W^{k,1}}\qquad \forall f\in \mathscr E_k, \quad \forall t\geq 0.$$
[*Proof of and of the claim $\widetilde f\in \mathscr E_k$*]{}. Let $f\in \mathscr E_k$. From , we clearly have $$\|e^{tL_0} f\|_{L^1} = \| f\|_{L^1}, \quad \forall f \in L^1, \quad \forall t\geq 0.$$ Moreover, we know from the analysis of the characteristics problem performed in Section \[section2\], that by conservation of the energy, for all $(\theta, v) \in \Omega_0$, we have $(\Theta(t, \theta,v), V(t, \theta,v)) \in \Omega_0$. Thus $$\mbox{Supp}(e^{tL_0}f) \subset \overline{\Omega_0}.$$ Let $k\geq 1$. By , to get an estimate of $e^{tL_0} f$ in $L^\infty(\RR_+, W^{k,1}(\TT,\RR))$, it is sufficient to estimate $\Theta$ and $V$ in $L^\infty(\RR,W^{k,\infty}(\TT,\RR))$ for $(\theta,v) \in \Omega_0$. Since the field $E_{f_0}(\theta)= -\pa_\theta \phi_0(\theta)$ and all its derivatives are bounded, one can apply Gronwall Lemma to get $$|\pa^j_\theta \pa^\ell_v \Theta |+ |\pa^j_\theta \pa^\ell_v V| \leq C_k e^{C_k t }, \quad \forall t\geq 0, \quad \forall (\theta,v)\in \Omega_0, \quad \mbox{for} \ j+\ell \leq k.$$ Now we observe that for $(\theta, v) \in \Omega_0$, we have $e=v^2/2 - m_0 \cos\theta <m_0$ and by Lemma \[lemperiod\], the solution $(\Theta,V)$ and its partial derivatives with respect to $\theta$ and $v$ are periodic in time, with a period $T_e$ given by . Hence $$|\pa^j_\theta \pa^\ell_v \Theta |+ |\pa^j_\theta \pa^\ell_v V| \leq C_k e^{C_k T_e}, \quad \forall t\geq 0, \quad \forall (\theta,v)\in \Omega_0, \quad \mbox{for} \ j+\ell \leq k.$$ By Lemma 2.2 [*(ii)*]{} in [@llmehats] the function $e\mapsto T_e$ is increasing on $(-m_0,m_0)$. Thus $T_e\leq T_{e_*}$ for all $(\theta, v)\in \Omega_0$. We conclude that $\Theta$ and $V$ are bounded in $L^\infty(\RR,W^{k,\infty}(\Omega_0))$, which yields the estimate .
Let us finally prove that $\widetilde f\in \mathscr E_k$. By Lemma \[lem1\], the function $\widetilde f$ is given by $$\widetilde f(\theta,v)=-F'(e_0(\theta,v))\cos\theta+F'(e_0(\theta,v))\int_{-\infty}^{0}\lambda e^{\lambda s}\cos\left({\Theta}(s,\theta,v)\right)ds.$$ Hence, the support of $F'(e_0(\theta,v))$ being in $\overline{\Omega_0}$, the support on $\widetilde f$ will also be contained in $\overline{\Omega_0}$. Moreover, by using that $\Theta$ is bounded in $L^\infty(\RR,W^{k,\infty}(\Omega_0))$, we obtain that, for some $C_k>0$, we have $$\forall j+\ell \leq k,\quad \forall (\theta,v)\in \TT\times \RR,\quad\left|\pa^j_\theta \pa^\ell_v\int_{-\infty}^{0}\lambda e^{\lambda s}\cos\left({\Theta}(s,\theta,v)\right)ds\right|\leq C_k.$$ This is sufficient to deduce from $F\in \mathcal C^\infty$ that $\widetilde f\in \mathscr E_k$.
An iterative scheme
-------------------
In this part, we prove Theorem \[mainthm2\] by following the strategy developed by Grenier in [@grenier], which has been also used in [@Han-Hauray; @Han-Nguyen] to analyse instabilities for homogeneous steady states of Vlasov-Poisson models. Let $N\geq 1$ be an integer to be fixed later. According to the previous subsection, we can consider an eigenvalue $\gamma$ of $L$ on $\mathscr E_N$ with maximal real part, $\operatorname{Re}\gamma >0$. Let $g\in \mathscr E_N$ be an associated eigenfunction. With no loss of generality, we may assume that $\|\operatorname{Re}g\|_{L^1}=1$. Let $$\label{f1}
f_1(t,\theta,v)= \operatorname{Re}\left( e^{\gamma t} g(\theta,v)\right)\chi_\delta(e_0(\theta,v)),$$ with $e_0(\theta,v)=\frac{v^2}{2}+\phi_0(\theta)$ and where $0\leq \chi_\delta(e) \leq 1$ is a smooth real-valued truncation function to be defined further, in order to ensure the positivity of $f_0+\delta f_1(0)$. Note that $f_1$ is almost a growing mode solution to the linearized HMF model since we have $$(\partial_t-L)f_1=\operatorname{Re}(e^{\gamma t}\widetilde R_\delta),$$ where $$\label{Rdelta}
\widetilde R_\delta=\left(-E_{(1-\chi_\delta)g}+(1-\chi_\delta)E_g\right)\pa_vf_0$$ will be small. We now construct an approximate solution $f^{N}_{app}$ to the HMF model (\[sis1\]) of the form $$f^{N}_{app}=f_0+\sum_{k=1}^{N}\delta^{k}f_k,$$ for sufficiently small $\delta>0$, in which $f_k$ ($k\geq2$) solves inductively the linear problem $$\label{eqfk}
(\partial_t-L)f_k+\sum_ {j=1}^{k-1}E_{f_j}\partial_v f_{k-j}=0$$ with $f_k(0)=0$. Then $f^{N}_{app}$ approximately solves the HMF model (\[sis1\]) in the sense that $$\partial_t f^{N}_{app}+v\partial_{\theta} f^{N}_{app}-\partial_{\theta} \phi_{f^{N}_{app}}\partial_v f^{N}_{app}=R_{N}+\delta \operatorname{Re}(e^{\gamma t}\widetilde R),$$ where the remainder term $R_{N}$ is given by $$R_{N}=\sum_{1\leq j,\ell \leq N;j+\ell\geq N+1}\delta^{j+\ell}E_{f_j}\partial_v f_{\ell}.$$
[*Step 1. Estimate of $f_k$.*]{} We claim that $f_k \in \mathscr E_{N-k+1}$ and, for all $1\leq k\leq N$, $$\label{estimatefk}
\|f_k\|_{W^{N-k+1,1}} \leq C_k e^{kt\operatorname{Re}\gamma}.$$ We proceed by induction. From , this estimate is a consequence, for $k=1$, of $$\label{condchi0}
\|g\chi_\delta\|_{W^{N,1}}\leq C_1,$$ which is proved below in Step 5. Let $k\geq 2$. We have $$f_k(t)=-\int_{0}^{t}e^{L(t-s)}\sum_ {j=1}^{k-1}E_{f_j}(s)\partial_v f_{k-j}(s)ds.$$ Therefore, for $\operatorname{Re}\gamma < \beta < 2 \operatorname{Re}\gamma$, $$\begin{aligned}
\|f_k\|_{W^{N-k+1,1}} & \leq \sum_ {j=1}^{k-1} \int_0^t \left\|e^{L(t-s)}\left( E_{f_j}(s)\partial_v f_{k-j}(s)\right)\right\|_{W^{N-k+1,1}} ds \\
&\leq M_{\beta, N-k+1} \sum_ {j=1}^{k-1} \int_0^t e^{\beta(t-s)} \left\|E_{f_j}(s)\right\|_{W^{N-k+1,\infty}} \left\|\partial_v f_{k-j}(s)\right\|_{W^{N-k+1,1}} ds \\
&\leq M_{\beta, N-k+1} \sum_ {j=1}^{k-1} \int_0^t e^{\beta(t-s)} \left\|f_j(s)\right\|_{L^{1}} \left\|f_{k-j}(s)\right\|_{W^{N-k+2,1}} ds\\
&\qquad \mbox{since }k-j\leq k-1,\\
&\leq M_{\beta, N-k+1} \left(\sum_ {j=1}^{k-1} C_j C_{k-j} \right)\int_0^t e^{\beta(t-s)} e^{ks \operatorname{Re}\gamma } ds\\
&\leq \frac{M_{\beta, N-k+1}}{k\operatorname{Re}\gamma -\beta} \left(\sum_ {j=1}^{k-1} C_j C_{k-j} \right) e^{k t\operatorname{Re}\gamma},\end{aligned}$$ where we used and the recursive assumption. This ends the proof of .
[*Step 2. Estimates of $f_{app}^N-f_0$ and $R_N$.*]{} The parameter $\delta$ and the time $t$ will be such that $$\label{cond0}\delta e^{t\operatorname{Re}\gamma}\leq \min\left(\frac 12,\frac{1}{2K_N}\right),\qquad K_N=\max_{1\leq k\leq N}C_k.$$ Hence, from we obtain $$\|f_{app}^N-f_0\|_{W^{1,1}}\leq \sum_{k=1}^N\delta^k C_k e^{kt\operatorname{Re}\gamma}\leq K_N \frac{\delta e^{t\operatorname{Re}\gamma}}{1-\delta e^{t\operatorname{Re}\gamma}}\leq 1$$ and $$\|R_N\|_{L^1}\leq \sum_{k=N+1}^{+\infty}\delta^k e^{kt\operatorname{Re}\gamma}\sum_{1\leq j,\ell \leq N;j+\ell =k}C_jC_{\ell}\leq
\widetilde C_N \left(\delta e^{t\operatorname{Re}\gamma}\right)^{N+1}.$$
[*Step 3. Estimate of $f-f_{app}^N$.*]{} Let $f(t)$ be the solution of with initial data $f_0+\delta \operatorname{Re}g\chi_\delta$ and let $h=f-f_{app}^N$. Note that the positivity of $f(t)$ is ensured by $f_0+\delta \operatorname{Re}g\chi_\delta\geq 0$ and that we have $$\|f(0)-f_0\|_{L^1}\leq \delta.$$ The function $h$ satisfies the following equation $$\partial_t h+v\partial_{\theta} h+ E_{f}\partial_v h = \left(E_{f_{app}^N} - E_{f}\right) \pa_v f_{app}^N- R_{N}-\delta \operatorname{Re}(e^{\gamma t}\widetilde R_\delta)$$ with $h(0)=0$. To get a $L^1$-estimate of $h$, we multiply this equation by $\mbox{sign}(h)$ and integrate in $(\theta, v)$. We get $$\begin{aligned}
\frac{d}{dt} \| h\|_{L^1}& \leq \left\|E_{f_{app}^N} - E_{f}\right\|_{L^\infty} \left\|\pa_v f_{app}^N\right\|_{L^1}+\|R_{N}\|_{L^1}+\delta e^{t\operatorname{Re}\gamma}\|\widetilde R_\delta\|_{L^1}\\
&\leq \| h\|_{L^1} \left\|\pa_v f_{app}^N\right\|_{L^1}+\|R_{N}\|_{L^1}+\delta e^{t\operatorname{Re}\gamma}\|\widetilde R_\delta\|_{L^1}.\end{aligned}$$ From Step 2 we have $\left\|\pa_v f_{app}^N\right\|_{L^1} \leq \left\|\pa_v f_0\right\|_{L^1}+1$, which implies that $$\| h(t)\|_{L^1}\leq \int_0^t e^{(t-s)(\left\|\pa_v f_0\right\|_{L^1}+1)} \left(\|R_{N}(s)\|_{L^1}+\delta e^{s\operatorname{Re}\gamma}\|\widetilde R_\delta\|_{L^1}\right) ds.$$ Again from Step 2, we then get $$\| h(t)\|_{L^1}\leq \int_0^t e^{(t-s)(\left\|\pa_v f_0\right\|_{L^1}+1)} \left( \widetilde C_N\left(\delta e^{s\operatorname{Re}\gamma}\right)^{N+1}+\delta e^{s\operatorname{Re}\gamma}\|\widetilde R_\delta\|_{L^1}\right) ds$$ We now fix $N$ as follows (with the notation $\lfloor \cdot \rfloor$ for the integer function) $$N:=\left\lfloor\frac{\left\|\pa_v f_0\right\|_{L^1}+1}{\operatorname{Re}\gamma}\right\rfloor+1\geq 1$$ and claim that $\chi$ may be chosen such that $$\label{condchi}
\|\widetilde R_\delta\|_{L^1}\leq \left(\delta e^{s\operatorname{Re}\gamma}\right)^{N},$$ see Step 5 for the proof. This yields $$\label{esti1} \| f-f_{app}^N\|_{L^1}(t)\leq \widecheck C_N\left(\delta e^{t\operatorname{Re}\gamma}\right)^{N+1}$$ with $\widecheck C_N=\frac{1+\widetilde C_N}{3\operatorname{Re}\gamma}$.
[*Step 4. End of the proof.*]{} Since $\operatorname{Re}g$ is not zero, we can choose a real valued function $\varphi(\theta,v)$ in $L^\infty$ such that $\|\varphi\|_{L^\infty}$=1 and $$\operatorname{Re}z_g>0 \quad \mbox{with}\quad z_g=\int_0^{2\pi}\int_\RR g \varphi d\theta dv.$$ Denoting $$z_{g,\delta}=\int_0^{2\pi}\int_\RR g \chi_\delta \varphi d\theta dv,$$ we have $$\begin{aligned}
\iint f_1\varphi d\theta dv&=e^{t\operatorname{Re}\gamma}\operatorname{Re}\left(e^{it\operatorname{Im}\gamma}z_{g,\delta}\right)\\
&\geq e^{t\operatorname{Re}\gamma}\operatorname{Re}\left(e^{it\operatorname{Im}\gamma}z_g\right)-e^{t\operatorname{Re}\gamma} |z_g-z_{g,\delta}|\\
&\geq e^{t\operatorname{Re}\gamma}\operatorname{Re}\left(e^{it\operatorname{Im}\gamma}z_g\right)-e^{t\operatorname{Re}\gamma} \|g(1-\chi_\delta)\|_{L^1}\end{aligned}$$ We claim that $$\label{condchi2}
\lim_{\delta\to 0}\|(1-\chi_\delta)g\|_{L^1}=0,$$ which again will be proved in Step 5. In order to end the proof of Theorem \[mainthm2\], we estimate from below, using and , $$\begin{aligned}
\|f-f_0\|_{L^1}& \geq \iint (f-f_0)\varphi d\theta dv=\iint (f_{app}^N-f_0)\varphi d\theta dv+\iint (f-f_{app}^N)\varphi d\theta dv\\
&\geq \delta \iint f_1\varphi d\theta dv-\sum_{k=2}^N\delta^k\|f_k\|_{L^1} - \widecheck C_N\left(\delta e^{t\operatorname{Re}\gamma}\right)^{N+1}\\
&\geq \delta \iint f_1\varphi d\theta dv-\sum_{k=2}^NC_k\left(\delta e^{t\operatorname{Re}\gamma}\right)^k - \widecheck C_N\left(\delta e^{t\operatorname{Re}\gamma}\right)^{N+1}\\
&\geq \delta \iint f_1\varphi d\theta dv-2K_N\left(\delta e^{t\operatorname{Re}\gamma}\right)^2- \widecheck C_N\left(\delta e^{t\operatorname{Re}\gamma}\right)^{N+1}\\
&\geq\delta e^{t\operatorname{Re}\gamma}\left(\operatorname{Re}\left(e^{it\operatorname{Im}\gamma}z_g\right)- \|(1-\chi_\delta)g\|_{L^1}-2K_N\delta e^{t\operatorname{Re}\gamma} - \widecheck C_N\left(\delta e^{t\operatorname{Re}\gamma}\right)^{N}\right)\end{aligned}$$ Assume for a while that $$\label{condition}\operatorname{Re}\left(e^{it\operatorname{Im}\gamma}z_g\right)\geq \frac{\operatorname{Re}z_g}{2}.$$ We have $$\|f-f_0\|_{L^1}\geq \frac{\delta e^{t\operatorname{Re}\gamma}\operatorname{Re}z_g}{2}\left(1-\frac{2\|(1-\chi_\delta)g\|_{L^1}}{\operatorname{Re}z_g}-\frac{4K_N\delta e^{t\operatorname{Re}\gamma}}{\operatorname{Re}z_g} - \frac{2\widecheck C_N\left(\delta e^{t\operatorname{Re}\gamma}\right)^{N}}{\operatorname{Re}z_g}\right)$$ Let $\delta_0>0$ be such that $$\frac{32K_N}{(\operatorname{Re}z_g)^2}\delta_0 + \frac{2 \widecheck C_N8^N}{(\operatorname{Re}z_g)^{N+1}}\delta_0^{N}\leq \frac14\quad \mbox{and}\quad \frac{8\delta_0 }{\operatorname{Re}z_g}\leq \min \left(\frac 12,\frac{1}{2K_N}\right)$$ (note that $N\geq 1$) and consider times $t$ such that $$\label{tdelta}\frac{4\delta_0 }{\operatorname{Re}z_g}\leq \delta e^{t\operatorname{Re}\gamma}\leq \frac{8\delta_0 }{\operatorname{Re}z_g}.$$ Owing to , we also choose $\delta$ small enough such that $$\frac{2\|(1-\chi_\delta)g\|_{L^1}}{\operatorname{Re}z_g}\leq\frac{1}{4}.$$ We conclude from these inequalities that $$\|f-f_0\|_{L^1}\geq \delta_0$$ and that is satisfied.
To end the proof, it remains to fix the time $t_\delta$ and to choose the truncation function $\chi_\delta$. Let us show that, for $\delta$ small enough, there exists a time $t_\delta$ satisfying both and . If $\operatorname{Im}\gamma =0$, then is clearly satisfied since $\operatorname{Re}z_g>0$: a suitable $t_\delta$ is then $$t_\delta=\frac{1}{\operatorname{Re}\gamma}\log \left(\frac{6\delta_0}{\delta \operatorname{Re}z_g}\right).$$ Assume now that $\operatorname{Im}\gamma\neq 0$. For $\delta$ small enough, the size of the interval of times $t$ satisfying becomes larger than $\frac{2\pi}{|\operatorname{Im}\gamma|}$. This means that it is possible to find a time $t_\delta$ in this interval satisfying .
[*Step 5. Choice of $\chi_\delta$.*]{} For all $\delta>0$, we have to fix the function $\chi_\delta\in \mathcal C^\infty(\RR)$ such that , , are satisfied and such that $f_0+\delta f_1(0)\geq 0$. First of all, proceeding as in the proof of Lemma \[lem1\], we obtain that $g$ takes the form $$\label{eqqg}
g(\theta,v)=-mF'(e_0)\cos\theta-mF'(e_0)\int_{-\infty}^{0}\gamma e^{\gamma s}\cos {\Theta}(s)ds,$$ with $m=\int^{2 \pi}_{0}\rho_f(\theta)\cos\theta d\theta$. Hence, $$\label{majg}
|g(\theta,v)|\leq |m|\left(1+\frac{|\gamma|}{\operatorname{Re}\gamma}\right)|F'(e_0)|.$$ The assumptions on $F$ and $F'$ in Theorem \[mainthm2\] imply that $$\label{alpha}\forall e<e_*,\qquad |F'(e)|\leq C(e_*-e)^{-\alpha}F(e)$$ with $\alpha\geq 1$. Since $F(e)>0$ for $e<e_*$, the local assumption becomes global. Let $\chi$ be a $\mathcal C^\infty$ function such that $0\leq \chi\leq 1$ and $$\left\{\begin{array}{ll}
\ds \chi(t)=0\quad &\mbox{for }t\leq 0,\\
\ds \chi(t)\leq 2t^\alpha\quad &\mbox{for }t\geq 0,\\
\ds \chi(t)=1\quad &\mbox{for }t\geq 1
\end{array}\right.$$ and let $$\label{chidelta}\chi_\delta(e)=\chi\left(\frac{e_*-e}{\delta^{1/{(2\alpha)}}}\right).$$ From $$\| (1-\chi_\delta)g\|_{L^1}\leq \|g\un_{e_*-\delta^{1/(2\alpha)}< e_0(\theta,v)<e_*}\|_{L^1}$$ and dominated convergence, we clearly have . By and , we have, for all $(\theta,v)\in \TT\times \RR$, $$\delta\left|\operatorname{Re}g(\theta,v)\chi_\delta(e_0(\theta,v))\right|\leq \delta C|e_*-e_0|^{-\alpha}F(e_0)\frac{|e_*-e_0|^{\alpha}}{\delta^{1/2}}=C\delta^{1/2}f_0(\theta,v),$$ so for $\delta$ small enough, we have $f_0+\delta f_1(0)\geq 0$.
By differentiating and using that $\Theta$ is bounded in $L^\infty(\RR,W^{k,\infty}(\Omega_0))$ (see Subsection \[subsecteL\]), we get $$\forall j+\ell \leq N,\qquad |\pa^j_\theta\pa^\ell_v g(\theta,v)|\leq C\max_{k\leq N+1}F^{(k)}(e_0)\leq C|e_*-e_0|^{N-1},$$ where we used Taylor formulas and the fact that $F\in \mathcal C^\infty$ with $F(e)=0$ for $e\geq e_*$. Besides, from , we obtain (if $\delta\leq 1$) $$\forall \ 1\leq j+\ell \leq N,\qquad |\pa^j_\theta\pa^\ell_v \chi_\delta(e_0(\theta,v))|\leq C\delta^{-N/(2\alpha)}\un_{e_*-\delta^{1/(2\alpha)}< e_0(\theta,v)< e_*}.$$ Therefore $$\begin{aligned}
\|g\chi_\delta\|_{W^{N,1}}&\leq \|g\|_{L^1}+C\delta^{-N/(2\alpha)}\int_0^{2\pi}\int_\RR|e_*-e_0(\theta,v)|^{N-1}\un_{e_*-\delta^{1/(2\alpha)}< e_0(\theta,v)< e_*}d\theta dv\\
&\leq \|g\|_{L^1}+C\delta^{-N/(2\alpha)}\int_{e_*-\delta^{1/(2\alpha)}}^{e_*}(e_*-e)^{N-1} \left(4\int_{0}^{\theta_{m_0}}\frac{d\theta}{\sqrt{2\left(e+m_0\cos\theta\right)}}\right)de\\
&\quad= \|g\|_{L^1}+C\delta^{-N/(2\alpha)}\int_{e_*-\delta^{1/(2\alpha)}}^{e_*}(e_*-e)^{N-1} \,T_e\,de,\end{aligned}$$ where $\theta_{m_0}=\arccos(-\frac{e}{m_0})$ and $T_e$ is given by . Now we recall that for $e\leq e_*$ we have $T_e\leq T_{e_*}$. This yields $$\|g\chi_\delta\|_{W^{N,1}}\leq \|g\|_{L^1}+CT_{e_*}\delta^{-N/(2\alpha)}\int_{e_*-\delta^{1/(2\alpha)}}^{e_*}(e_*-e)^{N-1}\,de=\|g\|_{L^1}+\frac{CT_{e_*}}{N}.$$ We have proved .
By , we have $$\begin{aligned}
\|\widetilde R_\delta\|_{L^1}&\leq C\left(\| (1-\chi_\delta)g\|_{L^1}+\|(1-\chi_\delta)\pa_vf_0\|_{L^1}\right)\\
&\leq C\left(\|g\un_{e_*-\delta^{1/(2\alpha)}< e_0(\theta,v)<e_*}\|_{L^1}+\|\pa_vf_0\un_{e_*-\delta^{1/(2\alpha)}<e_0(\theta,v)< e_*}\|_{L^1}\right),\end{aligned}$$ so by dominated convergence, $$\lim_{\delta\to 0}\|\widetilde R_\delta\|_{L^1}=0.$$ We now choose $\delta$ small enough such that $$\|R_\delta\|_{L^1}\leq \left(\frac{4\delta_0 }{\operatorname{Re}z_g}\right)^N.$$ From , we obtain , which ends the proof of Theorem \[mainthm2\].
Appendix. Existence of unstable steady states {#A1}
=============================================
In this section, we prove that the set of steady states satisfying the assumptions of Theorems \[mainthm1\] and \[mainthm2\] is not empty. More precisely, we prove the following
\[lemapp\] Let $m>0$. There exist $m>0$, $e_*<m$ and there exists a nonincreasing function $F$, $\mathcal {C} ^{\infty}$ on $\RR$, such that $F(e)>0$ for $e<e_*$, $F(e)=0$ for $e\geq e_*$ and $|F'(e)|\leq C|e_*-e|^{-\alpha}F(e)$ in the neighborhood of $e_*$, for some $\alpha\geq 1$, and such that the function $f(\theta,v)=F(\frac{v^2}{2}-m\cos \theta)$ is a steady state solution to the HMF model and such that $\kappa(m,F)>1$, where $\kappa(m,F)$ is given by $$\kappa(m,F)=\int^{2 \pi}_{0}\!\!\int^{+\infty}_{-\infty} \left|F'\!\left(e(\theta,v)\right)\right|\left(\frac{\ds \int_{\mathcal D_{e(\theta,v)}}(\cos \theta-\cos \theta')(e(\theta,v)+m\cos \theta')^{-1/2}d\theta'}{\ds\int_{\mathcal D_{e(\theta,v)}}(e(\theta,v)+m\cos \theta')^{-1/2}d\theta'} \right)^2d\theta dv,$$ with $$e(\theta,v)=\frac{v^2}{2}-m\cos \theta,\qquad {\mathcal D_e}=\left\{\theta'\in \TT\,:\,\,m\cos\theta'>-e\right\}.$$
Let $m>0$ and $F$ a nonincreasing $\mathcal {C} ^{\infty}$ function on $\RR$ supported in $(-\infty,m)$, which is not identically zero on $(-m,m)$. We first observe that $f(\theta,v)=F(\frac{v^2}{2}-m\cos \theta)$ is a steady state solution to the HMF model if and only if $m$ and $F$ satisfy $\gamma(m,F)=m$ with $$\gamma(m,F):=\int_0^{2\pi}\int_\RR F\left(\frac{v^2}{2}-m\cos\theta\right)\cos \theta d\theta dv>0.$$ By using the linearity of $\gamma$ in $F$ we deduce that $\frac{m}{\gamma(m,F)}F(\frac{v^2}{2}-m\cos \theta)$ is a steady state.
We proceed by a contradiction argument. Assume that $$\kappa\left(m,\frac{m}{\gamma(m,F)}F\right)\leq 1$$ for all $m>0$ and all nonincreasing $\mathcal {C} ^{\infty}$ function $F$ supported in $(-\infty,m)$ such that, denoting by $(-\infty,e_*]$ the support of $F$, we have $|F'(e)|\leq C|e_*-e|^{-\alpha}F(e)$ in the neighborhood of $e_*$, for some $\alpha\geq 1$. This is equivalent to $$\kappa\left(m,F\right)\leq \frac{\gamma(m,F)}{m},$$ or, after straightforward calculation and an integration by parts, $$\label{num}-\iint F'\left(e(\theta,v)\right)g_m(e(\theta,v))d\theta dv\leq 0$$ with $$g_m(e)=(\Pi_m \cos^2\theta)(e)-\left((\Pi_m\cos \theta)(e)\right)^2-(\Pi_m \sin^2\theta)(e)$$ and for all function $h(\theta)$, $$(\Pi_m h)(e)=\frac{\ds\int_{\mathcal D_e}(e+m\cos \theta)^{-1/2}h(\theta)d\theta}{\ds\int_{\mathcal D_e}(e+m\cos \theta)^{-1/2}d\theta}.$$ Now, we choose the functions $F$ as follows. We first pick a nonincreasing $\mathcal {C} ^{\infty}$ function $\Psi$ on $\RR$ with support $(\infty,e_\sharp]\subset(-\infty,m)$, then we set $e_*=\frac{e_\sharp+m}{2}$ and define $$F_\eps(e)=\Psi(e)+\eps \exp\left(-(e_*-e)^{-1}\right), \quad \mbox{for }e<e_*,$$ the parameter $\eps>0$ being arbitrary. Since $F_\eps$ satisfies the assumptions, it satisfies . Then, letting $\eps\to 0$, we get $$-\iint \Psi'\left(e(\theta,v)\right)g_m(e(\theta,v))d\theta dv\leq 0.$$ The function $\Psi$ being arbitrary, this is equivalent to $$g_m(e)\leq 0,\qquad \forall m>0,\quad \forall e\in (-m,m),$$ or, $$\label{g1}
g_1(e)\leq 0,\qquad \forall e\in (-1,1).$$ Let us now prove that the function $g_1(e)$ is in fact positive in the neighborhood of $e=1$, which contradicts .
Indeed, we introduce $$\alpha(e)=\int_{\mathcal D_e}(e+\cos\theta)^{-1/2}d\theta,\qquad \beta(e)=\int_{\mathcal D_e}(e+\cos\theta)^{-1/2}\sin^2\theta d\theta.$$ We have $$\begin{aligned}
\alpha(e) g_1(e)
&=\alpha(e)-2\beta(e)-\frac{1}{\alpha(e)}\left(\int_{\mathcal D_e}(e+\cos\theta)^{1/2}d\theta-e\alpha(e)\right)^2\\
&=(1-e^2)\alpha(e)-2\beta(e)+2e\int_{\mathcal D_e}(e+\cos\theta)^{1/2}d\theta-\frac{1}{\alpha(e)}\left(\int_{\mathcal D_e}(e+\cos\theta)^{1/2}d\theta\right)^2.\end{aligned}$$ From [@llmehats], we have $$\alpha(e)\sim-\sqrt{2}\log(1-e)\quad \mbox{as }e\to 1^-,$$ and direct calculations yield $$\int_0^{2\pi}(1+\cos\theta)^{1/2}d\theta=4\sqrt{2},\qquad \beta(1)=\frac{8\sqrt{2}}{3}.$$ This means that $$\alpha(e)g_1(e)\to \frac{8\sqrt{2}}{3}>0 \quad\mbox{as }e\to 1^-,$$ $$g_1(e)\sim \frac{8\sqrt{2}}{\alpha(e)}\mbox{as }e\to 1^-.$$ This proves the claim.
### Acknowledgments {#acknowledgments .unnumbered}
The authors wish to thank D. Han-Kwan for helpful discussions. A. M. Luz acknowledges support by the Brazilian National Council for Scientific and Technological Development (CNPq) under the program “Science without Borders’’ 249279/2013-4. M. Lemou and F. Méhats acknowledge supports from the ANR project MOONRISE ANR-14-CE23-0007-01, from the ENS Rennes project MUNIQ and from the INRIA project ANTIPODE.
M. Antoni, S. Ruffo, Clustering and relaxation in Hamiltonian long-range dynamics, Phys. Rev. E, [**52**]{} (1995), 2361. A. Antoniazzi, D. Fanelli, S. Ruffo, Y. Y. Yamaguchi, Nonequilibrium tricritical point in a system with long-range interactions, Phys. Rev. Lett. [**99**]{} (2007), 040601. J. Barré, F. Bouchet, T. Dauxois, S. Ruffo, Y. Y. Yamaguchi, The Vlasov equation and the Hamiltonian mean-field model, Physica A [**365**]{} (2006), 177. J. Barré, A. Olivetti, Y. Y. Yamaguchi, Dynamics of perturbations around inhomogeneous backgrounds in the HMF model, J. Stat. Mech. (2010), 08002. J. Barré, A. Olivetti, Y. Y. Yamaguchi, Algebraic damping in the one-dimensional Vlasov equation, J. Phys. A: Math. Gen. [**44**]{} (2011), 405502. J. Barré, Y. Y. Yamaguchi, Small traveling clusters in attractive and repulsive Hamiltonian mean-field models, Phys. Rev. E [**79**]{} (2009), 036208. J. Barré, Y. Y. Yamaguchi, On the neighborhood of an inhomogeneous stable stationary solution of the Vlasov equation – Case of an attractive cosine potential, J. Math. Phys. 56 (2015), 081502. E. Caglioti, F. Rousset, Long time estimates in the mean field limit, Arch. Ration. Mech. Anal. [**190**]{} (2008), no. 3, 517–547. E. Caglioti, F. Rousset, Quasi-stationary states for particle systems in the mean-field limit, J. Stat. Phys. [**129**]{} (2007), no. 2, 241–263. A. Campa, P.-H. Chavanis, Inhomogeneous Tsallis distributions in the HMF model, J. Stat. Mech. (2010), 06001. P.-H. Chavanis, Lynden-Bell and Tsallis distributions for the HMF model, Eur. Phys. J. B [**53**]{} (2006), 487. P.-H. Chavanis, J. Vatteville, F. Bouchet. Dynamics and thermodynamics of a simple model similar to self-gravitating systems : the HMF model, Eur. Phys. J. B, [**46**]{} (2005), 61. E. Faou, F. Rousset, Landau damping in Sobolev spaces for the Vlasov-HMF model, Arch. Ration. Mech. Anal. [**219**]{} (2016), no. 2, 887–902. E. Grenier, On the nonlinear instability of Euler and Prandtl equations, Comm. Pure Appl. Math. [**53**]{} (2000), no. 9, 1067–1091. Y. Guo, Z. Lin, Unstable and stable Galaxy models, Comm. Math. Phys. [**279**]{} (2008), 789–813. Y. Guo, W. Strauss, Nonlinear instability of double-humped equilibria, Ann. Inst. H. Poincaré Anal. Non Linéaire [**12**]{} (1995), 339–352. D. Han-Kwan, M. Hauray, Stability issues in the quasineutral limit of the one-dimensional Vlasov-Poisson equation, Comm. Math. Phys. [**334**]{} (2015), no. 2, 1101–1152. D. Han-Kwan, T. Nguyen, Instabilities in the mean field limit, J. Stat. Phys. [**162**]{} (2016), no. 6, 1639–1653. Z. Lin, Instability of periodic BGK waves, Math. Res. Letts. [**8**]{} (2001), 521–534. M. Lemou, F. Méhats, P. Raphaël, Structure of the linearized gravitational Vlasov-Poisson system close to a polytropic ground state, SIAM J. Math. Anal. [**39**]{} (2008), no. 6, 1711–1739. M. Lemou, F. Méhats, P. Raphaël, Orbital stability of spherical galactic models, Inventiones Math. [**187**]{} (2012), 145–194. M. Lemou, A. M. Luz, F. Méhats, Nonlinear stability criteria for the HMF Model, Arch. Rational Mech. Anal. [**224**]{} (2017), no. 2, 353–380. S. Ogawa, Spectral and formal stability criteria of spatially inhomogeneous solutions to the Vlasov equation for the Hamiltonian mean-field model, Phys. Rev. E [**87**]{} (2013), 062107. S. Ogawa and Y. Y. Yamaguchi, Precise determination of the nonequilibrium tricritical point based on Lynden-Bell theory in the Hamiltonian mean-field model, Phys. Rev. E, [**84**]{} (2011), 061140. Y. Shizuta, On the classical solutions of the Boltzmann equation, Comm. Pure Appl. Math. 36 (1983), no. 6, 705–754. F. Staniscia, P. H. Chavanis, G. De Ninno, Out-of-equilibrium phase transitions in the HMF model : a closer look, Phys. Rev. E. [**83**]{} (2011), 051111. Y. Y. Yamaguchi, Construction of traveling clusters in the Hamiltonian mean-field model by nonequilibrium statistical mechanics and Bernstein-Greene-Kruskal waves, Phys. Rev. E [**84**]{} (2011), 016211. Y. Y. Yamaguchi, J. Barré, F. Bouchet, T. Dauxois, S. Ruffo, Stability criteria of the Vlasov equation and quasi-stationary states of the HMF model, Physica A [**337**]{} (2004), 36.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We describe a field-driven domain wall creep-based method for the quantification of interfacial Dzyaloshinskii-Moriya interactions (DMI) in perpendicularly magnetized thin films. The use of only magnetic fields to drive wall motion removes the possibility of mixing with current-related effects such as spin Hall effect or Rashba field, as well as the complexity arising from lithographic patterning. We demonstrate this method on sputtered Pt/Co/Ir/Pt multilayers with a variable Ir layer thickness. By inserting an ultrathin layer of Ir at the Co/Pt interface we can reverse the sign of the effective DMI acting on the sandwiched Co layer, and therefore continuously change the domain wall (DW) structure from right- to the left-handed Néel wall. We also show that the DMI shows exquisite sensitivity to the exact details of the atomic structure at the film interfaces by comparison with a symmetric epitaxial Pt/Co/Pt multilayer.'
author:
- 'A. Hrabec'
- 'N. A. Porter'
- 'A. Wells'
- 'M. J. Benitez Romero'
- 'G. Burnell'
- 'S. McVitie'
- 'D. McGrouther'
- 'T. A. Moore'
- 'C. H. Marrows'
title: 'DMI meter: Measuring the Dzyaloshinskii-Moriya interaction inversion in Pt/Co/Ir/Pt multilayers'
---
The Dzyaloshinskii-Moriya interaction (DMI) [@dzyaloshinsky1958; @moriya1960] has recently returned to prominence due to recent findings in the field of magnetic domain wall (DW) motion. Initially, DWs in Permalloy nanowires were widely studied [@ganieee2000; @klauiprl2005; @yamaguchiprl2004; @hayashi2007; @meier2007; @lepadatu2009], but materials with out-of-plane (OOP) anisotropy promised even higher interaction between the current and DWs [@boulle2008; @alvarezprl2010]. It was subsequently shown that broken spatial symmetry plays an extremely important role in the current-induced DW propagation process in OOP materials [@Moore_APL2008; @Miron_Nature2010; @Buhrman_Science]. However, it has been pointed out that Bloch walls, which simple magnetostatic considerations predict to be the stable DW form in such materials [@malozemoff1979magnetic], do not have the appropriate spin texture for an efficient Slonczewski-like torque [@khvalkovskiy2013matching]. This has been demonstrated by an application of a longitudinal magnetic field which distorts the Bloch wall towards the Néel wall structure [@haazen2013domain] leading to much more efficient DW motion [@Martinez_APL]. This is of importance in the efficient and reliable operation of technologies such as racetrack memories [@parkin2008].
Soon after this demonstration, a series of theoretical [@Thiaville_DMI; @Kim_ChiralityFromSOI] and experimental works [@emori2013current; @ryu2013chiral; @lee2014; @torrejon2014] showed that a magnetic field that transforms a Bloch wall into a Néel wall can exist intrinsically due to the broken inversion symmetry at the interface. This effective field arises from the DMI as a result of high spin orbit coupling linking the broken inversion symmetry at the interface to the spin structure [@fert1991magnetic; @crepieux1998]. In contrast to the Heisenberg interaction (usually written as $- J \mathbf{S}_1 \cdot \mathbf{S}_2$ with $J$ being the exchange integral), which favours collinear alignment of neighbouring spins $\mathbf{S}_1$ and $\mathbf{S}_2$, the DMI can be expressed as $- \mathbf{D} \cdot {\mathbf{S}_{1}} \times \mathbf{S}_2$, thus preferring an orthogonal orientation of $\mathbf{S}_1$ and $\mathbf{S}_2$ with a given chirality depending on the direction of the DM vector $\mathbf{D}$. This interaction is equivalent to a magnetic field acting across the DW and establishes a Néel wall of fixed chirality which dictates the direction of DW motion under the influence of a spin Hall torque. This interfacial effect has been experimentally demonstrated by several *in situ* studies on epitaxial bilayers [@Kubetzka_2003; @Bode; @chen2013tailoring]. The DMI also plays a crucial role in bulk material systems with broken inversion symmetry producing exotic magnetization textures such as helices or skyrmions [@rossler2006; @nagaosa2013topological]. Skyrmions have been created on the atomic scale using the interfacial DMI in monolayer of Fe on Ir [@heinze2011spontaneous]. It has been predicted that skyrmions have a great potential for applications as magnetic memories due to their size and extremely low operational electric currents [@sampaio2013nucleation; @iwasaki2013current]. Therefore finding the means for *ex situ* studies of materials with high and tunable DMI is of a high interest.
Here we report a simple magnetic field-based method for DMI quantification in thin films with OOP magnetic anisotropy, and demonstrate its use by measuring the DMI inversion in Pt/Co/Ir/Pt multilayers with variable Ir thickness. Since crystallographically ordered Pt/Ni and Ir/Ni bilayers exhibit DMIs of opposite sign [@chen2013tailoring], the effective DMI in the Co layer can be potentially enhanced by placing the Pt and Ir layers on either side, i.e. using two DMI-active layers. Avoiding the use of currents to drive DW motion makes the method simple to implement, since it can be applied to sheet films and lithography is not required. Moreover, it makes the interpretation of the data much more straightforward, since the complexity of the interplay of spin-transfer, Rashba, and spin Hall torques, with their various field-like and Slonczewski-like components [@garello2013], does not enter the analysis. The power of magnetic field-based techniques has been already demonstrated by observing equi-speed contours in Pt/Co/Pt trilayers [@Choe]. It has also been suggested that the detection of the Walker breakdown can be used as a direct measure of the DMI [@Thiaville_DMI]. However, it is experimentally very difficult to observe the Walker breakdown field due to the fact that it is often not reached or hidden in the creep regime [@metaxas]. As will be seen below, the creep regime itself can be used to determine the strength of the DMI.
![(a) Sketch of the studied Ta///($t_{\mathrm{Ir}}$)/ layer stacks with a varying Ir thickness. (b) Polar Kerr hysteresis loops for samples with various Ir thickness. (c) Anisotropy field $\mu_0 H_\mathrm{K}$ and areal magnetization $M_\mathrm{s}t$ as a function of $t_{\mathrm{Ir}}$. \[Fig-KerrMicroscopy\]](Figure_1_4-eps-converted-to.pdf){width="8.7cm"}
The multilayers for our study were grown by room temperature dc sputtering at base pressures $\lesssim10^{-7}$ mbar on thermally oxidized Si substrates with a 3 nm thick Ta buffer layer. In order to reveal the effect of an Ir interface, we started from a stack of Pt(5 nm)/Co(0.7 nm)/Pt(3 nm) and inserted a thin layer of various Ir thicknesses $t_{\mathrm{Ir}}$ at the interface between the Co and top Pt layer, as depicted in Fig. \[Fig-KerrMicroscopy\](a). The films were consequently studied using polar Kerr microscopy. All the films exhibit a perpendicular anisotropy, as shown by the square OOP hysteresis loops presented in Fig. \[Fig-KerrMicroscopy\](b). The coercive field of about 20 mT in Pt/Co/Pt drops to about 9 mT as soon as the top surface is dusted with any thickness of Ir. The OOP anisotropy was measured by the vibrating sample magnetometry technique in an in-plane field configuration. Fig. \[Fig-KerrMicroscopy\](c) shows that the anisotropy field $\mu_0K_\mathrm{K}$ is about 1 T for all the films, which demonstrates that the anisotropy comes mostly from the bottom Pt/Co interface [@Lacour]. This is experimentally convenient, since it permits us to study changes in the DMI from the inclusion of the Ir layer without the complication of varying OOP anisotropy—and quantities that depend on it such as DW width—also varying.
![(a) Experimental setup in the Kerr microscope for DMI measurement. The magnetic field is tilted by a small angle $\delta$ with respect to the sample plane. (b) DW displacement in the case of $\delta=90^\circ$ after the application of a 1 s and $\mu_0H_z=7$ mT pulse. The initial DW position is indicated by the dashed line. (c) DW displacement in the case of $\delta=2.3^\circ$ after the application of a 1 s and $\mu_0H_x=+60$ mT pulse. (d) DW displacement after the application of a magnetic field pulse of the same length and $\mu_0H_x=-60$ mT. White arrows indicate the initial orientation of the magnetic moments within the DW. \[Fig\_2\]](Figure_2-eps-converted-to.pdf){width="7.5cm"}
The field-induced DW displacement was investigated by Kerr microscopy in the polar configuration. The experimental setup is shown in Fig. \[Fig\_2\](a). The magnetic field was applied in-plane with a small out-of-plane component. This is achieved by tilting the magnet by an angle $\delta$ with respect to the sample plane. This was needed due to the fact that an in-plane field alone is unable to move the DW. The role of the in-plane field is demonstrated in Fig. \[Fig\_2\](b)-(d). In each case, a reverse domain was nucleated and allowed to expand a little before switching off the field. Its shape was then recorded, indicated by the dashed line shown in Fig. \[Fig\_2\](b). Consequently we applied a 0.8-120 s long pulse of a magnetic field up to 350 mT, during which the domain expands as the DW propagates outwards. In the case of OOP field, i.e. $\delta=90^\circ$, the domain expansion is homogeneous (Fig. \[Fig\_2\](b)). The situation was very different in the case of an in-plane field component when $\delta \approx 2.3^\circ$, as shown in Fig. \[Fig\_2\](c). One can immediately see that the DW moving to the left and to the right moved with different velocities while the DWs moving in the directions perpendicular to the in-plane field moved with the same velocities. Our explanation for this observation is that the magnetic film contains Néel walls rather than Bloch walls. The in-plane magnetic field thus breaks the symmetry, and the magnetic moments within the DW on the right would be initially antiparallel, whereas the ones on the left parallel, to the magnetic field. To confirm this hypothesis, we have reversed the the sense of the in-plane magnetic field. The DW displacement after such a magnetic field pulse is shown in Fig. \[Fig\_2\](d) with the corresponding initial magnetic moment orientation within the DW.
![Differential Kerr image of the DW displacement in the case of (a) Ir (0 Å), (b) Ir (2.3 Å) and (c) Ir (4.6 Å) after the application of a 1 s and 320 mT, 1 s and 130 mT, 5 s and 60 mT field pulse, respectively.(d) DW velocity as a function of in-plane magnetic field $H_x$ in the case of Ir (0 Å), Ir (2.3 Å) and Ir (4.6 Å) for a DW creeping along the $x$ direction. The dashed curves show the fits of the creep model described by equation (\[equ\_creep\]) to the data. \[Fig\_3\]](Figure_3_7-eps-converted-to.pdf){width="7.3cm"}
The average DW velocity during a field pulse can be straightforwardly determined from the DW displacement and the pulse duration. We investigated systematically the DW velocities in the direction of in-plane magnetic field as a function of field pulse strength. A representative picture of the DW motion in a Pt/Co/Pt film is shown in Fig. \[Fig\_3\](a), showing the right-hand DW moving much faster than the left-hand one for a left pointing in-plane field component. We emphasise that the DW creep is driven by the small OOP component and the in-plane field component breaks the radial symmetry of the creep velocity. This is expressed by the asymmetry of the velocity-field curves in Fig. \[Fig\_3\](d). The detected asymmetry almost disappears in the film with 2.3 Å of Ir (Fig. \[Fig\_3\](b)), and has the opposite sign in the samples with no Ir (Fig. \[Fig\_3\](a)) and 4.6 Å of Ir (Fig. \[Fig\_3\](c)). The corresponding curves in Fig. \[Fig\_3\](d) reflect these asymmetries. The inverted asymmetry suggests an inversion of the spin texture within the DWs.
The DW displacement at low magnetic fields follows the creep law [@metaxas], which can be expressed as $$\label{equ_creep}
v = v_0 \exp\left[-\zeta \left(\mu_0H_z\right)^{-\mu}\right],$$ where $\mu=1/4$ is the creep scaling exponent, $v_0$ is the characteristic speed, and $\zeta$ is the scaling coefficient which can be expressed as [@Choe] $$\zeta=\zeta_{0} \left[\sigma(H_x)/\sigma_0 \right]^{1/4},$$ where $\zeta_{0}$ is a scaling constant, $\sigma$ is the DW energy density, which is dependent on the in-plane magnetic field $\mu_0H_x$ [@Thiaville_DMI]. This dependence can be written as $$\sigma(H_x)=\sigma_0-\frac{\pi^2\Delta \mu_0^2M_\mathrm{s}^2}{8 K_D} \left(H_x+H_{\mathrm{DMI}}\right)^2$$ for the case when the combination of the external magnetic field $\mu_0H_x$ and the intrinsic DM field $\mu_0H_{\mathrm{DMI}}$ is not able to fully transform the Bloch wall into the Néel wall, i.e. $\mid H_x + H_{\mathrm{DMI}} \mid <4K_D/\pi \mu_0M_\mathrm{s}\equiv \mu_0H_{\mathrm{N-B}}$ and $$\label{Neel_wall}
\sigma(H_x)=\sigma_0+2K_D\Delta -\pi \Delta \mu_0 M_\mathrm{s} \mid H_x+H_{\mathrm{DMI}}\mid$$ in the case of the Néel wall. In these expressions, $M_\mathrm{s}$ is the saturation magnetization, $\sigma_0$ is the Bloch wall energy density, $K_D$ is the DW anisotropy energy density, and $\Delta$ is the DW width. In this model we use $M_\mathrm{s}=1.1\times10^6$ A/m$^2$, $A=16$ pJ/m, $K_0=\mu_0 (H_\mathrm{K}M_\mathrm{s} - M_\mathrm{s}^2/2) = 3.4\times10^5$ J/m$^3$, $\Delta=\sqrt{A/K_0}=7.2$ nm and $\sigma_0 = 2\pi\sqrt{A K_0}=14$ mJ/m$^2$. The magnetostatic shape anisotropy term favoring the Bloch wall $K_D = N_x\mu_0M_\mathrm{s}^2/2 = 1.7\times10^4$ J/m$^3$ where $N_x$ is the demagnetizing coefficient of the wall [@Tarasenko]. As such, this model only requires three fitting parameters that are not determined by other experiments: the scaling parameters $v_0$ and $\zeta_0$, and $H_{\mathrm{DMI}}$ itself. This symmetry-breaking term is thus solely responsible for the asymmetry in the velocity-magnetic field plots.
![DM field and $D$ as a function of Ir thickness. The region between two dashed lines depicts the range where the DW structure changes continuously from a Néel wall to a Bloch wall and to a Néel wall of opposite chirality. Below this line (blue area) right-handed Néel wall is stable whereas above this line (red area) it is the left-handed Néel wall. The wall structures are depicted with sketches.[]{data-label="Fig_4"}](Figure_4_3-eps-converted-to.pdf){width="8.5cm"}
This model was fitted to the data for all our samples, with the fitted curves shown as the dashed lines in Fig. \[Fig\_3\](d), and the model can be seen to give an excellent description of the experimental results. The extracted DM fields as a function of Ir thickness are displayed in Fig. \[Fig\_4\]. One can see that the DM field sign reversal qualitatively agrees with the asymmetry reversal shown in Fig. \[Fig\_3\](a)-(c). The DM field is large and negative in the Pt/Co/Pt film, nearly compensated in the case of 2.3 Å of Ir and positive for $t_{\mathrm Ir}$ of 4.6 Å or greater. The calculated critical field separating the Néel wall stability region from the Bloch-Néel wall transition region is $|\mu_0H_{\mathrm{N-B}}| \approx 18$ mT. When $H_{\mathrm{DMI}}<-H_{\mathrm{N-B}}$, the DMI is able to stabilize the Néel wall structure of right-handed chirality, whilst for $H_{\mathrm{DMI}}>+H_{\mathrm{N-B}}$ the stable structure is the left-handed Néel wall, as depicted in Fig. \[Fig\_4\]. The region between two dashed lines denotes the transition region in which the DW is continuously distorted from the pure Bloch wall towards the Néel walls of the appropriate chirality. This behaviour is similar to the one observed in epitaxially grown films by Chen *et al*. [@chen2013tailoring], where the DM constant reverses sign on a similar length scale upon insertion of a thin Ir interlayer. We also emphasize that the suggested DW structure depicted in Fig. \[Fig\_2\] is no longer valid during the magnetic field pulse and all the magnetic moments eventually reorient into the field direction for sufficiently high magnetic fields. Such DWs, despite the similar magnetic moment orientation, have different energy expressed by equation (\[Neel\_wall\]). This is reflected in different resulting velocities in the creep regime.
We also estimate the effective DM constant $D$ by using the expression $D=\mu_0 H_{\mathrm{DMI}} M_\mathrm{s} \Delta$ [@Thiaville_DMI]. This is given on the right-hand ordinate axis of Fig. \[Fig\_4\]. It is apparent that the DMI in these samples is controlled largely by the top interface, in contrast to the OOP anisotropy, which we saw above to be dominated by the bottom interface. The strongest DMI, $D = 1.2\pm0.1$ mJ, is obtained in the case of Pt/Co/Ir which can be compared to the critical DMI $D_{\mathrm{crit}}$ resulting in a non-uniform magnetization state such as a cycloidal or skyrmionic phase. The critical DM constant can be estimated by using $D_{\mathrm{crit}}=4/\pi\sqrt{AK_0}$ [@heide2008dzyaloshinskii], which in this case is $D_{\mathrm{crit}}\sim 3$ mJ/m$^2$. However, the case of $D<D_{\mathrm{crit}}$ is very important for applications due to the coexistence of ferromagnetic and skyrmionic phases, so that isolated skyrmions can be used for information encoding [@sampaio2013nucleation].
![(a) High-angle annular dark-field in scanning transmission electron micrograph of the epitaxial Pt(3 nm)/Co(0.7 nm)/Pt(1 nm) trilayer. The darker Co layer is sandwiched between the two brighter Pt layers. (b) Differential Kerr image of the DW displacement in the epitaxial Co/Pt/Co sample after the application of a 1 s long, $\mu_0H_x=100$ mT field pulse. (c) Comparison of DW velocities as a function of magnetic field in the polycrystalline and epitaxial films. The dashed curves show the fits of the creep model described by equation (\[equ\_creep\]).[]{data-label="Fig_5"}](Figure_5_6-eps-converted-to.pdf){width="8.5cm"}
A strong DMI is also measured in the most structurally symmetric sample of Pt/Co/Pt, where one would not expect any DMI at all. In order to understand the origin of the strong DMI in the stack of Pt/Co/Pt, we grew a similar stack of Pt(3 nm)/Co(0.7 nm)/Pt(1 nm) epitaxially. The seed Pt layer was grown by the sputtering technique on a C-plane sapphire substrate at $500^\circ$C followed by the Co/Pt bilayer sputtering at $100^\circ$C, as described in Ref. . The epitaxial character of the grown film was confirmed by X-ray diffraction and high-angle annular dark-field imaging in a scanning transmission electron microscope. Fig. \[Fig\_5\](a) shows the high level of crystallographic ordering in the epitaxial trilayer. In order to study the DMI we have performed the same measurements as described above and Fig. \[Fig\_5\](b) shows a representative DW displacement for the epitaxial sample. One can directly see the striking difference from the picture obtained on the polycrystalline Ta/Pt/Co/Pt sample that was shown in Fig. \[Fig\_3\](a). The observed asymmetry is in this case suppressed and the DW displacement becomes radially symmetric. This is also expressed by the symmetric velocity-field curve shown in Fig. \[Fig\_5\](c) resulting in $D = 0.02\pm0.01$ mJ. The effective DMI thus vanishes in the case of the crystallographically symmetric interfaces on either side of the ferromagnet, just as expected. An important conclusion from the demonstrated experiment is that the DMI shows exquisite sensitivity to the atomic-scale details of the interfacial structure in these kinds of multilayer. Nevertheless, characterising the details of potentially asymmetric interface properties such as the roughness, degree of intermixing, density of stacking faults, remains an outstanding materials science challenge.
Besides the asymmetric metal composition and crystallographic structure around the ferromagnetic layer, the asymmetrically induced magnetic moment may play an important role. It has been shown that Pt and Ir exhibit strong proximity effect in the vicinity of a ferromagnet [@Fisher_XMCD] therefore one would expect different induced magnetic moment on either side of the Co layer. In our magnetometry data shown in Fig. \[Fig-KerrMicroscopy\](c) we see a significant drop of normalized magnetization once the Ir layer is inserted between the top Co/Pt interface indicating a decrease of induced magnetic moment in the top layer. The effect of this asymmetry on the DMI is not yet known.
In conclusion, we have demonstrated a simple-to-implement magnetic field-based method for the DMI detection and measurement in out-of-plane anisotropy materials. The DMI was quantified *ex situ* by Kerr microscopy in sputtered Pt/Co/Ir/Pt layers. We are able to control the DW chirality by changing the thickness of Ir film via in inversion of the effective intrinsic DM field. We also reveal the crucial importance of the exact nature of the ferromagnet/heavy metal interface for the DMI by comparing a polycrystalline multilayer of the type studied in most laboratories to a similar multilayer with controlled crystallographic order. The method we present opens the way for fast and convenient exploration of the DMI in new multilayer structures intended for use in DW and skyrmion racetrack memories.
This work was supported by the UK EPSRC (grant numbers EP/I011668/1, EP/I013520/1, EP/K003127/1 and EP/J007110/1), the Scottish Universities Physics Alliance and the University of Glasgow. The authors thank to Stefania Pizzini for helpful discussion.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
In this paper we study the existence of finite energy traveling waves for the Gross-Pitaevskii equation. This problem has deserved a lot of attention in the literature, but the existence of solutions in the whole subsonic range was a standing open problem till the work of Mariş in 2013. However, such result is valid only in dimension 3 and higher. In this paper we first prove the existence of finite energy traveling waves for almost every value of the speed in the subsonic range. Our argument works identically well in dimensions 2 and 3.
With this result in hand, a compactness argument could fill the range of admissible speeds. We are able to do so in dimension 3, recovering the aforementioned result by Mariş. The planar case turns out to be more difficult and the compactness argument works only under an additional assumption on the vortex set of the approximating solutions.
address:
- 'J. Bellazzini, Università di Sassari, Via Piandanna 4, 70100 Sassari, Italy'
- |
David Ruiz\
Universidad de Granada\
Departamento de Análisis Matemático\
Campus Fuentenueva\
18071 Granada, Spain
author:
- Jacopo Bellazzini
- David Ruiz
title: 'Finite energy traveling waves for the Gross-Pitaevskii equation in the subsonic regime'
---
[^1]
Introduction
============
In this paper we are concerned with the Gross-Pitaevskii equation
$$\label{eq:GP}
i \partial_t \Psi=\Delta \Psi+\Psi\left(1-|\Psi|^2\right) \text{on } {{\mathbb R}}^d \times {{\mathbb R}}$$
when $d=2$ or $d=3$. Observe that this is no more than a Nonlinear Schrödinger Equation with a Ginzburg-Landau potential. The Gross-Pitaevskii equation was proposed in 1961 ([@gross; @pita]) to model a quantum system of bosons in a Bose-Einstein condensate, via a Hartree-Fock approximation (see also [@b1; @b2; @jpr1; @jpr2]). It appears also in other contexts such as the study of dark solitons in nonlinear optics ([@k1; @k2]).
From the point of view of the dynamics, the Cauchy problem for the Gross-Pitaevskii equation was first studied in one space dimension by Zhidkov [@Z] and in dimension $d=2,3$ by Béthuel and Saut [@bs] (see also [@ge1; @ge2; @killip]). At least formally, equation presents two invariants, namely:
- *Energy:* $$\mathcal{E}= \int_{{{\mathbb R}}^d} \frac 12 |\nabla \Psi|^2 +\frac 14 \left(1-|\Psi|^2\right)^2,$$
- *Momentum:* $$\mathcal{\bf{P}}=\frac 12 \int_{{{\mathbb R}}^d} {\langle}i \nabla \Psi, \Psi {\rangle},$$ where $ {\langle}f,g {\rangle}=Re(f)Re(g)+Im(f)Im(g)$.
This paper is focused on the existence of traveling wave solutions to , that is, solutions in the form $$\label{eq:ansatz}
\Psi(x,t)=\psi(x_1-ct, \tilde{x}), \ \ \tilde{x}=(x_2 \dots x_d) \in {{\mathbb R}}^{d-1},$$ where the parameter $c\in {{\mathbb R}}$ characterizes the speed of the traveling wave. Without any lack of generality we will consider $c>0$ throughout the paper. By the ansatz the equation for the profile $\psi$ is given by $$\label{eq:GPellip}
i c \,\partial_{x_1}\psi +\Delta \psi+\left(1-|\psi|^2\right)\psi=0 \ \ \mbox{in } {{\mathbb R}}^d.$$
The study of finite energy traveling waves for has also implications in the dinamics of the equation. In particular, their pressence is an obstruction to scattering of solutions. Scattering of small energy solutions has been proved in [@gnt; @gnt2] for $d=3$, and such result is not true in dimension $d=2$. This latter fact may seem surprising for a defocusing Schrödinger equation; the reason is that finite energy solutions of do not vanish at infinity.
Nontrivial finite energy traveling waves in dimension $d=1$ are explicitly known, and they are uniquely given (up to rotation or translation) by the expression $$\psi_c(x)=\sqrt{\frac{2-c^2}{2}} \tanh \left( \frac{\sqrt{2-c^2} }{2}x \right)+i\frac{c}{\sqrt{2}},$$ if $c<\sqrt{2}$. In the literature the function $\psi_0$ is called black soliton whereas $\psi_c$ ($c \neq 0$) receives the name of dark soliton. Their orbital and asymptotic stability has been studied, see [@bgss; @bgs].
The problem of finding solutions to in dimension $d\geq 2$ has a long story. In the pioneer work of Jones, Putterman and Roberts ([@jpr1; @jpr2]), formal calculations and numerical analysis gave rise to a set of conjectures regarding existence, asymptotic behavior and stability of finite energy travelling waves: the so-called the Jones-Putterman-Roberts program. In particular, the existence of finite energy traveling waves is expected if and only if $c \in (0, \sqrt{2})$ (the sub-sonic case). The threshold value $c= \sqrt{2}$ comes from the linearization of the problem around the constant solutions of modulus 1. In a certain sense, those solutions correspond to local minima if $c< \sqrt{2}$.
In the last years much progress has been made to give rigorous proofs of those conjectures. Nontrivial finite energy traveling waves for supersonic speed $c>\sqrt{2}$ do not exist, see [@gravejat-CMP]. In dimension $d=2$ this nonexistence result holds also for $c=\sqrt{2}$, see [@gravejat-DIA]. For general nonlinearities analogous results have been proved in [@Maris-SIAM].
Concerning the asymptotics of finite energy solutions, for any $d \geq 2$, finite energy solutions of converge at infinity to a fixed complex number of modulus $1$. By the phase invariance of the problem, we can assume that $$\label{limit1} \psi(x) \to 1 \mbox{ as } |x| \to +\infty.$$ A more precise asymptotic description of $\psi$ is indeed available, see [@gravejat-AIHP; @gravejat-Asymp; @gravejat-Adv.Diff].
A very active field of research is the study of the location and dynamics of vortices, namely, the zeroes of the wave function $\psi$. The existence of multi-vortices traveling waves with small speed has been proved in dimension $d=2$, see [@lw; @chr2; @chr3]. In dimension $3$ there are traveling vortex rings ([@lwy]) as well as leapfrogging vortex rings, see [@jer2].
At least formally, the Lagrangian associated to is defined as:
$$\label{eq:ac}
I^c(\psi)= \mathcal{E}(\psi) - c \mathcal{P}(\psi)= \frac{1}{2} \int_{{{\mathbb R}}^d} |\nabla \psi|^2 -c\mathcal{P}(\psi)+ \frac 14 \int_{{{\mathbb R}}^d} \left(1-|\psi|^2\right)^2,$$
where $\mathcal{P}$ is the first component of the momentum $\mathcal{\bf P}$ that, under suitable integrability conditions (and taken into account ) can be written as: $$\label{eq:MO}
\mathcal{P}(\psi):= -\int_{{{\mathbb R}}^d} \partial_{x_1} (Im \Psi) (Re \Psi-1).$$
A classical approach to prove existence of traveling waves (starting from [@jpr1; @jpr2]) is a minimization procedure of the energy functional $\mathcal{E}$ under the constraint $\mathcal{P}(\psi)=p$ in a suitable functional space. This approach has been pursued in a number of papers, see for instance [@bgs-cmp; @bos] for the Gross-Pitaevskii equation and [@chr-mar] for more general nonlinearities. A major difficulty in this strategy is to find a natural definition of the momentum for functions with finite energy, since the integrand in \[eq:MO\] might be non-integrable (see [@bos]). This approach has the advantage of providing orbital stability of the solutions found (more precisely, of the set of minimizers). As a drawback, the speed $c$ appears as a Lagrange multiplier and is not under control. In particular the possibility of gaps in the subsonic range of velocity cannot be excluded with the constrained minimization approach (see [@bgs-cmp]).
We shall also quote existence results for small values of $c$, see [@bs] in dimension 2 and [@chr] in dimension 3, but a complete existence result in the sub-sonic case remained for many years as a standing open problem. Finally, Mariş proved in [@Ma] the existence result for any $c \in (0, \sqrt{2})$ in dimension $d \geq 3$. His approach is, summing up, to minimize $I^c(\psi)$ under a Pohozaev-type constraint. Once this is accomplished, Mariş proves that the corresponding Lagrange multiplier is 0, concluding the proof. This approach works also for more general nonlinearities with nonvanishing conditions at infinity, such as the cubic-quintic nonlinearity. As commented in [@Ma], this minimization approach breaks down in dimension 2 because of different scaling properties: the infimum is $0$ and is never attained.
One important tool in Mariş’ argument is the use of the fiber $t \mapsto u_t$, where $u_t(x_1, \tilde{x}) = u(x_1, t \tilde{x})$. For instance, in dimension $d \geq 4$ all solutions correspond to a maximizer of $I^c$ with respect to that fiber. In dimension $3$, $I^c(u_t)$ is independent of $t$ for any solution: the argument needs to be adapted, but still the use of the fiber is essential. Those cases have an analogy in the study of the Nonlinear Schödinger equation, see [@blions], [@bgk], respectively. However, in dimension $2$ this approach breaks down, and the fiber $u_t$ seems of no use; $I^c(u_t)$ attains a maximum at $t=1$ for any solution $u$.
One of the main motivations of this paper is to deal with the physically relevant $2D$ model where the existence of finite energy traveling waves in the full subsonic range is still an open problem. Our main result is the following:
\[teo:almost\] There exists a subset $E \subset (0,\sqrt{2})$ of plein measure such that, for any $c \in E$, there exists a nontrivial finite energy solution of $\psi_c$ such that:
1. For any $c_0 \in (0, \sqrt{2})$ there exists $\chi=\chi(c_0)>0$ such that $$0 <
I^c(\psi_c) \leq \chi \ \mbox{ for all } c \in E, \ c \geq c_0;$$
2. $ind (\psi_c) \leq 1$, where $ind (\psi_c)$ stands for the Morse index of $\psi_c$, that is,
$$\sup \{dim\, Y: \ Y \subset C_0^{\infty}({{\mathbb R}}^d) \mbox{ vector space, } (I^c)''(\psi_c)(\phi,\phi) <0 \ \forall \, \phi \in Y \} \leq 1.$$
The proof deals directly with the Lagrangian $I^c$ and is focused on searching critical points by using min-max arguments. Our proofs use several ingredients:
- Several regularization (or relaxation) techniques have been used in the literature to deal with the Gross-Pitaevskii equation ([@bs; @Ma]). Alternatively, some authors have proposed an approach by approximating domains, like flat tori, see [@bgs-cmp; @bos]. In this paper we choose the second approach, but we use as approximating domains the slabs: $$\label{omegaN} \Omega_N=\left\{(x_1,\tilde{x})\in {{\mathbb R}}\times {{\mathbb R}}^{d-1}, \ \ \ -N<x_1<N\right\}, \ N \in {{\mathbb N}}.$$
In other words, we first use a mountain-pass argument to address the question of existence of solutions to the problem: $$\label{eq:GPstrip}
\begin{array}{rcr} i c\partial_{x_1}\psi +\Delta \psi+\left(1-|\psi|^2\right)\psi & = & 0 \ \ \text{ on } \Omega_{N}, \\ \psi & = & 1 \text{ on } \partial \Omega_{N}. \end{array}$$ The boundary condition is motivated by . This approach has several advantages. First, as $\Omega_N$ is bounded in the $x_1$ direction, Poincaré inequality holds and we can work on the space $1+H_0^1(\Omega_N)$. As a consequence the momentum given by formula is well defined. Secondly, as $\Omega_N$ is invariant along the variable $\tilde{x}$, a Pohozaev type inequality is satisfied without boundary terms (see Lemma \[lem:poho2\]). This allows us to avoid the problem of unfolding choices of tori, as in [@bgs-cmp; @bos].
- A second fundamental tool is an energy bound argument via monotonicity in order to control the energy of (PS) sequences for almost all values of $c$. This idea has been used many times in literature starting from [@struwe]. The main point here is that we are able to obtain a *uniform bound on the energy for a subsequence of enlarging slabs $\Omega_{k(N)}$*. This is based in a key analytic argument, and it is fundamental in what follows. To the best of our knowledge, this abstract argument is completely new and could be of use in other frameworks where a monotonicity argument is used together with a relaxation procedure.
- The next step is to pass to the limit, and for that we need to deal with the problem of vanishing. Here we rely on arguments of [@bgs-cmp], and we use in an essential way that $\psi_N$ are solutions of . We can also exclude the concentration of solutions near the boundary of $\Omega_N$, since the problem posed in the half-space $$\begin{array}{rcl}i c\partial_{x_1}\psi +\Delta \psi+\left(1-|\psi|^2\right)\psi & = & 0 \ \ \text{ on } {{\mathbb R}}^d_+, \\
\psi & = & 1 \ \text{ on } \partial {{\mathbb R}}^d_+, \end{array}$$ does not admit nontrivial solutions. This is another reason for the choice of $\Omega_N$ as approximating domains (see Remark \[remark halfspace\]).
- Finally, we use the arguments of [@FG] to obtain a Morse index bound of the solutions obtained. Roughly speaking, since our solutions come from a mountain pass argument, their Morse index is at most 1. This will be used in an essential way in the proof of Theorem \[teo:all\].
With Theorem \[teo:almost\] in hand, one could ask whether we can pass to the limit and obtain a nontrivial solution for all values of $c \in (0,\sqrt{2})$. This is relatively easy, see Proposition \[ascoli\]. The problem here is to show that the limit solution has finite energy. Let us point out that the boundedness of the energy cannot be deduced only by using Pohozaev-type identities, and more delicate arguments are needed. We give two results on this aspect.
The only requirement of the next theorem is $d=3$:
\[teo:3\] Assume that $d=3$. Let $c \in (0,\sqrt{2})$, $c_n \in E$, $c_n \to c$, where $E$ is the set given by Theorem \[teo:almost\]. Let $\psi_n$ be the finite energy solutions with speed $c_n$ given by that theorem. Then there exists $\xi_n \in {{\mathbb R}}^d$ such that:
$$\psi_n(\cdot - \xi_n) \to \psi \ \mbox{ in } C^k_{loc}({{\mathbb R}}^d),$$
where $\psi$ is a nontrivial finite energy solution of with speed $c$.
Observe that Theorems \[teo:almost\] and \[teo:3\] give an alternative proof of the result of Mariş [@Ma] for the Gross-Pitaevskii equation.
Under minor changes, Theorems \[teo:almost\] and \[teo:3\] can be adapted to $d \geq 4$: the problem there is the fact that the term $(1-|\psi|^2)^2$ becomes critical or supercritical with respect to the Sobolev embedding. However, since this term has a positive sign in the functional, this issue could be fixed by changing suitably the functional setting, or, alternatively, by using a convenient truncation argument. For the sake of brevity we will not do so and restrict ourselves to the relevant spatial dimensions $d=2$ or $d=3$.
Regarding compactness of solutions, the case $d=2$ is, again, more involved. It presents analytical difficulties and also topological obstructions, see Remarks \[r1\], \[r2\]. In dimension 2 we are able to conclude only under some assumptions on the vortex set of the solutions:
\[teo:all\] Take $c \in (0,\ \sqrt{2})$, $c_n \in E$ with $c_n \to c$ and $\psi_n$ the finite energy solutions with speed $c_n$ given by Theorem \[teo:almost\]. Assume that
1. either $\psi_n$ are vortexless, that is, $\psi_n(x) \neq 0 $ for all $x \in {{\mathbb R}}^d$,
2. or there exists $R>0$, $\delta>0$ such that:
$$\label{control} \{x \in {{\mathbb R}}^d: \ \psi_n(x)=0 \} \subset B(0, R) \mbox{ and } |\psi_n(x)| \geq \delta \ \forall \, x \in \partial B(0,R).$$
Then there exists $\xi_n \in {{\mathbb R}}^d$ such that:
$$\psi_n(\cdot - \xi_n) \to \psi \ \mbox{ in } C^k_{loc}({{\mathbb R}}^d),$$
where $\psi$ is a nontrivial finite energy solution of with speed $c$.
The proofs of both Theorem \[teo:3\] and Theorem \[teo:all\] follow similar ideas, which include the following:
- A fundamental tool is the use of a lifting, that is, the existence of real functions $\rho_n(x)$, $\theta_n(x)$ such that $\psi_n = \rho_n e^{i \theta_n}$. This is always possible if the solutions are vortexless. If the solutions present vortices, one needs some information on the location of the vortex set. In Theorem \[teo:3\] one can show that the vortices are included in a set of disjoint balls, and that the number of balls and their radius is bounded. Generally speaking, a nonvanishing function $\psi$ admits a lifting if its domain is simply connected. Since the complement of a disjoint union of closed balls is simply connected if $d=3$, we can find a lifting outside those balls. In dimension 2 this is no longer true, though, and we can use a lifting only in the complement of one ball, since the total degree of a finite energy solution is 0 (see [@gravejat-AIHP]).
- We reason by contradiction assuming that $\mathcal{E}(\psi_n) \to +\infty $. A Pohozaev-type identity implies that $$\sum_{k=2}^d \int_{{{\mathbb R}}^d} |\partial_{x_k} \psi_n|^2 = (d-1) I^{c_n}(\psi_n),$$ and $I^{c_n}(\psi_n)$ is bounded by Theorem \[teo:almost\]. In our arguments we can pass to a limit (locally) which is a 1-D solution of the Gross-Pitaevskii equation (with finite or infinite energy). The knowledge of those 1-D solutions is essential at this point. For instance, in the proof of Theorem \[teo:all\] we are able to obtain in the limit a circular solution $\psi(x_1) = \rho_0 e^{i \omega_0 x_1}$, with $\rho_0^2 < \frac{2}{3} (1+c^2/4)$. But it turns out that such solution has infinite Morse index, and we reach a contradicion.
Under minor changes, it is possible to adapt the results of this paper to an equation with more general nonlinearities, namely:
$$i c\partial_{x_1}\psi +\Delta \psi+ F(|\psi|)\psi=0 \ \ \text{ on } {{\mathbb R}}^d.$$
Several assumptions on the nonlinearity $F$ would be in order. However, for the sake of brevity and clarity, we have preferred to focus on the prototype model of the Gross-Pitaevskii equation in this paper.
The rest of the paper is organized as follows. Section 2 is devoted to the setting of the notation and some preliminary results. In Section 3 we begin the proof of Theorem \[teo:almost\] by considering problem from a variational point of view. A main issue here is that we are not able to show that (PS) sequences have bounded energy. This problem is solved for almost all values of $c$ via the monotonicity trick of Struwe in Section 4. We are able to find sequences of slabs $\Omega_{k(N)}$ for which those solutions have uniformly bounded energy. In Section 5 we pass to the limit avoiding vanishing or concentration on the boundary, concluding the proof of Theorem \[teo:almost\]. Sections 6 and 7 are devoted to the proofs of Theorems \[teo:3\], \[teo:all\], respectively. The appendix deals with the Morse index computation of the 1-D circular solutions of the Gross-Pitaevskii equation, which is needed in the conclusion of Theorem \[teo:all\].
The authors wish to thank Rafael Ortega for many discussions on the 1-D solutions of the Gross-Pitaevskii equation, and also for his help in the elaboration of the Appendix.
Preliminaries
=============
In this section we collect some well-known properties of solutions of the Gross-Pitaevskii equation. We begin by stablishing the notation that we will use throughout the paper.
[**Notation:**]{} We denote by ${\langle}z_1, z_2 {\rangle}$ the real scalar product of two elements in $\mathbb{C}$, that is, ${\langle}z_1, z_2 {\rangle}= Re (z_1 \overline{z_2})$. We denote instead by $\xi_1 \cdot \xi_2$ the real scalar product in ${{\mathbb R}}^d$, to avoid confusion.
We shall use the letter $\psi$ for complex valued functions, and we will denote its real and imaginary part by $u$ and $v$, respectively, so that $\psi = u + i v$. Moreover, we will write $\rho$ to denote its modulus, that is, $\rho^2 = u^2 + v^2= {\langle}\psi, \overline{\psi} {\rangle}$.
We denote the partial derivatives by $\partial_{x_1} \psi$, but sometimes we will use $\psi_{x_1}$ for convenience.
In next lemma we are concerned with the regularity of solutions and the uniform boundedness of their derivatives.
\[lem:bound\] Any solution $\psi$ of or is of class $C^{\infty}$ and, for any $k \in {{\mathbb N}}$, there exists $C_k>0$ such that $\| D^k \psi (x) \| \leq C_k$ for any $x \in {{\mathbb R}}^d$.
The above result is well-known. The starting point is the $L^\infty$ estimate:
$$\| \psi \|_{L^\infty} \leq \sqrt{1+c^2/4}.$$
This was proved in [@farina] for all entire solutions of (not only those with finite energy). The argument works equally well for problem since the boundary condition is compatible with the $L^\infty$ bound. From this, one can obtain the result via local elliptic regularity estimates.
Indeed the solutions are analytic, see [@bgs-cmp]\[Theorem 2.1\] for more details.
Next lemma gives a Pohozaev identity:
\[lem:poho\]
Let ${\psi}$ be a finite energy solution of . Then:
$$\frac{d-2}{2} \int_{{{\mathbb R}}^d} |\nabla \psi|^2 -(d-1) c \mathcal{P}(\psi)+ \frac d 4 \int_{{{\mathbb R}}^d} \left(1-|\psi|^2\right)^2 =0.$$
See for instance [@gravejat-CMP], or [@bgs-cmp]\[Lemma 2.5 and following\].
Next identity is also of Pohozaev-type, but only uses the invariance of the domain by dilations in the $\tilde{x}$ variable:
\[lem:poho2\] Let ${\psi}$ be a finite energy solution of either or . Then the following identity holds: $$(d-3) A(\psi)+(d-1)B(\psi)=0,$$ where $$A(\psi)=\frac 12 \sum_{j=2}^d \int |\nabla_{x_j} \psi|^2$$ and $$B(\psi)=\frac{1}{2} \int |\partial_{x_1} \psi|^2 + \frac 14 \int \left(1-|\psi|^2\right)^2 -c\mathcal{P}(\psi).$$ Moreover, by the definition of the Lagrangian , we conclude that $$\label{eq:A bounded} I(\psi)=\frac{2}{d-1}A(\psi)\geq 0.$$ Finally, $I(\psi)=0$ if and only if $\psi$ is a constant function of modulus 1.
The case of has actually been proved in [@gravejat-CMP]\[Proposition 5\], taking into account [@bgs-cmp]\[Lemma 2.5\] (see also[@Maris-SIAM]\[Proposition 4.1\]). The case of the domain $\Omega_N$ is completely analogous and is based on the fact that the dilations $(x_1, \tilde{x}) \mapsto (x_1, \lambda \tilde{x} )$ leave the domain $\Omega_N$ invariant.
The following decay estimate has been proved in [@gravejat-AIHP]:
\[lem:decay\] Let ${\psi}$ be a finite energy solution of satisfying . Then the following asymptotics hold:
$$|v (x) | \leq \frac{K}{1+|x|^{d-1}}, \ \ |u(x)- 1 | \leq \frac{K}{1+|x|^{d}},$$
$$|\nabla v (x) | \leq \frac{K}{1+|x|^{d}}, \ \ |\nabla u (x) | \leq \frac{K}{1+|x|^{d+1}}.$$
Outside a ball $B(0,R)$ containing all vortices, ${\psi}$ can be lifted as $\psi = \rho e^{i\theta}$. Then the above decay estimates can be written as:
$$|\theta (x) | \leq \frac{K}{1+|x|^{d-1}}, \ \ | \rho(x) - 1 | \leq \frac{K}{1+|x|^{d}},$$
$$|\nabla \theta (x) | \leq \frac{K}{1+|x|^{d}}, \ \ |\nabla \rho (x) | \leq \frac{K}{1+|x|^{d+1}}.$$
In particular, the definition of the momentum is well defined for any finite energy solution of .
We now define the Morse index of a solution of :
\[Morse\] Let $\psi$ be a solution of (either with finite or infinite energy). We define its Morse index $ind(\psi)$ as:
$$\sup \{dim\, Y: \ Y \subset C_0^{\infty}({{\mathbb R}}^d) \mbox{ vector space, } Q(\phi)<0 \ \forall \, \phi \in Y \},$$ where $$\label{defQ} Q(\phi)=\int_{{{\mathbb R}}^d} |\nabla \phi|^2 - c {\langle}\phi, i \partial_{x_1} \phi {\rangle}- (1 - |\psi|^2) |\phi|^2 +2 {\langle}\phi, \psi {\rangle}^2.$$
If that set is not bounded from above, we will say that its Morse index is $+\infty$.
Observe that, at least formally, $Q(\phi) = I_c''(\psi)[\phi, \phi]$, and hence the Morse index is nothing but the maximal dimension for which $I_c''(\psi)$ is negative definite.
\[remark morse\] An useful property of the so-defined Morse index is that it is decreasing under convergence in compact sets. Being more specific, assume that $\psi_n$ is a sequence of solutions of or . Assume also that $ind(\psi_n) \leq m$ and $\psi_n$ converges to $\psi_0$ in $C^1_{loc}$ sense. Then $ind (\psi_0) \leq m$.
This property will be essential, in particular, in the proof of Theorem \[teo:all\].
The variational approach of Problem
====================================
We first recall the definition of $\Omega_N$ and observe that in the Sobolev Space $H_0^1(\Omega_N)$ the Poincaré inequality holds: $$\label{eq:poincare}
\int_{\Omega_N} |\phi|^2 \leq C_N \int_{\Omega_N} |\nabla \phi|^2 \ \ \forall \ \phi \in H_0^1(\Omega_N).$$
If we combine this with the Sobolev inequality we obtain that
$$\label{sob}
\| \phi \|_{L^p} \leq C_N \| \nabla \phi \|_{L^2}, \ \ \left \{ \begin{array}{ll} p\in [2, 6] & \mbox{if } d=3, \\ \\ p \geq 2 & \mbox{if } d=2. \end{array}\right.$$
Let us define the action functional $I^c_N$ as the Lagrangian $I^c$ defined in restricted to the affine space $ 1+H_0^1(\Omega_N)$, that is, $$I^c_N(\psi):= \mathcal{E}(\psi) - c \mathcal{P}(\psi)= \frac{1}{2} \int_{\Omega_N} |\nabla \psi|^2 -c\mathcal{P}(\psi)+ \frac 14 \int_{\Omega_N} \left(1-|\psi|^2\right)^2$$
We notice that thanks to the identities $$\mathcal{P}(u+i v)=- \int_{\Omega_N} (u(x)-1)\partial_{x_1} v(x),$$ $$(1-u^2-v^2)^2=(2(u-1)+(u-1)^2+v^2)^2$$ the action functional $I^c_N$ is $C^2$ in $H^1_0(\Omega_N)$. Our aim is to prove the existence of critical points of the action functional where the velocity parameter if fixed; these critical points correspond to solution to . Let us point out that $H_0^1(\Omega_N)$ is included in $H_0^1(\Omega_{N'})$ if $N'>N$ (up to extension by $0$).
Our strategy is to prove that $I^c_N$ has a mountain pass geometry on $1+ H_0^1(\Omega_N)$. More precisely we aim to prove that $$\label{gamma}
\gamma_N(c):= \inf_{g \in \Gamma} \max_{t\in [0,1]}I^c_N(g(t)) > 0,$$ where $$\label{Gamma}
\Gamma(N) =\{g \in C([0,1], (1+H_0^1(\Omega_N)): \ g(0)=1, \ g(1)=\psi_0\},$$ where $\psi_0$ is chosen so that $I^c_N(\psi_0)<0$.
\[min-max\] Given any $c_0 \in (0, \sqrt{2})$, there exist $N_0>0$, $\psi_0 \in 1 + H_0^1(\Omega_{N_0})$ and $\chi(c_0)>0$ such that $\forall N \geq N_0$, $c \in [c_0, \sqrt{2})$:
1. $I^c_N(\psi_0)<0$.
2. $0 < \gamma_N(c) \leq \chi(c_0)$.
We can write the action functional $I^c_N(\psi)$, as: $$I^c_N(\psi) = \int_{\Omega_N} \frac{1}{2} |\nabla u|^2 + \frac{1}{2} |\nabla v|^2 -c (1-u)\partial_{x_1} v +\frac 1 4 (2(u-1)+(u-1)^2+v^2)^2.$$ Moreover we have the elementary inequality $c x y\leq \frac{c^2}{4}x^2+y^2$, so that
$$\begin{aligned}
I^c_N(\psi) \geq \displaystyle \int_{\Omega_N} \frac{1}{2} |\nabla u|^2 + \left(\frac{1}{2} -\frac{c^2}{4} \right) |\nabla v|^2 - (u-1)^2 + \frac{(2(u-1)+(u-1)^2+v^2)^2}{4} \\
\geq \displaystyle\int_{\Omega_N} \frac{1}{2} |\nabla u|^2 + \left(\frac{1}{2} -\frac{c^2}{4} \right) |\nabla v|^2 - |u-1|^3 - |u-1|v^2. \qquad \qquad \qquad \quad \ \ \ \end{aligned}$$
By using Holder inequality and , we obtain: $$I^c_N(\psi) \geq \left(\frac{1}{2} -\frac{c^2}{4} \right) ||\psi-1||_{H^1_0(\Omega_N)}^2 - K||\psi-1||_{H^1_0(\Omega_N)}^3,$$ and hence $\psi=1$ is a local minimum of the action functional whenever $c^2<2$.
In [@Ma], Lemma 4.4, a compactly supported function $\phi_0$ is found so that $I^{c_0}(1+\phi_0)<0$. So it suffices to take sufficiently large $N_0$ such that $\Omega_N \supset supp \, \psi_0$, to obtain a).
Finally, define $\gamma_0(t) = 1 + t \phi_0$, which obviously belongs to $\Gamma(N)$ for all $N \geq N_0$. Observe that:
$$I_N^c(\gamma_0(t)) = \mathcal{E}(\gamma_0(t)) - c \, t^2 \mathcal{P}(\psi_0).$$
As commented above $I_N^c(\psi_0)<0$, which implies that $\mathcal{P}(\psi_0)>0$. Hence, for all $c \geq c_0$,
$$I_{N}^c(\gamma_0(t)) \leq I_N^{c_0}(\gamma_0(t)) \leq \max_{t \in [0,1]} I_{N_0}^{c_0} \circ \gamma_0(t)=\chi(c_0),$$ by definition. As a consequence, $\gamma_N(c) \leq \chi(c_0)$ for all $N \geq N_0$, $c \geq c_0$.
It is standard (see for instance [@AM; @willem]) that the mountain pass geometry induces the existence of a Palais-Smale sequence at the level $\gamma_N$. Namely, a sequence $\psi_n$ such that $$I_N^c(\psi_n)=\gamma_N(c)+o(1), \ \ \ ||(I_N^c)'(\psi_n)||_{H^{-1}_0(\Omega_N)}=o(1).$$
It is not clear if such Palais-Smale sequences are bounded or not; this is one of the main difficulties. The question of the existence of Palais-Smale sequences with bounded energy for almost all values of $c$ will be addressed in next section. In what follows we show that, if bounded, such sequences give rise to critical points of $I^c_N$.
\[eq:novan\] Let $d=2,3$ and $\left\{Q_j\right\}$ be the set of disjoint unitary cubes that covers $\Omega_N$. If $\psi_n=u_n+ i v_n$ is a bounded vanishing sequence in $1+ H^1_0(\Omega_n)$, i.e. such that $$\sup_j \int_{Q_j} |u_n-1|^p+|v_n|^p \rightarrow 0$$ for some $2\leq p<\infty$ if $d=2$, $2\leq p<6$ if $d=3$, then $$\int_{\Omega_N} |u_n-1|^r+|v_n|^r \rightarrow 0$$ for any $2< r<\infty$ if $d=2$, $2<r<6$ if $d=3.$
The proof is standard, see e.g [@lions2] Lemma I.1.
\[prop:12\] Given $0 <c<\sqrt{2}$ and $\psi_n$ a bounded Palais-Smale sequence at the energy level $\gamma_N(c)$. Then there exist $k$ sequences of points $\left\{y_n^j\right\} \subset \{0\} \times {{\mathbb R}}^{d-1}$, $1\leq j\leq k,$ with $|y_n^j - y_n^k| \to +\infty$ if $j \neq k$, such that, up to subsequence, $$\begin{aligned}
\psi_n -1 =w_n + \sum_{j=1}^k (\psi^j(\cdot + y_n^j)-1) \text{ with } w_n \rightarrow 0 \text { in } H^1_{0}(\Omega_N), \\
||\psi_n-1||_{ H^1_0(\Omega_N)}^2\rightarrow \sum_{j=1}^k ||\psi^j-1||_{H^1_0(\Omega_N)}^2, \qquad \qquad \qquad \label{eq:spl12} \\
I^c_N(\psi_n ) \rightarrow \sum_{j=1}^k I^c_N(\psi^j), \qquad \qquad \qquad \qquad \quad \label{eq:spl3}\end{aligned}$$ where $\psi^j$ are nontrivial finite energy solutions to .
In particular $I^c_N(\psi^j)\leq \gamma_N(c) \leq \chi(c_0)$, ${\mathcal{E}}({\psi^j}) \leq \limsup_{n \to +\infty} \, {\mathcal{E}}(\psi_n)$ for all $j =1, \dots k$.
In this proof, for the sake of clarity, we drop the dependence on $c$.
We first claim that $\psi_n$ is not vanishing. Reasoning by contradiction, by means of Lemma \[eq:novan\] we have $$\label{eq:consvan}
\int_{\Omega_N} |u_n-1|^r+|v_n|^r \rightarrow 0$$ for any $2< r<\infty$ if $d=2$, $2<r<6$ if $d=3.$\
Then, $$\int_{\Omega_N} (1-u_n^2-v_n^2)^2=\int_{\Omega_N} 4(u_n-1)^2+(u_n-1)^4+v_n^4+$$ $$+\int_{\Omega_N} 4(u_n-1)^3+4(u_n-1)v_n^2+2(u_n-1)^2v_n^2$$ and it follows by Hölder and that $$\label{eq:crucvan}
\int_{\Omega_N} (1-u_n^2-v_n^2)^2 =\int_{\Omega_N} 4(u_n-1)^2 +o(1).$$ Therefore $$\begin{aligned}
&I_N(u_n,v_n)=\frac{1}{2} \displaystyle \int_{\Omega_N} |\nabla u_n|^2 + \frac{1}{2} \int_{\Omega_N} |\nabla v_n|^2 -c \int_{\Omega_N} (1-u_n(x))\partial_{x_1} v_n(x) + \nonumber\\
&+ \displaystyle \int_{\Omega_N} (u_n-1)^2 +o(1) \label{eq:vanen}\end{aligned}$$ On the other hand direct computation gives $$o(1)=I_N'[\psi_n](1-\psi_n)=\int_{\Omega_N}|\nabla u_n|^2+|\nabla v|^2 -2c \int_{\Omega_N} (1-u_n)\partial_{x_1} v_n(x) -\\$$ $$+\int_{\Omega_N} (1-u_n^2-v_n^2)(u_n(1-u_n)-v_n^2).$$ Arguing as before we notice that $$\int_{\Omega_N} (1-u_n^2-v_n^2)v_n^2=o(1)$$ and, thanks to Hölder inequality and $$\int_{\Omega_N} (1-u_n^2-v_n^2)(u_n(1-u_n))=\int_{\Omega_N} (1-u_n^2-v_n^2)(u_n-1+1)(1-u_n))$$ $$=\int_{\Omega_N} (1-u_n^2-v_n^2)(1-u_n)+o(1)= \int_{\Omega_N} (1-u_n^2)(1-u_n)+o(1) =$$ $$\int_{\Omega_N} (2(1-u_n)-(1-u_n)^2)(1-u_n)+o(1)= 2||u_n-1||_{L^2(\Omega_N)}^2+o(1).$$ We get hence that $$\label{eq:vander}
I_N'[\psi_n](1-\psi_n)=\int_{\Omega_N}|\nabla u_n|^2+|\nabla v_n|^2 -2c \int_{\Omega_N} (1-u_n)\partial_{x_1} v_n(x) +2||u_n-1||_{L^2(\Omega_N)}^2+o(1)$$ Taken into account and we conclude $$\gamma(N)+o(1)=I_N(\psi_n)-\frac 12 I_N'[\psi_n](1-\psi_n)=o(1),$$ a contradiction.
Once vanishing is excluded, there exists a sequence $y^1_n \in \{0\} \times {{\mathbb R}}^{d-1}$ and $\psi^1 \in 1 + H_0^1(\Omega_N)$, $\psi^1 \neq 1$, such that $$\psi_n(\cdot + y^1_n) - \psi^1 \rightharpoonup 0 \mbox { in } H_0^1(\Omega_N)$$ up to a subsequence. From the definition of weak convergence we obtain that $$\frac 12 \int_{\Omega_N}|\nabla \psi_n|^2 -c \mathcal{P}(\psi_n)=\frac 12 \int_{\Omega_N}|\nabla \psi^1|^2 -c \mathcal{P}(\psi^1)+$$ $$+\frac 12 \int_{\Omega_N}|\nabla (\psi_n-\psi_0)|^2 -c \mathcal{P}(\psi_n-\psi^1)+o(1).$$ Now we notice that the nonlinear term fulfills the following splitting property $$\int_{\Omega_N} \left(1-|\psi_n|^2\right)^2 =\int_{\Omega_N} \left(1-|\psi^1|^2\right)^2 +\int_{\Omega_N} \left(1-|\psi_n-\psi^1|^2\right)^2 +o(1).$$ The proof of the splitting property is standard. A a consequence the action splits as $$\label{eq:split}
I_N(\psi_n ) = I_N(\psi^1)+ I_N((\psi_n-\psi^1))+o(1).$$
Clearly $\psi^1$ is a weak solution of . Now if $\psi_n-\psi^1 \rightarrow 0$ in $\tilde H^1_0(\Omega_N)$ the lemma is proved. Let us assume the contrary, i.e. that $z_n^1=\psi_n-\psi^1 \rightharpoonup 0$ and $z_n^1=\psi_n-\psi^1 \nrightarrow 0$ in $H^1_0(\Omega_N)$.
We aim to prove that there exists a sequence of points $y^2_n$ and $\psi^2 \in 1 + H_0^1(\Omega_N)$, $\psi^2 \neq 1$, such that $z_n^1(\cdot +y^2_n) - \psi^2 \rightharpoonup 0$. Let us argue again by contradiction assuming that the sequence $z_n^1$ vanishes which means by Lemma \[eq:novan\] that $$\int_{\Omega_N} |u_n-u^1|^r+|v_n-v^1|^r \rightarrow 0$$ for any $2< r<\infty$ if $d=2$, $2<r<6$ if $d=3,$ where $\psi^1 = u^1 + i v^1$. We have $$I_N'[\psi_n](1-\psi_n)=\int_{\Omega_N}|\nabla u_n|^2+|\nabla v_n|^2 -2c \int_{\Omega_N} (1-u_n)\partial_{x_1} v_n(x) +\\$$ $$+\int_{\Omega_N} (1-u_n^2-v_n^2)(u_n(1-u_n)-v_n^2)=o(1),$$ and $$I_N'[\psi^1](1-\psi^1)=\int_{\Omega_N}|\nabla u^1|^2+|\nabla v^1|^2 -2c \int_{\Omega_N} (1-u^1)\partial_{x_1} v^1(x) -\\$$ $$+\int_{\Omega_N} (1-(u^1)^2-(v^1)^2)(u^1(1-u^1)-(v^1)^2)=0.$$ Using the splitting property and using $I_N'[\psi_n](1-\psi_n)-I_N'[\psi^1](1-\psi^1)=o(1)$ we get $$\int_{\Omega_N}|\nabla z_n^1|^2 -2c \int_{\Omega_N} Re(z_n^1)\partial_{x_1}Im( z_n^1(x)) +o(1)=$$ $$\underbrace{\int_{\Omega_N} (1-u_n^2-v_n^2)(u_n(u_n-1)+v_n^2)-\int_{\Omega_N} (1-(u^1)^2-(v^1)^2)(u^1((u^1)-1)+(v^1)^2)}_{=I_4}.$$ Now, using the elementary inequality $c x y\leq \frac{c^2}{4}x^2+y^2$ we have $$\label{eq:import2}
\int_{\Omega_N}|\nabla Re(z_n^1)|^2+ (1-\frac{c^2}{2})\int_{\Omega_N}|\nabla Im(z_n^1)|^2 \leq 2 \int_{\Omega_N}|Re(z_n^1)|^2+ I_4+o(1).$$ Notice that $$(1-u_n^2-v_n^2)(u_n(1-u_n)-v_n^2)=(1-u_n^2-v_n^2)^2+(1-u_n^2-v_n^2)(u_n-1)$$ and that $$(1-u_n^2-v_n^2)^2=4(u_n-1)^2+(u_n-1)^4+v_n^4+ 4(u_n-1)^3+4(u_n-1)v_n^2+2(u_n-1)^2v_n^2.$$ On the other hand $$(1-u_n^2-v_n^2)(u_n-1)=-2(1-u_n)^2+(1-u_n)^3+v_n^2(1-u_n).$$ Therefore, assuming that $z_n^1=\psi_n-\psi^1\rightharpoonup 0$ we get $$2 \int_{\Omega_N}|Re(z_n^1)|^2+I_4=o(1)$$ and hence we get a contradiction with .\
We have hence proved the existence of a sequence $y_n^2 \in \{0\} \times {{\mathbb R}}^{d-1}$ and $\psi^2 \in 1 + H_0^1(\Omega_N)$, $\psi^2 \neq 1$, such that $$z_n^1(\cdot +y_n^2)- \psi^2 \rightharpoonup 0.$$
Clearly, $|y_n^1 - y_n^2| \to +\infty$, and $\psi^2$ is also a (weak) solution of .
Now we can iterate the splitting argument defining $z_n^2=z_n(x,y+y_n^2)-\psi^2$. We aim to show that we can have only a finite number of iterative steps.\
We claim that $$\inf_{\psi \in \mathcal{N}}||1-\psi||_{H^1_0(\Omega_N)}>0$$ where $$\mathcal{N}:=\left\{ \psi \in 1 + H^1_0(\Omega_N), \psi\neq 0, \ I_N'[\psi](1-\psi)=0 \right\}.$$ This allows us to conclude thanks to . In order to prove the claim we notice the identity $$\begin{aligned}
& \displaystyle \int_{\Omega_N} (1-u_n^2-v_n^2)(u_n(1-u_n)-v_n^2)=\int_{\Omega_N} 2(u-1)^2+3(u-1)^3+ \nonumber\\
&+ \displaystyle \int_{\Omega_N} \left( 3 v^2(u-1)+2(u-1)^2v^2+3(u-1)^3+ (u-1)^4+v^4\right)\end{aligned}$$ such that, thanks to the inequality $$-2c \int_{\Omega_N} (1-u)\partial_{x_1} v(x) \geq -\frac{c^2}{2} \int_{\Omega_N} |\nabla v|^2 - 2 \int_{\Omega_N} (u(x)-1)^2$$ we obtain $$\begin{aligned}
&0=I_N'[\psi](1-\psi)\geq \displaystyle \int_{\Omega_N} |\nabla u|^2 +(1-\frac{c^2}{2}) \displaystyle \int_{\Omega_N} |\nabla v|^2 + \label{eq:kfin}\\
& \displaystyle \int_{\Omega_N} \left(3(u-1)^3+3 v^2(u-1)+2(u-1)^2v^2+3(u-1)^3+ (u-1)^4+v^4\right) \nonumber. \end{aligned}$$ From we get $$\alpha||\psi -1||_{H^1_0(\Omega_N)}^3+\beta||\psi -1||_{H^1_0(\Omega_N)}^4\geq (1-\frac{c^2}{2})||\psi - 1||_{ H^1_0(\Omega_N)}^2$$ and hence $\inf_{\psi \in \mathcal{N}}||\psi-1||_{ H^1_0(\Omega_N)}>0$.\
Finally, recall that by , $I_n(\psi^j )>0$. Now, we have up to space translation $$\gamma_N+o(1)=I_N(\psi_n)=\sum_j I_N(\psi^j(\cdot +y_n^j)) +o(1) \geq I_N(\psi^j) +o(1)$$ and hence $I_N(\psi^j)\leq \gamma_N.$
Uniformly bounded energy solutions in approximating domains
===========================================================
In this section we prove shall prove the following result:
\[prop-crucial\] There exists a subset $E \subset (0,\sqrt{2})$ of plein measure satisfying that, for any $c \in E$, there exists a subsequence $k: {{\mathbb N}}\to {{\mathbb N}}$ strictly increasing such that:
1. There exists a nontrivial finite energy solution $\psi_N$ of the problem:
$$\begin{array}{rcr} i c\partial_{x_1}\psi_N +\Delta \psi+\left(1-|\psi_N|^2\right)\psi_N & = & 0 \ \ \text{ on } \Omega_{k(N)}, \\ \psi_N & = & 1 \text{ on } \partial \Omega_{k(N)}. \end{array}$$
2. $\mathcal{E}(\psi_N) \leq M$ for some positive constant $M=M(c)$ independent of $N \in {{\mathbb N}}$.
3. $I^c_{k(N)}(\psi_N) \leq \gamma_{k(N)}(c)$.
4. $ind(\psi_N) \leq 1.$
One of the key points here is that in (2) the energy is bounded uniformly in $N$. This will be essential later when passing to the limit as $N \to +\infty$.
In a first subsection we will give an abstract result, which is basically well-known but maybe not in this specific form. Later we will apply that result to prove Proposition \[prop-crucial\].
Entropy and Morse index bounds
------------------------------
Entropy bounds on Palais-Smale sequences via monotonicity (also called monotonicity trick argument), is a tool first devised in [@struwe] that has been used many times since then, applied to a wide variety of problems. Here we need to adapt this argument to obtain uniform bounds in $N$, for a subsequence $k(N)$. Moreover, we will also use Morse index bounds for Palais-Smale sequences, in the spirit of [@FG; @FG2]. For the sake of completeness, we state and give a proof of a general result in this subsection.
\[trick\] Let $X$ be a Banach space and $A$, $B: X \to {{\mathbb R}}$ two $C^1$ functionals. Assume that either $A({\psi}) \geq 0$ or $B({\psi}) \geq 0 $ for all ${\psi}\in X$. For any $c \in J \subset {{\mathbb R}}^+_0$, we define $I^c:X \to {{\mathbb R}}$,
$$I^c({\psi}) = A({\psi}) - c B({\psi}).$$
We assume that there are two points ${\psi}_0, {\psi}_1$ in $X$, such that setting $$\Gamma = \{g \in C([0,1], X),\ g(0) = \psi_0,\ g(1) = {\psi}_1\},$$ the following strict inequality holds for all $ c \in J$: $$\gamma(c) = \displaystyle \inf_{g \in \Gamma} \max_{t \in [0,1]} I^c (g(t)) > \max \{I({\psi}_0),I({\psi}_1)\}.$$
Then the following assertions hold true:
1. If $B \geq 0$, $\gamma$ is decreasing. If instead $A \geq 0$, then the map $\sigma(c) = \frac{\gamma(c)}{c}$ is decreasing. As a consequence, both the maps $ \gamma, \ \sigma $ are almost everywhere differentiable.
2. Let $c \in J$, $c>0$, be a point of differentiability of $\gamma$. Then, there exists a sequence $\{\psi_n\}$ such that
1. $I^c(\psi_n) \to \gamma(c)$,
2. $(I^c)'(\psi_n) \to 0$ in $X^{-1}$, and
3. $dist ({\psi}_n, G_n) \to 0$, where $$G_n=\{ \psi \in X:\ B(\psi) \leq \ - \gamma'(c) + 1/n, \ A(\psi) \leq \gamma(c) -\gamma'(c) c + \frac 1 n\}.$$
3. Let us define, for any $\delta>0$, the sets
$$\label{FGH} \begin{array}{c} F_{\delta} = \{\psi \in X: \ |I^c(\psi) - \gamma(c)|< 2\delta \}, \\ \\ G_{\delta}= \{\psi \in X: \ B(\psi) < \gamma'(c) + \delta, \ A(\psi) < - c^2 \sigma'(c) + \delta \}, \\ \\
H_\delta =\{\psi \in F_{\varepsilon}:\ dist(\psi, G_\delta) < 2 \delta \}. \end{array}$$
Let us assume that $A$ and $B$ are uniformly $C^{2, \alpha}$ functionals in $H_\delta$ for some $\delta>0$. Then in (2) we can choose $\psi_n$ satisfying also that:
1. There exists a sequence $\delta_n<0$, $\delta_n \to 0$, such that $$\sup \{dim \, Y: \ Y \subset X: I_c''(\psi_n)(\phi,\phi)\leq \delta_n \| \phi\|^2 \ \forall \ \phi \in Y \} \leq 1.$$
Observe that, in general, there exist (PS) sequences for $I^c$ for any $c \in J$; see for instance [@AM; @willem]. The above proposition shows that, for almost all values $c \in J$, there exist (PS) sequences for $I^c$ that satisfy also condition c). This extra condition c) can be useful in order to show convergence of the (PS) sequence. For instance, if either $A$ or $B$ is coercive, Proposition \[trick\] implies the existence of bounded (PS) sequences, which is an important information in order to derive convergence. This is the result of [@jeanjean].
Assertion (3) comes from [@FG] and gives also a Morse index bound of the (PS) sequence. The only novelty is that we have assumed uniform $C^{2,\alpha}$ regularity on the set $H_\delta$. If $A$ or $B$ is coercive, it suffices to have uniform $C^{2,\alpha}$ estimates on bounded sets.
To keep the ideas clear, we have stated the result under a mountain-pass geometric assumption. The same principle holds for other types of min-max arguments. What is essential is that the family $\Gamma$ does not depend on the parameter $c$.
The proof of (1) is inmediate. Indeed, if $B \geq 0$, $I^c(u)$ is decreasing in $c$. Since the family $\Gamma$ is independent of $c$, we have that $\gamma$ is decreasing. Instead, if $A \geq 0$, then the expression $\frac{I^c(u)}{c}$ is decreasing in $c$, and we conclude.
In any of the two cases, the maps $\gamma$, $\sigma$ are differentiable in a set $E \subset J$ of plein measure.
In order to prove (2), we are largely inspired by [@jeanjean]. We first state and prove the following lemma:
Let $c \in E$, $c>0$, then there exists $g_n \in \Gamma$ such that
1. $\max_{t \in [0,1]} I^c (g_n(t)) \to \gamma^c$.
2. There exists $\rho_n >0$, $\rho_n \to 0$ such that for all $t \in [0,1]$ with $I^c(g_n(t)) \geq \gamma^c - \frac 1 n$, we have:
$$B(g_n(t)) \leq -\gamma'(c) + \rho_n, \ \ \limsup_{n \to +\infty} A(g_n(t)) \leq - c^2 \sigma'(c) + \rho_n.$$
Take $c_n\in J$ an increasing sequence converging to $c$. For any $n \in {{\mathbb N}}$, there exists $g_n \in \Gamma$ such that $\max_{t \in [0,1]} I^{c_n}(g_n(t)) \leq \gamma (c_n) + |c_n-c|^2$.
If $B \geq 0$ we have that:
$$\max_{t \in [0,1]} I^{c}(g_n(t)) \leq \max_{t \in [0,1]} I^{c_n}(g_n(t)) \leq \gamma (c_n) + |c_n-c|^2 \to \gamma(c).$$
Instead, if $A \geq 0$,
$$\max_{t \in [0,1]} I^{c}(g_n(t)) \leq \frac{c}{c_n} \max_{t \in [0,1]} I^{c_n}(g_n(t)) \leq \frac{c}{c_n} (\gamma (c_n) + |c_n-c|^2) \to \gamma(c).$$
We now take $t \in [0,1]$ such that $I^c(g_n(t)) \geq \gamma(c) - |c-c_n|^2$. Then:
$$B(g_n(t)) = \frac{I^{c_n}(g_n(t)) - I^{c}(g_n(t)) }{c-c_n}$$$$\leq \frac{ \gamma(c_n) + |c_n-c|^2 - \gamma(c) + |c_n-c|^2 }{c-c_n} \to -\gamma'(c).$$
Moreover, $$\limsup_{n \to +\infty} A(g_n(t)) = \limsup_{n \to +\infty} I^c (g_n(t)) + c B(g_n(t)) \leq \gamma(c) - c \gamma'(c).$$
It suffices then to take $c_n = c- \frac{1}{\sqrt{n}}.$
Recall now the definitions of $F_\delta$, $G_\delta$ and $H_\delta$ given in . By the previous lemma the set $F_{\delta} \cap G_\delta$ is not empty: indeed, the curves $g_n$ pass through $F_{\delta} \cap G_\delta$ for sufficiently large $n$ . Proposition \[trick\], (2) is proved if we show that for any $\delta>0$, $$\inf \{ \| (I^c)'(\psi)\|: \ \psi \in H_{\delta} \}=0.$$
We argue by contradiction, and assume that there exists $\delta>0$ such that $\inf \{ \| (I^c)'(\psi)\|: \ \psi \in H_{\delta} \} \geq \delta >0$. A classical deformation argument shows that there exists ${\varepsilon}>0$, $\eta \in C([0,1] \times X:\ X)$ such that:
1. $\eta(s, \psi) = \psi$ if $s=0$, $ |I^c(\psi)- \gamma(c)| > 2{\varepsilon}$ or $dist (\psi, G_\delta) >2 \delta$.
2. $I^c(\eta(1, \psi)) \leq \gamma(c) - {\varepsilon}$ for all $\psi \in G_\delta$ with $I^c(\psi) \leq \gamma(c) + {\varepsilon}$.
3. $\eta(s, \cdot)$ is a homeomorphism of $X$.
4. $\|\eta(s,\psi) - \psi \| < \delta$,
5. $I^c(\eta(s, \psi)) \leq I^c(\psi)$ for all $\psi \in X$.
The existence of the above deformation can be found in [@willem Lemma 2.3], for instance. Actually our notation is compatible with that reference, setting $S= G_\delta$, and taking ${\varepsilon}= \delta^2/8$, for instance.
We now take $n$ large enough and the curve $\gamma_n$ given by the lemma. If $I^c(g_n(t)) < \gamma(c) - \frac 1 n$, by b), we have that $I^c(\eta (1, g_n(t))) < \gamma(c) - \frac 1 n$. In on the contrary, $I^c(g_n(t)) \geq \gamma(c) - \frac 1 n$, we can combine the lemma with ii) to conclude that $I^c(\eta(1, g_n(t))) \leq \gamma(c) - {\varepsilon}$. As a consequence,
$$\max_t I^c(\eta \circ g_n(t)) < \gamma(c),$$ a contradiction.
For the proof of (3) of Proposition \[trick\], we just use Theorem 1.7 of [@FG] to our sequence of paths $g_n$. It is important to observe that the uniform $C^{2,\alpha}$ regularity in [@FG] is required only in the set $H_\delta$ defined above (see, on that purpose, Lemma 3.7 of [@FG]).
Proof of Proposition \[prop-crucial\]
-------------------------------------
A direct application of the above results to our setting, combined with Proposition \[prop:12\], yields the existence of finite energy solutions in any domain $\Omega_N$, for almost all values of $c$. The problem here is that the energy of those solutions could diverge if we make $N \to +\infty$. In order to obtain uniform bounds independent of the parameter $N$, we need a more subtle application of Proposition \[trick\].
Define:
1. $X= 1+ H_0^1(\Omega_N)$, which is an affine Banach space, for which Proposition \[trick\] also holds;
2. $A(\psi)= {\mathcal{E}}({\psi})$, which is positive and coercive;
3. $B({\psi})= {\mathcal{P}}({\psi})$, the momentum;
4. $J= (c_0, \sqrt{2})$ for a fixed value $c_0>0$.
For $N\geq N_0$ the functional $I^c_N$ has a min-max geometry (see Proposition \[min-max\]); recall that $\gamma_N(c)>0$ is the function that associates to a speed $c \in J$ the min-max value of $I^c_N$. Clearly, $\sigma_N(c)= \frac{\gamma_N(c)}{c}$ is decreasing in $c$ as Proposition \[trick\] shows. By Proposition \[trick\], there exists a bounded (PS) sequence in $H^1_0(\Omega_N)$ at level $\gamma_N(c)$. Proposition \[prop:12\] yields then the existence of a solution $\psi_N$ with:
$$I_N^c(\psi_N) \leq \gamma_N(c), \ \mathcal{E}(\psi_N)=A(\psi_N) \leq -c^2 \sigma_N'(c).$$
Since $A$ is coercive, here the set $H_\delta$ is uniformly bounded, and $I$ is clearly uniformly $C^{2,\alpha}$ in bounded sets. By Proposition \[trick\], 3), we have that:
$$\sup \{dim \, Y: \ Y \subset H_0^1(\Omega_N): (I_N^c)''(\psi_N)(\phi,\phi) < 0 \ \forall \ \phi \in Y \} \leq 1.$$
We are now concerned with passing to the limit as $N \to +\infty$. In order to control the energy of the solutions $\psi_N$, we reason as follows.
Recall Proposition \[min-max\], b), and that $\sigma_N(c)$ is decreasing in $c$; then, for $N\geq N_0$,
$$\label{prima}
\frac{\chi(c_0)}{c_0} \geq \frac{\gamma_N(c_0)}{c_0} \geq \frac{\gamma_N(c_0)}{c_0} -\frac{\gamma_N(c)}{c} \geq \int_{c_0}^c |\sigma_N'(s)| \, ds .$$
Let us now define the sets $$D_{N,M}= \{c \in (c_0, \sqrt{2}): \sigma_N \mbox{ is not differentiable or } |\sigma_N'(c)|> M \},$$ for all $N$, $M \in {{\mathbb N}}$, $N \geq N_0$. Clearly the sets $D_{N,M}$ also depend on $c_0$, but we avoid to make that dependence explicit in the notation for the sake of clarity.
By , we have that $$|D_{N,M}| \leq \frac{\chi(c_0)}{c_0 M}.$$
The following claim is the key to be able to pass to the limit for enlargins slabs preserving bounded energy.
The set $D(c_0)$ defined as:
$$D(c_0) = \displaystyle \cap_{M \in {{\mathbb N}}} \cup_{N \geq N_0} \cap_{k \geq N} D_{k,M}$$
has $0$ measure.
Indeed, the sets $\cap_{k \geq N} D_{k,M}$ are increasing in $N$, and all of them satisfy that have measure smaller than $\frac{\chi(c_0)}{c_0 M}$. Hence the same estimate works also for the union in $N$. Now, $D(c_0)$ is a set given by an intersection of sets of measure $\frac{\chi(c_0)}{c_0 M}$, $M \in {{\mathbb N}}$, so that $D(c_0)$ has $0$ measure.
Finally, we can set $$D= \cup_{n=1}^{+\infty} \, D(1/n),$$ which has also $0$ measure.
Let us define $E= (0,\sqrt{2}) \setminus D$, and take $c \in E$. We can fix $n \in N$ such that $c_0=1/n< c$, and $c \notin D(c_0)$. Then, there exists $M(c)$ and a subsequence $k(N)$ such that $|\sigma_{k(N)}'(c)| \leq M(c)$. By Proposition \[trick\], for any of these slabs $\Omega_{k(N)}$ there exists a Palais-Smale sequence with bounded energy. According to Proposition \[prop:12\], this gives rise to a solution $\psi_{k(N)} \in 1+ H_0^1(\Omega_{k(N)})$ such that:
$$I_{k(N)}^c(\psi_{k(N)}) \leq \gamma_{k(N)}(c), \ {\mathcal{E}}(\psi_{k(N)} ) \leq M(c) c^2.$$
This concludes the proof of Proposition \[prop-crucial\].
Proof of Theorem \[teo:almost\]
===============================
In view of Proposition \[prop-crucial\], we aim to conclude the proof of Theorem \[teo:almost\] by passing to the limit. This is indeed possible thanks to Lemma \[lem:bound\]. However, we need to face two difficulties: vanishing of solutions (that is, the limit solution is trivial) and concentration near the boundary (that is, the limit solution is defined in a half-space). The purpose of this section is to exclude both scenarios.
Next result deals with the question of vanishing and is actually a version of Proposition 2.4 of [@bgs-cmp] adapted to problem .
\[lem:lemmacrucslab\] Let $\psi$ be a finite energy solution of with $0<c<\sqrt{2}$, then $$\|1-|\psi|\|_{L^{\infty}(\Omega_N)}\geq \frac{2}{5}(1-\frac{c}{\sqrt{2}}).$$
The proof is actually the same as in [@bgs-cmp], with one difference: when integrating by parts, the authors use the decay estimates of the solutions to avoid contributions from infinity, and those estimates are available only for the Euclidean space case. Instead, here we use integrability bounds.
In our argument we will use liftings of the solutions, that is, we write $\psi = \rho e^{i \theta}$. The existence of liftings is always guaranteed, for instance, if $|\psi(x)|\neq 0$ for all $x$.
Liftings for solutions in $\Omega_{N}$ without vortices
-------------------------------------------------------
We consider here solutions without vortices, i.e. that do not vanish. The energy density is given by the following formula $$e(\rho, \theta)=\frac 12 \left(|\nabla \rho|^2 + |\nabla \theta|^2\rho^2 \right)+\frac 14 \left(1-|\rho|^2\right)^2$$ and the associated energy is $$\mathcal{E}(\rho, \theta):=\int_{\Omega_N} e(\rho, \theta)$$ By using the fact that $\psi=\rho e^{i \theta}$ is a solution of , $\rho, \theta$ fulfill the following system of equations $$\label{eq:GPstripvarrp}
\left\{ \begin{aligned}
&\frac{c}{2}\partial_{x_1} \rho^2+\nabla \cdot(\rho^2\nabla \theta)=0,\\
& c \rho \partial_{x_1}\theta-\Delta \rho- \rho(1-\rho^2)+\rho|\nabla \theta|^2=0.
\end{aligned}\right..$$ The following pointwise inequality (Lemma 2.3 in BGS) $$\label{eq:pointcruc}
\left | (\rho^2-1)\partial_{x_1}\theta \right|\leq\frac{\sqrt{2}}{\rho} e(\rho, \theta)$$ that holds for arbitary $C^1$ scalar function that can be written as $\psi=\rho e^{i \theta}$ (not necessary being a solution) are crucial in the sequel.
\[prop:rhotheta\] Let $\psi$ be a vortexless finite energy solution in $\Omega_N$, then $1-\rho$ and $\theta$ belong to $H^1_0(\Omega_N)$.
Let us notice that for a vortexless finite energy solution
$$\label{energy-lifting} \mathcal{E}(\psi)=\int_{\Omega_N }|\nabla \rho|^2 + \rho^2 |\nabla \theta|^2 < +\infty$$
which implies, by means of Poincaré inequality, that $\rho-1\in H^1_0(\Omega_N)$. Since $\nabla \rho$ bounded in $L^{\infty}$ (by Lemma \[lem:bound\]), one concludes that $\rho(x) \to 1$ uniformly as $|x| \to +\infty$. Hence $\rho(x) \geq \rho_0 >0$ for all $x \in \Omega_N$. Again by , $\nabla \theta \in L^2(\Omega_N)$. To conclude the proof we shall prove that $\theta=0$ on $\partial \Omega_N$ which will allows to use Poincaré inequality.
Let us assume that $\theta=0$ if $x_1=-N$ and that $\theta=2 \pi$ if $x_1=N$. For any $\tilde{x} \in {{\mathbb R}}^{d-1}$ there exists $y \in (-N,N)$ such that $u(y, \tilde x)=0$. We get $$1=| u(y, \tilde x)-u(-N, \tilde x)|^2=\left| \int_{-N}^{y} \partial_{x_1} u(s,y) ds \right|^2\leq 2N \int_{-N}^{N} \left| \partial_{x_1} u(s,y)\right|^2 ds.$$ By Fubini we get $$\int_{\Omega_N} \left| \partial_{x_1} u\right|^2 =\int_{{{\mathbb R}}^{d-1}}\left( \int_{-N}^{N} \left| \partial_{x_1} u(s,y)\right|^2 ds\right) dy = + \infty,$$ which implies that the energy is infinity.
From we derive three useful identities that are important in the sequel. These identities have been stablished in [@bgs-cmp Lemmas 2.8, 2.10] for solutions in the whole euclidean space: here we adapt these arguments to the problem in the domain $\Omega_N$.
\[prop:stimmoment\] Let $\psi$ be a vortexless finite energy solution of . Then:
$$\label{eq:MOwv}
\mathcal{P}(\psi)=\frac 12 \int_{\Omega_N}(1-\rho^2)\partial_{x_1}\theta.$$
$$\label{eq:MOwv2}
c \mathcal{P}= \int_{\Omega_N}\rho^2|\nabla \theta|^2.$$
$$\label{eq:MOwv3}
\int_{\Omega_N} \left(2 \rho|\nabla \rho|^2+\rho(1-\rho^2)^2\right)= c \int_{\Omega_N} \rho(1-\rho^2)\partial_{x_1}\theta+ \int_{\Omega_N} \rho(1-\rho^2)| \nabla \theta|^2.$$
Straightforward computation gives $$\mathcal{P}(\psi)=\frac 12 \int_{\Omega_N}\partial_{x_1}(\rho \sin \theta)-\rho^2\partial_{x_1}\theta=\frac 12 \int_{\Omega_N}\partial_{x_1}(\rho \sin \theta -\theta)+(1-\rho^2)\partial_{x_1}\theta.$$ We will prove that $\int_{\Omega_N}\partial_{x_1}(\rho \sin \theta -\theta)=0$. Thanks to Lemma \[prop:rhotheta\] $\psi_1=(1-\rho^2)\partial_{x_1}\theta$ is integrable in $\Omega_N$ and hence we derive that $\psi_2=\partial_{x_1}(\rho \sin \theta -\theta)$ is integrable as well. By integration by parts together with Lemma \[prop:rhotheta\] we get $$\int_{\Omega_{N}} \psi_2 =\int_{\partial \Omega_{N} }\left(\rho \sin \theta -\theta\right)\eta_1 =0.$$
To get we multiply the first equation of by $\theta$ and we integrate in $\Omega_{N,M}$, defined as
$$\Omega_{N,M}=\left\{x\in {{\mathbb R}}^d, \ \ \ -N<x_1<N, \ |x_j|<M, \ \ 2 \leq j \leq d \right\} \subset \Omega_N.$$
By integrating by parts we obtain:
$$\frac{c}{2}\int_{\Omega_{N,M}} (1-\rho^2)\partial_{x_1}\theta-\int_{\Omega_{N,M}} \rho^2|\nabla \theta|^2$$$$=\int_{\partial \Omega_{N,M}} \theta \left ( \frac{c}{2} (1-\rho^2) \eta_1- \rho^2 \nabla \theta \cdot \eta \right ).$$
Observe that by Lemma \[prop:rhotheta\] all functions involved in the expression above belong to $L^1(\Omega_N)$, and recall that $\theta=0$ on $\partial \Omega_N$. Then, there exists a sequence $M_n$ such that $$\lim_{n \rightarrow \infty} \int_{\partial \Omega_{N,M_n}} \theta \rho^2 \nabla \theta \cdot \eta=0.$$ This proves .
By multiplying the second equation of by $\rho^2-1$ and integrating over $\Omega_{N,M}$ by parts we obtain $$\int_{\Omega_{N,M}} \left(2 \rho|\nabla \rho|^2+\rho(1-\rho^2)^2\right)+\int_{\partial \Omega_{N,M}} (1-\rho^2)\nabla \rho \cdot \eta =$$ $$=c \int_{\Omega_{N,M}} \rho(1-\rho^2)\partial_{x_1}\theta + \int_{\Omega_{N,M}} \rho(1-\rho^2)| \nabla \theta|^2.$$
Again by Lemma \[prop:rhotheta\], all functions involved in the above expression belong to $L^1(\Omega_N)$. Hence we can finde a sequence $M_n$ such that $$\lim_{n \rightarrow \infty} \int_{\partial \Omega_{N,M_n}}(1-\rho^2)\nabla \rho \cdot \eta=0.$$
This proves passing to the limit.
Proof of Proposition \[lem:lemmacrucslab\]
------------------------------------------
Let us call $\delta=||1-|\psi|||_{L^{\infty}(\Omega_N)}$. If $\delta>\frac 12>\frac{2}{5}(1-\frac{c}{\sqrt{2}})$ there is nothing to prove. Let us suppose hence that $\delta<\frac 12$ which implies that $\rho(x) \geq 1 - \delta >\frac 12 $ for any $x \in \Omega_N$. In particular $\psi$ admits a lifting $\psi = \rho e^{i \theta}$. We notice that $$4(1-\delta)\left(\int_{\Omega_N }\frac 12 |\nabla \rho|^2 +\frac 14 \left(1-|\rho|^2\right)^2\right)\leq\int_{\Omega_N } 2 \rho |\nabla \rho|^2 + \rho \left(1-|\rho|^2\right)^2$$ and thanks to we get $$\label{eq:stimenerg}
\int_{\Omega_N}e(\rho, \theta)\leq \frac{1}{4(1-\delta)} \int_{\Omega_N} \rho(1-\rho^2) \left( c \partial_{x_1}\theta + | \nabla \theta|^2 \right)+\frac 12 \int_{\Omega_N}\rho^2|\nabla\theta|^2$$ The strategy is to estimate r.h.s of using the pointwise bound given by . We have, thanks to and $$\frac{c}{4(1-\delta)} \int_{\Omega_N} \rho(1-\rho^2)\partial_{x_1}\theta+\frac 12 \int_{\Omega_N}\rho^2|\nabla \theta|^2\leq\left(\frac{\sqrt{2}c}{4(1-\delta)}+\frac{\sqrt {2}c}{4} \right)\int_{\Omega_N} e(\rho, \theta)$$ and hence $$\frac{c}{4(1-\delta)} \int_{\Omega_N} \rho(1-\rho^2)\partial_{x_1}\theta +\frac 12 \int_{\Omega_N}\rho^2|\nabla \theta|^2\leq \frac{c}{\sqrt{2}(1-\delta)}\int_{\Omega_N} e(\rho, \theta).$$ Now we claim that $$\label{eq:stimaimazz}
\left|\int_{\Omega_N}\rho(1-\rho^2)|\nabla \theta|^2 \right|\leq 6 \delta\int_{\Omega_N} e(\rho, \theta)$$ such that we obtain $$\int_{\Omega_N} e(\rho, \theta) \leq (\frac{c}{\sqrt{2}(1-\delta)}+\frac{3\delta}{2(1-\delta)})\int_{\Omega_N} e(\rho, \theta).$$ The fact that $e(\rho, \theta )\geq 0$ and that $1- (\frac{c}{\sqrt{2}(1-\delta)}+\frac{3\delta}{2(1-\delta)})\leq 0$ if $\delta \geq \frac{2}{5}(1-\frac{c}{\sqrt{2}})$ concludes the proof. Now we prove claim . Notice that $$\left|\int_{\Omega_N}\rho(1-\rho^2)|\nabla \theta|^2 \right| \leq \delta \int_{\Omega_N}\rho(1+\rho)|\nabla \theta|^2.$$ Now, $\rho(1+\rho)\leq 3 \rho^2$ if $\rho\geq \frac 12$, such that thanks to $$\left|\int_{\Omega_N}\rho(1-\rho^2)|\nabla \theta|^2 \right| \leq 3 \delta \int_{\Omega_N}\rho^2|\nabla \theta|^2 \leq \frac{3 \delta c}{2}\int_{\Omega_N} (1-\rho^2)\partial_{x_1}\theta \leq 3 \sqrt{2} \delta c\int_{\Omega_N} e(\rho, \theta).$$ The proof of the claim ends noticing that $0<c<\sqrt{2}$.
Conclusion of the proof of Theorem \[teo:almost\]
-------------------------------------------------
Take $c \in E$, $c_0 \in (0,c)$ and $N_0$ given by Proposition \[min-max\].
By Proposition \[lem:lemmacrucslab\], there exists $\xi_N \in \Omega_{k(N)}$ such that $|\psi_{k(N)}(\xi_N)-1| \nrightarrow 0 $, where $\psi_{k(N)}$ are the solutions given by Proposition \[prop-crucial\]. By the uniform bounds of Lemma \[lem:bound\], we can use Ascoli-Arzelà Theorem to obtain in the limit ($C^k$ locally) a nontrivial solution of the Gross-Pitaevskii equation $\psi_c$. By Fatou Lemma, $\mathcal{E}(\psi_c) $ is finite. Moreover, by Remark \[remark morse\], $ind(\psi_c) \leq 1$.
Finally, by Fatou lemma and ,
$$\begin{aligned}
I(\psi_c)= \frac{1}{d-1} \sum_{j=2}^d \int |\partial_{x_j} \psi_c |^2 \leq \frac{1}{d-1} \liminf_{N \to +\infty} \sum_{j=2}^d \int |\partial_{x_j} \psi_{k(N)} |^2 \\ = \liminf_{N \to +\infty} I_{k(N)} (\psi_{k(N)} ) \leq \liminf_{N \to +\infty} \gamma_{k(N)}(c) \leq \chi(c_0). \end{aligned}$$
If $d (\xi_N, \partial \Omega_{k(N)}) $ is bounded, up to a subsequence, our limit solution $\psi_c$ is defined in a half-space $\{x \in {{\mathbb R}}^d: \ x_1 > - m\}$ or $\{x \in {{\mathbb R}}^d: \ x_1 < m\}$, for some $m>0$, and $\psi_c=1$ on its boundary. Instead, if $d (\xi_N, \partial \Omega_{k(N)}) $ is unbounded, the solution $\psi_c$ is defined in the whole euclidean space ${{\mathbb R}}^d$. In next proposition we rule out the first possibility, and this concludes the proof of Theorem \[teo:almost\].
\[prop:half\] Let $\psi$ be a finite energy solution of the problem:
$$\label{eq:half}
\begin{array}{rcl}i c\partial_{x_1}\psi +\Delta \psi+\left(1-|\psi|^2\right)\psi & = & 0 \ \ \text{ on } {{\mathbb R}}^d_+, \\
\psi & = & 1 \ \text{ on } {{\mathbb R}}^d_+, \end{array}$$
where ${{\mathbb R}}^d_+ = \{x \in {{\mathbb R}}^d:\ x_1>0\}.$ Then $\psi=1$.
The proof follows well-known ideas that date back to [@el]. If $\psi$ is a finite energy solution, then $\nabla \psi$ and $(1-|\psi|^2)$ are functions in $L^2({{\mathbb R}}^d_+)$. Since $\psi$ is in $L^\infty({{\mathbb R}}^d_+)$ and is a strong solution, standard regularity results allow us to conclude that $D^2 \psi$ belongs to $L^2({{\mathbb R}}^d_+)$. Hence we can multiply equation by $\partial_{x_1}\psi$ and integrate by parts, obtaining:
$$c \int_{{{\mathbb R}}^d_+} {\langle}(i \partial_{x_1}\psi) , \partial_{x_1}\psi {\rangle}=0;$$
$$\int_{{{\mathbb R}}^d_+} {\langle}\Delta \psi, \partial_{x_1}\psi {\rangle}= \int_{\partial {{\mathbb R}}^d_+} {\langle}(\nabla \psi \cdot \nu ), \partial_{x_1} \psi {\rangle}- \int_{{{\mathbb R}}^d_+} \frac{1}{2} \partial_{x_1} \left ( |\nabla \psi|^2 \right )$$ $$= - \int_{\partial {{\mathbb R}}^d_+} | \partial_{x_1} \psi|^2 + \frac{1}{2} \int_{\partial {{\mathbb R}}^d_+} |\partial_{x_1} \psi|^2= -\frac{1}{2} \int_{\partial {{\mathbb R}}^d_+} |\partial_{x_1} \psi|^2;$$
$$\int_{{{\mathbb R}}^d_+} \left(1-|\psi|^2\right){\langle}\psi, \partial_{x_1}\psi {\rangle}= - \frac 1 4 \int_{{{\mathbb R}}^d_+} \partial_{x_1} \left ( (1- |\psi|^2)^2 \right )=0.$$
These computations imply that:
$$\int_{\partial {{\mathbb R}}^d_+} |\partial_{x_1} \psi|^2 =0.$$
In other words, $\partial_{x_1} \psi=0$ in $\partial {{\mathbb R}}^d_+$. By unique continuation, we conclude that $\psi=1$.
\[remark halfspace\]
The proof of Proposition \[prop:half\] breaks down if we consider the half-space $\{ x \in {{\mathbb R}}^d:\ x_j >0 \}$, if $j>1$. The reason is that we do not know if:
$$\int_{{{\mathbb R}}^d_+} {\langle}(i \partial_{x_1}\psi) , \partial_{x_j}\psi {\rangle}$$ may cancel. This is one of the reasons why we choose slabs as approximating domains, instead of expanding balls, for instance (a choice that would have had advantages from the point of view of compactness). The second reason is that in $\Omega_N$ the Pohozaev-type identity given in Lemma \[lem:poho2\] does not involve boundary terms.
Proof of Theorem \[teo:3\]
==========================
In this section we prove the compactness criterion given in \[teo:3\]. We start by the following result, which is independent of the dimension:
\[ascoli\] Let $d=2$ or $3$, $c_n \to c$, $c_n \in E$ where the set $E$ is given by Theorem \[teo:almost\]. Let $\psi_n$ be the sequence of solutions provided by that theorem. Then there exists $\xi_n \in {{\mathbb R}}^d$ such that $\psi_n(\cdot - \xi_n)$ converges locally in $C^k$ (up to a subsequence) to a nontrivial solution $\psi_0$ of .
Let $\psi$ be a finite energy solution of with $0<c<\sqrt{2}$. Then there exists ${\varepsilon}= {\varepsilon}(c) >0$ such that $$\label{puff} \|1-|\psi|\|_{L^{\infty}(\Omega_N)}\geq {\varepsilon}.$$
Statement is just Proposition 2.4 of [@bgs-cmp]. Compare it with Proposition \[lem:lemmacrucslab\], which is nothing but its version for problem (with a slight change of the constants).
Then, there exists $\xi_n$ such that $ | 1- |\psi_n(\xi_n)||> {\varepsilon}$ for some fixed ${\varepsilon}>0$. By Lemma \[lem:bound\] we can use Ascoli-Arzelà Theorem to obtain that $\psi_n(\cdot - \xi_n)$ converges locally in $C^k$ to a nontrivial solution $\psi_0$ of .
The main problem to conclude the proof of Theorem \[teo:3\] or \[teo:all\] is to assure that $\psi_0$ has finite energy. Let us point out that the boundedness of the energy cannot be deduced only by using the Pohozaev identities given in Lemmas \[lem:poho\], \[lem:poho2\].
Observe that since $I_{c_n}(\psi_n )$ is bounded, Lemma \[lem:poho2\] implies that $$\sum_{j=2}^3 \int_{{{\mathbb R}}^3} |\partial_{x_j} \psi_n|^2 = O(1).$$
The idea of the proof is to try to relate the behavior of $\psi_n$ with that of the 1-D solutions of the Gross-Pitaevskii equation. Next proposition is a first step in this line (see also Remark \[rrr\]).
Let $\psi_n$ be solutions of for $c_n$, $c_n \to c$, such that $I_{c_n}(\psi_n) \leq C$. Then,
$$\int_{{{\mathbb R}}^3} |\nabla g_n|^2 + \int_{{{\mathbb R}}^3} |\nabla h_n|^2 =O(1),$$
where
$$\label{g} g_n= (\partial_{x_1} u_n) v_n - (\partial_{x_1} v_n) u_n - \frac{c_n}{2} (\rho_n^2-1),$$
$$\label{h} h_n= \frac 1 2 |\partial_{x_1} \psi_n|^2 - \frac 1 4 (1-\rho_n^2)^2.$$
\[rrr\]
The two quantities defined above correspond to the invariants of the 1-D Gross-Pitaevskii equation. Indeed $h$ represents its hamiltonian, whereas $g$ is another invariant given by the fact that the problem, after a change of variables, is radially symmetric (see equations , ). On this aspect, see for instance [@bgs-survey], pages 3-4.
For the sake of clarity we drop the subscript $n$ in the proof of this proposition.
We first consider the function $g$, which is an $L^2$ function, but with $L^2$ norm out of control. Observe that equation implies that $\nabla \cdot G=0$, where
$$\label{defG} G=(g, u_{x_2} v - v_{x_2} u, u_{x_3} v - v_{x_3} u).$$
Straightforward computations give:
$$curl \, G =\left( \begin{array}{c} \displaystyle 2 u_{x_3} v_{x_2} - 2 v_{x_3} u_{x_2} \\ \displaystyle 2 v_{x_3} u_{x_1} - 2 u_{x_3} v_{x_1} - \frac{c}{2} (\rho^2-1)_{x_3} \\ \displaystyle
\displaystyle - 2 v_{x_2} u_{x_1} + 2 u_{x_2} v_{x_1} + \frac{c}{2} (\rho^2-1)_{x_2} \end{array} \right).$$
Observe that by , the derivatives with respect to $x_2$, $x_3$ are uniformly bounded (with respect to $n$) in $L^2$. Moreover, all factors involved are bounded in $L^\infty$ by Lemma \[lem:bound\]. As a consequence, $curl G$ is uniformly bounded in $L^2$. Observe now that:
$$G= curl (-\Delta^{-1} curl \, G ),$$
where $\Delta^{-1}$ is given by convolution with the Coulomb potential $\frac{1}{4 \pi |x|}$, see for instance [@bertozzi Subsection 2.4.1].
By using the Fourier Transform and Plancherel, all partial derivatives of $G$ are uniformly bounded in $L^2$, independently of $n$. This concludes the proof for $g$.
For $h$, the proof follows the same ideas. Let us define the vector field:
$$H =\left( h, u_{x_1} u_{x_2} + v_{x_1} v_{x_2}, u_{x_1} u_{x_3} + v_{x_1} v_{x_3} \right).$$
Let us recall here that $|\psi_{x_1}|^2 = u_{x_1}^2 + v_{x_1}^2$. Observe first that $H$ is an $L^2$ vector field, even if its $L^2$ norm could be unbounded as $n \to +\infty$. Taking into account , straightforward computations give:
$$\nabla \cdot H = u_{x_1 x_2} u_{x_2} + u_{x_1 x_3} u_{x_3} + v_{x_1 x_2} v_{x_2} + v_{x_1 x_3} v_{x_3}$$
which is uniformly bounded in $L^2$ norm, again, by and Lemma \[lem:bound\]. Moreover, we can compute:
$$curl \, H = \left( \begin{array}{c} \displaystyle u_{x_1 x_2} u_{x_3} + v_{x_1 x_2} v_{x_3} - u_{x_1 x_3} u_{x_2} -v_{x_1 x_3} u_{x_2} \\ \displaystyle - u_{x_1 x_1} u_{x_3} - v_{x_1 x_1} v_{x_3} - (1-\rho^2) (u u_{x_3} + v v_{x_3})\\ \displaystyle u_{x_1 x_1} u_{x_2} + v_{x_1 x_1} v_{x_2} + (1-\rho^2) (u u_{x_2} + v v_{x_2}) \end{array} \right).$$
which is also uniformly bounded in $L^2$ norm. We now recall that:
$$H = \nabla (\Delta^{-1} (\nabla \cdot H)) - curl (\Delta^{-1} \, curl H),$$
see again [@bertozzi Subsection 2.4.1]. By using the Fourier Transform and Plancherel, all partial derivatives of $H$ are uniformly bounded in $L^2$, finishing the proof.
\[r1\] Let us point out that the above result can be easily extended to any dimension. However, in dimension 3 it implies, by Sobolev inequality, that:
$$\label{eq:Sob} \int_{{{\mathbb R}}^3} |g_n|^6 + \int_{{{\mathbb R}}^3} |h_n|^6 = O(1)$$
In dimension $d>3$ the Sobolev exponent is $\frac{2d}{d-2}$. However we cannot deduce a similar expression in dimension 2. The lack of a Sobolev inequality in dimension 2 is one of the obstacles for this approach to work also in the planar case.
\[defS\] We define the set $S_n^r= \{ x \in {{\mathbb R}}^3: \rho_n(x) < r \}$. The behavior of these sets will be important in our arguments.
Next lemma is a key ingredient in our proof.
\[lem:51\] Under the assumptions of Theorem \[teo:3\], assume that for some $r\in (0, 1)$, $|S_n^{r}| \to +\infty$. Then, there exists $\xi_n \in S_n^r$ and $R_n \to +\infty$ such that:
$$\int_{B(\xi_n, R_n)} |g_n |^6 + |h_n |^6 + \sum_{i=2}^3 |\partial_{x_i} \psi_n |^2 \to 0.$$
Take $x^n_1 \in S_n^r$, and define $R_n = |S_n^r|^{\frac{1}{6}}$. Observe that
$$|B(x^n_1, 2 R_n)| = c_0 |S_n^r|^{\frac{1}{2}}, \ \ c_0 = \frac{32}{3} \pi.$$
As a consequence, there exists $x^n_2 \in S_n^r \setminus B(x^n_1, 2 R_n)$. Clearly,
$$B(x^n_1, R_n) \cap B(x^n_2, R_n) = \emptyset \mbox { and } |B(x^n_1, 2 R_n) \cup B(x^n_2, 2 R_n) | \leq 2 c_0 |S_n^r|^{\frac{1}{2}}.$$
We can choose then $x^n_3 \in S_n^r \setminus (B(x^n_1, 2 R_n) \cup B(x^n_2, 2 R_n))$, with
$$B(x^n_1, R_n) \cap B(x^n_2, R_n) \cap B(x^n_3, R_n)= \emptyset$$ and $$|B(x^n_1, 2 R_n) \cup B(x^n_2, 2 R_n) \cup B(x^n_3, R_n)| \leq 3 c_0 |S_n^r|^{\frac{1}{2}}.$$
In this way we find $x_n^1 \dots x_n^{j_n} \in S_n^r$ with:
$$B(x^n_j, R_n) \cap B(x^n_k, R_n) = \emptyset, \ \ j,\ k \in \{1, \dots j_n \}, \ j \neq k.$$
where $j_n = [ \frac{|S_n|^{1/2}}{c_0}]$ (here $[a]$ denotes the largest integer smaller or equal than $a$). Hence we can choose $\xi_n = x_n^k$ such that, taking into account and :
$$\int_{B(\xi_n, R_n)} | g_n |^6 + | h_n |^6 + \sum_{i=2}^d |\partial_{x_i} \psi_n |^2 \leq \frac{C}{j_n} \to 0.$$
The above result will be the key to prove next proposition, which allows us to have some control on the set of vortices of the solutions.
\[prop:vortices\] Let us fix $r \in (0, c/\sqrt{2})$. Then, there exists $N \in \mathbb{N}$ and $N$ sequences of disjoint closed balls $\overline{B_k^n}= \overline{B}(\xi_k^n, R_k)$ ($k=1 \dots n$) with $R_k \in (1, N)$ such that
$$S_n^r \subset \cup_{k=1}^N B_k^n.$$
The proof is divided into several steps:
[**Step 1:** ]{} $|S_n^r|$ remains bounded.
Assume by contradiction that $|S_n^r| \to +\infty$ as $n \to +\infty$. Take $\xi_n \in {{\mathbb R}}^3$ given by Lemma \[lem:51\], and define $\tilde{\psi}_n = \psi_n(\cdot - \xi_n)$. By Lemma \[lem:bound\] we can use Ascoli-Arzelà theorem to conclude that, up to a subsequence, $\tilde{\psi}_n$ converges $C^k$ locally to a solution $\psi$ of . By the choice of $\xi_n$, this solution satisfies that $\rho(0) \leq r$. Moreover, by Fatou lemma we have that:
$$\partial_{x_2} \psi =0, \ \partial_{x_3} \psi =0, \ g=0, \ h =0,$$ where $g$ and $h$ are the analogous of , , namely:
$$g= u_{x_1} v - v_{x_1} u - \frac{c}{2} (\rho^2-1),$$
$$h= \frac 1 2 |\psi_{x_1}|^2 - \frac 1 4 (1-\rho^2)^2.$$
As a consequence, $\psi$ is a 1-D solution to the Gross-Pitaevskii equation with $g=0$, $h=0$. But those are precisely the finite energy 1-D travelling waves (see [@bgs-survey pages 3, 4]); hence, after a rotation, $\psi$ has the explicit expression:
$$\psi(x_1)=\sqrt{\frac{2-c^2}{2}} \tanh \left( \frac{\sqrt{2-c^2} }{2}(x_1 + t) \right)+i\frac{c}{\sqrt{2}}, \ t \in {{\mathbb R}}.$$
But this is in contradiction with $|\psi(0)| \leq r < c/\sqrt{2}$, concluding the proof.
[**Step 2:**]{} There exists $N \in \mathbb{N}$, $\xi_k^n \in {{\mathbb R}}^3$ such that:
$$S_n^r \subset \cup_{k=1}^N B(\xi_k^n, 1).$$
Fix $s \in (r, \frac{c}{\sqrt{2}})$, and define $\xi_1^n$ as any point in $S_n^r$. Since $\nabla \rho_n$ is uniformly bounded (Lemma \[lem:bound\]), there exists $\delta >0$ such that $B(\xi_1^n, \delta) \subset S_n^s$. It suffices to take $$\delta \leq \frac{s-r}{sup_{n} \| \nabla \rho_n \|_{L^{\infty}}}.$$ Without loss of generality we can assume that $\delta < 1/2$.
Take now $\xi_2^n$ any point in $S_n^r \setminus B(\xi_1^n, 1)$; again, $B(\xi_2^n, \delta) \subset S_n^s$, and observe that $B(\xi_1^n, \delta) \cap B(\xi_2^n, \delta) = \emptyset$.
We follow by taking $\xi_3^n$ any point in $S_n^r \setminus \left( B(\xi_1^n, 1) \cup B(\xi_2^n, 1) \right)$, if there exists one. Since $|S_n^s|$ is bounded by the step 1, this procedure has to finish at a certain point, yielding the thesis of the proposition. Indeed we cannot find more than $N$ such points, where
$$N = \left [ \frac{\sup_n |S_n^s|}{\frac 4 3 \pi \delta^3} \right ].$$
Recall that $[a]$ stands for the largest integer smaller or equal than $a$.
[**Step 3:**]{} Conclusion
By step 2, we have already $S_n^r$ contained in $N$ balls of radius 1. The problem is that they might not be disjoint. We now make a procedure of aggregation of balls which is gererally described as follows:
Take a closed ball $\overline{B}(x, R_x)$. If it intersects a closed ball $\overline{B}(y,R_y)$, we replace both balls by $\overline{B}(x, R_x+R_y)$. We now repeat the procedure to the new set of balls.
We apply this procedure iteratively to the balls given in Step 2, and in this way we conclude.
The above proposition is the first milestone in our proof: it allows us to control the vortices of the solutions, as they are always contained in a fixed number of disjoint balls of bounded radii. Since ${{\mathbb R}}^3 \setminus \cup_{k=1}^N \overline{B}_k^n$ is a simply connected open set, we can guarantee the existence of a lifting of $\psi_n$ outside these balls. Being more specific, taking $\frac{c}{2}$ as the value $r$ (for instance), we can write:
$$\psi_n(x) = \rho_n(x) e^{i \theta_n(x)} \ \forall \ x \in {{\mathbb R}}^d \setminus \cup_{k=1}^N B_k^n$$ where the balls $B_k^n$ are given by Proposition \[prop:vortices\]. Since $\psi_n$ is a solution of , we have that $\rho_n$, $\theta_n$ satisfy equations .
\[r2\] This is a second crucial point in which the requirement $d\geq 3$ is crucial. If $d=2$ we can have liftings of finite energy solutions outside one ball (see [@gravejat-AIHP Lemma 15]), but this is not possible in the complement of two or more disjoint balls.
Next lemma is inspired in [@bgs-cmp]\[Lemmas 2.8, 2.10\], which are concerned with the case without vortices. Compare it with the identities , for the vortexless case.
\[lem:O(1)\] Take $c \in (0,\ \sqrt{2})$, $c_n \in E$ with $c_n \to c$ and $\psi_n$ the solutions given by Theorem \[teo:almost\]. Then $$\label{eq:stimmomassa}
\mathcal{P}(\psi_n)=\frac 12 \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n}(1-\rho_n^2)\partial_{x_1}\theta_n +O(1),$$ $$\label{eq:stimmomass2b}
c \mathcal{P}(\psi_n)= \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n}\rho_n^2|\nabla \theta_n|^2 +O(1),$$ $$\label{eq:stimmomass3c}
\int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n}|\nabla \rho_n|^2 =O(1).$$
First of all, observe that $$\nabla \theta = \frac{u \nabla v-v \nabla u}{\rho^2},$$ so that
$$\label{boundtheta} |\nabla \theta_n | = O(1) \mbox{ in } \partial B_k^n \Rightarrow |\theta_n(p) - \theta_n(q)| \leq C \ \forall \ p,\ q \in \partial B_{k}^n.$$
This is useful in what follows; observe that we do not know whether $\| \theta_n \|_{L^{\infty}}$ is bounded or not.
Direct computation gives $$\mathcal{P}(\psi_n)=\frac 12 \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n}\partial_{x_1}(\rho_n \sin \theta_n)-\rho_n^2\partial_{x_1}\theta +\frac 12 \int_{\cup_{k=1}^N B_k^n} {\langle}i \partial_{x_1} \psi_n, \psi_n-1{\rangle}$$ which implies that $$\label{eq:stimmomassz}\mathcal{P}(\psi_n)=\frac 12 \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n}\partial_{x_1}(\rho_n \sin \theta_n -\theta_n)+(1-\rho_n^2)\partial_{x_1}\theta_n +O(1).$$ In order to get it suffices hence to prove that $$\label{eq:stimmomasszz} \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n}\partial_{x_1}(\rho_n \sin \theta_n-\theta_n)=O(1).$$
We recall that $ \partial_{x_1} (\rho_n \sin \theta_n - \theta_n)$ is integrable thanks to and . By integration by parts, using the the decay estimates at the infinity, we get $$\int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n} \partial_{x_1}(\rho_n \sin \theta_n-\theta_n) =\sum_{k=1}^N \int_{\partial B_k^n} \left(\rho_n \sin \theta_n -\theta_n\right)\eta_1,$$ where $\eta_1$ is the first component of the inward unit normal vector to the spheres $B_k^n$. Relation follows now from together with the fact that the outward unit surface normal $\eta_1$ has zero average on the sphere, which implies that $$\int_{\partial B_k^n} \theta_n\eta_1=\int_{\partial B_k^n} \left(\theta_n-\theta_n(p_0)\right)\eta_1=O(1),$$ where $p_0$ is an arbitrary point on the sphere $B_k^n$.\
In order to prove we argue as in Lemma \[prop:stimmoment\], i.e. multiplying per first equation of by $\theta_n$ and then integrating on $R^3 \setminus \cup_{k=1}^N B_k^n$. By integration by parts we get $$\frac{c}{2}\int_{R^3 \setminus \cup_{k=1}^N B_k^n} (1-\rho_n^2)\partial_{x_1}\theta_n -\int_{R^3 \setminus \cup_{k=1}^N B_k^n} \rho_n^2|\nabla \theta_n|^2$$$$= \sum_{k=1}^N \int_{\partial B_k^n} \theta_n \left ( \frac{c}{2} (1-\rho_n^2) \eta_1- \rho_n^2 \nabla \theta \cdot \eta \right ).$$
Observe that $G_n(x)= (\frac{c}{2} (1-\rho_n^2), 0, 0 ) - \rho_n^2 \nabla \theta_n$, where $G_n$ is defined in . In particular it is defined in the whole euclidean space and $\nabla \cdot G_n=0$. By integrating by parts in $B_k^n$, we obtain that
$$\int_{\partial B_k^n} \frac{c}{2} (1-\rho_n^2) \eta_1- \rho_n^2 \nabla \theta \cdot \eta =0.$$
As a consequence, we can use to obtain:
$$\int_{\partial B_k^n} \theta_n \left ( \frac{c}{2} (1-\rho_n^2) \eta_1- \rho_n^2 \nabla \theta \cdot \eta \right )$$$$= \int_{\partial B_k^n} (\theta_n - \theta_n(p_0)) \left ( \frac{c}{2} (1-\rho_n^2) \eta_1- \rho_n^2 \nabla \theta \cdot \eta \right )=O(1).$$
Now we prove . From Lemma \[lem:poho\] we get
$$\frac{1}{2} \int_{{{\mathbb R}}^3} |\nabla \psi_n|^2 -2 c \mathcal{P}(\psi_n)+ \frac 3 4 \int_{{{\mathbb R}}^3} \left(1-|\psi_n|^2\right)^2 =0,$$ which implies, thanks to $$\frac{1}{2} \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n} |\nabla \rho_n|^2 +\frac{3}{2} \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n} \rho^2|\nabla \theta_n|^2 + \frac 3 4 \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n} \left(1-|\rho_n|^2\right)^2 =$$ $$=3 \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n} \rho^2|\nabla \theta_n|^2 +O(1)=3c \mathcal{P}(\psi_n)+O(1).$$ As a consequence we get $$3 \left(\mathcal{E}(\psi_n)-c \mathcal{P}(\psi_n)\right)=\int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n} |\nabla \rho_n|^2+O(1)$$ and hence follows by the fact that $\mathcal{E}(\psi_n)-c \mathcal{P}(\psi_n)=I(\psi_n)=O(1).$
For next proposition it is useful to recall the definition \[defS\].
\[prop:ma\] Take $c \in (0,\ \sqrt{2})$, $c_n \in E$ with $c_n \to c$ and $\psi_n$ the solutions given by Theorem \[teo:almost\]. Assume that:
$$\mathcal{E}(\psi_n) \to + \infty.$$
Then, for any $r \in (\frac{c}{\sqrt{2}}, 1)$, $|S_n^r| \to +\infty$.
We recall the form of the energy for functions $\psi$ given by a lifting $\psi = \rho e^{i \theta}$:
$$e(\rho, \theta)= \frac 12 \left(|\nabla \rho|^2 + |\nabla \theta|^2\rho^2 \right)+\frac 14 \left(1-|\rho|^2\right)^2.$$
The following function represents the lagrangian in the vortexless case, and is an approximation of the real lagrangian in view of :
$$l(\rho, \theta)= e(\rho, \theta)- \frac{c_n}{2} (1-\rho^2)\partial_{x_1} \theta.$$
Assume by contradiction that $|S_n^r|$ is bounded for some $r > \frac{c}{\sqrt{2}}$. Observe that:
$$I^{c_n} (\psi_n ) = \int_{{{\mathbb R}}^3 \setminus \cup_{k=1}^N B_k^n} l(\rho_n, \theta_n) + O(1)= \int_{\{\rho_n \geq r\}} l(\rho_n, \theta_n) + O(1).$$
We now use the inequality $|\frac c 2 (1-\rho_n^2)\partial_{x_1} \theta_n| \leq \frac{(1-\rho^2)^2}{4 (1+{\varepsilon})} + \frac{c^2}{4} (\partial_{x_1} \theta_n)^2(1+{\varepsilon})$ with suitable ${\varepsilon}>0$ to obtain:
$$\int_{\{\rho_n \geq r\}} l(\rho_n, \theta_n) \geq \int_{\{\rho_n \geq r \} } \frac 1 2 |\nabla \rho_n|^2 + \left[ \frac 1 2 - \frac{c^2 (1+{\varepsilon})}{4\rho_n^2} \right] |\nabla \theta_n|^2\rho_n^2 + \frac{{\varepsilon}}{1+{\varepsilon}} \frac{(1-\rho_n^2)^2}{4}$$ $$\geq {\varepsilon}_0 \int_{\{\rho_n \geq r \}} e(\rho_n, \theta_n) = {\varepsilon}_0 \, \mathcal{E}(\psi_n) + O(1),$$
for suitable ${\varepsilon}_0>0$. Then,
$$O(1) = I^{c_n}(\psi_n) \geq {\varepsilon}_0 \, \mathcal{E}(\psi_n) + O(1),$$ and this allows us to conclude.
Proof of Theorem \[teo:3\]
--------------------------
With all the results above we can inmediately conclude the proof of Theorem \[teo:3\]. Indeed, by and Sobolev inequality, we have that
$$\int_{{{\mathbb R}}^3} (1-\rho_n)^6 = O(1).$$
If $\mathcal{E}(\psi_n) \to +\infty$, Proposition \[prop:ma\] implies $|S_n^r|$ is unbounded for $r > \frac{c}{\sqrt{2}}$, and this is a contradiction with the above estimate. Hence $\mathcal{E}(\psi_n)$ is bounded. By Fatou Lemma, the solution $\psi_0$ given in Proposition \[ascoli\] has finite energy, concluding the proof.
Proof of Theorem \[teo:all\]
============================
In this section we prove the compactness criterion given in Theorem \[teo:all\]. The proof follows some of the ideas of the previous section, but with important differences. As previously, we will be done if we show that $\mathcal{E}(\psi_n)$ is bounded.
By , we have that $\psi_n \neq 0$ outside $B(0,R)$; as a consequence, $\psi_n$ admit a lifting $\psi_n(x) = \rho_n(x) e^{i \theta_n(x)}$ for all $x \in {{\mathbb R}}^2 \setminus B(0,R)$. This is a consequence of the fact that $\psi_n$ have finite energy, see [@gravejat-AIHP]\[Lemma 15\]. In the vortexless case, this lifting holds in the whole euclidean space.
Next lemma is a version of Lemma \[lem:O(1)\]:
\[jopeta\] Take $c \in (0,\ \sqrt{2})$, $c_n \in E$ with $c_n \to c$ and $\psi_n$ the solutions given by Theorem \[teo:almost\]. Assume also that there exists $R>0$ and $\delta>0$ such that is satisfied. Then $$\label{eq:stimmomass}
\mathcal{P}(\psi_n)=\frac 12 \int_{B(0,R)^c}(1-\rho_n^2)\partial_{x_1}\theta_n +O(1),$$ $$\label{eq:stimmomass2}
c \mathcal{P}(\psi_n)= \int_{B(0,R)^c}\rho_n^2|\nabla \theta_n|^2 +O(1),$$ $$\label{eq:stimmomass3}
\int_{B(0,R)^c}|\nabla \rho_n|^2 =O(1).$$
The proof is completely analogue to that of Lemma \[lem:O(1)\]. Observe that in the vortexless case we have exact identities in , , .
.
With Lemma \[lem:O(1)\] in hand, we can adapt the proof of Proposition \[prop:ma\] to our setting, obtaining the following result:
\[prop:ma2\] Take $c \in (0,\ \sqrt{2})$, $c_n \in E$ with $c_n \to c$ and $\psi_n$ the solutions given by Theorem \[teo:almost\]. Assume that:
$$\mathcal{E}(\psi_n) \to + \infty.$$
Then, for any $r \in (\frac{c}{\sqrt{2}}, 1)$, $|S_n^r| \to +\infty$.
Next result is analogue to Lemma \[lem:51\]. The only difference is that now we do not know that holds, but instead we have .
\[lem:61\] Under the assumptions of Theorem \[teo:all\], assume that for some $r\in (0, 1)$, $|S_n^{r}| \to +\infty$. Then, there exists $\xi_n \in S_n^r$ and $R_n \to +\infty$ such that:
$$\int_{B(\xi_n, R_n)} |\nabla \rho_n |^2 + |\partial_{x_2} \psi_n |^2 \to 0.$$
The proof is analogue to that of Lemma \[lem:51\].
Proof of Theorem \[teo:all\]
----------------------------
Assume by contradiction that $\mathcal{E}(\psi_n) \to +\infty$. By Proposition \[prop:ma2\], we can apply Lemma \[lem:61\] to a value $r$ satisfying that:
$$\frac{c}{\sqrt{2}} < r < \sqrt{ \frac{2}{3} (1+c^2/4)} <1.$$ Notice that this is possible if $c<\sqrt{2}$.
Let $\xi_n \in {{\mathbb R}}^d$ given by Lemma \[lem:61\] and define $\tilde{\psi}_n(x)= \psi_n(x- \xi_n)$. Up to a subsequence we have that:
$$\tilde{\psi}_n \to {\psi}_0 \mbox{ in } C^k_{loc}({{\mathbb R}}^d).$$
Taking into account Remark \[remark morse\], $ind(\psi_0) \leq 1$. By Lemma \[lem:51\], $\psi_0$ depends only of the $x_1$ variable. Moreover $\nabla \rho_0 =0$ where $\rho_0 = |{\psi}_0| \leq r$. That is, $\psi_0 (x_1)$ is a $1D$ circular solution,
$$\psi_0(x_1) = \rho_0 e^{i \omega (x_1 - t)},$$ where $\omega^2 + c \omega + \rho_0^2=1$. By the choice of $r$, we have that $\rho_0^2 < \frac{2}{3} (1+c^2/4)$. But those solutions have infinite Morse index, as shown in Proposition \[appendix\] (see Appendix). This contradiction shows that $\mathcal{E}(\psi_n)$ is bounded.
By Fatou Lemma, the solution $\psi_0$ given in Proposition \[ascoli\] has finite energy, concluding the proof.
Let us point out that Theorem \[teo:3\] does not need the information on the Morse index of the solutions. The main tool there is that $I^{c_n}(\psi_n) = O(1)$. Instead, Theorem \[teo:all\] requires in a essential way that the Morse index of the solutions obtained is bounded.
Appendix (by Rafael Ortega)
===========================
In this appendix we prove the following result:
\[appendix\] Given $t \in {{\mathbb R}}$, $\omega_0 \in {{\mathbb R}}$, $\rho_0 >0$ satisfying that $\omega_0^2 + c \omega_0 + \rho_0^2=1$, the function $\psi_0(x) = \rho_0 e^{i \omega_0 (x - t)}$ is a (infinite energy) solution of . Assume also that $\rho_0^2 < \frac{2}{3} (1+c^2/4)$. Then its Morse index, as defined in Definition \[Morse\], is infinity.
The problem is autonomous so that we can assume $t=0$. The proof is based on the study of the $1D$ problem:
$$\label{eq:GP1D}
\psi'' + i c \psi' + \left(1-|\psi|^2\right)\psi=0 \ \ \text{ on } {{\mathbb R}}.$$
By the change of variables $\phi= e^{i x c/2 } \psi$ we pass to a problem:
$$\label{eq:GP1Dbis}
\phi'' +\left(1+ c^2/4 -|\phi|^2\right)\phi=0 \ \ \text{ on } {{\mathbb R}}.$$
The Morse index of this problem depends on the existence of conjugate points to some solutions of the linearized equation, see for instance [@gelfand Chapter 5]. The function $\phi(x)= \rho_0 e^{i \omega_1 x}$ is a solution of , where , $\omega_1 = \omega_0 + c/2$. Observe that
$$\label{relation}
\omega_1^2 + \rho_0^2 = 1 + c^2/4$$
The linearized equation to around the solution $\phi$ is:
$$\zeta'' + (1+c^2/4) \zeta - 2 \overline{\phi(s)} \phi(s) \zeta - \phi(s)^2 \overline{\zeta}=0.$$
We will follow the lines of [@SM Section 21] to analyze the oscillatory properties of this equation.
$$\zeta'' + (1+c^2/4) \zeta - 2 \rho_0^2 \zeta - \rho_0^2 e^{2 i \omega_1 s }\overline{\zeta}=0.$$
We now make the change of variable $\zeta = e^{i \omega_1 s} \eta$, to obtain a constant coefficient linear system:
$$\label{constant}
\eta'' + 2 i \omega_1 \eta' - \rho_0^2 \eta - \rho_0^2 \overline{\eta}=0.$$
If $\rho_0^2 < 2 \omega_1^2$ (which, by , reduces to $\rho_0^2 < \frac{2}{3}(1+c^2/4)$) we can find the explicit solution to :
$$\eta(s)= \frac{ \sin \left( s \sqrt{4 \omega_1^2 - 2\rho_0^2} \right) }{\sqrt{4 \omega_1^2 - 2\rho_0^2}} + i \omega_1 \frac{ \cos \left( s \sqrt{4 \omega_1^2 - 2\rho_0^2} \right) -1}{2 \omega_1^2 - \rho_0^2}.$$
Clearly, $\zeta(s)= e^{i \omega_1 s} \eta(s)$ has infinitely many conjugate points $\frac{2 \pi n}{\sqrt{4 \omega_1^2 - 2\rho_0^2}}$, $n \in {{\mathbb N}}$.
Given any interval $I$, the quadratic functional $\tilde{Q}_{1,I}: H_0^1(I, \mathbb{C}) \to {{\mathbb R}}$,
$$\tilde{Q}_{1, I}(\sigma_k)= \int_{I} |\sigma_k'|^2 - (1+c^2/4 - |\phi|^2) |\sigma_k|^2 +2 ({\langle}\sigma_k, \phi {\rangle})^2 <0$$ is in the conditions of Section 29.2 of [@gelfand]. We can apply [@gelfand Theorem 3’ in page 122] to deduce that $\tilde{Q}_{1, I}$ takes negative values as soon as the length of the interval $I$ is greater than $\frac{2 \pi }{\sqrt{4 \omega_1^2 - 2\rho_0^2}}$. Then we can find infinitely many functions $\sigma_k \in C^{\infty}_0({{\mathbb R}})$ *with disjoint support* such that
$$\tilde{Q}_1(\sigma_k)= \int_{-\infty}^{\infty} |\sigma_k'|^2 - (1+c^2/4 - |\phi|^2) |\sigma_k|^2 +2 ({\langle}\sigma_k, \phi {\rangle})^2 <0.$$
We now want to pass to the original problem and estimate its Morse index. In order to do so, define $\tau_k(s)$ by $\sigma_k(s)= e^{ics/2}\tau_k(s)$. Simple computations give:
$$\sigma_k'(s)= (i \frac c 2 \tau_k(s) + \tau_k'(s) )e^{ics/2},$$
$$|\sigma_k'(s)|^2 = |\tau_k'(s)|^2 + \frac{c^2}{4} |\tau_k(s)|^2 + c {\langle}i \tau_k(s), \tau_k'(s) {\rangle}= |\tau_k'(s)|^2 + \frac{c^2}{4} |\tau_k(s)|^2 - c {\langle}\tau_k(s), i \tau_k'(s) {\rangle}.$$
Moreover,
$${\langle}\sigma_k(s), \phi(s) {\rangle}= {\langle}\tau_k(s), \psi(s) {\rangle}.$$
As a consequence $\tilde{Q}_1(\sigma_k)= Q_1(\tau_k)<0$, where
$$Q_1(\tau_k)=\int_{-\infty}^{\infty} |\tau_k'|^2 - c {\langle}\tau_k, i \tau_k' {\rangle}- (1 - |\psi|^2) |\tau_k|^2 +2 ({\langle}\tau_k, \psi {\rangle})^2.$$
Observe that this is the quadratic form associated to .
Take now a $C_0^\infty$ function $\chi_k: {{\mathbb R}}^{d-1} \to {{\mathbb R}}^+$, and let us estimate $Q$ on the function $\iota_k(x)=\chi_k(\tilde{x}) \tau_k(x_1)$, where $Q$ is defined in :
$$Q(\iota_k)= Q_1(\tau_k) \int_{{{\mathbb R}}^{d-1}} \chi_k(\tilde{x})^2 \, d\tilde{x} + \left( \int_{-\infty}^{+ \infty} |\tau_k(x_1)|^2 \, dx_1\right)
\left( \int_{{{\mathbb R}}^{d-1}} |\nabla \chi_k(\tilde{x})|^2 \, d\tilde{x} \right).$$
It suffices to take now $\chi_k$ such that $\int_{{{\mathbb R}}^{d-1}} \chi_k^2=1$ and $\int_{{{\mathbb R}}^{d-1}} |\nabla \chi_k|^2 $ is sufficiently small, to conclude that $Q(\iota_k )<0$.
Observe also that $supp \ \iota_k \cap supp \ \iota_{k'} = \emptyset$ if $k \neq k'$, since an analogue property holds for $\sigma_k$ and $\tau_k$. Hence, $Q$ is negative definite on the vector space generated by the linearly independent functions $\{\iota_1, \ \dots, \iota_k\}$ for any $k \in {{\mathbb N}}$, concluding the proof.
[99]{} A. Ambrosetti, A. Malchiodi, [Nonlinear analysis and semilinear elliptic problems,]{} Cambridge Studies in Advanced Mathematics, 104. Cambridge University Press, Cambridge, 2007.
, preprint , preprint
, [Vortex rings for the GrossÐ-Pitaevskii equation in $R^3$, ]{}[J. Math. Pures Appl. 100 (2013) 69Ð112]{}
, arXiv:1804.09875 (2018)
M. Mariş, [Traveling waves for nonlinear Schrödinger equations with nonzero conditions at infinity]{}, Ann. of Math. (2) 178 (2013), no. 1, 107Ð-182.
M. Mariş, [Nonexistence of supersonic traveling waves for nonlinear Schrodinger equations with nonzero conditions at infinity, ]{}[SIAM J. Math. Anal. 40 (2008), 1076-1103.]{}
[^1]: J. B. is partially supported by Project 2016 ÒDinamica di equazioni nonlineari dispersiveÓ of FONDAZIONE DI SARDEGNA. D. R. has been supported by the FEDER-MINECO Grants MTM2015-68210-P and PGC2018-096422-B-I00, and by J. Andalucia (FQM-116).
| {
"pile_set_name": "ArXiv"
} |
---
author:
- 'Javier M. Magán'
- and Joan Simón
bibliography:
- 'OpG-2.bib'
title: On operator growth and emergent Poincaré symmetries
---
Introduction {#sec:Introduction}
============
In the Schrödinger picture of quantum mechanics, unitary evolution mixes the initial state $\vert\psi \rangle$ with other quantum states as time evolves $$\label{schr}
\vert\psi (t)\rangle = e^{-iHt}\,\vert\psi \rangle =\sum_{n=0}^{\infty}\frac{(-iHt)^{n}}{n!}\,\vert\psi \rangle\equiv \sum_{n=0}^{\infty}\frac{(-it)^{n}}{n!}\,\vert\psi_{n} \rangle\;.$$ Hence, solving for the time evolution amounts to understanding the states $\vert\psi_{n} \rangle \equiv H^{n}\vert\psi \rangle$, an understanding which is definitely challenging for chaotic Hamiltonians. Similarly, in the Heisenberg picture, unitary evolution mixes the initial operator $\mathcal{O}$ with other operators according to $$\label{Heis}
\mathcal{O}(t)=e^{iHt}\,\mathcal{O}\, e^{-iHt}=\sum\limits_{n=0}^{\infty}\frac{(it)^{n}}{n!}\,[H,\cdots,[H,\mathcal{O}]\cdots]\equiv \sum\limits_{n=0}^{\infty}\frac{(it)^{n}}{n!}\,\mathcal{O}_{n}\;.$$ In this case, it is the understanding of the operators $\mathcal{O}_{n} \equiv [H,\cdots,[H,\mathcal{O}]\cdots]$ that allows to solve for the time evolution, and ultimately determines any notion of *operator growth* one might potentially define.
The structure of Heisenberg’s time evolution in chaotic systems has attracted some recent interest for several reasons. First, due to its expected connection to quantum chaos. It was found in [@Roberts:2018mnp; @Qi:2018bje], for the case of SYK [@Sachdev:1992fk; @Kitaev], that certain notion of operator size, to be reviewed in the main text, is related to out-of-time-ordered correlation (OTOC) functions [@1969JETP...28.1200L; @Shenker:2014cwa; @Maldacena:2015waa]. Second, because of the relation between operator growth, quantum complexity and the emergence of near horizon symmetries [@Susskind:2018tei; @Magan:2018nmu; @Lin:2019qwu; @Barbon:2019tuq]. Finally, due to the broader connection between complexity and operator growth, as discussed from different perspectives in [@Magan:2018nmu; @Parker:2018yvk; @Barbon:2019wsy; @Bueno2019], such as using Nielsen’s geometric approach to quantum circuit complexity [@2005quant.ph..2070N; @2007quant.ph..1004D] or the recursion method in many-body physics [@viswanath2008recursion].[^1]
The main goal of this work is twofold. First and most important, to discuss the structure of the operators $\mathcal{O}_{n}$ in large-N theories, broadly understood as whenever large-N factorization holds [@tHooft:1973alw], which include large-N holographic theories [@ElShowk:2011ag; @Papadodimas:2012aq; @Papadodimas:2013wnh; @Papadodimas:2013jku; @deBoer:2019kyr]. Second, to revisit some of the existent approaches to operator growth, apply them to large-N theories in the light of our previous analysis, and compare them with quantum circuit complexity and quantum chaos. Within the context of AdS/CFT [@Maldacena:1997re; @Gubser:1998bc; @Witten:1998qj], the present approach, stemming on the analysis of the operators $\mathcal{O}_{n}$, evades the bulk vs boundary dichotomy, and makes manifest that *any* notion of operator growth is the same at both sides of the duality, given the equivalence of Hilbert spaces, operator algebras, and Heisenberg time evolutions.
Albeit the first objective might seem a hopeless task, given the inherent complexity of a chaotic Hamiltonian, we will show how large-N factorization and generic finite temperature properties in relativistic QFTs completely determine the action of the operators $\mathcal{O}_{n}$ in most of the relevant states of the theory, at least if the Eigenstate Thermalization Hypothesis (ETH) [@PhysRevE.50.888] holds. Hence, we conclude that any notion of operator growth is determined by the behavior of the 2-pt functions alone, at leading order in the large-N limit.[^2]
As a byproduct of this discussion, we use the operators $\mathcal{O}_{n}$ in holographic theories to construct boundary CFT operators closing the *bulk* Poincaré algebra. Our construction is analogous to how the Poincaré algebra in free QFT is generated from the algebra of operators in the two Rindler wedges. It also uses the notion of mirror operators introduced in the context of holographic bulk reconstruction [@Papadodimas:2012aq; @Papadodimas:2013wnh; @Papadodimas:2013jku]. This emergent Poincaré algebra controls aspects of bulk infalling physics.
As for the second objective, we first note that quantum systems may have different notions of operator size depending on their nature and dynamics. However, all of them can be formulated as expectation values of simple operators within the Gelfand-Naimark-Segal (GNS) construction. The construction associates a Hilbert space to an algebra of operators, and maps Heisenberg’s evolution to Schrödinger’s evolution in the GNS Hilbert space. This allows extending notions of operator growth to QFT, and large-N theories in particular. As we will see, whenever large-N factorization holds, operator growth will be determined by 2-pt functions.
We then comment on natural size operators in these theories, mainly energy and number operators, stressing the exponential growth of the proper energy operator [@Magan:2018nmu]. We also study the recursion method [@viswanath2008recursion], typically used in many-body physics, in large-N theories. We find a closed solution for the basis of orthonormalized operators that solves the 1d diffusion equation that appears in this approach. Finally, we describe the relation between Nielsen’s quantum circuit complexity, operator growth, and chaos when one compares the complexity of formation between a pair of time evolved target states differing by an initial small perturbation.
This work is organized as follows. In section \[sec:structure\], we discuss the operators $\mathcal{O}_{n}$ in large-N theories. In section \[sec:algebra\], we first review the construction of the Poincaré algebra using the algebra of operators in the Rindler wedges in section \[sec:rindler\]. We then apply an analogous construction to holographic large-N theories in section \[sec:GFFtoPoincare\]. In section \[sec:GNS\], we reformulate different notions of operator size in the literature using the GNS construction reviewed in appendix \[App2\]. We discuss natural notions of operator growth and the many-body recursion method in large-N theories in sections \[sec:size-N\] and \[sec:recursion\], respectively. We close with a discussion connecting quantum circuit complexity and operator growth in section \[sec:chaos\]. A summary of our results and the logic purposed in this work are given in section \[Discussion\].
Operator evolution in large-N theories {#sec:structure}
======================================
The goal of this section is to evaluate the series of nested commutators $$\label{lno}
\mathcal{O}_{n}\equiv [H,\cdots,[H,\mathcal{O}]\cdots]\;.$$ controlling the Heisenberg time evolution in generic large-N gauge theories for a subset of initial operators $\mathcal{O}$ where large-N factorization of correlation functions holds [@tHooft:1973alw].
Consider a local gauge-invariant scalar operator[^3] $\mathcal{O}(t,\vec{x})$ in a gauge theory defined on $\mathbb{R}^{1,d-1}$. Its Fourier decomposition $$\mathcal{O}(t,\vec{x}) = \int_{\omega > 0} \frac{d\omega d^{d-1}\vec{k}}{(2\pi)^d} \left(\mathcal{O}_{\omega,\vec{k}}\,e^{-i\omega t + i\vec{k}\vec{x}} + \mathcal{O}^\dagger_{\omega,\vec{k}}\,e^{i\omega t - i\vec{k}\vec{x}}\right)\,,
\label{exparep}$$ defines the non-local Fourier mode operators $\mathcal{O}_{\omega,\vec{k}}$ as $$\mathcal{O}_{\omega,\vec{k}}=\int dt\,d^{d-1}\vec{x} \,\mathcal{O}(t,\vec{x})\,e^{i\omega t-i\vec{k}\vec{x}} \,,$$ with a similar expression for $\mathcal{O}^\dagger_{\omega,\vec{k}}$.
Even though the energy $\omega$ and momentum $\vec{k}$ labels are *not* related to each other by means of a dispersion relation, as it occurs for free quantum fields due to the classical equation of motion, the time evolution of $\mathcal{O}(t,\vec{x})$ remains trivial, as in the latter case, in this Fourier basis. Indeed, the nested commutators equal $$\label{no}
\mathcal{O}_{n}(t)= i^{-n}\frac{d^{n}}{dt^{n}}\mathcal{O}(t)=
\int\limits_{\omega > 0} \frac{d\omega d^{d-1}\vec{k}}{(2\pi)^d} \left((-\omega)^n\,\mathcal{O}_{\omega,\vec{k}}\,e^{-i\omega t + i\vec{k}\vec{x}} + \omega^n\,\mathcal{O}^\dagger_{\omega,\vec{k}}\,e^{i\omega t - i\vec{k}\vec{x}}\right)\,.$$ In particular, each mode satisfies $$[H,\mathcal{O}_{\omega,\vec{k}}]=-\omega\,\mathcal{O}_{\omega,\vec{k}} \quad \Longrightarrow \quad \mathcal{O}_{\omega,\vec{k}} (t)=e^{-i\omega t} \,\mathcal{O}_{\omega,\vec{k}}\,,$$ and similarly for $\mathcal{O}_{\omega\vec{k}}^{\dagger}$. Hence, these Fourier mode operators do not mix with other operators as time evolves. For this reason, they are specially suited to study operator growth.
The Fourier decomposition trades the problem of understanding operator growth, characterized by the operators $\mathcal{O}_{n}$, for the one of understanding $\mathcal{O}_{\omega,\vec{k}}$ and $\mathcal{O}_{\omega\vec{k}}^{\dagger}$. However, both these operators are well understood in theories admitting a large-N expansion, within the regime where factorization of higher point functions holds. In particular, working at finite temperature, large-N factorization of thermal higher point correlation functions guarantees that finite temperature dynamics is determined by the 2-pt functions $$\label{eq:exp}
\begin{aligned}
Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\mathcal{O}_{\omega\vec{k}}\,\mathcal{O}_{\omega^\prime\vec{k}^\prime}\right) &= Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\mathcal{O}^\dagger_{\omega\vec{k}}\,\mathcal{O}^\dagger_{\omega^\prime\vec{k}^\prime}\right) = 0\,,\\
Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\mathcal{O}_{\omega\vec{k}}\,\mathcal{O}^\dagger_{\omega^\prime\vec{k}^\prime}\right)&= G_\beta(\omega,\vec{k})\,\delta(\omega-\omega^\prime)\delta^{d-1}(\vec{k}-\vec{k}^\prime)\,,\\
Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\, \mathcal{O}_{\omega\vec{k}}^{\dagger}\,\mathcal{O}_{\omega^\prime\vec{k}^\prime}\right) &=G_\beta(-\omega,-\vec{k})\,\delta(\omega-\omega^\prime)\delta^{d-1}(\vec{k}-\vec{k}^\prime)
\end{aligned}$$ where $G_\beta(\omega,\vec{k})$ is the Fourier transform of the 2-pt function $$G_\beta(\omega,\vec{k}) \equiv \int dt\,d^{d-1}\vec{x}\, G_\beta(t,\vec{x})\,e^{i\omega t -i\vec{k}\vec{x}} \equiv\,Z_\beta^{-1}\,\int dt\,d^{d-1}\vec{x}\,\text{Tr}\left(e^{-\beta H}\,\mathcal{O}(t,\vec{x})\,\mathcal{O}(0,\vec{0})\right)\,e^{i\omega t -i\vec{k}\vec{x}}\,.$$ It follows from that the commutators of the Fourier mode operators are given by $$\begin{aligned}
Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\left[\mathcal{O}^\dagger_{\omega,\vec{k}}\,,\mathcal{O}^\dagger_{\omega^\prime,\vec{k}^\prime}\right]\right) &= Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\left[\mathcal{O}_{\omega,\vec{k}}\,,\mathcal{O}_{\omega^\prime,\vec{k}^\prime}\right]\right) = 0\,, \\
Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\left[\mathcal{O}_{\omega,\vec{k}}\,,\mathcal{O}^\dagger_{\omega^\prime,\vec{k}^\prime}\right]\right) &= \left(G_\beta(\omega,\vec{k})-G_\beta(-\omega,-\vec{k})\right)\,\delta(\omega-\omega^\prime)\delta^{(d-1)}(\vec{k}-\vec{k}^\prime)\,,
\end{aligned}
\label{eq:N2p}$$ up to $1/N$ corrections. These correlators assume that energy-momentum labels do not scale with $N$. Those high energy modes are not needed for the following reason. Local operators have an infinite amount of energy, as it is typical in QFT. They contain modes up to infinite $\omega$ but they are not well defined operators. The actual operators we should consider are smeared versions of these $$\mathcal{O}=\int dt d^{d-1}\vec{x} f(x,t)\mathcal{O}(t,\vec{x}) = \int_{\omega > 0} \frac{d\omega d^{d-1}\vec{k}}{(2\pi)^d} \,\left(\tilde{f}(\omega,\vec{k})\,\mathcal{O}_{\omega,\vec{k}} + \tilde{f}^{*}(\omega,\vec{k})\,\mathcal{O}^\dagger_{\omega,\vec{k}}\right)\;,$$ where $$\tilde{f}(\omega,\vec{k})\equiv\int dt\,d^{d-1}\vec{x} \,f(t,\vec{x})\,e^{i\omega t-i\vec{k}\vec{x}} \,.$$ To have a well defined (finite energy) operator $\mathcal{O}$ in the large-N limit we have to consider a smearing function $f(t,\vec{x})$ whose smoothness properties do not blow up in the limit. The Fourier transform of such function will exponentially suppress the modes $\mathcal{O}_{\omega,\vec{k}}$ with frequencies and wavelengths scaling with $N$ in the large-N limit. This smearing condition, together with large-N factorization, provides a good and precise definition of a “simple” operator in large-N QFT’s.[^4]
Coming back to the algebra of the modes, as stressed in [@ElShowk:2011ag] and further developed in [@Papadodimas:2012aq], due to large-N factorization, the same statements hold when inserting operators $\mathcal{P}_1$ and $\mathcal{P}_2$ $$\begin{aligned}
Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\mathcal{P}_1\left[\mathcal{O}^\dagger_{\omega,\vec{k}}\,,\mathcal{O}^\dagger_{\omega^\prime,\vec{k}^\prime}\right] \mathcal{P}_2\right) &= Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\mathcal{P}_1 \left[\mathcal{O}_{\omega,\vec{k}}\,,\mathcal{O}_{\omega^\prime,\vec{k}^\prime}\right] \mathcal{P}_2\right) = 0\,, \\
Z^{-1}_\beta \text{Tr}\left(e^{-\beta H}\,\mathcal{P}_1\left[\mathcal{O}_{\omega,\vec{k}}\,,\mathcal{O}^\dagger_{\omega^\prime,\vec{k}^\prime}\right] \mathcal{P}_2\right) &= \left(G_\beta(\omega,\vec{k})-G_\beta(-\omega,-\vec{k})\right)\,\delta(\omega-\omega^\prime)\delta^{(d-1)}(\vec{k}-\vec{k}^\prime)\cdot \\
&\cdot \,\text{Tr}\left(e^{-\beta H}\,\mathcal{P}_1\mathcal{P}_2\right)
\label{eq:N2pa}
\end{aligned}$$ involving a number of legs not scaling with $N$. Therefore, these commutators behave as c-numbers when inserted in correlation functions within this regime, a characteristic feature of free fields.
Correlators and , together with linearity, allow us to evaluate the expectation values of the nested commutators (\[no\]) in the thermal ensemble at any temperature. Furthermore, for the subset of large-N gauge theories satisfying the Eigenstate Thermalization Hypothesis (ETH) [@PhysRevE.50.888], the set of states in the Hilbert space where the previous correlation functions hold is much larger since ETH ensures the same expectation values apply to most energy eigenstates compatible with the physical temperature. Therefore, in the large-N limit, the previous equations define the action of the operators $\mathcal{O}_{\omega,\vec{k}}$ and consequently of the $\mathcal{O}_{n}$ in the basis of eigenstates of the theory, up to a set of atypical energy eigenstates.[^5] Notice that knowledge of this action and expectation values is enough to define the operators in this limit.[^6]
Before ending this section, it is worth making a couple of closing remarks. First, the “growth” of the operator in space, as defined by the nested commutators $$\begin{aligned}
\left[P^j,\cdots,\left[P^j,\mathcal{O}(t,\vec{x})\right] \cdots\right] & = (-i)^{n}\frac{d^{n}}{dx_{j}^{n}} \mathcal{O}(t,\vec{x}) \\
&= \int_{\omega > 0} \frac{d\omega d^{d-1}\vec{k}}{(2\pi)^d} \left[k_{j}^{n}\mathcal{O}_{\omega,\vec{k}}\,e^{-i\omega t + i\vec{k}\vec{x}} + (-k_{j})^{n}\mathcal{O}^\dagger_{\omega,\vec{k}}\,e^{i\omega t - i\vec{k}\vec{x}}\right]\,.
\end{aligned}
\label{pc}$$ is also determined by the same 2-pt functions above. Second, in the large-$N$ limit we are working with, the generators of time $(H)$ and space $(\vec{P})$ translations effectively reduce to $$H=\int\limits_{\omega> 0} \frac{d\omega d^{d-1}\vec{k}}{(2\pi)^d}\,\omega\,\mathcal{O}^{\dagger}_{\omega,\vec{k}}\mathcal{O}_{\omega,\vec{k}}\, \quad \text{and} \quad
\vec{P}=\int\limits_{\omega> 0} \frac{d\omega d^{d-1}\vec{k}}{(2\pi)^d}\, \vec{k}\,\mathcal{O}^{\dagger}_{\omega,\vec{k}}\mathcal{O}_{\omega,\vec{k}}\,.$$
#### Modular time evolution.
Besides unitary time evolution, there is a second natural notion of evolution, modular time evolution, when restricting physics to subregions of spacetime (see [@Haag:1992hx] for a review). Modular time evolution is defined as the unitary evolution generated by the modular hamiltonian $H_{\textrm{mod}}$ in the region of interest $$\mathcal{O}(s)\equiv e^{isH_{\textrm{mod}}}\mathcal{O} e^{-isH_{\textrm{mod}}}=\rho^{-is}\mathcal{O}\rho^{is}\,,$$ where $H_{\textrm{mod}}$ is related to the reduced density matrix[^7] in this region $\rho$, as $\rho=e^{-H_{\textrm{mod}}}$.
It is interesting to ask for the structure of operator evolution in this context.[^8] Since $$\mathcal{O}(s)=e^{isH_{\textrm{mod}}}\,\mathcal{O}\, e^{-isH_{\textrm{mod}}}=\sum\limits_{n=0}^{\infty}\frac{(is)^{n}}{n!}\,[H_{\textrm{mod}},\cdots,[H_{\textrm{mod}},\mathcal{O}]\cdots]\equiv \sum\limits_{n=0}^{\infty}\frac{(is)^{n}}{n!}\,\mathcal{O}^{\textrm{mod}}_{n}\;,$$ this structure is determined by the operators $$\mathcal{O}^{\textrm{mod}}_{n}\equiv [H_{\textrm{mod}},\cdots,[H_{\textrm{mod}},\mathcal{O}]\cdots]\;,$$ which are the only ones with whom the initial operator mixes through modular time evolution. Proceeding as before, we can Fourier transform the fields, but against modular time evolution $$\label{modular}
\mathcal{O}_{\omega}=\int ds\, \,\mathcal{O}(s)\,e^{i\omega s}$$ This allows for a simple computation of the action of $\mathcal{O}^{\textrm{mod}}_{n}$ in all eigenstates of the theory, in the vein of (\[no\]), if the correlators of the modular field modes are gaussian, as expected for holographic theories with free bulk duals.
Emergent Poincaré algebra in holographic theories {#sec:algebra}
=================================================
Previous observations hold in large-N holographic theories. The dual bulk description of the CFT at finite temperature is a black hole [@Witten:1998zw] and the spectrum of low energy excitations is given by a small (not scaling with $N$) number of generalized free fields whose Fourier modes satisfy similar commutation relations to the ones in (see below for a more precise account). Since the geometry near the black hole horizon is locally equivalent to a Rindler horizon, we can ask whether the exact relations between infalling and uniformly accelerated observers in free QFT in Rindler will continue to hold at the lowest order in a $1/N$ expansion in the holographic set-up.
In particular, free QFT in Rindler is Poincaré invariant. Motivated by the black hole scenario, we can ask how to build operators closing an exact Poincaré algebra out of the algebra of operators existing in the left and right Rindler wedges. We will review the construction of these operators in section \[sec:rindler\]. Since the latter only relies on the algebra satisfied by the Rindler creation/annihilation operators and this is the same algebra satisfied by the Fourier modes of the generalized free fields reconstructing the bulk excitations in the boundary CFT, it follows that applying the same Rindler construction to large-N holographic theories will give rise to boundary CFT operators closing the *bulk* Poincaré algebra at lowest order in a $1/N$ expansion. The implications of this construction for operator growth will be discussed in the next section.
From Rindler to Poincaré {#sec:rindler}
------------------------
Consider a free quantum scalar field of mass $m$ in $\mathbb{R}^{1,d}$. We want to review how to construct the relevant generators of the Poincaré algebra starting from the scalar description in both Rindler patches. To set some notation, let us decompose the space directions into $z$ and $\vec{x}$, with conjugate momentum $k_z$ and $\vec{k}$, respectively, so that $z$ corresponds to the direction along which the Rindler observer is uniformly accelerated. Local fields in the right Rindler patch can be expanded in terms of creation and annihilation operators $${\hat{a}^{\text{R}}_{\omega\vec{k}}}\,\,\,\,,\,\,\, {\hat{a}^{\text{R}\dagger}_{\omega\vec{k}}}$$ satisfying the standard commutation relations $$\left[{\hat{a}^{\text{R}}_{\omega\vec{k}}},\,{\hat{a}^{R\dagger}_{\omega^\prime\vec{k}^\prime}}\right] = \delta(\omega-\omega^\prime)\delta^{(d-1)}(\vec{k}-\vec{k}^\prime)\,.
\label{eq:right-rindler}$$ Since $\omega$ stands for the Rindler energy, time evolution in this Rindler wedge is generated by the operator $$\hat{H}^{\textrm{R}}=\int\limits d\omega\,d\vec{k}\, \omega\, {\hat{a}^{\text{R}\dagger}_{\omega\vec{k}}}{\hat{a}^{\text{R}}_{\omega\vec{k}}}\,.$$ It follows that each operator ${\hat{a}^{\text{R}}_{\omega\vec{k}}}$ evolves as $$[\hat{H}^{\textrm{R}},{\hat{a}^{\text{R}}_{\omega\vec{k}}}]= -\omega\,{\hat{a}^{\text{R}}_{\omega\vec{k}}}\quad \Rightarrow \quad {\hat{a}^{\text{R}}_{\omega\vec{k}}}(t)= e^{-i\omega t}\,{\hat{a}^{\text{R}}_{\omega\vec{k}}}$$ as usual in free quantum field theory, with a similar expression for ${\hat{a}^{\text{R}\dagger}_{\omega\vec{k}}}(t)$. Linearity extends these claims to local quantum fields. Notice Rindler time is labelled as $t$. Minkowski operators and coordinates will carry an $M$ superscript.
Consider a thermal state $\rho_{\beta}$ in the right Rindler wedge quantum field theory (QFT) with $\beta=\frac{2\pi}{a}$, where $a$ stands for the proper acceleration defining the Rindler frame. It is well known the latter is the reduced density matrix of the pure state $\vert 0_{\text{M}}\rangle$, i.e. the vacuum of the QFT in Minkowski (see [@Crispino:2007eb] for a review and more details about QFT in Rindler space). This purification involves duplication of the operator algebra giving rise to a new set of operators $${\hat{a}^{\text{L}}_{\omega\vec{k}}}\,\,\,\,,\,\,\, {\hat{a}^{\text{L}\dagger}_{\omega\vec{k}}}\label{eq:left-rindler}$$ which correspond to the creation and annihilation operators in the left Rindler wedge from the perspective of the full Minkowski spacetime. These operators commute with the original ones and altogether form a complete basis of operators in the Minkowski Hilbert space. This matches the general Gelfand-Naimark-Segal (GNS) representation of thermal states reviewed in appendix \[App2\], for later convenience. In this appendix, the origin of the duplication of the algebra responsible for the canonical purification of the thermal Rindler state $\rho_\beta$ in the current discussion is explained.
The Minkowski vacuum $|0_{\text{M}}\rangle$ is determined by the relations $$\left(\hat{a}^{\text{L}}_{\omega\vec{k}} - e^{-\pi\frac{\omega}{a}}\,\hat{a}^{{\text{R}}\dagger}_{\omega\,(-\vec{k})}\right)|0_{\text{M}}\rangle = \left(\hat{a}^{\text{R}}_{\omega\vec{k}} - e^{-\pi\frac{\omega}{a}}\,\hat{a}^{{\text{L}}\dagger}_{\omega\,(-\vec{k})}\right) |0_{\text{M}}\rangle = 0\,.
\label{trind}$$ These allow to write the action of both ${\hat{a}^{\text{L}}_{\omega\vec{k}}}$ and ${\hat{a}^{\text{L}\dagger}_{\omega\vec{k}}}$ on $|0_{\text{M}}\rangle$ in terms of operators acting on the right wedge. Notice that in our conventions, those of Ref. [@Crispino:2007eb], Rindler time in the left wedge also runs in the same direction as in the right Rindler wedge. To get the opposite conventions typically used in black hole physics and holography [@Maldacena:2001kr], one needs to perform the replacement $\hat{a}^{{\text{L}}\dagger}_{\omega\,(-\vec{k})}\to \hat{a}^{{\text{L}}\dagger}_{\omega\,\vec{k}}$.
Given this complete basis of operators, there are two different bases one can introduce which will be relevant in what follows. The first is the set of creation and annihilation operators associated with Minkowski time evolution $$\hat{a}^{\text{M}}_{k_z\vec{k}} = \int^\infty_0 \frac{d\omega}{\sqrt{2\pi a\,\omega_{\vec{k}}}}\frac{1}{\sqrt{1-e^{-2\pi\omega/a}}} \left[e^{i \vartheta(k_z) \frac{\omega}{a}}
\left(\hat{a}^{\text{L}}_{\omega\vec{k}} - e^{-\pi\frac{\omega}{a}}\,\hat{a}^{{\text{R}}\dagger}_{\omega\,(-\vec{k})}\right) + e^{-i\vartheta(k_z)\frac{\omega}{a}} \left(\hat{a}^{\text{R}}_{\omega\vec{k}} - e^{-\pi\frac{\omega}{a}}\,\hat{a}^{{\text{L}}\dagger}_{\omega\,(-\vec{k})}\right)\right]\,,
\label{eq:mink-op}$$ where $$\vartheta (k_z) = \frac{1}{2}\log \left(\frac{\omega_{\vec{k}}+k_z}{\omega_{\vec{k}}-k_z}\right)\,,$$ is the standard rapidity in relativistic physics and $\omega_{\vec{k}}^{2}=m^{2}+k_{z}^{2}+\vert\vec{k}\vert^{2}$ is the on-shell Minkowski frequency carried by each mode. In this basis, the generators of Minkoswki time and spatial $z$ translations are the standard expressions $$\begin{aligned}
\hat{H}^{{\text{M}}}&=\int\limits dk_{z}\, d^{d-1}\vec{k}\,\omega_{\vec{k}}\,\hat{a}^{{\text{M}}\dagger}_{k_z\vec{k}}\hat{a}^{\text{M}}_{k_z\vec{k}}\,,\\
\hat{P}_{z}^{{\text{M}}}&=\int\limits dk_{z}\,d^{d-1}\vec{k}\,k_{z}\,\hat{a}^{{\text{M}}\dagger}_{k_z\vec{k}}\hat{a}^{\text{M}}_{k_z\vec{k}}\;,
\end{aligned}
\label{eq:Mink-gen}$$ whereas the number operator equals $$\hat{N}^{{\text{M}}}=\int\limits dk_{z}\,d^{d-1}\vec{k}\,\hat{a}^{{\text{M}}\dagger}_{k_z\vec{k}}\hat{a}^{\text{M}}_{k_z\vec{k}}\;.$$ Notice that using , any Minkowski mode can be easily generated by acting with operators in the right wedge $$\label{ami3}
{\hat{a}^{\text{M}\dagger}_{k_z\vec{k}}}\vert 0_{M}\rangle = \int^\infty_0 \frac{d\omega}{\sqrt{2\pi a\,\omega_{\vec{k}}}} \sqrt{2\sinh \frac{\pi\omega}{a}} \left[e^{-i \frac{\omega}{a}\left(\vartheta(k_z) + i\frac{\pi}{2}\right)} \,\hat{a}^{R}_{\omega\,(-\vec{k})} + e^{i \frac{\omega}{a}\left(\vartheta(k_z) + i\frac{\pi}{2}\right)} \hat{a}^{R\dagger}_{\omega\vec{k}} \right] \vert 0_{\text{M}}\rangle\,.$$ The second basis is the Unruh basis defined by appropriate normalisation of the operators annihilating $|0_\text{M}\rangle$ in $$\label{eq:unruh1}
b_{+\omega,\vec{k}}= \frac{1}{\sqrt{1-e^{-2\pi\frac{\omega}{a}}}}\left( \hat{a}^{{\text{L}}}_{\omega\vec{k}} - e^{-\pi\frac{\omega}{a}}\,\hat{a}^{{\text{R}}\dagger}_{\omega\,(-\vec{k})}\right) \,, \quad
b_{-\omega,\vec{k}}=\frac{1}{\sqrt{1-e^{-2\pi\frac{\omega}{a}}}}\left( \hat{a}^{{\text{R}}}_{\omega\vec{k}} - e^{-\pi\frac{\omega}{a}}\,\hat{a}^{{\text{L}}\dagger}_{\omega\,(-\vec{k})}\right)\,,$$ together with $b^\dagger_{\pm \omega,\vec{k}}$. These operators satisfy standard commutation relations $$\left[\hat{b}_{\pm\omega\vec{k}}\,,\hat{b}^\dagger_{\pm\omega\vec{k}} \right] = \delta(\omega-\omega^\prime)\delta(\vec{k}-\vec{k^\prime})\,.$$ They are convenient to determine the thermal nature of the Rindler modes in $|0_{\text{M}}\rangle$ [@Unruh:1976db]. Indeed, using $${\hat{a}^{\text{R}}_{\omega\vec{k}}}= \frac{\hat{b}_{-\omega\vec{k}} + e^{-\pi\frac{\omega}{a}}\,\hat{b}^\dagger_{+\omega -\vec{k}}}{\sqrt{1-e^{-2\pi\frac{\omega}{a}}}}\,, \quad
{\hat{a}^{\text{L}}_{\omega\vec{k}}}= \frac{\hat{b}_{+\omega\vec{k}} + e^{-\pi\frac{\omega}{a}}\,\hat{b}^\dagger_{-\omega -\vec{k}}}{\sqrt{1-e^{-2\pi\frac{\omega}{a}}}}\,,
\label{eq:unruh}$$ and the fact that Unruh modes annihilate $|0_{\text{M}}\rangle$ (eq ), one easily finds $$\label{corr}
\begin{aligned}
\langle 0_{\text{M}}\vert\,{\hat{a}^{\text{R}}_{\omega\vec{k}}}{\hat{a}^{\text{R}}_{\omega\vec{k}}}\,|0_{\text{M}}\rangle &= \langle 0_{\text{M}}\vert\,{\hat{a}^{\text{R}\dagger}_{\omega\vec{k}}}{\hat{a}^{\text{R}\dagger}_{\omega\vec{k}}}\,|0_{\text{M}}\rangle = 0\,, \\
\langle 0_{\text{M}}\vert\,{\hat{a}^{\text{R}}_{\omega\vec{k}}}{\hat{a}^{\text{R}\dagger}_{\omega\vec{k}}}\,|0_{\text{M}}\rangle&= \frac{e^{\beta\omega}}{e^{\beta\omega}-1}\,\delta(\omega-\omega^\prime)\delta^{(d-1)}(\vec{k}-\vec{k}^\prime)\,, \\
\langle 0_{\text{M}}\vert\,{\hat{a}^{\text{R}\dagger}_{\omega\vec{k}}}{\hat{a}^{\text{R}}_{\omega\vec{k}}}\,|0_{\text{M}}\rangle &= \frac{1}{e^{\beta\omega}-1}\,\delta(\omega-\omega^\prime)\delta^{(d-1)}(\vec{k}-\vec{k}^\prime)\,.
\end{aligned}$$ The associated Unruh number and energy operators, labelled with a $U$ superscript, are $$\begin{aligned}
\hat{H}^{{\text{U}}}&=\int^\infty_0 d\omega\,\int\limits d^{d-1}\vec{k}\,\omega\left[\hat{b}^{\dagger}_{\omega\vec{k}} \hat{b}_{\omega\vec{k}} + \hat{b}^{\dagger}_{-\omega\vec{k}} \hat{b}_{-\omega\vec{k}}\right]\,, \\
\hat{N}^{{\text{U}}}&= \int^\infty_0 d\omega \int\limits d^{d-1}\vec{k}\left[\hat{b}^{\dagger}_{\omega\vec{k}} \hat{b}_{\omega\vec{k}} + \hat{b}^{\dagger}_{-\omega\vec{k}} \hat{b}_{-\omega\vec{k}}\right]\,.
\end{aligned}$$ It can be verified by direct computation that $N^{{\text{U}}}=N^{{\text{M}}}$.
Finally, starting from the expression for the boost generator (the total Rindler Hamiltonian in appropriate units) $$\label{Kren}
\hat{K}^{\textrm{M}}_\text{z} = \frac{1}{a}\left(\hat{H}_\text{R} - \hat{H}_\text{L}\right)\,,$$ and using the definition for the Minkowski creation/annihilation operators , together with the Rindler commutation relations , one can verify the expected commutation relations characteristic of the Poincaré algebra $$\left[ \hat{H}^{{\text{M}}},\hat{K}^{{\text{M}}}_\text{z}\right] = i\hat{P}^{{\text{M}}}_{\text{z}}\,, \quad \quad
\left[\hat{P}^{{\text{M}}}_{\text{z}},\hat{K}^{{\text{M}}}_\text{z}\right] =- i\hat{H}^{{\text{M}}}\,.$$
To sum up, using the free algebra of creation and annihilation operators in the left and right Rindler wedges together with the definition of the Minkowski vacuum , we constructed a set of operators $\hat{H}^{{\text{M}}},\,\hat{P}_{z}^{{\text{M}}}$ and $\hat{K}^{\textrm{M}}_\text{z}$, implicitly defined using and its hermitian conjugate, closing the exact Poincaré algebra.
From boundary CFT to bulk Poincaré algebra {#sec:GFFtoPoincare}
------------------------------------------
Let us consider large-N holographic theories. Due to large-N factorization, thermal correlation functions are determined by and . As noticed in [@Papadodimas:2012aq], a normalised version of the Fourier operators appearing in section \[sec:structure\] $$\hat{\mathcal{O}}_{\omega,\vec{k}} \equiv \frac{\mathcal{O}_{\omega,\vec{k}}}{\left(G_\beta(\omega,\vec{k})-G_\beta(-\omega,-\vec{k})\right)^{1/2}}= \frac{1}{\sqrt{G_\beta(t,\vec{k})}}\frac{\mathcal{O}_{\omega,\vec{k}}}{\sqrt{1-e^{-\beta\omega}}}\,,
\label{eq:GFFnormal}$$ have canonical commutations relations and their thermal expectation values satisfy $$\label{thermal}
\begin{aligned}
Z_\beta^{-1}\text{Tr}\left(e^{-\beta H}\hat{\mathcal{O}}^\dagger_{\omega,\vec{k}}\hat{\mathcal{O}}_{\omega^\prime,\vec{k}^\prime}\right) &= \frac{1}{e^{\beta\omega}-1}\delta(\omega-\omega^\prime)\delta^{d-1}(\vec{k}-\vec{k}^\prime)\,, \\
Z_\beta^{-1}\text{Tr}\left(e^{-\beta H}\hat{\mathcal{O}}_{\omega,\vec{k}}\hat{\mathcal{O}}^\dagger_{\omega^\prime,\vec{k}^\prime}\right) &= \frac{e^{\beta\omega}}{e^{\beta\omega}-1}\delta(\omega-\omega^\prime)\delta^{d-1}(\vec{k}-\vec{k}^\prime)\,.
\end{aligned}$$ Hence, these operators display the same algebra and expectation values as the right Rindler wedge creation/annihilation operators in and . We use this observation to explicitly construct boundary CFT operators closing the bulk Poincaré algebra, up to $1/N$ corrections, for both 2-sided and 1-sided holographic AdS black holes.
#### 2-sided AdS black holes.
If the state in the (right) CFT is exactly thermal, we can canonically purify it by the associated thermofield double state $$|\text{TFD}\rangle \equiv \frac{1}{\sqrt{Z(\beta)}}\sum_i e^{-\frac{\beta E_i}{2}} |E_i\rangle_{\text{L}}\otimes |E_i\rangle_{\text{R}}$$ belonging to the duplicated Hilbert space $\mathcal{H}_{\text{L}} \otimes \mathcal{H}_{\text{R}}$. This purification is precisely the GNS construction of the thermal state reviewed in appendix \[sec:GNS-thermal\]. It is holographically dual to the 2-sided eternal AdS black hole [@Maldacena:2001kr]. Since the algebra of operators is also duplicated, it gives rise to two sets of commuting modes $\mathcal{O}^\dagger_{\text{L}\,\omega,\vec{k}}$ and $\mathcal{O}^\dagger_{\text{R}\,\omega,\vec{k}}$ satisfying the same algebra and with the same expectation values as in and . They also verify the following relation $$\label{tomitO}
\left(\hat{\mathcal{O}}_{\text{L}\,\omega,\vec{k}} - e^{-\frac{\beta\omega}{2}}\,\hat{\mathcal{O}}^\dagger_{\text{R}\,\omega,\vec{k}}\right)|\text{TFD}\rangle = \left(\hat{\mathcal{O}}_{\text{R}\,\omega,\vec{k}} - e^{-\frac{\beta\omega}{2}}\,\hat{\mathcal{O}}^\dagger_{\text{L}\,\omega,\vec{k}}\right)|\text{TFD}\rangle = 0\,.$$ The origin of this equation is explained in appendix \[sec:GNS-thermal\]. It is an special case of eq. (\[tomit2\]), and it is basically equivalent to , the equation defining $|0_{\text{M}}\rangle$ in the Rindler discussion. As stressed there, our conventions involve time running in the same direction in both wedges. Furthermore, the Rindler acceleration $a$ is mapped to the black hole temperature using $\beta = \frac{2\pi}{a}$.
Given this algebraic equivalence, we can now proceed analogously to our discussion of the different bases of operators and Poincaré generators as in Rindler physics. In particular, the generator of time translations in the right CFT reduces in the large-N limit to $$\hat{H}^{\text{R}}=\int\limits d\omega\,d\vec{k}\,\omega\,\hat{\mathcal{O}}^\dagger_{\text{R}\,\omega,\vec{k}}\hat{\mathcal{O}}_{\text{R}\,\omega,\vec{k}}\,.$$ The Unruh creation/annihilation operators can be defined by $$\label{eq:unruh}
\mathcal{O}^{\textrm{U}}_{+\omega,\vec{k}}= \frac{1}{\sqrt{1-e^{-\beta\omega}}}\left( \hat{\mathcal{O}}_{\text{L}\,\omega,\vec{k}} - e^{-\beta\omega/2}\,\hat{\mathcal{O}}^\dagger_{\text{R}\,\omega,\vec{k}}\right) \,, \quad
\mathcal{O}^{\textrm{U}}_{-\omega,\vec{k}}=\frac{1}{\sqrt{1-e^{-\beta\omega}}}\left( \hat{\mathcal{O}}_{\text{R}\,\omega,\vec{k}} - e^{-\beta\omega/2}\,\hat{\mathcal{O}}^\dagger_{\text{L}\,\omega,\vec{k}}\right) \;,$$ while Minkowski annihilation modes can be defined by $$\mathcal{O}^{\text{M}}_{k_z\vec{k}} = \int^\infty_0 \frac{d\omega}{\sqrt{2\pi ak_0}}\frac{1}{\sqrt{1-e^{-\beta\omega}}} \left[e^{i\,2\pi\vartheta(k_z)\omega/\beta} \left(\hat{\mathcal{O}}_{\text{L}\,\omega,\vec{k}} - e^{-\beta\omega/2}\,\hat{\mathcal{O}}^\dagger_{\text{R}\,\omega,\vec{k}}\right) + e^{-i\,2\pi\vartheta(k_z)\omega/\beta} \left(\hat{\mathcal{O}}_{\text{R}\,\omega,\vec{k}} - e^{-\beta\omega/2}\,\hat{\mathcal{O}}^\dagger_{\text{L}\,\omega,\vec{k}}\right)\right]\,.$$ These operators allow us to define operators generating Minkowski time and spatial $z$ translations by $$\begin{aligned}
\hat{H}^{{\text{M}}}&=&\int\limits dk_{z}\, d^{d-1}\vec{k}\,\omega_{\vec{k}}\,\mathcal{O}^{{\text{M}}\dagger}_{k_z\vec{k}}\mathcal{O}^{\text{M}}_{k_z\vec{k}}\nonumber\\
\hat{P}_{z}^{{\text{M}}}&=&\int\limits dk_{z}\,d^{d-1}\vec{k}\,k_{z}\,\mathcal{O}^{{\text{M}}\dagger}_{k_z\vec{k}}\mathcal{O}^{\text{M}}_{k_z\vec{k}}\;,\end{aligned}$$ where $\omega_{\vec{k}}^{2}=m^{2}+k_{z}^{2}+\vert\vec{k}\vert^{2}$. The associated number operator is just given by $$\hat{N}^{\textrm{M}}=\int\limits dk_{z}\,d^{d-1}\vec{k}\,\mathcal{O}^{M\dagger}_{k_z\vec{k}}\mathcal{O}^M_{k_z\vec{k}}\;.$$ Since the algebra of the modes $\hat{\mathcal{O}}_{\text{R}\,\omega,\vec{k}}$ and $\hat{\mathcal{O}}_{\text{L}\,\omega,\vec{k}}$, together with the expectation values in the thermofield double, are equal to the ones in the Rindler discussion, up to $1/N$ corrections, we conclude these boundary CFT operators close the same bulk Poincaré algebra $$\label{pal}
\left[ \hat{H}^{{\text{M}}},\hat{H}^{\textrm{T}}\right] = i\hat{P}^{{\text{M}}}_{z}\,, \quad
\left[\hat{P}^{{\text{M}}},\hat{H}^{\textrm{T}}\right] =- i\hat{H}^{{\text{M}}}\,,$$ where the boost operator $\hat{H}^{\textrm{T}}$ is the total boost Hamiltonian of the two decoupled CFTs defined by $$\label{Hren}
\hat{H}^{\textrm{T}}=\frac{\beta}{2\pi}\left(\hat{H}^\textrm{R} - \hat{H}^\textrm{L}\right)$$ This is the same as in the Rindler discussion , after the replacement $a\to 2\pi/\beta$.
#### 1-sided AdS black holes.
The work in [@Papadodimas:2012aq; @Papadodimas:2013wnh; @Papadodimas:2013jku] extends the previous discussion to single AdS black holes involving a single boundary CFT. Observing that energy eigenstates $|E_i\rangle$ are well approximated by the thermal ensemble, one can work within the code subspace [@Almheiri:2014lwa; @Harlow:2018fse] and show that $|E_i\rangle$ is a *cyclic* and a *separating* vector in the Hilbert space with respect to the code subspace algebra [@deBoer:2018ibj; @deBoer:2019kyr]. The Tomita-Takesaki theorem[^9] guarantees the existence of a non-trivial “mirror” commutant. The mirror operators $\tilde{O}_{\omega,\vec{k}}$ generating this commutant play a similar role to the operators $\mathcal{O}_{\text{L}\,\omega,\vec{k}}$ in the 2-sided discussion, but they belong to the same boundary CFT dual to the single AdS black hole. They are defined by the following relations $$\begin{aligned}
\tilde{\mathcal{O}}_{\omega,\vec{k}} |\Psi_i\rangle &= e^{-\frac{\beta\omega}{2}}\,\mathcal{O}^\dagger_{\omega,\vec{k}}\, |\Psi_i\rangle \,, \\
\tilde{\mathcal{O}}_{\omega,\vec{k}} \mathcal{O}_{\omega_1,\vec{k}_1}\dots \mathcal{O}_{\omega_n,\vec{k}_n} |\Psi_i\rangle &= \mathcal{O}_{\omega_1,\vec{k}_1}\dots \mathcal{O}_{\omega_n,\vec{k}_n} \tilde{\mathcal{O}}_{\omega,\vec{k}} |\Psi_i\rangle \,, \\
\left[H, \tilde{\mathcal{O}}_{\omega,\vec{k}}\right] \mathcal{O}_{\omega_1,\vec{k}_1}\dots \mathcal{O}_{\omega_n,\vec{k}_n} |\Psi_i\rangle &= \omega\,\tilde{\mathcal{O}}_{\omega,\vec{k}} \mathcal{O}_{\omega_1,\vec{k}_1}\dots \mathcal{O}_{\omega_n,\vec{k}_n} |\Psi_i\rangle\,.
\end{aligned}
\label{eq:mirror-def}$$ These relations should be understood to hold only within the code subspace. The mirror operators are thus state dependent. As discussed in [@deBoer:2019kyr], there are some ambiguities regarding the $1/N$ extension of these operators, but for us it will be enough to work within the code subspace. This, together with microstates $|E_i\rangle$ being well approximated by the canonical ensemble, ensures the defining properties give rise to an algebra and correlation functions that are equivalent to the algebra of the $\mathcal{O}_{\text{L}\,\omega,\vec{k}}$ in the 2-sided discussion, and therefore to the algebra of annihilation operators in the left wedge. The previous formulas showing the existence of a Poincaré algebra can just be extended to this situation by the replacement $\mathcal{O}_{\text{L}\,\omega,\vec{k}}\rightarrow \tilde{O}_{\omega,\vec{k}}$.
Growth measures {#sec:growth}
===============
As argued in section \[sec:structure\], time evolution of simple perturbations in large-N theories at any finite temperature is simply described in terms of Fourier modes $\mathcal{O}_{\omega,\vec{k}}$ and $\mathcal{O}^\dagger_{\omega,\vec{k}}$. These modes allow for the evaluation of the series of nested commutators in any energy eigenstate compatible with the given temperature if the theory satisfies the ETH conjecture. Within the large-N limit,[^10] this operator structure should be enough to characterize any notion of operator growth.
In this section, we show this last expectation is indeed the case by describing some existent notions of operator growth, such as operator size in spin systems or the recursion method in many-body physics. We finish with some discussion regarding the relation between Nielsen’s geometric formulation of quantum circuit complexity, operator growth and quantum/classical chaos. In the way, we provide a generic framework to study operator growth in QFT, based on the GNS construction reviewed in appendix \[App2\]. Albeit our applications will be confined to large-N theories, the present GNS approach might help to understand the putative definitions of operator growth in generic QFT’s.
Operator growth as state mixing in the GNS construction {#sec:GNS}
-------------------------------------------------------
Talking about operator growth requires to be able to expand a given operator in different bases of the space of operators, to quantify how the support of the operator changes with time. Hence, given an operator algebra $\mathcal{A}$, we need an inner product endowing $\mathcal{A}$ with the structure of a Hilbert space. This is precisely the goal of the GNS construction, which is reviewed in appendix \[App2\]. Here we briefly summarize its main ingredients. Given a state[^11] $\phi$ acting on the algebra $\mathcal{A}$ satisfying $$A\in \mathcal{A}\, ,\,\,\,\,\, \phi (A^{\dagger}A) =0 \quad \Longleftrightarrow \quad A=0\,,$$ the GNS Hilbert space $\mathcal{H}_{\phi}$ and its inner product are defined by $$A\in\mathcal{A}\Rightarrow \vert A\rangle \in \mathcal{H}_{\phi}\,, \quad \langle B\vert A\rangle \equiv \phi (B^{\dagger}A) \,.$$ In this Hilbert space there are two equivalent representations $\pi$ and $\bar{\pi}$ of the algebra $\mathcal{A}$ acting on $\mathcal{H}_{\phi}$ $$A\in\mathcal{A}\Rightarrow \vert A\rangle \in \mathcal{H}_{\phi}\,, \quad \pi (A)\vert B\rangle \equiv \vert AB\rangle \, , \,\,\,\,\,\, \bar{\pi} (A)\vert B\rangle \equiv \vert BA^{\dagger}\rangle\,.$$ This construction is valid for any type of operator algebras, including the type III algebras relevant for QFT.
Consider states $|\kappa\rangle$ arising from abstract states in the algebra which are invariant under time evolution, such as thermofield double states.[^12] In this context, the GNS construction maps the Heisenberg time evolution $A(t)$ of any operator belonging to the algebra to the Schrödinger’s time evolution of the associated GNS state $$U(t) \pi(A)\vert \kappa\rangle\equiv \pi (A(t))\vert \kappa\rangle =\vert A(t)\kappa\rangle \equiv \vert \Psi (t)\rangle\,,$$ The same conclusion holds for the representation $\bar{\pi}$.
Since operator evolution is equivalent to state evolution in the associated GNS Hilbert space, any notion of operator growth should be characterized by expectation values of *size* operators in the GNS Hilbert space $\mathcal{H}_{\phi}$ $$\label{gen}
\langle \Psi (t)\vert \sum\limits_{ij}\bar{\pi}(B_i) \pi (B_j)\vert \Psi (t)\rangle\,,$$ as any other property attached to states in $\mathcal{H}_{\phi}$. In the previous relation $B_j$ runs over a basis of operators of the algebra. Below we show how this is the case in some recent examples in lattice systems.
#### Operator size for simple Majorana operators.
Before exploring this perspective for large-N theories, we briefly comment on how the case of Majorana spin systems at infinite temperature [@Roberts:2018mnp] and its extension to finite temperature [@Qi:2018bje] fit in this framework.
Consider a set of $N$ fundamental Majorana operators normalized by $\lbrace\psi_{a},\psi_{b}\rbrace^{2}=2\delta_{ab}$. Every operator $\psi$ in the algebra $\mathcal{A}$ can be expanded as $$\mathcal{O} =\sum\limits_{s=1}^{N}\,\sum\limits_{a_{1}\cdots a_{s}}c_{a_{1}\cdots a_{s}}\psi_{a_{1}}\cdots \psi_{a_{s}}\;.$$ The *size* of such operator was defined by [@Roberts:2018mnp] $$\label{size}
S_{\mathcal{O}}=\sum\limits_{s=1}^{N}\,s\sum\limits_{a_{1}\cdots a_{s}}\vert c_{a_{1}\cdots a_{s}}\vert^2\,.$$ This is natural if one thinks of the label $s$ as describing, either location in a 1d lattice, or directly in terms of the number of fundamental fermions building the operator.
The connection to the GNS construction is as follows. One assigns a vector $|\mathcal{O} \rangle \in \mathcal{H}_{\phi}$ to every operator $\mathcal{O} \in \mathcal{A}$ with GNS inner product $$\langle\mathcal{O} \vert\mathcal{O}'\rangle\equiv \frac{1}{Z}\textrm{Tr}(\mathcal{O}^{\dagger}\mathcal{O}')\;,$$ where $Z$ is the dimension of the Hilbert space $Z=\textrm{Tr}(\mathds{1})$ induced by the normalized inner product for finite matrices. It follows the identity element of the algebra $\mathds{1}\to \vert\mathds{1}\rangle$ and the inner product can be interpreted as an expectation value at infinite temperature. There is one natural representation of the algebra in $ \mathcal{H}_{\phi}$, defined by $\pi (\mathcal{O})\vert\mathcal{O}'\rangle =\vert\mathcal{O}\mathcal{O}'\rangle$. This allows to define unitary time evolution in the GNS Hilbert space by $$U_{t}\pi(\mathcal{O})\vert\mathds{1}\rangle\equiv\pi(\mathcal{O}(t))\vert\mathds{1}\rangle =\vert\mathcal{O}(t)\rangle \;.$$ This gives a concrete example on how operator evolution is seen as state evolution in the GNS Hilbert space.
Within the GNS construction, there exists an operator $\hat{S}$ acting on $\mathcal{H}_{\phi}$ $$\label{s1}
\hat{S}=\sum\limits_{s=1}^{N}\,s \sum\limits_{a_{1}\cdots a_{s}}\pi (\psi_{a_{1}}\cdots \psi_{a_{s}})\vert\mathds{1}\rangle\langle\mathds{1}\vert \pi (\psi_{a_{1}}\cdots \psi_{a_{s}})= \sum\limits_{s=1}^{N}\,s \sum\limits_{a_{1}\cdots a_{s}}\vert\psi_{a_{1}}\cdots \psi_{a_{s}}\rangle\langle\psi_{a_{1}}\cdots \psi_{a_{s}}\vert\;,$$ satisfying $$S_{\mathcal{O}}(t)=\langle\mathcal{O} (t)\vert \hat{S}\vert \mathcal{O} (t)\rangle\,.$$ Hence, the notion of size equals the expectation value of $\hat{S}$ in the GNS state $\vert \mathcal{O} (t)\rangle$ that is mapped to the time evolution of the original operator $\mathcal{O}(t)$. This matches our general expectation that any notion of operator growth should be computable by expectation values evaluated on the GNS state. In fact, introducing creation $c^{\dagger}_{i}$ and annihilation $c_{i}$ operators in $\mathcal{H}_{\phi}$ by $$\begin{aligned}
c_{i}\vert\psi_{a_{1}}\cdots \psi_{a_{s}}\rangle &=\delta_{i,a_{1}}\vert\psi_{a_{2}}\cdots \psi_{a_{s}}\rangle +\cdots +\delta_{i,a_{s}}\vert\psi_{a_{2}}\cdots \psi_{a_{s-1}}\rangle\,, \\
c^{\dagger}_{i}c_{i}\vert\psi_{a_{1}}\cdots \psi_{a_{s}}\rangle &=\delta_{i,a_{1}}\cdots \delta_{i,a_{s}}\,,
\end{aligned}
\label{s2}$$ the size operator can be reinterpreted as a *number* operator $$\hat{S}=\sum\limits_{i=1}^{N}c^{\dagger}_{i}c_{i}\;.
\label{eq:f-size}$$ Hence, we learn there is a basis of operators in $\mathcal{H}_{\phi}$ where a natural notion of size in this lattice system is a simple quadratic operator.
This notion of size in lattice systems was extended to finite temperature in [@Qi:2018bje] by purifying the thermal ensemble of the Majorana fermions. This approach is then directly on a GNS form, as reviewed in appendix \[sec:GNS-thermal\], except that using fermionic operators. Hence, our conclusions extend to this finite temperature case too.
#### Alternative notions of operator size
Depending on the dynamics and state of the system, operator size defined as in may not be a dynamical quantity. This can happen even if the complexity of the operator grows. Indeed, if the Hamiltonian preserves the number of particles, the previous definition of operator size will be a conserved charge for a natural class of initial states, as we review now.
Consider a bunch of spinless fermions whose hamiltonian conserves the number of particles. Any state can be expanded as $$\vert \psi\rangle =\sum\limits_{s=0}^{N}\,\sum\limits_{a_{1}\cdots a_{s}}\psi_{a_{1}\cdots a_{s}}c^{\dagger}_{a_{1}}\cdots c^{\dagger}_{a_{s}}\vert \downarrow_{1}\cdots \downarrow_{N}\rangle \;,$$ where $c^{\dagger}_{i}$ and $c_{i}$ create and destroy such fermions at site $i$. Besides operator size, one can ask about how many particles are being transported by the Hamiltonian as time evolves. This was considered in [@Magan:2016ehs]. To be definite, consider an starting state with the first $m$ particles excited. Unitary evolution mixes the state with other states in the $m$-particle sector $$\vert \psi (t)\rangle =\sum\limits_{a_{1}\cdots a_{m}}\psi_{a_{1}\cdots a_{m}} (t)\,c^{\dagger}_{a_{1}}\cdots c^{\dagger}_{a_{m}}\vert \downarrow_{1}\cdots \downarrow_{N}\rangle\;.$$ One proposed notion of operator growth is the average number of jumps (the average transport) the spins have performed due to Hamiltonian evolution. Such number can be measured by the expectation value of the number operator $$\label{n1}
\hat{T}\equiv\sum\limits_{i=m+1}^{N}c^{\dagger}_{i}c_{i}$$ counting the number of fermions in the sites that were not populated at $t=0$. It follows $$T(t)\equiv\langle\psi (t)\vert \hat{T}\vert\psi (t)\rangle$$ is the average particle transport. This is a sensible measure of the growth of the state $\vert\psi (t)\rangle$ and it is still the expectation value of a simple operator. Also, since $$\vert \psi (t)\rangle = U(t)c^{\dagger}_{1}\cdots c^{\dagger}_{m}U^{-1}(t)\vert \downarrow_{1}\cdots \downarrow_{N}\rangle \equiv \mathcal{O}(t) \vert \downarrow_{1}\cdots \downarrow_{N}\rangle\;,$$ this expectation value is equivalently studying the growth of the operator $\mathcal{O}=c^{\dagger}_{1}\cdots c^{\dagger}_{m}$ in the state $\vert \downarrow_{1}\cdots \downarrow_{N}\rangle $. Notice that while this notion of size is bound to grow, as analized in [@Magan:2016ehs], the previous notion of size, where we would add up all spins in the definition (\[n1\]), would be constant through time evolution due to particle number conservation.
Size, number operators and energies in large-N and holographic theories {#sec:size-N}
-----------------------------------------------------------------------
The observation that Heisenberg operator evolution is equivalent to Schrödinger’s time evolution in the GNS Hilbert space provides a hint to extend the notion of operator size to QFT. Such notion is based on *simple* operators, such as and in spin systems (see [@Magan:2016ehs; @Roberts:2018mnp; @Qi:2018bje; @Parker:2018yvk]).
Given the structure of time evolution for holographic CFTs in the large-N limit discussed in sections \[sec:structure\] and \[sec:GFFtoPoincare\], it may be natural to define any notion of operator size as being of the form $$\label{sizeQFT}
\hat{S}=\sum_\alpha F_{\alpha}\left[\left(\mathcal{O}_\alpha\right)_{\omega,\vec{k}},\left(\mathcal{O}^\dagger_\alpha\right)_{\omega,\vec{k}},
\left(\tilde{\mathcal{O}}_\alpha\right)_{\omega,\vec{k}},\left(\tilde{\mathcal{O}}^\dagger_\alpha\right)_{\omega,\vec{k}}\right]$$ in terms of the operator modes and its mirror partners for the different local low conformal dimension boundary operators indexed by $\alpha$. The additive nature on the spectrum of operators is due to the absence of mixing between operators when neglecting 1/N corrections.
Assuming the generic definition , large-N factorization ensures that any notion of size in large-N theories associated with a choice of the functionals $F_\alpha$ is completely determined by the two-point function (\[eq:exp\]), up to 1/N corrections. Indeed, to study the growth of the size of a certain field $\mathcal{O}(t)$ in the thermofield double $\vert\kappa_{\beta}\rangle$ at temperature $\beta$, we just need to take the expectation value of $\hat{S}$ in the evolving GNS state $\vert\mathcal{O}(t)\kappa_{\beta}\rangle$. Such expectation value can be computed by using expressions (\[exparep\]) and (\[eq:exp\]), together with large-N factorization.
Let us remark that in previous literature, starting with [@Roberts:2018mnp], the notion of size has been argued to be related to out-of-time ordered correlation functions. If this relation is extended to any large-N theory and any temperature, our analysis would imply a non-trivial relation between two and four-point functions in the large-N limit. These aspects will be studied elsewhere.
In analogy to the spin size , natural choices for the size $\hat{S}$ in large-N QFTs are simple operators like the number or energy operators associated to the different bases (Rindler, Unruh, and Minkowski) of creation and annihilation operators discussed in section \[sec:GFFtoPoincare\]. As in our discussion of the dynamics preserving the number of particles, not all choices for such operators will provide useful dynamical information. In what follows, the upper index in the size operator refers to the basis chosen, either Rindler (R), Unruh (U) or Minkowski (M), and the lower index to whether it is energy-based (H) or particle number based (N).
Let us start our discussion with the number and energy operator associated to the standard basis of operator modes $$\hat{S}^{\textrm{R}}_{\textrm{N}} = \hat{N}=\int\limits_{\omega> 0}d\omega d^{d-1}\vec{k}\, \mathcal{O}^{\dagger}_{\omega,\vec{k}}\mathcal{O}_{\omega,\vec{k}}\, \quad \text{and} \quad
\hat{S}^{\textrm{R}}_{\textrm{H}} = \hat{H}=\int\limits_{\omega> 0}d\omega d^{d-1}\vec{k}\,\,\omega\,\mathcal{O}^{\dagger}_{\omega,\vec{k}}\mathcal{O}_{\omega,\vec{k}}\,.
$$ Then the expectation value $\langle \kappa_\beta\mathcal{O}(t)|\hat{S}|\mathcal{O}(t)\kappa_\beta\rangle$ is constant for any operator $\mathcal{O}$ and provides no further dynamical information.
#### Unruh and Minkowski number operators.
Consider the Unruh and Minkowski number operators choice $$\hat{S}^{\textrm{U}}_{\textrm{N}}= \hat{N}^{\textrm{U}}= \int\limits dk_{z}\, d^{d-1}\vec{k}\,\mathcal{O}^{\textrm{U}\dagger}_{\omega\vec{k}}\mathcal{O}^\textrm{U}_{\omega\vec{k}}\,, \quad \text{and} \quad
\hat{S}^{\textrm{M}}_{\textrm{N}}= \hat{N}^{\textrm{M}}= \int\limits dk_{z}\, d^{d-1}\vec{k}\,\mathcal{O}^{\textrm{M}\dagger}_{k_z\vec{k}}\mathcal{O}^\textrm{M}_{k_z\vec{k}}\,.$$ Both choices are equal due to the algebra of field modes. This proposal is inspired by the relation (\[tomitO\]), which is a specific instance of the more general Tomita-Takesaki like equation . One basically defines the simplest operators annihilating the thermofield double[^13] and uses them to define a number operator. For Majorana fermions this was the path chosen in [@Qi:2018bje]. Here we see that such operators, in the black hole scenario, are the known Unruh creation/annihilation operators.
These notions of size give zero on the thermofield double, while positive sizes on the thermofield double with perturbations. However, if we consider a Minkowski mode excitation generated by $\mathcal{O}^{\textrm{M}\dagger}_{k_z\vec{k}}$, this notion of size will remain constant through time evolution. This is because unitary evolution acts like a boost on these excitations. Hence, it changes the value of their momentum but not the number of modes. This choice shows that depending on the basis of operators being considered, equating size with number operators may not provide useful dynamical information.
We remark that although the Minkowski creation/annihilation operators are state-dependent in the one-sided case since they make use of the mirror operators, exciting a Minkowski mode is not a state-dependent action. The reason is that the action of the mirror creation operator can be fully mimicked by their partners in the right wedge, as in .
#### Minkowski energy.
Consider the boundary CFT operator describing the analog of a bulk infalling Hamiltonian[^14] constructed in the previous section $$\hat{S}^{\textrm{M}}_{\textrm{H}}= H^{\textrm{M}}= \int\limits dk_{z}\, d^{d-1}\vec{k}\,\omega_{\vec{k}}\,\mathcal{O}^{\textrm{M}\dagger}_{k_z\vec{k}}\mathcal{O}^{\textrm{M}\dagger}_{k_z\vec{k}}\,.$$ This choice was studied in detail in [@Magan:2018nmu]. Since the CFT time evolution acts as a boost upon a Minkowski mode $\mathcal{O}^{\textrm{M}\dagger}_{k_z\vec{k}}$ $$e^{i\gamma\,K}\,\mathcal{O}^{\textrm{M}\dagger}_{k_z\vec{k}}\,e^{-i\gamma\,K} = \frac{\sqrt{\gamma(\omega_{\vec{k}})}}{\sqrt{\omega_{\vec{k}}}}\,\mathcal{O}^{\textrm{M}\dagger}_{\gamma(k_z)\vec{k}}\,,$$ it follows the size will exponentially grow at large times $(t^\text{M}\gg \beta)$, with Lyapunov exponent equal to $2\pi/\beta$, since evolving with the QFT Hamiltonian for time $t$ is equivalent to boosting the particle with rapidity $\frac{2\pi}{\beta} t$ (see ).
The recursion method at large-N {#sec:recursion}
-------------------------------
A standard approach in condensed matter physics to study the Heisenberg time evolution and complexity of operators is the recursion method (see [@viswanath2008recursion] for a detailed presentation). This perspective on operator growth was recently considered in SYK in [@Parker:2018yvk] and used to study long time scales of operator dynamics in [@Barbon:2019wsy]. In this section we explore what the structure of time evolution in large-N theories described in sections \[sec:structure\] and \[sec:GFFtoPoincare\] teaches us about this approach.
Let us first describe this method briefly. As before, the recursion method requires the definition of an inner product. The book [@viswanath2008recursion] considers a whole family of them, but we show in appendix \[inner-products\] that all choices can be related to one convenient representative in QFT. In the following we focus on such representative and, for simplicity, we only consider operators with vanishing one-point functions. Given two operators $A$ and $B$, the representative inner product $(A,B)$ is defined by $$(A,B)\equiv \langle e^{\beta H/2}A^{\dagger}e^{-\beta H/2} B\rangle_{\beta}\,,
\label{innerL}$$ where $\langle A \rangle_\beta = \text{Tr}(\rho_\beta\,A)$. Within the GNS framework, this inner product can be written as $$\langle \kappa\vert \bar{\pi}(A)\pi (B)\vert \kappa\rangle =\langle \kappa\vert B\kappa A^\dagger\rangle=\langle e^{\beta H/2}A^{\dagger}e^{-\beta H/2} B\rangle_{\beta}=(A,B) \,.
\label{eq:inner-GNS}$$
The inner product allows to expand any hermitian operator $\mathcal{O}(t)$ in an orthogonal basis of operators $$\label{exp}
\mathcal{O}(t)=\sum\limits_{k=0}^{\infty}C_{k}(t)f_{k}\,, \quad \text{with} \quad (f_{k},f_{k'})=(f_{k},f_{k})\,\delta_{kk'}$$ for some time dependent coefficients $C_{k}(t)$. The recursion method proposes to use an explicit basis $f_{k}$, the Lanczos basis, to study operator evolution. Basically, starting from the operators $\mathcal{O}_{n}$ defined previously, it provides a constructive algorithm, based on the Gram-Schmidt orthogonalization procedure, to determine such basis $$f_{k+1} = i[H,\,f_k] + \Delta_k\,f_{k-1}\,, \quad k=0,1,\dots \quad \text{with} \quad \Delta_k = \frac{\left(f_j,\,f_j\right)}{\left(f_{j-1},\,f_{j-1}\right)}\,, \,\,\,j=1,2,\dots
\label{eq:rec-basic}$$ with initial conditions $f_{-1}=0$ and $f_0=\mathcal{O}$ being the operator at initial time. The coefficients $\Delta_k$ are referred to as the Lanczos coefficients. The linear operator $\mathcal{L}$ generating the time evolution as $\mathcal{L}(\mathcal{O})\equiv [H,\mathcal{O}]$ is sometimes called the Liouvillian. It corresponds to the GNS hamiltonian in the GNS construction described in appendix \[App2\] generating the unitary evolution in -.
Plugging the expansion into Heisenberg’s equation of motion, one derives a 1d diffusion equation for the amplitudes $C_k(t)$ $$\frac{d\mathcal{O}}{dt} = i\mathcal{L}\,\mathcal{O} \quad \Leftrightarrow \quad \dot{C}_k(t) = C_{k-1}(t) - \Delta_{k+1}\,C_{k+1}(t)\,, \quad k=0,1,2\dots
\label{eq:diffusion}$$ with initial conditions $C_{-1}(t)=0$ and $C_k(0) = \delta_{k0}$. In this framework, the Lanczos operator complexity $L_{\mathcal{O}}$ of an operator $\mathcal{O}$ can be defined by the average position in this effective 1d chain[^15] $$L_{\mathcal{O}}\equiv \sum\limits_{k=0}^{\infty}k\,\frac{\vert (f_{k},\mathcal{O}(t))\vert^{2}}{(f_{k},f_{k})}\,.
\label{lc}$$ Because of , $L_{\mathcal{O}}$ equals the expectation value of the “Lanczos operator” in the GNS Hilbert space $$\hat{L}_{\mathcal{O}}\equiv\sum\limits_{k=0}^{\infty}k\, \frac{\bar{\pi}(f_{k})}{\sqrt{(f_{k},f_{k})}}\vert\kappa\rangle\langle \kappa\vert \frac{\bar{\pi}(f_{k})}{\sqrt{(f_{k},f_{k})}}\,,$$ so that $$L_{\mathcal{O}}=\langle \kappa\vert \pi (\mathcal{O}(t)) \hat{L}_{\mathcal{O}}\pi (\mathcal{O}(t))\vert\kappa\rangle\,.$$ Notice $\hat{L}_{\mathcal{O}}$ is a state dependent operator in the GNS Hilbert space, i.e. it is *not* linear in $\mathcal{O}$.
Knowledge of the Lanczos coefficients $\Delta_n$ determines the orthogonal basis $\{f_k\}$ and allows to determine the full-time evolution of the operator $\mathcal{O}(t)$ through the integration of the 1d diffusion equation . Interestingly, these coefficients $\Delta_n$ can be recursively extracted from the connected 2-pt function $$Q(t) \equiv \langle \mathcal{O}(t)\mathcal{O}(0)\rangle_\beta \,.
\label{eq:2ptbeta}$$ This can be seen as follows [@viswanath2008recursion]. Since $$C_{0}(t) \equiv \frac{\left(\mathcal{O}(t),\,\mathcal{O}\right)}{\left(\mathcal{O},\,\mathcal{O}\right)} = \frac{\Phi(t)}{\Phi(0)}\,,$$ where $\Phi(t) \equiv \left(Q(t) + Q(-t)\right)/2$, if one assumes the Taylor expansion $$C_0(t) = \sum_{k=0}^\infty \frac{(-1)^k}{(2k)!}\,M_{2k}\,t^{2k}\,,$$ is sensible, it follows the existence of a one–to–one reconstruction algorithm between the moments $M_{2k}$, which are determined solely by derivatives of the 2-pt function , and the Lanczos coefficients $\Delta_k$ $$M_{2k}^{(n)} = \frac{M_{2k}^{(n-1)}}{\Delta_{n-1}} - \frac{M_{2k-2}^{(n-2)}}{\Delta_{n-2}}\,, \quad \Delta_n=M_{2n}^{(n)}\,, \quad k=n,n+1,\dots ,K\,\,\,\,\,\text{and} \,\,\,\, n=1,2,\dots ,K
\label{eq:moment-Delta}$$ The initial conditions of this recursion are $M_{2k}^{(0)}=M_{2k}$ and $\Delta_{-1}=\Delta_0=1$ and $M_{2k}^{(-1)}=0$.
In section \[sec:structure\], the structure of time evolution in large-N theories was discussed. It is natural to ask whether we can learn anything about the recursion method given this structure. The first observation is that the operators $\mathcal{O}_{\omega,\vec{k}}$ diagonalize the Liouvillian $$\frac{d \mathcal{O}_{\omega,\vec{k}}(t)}{dt}=i\mathcal{L} (\mathcal{O}_{\omega,\vec{k}})=i[H,\mathcal{O}_{\omega,\vec{k}}]=-i\omega \,\mathcal{O}_{\omega,\vec{k}}\quad \Longrightarrow \quad \mathcal{O}_{\omega,\vec{k}}(t)=e^{-i\omega t}\,\mathcal{O}_{\omega,\vec{k}}(t)$$ Hence, there are two natural bases in the space of operators: the Lanczos basis, made of the orthogonal $f_k$ and the operator modes $\mathcal{O}_{\omega,\vec{k}}$ diagonalising the Liouvillian in the large-N limit. Finding the change of basis would immediately solve the diffusion equation , since the time evolution of the $\mathcal{O}_{\omega,\vec{k}}$ operators is known. Since properly normalised operator modes satisfy the canonical commutation relation associated to a set of of creation and annihilation operators, we analyse the latter from the recursion method perspective next.
Consider $K$ free harmonic oscillators, denoted by $a_{j}$ and $a^\dagger_j$, with $j=1,\dots ,K$. Take as our starting operator in the recursion method algorithm $$f_0 = \sum_j D_j (a_j+a^\dagger_j)\,, \quad \quad D_i\in \mathbb{R}$$ with $D_j$ any set of coefficients. This choice matches the $t=0$ expansion of the field in and accommodates their normalisation . Since $\mathcal{L}a^\dagger_j = \omega_j\,a^\dagger_j$ and $\mathcal{L} a_j = -\omega_j\,a_j$, it follows the vectors generated by the recursion method algorithm must be of the form $$\begin{aligned}
f_{2k} &= \sum_j D_j\,P_{k,j} (a_j + a^\dagger_j)\,, \,\,k=0,1,\dots K-1\quad \quad \text{with}\quad P_{0,j}=1\,\,\forall\,j \\
f_{2k+1} &= \sum_j D_j\,Q_{k,j}\,i (a^\dagger_j-a_j)\,, \,\,k=0,1,\dots K-1 \quad \quad \text{with}\quad Q_{0,j}=\omega_j \,\,\forall\,j
\end{aligned}
\label{eq:vec-ansatz}$$ for some unknown real coefficients $P_{k,j}$ and $Q_{k,j}$ satisfying the above initial conditions. Using the harmonic oscillator thermal correlators $$\langle a_i\,a^\dagger_j\rangle_\beta = e^{\beta\omega_i}\,n_\beta(\omega_i)\,\delta_{ij}\,, \quad \quad \langle a^\dagger_i\,a_j\rangle_\beta = n_\beta(\omega_i)\,\delta_{ij},$$ where $n_\beta(\omega) = (e^{\beta\omega}-1)^{-1}$, it follows $$\left(a_j + a_j^\dagger,\,a_k + a^\dagger_k\right) = \left(i(a_j^\dagger-a_j), i(a_k^\dagger - a_k)\right) = \frac{2}{\beta}\int^\beta_0 d\lambda\,e^{\lambda\omega_j}\,n_\beta(\omega_j) \delta_{jk} \equiv A_j\delta_{jk}\,,
\label{eq:norm-id}$$ with all other inner product combinations vanishing. Hence, the set of vectors $\{f_{2k},\,f_{2k+1}\}$ defines an orthogonal set, as it should. For $K$ finite oscillators, the algorithm will halt once we reach a basis of $2K$ orthogonal vectors, which matches the number of independent creation/annihilation operators. This explicitly confirms the relation between the two set of operators is indeed simply a change of basis. This change is only non-trivial for $K > 1$, as in large-N QFTs where there is an infinite set of operator modes when studying the growth of a local field.
Plugging the parameterisation into , we obtain the following recurrence relations $$\begin{aligned}
P_{s,j} &= -\omega_j\,Q_{s-1,j} + \Delta_{2s-1}\,P_{s-1,j}\,, \quad \quad s=1,2,\dots K-1\\
Q_{s,j} &= \omega_j\,P_{s,j} + \Delta_{2s}\,Q_{s-1,j}\,, \quad \quad s=1,2,\dots K-1
\end{aligned}
\label{eq:rec-rel}$$ These are solved by[^16] $$\begin{aligned}
P_{s,j} &= \sum_{m=0}^s (-1)^m\,\omega_j^{2m}\,\sum_{i_1=1}^{2m+1} \Delta_{i_1} \sum_{i_2=i_1+2}^{2m+3} \Delta_{i_2}\dots \sum_{i_{s-m}=i_{s-m-1}+2}^{2s-1} \Delta_{i_{s-m}}\,, \\
Q_{s,j} &= \omega_j \sum_{m=0}^s (-1)^m\,\omega_j^{2m}\,\sum_{i_1=1}^{2(m+1)} \Delta_{i_1} \sum_{i_2=i_1+2}^{2(m+2)} \Delta_{i_2}\dots \sum_{i_{s-m}=i_{s-m-1}+2}^{2s} \Delta_{i_{s-m}}\,.
\end{aligned}
\label{eq:rec-sol}$$ The proof is by induction and it is given in App (\[App5\]).
As a check of our formal solution to the recursion method, we show it satisfies the diffusion equation . As stressed above, the advantage of expressing our operators in the basis of field modes is that these modes diagonalize the Liouvillian operator $\mathcal{L}$ and therefore make the time dependence trivial. Hence, the exact time evolution of our initial operator just equals $$f_0(t) = \sum_j D_j\,\left(e^{it\omega_j}a^\dagger_j + e^{-it\omega_j}\,a_j\right)\,.$$ From this expression, to find the Lanczos expansion we just need to write the $a^\dagger_j$ and $a_j$ in terms of the $f_{k}$. We remark this is not a dynamical question, just a change of basis. To invert the relation we use that the amplitudes entering the diffusion equation are given by the projections $$C_{2k}(t) \equiv \frac{(f_{2k},\,f_0(t))}{(f_{2k},\,f_{2k})}\,, \quad\quad C_{2k+1}(t) \equiv \frac{(f_{2k+1},\,f_0(t))}{(f_{2k+1},\,f_{2k+1})}\,.$$ Explicit calculation yields $$\label{solution}
\begin{aligned}
C_{2k}(t) &= \sum_j D_j^2\,A_j\,\frac{P_{k,j}}{(f_{2k},\,f_{2k})}\,\cos\omega_jt\,, \\
C_{2k+1}(t) &= \sum_j D_j^2\,A_j\,\frac{Q_{k,j}}{(f_{2k+1},\,f_{2k+1})}\,\sin\omega_jt\,.
\end{aligned}$$ Computing the time derivatives and using the recursion relations to replace the $\omega_j$ dependent terms reproduces the diffusion equation .
To sum up, the 2-pt function determines the Lanczos coefficients $\Delta_{k}$ and these determine the orthonormal basis of operators $f_k$ controlling the Heisenberg time evolution of an initial operator $\mathcal{O}$. These are related by a change of basis to the operators used in section \[sec:structure\] to describe this same evolution in large-N gauge theories. This is consistent with our claim that 2-pt functions characterize operator growth in these theories at leading order. In this approximation, what makes these theories special, from the recursion method perspective, is that we can solve the diffusion equation analytically by relations , once we have the coefficients $\Delta_{k}$. In QFT, the set of modes is infinite and the recursion does not halt. The operator then grows indefinitely, even if we are dealing with a set of free harmonic oscillators.
The specific structure of the Lanczos basis unraveled here also differs from the general exponential growth of the Lanczos complexity in QFT. As remarked in [@Parker:2018yvk], in the typical QFT scenario in which the 2-pt function is exponentially decaying and its Fourier transform has poles, as dictated by the generic analyticity properties of thermal correlations, then the mean position in the one-dimensional diffusion equation , the Lanczos operator complexity , will grow exponentially fast with Lyapunov exponent $2\pi/\beta$.[^17] Such statement holds for any chaotic QFT, and it is not particular to large-N theories.
Chaos and quantum complexity {#sec:chaos}
----------------------------
In this section we stress the natural relation between operator growth and quantum circuit complexity, besides their connection through quantum chaos [@Magan:2018nmu; @Bueno2019; @Barbon:2019tuq]. Moreover, in the context of the AdS/CFT correspondence, we also comment on the relation between them and the emergence of classical bulk chaos.
In quantum complexity discussions, the complexity $\mathcal{C}_{\ket{\psi}}$ to prepare a particular target state $\ket{\psi}$ starting with a certain reference state $| \psi_{\textrm{\tiny R}} \rangle$ by applying a series of elementary gates $g_i$ $$\label{circuit}
\ket{\psi}= U_{\textrm{\tiny R}}\, \ket{\psi_{\textrm{\tiny R}}}= g_{n}\cdots g_{2}\,g_{1}\ket{\psi_{\textrm{\tiny R}}}\,,$$ is defined as the number of gates associated to the optimal protocol. Nielsen and collaborators [@2005quant.ph..2070N; @2006Sci...311.1133N; @2007quant.ph..1004D] mapped the problem of identifying this optimal circuit to the geometric problem of finding a geodesic in the space of unitaries acting on the Hilbert space. Given a one parameter family of states, labelled by $s$, the local driving hamiltonian $H(s)$ satisfies $$i \frac{d}{ds}\ket{\psi(s)}= H(s)\ket{\psi(s)}$$ and generates the unitary transformation acting on the state $$U(\sigma) = {\reflectbox{\ensuremath{\vec{ \reflectbox{\ensuremath{\mathcal{P}}}}}}} \exp \left[ -i \int^\sigma_0\!\!\! d s\, H(s)\right], \quad \text{with} \quad H(s)\equiv \sum_\text{I} Y^\text{I}(s)\,\mathcal{O}_\text{I}\,,$$ where the Hermitian operators $\mathcal{O}_I$ generate the individual gates $g_\text{I}$. Circuits satisfying eq. correspond to trajectories satisfying the boundary conditions $$U(\sigma=0)= \mathbbm{1}\,, \qquad U(\sigma=1)= U_{\textrm{\tiny R}}\,.$$ Optimal circuits minimise the cost defined as $$\mathcal{C}_{\ket{\psi_{\textrm{\tiny R}}}\rightarrow\ket{\psi}} \equiv \int^1_0 ds ~ F \left( H(s)\right)$$ where $F$ is a local cost function depending on the tangent vector $H(s)$.
Consider two states, $\ket{\psi}$, as above, and a perturbed state $\ket{\psi_{\mathcal{O}}}=e^{i\mathcal{O}}\ket{\psi}$, generated by the action of a simple unitary generated by certain local operator $\mathcal{O}$. The time evolution of both states is determined by the unitary action $U(t)=e^{-iHt}$ on them, where $H$ corresponds to the physical hamiltonian of the system, leading to the states $\ket{\psi (t)}$ and $\ket{\psi_{\mathcal{O}}(t)}$, respectively. In this set-up, one can define some notion of growth or size based on the circuit complexity to go from one evolved state to the other[^18] $$S_{\mathcal{O}(t)}\equiv \mathcal{C}_{\ket{\psi (t)}\rightarrow\ket{\psi_{\mathcal{O} }(t)} }$$
This relative complexity is simpler than expected in the limit of small perturbations. As shown in the original geometric complexity paper [@2005quant.ph..2070N], if the perturbation is small enough so that $\ket{\psi}$ and $\ket{\psi_{\mathcal{O}}}$ are sufficiently closed to each other, the geodesic connecting them is simply $$U(s)_{\ket{\psi}\rightarrow\ket{\psi_{\mathcal{O}}}}= e^{i\mathcal{O}s}\,, \qquad 0\leqslant s\leqslant 1$$ The key observation now is that the geodesic connecting the time evolved states $\ket{\psi (t)}$ and $\ket{\psi_{\mathcal{O} (t)}}$ is also going to be of the same type $$U(s)_{\ket{\psi (t)}\rightarrow\ket{\psi_{\mathcal{O}}(t)} }= e^{i\mathcal{O}(-t) s}\,, \qquad 0\leqslant s\leqslant 1\,,$$ by dialling the initial perturbation to be small enough. This argument allows to write the relative complexity of such geodesic as $$\label{circuitt}
S_{\mathcal{O}(t)}\equiv \mathcal{C}_{\ket{\psi (t)}\rightarrow\ket{\psi_{\mathcal{O} }(t)} }= \int^1_0 ds ~ F \left( \mathcal{O}(-t)\right) = F \left( \mathcal{O}(-t)\right)\;.$$ The precise evaluation requires a choice of the cost function. See [@Bueno2019] for a discussion and calculation of several possibilities.
#### Connection to previous notions of size.
Equation states that the circuit complexity is the computational cost of the time evolved operator responsible for the perturbation.[^19] The final value depends on the choice of a cost function, pretty much as in our earlier discussions on operator size, the latter depends on the definition of size. Crucially, stresses that, given some cost function, circuit complexity only depends on the time evolution of the operator, the same structure controlling any notion of operator growth or size. Hence, both notions are functionally dependent. In particular, if we were to define the cost as one of the previous notions of operator size, both would be equivalent. A convenient choice then is the Minkowski energy discussed in the previous section. With this choice, the relative complexity of large-N theories will grow exponentially fast with Lyapunov exponent $\lambda= 2\pi/\beta$. [^20]
#### Connection to chaos.
Chaotic behavior concerns the sensitivity of certain dynamical systems to small perturbations $\delta x_{i} $ of its initial conditions. Classically, such sensitivity is usually studied in a double scaling limit, where the size of the perturbation is taken to zero first and the limit of large times is taken afterward. The first limit ensures that a linearized equation of the type $$\label{chaosclas}
\delta x_{i} (t)=\sum\limits_{j}M_{ij}\delta x_{i}\;,$$ is a good approximation to the dynamics, where the Jacobian matrix $M_{ij}=\partial x_{i}(t)/\partial x_{j}$ encodes the dependence on the initial conditions $x_j$. The second limit ensures the solution to is dominated by the largest Lyapunov exponent.
A natural extension of the classical definition to the quantum domain has been recently developed in [@Bueno2019]. As shown in [@Ashtekar:1997ud], quantum dynamics can be formulated as classical dynamics on a “quantum phase space”, defined to be the Hilbert space itself. The symplectic form at point $\vert\psi\rangle$ along two infinitesimal directions, generated by Hamiltonian operators $H_{1}$ and $H_{2}$ is just the expectation value of the commutator $\Omega_{\vert\psi\rangle}(H_{1},H_{2})\equiv\langle\psi\vert [H_{1},H_{2}]\vert\psi\rangle$. It turns out that Schrödinger equations are seen as Hamilton equations in such a phase space with “classical Hamiltonian” $H(\vert\psi\rangle)=\langle\psi\vert \hat{H}\vert \psi\rangle$. Having framed quantum dynamics as a classical system, it is most natural to define quantum chaos by the usual classical definition (\[chaosclas\]) but applied to the quantum phase space. This definition has, by construction, the appropriate pullback to the classical definition on a semiclassical phase space, but it is otherwise valid through the whole quantum system.
The classical approach to quantum mechanics illuminates the relation between operator growth and chaos by showing the transparent relation between $\mathcal{O}(-t)$, the operator that generates the unitary interpolating between the nearby quantum states at time $t$, and the Jacobian matrix associated to the classical chaotic process. In the Hilbert space we can define generalized coordinates $\vert q_{i},p_{i}\rangle$ (at least locally) satisfying the canonical Poisson brackets with respect to the Hilbert space symplectic form. Generic infinitesimal perturbations of any state can be written as[^21] $$e^{i\mathcal{O}}\vert q_{i},p_{i}\rangle \equiv e^{i(\hat{p}_{i}\delta q_{i}-\hat{q}_{i}\delta p_{i})}\vert q_{i},p_{i}\rangle =\vert q_{i} +\delta q_{i},p_{i}+\delta p_{i}\rangle \,\,\,\,\,\Longrightarrow\,\,\,\,\, \mathcal{O}=\sum\limits_{i}\hat{p}_{i}\delta q_{i}-\hat{q}_{i}\delta p_{i}\;.$$ This equation just states that any small perturbation $\mathcal{O}$ can be expanded in the generators of translations along the local reference frame defined by $q_{i},p_{i}$. Evolving in time, one observes $$\mathcal{O} (-t)=\sum\limits_{i}\hat{p}_{i}(-t)\delta q_{i}-\hat{q}_{i}(-t)\delta p_{i}=\sum\limits_{i}\hat{p}_{i}\delta q_{i} (t)-\hat{q}_{i}\delta p_{i} (t)\;,$$ where $\delta q_{i} (t)$ and $\delta p_{i} (t)$ are determined by a linearized equation of the type (\[chaosclas\]) associated to the quantum phase space. This is analyzed in a specific generic example in [@Bueno2019]. It follows that the growth properties of the perturbation $\mathcal{O}$ are controlled by the Jacobian matrix defining the chaotic process in the quantum phase space, and vice versa.
#### Classical chaos in AdS/CFT.
In the context of the AdS/CFT correspondence, CFT perturbations $e^{i\mathcal{O}}\ket{\psi}$ generated by operators $\mathcal{O}$ with large conformal dimension can be described by freely falling particles in the bulk geometry dual to $\ket{\psi}$, in the semiclassical approximation [@Fitzpatrick:2011jn; @Nozaki:2013wia; @Goto:2016wme; @Goto:2017olq; @Terashima:2017gmc; @Berenstein:2019tcs]. At high energies, the bulk geometry involves a black hole with the temperature related to the energy of the state by the usual thermodynamic relation.
It was realized in [@Barbon:2011pn; @Barbon:2011nj; @Barbon:2012zv] that the optical metric capturing the near horizon physics has chaotic behavior, i.e. the optical metric is controlled by a hyperbolic space whose geodesics are known to have chaotic properties. The hyperbolic space (universally attached to any horizon) turns out to have a radius of curvature given by $R=\beta/2\pi$, and therefore the associated bulk geodesic deviation growth is compatible with the holographic Lyapunov exponent. Such deviation rests on the conformal transformation from the near horizon Rindler geometry to the hyperbolic one. It therefore secretly rests on the emergent Poincaré symmetries described above. The fact that Minkowski energies and radial momenta grow exponentially in the Rindler frame with Lyapunov exponent $\lambda=2\pi/\beta$ is mapped, in the optical frame, to the fact that perturbations in the transverse direction grow exponentially fast with the same Lyapunov exponent. The conformal transformation from the Rindler frame to the optical one was studied at the classical level and also in the QFT setup in [@Barbon:2012zv]. Having constructed the Poincaré symmetries above from the structure of time evolution, such conformal transformation to the optical frame can be constructed as well, as if we were in the Minkowski/Rindler scenario, and the chaotic properties of the optical metric are thus recovered.
Discussion {#Discussion}
==========
We have considered the problem of operator growth in large-N gauge theories at finite temperature. We framed the problem as that of understanding the operators $$\mathcal{O}_{n} \equiv [H,\cdots,[H,\mathcal{O}]\cdots]$$ where $H$ is the Hamiltonian. This is most natural since the expansion of Heisenberg time evolution in powers of time teaches us these are the only operators with whom the initial operator mixes over time.
In section \[sec:structure\], we argued that the expression $$\label{no2}
\mathcal{O}_{n}(t)=\int_{\omega > 0} \frac{d\omega d^{d-1}\vec{k}}{(2\pi)^d}\, \left((-\omega)^{n}\,\mathcal{O}_{\omega,\vec{k}}\,e^{-i\omega t + i\vec{k}\vec{x}} + \,\omega^{n}\,\mathcal{O}^\dagger_{\omega,\vec{k}}\,e^{i\omega t - i\vec{k}\vec{x}}\right)\,$$ together with linearity, large-N factorization, the fact that most eigenstates at a given temperature behave as thermal states (ETH) and the correlation functions of the field modes $\mathcal{O}_{\omega,\vec{k}}$ and $\mathcal{O}^\dagger_{\omega,\vec{k}}$, given by , completely specify the action of these operators in most interesting states, up to $1/N$ corrections. Hence, determines the time evolution of the operator $\mathcal{O}(t,\vec{x})$. A similar statement holds for modular time evolution as well.
A first interesting insight is that *any* notion of operator growth should be determined by the 2-pt function at this order. Given the relation found between operator growth and four-point functions at an infinite temperature in SYK [@Roberts:2018mnp], one might hope for a non-trivial relation between the two-point function and the connected four-point function in generic large-N theories. We leave this interesting observation for future work.
In section \[sec:algebra\], we constructed an emergent *bulk* Poincaré algebra as the first application of our proposed solution. This was achieved by the known doubling of the modes appearing in large-N theories [@Papadodimas:2012aq]. This algebra is related to the near horizon Rindler behaviour of thermal horizons. Albeit we have focused on the conventional modes, defined by means of the Hamiltonian of the large-N theory, the construction can be easily extended to modular time evolution, by using the modular modes (\[modular\]) instead of the conventional ones. If large-N factorization holds, we can again find renormalized modes satisfying the algebra of free creation and annihilation operators, and proceed with the construction of an emergent Poincaré algebra. It would be interesting to develop the arguments given here in relation to the recent results in [@deBoer:2019uem], based on prior work [@Czech:2019vih], where such Poincaré algebra emerges from a local bulk perspective due to the limiting modular evolution behaviour, making the latter much closer to Equivalence Principle considerations.
Albeit the simplicity of the previous solution contrasts with the expected complexity of the problem, in section \[sec:growth\], we analyzed several existent notions of operator growth and size existent in the literature from the perspective presented here. These include number operators, energy measures, the recursion method in condensed matter physics, and the approach to quantum chaos based on quantum circuit complexity. All of them can be considered in detail, and analytically, in the basis of field modes. We have seen that the different approaches are just variations over a common theme: the evolution in time of the initial operator $\mathcal{O}$, which is fully characterized by expression , at leading order. We made proposals for operator size in large-N QFTs by noticing that the GNS construction maps operator evolution to conventional state evolution in the GNS Hilbert space, and also derived an explicit relation between operator complexity and circuit complexity in eq . In the large-N limit, all such notions are functionals of the two-point function alone.[^22]
It is an important open problem to understand how to systematically incorporate $1/N$ corrections to our discussion. Given our approach, this is not an intrinsic problem attached to operator growth, but it is generic to large-N gauge theories including holographic ones if one is interested in a bulk interpretation of these statements. See [@Mousatov:2019xmc] for a recent discussion on how to incorporate these corrections in the bulk for some choices of operator size.
We finish by stressing a point made in the introduction: our work neatly shows that *any* notion of operator growth has an equivalent formulation both in the bulk and the boundary theories, in the context of large-N holographic theories. This is just a consequence of the equality between bulk and boundary Hilbert space and Hamiltonians. In our setup, on a technical level, this is transparent due to the work in bulk reconstruction relating bulk field modes with boundary modes $\mathcal{O}_{\omega,\vec{k}}$, $\mathcal{O}^\dagger_{\omega,\vec{k}}$, and their mirror partners, mainly following [@Papadodimas:2012aq; @Papadodimas:2013wnh; @Papadodimas:2013jku].
Acknowledgments {#acknowledgments .unnumbered}
===============
We thank José Barbón, Pablo Bueno, Horacio Casini and Simon Ross for useful discussions. We also want to thank the Kavli Institute for Theoretical Physics and the Higgs Centre for Theoretical Physics for hospitality and financial support during the workshops “Chaos and Order” and “Recent development in holography”, respectively. The work of JMM was supported by the Simons foundation through the It From Qubit Simons collaboration.
The Gelfand-Naimark-Segal (GNS) construction {#App2}
============================================
The GNS construction [@GelNeu43; @segal1947] generates a Hilbert space $\mathcal{H}_{\omega}$ from an abstract $C^\star$-algebra $\mathcal{A}$ and a linear functional (state) $\omega$ from $\mathcal{A}$ to $\mathbb{C}$, together with a representation of the algebra $\pi (\mathcal{A})$ acting on it. In this appendix, we review its main ideas following closely [@Haag:1992hx].
A $C^\star$-algebra $\mathcal{A}$ is a set of objects such that if $A,B\in \mathcal{A}$ and $a,b\in \mathbb{C}$, then the linear combination $aA+bB\in \mathcal{A}$. Furthermore, there exists a map $A\mapsto A^{\dagger}\,,$ $\forall\,A\in\mathcal{A}$ being an involution and satisfying $$(AB)^\dagger=B^\dagger A^\dagger\,, \quad (aA)^\dagger = a^\star\,A^\dagger \quad \quad \forall A,B \in \mathcal{A}\,, \quad \forall a\in\mathbb{C}$$ Up to topological requirements, see [@Haag:1992hx] for a more detailed account, $\mathcal{A}$ is called a von Neumann algebra if it further contains the identity. From now on we consider von Neumann algebras only.
States $\omega$ are positive and normalized linear functionals from $\mathcal{A}$ to $\mathbb{C}$ satisfying $$\begin{aligned}
\omega (aA+bB)&= a\,\omega(A)+b\,\omega (B)\,, \\
\omega (A^\star A)&\geq 0\,,\\
\omega (\mathds{1})&=1\,.
\end{aligned}
\label{eq:omega-state}$$
Hilbert spaces $\mathcal{H}$ are vector spaces with an inner product mapping any pair of elements $\vert w\rangle,\vert v\rangle\in\mathcal{H}$ to a complex number $\langle v\vert w\rangle \in \mathbb{C}$ satisfying $$\begin{aligned}
\langle v\vert w\rangle &=\langle w\vert v\rangle^\star\,, \\
\langle av_{1}+bv_{2}\vert w\rangle &=a^\star\langle v_{1}\vert w\rangle+b^\star \langle v_{2}\vert w\rangle\,, \\
\vert\langle v\vert v\rangle\vert^{2}&\geq 0\,,
\end{aligned}
\label{eq:inner-def}$$ where the last line is only saturated for $ \vert v\rangle =0$.
Let the algebra $\mathcal{A}$ be an algebra of operators. Since the algebra $\mathcal{A}$ is already a vector space over $\mathbb{C}$, to become a Hilbert space it requires an inner product. This can be defined using the state $\omega$ as $$\langle A\vert B\rangle = \omega (A^{\dagger}B)\,,
\label{eq:inner-map}$$ where we used the standard notation in quantum mechanics $|A\rangle$ to refer to the vector in $\mathcal{H}$ associated with the operator $A\in \mathcal{A}$.
This inner product satisfies all the requirements except for the existence of non-zero operators $W$ satisfying $\omega (W^{\dagger}W)=0$. The set of such operators $\mathcal{I}$ is a left ideal in $\mathcal{A}$, the so called Gelfand ideal of the state $\omega$, i.e a linear subspace of $\mathcal{A}$ that is stable under multiplication by any element $A\in \mathcal{A}$ from the left $$W\in \mathcal{I}\,\, ,\,\,\, A\in \mathcal{A}\Rightarrow AW\in\mathcal{I}\,.$$ The GNS construction defines the Hilbert space $\mathcal{H}_\omega$ as the quotient of $\mathcal{H}$ by the ideal $\mathcal{I}$, i.e. $\mathcal{H}_{\omega}\equiv \mathcal{A}/\mathcal{I}$ [^23]. Vectors $\vert [A]\rangle \in \mathcal{H}_{\omega}$ correspond to equivalence classes of operators in the algebra of the form $A+\mathcal{I}$ and such classes do not depend on the representative.
GNS induces a representation $\pi_{\omega}$ of $\mathcal{A}$ acting on $\mathcal{H}_{\omega}$ by the product in the algebra $\mathcal{A}$ $$\pi_{\omega}(A)\vert [B]\rangle =\vert [AB]\rangle\;.
\label{pi-repr}$$ A consequence of this construction is that the identity class $\vert \Omega\rangle\equiv\vert [\mathds{1}]\rangle$ vector can be associated to the starting state $\omega$ since $$\omega (A)=\langle \Omega\vert A \vert \Omega\rangle\,.$$
GNS of the thermal state {#sec:GNS-thermal}
------------------------
When the GNS construction is considered for finite-dimensional algebras $\mathcal{A}$ containing bounded operators, the identity operator $\mathds{1}$ can be understood as the maximally entangled density matrix (up to normalization) and the inner product can be taken as $$\langle A|B\rangle = \frac{1}{Z}\text{Tr}\left(A^\dagger B\right)\,,
\label{eq:infinite}$$ where $Z\equiv \text{Tr}(\mathds{1})$. There is no Gelfand ideal $\mathcal{I}$ in this case. Hence, there exists an isomorphism between $\mathcal{A}$ and $\mathcal{H}_\omega$. The same conclusion holds when one replaces $\mathds{1}$ with $\kappa = \rho_\beta^{1/2}$, where $\rho_\beta$ is the Boltzmann finite temperature density matrix (or any density matrix of full rank).
Besides the GNS representation $\pi(\mathcal{A})$ $$\pi (A)\vert \kappa\rangle \equiv\vert A\kappa\rangle\,,$$ there exists the conjugate representation $\bar{\pi}(\mathcal{A})$, defined by $$\bar{\pi}(A)\vert \kappa\rangle \equiv\vert \kappa A^{\dagger}\rangle\,.$$ These are equivalent because there exists an anti-unitary operator $J$ acting on $\mathcal{H}_{\omega}$ satisfying $$J\vert A\kappa\rangle = \vert \kappa A^{\dagger}\rangle \,\,\,\,\,\,\text{with} \,\,\,\,\,\,\, J^{2}=1\,,$$ implying $$J\pi (A)J= \bar{\pi}(A)\,.$$ It also follows from these expressions that $$\omega (A)=\langle\kappa\vert\pi (A)\vert \kappa\rangle =\langle\kappa\vert \bar{\pi} (A)\vert \kappa\rangle^{*}\,.
\label{eq:state-beta}$$ Since $\vert\kappa\rangle$ is invariant under time evolution[^24], we can now easily define unitary evolution $U_{t}$ in all states of the representation by $$\label{ugns1}
U_{t}\pi (A)\vert \kappa\rangle =\pi (A_{t})\vert \kappa\rangle \,, \,\,\,\,\,\,\,\,\,\,\,\,\, U_{t}\bar{\pi} (A)\vert \kappa\rangle =\bar{\pi} (A_{t})\vert \kappa\rangle\;.$$ It helps to disentangle the meaning of these definitions to explicitly write the unitary evolution as $$\label{ugns2}
U_{t}=\pi (e^{iHt})\bar{\pi} (e^{iHt})\;.$$ We can then check the definition $$U_{t}\pi (A)\vert \kappa\rangle =U_{t}\vert A\kappa\rangle= \vert e^{iHt}A\kappa e^{-iHt}\rangle=\vert e^{iHt}Ae^{-iHt}\kappa \rangle=\vert A_{t}\kappa \rangle=\pi (A_{t})\vert \kappa\rangle$$ is satisfied. Given this representation, it is natural to introduce the full hamiltonian as $H_{\textrm{F}}=\pi (H)-\bar{\pi} (H)$. Using $\kappa = \rho_\beta^{1/2} = Z^{-1/2} e^{-\beta H/2}$, it follows $$J e^{-\beta H_{\textrm{F}}/2}\pi (A)\vert\kappa\rangle =J e^{-\beta H_{\textrm{F}}/2}\vert A\kappa\rangle =J \vert\kappa A\rangle =J \bar{\pi}(A^{\dagger})\vert\kappa \rangle = \pi (A^{\dagger})\vert \kappa\rangle \;.
\label{tomit}$$ It is interesting to single out the equality from the first term to the fourth term. Multiplying from the left both side by $J$ and moving the right hand side to the left we obtain $$(e^{-\beta H_{\textrm{F}}/2}\pi (A)-\bar{\pi}(A^{\dagger}))\vert\kappa\rangle =0\;,
\label{tomit2}$$ which is the (generalized) origin of the known relations (\[trind\]) and (\[tomitO\]), associated to free QFT and large-N theories.
As stressed through the article, one general lesson is that operator evolution can be seen as a conventional state evolution through the GNS construction. The generator of the GNS unitary evolution, the GNS Hamiltonian, is what in the condensed matter community is called the Liouvillian [@viswanath2008recursion], see [@Parker:2018yvk; @Barbon:2019wsy] for recent applications of such approach. Furthermore, the GNS construction points out the subtlety of the Gelfand ideal, stressing why states with full rank are convenient, and allows a direct application into QFT since it is valid for all types of algebras.
Inner products in the space of operators {#inner-products}
----------------------------------------
In the previous GNS construction, one starts with a natural inner product on the space of operators, such as or . This choice of inner product is not unique. Denoting a general inner product by $(A,B)$, in [@viswanath2008recursion] the following family is considered $$\label{inner}
(A,B)=\frac{1}{\beta}\int\limits_{0}^{\beta}\,d\lambda\,g(\lambda)\, \langle e^{\lambda H}A^{\dagger}e^{-\lambda H} B\rangle_{\beta}-\langle A^{\dagger}\rangle_\beta \langle B\rangle_\beta\;,$$ where $\langle A \rangle_\beta \equiv \text{Tr}(\rho_\beta\,A)$ and $g(\lambda)$ is any function satisfying: $$g(\lambda)\geq 0 \,\,\,\,\,\,\,\, g(\beta-\lambda)=g(\lambda)\,\,\,\,\,\,\,\,\,\,\frac{1}{\beta}\int\limits_{0}^{\beta}d\lambda\, g(\lambda)=1\;.$$ Examples of such functions are $$g(\lambda)=\frac{1}{2}\beta\, [\,\delta (\lambda)+\delta (\beta-\lambda)]\, , \,\,\,\,\,\,\,\,\, g(\lambda)=\delta (\beta/2-\lambda)\;,$$ for which the inner product reduces to $$(A,B)=\frac{1}{2}\langle A^{\dagger} B+B A^{\dagger}\rangle-\langle A^{\dagger}\rangle \langle B\rangle\, , \,\,\,\,\,\,\,\,\, (A,B)= \langle e^{\beta H/2}A^{\dagger}e^{-\beta H/2} B\rangle_{\beta}-\langle A^{\dagger}\rangle \langle B\rangle\;,$$ respectively. In this section we want to describe the status of this big family (\[inner\]) of inner products. In particular, we show below how they change the specific functional describing the chaotic growth. Indeed faster growths than the chaos bound can be obtained, albeit there is no surprise here, since one can actually relate all the inner products (\[inner\]) to the one defined by $g(\lambda)=\beta\delta (\beta/2-\lambda)$, as we show below.
To test the dependence on the inner product we can analyze the basic quantity controlling the growth, which is the return probability $$p(t)\equiv\vert ( \mathcal{O}(t)\vert \mathcal{O}(0) )\vert^{2}\;.$$ Using the previous inner products we thus need to analyze[^25] $$(\mathcal{O}(t),\mathcal{O}(0))=\frac{1}{\beta}\int\limits_{0}^{\beta}\,d\lambda\,g(\lambda)\, \langle e^{\lambda H}\mathcal{O}(t)e^{-\lambda H} \mathcal{O}\rangle_{\beta}=\frac{1}{\beta}\int\limits_{0}^{\beta}\,d\lambda\,g(\lambda)\, \langle \mathcal{O}(t-i\lambda) \mathcal{O}\rangle_{\beta}\;,$$ and such function is completely determined equivalently by its Fourier transform $$R(\omega)=\int e^{i\omega t}\,(\mathcal{O}(t),\mathcal{O}(0))\;.$$ To find such Fourier transform first consider: $$\label{gl}
G_{\lambda}(\omega)\equiv \int e^{i\omega t}\,\langle \mathcal{O}(t-i\lambda) \,\mathcal{O}\rangle_{\beta}\;.$$ The expectation value $\langle \mathcal{O}(t) \,\mathcal{O}\rangle_{\beta}$ can be analytically continued to imaginary times for $0> \textrm{Im} (t)> -\beta$. Call $F(z)$ such a unique function in the strip. We can consider the following integrals of such function: $$\int\limits_{\mathcal{C}_{\lambda}} F(z) e^{iz\omega}dz\;,$$ where $\mathcal{C}_{\lambda}$ is a path parallel to the real time axis, shifted by $-i\lambda$. Due to the absence of poles or singularities in such a region, Cauchy’s theorem implies: $$\int\limits_{\mathcal{C}_{\lambda}} F(z) e^{iz\omega}dz = \int\limits_{\mathcal{C}_{\lambda'}} F(z) e^{iz\omega}dz \;.$$ This implies that: $$\int\limits_{-\infty}^{\infty} F(t-i\lambda)e^{i(t-i\lambda)\omega}dt=\int\limits_{-\infty}^{\infty} F(t-i\lambda')e^{i(t-i\lambda')\omega}\;.$$ Looking at (\[gl\]), we conlude that $$e^{\lambda\omega}G_{\lambda}(\omega)=e^{\lambda'\omega}G_{\lambda'}(\omega)\;.$$ This relation is important, since it says that knowing the exact Fourier transform at some imaginary time $\lambda$ is equivalent to knowing it at all complexified times in the holomorphic strip. In particular, for 2d CFT’s, we have the exact results for $\lambda=\beta/2$, so that for general $0<\lambda<\beta$: $$G_{\lambda}(\omega)=e^{-\lambda\omega}e^{\beta\omega/2}G_{\beta/2}(\omega)\;.$$ The Fourier transform of the autocorrelation function ends up being: $$R(\omega)=e^{\beta\omega/2}G_{\beta/2}(\omega)\frac{1}{\beta}\int\limits_{0}^{\beta}\,d\lambda\,g(\lambda)\,e^{-\lambda\omega}\;,$$ which is a simple functional of the information encoded in the inner product defined by $g(\lambda)=\beta\delta (\beta/2-\lambda)$.
The previous relation implies that all inner products have slower decay tails of $R(\omega)$ at large $\omega$ than the choice $g(\lambda)=\beta\delta (\beta/2-\lambda)$. The choice $g(\lambda)=\beta\delta (\beta/2-\lambda)$ is indeed the one considered in [@Parker:2018yvk], which in chaotic QFT leads to a exponential growth with Lyapunov exponent $\lambda=2\pi/\beta$. Therefore, we conclude that all other choices display stronger Lanczos growths than the one expected by the chaos bound. We point out the existence of the results found in [@Romero-Bermudez:2019vej], in the context of OTOCs, which show that different choices of euclidean separations in the 4-pt functions lead to faster growths than the chaos bound [@Maldacena:2015waa].
Solving the recurrence relation {#App5}
===============================
In this appendix we complete the proof of the solution of the Lanczos recursion method when applied to large-N theories. Given $P_{0,j}$ and $Q_{0,j}$ in , the ansatz reproduces correctly $P_{1,j}$ and $Q_{1,j}$. Assuming holds for $s$, we want to show $P_{s+1,j}$ and $Q_{s+1,j}$ satisfy . Let us use the first equation in to compute $P_{s+1,j}$ $$\begin{aligned}
P_{s+1,j} &= -\omega_j\,Q_{s,j} + \Delta_{2s+1}\,P_{s,j} \\
&= \sum_{m=0}^s (-1)^{s+1}\omega_j^{2(m+1)} \,\sum_{i_1=1}^{2(m+1)} \Delta_{i_1} \sum_{i_2=i_1+2}^{2(m+2)} \Delta_{i_2}\dots \sum_{i_{s-m}=i_{s-m-1}+2}^{2s} \Delta_{i_{s-m}} \\
& + \Delta_{2s+1}\,\sum_{r=0}^s (-1)^r\,\omega_j^{2r}\,\sum_{i_1=1}^{2r+1} \Delta_{i_1} \sum_{i_2=i_1+2}^{2r+3} \Delta_{i_2}\dots \sum_{i_{s-r}=i_{s-r-1}+2}^{2s-1} \Delta_{i_{s-r}} \\
&= (-1)^{s+1}\omega_j^{s+1} + \sum_{r=1}^s (-1)^r\,\omega_j^{2r} \,\sum_{i_1=1}^{2r} \Delta_{i_1} \sum_{i_2=i_1+2}^{2(r+1)} \Delta_{i_2}\dots \sum_{i_{s-m}=i_{s+1-r}+2}^{2s} \Delta_{i_{s-m}} \\
& + \Delta_{2s+1}\,\sum_{r=0}^s (-1)^r\,\omega_j^{2r}\,\sum_{i_1=1}^{2r+1} \Delta_{i_1} \sum_{i_2=i_1+2}^{2r+3} \Delta_{i_2}\dots \sum_{i_{s-r}=i_{s-r-1}+2}^{2s-1} \Delta_{i_{s-r}}\,.
\end{aligned}
\label{eq:rec-step}$$ To derive the third equality, we relabelled $m+1=r$ and wrote the $r=s+1$ term separately, i.e. the only one involving no $\Delta$s. Notice that the only $\omega_j^0$ term comes from $r=0$ in the last line and reproduces the right answer $\Delta_{2s+1}\Delta_1\Delta_3\dots \Delta_{2s-1}$. Hence, given an index $r=\{1,2,\dots s\}$, the task is to show $$\sum_{i_1=1}^{2r+1} \Delta_{i_1} \sum_{i_2=i_1+2}^{2r+3} \Delta_{i_2}\dots \sum_{i_{s+1-r}=i_{s-r-1}+2}^{2s+1} \Delta_{i_{s+1-r}}
\label{eq:rec-final}$$ equals the sum of the coefficients in the last two lines of for a given $r$. The key point is the different upper limit in the last sum in being $2s+1$. Indeed, when $i_{s+1-r}\neq 2s+1$, the contribution from equals the first line in . Hence, we are left to show the contribution from $i_{s+1-r} = 2s+1$ in $$\Delta_{2s+1} \sum_{i_1=1}^{2r+1} \Delta_{i_1} \sum_{i_2=i_1+2}^{2r+3} \Delta_{i_2}\dots \sum_{i_{s-r}=i_{s-r-1}+2}^{2s-1} \Delta_{i_{s-r}}$$ equals the last line in , which it does.
We are left to check $Q_{s+1,j}$ satisfies $$Q_{s+1,j} = \omega_j \sum_{m=0}^{s+1} (-1)^m\,\omega_j^{2m}\,\sum_{i_1=1}^{2(m+1)} \Delta_{i_1} \sum_{i_2=i_1+2}^{2(m+2)} \Delta_{i_2}\dots \sum_{i_{s+1-m}=i_{s-m}+2}^{2(s+1)} \Delta_{i_{s+1-m}}
\label{eq:rec-step2}$$ having assumed $P_{s+1,j}$ and $Q_{s,j}$ do satisfy . Using the second recursion relation , we can write $Q_{s+1,j}$ as $$\begin{aligned}
& \omega_j\,P_{s+1,j} + \Delta_{2(s+1)}\,Q_{s,j} = \\
&\omega_j \sum_{m=0}^{s+1} (-1)^m\,\omega_j^{2m}\,\sum_{i_1=1}^{2m+1} \Delta_{i_1} \sum_{i_2=i_1+2}^{2m+3} \Delta_{i_2}\dots \sum_{i_{s+1-m}=i_{s-m}+2}^{2s+1} \Delta_{i_{s+1-m}}\\
&+ \Delta_{2(s+1)}\,\omega_j \sum_{m=0}^{s} (-1)^m\,\omega_j^{2m}\,\sum_{i_1=1}^{2(m+1)} \Delta_{i_1} \sum_{i_2=i_1+2}^{2(m+2)} \Delta_{i_2}\dots \sum_{i_{s-m}=i_{s-m-1}+2}^{2s} \Delta_{i_{s-m}}\,.
\end{aligned}$$ Notice the second line already has the right functional dependence and the right number of sum terms, except for the fact that upper limits in the sums are off. In particular, the last index is not allowed to reach the maximal value $i_{s+1-m}=2(s+1)$. Consider the contribution from $\Delta_{2(s+1)}$ in $$\Delta_{2(s+1)}\,\omega_j \sum_{m=0}^{s+1} (-1)^m \omega_j^{2m} \sum_{i_1=1}^{2(m+1)} \Delta_{i_1} \dots \sum_{i_{s-m}=i_{s-m-1}+2}^{2s} \Delta_{i_{s-m}}\,.$$ In our conventions, the term $m=s+1$ vanishes, be definition. The remaining terms match $\Delta_{2(s+1)}\,Q_{s,j}$ above. The only remaining question is whether the terms in $Q_{s+1,j}$ not including $\Delta_{2(s+1)}$ $$\omega_j \sum_{m=0}^{s+1} (-1)^m\,\omega_j^{2m}\,\sum_{i_1=1}^{2(m+1)} \Delta_{i_1} \sum_{i_2=i_1+2}^{2(m+2)} \Delta_{i_2}\dots \sum_{i_{s+1-m}=i_{s-m}+2}^{2s+1} \Delta_{i_{s+1-m}}$$ equal $\omega_j\,P_{s+1,j}$. Since the upper limit in $i_{s+1-m}$ was reduced to $2s+1$, the remaining upper limits should also be decreased by one, finishing the proof.
[^1]: See [@Streicher:2019wek; @Lucas:2019cxr; @Lensky:2020ubw] for further work on operator growth in the context of SYK and [@Mousatov:2019xmc] for a more generic discussion in holographic theories.
[^2]: As we will properly discuss in section (\[sec:recursion\]), this statement should not be associated with a similar statement in the context of the recursion method.
[^3]: The analysis of more generic smeared operators just follow from linearity as we comment further below.
[^4]: In the context of spin systems, a “simple” operator is defined as one involving the product of an $\mathcal{O}(1)$ number of spins.
[^5]: For a discussion on large-N factorization in chaotic theories and its relevance to ETH, see [@Papadodimas:2012aq; @deBoer:2018ibj; @deBoer:2019kyr].
[^6]: This is typical in probability theory. We can define a random variable by its associated probability distribution, or equivalently by giving all its moments. From a physical perspective, the second option is better since the moments are the ones being measured.
[^7]: In QFT care has to be taken when defining these objects, but modular time evolution is well and unambiguously defined [@Haag:1992hx]. See [@Witten:2018lha] for a recent review.
[^8]: Operator growth in the context of modular time evolution has also been considered recently in [@deBoer:2019uem].
[^9]: See the book [@Haag:1992hx] for a physics introduction, the summary done in Ref. [@Papadodimas:2013wnh], or Ref. [@Witten:2018lha] for a recent review.
[^10]: Our discussion focuses on large-N theories, but it also applies to free QFT in Rindler space given the algebraic equivalence between Rindler operators and the Fourier modes $\mathcal{O}_{\omega,\vec{k}}$ and $\mathcal{O}^\dagger_{\omega,\vec{k}}$, as explicitly discussed in section \[sec:algebra\]. To our knowledge, operator growth in Rindler space has not been considered in the literature and it is useful to gauge away some of the confusions arising when defining the notion of operator growth in QFT.
[^11]: The word *state* refers to a linear functional acting on the algebra $\mathcal{A}$ as properly defined in . This is the standard terminology used in algebraic QFT [@Haag:1992hx].
[^12]: Notice that this can be straightforwardly generalized to states invariant under so-called modular time evolution. Also, notice that this definition is not restricted to time-independent states. Starting with the invariant one we can move to other states by using elements of the algebra. These states would then evolve as it is described.
[^13]: Notice there is an infinite number of choices that also annihilate the thermofield double, including the Minkowski annihilation/creation operators.
[^14]: Here we are referring to the Minkowski energy choice. The Unruh energy is not a sensible choice since it is infinite for any Minkowski mode.
[^15]: When adopting this definition, one is assigning some kind of locality interpretation to the 1d chain which was manifest in our Majorana fermion discussion . Here, the label $k$ technically accounts for the number of commutators with $H$ that have acted upon the starting operator. Thus, depending on the interactions in this hamiltonian, the support of the different $f_k$ operators will grow accordingly. Since one is interested in quantifying the evolution in the size of this support, one could consider more general functionals $\sum\limits_{k=0}^{\infty}g(k)\,\frac{\vert (f_{k},\mathcal{O}(t))\vert^{2}}{(f_{k},f_{k})}$ capturing this quantity more precisely. Our point here is that the existent inner product allows to define a notion of average size for any relevant choice of $g(k)$.
[^16]: It is understood that whenever the subindex labels $r$ in $i_r$ equal zero, such terms do *not* contribute to the solution.
[^17]: As shown in appendix \[inner-products\], this exponential growth with the right Lyapunov exponent only applies to the representative inner product considered in . For other inner products, one gets exponential growths with faster rates. This statement seems similar to the results found in [@Romero-Bermudez:2019vej] for out-of-time-ordered correlation functions, which show that different choices of euclidean separations in the OTOC 4-pt function might lead to faster growth than the chaos bound [@Maldacena:2015waa].
[^18]: This might be related to the complexity variation $\mathcal{C}_{\ket{\psi (t)}} - \mathcal{C}_{\ket{\psi_{\mathcal{O} }(t)}}$ considered in [@Barbon:2019tuq]. Variations in quantum circuit complexity have also been considered recently in [@Bueno2019; @Bernamonti:2019zyy].
[^19]: The connection between cost functions and operator size was recognized already in [@Magan:2018nmu] for spin systems, but equation shows it holds more generally.
[^20]: This construction and the Minkowski energy choice for the complexity cost provides a specific realization of the idea put forward in [@Magan:2018nmu], in which the cost function was argued to be related to the scaling dimension of the associated perturbation.
[^21]: There is a missing phase in the equation, due to the non-commutativity between $q$ and $p$. It is not included here because it is second order in the infinitesimal perturbations $\delta q ,\delta p$.
[^22]: The dependence of operator growth on the two-point function was noticed in Ref. [@Magan:2016ehs], for the case of the operator growth of a “complex” operator in the vacuum.
[^23]: More precisely $\mathcal{H}_{\omega}$ is defined as the completion of $\mathcal{A}/\mathcal{I}$ with respect to the norm topology.
[^24]: For a general full rank state $\rho$, the evolution which is naturally defined by this construction is the so-called modular evolution. If $\rho= e^{-H}$, with $H$ the modular Hamiltonian, then $H$ is the generator of modular time evolution.
[^25]: We assume the operator $\mathcal{O}$ to have vanishing one point function for visual clarity. Including in the discussion one-point functions is trivial since they do not depend on time.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
Recently, a new model of propagation of the light through the so-called weakly three-dimensional Cole-Cole nonlinear medium with short-range nonlocality has been proposed. In particular, it has been shown that in the geometrical optics limit, the model is integrable and it is governed by the dispersionless Veselov-Novikov (dVN) equation.
Burgers-Hopf equation can be obtained as 1+1-dimensional reduction of dVN equation. We discuss its properties in the specific context of nonlinear geometrical optics. An illustrative explicit example is considered.
PACS numbers: 02.30.Ik, 42.15.Dp\
Key words: Nonlinear Optics, Integrable Systems.
author:
- |
Boris Konopelchenko[^1] and Antonio Moro$^{\ast}$\
[*Dipartimento di Fisica dell’Università di Lecce*]{}\
[*and INFN, Sezione di Lecce, I-73100 Lecce, Italy*]{}\
title: 'Light propagation in a Cole-Cole nonlinear medium via Burgers-Hopf equation.'
---
Many recent studies concerning the dispersionless integrable systems have shown their relevance in a broad variety of fields in physics such as Laplacian growth, topological field theory, nonlinear optics [@Interface_tau]-[@Moro2], as well as their deep connection with various fields in mathematics, in particular, with the theory of asymptotic expansions, conformal and quasiconformal mappings and the study of integrable deformations of complex algebraic curves [@LaxLev]-[@Konalgebraic].
In the present paper we are interested in the application of the theory of dispersionless integrable systems in nonlinear geometrical optics. In particular, it has been shown in [@Moro1] that the Maxwell equations describing the propagation of the light through the Cole-Cole weakly three-dimensional nonlinear medium admit an integrable geometrical optics limit.
In this limit the system is governed by the dispersionless Veselov-Novikov (dVN) hierarchy, which is amenable by the quasiclassical $\bar{\partial}$-dressing method [@Moro2].
A reduction method based on the symmetry constraints [@Bogdanov1] appears to be an efficient approach to calculate solutions of the dVN equation [@Moro3].
In what follows we will focus on a 1+1-dimensional reduction when the refractive index does not depend on one coordinate. In this case dVN hierarchy is reduced to the so-called Burgers-Hopf (BH) hierarchy. For sake of simplicity we will focus on the first equations of the hierarchy, the so-called dVN and Burgers-Hopf equations, respectively.
In this regard, it is quite interesting to interpret some of the properties of BH equation within the specific context of nonlinear geometrical optics. One of them concerns the existence of breaking wave solutions. They could be useful to model dielectrics which present a kind of “impurities” inducing abrupt variations of the refractive index. Indeed, it can be seen that, at so-called breaking points, the curvature of the light rays blows up just like it happens at interface between different media.
Let us start with the Maxwell equations in a dielectric medium $$\begin{gathered}
\label{maxwell}
\begin{aligned}
&\nabla \wedge {\bf H} - {\frac{\partial {\bf D}}{\partial t}}=0 \quad{}\nabla \cdot {\bf D} = 0 \\
&\nabla \wedge {\bf E} + {\frac{\partial {\bf B}}{\partial t}}=0 \quad{}\nabla \cdot
{\bf B} = 0
\end{aligned}\end{gathered}$$ along with the [*constitutive equations*]{} $$\label{materials} {\bf D} = \varepsilon {\bf E},
\quad{}{\bf
B} = \mu {\bf H}.$$ Looking for monochromatic solutions $$\begin{gathered}
\label{full_fields}
\begin{aligned}
&{\bf E}\left(x,y,z,t\right) = {\bf
E_{0}}\left(x,y,z\right) e^{-i\omega t} \\
&{\bf H}\left(x,y,z,t\right) = {\bf
H_{0}}\left(x,y,z\right)e^{-i\omega t},
\end{aligned}\end{gathered}$$ one gets from (\[maxwell\]-\[materials\]) the following well-known second order equations $$\begin{gathered}
\label{second}
\begin{aligned}
\nabla^{2}{\bf E_{0}} + \omega^{2} \mu
\varepsilon {\bf E_{0}} + \left (\nabla \log~\mu \right) \wedge
\left(\nabla \wedge {\bf E_{0}} \right ) +
\nabla \left ({\bf E_{0}} \cdot \nabla \log~\varepsilon \right) &= 0 \\
\nabla^{2}{\bf H_{0}} + \omega^{2} \mu \varepsilon {\bf H_{0}} +
\left (\nabla \log~\varepsilon \right) \wedge \left(\nabla \wedge
{\bf H_{0}} \right ) + \nabla \left ({\bf H_{0}} \cdot \nabla
\log~\mu \right) &= 0.
\end{aligned}\end{gathered}$$ Medium under study is characterized by the following set of properties [@Moro1; @Moro2]:
1. $\varepsilon$ and $\mu$ obeys the [*Cole-Cole dispersion law*]{} $$\begin{gathered}
\label{ColeCole}
\begin{aligned}
\varepsilon = \varepsilon_{0} +
\frac{\tilde{\varepsilon}}{1+\left (i
\omega \tau_{0} \right)^{2\nu}}, \\
\mu = \mu_{0} +
\frac{\tilde{\mu}}{1+\left (i
\omega \tau_{0} \right)^{2\nu}},
\end{aligned}
\quad{}\quad{}0< \nu < \frac{1}{2}\end{gathered}$$ We highlight that the range of values of the exponent $\nu$ plays a crucial role in the construction of an integrable high frequency limit.
2. $\varepsilon_{0}$ and $\mu_{0}$ depend only on the coordinates $x$, $y$ and $z$, while $\tilde{\varepsilon}$ and $\tilde{\mu}$ are assumed to be depending on the coordinates, the fields and the spatial derivatives of the fields. The latter can be explained in terms of a mechanism of short-range non-locality by means of an integral constitutive relation among the electric field $\bf{E}$ and the displacement vector $\bf{D}$ [@Moro2].
3. All quantities, in high frequency limit show slow dependence on the variable $z$ formally given by $${\frac{\partial }{\partial z}} = \omega^{-\nu} {\frac{\partial }{\partial \xi}}.$$ where $\xi$ is a “slow” variable defined by $z = \omega^{\nu} \xi$.
Moreover, any of such a quantity $f(x,y,z)$ can be expanded in asymptotic series on the parameter $\omega^{\nu}$ $$f(x,y,z) = f(x,y,\xi) + \omega^{-\nu} f_{1}(x,y,\xi) + \omega^{-2 \nu}
f_{2}(x,y,\xi) + \dots .$$
Representing $$\label{fields}
{\bf E}_{0} = {\bf \tilde{E}}_{0} e^{i \omega S}, \quad{}{\bf H}_{0} =
{\bf \tilde{H}}_{0} e^{i \omega S},$$ and using the previous assumptions in one of equations (\[second\]), in high frequency limit ($\omega \to \infty$) one gets in the orders $\omega^{2}$ and $\omega^{2 - 2 \nu}$ the main non-trivial contributions [@Moro1; @Moro2] $$\begin{aligned}
\label{dVN_phase1}
&S_{x}^{2} + S_{y}^{2} = 4 u, \\
\label{dVN_phase2}
&S_{\xi} = \varphi \left (x,y,\xi,S_{x},S_{y} \right ).\end{aligned}$$ Equation (\[dVN\_phase1\]) is the standard eikonal equation in two dimensions, $\varphi$ is a certain function and $4 u = \varepsilon_{0} \mu_{0}$. As discussed in the paper [@Moro2] the equations (\[dVN\_phase1\]) and (\[dVN\_phase2\]) constitute an overdetermined system for the phase $S$. The compatibility condition imposes specific restrictions on the possible forms of the function $\varphi$ and the admissible refractive indices $u$. If $\varphi$ is a polynomial differential of $S$, the first non-trivial case is given by the third degree polynomial, i.e. $$\varphi = \frac{1}{4} S_{x}^{3} - \frac{3}{4} S_{x} S_{y}^{2} + V_{1}
S_{x} + V_{2} S_{y}.$$ The compatibility condition gives $$\begin{gathered}
\label{dVN}
\begin{aligned}
&u_{\xi} = \left(V_{1} u \right)_{x} + \left(V_{2} u \right)_{y} \\
&V_{1x} - V_{2y} = - 3 u_{x} \\
&V_{2x} + V_{1y} = 3 u_{y},
\end{aligned}\end{gathered}$$ which is the dVN equation. In the paper [@Moro3] it has been shown how symmetry constraints allow us to construct its 1+1-dimensional reductions of hydrodynamic type. A number of explicit solutions for the complex dVN equation have been discussed in [@Moro2].
Particular situation in which refractive index depend only on one coordinate on the plane $x$-$y$ is of physical interest too. So, let us assume that $$\label{Dy}
u_{y} = 0.$$ As a consequence of the eikonal equation (\[dVN\_phase1\]) the phase function $S$ must satisfy the condition $S_{y} = c$, with $c = \text{const}$, then $$S = c y + \tilde{S}(x,\xi)$$ where $\tilde{S}$ does not depend on $y$. Eikonal equation (\[dVN\_phase1\]) becomes $$\tilde{S}_{x}^{2} = 4 u - c^{2},$$ while, choosing $V_{2} = 0$, one gets $$V_{1} = - 3 u,$$ and the dVN equation (\[dVN\]) is reduced to the Burgers-Hopf (BH) equation $$\label{Burgers} u_{\xi} + 6 u u_{x} = 0.$$ Solutions of BH equation can be expressed in the following hodograph form [@whithambook] $$\label{hodograph} x - 6 u \xi + \psi (u) = 0,$$ where $\psi$ is an arbitrary function of its argument. Moreover, once calculated $u(x,\xi)$ by the (\[hodograph\]) the phase function is given explicitly $$\label{phase} S = c y \pm \int \sqrt{4 u - c^{2}} dx.$$ Shock structure of the Burgers-Hopf equation is connected to the existence of a value $\xi^{\ast}$ ([*breaking point*]{}), for which $$u_{x} \to \infty.$$
![[]{data-label="break"}](break.eps){width="12cm"}
In the theory of partial differential equations these solutions are called [*breaking waves*]{}. It is a textbook exercise to calculate the breaking point [@whithambook]. For instance, choosing the initial datum $u(x_{0},0) = (1 - \tanh x_{0})/6$ the breaking point is $\xi^{\ast}
= 1$ (see figure \[break\]).
![[]{data-label="realref"}](es1_DensReu.eps){width="6cm"}
![[]{data-label="surface"}](es1_RefrImu.eps "fig:"){width="6cm"} ![[]{data-label="surface"}](es1_phaseIm2.eps "fig:"){width="6cm"}
We remark that, as discussed in [@kodkon], the singular sector of BH hierarchy induces a stratification of the affine space of independent variables which gives rise to the integrable deformations of hyperelliptic curves.
The propagation of the light on the plane $\pi : \xi = \xi_{0}$ ($\xi_{0} = \text{const}$) is governed by the standard eikonal equation (\[dVN\_phase1\]). Denoting by $\Sigma$ the congruence on $\pi$ normal to the family of curves $S(x,y, \xi_{0}) =
\lambda$, parametrized by $\lambda$, the curvature of $\gamma \in
\Sigma$ can be represented by the well known formula [@Born] $$\label{curvature1}
\kappa = \frac{1}{2 u} \; {\bf \nu} \cdot \nabla u,$$ where $\nabla = \left(\partial_{x},\partial_{y} \right)$ and ${\bf
\nu} = (\nu_{1}, \nu_{2})$ is the unit principal normal to $\gamma$ on the plane $\pi$. Now, let us observe that the curve $\gamma$ coincides locally with the projection of the light ray on the plane $\pi$.
![[]{data-label="wave"}](es1_wave.eps){width="5cm"}
In virtue of the condition (\[Dy\]), the formula (\[curvature1\]) assumes the form $$\label{curvature2} \kappa = \frac{\nu_{1}}{2 u} \; u_{x}.$$ Then, if $u_{x}$ blows up we conclude that the curvature of the light ray blows up as well. This is the typical behavior which light rays exhibit on the interface between different materials. This type of solutions could be useful to describe situations where light rays propagate in the non-homogeneous medium and at certain point they cross some kind of impurity. These impurities with drastically different optical properties could be responsible of the abrupt change of direction.
As well known in electrodynamics, complex-valued refractive indices also have a physical meaning since they can be used to describe absorption effects of radiation. In this specific case, complex values of the refractive index $n$ are associated with a strong damping of the electromagnetic wave. We illustrate this fact by the following explicit example.
Let us consider a solution obtained from the hodograph relation (\[hodograph\]) setting $\psi(u) = u^{2}$. Expliciting $u$ from (\[hodograph\]) and using it in the expression (\[phase\]), one gets $$\begin{aligned}
&u = \frac{1}{2} \left(- 6 \xi + \sqrt{36 \xi^{2} - 4 x} \right), \\
&S = \frac{4}{15} \sqrt{2} \sqrt{- 6 \xi + \sqrt{36 \xi^{2} - 4
x}} \left(6 \xi \left(- 6 \xi + \sqrt{36
\xi^{2}- 4 x} \right) + 12 x \right).\end{aligned}$$ We note that the refractive index $n = 2 \sqrt{u}$ (we do not consider negative refractive indices) is not real-valued on whole plane $x$-$\xi$. Inside regions in which the refractive index takes a complex contribution even the phase function $S$ becomes complex-valued. Thus, writing $S = S_{1} + i S_{2}$, if $S_{2} > 0$ the electric field acquires a damping factor $$\label{damping}
{\bf E} = {\bf E}_{0} e^{- \omega S_{2}} \; e^{i \omega S_{1}}.$$ In this regions strong absorptions effects are occurring.
Figure \[realref\] shows the real part of refractive index $n = 2 \sqrt{u}$. The dashed line marks the region where the refractive index acquires an imaginary contribution. This has been visualized in the three-dimensional plot in the figure \[surface\]-a. Just where refractive index becomes complex-valued, the phase $S$ takes an imaginary contribution indicating a strong absorption (see figure \[surface\]-b).\
Figure \[wave\] shows the wave fronts configuration in the plane $x$-$\xi$. Shaded region represents the absorption region. For $\xi \to -\infty$ we have an almost plane wave front. Towards absorption region, wave fronts deform itself following level lines of refractive index. Just before absorption wave fronts present a deviation of an almost right angle.
[99]{}
I.K. Kostov, I. Krichever, M. Mineev-Weinstein, P.B. Wiegmann, A. Zabrodin, $\tau$-function for analytic curves, Proceedings of [*Introductory workshop on Random Matrix Models and their applications*]{}, Berkeley, California (1999).
[M. Mineev-Weinstein, P.B. Wiegmann and A. Zabrodin, Integrable structure of interface dynamics, [*Phys. Rev. Lett.*]{}, [**84(22)**]{}, 5106-5109, (2000)]{}.
[P.B. Wiegnamm and A. Zabrodin, Conformal maps and integrable hierarchies, [*Comm. Math. Phys.*]{}, [**213**]{}, 523-538, (2000)]{}.
I. Krichever, M. Mineev-Weinstein, P.B. Wiegmann and A. Zabrodin, [*Laplacian growth and Whitham equations of soliton theory*]{}, [arXiv:nlin.SI/0311005]{} (2003).
[I.M. Krichever, The dispersionless Lax equations and topological minimal models, [*Comm. Math. Phys.*]{}, [**143**]{}, 415-429, (1992)]{}.
[I.M. Krichever, The $\tau$-function of the universal Whitham hierarchy, matrix models and topological field theories, [*Commun. Pure Appl. Math.*]{}, [**47**]{}, 437-475, (1994)]{}.
[B.A. Dubrovin, Hamiltonian formalism of Whitham-type hierarchies and topological Landau-Ginzburg models , [*Comm. Math. Phys.*]{}, [**145**]{}, 195-207, (1992)]{}.
[B.A. Dubrovin, Integrable systems in topological field theory, [*Nucl. Phys. B*]{}, [**379**]{}, 627-689, (1992)]{}.
[Boris G. Konopelchenko and Antonio Moro, Geometrical optics in nonlinear media and integrable equations, [*J. Phys. A: Math. Gen.*]{}, [**37**]{}, L105-L111, (2004)]{}.
Boris Kononopelchenko and Antonio Moro, [*Integrable equation in nonlinear geometrical optics*]{}, to appear in Stud. Appl. Math., [arXiv:nlin.SI/0403051]{}.
[P.D. Lax and C.D. Levermore, The small dispersion limit on the Korteweg-de-Vries equation, [*Commun. Pure Appl. Math.*]{}, [**36**]{}, 253-290, 571-593, 809-830, (1983)]{}.
[B. Konopelchenko and L. Martinez Alonso, $\bar{\partial}$-equations, integrable deformations of quasi-conformal mappings and Whitham hierarchy, [*Phys. Lett. A*]{}, [**286**]{}, 161-166, (2001)]{}.
[B.G. Konopelchenko and L. Martinez Alonso, Dispersionless scalar integrable hierarchies, Whitham hierarchy and the quasi-classical $\bar{\partial}$-dressing method, [*J. Math. Phys*]{}, [**43(7)**]{}, 3807-3823, (2002)]{}.
B. Konopelchenko and L. Martinez Alonso, [ *Integrable quasiclassical deformations of algebraic curves*]{}, [ arXiv:nlin.SI/0403052]{} (2004).
L.V. Bogdanov, B.G. Konopelchenko, [*Symmetry constraints for dispersionless integrable equations and systems of hydrodynamic type*]{}, [arXiv:nlin.SI/0312013.SI]{}.
L. Bogdanov, B.G. Konopelchenko and A. Moro, [*Symmetry constraints for real dispersionless Veselov-Novikov equation*]{}, to appear on Fund. Prikl. Mat. (russian version), J. Math. Sci. (english version), preprint [arXiv:nlin.SI/0406023]{} (2004).
[G.B. Whitham, [*Linear and nonlinear waves*]{}, [Wiley Interscience Series]{}, (1974)]{}.
[Y. Kodama, B. Konopelchenko, Singular sector of the Burgers-Hopf hierarchy and deformations of hyperelliptic curves, [*J. Phys. A: Math. Gen.*]{}, [**35**]{}, L489-L500, (2002)]{}.
[M. Born and E. Wolf, [*Principles of Optics*]{}, [Pergamon Press, Oxford]{}, (1980)]{}.
[^1]: Supported in part by the COFIN PRIN “SINTESI” 2002.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Following the presentation and proof of the hypothesis that image features are particularly perceived at points where the Fourier components are maximally in phase, the concept of phase congruency (PC) is introduced. Subsequently, a two-dimensional multi-scale phase congruency (2D-MSPC) is developed, which has been an important tool for detecting and evaluation of image features. However, the 2D-MSPC requires many parameters to be appropriately tuned for optimal image features detection. In this paper, we defined a criterion for parameter optimization of the 2D-MSPC, which is a function of its maximum and minimum moments. We formulated the problem in various optimal and suboptimal frameworks, and discussed the conditions and features of the suboptimal solutions. The effectiveness of the proposed method was verified through several examples, ranging from natural objects to medical images from patients with a neurological disease, multiple sclerosis.'
author:
- 'Seyed Mohammad Mahdi Alavi, and Yunyan Zhang [^1] [^2]'
title: Phase Congruency Parameter Optimization for Enhanced Detection of Image Features for both Natural and Medical Applications
---
Optimization, Phase Congruency, Image Processing, Image Analysis, Image Recognition, Feature Detection, Biomedical image processing, Magnetic Resonance Imaging, Multiple Sclerosis.
Introduction
============
Motivation and Literature Survey
--------------------------------
Image Features Detection (IFD) is an important topic in the image-processing field [@Kovesi1996]. It aims at finding image features including lines, edges, Mach bands, corners, and blobs, by using quantitative methods. Many IFD methodologies have been proposed, which can be classified into two broad categories. First, the IFD methods that are based on dimensional metrics. The majority of the published IFD methods lie into this category including dilation-erosion residue operator [@Noble1989], gradient [@Canny1983], weak membrane [@Blake1987], anisotropic diffusion [@Perona1990], and univalue segment assimilating nucleus [@Smith1997] -based methods. The main issue of the dimensional IFD methods is that they are too sensitive to image contrasts and spatial magnifications [@Kovesi1996]. Second, the IFD methods that are based on non-dimensional metrics. Phase congruency (PC) is such a method [@Morrone1986]. References [@Kovesi1996], [@Ziou1998], [@Patel2014], and [@Noble1989] review many of the proposed IFD techniques in details.
The PC-IFD method is important for several reasons [@Kovesi1996; @Kovesi1999; @Kovesi2003]: 1) It is invariant to the image contrast, because it is not based on intensity gradient; 2) No assumption is made in the PC-IFD formulation and computation; 3) It can detect various features; and 4) It uses the image phase information. The latter might be one of the most significant features of the PC-IFD, because phase is shown to be more informative than magnitude in image processing [@Oppenheim1981]. PC is developed based on the hypothesis that image features are optimally perceived at points where the Fourier series components are maximally in phase, meaning that phases of Fourier series€™ components are similar. This hypothesis was first introduced and verified for Mach bands in [@Morrone1986]. Subsequently, its effectiveness for detecting other features types has been verified in [@Morrone1986; @Venkatesh1990; @Kovesi1996; @Kovesi1999; @Kovesi2003].
It is shown in [@Venkatesh1990] and [@Morrone1987] that the maxima of the local energy occur at the maxima of PC and vice versa. Thus, in practice, the PC measure is often obtained by computing the normalized local energy, [@Morrone1987; @Venkatesh1990; @Kovesi1999; @Kovesi2003]. Moreover, it has been illustrated that the peaks of PC outcome are higher and more distinct when local energy is computed using window-based approaches than the peaks obtained from local energy over the whole signal, [@Kovesi1996]. Consequently, Peter Kovesi proposed a multi-scale PC measurement method based on Gabor wavelets and extended it to solve two dimensions (2D), [@Kovesi1996; @Kovesi1999; @Kovesi2003]. Kovesi further modified the PC formulation to overcome noise and ill-conditioning issues, and introduced a weighting function to penalize PC measures at locations where the spread of frequencies is narrow. The proposed 2D multi-scale PC (2D-MSPC) method has been the basis of many IFD studies in various fields such as history, media, basic sciences, and medicine [@Cao2006; @Struc2009; @Zhang2012; @Obara2012; @Mouats2015; @Rijal2015; @Ziaei2014].
The current computation of 2D-MSPC requires several parameters to be tuned [*a priori*]{}. In [@Kovesi1996; @Kovesi1999; @Kovesi2003], several empirical hints have been provided for parameter tuning. But, the selection of parameters either is not explicitly discussed as in [@Cao2006; @Struc2009; @Zhang2012], or is performed manually based on trail and error [@Obara2012; @Mouats2015; @Rijal2015; @Ziaei2014]. In [@Obara2012] and [@Mouats2015], the parameters were manually optimized to increase the visualization of image features. In [@Rijal2015] the parameters were changed by trial and error, and the study found that an improvement in IFD occurred with parameters that maximized the signal-to-noise ratio of the image. In [@Ziaei2014], a set of experiments was also done based on the trial and error, to determine the best, fixed values for computing the maximum moment of PC covariance. The authors reported that there was no clear rule on how the 2D-MSPC parameters should be tuned. Collectively, there is a critical need for a standard parameter tuning method for 2D-MSPC. Finding a solution for this problem can enhance our understanding of the concept of PC and thereby promoting its applications.
Contribution of This Paper
--------------------------
We have classified the 2D-MSPC optimization problem as follows:
- What optimization criteria should be used for the tuning of 2D-MSPC parameters?
- How can we formulate the problem mathematically and compute the parameters of 2D-SMPC optimally and automatically?
This paper aims to address these questions. The 2D-MSPC, its definition, computational method and a list of its fundamental parameters are briefly described in section \[sec:2dmspc\]. The importance of parameter setting in 2D-MSPC IFD, which leads to Q1 and Q2, is discussed in section \[sec:pardisc\]. In section \[sec:opt\], we define a criterion for parameter optimization, based on the maximum and minimum moments of the 2D-MSPC. We formulate the problem in various optimal frameworks, and then describe about a suboptimal solution accordingly. Finally, the effectiveness of the proposed method is verified through several examples from both natural and medical images in section \[sec:ex\]. This includes two examples from magnetic resonance imaging (MRI) of patients with multiple sclerosis (MS). We illustrate that the proposed 2D-MSPC optimization is useful to MS lesions detection.
Nomenclatures
-------------
In this paper the following general notations are used. ${\ensuremath{\mathbb{N}}}$, ${\ensuremath{\mathbb{R}}}$ and ${\ensuremath{\mathbb{C}}}$ denote the integer, real and complex domains, respectively. $\|A\|$ denotes the norm of matrix $A$, where no subscript means that any norm can be used. $\|A\|_F$, $\|A\|_p$, $\|A\|_\infty$, $\|A\|_2$ and $\|A\|_1$ represent Frobenius-norm ($F-$norm), Schatten $p-$norm, $\infty-$norm, 2-norm and 1-norm of matrix $A$. The consistent norm of matrix $A$ is denoted by $\|A\|_c$. The determinant of matrix $A$ is denoted by $\mbox{det}(A)$. The optimal value of the parameter $a$ is denoted by $a^*$ (a superscript asterisk). It should not be confused with the convolution operator $*$, which is used in this paper as $A*B$. Note that there are several other specific notations also defined in the text, which will be introduced when they appear.
Two-Dimensional Multi-Scale Phase Congruency {#sec:2dmspc}
============================================
In this section, fundamental concepts of the 2D-MSPC are briefly described. Interested readers are directed to [@Kovesi1996], [@Kovesi1999] and [@Kovesi2003] for details. The 2D-MSPC is a combination of one-dimensional PC calculated over several orientations. The 2D-MSPC based on the measurement of local energy is given by $$\begin{aligned}
\label{eq:pc}
& \mbox{PC}(x)= \frac{ \sum_{o} W_o(x) \left \lfloor E_o(x)-T_o \right \rfloor}{ \sum_{o} \sum_{n}A_{no}(x) + \varepsilon}
\\ \nonumber & o=1,2,\ldots, O
\\ \nonumber & n=1,2,\ldots, N_o\end{aligned}$$ where, $o$ and $n$ denote the orientation and scale indexes, respectively. $O$ is the total number of orientations. $N_o$ is the total number of scales at the $o-$th orientation. $\lfloor \rfloor$ is the floor operator. Note that $\lfloor y \rfloor = y$ if $y>0$, otherwise zero. $E_o$ is the local energy at the $o-$th orientation, which is calculated by using the Hilbert transform as follows $$\begin{aligned}
\label{eq:e}
E_o(x)=\sqrt{F_o^2(x)+H_o^2(x)}\end{aligned}$$ where, $F_o(x)$ and $H_o(x)$ are the AC component and Hilbert transform of the image signal $I(x)$ at the $o-$th orientation. However, because the Hilbert transform operator is an improper integral, $F_o(x)$ and $H_o(x)$ are computed by convolving the image signal with a pair of even and odd wavelets filters in quadrature as follows [@Morrone1987; @Venkatesh1990]: $$\begin{aligned}
\label{eq:fs}
&F_o(x)= \sum_{n} I(x) * M_{no}^{e}\\
\label{eq:hs}
&H_o(s)= \sum_{n} I(x) * M_{no}^{o}$$ where, $*$ denotes the convolution operator. $M_{no}^e$ and $M_{no}^o$ are even and odd filters at the $o-$th orientation and $n-$th scale, generated by the logarithmic Gabor function $$\begin{aligned}
\label{eq:lgabor1}
G_{no}(f)=\exp \left(\frac{-\left(\log\frac{f}{\hat{f}_{no}}\right)^2}{2\left( \log \sigma_{no}\right)^2}\right)$$ where, $\hat{f}_{no}$ and $\sigma_{no}$ denote the centre frequency and bandwidth of the $n-$th Gabor filer at the $o-$th iteration. The filters are scaled as follows $$\begin{aligned}
\label{eq:lgabor2}
\hat{f}_{no}=\frac{1}{\lambda_{\min o} \times \eta_o^{(n-1)}}$$ where, $\lambda_{\min o}$ is the minimum wavelength (maximum centre frequency) of wavelets in banks of $\{M_{1o}^e, \ldots, M_{No}^e\}$ and $\{M_{1o}^o, \ldots, M_{No}^o\}$, and $\eta_o$ represents the distance between successive filters in the bank. It is noted that $M_{no}^e$ and $M_{no}^o$ are identical with a $90^{\circ}$ shift. Figure \[fig:gf\] shows the spectra of 4-scale logarithmic Gabor wavelets at the $o-$th orientation with $\lambda_{\min o}=3$, $\eta_o=3$ and $\sigma_{no}=0.55$, $n=1,\ldots,4$. Note that in the logarithmic frequency scale, the spectra of all Gabor functions are identical.
![Spectra of 4-scale Gabor wavelets.[]{data-label="fig:gf"}](gff.png)
Subsequently, the normalization factor in the 2D-MSPC is computed in the wavelet’s framework as follows $$\begin{aligned}
\label{eq:nf}
\sum_{n}A_{no}=\sum_{n} \sqrt{ (I(x) * M_{no}^{e})^2+(I(x) * M_{no}^{o})^2}.\end{aligned}$$
Parameter Tuning hints
-- --------------------- --------------------------------------------------------------------------------------------------------------------------------------------------------------------------
$c_o$ [Between 0 and 1, typical 0.4 or 0.55]{}
$g_o$ [Typical value, 10]{}
$ \lambda_{\min o}$ [The smallest value is the Nyquist wavelength of 2 pixels. Because of aliasing 3 pixels or above is suggested.]{}
$\sigma_{no}$ [The smaller $\sigma_{no}$, the larger the bandwidth of the filter. The following sets are suggested: $[\sigma_{no}=0.85,\eta_o=1.3]$, or ]{}
$\eta_o$ [$[\sigma_{no}=0.75,\eta_o=1.6]$ (Bandwidth $\approx$ 1 octave), or $[\sigma_{no}=0.65,\eta_o=2.1]$, or $[\sigma_{no}=0.55,\eta_o=3]$ (Bandwidth $\approx$ 2 octaves)]{}
$k_o$ [Typical value 2, up to 10 or 20 for noisy images]{}
$\varepsilon$ [Small scalar, typically 0.0001 ]{}
$N_o$ [Try values 3 to 6 ]{}
$O$ [A filter spacing of $30^{\circ}$ has found good. ]{}
\[table:par\]
As discussed in [@Kovesi1996], the calculation of PC makes sense only in locations where the spread of frequencies is significant. In 2D-MSPC formulation, $W_o(x)$ is a weighting function, which penalizes the information at locations where the spread of frequencies is narrow at the $o-$th orientation. It is defined by a sigmoid function $$\begin{aligned}
W_o(x)=\frac{1}{1+\exp \left(g_o(c_o-s_o(x))\right)}\end{aligned}$$ where, $c_o$ denotes the cutoff (mid-point) point of the sigmoid function; $g_o$ is a gain that controls the rate of weighting; $s_o$ is a measure of frequency spread, which ranges between 0 and 1, and is given by $$\begin{aligned}
s_o(x)=\frac{1}{N_o}\frac{\textstyle \sum_{n}A_{no}(x)}{A_{\max}(x)+\varepsilon}.\end{aligned}$$ where, $\textstyle \sum_{n}A_{no}(x)$ is computed through and $A_{\max}(x)$ is the amplitude of the filter pair having maximum response at $x$.
The parameter $T_o$ is an estimation of noise power that is subtracted from the energy of the signal. It is computed based on the assumptions that the image noise is additive, that the noise power spectrum is constant, and that the image features occur at isolated locations: $$\begin{aligned}
T_o=\mu_{Ro}+k_o \sigma_{Ro}\end{aligned}$$ where, $\mu_{Ro}$ and $\sigma_{Ro}^2$ are the mean and variance of the Rayleigh that describes the noise energy response; $k_o$ is a scaling factor used to estimate the maximum degree of the noise response.
When the spread of frequencies is narrow, $E_o$ and $\sum_n A_{no}$ become very small, making the computations to become ill-conditioned. The parameter $\varepsilon$ is to address this issues.
Table \[table:par\] summarizes the parameters discussed above which are fundamental and need to be fine-tuned in the 2D-MSPC calculations, including some hints from [@Kovesi1996; @Kovesi1999; @Kovesi2003; @Kovesi_url], mainly for manual tuning.
Hereafter, we denote all parameters in a vector format, $$\begin{aligned}
\label{parvec}
\upsilon_o= \left[ \begin{array}{ccccccccc}
c_o & g_o & \lambda_{\min o} & \sigma_{no} & \eta_o & k_o & \varepsilon & N_o & O \\
\end{array}\right]\end{aligned}$$ where, $\upsilon_o$ denote the parameter vector at orientation $o$.
As usual in the numerical optimization, we assume that upper and lower limits of the parameter vector are known [*a priori*]{}. The upper and lower limits of $\upsilon$ in the vector format are denoted by $\overline{\upsilon}_o$ and $\underline{\upsilon}_o$ respectively, such that $$\begin{aligned}
\underline{\upsilon}_o \leq \upsilon_o \leq \overline{\upsilon}_o\end{aligned}$$ In this paper, $\upsilon_o^*$ denotes the optimal parameter vector and $\mbox{PC}_o^*$ denotes the PC matrix computed with the optimal parameter vector $\upsilon_o^*$, both at orientation $o$.
Description of Q1 and Q2 {#sec:pardisc}
========================
It is illustrated that visualization of the image will significantly vary by changing only one parameter of the 2D-MSPC [@Mouats2015] and [@Rijal2015]. Figure \[fig:lena\] shows the visualization of a lena’s image by using 2D-MSPC, with different values of the cutoff point $c_o$. Based on information from Table \[table:par\], other parameters of the 2D-MSPC are fixed to: $k_o=2$; $\varepsilon=0.0001$; $N_o=4$; $O=6$; $\lambda_{\min o}=3$; $\eta_o=2.1$; and $g_o=10$, for $o=1,\dots, O$ and $n=1,\ldots,N_o$. Figure \[fig:lena-orig\] shows the lena’s grayscale image of lena that was used in this simulation. Figures \[fig:lena-co-p55\] and \[fig:lena-co-p10\] show the 2D-MSPC images for $c_o=0.55$ and $c_o=0.10$, respectively. It is seen that by decreasing the cutoff value, image features become increasingly detectable. There are several features that could not be detected with $c_o=0.55$ (Figure \[fig:lena-co-p55\]). This issue gives rise to questions Q1 and Q2.
[0.25]{} ![Visualization of a lena image by using the 2D-SMPC, with different values of the cutoff point $c_o$, while other parameters are fixed.[]{data-label="fig:lena"}](lena-orig "fig:"){width="\linewidth"}
\
[0.25]{} ![Visualization of a lena image by using the 2D-SMPC, with different values of the cutoff point $c_o$, while other parameters are fixed.[]{data-label="fig:lena"}](lena-pc-co-p55 "fig:"){width="\linewidth"}
[0.25]{} ![Visualization of a lena image by using the 2D-SMPC, with different values of the cutoff point $c_o$, while other parameters are fixed.[]{data-label="fig:lena"}](lena-pc-co-p10 "fig:"){width="\linewidth"}
In the next few sections, these questions are addressed.
Parameter Optimization {#sec:opt}
======================
In this paper, several optimal frameworks are proposed for parameter tuning for the 2D-MSPC, which are based on the maximization of the PC momentums. The maximum and minimum moments of the 2D-MSPC, $M$ and $m$ respectively, are given by [@Kovesi2003]: $$\begin{aligned}
\label{maxmo}
& M=\frac{1}{2}\left(\alpha+\gamma+\sqrt{\beta^2+(\alpha-\gamma)^2}\right)\\
\label{minmo}
& m=\frac{1}{2}\left(\alpha+\gamma-\sqrt{\beta^2+(\alpha-\gamma)^2}\right)\end{aligned}$$ with $$\begin{aligned}
\label{alpha}
& \alpha=\textstyle \sum_{o}\left(\mbox{PC}_o \cos(\theta_o)\right)^2\\
\label{gamma}
& \gamma=\textstyle \sum_{o}\left(\mbox{PC}_o \sin(\theta_o)\right)^2\\
\label{beta}
& \beta=2 \textstyle \sum_{o}\left(\mbox{PC}_o \cos(\theta_o)\right)\left(\mbox{PC}_o \sin(\theta_o)\right).\end{aligned}$$ where, $\mbox{PC}_o$ and $\theta_o$ represent the PC and axis angle at orientation $o$.
The calculation of momentums corresponds to performing a singular value decomposition to the PC covariance matrix, and thus momentums correspond to the singular values, [@Kovesi2003]. On the other hand, singular values indicate the level of increase in energy that can occur between the input and output of a given system. The larger the maximum singular value of a system, the greater the increase in energy, (see Definition 40.2 and descriptions in §40.2.1, p. 652,[@Levine1995]). Therefore, it may be deduced that momentums represent the maximum of local energies. It was also shown that the maxima of local energy are indications of features in an image, [@Venkatesh1990]. The answer to Q1 can now be simplified as such that a potential criterion for parameter tuning of 2D-MSPC should be given in terms of $M$ and $m$.
From the potential relationships between momentums, singular values, local energies, 2D-MSPC and image features, it may be deduced that the image features detection can be enhanced by increasing the maximum and minimum moments of the PC. In other words, optimal values of 2D-MSPC parameters are those that maximize $M$ and $m$. Both $M$ and $m$ are matrices and their size is equal to the size of $\mbox{PC}_o$. Thus, the optimal parameters of 2D-MSPC should be obtained by solving a matrix optimization problem (an answer to Q2).
In order to formulate the problem, we define a cost function $\mathcal{M}$:$$\begin{aligned}
\label{Mm}
\mathcal{M}=\mu_1 \times M+\mu_2 \times m\end{aligned}$$ where, $\mu_i \in [0,1]$, $i=1,2$. The cost function enables us to not only maximize both $M$ and $m$, but also to manage their interactions.
As mentioned above, the question of parameter tuning for 2D-MSPC is a matrix optimization problem. Matrix maximization has widely been used in optimal (experimental) designs [@Fredov1972; @Goodwin1977; @Atkinson1982; @Pronzato1985; @Berger1994; @Pukelsheim1993], where the Fisher information matrix [@Fisher1971] is maximized in order to reduce the Cram' er-Rao bound [@Cramer1946; @Rao1945]. The Fisher information matrix is maximized by maximizing some real-valued summary statistics (optimal) criteria), [@Atkinson2007]. The majority of matrix optimization methods are based on the maximization of the determinant of a matrix, referred to as D-optimal problem [@Berger1994; @StJohn1975]. The rest of the methods are mainly based on the maximization of the trace or maximization of the minimum eigenvalue of a matrix.
In the D-optimal framework, maximization of the cost function is given by $$\begin{aligned}
\begin{aligned}\label{D-opt}
& \underset{\upsilon_o}{\text{minimize}}
& & -\mbox{det}\Big(\mathcal{M}(\upsilon_o)\Big) \\
& \text{subject to}
& & \underline{\upsilon}_o \leq \upsilon_o \leq \overline{\upsilon}_o\\
\end{aligned}\end{aligned}$$ where $\det(\mathcal{M})$ denotes the determinant of $\mathcal{M}$. To be consistent with the optimization literature, the maximization of $z$ is written as the minimization of $-z$ throughout this paper. The solution to this optimization problem will provide the optimal parameters for 2D-MSPC.
For the PC-based image processing, we propose another optimization approach based on the norm of $\mathcal{M}$ as follows: $$\begin{aligned}
\begin{aligned}\label{norm-opt}
& \underset{\upsilon_o}{\text{minimize}}
& & -\left\| \mathcal{M}(\upsilon_o) \right\| \\
& \text{subject to}
& & \underline{\upsilon}_o \leq \upsilon_o \leq \overline{\upsilon}_o\\
\end{aligned}\end{aligned}$$ where $\| \mathcal{M}\|$ denotes the norm of $\mathcal{M}$. The matrix norm definition is given in Appendix \[matrixnorm\].
For $\mu_1=\mu_2=1$, $\mathcal{M}=M+ m= \textstyle \sum_o PC_o^2$, and the D-optimal and norm-optimal problems and are transferred respectively to $$\begin{aligned}
\begin{aligned}\label{D-opt2}
& \underset{\upsilon_o,~o=1,\ldots,O}{\text{minimize}}
& & -\mbox{det}\Big(\mbox{PC}_o^2(\upsilon_o)\Big) \\
& \text{subject to}
& & \underline{\upsilon}_o \leq \upsilon_o \leq \overline{\upsilon}_o\\
& & & \mathcal{M}=M+m\\
\end{aligned}\end{aligned}$$ and, $$\begin{aligned}
\begin{aligned}\label{norm-opt2}
& \underset{\upsilon_o,~o=1,\ldots,O}{\text{minimize}}
& & -\left\| \mbox{PC}_o^2(\upsilon_o) \right\| \\
& \text{subject to}
& & \underline{\upsilon}_o \leq \upsilon_o \leq \overline{\upsilon}_o\\
& & & \mathcal{M}=M+m\\
\end{aligned}\end{aligned}$$
Differences between the optimization sets { and } and { and } are the followings.
- In order to solve or , at each iteration of the optimization, the PC matrices of all orientations and $\mathcal{M}$ must be computed. Solving and does not require $\mathcal{M}$ at each iteration of the optimization.
- The solution to or results in one optimal parameter vector for all orientations. The solution to or results in $O$ numbers of optimal parameter vectors (one optimal vector for each orientation).
A Sub-Optimal Solution to the Norm-Optimal Problem
--------------------------------------------------
In order to avoid the computation of $\mathcal{M}$ at each iteration of the optimization, a sub-optimal criterion of is derived, which optimizes the parameter vector separately at each orientation. Note that the results of this subsection are valid if the matrix norm is consistent and $\mu_i \neq 0$ and $\mu_i \neq 1$ , $i=1,2$. Based on the literature on optimal designs [@Fredov1972; @Goodwin1977; @Atkinson1982; @Pronzato1985; @Berger1994; @Pukelsheim1993], the covariance matrix satisfies the following inequality $$\mbox{cov}(\upsilon) \geq \mathcal{F}^{-1}(\upsilon)$$ where, $\mbox{cov}$ denotes the covariance matrix, $\upsilon$ is the parameter vector and $\mathcal{F}$ denotes the Fisher information matrix. In this case, $\mathcal{F}^{-1}$ is the lower bound of the covariance of estimations (known as the Cram' er-Rao bound [@Cramer1946; @Rao1945]). It is a common approach that in order to reduce the covariance of estimations, the the Cram' er-Rao bound is reduced by maximizing the Fisher information matrix.
In this section, we obtain an upper bound of the norm of $M$ and $m$. We then derive a sub-optimal solution, which is based on the maximization of the upper bound of $M$ and $m$. First, we need the matrix norm to be consistent.
([@Lyche2012], §8.1.1, p. 178) A matrix norm is [*consistent*]{} if it is defined on ${\ensuremath{\mathbb{C}}}^{q\times r}$ for all $q,r\in {\ensuremath{\mathbb{N}}}$ and the sub-multiplicative property $$\begin{aligned}
\label{ableqab}
\| AB \| \leq \|A\| ~\|B\|\end{aligned}$$ holds for all matrices $A$ and $B$ for which the product $AB$ is defined.
In this paper, the consistent norm is denoted by a subscript, $\|.\|_c$.
([@Lyche2012], [@Schatten1960]) \[lemakf\] The Frobenius and all Schatten $p-$norms are sub-multiplicative, i.e., $$\begin{aligned}
\label{ableqab}
\| AB \|_F &\leq \|A\|_F ~\|B\|_F\\
\| AB \|_p & \leq \|A\|_p ~\|B\|_p\end{aligned}$$
\[lempc\] Let $\mbox{PC}_o$ denotes the PC at orientation $o$ and the 2D-MSPC is computed over $O$ orientations. Then, the following relationships hold between the maximum and minimum moments and $\mbox{PC}_o$. $$\begin{aligned}
M&=\frac{1}{2}\Big(\textstyle \sum_{o} \mbox{PC}_o^2 +\sqrt{\textstyle \sum_{o} \mbox{PC}_o^4 }\Big)\\
m&=\frac{1}{2}\Big(\textstyle \sum_{o} \mbox{PC}_o^2-\sqrt{\textstyle \sum_{o} \mbox{PC}_o^4 }\Big)\end{aligned}$$ [*Proof*]{}: Appendix \[p-lempc\].
\[thpcbnd\] Let $\mbox{PC}_o$ denotes the PC at orientation $o$. The 2D-MSPC is computed over $O$ orientations. If a consistent matrix norm is used, then $$\begin{aligned}
\label{Mbnd}
&\left\| M \right\|_c \leq \textstyle \sum_{o} \left\| \mbox{PC}_o^2 \right\|_c\\
\label{mbnd}
&\left\| m \right\|_c \leq \textstyle \sum_{o} \left\| \mbox{PC}_o^2 \right\|_c\end{aligned}$$ where, $c$ can take $F$ or $p$. If the subscript $c$ is replaced with $F$, the $F-$norm is applied. If the subscript $c$ is replaced with $p$, the Schatten $p-$norm is applied. : Appendix \[p-pcbnd\].
\[thsubopt\] A sub-optimal solution to the optimization problem with the consistent norm is obtained by solving the following optimization problem. $$\begin{aligned}
\begin{aligned}\label{norm-subopt}
& \underset{\upsilon_o,~o=1,\ldots,O}{\text{minimize}}
& & -(|\mu_1|+|\mu_2|)\left\| \mbox{PC}_o^2(\upsilon_o) \right\|_c \\
& \text{subject to}
& & \underline{\upsilon}_o \leq \upsilon_o \leq \overline{\upsilon}_o\\
\end{aligned}\end{aligned}$$ [*Proof*]{}: Appendix \[p-suboptlem\].
The differences between and are as follows:
- The solution to is obtained by the parameter vector optimization individually at each orientation, while the solution to is a parameter vector optimal for all orientations.
- The optimization problem does not require the computation of $\mathcal{M}$ at each iteration of optimization. To solve , the cost function $\mathcal{M}$ is computed at each iteration of optimization.
The differences between and are as follows:
- The norm in can be of any type, but in , it has to be consistent and satisfies the sub-multiplicative property.
- In , $\mathcal{M}=M+m$, but $\mathcal{M}$ can be any combinations of $M$ and $m$ in .
Norm Selection
--------------
Any norm type can be used in and . However, if $\mu_i \neq 0$ and $\mu_i \neq 1$, $i=1,2$, only can be applied, not . The prerequisite to apply is that the sub-multiplicative property must hold for the norm.
There are also some other features associated with the norm definitions that might be important. If 1- or $\infty$-norms of the matrix is chosen, because they are associated with the maximum absolute column and row of the matrix, certain features, which are not along that row or column, might be overlooked. The 2-norm, Frobenius-norm ($F-$norm), or Schatten $p-$norm might be more efficient, because they are based on the eigenvalues and singular values, which correspond to the energy of a signal. The 2-norm is the maximum singular value of a matrix: $\|A\|_2=\sigma_{\max}$ [@Lyche2012]. The $F-$norm is the square root of the summation of the singular values: $\|A\|_F=(\textstyle \sum_i \sigma_i^2)^{1/2}$ [@Lyche2012], and the Schatten $p-$norm is $\|A\|_p=(\textstyle \sum_i \sigma_i^p)^{1/p}$ [@Schatten1960]. Thus, it can be deduced that the 2-norm may amplify the single most-dominant feature, while the $F-$ and $p-$ norms may strengthen features by taking their summation.
In summary, the 2D-MSPC parameters can be optimally and automatically tuned by maximizing , through or . If $\mathcal{M}=M+m$ is going to be maximized, and or can be used. If $\mu_i \neq 0$ and $\mu_i \neq 1$, $i=1,2$, can provide a sub-optimal solution to . This answers Q2.
Orientation $c_o^*$ $g_o^*$ Cost function opt. value
-- ---------------- --------- --------- ------------------------------------------------------
$o=1,\dots,6$ 0.1212 45.4827 $\mbox{det}(\mathcal{M})^*=7.0088\times 10^{-55}$
$o=1,\dots,6$ 0.0100 50.0000 $\|\mathcal{M}\|_F^*=114.1408$
$o=1$ 0.1001 49.9997 $\mbox{det}(\mbox{PC}_1^2)^*=6.4589\times 10^{-241}$
$o=2$ 0.1021 39.9334 $\mbox{det}(\mbox{PC}_2^2)^*=1.7131\times 10^{-33}$
$o=3$ 0.1216 46.6111 $\mbox{det}(\mbox{PC}_3^2)^*=1.4946\times 10^{-43}$
$o=4$ 0.1946 42.9731 $\mbox{det}(\mbox{PC}_4^2)^*=3.2306\times 10^{-218}$
$o=5$ 0.1566 39.2954 $\mbox{det}(\mbox{PC}_5^2)^*=1.0452\times 10^{-37}$
$o=6$ 0.1492 46.3088 $\mbox{det}(\mbox{PC}_6^2)^*=1.7254\times 10^{-53}$
$o=1,\ldots,6$ 0.1374 44.1869 $\mbox{det}(\mathcal{M})^*=\mbox{inf}$
$o=1$ 0.0100 50.0000 $\|\mbox{PC}_1^2\|_F^*=8.3413\times 10^{3}$
$o=2$ 0.0100 50.0000 $\|\mbox{PC}_2^2\|_F^*=6.3400\times 10^{3}$
$o=3$ 0.0100 50.0000 $\|\mbox{PC}_3^2\|_F^*=3.2968\times 10^{3}$
$o=4$ 0.0100 49.9998 $\|\mbox{PC}_4^2\|_F^*=2.9181\times 10^{3}$
$o=5$ 0.0100 50.0000 $\|\mbox{PC}_5^2\|_F^*=4.7028\times 10^{3}$
$o=6$ 0.0100 50.0000 $\|\mbox{PC}_6^2\|_F^*=7.6949\times 10^{3}$
$o=1,\ldots,6$ 0.0100 50.0000 $\|\mathcal{M}\|_F^*=3.0379\times 10^{4}$
\[table:resultex1\]
Illustrative Examples {#sec:ex}
=====================
Example 1 {#sec:ex1}
---------
Consider the Lena’s image (Figure \[fig:lena-orig\]). The objective in this example is to find the optimal values for the parameters of the weighting function while keeping the other parameters constant, where $$\begin{aligned}
& \lambda_{\min o} =3,~ \sigma_{no} =0.55,~ \eta_o=2.1,~ k_o =2,\\& \varepsilon=0.0001 ,~ N_o =4,~ O=6. \end{aligned}$$
The parameter vector in this example contains only $c_o$ and $g_o$, thus $$\begin{aligned}
\upsilon_o= [ \begin{array}{cc}
c_o & g_o \end{array}]
\end{aligned}$$ We set fixed upper and lower limits $\upsilon_o$ as follows: $$\begin{aligned}
& \underline{\upsilon}_o=[ \begin{array}{cc}
0.1 & 1 \end{array}] \\
& \overline{\upsilon}_o=[ \begin{array}{cc}
0.9 & 50 \end{array}].\end{aligned}$$ With $c_o < 0.1$, almost all spreads of frequencies (even narrow ones) will be kept, but that is not desirable. The change in the slope of the weighting function is also not significant for $g_o > 50$.
In order to compare , , and , this example focuses on $\mu_1=\mu_2=1$, and therefore, $$\begin{aligned}
\mathcal{M}=M+m.\end{aligned}$$
Table \[table:resultex1\] shows the optimal values of the parameter vector and cost functions. Recall that the solution to and results in one optimal vector for all orientations, and the solution to or results in one optimal vector for each orientation. The results of Frobenius-norm optimization are given in Table \[table:resultex1\]. Recall that the superscript $^*$ represents the optimal value.
Figure \[fig:lena-opt\] shows the image of $\mathcal{M}^*$ for , , and . Features such as lines, edges, Mach bands, corners, and blobs are satisfactorily detected in all images. Visually, it is seen that all four optimization methods result in the same image.
Figure \[fig:diff-2par\] shows the comparison between determinant-based optimization methods and , and norm-based optimization methods and . Differences are demonstrated in the subtraction images between the two approaches. Notably, although the determinant and norm optimal frameworks result in a similar performance visually, their subtraction images are not empty, meaning not identical.
[0.25]{} ![Detection of features from the Lena’s image. This is done by optimizing the parameters of the weighting function for 2D-MSPC in Example 1. The image of $\mathcal{M}^*$ is obtained through methods , , and .[]{data-label="fig:lena-opt"}](pc_lena_det_Mpm_6or "fig:"){width="\linewidth"}
[0.25]{} ![Detection of features from the Lena’s image. This is done by optimizing the parameters of the weighting function for 2D-MSPC in Example 1. The image of $\mathcal{M}^*$ is obtained through methods , , and .[]{data-label="fig:lena-opt"}](pc_lena_Fronorm_Mpm_6or "fig:"){width="\linewidth"}
\
[0.25]{} ![Detection of features from the Lena’s image. This is done by optimizing the parameters of the weighting function for 2D-MSPC in Example 1. The image of $\mathcal{M}^*$ is obtained through methods , , and .[]{data-label="fig:lena-opt"}](pc_lena_det_PC_6or "fig:"){width="\linewidth"}
[0.25]{} ![Detection of features from the Lena’s image. This is done by optimizing the parameters of the weighting function for 2D-MSPC in Example 1. The image of $\mathcal{M}^*$ is obtained through methods , , and .[]{data-label="fig:lena-opt"}](pc_lena_Fronorm_PC_6or "fig:"){width="\linewidth"}
[0.25]{} ![Comparison between determinant-based optimization methods and , also between norm-based optimization methods and .[]{data-label="fig:diff-2par"}](pc_lena_det_Mpm_m_PC_2par "fig:"){width="\linewidth"}
\
[0.25]{} ![Comparison between determinant-based optimization methods and , also between norm-based optimization methods and .[]{data-label="fig:diff-2par"}](pc_lena_Fnorm_Mpm_m_PC_2par "fig:"){width="\linewidth"}
For 1-, 2-, $\infty-$ and $F-$ norms, the optimization method results in the optimal parameter vector $\upsilon_o^*=[0.1 ~ 50]$ with $\|\mathcal{M}\|^*_1=219.5310$, $\|\mathcal{M}\|^*_2=75.4579$, $\|\mathcal{M}\|^*_{\infty}=204.6513$, and $\|\mathcal{M}\|^*_F=114.1408$, respectively. The image of $\mathcal{M}^*$ with optimal settings for the weighting function is shown in Figure \[fig:lena-norm-opt\], obtained through based on 1-, 2-, $\infty-$ and $F-$ norms. Again, it is seen that all four norm types result in the same visual performance. This example shows that the norm-based optimization of 2D-MSPC for the Lena’s image is robust to the norm type.
[0.25]{} ![Impact of using different norm types for the detection of features in Lena’s image. The image of $\mathcal{M}^*$ and the value of the cost function (calculated for the optimal parameter vector) are achieved from based on 1-, 2-, $\infty-$ and $F-$ norms. []{data-label="fig:lena-norm-opt"}](pc_lena_1norm_Mpm_6or "fig:"){width="\linewidth"}
[0.25]{} ![Impact of using different norm types for the detection of features in Lena’s image. The image of $\mathcal{M}^*$ and the value of the cost function (calculated for the optimal parameter vector) are achieved from based on 1-, 2-, $\infty-$ and $F-$ norms. []{data-label="fig:lena-norm-opt"}](pc_lena_2norm_Mpm_6or "fig:"){width="\linewidth"}
\
[0.25]{} ![Impact of using different norm types for the detection of features in Lena’s image. The image of $\mathcal{M}^*$ and the value of the cost function (calculated for the optimal parameter vector) are achieved from based on 1-, 2-, $\infty-$ and $F-$ norms. []{data-label="fig:lena-norm-opt"}](pc_lena_Infnorm_Mpm_6or "fig:"){width="\linewidth"}
[0.25]{} ![Impact of using different norm types for the detection of features in Lena’s image. The image of $\mathcal{M}^*$ and the value of the cost function (calculated for the optimal parameter vector) are achieved from based on 1-, 2-, $\infty-$ and $F-$ norms. []{data-label="fig:lena-norm-opt"}](pc_lena_Fronorm_Mpm_6or "fig:"){width="\linewidth"}
Because the PC is a matrix with elements between 0 and 1, the determinant is expected to be small. Note that, the larger the image, the smaller the determinant is most likely achieved. Thus, from the numerical point of view, solving such an optimization problem should be more sophisticated than solving the norm-based optimization problem.
$c_o^*$ $g_o^*$ $\lambda_{\min o}^*$ $\sigma_{no}^*$ $\eta_o^*$ $N_o^*$ $O^*$ Cost function opt. value
-- --------- --------- ---------------------- ----------------- ------------ --------- ------- ---------------------------------------------
0.1 49.9994 3.055 0.4 3.8315 4 1 $\|\mathcal{M}\|_F^*=442.0584$
0.1 50.00 2.6894 0.4 4.0 4 1 $\|\mbox{PC}_o^2\|_F^*=6.2033\times 10^{4}$
$\|\mathcal{M}\|_F^*=1.2407\times 10^{5}$
\[table:resultex2\]
[0.25]{} ![The results of the optimization of 7 parameters by solving Frobenius norm-based methods in Example 2. Figures \[fig:pc-Fnorm-mpm-full\] and \[fig:pc-Fnorm-pc-full\] shows the images of $\mathcal{M}^*$ with optimal parameters obtained through and , respectively. Figure \[fig:diff\] shows the difference by subtracting the images. []{data-label="fig:lena-7par"}](pc_lena_Fnorm_Mpm_full "fig:"){width="\linewidth"}
[0.25]{} ![The results of the optimization of 7 parameters by solving Frobenius norm-based methods in Example 2. Figures \[fig:pc-Fnorm-mpm-full\] and \[fig:pc-Fnorm-pc-full\] shows the images of $\mathcal{M}^*$ with optimal parameters obtained through and , respectively. Figure \[fig:diff\] shows the difference by subtracting the images. []{data-label="fig:lena-7par"}](pc_lena_Fnorm_PC_full "fig:"){width="\linewidth"}
\
[0.25]{} ![The results of the optimization of 7 parameters by solving Frobenius norm-based methods in Example 2. Figures \[fig:pc-Fnorm-mpm-full\] and \[fig:pc-Fnorm-pc-full\] shows the images of $\mathcal{M}^*$ with optimal parameters obtained through and , respectively. Figure \[fig:diff\] shows the difference by subtracting the images. []{data-label="fig:lena-7par"}](pc_lena_Fnorm_Mpm_m_PC_full "fig:"){width="\linewidth"}
[0.5]{} {width="1\linewidth"}
[0.5]{} {width="1\linewidth"}
Example 2:
----------
The objective of this example is to find the optimal values for the number of scales, number of orientations, and the parameters of the weighting function parameters and bank of filters, with the noise and ill-conditioning parameters fixed to $$\begin{aligned}
k_o =2, \varepsilon=0.0001\end{aligned}$$
The parameter vector in this example is $$\begin{aligned}
\upsilon_o= [ \begin{array}{ccccccc}
c_o &g_o & \lambda_{\min o} & \sigma_{no} & \eta_o & N_o & O \\
\end{array}]
\end{aligned}$$ We set fixed upper and lower limits to $\upsilon_o$ as follows $$\begin{aligned}
&\underline{\upsilon}_o= [ \begin{array}{ccccccc}
0.1 & 1.0 & 2.0 & 0.4 & 1.0 & 1 & 1 \\
\end{array}]\\
&\overline{\upsilon}_o= [ \begin{array}{ccccccc}
0.9 & 50.0 & 5.0 & 1.0 & 4.0 & 4 & 6 \\
\end{array}]
\end{aligned}$$
Because the Lena’s IFD has shown robustness to the norm type as demonstrated above, we only test the Frobenius norm-based optimization problems and . The resulting optimal values are given in Table \[table:resultex2\]. Notably, the optimal number of orientations where $O^*=1$ is achieved. Another noticeable result is that the optimal values for the weighting function parameters do not change, for which the boundary value at the upper and lower limits are found.
Figures \[fig:pc-Fnorm-mpm-full\] and \[fig:pc-Fnorm-pc-full\] show the images of $\mathcal{M}^*$ with associated optimal parameters $\upsilon_{o}^*$. By optimizing additional parameters of the 2D-MSPC, it is seen that many features are now better detected than Example 1. This is also consistent with the outcome of cost functions: the value of cost functions is larger than that in Example 1. Figure \[fig:diff\] shows the difference between the images obtained through and .
Example 3:
----------
In this example, magnetic resonance imaging (MRI) from the brain of a patient having multiple sclerosis (MS) is studied. MRI plays a key role in diagnosis and management of MS, [@Rodriguez2013]. MS is a disease that causes nerve damage in the brain and spinal cord. Characteristically, multi-focal plaques (lesions) can be seen using MRI, showing areas of brightness compared to the surrounding tissue (Figure \[fig:IM-0004-0054\_patient6\_T2-orig\]) [@Cabezas2014; @Schmidt2012; @Lorenzo2013]. Our experiments show that maximization of the minimum moment can increase the brightness of brain white matter in MRI. Thus, we chose $\mu_2=0$, and the cost function became $\mathcal{M}=M$. We solved the 2D-MSPC optimization for the $F-$norm with $\underline{\upsilon}_o$ and $\overline{\upsilon}_o$ as given in Example 2. The following optimal parameters were achieved with the optimal cost function $\|\mathcal{M}\|^*_F=234.0708$. $$\begin{aligned}
\upsilon_{o}^{*}= [ \begin{array}{ccccccc}
49.9396 & 0.1001 & 4.7264 & 0.4 & 1.7499 & 3 & 1 \\
\end{array}]
\end{aligned}$$ Figure \[fig:IM-0004-0054\_patient6\_T2-pc\] shows the image of $\mathcal{M}$. It is seen that the location and size of lesions are clearly detectable by using the proposed optimal 2D-MSPC.
Example 4
---------
In this example, we show another MR image from a postmortem brain with MS obtained using a high-field MR scanner (Figure \[fig:NR01196s57R1-orig\]). Arrows indicate MS lesions, which have been confirmed by histological analysis.
Because maximization of the minimum moment increases the brightness of the non-lesion brain areas, we chose $\mu_2=0$. The cost function also became $\mathcal{M}=M$. We solve the 2D-MSPC optimization for the $F-$norm, with $\underline{\upsilon}_o$ and $\overline{\upsilon}_o$ as given in Example 2. The following optimal parameters were achieved with the optimal cost function $\|\mathcal{M}\|^*_F=262.1824$. $$\begin{aligned}
\upsilon_{o}^{*}= [ \begin{array}{ccccccc}
49.9998 & 0.1 & 2 & 0.4 & 3.998 & 4 & 1 \\
\end{array}]
\end{aligned}$$ Figure \[fig:NR01196s57R1-pc\] shows the image of $\mathcal{M}^*$. It is seen that the MS lesions are clearly detectable.
[0.5]{} {width="1\linewidth"}
[0.5]{} {width="1\linewidth"}
Conclusions
===========
In this paper, we have focused on the IFD using a 2D-MSPC method. 2D-MSPC is originally proposed by Peter Kovesi and has shown great potential for the detection of various image features, particularly, lines, edges, corners, Mach bands, and blobs. However, the parameter setting of 2D-MSPC is typically performed manually based on trial and error, and studies for tuning of such parameters are limited. To enhance the application of this method, we have proposed several optimization frameworks for optimal and automatic tuning of the 2D-MSPC parameters. Through demonstration of several examples including MR images from patients with MS, we show that the ability of IFD can significantly be enhanced.
Matrix Norm Definition {#matrixnorm}
======================
([@Lyche2012], §8.1, p.177) \[normdef\] A function $\|.\|: {\ensuremath{\mathbb{C}}}^{q\times r}$ is called a matrix norm on $ {\ensuremath{\mathbb{C}}}^{q\times r}$ if for all $A,B \in {\ensuremath{\mathbb{C}}}^{q\times r}$ and all $c \in {\ensuremath{\mathbb{C}}}$ $$\begin{aligned}
&\mbox{(Positivity)~}\|A\| \geq 0\\
&\mbox{(homogeneity)~}\| c A\| = |c| ~\|A\| \\
&\mbox{(subadditivity)~} \| A+B\| \leq \|A\|+\|B\|\end{aligned}$$
Proof of Lemma \[lempc\] {#p-lempc}
========================
The replacement of , and in results in $$\begin{aligned}
& M=\frac{1}{2}\Big(\textstyle \sum_o \mbox{PC}_o^2 \cos^2(\theta_o)+\sum_o \mbox{PC}_o^2 \sin^2(\theta_o)+\\
& \sqrt{4\textstyle \sum_o \mbox{PC}_o^4 \cos^2(\theta_o)\sin^2(\theta_o)+\textstyle \sum_o \mbox{PC}_o^4(\cos^2(\theta_o)-\sin^2(\theta_o))^2}\Big)\\
&\hspace{5em}=\frac{1}{2}\Big(\textstyle \sum_o \mbox{PC}_o^2 ( \cos^2(\theta_o)+ \sin^2(\theta_o))+\\
& \hspace{5em}~~ \sqrt{\textstyle \sum_o \mbox{PC}_o^4 (\cos^2(\theta_o)+\sin^2(\theta_o))^2}\Big)\\&\hspace{5em}=\frac{1}{2}\Big(\textstyle \sum_o \mbox{PC}_o^2+\sqrt{\textstyle \sum_o \mbox{PC}_o^4}\Big).\end{aligned}$$
The proof for the minimum moment is exactly the same and omitted.
Proof of Proposition \[thpcbnd\] {#p-pcbnd}
================================
We need the following Lemmas from linear algebra.
\[lemapb\] The following property holds for matrix norms. $$\begin{aligned}
\|A-B\| \leq \|A\| +\|B\|\end{aligned}$$ [*Proof:*]{} is straightforward using the matrix norm definition \[normdef\].
([@Lyche2012], §8.1.1, p.179) \[lemak\] For a consistent matrix norm $\|.\|_c$ on ${\ensuremath{\mathbb{C}}}^{q \times q}$, the following inequality holds $$\begin{aligned}
\| A^k\|_c \leq \|A\|_c^k, \mbox{~for~} k\in {\ensuremath{\mathbb{N}}}.\end{aligned}$$
The proof of Proposition \[thpcbnd\] is as follows. By using Lemma \[lempc\], we have $$\|M\|_c=\frac{1}{2}\left\| \mbox{PC}_1^2+\ldots+\mbox{PC}_O^2+\sqrt{\mbox{PC}_1^4+\ldots+\mbox{PC}_O^4}\right\|_c$$ From Lemma \[lemapb\], thus $$\|M\|_c \leq \frac{1}{2}\Big(\left\| \mbox{PC}_1^2\right\|_c+\ldots+\left\|\mbox{PC}_O^2\right\|_c+\left\|\sqrt{\mbox{PC}_1^4+\ldots+\mbox{PC}_O^4}\right\|_c\Big)$$ By using Lemma \[lemak\], it yields $$\begin{aligned}
\|M\|_c& \leq \frac{1}{2}\Big( \left\| \mbox{PC}_1^2\right\|_c+\ldots+\left\|\mbox{PC}_O^2\right\|_c+\sqrt{\left\|\mbox{PC}_1^4+\ldots+\mbox{PC}_O^4\right\|_c}\Big)\\
& \leq \frac{1}{2}\Big(\left\| \mbox{PC}_1^2\right\|_c+\ldots+\left\|\mbox{PC}_O^2\right\|_c+\sqrt{\left\|\mbox{PC}_1^4\right\|_c+\ldots+\left\|\mbox{PC}_O^4\right\|_c}\Big)\end{aligned}$$ Lemma \[lemak\] is again used, which yields $$\begin{aligned}
&\|M\|_c \leq \\
&\frac{1}{2}\Big(\left\| \mbox{PC}_1^2\right\|_c+\ldots+\left\|\mbox{PC}_O^2\right\|_c+\sqrt{\left\|\mbox{PC}_1^2\right\|_c^2+\ldots+\left\|\mbox{PC}_O^2\right\|_c^2}\big)\end{aligned}$$ From algebra, $\sqrt{a_1^2+\ldots+a_O^2}\leq a_1+\ldots+a_O$ for $a_i\geq0$, thus, $$\begin{aligned}
\|M\|_c &\leq \frac{1}{2}\Big(\left\| \mbox{PC}_1^2\right\|_c+\ldots+\left\|\mbox{PC}_O^2\right\|_c+\left\|\mbox{PC}_1^2\right\|_c+\ldots+\left\|\mbox{PC}_O^2\right\|_c\Big)\\
&=\frac{1}{2}\Big(2\left\| \mbox{PC}_1^2\right\|_c+\ldots+2\left\|\mbox{PC}_O^2\right\|_c\Big)\end{aligned}$$ and the proof of is complete.
By using Lemma \[lempc\], we have $$\|m\|_c=\frac{1}{2}\Big(\left\| \mbox{PC}_1^2+\ldots+\mbox{PC}_O^2-\sqrt{\mbox{PC}_1^4+\ldots+\mbox{PC}_O^4}\right\|_c\Big)$$ From the matrix norm property, thus $$\|m\|_c \leq \frac{1}{2}\Big(\left\| \mbox{PC}_1^2\right\|_c+\ldots+\left\|\mbox{PC}_O^2\right\|_c+\left\|\sqrt{\mbox{PC}_1^4+\ldots+\mbox{PC}_O^4}\right\|_c\Big)$$ and the rest of the proof is exactly the same as above.
Proof of Proposition \[thsubopt\] {#p-suboptlem}
=================================
From the norm definition, , and , it is deduced that $$\begin{aligned}
\left\| \mathcal{M}\right\|_c \leq (|\mu_1|+|\mu_2|) \textstyle \sum_{o} \left\| \mbox{PC}_o^2 \right\|_c\end{aligned}$$ A suboptimal solution to , with the consistent norm, is obtained by increasing the upper found of $\left\| \mathcal{M}\right\|_c$. This completes the proof.
Acknowledgment {#acknowledgment .unnumbered}
==============
First author would like to thank Shrushrita Sharma and Glen Pridham for some discussions. Funding from the MS Society of Canada, Natural Sciences and Engineering Council of Canada, and Alberta Innovates – Health Solutions is also acknewledged.
[1]{}
Peter Kovesi, [*Invariant Measures of Image Features From Phase Information*]{}, PhD Thesis, The University of Western Australia, 1996.
D. Ziou and S. Tabbone, [*Edge detection techniques: An overview*]{}, International Journal of Pattern Recognition and Image Analysis, 8(4), pp. 537–559, 1998.
T.P. Patel, S.R. Panchal, Corner Detection Techniques: An Introductory Survey, [*International Journal of Engineering Development and Research*]{}, Vol. 2, No. 4, pp. 3680–3686, 2014.
J.A. Noble, [*Descriptions of image surfaces*]{}, D.Phil thesis, Department of Engineering Science, University of Oxford, 1989.
J.F. Canny, [*Finding edges and lines in images*]{}, Master’s thesis, Massachusetts Institute of Technology, AI Lab. TR-720, 1983.
A. Blake, A. Zisserman, [*Visual Reconstruction*]{}, MIT Press, Cambridge, MA, 1987.
P. Perona, J. Malik, Scale-space and edge detection using anisotropic diffusion, [*IEEE Transactions on Pattern Analysis and Machine Intelligence*]{}, Volume 12, Issue 7, pp. 629 - 639, 1990.
S.M. Smith, J.M. Brady, SUSAN: A New Approach to Low Level Image Processing, [*International journal of computer vision*]{}, Volume 23, Issue 1, pp 45–78, 1997.
M.C. Morrone, D.C. Burr, J. Ross, R. Owens, Mach bands are phase dependent, [*Nature*]{}, 324, 250–253, 20 November 1986.
P. Kovesi, Image Features From Phase Congruency, [*Videre: Journal of Computer Vision Research, MIT Press*]{}. Volume 1, Number 3, pp. 1-25, Summer 1999.
P. Kovesi, Phase Congruency Detects Corners and Edges, [*The Australian Pattern Recognition Society Conference: DICTA 2003*]{}, pp 309–318, Sydney. December 2003.
A.V. Oppenheim, J.S. Lim, The importance of phase in signals, [*IEEE Proceedings*]{}, vol. 69, pp. 529–541, 1981.
S. Venkatesh, R. Owens, On the classification of image features, [*Pattern Recognition Letters*]{}, Volume 11, Issue 5, Pages 339–349, May 1990.
M.C. Morrone, R. Owens, Feature detection from local energy, [*Pattern Recognition Letters*]{}, 6, pp. 303–313, 1987.
G. Cao, P. Shi, B. Hu, Ultrasonic Liver Discrimination Using 2-D Phase Congruency, [*IEEE Transactions on Biomedical Engineering*]{}, Vol. 53, No. 10, pp. 2116 – 2119, 2006.
V. Štruc N. Paveši' c, Phase congruency features for palm-print verification, [*IET Signal Processing*]{}, Vol. 3, Issue 4, pp. 258 – 268, 2009.
L. Zhang, L. Zhang, D. Zhang, Z. Guo, Phase congruency induced local features for finger-knuckle-print recognition, [*Pattern Recognition*]{}, Volume 45, Issue 7, pp. 2522 – 2531, 2012.
B. Obara, M. Fricker, D. Gavaghan, V. Grau, Contrast-Independent Curvilinear Structure Detection in Biomedical Images, [*IEEE Transactions on Image Processing*]{}, Vol. 21, No. 5, pp. 2572–2581, 2012
T. Mouats, N. Aouf, M.A. Richardson, A Novel Image Representation via Local Frequency Analysis for Illumination Invariant Stereo Matching, [*IEEE Transactions on Image Processing*]{}, Vol. 24, No. 9, pp. 2685-2700, 2015.
O.M. Rijal, H. Ebrahimian, N.M. Noor, A. Hussin, A. Yunus, and A.A. Mahayiddin, Application of Phase Congruency for Discriminating Some Lung Diseases Using Chest Radiograph, [*Computational and Mathematical Methods in Medicine*]{}, Article ID 424970, 15 pages, 2015.
H. Ziaei Nafchi ; R. Farrahi Moghaddam ; M. Cheriet, Phase-Based Binarization of Ancient Document Images: Model and Applications, [*IEEE Transactions on Image Processing*]{}, Vol. 23, No. 7, 2916 – 2930, July 2014.
[ http://www.peterkovesi.com]( http://www.peterkovesi.com). W.S. Levine, [*The Control Handbook*]{}, CRC Press in Cooperation with IEEE Press, 1995.
V. V. Fedorov, [*Theory of optimal experiments*]{}, Elsevier, 1972.
G. C. Goodwin and R. L. Payne, [*Dynamic system identification: experiment design and data analysis*]{}, Academic press, 1977.
A. C. Atkinson, Developments in the design of experiments, [*International Statistical Review*]{}, vol. 50, pp. 161 – 177, 1982.
L. Pronzato and E. Walter, Robust experiment design via stochastic approximation, [*Mathematical Bio-sciences*]{}, vol. 75, no. 1, pp. 103–120, 1985.
M.P.F. Berger, D-optimal sequential sampling designs for item response theory models, [*Journal of Educational Statistics*]{}, vol. 19, no. 1, pp. 43–56, 1994.
F. Pukelsheim, [*Optimal design of experiments, ser. Prob- ability and mathematical statistic*]{}, Wiley, 1993.
R. A. Fisher, [*The Design of Experiments*]{}, 9th ed., Macmillan, 1971 (1935 1st ed.).
H. Cramér, [*Mathematical Methods of Statistics*]{}, Princeton Univ. Press. Princeton, NJ. 1946.
C.R. Rao, Information and the accuracy attainable in the estimation of statistical parameters. [*Bulletin of the Calcutta Mathematical Society*]{}, 37, p. 81–89, 1945.
A. Atkinson, A. Donev, R. Tobias, [*Optimum Experimental Designs, With SAS*]{}, Oxford University Press, 2007.
R.C. St.John and N. R. Draper, D-optimality for regression designs: a review, [*Technometrics*]{}, vol. 17, no. 1, pp. 15 – 23, 1975.
T. Lyche, [*Lecture Notes for Inf-Mat 4350, 2012*]{}, University of Oslo, August 2012.
R. Schatten, [*Norm Ideals of Completely Continuous Operators*]{}, Springer-Verlag, Berlin, Gottingen, Heidelberg, 1960.
M. Rodriguez, O.H. Kantarci, I. Pirko, [*Multiple Sclerosis*]{}, Oxford University Press, 2013.
M. Cabezas, A. Oliver, E. Roura, J. Freixenet, J. C. Vilanova, L. Ramió-Torrentà, À. Rovira, X. Lladó, Automatic multiple sclerosis lesion detection in brain MRI by FLAIR thresholding, [*Computer Methods and Programs in biomedicine*]{}, 115, pp. 147 – 161, 2014.
P. Schmidt, C. Gaser, M. Arsic, D. Buck, A. Förschler, A. Berthele, M. Hoshi, R. Ilga, V. J. Schmid, C. Zimmer, B. Hemmer, M. Mühlau An automated tool for detection of FLAIR-hyperintense white-matter lesions in Multiple Sclerosis, [*NeuroImage*]{}, Volume 59, Issue 4, pp. 3774–3783, 2012.
D. García-Lorenzo, S. Francis, S. Narayanan, D. L. Arnold, D. L. Collins, Review of automatic segmentation methods of multiple sclerosis white matter lesions on conventional magnetic resonance imaging, [*Medical Image Analysis*]{}, Volume 17, Issue 1, Pages 1–18, January 2013.
[^1]: S.M.M. Alavi is with the Hotchkiss Brain Institute and Department of Clinical Neurosciences at the Cumming School of Medicine, University of Calgary, Canada. Email: [email protected].
[^2]: Y. Zhang is with the Hotchkiss Brain Institute, Department of Clinical Neurosciences, and Department of Radiology at the Cumming School of Medicine, University of Calgary. Email: [email protected].
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
We consider the existence and the non-existence of a minimizer of the following minimization problems associated with an improved Hardy-Sobolev type inequality introduced by Ioku [@I]. $$\begin{aligned}
I_a := \inf_{u \in W_0^{1,p}(B_R ) \setminus \{ 0\} } \dfrac{\int_{B_R} |\nabla u |^{p} \,dx}{\( \int_{B_R} |u|^{p^*(s)} V_a(x) \,dx \)^{\frac{p}{p^*(s)}}}, \,\,\text{where}\,\, V_a (x) =\frac{1}{|x|^s \( 1- a \,\( \frac{|x|}{R} \)^{\frac{N-p}{p-1}} \)^\beta} \ge \frac{1}{|x|^s}.\end{aligned}$$ Only for radial functions, the minimization problem $I_a$ is equivalent to it associated with the classical Hardy-Sobolev inequality on $\re^N$ via a transformation. First, we summarize various transformations including that transformation and give a viewpoint of such transformations. As an application of this viewpoint, we derive [*an infinite dimensional form*]{} of the classical Sobolev inequality in some sense. Next, without the transformation, we investigate the minimization problems $I_a$ on balls $B_R$. In contrast to the classical results for $a=0$, we show the existence of non-radial minimizers for the Hardy-Sobolev critical exponent $p^* (s)=\frac{p (N-s)}{N-p}$ on bounded domains. Finally, we give remarks of a different structure between two nonlinear scalings which are equivalent to the usual scaling only for radial functions under some transformations.
address: 'Laboratory of Mathematics, Graduate School of Engineering, Hiroshima University, Higashi-Hiroshima, 739-8527, Japan'
author:
- Megumi Sano
title: 'Minimization problem associated with an improved Hardy-Sobolev type inequality'
---
Hardy-Sobolev inequality ,Optimal constant ,Extremal function
35A23 ,35J20 ,35A08
Introduction and main results {#intro}
=============================
Let $B_R \subset \re^N, 1 < p < N, 0 \le s \le p, p^* (s) = \frac{p (N-s)}{N-p}$ and $p^* = p^*(0)$. Then the classical Hardy-Sobolev inequality: $$\label{HS}
C_{N, p, s} \( \int_{B_R} \frac{|u|^{p^* (s)}}{|x|^s} dx \)^{\frac{p}{p^* (s)}} \le \int_{B_R} | \nabla u |^p dx$$ holds for all $u \in W^{1,p}_0(B_R)$, where $W_0^{1,p}(B_R)$ is the completion of $C_c^{\infty}(B_R)$ with respect to the norm $\| \nabla (\cdot )\|_{L^p(B_R)}$ and $C_{N,p,s}$ is the best constant of (\[HS\]). In the case where $s=0$ (resp. $s=p$), the inequality (\[HS\]) is called the Sobolev (resp. Hardy) inequality. The Hardy-Sobolev inequality (\[HS\]) is quite fundamental and important since it expresses the embeddings of the Sobolev spaces $W_0^{1,p}$. Furthermore the variational problems and partial differential equations associated with the Hardy-Sobolev inequality (\[HS\]) are well-studied by many mathematicians so far, see [@Au], [@T], [@L], [@BG], [@BV], [@VZ], to name a few.
Recently, Ioku [@I] showed the following improved Hardy-Sobolev inequality for radial functions via the transformation (\[trans\]), see §\[Trans\]. $$\label{IHS}
C_{N, p, s} \( \int_{B_R} \frac{|u|^{p^* (s)}}{|x|^s \( 1- \( \frac{|x|}{R} \)^{\frac{N-p}{p-1}} \)^{\beta}} dx \)^{\frac{p}{p^* (s)}} \le \int_{B_R} | \nabla u |^p dx\,\,\text{for} \,\, u \in W_{0, {\rm rad}}^{1,p}(B_R),$$ where $\beta = \beta (s)= \frac{(N-1)p - (p-1)s}{N-p}$. One virtue of (\[IHS\]) is that we can take a limit directly for the improved inequality (\[IHS\]) as $p \nearrow N$, differently from the classical one. Indeed, Ioku [@I] showed that the limit of the improved inequality (\[IHS\]) with $s=0$ is the Alvino inequality [@Al] which implies the optimal embedding of $W_0^{1,N}(B_R)$ into the Orlicz spaces, and also the limit of the improved inequality (\[IHS\]) with $s=p$ is the critical Hardy inequality which implies the embedding of $W_0^{1,N} (B_R)$ into the Lorentz-Zygmund spaces $L^{\infty, N}(\log L)^{-1}$ which is smaller than the Orlicz space. For some indirect limiting procedures for the classical Hardy-Sobolev inequalities, see [@Tru], [@BP], [@SS] and a survey [@S(RIMS)]. Based on the transformation (\[trans\]), the improved inequality (\[IHS\]) on $B_R$ equivalently connects to the classical one (\[HS\]) on the whole space $\re^N$. This yields that the improved inequality (\[IHS\]) has the scale invariance under a scaling to which the usual scaling is changed via the transformation (\[trans\]), and there exists a radial minimizer of (\[IHS\]) when $0 \le s < p$. For more details, see [@I] or §\[Trans\].
In this paper, without the transformation, we investigate the following extended minimization problems $I_a$ for $a \in [0,1]$ associated with improved Hardy-Sobolev inequalities. $$\begin{aligned}
I_{a} := \inf_{u \in W_{0}^{1,p}(B_R) \setminus \{ 0\} } \frac{\int_{B_R} | \nabla u |^p \,dx}{\( \int_{B_R} |u|^{p^* (s)} V_a (x) \,dx \)^{\frac{p}{p^* (s)}}}, \,\text{where}\, V_a(x) =\frac{1}{|x|^s \( 1- a \,\( \frac{|x|}{R} \)^{\frac{N-p}{p-1}} \)^\beta} \ge \frac{1}{|x|^s}.\end{aligned}$$ Note that the potential function $V_a(x)$ also has the boundary singularity when $a=1$. Due to the boundary singularity, $I_1 = 0$ if $a=1$ and $s < p$, see Proposition \[Posi I\_a\] in §\[Proof\]. Therefore we exclude the case where $a=1$ and $s < p$.
Our main results are as follows.
\[Main\] (i) Let $s=p$. Then for any $a \in [0, 1]$, $I_a =I_{a, {\rm rad}} = C_{N, p, p} = (\frac{N-p}{p})^p$ and $I_a$ is not attained.\
(ii) Let $s=0$. Then $I_a =I_{a, {\rm rad}} (1-a)^{\frac{N-1}{N} p}= C_{N, p, 0} (1-a)^{\frac{N-1}{N} p}$ and $I_a$ is not attained for any $a \in [0,1)$.\
(iii) Let $0 < s < p$. Then there exists $a_* \in (A, 1)$ such that $I_a < I_{a, {\rm rad}}$ for $a \in (a_*, 1)$, $I_a$ is attained for $a \in (a_*, 1)$, $I_a=I_{a, {\rm rad}}$ for $a \in [0, a_*]$ and $I_a$ is not attained for $a \in [0, a_*)$, where $A=1- \( \frac{s (p-1)}{p (N-1)} \)^{\frac{s}{\beta}} \left[ 1- \( \frac{s (p-1)}{p (N-1)} \)\right]$.
Note that $\frac{s (p-1)}{p(N-1)} < A$ and the potential function $V_a(x)$ is not monotone-decreasing with respect to $|x|$ for $a \in (\frac{s(p-1)}{p(N-1)}, 1]$. Therefore it seems difficult to reduce the radial setting due to the lack of the rearrangement technique, see the first part in §\[Proof\]. However we can show $I_a=I_{a, {\rm rad}}$ even for $a \in (\frac{s (p-1)}{p(N-1)}, a_*]$ thanks to the special shape of the potential function $V_a(x)$, see the last part of the proof of Theorem \[Main\] (iii).
Our minimization problem $I_a$ is related to the following nonlinear elliptic equation with the singular potential $V_a(x) \ge |x|^{-s}$. $$\begin{aligned}
\label{EL}
\begin{cases}
-\text{div} \,( \, |\nabla u|^{p-2} \nabla u \,) = b V_a(x) |u|^{p^*(s) -2}u \quad &\text{in} \,\, B_R, \\
\qquad u = 0 &\text{on} \,\, \pd B_R.
\end{cases}\end{aligned}$$ The minimizer for $I_a$ is a ground state solution of the Euler-Lagrange equation (\[EL\]) with a Lagrange multiplier $b$. Since the minimizer of Theorem \[Main\] (iii) is a non-radial function, we can observe that the symmetry breaking phenomenon of the ground states of the elliptic equation (\[Eq\]) occurs when $f(x)= V_a(x)$, $a$ is close to $1$, $s \in (0,p)$ and $p<q=p^*(s) < p^*$. $$\begin{aligned}
\label{Eq}
\begin{cases}
-\text{div} \,( \, |\nabla u|^{p-2} \nabla u \,) = f(x) \,|u|^{q -2}u \quad &\text{in} \,\, B_R, \\
\hspace{6.5em} u = 0 &\text{on} \,\, \pd B_R.
\end{cases}\end{aligned}$$ It is well-known that the symmetry breaking phenomenon of the ground states of (\[Eq\]) occurs when $f(x)=\( \frac{|x|}{R} \)^\alpha$, $\alpha >0$ is sufficiently large and $p<q < p^*$ (ref. [@SSW]). Theorem \[Main\] (iii) gives another example of the potential function $f(x)$, which is not monotone-increasing and is not bounded differently from the Hénon potential function $\( \frac{|x|}{R} \)^\alpha$.
In the case where $a=0$, Theorem \[Main\] is the same as the classical result which is the non-existence of the minimizer of the classical Hardy-Sobolev inequality on bounded domains. However, in Theorem \[Main\] (iii), we see that there exists a minimizer of $I_a$ even on bounded domains. This is different from the classical result. Concretely, we show that $I_{a, \text{rad}}$ is the concentration level of minimizing sequence of $I_a$ for $s \in (0,p)$, where $I_{a, \text{rad}}$ is a level of $I_a$ only for radial functions. By concentration-compactness alternative, we can obtain a minimizer of $I_a$ for $a \in (a_*, 1)$. Note that the continuous embedding: $W_0^{1,p}(B_R) \hookrightarrow L^{p^*(s)}(B_R; V_a(x)\,dx)$ related to our problem is not compact due to the existence of a non-compact sequence given by the nonlinear scaling (\[new scale\]) in §\[Trans\], see also §\[App\]. However, the parameter of $a$ plays a role of lowering the level of $I_a$ than $I_{a, \text{rad}}$, and thanks to it, we can remove the possibility of occurring such non-compact behavior.
This paper is organized as follows: In §\[Trans\], we summarize various transformations including the transformation and give a viewpoint of such transformations by which the classical Hardy-Sobolev inequality equivalently connects to another inequality. As an application of this viewpoint, we derive [*an infinite dimensional form*]{} of the classical Sobolev inequality in some sense. In §\[Proof\], we prepare several Lemmas and Propositions and show Theorem \[Main\]. In §\[App\], we give remarks of a different structure between two nonlinear scalings which are equivalent to the usual scaling only for radial functions under some transformations.
We fix several notations: $B_R$ or $B_R^N$ denotes a $N$-dimensional ball centered $0$ with radius $R$. As a matter of convenience, we set $B_\infty^N = \re^N$ and $\frac{1}{\infty} = 0$. $\omega_{N-1}$ denotes an area of the unit sphere $\mathbb{S}^{N-1}$ in $\re^N$. $|A|$ denotes the Lebesgue measure of a set $A \subset \re^N$ and $X_{{\rm rad}} = \{ \, u \in X \, | \, u \,\,\text{is radial} \, \}$. The Schwarz symmetrization $u^{\#} \colon \re^N \to [0, \infty]$ of $u$ is given by $$\begin{aligned}
u^{\#}(x) = u^{\#}(|x|)=\inf \left\{ \tau >0 \,: \, | \{ y \in \re^N \, :\, |u(y)| > \tau \} \,| \le |B_{|x|}(0) | \right\}.\end{aligned}$$ Throughout the paper, if a radial function $u$ is written as $u(x) = \tilde{u}(|x|)$ by some function $\tilde{u} = \tilde{u}(r)$, we write $u(x)= u(|x|)$ with admitting some ambiguity.
Various transformations and an infinite dimensional form of the classical Sobolev inequality {#Trans}
============================================================================================
First, we explain the transformation introduced by Ioku [@I] for readers convenience and give several remarks.
We use the polar coordinates: $\re^N \ni x = r \w \,(r \in \re_+, \, \w \in \mathbb{S}^{N-1})$. Let $T \in [R, \infty], y \in B_T^N, x \in B_R^N, r= |x|, t =|y|$, and $w \in C_c^1 (B_T)$. By using the fundamental solution of $p-$Laplacian with Dirichlet boundary condition, we consider the following transformation $$\begin{aligned}
\label{trans}
u(r \w) = w(t \w), \,\,\text{where}\,\, r^{-\frac{N-p}{p-1} } - R^{-\frac{N-p}{p-1} } = t^{-\frac{N-p}{p-1} } - T^{-\frac{N-p}{p-1} }.\end{aligned}$$ Note that in the case where $T = \infty$ the transformation (\[trans\]) is founded by Ioku[@I]. Then we see that $$\begin{aligned}
\int_{B_T} | \nabla w |^p \,dy
&= \int_{\mathbb{S}^{N-1}} \int_0^T \left| \frac{\pd w}{\pd t} \w + \frac{1}{t} \nabla_{\mathbb{S}^{N-1}} w \,\right|^p t^{N-1} \,dt dS_{\w}\\
&= \int_{\mathbb{S}^{N-1}} \int_0^R \left| \frac{\pd u}{\pd r} \w + \frac{1}{t} \frac{dt}{dr} \nabla_{\mathbb{S}^{N-1}} u \,\right|^p \( \frac{dr}{dt} \)^{p-1} t^{N-1} \,dr dS_{\w}.\end{aligned}$$ Since $$\begin{aligned}
\frac{dr}{dt} = \(\, \frac{r}{t} \,\)^{\frac{N-1}{p-1}}, \,\, t \, \frac{dr}{dt} = r^{\frac{N-1}{p-1}} \( r^{-\frac{N-p}{p-1} } - R^{-\frac{N-p}{p-1} } + T^{-\frac{N-p}{p-1} } \),\end{aligned}$$ we have $$\begin{aligned}
\label{nab non-rad}
\int_{B_T} | \nabla w |^p \,dy = \int_{B_R} | L_p u |^p \,dx,\end{aligned}$$ where $$\begin{aligned}
L_p u = \frac{\pd u}{\pd r} \w + \frac{1}{r} \nabla_{\mathbb{S}^{N-1}} u \left[ 1 - a \( \frac{r}{R} \)^{\frac{N-p}{p-1} } \right]^{-1}.\end{aligned}$$ where $a= 1- \( \frac{R}{T} \)^{\frac{N-p}{p-1}} \nearrow 1$ as $T \nearrow \infty$. Note that the differential operator $L_p$ is not $\nabla$ for $T > R$ due to the last term. However, if $u$ and $w$ are radial functions, then we can obtain the equality of two $L^p$ norms of $\nabla$ between $B_T$ and $B_R$ as follows: $$\begin{aligned}
\label{nab}
\int_{B_T} | \nabla w |^p \,dy = \int_{B_R} | \nabla u |^p \,dx \quad \text{if} \,\, u\,\,\text{and}\,\,w \,\, \text{are radial.}\end{aligned}$$ On the other hand, for the Hardy-Sobolev term, we have $$\begin{aligned}
\label{deno}
\int_{B_T} \frac{| w |^{p^*(s)}}{|y|^s} \,dy &= \int_{B_R} \frac{| u |^{p^*(s)}}{|x|^s\( 1-a \(\frac{|x|}{R}\)^{\frac{N-p}{p-1}}\)^{\beta}} \,dx,\end{aligned}$$ where $\beta = \beta (s)= \frac{(N-1)p - (p-1)s}{N-p}$. Set $$\begin{aligned}
I_{a,{\rm rad}} := \inf_{u \in W_{0, {\rm rad}}^{1,p}(B_R) \setminus \{ 0\} } \frac{\int_{B_R} | \nabla u |^p \,dx}{\( \int_{B_R} \frac{| u |^{p^*(s)}}{|x|^s \( 1-a \,\( \frac{|x|}{R} \) ^{\frac{N-p}{p-1}} \)^{\beta}} \,dx \)^{\frac{p}{p^*(s)}}}.\end{aligned}$$ From (\[nab\]) and (\[deno\]), we observe that the minimization problem $I_{a, {\rm rad}}$ can be reduced the classical Hardy-Sobolev minimization problem $C_{N, p, s}$: $$\begin{aligned}
C_{N, p, s}
= \inf_{w \in W_{0, {\rm rad}}^{1,p}(B_T ) \setminus \{ 0\} } \dfrac{\int_{B_T} | \nabla w |^p \,dy}{\( \int_{B_T} \frac{| w |^{p^*(s)}}{|y|^s} \,dy \)^{\frac{p}{p^*(s)}}}.
$$ It is well-known that $C_{N, p, s}$ is independent of the radius $T$ and $C_{N, p, s}$ is attained if and only if $T = \infty$ and $0\le s < p$. Moreover its minimizer is the family of $$\begin{aligned}
W_\la (t)= \la^{\frac{N-p}{p}} (1+(\la t)^{\frac{p-s}{p-1}})^{-\frac{N-p}{p-s}}\,\, \text{for}\,\,\la \in (0, \infty),\end{aligned}$$ see e.g. [@CRT]. Therefore we can obtain the following results for $I_{a, {\rm rad}}$ based on the transformation (\[trans\]).
\[I\_a rad\]([@I]) $I_{a, {\rm rad}}$ is independent of $a \in [0,1]$ and $I_{a, {\rm rad}} = C_{N, p, s}$. And $I_{a, {\rm rad}}$ is attained if and only if $a=1$ and $0\le s < p$. Moreover, the minimizer of $I_{1, {\rm rad}}$ is the family of $$\begin{aligned}
U^{\la} (r) = \la^{\frac{N-p}{p}} \left[ 1+ (\la r)^{\frac{p-s}{p-1}} \left\{ 1- \( \frac{r}{R}\)^{\frac{N-p}{p-1}} \right\}^{-\frac{p-s}{N-p}} \right]^{-\frac{N-p}{p-s}}\,\, \text{for}\,\,\la \in (0, \infty).\end{aligned}$$
It is well-known that thanks to the zero extension, the classical inequality (\[HS\]) has the scale invariance under the usual scaling for $\la \in [1, \infty)$. $$\begin{aligned}
\label{usual scale}
w_\la (t \w) =
\begin{cases}
\la^{\frac{N-p}{p}} w(\la t \w) \,\, &\text{for}\,\, t \in [0, \frac{T}{\la}], \\
0 &\text{for} \,\, t \in (\frac{T}{\la}, T].
\end{cases}\end{aligned}$$ Note that in the case where $T = \infty$, we can consider any $\la \in (0, \infty)$. By the transformation (\[trans\]), the usual scaling (\[usual scale\]) is changed its form to the following scaling for $\la \in [1, \infty)$. $$\begin{aligned}
\label{new scale}
&u^\la (r \w) =
\begin{cases}
\la^{\frac{N-p}{p}} u(\tilde{r} \w) \,\, &\text{for}\,\, t \in [0, \tilde{R}], \\
0 &\text{for} \,\, t \in (\tilde{R}, R],
\end{cases} \\
&\text{where}\,\,\tilde{r} = (\la r) \left[ 1 + a \( \la^{\frac{N-p}{p-1}} -1 \) \( \frac{r}{R} \)^{\frac{N-p}{p-1}} \right]^{\,-\frac{p-1}{N-p}},\, \tilde{R} = R \( \la^{\frac{N-p}{p-1}} (1-a) + a \)^{-\frac{p-1}{N-p}}. \notag\end{aligned}$$ Respectively, in the case where $a=1$, we can consider any $\la \in (0, \infty)$. Obviously from (\[nab non-rad\]) and (\[deno\]), we see that the improved Hardy-Sobolev inequality: $$\begin{aligned}
C_{N, p, s} \( \int_{B_R} \frac{| u |^{p^*(s)}}{|x|^s\( 1-a \(\frac{|x|}{R}\)^{\frac{N-p}{p-1}}\)^{\beta}} \,dx \)^{\frac{p}{p^*(s)}} \le \int_{B_R} | L_p u |^p \,dx\end{aligned}$$ with the differential operator $L_p$ is invariant under the scaling (\[new scale\]). The scaling in [@I] looks different from the scaling (\[new scale\]). However, taking $\la \mapsto \la^{-\frac{p-1}{N-p}}$, we observe that these scalings are same essentially. Note that $\| \nabla u \|_{L^p(B_R)}$ is not invariant under the scaling (\[new scale\]) for non-radial functions. Therefore, in §\[Proof\], we investigate our minimization problem $I_a$ without the transformation (\[trans\]).
Actually, the original transformation is given in Theorem 18. in [@F] for $u \in W_{0, {\rm rad}}^{1,2}(B_1)$, where $B_1, \Omega \subset \re^2$ as follows. $$\begin{aligned}
\label{Ftrans}
w(y)= u(x),\,\, {\rm where}\,\, G_{\Omega, z} (y) = G_{B_1, 0}(x) = -\frac{1}{2\pi} \log |x|\end{aligned}$$ and $G_{\Omega, z}(y)$ is the Green function in a domain $\Omega$, which has a singularity at $z \in \Omega$. Therefore, we can observe that the transformation (\[trans\]) is (\[Ftrans\]) in the case where $W_{0, {\rm rad}}^{1,p}(B_1), p < N$ and $z=0, B_1 \subset \re^N = \Omega$.
In addition to (\[Ftrans\]) and (\[trans\]), there are various transformations by which the classical Hardy-Sobolev type inequality equivalently connects to another inequality for radial functions (ref. [@Z], [@HK], [@II], [@ST], [@I], [@S]). Next, we explain them comprehensively. An unified viewpoint is [*to connect two typical functions on each world (e.g. fundamental solution of $p-$Laplacian, virtual minimizer of the Hardy type inequality)*]{}. In [@Z], [@HK] and [@II], the following transformation (\[HK trans\]) is considered by using two fundamental solutions of $p-$Laplacian and weighted $p-$Laplacian: div$( |x|^{p-N} |\nabla u |^{p-2} \nabla u)$. $$\begin{aligned}
\label{HK trans}
u(r) = w(t), \,\,\text{where}\,\, t^{-\frac{N-p}{p-1} } = \log \frac{R}{r}.\end{aligned}$$ Then they obtain the equality of two norms between the subcritical Sobolev space $W_0^{1, p} (\re^N) \,(p < N)$ and the weighted critical Sobolev space $W_0^{1, p} (B_R^N ; |x|^{p-N} \,dx)$ as follows. $$\begin{aligned}
\int_{\re^N} | \nabla w |^p \,dy = \int_{B^N_R} |x|^{p-N} | \nabla u |^p \,dx, \,\,
\int_{\re^N} \frac{| w |^{q}}{|y|^s} \,dy
= \frac{p-1}{N-p} \int_{B_R^N} \frac{| u |^{q}}{|x|^N \( \log \frac{R}{|x|} \)^{\beta (s)}} \,dx.\end{aligned}$$ On the other hand, in [@ST] and [@S], they consider the following transformation (\[ST trans\]) by using two fundamental solutions of $p \,(=N) -$Laplacian and $N-$Laplacian as follows. $$\begin{aligned}
\label{ST trans}
u(r) = w(t), \,\,\text{where}\,\, t^{-\frac{m-N}{N-1} } = \log \frac{R}{r} \,\, \text{is equivalent to} \, t^{-\frac{m-N}{N} } = \( \log \frac{R}{r} \)^{\frac{N-1}{N}} .\end{aligned}$$ A different point of the transformation (\[ST trans\]) from these transformations (\[Ftrans\]), (\[trans\]), (\[HK trans\]) is to consider [*the difference of dimensions on each world*]{}. Thanks to the difference of dimensions, we obtain the equality of two norms between the critical Sobolev space $W_0^{1, N} (B_R^N)$ and the higher dimensional subcritical Sobolev space $W_0^{1, N} (\re^m) \,(N < m)$ as follows. $$\begin{aligned}
\int_{\re^m} | \nabla w |^N \,dy = \int_{B^N_R} | \nabla u |^N \,dx,\,\,
\int_{\re^m} \frac{| w |^{q}}{|y|^{\alpha}} \,dy
=\frac{\w_{m-1} (N-1)}{\w_{N-1} (m-N)} \int_{B_R^N} \frac{| u |^{q}}{|x|^N \( \log \frac{R}{|x|} \)^{\gamma(\alpha )}} \,dx,\end{aligned}$$ where $\gamma(\alpha) = \frac{(m-1)N - (N-1) \alpha}{m-N}$. This gives an equivalence between the critical Hardy inequality and a part of the higher dimensional subcritical Hardy inequality (ref. [@ST]). Besides, this also gives an relationship between the embedding of the subcritical Sobolev space into the Lorentz spaces for $q>p$: $$\begin{aligned}
W_0^{1,p} (B^N_R) \hookrightarrow L^{p^*, p} \hookrightarrow L^{p^*, q} \hookrightarrow L^{p^*, \infty} \end{aligned}$$ and the embedding of the critical Sobolev space into the Lorentz-Zygmund spaces for $q>N$: $$\begin{aligned}
W_0^{1,N} (B^N_R) \hookrightarrow L^{\infty, N}(\log L)^{-1} \hookrightarrow L^{\infty, q}(\log L)^{-1+ \frac{1}{N} - \frac{1}{q}} \hookrightarrow L^{\infty, \infty}(\log L)^{-1+\frac{1}{N}} = {\rm Exp L}^{\frac{N}{N-1}}.\end{aligned}$$ Since almost all transformations are applicable only for radial functions, we can expect the different phenomena from classical results for any functions, for example the existence and the non-existence of a minimizer. In fact, the author in [@S] shows the existence of a non-radial minimizer of the inequality associated with the embedding: $W_0^{1,N} (B^N_R) \hookrightarrow L^{\infty, q}(\log L)^{-1+ \frac{1}{N} - \frac{1}{q}}$. In this paper, we study an analogue of this work [@S].
Finally, as an application of this unified viewpoint for the transformations, we derive some [*infinite dimensional form*]{} of the classical Sobolev inequality in a different way from [@BP] which is a study of the logarithmic Sobolev inequality. In order to consider a limit as the dimension $m \to \infty$, we reduce the dimension $m$ to $N$ by using the following transformation (\[trans dim\]) which connects two norms between the Sobolev space $W_0^{1,p}(\re^m)$ and the lower dimensional Sobolev space $W_0^{1,p}(\re^N)$, where $p < N < m$. $$\begin{aligned}
\label{trans dim}
u(r) = w(t), \,\,\text{where}\,\, t^{-\frac{N-p}{p-1} } = r^{-\frac{m-p}{p-1} }.\end{aligned}$$ Then we can see that $$\begin{aligned}
\int_{\re^m} | \nabla u |^p \,dx
&= \frac{\w_{m-1}}{\w_{N-1}} \( \frac{m-p}{N-p} \)^{p-1} \int_{\re^N} | \nabla w |^p \,dy,\\
\int_{\re^m} | u |^{\frac{mp}{m-p}} \,dx
&= \frac{\w_{m-1}}{\w_{N-1}} \,\frac{N-p}{m-p} \int_{\re^N} \frac{| w |^{\frac{mp}{m-p}}}{|y|^{\frac{m-N}{m-p}p}} \,dy.\end{aligned}$$ Therefore the Sobolev inequality (\[HS\]) for radial functions $u \in W_0^{1,p}(\re^m)$: $$C_{m, p, 0} \( \int_{\re^m} |u|^{\frac{mp}{m-p}} dx \)^{\frac{m-p}{m}} \le \int_{\re^m} | \nabla u |^p dx$$ is equivalent to the following inequality for radial functions $w \in W_0^{1,p}(\re^N)$. $$\begin{aligned}
\label{S another}
C_{m, p, 0} \( \frac{\w_{N-1}}{\w_{m-1}} \)^{\frac{p}{m}} \( \frac{N-p}{m-p} \)^{p-\frac{p}{m}} \( \int_{\re^N} \frac{| w |^{\frac{mp}{m-p}}}{|x|^{\frac{m-N}{m-p}p}} \,dy \)^{\frac{m-p}{m}} \le \int_{\re^N} | \nabla w |^p \, dy.\end{aligned}$$ Since $$\begin{aligned}
&C_{m,p,0}=\pi^{\frac{p}{2}} m \( \frac{m-p}{p-1} \)^{p-1} \( \frac{\Gamma (\frac{m}{p}) \Gamma (m+ 1 -\frac{m}{p})}{\Gamma (m) \Gamma(1+\frac{m}{2})} \)^{\frac{p}{m}}\,\,(\text{Sobolev's best constant}), \\
&\w_{N-1} = \frac{N \pi^{\frac{N}{2}}}{\Gamma \( 1+ \frac{N}{2} \)}, \,\,
\Gamma (t) = \sqrt{2\pi} \,t^{\,t-\frac{1}{2}} \,e^{-t} + o(1) \,\, \text{as}\,\, t \to \infty \,\,(\text{Stirling's formula}),\end{aligned}$$ we have $$\begin{aligned}
&C_{m, p, 0} \( \frac{\w_{N-1}}{\w_{m-1}} \)^{\frac{p}{m}} \( \frac{N-p}{m-p} \)^{p-\frac{p}{m}} \\
&= \frac{m}{m-p} \,\frac{(N-p)^p}{(p-1)^{p-1}} \( \frac{\w_{N-1} \,(m-p)}{N-p} \)^{\frac{p}{m}} \( \frac{\Gamma (\frac{m}{p}) \Gamma \( \frac{p-1}{p} m +1 \)}{\Gamma (m+1) } \)^{\frac{p}{m}} \\
&= \frac{(N-p)^p}{(p-1)^{p-1}} \( \frac{(\frac{m}{p})^{\frac{m}{p} -\frac{1}{2}} e^{-\frac{m}{p}} \( \frac{p-1}{p} m +1\)^{\frac{p-1}{p} m +\frac{1}{2}} e^{-\frac{p-1}{p} m -1}}{(m+1)^{m+\frac{1}{2}} e^{-(m+1)} } \)^{\frac{p}{m}} +o(1)\\
&= \( \frac{N-p}{p} \)^p+o(1) \,\,(m \to \infty).\end{aligned}$$ Hence we can obtain the limit of the left-hand side of (\[S another\]) as $m \to \infty$ as follows. $$\begin{aligned}
C_{m, p, 0} \( \frac{\w_{N-1}}{\w_{m-1}} \)^{\frac{p}{m}} \( \frac{N-p}{m-p} \)^{p-\frac{p}{m}} \( \int_{\re^N} \frac{| w |^{\frac{mp}{m-p}}}{|y|^{\frac{m-N}{m-p}p}} \,dy \)^{\frac{m-p}{m}} \to
\( \dfrac{N-p}{p} \)^p \int_{\re^N} \dfrac{|w|^p}{|y|^p} dy.\end{aligned}$$
From above calculations, we can observe an interesting new aspect of the classical Hardy inequality, that is [*an infinite dimensional form*]{} of the classical Sobolev inequality. And we also see that under the transformation (\[trans dim\]), the Hardy inequality on $W_0^{1,p}(\re^m)$ is equivalent to it on $W_0^{1,p}(\re^N)$, that is, the Hardy inequality is independent of the dimension in this sense.
Proof of Theorem \[Main\]: the existence and the non-existence of the minimizer {#Proof}
===============================================================================
In this section, we prepare several Lemmas and Propositions and show Theorem \[Main\].
Note that we can apply the rearrangement technique to our minimization problem $I_a$ for $a \in [0, \frac{s(p-1)}{p(N-1)}]$. More precisely, since the potential function $V_a(x)$ is radially decreasing on $B_R$ for $a \in [0, \frac{s(p-1)}{p(N-1)}]$, the Pólya-Szegö inequality and the Hardy-Littlewood inequality imply that $$\begin{aligned}
\dfrac{\int_{B_R} | \nabla u |^p \,dx}{\( \int_{B_R} |u|^{p^*(s)} V_a (x)\, dx \)^{\frac{p}{p^*(s)}}} \ge \dfrac{\int_{B_R} | \nabla u^{\#} |^p \,dx}{\( \int_{B_R} |u^{\#}|^{p^*(s)} V_a(x)\, dx \)^{\frac{p}{p^*(s)}}} \ge I_{a, {\rm rad}}\end{aligned}$$ for any $u \in W_0^{1,p}(B_R)$ and $a \in [0, \frac{s(p-1)}{p(N-1)}]$. Therefore we have $$\begin{aligned}
\label{rearrange}
I_a = I_{a, {\rm rad}} =C_{N, p, s} \quad \text{for any} \,\,a \in \left[0, \frac{s(p-1)}{p(N-1)} \right].\end{aligned}$$ However, we can not apply the rearrangement technique to $I_a$ for $a \in ( \frac{s(p-1)}{p(N-1)}, 1]$. Therefore it seems difficult to reduce the radial setting in general.
First, instead of rearrangement, we use the following lemma by which we can reduce the radial setting when $s=p$.
\[radial nomi\] Let $1 < q <\infty$, $f=f(x)$ be a radial function on $B_R$. If there exists $C>0$ such that for any radial functions $u \in C_c^1(B_R)$ the inequality: $$\label{rad}
C \int_{B_R} |u|^q f(x)\, dx \le \int_{B_R} | \nabla u |^q\, dx$$ holds, then for any functions $w \in C_c^1(B_R)$ the inequality : $$\label{non-rad}
C \int_{B_R} |w|^q f(x)\, dx \le \int_{B_R} \left| \nabla w \cdot \frac{x}{|x|} \right|^q\, dx$$ holds.
For any $w \in C_c^1(B_R)$, define a radial function $W$ as follows. $$\begin{aligned}
W(r) = \( \w_{N-1}^{-1} \int_{\mathbb{S}^{N-1}} | w(r\w ) |^q\,dS_{\w} \)^{\frac{1}{q}} \quad (0 \le r \le R).\end{aligned}$$ Then we have $$\begin{aligned}
|W \,'(r) | &= \w_{N-1}^{-\frac{1}{q}} \( \int_{\mathbb{S}^{N-1}} | w(r\w ) |^q\,dS_{\w} \)^{\frac{1}{q}-1} \int_{\mathbb{S}^{N-1}} | w |^{q-1} \left| \frac{\pd w}{\pd r} \right| \,dS_{\w} \\
&\le \w_{N-1}^{-\frac{1}{q}} \( \int_{\mathbb{S}^{N-1}} \left| \frac{\pd w}{\pd r} (r\w ) \right|^q \,dS_{\w} \)^{\frac{1}{q}}.\end{aligned}$$ Therefore we have $$\begin{aligned}
\label{right}
\int_{B_R} | \nabla W |^q\, dx &\le \int_{B_R} \left| \nabla w \cdot \frac{x}{|x|} \right|^q\, dx, \\
\label{left}
\int_{B_R} |W|^q f(x)\, dx &= \int_{B_R} |w|^q f(x)\, dx.\end{aligned}$$ From (\[rad\]) for $W$, (\[right\]), and (\[left\]), we obtain (\[non-rad\]) for any $w$.
Second, we give a necessary and sufficient condition of the positivity of $I_a$ for $a \in [0, 1]$. As we see Proposition \[I\_a rad\], $I_{a, {\rm rad}} = C_{N, p, s} >0$ for any $s \in [0,p]$ and any $a \in [0,1]$. However, $I_a$ is not so due to the boundary singularity. This is also mentioned by [@I]. For readers convenience, we give a sketch of the proof.
\[Posi I\_a\] $I_a = 0 \iff a =1$ and $0\le s < p$.
Let $a=1$. In the similar way to [@S], set $x_{\ep} = (R - 2\ep ) \frac{y}{R}$ for $y \in \pd B_R$ and for small $\ep >0$. Then we define $u_\ep$ as follows: $$\begin{aligned}
u_{\ep}(x) =
\begin{cases}
v\( \frac{|x-x_{\ep}|}{\ep} \) \,\,\,&\text{if} \,\,\, x \in B_{\ep}(x_{\ep}), \\
0 &\text{if} \,\,\, x \in B_R \setminus B_{\ep}(x_{\ep}),
\end{cases}
\,\, \text{where}\,\, v(t)=
\begin{cases}
1 \,\,\,&\text{if} \,\,\,0\le t \le \frac{1}{2}, \\
2(1-t) &\text{if} \,\,\, \frac{1}{2} < t \le 1.
\end{cases}\end{aligned}$$ Then we have $$\begin{aligned}
\int_{B_R} | \nabla u_{\ep} (x)|^p \,dx &= \ep^{N-p} \int_{B_1} |\nabla v (|z|)|^p \,dz = C \ep^{N-p}, \\
\int_{B_R} \frac{|u_{\ep}(x)|^{p^*(s)}}{|x|^s \( 1- \,\( \frac{|x|}{R} \)^{\frac{N-p}{p-1}} \)^\beta} \,dx
&\ge C \int_{B_\ep (x_\ep )} \frac{|u_{\ep}(x)|^{p^*(s)}}{(R-|x|)^{\beta}} dx \ge \frac{C}{(3\ep )^{\beta}} \int_{B_{\frac{\ep}{2}} (x_\ep )} dx =C \, \ep^{N-\beta}.\end{aligned}$$ Hence we see that $$\begin{aligned}
I_1 \le C \ep^{N-p- (N-\beta) \frac{p}{p^*(s)}} = C \ep^{\frac{N-1}{N-s} (p-s)} \to 0\,\,\text{as}\,\, \ep \to 0\,\,\text{if}\,\, 0\le s < p.\end{aligned}$$ Therefore $I_1 = 0$ if $a=1$ and $0\le s < p$. Conversely, we can easily show that $I_a > 0$ except for that case. Indeed, if $s=p$, then $I_a = I_{a, {\rm rad}} > 0$ for any $a \in [0,1]$ from Proposition \[I\_a rad\] and Lemma \[radial nomi\]. And also, if $a < 1$, then there is no boundary singularity. Thus $I_a > 0$ for any $a \in [0, 1)$. Therefore, we obtain the necessary and sufficient condition of the positivity of $I_a$.
Third, we show that $I_a$ is monotone decreasing and continuous with respect to $a \in [0,1]$. The potential function $V_a(x)$ is continuously monotone-increasing with respect to $a \in [0, 1)$. Thus it is easy to show the monotone-decreasing property of $I_a$ with respect to $a \in [0,1]$ and the continuity of $I_a$ with respect to $a \in [0, 1)$. Here, we give a proof of the continuity of $I_a$ at $a=1$ only.
\[conti I\_a\] $I_a$ is monotone-decreasing and continuous with respect to $a \in [0, 1]$.
From the definition of $I_1$, we can take $(u_m)_{m=1}^\infty \subset C_c^{\infty}(B_R)$ and $R_m < R$ for any $m$ such that supp $u_m \subset B_{R_m}, R_m \nearrow R$, and $$\begin{aligned}
\dfrac{\int_{B_{R_m}} | \nabla u_m |^p \,dx}{\( \int_{B_{R_m}} \frac{|u_{m}(x)|^{p^*(s)}}{|x|^s \( 1- \,\( \frac{|x|}{R} \)^{\frac{N-p}{p-1}} \)^\beta}\, dx \)^{\frac{p}{p^*(s)}}} = I_1 + o(1) \quad {\rm as }\,\, m \to \infty.\end{aligned}$$ Set $\la_m = \frac{R}{R_m} >1$ and $v(y)= \la_m^{-\frac{N-p}{p}} u_m(x)$, where $y=\la_m x$. Then $$\begin{aligned}
\dfrac{\int_{B_{R_m}} | \nabla u_m |^p \,dx}{\( \int_{B_{R_m}} \frac{|u_{m}(x)|^{p^*(s)}}{|x|^s \,\( 1- \,\( \frac{|x|}{R} \)^{\frac{N-p}{p-1}} \)^\beta}\, dx \)^{\frac{p}{p^*(s)}}}
= \dfrac{\int_{B_R} | \nabla v |^p \,dx}{\( \int_{B_R} \frac{|v|^{p^*(s)}}{|y|^s \,\( 1- a_m \( \frac{|y|}{R} \)^{\frac{N-p}{p-1}} \)^\beta}\, dy \)^{\frac{p}{p^*(s)}}} \ge I_{a_m},\end{aligned}$$ where $a_m = \la_m^{-\frac{N-p}{p-1}} \nearrow 1$ as $m \to \infty$. Therefore we have $I_{a_m} \le I_1 + o(1)$. Obviously, we have $I_1 \le I_{a_m}$ from the monotone-decreasing property of $I_a$. Hence we see that $\lim_{a \nearrow 1} I_a =I_1$.
Fourth, we show the following result for the Sobolev case where $s=0$, in the similar way to [@SSW] for the Hénon problem.
\[Sobo\] Let $R < \infty$ and $f: B_R \to \re$ be a nonnegative bounded continuous function with $f \nequiv 0$. Then $$\begin{aligned}
S:= \inf_{u \in W_0^{1,p}(B_R ) \setminus \{ 0\} } \dfrac{\int_{B_R} |\nabla u |^{p} \,dx}{\( \int_{B_R} |u|^{\frac{Np}{N-p}} f(x) \,dx \)^{\frac{N-p}{N}}} = \( \max_{x \in \overline{B_R}} f(x) \)^{-\frac{N-p}{N}} C_{N, p, 0}\end{aligned}$$ and there is no minimizers of the minimization problem $S$.
Since we easily obtain $S \ge \( \max_{x \in \overline{B_R}} f(x) \)^{-\frac{N-p}{N}} C_{N, p, 0}$, we shall show $S \le \( \max_{x \in \overline{B_R}} f(x) \)^{-\frac{N-p}{N}} C_{N, p, 0}$. Let $z \in \overline{B_R}$ be a maximum point of $f$. For simplicity, we assume that $z \in B_R$. For any $\ep > 0$ there exist $T>0$ and $v \in C_c^{\infty}(B_T)$ such that $$\begin{aligned}
C_{N, p, 0}
\ge \dfrac{\int_{B_T} |\nabla v |^{p} \,dx}{\( \int_{B_T} |v|^{\frac{Np}{N-p}} \,dx \)^{\frac{N-p}{N}}} - \frac{\ep}{2}. $$ Set $u_\la (x) = \la^{\frac{N-p}{p}} v (\la (x-z))$ for $\la >0$. Then for large $\la >0$ we have $$\begin{aligned}
C_{N, p, 0}
\ge \dfrac{\int_{B_{\la^{-1}T} (z)} |\nabla u_\la |^{p} \,dx}{\( \int_{B_{\la^{-1}T} (z)} |u_\la|^{\frac{Np}{N-p}} \,dx \)^{\frac{N-p}{N}}} - \frac{\ep}{2}
&\ge f(z)^{\frac{N-p}{N}} \dfrac{\int_{B_{\la^{-1}T} (z)} |\nabla u_\la |^{p} \,dx}{\( \int_{B_{\la^{-1}T} (z)} |u_\la|^{\frac{Np}{N-p}} f(x) \,dx \)^{\frac{N-p}{N}}} - \ep \\
&\ge f(z)^{\frac{N-p}{N}} S - \ep.\end{aligned}$$ Since $\ep$ is arbitrary, we obtain $S \le \( \max_{x \in \overline{B_R}} f(x) \)^{-\frac{N-p}{N}} C_{N, p, 0}$. The case where $z \in \pd B_R$ is also showed in the same way. We omit the proof in that case.
On the other hand, the non-attainability of $S$ comes from it of $C_{N, p, 0}$. Indeed, if we assume that $v \ge 0$ is a minimizer of $S$, then $$\begin{aligned}
S= \dfrac{\int_{B_R} |\nabla v |^{p} \,dx}{\( \int_{B_R} |v|^{\frac{Np}{N-p}} f(x) \,dx \)^{\frac{N-p}{N}}}
&\ge \( \max_{x \in \overline{B_R}} f(x) \)^{-\frac{N-p}{N}} \dfrac{\int_{B_R} |\nabla u |^{p} \,dx}{\( \int_{B_R} |u|^{\frac{Np}{N-p}} \,dx \)^{\frac{N-p}{N}}} \\
&> \( \max_{x \in \overline{B_R}} f(x) \)^{-\frac{N-p}{N}} C_{N, p, 0} = S,\end{aligned}$$ where the last inequality comes from the non-attainability of $C_{N, p, 0}$. This is a contradiction.
Fifth, we show the concentration level of minimizing sequences of $I_a$ is $I_{a, \text{rad}}$ when $0 < s < p$.
\[concentration level\] Let $0 < s < p$ and $0 \le a<1$. If $I_a < I_{a, {\rm rad}} = C_{N, p, s}$, then $I_a$ is attained by a non-radial function.
In order to show Lemma \[concentration level\] also in the case where $p \neq 2$, we prepare two Lemmas. Lemma \[BM\] is concerning with almost everywhere convergence of the gradients of a sequence of solutions. This guarantees to use Lemma \[BL\] in the proof of Lemma \[concentration level\].
([@Brezis-Lieb]) \[BL\] For $p \in (0, +\infty)$, let $(g_m)_{m=1}^{\infty} \subset L^p(\Omega, \mu)$ be a sequence of functions on a measurable space $(\Omega, \mu)$ such that
1. $\| g_m \|_{L^p(\Omega, \mu)} \le \ ^{\exists} C < \infty$ for all $m \in \N$, and
2. $g_m(x) \to g(x)$ $\mu$-a.e. $x \in \Omega$ as $m \to \infty$.
Then $$\lim_{m \to \infty} \( \| g_m \|_{L^p(\Omega, \mu)}^p - \| g_m -g \|_{L^p(\Omega, \mu)}^p \) = \| g \|_{L^p(\Omega, \mu)}^p.$$
Note that we can apply Lemma \[BL\] to $\mu(dx) = f(x) dx$, where $f$ is any nonnegative $L^1(\Omega)$ function.
([@Boccardo-Murat] Theorem 2.1.)\[BM\] Let $(u_m)_{m=1}^{\infty} \subset W^{1,p}_0(\Omega)$ be such that, as $m \to \infty$, $u_m \rightharpoonup u$ weakly in $W^{1,p}_0(\Omega)$ and satisfies $$- \lap_p u_m = g_m + f_m \quad \text{in} \,\,\,\D^{\prime}(\Omega),$$ where $f_m \to 0$ in $W^{-1,p^{\prime}}_0(\Omega)$ and $g_m$ is bounded in $\M (\Omega)$, the space of Radon measures on $\Omega$, i.e. $$|< g_m, \phi > | \leq C_K \| \phi \|_{\infty}$$ for all $\phi \in \D (\Omega)$ with $\text{supp} \; \phi \subset K$. Then there exists a subsequence, say $u_{m_k}$, such that $$u_{m_k} \to u \quad \text{in} \,\,\,W^{1,\gamma}_0(\Omega ) \quad ({}^{\forall}\gamma < p).$$
Before showing Lemma \[concentration level\], we apply Lemma \[BM\] for a minimizing sequence of $I_a$. Set $$\begin{aligned}
J(u) = \| \nabla u \|_{L^p(B_R)}^{p^*(s)} - I_a^{\frac{p^*(s)}{p}} \int_{B_R} |u|^{p^*(s)} V_a(x) \,dx \quad \text{for}\,\,u \in W_0^{1,p}(B_R).\end{aligned}$$ From the definition of $I_a$, we see that ${\rm inf}_{u \in W_0^{1,p}(B_R)} J(u) = 0$. And $$\begin{aligned}
J'(u)[\varphi] = p^*(s)\, \| \nabla u \|_{L^p(B_R)}^{p^*(s)-p} \int_{B_R} |\nabla u|^{p-2} \nabla u \cdot \nabla \varphi \,dx - I_a^{\frac{p^*(s)}{p}} p^*(s) \int_{B_R} |u|^{p^*(s)-2} u \varphi V_a(x) \,dx\end{aligned}$$ for $\varphi \in \( W_0^{1,p} \)^*$. Thus we observe that $J \in C^1 \( W_0^{1,p} \,; \re \)$. Let $( u_m )_{m=1}^\infty \subset W_0^{1,p}(B_R)$ be a minimizing sequence of $I_a$ with $\int_{B_R} |u_m|^{p^*(s)} V_a(x) \,dx =1$ for any $m \in \N$ and $\| \nabla u_m \|^p_{L^p(B_R)} = I_a + o(1)$ as $m \to \infty$. By Ekeland’s Variational Principle (see e.g. [@Struwe]), there exists $( w_m )_{m=1}^\infty \subset W_0^{1,p}(B_R)$ such that $$\begin{aligned}
&(i) \,\,0 \le J(w_m) \le J(u_m) = o(1) \quad (\,m \to \infty \,), \\
&(ii) \,\,\| J'(w_m) \|_{(W_0^{1,p})^*} =o (1) \quad ( \,m \to \infty\,),\\
&(iii)\,\, \| \nabla (w_m -u_m) \|_{L^p(B_R)} = o(1)\quad ( \,m \to \infty\,).\end{aligned}$$ From (iii), we see that $( w_m )_{m=1}^\infty$ is a minimizing sequence of $I_a$ with $\int_{B_R} |w_m|^{p^*(s)} V_a(x) \,dx =1 + o(1)$ and $\| \nabla w_m \|^p_{L^p(B_R)} = I_a + o(1)$ as $m \to \infty$. Let $w_m \rightharpoonup w$ in $W_0^{1,p}(B_R)$ as $m \to \infty$, passing to a subsequence if necessary. From (ii), for any $\varphi \in W_0^{1,p}(B_R)$ we have $$\begin{aligned}
\int_{B_R} |\nabla w_m|^{p-2} \nabla w_m \cdot \nabla \varphi \,dx - I_a^{\frac{p^*(s)}{p}} \| \nabla w_m \|_{L^p(B_R)}^{p-p^*(s)} \int_{B_R} |w_m|^{p^*(s)-2} w_m \varphi V_a(x) \,dx = o(1)\,$$ which yields that $w_m$ satisfies $$\begin{aligned}
-{\rm div} (\,|\nabla w_m|^{p-2} \nabla w_m ) = I_a^{\frac{p^*(s)}{p}} \| \nabla w_m \|_{L^p(B_R)}^{p-p^*(s)} |w_m |^{p-2} w_m V_a(x) + f_m \quad \text{in}\,\, \mathcal{D}'(B_R)\end{aligned}$$ and $f_m \to 0$ in $W^{-1,p^{\prime}}_0(B_R)$. From Lemma \[BM\], passing to a subsequence if necessary, we have $\nabla w_m \to \nabla w$ a.e. in $B_R$. As a consequence, we can apply Lemma \[BL\] for $\nabla w_m$ in the proof of Lemma \[concentration level\].
Take a minimizing sequence $(u_m)_{m=1}^{\infty} \subset W_0^{1,p}(B_R)$ of $I_a$. Without loss of generality, we can assume that $$\begin{aligned}
\int_{B_R} |u_m|^{p^*(s)} V_a(x) \,dx =1,\quad \int_{B_R} |\nabla u_m|^p \,dx = I_a + o(1) \,\,{\rm as}\,\, m \to \infty.\end{aligned}$$ Since $(u_m)$ is bounded in $W_0^{1,p}(B_R)$, passing to a subsequence if necessary, $u_m \rightharpoonup u$ in $W_0^{1,p}(B_R)$ as $m \to \infty$. Replacing $u_m$ with $w_m$ (we write $u_m$ again) and applying Lemma \[BL\], we have $$\begin{aligned}
I_a &= \int_{B_R} |\nabla u_m|^p \,dx + o(1)\\
&= \int_{B_R} |\nabla (u_m -u)|^p \,dx + \int_{B_R} |\nabla u|^p \,dx +o(1) \\
&\ge I_a \( \int_{B_R} |u_m-u|^{p^*(s)} V_a (x)\, dx \)^{\frac{p}{p^*(s)}} + I_a \( \int_{B_R} |u|^{p^*(s)} V_a (x)\, dx \)^{\frac{p}{p^*(s)}} +o(1) \\
&\ge I_a \( \int_{B_R} \( |u_m-u|^{p^*(s)} + |u|^{p^*(s)} \) V_a(x) \,dx \)^{\frac{p}{p^*(s)}} +o(1) \\
&= I_a \( \int_{B_R} |u_m|^{p^*(s)} V_a (x)\, dx \)^{\frac{p}{p^*(s)}} +o(1) =I_a\end{aligned}$$ which implies that either $u \equiv 0$ or $u_m \to u \nequiv 0$ in $L^{p^*(s)}(B_R; V_a(x) dx)$ holds true from the equality condition of the last inequality and the positivity of $I_a$ for $a \in [0,1)$. We shall show that $u \nequiv 0$. Assume that $u \equiv 0$. Then we claim that $$\begin{aligned}
\label{concentrate}
I_{a, {\rm rad}} \le \int_{B_R} |\nabla u_m|^p \,dx +o(1).\end{aligned}$$ If the claim (\[concentrate\]) is true, then we see that $I_{a, {\rm rad}} \le I_a$ which contradicts the assumption. Therefore $u \nequiv 0$ which implies that $u_m \to u \nequiv 0$ in $L^{p^*(s)}(B_R; V_a(x) dx)$. Hence we have $$\begin{aligned}
1= \int_{B_R} |u|^{p^*(s)} V_a(x)\, dx, \quad \int_{B_R} |\nabla u|^p \,dx \le \liminf_{m \to \infty} \int_{B_R} |\nabla u_m|^p \,dx = I_a.\end{aligned}$$ Thus we can show that $u$ is a minimizer of $I_a$. We shall show the claim (\[concentrate\]). Since $u_m \to 0$ in $L^r (B_R)$ for any $r \in [1, p^* (0))$ and the potential function $V(x)$ is bounded away from the origin, for any small $\ep > 0$ we have $$\begin{aligned}
1= \int_{B_R} |u_m|^{p^*(s)} V_a (x)\, dx = \int_{B_{\frac{\ep R}{2}}} |u_m|^{p^*(s)} V_a (x)\, dx +o(1).\end{aligned}$$ Let $\phi_\ep$ be a smooth cut-off function which satisfies the followings: $$\begin{aligned}
0 \le \phi_\ep \le 1, \,\,\phi_\ep \equiv 1 \,\,{\rm on}\,\,B_{\frac{\ep R}{2}}(0),\,\,{\rm supp}\, \phi_\ep \subset B_{\ep R} (0), \,\,|\nabla \phi_\ep | \le C\ep^{-1}. \end{aligned}$$ Set $\tilde{u_m}(y)=u_m(x)$ and $\tilde{\phi_\ep}(y)=\phi_\ep (x)$, where $y=\frac{x}{\ep}$. Then we have $$\begin{aligned}
1&= \( \int_{B_{\frac{\ep R}{2}}} |u_m|^{p^*(s)} V_a (x)\, dx \)^{\frac{p}{p^*(s)}} +o(1) \\
&\le \( \int_{B_{\ep R} } \frac{|u_m \phi_\ep |^{p^*(s)}}{|x|^s\( 1-a\, \(\frac{|x|}{R}\)^{\frac{N-p}{p-1}}\)^{\beta}}\, dx \)^{\frac{p}{p^*(s)}} +o(1) \\
&=\( \int_{B_R } \frac{|\tilde{u_m} \tilde{\phi_\ep} |^{p^*(s)}}{|x|^s\( 1-a\ep^{\frac{N-p}{p-1}} \, \(\frac{|x|}{R}\)^{\frac{N-p}{p-1}}\)^{\beta}}\, dx \)^{\frac{p}{p^*(s)}} +o(1)
\le I_{a\ep^{\frac{N-p}{p-1}}}^{-1} \int_{B_R} |\nabla (\tilde{u_m} \tilde{\phi_\ep } ) |^p \,dx +o(1).\end{aligned}$$ We see that $a \ep^{\frac{N-p}{p-1}} \le \frac{s(p-1)}{p(N-1)}$ for small $\ep$. Since $I_{a\ep^{\frac{N-p}{p-1}}} =I_{a,\,{\rm rad}}$ for small $\ep$ by (\[rearrange\]), we have $$\begin{aligned}
1&\le I_{a, {\rm rad}}^{-1} \int_{B_R} |\nabla (\tilde{u_m} \tilde{\phi_\ep } ) |^p \,dx +o(1) \\
&\le I_{a, {\rm rad}}^{-1} \( \int_{B_{\ep R}} |\nabla u_m |^p \,dx + C \int_{B_{\ep R}} |\nabla u_m |^{p-1} | \nabla \phi_\ep | |u_m| \phi_\ep^{p-1} + |u_m |^p | \nabla \phi_\ep |^p \,dx \) +o(1) \\
&\le I_{a, {\rm rad}}^{-1} \( \int_{B_{\ep R}} |\nabla u_m |^p \,dx + pC\ep^{-1} \| \nabla u_m \|_{L^p}^{p-1} \| u_m \|_{L^p} + C \ep^{-p} \| u_m \|_{L^p}^p \) + o(1)\\
&\le I_{a, {\rm rad}}^{-1} \int_{B_{\ep R}} |\nabla u_m |^p \,dx + o(1)
\le I_{a, {\rm rad}}^{-1} \int_{B_R} |\nabla u_m |^p \,dx +o(1).\end{aligned}$$ Therefore we obtain the claim (\[concentrate\]). The proof of Lemma \[concentration level\] is now complete.
Finally, we give a proof of Theorem \[Main\].
\
(i) Let $s=p$. From Lemma \[radial nomi\] and Proposition \[I\_a rad\], we easily obtain $I_a = I_{a, {\rm rad}} = C_{N, p, p} = (\frac{N-p}{p})^p$ and the non-attainability of $I_a$. We omit the proof.
\(ii) Let $s=0$. From Proposition \[Sobo\], we obtain $I_a = C_{N, p, 0} (1-a)^{\frac{N-1}{N} p}$ and the non-attainability of $I_a$.
\(iii) Let $0 < s < p$. Note that $I_1 = 0$ by Proposition \[Posi I\_a\] and $I_a = I_{a, {\rm rad}} = C_{N, p, s}$ at least for $a \in [0, \frac{s(p-1)}{p(N-1)}]$ by (\[rearrange\]). Since $I_a$ is continuous and monotone decreasing with respect to $a \in [0, 1]$ by Lemma \[conti I\_a\], there exists $a_* \in [\frac{s(p-1)}{p(N-1)}, 1)$ such that $I_a < I_{a, {\rm rad}} = C_{N, p, s}$ for $a \in (a_*, 1)$ and $I_a = I_{a, {\rm rad}} = C_{N, p, s}$ for $a \in [0, a_*]$. Hence $I_a$ is attained by a non-radial function for $a \in (a_*, 1)$ by Lemma \[concentration level\]. On the other hand, if we assume that there exists a nonnegative minimizer $u$ of $I_a$ for $a < a_*$, then we can show that at least, $u \in C^1(B_R \setminus \{ 0\})$ and $u >0$ in $B_R \setminus \{ 0\}$ by standard regularity argument and strong maximum principle to the Euler-Lagrange equation (\[EL\]), see e.g. [@D], [@PS]. Therefore we see that $$\begin{aligned}
I_{a,{\rm rad}} = I_a = \dfrac{\int_{B_R} | \nabla u |^p \,dx}{\( \int_{B_R} \frac{| u |^{p^*(s)}}{|x|^s\( 1-a\, \(\frac{|x|}{R}\)^{\frac{N-p}{p-1}}\)^{\beta}}\, dx \)^{\frac{p}{p^*(s)}}}
> \dfrac{\int_{B_R} | \nabla u |^p \,dx}{\( \int_{B_R} \frac{| u |^{p^*(s)}}{|x|^s\( 1-a_* \(\frac{|x|}{R}\)^{\frac{N-p}{p-1}}\)^{\beta}}\, dx \)^{\frac{p}{p^*(s)}}}
\ge I_{a, {\rm rad}}.\end{aligned}$$ This is a contradiction. Therefore $I_a$ is not attained for $a \in [0, a_*)$.
Finally, we show that $a_* > A > \frac{s(p-1)}{p(N-1)}$, where $A$ is defined in Theorem \[Main\] (iii). Note that the potential function $V_a$ is increasing with respect to $a \in [0, 1]$. Since $V_a(x)$ has one critical point at $|x|=R_a := \( \frac{s (p-1)}{a p (N-1)} \)^{\frac{p-1}{N-p}} R$, $V_a$ is decreasing for $|x| < R_a$ and increasing for $|x| > R_a$. Therefore we see that $R^{-s} ( 1-a)^{-\beta} = V_a (R) = V_1 (R_1) = R^{-s} \( \frac{s (p-1)}{p (N-1)} \)^{-s} \( 1- \frac{s (p-1)}{p (N-1)} \)^{-\beta}$ for $a = A$ by the shape of $V_a$. Let $\tilde{R} \in (0, R_1)$ satisfy $V_A(\tilde{R}) = V_A (R) = V_1 (R_1)$. Then we have $
V_A^{\#} (x) = V_A (x) \,\, \text{for}\,\, x \in B_{\tilde{R}}.
$ Since $V_A^{\#}(x)$ is decreasing for $x \in B_R \setminus B_{\tilde{R}}$, $V_1 (x)$ is increasing for $x \in B_R \setminus B_{R_1}$, and $\tilde{R} < R_1$, we see that $$\begin{aligned}
V_a^{\#}(x) < V_1 (x) \,\, \text{for any}\,\, x \in B_R\,\,\text{at least for}\,\, a \in [A, A+\ep].\end{aligned}$$ Then for any $u \in W_0^{1,p}(B_R)$ we have $$\begin{aligned}
\dfrac{\int_{B_R} | \nabla u |^p \,dx}{\( \int_{B_R} | u |^{p^*(s)} V_{A+\ep} (x) \, dx \)^{\frac{p}{p^*(s)}}}
&\ge \dfrac{\int_{B_R} | \nabla u^{\#} |^p \,dx}{\( \int_{B_R} | u^{\#} |^{p^*(s)} V_{A+\ep}^{\#} (x) \, dx \)^{\frac{p}{p^*(s)}}} \\
&\ge \dfrac{\int_{B_R} | \nabla u^{\#} |^p \,dx}{\( \int_{B_R} | u^{\#} |^{p^*(s)} V_1 (x) \, dx \)^{\frac{p}{p^*(s)}}} \ge I_{1, \text{rad}}.\end{aligned}$$ Hence we have $I_{A+\ep} \ge I_{1, \text{rad}} = C_{N,p,s}$ which implies that $a_* > A$.
We can also show Theorem \[Main\] for general bounded domains with Lipschitz boundary in the similar way in [@S], since we can generalize Proposition \[Posi I\_a\] to such domains.
Appendix {#App}
========
In this section, we give remarks of the following two nonlinear scalings (\[scale N\]), (\[scale p\]) for non-radial functions $v,w$ on the unit ball $B_1$. $$\begin{aligned}
\label{scale N}
v_\la (x) &= \la^{-\frac{N-1}{N}} v(y), \,{\rm where} \,\,y= \( \frac{|x|}{b} \)^{\la -1} x, \,\,b \ge 1,\,\, \la \le 1,\\
\label{scale p}
w_\la (x) &= \la^{\frac{N-p}{p}} w(y), \,{\rm where} \,\,y= \la \left[ 1 + a \( \la^{\frac{N-p}{p-1}} -1 \) |x|^{\frac{N-p}{p-1}} \right]^{\,-\frac{p-1}{N-p}} x, \,\,0\le a \le 1,\,\, \la \ge 1.\end{aligned}$$ Note that each $y$ in (\[scale N\]), (\[scale p\]) is in $B_1$ thanks to the restriction of the length of $\la$. If $b=1$ or $a=1$ in (\[scale N\]), (\[scale p\]), then we do not need to restrict the length of $\la$. From §\[Trans\], we see that two scalings (\[scale N\]), (\[scale p\]) are equivalent to the usual scaling: $$\begin{aligned}
\label{usual scale}
u_\la (x)=\la^{\frac{N-p}{p}} u(y), \,{\rm where} \,\,y= \la x,\,\, \la >0.\end{aligned}$$ only for radial functions $u$ by the transformations (\[ST trans\]), (\[trans\]). However each response of each derivative norm $\| \nabla (\cdot)\|_{L^N}$, $\| \nabla (\cdot)\|_{L^p}$ to each scaling (\[scale N\]), (\[scale p\]) is different for non-radial functions. This is a different structure between the minimization problem $I_a$ in this paper and it in [@S], which is related to the embedding: $W_{0}^{1,N}(B_1) \hookrightarrow L^{q} \( B_1; \frac{dx}{|x|^N (\log \frac{b}{|x|})^{\frac{N-1}{N}q +1}} \)$ with $q \ge N$. More precisely, we show the followings.
\[bdd\] Let $b \ge 1$ and $a \in [0, 1]$. Then $\{ v_{\frac{1}{m}} \}_{m=1}^\infty \subset W_{0}^{1,N}(B_1)$ is unbounded for $v \in C_c^{1}(B_1)$ with $\nabla_{\mathbb{S}^{N-1}}v \nequiv 0$. On the other hand, $\{ w_m \}_{m=1}^\infty \subset W_{0}^{1,p}(B_1)$ is bounded for any $w \in C_c^{1}(B_1)$.
We use the polar coordinates. Let $r=|x|, t= |y|, \w \in \mathbb{S}^{N-1}$. By the scaling (\[scale N\]), we have $$\begin{aligned}
t= b^{1-\la} r^\la \iff r= b^{1-\frac{1}{\la}} t^{\frac{1}{\la}}.\end{aligned}$$ Then we have $$\begin{aligned}
\frac{dt}{dr} &= \la \,\frac{t}{r} \to 0 \quad (\la \to 0).\end{aligned}$$ Therefore we see that for $\la \le 1$ $$\begin{aligned}
\int_{B_1} |\nabla v_\la (x)|^{N} \,dx
&= \int_{B_R} |\nabla v_\la (x)|^{N} \,dx \\
&= \int_{\mathbb{S}^{N-1}} \int_0^R \left| \frac{\pd v_\la}{\pd r} \w + \frac{1}{r} \nabla_{\mathbb{S}^{N-1}} v_\la \,\right|^N r^{N-1} \,dr dS_{\w}\\
&= \la^{1-N} \int_{\mathbb{S}^{N-1}} \int_0^1 \left| \frac{\pd v}{\pd t} \w + \( r \frac{dt}{dr} \)^{-1} \nabla_{\mathbb{S}^{N-1}} v \,\right|^N \( r \frac{dt}{dr} \)^{N-1} \,dt dS_{\w} \\
&= \int_{\mathbb{S}^{N-1}} \int_0^1 \left| \frac{\pd v}{\pd t} \w + \( r \frac{dt}{dr} \)^{-1} \nabla_{\mathbb{S}^{N-1}} v \,\right|^N t^{N-1} \,dt dS_{\w} \\
&= \int_{\mathbb{S}^{N-1}} \int_0^1 \left| \frac{\pd v}{\pd t} \w + \frac{1}{\la t} \nabla_{\mathbb{S}^{N-1}} v \,\right|^N t^{N-1} \,dt dS_{\w} \to \infty \quad (\la \to 0).\end{aligned}$$ Hence we observe that $\{ w_{\frac{1}{m}} \}_{m=1}^\infty \subset W_{0}^{1,N}(B_1)$ is unbounded.
On the other hand, by the scaling (\[scale p\]), we have $$\begin{aligned}
t= \la r \left[ 1 + a \( \la^{\frac{N-p}{p-1}} -1 \) r^{\frac{N-p}{p-1}} \right]^{\,-\frac{p-1}{N-p}} = \la \left[ r^{-\frac{N-p}{p-1}} + a \( \la^{\frac{N-p}{p-1}} -1 \) \right]^{\,-\frac{p-1}{N-p}}\end{aligned}$$ which is equal to $$\begin{aligned}
r^{\,-\frac{N-p}{p-1}} = \la^{\frac{N-p}{p-1}} t^{-\frac{N-p}{p-1}} - a (\la^{\frac{N-p}{p-1}} -1).\end{aligned}$$ Then we have $$\begin{aligned}
\frac{dt}{dr} &= \frac{t}{r} \,\left[ 1 + a \( \la^{\frac{N-p}{p-1}} -1 \) r^{\frac{N-p}{p-1}} \right]^{-1}\\
&= \frac{t}{r} \,\left[ 1 + \frac{a \( \la^{\frac{N-p}{p-1}} -1 \)}{ \la^{\frac{N-p}{p-1}} t^{-\frac{N-p}{p-1}} - a (\la^{\frac{N-p}{p-1}} -1) } \right]^{-1}\\
&= \frac{t}{r} \,\left[ 1 + \frac{a}{ \frac{\la^{\frac{N-p}{p-1}}}{\( \la^{\frac{N-p}{p-1}} -1 \)} t^{-\frac{N-p}{p-1}} - a } \right]^{-1}
\to \frac{t}{r} \,\left[ 1 + \frac{a}{ t^{-\frac{N-p}{p-1}} - a } \right]^{-1} \quad (\la \to \infty).\end{aligned}$$ Therefore we see that for $\la \ge 1$ $$\begin{aligned}
\int_{B_1} |\nabla w_\la (x)|^{p} \,dx
&= \int_{B_R} |\nabla w_\la (x)|^{p} \,dx \\
&= \int_{\mathbb{S}^{N-1}} \int_0^R \left| \frac{\pd w_\la}{\pd r} \w + \frac{1}{r} \nabla_{\mathbb{S}^{N-1}} w_\la \,\right|^p r^{N-1} \,dr dS_{\w}\\
&= \la^{N-p} \int_{\mathbb{S}^{N-1}} \int_0^1 \left| \frac{\pd w}{\pd t} \w + \( r \frac{dt}{dr} \)^{-1} \nabla_{\mathbb{S}^{N-1}} w \,\right|^p \( \frac{dt}{dr} \)^{p-1} r^{N-1} \,dt dS_{\w} \\
&= \int_{\mathbb{S}^{N-1}} \int_0^1 \left| \frac{\pd w}{\pd t} \w + \( r \frac{dt}{dr} \)^{-1} \nabla_{\mathbb{S}^{N-1}} w \,\right|^p t^{N-1} \,dt dS_{\w} \\
&\to \int_{\mathbb{S}^{N-1}} \int_0^1 \left| \frac{\pd w}{\pd t} \w + \frac{1}{t} \,\left[ 1 + \frac{a}{ t^{-\frac{N-p}{p-1}} - a } \right] \nabla_{\mathbb{S}^{N-1}} w \,\right|^p t^{N-1} \,dt dS_{\w}\,\,(\la \to \infty).\end{aligned}$$ Hence we observe that $\{ w_m \}_{m=1}^\infty \subset W_{0}^{1,p}(B_1)$ is bounded.
As a consequence of Proposition \[bdd\], we can also construct a non-compact non-radial sequence of the embedding: $W_0^{1,p}(B_1) \hookrightarrow L^{p^*(s)}(B_1; V_a(x)\,dx)$ which is related to our minimization problem $I_a$. On the other hand, in the view of the scaling, it seems difficult to construct a non-compact non-radial sequence of the embedding: $W_{0}^{1,N}(B_1) \hookrightarrow L^{q} \( B_1; \frac{dx}{|x|^N (\log \frac{b}{|x|})^{\frac{N-1}{N}q +1}} \)$ which is related to the minimization problem in [@S]. Therefore the following natural question arises.\
[*Let $N=2$ and $b>1$. Consider the orthogonal decomposition: $W_{0}^{1,2}(B_1) = W_{0, {\rm rad}}^{1,2}(B_1) \,\oplus \, \( W_{0, {\rm rad}}^{1,2}(B_1) \)^{\perp}$. Is the embedding: $\( W_{0, {\rm rad}}^{1,2}(B_1) \)^{\perp} \hookrightarrow L^{q} \( B_1; \frac{dx}{|x|^N (\log \frac{b}{|x|})^{\frac{N-1}{N}q +1}} \)$ compact?*]{}\
If the above question is affirmative, we may say that “Non-radial compactness” occurs in some sense. That is an opposite phenomenon of Strauss’s radial compactness (ref. [@St]). Thus it may be interesting if the question is affirmative.
Acknowledgment {#acknowledgment .unnumbered}
==============
This work was (partly) supported by Osaka City University Advanced Mathematical Institute (MEXT Joint Usage/Research Center on Mathematics and Theoretical Physics). The author was supported by JSPS KAKENHI Early-Career Scientists, No. JP19K14568.
[99]{} Alvino, A., [*A limit case of the Sobolev inequality in Lorentz spaces*]{}, Rend. Accad. Sci. Fis. Mat. Napoli (4) 44 (1977), 105-112 (1978).
Aubin, T., [*Problèmes isopérimétriques et espaces de Sobolev*]{}, J. Differential Geometry 11 (1976), no. 4, 573-598.
Baras, P., Goldstein, J. A., [*The heat equation with a singular potential*]{}, Trans. Amer. Math. Soc., 284 (1984), 121-139.
W. Beckner, W., Pearson, M., [*On sharp Sobolev embedding and the logarithmic Sobolev inequalitiy*]{}, Bull. London Math. Soc., 30 (1998), 80-84.
Boccardo, L., Murat, F., [*Almost everywhere convergence of the gradients of solutions to elliptic and parabolic equations*]{}, Nonlinear Anal. TMA. 19 (1992), 581-597.
H. Brezis, E. Lieb, [*A relation between pointwise convergence of functions and convergence of functionals*]{}, Proc. Amer. Math. Soc., 88 (1983), 486-490.
Brezis, H., Vázquez, J. L., [*Blow-up solutions of some nonlinear elliptic problems*]{}, Rev. Mat. Univ. Complut. Madrid 10 (1997), No. 2, 443-469.
Cassani, D., Ruf, B., Tarsi, C., [*Optimal Sobolev type inequalities in Lorentz spaces*]{}, Potential Anal. 39 (2013), no. 3, 265-285.
DiBenedetto, E., [*$C^{1+\alpha}$ local regularity of weak solutions of degenerate elliptic equations*]{}, Nonlinear Anal. 7 (1983), No. 8, 827-850.
Flucher, M., [*Extremal functions for the Trudinger-Moser inequality in 2 dimensions*]{}, Comment. Math. Helv. 67 (1992), no. 3, 471-497.
Horiuchi, T., Kumlin, P., [*On the Caffarelli-Kohn-Nirenberg-type inequalities involving critical and supercritical weights*]{}, Kyoto J. Math. 52 (2012), no. 4, 661-742.
Ioku, N., [*Attainability of the best Sobolev constant in a ball*]{}, Math. Ann. 375 (2019), no. 1-2, 1-16.
Ioku, N., Ishiwata, M., [*A note on the scale invariant structure of critical Hardy inequalities*]{}, Geometric properties for parabolic and elliptic PDE’s, 97-120, Springer Proc. Math. Stat., 176, Springer, \[Cham\], 2016.
Lieb, E. H., [*Sharp constants in the Hardy-Littlewood-Sobolev and related inequalities*]{}, Ann. of Math. (2), 118 (1983), no. 2, 349-374.
Pucci, P., Serrin, J., [*The maximum principle*]{}, Progress in Nonlinear Differential Equations and their Applications, **73**, Birkhauser Verlag, Basel, (2007).
Sano, M., [*Extremal functions of generalized critical Hardy inequalities*]{}, J. Differential Equations 267 (2019), no. 4, 2594-2615.
Sano, M., [*Two limits on Hardy and Sobolev inequalities*]{}, arXiv:1911.04105.
Sano, M., Sobukawa, T., [*Remarks on a limiting case of Hardy type inequalities*]{}, ArXiv: 1907.09609.
Sano, M., Takahashi, F., [*Scale invariance structures of the critical and the subcritical Hardy inequalities and their improvements*]{}, Calc. Var. Partial Differential Equations 56 (2017), no. 3, Art. 69, 14 pp.
Smets, D., Willem, M., Su, J., [*Non-radial ground states for the Hénon equation*]{}, Commun. Contemp. Math. 4 (2002), no. 3, 467-480.
Strauss, W. A., [*Existence of solitary waves in higher dimensions*]{}, Comm. Math. Phys. **55** (1977), no. 2, 149-162.
Struwe, M., [*Variational methods. Applications to nonlinear partial differential equations and Hamiltonian systems. Fourth edition*]{}, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics \[Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics\], 34. Springer-Verlag, Berlin, (2008).
Talenti, G., [*Best constant in Sobolev inequality*]{}, Ann. Mat. Pura Appl. (4) **110** (1976), 353-372.
Trudinger, N. S., [*On imbeddings into Orlicz spaces and some applications*]{}, J. Math. Mech. 17 (1967), 473-483.
Vázquez, J. L., Zuazua, E., [*The Hardy inequality and asymptotic behaviour of the heat equation with an inverse-square potential*]{}, J. Funct. Anal., 173 (2000), 103-153.
Zographopoulos, N. B. [*Existence of extremal functions for a Hardy-Sobolev inequality*]{}, J. Funct. Anal. 259 (2010), no. 1, 308-314.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We study the resonance spectroscopy of the proton-rich nucleus $^7$B in the $^4$He+$p$+$p$+$p$ cluster model. Many-body resonances are treated on the correct boundary condition as the Gamow states using the complex scaling method. We predict five resonances of $^7$B and evaluate the spectroscopic factors of the $^6$Be-$p$ components. The importance of the $^6$Be($2^+$)-$p$ component is shown in several states of $^7$B, which is a common feature of $^7$He, a mirror nucleus of $^7$B. For only the ground state of $^7$B, the mixing of $^6$Be($2^+$) state is larger than that of $^6$He($2^+$) in $^7$He, which indicates the breaking of the mirror symmetry. This is caused by the small energy difference between $^7$B and the excited $^6$Be($2^+$) state, whose origin is the Coulomb repulsion.'
author:
- 'Takayuki Myo[^1]'
- 'Yuma Kikuchi[^2]'
- 'Kiyoshi Katō[^3]'
title: 'Four-body resonances of $^7$B using the complex scaling method'
---
Introduction
============
The radioactive beam experiments have provided us with much information on unstable nuclei far from the stability. In particular, the light nuclei near the drip-line exhibit new phenomena of nuclear structures, such as the neutron halo structure found in $^6$He, $^{11}$Li and $^{11}$Be [@tanihata85]. The unstable nuclei can often be unbound states beyond the particle thresholds due to the weak binding nature. The resonance spectroscopy of unbound states beyond the drip-line has also been developed experimentally. In addition to the energies and decay widths, the configuration properties are important to understand the structures of the resonances. The spectroscopic factors ($S$-factors) give the useful information to know the configurations of extra nucleons in the resonances as well as in the weakly bound states. It is also interesting to compare the structures of resonances and weakly bound states between proton-rich and neutron-rich sides, which is related to the mirror symmetry in unstable nuclei.
Recently, the experiment on $^7$B have been reported [@charity11] in addition to the old observation[@mcgrath67]. The $^7$B nucleus is known as an unbound system beyond the proton drip-line and its ground state is naively considered to be the $3/2^-$ resonance. The ground state of $^7$B is observed at 2 MeV above the $^6$Be+$p$ threshold energy and the excited states have never observed yet. The $^7$B states can decay not only to two-body $^6$Be+$p$ channels, but also to many-body channels of $^5$Li+2$p$ and $^4$He+3$p$. This multi-particle decay condition makes difficulty to identify the states of $^7$B experimentally. The mirror nucleus of $^7$B is $^7$He, which is also unbound system with respect to the one neutron emission. Recent experiments of $^7$He have been reported [@Ko99; @Bo01; @Me02; @Bo05; @Sk06; @Ry06; @beck07; @Wu08] and confirmed that its ground state is assigned to be the $3/2^-$ resonance. The $S$-factor of $^6$He-$n$ component was reported for the ground state of $^7$He[@beck07]. The excited states of $^7$He can decay into the $^4$He+3$n$ channel, which also makes difficulty to observe experimentally. There still remain contradictions in the observed energy levels of $^7$He.
From the view point of the “$^4$He+three protons / neutrons” system, the information of $^7$B and $^7$He is important to understand the structures outside the drip-lines as a four-body picture. It is also interesting to examine the effect of Coulomb interaction and the mirror symmetry in the resonances of two nuclei. Structures of resonances generally depend on the existence of the open channels as the thresholds of the particle emissions. In this sense, the mirror symmetry of resonances can be related to the coupling behavior to the open channels. It is interesting to compare the effects of the couplings to the open channels for the resonances of $^7$B and $^7$He.
In the theoretical side to treat the unbound states explicitly, several methods have been developed, such as the microscopic cluster model [@adahchour06; @arai09], the continuum shell model [@volya05] and the Gamow shell model [@betan09; @michel07]. It is, however, difficult to satisfy the multiparticle decay conditions correctly for all open channels. For $^7$B, it is necessary to describe the $^4$He+3$p$ four-body resonances in the theory. So far, no theory describes the $^7$B nucleus as four-body resonances. It is also important to reproduce the threshold energies of subsystems for particle decays, namely, the positions of open channels. Emphasizing these theoretical conditions, in this study, we employ the cluster orbital shell model (COSM) [@suzuki88; @masui06; @myo077; @myo09] of the $^4$He+$3p$ four-body system. In COSM, the effects of all open channels are taken into account explicitly[@myo077], so that we can treat the many-body decaying phenomena. In our previous works of neutron-rich systems[@myo077; @myo09; @myo10], we have successfully described the He isotopes with the $^4$He+$4n$ model up to the five-body resonances of $^8$He including the full couplings with $^{5,6,7}$He. We have described many-body resonances using the complex scaling method (CSM) [@ho83; @moiseyev98; @aoyama06] under the correct boundary conditions for all decay channels. In CSM, the resonant wave functions are directly obtained by diagonalization of the complex-scaled Hamiltonian using the $L^2$ basis functions. The successful results of light nuclei using CSM have been obtained for energies, decay widths, spectroscopic factors and also for the breakup strengths induced by the Coulomb excitations[@myo01; @myo0711], monopole transition[@myo10] and one-neutron removal[@myo09]. Recently, CSM has been developed to apply to the nuclear reaction methods such as the scattering amplitude calculation [@kruppa07], Lippmann-Schwinger equation[@kikuchi10] and the CDCC method[@matsumoto10].
In this study, we proceed with our study of resonance spectroscopy to the proton-rich nucleus, $^7$B. It is interesting to examine how our model describes $^7$B as four-body resonances. We predict the resonances of $^7$B and investigate their configuration properties. We extract the $S$-factors of the $^6$Be-$p$ components for every $^7$B resonances. The above $S$-factors are useful for understanding the coupling behavior between $^6$Be and the last proton. For mirror nucleus, $^7$He, we have performed the same analysis of the $S$-factors of the $^6$He-$n$ components [@myo09], in which the large mixing of the $^6$He($2^+$) state is confirmed. From the viewpoint of the mirror symmetry, we compare the structures of $^7$B with those of $^7$He and discuss the effect of the Coulomb interaction on the mirror symmetry. Since two nuclei are both unbound, the coupling effect of the open channels is discussed.
In Sec. \[sec:model\], we explain the complex-scaled COSM wave function and the method of obtaining the $S$-factors using CSM. In Sec. \[sec:result\], we discuss the $^7$B structures and the $S$-factors of the $^6$Be-$p$ components. Summary is given in Sec. \[sec:summary\].
Complex-scaled COSM {#sec:model}
===================
COSM for the $^4$He+ systems
----------------------------
We use COSM of the $^4$He+$N_{\rm v} p$ systems, where $N_{\rm v}$ is a valence proton number around $^4$He, namely, $N_{\rm v}=3$ for $^7$B. The Hamiltonian form is the same as that used in Refs. [@myo077; @myo09]; $$\begin{aligned}
H
&=& \sum_{i=1}^{N_{\rm v}+1}{t_i} - T_G + \sum_{i=1}^{N_{\rm v}} V^{\alpha p}_i + \sum_{i<j}^{N_{\rm v}} V^{pp}_{ij}
\\
&=& \sum_{i=1}^{N_{\rm v}} \left[ \frac{\vec{p}^2_i}{2\mu} + V^{\alpha p}_i \right] + \sum_{i<j}^{N_{\rm v}} \left[ \frac{\vec{p}_i\cdot \vec{p}_j}{4m} + V^{pp}_{ij} \right] ,
\label{eq:Ham}\end{aligned}$$ where $t_i$ and $T_G$ are the kinetic energies of each particle ($p$ and $^4$He) and of the center of mass of the total system, respectively. The operator $\vec{p}_i$ is the relative momentum between $p$ and $^4$He. The reduced mass $\mu$ is $4m/5$ using a nucleon mass $m$. The $^4$He-$p$ interaction $V^{\alpha p}$ is given by the microscopic KKNN potential [@aoyama06; @kanada79] for nuclear part, in which the tensor correlation of $^4$He is renormalized on the basis of the resonating group method in the $^4$He+$N$ scattering. For the Coulomb part, we use the folded Coulomb potential using the density of $^4$He having the $(0s)^4$ configuration. We use the Minnesota potential [@tang78] as a nuclear part of $V^{pp}$ in addition to the Coulomb interaction. These interactions reproduce the low-energy scattering of the $^4$He-$N$ and the $N$-$N$ systems, respectively.
For the wave function, $^4$He is treated as the $(0s)^4$ configuration of a harmonic oscillator wave function, whose length parameter is 1.4 fm to fit the charge radius of $^4$He as 1.68 fm. The motion of valence protons around $^4$He is solved variationally using the few-body technique. We expand the relative wave functions of the $^4$He+$N_{\rm v} p$ system using the COSM basis states [@suzuki88; @masui06; @myo077; @myo09]. In COSM, the total wave function $\Psi^J$ with a spin $J$ is represented by the superposition of the configuration $\Psi^J_c$ as $$\begin{aligned}
\Psi^J
&=& \sum_c C^J_c \Psi^J_c,
\label{WF0}
\\
\Psi^J_c
&=& \prod_{i=1}^{N_{\rm v}} a^\dagger_{\alpha_i}|0\rangle,
\label{WF1}\end{aligned}$$ where the vacuum $|0\rangle$ is given by the $^4$He ground state. The creation operator $a^\dagger_{\alpha}$ is for the single particle state of a valence proton above $^4$He with the quantum number $\alpha=\{n,\ell,j\}$ in a $jj$ coupling scheme. Here, the index $n$ represents the different radial component. The coefficient $C^J_c$ represents the amplitude of the configuration and its index $c$ represents the set of $\alpha_i$ as $c=\{\alpha_1,\cdots,\alpha_{N_{\rm v}}\}$. We take a summation over the available configurations in Eq. (\[WF0\]), which give a total spin $J$.
![Sets of the spatial coordinates in COSM for the $^4$He+$N_{\rm v} p$ system.[]{data-label="fig:COSM"}](COSM2.eps){width="7.5cm"}
The coordinate representation of the single particle state corresponding to $a^\dagger_{\alpha}$ is given as $\psi_{\alpha}(\vc{r})$ as function of the relative coordinate $\vc{r}$ between the center of mass of $^4$He and a valence proton [@suzuki88], as shown in Fig. \[fig:COSM\]. Considering the angular momentum coupling, the explicit wave functions of the COSM configuration $\Psi^J_c$ in Eq. (\[WF1\]) are expressed as $$\begin{aligned}
\Psi^J_c
&=& {\cal A}^\prime \left\{\, [\Phi(^4{\rm He}), \chi^{J}_c(N_{\rm v} p)]^J\, \right\},
\label{eq:WF}
\\
\chi^{J}_c(p)
&=& \psi_{\alpha_1}^J,
\\
\chi^{J}_c(2p)
&=& {\cal A}\{ [\psi_{\alpha_1},\psi_{\alpha_2}]_J \},
\label{eq:WF6}
\\
\chi^{J}_c(3p)
&=& {\cal A}\{ [[\psi_{\alpha_1},\psi_{\alpha_2}]_{j_{12}},\psi_{\alpha_3}]_J \}.
\label{eq:WF7}\end{aligned}$$ Here, $\Phi(^4{\rm He})$ is the $^4$He wave function with spin $0^+$. The function $\chi^J_c(N_{\rm v} p)$ expresses the COSM wave functions for the valence protons. The spin $j_{12}$ is a coupled angular momentum of the first and second valence protons. The antisymmetrizers between valence protons and between a valence proton and nucleons in $^4$He are expressed as the symbols ${\cal A}$ and ${\cal A}^\prime$, respectively. The effect of ${\cal A}^\prime$ is treated in the orthogonality condition model[@myo09; @aoyama06], in which $\psi_{\alpha}$ is imposed to be orthogonal to the $0s$ state occupied by $^4$He. We employ a sufficient number of radial bases of $\psi_\alpha$ to describe the spatial extension of valence protons in the resonances, in which $\psi_\alpha$ are normalized. In this model, the radial part of $\psi_\alpha$ is expanded with the Gaussian basis functions for each orbit as $$\begin{aligned}
\psi_\alpha
&=& \sum_{k=1}^{N_{\ell j}} d^k_{\alpha}\ \phi_{\ell j}^k(\vc{r},b_{\ell j}^k),
\label{WFR}
\\
\phi_{\ell j}^k(\vc{r},b_{\ell j}^k)
&=& {\cal N}\, r^{\ell} e^{-(r/b_{\ell j}^k)^2/2} [Y_{\ell}(\hat{\vc{r}}),\chi^\sigma_{1/2}]_{j}.
\label{Gauss}\end{aligned}$$ The index $k$ is for the Gaussian basis with the length parameter $b_{\ell j}^k$. Normalization factor of the basis and a basis number are given by ${\cal N}$ and $N_{\ell j}$, respectively.
In the COSM using Gaussian expansion, the total wave function $\Psi^J$ contains two-kinds of the expansion coefficients $\{C_c^J\}$ in Eq. (\[WF0\]) for configuration and $\{d^k_\alpha\}$ in Eq. (\[WFR\]) for each valence proton. We determine them in the following procedure: First, we solve the eigenvalue problem of the norm matrix of the Gaussian basis set in Eq. (\[Gauss\]), which are non-orthogonal, with the dimension $N_{\ell j}$. The coefficients $\{d^k_\alpha\}$ are determined to construct the orthonormalized single-particle basis set $\{\psi_\alpha\}$ having different radial components with the number $N_{\ell j}$. Second, Hamiltonian matrix elements are constructed using $\{\psi_\alpha\}$ and diagonalized to determine $\{C_c^J\}$ from the variational principle. The relation $\sum_{c} \left(C_c^J\right)^2=1$ is satisfied due to the normalization of the total wave function. The same method of determining the expansion coefficients using Gaussian bases is used in the tensor-optimized shell model[@myo11].
The numbers of the radial bases $N_{\ell j}$ of $\psi_\alpha$ are determined to converge the physical solutions $\Psi^J$. The length parameters $b_{\ell j}^k$ are chosen in geometric progression [@myo09; @aoyama06]. We use at most 17 Gaussian basis functions by setting $b_{\ell j}^k$ from 0.2 fm to around 40 fm with the geometric ratio of 1.4 as a typical one. Due to the expansion of the radial wave function using a finite number of basis states, all the energy eigenvalues are discretized for bound, resonant and continuum states. For reference, in the Gamow shell model calculation [@betan09; @michel07], the single particle states $\psi_\alpha$ consist of the resonant and the discretized continuum states obtained with the single particle potential $V^{\alpha p}$ in Eq. (\[eq:Ham\]).
For $^7$B, all the channels of $^6$Be+$p$, $^5$Li+$2p$ and $^4$He+$3p$ are automatically included in the total COSM wave function $\Psi^J$. These components are coupled to each other via the interactions and the antisymmetrization. The couplings depend on the relative distances between $^4$He and a valence proton and between the valence protons. We explain the coupling behavior between $^4$He and valence protons in COSM. This is related to the boundary condition of the proton emission in $^7$B, which is important when the resonant and continuum states are treated[@myo077; @myo0711; @myo05]. As an example, we consider the coupling between $^{7}$B and the $^6$Be+$p$ configurations. Asymptotically, when the last proton is located far away from $^6$[Be]{}, namely, $\vc{r}_3\to\infty$ in Fig. \[fig:COSM\], any coupling between $^6$Be and a last proton disappears, and $^6$Be becomes its isolated eigenstate of the Hamiltonian in Eq. (\[eq:Ham\]) with $N_{\rm v}=2$.
$$\begin{aligned}
\Psi^J(^{7}{\rm B})
&=& \sum_c C_c^J {\cal A}^\prime\left\{ [ \Phi(^{4}\mbox{He}), \chi_c^J(3p) ]^J \right\}
\label{asympt0}
\\
&\mapleft{\vc{r}_3\to\infty}&
\left[ \Psi^{J^\prime}_\nu(^{6}\mbox{Be}), \psi_{\alpha_3} \right]^{J},
\label{asympt1}
\\
\Psi^{J^\prime}_\nu(^{6}\mbox{Be})
&=& \sum_c C_{c,\nu}^{J^\prime} {\cal A}^\prime \left\{ [ \Phi(^{4}\mbox{He}), \chi_{c,\nu}^{J^\prime}(2p) ]^{J^\prime} \right\},
\label{asympt2}\end{aligned}$$
where the spin $J$ and $J'$ are for $^7$B and $^6$Be, respectively, and the index $\nu$ indicates the eigenstate of $^6$He. The mixing coefficients $\{C^{J^\prime}_{c,\nu}\}$ and the wave function $\chi^{J^\prime}_{c,\nu}(2p)$ in Eq. (\[asympt2\]) are those of the $^6$Be eigenstates. Hence, the wave function $\chi^{J}_c(3p)$ in Eq. (\[asympt0\]) satisfies the following asymptotic forms $$\begin{aligned}
\sum_c C_c^J \chi^{J}_c(3p)
&\mapleft{\vc{r}_3\to\infty} & \left( \sum_c C_{c,\nu}^{J^\prime} \chi^{J^\prime}_{c,\nu}(2p)\right) \psi_{\alpha_3} .
\label{asympt3}\end{aligned}$$ This relation implies that the wave function of three valence protons of $^7$B is asymptotically decomposed into $^6$Be and a last proton. Equations (\[asympt0\])-(\[asympt3\]) determine the boundary condition of COSM. Contrastingly, when a last proton comes close to $^6$Be, the last proton dynamically couples to the $^6$Be eigenstates $\Psi_\nu^{J^\prime}$. This coupling depends on the relative distance between $^6$Be and a last proton, and changes the $^6$Be configurations from the isolated eigenstates of $^6$Be. In COSM, the structure change of $^6$Be inside $^7$B is determined variationally to optimize the $^7$B eigenstates. The same discussion is applied to the asymptotic conditions for the $^5$Li+$2p$ and $^4$He+$3p$ configurations. Hence, the proton emissions can be handled with the correct boundary conditions in COSM.
We explain the parameters of the model space of COSM and the Hamiltonian which are determined in the previous analyses of He isotope[@myo077; @myo09]. For the single-particle states, we take the angular momenta $\ell\le 2$ to keep the accuracy of the converged energy within 0.3 MeV of $^6$He with the $^4$He+$n$+$n$ model in comparison with the full space calculation[@aoyama06]. In this model, we adjust the two-neutron separation energy of $^6$He($0^+$) to the experiment of 0.975 MeV by taking the 173.7 MeV of the repulsive strength of the Minnesota potential instead of the original value of 200 MeV. The adjustment of the $NN$ interaction is originated from the pairing correlation between valence protons with higher angular momenta $\ell>2$ [@aoyama06]. Hence, the present model reproduces the observed energies of $^{6}$He and is applied to the proton-rich nuclei in this analysis.
Complex scaling method (CSM)
----------------------------
We explain CSM, which describes resonances and nonresonant continuum states [@ho83; @moiseyev98; @aoyama06]. Hereafter, we refer to the nonresonant continuum states as simply the continuum states. In CSM, we transform the relative coordinates of the $^4$He+$N_{\rm v} p$ system, as $\vc{r}_i \to \vc{r}_i\, e^{i\theta}$ for $i=1,\cdots,N_{\rm v}$, where $\theta$ is a scaling angle. The Hamiltonian in Eq. (\[eq:Ham\]) is transformed into the complex-scaled Hamiltonian $H_\theta$, and the corresponding complex-scaled Schrödinger equation is given as $$\begin{aligned}
H_\theta\Psi^J_\theta
&=& E\Psi^J_\theta .
\label{eq:eigen}\end{aligned}$$ The eigenstates $\Psi^J_\theta$ are obtained by solving the eigenvalue problem of $H_\theta$ in Eq. (\[eq:eigen\]). In CSM, we obtain all the energy eigenvalues $E$ of bound and unbound states on a complex energy plane, governed by the ABC theorem [@ABC]. In this theorem, it is proved that the boundary condition of resonances is transformed to one of the damping behavior at the asymptotic region. This condition makes it possible to use the same method of obtaining the bound states and resonances. For a finite value of $\theta$, every Riemann branch cut starting from the different thresholds is commonly rotated down by $2\theta$. Hence, the continuum states such as $^6$Be+$p$ and $^5$Li+2$p$ channels in $^7$B are obtained on the branch cuts rotated by the $-2\theta$ from the corresponding thresholds [@myo077; @myo09]. On the contrary, bound states and resonances are obtainable independently of $\theta$. We can identify the resonance poles with complex eigenvalues: $E=E_r-i\Gamma/2$, where $E_r$ and $\Gamma$ are the resonance energies and the decay widths, respectively. In the wave function, the $\theta$ dependence is included in the expansion coefficients in Eqs. (\[WF0\]) and (\[WFR\]) as $\{C_c^J(\theta)\}$ and $\{d_\alpha^k(\theta)\}$, respectively. The value of the angle $\theta$ is determined to search for the stationary point of each resonance in a complex energy plane[@aoyama06; @ho83; @moiseyev98].
The resonant state generally has a divergent behavior at asymptotic distance and then its norm is defined by a singular integral such as using the convergent factor method[@aoyama06; @homma97; @romo68]. In CSM, on the other hand, resonances are precisely described as eigenstates expanded in terms of the $L^2$ basis functions. The amplitudes of the resonances are finite and normalized as $\sum_{c} \left(C_c^J(\theta)\right)^2=1$. The Hermitian product is not applied due to the bi-orthogonal relation [@ho83; @moiseyev98; @berggren68]. The matrix elements of resonances are calculated using the amplitudes obtained in CSM.
In this study, we discretize the continuum states in terms of the basis expansion, as shown in the figures of energy eigenvalue distributions in Refs. [@myo01; @myo09; @aoyama06]. The reliability of the continuum discretization in CSM has already been shown using the continuum level density[@suzuki05] and the phase shift analysis[@kruppa07].
Spectroscopic factor of $^7$B
-----------------------------
We explain the $S$-factors of the $^6$Be-$p$ components for $^7$B. As was explained in the previous study [@myo09], since the resonant states generally give complex matrix elements, The $S$-factors of resonant states are not necessarily positive definite and defined by the squared matrix elements using the bi-orthogonal property [@berggren68] as $$\begin{aligned}
S^{J,\nu}_{J',\nu'}
&=& \sum_\alpha S^{J,\nu}_{J',\nu',\alpha}\, ,
\\
S^{J,\nu}_{J',\nu',\alpha}
&=& \frac{1}{2J+1} \langle \widetilde{\Phi}^{J'}_{\nu'}||a_\alpha ||\Psi^J_\nu \rangle^2\, ,
\label{eq:S}\end{aligned}$$ where the annihilation operator $a_\alpha$ is for single valence proton with the state $\alpha$. The spin $J$ and $J'$ are for $^7$B and $^6$Be, respectively. The index $\nu$ ($\nu'$) indicates the eigenstate of $^7$B ($^6$Be). The wave function $\Phi^{J'}_{\nu'}$ is for $^6$Be. In this expression, the values of $S^{J,\nu}_{J',\nu'}$ are allowed to be complex. In general, an imaginary part of the $S$-factors often becomes large relative to the real part for a resonance having a large decay width. Recently, the Gamow shell model calculation also discuss the $S$-factors of resonances [@michel10].
The sum rule value of $S$-factors, which includes resonance contributions of the final states, can be considered [@myo077]. When we count all the $S$-factors not only of resonances but also of the continuum states in the final states, the summed value of the $S$-factors is equal to the associated particle number, which is a real value and does not contain any imaginary part, as similar to the transition strength calculation[@myo01; @myo03]. For $^7$B into the $^6$Be-$p$ decomposition, the summed value of the $S$-factor $S^{J,\nu}_{J',\nu'}$ in Eq. (\[eq:S\]) by taking all the $^6$Be states, is given as $$\begin{aligned}
\sum_{J',\nu'}\ S^{J,\nu}_{J',\nu'}
&=& \sum_{\alpha,m}\
\langle \widetilde{\Psi}^{JM}_\nu|a^\dagger_{\alpha,m} a_{\alpha,m}| \Psi^{JM}_\nu \rangle
\nonumber
\\
&=& 3\ ,
\label{eq:sf-sum}\end{aligned}$$ where we use the completeness relation of $^6$Be as $$\begin{aligned}
1
&=& \sum_{J',M'}\sum_{\nu'}\hspace*{-0.5cm}\int |\Phi^{J'M'}_{\nu'}\rangle \langle \widetilde{\Phi}^{J'M'}_{\nu'}|.\end{aligned}$$ Here $M$ ($M'$) and $m$ are the $z$-components of the angular-momenta of the wave functions of $^7$B ($^6$Be) and of the creation and annihilation operators of the valence protons, respectively. It is found that the summed value of the $S$-factors for the $^6$Be states becomes the valence proton number $N_{\rm v}$ of $^7$B. This discussion of the $S$-factors is valid when the complex scaling is operated. It is also shown that $S$-factors of the resonances are invariant with respect to the scaling angle $\theta$ [@myo09; @homma97].
The present $S$-factors can be used to obtain the strengths of the proton removal reaction from $^7$B into $^6$Be as a function of the energy of $^6$Be. In the calculation, the $S$-factors not only of the resonances, but also of the many-body continuum states for $^7$B and $^6$Be are necessary. The complex-scaled Green’s function is also used to calculate the strength distribution [@myo09; @myo01; @myo98]. In fact, for neutron-rich case, we have shown the one-neutron removal strength distributions from $^7$He into the $^6$He states using CSM[@myo09]. The strength into the three-body scattering states of $^6$He as $^4$He+$n$+$n$ was successfully obtained by using the complex-scaled wave function of $^6$He. It was shown that the $^6$He($2^+$) resonance generates a sharp peak at around the resonance energy in the distribution.
In the numerical calculation, we express the radial part of the operator $a_\alpha$ in Eq. (\[eq:S\]) using the complete set expanded by 40 Gaussian basis functions with the maximum range of 100 fm for each orbit. This treatment is sufficient to converge the $S$-factor results.
Results {#sec:result}
=======
Energy spectra of $^5$Li, $^6$Be and $^7$B
------------------------------------------
We show the systematic behavior of level structures of $^5$Li, $^6$Be and $^7$B in Fig. \[fig:B7\]. It is found that the present calculations agree with the observed energy levels. We furthermore predict many resonances for $^6$Be and $^7$B. We first discuss the structures of $^6$Be, which are useful for the understanding of the $^7$B structures. The $^6$Be states together with a last proton compose the thresholds of the decay of $^7$B. It is also interesting to compare the $^6$Be structures with those of $^6$He, a mirror and a neutron halo nucleus.
![Energy levels of $^5$Li, $^6$Be and $^7$B measured from the $^4$He energy. Units are in MeV. Black and gray lines are theory and experiments, respectively. Small numbers are decay widths.[]{data-label="fig:B7"}](7B_lev3.eps){width="8.5cm"}
Energy \[MeV\] Width \[MeV\] Configuration
--------- ------------------- --------------- ---------------------- --
$0^+_1$ $1.383$ ($1.370$) 0.041(0.092) $(p_{3/2})^2$
$0^+_2$ $5.95$ 11.21 $(p_{1/2})^2$
$2^+_1$ $2.90$ ($3.04$) 1.05 (1.16) $(p_{3/2})^2$
$2^+_2$ $4.63$ 5.67 $(p_{3/2})(p_{1/2})$
$1^+$ $4.76$ 7.75 $(p_{3/2})(p_{1/2})$
: Energy eigenvalues of the $^{6}$Be resonances measured from the $^4$He+$p$+$p$ threshold. The values with parentheses are the experimental ones[@aj89]. Dominant configurations are listed.[]{data-label="ene6"}
Config. $^6$Be($0^+_1$) $^6$He($0^+_1$)
---------------- ----------------- ----------------- --
$(p_{3/2})^2$ $0.918-i0.006$ 0.917
$(p_{1/2})^2$ $0.041+i0.000$ 0.043
$(1s_{1/2})^2$ $0.010+i0.006$ 0.009
$(d_{5/2})^2$ $0.024+i0.000$ 0.024
$(d_{3/2})^2$ $0.007+i0.000$ 0.007
: Components of the ground states of $^{6}$Be and $^6$He.[]{data-label="comp6_0"}
Config. $^6$Be($2^+_1$) $^6$He ($2^+_1$)
---------------------- ----------------- ------------------ --
$(p_{3/2})^2$ $0.891+i0.030$ $0.898+i0.013$
$(p_{3/2})(p_{1/2})$ $0.097-i0.024$ $0.089-i0.013$
: Dominant components of the $2^+_1$ states of $^{6}$Be and $^6$He.[]{data-label="comp6_2"}
$^6$Be $^6$He $^6$He(exp.)
------------------------- ---------------- ------------ --------------------------------------------------------- --
$R_{\rm m}$ 2.80 + $i$0.17 2.37 2.33(4)$^{\rm a}$, 2.30(7)$^{\rm b}$, 2.37(5)$^{\rm c}$
$R_p$ 3.13 + $i$0.20 1.82
$R_n$ 1.96 + $i$0.08 2.60
$R_{\rm ch}$ 3.25 + $i$0.21 2.01 2.068(11)$^{\rm d}$
$r_{NN}$ 6.06 + $i$0.35 4.82
$r_{{\rm c}\mbox{-}2N}$ 3.85 + $i$0.37 3.15
$\theta_{NN}$ 75.3 74.6
: Radial properties of the ground states of $^6$Be and $^6$He in units of fm, in comparison with the experiments of $^6$He; a[@tanihata92], b[@alkazov97], c[@kiselev05], d[@mueller07].[]{data-label="radius"}
The resonance energies and the decay widths of $^6$Be are listed in Table \[ene6\] with dominant configurations. The components of each configuration for the $^6$Be and $^6$He ground states are listed in Table \[comp6\_0\], which are the square values of the amplitudes $\{C^J_c\}$ defined in Eq. (\[WF0\]). We show the summation of the components belonging to the same configurations with different radial components of a valence proton. It is noted that the amplitude of resonant wave function becomes a complex number and its real part can have a physical meaning when the imaginary part has relatively a small value. It is confirmed that two ground states show the similar trend of configurations, which is dominated by $p$-shell. The configurations of the $2^+_1$ states of $^6$Be and $^6$He are also shown in Table \[comp6\_2\], where the energy and decay width of $^6$He($2^+_1$) are obtained as ($E_r$, $\Gamma$)=(0.879, 0.132) in MeV, measured from the $^4$He+$n$+$n$ threshold. The good correspondence is seen for the dominant two configurations of the $2^+_1$ states. These results indicate that the mirror symmetry is kept well for the configurations between $^6$Be and $^6$He. Recently, Gamow shell model calculation discussed the $p$-shell contributions in the A=6 system[@michel10].
The radial properties of $^6$Be are interesting to discuss the effect of the Coulomb repulsion in comparison with $^6$He having a halo structure, although the radius of $^6$Be can be complex numbers because of the resonance. The results of the $^6$Be ground state are shown in Table \[radius\] for matter ($R_{\rm m}$), proton ($R_p$), neutron($R_n$) charge ($R_{\rm ch}$) parts, and the relative distances between valence nucleons ($r_{NN}$) and between the $^4$He core and the center of mass of two valence nucleons ($r_{{\rm c}\mbox{-}2N}$), and the opening angle between two nucleons ($\theta_{NN}$) at the center of mass of the $^4$He core. It is found that the values in $^6$Be are almost real, so that the real parts can be regarded to represent the radius properties of $^6$Be. The distances between valence protons and between core and $2p$ in $^6$Be are wider than those of $^6$He by 26% and 22%, respectively. This result comes from the Coulomb repulsion between three constituents of $^4$He+$p$+$p$ in $^6$Be. The Coulomb repulsion makes the energy of $^6$Be shift up to be a resonance in comparison with $^6$He, and also increases the relative distances between each constituent from the halo state of $^6$He.
![(Color online) Energy eigenvalue distribution of $^7$B in complex energy plane.[]{data-label="fig:ene_7B"}](ene_7B02.eps){width="8.0cm"}
We discuss the structures of $^7$B. The energy eigenvalues are listed in Table \[ene7\] measured from the $^4$He+3$p$ threshold. We obtain the five resonances which are all located above the $^6$Be($0^+_1$)+$p$ threshold, as shown in Fig. \[fig:B7\], and four-body resonances. In Fig. \[fig:ene\_7B\], we display the energy eigenvalues of the $^7$B resonances together with the many-body continuum cuts on the complex energy plane, which is useful to understand the positions of poles and the various thresholds relatively at glance. The $^6$Be resonances together with a last proton compose the thresholds of $^7$B, whose positions are located at the starting points of the $-2\theta$-rotated cuts in CSM. The energy of the $^7$B ground state is obtained as $E_r$=3.35 MeV and agrees with the recent experiment of $E_r=3.38(3)$ MeV[@charity11]. The decay width is 0.49 MeV, which is good but slightly smaller than the experimental value of 0.80(2) MeV. In the experiment, the decay width is determined from the $R$-matrix theory on the assumption of the decay into the $^6$Be($0^+_1$)+$p$ channel. On the other hand, our analysis shows that the $^6$Be($2^+_1$)-$p$ component is important in the $^7$B ground state, which is found from the $S$-factors of this channel and is suggested from the conventional shell model calculation[@charity11]. There is no experimental evidence for the excited states of $^7$B so far and it is desired that further experimental data are coming.
Energy \[MeV\] Width \[MeV\] Configuration
----------- ------------------ ---------------- ------------------------ --
$3/2^-_1$ $3.35$ (3.38(3)) 0.49 (0.80(2)) $(p_{3/2})^3$
$3/2^-_2$ $6.92$ 5.422 $(p_{3/2})^2(p_{1/2})$
$3/2^-_3$ $8.39$ 9.86 $(p_{3/2})(p_{1/2})^2$
$1/2^- $ $5.93$ 4.73 $(p_{3/2})^2(p_{1/2})$
$5/2^- $ $4.63$ 3.91 $(p_{3/2})^2(p_{1/2})$
: Energy eigenvalues of the $^7$B resonances measured from the $^4$He+3$p$ threshold. The values with parentheses are the experimental ones[@charity11]. Dominant configurations are listed.[]{data-label="ene7"}
------------------------ ---------------- ------------------------ ---------------- ------------------------ ---------------- --
$(p_{3/2})^3$ $0.923+i0.002$ $(p_{3/2})^2(p_{1/2})$ $0.795+i0.032$ $(p_{3/2})(p_{1/2})^2$ $0.770+i0.053$
$(p_{3/2})(p_{1/2})^2$ $0.020+i0.004$ $(p_{3/2})(p_{1/2})^2$ $0.195-i0.035$ $(p_{3/2})^2(p_{1/2})$ $0.182-i0.050$
$(p_{3/2})^2(p_{1/2})$ $0.021-i0.007$ $(d_{3/2})^2(p_{3/2})$ $0.006+i0.001$ $(p_{3/2})^3$ $0.003-i0.002$
------------------------ ---------------- ------------------------ ---------------- ------------------------ ---------------- --
------------------------- ---------------- ------------------------------- ---------------- --
$(p_{3/2})^2(p_{1/2})$ $0.969-i0.000$ $(p_{3/2})^2(p_{1/2})$ $0.957+i0.006$
$(d_{5/2})^2(p_{1/2})$ $0.018-i0.002$ $(d_{3/2})(d_{5/2})(p_{3/2})$ $0.015-i0.003$
$(1s_{1/2})^2(p_{1/2})$ $0.005+i0.002$ $(d_{3/2})^2(p_{3/2})$ $0.008-i0.001$
------------------------- ---------------- ------------------------------- ---------------- --
![Energy levels of He isotopes measured from the $^4$He energy. Units are in MeV. Black and gray lines are theory and experiments, respectively. Small numbers are decay widths.[]{data-label="fig:He7"}](7He_lev4.eps){width="8.5cm"}
![Excitation energy spectra of mirror nuclei of A=5,6,7 in the units of MeV.[]{data-label="fig:excite"}](7B_lev5.eps){width="8.0cm"}
We discuss the configuration properties of each resonance of $^7$B in detail. In Table \[conf\], we list the main configurations with their squared amplitudes $(C^J_c)^2$ in Eq. (\[WF0\]) for each $^7$B resonance. In general, the squared amplitude of resonant state can be a complex number, while the total of the squared amplitudes is normalized as unity. The interpretation of the imaginary part in the physical quantity of resonances is still an open problem[@homma97]. In the results of $^7$B, the amplitudes of the dominant components are almost real values. It is, hence, expected to discuss the physical meaning of the dominant components of the resonances in the same way as the bound state. It is furthermore found that the imaginary parts of the configurations are canceled to each other for every resonance and their summations have much smaller imaginary parts. When we consider all the available configurations, the summations conserve unity due to the normalization of the states.
For the $3/2^-$ ground state, the result indicates that the $(p_{3/2})^3$ configuration is dominant with a small mixing of the $p_{1/2}$ component. For the excited $3/2^-_2$ state, one proton occupies the $p_{1/2}$ orbit and the residual two protons in $p_{3/2}$ form the spin of $2^+$, which corresponds to the $^6$Be($2^+_1$) configuration as shown in Table \[comp6\_2\]. The importance of the $^6$Be($2^+_1$)-$p$ component in the $3/2^-_2$ state of $^7$B is discussed from the viewpoint of the $S$-factors. It is also found that the two-particle excitation into the $(p_{1/2})^2$ configuration is mixed by about 20%. The $3/2^-_3$ state is dominated by the $(p_{3/2})(p_{1/2})^2$ configuration, in which the $(p_{1/2})^2$ part is the same configuration of $^6$Be($0^+_2$).
The $1/2^-$ state of $^7$B corresponds to the one particle excitation from the ground state. Its decay width, 4.73 MeV is large and comparable to the resonance energy, 5.93 MeV among the five resonances of $^7$Be. This is confirmed from Fig. \[fig:ene\_7B\] as the large ratio of the imaginary part to the real one in the complex energy plane. The result of the large decay width is similar to the $^5$Li($1/2^-$) state in the $^4$He+$p$ system. In comparison with $^5$Li, whose resonance energy is 2.93 MeV with the decay width of 6.49 MeV, the $^7$B($1/2^-$) state of has a smaller decay width. This difference comes from the residual two protons occupying the $p_{3/2}$ orbit in $^7$B. The attractive contribution between the $p_{1/2}$ proton and other two protons makes the decay width of the $1/2^-$ state smaller. In the $5/2^-$ state, the $2^+$ component of $(p_{3/2})^2$ plus $p_{1/2}$ is dominant. This coupling scheme is similar to the $3/2^-_2$ case. In relation to the configuration properties of $^7$B, it is interesting to examine the $^6$Be-$p$ components in each $^7$B state, which is performed using the $S$-factors.
It is interesting to discuss the mirror symmetry between $^7$B and $^7$He consisting of $^4$He and three valence protons or neutrons. To do this, we show the energy spectra of He isotopes with COSM in Fig. \[fig:He7\], using the Hamiltonian in Eq. (\[eq:Ham\]) without the Coulomb term. The experimental data of $^7$He($1/2^-$) is not fixed[@Me02; @Bo05; @Sk06; @Ry06; @beck07; @Wu08], so that we do not put the data in the figure. From Figs. \[fig:B7\] and \[fig:He7\], it is found that the order of energy levels are the same between proton-rich and neutron-rich sides. In the proton-rich side, the whole spectra are shifted up due to the Coulomb repulsion in comparison with those of the neutron-rich side. The displacement energies are about 2.5 MeV for $^6$Be from $^6$He, and about 4 MeV for $^7$B from $^7$He, respectively. In Fig. \[fig:excite\], we compare the excitation energy spectra of proton-rich and neutron-rich sides. It is found that the good symmetry is confirmed between the corresponding nuclei. The differences of excitation energies for individual levels are less than 1 MeV. The properties of the configurations of $^7$B and $^7$He are discussed in terms of $S$-factors, next.
Spectroscopic factors of $^7$B {#sec:sfac}
------------------------------
We obtain the information of the structures of $^7$B via the $S$-factors. In this study, we extract the $S$-factors of the $^6$Be-$p$ components in $^7$B. This quantity is important to examine the coupling behavior between $^6$Be and a last proton including the excitations of $^6$Be. We choose the $0^+_1$ and $2^+_1$ states of $^6$Be, which are observed experimentally. In this analysis, both of initial ($^7$B) and final ($^6$Be) states are resonances, so that the $S$-factors become complex numbers. The present $S$-factors correspond to the components of $^6$Be in the $^7$B resonances and contain the imaginary parts. It is still difficult to derive the definite conclusion of the interpretation of the imaginary part in the $S$-factors, as was mentioned in the previous studies [@myo09]. The further theoretical and mathematical developments would be desired to solve this problem.
In Table \[sf\_B7\], we list the results of $S$-factors of $^7$B. For comparison, the results of $^7$He are shown in Table \[sf\_He7\]. It is found that most of the components show almost the real values in $^7$B and $^7$He. Hence, the comparison of the real parts of the $S$-factors for $^7$B and $^7$He is shown in Figs. \[fig:sfac0\] and \[fig:sfac2\].
$^6$Be($0^+_1$)-$p$ $^6$Be($2^+_1$)-$p$
----------- --------------------- --------------------- --
$3/2^-_1$ $0.51+i0.02$ $2.35-i0.15$
$3/2^-_2$ $0.02-i0.01$ $0.96-i0.01$
$3/2^-_3$ $0.00+i0.01$ $-0.01-i0.06$
$1/2^- $ $0.93-i0.02$ $0.10-i0.01$
$5/2^- $ $0.00+i0.00$ $1.04-i0.01$
: $S$-factors of the $^6$Be-$p$ components in $^7$B. Details are described in the text.[]{data-label="sf_B7"}
$^6$He($0^+_1$)-$n$ $^6$He($2^+_1$)-$n$
----------- --------------------- --------------------- --
$3/2^-_1$ $0.63+i0.08$ $1.60-i0.49$
$3/2^-_2$ $0.00-i0.01$ $0.97+i0.01$
$3/2^-_3$ $0.01+i0.00$ $0.04-i0.01$
$1/2^- $ $0.95+i0.03$ $0.07-i0.02$
$5/2^- $ $0.00+i0.00$ $1.00+i0.01$
: $S$-factors of the $^6$He-$n$ components in $^7$He. Details are described in the text.[]{data-label="sf_He7"}
![Real part of the $S$-factors of $^7$B and $^7$He, in which the daughter nuclei are the $0^+_1$ states. The experimental data of the $^7$He($3/2^-$) state[@beck07] is shown by the open circle.[]{data-label="fig:sfac0"}](sfac03_A70.eps){width="7.5cm"}
![Real part of the $S$-factors of $^7$B and $^7$He, in which the daughter nuclei are the $2^+_1$ states.[]{data-label="fig:sfac2"}](sfac03_A72.eps){width="7.5cm"}
In Table \[sf\_B7\], for the $3/2^-_1$ state, the $^6$Be($2^+_1$)-$p$ component is large, more than four times of that of the $^6$Be($0^+_1$)-$p$ component for real part. This means that the $^6$Be($2^+_1$) state is dominant in this state. The similar trend can be seen in $^7$He in Table \[sf\_He7\], where the real part of the $^6$He($0^+_1$)-$n$ component agrees with the observation of 0.64(9) [@beck07], as shown in Fig. \[fig:sfac0\]. For the $3/2^-_2$ state, the $^6$Be($2^+_1$)-$p$ component is selectively mixed from the dominant amplitude of $(p_{3/2})^2_{2^+}\otimes(p_{1/2})$. For the $3/2^-_3$ state, the $0^+_1$ and $2^+_1$ states of $^6$Be are hardly included because of the $(p_{3/2})\otimes(p_{1/2})^2$ configuration. Instead of the above two $^6$Be states, the $^6$Be($0^+_2$) state with $(p_{1/2})^2$ configuration and the $^6$Be($2^+_2$) state with $(p_{3/2})(p_{1/2})$ one may give large contributions for this state. For the $1/2^-$ state, the $S$-factor of $^6$Be($0^+_1$)-$p_{1/2}$ proton is close to unity with a small imaginary part and the $^6$Be($2^+_1$)-$p$ component is small. Hence, the $^6$Be($0^+_1$)-$p$ component is dominant in the $1/2^-$ state. The large mixing of the $0^+$ state of $A=6$ nuclei is also confirmed in the $^7$He($1/2^-$) state as shown in Table \[sf\_He7\]. In $^7$He($1/2^-$), we have suggested the weak coupling nature of the $p_{1/2}$ orbital neutron around $^6$He, which retains a two-neutron halo structure[@myo09]. For the $5/2^-$ state of $^7$B, the $^6$Be($2^+_1$)-$p$ component is included well, similar to $3/2^-_2$ as was explained. These two states have a similar structure of the configurations of valence protons. From the $S$-factor analysis, the most of the $^7$B states are not considered to be purely single particle states coupled with the $^6$Be ground state except for the $1/2^-$ state. The component of $^6$Be($2^+_1$) is important in several states. This conclusion is the same as that of $^7$He.
We consider the structure differences between $^7$B and $^7$He from the $S$-factors and discuss the mirror symmetry. From Fig. \[fig:sfac2\], the sizable difference between the components including the $A=6$($2^+$) states is seen in the ground states of $^7$B and $^7$He. The $^6$Be($2^+_1$)-$p$ component in $^7$B obtained as $2.35$ is larger than the $^6$He($2^+_1$)-$n$ component in $^7$He as $1.60$ by 47% for real part. The other four excited states show the similar values between two nuclei in Figs. \[fig:sfac0\] and \[fig:sfac2\]. In those excited states, either of the components of $0^+$ and $2^+$ of $A=6$ nuclei is selectively mixed. These results indicate that the breaking of the mirror symmetry is occurred only in their ground states. The reason of the difference in the $2^+$ coupling is that the $^7$B ground state is located closely to the $^6$Be($2^+_1$) state by 0.45 MeV for resonance energy, as shown in Fig. \[fig:B7\], where the decay widths of two states are rather small in comparison with other resonances. This situation is not occurred in $^7$He as shown in Fig. \[fig:He7\], in which the energy difference between $^7$He($3/2^-_1$) and $^6$He($2^+_1$) is 1.46 MeV. The small energy difference between $^7$B and $^6$Be($2^+_1$) enhances the $^6$Be($2^+_1$)-$p$ component in $^7$B as the coupling to the open channel of the $^6$Be($2^+_1$)+$p$ threshold. On the other hand, the $^6$Be($0^+_1$)-$p$ component in $^7$B becomes smaller than that of $^7$He by 24 % as shown in Fig. \[fig:sfac0\], because the energy difference between the ground states of $^7$B and $^6$Be is 1.97 MeV, larger than the case of $^7$He of 0.40 MeV. The origin of the difference of the $S$-factors in $^7$B and $^7$He is the Coulomb repulsion, which acts to shift the entire energies of the $^7$B states up. The well-known effect of the Coulomb interaction to break the mirror symmetry is the Thomas-Erhman shift, in which the $s$-wave dominant states suffer the different effect of Coulomb repulsion from the states having mainly other partial waves. On the other hand, the present result found in the $^7$B ground state is caused by the existence of the several open channels including the excitations of subsystems and is different from the Thomas-Erhman shift.
As conclusion, the mirror symmetry is broken only in the ground states of $^7$B and $^7$He, while the excited states of two nuclei keep the symmetry. This result is associated with the energies of the $A=6$ subsystem as the open channels of the one nucleon emission. It is experimentally desired to observe the $2^+$ components of $A=6$ nuclei in $^7$B and $^7$He and examine the mirror symmetry. In the present analysis, the $S$-factors represent the contributions of only the resonances of $^7$B and $^6$Be. By considering the additional contributions of the remaining continuum states of two nuclei, it is available to obtain the strength functions of the one-proton removal from $^7$B into $^6$Be and also into the $^4$He+$p$+$p$ final states, which are observable. It is interesting to obtain these strengths and compare them with the one-neutron removal strength from $^7$He into $^6$He [@myo09].
Summary {#sec:summary}
=======
We have investigated the resonance structures of $^7$B with the $^4$He+$3p$ four-body cluster model. The boundary condition for many-body resonances is accurately treated using the complex scaling method. The decay thresholds concerned with subsystems are described consistently. We have found five resonances of $^7$B, which are dominantly described by the $p$-shell configurations. The energy and the decay width of the ground state agree with the recent experiment. We also predict four excited resonances of $^7$B, which are desired to be confirmed experimentally.
We further investigate the spectroscopic factors of the $^6$Be-$p$ components in $^7$B to examine the coupling behavior between $^6$Be and a last proton. It is found that the $^6$Be($2^+_1$) state contributes largely in the ground and the several excited states of $^7$B. In comparison with $^7$He, the mirror nucleus of $^7$B, the $^6$Be($2^+_1$)-$p$ component in the $^7$B ground state is larger than the $^6$He($2^+_1$)-$n$ component in the $^7$He ground state. This difference comes from the fact that the $^7$B ground state is close to the $^6$Be($2^+_1$) state in energy by the Coulomb repulsion. This situation enhances the $^6$Be($2^+_1$)-$p$ component in $^7$B as the channel coupling. The different coupling of $A=6$ nuclei in $^7$B and $^7$He is occurred only in their ground states and indicates the breaking of the mirror symmetry. It is desired to observe the difference of the couplings in $^7$B and $^7$He experimentally.
Acknowledgments {#acknowledgments .unnumbered}
===============
We thank Professor Kiyomi Ikeda for fruitful discussions. This work was supported by a Grant-in-Aid for Young Scientists from the Japan Society for the Promotion of Science (No. 21740194).
References {#references .unnumbered}
==========
\#1\#2\#3\#4[ [[\#1]{}]{} **\#2**, \#4 (\#3)]{}
\[3\]
\[3\]
\[3\]
\[3\]
\[3\]
\[3\]
\[3\]
\[3\]
\[3\]
\[3\]
\[3\]
\[3\]
\[3\] [[*ibid.*]{}]{}
[00]{} I. Tanihata [*et al.*]{}, . R. J. Charity [*et al.*]{}, . L. R. McGrath and J. Cerny, . A. A. Korsheninnikov [*et al.*]{}, . H. G. Bohlen [*et al.*]{}, . M. Meister [*et al.*]{}, . P. Boutachkov [*et al.*]{}, . F. Skaza [*et al.*]{}, . N. Ryezayeva [*et al.*]{}, . F. Beck [*et al.*]{}, . A. H. Wuosmaa [*et al.*]{}, . A. Adahchour and P. Descouvemont, . K. Arai and S. Aoyama, . A. Volya and V. Zelevinsky, . R. I. Betan, A. T. Kruppa, and T. Vertse, . N. Michel, W. Nazarewicz and M. P[ł]{}oszajczak, . Y. Suzuki and K. Ikeda, . H. Masui, K. Katō, K. Ikeda, . T. Myo, K. Katō and K. Ikeda, . T. Myo, R. Ando and K. Katō, . T. Myo, R. Ando and K. Katō, . Y. K. Ho, . N. Moiseyev, . S. Aoyama, T. Myo, K. Katō, K. Ikeda, . T. Myo, K. Katō, S. Aoyama and K. Ikeda, . T. Myo, K. Katō, H. Toki and K. Ikeda, . A. T. Kruppa, R. Suzuki and K. Katō, . Y. Kikuchi, K. Katō, T. Myo, M. Takashina and K. Ikeda . T. Matsumoto, K. Katō, and M. Yahiro, . H. Kanada, T. Kaneko, S. Nagata, M. Nomoto, . Y. C. Tang, M. LeMere and D. R. Thompson, . T. Myo, A. Umeya, H. Toki and K. Ikeda, . T. Myo, K. Katō and K. Ikeda, . J. Aguilar and J.M.Combes, . E. Balslev and J.M. Combes, . W. J. Romo, . M. Homma, T. Myo and K. Katō, . T. Berggren, . R. Suzuki, T. Myo and K. Katō, . N. Michel, W. Nazarewicz, and M. P[ł]{}oszajczak, . T. Myo, S. Aoyama, K. Katō and K. Ikeda, . T. Myo, A. Ohnishi and K. Katō, . F. Ajzenberg-selove, . I. Tanihata [*et al.*]{}, . G. D. Alkhazov [*et al.*]{}, . O. A. Kiselev [*et al.*]{}, Eur. Phys. J. A[**25**]{}, Suppl. 1, 215 (2005). P. Mueller [*et al.*]{}, .
[^1]: [email protected]
[^2]: [email protected]
[^3]: [email protected]
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
Finite version of Random Domino Automaton (FRDA) - recently proposed in [@BiaCzAA] as a toy model of earthquakes - is investigated. Respective set of equations describing stationary state of the FRDA is derived and compared with infinite case. It is shown that for the system of big size, these equations are coincident with RDA equations. We demonstrate a non-existence of exact equations for size $N\geq5$ and propose appropriate approximations, the quality of which is studied in examples obtained within Markov chains framework.
We derive several exact formulas describing properties of the automaton, including time aspects. In particular, a way to achieve a quasi-periodic like behaviour of RDA is presented. Thus, based on the same microscopic rule - which produces exponential and inverse-power like distributions - we extend applicability of the model to quasi-periodic phenomena.
author:
- 'Mariusz Bia[ł]{}ecki'
title: Finite Random Domino Automaton
---
Introduction
============
The Random Domino Automaton, proposed in [@BiaCzAA], is a stochastic cellular automaton with avalanches. It was introduced as a toy model of earthquakes, but can be also regarded as an substantial extension of 1-D forest-fire model proposed by Drossel and Schwabl [@DSfire; @DCS1Dfire; @MMTfire].
The remarkable feature of the RDA is the explicit one-to-one relation between details of the dynamical rules of the automaton (represented by rebound parameters $\mu_i/\nu$ defined in cited article and also below) and the produced stationary distribution $n_i$ of clusters of size $i$, which implies distribution of avalanches. It is already shown how to reconstruct details of the “microscopic” dynamics from the observed “macroscopic” behaviour of the system [@BiaCzAA; @BiaFRDA1to1].
As a field of application of RDA we studied a possibility of constructing the Ito equation from a given time series and - in a broader sense - applicability of Ito equation as a model of natural phenomena. For RDA - which plays a role of a fully controlled stochastic natural phenomenon - the relevant Ito equation can be constructed in two ways: derived directly from equations and by histogram method from generated time series. Then these two results are compared and investigated in [@CzBiaTL; @CzBiaAG].
Note that the set of equations of the RDA in a special limit case reduces to the recurrence, which leads to known integer sequence - the Motzkin numbers, which establishes a new, remarkable link between the combinatorial object and the stochastic cellular automaton [@BiaMotzkin].
In the present paper a finite version of Random Domino Automaton is investigated. The mathematical formulation in finite case is precise and the presented results clarify which formulas are exact and allow to estimate approximations we impose in infinite case presented in [@BiaCzAA]. We also show, that equations of finite RDA can reproduce results of [@BiaCzAA], when size $N$ of the system is increasing and distributions satisfy an additional assumption ($ n_i \rightarrow 0$ for big $i$).
On the other hand, a time evolution of Finite RDA can exhibit a periodic-like behaviour (the assumption $ n_i \rightarrow 0$ for big $i$ is violated), which is a novel property. Thus, based on the same microscopic rules, depending on a choice of parameters of the model, a wide range of properties is possible to obtain. In particular, such behaviour is interesting in the context of recurrence parameters of earthquakes (see e.g. [@WeathRecCA; @Parsons]). For other simple periodic-like models, see [@PachMin; @Pach08].
The finite case makes an opportunity to employ Markov chains techniques to analyse RDA. Investigating the automaton in Markov chains framework we arrive at several novel conclusions, in particular related to expected waiting times for some specified behaviour.
This article completes and substantially extends previous studies of RDA on the level of mathematical structure. We analyse properties of the automaton related to time evolution and others, as a preparation for further prospective comparisons with natural phenomena, including earthquakes. A matter of adjusting the model to the real data is left for the forthcoming paper.
The plan of the article is as follows. Mimicking [@BiaCzAA] in Section \[sec:def\] we define the finite RDA. In Section \[sec:equations\] we derive respective equations for finite RDA. In Section \[sec:cases\] we will specify them for some chosen cases. In Section \[sec:Markov\] we will shortly describe Markov chains setting and describe time aspects of FRDA. Several examples are presented in Section \[sec:examples\]. The last Section \[sec:conclusions\] contain conclusions and remarks. In the Appendix we show non existence of exact equations for RDA as well as present supplementary formulas and Table \[tab:statesN10\] displaying all states of RDA of size $N=10$.
Finite RDA {#sec:def}
==========
The rules for Finite Random Domino Automaton are the same as in [@BiaCzAA]. We assume:\
- space is 1-dimensional and discrete – consists of $N$ cells;\
- periodic boundary conditions (the last cell is adjacent to the first one);\
- cell may be in one of two states: empty or occupied by a single ball;\
- time is discrete and in each time step an incoming ball hits one arbitrarily chosen cell (each cell is equally probable).
The state of the automaton changes according to the following rule:\
$\bullet$ if the chosen cell is empty it becomes occupied with probability $\nu$; with probability $(1-\nu)$ the incoming ball is rebounded and the state remains unchanged;\
$\bullet$ if the chosen cell is occupied, the incoming ball provokes an avalanche with probability $\mu$ (it removes balls from hit cell and from all adjacent cells); with probability $(1-\mu)$ the incoming ball is rebounded and the state remains unchanged.
The parameter $\nu$ is assumed to be a constant but the parameter $\mu$ is allowed to be a function of size of the hit cluster. The way in which the probability of removing a cluster depends on its size strongly influences evolution of the system and leads to various interesting properties, as presented in the following sections. We note in advance that in fact there is only one effective parameter $\mu/\nu$ which affects properties of the automaton. Changing of $\mu$ and $\nu$ proportionally in a sense corresponds to a rescaling of time unit.
A diagram shown below presents an automaton of size $N=12$, with three clusters (of size $1,2$ and $4$) in time $t$. An incoming ball provokes an relaxation of the size [**two**]{}, thus in time $t+1$ there are two clusters (of size $1$ and $4$).
[l c|c|c|c|c|c|c|c|c|c|c|c|c|c]{} & &\
time $= t $ & $\quad \quad \hookrightarrow$ & $ \bullet $ & $ \bullet $ & $ \ $ & $ \ $ & $ \bullet $ & $ \ $ & $\bullet$ & $ {\bullet} $ & $ \ $ & $ \ $ & $ \bullet $ & $ \bullet $ &$ \hookleftarrow \quad \quad $\
\
time $= t + 1 $ & $ \quad \quad \hookrightarrow$ & $\bullet$ & $\bullet$ & $ \ $ & $ \ $ & $ \bullet $ & $ \ $ & $\boldsymbol\downarrow $ & $\boldsymbol\downarrow$ & $ \ $ & $ \ $ & $\bullet$ & $ \bullet $ & $\hookleftarrow \quad \quad$\
& & &\
\
Denote by $n_i, i = 1, \ldots, N$ the number of clusters of length $i$, and by $n_i^0, i = 1, \ldots, N$ the number of empty clusters of length $i$. Due to periodic boundary conditions, the number of clusters is equal to the number of empty clusters in the lattice if two cases are excluded - when the lattice is full (single cluster of size $N$) and when the lattice is empty (single empty cluster of size $N$). Hence for $$n_R = \sum_{i=1}^{N-1} n_i, \quad \text{and} \quad n_R^0= \sum_{i=1}^{N-1} n_i^0
\label{eq:nR}$$ we have $$n_R =n_R^0.
\label{eq:nn0}$$ The density $\rho$ of the system is defined as $$\rho = \frac{1}{N}\sum_{i=1}^N n_i i.
\label{eq:def_rho}$$ In this article we investigate a stationary state of the automaton and hence the variables $n_i, n_R, \rho$ and others are expected values and do not depend of time.
Equations for finite RDA {#sec:equations}
========================
In this section we derive equations describing stationary state of finite RDA. The general idea of the reasoning presented below is: the gain and loss terms balance one another.
Balance of density $\rho$
-------------------------
The density $\rho$ may increase only if an empty cell becomes occupied, and the gain per one time step is $1/N$. It happens with probability $\sim \nu (1-\rho)$. Density losses are realized by avalanches and may be of various size. The effective loss is a product of the size $i$ of the avalanche and probability of its appearance $ \mu_i (n_i i)/N $. Any size $i$ contribute, hence the balance of $\rho$ reads $$\nu(1-\rho)= \frac{1}{N} \sum_{i=1}^N \mu_i n_i i^2.
\label{eq:rho_balance}$$ We emphasise, the above result is exact – no correlations were neglected. Its form is directly analogous to the respective formula in [@BiaCzAA].
Balance of the total number of clusters
----------------------------------------
[**Gain.**]{} A new cluster (can be of size $1$ only) can be created in the interior of empty cluster of size $\geq 3$. $$\hookrightarrow \cdots | \bullet | \underbrace{ \quad \ | \overbrace{\quad | \cdots \cdots| \quad }^{(i-2) \ \text{cells} \ = \ \text{interior}} | \ \quad }_{i} | \bullet | \cdots \hookleftarrow$$ If the empty cluster is of size $N$, then each cell is in interior. Summing up contributions for all empty clusters, the probability is $$\sim \sum_{i=3}^{N-1}\nu \left( \frac{i-2}{i}\right) \frac{n_i^0 i }{N} + \nu n_N^0,
\label{eq:ni_gains_p}$$ which can take a form (for $N\geq3$) $$\sim \nu(1-\rho) -2\nu \frac{n_R}{N} + \nu \frac{n_1^0}{N}.
\label{eq:ni_gains}$$
[**Loss.**]{} Two ways contribute: joining a cluster with another one and removing a cluster due to avalanche.
Joining of two clusters can occur if there exists an empty cluster of length $1$ between them. The exception is when the empty $1$-cluster is the only one empty cluster, and the system consists of a single cluster of length $N-1$. Hence, the probability of joining two clusters is $$\sim \nu \left( \frac{n_1^0}{N} - n_{N-1}\right).
\label{eq:balnjoin}$$ The probability of avalanche is just $$\sim \sum_{i=1}^N \mu_i \frac{n_i i}{N}.
\label{eq:balnav}$$
Gathering these terms one obtains equation for balance of the total number of clusters $n$ $$N(1-\rho) - \sum_{i=1}^N \frac{\mu_i}{\nu} n_i i+n_{N-1}=2n_R.
\label{eq:baln}$$ Again we emphasise that the above result is exact – no correlations were neglected. Finite size of the system reflects in the appearance of $(2n_R-n_{N-1})$ instead of $2n$ in the respective formula in [@BiaCzAA].
Balance of $n_i$s
-----------------
[**Loss.**]{} There are two modes.\
(a) Enlarging - an empty cluster on the edge of an $i$-cluster becomes occupied. There are two such empty clusters except for the case when system contains a single cluster of length $N-1$. Hence, the respective rates are $$\begin{aligned}
\sim & 2 \nu \frac{n_i}{N} \quad \quad \quad & i=1,\ldots,N-2, \\
\sim & \nu \frac{n_{N-1}}{N} \quad \quad & i=N-1. \end{aligned}$$ (b) Relaxation rate for any $i=1,\ldots,N$ is given by $$\sim \mu_i \frac{i n_i}{N}.$$
[**Gain.**]{} Again, there are two modes.\
(a) Enlarging. For $N\geq3$, there are following rates depending on the size $i$ of the cluster $$\begin{aligned}
\sim & \nu(1-\rho) - 2\nu\frac{n_R}{N} + \nu \frac{n_1^0}{N}, \quad & i=1, \\
\sim & 2\nu\frac{n_{i-1}}{N} \alpha^E_{i-1} \quad \quad \quad \quad & 2\leq i \leq N-1, \\
\sim & \nu\frac{n_{N-1}}{N}, \quad \quad \quad \quad \quad \quad \quad \quad & i=N, \label{eq:gecN}\end{aligned}$$ where $\alpha_E(i)$ is a probability that the size of empty cluster adjacent to the $i$-cluster is bigger than $1$. It is clear that $$\alpha^E_{N-2}=1 \quad \text{and} \quad \alpha^E_{N-1}=0.$$ Formula does not have a factor $2$, because there is only one empty cluster (of size $1$).
\(b) Merger of two clusters up to the cluster of size $i$. Two clusters: one of size $k\in \{1,2,\ldots,(i-2)\} $ and the other of size $((i-1)-k)$ will be combined if the ball fills an empty cell between them. $$\hookrightarrow \cdots | \quad | \underbrace{ \overbrace{\bullet \ | \cdots | \ \bullet }^k | \quad |
\overbrace{ \bullet \ | \bullet | \cdots | \ \bullet}^{(i-1-k)} }_{i} | \quad | \cdots \hookleftarrow$$ The probability is proportional to the number of empty $1$-clusters between $k$-cluster and $(i-1-k)$-cluster, $$\sim \nu\frac{n_{1}^0}{N} \gamma^E_{i} \quad \quad \quad 3 \leq i \leq N-1,$$ where $\gamma^E_{i}$ is a probability of such merger. For $i=N$ there is a single cluster in the lattice (there are no two clusters to merge) - filling the gap between ends of $(N-1)$-cluster is already considered in (a).
Gathering the terms, one obtains $$\begin{aligned}
n_1 &=&\frac{1}{\frac{\mu_1}{\nu}+2} \left( N(1-\rho) -2n_R + n_1^0 \right), \label{eq:baln1} \\
n_2 &=& \frac{1}{2\frac{\mu_2}{\nu}+2} 2n_1 \alpha^E_{1}, \label{eq:baln2}\\
n_i &=& \frac{1}{\frac{\mu_i}{\nu}i+2} \left( 2n_{i-1} \alpha^E_{i-1} + n_1^0 \gamma^E_{i} \right), \label{eq:balni}\\
n_{N-1} &=& \frac{1}{\frac{\mu_{N-1}}{\nu}(N-1)+1} \left( 2n_{N-2} + n_1^0 \gamma^E_{N-1} \right), \label{eq:balnN-1}\\
n_N &=& \frac{1}{\frac{\mu_N}{\nu}N} n_{N-1}, \label{eq:balnN}\end{aligned}$$ where $ \quad 3\leq i\leq(N-2)$.
The last equation has simple explanation. The state with all cells being occupied (corresponding to $n_N$) can be achieved only from the state with a single empty cell (corresponding to $n_{N-1}$) with probability $\nu (1/N) $. On the other hand, the automaton leaves the state with all cells being occupied with probability $\mu_N$.
Note that equations and are exact. Correlations in the systems reflect in appearing of multipliers $\alpha^E_{i}$ and $\gamma^E_{i}$. Their values depends on possible configurations of states of the automaton. As shown in the Appendix, for $N\geq 5$ exact formulas for $\alpha^E_{i}$ and $\gamma^E_{i}$ as functions of $n_i$s do not exist. Hence, it is necessary to propose approximated formulas.
A mean field type approximation for $\alpha^E_{i}$ is $$\alpha^E_{i} \approx \alpha^A_{i} = \left( 1- \frac{n_1^0}{\sum_{k=1}^{N-i}n_k^0} \right).
\label{eq:alphaEapprox}$$ For a given cluster of size $i$, the probability of appearance of an empty cluster of size $1$ is calculated as proportional to the number of empty $1$-clusters divided by the sum of the numbers of all empty clusters with size not exceeding $N-1$, because there is no room for larger.
When merger of two clusters up to a cluster of size $i$ is considered, the room denoted by $A$ is of size $(N-2-(i-1-k))$ and the room denoted by $B$ is of size $(N-2-k)$ - see a diagram below. $$\hookrightarrow \overbrace{ \underbrace{\cdots | }_{B} \quad | \underbrace{\bullet \ | \cdots | \ \bullet }_k}^A | \quad |
\underbrace{ \overbrace{ \bullet \ | \bullet | \cdots | \ \bullet}^{(i-1-k)} | \quad \overbrace{| \cdots }^A }_B \hookleftarrow$$ Hence a mean field type approximation for $\gamma^E_{i}$ is of the form $$\gamma^E_{i} \approx \gamma^A_{i} = \sum_{k=1}^{i-2} \left( \frac{n_k} {\sum_{j=1}^{N-(i-1-k+2)}n_j}
\cdot \frac{n_{i-1-k}}{\sum_{j=1}^{N-(k+2)}n_j} \right).
\label{eq:gammaEapprox}$$ It is also instructive to consider another approximation $$\gamma^E_{i} \approx \gamma^{AR}_{i} = \sum_{k=1}^{i-2} \left( \frac{n_k}{n_R} \cdot \frac{n_{i-1-k}}{n_R} \right).$$ Section \[sec:examples\] contains quantitative estimation of proposed approximations. Comparison of this approximation with exact results for small sizes $N$ is discussed in Section \[sec:conclusions\].
Thermodynamic limit
-------------------
In the paper [@BiaCzAA] an assumption of independence of clusters was considered. To have it adequate, it is required that there are no limitations in space, like those encountered when formulas and were considered. For systems that are big enough, i.e., when $N \longrightarrow \infty$, an empty cluster adjacent to a given $i$-cluster can be of any size, and thus $$\alpha^E_{i} \approx \alpha = \left( 1- \frac{n_1^0}{\sum_{k=1}^{\infty}n_k^0} \right) = \left( 1- \frac{n_1^0}{n} \right).$$ This is consistent with the requirement that $n_i \longrightarrow 0$ when $i \longrightarrow \infty$, which is required to have moments of the $n_i$s convergent. Similarly, $$\gamma^E_{i} \approx \gamma(i) = \sum_{k=1}^{i-2} \left( \frac{n_k}{n} \cdot \frac{n_{i-1-k}}{n} \right).$$ These formulas substituted into - give the respective set of equations considered in [@BiaCzAA]. The same reasoning can be applied to balance equations. The form of equation is left unchanged under the limit. For equation , $(2n_R-n_{N-1}) \longrightarrow 2n$, and it becomes of the form presented in [@BiaCzAA].
Special cases {#sec:cases}
=============
For fixed form of rebound parameters equations describing the automaton can be written in more specific form. This is the case for balance equations and , as well as for formulas for average cluster size $$\left\langle i \right\rangle = \frac{\sum_{i=1}^N n_i i }{\sum_{i=1}^N n_i} = \frac{N\rho}{n_R+n_N}
\label{eq:avcluster}$$ and average avalanche size $$\left\langle w \right\rangle = \frac{\sum_{i=1}^N \mu_i n_i i^2 }{\sum_{i=1}^N \mu_i n_i i}.
\label{eq:avavlanche}$$ We emphasize, these formulas are exact – correlations are encountered. We consider three special cases investigated in detail and illustrated by examples below.
$\mu=const.$
------------
For $\mu=const.$ and $\nu=const.$ equation is of the form $$(1-\rho)= \frac{1}{N} \frac{\mu}{\nu}\sum_{i=1}^N n_i i^2.
\label{eq:rho_balanceA}$$ and equation $$N(1-\rho(1+\frac{\mu}{\nu}))+n_{N-1}=2n_R.
\label{eq:balnA}$$ Also formulas for $\left\langle i \right\rangle$ and $\left\langle w \right\rangle$ are simplified only a little.
$\mu(i)=\delta/i$ where $\theta=\delta/\nu=const.$
--------------------------------------------------
Equation is of the form $$(1-\rho)= \theta \rho,
\label{eq:rho_balanceB}$$ hence the density is given by remarkably neat (end exact) formula $$\rho=\frac{1}{1+\theta}.
\label{eq:densityB}$$ Note that there is no dependence on the size of the system $N$; for $N \longrightarrow \infty$ it remains the same.
Equation can be written as $$N\frac{\theta}{1+\theta} = (2+\theta)n_R,
\label{eq:balnB}$$ where we use equations and . Hence the formula for $n_R$ is of the form $$n_R=N\frac{\theta}{(\theta+1)(\theta+2)}
\label{eq:nRB}$$ in direct analogy with $n$ in $N \longrightarrow \infty$ case [@BiaCzAA]. Thus, $n_R$ plays the role of $n$, as indicated also in balance of $n_1$ equation . The formula for $n$ is $$n=n_R+n_N= N\frac{\theta (1+\varepsilon)}{(\theta+1)(\theta+2)} \quad \text{where} \quad \varepsilon=\frac{n_N}{n_R}.
\label{eq:nB}$$
The average cluster size is given by $$\left\langle i \right\rangle = \frac{1}{1+\varepsilon}\left( 1+\frac{2}{\theta} \right).
\label{eq:avclusterB}$$ The average avalanche size is equal to the average cluster size $$\left\langle w \right\rangle = \left\langle i \right\rangle,
\label{eq:avavlancheB}$$ because each cluster has the same probability to be removed from the lattice.
The above formulas are exact (include correlations) and have good thermodynamic limit ($\varepsilon \longrightarrow 0$). Note also that variables $\rho$ and $n_R$ depend on single parameter $\theta$. Formulas with dependence on $\theta$ can be rewritten as functions of density $\rho$.
$\mu(i)=\eta/{i^2}$ and $\chi=\sigma/\nu=const.$
------------------------------------------------
Equation is of the form $$N(1-\rho)= \chi (n_R+n_N).
\label{eq:rho_balanceC}$$
Equation can be written as $$N(1-\rho)=2n_R-\chi\frac{1}{N}n_N+\chi\sum_{i=1}^N \frac{n_i}{i}
\label{eq:balnC}$$ where equation is used, namely $n_{N-1}=\chi\frac{1}{N}n_N$.
The average cluster size $$\left\langle i \right\rangle = \chi \frac{\rho}{1-\rho},
\label{eq:avclusterC}$$ and the average avalanche size $$\left\langle w \right\rangle = \frac{1}{\left( 1+ \frac{\sigma}{N} \right) - (1-\sigma)\frac{2}{\chi}},
\label{eq:avavlancheC}$$ where $$\sigma=\frac{n_N}{n_R+n_N}=\frac{\varepsilon}{\varepsilon+1}.$$ Note that also these formulas are exact.
Finite RDA as a Markov chain {#sec:Markov}
============================
General settings
----------------
state number example[^1] multiplicity contrib. to
-------------- -------------------------------------------------------------------- -------------- ----------------
1 $ \hookrightarrow | \ \ \ | \ \ \ | \ \ \ | \hookleftarrow $ 1 $ n_3^0 $
2 $ \hookrightarrow | \ \ \ | \ \ \ | \bullet | \hookleftarrow $ 3 $ n_1, n_2^0 $
3 $ \hookrightarrow | \ \ \ | \bullet | \bullet | \hookleftarrow $ 3 $ n_2, n_1^0 $
4 $ \hookrightarrow | \bullet | \bullet | \bullet | \hookleftarrow $ 1 $ n_3 $
: States for the size of the lattice $N=3$.[]{data-label="tab:N3"}
Finite Random Domino Automaton is a Markov chain, hence we use standard knowledge to solve several examples for small $N$ and derive a number of formulas for time aspects of the evolution of the system.
In general, for the lattice of size $N$ there are $2^N$ states, because each of $N$ cells may be empty or occupied. For $N=4$, an exemplary state is $$\hookrightarrow | \ \ \ | \bullet | \ \ \ | \bullet | \hookleftarrow$$ where assumed periodic boundary conditions are depicted by hook-arrows.
For periodic boundary conditions it is irrelevant to distinguish between states which differ by a translation only. Hence, in example, we consider the following states equivalent: $$\hookrightarrow | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \hookleftarrow
\quad \quad \equiv \quad \quad
\hookrightarrow | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \hookleftarrow$$ Thus states $a_i$ are defined up to translational equivalence (see Tables \[tab:N3\] and \[tab:N5\]). The label numbers are assigned to the states, as shown in tables - no exact rule is applied.
state number example[^2] multiplicity contrib. to
-------------- ---------------------------------------------------------------------------------------- -------------- -----------------------
1 $ \hookrightarrow | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \hookleftarrow $ 1 $ n_5^0 $
2 $ \hookrightarrow | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \hookleftarrow $ 5 $ n_1, n_4^0 $
3 $ \hookrightarrow | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \hookleftarrow $ 5 $ n_2, n_3^0 $
4 $ \hookrightarrow | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \hookleftarrow $ 5 $ n_1, n_1^0, n_2^0 $
5 $ \hookrightarrow | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \hookleftarrow $ 5 $ n_2, n_2^0 $
6 $ \hookrightarrow | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \hookleftarrow $ 5 $ n_1, n_2, n_1^0 $
7 $ \hookrightarrow | \ \ \ | \bullet | \bullet | \bullet | \bullet | \hookleftarrow $ 5 $ n_4, n_1^0 $
8 $ \hookrightarrow | \bullet | \bullet | \bullet | \bullet | \bullet | \hookleftarrow $ 1 $ n_5 $
: States for the size of the lattice $N=5$.[]{data-label="tab:N5"}
Further reduction of the number of states using reflections can be done, but it is not very efficient procedure. We do not perform it, keeping symmetrical states separate. They deliver a simple computation check - their probabilities are necessarily equal.
Such space of states for the finite random domino automaton is irreducible, aperiodic and recurrent. Transition matrix $\mathbf{P}$ is defined by $$[ \mathbf{P} ]_{ij} = \text{probability of transition} \quad a_i \longrightarrow a_j
\label{eq:transMat}$$
For $N=3$ the transition matrix is of the form $$\mathbf{P} =
\left( \begin{array}{cccc}
1-\nu & \nu & 0 & 0 \\
\frac{\mu_1}{3} & 1- \frac{\mu_1}{3}-\frac{2\nu}{3} & \frac{2\nu}{3} & 0 \\
\frac{2\mu_2}{3} & 0 & 1- \frac{2\mu_2}{3}-\frac{\nu}{3} & \frac{\nu}{3} \\
\mu_3 & 0 & 0 & 1-\mu_3
\end{array} \right)
\label{eq:transM3}$$ where entries are found from analysis of transition probability of all possible states $a_i$ (see Tab.\[tab:N3\]).
For $N=5$ the transition matrix is
$$\mathbf{P} = \frac{1}{5}
\left( \begin{array}{cccccccc}
5-5\nu & 5\nu& 0& 0& 0& 0& 0& 0 \\
\mu_1 & 5-\mu_1-4\nu & 2\nu & 2\nu& 0& 0& 0& 0\\
2\mu_2 & 0& 5-2\mu_2-3\nu& 0& 2\nu& \nu& 0& 0\\
0 & 2\mu_1& 0& 5-2\mu_1-3\nu& \nu& 2\nu& 0& 0\\
3\mu_3& 0& 0& 0& 5-3\mu_3-2\nu& 0& 2\nu& 0\\
0 & 2\mu_2& \mu_1& 0& 0& 5-2\mu_2-\mu_1-2\nu& 2\nu& 0\\
4\mu_4& 0& 0& 0& 0& 0& 5-4\mu_4-\nu& \nu\\
5\mu_5& 0& 0& 0& 0& 0& 0& 5-5\mu_5
\end{array} \right)
\label{eq:transM5}$$
Stationary distribution is given by $$v \cdot P=v.
\label{eq:MP}$$
The number of states increase rapidly with $N$: for $N=6$ there are $14$ states, for $N=7$ there are $20$ states and for $N=10$ there are $108$ states. The number of states for any $N$ is bigger than $2^N/N$, because translational symmetry of states is at most $N$, but always there are states with smaller symmetry, like empty state and fully occupied state. Thus practical usage of Markov chain settings for calculations is rather limited. This is one of the reasons for developing more “handy” framework, like presented in [@BiaCzAA] and here. On the other hand, Markov chains can be used for illustrations and justifications of some properties, as presented below.
Expected time of return
-----------------------
As system evolves, it hits a given state many times. Here we consider expected value of the time of return from state with density $\rho=0$ to itself and next from the state with $\rho=1$ to itself.
Starting from state $1$ (state with $\rho=0$) the next state (different from state $1$) contains a single $1$-cluster only. This state - denoted by label $2$ - has density $\rho=1/N$. Expected time for this change is $1/\nu$.
Let $\tau_i$ be the expected time to hit state $1$ starting in state $i$. Then $\tau_1=0$ and for $i\neq 1$ $$\begin{aligned}
\tau_i &=&\mathbb{E}( \text{time to hit} \ 1 \ | \ \text{start in} \ i ) \nonumber \\
&=& 1 + \sum_{k} p_{ik} \mathbb{E}(1 | k) = 1 + \sum_k p_{ik}\tau_k,
\label{eq:rettime}\end{aligned}$$ where $\mathbb{E}(1|k) =\mathbb{E}( \text{time to hit} \ 1 \ | \ \text{start in} \ k )$. After solving this system of equations, the return time is $$t_{1\rightarrow 1}= 1/\nu + \tau_2.
\label{eq:rettime11}$$
Similarly, for state with $\rho=1$ (state $L$) the next state (different from state $L$) is the empty state (with $\rho=0$) and $$t_{L \rightarrow L}= 1/\mu_N + \hat\tau_1,
\label{eq:rettimeL}$$ where $\hat\tau_1$ is the expected time to hit state $L$ starting in state $1$. The respective equation to determine $\hat\tau_i$ for $i\neq L$ reads $$\hat\tau_i = 1 + \sum_k p_{ik} \hat\tau_k,
\label{eq:hatrettime}$$ and obviously $\hat\tau_L=0$.
Note that the expected time $t_{L \rightarrow L}$ is equal to expected time of return from state $1$ to state $1$ through state $L$: $$t_{L \rightarrow L}=t_{1\rightarrow L \rightarrow 1}.
\label{eq:timell1l1}$$
probability of value
-------------------------- -------------------------------------------
rebound – occupied cell $\frac{1}{N}\sum_{i=1}^N n_i i (1-\mu_i)$
rebound – empty cell $(1-\rho)(1-\nu) $
occupation of empty cell $(1-\rho)\nu$
trigerring an avalanche $\frac{1}{N}\sum_{i=1}^N \mu_i n_i i$
: Probabilities of all four possibilities occurring in a single time step during evolution of the automaton.[]{data-label="tab:probrebocc"}
The expected time between two consecutive avalanches is $$t_{av}= \frac{\left\langle w \right\rangle + 1}{1-P_r},
\label{eq:timeavalanche}$$ where $P_r$ is the probability that the incoming ball is rebounded both form empty or occupied cell: $$P_r = (1-\rho)(1-\nu) + \frac{1}{N}\sum_{i=1}^N n_i i (1-\mu_i).
\label{eq:Pr}$$ Note that $(1-P_r)$ is equal to the sum of probability of triggering an avalanche and probability that an empty cell becomes occupied, hence $$P_r + \frac{1}{N}\sum_{i=1}^N \mu_i n_i i + (1-\rho)\nu=1.
\label{eq:Pz}$$ Formula can be derived as follows. In time between two consecutive avalanches, on average, $(t_{av} (1-P_r)-1)$ cells become occupied in the system – it receives one ball per a time step, part of them are rebounded and one ball triggers the avalanche. An avalanche is reducing the number of occupied cells by $\left\langle w \right\rangle$. These two quantities compensate each other, giving .
On the other hand, the expected time between two consecutive avalanches is equal to the inverse of the probability of triggering an avalanche $$t_{av} = \left( \frac{1}{N}\sum_{i=1}^N \mu_i n_i i \right)^{-1}.
\label{eq:tavav}$$ Both expressions given in and are equal to each other.
Frequency distribution of avalanches
------------------------------------
The probability of states obtained from condition allows to determine the distribution of frequency of avalanches. The frequency $f_i$ of the avalanche of size $i$ is given by the sum of products of probabilities $v_k$ of state $k$ and respective transition probability $p_{kj}$ to the appropriate states $j$ for all states that transition $k \longrightarrow j$ produce the avalanche of size $i$.
For example, for $N=5$, as can be seen in Table \[tab:N5\], transitions $2 \longrightarrow 1$, $4 \longrightarrow 2$ and $6 \longrightarrow 3$ result in an avalanche of size $1$, transitions $3 \longrightarrow 1$ and $6 \longrightarrow 2$ give an avalanche of size $2$, transition $5 \longrightarrow 1$ of size $3$, $7 \longrightarrow 1$ of size $4$ and $8 \longrightarrow 1$ of size $5$. Hence $$\begin{aligned}
f_1&=& v_2 \mu_1/5+v_4 2\mu_1/5+ v_6 \mu_1/5, \\
f_2&=& v_3 2 \mu_2/5+v_6 2 \mu_2/5, \\
f_3&=& v_5 3 \mu_3/5, \\
f_4&=& v_7 4 \mu_4/5, \\
f_5&=& v_8 \mu_5,
\label{eq:avalanchefreq}\end{aligned}$$ where respective $p_{kj}$ are taken from transition matrix .
The average time $t_i$ between two avalanches of size $i$ is given by $$t_i = 1/f_i,
\label{eq:avallanchetime}$$ in particular, for a maximum size $N$ $$t_{L \rightarrow L}=t_N.
\label{eq:tlltn}$$
The average time between (any) consecutive avalanches given by formula may be also calculated as $$t_{av}= \left( \sum_{i=1}^N t_i^{-1}\right)^{-1},
\label{eq:tavsumti}$$ because the probability of avalanche of any size is just a sum of probabilities of all possible avalanches. In this way one can calculate also average time between any two consecutive avalanches of prescribed size - for example, size $4$ and $5$ (or any other subset of possible sizes).
Examples {#sec:examples}
========
Below we present several examples to illustrate properties of finite RDA as well as to demonstrate application of the schemes outlined above.
$N=3$
-----
This is the simplest non-trivial, worm-up example. For $N=3$ the general results – i.e., for arbitrary $\mu_1$, $\mu_2$, $\mu_3$ and $\nu$ – can be calculated explicitly. Usage of equations - leads to exact results as presented below (see Appendix). The same can be also obtained from Markov chains framework. Equations , , and give $$\begin{aligned}
n_1 &=& {3\left(\frac{\mu_2}{\nu} + \frac{1}{2}\right)}/D, \\
n_2 &=& {3}/D, \\
n_3 &=& \left({\frac{\nu}{\mu_3}}\right)/D,
\label{eq:exactsolN3}\end{aligned}$$ where $$D = \frac{11}{2} + \frac{\mu_1}{2\nu} + 5\frac{\mu_2}{\nu} + \frac{\mu_1 \mu_2}{\nu^2} + \frac{\nu}{\mu_3}.$$ From inspecting of Table \[tab:N3\] it is evident that $n_1^0=n_2$, $n_2^0=n_1$ and $ n_3^0 = 1- n_1 -n_2 -n_3$ (all posibilities sum up to 1), hence $$n_3^0 = \left({1 + \frac{\mu_1}{2\nu} + 2\frac{\mu_2}{\nu} + \frac{\mu_1\mu_2}{\nu^2}}\right)/D.$$
General formulas for expected times of return are $$\begin{aligned}
t_{1 \rightarrow 1} &=&
\frac{1}{\nu} \left( 1+ \frac{2\nu^2+9\mu_3\nu+6\mu_2\mu_3}{\mu_3(\mu_1+2\nu)(2\mu_2+\nu)} \right), \label{eq:n3ret1} \\
t_{L \rightarrow L} &=&
\frac{1}{\nu} \left( \frac{\nu}{\mu_3} + \frac{11}{2} +\frac{\mu_1}{2\nu} + 5\frac{\mu_2}{\nu}+ \frac{\mu_1 \mu_2}{\nu^2} \right).
\label{eq:n3retL}\end{aligned}$$ The ratio $t_{L \rightarrow L}/t_{1 \rightarrow 1}$ is $$t_{L \rightarrow L}/t_{1 \rightarrow 1}= \frac{1}{2} \left(\frac{\mu_1}{\nu}+ 2\right)\left( \frac{2\mu_2}{\nu} + 1\right).
\label{eq:n3ttratio}$$ Note that it does not depend on $\mu_3$. If the probability of triggering an avalanche of size $1$ and $2$ is small comparing to the probability of occupation of an empty cell (i.e., $\mu_1/\nu \approx 0$ and $\mu_2/\nu \approx 0$) then $t_{L \rightarrow L} \approx t_{1 \rightarrow 1}$. The next stage after the lattice is fully occupied is the empty state; hence, if these two average waiting times are comparable, then they occur with comparable frequency. That means quasi-periodic like behaviour of the system: within average time $11/2\nu$ the lattice become fully occupied, then the triggering of an avalanche of maximal size $N$ occurs with average waiting time $1/\mu_3$. The same can be observed for bigger sizes $N$.
![Plot of the $Log_{10}$ of $n_i$s (left) and $n_i^0$s (right) vs. $i$ for $N=3$ in three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). Rebound parameters are chosen to have density $\rho=1/2$ in all cases (see main text for respective values).[]{data-label="fig:Fig1"}](n_i_prob3 "fig:"){width="4cm"} ![Plot of the $Log_{10}$ of $n_i$s (left) and $n_i^0$s (right) vs. $i$ for $N=3$ in three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). Rebound parameters are chosen to have density $\rho=1/2$ in all cases (see main text for respective values).[]{data-label="fig:Fig1"}](n_i0_prob3 "fig:"){width="4cm"}
${\mu_i}=const.$ $\mu_i=\delta/i$ $\mu_i=\sigma/i^2$
-------------------------------- ------------------ ------------------ --------------------
$\left\langle i \right\rangle$ $1.9281668$ $2$ $2.1134407$
$\left\langle w \right\rangle$ $2.2516538$ $2$ $1.7226121$
: Average cluster size $\left\langle i \right\rangle$ and average avalanche size $\left\langle w \right\rangle$ for three different rebound parameters. Density $\rho=1/2$, the size of the lattice $N=3$.[]{data-label="tab:N3av"}
Figure \[fig:Fig1\] and Table \[tab:N3av\] present examples of three types of dependence of rebound parameters on size $i$ of clusters considered in Section \[sec:cases\], each having the same density $\rho=1/2$ (with 8 digits accuracy). To obtain this density we put for these three cases $\mu/\nu=0.444118$ ($\mu=0.444118$, $\nu=1$), $\theta=1$ ($\delta=1$, $\nu=1$) and $\chi=2.113440690$ ($\eta=1$, $\nu=1/2.113440690$) respectively. As seen from Figure \[fig:Fig1\] it is possible to obtain flat distribution for $\mu_i=\delta/i$ – on that background, differences between the cases are clearly visible: $\mu_i=const.$ discriminate the existence of big clusters fostering big avalanches; the opposite is for $\mu_i=\sigma/i^2$. Average cluster size and avalanche size data presented in Table \[tab:N3av\] confirm this conclusion.
![Plot of the $Log_{10}$ of $n_i$s (left) and $n_i^0$s (right) versus $i$ for $N=5$ in three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). Rebound parameters are chosen to have density $\rho=1/4$ in all cases (see main text for respective values).[]{data-label="fig:Fig2a"}](n_i_prob5 "fig:"){width="4cm"} ![Plot of the $Log_{10}$ of $n_i$s (left) and $n_i^0$s (right) versus $i$ for $N=5$ in three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). Rebound parameters are chosen to have density $\rho=1/4$ in all cases (see main text for respective values).[]{data-label="fig:Fig2a"}](n_i0_prob5 "fig:"){width="4cm"}
${\mu_i}=const.$ $\mu_i=\delta/i$ $\mu_i=\sigma/i^2$
-------------------------------- ------------------ ------------------ --------------------
$\left\langle i \right\rangle$ $1.427017126$ $1.632218845$ $1.985611461$
$\left\langle w \right\rangle$ $1.845355789$ $1.632218845$ $1.41360643$
: Average cluster size $\left\langle i \right\rangle$ and average avalanche size $\left\langle w \right\rangle$ for three different rebound parameters. Density $\rho=1/4$, the size of the lattice $N=5$.[]{data-label="tab:N5av"}
![Ratio of return times ${t_{L \rightarrow L}}/{t_{1 \rightarrow 1}}$ for N=5 (left) and N=7 (right) for three cases: $\mu_i=const.$ (top), $\mu_i=\delta/i$ (middle) and $\mu_i=\eta/i^2$ (bottom). Parameter $t$ is equal to $\mu/\nu$, $\delta/\nu$ and $\eta/\nu$ respectively.[]{data-label="fig:FigRet5"}](N5retTs "fig:"){width="4cm"} ![Ratio of return times ${t_{L \rightarrow L}}/{t_{1 \rightarrow 1}}$ for N=5 (left) and N=7 (right) for three cases: $\mu_i=const.$ (top), $\mu_i=\delta/i$ (middle) and $\mu_i=\eta/i^2$ (bottom). Parameter $t$ is equal to $\mu/\nu$, $\delta/\nu$ and $\eta/\nu$ respectively.[]{data-label="fig:FigRet5"}](N7retTs "fig:"){width="4cm"}
${\mu_i}=const.$ $\mu_i=\delta/i$ $\mu_i=\sigma/i^2$
------------------------------------------------- ------------------ ------------------ --------------------
$R={t_{L \rightarrow L}}/{t_{1 \rightarrow 1}}$ $\approx52.212$ $\approx35.441$ $\approx34.801$
: Coefficient $R={t_{L \rightarrow L}}/{t_{1 \rightarrow 1}}$ for three different rebound parameters (see main text for details). Density for all cases $\rho=1/4$, the size of the lattice $N=5$.[]{data-label="tab:N5tavR"}
$N=5$
-----
For $N=5$ it is impossible to write down exact equations - depending on values of $n_i$s only – see Appendix for details. The case can be solved as a Markov process, but obtained general formulas are relatively complicated.
In this example we investigate properties of the system with density $\rho=1/4$. Figure \[fig:Fig2a\] and Table \[tab:N5av\] compare results in three cases: $\mu/\nu =16257/10000$ the density $\rho=0.2500003184$; for $\theta=3$ the density $\rho=0.25$ exactly; and $\chi=5.95682$ gives the density $\rho=0.2500004527$.
General expressions for return times ${t_{1 \rightarrow 1}}$ and ${t_{L \rightarrow L}}$ as well as their ratio (presented in Appendix) are relatively complex. Note that the return times – except of the dependence on $t$ – are proportional to $1/\nu$. Below we specify the ratio ${t_{L \rightarrow L}}/{t_{1 \rightarrow 1}}$ in three cases: for [$\mu_i=const$]{}, where $t=\mu/\nu$, it is equal to $$\frac{24 t^6 + 154 t^5 + 413 t^4 + 586 t^3 + 467 t^2 + 182 t + 24}{24 t^2 + 54 t + 24},
\label{eq:n5ttratioc}$$ for [$\mu_i=\delta/i$, where $\delta=const$]{} and $t=\delta/\nu$, it is equal to $$\frac{4 t^6 + 40 t^5 + 169 t^4 + 395 t^3 + 550 t^2 + 432 t + 144}{56 t^2 + 192 t + 144},
\label{eq:n5ttratioci}$$ and for [$\mu_i=\sigma/i^2$, where $\sigma=const$]{} and $t=\sigma/\nu$, is $$\frac{2 t^6 + 39 t^5 + 304 t^4 + 1232 t^3 + 2840 t^2 + 3744 t + 2304}{496 t^2 + 2208 t + 2304}.
\label{eq:n5ttratiocii}$$ In each case the ratio is a rational function of $t$, which is equal to $1$ for $t=0$ and asymptotically $\sim t^4$ for $t\longrightarrow \infty$. A generalisation of this observation is a Conjecture formulated in Section \[sec:conclusions\]. A comparison of these ratios is presented in left part of Figure \[fig:FigRet5\]. Table \[tab:N5tavR\] shows that for the cases discussed above with average density $\rho=1/4$ the highest value of $R$ is for $\mu_i=const.$ and the smallest for $\mu_i=\sigma/i^2$ (not much different from the value for $\mu_i=\delta/i$).
Average waiting times $t_i$ for avalanche of size $i$ can be also found. For example for $\mu_i=\delta/i$, where $\delta=const.$, they are presented in the Appendix (equations -). The average time between any two consecutive avalanches is $$t_{av}= \frac{4 t^5 + 48 t^4 + 237 t^3 + 603 t^2 + 762t + 360}{\nu t (4 t^4 + 36 t^3 + 121 t^2 + 168 t + 72)},
\label{eq:tav5ex}$$ where $t=\delta/\nu$. All these quantities are proportional to $1/\nu$. Figure \[fig:Figtimes\] in the left panel presents waiting times $t_i$ in for fixed density $\rho=1/4$ in three cases mentioned above. There are no big differences both in character of dependence of $t_i$ on $i$ and also values of $t_{av}$ do not differ much: for ${\mu_i}=const.$ average time is $t_{av}\approx 24.60$, for $\mu_i=\delta/i$ it is $\approx21.76$ and for $\mu_i=\sigma/i^2$ it is $\approx18.85$. (Choosing parameters to have density $\rho=1/4$ we put $\nu=1/10$ for all cases.)
Average waiting times $t_i, i=1,\ldots,5$ in the case $\mu_i=\delta/i$ for various densities are shown in the right panel of Figure \[fig:Figtimes\]. For small densities the maximal waiting time $t_i$ is for $i=5$, while for bigger densities the maximum is for $i=3$. Average waiting times range from $\approx13.57$ for $\rho={1}/{10}$ through $\approx21.76$, $\approx50.22$, $\approx145.01$ for densities ${1}/{4}$, $ {1}/{2}$, $ {3}/{4}$ respectively, up to $\approx441.60$ for density $\rho={9}/{10}$. (Again $\nu=1/10$ for all cases.)
![ [**Left.**]{} Plot of $Log_{10}$ of $t_i$s vs. $i$ for three rebound parameters for fixed density $\rho=1/4$ for $N=5$. Three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). Rebound parameters are chosen to have density $\rho=1/4$ in all cases (see main text for respective values). [**Right.**]{} Plot of $Log_{10}$ of $t_i$s versus $i$ for various densities for rebound parameter of the form $\mu_i=\delta/i$ for $N=5$. Densities are chosen as $\frac{1}{10}, \frac{1}{4}, \frac{1}{2}, \frac{3}{4}, \frac{9}{10}$; thinner line corresponds to smaller density. []{data-label="fig:Figtimes"}](N5_t_r "fig:"){width="4cm"} ![ [**Left.**]{} Plot of $Log_{10}$ of $t_i$s vs. $i$ for three rebound parameters for fixed density $\rho=1/4$ for $N=5$. Three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). Rebound parameters are chosen to have density $\rho=1/4$ in all cases (see main text for respective values). [**Right.**]{} Plot of $Log_{10}$ of $t_i$s versus $i$ for various densities for rebound parameter of the form $\mu_i=\delta/i$ for $N=5$. Densities are chosen as $\frac{1}{10}, \frac{1}{4}, \frac{1}{2}, \frac{3}{4}, \frac{9}{10}$; thinner line corresponds to smaller density. []{data-label="fig:Figtimes"}](N5_t_d "fig:"){width="4cm"}
$N=7$
-----
![Plot of the $Log_{10}$ of $n_i$s (left) and $n_i^0$s (right) versus $i$ for $N=7$ in three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). Rebound parameters are chosen to have density $\rho=3/4$ in all cases (see main text for respective values).[]{data-label="fig:Fig3a"}](n_i_prob7 "fig:"){width="4cm"} ![Plot of the $Log_{10}$ of $n_i$s (left) and $n_i^0$s (right) versus $i$ for $N=7$ in three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). Rebound parameters are chosen to have density $\rho=3/4$ in all cases (see main text for respective values).[]{data-label="fig:Fig3a"}](n_i0_prob7 "fig:"){width="4cm"}
${\mu_i}=const.$ $\mu_i=\delta/i$ $\mu_i=\sigma/i^2$
-------------------------------- ------------------ ------------------ --------------------
$\left\langle i \right\rangle$ $4.274328495$ $4.385371765$ $4.736665115$
$\left\langle w \right\rangle$ $5.767461682$ $4.385371765$ $2.671314107$
: Average cluster size $\left\langle i \right\rangle$ and average avalanche size $\left\langle w \right\rangle$ for three different rebound parameters. Density $\rho=3/4$, the size of the lattice $N=7$.[]{data-label="tab:N7av"}
${\mu_i}=const.$ $\mu_i=\delta/i$ $\mu_i=\sigma/i^2$
------------------------------------------------- ------------------ ------------------ --------------------
$R={t_{L \rightarrow L}}/{t_{1 \rightarrow 1}}$ $\approx1.4844$ $\approx1.6887$ $\approx 2.7001$
: Coefficient $R={t_{L \rightarrow L}}/{t_{1 \rightarrow 1}}$ for three different rebound parameters (see main text for details). Density for all cases $\rho=3/4$, the size of the lattice $N=7$.[]{data-label="tab:N7tavR"}
For $N=7$ we investigate properties of the system with the density $\rho=3/4$. Parameters are chosen as follows: $\mu=1$, $\nu=173024/10000$ gives the density $\rho=0.7500001621$, $\theta=1/3$ gives $\rho=3/4$ exactly, and $\mu=1$, $\nu:=1000000/1578886$ gives $\rho=0.7500002817$. Distributions of clusters are presented in Figure \[fig:Fig3a\] and average cluster and avalanche sizes in Table \[tab:N7av\]. Again differences in distributions $n_i$ are not big, but average avalanche size differs significantly between considered cases.
The novel property visible in the figure is that the highest probability is for the cluster of maximal size $i=N$. Thus, the system prefers merging clusters for high density.
A comparison of the ratios of return times $R={t_{L \rightarrow L}}/{t_{1 \rightarrow 1}}$ is presented in the right panel of Figure \[fig:FigRet5\], while formulas are presented in the Appendix. In each case the ratio is a rational function of $t$, which is equal to $1$ for $t=0$ and asymptotically $\sim t^6$ for $t\longrightarrow \infty$, which supports a Conjecture formulated in Section \[sec:conclusions\]. Table \[tab:N7tavR\] shows that for the cases discussed above, with average density $\rho=3/4$, the highest value of the ratio $R$ is for $\mu_i=\delta/i$ and the smallest for $\mu_i=const.$ (which does not differ much from the value for $\mu_i=\delta/i$). This is an opposite order comparing to the case with $\rho=1/5$ for $N=5$ considered above. Thus, for higher densities the automaton prefers more periodic-like behaviour when it is relatively easier to trigger big avalanches.
![Probability distributions of actual density of the system for the case $\mu_i/\nu=\theta/i$ for various average densities of $0.2,0.4,0.5,0.6,0.8$ – the respective parameters are $\theta=4,\frac{3}{2},1, \frac{2}{3},\frac{1}{4}$. On the plot: smaller average density corresponds to the higher probability rate for density equal to $0$ (and to the lower probability rate for density equal to $1$). The size of the system is $N=7$.[]{data-label="fig:Fig3d"}](N7dens){width="8cm"}
The size $N=7$ is big enough to notice how the actual density of the system (possible values are $0,\frac{1}{7},\frac{2}{7},\frac{3}{7},\frac{4}{7},\frac{5}{7},\frac{6}{7},1$) is distributed for various average densities. Results are shown in Figure \[fig:Fig3d\]. For small densities, like $\rho=0.2$, the maximum is for small $i$, that means that big densities and big avalanches are rare. Then, when the density increases, the bell-like shape distribution appears and its maximum is shifted to the bigger values. Next, for densities like $0.6$ or bigger, the maximum probability is for biggest possible size $i=N$ and the most probable state is that with $\rho=1$. To achieve big average density, the system must spend a substantial time being fully occupied. The evolution of such a system consists of two phases: filing up and waiting for avalanche of maximal size, as is described above while discussing the times of return for $N=3$.
For $N=500$ and constant parameters $\mu=\nu=1$, numerical experiments show that the density fits a Gaussian distribution [@BiaCz-Mon].
$N=10$
------
In the example with the biggest $N$ presented here we investigate in several cases influence of correlations and compare exact results with proposed approximations for $\alpha^A_{i}$, $\gamma^A_{i}$ and $\gamma^{AR}_{i}$. On the other hand, size $N=10$ requires relatively complex calculations – the transition matrix is of size $108 \times 108$ and has about $1000$ non-zero entries. All possible states are presented in Table \[tab:statesN10\].
The size $N=10$ is the smallest with states which consist of the same clusters, but in essentially different order. (For smaller $N$ states with different order of clusters were equivalent with respect to reflections.) Namely, the state $88$ $$\hookrightarrow | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \hookleftarrow$$ and the state $89$ $$\hookrightarrow | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \hookleftarrow$$ In this subsection we consider also the relative difference between probabilities of these two states, namely $\Delta=(p_{88}-2p_{89})/p_{88}$ for various rebound parameters as a measure of adequacy of independence of clusters assumption. The multiplier $2$ in the above formula is necessary because the multiplicity of state $89$ is equal to five, and the multiplicity of the state $88$ is equal to ten. This quantity reflects dependence of respective probabilities on specific order of clusters in the system. We assume there is no such dependence in order to write down approximations $\alpha^A_{i}$, $\gamma^A_{i}$ and $\gamma^{AR}_{i}$.
Other quantities analysed in examples below are $\frac{\alpha^E_{1}-\alpha^A_{1}}{\alpha^E_{1}}$, $\frac{\alpha^E_{4}-\alpha^A_{4}}{\alpha^E_{4}}$, $\frac{\gamma^E_{5} - \gamma^A_{5}}{\gamma^E_{5}}$, $\frac{\gamma^E_{5} - \gamma^{AR}_{5}}{\gamma^E_{5}}$, $\frac{\gamma^E_{9} - \gamma^A_{9}}{\gamma^E_{9}}$ and $\frac{\gamma^E_{9} - \gamma^{AR}_{9}}{\gamma^E_{9}}$. These quantities measure the quality of approximation formulas for $i=1$ and $4$ for $\alpha$ coefficients, and for $i=5$ and $9$ for $\gamma$ coefficients - just to test approximations for $n_2$ – the first approximate equation, $n_5$ – the middle one, and $n_9$ the last one (those for $n_1$ and $n_{10}$ are exact). A formula for exact value of $\alpha^E_{4}$, obtained from detailed analysis of states of the automaton, is presented in Appendix.
![Plot of the $Log_{10}$ of $n_i$s (left) and $n_i^0$ (right) for $N=10$ in three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). []{data-label="fig:Fig10_3"}](n_i_prob10_3 "fig:"){width="4cm"} ![Plot of the $Log_{10}$ of $n_i$s (left) and $n_i^0$ (right) for $N=10$ in three cases: ${\mu_i}=const.$ (dashed line), $\mu_i=\delta/i$ (solid line) and $\mu_i=\sigma/i^2$ (dotted line). []{data-label="fig:Fig10_3"}](n_i0_prob10_3 "fig:"){width="4cm"}
${\mu_i}=1$ $\mu_i=1/i$ $\mu_i=1/i^2$
------------------------------------------------------- ---------------- --------------- ---------------
$\rho$ 0.3076370614 0.5 0.8822697788
$\left\langle i \right\rangle$ $1.5985438$ $2.872872532$ $7.493995763$
$\left\langle w \right\rangle$ $2.250583644 $ $2.872872532$ $3.725820785$
$\frac{p_{88}-2p_{89}}{p_{88}}$ $0.00865$ $0.01899$ $0.01868$
$\frac{\alpha^E_{1}-\alpha^A_{1}}{\alpha^E_{1}}$ $0.00909$ $0.066784$ $0.19468$
$\frac{\alpha^E_{4}-\alpha^A_{4}}{\alpha^E_{4}}$ $0.08795 $ $0.07428$ $0.09239 $
$\frac{\gamma^E_{5} - \gamma^A_{5}}{\gamma^E_{5}}$ $-0.01141$ $-0.28676$ $-0.89662$
$\frac{\gamma^E_{5} - \gamma^{AR}_{5}}{\gamma^E_{5}}$ $-0.00110$ $-0.06842$ $-0.13292$
$\frac{\gamma^E_{9} - \gamma^A_{9}}{\gamma^E_{9}}$ $0.35717$ $0.29878$ $0.17045$
$\frac{\gamma^E_{9} - \gamma^{AR}_{9}}{\gamma^E_{9}}$ $0.48970$ $0.62573$ $0.71428$
$t_{1\rightarrow 1}$ 18.51 25.59 96.28
$t_{L\rightarrow L}/t_{1\rightarrow 1}$ 379.61 5.2988 1.4606
: Three cases: $\mu=1$, $\delta=1$ and $\sigma=1$ (and always $\nu=1$) for the size of the lattice $N=10$.[]{data-label="tab:N10-3"}
#### Cases with constants equal to $1$.
As a first set we consider three cases with the minimal possible rebounds factors, i.e., we put all constants equal to $1$. Cases with $\mu_i=1$, $\mu_i=1/i$ and $\mu_i=1/i^2$ with $\nu=1$ are presented in Figure \[fig:Fig10\_3\] and Table \[tab:N10-3\].
Three different rebound parameter types result in various average density values, and hence different distributions. In all cases, the assumption of independence of clusters is well satisfied; the respective error $\Delta$ does not exceed $2\%$. An approximation for $\alpha^E_{4}$ is less than $10\%$ for all cases, but $\alpha^E_{1}$ strongly depends on the case (in fact it depends on density, as will be seen below). Approximation formulas for $\gamma_E$ perform in diversified way – $\gamma_{AR}$ is better for mid $i$ terms, while $\gamma_A$ is better for big $i$ terms. Nevertheless, both cases provide rather roughly appropriate values. These examples also suggest that for higher densities the system exhibits a periodic-like evolution.
#### Big densities
![Plot of the $Log_{10}$ of $n_i$s and $n_i^0$ vs. $Log_{10}(i)$ for $N=10$ in two cases: ${\mu_i}=const.$ (solid line) and $\mu_i=\delta/i$ (dashed line) – upper panels, and in case $\mu_i=\delta/i^2$ – lower panels. []{data-label="fig:Fig10_big_rho_LL"}](ni_N10_big_rhoLL "fig:"){width="4cm"} ![Plot of the $Log_{10}$ of $n_i$s and $n_i^0$ vs. $Log_{10}(i)$ for $N=10$ in two cases: ${\mu_i}=const.$ (solid line) and $\mu_i=\delta/i$ (dashed line) – upper panels, and in case $\mu_i=\delta/i^2$ – lower panels. []{data-label="fig:Fig10_big_rho_LL"}](ni0_N10_big_rhoLL "fig:"){width="4cm"}\
![Plot of the $Log_{10}$ of $n_i$s and $n_i^0$ vs. $Log_{10}(i)$ for $N=10$ in two cases: ${\mu_i}=const.$ (solid line) and $\mu_i=\delta/i$ (dashed line) – upper panels, and in case $\mu_i=\delta/i^2$ – lower panels. []{data-label="fig:Fig10_big_rho_LL"}](ni_N10_bigLL "fig:"){width="4cm"} ![Plot of the $Log_{10}$ of $n_i$s and $n_i^0$ vs. $Log_{10}(i)$ for $N=10$ in two cases: ${\mu_i}=const.$ (solid line) and $\mu_i=\delta/i$ (dashed line) – upper panels, and in case $\mu_i=\delta/i^2$ – lower panels. []{data-label="fig:Fig10_big_rho_LL"}](ni0_N10_bigLL "fig:"){width="4cm"}
${\mu_i}\sim1$ $\mu_i\sim1/i$ $\mu_i\sim1/i^2$
------------------------------------------------------- ---------------- ---------------- ------------------
$\rho$ 0.9145269069 0.9145269069 0.9897960692
$\left\langle i \right\rangle$ 7.764600567 7.805017612 9.700144892
$\left\langle w \right\rangle$ 9.346154002 7.805017612 8.33021261
$\frac{p_{88}-2p_{89}}{p_{88}}$ 0.00161 0.00464 0.00191
$\frac{\alpha^E_{1}-\alpha^A_{1}}{\alpha^E_{1}}$ 0.30898 0.28421 0.31377
$\frac{\alpha^E_{4}-\alpha^A_{4}}{\alpha^E_{4}}$ -0.01511 0.01520 0.00330
$\frac{\gamma^E_{5} - \gamma^A_{5}}{\gamma^E_{5}}$ -0.96763 -0.93825 -1.05131
$\frac{\gamma^E_{5} - \gamma^{AR}_{5}}{\gamma^E_{5}}$ -0.15298 -0.14354 -0.14789
$\frac{\gamma^E_{9} - \gamma^A_{9}}{\gamma^E_{9}}$ 0.42749 0.39218 0.38965
$\frac{\gamma^E_{9} - \gamma^{AR}_{9}}{\gamma^E_{9}}$ 0.80211 0.78888 0.80019
$t_{1\rightarrow 1}$ 119.18 123.60[^3] 1002.44
$t_{L\rightarrow L}/t_{1\rightarrow 1}$ 1.1066 1.1339 1.0277
: Three cases with “big” $\rho$ for the size of the lattice $N=10$ (see main text for details).[]{data-label="tab:N10-bigrho"}
In order to investigate evolution of the system with high average density (and strong deviations in actual density) we consider case $\mu=const$ with $\mu_1=1/100$ and $\nu=1$, which gives the density $\rho\approx 0.91$, and case $\delta/i$ with $\mu_1=4673077001/5*10^{10}\approx0.093$ and $\nu=1$ to obtain the same density (with $10$ digits accuracy) for comparison. Also we consider case of $\sigma/i^2$ with $\mu_1=1/10$ and $\nu=1$ which gives the density $\rho\approx 0.99$. The results are presented in Figure \[fig:Fig10\_big\_rho\_LL\] and Table \[tab:N10-bigrho\].
Plots of respective distributions for $\mu_i\sim1/i$ and ${\mu_i}\sim1$ are overlapping each other. For relatively small size $N=10$, fixing the average density of the system strongly determines distributions, making the dependence on rebound parameters not essential. Their influence becomes more visible for larger sizes $N$ of the lattice. In case of high density, the system just spend much time being fully occupied.
For high densities, the assumption of independence of clusters is well satisfied; the respective error $\Delta$ does not exceed $0.5\%$. An approximation for $\alpha^E_{4}$ is fairly good ($\approx 1.5\%$ or less), but $\alpha^E_{1}$ has only accuracy $\approx 30\%$. Approximation formula for $\gamma_{AR}$ is much better for mid $i$ terms (though giving only $\approx 15\%$ accuracy), while $\gamma_A$ is better for big $i$ terms ($\approx40\%$). Thus, for high density cases the proposed set of equations for $n_i$s does not reproduce actual distribution. Note, however, that there are other exact equations valid for any density.
The parameter $t_{L\rightarrow L}/t_{1\rightarrow 1}$ for $\mu_i\sim1/i$ case is bigger than for ${\mu_i}\sim1$ case (both cases have the same “big” density), which agrees with the results for $N=7$ with $\rho=3/4$ presented in Table \[tab:N7tavR\].
${\mu_i}\sim1$ $\mu_i\sim1/i$ $\mu_i\sim1/i^2$
------------------------------------------------------- ------------------------ --------------------------- ------------------------
$\rho$ 0.0779280356 0.01031150521 0.01031150521
$\left\langle i \right\rangle$ 1.09149321 1.02083788 1.041894016
$\left\langle w \right\rangle$ 1.183235221 1.02083788 1.020408163
$\frac{p_{88}-2p_{89}}{p_{88}}$ $6.9231*10^{-5}$ $6.4609*10^{-4}$ $3.5663*10^{-3}$
$\frac{\alpha^E_{1}-\alpha^A_{1}}{\alpha^E_{1}}$ $1.7483*10^{-5}$ $2.7693 *10^{-7}$ $1.0585*10^{-6}$
$\frac{\alpha^E_{4}-\alpha^A_{4}}{\alpha^E_{4}}$ $0.14229$ $0.16313$ $0.16274$
$\frac{\gamma^E_{5} - \gamma^A_{5}}{\gamma^E_{5}}$ $0.0009069$ $0.0020519158$ $0.0040100$
$\frac{\gamma^E_{5} - \gamma^{AR}_{5}}{\gamma^E_{5}}$ 0.0009101 0.0020519262 0.0040124
$\frac{\gamma^E_{9} - \gamma^A_{9}}{\gamma^E_{9}}$ 0.09615 0.01473 0.00606
$\frac{\gamma^E_{9} - \gamma^{AR}_{9}}{\gamma^E_{9}}$ 0.12138 0.02356 0.03305
$t_{1\rightarrow 1}$ 22.32 106.35 110.57
$t_{L\rightarrow L}/t_{1\rightarrow 1}$ 2220903488.0 $1.666292752\cdot10^{14}$ 9971770329.0
$\sim 2\cdot 10^{10} $ $\sim 2\cdot 10^{14} $ $\sim 1\cdot 10^{11} $
: Three cases with “small” $\rho$ for the size of the lattice $N=10$ (see main text for details).[]{data-label="tab:N10-smallrho"}
#### Small densities
![Plot of the $Log_{10}$ of $n_i$s and $n_i^0$ for $N=10$ in case ${\mu_i}=const.$ – upper line, and in two cases: ${\mu_i}=\delta/i$ (solid line) and $\mu_i=\delta/i^2$ (dashed line) – lower line. []{data-label="fig:Fig10_small_rho"}](ni_N10_small "fig:"){width="4cm"} ![Plot of the $Log_{10}$ of $n_i$s and $n_i^0$ for $N=10$ in case ${\mu_i}=const.$ – upper line, and in two cases: ${\mu_i}=\delta/i$ (solid line) and $\mu_i=\delta/i^2$ (dashed line) – lower line. []{data-label="fig:Fig10_small_rho"}](ni0_N10_small "fig:"){width="4cm"}\
![Plot of the $Log_{10}$ of $n_i$s and $n_i^0$ for $N=10$ in case ${\mu_i}=const.$ – upper line, and in two cases: ${\mu_i}=\delta/i$ (solid line) and $\mu_i=\delta/i^2$ (dashed line) – lower line. []{data-label="fig:Fig10_small_rho"}](ni_N10_small_2 "fig:"){width="4cm"} ![Plot of the $Log_{10}$ of $n_i$s and $n_i^0$ for $N=10$ in case ${\mu_i}=const.$ – upper line, and in two cases: ${\mu_i}=\delta/i$ (solid line) and $\mu_i=\delta/i^2$ (dashed line) – lower line. []{data-label="fig:Fig10_small_rho"}](ni0_N10_small_2 "fig:"){width="4cm"}
To present system behaviour in small average density we choose $\mu_1=1$ and $\nu=1/10$ for case $\mu=const$ - it gives density $\rho\approx0.08$. Then for the remaining two cases we have the same density $\rho\approx0.01$ (with $10$ digits accuracy), with the following parameters: $\mu_1=1$ and $\nu=50000000/4798952601\approx 0.01$ - for case $\delta/i$ and $\mu_1=1$ and $\nu=1/100$ for case $\sigma/i^2$. The results are presented in Figure \[fig:Fig10\_small\_rho\] and Table \[tab:N10-smallrho\].
For small densities assumption of independence of clusters is well satisfied. In general, all proposed approximations are fairly good. An approximation for $\alpha^E_{4}$ is the worst; its accuracy is only $\approx 15\%$. As previously, approximation formula for $\gamma_A$ is better than $\gamma_{AR}$ for big $i$ terms, but it appears that for mid $i$ terms both formulas give almost the same values (because $n_i$s decrease rapidly). Thus, for small densities the set of equations for $n_i$s can be used to reproduce the actual distribution.
It is very improbable to find the lattice fully occupied for small average densities, which is reflected in high values of the parameter $R=t_{L\rightarrow L}/t_{1\rightarrow 1}$. The parameter $R$ for $\mu_i\sim1/i$ case is bigger than its for ${\mu_i}\sim1/i^2$ case (both cases have the same “small” density), which agrees with the results for $N=5$ with $\rho=1/4$ presented in Table \[tab:N5tavR\].
Conclusions {#sec:conclusions}
===========
In this article we investigated in detail a finite version of one-dimensional non-equilibrium dynamical system – Random Domino Automaton. It is a simple, slowly driven system with avalanches. The advantage of RDA (comparing to Drossel-Schwabl model) is the dependence of rebound parameters on the size of a cluster. This crucial extension allows for producing a wider class of distributions by the automaton, as well as leads to several exact formulas. Exponential type and inverse-power type distributions of clusters were studied in [@BiaCzAA]; the present work examines also V-shape distributions and quasi-periodic like behavior.
Detailed analysis of finite RDA, including finite size effects, extends and explains the previously obtained results for RDA. Moreover, we also analyzed approximations made when deriving equations for the stationary state of the automaton. This allows for the following conclusions.
The balance of $\rho$ equation and the balance of $N$ equations are exact – their forms incorporate all correlations present in the system. The first one has a form independent of the size of the lattice $N$, thus it is exactly the same as for RDA. The second one contains correction for finite size effect, namely a term $(2n_R-n_{N-1})$, which replaces the term $2n$ for RDA. When $n_{N-1}$ and $n_N$ are negligible, these two terms coincide. For finite RDA, balance of $n_i$s equations - contains two extra equatins, for $i=N-1$ and $i=N$, comparing to the those for RDA. The first (for $n_1$) and the last (for $n_N$) are exact. Note that all those equations are written for rebound parameter $\mu=\mu(i)$ being a function of cluster size and $\nu$ being a constant.
The most remarkable special case is when $\mu=\delta/i$, when any cluster has the same probability to be removed as an avalanche independently of its size $i$. It appears that the system depends on a single parameter $\theta=\delta/\nu$, or equivalently, due to neat exact formula (eq. ) $$\rho=\frac{1}{1+\theta},$$ the properties of the system may be characterized by the value of the average density. Note that the above expression does not depends on the size $N$, and is the same as for RDA. This specialization leads to more neat formulas, like the equation for $n_R$ (eq. ) $$n_R=N\frac{\theta}{(\theta+1)(\theta+2)}.$$ Note again that it has the same form as for RDA, except that $n$ is replaced by $n_R$ ($n=n_R+n_N$). Summarizing, the model allows to derive a number of explicit dependencies, as shown in Sections \[sec:equations\] and \[sec:cases\].
The Random Domino Automaton defines a discrete time Markov process of order 1 and, in principle, may be solved exactly. However, it turns out that computations are fairly complex and exact formulas are long, as visible from examples presented in the Appendix. Also, the exact numerical values are in the form of big numbers — in every considered example ($N=3,4,5,6,7,10$) significantly big prime numbers were encountered. For example, for the simplest possible rebound parameters ($\mu=1$ and $\nu=1$) the exact value of denominators of probabilities of states for $N=10$ (see Table \[tab:statesN10\]) is a 65-digit integer. Its prime factorization (presented in the Appendix) contains a 56-digit integer, which cannot be simplified with numerators. Thus, the usefulness of Markov chains for finding both formulas and values $n_i$s is limited in practice.
Nevertheless, Markov chains framework leads to interesting results concerning analysis of times of recurrence for specific states. A return time to the state with density $\rho=1$ (equation ) $$t_{L \rightarrow L}= 1/\mu_N + \hat\tau_1,$$ consists of two parts: waiting of fully occupied lattice for triggering a maximal avalanche and “loading” time, when the lattice is filled up, respectively. If the average density of the system is small, the second time is very long. The formula is more interesting for systems with relatively big average density, when the “loading” time is comparable to waiting time for triggering the biggest avalanche. Such a system exhibits a periodic like behavior. Dividing the waiting time $t_{L \rightarrow L}$ by the waiting time $t_{1 \rightarrow 1}$ (given by equation ) one has the following measure of quasi-periodicity $$R= t_{L \rightarrow L} / t_{1 \rightarrow 1} = t_{1 \rightarrow L \rightarrow 1} / t_{1 \rightarrow 1}.$$ If $R=1$ then the system is periodic.
Several considered examples lead to the following conjecture concerning the coefficient $R$.\
[**Conjecture.**]{} The ratio of return times $t_{L\longrightarrow L}/t_{1\longrightarrow 1}$ as a function of $t$ being the ratio of constants from rebound parameters ($\mu/\nu$, $\delta/\nu$, $\sigma/\nu$) are rational functions of $t$, $f(t)=t_{L\longrightarrow L}/t_{1\longrightarrow 1}$ with the following properties for any size $N$ of the system $$\begin{aligned}
f(t=0)&=& 1, \\
\lim_{t \rightarrow \infty} \frac{f(t)}{t^{N-1}} &=& const.
\label{eq:conj}\end{aligned}$$ The conjecture relates the size of the system $N$ with asymptotic behavior of ratio of waiting times.
There are big fluctuations (variations of actual density) during the evolution of systems with relatively big average densities. If the system is likely to achieve a fully occupied state, the next state is an empty state, and the variations in density are maximal. Nevertheless, some parameters of stationary state (more precisely, statistically stationary state) satisfy exact equations, as shown above. For big average densities, the system fluctuates within the whole possible range, and cannot be thought of as having approximately stationary values during the evolution. This aspect is easy to be overlooked (see [@PBfire]).
It is argued in the Appendix that no exact equations for $n_i$s exist for the size $N \geq 5$. Thus, to have compact equations for $n_i$s, some approximation formulas are proposed. The first general conclusion from the examples is that the approximations are acceptable for small densities, but for big densities the errors are substantial. The main reason is that for big densities correlations become more important and fluctuations makes actual values of the parameters substantially different from their stationary values, which are present in the formulas. These properties are particularly severe for small sizes of the system, where every avalanche changes the actual density considerably.
Table \[tab:comp\_alpha\] presents a dependence of a relative error of $\alpha^E_{1}$ with respect to $\alpha^A_{1}$ on size $N$ of the system. For bigger $N$ the accuracy of approximation is growing, which corresponds well with the remark in the last paragraph.
$N$ $\alpha^E_{1}=\frac{2n_2}{n_1}$ $\alpha^A_{1} = \left(1-\frac{n_1^0}{n_R} \right)$ $\frac{\alpha^E_{1} - \alpha^A_{1}}{\alpha^E_{1}}$ $\rho$
------------ --------------------------------- ---------------------------------------------------- ---------------------------------------------------- ------------------
$3$ $ 1.33(3) $ $0.6$ $0.55$ $\approx 0.3462$
$4$ $ 0.66(6) $ $\approx 0.6316 $ $\approx 0.053 $ $0.32(32)$
$5$ $ \approx 0.6829 $ $\approx 0.6565$ $\approx0.039$ $\approx 0.3139$
$6$ $ \approx 0.685296$ $\approx 0.669232$ $\approx0.023$ $\approx 0.3102$
$7$ $ \approx 0.685523$ $\approx 0.675066$ $\approx0.015$ $\approx 0.3086$
$10$ $\approx 0.685436 $ $\approx 0.679205$[^4] $\approx0.0091$ $\approx0.3076$
$4000$[^5] $\approx 0.677$ $\approx 0.677$ — $\approx 0.3076$
: Comparision of $\alpha^E_{1}$ with $\alpha^A_{1}$ for $N=3,4,5,6,7$ and $10$ for parameters $\mu=1$ and $\nu=1$. The last line presents value of $\alpha^A_{1}$ for $N=4000$ obtained from simulations and equations (respectively) in [@BiaCzAA].[]{data-label="tab:comp_alpha"}
It can be noticed from the distributions of $n_i$s of examples presented above that all $n_i$s except of the last two (namely $n_{N-1}$ and $n_N$) are placed on one “regular” curve, while the last two deviate from it. It may be regarded as a (correction of) finite size effect. Also in the respective set of equations -, the last two (for $i-N-1$ and $i=N$) have a form different from the previous ones. Thus, neglecting the size restriction, which in fact ignores the last two equations, is justified when the deviations of the last two $n_i$s from the “regular” curve are not big. That happens for small densities.
It appears also that for index $i$ in his middle range of values an approximation formula $\gamma_{AR}$ works better than $\gamma_{A}$, in spite of the fact that it looks to be more rough approximation. For distribution of $n_i$s vanishing rapidly (i.e., for small densities) both give comparable results.
All this justifies the form of equations for $n_i$s presented in [@BiaCzAA] as valid for small densities. A detailed examination of the RDA for big densities requires further investigations.
This article explores properties of FRDA in order prepare to modeling of real data. In this context, among others, formulas for waiting times can be used. We ephasize also a formula $$t_{av}= \frac{\left\langle w \right\rangle + 1}{1-P_r},$$ which relates the measure of scattering (dissipation) of balls $P_r$ with the average size of avalanche $\left\langle w \right\rangle$ and average time between any two consecutive avalanches $t_{av}$, which are a priori measurable quantities.
The Random Domino Automaton proved to be a stochastic dynamical system with interesting mathematical structure. It may be viewed as extension of Drossel-Schwabl model, and we showed that this is a substantial generalization with a wide range of novel properties. W expect it can also be applied to natural phenomena, including earthquakes and forest-fires. This is our aim for the future work.
Appendix. Exact equations for $N=3,4$ and their non-existence for $n\geq5$ {#appendix.-exact-equations-for-n34-and-their-non-existence-for-ngeq5 .unnumbered}
==========================================================================
For arbitrary size $N$, there are four exact equations: balance of $\rho$ - equation , balance of $n$ - equation , for $n_1$ - equation and for $n_N$ - equation .
[**Size $N=3$.**]{} Equation for $n_2$ is of the form , namely $$n_2 =\frac{1}{2\frac{\mu_1}{\nu}+1} (2 n_1 \alpha^E_{1} + n_1^0 \gamma^E_2).$$ In this case the only companion to single one-cluster is an empty two-cluster (see state $2$ in Tab.\[tab:N3\]), hence $$\alpha^E_{1}=1 \quad \text{and} \quad \gamma^E_{2}=0.$$ Thus, we arrived at the exact form of the equation for $n_2$.
[**Size $N=4$.**]{}
state number example[^6] multiplicity contrib. to
-------------- ------------------------------------------------------------------------------ -------------- ----------------
1 $ \hookrightarrow | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \hookleftarrow $ 1 $ n_4^0 $
2 $ \hookrightarrow | \ \ \ | \ \ \ | \ \ \ | \bullet | \hookleftarrow $ 4 $n_1, n_3^0$
3 $ \hookrightarrow | \ \ \ | \ \ \ | \bullet | \bullet | \hookleftarrow $ 4 $n_3, n_2^0$
4 $ \hookrightarrow | \ \ \ | \bullet | \ \ \ | \bullet | \hookleftarrow $ 2 $ n_1, n_1^0 $
5 $ \hookrightarrow | \ \ \ | \bullet | \bullet | \bullet | \hookleftarrow $ 4 $ n_3, n_1^0 $
6 $ \hookrightarrow | \bullet | \bullet | \bullet | \bullet | \hookleftarrow $ 1 $ n_4 $
: States for the size of the lattice $N=4$. []{data-label="tab:N4"}
All states of the automaton and their labels are presented in Tab.\[tab:N4\]. Equation for $n_2$ is of the form $$n_2 = \frac{2}{2\frac{\mu_2}{\nu}+2}n_1\alpha^E_{1},$$ where to $\alpha^E$ contributes only state $2$, not state $4$. Hence $$\alpha^E_{1}=\frac{p_2}{p_2+2p_4}=\left(1-\frac{2p_4}{n_1}\right)= \left(1-\frac{n_1^0-n_3}{n_1}\right),$$ where $p_i$ is probability of state $i$. Thus $\alpha^E_{1}$ is expressed as function of $n_i$s and $n_1^0$. The equation for $n_3$ is of the form $$n_3=\frac{1}{3\frac{\mu_3}{\nu}+1}\left( 2n_2 \alpha^E_{2} + n_1^0 \gamma^E_{3} \right).$$ In this case $$\alpha^E_{2}=1,$$ because only state $3$ contributes. The state $4$ (and not state $5$) contributes to $\gamma^E_3$, therefore $$\gamma^E_3=\frac{2p_4}{2p_4+p_5}=\left(1-\frac{p_5}{n_1^0}\right) =\left(1-\frac{n_3}{n_1^0}\right).$$ This completes the task of writing exact equations for $N=4$.
[**Size $N=5$.**]{} States and their labels are presented in Tab.\[tab:N5\]. In this case, the coefficients are as follows $$\alpha^E_{2}=\frac{p_2+p_4}{n_1}, \quad \quad \gamma^E_3=\frac{p_4-p_6-n_4}{n_1^0},$$
$$\alpha^E_{3}=\frac{p_3}{n_2}, \quad \quad \gamma^E_4=\frac{p_6-p_4-n_4}{n_1^0}.$$ Summing up the probabilities contributing to $n_1^0, n_1$ and $n_2$ one obtains $$\begin{aligned}
n_1^0 &=& p_4+ 2p_6+ n_4, \label{eq:n10fromstates}\\
n_1 &=& p_2+ 2p_4+p_6, \\
n_2 &=& p_3+ p_6. \label{eq:n2fromstates}\end{aligned}$$ The set cannot be solved for $p_2, p_3, p_4 ,p_6$. Since there are no more equations for those coefficients, respective $\alpha$s and $\gamma$s cannot be expressed as functions of $n_i$s only in an exact manner.
[**Sizes bigger than $5$.**]{} An argument for non-existence of exact set of equations -, i.e., non-existence of exact formulas for $\alpha^E_{i}$s and $\gamma^E_{i}$s as functions of $n_i$s and $n_1^0$ is based on the same impossibility of solving equations as presented above.
An increase of size of a grid $N$ by $1$ results in an increase of the set of $n_1, n_2, \ldots$ by one and much bigger increase of the number of states. An analog of the set of equations - will contain much more probabilities of states $p_1, p_2, \ldots$, on the right hand side – there will be more states containing $1$-clusters, $2$-clusters and so on, and contributing to $n_1,n_2,\ldots$ respectively. Thus, it is impossible to express those probabilities of states as functions of $n_1^0, n_1,n_2,\ldots$ only. As a consequence, there are no general exact formulas for $\alpha^E_{i}$s and $\gamma^E_{i}$.
Appendix. Formulas {#appendix.-formulas .unnumbered}
==================
The return times for $N=5$ for general values of the parameters $\mu_1,\mu_2,\mu_3,\mu_4,\mu_5$ and $\nu$ are
$t_{1\rightarrow1} = \frac{1}{\nu} +
\frac{1}{\mu_5 (4\mu_4 + \nu)(3\mu_3 + 2\nu)(2mu_1^3 \mu_2 + 2\mu_1^3\nu + 4\mu_1^2\mu_2^2 + 17\mu_1^2\mu_2\nu + 13\mu_1^2\nu^2 + 14\mu_1\mu_2^2\nu + 43\mu_1\mu_2\nu^2 + 33\mu_1\nu^3 + 16\mu_2^2\nu^2 + 42\mu_2\nu^3 + 36\nu^4)}
\times
(48\mu_1\nu^5 + 60\mu_2\nu^5 + 54\mu_3\nu^5 + 750\mu_5\nu^5 + 72\nu^6 + 8\mu_1^2\nu^4 + 8\mu_2^2\nu^4 + 20\mu_1\mu_2\nu^4 + 12\mu_1\mu_3\nu^4 + 24\mu_2\mu_3\nu^4 + 560\mu_1\mu_5\nu^4 + 760\mu_2\mu_5\nu^4
+ 720\mu_3\mu_5\nu^4 + 1560\mu_4\mu_5\nu^4 + 100\mu_1^2\mu_5\nu^3 + 160\mu_2^2\mu_5\nu^3 + 40\mu_1\mu_2^2\mu_5\nu^2 + 20\mu_1^2\mu_2\mu_5\nu^2 + 60\mu_1^2\mu_3\mu_5\nu^2 + 240\mu_1^2\mu_4\mu_5\nu^2
+ 150\mu_2^2\mu_3\mu_5\nu^2 + 480\mu_2^2\mu_4\mu_5\nu^2 + 340\mu_1\mu_2\mu_5\nu^3 + 390\mu_1\mu_3\mu_5\nu^3 +
1280\mu_1\mu_4\mu_5\nu^3 + 645\mu_2\mu_3\mu_5\nu^3 + 1840\mu_2\mu_4\mu_5\nu^3 + 1800\mu_3\mu_4\mu_5\nu^3 + 240\mu_1\mu_2^2\mu_3\mu_4\mu_5 + 120\mu1^2\mu_2\mu_3\mu_4\mu_5 + 285\mu_1\mu_2\mu_3\mu_5\nu^2 + 60\mu_1\mu_2^2\mu_3\mu_5\nu + 30\mu_1^2\mu_2\mu_3\mu_5\nu + 960\mu_1\mu_2\mu_4\mu_5\nu^2 + 160\mu_1\mu_2^2\mu_4\mu_5\nu + 80\mu_1^2\mu_2\mu_4\mu_5\nu + 1320\mu_1\mu_3\mu_4\mu_5\nu^2 + 240\mu_1^2\mu_3\mu_4\mu_5\nu + 2100\mu_2\mu_3\mu_4\mu_5\nu^2 + 600\mu_2^2\mu_3\mu_4\mu_5\nu + 1140\mu_1\mu_2\mu_3\mu_4\mu_5\nu),$
$t_{L\rightarrow L} =\frac{1}{\mu_5} + \frac{1}{2\nu^5(4\mu_1^2 + 10\mu_1\mu_2 + 24\mu_1\nu + 6\mu_3\mu_1 + 4\mu_2^2 + 30\mu_2\nu + 12\mu_3\mu_2 + 36\nu^2 + 27\mu_3\nu)}
\times
(626\mu_1\nu^5 + 844\mu_2\nu^5 + 828\mu_3\nu^5 + 1848\mu_4\nu^5 + 822\nu^6 + 126\mu_1^2\nu^4 + 4\mu_1^3\nu^3 + 192\mu_2^2\nu^4 + 8\mu_1^2\mu_2^2\nu^2 + 426\mu_1\mu_2\nu^4 + 489\mu_1\mu_3\nu^4 + 1544\mu_1\mu_4\nu^4 + 771\mu_2\mu_3\nu^4 + 2176\mu_2\mu_4\nu^4 + 2232\mu_3\mu_4\nu^4 + 68\mu_1\mu_2^2\nu^3 + 54\mu_1^2\mu_2\nu^3 + 4\mu_1^3\mu_2\nu^2 + 99\mu_1^2\mu_3\nu^3 + 6\mu_1^3\mu_3\nu^2 + 344\mu_1^2\mu_4\nu^3 + 16\mu_1^3\mu_4\nu^2 + 198\mu_2^2\mu_3\nu^3 + 608\mu_2^2\mu_4\nu^3 + 48\mu_1^2\mu_2^2\mu_3\mu_4 + 102\mu_1\mu_2^2\mu_3\nu^2 + 81\mu_1^2\mu_2\mu_3\nu^2 + 12\mu_1^2\mu_2^2\mu_3\nu + 272\mu_1\mu_2^2\mu_4\nu^2 + 216\mu_1^2\mu_2\mu_4\nu^2 + 32\mu_1^2\mu_2^2\mu_4\nu + 396\mu_1^2\mu_3\mu_4\nu^2 + 792\mu_2^2\mu_3\mu_4\nu^2 + 24\mu_1^3\mu_2\mu_3\mu_4 + 414\mu_1\mu_2\mu_3\nu^3 + 6\mu_1^3\mu_2\mu_3\nu + 1304\mu_1\mu_2\mu_4\nu^3 + 16\mu_1^3\mu_2\mu_4\nu + 1716\mu_1\mu_3\mu_4\nu^3 + 24\mu_1^3\mu_3\mu_4\nu + 2604\mu_2\mu_3\mu_4\nu^3 + 1656\mu_1\mu_2\mu_3\mu_4\nu^2 + 408\mu_1\mu_2^2\mu_3\mu_4\nu + 324\mu_1^2\mu_2\mu_3\mu_4\nu).$
Their ratio is
$ t_{L\rightarrow L}/ t_{1\rightarrow1} = \frac{(4\mu_4 + \nu)*(3\mu_3 + 2\nu)*(2\mu_1^3\mu_2 + 2\mu_1^3\nu + 4\mu_1^2\mu_2^2 + 17\mu_1^2\mu_2\nu + 13\mu_1^2\nu^2 + 14\mu_1\mu_2^2\nu + 43\mu_1\mu_2\nu^2 + 33\mu_1\nu^3 + 16\mu_2^2\nu^2 + 42\mu_2\nu^3 + 36\nu^4)}
{2\nu^4(4\mu_1^2 + 10\mu_1\mu_2 + 24\mu_1\nu + 6\mu_3\mu_1 + 4\mu_2^2 + 30\mu_2\nu + 12\mu_3\mu_2 + 36\nu^2 + 27\mu_3\nu)}.$
Average waiting times $t_i$ for avalanche of size $i$ in case $\mu_i=\delta/i$, where $\delta=const.$ and $t:=\delta/\nu$, for $N=5$, are
$$\begin{aligned}
t_1 &=& \frac{4 t^5 + 48 t^4 + 237 t^3 + 603 t^2 + 762 t + 360}{\nu t^2 (4 t^3 + 28 t^2 + 69 t + 60)}, \label{eq:wt1}\\
t_2 &=& \frac{4 t^5 + 48 t^4 + 237 t^3 + 603 t^2 + 762 t + 360}{ 2\nu t^2(4 t^2 + 16 t + 15)}, \\
t_3 &=& \frac{4 t^6 + 56 t^5 + 333 t^4 + 1077 t^3 + 1968 t^2 + 1884 t + 720}{ 2\nu t^2 (10 t^2 + 31 t + 18)},\\%
t_4 &=& \frac{4 t^7 + 60 t^6 + 389 t^5 + 1410 t^4 + 3045 t^3 + 3852 t^2 + 2604 t + 720}{8\nu t^2 (7 t^2 + 24 t + 18)}, \\
t_5 &=& \frac{4 t^7 + 60 t^6 + 389 t^5 + 1410 t^4 + 3045 t^3 + 3852 t^2 + 2604 t + 720}{8\nu t(7 t^2 + 24 t + 18)}. \label{eq:wt5}\end{aligned}$$
The ratio of return times for $N=7$ for three cases:
for $\mu=const.$
$t_{L\rightarrow L}/ t_{1\rightarrow1}= (5184000t^{16} + 90633600t^{15} + 734038560t^{14} + 3656624904t^{13} + 12543798852t^{12} + 31435490078t^{11} + 59579986661t^{10} + 87223274254t^9 + 99846813214t^8 + 89833419890t^7 + 63379753809t^6 + 34652851894t^5 + 14319281196t^4 + 4279417752t^3 + 859191840t^2 + 101520000t + 5184000)/(5184000t^{10} + 60393600t^9 + 306948960t^8 + 896350104t^7 + 1664901648t^6 + 2053477662t^5 + 1700206878t^4 + 930252240t^3 + 320428800t^2 + 62380800t + 5184000), $
for $\mu_i/\nu=\theta/i$
$ t_{L\rightarrow L}/ t_{1\rightarrow1}=(576t^{16} + 16800t^{15} + 229696t^{14} + 1956752t^{13} + 11645844t^{12} + 51472058t^{11} + 175326610t^{10} + 471411274t^9 + 1015867913t^8 + 1768373403t^7 + 2486683328t^6 + 2797983376t^5 + 2465006400t^4 + 1636404624t^3 + 767257920t^2 + 225504000t + 31104000)/(76032t^{10} + 1450368t^9 + 12336072t^8 + 61572600t^7 + 199652130t^6 + 439389384t^5 + 664690536t^4 + 682575840t^3 + 455457600t^2 + 178329600t + 31104000),$
and for $\mu=\sigma/i^2$
$ t_{L\rightarrow L}/ t_{1\rightarrow1}= (27000t^{16} + 1648350t^{15} + 46021545t^{14} + 781598610t^{13} + 9060806565t^{12} + 76286696592t^{11} + 484749056302t^{10} + 2385421175676t^9 + 9253317988496t^8 + 28615082281632t^7 + 70836261328608t^6 + 139636245477312t^5 + 215233793554176t^4 + 250177275371520t^3 + 205696375603200t^2 + 106205478912000t + 25798901760000)/(1032264000t^{10} + 28816738200t^9 + 362493838440t^8 + 2695926522960t^7 + 13095290178720t^6 + 43360671643200t^5 + 99111840724224t^4 + 154547465674752t^3 + 157566016143360t^2 + 95025954816000t + 25798901760000). $
The exact value of $\alpha^E_{4}$ for $N=10$: $$\alpha^E_{4}= \frac{2v_{20}+v_{43}+2v_{44}+2v_{45}+v_{46}+v_{72}+2v_{73}+v_{74}+v_{77}+2v_{79}+v_{95}+v_{96}}
{
2(v_{20}+v_{43}+v_{44}+v_{45}+v_{46}+v_{72}+v_{73}+v_{74}+v_{77}+v_{78}+v_{79}+v_{95}+v_{96}+v_{98}+v_{99}+2v_{106} )
}.$$
A prime factorization of the common factor of probabilities of states of FRDA for $N=10$ (presented in Table \[tab:statesN10\]) for rebound parameters $\mu=1$ and $\nu=1$: $$2^2*3^2*607*66617*3 56222 1887802308 6289926346 2293218927 8661478627 7415883867.$$ The biggest prime has 56 digits.
states probability sym
---- ------------------------------------------------------------------------------------------------- ------------- ----------
1 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | $ 0.05402 $^{(1)}$
2 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | $ 0.13914
3 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.06060
4 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.04841
5 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | $ 0.04075
6 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | $ 0.03777
7 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | $ 0.01853 $^{(5)}$
8 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.03074
9 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet| \ \ \ | \bullet | \bullet | $ 0.02256 14
10 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.01807 13
11 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.01657 12
12 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.01657 11
13 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.01807 10
14 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.02256 9
15 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.01744
16 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.01450 18
17 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.01371
18 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.01450 16
19 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | $ 0.01252
20 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.01641
21 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.01168 25
22 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00926 24
23 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00864
24 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00926 22
25 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.01168 21
26 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet| \ \ \ | \bullet | \bullet | $ 0.01075
27 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00808
28 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00372 $^{(5)}$
29 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | $ 0.00825 38
30 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | $ 0.00697 37
31 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | $ 0.00703 35
32 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | $ 0.00857
33 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00663 36
34 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00612
35 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00703 31
36 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00663 33
37 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00697 30
38 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00825 29
39 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.00656
40 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.00590
41 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.00290 $^{(5)}$
42 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00893
43 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00632 46
44 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00510 45
45 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00510 44
46 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00632 43
47 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00563 50
48 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00424 49
49 $ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00424 48
50 $ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00563 47
51 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00441 56
52 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00399 55
53 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00463
54 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00376
states probability sym
----- ----------------------------------------------------------------------------------------------------------- ------------- ----------
55 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00399 52
56 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00441 51
57 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | $ 0.00422 59
58 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | $ 0.00370
59 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | $ 0.00422 57
60 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00348 61
61 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00348 60
62 $ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00401
63 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | $ 0.00355 66
64 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | $ 0.00344 65
65 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | $ 0.00344 64
66 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | $ 0.00355 63
67 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | $ 0.00066 $^{(2)}$
68 $ | \ \ \ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00492
69 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00355 71
70 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00308
71 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00355 69
72 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00313 74
73 $ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00256
74 $ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00313 72
75 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00303
76 $ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00122 $^{(5)}$
77 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00266 79
78 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00277
79 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00266 77
80 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet| \ \ \ | \bullet | \bullet | \bullet | $ 0.00246 85
81 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00249 84
82 $ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00236 83
83 $ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00236 82
84 $ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00249 81
85 $ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | $ 0.00246 80
86 $ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | $ 0.00228
87 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00236
88 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | $ 0.00224
89 $ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | $ 0.00111 $^{(5)}$
90 $ | \ \ \ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00279
91 $ | \ \ \ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00219 92
92 $ | \ \ \ | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00219 91
93 $ | \ \ \ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00195 94
94 $ | \ \ \ | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00195 93
95 $ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00186 96
96 $ | \ \ \ | \bullet | \bullet | \bullet | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00186 95
97 $ | \ \ \ | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00194
98 $ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00178 99
99 $ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00178 98
100 $ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00174
101 $ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | $ 0.00167
102 $ | \ \ \ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00176
103 $ | \ \ \ | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00164
104 $ | \ \ \ | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00152
105 $ | \ \ \ | \bullet | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00147
106 $ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \ \ \ | \bullet | \bullet | \bullet | \bullet | $ 0.00073 $^{(5)}$
107 $ | \ \ \ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00142
108 $ | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | \bullet | $ 0.00014 $^{(1)}$
Acknowledgement {#acknowledgement .unnumbered}
===============
The author would like to express his gratitude to Professors Zbigniew Czechowski, Adam Doliwa and Maciej Wojtkowski for inspiring comments and discussions.
[10]{}
M. Bia[ł]{}ecki and Z. Czechowski. Analytic approach to stochastic cellular automata: exponential and inverse power distributions out of random domino automaton. , 2010.
B. Drossel and F. Schwabl. Self-[O]{}rganized [C]{}ritical [F]{}orest-[F]{}ire [M]{}odel. , 69:1629–1632, 1992.
B. Drossel, S. Clar, and F. Schwabl. Exact [R]{}esults for the [O]{}ne-[D]{}imensional [S]{}elf-[O]{}rganized [C]{}ritical [F]{}orest-[F]{}ire [M]{}odel. , 71:3739–3742, 1993.
B. D. Malamud, G. Morein, and D. L. Turcotte. Forest [F]{}ires: [A]{}n [E]{}xample of [S]{}elf-[O]{}rganized [C]{}ritical [B]{}ehavior. , 281:1840–1842, 1998.
M. Bia[ł]{}ecki. Random domino automaton: from avalanches to rebound parameters. in prep., 2012.
Z. Czechowski and M. Bia[ł]{}ecki. Three-level description of the domino cellular automaton. , 45:155101, 2012.
Z. Czechowski and M. Bia[ł]{}ecki. Ito equations out of domino cellular automaton with efficiency parameters. , 60(3):846–857, 2012.
M. Bia[ł]{}ecki. Motzkin numbers out of random domino automaton. , 2011.
D. Weatherley. Recurrence [I]{}nterval [S]{}tatistics of [C]{}ellular [A]{}utomaton [S]{}eismicity [M]{}odels. , 163:1933–1947, 2006.
T. Parsons. Monte [C]{}arlo method for determining earthquake recurrence parameters from short paleoseismic catalogs: [E]{}xample calculations for california. , 113:B03302 (14pp), 2008.
M. Vazquez-Prada, A. Gonzalez, J. B. Gomez, and A. F. Pacheco. A minimalist model of characteristic earthquakes. , 9:513–519, 2002.
A. Tejedor, S. Ambroj, J. B. Gomez, and A. F. Pacheco. Predictability of the large relaxations in a cellular automaton model. , 41:375102 (16pp), 2008.
M. Bia[ł]{}ecki and Z. Czechowski. On a simple stochastic cellular automaton with avalanches: simulation and analytical results. In V. De Rubeis, Z. Czechowski, and R. Teisseyre, editors, [ *Synchronization and triggering: from fracture to earthquake processes*]{}, pages 63–75. Springer, 2010.
M. Paczuski and P. Bak. Theory of the one-dimensional forest-fire model. , 48:R3214–R3216, 1993.
[^1]: Other states differ by translations.
[^2]: Other states differ by translations.
[^3]: The system stays in fully occupied state $1/\mu_{10}\approx107,5$, which is longer than $100$ as in $\mu=const.$ case (previous column). The respective average times for filling up the lattice are $\approx 19$ and $\approx16$.
[^4]: $(1-{n_1^0}/{n})\approx0.679229$
[^5]: Results from [@BiaCzAA]
[^6]: Other states differ by shifts.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The notion of *guessing cost* (also referred to as cost of guessing) is introduced and an optimal strategy for the $\rho$-th moment of guessing cost is provided for a random variable taking values on a finite set whereby each choice is associated with a positive finite cost value. Moreover, we drive asymptotically tight upper and lower bounds on the moments of cost of guessing problem. Similar to previous studies on the standard guesswork, established bounds on the moments of guessing cost quantify the accumulated cost of guesses required for correctly identifying the unknown choice and are expressed in terms of the Rényi’s entropy. Moreover, a new random variable is introduced to establish connections between the guessing cost and the standard guesswork. Based on this observation, further improved bounds are conjectured in the non-asymptotic region by borrowing ideas from recent works given for the standard guesswork. Finally, we establish the *guessing cost exponent* in terms of Rényi *entropy rate* on the moments of the optimal guessing cost by considering a sequence of independent random variables. Finally, these bounds are shown to be very useful for bounding the overall repair bandwidth for distributed data storage systems. Our particular application utilizes a sparse graph code (low density parity check codes) in conjunction with a back-up master which is pretty common in cellular networks.'
author:
- 'Suayb S. Arslan, and Elif Haytaoglu, [^1][^2][^3]'
title: 'Guessing Cost: Applications to Distributed Data Storage and Repair in Cellular Networks'
---
[Shell : Bare Demo of IEEEtran.cls for IEEE Journals]{}
Guessing, entropy, moments, bounds, cellular networks, sparse graph codes, repair bandwidth.
Introduction
============
typical guessing framework involves finding the value of a realization of a random variable $X$ from a finite or countable infinite set $\mathcal{X}$ by asking a series of questions “Is $X$ equal to $x \in \mathcal{X}$?” until the answer becomes “Yes”. In fact, the ultimate objective of guessing is to find the distribution of the number of such questions (guesses). In an attempt to optimize the order of these questions, an optimal guessing strategy i.e., a bijective function from $\mathcal{X}$ to a finite or countably infinite set $[|\mathcal{X}|] \triangleq \{1,\dots,|\mathcal{X}|\}$ is adapted to typically minimize the average number of guesses, also known as the average guessing number. In [@pliam99], this problem is named as *Guesswork* and appropriate lower and upper bounds are investigated on the guessing number in terms of Shannon’s entropy both by Massey [@massey1] and McElice and Yu [@mceliceyu] for the first time. Later, a sequence of independent and identically distributed random variables $X_1,\dots,X_n$ are considered for practical applications and asymptotically tight bounds are derived on the moments of the expected number of guesses for a typical guesswork [@arikan1996]. This study has related the asymptotic exponent of the best achievable guessing moment to the Rényi’s entropy. In fact, Rényi’s entropy was a frequently used information measure in different contexts such as source coding to be able to generalize coding theorems in the past [@cambell56]. Such findings are successfully applied to various recent applications of data compression [@kuzuoka2019], channel coding [@duffy2019], networking and data storage security [@bracher2019] by tweaking the original problem so that it fits within the requirements of the application at hand.
In many practical cases, making a guess about the unknown value of a random variable (even in presence of a side information [@arikan1996]) leads to a certain amount of cost. Therefore, making a choice among multiple possibilities may lead to different types and amounts of costs overall where we refer it as the *cost of guessing* or simply *guessing cost*. In fact, these costs may dynamically be changing after making subsequent guesses about a series of random variables $\{X_i\}_{i=1}^n$ distributed identically but not independently. Independent and identically distributed random variable case is thoroughly studied in the literature and some extensions to Markovian dependencies are also considered [@malone2004]. To our best of knowledge, the cost of guessing is only mentioned recently in [@kuzuoka2019] in a limited context whereby the guesser is allowed to stop guessing and declare an error and only then a fixed amount of cost is applied, otherwise the mechanism is identical to the standard guesswork. Applications of guessing are abundant. For instance, it is shown that the cut-off rate of sequential decoding can easily be characterized if guessing is applied to the general idea of decoding of a tree code [@arikan1996]. Moreover, capacity-achieving maximum likelihood decoding algorithms are developed in a communication context based on guessing [@duffy2019].
In distributed systems, cost of data communication depends on the link loads, node availabilities and current traffic at the time of communication etc. These costs can be expressed in terms of latency, bandwidth used to transfer information or computation complexity. Inspired from these observations in this study, we proposed a generalization of the guessing framework, namely guessing cost, and derive optimal guessing strategies as well as asymptotically tight bounds by using a quantity related to the Rényi’s entropy. Note that numerical calculation of moments of guessing cost might be computationally efficient for small $|\mathcal{X}|$, however as we consider a series of random variables $\{X_i\}_{i=1}^n$, finding the optimal solution as well as the moments would be computationally intractable (exponential in $n$). We also have shown that guessing cost of a sequence of independent random variables (not necessarily identically distributed) can be expressed in terms of Rényi entropy rate. On the other hand, the computation of the moment of guessing cost for independent $\{X_i\}_{i=1}^n$ would only be linear in $n$. Our results are asymptotically tight i.e., as $n \rightarrow \infty$ we characterize the exponential growth rate of the moments of guessing cost. Several improved bounds are conjectured for the non-asymptotic region based on an established connection with the standard guesswork. Finally, we have observed that our findings for guessing cost can readily be applied to a distributed data storage scenario in new generation mobile networks where nodes are repaired in case of node failures or unexpected node departures from the cell coverage using graph-based codes such as low density parity check (LDPC) codes [@LDPC1963].
The rest of the paper is organized as follows. In section II, the problem is formally stated and conditions are laid out for an optimal guessing strategy that minimizes the moments of guessing cost. In addition distinct determination times for costs are discussed along with an algorithm that finds the best strategy. In section III, asymptotically tight upper and lower bounds are given. While deriving the upper bound, connection with the standard guesswork is established. This connection helped us use previous works to find tighter bounds in the non-asymptotic region. In section IV, a distributed storage scenario is considered where a long block LDPC code is utilized. It is shown that the data repair problem of an LDPC code can be considered within the context of cost of guessing. Several numerical results are provided before we conclude our paper in section V.
Problem Statement and Guessing Strategy
=======================================
Let us use $C_\mathcal{G}(x)$ to denote the total cost of guessing required by a particular guessing strategy $\mathcal{G}$ when $X = x$. If the cost of making each guess $X = x$ is independent of other guesses and amounts to 1 (unity), then this problem would be the same as the characterization of the average number guesses (average guessing number) and is identical to Massey’s original guessing problem [@massey1]. Later, bounds on the moments of optimal guessing are derived [@arikan1996] and improved [@Bozdas1997], [@sason2018]. Particularly, the relationship between Rényi’s entropy and expected guessing number is interesting and useful in different engineering contexts.
Let us assume that the random variable $X$ can take on values from a finite set $\mathcal{X} = \{x_1,\dots,x_M\}$ according to a distribution $P_X(x)$ with the finite set of costs $\mathcal{C} = \{c_{x_1},\dots,c_{x_M}\}$. Without loss of generality, sets are assumed to have cardinalities $|\mathcal{X}|=|\mathcal{C}|=M$ in which using a particular guessing strategy $\mathcal{G}$, the probability that a randomly selected element of $\mathcal{X}$ can be found in the $i$-th guess is $p_i = P_X({\mathcal{G}^{-1}(i)})$ with cost $c_i = c_{\mathcal{G}^{-1}(i)}$, independently of already made guesses. Then, the average cost of guessing can be expressed as follows $$\begin{aligned}
\mathbb{E}[C_\mathcal{G}(X)] = \sum_{i=1}^M \sum_{j=1}^{i} c_j p_i = \sum_{i=1}^M f_i p_i = \sum_{i=1}^M c_i \left(1 - g_{i-1} \right)\end{aligned}$$ where $f_i=\sum_{j=1}^{i} c_j$ and $g_{i} = \sum_{j=1}^{i}p_j$ are cumulative cost and probability distributions. The minimization of this value is a function of both guessing strategy $\mathcal{G}$ and the probability distribution of $X$.
One of the classical questions about guessing is the best strategy that would minimize $\mathbb{E}[C_\mathcal{G}(X)]$ i.e., $\mathcal{G}^* = \operatorname*{arg\,min}_{\mathcal{G}} \mathbb{E}[C_\mathcal{G}(X)]$. In case of $c_i = 1, \forall i \in [M]$, the strategy is simple i.e., guess the possible values of $X$ in the order of non-increasing probabilities [@massey1]. In other words, without loss of generality, we can assume $\mathbf{p} = (p_1,p_2,\dots,p_M)$ with $p_1 \geq p_2 \geq \dots \geq p_{M}$ are the probabilities of choosing values from $[M]$ and $\mathcal{G}(x_i) = i$. Then with this choice, $\sum_i i p_i$ would be minimized. However, the same conclusion cloud not be drawn where an arbitrary vector of costs $\mathbf{c}
^{(i)} = (c_1, \dots, c_i)$ is present for any $i\in \{1,\dots,M\}$. Let us consider two possible scenarios.
Configurable Costs Determined A Posteriori
------------------------------------------
In this particular scenario, although the cost vector is given i.e., $\{c_i\}$, the assignments are not made i.e., costs can be associated with each choice as it fits in the beginning. In this case, the best strategy is to guess the possible values of $X$ in the order of non-increasing probabilities and associate the more probable choice with the least cost value. In other words, for the assignment (a permutation of $\mathbf{c}^{(M)}$) $\Tilde{\mathbf{c}}^{(M)} = (\tilde{c}_1,\tilde{c}_2,\dots,\tilde{c}_M)$ with $\tilde{c}_1 \leq \tilde{c}_2 \leq \dots \leq \tilde{c}_M$ and $\tilde{c}_i \in \mathbf{c}$, it is easy to see that for any $\rho > 0$, we have $$\begin{aligned}
\sum_{i=1}^M \left( \sum_{j=1}^{i} c_j \right)^\rho p_i \geq \sum_{i=1}^M \left( \sum_{j=1}^{i} \tilde{c}_j \right)^\rho p_i.\end{aligned}$$
In case the cumulative costs are given by the moments of the guessing number i.e., $f_{i}=i^\rho$ for any $\rho \geq 1$, then it is easy to see that $c_i = i^\rho - (i-1)^\rho$ which implies that $c_1 \leq c_2 \leq \dots \leq c_M$ is satisfied. Thus, the best strategy would again be to guess the possible values of $X$ in the order of non-increasing probabilities as argued in [@arikan1996].
Non-configurable Costs Determined A Priori
------------------------------------------
In this case, cost associated with each choice is determined externally. In other words, costs and choices are bound and determined prior to guessing. In this case, the best strategy would not necessarily be guessing the possible values of $X$ in the order of non-increasing probabilities. Consider for instance three choices i.e., $M=3$ with $(1,p_1=0.5,c_1=20)$, $(2,p_2=0.4,c_2=2)$ and $(3,p_3=0.1,c_3=1)$. In that case the guessing order $(2,3,1)$ would be preferable compared to the order $(1,2,3)$ with the average costs 12.6 and 21.1, respectively, where the latter choice, based on the order of non-increasing probabilities, is clearly not optimal. The following proposition establishes a necessary condition for the optimal guessing strategy $\mathcal{G}^*$.
\[prop0\] For a given $\rho > 0$ and an optimal guessing strategy, namely $\mathcal{G}^*$ for the $\rho$-th moment of guessing cost, we have the following necessary condition for all $i, j \in\{1,\dots,M\}$ satisfying $i \leq j$. $$\begin{aligned}
\frac{\left(c_i\sum_{k=1}^j c_k\right)^\rho}{\sum_{k=1}^j c_k^\rho} p_j \leq \frac{\left(c_j\sum_{k=1}^i c_k\right)^\rho}{\sum_{k=1}^i c_k^\rho} p_i \label{condition_new}\end{aligned}$$ Furthermore, let $\mu_k^{1/\rho} = \frac{\sum_j^k c_j}{ || \mathbf{c}^{(k)} ||_{\rho}}$ for all $j \leq k$. if the condition $\mu_{k-1} \leq \mu_{k}$ is also satisfied for $k>1$, then the necessary condition can be simplified to $c_i^\rho p_j \leq c_j^\rho p_i$.
\[prop1\] For a given optimal guessing strategy for the mean guessing cost ($\rho=1$), denoted as $\mathcal{G}^*$, we must have $c_{i}p_{j} \leq c_{j}p_{i}$ for all $i, j \in\{1,\dots,M\}$ satisfying $i \leq j$.
Let us start with $\rho=1$ i.e., mean guessing cost mentioned in the corollary \[prop1\]. Consider swapping the $i$-th and $(i+1)$-th guessed values. Let $\mathcal{G}_{i,i+1}$ be the original guessing strategy and $\mathcal{G}_{i+1,i}$ be the swapped version. Then it is straightforward to show that the difference is $$\begin{aligned}
\mathbb{E}[C_{\mathcal{G}_{i,i+1}}(x)] - \mathbb{E}[C_{\mathcal{G}_{i+1,i}}(x)] &= c_i(1-g_{i-1}) \\
&+ c_{i+1}(1-g_{i-1}-p_i) \\
&- c_{i+1}(1-g_{i-1}) \\
&- c_i(1-g_{i-1}-p_{i+1}) \\
&= c_ip_{i+1} - c_{i+1}p_i\end{aligned}$$ which implies that if $c_ip_{i+1} > c_{i+1}p_i $, then we swap $i$-th and $(i+1)$-th guessed values in order to reduce the average cost of guessing, otherwise no swapping is performed. Since each swapping leads to lower cost, for any $i, j \in\{1,\dots,M\}$ with $i \leq j$, the optimal guessing strategy $\mathcal{G}^*$ would satisfy the following series of inequalities $$\begin{array}{ll}
c_ip_{i+1} &\leq c_{i+1}p_i \\
c_{i+1}p_{i+2} &\leq c_{i+2}p_{i+1} \\ &\vdots \\
c_{j-1}p_{j} &\leq c_{j}p_{j-1}
\end{array} {\bBigg@{5}}\} \Rightarrow c_i p_j
\prod_{k=i+1}^{j-1} c_kp_k \leq c_j p_i
\prod_{k=i+1}^{j-1} c_kp_k$$ where continuing deriving the inequality and multiplying left-hand and right-hand consecutive terms individually would give us the desired result since all $p_i$s and $c_i$s are non-negative.
Now, let us consider the general case i.e., for any real $\rho>0$, we have $$\begin{aligned}
\mathbb{E}[C_\mathcal{G}(X)^\rho] &= \sum_{i=1}^M \left(\sum_{j=1}^{i} c_j\right)^\rho p_i \\ &= \sum_{i=1}^M \sum_{j=1}^{i} \widetilde{c}_j^\rho p_i =
\sum_{i=1}^M f_i^\rho \widetilde{p}_i = \sum_{i=1}^M c_i^\rho \left(1 - \widetilde{g}_{i-1} \right) \label{newtildaexp}\end{aligned}$$ where for a given $k$, $\widetilde{c}_j = \mu_k^{1/\rho} c_j$ for all $j \leq k$ and some positive real constant $\mu_k^{1/\rho} = \frac{\sum_j^k c_j}{ || \mathbf{c}^{(k)} ||_{\rho}}$. Also, $\widetilde{p}_i = \mu_ip_i$ and $\widetilde{g}_i = \sum_{j=1}^i \widetilde{p}_j$. Since due to being able to express the $\rho$-th moment as in , we can apply the Corollary 2.1.1 for all $i,j \in \{1,2,\dots,M\}$ satisfying $i \leq j$ to come up with the following necessary condition $$\begin{aligned}
c_i^\rho \mu_j p_j = c_i^\rho \widetilde{p}_j \leq c_j^\rho \widetilde{p}_i = c_j^\rho \mu_i p_i \end{aligned}$$ which is identical to the condition given in . Note that if defined real constants satisfy the following set ordering due to increasing number of cross terms with appropriate cost values $$\begin{aligned}
1 = \mu_1 \leq \mu_2 \leq \dots \leq \mu_{M} \label{muorder}\end{aligned}$$ then, we shall also have the inequality $c_i^\rho \mu_i p_j \leq c_i^\rho \mu_j p_j$ since $\mu_i \leq \mu_j$ for all $i\leq j$. Thus, the more simplified inequality $c_i^\rho p_j \leq c_j^\rho p_i $ would follow.
Note that $c_ip_j \leq c_jp_i$ does not necessarily imply the condition in (and hence the condition $c_i^\rho p_j \leq c_j^\rho p_i$) for any $\rho>0$. Thus, this general necessary condition is more strict in the sense that the strategy that is satisfying $c_ip_j \leq c_jp_i$ (condition in with $\rho=1$) would be an upper bound on the optimal average guessing cost. For instance, if for all $j \geq i$, $$\begin{aligned}
\sum_{k=1}^j \frac{c_k}{\left(\sum_{k=1}^j c_k^\rho\right)^\frac{1}{\rho}} \geq \sum_{k=1}^i \frac{c_k}{\left(\sum_{k=1}^i c_k^\rho\right)^\frac{1}{\rho}}\end{aligned}$$ is satisfied (which typically holds), then $c_ip_j \leq c_jp_i$ needs to hold otherwise, the inequality would not hold. Hence, our argument in Eqn. is still valid since we are generating an upper bound for the optimal guessing strategy. However with the new condition the upper bound can be tightened.
We note that there may be more than one optimal guessing strategy that would satisfy the condition of Corollary \[prop1\]. However all of them will result in minimum average cost. Hence it is sufficient to find one of such guessing strategies that this necessary condition would hold.
In observation of Proposition \[prop0\], let us provide an algorithmic solution to finding optimal guessing cost. We notice that if the swapping in proposition is executed within a Bubble-sort[^4] style for the given strategy, the convergence to an optimal guessing would be guaranteed and we can find the optimal solution with the best and the worst time complexities of $\Omega(M)$ and $\Theta(M^2)$, respectively. The naive algorithm for finding an optimal cost of guessing is provided in Algorithm \[alg1\] where swap(.,.) function swaps the entries of a given array in the argument. Consider three different items, $i,j,z$ out of $n$ such that $i\leq j \leq z \leq n$. Due to the proposition \[prop0\] with $\rho = 1$, we have $c_ip_j \leq c_jp_i$ and $c_jp_z \leq c_zp_j$. If so, Since, $c_i\frac{p_j}{p_i}p_z\leq c_z p_j$then the inequality $c_ip_z\leq c_zp_i$ must also be true. In other words, the inequality has the transitive property. Thus, the optimal cost of guessing order can be achieved by applying Merge-sort or Heap-sort on $c_ip_j$’s, which results $\Theta(nlogn)$ as the worst and the average case complexities. In the next section, we focus on the moments of the cost of guessing whereby the average cost would a special case. Furthermore, lower and upper bounds are derived in terms of a popular information theoretic measure, namely Rényi’s entropy.
\[alg1\]
**function** **OptimalCostGuess**($\mathbf{p}, \mathbf{c}, \rho$)
$M \gets |\mathbf{p}|$ $\mathcal{I} \gets \lbrace 1,2,3...,M\rbrace$ $swapped \gets true$ $i \gets 1$ $swapped \gets false$ swap($c_{j},c_{j+1}$),swap($p_{j},p_{j+1}$),swap($\mathcal{I}_j,\mathcal{I}_{j+1})$ $swapped \gets true$ $i \gets i+1$ return $\mathcal{I}$
Bounds on Moments of the Cost of Guessing
=========================================
Throughout this section, we assume static costs determined a priori and focus on moments of guessing as the average cost of guessing would be a special case.
Lower and Upper Bounds
----------------------
Let $P_X(x)$ to denote the probability distribution of $X$ and define the moments of the cost of guessing using a particular guessing function $\mathcal{G}$ as $$\begin{aligned}
\mathbb{E}[C_\mathcal{G}(X)^\rho] = \sum_i P_X({\mathcal{G}^{-1}(i)}) \left[\sum_j^i c_{\mathcal{G}^{-1}(j)}\right]^\rho \label{orgeqn}\end{aligned}$$ where the costs are not necessarily integers. Let us use the previous notation $c_i = c_{\mathcal{G}^{-1}(i)}$ and define $\mathbf{c}^*=\{c^*_1,c^*_{2},\dots,c^*_M\}$ which is the order obtained by running Algorithm 1 for optimal guessing strategy $\mathcal{G}^*$. This shall be useful in expressing the lower and upper bounds in the following.
\[thm31\] For any guessing function $\mathcal{G}$, $\rho > 0$ and costs $c_j>1$, $\rho$-th moment of the cost of guessing is lower bounded by $$\begin{aligned}
\mathbb{E}[C_\mathcal{G}(X)^\rho] \geq \mathbb{E}[C_{\mathcal{G}^*}(X)^\rho]
\geq \left(\frac{M}{1+\gamma^*}\right)^{-\rho} \exp\left\{\rho H_{\frac{1}{1+\rho}}(X) \right\} \label{eqn11}\end{aligned}$$ where $\gamma^*$ is the harmonic mean of $\{\sum_j^i c^*_j-1\}'s$ for $i=\{1,2,\dots,M\}$ and $H_{\alpha}(X)$ is Rényi’s entropy of order $\alpha$.
Before giving the formal proof let us state a well known lemma.
Let $a_i$ and $b_i$ for $(i = 1, . . . , n)$ be positive real sequences. If $q > 1$ and $1/q + 1/r = 1$, then $$\begin{aligned}
\left(\sum_{i=1}^n a_i^q\right)^{1/q}\left(\sum_{i=1}^n b_i^r\right)^{1/r} \geq \sum_{i=1}^{n} a_ib_i\end{aligned}$$
Let $a_i$ be a positive real number for all $i$, $M$ be a natural number, and $\gamma$ be the harmonic mean of $\{a_1,\dots,a_n\}$, then we have $$\begin{aligned}
\sum_{i=1}^M \frac{1}{1+a_i} \leq \frac{M}{1+\gamma} \label{eqn111}\end{aligned}$$ which can easily be proved using Radon’s inequality [@lai2016]. Now, let us express the lower bound of the moments of the cost of guessing as follows, $$\begin{aligned}
\mathbb{E}[C_\mathcal{G}(X)^\rho] \geq \mathbb{E}[C_\mathcal{G^*}(X)^\rho]
\geq \left[\sum_i \frac{1}{\sum_j^i c_j^*}\right]^{-\rho} \left[\sum_i P_X({\mathcal{G}^{-1}(i)})^\frac{1}{1+\rho}\right]^{1+\rho} \label{eqn133}\end{aligned}$$ which easily follows from a direct application of Hölder’s inequality. To see this, let us set $r=1+ \rho$, $q=(1+\rho)/\rho$ in Hölder’s inequality so that $1/q+1/r =1$ is satisfied for $\rho > 0$. We also let $$\begin{aligned}
a_i = \left[\sum_j^i c_{\mathcal{G}^{-1}(j)}\right]^{-\rho/(1+\rho)} \mathrm{and} \ \ \ \
b_i = \left[\sum_j^i c_{\mathcal{G}^{-1}(j)}\right]^{\rho/(1+\rho)}P_X({\mathcal{G}^{-1}(i)})^{1/(1+\rho)}.\end{aligned}$$ Now, using Hölder’s inequality, it would be easy to obtain $$\begin{aligned}
\left[ \sum_i \frac{1}{\sum_j^ic_{\mathcal{G}^{-1}(j)}}\right]^{\rho/(1+\rho)} \left(\mathbb{E}[C_\mathcal{G}(X)^\rho]\right)^{1/(1+\rho)} \geq \sum_i P_X({\mathcal{G}^{-1}(i)})^{1/(1+\rho)}\end{aligned}$$ from which inequality follows for the optimal strategy $\mathcal{G^*}$.
Now, considering the ordering of costs that minimizes the right hand side, we shall have, $$\begin{aligned}
\mathbb{E}[C_\mathcal{G}(X)^\rho] &\geq \left[\sum_i \frac{1}{\sum_j^i c^*_j}\right]^{-\rho} \left[\sum_i P_X({\mathcal{G}^{-1}(i)})^\frac{1}{1+\rho}\right]^{1+\rho} \nonumber \\
&\geq \left(\frac{M}{1+\gamma^*}\right)^{-\rho} \left[\sum_i P_X(x_i)^\frac{1}{1+\rho}\right]^{1+\rho} \label{eqn14} = \left(\frac{M}{1+\gamma^*}\right)^{-\rho} \exp\left\{\rho H_{\frac{1}{1+\rho}}(X) \right\} \end{aligned}$$ where $\gamma^*$ is the harmonic mean of $\{\sum_j^i c^*_j-1\}'s$ for $i=\{1,2,\dots,M\}$ and $H_{\alpha}(X)$ is Rényi’s entropy of order $\alpha$ ($\alpha>0, \alpha \not=1$) for random variable $X$ defined as, $$\begin{aligned}
H_\alpha(X) = \frac{\alpha}{1-\alpha}\ln \left[\sum_x P_X(X)^\alpha \right]^{1/\alpha}\end{aligned}$$
Note that inequality followed from the inequality .
Let us demonstrate that the bound given in Theorem \[thm31\] is tight within a factor of $\left( M / (1+\gamma^*) \right)^\rho$.
\[thm31\] For the optimal guessing function $\mathcal{G}^*$, and $\rho \geq 0$, $\rho$-th moment of the cost of guessing is upper bounded by $$\begin{aligned}
\mathbb{E}[C_\mathcal{G^*}(X)^\rho] \leq \exp\{ \rho H_{\frac{1}{1+\rho}}(Y)\} \label{eqn2222}\end{aligned}$$ where the random variable $Y$ is defined to take on values from a finite set $\mathcal{Y}=\{y_1,y_2,\dots,y_{\sum_x^M \lceil c_x \rceil }\}$ with probabilities $P_Y(y) = P_X(x)/ \lceil c_x \rceil$ for all $c_x > 1$ and $y$ satisfying $$\begin{aligned}
\sum_{x^\prime}^{x-1} \lceil c_{x^\prime} \rceil < y \leq \sum_{x^\prime}^{x} \lceil c_{x^\prime} \rceil \end{aligned}$$ and $H_{\alpha}(X)$ is Rényi’s entropy of order $\alpha$.
Let us first observe that with the optimal guessing strategy $\mathcal{G}^*$ that minimizes the expected cost of guessing $x$, $$\begin{aligned}
C_{\mathcal{G}^*}(x) &=& \sum_{x^\prime:C_{\mathcal{G}^*}(x^\prime) \leq C_{\mathcal{G}^*}(x)} \sum_{x^{\prime \prime}}^{c_{x^\prime}} 1\\
&\leq& \sum_{x^\prime:C_{\mathcal{G}^*}(x^\prime) \leq C_{\mathcal{G}^*}(x)} \sum_{x^{\prime \prime}}^{c_{x^\prime}} \left(\frac{c_xP_X(x^\prime)}{c_{x^\prime}P_X(x)} \right)^{\frac{1}{1+\rho}} \label{eqn20} \\
&=& \sum_{x^\prime:C_{\mathcal{G}^*}(x^\prime) \leq C_{\mathcal{G}^*}(x)}
c_{x^\prime}^{\frac{\rho }{1+\rho}}\left(\frac{c_xP_X(x^\prime)}{P_X(x)} \right)^{\frac{1}{1+\rho}} \\
&\leq & \sum_{x^\prime}
c_{x^\prime}^{\frac{\rho}{1+\rho}}\left(\frac{c_xP_X(x^\prime)}{P_X(x)} \right)^{\frac{1}{1+\rho}} \label{eqn222}\end{aligned}$$ where the inequality follows from the necessary condition $c_{x^\prime} P_X(x) \leq c_x P_X(x^{\prime})$ for all $\{x^{\prime}:C_{\mathcal{G}^*}(x^\prime) \leq C_{\mathcal{G}^*}(x)\}$ that needs to hold for the optimal strategy $\mathcal{G}^*$. Also, although the exponent $1/(1+\rho)$ decreases the value, it is still greater than 1 due to ${\frac{c_xP_X(x^\prime)}{c_{x^\prime}P_X(x)}} \geq 1$. Using the inequality given in in equation , we get $$\begin{aligned}
\mathbb{E}[C_\mathcal{G^{*}}(X)^\rho] & = \sum_x P_X(x) C_{\mathcal{G}^*}(x)^\rho \\
& \leq \sum_x P_X(x) \left( \sum_{x^{\prime}} c_{x^\prime}^{\frac{\rho}{1+\rho}} \left(\frac{c_xP_X(x^\prime)}{P_X(x)}\right)^{\frac{1}{1+\rho}}
\right)^\rho = \left[\sum_x c_x^{\frac{\rho}{1+\rho}} P_X(x)^{\frac{1}{1+\rho}}\right]^{1+\rho} \\
& = \left[\sum_x c_x (P_X(x)/c_x)^\frac{1}{1+\rho}\right]^{1+\rho} \label{eqnfinal}
$$
On the other hand, we notice that $$\begin{aligned}
\frac{P_X(x)}{\lceil c_x \rceil} = \frac{c_x P_X(x)}{\lceil c_x \rceil c_x} \geq \frac{P_X(x)}{c_x} \left( \frac{c_x}{\lceil c_x \rceil} \right)^{1+\rho}\end{aligned}$$ from which the following inequality follows for $\rho \geq 0$, $$\begin{aligned}
\lceil c_x \rceil (P_X(x)/\lceil c_x \rceil)^{\frac{1}{1+\rho}} \geq c_x (P_X(x)/c_x)^\frac{1}{1+\rho}. \label{eqn311}\end{aligned}$$
Thus, using the inequality and the defined random variable $Y$ earlier, we finally express the upper bound in a more compact form $$\begin{aligned}
\mathbb{E}[C_\mathcal{G^{*}}(X)^\rho] &\leq \left[\sum_x c_x (P_X(x)/c_x)^\frac{1}{1+\rho}\right]^{1+\rho} \\
& \leq \left[ \sum_x \lceil c_x \rceil (P_X(x)/\lceil c_x \rceil)^{\frac{1}{1+\rho}} \right]^{1+\rho} = \left[ \sum_y P_Y(y)^{\frac{1}{1+\rho}} \right]^{1+\rho} \\
&= \exp\{ \rho H_{\frac{1}{1+\rho}}(Y) \}
$$
Notice that this upper bound will reduce to Arikan’s upper bound i.e., $\exp(\rho H_{\frac{1}{1+\rho}}(X))$ with all costs set to unity.
Relation to Guesswork and Guessing Cost Exponent
------------------------------------------------
Introduction of a random variable $Y$ is useful for establishing a relationship with the standard guesswork. From the earlier discussions on the random variable $Y$, we can express a looser lower bound (compared to ) for any guessing function $\mathcal{G}(.)$ by observing the following for $c_j >0$, $$\begin{aligned}
\mathbb{E}[C_\mathcal{G}(X)^\rho]
& = \sum_{x} P_X(x) C_{\mathcal{G}}(x)^\rho = \sum_i P_X({\mathcal{G}^{-1}(i)}) \left[\sum_j^i c_{\mathcal{G}^{-1}(j)}\right]^\rho \\
& = \sum_i \sum_j^{\lceil c_{\mathcal{G}^{-1}(i)} \rceil} \frac{P_X(\mathcal{G}^{-1}(i))}{\lceil c_{\mathcal{G}^{-1}(i)} \rceil} \left[\sum_j^i c_{\mathcal{G}^{-1}(j)}\right]^\rho \\
& \geq \sum_i \sum_j^{\lceil c_{\mathcal{G}^{-1}(i)} \rceil} \frac{P_X(\mathcal{G}^{-1}(i))}{\lceil c_{\mathcal{G}^{-1}(i)} \rceil} \left(\sum_{k}^{i-1}\lceil c_{\mathcal{G}^{-1}(k)} \rceil+j \right)^\rho \\
& = \mathbb{E}[C_\mathcal{H}(Y)^\rho] \geq \left(1+ \ln \left(\sum_x \lceil c_x \rceil \right)\right)^{-\rho} \exp \left\{ \rho H_{\frac{1}{1+\rho}}(Y) \right\} \label{lowerboundalter}\end{aligned}$$ where the last inequality is due to standard Guesswork and follows directly from [@arikan1996] based on the definition of the random variable $Y$. Better lower bounds can be given, however this loose lower bound is enough to prove the following asymptotically tight result. Here the guessing function $\mathcal{H}(Y)$ for the random variable $Y$ defined earlier is directly induced from $\mathcal{G}(X)$. The guessing cost exponent is given by the following theorem.
\[thmexponent\] Let $\boldsymbol{X}= (X_1,\dots,X_n)$ be a sequence of independent (not necessarily identically distributed) random variables where each is defined over the set $\mathcal{X}_i$ with the associated cost distribution $\mathcal{C}_i$. Let $\mathcal{G}^* (X_1,\dots, X_n)$ be an optimal guessing function for $\boldsymbol{X}$. Then, for any $\rho>0$, we have $$\begin{aligned}
\lim_{n \rightarrow \infty} \frac{1}{n} \ln ( \mathbb{E}[C_{\mathcal{G}^*}(X_1,X_2,\dots,X_n)^\rho])^{1/\rho} =\mathcal{R}_{\frac{1}{1+\rho}}(\{Y_i\}) \end{aligned}$$ where $\mathcal{R}_{\frac{1}{1+\rho}}(.)$ denotes the order-$1/(1+\rho)$ Rényi rate and $Y_i$s are random variables induced from random variables $X_i$s.
Let us consider the general case and define random variables $Y_i \sim Y$ as found in Theorem \[thm31\] for the corresponding random variables $X_i$ for $i=\{1,\dots,n\}$ each with cost distributions $\mathcal{C}_i$. Also let $\mathcal{H}^*$ be the induced optimal guessing strategy from $\mathcal{G}^*$ for random variables $\{Y_1, \dots, Y_n\}$. Now, consider the upper bound for i.i.d. random variables and observe $$\begin{aligned}
C_{\mathcal{G}^*}(x_1,\dots,x_n)
&\leq& \sum_{\substack{
x_1^\prime, x_2^\prime, \dots, x_n^\prime: \\
C_{\mathcal{G}^*}(x_1^\prime,x_2^\prime,\dots,x_n^\prime) \leq C_{\mathcal{G}^*}(x_1,x_2,\dots,x_n) } }
\sum_{x_1^{''}}^{c_{x_1^\prime}} \dots \sum_{x_n^{\prime \prime}}^{c_{x_n^\prime}} \left(\prod_i \frac{c_{x_i}P_{X_i}(x_i^\prime)}{c_{x_i^\prime}P_{X_i}(x_i)} \right)^{\frac{1}{1+\rho}} \\
&\leq& \prod_i \sum_{x_i^\prime}
c_{x_i^\prime}^{\frac{\rho}{1+\rho}}\left(\frac{c_{x_i}P_{X_i}(x_i^\prime)}{P_{X_i}(x_i)} \right)^{\frac{1}{1+\rho}} = \left[ \sum_{x_1^\prime}
c_{x_1^\prime}^{\frac{\rho}{1+\rho}}\left(\frac{c_{x_1}P_{X_1}(x_1^\prime)}{P_{X_1}(x_1)} \right)^{\frac{1}{1+\rho}} \right]^n\end{aligned}$$ due to independence and series of inequalities $c_{x_1^\prime} P_X(x_1) \leq c_{x_1} P_X(x_1^{\prime}), c_{x_2^\prime} P_X(x_2) \leq c_{x_2} P_X(x_2^{\prime}), \dots, c_{x_n^\prime} P_X(x_n) \leq c_{x_n} P_X(x_n^{\prime})$ for all $\{x_i^{\prime}:C_{\mathcal{G}^*}(x_i^\prime) \leq C_{\mathcal{G}^*}(x_i)\}$ where $i=1,\dots,n$ that needs to hold for the optimal strategy $\mathcal{G}^*$ required by the necessary condition. Finally, we can upper bound the expected guessing cost for a sequence of i.i.d. random variables as $$\begin{aligned}
\mathbb{E}[C_{\mathcal{G}^*}(X_1,\dots,X_n)^\rho] & = \sum_x P_{\boldsymbol{X}}(x_1,\dots,x_n) C_{\mathcal{G}^*}(x_1,\dots,x_n)^\rho \\
& \leq \prod_i \sum_{x_i} P_{X_i}(x_i) \left[\sum_{x_i^\prime}
c_{x_i^\prime}^{\frac{\rho}{1+\rho}}\left(\frac{c_{x_i}P_{X_i}(x_i^\prime)}{P_{X_i}(x_i)} \right)^{\frac{1}{1+\rho}}\right]^\rho = \prod_i \left[\sum_{x_i} c_{x_i} (P_{X_i}(x_i)/c_{x_i})^\frac{1}{1+\rho}\right]^{(1+\rho)}
\\
& \leq \prod_i \left[ \sum_{y_i} P_{Y_i}(y_i)^{\frac{1}{1+\rho}} \right]^{(1+\rho)} = \exp\left\{ \rho \sum_{i} H_{\frac{1}{1+\rho}}(Y_i) \right\} \label{eqn39}\end{aligned}$$ where the last inequality follows due to inequalities similar to for each random variable $X_i$. If the cost and probability distributions of $X_i$’s are arranged such that the induced $Y_i$’s are identically distributed (for instance $X_i$’s are i.i.d. with identical cost distributions i.e., $\mathcal{C}_1 \equiv \mathcal{C}_2 \equiv \dots \equiv \mathcal{C}_n \triangleq \mathcal{C}$) then we can further simplify as $$\begin{aligned}
\mathbb{E}[C_{\mathcal{G}^*}(X_1,\dots,X_n)^\rho] \leq \exp\{ \rho n H_{\frac{1}{1+\rho}}(Y_1) \} \label{finupb}\end{aligned}$$
In addition to this upper bound, we can extend the lower bound given in for a sequence of random variables as $$\begin{aligned}
\mathbb{E}[C_{\mathcal{G}^*}(X_1,\dots,X_n)^\rho] \geq \mathbb{E}[C_{\mathcal{H}^*}(Y_1,\dots,Y_n)^\rho] \geq \left(1+ \ln \left( \prod_i \sum_{x_i} \lceil c_{x_i} \rceil \right)\right)^{-\rho} \exp \left\{ \rho \sum_i H_{\frac{1}{1+\rho}}(Y_i) \right\} \label{eqn47}\end{aligned}$$ where the first inequality can be shown to be true through induction and the second inequality follows from [@arikan1996] through a bit of generalization. Note that $\mathcal{H}^*$ is the optimal induced strategy from $\mathcal{G}^*$. As a consequence, using we have $$\begin{aligned}
\lim_{n \rightarrow \infty} \frac{1}{n} \ln ( \mathbb{E}[C_{\mathcal{G}^*}(X_1,X_2,\dots,X_n)^\rho])^{1/\rho} &\geq \lim_{n \rightarrow \infty} \ln \left(1 + \ln \left(\prod_i \sum_{x_i} \lceil c_{x_i} \rceil \right) \right)^{-1/n} + \lim_{n \rightarrow \infty} \frac{1}{n} \sum_i H_{\frac{1}{1+\rho}}(Y_i) \label{eqn49} \\
&= \lim_{n \rightarrow \infty} \frac{1}{n} \sum_i H_{\frac{1}{1+\rho}}(Y_i) = \mathcal{R}_{\frac{1}{1+\rho}}(\{Y_i\}) \label{finaleqn}\end{aligned}$$ which is defined to be the order-$1/(1+\rho)$ Rényi entropy rate [@Bunte2016] as long as the limit exists. If $\{X_i\}$ are indentically distributed with the same cost distribution $\mathcal{C}$, then this rate would be equal to $H_{\frac{1}{1+\rho}}(Y)$. Combining equations with , the intended equality result follows.
These results indicate that the complexity of guessing cost of a random variable $X$ with strategy $\mathcal{G}$ can be tied to the complexity of guessing another random variable $Y$ with induced strategy $\mathcal{H}$ based on the cost distribution $\mathcal{C}$ as defined earlier.
Improved Bounds: Non-asymptotic regime
--------------------------------------
One of the observations is that the provided bounds have the potential for improvement particularly in the non-asymptotic regime similar in spirit to works such as [@Bozdas1997] and [@sason2018]. These improvements can easily be made after we recognize the relationship between cost of guessing and the standard Guesswork. In particular we have the following improved lower bounds that show better performance in the non-asymptotic regime, $$\begin{aligned}
\mathbb{E}[C_\mathcal{G}(X)^\rho] &\geq \sup_{\beta \in (-\rho,\infty)-\{0\}} \exp\left\{\frac{\rho}{\beta}\left[H_{\frac{\beta}{\beta+\rho}}(Y) - \log u_{\sum_x \lceil c_x \rceil} (\beta) \right]\right\} \\
&= \sup_{\beta \in (-\rho,\infty)-\{0\}} \left[u_{\sum_x \lceil c_x \rceil} (\beta)\right]^{-\frac{\rho}{\beta}}\exp\left(\frac{\rho}{\beta}H_{\frac{\beta}{\beta+\rho}} (Y)\right) \label{sasonlb}\end{aligned}$$
where $u_{\sum_x \lceil c_x \rceil}(\beta)$ is as defined in [@sason2018] with $M$ is replaced with $\sum_x \lceil c_x \rceil$. The proof is quite similar and therefore omitted. As for the upper bound, the following conjecture can be made $$\begin{aligned}
\mathbb{E}[C_\mathcal{G^*}(X)^\rho]
\leq \frac{1}{1+\rho} \left[\exp \left\{ \rho H_{\frac{1}{1+\rho}}(Y)\right\}-1\right] + \exp \left\{(\rho-1)^+H_{1/\rho}(Y)\right\} \label{eqnupimp}\end{aligned}$$ where $(z)^+ \triangleq \max\{z, 0\}$ for $z \in \mathbb{R}$. As shall be seen, we can tighten up the upper bound with this conjecture, particularly for small $\rho$. In fact, this bound can be tightened for different ranges of $\rho$ as follows, $$\label{lowerboundlast} \mathbb{E}[C_\mathcal{G^*}(X)^\rho]
\leq
\begin{cases}
\frac{1}{1+\rho} \exp \left\{\rho H_{1/(1+\rho)}(Y)\right\} + \frac{\rho-(1-\rho)(2^\rho-1)(1-p_{max})}{1+\rho} & \text{for } 0 \leq \rho \leq 1\\
\frac{1}{1+\rho} \exp \left\{\rho H_{1/(1+\rho)}(Y)\right\} + \frac{1}{\rho} \exp \left((\rho-1)H_{1/\rho}(Y)\right) + \frac{\rho^2-\rho+1}{\rho(1+\rho)} & \text{for } 1 \leq \rho \leq 2\\
1 + \sum_{j=0}^{\lfloor\rho\rfloor}\xi_j(\rho) \left[\exp\left((\rho-j)H_{\frac{1}{1+\rho-j}}(Y)\right)-1\right] & \text{for } \rho \geq 2
\end{cases}$$ where $\{\xi_j(\rho)\}$ is given by $$\label{cjs}
\xi_j(\rho)
=
\begin{cases}
\frac{1}{1+\rho} & \text{for } j = 0 \\
\frac{1}{2} & \text{for } j = 1 \\
\frac{\rho \dots (\rho-j+2)}{2^j} & \text{for } j \in \{2,3,\dots,\lfloor\rho\rfloor-1\} \\
\frac{\rho \dots (\rho-j+2)}{2^{j-1}(\rho-j+1)} & \text{for } j = \lfloor\rho\rfloor
\end{cases}$$
An Application: Distributed Data Repair
=======================================
Long Block Length Sparse Graph Codes With A Back-up Master
----------------------------------------------------------
Let us consider a master-slave configuration for a distributed data storage scenario in which the data protection is provided by a long block length $(n,k)$ sparse graph code whereby each slave node is assumed to store a single symbol. In addition, a master node constitutes a back-up system and keeps the copy of all coded symbols. In case one of the slave nodes fails, departs the network or becomes unavailable for some reason i.e., a symbol is lost in the system, due to multiple check relations defined for that lost symbol in the sparse graph code there would be multiple options of repair. To be able to maintain instantaneous reliability, the symbol needs to be repaired as soon as possible. However, it may not be possible to obtain the status of all nodes within the same network (due to other unexpected failures or network link breakages) within a short time or else it may be too time and bandwidth costly to contact the master directly for that information. Therefore in that case, the failed node needs to adapt the best guessing strategy and choose among the multiple repair options to complete the repair process (either exactly or functionally) using minimum network resources and as quickly as possible.
![An example repair process using an LDPC code Tanner graph. $d_v$ represents the degree number of the lost symbol whereas the $d_{c_1},\dots,d_{c_{d_v}}$ are the degrees of the potential repair check relations.[]{data-label="fig:sgc2"}](ldpc_guessing.pdf){width="0.5\linewidth"}
Let us suppose one of the degree-$d_v$ symbols of an irregular LDPC code, shown as gray-colored node in Fig. \[fig:sgc2\] is to be exactly repaired. Suppose it is connected to check nodes of degrees $d_{c_1},d_{c_2},\dots,d_{c_{d_v}}$, as shown in the same figure. As a natural choice, we define the costs associated with each choice to be the number of downloaded symbols, i.e., $c_j\triangleq d_{c_j}-1$[^5] for all $j$ satisfying $1 \leq j \leq M-1$. The following proposition establishes the condition for contacting the master node if no neighboring node is able to help in an optimal guessing context.
Let $c_M$ be the cost of contacting the back-up node and $c_{max} = \max\{c_1,c_2,\dots,c_{d_v}\}$ satisfying $c_M \geq c_{max}((1-q)^{-c_{max}}-1) \geq c_{max}$ where $M=d_v+1$. Let also each slave node to be independently unavailable/failed with probability $q > 0$, then guessing check relations as well as the back-up master in the order of non-decreasing costs minimizes the average number of downloaded symbols in the repair process.
\[prop2\]
Assuming independence, The probability that $j$-th check node will successfully repair the gray-colored node can be shown to be of the form $$\begin{aligned}
p_j = (1-q)^{c_j} \prod_{i=1}^{j-1} (1-(1-q)^{c_i}), \ \ \ \ p_{M} = 1 - \sum_{j=1}^{d_v} p_j \label{eqn43}
\end{aligned}$$ from which we realize that the probabilities are dependent on the costs. In a more general version of the problem, the costs of the check nodes may take values independent of the degrees (e.g., the communication cost required for obtaining a variable node may be different). In search of an optimal strategy, we need to think about $p_j$’s and $c_j$’s at the same time. From equation (\[eqn43\]), we can express $p_j$’s recursively for $j \leq M-1$, $$\begin{aligned}
p_j = p_{j-1} \left[(1-q)^{c_j-c_{j-1}}-(1-q)^{c_j}\right]\end{aligned}$$ which implies that if $c_{j-1} \leq c_j$, due to $0 < (1-q)^s \leq (1-q)^t \leq 1$ for all positive $t \leq s$, we shall have $p_{j} \leq p_{j-1}$. Therefore, ordering costs in ascending order leads to ordering probabilities in descending order. But this result implies that the necessary condition of Corollary \[prop1\] i.e., $c_ip_j \leq c_jp_i$ (by the same rationale the necessary condition in Proposition \[prop0\]) is satisfied for all $i,j \in \{1,2,\dots,M-1\}$ and $i \leq j$. Note that if $c_{j-1} > c_j$, we would not be able to satisfy the necessary condition for all cases. In order to contact the master(back-up) node when no neighboring nodes are able to help we then have to satisfy the necessary condition $c_Mp_{M-1} \geq c_{M-1}p_M$. Using equation , this condition can be re-expressed as $c_M(1-q)^{c_{M-1}}\geq c_{M-1}(1-(1-q)^{c_{M-1}})$ which is the first inequality in the condition of the proposition with $c_{max} = c_{M-1}$. We finally recognize that the second inequality in the condition while costs satisfying $c_i > 1$ means $(1-q)^{c_{max}} \leq 1/2$. Considering it with the first inequality, this condition reduces to $c_M \geq c_{M-1}$ which completes the proof of the optimality of the non-decreasing cost order. We finally note that the second inequality need not to be satisfied for the back-up master to be the last resort. However $c_M \geq c_{M-1}$ becomes only necessary if the second inequality is satisfied.
Now, for any code symbol (also the node for that matter) we associate a random variable $X_v$ that will characterize the identification of the right check node for a successful repair. For instance $X_v = 3$ indicates that the $3^{rd}$ check relation with the minimum cost is the first possibility (due to proposition 4.1) for a successful recovery. Let $\mathcal{G}^*(X_1,X_2,\dots,X_n)$ denote the optimal guessing function for the value of a joint realization of independent random variables $X_1,X_2,\dots,X_n$ where each represents one of the code symbols. Then due to Theorem \[thmexponent\], for large enough block length (number of nodes $n$ tends large), the moments of repair bandwidth (cost in terms of downloaded symbols) using the optimal guessing strategy can be well approximated by the Rényi entropy rate, $$\begin{aligned}
\mathbb{E}[C_{\mathcal{G}^*}(X_1,X_2,\dots,X_n)^\rho] \approx \prod_i \exp \{ \rho H_{\frac{1}{1+\rho}}(Y_i) \} = \exp\{\rho \mathcal{R}_{\frac{1}{1+\rho}} (\{Y_i\})\} \label{eqn61}\end{aligned}$$
Numerical Results
-----------------
First, let us provide several numerical results to be able to illustrate how close the provided bounds are for finite values of costs, $\rho$ and $M$. The exact moments for the optimal guessing strategy are calculated using Algorithm 1. The results are provided in Table 1 and these results indicate that lower and upper bounds approximate the actual results well. More specifically, we consider the second and third moments where the exact values of $\frac{1}{2}\ln\mathbb{E}[{C_\mathcal{G*}(X)}^2 ]$ and $\frac{1}{3}\ln\mathbb{E}[{C_\mathcal{G*}(X)}^3 ]$ and their lower and upper bounds are calculated and compared. The probability of each choice is generated using geometric distribution as assumed in [@sason2018] with the restricted probability distribution $P_X(x) = (1-a)a^{x-1}/(1-a^M)$ with $M=32$ and the parameter $a=0.9$. The non-integer cost values are generated based on a truncated normal distribution defined in the range $(1,100)$ with mean and variance $\mu=\sigma^2=16$.
LB (Eq. ) LB (Eq. ) LB (Eq. ) **Exact Value** UB (Eq. ) UB (Eq. ) UB (Eq. )
------------------------------------------------------------- ----------- ----------- ----------- ----------------- ----------- ----------- -----------
$\frac{1}{2}\ln\mathbb{E}[{C_{\mathcal{G}^*}(X)}^2 ]$ 4.1430 4.5053 5.1598 5.2062 5.5799 5.5814 6.1277
$\frac{1}{3}\ln\mathbb{E}[{C_{\mathcal{G}^*}(X)}^3 ]$ 4.1794 4.5529 5.3622 5.3818 5.7035 5.7048 6.1643
$\frac{1}{100}\ln\mathbb{E}[{C_{\mathcal{G}^*}(X)}^{100} ]$ 4.3324 4.9383 6.2064 6.2512 6.2828 6.2835 6.3281
: Lower and Upper bounds for $\frac{1}{2}\ln\mathbb{E}[{C_{\mathcal{G}^*}(X)}^{2} ]$, $\frac{1}{3}\ln\mathbb{E}[{C_{\mathcal{G}^*}(X)}^{3} ]$ and $\frac{1}{100}\ln\mathbb{E}[{C_{\mathcal{G}^*}(X)}^{100} ]$. UB: Upper Bound, LB: Lower Bound. In equation , $\beta$ values for this bound are searched discretely in $(-\rho,100]$[]{data-label="tab:numerics"}
Conclusions and Future work
===========================
In this study the general notion of guessing cost is introduced and an optimal strategy is provided to minimize the moments of guessing a random variable defined on a finite set with each choice may be associated with a positive finite cost. Asymptotically tight upper and lower bounds on the moments of cost of guessing are derived and expressed in terms of the Rényi’s entropy. We have also have laid out connections with the standard guesswork through a new random variable which can be efficiently calculated from the original random variable. Thanks to this connection, previous works on the bound improvements (especially in the non-asymptotic region) of the standard guesswork become readily applicable and hence we provided improved bounds on the cost of guessing without rigorous proofs. Finally, we establish the guessing cost exponent on the moments of the optimal guessing by considering a sequence of random variables and expressed it in terms of Rényi entropy rate. Finally, these bounds are shown to serve quite useful for bounding the overall repair latency cost (data repair complexity and bandwidth) for distributed data storage systems in which sparse graph codes may be utilized.
We note that it is of great interest in a distributed storage protocol design to consider the possibility of giving up on the guessing the next possible repair option based on a condition such as the total accumulated cost. Characterization of the cost of guessing in that case would have to be expressed in terms of smooth Rényi’s entropy. Recent studies such as [@kuzuoka2019] and [@duffy2019] considered similar constraints for the standard guesswork within the context of source coding and communications, respectively.
[1]{}
J. O. Pliam. (1999) The Disparity Between Work and Entropy in Cryptology. Available Online: http://philby.ucsd.edu/cryptolib/1998/98-24.html
J. L. Massey, “Guessing and entropy,” *in Proc. IEEE Int. Symp. on Information Theory* Trondheim, Norway, 1994, pp. 204.
R. J. McEliece and Z. Yu, “An inequality on entropy,” *Proceedings of the 1995 IEEE International Symposium on Information Theory*, p. 329, Whistler, Canada, September 1995.
E. Arikan, “An inequality on guessing and its application to sequential decoding,” *IEEE Trans. Inform. Theory,* vol. 42, pp. 99–105, Jan. 1996.
L. L. Campbell, “A coding theorem and Rényi’s entropy." *Information and control* 8.4: 423-429, 1965.
S. Kuzuoka, “On the Conditional Smooth Rényi Entropy and Its Application in Guessing," *IEEE International Symposium on Information Theory (ISIT),* Paris, France, 2019, pp. 647-651.
K. R. Duffy, J. Li and M. Médard, “Capacity-Achieving Guessing Random Additive Noise Decoding," in *IEEE Transactions on Information Theory,* vol. 65, no. 7, pp. 4023-4040, July 2019.
Bracher, Annina, Eran Hof, and Amos Lapidoth. “Guessing attacks on distributed-storage systems." *IEEE Transactions on Information Theory* 65.11 (2019): pp. 6975-6998.
D. Malone and W. G. Sullivan, “Guesswork and entropy," in *IEEE Transactions on Information Theory,* vol. 50, no. 3, pp. 525-526, March 2004.
K. R. Duffy, J. Li and M. Médard, “Capacity-Achieving Guessing Random Additive Noise Decoding,” in IEEE Transactions on Information Theory, vol. 65, no. 7, pp. 4023-4040, July 2019,
R. Gallager, “Low-density parity-check codes," in *IRE Transactions on Information Theory,* vol. 8, no. 1, pp. 21-28, January 1962.
S. Boztas, “Comments on “An inequality on guessing and its application to sequential decoding”," in *IEEE Transactions on Information Theory,* vol. 43, no. 6, pp. 2062-2063, Nov. 1997.
C. Bunte and A. Lapidoth, “Maximum Rényi Entropy Rate,” in IEEE Transactions on Information Theory, vol. 62, no. 3, pp. 1193-1205, March 2016.
I. Sason and S. Verdú, “Improved Bounds on Guessing Moments via Rényi Measures," *IEEE International Symposium on Information Theory* (ISIT), Vail, CO, 2018, pp. 566-570.
W. K. Lai and E. Kim, “Some inequalities involving geometric and harmonic means," *In International Mathematical Forum,* (2016), Vol. 11, No. 4, pp. 163-169.
Park, H., Lee, D., & Moon, J. (2017). LDPC code design for distributed storage: Balancing repair bandwidth, reliability, and storage overhead. *IEEE Transactions on Communications,* 66(2), 507-520.
[^1]: S. S. Arslan is with the Department of Computer Engineering, MEF University, Maslak, Istanbul, Turkey e-mail: [email protected].
[^2]: E. Haytaoglu is with the Department of Computer Engineering, Pamukkale University, Turkey.
[^3]: Manuscript draft version prepared in June 2020, received; revised, in NA.
[^4]: Bubble-sort is a sorting algorithm that works by repeatedly swapping the adjacent elements in a given list based on a condition.
[^5]: Here, due to large block length assumption, it is assumed that subsequent guesses cannot help each other. In addition, other cost metrics can be used.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Statistical mechanical concepts and processes such as decoherence, correlation, and dissipation can prove to be of basic importance to understanding some fundamental issues of quantum cosmology and theoretical physics such as the choice of initial states, quantum to classical transition and the emergence of time. Here we summarize our effort in 1) constructing a unified theoretical framework using techniques in interacting quantum field theory such as influence functional and coarse-grained effective action to discuss the interplay of noise, fluctuation, dissipation and decoherence; and 2) illustrating how these concepts when applied to quantum cosmology can alter the conventional views on some basic issues. Two questions we address are 1) the validity of minisuperspace truncation, which is usually assumed without proof in most discussions, and 2) the relevance of specific initial conditions, which is the prevailing view of the past decade. We also mention how some current ideas in chaotic dynamics, dissipative collective dynamics and complexity can alter our view of the quantum nature of the universe.'
author:
- |
B. L. Hu [^1]\
[Department of Physics, University of Maryland, College Park, MD 20742, USA ]{}
title: 'Statistical Mechanics and Quantum Cosmology [^2]'
---
=8.5in =6.5in =-0.5in =0.in =0.in
Introduction
============
In this talk I would like to present some general personal views on how the concepts and methodologies in statistical mechanics can be of use to facilitate a better understanding and a clearer formulation of some basic issues of quantum cosmology.
In keeping with the interdisciplinary nature of this workshop I will only discuss ideas here and avoid technicalities, knowing that one can always find the details in the original papers. Because of the general nature of discussions the viewpoints presented here could also appear to be tentative, sketchy and partial. I will, however, try to mention the important papers, as well as recent reviews and conference proceedings in this field for those who would like to get a better overall perspective. In this connection I may suggest Proceedings of the 1988 Osgood Hill Conference$^{1}$, the 1989 Jerusalem Winter School$^{2}$, and selected papers in the 1989 Santa Fe Institute Studies$^{3}$. Quantum Cosmology began in the sixties with the work of DeWitt$^{4}$ Wheeler$^{5}$ and Misner$^{6}$. The revival of interest in the eighties was brought about primarily by the works of Hartle, Hawking$^{7}$ and Vilenkin$^{8}$. A very complete bibliography of recent papers on quantum cosmology was compiled by Halliwell$^{9}$, who himself contributed extensively to its recent development.
Information theory and quantum measurement ideas useful in quantum cosmology were introduced by Wheeler, Zurek and Unruh$^{10}$, Joos, Zeh and Kiefer$^{11}$, Griffith and Omnes$^{12}$, Gell-Mann and Hartle$^{13}$. Since this particular aspect of statistical mechanics in relevance to quantum cosmology is quite well noticed (see Ref. 3), I will not pursue it here. Neither will I belabor on the Hamiltonian$^{4-6}$ nor the path integral$^{7,8}$ formulations, which make up most of the formal work in this field. Because I have difficulty understanding the physical meaning of results derived from Euclidean techniques I prefer to start from a firmer ground (quantum field theory in curved spacetime, or quantum mechanics) and analyze smaller problems (Brownian motion and the ubiquitous harmonic oscillator!) which I can understand and trust the results. In this sense I’ll actually not be doing quantum cosmology (as we shall see these two words given their present connotation can be intrinsically contradictory), but discuss quantum statistical processes like dissipation, fluctuation, correlation and decoherence in a much simpler context. These problems, which are basic to many physical phenomena and are believed to play an important role in quantum cosmology, have simply not been understood well enough in the more complicated conditions characteristic of quantum cosmology to warrant liberal generalization and excessive claims. In essence, then, I will take the layman’s approach, asking some intuitive and rudimentary questions, and trying to see if the basic tenets of quantum cosmology are sound and are compatible with what we understood in statistical and quantum physics. One of these sample problems, specifically on the stochastic properties of interacting quantum fields, has recently been treated in detail from first principles by Juan Pablo Paz, Yuhong Zhang and myself$^{14-17}$. It was reported in Dr. Paz’s talk. We are now in the process of trying to understand its implications in the context of stochastic inflationary universe$^{18-22}$ semiclassical gravity$^{23.24}$, and quantum cosmology. The study of dissipation in quantum fields, semiclassical gravity and in quantum cosmology has been pursued by Esteban Calzetta and myself over the years$^{25-33}$ and furthered by Paz$^{34,35}$ recently. An earlier account of some views on the basic issues of minisuperspace cosmology can be found in my 1989 Erice lectures$^{31}$.
Viewpoint and Issues in Quantum Cosmology
=========================================
Classical cosmology as a discipline, from Newtonian to relativistic, is based on the continuum notion of spacetime described by Riemannian geometry. There is no valid picture of quantum geometry as yet. When gravity is conceived as the force mediated by a spin-two particle, the graviton, one can talk about quantum gravity, even though grave problems exist in relation to its renormalizability; and for its higher derivative extensions, the $R^{2}$ theories, also unitarity and causality problems. Einstein’s relativity theory based on the Hilbert action $\int R \sqrt {-g} d^{4}x$ from which cosmology is derived, is in this view believed to be the long wavelength or low energy limit of a more complete and intricate theory, superstring theory being one serious candidate. Above the Planck scale quantum properties of spacetime become important. Here geometries with non-trivial topology can spontaneously be created and annihilated, rendering the smooth continuum picture of spacetime totally inadequate. At this time there are many attempts at constructing a picture of quantum spacetime but no success is yet in sight. The goal is however shared by workers on many fronts, including superstring theory$^{36}$, conformal field theory$^{37}$,topological field theory$^{38}$, wormholes and baby universes$^{2}$, and random geometry$^{39}$.
Therefore, according to this view, any talk on, say, superstring cosmology, which starts with a Robertson-Walker or de Sitter metric does not make much sense, because the spacetime constructed from and governing the superstrings cannot be the simple manifolds. One can at best be looking at the low energy limit (about, but not above the Planck scale) of such a theory, which can equally well be treated by the well-known semi-classical theories. If one views gravity as an effective theory resulting from elementary particle interactions, as Sakharov$^{40}$ had proposed, the notion of continuum spacetime only makes sense in the long-range limit, much as elasticity is to electrodynamics. Then quantum cosmology is as meaningful as quantum elasticity. It is almost just as crazy (though doable in principle) to try to derive elasticity from QED, as is it hopeless (though we all seem to be trying) to deduce QED from elasticity. The incongruence ingrained in the words “quantum cosmology” is a reflection of the deeper dilemma one faces today in probing the quantum structure of spacetime.
So just what exactly does one mean when one talks about quantum cosmology? and what does one gain in asking such almost impossible questions? Different people have different ideas about quantum cosmology. If one believes that the metric is a basic observable depicting a fundamental physical field quite at the opposite spectrum from the effective theory, then one would quantize the three-geometry $^{3}g_{ij}$ by imposing commutation relations between $^{3}g_{ij}$ and $\pi _{ij}$, its conjugate momentum, like what one usually does with any classical physical observable. The difficulties one encounters are that of quantum gravity proper. Quantum cosmology then refers to the quantization of a restricted set of the degrees of freedom allowed in quantum gravity. The geometries commonly encountered in (classical) cosmology like Robertson-Walker, de Sitter, and mixmaster universes are only the homogeneous anisotropic solutions to Einstein’s equations. In a perturbative sense (e.g. the Lifshitz operator) one can view these geometries as the lowest modes of spacetime excitations$^{31}$, the higher modes corresponding to the inhomogeneous universes are being ignored. In the superspace picture$^{4,5} ($the space of all 3-geometries) homogeneous cosmology constitutes a highly truncated class, the so-called minisuperspace $^{6}$. The problem is now simplified from quantizing an infinite to only a few degrees of freedom. Most discussions of quantum cosmology make such a simplification. Whether this is acceptable has been an open question since the Hamiltonian cosmology of the sixties. We shall address this issue concerning the validity of minisuperspace approximation $^{41}$ from the statistical mechanical viewpoint later.
Quantum cosmology in the eighties works also mainly with minisuperspace, i.e., makes the same assumption of truncated degrees of freedom, but the attention was shifted to the issue of initial states or boundary conditions on the Euclidean path integrals. This opened up new avenues because a simple prescription$^{7,8}$ on what 4-geometries to sum over in the Euclidean path integral leads to physically interesting results. For example, Hartle and Hawking’s$^{7}$ no boundary condition of summing over all compact 4-geometries with boundary on $^{3}g$ and Vilenkin’s$^{8}$ choice of outgoing modes on regions of the boundary where the 4-geometry is singular both have the following desirable features:
\(1) There are regions where the wave function of the universe $\Psi $ is oscillatory, corresponding to classically allowed solutions.
\(2) There exist solutions with inflationary behavior (see, however, Ref. 2)
\(3) These boundary conditions on $\Psi $ select a particular solution to the functional Schrödinger equation and define a particular vacuum state in the semiclassical limit of quantum field theory in curved spacetime.
Simply put, these boundary conditions admit classical solutions, some induce inflation, and give the correct vacuum state in the semiclassical limit. The prediction of classical spacetimes corresponding to the existence of oscillatory wave function is a very important result, because many features of our physical world$^{18.19}$ are connected with the existence of a late universe (the flatness, age and entropy problems in cosmology are all related to this fact$)^{20}$ described by classical spacetimes. Time is one of such features, as it is the only observable in quantum mechanics not represented by an operator, but enters as a parameter, thus lacking the interference effect intrinsic in all quantum phenomena. Because of its preferred status, time plays a special role in quantum mechanics and in general relativity, and brings about special problems in quantizing gravity. In this view, the issue of time is naturally linked to the issue of quantum to classical transition. In addressing these two issues statistical mechanical considerations enter in a fundamental way. Though occuping a central position, they are not, however, exclusive concerns of quantum cosmology. In fact the basic mechanisms which can bring about quantum to classical transitions like decoherence and correlation are common to all quantum phenomena and, in my view, should be better explored first outside of quantum cosmology without its particular problems. What makes these issues particularly relevant in quantum cosmology are 1) the appearance of classical spacetime and the associated special features of time in classical physics, and 2) the existence of a late classical universe as “a consequence of the specific condition in a more general sum-over-history framework of quantum prediction”$^{13}$. Thus according to this view$^{7,13}$ quantum cosmology provides one with a pathway to connect these issues (of time and classicality) back to the issue of initial conditions. A parallel development also claims the usefulness of these initial conditions to predict the value of universal coupling constants in the context of summing over non-trivial topologies (wormholes and baby universes$)^{2}$. The overwhelming emphasis in recent work on quantum cosmology is on the specificity of initial conditions. This is where our view differs: Put succinctly, we attach equal importance to these issues (time and classicality) which are amplified in the context of quantum cosmology, but we don’t see the choice of initial conditions as the most natural way to resolve these issues. Rather, we view the emergence of time and the quantum to classical transition more as consequences of dynamics and interactions of the system of interest with its environment, and attribute more importance to the working of statistical mechanical effects in the broadest senses, including the guiding principles of information-theory and chaotic dynamics. Let me explain what we mean by this.
How are Statistical Considerations Relevant?
============================================
In a broad sense, statistical mechanics deals with the issue of how to extract from a large, often infinite, degrees of freedom, a few variables which can capture the most essential physics of the whole system. Examples in physics are abound: Thermodynamics describes a system in equilibrium with its surrounding by a few macrovariables like temperature, pressure, entropy, etc; hydrodynamics with the transport functions, and critical phenomena with the critical exponents and universality classes. In all cases the skill rests upon 1) identifying (separation) the relevant variables describing the system of interest (often this involves not just taking a subset of the primitive microvariables but taking the average in some approximation and working in a different level of structure), 2) averaging away some information of the irrelevant variables (coarse-graining) which make up the environment, and 3) consideration of its overall effect on the system (backreaction). This schema of isolating a system of interest and the effective accounting for its interaction with the surroundings is not exclusive to statistical mechanics. In field theory renormalization and coupling hierarchy problems share the same spirit$^{29}$.
Such a reduction procedure applied to the overall system and environment (closed system) results in the appearance of some unusual but commonly accepted behavior for the system alone (open system) which would otherwise be completely absent. The most salient feature of the dynamics of an open system$^{42}$ is dissipation, and the associated appearance of an (thermodynamic) arrow of time for the system observer. Formally one sees this difference arises when one goes from the density matrix description of the system-environment obeying the unitary Louiville-von Neumann equation,to the reduced density matrix for the system alone obeying a non-unitary master equation$^{43}$. This is the basic reason for the apparent contradiction between the unitarity of microphysics and the apparent breakdown of macrophysics. This schema can be used to describe the so-called “collapse of the wave function”, or quantum to classical transition in quantum measurement theory$^{10,11}$. The disappearance of interference between different branches of the wave functions associated with the vanishing of the off-diagonal components of the reduced density matrix is called decoherence. It is a consequence of information loss in the system through its coupling to the environment. In most practical cases, the environment can be simplified as a stochastic source. In fact, dissipative effects are oftentimes depicted by phenomenological equations such as the Langevin equation. Instead of an explicit description of the environment, one replaces it with a noise source which drives the dynamics of the system. The noise source can be from thermal or vacuum fluctuations which are related to dissipation through the fluctuation-dissipation relation.
This general scheme in statistical mechanics, often modeled by the motion of a Brownian particle (system) interacting with a collection of harmonic oscillators (bath) and depicted by a master equation for the reduced density matrix$^{42-47}$ show clearly 1) the interconnection of fluctuation, noise, dissipation, decoherence and correlation; and as a consequence of these processes, 2) the nonspecificity of initial conditions in determining the long time behavior of the system.
We have applied this scheme to interpret quantum dissipative processes in semiclassical gravity$^{29}$. Let us now examine the issues of quantum cosmology in this light. The issues I raised earlier can be grouped into three distinct levels:
\(1) Is gravity an effective theory? If yes, how do statistical concepts enter?
\(2) Assuming that it is not, i.e., that the 3-geometries can be viewed as fundamental physical variables, then within this (conventional) framework of quantum cosmology,
a\) How valid is the minisuperspace approximation in describing the full dynamics of quantum cosmology?
$b)$ How specific are the initial conditions in yielding classical limits?
\(3) What brings about decoherence and classicality?
Note that questions in level (3) are actually questions in quantum mechanics proper, without a proper understanding of which one cannot get a satisfactory answer to the corresponding questions in quantum cosmology. A thorough understanding of (2) alone may not help answering questions in level (1), but there is reason to believe that answers to questions in level (1) could help to resolve most questions on level (2). Cracking the mystery of (1) is, however, very difficult.
I will give here a general description of how statistical considerations enter into quantum cosmology and leave the more technical discussions on methodology and sample calculations to the next section.
As we mentioned early, cosmology is a study of spacetime dynamics and structure based on the observed facts in our universe. From the observed high degree of isotropy and homogeneity one constructs the standard model $^{18}$ based on the Friedmann-Robertson-Walker universe, which is a fairly good depiction of the late history of the universe back to at least $10^{-43}$ sec. from the big bang. The flatness and entropy problems prompted one into speculating that at some very early time (the grand unification time ${\char"7E} 10^{-35} \sec )$ the universe might have undergone a stage of inflation$^{20}$. This ushered in renewed interest of the de Sitter universe. Mathematical analysis of Einstein’s equation indicates that near the singularity the universe could have been highly anisotropic and inhomogeneous, thus pointing to the relevance of the Bianchi models (spatially homogeneous) and the mixmaster universe (Type IX rotation group of motion) in particular$^{19}$. There is no a priori reason why the inhomogeneous cosmologies are usually ignored, except that they are very difficult to study, and for the belief that inhomogeneities and anisotropies in the very early universe could be dissipated away through various mechanisms, vacuum particle production near the Planck time being the most powerful$^{24,25}$.
So cosmology is the study of the dynamics of a very restricted set of spacetime (with high symmetries) to begin with. This is similar to the first task of statistical mechanics, i.e., selecting a few relevant parameters which can capture the physical essence of the whole system. In late cosmology, these parameters are given by observation (e.g. Hubble expansion, microwave background), but in early cosmology especially in the realm where quantum phenomena are dominant, which parameters are important is not as straightforward. (It may be that spacetime is an averaged, composite concept and even the simple scale factor in the metric function loses its obvious meaning. See ideas from Regge calculus$^{48}$, random geometry$^{39}$, cellular automata$^{49}$, causal sets$^{50}$, etc). Within the conventional framework, the minisuperspace approximation in quantum cosmology assumes that the infinite degrees of freedom ignored has little effect on the selected ones (see however calculation of Halliwell and Hawking$^{7})$. Even so, using the schema in statistical mechanics I described above, this is highly questionable. Viewing the scale factor a and the anisotropy parameters $\beta
\pm $ as our system, say, in a Bianchi Type IX universe$^{19}$ and treating the remaining infinite degrees of freedom corresponding to other anisotropic $(\theta _{1,2,3})$ and inhomogeneous universes as the environment, one would get additional terms in the (non-unitary) master equation associated with the Wheeler-DeWitt equation (an energy constraint condition$)^{31}$. With an eye towards matching the classical limit, one can use the Wigner functional $f(g,\pi )$ for the 3-geometries $g$ and the conjugate momenta $\pi
($which describes the distribution of states in a superphase space) and obtain a Wheeler-DeWitt Vlasov equation$^{28}$, or a Fokker-Planck equation$^{11,22,19}$. These equations embody the dissipative and diffusive effects in the dynamics of the minisuperspace variables. An immediate consequence is that the late time behavior will no longer depend sensitively on the specific choice of initial conditions. One can show that for reasonable conditions on the bath and the coupling, memory loss (near Markovian behavior) is a rather general phenomena$^{30-33}$. This viewpoint supports the chaotic cosmology philosophy, which the mixmaster$^{19}$ and the inflationary$^{20}$ programs both share - i.e., that the present state of the universe depends only weakly on the stipulation of initial conditions, but results largely from its own dynamics and interactions.
For cosmology, splitting the system and the bath may pose some conceptual difficulty if one envisages the Universe as containing “everything” - spacetime and matter. Though this is by definition true, it is in reality often not so stringently applied. The observable “Universe” often refers to the causally connected parts which are much smaller, and relevant physics goes on largely within the particle horizons. One can take the part beyond as the environment. For spacetimes with event horizons (e.g. the de Sitter universe) physics within and without can be quite different (e.g., Starobinsky’s stochastic inflation$)^{21}$. One can also view topological fluctuations such as wormholes and baby universes as making up the environment, which can have interesting effects on the universe proper$^{51}$. For an earlier discussion on how nontrivial global structures of spacetime and quantum processes in the universe can lead to entropy generation see Ref. 52.
Another way statistical considerations enter into cosmology is from chaos and dynamical systems. As is well-known, the Einstein equations for some classical cosmological models (e.g. Bianchi Type IX) admit chaotic behavior$^{53}$. The criteria of chaos in this context (e.g. Liapunov exponent) are still under investigation$^{54}$, but the existence of even slight chaos will render the specificity of initial conditions highly unstable in its prediction of late time behavior. Whether quantum dynamics exhibits the same degree of chaos is unknown but it is unlikely that the initial conditions could remain highly regulative and predictive, as the prevailing view in quantum cosmology seeks to establish.
A related criticism of a statistical nature of the “initial condition” school of thought comes from complexity and information theory considerations. C. H. Woo$^{55}$ pointed out that in addition to a simple and elegant initial condition for the wave function of the universe, the macroscopic variables require arbitrary inputs for a more detailed description of the classical history of the universe. He estimated that “the number of bits needed to encode the algorithm for the wave function and its simple initial condition is relatively small, of the order of $10^{3}$”. Thus, “such a wave function can predict at most the specific behavior of a few hundred macroscopic variables”, a far cry from the sweeping optimism implied by the initial conditions school of thought.
So far we have assumed that gravity is viewed as a fundamental field. If instead gravity is regarded as a composite force or an effective theory, then statistical considerations would enter in an even more basic and familiar way. Think about how we construct molecules from atoms, then gas, fluids and solids. Statistical mechanics is an almost indispensable tool when one wants to extract meaning and structure from a collection of more elementary constituents. In addition to the effective theory of Sakharov mentioned above, I have also toyed with the idea of viewing gravity as the result of a “time-dependent” Hartree-Fock$^{56}$ interaction among the unspecified basic constituents, similar to the nuclear collective model, where the “normal modes” of nuclear rotation and vibration are regarded as the dynamical variables. Note that for the description of many gross features one does not need to know the details of the nucleons, which are the more elementary constituents, but only their collective motion. This is in contradistinction to the independent particle model, where the starting point is the individual nucleon. From this maybe we can learn something about the relationship of the two approaches to quantum gravity, vis, starting with a collective structure like geometry or starting with the more basic constituents like superstrings or pre-geometry$^{18}$. The other idea I have toyed with somewhat is to view Einstein’s theory as the hydrodynamic (long wavelength) limit of a micro-theory of gravity. There is of course no implied “collision” of the basic constituents in the corresponding gravitational “kinetic theory” other than their nonlinear interaction. But the reduction of the unitarity dynamics of a microsystem to a macrosystem exhibiting irreversibility like the Boltzmann equation is an interesting analogous scheme. The incorporation of hydrodynamic fluctuations and phase transition ideas$^{57}$ into the consideration of transition from quantum to classical gravity may bring forth some new insight. In the more main stream recent developments, how statistical mechanics enters into random geometry, conformal field theory and superstring as ways to relate to quantum gravity should be quite apparent. It is not my intention to delve into these areas here but I hope at least I have convinced you that a good knowledge of non-equilibrium statistical mechanics (field theory) is essential for understanding the basic issues of quantum cosmology.
Let me now start over from the beginning and discuss some techniques which we find useful to treat problems of this nature.
Formalisms and Sample Problems
==============================
As I mentioned in the beginning, because there exist many subtleties and conceptual pitfalls in quantum cosmology and our current understanding of the statistical meaning of quantum mechanics is still vague, it is perhaps more fruitful to begin the investigation on a more familiar and firmer ground before putting them to test in the volatile setting of quantum cosmology. These phenomena are: fluctuation, noise, dissipation, particle creation, decoherence and correlation; the processes are: separation (definition of open system), coarse-graining, backreaction semiclassical limit, and quantum to classical transition. Various authors have tried to attack individual phenomenon without due consideration of the interrelation with others - e.g. decoherence without taking into account dissipation$^{58}$, semiclassical limit without taking into account the nonlinearity of fields and nonseparability of the background$^{59}$. And because they want to do these in the complexity of quantum cosmology and to see the result in one strike, they have to make many simplifying assumptions which may actually wash away the many features special to quantum cosmology (e.g. the Born-Oppenheimer separability for geometry and matter, the WKB approximation for the wave function, the linearity of gravitational normal modes).
The platform we choose to work on in this first stage of investigation is quantum field theory, first in flat space and then in curved space, the latter being the semiclassical limit of quantum gravity. We have some previous knowledge of how dynamical excitation of the vacuum in the form of particle creation can act as a dissipative force in changing the dynamics of spacetimes$^{24,25}$. We know how the closed-time-path formalism can yield a real and causal equation of motion for the geometry$^{25}$ and provide a correct statistical interpretation of particle creation as a dissipative process$^{29}$. We recognize the effective action as a suitable object for studying backreaction effects$^{25}$. We also gained some experience in the properties of Wigner function as a classical distribution function$^{26}$, and in dealing with near-uniform kinetic systems. These were the ground posts we have established to explore the statistical properties of quantum fields with an eye on the cosmological applications.
The one big missing piece in this mosaic is noise and fluctuation. We believe there must exist some fluctuation-dissipation relation even for nonequilibrium systems (not just for close to equilibrium linear response systems), as suggested in the particle creation and backreaction problems $^{29}$. Even though the effect of noise, fluctuation and dissipation are rarely mentioned in quantum cosmology$^{60}$, we believe they should play as essential a role as decoherence and correlation, as they are intrinsically related, although they address different aspects of the basic issues.
Thus the first task we set for ourselves in this program is to formulate a stochastic theory of quantum fields from first principle. The criterion is that it should contain the interconnection of all the above-mentioned processes and can address all statistical properties of quantum fields in curved spacetime. The two major useful ingredients we found are the influence functional and the coarse-grained effective action, which I will now briefly describe.
These ideas are of course not new. The former was established by Feynman and Vernon$^{45}$ , the latter is a recasting of the Zwanzig-Mori$^{43}$ projection operator formalism in effective action forms. The influence functional captures the overall averaged effect of the environment on the system, from which one can obtain the quantum master equation for the reduced density matrix. This formalism was lately popularized by Caldeira and Leggett$^{41}$ in their work on dissipative tunneling. It has also seen application to the inflationary cosmology$^{61}$. We have worked with the closed-time-path integral formalism of Schwinger and Keldysh$^{46}$ for particle creation and backreaction problems$^{25}$ and we know that it is suitable for treating non-equilibrium quantum fields$^{26}$, but did not realize that it is just the influence functional formalism of Feynman-Vernon (in fact Schwinger’s paper was on Brownian motion$)^{62}$. One can interpret the $x$ and $x'$ paths as one propagating forward and the other backwards in time. After we discovered this connection all of our previous results on CTP formalism can be carried over easily. The other pivotal point is the incorporation of noise and stochastic sources in purely quantum field theory terms without the presence of a thermal bath. This lifts the restriction of near-equilibrium conditions customarily imposed in the treatment (e.g. linear response theory) of noise, fluctuation and dissipation. Indeed one major result we obtained from this program is to show that a general fluctuation-dissipation relation exists for non-equilibrium quantum systems. We have used this to improve on a recent result of Unruh and Zurek, as well as making new predictions.
Our program was executed in successive stages$^{14-16}$, starting with the quantum mechanical problem of a Brownian particle bilinearly (Cxq) coupled to a collection of harmonic oscillators as bath, generalizing the result of Caldeira-Leggett and Unruh-Zurek to non-local dissipation and colored noise. (We call the dissipation local and the noise white if the dissipation and noise kernels are both delta functions.) We obtain (for the first time, we believe) an exact quantum master equation for these more general cases. Contrary to what we originally anticipated, this equation describing a non-Markoffian process turns out to be not so complicated as an integro-differential equation, but only an ordinary differential equation with complicated time dependent coefficients, and is exact (see Eq. 7 of Ref. 63). When we apply the Wigner transform to this quantum master equation, we obtain for the Wigner distribution function the quantum Fokker-Planck equation with time-dependent diffusion coefficients. The second problem we looked at was a biquadratic $(\lambda x^{2}q^{2})$ coupling between the system and bath variables, still in the quantum mechanical context. Here we carried out a perturbation analysis (in preparation for the $\lambda \phi ^{4}$ theory) up to quadratic order in $\lambda $ and derived a nonstationary quantum master equation with nonlinearly generated dissipation and colored noise. The third problem was to do everything in field theory, first in flat space then in de Sitter space making use of its conformally flat property, thus deriving for the Brownian field with nonlinear nonlocal dissipation and colored noise a quantum functional master equation and the associated Fokker-Planck equation for the Wigner distribution functional.
This completes the first stage of our program. We have used this equation to study the loss of coherence in some interesting quantum mechanical problems and cosmological models$^{63}$. We are now in the process of applying this formalism to problems in semiclassical gravity (connecting dissipation due to particle creation to noise and fluctuation), stochastic inflation (noise generation) and quantum cosmology (may need a different framework). In this path-integral framework one recovers the nice results of Unruh and Zurek$^{10}$ on decoherence$^{58}$, and can also see its relation with dissipation and other processes. One can also relate the degree of decoherence with correlation (between the coordinate and momenta variables) as one set of criteria for classicality$^{64}$. We have indicated before that particle creation in quantum fields can bring about both dissipation and decoherence. Calzetta and Mazitteli$^{65}$ have recently shown in the context of quantum field theory in curved spacetime that particle creation is a necessary and sufficient condition for decoherence. For stochastic inflation we have some doubts on the validity of Starobinsky’s scheme$^{21}$ in generating white noise for a linear field via dynamic truncation at the event horizon. Instead, with nonlinear mode coupling, one can generate noise without assuming a dynamic truncation. We applied the coarse-grained effective action method$^{66}$ to calculate the effect of bath (high frequency modes) on the system (low-frequency modes) which are nonlinearly coupled. Indeed, colored noise is generated with nonlocal dissipation in this problem, which is what our general program predicts$^{16}$.
The problem of backreaction and of semiclassical approximation was brought up recently in the context of quantum cosmology$^{59}$. We feel that without incorporating dynamical fluctuations both in the fields and the geometry, and dealing with the nonlinearity and nonadiabaticity condition squarely one cannot bring forth too much new beyond what we already know from quantum field theory in curved spacetime. Quantum gravity is an intrinsically nonlinear theory. Backreaction is only an approximate concept - it is meaningful only if one can separate some background geometry apart from the remaining (field or gravitational) degrees of freedom which is only possible in linearized theory and at energies lower than the Planck energy. For quantum gravity at the Planck energy where full nonlinearity is at work, this separation is not easily attainable. There is of course still particle creation - not only of quantum fields, but also gravitons - but the way to treat these quantum processes is not by the background field method to which quantum field theory in curved spacetime belongs, but rather by dealing directly with the nonlinear interaction of fluctuations of both gravity and fields and their dynamical excitations (which give rise to particle creation). Our formulation can shed some light on the nonlinearity aspects if one models the gravity-field interaction in the form of field-field interaction, although factorizability assumption for the density matrix of a closed system in the influence functional method is still a limitation (see however, Grabert [*et al*]{} in Ref. 47).
We have not yet embarked on a full scale investigation of the basic issues in quantum cosmology using the framework we constructed, as the issue of time still poses a nontrivial problem for the path-integral formalism (see Ref. 13 and Kiefer in Ref. ll). However we have earlier done a few smaller pilot problems to show 1) how dissipation in quantum cosmology can efface the memory of initial conditions and 2) how the higher gravitational modes usually discarded can introduce an effective dissipative term in the equation of motion for the lower (minisuperspace) modes. The first problem was illustrated by Calzetta$^{30}$ with the example of a linear scalar field in a Robertson-Walker universe. The reduced density matrix obeys an equation of motion with a viscosity term containing a nonlocal kernel, signifying the existence of non-Markovian dissipation processes. In the second problem Sinha$^{67}$ used a $\phi ^{4}$ scalar field in a Robertson-Walker universe to mimic the nonlinear gravitational interaction. She viewed the scale factor $a$ and the lowest (conformal) scalar field mode $\chi _{0}$ as the background and studied the coupling of the higher scalar field modes $\chi _{n}$ with themselves and their backreaction on the background modes via the coarse-grained effective action method. Backreaction shows up as a dissipative term in the equation of motion for the lowest modes, in addition to the usual renormalized mass and a shifted natural frequency. These results exemplify our earlier claim that statistical effects can alter ones view on the importance of initial conditions and the validity of the minisuperspace truncation. We still need to understand better how the dissipative effect manifests itself in the dynamics described by different “times” chosen in quantum cosmology. We would also like to see how the “thermodynamic” arrow of time emerges from the dissipation and decoherence processes in a quantum to classical transition. These are part of our future work.
Summary
=======
I have arranged the issues, processes and methodologies discussed in this talk into a Table below. Also noted therein are some related problems in gravity and cosmology where statistical mechanical considerations can be fruitful. To conclude we think more attention need be paid to the following aspects:
1\) Relevance of statistical considerations in quantum cosmology.
2\) Their role in addressing basic issues in theoretical physics.
3\) Interconnectedness of statistical processes such as
a\) decoherence, dissipation and correlation
b\) noise, fluctuation and
c\) particle creation, backreaction and semiclassical approximation.
For example, quantum to classical transition involves all processes in $a)$ and requires considerations of $c)$, but $b)$ also affects $a)$ and engenders $c)$.
Inquires on the statistical properties of the vacuum (noise, fluctuation, excitation by dynamics, constraint by event horizon) in curved spacetime and quantum gravity can also shed light on the interconnection of basic issues in quantum mechanics, general relativity and statistic mechanics (as manifested in the Hawking effect$)^{68}$.
[*Issues:*]{}
1\) How large is the class of initial conditions which can admit classical spacetimes as solutions? How regulative and predictive are the specific initial conditions?
2\) How valid is the minisuperspace approximation?
3\) How does time emerge? Is classical spacetime a necessary condition for rendering time as we perceive it?
4\) Quantum to classical transition - criteria for classicality.
5\) Semiclassical limit - relation of quantum field theory in curved spacetime with quantum cosmology.
6\) Separability of background and field, validity of the adiabatic condition, backreaction and consistency.
[*Processes*]{}
1\) Coarse-graining: how sensitive are the final results to the averaging measure.
2\) Decoherence, correlation and dissipation
3\) Noise, fluctuation and dissipation
4\) Particle creation
[*Frameworks*]{}
1\) Nonunitary evolution equations: master equation, Fokker-Planck equation and Langevin equation.
2\) Closed-time-path integral formalism, influence-functional formalism
3\) Superscattering (\$) matrix formalism$^{69}$.
[*Techniques*]{}
1\) Subdynamics and projection operator formalism; coarse-grained effective action (for coarse-graining).
2\) Wigner distribution function, coherent state representation (for classical limits).
3\) BBGKY hierarchy, nth order correlation function (for correlation).
[*Other Related Problems*]{}
1\) Gravitational entropy$^{70}$ and Hawking effect: a stochastic field theoretical interpretation.
2\) Tunneling, decoherence and dissipation.
3\) Dynamical critical phenomena and noise-induced transition.
[**Acknowledgement**]{}
The work I described in this talk was done jointly in stages with Esteban Calzetta, Juan Pablo Paz, Sukanya Sinha, and Yuhong Zhang with whom I have enjoyed many interesting discussions and correspondences. (They should however not be held responsible for any outlandish comment or crazy idea I advanced here.) I would like to thank the organizers of this workshop, Professor Arimatsu in particular, for the effort they put in, and the warm hospitality they extended to us. This work is supported in part by the National Science Foundation under grant No. PHY-8717155
**REFERENCES**
1\) A. Ashtekar and J. Stachel (ed.) Conceptual Problems in Quantum Gravity, Proceedings Second Osgood Hill Conference (Birkhaüser, Boston, 1989).
2\) S. Coleman, J. Hartle, T. Piran and S. Weinberg (ed.) Quantum Cosmology and Baby Universes, Proceedings 7th Jerusalem Winter School for Theoretical Physics (World Scientific, Singapore, 1990).
3\) W. H. Zurek (ed.) Complexity, Entropy and the Physics of Information, Proceedings Santa $Fe$ Institute Studies in the Sciences of Complexity, Vol. IX. (Addison-Wesley, Reading, 1990).
4\) B. S. DeWitt, Phys. Rev. [*160*]{} (1967) 1113.
5\) J. A. Wheeler, in: Battelle Recontres, ed. C. DeWitt and J. A. Wheeler (Benjamin, New York, 1968).
6\) C. W. Misner, in: Magic Without Magic, ed. J. Klauder (Freeman, San Francisco, 1972).
7\) J. B. Hartle and S. W. Hawking, Phys. Rev. ${\sl D28} (1983)
2960$; J. J. Halliwell and S. W. Hawking, ibid ${\sl D31} (1985) 1777.$
8\) A. Vilenkin, Phys. Lett. ${\sl ll7B} (1982) 25,$ Phys. Rev. ${\sl D27} (1983) 2848; {\sl D30} (1984) 509.$
9\) J. J. Halliwell, Bibliography on Quantum Cosmology, Int. J. Mod. Phys. (1990).
10\) J. A. Wheeler and W. H. Zurek (ed.) Quantum Theory and Measurement (Princeton Univ. Press, Princeton, 1983); W. H. Zurek, Phys. Rev. ${\sl D24}
(1981) 1516; {\sl D26} (1982) 1862$; W. G. Unruh and W. H. Zurek, Phys. Rev. [*D40*]{} (1989) 1071.
11\) H. D. Zeh, Found. Phys. [*1*]{} (1970) 69; Phys. Lett. [*A116*]{} (1986) 9. E. Joos and H. D. Zeh, Z. Phys. ${\sl B59} (1985) 223$; E. Joos, Phys. Rev. ${\sl D36} (1987) 3285.$ C. Kiefer, Class. Quant. Grav. [*4*]{} (1987) 1369; [*6*]{} (1989) 651; Phys. Lett. [*A139*]{} (1989) 201; “Interpretation of the Decoherence Functional in Quantum Cosmology” Zurich Preprint 1990.
12\) R. Griffith, J. Stat. Phys. [*36*]{} (1984) 219; R. Omnès, ibid [*53*]{} (1988) 893, 933, 957.
13\) M. Gell-Mann and J. B. Hartle in Ref. 3; J. B. Hartle in Ref. 2.
14\) Yuhong Zhang, Ph.D. Thesis, University of Maryland 1990.
15\) B. L. Hu, J. P. Paz and Y. Zhang, “Quantum Brownian Motion with Nonlocal Dissipation and Colored Noise”, Paper I.
16\) B. L. Hu, J. P. Paz and Y. Zhang, “Stochastic Properties of Interacting Quantum Fields”, Paper II.
17\) B. L. Hu, J. P. Paz and Y. Zhang, “Noise and Decoherence in Stochastic Inflation”, Paper III.
18\) See, e.g. J. Peebles, Physical Cosmology (Princeton University, Princeton, 1971) S. Weinberg, Gravitation and Cosmology (Wiley, New York, 1982); C. W. Misner, K. S. Thorne and Wheeler, Gravitation (Freeman, San Francisco, 1973).
19\) C. W. Misner, Phys. Rev. Lett. [*22*]{} (1969) 1071; V. A. Belinsky, I. M. Khalatnikov and E. M. Lifshitz, Adv. Phys. [*19*]{} (1970) 525, [*31*]{} (1982) 639; D. M. Eardley, E. P. T. Liang and R. K. Sachs, J. Math. Phys. [*13*]{} (1972) 99; M. P. Ryan and L. C. Shepley, Homogeneous Relativistic Cosmology (Princeton University, Princeton, 1975).
20\) A. Guth, Phys. Rev. ${\sl D23} (1981) 347$; A. Albrecht and P. J. Steinhardt, Phys. Rev. Lett [*48*]{} (1982) 1220; A. Linde, Phys. Lett. ${\sl 108B} (1982) 389.$
21\) A. A. Starobinsky, in: Field Theory, Quantum Gravity and Strings, eds. H. J. de Vega and N. Sanchez (Springer, Berlin, 1986); S. J. Rey, Nucl. Phys. ${\sl B284} (1987) 706$; J. M. Bardeen and G. J. Bublik, Class. Quan. Grav. [*4*]{} (1987) 573; F. Graziani, Phys. Rev. $D38 (1988) 1122, 1131, 1802.$
22\) A. A. Starobinsky, in: Quantum Mechanics in Curved Spacetime, eds. J. Audretsch and V. de Sabbata (Plenum, London, 1990).
23\) See, e.g., N. Birrell and P. C. W. Davies, Quantum Fields in Curved Space (Cambridge University, Cambridge, 1982); B. L. Hu, L. Parker and D. J. Toms, Gravitation, Quantum Fields and Curved Spacetime (Cambridge University, Cambridge, 199?).
24\) Ya. B. Zel’dovich and A. A. Starobinsky, JETP [*34*]{} (1972) 1159; B. L. Hu and L. Parker, Phys. Rev. ${\sl D17} (1978) 933$; J. B. Hartle and B. L. Hu, ibid ${\sl D20} (1979) 1757, 1772$; J. Frieman, ibid ${\sl D39} (1989) 389.$
25\) E. Calzetta and B. L. Hu, Phys. Rev. ${\sl D35} (1987) 495.$
26\) E. Calzetta and B. L. Hu, Phys. Rev. ${\sl D37} (1988) 2878.$
27\) E. Calzetta and B. L. Hu, Phys. Rev. ${\sl D40} (1989) 656.$
28\) E. Calzetta and B. L. Hu, Phys. Rev. ${\sl D40} (1989) 380.$
29\) B. L. Hu, Physica [*A158*]{} (1989) 399.
30\) E. Calzetta, Class. Quant. Grav. ${\sl 6} (1989) L227.$
31\) B. L. Hu, “Quantum and Statistical Effects in Superspace Cosmology” in: Quantum Mechanics in Curved Spacetime, Proceedings of the 11th Course of the International School on Cosmology and Gravitation, Erice 1989, ed. by J. Audretsch and V. de Sabbata (Plenum, London, 1990).
32\) E. Calzetta, “Anisotropy Dissipation in Quantum Cosmology”, IAFE preprint, Buenos Airs, 1990.
33\) E. Calzetta and B. L. Hu “Dissipation in Quantum Cosmology”, in preparation.
34\) J. P. Paz, Phys. Rev. ${\sl D40} (1990) 1054.$
35\) J. P. Paz, Phys. Rev. ${\sl D42} (1990) 2.$
36\) See, e.g., M. Green, C. Schwarz and E. Witten, Superstrings Vol. $I \&$ II (Cambridge University, Cambridge, 1987); M. Kaku, Superstrings (Springer, Berlin, 1988); W. Siegel, String Field Theory (World Scientific, Singapore, 1989).
37\) See, e.g., B. E. Baaquie [*et al*]{} (ed.) Conformal Field Theory, Anomalies and Superstrings (World Scientific, Singapore, 1988).
38\) See, e.g., E. Witten, Comm. Math. Phys. [*117*]{} (1988) 353.
39\) See, e.g., D. Gross and A. A. Migdal, Phys. Rev. Lett. [*64*]{} (1990) 127; E. Brezin and V. Kozakov, Phys. Lett. ${\sl B236} (1990) 144$; M. R. Douglas and S. H. Shenkar, Nucl. Phys. ${\sl B335} (1990) 635$; M. R. Douglas, Phys. Lett. ${\sl B238} (1990) 176.$
40\) A. D. Sakharov, Dok. Akad. Nauk. SSSR [*177*]{} (1967) 70 \[Sov. Phys.- Doklady [*12*]{} (1968) 1040\]; S. L. Adler, Phys. Rev. Lett. [*44*]{} (1980) 1567; Rev. Mod. Phys. [*54*]{} (1982) 719; A. Zee, Phys. Rev. Lett. [*42*]{} (1979) 417.
41\) K. Kuchar and M. P. Ryan, Jr., in: Proc. Yamada Conference XIV. eds. H. Sato and T. Nakamura (World Scientific, Singapore, 1986), Phys. Rev. ${\sl D40} (1989) 3982.$
42\) E. B. Davies, Quantum Theory of Open Systems (Academic, London, 1976).
43\) R. Zwangzig in: Lectures in Theoretical Physics III, eds. W. E. Britten, B. W. Downes and J. Downes (Interscience, New York, 1961) 106-141; H. Mori, Prog. Theor. Phys. [*33*]{} (1965) 1338; C. R. Willis and R. H. Picard, Phys. Rev. [*A9*]{} (1974) 1343; R. Balescu, Equilibrium and Nonequilibrium, Statistical Mechanics (Wiley, New York, 1975); H. Grabert, Projection Operator Techniques in Nonequilibrium Statistical Mechanics (Springer-Verlag, Berlin, 1982); B. L. Hu and H. E. Kandrup, Phys. Rev. ${\sl D36} (1987) 1776$ and references therein.
44\) R. Rubin, J. Math Phys. [*1*]{} (1960) 309, [*2*]{} (1961) 373; G. W. Ford, M. Kac and P. Mazur, J. Math. Phys. [*6*]{} (1963) 504.
45\) R. P. Feynman and F. L. Vernon, Ann. Phys. (N.Y.) [*24*]{} (1963) 118; R. P. Feynman and A. R. Hibbs, Quantum Mechanics and Path Integrals (McGraw-Hill, New York, 1965).
46\) J. Schwinger, J. Math. Phys. [*2*]{} (1961) 407; L. V. Keldysh, Zh. Eksp. Teor. Fiz. [*47*]{} (1964) 1515 \[Sov. Phys. JETP 20 (1965) 1018\]. K. C. Chou, Z. B. Su, B. L. Hao and L. Yu, Phys. Rep. 118 (1985) 1.
47\) A. O. Caldeira and A. J. Leggett, Physica (Utrecht) $121A (1983)
587$; H. Grabert, P. Schramm and G. Ingold, Phys. Rep. [*168*]{} (1988) 115.
48\) T. Regge, Nuovo Cimento 19 (1961) 558; R. Sorkin, Phys. Rev. ${\sl D12}
(1975) 385$; H. Hamber and R. M. Williams, Nucl. Phys. ${\sl B248} (1984)
392, {\sl B267} (1986) 482, {\sl B269} (1986) 712$; R. Friedberg and T. D. Lee, Nucl. Phys. ${\sl B242} (1984) 145$; J. B. Hartle, J. Math. Phys. [*26*]{} (1985) 804, [*27*]{} (1986) 287.
49\) An example of ideas in cellular automata applied to inflationary cosmology is Linde’s eternal, self-reproducing universe: A. Linde, Physica Scripta ${\sl T15} (1987) 169.$
50\) See, e.g., L. Bombelli, J. Lee, D. Mayer and R. D. Sorkin, Phys. Rev. Lett. [*59*]{} (1988) 521.
51\) J. Ellis, S. Mohanty and D. V. Nanopoulos, Phys. Lett. ${\sl B211} (1989) 113.$
52\) B. L. Hu, in: Cosmology of the Early Universe eds. H. Sato and R. Ruffini (World Scientific, Singapore, 1984) and references therein.
53\) J. D. Barrow, Phys. Rep. [*85*]{} (1982) 1; I. M. Kalatnikov [*et al*]{} in: General Relativity and Gravitation $(GR10,$ Padova, 1983) ed. B. Bertotti, F. de Felice and A. Pascolini (Reidel, Dordrecht, 1984); O. I. Bogoiavlenskii, Methods in the Qualitative Theory of Dynamical Systems in Astrophysics and Gas Dynamics (Springer, Berlin, 1985).
54\) A. Ashtekar and J. Pullin, Syracuse Preprint 1990; K. Schleich, talk at Quantum Cosmology Workshop, UBC, May 1990; E. Calzetta, “Cosmology, Entropy and Chaos” IAFE Preprint, Buenos Airs 1990.
55\) C. H. Woo, Phys. Rev. ${\sl D39} (1989) 3174, {\sl D41} (1990)
1355,$ and in Ref. 3; K. Svozil, Phys. Rev. ${\sl D41} (1990) 1353.$
56\) A. Bohr and B. R. Mottelson, Nuclear Structure, Vol. 2 (Benjamin, New York, 1975); K. Goeke and P. -G. Reinhard (eds.) Time-Dependent Hartree-Fock and Beyond (Lecture Notes in Physics 171) (Springer, Berlin, 1982).
57\) D. Forster, Hydrodynamics Fluctuations, Broken Symmetry and Correlation Functions (Benjamin, New York, 1975).
58\) See e.g., J. J. Halliwell, Phys. Rev. ${\sl D39} (1989) 2912$; T. Padmanabhan ibid ${\sl D39} (1989) 2924$ and references therein.
59\) See e.g., T. P. Singh and T. Padmanabhan, Ann. Phys. (N.Y.) [*196*]{} (1989) 296 and references therein.
60\) There are scattered work on this aspect but are mostly incomplete or incorrect, see, e.g., M. Morikawa, Phys. Rev. ${\sl D33} (1986) 3607,$ ibid ${\sl D40} (1989) 4023$; ibid ${\sl D42} (1990) 1027.$
61\) J. M. Cornwall and R. Bruinsma, Phys. Rev. ${\sl D38} (1988)
3146.$
62\) In this workshop Professor Su kindly quoted me a paper he wrote \[Z. B. Su, L. Y. Chen, X. T. Yu, K. C. Chou, Phys. Rev. ${\sl B37} (1988)
9810]$ which shows the equivalence of these two formalisms.
63\) J. P. Paz, Nonequilibrium Quantum Fields in Cosmology, this volume.
64\) J. J. Halliwell, Phys. Rev. ${\sl D36} (1987) 3626$; H. Kodama, Prog. Theor. Phys. (1989); S. Habib and R. Laflamme, “Wigner Function and Decoherence in Quantum Cosmology”, UBC Preprint (1990); A. Anderson, Univ. Utah Preprint (1990); J. P. Paz and S. Sinha, “On Decoherence and Correlation” in preparation.
65\) E. Calzetta and F. D. Mazzitelli, “On Decoherence and Particle Creation” IAFE Preprint (1990).
66\) B. L. Hu and Y. Zhang, “Coarse Grained Effective Action, Renormalization Group Transformation and Inflationary Cosmology” University of Maryland Preprint (1990).
67\) B. L. Hu and S. Sinha, “Coarse-Grained Effective Action, Backreaction and Minisuperspace Approximation” University of Maryland preprint (1990).
68\) S. W. Hawking, Nature [*248*]{} (1974) 30.
69\) S. W. Hawking, Comm. Math. Phys. [*87*]{} (1982) 395; S. W. Hawking, Physica Scripta ${\sl T15} (1986)$; D. Page, Phys. Rev. ${\sl D34}
(1986) 2267.$
70\) R. Penrose, in: General Relativity, An Einstein Centenary (eds.) S. W. Hawking and W. Israel (Cambridge Univ., Cambridge, 1979); B. L. Hu, Phys. Lett. ${\sl 97A} (1983) 368.$
[^1]: Work supported in part by the National Science Foundation under Grant No. PHY87-17155.
[^2]: Invited talk given at the Second International Workshop on Thermal Field Theories and Their Applications, Tuskuba, Japan, July 1990. Published in [*Thermal Field Theories*]{}, edited by H. Ezawa, T. Aritmitsu and Y. Hashimoto (North-Holland, Amsterdam, 1991) p. 233-252.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Fermi-LAT analyses show that the $\gamma$-ray photon spectral indices $\Gamma_\gamma$ of a large sample of blazars correlate with the $\nu F_\nu$ peak synchrotron frequency $\nu_{s}$ according to the relation $\Gamma_\gamma = d - k \log\nu_{s}$. The same function, with different constants $d$ and $k$, also describes the relationship between $\Gamma_\gamma$ and peak Compton frequency $\nu_{\rm C}$. This behavior is derived analytically using an equipartition blazar model with a log-parabola description of the electron energy distribution (EED). In the Thomson regime, $k = k_{EC} = 3b/4$ for external Compton processes and $k = k_{SSC} = 9b/16$ for synchrotron self-Compton (SSC) processes, where $b$ is the log-parabola width parameter of the EED. The BL Lac object Mrk 501 is fit with a synchrotron/SSC model given by the log-parabola EED, and is best fit away from equipartition. [ Corrections are made to the spectral-index diagrams for a low-energy power-law EED and departures from equipartition, as constrained by absolute]{} jet power. Analytic expressions are compared with numerical values derived from self-Compton and external Compton scattered $\gamma$-ray spectra from Ly $\alpha$ broad-line region and IR target photons. The $\Gamma_\gamma$ vs. $\nu_s$ behavior in the model depends strongly on $b$, with progressively and predictably weaker dependences on $\gamma$-ray detection range, variability time, and isotropic $\gamma$-ray luminosity. Implications for blazar unification and blazars as ultra-high energy cosmic-ray sources are presented. [ Arguments by Ghisellini et al. (2014) that the jet power exceeds the accretion luminosity depend on the doubtful assumption that we are viewing at the Doppler angle.]{}'
author:
- 'Charles D. Dermer, Dahai Yan, Li Zhang, Justin D. Finke, & Benoit Lott'
title: 'Near-Equipartition Jets with Log-Parabola Electron Energy Distribution and the Blazar Spectral-Index Diagrams '
---
Introduction
============
Searches for an ordering principle in blazar science have met with limited success. One of the most debated is the [*blazar sequence*]{}, in which blazar data seem to show an inverse correlation between apparent isotropic synchrotron luminosity $L_{syn}$ and peak synchrotron frequency $\nu_s$ of the blazar $\nu F_\nu$ spectral energy distribution (SED) [@fos98; @smu96]. This behavior, which is mirrored in the $\gamma$-ray regime, has been interpreted in terms of cooling processes [@ghi98; @bd02; @fin13]. The validity of the blazar sequence has, however, been criticized [@gio12; @gpp13] as possibly resulting from spurious correlations introduced by combining samples from radio and X-ray blazar surveys, problems from redshift incompleteness, and confusing lineless BL Lac objects that lack accretion disk with those where the BLR radiation is overwhelmed by beamed emission. Contrary to the simple blazar sequence, @Meyer present evidence for the existence of two separate tracks in the $L_{syn}$ vs. $\nu_s$ plane, including radio galaxies in the blazar-sequence plot.
A second strong correlation is the [*blazar divide*]{}. From the first three months of Fermi Large Area Telescope (LAT) blazar data, [@gmt09] argued that hard ($\Gamma_\gamma <2$) $\gamma$-ray spectrum blazars are associated with sources radiating isotropic $\gamma$-ray luminosities $L_\gamma\lesssim 5\times 10^{46}$ erg s$^{-1}$, while soft ($\Gamma_\gamma > 2$) $\gamma$-ray blazars are more likely to be at larger values of $L_\gamma$. From the Second Fermi LAT AGN (2LAC) data [@2LAC], a broad divide is evident in the direct data at $L_\gamma\cong 10^{46}$ erg s$^{-1}$ [Figs. 37 and 38 in @2LAC], though with no other apparent dependence of $\Gamma_\gamma$ on $L_\gamma$ in the ranges $10^{44} \lesssim L_\gamma\lesssim 10^{46}$ erg s$^{-1}$ and $10^{46} \lesssim L_\gamma\lesssim 10^{49}$ erg s$^{-1}$. In terms of a beaming-corrected Eddington ratio $\ell_{\rm Edd}$ for a black hole with mass $\sim 10^9$ M$_\odot$, this could imply a transition from an inefficiently radiating ADAF-type flow at $\ell_{\rm Edd} \lesssim 0.01$ to a thick disk when $\ell_{\rm Edd} \gtrsim 0.01$ [@gmt09]. The $\gamma$-ray Compton dominance ${\cal A}_{\rm C}$, which is essentially the ratio of the bolometric $\gamma$-ray and synchrotron luminosities, also strongly correlates with $\nu_s$ [@fos98; @fin13]. Definitive interpretations of blazar sequence and blazar divide data are hampered by redshift incompleteness. BL Lac objects without redshift information may themselves constitute separate populations in the $L_{syn}$ vs. $\nu_s$ or $\Gamma_\gamma$ vs. $L_\gamma$ planes, though large efforts have been made to provide complete, or at least redshift-constrained samples of blazar data [@sha13; @aje14]. [ The ${\cal A}_{\rm C}$ vs. $\nu_s$ distributions of 2LAC blazars with and without redshift do not significantly differ [@fin13].]{}
A third robust correlation in blazar physics relates $\gamma$-ray spectral index $\Gamma_\gamma$ with peak synchrotron ($\nu_s$) or peak Compton $\gamma$-ray ($\nu_{\rm C}$) frequencies (in this study, we assume that the blazar SEDs are made by leptonic processes only).[^1] These [*spectral-index diagrams*]{} for FSRQ and BL Lac blazars have been reported in the First LAT AGN Catalog [1LAC, Fig. 13 in @1LAC], the Fermi Bright Blazar SED paper [Fig. 29 in @SED], the 2LAC [Fig. 17 in @2LAC], and the 3LAC [Fig. 10 in @3LAC]. The distributions of spectral indices of the entire BL Lac and FSRQ blazar samples follow a pattern, with large scatter, described by the relation $\Gamma_\gamma = d - k \log \nu_{14}$, where $\nu_{s} = 10^{14}\nu_{14}$ Hz. For the entire sample of FSRQs and BL Lac objects, the value $k = 0.18\pm0.03$ is found in @3LAC. A similar function, with different values of $d$ and $k$, apply to the $\Gamma_\gamma$ vs. $\nu_{\rm C}$ data. The spectral-index distribution of BL Lac objects with unknown redshift is generally consistent with the distribution of BL Lac objects with known redshift [@2LAC; @3LAC].
In this paper, we use an equipartition blazar modeling approach [@cer13; @der14a] [bf assuming a log-parabole electron energy distribution (EED)]{} to explain the blazar spectral-index diagrams. In Section 2 we derive analytic Thomson-regime expressions for the relationship between $\Gamma_\gamma$ and $\nu_s$, depending on whether the $\gamma$ rays are made through external Compton (EC) or synchrotron self-Compton (SSC) processes. Because of the equipartition relations, the expressions depend on $b$, $\nu_{s}$, variability time $t_{var}$, bolometric isotropic synchrotron luminosity $L_{syn}$, an equipartition parameter $\zeta_e$ and a radiation parameter $\zeta_s$. The simpler $\Gamma_\gamma$ vs. $\nu_{\rm C}$ expressions are also obtained. The derived analytic relations, [ confirmed by numerical modeling,]{} are shown in Section 3 to be in [ general]{} accord with the blazar spectral-index diagram data, whether external radiation fields in the jet environment are present or absent. [ The effects of a log-parabola EED with a low-energy power-law component are also considered.]{}
In Section 4, application of the equipartition model to the BL Lac object Mrk 501 is demonstrated, and effects of departures from equipartition are evaluated. Trends in spectral-index behavior with other observables constrained can be tested with correlated Fermi-LAT and multiwavelength data, and how this work relates to the blazar sequence and blazar divide, and blazars as UHECR sources, are discussed in Section 5. [ The work is summarized in Section 6.]{}
Appendix A gives a Thomson-regime derivation of the SSC spectrum with a log-parabola electron distribution, and Appendix B gives a jet power-analysis. [ There we show that the assumption that blazars are typically observed at the Doppler beaming angle may have led @ghi14 to overestimate the absolute jet power.]{} Indeed, out-of-equipartition models are ultimately constrained by demands for power.
Equipartition Blazar Modeling with Log-Parabola Electron Energy Distribution
============================================================================
A standard blazar-jet model, treated in innumerable blazar spectral modeling papers [see @bhk12 for review], starts with magnetized plasma that is ejected at relativistic speeds along the poles of a rotating black hole.[^2] The jet plasma, which entrains thermal and nonthermal particles in a hypothetical tangled and randomly oriented magnetic field, is a source of escaping photons, and potentially also of escaping cosmic rays and neutrinos. The jet power is extracted from the mass energy of accreting matter and/or the rotational energy of the black hole itself. The collimated relativistic plasma outflow, an exhaust byproduct of the energy generated by the black-hole engine, is usually attributed to processes taking place in the magnetosphere of the rotating black hole. The polarized broad-band synchrotron radiation emitted by an energetic EED (which could also contain positrons) is boosted by the Doppler effect along the jet axis, so that rapidly variable jet synchrotron radiation can be detected by Earth-based observatories from large redshift ($z\gg 1$) sources.
The jet electrons also Compton scatter ambient photons to $\gamma$-ray energies. Besides the accompanying SSC emission from target synchrotron photons [e.g., @mgc92; @bm96], EC $\gamma$ rays are made when the nonthermal jet electrons scatter photons from external radiation fields. Depending on jet Doppler factor $\delta_{\rm D}$ and BLR cloud parameters, the direct accretion-disk radiation field dominates the external radiation field of a powerful FSRQ at $\ll 10^3$ Schwarzschild radii, BLR fields are strongest within $\sim 0.3$ pc [@dsm92; @sbr94; @ds02], while at the pc scale and beyond, infrared radiation from a surrounding IR-emitting dust torus would have the largest energy density of all ambient radiation fields [@bla00; @sik09; @gt09] in the inner blazar-jet environment.
The form of the nonthermal EED is often treated by either assuming a nonthermal injection spectrum of leptons that evolves in response to adiabatic and radiative losses, or by assuming a form for the average steady-state EED in the radiating jet plasma. Adopting the latter approach, we assume that the 3 parameter log-parabola function $${\gamma}^{\prime 2} N_e^\prime({{\gamma^\prime}})=[\gamma_{pk}^{\prime 2}N^\prime_e({\gamma}^\prime_{pk})]({{\gamma}^\prime\over{\gamma}^\prime_{pk}})^{-b\log({{\gamma}^\prime\over \gamma_{pk}^\prime})}
\equiv K^\prime y^{- b\log y}\;
\label{g2Ng}$$ provides an approximate description of the nonthermal lepton spectrum. Here $y\equiv \gamma^\prime/\gamma^\prime_{pk}$, $\gamma_{pk}^\prime$ is the peak, or principal, Lorentz factor of the fluid-frame EED, eq. (\[g2Ng\]). The value of $K^\prime$ can be related to either the total particle number or total comoving particle energy [@der14a]; in the latter case, $K^\prime = {\cal E}^\prime_e/m_ec^2 \sqrt{\pi\ln 10/b}$, where ${\cal E}^\prime_e$ is the nonthermal electron energy of the blob. The continuously curving EED given by a log-parabola function derives from stochastic acceleration processes with radiation and escape [see, e.g., @mas04; @bld06; @tra07; @tra11; @sp08]. With this form of the EED, GeV breaks in FSRQs and blazars with $\nu_{pk}^{syn} \lesssim 10^{14}$ Hz are shown to arise from the onset of Klein-Nishina effects when scattering BLR photons [@cer13; @ack10; @tg08], and to give [@der14a] reasonable fits to four epochs of quasi-simultaneous multiwavelength observations of 3C 279 [@hay12]. As we show below, this approach also gives good fits to the SED of Mrk 501, though the best fits are achieved with an electron distribution out of equipartition with the magnetic field.
The comoving synchrotron luminosity $$L_{syn}^\prime = c \sigma_{\rm T} {B^{\prime 2}\over 6\pi} \int_1^\infty d\gamma^\prime \gamma^{\prime 2} N_{e}( \gamma^\prime)\;
\label{Lsyn}$$ implies, using Eq. (\[g2Ng\]) and a $\delta$-function approximation for the synchrotron photon with average dimensionless energy $\e_{syn} = (3/2)\delta_{\rm D} (B^\prime/B_{cr})\gamma^{\prime 2}$ [@dm09], the received synchrotron luminosity spectrum $$\e L_{syn}(\e ) = \upsilon x^{1-\hat b \ln x}= \upsilon ({\epsilon\over\e_{pk}})^{{1\over 2}-{b\over 4}\log ({\e/\e_{pk}})}\;,
\label{Lsyne}$$ where $x = \sqrt{\e/\e_{pk}}$, $\upsilon = f_3 L_{syn}$, and $f_3^{-1} = 2\cdot 10^{1/4b}\sqrt{\pi \ln 10/b}$ [@der14a]. Thus the effective log-parabola width parameter $b_{sy}$ for the synchrotron spectrum is given by $b_{sy} = b/4$ in the $\delta$-function approximation, and $b_{sy} \cong b/5$ when using the full Thomson cross section [@mas06; @pag09]. The peak synchrotron frequency $\epsilon_{pk} = (3/2)\delta_{\rm D}(B^\prime/B_{cr})\gamma_{pk}^{\prime 2}$. The slope of the $\e L_{syn}(\e )$ spectrum is $$\alpha_\nu \equiv {d \ln[\e L_{syn}(\e )]\over d\ln \e } = {1\over 2} [1- b \log (\e/\e_{pk})]\;.
\label{alphanu}$$ [@mas04a]. Because the nonthermal electron energy-loss rate from synchrotron processes scales quadratically with electron Lorentz factor $\gamma^\prime$, the synchrotron spectrum from a log-parabola distribution of electrons has a $\nu L_\nu$ synchrotron peak energy $\epsilon_s = h\nu_s/m_ec^2$ that is shifted to higher values than $\epsilon_{pk}$. Eq. (\[alphanu\]) shows that $\epsilon_s = 10^{1/b} \e_{pk}$ [@mas06].
[cccccccccccc]{} ${\rm }$ & Coef. & $L_{48}$ & $\nu_{14}$ & $t_4$ & & $\zeta_s$ & $\zeta_e$ & & $f_0$ & $f_1$ & $f_2$\
$$ & & $ $ & & & & & & & &\
& & & & & & & & & & &\
$\delta_{\rm D}$ & 17.5 & $~~3/16$ & $~~1/8$ & $-1/8$ & & $-7/16$ & $~~1/4$ & & $-7/16$ & $-1/4$ & $-1/8$\
$B^\prime({\rm G})$ & 5.0 & $-1/16$ & $-3/8$ & $-5/8$ & & $~13/16$ & $-3/4$ & & $~13/16$ & $~~3/4$ & $~~3/8$\
$\gamma_{pk}^\prime$ & 523 & $-1/16$ & $~~5/8$ & $~~3/8$ & & $-3/16$ & $~~1/4$ & & $-3/16$ & $-1/4$ & $-5/8$\
& & & & & & & &\
${\cal E}^b$ & $1.4$ & $5/16$ & $-1/8$ & $1/8$ & & $-1/16$ & $-1/4$ & & $-11/16$ & $~~1/4$ & $~~1/8$\
$L_{jet,B}^c$ & $4$ & $5/8$ & $-1/4$ & $1/4$ & & $-1/8$ & $-1/2$ & & $-1/8$ & $~~1/2$ & $~~1/4$\
\
\[table1\]
$^a$ So, e.g., $\delta_{\rm D} \cong 17.5 L_{48}^{3/16}(\nu_{14}/f_2t_4)^{1/8} (f_0\zeta_s)^{-7/16}(\zeta_e/f_1)^{1/4}$, etc.\
$^b$ ${\cal E} =E_{max}(10^{20}~eV)/Z$\
$^c$ Absolute power in magnetic field, units of $10^{44}~$erg s$^{-1}$\
Table \[table1\] shows the various dependencies of blob properties on the observables $L_{48}$, $\epsilon_s$ (or $\nu_{14}$), $t_{var}$ and $b$, and on the equipartition factor $\zeta_e$ and radiative factor $\zeta_s$. The factor $\zeta_e$ is the ratio of nonthermal electron energy density $u^\prime_e$ to magnetic-field energy density $u^\prime_{B^\prime}= B^{\prime 2}/8\pi$, and $\zeta_s$ is the ratio the jet-frame synchrotron photon energy density and $u^\prime_{B^\prime}$. In the blob scenario, the geometry factor $f_0 = 1/3$. The $b$-dependent factors are $f_1 = 10^{-1/4b}$, $f_2 = 10^{1/b}$, and $f_3 = (2\cdot 10^{1/4b} \sqrt{\pi \ln 10/b})^{-1}$ [@der14a], so $\epsilon_s = f_2 \e_{pk}$.
The similarity of the underlying physics of the synchrotron and Compton processes [@bg70] means that an expression like eq. (\[alphanu\]) holds for Compton scattering in the Thomson regime, except now $\epsilon_{pk}$ is replaced by a corresponding peak photon energy for EC and SSC processes [e.g., @pag09]. In the EC case, $\epsilon_{pk, EC} = (4/3)\delta_{\rm D}^2 \epsilon_0 \gamma_{pk}^{\prime 2}$, assuming an isotropic monochomatic external radiation field with energy $\e_0$ and energy density $u_0$. From the equipartition relations [@der14a] shown in Table 1 for the photon spectral index $\Gamma_\gamma = 2-\alpha_\nu$, we find that the photon index for EC processes is given by $$\Gamma_{\gamma}^{EC} \cong {17\over 8} + {b\over 2}\,\log\left({f_0^{5/4}E_{GeV} \zeta_s^{5/4}}\over \epsilon_{Ly\alpha}t_4^{1/2} \zeta_e L_{48}^{1/4} \right )
- {3b\over 4}\log \nu_{14}\;. \\
\label{GammagammaEC}$$ Here $E_{GeV}$ is the effective detection energy in GeV, and $\e_0 = 2\times 10^{-5}\epsilon_{Ly\alpha }$ for Ly $\alpha$/BLR scattering. A nominal value of $E_{GeV} = 1$ is chosen because the Fermi-LAT is most sensitive at $\approx 1$ GeV [Fig. 18 in @1FGL for a $\Gamma_\gamma= 2.2$ source spectrum]. The dependence of $\Gamma_\gamma$ on $E_{GeV}$ can be studied by analyzing Fermi-LAT data in discrete energy ranges.
Scattering the dusty torus emission, with IR photon energies corresponding to $\epsilon_{Ly\alpha }\sim 0.02$, implies a Thomson spectrum softer by $\Delta\Gamma_\gamma \cong 0.85 b$, because $\gamma$-ray photons at a given observing energy are produced in the softer part of the Compton-scattered spectrum when the target photons have lower energies. If equipartition is instead made to total particle energy density $u^\prime_{tot}$ according to the factor $\zeta_{eq} =u^\prime_{tot}/u^\prime_{B^\prime}$, then $\zeta_e = \zeta_{eq}/(1+\eta_{bl})$ and $\eta_{bl} = u^\prime_{baryons}/u^\prime_{e}$ is the baryon loading, and $u^\prime_{baryons}$ is the internal energy density in protons and ions [@dmi14 and Appendix B].
The specific spectral synchrotron luminosity, from eq. (\[Lsyne\]) in the $\delta$-function approximation and results of @der14a, is given by $$\e L_{syn}(\e,\Omega) = f_3 N_e \, {4\over 3}\, c \sigma_{\rm T} {B^{\prime 2}\over 8\pi }\,\gamma_{pk}^{\prime 2} \delta_{\rm D}^4
x^{1-b\log x}\;,
\label{eLsynesOmegas}$$ where $x = \sqrt{\e/\e_{pk}}$, and $\e_{pk} = 4 \delta_{\rm D} B^\prime \gamma_{pk}^{\prime 2}/3B_{cr}$. The specific spectral $\gamma$-ray luminosity in the Thomson regime for a jet traveling through an external isotropic, monochromatic radiation field with frequency $m_e^2 \epsilon_0/h$ and energy density $u_0$, in units of $m_ec^2$ cm$^{-3}$, using a $\delta$-function approximation for Thomson scattering, is $$\e L_{EC}(\e,\Omega) \cong \,f_3 N_e \, {4\over 3}\, c \sigma_{\rm T} u_0\,\gamma_{pk}^{\prime 2} \delta_{\rm D}^6
{\tt v}^{1-b\log {\tt v}}\;,
\label{eLECesOmegas}$$ where ${\tt v} \equiv \sqrt{\e/\e_{pk,EC}}$ and $\e_{pk,EC} = (4/3)\delta_{\rm D}^2 \gamma_{pk}^{\prime 2}\e_0$. The technique of @gkm01 is used to derive this expression. The ratio of the spectral synchrotron and Thomson luminosities at their respective peak frequencies is $\delta_{\rm D}^2 u_0/u^\prime_{B^\prime }$.
For the SSC process, the combined effects of the widths of both the EED and the target synchrotron photon spectrum will broaden the Compton-scattered photon spectrum such that its effective width in the Thomson regime is obtained by replacing $b$ by $b_{SSC} = b/2$ in eq. (\[alphanu\]) [@pag09] and replacing $\e_{pk}$ by $\e_{pk,SSC} = 2\delta_{\rm D} (B^\prime/B_{cr}) \gamma_{pk}^{\prime 4}$, giving $$\Gamma_{\gamma}^{SSC} = {65\over 32} + {b\over 4}\,\log\left( {6.5\times 10^3 E_{GeV}f_{0}^{3/8} \zeta_{s}^{3/8}L_{48}^{1/8}\over \zeta_e^{1/2} t_4^{3/4} }\right)$$ $$- {9b\over 16}\log \nu_{14}\;,
\label{GammagammaSSC}$$ from Table 1. Note that the $\nu L_\nu$ peak SSC frequency is a factor $10^{2/b}$ larger than $\e_{pk,SSC}$. The SSC expression is justified by a more detailed derivation in Appendix A. The uncertainty $\Delta \Gamma_\gamma$ in the spectral index related to geometrical uncertainties can be estimated by letting $f_0$ range from unity for a blast-wave shell geometry to $f_0 = 1/3$ for a comoving spherical-blob geometry. From eqs. (\[GammagammaEC\]) and (\[GammagammaSSC\]), one can see that this translates into an uncertainty $\Delta \Gamma_\gamma^{EC} \cong 0.30 b$ for EC processes and an uncertainty $\Delta \Gamma_\gamma^{SSC} \cong 0.04 b$ for SSC processes.
We also derive the Thomson-regime expressions $${\Gamma^{EC,\gamma}_\gamma = {2} +{b\over 2}\log (2.4 E_{GeV}) - {b\over 2}\log \nu_{23}\;}
\label{Gg}$$ for $\Gamma_\gamma$ vs. $\nu_C$ in EC processes, and $${\Gamma^{SSC,\gamma}_\gamma = 2 +{b\over 4}\log (2.4 E_{GeV}) - {b\over 4}\log \nu_{23}\;}
\label{GgSSC}$$ for $\Gamma_\gamma$ vs. $\nu_C$ in SSC processes. Here $\nu_{23} = \nu_C/10^{23}$ Hz is the peak frequency of the Compton component of the $\nu L_\nu$ SED. Note that eq. (\[Gg\]) is independent of the target photon energy, because the expression assumes that the EED and Doppler factor are adjusted to produce a Compton-scattered $\gamma$-ray spectrum that peaks at $\nu_{\rm C}$.
![Data are the $> 100$ MeV photon spectral index values $\Gamma_\gamma$ as a function of peak synchrotron frequency $\nu_s$ for blazars from the 2LAC [@2LAC]. Red, blue, green, and black symbols identify, respectively, FSRQs, BL Lac objects with redshifts, BL Lac objects without redshifts, and blazars with data too poor to determine if the source is an FSRQ or a BL Lac object. [*Left:*]{} Curves labeled by EC BLR, EC IR and SSC for EC processes with BLR photons, EC processes with IR photons and SSC processes, respectively, show $\Gamma_\gamma$ vs. $\nu_s$ predictions of the log-parabola equipartition model using standard parameters given by eq. (\[stanpar\]). Also, $\epsilon_0 = 2\times 10^{-5}$ and $u_0= 10^{-2}$ erg cm$^{-3}$ for Ly$\alpha$, and $\epsilon_0 = 4.6\times 10^{-7}$ and $u_0= 10^{-3}$ erg cm$^{-3}$ for the $\sim 1000$ K IR radiation. Thick curves give numerical calculations, and thin curves show analytic results, from eqs. (\[GammagammaEC\]) and (\[GammagammaSSC\]). [ The thick curves that approach constant values at large $\nu_s$ are numerical predictions for the power-law, log-parabola model, eq. (\[PLLP\]). ]{} [*Right:*]{} Compton-dominance ${\cal A}_{\rm C}$ as a function of $\nu_s$ for EC BLR, EC IR, and SSC processes, as labeled. The line with arrows has a slope of $+1$ in the ${\cal A}_{\rm C}$ vs. $\nu_s$ plane. []{data-label="fig1"}](f1.eps){width="3.5in"}
Modeling the Blazar Spectral-Index Diagram
==========================================
Fig. \[fig1\] shows measured values of Fermi-LAT spectral index $\Gamma_\gamma$ from the 2LAC [@2LAC] derived from a single power-law fit to the complete data set in the 0.1 – 100 GeV range for sources with $TS>25$.[^3] The red, blue, green, and black data symbols correspond, respectively, to $\gamma$-ray sources detected with the Fermi-LAT that have been associated with FSRQs, BL Lac objects with and without redshifts, and blazars with optical data too poor to determine if the source is an FSRQ or BL Lac.
From inspection of the plot, it is clear that a function of the form $\Gamma_\gamma = d - k \log \nu_{14}$ will provide a reasonable description of the data. For the entire FSRQ and BL Lac sample, but excluding other blazar candidates, values of $k = 0.18\pm 0.03$ and $d =2.25\pm 0.04$ are deduced in the 3LAC [@3LAC]. Comparing this value with the analytic expressions, eqs. (\[GammagammaEC\]) and (\[GammagammaSSC\]), a larger value of $b$ is implied for SSC processes compared to EC processes, but in both cases consistent with $b \approx 1/3$. The typical value of $b$ can also be deduced from the average nonthermal blazar synchrotron SED, when fit with an expression of the form of Eq. (\[Lsyne\]). From X-ray analysis of Beppo-SAX data on Mrk 501, @mas04 finds values of $b_{sy}$ ranging from 0.12 – 0.33, implying a corresponding log-parabola width parameter $b\gtrsim 0.5$. Narrow bandwidth modeling of X-ray synchrotron emission from Mrk 421 gives $b_{sy} \cong 0.3$ – 0.5 [@tra07], though values of $b_{sy} = 0.17 \pm 0.02$ (2006 15 July pointing), $b_{sy} = 0.11 \pm 0.02$ (2006 April 22 pointing), and $b_{sy} = 0.08 \pm 0.03$ (2006 June 23 pointing) are obtained in more complete joint XRT-BAT analysis [@tra09], consistent with an electron distribution with $b \cong 5 b_{sy} \approx 0.5$. @chen14 finds that $b_{sy}$ is distributed in the range $0.05 \lesssim b_{sy} \lesssim 0.25$, implying $0.25 \lesssim b \lesssim 1.25$. More importantly, he finds a dependence of $b_{sy}$ on $\nu_s$, which we discuss further in Section \[sec:ds\]. The values of $b$ deduced from spectral modeling tend to be larger than obtained from the slope implied by the spectral-index diagram.
Standard parameters in log-parabola model
------------------------------------------
To compare the log-parabola equipartition model with data, we adopt a [standard]{} parameter set, and take $$b=1/2,\;t_4 = L_{48} = \zeta_{e} = \zeta_{s} = E_{GEV} = 1\;.
\label{stanpar}$$ The reasoning driving the choice of the standard variability time scale is that the masses of supermassive black holes powering blazars—both FSRQs and BL Lacs—are typically of the order $\sim 10^9 M_\odot$. The value $t_4 \cong 1$ or $t_{var} \cong 3$ hr corresponds to the light-crossing time across a size equal to the Schwarzschild radius of a $\sim 10^9 M_\odot$ black hole, though of course shorter variability time scales have been recorded during spectacular outbursts of BL Lac objects, including Mrk 421 [@fos08MRK421], Mrk 501 [@alb07Mrk501], and PKS 2155-304 [@aha07PKS2155], not to mention the extraordinary VHE outburst observed with the MAGIC telescope from the FSRQ PKS 1222$+$216 with $t_{var} \sim 10$ m [@PKS1222]. The isotropic synchrotron luminosity $L_{syn}$ can exceed the Eddington limit $L_{\rm Edd}$, though $L_{\rm Edd}$ is presumably the upper limit to the persistent absolute jet power (see App. B). Standard values $L_{48}\sim 0.1$ – 1 and $L_{48}\sim 10^{-2}$ – $10^{-3}$ are typical of powerful FSRQs and BL Lac objects, respectively. At the other side of the time domain, $t_{var} \sim 10^5$ – $10^6$ s may be compatible with quiet times of blazars.
Fig. 1 shows analytic results of Eqs. (\[GammagammaEC\]) and (\[GammagammaSSC\]) for $\Gamma_\gamma$ as a function of $\nu_{\rm pk}^s = \nu_s$, using the standard parameter set. Results of numerical calculations, obtained by modifying the code used in @der14a, are also shown. The dimensionless photon energies for the BLR and IR photons used in the model are $\epsilon_0 = 2\times 10^{-5}$ (i.e., 10.2 eV) for BLR photons and $\epsilon_0 = 4.6\times 10^{-7}$ for warm IR torus dust emission described by an $\approx 1000$ K greybody spectrum with $\approx 15$% covering factor, giving an energy density of $\approx 10^{-3}$ erg cm$^{-3}$. The analytic results are shown by the thin lines. The numerical results are shown by the thick curves. As can be seen, the analytic SSC and EC IR results are in reasonable agreement with the numerical calculations, whereas the analytic EC BLR results do not agree with the numerical results. Klein-Nishina effects already make themselves felt strongly for target BLR photons scattered to 1 GeV, but only weakly for target IR photons scattered to 1 GeV, as is clear by noting that KN effects set in at photon energies $E_\gamma\gtrsim m_ec^2/12\epsilon_0 \approx 100$ GeV for 1000 K photons, and $E_\gamma \approx 2$ GeV for Ly $\alpha$ photons. The Thomson-regime expressions are harder than the numerical curves because of the Klein-Nishina softening.
[ Fig. 1 also shows the effects of a low-energy power-law extension of the EED on the spectral-index diagrams. In such a power-law log-parabola (PLLP) model with a low-energy cutoff Lorentz factor $\gamma^\prime_{min}$ [@yan13; @pyz14], the EED distribution extends eq. (\[g2Ng\]) by two parameters to take the form $$\gamma^{\prime 2}N_e^\prime(\gamma^\prime) = K_e^\prime \,[y^{2-s}H(y;y_{\ell},1)+y^{2-s-r\log y} H(y-1)]\;.
\label{PLLP}$$ Here $s$ is the power-law spectral index of the low-energy component, $r$ is a log-parabola width parameter, and $y_{\ell} = \gamma^\prime_{min}/\gamma_{pk}$. The Heaviside functions are defined such that $H(u) = 1$ when $u\geq 0$ and $H(u) = 0$ otherwise, and $H(u;a,b) = H(u-a)H(b-u)$. The theoretical basis for the form of eq. (\[PLLP\]) is discussed below. Results are shown for $s = 2$ and $y_\ell \ll 1$, in which case $r\rightarrow b$, reducing the PLLP model to a 3-parameter model. ]{}
Compton dominance
------------------
The numerical results for this particular set of parameters are seen to follow the trend of much of the data. Virtually no FSRQs are observed, however, with $\nu_{14} > 1$. To obtain some insight into this, we calculate the Compton dominance ${\cal A}_{\rm C}$ for our model, defined here as the ratio of the 100 MeV – 100 GeV $\gamma$-ray luminosity to the bolometric synchrotron luminosity. It is calculated from the relation $${\cal A}_{\rm C} \equiv {L_\gamma(100~{\rm MeV})\over \alpha_\nu L_{syn}}\;[({100\over E_{GeV}})^{\alpha_\nu} - ({0.1\over E_{GeV}})^{\alpha_\nu}]
\label{AC}$$ where $\alpha_\nu$ and $L_\gamma(100~{\rm MeV})$ are, respectively, the $\nu L_\nu$ spectral index and luminosity calculated at $E_{GeV}$ GeV. Note that a more detailed and time-intensive calculation would integrate the blazar SED to determine ${\cal A}_{\rm C}$. The Compton dominance depends on the energy density of the surrounding radiation fields. For definiteness, we have taken $u_{BLR} = 10^{-2}$ erg cm$^{-3}$ and $u_{IR} = 10^{-3}$ erg cm$^{-3}$ in our calculations. Note that ${\cal A}_{\rm C}$ scales approximately linearly with $u_0$. As ${\cal A}_{\rm C}$ becomes progressively smaller, the corresponding blazars becomes progressively less detectable as $\gamma$-ray sources. So solutions should be restricted to a minimum value of ${\cal A}_{\rm C}$. Solutions should also be restricted at large values of ${\cal A}_{\rm C}$, because Compton drag on the jet becomes a strongly limiting factor, as discussed more in Section \[sec:ds\]. Regions where $0.1 \lesssim {\cal A}_{\rm C} \lesssim 30$ may favor LSP blazars to be FSRQs, ISP blazars to the EC BLR, EC IR, and SSC solutions in Fig. \[fig1\], as these values bracket measured values of the Compton dominance [Fig. 7 in @fin13].
![Same as Fig. \[fig1\], except that $b = 1$. Heavy and light downward-going curves are the numerical and analytic equipartition model predictions, respectively, and upward going curves show Compton dominance for EC BLR, EC IR and SSC processes [ for the log-parabola EED, eq. (\[g2Ng\]).]{} [ The thick curves approaching constant values at large values of $\nu_s$ correspond to spectral-index predictions of the PLLP model, eq. (\[PLLP\]), with a $-2$ number index of the low-energy EED.]{} []{data-label="fig2"}](f2.eps){width="3.5in"}
![Same as Figs. \[fig1\] and \[fig2\], except that $E_{GeV} = 0.1$. []{data-label="fig3"}](f3.eps){width="3.5in"}
![Same as Figs. \[fig1\] and \[fig2\], except that $t_{4} = 100$. []{data-label="fig4"}](f4.eps){width="3.5in"}
In Fig. \[fig1\], we calculate three models in the $\Gamma_\gamma$ vs. $\nu_s$ plane corresponding to complete dominance either of Ly $\alpha$ BLR radiation (EC BLR), IR radiation from the dusty torus (EC IR), or internal synchrotron radiation (SSC) as the target photon source. Restricting the Compton dominance to $0.1\lesssim {\cal A}_{\rm C} \lesssim 30$ suggests that most blazars with $\nu_{14} < 0.1$ have $\gamma$ rays that result from scattered BLR radiation, while blazars with $0.1 \lesssim \nu_{14}\lesssim 1$ would have a mix of blazars with $\gamma$ rays made by Compton scattering of either BLR or IR photons, or both. At higher peak synchrotron frequencies, SSC-dominated sources would be most plentiful.
[ The use of a different model, the PLLP EED, eq. (\[PLLP\]) with $s=2$ and $y_\ell \ll 1$, is displayed in Fig. \[fig1\] and subsequent figures by the numerically calculated spectral index curves that approach constant values of spectral index at $\nu_{14}\gg 1$. Klein-Nishina effects, described more below, soften the spectral index below the Thomson regime value of $\Gamma_\gamma = 1.5$. It is interesting that essentially all data are softer than $\Gamma_\gamma = 1.5$, and that smaller values of external radiation energy density could yield typical measured Compton dominance values for ISP and HSP blazars with an external Compton $\gamma$-ray component making a significant contribution to the SED.]{}
Figs. \[fig2\] – \[fig4\] show how changes in the model parameters affect results. Fig. \[fig2\] shows that a value of $b = 1$ is incompatible with the combined trend of the data, though a values of $b\cong 1$ may be consistent with sub-populations, e.g., FSRQs. Returning to $b = 1/2$, Fig. \[fig3\] shows the effects of calculating the spectral index at $E_{GeV} = 0.1$, that is, at 100 MeV rather than 1 GeV. Because the Compton-scattered $\gamma$-ray SED becomes progressively softer at larger $\gamma$-ray energies, the model results in Fig. \[fig3\] are uniformly harder than in Fig. \[fig1\]. The discrepancy between the analytic and numerical results decreases when scattering Ly $\alpha$ radiation because the Klein-Nishina effects on the Compton cross section are not so great when scattering to 100 MeV as compared to 1 GeV. The dependence on detector energy $E_{GeV}$ should clearly show up in Fermi-LAT spectral index diagrams calculated in discrete energy ranges, e.g., 0.3 – 3 GeV and 3 – 30 GeV, and should, in a statistical study, discriminate between EC and SSC processes, though correlations between $b$ and $\nu_s$ can hide the effect.
Fig. \[fig4\] shows how a slower variability time, with $t_4 = 100$, affects the equipartition spectral-index diagram. Compared to the results in Fig. \[fig1\], the effect of longer variability times is to harden the spectrum. From eqs. (\[GammagammaEC\]) and (\[GammagammaSSC\]), the hardening for a factor of 10 longer variability time is $\Delta\Gamma_{\gamma}^{EC} = -b/4$ for EC processes and $\Delta \Gamma_{\gamma}^{SSC} = - 3b/16$ for SSC processes. Because of the difficulty in measuring $t_{var}$, the variability effect on spectral index may be too subtle to discriminate between EC and SSC processes. In a statistical sample, however, more rapidly variable sources at equipartition would in general be softer, assuming that there are no underlying correlations between $b$ and $t_{var}$, and that equipartition holds in the various states.
![Data points show the Fermi-LAT $\gamma$-ray spectral index evaluated in the range $0.1$ 1– 100 GeV as a function of $\nu F_\nu$ peak Compton frequency $\nu_{\rm C} = \nu_{pk}^{\rm C}$ of the blazar $\gamma$-ray SED [@SED]. In both panels, the range to calculate the Fermi-LAT spectral index is at 0.3 GeV and 3 GeV for the solid and dashed curves, respectively. [*(a), upper*]{}: Equipartition Thomson-model EC predictions (black) are shown along with numerical predictions evaluated for external 1000 K radiation fields from a dusty torus (magneta curves) and from Ly$\alpha$ radiation (orange), using parameters of Fig. \[fig1\] but with $b = 1/2$. [*(b), lower*]{}: Equipartition Thomson-model SSC predictions (black) are shown along with numerical SSC predictions resulting from synchrotron emission with $\nu_s = 10^{12}$ Hz and $10^{15}$ Hz, as labeled. []{data-label="fig5"}](f5a.eps "fig:"){width="3.3in"} 0.4in ![Data points show the Fermi-LAT $\gamma$-ray spectral index evaluated in the range $0.1$ 1– 100 GeV as a function of $\nu F_\nu$ peak Compton frequency $\nu_{\rm C} = \nu_{pk}^{\rm C}$ of the blazar $\gamma$-ray SED [@SED]. In both panels, the range to calculate the Fermi-LAT spectral index is at 0.3 GeV and 3 GeV for the solid and dashed curves, respectively. [*(a), upper*]{}: Equipartition Thomson-model EC predictions (black) are shown along with numerical predictions evaluated for external 1000 K radiation fields from a dusty torus (magneta curves) and from Ly$\alpha$ radiation (orange), using parameters of Fig. \[fig1\] but with $b = 1/2$. [*(b), lower*]{}: Equipartition Thomson-model SSC predictions (black) are shown along with numerical SSC predictions resulting from synchrotron emission with $\nu_s = 10^{12}$ Hz and $10^{15}$ Hz, as labeled. []{data-label="fig5"}](f5b.eps "fig:"){width="3.3in"}
Spectral index vs. peak Compton frequency
-------------------------------------------
Fig. \[fig5\] shows data from multiwavelength spectral analysis [@SED] of 48 bright blazars in the Fermi-LAT Bright AGN Sample [LBAS; @LBAS], separated into FSRQs, and low, intermediate, and high synchrotron-peaked (LSP, ISP, and HSP, respectively, defined by whether $\nu_s<10^{14}$ Hz, $10^{14}<\nu_s$(Hz)$<10^{15}$ Hz, or $\nu_s>10^{15}$ Hz) BL Lac objects. The upper and lower panels gives predictions for the dependence of $\Gamma_\gamma$ on $\nu_{\rm C}$ for the equipartition EC and SSC models. The Thomson-regime predictions, eqs. (\[Gg\]) for EC processes and eq. (\[GgSSC\]) for SSC processes, are plotted in black, depending on whether the $\gamma$-ray spectral index is measured at 0.3 GeV (solid curves) or 3 GeV (dashed curves). The index is softer when the $\gamma$-ray energy range used to determine the spectral index is larger, as noted above.
It is worth taking a moment to explain the deviations of the numerical curves from the Thomson-regime expressions. Suppose the detector waveband $E_{GeV}\gg h\nu_{\rm C}$, corresponding to the left potions of the figures for $E_{GeV} = 0.3$ – 3. Consider two $\gamma$-ray SEDs aligned at the same value of $\nu_{\rm C}$, one with strong Klein-Nishina effects and one in the Thomson regime. The SED with strong KN effects will be much softer at frequencies $\nu \gg \nu_{\rm C}$ by comparison with the one in the Thomson regime, causing the softer spectra when $h\nu_{\rm C}\ll E_{GeV}$, sometimes dramatically so, compared to SEDs formed by scattering in the Thomson regime.
At the other extreme $h\nu_{\rm C}\gg E_{GeV}$, corresponding to the right portions of the figures, the effects of strong KN losses is to harden the low-energy portion of the $\gamma$-ray SED compared to an SED formed by scattering in the Thomson regime (and with the same peak Compton frequency). Consequently, Klein-Nishina effects will produce harder spectra when the detector energy range is less than the peak Compton frequency compared to Thomson scattering.
Fig. \[fig5\] shows that an EC origin in either BLR or IR radiation is consistent with LSP FSRQ data, but is inconsistent with an SSC origin. A similar conclusion was reached earlier by examining the correlation of Compton dominance with core dominance in FSRQs and BL Lac objects [@mey12]. At values of $\nu_{\rm C} \gg 10^{23}$ Hz, or $E_{\rm C} \gg 1$ GeV, the sources are all ISP and HSP BL Lac objects, and are compatible with either an SSC or EC origin of the emission, given the uncertainties in $t_{var}$. In principle, however, an EC origin can be distinguished from an SSC origin by comparing the curvature of the $\gamma$-ray component with that of the synchrotron component.
Non-Equipartition Model for BL Lac Objects
==========================================
Blazars may be out of equipartition, though extremely out-of-equipartition blazars would be less favored because of the additional power required. Up to now, we have assumed that the equipartition parameter $\zeta_e = 1$, which minimizes jet power for a given synchrotron SED and variability time, assuming small baryon-loading. Modeling of 3C 279 with $\zeta_e = 1$ was possible in @der14a, though the very highest energy $\gamma$ rays were only successfully fit by using long variability times with $t_{var} \approx 10^5$ – $10^6$ s, in which case the X-ray emission was not well fit [cf. @hay12]. Better fits were found in the modeling of 3C 454.3 by taking $\zeta_e$ between 0.6 and 3.5 [@cer13], which has a minor effect on the spectral slope relation.[^4]
![Best-fit models for Mrk 501 for fixed $\zeta_e =1$ and letting $\zeta_e$ vary. Data from @abd11MRK501, with galactic feature removed.[]{data-label="fig6"}](f6.eps){width="3.8in"}
It is worth asking if an equipartition situation applies to BL Lac objects, which would be simpler than FSRQs by lacking significant external radiation fields. We apply the near-equipartition log-parabola (NELP) modeling technique to the 15 March 2009 – 1 Aug 2009 multiwavelength data of the HSP BL Lac object Mrk 501 [@abd11MRK501]. The data in Fig. \[fig6\] include OVRO radio observations, optical data, Swift UVOT and XRT data, GeV $\gamma$-ray data from Fermi-LAT, and VHE data from MAGIC. Parameter values are derived using the Markov Chain Monte Carlo (MCMC) technique of [@yan13] for Mrk 421 and [@pyz14] for Mrk 501, and using the 3-parameter log-parabola electron spectrum, eq. (\[g2Ng\]). The fit to the TeV data is always bad in the $\zeta_e=1$ case. The fit with $\zeta_e$ allowed to vary is obviously far better.
![Distribution of parameter values for the varying $\zeta_e$ case. The dashed curves represent the mean likelihoods of samples and the solid curves are the marginalized probabilities. []{data-label="fig7"}](f7.eps){width="3.5in"}
The distribution of parameter values derived from the MCMC technique for the data of Mrk 501 is shown in Fig. \[fig7\]. The dashed curves are the mean likelihoods of samples and the solid curves are the marginalized probabilities.[^5] In the fits, we run single chains and assume flat priors in the model parameter spaces. Since the MCMC code we used in this paper [@liu12; @yuan11] is adapted from COSMOMC, we refer the reader to @lewis02 for a detailed explanation of the code about sampling options, convergence criteria, and statistical quantities. According to the results of @yan13, @pyz14, and @zhou14, the MCMC method is well suited to systematically investigate the high-dimensional model parameter spaces in fits to blazar SEDs.
![Two-dimensional probability contours of parameters. []{data-label="fig8"}](f8.eps){width="3.5in"}
Pairs of values of $\zeta_s \cong 2, \zeta_e \cong 70$, and $\zeta_s \cong 3, \zeta_e \cong 30$, from the fitting results shown in Fig. \[fig8\] correspond to a change in index compared to an equipartition circumstance of $\Delta\Gamma_\gamma\cong -0.2b$ and $\Delta\Gamma_\gamma\cong -0.14b$, respectively. Even the large deviation from equipartition causes a spectral-index change $\lesssim 0.1$ unit for $b\cong 0.5$, and even less for $b\cong 1/3$. The typical fluid-frame magnetic field derived from the fits has $B^\prime \approx 10$ mG. Synchrotron self-absorption is included in the fit. Considerations about allowed jet power (see App. \[sec:AppB\]) restrict the departure from equipartition further. Thus deviations from equipartition do not, on the basis of the Mrk 501 case, affect the spectral-index relation significantly.
The inability of the numerical MCMC model to find a most favored value for $t_{var}$ may reflect limitations of the log-parabola EED used to model the Mrk 501 spectrum. Using a model joining a power-law at low electron energies with a log-parabola function at high electron energies, @pyz14 fit radio data down to $\approx$ GHz frequencies, and obtain preferred variability times of $t_{var} \approx 5\times 10^5$ s.
For given values of $L_{48}$ and $\nu_{14}$, production of the highest energy $\gamma$-ray photons is assisted by going to an electron-dominated regime, where $\zeta_e\gg 1$ and $u_e^\prime \gg u^\prime_{B^\prime}$. The larger Lorentz factor electrons required to produce the same value of $\nu_s$ in a weaker magnetic field can Compton scatter ambient photons to the highest energies.
Discussion and Summary {#sec:ds}
======================
The nonthermal synchrotron paradigm pervades thinking in blazar physics, yet is incapable of explaining some of the most elementary facts, e.g., why synchrotron-radiating nonthermal electrons are apparently accelerated so inefficiently. Rather than reaching values of $\approx 100\Gamma$ MeV [e.g., @gfr83; @jag96], the peak synchrotron frequencies of FSRQs with $\nu_s \cong 10^{13}$ Hz are $\approx 10^{10}$ times less than the highest energy synchrotron photon in the maximally efficient electron Fermi-acceleration scenario. Even the highest energy synchrotron photons from HSP BL Lac objects rarely exceed $\approx 10$ – 100 keV, orders of magnitude below the radiation-reaction limit. It is crucial to understand the reason for the low peak synchrotron frequencies (smaller, of course, than the maximum synchrotron frequency), and how they relate to source luminosities and SEDs, which are the basis of the blazar sequence and blazar divide.
Near-equipartition, log-parabola (NELP) model
---------------------------------------------
The astrophysics developed here may point a way to the solutions of these puzzles by first explaining the spectral-index diagrams. If the radiating electrons are near equipartition and approximately described by a log-parabola EED because of the underlying acceleration and radiation physics, then the relationships between the $\gamma$-ray spectral index $\Gamma_\gamma$ and $\nu_s$ and $\nu_{\rm C}$ are precisely defined in the Thomson regime by functions of the form $\Gamma_\gamma = d - k\log\nu_{s({\rm C})}$, namely eqs. (\[GammagammaEC\]) – (\[GgSSC\]). The slope is accurately reproduced even when Klein-Nishina effects are important. Moreover, the model inputs are all in principle observable from near-simultaneous multi-wavelength blazar campaigns: $L_{48}$, $\nu_s$, $\nu_{\rm C}$ and ${\cal A}_{\rm C}$ from spectral observations, $b$ and $\zeta_s$ from SED modeling, and $t_{var}$ from temporal analysis. As shown here for Mrk 501, $\zeta_e$ and $\zeta_s$ can also be deduced from SED modeling, leaving only the baryon-loading $\eta_{bl}$ as a major uncertainty, which affects the jet power (Appendix \[sec:AppB\]).
The near-equipartition approach using a 3-parameter log-parabola EED furthermore makes quantitative predictions about the dependence of observables on $\Gamma_\gamma$ for statistical quantities of blazars, or for different states of a single blazar. A specific example that can be performed with Fermi-LAT data is to determine $\gamma$-ray spectral indices of a large sample of blazars of specific types, e.g., LSP FSRQs and HSP BL Lac objects, in adjacent energy bands, giving the spectral curvature. The curvature of the $\gamma$-ray SED is uniquely related to the curvature of the synchrotron SED, depending on whether the $\gamma$ rays have an SSC or EC origin. The difficulty of performing this test, of course, is the requirement of quasi-simultaneous observations over a large energy range in order to provide a good characterization of the synchrotron SED peak and curvature.
For the synchrotron spectral-index diagram, our analysis shows that $k = k_{EC} = 3b/4$ for EC scattering in the Thomson regime, and $k = k_{SSC} = 9b/16$ for synchrotron self-Compton (SSC) radiation. Numerical results show that this dependence is even valid when Klein-Nishina effects are important. Analysis of the combined sample of FSRQs and BL Lac objects in the 3LAC [@3LAC], $k_{3LAC} = 0.18\pm0.03$, implying curvatures of $b \cong 0.24$ if the emission arises from EC processes, and $b \cong 0.32$ if the $\gamma$ rays are SSC. However, it may not be correct to combine the two samples with different typical values of $b$ in their populations. Specific predictions for the slope of the $\Gamma_\gamma$ vs. $\nu_s$ behavior, depending on whether the emission has an EC and SSC origin, should be studied for samples of blazars binned in ranges of $b$, because the two variables are correlated, with larger curvatures, $b \approx 1$ for FSRQs, compared to $b\lesssim 0.5$ for BL Lac objects [@chen14]. Insofar as the SSC component seems less dominant in FSRQs ($\zeta_s \approx 0.2$) than in BL Lac objects ($\zeta_s \approx 1$), the effect of this correlation on the spectral-index diagrams also has to be considered.
In principle, underlying correlations of $\Gamma_\gamma$ with $t_{var}$ can be examined with the increasing number of simultaneous multiwavelength blazar SEDs. Limitations of the log-parabola function to describe the EED remains a central assumption that can be relaxed, though not without associated theoretical or numerical efforts.
Departures from equipartition
-----------------------------
One of the uncertain parameters is the electron equipartition parameter $\zeta_e$, here defined as the ratio of nonthermal electron and positron energy to magnetic-field energy throughout the volume of the radiating region. Assuming $\zeta_e \cong 1, \zeta_s = 1$ gives the model predictions shown in the Figs. \[fig1\] – \[fig4\]. As shown in App. \[sec:AppB\], large departures from equipartition are not allowed if the absolute jet power is required to be less than the accretion power, which in turn is assumed to be bounded by the Eddington luminosity.[^6] From the results of App. \[sec:AppB\], one possibility is that the SSC bolometric luminosity in the SEDs of large Compton-dominance FSRQ flaring events should be small compared to the bolometric synchrotron luminosity (that is, $\zeta_s\ll 1$) for compatibility with sub-Eddington jet powers.
Spectral modeling of the FSRQs 3C 279 and 3C 454.3 is possible for $\zeta_e \cong 1$ [@cer13; @der14a]. For the BL Lac Mrk 501, a large departure from equipartition is required to get a good spectral fit, as we have shown, but even in this case, the effect from this out-of-equipartition condition on $\Gamma_\gamma$ is small. The deviations from equipartition giving the best fits to the SEDs of Mrk 501 show that large $\zeta_e$, electron-particle–dominated fits (with correspondingly weak magnetic fields) are favored to fit HSP BL Lacs extending into the TeV regime.
Extensions of the log-parabola model
-------------------------------------
The log-parabola function, eq. (\[g2Ng\]), is motivated by second-order Fermi acceleration theory where MHD turbulence in the emitting fluid systematically accelerates particles to form a curving EED (see Section 2). An equally compelling scenario combining first- and second-order processes considers a power-law distribution of particles injected downstream of a shock into a turbulent region where second-order processes broaden the distribution, so that the EED approximates the PLLP function, eq. (\[PLLP\]).
The full PLLP model has 5 parameters, but we have treated in Figs. 1 – 4 the important case of an EED with a $-2$ number index extending to low energies without cutoff. This EED makes a low-energy boundary to the spectral-index data near the Thomson value of $\Gamma_\gamma = 3/2$. Remarkably, this is as hard as the hardest Fermi-LAT blazar spectral indices measured. So if Compton scattering is responsible for the formation of the $\gamma$-ray SEDs of HSP blazars, as is undoubtedly true for the bulk of the radiation, then the PLLP model would give a simple explanation for the lack of blazars harder than $\Gamma_\gamma = 3/2$.
The situation is complicated, however, by the possibility that if the EEDs had low-energy cutoffs rather than power-law extensions to low energies, then HSP blazars in the LAT band would tend to be dim and hard to detect. So the apparent lack of blazars harder than $\Gamma_\gamma
= 3/2$ could be a selection effect rather than a limit imposed by the radiation physics. Searches in the Fermi-LAT $\gamma$-ray data for blazars harder than $\Gamma_\gamma
= 3/2$ would test whether a PLLP model is preferred over a LP model; the detection of such hard blazars would rule out the form of the PLLP considered here.
Except when $\nu_s\gg 10^{16}$ Hz, the existence of a low-energy power law in the EED makes only a small difference to the SSC predictions compared to the pure LP model. Figs. 1 – 4 show that the SSC predictions tend to be slightly softer than the data. A low-energy cutoff in the EED could harden the SEDs, making it possible to attribute an SSC origin to the $\gamma$-ray components to all HSP blazars. The discovery of blazars harder than $\Gamma_\gamma = 3/2$ in the Fermi-LAT energy range would support this interpretation.
Blazar types in the $\Gamma_\gamma$ vs. $\nu_s, \nu_{\rm C}$ plane
-------------------------------------------------------------------
We now ask why there are [ essentially]{} no FSRQ blazars with $\nu_{s}\gtrsim 10^{14}$ Hz. The answer is likely to involve the dynamics of increasingly higher synchrotron-peaked near-equipartition jets which, when finding themselves in an external radiation field, are subject to a radiation force opposite to the direction of motion that acts on the nonthermal electron population. In the ideal one-zone model considered here, there is no radiative drag from synchrotron and SSC processes, only from EC processes [@tra11]. The larger values of $\delta_{\rm D}$ and $\gamma_{pk}^\prime$ for increasing $\nu_{s}$ implies a correspondingly larger radiative drag when external radiation fields are present that would either slow the jet plasma down [@gt10] or prevent it from reaching such large $\Gamma$ factors in the first place.
Rather than treating the jet dynamics, which is beyond the scope of the present enquiry, we quote simple analytic expressions for the synchrotron and external Compton SEDs from log-parabola distribution, eqs. (\[eLsynesOmegas\]) and (\[eLECesOmegas\]), which effectively answers the question of how the Compton dominance ($\propto$ radiative drag) grows with increasing $\nu_s$. From eq. (\[eLsynesOmegas\]), to keep the apparent synchrotron luminosity constant requires $\delta_{\rm D}^4B^{\prime 2} \gamma_{pk}^{\prime 2} \sim $ constant. Eq. (\[eLECesOmegas\]) shows that the apparent external Compton component grows $\propto \delta_{\rm D}^6 u_0 \gamma_{pk}^{\prime 2}$. The ratio of the EC to synchrotron component is essentially the Compton dominance, which goes $\sim \delta_{\rm D}^2u_0/B^{\prime 2}\propto \nu_{14} L_{48}^{1/2}u_0$. So the Compton dominance and radiation drag grow $\propto\nu_{14}$, all other things being equal. This confirms the behavior shown in Figs. \[fig1\] – \[fig4\], which deviates at large values of $\nu_{14}$ due to Klein-Nishina effects. Depending on baryon loading, the jet could be quenched before escaping the BLR, making an unusual, short flaring event. In such an unstable situation, persistent emissions with large $\nu_{14}$ in dense external radiation environments might not be possible.
Maximum particle energy
-----------------------
The near-equipartition log parabola model can be used to derive expressions for maximum escaping proton or ion energy $E_{max}$. Starting with the @hil84 condition in the form $E_{\max} = Ze c B^\prime \delta_{\rm D}^2 t_{var}$ implies $$E_{\max}({\rm eV}) = 1.4\times 10^{20}Z L_{48}^{5/16}\;({t_4\over \nu_{14}})^{1/8} {f_1^{1/4} f_2^{1/8}\over \zeta_e^{1/4} \zeta_s^{1/16}f_0^{11/16}}\;,
\label{Emax}$$ using the dependences given in Table \[table1\] for the NELP model. For a BL Lac object with $L_{48}\lesssim 0.01$, the only way to accelerate ultra-high energy cosmic-ray protons to $\gtrsim 10^{20}$ eV occurs when $\zeta_e\ll 1$, that is, in a magnetically-dominated jet. It is interesting to compare this expression with the formula $$E_{max}({\rm eV}) = 2\times 10^{20} Z {\sqrt{\epsilon_B (L_{ph}/10^{48} {\rm ~erg~s}^{-1})/\epsilon_e}\over \Gamma/10 }\;
\label{Emax1}$$ [@wax04; @fg09; @dr10], which was also derived from the Hillas condition, where $L_{ph}$ is the isotropic bolometric photon luminosity, and $\epsilon_B$ and $\epsilon_e$ are the fractions of jet power going into magnetic field and electrons, respectively.
An interesting feature of a combined lepto-hadronic blazar model using log-parabola functions for the particle distributions is that the synchrotron radiation-reaction limit for protons is $\approx 200\Gamma$ GeV, a factor $m_p/m_e$ greater than the electron limit. The evolution of the combined lepton synchrotron/SSC and proton synchrotron SEDs with $\zeta_e$ would favor a proton synchrotron component in the same large magnetization ($\zeta_e\ll 1$) regime where the electrons are incapable of making high-energy radiation.
Blazar sequence and blazar divide
---------------------------------
Multiwavelength data from any given blazar display a rich array of spectral and variability properties. The spectral properties of blazars in this analysis are reduced to $\nu_s$, $L_{syn}$, $\nu_{\rm C}$, $\Gamma_\gamma$ and $b$, while the variability properties are reduced to $t_{var}$.
The spectral-index diagrams show robust correlations, which we explain as a consequence of relativistic blazar jets with different powers and in different environments, within which are entrained relativistic electrons that can be described by log-parabola EEDs. By relating the synchrotron peak frequency and synchrotron SED, which mirrors the EED, to the spectral index of the $\gamma$-ray SED formed through EC or SSC processes, the dependence of $\Gamma_\gamma = d - k\log \nu_s$ is easily derived in the Thomson regime. Moreover, the equipartition relations imply specific predictions for underlying correlations.
The spectral-index diagrams are one side of a triangle relating $\Gamma_\gamma$, $\nu_s$ (or $\nu_c$), and $L^{iso}$. The term $L^{iso}$ can either be the apparent isotropic synchrotron, $\gamma$-ray, or total bolometric luminosity. The other two sides of the triangle are $L^{iso}$ vs. $\nu_s$ or $\nu_{\rm C}$, the blazar-sequence relations, and $\Gamma_\gamma$ vs. $L_\gamma$, the blazar-divide relation.
Our work illuminates one side of the triangle, namely $\Gamma_\gamma$ vs.$\nu_s$ or $\nu_{\rm C}$. Regarding the blazar divide, suppose as a first approximation that the typical mass of a supermassive black hole is $10^9 M_\odot$, then Fermi-LAT data shows a significant change of spectral index at the Fermi divide of $L_\gamma \cong 10^{46}$ erg s$^{-1}$. If the apparent $\gamma$-ray luminosity is 10% of the apparent jet power, and the beaming correction is $\sim 100$, then the divide is at $L/L_{\rm Edd} \cong 0.01$ [@gmt09], and this would also represent the Eddington ratio below which the external radiation field energy density becomes small.
Extremely weak dependences, if any, are seen in the $\Gamma_\gamma$ vs. $L_\gamma$ blazar-divide plots on either side of the divide. Within blazar subpopulations [see Fig. 39 in @2LAC], the dependences of $\Gamma_\gamma$ on $L_\gamma$ are also weak. Other than near the divide itself, there is no clear dependence of $\Gamma_\gamma$ on blazar luminosity. Indeed, any such dependence is predicted to be weak, as can be seen from eqs. (\[GammagammaEC\]) and (\[GammagammaSSC\]), which show that $\Gamma_\gamma\propto -b\log L_{syn}/8$ for EC processes, and $\Gamma_\gamma \propto b\log L_{syn}/32$ for SSC processes.
To explain the blazar sequence relating $L^{iso}$ and $\nu_s$ or $\nu_{\rm C}$ requires jet physics outside the scope of the present investigation. Rather than saying why blazars of a certain type can exist, however, we can suggest why blazars dominated by EC emission require low-synchrotron peaks. The presence of any appreciable external radiation field would produce a Compton drag that decelerates the bulk flow or prevents such a near-equipartition situation that would produce a synchrotron SED peaking at such large $\nu_s$ from forming.
Summary
=======
To conclude, we have used an equipartition blazar modeling approach [@cer13; @der14a] to explain the correlations of Fermi-LAT $\gamma$-ray number spectral index $\Gamma_\gamma$ with peak synchrotron frequency $\nu_s$ and peak Compton frequency $\nu_{\rm C}$. This approach assumes a one-zone model fit to the broadband emission, so that emissions from, e.g., extended VHE jets [@bdf08; @yzz12], spine-sheath structures [@gtc05], decelerating jets [@gk03], or VHE emissions induced by UHECRs produced by the jet [@ess10; @tmd13], [ are assumed not to]{} affect the $\gamma$-ray spectral indices or peak frequencies. Within this framework, the trends in the spectral-index diagrams are reproduced in a model with equipartition conditions and a log-parabola electron distribution with $b \cong 1/2$. This conclusion holds even for out-of-equipartition conditions limited by absolute jet power to be sub-Eddington.
[ The broadly distributed data in the spectral-index diagrams suggest that a better model comparison would consider a distribution of parameter values to define a preferred model region in the $\Gamma_\gamma$ vs. $\nu_s$ and $\nu_{\rm C}$ diagrams. Such an approach depends on knowing whether the correlation of $b$ with $\nu_s$ [@chen14] is robust, if one is to sample from a distribution in $b$ values. Nevertheless, allowed regions in the spectral-index diagrams in Figs. 1 – 5 are already defined by the heavy solid curves, depending on whether internal SSC or external EC BLR or EC IR processes dominate the formation of the $\gamma$-ray SED. A distinct trend in the boundaries on the spectral index diagrams in Figs. 1 and 3 are found for the PLLP model that can be tested with Fermi-LAT analyses in different energy ranges. The $\Gamma_\gamma$ vs. $\nu_s$ boundaries defined by the dominance of internal SSC or external EC IR or EC BLR processes have, furthermore, a very different shape when $b = 1$ (Fig. 2) compared to $b = 0.5$ (Fig. 1). This can be tested by subdividing the Fermi-LAT $\gamma$-ray spectral indices in different ranges of $b$. Whether the log-parabola function or the PLLP model, eq. (\[PLLP\]) with a low-energy electron index $s=2$, better approximates the EEDs can be tested by searching for Fermi-LAT sources with $\Gamma_\gamma < 1.5$. ]{}
The weak dependences of $\Gamma_\gamma$ on changes in $L_{syn}$ found in eqs. (\[GammagammaEC\]) and (\[GammagammaSSC\]) are consistent with the weak dependences of $\gamma$-ray spectral index $\Gamma_\gamma$ on $L_\gamma$ on either side of the blazar divide. A physical explanation for the change of the radiation environment of blazars at $\approx 0.01 L_{\rm Edd}$, though a reasonable model assumption, would make sense of the blazar divide.
This leaves open the blazar sequence relations, which can ultimately only be understood from the physics occurring in the magnetospheres of the supermassive black holes powering the blazars. Near-equipartition blazar synchrotron sources with $\nu_s\gg 10^{15}$ Hz would suffer increasingly strong radiation pressure in an environment with dense external radiation fields, which could explain the absence of HSP FSRQs. Supermassive black-hole jets are most luminous when their emissions are coolest, that is, when their peak synchrotron and Compton frequencies are lowest. The near-equipartition log parabola blazar model provides a constrained system that explains the spectral-index diagrams, and points to studies that could allow for a deeper understanding of the blazar sequence and blazar divide.
The work of C.D.D. and J.D.F. is supported by the Chief of Naval Research. We thank Dr. Matteo Cerruti for discussions about spectral fitting, [ and the anonymous referee for constructive questions and the recommendation to consider the PLLP model]{}.
$\delta$-function Thomson-Regime SSC Derivation with Log-Parabola EED
=====================================================================
We derive the form of the $\nu L_\nu$ SED for the SSC component in the Thomson regime assuming a log-parabola function of the EED and employing $\delta$-function approximations for synchrotron and Thomson scattering. The comoving Thomson-scattered synchrotron self-Compton spectrum for isotropic distributions of photons and nonthermal relativistic electrons is given by $$\e_1^\prime L^\prime_{SSC}(\e_1^\prime;\Omega^\prime) = {1\over 2} m_ec^3 \e_1^{\prime 2}
\int_1^\infty d\gamma^\prime \int_0^\infty d\ep\int_{-1}^1 d\mu^\prime (1-\mu^\prime )\; N^\prime_e({{\gamma^\prime}})\; n_{ph}^\prime(\ep )\;{d\sigma(\bar\epsilon )\over d\e^\prime_1 }\;.
\label{esLSSC}$$ Here, $\mu^\prime $ is the cosine of the angle between the directions of the interacting electron and photon, $\bar \e = {\gamma}\ep(1-\mu^\prime )$ is the invariant collision energy, and $d\sigma(\bar\e )/d\e^\prime_1$ is the differential scattering cross section. We use the $\delta$-function Thomson scattering cross section $d\sigma(\bar\e )/d\e^\prime_1 = \sigma_{\rm T} \delta [\e^\prime_1 -\gamma^{\prime 2}\ep (1-\mu^\prime )]$ [eq. (6.44); @dm09]. From eq. (\[g2Ng\]), $N_e^\prime({{\gamma^\prime}}) = K^\prime y^{-2-b\log y}/ \gamma_{pk}^{\prime 2}$, and the photon spectral density $n^\prime_{ph}(\ep ) = \ep L^\prime (\ep ) / 4\pi f_0 R_b^{\prime 2}\e^{\prime 2}m_ec^3$, where $f_0$ is a geometry factor, $R_b^\prime = c \delta_{\rm D} t_{var}$, and $t_{var}(1+z)$ is the measured variability time [@der14a]. For the synchrotron target photon spectrum, $\ep L_{syn}^\prime (\ep ) = \e L_{syn}(\e )/\delta_{\rm D}^4 = \upsilon x^{1-b\log x}/\delta_{\rm D}^4$. Plugging these expressions into eq. (\[esLSSC\]), and using the $\delta$-function to solve the $\mu^\prime$ integral, we find $$\e_1^\prime L^\prime_{SSC}(\e_1^\prime;\Omega^\prime) = {\sigma_{\rm T} \upsilon K^\prime \over \delta_{\rm D}^4 4\pi f_0 R_b^{\prime 2} \gamma_{pk}^{\prime 5} \epsilon_{pk}^{\prime 3}} \int_{1/\gamma_{pk}^\prime }^\infty dy \; y^{-6-b\log y}\;
\int_{x_\ell}^\infty dx \;x^{-6-b\log x}\;.
\label{esLSSC1}$$ Here $x_\ell \equiv \sqrt{A}/y$, where $A = \e_1^\prime /2\gamma_{pk}^{\prime 2} \e_{pk}^\prime = \epsilon/\e_{pk,SSC}$. The interior integral can be solved by noting, to good approximation, the logarithmic term is slowly varying compared to the $x^{-6}$ term. The value of this integral is then $x_\ell^{-5-b\log x_\ell}/5$. After some manipulations, we obtain $$\e L_{SSC}(\e;\Omega ) \cong{2\sigma_{\rm T}\upsilon K^\prime\gamma_{pk}^\prime\over 5\pi f_0 R_b^{\prime 2}}\;
\sqrt{{\pi \ln 10\over 2 b}}\; A^{{1\over 2} - {b\over 8}\log A}\;.
\label{eLSSC}$$ Comparing with eqs. (\[Lsyne\]) and (\[alphanu\]) shows that the SSC spectral index is given by eq. (\[alphanu\]) with $b$ replaced by $b/2$ and $\e_{pk}$ by $\e_{pk,SSC}$, leading to eq. (\[GammagammaSSC\]).
Jet Power in the Near-Equipartition Log-Parabola Model {#sec:AppB}
======================================================
We consider jet power with the addition of baryons and photons. The baryon-loading factor $\eta_{bl} \equiv u^\prime_{p/i}/u^\prime_{e}$, and $\zeta_e \equiv u_e^\prime/u^\prime_{B^\prime}$, where $u^\prime_{p/i}$ is the fluid energy density of both thermal and nonthermal protons and ions, and $u^\prime_e$ is the nonthermal lepton energy density, including both electrons and positrons, For convenience, the thermal electron and positron energy density is assumed small. The absolute jet power for a two-sided jet is given by [@cf93; @cg08; @gt10] $$L_{jet} = 2\pi r_b^{\prime 2} \beta \Gamma^2 c u^\prime_{B^\prime} [1+\zeta_e(1+\eta_{bl})] + L_{ph}\;,
\label{Ljet}$$ where the absolute photon power $L_{ph}$ comprises synchrotron and SSC radiations, each assumed to be emitted isotropically in the jet frame, and EC radiations, with its comparatively narrower beaming [@der95]. The absolute photon powers depend on the observing angle $\theta$ through the Doppler factor $\delta_{\rm D}$ [@gt10; @der12], giving the absolute jet power in the NELP model: $$L_{jet} ({\rm erg~s}^{-1}) = (1+N_\Gamma^2)^2\,
\{ 4.0\times 10^{44} \sqrt{f_1 \sqrt{ {f_2 t_4\over \nu_{14}} \sqrt{{L_{48}^5\over f_0\zeta_s} } } }
\; [{1\over \sqrt{\zeta_e}} +
\sqrt{\zeta_e}(1+\eta_{bl}) ]+ {2 (L_{SSC}^{iso}+L_{syn}^{iso})\over 3\delta_{\rm D}^2} + {2 L_{EC}^{iso}\over 5\delta_{\rm D}^2} (1+N_\Gamma^2)^2
\}
\label{Ljet1}$$ where the observing angle $\theta \equiv N_\Gamma/\Gamma \ll 1$ and $\Gamma \gg 1$. In this expression, $L_{syn}^{iso}$, $L_{SSC}^{iso}$, and $L_{EC}^{iso}$ are the measured apparent isotropic bolometric synchrotron, SSC, and EC luminosities, respectively. The other two terms in eq. (\[Ljet1\]) correspond to the magnetic-field, $\propto 1/\sqrt{\zeta_e}$, and the particle power, $\propto \sqrt{\zeta_e(1+\eta_{bl})}$. The additional factor of $(1+N^2_\Gamma)^2$ narrows the focus of the $\gamma$-ray beam. Eq. (\[Ljet1\]) can be rewritten as $$L_{jet} ({\rm erg~s}^{-1}) = 4.0\times 10^{44}\,(1+N_\Gamma^2)^2\,{L_{48}^{5/8} f_1^{1/2}\over (f_0\zeta_s)^{1/8}}\;
({f_2 t_4\over \nu_{14}})^{1/4}\;\sqrt{(1+\eta_{bl})(1+\eta_{ph})}\,
\large( {1\over w}+ w \large)\;,
\label{Ljet2}$$ where $$w \equiv \sqrt{\zeta_e(1+\eta_{bl})\over 1+\eta_{ph}}\;,
\label{what}$$ and the radiation loading $$\eta_{ph} \equiv {u^\prime_{rad}\over u^\prime_{B^\prime}} =
{5.4 f_0 \zeta_s} \;[1+\zeta_s + 0.6{\cal A}_{\rm EC}(1+N_\Gamma^2)^2]\;.
\label{Aph}$$ The isotropic bolometric photon luminosity $L_{ph}^{iso} = L_{syn}^{iso} + L_{EC}^{iso}+L_{SSC}^{iso} = \eta_{ph}L_{syn}^{iso}$, where the external Compton dominance ${\cal A}_{\rm EC} = L_{EC}^{iso}/L_{syn}^{iso}$.
From eq. (\[Ljet2\]), the minimum power condition is defined by the condition $w = 1$. When the baryon loading factor $\eta_{bl}\ll 1$ and the radiation loading $\eta_{ph}\ll 1$, the minimum power condition is defined by $\zeta_e = 1$. If the baryon-loading is arbitrary, but $\eta_{ph}\ll 1$, the minimum power condition is defined by $\zeta_e = 1/(1+\eta_{bl})$. When $\eta_{bl}\gg 1$, the minimum power condition also corresponds to a highly magnetized jet (in terms of the electron energy density), with a larger jet luminosity by a factor $\sqrt{1+\eta_{bl}}$ at minimum jet power compared to a pure electron/positron jet. When $\eta_{bl}$ and $\eta_{ph}$ take arbitrary values, the minimum power condition is defined by $\zeta_e = (1+\eta_{ph})/(1+\eta_{bl})$, and the minimum jet power increases $\propto \sqrt{(1+\eta_{bl})(1+\eta_{ph})}$.
Eq. (\[Ljet2\]) gives the absolute minimum power to make the observed radiations from a blazar jet. For example, if the angular extent $\theta_j$ of the jet exceeds $1/\Gamma$, the power is increased by $\approx (\Gamma\theta_j)^2$. When observing at $\theta\approx 1/\Gamma$, the minimum jet power, $L_{jet} \approx 3\times 10^{45} L_{48}^{5/8}$ erg s$^{-1}$ can only be increased by one to to orders of magnitude before exceeding $L_{\rm Edd} \approx 1.3\times 10^{47}M_9$ erg s$^{-1}$ for a $10^9 M_\odot$ black hole, unless one demands unusually high photon efficiencies. For HSP BL Lac objects, with $L_{syn}\lesssim 10^{46}$ erg s$^{-1}$, there is no great difficulty in satisfying the Eddington limit, even far from equipartition. However, these very same objects are believed to be accreting at a rate $\lesssim 0.01 L_{\rm EDD}$, so even in this case, large departures from equipartition cannot be tolerated.
Based on Fermi-LAT data, @ghi14 argue that the absolute jet power $P_{jet}$ is larger than the accretion-disk luminosity, which is approximated as $10$ times the BLR luminosity. They also approximate $P_{jet} \approx 10P_{rad}$, with the absolute radiation power $P_{rad} \approx k_fL^{iso}_{\gamma}/\Gamma^2$, where the factor $k_f = 8/3$ for synchrotron/SSC processes and $k_f = 32/5$ for EC processes, and the bulk Lorentz factor $\Gamma$ of the radiating jet’s plasma outflow is stated to be in the range $10\lesssim \Gamma \lesssim 15$. The underlying assumption is that the observer is looking at $\theta_0 = 1/\Gamma$ to the jet axis. Inspection of the photon power in eq. (\[Ljet2\]) shows how uncertain this assumption is given how much brighter fluxes are along the jet axis compared to fluxes from sources at $\theta_0\approx 1/\Gamma$. Using the relation $1+N_\Gamma^2 =2\Gamma/\delta_{\rm D}$, the photon power in eq. (\[Ljet1\]) for a 2-sided jet is $$P_{rad} = (1+N_\Gamma^2)^4 \,{L_{syn}+L_{SSC}\over 6\Gamma^2} + (1+N_\Gamma^2)^6 \, {L_{EC}\over 10\Gamma^2}.
\label{Prad}$$ When viewing down the jet axis, these powers are $\approx 16$ (synchrotron/SSC) and $\approx 64$ (for EC) times less than the values used in the expression for $P_{rad}$ by @ghi14. The most powerful $\gamma$-ray sources have the largest core dominances and brightness temperatures [e.g., @pus09; @kov09; @lfw14 however, see @sav10], suggesting that these sources are also the ones viewed almost along the jet axis. [ Indeed, @jor05, Fig. 25, find no blazar with viewing angle $\theta_0 > 2/\Gamma$, whereas a large number of both BL Lacs and FSRQs have $\theta_0 < 1/2\Gamma$. A severe overestimation of the radiation and therefore jet power is made by not taking this effect into account.]{}
For very powerful FSRQs like 3C 454.3, which has $L_{48}\sim 1$, a curious feature arises. Great flares exceeding $L_{EC}^{iso} \gtrsim 10^{50}$ erg s$^{-1}$ with large Compton dominance $\gtrsim 100$ can be allowed while maintaining absolute jet power $L_{jet} \lesssim L_{\rm Edd}$ only if $\theta\ll 1/\Gamma$ and $\zeta_s\ll 1$, that is, $L^{iso}_{syn} \gg L^{iso}_{SSC}$, implying a small SSC component relative to the synchrotron component. The effect of decreasing $\zeta_s$ is to increase $\delta_{\rm D}$ and narrow the Doppler cone, making the beaming factor even smaller, so that [ extreme apparent EC $\gamma$-ray powers lead to absolute jet powers that are sub-Eddington]{}. Spectral modeling to give the relative SSC and EC powers depends on X-ray observations, for example, Swift and NuSTAR, Fermi-LAT observations at GeV energies, and ground-based VHE air Cherenkov arrays.
Abdo, A. A., Ackermann, M., Ajello, M., et al. 2009, , 700, 597
Abdo, A. A., Ackermann, M., Ajello, M., et al. 2010a, , 715, 429
Abdo, A. A., Ackermann, M., Agudo, I., et al. 2010b, , 716, 30 Abdo, A. A., Ackermann, M., Ajello, M., et al. 2010c, , 188, 405
Abdo, A. A., Ackermann, M., Ajello, M., et al. 2011, , 727, 129
Ackermann, M., Ajello, M., Baldini, L., et al. 2010, , 721, 1383
Ackermann, M., Ajello, M., Allafort, A., et al. 2011, , 743, 171
Ackermann, M., Ajello, M., Atwood, W., et al. 2015, submitted to , arXiv:1501.06054
Aharonian, F., Akhperjanian, A. G., Bazer-Bachi, A. R., et al. 2007, , 664, L71
Ajello, M., Romani, R. W., Gasparrini, D., et al. 2014, , 780, 73
Albert, J., Aliu, E., Anderhub, H., et al. 2007, , 669, 862
Aleksi[ć]{}, J., Antonelli, L. A., Antoranz, P., et al. 2011, , 730, L8
Becker, P. A., Le, T., & Dermer, C. D. 2006, , 647, 539
B[ł]{}a[ż]{}ejowski, M., Sikora, M., Moderski, R., & Madejski, G. M. 2000, , 545, 107
Bloom, S. D., & Marscher, A. P. 1996, , 461, 657
Blumenthal, G. R., & Gould, R. J. 1970, Reviews of Modern Physics, 42, 237
B[ö]{}ttcher, M., & Dermer, C. D. 2002, , 564, 86
B[ö]{}ttcher, M., Dermer, C. D., & Finke, J. D. 2008, , 679, L9
Böttcher, M., Harris, D. E., & Krawczynski, H., eds. 2012, Relativistic Jets from Active Galactic Nuclei (Berlin: Wiley)
Cavaliere, A., & D’Elia, V. 2002, , 571, 226
Celotti, A., & Fabian, A. C. 1993, , 264, 228 Celotti, A., & Ghisellini, G. 2008, , 385, 283
Cerruti, M., Dermer, C. D., Lott, B., Boisson, C., & Zech, A. 2013, , 771, LL4
Chen, L. 2014, , 788, 179
Dermer, C. D. 1995, , 446, L63
Dermer, C. D., Murase, K., & Takami, H. 2012, , 755, 147
Dermer, C. D., & Razzaque, S. 2010, , 724, 1366
Dermer, C. D., Cerruti, M., Lott, B., Boisson, C., & Zech, A. 2014a, , 782, 82
Dermer, C. D., Murase, K., & Inoue, Y. 2014b, Journal of High Energy Astrophysics, 3, 29
Dermer, C. D., & Schlickeiser, R. 2002, , 575, 667
Dermer, C. D., & Menon, G. 2009, High Energy Radiation from Black Holes (Princeton University Press)
Dermer, C. D., Schlickeiser, R., & Mastichiadis, A. 1992, , 256, L27
Essey, W., Kalashev, O. E., Kusenko, A., & Beacom, J. F. 2010, Physical Review Letters, 104, 141102
Farrar, G. R., & Gruzinov, A. 2009, , 693, 329
Finke, J. D. 2013, , 763, 134
Fossati, G., Maraschi, L., Celotti, A., Comastri, A., & Ghisellini, G. 1998, , 299, 433
Fossati, G., Buckley, J. H., Bond, I. H., et al. 2008, , 677, 906
Georganopoulos, M., Kirk, J. G., & Mastichiadis, A. 2001, , 561, 111 Georganopoulos, M., & Kazanas, D. 2003, , 594, L27
Ghisellini, G., Maraschi, L., & Tavecchio, F. 2009, , 396, L105
Ghisellini, G., Celotti, A., Fossati, G., Maraschi, L., & Comastri, A. 1998, , 301, 451
Ghisellini, G., & Tavecchio, F. 2009, , 397, 985
Ghisellini, G., & Tavecchio, F. 2010, , 409, L79
Ghisellini, G., Tavecchio, F., Maraschi, L., Celotti, A., & Sbarrato, T. 2014, , 515, 376
Ghisellini, G., Tavecchio, F., & Chiaberge, M. 2005, , 432, 401
Giommi, P., Padovani, P., & Polenta, G. 2013, , 431, 1914
Giommi, P., Padovani, P., Polenta, G., et al. 2012, , 420, 2899
Guilbert, P. W., Fabian, A. C., & Rees, M. J. 1983, , 205, 593
Hayashida, M., Madejski, G. M., Nalewajko, K., et al. 2012, , 754, 114
Hillas, A. M. 1984, , 22, 425
de Jager, O. C., Harding, A. K., Michelson, P. F., et al. 1996, , 457, 253
Jorstad, S. G., Marscher, A. P., Lister, M. L., et al. 2005, , 130, 1418
Kovalev, Y. Y., Aller, H. D., Aller, M. F., et al. 2009, , 696, L17
Lewis, A., & Bridle, S. 2002, PhRvD, 66, 103511
Li, S. H., Fan, J. H., & Wu, D. X. 2014, Journal of Astrophysics and Astronomy, 35, 467
Liu, J., Yuan, Q., Bi, X. J., Li, H., & Zhang, X. M. 2012, PhRvD, 85, d3507
Maraschi, L., Ghisellini, G., & Celotti, A. 1992, , 397, L5
Marscher, A. P., & Gear, W. K. 1985, , 298, 114
Massaro, E., Perri, M., Giommi, P., & Nesci, R. 2004, , 413, 489
Massaro, E., Perri, M., Giommi, P., Nesci, R., & Verrecchia, F. 2004a, , 422, 103
Massaro, E., Tramacere, A., Perri, M., Giommi, P., & Tosti, G. 2006, , 448, 861
Meyer, E. T., Fossati, G., Georganopoulos, M., & Lister, M. L. 2011, , 740, 98
Meyer, E. T., Fossati, G., Georganopoulos, M., & Lister, M. L. 2012, , 752, LL4
Paggi, A., Massaro, F., Vittorini, V., et al. 2009, , 504, 821
Peng, Y., Yan, D., & Zhang, L. 2014, , 442, 2357
Pushkarev, A. B., Kovalev, Y. Y., Lister, M. L., & Savolainen, T. 2009, , 507, L33
Sambruna, R. M., Maraschi, L., & Urry, C. M. 1996, , 463, 444
Savolainen, T., Homan, D. C., Hovatta, T., et al. 2010, , 512, AA24
Shaw, M. S., Romani, R. W., Cotter, G., et al. 2013, , 764, 135
Sikora, M., Begelman, M. C., & Rees, M. J. 1994, , 421, 153
Sikora, M., Stawarz, [Ł]{}., Moderski, R., Nalewajko, K., & Madejski, G. M. 2009, , 704, 38
Stawarz, [Ł]{}., & Petrosian, V. 2008, , 681, 1725
Takami, H., Murase, K., & Dermer, C. D. 2013, , 771, LL32
Tavecchio, F., & Ghisellini, G. 2008, , 386, 945
Tramacere, A., Massaro, F., & Cavaliere, A. 2007, , 466, 521
Tramacere, A., Giommi, P., Perri, M., Verrecchia, F., & Tosti, G. 2009, , 501, 879
Tramacere, A., Massaro, E., & Taylor, A. M. 2011, , 739, 66
Waxman, E. 2004, New Journal of Physics, 6, 140
Yan, D., Zeng, H., & Zhang, L. 2012, , 424, 2173
Yan, D., Zhang, L., Yuan, Q., Fan, Z., & Zeng, H. 2013, , 765, 122
Yuan, Q., Liu, S., Fan, Z., Bi, X., & Fryer, C. 2011, , 735, 120
Zhou, Y., Yan, D., Dai, B., & Zhang, L. 2014, PASJ, 66, 12
[^1]: [ A correlation of the log-parabola width parameter $b$ and $\nu_s$ is apparent in SED modeling studies [@chen14], but is based on only 5 or 6 high-synchrotron peaked blazars.]{}
[^2]: The shock-in-jet model of @mg85 provides an alternate approach that could apply to the $\ll 10^{12}$ Hz radio regime that often remains unfit in the standard model described here.
[^3]: Energy flux is derived in 5 energy bands in intervals defined by 0.1, 0.3, 1, 3, 10 and 100 GeV.
[^4]: The fitting published in @cer13 lacked log-parabola $b$-dependent factors derived in reply to the referee of @der14a. Updated values have $\zeta_e\sim 1$ and $\zeta_s\gtrsim 0.2$.
[^5]: See http://cosmologist.info/cosmomc/readme.html
[^6]: Counter-examples to this assumption are claimed [@ghi14]. See App. \[sec:AppB\].
| {
"pile_set_name": "ArXiv"
} |
---
author:
- |
Angela Dai$^{1,3,5}$ Daniel Ritchie$^{2}$ Martin Bokeloh$^{3}$ Scott Reed$^{4}$ Jürgen Sturm$^{3}$ Matthias Nie[ß]{}ner$^{5}$\
$^{1}$Stanford University $^{2}$Brown University $^{3}$Google $^{4}$DeepMind $^{5}$Technical University of Munich\
bibliography:
- 'main.bib'
title: |
[ScanComplete]{}: Large-Scale Scene Completion and\
Semantic Segmentation for 3D Scans
---
Acknowledgments {#acknowledgments .unnumbered}
===============
This work was supported by a Google Research Grant, a Stanford Graduate Fellowship, and a TUM-IAS Rudolf M[ö]{}[ß]{}bauer Fellowship. We would also like to thank Shuran Song for helping with the SSCNet comparison.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
I present a comprehensive and uniform analysis of 25 [*ASCA*]{} observations from 23 Narrow-line Seyfert 1 galaxies. The spectral analysis and correlations are presented in this paper, Part 2; the reduction and time series analysis is presented in the companion paper, Part 1.
A maximum likelihood analysis confirms that the hard X-ray photon index is significantly steeper at $>$90% confidence in this sample of NLS1s compared with a random sample of Seyfert galaxies with broad optical lines. Soft excess emission was detected in 17 of the 19 objects which had no significant absorption, a result that demonstrates that soft excesses appear considerably more frequently in NLS1s than in Seyfert 1 galaxies with broad optical lines. The strength of the soft excess, parameterized using a flux ratio obtained from the black body plus power law model, has a wide range of values in these objects (a factor of 50). The photon index was found to be correlated with the H$\beta$ FWHM, despite the small range of the latter parameter; however, neither parameter is correlated with the strength of the soft excess or [*ROSAT*]{} slope. Therefore, assuming that an excess of soft photons results in the steep photon index and narrow H$\beta$ FWHM, that excess may lie primarily in the unobservable EUV. The strength of the soft excess is correlated with the variability parameters, so that objects with strong soft excesses show higher amplitude variability; this potentially important result is not easily explained. While a range of two orders of magnitude in luminosity is represented, the temperature of the soft excess is approximately consistent throughout the sample, in contrast with expectations of simple accretion disk models. The presence of ionized absorption was sought using a two-edge model. It was found that this component appears to be typically less common in NLS1s and some evidence was found that the typical ionization state is lower compared with broad-line Seyfert galaxies. This fact, plus evidence for a correlation between the presence of the warm absorber and significant optical polarization, may imply that the inner warm absorber is missing or is too highly ionized to be detected, and only the outer, dusty, less ionized warm absorber is present in many cases. The iron line equivalent width appears to be similar among narrow and broad-line Seyfert galaxies. This could mean that reprocessing occurs with similar geometry in both classes of objects; however, the detection of ionized iron lines in a few objects, implying possibly altered fluorescence yields, and poor statistics, makes this conclusion tentative. Constraints on physical processes and models based on extreme values of orientation and accretion rate for NLS1s are examined in light of the observational results.
author:
- 'Karen M. Leighly'
title: 'A Comprehensive Spectral and Variability Study of Narrow-Line Seyfert 1 Galaxies Observed by ASCA: II. Spectral Analysis and Correlations'
---
\#1\#2[\#1$\;$[[\#2]{}]{}]{}
Introduction
============
Narrow-line Seyfert galaxies are identified by their optical line properties: H$\beta$ FWHM is $<2000 \rm\,km/s$, the \[\] $\lambda 5007$ to H$\beta$ ratio is $<3$ and there are high ionization lines and frequently strong emission present in the spectrum (Osterbrock & Pogge 1985; Goodrich 1989). Although their permitted lines may be only slightly broader than their forbidden lines, they can be clearly distinguished from Seyfert 2 galaxies. For example, the permitted lines are polarized differently than the forbidden lines, indicating two emission regions (Goodrich 1989). As shown by Boroson & Green (1992) these emission line properties are strongly correlated; a principal component analysis (PCA) shows that much of the variance in line properties can be traced to a single eigenvector, despite the fact that different emission lines are thought to originate in widely separated parts of the AGN. Therefore, these correlations must have an origin in a primary intrinsic physical parameter. The narrow-line Seyfert 1 galaxies are located at the extreme end of the Boroson and Green eigenvector; therefore, as a class they exemplify an extreme value of this physical parameter. It is important to identify this physical parameter and to understand how it drives these correlations.
This paper is the second of a two part series that presents the first uniform analysis of the X-ray variability and spectral properties from [*ASCA*]{} observations of a sizable sample of Narrow-line Seyfert 1 galaxies. Part 1 presents the data reduction and the time series analysis of the [*ASCA*]{} data as well as a preliminary search for spectral variability and a general introduction to this subclass of Seyfert 1 galaxy. This part presents the spectral analysis of the [*ASCA*]{} data and studies of correlations between the spectral and variability parameters and optical emission line information from the literature as well as from analysis of spectra in hand. The results are discussed in terms of current models for the X-ray emission from NLS1s and Seyfert galaxies; in particular, effects of orientation and accretion rate are discussed. Some of these results have been previously presented in Leighly 1998.
[*ROSAT*]{} observations demonstrated that the soft X-ray spectra from NLS1s is systematically steeper than those from Seyfert 1 galaxies with broad optical lines (Boller, Brandt & Fink 1996; Forster & Halpern 1996; Laor et al. 1997a). [*ASCA*]{} spectra from NLS1s have the advantage of a broader band pass and better energy resolution than the [*ROSAT*]{} PSPC spectra. A breakthrough in the understanding of NLS1s occurred with the observation of RE 1034+39 using [*ASCA*]{} (Pounds, Done & Osborne 1995). The [*ASCA*]{} spectrum revealed a very strong soft excess component dominating the spectrum below 1–2 keV, and a very steep hard X-ray power law with photon index $\sim 2.6$. This spectrum is very different than the typical Seyfert 1 spectrum, which tends to have a flatter power law with photon index typically around 1.7–1.9 and the soft excess is generally not observed in the [*ASCA*]{} band pass. This dichotomy resembles that observed between Galactic black hole candidates (GBHC) in the hard (or low) and soft (or high) states (e.g.Nowak 1995), a fact that prompted Pounds, Done & Osborne (1995) to postulate that NLS1s are the supermassive black hole analogs of Galactic black hole candidates in the soft state. Soft state GBHC are thought to be accreting at a larger fraction of the Eddington limit than hard state GBHC; this analogy supports the idea that overall behavior of NLS1s originates from an accretion rate that is a higher fraction of Eddington than in Seyfert 1 galaxies with broader optical lines. More recently, Brandt, Mathur & Elvis (1997) presented a collection of photon indices from [*ASCA*]{} observations of Seyfert galaxies obtained from the literature. They found that the steep hard X-ray spectrum from RE 1034+39 is not anomalous; NLS1s have generally steeper hard (2–10 keV) X-ray spectra than broad-line Seyfert 1 galaxies.
Narrow-line Seyfert 1 galaxies have been also observed to exhibit peculiar spectral features in their [*ASCA*]{} spectra that have not been observed in Seyfert 1 galaxies with broad optical lines. Three NLS1s which have very strong soft excesses were discovered to have peculiar absorption features near 1 keV (Leighly et al. 1997a; see also Otani, Kii & Miya 1996; Hayashida 1996; Comastri, Molendi & Ulrich 1997). Absorption by highly ionized oxygen is common in Seyfert 1 galaxies with broad optical lines (e.g. Reynolds 1997). However, the energies of the features in these three objects is far too high to be interpreted as absorption by oxygen in the rest frame of the galaxy. Citing the similarities between many aspects of NLS1s and broad-absorption line quasars, which are known for their UV absorption lines that indicate high-velocity outflows, Leighly et al. (1997a) suggested that these features are due to relativistically outflowing gas with speeds of 0.2–0.6 c. More recently, Nicastro, Fiore & Matt 1999 presented an alternative interpretation: they suggest that these features are due to absorption by large columns of highly ionized gas, and the features are due to resonant absorption primarily by Fe L.
Other unusual features have also been found. Fiore et al. 1998 discovered a feature near 1 keV in the spectrum of PG 1244+026 that is best modeled by an emission line. Comastri et al. 1998 discovered that the iron line in the [*BeppoSAX*]{} spectrum of Ton S180 has an energy near 6.7 keV indicating that the iron is ionized. Both of these features were interpreted as evidence for reprocessing in an accretion disk which is highly ionized, a property predicted if the accretion rate is high (e.g.Matt, Fabian & Ross 1993a). However, it has been shown more recently that 6.7 keV iron lines are not restricted to NLS1s but occur also in Seyfert 1 galaxies with broad optical lines (e.g.Guainazzi et al. 1998a).
Data Reduction
==============
Spectra
-------
The data reduction and filtering are described in Part 1, and this section describes treatment of the data specific to the spectral fitting. Dark-frame error and echo corrections were made on data taken in Faint mode. As the Rev. 2 reprocessing had been performed on all of the data, no corrections for the gain in GIS3 during later time periods was made.
To obtain the best signal to noise, a range of three different-sized extraction regions were used. For brightest objects, the nominal extraction regions ($4^\prime$ and $6^\prime$ for SIS and GIS respectively) were used. For fainter objects, regions 87.5% and 75% of the nominal ones were used. Background spectra were accumulated from source-free parts of the detector.
In order to account for the decline in the energy resolution as a function of time for the CCD detectors, response matrices appropriate for the date of the observation were made using the script [*sisrmg*]{}. Ancillary response files (arf files) were made for each detector using [*ascaarf*]{}. Several early observations were made in 4CCD or 2CCD Bright mode and significant flux is spread out over multiple chips. For the most part, these were reduced and combined as recommended in [*The ASCA Data Reduction Guide*]{}. Spectra and response files were made for each separate chip. Response files were averaged together using [*addrmf*]{} weighted by the number of source photons in each chip. A spectrum was extracted from the entire source region and the arf file was made using that.
Spectral fitting was performed using XSPEC v. 10. The energy bands over which the spectra were fitted varied. For observations before January 1996, the lower limit was constrained only by the presence of a lower level discriminator applied during the observation. In fact, a level discriminator was applied only during the I Zw 1, RX J0439$-$45, 1H 0707$-$495 and Ark 564 observations. The GIS spectra were fit above 0.8 keV, since the GIS detectors are not well calibrated below that energy. Although Rev. 2 reprocessing should account for the gain changes in GIS3, better behavior of the iron line was obtained from some spectra when the region around the iron line was excluded from spectral fitting. After January 1996, differences between the SIS0 and SIS1 detectors at low energies are marked. This is thought to be due to Residual Dark Distribution (RDD) effects (Dotani 1998), and this problem most seriously affects the spectra below about 0.8-1.0 keV. It is thought that this effect more seriously affects the SIS1 detector because of its average higher temperature. Therefore, during observations from this time period, the SIS1 spectra below 0.8 keV were ignored, and GIS spectra below 1.0 keV were ignored.
The spectra from each observation were fit separately, except for IRAS 13349+2438. There were two observations of this object, separated by about 4 days. It has previously been shown that no detectable spectral variability had occurred between these two observations (Brinkmann et al. 1996; Brandt et al. 1997); this was verified and the data from these two observations were combined. For statistical tests, the properties of NGC 4051 were taken from the second, longer observation because of the better statistics offered.
ROSAT data
----------
Nearly all of these objects considered here had pointed [*ROSAT*]{} observations. These were reduced using [*Xselect*]{} and fit between 0.1–2 keV. The intent of using [*ROSAT*]{} spectra in this paper is to investigate the relationship between the slope of the soft X-ray spectrum and the soft excess observed in the [*ASCA*]{} spectra. The poor energy resolution of [*ROSAT*]{} prevents robust deconvolution of complex models. Therefore, we quote only the results from a power law fit with Galactic absorption, or with extra absorption if that improves the fit at greater than 90% confidence.
Three objects did not have pointed observations: RX J0439$-$45, PKS 0558$-$504 and 1H 0707$-$495. The information for these was taken from the [*ROSAT*]{} All Sky Survey data: for the first two, information from the spectra was available in the literature and for the last one, an estimation was made for the slope using the hardness ratios available in the [*ROSAT*]{} Bright Source Catalog (Grupe 1997 P. comm.). Finally, IRAS 20181$-$2244 was detected in the [*ROSAT*]{} All Sky Survey (Boller et al. 1992; Boller et al. 1998) but the flux was too low for it to be included in the [*ROSAT*]{} Bright Source Catalog. Therefore, no photon index estimate is possible.
Emission Line Data
------------------
Table 1 lists some of the optical emission line properties from NLS1s, obtained both from the literature and from spectra in-hand. The properties include the FWHM of H$\beta$, the H$\beta$ equivalent width, the ratio of to H$\beta$, the FWHM of \[\] when available, and the ratio of \[O III\] to H$\beta$. Measurement of these parameters is complicated by the fact that NLS1s often exhibit strong emission from blends of optical . This contaminates the continuum around the H$\beta$/\[\] complex and furthermore the multiplet 42 occurs so close to \[\]$\lambda 5007$ that it can significantly distort the profile. One of the more effective ways to deal with this problem is to subtract the using a scaled and broadened template formed from the spectrum of the quintessential NLS1 I Zw 1. This method was used in the seminal paper by Boroson & Green (1992) and has subsequently been used by several other authors including Grupe 1996 (also Grupe et al. 1999a). While the I Zw 1 template method does not provide a perfect iron subtraction (e.g. Boroson & Green 1992), it is useful also because it provides a consistent way to estimate the contribution of in the spectra. Seeking consistency in Table 1 parameters, the numbers from papers in which the template method was applied are preferentially quoted.
Spectra were available for analysis from the following objects: PHL 1092, 1H 0707$-$495, NGC 4051, Mrk 507 (kindly provided by J. Halpern), RE 1034+39 (kindly provided by D. Grupe), IRAS 13224$-$3809, IRAS 20181$-$2244 (kindly provided by L. Kay), IRAS 17020+4544 (Leighly et al. 1997b), Ark 564 (obtained from the [*HST*]{} archive), Mrk 142 and Kaz 163. The I Zw 1 template method was used to subtract the and the spectra were fitted using the [*LINER*]{} spectral fitting package (Pogge & Owen 1993). Deblending of the H$\beta$ line can contribute uncertainty to the derived parameters. Some fraction of H$\beta$ originates in low ionization material from the narrow-line region (NLR); in NLS1s it is difficult to estimate how much arises from the NLR because the width of the H$\beta$ is similar to that of the \[\] lines and therefore the narrow component cannot be easily recognized (e.g.Goncalves, Veron & Veron-Cetty 1998). As it is beyond the scope of this paper and the signal to noise of many of the spectra to attempt a detailed deblending, I assume that the NLR contributes H$\beta$ with $1\over 10$ the flux of \[\]$\lambda$ 5007, and I subtract that from the H$\beta$ profile before modeling. The assumption is based on the results obtained from narrow-line emission from Seyfert 2 and intermediate Seyfert galaxies (e.g.Koski 1978; Cohen 1983).
The \[\]$\lambda$ 5007 line was modeled using one or two Gaussians. Two were required in the following objects; the second broader component was blueshifted with respect to the narrow component: RE 1034+39 (see also Mason, Puchnarewicz & Jones 1996 and Goncalves, Veron & Veron-Cetty 1998), NGC 4051, IRAS 17020+4544, Mrk 507 and IRAS 20181$-$2244 (see also Halpern & Moran 1998). H$\beta$ and \[\] line fluxes were obtained by integrating the flux above the continuum, a procedure which has the advantage of being independent of the line model. The H$\beta$ was finally fit with a Lorentzian profile, which provided a substantially better description in most cases than a Gaussian in the NLS1 spectra (also Goncalves, Veron & Veron-Cetty 1998). In a few cases, subtraction of strong distorted the \[\]$\lambda$ 5007 profile, making the FWHM unreliable (1H 0707$-$495). These results were merged with those from the literature to obtain the values in Table 1.
Spectral Properties
===================
The spectral analysis is limited to the time averaged spectral properties of the sample. A preliminary investigation of spectral variability was discussed in Part 1. Detailed analysis of spectral variability has been presented already for Mrk 766 (Leighly et al. 1996) and NGC 4051 (Guainazzi et al. 1996), and will be discussed for Ark 564 in Leighly et al. in prep.
The spectra were first fitted with an absorbed power law with the absorption fixed at the Galactic value, and the ratio of the data to this model are shown for all the spectra in Figure 1. Significant deviations from the power law in nearly all of the spectra can be seen and the $\chi^2$ for this model are given in Table 2. An acceptable $\chi^2$ ($\chi^2_\nu<1.3$) was obtained from only 10 of the 24 spectral fits. Because of the wide range in fluxes represented by this sample, an acceptable fit may be obtained either when the model is adequate or when the statistics are poor. The shape of the residuals is generally concave upward and thus appears quite different than many of those from Seyfert galaxies with broad optical lines which tend to be modified by absorption (e.g. Reynolds 1997).
To investigate the origin of the residuals, initially two phenomenological models were tried. These consisted of a power law with a photon index $\Gamma$ and either a soft excess modeled as a black body with characteristic temperature $kT$ (the “soft excess” model) or an ionized absorber modeled as two absorption edges at $0.74$ and $0.87\,\rm keV$ in the galaxy rest frame (the “warm absorber” model). These models have the same number of degrees of freedom. The $\chi^2$ for these preliminary models are given in Table 2. The fit was improved substantially for 20 of the 24 spectra; only spectral fits to I Zw 1 and Kaz 163 were not improved at all, and the fit to IRAS 20181$-$2244 was improved only marginally. (Note that the following criteria were used to evaluate the spectral fits: the improvement was [*significant*]{} ($P<0.01$) when $\Delta\chi^2>6.34$, 9.21 and 11.3 for 1,2 and 3 additional parameters respectively; the improvement was marginal ($P<0.05$) when $\Delta\chi^2>3.84$, 5.99, and 7.82 for 1,2 and 3 additional parameters respectively.) Nineteen of the spectra (18 objects) are modeled substantially better by the soft excess model ($\Delta
\chi^2=10$, corresponding to a likelihood ratio of 140, e.g. Mushotzky 1982), and the fit was acceptable ($\chi^2<1.3$) for 18 of these (all but 1H 0707$-$495). Three objects (IRAS 13349+2438, IRAS 17020+4544, Mrk 507) were fit substantially better using the warm absorber model, and IRAS 20181$-$2244 was fit marginally better ($\Delta\chi^2=6$).
Some ambiguity remains after this preliminary modeling, as it is possible that the two-edge model may not be able to adequately describe the warm absorber. Therefore, a third preliminary model was applied, consisting of a power law, Galactic absorption and the [*absori*]{} model available in [*XSPEC*]{}; the resulting $\chi^2$ are listed in Table 2. The [*absori*]{} model describes the warm absorber and is characterized chiefly by two parameters: the column density $N_w$ and the ionization parameter $\xi$. The temperature of the ionized plasma is also an adjustable parameter; however, because the spectral fitting results are not very sensitive to this parameter, the temperature was fixed at either $3
\times 10^4\rm\, K$ or $1 \times 10^6\rm\, K$, and the lower value of $\chi^2$ is quoted. Note that the [*absori*]{} model does not include resonance line absorption. In 18 spectra from 17 objects, the $\chi^2$ for the [*absori*]{} model is more than 10 less than the $\chi^2$ for the two-edge model, indicating a likelihood ratio $>140$. In many of these objects, detailed analysis indicates the presence of a warm absorber, as discussed below. Alternatively, the ionization parameter becomes very low and neutral absorption is modeled (Mrk 507; IRAS 20181$-$2244). However, for the others, the ionization parameter becomes pegged at the maximum value of $\xi=5000$, where the model is not considered to be very accurate (P. Magdziarz 1997, P. comm.). When this is the case, the [*absori*]{} model appears to be trying to model the continuum. However, a soft excess models the continuum better. In nearly all cases, the difference in $\chi^2$ between the soft excess and [*absori*]{} models is larger than 10, indicating a likelihood ratio of $>140$. It is less than 10 only for Mrk 142; this is likely to be the fault of poor statistics, as the source is relatively faint and the exposure for the SIS0 is only 19 ks. Further evidence that the complexity of the spectrum is better described by a soft excess model is given by the lower panels in Figure 1. These show the ratio of the data to a power law (plus iron line as necessary) model fit above 2–3 keV, after the iron line component is removed from the model. The signature of absorption should be characterized by a deficit or localized features upon the continuum; in contrast, in many of the objects, a broad continuum excess is seen.
These preliminary results show that a soft excess is generally present in narrow-line Seyfert 1 galaxy [*ASCA*]{} spectra. This situation contrasts with that from Seyfert galaxies with broad optical emission lines; Reynolds (1997) finds that the warm absorber fitted most of the spectra adequately, and a soft excess was required in only 3 cases.
These simple models do not adequately model all of the spectral residuals; evidence is also present for iron $\rm K\alpha$ lines and additional absorption. I follow Reynolds (1997) and proceed to fit with a phenomenological model with the following components, where all parameter values are obtained in the rest frame:
- a power-law with photon index $\Gamma$ which models the high energy continuum;
- a soft excess modeled by a black body with temperature $kT$, which provides a generic model of the soft excess;
- two absorption edges with energies 0.74 and 0.87 keV in the rest frame of the galaxy, which represent absorption by O VII and O VIII in ionized gas;
- Galactic absorption as listed in Table 3;
- intrinsic neutral absorption in the rest frame of the galaxy; in some cases, information from nonsimultaneous [*ROSAT*]{} spectra is used to constrain this parameter.
- an iron $K\alpha$ line, represented by a Gaussian with central energy $E$ and width $\sigma$.
The results are listed in Tables 3a, b, and c. Note that the values of the parameters and their errors ($\chi^2=2.71$, that is, 90% for one parameter of interest) are listed only when there is a significant or marginal detection of that component. For the absorption edges and iron lines, upper limits are also listed.
Because the two-edge model may not adequately model the warm absorber in all cases, a complementary set of fits using the [*absori*]{} model was performed, and the results are listed in Table 4. These results are discussed in Section 3.3.
After applying this simple phenomenological model uniformly to the spectra, evidence for further complexity was discovered in 5 objects. There is additional evidence for an edge-like structure near 1 keV in 1H 0707$-$495, IRAS 13224$-$3809 and PG 1404+226. These features were discussed extensively in Leighly et al. 1997a; these features were modeled as edges here and will not be discussed. An additional soft X-ray emission line feature is seen around 1 keV in PG 1244+026 and Ark 564. These features will be discussed in Section 3.4.
The Photon Index in NLS1s
-------------------------
The photon index from narrow-line Seyfert 1 galaxies appears to be on average steeper than that from Seyfert galaxies with broad optical lines. This result was first reported by Brandt, Mathur & Elvis (1996), and it is confirmed by the spectra in this sample. Histograms of the photon indices for the NLS1s and for broad-line Seyfert 1 galaxies taken from Reynolds (1997) are plotted in Figure 2. The broad-line radio galaxies and 3C 273 are excluded; radio-loud objects often have flatter spectra than radio-quiet objects, possibly as the result of a flat component from the jet. Note that all of our narrow-line Seyfert 1 galaxies are radio quiet except PKS 0558$-$504 (Remillard et al. 1986). Also note that with or without the radio-loud objects, the distributions of the 2–10 keV luminosities of the broad and narrow-line AGN are consistent as determined by a two-sample KS test. Figure 2 shows clearly that the photon indices are consistently steeper for the NLS1s. To test whether the photon indicies could have been drawn from the same parent population, the statistical package [*ASURV*]{} Rev. 1.2 (Isobe & Feigelson 1990) which implements methods presented in Feigelson & Nelson (1985) was used. A variety of tests are available in this package which make different assumptions about the underlying distributions. The differences were found to be highly significant ($P<0.0001$) regardless of the assumed distribution.
The histogram shows an apparent overlap and spread in the values for both types of objects; however, it is difficult to determine whether this is significant from the histogram alone since the uncertainties in the photon indices are not taken into account. Therefore, the maximum likelihood method was used to determine jointly the best estimate of the mean and dispersion of both sets of indices (Figure 2; e.g. Maccacaro et al. 1988). These were $2.19 \pm 0.10$ and $0.30^{+0.07}_{-0.06}$ for the NLS1s and $1.78 \pm 0.11$ and $0.29^{+0.09}_{-0.07}$ for the broad line objects (errors are 68% for two parameters of interest). In both cases the dispersions are significantly greater than zero (99% confidence $>0.19$ and $0.17$ for the narrow and broad-line objects respectively; Figure 2). A significant spread in Seyfert galaxy X-ray 2–10 keV photon indices has been recognized previously by several investigators (e.g. Turner & Pounds 1989; Nandra & Pounds 1994). The contours representing the mean and dispersion exclude each other at greater than 90% significance. The validity of the maximum likelihood test depends on the assumption that the photon indices are normally distributed. This assumption is justified if the distributions are symmetric; however, the distribution of the broad-line objects appears to be significantly skewed ($S=-0.82\pm
0.50$). There is an interesting suggestion that high values are preferred for both types of objects, and the distributions tail toward the lower values.
The Soft Excess in NLS1s
------------------------
A soft excess component is required to model 18 of the 24 spectra (17 of 23 objects). The most striking difference is in the degree of prominence of the soft excess in the spectra. This can be seen in the lower panels of Figure 1, which show the ratio of the data to a power law plus iron line (as necessary) model above 2 keV, and extrapolated to lower energies. This figure shows that the soft excess component in NLS1s stretches well into the [*ASCA*]{} band pass and that it can dominate the spectrum up to at least 1 keV. This situation contrasts with that from broad-line Seyfert galaxies; in that case, the soft excess is seen in only in the lowest channels in the [*ASCA*]{} band, if at all (e.g. Guainazzi et al. 1994).
The measured temperatures of the soft excess component in the galaxy rest frame when modeled by a black body range from 0.1 to 0.25 keV. The maximum likelihood method gives an average of $0.15\pm 0.05 \rm\,
keV$ and dispersion of $0.12^{+0.09}_{-0.05}\rm\, keV$ (uncertainties are 68% confidence). The dispersion is significantly greater than zero at $>99$% confidence; however, it must be kept in mind that there may be some model dependence in the temperatures. The measured temperature is likely to be correlated with the energy of the lowest spectral channel, which is not the same for all of the spectra. Observations from later in the mission generally had a higher event threshold than observations from early in the mission, and some observations were performed with a level discriminator. Also, the objects represent a range of redshifts. Furthermore, application of warm absorber edges and additional absorption can alter the measured temperature. Finally, physical models for the soft excess (e.g. an accretion disk spectrum) are likely to be characterized by a range of temperatures, although the inner edge temperature may be well defined.
To study the variation in strength of the soft excess, I borrowed the hardness ratio plot technique used in the study of Galactic X-ray sources. Figure 3 shows the ratio from the summed SIS0 and SIS1 spectra of the medium (1.0–2.0 keV) to soft (0.6–0.9 keV) energy bands on the x axis, and the ratio of the medium to the hard (2.0–6.0 keV) on the y axis. There are two problems with applying this method directly. The first is that the Galactic column densities differ among the sample enough to seriously affect the flux in the soft band. Therefore, before accumulating the ratios, the spectra were corrected channel by channel for Galactic absorption using the analytic formulation from Morrison & McCammon (1983). The second problem is that the redshifts of the sources differ significantly. This can be corrected using the [*ASCA*]{} SIS0 effective area curve; however, the correction becomes increasingly less accurate as the redshift increases as large changes in the effective area become shifted from one band to the other. The accuracy of this method is checked by simulating the spectra using the best fit models shifted to zero redshift. The result is that the correction gives acceptable results for all but the two highest redshift objects (PHL 1092 and RX J0493$-$45, with redshifts 0.396 and 0.224 respectively); therefore they are not included in the top panel of Figure 3. Finally, the hardness ratios are made from the luminosities predicted by the best fit models with the Galactic $N_H$ removed. These are shown in the bottom panel of Figure 3. The advantage here is that no $N_H$ or redshift correction is necessary; the disadvantage of this is that the luminosities are model dependent. Thus the top and bottom panels are complementary.
In Figure 3, the x axis, being the ratio of the medium to soft X-rays, displays the strength of the soft excess, where objects with [*strong*]{} soft excesses are on the [*left*]{} side of the plot and correspond to [*lower*]{} values of the ratio. Henceforth, this ratio will be referred to as the “color ratio”. Six points marked by stars have very strong soft excesses: PHL 1092, RX J0439$-$45, 1H 0707$-$495, RE 1034+39, IRAS 13224$-$3809 and PG 1404+226. This group will be collectively referred to a “strong soft-excess objects”. In the middle of the plot marked by circles are all the other objects with detectable soft excesses; note that NGC 4051 (2 observations) and Mrk 766 are separately marked with open circles. These will be referred to as “weak soft-excess objects”. Note that although the strength or contrast of the soft excess is weak in these objects, the soft excess is required with high confidence in the spectral fit. On the left side of the plot marked by triangles are objects with no detected soft excess: I Zw 1, IRAS 13349+2438, IRAS 17020+4544, and Kaz 163. Note that absorption may be hiding a weak soft excess in some of these objects. Finally, the more heavily absorbed objects Mrk 507 and IRAS 20181$-$2244 are off the plot toward the right.
The y axis on this graph, displaying the ratio of the 1.0-2.0 keV medium band to the 2.0–6.0 keV hard band, illustrates the steepness of the hard X-ray spectrum or power law, where large values correspond to steeper spectra. This assertion was verified by comparing the ratios with the measured photon indices. This graph shows that the strong and weak soft excess objects have almost the same distribution of photon indices. This means that the large range of soft excess strengths cannot be attributed to differences in the slope of the power law, but rather must be due to differences in the relative normalizations of the soft excess component and the power law.
The color ratio, as defined above, provides a measure of the strength of the soft excess. However, since only the effect of the Galactic absorption has been removed, the color ratio is also sensitive to intrinsic absorption, either ionized or neutral. Two other complementary parameters to measure the strength of the soft excess can be defined. These have the advantage that they should not be affected by the presence of intrinsic absorption; the disadvantage is that they are model dependent. The first, called the “black body parameter” is defined as the model flux color ratio between the black body component and the power law at 0.45 keV plus the measured temperature of the black body in the rest frame \[$(F(pl)-F(bb))/(F(pl)+F(bb))$\]. This parameter describes the prominence of the black body over the power law and it should be approximately directly proportional to the color ratio in the case where there is no intrinsic absorption. The values of the black body parameter are listed in Table 3. Because this parameter potentially depends most sensitively on the power law index, the uncertainties were evaluated by fitting the spectra with the power law index fixed at its 90% error limits. Values near $-1$ and $1$ imply very strong and very weak soft excesses, respectively. The black body parameters range from $-0.72$ for IRAS 13224$-$3809 to $0.77$ for PKS 0558$-$508 (not counting objects with no detectable soft excess, for which the black body parameter equals 1). These values imply that the black body component is a factor of $\sim 50$ stronger in IRAS 13224$-$3809 than in PKS 0558$-$508.
The final parameter is $\alpha_{xx}$. Defined as the slope between 0.7 and 4 keV in the rest frame, it measures the overall steepness of the continuum spectrum. This slope is computed from the modeled continuum and therefore should also be independent of intrinsic absorption; however, it is model dependent. This parameter indirectly measures the strength of the soft excess: when there is no soft excess or it is weak, $\alpha_{xx}$ is close to the photon index; when the soft excess is strong, $\alpha_{xx}$ is much larger than the photon index.
Neutral and Ionized Absorption in NLS1s
---------------------------------------
Evidence for an ionized “warm” absorber was found in more than half of a sample of [*ASCA*]{} spectra from bright, predominantly broad-line Seyfert galaxies (Reynolds 1997; also George et al. 1998), a result which supports the idea that absorption by highly ionized material may be ubiquitous in Seyfert galaxies. In contrast, significant detections of a warm absorber as modeled by two absorption edges with fixed energies were found in only 6 spectra from 5 of the 23 objects; marginal detections were obtained from 5 others. It is conceivable that this is a consequence of the low flux level of our objects; however, there is also the impression that when a warm absorber is present, it has a lower optical depth, especially in the higher ionization, O VIII edge. The detections and upper limits are plotted in Figure 4, along with the results from broad-line AGN given in Reynolds (1997). Note that several of Reynold’s detections were changed to upper limits for consistency with the detection criteria adopted here. This figure shows that many of the O VIII upper limits from the NLS1s are below the detections from the broad-line objects.
The possibility of a difference in warm absorber parameters for the broad and narrow-line objects was investigated by running statistical tests on the measured edge depths and upper limits from the sample of NLS1s presented here and from the comparison sample of broad-line AGNs from Reynolds (1997). Note that because the warm absorber is not detected in all objects, the upper limits should be taken into account properly using survival analysis statistical methods (e.g. Feigelson & Nelson 1985); and here they were accounted for using the Kaplan-Meier estimator as implemented in the [*ASURV*]{} package. The resulting average optical depths for O VII and O VIII are $0.19\pm 0.04$ and $0.053 \pm 0.020$ respectively for the NLS1s, and $0.29\pm 0.07$ and $0.18 \pm 0.06$ for 21 broad-line objects. Note that in the algorithm when the minimum optical depth is an upper limit, it must be changed to a detection; this occurred in both data sets so the mean could be biased. Nevertheless, these results suggest that the O VII optical depths are consistent between broad and narrow-line objects, while the O VIII optical depths are smaller for the narrow-line objects compared with the broad. To test this further, [*ASURV*]{} two-sample tests which assume a variety of different intrinsic distributions were run. No difference between O VII optical depths was found (probability of a significant difference of 29–43%; a range is found for the different assumed intrinsic distributions), but a difference between the O VIII optical depths at 97–98% confidence was found. Reynolds (1997) found some evidence for a difference between high and low luminosity objects that could possibly be attributed to different behavior of radio-loud objects. If the radio-loud objects are excluded from the broad-line sample, there is still no difference in $\tau_{O\,VII}$, but the significance of the difference in $\tau_{O\,VIII}$ becomes $>99$%. Finally, warm absorber in the broad-line Seyfert 1 galaxy NGC 3783 has a very high optical depth; it is possible that this galaxy skews the results. If it is excluded, no difference in $\tau_{O\,VII}$ is obtained, but there is still a significant difference in $\tau_{O~VIII}$ of $>98$%.
The fact that the average O VIII optical depth is lower, that many of the upper limits on the NLS1 $\tau_{O~VIII}$ are smaller than the broad-line object detections, and that there is an apparently robust significant difference in the distribution of $\tau_{O~VIII}$ suggests that the ionization state in NLS1s is lower on average than in broad-line Seyfert galaxies. A better way to test this supposition would be to compare ratios of edge optical depths; however, it is difficult to determine a sensible way to derive the ratio of the edges when both values are upper limits. However, the joint distributions of $\tau_{O~VII}$ and $\tau_{O~VIII}$ can be compared using a two sample test for multivariate data containing upper limits (Makuch, Escobar & Merrill 1991). Testing against the hypothesis that the joint $\tau_{O~VII}$ and $\tau_{O~VIII}$ distributions are equal, I find mixed results. If the generalized Gehan distribution is appropriate, the warm absorber properties from the NLS1s are not significantly different from those of the broad-line Seyferts at 84–91% confidence. However, if the logrank distribution is appropriate, indications are that the difference may be significant (95–98%). Since it is not clear which distribution is appropriate, I conservatively conclude that there is a suggestion rather than a discovery that the distribution of warm absorber properties [*as a whole*]{} is different in NLS1s than broad-line AGN. However, the difference in distribution of the optical depths remains significant.
All of the spectra were also fit with warm absorber modeled using the [*absori*]{} model in [*XSPEC*]{}, and the results are listed in Table 4. A warm absorber was detected using this model in all of the objects in which one was detected using the two-edge model. A warm absorber was detected with low significance in several other objects as well (marginal detections in PG 1244+026, IRAS 13224$-$3809, Kaz 163; $\Delta\chi^2=10$ in Ark 564).
Additional neutral absorption in the rest frame of the galaxy was detected in 4 objects: I Zw 1, IRAS 17020+4544, Mrk 507 and IRAS 20181$-$2244. I Zw 1 is a moderately bright object, and thus since the detection of absorption is not very significant, the necessity of this component is somewhat doubtful. Furthermore there is marginal detection of a warm absorber and the two components are probably coupled in the spectral fitting. Significant absorption is indicated in the [*ROSAT*]{} spectrum, but at a much lower level ($N_H=2.3\times 10^{20}\rm\,cm^{-2}$). Further broad band observations are are necessary to determine whether this is significant or not. The absorption in Mrk 507 has been discussed by Iwasawa, Brandt & Fabian (1998); it is likely to be the reason that the [*ROSAT*]{} photon index is so flat in this object. Significant neutral absorption as well as ionized absorption was reported in IRAS 17020+4544, and has been linked to the high optical polarization in this object (Leighly et al. 1997b). The absorption in IRAS 20181$-$2244 has been discussed by Halpern & Moran (1998), and is probably the reason that this object was so faint in the [*ROSAT*]{} All Sky survey.
It is important to note that soft excesses were not detected in several of the objects with warm or neutral absorbers. This may be because weak soft excesses cannot be distinguished when there is also absorption. Thus, soft excesses may be even more prevalent in NLS1s than is indicated here. It is possible that the presence of a soft excess can also make detection of a warm absorber more difficult, since the absorption structure may be easier to see when the continuum is a power law. However, the moderate energy resolution provided by [*ASCA*]{} should be adequate to allow us to detect features from the warm absorber even if they fall upon a multicomponent continuum.
Leighly et al. 1997b discovered that highly polarized Seyfert 1 galaxies almost always had a warm absorber. This result supports the hypothesis of dusty warm absorbers (Brandt, Fabian & Pounds 1996; Reynolds et al. 1997); that is, dust which reddens and polarizes the optical emission is coincident with ionized gas. Several of the NLS1s discussed here have high optical polarization, including I Zw 1 (Smith et al. 1997), NGC 4051, Mrk 766 (Goodrich 1989), IRAS 13349+2438 (Wills et al. 1992), IRAS 17020+4544 (Leighly et al. 1997b), Mrk 507 (Goodrich 1989) and IRAS 20181$-$2244 (Kay et al. 1999). All of these objects show evidence for either neutral or ionized absorption. The following objects have weak or no polarization: Mrk 335, Mrk 142, PG 1211+143, PG 1244+026, PG 1404+226, Mrk 478 (Berriman et al. 1990), IRAS 13224$-$3809 (Kay et al. 1999), and Ark 564 (Goodrich 1989). There is either no or only marginal evidence for a warm absorber in these, except for Mrk 478. Polarization properties of the following objects have not yet been reported: Ton S180, PHL 1092, RX J0439$-$45, NAB 0205+024, PKS 0558$-$504, 1H 0707$-$495 and Kaz 163. Thus the NLS1s appear to show the same trend of absorption and polarization as we found for Seyfert galaxies in general (Leighly et al. 1997b).
Soft X-ray Features in NLS1s
----------------------------
Features resembling absorption edges near 1 keV were found in the [*ASCA*]{} spectra from the three NLS1s: 1H 0707$-$495, IRAS 13224$-$3809 and PG 1404+226. This work is discussed in detail in Leighly et al. 1997a. These features could be fit either by a one or two absorption edges or two or three unresolved absorption lines. Oxygen offers the primary opacity in photoionized gas; however, the energies of the absorption features are much too high to be attributed to oxygen absorption. If interpreted as absorption due to oxygen, the energies of these features imply a high blueshift of the absorbing material: 0.2–0.3 c for the edge model and near 0.57 c for the line model. It is difficult to accelerate ionized gas to such high velocities; however, the notable similarities between NLS1s and low-ionization broad absorption line quasars (BALQSOs) prompt us to take this hypothesis seriously. Both NLS1s and BALQSOs show strong or extreme and weak \[\] emission; many NLS1s and low-ionization BALQSOs have red optical spectra and strong infrared emission; finally, both classes are predominantly radio-quiet. See Leighly et al. 1997a for more details. An alternative interpretation in terms of a highly ionized warm absorber dominated by resonance absorption lines dominated by Fe L has been proposed by Nicastro, Fiore & Matt 1999. For this model, it is probably relevant that these absorption features near 1 keV are found exclusively in the objects with the strongest soft excesses in the sample.
The spectra of two other NLS1s, PG 1244+226 and Ark 564 revealed excess emission near 1 keV (Figure 1). The presence of this feature in PG 1244+226 has been discussed by Fiore et al. 1998. The residual can be most simply modeled as an unresolved Gaussian emission line. The features are clearly significant [*in addition*]{} to a weak soft excess component. This is demonstrated by energy versus flux $\chi^2$ contours shown in Figure 5, where the narrow width has been fixed at the best fit value to better constrain the fit; when the line width is left free, it becomes confused with the soft excess when at large fluxes and widths.
Fiore et al. 1998 discuss possible origins of this line. Emission in an optically thin thermal plasma is highly unlikely because the large luminosity requires a large amount of gas, and the rapid variability implies a small emission region. The combination of these two constraints implies the gas must be optically thick, in contrast with the assumption that the gas is optically thin. Rather, they support the hypothesis that this feature may arise from reflection in a photoionized accretion disk. This hypothesis is supported by the idea that NLS1s are accreting at a higher fraction of the Eddington rate, since under that condition, the accretion disk is expected to become ionized (e.g. Matt, Fabian & Ross 1993a). The energy near 1 keV supports interpretation as a blend of and at 0.92 and 1.02 keV respectively, as well as lines at around 0.8 keV. The lack of a K$\alpha$ line near 0.65 keV is puzzling however as it is expected to be strong in photoionized plasmas (e.g. Netzer 1996).
Fiore et al. 1998 also note that in the Matt et al. (1993a) models, the disk ionization parameter $\xi$ depends on the accretion rate as $\dot m^3$. The sensitivity of this dependence means that while the accretion rate may be large in many NLS1s and therefore while the disk may be ionized in many NLS1s, observation of these features would not necessarily be expected in all of them. Note that the $\xi \propto \dot m^3$ dependence is based on several assumptions; two primary ones are that the fraction of accretion energy going into X-rays which illuminate the disk remains the same as the accretion rate increases and that the geometry also remains the same.
The Iron Line in NLS1s
----------------------
An iron K$\alpha$ line with energy consistent with emission from neutral or low-ionization-state iron is ubiquitous in [*ASCA*]{} spectra from low-luminosity broad-line Seyfert 1 galaxies (e.g. Nandra et al. 1997a; Reynolds 1997). Frequently the line is found to be significantly broader than the instrument resolution (e.g. Mushotzky et al. 1995). The signal to noise is sufficient in at least one observation that the line profile can be examined; a clearly identified red wing provides the best evidence so far for emission from a relativistic accretion disk near a black hole (MCG–6$-$30$-$15; Tanaka et al. 1995). At higher luminosities, however, the iron line appears to be less frequent, and there is some indication that the characteristic ionization state of the iron is observably higher (Nandra et al. 1997b).
It is important to compare the iron line properties from narrow-line Seyfert 1 galaxies with those from broad-line Seyfert galaxies in order to place constraints on the geometry and importance and conditions of reprocessing. This is a difficult task using these data, however. The low flux combined with the steepness of the spectrum means that the signal to noise at the iron line is generally very poor. In fact, some of the sample were not significantly detected above 6.0 keV. For this reason, no useful constraints could be obtained from the following: 1H 0707$-$495, PG 1404+226 and RE 1034$-$3809.
To examine the iron line in the remaining 20 objects, the spectra above 2 keV were fit, rather than the whole band pass, so that potential broad band spectral curvature would minimally affect the results. Some evidence for an iron line was found in 13 of the 20 objects detected in the iron line energy range. An iron line was not statistically necessary in the models for PHL 1092, RX J0439$-$45, NAB 0205+024, Mrk 142, Mrk 507, Kaz 163 and IRAS 20181$-$2244. It is interesting to note that in PHL 1092 and IRAS 20181$-$2244 an iron line is found just below the detectability threshold (PHL 1092: $\Delta\chi^2=5.3$; IRAS 20181$-$2244: $\Delta\chi^2=5.6$) and in both cases the best fitting line energy is high, near 6.8–6.9 keV. An upper limit was obtained from the nondetections by fitting with an narrow Gaussian ($\sigma=0.05\rm\,
keV$). The rest frame energy was fixed at either the best fitting energy or at $6.4$ and $6.7$ keV and the result from the better fit was reported. These results show that the upper limits on the equivalent widths for these 7 objects are commensurate with those from the detections, and the line fluxes are on the low end of the distribution. This implies that lines were not detected in these objects simply because the statistics are poor.
In 6 spectra from 6 objects the uncertainty on the width of the line excludes zero at 90% confidence for 1 parameter of interest. $\chi^2$ contours were constructed from fits in which the line energy, width and flux were allowed to vary and a large region of parameter space was explored. These show that for two parameters of interest, the iron line is broad at the 90% confidence level in all 6 objects: I Zw 1, Ton S180, NGC 4051, Mrk 478, IRAS 17020+4544 and Ark 564. Not surprisingly, those are among the brightest objects at high energies. Interestingly, the line was unresolved in Mrk 335 and Mrk 766, both bright objects. Evidence for both broad and narrow iron lines was found in the first observation of NGC 4051.
In 6 of the spectra, the best fit line energy and 90% confidence interval for one parameter of interest exclude neutral iron emission at 6.4 keV. However, note that the line detection is marginal or nearly marginal in 4 of these. $\chi^2$ contours show that, for two parameters of interest, the line energy excludes 6.4 keV in only two of the objects (I Zw 1 and Ton S180) at greater than 90% confidence level when the line was allowed to be broad. The line energy in Ton S180 excludes 6.4 keV at 99% confidence (Figure 6; see below). The line energy excludes 6.4 keV when the line width was constrained to be narrow in three other objects: PKS 0558-504, PG 1244+026, and IRAS 13349+2438.
The average equivalent width, evaluated using the Kaplan-Meier estimator in [*ASURV*]{} which takes into account upper limits, was computed using the results from the 20 objects in which the iron line properties could be investigated. The resulting average was $352 \pm
77\rm\, eV$. This is consistent with the average of $362
\pm 56 \,\rm eV$ for radio-quiet broad-line objects from Reynolds (1997). The distributions of equivalent widths are not found to be significantly different using two sample tests implemented in [*ASURV*]{}. There are some difficulties with a direct comparison, however. A difference in equivalent width should be expected based solely on the difference in photon index: the equivalent width of the line is expected to be $\sim 20$% larger when the photon index is 1.9 compared with 2.3, assuming a face-on viewing angle, because there are relatively fewer photons above the iron K-edge in the steeper spectrum (George & Fabian 1991). The equivalent widths of detected lines may be artificially large in the NLS1s because the statistics are poorer in those spectra (a Malmquist bias). The energies of some of the lines are consistent with emission from ionized iron and under these conditions the fluorescence yield can be either enhanced or depressed depending on the ionization state (e.g. Matt, Fabian & Ross 1993b).
Since the equivalent width relative uncertainty is approximately the same as the line flux relative uncertainty (Yaqoob 1998), a maximum likelihood test on the 12 significant and marginal detections could be done. This revealed a significant dispersion on the equivalent widths at 90% confidence for two parameters of interest: the mean and dispersion are $270^{+140}_{-110}\,\rm keV$ and $190^{+130}_{-110}\rm\, keV$, respectively.
The inclination of the disk is also important. The five of the six objects in which a broad line was detected with 90% confidence were fit with a model representing a emission line emerging from a planar geometry near a Schwarzschild black hole (Fabian et al. 1989); note that the detection in Mrk 478 was too nearly marginal to provide any useful constraints). Poor statistics prevented many of the parameters from being constrained; therefore, the inner and outer radii were fixed at 6 and 1000 gravitational radii, respectively, and the emissivity parameter was fixed at its best fit value between $-2$ and $-3$. The results shown in Figure 1 illustrate that useful constraints could only be obtained for NGC 4051. Since the inclination is determined by the presence of a blue wing, the steep spectrum, conspiring with the low flux, make the inclination especially difficult to constrain in [*ASCA*]{} data from NLS1s.
Two objects present particularly interesting iron line behavior. I Zw 1 has a strong hard excess in its spectrum (Figure 7) which appears to be a very large equivalent width iron line. This result is interesting in light of the fact that I Zw 1 is the prototype narrow-line quasar with very strong optical emission (e.g. Phillips 1976). The iron emission in the UV is also very strong (Laor et al. 1997b). Thus it may not be surprising that the iron emission in X-rays is also strong, although if it is connected with the optical and UV results it may imply a nucleus-wide overabundance of iron, since the emission mechanisms for and iron K$\alpha$ are much different. Ton S180 has the clearest example of a broad, ionized iron line among the objects in our sample. Such a line was previously detected in the [*SAX*]{} data on this object (Comastri et al. 1998). The $\chi^2$ contours of the energy versus the width are shown in Figure 6.
In summary, it appears that the equivalent width of iron line is the same in NLS1s as in broad-line objects; this would imply that the covering fraction of optically thick material is the same in both types of objects. This result is severely limited by the statistics of the spectra, however.
Correlation Analysis
--------------------
This paper and Part 1 present the X-ray spectral and variability results as well as some optical emission line properties from a sample of narrow-line Seyfert 1 galaxies observed by [*ASCA*]{}. It is appropriate to search for correlations among these properties. Nonparametric correlation coefficients are more robust against outliers and more appropriate for these data which are not expected to be distributed normally. The Kendall’s Tau statistic was used (e.g. Press et al. 1992), primarily because there is an implementation with reliable confidence estimates available in [*ASURV*]{}. Tests for correlations were also done using the more familiar Spearman rank statistic, and the results were not substantially different.
Two sets of correlations were performed. One set included data from all 23 objects, and the other used data from a subsample consisting of only the 17 objects that possessed an observable soft excess. There were two observations of NGC 4051, and the data from the second observation was used for the correlations owing to the longer exposure and better statistics. The correlation matrix, listing the probabilities that the data are uncorrelated, is given in Table 5. Plots of some of the correlations are given in Figure 8. I refer to those correlations having probabilities of no correlation of $<0.06$ to be [*significant*]{} or suggestive, and those with probabilities of $<0.01$ to be [*strong*]{}. Significant and strong correlations are marked in Table 5 by open and solid circles respectively.
Correlations between the optical emission line properties, including H$\beta$ FWHM, H$\beta$ equivalent width, the ratio of to the H$\beta$ emission and the ratio of \[\] to H$\beta$ emission, were examined. Note that detection of a significant correlation between H$\beta$ FWHM and any parameter is somewhat surprising, considering that the sample of objects is selected on the basis of small H$\beta$ FWHM and therefore the dynamic range for this parameter is small. A strong correlation was found between the H$\beta$ FWHM and the H$\beta$ equivalent width. This correlation has been discussed by Goodrich (1989) and references therein, and indicates a connection between the kinematics of the broad-line region and the physical conditions within the emitting gas. This correlation was also found to be strong in a sample of optically selected NLS1s observed by [*ROSAT*]{} that was considered by Forster (1998). Correlations were also performed between the equivalent width, which can be approximated as the product of the /H$\beta$ ratio and the H$\beta$ equivalent width, assuming the continuum is approximately flat. An anticorrelation between equivalent width and H$\beta$ FWHM has been previously reported by many investigators including Zheng & O’Brien 1990, Puchnarewicz et al. 1992, Grupe et al. 1999a, and Boroson & Green 1992. Interestingly, I find a positive [*correlation*]{} between these properties. Forster (1998) also finds this result; specifically, he finds that for low H$\beta$ FWHM, there is a correlation with equivalent width which shifts to an anticorrelation for higher H$\beta$ FWHM.
Correlations were sought between the [*ASCA*]{} and [*ROSAT*]{} photon indices and the optical line properties. A significant correlation was found between the H$\beta$ FWHM and the [*ASCA*]{} hard X-ray photon index (Figure 8). There is also a strong correlation between H$\beta$ equivalent width and the [*ASCA*]{} photon index which is probably a consequence of the strong correlation between the H$\beta$ parameters discussed above. These correlations are important because they imply a direct connection between the conditions in the central engine and the kinematics of the broad-line region. This result is also important because it challenges some models proposed for the production of the strong characteristic of NLS1s, as has been noted previously. The large to H$\beta$ ratio is generally difficult to understand because both emission lines are predicted to be produced under the same conditions in photoionization models and those models generically underestimate (for a review, Joly 1993). It has been proposed that strong emission would be produced if the X-ray spectrum is flat because hard X-ray photons can penetrate deeper into emission line clouds producing heating and leading to strong emission. However, an anticorrelation between H$\beta$ FWHM and the X-ray photon index has always been found (Puchnarewicz et al. 1992, Boller, Brandt & Fink 1996, Grupe 1996 and Grupe et al. 1999a, and Brandt, Mathur & Elvis 1997, Wilkes, Elvis & McHardy 1987). Most of these studies used soft X-ray spectral indices obtained from [*ROSAT*]{} or [*Einstein*]{} IPC. Therefore it is interesting that the correlation between H$\beta$ FWHM and the soft X-ray photon index measured using [*ROSAT*]{} is not significant. Conceivably this could be an effect of absorption which would tend to flatten the soft X-ray spectrum. However, the subsample of objects that have detectable soft excesses are minimally affected by absorption, and a significant correlation is not found within that sample either. Figure 8 shows that the [*ROSAT*]{} photon index is clustered around 3 for most of the sample except for the objects with the strongest soft excesses, and those with significant absorption.
No correlation is observed between the [*ASCA*]{} and [*ROSAT*]{} photon indices, even for the relatively unabsorbed, soft excess subsample. Again, this is apparently because $\Gamma_{ROSAT}$ is clustered between 3 and 4 for a broad range of [*ASCA*]{} photon indices. The lack of correlation between soft and hard photon indices is interesting because it appears to not support the most direct interpretation of the model proposed by Pounds, Done & Osborne (1995) for NLS1s. They proposed that the steep hard X-ray photon index is produced when the Comptonizing electron cloud is cooled by strong soft excess emission, in analogy with Galactic black hole candidates in the high state (e.g.Nowak 1995). The steep [*ROSAT*]{} spectrum implies a strong soft excess (see below); therefore, one might expect that the abundance of soft photons available when the soft excess is strong would predict that the strongest soft excesses would be associated with the steepest hard X-ray spectra. However, the soft photons necessary to cool the electrons may be emitted in the unobservable EUV and therefore a correlation between these indices is not required for the model to be viable.
Correlations between the overall shape of the [*ASCA*]{} spectra and the strength of the soft excess with other parameters may be important to examine. Three complementary parameters as defined in Section 3.2 (the color ratio parameter, the black body parameter and $\alpha_{xx}$) were used to to describe the overall shape of the [*ASCA*]{} spectra. These parameters are very strongly correlated, as expected. A strong correlation was found between $\alpha_{xx}$ and the [*ASCA*]{} photon index. This correlation is expected since these parameters are not independent: $\alpha_{xx}=\Gamma_{ASCA}-1$ when there is no soft excess. There is no correlation between the color ratio and the [*ASCA*]{} photon index; this appears to be because the color ratio includes intrinsic absorption, which would not be expected to be correlated with intrinsic continuum parameters. The correlation is weak between the black body parameter and the [*ASCA*]{} photon index; again, this supports the idea that the hard X-ray photon index is independent of the strength of the soft excess in the [*ASCA*]{} band. The color ratio and $\alpha_{xx}$ are both correlated with the [*ROSAT*]{} photon index, a result that may originate in two effects. The color ratio correlation appears to be dominated by absorption; that is, objects with high color ratio and low [*ROSAT*]{} photon index are also the ones that were found to have warm or neutral absorption in their [*ASCA*]{} spectra. On the other hand, $\alpha_{xx}$ should be largely independent of absorption and thus this correlation is due to the fact that objects with steeper [*ROSAT*]{} spectra have overall steeper [*ASCA*]{} spectra. Figure 8 shows that most of the strong soft excess objects are located in the upper right corner of the $\alpha_{xx}$–$\Gamma_{ROSAT}$ plot.
Correlations between the variability parameters and other parameters are also important; however, they must be examined carefully. Figure 3 in Part 1 shows that the excess variance appears correlated with the luminosity. However, the measured correlation is weak, a result which may be caused by the large scatter. Recall from Part 1 that the skew parameter is defined as the number of sigma from the predictions of a very specific model; it is therefore more reliable for high values rather than low values which would be dominated by fluctuations.
A most interesting result is the evidence for correlations between the variability parameters and the overall shape of the [*ASCA*]{} spectra. Consistently, more variable objects have steeper spectra. The parameter $\alpha_{xx}$ was chosen to be most representative of the intrinsic steepness and is plotted versus excess variance and skew in Figure 8. This figure illustrates that a primary reason for this correlation is that the objects with strong soft excesses are more variable than the ones with weak soft excesses. To investigate this further, the distribution of properties shown in Figure 8 for the strong (6 objects) and weak (9 objects) soft excess objects were investigated using several two-sample tests implemented in [*ASURV*]{}. The emission line parameters, including H$\beta$ FWHM, H$\beta$ equivalent width, equivalent width and ratio of to H$\beta$, are distributed equally among the strong and weak soft excess objects. This is important because it suggests that the correlation between the overall shape of the [*ASCA*]{} spectra and the variability properties may not be associated with the Boroson & Green (1992) Eigenvector 1. There is no difference in distributions of the [*ASCA*]{} photon index or 2–10 keV luminosity either. There are only strong differences in the distributions of the variability parameters, as indicated by the plots.
Correlations between the iron line equivalent width and all of the other parameters listed in Table 5 were also considered. Only one significant correlation was found: an anticorrelation between color ratio and the equivalent width for the soft excess objects which was significant at 0.058.
Several other significant correlations were found. The color ratio is strongly correlated with the optical depth of O VII. This is expected, based on the sensitivity of the color ratio to absorption. Similarly, the [*ROSAT*]{} photon index is correlated with $\tau_{O~VII}$, a further indication that the measured [*ROSAT*]{} photon index is influenced by unmodeled absorption. The temperature of the black body model for the soft excess is not significantly correlated with any parameter. It is weakly anticorrelated with the black body parameter, a result that may reflect model dependence. There is an intriguing correlation between the to H$\beta$ ratio and the variability parameters for the soft excess objects. This appears to be due to the fact that some of the strong soft excess objects have high to H$\beta$ ratios.
In summary, the most interesting correlations are as follows: a significant correlation between [*ASCA*]{} photon index and H$\beta$ FWHM was found; however, no correlation between [*ROSAT*]{} slope and H$\beta$ FWHM was found. The strength of the soft excess was not correlated with the hard X-ray photon index, as would be naively expected from Compton cooling models, but is correlated with the variability parameters. This is especially interesting because the latter correlation does not appear to be associated with the Boroson & Green (1992) Eigenvector 1.
Discussion
==========
The discussion is organized as follows. First, the constraints imposed by the results on current models for X-ray emission and absorption are discussed in a general way. Then, the consistency of the results on general models based on the orientation and accretion rate are discussed.
The Photon Index
----------------
The photon indices in the sample of NLS1s presented here was found to be systematically steeper than those in the broad-line Seyfert 1 galaxies analyzed by Reynolds (1997). A maximum likelihood analysis showed that the difference is significant at 90% confidence, and there is significant dispersion among the photon indices of both types of objects. Brandt, Mathur & Elvis (1997) previously reported this result. The advantage of the analysis presented here is that it is uniform, whereas Brandt et al. took their results from the literature and therefore the continuum model or fitting range could be different from object to object. This is conceptually important, because the soft excesses in some objects are strong enough and extend to high enough energy that they could modify the spectrum above 2 keV and skew the photon index obtained from a 2–10 keV fit. In practice, however, the results are probably within the errors obtained from both approaches.
### Thermal Comptonization
Pounds, Done & Osborne (1995) proposed that, in analogy with the soft state of Galactic black hole candidates, the steep hard X-ray photon index in the NLS1 RE 1034+39 can be understood in terms of a thermal Comptonization model. The energetic electrons are cooled by the copious soft photons from the soft excess, and the resulting cooler temperature leads to a steeper spectrum. This scenario was also adopted by Brandt, Mathur & Elvis (1997) to explain the anticorrelation between H$\beta$ FWHM and [*ASCA*]{} photon index.
However, the simplest thermal Comptonization models (e.g. Rybicki & Lightman 1979), will not work for AGN, because they predict a far more sensitive photon index dependence on the electron temperature and optical depth than is observed (e.g. Haardt & Maraschi 1991). Feedback between the soft photon source (cold phase) and the energetic electrons (hot phase) is required to stabilize the models and produce photon indices fairly narrowly distributed around a single value (e.g. Figure 2). Feedback models have been developed by Haardt & Maraschi 1991, 1993, Ghisellini & Haardt 1994, and Pietrini & Krolik 1994. Generally, the feedback consists of reprocessing of radiation from the hot phase by the cool phase, which increases the luminosity of the cool phase, as well as pair creation in the hot phase which alters the optical depth. Pietrini & Krolik (1994) find that the plasma parameters can be described by two scaling laws which are weak functions of the ratio of hard and soft luminosity: $\theta \tau_T
\simeq 0.1 (l_h/l_s)^{1/4}$ and $\alpha \approx 1.6(l_s/l_h)^{1/4}$, where $\theta=kT/m_e c^2$ is the plasma thermal energy in units of electron rest energy, $\tau_T$ is the scattering optical depth, $\alpha$ is the power law energy index and the compactness is $l=(\sigma_{T}/m_e c^3)(L/R)$. A difficulty with these feedback models is that if the hot phase completely covers the cool phase, and if all of the dissipation occurs in the hot phase, the predicted photon index will be far steeper than observed in broad-line Seyfert 1 galaxies. This problem has been addressed by proposing that covering is not complete; instead, the hot phase is confined to isolated hot spots (Haardt, Maraschi & Ghisellini 1994; Stern et al. 1995). If the spots hug the surface of the accretion disk, then the hot phase will intercept few photons from the cool phase, and the resulting photon index will be flat. If instead the hot spot is tall and thin, many soft photons will be intercepted, and the resulting photon index will be steeper.
The scaling law can be used to estimate the difference in $l_s/l_h$ ratio in NLS1s and broad-line Seyfert 1 galaxies. Assuming that the intrinsic indices are $\sim 2.4$ and $\sim 2.0$ for NLS1s and broad-line Seyfert 1 galaxies, respectively, means that $l_s/l_h$ should be a factor of 3.8 larger in NLS1s. Pietrini & Krolik (1994) point out that $l_s$ can be interpreted as a combination of intrinsic soft X-ray luminosity and of geometry (efficiency of reprocessing); therefore, if the geometry and therefore reprocessing is the same in NLS1s and broad-line Seyfert 1s, then the intrinsic soft X-ray luminosity must be higher to explain the difference in index. However, the amount of reprocessing and therefore geometry is not well constrained in NLS1s. The iron line equivalent width potentially gives some information about the amount of reprocessing. In the sample of NLS1s presented here, the iron emission line has approximately the same equivalent width as the lines in broad-line Seyfert 1s, overall, arguing that the covering fraction of reprocessing material in NLS1s is about the same as in broad-line objects. However, the iron line measurements have very large uncertainties. Also, there is evidence for an ionized iron line in a few objects; ionization of iron can result in either enhanced or reduced equivalent widths compared with neutral iron, depending on its ionization state (e.g. Matt, Fabian & Ross 1993b). Because there is very little information above 10 keV from NLS1s, the amount of Compton reflection cannot be determined. NGC 4051 and Mrk 335, the only NLS1s considered by Nandra & Pounds (1994), both have evidence for a reflection component in their [*Ginga*]{} spectra.
Spectral variability provides a potential diagnostic for this model. Assuming the geometry and plasma parameters remain the same, because of the shallow dependence of $\alpha$ on $l_s/l_h$, little spectral variability is expected for moderate changes in flux. It is also worth noting that these thermal comptonization models cannot explain the spectral variability exhibited by Mrk 766 as a change in $l_s/l_h$ (Leighly et al. 1996). In this case, a change in geometry would be required also.
A simple interpretation of the Pietrini & Krolik (1994) scaling law indicates that there should be a correlation between the strength of the soft excess, if it represents $l_s$, and the hard X-ray photon index, if it represents $l_s/l_h$. The results presented here do not support this, as there is no significant correlation between the hard X-ray photon index and any of the soft excess indicators (except $\alpha_{xx}$ which is not independent of the photon index). Figure 8 shows that strong soft excess objects are distributed more or less equally above and below the maximum likelihood average photon index of 2.19, and there are several weak soft excess objects which have very steep hard X-ray spectra. On the other hand, the soft photons which comprise $l_s$ may be emitted predominately in the unobservable EUV, and may not be directly related to the [*ASCA*]{} band soft excess. It is interesting to recall that the EUV photons are those that should be responsible for photoionization of the broad-line region clouds, rather than the soft excess photons seen in the [*ASCA*]{} or [*ROSAT*]{} band. If we accept the model proposed by Wandel & Boller (1998), that the narrowness of the lines is a consequence of a very strong but unobservable EUV emission, the observed correlation between hard X-ray photon index and H$\beta$ FWHM would support a difference in $l_s$ as the origin of the steep hard photon index, rather than in geometry of the Comptonizing electrons.
A yet unresolved point is the fact that soft X-ray selected samples are composed of half NLS1s but the other half are broad-line Seyfert 1 galaxies (Grupe et al. 1999a). The presence of broad-line Seyfert 1 galaxies in soft X-ray selected samples means that they too have strong soft excess components. It is not yet known whether or not these the soft X-ray selected broad-line objects have systematically steeper or flatter hard X-ray spectra than hard X-ray selected Seyfert galaxies; however, there is evidence that at least some of them have very flat hard X-ray spectra (e.g. 1H 0419$-$577 has a hard X-ray photon index $\sim 1.5$; Guainazzi et al. 1998b). Again, this may mean that the bulk of the soft photons are emitted in the unobservable EUV and we don’t know how strongly peaked the spectrum is in that region. Some information may come from examination of UV line ratios from soft X-ray selected AGNs as they may be sensitive to the shape of the EUV ionizing continuum (Krolik & Kallman 1988; Zheng, Kriss & Davidsen 1995; Korista, Ferland, & Balwin 1997). This work must be done carefully, however, as difference in radial or density distribution of emitting clouds can also modify line ratios.
### Nonthermal Comptonization
It has recently been discovered that in the soft state, the gamma-ray spectrum of the Galactic black hole candidate Cyg X-1 extends as a power law to at least 1 MeV without a break (Phlips et al. 1996). Such high energy emission is unlikely to be produced in a thermal plasma, it has therefore been asserted that this is evidence for nonthermal Comptonization in the soft state of Cyg X-1 (Gierliński et al. 1998). Following the analogy that NLS1s may be Seyfert galaxies in the soft state implies that nonthermal Comptonization may be occurring in NLS1s.
Gierliński et al. (1998) apply a hybrid thermal/nonthermal model to Cyg X-1 (see also Coppi 1999 for a review of hybrid models). The nonthermal part is comprised of electrons with a steep power law distribution, and the optical depth is sufficiently low that the soft photons experience at most 1 scattering. Therefore, the photon index of the radiation will be the same as that of the electrons, and in principle a photon index of $2.4$ can be produced easily. The required electron distribution may be natural as it is the same one that cosmic rays have. This model is to be distinguished from the saturated nonthermal Comptonization, which produces photon indices $\leq 2$ (e.g. Swensson 1994). Thermal electrons can be present also, if the time scale for energy loss due to Coulomb interactions is shorter than the Comptonization time scale. They can also Compton-scatter soft photons and produce the usual thermal spectrum.
### Bulk Motion Comptonization
A final model that has been suggested for NLS1s is that of bulk motion Comptonization (Shrader & Titarchuk 1998). This model assumes a three part accretion flow consisting of a cold Keplerian optically thick accretion disk and a sub-Keplerian optically thin corona, both of which suffer a shock at 10–30 $R_S$. Interior to the shock, the flow is hot and the optical depth is $\sim 1$. The hot plasma interior to the shock will have a high inflow velocity. It will intercept soft photons from the optically thick material exterior to the shock and upscatter them by transferring kinetic energy rather than thermal energy. The predicted hard X-ray spectrum will have a steep slope with $\alpha \sim 1.5$ (Chakrabarti & Titarchuk 1995; also Ebisawa, Titarchuk & Chakrabarti 1996). Bulk motion Comptonization is also discussed by Colpi (1988). A diagnostic of this model is that the spectrum should not extend beyond 511 keV; thus it cannot be applied to Cyg X-1 in the soft state because the observed spectrum extends to 1 MeV.
In this model, it is asserted that the hard X-ray spectrum will be weak. This is at least partially because the intensity of the soft X-ray photons that the infalling electrons see will be decreased by special relativistic effects (e.g. Pietrini & Krolik 1994). Also, some of the upscattered photons may possibly be trapped by the flow and not be able to diffuse out (e.g. Colpi 1988). This scenario may not be appropriate for NLS1s; Grupe et al. (1998a) finds that soft X-ray selected AGN, which are comprised of 50% NLS1s, are stronger in soft X-rays rather than being weaker in hard X-rays. However, there is some evidence that the hard power law is weaker in optically selected samples (Laor et al. 1997a).
The Soft Excess
---------------
This paper is the first to systematically study the soft excess in a moderate number of [*ASCA*]{} observations from NLS1s. Evidence for a soft excess was found in 17 of the 19 objects without significant absorption, implying that it is nearly ubiquitous in NLS1 [*ASCA*]{} spectra. The temperature of the black body used to parameterize the soft excess was nearly the same for all objects. However, an important result is that there is a wide range of soft excess strengths. At 0.6 keV, a factor of 50 is spanned, as measured by the black body parameter (Section 3.2). Note that in principle this result could be the effect of a hard X-ray power law with a range of strengths, or an inverse correlation between strength of the soft excess and weakness of the hard power law. These possibilities cannot be distinguished on the basis of the [*ASCA*]{} data alone. Another important result is that the strength of the soft excess is correlated with measures of X-ray variability: objects with strong soft excesses tend to have higher excess variance and skew parameter than objects with weak soft excesses. There appears to be no correlation between the soft excess strength and optical emission line parameters, suggesting that this behavior is not connected with Boroson & Green (1992) Eigenvector 1.
### The Soft Excess and the Optical Emission Lines
A strong soft excess is predicted to affect the thermal balance in the two phase model of the broad-line region, resulting in a lower temperature of the hot intercloud medium (Fabian et al. 1986). They predict that this will result in a very small region of the thermal stability curve being amenable to two phases. While a stationary two-phase medium in pressure balance may not be a completely appropriate description for the broad-line region clouds, this result may be partially responsible for the lower equivalent widths of permitted emission lines (e.g. Goodrich 1989) and the correlation between photon index and equivalent width observed in NLS1s.
The strong soft excess may also affect the location of the broad line region. Wandel & Boller (1998) note that if the steep soft [*ROSAT*]{} spectra are extrapolated into the unobservable EUV, a stronger ionizing continuum is inferred. If the permitted lines form at a particular ionization parameter, a larger radius for the broad-line clouds would result. The Keplerian velocities at larger radii will be smaller, and therefore the lines will be narrower. Alternatively, the densities in the broad line region clouds could be larger. Such an idea may explain correlation of the density-indicating ratio \]/\] with /H$\beta$ recently presented by Wills et al. (1999).
### The Soft Excess Temperature
The soft excess is conventionally thought to be the high energy tail of the accretion disk spectrum. If so, a soft excess at higher temperatures is expected for larger accretion rates and smaller black hole masses (Ross, Fabian & Mineshige 1992). Drawing upon an analogy with Galactic Black Hole candidates in their high state, and using RE 1034+39 for their example, Pounds, Done & Osborne (1995) claim that the soft excess could be primary emission from an accretion disk at near Eddington luminosity. They note that the relation of inner disk temperature and black hole mass for two objects radiating at near the Eddington limit is $T_1/T_2=(M_2/M_1)^{1 \over 4}$, and that the observed temperature ratio is consistent with plausible differences in the black hole mass of RE 1034+39 and the BHC Cyg X-1.
As noted above and in Section 3.2, the temperature of the soft excess measured among NLS1 is nearly constant, despite large differences in luminosity. RE 1034+39 and PHL 1092 have, respectively, the lowest and highest luminosities of the six strong soft excess objects in this sample. The difference in luminosity is a factor of 30, suggesting that the black hole mass is a factor of 30 higher in PHL 1092. If so, according to thin disk models, the temperature at the inner edge of the accretion disk should be a factor of 2.3 lower in PHL 1092 compared with RE 1034+39; in fact, it is slightly higher. A soft excess with $kT$ a factor of 2.3 lower would not appear in the [*ASCA*]{} band pass. However, it must be noted that there could be systematic uncertainties with the measurement of the temperature of the soft excess because it is a continuum component appearing at the edge of the [*ASCA*]{} bandpass, and there could be some model dependence. Furthermore, simulations show that there can be a tendency for very weak soft excesses to have a higher measured temperature.
This temperature constraint becomes even stiffer when [*ROSAT*]{} results are considered. In this study, a significant correlation was found between the [*ROSAT*]{} photon index and the soft excess strength, implying that objects with strong soft excesses generally have very steep [*ROSAT*]{} spectra. Grupe (1996) found a significant correlation between the photon index in a soft X-ray selected sample and the redshift. High redshift objects are more luminous in count rate limited samples. All of may imply that there could be objects which are even more luminous than PHL 1092 which have soft excesses observable in the [*ASCA*]{} band; however, spectroscopic observations would be necessary to confirm the temperature.
### Normalization of the Soft Excess
The soft excesses were modeled using a blackbody. Assuming that the emission originates in an accretion disk, one may ask whether the size of the emission region implied by the strength of the black body is commensurate or at least smaller than the source size inferred using variability arguments.
Using $H=50\rm km/s/Mpc$, the area of the emission region was computed from the black body model fits for all of the objects which had detectable soft excess component. The areas ranged from $10^{+20.7}\rm \,cm^{-2}$ for NGC 4051 to $10^{24.8}\rm \,cm^2$ for RX J0439$-$45. The areas were also quite large for PHL 1092, PG 1211+143 and PG 1404+226, which were among the most luminous objects in the sample. Rapid variability was observed from RX J0439$-$45: a change in flux by a factor of $>2$ was detected in about 2700 seconds. If the emission occurs within $7\,R_S$, then the upper limit on the black hole mass would be $4\times 10^7 M\odot$. However, this exercise implicitly assumes that the emission is not coming from the entire central region but rather a localized emission region, and therefore, the light crossing time of the source may overestimate the size of the emission region. Nevertheless, for illustration, consider the emitting region for the [*ASCA*]{} soft excess to be between 7 and 8 $R_S$. Then the area of that emitting region for a $4\times 10^7 M\odot$ black hole would be $6.8\times
10^{27}\rm \, cm^2$. This means that the time scale of variability for RX J0439$-$45, combined with its soft excess normalization, imply that only 0.1% of the region need be emitting if the whole thing is optically thick.
However, in the classical the thin disk model, the temperature of the disk at $7\,R_S$ should be much cooler than the temperatures of the soft excesses observed here. This may be a consequence of the fact that the temperatures in the thin disk model are obtained by averaging in the vertical direction. When realistic radiative transfer is taken into account, it turns out that the underlying disk emission will be Compton upscattered by the hotter upper layers of the disk and the observed emission should be hotter (Czerny & Elvis 1987; Ross, Fabian & Mineshige 1992; Shimura & Takahara 1993). Compton scattering will smear the local disk black body over a larger band pass, and therefore a larger area will be required to emit the observed soft excesses. An estimation of the increased area needed can be obtained from Figure 6 in Ross, Fabian & Mineshige (1992), which shows the fraction of total disk emission in soft X-rays above 150 eV as a function of black hole mass and accretion rate. For a $4\times 10^7\, M_\odot$ black hole, the fraction of flux in soft X-rays is order 10% if the Eddington fraction is as high as 1/3, implying that an area at least ten times larger than the previous estimate is required; this would still be only 1% of the 7–8$R_S$ annulus. For larger black hole masses and smaller Eddington fractions, the fraction emitted in soft X-rays is smaller.
### Models of the Soft Excess
Accretion disk models predict more intense X-ray emission as well as higher inner edge temperature when the accretion rate is high (e.g. Shakura & Sunyaev 1973). However, while standard models of accretion disk emission find a good fit to the optical-UV continuum, they predict far too cool temperatures at the inner edge to explain the soft X-ray emission. Many of these models ignore the vertical structure in the accretion disk and the radiation transfer. More sophisticated treatments of these corrections indeed do find that the accretion disks can emit strongly in soft X-rays, especially when the accretion rate is high and/or the black hole mass is low (e.g.Czerny & Elvis 1987; Ross, Fabian & Mineshige 1992; Shimura & Takahara 1993). A shifted and strengthened accretion disk spectrum can explain the steeper [*ROSAT*]{} spectra observed in NLS1s, and therefore this was interpreted as evidence for a high accretion rate (e.g. Boller, Brandt & Fink 1996; Laor et al. 1997a; Grupe et al. 1999a). This can also generally explain the high temperatures and higher frequency of X-ray soft excesses observed in the [*ASCA*]{} spectra from NLS1s (e.g. Pounds, Done & Osborne 1996).
More realistic accretion disk models can also potentially explain the fact that the temperature of the X-ray soft excess remains constant over a large range of source luminosities and inferred black hole masses (Section 4.2.2). Examining a wide range of parameters, Shimura & Takahara (1993) solve for hydrodynamic equilibrium and radiative transfer self-consistently . They find that the emission spectrum gradually shifts toward the soft X-rays as the accretion rate increases, confirming the results of Ross, Fabian & Mineshige (1992). However, when the accretion rate is very large, nearly the Eddington rate, the derived temperature, near 100–200 eV, has only a weak dependence on the black hole mass. This constancy in spectral shape occurs because at such high accretion rates, the scattering optical depth becomes independent of black hole mass, Comptonization is the dominant process at the inner edge of the disk, and the Compton $y$ parameter in the upper layers of the disk becomes larger than 1 and nearly independent of black hole mass. This behavior is not expressed for lower accretion rates, where the soft X-ray cutoff is a stronger function of temperature.
The large range in soft excess strengths may suggest that the soft excess does not have a uniform origin in NLS1s. Another possible origin besides primary disk emission is suggested by the discovery of soft X-ray line emission near 1 keV in two of the weak soft excess objects, PG 1244+026 (Fiore et al. 1998) and Ark 564, as well as the ionized Fe K$\alpha$ line in Ton S180. The soft excess in these objects could be due to reflection from an ionized accretion disk. When the surface of the accretion disk becomes ionized, the reflectivity to soft X-rays is increased as the light elements are ionized and the absorption in the upper layers is decreased. This situation has been explicitly modeled by Matt, Fabian & Ross (1993a) and Życki et al. (1994) and discussed by Czerny & Życki (1994) and Fiore, Matt & Nicastro (1997). Fiore, Matt & Nicastro (1997) mention that the effect of the reflection by ionized material would be an increase in the steepness of the 0.2–2 keV spectrum by about $\Delta \Gamma \approx 0.3$.
Another possibility is that the weak soft excesses seen in some NLS1s is the signature of Compton scattering of blackbody emission (e.g. Nishimura, Mitsuda & Itoh 1986). Evidence for this feature was found in the high state spectrum from Cyg X-1 (Cui et al. 1998; Gierliński et al. 1999). The spectrum was not adequately modeled with a black body plus power law; an additional soft excess, a signature of the fact that a black body is the source of photons, was required for intermediate energies. If the weak soft excess observed in some objects is in fact Comptonized blackbody emission, then the primary black body must not be very far below the [*ASCA*]{} bandpass.
The Warm Absorber
-----------------
This paper presents the first systematic survey of the warm absorber properties of NLS1s. The first result is that there is evidence that the incidence of warm absorbers, as indicated by the presence of and edges, is lower in NLS1s. The second result is that when warm absorbers are detected, they appear to have a lower ionization state than those in Seyfert 1 galaxies with broad optical lines. Both of these conclusions, however, are subject to restrictions based on the methods used to search for the warm absorber.
### Biases
Neither the narrow-line Seyfert 1 sample considered here, nor the broad-line Seyfert 1 comparison sample presented in Reynolds 1997, can be considered complete in any way. Therefore, it is important to determine whether biases in the sample selection could influence the result.
The NLS1s discussed here are selected in several ways, all of which may be biased against discovering highly absorbed objects. Soft X-ray selected objects present the strongest bias. They are drawn from Grupe’s [*ROSAT*]{} soft X-ray selected AGN (e.g. Ton S180, RX J0439$-$45, RE 1034+39, IRAS 13349+2438); the [*ROSAT*]{} spectra from these objects indicates little evidence for heavy absorption (Grupe et al. 1998b). They are also drawn from the Moran, Halpern & Helfand (1996) [*IRAS*]{}/[*ROSAT*]{} cross correlation sample (e.g. IRAS 13224$-$3809, IRAS 17020+4544, IRAS 20181$-$2244). Two objects were selected by the [*HEAO*]{} A-1 experiment (PKS 0558$-$504 and 1H 0707$-$495). Other objects are drawn from PG quasars and Markarian galaxies, including Mrk 335, I Zw 1, Mrk 142, PG 1211+143, Mrk 766, PG 1404+226 and Mrk 478. These objects are chosen by their blue optical spectra, which means that objects with moderate reddening may be present, but objects with very heavy absorption should not be.
In contrast, many of the broad-line Seyfert 1 galaxies considered by Reynolds 1997 are members of the HEAO-A2 Piccinotti sample (Piccinotti et al. 1982). HEAO-A2 was sensitive between 2 and 10 keV, so many more absorbed sources are present in the Piccinotti sample. Thus the difference in selection could lead to a bias in the warm absorber result. On the other hand, no NLS1s are present in the Piccinotti sample, suggesting that hard X-ray selection is biased against these objects, whether or not they are absorbed. That may be because the steeper hard X-ray spectrum of NLS1s make them somewhat weaker in the 2–10 keV band compared with broad-line Seyfert galaxies.
There is also a possibility of a spectroscopic bias. It is possible, although it has not yet been demonstrated, that warm absorbers and associated polarization is found more often among Seyfert 1.5 galaxies compared with Seyfert 1s. Intermediate-type narrow-line Seyfert 1 galaxies might be difficult to identify without good signal-to-noise spectra. A case in point is IRAS 20181-2244; it was thought to be a Seyfert 2 galaxy until good signal to noise spectroscopy allowed clear identification of the emission (Halpern & Moran 1998).
It is difficult to determine what kind of sample would be the best one in which to study the absorption properties in NLS1s and broad-line objects in an unbiased way. Hard X-ray selected samples should be ideal, since they choose objects whether or not they are absorbed. However, the lack of NLS1 in the Piccinotti sample make this impossible. Spectroscopically identified Seyfert galaxies may present the best sample, since both absorbed and unabsorbed Seyfert 1s and NLS1s are present.
### The Frequency of the Warm Absorber in NLS1s
Despite the potential selection bias, there are also some physical reasons to explain why a warm absorber may be less frequent in NLS1s. The simplest explanation is that it is a consequence of the steep X-ray spectrum. There will be fewer photons above the threshold energy for oxygen ionization if the spectrum is steep, and therefore, fewer appropriate ions may be created. Note that this may not be the whole story; if the warm absorber occurs in equilibrium gas, then the thermal and ionization balance would be more important for determining the properties of the warm absorber.
### The Ionization State of Warm Absorbers in NLS1s
The potential biases explored above should not affect the second result of this study, that the ionization appears to be lower in warm absorbers in NLS1s compared with Seyfert 1 galaxies with broader optical lines. Recall that this inference is based on the comparison of distributions of and edges between the NLS1s presented here and a sample of broad-line objects from Reynolds (1997): while the distributions of edge depths were indistinguishable, detection of significant optical depth in the edge was less frequent in the NLS1s. This can be inferred to be evidence that the ionization parameter is lower in NLS1s. However, there are several weaknesses in this argument. In a single-zone model, the oxygen edges are not independent. Therefore, it is conceivable that a detailed study of the fractional ionization of the gas as a function of ionization parameter will reveal that for NLS1s the $\rm O^{+7}$ inhabits only a very narrow range of ionization parameter, and the gas is more typically characterized by $\rm O^{+6}$ and fully ionized oxygen. Furthermore, assuming that the gas is in multiphase equilibrium, it may happen that $\rm O^{+7}$ is the dominate ionization state under conditions when the gas is unstable. Both of these questions require computations beyond the scope of this paper to address. Finally, the two-edge model will fail to detect very high ionization warm absorbers which may be dominated by Fe L resonance absorption and have been postulated to be present in several strong soft excess NLS1s by Nicastro, Fiore & Matt 1999.
Among Seyfert 1 galaxies with broad optical lines, there is some evidence that the ionization of the warm absorber is about the same for all objects. George et al. 1998 found, in their study of a sample of 23 [*ASCA*]{} observations of 18 objects dominated by broad-line Seyfert 1 galaxies, that the ionization parameter appears to be strongly peaked around one value, although a maximum likelihood analysis indicates a significant dispersion. This result may be due either to a sample selection effect or a physical selection effect. Possible reasons why it may be a selection effect were argued by George et al. They postulate that if the absorbing gas with lower ionization parameter happens to always be coincident with dust, the dust may block the view to the broad line region clouds causing the objects with low ionization parameters to be classified as Seyfert 2s which would have been excluded from the George et al.sample. The results presented here suggest that Seyfert 1 galaxies with less ionized warm absorbers exist.
The typically steeper X-ray spectrum may affect the conditions of the warm absorber in NLS1s. Reynolds & Fabian (1995) consider models in which the warm absorber is in thermal equilibrium with the broad-line region clouds. They show that the presence of a soft excess or a steeper X-ray spectrum appears to modify the thermal stability curve such that the warm absorber is stable over a broader range of parameters. Everything else being equal, this would predict that warm absorbers should be more common in NLS1s. At a given temperature, the values of $\xi/T$ are generally larger for steeper spectra and stronger soft excesses (see Reynolds & Fabian 1995), suggesting that the ionization should be larger in NLS1s rather than smaller. This is in contrast to what is inferred from the oxygen edge measurements but may be consistent with the presence of a very highly ionized warm absorber.
A similar result is presented by Shields, Ferland & Peterson (1995). They investigated the contribution of optically-thin broad-line region gas in photoionization but not thermal equilibrium in Seyfert 1 galaxies, and find that optically thin gas may contribute substantially to the high ionization line emission in AGN. The predicted opacity and some ionization states of this gas is also appropriate to produce the warm absorber phenomenon. Shields et al. find that when there are more soft X-rays relative to UV, the optically thin material is in general more highly ionized and will more likely appear as a warm absorber. Again, this predicts a more prevalent warm absorber in NLS1s, in contrast to what is inferred.
These models assume that the warm absorber gas is in ionization and sometimes also thermal equilibrium. That may not be the case. Krolik & Kriss (1995) and Nicastro et al. (1999) consider in some detail the situation in which the warm absorber gas is subject to a time variable photoionizing source. Since each ion has a different recombination rate, the details can be complicated. However, when the time scale of variability is shorter than the recombination time scale, the warm absorber will be overionized with respect to ionization equilibrium, generally speaking. This would predict an even higher ionization state than predicted in equilibrium models.
This discussion shows that many simple one-zone models cannot easily explain the lower ionization inferred from the oxygen edge measurements in the sample of NLS1s. However, two-zone models may be able to explain this result naturally. Evidence for more than one zone of ionized material has been found in a few well-studied broad-line Seyfert galaxies. Difference in variability behavior of the and edges in response to flux changes in MCG–6$-$30$-$15 provides one example (Otani et al. 1996), and discrepancies between absorption profiles in high optical depth warm absorbers and detailed single zone models was also interpreted as evidence for two absorbers (George et al. 1998). Evidence of an association of the outer warm absorber with dust via reddening (Reynolds 1997; Reynolds et al. 1997) and polarization (Leighly et al. 1997b) placed the outer warm absorber coincident with or exterior to the broad line region. Therefore, one possibility is that the inner warm absorber in NLS1s may be too highly ionized to be observed. Alternatively, it may not exist. For example, it may have been blown from the nucleus by radiation pressure. If only the outer warm absorber is detected, the cumulative ionization state will appear to be low. This somewhat fits our expectations because a fairly good correlation between polarization and presence of a warm absorber is observed in this sample; the dusty warm absorber should be the outer one, as dust is not expected to survive the intense heating interior to the broad-line region.
We may be able to test whether or not there is warm absorber gas present in NLS1s, and in similar quantities as in broad-line Seyfert 1 galaxies, by measuring emission lines, such as coronal lines, that are thought to associated with warm absorber gas (e.g. Porquet et al. 1998; Erkens, Appenzeller & Wagner 1997). A complicating factor may be that the ionization potentials for the common coronal line ions are quite high, around 0.2 keV. NLS1s generally have steeper soft X-ray spectra than Seyfert 1s with broader optical lines, and there is evidence that NLS1s are stronger on average at 0.2 keV than broad-line Seyfert 1s compared with the rest of the spectrum (Grupe et al. 1998a). Therefore, stronger coronal line emission might be expected from NLS1s, independent of whether warm absorber gas is present. This expectation appears to be roughly borne out in simulations by Porquet et al. Of the two continuum models used, the one that is stronger in soft X-rays requires a lower column density to produce the observed coronal line equivalent widths, at least for the lower densities. Interestingly, little dependence of the oxygen edges on the continuum shape is found.
Models for NLS1s
----------------
### Models Based on Orientation
NLS1s have narrower optical emission lines than Seyfert 1 galaxies with broad lines. It has been suggested that the origin of this behavior is that NLS1s are viewed with a face-on (pole-on) orientation. Narrower emission lines are seen because the optical line emitting clouds are confined to a plane and therefore a smaller degree of Doppler broadening is seen (e.g. Osterbrock & Pogge 1985; Puchnarewicz et al. 1992; Boller, Brandt & Fink 1996). Stronger soft X-ray emission is expected for face-on orientations from some geometrically thick accretion disk models (Madau 1988). Models in which observed line width is based on orientation have recently been supported by evidence for a correlation between radio power and the monochromatic optical luminosity (an indication of the orientation angle) and the width of the H$\beta$ lines (Wills & Brotherton 1995). It has been postulated that the emission may arise in the accretion disk (Collin-Souffrin, Hameury & Joly 1988; Kwan et al. 1995), so a face-on viewing angle would enhance the amount of observed.
To be consistent with the [*ASCA*]{} results, the observed hard X-ray photon indices should be orientation dependent. The slab thermal Comptonization model as implemented by Haardt & Maraschi (1993) predicts softer spectra as the inclination angle increases. This effect originates in the anisotropy of the first order scattering, resulting in more photons directed downward toward the slab. A similar result was obtained by Dove, Wilms & Begelman (1997) using a similar but improved slab model. This trend is opposite of what the observations require. When this model is generalized, and the hot phase is assumed to be in localized regions on the disk, the dependence on inclination decreases, and the index depends rather on the height of the hot phase blobs above the disk.
A face-on geometry might explain the low frequency of warm absorbers in NLS1s inferred from this sample. It has been suggested that intermediate Seyfert galaxy spectra are produced when the line of sight to the broad line region is partially blocked by absorbing material (e.g. Lawrence & Elvis 1982). The absorbing material may be the molecular torus thought to lie at a distance intermediate to the broad and narrow-line regions; the warm absorbing material and associated dust may be ablated from the molecular torus. Thus, warm absorbers may be more likely to be observed when the inclination is moderate rather than face-on.
The enhanced variability observe in NLS1s should also be orientation dependent. These aspects are discussed in Part 1, and both face-on and edge-on scenarios may be able to explain the enhanced variability. The edge-on model may naturally explain the fact that the variability parameters are correlated with the strength of the soft excess, since in some disk models, enhanced soft X-ray emission is expected when the viewing angle is edge-on (Laor & Netzer 1989).
### Models Based on Accretion Rate and Black Hole Mass
It has been suggested that a higher accretion rate relative to Eddington in NLS1s can explain many of their characteristic properties. If so, assuming that the efficiency of conversion of accretion energy to radiation is the same in both types of objects, the black hole mass should be smaller in NLS1s. The optical lines then will be narrower, assuming that the broad-line region emission is the same and that the motions of the broad line emitting clouds are dominated by Keplerian velocities. However, it may be more important that a high accretion rate implies an enhanced photoionizing continuum (Wandel & Boller 1998).
Several models of enhanced accretion predict a strong and hot soft excess, as discussed in Section 4.2.3. That an enhanced accretion rate should be associated with a steeper hard X-ray spectral index was first introduced first by Pounds, Done & Osborne 1995. If the geometry and high energy cutoff are assumed to be the same in all Seyfert 1 galaxies, then we can determine how much larger intrinsic luminosity of the soft component would have to be. It was shown in Section 4.1.1 that $l_s/l_h$ should be a factor of 3.8 larger in NLS1s to explain the difference in slope. Assuming a covering fraction of 0.5 and using Eq. 29 from Pietrini & Krolik (1995), the intrinsic soft X-ray luminosity would be a factor of 5.5 larger in NLS1s. However, the steeper hard X-ray spectrum may also mean that the hot phase luminosity is smaller in NLS1s. Assuming that the normalization at 1 keV is the same in both types of objects (Grupe et al. 1998a), and the spectra show an exponential cutoff at 100 keV, the hot phase luminosity in NLS1s may be 70% that of broad-line objects. Therefore, the intrinsic luminosity in the soft component would have to be about 7.5 times greater in NLS1s, assuming that everything else remains the same. If the efficiency of conversion of gravitational potential energy to radiation is the same in both types of objects, this would mean that the accretion rate should be about 7.5 times greater in NLS1s compared with broad-line objects.
As discussed in Part 1, the excess variance versus luminosity plot can also be explained by a higher accretion rate and correspondingly smaller black hole mass, assuming that the variability has the same structure in NLS1s as in Seyfert galaxies with broad optical lines. The inferred difference is a factor of 10, interestingly near the enhancement required to produce the steep hard X-ray spectrum discussed above.
The high accretion rate may also affect the properties of the warm absorber. Reynolds & Fabian 1995 hypothesize a dynamical model for the warm absorber in which neutral gas is accelerated by radiation pressure on the resonance lines. In the process, the gas is partially ionized. When the gas becomes ionized, the acceleration ceases, the inward gravitational acceleration takes over until it becomes compressed enough to recombine. The critical luminosity required to balance these two effects is estimated by Reynolds & Fabian (1995) to be about $0.05 L_{Edd}$. They speculate that if the luminosity exceeds this critical luminosity, an outflow of highly ionized material may be formed, and this may explain the lower frequency of warm absorbers in quasars, as they are thought to be radiating at a higher Eddington fraction than Seyfert galaxies.
### Other Comments on Physical Models
The two phase model has been a very successful scenario to describe the behavior of broad-line Seyfert galaxies. In this model, most of the accretion energy, rather than being released in the disk, is dissipated instead in the corona above the disk (Haardt & Maraschi 1991, 1993; Stern et al. 1995; Svensson & Zdziarski 1994). An important feature of this model is that if a sufficient amount of energy is released in the corona, the thin disk solution can exist close to the black hole (Svensson & Zdziarski 1994); otherwise, a transition to another type of accretion is expected.
This model assumes that the source of power for the disk emission is reprocessing of the coronal emission by the optically thick material. However, in NLS1s, the soft excess is sufficiently strong and the hard X-ray photon index is sufficiently steep that the soft excess cannot be powered by reprocessing (e.g. Pounds, Done & Osborne 1996). Although it has been shown that a thin disk with intrinsic emission can exist close to the black hole in Galactic black hole candidates (Gierliński et al. 1999), because of the dependence of the stability criterion on black hole mass as well as accretion rate, there can be no stable AGN thin disk solution in this situation. This fact marks a limitation of the BHC/AGN analogy as an explanation for NLS1 behavior. However, it may alleviate one of the more uncomfortable problems with this analogy. As has been widely demonstrated, NLS1s are more variable than broad-line Seyfert galaxies; in contrast, BHC in the soft state are less variable than in the hard state.
What kind of accretion flow might be expected at small radii when the radiation pressure becomes too large for the thin disk solution? One possibility that has been proposed for hard state Galactic black hole candidates and broad-line radio galaxies is that the inner regions form a hot, optically thin, geometrically thick torus, either the two-temperature torus (SLE; e.g. Shapiro, Lightman & Eardley 1976) or an advection dominated flow (ADAF; for a review, Narayan, Mahadevan & Quataert 1999). These flows are typified by rather hard X-ray power-law emission and no soft excess; therefore they are likely not to be applicable to NLS1s.
It was originally thought that when the accretion rate is high, there would be a radiation pressure dominated disk (e.g. Shakura & Sunyaev 1973). However, this solution is shown to be unstable if the viscosity is proportional to the radiation pressure, although it may be stable if the viscosity is proportional to the gas pressure only (Lightman & Eardley 1974). At higher accretion rates lies the slim disk solution, in which part of the energy is advected into the black hole (e.g. Abramowicz et al. 1988). In NLS1s, it is possible that the accretion rate is in the radiation dominated region and it instead develops an unsteady flow. Such a scenario has recently been proposed for the superluminal black hole binary GRS 1915+105 by Belloni et al. 1997. This object may undergo unsteady accretion in a similar way as dwarf novae, except the Lightman-Eardley instability (which operates between the thin disk and slim disk solutions, with the unstable radiation pressure dominated disk in between) operates rather than the thermal-viscous instability. The time scales for the limit cycle are much longer for AGN (on the order of years); however, on shorter time scales, during the accreting portion of the cycle, the accretion happens unsteadily, and rapid variability is observed as heating fronts move through the disk at the sound speed, and material is removed on the infall time scale (Belloni et al. 1997).
Such a scenario may be attractive for NLS1s. Slim disks can produce soft X-rays (Szuszkiewicz, Malkan & Abramowicz 1996). The sound crossing speed and infall time scales in the inner region may be appropriate for the rapid variability observed in NLS1s. The heating waves could conceivably be coherent, and therefore produce large amplitude flares. The longer time scale variability due to the limit cycle behavior may provide a way to explain the very large amplitude variability observed from some NLS1s (IC 3599: Grupe et al. 1995a; WPVS 007: Grupe et al. 1995b; RX J0134$-$42: Grupe et al. 1999b).
The correlation between the excess variance and strength of the soft excess might be explained as a consequence of a variable transition radius $r_t$ from the usual AGN geometry, that is, a thin disk plus corona from magnetic flares existing at $r>r_t$, to the proposed highly variable unsteady inner disk at $r<r_t$. The location of this transition radius may be a function of specific accretion rate. When $r_t$ is small, only a relatively small region is characterized by the highly variable unstable disk, while there is a relatively large region containing the usual thermal corona present in broad-line objects. Therefore, the soft excess is moderately strong, and the hard power law is relatively strong, and, while the excess variance is higher than that of broad-line Seyfert 1 galaxies, the variability is not detectably non-Gaussian because the coherent variability from the inner region is diluted by stochastic variability from the flaring outer region. When $r_t$ is large, the highly variable unstable inner disk dominates the central parts of the AGN, leading to a very strong soft excess. The hard tail is weaker, because less of the accretion energy goes into the usual thermal corona. Because the variability is dominated by the very variable inner region, the excess variance is again large, but since the coherent variability is minimally diluted, non-Gaussianity is detected. Note that because the region sizes change approximately in concordance with the emission from each region, the compactnesses may stay approximately the same.
Summary and Future Observations
===============================
I present a comprehensive and uniform analysis of 25 [*ASCA*]{} observations from 23 Narrow-line Seyfert 1 galaxies. The results of spectral analysis and correlations are reported here; the results of the time series analysis are reported in Part 1. The primary results of this paper are the following:
- A maximum likelihood analysis confirms that the hard X-ray photon index is significantly steeper at $>$90% confidence in this sample of NLS1s than in a random sample of Seyfert galaxies with broad optical lines. As previously noted, this can be explained by two-phase thermal Comptonization models if the soft photon input is greater in NLS1s than in broad-line objects. This view may be supported by the observed correlation between the photon index and the H$\beta$ FWHM among the objects in this sample. Alternatively, the steep photon indices may originate in a single scattering from a power law distribution of electrons.
- Soft excess emission was detected in 17 of the 19 objects which had no significant absorption. While the luminosities span two orders of magnitude, the temperatures are roughly consistent, a result that contradicts predictions of the simplest thin disk model. The soft excess strength above the power law spans a large range (a factor of 50), and the strength of the soft excess was found to be correlated with the variability parameters as well as the [*ROSAT*]{} photon index. Possible origins for the soft excess include primary emission from an accretion disk characterized by a high accretion rate, reprocessing in an ionized disk, and the signature of Comptonization of black body emission.
- Evidence was found that the warm absorber is less likely to be observed in NLS1s and when present has typically lower ionization than in broad-line Seyfert 1 galaxies. However, this result was obtained using a two-edge parameterization which would not necessarily be sensitive to warm absorbers with very high ionization. This result may imply that the inner warm absorber is either absent or too highly ionized to be detected in these objects, but the outer warm absorber, typified by a lower ionization parameter, remains.
- The iron line equivalent width average and distribution was found to be the same in the sample of NLS1s as in a random sample of Seyfert 1 galaxies with broad optical lines, a result which may imply that the amount of reprocessing is the same in both types of objects. However, this result is qualified by the generally poorer statistics at the iron line in the NLS1 sample, and the fact that emission from ionized iron, which may have either enhanced or reduced fluorescence yield, was found in a few objects.
Models based on orientation and high accretion rate were examined in the light of these results. A model based a face-on orientation possibly explains the narrow emission lines, stronger soft excess and less frequent warm absorbers, but cannot explain the steep power law, the variability results or the low ionization of the warm absorber. A model based on an edge on viewing angle may be able to explain the variability results, the steep soft excess and perhaps also the steep hard X-ray spectrum but requires NLS1 to consistently be inclined to their host galaxies more than Seyfert 1 galaxies with broad optical lines, and the dominant motion of broad line region clouds would be required to be along the symmetry axis.
A model based on a higher accretion rate relative to Eddington generally fairs the best. It can explain the steep soft excess due to a shift of the accretion disk spectrum to high energies; this also explains the narrow optical emission lines through stronger ionization and also the steep hard X-ray photon index through stronger Comptonization cooling. The rapid variability is naturally explained by the requirement of a smaller mass black hole. The nearly constant soft excess temperature is predicted at high accretion rates in some accretion disk models. The less frequent warm absorber may be explained if it is too highly ionized to be seen or has been blown out of the system.
The [*ASCA*]{} observations and results presented here lead to further questions which may be answerable in the future.
- Observations at higher energies are necessary to make significant progress on understanding the origin of the steep hard X-ray spectrum in NLS1s. Observations up to 20 keV will allow us to search for the Compton reflection component and obtain information on the amount of reprocessing. Observations to higher energies will allow us to determine whether the hard X-ray emission process is thermal or nonthermal. In addition, studies of the photon index spectral variability may yield some constraints on these processes.
- The nature of the soft excess in the [*ASCA*]{} band remains a mystery. Observations with a broader band pass down to 0.1 keV and better statistics would allow us to better constrain the shape. Observations with better energy resolution may reveal line emission characteristic of reprocessing in a ionized disk. Observations of variations in the shape of this component as a function of time also may be valuable.
- Better statistics in the iron line region are needed to determine the importance of reprocessing and also to constrain the inclination and properties of the reprocessing material including ionization and distance from the black hole.
- Observations with good statistics and energy resolution will allow us to determine the properties of the warm absorber in NLS1s with less ambiguity.
KML gratefully thanks all those people who built and operate [*ASCA*]{}. This project could have never attained this form without the help and support of Jules Halpern. Many thanks go to Dirk Grupe for many various kinds of help and advice. Useful discussions with Karl Forster, Julian Krolik, Herman Marshall, Tahir Yaqoob and Andrzej Zdziarski are acknowledged. The following are thanked for a critical reading of a draft: Joachim Siebert, Jules Halpern & Tahir Yaqoob. This research has made use of the NASA/IPAC extragalactic database (NED) which is operated by the Jet Propulsion Laboratory, Caltech, under contract with the National Aeronautics and Space Administration. This research has made use of data obtained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center. KML gratefully acknowledges support through NAG5-3307 and NAG5-7261 ([*ASCA*]{}) and NAG5-7971 (LTSA) .
Appendix – Notes on Individual Objects
======================================
### Ton S180:
GIS3 5.5–7.5 keV was excluded from spectral fitting.
Turner, George & Nandra (1998) report evidence for soft X-ray line emission in this [*ASCA*]{} observation of Ton S180. The fact that Turner et al. model the spectra from both CCD detectors and also do not model the soft excess component leads to some doubt about this result. As discussed in Section 2.1, data taken after the beginning of 1996 tended to show large discrepancies between the SIS0 and SIS1 detectors. The Ton S180 observation was made in June 1996 and displays one of the worst examples of this discrepancy found in this sample of data. The SIS1 residuals fall markedly below the SIS0 residuals for energies less than 0.8 keV (Leighly 1998). Thus, spectral complexity appears to be present in the SIS1 detector but not in SIS0. Spectral fitting confirms this conjecture; a soft X-ray line appears significant in the SIS1 spectrum, but there is no evidence for it in SIS0. When a soft X-ray line is added to the best fit model presented here, the improvement in $\chi^2$ was 6.2 for 2 d.o.f. and the equivalent width is 7.4 eV; therefore this feature is not significant.
### PKS 0558$-$504:
GIS3 5.0–7.0 keV was excluded from spectral fitting.
### Mrk 766:
The spectral classification of Mrk 766 has been debated. Osterbrock & Pogge (1985) classify this object as a Seyfert 1.5; later, however, Goodrich (1989) includes it in his NLS1 spectropolarimetry sample. Goodrich (1989) published a spectrum near H$\beta$ from this object. The typical Lorentzian NLS1 line profile can clearly be seen, and measured FWHM from the plot is 1700 km/s. I speculate that the Seyfert 1.5 classification arose from a failure of a single Gaussian fit to the Lorentzian profile; this is a common situation in NLS1 optical spectra NLS1s (e.g. Goncalves, Veron & Veron-Cetty 1998). Furthermore, the H$\beta$ appears weak compared with \[\] in Goodrich 1989; however, this could be a consequence of a Balmer decrement arising from a dusty warm absorber responsible for the ionized absorption in the X-ray band (e.g. Leighly et al. 1997b) and the red optical spectrum (Molendi & Maccacaro 1994) and not intrinsic to the emitted spectrum.
### IRAS 13349+2438:
The FWHM H$\beta$ of IRAS 13349+2438 is $2200\,\rm km\,s^{-1}$ and therefore it is not an NLS1, technically speaking. However, as noted by Brandt, Mathur & Elvis 1997, the steep spectrum and variability make it more similar to NLS1s than to Seyfert 1s or quasars with broad optical lines. Therefore it is also included in this sample.
### PG 1244+226:
The soft X-ray line feature reported here is qualitatively the same as that reported by Fiore et al. 1998; however, a smaller equivalent width is reported here (35 compared with 64 eV). The origin of this difference is that here the spectra over the full [*ASCA*]{} band pass is modeled and a soft excess component is included whereas in Fiore et al., only 0.4–4.0 keV region is modeled and the continuum consists only of a power law. The soft excess component models part of the positive residuals resulting in smaller equivalent width.
The feature in the PG 1244+226 spectrum was alternatively modeled as an edge at 1.17keV by Fiore et al. 1998. I found that an edge at 1.17 keV also modeled the spectrum adequately ($\chi^2=487$/474 d.o.f.). The unresolved line model yielded a $\chi^2$ of 489 for the same number of degrees of freedom, indicating a difference in $\chi^2$ of 2. The likelihood ratio then is just 2.71 and therefore the difference is not significant.
### IRAS 17020+4544:
GIS3 6.0–8.0 keV was excluded from spectral fitting.
### Kaz 163:
The ratio of data to model indicates that there may be a weak soft excess in this object. However, addition of this component leads to only a small reduction in $\chi^2$ ($\Delta\chi^2=8.7$) and therefore it is required with less than 99% confidence.
### IRAS 20181$-$2244:
Marginal evidence for a warm absorber was found during this analysis which was not reported by Halpern & Moran 1998.
### Ark 564:
Analysis of this object was complicated by the fact that the observation was made partly with a level discriminator and partly without, and the level discriminator was set at different values during the time that it was on. This required different response matrices for the different level discriminator states. Spectra were accumulated excluding the data with the high level discriminator. The spectra were fit simultaneously with spectra from a coordinate [*RXTE*]{} observation. The details will be reported in Leighly et al. in prep.
Abramowicz, M. A., Czerny, B., Lasota, J. P., & Szuszkiewicz, E., 1988, , 332, 646
Belloni, T., Méndez, M., King, A. R., van der Klis, M., & van Paradijis, J., 1997, ApJL, 145
Bergeron, J., & Knuth, D., 1980, A&A, 85, 11
Bergeron, J., & Knuth, D., 1984, , 207, 263
Berriman, G., Schmidt, G. D., West, S. C. & Stockman, H. S. 1990, ApJS, 74, 869
Boller, Th., Bertoldi, F., Dennefeld, M., & Voges, W. 1998, A&AS, 129, 87
Boller, Th., Brandt, W. N., & Fink, H. 1996, A&A, 305, 53
Boller, Th., Meurs, E. J. A., Brinkmann, W., Fink, H., Zimmermann, U., & Adorf, H.-M. 1992, A&A, 261, 57
Boroson, T. A. & Green, R. F. 1992, ApJS, 80, 109
Brandt, W. N., Fabian, A. C., & Pounds, K. A., 1996, , 278, 326
Brandt, W. N., Mathur, S., & Elvis, M., 1997, , 285, 25p
Brandt, W. N., Mathur, S., Reynolds, C. S., & Elvis, M. 1997, , 292, 407
Brinkmann, W., Kawai, N., Ogasaka, Y., & Siebert, J., 1996, A&A, 316, 9
Chakrabarti, S., & Titarchuk, L. G. 1995, , 455, 623
Cohen, R. D. 1983, , 273, 489
Collin-Souffrin, S., Hameury, J.-M., & Joly, M., 1988, A&A, 205, 19
Colpi, M. 1988, , 326, 223
Comastri, A., et al. 1998, A&A, 333, 31
Comastri, A., Molendi, S., & Ulrich, M. H., 1997, “X-ray Imaging and Spectroscopy of Cosmic Hot Plasmas”, ed. F. Makino & K. Mitsuda (Tokyo: University Academy Press), 279
Coppi, P. S., 1999, in proc. “High Energy Processes in Accreting Black Holes”, eds. J. Poutanen & R. Svensson, ASP Conf. Series, Vol. 161, p. 375
Corbin, M. R., 1997, ApJS, 113, 245
Cui, W., Ebisawa, K., Dotani, T., & Kubota, A., 1998, ApJL, 493, 75
Czerny, B., & Elvis, M., 1987, , 321, 305
Czerny, B., & Życki, P. T., 1994, ApJL, 431, 5
Dickey, J. M., & Lockman, F. J., 1990, Ann. Rev. Astron. & Astrophys. 28, 215
Dotani T. 1998, ASCA Calibration Report, http://www.astro.isas.ac.jp/ dotani/rdd.html
Dove, J. B., Wilms, J., & Begelman, M. C., 1997, , 487, 747
Ebisawa, K., Titarchuk, L., & Chakrabarti, S. K., 1996, PASJ, 48, 59
Elvis, M., Lockman, F. J., & Wilkes, B. J., 1989, , 97, 777
Erkens, U., Appenzeller, I., & Wagner, S., 1997, A&A, 323, 707
Fabian, A. C., Guilbert, P. W., Arnaud, K. A., Shafer, R. A., Tennant, A. F., & Ward, M. J., 1986, , 218, 457
Fabian, A. C., Rees, M. J., Stella, L., & White, N. E., 1989, , 238, 729
Feigelson, E. D., & Nelson, P. I., 1985, , 293, 192
Fiore, F., Matt, G., Cappi, M., Elvis, M., Leighly, K. M., Nicastro, F., Piro, L., Siemiginowska, A., & Wilkes, B. J., 1998, , 298, 103
Fiore, F., Matt, G., & Nicastro, F., 1997, , 284, 731
Forster, K. 1998, PhD Thesis, Columbia University
Forster, K. & Halpern, J. P., 1996, , 468, 565
Gelderman, R., & Whittle, M., 1994, ApJS, 91, 491
George, I. M. & Fabian, A. C., 1991, , 249, 352
George, I. M., Turner, T. J., Netzer, H., Nandra, K., Mushotzky, R. F., & Yaqoob, T., 1998, ApJS, 114, 73
Ghisellini, G., & Haardt, F., 1994, ApJL, 429, 53
Gierliński, M., Zdziarski, A. A., Poutanen, J., Coppi, P., Ebisawa, K., & Johnson, W. N., 1999, MNRAS, in press
Goncalves, A. C., Véron, P., & Véron-Cetty, M.-P. 1998, in Proc. “Structure and Kinematics of Quasar Broad Line Regions”, eds. C. M. Gaskell, W. N. Brandt, M. Dietrich, D. Dultzin-Hacyan & M. Eracleous, in press
Goodrich, R. W. 1989, , 342, 224
Grupe, D. 1996, PhD Thesis, University of Göttingen
Grupe, D., Beuermann, K., Mannheim, K., Bade, N., Thomas, H.-C., De Martino, D., & Schwope, A., 1995a, A&A, 299, 5
Grupe, D., Beuermann, K., Mannheim, K., Thomas, H.-C., Fink, H. H., & De Martino, D., 1995b, A&A, 300, 21
Grupe, D., Beuermann, K., Thomas, H.-C., Mannheim, K. & Fink, H. H. 1998a, A&A, 330, 25
Grupe, D., Wills, B. J., Wills, D., & Beuermann, K., 1998b, A&A, 333, 827
Grupe, D., Beuermann, K., Mannheim, K., & Thomas, H.-C., 1999a, A&A, in press
Grupe, D., Leighly, K. M., Thomas, H.-C., Laurent-Muehleisen, S. A., 1999b, A&A, submitted
Guainazzi, M., Matsuoka, M., Piro, L., Mihara, T. & Yamauchi, M., 1994, ApJL, 436, 35
Guainazzi, M., Mihara, T., Otani, C., & Matsuoka, M., 1996, PASJ, 48, 781
Guainazzi, M., Piro, L., Capalbi, M., Parmar, A. N., Yamauchi, M., & Matsuoka, M., 1998a, A&A, 339, 337
Guainazzi, M., et al., 1998b, A&A, 339, 327
Haardt, F., & Maraschi, L., 1991, ApJL, 380, 51
Haardt, F., & Maraschi, L., 1993, , 413, 507
Haardt, F., Maraschi, L., & Ghisellini, G., 1994, , 432, 95
Halpern, J. P., & Moran, E. C. 1998, , 494, 194
Hayashida, K. 1996, Proc. “Emission Lines in AGN: New Methods and Techniques”, ed. B. M. Peterson, F.-Z. Cheng & A. S. Wilson, (San Fransisco: ASP), 40
Isobe, T., & Feigelson, E. D., 1990, BAAS, 22, 917
Iwasawa, K., Brandt, W. N., & Fabian, A. C., 1998, , 293, 251
Joly, M., 1993, Ann. Phys. Fr., 18, 241
Kay, L. E., Magalhães, A. M., Elizalde, F., Rodrigues, C., 1999, ApJ, 518, 219
Korista, K. T. 1991, , 102, 41
Korista, K., Ferland, G., & Baldwin, J., 1997, ApJ, 487, 555
Koski, A. T. 1978, , 223, 56
Krolik, J. H., & Kallman, T. R. 1988, , 324, 714
Krolik, J. H., & Kriss, G. A., 1995, , 447, 512
Kwan, J., Cheng, F.-Z., Fang, L.-Z., Zheng, W., & Ge, J., 1995, , 440, 628
Laor, A., Fiore, F., Elvis, M., Wilkes, B. J., & McDowell, J. C., 1997a, , 477, 93
Laor, A., Jannuzzi, B. T., Green, R. F., & Boroson, T. A., 1997b, , 489, 656
Laor, A., & Netzer, H., 1989, , 238, 897
Lawrence, A., & Elvis, M., 1982, , 256, 410
Lawrence, A., Elvis, M., Wilkes, B. J., McHardy, I., & Brandt, N. 1997, , 286, 879
Leighly, K. M. 1998, in proc. “Accretion Processes in Astrophysical Systems: Some Like it Hot!”, eds. S. S. Holt, T. R. Kallman, (AIP: Woodbury, New York), p. 199
Leighly, K. M., Kay, L. E., Wills, B. J., Wills, D., & Grupe, D. 1997b, ApJL, 489, 137
Leighly, K. M., Mushotzky, R. F., Yaqoob, T., Kunieda, K., & Edelson, R., 1996, , 469, 14
Leighly, K. M., Mushotzky, R. F., Nandra, K., Forster, K., 1997a, ApJL, 489, 25
Leighly, K. M., & O’Brien, P. T., 1997, ApJL, 481, 15
Lightman, A. P. & Eardley, D. M. 1974, ApJL, 187, 1
Maccacaro, T., Gioia, I. M., Wolter, A., Zamorani, G., & Stocke, J. T., 1988, , 326, 680
Madau, P. 1998, , 327, 116
Makuch, R. W., Escobar, M., & Merrill III, S., 1991, Appl. Statist. 40, 19
Mason, K. O., Puchnarewicz, E. M., & Jones, L. R., 1996, , 283, 26
Matt, G., Fabian, A. C., & Ross, R. R., 1993b, , 262, 179
Matt, G., Fabian, A. C., & Ross, R. R. 1993a, , 264, 839
Molendi, S., & Maccacaro, T., 1994, A&A, 291, 420
Moran, E. C., Halpern, J. P., & Helfand, D. J., 1996, ApJS, 106, 341
Morrison, R., & McCammon, D., 1983, , 270, 119
Murphy, E. M., Lockman, F. J., Laor, A., & Elvis, M. 1996, ApJS, 105, 369
Mushotzky, R. F. 1982, , 256, 92
Mushotzky, R. F., Fabian, A. C., Iwasawa, K., Kunieda, H., Matsuoka, M., Nandra, K., & Tanaka, Y. 1995, , 272, 9
Nandra, K., George, I. M., Mushotzky, R. F., Turner, T. J., & Yaqoob, T., 1997a, , 477, 602
Nandra, K., George, I. M., Mushotzky, R. F., Turner, T. J., & Yaqoob, T., 1997b, ApJL, 488, 91
Nandra, K., & Pounds, K. A., 1994, , 267, 974
Narayan, R., Mahadevan, R., & Quataert, E., 1999, in “Theory of Black Hole Accretion Disks”, eds. M. A. Abramowicz, G. Bjornsson, & J. E. Pringle, in press
Netzer, H., 1996, , 473, 781
Nicastro, F., Fiore, F., Perola, G. C., & Elvis, M., 1999, , 512, 184
Nicastro, F., Fiore, F., & Matt, G., 1999, , 517, 108
Nishimura, J., Mitsuda, K, & Itoh, M., 1986, PASJ, 38, 819
Nowak, M. A. 1995, PASP, 107, 1207
Osterbrock, D. E., & Pogge, R. W. 1985, , 297, 166
Otani, C., Kii, T. & Miya, K. 1996, Proc. “Röntgenstrahlung from the Universe”, eds. Zimmermann, H. U., Trümper, J., and Yorke H., 1996, MPE Report 263, 491
Otani, C. et al. 1996, PASJ, 48, 211
Phillips, M. M., 1976, , 208, 37
Phlips, B. F. et al., 1996, ApJ, 465, 907
Piccinotti, G., Mushotzky, R. F., Boldt, E. A., Holt, S. S., Marshall, F. E., Serlemitsos, P. J., & Shafer, R. A., 1982, , 253, 485
Pietrini, P., & Krolik, J. H. 1995, , 447, 526
Pogge, R. W., & Owen, J. M. 1993, OSU Internal Report 93-01
Porquet, D., Dumont, A.-M., Collin, S., & Mouchet, M., 1999, A&A, 341, 58
Pounds, K. A., Done, C., & Osborne, J., 1996, , 277, L5
Press, W. H., Teukolsky, S. A., Vettering, W. T., & Flannery, B. P. 1992, [*Numerical Recipes*]{}, (Cambridge University Press: Cambridge)
Puchnarewicz, E. M. et al. 1992, , 256, 589
Remillard, R. A., Bradt, H. V., Buckley, D. A. H., Roberts, W., Schwartz, D. A., Tuohy, I. R., & Wood, K. 1986, , 301, 742
Reynolds, C. S., 1997, , 286, 513
Reynolds, C. S., & Fabian, A. C., 1995, , 273, 1167
Reynolds, C. S., Ward, M. J., Fabian, A. C., & Celotti, A. 1997, , 291, 403
Ross, R. R., Fabian, A. C., & Mineshige, S., 1992, , 258, 189
Rybicki, G. B., & Lightman, A. P., 1979, “Radiative Processes in Astrophysics” (Wiley: New York)
Schartel, N., Walter, R., Fink, H. H., & Trümper, J., 1996, A&A, 307, 33
Shakura, N. I., & Sunyaev, R. A., 1973, A&A, 24, 337
Shapiro, S. L., Lightman, A. P., & Eardley, D. M., 1976, , 204, 187
Shields, J. C., Ferland, G. J., & Peterson, B. M., 1995, , 441, 507
Shimura, T., & Takahara, F., 1993, , 419, 78
Shrader, C., & Titarchuk, L., 1998, ApJL, 499, 31
Smith, P. S., Schmidt, G. D., Allen, R. G. & Hines, D. C. 1998, , 448, 202
Stern, B. E., Poutanen, J., & Svensson, R. 1995, ApJL, 449, 13
Swensson, R., 1994, ApJS, 92, 585
Svensson, R., & Zdziarski, A. A., 1994, , 436, 599
Szuszkiewicz, E., Malkan, M. A., & Abramowicz, M. A., 1996, , 458, 474.
Tanaka, Y., et al. 1995, Nature, 375, 659
Turner, T. J., George, I. M., & Nandra, K., 1998, , 508, 648
Turner, T. J., & Pounds, K. A., 1989, , 240, 833
Wandel, A., & Boller, T., 1998, A&A, 331, 884
Wilkes, B. J., Elvis, M. & McHardy, I. 1987, , 321, 23
Wills, B. J., & Brotherton, M. S., 1995, ApJL, 448, 81
Wills, B., J., Laor, A., Brotherton, M. S., Wills, D., Wilkes, B. J., Ferland, G. J., & Shang, Z., 1999, ApJL, 515, 53
Wills, B. J., Wills, D., Evans, N. J., Natta, A., Thompson, K. L., Breger, M., & Sitko, M. L. 1992, , 400, 96
Yaqoob, T., 1998, ApJ, 500, 893
Zheng, W., Kriss, G. A., & Davidsen, A. F. 1995, ApJ, 440, 606
Zheng, W., & O’Brien, P. T. 1990, , 353, 433
Życki, P. T., Krolik, J. H., Zdziarski, A. A., & Kallman, T. R., 1994, , 437, 597
[lllllllll]{}
Mrk 335 & 0.025 & 1640 & 95 & 0.62 & & 0.23 & BG92\
I Zw 1 & 0.061 & 1240 & 51 & 1.47 & & 0.43 & BG92\
Ton S180 & 0.062 & 980 & 45 & 0.8 & 640 & 0.12 & G96\
PHL 1092 & 0.396 & 1790 & 63 $^1$ & 1.8 & & 0.91 & BK80 LEWMB97 this paper\
RX J0439$-$45 & 0.224 & 1010 & 65 & 0.55 & 1020 & 0.17 & G96\
NAB 0205+024 & 0.155 & 1050 & 58 & 0.62 & & 0.36 & ZO90 GW94 BK84 K91\
PKS 0558$-$504 & 0.137 & 1250 & 45 & 1.56 $^2$ & & 0.04 & C97 R86\
1H 0707$-$495 & 0.0411 & 1050 & 32 & 1.36 & 1516 & 0.19 & this paper\
Mrk 142 & 0.04494 & 1470 & 67 & 1.11 & 404 & 0.16 & this paper; G96\
RE 1034+39 & 0.04244 & 840 & 20 & 0.32 & 543 & 1.25 & this paper MPJ96 GVV98\
NGC 4051 & 0.00242 & 1150 & 28 & 0.77 & 325 & 0.92 & this paper\
PG 1211+143 & 0.0809 & 1860 & 84 & 0.52 & & 0.14 & BG92\
Mrk 766 & 0.01293 & 850 $^3$ & 71 & 0.52 & 360 & 1.85 & OP85 G89 GWWB98\
PG 1244+026 & 0.048 & 830 & 41 & 1.20 & & 0.41 & BG92\
IRAS 13224$-$3809 & 0.0667 & 650 & 23 & 2.42 & 810 & 0.60 & this paper; K98\
IRAS 13349+2438 & 0.108 & 2200 & 72 & 6.5 $^4$ & & 0.13 & GWWB98\
PG 1404+226 & 0.098 & 880 & 54 & 1.01 & & 0.12 & BG92\
Mrk 478 & 0.079 & 1450 & 64 & 1.19 & & 0.15 & BG92\
IRAS 17020+4544 & 0.0604 & 1040 & 23 & 1.86 & 1021 & 2.48 & this paper; LKWWG97\
Mrk 507 & 0.055 & 1150 & 19 & 1.45 & 454 & 0.43 & this paper\
KAZ 163 & 0.063 & 1620 & 76 & 0.57 & 534 & 0.86 & this paper\
IRAS 20181$-$2244 & 0.185 & 370 & 22 & 0.89 & 537 & 6.12 & this paper; K98; HM98\
Ark 564 & 0.024 & 950 & 43 & 0.95 & 350 & 0.96 & this paper\
[lccccc]{}
Mrk 335 & 739 & 969 & 749 & 939 & 786 I Zw 1 & 549 & 627 & 627 & 626 & 623 Ton S180 & 856 & 1109 & 888 & 1109 & 1011 PHL 1092 & 246 & 335 & 239 & 335 & 265 RX J0439$-$45 & 330 & 475 & 320 & 475 & 341 NAB 0205+024 & 599 & 638 & 602 & 638 & 612 PKS 0558$-$504 & 1093 & 1126 & 1087 & 1126 & 1113 1H 0707$-$495 & 368 & 955 & 531 & 897 & 623 Mrk 142 & 365 & 364 & 336 & 364 & 345 RE 1034+39 & 324 & 415 & 315 & 400 & 330 NGC 4051(1) & 1177 & 2759 & 1313 & 2329 & 1698 NGC 4051(2) & 1850 & 4191 & 2288 & 3716 & 2646 PG 1211+143 & 582 & 1047 & 544 & 903 & 625 Mrk 766 & 1385 & 2102 & 1506 & 1535 & 1425 PG 1244+026 & 480 & 560 & 512 & 560 & 552 IRAS 13224$-$3809 & 404 & 1207 & 509 & 1207 & 1207 IRAS 13349+2438 & 568 & 583 & 569 & 541 & 549 PG 1404+226 & 230 & 565 & 276 & 565 & 371 Mrk 478 & 530 & 770 & 590 & 677 & 620 IRAS 17020+4544 & 939 & 1347 & 1347 & 1327 & 1080 Mrk 507 & 152 & 150 & 150 & 142 & 137 KAZ 163 & 249 & 221 & 212 & 220 & 218 IRAS 20181$-$2244 & 400 & 663 & 663 & 656 & 456 Ark 564 & 1402 & 2185 & 1590 & 2185 & 1819
[lllllllllllllllll]{}
Mrk 335 & $2.08 \pm 0.03$ & $2.6^{+2.3}_{-1.2}$ & $570^{+440}_{-250}$ & $1.0 \times 10^6$ & 19 & 705/732 I Zw 1 & $2.42^{+0.07}_{-0.08}$ & $0.06 \pm 0.04$ & $0.03^{+0.24}_{-0.03}$ & $3.0 \times 10^4$ & 10 & 598/544 NAB 0205+024 & $2.24\pm 0.07$ & $0.28^{+0.32}_{-0.18}$ & $33^{+120}_{-33}$ & $3.0\times 10^4$ & 8(m) & 594/595 NGC 4051(1) & $2.05^{+0.04}_{-0.03}$ & $0.11^{+0.11}_{-0.05}$ & $1.9^{+16}_{-0.9}$ & $1.0 \times 10^6$ & 15 & 1243/1167 & NGC 4051(2) & $2.00 \pm 0.02$ & $0.21 \pm 0.04$ & $1.7^{+0.6}_{-0.4}$ & $1.0 \times 10^6$ & 90 & 2058/1844 Mrk 766 & $1.98 \pm 0.03$ & $0.30^{+0.08}_{-0.06}$ & $9.8^{+7.9}_{-4.0}$ & $3.0 \times 10^4$ & 154 & 1338/1378 PG 1244+026 & $2.40^{+0.12}_{-0.13}$ & $0.82^{+0.88}_{-0.56}$ & $260^{+680}_{-210}$ & $3.0 \times 10^4$ & 6(m) & 483/471 IRAS 13224$-$3809 & $2.07^{+0.27}_{-0.30}$ & $1.12^{+0.48}_{-0.51}$ & $86^{+104}_{-35}$ & $3.0 \times 10^4$ & 6(m) & 401/396 IRAS 13349+2438 & $2.39 \pm 0.05$ & $0.18^{+0.07}_{-0.05}$ & $2.1^{+2.7}_{-1.1}$ & $1.0 \times 10^6$ & 38 & 540/563 Mrk 478 & $2.22^{+0.16}_{-0.13}$ & $0.54^{+0.38}_{-0.28}$ & $21^{+27}_{-16}$ & $1.0\times 10^6$ & 14 & 567/523 IRAS 17020+4544 & $2.45\pm 0.05$ & $0.22^{+0.06}_{-0.05}$ & $24^{+24}_{-15}$ & $3.0 \times 10^4$ & 80 & 1043/935 KAZ 163 & $1.50^{+0.28}_{-0.30}$ & $1.2^{+0.8}_{-0.7}$ & $60^{+130}_{-40}$ & $3.0 \times 10^4$ & 7(m) & 205/245 IRAS 20181$-$2244 & $2.55^{+0.12}_{-0.11}$ & $0.24^{+0.21}_{-0.16}$ & $60^{+290}_{-60}$ & $3.0 \times 10^4$ & 7(m) & 449/395 Ark 564 & $2.59^{+0.01}_{-0.02}$ & $1.0 \pm 0.4$ & $750^{+280}_{-160}$ & $1.0 \times 10^6$ & 10 & 1439/1393
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to3.5in
------------------------------------------------------------------------
to4.0in
------------------------------------------------------------------------
to3.0in
------------------------------------------------------------------------
to5.0in
------------------------------------------------------------------------
to5.0in
------------------------------------------------------------------------
to8.0in
------------------------------------------------------------------------
| {
"pile_set_name": "ArXiv"
} |
=15.5pt
-.8cm
2 cm
1.5cm [Takuya Okuda [^1]]{}
.5cm
[California Institute of Technology 452-48,\
Pasadena, CA 91125, USA]{}
1.0cm [Shigeki Sugimoto [^2]]{}
.5cm
1.5cm
[**Abstract**]{}
We study the late time behavior of the boundary state representing the rolling tachyon constructed by Sen. It is found that the coupling of the rolling tachyon to massive modes of the closed string grows exponentially as the system evolves. We argue that the description of rolling tachyon by a boundary state is valid during the finite time determined by string coupling, and that energy could be dissipated to the bulk beyond this time. We also comment on the relation between the rolling tachyon boundary state and the spacelike D-brane boundary state.
Introduction
============
Initiated by a series of papers [@Sen:2002nu; @Sen:2002in; @Sen:2002an] by A. Sen, the dynamics of a time-dependent tachyon background has been intensely studied recently. In [@Sen:2002nu], he has presented an iterative scheme to construct a solution in Witten’s Open String Field Theory that describes the tachyon rolling up and down its potential in unstable D-brane systems (see also [@Sen:2002vv; @Moeller:2002vx; @Kluson:2002te]). In [@Sen:2002nu; @Sen:2002in; @Sen:2002vv], he has also constructed a boundary state corresponding to the solution. The boundary state representing the rolling tachyon is constructed by deforming the boundary state of an unstable D-brane by an exactly marginal operator. Therefore, it is expected to define an open string background which is an exact solution in the weak coupling limit, where the back reaction of the closed string is ignored. The boundary state is also a source for closed strings. Once the boundary state is obtained, one can extract the time evolution of the source for each closed string mode. In particular, the energy momentum tensor can be obtained from the coefficients of the level $(-1,-1)$ states in the boundary state and it has been argued that the system will evolve to a pressureless gas with non-zero energy density called tachyon matter [@Sen:2002in]. A similar behavior has been found in the context of Boundary String Field Theory [@Sugimoto:2002fp; @Minahan:2002if].
Tachyon matter appears to be a mysterious object in string theory. It is supposedly the decay product of the unstable D-brane. While the energy remains preserved and localized in the world-volume, there are no physical open string excitations [@Sen:2002an; @Ishida:2002fr]. Furthermore, since the energy momentum tensor does not oscillate, tachyon matter does not seem to turn into the radiation of massless closed string modes. These features are crucial in the investigation of cosmology based on rolling tachyon [@cosmology; @dark; @matter; @inflation]. For example, from the observation that tachyon matter does not decay into light particles, tachyon matter has been regarded as a candidate of dark matter in several works [@Sen:2002in; @dark; @matter]. It also implies the reheating problem in tachyon inflation scenarios [@inflation] as pointed out in [@reheating]. However, these analyses have only taken into account the behavior of the source for massless closed string modes in the rolling tachyon boundary state and have not seriously considered the effect of infinitely many massive closed string modes.
In this paper, we examine the rolling tachyon boundary state further to higher levels and compute the coefficients of these terms, which give the couplings between rolling tachyon and closed string massive modes. One might expect that the couplings would converge to zero or some finite values in the far future. We will see, however, that they become exponentially strong as the system evolves. In contrast to the ordinary case where massive modes decouple from the low energy physics, massive modes could play a significant role in our case, as the couplings soon become very large. Therefore, the analysis of the rolling tachyon using low energy effective theory such as supergravity [^3] is not good enough to obtain a correct physical picture. This result also suggests that the tachyon matter could decay through the massive closed string states. So, it may affect the cosmological scenarios that involve the rolling tachyon.
This paper is organized as follows. In section \[review\], we review the rolling tachyon boundary state constructed in [@Sen:2002nu; @Sen:2002in]. In section \[massive\], we explicitly compute the couplings to some of the closed string massive modes and show that they become large exponentially at late times. We also make an argument that the couplings to the massive closed string states blow up quite generally. In section \[estimate\], we consider the effect of the back reaction of the closed string fields and an estimate is given on the time scale in which the boundary state description is valid. Section \[s-brane\] is devoted to exploring the exceptional cases in which all the couplings remain finite. We argue that the S-brane (spacelike D-brane) boundary state constructed in [@Gutperle:2002ai] can be obtained as a limit of the rolling tachyon boundary state. In section \[discussions\], we summarize our results and conclude with discussions.
Review of the Rolling Tachyon Boundary State {#review}
============================================
In this section, we review the boundary state description of the rolling tachyon constructed in [@Sen:2002nu; @Sen:2002in]. Throughout the paper, we use the convention $\alpha'=1$.
Let us first consider a D25-brane in bosonic string theory. The configuration of the open string tachyon field on the D25-brane is described in the boundary CFT by turning on a boundary interaction. In [@Sen:2002nu], Sen has considered the tachyon field rolling up and down the potential as $$\begin{aligned}
T(x^0)\propto\cosh(x^0),\end{aligned}$$ for which the boundary interaction is of the form $$\tilde{\lambda}\int dt\,\cosh X^0(t),
\label{bdry}$$ where $t$ is the parameter which parameterizes the boundary of the world-sheet. The advantage of considering this configuration is that the Wick rotated theory, which is described by the world-sheet action $$S=\frac{1}{2\pi}\int d^2 z \partial X \bar{\partial} X
+\tilde{\lambda}\int dt\, \cos X(t),
\label{wick}$$ is a solvable boundary conformal field theory [@Callan:1994ub; @Polchinski:my; @Recknagel]. Therefore, we can use some of the exact results obtained in the Wick rotated theory in the analysis of the rolling tachyon by performing inverse Wick rotation.
The boundary state for the D25-brane with the boundary interaction (\[bdry\]) takes the form $$|B\rangle=|B\rangle_{X^0} \otimes |B\rangle_{\vec{X}} \otimes |B\rangle_{bc},$$ where $|B\rangle_{\vec{X}}$ and $ |B\rangle_{bc}$ are the usual boundary states for flat D25-branes: $$\begin{aligned}
|B\rangle_{\vec{X}} &\propto&\exp\left(-\sum_{n=1}^\infty
\frac{1}{n}\alpha_{-n}^i \bar{\alpha}_{-n}^i\right)|0\rangle,
~~~~~(i=1,\dots,25)\\
|B\rangle_{bc}&\propto&\exp\left(-\sum_{n=1}^{\infty}(\bar{b}_{-n}c_{-n}
+b_{-n}\bar{c}_{-n})\right)(c_0+\bar{c_0})c_1\bar{c}_1|0\rangle.\end{aligned}$$
$|B\rangle_{X^0}$ is the part of the boundary state that describes the dynamics of the rolling tachyon. The corresponding boundary state in the Wick rotated theory (\[wick\]) has been constructed in [@Callan:1994ub; @Polchinski:my; @Recknagel]. Since the boundary interaction in (\[wick\]) introduces integer momenta, one can restrict oneself to the subspace spanned by the states carrying integer momenta. This allows one to work at the self-dual radius $R=1$. The boundary state in the compactified theory is given by acting an $SU(2)$ rotation on the unperturbed boundary state. The boundary state in the uncompactified theory is then obtained by projecting onto the unwinding states [@Callan:1994ub; @Recknagel]. The result is $$|B\rangle_{X} =
\sum_{j=0,\frac{1}{2},1,\cdots}\sum_{m=-j}^{j}D^j_{m,-m}(R)
|j;m,m\rangle\rangle.
\label{bdrystate}$$ Here $R$ is the $SU(2)$ rotation matrix $$\label{R matrix}
R= \left(
\begin{array}{ll}
\cos(\pi\tilde{\lambda}) & i\sin(\pi\tilde{\lambda})\\
i\sin(\pi\tilde{\lambda}) &\cos(\pi\tilde{\lambda})
\end{array}
\right),$$ and $D^j_{m,-m}(R)$ is a corresponding spin $j$ representation matrix element. $|j;m,m\rangle\rangle$ is the Virasoro Ishibashi state built over the primary state $|j;m,m\rangle$ (see section \[boso\] for more details).
In [@Sen:2002nu], the oscillator-free part and the part proportional to $\alpha_{-1}\bar{\alpha}_{-1}|0\rangle$ have been explicitly computed: $$\begin{aligned}
|B\rangle_X&=&
\left[ 1+2\sum_{n=1}^{\infty}(-\sin(\tilde{\lambda}\pi))^n \cos(nX(0))
\right]|0\rangle\nonumber\\
&&-\alpha_{-1}\bar{\alpha}_{-1} \left[ \cos(2\pi\tilde{\lambda})
-2\sum_{n=1}^{\infty}(-\sin(\tilde{\lambda}\pi))^n \cos(n X(0))
\right]|0\rangle +\cdots.
\label{SenBX}\end{aligned}$$ Performing the inverse Wick rotation $X\rightarrow -iX^0$, we obtain [@Sen:2002nu] $$\begin{aligned}
|B\rangle_{X^0}&=& f(X^0(0))|0\rangle +
\alpha^0_{-1}\bar{\alpha}^0_{-1}\, g(X^0(0))|0\rangle +\cdots,
\label{SenBX0}\end{aligned}$$ where $$\begin{aligned}
f(x^0)&=&\frac{1}{1+e^{x^0}\sin(\tilde\lambda\pi)}
+\frac{1}{1+e^{-x^0}\sin(\tilde\lambda\pi)}-1,
\label{f(x)}\\
g(x^0)&=&\cos(2\tilde\lambda\pi)+1-f(x^0).\end{aligned}$$
The boundary state describing the rolling of the tachyon on unstable D-brane systems in superstring theory can also be obtained similarly. We consider a pair of D9 and ${\rm \bar{D}9}$ branes in type IIB theory or a single non-BPS D9-brane in type IIA theory. The D9-${\rm \bar{D}9}$ brane pair is described by the boundary CFT with $2\times 2$ Chan-Paton factors and with an appropriate GSO projection ($(-1)^F =1$ for DD and ${\rm
\bar{D}\bar{D}}$ strings, $(-1)^F =-1$ for ${\rm D\bar{D}}$ and ${\rm \bar{D}D}$ strings). A single non-BPS D9-brane is described as the orbifold of the D9-${\rm \bar{D}9}$ system by $(-1)^{F_L}$, where $F_L$ is the space-time fermion number from left-movers (see [@Sen:1999mg] for a review).
Following [@Sen:2002in], we turn on a tachyon field of the form $$\begin{aligned}
T(x^0)\propto \cosh(x^0/\sqrt{2}),\end{aligned}$$ for which we should assign a Chan-Paton factor $\sigma_1$ [@Sen:1999mg]. This is represented as the perturbation of the world-sheet action by the boundary interaction $$\tilde{\lambda} \int dt\, \psi^0(t) \sinh\left(
\frac{X^0 (t)}{\sqrt{2}}\right)\otimes\sigma_1,$$ i.e., the integral of the corresponding zero-picture vertex operator in the $\sigma_1$ sector.
The boundary state takes the form $$|B\rangle=|B,+\rangle-|B,-\rangle,$$ where $$|B,\epsilon\rangle\propto|B,\epsilon\rangle_{X^0,\psi^0}
\otimes
|B,\epsilon\rangle_{\vec{X},\vec{\psi}}
\otimes
|B,\epsilon\rangle_{ghost},~~~~(\epsilon=\pm).$$ Again, the spatial and the ghost parts take the usual expressions: $$\begin{aligned}
|B,\epsilon\rangle_{\vec{X},\vec{\psi}}&\propto&
\exp\left(
-\sum_{n=1}^{\infty}\frac{1}{n}\alpha^{j}_{-n}\bar{\alpha}^{j}_{-n}
\right)
\exp\left(
-i\epsilon\sum_{n=1}^{\infty}\psi^{j}_{-n-1/2}\bar{\psi}^{j}_{-n-1/2}
\right)
|0\rangle,\\
|B,\epsilon\rangle_{ghost}&\propto&\exp\left(-\sum_{n=1}^{\infty}(\bar{b}_{-n}c_{-n}
+b_{-n}\bar{c}_{-n})\right)\nonumber\\
&&
\times
\exp\left(
-i\epsilon\sum_{n=1}^{\infty}(\bar{\beta}_{-n-1/2}\gamma_{-n-1/2}
-\beta_{-n-1/2}\bar\gamma_{-n-1/2})
\right)\nonumber\\
&&\times
(c_0+\bar{c_0})c_1\bar{c}_1
e^{-\phi(0)}e^{-\bar{\phi}(0)}|0\rangle.\end{aligned}$$
We are interested in the temporal part $|B\rangle_{X^0,\psi^0}$. This part of the boundary state is obtained in a similar way to the bosonic case above. Namely, we first construct the boundary state in the Wick rotated theory, which is obtained by the analytic continuation $X^0\rightarrow iX$, $\psi^0\rightarrow
i\psi$ and $\bar\psi^0\rightarrow i\bar\psi$, and inverse Wick rotate the result.
In the Wick rotated theory, the boundary interaction is $$-\tilde{\lambda}\int dt\,
\psi(t)\sin\left(\frac{X(t)}{\sqrt{2}}\right) \otimes\sigma_1.$$ Since this boundary interaction is invariant under $$X\rightarrow X+2\pi/\sqrt{2},~~~ \psi\rightarrow -\psi,$$ and $\psi$ appears only as byliners in the boundary state, the boundary state is a sum of states with momenta $p=\sqrt{2}n$ with integer $n$. Thus we can restrict ourselves to the space spanned by such states. This space is a subspace of the theory compactified at the free fermion radius $R=1/\sqrt{2}$. Thus we can fermionize $X$ through $$\begin{aligned}
&&X\equiv X_R(z)+X_L(\bar{z}),\\
&&e^{i\sqrt{2}X_R(z)}=\frac{1}{\sqrt{2}} (\xi+i\eta),~~~
e^{i\sqrt{2}X_L(\bar{z})}=\frac{1}{\sqrt{2}} (\bar{\xi}+i\bar{\eta})\end{aligned}$$ This shows that the compactified theory admits the $SO(3)$ (or $SU(2)$) current algebra. The super-Virasoro primaries can be classified according to the $SU(2)$ quantum numbers.
Such a boundary state takes the form $$|B,\epsilon\rangle_{X,\psi}=
\sum_{j=0,1,\cdots}\sum_{n=-j}^j D^j_{n,-n}(R)|j;n,n;\epsilon\rangle\rangle,$$ where $R$ is the same matrix as above, but $n$ takes integer values here and $|j;n,n;\epsilon\rangle\rangle$ is the super-Virasoro Ishibashi state built over the super-Virasoro primary $|j,n,n\rangle$. Explicit examples will be given in section \[superstring boundary\]. In [@Sen:2002in], Sen has computed the parts of the boundary state relevant to the coupling to massless modes of closed string: $$\begin{aligned}
\label{BXpsi}
&& |B,\epsilon\rangle_{X,\psi}\nonumber\\
&=&
\left(1+2\sum_{n=1}^{\infty}(-1)^n\sin^{2n}(\pi\tilde{\lambda})
\cos(\sqrt{2}n X(0))\right)|0\rangle -i\epsilon
\psi_{-1/2}\bar{\psi}_{-1/2}
\nonumber\\
&& \times\left(\cos(2\pi\tilde{\lambda})
-2\sum_{n=1}^{\infty}(-1)^n\sin^{2n}(\pi\tilde{\lambda})
\cos(\sqrt{2}n X(0))\right)|0\rangle+\cdots.\end{aligned}$$ The inverse Wick rotation leads to $$\label{BX0psi}
|B,\epsilon\rangle_{X^0,\psi^0}= \left( \hat f(X^0(0))+ i \epsilon
\psi_{-1/2}^0\bar{\psi}_{-1/2}^0 \hat g(X^0(0))\right)|0\rangle,$$ where $$\begin{aligned}
\hat f(x^0)&=&\frac{1}{1+\sin^2(\pi\tilde{\lambda} )e^{\sqrt{2}x^0}}
+\frac{1}{1+\sin^2(\pi\tilde\lambda) e^{-\sqrt{2}x^0}}-1,\\
\hat g(x^0)&=&\cos(2\pi\tilde{\lambda})+1-\hat f(x^0).\end{aligned}$$
Computation of the Massive Mode Coupling {#massive}
========================================
In this section, we compute several higher level states in the rolling tachyon boundary state constructed in [@Sen:2002nu; @Sen:2002in]. We will find that the couplings of the tachyon to the massive modes exponentially diverge as $x^0 {\rightarrow}\infty$.
Bosonic String {#boso}
--------------
We write down the Virasoro Ishibashi states which are relevant in the calculation of the boundary state up to level (2,2) (see appendix \[discrete\] for details). $$\begin{aligned}
|0;0,0\rangle\rangle&=&
\left(
1+\frac{1}{2}\alpha_{-1}^2\bar{\alpha}_{-1}^2+\cdots
\right)|0\rangle,
\label{Ishi1} \\
\left.\left|\frac{1}{2};\pm\frac{1}{2},\pm\frac{1}{2}
\right\rangle\!\right\rangle &=&
\left(1+\alpha_{-1}\bar{\alpha}_{-1}+
\frac{1}{6}
(\alpha_{-1}^2\pm\sqrt{2}\alpha_{-2})
(\bar{\alpha}_{-1}^2\pm\sqrt{2}\bar{\alpha}_{-2})+
\cdots\right)\nonumber\\
&&\times e^{\pm i X(0)}|0\rangle,\\
|1;0,0\rangle\rangle&=&
\left(\alpha_{-1}\bar{\alpha}_{-1}+
\frac{1}{2}\alpha_{-2}\bar{\alpha}_{-2}+\cdots\right)
|0\rangle,\\
|j;\pm j,\pm j\rangle\rangle&=&\left(
1+\alpha_{-1}\bar{\alpha}_{-1}+\frac{1}{2}\alpha_{-1}^2\bar{\alpha}_{-1}^2
+\frac{1}{2}\alpha_{-2}\bar{\alpha}_{-2}+\cdots
\right)\nonumber\\
&&\times e^{\pm 2i j X(0)}|0\rangle\ (j\ge 1),
\label{Ishi2}\\
\left.\left|\frac{3}{2};\pm\frac{1}{2},\pm\frac{1}{2}
\right\rangle\!\right\rangle&=&\left(\frac{1}{6}
(\alpha_{-2}\mp\sqrt{2}\alpha_{-1}^2)
(\bar{\alpha}_{-2}\mp\sqrt{2}\bar{\alpha}_{-1}^2)+\cdots\right)
e^{\pm iX(0)}|0\rangle.
\label{Ishi3}\end{aligned}$$ Now, we are ready to compute the boundary state (\[bdrystate\]). Relevant matrix elements $D^j_{m,-m}(R)$ of the $SU(2)$ rotation $R$ are given in Appendix \[matrix elements\]. Note that we have not determined possible phase factors which could appear in the above expressions (\[Ishi1\])$-$(\[Ishi3\]). In [@Sen:2002nu], the phase factors for the Ishibashi states (\[Ishi1\])$-$(\[Ishi2\]) have been determined by demanding that the boundary state $|B_X\rangle$ represents an array of D-branes localized at $X=(2n+1)\pi$ when $\tilde{\lambda}=1/2$ [@Callan:1994ub; @Recknagel; @Sen:1999]. They can be read off from (\[SenBX\]). The phase factor for (\[Ishi3\]) can be obtained similarly and it turns out to be $-i$.
Collecting these together, we obtain $$\begin{aligned}
&&|B\rangle_X \nonumber\\
&=&|0;0,0\rangle\rangle+\sum_{n=1}^\infty
(-1)^n\sin^n(\pi\tilde{\lambda})\left(
{\left.\left|\, {n\over2};{n\over2},{n\over2}\,\right\rangle\!\right\rangle} +
{\left.\left|\, {n\over2};-{n\over2},-{n\over2}\,\right\rangle\!\right\rangle}\right)\nonumber\\
&&-\cos(2\pi{\tilde{\lambda}}){\left.\left|\, 1;0,0\,\right\rangle\right\rangle}\nonumber\\
&&+\sin(\pi{\tilde{\lambda}})(3\cos^2(\pi{\tilde{\lambda}})-1) \left(
{\left.\left|\, {3\over2};{1\over2},{1\over2}\,\right\rangle\!\right\rangle} +
{\left.\left|\, {3\over2};-{1\over2},-{1\over2}\,\right\rangle\!\right\rangle} \right)+\cdots.\end{aligned}$$
After the inverse Wick rotation, we find $$\begin{aligned}
|B\rangle_{X^0} &=&
|B_{0}\rangle+
|B_{-1;-1}\rangle\nonumber\\
&&+|B_{-2;-2}\rangle+|B_{(-1)^2;(-1)^2}\rangle
+{\left|\, B_{(-1)^2;-2}\,\right\rangle}+{\left|\, B_{-2;(-1)^2}\,\right\rangle}+\cdots.\end{aligned}$$ Here $|B_{0}\rangle$ and $|B_{-1,-1}\rangle$ are level $(0,0)$ and $(-1,-1)$ states given in (\[SenBX0\]).
We further find $$\begin{aligned}
&& |B_{-2;-2}\rangle\nonumber\\
&=&-\frac{1}{2}\alpha^0_{-2}\bar{\alpha}^0_{-2}\int dx^0
\left[-(1+\cos(2\pi\tilde{\lambda}))
+2\sin(\pi\tilde{\lambda})\cos^2(\pi\tilde{\lambda})\cosh(x^0)
+f(x^0)
\right]|x^0\rangle,\nonumber
\\
\label{22}\end{aligned}$$ $$\begin{aligned}
|B_{(-1)^2;(-1)^2}\rangle
&=&\frac{1}{2}(\alpha^0_{-1})^2(\bar{\alpha}^0_{-1})^2\int dx^0
\left[
4\sin(\pi\tilde{\lambda}) \cos^2(\pi \tilde{\lambda}) \cosh(x^0)
+f(x^0)
\right]|x^0\rangle,\nonumber\\
\label{11}\end{aligned}$$ and $$\begin{aligned}
&&{\left|\, B_{(-1)^2;-2}\,\right\rangle}+{\left|\, B_{-2;(-1)^2}\,\right\rangle}\nonumber\\
&=&-i\sqrt{2}((\alpha_{-1}^0)^2\bar\alpha_{-2}^0
+\alpha_{-2}^0(\bar\alpha_{-1}^0)^2)
\sin(\pi\tilde\lambda)\cos^2(\pi\tilde\lambda)
\int dx^0\sinh(x^0){\left|\, x^0\,\right\rangle}.
\label{12}\end{aligned}$$
All these states contain terms proportional to either $\cosh(x^0)$ or $\sinh(x^0)$ which exponentially blow up when $|x^0|$ becomes large. This result shows that the coupling between tachyon matter and massive closed string modes will become strong at late times.
We will not explicitly compute the states of level higher than 2, but we generally expect this kind of behavior in the higher level states. Note that the primary state ${\left|\, j;m,m\,\right\rangle}$ is of the form $$\begin{aligned}
{\left|\, j;m,m\,\right\rangle}={\cal O}_{j,m}\,e^{2imX(0)}{\left|\, 0\,\right\rangle},\end{aligned}$$ where ${\cal O}_{j,m}$ is an operator of level $(j^2-m^2,j^2-m^2)$. When we are interested in a level $(k,k)$ state in the boundary state, we should take into account the Ishibashi states ${\left|\, j;m,m\,\right\rangle}\rangle$ with $j^2-m^2\le k$. Then, apart from the case $m=\pm j$, there are only a finite number of choices for $j$ and $m$ satisfying $j^2-m^2\le k$. The contribution from the infinite sum $\sum_j D_{\pm j,\mp
j}^j{\left|\, j;\pm j,\pm j\,\right\rangle}\rangle$ managed to sum up to a harmless function like $f(x^0)$ in (\[f(x)\]). But, unless there is an accidental cancellation among the other terms, the inverse Wick rotation of the contribution from momentum $2m$ states will blow up as $e^{2m x^0}$.
Superstring {#superstring boundary}
-----------
We again begin by listing the super-Virasoro Ishibashi states for low-level primaries: $$\begin{aligned}
|0;0,0;\epsilon\rangle\rangle&=&
(1+i\epsilon\alpha_{-1}\psi_{-1/2}
\bar{\alpha}_{-1}\bar{\psi}_{-1/2}
+\cdots)|0\rangle ,\\
|1;\pm 1,\pm 1;\epsilon\rangle\rangle&=&
[1+i\epsilon\psi_{-1/2}\bar{\psi}_{-1/2}
+\alpha_{-1}\bar{\alpha}_{-1}\nonumber\\
&&+\frac{1}{2}i\epsilon(\psi_{-3/2}\pm\alpha_{-1}\psi_{-1/2})
(\bar{\psi}_{-3/2}\pm\bar{\alpha}_{-1}\bar{\psi}_{-1/2})+\cdots]\nonumber\\
&&\times e^{\pm i\sqrt{2}X(0)}|0\rangle ,\\
|1;0,0;\epsilon\rangle\rangle&=&(i\epsilon\psi_{-1/2}
\bar{\psi}_{-1/2}
+\alpha_{-1}\bar{\alpha}_{-1}+
i\epsilon\psi_{-3/2}\bar{\psi}_{-3/2}+
\cdots)|0\rangle,\\
|j;\pm j,\pm j;\epsilon\rangle\rangle &=&
(1+i\epsilon \psi_{-1/2}\bar{\psi}_{-1/2}
+\alpha_{-1}\bar{\alpha}_{-1}\nonumber\\
&&+i\epsilon\psi_{-3/2}\bar{\psi}_{-3/2}
+i\epsilon\psi_{-1/2}\alpha_{-1}
\bar{\psi}_{-1/2}\bar{\alpha}_{-1}
+\cdots)
e^{\pm \sqrt{2}i j X(0)}|0\rangle, \nonumber\\
&&(j=2,3,4,\cdots),\\
{\left.\left|\, 2;\pm 1,\pm 1;\epsilon\,\right\rangle\right\rangle}&=&\frac{1}{2}
(\psi_{-3/2}\mp\alpha_{-1}\psi_{-1/2})
(\bar{\psi}_{-3/2}\mp\bar{\alpha}_{-1}
\bar{\psi}_{-1/2}) e^{\pm i\sqrt{2}X(0)}|0\rangle+\cdots.
\nonumber\\\end{aligned}$$ The coefficient of $|2;\pm 1,\pm 1;\epsilon\rangle\rangle$ is $D^2_{\pm 1,
\mp 1}(R)$ times a phase factor which is determined to be $-i\epsilon$ as in the bosonic case. Other coefficients can be read off from (\[BXpsi\]). Therefore the boundary state is $$\begin{aligned}
&& |B,\epsilon\rangle_{X,\psi} \nonumber\\
&=&{\left.\left|\, 0;0,0;\epsilon\,\right\rangle\right\rangle}+\sum_{n=1}^\infty (-1)^n
\sin^{2n}(\pi{\tilde{\lambda}})\left(
{\left.\left|\, n;n,n;\epsilon\,\right\rangle\right\rangle}+{\left.\left|\, n;-n,-n;\epsilon\,\right\rangle\right\rangle} \right)\nonumber\\
&&-\cos(2\pi{\tilde{\lambda}}){\left.\left|\, 1;0,0;\epsilon\,\right\rangle\right\rangle}\nonumber\\
&&+i\epsilon\sin^2(\pi{\tilde{\lambda}})
\cos(2\pi{\tilde{\lambda}})({\left.\left|\, 2;1,1;\epsilon\,\right\rangle\right\rangle}+{\left.\left|\, 2;-1,-1;\epsilon\,\right\rangle\right\rangle})+\cdots.\end{aligned}$$ After analytic continuation, we get $$\begin{aligned}
|B\rangle_{X^0,\psi^0}&=&|B_{0;0}\rangle+|B_{-1/2;-1/2}\rangle
+|B_{-1;-1}\rangle+|B_{-3/2;-3/2}\rangle \\
&& +|B_{-1,-1/2;-1,-1/2}\rangle +|B_{-3/2;-1,-1/2}\rangle
+|B_{-1,-1/2;-3/2}\rangle + \cdots,\end{aligned}$$ where the first two terms are given in (\[BX0psi\]) and $$\begin{aligned}
|B_{-1;-1}\rangle&=& -\alpha^0_{-1}\bar{\alpha^0_{-1}}
(-\cos(2\pi\tilde{\lambda})-1+\hat f(X^0(0)))|0\rangle, \\
|B_{-3/2;-3/2}\rangle &=&
i\epsilon\psi_{-3/2}^0\bar{\psi}_{-3/2}^0[\, 1+\cos(2\pi{\tilde{\lambda}})
-\hat f(X^0(0))\nonumber \\
&&-\sin^2(\pi{\tilde{\lambda}}) (\cos (2\pi{\tilde{\lambda}})+1) \cosh(\sqrt{2}X^0(0))]
{\left|\, 0\,\right\rangle},\\
|B_{-1,-1/2;-1,-1/2}\rangle &=&
i\epsilon\alpha_{-1}^0\psi_{-1/2}^0 \bar{\alpha}_{-1}^0
\bar{\psi}_{-1/2}^0 [\,\hat f(X^0(0))\nonumber\\
&&+ \sin^2 (\pi{\tilde{\lambda}})(1+\cos(2\pi{\tilde{\lambda}}))\cosh(\sqrt{2}X^0(0))
]{\left|\, 0\,\right\rangle},\end{aligned}$$ and $$\begin{aligned}
&&{\left|\, B_{-3/2;-1/2,-1}\,\right\rangle}+{\left|\, B_{-1/2,-1;-3/2}\,\right\rangle}\nonumber\\
&=&
\epsilon\sin^2(\pi{\tilde{\lambda}})(1+\cos(2\pi{\tilde{\lambda}}))\nonumber\\
&&\times\left( \psi_{-3/2}^0\bar{\alpha}_{-1}^0\bar{\psi}_{-1/2}^0
+\alpha_{-1}^0\psi_{-1/2}^0\bar{\psi}_{-3/2}^0
\right)\sinh(\sqrt{2} X^0(0)){\left|\, 0\,\right\rangle}.\end{aligned}$$ We see that as in the bosonic case, the boundary state contains exponentially divergent couplings to massive modes.
On the Back Reaction of Closed Strings {#estimate}
======================================
So far, we have ignored the dynamics of the closed string and focused on the classical equation of motion for the open string. This analysis is only justified in the weak coupling limit. What happens when we slightly turn on the string coupling $g_s$? As we have shown in the previous section, even if the string coupling is small, the system with rolling tachyon will become strongly coupled to closed string massive modes at late times and we expect a large back reaction of the closed string fields. Let us make this point more explicit in this section.
In string field theory, the tree-level action in terms of the canonically normalized string fields looks like [@Zwiebach:1997fe] $$\label{eq:action}
S\sim\tilde{\psi} Q_o \tilde{\psi} +g_s^{1/2} \tilde{\psi}^3+
\tilde{\phi} Q_c \tilde{\phi} + g_s \tilde{\phi}^3+
\tilde{\phi}B(g_s^{1/2}\tilde{\psi})+\cdots.$$ Here $\tilde{\psi}$ is an open string field and $\tilde{\phi}$ is a closed string field. $\tilde\phi B(g_s^{1/2}\tilde\psi)$ symbolically denotes the coupling among one closed string field and open string fields. The power of $g_s$ in each term is determined from the Euler characteristic of the disk or sphere, and the powers of open and closed string fields.
The boundary state considered in the previous sections corresponds to $B(g_s^{1/2}\tilde\psi_0)$ and plays a role of a source for the closed string. Here, $\tilde\psi_0$ denotes a solution of the equations of motion for the open string obtained by turning off the closed strings. To see this structure more clearly, it is convenient to change the normalization for the open string and absorb the coupling constant as $$\label{eq:normalization}
\tilde{\psi}=g_s^{-1/2}\psi,$$ Then, the action becomes $$\label{eq:action2}
S=g_s^{-1}(\psi Q_o \psi + \psi^3)+
\tilde{\phi} Q_c \tilde{\phi} + g_s \tilde{\phi}^3+
\tilde{\phi}B(\psi)+\cdots.$$ The equation of motion for the closed string field is $$\label{closedEOM}
Q_c\tilde\phi+g_s\tilde\phi^2+ B(\psi) +\cdots=0,$$ which imply the BRST invariance of the boundary state $Q_c B=0$ in the weak coupling limit. On the other hand, the equation of motion for the open string field is $$\label{eq:eom}
Q_o\psi+\psi^2+g_s \tilde\phi B'(\psi)+\cdots=0,$$ which shows that the closed strings are decoupled from the open strings in the weak coupling limit.
However, once we turn on the non-zero string coupling $g_s$, the coupling of the rolling tachyon boundary state to the closed string massive modes exponentially grows when $x^0 {\rightarrow}\infty$, as we have shown in the previous section, and hence it is not consistent to ignore the dynamics of closed string at late times, even if $g_s$ is small. In fact, since there are exponentially growing source terms for the massive closed string modes in the equations of motion (\[closedEOM\]), the fluctuation will also blow up exponentially. Then, according to (\[eq:eom\]), this back reaction will alter the equations of motion for the open string. Therefore, even if the effect of merely the lowest diverging modes are included, the boundary state description of the rolling tachyon given in the previous sections is only reliable when $$\label{eq:validity}
g_s e^{|t|/l_s} \ll 1.$$ Since higher levels have couplings which grow as $e^{m|t|}$ with arbitrary $m$, the boundary state description is expected to be valid in the time-scale shorter than that given by this inequality.
When the time $t$ is large enough, we need to take the closed strings into consideration, and the whole picture will drastically change. For example, from the fact that the energy momentum tensor does not oscillate, people anticipated that the tachyon matter will not decay into closed string modes. However, since the coupling to the massive modes will exponentially grow, we can expect that the tachyon matter will start to emit massive closed string modes and eventually decay into massless particles. Moreover, there is a large back reaction of closed strings, especially from the massive fields. This is in contrast to the usual static D-brane cases, in which supergravity analysis essentially gives the precise low energy description. It would be interesting to analyze the system further using the equations of motion (\[closedEOM\]) and (\[eq:eom\]).
Relation to the S-brane Boundary States {#s-brane}
=======================================
In section \[massive\], we explicitly computed the coefficients of some higher level states in the rolling tachyon boundary state and found that they become large at late times. However, there are special values of $\tilde\lambda$ at which the couplings remain finite. In the bosonic string case, since $D^j_{m,-m}(R)$ is invariant under ${\tilde{\lambda}}{\rightarrow}{\tilde{\lambda}}+2$ and ${\tilde{\lambda}}{\rightarrow}1-{\tilde{\lambda}}$, the whole system has the same symmetries. Thus we can restrict ${\tilde{\lambda}}$ to $-1/2\le {\tilde{\lambda}}\le 1/2$. If we take $\tilde\lambda$ to be $\pm 1/2$ or $0$, the coefficients of $\cosh(x^0)$ and $\sinh(x^0)$ in (\[22\])$-$(\[12\]) vanish. This fact can be easily understood from the known behavior of the boundary state (\[bdrystate\]) in the Wick rotated theory at those values of ${\tilde{\lambda}}$. When $\tilde\lambda$ is zero, the boundary state becomes a static boundary state with the Neumann boundary condition, which represents a static D25-brane with a tachyon sitting on top of the maximum of its potential. For $\tilde{\lambda}=1/2$ or $\tilde{\lambda}=-1/2$, the boundary state (\[bdrystate\]) represents an array of D-branes localized at $X=(2n+1)\pi$ or $X=2n\pi$ with $n\in{\bf Z}$, respectively, in the Wick rotated theory [@Callan:1994ub; @Recknagel; @Sen:1999]. Therefore, we obtain $$\begin{aligned}
\label{lambdapm1/2}
{\left|\, B\,\right\rangle}_{X;{\tilde{\lambda}}=\pm 1/2}=\exp\left(
\sum_{n=1}^\infty\frac{1}{n}\alpha_{-n}\bar\alpha_{-n}
\right)\sum_{n\in{\bf Z}}(\mp)^n e^{in X(0)}{\left|\, 0\,\right\rangle}.\end{aligned}$$ The zero mode part can be formally written as $$\begin{aligned}
\label{formalsum}
\sum_{n \in {\bf Z}}(\mp)^n e^{in X(0)}|0\rangle
&=&\left(
\frac{1}{1\pm e^{iX(0)}}
+\frac{1}{1\pm e^{-iX(0)}}-1
\right)|0\rangle,\end{aligned}$$ whose inverse Wick rotation is given as $$\begin{aligned}
\left(
\frac{1}{1\pm e^{X(0)}}
+\frac{1}{1\pm e^{-X(0)}}-1
\right){\left|\, 0\,\right\rangle}
=
\lim_{\tilde\lambda{\rightarrow}\pm1/2} f(X^0(0)){\left|\, 0\,\right\rangle},\end{aligned}$$ where $f(x^0)$ is given in (\[f(x)\]). In fact, we can show that the boundary state is of the form [^4] $$\begin{aligned}
{\left|\, B\,\right\rangle}_{X^0;{\tilde{\lambda}}\sim\pm 1/2}\sim\exp\left(
\sum_{n=1}^\infty\frac{1}{n}\alpha_{-n}\bar\alpha_{-n}
\right) f(X(0)){\left|\, 0\,\right\rangle}.\end{aligned}$$ up to the terms which are cancelled at $\tilde\lambda=\pm 1/2$. As a check, we can see in (\[SenBX0\]) and (\[22\])$-$(\[12\]) that every term which does not include $f(x^0)$ vanishes in the limit $\tilde\lambda{\rightarrow}\pm 1/2$.
In the limit $\tilde\lambda{\rightarrow}1/2$, the function $f(x^0)$ also vanishes and the whole boundary state becomes zero. This implies that the system is at the closed string vacuum, which corresponds to placing the tachyon at the minimum of its potential [@Sen:2002nu]. We should be more careful in taking the limit $\tilde\lambda{\rightarrow}-1/2$, since the function $f(x^0)$ is singular at $x^0=\pm\log(-\sin(\tilde\lambda\pi))$ for $-1/2<\tilde\lambda<0$. This behavior of $f(x^0)$ with negative $\tilde\lambda$ has been interpreted in [@Sen:2002nu] that the tachyon is rolling on the wrong side of the hill where the potential is unbounded from below. Though it is not quite clear whether it is physically meaningful to take the limit $\tilde\lambda{\rightarrow}-1/2$, here we would like to point out a suggestive relation between the rolling tachyon boundary state in this limit and the S-brane boundary state constructed in [@Gutperle:2002ai].
Because of the limit ${\tilde{\lambda}}{\rightarrow}-1/2$, the summation in (\[formalsum\]) should be understood as being regularized as $$\sum_{n\in{\bf Z}}e^{inx} =\lim_{\epsilon \rightarrow +0}
\sum_{n\in{\bf Z}} e^{-\epsilon|n|}e^{inx} =\lim_{\epsilon
\rightarrow +0} \left(\frac{1}{1-e^{-\epsilon+ix}}+
\frac{1}{1-e^{-\epsilon-ix}}-1\right).$$ After the inverse Wick rotation $x$ is replaced by $-i
e^{i\delta}x^0$: $$\lim_{\epsilon,\delta \rightarrow +0}
\left(\frac{1}{1-e^{-\epsilon+e^{i\delta}x^0}}+
\frac{1}{1-e^{-\epsilon-e^{i\delta}x^0}}-1\right).$$ The limit vanishes except at $x^0=0$. Integrating this function with the use of residue theorem, one finds that this function equals $2\pi i\delta(x^0)$. The boundary state now becomes $$\begin{aligned}
{\left|\, B\,\right\rangle}_{X^0;\tilde\lambda{\rightarrow}-1/2}=2\pi i \exp\left(
-\sum_{n=1}^\infty\frac{1}{n}\alpha^0_{-n}\bar\alpha^0_{-n}
\right)\delta(X^0(0)){\left|\, 0\,\right\rangle}.\end{aligned}$$ This is nothing but the S-brane boundary state, which represents the Dirichlet boundary condition in the time direction.
Similarly, in the superstring case, the system is invariant under ${\tilde{\lambda}}{\rightarrow}{\tilde{\lambda}}+1$ and ${\tilde{\lambda}}{\rightarrow}-{\tilde{\lambda}}$, so we can restrict ${\tilde{\lambda}}$ to $0\le{\tilde{\lambda}}\le1/2$. The system does not evolve at $\tilde{\lambda}=0, 1/2$. When $\tilde{\lambda}$ is zero, the system is the original D9-brane system. When $\tilde{\lambda}= 1/2$, the Wick rotated NS-NS boundary state considered here represents an array of D-branes placed at $x=\frac{2\pi}{\sqrt{2}}(n+\frac{1}{2})$: $$|B\rangle_{X,\psi;\tilde{\lambda}=1/2}
=\exp\left(
\sum_{n=1}^{\infty}\frac{1}{n}\alpha_{-n}\bar{\alpha}_{-n}
+i\epsilon\sum_{r=1/2}^{\infty}\psi_{-r}\bar{\psi}_{-r}
\right)\sum_{n \in {\bf Z}}(-1)^n
e^{i\sqrt{2}n X(0)}
|0\rangle.$$ Now if we shift $X\rightarrow X+\pi/\sqrt{2}$ and then inverse Wick rotate, one obtains as in the bosonic case $$\sqrt{2}\pi i|x^0=0\rangle$$ for the zero mode part. This gives the boundary state for a spacelike brane. This suggests that the spacelike brane corresponds to the boundary interaction $$\label{s-brane action}
S=
\frac{i}{2}\int dt\, \psi^0(t)\cosh\left(\frac{X^0(t)}{\sqrt{2}}\right)
\otimes\sigma_1.$$ Unfortunately, comparing this with the boundary interaction before the shift in $X$, one sees that this corresponds to an imaginary value of the tachyon field, which should be real to be physical. Thus, we conclude that the S-brane boundary state in superstring can be formally thought of as a rolling of imaginary tachyon, though its physical relevance is unclear.
Conclusions and Discussions {#discussions}
===========================
In this paper, we analyzed the higher level states in the rolling tachyon boundary state for bosonic string as well as superstring theory. We explicitly calculated some of the coefficients of the higher level states which correspond to the coupling between the tachyon matter and the massive closed string modes and found that they include terms that blow up like $\cosh(x^0)$ or $\sinh(x^0)$. We also argued that the S-brane boundary state given in [@Gutperle:2002ai] can be obtained as a special limit of the rolling tachyon boundary state.
These results suggest that the massive closed string modes play an important role in string theory, when we take into account the interaction between closed strings and open strings. The tachyon matter may decay through the massive closed string modes and this effect may become important in the cosmological scenarios using the rolling tachyon. Note that this is truly a stringy effect which cannot be seen using the low energy effective theory. It would be interesting to make a systematic analysis of the open-closed mixed system to obtain a better picture about the fate of the unstable D-brane.
It is well-known that in the presence of the space-time filling D-branes, the closed string background should be shifted in order to cancel the divergence due to the massless tadpole. This argument is usually given in the static configuration, in which the tachyon is not rolling. In the static case, the boundary state represents a constant source for the closed strings. This constant source is important for the massless fields, but not for the massive fields, since it only causes a small constant shift for the massive fields. On the other hand, in the case with time evolution, the massive fields could also receive a large back reaction. Moreover, since the boundary state carries non-zero energy, it will become possible to create some particles and transfer the energy to closed string modes. In our rolling tachyon case, since the coupling between the closed string and the open string will become large and we should take into account the huge back reaction of the closed string fields beyond the time range given in (\[eq:validity\]), the perturbative analysis of the string theory may not be practical. We hope we can make these issues clear in the near future.
Acknowledgments {#acknowledgments .unnumbered}
===============
We would like to thank Sanefumi Moriyama, Paul Mukhopadhyay, Yuji Okawa, Ashoke Sen, and Barton Zwiebach for useful discussions. T. O. thanks Hiroshi Ooguri for helpful conversations. S. S. is also grateful to Shinji Mukohyama, Kazumi Okuyama, Soo-Jong Rey, and Seiji Terashima for useful discussions.
[**Note added:**]{} While preparing the paper, we received a related paper [@Mukhopadhyay:2002en] in which some of the higher level states in the rolling tachyon boundary state in bosonic string theory are also calculated.
Discrete Primaries and Ishibashi States {#discrete}
=======================================
Bosonic String {#bosonic-string}
--------------
Let us consider the theory of a single free boson $X$. The state ${\left|\, j;m,m\,\right\rangle}$ used in section \[review\] is a primary state of momentum $2m$ and conformal weight $(j^2,j^2)$, where $j=0,1/2,1,\dots$ and $m=-j,-j+1,\dots,j$. It can be factorized into right and left moving parts as ${\left|\, j;m,m\,\right\rangle}={\left|\, j,m\,\right\rangle}{\overline}{{\left|\, j,m\,\right\rangle}}$. The state ${\left|\, j,m\,\right\rangle}$ belongs to the spin $j$ representation of the $SU(2)$ current algebra defined by $$\begin{aligned}
J^\pm=\oint\frac{dz}{2\pi i}\,e^{\pm 2iX_R(z)},~~
J^3=\oint\frac{dz}{2\pi i}\,i\partial X_R(z),\end{aligned}$$ and it can be explicitly given as [@Klebanov:1991hx] $$\begin{aligned}
{\left|\, j,j\,\right\rangle}&=& e^{2ijX(0)}{\left|\, 0\,\right\rangle},\\
{\left|\, j,m\,\right\rangle}&=&N_{j,m}(J^-)^{j-m}{\left|\, j,j\,\right\rangle},\end{aligned}$$ where $N_{j,m}$ is the normalization constant. In practice, primaries are computed as the lowest null state in the Verma module of lower-weight primaries. Examples of the discrete primary states can be found in (\[Ishi1\])-(\[Ishi3\]) as the first terms.
Given a primary state $|h\rangle$, the Virasoro Ishibashi state $|h\rangle\rangle$ is defined as [@Ishibashi:1988kg] $$\label{ishibashi}
|h\rangle\rangle\equiv{\left|\, n\,\right\rangle}\otimes {\overline}{U{\left|\, n\,\right\rangle}},$$ where $\{{\left|\, n\,\right\rangle}\}$ is an arbitrary orthonormal basis of the space spanned by Virasoro descendant states of $|h\rangle$ and $U$ is an antiunitary operator such that $$\label{U 1}
U L_n U^{-1}=L_n,\ U |h\rangle=|h\rangle.$$ Such a state preserves half the conformal symmetries: $$(L_n-\bar{L}_{-n})|h\rangle\rangle=0.$$
Superstring {#superstring}
-----------
Let us next consider the superconformal theory of $(X,\psi)$ in the NS-NS sector. This theory has the symmetry of the NS-algebra with $\hat{c}=1$ and its antiholomorphic copy. The super-Virasoro primaries $|j;m,m\rangle$ in section \[review\] are given in a similar way to the bosonic case. This time, $j$ takes only integer values and the $SO(3)$ generators are $$J^\pm =\oint \frac{dz}{2\pi i}\sqrt{2}\psi(z)e^{\pm
i\sqrt{2}X_R(z)},\ J^3=\oint \frac{dz}{2\pi i} \sqrt{2}
\partial X_R(z),$$ which commute with super-Virasoro generators. Given a super-Virasoro primary $|h\rangle$, the super-Virasoro Ishibashi state $|h\rangle\rangle$ is defined as (\[ishibashi\]), where $\{ {\left|\, n\,\right\rangle}\}$ is any orthonormal basis of the space spanned by the descendants $L_{-n_1}\cdots L_{-n_p} G_{-r_1}\cdots
G_{-r_q}{\left|\, h\,\right\rangle}$. $U$ is an antiunitary operator satisfying (\[U 1\]) as well as $$\label{U 2}
U G_r U^{-1} = i\epsilon G_r (-1)^F,$$ where $F$ counts the number of $G_r$ acting on ${\left|\, h\,\right\rangle}$ [@Ishibashi:1988kg]. The super-Virasoro Ishibashi state preserves half the superconformal symmetries: $$\label{gluing2}
(G_r-i\epsilon\bar{G}_{-r}){\left|\, h\,\right\rangle}\rangle
=(L_n-\bar{L}_{-n}){\left|\, h\,\right\rangle}\rangle=0.$$
Matrix Elements for rotation $R$ {#matrix elements}
================================
Here, we list the relevant matrix elements of the $SU(2)$ rotation $R$ in (\[R matrix\]). See for example [@Recknagel] for general matrix elements. $$\begin{aligned}
D^j_{\pm j,\mp j}(R)&=&(i\sin(\pi\tilde\lambda))^{2j}~~~(j=0,1/2,1,\cdots),\\
D^1_{0,0}(R)&=&\cos(2\pi\tilde\lambda),\\
D^{3/2}_{\pm 1/2,\mp 1/2}&=&
i\sin(\pi\tilde\lambda)(3\cos^2(\pi\tilde\lambda)-1),\\
D^2_{\pm1,\mp1}(R)&=&-\sin^2(\pi{\tilde{\lambda}})\cos(2\pi{\tilde{\lambda}}) .\end{aligned}$$
[10]{}
A. Sen, “Rolling tachyon,” JHEP [**0204**]{}, 048 (2002) \[arXiv:hep-th/0203211\]. A. Sen, “Tachyon matter,” arXiv:hep-th/0203265. A. Sen, “Field theory of tachyon matter,” arXiv:hep-th/0204143. A. Sen, “Time evolution in open string theory,” arXiv:hep-th/0207105. N. Moeller and B. Zwiebach, “Dynamics with infinitely many time derivatives and rolling tachyons,” arXiv:hep-th/0207107. J. Kluson, “Time Dependent Solution in Open Bosonic String Field Theory,” arXiv:hep-th/0208028. S. Sugimoto and S. Terashima, “Tachyon matter in boundary string field theory,” JHEP [**0207**]{}, 025 (2002) \[arXiv:hep-th/0205085\]. J. A. Minahan, “Rolling the tachyon in super BSFT,” JHEP [**0207**]{}, 030 (2002) \[arXiv:hep-th/0205098\]. A. Ishida and S. Uehara, “Gauge fields on tachyon matter,” arXiv:hep-th/0206102.
G. W. Gibbons, “Cosmological evolution of the rolling tachyon,” Phys. Lett. B [**537**]{}, 1 (2002) \[arXiv:hep-th/0204008\]; S. Mukohyama, “Brane cosmology driven by the rolling tachyon,” Phys. Rev. D [**66**]{}, 024009 (2002) \[arXiv:hep-th/0204084\]; T. Padmanabhan, “Accelerated expansion of the universe driven by tachyonic matter,” Phys. Rev. D [**66**]{}, 021301 (2002) \[arXiv:hep-th/0204150\]. A. Frolov, L. Kofman and A. A. Starobinsky, “Prospects and problems of tachyon matter cosmology,” arXiv:hep-th/0204187;
D. Choudhury, D. Ghoshal, D. P. Jatkar and S. Panda, “On the cosmological relevance of the tachyon,” arXiv:hep-th/0204204; X. z. Li, J. g. Hao and D. j. Liu, “Can quintessence be the rolling tachyon?,” arXiv:hep-th/0204252; H. B. Benaoum, “Accelerated universe from modified Chaplygin gas and tachyonic fluid,” arXiv:hep-th/0205140; T. Mehen and B. Wecht, “Gauge fields and scalars in rolling tachyon backgrounds,” arXiv:hep-th/0206212; G. Shiu, S. H. Tye and I. Wasserman, “Rolling tachyon in brane world cosmology from superstring field theory,” arXiv:hep-th/0207119; G. N. Felder, L. Kofman and A. Starobinsky, “Caustics in tachyon matter and other Born-Infeld scalars,” arXiv:hep-th/0208019; S. Mukohyama, “Inhomogeneous tachyon decay, light-cone structure and D-brane network problem in tachyon cosmology,” arXiv:hep-th/0208094. G. Shiu and I. Wasserman, “Cosmological constraints on tachyon matter,” Phys. Lett. B [**541**]{}, 6 (2002) \[arXiv:hep-th/0205003\]; T. Padmanabhan and T. R. Choudhury, “Can the clustered dark matter and the smooth dark energy arise from the same scalar field?,” arXiv:hep-th/0205055. M. Fairbairn and M. H. Tytgat, “Inflation from a tachyon fluid?,” arXiv:hep-th/0204070; A. Feinstein, “Power-law inflation from the rolling tachyon,” arXiv:hep-th/0204140. M. Sami, “Implementing power law inflation with rolling tachyon on the brane,” arXiv:hep-th/0205146; M. Sami, P. Chingangbam and T. Qureshi, “Aspects of tachyonic inflation with exponential potential,” arXiv:hep-th/0205179; Y. S. Piao, R. G. Cai, X. m. Zhang and Y. Z. Zhang, “Assisted tachyonic inflation,” arXiv:hep-ph/0207143; X. z. Li, D. j. Liu and J. g. Hao, “On the tachyon inflation,” arXiv:hep-th/0207146. B. Wang, E. Abdalla and R. K. Su, “Dynamics and holographic discreteness of tachyonic inflation,” arXiv:hep-th/0208023. M. C. Bento, O. Bertolami and A. A. Sen, “Tachyonic inflation in the braneworld scenario,” arXiv:hep-th/0208124.
L. Kofman and A. Linde, “Problems with tachyon inflation,” JHEP [**0207**]{}, 004 (2002) \[arXiv:hep-th/0205121\]. J. M. Cline, H. Firouzjahi and P. Martineau, “Reheating from tachyon condensation,” arXiv:hep-th/0207156.
K. Ohta and T. Yokono, “Gravitational approach to tachyon matter,” arXiv:hep-th/0207004. A. Buchel, P. Langfelder and J. Walcher, “Does the tachyon matter?,” arXiv:hep-th/0207235. M. Gutperle and A. Strominger, “Spacelike branes,” JHEP [**0204**]{}, 018 (2002) \[arXiv:hep-th/0202210\]. C. G. Callan, I. R. Klebanov, A. W. Ludwig and J. M. Maldacena, “Exact solution of a boundary conformal field theory,” Nucl.Phys. B [**422**]{}, 417 (1994) \[arXiv:hep-th/9402113\]. J. Polchinski and L. Thorlacius, “Free Fermion Representation Of A Boundary Conformal Field Theory,” Phys. Rev. D [**50**]{}, 622 (1994) \[arXiv:hep-th/9404008\]. A. Recknagel and V. Schomerus, “Boundary deformation theory and moduli spaces of D-branes,” Nucl. Phys. B [**545**]{}, 233 (1999) \[arXiv:hep-th/9811237\]. A. Sen, “Non-BPS states and branes in string theory,” arXiv:hep-th/9904207. A. Sen, “Descent relations among bosonic D-branes,” Int. J. Mod. Phys. A [**14**]{}, 4061 (1999) \[arXiv:hep-th/9811237\]. B. Zwiebach, “Oriented open-closed string theory revisited,” Annals Phys. [**267**]{}, 193 (1998) \[arXiv:hep-th/9705241\]. P. Mukhopadhyay and A. Sen, “Decay of unstable D-branes with electric field,” arXiv:hep-th/0208142. I. R. Klebanov and A. M. Polyakov, “Interaction of discrete states in two-dimensional string theory,” Mod. Phys. Lett. A [**6**]{}, 3273 (1991) \[arXiv:hep-th/9109032\]. N. Ishibashi, “The Boundary And Crosscap States In Conformal Field Theories,” Mod. Phys. Lett. A [**4**]{}, 251 (1989).
[^1]: E-mail: [[email protected]]{}
[^2]: E-mail: [[email protected]]{}
[^3]: See [@Ohta:2002ac; @Buchel:2002tj] for the study of the tachyon matter in the supergravity approach.
[^4]: At each level, an infinite number of terms from ${\left.\left|\, j;\pm j,\pm j\,\right\rangle\right\rangle}$ sum up to $f(x^0)$, leaving a finite number of terms that vanish in the limit ${\tilde{\lambda}}{\rightarrow}\pm 1/2$.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Integrable defects in two-dimensional integrable models are purely transmitting thus topological. By fusing them to integrable boundaries new integrable boundary conditions can be generated, and, from the comparison of the two solved boundary theories, explicit solutions of defect models can be extracted. This idea is used to determine the transmission factors and defect energies of topological defects in sinh-Gordon and Lee-Yang models. The transmission factors are checked in Lagrangian perturbation theory in the sinh-Gordon case, while the defect energies are checked against defect thermodynamic Bethe ansatz equations derived to describe the ground-state energy of diagonal defect systems on a cylinder. Defect bootstrap equations are also analyzed and are closed by determining the spectrum of defect bound-states in the Lee-Yang model.'
author:
- 'Z. Bajnok and Zs. Simon'
title: Solving topological defects via fusion
---
*Theoretical Physics Research Group of the Hungarian Academy of Sciences,*\
*H-1117 Pázmány s. 1/A, Budapest, Hungary*
Introduction
============
Recently, there has been an increasing interest in integrable quantum field theories including defects or impurities. This is motivated both by the realistic physical applications in statistical and solid state physics and also by the need of theoretical understanding of this so-far unexplored field.
The community of integrable systems have not payed much attention to defect theories at the beginning due to the no-go theorem formulated by Delfino, Mussardo and Simonetti in [@DMS1; @DMS2]. The theorem, formulated originally for diagonal theories and extended later for a large class of non-diagonal ones in [@Defboot], states that a relativistically invariant theory with a non-free integrable interaction in the bulk can allow only two types of integrable defects: the purely reflecting and the purely transmitting ones. (Although some effort has been made to overcome this obstacle by giving up Lorentz invariance, see for instance [@CMRS] and references therein, in the present paper we restrict ourselves to the relativistically invariant case.)
The analysis of boundary integrable theories was initiated in [@GZ] by formulating, in an axiomatic way, the properties of the reflection matrix: unitarity, boundary crossing unitarity and boundary bootstrap equation. The boundary bootstrap framework was completed by introducing boundary Coleman-Thun mechanism [@BCT] and the bulk bootstrap equations [@FK]. Later this framework got a sound basis by developing boundary quantum field theories from first principles in [@BRF; @BBT]. The success of the boundary bootstrap approach resulted in a large class of closed bootstrap theories in which the boundary reflection factors together with the spectrum of boundary excited states were determined [@GZ; @BCT; @Valentina; @CT; @MD; @BPTT; @TGZS]. The solutions, obtained by the bootstrap method, are not connected, however, to other formulations of the model such as to the classical field theory or to the perturbed conformal field theory (both defined by a Lagrangian). To connect the different descriptions one either has to check the reflection factors perturbatively, like in [@Edpert], or solve the theories in finite volume. The boundary thermodynamic Bethe ansatz (BTBA), developed in [@LMSS], systematically sums up the finite size corrections by taking into account the scatterings and reflections. By analyzing its small volume limit the needed link between the bootstrap and perturbed conformal field theoretical descriptions can be established.
As the no-go theorem showed non-free integrable defect theories are purely transmitting. This fact kept back the researchers for some time to analyze these models until new life was put into the subject due to their explicit Lagrangian realizations [@ClDef]. Following the original idea many integrable defect theories were constructed at the classical level [@DAT; @GYZ1; @Vincent]. The basis for the quantum formulation of defect theories is provided by the folding trick [@DefBound] by which one can map any defect theory into a boundary one. As a consequence defect unitarity, defect crossing symmetry and defect bootstrap equations together with defect Coleman-Thun mechanism are derived. Despite of these results the explicitly solved relativistically invariant defect quantum field theories are quite rare, containing basically the sine-Gordon and affine Toda field theories [@KL; @sgdef; @QAT] and even in these cases the explicit relation to their Lagrangian have not been worked out yet.
One may think that purely transmitting theories are too simple and there is no point to analyze them, but we would like to argue that they carry very important information about an integrable quantum field theory, without which our knowledge cannot be complete. Purely transitivity implies the conservation of momentum from which the topological nature of the defect follows. Thus, such defects can freely be transported in space without affecting the physics of the theory. We can either move them close to each other or move them to integrable boundary conditions and, as a result, new integrable boundary conditions can be generated. These ideas were successfully applied in conformal field theories [@PZ; @GW; @Ingo], in integrable lattice models [@BP; @CMOP] and the aim of the present paper is to exploit it in solving integrable defects in the sinh-Gordon and Lee-Yang theories.
The paper is organized as follows: In section 2, on the example of the sinh-Gordon theory, we show how new boundary conditions can be obtained by fusing integrable defects to boundaries at the classical level. We focus on two cases in detail: fusing the integrable defect to Dirichlet boundary condition (DBC) the perturbed Neumann boundary condition (PNBC) can be obtained, while fusing it to the Neumann BC a new class of time-dependent integrable BCs can be generated. Section 3 recalls the quantum version of the fusion method together with the properties of transmission factors. By comparing the already known reflection factors and boundary energies of the DBC to those of the PNBC solutions for the defect transmission factor and defect energy can be extracted. The same method is then used to determine defect energies and transmission factors of the scaling Lee-Yang model. This method not only provides the explicit solutions of the sinh-Gordon defect theory but also relates its parameter to that of the Lagrangian. Since the relation obtained here is different from the suggestion of [@sgdef] we perform a perturbative analysis at one-loop level in section 4. (In the subsequent paper [@QAT] the same authors raised the possibility of the quantum renormalization of the transmission parameter, which is confirmed at one-loop level here). In section 5 the calculated defect energies are subject to another consistency check. For this we derive a TBA equation to describe the groundstate energy of a diagonal defect system on a cylinder. From a careful ultra-violet (UV) analysis defect energies are extracted and the previous results are verified. By completing the defect bootstrap program we analyze the singularity structure of the transmission factors in section 6. Since the sinh-Gordon transmission factor does not contain any singularity in the physical strip we analyze the Lee-Yang model only. For each pole of the transmission factor in the physical strip we associate either a defect boundstate or a defect Coleman-Thun diagram and calculate the excited transmission factors in the former case from the defect bootstrap equation. Finally, we conclude in section 7 and give directions for further research.
Fusion method at the Lagrangian level\[sec:Lagrangianshg\]
==========================================================
In this section we demonstrate, on the example of the sinh-Gordon (ShG) theory, how new integrable boundary conditions can be obtained by the fusion method at the Lagrangian or classical level [@Peter].
The ShG theory is defined on the whole line by the following Lagrangian$$\mathcal{L}_{\mathrm{ShG}}(\Phi)=\frac{1}{2}(\partial_{t}\Phi)^{2}-\frac{1}{2}(\partial_{x}\Phi)^{2}-\frac{m_{\mathrm{cl}}^{2}}{b^{2}}\cosh b\Phi\,\,\quad,\label{ShGbulkL}$$ It describes an integrable field theory and by restricting the theory to the half line this property can be only maintained, whenever the following boundary potential is introduced [@GZ]$$\mathcal{L}_{\mathrm{BShG}}=\Theta(-x)\mathcal{L}_{\mathrm{ShG}}(\Phi)-\delta(x)B(\Phi)\quad;\qquad B(\Phi)=M_{0}\cosh\frac{b}{2}(\Phi-\varphi_{0})\label{eq:shGBC}$$ By varying $M_{0}$ from $0$ to $\infty$ the arising boundary condition interpolates between the Neumann $\partial_{x}\Phi\vert_{x=0}=0$ and the Dirichlet $\Phi(0,t)=\varphi_{0}$ ones.
The most general integrable defect condition can be obtained by the analytical continuation of the sine-Gordon result [@sgdef]. The Lagrangian in the ShG case reads as $$\mathcal{L}_{\mathrm{DShG}}=\Theta(-x)\mathcal{L}_{\mathrm{ShG}}(\Phi_{-})-\delta(x)D(\Phi_{-},\Phi_{+})+\Theta(x)\mathcal{L}_{\mathrm{ShG}}(\Phi_{+})\label{eq:ShGdefL}$$ where the defect potential contains just one single parameter: $$\begin{aligned}
2D(\Phi_{-},\Phi_{+}) & = & \Phi_{+}\dot{\Phi}_{-}-\Phi_{-}\dot{\Phi}_{+}+M_{\mathrm{cl}}e^{\mu}\cosh\frac{b}{2}(\Phi_{+}+\Phi_{-})+M_{\mathrm{cl}}e^{-\mu}\cosh\frac{b}{2}(\Phi_{+}-\Phi_{-})\end{aligned}$$ Here $\Phi_{\mp}$ are the fields living on the left/right half-line, respectively and $M_{\mathrm{cl}}=\frac{4m_{\mathrm{cl}}}{b^{2}}$.
The fusion idea is based on the integrability of the defect: Integrability guaranties the existence of an infinite number of commuting conserved charges which results in the possibility of shifting the trajectories of particles, without changing the amplitude of any scattering process. The shifting of all the trajectories can alternatively be described by shifting the location of the defect, which then, does not alter the physics.
At the level of the Lagrangian this observation can be formulated in the following way: The spectrum of the system, which contains a defect in the origin, $x=0$, in front of a boundary, located at $x=a$, $$\mathcal{L}_{\mathrm{DBShG}}=\Theta(-x)\mathcal{L}_{\mathrm{ShG}}(\Phi_{-})-\delta(x)D(\Phi_{-},\Phi_{+})+\Theta(x)\Theta(a-x)\mathcal{L}_{\mathrm{ShG}}(\Phi_{+})-\delta(x-a)B(\Phi_{+})$$ does not actually depend on $a$. Thus we can perform the $a\to0$ limit and represent the same system as a boundary one, but with a different (dressed) boundary condition:$$\mathcal{L}_{\mathrm{DBShG}}=\Theta(-x)\mathcal{L}_{\mathrm{ShG}}(\Phi_{-})-\delta(x)B^{'}(\Phi_{-},\Phi_{+})\quad;\qquad B^{'}(\Phi_{-},\Phi_{+})=D(\Phi_{-},\Phi_{+})+B(\Phi_{+})$$ The field $\Phi_{+}$ lives on the boundary only and can be thought naively to be a boundary degree of freedom. It does not have, however, any kinetic term so it merely implements a new time-dependent integrable boundary condition. Let us specify these findings in two concrete examples that will be used later on.
If the original boundary condition is the Dirichlet one with $\Phi(a,t)=\phi_{0}$, then the arising dressed boundary condition is$$B^{'}(\Phi_{-},\Phi_{+})=D(\Phi_{-},\varphi_{0})=\frac{M_{\mathrm{cl}}e^{\mu}}{2}\cosh\frac{b}{2}(\Phi_{-}+\phi_{0})+\frac{M_{\mathrm{cl}}e^{-\mu}}{2}\cosh\frac{b}{2}(\Phi_{-}-\phi_{0})\label{eq:DNfusion}$$ where we dropped the total time derivatives. This BC is exactly of the form of (\[eq:shGBC\]) with parameters $$M_{\mathrm{cl}}\cosh(\mu\pm\frac{b}{2}\phi_{0})=M_{0}e^{\mp\frac{b}{2}\varphi_{0}}$$ Thus by fusing the integrable defect to the DBC we can reconstruct the most general (two parameter family of) PNBCs. Interestingly $\phi_{0}$ and $\mu$ are the classical analogues of the parameters in which the boundary reflection is factorized [@BRZ], see also (\[eq:RpN\],\[eq:UVIRpN\]) in section 3.
By fusing the defect to the NBC we obtain the boundary potential$$B^{'}(\Phi_{-},\Phi_{+})=D(\Phi_{-},\Phi_{+})$$ Variation of action provides BCs in the form:$$\partial_{t}\Phi_{-}\vert_{x=0}=-\frac{\partial B^{'}(\Phi_{+},\Phi_{-})}{\partial\Phi_{+}}\quad;\qquad(\partial_{t}\Phi_{+}-\partial_{x}\Phi_{-})\vert_{x=0}=\frac{\partial B^{'}(\Phi_{+},\Phi_{-})}{\partial\Phi_{-}}\label{eq:Nfusion}$$ By expressing $\Phi_{+}$ in terms of $\Phi_{-}$ and $\partial_{t}\Phi_{-}$ then plugging back to the second equation we obtain a highly non-trivial boundary condition for $\Phi_{-}$ containing its second time derivative, which is nevertheless integrable as it follows from the construction. Obviously, this solution was not covered by the two parameter family of (time-independent) integrable boundary conditions determined in [@GZ], thus by the fusion method we were able to construct a new type of integrable BC. By fusing other integrable defects with new free parameters to this dressed boundary we can generate integrable BCs with as many parameters as we want. What is nice in the construction, that the solution of the defects transmission factor will provide, via the fusion method, solutions for these general integrable BCs, too, as we will show in the next section.
Finally, we note that similar construction can be used in the case of the Lee-Yang model, however, the explicit form of the integrable defect perturbation has not been identified at the Lagrangian level yet ( for details see the next section).
Fusion method in the bootstrap
==============================
For simplicity we present the fusion idea in the case of an integrable diagonal scattering theory with one particle type of mass $m$. The general discussion can be found in [@DefBound].
In integrable bulk theories multi-particle scattering processes factorize into the product of two particle scatterings, $S(\theta_{12})$, where $\theta_{12}=\theta_{1}-\theta_{2}$ is the rapidity difference of the scattering particles whose momenta are parametrized as $p_{i}=m\sinh\theta_{i}$. In relativistically invariant theories the two particle $S$-matrix satisfies unitarity and crossing symmetry $$S(-\theta)=S^{-1}(\theta)\quad;\qquad S(i\pi-\theta)=S(\theta)$$
Once boundaries are introduced the basic process is the multi-particle reflection. Integrability ensures its factorization into pairwise scatterings $S(\theta_{ij})$ and individual reflections $R(\theta_{i})$, where $\theta_{i}$ is the rapidity of the reflected particle. The reflection matrix satisfies unitarity and boundary crossing unitarity [@GZ]$$R(-\theta)=R^{-1}(\theta)\quad;\qquad R(\frac{i\pi}{2}-\theta)=S(2\theta)R(\frac{i\pi}{2}+\theta)$$
Integrable non-free defects are severely restricted: they are either purely reflecting (thus boundaries, like above) or purely transmitting. The latter case can be described by the two, left ($-$) and right ($+$), transmission matrices $T_{-}(\theta)$ and $T_{+}(-\theta)$. We parametrize $T_{+}$ such a way that for its physical domain ($\theta<0$) its argument is always positive. Transmission factors satisfy unitarity and defect crossing symmetry [@DefBound]$$T_{+}(-\theta)=T_{-}^{-1}(\theta)\quad;\qquad T_{-}(\theta)=T_{+}(i\pi-\theta)\label{Defprop}$$
If we place a defect with transmission matrices $T_{\pm}(\theta)$ in front of a boundary with reflection matrix $R(\theta)$ then the fused boundary will also be integrable and have reflection factor $R^{'}(\theta)$:
$$R^{'}(\theta)=T_{+}(\theta)R(\theta)T_{-}(\theta)\label{eq:Qfusion}$$
The correspondence (\[eq:Qfusion\]) between the original $R(\theta)$ and the fused $R^{'}(\theta)$ reflection factors can be used either to generate new BCs or, if the two BCs are already known, to solve defect transmission factors. This will be illustrated in the next subsections for the sinh-Gordon and Lee-Yang models.
Solution of defect sinh-Gordon theory
--------------------------------------
The spectrum of ShG theory defined by (\[ShGbulkL\]) consists of one particle type with mass [@masscale] $$m=\frac{4\sqrt{\pi}}{\Gamma(\frac{1-B}{2})\Gamma(1+\frac{B}{2})}\left(\frac{-\pi m_{\mathrm{cl}}^{2}\Gamma(1+b^{2})}{b^{2}\Gamma(-b^{2})}\right)^{\frac{1}{2+2b^{2}}}\qquad;\qquad B=\frac{b^{2}}{8\pi+b^{2}}\label{eq:mandB}$$ The two particle scattering matrix is given by $$S=\frac{\sinh\theta-i\sin B\pi}{\sinh\theta+i\sin B\pi}=-(-B)(1+B)\quad,\qquad(x)=\frac{\sinh(\frac{\theta}{2}+\frac{i\pi x}{2})}{\sinh(\frac{\theta}{2}-\frac{i\pi x}{2})}$$ It has no poles in the physical strip: $0\leq\theta<i\pi$ and is invariant under the weak-strong duality, $\frac{b^{2}}{8\pi}\rightarrow\frac{8\pi}{b^{2}}$. The bulk energy density turns out to be [@Ebulk] $$\epsilon_{\mathrm{bulk}}=\frac{m^{2}}{8\sinh\pi B}\label{eq:becshg}$$
Integrable boundary conditions can be either Dirichlet type with $\Phi(0,t)=\phi_{0}$ or PN type (\[eq:shGBC\]). The corresponding reflection factors can be obtained from the analytical continuation of the sine-Gordon’s first breather’s one [@GZ; @Ghoshal]. In the Dirichlet case it reads as
$$R_{\mathrm{Dir}}(\theta,\eta_{\mathrm{Dir}})=\frac{\left(\frac{1}{2}\right)\left(1-\frac{B}{2}\right)}{\left(\frac{3}{2}-\frac{B}{2}\right)}\frac{\left(\frac{iB\eta_{\mathrm{Dir}}}{\pi}-\frac{1}{2}\right)}{\left(\frac{iB\eta_{\mathrm{Dir}}}{\pi}+\frac{1}{2}\right)}\label{eq:RDir}$$
where the reflection parameter $\eta_{\mathrm{Dir}}$ is related to $\phi_{0}$ as $$\eta_{\mathrm{Dir}}=\frac{4\pi}{b}\phi_{0}\label{eq:UVIRDir}$$ The boundary energy has been also calculated [@LMSS]$$\epsilon_{\mathrm{bdry}}^{\mathrm{Dir}}(\eta_{\mathrm{Dir}})=\frac{m}{4\sin B\pi}\left(2\cosh B\eta_{\mathrm{Dir}}-\sin\frac{\pi B}{2}-\cos\frac{\pi B}{2}-1\right)\label{eq:EDir}$$ In the PN case the reflection factor turns out to be $$R_{\mathrm{PN}}(\theta,\eta,\vartheta)=\frac{\left(\frac{1}{2}\right)\left(1-\frac{B}{2}\right)}{\left(\frac{3}{2}-\frac{B}{2}\right)}\frac{\left(\frac{iB\eta}{\pi}-\frac{1}{2}\right)}{\left(\frac{iB\eta}{\pi}+\frac{1}{2}\right)}\frac{\left(\frac{iB\vartheta}{\pi}-\frac{1}{2}\right)}{\left(\frac{iB\vartheta}{\pi}+\frac{1}{2}\right)}\label{eq:RpN}$$ while the relation of $\eta,\vartheta$ to the parameters of the Lagrangian (UV-IR relation) is [@BRZ]$$M\cosh\frac{b^{2}}{8\pi}(\eta\pm\vartheta)=M_{0}e^{\mp\frac{b\varphi_{0}}{2}}\quad;\qquad M=m_{\mathrm{cl}}\sqrt{\frac{2}{b^{2}\sin(b^{2}/8)}}\label{eq:UVIRpN}$$ The boundary energy has been also determined [@BUVIR; @BRZ] as $$\epsilon_{\mathrm{bdry}}^{\mathrm{PN}}(\eta,\vartheta)=\frac{m}{4\sin B\pi}\left(2\cosh B\eta+2\cosh B\vartheta-\sin\frac{\pi B}{2}-\cos\frac{\pi B}{2}-1\right)\label{eq:EpN}$$ We note that the results - both for the UV-IR relation and for the boundary energy - were obtained in the framework of perturbed BCFT in which the perturbing operator is normal-ordered to have a definite scaling dimension.
The integrable defect potential for the sinh-Gordon model can be written as in (\[eq:ShGdefL\]). Let us denote the transmission factors by $T_{\pm}(\theta,\mu)$. Fusing classically this defect to a DBC a PNBC can be obtained (\[eq:DNfusion\]). The quantum analogue of this statement in view of (\[eq:Qfusion\]) is $$R_{\mathrm{PN}}(\theta,\eta,\vartheta)=T_{+}(\theta,\mu)R_{\mathrm{Dir}}(\theta,\eta_{\mathrm{Dir}})T_{-}(\theta,\mu)$$ Comparing the reflection factor of the PNBC (\[eq:RpN\]) to that of the Dirichlet one (\[eq:RDir\]) and taking into account defect unitarity and defect crossing symmetry (\[Defprop\]) we can extract the transmission factors for the defect. The simplest possible solution corresponds to $\eta=\eta_{\mathrm{Dir}}$ and
$$T_{-}(\theta)=-i\frac{\sinh\left(\frac{\theta}{2}-\frac{i\pi}{4}+\frac{B\vartheta}{2}\right)}{\sinh\left(\frac{\theta}{2}+\frac{i\pi}{4}+\frac{B\vartheta}{2}\right)}\quad;\qquad T_{+}(\theta)=i\frac{\sinh\left(\frac{\theta}{2}-\frac{i\pi}{4}-\frac{B\vartheta}{2}\right)}{\sinh\left(\frac{\theta}{2}+\frac{i\pi}{4}-\frac{B\vartheta}{2}\right)}\label{Defsol}$$
All other solutions contain additional CDD type factors satisfying (\[Defprop\]). Actually the solution (\[Defsol\]) itself is a CDD factor, therefore it is the simplest non-trivial solution of (\[Defprop\]).
To find the correspondence between the parameter of the quantum transmission factor $\vartheta$ and the Lagrangian parameter $\mu$ we follow the following strategy: Since the boundary results are derived in the perturbed BCFT normalization we allow not only the parameter $\mu$ but also $M_{{\rm \mathrm{cl}}}$ to renormalize. Their renormalized quantum values are determined from the requirement that when fusing the defect to the DBC with (\[eq:UVIRDir\]) we obtain the PNBC with (\[eq:UVIRpN\]). The unique solution turns out to be $$Me^{\pm\frac{b^{2}\vartheta}{8\pi}}=M_{\mathrm{cl}}e^{\pm\mu}\label{eq:UVIRdef}$$ The renormalization of $M_{\mathrm{cl}}$ may depend on the scheme in which the quantum potential is defined. The $b\to0$ limit, in which $M\to M_{\mathrm{cl}}$, shows that $\vartheta$ is the quantum renormalized version of $\mu$.
Also the defect energy can be extracted as the difference of the boundary energies corresponding to the PN (\[eq:EpN\]) and to the Dirichlet (\[eq:EDir\]) one: $$\epsilon_{\mathrm{Def}}(\vartheta)=\epsilon_{\mathrm{bdry}}^{\mathrm{PN}}(\eta,\vartheta)-\epsilon_{\mathrm{bdry}}^{\mathrm{Dir}}(\eta)=\frac{m\cosh B\vartheta}{2\sin B\pi}\label{eq:Edef}$$ Summarizing, by the fusion method we were able to solve the defect theory defined by the Lagrangian (\[eq:ShGdefL\]): The transmission factors are (\[Defsol\]), the defect energy is (\[eq:Edef\]), and the bootstrap parameter $\vartheta$ parametrizes the Lagrangian as (\[eq:UVIRdef\]). We spend the next section to provide consistency checks of this solution.
Once the defect theory is solved we can use it to generate new integrable BCs from known ones. In the example presented in section \[sec:Lagrangianshg\] the defect with parameter $\mu$ was fused to the NBC to generate a more general integrable BC (\[eq:Nfusion\]). The quantum version of this fusion dresses up the Neumann reflection factor $$R_{\mathrm{N}}(\theta)=\frac{\left(\frac{1}{2}\right)\left(1-\frac{B}{2}\right)}{\left(\frac{3}{2}-\frac{B}{2}\right)}\frac{\left(\frac{1}{2}-\frac{B}{2}\right)}{\left(\frac{1}{2}+\frac{B}{2}\right)}\label{eq:RNeumann}$$ to the reflection factor$$R(\theta,\vartheta)=T_{+}(\theta,\mu(\vartheta))R_{\mathrm{N}}(\theta)T_{-}(\theta,\mu(\vartheta))=\frac{\left(\frac{1}{2}\right)\left(1-\frac{B}{2}\right)}{\left(\frac{3}{2}-\frac{B}{2}\right)}\frac{\left(\frac{1}{2}-\frac{B}{2}\right)}{\left(\frac{1}{2}+\frac{B}{2}\right)}\frac{\left(\frac{iB\vartheta}{\pi}-\frac{1}{2}\right)}{\left(\frac{iB\vartheta}{\pi}+\frac{1}{2}\right)}\label{eq:RNdressed}$$ Thus we solved the more general integrable BC defined by (\[eq:Nfusion\]) without doing any serious calculation. It is important to note, that the extra factor appearing in (\[eq:RNdressed\]) compared to (\[eq:RNeumann\]) is a CDD factor. Consequently, we have determined the physical meaning of the CDD factors appearing in the reflection factors: they represent integrable defects of the form of (\[eq:Nfusion\]) standing in front of integrable boundaries. In principle, we can fuse as many integrable defects with various parameters as we want, the resulting theory can be solved and its reflection factor contains the corresponding boundary CDD factors.
Finally, we note that placing two defects with parameters $\vartheta_{\pm}=\pm i(1-\frac{\pi}{2B})$ after each other both the scattering matrix and the energy of a standing particle can be reproduced. Thus the defect with imaginary parameter can be considered as a ’half’ particle. Similar phenomena was observed in [@sgdef] at the classical level.
Solution of defect scaling Lee-Yang model
-----------------------------------------
The scaling Lee-Yang model can be defined as the perturbation of the $\mathcal{M}_{(2,5)}$ conformal minimal model with central charge $c=-\frac{22}{5}$. It contains two modules of the Virasoro algebra corresponding to the $Id$ and the $\varphi(z,\bar{z})$ primary fields with weight $(0,0)$ and $(-\frac{1}{5},-\frac{1}{5})$, respectively. The only relevant perturbation by the field $\varphi$ results in the simplest scattering theory with one neutral particle of mass $m$ and scattering matrix [@SYLbulk] $$S(\theta)=\frac{\sinh\theta+i\sin\frac{\pi}{3}}{\sinh\theta-i\sin\frac{\pi}{3}}=-\left(\frac{1}{3}\right)\left(\frac{2}{3}\right)$$ The pole at $\theta=\frac{i\pi}{3}$ shows that the particle can form a bound-state. The relation$$S(\theta+i\frac{\pi}{3})S(\theta-i\frac{\pi}{3})=S(\theta)$$ however, implies that the bound-state is the original particle itself and the bulk bootstrap is closed. The bulk energy constant is given by $\epsilon_{\mathrm{bulk}}=-\frac{1}{4\sqrt{3}}m^{2}$.
We can impose two conformal invariant boundary conditions in the model [@BYLTBA; @BYL1pt]. They can be labeled by $\mathbb{I}$ and $\Phi$ and correspond to the highest weight representations of a single copy of the Virasoro algebra with weight $0$ and $-\frac{1}{5}$, respectively. Introducing the integrable bulk perturbations with the $\mathbb{I}$ conformal invariant boundary condition the integrability is maintained and the reflection factor of the particle can be written as $$R_{\mathbb{I}}(\theta)=\left(\frac{1}{2}\right)\left(\frac{1}{6}\right)\left(-\frac{2}{3}\right)$$ while the boundary energy is given by $\epsilon_{\mathrm{bdry}}^{\mathbb{I}}=\frac{m}{2}\left(\sqrt{3}-1\right)$. If the conformal invariant boundary condition corresponds to $\Phi$ then additionally to the bulk perturbation we can introduce a one-parameter family of integrable boundary perturbations and the corresponding reflection factor turns out to be$$R_{b}(\theta)=\left(\frac{1}{2}\right)\left(\frac{1}{6}\right)\left(-\frac{2}{3}\right)\left(\frac{b-1}{6}\right)\left(\frac{b+1}{6}\right)\left(\frac{5-b}{6}\right)\left(\frac{-5-b}{6}\right)$$ while the boundary energy is $\epsilon_{\mathrm{bdry}}=\frac{m}{2}\left(\sqrt{3}-1+2\sin\frac{b\pi}{6}\right)$. The boundary bound-states were analyzed in [@BCT] where the boundary bootstrap program was carried out.
The Lee-Yang model has two types of conformal defects [@YLD], but only one of them admits relevant chiral defect fields. They have weights $(-\frac{1}{5},0)$, $(0,-\frac{1}{5})$. We conjecture that perturbing in the bulk and with a certain combination of these defect fields we can maintain integrability and arrive at a purely transmitting theory. We plan to analyze this issue systematically in a forthcoming publication. Let us denote the transmission factors of this integrable defect by $T_{\pm}(\theta)$. Using that the fusion of the defect to the $\mathbb{I}$ boundary results in the perturbed $\Phi$ boundary we have$$R_{b}(\theta)=T_{+}(\theta,b)R_{\mathbb{I}}(\theta)T_{-}(\theta,b)$$ This is supported by the fact that fusing the conformal defect to the $\mathbb{I}$ boundary we obtain the $\Phi$ boundary. Since the particle appears as a bound-state in the two particle scattering process the transmission matrix satisfies the defect bootstrap equation [@Defboot]: $$T_{-}(\theta+\frac{i\pi}{3})T_{-}(\theta-\frac{i\pi}{3})=T_{-}(\theta)\label{eq:LYTboot}$$ Using this relation together with the defect unitarity and defect crossing symmetry (\[Defprop\]) we can fix the transmission factor as $$T_{-}(\theta)=[b+1][b-1]\quad;\qquad[x]=i\frac{\sinh(\frac{\theta}{2}+i\frac{\pi x}{12})}{\sinh(\frac{\theta}{2}+i\frac{\pi x}{12}-i\frac{\pi}{2})}\label{eq:YLdefsol}$$ (Actually the inverse of the solution is also a solution but the two are related by the $b\to6+b$ transformation). The defect energy, as in the sinh-Gordon case, can be obtained as $$\epsilon_{\mathrm{def}}=\epsilon_{\mathrm{bdry}}-\epsilon_{\mathrm{bdry}}^{\mathbb{I}}=m\sin\frac{b\pi}{6}$$ We are going to recover this expression from the UV analysis of defect TBA in section 5. We also note that the defect with parameter $b=3$ behaves as a standing particle both from the energy and from the scattering point of view.
Peturbative calculations
========================
In this section we check the exact solution of the ShG defect system in the free/classical ($b\to0$) limit, and develop a systematic perturbative expansion. ** The parameter $b^{2}$ plays the same role as $\hbar$ which can be seen by scaling it out from the Lagrangian via $\Phi\to b\Phi$. Since the classical groundstate $\Phi=0$ is invariant under this scaling, the $b^{2}\to0$ limit corresponds both to the free and also to the classical limit. Moreover, the perturbative expansion in $b^{2}$ is equivalent both to the loop expansion and to the semi-classical approximation.
Classical/free limit
--------------------
As a first step we identify the $b\to0$ limit of the defect Lagrangian (\[eq:ShGdefL\]) as: $$\begin{aligned}
\mathcal{L} & = & \Theta(-x)\left[\frac{1}{2}(\partial_{\mu}\Phi_{-})^{2}-\frac{m_{\mathrm{cl}}^{2}}{2}\Phi_{-}^{2}\right]+\Theta(x)\left[\frac{1}{2}(\partial_{\mu}\Phi_{+})^{2}-\frac{m_{\mathrm{cl}}^{2}}{2}\Phi_{+}^{2}\right]\\
& & -\frac{\delta(x)}{2}\left(\Phi_{+}\dot{\Phi}_{-}-\Phi_{-}\dot{\Phi}_{+}+m_{\mathrm{cl}}\left[\cosh\mu\left(\Phi_{+}^{2}+\Phi_{-}^{2}\right)+2\sinh\mu\,\Phi_{+}\Phi_{-}\right]\right)\end{aligned}$$ Then we expand the fields on the two sides of the defect in terms of creation/annihilation operators $$\Phi_{\pm}(x,t)=\int_{-\infty}^{\infty}\frac{dk}{2\pi}\frac{1}{2\omega(k)}\left(a_{\pm}(k)e^{ikx-i\omega(k)t}+a_{\pm}^{+}(k)e^{-ikx+i\omega(k)t}\right)\quad;\quad\omega(k)=\sqrt{k^{2}+m_{\mathrm{cl}}^{2}}$$ where the $a,a^{+}$ operators are adjoint of each other with commutators:$$[a_{\pm}(k),a_{\pm}^{+}(k^{'})]=2\pi2\omega(k)\delta(k-k^{'})$$ Imposing the defect condition (obtained by varying the action) at the origin $$\begin{aligned}
\pm\partial_{t}\Phi_{\pm}\mp\partial_{x}\Phi_{\mp} & = & m_{\mathrm{cl}}(\sinh\mu\,\Phi_{\pm}+\cosh\mu\,\Phi_{\mp})\end{aligned}$$ we can connect the creation/annihilation operators as $$a_{\pm}(\pm k)=T_{\mp}(k)a_{\mp}(\pm k)\quad;\qquad T_{\mp}(k)=-\frac{m_{\mathrm{cl}}\sinh\mu\mp i\omega(k)}{m_{\mathrm{cl}}\cosh\mu-ik}\quad;\quad k>0$$ This shows that the defect is purely transmitting, that is we do not have any reflected wave. The transmission factor in the rapidity parametrization ($k=m_{\mathrm{cl}}\sinh\theta)$ can be written also in the following form: $$T_{-}(\theta)=-i\frac{\sinh(\frac{\theta}{2}-\frac{i\pi}{4}+\frac{\mu}{2})}{\sinh(\frac{\theta}{2}+\frac{i\pi}{4}+\frac{\mu}{2})}$$ Clearly it has exactly the same form as the exact quantum one (\[Defsol\]) except the $B\vartheta\leftrightarrow\mu$ replacement. Having observed this coincidence the authors in [@sgdef] suggested that they might be the same $B\vartheta=\mu$. Using our defect UV-IR relation (\[eq:UVIRdef\]) we can perform the expansion: $$B\vartheta=\mu(1-\frac{b^{2}}{8\pi}+\dots)\label{eq:murenb2}$$ The term of first order shows that our exact solution is correct in the classical limit, i.e. for $b\to0$. The term of second order shows that the $B\vartheta=\mu$ relation suggested in [@sgdef] is not valid: $B\vartheta$ acquires nontrivial quantum correction. Since the renormalization of the parameter $\mu$ is crucial to decide about the two proposals we perform a perturbative check at order $b^{2}$.
Perturbation theory
-------------------
As a first step we collect the free propagators. If the fields are on the same side of the defect we have
-- --------------------------------
{height="2cm"}
-- --------------------------------
where $q=(k,\omega)$ and $y=(x,t)$. The absence of an $e^{ik(x+x^{'})}$ term shows the absence of reflection. The other two point functions are
-- --------------------------------
{height="2cm"}
-- --------------------------------
where $$T_{\pm}(\omega,k)=-\frac{m_{\mathrm{cl}}\sinh\mu\pm i\omega}{m_{\mathrm{cl}}\cosh\mu-ik}$$ In the final equations we used the fact that the $\omega$ contour can be closed on the upper/lower half plane. As it was shown in [@BRF] the reflection factor can be read off from the $\langle T(\Phi_{\pm}\Phi_{\pm})\rangle$propagator of the fields. The defect/boundary equivalence [@DefBound] then implies that the transmission factor can be read off from the $\langle T(\Phi_{\mp}\Phi_{\pm})\rangle$ propagator.
The perturbation at order $b^{2}$ follows from (\[ShGbulkL\]): $$\begin{aligned}
\delta\mathcal{L} & = & -\Theta(-\zeta)\left[\frac{m_{\mathrm{cl}}^{2}b^{2}}{4!}\Phi_{-}^{4}\right]-\Theta(\zeta)\left[\frac{m_{\mathrm{cl}}^{2}b^{2}}{4!}\Phi_{+}^{4}\right]\\
& & -\delta(\zeta)\frac{m_{\mathrm{cl}}b^{2}}{4\cdot4!}\left[\cosh\mu\left(\Phi_{+}^{4}+6\Phi_{+}^{2}\Phi_{-}^{2}+\Phi_{-}^{4}\right)+4\sinh\mu\,\Phi_{+}\Phi_{-}\left(\Phi_{+}^{2}+\Phi_{-}^{2}\right)\right]\end{aligned}$$ We calculate the propagators upto first order in $b^{2}$: $$\langle0\vert T\left(\Phi_{\mp}(x,t)\Phi_{\pm}(x^{'},t^{'})\right)\vert0\rangle=\,\,_{0}\langle0\vert T\left(\Phi_{\mp}(x,t)\Phi_{\pm}(x^{'},t^{'})(1-i\int d\zeta\int d\tau\,\delta\mathcal{L}+\dots)\right)\vert0\rangle_{0}$$ Using Wick’s theorem we obtain the contribution of the following diagrams:
We have two bulk diagrams presented on Figure 1:
{height="2.5cm"}{height="2.5cm"}
where the bulk interaction point, denoted by an empty circle, represents $z=(\zeta,\tau)$ and we have to integrate over the whole left/right space-time. Thus the contribution of the first diagram is $$\frac{m_{\mathrm{cl}}^{2}b^{2}}{2}\int_{-\infty}^{\infty}d\tau\int_{-\infty}^{0}dz\, G_{-}^{-}(y,z)G_{-}^{-}(z,z)G_{-}^{+}(z,y')$$ Clearly $G_{-}^{-}(z,z)$ is divergent and we have to regularize it by introducing a cutoff $\Lambda$: $$G_{-}^{-}(z,z)=\int\frac{d^{2}q}{(2\pi)^{2}}\frac{i}{q^{2}-m_{\mathrm{cl}}^{2}+i\epsilon}=\int_{0}^{\Lambda}\frac{1}{\sqrt{k^{2}+m_{\mathrm{cl}}^{2}}}\frac{dk}{2\pi}=\Delta(m_{\mathrm{cl}})$$ Its contribution can be absorbed into the renormalization of the mass parameter $m_{\mathrm{cl}}^{2}\to m_{\mathrm{cl}}^{2}-m_{\mathrm{cl}}^{2}\frac{b^{2}}{2}\Delta(m_{\mathrm{cl}})$ which results in extra counter term diagrams presented on Figure 2:
{height="2.5cm"}{height="2.5cm"}
The contributions from the defect terms can be grouped in two sets of diagrams. The first contains the same divergent loop integral and consists of those on Figure 3:
{height="2.5cm"}{height="2.5cm"}{height="2.5cm"}{height="2.5cm"}
where the interaction vertex is even together with the odd diagrams presented on Figure 4:
{height="2.5cm"}{height="2.5cm"}{height="2.5cm"}{height="2.5cm"}
We have to integrate in time $\tau$ over the real axis and the left/right part of the defect represents the contraction with the operators $\Phi_{-}$ and $\Phi_{+}$, respectively. They all contain the divergent and regularized $\Delta(m_{\mathrm{cl}})$ loop integral which can be absorbed into the renormalization of the defect parameter $m_{\mathrm{cl}}\to m_{\mathrm{cl}}-m_{\mathrm{cl}}\frac{b^{2}}{4}\Delta(m_{\mathrm{cl}})$, which is consistent with the bulk renormalization. The resulting counter-terms produce the diagrams on Figure 5:
{height="2.5cm"}{height="2.5cm"}{height="2.5cm"}{height="2.5cm"}
The fact that all these singularities can be absorbed into the renormalization of $m_{\mathrm{cl}}$ is a nontrivial statement, since we have eight divergent diagrams having different propagators on the outer legs those we canceled just by renormalizing one single parameter in the original Lagrangian. The form of the renormalized Lagrangian is the same as the original one thus the integrable/topological nature of the defect is not spoiled by quantum effects, there is no anomaly. Observe also that the bulk mass term $m_{\mathrm{cl}}^{2}$ and the boundary term $m_{\mathrm{cl}}$ renormalizes the same way so the bulk is the square of the other.
The last group of the diagrams is the one which really contributes to the transmission factor. They are presented on Figure 6:
{height="2.5cm"}{height="2.5cm"}{height="2.5cm"}{height="2.5cm"}
The contribution of the first diagram is $$\frac{m_{\mathrm{cl}}b^{2}}{4}\sinh\mu\int_{-\infty}^{\infty}d\tau\, G_{-}^{-}(y,z)G_{-}^{+}(z,z)G_{-}^{+}(z,y')$$ Each of the terms on Figure 6 contains the finite contribution of the propagator $$G_{\pm}^{\mp}(z,z)=\int\frac{d^{2}q}{(2\pi)^{2}}\frac{i}{q^{2}-m_{\mathrm{cl}}^{2}+i\epsilon}T_{\pm}(\omega,k)=-\frac{\mu}{2\pi}$$ This term together with the prefactor can be interpreted as the finite renormalization of the parameter $\mu\to\mu+\delta\mu$, where $\delta\mu=-\frac{\mu b^{2}}{8\pi}$. Summing up all the contributions and taking into account the different transmission factor dependent contributions on the outer legs we obtain the correction of order $b^{2}$ to the transmission factor as $$T_{-}(\theta,B\vartheta(\mu))=T_{-}(\theta,\mu)+\frac{\mu b^{2}}{8\pi}\frac{1}{1-i\sinh(\theta+\mu)}+\mathcal{O}(b^{4})$$ which is in complete agreement with (\[eq:murenb2\]). Thus we confirmed the renormalization of the parameter $\mu$, but we have not checked the renormalization of the parameter $M$ which has not shown up at this order. We suspect, however, that in this perturbative scheme the boundary parameter $m_{\mathrm{cl}}$ renormalizes as the square-root of the $m_{\mathrm{cl}}^{2}$ term and only $\mu$ renormalizes as $B\vartheta$. It would be interesting to perform a two-loop perturbative calculation to decide about the renormalization of $M_{\mathrm{cl}}$ in the perturbative scheme.
We note that we have also performed a perturbative calculation of order $b^{2}$ of the reflection factor, which can be extracted from $G_{+}^{+}(y,y')$, and confirmed the absence of reflection at this order. In [@Patrick] the form of the sine-Gordon Lagrangian was fixed at order $b^{6}$ by demanding the absence of particle creation. Following a similar line it would be tempting to see how the absence of reflection restricts the form of the defect potential in perturbation theory.
Defect thermodynamic Bethe ansatz
=================================
In this section we would like to check the consistency between the transmission factors and defect energies. In doing so we derive a DTBA to describe the ground state energy of a purely transmitting diagonal integrable defect on the circle of perimeter $L$. In boundary and defect systems there are two inequivalent ways to derive TBA equations. We can either focus on the groundstate energy or on the so-called $g$-factors which is related to the finite volume normalization of defect/boundary states. Both BTBA equations have been analyzed in [@LMSS] although the $g$-function type required further refinement [@gBTBA]. Some details of the $g$-function type DTBA can be found in [@gDTBA] and references therein. Here we consider the groundstate DTBA in the periodic setting as opposed to the strip geometry analyzed in [@DefBound].
Derivation of DTBA
------------------
In order to derive the DTBA equation for the ground state energy we compactify the time-like direction with period $R$ and calculate the partition function in two inequivalent ways by changing the role of the space and time coordinates.
In the original description the defect is located in space as drawn in Figure 7:
{height="5cm"}
By taking the $R\to\infty$ limit the groundstate energy can be extracted from the partition function as $$\lim_{R\to\infty}Z(L,R)=\lim_{R\to\infty}Tr\left(e^{-H(L)R}\right)=e^{-E_{0}(L)R}+\dots$$ In the alternative description when the role of time and space is exchanged as shown on Figure 8:
{width="5cm"}
the defect becomes an operator of the form [@DefBound] $$D=\exp\left\{ \int_{-\infty}^{\infty}\frac{d\theta}{2\pi}T_{+}(\frac{i\pi}{2}-\theta)a^{+}(\theta)a(\theta)\right\}$$ which acts on the Hilbert space of the periodic model, $\mathcal{H}$. The partition function can be calculated as $$Z(L,R)=Tr\left(e^{-H(R)L}D\right)=\sum_{\vert n\rangle\in\mathcal{H}}\frac{\langle n\vert D\vert n\rangle e^{-E_{n}(R)L}}{\langle n\vert n\rangle}\label{eq:ZRL}$$ The Hilbert space consists of multi-particle states $$\vert\theta_{1},\theta_{2},\dots,\theta_{n}\rangle=a^{+}(\theta_{1})a^{+}(\theta_{2})\dots a^{+}(\theta_{n})\vert0\rangle\quad;\qquad\theta_{1}>\theta_{2}>\dots>\theta_{n}$$ on which the defect operator collects nontrivial diagonal matrix elements from $$D\vert\theta_{1},\theta_{2},\dots,\theta_{n}\rangle=T_{+}(\frac{i\pi}{2}-\theta_{1})T_{+}(\frac{i\pi}{2}-\theta_{2})\dots T_{+}(\frac{i\pi}{2}-\theta_{n})\vert\theta_{1},\theta_{2},\dots,\theta_{n}\rangle+\dots$$ while the energy operator acts as$$H\vert\theta_{1},\theta_{2},\dots,\theta_{n}\rangle=\left(m\cosh(\theta_{1})+m\cosh(\theta_{2})+\dots+m\cosh(\theta_{n})\right)\vert\theta_{1},\theta_{2},\dots,\theta_{n}\rangle$$ We can introduce another energy operator via $\hat{H}=H-\frac{1}{L}\log D$ such that the partition function can be written as $$Z(L,R)=Tr\left(e^{-H(R)L}D\right)=Tr\left(e^{-\hat{H}(R)L}\right)=\sum_{\vert n\rangle\in\mathcal{H}}e^{-\hat{E}_{n}(R)L}$$ This partition function can be calculated in the $R\to\infty$ limit by standard saddle point approximation taking into account the scattering of the particles. The calculation follows the usual route of TBA calculations, but now the kinetic term is shifted $m\cosh\theta\to m\cosh\theta-\frac{1}{L}\log T_{+}(\frac{i\pi}{2}-\theta)$. As a consequence we obtain the following DTBA equations for the pseudo energy $$\tilde{\epsilon}(\theta)=mL\cosh\theta-\log T_{+}(\frac{i\pi}{2}-\theta)-\int_{-\infty}^{\infty}\frac{d\theta^{'}}{2\pi}\phi(\theta-\theta^{'})\log(1+e^{-\tilde{\epsilon}(\theta^{'})})\label{eq:DTBA1}$$ where $\phi(\theta)=-i\frac{d}{d\theta}\log S(\theta)$. Once $\tilde{\epsilon}$ is known the ground state energy can be expressed as $$E_{0}(L)=-m\int_{-\infty}^{\infty}\frac{d\theta}{2\pi}\,\cosh\theta\,\log(1+e^{-\tilde{\epsilon}(\theta)})$$ Here we do not have to shift the $\cosh\theta$ term since it comes from the derivative of the momentum ($\sinh\theta$), which appears in the quantization condition. For the result in this generality see e.g. [@O(1)].
Alternatively, we can redefine the pseudo energy as $\tilde{\epsilon}(\theta)=\epsilon(\theta)-\log T_{+}(\frac{i\pi}{2}-\theta)$ to obtain $$\epsilon(\theta)=mL\cosh\theta-\int_{-\infty}^{\infty}\frac{d\theta^{'}}{2\pi}\phi(\theta-\theta^{'})\log\left(1+T_{+}(\frac{i\pi}{2}-\theta^{'})e^{-\epsilon(\theta^{'})}\right)$$ from which the ground state energy turns out to be $$E_{0}(L)=-m\int_{-\infty}^{\infty}\frac{d\theta}{2\pi}\,\cosh\theta\,\log\left(1+T_{+}(\frac{i\pi}{2}-\theta)e^{-\epsilon(\theta)}\right)$$ The ground state energy is real which can be easily seen form (\[Defprop\]) since $T_{+}^{*}(\frac{i\pi}{2}-\theta)=T_{+}(\frac{i\pi}{2}+\theta)$. As a simple consistency check we can see that the DTBA equation for the trivial defect, $T_{+}=1$, reduces to the periodic TBA equation [@TBA]. We analyze the large and small volume limits separately in the next two subsections.
Lüscher type correction in defect systems
-----------------------------------------
If the volume $L$ is large then $\epsilon(\theta)\cong mL\cosh\theta$ is large and we can expand the logarithm to obtain$$E_{0}(L)=-m\int_{-\infty}^{\infty}\frac{d\theta}{2\pi}\,\cosh\theta\,\, T_{+}(\frac{i\pi}{2}-\theta)e^{-mL\cosh\theta}+O(e^{-2mL})$$ This result can be calculated directly from (\[eq:ZRL\]) by taking the large $L$ limit there. In that case, however, only the one-particle transmission term contributes which is universal for any quantum field theory. The result obtained is the analogue of the boundary Lüscher type correction to the ground state energy [@BLusch] and is valid in any theory even in non-integrable ones.
Defect energy
-------------
Here we analyze the small volume behavior of the ground state energy. Its normalization depends on the scheme in which the quantum field theory is defined. If we would like to compare the DTBA normalization to that of a perturbed defect conformal field theory, in which the perturbing operators have dimensions $h$ on the defect and $(h,h)$ in the bulk, then we have: $$E_{0}(L)=-\epsilon_{\mathrm{def}}-\epsilon_{\mathrm{bulk}}L+\frac{2\pi}{L}\sum_{n=0}^{\infty}c_{n}\,{l}^{n(1-h)}\quad;\qquad l=mL$$ Only the perturbative terms, $c_{n}$, are present in a perturbed rational defect CFT. (In non-rational CFT-s, like the UV limit of the boundary sinh-Gordon theory, we expect terms with logarithmic behaviour, see [@BRZ] for the details). By calculating the small volume limit of $E_{0}(L)$ from DTBA $\epsilon_{\mathrm{def}}$ and $\epsilon_{\mathrm{bulk}}$ can be extracted exactly. The computation is analogous to the boundary one [@BUVIR; @BYLTBA], so we sketch only here. In the $L\to0$ limit the solution for $\tilde{\epsilon}$ in (\[eq:DTBA1\]) develops two kink regions around $\theta=\pm\log\frac{2}{l}$ and a breather region around the origin. The behaviour of the solutions are determined by the $\theta\to\pm\infty$ asymptotics of the integral kernel and defect source term: $$\phi(\theta)=Ce^{-\vert\theta\vert}+O(e^{-2\vert\theta\vert})\quad;\quad\log(T_{+}(\frac{i\pi}{2}-\theta))=A_{\pm}e^{\mp\theta}+O(e^{\mp2\theta})\qquad\mathrm{as}\quad\theta\to\pm\infty$$ The two kink functions are responsible for the terms giving the central charge and the bulk energy constant, while the central/breather part gives the defect energy in the following form: $$\epsilon_{\mathrm{bulk}}=\frac{m^{2}}{2C}\quad;\qquad\epsilon_{\mathrm{def}}=-\frac{m(A_{+}+A_{-})}{2C}\label{eq:becdefe}$$ We note that the kink type behaviour does not exists for the whole parameter range of $C$ and $A_{\pm}$. The results is understood that we analytically continued it from a range where the calculation is reliable.
Let us concretes the result for the two cases in question. In the sinh-Gordon model $$C=4\sinh\pi B\quad;\qquad A_{\pm}=-2e^{\mp B\vartheta}$$ so using (\[eq:becdefe\]) we recover (\[eq:becshg\]) and (\[eq:Edef\]).
In the Lee-Yang case we have $$C=-2\sqrt{3}\quad;\qquad A_{\pm}=\mp2i(e^{\pm i\pi\frac{b+1}{6}}+e^{\pm i\pi\frac{b-1}{6}})$$ Plugging these expressions back to (\[eq:becdefe\]) the results confirms the bulk energy density and the defect energy.
We emphasize that the agreement obtained in the two cases confirm the solutions on one side and the DTBA equation on the other.
The other perturbative coefficients $c_{n}$ can be calculated from the DTBA only numerically. In the Lee-Yang case, however, one can gain further analytical information. One has to define $Y(\theta)=e^{-\epsilon(\theta)}$ and to show from (\[eq:LYTboot\]) that it satisfies the Lee-Yang $Y$-system relation$$Y(\theta-\frac{i\pi}{3})Y(\theta+\frac{i\pi}{3})=1+Y(\theta)$$ from which the $Y(\theta)=Y(\theta+\frac{5i\pi}{3})$ periodicity follows. Similarly to the boundary case this gives the exponent of the perturbative expansion to be $\frac{6}{5}$ showing that the dimension of the perturbing operator is $h=-\frac{1}{5}$. **
Defect bound-states and bootstrap closure
=========================================
In this section we analyze the analytic structure of the transmission factors for the whole range of their parameters. Since the sinh-Gordon transmission factors (\[Defsol\]) never have poles in the physical strip we focus on the Lee-Yang model only. Recall that in our convention the physical strip of the transmission factors $T_{\mp}(\theta)$ are $\Im m\theta\in[0,\frac{\pi}{2}]$.
Pole analysis on the ground-state defect
----------------------------------------
We analyze the pole structure of both $$T_{-}(\theta)=[b+1][b-1]\quad\mbox{and}\quad T_{+}(\theta)=[5-b][-5-b]$$ as the function of the parameter $b$, simultaneously. We note that by folding the theory to a boundary one (with two particles) we could analyze its bootstrap in the usual boundary formulation. Here, however, we present the results in the defect language since the diagrams are more clear-cut. (The whole procedure, going from the reflection factor to the defect transmission factor, can be interpreted as taking a sort of square root of the boundary theory. By closing the defect bootstrap we would like to show that such a theory is indeed sensible. Since on the $\mathbb{I}$ boundary there is no boundstate any pole of the reflection factor appears either in $T_{-}$ or in $T_{+}$ so the bootstrap will be very similar to the boundary one [@BCT]).
In determining the fundamental range of the parameter $b$ we can see that $b\to b+12$ is a symmetry. Moreover $b\leftrightarrow6-b$ exchanges $T_{-}\leftrightarrow T_{+}$ so we can restrict ourselves to the range $b\in[-3,3]$. We will see by analyzing the defect excited states that the fundamental range is even smaller, only $b\in[-3,2]$, as in the boundary case, since at $b=2$ the role of the ground-state and the first excited state is exchanged.
The poles and zeros of the transmission factor $T_{-}(\theta)$ are $$\textrm{poles of $T_{-}$ are at }\theta=-i\frac{\pi}{6}(b\pm5)\quad;\qquad\textrm{zeros of $T_{-}$ are at }\theta=-i\frac{\pi}{6}(b\pm1)$$ The analogous expressions for $T_{+}$ are $$\textrm{poles of $T_{+}$ are at }\theta=i\frac{\pi}{6}(b\pm1)\quad;\qquad\textrm{zeros of $T_{+}$ are at }\theta=i\frac{\pi}{6}(b\pm5)$$ They can be drawn as the function of the parameter $b$ as shown on Figure 9:
{width="6cm"}
For $b\in[-1,2]$ there is a pole in the transmission factor $T_{+}$ at $\theta=iu=i\frac{\pi}{6}(b+1)$, for which we associate a defect boundstate and denote it by $\vert1+\rangle$. Its energy is $m\cos\frac{\pi}{6}(b+1)$ and the corresponding excited transmission factor can be calculated from the defect bootstrap equation shown on Figure 10:
{height="4cm"}{height="4cm"}
$$T_{-}^{\vert1+\rangle}(\theta)=T_{-}(\theta)S(\theta+iu)$$ From the defect crossing symmetry (\[Defprop\]) we can calculate $T_{+}^{\vert1+\rangle}(\theta)$ as $$T_{+}^{\vert1+\rangle}(\theta)=T_{-}^{\vert1+\rangle}(i\pi-\theta)=T_{-}(i\pi-\theta)S(i\pi-\theta+iu)=T_{+}(\theta)S(\theta-iu)$$ which is consistent with the other bootstrap equation where the second particle arrives from the right. The resulting transmission factors are $$T_{-}^{\vert1+\rangle}(\theta)=[b+1][b+3]\quad;\qquad T_{+}^{\vert1+\rangle}(\theta)=[5-b][3-b]$$ They are related to the groundstate ones as $T_{\pm}^{\vert1+\rangle}(b\to4-b,\theta)=T_{\mp}(b,\theta)$. This symmetry together with the defect energies indicate that when $b$ exceeds $2$ the role of the ground-state and the excited state $\vert1+\rangle$ are exchanged. This confirms that the fundamental range is indeed $b\in[-3,2]$.
In the range $b\in[1,2]$ the transmission factor $T_{+}(\theta)$ has another pole at $\theta=i\frac{\pi}{6}(b-1)$ for which we associate the defect boundstate $\vert2+\rangle$. It has energy $m\cos\frac{\pi}{6}(b-1)$ and transmission factor $$T_{\pm}^{\vert2+\rangle}(\theta)=T_{\pm}(\theta)S(\theta\mp i\frac{\pi}{6}(b-1))$$ Now we turn to the pole analysis of excited defect states.
Pole analysis on the exited defect state $\vert1+\rangle$
---------------------------------------------------------
The poles and zeros of the transmission factors on the state $\vert1+\rangle$ are indicated on Figure 11.
{width="6cm"}
The state exist in the $b\in[-1,2]$ domain so we have to explain the poles in this range only.
The pole of $T_{+}^{\vert1+\rangle}$ labeled by $1$ on Figure 11 is at the same location as the one which creates the excited state itself, namely at $\theta=i\frac{\pi}{6}(b+1)$ in the full range $b\in[-1,2]$. It can be explained by the first of the defect Coleman-Thun diagrams on Figure 12
{height="5cm"}{height="5cm"}{height="5cm"}
The pole of $T_{+}^{\vert1+\rangle}$ labeled by $2$ on Figure 11 is at $\theta=i\frac{\pi}{6}(b+3)$ and can be explained by the second diagram on Figure 12. Observe that by applying the Cutkosky rules [@BBT] we would obtain a pole of second order but the transmission factor $T_{-}$ has a first order zero at $\theta=i\frac{\pi}{6}(1-b)$ which, in this way, reduces the order of the pole to one.
The pole of $T_{-}^{\vert1+\rangle}$ labeled by $3$ on Figure 11 is at $\theta=i\frac{\pi}{6}(3-b)$. In the range $b\in[0,1]$ it can be explained by the third diagram on Figure 12. Since the transmission factor $T_{-}$ on the ground state has a zero at $\theta=i\frac{\pi}{6}(1-b)$ the order of the diagram is reduced to one again. In order for the diagram to exist the particle has to travel towards the defect, that is $1-b>0$. This explains the pole in the range $b\in[0,1]$. In the range $b\in[1,2]$ the particle creates a defect boundstate which is nothing but $\vert2+\rangle$. This can be seen both from the energy of the excited state $m\cos\frac{\pi}{6}(b+1)+m\cos\frac{\pi}{6}(3-b)=m\cos\frac{\pi}{6}(b-1)$ and from the transmission factor. If the left particle creates a defect boundstate at rapidity $\theta=iu$ then the excited states transmission factors are $T_{\pm}^{ex}(\theta)=T_{\pm}(\theta)S(\theta\pm iu)$. Now we can see from the bulk bootstrap equation that $$T_{\pm}^{\vert1+\rangle}(\theta)S(\theta\pm i\frac{\pi}{6}(3-b))=T_{\pm}(\theta)S(\theta\mp i\frac{\pi}{6}(b+1))S(\theta\pm i\frac{\pi}{6}(3-b))=T_{\pm}(\theta)S(\theta\mp i\frac{\pi}{6}(b-1))=T_{\pm}^{\vert2+\rangle}(\theta)$$ that is the transmission factors also supports the identification.
The pole analysis on the excited defect state $\vert2+\rangle$
--------------------------------------------------------------
The defect boundstate labeled by $\vert2+\rangle$ has transmission factor $$T_{-}^{\vert2+\rangle}(\theta)=[b-1][b+1]^{2}[b+3]\quad;\qquad T_{+}^{\vert2+\rangle}(\theta)=[3-b][5-b]^{2}[7-b]$$ The singularity structure can be summarized as follows.
{width="6cm"}
The pole labeled by $5$ on Figure 13 is in $T_{+}^{\vert2+\rangle}(\theta)$ at $\theta=i\frac{\pi}{6}(b-1)$ and can be explained in the full range $b\in[1,2]$ by the first diagram on Figure 12, if we replace $\vert1+\rangle$ by $\vert2+\rangle$.
The pole labeled by $6$ on Figure 13 is in $T_{-}^{\vert2+\rangle}(\theta)$ at $\theta=i\frac{\pi}{6}(3-b)$ and can be explained by a diagram similar to the third one of Figure 12 in which the $\vert1+\rangle$ state is replaced by $\vert2+\rangle$ and the vacuum $\vert0\rangle$ is replaced by $\vert1+\rangle$.
{height="5cm"}{height="5cm"}
The pole labeled by $7$ on Figure 13 is a second order one in $T_{+}^{\vert2+\rangle}(\theta)$ at $\theta=i\frac{\pi}{6}(b+1)$ and can be explained by the two diagrams on Figure 14. Clearly the transmission factor $T_{-}(\theta)$ does not have zeros neither at $\theta=i\frac{\pi}{6}(3-b)$ nor at $\theta=i\frac{\pi}{6}(b-1)$ so the pole is of second order.
By now we explained all the poles of all the transmission factors of the ground and excited defect states. We used either the creation of a new defect boundstate or presented the appropriate defect Coleman-Thun diagram which was responsible for the singularity. By finishing this procedure the spectrum become complete and we managed to define a sensible defect theory. It would be nice to check these findings by the defect truncated conformal space approach (TCSA).
Conclusions
===========
We have demonstrated how the fusion idea can be used to solve topological defects in the sinh-Gordon and Lee-Yang models. In the sinh-Gordon case we determined the transmission factors and the defect energy as a function of a bootstrap parameter whose relation to the Lagrangian was also given. We checked these results in perturbation theory and against the newly derived DTBA.
In the Lee-Yang case we determined the transmission factors together with the defect energy and checked them in DTBA. For certain range of the parameter the transmission factor admits poles in the physical strip. We closed the defect bootstrap programme: we explained all poles either by associating new defect boundstates or by giving the appropriate defect Coleman-Thun mechanism both for the groundstate and for excited defect states.
The relation obtained between the transmission parameter and that of the Lagrangian in the sinh-Gordon theory can be analytically continued to describe the analogues relation in the sine-Gordon theory. This result also passes the test of first order perturbation theory and together with the transmission factors obtained in [@KL; @sgdef] gives the complete solution of defect sine-Gordon model. We have checked this solution by performing the fusing procedure on the solitonic transmission factors. This is analogous to the dressing procedure in the XXZ spin chain developed in [@XXZ].
The perturbation theory developed here can also be used in higher rank affine Toda theories to connect the parameters of the transmission factor of the bootstrap solution [@QAT] to the parameters of their Lagrangians [@DAT].
The derivation of the DTBA generalizes to any diagonal scattering theory. A large and small volume analysis analogous to the one presented in the paper will provide the leading finite size correction to the groundstate energy and give the bulk/defect energies, respectively.
In the present paper we were concerned with the bootstrap (IR) description of our models. There is a need, however, to understand their UV behavior which probably can be described by perturbed defect CFTs. To connect these alternative descriptions we can use methods starting either from the IR side, like DTBA, or starting from the UV side, like defect TCSA. There are works in progress in both directions.
Acknowledgments {#acknowledgments .unnumbered}
---------------
ZB thanks Laszló Palla, Gábor Takács, Gerard Watts and Cristina Zambon for the useful discussions. ZB was supported by a Bolyai Scholarship, OTKA K60040 and the EC network “Superstring”.
[10]{} G. Delfino, G. Mussardo and P. Simonetti, *Statistical Models with a Line of Defect*, Phys.Lett. B **328** (1994) 123.
G. Delfino, G. Mussardo and P. Simonetti, *Scattering Theory and Correlation Functions in Statistical Models with a Line of Defect*, *Nucl. Phys. B* **432** (1994) 518.
O.A. Castro-Alvaredo, A. Fring and F. Göhmann, *On the absence of simultaneous reflection and transmission in integrable impurity systems*, hep-th/0201142.
V. Caudrelier, M. Mintchev, E. Ragoucy, P. Sorba, *Reflection-Transmission Quantum Yang-Baxter Equations* J.Phys. **A38** (2005) 3431-3442.
S. Ghoshal and A. Zamolodchikov, *Boundary S-Matrix and Boundary State in Two-Dimensional Integrable Quantum Field Theory*, Int.J.Mod.Phys. **A9** (1994) 3841-3886; Erratum-ibid. **A9** (1994) 4353.
Patrick Dorey, Roberto Tateo, Gerard Watts, *Generalisations of the Coleman-Thun mechanism and boundary reflection factors*, Phys. Lett. **B448** (1999) 249-256.
A. Fring, R. Köberle, *Factorized Scattering in the Presence of Reflecting Boundaries,* Nucl.Phys. **B421** (1994) 159-172.
Z. Bajnok, G. Böhm, G. Takács, *Boundary reduction formula*, J.Phys. **A35** (2002) 9333-9342.
Z. Bajnok, G. Böhm, G. Takács, *On perturbative quantum field theory with boundary,* Nucl.Phys. **B682** (2004) 585-617.
Valentina Riva, *Boundary effects in two-dimensional critical and off-critical systems,* hep-th/0106268.
E. Corrigan, A. Taormina, *Reflection factors and a two-parameter family of boundary bound states in the sinh-Gordon model*, J.Phys. **A33** (2000) 8739-8754.
P. Mattsson, P.E. Dorey, *Boundary spectrum in the sine-Gordon model with Dirichlet boundary conditions*, J.Phys. **A33** (2000) 9065-9094.
Z. Bajnok, L. Palla, G. Takács, G.Zs. Tóth, *The spectrum of boundary states in sine-Gordon model with integrable boundary conditions*, Nucl.Phys. **B622** (2002) 548-564
Gábor Zsolt Tóth, *Investigations in Two-Dimensional Quantum Field Theory by the Bootstrap and TCSA Methods*, arXiv:0707.0015.
E Corrigan, *On duality and reflection factors for the sinh-Gordon model,* Int.J.Mod.Phys. **A13** (1998) 2709-2722.
A. LeClair, G. Mussardo, H. Saleur, S. Skorik, *Boundary energy and boundary states in integrable quantum field theories,* Nucl.Phys. **B453** (1995) 581-618.
P. Bowcock, E. Corrigan, C. Zambon, *Classically integrable field theories with defects*, Int.J.Mod.Phys. **A19S2** (2004) 82-91.
P. Bowcock, E. Corrigan, C. Zambon, *Affine Toda field theories with defects,* JHEP **0401** (2004) 056.
J.F. Gomes, L.H. Ymai, A.H. Zimerman, *Classical Integrable Super sinh-Gordon equation with defects,* J.Phys. **A39** (2006) 7471-7483, *Classical Integrable $N=1$ and $N=2$ Super Sinh-Gordon Models with Jump Defects,* arXiv:0708.2407.
V. Caudrelier, *On a systematic approach to defects in classical integrable field theories*, arXiv:0704.2326.
Z. Bajnok, A. George, *From Defects to Boundaries,* Int.J.Mod.Phys. **A21** (2006) 1063-1078.
Robert Konik, Andre’ LeClair*, Purely Transmitting Defect Field Theories,* Nucl.Phys. **B538** (1999) 587-611.
P. Bowcock, E.Corrigan, C. Zambon, *Some aspects of jump-defects in the quantum sine-Gordon model*, JHEP **0508** (2005) 023.
E. Corrigan, C. Zambon, *On purely transmitting defects in affine Toda field theory*, JHEP **07** (2007) 001.
V.B. Petkova, J.-B. Zuber, *The many faces of Ocneanu cells,* Nucl.Phys. **B603** (2001) 449-496.
K. Graham, G. M. T. Watts, *Defect Lines and Boundary Flows,* JHEP **0404** (2004) 019.
Ingo Runkel, *Perturbed Defects and T-Systems in Conformal Field Theory*, arXiv:0711.0102.
Roger E. Behrend, Paul A. Pearce, *Integrable and Conformal Boundary Conditions for sl(2) A-D-E Lattice Models and Unitary Minimal Conformal Field Theories*, hep-th/0006094.
C. H. Otto Chui, Christian Mercat, Will Orrick, Paul A. Pearce, *Integrable Lattice Realizations of Conformal Twisted Boundary Conditions*, Phys.Lett. **B517** (2001) 429-435.
P. Bowcock, *Boundaries and defects*, 3rd EUCLID Network School, 2006 Trieste, Italy.
Al. Zamolodchikov, *unpublished,* Z. Bajnok, L. Palla, G. Takacs, *Finite size effects in boundary sine-Gordon theory*, Nucl.Phys. **B622** (2002) 565-592
Al. Zamolodchikov, *Mass scale in sin-Gordon and its reductions*, Int. J. Mod. Phys. **A10** (1995) 1125–1150.
C. Destri and H. deVega, *New exact results in affine Toda field theories: Free energy and wave function renormalizations*, Nucl. Phys. **B358** (1991) 251–294.
S. Ghoshal, *Bound state boundary S matrix of the Sine-Gordon model*, Int. J. Mod. Phys. **A9** (1994) 4801–4810 [\[]{}Archive: hep-th/9310188\].
J. L. Cardy and G. Mussardo, *S matrix of the Lee-Yang edge singularity in two dimensions*, Phys. Lett. **B225** (1989) 275-278.
Patrick Dorey, Andrew Pocklington, Roberto Tateo, Gerard Watts, *TBA and TCSA with boundaries and excited states*, Nucl.Phys. **B525** (1998) 641-663, hep-th/9712197.
P. Dorey, M. Pillin, R. Tateo, G.M.T. Watts, *One-point functions in perturbed boundary conformal field theories,* Nucl.Phys. **B594** (2001) 625-659, hep-th/0007077.
Thomas Quella, Ingo Runkel, Gerard M.T. Watts, *Reflection and Transmission for Conformal Defects,* hep-th/0611296.
Patrick Dorey, *Exact S-matrices*, hep-th/9810026.
Patrick Dorey, Davide Fioravanti, Chaiho Rim, Roberto Tateo, *Integrable quantum field theory with boundaries: the exact g-function*, Nucl.Phys. **B696** (2004) 445-467.
Olalla Castro-Alvaredo, Andreas Fring, *From integrability to conductance, impurity systems,* Nucl.Phys. **B649** (2003) 449-490.
F. Woynarovich, *O(1) contribution of saddle point fluctuations to the free energy of Bethe Ansatz systems,* Nucl.Phys. **B700** (2004) 331-360.
Al.B. Zamolodchikov, *Thermodynamic Bethe Ansatz In Relativistic Models. Scaling Three State Potts And Lee-Yang Models,* Nucl.Phys. **B342** (1990) 695-720.
Z. Bajnok, L. Palla, G. Takács, *Finite size effects in quantum field theories with boundary from scattering data,* Nucl.Phys. **B716** (2005) 519-542.
Z. Bajnok, Chaiho Rim, Al. Zamolodchikov, *Sinh-Gordon Boundary TBA and Boundary Liouville Reflection Amplitude,* arXiv:0710.4789.
Z. Bajnok*, Equivalences between spin models induced by defects*, J.Stat.Mech. **0606** (2006) P010.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Surface level instability when tube is injected into vibrating bed of powder, which was originally found in experiments, is investigated numerically. We find that thicker (thiner) tube makes surface level inside tube higher (lower) than surface level outside tube. With fixed acceleration amplitude of vibration, surface level inside tube becomes higher as amplitude of vibration increases, which can be explained by considering the dependence upon strength of convective flow.'
address: |
c/o Dr. Y-h. Taguchi, Department of Physics, Tokyo Institute of Technology\
Oh-okayama, Meguro-ku, Tokyo 152, Japan
author:
- 'Yasushi Maeno[^1]'
title: Numerical investigation of surface level instability due to tube in vibrating bed of powder
---
Introduction
============
The dynamics of granular material attracted many attentions of physicists[@review]. Among them, vibrating bed of powder was studied by Faraday half and a century ago[@Faraday], and many papers were written about its behavior. This is because vibrating bed can exhibit many interesting phenomena like convection[@conv], surface heaping[@heap], surface fluidization[@surface], size segregation[@seg], and turbulence[@turb]. Other than those phenomena, surface level instability due to injection of tube into vibrating bed of powder was observed in experiments[@akiyama]. The main purpose of this paper is to reproduce this phenomenon numerically.
The organization of this paper is as follows. In Sec. \[sec:exp\], we briefly summarized experimental findings by Akiyama and Shimomura. Numerical modeling used in this paper is explained in Sec. \[sec:num\] and results will be presented in Sec. \[sec:res\]. Summary and discussion can be found in Sec. \[sec:sum\].
Experiments {#sec:exp}
===========
Vibrating bed of powder is a vessel filled with granular matter, typically mono-disperse glass beads, and vessel is shaken vertically as strong as gravity acceleration. When acceleration amplitude of vibration exceeds critical value, which is usually a little bit larger than gravity acceleration, the bed exhibits several instabilities, e.g., surface heaping[@heap], surface fluidization[@surface], and convection[@conv]. Further increase of acceleration amplitude results in disappearance of heap, and surface starts to fluctuate violently. In order to measure shear friction in this vibrating bed, Akiyama and Shimomura[@akiyama] injected thick tube into vibrating bed and found vary surprising effect. With tube fixed in space, i.e., tube does not vibrate, the surface level inside tube differs from surface level outside tube (Fig. \[fig:schem\]), even if heaping is not observed without tube. Thus, this instability may be different from heaping instability observed in vibrating bed. The surface level difference depends upon several physical parameters as diameter of tube, acceleration amplitude, and particle diameter. It was rather difficult to understand these dependences because experiment of powder can be affected by many other fine differences of conditions, e.g., moisture, temperature, and so on. Thus, numerical investigation is much suitable to understand these dependences in detail.
Numerical model {#sec:num}
===============
Although there are many numerical schemes for investigating dynamics of powder, we employ here distinct element method (DEM)[@cundal] to reproduce Akiyama’s experiments. In DEM, granular particle is models as visco-elastic particle, whose interaction is limited within short range. Particularly we employ here non-spherical models introduced by Pöschel and Buchholtz[@PRL.Poeschel], which is known to reproduce static friction effect better than conventional models. In their non-spherical model, each granular particle is modeled as a set of five sub-particles (See Fig. \[fig:non-spherical\]). Although tangential force is ignored, static friction can be considered as interaction among surrounding four small sub-particles. Since relative position between center sub-particle and surrounding small sub-particles are considered, rotation of non-spherical particle can be considered effectively. Each sub-particle obeys the following equation $$m_i \frac{d^2 \r_i}{dt^2} = \sum_j \F^{ij}_{inner} +
\sum_j\F^{ij}_{outer} \times \Theta ( \ell'_{ij}-\mid \r_j-\r_i\mid)$$ where $F^{ij}_{inner}$ and $F^{ij}_{outer}$ are interaction within a non-spherical particle and interaction between non-spherical particles respectively. Here subscript $i$ is $c$ or $r$ depending upon whether $i$ sub-particle is center (c) sub-particle or surrounding (r) sub-particle. $m_i$ is mass of $i$ sub-particle and $\r_i$ is position vector of $i$ sub-particle. $\Theta$ is step function, and $\ell'_{ij}$ is distance between sub-particles which belong to different non-spherical particles, $$\l'_{ij}= \left \{
\begin{array}{lc}
2R_r & (i,j=r,r)\\
2R_c & (i,j=c,c)\\
R_c+R_r & (i,j=c,r)\\
\end{array}
\right.$$ where $R_i$ is radius of $i$ sub-particle. $\F^{ij}_{inner}$ is defined as, $$\F^{ij}_{inner} = \left [ -k
(\mid \r_i-\r-j\mid-\ell_{ij})
- - -\gamma \left( \frac{d\r_j}{dt}-\frac{d\r_i}{dt} \right )
\cdot \n_{ij} \right ]\n_{ij}$$ with $\n_{ij} \equiv (\r_j-\r_i)/\mid \r_j-\r_i\mid$. $k$ and $\gamma$ are elastic and viscosity constant respectively, which are related to coefficient of restitution, $e = \exp \left( - \frac{ \pi \gamma}
{2 \sqrt{k-(\gamma/2)^2}} \right )$. $\ell_{ij}$ is distance between sub-particles among a non-spherical particle, $$\ell_{ij} = \left \{
\begin{array}{lc}
R_c + R_r & (i,j =c,r) \\
\sqrt{2} (R_c + R_r) & (i,j =r,r) \\
\end{array}
\right.$$ This means that five sub-particles have interaction only when their relative positions deviate from that shown in Fig.\[fig:non-spherical\]. $\F_{outer}^{ij}$ is defined as $$\F_{outer}^{ij} = \left [
- - -k ( \mid \r_i -\r_j \mid -\ell'_{ij} )
- - -\gamma \left ( \frac{d \r_j}{dt} -\frac{d\r_i}{dt}
\right ) \cdot \n_{ij} \right ] \n_{ij}$$
In order to simulate vibrating bed of powder, we have to introduce vessel composed of sub-particles. Interaction between sub-particles which construct non-spherical particles and sub-particles which construct vessels employs the same functional form as $\F^{ij}_{outer}$, but $\ell'_{ij}$ is replaced with $$\l'_{ij}= \left \{
\begin{array}{lc}
R_c+R_v & (i,j=c,v)\\
R_r+R_v & (i,j=r,v)\\
\end{array}
\right.$$ Motion of sub-particles which construct vessel is not affected by the interaction with sub-particles which constructs non-spherical particle, but follows given motion of vessel itself, e.g., vibration or static state. In addition to this, ’solid plate’ is added to the bottom to prevent particles from falling out of the vessel through bottom. This plate causes vertical force applied to $i$th sub-particle, $$F^i_{bottom}=k'[z_{bottom}-z_i],$$ where $z_{bottom}$ and $z_i$ are vertical components of bottom and $i$th sub-particle, respectively.
The results {#sec:res}
===========
In the simulations, we employ following parameters (See Fig.\[fig:schem.num\]). Radius of sub-particles are $R_c=3.0, R_r=0.5$, and $R_v=1.5$. We ignore distribution of radius and use identical non-spherical particles. Number of non-spherical particles are 200, this means, number of sub-particles is 1000. Interaction parameters are taken as $k/m_i=600.0$ and $\gamma/m_i=2.0$ independent of the kind of sub-particles. Thus coefficient of restitution $e=0.88$. $k'$ is take to be $2k$. For size of vessel, its diameter $D_v=90$ and its height $H_v=135$. Tube whose height $H_c=10.5$ is separated from bottom of vessel by $H_s=45$. Gravity acceleration is taken to be 9.8.
Control parameters of simulation are the acceleration amplitude of vibration $\Gamma$, angular frequency of vibration $\omega$, and diameter of tube, $D_c$. Here $\Gamma$ is related with $\omega$ as $\Gamma= a\omega^2/g$ where $a$ is amplitude of vibration. The values of these parameters used in simulations are, $\Gamma=5.0, 6.0,$ and $7.0$, $\omega=3.0, 4.0,$ and $5.0$, $D_c=30,24$, and $18$. Total length of simulation is over $40$ periods.
Figure \[fig:snapshot\] shows typical snapshot of simulation. In order to measure surface level differences, we define height of surface as follows. For example, surface height inside tube $z_i$ is taken such that number of non-spherical particles above $z_i$ is half as much non-spherical particles as aligned along section of tube. (For detail, see Fig. \[fig:height\].) The height $z_o$ outside tube is defined in similar way. Then height difference between inside and outside tube $\Delta z = z_i-z_o$. In addition to this, we measure flow of non-spherical particles within vessel following ref. [@conv], i.e., by dividing vessel into cells as large as non-spherical particle, and flow is defined as number of particle transport between cells per period of vibration.
Figure \[fig:del z\] shows the dependence of $\Delta z$ upon several parameters, and Figure \[fig:flow\] shows flow patterns and strength for several parameter values.
Summary and discussion {#sec:sum}
======================
In this paper, we numerically investigated height difference $\Delta z$ between inside and outside injected tube. $\Delta z$ increases as diameter $D_c$ of injected tube increases except for a few cases (See Fig. \[fig:del z\]). When accelerated amplitude of vibration $\Gamma$ and $D_c$ are fixed, $\Delta z$ decreases as angular frequency of vibration $\omega$ increases. Since convection becomes weak as $\omega$ increases except for a few cases (See Fig. \[fig:del z\]), $\Delta z$ can be regarded as a function of convection indirectly. These results are schematically shown in Fig.\[fig:summary\]. Upper raw corresponds to larger $\omega$ and lower raw to smaller. Left column shows the results for smaller $D_c$ and right column shows those for larger. Thus, in upper left case, $\Delta z$ takes minimum, and in lower right case, $\Delta z$ takes maximum. Also convection which occurs when $\omega$ is small enough is drawn in lower raw.
Possible qualitative explanation of these results are as follows. Granular material inside tube is affected by both upward and downward forces. Upward force is due to convection. Convection is upward at the center of vessel, thus it push up granular material inside tube. Difference between upper raw and lower raw is the difference of strength of convection. Lower raw has stronger upward force due to convection, thus surface level inside tube is higher in lower raw than in upper raw.
Downward force is possibly due to difference of density between inside and outside tube. Generally, fluidized granular matter has smaller density than that of fixed bed. However, granular matter inside tube is hard to flow due to friction with tube, thus density is relatively high. This effect pushes down granular matter into tube. This conjecture explains why thiner tube has lower level of granular matter inside tube. Thiner tube causes higher friction which prevents granular matter from being fluidized, thus heavier granular matter is hard to rise by convection. It is schematically illustrated in Fig.\[fig:summary\], where left column has lower level of surface inside tube than right column. This conjecture should be confirmed in future.
Acknowledgement
===============
The author thanks Dr. Hiraku Nishimori at Ibaraki University for providing his experimental results before publication. Prof. J. Rajchenbach is also acknowledged for helpful discussion. The author also thanks Dr. Y-h. Taguchi for his translating Japanese version of this paper into English and for helpful discussions.
For general reviews about powder, see e.g., H. M. Jaeger and S. R. Nagel, Science [**255**]{}, 1523 (1992); H. Hayakawa, H. Nishimori, S. Sasa, and Y-h. Taguchi, Jpn. J. Appl. Phys., 34 (1995) 397; Y-h. Taguchi, H. Hayakawa, S. Sasa, and H. Nishimori eds., [*Dynamics of Powder Systems*]{}, Int. J. Mod. Phys. B [**7**]{} Nos. 9 & 10 (1993); D. Bideau and A. Hansen eds., [*Disorder and Granular Media*]{}, (North-Holland, Amsterdam, 1993); A. Mehta ed., [*Granular Matter*]{}, (Springer, Berlin, 1993); C. Thornton ed., [*Powders and Grains ’93*]{} (A.A.Balkema Publishers, Rotterdam, 1993). M. Faraday, Philos. Trans. R. Soc. London 52 (1831) 299. Y-h. Taguchi, Phys. Rev. Lett. 69 (1992) 1367. J. A. C. Gallas, H. J. Herrmann, S. Sokołowsky, Phys. Rev. Lett. 69 (1992) 1371; P. Evesque and J. Rajchenbach, Phys. Rev. Lett. 62 (1989) 44; C. Laroche, S. Douady, and S. Fauve, J. Phys. (France) 50 (1989) 44. P. Evesque, E. Sznatula, and J.-P. Dennis, Europhys. Lett. 12 (1990) 623. J. B. Knight, H. M. Jaeger, and S. R. Nagel, Phys. Rev. Lett. 70 (1993) 3728; J. Duran, J. Rajchenbach, and E. Clement, Phys. Rev. Lett. 70 (1993) 2431; A. D. Rosato, F. Prinz, K. J. Stanburg, and R. H. Swendsen, Phys. Rev. Lett. 58 (1987) 1038. Y-h. Taguchi, J. Phys. (France) II 2 (1992) 2103; Europhys. Lett. 24 (1993) 203; Fractals 1 (1993) 1080;Physica D 80 (1995) 61. T. Akiyama and T. Shimomura, Powder Technol. 66 (1991) 243; Adv. Powder Technol. 4 (1993) 129. P. A. Cundal and O. D. L. Strack, Geotechnique 29-1 (1979) 47. T. Pöshcel and V. Buchholtz, Phys. Rev. Lett. 71 (1993) 3963; V. Buchholtz and T. Pöshcel, Physica A 202 (1994) 390.
[^1]: All correspondence should be sent to Y-h. Taguchi, at the same address as above. Electronic Address: [email protected]
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Experiments on the violation of equivalence principle (EP) and solar system give a number of constraints in which any modified gravity model must satisfy them. We study these constraints on a kind of $f(R)$ gravity as $f(R) = R(1\pm \epsilon \ln({R \over R_c}))$. For this investigation we use of chameleon mechanism and show that a spherically body has thin-shell in this model. So that we obtain an effective coupling of the fifth force which is suppressed through a chameleon mechanism. Also, we obtain $\gamma_{PPN} = 1 \pm 1.13 \times10^{-5}$ which is agreement with experiment results. At last, we show that for $R_c \thickapprox \rho_c$ this model is consistent with EP, thin shell condition and fifth force of chameleon mechanism for $\epsilon \backsimeq 10^{-14}$.'
author:
- '$^a$Kh. Saaidi'
- '$^b$A. Aghamohammadi'
title: 'Equivalence Principle (EP) and Solar System Constraints on $R(1\pm \epsilon \ln({R \over R_c}))$ model of Gravity'
---
Introductions
=============
In the recent decade, the acceleration of the universe expansion was discovered and is still a deep mystery, (for review see e.g ([@Brax], [@3], [@4],[@5] )). Two approaches have been introduced for interpret the present acceleration. One can either introduce an unknown form of energy, dubbed dark energy, that suggest about %70 of energy density of the present universe is composed by it or modify the behavior of gravity at cosmological distances. In the first approach, the most relevant candidate for the role of dark energy is Einstein’s cosmological constant, which interpret the cosmic expansion in the $\Lambda CDM$ model ([@Bisabr], [@7], [@8], [@9]) but in order to overcome its intrinsic shortcomings associated with the energy scale, several alternative models such as quintessence, etc have been proposed ([@10]). Most of these models have the common feature to introduce new sources in to the cosmological dynamics, but these models have the cosmological constant problem, the coincidence problem and the value of equation of state. On the other hand, in the second approach, various attempts to modify gravity have been presented([@11],[@12] ,[@14],[@15]), ([@18], [@19], [@20]) ([@21],[@Ali]), ([@16]). Among different approaches, there are modified gravity models so called $f(R)$, namely (we replace the Ricci scalar $R$ in the Einstein-Hilbert lagrangian density to some function $f(R)$), which do not seem to introduce new type of matter and can lead to late time acceleration. In fact, these theories can be reformulated in terms of scalar tensor theories with a stabilized coupling of the additional scalar degree of freedom to matter. As theories of dark energy, they suffer from the usual problems and are also potentially ruled out by gravitational tests of Newton’s law. Among all cosmologically viable $f(R)$ theories there is still an important issue to be pursued, they must be probed at solar system scale. This claim was based on the fact that $f(R)$ theories are equivalent to Brans-Dicke theory $\omega=0$, while observations set the constraint $\omega>40000$ ([@22]), and the post-Newtonian parameter satisfies $\gamma_{PPN}=\frac{1}{2}$ instead of being equal to unity as needed by observations([@a1], [@a2], [@Bisabr]). The only way-out for these models is that behave as chameleon theories ([@Khoury]), i.e. evolve a field mass dependent on the local matter density ([@24], [@26]) ([@25]). The scalar field mediates a fifth force which is suppressed in the laboratory and in interactions between large bodies such as planets, but which may be detectable between small test masses in space ([@27]). In the massive bodies, the fifth force is attenuated as the chameleon is trapped inside very massive bodies (the sun for instance) ([@10]). It has argued that the existence of thin shell is usually adequate to salvage $f(R)$ gravity models ([@27]). Meanwhile, in Ref. ([@16]) the authors derived the conditions under which a successful sequence of radiation, matter and accelerated epochs can be realized. In addition the stability conditions $f_{,R}>0$ and $f_{,R,R}>0$ are required to avoid ghosts and tachyon for $R\geq R_1$, where $R_1$ is the Ricci scalar de-Sitter point([@28]). There are viable $f(R)$ models that can satisfy both cosmological constraints and stability conditions([@su1], [@29], [@30], [@31], [@32]).\
In this work we have used chameleon mechanism to place constraints as a local experiments on a viable $f(R)$ gravity model, that in the our previous work ([@1]), have been proposed. Whereas some proposal of $f(R)$ model of gravity could not consistent with astrophysical and experimental data, we want to find a model of $f(R)$ gravity which has been the most agreement with experimental data.\
The paper is organized as follows. In first section, we introduce a $f(R)$ model, then we apply the local gravity experiment constraints, as thin shell condition, equivalence principle, fifth force on viability of the model. Finally, the latter section is devoted to conclusions.
Primary Calculations
====================
In a static and spherically symmetric space time, with the metric $$\bar{g}_{\mu\nu}=diag(1,-1,-\bar{r}^2,-\bar{r}^2\sin ^2\theta),$$ the field equation ([@Brax]) gives: $$\begin{aligned}
\label{9}
\frac{d^2 \phi}{{d\bar{r}}^2}+\frac{2}{\bar{r}}\frac{d\phi}{d\bar{r}}=\frac{dV_{eff}}{d\phi},\end{aligned}$$ where $\bar{r}$ is the distance from the center of symmetry and $$\begin{aligned}
V_{eff}(\phi)=V(\phi)+e^{\beta \phi}\bar{\rho}.\label{10}\end{aligned}$$ Here $\bar{\rho}$ is the energy density in the Einstein frame, which is connected to the energy density $\rho$ in the Jordan frame via the relation $\bar{\rho}=e^{3\beta\phi}\rho$. We assume that a spherically symmetric body with radius $\bar{r_c}$ has a constant energy density $\bar{\rho}_{in}$ inside body ($\bar{r}<\bar{r}_c$) and the energy density $\bar{\rho}_{out}$ outside the body ($\bar{r}>\bar{r}_c$). The mass $M_c$ and the gravitational potential $\Phi_c$ of the body with radius $\bar{r}_c$ are given by $M_c=\frac{4\pi}{3}\bar{r}^3_c\bar{\rho}$ and $\Phi_c=\frac{M_c}{8\pi\bar{r}_c}$, respectively. For obtain $V_{eff}$, originally, we introduce a $f(R)$ model that can satisfy local gravity constraints as well as cosmological and stability condition which we have proposed that in the our previous work as follows $$\label{11}
f(R)=R+R\ln {[\frac{R}{R_{c}}]}^{\mp\epsilon},$$ where $R_c$ is positive constants and $\epsilon$ is a small and dimensionless constant. It is clear $f(R)\vert_{R=0}=0$, on the flat space time and in the $\epsilon\ll 1 $, Eq. (\[11\]) reduced to $$\label{12}
f(R)\backsimeq R({\frac{R}{R_c}})^{\mp\epsilon}.$$ Also, in ([@Brax]) have been given that $$\begin{aligned}
V(\phi)=\frac{Rf'(R)-f(R)}{2{f'(R)}^2}.\label{5}\end{aligned}$$ Then, with substituting (\[12\]) into (\[5\]) the function $V_{eff}(\phi)$, will become $$\label{13}
V_{eff}=\mp\frac{\epsilon R_c}{2}e^{2\beta\phi(1\pm\frac{1}{\epsilon})}+\bar{\rho}e^{\beta\phi}.$$ It is obviously seen that the potential $$V(\phi)=\mp\frac{\epsilon R_c}{2}e^{2\beta\phi(1\pm\frac{1}{\epsilon})}$$ satisfy the conditions of chameleon mechanism $$\label{8a}
\frac{dV}{d\phi}<0, \qquad \frac{d^2 V}{{d \phi}^2}>0,\qquad \frac{d^3 V}{d^3\phi}<0.$$ Assuming $\phi\ll\epsilon\ll 1$, one can find the solution of $V'_{eff}=0$, with hypothesis $\bar{\rho}\ll \frac{2R_c}{\epsilon}$ as $$\label{14}
\phi_{min}=[\frac{-\bar{\rho}+R_c}{2\beta R_c}] \epsilon.$$ The effective potential $V_{eff}$ has two minima at the field value $\phi_{in}$ and $\phi_{out}$ according to (\[14\]), established upon $\bar{\rho}_{in}$ and $\bar{\rho}_{out}$, respectively. Here, the $\phi_{in}$ correspond to the region with a high density that gives rise to a heavy mass squared, whereas the $\phi_{out}$ to the lower density region with a lighter mass. Generally, the masses of scalar fields about these minima are given by $m_{in}^2=\frac{d^2V_{eff}(\phi_{in})}{{d\phi}^2}$ and $m_{out}^2=\frac{d^2V_{eff}(\phi_{out})}{{d\phi}^2}$. In this regard, the near of massive bodies with a heavy field mass, it is known that the spherically symmetric body has a thin-shell under the chameleon mechanism, i.e. $$\label{15}
\frac{\Delta \bar{r}_c}{\bar{r}_c}=\frac{\phi_{out}-\phi_{in}}{6\beta \phi_c}\ll 1.$$ Solution, equation (\[9\]) with appropriate boundary conditions $\phi_{out}=\phi(r=\infty)$ gives the field profile of the body($\bar{r}>\bar{r}_c$) $$\label{16}
\phi(\bar{r})\simeq -\frac{\beta}{4\pi}\frac{3\Delta \bar{r}_c}{\bar{r}_c}\frac{M_ce^{-m_{out}(\bar{r}-\bar{r}_c)}}{\bar{r}}+\phi_{out}.$$
Thin-Shell Condition
====================
In the chameleon mechanism, the chameleon field is trapped inside massive bodies and its influence on the other bodies is alone caused by a thin-shell close to the surface of the body ([@Khoury]). Hence, by substituting (\[14\]) and (\[16\]) into the thin-shell condition,(\[15\]), we have $$\label{17}
\frac{\Delta \bar{r}_c}{\bar{r}_c}=[\frac{\bar{\rho}_{in}-\bar{\rho}_{out}}{12{\beta}^2\Phi_c R_c}]\epsilon.$$ Where $\bar\rho_{in}$ and $\bar\rho_{out}$ are energy densities inside and outside of the body in the Jordan frame. Note that $R_c$ is not very different from $\rho_{in}$. Since $ \rho_{out}\ll \rho_{in}\ll \frac{\Phi_c R_c}{\epsilon}$, also $m_{out}r\ll 1$, means that the Compone wavelength $m_{out}^{-1}$ is very larger than Solar System scales, then in the relation $\bar{r}=e^{-2\beta \phi}r $ we can apply $\bar{r}\simeq (1-2\beta\phi+\cdots)\simeq r$. In the following we eliminate the bar upon the quantity $r$. We will first discuss post-Newtonian solar- system constraints on the model (\[11\]) or (\[12\]). In the weak-field approximation the spherically symmetric in the Jordan frame is as $$\label{18}
ds^2=[1+2a(r)]dt^2-[1+2b(r)]^{-1}dr^2-r^2d\Omega.$$ Where $a(r)$ and $b(r)$ are the functions of $r$. It was shown in ([@19]) that under the chameleon mechanism the post-Newton parameter, $\gamma_{PPN}=\frac{ b(r)}{a(r)}$, is approximately given by $$\label{19}
\gamma_{PPN}=\frac{3-\frac{\Delta r_c}{r_c}}{3+\frac{\Delta r_c}{r_c}}\simeq 1-\frac{2}{3}\frac{\Delta r_c}{r_c},$$ presuming that the condition $m_{out}r\ll 1$ holds on the solar-system. At the moment, if we apply (\[17\]) on the Earth and obtain the condition the thin-shell on the it, then to presumption that the Earth is a solid sphere with radius $R_e=6.37\times 10^3km$ and a mean density $\rho_e=5.52g/cm^3$ also is surrounded by an atmosphere with homogenous density $\rho_{atm}\simeq10^{-3}g/cm^3$ and $r_{atm}\simeq r_e$ as well as with the aid of the present tightest constraint on $\gamma_{PPN}$, $\vert \gamma_{PPN}-1\vert<2.3\times 10^{-5}$ ([@20]), we obtain $$\label{20}
\frac{\Delta r_c}{r_c}<3.45\times 10^{-5}.$$ By substituting (\[17\]) into (\[20\]) and replace the dimensionless Earth potential $\Phi_c=\Phi_e=\frac{M_eG}{R_ec^2}=6.95\times 10^{-10}$ ([@34]), $$\label{21}
\epsilon<4.79\times 10^{-14}\frac{R_c}{\rho_{in}}$$ By set $\rho_{in}=\rho_e=\frac{6\Phi_e}{r^2_e}=1.02\times 10^{-26}{cm}^{-2}$, the equation (\[21\]) gives $\epsilon<4.79\times10^{12}R_c$, so taking $Rc\sim \rho_{in}$, we have $\epsilon<4.79\times 10^{-14}$. It is observed that the deviation from the general relativity is very small. It is notable that this model of $f(R)$ gravity has been studied in ([@35]), for a weak field limit. In this regard, the line element was obtained as for $f(R)=R(\frac{R}{R_c})^{\mp \epsilon}$ model from $f(R)$ gravity ([@1]), given as $$\label{211}
ds^2= (1-\frac{2M}{r}\mp2\epsilon \ln(r))dt^2-\frac{dr^2}{(1-\frac{2M}{r}\mp2\epsilon)}-r^2d\Omega$$ This shows that $$\begin{aligned}
a&=&-\frac{M}{r}\mp \epsilon \ln (r),\label{212}\\
b&=&-\frac{M}{r}\mp \epsilon. \label{213}\end{aligned}$$ One can obtain $\gamma_{PPN},$ up to second order of $\epsilon$ as $$\label{124}
\gamma_{PPN}=\frac{b}{a}\simeq1 \pm\frac{\epsilon r}{M}[1-\ln(r)]+{\cal O}({\epsilon}^2).$$ Whereas, the most distance in the solar system is around $5.5\times 10^{12}m,$ so that $\frac{M}{r}=\frac{GM}{rc^2}\simeq 2.5\times 10^{-8}$. Therefore one can obtain $\gamma_{PPN}\simeq1\mp 1.13\times 10^{-5}.$ This measure of $\gamma_{PPN}$, was obtained in the Jordan frame of $f(R)$ gravity is agreement with the experimental consequence of $\gamma_{PPN}.$
Equivalence Principle
=====================
Now, we will consider experimental bounds from a probable violation of the equivalence principle(EP). For this doing, we suppose the Earth and its atmosphere to be an isolated body far away from the effect of the other objets such as Moon, the Sun and the next planets. The same as the previous section that specified, the Earth has a mean density $\rho_e\simeq5.5 g/{cm}^3$ with radius $r_e=6.37\times 10^3km$ and the atmosphere density $\rho_{atm}\simeq 10^{-3}g/{cm}^3$. The region outside the atmosphere($r>r_{atm}$) has a homogeneous density $\rho_G\simeq10^{-24}g/{cm}^3$ ([@10]). Clearly, we know the gravitational potentials inside a spherically symmetric body, such as the Earth and the atmosphere as $\Phi_e\propto \rho_e r^2_e$, $\Phi_{atm}\propto \rho_{atm}r^2_{atm}$ respectively. As a result, we have $\Phi_e \simeq 5.5\times 10^3\Phi_{atm}$ with assume $r_{atm}\simeq r_e $. From Eq. (\[15\]), for the field values $\Phi_{atm}$ and $\Phi_G$ correspond with the regions $r_e<r<r_{atm}$ and $r>r_{atm} $ at their local minima of the effective potential respectively we have $\frac{\Delta r_{atm}}{r_{atm}}=\frac{(\Phi_G-\Phi_{atm})}{6\beta\Phi_{atm}}<1 .5\times 10^{-2 }$. Where $\Delta r_{atm}=10^2km$, i.e. (when the atmosphere has a thin-shell then the thickness of the shell ($\Delta r_{atm}$ ) is smaller than the atmosphere $r_s=10-10^2km$) and $r_{atm}=6.5\times 10^3km$. Therefore, $\frac{\Delta r_e}{r_e}=\frac{(\Phi_{atm}-\Phi_{e})}{6\beta\phi_e}\simeq 2\times 10^{-4}\frac{\Delta r_{atm}}{r_{atm}}$. Where, we have used from $\Phi_e=5.5\times10^{3}\Phi_{atm}$. As a result, the condition have a thin-shell for the atmosphere as $$\label{22}
\frac{\Delta r_e}{r_e}<3\times 10^{-6}.$$ By making use of the EP test, we want to measure the difference of the free-fall acceleration of the Moon and the Earth toward the Sun. The constraint on the difference of two accelerations is given by ([@10]) $$\label{23}
\eta\equiv2\frac{\mid a_{Moon}-a_e\mid}{a_{Moon}+a_e}<10^{-13}.$$ The acceleration induced by fifth force with the field profile $\phi(r)$ and the effective coupling $\beta_{eff}$ is $ a^{fifth}=\mid\beta_{eff}\phi(r)\mid$. Then the acceleration $a_e$ and $a_{Moon}$ are ([@Khoury]) $$\begin{aligned}
a_e&\simeq&\frac{GM_\odot}{r^2}\left[ 1+3{(\frac{\Delta r_e}{r_e})}^2\frac{\Phi_e}{\Phi_\odot}\right] \label{24}\\
a_{Moon}&\simeq&\frac{GM_\odot}{r^2}\left[ 1+3{(\frac{\Delta r_e}{r_e})}^2\frac{{\Phi_e}^2}{\Phi_\odot\Phi_{Moon}}\right] \label{25}\end{aligned}$$ Substituting the dimensionless potentials, $\Phi_\odot\simeq 2.1\times 10^{-6},\Phi_e\simeq6.95\times 10^{-10}$ and $\Phi_{Moon}\simeq 3.1\times 10^{-11}$ into the Eq (\[24\], \[25\]) and combine those with Eq (\[23\]) gives $$\begin{aligned}
\frac{\Delta r_e}{r_e}<2.1\times10^{-6}.\label{26}\end{aligned}$$ Which is the same order of the condition (\[22\]) as in the thin-shell condition for the atmosphere. Taking Eq. (\[26\]) as the constraint of violation of EP and combining with (\[17\]) one can obtain $$\begin{aligned}
\epsilon<\frac{0.3R_c}{\rho_{in}}\times10^{-14},\label{27}\end{aligned}$$ which is not different from (\[21\])
Fifth Force
===========
We want to consider the fifth force mediated by $\phi$. The dynamics of $\phi$ is still governed by the effective potential which is important when there is a component of matter whose energy-momentum tensor has nonzero trace. The solar system constrains get simply by making the large field mass associated with the field $\phi$, is given by $m^2_{min}=V_{eff,\phi\phi}(\phi_{min})$. The profile for a potential connected with a fifth force interaction is given by a Yukawa potential between two tests masses $m_1$ and $m_2$ by distance $r$ as $$\label{28}
V(r)=-\alpha\frac{m_1m_2}{8\pi}\frac
{e^{-m_{\phi}r}}{r}$$ Where $\alpha$ is strength of the interaction and $m_{\phi}^{-1}$ is the range. Therefore the fifth force experiment constrains regions of ( $\epsilon,m_{\phi}^{-1}$) parameter space. The experiments are generally fulfilled in a vacuum chamber that the range of the interaction is the same order of the spatial dimension of the chamber, namely ${m_{\phi}}^{-1}\sim R_{vac}$. The tightest bound of the strength of the interaction is $\alpha<10^{-3} ($[@35]). We consider two identical bodies with radius, mass, uniform density $r_c,m_c,\rho_c$, respectively, in to the chamber. Assuming the thin-shell condition is satisfied by the two bodies, their field profile out side the bodies are given by ([@26]) $$\label{29}
\phi(r)=-\frac{\beta}{4\pi}\frac{3\Delta r_c}{r_c}\frac{m_ce^{-\frac
{r}{R_{vac}}}}{r}+\phi_{vac}.$$ Hence the corresponding potential energy of the interaction is $$\label{30}
V(r)=-2\beta ^2{(\frac{3\Delta r_c}{r_c})}^2\frac{{m_c}^2}{8\pi}\frac{e^{-r/R_{vac}}}{r}.$$ The bound on the strength of the interaction to be given as follows $$\label{31}
2{\beta}^2{(\frac{3\Delta r_c}{r_c})}^2<10^{-3}.$$ Writing Eq. (\[17\]) for every one of the test bodies, we get $$\label{32}
\frac{\Delta r_c}{r_c}\approx[ \frac{\rho_{c}-\rho_{vac}}{12{\beta}^2\phi_c R_c}]\epsilon$$ Here $\rho_{vac}$ is energy density of the vacuum inside the chamber. In the experiment performed in ([@35]), for a typical test body with mass $m_c\approx 40gr$ and radius $r_c\approx 1cm$ the density $\rho_c$ and the potential $\Phi_c$ have been obtained $9.5gr/{cm}^3,\, 10^{-27}$ respectively. Furthermore, the pressure in the vacuum chamber have been reported $3\times 10^{-8}mmHg$ which is equivalent to $\rho_{vac}\approx4.8\times 10^{-14}gr/{cm}^3$. Substituting the upon values in to (\[32\]) and combining with (\[31\]) result in the bound $$\label{33}
\epsilon <3.6\times 10^{-29}\frac{R_c}{\rho_c}.$$ Which if we replace $\rho_c=1.7\times 10^{-26}{cm}^{-2}$, with substitute in to Eq. (\[33\]) we can obtain $\epsilon<2\times10^{-3}R_c $
Conclusion
==========
In this paper we have constrained $f(R)$ model of gravities, using the renowned equivalence between these models and scaler tensor theories. Mostly, in this representation from $f(R)$ theory there is a strong coupling of the scalar field with the matter sector. We have used from the chameleon mechanism to suppress this coupling. We have shown that in order to the model (\[11\]) be consistent with the present local gravity experiments, the parameters $\epsilon$ and $R_c$ should satisfy the following conditions.
- The thin-shell condition of chameleon mechanism for model (\[11\]), is satisfied for $\epsilon\lesssim 4.79\times 10^{-14}\frac{R_c}{\rho_e}$
- This model is consistent with EP experiment for $\epsilon\lesssim 0.3 \times 10^{-14}\frac{R_c}{\rho_e}$
Whereas, the structure of this paper is perturbation state of general relativity, hence $\epsilon$ should be small. Therefore, we suggest $R_c$ must be the same order of $\rho_e = 10^{-26} cm^{-2}$, where with this value of $R_c$, we have found that the model (\[11\]) is consistent with the thin-shell and EP conditions for $\epsilon\lesssim 10^{-14}$, as well as it is consistent with fifth force. At the moment, as a result applying the above values from $\epsilon$, we obtain $\gamma_{PPN}\simeq1\mp1.13\times 10^{-5}$. That is remarkable that this value of $\gamma_{PPN}$ are nearly equal to what is required by observation. It should be implied that the above result is achieved under the assumption that $\rho_{out}\ll\rho_{in}\ll \frac{\Phi_c R_c}{\epsilon}$. Therefore, the studied model is a viable $f(R)$ model which satisfied EP and Solar System bounds and it has very small deviation from the general relativity.
A. Aghamohamadi and Kh. Saaidi, [*et al*]{}., . [**80**]{}, 065008 (2009). Int. J. Theor. Phys. 10.1007/s10773-0100250-4 (2010). A. Chiavassa, [*et al*]{}., Arxive: astro-ph/1012.5234 A. A. Starobinsky, JETP Lett. [**86**]{}, 157 (2007). B. Li and J. D. Barrow, , 084010 (2007). C. M. Will, Liv. Rev. Rel. [**9**]{}, 3 (2005). E. J. Copeland, M. Sami and S.Tsujikawa, Int. J. mod. phys. D[**15**]{}, 1753(2006). I. Navarro and K. Van Acoleyen, , 022(2007). J. D. Evans, L. M. H. Hall and P. Caillol, , 083514 (2008). J. K. Hoskins, R. D. Newman, R. Spero and J. Schultz, Ph ys. [**D 32**]{}, 3084 (1985). J. Khoury and A. Weltman, Phys. Rev. Lett. [**93**]{}, 1711o4 (2004); , 044026 (2004) L. Amendola, R. Gannouji, D. Polarski and S. Tsujikawa, , 083504 (2007). L. Amendola, D. Polarski and S. Tsujikawa, Int. J. Md. Phys. D [**16**]{}, 1555 (2007). L. Amendola and S. Tsujikawa, Phys. Lett. B[**660**]{}, 125 (2008). L. Pogosian and A. Silvestri, , 023503 (2008). P. Brax,C. van de Bruck. \[ESSNCE Collaboration\] arXiv 0806. 3415v2. Astrophys (2008). R. Bean, D. Bernat, L. Pogoosian, A. Silvestri and M. Trodden, Phys. Rev. D[**75** ]{}, 064020 (2007). R. Durrer and R. Maartens, Gen. Rel. Grav. [**40**]{} (2008). S. Capozziello and S. Tsujikawa. Phys.Rev.D[**77**]{}, 107501 (2008). S. A. Appleby and R. A. Battye, Phy. Lett. B[**654** ]{}, 7 (2007). S. Nojiri and S. D. Odintsov, Phys. Rev. D[**74**]{}, 086005 (2006) S. M. Carroll, Liv. Rel. [**4**]{}, 1 (2001) S. Weinberg, Rev. Mod. Phy. [**61**]{}, 1 (1989). S. A. Appleby and R. A. Battye, Phy. Lett. [**B 654**]{}, 7 (2007) S. Tsujikawa, Phy. Rev. D [**77**]{}, 023507 (2008) S. Tsujikawa, K. Uddin and R. Tavakol, , 043007 (2008). S. Weinberg, Gravitation and Cosmology, Johan Wiley and Sons, 1972 T. Padmanaban, Phys. Rept. [**380**]{}, 235 (2003) T. Zwitter, U. Mumari, Arxiv: astro-ph/0306019. T. Faulkner, M. Tegmark, E. F. Bunn and Y. Mao, , 063505 (2007). T. P. Waterhouse. astro-Phys/061181v1. (2006). W.Hu and I. Sawicki, , 064004 (2007). W. M. Wood-Vasey [*et al*]{}. \[ESSNCE Collaboration\], . [**666**]{}, 694 (2007). Y. Bisabr, Phys.lett. B[**683**]{}, 96 (2010). Y. Sobouti, ArXiv: astro-ph/0603302. Y. S. Song, H. Peiris and W. Hu, , 063517 (2007).
| {
"pile_set_name": "ArXiv"
} |
---
author:
- 'Gilbert Moultaka [^1]'
title: The dark matter as a light gravitino
---
Introduction
============
In some instances, the requirement that supersymmetric particle dark matter scenarios aught to be the simplest to handle cosmologically and the least model-dependent, seems occasionally to take over the more fundamental question for supersymmetry, namely the origin of supersymmetry breaking. Since there is to date no particularly compelling susy breaking mechanism/model to be preferred to all the others, one should also consider the dark matter issue from a particle physics standpoint which offers different classes of susy breaking mechanisms, irrespective of whether the ensuing cosmological context is “simple" or not.
Recent developments [@ISS], [@murayama] stressing the existence of metastable susy breaking vacua, have renewed the interest in gauge-mediated susy breaking (GMSB) scenarios opening new possibilities for the model-building [@GR99], and appear to be very interesting from the early Universe point of view as well [@AK]. On the other hand, the gravitational interactions which play a minor role for susy breaking in GMSB models remain physically relevant through the coupling to supergravity, at least in order to absorb the unphysical goldstino component, to adjust the cosmological constant to a small value and to avoid a massless R-axion. In such scenarios where supersymmetry breaking is triggered by non-perturbative dynamics of some (secluded) gauge sector and communicated to the MSSM by a messenger sector through perturbative gauge interactions, the susy breaking scale $\sqrt{F}$ and the mass scale $\Lambda$ of the secluded gauge sector can be well below the Planck scale. Moreover, if these two scales combine to trigger the electroweak symmetry breaking yielding $G_F^{-1/2} \sim (\alpha/4 \pi) k {F / \Lambda } $, where $G_F$ is Fermi’s constant (and $0< k \le 1$ measures the secludedness of the secluded sector), then the gravitino mass $m_{3/2} \simeq F/(\sqrt{3}
m_{\rm Pl}) \sim \left(4 \pi/ \alpha)(\Lambda /\sqrt{3} k m_{\rm Pl}\right)
G_F^{-1/2}$ where $m_{\rm Pl}$ is the reduced Planck mass, is expected to be very small (${\raisebox{-0.13cm}{~\shortstack{$<$ \\[-0.07cm] $\sim$}}~}{\cal O}(1)$ GeV) and is the lightest supersymmetric particle (LSP). The question then arises as to which particle can be a good candidate for the cold dark matter (CDM) in this case? To answer this question requires an unconventional treatment as compared to the Neutralino “vanilla” candidate or even to the heavy gravitino candidate in the context of gravity mediated susy breaking models. Indeed, in contrast with the latter where the hidden sector is typically too heavy to be produced at the end of inflation, the secluded and messenger sectors of GMSB provide stable particles that may be present in the early Universe for a sufficiently heavy reheat temperature $T_{RH}$. We consider hereafter such configurations assuming that only the messenger (including the spurion) sector can be produced and illustrate its relevance to the issue of the CDM.
Curing a Messenger Problem
==========================
The mass degeneracy within a supermultiplet of messenger fields is lifted by susy breaking leading to a lighter and a heavier scalar messengers with masses $M_{A \pm} =M_X(1 \pm {k F/M_X^2})^{1/2}$ and a fermionic partner with mass $M_X$ (where $F$ and $M_X$ are related to the dynamical scale $\Lambda$). Thus ${k F/M_X^2} < 1$. Moreover, one has to require ${k F/M_X} {\raisebox{-0.13cm}{~\shortstack{$<$ \\[-0.07cm] $\sim$}}~}10^5 \mbox{GeV}$ to ensure an MSSM spectrum ${\raisebox{-0.13cm}{~\shortstack{$<$ \\[-0.07cm] $\sim$}}~}{\cal O}(1) \mbox{TeV}$. One then expects typically
$M_X {\raisebox{-0.13cm}{~\shortstack{$>$ \\[-0.07cm] $\sim$}}~}10^5 \mbox{GeV}$. In GMSB models the lightest messenger particle (LMP) with mass $M_{-}$ is stable due to the conservation of a messenger quantum number. If present in the early Universe the messenger particles are thermalized through their gauge interactions with the thermal bath. The corresponding LMP relic density is calculable similarly to the case of Neutralino LSP albeit an extended particle content and couplings. However, it turns out to be typically too large to account for the CDM even in the most favorable case of the electrically neutral component of a $\mathbf{5}+\mathbf{\overline{5}}$ representation of $SU(5)$ where it is found to scale as $\Omega_M h^2 \simeq 10^5 \left({M_{-}/10^3 TeV}\right)^2$ with the LMP mass [@DGP]. The situation is even worse in the case of $SO(10)$ where the LMP is an MSSM singlet with a suppressed annihilation cross-section leading to a very large relic density. One possible cure to this messenger overcloser problem is to allow the LMP to decay to MSSM particles. This can be easily achieved by adding renormalizable but messenger number violating operators to the superpotential, however, such low-scale operators would tend to ruin the nice FCNC suppression of GMSB models. A more appealing approach is to insist on the messenger number conservation as a consequence of a discrete accidental symmetry at low energy and invoke the typical violation of such a (non gauge) symmetry by gravitational interactions [@ross] once the GMSB model is coupled to supergravity. The LMP decay would then occur via Planck mass suppressed operators in the Lagrangian, which can originate either directly from effective non-renormalizable operators in the Kähler or the superpotential, or indirectly after susy breaking through (holomorphic) renormalizable operators in the Kähler potential. In the latter case the suppression is controlled by the gravitino mass. In the case of $SU(5)$, an exhaustive study of these operators was carried out in [@JLM04] for messengers transforming as $\mathbf{5}+\mathbf{\overline{5}}$ or $\mathbf{10}+\mathbf{\overline{10}}$. In $SO(10)$ with one set of messengers transforming as $\mathbf{16} + \overline{\mathbf{16}}$, the LMP decay can be induced by non-renormalizable operators in the Kähler potential, e.g. $K \supset \mathbf{16}_F
\overline{\mathbf{16}}_M^\dag \mathbf{10}_H/m_{\rm Pl}$, or in the superpotential, e.g. $W \supset
\overline{\mathbf{16}}_M {\mathbf{16}}_F {\mathbf{16}}_F \mathbf{10}_H
\times m_{\rm Pl}^{-1}$, leading respectively to $2$- and $3$-body decays, where $ \mathbf{16}_M (\overline{\mathbf{16}}_M), \mathbf{16}_F$ and $\mathbf{10}_H$ denote respectively the messenger, the standard matter and the electroweak Higgs supermultiplets. We assume a typical decay width
$\Gamma_{\rm LMP} = (1/16\pi) f' M_X^3/m_{\rm Pl}^2$ where $f'$ parameterizes our ignorance of the couplings and possible further phase space suppression. For couplings of ${\cal O}(1)$, $f'\simeq 1 ( 3 \times 10^{-3})$ for $2$- ($3$-body) decays into essentially massless particles. On the one hand, such suppressed decays would not upset the FCNC suppression, and on the other, will turn out actually to be a blessing regarding a solution to the gravitino overproduction in the early universe, eventually allowing the gravitino to account for the [*cold*]{} dark matter in the context of GMSB models.
Gravitino abundance
===================
the cosmological set-up
-----------------------
In favorable parts of the parameter space, the LMP late decay into MSSM particles can release enough entropy to dilute the initial gravitino relic density down to a level which can account for the CDM in the Universe even for very high $T_{RH}$, [@FY02], [@JLM04], [@JLM05]. For this to work, though, the LMP should dominate the Universe energy density before it decays, and should decay after the gravitino has freezed-out from the thermal bath. The necessary conditions $T_d < T_{MD}, T^f_{3/2}$ \[where $T_d, T_{MD},
T^f_{3/2}$ denote respectively the LMP decay and messenger matter domination temperatures, and the gravitino freeze-out temperature\] is then determined by the particle properties and annihilation cross-section and decay width of the LMP, delineating the favorable parts of the parameter space. We have studied this scenario in detail for the case of $SU(5)$ [@JLM04] and $SO(10)$, [@JLM05], [@CKLM]. Here we concentrate on the latter case with one set of messengers transforming as $\mathbf{16} + \overline{\mathbf{16}}$. The entropy release $\Delta S \equiv S_{\rm after}/S_{\rm before}$, diluting the initial gravitino density, is determined by the temperatures before and after LMP decay and can be approximated to $T_{MD}/T_d$. $T_{MD}$ is given by the LMP yield and mass ($T_{MD}\simeq (4/3)M_{-} \times Y_{\rm LMP}$) and $T_d$ is determined by the LMP width ($\Gamma_{\rm LMP} \simeq H(T_d)$). The LMP yield $Y_{\rm LMP}$ is controlled by the LMP annihilation into MSSM particles which enters the corresponding Boltzmann equation. Since in our case the LMP is an $SU(5)$ singlet [@DGP], [@JLM05], this annihilation proceeds via loop effects of virtual messengers ($A_M, \psi_M$) and spurion ($S$) exchange. We consider here the leading annihilation cross-section into $2$ gluons, fig.1, and parameterize its thermal averaged as $\langle \sigma_{1{\rm loop}}v\rangle \sim f \times (\alpha_s/4\pi)^2
\kappa^4/s$ where $\kappa$ is the spurion-messenger coupling ($W \supset
\kappa \hat{S} \mathbf{16}_M\overline{\mathbf{16}}_M$), $\alpha_s$ the strong coupling constant, $\sqrt{s}$ the C.M. energy and $f$ a form factor depending on the internal masses and couplings. Neglecting the very heavy GUT sector contributions which typically decouple, one finds $$\begin{aligned}
&& f = \frac{32}{\pi} |(3/8) \bar{g}^2 + M_{A_{-}}^2 C_{-} +
( (3/4) \bar{g}^2 - 1)
M_{A_{+}}^2 C_{+} \nonumber \\ && + \; 2 M_X^2( M_{A_{-}}^2 C_{-} +
M_{A_{+}}^2 C_{+} + ( s - 4 M_X^2) C_X -1 ) \; D[s]|^2 \nonumber \end{aligned}$$
where $D[s]$ denotes the spurion propagator $(s - M_S^2 + i \Gamma_S M_S)^{-1}$, $C_{\pm, X}$ are $C_0$ functions scaling as $s^{-1}$, and $\bar{g}^2 \equiv 4 \pi \alpha_s/\kappa^2$. Since the messengers in the loops carry color charges, a substantial mass splitting occurs between the contributing $A_{-}$ states and the LMP due to RGE running from the GUT scale down to $Q^2=M_{-}^2$, as well as from genuine loop corrections [@DGP96], [@JLM05], leading typically to $M_{A_{-}}
\simeq 3 M_{-}$. Such effects, as well as the running of $\alpha_s$ should be taken into account for a precise determination of $Y_{LMP}$. On the other hand, the contribution of the scalar spurion depends on its mass and width. The spurion can be either heavier or lighter than the messenger sector. Here we consider only the former case, where the spurion decays at tree-level into pairs of messengers or at one-loop into MSSM particles.[^2] We find that the decay into MSSM particles dominates irrespective of the value of the coupling $\kappa$, the total width $\Gamma_S$ remaining small though ($\Gamma_S/M_S
\simeq (1 - 4) \times 10^{-3}$). A careful treatment (beyond the relative velocity expansion) of the thermally averaged annihilation cross-section is thus required close to this narrow resonance, i.e. typically when $M_{-} \simeq
M_S/2$ and assuming a non-relativistic decoupling of the LMP from the thermal bath.
relic gravitinos
----------------
When the necessary temperature conditions discussed in the previous section are met, the final gravitino relic density is given by $\Omega_{{}_{grav}} = \Omega_{{}_{grav}}^{th}/\Delta S +
\Omega_{{}_{grav}}^{{}^{Mess}} + \Omega_{{}_{grav}}^{{}^{NLSP}}$ where the last two contributions denote non-thermal production through late decays or scattering. One should also consider various cosmological constraints (hotness/warmness, BBN, species dilution, etc...). Let us illustrate the case first with some effective fixed values for $f$ and $f'$. This is shown in fig.2 taking $T_{RH} \simeq 10^{12}$ GeV, see also [@JLM04]. The red horizontal shading shows the theoretically excluded region where $ k > 1$; the other red shading indicates the region excluded by BBN constraints. If the spurion is heavier than the LMP, gravitino cold DM (green region) occurs for relatively light LMPs and $m_{3/2} \sim 1 \mbox{kev} - 10 \mbox{MeV}$. Note that without the LMP induced entropy dilution, the reheat temperature would have been constrained to be several orders of magnitude lower than $10^{12}$ GeV in order to avoid overcloser for the range of gravitino masses found here.
More generally, in the models of ref. [@GMSB] one finds [@JLM05]
$\Omega_{grav}h^2 \,\simeq\, 10^3
f^{0.8} \kappa^{3.2} f'^{1/2}\left({M_{-} \over 10^6\,{\rm
GeV}}\right)^{-0.3} \times \left({m_{3/2} \over 1\,{\rm MeV}}\right)$ for non-relativistic LMP freeze-out, putting the gravitino relic abundance in the ballpark of WMAP results, for $\kappa \sim {\cal O}(10^{-1})$ and typical ranges for $f$ and $f'$. Considering now the specific one-loop form-factor $f$ given in the previous section and assuming for the sake of the illustration a spurion much heavier than the messenger sector, we still find regions consistent with WMAP results for $\Omega h^2$ at the 99% confidence level, e.g. for $M_X = 10^6 - 10^8$ GeV, one has $1.1 \mbox{MeV} < m_{3/2} < 4$ MeV for a three body decay LMP, and $65 \mbox{keV} < m_{3/2} < 230$ keV, for a two body decay LMP.
For heavier spurions the annihilation into 2 gravitinos through tree-level gravitational interactions (see fig.1) becomes significant as its cross-section scales like $\langle \sigma v\rangle\,\simeq\, (1/ 24\pi) k^2 M_-^2/(m_{3/2} m_{\rm Pl})^2
$. It can dominate the 1-loop annihilation, eventually saturating the unitarity limit (the black dashed line in fig.2), thus disfavouring gravitino CDM solutions for $M_{-} {\raisebox{-0.13cm}{~\shortstack{$>$ \\[-0.07cm] $\sim$}}~}10^8$ GeV.
Conclusion
==========
Light gravitino can account for CDM irrespective of $T_{RH}$, making it a good DM candidate in GMSB: typically if $T_{RH} {\raisebox{-0.13cm}{~\shortstack{$<$ \\[-0.07cm] $\sim$}}~}10^5 \mbox{GeV}$ then the messengers are not produced and thermal gravitinos with $m_{3/2} {\raisebox{-0.13cm}{~\shortstack{$<$ \\[-0.07cm] $\sim$}}~}1 \mbox{MeV}$ provide the right CDM density, while for $T_{RH} {\raisebox{-0.13cm}{~\shortstack{$>$ \\[-0.07cm] $\sim$}}~}10^5 \mbox{GeV}$ the messenger can be present and should be unstable, thus providing a source of entropy production that can reduce a thermally overproduced gravitino to a cosmologically acceptable level. Moreover, various constraints (e.g. on $T_{RH}$, [@SP], or on the gravitino mass [@VLHMR]) simply do not apply in the scenarios we have illustrated, thus escaping possible tension with thermal leptogenesis. Finally, let us mention that such scenarios can potentially allow to avoid the recently studied cosmological gravitino problem due to inflaton decay [@ETY].
\[fig1\]
\[fig2\] {height=".3\textheight"}
[999]{} K. Intriligator, N. Seiberg and D. Shih *JHEP 0604:021,2006*.
see also, H. Murayama these proceedings.
for a review of the earlier GMSB model-building and phenomenology see for instance G.F. Giudice, R. Rattazzi, *Phys.Rept 322: 419, 1999*, and references therein.
S.A. Abel, C.-S. Chu, J. Jaeckel, V.V. Khoze, *JHEP 0701:089,2007*, *JHEP 0701:015,2007*; N.J. Craig, P.J. Fox, J.G. Wacker *PRD75:085006,2007*; W. Fischler, V. Kaplunovsky, C. Krishnan, L. Mannelli, M.A.C. Torres, *JHEP 0703:107,2007*.
S. Dimopoulos, G.F. Giudice, A. Pomarol, *PLB 389:37,1996*.
L.E. Ibanez, G.G. Ross, *NPB 368:3-37, 1992.*
K. Jedamzik, M. Lemoine, G. Moultaka, *PRD 73:043514,2006*; see also G. Moultaka, *Acta Phys.Polon.B38:645,2007*.
M. Fujii, T. Yanagida, *PLB 549:273,2002*; see also E. A. Baltz, H. Murayama, *JHEP 0305:067, 2003*.
M. Lemoine, G. Moultaka, K. Jedamzik, *PLB 645:222,2007*.
M. Capdequi-Peyranère, M. Kuroda, M. Lemoine, G. Moultaka, *to appear*.
S. Dimopoulos, G. F. Giudice, A. Pomarol, *PLB 389:37, 1996*; T. Hahn, R. Hempfling, *PLB 415:161, 1997*.
M. Dine, A.E. Nelson, *PRD 48:1277, 1993*; M. Dine, A.E. Nelson, Y. Shirman, *PRD 51:1362, 1995*; M. Dine [*et al.*]{} *PRD 53:2658, 1996*.
J. Pradler, F.D. Steffen, *PLB 648:224,2007*. see also F.D. Steffen, these proceedings.
M. Viel, J. Lesgourgues, M.G. Haehnelt, S. Matarrese, A. Riotto, *PRD 71:063534,2005*.
M. Endo, F. Takahashi, T.T. Yanagida, *arXiv:hep-ph/0701042*, and *PRD 76:083509,2007* \[arXiv:hep-ph/0706.0986v2\].
[^1]: based on work in collaboration with K. Jedamzik ([*LPTA-Montpellier*]{}), M. Lemoine ([*IAP-Paris*]{}) [@JLM04], [@JLM05], and work in progress, M. Kuroda (Meiji-Gakuin), M. Lemoine (Paris), M. Capdequi-Peyranère (Montpellier).
[^2]: If the spurion is lighter than the LMP, an efficient tree-level annihilation of the latter into a pair of spurions would lead to a too small $Y_{LMP}$.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We present the general structure of proper Ricci Collineations (RC) for type B warped space-times. Within this framework, we give a detailed description of the most general proper RC for spherically symmetric metrics. As examples, static spherically symmetric and Friedmann-Robertson-Walker space-times are considered.'
author:
- |
J.Carot[^1]\
Departament de Física\
Universitat de les Illes Balears\
E-07071 Palma de Mallorca\
Spain
- |
L. A. Núñez [^2] and U. Percoco [^3]\
[*Centro de Astrofísica Teórica*]{}\
[*Departamento de Física, Facultad de Ciencias,* ]{}\
Universidad de los Andes, Mérida 5101, Venezuela.
title: |
Ricci Collineations\
for\
type B warped space-times
---
Introduction
============
The purpose of this paper is to study Ricci Collineations (RC) for a certain class of space-times, namely type B warped space-times and in particular spherically symmetric space-times. Collineations are symmetry properties of space times. Katzin[* et al.* ]{}[@KatzinEtal69] define them as those vector fields, $X$, such that leave the various relevant geometric quantities in General Relativity invariant under Lie dragging in their direction. The best known examples of collineations are the [*Killing vectors*]{} ([*Motions*]{}), i.e. vectors that satisfy: $$\pounds _X\ g_{ab}=0 \label{collkill}$$ Other interesting symmetries are defined in analogously and the more frecuent cases of study have been:\
Conformal Motions: $$\pounds _X\ g_{ab}=2\sigma g_{ab} \label{collconf}$$ Affine Collineations: $$\pounds _X\ \Gamma _{ab}^c=0 \label{collafin}$$ Curvature Collineations: $$\pounds _X\ R_{abcd}=0 \label{collcurv}$$ Ricci Collineations: $$\pounds _X\ R_{ab}=0 \label{collricc}$$ Contracted Ricci Collineations: $$g^{ab}\pounds _X\ R_{ab}=0 \label{collcric}$$ Here $\pounds _X$ stands for the Lie derivative operator and the indices $%
a,b,...$ run from 1 to 4.
The well established connection between [*Killing Vectors* ]{} and constants of the motion has encouraged the search for general relations between collineations and conservation laws. Collineations, other than [*Motions*]{}, can be considered as non-noetherian symmetries and can also be associated to constants of the motion. [*Affine Collineations*]{} have been shown to be related to conserved quantities [@HojmanEtal86], and this property has been used to integrate geodesics of the Robertson-Walker metric [@BedranLesche86]. As far as we know, the first [*Curvature Collineation*]{} was found by Aichelburg [@Aichelburg70] for pp-wave metrics, and their relationships to first integrals of the geodesic equations extensively studied in Ref. [@KatzinEtal69]. Particular types of [*Ricci*]{} and [*Contracted Ricci Collineations*]{}, for the Robertson-Walker metric have also been found and shown to be related to the particle number conservation [@GreenEtal77]. Also, considerable attention is being paid to the related problem of symmetry inheritance in General Relativity [@ColeyTupper89]. Collineations have been studied in connection with fluid spacetimes [@GreenEtal77], [@OliverDavis77], [@TsamparlisMason90], [@Duggal92] and some specific examples have been given for the [*C-metric* ]{} [@AulestiaEtal84], Robertson-Walker Spacetimes [@NunezEtal90], and Gödel-type manifolds [@MelfoEtal92].
It is clear from the above definitions that [*Motions*]{} are particular cases of [*Affine Collineations*]{}, [*Affine Collineations*]{} are particular cases of [*Curvature Collineations*]{}, and so on. It is therefore possible to construct an “inclusion diagram” connecting these symmetries. One such diagram, that includes these and other related symmetries, is presented in Ref. [@KatzinEtal69]. A collineation of a given type is said to be[* proper*]{} if it does not belong to any of the subtypes. Clearly, in solving any collination equation, with the obvious exception of the [*Killing equation*]{}, solutions representing improper collineations can be found. Therefore, in order to relate a symmetry to a particular conservation law, and its corresponding constant of the motion, the “properness” of that collineation must be assured. Some computer algebra tools have been developed to check the properness of Ricci and other collineation vectors are under development [@MelfoNunez92] [@BertolotiEtal95a].
We assume that RCs are smooth vector fields. Although this is not necessarily so, by restricting ourselves to this case, we ensure that they form a Lie algebra with the usual bracket operation. Such an algebra naturally contains that of Special Conformal Killing Vectors (SCKV) (see reference [@ColeyTupper89]) which in turn contains that of Homothetic Vector Fields (HVF) and therefore the isometry algebra of all Killing Vector Fields (KV).
Regarding the Ricci tensor, we shall consider that it is non-degenerate (i.e.: rank 4) and this in turn ensures that the Lie algebra of RC is finite dimensional, its maximal dimension being 10 (9 being forbidden by Fubini ’s theorem). For further information on issues concerning dimensionality and degenerate Ricci tensor see, for instance, references [@CarotEtal94] and [@HallEtal95].
The paper is organized as follows: in section 2 we describe the basic features of the RC in type B warped space-times, then, in section 3, we consider spherically symmetric space-times as a particular case of these, studying two distinct cases, namely: static solutions and Friedmann-Robertson-Walker space-times.
Type B warped space-times
=========================
Suppose that $(M_1,h_1)$ and $(M_2,h_2)$ are a pair of pseudo-Riemannian manifolds, and $\Phi$ is a real valued function on $M_1$ (warping function’), one can then build a Lorentz manifold, $(M,g)$ by setting $M=M_1
\times M_2$ and $g=\pi_1^\ast h_1 \otimes \Phi^2 \pi_2^\ast h_2$, where the functions $\pi_1$ and $\pi_2$ are the canonical projections onto the factors of the product. $(M,g)$ is then called a warped product manifold’. If dim $M=4$, we say that $(M,g)$ is a warped space-time’ and one can classify them according to the respective dimensions of the factor (sub-)manifolds $M_1$ and $M_2$. We shall refer the reader to [@CarotDaCosta93] and references cited therein for a general discussion, restricting ourselves hereafter to the case dim $M_1=$ dim $M_2=2$, namely; warped space-times of the class $B$. Although all our considerations will be local, see [@Haddow] for some remarks on globally warped space-times. It can be shown that for type $B$ warped space-times, a coordinate chart exists (adapted to the manifold product structure), such that the metric takes the form
$$\label{warped}
ds^2=h_{AB}\left( x^D\right) \ {\rm d}x^A\ {\rm d}x^B+ \Phi ^2\left(
x^D\right) \ h_{\alpha \beta }\left( x^\gamma \right) \ {\rm d}x^\alpha \
{\rm d}x^\beta$$
where the indices $A,B,\ldots $ run from 1 to 2 and $\alpha ,\beta , \ldots $ from 3 to 4. The functions $h_{AB}$ and $h_{\alpha \beta }$ are the component forms of $\pi_1^\ast h_1$ and $\pi_2^\ast h_2$ in the local charts $\{ x^A \}$ and $\{ x^\alpha \}$, which are in turn adapted to $M_1$ and $%
M_2 $ respectively. The Ricci tensor of such a space-time takes then the following component form in the above chart:
$$\label{ricci1}
R_{AB}=\frac 12R_1\ h_{AB}-\frac 2\Phi \ \Phi _{A;B}\ ,$$
$$\label{ricci2}
R_{A\alpha }=0\ ,$$
$$\label{ricci3}
R_{\alpha \beta }=\frac 12\left( R_2-\left( \Phi ^2\right) _{;A}^A\right)
h_{\alpha \beta }\ \equiv \ Fh_{\alpha \beta } \ ;$$
where $F=\frac 12\left( R_2-\left( \Phi ^2\right) _{;A}^A\right)$ and $R_1$ and $R_2$ are the Ricci scalars associated to the 2-metrics $h_1$ and $h_2$. The semi-colon indicates, as usual, the covariant derivative with respect to the space-time metric.
Let now $X$ be a RC on $M$, and define its vertical and horizontal components, $X_1$ and $X_2$, as follows (see [@CarotDaCosta93]):
$$X_1^a\equiv g^{ab}\left( \pi _1^{*}h_1\right) _{bd}X^d\ \ \ \ X_2^a\equiv
X^a-X_1^a \label{components}$$
In the above adapted chart, one readily sees that $X_1^A=X^A,$ $\ X_1^\alpha
=0\ $, and $X_2^A=0,\ X_2^\alpha =X^\alpha \ $.
On account of (\[ricci1\]), (\[ricci2\]) , ( \[ricci3\]) and (\[components\]), equation (\[collricc\]) is now equivalent to:
$$\label{set}
R_{AB,D}X_1^D + R_{AD}X^D_{1,B} + R_{DB}X^D_{1,A}=0 \ ,$$
$$\label{vuit}
R_{AD}X^D_{1,\alpha} + Fh_{\alpha \beta}X^\beta_{2,A}=0 \ ,$$
$$\pounds _{X_2}h_{\alpha \beta }=2\Psi h_{\alpha \beta } \label{nou}$$
where $$\Psi =-\frac 12{\frac{F_{,D}X_1^D+F_{,\gamma }X_2^\gamma }F} \label{deu}$$ Take now $p_1\in M_1$ and consider the manifold ${\tilde{M}}_2\equiv
\{p_1\}\times M_2\cong M_2$ (see [@CarotDaCosta93]), equation (\[nou\]) is then a statement that $X_2$ is a Conformal Killing Vector (CKV) of $({%
\tilde{M}}_2,h_2)$, and therefore it can be re-written as $$X_{2\alpha /\beta }+X_{2\beta /\alpha }=2\Psi \ h_{\alpha \beta }
\label{x2conf1}$$ where a stroke denotes the covariant derivative associated with the metric $%
h_2$.
Furthermore, it is possible to write [@Hall90], $$\pounds _{X_2}R_{2\alpha \beta }=-2\Psi _{\alpha /\beta }-\left( h^{\mu \nu
}\Psi _{\mu /\nu }\right) h_{\alpha \beta } \label{x2conf2}$$ where $$R_{2\alpha \beta }~=~\frac{R_2}2\ h_{\alpha \beta } \label{Escricci2}$$ is the Ricci tensor of the metric $h_2$ . In addition, the Conformal Bivector associated to $X_2$ , i.e. $$F_{\alpha \beta }\equiv X_{2\alpha /\beta }-X_{2\beta /\alpha }
\label{Bivector}$$ satisfies $$F_{\alpha \beta /\gamma }=\frac{R_2}2(h_{\alpha \gamma }\ X_{2\beta
}-X_{2\alpha }\ h_{\beta \gamma })-\Psi _\alpha \ h_{\beta \gamma }+\Psi
_\beta \ h_{\alpha \gamma } \label{x2conf3}$$ Now, from (\[x2conf2\]) one obtains $$\Psi _{\alpha /\beta }=\lambda h_{\alpha \beta }\qquad {\rm and}\qquad
\lambda \equiv -\frac 18(\pounds _{X_2}\ R_2+2\Psi \ R_2) \label{x2conf4}$$ and from the Bianchi identities (on $({\tilde{M}}_2,h_2)$) for $\Psi \ $, it readily follows: $$\lambda _{,\gamma }=-\frac{R_2}2\Psi _{,\gamma } \label{lambdaderiv}$$ Furthermore, taking a further covariant derivative in the above expression, skewsymmetrising, and equating to zero, one has $$-\left( \frac{R_2}2\right) _{,\alpha }=\sigma \Psi _{,\alpha }
\label{R2deriv}$$ for some function $\sigma $.
To proceed with our study, it is useful to consider now the following decomposition of ${\tilde M}_2$; ${\tilde M}_2= H \cup K \cup C$, where $H$ is that open submanifold of ${\tilde M}_2$ on which $\Psi_{\alpha / \beta}
\not{= }0$ (hence $\lambda \not{= }0$ and $\Psi^\alpha \Psi_\alpha \not{=}0$ on $H$), $K$ is the interior of that set of points for which $\Psi_{\alpha /
\beta} = 0 \ $, and $C$ is a set with no interior defined by the decomposition itself.
We shall first study what happens in $K$. Since $\Psi_{\alpha / \beta} = 0 $ there, it follows that $\Psi_{, \alpha}$ is either zero on $K$ (in which case $X_2$ is homothetic), or else it is a (gradient) Killing vector (and then $X_2$ is an SCKV), the Bianchi identities then implying $R_2=0 \ $, i.e.; $h_2 \vert_K$ is flat. In the latter case ($h_2$ flat), one can always choose coordinates on $K$, say $\{ x,y \} \ $, such that $\Psi \vert_K = A x$ ($A =$ constant), and integrating out the conformal equations (\[nou\]) for $X_2$ on $K$ it follows
$$\label{sckvK}
X_2= \left( \frac 12 A(x^2-y^2) - D y + L \right) \partial_x + \left( Axy +
Dx + E \right) \partial_y$$
where $A, \ D, \ E$ and $L$ are constants on $K$ which will, in general, depend on the chosen $p_1 \in M_1$, and therefore, when considering $X_2$ on $M$, one will have that all of them are functions of the coordinates set up in $M_1$, thus $A=A(x^B), \ D=D(x^B), \cdots$ to be determined, along with the vertical component $X_1$ of $X$, from (\[set\]) and (\[vuit\]). In fact, it is easy to see from (\[vuit\]) that $A$ and $D$ must be constants, say $A=A_0$ and $D=D_0$, and from the expression (\[deu\]) of $%
\Psi$ with $R_2=0 \ $, together with $\Psi=A_0x $, it follows that $E$ must also be constant (which can be set equal to zero without loss of generality), then from (\[set\]) $X_1^A=P^A(x^B)x+Q^A(x^B)$, and therefore one has, on $M \cap K$ and if $R_2=0$: $$\label{2flat1}
X= ( P^Ax+Q^A ) \partial_A + ( \frac {A_0}2 (x^2-y^2) - D_0 y + L)
\partial_x + ( A_0 xy + D_0 x ) \partial_y$$ $$\label{2flat2}
\Psi=A_0x$$ $P$, $Q$ and $L$ being functions of the coordinates $\{ x^B \} $ on $M_1$ to be determined from (\[set\]) and (\[vuit\]).
If $\Psi _{,\alpha }|_K=0$, $X_2$ is an HVF, and therefore ([@Hall90]) $%
\pounds _{X_2}R_2=-2\Psi R_2$ if $R_2\neq $ constant or $\pounds _{X_2}R_2=0$ if $R_2=$ constant, hence (\[deu\]) implies $$\Psi =-\frac 12{\frac{(\Phi ^2)_{;AD}^AX_1^D}{(\Phi ^2)_{;A}^A}}
\label{psinova}$$ thus, given a basis of the homothetic algebra of $(M_2,h_2)$, say $\{\zeta
_I\}$ with $I\leq 4$, one will have $X_2=C^I\zeta _I$ on $({\tilde{M}}%
_2,h_2) $, the $C^I$ being constants which will in general depend on the chosen $p_1\in M_1$, and again, when considering $X_2$ on $M$, they will become functions of the coordinates in $M_1$, to be determined as before from (\[set\]) and (\[vuit\]). It is worth noticing that, whenever a proper HVF exists in $({\tilde{M}}_2,h_2)$, say $\zeta _1$ then (\[nou\]) implies that $C^1=\Psi $. It will be shown later on that, in all cases but one, the functions $C^I$ must in fact be constants (and (\[vuit\]) then implies that $X_1$ is just a vector field on $M_1$). Thus, we conclude that whenever $\Psi _{,\alpha }=0$, one has $$X=X_1^A(x^B,x^\gamma )\partial _A+C^I(x^B)\zeta _I \label{homot}$$ where $C^I$ and $X_1^A$ are functions of their arguments, to be determined from (\[set\]) and (\[vuit\]), and $\{\zeta _I\}$ with $I\leq 4$ form a basis of the homothetic algebra of $({\tilde{M}}_2,h_2)$.
Let us next study what happens on $H$. Notice that (\[x2conf4\]) can be rewritten as $\pounds _Yh_{\alpha \beta }=2\lambda h_{\alpha \beta }$ with $%
Y_\alpha =\Psi _{,\alpha }$; thus, $Y$ is also a CKV of $({\tilde{M}}_2,h_2)$ with conformal factor $\lambda $, and one therefore has [@Hall90]: $$\pounds _YR_{2\alpha \beta }=-2\lambda _{\alpha /\beta }-\left( h^{\mu \nu
}\lambda _{\mu /\nu }\right) h_{\alpha \beta } \label{lambdaH}$$ which, on account of (\[Escricci2\]) and (\[x2conf4\]), can be rewritten as $$\lambda _{\alpha /\beta }=-\frac 18\left( R_{2,\alpha }\Psi ^\alpha
+2\lambda R_2\right) h_{\alpha \beta }\equiv \Sigma h_{\alpha \beta }
\label{asterisc}$$ that is: $Z$ such that $Z_\alpha \equiv \lambda _{,\alpha }$ is a (gradient) CKV, colinear with another CKV, namely $Y$; it is then immediate to show, taking into account (\[x2conf4\]), (\[lambdaderiv\]), (\[R2deriv\]) and (\[asterisc\]), that $R_2$ must be constant ($\sigma =0$ in (\[R2deriv\])), (\[x2conf4\]) then reading $$\Psi _{\alpha /\beta }=-\frac{R_2}4\Psi h_{\alpha /\beta }$$ The Bianchi identities specialized to $\Psi _{,\alpha }$ then imply one of the following:
1. $R_2=0$ and $\Psi _{,\alpha }\not{=}0\ $, one then has the expression (\[sckvK\]) for $X_2$, etc.
2. $R_2=$ constant ($\not{=}0$) and $\Psi _{,\alpha }=0\ $, $X_2$ is then an HVF of $\left( {\tilde{M}}_2,h_2\right) \ $, but since $R_2$ is constant and non-zero, it must be a KV, i.e.; $\Psi =0$.
3. $R_2=0$ and $\Psi _{,\alpha }=0\ $, $X_2$ is an HVF of $\left( {%
\tilde{M}}_2,h_2\right) \ $, possibly non-Killing.
Notice that whenever $\Psi_{,\alpha} = 0 \ $, one gets the same results as when studying this case in $K \subset {\tilde M}_2$, i.e.; equations (\[psinova\]) and (\[homot\]) hold.
We can roughly summarize the results so far obtained as follows:
The horizontal component $X_2$ of a RC $X$ is either an HVF of $({\tilde M}%
_2,h_2)$ (i.e.; $\Psi_{,\alpha}=0$) and $X$ is then given by (\[homot\]), or else it is a proper SCKV of $({\tilde M}_2,h_2)$ (that is; $%
\Psi_{,\alpha} \not{= }0$, $\Psi_{\alpha / \beta}=0$), this being possible only when $R_2=0$ (i.e.; $({\tilde M}_2,h_2)$ flat), and in that case $X$ takes the form given by (\[sckvK\]). In both cases, the functions appearing in (\[homot\]) and (\[sckvK\]) must satisfy (\[set\]) and (\[vuit\]).
We shall next focus our attention on the case $X_2$ homothetic, studying the various cases that may arise in connection with the different structures and dimensions of the homothetic algebra of $({\tilde M}_2, h_2)$.
To this end, let ${\cal H}_r$ be the homothetic algebra of $({\tilde M}_2,
h_2)$, $r$ being its dimension. Since dim $M_2=2$ it follows that $r$ can only be $0, \, 1, \, 2, \, 3$ or $4$. We shall deal separately with all these cases assuming, for the sake of simplicity, that $h_2$ is Riemannian (similar conclusions hold if $h_2$ is Lorentz).
1. $r=0$. In this case no HVFs exist (including KVs), and therefore $%
\Psi =C^I=0$, i.e.; $X_2=0$ and $X=X_1$ with $X_{1,\alpha }^D=0$ as a consequence of (\[vuit\]), that is: $X$ is a vector field on $M_1$ which must satisfy (\[set\]) and (\[psinova\]) with $\Psi =0$.
2. $r=1$. There are now two cases to be distinguished, depending on whether a proper HVF exists or not.
1. A proper HVF $\zeta $, exists in $({\tilde{M}}_2,h_2)$. It is easy to see that one can then always choose coordinates, say $x$ and $y$, such that the line element $d\sigma ^2$ associated with $h_2$, and the HVF $\zeta $ read in these coordinates $$d{\sigma }^2=e^{2y}(dx^2+h^2(x)dy^2)\,\,\,{\rm and}\,\,\zeta =\partial _y$$ the associated Ricci scalar is $R_2=-2e^{-2y}h^{-1}h^{\prime \prime }$ (a prime denoting derivative with respect to $x$), and (\[vuit\]) then implies:
$$R_{AD}X_{1,x}^D=0\,\,\,\,\,\,{\rm and}\,\,\,\,\,\,R_{AD}X_{1,y}^D=-F\Psi
_{,A}e^{2y}h^2(x)$$
which cannot be fulfilled unless $(Fh^2(x))_{,x}=0$, i.e.; $h(x)={\rm %
constant}$, in which case $R_2=0$ and therefore $r=4$. Thus, $\Psi _{,A}=0$ ($\Psi =$ constant $\not{=}0$), $X_{1,\alpha }^D=0$ and then $%
X=X_1^A(x^D)\partial _A+\Psi \zeta $ with $X_1$ satisfying (\[set\]) and (\[psinova\]) with $\Psi =$ constant ($\not{=}0$).
2. No proper HVF exists in $({\tilde{M}}_2,h_2)$, just a KV, say $\xi $. It then follows that $\Psi =0$ necessarily, and again coordinates may be chosen such that $$d{\sigma }^2=dx^2+h^2(x)dy^2\,\,\,\,\,\,{\rm and}\,\,\,\,\,\xi =\partial _y$$ the Ricci scalar is then $R_2=-2h^{-1}h^{\prime \prime }$, and (\[vuit\]) implies, as in the previous case, $(Fh^2(x))_{,x}=0$, which in turn can be seen to imply
$$\left( \Phi ^2\right) _{;A}^A=2a\,\,\,\,,\,\,\,\,(a={\rm constant})$$
$$h^{\prime \prime }+ah^2=b\,\,\,\,,\,\,\,\,(b={\rm constant})$$
Performing now the coordinate change $h(x)\equiv z$, the above line element reads $$d{\sigma }^2={\frac{dz^2}{2C+z^2+2\log z}}+z^2dy^2$$ Hence (\[vuit\]) implies $X_1^A=P^A(x^D)y+Q^A(x^D)$ and then $$X=\left( P^A(x^D)y+Q^A(x^D)\right) \partial _A+C(x^D)\xi
\label{excepcional1}$$ where $P^A(x^D)$ and $Q^A(x^D)$ must both satisfy (\[set\]) separately, and $C(x^D)$ must be such that $R_{AD}P^D=-bC_{,A}$.
3. $r=2$, ${\cal H}$ must then contain at least one proper HVF, since otherwise (${\cal H}$ spanned by two KVs) a third KV would necessarily exist, hence dim ${\cal H}=3$. Suppose then that a proper HVF, $\zeta $, exists; the other vector in the basis of ${\cal H}$ can always be chosen to be a KV, say $\xi $, and there are two possible, non-isomorphic, Lie algebra structures for ${\cal H}$, namely $[\xi ,\zeta ]=0$ (abelian), and $[\xi
,\zeta ]=\xi $ (non abelian). In the abelian case, coordinates may be chosen such that the line element, $\zeta $ and $\xi $ read respectively $$d\sigma ^2=dx^2+x^2dy^2\,\,,\,\,\zeta =x\partial _x\,\,,\,\,\xi =\partial _y$$ but it then follows that $R_2=0$ and therefore two other KVs exist, $r$ thus being 4, therefore this case cannot arise.
In the non-abelian case, and again by means of a suitable choice of coordinates, one has: $$d\sigma ^2=dx^2+x^{2{\frac{n-1}n}}dy^2\,\,,\,\,\zeta =nx\partial
_x+y\partial _y\,(n\not{=}1)\,\,,\,\,\xi =\partial _y$$ but then (\[vuit\]) implies, as in previous cases, that $\left( Fx^{2{%
\frac{n-1}n}}\right) _{,x}=0$, which can not be satisfied, therefore $\Psi
_{,A}=C_{,A}=0$ (i.e.; $\Psi $ and $C$ constants) and then $X_{1,\alpha
}^D=0 $, and again $X=X_1^A(x^D)\partial _A+\Psi \zeta $ with $X_1$ satisfying (\[set\]) and (\[psinova\]) with $\Psi =$ constant ($\not{=}0$).
4. $r=3$ If a proper HVF exists in $({\tilde{M}}_2,h_2)$, the associated Killing subalgebra is then of dimension 2, and therefore a third KV exists, hence dim ${\cal H}=4$ and therefore this case is not possible. If, on the other hand, no proper HVFs exist, $({\tilde{M}}_2,h_2)$ is of constant curvature and $\Psi =0$ necessarily. Let $\{\xi _J\}\,,\,\,J=1,2,3$ be three KVs spanning ${\cal H}$, from (\[vuit\]) it follows $R_{AD}X_{1,\alpha
}^D=-FC_{,A}^J\xi _{J\alpha }$, differentiating with respect to $x^\beta $, skewsymmetrising and equating to zero, one has $$C_{,A}^J\xi _{J[\alpha ,\beta ]}=0$$ that is; either $C_{,A}^J=0$ or else $({\tilde{M}}_2,h_2)$ contains a gradient KV. From [@KramerEtal80], it is easy to see that the later is only possible if $R_2=0$, but in that case a proper HVF is always admitted (namely $\zeta =x\partial _x$ in the coordinates used in [@KramerEtal80]), and therefore dim ${\cal H}=4$.
5. $r=4$. In this case $({\tilde{M}}_2,h_2)$ is flat, the line element and KVs being those given in [@KramerEtal80] and the proper HVF $\zeta
=x\partial _x$. Proceeding as before, one can readily see from (\[vuit\]) that $X_1^A=M^A(x^D)x^2+(P_1^A(x^D)\cos y+P_3^A(x^D)\sin y)x+Q^A(x^D)$, but since $\Psi \not{=}0$ and $\Psi _{,\alpha }=0$, it follows that $%
P_1^A=P_3^A=M^A=0$, which in turn imply $\Psi
_{,A}=C_{,A}^1=C_{,A}^2=C_{,A}^3=0$, hence $X_{1,\alpha }^A=0$, that is: $%
X_1 $ is a vector field on $M_1$ that has to satisfy (\[set\]) and (\[psinova\]) with $\Psi =$ constant ($\not{=}0$), and $X=X_1^A(x^D)\partial
_A+\Psi \zeta +C^J\xi _J$.
Our purpose in the next sections is to apply the results so far obtained to the case of spherically symmetric space-times which are also static, as well as to Friedmann-Robertson-Walker (FRW) models.
Spherically symmetric space-times
=================================
We next specify the above results to the case of a general spherically symmetric spacetime whose metric, in the local chart $\{x^{0,1,2,3}~= ~t,r,
\vartheta ,\phi \}$ takes the form [@KramerEtal80] $$\label{MetricShearF}
ds^2=\ -{\bf e}^{2\nu (t,r)}\,{\rm d}t^2+\,{\bf e}^{2\lambda (t,r)}\, {\rm d}
r^2+\,r^2({\rm d}\vartheta ^2+\,\sin ^2\phi {\rm \,d}\phi ^2)$$ Comparing the metric (\[warped\]) with the above (\[MetricShearF\]), we have $\{x^A~=~t,\ r\ ;\ x^\alpha =\vartheta ,\ \phi \}$ ; $\Phi ~=~r$ ;
$$h_{AB}\left( t,r\right) {\rm d}x^A{\rm d}x^B~=~\,-{\bf e}^ {2\nu (t,r)} {\rm %
d}t^2~~+~~\,{\bf e}^{2\lambda (t,r)} \,{\rm d}r^2$$
and
$$h_{\alpha \beta }{\rm d}x^\alpha {\rm d}x^\beta ~=~{\rm d} \vartheta
^2~+~\sin ^2\phi {\rm \,d}\phi ^2\$$ Thus, the Ricci tensor can be written as
$$\label{sphricci1}
R_{tt}=-\frac 12R_1{\bf e}^{2\nu (t,r)}\ +\frac{2\nu ^{\prime }}r\ {\bf e}
^{2\left( \nu (t,r)-\lambda (t,r)\right) }$$
$$\label{sphriccitr}
R_{t\ r}=\frac 2r\ \dot{\lambda}$$
$$\label{sphricci2}
R_{rr}=\frac 12\ R_1\ {\bf e}^{2\lambda }+\frac{2\lambda ^{\prime }}r$$
and $$\label{sphricci3}
R_{\alpha \beta }=\left\{ 1-{\bf e}^{-2\lambda }\left[ 1+r\ \left( \nu
^{\prime }-\lambda ^{\prime }\right) \right] \right\} h_{\alpha \beta }$$ where a dash and a dot indicate, as usual, partial derivatives with respect to $r$ and $t$ respectively. As above, $R_1$ is the Ricci scalar associated with the 2-dimensional metric $h_{AB}$, and now $\frac 12~R_2~=~1$.
According to our foregoing discussion, any RC $X$ must be of the form $$\label{rc}
X=X_1+C^J \xi _J$$ where $\left\{ \xi _J,\ \ J=1,2,3\right\} $ are the KV ’s that implement the spherical symmetry, $X_1=X^A(t,r)\partial _A$ and $C^J$ are constants, $%
J=1,\ 2, \ 3$, which can be set equal to zero without loss of generality (since $C^J \xi_J$ is a KV of the space-time and therefore a trivial RC). On the other hand, since $\Psi=0$ and $(\Phi^2)^A_{;A}= 2{\bf e}^{-2\lambda }
\left[ 1+r\ \left( \nu ^{\prime }-\lambda ^{\prime }\right) \right]$, (\[psinova\]) implies $$\label{psi1}
\{{\bf e}^{-2\lambda } \left[ 1+r\ \left( \nu ^{\prime }-\lambda ^{\prime
}\right) \right] \}_{,D} X^D =0$$ therefore, the proper RCs of a spherically symmetric space-time whose Ricci tensor is non-degenerate, are of the form $$X=X^t(t,r) \partial_t + X^r(t,r) \partial_r$$ and they must satisfy (\[psi1\]) in addition to (\[set\]) specialised to the Ricci tensor components given by (\[sphricci1\]), (\[sphriccitr\]) and (\[sphricci2\]).
We shall next present two examples: static spherically symmetric space-times, and FRW spacetimes.
Static spherically symmetric space-times
----------------------------------------
Let us consider first the case of static spherically symmetric spacetimes, these are described by (\[MetricShearF\]) where the functions $v$ and $%
\lambda $ appearing in it depend just on $r$ , $\partial _t$ thus being a KV. For the purpose of this paper it is convenient to write the components of the Ricci tensor for this metric as follows [@BokhariQadir93], [@JamilEtal94], [@FaridEtal95] $$R_{tt}\equiv A(r)\quad R_{rr}\equiv B(r)\quad R_{\theta \theta }\equiv
C(r)\quad {\rm and}\quad R_{\phi \phi }\equiv \sin ^2\theta \ R_{\theta
\theta } \label{Ricciform}$$
Taking now into account the results of the previous section one has $$\label{GeneralColl}
X =X ^t\left( t,r\right) \partial _t+X ^r\left( t,r \right) \partial _r$$ and the (non-trivial) equations arising from (\[set\]), are simply: $$\label{LieRicci00}
A^{\prime }(r)X ^r+2A(r)X _{,t}^t=0$$ $$\label{LieRicci01}
A(r)X _{,r}^t+B(r)X _{,t}^r=0$$ $$\label{LieRicci11}
B^{\prime }(r)X ^r+2B(r)\ X _{,r}^r=0$$ $$\label{LieRicci22}
C^{\prime }(r)X ^r=0$$ Equation (\[LieRicci22\]) directly implies: $$\label{c'=0}
C^{\prime }(r)=0$$ since otherwise one would have $X ^r=0$ that would imply, from the remaining equations, $X ^t=$ constant, thus being a KV and not a proper RC.
A direct integration of equation (\[LieRicci11\]) gives $$\label{tempXi1}
X ^r=\frac{{\cal K}(t)}{\sqrt{\left| B(r)\right| }}$$ Now, substituting this result back into eq. (\[LieRicci00\]) and (\[LieRicci01\]), differentiating them with respect to $t$ and $r$, respectively; and equating the crossed derivatives of $X ^t$, we obtain $$\label{temp11Xi}
{\cal K}_{,tt}\ \frac{\sqrt{\left| B(r)\right| }}{A(r)}=\frac 12 {\cal K}
\left( \frac{A^{\prime }(r)\ }{A(r)\sqrt{\left| B(r)\right| }} \right)
^{\prime }$$ and the following two cases arise:
**Case I**
----------
$$\label{temp12Xi}
{\cal K}_{,tt}\ -\epsilon \ k^2{\cal K}=0;\qquad k={\it Const},\quad
\epsilon =\pm 1$$
therefore $$\label{temp13Xi}
{\cal K}(t)=\left\{
\begin{array}{cc}
a{\bf e}^{kt}+b{\bf e}^{-kt} & \epsilon =+1 \\
a\sin kt+b\cos kt & \epsilon =-1
\end{array}
\right| \quad$$ and $$\label{temp14Xi}
2\epsilon \ k^2\ \frac{\sqrt{\left| B(r)\right| }} {A(r)}=\left( \frac{
A^{\prime }(r)\ }{A(r)\sqrt{\left| B(r)\right| }} \right) ^{\prime }$$ Substituting these results back into (\[LieRicci00\]), integrating and plugging them back into (\[LieRicci01\]), we find $$\label{temp15Xi}
X ^t=-\frac 12 \left( \frac{A^{\prime }(r)\ }{A(r)\sqrt{\left| B(r) \right| }
}\right) M(t)$$ where $M(t)= \int {\cal K}(t) dt$. and the constant of integration has been set equal to zero without loss of generality.
Thus, for this case a proper RC is of the form:
$$\label{Xisolcas1}
X =-\frac 12\left( \frac{A^{\prime }(r)\ }{A(r)\sqrt{\left| B(r) \right| }}
\right) \left( \int {\cal K}(t){\rm d}t\right) \partial _t+ \frac{{\cal K}%
(t) }{\sqrt{\left| B(r)\right| }}\partial _r$$
where ${\cal K}(t)$ is given by (\[temp13Xi\]), and the components of the Ricci tensor must satisfy (\[c’=0\]) and (\[temp14Xi\]).
**Case II**
-----------
$$\label{II-1}
{\cal K}={\cal S}_1t+{\cal S}_2\qquad {\cal S}_1, \ {\cal S}_2 = {\it Const}$$
and $$\label{II-2}
\frac 12\frac{A^{\prime }(r)\ }{A(r)\sqrt{\left| B(r)\right| }}= \sigma =%
{\it Const}$$ Then from (\[tempXi1\]) and (\[LieRicci00\]-\[LieRicci11\]), one gets after some straightforward calculations:
$$\label{temp30Xi}
X =\left\{ -\sigma \left( \frac 12{\cal S}_1\ t^2+{\cal S}_2 \ t \right) +
\frac{{\cal S}_1}{2\,\sigma }\frac 1{A(r)}\right\} \, \partial _t+\left(
{\cal S}_1\ t+{\cal S}_2\ \right) \frac 1 {\sqrt{B(r)}}\,\partial _r$$
As an example of a space-time satisfying the above requirements [@BertolotiEtal95b], take for instance $$\nu (r)=\frac 12\left( \frac{r^4}{8r_0^2}+h\,\ln \frac r{r_0}+k\right) \quad
{\rm and\quad }\lambda (r)=\nu (r)+\,\ln \frac r{r_0} \label{metricelem}$$ therefore the Ricci components can be written $$B(r)=2\frac{h+1}{r^2}\qquad C(r)=1\quad {\rm and}\quad A(r)={\em Const}
\label{b}$$ where $r_0$, $h$ and $k$ are constants, and we have for the RC: $$X^t=-\,c_4\sqrt{2\,\left( h+1\right) }\,\ln r+c_0\quad {\rm and}\quad X^r=%
\frac{c_4\,t+c_5}{\sqrt{2\,\left( h+1\right) }}r \label{Collstatic}$$ This result invalidates a misleading theorem stated in reference [@JamilEtal94], and used in [@FaridEtal95]. According to this “theorem”, this collineation vector (\[Collstatic\]) should represent an isometry; however it is easy to see that $X$ does not reduce to a KV unless $%
c_4~=~c_5~=~0$.
Since all KV’s are naturally RC’s and these (if assumed smooth) form a Lie algebra under the usual bracket operation, the Lie bracket of the above RC’s with the four KV’s the metric admits, must yield in turn RC’s; thus $$\left[ \xi _I,\xi \right] =0\qquad \forall I=1,2,3$$ where $\xi _I$ designate the KV’s implementing the spherical symmetry, and $$\left[ \partial _t,X\right] =X^{\prime }\left( \neq 0\right)$$ where $X^{\prime }=\frac{c_{4\,}r}{\sqrt{2(h+1)}}$ is also a proper RC.
A more detailed account of RCs for non-static spherically symmetric space-times will be given in a forthcoming paper.
FRW space-times.
----------------
As an example of RC for non-static spherically symmetric metrics, we consider FRW space-times described by [@Weinberg72]:
$$\label{frw}
{\rm d}s^2=-\ {\rm d}t^2+R\left( t\right) ^2\left( \frac{{\rm d}r^2}
{1-k\,r^2 }+r^2\,{\rm d}\vartheta ^2+r^2\sin ^2\vartheta \ {\rm d} \phi
^2\right)$$
Again, using the above notation, we have $\Phi =rR(t)$, $$\label{subspace1}
h_{AB}\,{\rm d}x^A\,{\rm d}x^B=-\ {\rm d}t^2+R\left( t\right) ^2 \frac{{\rm d%
} r^2}{1-k\,r^2}$$ and $$\label{subspace2}
h_{\alpha \beta }\,{\rm d}x^\alpha \,{\rm d}x^\beta =\, {\rm d}\vartheta
^2+\sin ^2\vartheta \ {\rm d}\phi ^2$$ Then the Ricci tensor takes the form $$\begin{aligned}
&&R_{tt}=-3\frac{\ddot{R}}R \nonumber \\
&&R_{rr}=g_{rr}\frac \Delta {R^2} \label{ricci-frw} \\
&&R_{\alpha \beta }=g_{\alpha \beta } \frac \Delta {R^2} \nonumber \\
&&\Delta =2k+2\dot{R}^2+R\ddot{R} \nonumber\end{aligned}$$ Specializing (\[set\]) to the present case, we obtain $$\begin{aligned}
&&X ^tR_{rr,t}+X ^rR_{rr,r}+2R_{rr}X _{,r}^r=0 \nonumber \\
&&X ^tR_{tt,t}+2R_{tt}X _{,t}^t=0 \label{rc-frw} \\
&&R_{tt}X _{,r}^t+R_{rr}X _{,t}^r=0 \nonumber \\
&&X ^tR_{\theta \theta ,t}+X ^rR_{\theta \theta ,r}=0 \nonumber\end{aligned}$$ Thus, we get [@NunezEtal90] $$\begin{aligned}
X ^t &=&c\left( 1-kr^2\right) ^{1/2}\left| R_{00}\right| ^{-1/2} \nonumber
\\
X ^r &=&g(t)\,r\,\,\left( 1-kr^2\right) ^{1/2} \label{psi-frw}\end{aligned}$$ where $g(t)=-\,c\,\left| R_{00}\right| ^{-1/2}\left( \dot \Delta /2\Delta
\right) $ , and $c$ is a constant.
ACKNOWLEDGEMENTS
================
Two of us (J.C. and U. P.) gratefully acknowledge funding from Postgrado en Astronomía y Astrofísica as well as the warm hospitality of the Laboratorio de Física Teórica, Universidad de Los Andes Mérida, Venezuela, where most of this work was carried out. J.C. acknowledges partial financial support from STRIDE program, Research Project No. STRDB/C/CEN/509/92. The authors wish also to thank the staff of the [*SUMA*]{}[* *]{}, the computational facility of the Faculty of Science (Universidad de Los Andes).
[99]{} G.H. Katzin, J. Levine and W.R. Davis, [*J. Math. Phys.*]{} [**10**]{}, 617 (1969).
Hojman, S., Núñez, L., Patiño, A., and Rago, H. [*J. Math. Phys*]{} [**27**]{}, 281 (1986).
Bedran, M.L.,and Lesche, B. [*J. Math. Phys*]{} [**27**]{}, 2360 (1986).
Aichelburg, P. (1970) [*J. Math. Phys.*]{} [**11**]{}, 2458.
L.H. Green, L.K. Norris, D.R. Oliver and W.R, Davis, [*Gen. Rel. Grav.*]{} [**8**]{}, 731 (1977).
Coley, A.A., and Tupper, B.O.J. (1989) [*J. Math. Phys.*]{} [**30**]{}, 2616.
D.R. Oliver and W.R, Davis, [*Gen. Rel. Grav.*]{} [**8**]{}, 905 (1977).
M. Tsamparlis and D.P. Mason, [*J. Math. Phys.*]{} [**31**]{}, 1707 (1990).
K.L. Duggal, [*J. Math. Phys.*]{} [**33**]{}, 2989 (1992).
L. Aulestia, L. Núñez, A. Patiño, H. Rago and L. Herrera, [*Nuov. Cim.*]{} [**B 80**]{}, 133 (1984).
L. Núñez, U. Percoco and V. M. Villalba,[* J. Math. Phys.*]{} [**31**]{}, 137 (1990).
A. Melfo, L. A. Núñez, U. Percoco and V. Villalba, [*J. Math. Phys.*]{} [**33**]{}, 2558 (1992).
A. Melfo and L.A. Núñez, [*Gen. Rel. Grav.* ]{} [**24**]{}, 1125 (1992).
R. Bertolotti, L. A. Núñez and U. Percoco “Computer Algebra and Collineation Vectors in General Relativity” [*Preprint*]{} Laboratorio de Física Teórica, Universidad de los Andes (1995).
J. Carot, J. da Costa and E. G. L. R. Vaz, [*J. Math. Phys.*]{} [**35**]{}, 4832 (1994).
G.S. Hall, [*J. Math. Phys.*]{}, [**31**]{}, 1198, (1990)
G.S. Hall, I. Roy and E. G. L. R. Vaz, “Ricci and Matter Collineations in Spacetimes” [*Preprint*]{} University of Aberdeen (1995).
J. Carot and J. da Costa [*Class. Quantum Grav.*]{} [**10**]{}, 461 (1993).
B. Haddow and J. Carot [*Class. Quantum Grav.*]{}, [**13**]{} (1996) 289
D. Kramer, H. Stephani, E. Hearlt, and M. A. H. MacCallum,[* Exact Solutions of Einstein Field Equations*]{} (Cambridge University, Cambridge, 1980).
A. H. Bokhari, and A. Qadir, [*J. Math. Phys.*]{} [**34**]{}, 3543 (1993).
M. Jamil Amir, A. H. Bokhari, and A. Qadir, [*J. Math. Phys.*]{} [**35**]{}, 3005 (1994).
T. B. Farid, A. Qadir and M. Ziad [*J. Math. Phys.*]{} [**36**]{}, 5812 (1995)
R. Bertolotti, G. Contreras, L. A. Núñez, U. Percoco and J. Carot [** **]{}[*J. Math. Phys.,*]{}[* *]{}[**37**]{},1086-1088 (1996).
S. Weinberg, [*Gravitation and Cosmology*]{} (Wiley, New York 1972).
[^1]: Email: [email protected]
[^2]: Email:[email protected]
[^3]: Email: [email protected]
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Taking the collisionless damping of geodesic acoustic mode (GAM) as an example, the physics processes underlying wave particle resonances in the short wavelength limit are clarified. As illustrative application, GAM excitation by energetic particles in short wavelength limit is investigated assuming a single pitch angle slowing-down fast ion equilibrium distribution function. Conditions for this energetic particle-induced GAM (EGAM) to be unstable are discussed.'
author:
- 'Liu Chen$^{1, 2}$, Zhiyong Qiu$^{1}$ and Fulvio Zonca$^{3, 1}$'
title: Short Wavelength Geodesic Acoustic Mode Excitation by Energetic Particles
---
Recently, due to geodesic acoustic mode (GAM) [@NWinsorPoF1968] excitation by energetic particles (EPs) in the large drift orbit limit [@MSasakiPoP2016], there has been renewed interest in wave-particle resonances at short wavelength, which was firstly investigated in Ref. for the collisionless damping of GAM, and presented later providing detailed derivation and physics interpretation [@ZQiuPPCF2009]. The same approach was also applied to study quasi-linear transport of EPs by drift wave turbulence [@WZhangPRL2008; @ZFengPoP2013]. However, the understanding of the underlying physics processes proposed in recent literature, e.g. [@MSasakiPoP2016], may yield to some mis-interpretation and inconsistency with the existing theoretical framework. In this brief communication, our aim is to clarify the underlying physics processes for wave-particle resonance in the short wavelength limit and, as illustrative application, investigate EP-induced GAM (EGAM) [@RNazikianPRL2008; @GFuPRL2008; @ZQiuPPCF2010] excitation by fast ions with large magnetic drift orbits.
To discuss the physics picture of wave-particle resonance in the short wavelength limit, we take electrostatic GAM collisionless damping originally discussed in [@FZoncaEPL2008; @ZQiuPPCF2009] as example. For the clarity of discussion, we assume small but finite electron temperature, i.e., $\tau\equiv T_e/T_i\ll1$ such that $|\widetilde{\delta\phi}_G/\overline{\delta\phi}_G|\sim \tau k_r\rho_{ti}\ll1$ while one still has $|\omega_{tr,e}|\gg|\omega_G|$. Here, $\widetilde{\delta\phi}_G $ and $\overline{\delta\phi}_G$ are respectively the $m\neq0$ and $m=0$ components of the perturbed scalar potential; $\omega_{tr}\equiv v_{\parallel}/(qR_0)$ is the transit frequency, $k_r$ is the radial wavenumber and $\rho_{ti}$ is the ion Larmor radius at thermal velocity. In this limit, consistent with the short wavelength assumption of interest here, the perturbed electron response (distribution function) to GAM is $\delta f_e=0$, and the GAM dispersion relation can be derived from the quasi-neutrality condition: $$\begin{aligned}
\sum_{s}\langle \delta f_s\rangle=0.\label{eq:QN}\end{aligned}$$ Here, $\langle\cdots\rangle$ denotes velocity space integration, subscript $s$ denotes different ions species and, thus, equation (\[eq:QN\]) can also be applied to study EGAM excitation by EPs. $\delta f_s$ can be expressed as $\delta f_s=e\partial_EF_0\delta\phi/m+\exp[i(m_ic)/(eB^2)\mathbf{k}\times\mathbf{B}\cdot\mathbf{v}]\delta H$, and the nonadiabatic response can be derived from the following linear gyrokinetic equation [@PRutherfordPoF1968; @JTaylorPP1968]: $$\begin{aligned}
\left(-i\omega+\omega_{tr}\partial_{\theta}+i\omega_d\right)\delta H_k=i\omega(e/m_i)\partial_E F_0 J_k\overline{\delta\phi}_G,\label{eq:LinearGKE}\end{aligned}$$ with $\omega_d=\hat{\omega}_d\sin\theta=k_r\rho_{ti}v_{ti}(v^2_{\perp}/2+v^2_{\parallel})/(v^2_{ti}R_0)\sin\theta$ being the magnetic drift frequency due to geodesic curvature, $J_k\equiv J_0(k_r\rho_{ti})$ with $J_0$ being Bessel function of zero-order accounting for finite Larmor radius effects, $v_{ti}\equiv \sqrt{2T_i/m_i}$ being the ion thermal velocity, $E=v^2/2$ and other notations are stardard.
Noting that $\omega_G\simeq v_{ti}/R_0\sim q\omega_{tr,i}\gg\omega_{b,i}\simeq \sqrt{\epsilon}\omega_{tr,i}$, and assuming well circulating particles in the large aspect ratio limit, equation (\[eq:LinearGKE\]) can be solved and yields, for $v_{\parallel}>0$, $$\begin{aligned}
\delta H_s=\frac{\omega}{\omega_{tr}}\hat{S}e^{-\psi(\theta)}\int^{\theta}_{-\infty} e^{\psi(\theta')}d\theta'.\label{eq:deltaH_general}\end{aligned}$$ Here, $\hat{S}\equiv -i(e/m_i)\partial_E F_0J_G\overline{\delta\phi}_G$, $\psi(\theta)\equiv -i(\omega\theta+\hat{\omega}_d\cos\theta)/\omega_{tr}$, $\omega_b$ is the bounce frequency of trapped particles, and $\epsilon\equiv r/R_0$ is the inverse aspect ratio. Similar expression can also be obtained for $v_{\parallel}<0$.
Noting that $$\begin{aligned}
e^{i\hat{\Lambda}cos\theta}=\sum_l i^lJ_l(\hat{\Lambda})e^{il\theta},\nonumber\end{aligned}$$ the integration in $\theta'$ in equation (\[eq:deltaH\_general\]) can be carried out by transforming into transit harmonics, and one obtains $$\begin{aligned}
\delta H_s=i\omega\hat{S}\sum_p i^p J_p(\hat{\Lambda})e^{ip\theta}\sum_l \frac{(-i)^lJ_l(\hat{\Lambda})e^{il\theta}}{\omega-l\omega_{tr}}.\label{eq:Hi_harmonic}\end{aligned}$$ Here, $\hat{\Lambda}\equiv \hat{\omega}_d/\omega_{tr}$ and $\exp{(-i\hat{\Lambda}\cos\theta)}$ is the “pullback" (coordinate transformation) from drift orbit center to particle guiding center coordinates. The resonance condition is $\omega-l\omega_{tr}=0$, with $l$ being integer, and resonant particles satisfying $|v_{\parallel,res}/v_{ti}|\sim O(q/l)$ due to the GAM/EGAM frequency ordering. The subscript “res" denotes resonant particles. Furthermore, the “population" of particles for each transit resonances is proportional to $J^2_l(\hat{\Lambda})\partial_E F_0|_{v_{\parallel,res}}$. Noting that $\hat{\Lambda}_{res}\sim k_r\rho_i q^2/l$ and the properties of Bessel functions, one can truncate the summation in equation (\[eq:Hi\_harmonic\]) at finite $l$ [@FHintonPPCF1999; @HSugamaJPP2006] in the small drift orbit limit with $k_r\rho_i q^2\ll1$. GAM collisionless damping due to the primary transit resonance ($|\omega|=|\omega_{tr}|$) only was investigated in Ref. . It was shown by Sugama et al [@HSugamaJPP2006] that, for increasing $k_r\rho_{ti}q^2$, GAM collisionless damping can be significantly enhanced by the increasing weight of higher order transit resonances due to the finite orbit width effect; and the analytical expression including $|\omega|=2|\omega_{tr}|$ resonance was derived. By further increasing $\hat{\Lambda}_{res}$ due to larger $k_r$ or $q$, however, more and more transit resonances are needed for the accurate description of GAM collisionless damping [@XXuPRL2008], and the analytical expression is very difficult to obtain due to the non-trivial task of summing up all the transit resonances.
An alternative approach was developed in Ref. [@FZoncaEPL2008], to derive the analytical expression of GAM collisionless damping rate in the short wavelength limit ($k_r\rho_iq^2\gg1$), with all the transit resonances taken into account. Here, we will first show that, the perturbed distribution function for resonant particles derived in Refs. [@FZoncaEPL2008; @ZQiuPPCF2009] are equivalent to the general solution of equations (\[eq:deltaH\_general\]) or (\[eq:Hi\_harmonic\]) in the proper limit, and then briefly summarize the main idea of this approach [@FZoncaEPL2008]; while interested readers may refer to Ref. [@ZQiuPPCF2009] for the detailed derivation.
In the large orbit limit, equation (\[eq:deltaH\_general\]) can be expanded using the smallness parameter $1/\dot{\psi}$, with $|\dot{\psi}|\sim |\hat{\omega}_d/\omega_{tr}|\gg1$ in the large orbit limit and having denoted derivation of $\psi(\theta)$ with respect to $\theta$ as $\dot \psi$ for brevity. Noting that $$\begin{aligned}
\int^{\theta}_{-\infty} e^{\psi(\theta')}d\theta'&=&\frac{e^{\psi}}{\dot{\psi}}-\frac{e^{\psi}}{\dot{\psi}}\frac{\partial}{\partial\theta}\frac{1}{\dot{\psi}}+\frac{e^{\psi}}{\dot{\psi}}\frac{\partial}{\partial\theta}\left(\frac{1}{\dot{\psi}}\frac{\partial}{\partial\theta}\frac{1}{\dot{\psi}}\right)\nonumber\\
&-& \int^{\theta}_{-\infty}e^{\psi(\theta')}\frac{\partial}{\partial\theta'}\left(\frac{1}{\dot{\psi}}\frac{\partial}{\partial\theta'}\left(\frac{1}{\dot{\psi}}\frac{\partial}{\partial\theta'}\frac{1}{\dot{\psi}}\right)\right)d\theta',\nonumber\end{aligned}$$ one then has $$\begin{aligned}
\delta H_s&=&\frac{\omega}{\omega_{tr}}\hat{S}\left[\frac{1}{\dot{\psi}} -\frac{1}{2}\frac{\partial}{\partial\theta}\left(\frac{1}{\dot{\psi}}\right)^2+\frac{1}{2\dot{\psi}}\frac{\partial^2}{\partial \theta^2}\left(\frac{1}{\dot{\psi}}\right)^2\right.\nonumber\\
&&\hspace{11em}\left.+ O(\dot{\psi}^{(-4)})\right].\label{eq:deltaH_asym}\end{aligned}$$
Noting that $\dot{\psi}=-i (\omega-\hat{\omega}_d\sin\theta)/\omega_{tr}$, the three terms in the square bracket of equation (\[eq:deltaH\_asym\]) corresponds, respectively, to $\delta H^{(0)}_{res}$, $\delta H^{(1)}_{res}$ and $\delta H^{(2)}_{res}$ in equations (16), (21) and (23) of Ref. [@ZQiuPPCF2009], in the $T_e/T_i\ll1$ limit assumed here. Thus, the $\delta H_{res}$’s in Ref. [@ZQiuPPCF2009] are equivalent to the general solution of equation (\[eq:Hi\_harmonic\]) by summing up all the transit harmonics, and the underlying wave-paricle interactions in the short wavelength limit are indeed through transit resonances, as pointed out in Ref. [@ZQiuPPCF2009]. The first term in equation (\[eq:deltaH\_asym\]) corresponds to the perturbed resonant particle distribution function in the $q\rightarrow\infty$ limit; the third term gives the $O(1/q^2)$ corrections while the second term vanishes in the surface average.
Since we are interested in the collisionless damping due to thermal ion contribution, a single thermal ion species with Maxwellian distribution function can be assumed, and the GAM dielectric function is derived from the surface averaged quasi-neutrality condition $$\begin{aligned}
D_G\equiv \left.\left\langle -\frac{e}{T_i}F_0\overline{\delta\phi}_G+J_G\overline{\delta H_i} \right\rangle\right/\left(\frac{e}{T_i}n_0\overline{\delta\phi}_G\right).\nonumber\end{aligned}$$ The imaginary part of $D_G$ due to resonant particle contribution, to the leading order, is then $$\begin{aligned}
D^{(0)}_i=\mathbb{I}{\rm m}\left\langle \frac{F_0}{n_0}J^2_G\omega\int \frac{d\theta}{2\pi}\frac{1}{\omega-\omega_d}\right\rangle.\label{eq:GAM_Di}\end{aligned}$$
We note that, even though in equation (\[eq:GAM\_Di\]) the anti-Herimitian part comes from the imaginary part of $1/(\omega-\omega_d)$, the underlying interaction is not a “drift resonance" [@MSasakiPoP2016], since $\omega_d\propto\sin\theta$ is temporally fast varying and the effective energy exchange is due to transit resonances as shown in equation (\[eq:Hi\_harmonic\]). The surface average is then carried out by expanding $\omega_d$ round $\theta=\pm \pi/2$ where $|\omega_d|$ is maximized and the integration in $\theta$ is performed by the method of steepest descent. Again, readers interested in the details of the algebra can consult Ref. . Here, we will briefly summarize the main ideas underlying the derivation:
- considering the wave-particle interaction on the time scale of $|\omega_d|^{-1}$, which is much shorter than the transit time $|\omega_{tr}|^{-1}$ in the large orbit limit, corresponds to the inclusion of a broad spectrum in frequency, i.e., all the transit harmonics are taken into account;
- for resonant particles, the dominant energy exchange with GAM is captured noting that the wave- particle energy exchange is caused by the acceleration in the radial direction associated with the radial magnetic drift, i.e., $\dot{E}=(e/m)\mathbf{V}_d\cdot\delta\mathbf{E}_r$, which maximises around $|\theta|=\pi/2$. Here, $V_d\equiv (v^2_{\perp}/2+v^2_{\parallel})\sin\theta \mathbf{e}_r/(\Omega_i R_0)$ is the radial component of magnetic drift velocity.
- Noting again that $\omega_d\propto\sin\theta$ is maximized around $|\theta|=\pi/2$, ions with lower energy and thus, proportionally (exponentially for a Maxwellian distribution with typical parameters) larger population, will contribute to the resonance.
As a further application, EGAM excitation by EPs in the large magnetic drift orbit limit will be investigated; which is part of the motivation of this communication. To focus on the wave-particle resonance in the short wavelength limit considering the effect of finite magnetic drift orbit averaging, we take $T_e/T_i\ll1$ and further neglect the finite Larmor radius effect of EPs. Thus, the leading order EP response to GAM can be derived as $$\begin{aligned}
\delta H_h=-\frac{e}{m}\partial_E F_{0h}\overline{\delta\phi}_G\frac{\omega}{\omega-\omega_d},\nonumber\end{aligned}$$ and the linear dispersion relation of EGAM can be obtained from the quasi-neutrality condition $$\begin{aligned}
\hat{\mathscr{E}}_{EGAM}\equiv\left.\left(\overline{\delta n_i}+\overline{\delta n_h}\right)\right/\left(en_0\overline{\delta\phi}_G/T_i\right).\nonumber\end{aligned}$$
As the expression of thermal ion density perturbation can be found in Ref. , we will focus on the EP density perturbation, $$\begin{aligned}
\overline{\delta n_h}
&=& -\frac{e}{m}B_0\sum_{\sigma=\pm1}\int \frac{E dE d\Lambda}{|v_{\parallel}|}\int d\theta \frac{\partial F_{0h}}{\partial E}\overline{\delta\phi}_G\frac{\omega_d}{\omega-\omega_d}.\nonumber\end{aligned}$$ Here, $\Lambda = \mu/E$ is the usual definition of the particle pitch angle in velocity space, with $\mu=v_\perp^2/(2B)$ the magnetic moment. Noting that $\omega_d=\hat{\omega}_d\sin\theta$ maximizes at $\theta\simeq \pi/2$, the contribution around $\theta\simeq \pm\pi/2$ dominates where wave-particle power exchange maximizes. Taking $x=\theta-\mbox{sign}(\theta) \pi/2$, one then has $$\begin{aligned}
\int d\theta\frac{\omega_d}{\omega-\omega_d}&=&-2\pi+\omega\int^{\infty}_{-\infty} dx \frac{1}{\omega-\hat{\omega}_d(1-x^2/2)}\nonumber\\
&=&-2\pi\frac{i}{\sqrt{(2\hat{\omega}_d/\omega)(\hat{\omega}_d/\omega-1)}}.\label{eq:surface_averaged}\end{aligned}$$ In equation (\[eq:surface\_averaged\]), the contribution of non-resonant adiabatic particle response is neglected, and the perturbed EP density is then $$\begin{aligned}
\overline{\delta n_h}&=&2\pi i B_0\frac{e}{m}\overline{\delta\phi}_G\nonumber\\
&\times& \sum_{\sigma=\pm1}\int\frac{EdEd\Lambda}{|v_{\parallel}|}\frac{\partial_EF_{0h}}{\sqrt{(2\hat{\omega}_d/\omega)(\hat{\omega}_d/\omega-1)}}.\end{aligned}$$ Taking a single-pitch angle slowing down EP distribution function [@ZQiuPPCF2010] as that for neutral beam injection, i.e., $F_{0h}=c_0\delta(\Lambda-\Lambda_0)H_E$, with $c_0=n_b\sqrt{2(1-\Lambda_0B_0)}/(4\pi B_0\ln(E_b/E_c))$, $n_b$ is the density of the EP beam, $E_b$ and $E_c$ being respectively the EP birth and critical energies, $\delta(x)$ is the Dirac delta function, and $H_E=1/(E^{3/2}+E^{3/2}_c)\Theta(1-E/E_b)$ with $\Theta(1-E/E_b)$ being the Heaviside step function. The integration in velocity space can then be carried out, and yields the short wavelength EGAM dispersion relation: $$\begin{aligned}
\hat{b}_i\left(-1+\frac{\omega^2_G}{\omega^2}\right)+\Delta_f&+& i n_b\left[\frac{-2+3\Lambda_0B_0}{1-\Lambda_0B_0}\frac{\Omega_b}{\omega}\sqrt{\frac{\Omega_b}{\omega}-1}\right.\nonumber\\
&&\left.-\Lambda_0B_0\frac{(\omega/\Omega_b)^{1/2}}{\sqrt{\Omega_b/\omega-1}}\right]=0,\label{eq:DR_final}\end{aligned}$$ with $\Delta_f$ being the non-resonant EP contribution $$\begin{aligned}
\Delta_f=n_b\left[\frac{-2+3\Lambda_0B_0}{1-\Lambda_0B}\frac{\Omega_b}{\omega}+\Lambda_0B_0\left(\frac{\omega}{\Omega_b}\right)^{1/2}\left(\frac{E_b}{E_c}\right)^{3/2}\right],\nonumber\end{aligned}$$ $\Omega_b\equiv\hat{\omega}_d(E=E_b)$, $\hat{b}_i=k^2_r\rho^2_{ti}/2$, and $\omega_G\simeq\sqrt{7/4+\tau}v_{ti}/R_0$ is the GAM frequency.
![Real frequency v.s. $\Omega_b/\omega_G$[]{data-label="fig:RF"}](EGAM_short_RF.eps){width="9cm"}
The first term in the square bracket of equation (\[eq:DR\_final\]) ($\propto\sqrt{\Omega_b/\omega-1}$) could be the destabilizing term depending on the value of $\Lambda_0B_0$, while the second term ($\propto(\Omega_b/\omega-1)^{-1/2}$) is stabilizing. As a result, EGAM excitation in the large orbit limit requires, first, $$\begin{aligned}
\Lambda_0B_0>2/3,\label{eq:drive}\end{aligned}$$ for the first term of EP contribution in equation (\[eq:DR\_final\]) to be destabilizing; and second, $\Omega_b/\omega$ being sufficiently large for the short wavelength EGAM to be unstable.
![Growth rate v.s. $\Omega_b/\omega_G$[]{data-label="fig:GR"}](EGAM_short_GR.eps){width="9cm"}
The dispersion relation is solved numerically as a function of $\Omega_b/\omega_G$. Note that from our previous analysis, the drive due to wave-particle resonance exists only for $|\Omega_b/\omega|>1$; and that the destabilizing term increases with $|\Omega_b/\omega|$ while the stabilizing term decreases with $|\Omega_b/\omega|$, so the destabilizing term is neglected, which gives negligible contribution for $|\Omega_b/\omega|>1$ where wave-particle energy exchange exists. The other parameters are taken as follows: $n_b/\hat{b}_i=0.3$, $\Delta_f=0$, and the obtained short wavelength EGAM real frequency and growth rate dependences on $\Omega_b/\omega_G$ are shown in Figs. \[fig:RF\] and \[fig:GR\], respectively. It is shown that, the EGAM real frequency decreases slightly with increasing $\Omega_b/\omega_G$, and the unstable EGAM frequency is always smaller than local GAM frequency. On the other hand, as the EGAM is unstable for $\Omega_b/\omega_G>1$, the growth rate increases with $\Omega_b$. For $\Omega_b/\omega_G$ significantly larger than unity, the growth rate increases almost linearly with $\Omega_b/\omega_G$, and thus, $E_b$, as is clearly seen from the destabilizing term of equation (\[eq:DR\_final\]). This is due to the increasingly dense high order transit resonances associated with increasing $E_b$. Whereas, in the long wavelength limit, the growth rate will be peaked when the EP parallel velocity at birth energy satisfies a certain harmonic resonance, similar to the case for GAM Landau damping discussed in Ref. [@HSugamaJPP2006].
In conclusion, the underlying physics picture of wave-particle resonances at short wavelength is clarified, taking short wavelength GAM collisionless damping as an example. Assuming large aspect ratio tokamak and well circulating particles, the ion response to GAM is derived from linear gyrokinetic equation by integration along unperturbed guiding-center orbit. The general solution is then obtained by expansion into transit harmonics, with the “population" of resonant particles to each transit harmonic proportional to $J^2_l(\hat{\Lambda}_{res})\partial_E F_0|_{v_{\parallel,res}}$. As a result, to obtain the GAM collisionless damping in the short wavelength limit, all the transit resonances must be kept. It is then shown that, the result obtained in Ref. based on large orbit width expansion, is equivalent to the general solution up to $O(1/(k_r\rho_{ti}q^2))$; and the underlying physics for wave-particle interactions at short wavelength consists indeed in the summation of all the transit resonances.
As a further application, the EGAM excitation at short wavelengths is also investigated, and the analytical dispersion relation is derived assuming a single pitch angle slowing down EP distribution function. Our results indicates that the short wavelength EGAM dispersion relation depends algebraically on the EP characteristic frequency, instead of the logarithmic dependence characterizing the long wavelength limit, which is typical for a slowing down EP distribution function. The short wavelength EGAM is unstable for $\Omega_b>\omega_G$, and $\Lambda_0B_0>2/3$. For $\Omega_b$ significantly larger than GAM frequency, the short wavelength EGAM growth rate is proportional to $\Omega_b$, and thus, EP birth energy $E_b$ due to the increasingly denser high order transit resonances as $\Omega_b\gg\omega_G$.
This work is supported by US DoE GRANT, the National Magnet Confinement Fusion Research Program under Grants Nos. 2013GB104004 and 2013GB111004, the National Science Foundation of China under grant Nos. 11575157 and 11235009, Fundamental Research Fund for Chinese Central Universities under Grant No. 2017FZA3004 and EUROfusion Consortium under grant agreement No. 633053.
[14]{} natexlab\#1[\#1]{}bibnamefont \#1[\#1]{}bibfnamefont \#1[\#1]{}citenamefont \#1[\#1]{}url \#1[`#1`]{}urlprefix\[2\][\#2]{} \[2\]\[\][[\#2](#2)]{}
, , , ****, ().
, , , , , , , ****, ().
, ****, ().
, , , ****, ().
, , , ****, ().
, , , ****, ().
, , , , , , , , , , , , , , , ****, ().
, ****, ().
, , , **** ().
, ****, ().
, ****, ().
, ****, ().
, ****, ().
, , , , , ****, ().
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The Relativistic Random Phase Approximation (RRPA) is derived from the Time-dependent Relativistic Mean Field (TD RMF) theory in the limit of small amplitude oscillations. In the [*no-sea*]{} approximation of the RMF theory, the RRPA configuration space includes not only the ususal particle-hole $ph$-states, but also $\alpha h$-configurations, i.e. pairs formed from occupied states in the Fermi sea and empty negative-energy states in the Dirac sea. The contribution of the negative energy states to the RRPA matrices is examined in a schematic model, and the large effect of Dirac sea states on isoscalar strength distributions is illustrated for the giant monopole resonance in $^{116}$Sn. It is shown that, because the matrix elements of the time-like component of the vector meson fields which couple the $\alpha h$-configurations with the $ph$-configurations are strongly reduced with respect to the corresponding matrix elements of the isoscalar scalar meson field, the inclusion of states with unperturbed energies more than 1.2 GeV below the Fermi energy has a pronounced effect on giant resonances with excitation energies in the MeV region. The influence of nuclear magnetism, i.e. the effect of the spatial components of the vector fields is examined, and the difference between the non-relativistic and relativistic RPA predictions for the nuclear matter compression modulus is explained.'
address: |
1. Physics Department, TU Munich, D-85748 Garching, Germany\
2. China Institute of Atomic Energy, Beijing, P.R. of China\
3. Institut de Physique Nucléaire, IN2P3-CNRS, F-91406 Orsay Cedex, France\
author:
- |
P. Ring$^{1}$, Zhong-yu Ma$^{2}$[^1] , Nguyen Van Giai$^3$, D. Vretenar$^{1}$[^2] ,\
A. Wandelt$^{1}$, and Li-gang Cao$^{2}$
date: 'January 25, 2001'
title: 'The time-dependent relativistic mean-field theory and the random phase approximation'
---
Introduction
============
Models based on the Relativistic Mean Field (RMF) approximation provide a microscopic self-consistent description of nuclear structure phenomena (for reviews, see Refs. [@SW.86; @Rei.89; @Ser.92; @Rin.96]). In this framework classical equations of motion are derived self-consistently from a fully relativistic Lagrangian. Vacuum polarization effects, as well as Fock exchange terms, are usually not taken into account explicitly. This framework, however, is based on an effective theory and the parameters of effective interactions are determined from a set of experimental data. In adjusting the parameters of the effective Lagrangian, a large part of vacuum polarization effects and effects of exchange terms are already taken into account. In fact, both contributions have been treated explicitly in nuclear matter and in some finite spherical nuclei [@HS.83; @BMG.87; @BFG.93; @HS.84; @Was.91], and it has been shown that these contributions are not small. However, by renormalizing the parameters of the effective Lagrangian, virtually identical results for ground-state properties have been obtained without the inclusion of vacuum polarization and exchange terms. Essential for a quantitative description of properties of complex nuclei are the non-linear terms in the meson sector [@BB.77], which in a simple way include an effective density dependence of the meson coupling parameters.
The relativistic mean-field models have been mostly applied in the description of ground-state properties of nuclei all over the periodic table. In several cases this framework has also been very successfully applied to excited states. For example, the cranked relativistic mean-field model [@KR.89] describes a large variety of phenomena in rotational bands of superdeformed nuclei [@AKR.96; @ALR.98; @AKR.99]. Another example is the Time-dependent RMF model which has been used to describe the dynamics of giant resonances in nuclei [@VBR.95; @PVR.96; @VLB.97]. From the time evolution of multipole moments of the single-particle density, the excitation energies of giant resonances can be determined. Excellent agreement with experimental values of isoscalar and isovector giant monopole, giant quadrupole and the isovector giant dipole resonances has been obtained. The disadvantage of the time-dependent approach is that in some cases the large computational effort prevents an accurate description of excited states as, for instance, the dynamics of low-lying collective excitations.
The Relativistic Random Phase Approximation (RRPA) represents the small amplitude limit of the time-dependent relativistic mean-field theory. Some of the earliest applications of the RRPA [@Fur.85; @HG.89; @BC.88; @DF.90; @HP.90] to finite nuclei include the description of low-lying negative parity excitations in $^{16}$O [@Fur.85], and studies of isoscalar giant resonances in light and medium nuclei [@HG.89]. These RRPA calculations, however, were based on the most simple, linear $\sigma - \omega$ relativistic mean field model. Only recently non-linear meson self-interaction terms have been taken included in RRPA calculations [@MGT.97; @MTG.97]. However, in these calculations the RRPA configuration space included only ordinary particle-hole pairs. This seems a reasonable approximation, since the states formed from occupied positive-energy states in the Fermi sea and empty negative energy states in the Dirac sea, have unperturbed energies more than 1.2 GeV below the Fermi level. It turned out, however, that excitation energies of isoscalar resonances calculated in this way were very different from those obtained in the TD RMF approach with the same effective interactions [@GMT.99; @VRL.99].
Since it is well known that in the non-relativistic framework the RPA corresponds to the small amplitude limit of time-dependent Hartree-Fock (TDHF) (see, e.g., Ref. [@RS.80]), these discrepancies remained an open puzzle for a couple of years. Only recently it has been shown [@VWR.00; @MGW.00] that, in order to reproduce results of time-dependent relativistic mean-field calculations for giant resonances, the RRPA configuration space must contain negative-energy states from the Dirac sea. In principle, if the Dirac sea were fully occupied, these configurations would be forbidden by the Pauli principle. However, in the [*no-sea*]{} approximation the negative-energy states do not contribute to the nucleon densities, i.e. these states are not occupied. It is, thus, possible to form $\alpha h$ pairs ($\alpha$ empty state in the Dirac sea, $h$ occupied state in the Fermi sea) and include them in the RRPA configuration space. Although formally possible, and also necessary in order to preserve symmetries, this procedure raises some serious conceptual problems because the $\alpha h$ configurations have negative unperturbed excitations energies. This means that the energy surface is no longer positive definite. The static solutions of the RMF equations correspond to saddle points on the energy surface, rather than to minima as in non-relativistic Hartree-Fock calculations. It is not, therefore, a priori clear that small amplitude oscillations around these stationary solutions will be stable. Furthermore, it is not obvious why configurations with unperturbed energies more than 1.2 GeV below the Fermi level have such a pronounced effect on the excitation energies of giant resonances in the MeV region. In non-relativistic calculations, for instance, it has been noted [@RS.74] that $ph$-configurations with very large excitation energies affect the position of the spurious mode, but they have no effect on the excitation energies of giant resonances.
The purpose of the present investigation is to clarify some of these problems and to understand in a better way the relation between the RRPA and the TD RMF in the [*no-sea*]{} approximation. The paper is organized as follows: the time-dependent RMF model is analyzed in Sec. II. In Sec. III the RRPA is derived from the TDRMF equations in the limit of small amplitude motion. In particular, we discuss the importance of $\alpha h$-pairs in the RRPA configuration space. In Sec. IV we introduce a relativistic extension of the Brown-Bolsterli model and solve it in the linear response approximation. It is shown how large matrix elements, which couple the $\alpha h$-sector with the $ph$-configurations, can arise in the RRPA matrices. The large effect of Dirac sea states on isoscalar strength distributions is illustrated in Sec. V. The results are summarized in Sec. VI.
The time-dependent relativistic mean-field model
================================================
In quantum hadrodynamics the nucleus is described as a system of Dirac nucleons which interact through the exchange of virtual mesons and photons. The model is based on the one-boson exchange description of the nucleon-nucleon interaction. The Lagrangian density reads [@SW.86] $$\begin{aligned}
{\cal L} &=&\bar{\psi}\left( i\gamma \cdot \partial -m\right) \psi ~+~\frac{1}{2}\partial \sigma \cdot \partial \sigma -\frac{1}{2}m_{\sigma }\sigma ^{2}
\nonumber \\
&&-~\frac{1}{4}\Omega _{\mu \nu }\Omega ^{\mu \nu }+\frac{1}{2}m_{\omega
}^{2}\omega ^{2}~-~\frac{1}{4}{\vec{{\rm R}}}_{\mu \nu }{\vec{{\rm R}}}^{\mu
\nu }+\frac{1}{2}m_{\rho }^{2}\vec{\rho}^{\,2}~-~\frac{1}{4}{\rm F}_{\mu \nu
}{\rm F}^{\mu \nu } \nonumber \\
&&-~\bar{\psi}[g_{\sigma }\sigma ~+~g_{\omega }\gamma \cdot \omega
~+~g_{\rho }\gamma \cdot \vec{\rho}\vec{\tau} ~+~e\gamma \cdot A \frac{(1-\tau_{3})}{2}]\psi~. \label{E2.1}\end{aligned}$$ Vectors in isospin space are denoted by arrows, and bold-faced symbols will indicate vectors in ordinary three-dimensional space; a dot denotes the scalar product in Minkowski space ($\gamma \cdot
\omega =\gamma^\mu\omega_\mu
=\gamma _{0}\omega _{0}-{{\mbox{\boldmath $\gamma$}}}{{\mbox{\boldmath $\omega$}}}$). The Dirac spinor $\psi$ denotes the nucleon with mass $m$. $m_\sigma$, $m_\omega$, and $m_\rho$ are the masses of the $\sigma$-meson, the $\omega$-meson, and the $\rho$-meson, and $g_\sigma$, $g_\omega$, and $g_\rho$ are the corresponding coupling constants for the mesons to the nucleon, and $e^{2}/\hbar c =1/137.036$. Eq.(\[E2.1\]), $\Omega ^{\mu \nu }$, $\vec{R}^{\mu \nu }$, and $F^{\mu \nu }$ denote the field tensors of the vector fields $\omega $, $\rho $, and of the photon, respectively: $$\begin{aligned}
\Omega ^{\mu \nu } &=&\partial ^{\mu }\omega ^{\nu }-\partial ^{\nu }\omega
^{\mu }~, \nonumber \\
\vec{R}^{\mu \nu } &=&\partial ^{\mu }\vec{\rho}^{\,\nu }-\partial ^{\nu }\vec{\rho}^{\,\mu }~, \nonumber \\
F^{\mu \nu } &=&\partial ^{\mu }A^{\nu }-\partial ^{\nu }A^{\mu }~.
\label{E2.2}\end{aligned}$$ If the bare masses $m$, $m_{\omega }$, and $m_{\rho }$ are used for the nucleons and the $\omega $ and $\rho $ mesons, there are only four free model parameters: $m_{\sigma }$, $g_{\sigma }$, $g_{\omega }$ and $g_{\rho }$. Their values can be adjusted to experimental data of just few spherical nuclei. This simple model, however, is not flexible enough for a quantitative description of properties of complex nuclear systems. An effective density dependence has been introduced [@BB.77] by replacing the quadratic $\sigma $-potential $\frac{1}{2}m^2_{\sigma }\sigma ^{2}$ with a quartic potential $U(\sigma )$ $$U(\sigma )~=~\frac{1}{2}m_{\sigma }^{2}\sigma ^{2}+\frac{1}{3}g_{2}\sigma
^{3}+\frac{1}{4}g_{3}\sigma ^{4}~. \label{E2.3}$$ The potential includes the nonlinear $\sigma $ self-interaction, with two additional parameters $g_{2}$ and $g_{3}$. The corresponding Klein-Gordon equation becomes nonlinear, with a $\sigma $-dependent mass $m^2_{\sigma }(\sigma )=$ $m^2_{\sigma }+g_{2}\sigma
+g_{3}\sigma ^{2}$. More details on the relativistic mean-field formalism can be found in Refs. [@SW.86; @Rei.89; @Ser.92; @Rin.96].
From the Lagrangian density the set of coupled equations of motion is derived. The Dirac equation for the nucleons $$\begin{aligned}
i\partial _{t}\psi _{i} &=&\hat{h}(t)\psi _{i} \nonumber \\
&=&\left\{{{\mbox{\boldmath $\alpha$}}}\lbrack-i{{\mbox{\boldmath $\nabla$}}}-
{\bf V(r},t{\bf )]+}V({\bf r},t)+
{\bf \beta }\big(m-S({\bf r},t)\big)\right\} \psi _{i}.
\label{E2.4}\end{aligned}$$ If one neglects retardation effects for the meson fields, the time-dependent mean-field potentials $$\begin{aligned}
S({\bf r},t) &=&-g_{\sigma }\sigma ({\bf r},t)~, \nonumber \\
V_{\mu }({\bf r},t) &=&g_{\omega }\omega _{\mu }({\bf r},t)+g_{\rho }\vec{\tau}\vec{\rho}_{\mu }({\bf r},t)+eA_{\mu }({\bf r},t)\frac{(1-\tau _{3})}{2}~, \label{E2.5}\end{aligned}$$ are calculated at each step in time from the solution of the stationary Klein-Gordon equations $$\begin{aligned}
\left[ -\Delta +m_{\sigma }^{2}\right] \,\sigma ({\bf r,}t) &=&-g_{\sigma
}\,\rho _{s}({\bf r,}t)-g_{2}\,\sigma ^{2}({\bf r,}t)-g_{3}\,\sigma ^{3}({\bf r,}t)~, \nonumber \\
\left[ -\Delta +m_{\omega }^{2}\right] \,\omega _{\mu }({\bf r,}t)
&=&g_{\omega }\,j_{\mu }({\bf r,}t)~, \nonumber \\
\left[ -\Delta +m_{\rho }^{2}\right] \,\vec{\rho}_{\mu }({\bf r,}t)
&=&g_{\rho }\,\,\vec{j}_{\mu }({\bf r},t)~, \nonumber \\
-\Delta \,A_{\mu }({\bf r,}t) &=&e\,j_{c\mu }({\bf r},t)~. \label{E2.6}\end{aligned}$$ This approximation is justified by the large masses in the meson propagators. Retardation effects can be neglected because of the short range of the corresponding meson exchange forces. In the mean-field approximation only the motion of the nucleons is quantized, the meson degrees of freedom are described by classical fields which are defined by the nucleon densities and currents. The single-particle spinors $\psi_i~(i=1,2,...,A)$ form the A-particle Slater determinant $|\Phi(t)\rangle$. The nucleons move independently in the classical meson fields, i.e. residual two-body correlations are not included, and the many-nucleon wave function is a Slater determinant at all times. The sources of the fields in the Klein-Gordon equations are the nucleon densities and currents calculated in the [*no-sea*]{} approximation $$\begin{aligned}
\rho _{s}({\bf r},t) &=&\sum\limits_{i=1}^{A}\bar{\psi}_{i}^{{}}({\bf r},t)\psi _{i}^{{}}({\bf r},t)~, \nonumber \\
j_{\mu }({\bf r},t) &=&\sum\limits_{i=1}^{A}\bar{\psi}_{i}^{{}}({\bf r},t)\gamma _{\mu }\psi _{i}^{{}}({\bf r},t)~, \nonumber \\
\vec{j}_{\mu }({\bf r},t) &=&\sum\limits_{i=1}^{A}\bar{\psi}_{i}^{{}}({\bf r},t)\vec{\tau}\gamma _{\mu }\psi _{i}^{{}}({\bf r},t)~, \nonumber \\
j_{c\mu }({\bf r},t) &=&\sum\limits_{i=1}^{Z}\bar{\psi}_{i}^{{}}({\bf r},t)\gamma _{\mu }\psi _{i}^{{}}({\bf r},t)~. \label{E2.7}\end{aligned}$$ where the summation is over all occupied states in the Fermi sea. In the [*no-sea*]{} approximation the negative-energy states do not contribute to the densities and currents, i.e. vacuum polarization is explicitly neglected. However, as already discussed in the Introduction, this is an effective theory with the parameters of the Lagrangian determined from a set of experimental data. In adjusting the parameters of the effective Lagrangian, a large part of vacuum polarization effects is therefore already taken into account. It should be emphasized that the [*no-sea*]{} approximation is essential for practical applications of the relativistic mean-field model.
The stationary solutions of the relativistic mean-field equations describe the ground-state of a nucleus. They correspond to stationary points on the relativistic energy surface. The Dirac sea, i.e. the negative energy eigenvectors of the Dirac hamiltonian, is different for different nuclei. This means that it depends on the specific solution of the set of non-linear RMF equations. The Dirac spinors which describe the ground-state of a finite nucleus (positive energy states) can be expanded, for instance, in terms of vacuum solutions, which form a complete set of plane wave functions in spinor space. This set is only complete, however, if in addition to the positive energy states, it also contains the states with negative energies, in this case the Dirac sea of the vacuum. Positive energy solutions of the RMF equations in a finite nucleus automatically contain vacuum components with negative energy. In the same way, solutions which describe excited states, as for instance states with different angular momenta which are solutions of the cranked RMF equations, contain negative energy components which correspond to the ground-state solution.
This is also true for the solutions of the time-dependent problem. Although for the stationary solutions the negative-energy states do not contribute to the densities in the [*no-sea*]{} approximation, their contribution is implicitly included in the time-dependent calculation. The coupled system of RMF equations describes the time-evolution of A nucleons in the effective mean-field potential. Starting from the self-consistent solution which describes the ground-state of the nucleus, initial conditions can be defined which correspond, for instance, to excitations of giant resonances in experiments with electromagnetic or hadron probes. For example, the one-body proton and neutron densities can be initially deformed and/or given some initial velocities. The resulting mean-field dynamics can be described by the time-evolution of the collective variables. In coordinate space for example, these will be the multipole moments of the density distributions. At each time $t$, the Dirac spinors $\psi _{i}(t)$ can be expanded in terms of the complete set of solutions of the stationary Dirac equation $\psi _{k}^{(0)}$ $$\psi _{i}({\bf r},t)=\sum_{k}c_{k}(t)\psi _{k}^{(0)}({\bf r})\text{e}^{-i\varepsilon _{k}t}+\sum_{\alpha}c_{\alpha }(t)\psi _{\alpha }^{(0)}({\bf r})\text{e}^{-i\varepsilon _{\alpha }t}~, \label{E2.9}$$ where the index $k$ runs over all positive energy eigen-solutions $\varepsilon _{k}>0$ (hole states $h$ in the Fermi sea, and particle states $p$ above the Fermi sea), and the index $\alpha $ denotes eigen-solutions with negative energy $\varepsilon _{\alpha }<0$. We follow the time evolution of $A$ Dirac spinors which at time $t=0$ form the Fermi sea of the stationary solution. This means that at each time we have a [*local*]{} Fermi sea of $A$ time-dependent spinors which, of course, contain components of negative-energy solutions of the stationary Dirac equation. One could also start with the infinitely many negative energy solutions $\psi _{\alpha}({\bf r},t=0)$ ($\varepsilon_\alpha <0$), and propagate them in time with the same hamiltonian $\hat{h}(t)$. Since the time-evolution operator is unitary [@RS.80] $$i\partial _{t}\left\langle \psi _{i}|\psi _{a}\right\rangle =\left\langle
\psi _{i}|h^{\dagger }-h|\psi _{a}\right\rangle =0, \label{E2.9a}$$ the states which form the [*local*]{} Dirac sea are orthogonal to the [*local*]{} Fermi sea at each time. This is the meaning of the [*no-sea*]{} approximation in the time-dependent problem. For small-amplitude oscillations around the stationary solution, the coefficients $c_{\alpha }(t)$ of the negative energy components in (\[E2.9\]) are, of course, small.
We will first consider linear relativistic mean-field models ($g_{2}=g_{3}=0$ in (\[E2.3\])). In the instantaneous approximation, i.e. neglecting the time derivatives $\partial_t^2$ in the Klein-Gordon equations, the solutions for the mean-fields are calculated from $$\begin{aligned}
\sigma ({\bf r},t) &=&g_{\sigma }\int D_{\sigma }({\bf r,r}^{\prime })\rho
_{s}({\bf r}^{\prime },t)d^{3}r~, \nonumber \\
\omega _{\mu }({\bf r},t) &=&g_{\omega }\int D_{\omega }({\bf r,r}^{\prime
})j_{\mu }({\bf r}^{\prime },t)d^{3}r~, \nonumber \\
\vec{\rho}_{\mu }({\bf r},t) &=&g_{\rho }\int D_{\rho }({\bf r,r}^{\prime })\vec{j}_{\mu }({\bf r}^{\prime },t)d^{3}r~, \nonumber \\
A_{\mu }({\bf r},t) &=&e\int D_{photon}({\bf r,r}^{\prime })j_{c\mu }({\bf r}^{\prime },t)d^{3}r~. \label{E2.10}\end{aligned}$$ The propagators have the Yukawa form $$D_{\phi }({\bf r},{\bf r}^{\prime })\,=\pm \frac{1}{4\pi }\frac{\text{e}^{-m_{\phi }|{\bf r}-{\bf r}^{\prime }|}}{|{\bf r}-{\bf r}^{\prime }|}~,
\label{E2.11}$$ where $\phi $ denotes the mesons $\sigma $, $\omega $, $\rho $, and the photon. The plus (minus) sign is for vector (scalar) fields. In the non-linear case an analytic solution of the Klein-Gordon equation is, of course, no longer possible. The corresponding meson field is a non-linear functional of the density and currents.
The relativistic single-particle density matrix reads $$\hat{\rho}({\bf r},{\bf r}^{\prime },t)=\sum\limits_{i=1}^{A}|\psi _{i}^{{}}({\bf r},t)\rangle \langle \psi _{i}^{{}}({\bf r}^{\prime },t)|~.
\label{E2.12}$$ If the Dirac spinor is written in terms of large and small components $$|\psi _{i}^{{}}({\bf r},t)\rangle =\left(
\begin{array}{c}
\,\,\,\,f_{i}({\bf r},t) \\
ig_{i}({\bf r},t)
\end{array}
\right), \label{E2.13}$$ the density matrix takes the form $$\rho ({\bf r},{\bf r}^{\prime },t)=\left(
\begin{array}{cc}
\,\,\,\sum\limits_{i=1}^{A}f_{i}^{{}}({\bf r},t)f_{i}^{\dagger }({\bf r}^{\prime },t) & -i\sum\limits_{i=1}^{A}f_{i}^{{}}({\bf r},t)g_{i}^{\dagger }({\bf r}^{\prime },t) \\
i\sum\limits_{i=1}^{A}g_{i}^{{}}({\bf r},t)f_{i}^{\dagger }({\bf r}^{\prime
},t) & \,\,\,\,\,\,\sum\limits_{i=1}^{A}g_{i}^{{}}({\bf r},t)g_{i}^{\dagger
}({\bf r}^{\prime },t)
\end{array}
\right) ~. \label{E2.14}$$ Further, a relativistic two-body interaction is defined $$\hat{V}=\int d^{3}r_{1}d^{3}r_{2}\hat{\psi}^{\dagger }({\bf r}_{1})\hat{\psi}^{\dagger }({\bf r}_{2})V({\bf r}_{1},{\bf r}_{2})\hat{\psi}({\bf r}_{1})\hat{\psi}({\bf r}_{2})~, \label{E2.15}$$ where $\hat{\psi}^{\dagger }$and $\hat{\psi}$ are the Dirac field creation and annihilation operators, and $$V({\bf r}_{1},{\bf r}_{2})=D_{\sigma }({\bf r}_{1},{\bf r}_{2})\,\beta
^{(1)}\beta ^{(2)}+D_{\omega }^{{}}({\bf r}_{1},{\bf r}_{2})
\left( 1-{{\mbox{\boldmath $\alpha$}}}^{(1)}{{\mbox{\boldmath $\alpha$}}}^{(2)}\right) ~.
\label{E2.16}$$ In order to simplify the notation, we omit the $\rho $-meson and the photon, though they are, of course, included in actual applications of the relativistic mean-field model. Their contribution to the matrix elements of $V({\bf r}_{1},{\bf r}_{2})$ is, however, much smaller than that of the $\sigma$ and $\omega$ mesons.
By introducing an arbitrary complete spinor basis (the indices $k,l,\dots $ denote both positive and negative energy states), the two-body interaction operator can be written in the form $$\hat{V}=\frac{1}{2}\sum_{kk^{\prime }ll^{\prime }}V_{klk^{\prime }l^{\prime
}}\hat{\psi} _{k}^{\dagger }\hat{\psi} _{l}^{\dagger }
\hat{\psi} _{l^{\prime }}^{{}}\hat{\psi}
_{k^{\prime }}^{{}}~. \label{E2.17}$$ The single-particle equation of motion corresponds to the time-dependent relativistic Hartree problem $$i\partial _{t}\psi _{i}=\hat{h}(\hat{\rho})\psi _{i}~, \label{E2.18}$$ with the Dirac Hamiltonian $$\hat{h}(\hat{\rho})=
{{\mbox{\boldmath $\alpha$}}}\,{\bf p} + \beta(m+\,\Sigma(\hat{\rho})),
\label{E2.19}$$ and the mass operator $$\Sigma _{kl}(\hat{\rho})=\sum_{k^{\prime }l^{\prime }}V_{kl^{\prime
}lk^{\prime }}\rho _{k^{\prime }l^{\prime }}~. \label{E2.20}$$ The corresponding equation of motion for the density operator reads $$i\partial _{t}\hat{\rho}=\left[ \hat{h}(\hat{\rho}),\hat{\rho}\right] ~,
\label{E2.21}$$ in full analogy with the non-relativistic Hartree-Fock problem (see, e.g., Ref. [@RS.80]).
In expressing the TD RMF equations (\[E2.4\]-\[E2.6\]) in terms of a relativistic two-body interaction, we have eliminated the meson degrees of freedom by using the Yukawa form (\[E2.11\]) of the meson propagators. This applies, of course, only to Lagrangians that do not contain non-linear meson self-interactions. The non-linear couplings are, however, essential for a realistic description of nuclear properties. Formally, also in this case the Klein-Gordon equations can be solved at each step in time, and the resulting meson fields are non-linear functionals of the densities and currents. The Dirac operator has still the form of Eq. (\[E2.19\]), but the mass-operator $\Sigma _{kl}(\hat{\rho})$ becomes a much more complicated functional of the single-particle density.
The numerical solution of the full time-dependent problem with non-linear meson self-interactions does not present particular difficulties (see Refs. [@VBR.95; @PVR.96; @VLB.97]). Much more difficult, however, is to eliminate the meson degrees of freedom and to derive a relativistic two-body interaction in the general case of large amplitude motion. This has only been done in the small amplitude limit [@MGT.97]. The $\sigma $-field and the scalar density $\,\rho _{s}$ are expanded in the neighborhood of the stationary ground-state solutions $$\begin{aligned}
\sigma ({\bf r,}t{\bf )} &=&{\bf \,}\sigma ^{(0)}({\bf r)+}\delta \sigma ({\bf r,}t{\bf )}~, \label{E2.22} \\
\,\rho _{s}({\bf r,}t{\bf )} &=&{\bf \,}\rho _{s}^{(0)}({\bf r)+}\delta \rho
_{s}({\bf r,}t{\bf )}~.\end{aligned}$$ The corresponding Klein-Gordon equation (\[E2.6\]) for the $\sigma $-field is solved by linearization, i.e. up to terms linear in $\delta \sigma$ we obtain $$\left[ -\Delta +m_{\sigma }^{2}({\bf r})\right] \delta \,\sigma ({\bf r,}t)=-g_{\sigma }\delta \,\rho _{s}({\bf r,}t)~, \label{E2.23}$$ with $$m_{\sigma }^{2}({\bf r})=\left. \frac{\partial ^{2}U}{\partial \sigma ^{2}}\right| _{\sigma =\sigma ^{(0)}({\bf r})}~. \label{E2.24}$$ The $\sigma $-meson propagator is defined by the equation $$\left[ -\Delta +m_{\sigma }^{2}({\bf r})\right] D_{\sigma }({\bf r},{\bf r}^{\prime })=-\delta ({\bf r-r}^{\prime })~, \label{E2.25}$$ and it has been determined numerically in the RRPA calculations of Refs. ( [@MGT.97; @MTG.97; @GMT.99]). In the following only the small amplitude limit will be studied, and therefore we do not need to worry about the more general problem of large amplitude motion.
The small amplitude limit of TD RMF and the relativistic RPA
============================================================
In this section we study the response of the density matrix $\hat{\rho}(t)$ to an external one-body field $$\hat{F}(t)=\hat{F}\text{e}^{-i\omega t}+h.c.~, \label{E3.1}$$ which oscillates with a small amplitude. Assuming that in the single-particle space this field is represented by the operator $$\hat{f}(t) = \sum_{kl} \, f_{kl}(t) \; \hat{a}^{\dagger}_k \hat{a}^{}_l ,$$ the equation of motion for the density operator is $$i\partial _{t}\hat{\rho}=\left[ \hat{h}(\hat{\rho})+\hat{f}(t),\hat{\rho}\right] ~, \label{E3.2}$$ In the linear approximation the density matrix is expanded $$\hat{\rho}(t)=\hat{\rho}^{(0)}+\delta \hat{\rho}(t)~, \label{E3.3}$$ where $\hat{\rho}^{(0)}$ is the stationary ground-state density. From the definition of the density matrix (\[E2.12\]), it follows that $\hat{\rho}\left( t\right) $ is a projector at all times, i.e. $\hat{\rho}\left( t\right) ^{2}=$ $\hat{\rho}\left( t\right) $. In particular, this means that the eigenvalues of $\hat{\rho}^{(0)}$ are 0 and 1. In the non-relativistic case particle states above the Fermi level correspond to the eigenvalue 0, and hole states in the Fermi sea correspond to the eigenvalue 1. In the relativistic case, one also has to take into account states from the Dirac sea. In the [*no-sea*]{} approximation these states are not occupied, i.e. they correspond to the eigenvalue 0 of the density matrix. We will work in the basis which diagonalizes $\hat{\rho}^{(0)}$ $$\rho _{kl}^{(0)}=\delta _{kl}\rho _{k}^{(0)}=\left\{
\begin{array}{ll}
0 & \text{for unoccupied states above the Fermi level (index }p\text{)}
\\
1 & \text{for occupied states in the Fermi sea (index }h\text{)\quad } \\
0 & \text{for unoccupied states in the Dirac sea (index }\alpha \text{)}
\end{array}
\right. \label{E3.4}$$ Since $\hat{\rho}(t)$ is a projector at all times, in linear order $$\hat{\rho}^{(0)}\delta \hat{\rho}+\delta \hat{\rho}\hat{\rho}^{(0)}=\delta
\hat{\rho}~. \label{E3.5}$$ This means that the non-vanishing matrix elements of $\delta \hat{\rho}$ are: $\delta \rho _{ph}$, $\delta \rho_{hp}$, $\delta \rho _{\alpha h}$, and $\delta \rho _{h\alpha }$. These are determined by the solution of the TD RMF equation (\[E3.2\]). In the linear approximation the equation of motion reduces to $$i\partial _{t}\delta \hat{\rho}=\left[ \hat{h}^{(0)},\delta \hat{\rho}\right]
+\left[ \frac{\partial \hat{h}}{\partial \rho }\delta \rho ,\hat{\rho}^{(0)}\right] +\left[ \hat{f},\hat{\rho}^{(0)}\right] ~, \label{E3.6}$$ where $$\frac{\partial \hat{h}}{\partial \rho }\delta \rho =\sum_{ph}\frac{\partial
\hat{h}}{\partial \rho _{ph}}\delta \rho _{ph}+\frac{\partial \hat{h}}{\partial \rho _{hp}}\delta \rho _{hp}+\sum_{\alpha h}\frac{\partial \hat{h}}{\partial \rho _{\alpha h}}\delta \rho _{\alpha h}+\frac{\partial \hat{h}}{\partial \rho _{h\alpha }}\delta \rho _{h\alpha }~. \label{E3.7}$$ In the small amplitude limit $\delta \rho$ will, of course, also display a harmonic time dependence e$^{-i\omega t}$. Taking into account the fact that $\hat{h}_{kl}^{(0)}=\delta _{kl}\epsilon
_{k}$ is diagonal in the stationary basis, we obtain $$\begin{aligned}
(\omega -\epsilon _{p}+\epsilon _{h})\delta \rho _{ph}
&=&f_{ph}+\sum_{p^{\prime }h^{\prime }}V_{ph^{\prime }hp^{\prime }}\delta
\rho _{p^{\prime }h^{\prime }}+V_{pp^{\prime }hh^{\prime }}\delta \rho
_{h^{\prime }p^{\prime }}+\sum_{\alpha^{\prime }h^{\prime }}V_{ph^{\prime
}h\alpha^{\prime }}\delta \rho _{\alpha^{\prime }h^{\prime }}+
V_{p\alpha^{\prime }hh^{\prime }}
\delta \rho_{h^{\prime }\alpha^{\prime }}
\nonumber
\\
(\omega -\epsilon_{\alpha }+\epsilon _{h})\delta \rho _{\alpha h}
&=&f_{\alpha h}+\sum_{p^{\prime }h^{\prime }}V_{\alpha h^{\prime }hp^{\prime
}}\delta \rho _{p^{\prime }h^{\prime }}+V_{\alpha p^{\prime }hh^{\prime
}}\delta \rho _{h^{\prime }p^{\prime }}+\sum_{\alpha ^{\prime }h^{\prime
}}V_{\alpha h^{\prime }h\alpha ^{\prime }}\delta \rho _{\alpha ^{\prime
}h^{\prime }}+V_{\alpha \alpha ^{\prime }hh^{\prime }}\delta \rho
_{h^{\prime }\alpha ^{\prime }} \nonumber \\
(\omega -\epsilon _{h}+\epsilon _{p})\delta \rho _{hp}
&=&f_{hp}+\sum_{p^{\prime }h^{\prime }}V_{hh^{\prime }pp^{\prime }}\delta
\rho _{p^{\prime }h^{\prime }}+V_{hp^{\prime }ph^{\prime }}\delta \rho
_{h^{\prime }p^{\prime }}+\sum_{\alpha ^{\prime }h^{\prime }}V_{hh^{\prime
}p\alpha ^{\prime }}\delta \rho _{\alpha ^{\prime }h^{\prime }}+V_{h\alpha
^{\prime }ph^{\prime }}\delta \rho _{h^{\prime }\alpha ^{\prime }}
\nonumber
\\
(\omega -\epsilon _{h}+\epsilon_{\alpha })\delta \rho _{h\alpha }
&=&f_{h\alpha }+\sum_{p^{\prime }h^{\prime }}V_{hh^{\prime }
\alpha p^{\prime}}\delta \rho _{p^{\prime }h^{\prime }}+
V_{hp^{\prime }\alpha h^{\prime}}
\delta \rho _{h^{\prime }p^{\prime }}+
\sum_{\alpha ^{\prime }h^{\prime}}
V_{hh^{\prime }\alpha \alpha ^{\prime }}
\delta \rho _{\alpha ^{\prime}h^{\prime }}+
V_{h\alpha ^{\prime }\alpha h^{\prime }}
\delta \rho_{h^{\prime }\alpha ^{\prime }}
\label{E3.8}\end{aligned}$$ or, in matrix form $$\left[ \omega \left(
\begin{array}{cc}
1 & 0 \\
0 & -1
\end{array}
\right) -\left(
\begin{array}{cc}
A & B \\
B^{\ast } & A^{\ast }
\end{array}
\right) \right] \left(
\begin{array}{c}
X \\
Y
\end{array}
\right) =\left(
\begin{array}{c}
F \\
\bar{F}
\end{array}
\right) ~, \label{E3.9}$$ The RRPA matrices $A$ and $B$ read $$\begin{aligned}
A &=&\left(
\begin{array}{cc}
(\epsilon _{p}-\epsilon _{h})\delta _{pp^{\prime }}\delta _{hh^{\prime }} &
\\
& (\epsilon _{\alpha }-\epsilon _{h})\delta _{\alpha \alpha ^{\prime
}}\delta _{hh^{\prime }}
\end{array}
\right) +\left(
\begin{array}{cc}
V_{ph^{\prime }hp^{\prime }} & V_{ph^{\prime }h\alpha ^{\prime }} \\
V_{\alpha h^{\prime }hp^{\prime }} & V_{\alpha h^{\prime }h\alpha ^{\prime }}
\end{array}
\right) \label{E3.10} \\
B &=&\left(
\begin{array}{cc}
V_{pp^{\prime }hh^{\prime }} & V_{p\alpha ^{\prime }hh^{\prime }} \\
V_{\alpha p^{\prime }hh^{\prime }} & V_{\alpha \alpha ^{\prime }hh^{\prime }}
\end{array}
\right) \label{E3.11}\end{aligned}$$ and the amplitudes $X$ and $Y$ are defined $$X=\left(
\begin{array}{c}
\delta \rho _{ph} \\
\delta \rho _{\alpha h}
\end{array}
\right) ,\quad Y=\left(
\begin{array}{c}
\delta \rho _{hp} \\
\delta \rho _{h\alpha }
\end{array}
\right) ~. \label{E3.12}$$ The vectors which represent the external field contain the matrix elements $$F=\left(
\begin{array}{c}
f_{ph} \\
f_{\alpha h}
\end{array}
\right) ,\quad \bar{F}=\left(
\begin{array}{c}
f_{hp} \\
f_{h\alpha }
\end{array}
\right) ~. \label{E3.13}$$ In conventional linear response theory (see, e.g., Ref. [@RS.80]) the polarization function $\Pi _{pqp^{\prime }q^{\prime
}}(\omega )$ is defined by the response of the density matrix to an external field with a harmonic time dependence $$\delta \rho _{pq}=\sum_{p^{\prime }q^{\prime }}\Pi _{pqp^{\prime }q^{\prime
}}(\omega )\,f_{p^{\prime }q^{\prime }}~. \label{E3.14}$$ Its spectral representation reads $$\Pi _{pqp^{\prime }q^{\prime }}(\omega )=\sum_{\mu }\frac{\langle 0|\psi
_{q^{{}}}^{\dagger }\psi _{p^{{}}}^{{}}|\mu \rangle \langle \mu |\psi
_{p^{\prime }}^{\dagger }\psi _{q^{\prime }}^{{}}|0\rangle }{\omega -E_{\mu
}+E_{0}+i\eta }-\frac{\langle 0|\psi _{p^{\prime }}^{\dagger }\psi
_{q^{\prime }}^{{}}|\mu \rangle \langle \mu |\psi _{q^{{}}}^{\dagger }\psi
_{p^{{}}}^{{}}|0\rangle }{\omega +E_{\mu }-E_{0}+i\eta }~, \label{E3.15}$$ where the index $\mu $ runs over all excited states $\mu \rangle $ with energy $E_{\mu }$. In the RPA approximation the polarization function is obtained by inverting the matrix $$\Pi (\omega )=\left[ \left(
\begin{array}{cc}
\omega +i\eta & 0 \\
0 & -\omega -i\eta
\end{array}
\right) -\left(
\begin{array}{cc}
A & B \\
B^{\ast } & A^{\ast }
\end{array}
\right) \right] ^{-1}~. \label{E3.16}$$ $\Pi (\omega )$ is the solution of the linearized Bethe-Salpeter equation $$\Pi (\omega )=\Pi ^{0}(\omega )+\Pi ^{0}(\omega )V\,\Pi (\omega )~,
\label{E3.17}$$ where the free polarization function is given by $$\Pi _{klk^{\prime }l^{\prime }}^{0}(\omega )=\frac{\rho _{l}^{(0)}-\rho
_{k}^{(0)}}{\omega -\epsilon _{k}+\epsilon _{l}+i\eta }\delta _{kk^{\prime
}}\delta _{ll^{\prime }}~. \label{E3.18}$$ The eigenmodes of the system are determined by the RPA equation $$\left(
\begin{array}{cc}
A & B \\
-B^{\ast } & -A^{\ast }
\end{array}
\right) \left(
\begin{array}{c}
X \\
Y
\end{array}
\right) _{\mu }=\left(
\begin{array}{c}
X \\
Y
\end{array}
\right) _{\mu }\Omega _{\mu }~. \label{E3.19}$$ In principle, this is a non-Hermitian eigenvalue problem. In the non relativistic case, however, it can be reduced to a Hermitian problem of half dimension, if the RPA matrices are real and if $(A+B)$ is positive definite. In this case one can also show that the eigenvalues $\Omega _{\mu }^{2}$ are positive, i.e., the RPA eigenfrequencies $\Omega
_{\mu }$ are real (see [@RS.80]).
The relativistic case is much more complicated. From Eq. (\[E3.10\]) we notice that the matrix $(A+B)$ is not positive definite. The $\alpha h$ configurations have large negative diagonal matrix elements $\epsilon _{\alpha h}=\epsilon _{\alpha}-\epsilon _{h}\le -1.2$ GeV, and the RRPA equation can no longer be reduced to a Hermitian problem of half dimension. In this case it is also not clear whether the eigenfrequencies are necessarily real, because the stability matrix $${\cal S}=\left(
\begin{array}{cc}
A & B \\
B^{\ast } & A^{\ast }
\end{array}
\right) \label{E3.20}$$ is no longer positive definite. Rather than minima, the solutions of the RMF equations are saddle points[@Providencia] in the multi-dimensional energy surface, and the Thouless theorem [@Th.61], which states that a positive definite stability matrix ${\cal \ S}$ leads to a stable RPA equation with real frequencies, does not apply.
However, the opposite is not true: if the stability matrix is not positive definite, it does not automatically follow that the eigenvalues of the corresponding RPA matrix are not real. In fact, cases like this occur also in the non relativistic RPA in the neighborhood of phase transitions, where the interaction $V$ is very large and attractive. The positive energies $\varepsilon_p - \varepsilon_h$ on the diagonal of the stability matrix are not large enough, as compared to the matrix elements of $V$, to guarantee positive eigenvalues of ${\cal S}$. In the relativistic case the energies on the diagonal $\varepsilon_\alpha - \varepsilon_ h$ are negative. Even for small matrix elements of V the stability matrix ${\cal S}$ will have negative eigenvalues. However, as long as the diagonal part dominates, i.e. as long as we are not in a neighborhood of a phase transition, the RRPA eigenfrequencies are real. This can be easily demonstrated if instead of the RPA amplitudes $X$ and $Y$, we define the generalized coordinates $Q$ and momenta $P$ $$Q=\frac{1}{\sqrt{2}}(X-Y^{\ast }),\;\;\;\;\;\;\;P=\frac{i}{\sqrt{2}}(X+Y^{\ast })~. \label{E3.21}$$ In the small amplitude limit the time-dependent mean field equations take the form of classical Hamiltonian equations (for details see Ref. [@RS.80], Chapt. 12) for the Hamiltonian function $${\cal H}(P,Q)=\frac{1}{2}\left(
\begin{array}{cc}
P^{\ast } & -P
\end{array}
\right) {\cal M}^{-1}\left(
\begin{array}{c}
P \\
-P^{\ast }
\end{array}
\right) +\frac{1}{2}\left(
\begin{array}{cc}
Q^{\ast } & Q
\end{array}
\right) {\cal S}\left(
\begin{array}{c}
Q \\
Q^{\ast }
\end{array}
\right), \label{E3.22}$$ with the inertia tensor $${\cal M}=\left(
\begin{array}{cc}
A & -B \\
-B^{\ast } & A^{\ast }
\end{array}
\right) ^{-1}~. \label{E3.23}$$ The large negative diagonal matrix elements are also present in the inertia tensor. If the off-diagonal matrix elements are not too large, a negative inertia and a negative curvature will again result in real frequencies. In all applications of RRPA we have found real frequencies, though in none of these cases the stability matrix ${\cal S}$ was positive definite. This also explain why the time-dependent RMF equations have stable solutions which describe oscillations with real frequencies around the static solution, although the static solution itself corresponds to a saddle point.
The solution of the RPA equations in configuration space is much more complicated in the relativistic case. Firstly, because in addition to the usual $ph$-states, the configuration space includes a large number of $\alpha h$-states. A further complication arises because the full non-Hermitian RPA matrix has to be diagonalized, even in cases when the matrix elements are real. The usual method [@RS.80], which reduces the dimension of the RPA equations by half does not apply.
Summarizing the results of this section, we have shown that the relativistic RPA represents the small amplitude limit of the time-dependent RMF theory. However, because the RMF theory is based on the [*no-sea*]{} approximation, the RRPA configuration space includes not only the ususal $ph$-states, but also $\alpha h$-configurations, i.e. pairs formed from occupied states in the Fermi sea and empty negative-energy states in the Dirac sea. At each time $t\neq 0$ the occupied positive energy states can have non-vanishing overlap with both positive and negative energy solutions calculated at $t=0$. If the density matrix $\hat{\rho}(t)$ is represented in the basis which diagonalizes the static solution $\hat{\rho}^{(0)}$, it contains not only the usual components $\delta \hat{\rho}_{ph}$ with a particle above the Fermi level and a hole in the Fermi sea, but also components $\delta \hat{\rho}_{\alpha h}$ with a particle in the Dirac sea and a hole in the Fermi sea.
One of the important advantages of using the time-dependent variational approach is that it conserves symmetries. It is well known from non-relativistic time-dependent mean field theory that symmetries are connected with zero energy solutions of the RPA, i.e. the Goldstone modes, and it is one of the advantages of RPA that it restores the symmetries broken by the mean field. This has already been realized in the early studies of symmetry conservation in RRPA, and it has been emphasized by Dawson and Furnstahl in Ref. [@DF.90], that it is essential to include the $\alpha h$ configuration space in order to bring the Goldstone modes to zero energy.
However, it was not anticipated that negative energy states in the RRPA configuration space could have dramatic effects on the excitation energies of giant resonances, as we will show in the following sections. It is not obvious that basis states with unperturbed energies more than 1.2 GeV below the Fermi energy, can have a big influence on giant resonances with excitation energies in the MeV region. In the following section we will study a simple model which provides a deeper insight into this problem.
A separable model
=================
The model studied in this section represents a relativistic extension of the Brown and Bolsterli model[@BB.59], which has played an essential role in the understanding of the microscopic picture of collective excitations.
The single-particle basis consists of 4 states, each of them $\Omega $-fold degenerate ($\nu =1, \ldots \Omega $). The first two states (1 and 2) correspond to particle levels with the free mass $m$, the states 3 and 4 correspond to the negative energy levels with free mass $-m$. The model Hamiltonian reads $$H=H_{0}-\lambda _{s}^{{}}S^{\dagger }S+\lambda _{v}^{{}}V^{\dagger }V~,
\label{E4.1}$$ where $H_0$ is the Hamiltonian which describes free Dirac particles $$H_{0}=\sum_{i=1}^{A}
\Big(\bf{\alpha}{\bf p}_{i }+ \beta m_{i }+ \frac{1}{2} \sigma
\varepsilon_{i }\Big)~,
\label{E4.2}$$ with $$\alpha =\left(
\begin{array}{cccc}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1 \\
1 & 0 & 0 & 0 \\
0 & 1 & 0 & 0
\end{array}
\right) ,\quad \beta =\left(
\begin{array}{cccc}
1 & 0 & 0 & 0 \\
0 & 1 & 0 & 0 \\
0 & 0 & -1 & 0 \\
0 & 0 & 0 & -1
\end{array}
\right) ,\quad \sigma =\left(
\begin{array}{cccc}
1 & 0 & 0 & 0 \\
0 & -1 & 0 & 0 \\
0 & 0 & 1 & 0 \\
0 & 0 & 0 & -1
\end{array}
\right) \label{E4.3}$$ In Eq.(\[E4.2\]) $m_i = m$ is the free mass of particle $i$, $p_i = p$ denotes its momentum, and $\varepsilon_{i} = \varepsilon_{0} \ll m$ induces a small splitting between the levels 1 and 2 (and, of course, between 3 and 4). The interaction consists of an attractive scalar field $S$ and a repulsive vector field $V$ $$S=\sum_{i=1}^{A}\left(
\begin{array}{cccc}
0 & 1 & 0 & 0 \\
1 & 0 & 0 & 0 \\
0 & 0 & 0 & -1 \\
0 & 0 & -1 & 0
\end{array}
\right) _{i },\quad \quad V=\sum_{i=1}^{A}\left(
\begin{array}{cccc}
0 & 1 & 0 & 0 \\
1 & 0 & 0 & 0 \\
0 & 0 & 0 & 1 \\
0 & 0 & 1 & 0
\end{array}
\right) _{i}~, \label{E4.4}$$ with the strength parameters $\lambda _{s}$ and $\lambda _{v}.$ In the formalism of second quantization the operators $H_{0\text{,}}$ $S$ and $V$ take the forms $$\begin{aligned}
H_{0} &=&p\sum_{\nu }c_{1\nu }^{\dagger }c_{3\nu }^{{}}+c_{2\nu }^{\dagger
}c_{4\nu }^{{}}+h.c. \nonumber \\
&& + m\sum_{\nu }c_{1\nu }^{\dagger }c_{1\nu }^{{}}+c_{2\nu }^{\dagger
}c_{2\nu }^{{}}-c_{3\nu }^{\dagger }c_{3\nu }^{{}}-c_{4\nu }^{\dagger
}c_{4\nu }^{{}} \nonumber \\
&& + \frac{\varepsilon _{0}}{2}\sum_{\nu }c_{1\nu }^{\dagger }c_{1\nu
}^{{}}-c_{2\nu }^{\dagger }c_{2\nu }^{{}}+c_{3\nu }^{\dagger }c_{3\nu
}^{{}}-c_{4\nu }^{\dagger }c_{4\nu }^{{}}~, \label{E4.5} \\
S &=&\sum_{\nu }c_{1\nu }^{\dagger }c_{2\nu }^{{}}-c_{3\nu }^{\dagger
}c_{4\nu }^{{}}+h.c.~, \label{E4.6} \\
V &=&\sum_{\nu }c_{1\nu }^{\dagger }c_{2\nu }^{{}}+c_{3\nu }^{\dagger
}c_{4\nu }^{{}}+h.c. \label{E4.7}\end{aligned}$$
At the mean field level the diagonalization of the Dirac operator $$H_{0}=\left(
\begin{array}{cccc}
m+\frac{\varepsilon _{0}}{2} & 0 & p & 0 \\
0 & m-\frac{\varepsilon _{0}}{2} & 0 & p \\
p & 0 & -m+\frac{\varepsilon _{0}}{2} & 0 \\
0 & p & 0 & -m-\frac{\varepsilon _{0}}{2}
\end{array}
\right) ~, \label{E4.10}$$ result in the eigenvalues $$\epsilon _{p}=E+\frac{\varepsilon _{0}}{2},\quad \quad \epsilon _{h}=E-\frac{\varepsilon _{0}}{2},\quad \quad \epsilon _{\alpha }=-E+\frac{\varepsilon
_{0}}{2},\quad \quad \epsilon _{\alpha ^{\prime }}=-E-\frac{\varepsilon _{0}}{2}, \label{E4.12}$$ and the corresponding eigenvectors are $$\psi _{p}=\left(
\begin{array}{c}
f \\
0 \\
g \\
0
\end{array}
\right) ,\,\,\,\,\psi _{h}=\left(
\begin{array}{c}
0 \\
f \\
0 \\
g
\end{array}
\right) ,\,\,\,\,\,\psi _{\alpha }=\left(
\begin{array}{c}
-g \\
0 \\
f \\
0
\end{array}
\right) ,\,\,\,\,\,\,\psi _{\alpha ^{\prime }}=\left(
\begin{array}{c}
0 \\
-g \\
0 \\
f
\end{array}
\right) ~, \label{E4.13}$$ respectively. We use the notation $$E=\sqrt{p^{2}+m^{2},},\,\quad \quad \quad f=\cos \frac{\phi }{2},\,\quad
\quad \quad g=\sin \frac{\phi }{2}~, \label{E4.14}$$ with $$\tan \frac{\phi }{2}=\frac{p}{m+E}~. \label{E4.16}$$ A realistic choice of single particle energies is $$\begin{aligned}
\epsilon _{ph} &=&\epsilon _{p}-\epsilon _{h}=\varepsilon _{0}\approx 10\text{ MeV}~, \nonumber \\
\epsilon _{\alpha h} &=&\epsilon _{\alpha }-\epsilon _{h}=-2E+\varepsilon
_{0}\simeq -2\text{ GeV}~, \nonumber \\
\epsilon _{\alpha ^{\prime }h} &=&\epsilon _{\alpha ^{\prime }}-\epsilon
_{h}=-2E\simeq -2\text{ GeV}~. \label{E4.18}\end{aligned}$$ In realistic calculations the ratio between large and small components of the Dirac spinors is approximately $f/g\approx 30$, i.e., $\phi \simeq 33^{0}$. In the basis (\[E4.13\]) the matrices of the operators $S$ and $V$ are $$S=\sum_{i}\left(
\begin{array}{cccc}
0 & \cos \phi & 0 & -\sin \phi \\
\cos \phi & 0 & -\sin \phi & 0 \\
0 & -\sin \phi & 0 & -\cos \phi \\
-\sin \phi & 0 & -\cos \phi & 0
\end{array}
\right) _{i}~, \label{E4.19}$$ $$V=\sum_{i}\left(
\begin{array}{cccc}
0 & 1 & 0 & 0 \\
1 & 0 & 0 & 0 \\
0 & 0 & 0 & 1 \\
0 & 0 & 1 & 0
\end{array}
\right) _{i}~. \label{E4.20}$$ We notice the essential matrix elements $$\begin{array}{ll}
S_{ph}=\cos \phi & V_{ph}=1 \\
S_{\alpha h}=-\sin \phi & V_{\alpha h}=0 \\
S_{\alpha ^{\prime }h}=0 & V_{\alpha ^{\prime }h}=0~.
\end{array}
\label{E4.21}$$ In analogy to the Brown-Bolsterli model, the unperturbed polarization function is $$\Pi _{FF^{\prime }}^{0}(\omega )=\frac{2\varepsilon _{0}^{{}}\Omega
F_{ph}^{{}}F_{ph}^{\prime }}{\omega _{{}}^{2}-\varepsilon _{0}^{2}+i\eta }+\frac{2\varepsilon _{\alpha h}^{{}}\Omega F_{\alpha h}^{{}}F_{\alpha
h}^{\prime }}{\omega _{{}}^{2}-\varepsilon _{\alpha h}^{2}+i\eta }~,
\label{E4.22}$$ where the operators $F,$ $F^{\prime }$ $\in$ {$S,\ V$}. In particular, $$\begin{aligned}
\Pi _{SS}^{0}(\omega ) &=&\frac{2\varepsilon _{0}^{{}}\Omega \cos ^{2}\phi }{\omega _{{}}^{2}-\varepsilon _{0}^{2}+i\eta }-\frac{2E\Omega \sin ^{2}\phi }{\omega _{{}}^{2}-E_{{}}^{2}+i\eta }~, \label{E4.23} \\
\Pi _{VV}^{0}(\omega ) &=&\frac{2\varepsilon _{0}^{{}}\Omega }{\omega
_{{}}^{2}-\varepsilon _{0}^{2}+i\eta }~, \label{E4.24} \\
\Pi _{VS}^{0}(\omega ) &=&\Pi _{VS}^{0}(\omega )=\frac{2\varepsilon
_{0}^{{}}\Omega \cos \phi }{\omega _{{}}^{2}-\varepsilon _{0}^{2}+i\eta }~.
\label{E4.25}\end{aligned}$$ The RRPA frequencies are the roots of the determinant $$\det \left( 1-
\begin{array}{cc}
-\Pi _{SS}^{0}(\omega )\lambda _{s} & \Pi _{SV}^{0}(\omega )\lambda _{v} \\
-\Pi _{SV}^{0}(\omega )\lambda _{s} & \Pi _{VV}^{0}(\omega )\lambda _{v}
\end{array}
\right) =0 \label{E4.26}$$ The essential difference with respect to the non-relativistic Brown-Bolsterli model is the additional term $2E\Omega \sin ^{2}\phi /(\omega ^{2}-E^{2}+i\eta )$ in the scalar polarization $\Pi _{SS}^{0}(\omega )$. Without this term (i.e. $\phi =0)$, the eigenfrequencies would be determined by $$\omega _{{}}^{2}=\varepsilon _{0}^{2}-2\varepsilon _{0}^{{}}\Omega (\lambda
_{s}-\lambda _{v})~, \label{E4.28}$$ with the usual cancellation of scalar and vector interactions. With the additional term, states with unperturbed energies at $\simeq -2E$ are included in the RPA configuration space. The interaction between these states and the $ph$-states is determined by the matrix elements of the scalar interaction $$v_{\alpha h^{\prime }hp}=-\lambda _{s}\cos \phi \sin \phi ~. \label{E4.29}$$ These matrix elements are not reduced by a similar term coming from the vector interaction. In our simplified model this vector-induced term vanishes as a consequence of the relativistic structure of the equations: while for a state from the Fermi sea the large component is the upper component of the spinor, a state from the Dirac sea has a large lower component of the spinor. Due to the $\gamma$-matrix structure of the vertex, the matrix elements of the vector interaction vanish. In realistic calculations these matrix elements do not vanish identically, but they are reduced by an order of magnitude as compared to the corresponding scalar terms.
At excitation energies in the MeV region, $\omega \ll E$ in the denominator of the second term of Eq.(\[E4.23\]), and we obtain an energy-independent term $(2\Omega \lambda_{s}\sin ^{2}\phi )/E$ in the dispersion relation. The eigenfrequencies are now determined by $$\omega _{{}}^{2}=\varepsilon _{0}^{2}-\varepsilon _{0}^{{}}2\Omega (\lambda
_{s}-\lambda _{v})+\varepsilon _{0}^{{}}2\Omega \lambda _{s}\sin ^{2}\phi
\frac{E+2\Omega \lambda _{v}}{E+2\Omega \lambda _{s}\sin ^{2}\phi }~,
\label{E4.30}$$ i.e. we find an additional repulsion for the collective state.
An illustrative case: the giant monopole resonance
==================================================
The RPA equations can be solved either by diagonalizing the RPA matrix in configuration space (see Eq.(\[E3.19\])), or the response function can be calculated by solving the Bethe-Salpeter equation (\[E3.17\]) in momentum space[@HG.89; @MGT.97]. In both cases, of course, one first has to determine the single-nucleon spinors and the mean-fields which correspond to the stationary solution for the ground-state. The Dirac-Hartree equations and the equations for the meson fields are solved self-consistently in the mean-field approximation. The eigenvalue problem is solved, for instance, by diagonalization in a spherically symmetric harmonic oscillator basis [@GRT.90]. From the spectrum of single-nucleon states the RPA configuration space is built: particle-hole ($ph$) and antiparticle-hole ($\alpha h$) pairs which obey the selection rules for angular momentum, parity and isospin. The number of basis states is also determined by two cut-off parameters: the maximal $ph$-energy ($\epsilon _{m}-\epsilon
_{i}<E_{\max }$) and the minimal $\alpha h$-energy ($\epsilon _{\alpha}-\epsilon
_{i}>E_{\min }$). With this basis the RPA matrix is calculated for the same effective interaction that determines the ground-state, or the free polarization function $\Pi ^{(0)}$ is calculated in the response function method. Both methods require that the single-particle continuum is discretized. In order to smooth out the RPA strength function, the discrete strength distribution is folded by a Lorentzian of width $\Gamma $. In the response function method the folding is automatic if a finite value parameter $i\Gamma $ is used in the denominators of Eqs.(\[E3.15\]-\[E3.18\]), instead of the infinitesimal parameter $i\eta $. We have verified that identical results are obtained with both methods.
The large effect of Dirac sea states on isoscalar strength distributions is illustrated in Fig. 1, where we display the isoscalar monopole RRPA strength in $^{116}$Sn calculated with the NL3 effective interaction [@LKR.97] and the width of the Lorentzian is $\Gamma $ = 2 MeV. Recent experimental data are available for the isoscalar giant monopole resonance in $^{116}$Sn [@YCL.99]. The solid curve represents the full RRPA strength and it displays a pronounced peak at 16 MeV, in excellent agreement with the measured value of 15.9 MeV[@YCL.99]. Giant monopole resonances in spherical nuclei are in best agreement with experimental data, when calculated with effective Lagrangians with a nuclear matter compression modulus in the range 250-270 MeV [@VLB.97; @VWR.00; @MGW.00]. The nuclear matter incompressibility of the NL3 effective interaction is 272 MeV.
The long-dashed curve in Fig. 1 corresponds to the to the case with no ${\alpha}h$ pairs in the RRPA configuration space. We notice that, without the contribution from Dirac sea states, the strength distribution is shifted to lower energy. The position of the peak is shifted from $\approx 16$ MeV to below 10 MeV if ${\alpha}h$ pairs are not included in the RRPA basis. Quantitatively similar results are also obtained with other effective interactions. In Fig. 1 we have also separated the contributions of vector and scalar mesons to the ${\alpha}h$ matrix elements. The dash-dot-dot (dash-dot) curve corresponds to calculations in which only vector mesons (scalar mesons) were included in the coupling between the Fermi sea and Dirac sea states. Both interactions were included in the positive energy particle-hole matrix elements. The resulting strength distributions nicely illustrate the dominant contribution of the isoscalar scalar sigma meson to the ${\alpha}h$ matrix elements, in complete agreement with the result obtained in the previous section for the schematic Brown-Bolsterli model.
It is also interesting to examine the effect of the coupling via the spatial components of the vector meson fields, i.e. the term $-{{\mbox{\boldmath $\alpha$}}}^{(1)}{{\mbox{\boldmath $\alpha$}}}^{(2)}$ in the interaction of Eq. (\[E2.16\]). In time-dependent calculations this coupling results from the nucleon currents. In Fig.2 we display the isoscalar monopole RRPA strength in $^{116}$Sn calculated as follows: a) full RRPA (solid curve); b) without the matrix elements of the spatial components of the vector meson fields (dot-dashed curve); c) without the contribution of the Dirac sea to the matrix elements of the spatial components of the vector meson fields (dashed curve); and d) the free Hartree response function. The currents do not contribute to the static polarizability and to the $M_{-1}$ moment. At finite frequencies, however, their contribution is attractive and it lowers the ISGMR energy by $\approx 2$ MeV.. Therefore, if the contribution of the spatial components of the vector fields is neglected, a better agreement with experimental values would be obtained with a lower nuclear matter incompressibility: $K_{\infty }\simeq 230$ MeV. Incidentally, this lower value for the nuclear matter incompressibility is the one advocated by non-relativistic RPA calculations [@BBD.95; @CGB.00]. It has already been noted in time-dependent RMF calculations [@VLB.97], as well as in recent relativistic RPA studies [@MGW.00], that effective interactions which reproduce the IS GMR excitation energies in finite nuclei have a somewhat higher nuclear matter incompressibility than the corresponding non-relativistic Skyrme or Gogny interactions. Here we point to a possible solution to this puzzle: the current terms in the matrix elements of the particle-hole interaction (\[E3.10\],\[E3.11\]) are given by $$\langle p||j_{1}(kr)[{{\mbox{\boldmath $\alpha$}}}Y_{1}({\bf \hat{r}})]_{J=0}||h\rangle
=\langle p||j_{1}(kr)\left(
\begin{array}{cc}
0 & [{{\mbox{\boldmath $\sigma$}}}Y_{1}]_{J=0} \\
\lbrack {{\mbox{\boldmath $\sigma$}}}Y_{1}]_{J=0} &
\end{array}
\right) ||h\rangle~, \label{E5.1}$$ which is a typical relativistic term because it couples large and small components of a Dirac spinor. Since they change parity, terms of the type $[{\bf \sigma }Y_{1}]_{J=0}$ cannot contribute to the giant monopole resonance in a non-relativistic calculation.
Conclusions
===========
In the last couple of years, several discrepancies have been reported between the results obtained with the Relativistic Random Phase approximation and the Time-Dependent Relativistic Mean Field theory, when applied to the the description of small amplitude collective motion in atomic nuclei.
In order to resolve this puzzle, in the present work we have derived the RRPA from the TDRMF equations in the limit of small amplitude motion. The relativistic single particle density matrix $\hat{\rho}(t)$ has been expanded in terms of the stationary solutions of the ground-state. We have shown that the [*no-sea*]{} approximation, which is essential for practical application of the RMF theory in finite nuclei, leads to a fundamental difference between the relativistic and non-relativistic approaches. While in the non-relativistic case the time-dependent variation of the density $\delta \hat{\rho}(t)=\hat{\rho}(t)-\hat{\rho}^{(0)}$ has only $ph$-matrix elements (particle ($p$) above the Fermi surface, hole ($h$) in the Fermi sea), in the relativistic case $\delta \hat{\rho}$ contains also $\alpha h$-matrix elements, where $\alpha$ denotes unoccupied states in the Dirac sea. The fact that states in the Dirac sea can be occupied is a direct consequence of the [*no-sea*]{} approximation. In constructing the matrix $\delta \hat{\rho}$ one has to take into account that a complete basis of single particle states contains both positive and negative energy solutions of the Dirac equation. Already in Ref. [@DF.90] it has been shown that an RRPA calculation, consistent with the mean-field model in the $no-sea$ approximation, necessitates configuration spaces that include both particle-hole pairs and pairs formed from occupied states and negative-energy states. The contributions from configurations built from occupied positive-energy states and negative-energy states are essential for current conservation and the decoupling of the spurious state.
What is less obvious, however, is that the inclusion of negative-energy single particle states in the RRPA configuration space has such a dramatic effect on the calculated excitation energies of isoscalar giant resonances. In a schematic model we have shown that, due to the relativistic structure of the RPA equations, the matrix elements of the time-like component of the vector meson fields which couple the $\alpha h$-configurations with the $ph$-configurations vanish. In realistic calculations these matrix elements do not vanish identically, but they are strongly reduced with respect to the corresponding matrix elements of the isoscalar scalar meson field. As a result, the well known cancellation between the contributions of the $\sigma $ and $\omega $ fields, which, for instance, leads to ground-state solution, does not take place and we find large matrix elements coupling the $\alpha h$-sector with the $ph$-configurations. In addition, the number of $\alpha h$-configurations which can couple to the $ph$-configurations in the neighborhood of the Fermi surface is much larger than the number of $ph$-configurations. This can increase the effect by enhancing the collectivity of the $\alpha h$-configuration space.
The large effect of Dirac sea states on isoscalar strength distributions has been illustrated for the giant monopole resonance in $^{116}$Sn. We have also shown that currents cannot be neglected in the calculation of giant resonances. Of course they do not occur in the static case, i.e. in the calculations of the static polarizability or the $M_{-1}$ moment. At finite frequencies, however, time reversal invariance is broken and spatial components of the vector meson fields play an important role. This effect is known as [*nuclear magnetism*]{}. It is a genuine relativistic effect, because the matrix elements couple the large and small components of a Dirac spinor. Since the spatial components of the vector fields have the form $-{{\mbox{\boldmath $\alpha$}}}^{(1)}{{\mbox{\boldmath $\alpha$}}}^{(2)}$ $D_{\omega }({\bf r}_{1},{\bf r}_{2})$, they result in an attractive contribution which lowers the value of the calculated excitation energies of giant resonances. This explains the difference between the non-relativistic and relativistic RPA results for the isoscalar giant monopole resonances in spherical nuclei, and the corresponding predictions for the nuclear matter compression modulus.
[**ACKNOWLEDGMENTS**]{}
P.R. acknowledges the support and the hospitality extended to him during his stay at the IPN-Orsay, where a large part of this work was completed. This work has been supported in part by the Bundesministerium für Bildung und Forschung under the project 06 TM 979 and by the Deutsche Forschungsgemeinschaft. It was also partially supported by the National Natural Science Foundation of China under grant No. 19847002, 19835010-10075080 and Major State Basic Research Development Program under contract No. G200077407.
B.D. Serot and J.D. Walecka, Adv. Nucl. Phys. [**16**]{}, 1 (1986).
P.-G. Reinhard, Rep. Progr. Phys. [**52**]{}, 439 (1989).
B.D. Serot, Rep. Prog. Phys. [**55**]{}, 1855 (1992).
P. Ring, Progr. Part. Nucl. Phys. [**37**]{}, 193 (I996).
C.J. Horowitz and B.D. Serot, Nucl. Phys. [**A399**]{}, 529 (1983).
A. Bouyssy, J.F. Mathiot, N. Van Giai, and S. Marcos, Phys. Rev. [**C36**]{}, 380 (1987).
P. Bernados, V.N. Fomenko, N. Van Giai, M.L. Quelle, S. Marcos, R. Niembro, and L.N. Savushkin, Phys. Rev. [**C48**]{}, 2665 (1993).
C.J. Horowitz and B.D. Serot, Phys. Lett. [**140B**]{}, 181 (1984).
D.A. Wasson, Nucl. Phys. [**A535**]{}, 456 (1991).
J. Boguta and A. R. Bodmer, Nucl. Phys. [**A292**]{}, 413 (1977).
W. Koepf and P. Ring, Nucl. Phys. [**A493**]{}, 61 (1989).
A.V. Afanasjev, J. König, and P. Ring, Nucl. Phys. [**A608**]{}, 107 (1996).
A.V. Afanasjev, G.A. Lalazissis, and P. Ring, Nucl. Phys. [**A634**]{}, 395 (1998).
A.V. Afanasjev, J. König, and P. Ring, Phys. Rev. [**C60**]{}, 051303 (1999).
D. Vretenar, H. Berghammer, and P. Ring, Nucl. Phys. [**A581**]{}, 679 (1995).
B. Podobnik, D. Vretenar, and P. Ring, Z. Phys. [**A354**]{}, 375 (1996).
D. Vretenar, G.A. Lalazissis, R. Behnsch, W. Pöschl, and P. Ring, Nucl. Phys. [**A621**]{}, 853 (1997).
R.J. Furnstahl, Phys. Lett. [**152B**]{}, 313 (1985).
M. L’Huillier and N. Van Giai, Phys. Rev. [**C39**]{}, 2022 (1989).
P.G. Blunden and M. McCorquodale, Phys. Rev. [**C38**]{}, 1861 (1988).
J.F. Dawson and R.J. Furnstahl, Phys. Rev. [**C42**]{}, 2009 (1990).
C.J. Horowitz and J. Piekarewicz, Nucl. Phys. [**A511**]{}, 461 (1990).
Z.Y. Ma, N. Van Giai, H. Toki, and M. L’Huillier, Phys. Rev. [**C42**]{}, 2385 (1997).
Z.Y. Ma, H. Toki, and N. Van Giai, Nucl. Phys. [**A627**]{}, 1 (1997).
N. Van Giai, Z.Y. Ma, H. Toki, and B.Q. Chen, Nucl. Phys. [**A649**]{}, 37c (1999).
D. Vretenar, P. Ring, G.A. Lalazissis, and N. Paar, Nucl. Phys. [**A649**]{}, 29c (1999).
P. Ring and P. Schuck, [*The Nuclear Many-Body Problem*]{}, Springer Verlag, New York 1980.
D. Vretenar, A. Wandelt, and P. Ring, Phys. Lett. [**B487**]{}, 334 (2000).
Z.Y. Ma, N. Van Giai, A. Wandelt, D. Vretenar, and P. Ring, Nucl. Phys. [**A 685**]{} (2001) (in press).
P. Ring and J. Speth, Nucl. Phys. [**A235**]{}, 315 (1974).
G.E. Brown and M. Bolsterli, Phys. Rev. Lett. [**3**]{}, 472 (1959).
T. Kohmura, Y. Miyama, T. Nagai, S. Ohnaka, J. da Providencia, and T. Kodama, Phys. Lett.[**B 226**]{}, 207 (1989).
D.J. Thouless, Nucl. Phys. [**22**]{}, 78 (1961).
Y.K. Gambhir, P. Ring, and A. Thimet, Ann. Phys. (N.Y.) [**198**]{}, 132 (1990).
D.H. Youngblood, H.L. Clark, and Y.W. Lui, Phys. Rev. Lett. [**82**]{}, 691 (1999).
G.A. Lalazissis, J. König, and P. Ring, Phys. Rev. [**C 55**]{}, 540 (1997).
J.P. Blaizot, J.F. Berger, J. Dechargé, and M. Girod, Nucl. Phys. [**A 591**]{}, 435 (1995).
G. Colò, N. Van Giai, P.F. Bortignon, and M.R. Quaglia, Phys. Lett. [**B 485**]{}, 362 (2000).
![ISGMR strength distributions in $^{116}$Sn calculated with the NL3 effective interaction. The solid and long-dashed curves are the RRPA strengths with and without the inclusion of Dirac sea states, respectively. The dash-dot-dot (dash-dot) curve corresponds to calculations in which only vector mesons (scalar mesons) are included in the coupling between the Fermi sea and Dirac sea states.[]{data-label="Fig.1"}](ringetal_fig1.eps)
![Effects of nuclear magnetism on the IS GMR strength distribution in $^{116}$Sn. Solid curve: full RRPA calculation; dash-dotted curve: without the matrix elements of the spatial components of the vector meson fields; dashed curve: without the contribution of the Dirac sea to the matrix elements of the spatial components of the vector meson fields. The free Hartree response is also displayed.[]{data-label="Fig.2"}](ringetal_fig2.eps)
[^1]: also Institute of Theoretical Physics, Beijing, P.R. of China
[^2]: On leave from University of Zagreb, Croatia
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Despite the growing popularity of modern machine learning techniques (e.g. Deep Neural Networks) in cyber-security applications, most of these models are perceived as a black-box for the user. Adversarial machine learning offers an approach to increase our understanding of these models. In this paper we present an approach to generate explanations for incorrect classifications made by data-driven Intrusion Detection Systems (IDSs) An adversarial approach is used to find the minimum modifications (of the input features) required to correctly classify a given set of misclassified samples. The magnitude of such modifications is used to visualize the most relevant features that explain the reason for the misclassification. The presented methodology generated satisfactory explanations that describe the reasoning behind the mis-classifications, with descriptions that match expert knowledge. The advantages of the presented methodology are: 1) applicable to any classifier with defined gradients. 2) does not require any modification of the classifier model. 3) can be extended to perform further diagnosis (e.g. vulnerability assessment) and gain further understanding of the system. Experimental evaluation was conducted on the NSL-KDD99 benchmark dataset using Linear and Multilayer perceptron classifiers. The results are shown using intuitive visualizations in order to improve the interpretability of the results.'
author:
-
bibliography:
- 'my\_bibliography.bib'
title: An Adversarial Approach for Explainable AI in Intrusion Detection Systems
---
Adversarial Machine Learning, Adversarial samples, Explainable AI, cyber-security,
(current page.north west) ++(1.5,-0.3) node\[below right,text width=19cm\] [ ©2018 IEEE. Personal use of this material is permitted. Permission from IEEE must be obtained for all other uses, in any current or future media, including reprinting/republishing this material for advertising or promotional purposes, creating new collective works, for resale or redistribution to servers or lists, or reuse of any copyrighted component of this work in other works.]{};
(current page.south west) ++(1.7,2.3) node\[below right,text width=9.1cm\]
------------------------------------------------------------------------
\
Accepted version of the paper appearing in the proceedings of the 44th Annual Conference of the IEEE Industrial Electronics Society, IECON 2018.
;
Introduction {#section:introduction}
============
Adversarial Machine Learning {#section:adversarial_machine_learning}
============================
Adversarial machine learning has been extensively used in cyber-security research to find vulnerabilities in data-driven models [@Lowd2005] [@Barreno2010] [@Wagner2002] [@DBLP:journals/corr/LiMSRJ17] [@barreno2006can] [@frederickson2018attack]. Recently, there has been an increasing interest on adversarial samples given the susceptibility of Deep Learning models to these type of attacks [@goodfellow6572explaining] [@kurakin2016adversarial].
Adversarial samples are samples crafted in order to change the output of a model by making small modifications into a reference (usually real) sample [@goodfellow6572explaining]. These samples are used to detect blind spots on ML algorithms.
Adversarial samples are crafted from an attacker perspective to evade detection, confuse the classifier [@goodfellow6572explaining], degrade performance [@papernot2017practical] and/or gain information about the model or the dataset used to train the model [@frederickson2018attack]. Adversarial samples are also useful from a defender point of view given that they can be used to perform vulnerability assessment [@barreno2006can], study the robustness against noise, improve generalization and debug the machine learning model [@papernot2017practical].
In general, the problem of crafting an adversarial sample is stated as follows [@frederickson2018attack]: $$\begin{aligned}
\max_{{\hat{{{ {\boldsymbol{x}} }}}}}\quad & {\mathcal{L}}({{\hat{{{ {\boldsymbol{x}} }}}}}) - \Omega({{\hat{{{ {\boldsymbol{x}} }}}}}) \label{eq:general_attack}\\
s.t. \quad & {{\hat{{{ {\boldsymbol{x}} }}}}}\in \phi({{ {\boldsymbol{x}} }}) \nonumber\end{aligned}$$ where:
- ${\mathcal{L}}({{ {\boldsymbol{x}} }})$ measures the impact of the adversarial sample in the model,
- $\Omega({{ {\boldsymbol{x}} }})$ measures the capability of the defender to detect the adversarial sample.
- ${{\hat{{{ {\boldsymbol{x}} }}}}}\in \phi({{ {\boldsymbol{x}} }})$ ensures the crafted sample ${{\hat{{{ {\boldsymbol{x}} }}}}}$ is inside the domain of valid inputs. It also represents the capabilities of the adversary.
The objective of Eq. (\[eq:general\_attack\]) can be interpreted as crafting an attack point ${{\hat{{{ {\boldsymbol{x}} }}}}}$ that maximizes the impact of the attack, while minimizing the chances of the attack to be detected. The definition of ${\mathcal{L}}({{ {\boldsymbol{x}} }})$ will depend on the intent of the attack and the available information about the model under attack. For example: 1) ${\mathcal{L}}({{ {\boldsymbol{x}} }})$ can be a function that measures the difference between a target class ${{\hat{{{ {\boldsymbol{y}} }}}}}$ and the output from the model; 2) $\Omega({{ {\boldsymbol{x}} }})$ measures the discrepancy between the reference ${{{{ {\boldsymbol{x}} }}_0}}$ and the modified sample ${{\hat{{{ {\boldsymbol{x}} }}}}}$
In this paper, instead of deceiving the classifier, we use Equation \[eq:general\_attack\] to find the minimum number of modifications needed to correctly classify a misclassified sample. The methodology is described in detail in section \[section:methodology\].
Explaining misclassifications using adversarial machine learning {#section:methodology}
================================================================
Experiments and Results {#section:results}
=======================
Conclusion {#section:conclusion}
==========
In this paper, we presented an approach for generating explanations for the incorrect classification of a set of samples. The methodology was tested using an Intrusion Detection benchmark dataset.
The methodology uses an adversarial approach to find the minimum modifications needed in order to correctly classify the misclassified samples. The modifications are used to find and visualize the relevant features responsible for the misclassification. Experiments were performed using Linear and Multilayer perceptron classifiers. The explanations were presented using intuitive plots that can be easily interpreted by the user.
The proposed methodology provided insightful and satisfactory explanations for the misclassification of samples, with results that match expert knowledge. The relevant features found by the presented approach showed that misclassification often occurred on samples with conflicting characteristics between classes. For example, normal connections with low duration and low login success are misclassified as attacks, while attack connections with low error rate and higher login success are misclassified as normal.
Discussion {#section:discussion}
==========
An advantage of the presented approach is that it can be used for any differentiable model and any classification task. No modifications of the model are required. The presented approach only requires the gradients of the model cross-entropy w.r.t. the inputs. Non-continuous functions can also be incorporated into the approach, for example rounding operations for integer and categorical features.
The presented adversarial approach can be extended to perform other analysis of the model. For example, instead of finding modifications for correct the predicted class, the modifications can be used to deceive the classifier, which can be used for vulnerability assessment. Future research will be conducted to incorporate the explanations to improve the accuracy of the model without having to include new data into the training procedure.
0.2in
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We consider the problem of modeling, estimating, and controlling the latent state of a spatiotemporally evolving continuous function using very few sensor measurements and actuator locations. Our solution to the problem consists of two parts: a predictive model of functional evolution, and feedback based estimator and controllers that can robustly recover the state of the model and drive it to a desired function. We show that layering a dynamical systems prior over temporal evolution of weights of a kernel model is a valid approach to spatiotemporal modeling that leads to systems theoretic, control-usable, predictive models. We provide sufficient conditions on the number of sensors and actuators required to guarantee observability and controllability. The approach is validated on a large real dataset, and in simulation for the control of spatiotemporally evolving function.'
author:
- 'Hassan A. Kingravi, Harshal Maske and Girish Chowdhary [^1]'
title: ' Kernel Controllers: A Systems-Theoretic Approach for Data-Driven Modeling and Control of Spatiotemporally Evolving Processes '
---
=1
Introduction {#sec:intro}
============
Modeling, control, and estimation of spatiotemporally varying systems is a challenging area in controls research. These systems are characterized by dynamic evolution in both the spatial and temporal variables. Some examples of relevant problems include active wing-shaping based control of flexible aircraft, control of heat or particulate diffusion in manufacturing processes, control of rumor spreading across a social network, and tactical asset allocation and control problems in dynamically varying battlespaces. The traditional approach to modeling and control of spatiotemporal systems have relied on Partial Differential Equations (PDEs) [@Alabau2012controlpde], solutions to which are functions that evolve in both space and time. However, PDE models can be limited in situations where exact physics based models of the functional evolution are difficult to formulate, or are inherently limited due to the physical understanding of the process or unknown spatiotemporal interactions [@cressie2011statistics]. Furthermore, the control of PDEs is fundamentally more challenging than the control of finite-dimensional state-space systems because the evolution and control spaces are infinite dimensional Hilbert spaces, as opposed to $\mathbb{R}^n$ [@Alabau2012controlpde].
Accordingly, there has been significant work in approximate modeling of spatiotemporally evolving functions using data-driven or distributed parameter based approximations of PDEs [@cressie2011statistics; @wikle2001kernel]. One way to model spatiotemporally evolving functions is to approximate the function at several sampling locations and build an autoregressive model of the evolution of the function’s output over that grid [@baker2000finite]. The fidelity of these models heavily depends on the number of sampling (equivalently Euclidean grid locations in the independent variable space) locations employed, with a large number of grid locations leading to large-scale state-space models that are difficult to manage. An alternative approach to modeling spatiotemporal functional evolution relies on modeling the correlation between any two sampling locations through a smooth covariance kernel [@cressie2011statistics]. The model of the evolution is then formed through a linear, weighted combination of the kernels, and the hyperparameters of the spatiotemporal covariance kernel and the weights are learned by solving an optimization problem. The power and flexibility of this approach lies in the fact that kernels can be defined over abstract objects, and not just Euclidean grid locations, leading to a modeling technique that is domain agnostic. For example, kernel embeddings are available for graphical models studied in decentralized control [@johansson2014global], images [@ren2012coupled], and many other domains. However, formulating control-usable kernel-based models of spatiotemporal phenomena can be challenging due to the need to take into account the spatiotemporal dependence. Many recent techniques in spatiotemporal modeling have focused on covariance kernel design and associated hyperparameter learning algorithms [@garg2012AAAI; @ma2003nonstationary; @RasmussenWilliams2005; @plagemann2008nonstationary]. The main benefit of careful design of covariance kernels over approaches that simply include time in as an additional input variable [@perez:13:gaussian; @Chowdhary13_ACC2] is that they can account for intricate spatiotemopral couplings. However, there are two key challenges with these approaches: the first challenge is in ensuring the scalability of the model to large scale phenomena. This is difficult due to the fact that the hyperparameter optimization problem is not convex in general, and because when time is used as a kernel input, it is nontrivial to restrict the number of kernels used without losing modeling fidelity [@garg2012AAAI; @ma2003nonstationary; @plagemann2008nonstationary]. The second very important challenge is concerned with the formulation of feasible control strategies utilizing predictive kernel-based models of spatiotemporal phenomena. In particular, when the spatiotemporal evolution is embedded in the design of complex covariance kernel, the resulting model of functional evolution can be highly nonlinear and difficult to utilize in control design.
In this paper, we pursue an alternative systems-theoretic approach to the modeling, control, and estimation of spatiotemporally varying functions that fuses the strengths of kernel methods with systems theory. Our main contribution is to provide a systems-theoretic formulation for approximating, with very high accuracy, spatiotemporal functional evolution by layering a linear dynamical systems prior over temporal evolution of weights of a kernel model. For a class of linearly evolving PDEs, such as the heat diffusion and the wave equation, our approach can lead to a very high-accuracy approximation. This modeling approach is also applicable to data-driven modeling of real-world phenomena, which we demonstrate on a challenging inference problem on satellite data of sea surface temperatures. One benefit of our model is that it can encode spatiotemporal evolution of complex nonlinear surfaces through an Ordinary Differential Equation (ODE) evolving in a Hilbert space induced by the specific kernel choice. Yet, the main benefit of our systems-theoretic approach is that it is highly conducive to control synthesis. To illustrate this fact, we demonstrate that feasible control strategies for a class of spatiotemporally evolving systems can be found using linear control synthesis. In particular, we derive sufficient conditions on the kernel selection to guarantee observability and controllability of the presented model. Furthermore, we demonstrate control synthesis for a diffusion PDE using simple Gaussian kernels distributed uniformly in the input domain.
The outline of this paper is as follows, Section \[sec:observers\] focuses on the development of a systems-theoretic kernel-based model of spatiotemporal evolution, Section \[sec:theory\_results\] presents the main theoretical results, Section \[sec:experimental\_results\] presents modeling results on a real-world large dataset and control synthesis results for a diffusion PDE.
Kernel Controllers {#sec:observers}
==================
This section outlines our modeling framework and presents theoretical results associated with the number of sampling locations required for monitoring functional evolution.
Problem Formulation {#sec:formulation}
-------------------
We focus on predictive inference and control over a time-varying stochastic process, whose mean $f$ is temporally evolving: $$\begin{aligned}
{\label{eq:e:bnp_model1}}
& f_{k+1}& \sim \mathbb{F}(f_k,\eta_k) $$ where $\mathbb{F}$ is a distribution varying with time $t$ and exogenous inputs $\eta$. The theory of reproducing kernel Hilbert spaces (RKHSs) provides powerful tools for generating flexible classes of functions with relative ease, and is thus a natural choice for modeling complex spatial functions [@scholkopf:bk:2002]. Therefore, our focus will be on spatiotemporally evolving kernel-based models, such as Gaussian Processes (GPs). In a kernel-based model, ${k}:{\Omega}\times{\Omega}\to{\mathbb{R}}$ is a positive definite kernel on some compact domain ${\Omega}$ that models the covariance between any two points in the input space. A Mercer kernel [@scholkopf:bk:2002] implies the existence of a smooth map ${\psi}:{\Omega}\to{\mathcal{H}}$, where ${\mathcal{H}}$ is an RKHS with the property $$\begin{aligned}
{k}(x,y) &= {\langle {\psi}(x),{\psi}(y) \rangle_{{\mathcal{H}}}} = {\langle {\psi}({k}(x,\cdot)),{\psi}({k}(y,\cdot)) \rangle_{{\mathcal{H}}}}. \end{aligned}$$ There is a large body of literature on modeling spatiotemporal evolution in ${\mathcal{H}}$ [@wikle2002kernel; @cressie2011statistics]. A simple approach for spatiotemporal modeling is to utilize both spatial and temporal variables as inputs to the kernel [@perez:13:gaussian; @Chowdhary13_ACC2]. However, this technique leads to an ever-growing kernel dictionary, which is computationally taxing. Furthermore, constraining the dictionary size or utilizing a moving window will occlude the learning of long-term patterns. Periodic or nonstationary covariance functions and nonlinear transformations have been proposed to address this issue [@ma2003nonstationary; @RasmussenWilliams2005]. Furthermore, work in the design of nonseparable and nonstationary covariance kernels seeks to design kernels optimized to environment-specific dynamics, and optimize their hyperparameters in local regions of the input space [@garg2012AAAI; @das2014nonstationary; @plagemann2008nonstationary]. The model of spatiotemporal functional evolution proposed in this paper builds on the idea that modeling the temporal evolution of mixing weights of a kernel model is a valid approach to spatiotemporal modeling. The key idea behind our approach is that the spatiotemporal evolution of a kernel-based model can be directly modeled by tracing the evolution of the mean embedded in a RKHS using switched ordinary differential equations (ODE) when the evolution is continuous, or switched difference equations when it is discrete (Figure \[fig:hilbert\_evolution\]). The advantage of this approach is that it allows us to utilize powerful ideas from systems theory for knowing necessary conditions for functional convergence; furthermore, it offers a natural framework for designing control mechanisms as well.
![Two types of Hilbert space evolutions. Left: the model, represented by the functions $m_i$, switches discretely in the Hilbert space ${\mathcal{H}}$; Right: the evolution of the function $m_t$ is smooth, represented by a solution to an ordinary differential equation in ${\mathcal{H}}$.[]{data-label="fig:hilbert_evolution"}](model.png){width="0.8\columnwidth"}
In this paper, we restrict our attention to the class of functional evolutions $\mathbb{F}$ defined by linear Markovian transitions in an RKHS. While extension to the nonlinear case is possible (and non-trivial), it is not pursued in this paper to help ease the exposition of key ideas. Let $y\in{\mathbb{R}}^{{N}}$ be the measurements of the function available from ${N}$ sensors, ${\mathcal{A}}:{\mathcal{H}}\to{\mathcal{H}}$ be a linear transition operator in the RKHS ${\mathcal{H}}$, and ${\mathcal{K}}:{\mathcal{H}}\to{\mathbb{R}}^{{N}}$ be a linear measurement operator, the model for the infinite-dimensional functional evolution and measurement studied in this paper is: $$\begin{aligned}
{\label{eq:ideal_lin_evol}}
f_{k+1} &= {\mathcal{A}}f_k + \eta_k\\
y_k &= {\mathcal{K}}f_k + \zeta_k,\end{aligned}$$ where $\eta_k$ is a zero-mean stochastic process in ${\mathcal{H}}$, and $\zeta_k$ is a Wiener process in ${\mathbb{R}}^{{N}}$. For many kernels, the feature map ${\psi}$ is unknown, and therefore it is necessary to work in the dual space of ${\mathcal{H}}$. For concreteness, we work with an approximate space as follows: given points ${\mathcal{C}}= {\{c_1,\dots,c_{{M}}\}}$, $c_i\in{\Omega}$, we have a dictionary of atoms ${\mathcal{F}_{{C}}}= \begin{bmatrix}{\psi}(c_1) &\cdots & {\psi}(c_{{M}})
\end{bmatrix}$, ${\psi}(c_i)\in{\mathcal{H}}$, the span of which is a strict subspace of the RKHS generated by the kernel. Formally, we have $$\begin{aligned}
{\mathcal{C}}\mapsto {\ensuremath{{\mathcal{H}}_{{C}}}}:= \operatorname{span}\begin{bmatrix}{\psi}(c_1) &\cdots & {\psi}(c_{{M}})
\end{bmatrix}
\subset {\mathcal{H}}.\end{aligned}$$ This regime, which trades off the flexibility of a truly nonparametric approach for computational realizability, still allows for the representation of rich phenomena. Let ${N}$ represent the number of sampling locations, and ${M}$ be the number of bases generating ${\ensuremath{{\mathcal{H}}_{{C}}}}$. Note that every function $f\in{\ensuremath{{\mathcal{H}}_{{C}}}}$ has an expansion of the form $$\begin{aligned}
{\label{eq:expansion}}
f(x) = \sum_{i=1}^{{M}} w_i k(c_i,x).\end{aligned}$$ This expansion allows us to write the $w_i$ coordinates in the dual space as vectors $w\in{\mathbb{R}}^{{M}}$. We can show the relation of the function spaces to their Euclidean counterparts via commutative diagrams. Define ${\ensuremath{\mathcal{W}}}:{\ensuremath{{\mathcal{H}}_{{C}}}}\to{\mathbb{R}}^{{M}}$ as the operator that maps the coordinates $w_i$ in [(\[eq:expansion\])]{} to vectors ${w}\in{\mathbb{R}}^{{M}}$, and let ${\ensuremath{\mathcal{W}^{-1}}}:{\mathbb{R}}^{{M}}\to{\ensuremath{{\mathcal{H}}_{{C}}}}$. Note that for finite-dimensional spaces, this inverse map always exists. These definitions allow us to outline the relations between the dynamics operators ${\mathcal{A}}$ and $A$, and the measurement operators ${\mathcal{K}}$ and ${\ensuremath{K}}$ using the commutative diagrams in Figure \[fig:commute\_sysop\_dyn\] and Figure \[fig:commute\_sysop\_meas\] respectively.
The finite-dimensional evolution equations equivalent to [(\[eq:ideal\_lin\_evol\])]{} in the dual space can be formulated as $$\begin{aligned}
{w}_{k+1} &= A{w}_k + \eta_k {\label{eq:k_measure}}\\
y_{k} &= {\ensuremath{K}}_{k} w_{k} +\zeta_k, {\label{eq:k_measure1}}\end{aligned}$$ where we have matrices $A\in {\mathbb{R}}^{{M}\times{M}}, \ {\ensuremath{K}}_{k}\in {\mathbb{R}}^{{N}\times{M}}$, the vectors ${w}_k, {w}\in{\mathbb{R}}^{{M}}$, and we have slightly abused notation to let $\eta_k$ and $\zeta_k$ denote their ${\ensuremath{{\mathcal{H}}_{{C}}}}$ counterparts. Note that the measurement operator ${\mathcal{K}}$ is simply a sampling of the function $f$ at an arbitrary set of sensing locations ${\mathcal{X}}= {\{x_1, \dots, x_{{N}}\}}$, where $x_i\in{\Omega}$: we will see how this affects the structure of ${\ensuremath{K}}_k$ momentarily. The equations [(\[eq:ideal\_lin\_evol\])]{} suggest an immediate extension to functional control problems. Pick another dictionary of atoms ${\mathcal{F}_{{D}}}= \begin{bmatrix}{\psi}(d_1) &\cdots & {\psi}(d_{{\ell}})
\end{bmatrix}$, ${\psi}(d_j)\in{\mathcal{H}}$, $d_j\in{\Omega}$, the span of which, denoted by ${\ensuremath{{\mathcal{H}}_{{D}}}}$, is a strict subspace of the RKHS ${\mathcal{H}}$ generated by the kernel. The functional evolution equation is then as follows: $$\begin{aligned}
{\label{eq:ideal_lin_evol_control}}
f_{k+1} &= {\mathcal{A}}f_k + {\mathcal{B}}{\delta}_k + \eta_k\\
y_k &= {\mathcal{K}}_k f_k + \zeta_k,\end{aligned}$$ where the control functions ${\delta}_k$ evolve in ${\ensuremath{{\mathcal{H}}_{{D}}}}$, and ${\mathcal{B}}:{\ensuremath{{\mathcal{H}}_{{D}}}}\to{\ensuremath{{\mathcal{H}}_{{C}}}}$. To derive the finite-dimensional equivalent of ${\mathcal{B}}$, we have to work out the structure of the matrix $B$: since ${\ensuremath{{\mathcal{H}}_{{C}}}}$ is not, in general, isomorphic to ${\ensuremath{{\mathcal{H}}_{{D}}}}$, this imposes strict restrictions on $B$. We derive $B$ using least squares using the inner product of ${\mathcal{H}}$. Let ${\delta}= \sum_{j=1}^{{\ell}}{\acute{{w}}}_j{k}(d_j,x)$, and let ${\mathcal{F}_{{C}}}=
\begin{bmatrix}{\psi}(c_1) &\cdots & {\psi}(c_{{M}})
\end{bmatrix}$ be the basis for ${\ensuremath{{\mathcal{H}}_{{C}}}}$. Then the projection of $\delta$ onto ${\ensuremath{{\mathcal{H}}_{{C}}}}$ can be derived as $$\begin{aligned}
\begin{bmatrix}
\l \delta, {\psi}(c_1) \r_{{\mathcal{H}}}\\
\vdots\\
\l \delta, {\psi}(c_{{M}}) \r_{{\mathcal{H}}}
\end{bmatrix}
&=
\underbrace{
\begin{bmatrix}
{k}(d_1,c_1) & \cdots & {k}(d_{{\ell}},c_1)\\
\vdots &\ddots &\vdots\\
{k}(d_1,c_{{M}}) & \cdots & {k}(d_{{\ell}},c_{{M}})
\end{bmatrix}}_{{{\ensuremath{K}}_{{C}{D}}}}
\begin{bmatrix}
{\acute{{w}}}_1\\
\vdots\\
{\acute{{w}}}_{{\ell}}
\end{bmatrix},\end{aligned}$$ using the reproducing property. This derivation shows that the operator $B = {{\ensuremath{K}}_{{C}{D}}}\in{\mathbb{R}}^{{M}\times{\ell}}$, the kernel matrix between the data ${C}$ generating the atoms ${\mathcal{F}_{{C}}}$ of ${\ensuremath{{\mathcal{H}}_{{C}}}}$ and the data ${D}$ generating the atoms ${\mathcal{F}_{{D}}}$ of ${\ensuremath{{\mathcal{H}}_{{D}}}}$. Using similar arguments, it can be shown that, given sensing locations ${X}= \{x_1,x_2,\dots, x_{{N}}\}$, ${{\ensuremath{K}}_{{D}}}\in{\mathbb{R}}^{{N}\times{\ell}}$ is the kernel matrix between ${X}$ and ${D}$. Thus the finite-dimensional evolution equations equivalent to [(\[eq:ideal\_lin\_evol\_control\])]{} are $$\begin{aligned}
{w}_{k} &= A{w}_k + {{\ensuremath{K}}_{{C}{D}}}{\acute{{w}}}_k{\label{eq:k_measure_c}}\\
y_{k} &= {\ensuremath{K}}_{k} w_{k} {\label{eq:k_measure_c1}}.\end{aligned}$$ We pause here to point out just how flexible the kernel-based framework is. First of all, the choice of kernel completely determines the space ${\mathcal{H}}$, which may allow wildly different functional outputs for the same dynamics matrix, as shown in Figure \[fig:kernel\_variation\]. Note also that the dynamical equations [(\[eq:k\_measure\_c\])]{} and [(\[eq:k\_measure\_c1\])]{} are *independent of the choice of domain ${\Omega}$*: different domains with different kernels may result in the same sequence of matrices ${\ensuremath{K}}_k$. This allows our results to hold for any domain over which a kernel can be defined, including examples like graphs, hidden Markov models, and strings, which are not typically studied in the controls literature, at virtually no extra complexity in implementation beyond the design of the actual sensors and actuators. This remarkable fact is why we denote our method to be *domain agnostic*.
\
Since ${\ensuremath{K}}_{k+1}$ is the kernel matrix between the data points and basis vectors, its rows are of the form ${\ensuremath{K}}_{(i)} =
\begin{bmatrix}
{k}(x_i, c_1) & {k}(x_i, c_2) & \cdots & {k}(x_i, c_{{M}})
\end{bmatrix}$. In systems-theoretic language, each row of the kernel matrix corresponds to a *measurement* at a particular location, and the matrix itself acts as a measurement operator. We define the *generalized observability matrix* [@zhou:bk:96] as $$\begin{aligned}
{\label{eq:obs_mat}}
{\mathcal{O}}_{{\Upsilon}} =
\begin{bmatrix}
{\ensuremath{K}}_{t_1} A^{t_1}\\
\cdots\\
{\ensuremath{K}}_{t_{L}} A^{t_{L}}
\end{bmatrix},\end{aligned}$$ where ${\Upsilon}= \{t_1, t_2, \dots, t_{{L}}\}$ are the set of instances $t_i$ when we apply the measurement operators ${\ensuremath{K}}_{t_i}$. Note that ${\mathcal{O}}_{{\Upsilon}}\in{\mathbb{R}}^{{N}{L}\times{M}}$. Similarly, we can define the *generalized controllability matrix* as $$\begin{aligned}
{\label{eq:control_mat}}
{\Psi}_{{\Upsilon}} =
\begin{bmatrix}
{A^{t_1}}^T {{{\ensuremath{K}}_{{D}}}}_{t_1} & {A^{t_2}}^T {{{\ensuremath{K}}_{{D}}}}_{t_2} & \cdots {A^{t_{L}}}^T {{{\ensuremath{K}}_{{D}}}}_{t_{L}}
\end{bmatrix},\end{aligned}$$ ${\Psi}_{{\Upsilon}}\in{\mathbb{R}}^{{M}\times{L}{\ell}}$ A linear system is said to be observable if ${\mathcal{O}}_{{\Upsilon}}$ has full column rank (i.e. $\mathrm{Rank}~{\mathcal{O}}_{{\Upsilon}}={M}$) and is controllable if ${\Psi}_{{\Upsilon}}$ has full row rank, for ${\Upsilon}= \{0, 1, \dots, {M}-1\}$ [@zhou:bk:96].
Observability guarantees that a feedback-based observer can be designed such that the estimate of $w$ denoted by $\hat {w_k}$ converges exponentially fast to the true state $w_k$. In particular, observability is the necessary condition for the existence of a unique solution to the Riccatti equation required in designing a Kalman filter. Therefore, when $\eta,\zeta$ have a zero mean Gaussian distribution, a Bayes optimal filter can be designed for estimating $w$ if and only if $\mathrm{Rank}~{\mathcal{O}}_{{\Upsilon}}={M}$. Similarly, controllability guarantees that a feedback-based controller can drive the current functional state of the system $f_{k}$ to a reference function $f_{\text{ref}}$, as long as $f_{\text{ref}}\in{\ensuremath{{\mathcal{H}}_{{C}}}}$.
We are now in a position to formally state the spatiotemporal monitoring and control problem considered: Given a spatiotemporally evolving system modeled using [(\[eq:ideal\_lin\_evol\_control\])]{}, choose a set of ${N}$ sensing locations ${\mathcal{X}}= {\{x_1, \dots, x_{{N}}\}}$ and ${\ell}$ actuating locations ${\mathcal{D}}= {\{d_1,\dots,d_{{\ell}}\}}$ such that even with ${N}\ll {M}$ and ${\ell}\ll {M}$, the functional evolution of the spatiotemporal model can be estimated robustly, and driven (controlled) to a reference function $f_{\text{ref}}$. Our approach to solve this problem relies on the design of the measurement operator ${\ensuremath{K}}$ such that the pair $(A,{\ensuremath{K}})$ is observable, and the control operator ${{\ensuremath{K}}_{{D}}}$ such that the pair $(A,{{\ensuremath{K}}_{{D}}})$ is controllable.
Theoretical Results {#sec:theory_results}
-------------------
In this section, we prove results concerning the observability of spatiotemporally varying functions modeled by the functional evolution and measurement equations [(\[eq:k\_measure\])]{} and [(\[eq:k\_measure1\])]{} formulated in Section \[sec:formulation\]. In particular, observability of the system states implies that we can recover the current state of the spatiotemporally varying function using a small number of sampling locations ${N}$, which allows us to 1) track the function, and 2) predict its evolution forward in time. It should be noted that the results are also applicable to controllability of the system in [(\[eq:k\_measure\_c1\])]{} since the structure of the control matrix $K_{CD}$ is also that of a Kernel matrix. We first show in Proposition \[prop:1\] that if $A$ has a full-rank Jordan decomposition, the kernel matrix meeting a condition called *shadedness* (to be defined below) is sufficient for the system to be observable. In Proposition \[prop:2\], we prove a lower bound on the number of sampling locations required for observability which holds for more general $A$. Finally, in Proposition \[prop:3\], we outline a method that achieves this lower bound for certain kernels. Since both ${\ensuremath{K}}$ and ${{\ensuremath{K}}_{{C}{D}}}$ are kernel matrices generated from a shared kernel, these observability results translate directly into controllability results.
To prove our results, we will leverage the spectral decomposition of $A$. Specifically, recall that any matrix $A\in{\mathbb{R}}^{{M}\times{M}}$ is similar to a unique block diagonal matrix ${\Lambda}$ (i.e. $\exists P\in{\mathbb{R}}^{{M}\times{M}}$ invertible such that $A = P{\Lambda}P^{-1}$) whose diagonal blocks are matrices of the form $$\begin{aligned}
{\Lambda}_k(\lambda_i, \lambda_i^*) :=
\begin{bmatrix}
{M}& {I_2}& \cdots & {0}\\
\vdots & \vdots & \ddots & {I_2}\\
{0}& {0}& \cdots & {M}{\label{eq:jor_com}}
\end{bmatrix}.\end{aligned}$$ where $(\lambda_i, \lambda_i^*)$ is a complex conjugate eigenvalue of $A$, and $
{M}= \begin{bmatrix}
\mu_1 & \mu_2\\
-\mu_2 & \mu_1
\end{bmatrix}$ and $ {I_2}= \begin{bmatrix}
1 & 0\\
0 & 1
\end{bmatrix}$. Real eigenvalues ${\lambda}_i$ correspond to the case ${M}= {\lambda}_i$ and ${I_2}= 1$. Thus the complete real Jordan form of $A$ will be the appropriate diagonal array of these blocks. If all the eigenvalues $\lambda_i$ are nonzero and real, we say the matrix has a *full-rank Jordan decomposition*.
\[def:shaded\] **(Shaded Kernel Matrix)** Let ${k}:{\Omega}\times{\Omega}\to{\mathbb{R}}$ be a positive-definite kernel on a compact domain ${\Omega}$. Let $C = [c_1, c_2, \cdots , c_{{M}}\}$, $c_j\in{\Omega}$ be the points generating a finite-dimensional covering of the reproducing kernel Hilbert space ${\mathcal{H}}$ associated to ${k}(x,y)$, and let ${\mathcal{X}}= {\{x_1, \dots, x_{{N}}\}}$, $x_i\in{\Omega}$ Let ${\ensuremath{K}}\in{\mathbb{R}}^{{N}\times{M}}$ be the kernel matrix, where $ {\ensuremath{K}}_{ij} := {k}(x_i,c_j)$. For each row ${\ensuremath{K}}_{(i)} := [k(x_i,c_1), k(x_i,c_2),\dots, k(x_i,c_{{M}})]$, define the set ${\mathcal{I}}_{(i)} := \{\iota_1^{(i)},\iota_2^{(i)},\dots, \iota_{{M}_i}^{(i)}\}$ to be the indices in the kernel matrix row $i$ which are nonzero. Then if $$\begin{aligned}
{\label{eq:shaded_cond}}
\bigcup_{1 \leq i \leq {N}} {\mathcal{I}}^{(i)} = \{1,2,\dots, {M}\},
\end{aligned}$$ we denote ${\ensuremath{K}}$ as a *shaded kernel matrix* (see figure \[shaded\_matrix\]).
This condition implies that the null space of the adjoint of ${\ensuremath{K}}$ as a linear operator between Euclidean spaces, i.e. ${\ensuremath{K}}^T:{\mathbb{R}}^{{N}}\to{\mathbb{R}}^{{M}}$ is trivial. Note that, in principle, for the Gaussian kernel, a single row generates a shaded kernel matrix, although this matrix can have many entries that are extremely close to zero. With this definition in place, we can prove the following proposition, which shows that if $A$ has a full-rank Jordan decomposition, a shaded kernel matrix is sufficient to prove observability.
\[prop:1\] Let ${k}:{\Omega}\times{\Omega}\to{\mathbb{R}}$ be a positive definite kernel on a domain ${\Omega}$. Let $C = [c_1, c_2, \cdots , c_{{M}}\}$, $c_j\in{\Omega}$ be the points generating a finite-dimensional covering of the reproducing kernel Hilbert space ${\mathcal{H}}$ associated to ${k}(x,y)$, and consider the discrete linear system on ${\mathcal{H}}$ given by the evolution and measurement equations [(\[eq:k\_measure\])]{} and [(\[eq:k\_measure1\])]{}. Let $A\in{\mathbb{R}}^{{M}\times{M}}$ be a full-rank Jordan decomposition of the form $A = {P}{\Lambda}{P}^{-1}$, where ${\Lambda}= \operatorname{diag}(\begin{bmatrix}{\Lambda}_1 & {\Lambda}_2 &\cdots & {\Lambda}_{{O}}\end{bmatrix})$, and there are no repeated eigenvalues. Given a set of time instances ${\Upsilon}= \{t_1,t_2,\dots,t_{{L}}\}$, and a set of sampling locations ${\mathcal{X}}={\{x_1, \dots, x_{{N}}\}}$, the system [(\[eq:k\_measure\])]{} is observable if the kernel matrix ${\ensuremath{K}}_{ij} := {k}(x_i,c_j)$ is shaded, ${\ensuremath{K}}^D$, the row vector generated by summing the rows of ${\ensuremath{K}}$, has all nonzero entries, ${\Upsilon}$ has distinct values, and $|{\Upsilon}| \geq {M}$.
To begin, consider a system where $A = {\Lambda}$, with Jordan blocks $\{{\Lambda}_1, {\Lambda}_2, \dots, {\Lambda}_{{O}}\}$ along the diagonal. Then $A^{t_i} = \operatorname{diag}(\begin{bmatrix}{\Lambda}_1^{t_i} & {\Lambda}_2^{t_i} & \cdots & {\Lambda}_{{O}}^{t_i}\end{bmatrix})$. We have that $${\mathcal{O}}_{{\Upsilon}} =
\begin{bmatrix}
{\ensuremath{K}}A^{t_1}\\
\cdots\\
{\ensuremath{K}}A^{t_{L}}
\end{bmatrix}
=
\underbrace{
\begin{bmatrix}
{\ensuremath{K}}& \cdots & {\ensuremath{K}}\end{bmatrix}}_{{\widehat{\mathbf{K}}}\in{\mathbb{R}}^{{N}\times{M}{L}}}
\underbrace{
\begin{bmatrix}
{\Lambda}_1^{t_1} & \cdots & 0\\
\vdots & \ddots & \vdots\\
0 & \cdots & {\Lambda}_{{O}}^{t_1}\\
\hline
\vdots & \ddots & \vdots\\
\hline
{\Lambda}_1^{t_{{L}}} & \cdots & 0\\
\vdots & \ddots & \vdots\\
0 & \cdots & {\Lambda}_{{O}}^{t_{L}}
\end{bmatrix}}_{{\widehat{\mathbf{A}}}\in{\mathbb{R}}^{{M}{L}\times{M}}}$$ Recall that a matrix’s rank is preserved under a product with an invertible matrix. Design a matrix $U\in{\mathbb{R}}^{{N}\times{N}}$ s.t. ${\widetilde{{\ensuremath{K}}}}:= U{\ensuremath{K}}$ is a matrix with one row vector of nonzeros, and all of the remaining rows as zeros. Then $\operatorname{rank}({\widehat{\mathbf{K}}}{\widehat{\mathbf{A}}}) = \operatorname{rank}(U{\widehat{\mathbf{K}}}{\widehat{\mathbf{A}}})$. Therefore, we have that $${\widetilde{{\ensuremath{K}}}}A^{t_j} =
\begin{bmatrix}
{\widetilde{{\ensuremath{K}}}}_{(1)}\\
0\\
\vdots\\
0
\end{bmatrix}
A^{t_j}\\
=
\begin{bmatrix}
k_{11}{\lambda}_1^{t_j} & \binom{t_j}{1}{\lambda}_1^{t_j-1} + k_{12}{\lambda}_1^{t_j} & \cdots & k_{1{M}}{\lambda}_{{O}}^{t_j}\\
0 & 0 & \cdots & 0\\
\vdots & \vdots & \ddots & 0\\
0 & 0 & \cdots & 0
\end{bmatrix}$$ Therefore, following some more elementary row operations encoded by $V\in{\mathbb{R}}^{{M}{L}\times{M}{L}}$, we get that $$V \begin{bmatrix}
{\widetilde{{\ensuremath{K}}}}& \cdots & {\widetilde{{\ensuremath{K}}}}\end{bmatrix}
\begin{bmatrix}
A^{t_1}\\
\vdots\\
A^{t_{{L}}}
\end{bmatrix}
=
\begin{bmatrix}
\tilde{k}_{11}{\lambda}_1^{t_1} & \cdots & \tilde{k}_{1{M}}{\lambda}_{{O}}^{t_1}\\
\tilde{k}_{11}{\lambda}_1^{t_2} & \cdots & \tilde{k}_{1{M}}{\lambda}_{{O}}^{t_2}\\
\vdots & \ddots & 0\\
\tilde{k}_{11}{\lambda}_1^{t_{{L}}} & \cdots
& \tilde{k}_{1{M}}{\lambda}_{{O}}^{t_{{L}}}\\
\mathbf{0} & \cdots & \mathbf{0}
\end{bmatrix}\\
=
\begin{bmatrix}
\boldsymbol{\Phi}\\
\widehat{\mathbf{0}}
\end{bmatrix}.$$ If the individual entries $\tilde{k}_{1i}$ are nonzero, and the Jordan block diagonals have nonzero eigenvalues, the columns of $\boldsymbol\Phi$ become linearly independent. Therefore, if ${L}\geq {M}$, the column rank of ${\mathcal{O}}_{{\Upsilon}}$ is ${M}$, which results in an observable system.
To extend this proof to matrices $A = {P}{\Lambda}{P}^{-1}$, note that $${\mathcal{O}}_{{\Upsilon}} =
\begin{bmatrix}
{\ensuremath{K}}A^{t_1}\\
\cdots\\
{\ensuremath{K}}A^{t_{L}}
\end{bmatrix}
=
\begin{bmatrix}
{\ensuremath{K}}{P}{\Lambda}^{t_1}{P}^{-1}\\
\cdots\\
{\ensuremath{K}}{P}{\Lambda}^{t_{L}}{P}^{-1}.
\end{bmatrix}
=
\begin{bmatrix}
{\ensuremath{K}}& \cdots & {\ensuremath{K}}\end{bmatrix}
\boldsymbol{{P}}
\boldsymbol{{\Lambda}}^t
\boldsymbol{{P}^{-1}},$$ where $\boldsymbol{{P}}\in{\mathbb{R}}^{{M}{L}\times{M}{L}}$, $\boldsymbol{{\Lambda}}^t\in{\mathbb{R}}^{{M}{L}\times{M}{L}}$, and $\boldsymbol{{P}^{-1}}\in{\mathbb{R}}^{{M}{L}\times{M}{L}}$ are the block diagonal matrices associated with the system. Since $\boldsymbol{{P}}$ is an invertible matrix, the conclusions about the column rank drawn before still hold, and the system is observable.
When the eigenvalues of the system matrix are repeated, it is not enough for ${\ensuremath{K}}$ to be shaded. The next proposition proves a lower bound on the number of observations required.
\[prop:2\] Suppose that the conditions in Proposition \[prop:1\] hold, with the relaxation that the Jordan blocks $\begin{bmatrix}{\Lambda}_1 & {\Lambda}_2 &\cdots & {\Lambda}_{{O}}\end{bmatrix}$ may have repeated eigenvalues. Let ${r}$ be the number of unique eigenvalues of $A$, and let ${\gamma}({{\lambda}}_i)$ denote the geometric multiplicity of eigenvalue ${{\lambda}}_i$. Then there exist kernels ${k}(x,y)$ such that the lower bound ${l}$ on the number of sampling locations ${N}$ is given by the cyclic index of $A$, which can be computed as $$\begin{aligned}
{\label{eq:geom_mult}}
{l}= \max_{1\leq i\leq{r}}{\gamma}({{\lambda}}_i).
\end{aligned}$$
We first prove the lower bound. WLOG, let ${\mathbf{{\ensuremath{K}}}}$ have ${l}-1$ fully shaded, linearly independent rows, and write it as $$\begin{aligned}
{\mathbf{{\ensuremath{K}}}}&= \begin{bmatrix}
k_{11} & k_{12} & \cdots & k_{1{M}} \\
\vdots & \vdots & \cdots & \vdots \\
k_{({l}-1)1} & k_{({l}-1)2} & \cdots & k_{({l}-1){M}}
\end{bmatrix}.
\end{aligned}$$ Since the cyclic index is ${l}$, this implies that at least one eigenvalue, say ${{\lambda}}$, has ${l}$ Jordan blocks. Define indices $j_1, j_2, \dots, j_{{l}} \in \{1,2,\dots,{M}\}$ as the columns corresponding to the leading entries of the ${l}$ Jordan blocks corresponding to ${{\lambda}}$. WLOG, let $j_1 = 1$. Using ideas similar to the last proof, we can write the observability matrix as $$\begin{aligned}
{\mathcal{O}}_{{\Upsilon}}
&:=
\begin{bmatrix}
{k}_{11}{{\lambda}}^{t_1} & \cdots & {k}_{1j_{{l}}}{{\lambda}}^{t_1} & \cdots\\
\vdots & \ddots &\vdots & \ddots\\
{k}_{11}{{\lambda}}^{t_{{L}}} & {k}_{1j_{{l}}}{{\lambda}}^{t_{{L}}} & \cdots\\
\vdots & \ddots & \vdots & \ddots\\
{k}_{({l}-1)1}{{\lambda}}^{t_1} \cdots & {k}_{({l}-1)j_{{l}}}{{\lambda}}^{t_1} & \cdots\\
\vdots & \ddots &\vdots & \ddots\\
{k}_{({l}-1)1}{{\lambda}}^{t_{{L}}} & \cdots & {k}_{({l}-1)j_{{l}}}{{\lambda}}^{t_{{L}}} & \cdots
\end{bmatrix}.
\end{aligned}$$ Define ${\boldsymbol{{{\lambda}}}}:= \begin{bmatrix}{{\lambda}}^{t_1} & {{\lambda}}^{t_2} & \cdots {{\lambda}}^{t_{{L}}}\end{bmatrix}^T$. Then the above matrix becomes $$\begin{aligned}
{\mathcal{O}}_{{\Upsilon}}
&:=
\begin{bmatrix}
{k}_{11}{\boldsymbol{{{\lambda}}}}& \cdots & {k}_{1j_2}{\boldsymbol{{{\lambda}}}}& \cdots & {k}_{1j_{{l}}}{\boldsymbol{{{\lambda}}}}& \cdots\\
\vdots & \ddots & \vdots & \ddots &\vdots & \ddots\\
{k}_{({l}-1)1}{\boldsymbol{{{\lambda}}}}& \cdots & {k}_{({l}-1)j_2}{\boldsymbol{{{\lambda}}}}& \cdots & {k}_{({l}-1)j_{{l}}}{\boldsymbol{{{\lambda}}}}& \cdots
\end{bmatrix}.
\end{aligned}$$ We need to show that one of the columns above can be written in terms of the others. This is equivalent to solving the linear system $$\begin{aligned}
\begin{bmatrix}
{k}_{1j_1}\\
{k}_{2j_1}\\
\vdots\\
{k}_{({l}-1)j_1}
\end{bmatrix}
&=
\begin{bmatrix}
{k}_{1j_2} & \cdots & {k}_{1j_{{l}}}\\
{k}_{2j_2} & \cdots & {k}_{2j_{{l}}}\\
\vdots & \ddots & \vdots\\
{k}_{({l}-1)j_2} & \cdots & {k}_{({l}-1)j_{{l}}}\\
\end{bmatrix}
\begin{bmatrix}
c_1\\
c_2\\
\vdots\\
c_{({l}-1)}
\end{bmatrix}.
\end{aligned}$$ Suppose the kernel matrix on the RHS is generated from the Gaussian kernel. From [@micchelli1984interpolation], it’s known that every principal minor of a Gaussian kernel matrix is invertible, which implies that ${\mathcal{O}}_{{\Upsilon}}$ cannot be observable.
We now prove a sufficient condition for the observability of a system with repeated eigenvalues, but with the condition that the Jordan blocks are trivial.
\[prop:3\] Suppose that the conditions in Proposition \[prop:1\] hold, with the relaxation that the Jordan blocks $\begin{bmatrix}{\Lambda}_1 & {\Lambda}_2 &\cdots & {\Lambda}_{{O}}\end{bmatrix}$ may have repeated eigenvalues, and where ${\Lambda}_i$ are single-dimensional. Let ${l}$ be the cyclic index of $A$. We define $$\begin{aligned}
{\label{eq:empKShadFull}}
{\mathbf{{\ensuremath{K}}}}= \begin{bmatrix}
{\ensuremath{K}}^{(1)}\\
\vdots\\
{\ensuremath{K}}^{({l})}
\end{bmatrix}
\end{aligned}$$ as the *${l}$-shaded matrix* which consists of ${l}$ shaded matrices with the property that any subset of ${l}$ columns in the matrix are linearly independent from each other. Then system [(\[eq:k\_measure\])]{} is observable if ${\Upsilon}$ has distinct values, and $|{\Upsilon}| \geq {M}$.
A cyclic index of ${l}$ for this system implies that there exists an eigenvalue ${{\lambda}}$ that’s repeated ${l}$ times. WLOG, let ${\mathbf{{\ensuremath{K}}}}$ have ${l}$ fully shaded, linearly independent rows, and, assume that the column indices corresponding to this eigenvalue are $\{1,2,\dots,{l}\}$. Define ${\boldsymbol{{{\lambda}}}}_i:= \begin{bmatrix}{{\lambda}}_i^{t_1} & {{\lambda}}_i^{t_2} & \cdots {{\lambda}}_i^{t_{{L}}}\end{bmatrix}^T$. Then $$\begin{aligned}
{\mathcal{O}}_{{\Upsilon}}
&:=
\begin{bmatrix}
k_{11} {\boldsymbol{{{\lambda}}}}_1 & k_{12} {\boldsymbol{{{\lambda}}}}_2 & \cdots & k_{1{M}} {\boldsymbol{{{\lambda}}}}_{{M}}\\
\vdots & \vdots & \ddots & \vdots\\
k_{{l}1} {\boldsymbol{{{\lambda}}}}_1 & k_{{l}2} {\boldsymbol{{{\lambda}}}}_2 & \cdots & k_{{l}{M}} {\boldsymbol{{{\lambda}}}}_{{M}}.
\end{bmatrix}
\end{aligned}$$ Let ${\boldsymbol{{{\lambda}}}}_1 = {\boldsymbol{{{\lambda}}}}_2 = \cdots {\boldsymbol{{{\lambda}}}}_{{l}} := {\boldsymbol{{{\lambda}}}}$. Focusing on these first ${l}$ columns of this matrix, this implies that we need to find constants $c_1,c_2,\dots, c_{{l}-1}$ s.t. $$\begin{aligned}
\begin{bmatrix}
k_{11}\\
\vdots\\
k_{{l}1}
\end{bmatrix}
&=
c_1
\begin{bmatrix}
k_{12}\\
\vdots\\
k_{{l}2}
\end{bmatrix}
+ \cdots +
c_{{l}-1}
\begin{bmatrix}
k_{1{l}}\\
\vdots\\
k_{{l}{l}}.
\end{bmatrix}
\end{aligned}$$ However, these columns are linearly independent by assumption, and thus no such constants exist, implying that ${\mathcal{O}}_{{\Upsilon}}$ is observable.
An example of a kernel such that any subset of ${l}$ columns in ${\mathbf{{\ensuremath{K}}}}$ are linearly independent of each other is the Gaussian kernel evaluated on sampling locations ${\{x_1, \dots, x_{{N}}\}}$, where $x_i\in{\Omega}\subset{\mathbb{R}}^d$, and $x_i\neq x_j$.
We can reuse Propositions \[prop:1\], \[prop:2\], and \[prop:3\] to prove kernel controllability results, because the structure of the control matrix ${{\ensuremath{K}}_{{C}{D}}}$ in [(\[eq:k\_measure\_c\])]{} is also that of a kernel matrix.
Kernel ${k}$, basis points ${\mathcal{C}}$, final time step $T_f$. $1)$ Sample data $\{y^i_k\}_{i=1}^{{N}}$ from $f(x,k)$. $2)$ Estimate ${\widehat{w}}_k$ via standard kernel inference procedure. $3)$ Store weights ${\widehat{w}}_k$ in matrix ${\mathcal{W}}\in{\mathbb{R}}^{{M}\times T_f}$. Infer ${\widehat{A}}$ using method of choice (e.g. matrix least squares). Compute the covariance matrix ${\widehat{B}}$ of the observed weights ${\mathcal{W}}$. estimated transition matrix ${\widehat{A}}$, predictive covariance matrix ${\widehat{B}}$.
Kernel ${k}$, basis points ${\mathcal{C}}$, estimated system matrix ${\widehat{A}}$, estimated covariance matrix ${\widehat{B}}$. Compute the cyclic index ${l}$ of ${\widehat{A}}$, and compute [(\[eq:empKShadFull\])]{}, by possibly iterating over ${\mathcal{X}}= {\{x_1, \dots, x_{{N}}\}}$. Use ${\widehat{A}}$, ${\widehat{B}}$, and ${\mathbf{{\ensuremath{K}}}}$ to initialize a state-observer (e.g. Kalman filter (KF)) on ${\ensuremath{{\mathcal{H}}_{{C}}}}$. 1) Sample data $\{y^i_k\}_{i=1}^{{N}}$ from $f(x,k)$. 2) Propagate KF estimate ${\widehat{w}}_{k+1}$ forward to time $t_f$, correct using measurement feedback with $\{y^i_k\}_{i=1}^{{N}}$. 3) Output predicted function $\widehat{f}(x,k+1)$ and predictive covariance of KF.
Kernel ${k}$, basis points ${\mathcal{C}}$, estimated system matrix ${\widehat{A}}$, estimated covariance matrix ${\widehat{B}}$, and function $f_{\text{ref}}$ to drive initial function to. (see Algorithm \[alg:egp\_inf\]). Use Jordan decomposition of ${\widehat{A}}$ to obtain control locations ${\mathcal{D}}$, compute kernel matrix ${{\ensuremath{K}}_{{C}{D}}}\in{\mathbb{R}}^{{\ell}\times{M}}$ between ${\mathcal{D}}$ and ${\mathcal{C}}$, and initialize controller (e.g. LQR) utilizing $({\widehat{A}}, {\widehat{B}})$. 1) Sample data $\{y^i_k\}_{i=1}^{{N}}$ from $f(x,k)$. 2) Utilize observer to estimate ${\widehat{w}}_{k+1}$. 3) Use ${\widehat{w}}_{k+1}$ and $f_{\text{ref}}$ as input to controller to get feedback.
Experimental Results {#sec:experimental_results}
====================
We report experimental results on controlling synthetic and modeling real-world data. All experiments were performed using MATLAB on a laptop running Ubuntu 14.04 with $8$ GB of RAM, and an Intel core i7 processor.
Prediction of global ocean surface temperature
----------------------------------------------
We first analyzed the feasibility of this modeling approach on a large dataset: the $4$ km AVHRR Pathfinder project, which is a satellite monitoring global ocean surface temperature. This data was obtained from the National Oceanographic Data Center. The data consists of longitude-latitude measurements on a 2D domain ${\Omega}\subset[-180,180]\times[-90,90]$; this dataset is challenging, with measurements at over $37$ million coordinates, and several missing pieces of data. The goal was to learn the day and night temperature models $f_k(x,y)\in{\ensuremath{{\mathcal{H}}_{{C}}}}$, where ${\ensuremath{{\mathcal{H}}_{{C}}}}$ was generated using the Gaussian kernel ${k}(x,y) = e^{-(\|x-y\|^2/2{\sigma}^2)}$. We first did a search for the ideal bandwidth ${\sigma}$ for a $304$-dimensional sparse Gaussian process model with a Gaussian kernel. The set of atoms ${\mathcal{F}_{{C}}}$ was determined through a linear independence test based sparsification algorithm [@csato2002sparse]. Once the parameters were chosen, a budgeted GP was learned for each date, resulting in weight vectors ${w}_i, \ i\in\{1,2,\dots,365\}$. We used Algorithm \[alg:egp\_trans\] to infer ${\widehat{A}}$, and applied Algorithm \[alg:egp\_inf\] with ${N}\in \{280,500,1000,2000\}$ chosen randomly in the ${\Omega}$ to track the system state given a random initial condition ${w}_0$. Figures \[fig:pathfinder\_errors\_boxplots\_day\] and \[fig:pathfinder\_errors\_boxplots\] show a comparison of the deviation in percentage of the estimated values from the real data, averaged over all the days. As can be seen, the observer enables the prediction of functional evolution *without needing all the measurements (37 million)*, and performance comparable to sampling over all locations is obtained with sampling only over $2,000$ locations. Note that here, even though the system model is observable at ${N}=280$, since the dynamics are not truly linear in ${\ensuremath{{\mathcal{H}}_{{C}}}}$, we get better performance with more sampling locations. Finally, \[fig:pathfinder\_tr\_times\_boxplots\_day\] and \[fig:pathfinder\_tr\_times\_boxplots\] show that the time required to estimate the state during function tracking with kernel observer are an order of magnitude better than retraining the model every time step (“original” in the figure), with comparable performance.
\
Control of a linear PDE
-----------------------
We then employed kernel controllers for controlling an approximation to the scalar diffusion equation $u_t = bu_{xx}$ on the domain ${\Omega}=[0,1]$, with $b=0.25$. The solution to this equation is infinite-dimensional, so we chose a kernel ${k}(x,y) = e^{-(\|x-y\|^2/2{\sigma}^2)}$, and a set of atoms ${\mathcal{F}_{{C}}}={\{c_1,\dots,c_{{M}}\}}$, $c_i\in{\Omega}$, with ${M}= 25$ generating ${\ensuremath{{\mathcal{H}}_{{C}}}}$, the space approximating ${\mathcal{H}}$, and another set of atoms ${\mathcal{F}_{{D}}}=\{{\psi}(d_1),\dots,{\psi}(d_{{\ell}})\}$, $d_j\in{\Omega}$, ${\ell}=13$, generating the control space ${\ensuremath{{\mathcal{H}}_{{D}}}}$. The number of, and the location of the observations was chosen to be the same as that of the actuation locations $d_j$. First, tests (not reported here) were conducted to ensure that the solution to the diffusion equation is well approximated in ${\ensuremath{{\mathcal{H}}_{{C}}}}$. Algorithm \[alg:egp\_trans\] was then used to infer ${\widehat{A}}$. Figure \[fig:uncontrolled\_pde\] shows an example of an initial function $f_{\text{init}}$ evolving according to the PDE. A reference function $f_{\text{ref}}\in{\ensuremath{{\mathcal{H}}_{{C}}}}$ was chosen to drive $f_{\text{init}}$ to $f_{\text{ref}}$ under the action of the PDE. Finally, Algorithm \[alg:egp\_control\] was used to control the PDE. Figure \[fig:controlled\_pde\] shows $f_{\text{init}}$ being driven to $f_{\text{ref}}$, while Figure \[fig:controlled\_pde\_error\] shows the absolute value of the error between $f_k $ and $f_{\text{ref}}$ as a function of time.
Conclusions
===========
In this paper we presented a systems theoretic approach to the problem of modeling, estimating, and controlling complex spatiotemporally evolving phenomena. Our approach focused on developing a predictive model of spatiotemporal evolution by layering a dynamical systems prior over temporal evolution of weights of a kernel model. The resulting model can approximate PDE evolution, while it has the form of a finite state linear dynamical system. The lower bounds on the number of sampling and actuation locations provided in this paper are non-conservative, as such they provide direct guidance in ensuring robust real-world sensor network and actuation matrix design that must also account for fault-tolerance and reliability considerations.
[^1]: This work was supported in parts by DOE Award Number DE-FE0012173 and AFOSR Award Number FA9550-14-1-0399. Hassan Kingravi is with Pindrop Security, Harshal Maske, and Girish Chowdhary are with the Distributed Autonomous Systems (DAS) laboratory Oklahoma State University,[{[email protected], [email protected], [email protected]}]{}
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We demonstrate a state reconstruction technique which provides either the Wigner function or the density matrix of a field mode and requires only avalanche photodetectors, without any phase or amplitude discrimination power. It represents an alternative to quantum homodyne tomography of simpler implementation.'
author:
- Alessia Allevi
- Alessandra Andreoni
- Maria Bondani
- Giorgio Brida
- Marco Genovese
- Marco Gramegna
- Stefano Olivares
- 'Matteo G. A. Paris'
- Paolo Traina
- Guido Zambra
title: 'State reconstruction by on/off measurements'
---
Introduction
============
The characterization of states and operations at the quantum level plays a leading role in the development of quantum technology. A state reconstruction technique is a method that provides the complete description of a physical system upon the measurements of an observable or a set of observables [@gentomo]. An effective reconstruction technique gives the maximum possible knowledge of the state, thus allowing one to make the best, at least the best probabilistic, predictions on the results of any measurement that may be performed on the system. At a first sight, there is an unavoidable tradeoff between the complexity of the detection scheme and the amount of extractable information, which can be used to reconstruct the quantum state [@mg]. Currently, the most effective quantum state reconstruction technique for the radiation field is quantum homodyne tomography [@wv99; @sp], which requires the measurement of a continuous set of field quadrature and allows for the reliable reconstruction of any quantity expressible in terms of an expectation value [@rec1; @rec2; @rec3; @rec4; @rec5; @rec6]. A question arises on whether the tradeoff may be overcome by a suitable experimental configuration or it corresponds to some fundamental limitations. Here we demonstrate that no specific discrimination power is required to the detector in either amplitude or phase, and that full state reconstruction is possible by a suitable processing of the data obtained with detectors revealing light in the simplest way, i.e. on/off detectors, such as single-photon avalanche photodiodes. Of course, some form of phase and/or amplitude modulation is necessary, which, in our scheme, is imposed to the field before the detection stage. In fact, our technique is built on the completeness of any set of displaced number states [@WB96; @WV96; @L96; @TO971; @TO972; @KB99] and the reliable maximum likelihood reconstruction of arbitrary photon-number distributions [@CVP] from on/off data.
The paper is structured as follows. In Section \[s:rec\] we describe our reconstruction method, whereas in Section \[s:exp\] the experimental setup used in the reconstruction is described in some details. Results are illustrated in Section \[s:res\] and the error analysis is reported in Section \[s:err\]. In Section \[s:dis\] we discuss few additional topics while Section \[s:out\] closes the paper with some concluding remarks.
State reconstruction by on/off measurements {#s:rec}
===========================================
We start to describe our reconstruction technique by observing that the modulation of a given signal, described by the density matrix $\varrho$, corresponds to the application of a coherent displacement (probe) $\varrho_\alpha = D(\alpha) \varrho D^\dag (\alpha)$, $\alpha\in
{\mathbb C}$. In practice, it can be easily obtained by mixing the state under investigation with a known coherent reference in a beam-splitter or a Mach-Zehnder interferometer [@bs96]. Upon varying amplitude and phase of the coherent reference and/or the overall transmissivity of the interferometer, the modulation may be tuned in a relatively broad range of values. The main idea behind our method is simple: the photon distributions of coherently modulated signals, i.e. the diagonal elements $p_n(\alpha)=\langle n |\varrho_\alpha |n\rangle$ of the density matrix $\varrho_\alpha$, contain relevant information about the complete density matrix of the original signal $\varrho$. Upon measuring or reconstructing the photon distribution $p_n(\alpha)$ for different values of the modulation one has enough information for full state reconstruction. By re-writing the above relation as $p_n(\alpha) = \sum_{km} D_{nk}(\alpha) \varrho_{km} D_{mn} (\alpha)$, the off diagonal matrix elements may be recovered upon inversion by least square method, i.e. [@TO971] $$\langle m+s | \varrho |m\rangle = N_\phi^{-1} \sum_{l=1}^{N_\phi}
\sum_{n=0}^{\bar n} F^{(s)}_{nm} (|\alpha|) p_n (|\alpha|
e^{i \phi_l}) e^{i s \phi_l}$$ where $N_\phi$ is the number of modulating phases, $\bar n$ the truncation dimension of the Fock space, and $F$ depends only on $|\alpha|$ [@TO971]. State reconstruction by the above formula requires, in principle, only phase modulation of the signal under investigation. Maximum likelihood methods and iterative procedures may be also used [@ZJ06]. On the other hand, the Wigner function may be reconstructed using its very definition in terms of displacement [@Cahill] $$W(\alpha) = \hbox{Tr}[D(\alpha)\varrho D^\dag (\alpha)\, (-)^{a^\dag
a}] = \sum_n (-)^n\, p_n (\alpha)\:.$$ As a matter of fact, the measurement of the photon distribution is challenging as photo-detectors that can operate as photon counters are rather rare and affected either by a low quantum efficiency [@burle] or require cryogenic conditions, thus impairing common use [@xxx; @serg]. Therefore, a method with displacement but without photo-counting has been used so far only for states in the 0-1 subspace of the Fock space [@KB99]. On the other hand, the experimental reconstructions of photon-number distributions for both continuous-wave and pulsed light beams is possible using simple on/off single-photon avalanche photodetectors. This requires the collection of the frequencies of the [*off*]{} events, $P_{0,k}=\sum_{n = 0}^\infty{( {1 - \eta_k})^n p_n}$ at different quantum efficiencies of the detector, $\eta_k$. The data are then used in a recursive maximum likelihood reconstruction algorithm that yields the photon-number distributions as $$p_n^{i + 1} = p_n^i \sum_{k = 1}^K
(A_{kn}/\sum_j {A_{kj}})(P_{0,k}/p_{0,k}[\{p_n^i\}]),$$ where $A_{kn} = \left({1 - \eta _k } \right)^n$ and $p_{0,k}$ is the probability of [*off*]{} events calculated from the reconstructed distribution at the $i$th iteration [@pcount]. The effectiveness of the method has been demonstrated for single-mode [@CVP] and multimode fields [@bip], and also applied to improve quantum key distribution [@TM08].
Since the implementation of the modulation is relatively easy, we have thus a reconstruction technique which provides the quantum state of radiation modes and requires only avalanche detectors, without any phase or amplitude discrimination power. Here, we develop the idea into a proper reconstruction technique and demonstrate the reconstruction of the Wigner function [@silb] and the density matrix for different states of the optical field.
Experimental setups {#s:exp}
===================
We have performed two experiments for the reconstruction of the Wigner function and the density matrix respectively. In Fig. \[f:setup\] we sketch the corresponding experimental setups: the upper panel for the measurement of the Wigner function and the lower panel for the density matrix (lower panel). The first set of measurements was performed on ps-pulsed light fields at 523 nm wavelength. The light source was the second-harmonic output of a Nd:YLF mode-locked laser amplified at 5 kHz (High Q Laser Production) delivering pulses of $\sim 5.4$ ps duration.
![(Color online) Schematic diagram of the two experimental setups. Upper panel: Nd:YLF, pulsed laser source; P, polarizer; HPD, hybrid photodetector; SGI, synchronous gated integrator; ADC, analog-to-digital converter. Lower panel: He-Ne, cw laser source; IF, interference filter; APD, single-photon avalanche photodiode. F, neutral-density filter; BS, beam splitter; Pz, piezoelectric movement; R, rotating ground glass plate. Components in dotted boxes are inserted/activated when necessary.[]{data-label="f:setup"}](f1_setup.ps){width="0.8\columnwidth"}
The field, spatially filtered by means of a confocal telescope, was split into two parts to give both signal and probe. The photon-number distribution of the probe was kept Poissonian, while the coherent photon-number distribution of the signal field was modified in order to get suitable states of light. In particular, we have generated two phase-insensitive classical states, namely a phase-averaged coherent state and a single-mode thermal state. The first one was obtained by changing the relative phase between signal and probe fields at a frequency of $\sim300$ Hz by means of a piezoelectric movement (Pz in the upper panel of Fig. \[f:setup\]), covering 1.28 $\mu$m span. On the other hand, to get the single-mode thermal state we inserted an Arecchi’s rotating ground glass disk on the pathway of the signal field followed by a pin-hole that selected a single coherence area.
Signal and probe fields were mixed in an unpolarizing cube beam-splitter (BS) and a portion of the exiting field was sent, through a multimode optical fiber (600 $\mu$m core-diameter), to a hybrid photodetector module (HPD, mod. H8236-40, Hamamatsu, maximum quantum efficiency $\eta_\mathrm{HPD}$=0.4 at 550 nm). Although the detector is endowed with partial photon resolving capability, we used it as a simple on/off detector by setting a threshold at the value corresponding to zero detected photons. Its output current pulses were suitably gate-integrated by a SR250 module (Stanford Research Systems, CA) and sampled to produce a voltage, which was digitized and recorded at each shot. In order to modulate the amplitude of the probe field, a variable neutral density filter was placed on its pathway. The maximum overall detection efficiency, calculated by including the losses of the collection optics, was $\eta_{max}=0.29$. We used a polarizer put in front of the fiber to vary the quantum efficiency of the detection chain from $\eta_{max}$ to 0.
Results {#s:res}
=======
Here we illustrate the reconstruction obtained for the Wigner function and the density matrix of phase-averaged coherent states and thermal states, as obtained by our method after recording the on/off statistics of amplitude- and/or phase-modulated signals.
![(Color online) State reconstruction by amplitude-modulation and on/off measurements. In the main plot we report the [*off*]{} frequencies as a function of the quantum efficiency as obtained when the signal under investigation is a phase-averaged coherent state and for different values of the probe intensity $|\alpha|^2$. The two insets show the reconstructed photon distributions for $|\alpha|^2=5.02$ and $|\alpha|^2=10.69$, corresponding to the off distributions indicated by the arrows. The vertical black bars denote the mean value of the photon number for the two distributions ($\langle a^\dag a\rangle = 11.3$, upper, and $\langle a^\dag a\rangle = 18.0$, lower). In the lower left plot we report the corresponding reconstructed Wigner function. In the lower right plot we report the Wigner functions for signals in (blue) thermal state and (yellow, with sharper peak) vacuum . \[f:wsamm\]](f2_wsamm.ps){width="0.88\columnwidth"}
In Fig. \[f:wsamm\] we report the reconstructed Wigner functions for a phase-averaged coherent state with real amplitude $z \simeq2.1$ and a thermal state with average number of photons $n_{th}\simeq2.4$. The Wigner function of the vacuum state is also reported for comparison. As it is apparent from the plots all the relevant features of the Wigner functions are well recovered, including oscillations and the broadening due to thermal noise. In this case raw data are frequencies of the [*off*]{} event (collected over 30000 laser shots) as a function of detector efficiency, taken at different amplitudes of the modulating field, whereas the intermediate step corresponds to the reconstruction of the $p_n(\alpha)$. The insets of Fig. \[f:wsamm\] display $p_n(\alpha)$ for the phase-averaged coherent state at two values of the modulation.
![(Color online) State reconstruction by phase-modulation and on/off measurements. In the upper plot we report the [*off*]{} frequencies as a function of the quantum efficiency as obtained when the signal under investigation is a coherent state and for different phase-shifts. The two insets show the reconstructed photon distributions for the two phase-modulated versions of the signal corresponding to maximum and minimum visibility at the output of the Mach-Zehnder interferometer. The vertical black bars denote the mean value of the photon number for the two distributions, $\langle a^\dag a\rangle = 3.5$ and $\langle a^\dag a\rangle = 2.9$. In the lower left plot (left) we report the corresponding reconstructed density matrix in the Fock representation (diagonal and subdiagonal elements). In the lower right plot we report the density matrix for the signal excited in a thermal state. \[f:dmphm\]](f3_dmphm.ps){width="0.88\columnwidth"}
The second set of measurements has been performed to achieve state reconstruction with phase modulation. Here the light source was a He-Ne laser beam attenuated to single-photon regime by neutral density filters. The spatial profile of the beam was purified from non-Gaussian components by a spatial filter. Beyond a beam-splitter, part of the beam was addressed to a control detector in order to monitor the laser amplitude fluctuations, while the remaining part was sent to a Mach-Zehnder interferometer. A piezo-movement system allowed changing the phase between the “short” and “long” paths by driving the position of the reflecting prism with nanometric resolution and high stability. The beam on the short arm represented the probe, while the beam on the long arm was the state to be reconstructed. In the first acquisition the signal was the coherent state itself while in the second acquisition it was a pseudo-thermal state generated as described above. The detector, a Perkin-Elmer Single Photon Avalanche Photodiode (SPCM-AQR) with quantum efficiency $\eta_{max}=0.67$, was gated by a 20 ns time window at a repetition rate of 200 kHz. To obtain a reasonable statistics, a single run consisted of 5 repetitions of acquisitions lasting 4 s each. In the bottom part of Fig. \[f:dmphm\] we report the reconstructed density matrix in the Fock representation (diagonal and subdiagonal) for a coherent state with real amplitude $z\simeq1.8$ and a thermal state with average number of photons equal to $n_{th}\simeq1.4$.
As it is apparent from the plots the off-diagonal elements are correctly reproduced in both cases despite the limited visibility. Here the raw data are frequencies of the [*off*]{} event as a function of the detector efficiency, taken at different phase modulations, $\phi$, whereas the intermediate step corresponds to the reconstruction of the photon distribution for the phase-modulated signals. In our experiments we used $N_\phi=12$ and $|\alpha|^2=0.01$ for the coherent state and $|\alpha|^2=1.77$ for the thermal state. The use of a larger $N_\phi$ would allow the reliable reconstruction of far off-diagonal elements, which is not possible in the present configuration. In the insets of Fig. \[f:dmphm\] we report the reconstructed distributions at the minimum and maximum of the interference fringes.
Error analysis {#s:err}
==============
The evaluation of uncertainties on the reconstructed states involves the contributions of experimental fluctuations of on/off frequencies as well as the statistical fluctuations connected with photon-number reconstruction. It has been argued [@KER97; @KER99] that fluctuations involved in the reconstruction of the photon distribution may generally result in substantial limitations in the information obtainable on the quantum state, e.g. in the case of multipeaked distributions [@zz]. For our purposes this implies that neither large displacement amplitudes may be employed, nor states with large field and/or energy may be reliably reconstructed, although the mean values of the fields measured here are definitely non-negligible. On the other hand, for the relevant regime of weak field or low energy, observables characterizing the quantum state can be safely evaluated from experimental data, including e.g. the parity operator used to reconstruct the Wigner function in the phase-space. In our experiments, the errors on the reconstructed Wigner function are of the order of the size of the symbols in Fig. \[f:dmphm\] whereas the absolute errors $\Delta_{nm}=|\varrho^{exp}_{nm}-\varrho^{th}_{nm}|$ on the reconstruction of the density matrix in the Fock basis are reported in Fig. \[f:err\].
![(Color online) Absolute difference $\Delta_{nm}=|\varrho^{exp}_{nm}
-\varrho^{th}_{nm}|$ between reconstructed and theoretical values of the density matrix elements for the coherent (left) and thermal (right) states used in our experiments.[]{data-label="f:err"}](f4_err.eps){width="0.88\columnwidth"}
Discussions {#s:dis}
===========
We have so far reconstructed the Wigner function and the density matrix for coherent and thermal states. The extension to highly nonclassical states does not require qualitative changes in the setups. The only difference stays in the displacement, which should be obtained with high transmissivity beam splitter in order to avoid mixing of the signal [@bs96].
As our method involves a beam splitter where the signal interferes with a reference state in order to obtain the displacement, we have optimized the effectiveness of mode matching and of the overall scheme by standard visibility test. We note that in this point our technique is similar to quantum homodyne tomography (QHT), where the signal is amplified by the mixing at a beam splitter with a strong local oscillator. The main difference with standard QHT, however, is the spectral domain of the measurement, which for QHT is confined to the sideband chosen for the measurement, while it is not in our case. The use of pulsed temporal homodyning [@ht1; @ht1a; @ht2; @ht3; @ht4] would remove this limitation of QHT. However, this technique is still challenging from the experimental point of view and thus of limited use. The effect of photodetectors efficiency should be also taken into account. This is a crucial issue for QHT, which may even prevent effective reconstruction [@prec]. For the present on/off reconstrucion, it does not dramatically affect the accuracy [@pcount]. Notice also that any uncertainty in the nominal efficiency $\eta_{max}$ of the involved photodetectors does not substantially affect the reconstruction [@pcount].
We stress that our method is especially suited for low excited states, as it does not involve intense fields and delicate balancing to reveal the relevant quantum fluctuations.
Conclusions {#s:out}
===========
In conclusion, we have demonstrated a state reconstruction technique providing Wigner function and density matrix of a field mode starting from on/off photodetection of amplitude- and/or phase-modulated versions of the signal under investigation. Our scheme is little demanding as to the detectors, with the amplitude and phase control required for full state characterization transferred to the optical setup, and appears to be reliable and simple especially for states with small number of photons. We foresee a possible widespread use in emerging quantum technologies like quantum information, metrology and lithography.
Acknowledgments {#acknowledgments .unnumbered}
===============
This work has been partially supported by the CNR-CNISM agreement, EU project QuCandela, Compagnia di San Paolo, MIUR-PRIN-2007FYETBY (CCQOTS), and Regione Piemonte (E14).
[99]{} G. M. D’Ariano, L. Maccone, and M. G. A. Paris, J. Phys. A, [**34**]{}, 93 (2001). Y. Bogdanov et al., JETP Lett. [**82**]{}, 180 (2005); C. Marquardt et al., Phys. Rev. Lett. [**99**]{}, 220401 (2007). D. G. Welsch et al., Prog. Opt. [**39**]{}, 63 (1999). , Lect. Not. Phys. [**649**]{}, M. G. A. Paris and J. Rehacek Eds. (Springer, Berlin, 2004); A. I. Lvovsky and M. G. Raymer, Rev. Mod. Phys. [**81**]{}, 299 (2009). D.T. Smithey et al., Phys. Rev. Lett. [**70**]{}, 1244 (1993); G. Breitenbach et al., Nature [**387**]{}, 471 (1997); M. Vasilyev et al., Phys. Rev. Lett. [**84**]{}, 2354 (2000); I. Lvovsky et al., Phys. Rev. Lett. [**87**]{}, 050402 (2001); V. Parigi et al., Science [**317**]{}, 1891 (2008); V. D’Auria et al., Phys. Rev. Lett. [**102**]{}, 020502 (2009). S. Wallentowitz and W. Vogel, Phys. Rev. A [**53**]{}, 4528 (1996). K. Banaszek and K. Wodkiewicz, Phys. Rev. Lett. [**76**]{}, 4344 (1996). D. Leibfried et al., Phys. Rev. Lett. [**77**]{}, 4281 (1996). T. Opatrny and D. G. Welsch, Phys. Rev A [**55**]{}, 1462 (1997). T. Opatrny, D. G. Welsch, and W. Vogel, Phys. Rev A [**56**]{}, 1788 (1997). K. Banaszek et al., Phys. Rev. A [**60**]{}, 674 (1999); Act. Phys. Slov. [**49**]{}, 643 (1999). G. Zambra et al., Phys. Rev. Lett. [**95**]{}, 063602 (2005); G. Brida et al., Las. Phys. [**16**]{}, 385 (2006); G. Brida et al., Open Syst. & Inf. Dyn. [**13**]{}, 333 (2006); M. Bondani et al., Adv. Sci. Lett. (in press) ArXiv:0810.4026. M. G. A. Paris, Phys. Lett. A [**217**]{}, 78 (1996). Z. Hradil, D. Mogilevtsev, and J. Řeháček, Phys. Rev. Lett. [**96**]{}, 230401 (2006); New J. Phys. [**10**]{}, 043022 (2008). K. E. Cahill and R. J. Glauber, Phys. Rev. **177**, 1882 (1969). G. Zambra, M. Bondani, A. S. Spinelli, and A. Andreoni, Rev. Sci. Instrum. [**75**]{}, 2762 (2004). J. Kim, S. Takeuchi, Y. Yamamoto, and H.H. Hogue, Appl. Phys. Lett. [**74**]{}, 902 (1999); A. Peacock et al., Nature [**381**]{}, 135 (1996); A.E.Lita et al., Opt. Exp. 16 (2008) 3032. G. Di Giuseppe, A. V. Sergienko, B. E. A. Saleh, and M. C. Teich in [*Quantum Information and Computation*]{}, E. Donkor, A. R. Pirich, and H. E. Brandt Eds., Proceedings of the SPIE [**5105**]{}, 39 (2003). A. R. Rossi, S. Olivares, and M. G. A. Paris, Phys. Rev. A [**70**]{}, 055801 (2004). G. Brida et al., Opt. Lett. [**31**]{}, 3508 (2006). T. Moroder et al., ArXiv:0811.0027v1 K. Laiho, M. Avenhaus, K. N. Cassemiro, Ch. Silberhorn, New J. Phys. [**11**]{}, 043012 (2009). K. Banaszek and K. Wodkiewicz, J. Mod. Opt. [**44**]{}, 2441 (1997). K. Banaszek, J. Mod. Opt. [**46**]{}, 675 (1999). G. Zambra and M. G. A. Paris, Phys. Rev. A [**74**]{}, 063830 (2006). H. Hansen, T. Aichele, C. Hettich, P. Lodahl, A. I. Lvovsky, J. Mlynek, S. Schiller, Opt. Lett. [**26**]{}, 1714 (2001). A. Zavatta, M. Bellini,P. L. Ramazza, F. Marin, F. T. Arecchi, J. Opt. Soc. Am. B [**19**]{}, 1189 (2002). T. Hirano, H. Yamanaka, M. Ashikaga, I. Konishi, R. Namiki, Phys. Rev. A[68]{}, 042331 (2003). J. Wenger, R. Tualle-Brouri, P. Grangier, Opt. Lett. [**29**]{} 1267 (2004). Y. Eto, T. Tajima, Y. Zhang Y, T. Hirano, Opt. Lett. [**32**]{}, 1698 (2007). G. M. D’Ariano, M. G. A Paris, and M. F. Sacchi, Adv. Im. Electr. Phys. [**128**]{}, 205 (2003).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Many real-world applications involve multivariate, geo-tagged time series data: at each location, multiple sensors record corresponding measurements. For example, air quality monitoring system records PM2.5, CO, etc. The resulting time-series data often has missing values due to device outages or communication errors. In order to impute the missing values, state-of-the-art methods are built on Recurrent Neural Networks (RNN), which process each time stamp sequentially, prohibiting the direct modeling of the relationship between distant time stamps. Recently, the self-attention mechanism has been proposed for sequence modeling tasks such as machine translation, significantly outperforming RNN because the relationship between each two time stamps can be modeled explicitly. In this paper, we are the first to adapt the self-attention mechanism for multivariate, geo-tagged time series data. In order to jointly capture the self-attention across multiple dimensions, including time, location and the sensor measurements, while maintain low computational complexity, we propose a novel approach called Cross-Dimensional Self-Attention (CDSA) to process each dimension sequentially, yet in an order-independent manner. Our extensive experiments on four real-world datasets, including three standard benchmarks and our newly collected NYC-traffic dataset, demonstrate that our approach outperforms the state-of-the-art imputation and forecasting methods. A detailed systematic analysis confirms the effectiveness of our design choices.'
author:
- 'Jiawei Ma$^1$[^1]'
- 'Zheng Shou$^{1*}$'
- Alireza Zareian$^1$
- Hassan Mansour$^2$
- Anthony Vetro$^2$
- 'Shih-Fu Chang$^1$'
- |
\
$^{1}$Columbia University $^{2}$Mitsubishi Electric
bibliography:
- 'Section/reference.bib'
title: 'CDSA: Cross-Dimensional Self-Attention for Multivariate, Geo-tagged Time Series Imputation'
---
Conclusion
==========
In this paper, we have proposed a cross-dimensional self-attention mechanism to impute the missing values in multivariate, geo-tagged time series data. We have proposed and investigated three methods to model the crossing-dimensional self-attention. Experiments show that our proposed model achieves superior results to the state-of-the-art methods on both imputation and forecasting tasks. Given the encouraging results, we plan to extend our CDSA mechanism from multivariate, geo-tagged time series to the input that has higher dimension and involves multiple data modalities. Furthermore, we will publicly release the collected NYC-traffic dataset for future research.
Acknowledgments
===============
This work is supported by Mitsubishi Electric.
[^1]: indicates equal contributions.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Low-lying nuclear states of Sm isotopes are studied in the framework of a collective Hamiltonian based on covariant energy density functional theory. Pairing correlation are treated by both BCS and Bogoliubov methods. It is found that the pairing correlations deduced from relativistic Hartree-Bogoliubov (RHB) calculations are generally stronger than those by relativistic mean-field plus BCS (RMF+BCS) with same pairing force. By simply renormalizing the pairing strength, the diagonal part of the pairing field is changed in such a way that the essential effects of the off-diagonal parts of the pairing field neglected in the RMF+BCS calculations can be recovered, and consequently the low-energy structure is in a good agreement with the predictions of the RHB model.'
author:
- 'J. Xiang'
- 'Z. P. Li'
- 'J. M. Yao'
- 'W. H. Long'
- 'P. Ring'
- 'J. Meng'
title: 'Effect of pairing correlations on nuclear low-energy structure: BCS and general Bogoliubov transformation'
---
The study of nuclear low-lying states is of great importance to unveil the low-energy structure of atomic nuclei and turns out to be essential to understand the evolution of shell structure and collectivity [@Meng98; @Hagen12; @Kshetri06], nuclear shape phase transitions [@Meng05; @Casten06; @Cejnar10], shape coexistence [@Heyde11], the onset of new shell gaps [@Ozawa2000], the erosion of traditional magic numbers [@Sorlin08], etc. The understanding and the quantitative description of low-lying states in nuclei necessitate an accurate modeling of the underlying microscopic nucleonic dynamics.
Density functional theory (DFT) is a reliable platform for studying the complicated nuclear excitation spectra and electromagnetic decay patterns [@BHR.03; @JacD.11; @Vretenar05; @Meng06; @Meng2013FrontiersofPhysics55]. Since the DFT scheme breaks essential symmetries of the system, this requires to include the dynamical effects related to the restoration of broken symmetries, as well as the fluctuations in the collective coordinates. In recent years several accurate and efficient models and algorithms, based on microscopic density functionals or effective interactions, have been developed that perform the restoration of symmetries broken by the static nuclear mean field, and take the quadrupole fluctuations into account [@NVR.06a; @NVR.06b; @BH.08; @Yao08; @Yao.09; @Yao.10; @RE.10]. This level of implementation is also referred as the multi-reference (MR)-DFT [@Lacroix09]. Compared with MR-DFT, the model of a collective Hamiltonian with parameters determined in a microscopic way from self-consistent mean-field calculations turns out to be a powerful tool for the systematical studies of nuclear low-lying states [@PR.04; @Nik.09; @Nik.11], with much less numerical demanding. Even for the heavy nuclei full triaxial calculations can be relatively easily carried out with a five-dimension collective Hamiltonian [@Li.10]. It has achieved great success in describing the low-lying states in a wide range of nuclei, from $A\sim 40$ to superheavy nuclei including spherical, transitional, and deformed ones [@Nik.09; @Li.09; @Li.09b; @Li.10; @Li.11; @Nik.11; @Yao11-lambda; @Mei12; @Del10].
For open-shell nuclei, pairing correlations between nucleons have important influence on low-energy nuclear structure [@Dean03]. In the relativistic scheme they could be taken into account using the BCS ansatz [@GRT.90] or full Bogoliubov transformation [@KR.91; @Rin.96]. Compared with the simple BCS method, the consideration of pairing correlations through the Bogoliubov transformation is numerically demanding for heavy triaxial deformed nuclei. It has been demonstrated that there is no essential difference between BCS and Bogoliubov methods for the descriptions of the ground-state of stable nuclei [@Ring80]. Girod [*et al.*]{} have compared the results obtained from Hartree-Fock-Bogoliubov (HFB) and Hartree-Fock plus BCS (HF+BCS) calculations, including the potential energy surfaces (PESs), pairing gaps, and pairing energies as functions of the axial deformation [@Girod83]. It has been shown that the PESs given by these two methods are very similar. Moreover, the pairing gaps and energies from the HF+BCS calculations are slightly smaller than those from the HFB calculation. In view of these facts, it is natural to test the validity of the BCS ansatz in describing the the low-energy structure of nuclei, as referred to [the RHB method]{}. Aiming at this point, the comparisons are performed within the covariant density functional based 5DCH model, specifically between the triaxial deformed RMF+BCS and RHB calculations. Due to the emergence of an abrupt shape-phase-transition [@Li.09], the even-even Sm isotopes with $134\leqslant A\leqslant 154$ are taken as the candidates in this study.
Practically nuclear excitations determined by quadrupole vibrational and rotational degrees of freedom can be treated by introducing five collective coordinates, i.e., the quadrupole deformations $(\beta,\gamma)$ and Euler angles ($\Omega=\phi,\theta,\psi$) [@Pro.99]. The quantized 5DCH that describes the nuclear excitations of quadrupole vibration, rotation and their couplings can be written as, $$\label{hamiltonian-quant}
\hat{H} =
\hat{T}_{\textnormal{vib}}+\hat{T}_{\textnormal{rot}}
+V_{\textnormal{coll}} \; ,$$ where $V_{\textnormal{coll}}$ is the collective potential, and $\hat{T}_{\textnormal{vib}}$ and $\hat{T}_{\textnormal{rot}}$ are respectively the vibrational and rotational kinetic energies, $$\begin{aligned}
\label{Vcoll}
{V}_{\text{coll}} =& E_{\text{tot}}(\beta,\gamma) - \Delta V_{\text{vib}}(\beta,\gamma) - \Delta V_{\text{rot}}(\beta,\gamma),\\
\hat{T}_{\textnormal{vib}} =&-\frac{\hbar^2}{2\sqrt{wr}} \left\{\frac{1}{\beta^4} \left[\frac{\partial}{\partial\beta}\sqrt{\frac{r}{w}}\beta^4
B_{\gamma\gamma} \frac{\partial}{\partial\beta}\right.\right.\nonumber\\
& \left.\left.- \frac{\partial}{\partial\beta}\sqrt{\frac{r}{w}}\beta^3 B_{\beta\gamma}\frac{\partial}{\partial\gamma} \right]+\frac{1}{\beta\sin{3\gamma}} \left[
-\frac{\partial}{\partial\gamma} \right.\right.\label{Tvib}\\
& \left.\left.\sqrt{\frac{r}{w}}\sin{3\gamma} B_{\beta \gamma}\frac{\partial}{\partial\beta} +\frac{1}{\beta}\frac{\partial}{\partial\gamma} \sqrt{\frac{r}{w}}\sin{3\gamma} B_{\beta \beta}\frac{\partial}{\partial\gamma}
\right]\right\}\nonumber,\\
\hat{T}_{\textnormal{\textnormal{\textnormal{rot}}}}
=&\frac{1}{2}\sum_{k=1}^3{\frac{\hat{J}^2_k}{\mathcal{I}_k}}.\label{Trot}\end{aligned}$$ In eq. (\[Vcoll\]), $E_{\textnormal{tot}}(\beta,\gamma)$ is the binding energy determined by the constraint mean-field calculations, and the terms $\Delta V_{\textnormal{vib}}$ and $\Delta V_{\textnormal{rot}}$, calculated in the cranking approximation [@Ring80], are zero-point-energies (ZPE) of vibrational and rotational motions, respectively. In eq. (\[Trot\]), $\hat{J}_k$ denotes the components of the angular momentum in the body-fixed frame of the nucleus. Moreover the mass parameters $B_{\beta\beta}$, $B_{\beta\gamma}$, $B_{\gamma\gamma}$ in eq. (\[Tvib\]), as well as the moments of inertia $\mathcal{I}_k$ in eq. (\[Trot\]), depend on the quadrupole deformation variables $\beta$ and $\gamma$, $$\begin{aligned}
\mathcal{I}_k =& 4B_k\beta^2\sin^2(\gamma-2k\pi/3),&k=&1, 2, 3,\end{aligned}$$ where $B_k$ represents inertia parameter. In eq. (\[Tvib\]), the additional quantities $r=B_1B_2B_3$ and $w=B_{\beta\beta}B_{\gamma\gamma}-B_{\beta\gamma}^2 $ define the volume element of the collective space. The corresponding eigenvalue problem is solved [by expanding the eigenfunctions on]{} a complete set of basis functions in the collective space of the quadrupole deformations $(\beta, \gamma)$ and Euler angles $(\Omega =\phi, \theta, \psi)$.
The dynamics of the 5DCH is governed by seven functions of the intrinsic deformations $\beta$ and $\gamma$: the collective potential $V_{\rm coll}$, three mass parameters $B_{\beta\beta}$, $B_{\beta\gamma}$, $B_{\gamma\gamma}$, and three moments of inertia $\mathcal{I}_k$. These functions are determined using the cranking approximation formula based on the intrinsic triaxially deformed mean-field states. The diagonalization of the Hamiltonian (\[hamiltonian-quant\]) yields the excitation energies and collective wave functions that are used to calculate observables [@Nik.09].
The fact that, the 5DCH model using the collective inertia parameters calculated based on the cranking approximation can reproduce the structure of the experimental low-lying spectra [@Nik.09] up to an overall renormalization factor, demonstrates such approximation is fair enough for the present study. As it has been shown in Ref. [@LNRVYM12], this factor takes into account the contributions of the time-odd fields. A microscopic calculation of this factor would go far beyond the scope of the present investigation.
The intrinsic triaxially deformed mean-field states are the solutions of the Dirac (RMF+BCS) or RHB equations. The point-coupling energy functional PC-PK1 [@Zhao10] and the separable pairing force [@TMR.09a] are used in the particle-hole and particle-particle channels, respectively. In solving the Dirac and RHB equations, the Dirac spinors are expanded on the three-dimension harmonic oscillator basis with 14 major shells [@KR.89; @Peng08]. A quadratic constraint on the mass quadrupole moments is carried out to obtain the triaxially deformed mean-field states with $\beta\in[0.0, 0.8]$ and $\gamma\in[0^\circ, 60^\circ]$,and the step sizes $\Delta\beta=0.05$ and $\Delta\gamma=6^\circ$. More details about the calculations can be found in Refs. [@NRV.10; @Xiang12].
Figure \[fig1\] displays the comparison between the RHB and RMF+BCS calculations for the binding energy per nucleon $E/A$ \[plot (a)\], quadrupole deformation $\beta$ \[plot (b)\], neutron \[plot (c)\] and proton \[plot (d)\] average pairing gaps weighted by the occupation probabilities $v^2$ [@Bender00] of even-even Sm isotopes with $134\leqslant A\leqslant 154$. The binding energies and deformations found in the two calculations are close to each other. However, the average neutron and proton pairing gaps provided by the RHB calculations are generally larger than those by the RMF+BCS ones. This is consistent with the observations in Ref. [@Girod83], which indicates that the BCS ansatz gives slightly weaker pairing correlations with same pairing force. The underlying reason is well-known that the BCS ansatz corresponds to a special Bogoliubov transformation, which only considers pairing correlation between two nucleons in time-reversed conjugate states [@Ring80], and the off-diagonal matrix elements of the pairing field $\Delta$ are neglected in this approach.
In the following we have to consider that neglecting the off-diagonal matrix elements of the pairing field leads i) to a reduced configuration mixing and ii) as a consequence of self-consistency also to an overall reduction of the pairing strength in the diagonal matrix elements of the pairing field. Therefore it is interesting to address two points: i) whether the additional configuration mixing induced by the off-diagonal matrix elements of the pairing field is really essential and ii) whether the reduced strength of pairing caused by neglecting the off-diagonal matrix elements in the RMF+BCS approach can recovered simply by multiplying a strength factor $R_\tau$ to the diagonal pairing, i.e. whether the enhanced pairing strength is also able to reproduce the low-lying structure properties, e.g. the PESs, inertia parameters, as well as the low-lying spectra. Taking $^{152}$Sm as the example, Fig. \[fig2\] shows the neutron and proton average pairing gaps of the global minimum calculated by RMF+BCS as the functions of the pairing strength factor $R_\tau$, as referred to the horizontal lines denoting the RHB results with original pairing force. It is shown that the average pairing gaps increase almost linearly with respect to the pairing strength factor $R_\tau$ and cross the RHB results at $R_\tau\sim1.06$. Moreover, as shown in Fig. \[fig2\] (c) and (d) the RMF+BCS calculations with 6% enhanced pairing strength provide nearly identical average pairing gaps with the RHB results for the selected even-even Sm isotopes, with a relative deviation less than 5%.
As the further clarification, Fig. \[fig3\] displays the PESs, neutron and proton average pairing gaps, moments of inertia ${\cal I}_x$, collective masses $B_{\beta\beta}$ and $B_{\gamma\gamma}$ for $^{152}$Sm as functions of the quadrupole deformation parameter $\beta$, where the results are calculated by RHB with the original, and by RMF+BCS with the original and the enhanced (by 6%) pairing strength. It is well demonstrated that for the selected Sm isotopes the deviations on the low-lying structure properties described by RMF+BCS and RHB models can be eliminated by simply enhancing the pairing force about 6% in the BCS ansatz. Specifically, as the pairing strength increases, the average pairing gaps become larger, which leads to lower spherical barrier of PES [@Rutz99] and reduced inertia parameter [@Sobiczewski69].
In Fig. \[fig4\] we also compare the theoretical low-lying spectra of $^{152}$Sm calculated by RMF+BCS with the original and the enhanced (by 6%) pairing strength, to the RHB results. As seen from the left two panels,when the pairing strength is enhanced by 6%, the low-lying spectrum is extended, and systematically the intraband $B(E2)$ transitions become weaker, and the interband transitions are strengthened, finally leading to an identical prediction as the full RHB calculations (right panel). Quantitatively, the relative deviations between the RHB and RMF+BCS predications are reduced to less than 4% for the intraband transitions, and the main interband transitions agree with each other within $\sim 2$ W.u.. We have also checked the results for the other Sm isotopes, and very similar spectra are predicted by RHB with the original and RMF+BCS with enhanced (6%) pairing forces.
The similarity on the low-lying structure can be understood by analyzing the underlying shell structure predicted by the two mean-field calculations. Taking $^{152}$Sm as an example, in Fig. \[fig5\] we plot the single-particle configurations (energy and occupation probability) around the Fermi surface corresponding to the mean-field states of the global minimum in the PESs determined by the calculations of RHB with the original and RMF+BCS calculations with both original and enhanced (by 6%) pairing strength. Notice that the RHB results correspond to the canonical single-particle configurations, which are determined from the diagonalization of the density matrix [@Ring80]. Consistent with the agreement on the low-lying structure properties, the RMF+BCS calculations with the enhanced pairing strength also provide nearly identical single-particle configurations as the RHB ones.
In conclusion, we have taken Sm isotopes as examples to carry out a detailed comparison between the 5DCH calculations based on the RMF+BCS and the RHB approaches for the nuclear low-lying structure properties. It has been shown that the pairing correlations resulting from the RHB method are generally stronger than those from the RMF+BCS method with the same effective pairing force. However, by simply increasing the pairing strength by a factor 1.06 in the RMF+BCS calculations, the low-energy structure becomes very close to that of the full RHB calculations with the original pairing force. We have also carried out similar calculations in other regions of the nuclear chart and found that the necessary renormalization factor stays roughly constant up to heavy nuclei (1.06 in the Pu region) and increases slightly for light ones (1.10 in the Mg region).
This work was supported in part by the Major State 973 Program 2013CB834400, the NSFC under Grant Nos. 11335002, 11075066, 11175002, 11105110, and 11105111, the Research Fund for the Doctoral Program of Higher Education under Grant No. 20110001110087, the Natural Science Foundation of Chongqing cstc2011jjA0376, the Fundamental Research Funds for the Central Universities (XDJK2010B007, XDJK2011B002, and lzujbky-2012-k07), the Program for New Century Excellent Talents in University of China under Grant No. NCET-10-0466, and the DFG cluster of excellence Origin and Structure of the Universe (www.universe-cluster.de).
[99]{} J. Meng, I. Tanihata, and S. Yamaji, Phys. Lett. **B419**, 1 (1998).
G. Hagen, M. Hjorth-Jensen, G. R. Jansen, R. Machleidt, and T. Papenbrock, Phys. Rev. Lett. **109**, 032502 (2012).
R. Kshetri, M.S. Sarkar and S. Sarkar, Phys. Rev. C **74**, 034314 (2006) .
J. Meng, W. Zhang, S. G. Zhou, H. Toki and L. S. Geng, Eur. Phys. J. A **25**, 23 (2005). R. F. Casten, Nature Physics **2**, 811 (2006).
P. Cejnar, J. Jolie, and R. F. Casten, Rev. Mod. Phys. **82**, 2155 (2010).
K. Heyde and J. L. Wood, Rev. Mod. Phys. **83**, 1467 (2011).
A. Ozawa, T. Kobayashi, T. Suzuki, K. Yoshida, and I. Tanihata, Phys. Rev. Lett. **84**, 5493 (2000).
O. Sorlin and M.-G. Porquet, Prog. Part. Nucl. Phys. **61**, 602 (2008).
M. Bender, P.-H. Heenen, and P.-G. Reinhard, Rev. Mod. Phys. **75**, 121 (2003).
J. Meng, H. Toki, S. G. Zhou, S. Q. Zhang, W. H. Long, and L. S. Geng, Prog. Part. Nucl. Phys. **57**, 470 (2006).
D. Vretenar, A. V. Afanasjev, G. A. Lalazissis, and P. Ring, Phys. Rep. **409**, 101 (2005).
J. Meng, J. Peng, S. Q. Zhang, and P.W. Zhao, Frontiers of Physics **8**, 55 (2013).
J. Dobaczewski, J. Phys.: Conf. Ser. **312**, 092002 (2011).
T. Nikšić, D. Vretenar, and P. Ring, Phys. Rev. C **73**, 034308 (2006).
T. Nikšić, D. Vretenar, and P. Ring, Phys. Rev. C **74**, 064309 (2006).
M. Bender and P.-H. Heenen, Phys. Rev. C **78**, 024309 (2008).
J. M. Yao, J. Meng, D. P. Arteaga, and P. Ring, Chin. Phys. Lett. **25**, 3609 (2008).
J. M. Yao, J. Meng, P. Ring, and D. Pena Arteaga, Phys. Rev. C **79**, 044312 (2009).
J. M. Yao, J. Meng, P. Ring, and D. Vretenar, Phys. Rev. C **81**, 044311 (2010).
T. R. Rodr' iguez and J. L. Egido, Phys. Rev. C **81**, 064323 (2010).
D. Lacroix, T. Duguet, and M. Bender, Phys. Rev. C **79**, 044318 (2009).
L. Pr[ó]{}chniak and P. Ring, Int. J. Mod. Phys. E **13**, 217 (2004).
T. Nikšić, Z. P. Li, D. Vretenar, L. Pr[ó]{}chniak, J. Meng, and P. Ring, Phys. Rev. C **79**, 034303 (2009).
T. Nikšić, D. Vretenar, and P. Ring, Prog. Part. Nucl. Phys. **66**, 519 (2011).
Z. P. Li, T. Nikšić, D. Vretenar, P. Ring, and J. Meng, Phys. Rev. C 81, 064321 (2010). Z. P. Li, T. Nikšić, D. Vretenar, J. Meng, G. A. Lalazissis, and P. Ring, Phys. Rev. C **79**, 054301 (2009).
Z. P. Li, T. Nikšič, D. Vretenar, and J. Meng, Phys. Rev. C **80**, 061301 (2009).
Z. P. Li, J. M. Yao, D. Vretenar, T. Nikšič, H. Chen, and J. Meng, Phys. Rev. C **84**, 054304 (2011).
J.M. Yao, Z.P. Li, K. Hagino, M.Thi Win, Y. Zhang, J. Meng, Nucl. Phys. A **868**, 12 (2011).
H. Mei, J. Xiang, J. M. Yao, Z. P. Li, and J. Meng, Phys. Rev. C **85**, 034321 (2012).
J.-P. Delaroche, M. Girod, J. Libert, H. Goutte, S. Hilaire, S. Péru, N. Pillet, G. F. Bertsch, Phys. Rev. C **81**, 014303 (2010).
D. J. Dean and M. Hjorth-Jensen, Rev. Mod. Phys. **75**, 121(2003). Y. K. Gambhir, P. Ring, A. Thimet, Ann. Phys. (N.Y.) **198**, 132 (1990).
H. Kucharek and P. Ring, Z. Phys. A **339**, 23 (1991)
P. Ring, Prog. Part. Nucl. Phys. **37**, 193 (1996)
P. Ring and P. Schuck. [*The nuclear many-body problem*]{}. Springer-Verlag, New York, 1980.
M. Girod and B. Grammaticos Phys. Rev. C **27**, 2317(1983).
L. Próchniak, K. Zajac, K. Pomorski, S.G. Rohoziński, and J. Srebrny, Nucl. Phys. A **648**, 181 (1999).
Z. P. Li, T. Nikšić, P. Ring, D. Vretenar, J. M. Yao, and J. Meng, Phys. Rev. C **86**, 034334 (2012).
P. W. Zhao, Z. P. Li, J. M. Yao, and J. Meng, Phys. Rev. C **82**, 054319 (2010).
Y. Tian, Z. Y. Ma, and P. Ring, Phys. Lett. B **676**, 44 (2009).
W. Koepf and P. Ring, Nucl. Phys. A **493**, 61 (1989).
J. Peng, J. Meng, P. Ring, and S. Q. Zhang, Phys. Rev. C **78**, 024313 (2008).
T. Nikšić, P. Ring, D. Vretenar, Y. Tian, and Z. Y. Ma, Phys. Rev. C **81**, 054318 (2010).
J. Xiang, Z. P. Li, Z. X. Li, J. M. Yao, and J. Meng, Nucl. Phys. A **873**, 1 (2012).
M. Bender, K. Rutz, P.-G. Reinhard, J.A. Maruhn, Eur. Phys. J. A **8**, 59 (2000).
K. Rutz, M. Bender, P.-G. Reinhard, and J.A. Maruhn, Phys. Lett. B **468** 1 (1999).
A. Sobiczewski, Z. Szymański, S. Wycech, et al. Nucl. Phys. A **131** 67 (1969)
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The role played by the effective residual interaction in the transverse nuclear response for quasi–free electron scattering is discussed. The analysis is done by comparing different calculations performed in the Random–Phase Approximation and Ring Approximation frameworks. The importance of the exchange terms in this energy region is investigated and the changes on the nuclear responses due to the modification of the interaction are evaluated. The calculated quasi–elastic responses show clear indication of their sensibility to the details of the interaction and this imposes the necessity of a more careful study of the role of the different channels of the interaction in this excitation region.'
address:
- |
Departamento de Física, Facultad de Ciencias Exactas, Universidad Nacional de La Plata,\
La Plata, 1900, Argentina.
- 'Departamento de Física Moderna, Universidad de Granada, E–18071 Granada, Spain.'
author:
- 'Eduardo Bauer [@AAAuth]'
- 'Antonio M. Lallena'
title: 'ON THE ROLE OF THE EFFECTIVE INTERACTION IN QUASI–ELASTIC ELECTRON SCATTERING CALCULATIONS'
---
INTRODUCTION
============
An aspect of great importance in nuclear structure calculations at any excitation energy concerns with the role of the effective interaction. At low energies this problem has generated a considerable body of work in the last twenty years. On the contrary, this is a question not studied yet in deep in the literature for higher energies.
Giant resonances show an intricate mixture of multipolarities and the study of how the interaction affects it is a difficult task. In the quasi–elastic peak region, the problem of the longitudinal and transverse separation has occupied most of the investigations carried out till now and the discussion of the effects due to changes in the effective interaction have not been considered in detail.
As an example, we mention the considerable number of Random–Phase Approximation (RPA) type calculations performed in this energy region, much of them using residual interactions which include basically a zero–range term plus meson–exchange potentials corresponding to $\pi$, $\rho$ and, eventually, other mesons [@al84]–[@gi97]. An important point concerning the interaction refers to the values chosen for the parameters entering in the zero–range piece. However, and to the best of our knowledge, only in Ref. [@gi97] a certain discussion relative to the effects of varying these parameters can be found. In fact, the common practice is to pick an interaction from the literature, which usually corresponds to a parameterization fixed for low energy calculations, and afterwards use it to evaluate quasi–free observables sometimes without taking care of the effective theory in which the interaction was adjusted. It is obvious that, to a certain level, doubtful results are possible because of the known link between effective theory and interaction.
In this work we want to address this question and investigate if different parametrizations of the interaction can produce noticeable differences in the results and the extent to which the use of an interaction fixed for a given effective theory affects the results obtained within a different one. In Sec. II we give the details about the effective theories and interactions used to perform the calculations. In Sec. III we show and discuss the results we have obtained. Finally, we present our conclusions in Sec. IV.
EFFECTIVE THEORIES AND INTERACTIONS
===================================
The first interaction we consider in this work is the so–called Jülich–Stony Brook interaction [@sp80] which is an effective force widely used for calculations in the quasi–elastic peak. It is given as follows: $$\label{int-I}
\displaystyle
V^{\rm I}_{\rm res} \, = \, V_{\rm LM} \, + \, V_\pi \, + \,
\tilde{V}_\rho \, .$$ Here $V_{\rm LM}$ is a zero–range force of Landau–Migdal type, which takes care of the short–range piece of the NN interaction: $$V_{\rm LM} \, = \,C_0 \,[ g_0\,
{\mbox{\boldmath $\sigma$}}^1\cdot {\mbox{\boldmath $\sigma$}}^2
+\, g_0^\prime\,
{\mbox{\boldmath $\sigma$}}^1\cdot {\mbox{\boldmath $\sigma$}}^2
{\mbox{\boldmath $\tau$}}^1 \cdot {\mbox{\boldmath $\tau$}}^2]
\, .$$ On the other hand, a finite–range component generated by the ($\pi$+ $\rho$)–meson exchange potentials is also included. The tilde in $\tilde{V}_\rho$ means that the bare $\rho$–exchange potential is slightly modified in order to take into account the effect of the exchange of more massive mesons. In particular, a factor $r=0.4$ is multiplying the finite–range non–tensor piece of the potential (see Ref. [@sp80] for details). This force was fitted to reproduce low energy magnetic properties in the lead region (specifically, magnetic resonances in $^{208}$Pb and magnetic moments and transition probabilities in the neighboring nuclei). The calculations were performed in the framework of the RPA and Woods–Saxon single–particle wave functions were used in the configuration space. The values $g_0=0.57$ and $g_0^\prime=0.717$ (with $C_0=386.04$ MeV fm$^3$) were found to be adequate to describe the properties studied.
As previously stated, this interaction has been considered in different calculations in the quasi–elastic peak (see e.g. [@gi97]). The problem arise because some of them have been done within the Fermi gas (FG) formalism, with local density approximation to describe finite nuclei, in the ring approximation (RA), where the exchange terms are not taken into account, and with the full unmodified $\rho-$exchange potential. Under these circumstances, the possible effects in the nuclear responses due to the inconsistency between the model and the effective interaction could be non–negligible. This is precisely the first aspect we want to investigate. To do that we compare the responses obtained with the Jülich–Stony Brook interaction with those calculated with a second effective force of the form: $$\label{int-II}
\displaystyle
V^{\rm II}_{\rm res} \, = \, V_{\rm LM} \, + \, V_\pi \, + \,
V_\rho \, ,$$ by considering the same values for the zero–range parameters in both cases. The force in Eq. (\[int-II\]) only differs from $V^{\rm I}_{\rm res}$ in the $\rho$-potential which, in this case, does not include any reduction factor. Both RPA and RA effective theories are used to analyze the results.
A second question of interest to us is to determine how the change of the zero–range parameters affects the responses calculated within a given theory. This will inform us about the necessity of considering or not in detail the role of these parameters. This aspect is analyzed by considering $V^{\rm II}_{\rm res}$ with parameters $g_0$ and $g_0^\prime$ fixed, as in the case of the Jülich–Stony Brook interaction, to reproduce some low energy properties in the lead region (see details in the next section). It is worth to point out that $V^{\rm II}_{\rm res}$ is precisely the interaction used in practice in much of the calculations mentioned above and that is why we want to use it for this analysis.
Our analysis focuses on the transverse response functions in the quasi–elastic peak. We will not consider the longitudinal ones because they are strongly influenced by the spin independent pieces of the interaction (in particular, the $f_0$ and $f_0^\prime$ channels) and these are difficult to fix at low energy because of the role played by the scalar mesons not usually taken into account.
RESULTS OF THE CALCULATIONS
===========================
The investigation of the various questions we are interested in has been carried out by comparing different calculations of the transverse (e,e’) responses in $^{40}$Ca for three different momentum transfer ($q=300$, 410 and 550 MeV/$c$).
(100,180)(35,0) (0,0)
[. Transverse nuclear responses for $^{40}$Ca, calculated for the three momentum transfers we have considered in this work. Dotted lines correspond to an RPA calculation with $V^{\rm I}_{\rm res}$, while solid curves represent the RA results for $V^{\rm II}_{\rm res}$. In both cases the values $g_0=0.57$ and $g_0^\prime=0.717$ (with $C_0=386.04$ MeV fm$^3$) have been used. Dashed curves give the free FG responses. In all the calculations a value of $k_{\rm F}=235$ MeV/$c$ has been used.]{}
First we have study the effects produced when an effective interaction, which has been determined in a given effective theory (e.g. RPA), is used to calculate (e,e’) transverse responses in a different framework (e.g. RA).
By considering the parameterization of Ref. [@sp80] (that is $g_0=0.57$, $g_0^\prime=0.717$ and $C_0=386.04$ MeV fm$^3$), we have carried out two different calculations, the results of which are shown Fig. 1. Therein, solid curves correspond to the calculations performed in the FG approach within the RA and with the interaction $V^{\rm II}_{\rm res}$ in Eq. (\[int-II\]). On the other hand, dotted lines have been obtained within the RPA, also for the FG. The model used in this case is the one developed in Ref. [@ba96], which, contrary to what happens for the RA approach, includes explicitly the exchange terms in the RPA expansion. In this case we have used the interaction $V^{\rm I}_{\rm res}$ in Eq. (\[int-I\]) and we have adopted the factor $r=0.4$, which is consistent with the parameterization used. Also in Fig. 1, we have plotted the free FG responses for comparison (dashed curves).
The first comment one can draw from these results is that the use of the interaction, as it was fixed at low energy, leads to transverse responses which are quite different from those obtained in the RA (with $V^{\rm II}_{\rm res}$) calculation, though the differences reduce with increasing momentum transfer. As we can see, the results obtained in the RPA are peaked at lower energies and this is a clear evidence of a more attractive residual interaction. It is straightforward to check this point because the central piece of the $V^{\rm I}_{\rm res}$ is attractive, while the contrary happens for $V^{\rm II}_{\rm res}$, at least for $q\leq 2k_{\rm F}$. On the other hand it is interesting to note how the RA results are more similar to the free response as long as $q$ increases, while the same does not occurs for the RPA responses.
Obviously, the reason for the discrepancies between both calculations can be ascribed to the two basic ingredients of the effective theories used in each case: the exchange terms, which are included in the RPA calculations but not in the RA ones, and the reduction factor $r$ modifying the $\rho-$exchange potential.
Before going deeper in this question, it is worth to comment on the nuclear wave functions used in the calculations discussed above. As in any FG type calculation, plane–waves have been considered here to describe the single–particle states. The fact that the interaction was fixed in a framework which considered microscopic RPA wave functions, based on Woods–Saxon single–particle states, is an obvious inconsistency. Despite that, it has been shown [@am94; @am96] that, in this energy region, the details concerning the nuclear wave functions are not extremely important and, at least to some extent, the shell–model response can be reasonably described with the FG model, provided an adequate value of the Fermi momentum, $k_{\rm F}$, is used. In the present work, where we study the response in $^{40}$Ca, we have taken $k_{\rm F}=235$ MeV/$c$ which gives a good agreement between FG and finite nuclei calculations [@am94].
We come back to investigate the reasons for the large discrepancy between the RPA and RA calculations presented above. To do that we have done two new calculations: RA with $V^{\rm I}_{\rm res}$ and RPA with $V^{\rm II}_{\rm res}$. These calculations have been compared with the two previous ones by means of the two following quantities: $$\label{relexc}
\displaystyle
\gamma_{\rm exc}^r(q,\omega) \, = \,
\displaystyle
\frac{R_T^{{\rm RPA}(r)}(q,\omega) \, - \,
R_T^{{\rm RA}(r)}(q,\omega) }
{R_T^{{\rm RA}(r)}(q,\omega)}$$ and $$\label{relfac}
\displaystyle
\gamma_r^{\rm mod}(q,\omega) \, = \,
\displaystyle
\frac{R_T^{{\rm mod}(r=1.0)}(q,\omega) \, - \,
R_T^{{\rm mod}(r=0.4)}(q,\omega) }
{R_T^{{\rm mod}(r=0.4)}(q,\omega) } \, .$$ The first one gives us information about the effect of the consideration of the exchange terms in the calculation. The corresponding results have been plotted in Fig. 2 (left panels). The first aspect to be noted is that the exchange terms produce effects considerably larger for $V^{\rm I}_{\rm res}$ (solid lines) than for $V^{\rm II}_{\rm res}$ (dashed curves). These effects reduce with increasing momentum transfer and they are rather small for $V^{\rm II}_{\rm res}$ above $q=410$ MeV/$c$.
On the other hand, the effect of the reduction factor $r$ in the $\rho$–exchange potential is measured with the parameter $\gamma_r$. The values of this parameter for the two effective models considered, these are RPA and RA, are shown in Fig. 2 (right panels), with solid and dashed curves respectively. It is apparent that the effects of considering the $r$ factor are much larger than those due to the exchange. In general they are more important for the RA calculations than for the RPA ones, and reduce the higher $q$ is.
(100,180)(30,0) (0,0)
[. Left panels: $\gamma_{\rm exc}$, in percentage, as defined in Eq. (\[relexc\]). Dashed (solid) curves give the results obtained for $r=1.0 (0.4)$. Right panels: $\gamma_r$, in %, calculated as in Eq. (\[relfac\]), for RPA (solid curves) and RA (dashed curves).]{}
The first conclusion to be noted is that when using a given interaction is mandatory to take care of the effective theory where its parameterization was fixed. The change of the framework produces results which could not be under control.
The open question in this respect is how different becomes the responses calculated within different effective theories but with an interaction fixed consistently with the theory. This is the second aspect we investigate. To do that we have considered the $V^{\rm II}_{\rm res}$ and have determined the parameters $g_0$ and $g_0^\prime$ of the Landau–Migdal piece in such a way that the energies and B-values of the two $1^+$ states in $^{208}$Pb at 5.85 and 7.30 MeV are reproduced. This has been done both in RPA and RA. The reason for choosing these two states lies in their respective isoscalar and isovector character, what makes them particularly adequate to permit the determination of both parameters almost independently. The values obtained in this procedure are shown in Table I. It is remarkable the small value of $g_0$ needed for the RA calculation. A similar result is found when a pure zero–range Landau–Migdal interaction is adjusted, with the same criterion, in RPA type calculations (see Refs. [@co90; @hi92]). This points out the importance of the exchange, at least at low energy.
[ Values of the Landau–Migdal parameters $g_0$ and $g_0^\prime$ obtained in the procedure of fixing the effective interaction $V^{\rm II}_{\rm res}$ (see text). The values quoted “RPA” (“RA”) correspond to calculations performed with (without) the consideration of the exchange terms.]{}
Effective theory $g_0$ $g_0^\prime$
------------------ -- ---------- --------------
RPA $~0.470$ $~0.760$
RA $~0.038$ $~0.717$
With the interaction fixed in this way we have evaluated the transverse responses for the three momentum transfer we are considering throughout this work. The results are shown in Fig. 3 where dotted (solid) curves correspond to the RPA (RA) calculations. Dashed lines represent the free FG responses. As we can see, the differences between the results obtained with the two effective theories are now much smaller than in Fig. 1.
(100,180)(35,0) (0,0)
[. Transverse nuclear responses for $^{40}$Ca, calculated with the $V^{\rm II}_{\rm res}$ interaction. Dotted lines correspond to an RPA calculation while solid curves represent the RA results. The values of $g_0$ and $g_0^\prime$ in Table I have been used. Dashed curves give the free FG responses. In all the calculations a value of $k_{\rm F}=235$ MeV/$c$ has been used.]{}
Two points deserve a comment. First, it is clear that the large differences observed between the RPA calculation here discussed and that shown in Fig. 1 are mainly due to the presence of the reduction factor $r=0.4$ in the $V^{\rm I}_{\rm res}$ interaction. Second, the similitude of the results obtained with the two calculations done now with $V^{\rm II}_{\rm res}$, shows up the relevance of the link between effective theories and interactions.
The last aspect we want to analyze is how the responses calculated in a given approach change when the zero–range parameters are modified. In other words, we want to determine what is the role of these parameters. How $g_0^\prime$ affects the responses is a point which has been investigated with a certain detail in different previous works (see e.g., Ref. [@ba96]) and then we focus here in $g_0$. Its influence can be seen in Fig. 4, where we compare the responses plotted in Fig. 3 (solid curves), with those obtained by changing the $g_0$ parameter in order to use values considered by different authors. Dashed–dotted curves correspond to $g_0=0$. Dashed lines represent the responses obtained with $g_0=0.70$ (0.57) for the RPA (RA) calculation. The values of $g_0^\prime$ have not been changed. The first point to be noted is the insensibility of the RA responses to the changes in $g_0$. As we can see, strong changes in $g_0$ produce almost no effect on the RA result. This can be easily understood because in the ring series the $g_0$ contribution is weighted with the magnetic moment $\mu_s^2$ while the $g_0^\prime$ piece appears with $\mu_v^2$. That means, the $g_0$ contribution is $\mu_s^2/\mu_v^2 \, \approx \, 1/28$ of the $g_0^\prime$ contribution. The situation is different in the RPA case, where the $g_0$ contribution is as important as the $g_0^\prime$ one because of the presence of the exchange terms (see Ref. [@ba96]). This makes that some of the RA calculations performed by other authors can be considered as “consistent” in practice, of course despite the fact that these parametrizations are unable to reproduce low energy properties. For example, in Ref. [@gi97], the parameterization of the Jülich–Stony Brook interaction was considered and this coincides with one of those used here ($g_0=0.57$ and $g_0^\prime=0.717$).
The results obtained in this work open a series of questions which we consider worth for nuclear calculations in this energy region. In the following we enumerate and comment them:
(100,170)(24,0) (0,0)
[. $R_T$ responses calculated in the RPA (left panels) and RA (right panels). Solid curves correspond to the parametrizations of Table I. Dashed–dotted curves have been obtained with $g_0=0$, while the dashed ones correspond to $g_0=0.70$ (0.57) for the RPA (RA) calculation, with the same values of $g_0^\prime$ as for the solid curves.]{}
1. It has been shown that the strength of the tensor piece of $V^{\rm I}_{\rm res}$ is too strong to describe low energy properties (see, e.g., Ref. [@co90]) and different mechanisms have been proposed to cure this problem (core–polarization effects [@na85], two–particles two–holes excitations [@dr86], [*in–medium*]{} scaling law [@br91], etc.) The role of the tensor part of the interaction in the quasi-elastic peak should be investigated in order to establish the effective force to be used.
2. The presence of the exchange terms increase the sensitivity of the responses to the details of the interaction. How important can be the interference between these terms and other physical mechanisms basic in this energy region (such as, e.g., meson–exchange currents, final state interactions, short–range correlations, etc.) is a matter of relevance in order to fully understand the nuclear response. The analysis of the possible differences between RA and RPA with respect to these effects is of special interest in view of the fact that RA calculations are the most usual in the quasi–elastic peak.
3. The procedure of fixing the interaction is basic in order to deal with the possibility of having an unique framework to calculate the nuclear response at any momentum transfer and excitation energy. The problem of developing such “unified” model is still unsolved, but the cross analysis of low energy nuclear properties and quasi–elastic peak responses could give valuable hints.
CONCLUSIONS
===========
In this work we have analyzed the role of the effective interaction in the quasi–elastic peak region by comparing the results obtained with different effective theories and forces previously fixed in order to give a reasonable description of several low energy nuclear properties.
Some conclusions can be drawn after our analysis. First, it has been found that the interaction plays a role that, similarly to what happens at low excitation energy, cannot be neglected. The particular point to be noted is the necessity of using effective interactions which have been fixed within an effective theory.
Second, the procedure we have followed to perform the calculations, that is to determine the interaction at low energy before calculating at the quasi–elastic peak, seems to be adequate to look for an “unique” framework to calculate the nuclear response in different energy and momentum regimes.
The role of the tensor piece of the interaction must be investigated. At low energy is a basic ingredient of the nuclear structure calculations. Thus it is important to disentangle its contribution in other excitation energy regions. Additionally, it seems encouraging to analyze the problem by including other physical mechanisms (meson–exchange currents, short–range correlations, final state interactions, etc.) which are known to be important in the description of the nuclear response and which depend on the interaction.
Discussions with G. Co’ are kindly acknowledged. This work has been supported in part by the DGES (Spain) under contract PB95-1204 and by the Junta de Andalucía (Spain).
Fellow of the Consejo Nacional de Investigaciones Científicas y Técnicas, CONICET.
W.M. Alberico, M. Ericson and A. Molinari, Ann. Phys. (N.Y.) [**154**]{}, 356 (1984); W.M. Alberico, A. De Pace, A. Drago and A. Molinari, Rivista del Nuovo Cimento [**14**]{}, 1 (1991).
J. Ryckebusch, M. Warroquier, K. Heyde, J. Moreau and D. Ryckbosch, Nucl. Phys. [**A476**]{}, 237 (1988).
A. De Pace and M. Viviani, Phys. Rev. C [**48**]{}, 2931 (1993).
M. Buballa, S. Drożdż, S. Krewald and J. Speth, Ann. Phys. (N.Y.) [**208**]{}, 346 (1991); M. Buballa, S. Drożdż, S. Krewald and A. Szczurek, Phys. Rev. C [**44**]{}, 810 (1991); S. Jeschonnek, A. Szczurek, G. Co’ and S. Krewald, Nucl. Phys. [**A570**]{}, 599 (1994).
E. Bauer, Nucl. Phys. [**A589**]{}, 669 (1995).
V. Van der Sluys, J. Ryckebusch and M. Warroquier, Phys. Rev. C [**51**]{}, 2664 (1995).
E. Bauer, A. Ramos and A. Polls, Phys. Rev. [**C54**]{}, 2959 (1996).
M.B. Barbaro, A. De Pace, T.W. Donnelly and A. Molinari, Nucl. Phys. [**A596**]{}, 553 (1996); [*ibid.*]{} [**A598**]{}, 503 (1996).
A. Gil, J. Nieves and E. Oset, Nucl. Phys. [**A**]{} (1997) (in press); A. Gil, Ph. D. Thesis, Universitat de València, 1996.
J. Speth, V. Klemt, J. Wambach and G. E. Brown, Nucl. Phys. [**A343**]{}, 382 (1980).
J.E. Amaro, A.M. Lallena and G. Co’, Int. J. Mod. Phys. [**E3**]{}, 735 (1994).
J.E. Amaro [*et al.*]{}, Nucl. Phys. [**A602**]{}, 263 (1996).
G. Co’ and A.M. Lallena, Nucl. Phys. [**A510**]{}, 139 (1990).
N.M. Hintz, A.M. Lallena and A. Sethi, Pys. Rev. C [**45**]{}, 1098 (1992).
K. Nakayama, Phys. Lett. [**B165**]{}, 239 (1985).
S. Drożdż, J.L Taín and J. Wambach, Phys. Rev. C [**34**]{}, 345 (1986).
G.E Brown and M. Rho, Phys. Rev. Lett. [**66**]{}, 2720 (1991).
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'In this paper we establish the existence in low dimensions of solutions to the constraint equations in the case of the conformal system recently proposed by David Maxwell [@Max14], with the added presence of a scalar field and under suitable smallness assumptions on its parameters.'
author:
- 'Caterina Vâlcu[^1]'
title: 'The constraint equations in the presence of a scalar field - the case of the conformal method with volumetric drift'
---
Introduction
============
The field of general relativity deals with the study of spacetime, an object defined as the equivalence class, up to an isometry, of Lorentzian manifolds $(\tilde{M},\tilde{g})$ of dimension $n+1$ satisfying the Einstein field equations $$Ric_{\alpha\beta}(\tilde{g})-\frac{1}{2}R(\tilde{g})\tilde{g}_{\alpha\beta}=8\pi T_{\alpha\beta},$$ $\alpha,\beta=\overline{1,n+1}$. Here, $R(\tilde{g})$ is the scalar curvature of $\tilde{g}$, $Ric$ the Ricci curvature and $T_{\alpha\beta}$ the stress-energy tensor describing the presence of matter and energy. For example, $T_{\alpha\beta}=0$ describes vacuum. Our interest focuses on the more general case $$T_{\alpha\beta}=\nabla_\alpha\tilde\psi\nabla_\beta\tilde\psi-\left(\frac{1}{2}|\nabla\tilde\psi|^2_{\tilde{g}}+V(\tilde\psi)\right)\tilde{g}_{\alpha\beta},$$ which models the existence within the spacetime of a scalar field $\tilde\psi\in\mathcal{C}^\infty(M)$ having potential $V\in\mathcal{C}^\infty({\mathbb{R}})$. Thus, $\tilde\psi=0$ and $V=\Lambda$ yield the vacuum with cosmological constant $\Lambda$, while $V=\frac{1}{2}m\tilde\psi^2$ corresponds to the Einstein-Klein-Gordon setting.
For a globally hyperbolic spacetime, we define its initial data $(M,\hat{g},\hat{K},\hat{\psi},\hat{\pi})$. They consist of an $n$-dimensional Riemannian manifold $(M,\hat{g})$, which models the spacetime at a particular moment in time, a symmetric 2-tensor $\hat{K}$, corresponding to its second fundamental form, the scalar field $\hat{\psi}$ in $M$, and its temporal derivative $\hat{\pi}$. The associated spacetime development takes the form $(M\times {\mathbb{R}},\tilde{g}, \tilde{\psi})$, where $\tilde{g}$ is a Lorentzian metric that verifies $\tilde{g}|_{M}=\hat{g}$ and $\tilde\psi$ is a scalar field such that $\tilde{\psi}|_{M}=\hat{\psi}$ and $\partial_t\tilde{\psi}|_{M}=\hat{\pi}$.
Initial data in general relativity may not be freely specified, unlike their Newtonian counterparts. Instead, they must verify the Gauss and Codazzi equations, $$\begin{cases}
\displaystyle R(\hat{g})+(tr_{\hat{g}}\hat{K})^2-||\hat{K}||^2_{\hat{g}} &\displaystyle= \hat{\pi}^2+|\hat{\nabla}\hat{\psi}|^2_g+2V(\hat{\psi})\\
\partial_i(tr_{\hat{g}} \hat{K})-\hat{K}_{i,j}^j&\displaystyle= \hat{\pi}\partial_i\hat{\psi},
\end{cases}$$ which are referred to as the *constraint equations*. The work of Choquet-Bruhat [@Cho] establishes, once and for all, that the constraint equations are not only necessary but sufficient conditions for the (local) existence of a solution. Later, Choquet-Bruhat and Geroch [@ChoGer] prove that the maximal development of initial data is unique, up to an isometry. Globally hyperbolic spacetimes may rigorously be studied in the context of mathematical analysis as the result of an evolution problem. The above system is clearly under-determined, which allows for considerable freedom in choosing a solution $(\hat{g},\hat{K},\hat{\psi},\hat{\pi})$.
Using the conformal method introduced by Lichnerowicz [@Lic], the constraint equations may be transformed into a determined system of equations by fixing well-chosen quantities (see Choquet-Bruhat, Isenberg and Pollack [@ChoIsePol]). The appeal of such a method lies in that it provides a characterisation of the resulting initial data by fixed quantities. Essentially, it maps a space of parameters to the space of solutions.
Given an initial data set $(\hat{g},\hat{K},\hat{\psi},\hat{\pi})$, the classical choice of parameters is $(\mathbf{g},\mathbf{U},\tau,\psi,\pi;\alpha)$: in this case, the conformal class $\mathbf{g}$ is represented by a Riemannian metric $g$, the smooth function $\tau=\hat{g}^{ab}\hat{K}_{ab}$ is a mean curvature and the conformal momentum $\mathbf{U}$ measured by a volume form $\alpha$ (volume gauge) is a 2-tensor that is both trace-free and divergence-free with respect to $g$ (a transverse-traceless tensor). We sometimes prefer to indicate the volume gauge by the densitized lapse $$\tilde{N}_{g,\alpha}:=\frac{\alpha}{dV_g}.$$ Note that this quantity depends on the choice of representative $g$, unlike the volume gauge $\alpha$ which does not. The standard conformal method implicitly fixes $\tilde{N}_{g,\alpha}=2$; in the present paper, we prefer to make use of the freedom of choosing $\tilde{N}_{g,\alpha}$ as needed. We often refer to a parameter set by indicating the representative metric $g$ and the corresponding densitized lapse $\tilde{N}_{g,\alpha}$ instead of giving the conformal class and volume gauge. However, these quantities can immediately be reconstructed from our data. We refer to Maxwell [@Max14] for an introduction to the conformal method in our context.
Starting from the parameter set $(g,U,\tau,\psi,\pi;\tilde{N})$, the corresponding (physical) initial data is pinpointed by solving a resulting system, comprising the Lichnerowicz-type equation and the momentum constraints, for a smooth positive function (or conformal factor) $u$ and a smooth vector field $W$ in $M$, $$\begin{cases}\label{syst classical conformal method}
\displaystyle \Delta_g u+\mathcal{R}_\psi=&\displaystyle-\mathcal{B}_{\tau,\psi,V}u^{q-1}+\frac{\mathcal{A}_{\pi,U}(W)}{u^{q+1}},\\
\Delta_{g, conf}W=&\displaystyle \frac{n-1}{n}u^{q}\nabla\tau+\pi\nabla\psi,
\end{cases}$$ where $$\begin{array}{l}
\displaystyle \mathcal{R}_\psi=\frac{n-2}{4(n-1)} \left(R(g)-|\nabla\psi|^2_g\right),\\
\displaystyle \mathcal{B}_{\tau,\psi,V}=\frac{n-2}{4(n-1)}\left(\frac{n-1}{n}\tau^2-2V(\psi)\right),\\
\displaystyle \mathcal{A}_{\pi,U}(W)=\frac{n-2}{4(n-1)}\left(|U+\mathcal{L}_gW|^2_g+\pi^2\right).
\end{array}$$ If $(u,W)$ solves the above system, then the initial data we’ve been searching for are $$\hat{g}=u^{q-2}g,\quad \hat{K}=u^{-2}\left(U+\frac{\tilde{N}}{2}\mathcal{L}_g W\right)+\frac{\tau}{n}\hat{g},\quad \hat{\psi}=\psi,\quad\hat{\pi}=u^{-q}\pi.$$ Note also that the solutions generated by $(g,U,\tau;\tilde{N})$ and $$(\varphi^{q-2}g,\varphi^{-2}U,\tau,\psi,\varphi^{-q}\pi; \varphi^q \tilde{N})$$ are the same, where $\varphi$ is a smooth positive function. The notations above are similar to those of Choquet-Bruhat, Isenberg and Pollack ([@ChoIsePol], [@ChoIsePol2]). The following quantities often appear throughout the present paper: $q=\frac{2n}{n-2}$ is the critical Sobolev exponent for the embedding of $H^1$ in Lebesgue spaces, $\Delta_g=-div_g\nabla$ denotes the Laplace-Beltrami operator taken with non-negative eigenvalues, $\Delta_{g,conf}W=div_g(\mathcal{L}_g W)$ is the Lamé operator and $\mathcal{L}_g$ is the conformal Killing operator with respect to $g$, $$\mathcal{L}_g W_{ij}=W_{i,j}+W_{j,i}-\frac{2}{n}div_g W g_{ij}.$$ Conformal Killing fields are defined as vector fields in the kernel of $\mathcal{L}_g.$
The conformal method is particularly successful in finding solutions when the mean curvature $\tau$ is constant as the system (\[syst classical conformal method\]) becomes uncoupled, but it is unclear how well the method functions when the mean curvature is far from being constant: see Maxwell [@Max11] and [@Max14c], where a given set of parameters point to no or to an infinite number of solutions. We emphasize that any failing of the system does not necessarily translate to a singularity in the space of solutions to the constraints system, but may instead derive from a poor choice of mapping. This motivates the study of variations to standard conformal methods.
The drift method introduced by Maxwell replaces the mean curvature $\tau$ as a parameter by a pair $(\tau^*,\tilde{V})$, where $\tau^*$ is a unique constant called volumetric momentum and $\tilde{V}$ a vector field related to the drift. They verify an analogue of York splitting, namely $$\label{volumetric momentum}
\tau=\tau^*+\tilde{N}_{\hat{g},\alpha}div_{\hat{g}}\tilde{V}=\tau^*+\frac{\tilde{N}_{g,\alpha}}{u^{2q}}div_g(u^q\tilde{V}),$$ the notation $\tilde{V}$ being specific to this paper in order to avoid confusion with the potential $V$. Interestingly, $\tau^*=0$ holds true for all counterexamples found by Maxwell ([@Max11], [@Max14c]). This suggests that the volumetric momentum may play an important role in characterizing the space of initial data. Ideally, we would like to know as soon as we fix a set of parameters $(g,U,\tau,\psi,\pi;\tilde{N})$ if we find ourselves in the case $\tau^*=0$. However, $\tau^*$ cannot be directly calculated by (\[volumetric momentum\]) without first solving (\[syst classical conformal method\]), which somewhat defeats the purpose. This motivates a new choice of parameters, even at the risk of working with an analytically more complicated system. The idea of Maxwell ([@Max11], [@Max14c]) is thus to choose $\tau^*$ as an additional parameter that is to be fixed in the place of $\tau$: the hope is thus to avoid the aforementioned problem. As well as $\tau^*$, Maxwell added $\tilde{N}$ and $\tilde{V}$ as parameters, for geometric and physical reasons. Therefore, instead of fixing $\tau$, we fix $\tau^*$, $\tilde{N}$ and $\tilde{V}$.
Intuitively, the drift is a geometric quantity describing infinitesimal motion in the space of metrics modulo the group of diffeomorphisms connected to the identity such that the conformal class and volume are preserved. For any given drift, the choice of a representative vector field $\tilde{V}$ is unique up to conformal Killing fields and vector fields which are divergence-free with respect to the initial metric $\hat{g}$. Given $g$ an arbitrary representative of the conformal class, it is not clear whether two vector fields are indicative of the same drift class defined for $\hat{g}$; this problem is discussed at length in the paper of Mike Holst, David Maxwell and Rafe Mazzeo [@HolMaxMaz] for conformal systems where the critical non-linearity is non-focusing, or negative. Our analysis treats systems with focusing (that is to say positive) non-linearities stemming from the presence of a scalar field with positive potential.
The following system corresponds to Problem 12.1 of [@Max14] in the presence of a scalar field, where $g$ admits no non-trivial conformal Killing fields: $$\label{syst of Maxwell}
\begin{cases}
\Delta_g u+\frac{n-2}{4(n-1)}(R(g)-|\nabla\psi|_g^2)u -\frac{(n-2)|U+\mathcal{L}_g W|^2+\pi^2}{4(n-1)u^{q+1}}\\
\displaystyle\quad-\frac{n-2}{4(n-1)}\left[2V(\psi)-\frac{n-1}{n}\left(\tau^*+\frac{\tilde{N}div_g(u^q\tilde{V})}{u^{2q}}\right)^2\right]u^{q-1}=0\\ \\
div_g\left(\frac{\tilde{N}}{2}\mathcal{L}_g W\right)-\frac{n-1}{n}u^q \mathbf{d} \left(\frac{\tilde{N}div_g(u^{q}\tilde{V})}{2u^{2q}}\right)-\pi\nabla\psi=0.
\end{cases}$$ We denote the exterior derivative by $\mathbf{d}$. The unknowns are a smooth positive scalar function $u$ defined on $M$ and a smooth vector field $W$ on $M$. The parameters are $(g,U,\tau^*,\tilde{V},\psi,\pi;\tilde{N})$. Maxwell’s new set of parameters include $\tau^*$, which could not be calculated *a priori* in the classical method. The initial data of the constraint equations verify $$\begin{array}{c}
\hat{g}=u^{q-2}g,\quad \hat{K}=u^{-2}\left(U+\frac{\tilde{N}}{2}\mathcal{L}_g W\right)+\frac{1}{n}\left(\tau^*+\frac{\tilde{N}}{u^{2q}}div(u^q\tilde{V})\right)\hat{g},\\ \hat{\psi}=\psi,\quad\hat{\pi}=u^{-q}\pi.
\end{array}$$
The following is a more general system than (\[syst of Maxwell\]). The central result of the paper consists in showing that it admits solutions. Let $(M,g)$ be a closed Riemannian manifold of dimension $n\in\{3,4,5\}$, and $g$ has no non-trivial conformal Killing fields. Let $b$, $c$, $d$, $f$, $h$, $\rho_1$, $\rho_2$, $\rho_3$ be smooth functions on $M$ and let $Y$ and $\Psi$ be smooth vector fields defined on $M$. Let $0<\gamma<1$. Assume that $\Delta_g+h$ is coercive, in the sense that its first eigenvalue is positive. Assume that $f>0$, $\rho_1>0$ and $|\nabla \rho_3|<(2C_1)^{-1}$, where $C_1$ is a dimensional constant - see (\[lxest\]). Consider the system $$\label{systemshort}
\begin{cases}
\displaystyle \Delta_g u+hu
&\displaystyle= fu^{q-1}+\frac{\rho_1+|\Psi+\rho_2\mathcal{L}_gW|^2_g}{u^{q+1}}\\
&\displaystyle -\frac{b}{u}-c\langle\nabla u,Y\rangle \left(\frac{d}{u^2}+\frac{1}{u^{q+2}}\right)-\frac{\langle\nabla u,Y\rangle^2}{u^{q+3}}\\
\displaystyle div_g\left(\rho_3\mathcal{L}_g W\right)
&\displaystyle =\mathcal{R}(u).
\end{cases}$$ Here $\mathcal{R}$ is an operator verifying $$\mathcal{R}(u)\leq C_\mathcal{R}\left(1+\frac{||u||_{\mathcal{C}^2}^2}{(\inf_Mu)^2}\right)$$ for a constant $C_\mathcal{R}>0$.
A supersolution of the Lichnerowicz-type equation is a smooth function $u$ verifying that $$\begin{array}{r l}
\Delta_g u+hu\geq &fu^{q-1}+\frac{\rho_1+|\Psi+\rho_2\mathcal{L}_gW|^2_g}{u^{q+1}} -\frac{b}{u}\\&
-c\langle\nabla u,Y\rangle \left(\frac{d}{u^2}+\frac{1}{u^{q+2}}\right)-\frac{\langle\nabla u,Y\rangle^2}{u^{q+3}}.
\end{array}$$ Similarly, a subsolution satisfies an inequality of opposite sign. Whenever the inequality is strict, we say $u$ is a strict subsolution or a strict supersolution respectively.
We fix $$\label{th main}
\theta=\min(\inf_M\rho_1,\inf_M f),$$ and $$\label{T main}
T=\max(||f||_{\mathcal{C}^{1,\gamma}}, ||\rho_1||_{\mathcal{C}^{0,\gamma}}, ||c||_{\mathcal{C}^{0,\gamma}}, ||d||_{\mathcal{C}^{0,\gamma}}, ||h||_{\mathcal{C}^{0,\gamma}}).$$
Here is the main result of our paper:
\[thm 1\] There exists a constant $C=C(n,h)$, $C>0$ such that if $\rho_1$ verifies $$||\rho_1||_{L^1(M)}\leq C(n,h)\left(\max_M |f|\right)^{1-n},$$ and there exists a constant $$\delta=\delta(\theta, T )>0$$ such that, if $$||b||_{\mathcal{C}^{0,\gamma}}+ ||Y||_{\mathcal{C}^{0,\gamma}}+ ||\Psi||_{\mathcal{C}^{0,\gamma}}+||\rho_2||_{\mathcal{C}^{0,\gamma}}+ C_\mathcal{R}\leq \delta,$$ then the system (\[systemshort\]) admits a solution.
For a slightly more detailed expression of the smallness assumptions, see Section \[theo: system\]. The constant $\displaystyle C(n,h)=\frac{C(n)}{S_h^{n-1}}$ appears explicitly in a paper by Hebey, Pacard and Pollack ([@HebPacPol08], Corollary 3.1). By $S_h$ we understand the Sobolev constant which is defined as the smallest constant $S_h>0$ such that $$\int_M |v|^q\,dv_g\leq S_h\left(\int_M(|\nabla v|^2+hv^2)\,dv_g\right)^\frac{q}{2}$$ for all $v\in H^1(M)$.
The following corollary deals with the existence of solutions to the conformal system. It suffices to take $$\begin{cases}
h=\frac{n-2}{4(n-1)}\left(\mathcal{R}_g-|\nabla\psi|^2_g\right),\quad f=\frac{n-2}{4(n-1)}\left[2V(\psi)-\frac{n-1}{n}(\tau^*)^2\right],\\
\rho_1=\frac{n-2}{4(n-1)}\left(\pi-\frac{n-1}{n}(\tilde{N})^2div_g(\tilde{V})\right),\quad \rho_2=\sqrt{\frac{n-2}{4(n-1)}}\frac{\tilde{N}}{2}, \quad \Psi=\sqrt{\frac{n-2}{4(n-1)}}U,\\
b=\frac{n-2}{2n}\tau^*\tilde{N}div_g(\tilde{V}),\quad c=2\sqrt{\frac{n-2}{4n}},\quad d=\tau^*\\
Y=\sqrt{\frac{n}{n-2}}\tilde{N}\tilde{V},\quad \rho_3=\ln\tilde{N},\\
\mathcal{R}=\frac{n-1}{n}div_g(\tilde{V})\nabla\ln\tilde{N}+\frac{n-1}{n}\nabla(div_g(\tilde{V}))+\frac{\pi\delta_i\psi}{\tilde{N}}\\\quad+2\langle\tilde{V},\frac{\nabla u}{u}\rangle\nabla\ln\tilde{N}-2\frac{n-1}{n+1}\frac{\langle\tilde{V},\nabla u\rangle\nabla u}{u^2}-\frac{n-1}{n}\langle\tilde{V},\frac{\Delta_g u}{u}\rangle
\end{cases}$$ in (\[systemshort\]). It is a direct application of Theorem \[thm 1\].
\[physical result\] Let $\Delta_g + \frac{n-2}{4(n-1)}\left(\mathcal{R}_g-|\nabla\psi|^2_g\right)$ be a coercive operator. Assume that $$2V(\psi)>\frac{n-1}{n}(\tau^*)^2,\quad \pi>\frac{n-1}{n}(\tilde{N})^2div_g(\tilde{V})\quad\text{and}\quad|\nabla \ln\tilde{N}|<C_1^{-1},$$ where $C_1$ depends on $n$ and $g$ (see (\[lxest\]) for more details). Moreover, assume that $$||\pi-\frac{n-1}{n}(\tilde{N})^2div_g(\tilde{V})||_{L^1}\leq C(n,g,h)||2V(\psi)-\frac{n-1}{n}(\tau^*)^2||^{1-n}_{L^\infty}.$$ Then there exists a constant $$\begin{array}{c}
\delta=\delta\Big(\inf_M\frac{n-2}{4(n-1)}\left[2V(\psi)-\frac{n-1}{n}(\tau^*)^2\right],
\inf_M\frac{n-2}{4(n-1)}\left(\pi-\frac{n-1}{n}(\tilde{N})^2div_g(\tilde{V})\right)\\
\tau^*,||\pi||_{\mathcal{C}^{0,\gamma}},
||\mathcal{R}_g-|\nabla\psi|^2_g||_{\mathcal{C}^{0;\gamma}},||2V(\psi)||_{\mathcal{C}^{1,\gamma}}\Big)>0
\end{array}$$ such that, if $$||U||_{\mathcal{C}^{0,\gamma}}+||\pi||_{\mathcal{C}^{0,\gamma}}+||\nabla\psi||_{\mathcal{C}^{0,\gamma}}+||\ln \tilde N||_{\mathcal{C}^{1,\gamma}}+||\tilde{V}||_{\mathcal{C}^{1,\gamma}}\leq \delta,$$ then (\[syst of Maxwell\]) admits a solution $(u,W)$, where $u$ is a smooth positive function on $M$ and $W$ a smooth vector field on $M$.
A few remarks on the results of the present paper. The classical system of constraint equations obtained by the conformal method (without the modifications proposed by Maxwell [@Max14]) was studied by Bruno Premoselli ([@Pre14], [@Pre14a]) in the presence of a scalar field. Second, the above system is the subject of a paper by Mike Holst, David Maxwell and Rafe Mazzeo [@HolMaxMaz] - in their case, certain conditions are imposed on the presence of the matter field. We treat the separate and delicate case wherein the dominant non linearity is focusing and leads to possible loss of compactness. It is interesting to note that the size of $n$ plays a role; as Premoselli proves in his paper, while his system may be well-behaved in low dimensions ($3\leq n\leq 5$), it most certainly fails to do so in higher dimensions $(n\geq 6)$. Even if our results are similar to those of Premoselli, they are considerably more difficult to obtain. This is mainly due to the presence of a $|\nabla u|^2$ term in the scalar equation, a term which is not compact a priori.
**Outline of the paper**. Section 2 is devoted to the study of the first equation in (\[systemshort\]), the so-called Lichnerowicz equation. We prove the existence of stable solutions under suitable assumptions.
Section 3 deals with a priori estimates for solutions of the Lichnerowicz equation. A careful blow-up analysis is carried out. As already mentioned, the term $|\nabla u|^2$ poses additional difficulty: blow-up can occur at the $\mathcal{C}^1$ level, even if the solution is bounded in $L^\infty$.
Section 4 is devoted to the proof of Theorem \[thm 1\] and Corollary \[physical result\], which relies heavily on the a priori estimates obtained in Section 3. At the end of Section 4, we also explain how to extend Corollary \[physical result\] in the presence of conformal Killing vector fields.
**Aknowledgements.** It is a pleasure to express my sincere gratitude to Olivier Druet for many helpful discussions and suggestions.
Existence of minimal solutions of the scalar equation
=====================================================
We study the Lichnerowicz-type scalar equation in (\[systemshort\]). The following theorem states that, given the existence of supersolutions, one may use an iterative procedure to obtain a sequence which converges in $\mathcal{C}^1$ norm to a solution. We draw the reader’s attention to the fact that this solution is uniquely determined by its construction. The proof contains some similarities with that of Premoselli [@Pre14], but some new difficulties appear. The main difference here comes from the presence of non-linearities containing gradient terms, which force us to further refine the analysis. These gradient terms lead to difficulties in obtaining a priori estimates on solutions of the equation, which in turn lead to problems of stability. The existence result we prove in this section reads as follows:
\[theo: existence of sol to Lich type eq\] Let $(M,g)$ be a closed Riemannian manifold. Let $a$, $b$, $c$, $d$, $f$, $h$ be smooth functions on $M$ and $Y$ be a smooth vector field on $M$. Assume that $a>0$ and $f>0$. The equation $$\label{EM}
\Delta_g u+ hu-fu^{q-1}-\frac{a}{u^{q+1}}+\frac{b}{u}+\frac{\langle\nabla u, Y\rangle^2}{u^{q+3}}+c\langle\nabla u,Y\rangle\left(\frac{d}{u^2}+\frac{1}{u^{q+2}}\right)=0$$ admits a smooth positive solution $u$ as soon as it admits a supersolution.
The solution obtained by the construction below is unique. Moreover, it is stable (see Lemma \[lem weak stability\] at the end of this section.)
We begin by fixing a supersolution and a subsolution to serve as upper and lower bounds respectively for the iterative process. Let $\psi$ be a positive supersolution of (\[EM\]). Let $\varepsilon_0>0$ be a small constant such that $$\varepsilon_0<\inf_M\psi,\quad \left(\sup_M h\right)\varepsilon_0^{q+2}<\frac{\inf_M a}{2}\quad \text{and}\quad \left(\sup_M b\right)\varepsilon_0^{q}<\frac{\inf_M a}{2}.$$ The last two bounds ensure that $u_0=\varepsilon_0$ is a strict subsolution of (\[EM\]) since $f>0$. We let $$F(t,x)=-f(x)t^{q-1}-\frac{a(x)}{t^{q+1}}+\frac{b(x)}{t}-Kt$$ for $x\in M$ and $t>0$. We choose $K>0$ large enough such that $$\label{monotonicity of F}
\frac{\partial}{\partial t}\left[F(t,x)+\frac{A(x)^2}{t^{q+3}}+c(x)A(x)\left(\frac{d(x)}{t^2}+\frac{1}{t^{q+2}}\right)\right]\leq 0$$ for all $x\in M$ and all $\varepsilon_0\leq t\leq \sup_M\psi$, whatever $A(x)$ is. It is sufficient to take $$\label{K estimate}
\begin{array}{c}
K\geq \sup_{x\in M, \varepsilon_0\leq t\leq \sup_M\psi}\Big[-(q-1)f(x)t^{q-2}+(q+1)\frac{a(x)}{t^{q+2}}-\frac{b(x)}{t^2}\\
+\frac{c(x)^2\left(\frac{2d(x)}{t^3}+\frac{d+2}{t^{q+3}}\right)^2t^{q+4}}{4(q+3)}\Big].
\end{array}$$ Up to choosing $K$ larger, we may also assume that $h+K>0$ and that $F(t,x)$ is negative for $\varepsilon_0\leq t\leq \sup_M\psi$.
We shall now consider a sequence $(u_i)_{i\in\mathbb{N}}$ defined by induction by $u_0\equiv \varepsilon_0$ and $$\label{eq: the Ei eqs}
\begin{array}{c}
\displaystyle (E_i):\quad \Delta_g u_i+(h+K)u_i+F(u_{i-1}(x),x)+\frac{\langle\nabla u_i, Y\rangle^2}{u_{i-1}^{q+3}}\\
\displaystyle \quad +c\langle \nabla u_{i},Y\rangle\left(\frac{d}{u^2_{i-1}}+\frac{1}{u^{q+2}_{i-1}}\right)=0.
\end{array}$$
We prove in Step 1 below that the sequence is well defined. In Step 2, we prove that the sequence if pointwise increasing and uniformly bounded. At last, Step 3 is devoted to the proof that the sequence $(u_i)$ converges to a solution of (\[eq: the Ei eqs\]).
**Step 1:** We prove that $(u_i)$ is well defined. We consider the more general equation $$\label{gen eq}
\Delta_g u+Hu+\theta_1\langle\nabla u,Z\rangle^2+\theta_2\langle\nabla u,Z\rangle+\theta_3=0,$$ with $H$, $\theta_1$, $\theta_2$, $\theta_3$ smooth functions on $M$ and $Z$ a smooth vector field on $M$ such that $\theta_1>0$, $H>0$, $\theta_3<0$. We claim that (\[gen eq\]) admits a unique smooth positive solution.
We shall use the fixed point theorem as stated in Evans [@Eva], Section 9.2.2, Theorem 4. Let us define the operator $T:\mathcal{C}^{1,\gamma}(M)\to\mathcal{C}^{1,\gamma}(M)$ such that $$\Delta_g T(u)+HT(u)+\theta_1\langle\nabla u,Z\rangle^2+\theta_2\langle\nabla u,Z\rangle+\theta_3=0.$$ If we can prove that there exists $C>0$ such that $$\label{fixed point condition}
\forall\quad 0\leq \tau\leq 1,\quad w=\tau T(w)\quad\Rightarrow\quad ||w||_{\mathcal{C}^{1,\gamma}(M)}\leq C,$$ then the operator $T$ will have a fixed point, leading to a solution of (\[gen eq\]). Note that this solution will be unique. Indeed, assume that $w_1$ and $w_2$ are two solutions of (\[gen eq\]), then at a point of maximum $x_0$ of $w_1-w_2$, we have that $\nabla w_1(x_0)=\nabla w_2(x_0)$ and $\Delta_g w_1(x_0)\geq \Delta_g w_2(x_0)$ so that (\[gen eq\]) gives $$H(x_0)\left(w_1(x_0)-w_2(x_0)\right)\leq 0.$$ Since $H>0$, this leads to $w_1\leq w_2$. By symmetry, uniqueness is proved. Note that the fixed point of $T$ is smooth and positive by the standard regularity theory and the maximum principle.
Thus we are left with the proof of (\[fixed point condition\]). Let $0\leq\sigma_m\leq 1$ and let $w_m\in\mathcal{C}^{1,\gamma}(M)$ be such that $$w_m=\sigma_m T(w_m).$$ Multiplying (\[gen eq\]) by $\sigma_m$, we obtain that $$\Delta_g w_m+Hw_m+\sigma_m\theta_1\langle\nabla w_m,Z\rangle^2+\sigma_m\theta_2\langle\nabla w_m,Z\rangle+\sigma_m\theta_3=0.$$ First, the $L^\infty$ bounds on $w_m$ exist *a priori*. Indeed, consider $x_0\in M$ a minimum of $w_m$. Since $\Delta_g w_m(x_0)\leq 0$ and $\nabla w_m(x_0)=0$, which holds true for all minima, then $Hw_m(x_0)\geq -\sigma_m \theta_3(x_0).$ By applying the same procedure to the study of maxima, we obtain that $$\label{a priori b}
\inf_M\frac{-\sigma_m\theta_3}{H}\leq w_m\leq \sup_M\frac{-\sigma_m\theta_3}{H}.$$ Assume now that $||\nabla w_m||_{L^\infty(M)}\to\infty$. Let $$\mu_m:=\frac{1}{||\nabla w_m||_{L^\infty(M)}}\to 0\quad\text{ as }\quad m\to\infty,$$ and $(x_m)_m\subset M$ be such that $$||\nabla w_m||_{L^\infty(M)}=|\nabla w_m(x_m)|.$$ Consider the domains $\displaystyle \Omega_m:=B_{
0}\left(\frac{i_g (M)}{2\mu_m}\right)$, where $i_g(M)$ is the injectivity radius of $M$, and the rescaled quantities $$v_m(x):=w_m\left(\exp_{x_m}(\mu_m x)\right)\quad\text{and}\quad g_m(x):=\left(\exp_{x_m}^*g\right)(\mu_m x).$$ Clearly, $||\nabla v_m||_{L^\infty}\leq 1$ and $|\nabla v_m(0)|=1$. The $L^\infty$ bounds remain unchanged. In $(\Omega_m)_{m\geq 1}$, we have that $$\begin{array}{c}
\Delta_{g_m}v_m+\mu_m^2H\left(\exp_{x_m}(\mu_m\cdot)\right)v_m+\sigma_m\mu_m^2\theta_3\left(\exp_{x_m}(\mu_m\cdot)\right)\\ \\
+\sigma_m\theta_1\left(\exp_{x_m}(\mu_m\cdot)\right)\langle\nabla v_m,Z\left(\exp_{x_m}(\mu_m\cdot)\right)\rangle^2\\ \\
+\mu_m\sigma_m\theta_2\left(\exp_{x_m}(\mu_m\cdot)\right)\langle\nabla v_m, Z\left(\exp_{x_m}(\mu_m\cdot)\right)\rangle=0
\end{array}$$ Note that $g_m\to\xi$ in $\mathcal{C}^2_{loc}({\mathbb{R}}^n).$ By standard elliptic theory, $(v_m)_m$ is bounded in $\mathcal{C}^{1,\eta}_{loc}({\mathbb{R}}^n)$, with $\eta\in (0,1)$. We may extract, up to a subsequence, $v_\infty=\lim_{m\to\infty}v_m$, $x_\infty=\lim_{m\to\infty}x_n$ and $\sigma_\infty:=\lim_{m\to\infty}\sigma_m$. From this is follows that $||\nabla v_\infty||_{L^\infty}=1$ and that the [*a priori* ]{} bounds (\[a priori b\]) become $$\label{bounds on v infinity}
\inf_M\frac{-\sigma_{\infty}\theta_3}{H}\leq v_\infty\leq \sup_M\frac{-\sigma_{\infty}\theta_3}{H}.$$ Moreover, $v_\infty$ solves the limit equation $$\Delta v_\infty+\left(\partial_1 v_\infty\right)^2=0$$ in ${\mathbb{R}}^n$, where we have let $$\partial_1 v_\infty:=\sqrt{\sigma_\infty\theta_1(0)}\nabla v_\infty\cdot Z(x_0).$$ If $\sigma_\infty=0$, then $v_\infty$ is a bounded harmonic function, and thus a constant. Let us assume that $\sigma_\infty\not = 0$. Note that, for $\alpha\in{\mathbb{R}}$, $$\begin{array}{r l}
\displaystyle \Delta v_\infty^{-\alpha} &\displaystyle =-\frac{\alpha}{v_\infty^{\alpha+1}}\Delta v_\infty-\frac{\alpha(\alpha+1)|\nabla v_\infty|^2}{v_\infty^{\alpha+2}}\\
&\displaystyle \leq\frac{\alpha|\nabla v_\infty|^2}{v_\infty^{\alpha+1}}\left(\sigma_\infty\theta_1(0)|Z(0)|^2-\frac{\alpha+1}{v_\infty}\right).
\end{array}$$ This and (\[bounds on v infinity\]) imply that, for $\alpha$ sufficiently large, $v_\infty^{-\alpha}$ is subharmonic. We then apply Lemma \[lem: A1\] (see annex) to get that $v_\infty$ must be constant. Whichever the case, $\nabla v_\infty\equiv 0$ leads to a contradiction. The $\mathcal{C}^{1,\gamma}$ bound follows from an elliptic regularity argument. This ends the proof of Step 1.
**Step 2:** We claim that $$\varepsilon_0\leq u_i(x)\leq u_{i+1}(x)\leq \psi(x)$$ for all $x\in M$ and all $i\leq 0$.
We proceed by induction. We prove first that $$\label{inductive hyp}
\forall\, i\geq 0,\quad u_i \text{ is a subsolution of } (E_{i+1}) \text{ and } u_i\leq \psi.$$ Note that $$\label{induction}
(\ref{inductive hyp}) \Rightarrow u_i\leq u_{i+1}.$$ Indeed, let $x_0\in M$ be a maximum point of $u_i-u_{i+1}.$ Then $\nabla u_i(x_0)=\nabla u_{i+1}(x_0)$ and we can use the fact that $u_i$ is a subsolution of $(E_{i+1})$ and $u_{i+1}$ a solution of $(E_{i+1})$ to write that $$\Delta_g(u_i-u_{i+1})(x_0)+(h(x_0)+K)(u_i-u_{i+1})(x_0)\leq 0$$ which implies that $u_i(x_0)\leq u_{i+1}(x_0)$ since $\Delta_g(u_i-u_{i+1})(x_0)\geq 0$ and $h+K>0$. This proves (\[induction\]).
We now prove (\[inductive hyp\]) by induction. For $i=0$, it follows from the choice of $\varepsilon_0$ we made. Assume that (\[inductive hyp\]) holds for some $i\geq 0$. We need to prove that $u_{i+1}$ is a subsolution of $(E_{i+2})$. It suffices to show that $$\begin{array}{c}
\Delta_g u_{i+2}+(h+K)u_{i+2}+F(u_{i+1}(x),x)+\frac{\langle\nabla u_{i+2}, Y\rangle^2}{u_{i+1}^{q+3}}\\
+c\langle \nabla u_{i+2},Y\rangle\left(\frac{d}{u^2_{i+1}}+\frac{1}{u^{q+2}_{i+1}}\right)\leq \Delta_g u_{i+1}+(h+K)u_{i+1}+F(u_{i+1}(x),x)\\+\frac{\langle\nabla u_{i+1}, Y\rangle^2}{u_{i+1}^{q+3}} +c\langle \nabla u_{i+1},Y\rangle\left(\frac{d}{u^2_{i+1}}+\frac{1}{u^{q+2}_{i+1}}\right),
\end{array}$$ since $u_{i+1}$ is defined as a solution of $(E_{i+1})$. This is equivalent to showing that $$\begin{array}{c}
F(u_{i+1})+c\langle \nabla u_{i+1},Y\rangle\left(\frac{d}{u_{i+1}^2}+\frac{1}{u_{i+1}^{q+2}}\right)+\frac{\langle\nabla u_{i+1},Y\rangle^2}{u_{i+1}^{q+3}}\\ \leq F(u_{i})+c\langle\nabla u_{i+1},Y\rangle\left(\frac{d}{u_{i}^2}+\frac{1}{u_{i}^{q+2}}\right)+\frac{\langle\nabla u_{i+1},Y\rangle^2}{u_{i}^{q+3}}.
\end{array}$$ And this is a consequence of (\[monotonicity of F\]) with $A(x)=\langle\nabla u_{i+1},Y\rangle$, since (\[induction\]) implies that $u_{i+1}\geq u_i$ by induction hypothesis. Thus, $u_{i+1}$ is a subsolution of $(E_{i+2})$.
Finally, so as to check the last point, assume there exists $x_0\in M$ such that $u_{i+1}(x_0)>\psi(x_0)$ and that it corresponds to $\max_M\left(u_{i+1}(x)-\psi(x)\right).$ Since $\nabla u_{i+1}(x_0)=\nabla \psi (x_0)$ and $\Delta_g u_{i+1}(x_0)\geq \Delta_g\psi(x_0),$ we obtain that $$\Delta_g(u_{i+1}-\psi)(x_0)+(h+K)(u_{i+1}-\psi)(x_0)>0.$$ But $\psi$ is a supersolution for (\[EM\]), so we get that $$\begin{array}{c}
0<\Delta_g(u_{i+1}-\psi)(x_0)+(h(x_0)+K)(u_{i+1}-\psi)(x_0)\\
\leq F(\psi(x_0),x_0)-F(u_i(x_0),x_0)-\langle\nabla\psi(x_0),Y(x_0)\rangle^2\left(\frac{1}{u_i^{q+3}}-\frac{1}{\psi^{q+3}}\right)(x_0)\\
-\langle\nabla\psi(x_0),Y(x_0)\rangle\left(\frac{d}{u_i^2}-\frac{d}{\psi^2}+\frac{1}{u_i^{q+2}}-\frac{1}{\psi^{q+2}}\right)(x_0).
\end{array}$$ Thanks to (\[monotonicity of F\]) with $A(x)=\langle\nabla\psi(x),Y(x)\rangle$ and to the induction hypothesis which says that $u_i\leq\psi$, we obtain a contradiction. This wraps up the induction argument and the proof of Step 2.
**Step 3:** The sequence $(u_i)_{i\in\mathbb{N}}$ is uniformly bounded in $\mathcal{C}^1(M)$.
Thanks to Step 2, we know that $(u_i)_{i\in\mathbb{N}}$ is an increasing sequence bounded by $\psi$. Thus there exists $u\in\mathcal{C}^0(M)$ such that $u_i\to u_0$ in $\mathcal{C}^0(M).$
Assume by contradiction that exists a subsequence $(u_{\phi(m)})_{m\in N}$ such that $||\nabla u_{\phi(m)}||_{L^\infty}\to\infty.$ Let $$\displaystyle \mu_m:=\frac{1}{||\nabla u_{\phi(m)}||_{L^\infty}}$$ and let $(x_m)_m\subset M$ be such that $$|\nabla u_{\phi(m)}(x_m)|=||\nabla u_{\phi(m)}||_{L^\infty}.$$ Consider the domains $\displaystyle \Omega_m=B_{0}\left(\frac{i_g M}{2\mu_m}\right)$ and the rescaled quantities $$v_m(x):=u_{\phi(m)}\left(\exp_{x_m}(\mu_m x)\right)\quad\text{ and }\quad g_m(x):=\left(\exp_{x_m}^* g\right)(\mu_m x)$$ in $\Omega_m$. We get $$\label{eq: scaling in exist thm}
\begin{array}{c}
\Delta_{g_m}v_m+\mu_m^2(h\left(\exp_{x_m}(\mu_m \cdot)\right)+K)v_m\left(\exp_{x_m}(\mu_m \cdot)\right)\\
\displaystyle +\mu_m^2F\left(u_{\phi(m)-1}\left(\exp_{x_m}(\mu_m \cdot)\right)\right)
\displaystyle +\frac{\langle\nabla v_m,Y\left(\exp_{x_m}(\mu_m \cdot)\right)\rangle^2}{u_{\phi(m)-1}^{q+3}\left(\exp_{x_m}(\mu_m \cdot)\right)}\\
\displaystyle +\mu_m\langle\nabla v_m, Y\left(\exp_{x_m}(\mu_m \cdot)\right)\rangle c\left(\exp_{x_m}(\mu_m \cdot)\right)\Big[\frac{d\left(\exp_{x_m}(\mu_m \cdot)\right)}{u_{\phi(m)-1}^2\left(\exp_{x_m}(\mu_m \cdot)\right)}\\
\displaystyle +\frac{1}{u_{\phi(m)-1}^{q+2}\left(\exp_{x_m}(\mu_m \cdot)\right)}\Big]=0
\end{array}$$ with $(v_m)_{m\in N}$ bounded in $L^\infty$, $||\nabla v_m||_{L^\infty}=1$ and $\varepsilon_0\leq v_m$. By the Sobolev embedding theorem and standard elliptic regularity, there exists a smooth positive limit $v_\infty$ of $(v_m)_{m\in N}$, up to a subsequence. Recall that $(u_i)_{i\in N}$ converges everywhere, so $u_{\phi(m)-1}(\mu_m x)\to v_\infty(x)$ in $M$. By taking $m\to\infty$ in (\[eq: scaling in exist thm\]), $$\Delta v_\infty+\frac{\left(\nabla v_\infty\cdot Y(0)\right)^2}{v_\infty^{q+3}}=0.$$ Note also that $$\Delta v_\infty^{-\alpha}\leq \frac{\alpha|\nabla v_\infty|^2}{v_\infty^{\alpha+2}}\left(\frac{|Y(0)|^2}{v_\infty^{q+2}}-(\alpha+1)\right).$$ For $\alpha$ large enough, $v_\infty^{-\alpha}$ is subharmonic. Using Lemma \[lem: A1\] (see Annex), we find that $v_\infty$ is constant, which contradicts the fact that $||\nabla v_\infty||_{L^\infty}=1.$ This ends the proof of Step 3.
Since $(u_i)_{i\in N}$ is uniformly bounded in $\mathcal{C}^1$, we conclude by standard elliptic theory that its limit $u$ is a positive smooth function solving equation (\[EM\]). This ends the proof of the theorem.
The solution constructed in the previous proof is uniquely determined as the pointwise limit of $(u_i)_{i\in N}$, where each $u_i$ is the unique solution of (\[eq: the Ei eqs\]). Furthermore, the solution is minimal among all supersolutions (including solutions) of (\[EM\]) with values between $\varepsilon_0$ and $\sup_M\psi$. These bounds were explicitly used in the inductive argument. By construction, $u\leq\psi$, where $\psi$ is the supersolution fixed at the very beginning. Note that the constant $K$ appearing in (\[eq: the Ei eqs\]) depend on $\sup_M\psi$ and $\varepsilon_0$. We would obtain the same iteration were we to use another supersolution $\tilde{\psi}$ and the same $K$, given that $\varepsilon_0<\tilde{\psi}<\sup_M\psi$. Therefore, $u$ is smaller than any supersolution between $\varepsilon_0$ and $\sup_M\psi$.
As an immediate consequence of the minimality discussed above, the solutions we found corresponding to different functions $a$ are ordered. Let $0<a<\tilde{a}$ be two functions, and assume that the equation associated to $\tilde{a}$ admits a solution $\tilde{u}$. Then $\tilde{u}$ is a supersolution for (\[EM\]) corresponding to $a$, and by the previous proof we find a solution $u\leq\tilde{u}$. Moreover, given that $\tilde{u}$ may be viewed as a supersolution to all (\[EM\]) with $a\leq\tilde{a}$, we obtain a monotonicity of $u$ in $a$: for $a_1\leq a_2\leq \tilde{a}$, then $u_1\leq u_2\leq \tilde{u}$.
Finally, the solution $u$ is stable, as defined in the following lemma.
\[lem weak stability\] The operator $L$ resulting from the linearization of (\[EM\]) at the minimal solution $u$ admits a real, simple eigenvalue $\lambda_0\geq 0$ such that $$\begin{array}{c}
\displaystyle L\varphi_0=\Delta_g\varphi_0+\Big[h-(q-1)fu^{q-2}+(q+1)au^{-q-2}-bu^{-2}\\
-(q+3)\frac{\langle\nabla u,Y\rangle^2}{u^{q+4}}-c\langle\nabla u,Y\rangle\left(\frac{2d}{u^3}+\frac{q+2}{u^{q+3}}\right)\Big]\varphi_0\\+\langle\nabla\varphi_0,Y\rangle\Big[c\left(\frac{d}{u^2}+\frac{1}{u^{q+2}}\right)+\frac{2\langle\nabla u,Y\rangle}{u^{q+3}}\Big]\\=\lambda_0\varphi_0,
\end{array}$$ where $\varphi_0$ is the corresponding positive eigenfunction. Furthermore, if $\lambda\in {\mathbb{C}}$ is any other eigenvalue, then $Re(\lambda)\geq \lambda_0$.
Notice that $L$ is nonsymmetric; moreover, one may find a large enough constant $K$ such that $$\begin{array}{c}
h-(q-1)fu^{q-2}+(q+1)au^{-q-2}-bu^{-2}-(q+3)\frac{\langle\nabla u,Y\rangle^2}{u^{q+4}}\\
-c\langle\nabla u,Y\rangle\left(\frac{2d}{u^3}+\frac{q+2}{u^{q+3}}\right)+K\geq 0.
\end{array}$$ According to ([@Eva], Section 6.5, Theorem 1) there exists a real, positive eigenvalue $\lambda_K>0$ of $L+K$, such that any other complex eigenvalue of $L+K$ has a greater real part. Consequently, the operator $L$ admits a minimal real eigenvalue $\lambda_0>-K$. We now assume that $\lambda_0<0$. Let $u_\delta:=u_0-\delta\varphi_0$, $\delta>0$. By taking $\delta$ small enough, we may ensure that $\varepsilon_0< u_\delta$. Then $$\begin{array}{c}
\displaystyle \Delta_g u_\delta+hu_\delta-fu_\delta^{q-1}-\frac{a}{u_\delta^{q+1}}+\frac{b}{u_\delta}+\frac{\langle\nabla u_\delta,Y\rangle^2}{u_\delta^{q+3}}+c\langle\nabla u_\delta, Y\rangle\left(\frac{d}{u_\delta^2}+\frac{1}{u_\delta^{q+2}}\right)\\ \\
\displaystyle =-\delta\lambda_0\varphi_0+o(\delta).
\end{array}$$ This implies that $\varepsilon_0< u_\delta<u$ is a supersolution of (\[EM\]), which cannot be the case, as discussed above. Thus, $\lambda_0\geq 0.$
A priori estimates on solutions of the scalar equation in low dimensions
========================================================================
The $\mathcal{C}^1$ estimates obtained in this section will play a crucial role in the proof of Theorem \[thm 1\], which is based on a fixed-point argument. This section is devoted to the proof of the following theorem:
\[theo stability eq\] Let $(M,g)$ be a closed Riemannian manifold of dimension $n=3,4,5$. Let $\frac{1}{2}<\eta<1$ and $0<\alpha<1$. Let $a$, $b$, $c$, $d$, $f$, $h$ be smooth functions on $M$, let $Y$ be a smooth vector field on the $M$.
For any $0<\theta<T$, there exists $S_{\theta,T}$ such that any smooth positive solution $u$ of (\[EM\]) with parameters within $$\begin{array}{c}
\mathcal{E}_{\theta,T}:=\Big\{(f,a,b,c,d,h,Y),\quad f\geq \theta,\quad a\geq \theta, \\
\quad ||f||_{\mathcal{C}^{1,\eta}}\leq T,\quad ||a||_{\mathcal{C}^{0,\alpha}},||b||_{\mathcal{C}^{0,\alpha}}, ||c||_{\mathcal{C}^{0,\alpha}}, ||d||_{\mathcal{C}^{0,\alpha}}, ||h||_{\mathcal{C}^{0,\alpha}}, ||Y||_{\mathcal{C}^{0,\alpha}}\leq T\Big\},
\end{array}$$ satisfies $||u||_{\mathcal{C}^2}\leq S_{\theta,T}$.
For the sake of clarity, we’ve taken the bounds on the parameters to be of the form $\theta$ and $T$. They can of course be individually specified.
We proceed by contradiction. We assume the existence of a sequence $(u_\alpha)_{\alpha\in{\mathbb{N}}}$ of smooth positive solutions of equations ($EL_\alpha$) $$\label{eq: EL alpha}
\begin{array}{c}
\Delta_g u_\alpha+ h_\alpha u_\alpha-f_\alpha u_\alpha^{q-1}- \frac{a_\alpha}{u_\alpha^{q+1}}+\frac{b_\alpha}{u_\alpha}+\frac{\langle\nabla u_\alpha, Y_\alpha\rangle^2}{u_\alpha^{q+3}}\\
+c_\alpha\langle\nabla u_\alpha,Y_\alpha\rangle\left[\frac{d_\alpha}{u_\alpha^2}+\frac{1}{u_\alpha^{q+2}}\right]=0
\end{array}$$ with parameters $(a_\alpha, b_\alpha, c_\alpha, d_\alpha, f_\alpha,h_\alpha, Y_\alpha)$ in $\mathcal{E}_{\theta,T}$ such that $$\label{est: explosion}
||u_\alpha||_{\mathcal{C}^1(M)}\to\infty,\quad\text{ as }\alpha\to\infty.$$ A *concentration point* is the limit in $M$ of any sequence $(x_\alpha)_\alpha$ where (\[est: explosion\]) holds. Note that a $\mathcal{C}^1$-bound on $(u_\alpha)_\alpha$ automatically gives a $\mathcal{C}^2$-bound by elliptic theory. Note also that, up to a subsequence, all parameters converge in $\mathcal{C}^0(M).$
Let $m_\alpha=\min_{x\in M}u_\alpha(x)=u_\alpha(x_\alpha)>0.$ Since $\nabla u_\alpha(x_\alpha)=0$ and since $\Delta_g u_\alpha(x_\alpha)\leq 0$, we have thanks to (\[eq: EL alpha\]) that $$h_\alpha(x_\alpha)m_\alpha-f_\alpha(x_\alpha)m_\alpha^{q-1}-\frac{a_\alpha(x_\alpha)}{m_\alpha^{q+1}}+\frac{b_\alpha(x_\alpha)}{m_\alpha}\geq 0.$$ Thanks to the definition of $\mathcal{E}_{\theta,T},$ it follows that $$\frac{\theta}{m_\alpha^{q+1}}\leq T(m_\alpha+\frac{1}{m_\alpha}).$$ Then there exists $\varepsilon=\varepsilon(\theta,T,n)>0$ such that $m_\alpha\geq \varepsilon$, meaning that $$\label{varepsilon}
u_\alpha>\varepsilon>0\quad \text{ for all } x\in M \text{ and all } \alpha.$$
The scheme of the proof follows the work of Druet and Hebey [@DruHeb], with the added difficulty consisting in the gradient terms in (\[eq: EL alpha\]).
Concentration points
--------------------
The first step in finding *a priori* estimates for $(u_\alpha)_\alpha$ is to find all potential concentration points.
\[lem: former lemma 2\] There exists $N_\alpha\in {\mathbb{N}}^*$ and $\displaystyle \mathcal{S}_\alpha:=\left(x_{1,\alpha},\dots x_{N_\alpha, \alpha}\right)$ a set of critical points of $(u_\alpha)_\alpha$ such that $$\label{est: 1 of former lemma 1}
d_g(x_{i,\alpha},x_{j,\alpha})^{\frac{n-2}{2}}u_\alpha(x_{i,\alpha})\geq 1$$ for all $i,j\in\{1,\dots,N_\alpha\}$, $i\not =j$, and $$\label{est: add fact}
\left(\min_{i=1,\dots, N_\alpha}d_g(x_{i,\alpha},x)\right)^{\frac{n-2}{2}}u_\alpha(x)\leq 1$$ for all critical points of $u_\alpha$ and such that there exists $C_1>0$ such that $$\label{est: initial estimate}\left(\min_{i=1,\dots N_\alpha} d_g(x_{i,\alpha},x)\right)^{\frac{n-2}{2}}\left(u_\alpha(x)+\left|\frac{\nabla u_\alpha(x)}{u_\alpha(x)}\right|^{\frac{n-2}{2}}\right)\leq C_1$$ for all $x\in M$ and all $\alpha\in {\mathbb{N}}$.
Estimate (\[est: initial estimate\]) implies that any concentration point of $(u_\alpha)_{\alpha\in N}$ calls for the existence of a sequence $(x_\alpha)_\alpha\subset (\mathcal{S}_\alpha)_\alpha$ converging to it. We shall focus our analysis in the neighbourhood of $S_\alpha$ as $\alpha\to\infty$ to find concentration points.
In order to choose $(S_\alpha)_\alpha$, we make use of a simple result describing any sufficiently regular function on a compact manifold.
\[lem: former lemma 1\] Let $u$ be a positive real-valued $\mathcal{C}^2$ function defined in a compact manifold $M$. Then there exists $N\in {\mathbb{N}}^*$ and $\left(x_1,\, x_2,\dots x_N\right)$ a set of critical points of $u$ such that $$d_g(x_i,x_j)^{\frac{n-2}{2}}u(x_i)\geq 1$$ for all $i,\,j\in\{1, \dots, N\}$, $i\not= j$, and $$\left(\min_{i=1,\dots, N}d_g(x_i,x)\right)^{\frac{n-2}{2}}u(x)\leq 1$$ for all critical points $x$ of $u$.
The lemma and its proof may be found in Druet and Hebey’s paper [@DruHeb]. Applying this lemma to $(u_\alpha)$ gives $N_\alpha$ and $\mathcal{S}_\alpha$ as in Lemma \[lem: former lemma 2\] such that (\[est: 1 of former lemma 1\]) and (\[est: add fact\]) hold. We need to prove (\[est: initial estimate\]). Proceeding by contradiction, assume that there exists a sequence $(x_\alpha)_\alpha$ such that $$\label{est: contradiction initial estimate}\left(\min_{i=1,\dots N_\alpha} d_g(x_{i,\alpha},x_\alpha)\right)^{\frac{n-2}{2}}\left(u_\alpha(x_\alpha)+\left|\frac{\nabla u_\alpha(x_\alpha)}{u_\alpha(x_\alpha)}\right|^{\frac{n-2}{2}}\right)\to\infty$$ as $\alpha\to\infty$, where $$\label{contr}
\begin{array}{c}
\displaystyle \left(\min_{i=1,\dots N_\alpha} d_g(x_{i,\alpha},x_\alpha)\right)^{\frac{n-2}{2}}\left(u_\alpha(x_\alpha)+\left|\frac{\nabla u_\alpha(x_\alpha)}{u_\alpha(x_\alpha)}\right|^{\frac{n-2}{2}}\right)\\ \\
\displaystyle =\sup_{x\in M}\left(\min_{i=1,\dots N_\alpha} d_g(x_{i,\alpha},x)\right)^{\frac{n-2}{2}}\left(u_\alpha(x)+\left|\frac{\nabla u_\alpha(x)}{u_\alpha(x)}\right|^{\frac{n-2}{2}}\right).
\end{array}$$
Denote $$\nu_\alpha^{1-\frac{n}{2}}:=u_\alpha(x_\alpha)+\left|\frac{\nabla u_\alpha(x_\alpha)}{u_\alpha(x_\alpha)}\right|^{\frac{n-2}{2}}$$ and see that (\[est: contradiction initial estimate\]) translates to $$\label{est: distance to s alpha}
\frac{d_g(x_\alpha, \mathcal{S}_\alpha)}{\nu_\alpha}\to\infty\quad\text{as}\quad \alpha\to\infty.$$ Also, since $M$ is compact, $$\label{nu goes to zero}
\nu_\alpha\to 0 \text{ as }\alpha\to\infty.$$ Consider the rescaled quantities $$v_\alpha(x):=\nu_\alpha^{\frac{n-2}{2}}u_\alpha\left(\exp_{x_\alpha}(\nu_\alpha x)\right)\quad\text{and}\quad g_\alpha(x):=\left(\exp_{x_\alpha}^* g\right)(\nu_\alpha x)$$ defined in $\Omega_\alpha:=B_0\left(\frac{\delta}{\nu_\alpha}\right)$, with $ 0<\delta<\frac{1}{2}i_g(M).$ We emphasize that, for any $R>0$, $$\label{est: size of v alpha and co}
\limsup_{\alpha\to\infty}\sup_{B_0(R)} \left(v_\alpha+\left|\frac{\nabla v_\alpha}{v_\alpha}\right|^{\frac{n-2}{2}}\right) =1$$ thanks to (\[contr\]) and (\[est: distance to s alpha\]). However, unlike $(u_\alpha)_\alpha$, the sequence $(v_\alpha)_\alpha$ is not necessarily bounded from below by a small positive constant $\varepsilon$. Instead, we deduce from (\[est: size of v alpha and co\]) that $$|\nabla \ln v_\alpha|\leq 1+o(1)\quad\text{ in } B_0(R)$$ for all $R>0$ so that $$\label{size of v alpha}
v_\alpha(0)e^{-2|x|}\leq v_\alpha(x)\leq v_\alpha(0)e^{2|x|}\quad\text{ in } B_0(R)$$ for all $R>0$ as soon as $\alpha$ is large enough. We rewrite (\[eq: EL alpha\]) in $\Omega_\alpha$ as $$\label{eq: Ev alpha eqs}
\begin{array}{c}
\Delta_{g_\alpha}v_\alpha = f_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)v_\alpha^{q-1}+\nu_\alpha^{\frac{n+2}{2}}\frac{a_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}{u_\alpha^{q+1}\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)} \\
-\nu_\alpha^2 h_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)v_\alpha-\nu_\alpha^{\frac{n+2}{2}}\frac{b_\alpha \left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}{u_\alpha \left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}\\
-\frac{\langle\nabla v_\alpha,Y_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)\rangle^2}{v_\alpha}\frac{1}{u_\alpha^{q+2}\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}\\
-\nu_\alpha c_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)\langle\nabla v_\alpha,Y_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)\rangle\Big[\frac{d_\alpha \left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}{u_\alpha^2\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}\\
+\frac{1}{u_\alpha^{q+2}\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}\Big]
\end{array}$$ Note that the metrics $g_\alpha\to\xi$ in $\mathcal{C}^2_{loc}$ as $\alpha\to\infty$. Because of (\[varepsilon\]) and (\[est: size of v alpha and co\]), the right hand side is bounded, so by standard elliptic theory there exists up to a subsequence a $\mathcal{C}^1$ limit $U:=\lim_{\alpha\to\infty}v_\alpha$ and $x_0:=\lim_{\alpha\to\infty}x_\alpha$.
**First case:** If $u_\alpha(x_\alpha)\to\infty$ as $\alpha\to\infty$, then we can pass to the limit in equation (\[eq: Ev alpha eqs\]) to get $$\Delta U=f(x_0)U^{q-1}$$ in ${\mathbb{R}}^n$. The exact form of these solutions is found in a paper by Caffarelli, Gidas and Spruck [@CaffGidSpr]: $$U(x)=\left(1+\frac{f(x_0)|x-y_0|^2}{n(n-2)}\right)^{1-\frac{n}{2}}$$ with $y_0\in {\mathbb{R}}^n$ the unique maximum point of the function. There exist therefore $(y_\alpha)_\alpha$ local maxima of $(u_\alpha)_\alpha$ such that $$\label{est: blow up analysis, distances}
d_g(x_\alpha,y_\alpha)=O(\nu_\alpha)$$ and $$\label{est lim 1}
\nu_\alpha^{\frac{n-2}{2}}u_\alpha(y_\alpha)\to 1\quad \text{ as }\quad \alpha\to\infty.$$ Since $(y_\alpha)_\alpha$ are critical points, (\[est: add fact\]) implies that $$d_g(\mathcal{S}_\alpha, y_\alpha)^{\frac{n-2}{2}}u_\alpha(y_\alpha)\leq 1$$ for all $\alpha\in {\mathbb{N}}$, so by (\[est lim 1\]), $\displaystyle d_g(\mathcal{S}_\alpha,y_\alpha)=O(\nu_\alpha)$; together with (\[est: blow up analysis, distances\]), this leads to $\displaystyle d_g(\mathcal{S}_\alpha,x_\alpha)=O(\nu_\alpha)$, which contradicts (\[est: distance to s alpha\]).
**Second case:** Assume that, up to a subsequence, $u_\alpha(x_\alpha)\to l<\infty$. From (\[nu goes to zero\]) we deduce that $|\nabla u_\alpha (x_\alpha)|\to\infty$ and that $v_\alpha(0)\to 0$ as $\alpha\to\infty$. Let us set $$w_\alpha(x):=\frac{v_\alpha(x)}{v_\alpha(0)}.$$ These functions are bounded from below, since by (\[varepsilon\]), $$\label{est w}
w_\alpha(x)=\frac{u_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}{u_\alpha(x_\alpha)}\geq\frac{\varepsilon}{l}+o(1)$$ in $B_0(R)$ for all $R>0$. Moreover, (\[size of v alpha\]) implies that $$w_\alpha(x)\leq e^{2|x|}$$ in $B_0(R)$ for $\alpha$ large. Multiply (\[eq: Ev alpha eqs\]) by $v_\alpha(0)^{-1}$ to get $$\begin{array}{r l}
\displaystyle \Delta_g w_\alpha= &\displaystyle f_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)w_\alpha^{q-1}v_\alpha^{q-2}(0)+\nu_\alpha^2\frac{a\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}{u_\alpha^{q+1}\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)u_\alpha(x_\alpha)} \\
& \displaystyle -\nu_\alpha^2h_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)w_\alpha-\nu_\alpha^2\frac{b_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}{u_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)u_\alpha(x_\alpha)}\\
& \displaystyle-\frac{\langle\nabla w_\alpha,Y_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)\rangle^2}{w_\alpha^{q+2}}\frac{1}{u_\alpha(x_\alpha)^{q+2}}\\
&\displaystyle -\nu_\alpha c_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)\langle\nabla w_\alpha,Y_\alpha\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)\rangle\Big[\frac{d_\alpha \left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}{u_\alpha^2\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}\\
&\displaystyle +\frac{1}{u_\alpha^{q+2}\left(\exp_{x_\alpha}(\nu_\alpha\cdot)\right)}\Big].
\end{array}$$ By standard elliptic theory, we find that there exists $w:=\lim_{\alpha\to\infty}w_\alpha$ in $\mathcal{C}^1$ solving: $$\Delta w=-\frac{1}{l^{q+2}}\frac{\langle\nabla w, Y(x_0)\rangle^2}{w^{q+3}}$$ in ${\mathbb{R}}^n$. Note that, since $w\geq \frac{\varepsilon}{l}$ by (\[est w\]), $$\Delta w^{-\alpha}\leq \alpha\frac{|\nabla w|^2}{w^{\alpha+2}}\left[\frac{|Y(x_0)|^2}{\varepsilon^{q+2}}-(\alpha+1)\right],$$ so $w^{-\alpha}$ is subharmonic for $\alpha$ large. By applying Lemma \[lem: A1\] (see the Annex), we deduce that $w$ is constant, which in turn implies that $U=0$, and so $\nabla U=0$. This implies that $$\lim_{\alpha\to\infty}\nu_\alpha\frac{\nabla u_\alpha(x_\alpha)}{u_\alpha(x_\alpha)}=0,$$ as it contradicts the choice of $\nu_\alpha$ above, which may be rewritten as $$\nu_\alpha^\frac{n-2}{2}\left(u_\alpha(x_\alpha)+\left|\frac{\nabla u_\alpha(x_\alpha)}{u_\alpha(x_\alpha)}\right|\right)=1,$$ since we are in the case where $\nu_\alpha^{\frac{n-2}{2}}u_\alpha(x_\alpha)=v_\alpha(0)\to 0$ as $\alpha\to 0.$
The following is a Harnack-type inequality. It holds whenever an estimate like (\[est: initial estimate\]) is verified, that is when there exists a constant $C_2$ and a sequence $(x_\alpha,\rho_\alpha)_\alpha$ such that $$\label{est: for Harnack}
d_g(x_\alpha,x)^{\frac{n-2}{2}}\left[u_\alpha(x)+\left|\frac{\nabla u_\alpha(x)}{u_\alpha(x)}\right|^{\frac{n-2}{2}}\right]\leq C_2,\quad\forall x\in B_{x_\alpha}(7\rho_\alpha).$$
\[lem: former lemma 3\] Let $(x_\alpha, \rho_\alpha)_\alpha$ be a sequence such that (\[est: for Harnack\]) holds. Then there exists a constant $C_3>1$ such that for any sequence $0<s_\alpha\leq\rho_\alpha$, we get $$s_\alpha||\nabla u_\alpha||_{L^\infty(\Omega_\alpha)}\leq C_3\sup_{\Omega_\alpha}u_\alpha\leq C_3^2\inf_{\Omega_\alpha}u_\alpha,$$ where $\Omega_\alpha=B_{x_\alpha}(6s_\alpha)\backslash B_{x_\alpha}(\frac{1}{6}s_\alpha).$
Estimate (\[est: for Harnack\]) implies that $$\label{e1}
\left|\frac{\nabla u_\alpha(x)}{u_\alpha(x)}\right|\leq C_2 d_g(x_\alpha,x)^{-1}$$ in $\Omega_\alpha,$ and therefore $$\label{e2}
s_\alpha|\nabla\ln u_\alpha(x)|\leq 6C_2$$ in $\Omega_\alpha$. Taking $C_3\geq 6C_2$, we get the first inequality from (\[e1\]). Then, from (\[e2\]) and from the fact that the domain is an annulus $\Omega_\alpha=B_{x_\alpha}(6s_\alpha)\backslash B_{x_\alpha}(\frac{1}{6}s_\alpha)$, we estimate that $$\sup_{\Omega_\alpha}\ln u_\alpha- \inf_{\Omega_\alpha}\ln u_\alpha\leq l_\alpha(\Omega_\alpha)||\nabla \ln u_\alpha||_{L^\infty(\Omega_\alpha)}\leq 42C_2,$$ where $l_\alpha(\Omega_\alpha)$ is the infimum of the length of a curve in $\Omega_\alpha$ drawn between a maximum and a minimum of $u_\alpha$. Equivalently $$\sup_{\Omega_\alpha} u_\alpha\leq e^{42C_2}\inf_{\Omega_\alpha} u_\alpha,$$ so it suffices to take $C_3=e^{42C_2}$.
Local blow-up analysis
----------------------
In order to show that $u_\alpha$ is bounded in $\mathcal{C}^1$, we define a *blow-up sequence* $(x_\alpha)_\alpha$ with $(\rho_\alpha)_\alpha$ as follows: let $(x_\alpha)_\alpha$ be critical points of $(u_\alpha)_\alpha$ and $(\rho_\alpha)_\alpha$ positive numbers such that they verify the following three conditions: $$\label{est: blow up no 1}
0<\rho_\alpha<\frac{1}{7}i_g(M),$$ $$\label{est: blow up no 2}
\rho_\alpha^{\frac{n-2}{2}}\sup_{B_{x_\alpha}(6\rho_\alpha)}u_\alpha\to\infty,$$ and $$\label{est: blow up no 3}
d_g(x_\alpha,x)^{\frac{n-2}{2}}\left[u_\alpha(x)+\left|\frac{\nabla u_\alpha(x)}{u_\alpha(x)}\right|^{\frac{n-2}{2}}\right]\leq C_2\quad\forall x\in B_{x_\alpha}(7\rho_\alpha),$$ where $C_2$ is a constant. In the rest of the section, we denote $$\mu_\alpha^{1-\frac{n}{2}}:=u_\alpha(x_\alpha).$$
The limit as $\alpha\to \infty$ of a blow-up sequence is a concentration point, as seen from (\[est: blow up no 2\]).
Any sequence $(x_\alpha)_\alpha\subset (S_\alpha)_\alpha$ qualifies as a blow-up sequence as soon as (\[est: blow up no 2\]) is verified. In this case, $(\rho_\alpha)_\alpha$ can be chosen as $$\rho_\alpha:=\min\left(\frac{1}{7}i_g(M),\frac{1}{2}\min_{1\leq i<j\leq N_\alpha}d_g(x_{i,\alpha},x_{j,\alpha})\right)$$ and $C_2=C_1$.
Given any blow-up sequence, the following proposition gathers the central results of our local analysis for the reader’s convenience: namely, it states the exact asymptotic profile of $(u_\alpha)_\alpha$ at distance $(\rho_\alpha)_\alpha$ of $(x_\alpha)_\alpha$ and it gives sharp pointwise asymptotic estimates on balls of radius $\rho_\alpha$. This is the result we shall point to whenever we want to describe the local asymptotic behaviour of $(u_\alpha)_\alpha$ around a concentration point corresponding to local maximum points $x_\alpha$. Note that in this case $\nabla u_\alpha(x_\alpha)=0$.
\[prop\] Let $(x_\alpha)_\alpha$ and $(\rho_\alpha)_\alpha$ be a blow-up sequence. Then there exists $C_4>0$ such that $$u_\alpha(x)+d_g(x_\alpha,x)|\nabla u_\alpha(x)|\leq C_4\mu^{\frac{n-2}{2}}_\alpha d_g(x_\alpha,x)^{2-n}$$ for all $x\in B_{x_\alpha}(6\rho_\alpha)\backslash\{x_\alpha\}.$ Moreover, we see that up to a subsequence, the asymptotic profile of $(u_\alpha)_\alpha$ is $$u_\alpha(x_\alpha)\rho^{n-2}_\alpha u_\alpha(\exp_\alpha(\rho_\alpha x))\to\frac{R_0^{n-2}}{|x|^{n-2}}+H(x)$$ in $C^2_{loc}(B_0(5)\backslash\{0\})$, where $H$ is some harmonic function in $B_0(5)$ satisfying $H(0)=0.$ Here $R_0^2=\frac{n(n-2)}{f(x_0)}$ where $x_0=\lim_{\alpha\to\infty} x_\alpha$.
The proof of this proposition is the subject of this section. It will follow from Lemma \[lem: former lemma 6\] and Lemma \[lem: former lemma 7\] below. We first describe the asymptotic profile at distance $(\mu_\alpha)_\alpha$ of $(x_\alpha)_\alpha$ as $\alpha\to\infty$ of any blow-up sequence.
\[lem: former lemma 4\] Let $(x_\alpha)_\alpha$ with $(\rho_\alpha)_\alpha$ be a blow-up sequence. We have $$\label{est: estimate of old lemma 4}
\mu_\alpha^{\frac{n-2}{2}}u_\alpha\left(\exp_{x_\alpha}(\mu_\alpha x)\right)\to\left(1+\frac{f(x_0)|x|^2}{n(n-2)}\right)^{1-\frac{n}{2}}$$ in $\mathcal{C}^1_{loc}({\mathbb{R}}^n)$ as $\alpha\to\infty$, with $\mu^{1-\frac{n}{2}}_\alpha:=u_\alpha(x_\alpha)$ and $x_0:=\lim_{\alpha\to\infty}x_\alpha$, up to a subsequence.
The proof involves similar arguments to the ones used for Lemma $\ref{lem: former lemma 2}$. Let $y_\alpha\in B_{x_\alpha}(6\rho_\alpha)$ be such that $$u_\alpha(y_\alpha)+\left|\frac{\nabla u_\alpha(y_\alpha)}{u_\alpha(y_\alpha)}\right|^{\frac{n-2}{2}}=\sup_{B_{x_\alpha}(6\rho_\alpha)}\left(u_\alpha(x)+\left|\frac{\nabla u_\alpha(x)}{u_\alpha(x)}\right|^{\frac{n-2}{2}}\right)$$ and let $$\nu_\alpha^{1-\frac{n}{2}}:=u_\alpha(y_\alpha)+\left|\frac{\nabla u_\alpha(y_\alpha)}{u_\alpha(y_\alpha)}\right|^{\frac{n-2}{2}}.$$ Conditions (\[est: blow up no 2\]) and (\[est: blow up no 3\]) imply that $$\frac{\rho_\alpha}{\nu_\alpha}\to \infty$$ and $$d_g(x_\alpha,y_\alpha)\leq C_2^{\frac{2}{n-2}}\nu_\alpha.$$ It follows that the coordinates of $y_\alpha$ in the exponential chart around $x_\alpha$ defined as $\tilde{y}_\alpha:=\nu^{-1}\exp^{-1}_{x_\alpha}(y_\alpha)$ are bounded by $C_2^{\frac{2}{n-2}}$. Up to a subsequence, we may choose a finite limit $\tilde{y}_0:=\lim_{\alpha\to\infty}\tilde{y}_\alpha$. We denote $$v_\alpha(x)=\nu_\alpha^{\frac{n-2}{2}}u_\alpha\left(\exp_{x_\alpha}(\nu_\alpha x)\right)\quad\text{ and}\quad g_\alpha(x)=\left(\exp^*_{x_\alpha}g\right)\left(\exp_{x_\alpha}(\nu_\alpha x)\right)$$ for $\displaystyle x\in\Omega_\alpha:=B_0\left(\frac{\rho_\alpha}{\nu_\alpha}\right).$ As before, $g_\alpha\to\xi$ in $\mathcal{C}^2_{loc},$ $v_\alpha=O(1),$ and $\left|\frac{\nabla v_\alpha}{v_\alpha}\right|=O(1)$. By applying the same analysis as in the proof of Lemma $\ref{lem: former lemma 2}$, we get that, up to passing to a subsequence, there exists $U:=\lim_{\alpha\to\infty}v_\alpha$ in $\mathcal{C}^1_{loc}({\mathbb{R}}^n)$, with $x_0:=\lim_{\alpha\to\infty}x_\alpha$, $$\Delta U= f(x_0) U^{q-1}.$$ where $$U(x)=\left(1+\frac{f(x_0)|x-\tilde{y_0}|^2}{n(n-2)}\right)^{1-\frac{n}{2}}.$$ We know that $x_\alpha$ are local maxima for $u_\alpha$, so both $0$ and $\tilde{y}_0$ are maxima of $U$. However, since $U$ admits a unique maximum, we conclude that $\tilde{y}_0=0$.
We recall the aim of this section is to show that concentration points do not exist for the system (\[systemshort\]). So far, we have obtained a pointwise estimate (\[est: initial estimate\]) that holds everywhere on $M$ and an asymptotic profile in the neighbourhood of $x_\alpha$, a blow-up sequence; we aim to also find estimates around $x_\alpha$. We defined $(\rho_\alpha)_\alpha$ as the quantity describing the sphere of dominance of the blow-up sequence $(x_\alpha)_\alpha$. However, the influence of other blow-up sequences may be felt earlier. Let $\varphi_\alpha : (0, \rho_\alpha)\to {\mathbb{R}}^+$ be the average of $u_\alpha$ defined as $$\varphi_\alpha(r) :=\frac{1}{|\partial B_{x_\alpha}(r)|_g}\int_{\partial B_{x_\alpha}(r)} u_\alpha d\sigma_g.$$ It follows from Lemma \[lem: former lemma 4\] that $$\label{est: former lemma 4 but for averages}
(\mu_\alpha r)^{\frac{n-2}{2}}\varphi_\alpha(\mu_\alpha r)\to r^{\frac{n-2}{2}}\left(1+\frac{f(x_0) r^2}{n(n-2)}\right)^{1-\frac{n}{2}}$$ in $\mathcal{C}^1_{loc}([0,+\infty))$. We define $$\label{def: r alpha}
r_\alpha:=\sup_{r\in(2R_0\mu_\alpha,\rho_\alpha)}\left\{s^{\frac{n-2}{2}}\varphi_\alpha(s)\text{ is non-increasing in } (2R_0\mu_\alpha, r)\right\}$$ with $$R_0^2:=\frac{n(n-2)}{f(x_0)}.$$ Note that $$\text{ if }r_\alpha<\rho_\alpha, \quad\text{ then } \left(r^{\frac{n-2}{2}}\varphi_\alpha(r)\right)'(r_\alpha)=0.$$ Since $r^{\frac{n-2}{2}}U$ is non-increasing in $[2R_0,\infty)$, then (\[est: former lemma 4 but for averages\]) implies $$\label{ess: between r alpha and mu alpha}
\frac{r_\alpha}{\mu_\alpha}\to\infty.$$ so $\mu_\alpha=o(r_\alpha)$.
Considering that the asymptotic profile of blow-up sequences $(x_\alpha)_\alpha$, which is a bump function with a unique maximum, the quantity $r_\alpha$ is an indicator of the beginning of the influence of neighbouring blow-up sequences within the sphere of dominance, as the average of $u_\alpha$ is no longer decreasing.
Let $$\label{def: eta}
\eta_\alpha:=\sup_{B_{x_\alpha}(6r_\alpha)\backslash B_{x_\alpha}(\frac{1}{6}r_\alpha)}u_\alpha.$$ Note that, by Lemma \[lem: former lemma 3\], $$\frac{1}{C_3}\sup_{B_{x_\alpha}(6s_\alpha)\backslash B_{x_\alpha}\left(\frac{1}{6}s_\alpha\right)}u_\alpha\leq\varphi_\alpha(s_\alpha)\leq C_3\inf_{B_{x_\alpha}(6s_\alpha)\backslash B_{x_\alpha}\left(\frac{1}{6}s_\alpha\right)}u_\alpha$$ for $0<s_\alpha\leq r_\alpha$ and all $\alpha$. By (\[est: former lemma 4 but for averages\]), we obtain the estimate $$\label{est: using the harnack}
\lim_{R\to\infty}\limsup_{\alpha\to\infty}\sup_{B_{x_\alpha}(6r_\alpha)\backslash B_{x_\alpha}(R\mu_\alpha)}d_g(x_\alpha,x)^{\frac{n-2}{2}}u_\alpha=0.$$ Thus, $$\label{est: rel size to max}
r_\alpha^2\eta^{q-2}_\alpha\to 0$$ as $\alpha\to\infty.$ It is important to note that this implies that $$\label{ess: ok for former lemma 5}
r_\alpha\to 0$$ as $\alpha\to\infty$ since $u_\alpha\geq\varepsilon$ by (\[varepsilon\]). We now prove a pointwise asymptotic estimate for $u_\alpha$ in $B_{x_\alpha}(6r_\alpha)\backslash\{x_\alpha\}.$
\[lem: former lemma 5\] Let $(x_\alpha)_\alpha$ with $(\rho_\alpha)_\alpha$ be a blow-up sequence. Then, for any $0<\varepsilon<\frac{1}{2},$ there exists $C_\varepsilon>0$ such that $$u_\alpha(x)\leq C_\varepsilon\left(\mu_\alpha^{\frac{n-2}{2}(1-2\varepsilon)}d_g(x_\alpha,x)^{(n-2)(1-\varepsilon)}+\eta_\alpha\left(\frac{r_\alpha}{d_g(x_\alpha,x)}\right)^{(n-2)\varepsilon}\right)$$ for all $x\in B_{x_\alpha}(6r_\alpha)\backslash\{x_\alpha\}.$
Let $G$ be a Green function for the Laplace operator $\Delta_g$ on $M$ with $G> 0$. Recall the following estimates, that can be found in Aubin [@Au]: $$\label{ess: G on M}
\begin{array}{c}
\left|d_g(x_\alpha,y)^{n-2}G(x,y)-\frac{1}{(n-2)\omega_{n-1}}\right|\leq\tau\left(d_g(x,y)\right)\\
\left|d_g(x_\alpha,y)^{n-1}|\nabla G(x,y)|-\frac{1}{\omega_{n-1}}\right|\leq\tau\left(d_g(x,y)\right)
\end{array}$$ where $\tau:{\mathbb{R}}^+\to {\mathbb{R}}^+$ is a continuous function satisfying $\tau(0)=0$. For a fixed $\varepsilon$, let $$\label{def: uhm monster function i don't know it's late}
\Phi^{\varepsilon}_\alpha(x):=\mu_\alpha^{\frac{n-2}{2}(1-2\varepsilon)}G(x_\alpha,x)^{1-\epsilon}+\eta_\alpha r_\alpha^{(n-2)\varepsilon}G(x_\alpha,x)^\varepsilon$$ and let $y_\alpha\in\overline{B_{x_\alpha}(6 r_\alpha)}\backslash\{x_\alpha\}$ be such that $$\label{def of y}
\sup_{B_{x_\alpha}(6r_\alpha)}\frac{u_\alpha}{\Phi_\alpha^{\varepsilon}}=\frac{u_\alpha(y_\alpha)}{\Phi_\alpha^\varepsilon(y_\alpha)},$$ We continue by studying the following two cases, separately.
**First case:** Assume that the relative size of $d_g(x_\alpha,y_\alpha)$ with respect to $\mu_\alpha$ is $$\label{ess: position of y}
R:=\lim_{\alpha\to\infty}\frac{d_g(x_\alpha,y_\alpha)}{\mu_\alpha}\quad\text{with }R\in[0,\infty).$$ Thanks to Lemma \[lem: former lemma 4\], $$\mu_\alpha^{\frac{n-2}{2}}u_\alpha(y_\alpha)=\left(1+\frac{R^2}{R_0^2}\right)^{1-\frac{n}{2}}+o(1),$$ so that, whenever $R\in[0,\infty)$, using (\[ess: between r alpha and mu alpha\]), (\[ess: G on M\]) and (\[ess: position of y\]), it is easily shown that $$\frac{u_\alpha(y_\alpha)}{\Phi_\alpha^\varepsilon(y_\alpha)}\to((n-2)\omega_{n-1})^{1-\varepsilon}R^{(n-2)(1-\varepsilon)}\left(1+\frac{R^2}{R_0^2}\right)^{1-\frac{n}{2}}$$ as $\alpha\to\infty.$
**Second case:** It remains to study the case $$\lim_{\alpha\to\infty}\frac{d_g(x_\alpha,y_\alpha)}{\mu_\alpha}\to\infty\quad\text{ as }\alpha\to\infty.$$ If $(y_\alpha)_\alpha$ sits on the outer boundary $\partial B_{x_\alpha}(6r_\alpha)$, then by (\[ess: ok for former lemma 5\]), (\[ess: G on M\]), and (\[def: uhm monster function i don’t know it’s late\]), $$\frac{u_\alpha(y_\alpha)}{\Phi^{\varepsilon}_\alpha(y_\alpha)}\leq \left(6^{n-2}(n-2)\omega_{n-1}\right)^\varepsilon + o(1).$$ Otherwise, if up to a subsequence $y_\alpha\in B_{x_\alpha}(6r_\alpha)$, then $$\frac{\Delta_{g}u_\alpha(y_\alpha)}{u_\alpha(y_\alpha)}\geq\frac{\Delta_g\Phi^\varepsilon_\alpha(y_\alpha)}{\Phi_\alpha^{\varepsilon}(y_\alpha)}$$ as a consequence of the fact that $y_\alpha$ is the maximum of $\displaystyle\frac{u_\alpha}{\Phi_\alpha^\varepsilon}$. On the other hand, taking note of the sign of the dominant gradient term in equations (\[eq: EL alpha\]), we see that $$\label{because of the gradient sign}
\begin{array}{r l}
\Delta_g u_\alpha&= -h_\alpha u_\alpha+f_\alpha u_\alpha^{q-1}+ \frac{a_\alpha}{u_\alpha^{q+1}}-\frac{b_\alpha}{u_\alpha}-\frac{\langle\nabla u_\alpha, Y_\alpha\rangle^2}{u_\alpha^{q+3}}\\
&\quad-c_\alpha\langle\nabla u_\alpha,Y_\alpha\rangle\left[\frac{d_\alpha}{u_\alpha^2}+\frac{1}{u_\alpha^{q+2}}\right]\\
&\leq C u_\alpha^{q-1},
\end{array}$$ where $C$ is a constant depending on $\theta$ and $T$. Here we used (\[varepsilon\]). Finally, thanks to (\[est: using the harnack\]), $$d_g(x_\alpha,y_\alpha)^2\frac{\Delta_g u_\alpha(y_\alpha)}{u_\alpha(y_\alpha)}\leq C d_g(x_\alpha,y_\alpha)^2 u^{q-2}_\alpha(y_\alpha)\to 0.$$ To conclude, (\[ess: ok for former lemma 5\]) and (\[ess: G on M\]) imply that $$d_g(x_\alpha,y_\alpha)^2\frac{\Delta_g \Phi^\varepsilon_\alpha(y_\alpha)}{\Phi^\varepsilon_\alpha(y_\alpha)}=\varepsilon(1-\varepsilon)(n-2)^2+o(1).$$ We deduce that $u_\alpha(y_\alpha)=O(\Phi^\varepsilon_\alpha(y_\alpha)).$ The study of the previous two cases ends the proof of the lemma.
The following lemma improves the estimate we’ve just obtained and gives a very important bound on the size of $r_\alpha$.
\[lem: former lemma 6\] Let $(x_\alpha)_\alpha$ with $(\rho_\alpha)_\alpha$ be a blow-up sequence. Then there exists $C_4>0$ such that $$\label{est: for former lemma 6}
u_\alpha(x)+d_g(x_\alpha,x)|\nabla u_\alpha(x)|\leq C_4\mu^{\frac{n-2}{2}}_\alpha \left(d_g(x_\alpha,x)+\mu_\alpha\right)^{2-n}$$ for all $x\in B_{x_\alpha}(6r_\alpha)\backslash\{x_\alpha\}.$ Moreover, $r_\alpha^2=O(\mu_\alpha)$.
It suffices to prove the estimate for $u_\alpha$; the rest follows as an immediate consequence of Lemma \[lem: former lemma 3\]. We start by showing that for any sequence $z_\alpha\in \overline{B_{x_\alpha}(6r_\alpha)}\backslash\{x_\alpha\}$, there holds $$\label{est: form lem 6 most of the work}
u_\alpha(z_\alpha)=O\left(\mu_\alpha^{\frac{n-2}{2}}d_g(x_\alpha,z_\alpha)^{2-n}+\eta_\alpha\right).$$ First, if $d_g(x_\alpha,z_\alpha)=O(\mu_\alpha)$, it falls within the range described in Lemma \[lem: former lemma 4\]. On the other hand, when $r_\alpha=O\left(d_g(x_\alpha,z_\alpha)\right),$ we use Lemma \[lem: former lemma 3\] together with (\[def: eta\]). It remains to consider the intermediary case: $$\frac{d_g(x_\alpha,z_\alpha)}{\mu_\alpha}\to\infty\quad\text{ and }\quad\frac{d_g(x_\alpha,z_\alpha)}{r_\alpha}\to 0\quad\text{ at }\alpha\to\infty.$$ According to the Green representation formula, $$u_\alpha(z_\alpha)=O\left(\int_{B_{x_\alpha}(6r_\alpha)}d_g(z_\alpha,x)^{2-n}\Delta_g u_\alpha(x)\,dv_g\right)+O(\eta_\alpha),$$ where the second term corresponds to the boundary element. Recall that $$\Delta_g u_\alpha\leq C u_\alpha^{q-1}$$ because of the sign of the dominant gradient term, see (\[because of the gradient sign\]). Using (\[varepsilon\]), (\[est: rel size to max\]) and Lemma \[lem: former lemma 5\], we can write that $$\begin{array}{l}
\int_{B_{x_\alpha}(6r_\alpha)}d_g(z_\alpha,x)^{2-n}u_\alpha^{q-1}(x)\,dv_g\\
\displaystyle =O\left(\mu_\alpha^{\frac{n}{2}-1}d_g(x_\alpha,z_\alpha)^{2-n}\right)\\
+O\left(\mu_\alpha^{\frac{n+2}{2}(1-2\varepsilon)}\int_{B_{x_\alpha}(6r_\alpha)\backslash B_{x_\alpha}(\mu_\alpha)}d_g(z_\alpha,x)^{2-n}d_g(x_\alpha,x)^{-(n+2)(1-\varepsilon)}\,dv_g\right)\\
+O\left(\eta_\alpha^{q-1}r_\alpha^{(n+2)\varepsilon}\int_{B_{x_\alpha}(6r_\alpha)\backslash B_{x_\alpha}(\mu_\alpha)}d_g(z_\alpha,x)^{2-n}d_g(x_\alpha,x)^{-(n+2)\varepsilon}\,dv_g\right)\\
= O\left(\mu^{\frac{n}{2}-1}_\alpha d_g(x_\alpha,z_\alpha)^{2-n}\right)+O\left(\eta_\alpha^{q-1}r_\alpha^2\right)\\
= O\left(\mu^{\frac{n}{2}-1}_\alpha d_g(x_\alpha,z_\alpha)^{2-n}\right)+O\left(\eta_\alpha\right).
\end{array}$$ In order to get estimate (\[est: for former lemma 6\]), it suffices to show that $$\label{est: of eta}
\eta_\alpha=O\left(\mu_\alpha^{\frac{n-2}{2}}r_\alpha^{2-n}\right).$$ For any fixed $0<\delta<1$, taking $\alpha$ large enough, then the monotonicity of $r^{\frac{n-2}{2}}\varphi_\alpha(r)$ expressed in the definition of $r_\alpha$, see (\[def: r alpha\]), and the fact that $\mu_\alpha=o(r_\alpha)$, see (\[ess: between r alpha and mu alpha\]), imply that $$r_\alpha^{\frac{n-2}{2}}\varphi_\alpha(r_\alpha)\leq (\delta r_\alpha)^{\frac{n-2}{2}}\varphi_\alpha(\delta r_\alpha)\quad\text{ for all }0<\delta<1,$$ so by Lemma \[lem: former lemma 3\], $$\frac{1}{C_3}\eta_\alpha\leq \delta^{\frac{n-2}{2}}\sup_{\partial B_{x_\alpha}(\delta r_\alpha)}u_\alpha.$$ According to estimate (\[est: form lem 6 most of the work\]), this leads to $$\eta_\alpha\leq C\left(\mu_\alpha^{\frac{n-2}{2}}\delta^{2-n}r_\alpha^{2-n}+\eta_\alpha\right)\delta^{\frac{n-2}{2}},$$ where $C$ is independent of $\delta$ and $\alpha$. Choosing $\delta$ small enough leads to (\[est: of eta\]).
Estimates (\[varepsilon\]) and (\[est: of eta\]) imply that $$\label{size of r alpha}
r_\alpha^2= O(\mu_\alpha).$$ This ends the proof of the lemma.
Given a blow-up sequence $(x_\alpha)_\alpha$ with $(\rho_\alpha)_\alpha$, the following lemma gives the exact asymptotic profile of $(u_\alpha)_\alpha$ at distance $(\rho_\alpha)_\alpha$ of $(x_\alpha)_\alpha$.
\[lem: former lemma 7\] Let $(x_\alpha)_\alpha$ and $(\rho_\alpha)_\alpha$ be a blow-up sequence. Then we have that $r_\alpha=\rho_\alpha$, where $r_\alpha$ is as in (\[def: r alpha\]). Up to a subsequence, we have $$\label{est final profile in r alpha}
u_\alpha(x_\alpha)\rho^{n-2}_\alpha u_\alpha(\exp_\alpha(\rho_\alpha x))\to\frac{R_0^{n-2}}{|x|^{n-2}}+H(x)$$ in $C^2_{loc}(B_0(5)\backslash\{0\})$, where $H$ is some harmonic function in $B_0(5)$ satisfying $H(0)=0.$
First, we prove that, up to a subsequence, $$\label{prefinal profile}
u_\alpha(x_\alpha)r^{n-2}_\alpha u_\alpha(\exp_\alpha(r_\alpha x))\to\frac{R_0^{n-2}}{|x|^{n-2}}+H(x).$$ Let us define the following rescaled quantities: $$\hat{u}_\alpha(x)=\mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n-2}u_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)\text{ and }\hat{g}_\alpha(x)=\exp_{x_\alpha}^*g\left(\exp_{x_\alpha}(r_\alpha x)\right).$$ Then $$\Delta_{\hat{g}_\alpha}\hat{u}_\alpha=\hat{F}_\alpha,$$ in $B_0(\delta r_\alpha^{-1})$ for some $\delta>0$ small enough, with $$\label{eq: form lemma 7's hat F}
\begin{array}{r l}
\hat{F}_\alpha=& \displaystyle -\mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n}h_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)u_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right) \\
& +\mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n}f_\alpha \left(\exp_{x_\alpha}(r_\alpha x)\right)u_\alpha^{q-1}\left(\exp_{x_\alpha}(r_\alpha x)\right) \\
& +\mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n}\frac{a_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)}{u^{q+1}_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)} -\mu_\alpha^{1-\frac{n}{2}}r_\alpha^n\frac{b_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)}{u_\alpha \left(\exp_{x_\alpha}(r_\alpha x)\right)}\\
& -\mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n} \frac{\langle\nabla u_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right), Y_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)\rangle^2}{u_\alpha^{q+3}\left(\exp_{x_\alpha}(r_\alpha x)\right)}\\
& -\mu_\alpha^{1-\frac{n}{2}}r_\alpha^n c_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)\langle\nabla u_\alpha,Y_\alpha\rangle\left(\exp_{x_\alpha}(r_\alpha x)\right) \Big[ \frac{d_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)}{u_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)^2}\\
& + \frac{1}{u_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)^{q+2}}\Big]\\
\end{array}$$ Thanks to Lemma \[lem: former lemma 6\], we know that $|\hat{u}_\alpha|\leq C_K$ and $|\hat{F}_\alpha|=O(1)$ on any compact $K\subset B_0(5)\backslash\{0\}$. By standard elliptic theory, up to a subsequence, $$\hat{u}_\alpha\to\hat{U}\text{ in }\mathcal{C}^1_{loc}(B_0(5)\backslash\{0\})$$ with $$\Delta_\xi\hat{U}=0\text{ in }B_0(5)\backslash\{0\}.$$ Separate $\hat{U}$ into the sum of a regular harmonic function and a singular part $$\hat{U}=\frac{\lambda}{|x|^{n-2}}+H(x),$$ where $\lambda\geq 0.$
To get (\[prefinal profile\]), it remains to show that $\lambda=R_0^{n-2}.$ For any $\delta>0$: $$\label{eq: lemma 7 minutia}
\int_{B_0(\delta)}\hat{F}_\alpha\,dv_{\hat{g}_\alpha}=-\int_{\partial B_0(\delta)}\partial_\nu\hat{u}_\alpha\,d\sigma_{\hat{g}_\alpha}.$$ Using the equation (\[eq: form lemma 7’s hat F\]), we estimate the left hand side of (\[eq: lemma 7 minutia\]). In particular, $$\begin{array}{c}
\displaystyle \int_{B_0(\delta)}f_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)\mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n}u_\alpha^{q-1}\left(\exp_{x_\alpha}(r_\alpha x)\right)\,dx \\ \\
\displaystyle =\int_{B_0(\delta\frac{r_\alpha}{\mu_\alpha})}f_\alpha\left(\exp_{x_\alpha}(\mu_\alpha z)\right)\mu_\alpha^{\frac{n+2}{2}}u_\alpha^{q-1}\left(\exp_{x_\alpha}(\mu_\alpha z)\right)\,dz
\end{array}$$ where $\displaystyle z=\frac{r_\alpha}{\mu_\alpha}x$, and by Lemma \[lem: former lemma 4\] and Lemma \[lem: former lemma 6\], $$\begin{array}{c}
\displaystyle \lim_{\alpha\to\infty}\int_{B_0(\delta)}f_\alpha\left(\exp_{x_\alpha}(r_\alpha x)\right)\mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n}u_\alpha^{q-1}\left(\exp_{x_\alpha}(r_\alpha x)\right)\,dx \\ \\
\displaystyle =f(x_0)\int_{R^n}\left(1+\frac{|x|^2}{R_0^2}\right)^{-1-\frac{n}{2}}\,dx.
\end{array}$$ The gradient terms are controlled with the estimate (\[est: initial estimate\]), and together with (\[varepsilon\]), we obtain that the dominant gradient term of (\[eq: form lemma 7’s hat F\]) verifies $$\label{es 1}
\begin{array}{r l}
\displaystyle \int_{B_0(\delta)}\frac{|\nabla u_\alpha|^2}{u_\alpha^{q+3}}\left(\exp_{x_\alpha}(r_\alpha x)\right)\mu_\alpha^{1-\frac{n}{2}}r_\alpha^n\,dx\leq
& \displaystyle C \mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n-2}\int_{B_0(\delta)}|x|^{-2}\,dx \\ \\
& \leq \displaystyle C\omega_{n-1} \mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n-2}\delta^{n-2}.
\end{array}$$ As $\mu_\alpha^{1-\frac{n}{2}}r_\alpha^{n-2}=O(1),$ the integral does not vanish as $\alpha\to\infty$; its size depends on $\delta$. The remaining terms in (\[eq: form lemma 7’s hat F\]) are negligible. Thus $$\int_{B_0(\delta)}\hat{F}_\alpha\, dv_{\hat{g}_\alpha}= f(x_0)\int_{R^n}\left(1+\frac{|x|^2}{R_0^2}\right)^{-1-\frac{n}{2}}\,dx+o(1)+O(\delta^{n-2})$$ for any $\delta>0$. It follows that $$f(x_0)\int_{{\mathbb{R}}^n}\left(1+\frac{|x|^2}{R_0^2}\right)^{-1-\frac{n}{2}}\,dx=(n-2)\omega_{n-1}R_0^{n-2}.$$ Note also that the right hand side of (\[eq: lemma 7 minutia\]) verifies $$\begin{array}{r l}
-\int_{\partial B_0(\delta)}\partial_\nu \hat{u}_\alpha\,d\sigma_{\check{g}_\alpha}&=-\int_{\partial B_0(\delta)}\partial_\nu\hat{U}+o(1)\\
&=\lambda(n-2)\omega_{n-1}+o(1).
\end{array}$$ since $H$ is smooth and harmonic. Since $$\lambda(n-2)\omega_{n-1}=R^{n-2}_0(n-2)\omega_{n-1}+O(\delta^{n-2})+o(1),$$ for any $\delta>0$, we get that $\lambda=R_0^{n-2}.$
Finally, let us prove that $H(0)=0.$ The equation’s dominant terms are invariant by rescaling, which leads us to use a Pohozaev identity to obtain new estimates for the remaining terms. Let $\Omega_\alpha$ correspond to $B_0(\delta r_\alpha)$ in the exponential chart at $x_\alpha\in M$ and let $X_\alpha=\frac{1}{2}\nabla d_g(x_\alpha,x)^2$ be the vector field of coordinates. Using integration by parts, $$\begin{array}{r l}
\int_{\Omega_\alpha}\nabla u_\alpha(X_\alpha)\Delta_gu_\alpha\,dv_g=& \int_{\Omega_\alpha}\langle\nabla \left(\nabla u_\alpha(X_\alpha)\right),\nabla u_\alpha\rangle\, dv_g\\
& - \int_{\partial \Omega_\alpha}\nabla u_\alpha(X_\alpha)\partial_{\nu}u_\alpha\,d\sigma_g\\
=
& \int_{\Omega_\alpha}\nabla^\#\nabla u_\alpha( X_\alpha, \nabla u_\alpha) + \nabla^\# X_\alpha\left(\nabla u_\alpha, \nabla u_\alpha\right)\,dv_g\\
& - \int_{\partial \Omega_\alpha}\nabla u_\alpha(X_\alpha)\partial_{\nu}u_\alpha\,d\sigma_g,
\end{array}$$ where $(\nabla^\# X_\alpha)=(\nabla^i X_\alpha)^j$. Since $$\begin{array}{c}
\int_{\Omega_\alpha}|\nabla u_\alpha|^2div_g X_\alpha\,dv_g+\int_{\Omega_\alpha}\langle\nabla\left(|\nabla u_\alpha|^2\right),X_\alpha\rangle\,dv_g\\
=\int_{\partial \Omega_\alpha}|\nabla u_\alpha|^2 \langle X_\alpha, \nu\rangle \,d\sigma_g,
\end{array}$$ we can write that $$\label{eq: Pohozaev}
\begin{array}{r l}
\int_{\Omega_\alpha}
&\displaystyle\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)\Delta_g u_\alpha\,dv_g \\
& =\int_{\Omega_\alpha}\left(\nabla^\# X_\alpha\left(\nabla u_\alpha, \nabla u_\alpha\right)-\frac{1}{2}\left(div_g X_\alpha\right)|\nabla u_\alpha|^2\right)\,dv_g\\
& + \int_{\partial\Omega_g}\left(\frac{1}{2}\langle X_\alpha,\nu\rangle |\nabla u_\alpha|^2-\nabla u_\alpha(X_\alpha)\partial_\nu u_\alpha-\frac{n-2}{2}u_\alpha\partial_\nu u_\alpha\right)\,d\sigma_g.
\end{array}$$ We begin by analyzing the right-hand side of (\[eq: Pohozaev\]). By our choice of $X_\alpha$, $(\nabla^\# X_\alpha)^{ij}=g^{ij}+O(d_g(x_\alpha,x)^2),$ and consequently $$\begin{array}{rl}
\int_{\Omega_\alpha} &\displaystyle \left(\nabla^\# X_\alpha\left(\nabla u_\alpha, \nabla u_\alpha\right)-\frac{1}{2}\left(div_g X_\alpha\right)|\nabla u_\alpha|^2\right)\,dv_g \\
& =O\left(\int_{\Omega_\alpha}d_g(x_\alpha,x)^2|\nabla u_\alpha|^2dv_g\right).
\end{array}$$ According to (\[est: for former lemma 6\]), $$\begin{array}{c}
\displaystyle \int_{\Omega_\alpha}d_g(x_\alpha,x)^2|\nabla u_\alpha|^2\,dv_g\leq C\int_{\Omega_\alpha}\mu_\alpha^{n-2}\left(d_g(x_\alpha,x)+\mu_\alpha\right)^{4-2n}\,dv_g,
\end{array}$$ so $$\displaystyle \int_{\Omega_\alpha}d_g(x_\alpha,x)^2|\nabla u_\alpha|^2\,dv_g \leq\left\{ \begin{array}{l c}
\displaystyle O(\mu_\alpha r_\alpha) & \displaystyle \text{ if }n=3 \\ \\
\displaystyle O\left(\mu_\alpha^2\ln\frac{1}{\mu_\alpha}\right) & \displaystyle \text{ if }n=4 \\ \\
\displaystyle O\left(\mu_\alpha^2\right) & \displaystyle \text{ if }n=5
\end{array}\right.$$ In all these three cases, thanks to (\[size of r alpha\]), the integral is of the order $o(\mu_\alpha^{n-2} r_\alpha^{2-n})$. From (\[est final profile in r alpha\]), $$\begin{array}{r l}
\displaystyle \int_{\partial\Omega_\alpha} & \displaystyle \left(\frac{1}{2}\langle X_\alpha,\nu\rangle |\nabla u_\alpha|^2-\nabla u_\alpha(X_\alpha)\partial_\nu u_\alpha-\frac{n-2}{2}u_\alpha\partial_\nu u_\alpha\right)\,d\sigma_g \\
&=\left(\frac{(n-2)^2}{2}\omega_{n-1}R_0^{n-2}H(0)+o(1)\right)\mu_\alpha^{n-2}r_\alpha^{2-n}.
\end{array}$$ Note that the boundary term does not depend on $\delta$, and as a result $$\label{eq: pohozaev id 1 in lem 7 former}
\begin{array}{r l}
\int_{\Omega_\alpha}&\displaystyle\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)\Delta_g u_\alpha\,dv_g \\
& =\left(\frac{(n-2)^2}{2}\omega_{n-1}R_0^{n-2}H(0)+o(1)\right)\mu_\alpha^{n-2}r_\alpha^{2-n}
\end{array}$$ We now analyse the right hand side of (\[eq: Pohozaev\]) by using (\[eq: EL alpha\]): $$\label{eq: from the eq}
\begin{array}{r l}
\int_{\Omega_\alpha}
&\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)\Delta_g u_\alpha\,dv_g \\
=& \int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)f_\alpha u_\alpha^{q-1}\,dv_g\\
& - \int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)\langle\nabla u_\alpha, Y_\alpha\rangle^2 u_\alpha^{-q-3}\,dv_g\\
& - \int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)c_\alpha\langle\nabla u_\alpha, Y_\alpha\rangle(d_\alpha u_\alpha^{-2}+u_\alpha^{-q-2})\,dv_g\\
& + \int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)\left(a_\alpha u_\alpha^{-q-1}-b_\alpha u_\alpha^{-1}- h_\alpha u_\alpha\right)\,dv_g
\end{array}$$ and we look at each term in turn. By the estimates (\[varepsilon\]) and (\[est: initial estimate\]), we get $$\label{for dominant nabla term}
\begin{array}{r l}
\left|\int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)\langle\nabla u_\alpha,Y_\alpha\rangle^2 u_\alpha^{-q-3}\,dv_g \right|
& \leq C\int_{\Omega_\alpha}d_g(x_\alpha,x)^{-2}dv_g\\
& \leq C(\delta r_\alpha)^{n-2}
\end{array}$$ and, similarly, $$\label{for nondominant nabla term}
\begin{array}{c}
\left|\int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)c_\alpha\langle\nabla u_\alpha, Y_\alpha\rangle(d_\alpha u_\alpha^{-2}+u_\alpha^{-q-2})\,dv_g\right|\\
\leq C(\delta r_\alpha)^{n-1}.
\end{array}$$ We also have that $$\label{for a and b}
\left|\int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)\left(a_\alpha u_\alpha^{-q-1}-b_\alpha u_\alpha^{-1}\right)\right|\leq Cr_\alpha^n.$$ From (\[est: for former lemma 6\]), we obtain that $$\label{for h}
\begin{array}{c}
\left|\int_{\Omega_\alpha}h_\alpha\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)u_\alpha\,dv_g\right|\\=O\left(\int_{\Omega_\alpha}\mu_\alpha^{n-2}(\mu_\alpha+d_g(x_\alpha,x))^{4-2n}\,dx\right)\\
=o(\mu_\alpha^{n-2}r_\alpha^{2-n})
\end{array}$$ for $3\leq n\leq 5.$ Using integration by parts, $$\begin{array}{r l}
\int_{\Omega_\alpha}\nabla u_\alpha(X_\alpha)f_\alpha u_\alpha^{q-1}\,dv_g=&\frac{1}{q} \int_{\partial \Omega_\alpha}f_\alpha r_\alpha u_\alpha^q\,d\sigma_g\\
& -\frac{1}{q}\int_{\Omega_\alpha}div_g X_\alpha f_\alpha u_\alpha^q\, dv_g \\ &- \frac{1}{q}\int_{\Omega_\alpha}\nabla f_\alpha(X_\alpha)u_\alpha^q\,dv_g.
\end{array}$$ Thus we can write that $$\begin{array}{r l}
\int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)f_\alpha u_\alpha^{q-1}\,dv_g&=\frac{1}{q}r_\alpha\int_{\partial \Omega_\alpha}f_\alpha u_\alpha^q\,d\sigma_g\\
&\,+\int_{\Omega_\alpha}\left(-\frac{1}{q}div_g(X_\alpha)+\frac{n-2}{2}\right)f_\alpha u_\alpha^q\,dv_g\\
&\,-\frac{1}{q}\int_{\Omega_\alpha}\nabla f_\alpha(X_\alpha)u_\alpha^q\,dv_g.
\end{array}$$ Since $div_g(X_\alpha)=n+O\left(d_g(x_\alpha,x)^2\right)$, this leads to $$\begin{array}{r l}
\int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)f_\alpha u_\alpha^{q-1}\,dv_g&=\frac{1}{q}r_\alpha\int_{\partial \Omega_\alpha}f_\alpha u_\alpha^q\,d\sigma_g\\
&\,+O\left(\int_{\Omega_\alpha} d_g(x_\alpha,x)^2 u_\alpha^q\,dv_g\right)\\
&\,-\frac{1}{q}\int_{\Omega_\alpha}\nabla f_\alpha(X_\alpha)u_\alpha^q\,dv_g.
\end{array}$$ Using lemmas \[lem: former lemma 4\] and \[lem: former lemma 6\], this leads to $$\begin{array}{r l}
\int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)f_\alpha u_\alpha^{q-1}\,dv_g&= O\left(\mu_\alpha^n r_\alpha^{-n}\right)+O(\mu_\alpha^2)\\
&-\frac{1}{q}\int_{\Omega_\alpha}\nabla f_\alpha(X_\alpha)u_\alpha ^q\,dv_g
\end{array}$$ so that, thanks to (\[size of r alpha\]), $$\label{for f}
\begin{array}{c}
\int_{\Omega_\alpha}\left(\nabla u_\alpha(X_\alpha)+\frac{n-2}{2}u_\alpha\right)f_\alpha u_\alpha^{q-1}\,dv_g=o\left(\mu_\alpha^{n-2}r_\alpha^{2-n}\right)\\-\frac{1}{q}\int_{\Omega_\alpha}\nabla f_\alpha(X_\alpha)u_\alpha^q\,dv_g
\end{array}$$ if $3\leq n\leq 5.$
We claim that $$\label{for nabla f}
\int_{\Omega_\alpha}\nabla f_\alpha (X_\alpha)u_\alpha^q\,dv_g=o\left(\mu_\alpha^{n-2}r_\alpha^{2-n}\right).$$ Thanks to (\[eq: pohozaev id 1 in lem 7 former\]), (\[for dominant nabla term\]), (\[for nondominant nabla term\]), (\[for a and b\]), (\[for h\]), (\[for f\]) and (\[for nabla f\]), we see that $$\label{Hdelta}
H(0)=o(1)+\delta^4$$ for any $\delta>0$, so by taking $\delta\to 0$ wee see that $H(0)=0$.
In order to prove (\[for nabla f\]), we can first use Lemma \[lem: former lemma 6\] to write that $$\int_{\Omega_\alpha}\nabla f_\alpha(X_\alpha)u_\alpha^q\,dv_g=O(\mu_\alpha)$$ which leads to (\[for nabla f\]) if $n=3,4$ thanks to (\[size of r alpha\]), but is not enough for $n=5$. In order to improve the estimate in the case $n=5$, note that $$\begin{array}{r l}
\displaystyle \int_{\Omega_\alpha}\nabla f_\alpha(X_\alpha)u_\alpha^q\,dv_g
&\displaystyle =\partial_i f(x_\alpha)\int_{\Omega_\alpha}x^iu_\alpha^q\,dv_g\\ \\
&\displaystyle\quad +O\left(\int_{\Omega_\alpha}d_g(x_\alpha,x)^{1+\eta}u_\alpha^q\,dv_g\right)\\ \\
&\displaystyle =o\left(\mu_\alpha|\nabla f_\alpha(x_\alpha)|\right)+ O\left(\mu_\alpha^{1+\eta}\right)\\ \\
&\displaystyle =o\left(\mu_\alpha|\nabla f_\alpha(x_\alpha)|\right)+ o\left(\mu_\alpha^{3}r_\alpha^{-3}\right).
\end{array}$$ with $\eta>\frac{1}{2}.$ Thus it remains to prove that $$\label{est: gjgjhvj}
|\nabla f_\alpha(x_\alpha)|=O\left(\mu_\alpha^2 r_\alpha^{-3}\right).$$ As before, we use a Pohozaev-type identity. We make use of the equation’s symmetry by translation, with $Z=Z^i$ a constant vector field in the exponential chart of $x_\alpha$. We can write that $$\label{newpoz}
\int_{\Omega_\alpha}\nabla u_\alpha(Z_\alpha)\Delta_g u_\alpha\,dv_g=O\left(\int_{\Omega_\alpha}d_g(x_\alpha,x)|\nabla u_\alpha|^2\,dv_g +\int_{\partial \Omega_\alpha}|\nabla u_\alpha|^2\,d\sigma_g\right),$$ which is $o(\mu_\alpha^2 r_\alpha^{-3})$. On the left-hand side, we use (\[eq: EL alpha\]). Lemma \[lem: former lemma 3\] and (\[varepsilon\]) imply that $$\label{newpoz1}
\begin{array}{r l}
\displaystyle \int_{\Omega_\alpha}\nabla u_\alpha(Z_\alpha)\frac{\langle\nabla u_\alpha,Y_\alpha\rangle^2}{u_\alpha^{q+3}}
\,dv_g &\displaystyle \leq \frac{1}{\varepsilon^q}\int_{\Omega_\alpha}\left|\frac{\nabla u_\alpha}{u_\alpha}\right|^3\,dv_g\\ \\
&\displaystyle
\leq \frac{1}{\varepsilon^q}\int_{\Omega_\alpha}d_g(x_\alpha,x)^{-3}\,dv_g\\ \\
&\displaystyle =O(r^2_\alpha)=o(\mu_\alpha^2r_\alpha^{-3}).
\end{array}$$ We see that $$\label{newpoz2}
\int_{\Omega_\alpha}\nabla u_\alpha(Z_\alpha)c_\alpha\langle\nabla u_\alpha, Y_\alpha\rangle\left(\frac{d_\alpha}{u_\alpha^2}+\frac{1}{u_\alpha^{q+2}}\right)
\,dv_g=O(r^2_\alpha)=o(\mu_\alpha^2r_\alpha^{-3}),$$ and that the same holds for the terms corresponding to $h_\alpha$, $b_\alpha$ and $c_\alpha$. So (\[newpoz\]), (\[newpoz1\]) and (\[newpoz2\]) imply that $$\int_{\Omega_\alpha}\nabla u_\alpha(Z_\alpha)f_\alpha u_\alpha^{q-1}\,dv_g=o(\mu_\alpha^2 r_\alpha^{-3}).$$ Furthermore, $$\begin{array}{r l}
\displaystyle \int_{\Omega_\alpha}\nabla u_\alpha(Z_\alpha)f_\alpha u_\alpha^{q-1}\,dv_g=
&\displaystyle O\left(\int_{\partial\Omega_\alpha}u_\alpha^q\,d\sigma_g\right)\\ \\
&\displaystyle-\frac{1}{q}\int_{\Omega_\alpha}div_g(Z_\alpha)f_\alpha u_\alpha^q\,dv_g\\ \\
&\displaystyle -\frac{1}{q}\nabla f_\alpha(Z_\alpha)\int_{\Omega_\alpha}u_\alpha^q\,dv_g,
\end{array}$$ which leads us to conclude the claim in (\[for nabla f\]). Note that $$div_g(Z_\alpha)=O\left(d_g(x_\alpha,x)^\eta\right)$$ and $$\int_{\Omega_\alpha}u_\alpha^q\,dv_g\rightarrow\int_{R^n}\left(1+\frac{|x|^2}{R_0^2}\right)^{-5}\,dx<\infty.$$
Finally, we are in the position to remark that $\rho_\alpha=r_\alpha.$ Remember that $$\varphi(r)=\frac{1}{\omega_{n-1}r^{n-1}}\int_{\partial B_0(r)}\hat{U}=\left(\frac{R_0}{r}\right)^{n-2}+H(0)$$ and that $\left(r^{\frac{n-2}{2}}\varphi(r)\right)'(1)=0$, so if $r_\alpha<\rho_\alpha$, then $H(0)=R_0^{n-2}$, which contradicts (\[Hdelta\]). Thus (\[prefinal profile\]) implies (\[est final profile in r alpha\]), and this wraps up the proof of the lemma.
Moreover, $\rho_\alpha=r_\alpha$ means that $\rho_\alpha\to 0$ because $r_\alpha=O\left(\mu_\alpha^\frac{1}{2}\right)$ thanks to (\[size of r alpha\]). As an important consequence, there do not exist any isolated bubbles. Otherwise, if a bubble were isolated, then we could choose a blow-up sequence with $0<\delta<\rho_\alpha$, contradicting the previous result.
Proof of the stability theorem
------------------------------
We are now in the position to prove Theorem \[theo stability eq\]. Let $$\delta_\alpha:=\min_{1\leq i<j\leq N_\alpha}d_g(x_{i,\alpha}, x_{j,\alpha}).$$
For any $R>0$, let $1\leq M_{R,\alpha}$ be such that $$d_g(x_{1,\alpha}, x_{i_\alpha,\alpha})\leq R\delta_\alpha\quad\text{ for }\quad i_\alpha\in\{1,\dots,M_{R,\alpha}\}, \text{ and}$$ $$d_g(x_{1,\alpha}, x_{j_\alpha,\alpha})> R\delta_\alpha \quad\text{ for }\quad j_\alpha\in\{M_{R,\alpha}+1,\dots, N_\alpha\}.$$ We consider the rescaled quantities $$\check{u}_\alpha(x):=\delta_\alpha^{\frac{n-2}{2}}u_\alpha(\exp_{x_{1,\alpha}}(\delta_\alpha x))\quad\text{ and }\quad \check{g}_\alpha(x):=\left(\exp^*_{x_{1,\alpha}}g\right)(\delta_\alpha x)$$ and the coordinates $\check{x}_{i,\alpha}:=\delta_\alpha^{-1}\exp_{x_{1,\alpha}}^{-1}(x_{i,\alpha})$ in the exponential chart. It’s obvious that $|\check{x}_{2,\alpha}|=1$ and $|\check{x}_{i,\alpha}|\geq 1$.
The following lemma is a direct consequence of Lemma \[lem: former lemma 3\].
\[lem former lemma 9\] For all $R>0$, there exists $C_R>0$ such that the Harnack-type inequality $$||\nabla \check{u}_\alpha||_{L^\infty(\Omega_R)}\leq C_R \sup_{\Omega_R}\check{u}_\alpha\leq C_R^2\inf_{\Omega_R}\check{u}_\alpha$$ holds, where $\Omega_R=B_0(R)\backslash\bigcup^{M_{2R,\alpha}}_{i=1}B_{\check{x}_{i,\alpha}}\left(\frac{1}{R}\right).$
Note that, for $1\leq i<j\leq M_{R,\alpha}$, $B_{x_{i,\alpha}}\left(\frac{\delta_\alpha}{4}\right)$ and $B_{x_{j,\alpha}}\left(\frac{\delta_\alpha}{4}\right)$ are disjoint, which is equivalent to saying that $B_{\check{x}_{i,\alpha}}\left(\frac{1}{4}\right)$ and $B_{\check{x}_{j,\alpha}}\left(\frac{1}{4}\right)$ are also disjoint.
At this point, we are finally able to prove Theorem \[theo stability eq\], which we stated at the very beginning of this section. We define two possible types of concentration points, according to how $\check{u}_\alpha$ explodes. We prove that, within a cluster, we can only find one type or the other, but never both. Finally, we see that the existence of either type leads to contradictions, which implies that $\check{u}_\alpha$ admits no concentration points whatsoever.
Consider the cluster around $(x_{1,\alpha})_\alpha$, for some $R>0$. There are two possible cases. The first type of concentration point corresponds to $$\label{first case concentration point}
\sup_{B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right)}\left(\check{u}_\alpha(x)+\left|\frac{\nabla\check{u}_\alpha(x)}{\check{u}_\alpha(x)}\right|^{\frac{n-2}{2}}\right)=O(1).$$ In this case, note that $(\check{u}_\alpha)_\alpha$ is uniformly bounded in $\mathcal{C}^1_{loc}$. Moreover, we find a lower bound, as by (\[est: 1 of former lemma 1\]) from Lemma \[lem: former lemma 2\], $$|\check{x}_{i,\alpha}|^{\frac{n-2}{2}}\check{u}_\alpha(\check{x}_{i,\alpha})\geq 1.$$ There exists $\delta_i>0$ such that $$\inf_{B_{\check{x}_{i,\alpha}}(\delta_i)}\check{u}_\alpha\geq \frac{1}{2}|\check{x}_{i,\alpha}|^{1-\frac{n}{2}},$$ which leads to the existence of $\delta_0>0$ where $$\label{lower bound}
\inf_{B_{\check{x}_{i,\alpha}}(\delta_0)}\check{u}_\alpha\geq \frac{1}{2}.$$ The second type is defined by $$\label{second case concentration point}
\sup_{B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right)}\left(\check{u}_\alpha(x)+\left|\frac{\nabla\check{u}_\alpha(x)}{\check{u}_\alpha(x)}\right|^{\frac{n-2}{2}}\right)\to\infty.$$ In this case, either $$\label{nabla explodes}
\sup_{B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right)}\check{u}_\alpha(x)\leq M\quad \text{ and }\quad \sup_{B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right)}|\nabla\check{u}_\alpha(x)|\to\infty,$$ or $$\label{l infty explodes}
\sup_{B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right)}\check{u}_\alpha(x)\to\infty.$$ We show (\[nabla explodes\]) is not actually possible. Assume it holds true. Then there exist $(\check{x}_\alpha)_\alpha\subset (B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right))_\alpha$ and $(\check{\nu}_\alpha)_\alpha$ such that $$\check{\nu}_\alpha^{1-\frac{n}{2}}:= \check{u}_\alpha(\check{x}_\alpha)+\left|\frac{\nabla \check{u}_\alpha(\check{x}_\alpha)}{\check{u}_\alpha(\check{x}_\alpha)}\right|^{\frac{n-2}{2}}=\sup_{x\in B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right)}\left(\check{u}_\alpha(x)+\left|\frac{\nabla \check{u}_\alpha(x)}{\check{u}_\alpha(x)}\right|^{\frac{n-2}{2}}\right)$$ with $$\check{\nu}_\alpha\to 0.$$ We define the rescaled quantities $$\check{v}_\alpha(x):=\check{\nu}_\alpha^{\frac{n-2}{2}}\check{u}_\alpha\left(\exp_{\check{x}_\alpha}(\check{\nu}_\alpha x)\right)\quad\text{and}\quad \check{h}_\alpha(x):=\left(\exp_{\check{x}_\alpha}^* \check{g}\right)(\check{\nu}_\alpha x)$$ respectively, defined in $\Omega_\alpha:=B_0\left(\frac{1}{2\check{\nu}_\alpha}\right)$. For any $R>0$ and $\alpha$ large enough so that $R<\frac{1}{2\check{\nu}_\alpha}$, $$\limsup_{\alpha\to\infty}\sup_{B_0(R)} \left(\check{v}_\alpha+\left|\frac{\nabla \check{v}_\alpha}{\check{v}_\alpha}\right|\right)=1.$$ Thus $$|\nabla \ln \check{v}_\alpha|\leq 1$$ and $$\check{v}_\alpha(0)e^{-x}\leq \check{v}_\alpha(x)\leq \check{v}_\alpha(0)e^{x}.$$ Note that the metrics $\check{h}_\alpha\to\xi$ in $\mathcal{C}^2_{loc}$ as $\alpha\to\infty$. Assume that, up to a subsequence, $u_\alpha(\check{x}_\alpha)\to l<\infty$. We also deduce that $\check{v}_\alpha(0)\to 0.$ Let $\check{x}_0:=\lim_{\alpha\to\infty}\check{x}_\alpha$ and let us denote $$\check{w}_\alpha(x):=\frac{\check{v}_\alpha(x)}{\check{v}_\alpha(0)}.$$ These functions are bounded from below, $$\label{est w again}
\check{w}_\alpha(x)\geq\frac{\varepsilon}{l}+o(1)>0.$$ Moreover, $$\check{w}_\alpha(x)\leq e^{|x|}.$$ By standard elliptic theory, we find that there exists $\check{w}:=\lim_{\alpha\to\infty}\check{w}_\alpha$ in $\mathcal{C}^1$ solving: $$\Delta \check{w}=-\frac{1}{l^{q+2}}\frac{\langle\nabla \check{w}, Y(\check{x}_0)\rangle^2}{\check{w}^{q+3}}.$$ Note that $$\Delta \check{w}^{-\alpha}\leq \alpha\frac{|\nabla \check{w}|^2}{\check{w}^{\alpha+2}}\left[\frac{||Y(\check{x}_0)||_{L^\infty}^2}{\varepsilon^{q+2}}-(\alpha+1)\right],$$ so $\check{w}^{-\alpha}$ is subharmonic for $\alpha$ large, and so Lemma \[lem: A1\] (see the Annex) implies that $\check{w}$ is constant, which in turn implies that $\check{U}=0$ and $\nabla \check{U}=0$, which is false (see proof of Lemma \[lem: former lemma 1\]). Therefore, the second subcase cannot be true. This essentially means that when a concentration point is of the second type, then $$\sup_{B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right)}\check{u}_\alpha(x)\to\infty$$ and so $$\check{u}_\alpha(\check{x}_{i,\alpha})\to\infty.$$
Let us denote $\check{x}_i:=\lim_{\alpha\to\infty}\check{x}_{i,\alpha}$ up to a subsequence. According to Proposition \[prop\], $$\label{sectype}
\check{u}_\alpha(\check{x}_{i,\alpha})\check{u}_\alpha(x)\mapsto\frac{\lambda_i}{|x-\check{x}_i|^{n-2}}+H_i(x)$$ in $\mathcal{C}^1$ in $B_{\check{x}_i}\left(\frac{1}{2}\right)\backslash \{\check{x_i}\}$, with $\lambda_i>0$, where $H_i$ is a harmonic function in $B_{\check{x}_{i,\alpha}}\left(\frac{1}{2}\right)$, $H(\check{x}_i)=0$.
Let $U$ be a connected open set of ${\mathbb{R}}^n$, $U_R\subset B_0(R+1)$, containing no other point of the cluster apart from $\check{x}_i$ and $\check{x}_j$. For any $0<r<\frac{1}{8}$, we set $$V_{r,R}=U_R\backslash\left(\overline{B_{\check{x_i}}(r)\bigcup B_{\check{x_j}}(r)}\right).$$
For a fixed $x\in B_{\check{x}_i}\left(\frac{1}{4}\right)\backslash V_{r,R}$, (\[sectype\]) implies that $\check{u}_\alpha(x)\to 0$ as $\alpha\to\infty$. It follows from Lemma \[lem former lemma 9\] and (\[lower bound\]) that all points of a cluster must be of the same type.
Assuming all points in the cluster are of the first type, then $$\check{u}_\alpha(0)+\left|\frac{\nabla\check{u}_\alpha(0)}{\check{u}_\alpha(0)}\right|^{\frac{n-2}{2}}=O(1),$$ then by standard elliptic theory there exists $\check{u}:=\lim_{\alpha\to\infty}\check{u}_\alpha$ in $\mathcal{C}^1(B_0(R))$, $R>0$. Repeating the reasoning of Lemma \[lem: former lemma 2\] or Lemma \[lem: former lemma 4\], we know that $$\Delta_\xi \check{u}=f(x_1)\check{u}^{q-1}.$$ However, $\check{u}$ must have at least two separate maxima, at 0 and $\check{x}_2$, which leads to a contradiction by the classification result of Caffarelli, Gidas and Spruck [@CaffGidSpr].
Therefore $\check{u}_\alpha(\check{x}_{i,\alpha})\to\infty$, for any $i=\overline{1,M_{2R,\alpha}}$. Up to a subsequence $$\frac{\check{u}_\alpha(\check{x}_{i,\alpha})}{\check{u}_\alpha(0)}\to\mu_i>0\quad\text{as}\quad\alpha\to\infty.$$ We fix $R>0$ and assume, without loss of generality, that $(M_{2R,\alpha})_\alpha$ is a constant denoted by $M_{2R}$. Using Lemma \[lem former lemma 9\] and standard elliptic theory, we pass to a subsequence and get $$\check{u}_\alpha(0)\check{u}_\alpha(x)\to\check{G}(x)$$ in $\mathcal{C}^1_{loc}\left(B_0(R)\backslash\{\check{x}_i\}_{i=\overline{1,N_{2R}}}\right)$ for $\alpha\to\infty$, with $$\begin{array}{r l}
\check{G}(x)&=\sum_{i=1}^{p}\frac{\lambda_i}{\mu_i|x-\check{x}_i|^{n-2}}+\check{H}(x)\\&=\frac{\lambda_1}{|x|^{n-2}}+\left(\sum_{i=2}^{p}\frac{\lambda_i}{\mu_i|x-\check{x}_i|^{n-2}}+\check{H}(x)\right)
\end{array}$$ Here, $\check{H}$ is harmonic on $B_0(R)$, and $2\leq p\leq M_{2R}$ such that $|\check{x}_p|\leq R$ as $|\check{x}_{p+1}|>R$. If we apply Proposition \[prop\] to the blow-up sequence $x_\alpha=x_{1,\alpha}$ with $\rho_\alpha=\frac{1}{16}d_\alpha$, we obtain $$\hat{H}(0):=\sum_{i=2}^{p}\frac{\lambda_i}{\mu_i|\check{x}_i|^{n-2}}+\check{H}(0)=0$$ Since $\hat{H}(x)-\frac{\lambda_2}{\mu_2|x-\check{x}_{2}|}=\check{G}(x)-\frac{\lambda_1}{|x|^{n-2}}-\frac{\lambda_2}{\mu_2|x-\check{x}_2|^{n-2}}$ is harmonic in the ball $B_0(R)\backslash\{\check{x}_i\}_{i\in\overline{2,N_{2R}}}$ and $\check{G}\geq 0$, then as a consequence of the maximum principle, by considering a minimum on $\partial B_0(R),$ we see that $$\hat{H}(0)\geq\frac{\lambda_2}{\mu_2}-\frac{\lambda_1}{R^{n-2}}-\frac{\lambda_2}{\mu_2(R-1)^{n-2}}.$$ Choosing $R>0$ large enough, we ensure that $\hat{H}>0,$ which contradicts Theorem \[lem: former lemma 7\]. Consequently, $u_\alpha$ admits no concentration points and is therefore uniformly bounded in $\mathcal{C}^1$.
\[lem strong stability\] Assuming equation (\[EM\]) associated to $\tilde{a}$ admits a supersolution and that $\Delta_g+h$ is coercive, then for any $0<T<\inf_M\tilde{a}$ and any equation with parameters in $\mathcal{E}_{\theta,T}$ (as in Theorem \[theo stability eq\]), there exists a constant $C_{\theta,T}=C(n,\theta,T)>0$ such that, for any $||Y||_{L^\infty}\leq C_{\theta,T}$ and $||b||_{L^\infty}\leq C_{\theta,T}$, we may find a smallest real eigenvalue $\lambda_0>0$, where $\lambda_0$ is as in Lemma \[lem weak stability\].
Given any parameters $(f,a,b,c,d,h,Y)$ in $\mathcal{E}_{\theta,T}$ and additionally asking for $Y$ and $b$ to be sufficiently small in $L^\infty$ norm with respect to $\theta$ and $T$, we aim to prove that minimal solutions to the Lichnerowicz-type equation change continuously with their parameters. In order to do this, we study the sign of the smallest real eigenvalue associated to the linearisation around a minimal solution and show that it is positive by comparing it to the smallest real eigenvalue at $b=0$ and $Y=0$. Indeed, let $s>0$ a real number and $E_s$ the equation $$\label{eq theta}
\begin{array}{c}
E_s(u_s):=\Delta_g u_s+hu_s-fu_s^{q-1}-\frac{a}{u_s^{q+1}}+\frac{s b}{u_s}+c\langle\nabla u_s,s Y\rangle\left(\frac{d}{u_s^2}+\frac{1}{u_s^{q+2}}\right)\\+\frac{\langle\nabla u_s,s Y\rangle^2}{u_s^{q+3}}=0,
\end{array}$$ with $u_s$ its minimal solution. Let $L_s$ be the linearisation of $E_s$ around $u_s$, $$\begin{array}{c}\label{eq varphi}
\Delta_g\varphi_s+\Big[h-(q-1)fu_s^{q-2}+(q+1)\frac{a}{u_s^{q+2}}-\frac{s b}{u_s}-c\langle\nabla u_s,s Y\rangle\left(\frac{2d}{u_s^3}+\frac{q+2}{u_s^{q+3}}\right)\\-(q+3)\frac{\langle\nabla u_s,s Y\rangle^2}{u_s^{q+4}}\Big]\varphi_s+\langle\nabla\varphi_s,s Y \rangle\Big[c\left(\frac{d}{u_s^2}+\frac{1}{u_s^{q+2}}\right)+\frac{2\langle\nabla u_s,s Y\rangle}{u_s^{q+3}}\Big]=\lambda_s\varphi_s,
\end{array}$$ with $\lambda_s\geq 0$ the smallest real eigenvalue, $\varphi_s>0$ the associated eigenfunction, normalised such that $||\varphi_s||_{L^2}=1.$ Note that the linear equations $L_s$ are stable, in the sense that $\varphi_s$ is *a priori* uniformly bounded in $\mathcal{C}^1$. This follows from the fact that the $u_s$ is uniformly bounded. We may also suppose that $\lambda_s$ is uniformly bounded, because if $\lambda_s\to\infty$, then it is clear that $\lambda_s>0$.
As Premoselli proved by way of a variational argument [@Pre14], the equation $E_0$ is strictly stable, in the sense that its corresponding smallest real eigenvalue is positive. It uses the coerciveness of $\Delta_g+h$. We emphasize that his argument makes use of the fact that $E_0$ is symmetric, which is not the case for our more general equations. The strict stability implies continuity, *i.e.* that $u_s\to u_0$, with $u_0$ the minimal value. Indeed, let $u_s\to \tilde{u}$ another solution of $E_0$. Clearly, $\tilde{u}>u_0$. Let $\tilde{u}_\delta=u_0+\delta\varphi_0$. Note that $$\begin{array}{r l}
E_s(\tilde{u}_\delta)&=E_0(\tilde{u}_\delta)+\frac{s b}{u_\delta}+c\langle\nabla u_\delta,s Y\rangle\left(\frac{d}{u_\delta^2}+\frac{1}{u_\delta^{q+2}}\right)+\frac{\langle\nabla u_\delta,s Y\rangle^2}{u_\delta^{q+3}}\\
&=E_0(u_0)+\lambda_0\delta\varphi_0+o(\delta)+\frac{s b}{u_\delta}+s c\langle\nabla u_\delta, Y\rangle\left(\frac{d}{u_\delta^2}+\frac{1}{u_\delta^{q+2}}\right)\\
&\quad +s^2\frac{\langle\nabla u_\delta, Y\rangle^2}{u_\delta^{q+3}}
\end{array}$$ If we fix $\delta>0$ sufficiently small, the error terms $|o(\delta)|\leq \frac{\lambda_0\delta\varphi_0}{3}$ and $\tilde{u}_\delta<\tilde{u}\leq u_s$, $\forall s$. Then, by taking $s$ sufficiently close to $0$, we get that the rest of the terms are also smaller in absolute size than $\frac{\lambda_0\delta\varphi_0}{3}$. Consequently, $E_s(\tilde{u}_\delta)>0$, so $\tilde{u}_\delta$ is a supersolution of $E_s$ that is smaller than the minimal solution $u_s$.
Since $u_s\to u_0$, we also get that $\lambda_s\to\lambda_0$, so for $s$ small, the first eigenvalue $\lambda_s>0$. We would like to obtain that there exists $s_{\theta,T}>0$ such that, for any $0\leq s<s_{\theta,T}$, the minimal eigenvalue corresponding to $L_s$ is positive, where $(a,b,c,d,f,h,Y)\in\mathcal{E}_{\theta,T}$. In other words, we attempt to set a size for $Y$ and $b$, depending on $\theta$ and $T$ (and $n$), such that the resulting equations are strictly stable.
First, there exists $\delta_{\theta,T}>0$ such that if $Y=0$, $b=0$ and the equation’s parameters are found in $\mathcal{E}_{\theta,T}$, then $\lambda_0>\delta_{\theta,T}.$ We let $u_s=u_0+\varepsilon_s v_s$ such that $||v_s||_{L^2}=1$, $\varepsilon_s\in{\mathbb{R}}$. Note that $\varepsilon_s\to 0$ as $s\to 0$.
We begin by analyzing the difference in size between $\varepsilon_s$ and $s$, or equivalently between $||u_s-u_0||_{L^\infty}$ and $s$. Let $$E_s=E_0+s M_s,$$ where $$M_s(u_s)=\frac{b}{u_s}+c\langle\nabla u_s, Y\rangle\left(\frac{d}{u_s^2}+\frac{1}{u_s^{q+2}}\right)+s\frac{\langle\nabla u_s, Y\rangle^2}{u_s^{q+3}}.$$ Recall that $$\label{Eq theta develop}
E_0(u_s)=-s M_s(u_s)$$ where $$E_0(u_s)=L_0(u_s-u_0)+O(|u_s-u_0|^2).$$ Since $u_0$ is a solution of $E_0$ and the operator $L_0$ is coercive, with minimal eigenvalue $\lambda_0$, then by testing (\[Eq theta develop\]) against $(u_s-u_0)$, we see that $$\lambda_0\left(1+o(1)\right)||u_s-u_0||_{L^2}^2\leq -s \int_M M_s(u_s)(u_s-u_0)\leq s||M_s(u_s)||_{L^2}||u_s-u_0||_{L^2}.$$ The size of $M_s(u_s)$ is determined by a constant depending on $\theta$ and $T$. Therefore, we may write $$(1+o(1))\varepsilon_s =(1+o(1))||u_s-u_0||_{L^2}\leq s \frac{C}{\lambda_0}$$ Finally, in order to compare $\lambda_s$ to $\lambda_0$, extract the terms of order $s$ from the quantity $\int_M \varphi_0L_s(\varphi_s)-\varphi_s L_0(\varphi_0)$, $$\label{eq L theta order varepsilon theta}
\begin{array}{c}
-\int_M\frac{s b}{u_0}\varphi_0\varphi_s-\int_Mc\langle\nabla u_0,s Y\rangle\left(\frac{2d}{u_0^3}+\frac{q+3}{u_0^{q+3}}\right)\varphi_s\varphi_0-\int_M(q+3)\frac{\langle\nabla u_s,s Y\rangle^2}{u_s^{q+4}}\varphi_s\varphi_0\\
\varepsilon_s\int_M\left[(q-1)(q-2)fu_0^{q-3}-(q+1)(q+2)\frac{a}{u_0^{q+3}}\right]v_s \varphi_s\varphi_0\\
+O(s^2)=(\lambda_\theta-\lambda_0)\int_M\varphi_s\varphi_0,
\end{array}$$ so there exists a constant $C$ depending on $\theta$ and $T$ such that $$|\lambda_s-\lambda_0|\leq s C\left(\int_M\varphi_s\varphi_0\right)^{-1}.$$ As $\varphi_s=\varphi_0+o(1)$ and the $L^2$ norm of $\varphi_0$ is $1$, we may choose $s$ small enough so that $|\lambda_s-\lambda_0|\geq \frac{\delta_{\theta,T}}{2}$, and thus $\lambda_s>0$.
Existence of solutions to the system {#theo: system}
====================================
The proof of the main theorem
-----------------------------
The following is a useful estimate we can find in [@IseMur]; it plays a crucial role in ensuring the necessary compacity of the sequence $W_\alpha$ in the main theorem.
\[prop isemur\] Let $(M,g)$ be a closed Riemannian manifold of dimension $n\geq 3$ such that $g$ has no conformal Killing fields. Let $X$ be a smooth vector field in $M$. Then there exists a unique solution $W$ of $$\Delta_{g, conf}W=X.$$ Also, for $0<\gamma<1$, there exists a constant $C_0>0$ that depends only on $n$ and $g$ such that $$||W||_{\mathcal{C}^{1,\gamma}}\leq C_0||X||_\infty.$$
As a consequence, there exists a constant $C_1=C_1(n,g)$ such that $$\label{lxest}
||\mathcal{L}_g W||_{\mathcal{C}^{0,\gamma}}\leq C_1||X||_\infty$$
Let $(M,g)$ be a closed Riemannian manifold of dimension $n\in\{3,4,5\}$ such that $g$ has no conformal Killing fields. Let $b$, $c$, $d$, $f$, $h$, $\rho_1$, $\rho_2$, $\rho_3$ be smooth functions on $M$ and let $Y$ and $\Psi$ be smooth vector fields defined on $M$. Let $0<\gamma<1$.
Assume that $\Delta_g+h$ is coercive. Assume that $f>0$, $\rho_1>0$ and $|\nabla \rho_3|<(2C_1)^{-1}$, where $C_1$ is defined in (\[lxest\]).
Consider the coupled system $$\label{system in proof}
\begin{cases}
\displaystyle \Delta_g u+hu
&\displaystyle= fu^{q-1}+\frac{\rho_1+|\Psi+\rho_2\mathcal{L}_gW|^2_g}{u^{q+1}}\\
&\displaystyle -\frac{b}{u}-c\langle\nabla u,Y\rangle \left(\frac{d}{u^2}+\frac{1}{u^{q+2}}\right)-\frac{\langle\nabla u,Y\rangle^2}{u^{q+3}}\\
\displaystyle div_{g}\left(\rho_3\mathcal{L}_g W\right)
&\displaystyle =\mathcal{R}(u),
\end{cases}$$ where $\mathcal{R}$ is an operator verifying $$\label{CR}
\mathcal{R}(u)\leq C_\mathcal{R}\left(1+\frac{||u||_{\mathcal{C}^2}^2}{(\inf_Mu)^2}\right)$$ for a constant $C_\mathcal{R}>0$.
We fix $$\label{theta}
\theta=\min(\inf_M\rho_1,\inf_M f),$$ and $$\label{T2}
T=\max(||f||_{\mathcal{C}^{1,\eta}}, ||\rho_1||_{\mathcal{C}^{0,\gamma}}, ||c||_{\mathcal{C}^{0,\gamma}}, ||d||_{\mathcal{C}^{0,\gamma}}, ||h||_{\mathcal{C}^{0,\gamma}}).$$ Let $$\label{bound m}
M=\ln S_{\theta,2T},$$ with $S_{\theta,2T}$ a constant as in Theorem \[theo stability eq\]. The following theorem is the main result of the present paper.
\[thm plus 1\] Assume there exists a smooth positive function $\tilde{a}$ for which $$\label{licheq}
\Delta_g \tilde{u}+ h\tilde{u}=f\tilde{u}^{q-1}+\frac{\tilde{a}}{\tilde{u}^{q+1}}$$ admits a positive supersolution $\tilde{u}$. Assume that $\rho_1<\tilde{a}$ and let $\omega=\inf_M(\tilde{a}-\rho_1)$. Then there exists $$\delta=\delta(\omega,\theta,T)>0$$ such that if $$||b||_{\mathcal{C}^{0,\gamma}}+ ||Y||_{\mathcal{C}^{0,\gamma}}+ ||\Psi||_{\mathcal{C}^{0,\gamma}}+||\rho_2||_{\mathcal{C}^{0,\gamma}}+ C_\mathcal{R}\leq \delta,$$ the system (\[system in proof\]) admits a solution $(u,W)$, with $u$ a smooth positive function and $W$ a smooth vector field.
We can use a result by Hebey, Pacard and Pollack ([@HebPacPol08], Corollary 3.1) in order to ensure the existence of a supersolution $\tilde{u}$. There exists a constant $C=C(n,h)$, $C>0$ such that if $\tilde{a}$ is a smooth positive function verifying $$||\tilde{a}||_{L^1(M)}\leq C(n,h)\left(\max_M |f|\right)^{1-n},$$ then (\[licheq\]) accepts a smooth positive solution.
The proof of the theorem consists of a fixed-point argument. Formally, we define the operator $$\Phi:\varphi\to\ln u\left(\mathcal{L}_gW(e^\varphi)\right),$$ where $W(e^\varphi)$ solves the second equation of (\[system in proof\]) for a fixed $u=e^\varphi$ and where $u\left(\mathcal{L}_gW(e^\varphi)\right)$ is the solution of the scalar equation of (\[system in proof\]) constructed in Section 2 for a fixed $W(e^\varphi)$. In order to apply Schauder’s fixed point theorem, we show that $\Phi:B_M\to B_M$, $B_M:=\{\varphi\in\mathcal{C}^2(M), ||\varphi||_{\mathcal{C}^2}\leq M\}$, where $M$ is defined as in (\[bound m\]), and that $\Phi$ is continuous and compact.
We first want to prove that $$\label{smaller than a tilde}
\rho_1+|\Psi+\rho_2\mathcal{L}_gW(e^\varphi)|^2_g<\tilde{a}$$ to ensure that $\Phi(\varphi)$ is well defined, with $\tilde{u}$ from (\[licheq\]) a supersolution. By (\[lxest\]), we have $$\label{l c gamma}
||\mathcal{L}_g W(e^\varphi)||_{\mathcal{C}^{0,\gamma}}\leq C_1\left(||\nabla\rho_3||_{L^\infty}||\mathcal{L}_g W(e^\varphi)||_{L^\infty}+||\mathcal{R}(e^\varphi)||_{L^\infty}\right)$$ and thanks to (\[CR\]) we see that $$\begin{array}{c}
\rho_1+|\Psi+\rho_2\mathcal{L}_gW(e^\varphi)|^2_g\leq \rho_1+ 2||\Psi||_{L^\infty}^2\\
+2\left(\frac{C_1 C_{\mathcal{R}}||\rho_2||_{L^\infty}}{1-C_1||\nabla \rho_3||_{L^\infty}}\right)^2 \left(1+\frac{M^2e^{2M}}{e^{2\varepsilon}}\right)^2,
\end{array}$$ where $\varepsilon$ is the lower bound of any solution corresponding to $\mathcal{E}_{\theta,2T}$ from Theorem \[theo stability eq\]. There exists $$\label{delta1}
\delta_1=\delta_1(\omega,\theta,T)>0$$ such that if $$||\Psi||_{\mathcal{C}^{0,\gamma}}+||\rho_2||_{\mathcal{C}^{0,\gamma}}+C_{\mathcal{R}}\leq \delta_1,$$ then (\[smaller than a tilde\]) holds.
In order to use the *a priori* estimate of Section 3 to see that $\Phi:B_M\to B_M$, we need to prove that $$\label{ee}
\theta\leq \rho_1+|\Psi+\rho_2\mathcal{L}_gW(e^\varphi)|^2_g \quad\text{ and }\quad ||\rho_1+|\Psi+\rho_2\mathcal{L}_gW(e^\varphi)|^2_g||_{\mathcal{C}^{0,\gamma}}\leq 2T.$$
From (\[theta\]) we deduce that $$\label{ee1}
\theta\leq \rho_1+|\Psi+\rho_2\mathcal{L}_gW(e^\varphi)|^2_g.$$ and thanks to (\[l c gamma\]) we see that $$\begin{array}{c}
||\rho_1+|\Psi+\rho_2\mathcal{L}_gW(e^\varphi)|^2_g||_{\mathcal{C}^{0,\gamma}}\leq ||\rho_1||_{\mathcal{C}^{0,\gamma}}+ 2||\Psi||_{\mathcal{C}^{0,\gamma}}^2\\
+2\left(\frac{C_1 C_{\mathcal{R}}||\rho_2||_{\mathcal{C}^{0,\gamma}}}{1-C_1||\nabla \rho_3||_{L^\infty}}\right)^2 \left(1+\frac{M^2e^{2M}}{e^{2\varepsilon}}\right)^2.
\end{array}$$ There exists $$\label{delta2}
\delta_2=\delta_2(\omega,\theta, T)>0$$ such that if $$||\Psi||_{\mathcal{C}^{0,\gamma}}+||\rho_2||_{\mathcal{C}^{0,\gamma}}+C_{\mathcal{R}}\leq \delta_2,$$ then $$\label{ee2}
||\rho_1+|\Psi+\rho_2\mathcal{L}_gW(e^\varphi)|^2_g||_{\mathcal{C}^{0,\gamma}}\leq 2T.$$ Thanks to (\[ee1\]) and (\[ee2\]), the *a priori* estimates in Section 3 imply that $$||u\left(\mathcal{L}_gW(e^\varphi)\right)||_{\mathcal{C}^2}\leq S_{\theta,2T},$$ so $$\Phi(\varphi)\leq M,$$ where $M$ is as in (\[bound m\]). We have thus proved that $\Phi$ is well-defined and that $\Phi:B_M\to B_M$.
In order to show that $\Phi$ is continuous, we want to check that it holds true for $a\mapsto u(a)$, where $u(a)$ is the minimal solution constructed in Section 2. For all $a<\tilde{a}$, we’ve established monotony, which ensures that the minimal solutions exist. For $t>0$ small, let us denote by $u_t$ the solutions corresponding to $a(1+t)<\tilde{a}$. Let $u_0$ be the limit of $u_t$ as $t\to 0$; it is also a solution of the Lichnerowicz-type equation associated to $a$. If $u_0\not=u,$ then $u<u_0$. According to Section 3, there exists $C_{\theta,2T}>0$ such that $$||b||_{\mathcal{C}^{0,\gamma}}+||Y||_{\mathcal{C}^{0,\gamma}}\leq C_{\theta,2T}$$ implies that $u$ is strictly stable. We ask that $$\delta\leq\min\left( \delta_1, \delta_2, C_{\theta,2T}\right)$$ where $\delta_1$ is defined in (\[delta1\]) and $\delta_2$ is defined in (\[delta2\]). We choose $\mu>0$ small enough such that $u< \hat{u}_\mu< u_0$, where $\hat{u}_\mu:=u+\mu\psi$, $\psi$ a positive eigenfunction at $u$ corresponding to the smallest real eigenvalue. But $\hat{u}_\mu$ is a supersolution for $a(1+\epsilon)$, $\epsilon>0$ small, which contradicts the monotonicity. Therefore, $\Phi$ is continuous.
Lastly, $B_M$ being a closed convex set in $\mathcal{C}^2$, it remains to show that $\Phi(B_M)$ is compact to conclude. From the previous discussion, $\Phi(B_M)\subset B_M$, and is thus bounded in $\mathcal{C}^2$. By standard elliptic theory, we conclude the proof of Theorem \[thm 1\].
The case of a metric with conformal Killing fields
--------------------------------------------------
Let us consider the case of a metric $g$ with non-trivial conformal Killing fields associated to it. For $\tilde{V}$ a representative of the drift, the equation $$div_{\bar{g}}\left(\frac{\tilde{N}}{2}\mathcal{L}_{g}W\right)=\frac{n-1}{n}u^q\mathbf{d}\left(u^{-2q}\tilde{N}div_{g}(u^q\tilde{V})\right)+\pi\nabla\psi$$ admits a solution $W$ if and only if $$\frac{n-1}{n}\int_Mu^{-2q}\tilde{N}div_{g}(u^q\tilde{V})div_{g}(u^q P)=\int_M\langle\pi\nabla\psi,P\rangle$$ for all $P$ conformal Killing fields. Moreover, the solution $W$ is unique up to the addition of a conformal Killing field. Note that the drift is defined modulo conformal Killing fields, so $\tilde{V}$ and $\tilde{V}+P$ are representatives of the same drift for all $P$ conformal Killing fields. We claim that given a vector field $\tilde{V}$ there exists a conformal Killing field $\tilde{Q}$ which is unique up to a true Killing field and such that $$\label{Q eq}
\frac{n-1}{n}\int_M u^{-2q}\tilde{N}div_{g}\left(u^q(\tilde{V}+\tilde{Q})\right)div_{g} (u^qP)=\int_M\langle\pi\nabla\psi,P\rangle.$$ By analyzing the homogeneous operator associated to the equation above, $$\int_Mu^{-2q}\tilde{N}div_{g}\left(u^q(\tilde{V}+\tilde{Q}')\right)div_{g}(u^q P)=0,$$ we check that it is positive definite, thus invertible. Consider the functional $$F(P)=\int_Mu^{-2q}\tilde{N}div_{g}\left(u^q(\tilde{V}+P)\right)^2\,dv_{g}$$ on the finite-dimensional space of conformal Killing fields and note that $\tilde{Q}'$ is stationary for $F$. Since $F$ is quadratic and non-negative definite, stationary points are associated to minimizers. If $\bar{g}$ does not admit any nontrivial true Killing fields, then every conformal Killing field $P$ satisfies $div P\not = 0$ and the quadratic term of $F$ is positive definite. On the other hand, if $g$ admits proper Killing fields, then $F$ descends to a functional on the quotient space and its quadratic order term is again positive definite. So the minimum of $F$ is unique up to a true Killing field.
The conformal system proposed by Maxwell [@Max14] becomes in this framework $$\label{syst of Maxwell - with CKF}
\begin{cases}
\Delta_g u+\frac{n-2}{4(n-1)}(R(g)-|\nabla\psi|_g^2)u -\frac{(n-2)}{4(n-1)}\frac{|U+\mathcal{L}_g W|^2+\pi^2}{u^{q+1}}\\
\displaystyle\quad-\frac{n-2}{4(n-1)}\left[2V(\psi)-\frac{n-1}{n}\left(\tau^*+\frac{\tilde{N}div_g\left(u^q(\tilde{V}+\tilde{Q})\right)}{u^{2q}}\right)^2\right]u^{q-1}=0\\ \\
div_g\left(\frac{\tilde{N}}{2}\mathcal{L}_g W\right)=\frac{n-1}{n}u^q \mathbf{d} \left(\frac{\tilde{N}div_g(u^q\left(\tilde{V}+\tilde{Q}\right))}{2u^{2q}}\right)+\pi\nabla\psi,
\end{cases}$$ whose solution $(u,W,\tilde{Q})$ is a smooth positive function $u$, a smooth vector field $W$, defined up to a conformal Killing field, and $\tilde{Q}$ a conformal Killing field defined up to a true Killing field.
The existence of solutions to (\[syst of Maxwell - with CKF\]) follows from Theorem \[thm 1\] and is similar to Corollary \[physical result\], with slight modifications. Here, $$\begin{array}{c}
\rho_1=\frac{n-2}{4(n-1)}\left[\pi-\frac{n-1}{n}\tilde{N}^2 div_g(\tilde{Q}+\tilde{V})\right],\quad b=-\tau^*\tilde{N}div_g(\tilde{V}+\tilde{Q}),\\
Y=\sqrt{\frac{n}{n-2}}\tilde{N}(\tilde{V}+\tilde{Q})
\end{array}$$ and $$C_\mathcal{R}=C_\mathcal{R}(||\tilde{Q}||_{\mathcal{C}^2}).$$ Moreover, we define $\theta$ and $T$ as in (\[th main\]) and (\[T main\]) respectively, but without the dependency on $\rho_1=\rho_1(\tilde{Q})$, *i.e.* $$\theta=\min(\inf_M f),$$ and $$T=\max(||f||_{\mathcal{C}^{1,\eta}}, ||c||_{\mathcal{C}^{0,\gamma}}, ||d||_{\mathcal{C}^{0,\gamma}}, ||h||_{\mathcal{C}^{0,\gamma}}).$$ First of all, the stability of the first equation still holds, as in Lemma \[lem weak stability\] and Lemma \[lem strong stability\]. In order to apply the last theorem, we need to check that: $\rho_1(\tilde{Q})>\theta$, $||\rho_1(\tilde{Q})||_{\mathcal{C}^{0,\gamma}}<2T$, $||b||_{\mathcal{C}^{0,\gamma}}\leq C_{\theta,2T}$ and $||Y||_{\mathcal{C}^{0,\gamma}}\leq C_{\theta, 2T}$. This translates to $$div_g\tilde{Q}<\left(\frac{n-2}{4(n-1)}\pi-\theta\right)\frac{n}{n-1}\tilde{N}^{-2}-div_g\tilde{V},$$ $$\frac{n-2}{4(n-1)}||\pi||_{\mathcal{C}^{0,\gamma}}+\frac{n-1}{n}||\tilde{N}div_g(\tilde{V}+\tilde{Q})||_{\mathcal{C}^{0,\gamma}}\leq 2T,$$ $$||\tau^*\tilde{N}div_g(\tilde{V}+\tilde{Q})||_{\mathcal{C}^{0,\gamma}}\leq C_{\theta,2T},$$ and $$\sqrt{\frac{n}{n-2}}||\tilde{N}(\tilde{V}+\tilde{Q})||\leq C_{\theta,2T}.$$ We find bounds on $\tilde{Q}$ depending on $\pi$, $\psi$, $\tilde{N}$, $\tilde{V}$ from (\[Q eq\]), thereby proving the necessary compactness. Finally, the continuity $(a,b,Y)\to u_{a,b,Y}$ doesn’t pose any problem, and the proof mirrors our previous argument for the continuty of $a\to u(a).$ This shows the existence of solutions $(u,W,\tilde{Q})$.
Annex
=====
We used the following result repeatedly throughout the paper.
\[lem: A1\] Let $u$ be a bounded subharmonic function defined on ${\mathbb{R}}^n$. If there exists $0<\varepsilon\leq u$ which bounds $u$ from below and $\alpha>0$ such that $u^{-\alpha}$ is a subharmonic function, then $u$ is a constant.
Let us denote $$\bar{u}_{x}(R):=\frac{1}{\omega_{n-1}R^{n-1}}\int_{\partial B_{x}(R)}u(y)\,dy$$ the average of a smooth function $u$ over the sphere $\partial B_{x}(R)$. We will sometimes use the simplified notation $\bar{u}(R)$. Recall that, given any subharmonic function $u$, $x\in{\mathbb{R}}^n$ and for any two radii $R\leq\tilde{R}$, then $$\label{decreasing averages}
\bar{u}_{x}(R)\leq \bar{u}_x(\tilde{R}).$$ This follows from $$r^{n-1}\bar{u}'(r)=\frac{1}{\omega_{n-1}}\int_{\partial B_x{(r)}}\partial_\nu u(y)\,dy=-\frac{1}{\omega_{n-1}}\int_{B_x{(r)}}\Delta u(y)\,dy\geq 0$$ where $r>0$ and $\nu$ is the exterior normal.
Note that $u^{-\alpha}\leq \varepsilon^{-\alpha}$ implies that the average of $u^{-\alpha}$ on arbitrary subsets is uniformly bounded. Let us fix $x\in{\mathbb{R}}^n$. Since $u^{-\alpha}$ is bounded, there exists a constant $M>0$ and a sequence of radii $R_i\to\infty$ as $i\to\infty$ such that $$\label{def M}
M^{-\alpha}:=\lim_{i\to\infty}\overline{u^{-\alpha}}_{x}(R_i).$$ In fact, because the averages are decreasing (\[decreasing averages\]), any sequence $R\to\infty$ around any point in ${\mathbb{R}}^n$ leads to the same limit $M$, since one may always find a subsequence of $R_i$ such that $B_x(R_i)$ includes the new sequence.
As $u^{-\alpha}$ is subharmonic, $$u^{-\alpha}(x)\leq \overline{u^{-\alpha}}_x(R)$$ and therefore $u^{-\alpha}(x)\leq M^{-\alpha}$, or equivalently $$\label{M smaller than u x}
M\leq u(x).$$
For $z\in {\mathbb{R}}^n$, let $R:=|z-x|$ and $\tilde{R}>R$. By Green’s representation theorem, we get $$\label{est: from the annex}
\begin{array}{r l}
\displaystyle u(z) &\displaystyle \leq \int_{\partial B_x(\tilde{R})}u(y)\frac{\tilde{R}^2-R^2}{\omega_{n-1}\tilde{R}|z-y|^n}\, dy\\ \\
&\displaystyle\leq \frac{(\tilde{R}+R)\tilde{R}^{n-2}}{(\tilde{R}-R)^{n-1}}\overline{u}_x(\tilde{R}).
\end{array}$$ For $\delta>0,$ we denote $$\Omega_{\delta,R}:=\{z\in\partial B_x(R), u(z)\geq M+\delta\}$$ a subset of $\partial B_x(R)$ and let $$\theta_{\delta, R}:=\frac{|\Omega_{\delta, R}|}{|\partial B_x(R)|}\in[0,1]$$ be the corresponding relative size of its volume. Note that $\theta_{\delta,R}\to 0$ as $R\to\infty$. Otherwise, if there exists $\varepsilon\in(0,1]$ such that $$\limsup_{R\to\infty}\frac{|\{z\in\partial B_x(R), u(z)\geq M+\delta\}|}{|\partial B_x(R)|}=\varepsilon$$ then $$\limsup_{R\to\infty}\overline{u^{-\alpha}}_x(R) \leq \varepsilon(M+\delta)^{-\alpha}+(1-\varepsilon)M^{-\alpha}<M^{-\alpha}$$ which contradicts our definition (\[def M\]) of $M$.
By choosing $R$ large, $\theta_{\delta,R}\leq \delta$. Let $$\lambda_{\delta,i}:=\bar{u}_x(2^iR)$$ Note that, by (\[est: from the annex\]), $\lambda_{\delta,i}\leq 3\times 2^{n-2}\lambda_{\delta, i+1}.$ Since $$u(x)\leq\lambda_{\delta,i}\leq(M+\delta)(1-\theta_{\delta,2^iR})+\lambda_{\delta,i+1}\times\theta_{\delta,2^iR}$$ then, by induction, $$u(x)\leq (M+\delta)\frac{1-\delta^l}{1-\delta}+\lambda_l\delta^l$$ for all $l\in {\mathbb{N}}$. As we take $l\to\infty$, $$u(x)\leq(M+\delta)\frac{1}{1-\delta}$$ for any $\delta>0$, and therefore $u(x)\leq M.$ By (\[M smaller than u x\]), $u(x)\equiv M.$
We may apply the same argument to any other $\tilde{x}\in{\mathbb{R}}^n$ and obtain the same value $u(\tilde{x})=M$. Indeed, assuming that $$\tilde{M}^{-\alpha}:=\lim_{\tilde{R}\to\infty}\overline{u^{-\alpha}}_{\tilde{x}}(\tilde{R})$$ so that $\tilde{M}^{-\alpha}\geq M^{-\alpha}$, then for $\tilde{R}$ large, $\overline{u^{-\alpha}}_{\tilde{x}}(\tilde{R})\geq M^{-\alpha}$. But, at the same time, given any fixed $\tilde{R}$, then for $R$ sufficiently large, by (\[est: from the annex\]), $\overline{u^{-\alpha}}_{\tilde{x}}(\tilde{R})\leq \overline{u^{-\alpha}}_{x}(R)$. Thus we obtain that $u\equiv M$ in ${\mathbb{R}}^n$.
[99]{} Aubin, T., *Nonlinear analysis on manifolds. Monge-Amp[è]{}re equations. Grundlehren der Mathematischen Wissenschaften*, vol. 252. Springer, New York (1982) Caffarelli, L.A., Gidas, B., Spruck, J. *Asymptotic symmetry and local behavior of semilinear elliptic equations with critical Sobolev growth* *Comm. Pure Appl. Math.*, 271–297 (1989) Yvonne Choquet-Bruhat and Robert Geroch, *Global aspects of the Cauchy problem in general relativity*, Comm. Math. Phys. **14** (1969), no. 4, 329–335. Yvonne Choquet-Bruhat, James Isenberg, and Daniel Pollack, *Applications of theorems of Jean Leray to the Einstein-scalar field equations*, J. Fixed Point Theory Appl. 1 (2007), no. 1, 31–46. Yvonne Choquet-Bruhat, James Isenberg, and Daniel Pollack, *The constraint equations for the Einstein-scalar field system on compact manifolds*, Classical Quantum Gravity 24 (2007), no. 4, 809–828. MR 2297268 (2008a :83012) Olivier Druet, Emmanuel Hebey, *Stability and instability for Einstein-scalar field Lichnerowicz equations on compact Riemannian manifolds*, *Math. Z* (2009) 263:33-67. Lawrence C. Evans, *Partial Differential Equations*, 2nd edition, GSM19, *American Mathematical Society* (2010). Y. Four[è]{}s-Bruhat, *Th[é]{}or[è]{}me d’existence pour certains syst[è]{}mes d’[é]{}quations aux d[é]{}riv[é]{}es partielles non lin[é]{}aires*, Acta Math. 88 (1952), 141–225. MR 0053338 (14,756g) Emmanuel Hebey, Frank Pacard, Daniel Pollack, *A Variational Analysis Of Einstein–Scalar Field Lichnerowicz Equations On Compact Riemannian Manifolds* arXiv:gr-qc/0702031, Commun.Math.Phys.278:117-132,2008 Michael Holst, Gabriel Nagy, and Gantumur Tsogtgerel, *Rough Solutions of the Einstein Constraints on Closed Manifolds without Near-CMC Conditions*, Comm. Math. Phys. 288 (2009), no. 2, 547–613. MR 2500992 (2010k :53053) Mike Holst, David Maxwell, Rafe Mazzeo, *Conformal Fields and the Structure of the Space of Solutions of the Einstein Constraint Equations* arXiv:1711.01042, 2017 James Isenberg and Niall Ó’Murchadha, *Non-CMC conformal data sets which do not produce solutions of the Einstein constraint equations*, *Classical and Quantum Gravity* 21(2004), no. 3, S233–S241, A spacetime safari: essays in honour of Vincent Moncrief. MR 2053007 (2005c :83003) André Lichnerowicz, *L’int[é]{}gration des [é]{}quations de la gravitation relativiste et le probl[è]{}me des n corps*, J. Math. Pures Appl. (9) 23 (1944), 37–63. MR 0014298 (7,266d) David Maxwell, *Initial Data in General Relativity Described by Expansion, Conformal Deformation and Drift* arXiv:1407.1467v1 David Maxwell, *A model problem for conformal parametrizations of the Einstein constraint equations*, Communications in Mathematical Physics **302** (2011), no. 3, 697-736 David Maxwell, *Conformal parameterizations of slices of flat Kasner spacetimes*, arXiv:1404.7242, 2014. David Maxwell, *A class of solutions of the vacuum Einstein constraint equations with freely specified mean curvature*, Math. Res. Lett. 16 (2009), no. 4, 627–645. MR 2525029 (2010j :53057) Louis Nirenberg, *Topics in nonlinear functional analysis*, volume 6 of Courant Lecture Notes in Mathematics. New York University, Courant Institute of Mathematical Sciences, New York ; American Mathematical Society, Providence, RI, 2001. Chapter 6 by E. Zehnder, Notes by R. A. Artino, Revised reprint of the 1974 original. Bruno Premoselli, *The Einstein-Scalar Field Constraint System in the Positive Case*, *Communications in Mathematical Physics* **326** (2014), no. 2, 543-557. Bruno Premoselli, *Effective multiplicity for the Einstein-scalar field Lichnerowicz equation* *Calculus of Variations and Partial Differential Equations*, (2014), 10.1007/s00526-014-0740-y.
[^1]: Université Claude Bernard Lyon 1, 43 Boulevard du 11 Novembre 1918, 69100 Villeurbanne, `[email protected]`
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'In this paper we examine some aspects of the field of a scalar point charge in curved spacetimes. First we find the closed form solution for the scalar field due to a point charge in Schwarzschild spacetime. Then we expand it locally in powers of r (coordinate distance from the charge) and compare it to Quinn’s local expansion for the field of scalar charge. We show a term by term match, except for a mismatch in one term arising due to an error in a Riemann tensor term in Quinn’s expression. We show the correct expression, the detailed derivation of which will be subject of a later paper.'
author:
- Swapnil Tripathi
title: Examining the Field of Static Point Scalar Charge in Schwarzschild Spacetime
---
Introduction
============
Notations and Conventions
=========================
Field Equation
===============
Solution of field equation
==========================
Quinn’s Local Expansion of Scalar Field
=======================================
Conversion of Coordinates
=========================
Comparison
==========
Conclusion
==========
Acknowledgments
===============
It is a pleasure to thank Alan Wiseman for suggesting this problem and providing guidance in solving the problem. I would also like to thank John Friedman for many useful conversations and helpful suggestions. I would like to thank Gonzalo Olmo for providing help in overcoming some mathematical hurdles and Tobias Keidl for proofreading the manuscript and suggesting ways to improve the paper.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'A relativistic extension of our pseudo-shifted $\ell $–expansion technique is presented to solve for the eigenvalues of Dirac and Klein-Gordon equations. Once more we show the numerical usefulness of its results via comparison with available numerical integration data.'
address: 'Eastern Mediterranean University,'
author:
- Omar Mustafa
title: 'Perturbed Coulombic potentials in Dirac and Klein-Gordon equations'
---
\[theorem\][Acknowledgement]{} \[theorem\][Algorithm]{} \[theorem\][Axiom]{} \[theorem\][Claim]{} \[theorem\][Conclusion]{} \[theorem\][Condition]{} \[theorem\][Conjecture]{} \[theorem\][Corollary]{} \[theorem\][Criterion]{} \[theorem\][Definition]{} \[theorem\][Example]{} \[theorem\][Exercise]{} \[theorem\][Lemma]{} \[theorem\][Notation]{} \[theorem\][Problem]{} \[theorem\][Proposition]{} \[theorem\][Remark]{} \[theorem\][Solution]{} \[theorem\][Summary]{}
Introduction
============
Dirac and Klein-Gordon (KG) equation are not exactly soluble for most of Lorentz vector ( coupled as the 0-component of the 4-vector) and/or Lorentz scalar ( added to the mass term) potentials\[1-14\]. One, therefore, has to resort to some approximation schemes \[1-6\]. Yet, in between non-numerical and purely numerical non-relativistic (Schrödinger) and relativistic (KG and Dirac) wave equations there exists a broad [*gray zone*]{} of potentials tractable via various systematic semi-numerical (or semi-analytical) power-series expansions.
In numerous methodical predecessors of a subset of papers, Mustafa and co-workers \[7\] have sought a possibility in the power-law asymptotic expansions using some [*small*]{} parameter to solve for Schrödinger equation. It has been noted that the presence of the central spike $\ell
_{d}\,(\ell _{d}+1)\,/\,r^{2}$ (where $\ell _{d}=\ell +(d-3)/2$ and the dimensions $d\geq 2$) in the radial Schrödinger equation, just copies the effect of the centrifugal and/or centripetal force and immediately inspires the use of [*small shifted inverse angular momentum quantum number*]{}. Their PSLET ( pseudo-perturbative shifted-$\ell $ expansion technique) has provided persuasive numerical verifications by immediate comparison of its results with available [*brute force*]{} numerical data \[7\]. PSLET simply consists of using $1/\bar{l}$ as a perturbative expansion parameter, where $\bar{l}=\ell -\beta _{o}$, $\ell $ is a quantum number, and $\beta _{o}$ is a suitable shift introduced to avoid the trivial case $%
\ell =0$ \[7\].pagebreak
In this paper, we extend PSLET recipe to solve for Dirac and KG equations with Lorentz scalar and/or Lorentz vector radially symmetric potentials (in section 2). In section 3 we apply this relativistic recipe to some exactly solvable potentials ( e.g., $V(r)=S(r)=-A/r$, $V(r)=-A_{1}/r$ and $S(r)=0$, $%
V(r)=0$ and $S(r)=-A_{2}/r$, and the Dirac oscillator) and non-exactly solvable (by PSLET) potentials ( e.g., the pure scalar linear, the [*funnel-shapped*]{}, and the power-law potentials) to study the usefulness of its numerical results. We conclude in section 4.
PSLET recipe for Dirac and KG equations
=======================================
The Dirac equation with the Lorentz scalar ( added to the mass term) and Lorentz vector ( coupled as the 0-component of the 4-vector potential) potentials reads (in $\hbar =c=1$ units) $$\left\{ \vec{\alpha}.\vec{p}+\beta \,\left[ m+S(r)\right] \right\} \,\Psi (%
\vec{r})=\left\{ E-V(r)\right\} \,\Psi (\vec{r}).$$ Which decouples into $$\xi _{1}(r)\,G(r)+\frac{dF(r)}{dr}-\frac{\kappa }{r}F(r)=0,$$ $$\xi _{2}(r)\,F(r)-\frac{dG(r)}{dr}-\frac{\kappa }{r}G(r)=0.$$ where $\kappa =-\,(\ell +1)$ for $j=\ell +1/2$, $\kappa =\ell $ for $j=\ell
-1/2$, and $$\xi _{1}(r)=E-V(r)-\left[ m+S(r)\right] ,$$ $$\xi _{2}(r)=E+m-y(r);\;\;y(r)=V(r)-S(r).$$ $E$ is the relativistic energy, and $G(r)$ and $F(r)$ are the large and small radial components of the Dirac spinor, respectively. In terms of the large component $G(r),$ equation (2) reads $$\left[ \frac{d^{2}}{dr^{2}}-\frac{\kappa (\kappa +1)}{r^{2}}+\frac{1}{\xi
_{2}(r)}\,(\,y^{^{\prime }}(r)\,[\frac{d}{dr}+\frac{\kappa }{r}]\,)+\xi
_{1}(r)\,\xi _{2}(r)\right] G(r)=0,$$ where the prime denotes $d/dr.$ It can be shown that with the ansatz $$G(r)=\Phi (r)\,\,exp(-p(r)\,/\,2)~;~~\,~~p^{^{\prime }}(r)=y^{^{\prime
}}(r)\,/\,\xi _{2}(r),$$ equation (4) reads $$\left\{ -\frac{d^{2}}{dr^{2}}+\frac{\kappa (\kappa +1)}{r^{2}}+U(r)-\xi
_{1}(r)\,\xi _{2}(r)\right\} \Phi (r)=0,$$ where $$U(r)=\frac{y^{^{\prime \prime }}(r)}{2\,\xi _{2}(r)}-\frac{\kappa }{r}\frac{%
\,y^{^{\prime }}(r)}{\,\xi _{2}(r)}+\frac{3}{4}\,\left( \frac{y^{^{\prime
}}(r)}{\xi _{2}(r)}\right) ^{2}.$$ Obviously, equation (6) reduces to KG-equation with $\kappa \,(\kappa
+1)=\ell \,(\ell +1)$, for any $\kappa $, if $U(r)$ is set zero. It is therefore convenient to introduce a parameter $\lambda =0,\,1$ in $U(r)$ so that $\lambda =0$ and $\lambda =1$ correspond to KG and Dirac equations respectively. Also, we shall be interested in the problems where the rest energy $mc^{2}$ is large compared to the binding energy $E_{bind.}=E-mc^{2}$. This would manifest the approximation $$\frac{1}{\xi _{2}(r)}=\frac{1}{E_{bind.}+2m-y(r)}\simeq \frac{1}{2m}%
-O(1/m^{2}),$$ which in turn implies $$U(r)=\frac{\lambda }{4\,m}\left[ y^{^{\prime \prime }}(r)-\frac{2\,\kappa
\,y^{^{\prime }}(r)}{r}+\frac{3\,y^{^{\prime }}(r)^{2}}{4\,m}\right] .$$
For Coulombic-like potentials (i.e., a Lorentz-vector $V(r)=-A_{1}/r$ and a Lorentz-scalar $S(r)=-A_{2}/r$ potentials) one may re-scale the potentials and use the substitutions $$V_{r}(r)=V(r)^{2}-\frac{A_{1}^{2}}{r^{2}},$$ $$S_{r}(r)=S(r)^{2}-\frac{A_{2}^{2}}{r^{2}},$$ to recast equation (6) as $$\left[ -\frac{d^{2}}{dr^{2}}+\frac{[\,\bar{l}^{2}+\bar{l}\,(2\,\beta
_{o}+1)+\beta _{o}\,(\beta _{o}+1)]}{r^{2}}+\Gamma (r)+2\,E\,V(r)\right]
\,\Phi (r)=E^{2}\Phi (r).$$ where $$\Gamma (r)=-V_{r}(r)+S_{r}(r)+2\,m\,S(r)+m^{2}+U(r),$$ $$\bar{l}=\ell ^{\prime }-\beta _{o};\,\,\,\,\ell ^{\prime }=-\frac{1}{2}+%
\sqrt{\left( \ell +1/2\right) ^{2}-A_{1}^{2}+A_{2}^{2}},$$ and $\beta _{o}$ is a suitable shift to be determined below. Next, we shift the origin of the coordinate system $x=\bar{l}^{1/2}(r-r_{o})/r_{o}$, where $%
r_{o}$ is currently an arbitrary point to be determined through the minimization of the leading energy term below. It is therefore convenient to expand about $x=0$ (i.e., $r=r_{o}$) and use the following expansions $$r^{-2}=\sum_{n=0}^{\infty }\,\frac{a_{n}}{r_{o}^{2}}\,x^{n}\,\bar{l}%
^{-n/2};\,\,a_{n}=\left( -1\right) ^{n}\,(n+1),$$ $$\Gamma (x(r))=\frac{\bar{l}^{2}}{Q}\,\sum_{n=0}^{\infty }\,b_{n}\,x^{n}\,%
\bar{l}^{-n/2};\,\,b_{n}=\frac{d^{n}\,\Gamma (r_{o})}{dr_{o}^{n}}\,\frac{%
r_{o}^{n}}{n!}$$ $$V(x(r))=\frac{\bar{l}}{\sqrt{Q}}\,\sum_{n=0}^{\infty }\,c_{n}\,x^{n}\,\bar{l}%
^{-n/2};\,\,c_{n}=\frac{d^{n}\,V(r_{o})}{dr_{o}^{n}}\,\frac{r_{o}^{n}}{n!}$$ $$E=\frac{1}{\sqrt{Q}}\,\sum_{n=-1}^{\infty }\,E^{n}\,\bar{l}^{-n},$$ where $Q$ is set equal to $\,\bar{l}^{2}$ at the end of the calculations. With the above expressions into (12), one may collect all $x$-dependent terms of order $\bar{l}$ to imply the leading-order approximation for the energies $$E^{\left( -1\right) }=V(r_{o})\pm \sqrt{V(r_{o})^{2}+\Gamma (r_{o})+\frac{Q}{%
r_{o}^{2}}}.$$ Which upon minimization, i.e., $dE^{\left( -1\right) }/dr_{o}=0$ and $%
d^{2}E^{\left( -1\right) }/dr_{o}^{2}>0$, yields $$2\,Q=h\left( r_{o}\right) +\sqrt{h\left( r_{o}\right) ^{2}-g\left(
r_{o}\right) }$$ where $$h(r_{o})=r_{o}^{3}\left[ 2\,V(r_{o})\,V^{^{\prime }}(r_{o})+\Gamma
^{^{\prime }}(r_{o})+r_{o}V^{^{\prime }}(r_{o})^{2}\right] ,$$ $$g(r_{o})=r_{o}^{6}\left[ \Gamma ^{^{\prime
}}(r_{o})^{2}+4V(r_{o})V^{^{\prime }}(r_{o})\Gamma ^{^{\prime
}}(r_{o})-4\Gamma (r_{o})V^{^{\prime }}(r_{o})^{2}\right]$$ and primes denote derivatives with respect to $r_{o}$. This implies that $x%
\bar{l}^{-1}$-coefficients vanish, i.e.; $$Q\,a_{1}+r_{o}^{2}\,b_{1}+2\,r_{o}^{2}\,E^{\left( -1\right) }\,c_{1}=0.$$ Equation (12) therefore reduces to
$$\begin{aligned}
&&\left[ -\frac{d^{2}}{dx^{2}}+\sum_{n=2}^{\infty }T_{n}\,x^{n}\bar{l}%
^{-\left( n-2\right) /2}+\left( 2\,\beta _{o}+1\right) \sum_{n=0}^{\infty
}a_{n}\,x^{n}\,\bar{l}^{-n/2}\right. \nonumber \\
&&+\beta _{o}\,\left( \beta _{o}+1\right) \,\sum_{n=0}^{\infty
}a_{n}\,x^{n}\,\bar{l}^{-\left( n+2\right) /2} \nonumber \\
&&\left. +\frac{2\,r_{o}^{2}}{Q}\,\sum_{n=0}^{\infty
}\sum_{p=0}^{n+1}E^{\left( n-p\right) }\left( c_{2p}\,x^{2p}\,\bar{l}%
^{-n}+c_{2p+1}\,x^{2p+1}\bar{l}^{-\left( n+1/2\right) }\right) \right]
\,\Phi _{k,\ell }(x) \nonumber \\
&=&\left[ \frac{\,r_{o}^{2}}{Q}\,\,\sum_{n=-1}^{\infty
}\sum_{p=-1}^{n+1}E^{\left( n-p\right) }\,E^{\left( p\right) }\,\bar{l}%
^{-\left( n+1\right) }\right] \,\Phi _{k,\ell }(x)\end{aligned}$$
where $$T_{n}=a_{n}+\frac{r_{o}^{2}}{Q}\,b_{n}.$$ One may now compare equation (24) with the Schrödinger equation for the one-dimensional anharmonic oscillator $$\left[ -\frac{d^{2}}{dy^{2}}+\frac{1}{4}\,\omega ^{2}y^{2}+{\Large %
\varepsilon }_{o}+B(y)\right] \,Y_{k}(y)={\Large \mu }_{k}\,Y_{k}(y)$$ where $\varepsilon _{o}$ is constant, $B(y)$ is a perturbation like term and $${\Large \mu }_{k}={\Large \varepsilon }_{o}+(k+1/2)\omega
+\sum_{n=1}^{\infty }{\Large \mu }^{\left( n\right) }\,\bar{l}^{-n},$$ with $k=0,1,2,\cdots $ and $$\omega =\sqrt{12+\frac{2r_{o}^{4}}{Q}\,\Gamma ^{\prime \prime }(r_{o})+\frac{%
4\,r_{o}^{4}}{Q}\,E^{\left( -1\right) }\,V^{\prime \prime }(r_{o}).}$$ In a straightforward manner, one can show that $$E^{\left( 0\right) }=\frac{Q}{2r_{o}^{2}\left[ E^{\left( -1\right) }-c_{0}%
\right] }\left[ \left( 2\,\beta _{o}+1\right) +\left( k+1/2\right) \omega %
\right] ,$$ and choose $\beta _{o}\,$ so that $E^{\left( 0\right) }=0$ to obtain $$\beta _{o}=-\frac{1}{2}\left[ 1+\left( k+1/2\right) \omega \right] .$$ equation (24) then becomes
$$\left[ -\frac{d^{2}}{dx^{2}}+\sum_{n=0}^{\infty }{\Huge (}{\Large \upsilon }%
^{\left( n\right) }(x)\,\bar{l}^{-n/2}+J^{\left( n\right) }(x)\,\bar{l}%
^{-n}+K^{\left( n\right) }(x)\,\bar{l}^{-\left( n+1/2\right) }+{\LARGE %
\epsilon }^{\left( n\right) }\,\bar{l}^{-\left( n+1\right) }{\Huge )}\right]
\,\Phi _{k,\ell }(x)=0,$$
reak where $${\Large \upsilon }^{\left( 0\right) }(x)=T_{2}\,x^{2}+\left( 2\,\beta
_{o}+1\right) \,a_{0},$$
$${\Large \upsilon }^{\left( 1\right) }(x)=T_{3}\,x^{3}+\left( 2\,\beta
_{o}+1\right) \,a_{1}x,$$
$${\Large \upsilon }^{\left( n\right) }(x)=T_{n+2}\,x^{n+2}+\left( 2\,\beta
_{o}+1\right) \,a_{n}\,x^{n}+\beta _{o}\,(\beta
_{o}+1)\,a_{n-2}\,x^{n-2};\,\ n\geq 2,$$
$$J^{\left( n\right) }(x)=\left( \frac{2\,r_{o}^{2}}{Q}\right)
\sum_{p=0}^{n+1}E^{\left( n-p\right) }\,c_{2p}\,x^{2p},$$
$$K^{\left( n\right) }(x)=\left( \frac{2\,r_{o}^{2}}{Q}\right)
\sum_{p=0}^{n+1}E^{\left( n-p\right) }\,c_{2p+1}\,x^{2p+1}\,$$
$${\Huge \epsilon }^{\left( n\right) }=\left( \frac{\,r_{o}^{2}}{Q}\right)
\sum_{p=-1}^{n+1}E^{\left( n-p\right) }\,E^{\left( p\right) }.$$
Now we may closely follow PSLET recipe for the $k$-nodal wavefunction and define
$$\Phi _{k,\ell }(x)=F_{k,\,\ell }(x)\,\exp (U_{k,\,\ell }(x))$$
where $$F_{k,\,\ell }(x)=x^{k}+\sum_{n=0}^{\infty }\sum_{p=0}^{k-1}A_{p,k}^{\left(
n\right) }\,x^{p}\,\bar{l}^{-n/2},$$ $$U_{k,\,\ell }^{\prime }(x)=\sum_{n=0}^{\infty }\left( U_{k,\,\ell }^{\left(
n\right) }(x)\,\,\bar{l}^{-n/2}+G_{k,\,\ell }^{\left( n\right) }(x)\,\,\bar{l%
}^{-\left( n+1\right) /2}\right) ,$$ with $$U_{k,\,\ell }^{\left( n\right)
}(x)=\sum_{p=0}^{n+1}D_{p,n,k}\,x^{2p-1};\;\;D_{0,n,k}=0,$$
$$G_{k,\,\ell }^{\left( n\right) }(x)=\sum_{p=0}^{n+1}C_{p,n,k}\,x^{2p}.$$
Equation (31) then reads
$$\begin{aligned}
&&F_{k,\,\ell }(x)\sum_{n=0}^{\infty }{\huge [}{\Large \upsilon }^{\left(
n\right) }(x)\,\bar{l}^{-n/2}+J^{\left( n\right) }(x)\,\bar{l}%
^{-n}+K^{\left( n\right) }(x)\,\bar{l}^{-\left( n+1/2\right) }-{\Huge %
\epsilon }^{\left( n\right) }\,\bar{l}^{-\left( n+1\right) }{\huge ]}
\nonumber \\
&&-F_{k,\,\ell }(x)\,{\huge [}U_{k,\,\ell }^{^{\prime \prime
}}(x)+U_{k,\,\ell }^{^{\prime }}(x)U_{k,\,\ell }^{^{\prime }}(x){\huge ]}%
-2\,F_{k,\,\ell }^{\prime }(x)\,U_{k,\,\ell }^{^{\prime }}(x)-F_{k,\,\ell
}^{^{\prime \prime }}(x)=0\end{aligned}$$
where primes denote derivatives with respect to $x$. One may also eliminate $%
\,\bar{l}$-dependance from equation (43) to obtain four exactly solvable recursive relations ( see Appendix for details). Once $r_{o}$ is determined, through equation (20), one may then calculate the energy eigenvalues and eigenfunctions from the knowledge of $C_{p,n,k}$, $D_{p,n,k}$ and $%
A_{p,k}^{\left( n\right) }$ in a hierarchical manner.
Illustrative examples
=====================
In this section we illustrate the applicability of the above relativistic PSLET recipe through some examples covering Dirac and KG equations.
An equally mixed Coulomb potentials
-----------------------------------
For an equally mixed Coulombic potentials, i.e. $V(r)=S(r)=-A/r$, $U(r)$ in (9) vanishes. Consequently $\Gamma (r)=-2mA/r+m^{2}$, $Q=-A^{2}+2mAr$, $%
\omega =2$, $\beta _{o}=-\left( k+1\right) $, $r_{o}=\left[ \left( \ell
^{\prime }+k+1\right) ^{2}+A^{2}\right] \,/\left( 2mA\right) $, and the leading-order approximation reads $$E^{\left( -1\right) }=m\,\left[ 1-\frac{2\,A^{2}}{\left( k+\ell +1\right)
^{2}+A^{2}}\right] ,$$ which is the well known exact result for the generalized Dirac- and KG-Coulomb problems, where higher-order corrections vanish identically.
Vector Coulomb or scalar Coulomb potential
------------------------------------------
For $V(r)=-A_{1}/r$ and $S(r)=0$ or $V(r)=0$ and $S(r)=-A_{2}/r$ in KG equation one would obtain the well known exact results $$E^{\left( -1\right) }=m\left[ 1+\frac{A_{1}^{2}}{n_{1}^{2}}\right]
^{-1/2};\;\;n_{1}=k+\frac{1}{2}+\sqrt{\left( \ell +1/2\right) ^{2}-A_{1}^{2}}$$ or $$E^{\left( -1\right) }=\pm \,m\left[ 1-\frac{A_{2}^{2}}{n_{2}^{2}}\right]
^{1/2};\;\;n_{2}=k+\frac{1}{2}+\sqrt{\left( \ell +1/2\right) ^{2}+A_{2}^{2}}$$ respectively. Again higher-order corrections vanish identically.
Dirac oscillator
----------------
Following the work of Romero et al \[13\], the Dirac oscillator \[14\] eigenvalue problem ( see equation (30) in \[13\]) reduces to $$\left[ -\frac{d^{2}}{dr^{2}}+\frac{\Lambda (\Lambda +\epsilon \,\beta )}{%
r^{2}}+m^{2}B^{2}r^{2}+m^{2}+mB\left( \epsilon \left[ 2j+1\right] -\beta
\right) \right] \,\Phi (r)=E^{2}\Phi (r),$$ where $\Lambda =j+1/2$, $B$ is the oscillator frequency, and $\epsilon =\pm
1 $. In this case our $\Gamma (r)=m^{2}B^{2}r^{2}+m^{2}+mB\left( \epsilon %
\left[ 2j+1\right] -\beta \right) $, $\omega =4$, $r_{o}^{2}=$ $\bar{l}\,/mB$, and our leading term reads $$E^{\left( -1\right) }=\pm \,\left[ 2mB\left( 2k+\ell +3/2\right)
+m^{2}+mB\left( \epsilon \left[ 2j+1\right] -\beta \right) \right] ^{1/2}$$ with heigher-order terms identical zeros. Thus, if we take $N=2k+\ell $ ( the harmonic oscillator principle quantum number) we come out with the exact Dirac oscillator’s closed form solution (see equation (35) in \[13\]) $$E^{2}-m^{2}=\,\left[ 2mB\left( N+3/2\right) +mB\left( \epsilon \left[ 2j+1%
\right] -\beta \right) \right] .$$
Pure scalar linear potential
----------------------------
A pure scalar linear potential, i.e., $S(r)=Ar$ and $V(r)=0$, is precisely a quark confining potential. It has been used by Gunion and Li \[5\] in Dirac equation to find, numerically, part of Dirac $J/\Psi $ mass spectra.
Obviously, for this potential equation (20) has to be solved numerically. Then one can proceed to obtain the mass spectra for $A=0.137\,GeV^{2}$, $%
m=1.12\,GeV$, $\kappa =-\left( \ell +1\right) $, and $\kappa =\ell $ through the prescription $M=2E$.
In tables 1 and 2 we report our results for $J/\Psi $ mass ( in $GeV$) for $%
\kappa =-\left( \ell +1\right) $ and $\kappa =\ell $, respectively. To show the trends of convergence of our results, we report them as $M(N)=2E(N)$, with $N$ denoting the number of corrections added to the leading-order approximation $E^{\left( -1\right) }$. Our results are also compared with the numerically predicted ones of Gunion and Li \[5\]. Evidently, the accuracy and trend of convergence are satisfactory.
[*Funnel-shaped*]{} potential
-----------------------------
The [*funnel-shaped* ]{}potential is widely used in quarkonium physics. It has both vector and components, $V(r)$ and $S(r)$, respectively.
In Dirac equation, Stepanov and Tutik \[4\] have used numerical integrations and $\hbar $-expansion formalism without the traditional conversion of Dirac equation into a Schrődinger-like form (unlike what we have already done in section II above). They have obtained the Charmonium masses for $%
V(r)=-2\alpha /3r$ and $S(r)=br/2$, where $m=1.358\,GeV$, $\alpha =0.39$, and $b=0.21055\,GeV^{2}$.
In table 3 we show our results for the Charmonium masses for $\kappa =\ell $ and compare them with those of numerical integration and $\hbar $-expansions of Stepanov and Tutik \[4\]. They are in good agreement and the trend of convergence of our results is also satisfactory. However, in table 4 we report the Charmonium masses, for $\kappa =-\left( \ell +1\right) $. Therein, we only list our results where the mass series and Padé approximants stabilize.
In KG equation Kobylinsky, Stepanov, and Tutik \[3\] have used $V(r)=-a/r$ and $S(r)=b\,r$ with $m=1.370\,GeV$, $b=0.10429\,GeV^{2}$ and $a=0.26\,$ to obtain the energies for this [*funnel-shaped* ]{}potential. They have also used $\hbar $-expansions and numerical integrations. In table 5 we list our results and compare them with those of $\hbar $-expansions, $E_{\hbar }$, and numerical integrations, $E_{num}$, reported in \[3\]. They are in excellent agreement.
Power-law potential
-------------------
An equally mixed scalar and vector power-law potential $$V(r)=S(r)=Ar^{\nu }+B_{o}$$ where $\nu =0.1$ ( the Martin potential \[6\] ) and $A>0$ is found most successful in describing the entire light and heavy meson spectra in the Dirac equation (cf. e.g., Martin 1980, 1981, and Jena and Tripati in \[6\]). Once such potential setting is used in Eq.(6) along with the substitution $%
q=r/\varrho $, with $$\varrho =\left[ 2\left( E+m\right) A\right] ^{-1/\left( \nu +2\right) },$$ one gets a simple Schrödinger-type form $$\left[ -\frac{d^{2}}{dq^{2}}+\frac{\ell \left( \ell +1\right) }{q^{2}}%
+q^{\nu }\right] \,\Omega \left( q\right) =\check{E}\,\Omega \left( q\right)
,$$ where $$\check{E}=\left( E-m-2B_{o}\right) \left[ \left( E+m\right) \left( 2A\right)
^{-2/\nu }\right] ^{\nu /\left( \nu +2\right) }.$$ Therefore, one better solve (44) for $\check{E}\,(N)$ and then to find the Dirac quark binding energies $E$ from (45). In table 6 we compare PSLET results with those obtained numerically, $E_{num},$ by Jena and Tripati \[6\]. The results from the shifted $1/N$ - expansion technique by Roy and Roychoudhury \[6\], $E_{1/N},$ are also listed.
Concluding remarks
==================
In this work we presented a straightforward extension of our PSLET recipe \[7\] to solve for the eigenvalues of Dirac and KG equations, with Lorentz vector and/or Lorentz scalar potentials. We have, again, documented (through tables 1-5) the usefulness of this recipe by immediate comparisons with available numerical integration data.
Nevertheless, one should notice that our results from KG equation are only partially better , compared with those from $\hbar $-expansions and numerical integrations, than our results from Dirac equation. The reason is obviously, and by large, manifested by our approximation in equation (8). For Klein-Gordon equation $\lambda =0$ in (9) while for Dirac equation $%
\lambda =1$.
In the process, moreover, there still remain some issues of delicate nature. Namely, one can not obtain (using our PSLET above) the exact eigenvalues for $V(r)=-A_{1}/r$ and $S(r)=0$, $V(r)=0$ and $S(r)=-A_{2}/r$, or even for $%
V(r)=-A_{1}/r$ and $S(r)=-A_{2}/r$ in Dirac equation.
The remedy seems to be feasible in a sort of combination between the current relativistic PSLET and a similarity transformation (cf., e.g. ref.\[8\] and related references therein). Preliminary results show that if $%
S(r)\longrightarrow -A_{2}/r+S_{o}(r)$ and $V(r)\longrightarrow
-A_{1}/r+V_{o}(r)$ in (1) such that $S_{o}(r)\longrightarrow 0$ and $%
V_{o}(r)\longrightarrow 0$ as $r\longrightarrow 0$, then a similarity transformation could accompany our relativistic PSLET to obtain exact results for the generalized Dirac-Coulombic problem and better results for potentials of the type $S(r)\longrightarrow -A_{2}/r+S_{o}(r)$ and $%
V(r)\longrightarrow -A_{1}/r+V_{o}(r)$. That is, one may carefully follow Mustafa’s work \[8\] to obtain $$\left[ -\frac{d^{2}}{dr^{2}}+\frac{(\,\gamma ^{2}+s\,\gamma )}{r^{2}}%
+U(r)+m^{2}+\frac{2}{r}\,\left( A_{2}\,m(r)+E(r)\,A_{1}\right) \right]
\,\Phi (r)=E^{2}\Phi (r).$$ where $\gamma =\sqrt{\kappa ^{2}-A_{1}^{2}+A_{2}^{2}}$ , $s=\pm 1$, $%
E(r)=E-V_{o}(r)$, $m(r)=m+S_{o}(r)$, and $U(r)\longrightarrow 0$ as $%
V_{o}(r)\longrightarrow 0$ and $S_{o}(r)\longrightarrow 0$ (hence, $%
m(r)\longrightarrow m$ and $E(r)\longrightarrow E$). Therefore, one would replace our $\ell ^{\prime }$, in (12), by $\ell ^{\symbol{126}}=-1/2+\gamma
+s\,/\,2$ and obtain ( following PSLET recipe above) for (48), $\Gamma
(r_{o})=m^{2}-2mA_{2}r_{o}$, $V(r)=-A_{1}/r$, $\omega =2$, $\beta
_{o}=-\left( k+1\right) $, and $\bar{l}=\ell ^{\symbol{126}}-\beta
_{o}=(k+1/2+s\,/\,2+\,\gamma )$. In turn, equation (23) yields $$4Q\left[ Q-2mA_{2}r_{o}+A_{1}^{2}\right] +4m^{2}r_{o}^{2}\left[
A_{2}^{2}-A_{1}^{2}\right] =0$$ to solve for $r_{o}$. This would lead to $$E^{\left( -1\right) }=m\left[ 1+\frac{A_{1}^{2}}{Q}\right]
^{-1/2};\;\;Q=\left( k+1/2+s\,/\,2+\,\sqrt{\kappa ^{2}-A_{1}^{2}}\right) ^{2}$$ and $$E^{\left( -1\right) }=\pm \,m\left[ 1-\frac{A_{2}^{2}}{Q}\right]
^{1/2};\;\;Q=\left( k+1/2+s\,/\,2+\,\sqrt{\kappa ^{2}+A_{2}^{2}}\right) ^{2}$$ for $A_{2}=0$, $A_{1}\neq 0$ and $A_{1}=0$, $A_{2}\neq 0$, respectively. It should be noted that these are the well known exact results (cf., e.g.; ref \[8\]) with the heigher-order terms vanish identically.
Before we conclude it should be noted that if our results are to be generalized to $d$-dimensions we may incorporate interdimensional degeneracies associated with the isomorphism between the angular momentum $%
\ell $ and dimensionality $d$ (cf., e.g., Mustafa and Odeh (2000) \[7\]). This would replace our $\kappa $ by $\kappa _{d}=s(2j+d-2)/2$, where $\ell
_{d}=\ell +(d-3)/2$. In this way we reproduce Stepanov and Tutik’s \[4\] and Dong’s \[9\] results in $d$-dimensions.
Finally, the above has been a very limited review and a number of other useful and novel approaches such as those of Franklin \[10, and references therein\], Njock et al \[11, and references therein\], $\cdots $etc., have not been touched on.
Eliminating $\bar{l}$-dependence, equation (42) can be simplified into four recursive relations to read $$k(k-1)\,x^{k-2}+T_{k,\ell }^{\left( 0\right) }(x)-N_{k,\ell }^{\left(
0\right) }(x)=0$$ $$T_{k,\ell }^{\left( 1\right) }(x)+S_{k,\ell }^{\left( 0\right)
}(x)-O_{k,\ell }^{\left( 0\right) }(x)=0$$ and for $n\geq 0$$$T_{k,\ell }^{\left( 2n+2\right) }(x)-N_{k,\ell }^{\left( n+1\right)
}(x)+S_{k,\ell }^{\left( 2n+1\right) }(x)+M_{k,\ell }^{\left( 2n\right)
}(x)+\Lambda _{k,\ell }^{\left( n\right) }(x)=0$$ $$T_{k,\ell }^{\left( 2n+3\right) }(x)+S_{k,\ell }^{\left( 2n+2\right)
}(x)-O_{k,\ell }^{\left( n+1\right) }(x)+M_{k,\ell }^{\left( 2n+1\right)
}(x)+{\Large \zeta }_{k,\ell }^{\left( n\right) }(x)=0$$ where $$\begin{aligned}
T_{k,\ell }^{\left( n\right) }(x) &=&L_{k,\ell }^{\left( n\right) ^{\prime
\prime }}(x)+2\,k\,x^{k-1}U_{k,\,\ell }^{\left( n\right) }(x) \nonumber \\
&&+\sum_{p=0}^{n}2L_{k,\ell }^{\left( p\right) ^{\prime }}(x)\,U_{k,\,\ell
}^{\left( n-p\right) }(x)+x^{k}\,\left[ U_{k,\,\ell }^{\left( n\right)
^{\prime }}(x)+R_{k,\,\ell }^{\left( n\right) }(x)-{\Large \upsilon }%
^{\left( n\right) }(x)\right] \nonumber \\
&&+\sum_{p=0}^{n}L_{k,\ell }^{\left( p\right) }(x)\left( U_{k,\,\ell
}^{\left( n-p\right) ^{\prime }}(x)+R_{k,\,\ell }^{\left( n-p\right) }(x)-%
{\Large \upsilon }^{\left( n-p\right) }(x)\right)\end{aligned}$$ $$\begin{aligned}
S_{k,\ell }^{\left( n\right) }(x) &=&2\,k\,x^{k-1}G_{k,\,\ell }^{\left(
n\right) }(x)+\sum_{p=0}^{n}2L_{k,\ell }^{\left( p\right) ^{\prime
}}(x)\,G_{k,\,\ell }^{\left( n-p\right) }(x) \nonumber \\
&&+x^{k}\,\left[ G_{k,\,\ell }^{\left( n\right) ^{\prime }}(x)+Q_{k,\,\ell
}^{\left( n\right) }(x)\right] +\sum_{p=0}^{n}L_{k,\ell }^{\left( p\right)
}(x)\left( G_{k,\,\ell }^{\left( n-p\right) ^{\prime }}(x)+Q_{k,\,\ell
}^{\left( n-p\right) }(x)\right)\end{aligned}$$ $$M_{k,\ell }^{\left( n\right) }(x)=x^{k}\,P_{k,\ell }^{\left( n\right)
}\left( x\right) +\sum_{p=0}^{n}L_{k,\ell }^{\left( p\right)
}(x)\,\,P_{k,\ell }^{\left( n-p\right) }\left( x\right)$$ $$N_{k,\ell }^{\left( n\right) }(x)=x^{k}\,J^{\left( n\right) }\left( x\right)
+\sum_{p=0}^{n}L_{k,\ell }^{\left( p\right) }(x)\,\,J^{\left( n-p\right)
}\left( x\right)$$ $$O_{k,\ell }^{\left( n\right) }(x)=x^{k}\,K^{\left( n\right) }\left( x\right)
+\sum_{p=0}^{n}L_{k,\ell }^{\left( p\right) }(x)\,\,K^{\left( n-p\right)
}\left( x\right)$$ $$\Lambda _{k,\ell }^{\left( n\right) }(x)=x^{k}\,{\Huge \epsilon }^{\left(
n\right) }+\sum_{p=0}^{n}L_{k,\ell }^{\left( 2p\right) }(x)\,{\Huge \epsilon
}^{\left( n-p\right) }$$ $${\Large \zeta }_{k,\ell }^{\left( n\right) }(x)=\sum_{p=0}^{n}L_{k,\ell
}^{\left( 2p+1\right) }(x)\,{\Huge \epsilon }^{\left( n-p\right) }$$ $$R_{k,\ell }^{\left( n\right) }(x)=\sum_{p=0}^{n}U_{k,\,\ell }^{\left(
p\right) }(x)U_{k,\,\ell }^{\left( n-p\right) }(x)$$ $$P_{k,\ell }^{\left( n\right) }(x)=\sum_{p=0}^{n}G_{k,\,\ell }^{\left(
p\right) }(x)G_{k,\,\ell }^{\left( n-p\right) }(x)$$ $$Q_{k,\ell }^{\left( n\right) }(x)=\sum_{p=0}^{n}2U_{k,\,\ell }^{\left(
p\right) }(x)G_{k,\,\ell }^{\left( n-p\right) }(x)$$ $$L_{k,\ell }^{\left( n\right) }(x)=\sum_{p=0}^{k-1}A_{k,\,p}^{\left( n\right)
}(x)\,x^{p}$$
Au C K, and Aharonov Y 1981, J Math Phys [**22** ]{}1428
Lai C S 1982, J Phys [**A15**]{} L155
Rogers G W 1985, J Math Phys [**27**]{} 567
Chatterjee A 1986, J Math Phys [**27**]{} 2331
Papp E 1991, Phys Lett [**B259**]{} 19
Mustafa O and Sever R 1991, Phys Rev [**A44** ]{}4142
Kobylinsky N A, Stepanov S S and Tutik R S 1990, J Phys [**A23**]{} L237
Stepanov S S and Tutik R S 1992, Phys Lett [**A163**]{} 26
Critchfield C 1976, J Math Phys [**17**]{} 261
Gunion J and Li L 1975, Phys Rev [**D12**]{} 3583
Magyari E 1980, Phys Lett [**B95**]{} 295
Martin A 1980, Phys. Lett. [**B93**]{}, 3381
Martin A 1981, Phys. Lett. [**B100**]{}, 511
Jena S and Tripati T 1983, Phys Rev [**D28**]{} 1780
McQuarrie B R and Vrscay E R 1993, Phys Rev [**A47**]{} 868
Panja M, Dutt R and Vatshni Y P 1990, Phys Rev [**A42**]{} 106
Roy B and Roychoudhury 1990, J. Phys. [**A23**]{} 3555
Barakat T, Odeh M and Mustafa O 1998, J Phys [**A31**]{} 3469
Villalba V M and Maggiolo A R 2001, Eur Phys J [**B22**]{} 31
Mustafa O and Odeh M 1999, J Phys [**A32**]{} 6653
Mustafa O and Odeh M 2000, J Phys [**A33** ]{}5207
Mustafa O and Odeh M 2002, Czech J Phys [**52**]{} 795
Mustafa O and Znojil M 2002, J Phys [**A35**]{} 8929
Mustafa O 2003, Czech J Phys [**53**]{} 433
Mustafa O 2003, J Phys [**A36**]{} 5067
Mustafa O and Barakat T 1998, Commun Theor Phys [**30**]{} 411
Dong S H 2003, J. Phys [**A36** ]{}4977
Franklin J 1999, Mod. Phys. Lett [**A14**]{} 2409
Franklin J 2000, Int. J. Mod. Phys.[** A12**]{} 4355
Njock M G K 2000, Phys Rev [** A61**]{} 042105
Mustafa O 2002, J. Phys. [**A35**]{} 10671
Romero R P M, Yepez H N N, and Brito A L S 1995, Eur. J. Phys. [**16**]{} 135
Moshinsky M, Quesne C, and Smirnov Y F 1995, J. Phys. [**A28**]{} 6447
Rozmej P and Arvieu R 1999, J. Phys. [**A32**]{} 5367
Villalba V M 1994, Phys. Rev. [**A49**]{} 586
[**Table1:** ]{}PSLET results for part of Dirac $J/\Psi $ spectra (in GeV) for $S(r)=Ar,$ $A=0137GeV^{2},$ $V(r)=0,$ $\kappa =-(\ell +1),$ and $%
m=1.12GeV.$ Where $M(N)=2E(N)$ with $N$ denoting the number of corrections added to the leading - order term $E^{(-1)}$ and $M_{num}$ are the numerical values reported by Gunion and Li \[5\].
[cccccc]{} $k$ & $\ M(N)$ & $\ell =0$ & $\ \ell =1$ & $\ \ell =2$ & $\ \ell =3$\
$
\begin{tabular}{c}
$0$%
\end{tabular}
$ & $
\begin{tabular}{c}
$M(1)$ \\
$M(2)$ \\
$M(3)$ \\
$M(4)$ \\
$M(5)$ \\
$[****]{}$ \\
$M(14)$ \\
$M\_[num]{}$%
\end{tabular}
$ &
-----------------
3.0919
3.0961
3.0963
3.0961
3.0961
${\bf \vdots }$
3.0961
3.103
-----------------
&
-----------------
3.43078
3.43252
3.43259
4.43256
3.43256
${\bf \vdots }$
3.43256
3.442
-----------------
&
-----------
3.711960
3.712914
3.712947
3.712940
3.712939
$\vdots $
3.712939
3.725
-----------
&
-----------------
3.9581219
3.9587277
3.9587465
3.9587436
3.9587434
${\bf \vdots }$
3.9587434
3.973
-----------------
\
-----
$2$
-----
&
-------------------
$M(1)$
$M(2)$
$M(3)$
$M(4)$
$M(5)$
$M(6)$
$M(7)$
$\ {\bf \vdots }$
$M(14)$
$M_{num}$
-------------------
&
---------------------
4.131
4.142
4.148
4.150
4.151
4.152
4.152
$\ \ {\bf \vdots }$
4.152
4.158
---------------------
&
-----------------------
4.3395
4.3468
4.3502
4.3515
4.3519
4.3520
4.3521
$\ \ \ {\bf \vdots }$
4.3521
4.36
-----------------------
&
---------------------
4.5325
4.5378
4.5401
4.5408
4.5410
4.5411
4.5411
$\ \ {\bf \vdots }$
4.5411
4.551
---------------------
&
-----------------------
4.71334
4.71739
4.71905
4.71950
4.71961
4.71965
4.71966
$\ \ \ {\bf \vdots }$
4.71966
4.732
-----------------------
\
[**Table 2:** ]{}Same as table 1 for $\kappa =\ell .$
[cccccc]{} $k$ & $M(N)$ & $\ell =1$ & $\ell =2$ & $\ell =3$ & $\ell =4$\
-----
$0$
-----
&
-----------------
$M(1)$
$M(2)$
$M(3)$
$M(4)$
$M(5)$
${\bf \vdots }$
$M(14)$
$M_{num}$
-----------------
&
-----------------
3.47090
3.47183
3.47188
3.47186
3.47186
${\bf \vdots }$
3.47186
3.47
-----------------
&
-----------------
3.760125
3.760677
3.760700
3.760696
3.760695
${\bf \vdots }$
3.760695
3.757
-----------------
& $
\begin{tabular}{c}
4.0111817 \\
4.0115488 \\
4.0115615 \\
4.0115597 \\
4.0115595 \\
$[****]{}$ \\
4.0115595 \\
4.006
\end{tabular}
$ &
-----------------
4.2364739
4.2367365
4.2367444
4.2367435
4.2367435
${\bf \vdots }$
4.2367435
4.23
-----------------
\
---
1
---
&
-----------------
$M(1)$
$M(2)$
$M(3)$
$M(4)$
$M(5)$
$M(6)$
${\bf \vdots }$
$M(14)$
$M_{num}$
-----------------
&
-----------------
3.9570
3.9624
3.9640
3.9644
3.9646
3.9646
${\bf \vdots }$
3.9646
3.965
-----------------
&
-----------------
4.19083
4.19451
4.19537
4.19557
4.19561
4.19563
${\bf \vdots }$
4.19563
4.194
-----------------
&
-----------------
4.40310
4.40578
4.40631
4.40642
4.40644
4.40644
${\bf \vdots }$
4.40644
4.403
-----------------
&
-----------------
4.599080
4.601126
4.601482
4.601542
4.601552
4.601554
${\bf \vdots }$
4.601554
4.597
-----------------
\
[**Table 3:** ]{}PSLET Charmonium masses $M(N)=2E(N)$ for the [*funnel-shaped*]{} potential, $V(r)=-2\alpha /3r$ and $S(r)=br/2$, with $%
m=1.358\,GeV$, $b=0.21055\,GeV^{2}$, $\alpha =0.39$ and $\kappa =\ell .$ The quantum numbers in parentheses are ( $k,\ell $). $M_{num}$ is the numerical integration and $M_{\hbar }$ is the $\hbar $-expansion result ( up to the tenth-order correction) reported by Stepanov and Tutik \[4\].
[ccccccc]{} $M(N)$ & $(0,1)$ & $(0,2)$ & $(1,1)$ & $(1,3)$ & $(2,1)$ & $(2,3)$\
-----------------
$M(1)$
$M(2)$
$M(3)$
$M(4)$
$M(5)$
$M(6)$
$M(7)$
$M(8)$
${\bf \vdots }$
$M(14)$
$M_{\hbar }$
$M_{num}$
-----------------
&
-----------------
3.5071
3.5062
3.5056
3.5055
3.5055
3.5055
3.5055
3.5055
${\bf \vdots }$
3.5055
3.4998
3.4998
-----------------
&
-----------------
3.8012
3.8007
3.8006
3.8005
3.8005
3.8005
3.8005
3.8005
${\bf \vdots }$
3.8005
3.7974
3.7974
-----------------
&
-----------------
3.966
3.963
3.961
3.959
3.959
3.958
3.958
3.958
${\bf \vdots }$
3.958
3.9501
3.9499
-----------------
&
-----------------
4.3862
4.3857
4.3853
4.3852
4.3851
4.3851
4.3850
4.3850
${\bf \vdots }$
4.3850
4.3812
4.3812
-----------------
&
-----------------
4.333
4.331
4.329
4.327
4.326
4.325
4.325
4.324
${\bf \vdots }$
4.324
4.316
4.315
-----------------
&
-----------------
4.6906
4.6908
4.6905
4.6901
4.6899
4.6898
4.6897
4.6897
${\bf \vdots }$
4.6897
4.6858
4.6858
-----------------
\
[**Table 4:** ]{}Same as table 3 for $\kappa =-\,(\ell +1).$ Here we report the Charmonium masses where the mass series and Padé approximants $M_{i,\,j}$ stabilize.
$k,\,\ell $ $M(N)$ $M[i,j]$ $k,\,\ell $ $M(N)$ $M[i,j]$
------------- ------------- ----------------- ------------- ------------- -----------------
0,0 M(6)=3.0333 M\[4,4\]=3.0333 1,0 M(7)=3.65 M\[5,5\]=3.6502
0,1 M(5)=3.4918 M\[2,3\]=3.4918 1,1 M(7)=3.946 M\[4,4\]=3.9462
0,2 M(4)=3.7787 M\[2,3\]=3.7787 1,2 M(7)=4.1690 M\[4,4\]=4.1690
0,3 M(4)=4.0129 M\[2,3\]=4.0129 2,0 M(9)=4.08 M\[6,6\]=4.0789
0,4 M(4)=4.2177 M\[2,3\]=4.2177 2,1 M(9)=4.314 M\[4,5\]=4.3139
[**Table 5:** ]{}KG results for the [*funnel-shaped* ]{}potential $%
S(r)=br$ and $V(r)=-a/r$, with $m=1.370\,GeV$, $b=0.10429\,GeV^{2}$, $%
a=0.26, $ $E_{\hbar }$ represents the results of Kobylinsky et al \[3\] via $%
\hbar $-expansion (up to the third-order correction), and $E_{num}$ is the numerical integration value reported in \[3\].
[cccc]{} $E(N)$ & $\ \ \ \ \ \ \ k=0,\,\ell =0$ & $\ \ \ \,\ \ \ \ \ \ \ k=0,\,\ell
=1 $ & $\ \ \ \ \ \ k=0,\,\ell =2$\
-----------------
$E(1)$
$E(2)$
$E(3)$
$E(4)$
${\bf \vdots }$
$E(14)$
$E_{\hbar }$
$E_{num}$
-----------------
&
-----------------
1.541
1.535
1.534
1.533
${\bf \vdots }$
1.533
1.536
1.533
-----------------
&
-----------------
1.76167
1.76064
1.76037
1.76033
${\bf \vdots }$
1.76033
1.7604
1.760
-----------------
&
-----------------
1.90420
1.90388
1.90380
1.90379
${\bf \vdots }$
1.90379
1.9038
1.904
-----------------
\
[**Table 6:** ]{}PSLET results for $\check{E}$ of equation (44) along with the numerically predicted ones by Jena and Tripati \[6\] and the $1/N$-expansion method $E_{1/N}$ by Roy and Roychoudhury \[6\]. $N$ in $\check{E}(N)
$ denotes the number of corrections added to the leading term where the series stabilizes.
[cccc]{} $k,\,\ell $ & $\check{E}(N)$ & $E_{num}$ & $\ \ E_{1/N}$\
------
0, 0
1, 0
2, 0
0, 1
1, 1
0, 2
------
&
-----------------------
$\check{E}(2)=$1.2358
$\check{E}(7)=$1.3347
$\check{E}(4)=$1.3922
$\check{E}(1)=$1.3072
$\check{E}(4)=$1.3731
$\check{E}(1)=$1.3540
-----------------------
&
--------
1.2364
1.3347
1.3923
1.3071
1.3731
1.3544
--------
&
-------
1.240
1.340
1.398
1.309
1.411
1.358
-------
\
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We examine the effects of charge transfer inefficiency (CTI) during CCD readout on the demanding galaxy shape measurements required by studies of weak gravitational lensing. We simulate a CCD readout with CTI such as that caused by charged particle radiation damage in space-based detectors. We verify our simulations on real data from fully-depleted p-channel CCDs that have been deliberately irradiated in a laboratory. We show that only charge traps with time constants of the same order as the time between row transfers during readout affect galaxy shape measurements. We simulate deep astronomical images and the process of CCD readout, characterizing the effects of CTI on various galaxy populations. Our code and methods are general and can be applied to any CCDs, once the density and characteristic release times of their charge trap species are known. We baseline our study around p-channel CCDs that have been shown to have charge transfer efficiency up to an order of magnitude better than several models of n-channel CCDs designed for space applications. We predict that for galaxies furthest from the readout registers, bias in the measurement of galaxy shapes, $\Delta e$, will increase at a rate of $(2.65\pm 0.02)\times10^{-4}\textrm{yr}^{-1}$ at L2 for accumulated radiation exposure averaged over the solar cycle. If uncorrected, this will consume the entire shape measurement error budget of a dark energy mission surveying the entire extragalactic sky within about 4 years of accumulated radiation damage. However, software mitigation techniques demonstrated elsewhere can reduce this by a factor of $\sim10$, bringing the effect well below mission requirements. This conclusion is valid only for the p-channel CCDs we have modeled; CCDs with higher CTI will fare worse and may not meet the requirements of future dark energy missions. We also discuss additional ways in which hardware could be designed to further minimize the impact of CTI.'
author:
- 'Jason Rhodes, Alexie Leauthaud, Chris Stoughton, Richard Massey, Kyle Dawson, William Kolbe, Natalie Roe'
title: 'The effects of charge transfer inefficiency (CTI) on galaxy shape measurements'
---
Introduction
============
The past decade has seen enormous changes in the field of cosmology. A concordance cosmology in which the expansion of the universe is accelerating has been accepted (Spergel [et al.]{} 2007). This accelerated expansion was first demonstrated by observations of SN Ia (Riess [et al.]{} 1998; Perlmutter [et al.]{} 1998) and has been confirmed with other probes in recent years. This startling aspect of our Universe has prompted a wide variety of possible explanations (see, e.g., Caldwell 2004) and considerable effort has gone into developing concepts for dedicated space missions to probe the mysterious dark energy thought to be causing this accelerated expansion. These missions, which include the NASA/DOE Joint Dark Energy Mission and ESA’s Euclid mission, plan to use a variety of probes in order to constrain the properties of dark energy. It has become widely accepted that weak gravitational lensing, the small distortion of the observed shapes of background galaxies by foreground dark matter, is one of the most powerful probes of dark energy, provided that systematic effects can be controlled (Albrecht [et al.]{}2006). Space missions are attractive in large part due the greater ability to control many systematic effects (Rhodes [et al.]{} 2004a).
The field of weak lensing by large-scale structure, or cosmic shear, developed in parallel to the study of dark energy as the dominant component of the Universe. From the first detections of cosmic shear a decade ago (Wittman [et al.]{} 2000; Bacon, Refregier & Ellis, 2000; Kaiser, Wilson, & Luppino, 2000; Van Waerbeke [et al.]{} 2000), surveys have grown in size and thus information content (see Hoekstra & Jain 2008 for a recent review). The culmination of this effort will be in the execution of the above-mentioned dedicated dark energy missions, which plan to survey up to 20000 square degrees, the entire extragalactic sky. It has become clear that control of systematic effects, both observational and astrophysical, is of paramount importance in making use of weak lensing as a probe of dark energy. The subtle shape changes induced by weak lensing require exquisite control of observational systematic effects, especially knowledge of the telescope’s point spread function (PSF). From the ground, thermal and gravity load-induced fluctuations in the telescope can change the PSF, and the atmospheric seeing both broadens the PSF and makes the PSF unstable on the timescale of astronomical exposures. These effects can be largely or completely mitigated by making observations in a thermally stable environment above the atmosphere.
The control of systematics is a large factor in the drive towards a dedicated space mission. However, the harsh radiation environment of space has the potential to introduce an observational systematic effect due to charge traps created in CCD detectors by impacts from charged particles. These defects trap charge (either electrons in so-called n-channel CCDs or holes in p-channel CCDs) as charge is clocked across pixels toward the readout registers. When the charge is subsequently released from the trap, it shows up in a neighboring pixel, thus creating a trail along the readout direction. These trails obviously change the observed shapes of the galaxies in the images. These shape changes are coherent across the image, thus mimicking a weak lensing signal caused by dark matter. The degradation of charge transfer efficiency (CTE, this quantity is one minus the CTI, or charge transfer *inefficiency*) due to these radiation-damage induced traps has been observed in all Hubble Space Telescope (HST) imaging cameras: the Wide Field Planetary Camera 2 (WFPC2; Dolphin 2009), the Space Telescope Imaging Spectrograph (STIS; Goudfrooij [et al.]{} 2006), and the Advanced Camera For Surveys (ACS; Sirianni [et al.]{} 2005) and has hampered the sensitive shape measurements needed for weak lensing with those cameras (Rhodes [et al.]{}2004b; Schrabback [et al.]{} 2007; Rhodes [et al.]{} 2007). CTE degradation is a particularly difficult effect to correct for because its non-linear nature means that high signal-to-noise (S/N) stars which are typically used in weak lensing for PSF modeling will be affected less than the low S/N galaxies whose shapes are being measured. Thus, typical PSF deconvolution techniques are complicated by the effects of CTI. Thus, it is clear that future space weak lensing missions will need to minimize CTI due to radiation damage and have CCDs that are sufficiently well understood to allow for mitigation of the CTE degradation that does occur.
In this paper we carry out a quantitative analysis of the effects of CTI in CCDs on the analysis of galaxy shapes for weak lensing, and explore techniques to mitigate the shape distortions due to trailing charge. We have developed a detailed model for the effect of charge traps based on data from irradiated CCDs, and applied it to simulated galaxies. We quantify the effects on measured galaxy shapes in simulated data as a function of galaxy size, CTI, and S/N. We base our analysis on data from p-channel CCDs fabricated at Lawrence Berkeley National Laboratory (LBNL) and irradiated with protons at the LBNL 88” cyclotron (Dawson et al 2008; hereafter D08). However, our methods are general and can be applied to any CCDs if the density and time constants of the charge traps are known.
This paper is organized as follows. §\[sec:cte\] gives a brief overview of how charge is transferred between pixels during CCD readout and how different types of CCDs have performed after radiation damage. §\[sec:code\] describes the code we have developed to mimic the effects of CTI on CCD readout. In §\[sec:validation\] we validate that code by showing that it can reproduce the effects of CTE degradation as measured on real, irradiated LBNL CCDs. We apply our code to simulated astronomical images and detail the effects of CTI on the shapes of galaxies in §\[sec:simulations\]. We examine how this will effect the future space missions and give recommendations for mitigating the effects in §\[sec:future\]. Finally, we offer concluding remarks in §\[sec:conclusions\].
Background {#sec:cte}
==========
CCD readout and charge transfer
-------------------------------
During exposure, a photon incident on a CCD generates an electron-hole pair. In n-channel CCDs, the electrons drift into the potential well of the nearest pixel, which is created by an electrostatic potential gradient within the substrate. In p-channel CCDs, holes are collected instead, but we shall not distinguish between the two mechanisms hereafter. Charge (electrons or holes) accumulates in a well-defined volume, outside which the density falls rapidly to zero (Hardy, Murowinski & Deen 1998, Seabroke [et al.]{}2008). We are most concerned with the cross-sectional area of this cloud, which expands as a monotonic function of the amount of charge $n_e$. As illustrated in Figure \[fig:cte\_diagram\], we parameterize this as an effective height $h$ within the pixel; the variable need not really be the height, but that is a useful one-dimensional function for the purposes of explanation.
Some CCDs contain a supplementary buried channel or “notch” constructed at the bottom of the potential well. This is a small region of slightly lower potential and, like the channel in an artificial river bed designed to improve the flow of a small amount of water, it will concentrate the first few electrons (or holes). The notch reduces the number of traps the charge packet is exposed to, by confining the charge to a smaller region. We model the notch by setting the height $h$ to zero below some notch depth $d$, and $$h(n_e)=\left( \frac{n_e-d}{w-d} \right) ^{\alpha}$$ above the notch, where $w$ is the full well depth (the total amount of charge that will fit in a pixel before it overflows) and $\alpha$, which depends on the construction of the potential but which is typically $\sim0.5$ (see for instance, Chiaberge [et al.]{} 2009 and Mutchler & Sirianni, 2005). Setting $h$ to zero below the notch is an approximation because it is possible that there are some traps in the notch. At any rate, the CCDs we model in this paper have no notch in the imaging region and thus this approximation does not affect the results presented here.
At the end of the exposure, the charge must be moved to the amplifier electronics at the edge of the CCD to be read out and digitized. Each row of charge is first shuffled one pixel in the parallel direction, towards the serial readout register. This is typically accomplished using a 3-phase clock, as is the case for the CCDs described here (Janesick 2001), but can be approximated as a single operation (for a discussion of the consequences of this approximation, see Massey [et al.]{} 2009). Charge from the bottom-most row is transferred into the serial readout register and then shuffled using the same technique but in a perpendicular direction. The charge from each pixel is shifted onto a capacitor connected to an amplifier and then ‘counted’ by being sensed as a voltage and digitized. Read noise is the shot noise on this voltage and can be reduced by lengthening the sampling time, and thus slowing the readout rate. This process is then repeated once for each additional row of pixels: shuffling one row in the parallel direction and then through the serial readout register.
As the charge is transferred from pixel to pixel, charge traps due to defects or impurities in the Si lattice temporarily capture passing charge, and release it after some delay. The typical capture time is effectively instantaneous (and if it is not, the lower capture rate can be equivalently modeled as a lower density of charge traps). The probability of release is governed by an exponential decay, with a characteristic time constant that depends upon the properties of the lattice and impurities, as well as the operating temperature of the detector (Shockley & Read 1952, Hall 1952). Several different species of charge traps may be present in any given device, with different characteristic release times $\tau$, and densities $\rho$. If the captured charge is released after its charge cloud has been shuffled along, the released charge appears as a faint trail behind the main charge packet.
Trapping and trailing can occur in both the parallel and serial directions. Each column of pixels is independent and has a unique set of charge traps, while all of the rows share the same traps along the serial register. Note that even though the pixels and transfer mechanisms are physically similar in the two directions, the time per pixel transfer is typically $\sim 10^3$ shorter in the serial direction (D08). We demonstrate in §\[sec:validation\] that only charge traps with release times roughly similar to the clocking (time between transfers) affect galaxy shapes. Thus, separate species of charge traps can be important for parallel and serial transfers.
CTE Effects in Irradiated CCDs
------------------------------
Thick, fully depleted p-channel CCDs have several advantages over conventional thin, n-channel CCDs, including enhanced quantum efficiency at near-infrared wavelengths, reduced fringing at near-infrared wavelengths, and significantly less degradation of CTE with a given accumulated radiation exposure. This last effect is particulary important for the measurement of subtle weak lensing-induced galaxy shape distortions. It arises due to the fact that the divacancy traps that are primarily responsible for CTI in p-channel devices are more difficult to form than the phosphorous vacancy traps that occur due to radiation damage in n-channel CCDs (Bebek [et al.]{} 2002; Janesick & Elliott 1992; D08; Spratt [et al.]{} 2005). We focus our analysis on p-channel CCDs precisely because it is more difficult to form traps in these types of CCDs than in existing n-channel CCDs.
Marshall [et al.]{} (2004) compared the CTE responses of irradiated p-channel LBNL CCDs with n-channel CCDs. These n-channel devices are designed for space applications and are the ones used for the recently installed Wide Field Camera 3 on the HST. They found that a notch implant in the channels improved the CTE performance by a factor of 2 for both p-channel and n-channel devices. More importantly, they found that the CTE performance of p-channel devices is about an order of magnitude better than that of n-channel devices after irradiation. A re-analysis of the Marshall [et al.]{} data (Lumb 2009) indicates that the p-channel devices may only have about a factor of 3-8 better CTE (depending on the signal level, with the p-channel advantage being greater at low signal levels such as those expected in the images of faint galaxies). Likewise, a comparison by Gow [et al.]{}(2009) found a factor of 7 improvement in tolerance for parallel CTI (and 15 for serial CTI) in otherwise similar p and n-channel devices. Thus, thick, fully depleted p-channel CCDs are particularly attractive for a weak lensing space mission and we use them as the baseline for this study. We do, however, note that p-channel devices do not have the rich heritage that n-channel devices do, particularly in space applications.
LBNL has developed radiation-hardened CCDs with the specific application of dark energy missions in mind (Holland [et al.]{}2006). These CCDS are composed of $3512\times3512$ $10.5\mu$m pixels with 4 readout registers and were baselined for the SuperNova Acceleration Probe (SNAP), a JDEM concept (Bebek 2007). The SNAP mission, like other candidate dark energy missions, would be at the L2 Earth-Sun Lagrange point, and we use the radiation flux there as the baseline for the flux experienced by a dark energy mission. A dark energy observatory will experience significant radiation exposure at L2, primarily from solar protons. Exposure to energetic protons leads to degraded CCD performance due to bulk damage from non-ionizing energy loss (NIEL) and charging of oxide layers from ionizing radiation. Bulk damage in the Si lattice is the dominant effect that manifests itself through increased CTI, increased dark current, and isolated hot pixels. Of these effects, an increase in CTI is the most likely to introduce systematic errors to a weak lensing survey.
Software Algorithm to Mimic CTI {#sec:code}
===============================
Our model for CTI is that inefficient charge transfers are caused by discrete charge traps embedded in pixels. These charge traps can capture and release single electrons (or holes). Each pixel of the detector contains a number of charge traps, that have a mean density $\rho$. Each trap is characterized by $h_{trap}$, its vertical location in the pixel and $\tau$, its characteristic release time constant. Different species of charge traps have different decay time constants and thus different values of $\tau$ (and $\rho$). Whenever a trap in a pixel containing $n_e$ electrons (or holes) is within the electron cloud, [[*i.e.*]{}]{} $h_{trap}<h(n_e)$, we assume that it immediately absorbs one electron from the free charge. When a full charge trap is above the free charge height (that is, not within the electron cloud), it is given an opportunity to decay.
We have developed a code that mimics the readout of a CCD with imperfect CTI. For each column of pixels, we use the following procedure to “read out” the image and determine the observed charge in each pixel. This procedure is also illustrated graphically in Figure \[fig:flowchart\].
1. Populate each pixel with free charge, as accumulated during an exposure, and calculate $h(n_e)$.
2. Define the locations of charge traps throughout the pixel array. This is done for multiple species of charge traps, each with a different $\rho$ and $\tau$.
3. For charge traps with $h_{trap}<h(n_e)$ fill the trap with one electron, and subtract one electron from the free charge in its pixel.
4. Read out the charge in pixel row $n=1$.
5. For $n=1$ to $n=n_{max}$ set the free charge in pixel $n$ to the free charge in pixel $n+1$.
6. For full traps with $h^{trap} > h(n_e)$, calculate the probability that the trap will decay (that is, release the charge), based on an exponential decay with time constant $\tau$. Generate a random number in the range \[0,1\] and, if the probability is less than the random number chosen, empty the trap and increase the free charge by 1 electron.
7. Repeat the previous four steps to calculate the measured charge in pixel rows $n=2$, $n=3$, .... Note that $n_{max}$ is decreased by one for each iteration.
The image is then rotated by ninety degrees and exactly the same process is repeated to simulate serial transfers. The total number of operations to read out one row or column of $n_{pix}$ pixels is $(n_{pix})(n_{pix}+1)/2$. Thus, reading out out an entire $n_{pix} \times n_{pix}$ array scales as $n_{pix}^3$ and is computationally intensive. During a readout, a charge can be trapped multiple times in different pixels.
This process inevitably adds noise to an image because we do not know the true locations of individual traps within the detector, but simply model the traps as a uniform density. We therefore ought to perform many iterations and average the results. As a more computationally efficient solution to this problem, we instead introduce “fractional traps”. That is, we place the same number of traps in each pixel and let each trap capture a “fractional electron” (with the fraction being $\rho$, the trap density per pixel, which can be less than $1$). We further divide the fractional trap in each pixel into $n_{levels}$ fractional traps, each located at a vertical position in multiples of $1/n_{levels}$. These traps release charge exactly as described above. We have found that setting $n_{levels}=10000$ allows us to reproduce the averaged results of many iterations with full traps placed randomly within the pixels (mimicking a real, physical CCD). However, the fractional trap method saves considerable computational overhead when simulating the effects of CTI on many thousands of images as described in §\[sec:simulations\]. Our code allows each trap to have a different $\tau$. However, as discussed below, we find that in each charge transfer direction, a small number of $\tau$ values (trap species) describe the behavior of physical detectors.
Validation of the Charge Transfer Code {#sec:validation}
======================================
The charge transfer code has been tested and validated by creating simulated images designed to mimic the irradiated data used in D08. Using the same software as D08, we show that the code is able to reproduce the observed radiation damage effects as measured in the D08 data.
Irradiation in the Lab to simulate Space Radiation
--------------------------------------------------
The damaging particle radiation flux incident on a spacecraft depends on exactly when the mission occurs within the $\sim11$ year solar cycle (Barth [et al.]{} 2000). There is an order of magnitude difference in the flux of solar protons during the heaviest and lightest parts of the solar cycle (see, e.g. Figure 1 of Barth [et al.]{} 2000). In D08, the solar proton flux was modeled using the European Space Agency’s Space Environment Information System (SPENVIS). In SPENVIS, a simplified solar cycle consisting of 7 years near the maximum flux level and 4 years at zero flux is used. The total displacement damage (energy deposited in the silicon) is predicted from a uniform 4$\pi$ steradian spatial distribution of solar protons with an energy distribution derived from the Xapsos [et al.]{} (1999) model for solar proton emission. SPENVIS employs a statistical model based on data from previous solar cycles to predict the dose at $95\%$ confidence level (CL); that is, the prediction will underestimate the dose only $5\%$ of the time. D08 used the SPENVIS model to calculate the 95% CL solar proton flux incident on the CCDs, after passing through the shielding provided by the SNAP spacecraft and telescope, yielding an integrated NIEL exposure of $2.54 \times 10^6$ MeV/g (Si) for one year at solar maximum.
D08 then characterized the CTE performance of thick, fully depleted LBNL p-channel CCDs by irradiating several CCDs at the LBNL 88-Inch Cyclotron with $12.5$ and $55$ MeV protons. Although a variety of irradiation levels were used, we only consider the $12.5$ MeV data with an irradiation level of $2\times10^{10}$ protons/cm$^2$. The NIEL factor for $12.5$ MeV protons is $8.9\times10^{-3} \textrm{MeV}\textrm{g}^{-1}\textrm {cm}^{-2}$ per proton (Jun [et al.]{} 2003). Thus, the $2\times10^{10}$ protons/cm$^2$ flux of 12.5 MeV protons used to irradiate the CCDs corresponds to $1.78 \times 10^{10}$ MeV/g (Si,) a total dose equivalent to ten solar cycles at 95% CL, using the SPENVIS approximation detailed here, or 110 years at L2.
Although this accumulated radiation exposure is significantly higher than any proposed dark energy mission would encounter, the exaggerated radiation exposure makes it easier to characterize the detailed effects of CTE degradation. Since the number of traps (and thus the degradation of CTE) is linear with radiation exposure and the NIEL dose, the D08 accumulated radiation exposure can be used to estimate CCD performance over the course of a dark energy mission lifetime.
Analysis of CTI Due to Radiation Damage
---------------------------------------
The CTE of irradiated CCDs was measured using a $^{55}Fe$ x-ray source that emits K-alpha photons with energies of 5.9KeV. At the operating temperature of 133K, a single K-alpha x-ray will generate 1580 electron/hole pairs, which, depending on the location of the x-ray relative to the pixel potential wells, may be localized in a single pixel or shared among two or more pixels.
In D08, CTE was characterized using single pixel events from the K-alpha peak, and the results showed that the irradiated LBNL CCDs are three times more affected by charge trailing in the parallel readout direction that in the serial readout direction. In this article, we disregard the serial CTE and only consider the trailing in the parallel direction since this will most affect galaxy shape measurements.
The effects of irradiation on CTE was studied in two ways in D08. In the first method, called the $\it stacking$ method, CTE is characterized by the average charge collected for single pixel x-ray events as a function of the number of pixel transfers. Those x-ray events that experience more transfers lose a larger amount of charge due to CTI, as shown in Figure \[fig:test\_cte2\]. The serial and parallel CTE components are determined independently by fitting the fractional loss of each transfer to the data. The same single pixel x-ray events are used for a measurement of CTE using the $\it trailing$ method, in which the charge is counted in each trailing pixel as a fraction of the charge in the primary charge packet. The fractional trailing charge in each event is divided by the total number of transfers, and the results are averaged over all x-ray events. In other words, the averaged trails represent the fraction of charge left behind the primary charge packet for a single transfer. The effect of this trailing with a best fit to the data after irradiation and 1650 transfers is shown in Figure \[fig:test\_cte1\]. The total fractional charge integrated over these trails represents the CTI. To summarize and compare the two methods, the *stacking method offers a direct measurement of CTE*, measuring charge that is successfully transferred relative to the expected x-ray charge deposition, while the *trails method offers a direct measurement of CTI* by measuring the trailing charge relative to the charge in the leading pixel. The total fractional charge integrated over these trails represents the CTI.
In practice, the analysis of the trails following x-ray events is limited by the ability to measure the faint trails after a large number of transfers in the presence of non-zero read noise. Because of the low S/N at large distances from the primary x-ray event, D08 fit only the first 45 pixels with a two term exponential. The charge in this two-term exponential represents approximately $2/3$ of the total charge lost due to CTE effects as identified in the stacking method. In a re-analysis of the D08 data, in which stacking plots were made with varying x-ray flux, we find strong evidence that the remaining charge must be attributed to one or more populations of traps with a much longer time constant. One such candidate is the C-O trap which has been independently identified in the LBNL CCDs in a previous analysis (Bebek [et al.]{} 2002) with a de-trapping time constant of many seconds compared to the typical time between pixel row transfers of 25 ms at the $70,000$ pixel/sec readout speed employed in D08.
We therefore choose to model the CTI using three distinct trap populations instead of the two used by D08. We assign the third trap a time constant corresponding to 200 pixels in our analysis. The results of our fit are shown in Fig. \[fig:test\_cte1\] and the best-fit parameters are found to be: $\rho_1=0.35$, $\rho_2=0.49$, $\rho_3=0.7$, $\tau_1=10.0$, $\tau_2=0.486$, $\tau_3=200.0$, where $\tau$ is in units of ‘pixels’. These best fit parameters are tabulated in Table \[tab:params\_tab\]. There are, on average, about $1.5$ traps per pixel when summed over the three species; a mission that was at L2 for only about 5 years would thus have only about one twentieth that number of traps. We created a set of simulated images with the same characteristics of the D08 images to test our models. Overall, the agreement is good, as shown in Figure \[fig:test\_cte1\].
[ccccc]{}
Trap Species & $\rho$ (traps per pixel) & $\tau$ (pixels) & $\tau$ (ms) & probable defect type\
1 & 0.35 & 10.0 & 250 & Carbon-interstial (Ci)\
2 & 0.49 & 0.486 & 12 & divacancy (VV)\
3 & 0.7 & 200.0 & 5000 & carbon-oxygen (CO)\
As we show in §\[sec:simulations\], charge traps with time constants that are long relative to the time between parallel transfers in readout have negligible effect on galaxy shapes (i.e. they do not leave trails). The trapped charge is, however, removed from the object, and will thus affect photometry. Because our primary motivation is to study the impact of charge trailing on galaxy shapes, we chose to mimic the measured trails in D08 instead of matching the CTE as measured by the stacking method. We note that the opposite approach would be appropriate if our aim was to measure changes in photometry instead of shapes.
Application of the CTE Code to Simulated Galaxy images {#sec:simulations}
======================================================
We create simulated galaxy images with de Vaucouleurs and exponential profiles and use the code described in §\[sec:code\] to introduce the effects of CTI on the galaxy images. We create galaxies as they would appear in the I-band of the proposed SNAP mission, which has a 2 meter mirror, $0.1$” pixels and 400 second exposures. The background level is chosen to be the average background for extragalactic observations taken from L2. The measured background is slightly lower than the input background because the CTI causes the flux to be dragged out into the overscan region of the CCD during readout. We create single galaxies in each image to avoid having the traps in a pixel be filled by charge from an object that has already passed through that pixel during readout. All objects are placed 1650 pixels from the readout register (close to the maximum of 1712 pixels from the readout register). This is done because objects farthest from the readout registers will encounter the most traps and suffer the worst CTI; we are trying to estimate the worst case scenario for galaxy shape measurement.
We assume that the trap densities $\rho$ increase linearly over time, as traps accumulate due to radiation damage, and that the time release constants $\tau$ do not change because they are properties of the detector material itself. The assumption of linearity in trap density is not entirely accurate because the proton flux is dominated by solar radiation, which varies over the solar cycle. A specific analysis of any future dark energy mission will need to take into account the portion of the solar cycle in which the mission occurs. We make a further simplification by assuming that at the start of a mission a CCD will have no charge traps, so only traps accumulated during the time spent at L2 affect readout. This is of course not true, because real CCDs always have some imperfections even immediately after their production. In the case of the LBNL CCDs we are simulating, however, this turns out to be a good approximation because the pre-irradiation CTE is so high and the number of traps so low (see Table IV of D08). However, our results represent a best-case scenario for the number of traps (and thus CTI) as a function of time; the real CTI will be slightly worse.
For each galaxy, we measure the shape both before and after the image is degraded with imperfect CTE. The shape is parameterized in the typical weak lensing fashion by a two component ellipticity $e_i$, where $e_1=\frac{I_{xx}-I_{yy}}{I_{xx}+I_{yy}}$ corresponds to elongation along the $x$ axis (for positive $e_1$) or the $y$ axis (for negative $e_1$), and $e_2=\frac{2I_{xy}}{I_{xx}+I_{yy}}$ corresponds to elongation at $\pm45$ degrees. Here, the shapes are described in terms of the second order moments of the pixel intensity $I$ such that $I_{ij}=\frac{\sum I w x_{i} x_{j}}{\sum Iw}$ where $x_i$ is the distance in pixels from the object centroid and $w$ is a Gaussian weight function. The ellipticities were measured using the method of Rhodes, Refregier, & Groth (2000; hereafter RRG). This method has been well-tested on real and simulated space-based data (see Leauthaud [et al.]{} 2007). Since we are only interested in perturbations to the galaxy shapes, we do not go through the somewhat complicated steps of point spread function deconvolution, which can introduce biases in shape measurements (Heymans [et al.]{} 2006; Massey [et al.]{} 2007). Instead, we only concern ourselves with $\Delta
e$, which is relatively independent of the particularly shear measurement method. We explore the effects of S/N and galaxy size on $\Delta e$ but for the bulk of our analysis, we consider small, faint galaxies; any weak lensing survey will be dominated by galaxies that are faint and small relative to the PSF size.
For the purposes of this paper, we only introduce parallel CTI into the simulated images and we set the serial CTE equal to 100%. We do this for two reasons. First, the parallel CTI smears objects in the vertical direction (negative $e_1$), but serial CTI smears them horizontally (positive $e_1$, the serial readout direction). Thus, just using the ellipticity $e$ as an indicator of CTI-induced galaxy shape changes means that the effects of serial and parallel CTI partially cancel. The combined effects change the *size* of the PSF, and thus must be corrected for in real images, but would provide an unfair test for these purposes. The second reason we concentrate just on parallel CTI is that the parallel CTE degradation is three times worse for a given radiation exposure (D08) and different charge trap species affect the parallel and serial CTE because of the different clocking times in the parallel and serial directions. Thus, we seek only to demonstrate that we understand the more influential effects of the parallel CTI on shape measurement in this paper.
Effect of charge trap release time {#sec:tau}
----------------------------------
Figure \[fig:tau\_effect\] demonstrates the effect of charge trap release time $\tau$ on the measurement of photometry as measured by S/N (calculated via SExtractor; Bertin & Arnouts 1996, S/N =flux\_auto/fluxerr\_auto), astrometry (the $y$ centroid of an object), and shapes (rms size $d_{rms}=\sqrt{0.5(I_{xx}+I_{yy})}$ and ellipticity $e_{1}$). We measure the release time $\tau$ in units of “pixels,” the amount of time it takes to clock the charge by a certain number of pixels in the parallel readout direction (i.e., one “pixel” is the time between row shifts during readout).
In terms of astrometry, photometry, and size, there are two limiting regimes. Charge traps with very short release times (or slow CCD readout) push charge from an object’s leading edge onto its core, and drag core charge into a short tail. Both effects shift the object away from the readout register. The net effect also increases the object’s size, because the core contains more charge than the wings. A small amount of flux can be lost from the wings into a tail, so $\Delta$flux is always slightly negative. However, the smoothing inherent in trailing correlates adjacent pixels and has the perverse effect of [*increasing*]{} the S/N. Note that the limiting behavior at low $\tau$ is as expected: in our model, all traps inside a charge cloud carry an electron to the adjacent pixel at every clock cycle. In a real CCD, some charge may be released from very fast charge traps part-way through the 3-stage clocking cycle and returned to their original pixel. This process would lower the effective density of charge traps with low $\tau$.
Charge traps with long release times (or fast CCD readout) steal flux primarily from an object’s leading edge, and return it to the image in pixels well separated from the object. This stolen flux lowers the detection S/N. It also shifts the centroid as before, and decreases the size. For intermediate $\tau$, these effects are dominated by the addition of a tail, which increases the overall size. One curious dependency upon measurement method is that, while the rms size $d_{rms}$ decreases with $\tau$, the FWHM fitted by [SExtractor]{} [*increases*]{}: for example, $
\Delta$FWHM is negative for small $\tau$. This is presumably related to the net increase in detection S/N, and the segmentation of the image into fewer pixels that [SExtractor]{} determines belong to a given object. Note that if high $\tau$ were achieved by dramatically speeding the CCD readout, our assumption of instantaneous capture times may become invalid. A probabilistic capture mechanism over a finite time would result in lower effective densities of all traps, and potentially increased sensitivity to the density of charge throughout a pixel potential, changing the well filling parameters $\alpha$ and $d$.
The spurious ellipticity induced in an object is interestingly different. The tail and the centroid shift induced by charge traps with short release times both elongate an object in the readout direction. As the tail lengthens, the spurious ellipticity initially increases. However, once the charge in the tail is sufficiently disconnected from the object and the object’s centroid shifts back towards the correct position, the spurious ellipticity begins to decrease. In the limiting case of charge traps with very long release times, charge missing from the object’s leading edge could potentially elongate the object perpendicular to the readout direction; however, the residual centroid shift in this case is sufficient to maintain a small ellipticity in the readout direction (this result may depend upon the object’s radial profile).
Thus, we find that, in terms of weak lensing shear measurement, not all CTI is equally bad. Furthermore, there is a worst possible case, in which traps with release times corresponding to 3–4 clock cycles induce the most spurious ellipticity. This value depends upon the shape measurement method: with KSB (Kaiser, Squires & Broadhurst 1995) and RRG, it depends upon the size of the Gaussian weight function. The bump in $\Delta d(\tau)$ around this value is real and also depends upon this scale. However, from a more general argument about the dissociation of flux from an object in a very extended trail, it is clear that a local maximum in $|\Delta
e_1|$ will be inevitable for all shear measurement methods. The clock speed is a parameter that can be tuned in the hardware. We discuss this possibility in §\[sec:future\].
Effects on galaxy morphology
----------------------------
We create a series of images at the fiducial irradiation level of D08 (110 years at L2). We created galaxies with De Vaucouleurs (DvC) profiles and exponential profiles. For the DvC galaxies, we varied S/N, size, and input ellipticity. For each different set of simulation parameters, we create 1000 simulated galaxies, each with a different, random sub-pixel position of the galaxy centroid and different background noise realization. We measure the size in terms of the rms size $d$. Small galaxies have a size close to that of the PSF, representing the typical galaxies that will dominate a lensing survey; large galaxies are significantly bigger than the PSF. The values of S/N, size, and ellipticity for the different simulations are shown in Table \[tab:simstable\]. The key result for weak lensing, the change in measured ellipticity, is illustrated in Figure \[fig:trap\_density\].
[llccccc]{}
Profile & S/N & $\Delta$(S/N) & $e$ & $\Delta e$ & rms size $d_{rms}$ \[pixels\] & $|\Delta y |$\
DvC & low (13) & 1.1 & 0 & -0.028 & (small) 1.8 & 0.19\
DvC & low (13) & 1.1 & -0.17 & -0.024 & (small) 1.8 & 0.19\
DvC & low (13) & 1.2 & +0.17 & -0.031 & (small) 1.8 & 0.19\
DvC & high (50) & 0.7 & 0 & -0.024 & (small) 1.9 & 0.17\
DvC & low (20) & 0.3 & 0 & -0.010 & (large) 3.1 & 0.20\
DvC & high (50) & 3.0 & 0 & -0.010 & (large) 3.6 & 0.20\
Exponential & high (49) & 0.9 & 0 & -0.022 & (small) 1.9 & 0.18\
As expected, the small galaxies are significantly more affected by CTI than large galaxies. We also show that for small galaxies, brightness (S/N) is a mitigating factor (but not for large galaxies); small bright objects are slightly less affected by CTI than small faint ones. Another interesting feature recovered from these simulations is the dependence of $|\Delta e|=|\Delta e_{1}|$ on galaxy ellipticity. Galaxies that are already aligned along the readout direction ($e_{1}<0$ in the case of our simulations) are less affected by CTI than galaxies that are aligned perpendicular to the readout direction ($e_{1}>0$). Perturbations in the $y$ direction (such as CTI) affect galaxies that are already aligned in the $y$ direction less than galaxies aligned in the $x$ direction.
Effects of trap density {#sec:plots}
-----------------------
Figure \[fig:trap\_density\] also shows the results of simulations with varying trap densities. We increment the trap density (in all three species) from zero to the density that would be found after 220 years at L2 (twice the fiducial value from D08). At each of 50 evenly spaced points along this timeline, we create 100 simulated galaxies, each with a DvC profile (small, low S/N, $e=0$). As expected, the degradation of shape measurement increases linearly with trap density. We find that $$\label{eqn:delta_e_rate}
\frac{{\mathrm d}\Delta e}{{\mathrm d}t}=(2.65\pm 0.02)\times10^{-4} \textrm{~[yr}^{-1}\textrm{~at L2]}$$ for a radiation dosage averaged over an 11-year solar cycle and yearly displacement damage dose of $1.6 × 10^6$ MeV per gram of Si. We make the simplifying assumption of zero traps (CTI$=0$) at time $t=0$. However, D08 show that the number of traps in a new LBNL CCD is smaller than the measurement error, so the approximation we make here is a good one.
Consequences for the Future Space Missions {#sec:future}
==========================================
Amara & Refregier (2008) ascertain that the multiplicative error on measured shear needs to be kept below one part in $10^{-3}$ in order that future dark energy missions not be dominated by systematic errors. This means that *all* sources of shear measurements error (not just the portion due to CTI) must be kept below $\Delta e<10^{-3}$ throughout the mission lifetime. This level of shape measurement accuracy is represented in Figure \[fig:trap\_density\] by a horizontal dotted line. The prediction for $\Delta e$ after a fiducial 5 year mission can be calculated from Equation (\[eqn:delta\_e\_rate\]). To avoid any assumption of linearity with $\rho$, however, we have also run a larger number of simulations at 2 and 5 years of mean L2 exposure. In each case, we created 3000 simulations to reduce measurement noise due to the sub pixel galaxy position and sky noise. We find that $\Delta e_{\textrm{2
years}}=0.490\pm0.01\times10^{-3}$ and $\Delta e_{\textrm{5 years}}=1.56\pm
0.02\times 10^{-3}$. That is, without any correction, a 5 year weak lensing mission will have the entire shape measurement error budget consumed by CTI-induced effects before the end of the mission, even with specially designed, fully depleted, radiation hardened, p-channel CCDs.
Fortunately, recent work using data from the HST’s Advanced Camera for Surveys Wide Field Camera (ACS/WFC) has shown that software postprocessing can correct the effects of CTI on galaxy shapes by about a factor of 10 (Massey [et al.]{} 2009). Using the same code described here, trailed charge could be moved back to where it belonged in an iterative procedure restoring images to their true appearance. Massey [et al.]{} (2009) measured the time constants and number density of traps in ACS/WFC as a function of time, using extragalactic survey imaging that would be naturally available in any future survey without additional overhead. For this software, the factor of ten level of correction will be maintained down to the regime of future missions with much lower trap densities. One component of noise (due to variations in the number of traps in a given pixel, which we treat as a constant density) will be improved, but this affects only scatter in $\Delta e$ rather than the level itself. We therefore conclude that, *using CCDs with the characteristics of those in our study, and proven software mitigation techniques to achieve an additional factor of 10 correction, CTI in a future dark energy mission would be satisfactorily controlled at only 10% of the total shape error budget.*\
One way in which CTI models (and mitigation techniques) could potentially be pushed beyond correction by a factor of 10 would be to precisely locate individual charge traps, rather than treating them statistically. This would be most beneficial in the early years of a dark energy mission, when $\rho << 1$. Designing flexibility into the clocking speed, waveform, and voltage in CCD electronics provides the ability to locate traps via “pocket pumping” (Janesick 2001). Pocket pumping is a process in which a uniform level of charge introduced by a flat field lamp is rapidly shuffled back and forth thousands of times in the parallel direction. Following the charge shuffle, the charge is transferred to the readout transistor in the normal manner. The resulting image reveals accumulated charge captured and then released by each trap in the shape of a dipole of an overdensity neighboring an underdensity of charge. The orientation and strength of this dipole reveal the location of the trap within the pixel and the effectiveness of the trap. This can be repeated with different levels of initial charge to map out traps in 3 dimensions within the CCD. Because the readout time of the CCD is dominated by the clocking of serial charge, the pumping of charge in the parallel direction does not introduce a significant amount of overhead to the survey. For example, it takes approximately 45 seconds to record a normal image from the LBNL CCDs using the D08 clocking parameters. Image acquisition takes an additional 45 seconds for 20000 cycles of pocket pumping with a five pixel shift. Acquisition of five successive pocket-pumping images every week to reject cosmic rays and model the traps therefore would account for less than ten minutes of additional calibration time.
Another suggested mitigation technique is “charge injection” (also called “fat zero” or “pre-flash”) in which charge is placed into the pixels in order to fill the traps. The charge injection is simply a method to increase the overall background level and fill the volume. However, this has the effect of increasing the background in an image (much like increasing the zodiacal background) and will reduce the S/N of the detected objects. This is obviously undesirable for a weak lensing experiment in which the observer is attempting to measure shapes of faint galaxies.
We showed in §\[sec:tau\] that there exist clocking time scales that are maximally bad for shape measurement. If the clock cycle is 3–4 times the trap release time, then $\Delta e$ is maximized. Thus, future missions with weak lensing as a primary science driver should include an optimization of the charge clocking time in their CCD readout electronics. Increasing the rate at which charge is clocked serially, and thus increasing the rate at which parallel transfers of rows can be made, increases readout noise, resulting in an effective loss of survey depth; decreasing the rate at which charge is serially clocked increases the readout time and, if that dominates over factors like slew and settle time between exposures, will reduce survey area. Thus, careful consideration must be paid to the trade-offs in any such optimization. Furthermore, we are only making recommendations for how to minimize shape measurement errors due to CTI. Photometry is also degraded by traps with large values of $\tau$ relative to the charge clock period. This can remove charge from objects but place it far enough away that the shape is not significantly affected. Thus, trade-offs in charge clocking time must also take into account the photometric accuracy requirements of the mission.
We finally note that the temperature at which detectors are operated at has significant effect on CTE, and thus future missions should be tested and optimized with this in mind.
Conclusions {#sec:conclusions}
===========
We have quantified the effect of CTI on measurements of weak gravitational lensing. We first simulated the transfer of charge within LBNL p-channel CCDs using a model for charge traps with three characteristic release times to reproduce the experimental results of D08. Using this model, we then simulated deep exposures of galaxies for weak lensing measurements from a space-based telescope subject to radiation damage. The resulting simulated data were used to quantify the effects of radiation damage on shape measurements of galaxies of various sizes and S/N levels, the true shapes of which were precisely known. Most galaxies in any weak lensing survey will be small and faint; as expected, we find that these suffer worst from the effects of CTI.
The level of CTI-induced shape error $\Delta e$ will approach the total shape error budget of a dark energy mission (1 part in $10^{-3}$; Amara & Refregier 2008) after less than 4 years of radiation exposure at L2. Software mitigation techniques in image postprocessing, already proven on HST data (Massey [et al.]{} 2009), will be able to reduce the levels of shape error well below mission requirements. We have also suggested hardware capabilities, such as “pocket pumping” and adjustments to the readout speed, that may provide additional help. However, our numerical results are only valid for p-channel devices, whose CTE characteristics after radiation exposure have been shown to be superior to more common n-channel devices (Lumb 2009; Marshall [et al.]{} 2004). *Given the necessity of both hardware and software mitigation of CTI effects for successful mission operation, we recommend that future spacecraft be designed with detectors and mission parameters that ensure CTE characteristics no worse than the LBNL p-channel devices we simulated for this work.*\
There are several caveats to our results. First, we assumed that all galaxies are small, faint, and lie far from the readout register. Galaxies that are bright, large, or nearer the readout register will suffer less from CTI. Indeed, the average distance to the readout register will be exactly half of the worst case scenario we outlined, so the mean $\Delta e$ will be a factor of two lower. In a real mission, there will be several dithered exposures of each galaxy, with each dither placing the galaxy a different distance from the readout register; this may allow us to further model the effects of CTI on shapes and partially mitigate those effects. We also assumed a high level of radiation from the Sun for at least a portion of the mission due to modeling radiation flux in the heaviest part of a typical 11-year solar cycle. A mission flown during the level of minimum particle radiation at L2 would suffer less radiation damage, but a mission flown entirely during the maximum of the solar radiation flux would suffer more damage. This is a large uncertainty because the radiation flux due to the Sun can vary by an order of magnitude over the solar cycle (Barth [et al.]{} 2000). Furthermore, any real mission would need to adjust the flux according to the planned shielding on the spacecraft (D08 assume the SNAP design) and should take into account secondary particle cascades from reactions of high energy radiation with the shielding material (D08 ignored such secondary radiation). The ‘stacking’ and ‘trails’ methods probe different properties of CTE and we have chosen the fits to trails in D08 to model the effects of radiation exposure on shape measurements. We have only explored the charge transfer and radiation tolerance properties of a certain model of CCD operating at a single temperature; any future space-based weak lensing missions should undertake a similar analysis using the CCDs planned for that mission. Finally, if the CCDs contain a significant density of traps even before they are launched into the harsh radiation environment of space, the CTI will be worse throughout, and the useful mission lifetime reduced. Clearly, then, this paper is a first step, and any future mission should use a procedure similar to the one we have developed in this paper to meet specific mission requirements by fully optimizing its choice of CCDs, clocking rate, and shielding.
[**Acknowledgments**]{}
This work was supported in part by the Jet Propulsion Laboratory, operated by the California Institute of Technology under a contract with NASA. This work was also supported by the United States Department of Energy under contract No. DE-AC02-05CH11231. We thank Chris Bebek, Mike Lampton, Michael Levi, and Roger Smith for useful discussions about CCDs and CTE. AL acknowledges support from the Chamberlain Fellowship at LBNL and from the Berkeley Center for Cosmological Physics. RM is supported by STFC Advanced Fellowship \#PP/E006450/1 and FP7 grant MIRG-CT-208994. CS was supported by funding from the Office of Science at LBNL and Fermilab.
Albrecht, A., [et al.]{}, 2006, astro.ph..9591A Amara, A., & Refregier, A., 2008, MNRAS, 391, 228 Bacon, D., Refregier, A., & Ellis, R., 2000, MNRAS, 318, 625 Barth, J., Isaacs, J., & Poivey, C., 2000,“The Radiation Environment for the next generation space telescope,” GSFC Bebek, C., 2007, NIMPA, 579, 848 Bebek, C., [et al.]{}, 2002, SPIE 4669, 161-171 Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393 Caldwell, R., 2004, MPLA, 19, 1063 Chiaberge, M., [et al.]{}, 2009, Instrument Science Report ACS 2009-01, STScI Dawson, K., [et al.]{}, 2008, IEEE Transactions on Nuclear Science, 55, 1725 Dolphin, A., 2009, PASP, 121, 655 Goudfrooij, P., [et al.]{}, 2006, PASP 118, 1455 Gow, J., [et al.]{}, 2009, SPIE,7435, 74350F-2 Hall R., 1952, Phys. Rev. 87, 387 Hardy T., Murowinski R. & Deen M., 1998, IEEE Trans. Nuclear Sci. 45, 154 Heymans, C., [et al.]{}, 2006, MNRAS, 368, 1323 Hoekstra, H., & Jain, B., 2008, ARNPS, 58, 99 Holland, S.E., [et al.]{}, 2006, SPIE, 6276, 10H Dawson, K., [et al.]{}, 2003, IEEE Trans. Electronics Dev., 50, 225 Janesick, J., [et al.]{}, 1991, SPIE, 1447, 87 Janesick, J., & Elliott, T., 1992, ASPC, 23, 1J Janesick J., “Scientific Charge Coupled Devices”, 2001, SPIE (ISBN 0-8194-3698-4) Jun, I. [et al.]{}, 2003, NSS-IEEE, 50, 192. Kaiser, N., Squires, G., & Broadhurst, T., 1995, ApJ, 449, 460 Kaiser, N., Wilson, G., & Luppino, G., astro-ph/0003338 Leauthaud, A., [et al.]{}, 2007, ApJS, 172, 219 Lumb, D., 2009, SPIE, 743904 Massey, R., [et al.]{}, 2007, MNRAS, 376, 13 Massey, R., [et al.]{}, 2009, submitted Marshall, C., [et al.]{}, 2004, SPIE, 5499,542 Mutchler, M., & Siriani, M. 2005, Instrument Science Report ACS 2005-03, STScI Perlmutter, S., [et al.]{}, 1998, ApJ, 517, 565 Rhodes, J., [et al.]{}, 2007, ApJS, 172, 203 Rhodes, J., [et al.]{}, 2004, APh, 20, 377 Rhodes, J., [et al.]{}, 2004, ApJ, 605, 29 Rhodes, J., Refregier, A., & Groth, E., 2000, ApJ, 536, 79 Riess, A., [et al.]{}, 1998, AJ, 116, 1009 Schrabback, T., [et al.]{}, A&A, 468, 823 Seabroke G., [et al.]{}, 2008, SPIE, 7021, 49S Shockley W. & Read W., 1952, Phys. Rev. 87, 835 Sirianni, M., [et al.]{}, 2005, PASP 117, 1049 Spergel, D., [et al.]{}, 2007, ApJS, 170, 377s Spratt, J.P., [et al.]{} 2005, IEEE Trans Nuc Sci, 52, 6, 2695 Van Waerbeke, L., [et al.]{}, 2000, A&A, 538, 30 Wittman, D., [et al.]{}, 2000, Nature, 405, 143 Xapsos M.A., [et al.]{}, 1999, NASA/TP-1999-209763.
| {
"pile_set_name": "ArXiv"
} |
---
author:
- 'A. Antoniadis-Karnavas , S. G. Sousa, E. Delgado-Mena, N. C. Santos, G. D. C. Teixeira, V. Neves'
date: 'Received / Accepted'
title: 'ODUSSEAS: A machine learning tool to derive effective temperature and metallicity for M dwarf stars'
---
[The derivation of spectroscopic parameters for M dwarf stars is very important in the fields of stellar and exoplanet characterization. The goal of this work is the creation of an automatic computational tool, able to derive quickly and reliably the T$_{\mathrm{eff}}$ and \[Fe/H\] of M dwarfs by using their optical spectra, that can be obtained by different spectrographs with different resolutions.]{} [ODUSSEAS (Observing Dwarfs Using Stellar Spectroscopic Energy-Absorption Shapes) is based on the measurement of the pseudo equivalent widths for more than 4000 stellar absorption lines and on the use of the machine learning Python package “scikit-learn” for predicting the stellar parameters.]{} [We show that our tool is able to derive parameters accurately and with high precision, having precision errors of $\sim$30 K for T$_{\mathrm{eff}}$ and $\sim$0.04 dex for \[Fe/H\]. The results are consistent for spectra with resolutions between 48000 and 115000 and SNR above 20. ]{}
Introduction
============
Spectra can be used to reveal the chemical composition of the stars, as well as important stellar atmospheric parameters, such as effective temperature (T$_{\mathrm{eff}}$) and \[Fe/H\]. These parameters are crucial for the characterization of the stars and therefore fundamental to understand their formation and evolution. Furthermore, they influence the properties of the planets forming and orbiting around them [@ever13]. However, the spectroscopic analysis to derive these parameters has some difficulties to overcome. One of the main problems is the correct determination of the spectral continuum, which is more problematic in cool and faint stars, such as M dwarfs. Their study is quite difficult and complicated, compared to FGK stars, since in M dwarfs, molecules are the dominant sources of opacity. These molecules create thousands of lines that are poorly known and moreover many of them blend with each other. Therefore, the position of the continuum is hardly identified in their spectra.
Methods which rely on the correct determination of the continuum, work better only for the metal poor and earliest types of M dwarfs [@woolf05]. Methods using spectral synthesis have not achieved as precise results as in FGK cases, because of the poor knowledge of many molecular line strengths. Recently, spectral synthesis in the near infrared has presented advances, as shown by several studies. [@one12; @lind16; @raj18; @pass19].
Regarding these limitations, most attempts for determining effective temperature and metallicity, are done with photometric calibrations [@bon05; @johnapps09; @neves12] or spectroscopic indices [@rojas10; @rojas12; @mann13a]. Metallicity uncertainties range from 0.20 dex using photometric calibrations, to 0.10 dex by using spectroscopic scales in the infrared [@rojas12]. For T$_{\mathrm{eff}}$, precisions of 100 K are reported, but significant uncertainties and systematics are still present, ranging from 150 to 300 K. [@casag08; @rojas12].
One of the most popular methods to derive atmospheric stellar parameters for FGK stars is by measuring the equivalent widths (EW) of many metal lines of the spectrum. @neves14 using the MCAL code, measured pseudo EWs in the optical part of the spectrum for 110 M dwarfs observed in the HARPS GTO M dwarf program, by setting a pseudo continuum for each line. They proceeded to the derivation of T$_{\mathrm{eff}}$ and \[Fe/H\] of these stars applying a calibration based on reference photometric T$_{\mathrm{eff}}$ and \[Fe/H\] scales that exist for 65 of them from @casag08 and @neves12 respectively. In the first case, the reference T$_{\mathrm{eff}}$ is the average value of the V - J, V - H, and V - K photometric scales as seen in @casag08, while for \[Fe/H\] the calculation of its reference values was done using stellar parallaxes, V and Ks magnitudes as described in @neves12.
Machine learning is an increasingly popular concept in several fields of science. It can be accurate in predicting outcomes without the need of the user explicitly creating a specific model to the problem at hand. The algorithms in machine learning receive input data and by applying statistical analysis, they predict an output value within a reasonable range. The interest for machine learning algorithms and automatic processes in astronomy is emerging from the increasing volume of survey data [@how17]. It can be applied to a wide range of studies, with the input attributes being for example the photometric properties of the sources [@dasa19; @akras19; @rau19; @ucci19].
In our work, we follow the pseudo-EW approach. We present our tool ODUSSEAS (Observing Dwarfs Using Stellar Spectroscopic Energy-Absorption Shapes), which makes use of the machine learning “scikit learn” package of Python. It offers a quick automatic derivation of T$_{\mathrm{eff}}$ and \[Fe/H\] for M dwarf stars, by being provided with their 1D spectra and their resolutions. The main advantage of this tool, compared to other ones that derive stellar parameters such as the MCAL code by @neves14 (which is limited to HARPS range and needs manual adjustment of results for different resolutions), is that it can operate simultaneously in an automatic fashion for spectra of different resolutions and different wavelength ranges in the optical. It is based on a supervised machine learning algorithm, meaning that it is provided with both input and expected output for creating a model. This input to the machine learning function are the values of the pseudo EWs for 65 HARPS spectra and the expected output are the values of their reference T$_{\mathrm{eff}}$ and \[Fe/H\] from @casag08 and @neves12 respectively. After training with a part of these HARPS data, the algorithm produces a model and tests it on the rest of the HARPS data. It predicts their values and compares them with the reference ones given as expected output. Thus, it examines the accuracy and the precision of the model by using several regression metrics described later. Finally, it applies the model to unknown spectra and estimates their stellar parameters.
In Sect. \[EWsection\] we describe how the tool computes the pseudo EWs. In Sect. \[mlsection\] we describe our tool and the flow of its process. We explain the characteristics of the machine learning function and its efficiency regarding different regression types, resolutions and wavelength areas. In Sect. \[otherspec\] we apply our tool to spectra obtained by several spectrographs of various resolutions and we examine the results. Finally, Sect. \[sum\] summarizes the work presented in this paper.
Pseudo-EW measurements {#EWsection}
======================
Since the identification of the continuum is very difficult in the spectra of M dwarfs, we follow the way of setting a pseudo continuum in each absorption line. The method is based on measurements of the pseudo EWs of absorption lines and blended lines in the range between 530 and 690 nm. We have excluded the parts where the activity-sensitive Na doublet and H$\alpha$ lines and strong telluric lines reside. The linelist consists of 4104 features. It is given in the form of left and right boundaries, between which these absorption features are supposed to be created. This method, based on pseudo EWs and the specific linelist, was used by @neves14.
We have created our own Python version of the method to compute the pseudo EWs. Our code reads the linelist and the 1D fits files of the stellar spectra. We have set an option for radial-velocity correction of the input spectra by our code, in the case they are shifted. Then, for each line, it identifies the position of the minimum flux of the feature, which is the central absorption wavelength. Starting from it, the code identifies the maximum in each side of this absorption feature, after having cut this spectral area at the range defined by the respective boundaries provided in the linelist. Eventually, it fits the pseudo continuum along the edges of the absorption feature with a straight line and it obtains the pseudo EW by calculating the area between the pseudo continuum and the flux. Mathematically, the pseudo EW is defined as following, where F$_{pp}$ is the value of the flux between the peaks of the feature (i.e. the pseudo continuum) and F$_{\lambda}$ is the flux of the line at each integration step.
$${\mathrm{pseudoEW}} = \Sigma \frac{(F_{pp}-F_{\lambda})}{F_{pp}} \Delta{\lambda}$$
We present such example in Fig. \[EWs\] where we use the star Gl176 and an absorption line at the region around 6530 $\AA$.
\
The evaluation of our pseudo-EW measurements, by comparing them with the ones obtained from MCAL code, is presented at Appendix \[A\].
Machine learning on M dwarfs {#mlsection}
============================
We base our tool for the derivation of T$_{\mathrm{eff}}$ and \[Fe/H\] on the machine learning concept. The user needs to run two codes. The “HARPS$\_$dataset.py” creates the databases which contain pseudo-EW measurements in different resolutions and the reference stellar parameters. The “ODUSSEAS.py” measures the pseudo EWs of new stellar spectra and derives their unknown T$_{\mathrm{eff}}$ and \[Fe/H\] via machine learning. Below, we explain the details of their structure, describing the input parameters and how to use the codes.
The HARPS dataset
-----------------
Each time the code “HARPS$\_$dataset.py” runs, the outcome is a file which is used later as input to the machine learning algorithm when running “ODUSSEAS.py” for training the machine and testing the generated model. It contains the names of 65 stars of the HARPS M dwarf sample, the central wavelengths of the 4104 absorption features from 530 to 690 nm, their pseudo-EW values according to the resolution we convolve the spectra and their reference values of T$_{\mathrm{eff}}$ and \[Fe/H\] from @casag08 and @neves12 respectively. All of these 65 spectra have SNR above 100, as reported by @neves14. They are presented in Table \[refparam\]. The range of the reference stellar parameters is presented in Fig \[range\]. Their photometric derivations have uncertainties of 100 K for T$_{\mathrm{eff}}$ and 0.17 dex for \[Fe/H\], as reported by @casag08 and @neves12 respectively.
\
The convolution function we use is the “instrBroadGaussFast” of “pyAstronomy” (<https://github.com/sczesla/PyAstronomy>), which applies Gaussian instrumental broadening. The width of the kernel is determined by the resolution. A description of it can be found at <https://www.hs.uni-hamburg.de/DE/Ins/Per/Czesla/PyA/PyA/pyaslDoc/aslDoc/broad.html>.
Since the HARPS spectra have a specific finite resolution, our code calculates the actual resolution to which they need to be convolved by the function, in order to get spectra to the final resolution we really want. This calculation is done considering the following relation: $$\sigma _{conv} = \sqrt{\sigma _{final}^{2} - \sigma _{orig}^{2}}$$
where $\sigma_{conv}$ corresponds to the resolution to which we need to convolve a spectrum with original resolution of $\sigma_{orig}$, in order to get a final resolution of $\sigma_{final}$.
The settings input by the user are two. a) Choose whether or not to convolve the reference HARPS spectra to the spectral resolution of our new data. We already provide precomputed pseudo EWs for a range of spectral resolutions in widely used spectrographs. In that case there is no need to convolve again the spectra and recalculate the pseudo EWs. b) The resolution of the data we want to analyse.
The “HARPS$\_$dataset.py” is presented schematically in Fig. \[HARPSdia\].
$\begin{array}{c}
\includegraphics[width=11cm]{HARPS_flow.png} \\
\end{array}$
ODUSSEAS tool
-------------
“ODUSSEAS.py” makes use of two algorithms that we developed: the “New$\_$data.py”, for measuring the pseudo EWs of new spectra to analyze, and the “MachineLearning.py” for the derivation of their T$_{\mathrm{eff}}$ and \[Fe/H\]. The innovative aspect of this tool is the simultaneous predictions for spectra of different resolutions and wavelength ranges.
The user has the option to activate the automatic radial velocity correction for the spectra if they are shifted. In addition, the user can set the regression type to be used by the machine learning process. The “ridge” is recommended, but also “ridgeCV” and “linear” work at similar level of efficiency as well. We present the efficiency of all the regression types used in Sect. \[mleff\].
The workflow of “New$\_$data.py” is similar to the “HARPS$\_$dataset.py”. It reads the files and resolutions of new spectra and, if needed, it calculates and corrects their radial velocity shift. In addition, if the original step of a spectrum is not 0.010, i.e. equal to that of the HARPS dataset, it is changed with linear interpolation to this value. Thus, the pseudo EWs are measured in a consistent way. The files containing the pseudo-EW measurements of each spectrum are then used during the operation of “MachineLearning.py”, which returns the values of T$_{\mathrm{eff}}$ and \[Fe/H\] along with the regression metrics of the models that predicted them.
The diagram of “ODUSSEAS.py” is presented in Fig. \[Mdwarfsdia\] showing concisely its inputs, operations and output.
$\begin{array}{c}
\includegraphics[width= 11cm ]{ODU_flow.png} \\
\end{array}$
Machine learning function {#mlf}
-------------------------
Here we present in more detail the machine learning function. The machine learning algorithm operates in a loop for each star separately, as each star may have different wavelength range and different resolution. For each star in the filelist, it loads automatically two files: the HARPS dataset of respective resolution, for training and testing the model, and the pseudo EWs of the star for which we want to derive T$_{\mathrm{eff}}$ and \[Fe/H\], in order to apply the model and return the stellar parameters. Based on the wavelength range that each spectrum has, a mask is applied on the HARPS dataset for considering the absorption lines in common.
The 65 HARPS stars split into training group consisting of the 70% of the sample (45 stars) and into testing group consisting of the remaining 30% of the population (20 stars). With these numbers selected, the machine learning model can be both trained accurately and tested on a sufficient number of stars.
We provide the algorithm with different regression types that can be used: the “linear”, the “ridge”, the “ridgeCV”, the ”multi-task Lasso“ and the ”multi-task Elastic Net". All these kinds of models provide an output value by fitting a linear regression to the input values. The relation between the predicted value *y* (the stellar parameter), the input variables *x* (the pseudo EWs) and the coefficients *w* is expressed as $$y(w,x)=w_{o}+w_{1}x_{1}+...+w_{p}x_{p}$$
The mathematical details of each regression type are described in the official online documentation at <https://scikit-learn.org/stable/modules/linear_model.html>.
The performance of machine learning is indicated by the following three kinds of regression metrics that are returned. The mean absolute error is computed when the model is applied on the test dataset. It corresponds to the expected value of the absolute error loss in the predictions. In addition, the “explained variance score” is calculated. The best possible value of this score is 1.0. Variance is the expectation of the squared deviation of a random variable from its mean. It measures how far a set of numbers are spread out from their average value. Furthermore, the “r2 score” computes the coefficient of determination, defined as R$^2$. The coefficient of determination is the proportion of the variance in the dependent variable that is predictable from the independent variables. This score provides a measure of how well future samples are likely to be predicted by the model. Best possible score is 1.0 too. A constant model that always predicts the expected value, disregarding the input features, would get a score of 0.0. In our case of multi-output, the resulting “explained variance” and “r2” scores are by default the averages with uniform weight of the respective scores for T$_{\mathrm{eff}}$ and \[Fe/H\]. The mathematical types of those regression metrics are described in their official online address at <https://scikit-learn.org/stable/modules/model_evaluation.html#regression-metrics>.
For each star, the tool makes 100 determinations by splitting randomly the train and test groups each time. After these determinations, it returns the average values of T$_{\mathrm{eff}}$ and \[Fe/H\], the average values of the mean absolute errors of the models, the average scores of machine learning and the dispersion of T$_{\mathrm{eff}}$ and \[Fe/H\] (measured as the standard deviation). This iterative process minimizes the possible dependence of the resulting parameters on how the stars from the HARPS dataset are split for training and testing in one single measurement. Since the reference stars are only 65, which stars end up in the training set could change the results in a measurement. This is the reason we do these multiple runs with shuffling and splitting the reference stars in different train and test groups, and finally we calculate the average values and the dispersion. The final results are automatically saved in the file called “Parameter$\_$Results.dat”. Moreover, it saves a group of plots with the reference and the predicted parameters of model testing, as well as their differences, as a visualization of the model accuracy. An example is presented at Fig. \[ml\].
$\begin{array}{c}
\includegraphics[width=\hsize]{Teff_test_comparison.pdf} \\
\includegraphics[width=\hsize]{Diff_Teff_test_comparison.pdf} \\
\\
\includegraphics[width=\hsize]{FeH_test_comparison.pdf} \\
\includegraphics[width=\hsize]{Diff_FeH_test_comparison.pdf} \\
\end{array}$
Machine learning efficiency {#mleff}
---------------------------
Firstly, we test the regression models mentioned above to find the best one. We use the original spectra of the HARPS dataset to their real resolution of 115000. For 100 runs with each regression type, we measure the scores and the absolute mean errors of the stellar parameters on the test set. We report the average values around which each model tends to result in Table \[reg\]. The “linear”, “ridge” and “ridgeCV” work very well in general, having “r2” and “explained variance” scores with average values around 0.93 and 0.94 respectively. The range of these scores, in the 100 runs, is usually from 0.87 to 0.99. The average uncertainties of those regression types are $\sim$27 K for T$_{\mathrm{eff}}$ and $\sim$0.04 for \[Fe/H\]. The “ridge” model has slightly greater scores than the “linear” one. “RidgeCV”, which has a built-in cross validation function that applies “leave-one-out” or “k-fold” strategies, does not seem to work better than the classic “ridge” one, at least in this sample of M dwarf measurements. Furthermore, “multi-task Elastic Net” and “multi-task Lasso” give considerably lower scores and higher mean absolute errors. Thus, we suggest “ridge” regression, as it operates best on the spectral values of the M dwarfs.
Secondly, we evaluate the “explained variance” and “r2” scores and the mean absolute errors of the algorithm for different resolutions of the spectra. We do it for the HARPS dataset at its actual resolution of 115000 and we repeat this test for convolved datasets at resolutions of other broadly used spectrographs: 110000 (UVES), 94600 (CARMENES) , 75000 (SOPHIE) and 48000 (FEROS). This is done to examine the level of machine learning precision towards lower resolutions. After 100 measurements of each case, we present the average values at Table \[reso\].
To further test the reliability of the method, we examine the efficiency of the machine learning in different wavelength ranges of the spectrum. We divide the linelist, which is from 530 to 690 nm, in four spectral regions and we calculate the respective scores and mean absolute errors. We do this test to check if machine learning works better using the full range or a specific part of the wavelengths. For this test, we use the case of the convolved data at the resolution of 110000. The machine learning operates at its best while using the full range of the initial linelist. In addition, regarding the divided areas, we notice that the bluer the part the higher the scores and the lower the mean absolute errors respectively. In general, the results show that we can get highly precise predictions for stars observed at any part of the 530-to-690 nm spectrum. These results are presented in Table \[linelist\].
[ccccc]{}\
Regression & r2 score & E.V. score & M.A.E. T$_{\mathrm{eff}}$ & M.A.E. \[Fe/H\]\
& & & \[K\] & \[dex\]\
\
Ridge & 0.93 & 0.94 & 27 & 0.037\
RidgeCV & 0.93 & 0.94 & 27 & 0.038\
Linear & 0.93 & 0.93 & 27 & 0.039\
Multi-task Elastic Net & 0.91 & 0.92 & 35 & 0.045\
Multi-task Lasso & 0.88 & 0.89 & 41 & 0.056\
\
\[reg\]
[ccccc]{}\
Resolution & r2 score & E.V. score & M.A.E. T$_{\mathrm{eff}}$ & M.A.E. \[Fe/H\]\
& & & \[K\] & \[dex\]\
\
real 115000 & 0.93 & 0.94 & 27 & 0.037\
conv. 110000 & 0.93 & 0.94 & 28 & 0.038\
conv. 94600 & 0.93 & 0.93 & 28 & 0.039\
conv. 75000 & 0.93 & 0.93 & 29 & 0.041\
conv. 48000 & 0.92 & 0.93 & 30 & 0.043\
\
\[reso\]
[cccccc]{}\
Wavelength range & Number of lines & r2 score & E.V. score & M.A.E. T$_{\mathrm{eff}}$ & M.A.E. \[Fe/H\]\
(nm) & & & & \[K\] & \[dex\]\
\
530 - 690 & 4104 & 0.93 & 0.94 & 28 & 0.038\
530 - 580 & 1300 & 0.92 & 0.93 & 31 & 0.039\
580 - 630 & 1300 & 0.91 & 0.91 & 48 & 0.044\
630 - 690 & 1504 & 0.89 & 0.90 & 56 & 0.048\
\
\[linelist\]
Derivation of stellar parameters {#otherspec}
================================
We apply our tool to spectra obtained by five widely used instruments of different resolutions: HARPS of 115000, UVES of 110000, CARMENES of 94600, SOPHIE of 75000 and FEROS of 48000. The spectra were taken from the respective public data archives. To test the efficiency of our tool on other-than-HARPS instruments, we use spectra from stars in common with the HARPS dataset, so we can compare their results with the reference parameters of the respective HARPS spectra. To validate further the accuracy of our tool, we proceed to determinations and comparisons on more stars. Finally, we discuss about possible future improvements of our determinations.
Resolution and spectral shape {#s1}
-----------------------------
We examine the spectral change of M dwarfs according to convolution in different resolutions. The shapes of M dwarf spectra are different when obtained in lower resolutions. In general, the lower the resolution, the shallower the absorption lines. This is illustrated in Fig \[convall\] where three lines of Gl176 are shown in detail, for the original HARPS spectrum and the convolved ones to several resolutions. We also measure these lines and we report their pseudo-EW values in Table \[EWvalues\], to show their differences. The relative differences can vary, as not only the depth changes but also the location of the pseudo continuum is different in each case. They all confirm that the lower resolution always has lower pseudo-EW values. This is why we need to convolve the HARPS spectra to the respective resolutions of the new spectra. Consequently, machine learning compares the pseudo EWs of the same resolution and predicts accurately the stellar parameters. In Fig \[shapeferos\] we show the spectral shapes of Gl674 for three different cases: the original HARPS spectrum with resolution 115000, the convolved HARPS spectrum to the resolution of FEROS (48000) and the original FEROS spectrum that is the lowest resolution we examine. We notice that the convolved HARPS spectrum follows the shape of the FEROS one in a consistent way. In Fig \[hsEW\], we show the comparison of the pseudo EWs of SOPHIE spectrum for Gl908 and the spectrum of the same star by HARPS before and after its convolution. The SOPHIE spectrum, which is of lower resolution, has consistently lower pseudo-EW values than the HARPS one, as expected. After the convolution of HARPS spectrum to the respective resolution, the overall trend of their values become highly compatible.
$\begin{array}{c}
\includegraphics[width=\hsize]{All5resolutions.pdf} \\
\end{array}$
$\begin{array}{c}
\includegraphics[width=\hsize]{2HARPS1FEROS.pdf} \\
\end{array}$
$\begin{array}{c}
\includegraphics[width=\hsize]{SH_original.pdf} \\
\includegraphics[width=\hsize]{SH_convolved.pdf} \\
\end{array}$
[cccc]{}\
Resolution & p.EW of $\lambda$ 6536.67 & p.EW of $\lambda$ 6537.08 & p.EW of $\lambda$ 6537.64\
& \[m$\AA$\] & \[m$\AA$\] & \[m$\AA$\]\
\
original 115000 & 14.19 & 29.08 & 27.31\
convolved 110000 & 14.04 & 28.89 & 26.98\
convolved 94600 & 13.39 & 28.27 & 26.35\
convolved 75000 & 11.99 & 27.45 & 25.08\
convolved 48000 & 7.50 & 22.50 & 20.04\
\
\[EWvalues\]
Measurements on different spectrographs {#s2}
---------------------------------------
We examine the performance of our tool in new spectra. We show the accuracy of the stellar parameters predicted and the precision for each resolution, by presenting the mean absolute errors of the models and the dispersion of the results, as calculated after the 100 determinations for each spectrum.
For the case of HARPS, we use a HARPS spectrum of Gl643 with SNR = 83, which is not part of the HARPS dataset used in the machine learning. As reference values for this star, we consider its parameters reported by @neves14. For the cases of the other instruments, we use a UVES spectrum of Gl846 with SNR = 149, a CARMENES spectrum of Gl514 with SNR = 191, a SOPHIE spectrum of Gl908 with SNR = 90 and a FEROS spectrum of Gl674 with SNR = 61. As reference values to those spectra, we consider the values of the respective HARPS ones in the dataset.
The results of T$_{\mathrm{eff}}$ and \[Fe/H\] are presented in Table \[newstarspar\]. We notice that the parameters of the new spectra are very close to the respective reference values. The differences in T$_{\mathrm{eff}}$ vary up to $\sim$50 K and the differences in \[Fe/H\] vary up to 0.03 dex. The mean absolute errors of models and the dispersions of values are slightly growing towards lower resolutions.
Measurements on different SNR’s {#snr}
-------------------------------
Here we examine the possible variation of the results regarding different signal-to-noise ratios (SNR) for a given spectrum. We take the spectrum Gl514 of CARMENES, which has the highest SNR of the ones we examine (equal to 191, as reported in the CARMENES data archive) and we inject amounts of noise which correspond to lower SNR values that we set. Since the final noise is obtained by the quadratic sum of the initial noise and the injected noise, the final SNR values are calculated using the relation below. $$(\frac{1}{SNR})^{2}_{final}=(\frac{1}{SNR})^{2}_{initial} + (\frac{1}{SNR})^{2}_{injected}$$
We create new spectra with final SNR values ranging from 100 to 9. For each spectrum, we measure the stellar parameters and their dispersion. Fig. \[snrcomp\] illustrates the measurements of the CARMENES spectrum while degrading its SNR. Overall, the results are similar to the ones of the original spectrum and the differences are kept roughly constant with respect to the reference values. For SNR values down to 20, we notice that the dispersions are between 17 and 27 K for T$_{\mathrm{eff}}$ and between 0.03 and 0.04 dex for \[Fe/H\], i.e. at similar levels as those of the original spectrum. For SNR values below 20, the dispersions start to increase up to $\sim$50 K and up to $\sim$0.07 dex respectively. Moreover, it seems that there is a slight decrease of the order of 20 K in T$_{\mathrm{eff}}$ and a slight increase of the order of 0.02 dex in \[Fe/H\] for the spectra with SNR below 20. However, these results are within the uncertainties of the tool. Therefore, we conclude that our tool works consistently for spectra with SNR above 20. Bellow this SNR, the errors increase significantly.
$\begin{array}{c}
\includegraphics[width=9cm]{multiSNR_Teff.pdf} \\
\includegraphics[width=9cm]{multiSNR_FeH.pdf} \\
\end{array}$
Comparison of results between our tool and @neves14 {#oduvn}
---------------------------------------------------
Now, we make an overall comparison of our results on a group of HARPS spectra with the ones presented by @neves14. For this purpose, we measure 30 HARPS spectra from the initial GTO sample, for which we do not know their parameters from photometry and are not part of the machine learning dataset we use. Based on the information from @neves14, we have excluded very active stars and stars with SNR lower than 25, below which that method does not apply. Both methods have been tested and do not work properly for very active or young stars, since the pseudo EWs of such spectra are affected and their parameters can not be determined accurately with the pseudo-EW approach we follow. Then, we compare the results we get by our tool with the results presented by @neves14.
The errors of the stellar parameters derived using our tool, are 27 K for T$_{\mathrm{eff}}$ and 0.04 dex for \[Fe/H\], as the mean absolute errors are measured when the machine learning model is applied on the test dataset. The errors of the calibration by @neves14, which are quantified from the root mean squared error (RMSE) in that work, are equal to 91 K and 0.08 dex respectively. It is reminded that both methods are tied to the same initial systematic uncertainties of the reference parameters used, which are 100K for T$_{\mathrm{eff}}$ and 0.17 dex for \[Fe/H\].
The results and their differences are presented in Table \[parall\] and Fig. \[paracomp\]. The mean and median difference of T$_{\mathrm{eff}}$ is 11 and 22 K respectively, with a standard deviation of 101 K. Regarding \[Fe/H\], the mean and median difference is -0.04 dex, with a standard deviation of 0.06 dex.
Work by @neves14 follows a traditional approach, using a least-squares weighted fit to determine parameters. The regression of our tool reduces those errors of T$_{\mathrm{eff}}$ and \[Fe/H\] from 91 to 27 K and from 0.08 to 0.04 dex respectively. So, our machine learning approach increases significantly the precision of parameter determinations. In terms of speed, the determination for a star by machine learning, even after the multiple runs with shuffling and splitting again the train/test samples, is a matter of few seconds.
$\begin{array}{c}
\includegraphics[width=\hsize]{TDN.pdf} \\
\includegraphics[width=\hsize]{FDN.pdf}\\
\end{array}$
[ccccccc]{}\
Star & T$_{\mathrm{eff}}$ (AA) & T$_{\mathrm{eff}}$ (Ne14) & T$_{\mathrm{eff}}$ Diff. & \[Fe/H\] (AA) & \[Fe/H\] (Ne14) & \[Fe/H\] Diff.\
& \[$\pm$27 K\] & \[$\pm$91 K\] & \[K\] & \[$\pm$0.04 dex\] & \[$\pm$0.08 dex\] & \[dex\]\
\
CD-44-836A & 3104 & 3032 & 72 & -0.07 & -0.07 & 0.00\
G108-21 & 3214 & 3186 & 28 & -0.02 & -0.02 & 0.00\
GJ1057 & 2926 & 2916 & 10 & -0.11 & -0.10 & -0.01\
GJ1061 & 2772 & 2882 & -110 & -0.25 & -0.09 & -0.16\
GJ1065 & 3106 & 3082 & 24 & -0.32 & -0.23 & -0.09\
GJ1123 & 2971 & 2779 & 192 & -0.02 & 0.15 & -0.17\
GJ1129 & 3037 & 3017 & 20 & -0.02 & 0.05 & -0.07\
GJ1236 & 3225 & 3280 & -55 & -0.44 & -0.47 & 0.03\
GJ1256 & 2964 & 2853 & 111 & -0.02 & 0.06 & -0.08\
GJ1265 & 3020 & 2941 & 79 & -0.28 & -0.20 & -0.08\
Gl12 & 3245 & 3239 & 6 & -0.31 & -0.29 & -0.02\
Gl145 & 3297 & 3270 & 27 & -0.27 & -0.28 & 0.01\
Gl203 & 3174 & 3138 & 36 & -0.31 & -0.22 & -0.09\
Gl299 & 3078 & 3373 & -295 & -0.53 & -0.53 & 0.00\
Gl402 & 3052 & 2943 & 109 & 0.00 & 0.03 & -0.03\
Gl480.1 & 3214 & 3211 & 3 & -0.48 & -0.48 & 0.00\
Gl486 & 3096 & 2941 & 155 & -0.02 & 0.03 & -0.05\
Gl643 & 3113 & 3102 & 11 & -0.29 & -0.26 & -0.03\
Gl754 & 2988 & 3005 & -17 & -0.23 & -0.14 & -0.09\
L707-74 & 3250 & 3353 & -103 & -0.39 & -0.38 & -0.01\
LHS1134 & 3007 & 2950 & 57 & -0.20 & -0.13 & -0.07\
LHS1481 & 3342 & 3510 & -168 & -0.66 & -0.76 & 0.10\
LHS1723 & 3031 & 3167 & -136 & -0.29 & -0.24 & -0.05\
LHS1731 & 3229 & 3273 & -44 & -0.22 & -0.19 & 0.03\
LHS1935 & 3222 & 3181 & 41 & -0.20 & -0.22 & 0.02\
LHS337 & 3003 & 3007 & -4 & -0.33 & -0.27 & -0.06\
LHS3583 & 3205 & 3236 & -31 & -0.13 & -0.22 & 0.09\
LHS3746 & 3111 & 3013 & 98 & -0.17 & -0.13 & -0.04\
LHS543 & 3042 & 2872 & 170 & 0.17 & 0.23 & -0.06\
LP816-60 & 3030 & 2960 & 70 & -0.11 & -0.07 & -0.04\
\
\[parall\]
Estimating total uncertainties {#gauss}
-------------------------------
Intrinsic uncertainties exist in the T$_{\mathrm{eff}}$ and \[Fe/H\] reference values of the HARPS dataset, since their initial photometric derivations have average uncertainties of 100 K and 0.17 dex respectively. Since these parameters are used as the training values for the machine learning process, we decide to inject these uncertainties by perturbing their values accordingly, in order to see how the final results of the predictions will vary.
Therefore, we create gaussian distributions on the parameters for each HARPS training dataset, increasing the dispersion of distribution on the reference parameters each time with step of 10 K and 0.02 dex, until the uncertainties of 100 K and 0.17 dex. This adds different training values to the machine learning algorithm each time. For each step, we create 100 gaussian-distributed training datasets. After these runs of machine learning, we calculate the average values of predicted parameters and their dispersion.
In Fig. \[gd\], we present the variations for spectra from the highest resolution (HARPS), the lowest resolution (FEROS) and an intermediate resolution (CARMENES). The datapoints represent the average difference between the resulting parameters and the reference values, after being calculated with the 100 different datasets. The errorbars are the dispersion of it. We notice that the average differences from the reference values are almost the same among them, regardless the amount of uncertainty injected to the gaussian distribution.
The average results of T$_{\mathrm{eff}}$ and \[Fe/H\] for the spectra from all the instruments are presented in Table \[gdstarpar\]. We report their maximum errors after considering the maximum gaussian distribution with 100 K and 0.17 dex. Overall, the average values of the parameters remain roughly the same as the ones calculated with no gaussian distribution at all. The mean absolute errors (M.A.E.) of the machine learning models have grown to values between 65 and 80 K for T$_{\mathrm{eff}}$ and between 0.10 to 0.13 dex for \[Fe/H\], depending on the resolution of the HARPS dataset. The dispersion of the derived parameters grows as the resolution of the spectra becomes lower. Specifically, it is smaller than the injected uncertainties for the HARPS spectrum ($\sim$60 K and $\sim$0.10 dex), while for the spectra from other instruments, it is slightly higher than the uncertainties injected ($\sim$110 to $\sim$130 K and $\sim$0.18 to $\sim$0.22 dex respectively).
In all the cases though, the resulting average values of stellar parameters are very close to their expected values. Differences of T$_{\mathrm{eff}}$ are up to $\sim$40 K and differences of \[Fe/H\] are up to 0.03 dex, regarding to the expected values.
[ccccccccccc]{}\
Star & Spec. & Res. & Ref. & M.L. & M.A.E. & Disp. & Ref. & M.L. & M.A.E. & Disp.\
& & & T$_{\mathrm{eff}}$ & T$_{\mathrm{eff}}$ & T$_{\mathrm{eff}}$ & T$_{\mathrm{eff}}$ & \[Fe/H\] & \[Fe/H\] & \[Fe/H\] & \[Fe/H\]\
& & & \[K\] & \[K\] & \[K\] & \[K\] & \[dex\] & \[dex\] & \[dex\] & \[dex\]\
\
Gl643 & HARPS & 115000 & 3102 & 3113 & 27 & 10 & -0.26 & -0.28 & 0.04 & 0.01\
Gl846 & UVES & 110000 & 3682 & 3691 & 28 & 13 & -0.08 & -0.05 & 0.04 & 0.02\
Gl514 & CARMENES & 94600 & 3574 & 3547 & 28 & 17 & -0.13 & -0.13 & 0.04 & 0.03\
Gl908 & SOPHIE & 75000 & 3587 & 3580 & 28 & 18 & -0.38 & -0.35 & 0.04 & 0.03\
Gl674 & FEROS & 48000 & 3284 & 3338 & 30 & 24 & -0.18 & -0.16 & 0.04 & 0.03\
\
\[newstarspar\]
[ccccccccccc]{}\
Star & Spec. & Res. & Ref. & M.L. & M.A.E. & Disp. & Ref. & M.L. & M.A.E. & Disp.\
& & & T$_{\mathrm{eff}}$ & T$_{\mathrm{eff}}$ & T$_{\mathrm{eff}}$ & T$_{\mathrm{eff}}$ & \[Fe/H\] & \[Fe/H\] & \[Fe/H\] & \[Fe/H\]\
& & & \[K\] & \[K\] & \[K\] & \[K\] & \[dex\] & \[dex\] & \[dex\] & \[dex\]\
\
Gl643 & HARPS & 115000 & 3102 & 3126 & 65 & 60 & -0.26 & -0.29 & 0.10 & 0.10\
Gl846 & UVES & 110000 & 3682 & 3678 & 68 & 109 & -0.08 & -0.06 & 0.10 & 0.18\
Gl514 & CARMENES & 94600 & 3574 & 3545 & 77 & 113 & -0.13 & -0.13 & 0.12 & 0.19\
Gl908 & SOPHIE & 75000 & 3587 & 3585 & 78 & 120 & -0.38 & -0.36 & 0.13 & 0.21\
Gl674 & FEROS & 48000 & 3284 & 3324 & 80 & 138 & -0.18 & -0.15 & 0.13 & 0.22\
\
\[gdstarpar\]
$\begin{array}{cc}
\includegraphics[width=9cm]{G_Teff_Harps.pdf} &
\includegraphics[width=9cm]{G_Feh_Harps.pdf} \\
\includegraphics[width=9cm]{G_Teff_Carmenes.pdf} &
\includegraphics[width=9cm]{G_Feh_Carmenes.pdf} \\
\includegraphics[width=9cm]{G_Teff_Feros.pdf} &
\includegraphics[width=9cm]{G_Feh_Feros.pdf} \\
\end{array}$
Validation of \[Fe/H\] determinations by measuring binary systems {#binaries}
-----------------------------------------------------------------
Here, we measure \[Fe/H\] in binary systems, containing M dwarfs which are not part of the reference sample used for machine learning. Thus, we validate our method of \[Fe/H\] prediction in an independent way. We determine \[Fe/H\] both in FGK+M and in M+M systems for an even more intrinsic test of \[Fe/H\] agreement.
The \[Fe/H\] determinations of eight FGK+M binary systems, from spectra obtained by UVES and FEROS spectrographs, are presented in Table \[FGKM\]. Regarding the FGK stars, their \[Fe/H\] and respective uncertainties were derived using the methodology described in @sousa and @santos13. The method measures the equivalent widths of FeI and FeII lines and assumes ionization and excitation equilibrium. It makes use of the radiative transfer code MOOG [@sneden] and a grid of Kurucz model atmospheres [@kurucz]. The \[Fe/H\] values of the respective M dwarf secondaries, derived by ODUSSEAS, are presented along with the total uncertainties of our tool at the resolutions of UVES (0.10 dex) and FEROS (0.13 dex). All binaries have differences within the uncertainties of the methods.
Furthermore, we proceed to \[Fe/H\] determinations of stars in five M+M binary systems, measuring their available spectra from the CARMENES public archive. In Table \[MM\], we present these results along with their own dispersions, since both are estimated by our tool based on the same reference values with the same initial uncertainties. We notice agreement between the respective members of all the M+M binaries, within the dispersions of their \[Fe/H\] determinations. This is a validation that our tool predicts \[Fe/H\] in a consistent and accurate way.
[ccccccc]{}\
Primary & \[Fe/H\] & $\sigma$\[Fe/H\] & Secondary & \[Fe/H\] & $\sigma$\[Fe/H\] & \[Fe/H\] Difference\
& \[dex\] & \[dex\] & & \[dex\] & \[dex\] & \[dex\]\
\
Gl100A & -0.29 & 0.05 & Gl100C & -0.29 & 0.13 & 0.00\
Gl118.1A & 0.02 & 0.05 & Gl118.1B & 0.05 & 0.10 & 0.03\
Gl173.1A & -0.37 & 0.03 & Gl173.1B & -0.29 & 0.10 & 0.08\
Gl157A & -0.08 & 0.04 & Gl157B & 0.02 & 0.13 & 0.10\
NLTT19073 & 0.08 & 0.03 & NLTT19072 & -0.07 & 0.13 & -0.15\
NLTT29534 & 0.00 & 0.03 & NLTT29540 & -0.03 & 0.10 & -0.03\
NLTT34137 & -0.12 & 0.05 & NLTT34150 & -0.13 & 0.10 & -0.01\
NLTT34353 & -0.10 & 0.03 & NLTT34357 & -0.11 & 0.10 & -0.01\
\
\[FGKM\]
[ccccccc]{}\
Primary & \[Fe/H\] & Disp. & Secondary & \[Fe/H\] & Disp. & \[Fe/H\] Difference\
& \[dex\] & \[dex\] & & \[dex\] & \[dex\] & \[dex\]\
\
Gl553 & -0.07 & 0.06 & Gl553.1 & -0.10 & 0.07 & 0.03\
Gl875 & -0.15 & 0.05 & Gl875.1 & -0.17 & 0.06 & 0.02\
Gl617A & 0.04 & 0.04 & Gl617B & -0.08 & 0.08 & 0.12\
Gl745A & -0.48 & 0.04 & Gl745B & -0.53 & 0.06 & 0.05\
Gl752A & 0.01 & 0.03 & Gl752B & -0.07 & 0.08 & 0.08\
\
\[MM\]
Discussion on the reference parameter scales {#referencescales}
--------------------------------------------
Since supervised machine learning determines the parameters based on reference values given to it, their systematics will apply to the results of new stars too. In this work, we have used the reference T$_{\mathrm{eff}}$ and \[Fe/H\] photometric scales of @casag08 and @neves12 respectively, as they are derived in a homogeneous way for a sufficiently big number of spectra available to us. It is important to make a comparison between the reference values we use and values of same stars derived by other recent works, which may be subject to different systematics. Such is @mann15, with which we share 26 common stars of the 65 ones we use as our reference dataset. In Table \[scales\], we compare our reference parameters with determinations by @mann15 and report the differences. These differences are illustrated in Figure \[scalecomp\]. Regarding T$_{\mathrm{eff}}$, we notice that our reference values have a systematic underestimation of 178 K on average with a standard deviation of 73 K. This systematic difference roots back to the different methods of derivation followed. Work by @casag08 is based on the multiple optical-infrared technique (MOITE) for M dwarfs, which is an extension of the infrared flux method (IRFM) as described in @casag06. On the other hand, determinations by @mann15 are done by comparing the optical spectra with the CFIST suite of the BT-SETTL version of the PHOENIX atmosphere models [@allard13]. The detailed description of this method can be found in @mann13b. Regarding \[Fe/H\], we notice no significant systematic difference between the methods of calibration by @neves12 and @mann15. The average difference is 0.06 dex with a standard deviation of 0.11 dex for the sample of stars in common.
As a potential future improvement of our determinations, we consider the possibility of replacing our reference dataset. Since new techniques of parameter determination become more accurate and precise and as more spectra will become available to us, their homogeneously derived parameters can be correlated with their pseudo EWs. Thus, we take into account the creation of an improved reference dataset for our machine learning tool.
[ccccccc]{}\
Star & T$_{\mathrm{eff}}$ (Ref.) & T$_{\mathrm{eff}}$ (Mann15) & T$_{\mathrm{eff}}$ Diff. & \[Fe/H\] (Ref.) & \[Fe/H\] (Mann15) & \[Fe/H\] Diff.\
& \[$\pm$100 K\] & \[$\pm$60 K\] & \[K\] & \[$\pm$0.17 dex\] & \[$\pm$0.08 dex\] & \[dex\]\
\
Gl54.1 & 2091 & 3056 & -65 & -0.40 & -0.26 & -0.14\
Gl87 & 3565 & 3638 & -73 & -0.30 & -0.36 & 0.06\
Gl105B & 3054 & 3284 & -230 & -0.14 & -0.12 & -0.02\
Gl176 & 3369 & 3680 & -311 & 0.02 & 0.14 & -0.12\
Gl205 & 3497 & 3801 & -304 & 0.17 & 0.49 & -0.32\
G213 & 3026 & 3250 & -224 & -0.19 & -0.22 & -0.03\
Gl250B & 3369 & 3481 & -112 & -0.09 & 0.14 & -0.25\
Gl273 & 3107 & 3317 & -210 & -0.05 & -0.11 & 0.06\
Gl382 & 3429 & 3623 & -194 & 0.04 & 0.13 & -0.09\
Gl393 & 3396 & 3548 & -154 & -0.13 & -0.18 & 0.05\
Gl436 & 3277 & 3479 & -202 & 0.01 & 0.01 & 0.00\
Gl447 & 2952 & 3192 & -240 & -0.23 & -0.02 & -0.21\
Gl514 & 3574 & 3727 & -153 & -0.13 & -0.09 & -0.04\
Gl526 & 3545 & 3649 & -104 & -0.18 & -0.31 & 0.13\
Gl555 & 2987 & 3211 & -224 & 0.13 & 0.17 & -0.04\
Gl581 & 3203 & 3395 & -192 & -0.18 & -0.15 & -0.03\
Gl686 & 3542 & 3657 & -115 & -0.29 & -0.25 & -0.04\
Gl699 & 3094 & 3228 & -134 & -0.59 & -0.40 & -0.19\
Gl701 & 3535 & 3614 & -79 & -0.20 & -0.22 & 0.02\
Gl752A & 3336 & 3558 & -222 & 0.04 & 0.10 & -0.06\
Gl846 & 3682 & 3848 & -166 & -0.08 & 0.02 & -0.10\
Gl849 & 3200 & 3530 & -330 & 0.24 & 0.37 & -0.13\
Gl876 & 3059 & 3247 & -188 & 0.14 & 0.17 & -0.03\
Gl880 & 3488 & 3720 & -232 & 0.05 & 0.21 & -0.16\
Gl887 & 3560 & 3688 & -128 & -0.20 & -0.06 & -0.14\
Gl908 & 3587 & 3646 & -59 & -0.38 & -0.45 & -0.07\
\
\[scales\]
$\begin{array}{c}
\includegraphics[width=\hsize]{TRM.pdf} \\
\includegraphics[width=\hsize]{FRM.pdf}\\
\end{array}$
Summary {#sum}
=======
We present our machine learning tool ODUSSEAS for the derivation of T$_{\mathrm{eff}}$ and \[Fe/H\] in M dwarf stars, whose spectra can have different resolutions and wavelength ranges inside the area from 530 to 690 nm. We explain in detail the way it is built and works. We present the results of the tests we perform and we examine its accuracy and precision from very high resolution of 115000 to resolution of 48000. Our tool seems to be reliable, as it operates with high machine learning scores around 0.94 and achieves excellent predictions of significantly high precision with mean absolute errors of $\sim$30 K for T$_{\mathrm{eff}}$ and $\sim$0.04 dex for \[Fe/H\]. Taking into consideration the intrinsic uncertainties of the reference parameters and perturbing them accordingly, our models have maximum uncertainties of $\sim$80 K for T$_{\mathrm{eff}}$ and $\sim$0.13 dex for \[Fe/H\], which are within the typical uncertainties for M dwarfs. Our parameters for spectra from different spectrographs, occurring from the average of 100 determinations, have consistent values with differences within $\sim$50 K and $\sim$0.03 dex from the expected ones. Spectra should have SNR above 20 for optimal predictions. Our tool is valid for M dwarfs in the intervals 2800 to 4000 K for T$_{\mathrm{eff}}$ and -0.83 to 0.26 dex for \[Fe/H\], except from very active or young stars. It can be tested by downloading the files in the webpage <https://github.com/AlexandrosAntoniadis/ODUSSEAS>, after reading the README instructions for clarifying the technical details.
This work was supported by FCT/MCTES through national funds and by FEDER - Fundo Europeu de Desenvolvimento Regional through COMPETE2020 - Programa Operacional Competitividade e Internacionalização by these grants: UID/FIS/04434/2019, UIDB/04434/2020 and UIDP/04434/2020; PTDC/FIS-AST/32113/2017 and POCI-01-0145-FEDER-032113; PTDC/FIS-AST/28953/2017 and POCI- 01-0145-FEDER-028953. A.A.K., S.G.S., E.D.M. and G.D.C.T. acknowledge the support from FCT in the form of the exploratory projects with references IF/00028/2014/CP1215/CT0002, IF/00849/2015/CP1273/CT0003 and IF/00956/2015/CP1273/CT0002. S.G.S. and E.D.M. further acknowledge the support from FCT through the Investigador FCT contracts IF/00028/2014/CP1215/CT0002, IF/00849/2015/CP1273/CT0003 and POCH/FSE (EC). G.D.C.T. further acknowledges the support from an FCT/Portugal PhD grant with reference PD/BD/113478/2015.
[200]{}
Akras, S., Leal-Ferreira, M. L., Guzman-Ramirez, L., & Ramos-Larios, G. 2019, , 483, 5077
Allard, F., Homeier, D., Freytag, B., et al. 2013, Memorie della Societa Astronomica Italiana Supplementi, 24, 128
Bonfils, X., Delfosse, X., Udry, S., et al. 2005, , 442, 635
Bonfils, X., Delfosse, X., Udry, S., et al. 2013, , 549, A109
Casagrande, L., Portinari, L., & Flynn, C. 2006, , 373, 13
Casagrande, L., Flynn, C., & Bessell, M. 2008, , 389, 585
Das, P., & Sanders, J. L. 2019, , 484, 294
Everett, M. E., Howell, S. B., Silva, D. R., et al. 2013, , 771, 107
Howard, E. M. 2017, Astronomical Data Analysis Software and Systems XXV, 512, 245
Johnson, J. A., & Apps, K. 2009, , 699, 933
Kurucz, R. 1993, ATLAS9 Stellar Atmosphere Programs and 2 km/s grid. Kurucz CD-ROM No. 13. Cambridge, 13
Lindgren, S., Heiter, U., & Seifahrt, A. 2016, , 586, A100
Mann, A. W., Brewer, J. M., Gaidos, E., L[é]{}pine, S., & Hilton, E. J. 2013a, , 145, 52
Mann, A. W., Gaidos, E., & Ansdell, M. 2013b, , 779, 188
Mann, A. W., Feiden, G. A., Gaidos, E., et al. 2015, , 804, 64
Neves, V., Bonfils, X., Santos, N. C., et al. 2012, , 538, A25
Neves, V., Bonfils, X., Santos, N. C., et al. 2014, , 568, A121
nehag, A., Heiter, U., Gustafsson, B., et al. 2012, , 542, A33
Passegger, V. M., Schweitzer, A., Shulyak, D., et al. 2019, , 627, A161
Rajpurohit, A. S., Allard, F., Rajpurohit, S., et al. 2018, , 620, A180
Rau, M. M., Koposov, S. E., Trac, H., & Mandelbaum, R. 2019, , 484, 409
Rojas-Ayala, B., Covey, K. R., Muirhead, P. S., & Lloyd, J. P. 2010, , 720, L113
Rojas-Ayala, B., Covey, K. R., Muirhead, P. S., & Lloyd, J. P. 2012, , 748, 93
Santos, N. C., Sousa, S. G., Mortier, A., et al. 2013, , 556, A150
Sneden, C. A. 1973, Ph.D. Thesis
Sousa, S. G., Santos, N. C., Mayor, M., et al. 2008, , 487, 373
Ucci, G., Ferrara, A., Gallerani, S., et al. 2019, , 483, 1295
Woolf, V. M., & Wallerstein, G. 2005, , 356, 963
Evaluation of our pseudo-EW measurements {#A}
========================================
We calculated the pseudo EWs of 4104 lines for the 110 stars of the total HARPS sample. Here we compare our values with the ones obtained by MCAL code. In the upper panel of Fig. \[comp\], we present the comparison of all the pseudo-EW values for the star Gl176 as an example. The units of pseudo EWs are m$\AA$. Inside the plots, AA stands for our measurements and VN stands for the measurements by @neves14. The slope in the diagrams of most stars is almost identical with the identity line, with only few pseudo EWs having considerably different values. In the lower panel of Fig. \[comp\] we show the relative difference (the percentage) of the values against our values. A scatter appears for the pseudo-EW values smaller than 30 m$\AA$, which is normal, as the relative difference for these narrow lines is greater. In contrast, nearly all lines broader than 50 m$\AA$ are measured with significant agreement. The actual quality test for measuring the pseudo EWs, comes from the following comparison. We plot the mean differences and mean relative differences between our method and the code by @neves14 for each line averaged by all the stars, to see how all the lines are measured. The result is very good as it can be seen in Fig. \[d\], with only 184 lines out of 4104 showing a mean relative difference greater than $\pm$15%.
$\begin{array}{c}
\includegraphics[width=8cm]{pew_AAVN.pdf} \\
\includegraphics[width=8cm]{pew_relAAVN.pdf} \\
\end{array}$
$\begin{array}{c}
\includegraphics[width=8cm]{Ave_Dplot.pdf} \\
\includegraphics[width=8cm]{RelAve_Dplot.pdf} \\
\end{array}$
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Every prism manifold can be parametrized by a pair of relatively prime integers $p>1$ and $q$. In our earlier papers, we determined a complete list of prism manifolds $P(p, q)$ that can be realized by positive integral surgeries on knots in $S^3$ when $q<0$ or $q>p$; in the present work, we solve the case when $0<q<p$. This completes the solution of the realization problem for prism manifolds.'
address:
- 'Department of Mathematics, Princeton University, Princeton, NJ 08544'
- 'Department of Mathematics, California Institute of Technology, Pasadena, CA 91125'
- 'Department of Mathematics, University of Texas, Austin, TX 78712'
- 'Department of Mathematics, Duke University, Durham, NC 27708'
author:
- William Ballinger
- Yi Ni
- Tynan Ochse
- Faramarz Vafaee
bibliography:
- 'Reference.bib'
title: The prism manifold realization problem III
---
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The broad line region (BLR) of luminous active galactic nuclei (AGN) is a prominent observational signature of the accretion flow around supermassive black holes, which can be used to measure their masses ($\mbh$) over cosmic history. Due to the $\lesssim100\muas$ angular size of the BLR, current direct constraints on BLR kinematics are limited to those provided by reverberation mapping studies, which are most efficiently carried out on low-luminosity $L$ and low-redshift $z$ AGN. We analyze the possibility to measure the BLR size and study its kinematic structure using *spectroastrometry*, whereby one measures the spatial position centroid of emission line photons as a function of velocity. We calculate the expected spectroastrometric signal of a rotation-dominated BLR for various assumptions about the ratio of random to rotational motions, and the radial distribution of the BLR gas. We show that for hyper-luminous quasars at $z < 2.5$, the size of the low-ionization BLR can already be constrained with existing telescopes and adaptive optics systems, thus providing a novel method to spatially resolve the kinematics of the accretion flow at $10^3 - 10^4$ gravitational radii, and measure $\mbh$ at the high-$L$ end of the AGN family. With a 30m-class telescope, BLR spectroastrometry should be routinely detectable for much fainter quasars out to $z\sim 6$, and for various emission lines. This will enable kinematic $\mbh$ measurements as a function of luminosity and redshift, providing a compelling science case for next generation telescopes.'
author:
- 'Jonathan Stern, Joseph F. Hennawi, J[" o]{}rg-Uwe Pott'
title: |
Spatially Resolving the Kinematics of the $\lesssim100\muas$ Quasar Broad Line Region\
Using Spectroastrometry
---
Introduction {#sec: intro}
============
Quasars are the most luminous compact objects in the Universe, with luminosities $L$ reaching up to $10^{48}\ergs$. Such immense energy outputs are believed to be the result of gas accretion onto black holes (BHs) with masses of up to $\mbh\approx10^{10}\msun$ (@Lynden-Bell69), shining at luminosities reaching their Eddington luminosity ($1.3\times10^{38}\,\mbh/\msun\ergs$).
Arguably the most distinct property of quasar spectra is the prominence of broad emission lines, with typical widths of $\sim3,000\kms$ (e.g. @VandenBerk+01), and maximum widths of $10-20,000\kms$ (@Laor03 [@Fine+08; @SternLaor12a]). The ionization structure of the broad line emitting gas, known as the broad line region (BLR), is well described by photo-ionization by the central continuum source (@DavidsonNetzer79 and citations thereafter). However, the spatial structure and dynamics of the BLR are not well understood. The large observed velocities suggest that the kinematics of the BLR are dominated by the BH gravity, and that the BLR resides at $\sim10^3-10^4\,\rg$, where $\rg$ is the gravitational radius.
Assuming the observed BLR velocities are of the order of the local Keplerian velocity, then a measure of the BLR distance from the BH, ${r_{\rm BLR}}$, provides a measure of $\mbh$. Since luminous quasars are observable to high redshift $z$, then estimates of $\mbh$ in quasars can be utilized to track the buildup of massive BHs over cosmic time. In low-$L$ and low-$z$ Active Galactic Nuclei (AGN), the family of objects of which quasars are the luminous subset, ${r_{\rm BLR}}$ can be measured with reverberation mapping (RM). In RM, one derives ${r_{\rm BLR}}$ from the time lag between changes in the continuum luminosity and the corresponding variation in the line luminosity (@BlandfordMcKee82 [@Peterson93; @Peterson+04]). RM has been widely applied in the past two decades, and has yielded three main results. First, the response of to changes in continuum luminosity in single objects suggests that the BLR resides in a narrow range of $r$, where the bulk of the emission originates from a dynamical range of $5-10$ in $r$ (@Maoz+91 [@Pancoast+13]). Second, the characteristic response-weighted ${r_{\rm BLR}}$ of $\Hb$ scales as $\sim L^{1/2}$ (@Kaspi+05 [@Bentz+09b; @Bentz+13]), with a typical value of $0.05\pc$ at $L_{45}=1$, where $L_{45}$ is the monochromatic luminosity at 1450Å ${L_{1450\AA}}$ in units of $10^{45}\ergs$. And third, different broad emission lines are emitted from somewhat different $r$, where generally higher ionization lines originate from smaller $r$ than the Balmer lines.
The observed value of ${r_{\rm BLR}}$, its scaling with $L$, and its small dynamical range, are all aptly explained by a combination of two effects which suppress the line emission at $r\ll 0.05\,L_{45}^{1/2}\pc$ and $r> 0.1\,L_{45}^{1/2}\pc$. At $r\gtrsim\rsub\approx 0.1\,L_{45}^{1/2}\pc$, where $\rsub$ is the dust sublimation radius, dust grains can survive and hence suppress the line emission (@NetzerLaor93). At $r\ll 0.05\,L_{45}^{1/2}\pc$, the equilibrium of BLR gas pressure with radiation pressure suggested by [@Baskin+14a] implies that the gas volume densities are so high ($\gg10^{11}\cm^{-3}$), that emission of the permitted lines is collisionally suppressed. The equilibrium with radiation pressure can also explain the stratification of lines with ionization (see fig. 5 and §4.6 in @Baskin+14a).
Do the aforementioned properties of ${r_{\rm BLR}}$ apply also to high-$L$ and high-$z$ quasars? RM of a couple of $L\approx10^{47}\ergs$ quasars at $z\approx2$ suggests that the answer is yes (@Kaspi+07 [@Chelouche+12]). However, applying RM to luminous quasars is problematic since in high-$L$ quasars ${r_{\rm BLR}}$ is expected to be on the scale of light-years, with correspondingly long response times. Also, since quasar evolution dictates that high-$L$ quasars are at high-$z$, cosmological time dilation makes the observed time lags even longer. Furthermore, high luminosity quasars are less variable (@AngioneSmith72 [@VandenBerk+04]) and therefore have a weaker RM signal. The combination of a weak signal with RM timescales of several years in the observed frame makes RM of high-$L$ high-$z$ quasars extremely challenging, and hence to date RM studies of only a couple of such objects have been published. Therefore, an independent method to constrain ${r_{\rm BLR}}$ (and $\mbh$) at high-$L$ and high-$z$ would be a big advantage.
Another open question concerns the kinematic structure of the BLR, which given that it resides at $\sim10^3-10^4\,\rg$, is likely an integral part of the accretion flow. Is the BLR in some kind of ordered flow, such as an extension of the accretion disk, or are the motions disordered, such as a population of clouds in random virial motion about the BH? There are some indirect observations which favor an ordered flow, such as the rotation of the polarization angle across the spectral profile seen in some objects (@Smith+05), and the distinct variability characteristics of the red and blue line wings which is sometimes observed (@VeilleuxZheng91). Both the rotation of the polarization angle and the distinct variability patterns as a function of velocity suggest that the red-wing emitting gas and the blue-wing emitting gas are spatially distinct, as expected in an ordered flow. Also, an ordered flow is favored due to the lack of ‘spikes’ in the broad emission line profile, which are expected if the BLR is composed of individual clouds in a random velocity field (@Arav+98 [@Dietrich+99; @Laor+06]). However, a direct observation which confirms the existence of an ordered velocity field in the BLR remains elusive.
In this study, we analyze the possibility to constrain the BLR size and velocity field in high-$L$ and high-$z$ quasars using spectroastrometry (e.g. @Bailey+98). In spectroastrometry, one measures the position centroid of photons as a function of photon wavelength, which in principle can be pinpointed to an accuracy which is $\sim N_{\rm ph}^{1/2}$ higher than the resolution limit of the telescope, where $N_{\rm ph}$ is the number of photons. Extrapolation of the ${r_{\rm BLR}}-L$ relation mentioned above to the most luminous quasars implies BLR angular sizes of $\approx100\muas$, which is a factor of $\sim 10^3$ below the resolution of $\approx50\mas$ of diffraction-limited observations on 8m telescopes using adaptive optics at near-infrared wavelengths. Therefore, with $\sim 10^{6}$ photons, the $100\muas$-scale photocenter offset between the red-wing and blue-wing photons can in principle be detected. As spectroastrometry is most effective in luminous quasars where ${r_{\rm BLR}}$ is large and the photon flux is high, it is complementary to RM which is most easily applied to low-$L$ AGN. Additionally, spectroastrometry measures an $r$-weighted function of the BLR (see below), compared to the response-weighted function of the BLR measured by RM, hence the two methods give independent constraints on the size, kinematics, and distribution of material in the BLR.
[@Bailey+98] applied spectroastrometry to seeing limited observations of binary stars and of the narrow line region (NLR) in AGN. Spectroastrometry was also applied to nuclear gas in the Circinus galaxy (@Gnerucci+13), to molecular disks in young stellar objects (@Pontoppidan+08 [@Pontoppidan+11; @Joergens+13]), and to planetary nebulae (@Blanco+14). Using Adaptive Optics, @Pontoppidan+11 succeeded in overcoming atmospheric and technical issues and achieved photon-limited angular resolutions of $100-500\muas$, suggesting that resolving the BLR with spectroastrometry is feasible with existing telescopes. Conceptually similar to BLR spectroastrometry, [@Shen12] discusses the possibility to exploit the astrometric position of a variable broad line emission, via imaging with a narrow band filter with wavelength tuned to the broad emission line. As discussed in [@Trippe+10], measuring the 2D-astrometric position is dominated by optical aberrations in the imaging instrument, and current imagers show astrometric biases at the $400\muas$ level, a factor of $>$4 too large for BLR spectroastrometry. Hence, [@Shen12] concluded that detecting the astrometric signal of the BLR will be delayed to next-generation telescopes. However, in spectroastrometry the optical abberation bias is circumvented, since all photons practically travel along the same path through the optics (see below), and therefore the astrometric signal can in principle be detected already by existing telescopes. Additionally, since the technique suggested by [@Shen12] requires a variable BLR, it has the same observational resource requirements as RM, and is therefore practically limited to focus on the same highly and shortly variable low-$L$ low-$z$ objects. Therefore, in this paper we limit the discussion to the simpler possibility of applying spectroastrometry to the BLR without any temporal information.
This paper is structured as follows. In §2 we calculate the BLR angular sizes for the most luminous quasars at each $z$. In §3 we estimate the expected spectroastrometric signal for a rotation-dominated BLR, for different assumptions on the radial distribution of the BLR gas and on the ratio of random to rotational motion. We present several observational considerations in §4, and simulate the expected spectroastrometric signal with $8{\rm m}$ and $30{\rm m}$-class telescopes in §5. We discuss our results in §6, and summarize in §7. Throughout the paper, we assume a FRW cosmology with $\Omega$ = 0.3, $\Lambda$ = 0.7 and $H_0 = 70\ \kms$ Mpc$^{-1}$.
The Angular Size of the Broad Line Region {#sec: angular size}
=========================================
[@Kaspi+05] derived the relation between ${L_{1450\AA}}$ and ${r_{\rm RM}}$, the distance of the BLR measured by reverberation mapping of , at a luminosity range $10^{41}<{L_{1450\AA}}<10^{46}\ergs$. To estimate ${r_{\rm BLR}}$ in spectroastrometry candidates, we extrapolate their relation to the higher luminosities considered here ${L_{1450\AA}}\gtrsim10^{47}\ergs$: $$\label{eq: Kaspi05}
{r_{\rm RM}}= 0.5 \left(\frac{{L_{1450\AA}}}{10^{47}\ergs}\right)^{0.5} \pc ~~~.$$
We can convert eqn. (\[eq: Kaspi05\]) to a cosmology-independent relation based on angular size $\theta$ and observed flux. Define ${F_{\nu;\,1450{\rm \AA}}}$ as the flux density at rest-frame wavelength $\lrest=1450\AA$, and defining $\dL$ as the luminosity distance, we have $$\label{eq: nln1450}
{L_{1450\AA}}= \frac{\nu_{1450\AA}}{1+z} {F_{\nu;\,1450{\rm \AA}}}\cdot 4\pi \dL^2 ~~~.$$ Eq. (\[eq: nln1450\]), combined with the relation between angular distance and luminosity distance ${r_{\rm RM}}/{\theta({r_{\rm RM}})}=\dL/(1+z)^2$, gives $$\label{eq: tblr vs f_nu}
{\theta({r_{\rm RM}})}= 51 \left(\frac{{F_{\nu;\,1450{\rm \AA}}}}{10\mJy}\right)^{1/2} \left(1+z\right)^{3/2} \muas ~~~.$$ It is convenient to express eqn. (\[eq: tblr vs f\_nu\]) also with the commonly used $\mi(z=2)$, the apparent $i$-band magnitude after K-correcting to $z=2$ (i.e. the apparent magnitude at $\lrest=2500\AA$). Therefore, $${F_{\nu;\,1450{\rm \AA}}}=0.74\cdot 10^{-0.4(\mi(z=2) + 48.6)} ~~~,$$ where the factor of $0.74$ follows from a [@Richards+06] quasar template. Together with eqn. (\[eq: tblr vs f\_nu\]), we get $$\label{eq: tblr vs mi}
{\theta({r_{\rm RM}})}= 22\cdot 10^{\left(15-\mi(z=2)\right)/5} \left(1+z\right)^{3/2} \muas ~~~.$$
![ The apparent magnitudes and expected BLR angular sizes of the brightest AGN at each $z$, complied from BQS, SDSS, the $z>5.7$ quasar sample of [@Banados+14], and a sample of local Seyferts. [**(Top)**]{} For each 0.5-wide bin in $z$, the values of $\mi$ (K-corrected to $z=2$) of the brightest quasar, the 10 brightest, and the 100 brightest quasar, are shown. The ranges in $z$ where , , and are observed in NIR bands are noted on top. [**(Bottom)**]{} The expected BLR angular size of the quasars shown in the top panel, derived from ${L_{1450\AA}}$ by assuming the RM-based relation between ${r_{\rm RM}}$ and ${L_{1450\AA}}$. The implied cosmology-independent relation between $\theta$ and $F_\nu$ (eqn. \[eq: tblr vs f\_nu\]) is noted. The dependence of the maximum ${\theta({r_{\rm RM}})}$ on $z$ is remarkably weak, spanning merely a factor of three over the entire $z$ range. []{data-label="fig: angular size"}](muasVSzQC10.eps)
We search for the brightest quasars at each $z$ in the Bright Quasar Survey (BQS, @SchmidtGreen83), the quasar catalogs of the Sloan Digital Sky Survey 7 and 10 data releases (SDSS-DR7, @Schneider+10; SDSS-DR10, @Paris+14), and the $z>5.7$ quasar sample compiled by [@Banados+14]. We supplement this quasar list with local bright Seyferts from [@Bentz+09a], who corrected the AGN continuum emission for the host galaxy contribution, which may be significant in the relatively low luminosity Seyferts. We derive $\mi(z=2)$ for SDSS quasars by K-correcting the PSF magnitude of the SDSS-band which is closest to 2500Å in rest-frame wavelength. For BQS quasars we derive $\mi(z=2)$ from the B-magnitude, and for the high-$z$ quasars we use the J-band magnitudes, if available, or otherwise the y-band magnitude. The luminosity of the local Seyferts is based on the luminosity at 5100Å listed in table 9 of [@Bentz+09a]. All K-corrections are performed assuming a [@Richards+06] quasar template. We divide the quasars into bins of $0.5$ in $z$, and sort them by brightness. The brightest[^1], 10 brightest, and 100 brightest quasar in each $z$-bin are shown in Figure \[fig: angular size\]. For reference, we note the ranges in $z$ where , , and are observed in NIR bands. As expected, $\mi(z=2)$ of the brightest quasars generally decreases with $z$. However, due to the factor of $(1+z)^{3/2}$ in eqn. (\[eq: tblr vs mi\]), the dependence of the maximum ${\theta({r_{\rm RM}})}$ on $z$ is remarkably weak, spanning merely a factor of three ($50-150\muas)$ over the entire $z$ range.
The spectroastrometric signal of the BLR {#sec: spectroastrometric signal}
========================================
[@Chen+89] and @ChenHalpern89 (1989, hereafter CH89) derived the spectral profile of a broad emission line originating from a rotating disk, including both special and general relativistic effects to first order in $\rg/r$. The two main parameters of their model are the distribution of line emission as a function of $r$, and the amount of line broadening at each location of the disk, due to unordered motions which could be sourced by turbulence or other physical processes. Below, we discuss the physical and observational constraints on these two parameters. Here, we use the CH89 formulation to derive the expected spectroastrometric signal of a rotating BLR as a function of these two parameters.
We note that the flat disk assumed by CH89 is probably an oversimplification even for a rotation-dominated BLR, since the BLR characteristic height above the disk mid-plane may change with $r$. Therefore, our analysis is based on the assumption that as long as the dominant motion is rotation, the CH89 formulation should give a reasonable approximation of the expected spectroastrometric signal.
Following CH89, we define $r$ and $\varphi'$ as the polar coordinates in the disk frame. We assume that the orientation on the sky of the spatial direction of the telescope slit is at an angle $j$ from the major axis of the projected BLR ellipse. These definitions are pictured in the top panel of Figure \[fig: setup\], which depicts an idealized BLR, which originates from a single $r={r_{\rm BLR}}$.
![[**(Top)**]{} An illustration of a spectroastrometric observation of a simplified BLR, which originates from a single $r={r_{\rm BLR}}$ and has negligible local line broadening. The ellipse depicts the BLR ring, with coordinate $\varphi'$, projected onto the plane of the sky. Color denotes line-of-sight velocity. The physical size of the ellipse axes, up to relativistic effects, are ${r_{\rm BLR}}$ and ${r_{\rm BLR}}\cos i$. The angle between the spatial direction of the slit and the major axis of the projected ring is defined as $j$. [**(Middle)**]{} The implied emission-line profile of the rotating ring on top of the underlying quasar continuum, for an assumed ${v_{\rm rot}}=10\,000\kms$, $\sin i=0.5$, and ${\rm EW}=26\,000\kms$. Normalization assumes an line of a $L=10^{48}\ergs$ quasar at $z=2$. [**(Bottom)**]{} The implied photocenter offsets of the BLR, assuming no contamination by non-BLR sources, for the ${\theta({r_{\rm BLR}})}$ and $j$ noted in the panel. The centroid of the reddest line photons are offset from the centroid of the bluest line photons by $150\muas$. []{data-label="fig: setup"}](slitposition.eps)
We define ${\Phi^\ast}_v(r, \varphi')\d r\d\varphi' ~~\left[\cm^{-2}\s^{-1} \left(\kms\right)^{-1}\right]$ as the observed flux density of photons per unit velocity which originate from the disk coordinate $(r,\varphi')$. Similarly, the photocenter position of these photons along the spatial direction of the slit is defined as ${S^\ast}_v(r,\varphi')\ \left[\muas\right]$. The value of ${S^\ast}_v(r,\varphi')$ can be calculated from the impact parameter of a photon at infinity $b$, which equals (eqn. 6 in CH89) $$\label{eq: b}
b=r\left(1-\sin^2 i\cos^2\varphi'\right)^{1/2}\left(1+ O(\frac{\rg}{r})\right) ~~~,$$ where $i$ is the inclination of the disk normal to the line of sight. The first term in eqn. (\[eq: b\]) is the Newtonian projection of $(r,\varphi')$ on the sky, and the second term is a correction due to light bending by the central mass. Projecting $b$ on the slit spatial direction yields, with some trigonometry, $$\begin{aligned}
\label{eq: Sphs long}
{S^\ast}_v(r,\varphi') = r\left(\sin j \cos i \cos \varphi' + \cos j \sin \varphi'\right)\left(1+ O(\frac{\rg}{r})\right) ~. \nonumber \\\end{aligned}$$ The observable photocenter position at velocity $v$, $S_v$, is the photon-flux-average of ${S^\ast}_v(r,\varphi')$: $$\label{eq: Sphs}
S_v = \frac{\iint {S^\ast}_v(r,\varphi') {\Phi^\ast}_v(r, \varphi')\d \varphi'\d r}{\Phi_v} ~~~,$$ where $\Phi_v$ is the observed photon flux density: $$\label{eq: Nphs}
\Phi_v = \iint {\Phi^\ast}_v(r, \varphi')\d \varphi'\d r ~~~.$$ Now, since ${\Phi^\ast}_v(r, \varphi') = {\Phi^\ast}_v(r, \pi-\varphi')\left(1+ O(\rg/r)\right)$ (see CH89), then by summing the contribution from $\varphi'$ and $\pi-\varphi'$ to the integral in eqn. (\[eq: Sphs\]), the term in eqn. (\[eq: Sphs long\]) with $\cos \varphi'$ cancels out to zeroth order in $\rg/r$, and we get: $$\begin{aligned}
\frac{1}{2}\left[ {S^\ast}_v(r,\varphi') {\Phi^\ast}_v(r, \varphi') + {S^\ast}_v(r,\pi-\varphi') {\Phi^\ast}_v(r, \pi-\varphi')\right] = \nonumber \\
r\cos j \sin \varphi'{\Phi^\ast}_v(r, \varphi')\left(1+ O(\frac{\rg}{r})\right) ~~~.\end{aligned}$$ Hence, eqn. (\[eq: Sphs\]) simplifies to $$\label{eq: S_v0}
S_v = \cos j\cdot \frac{\int r\d r\int \d \varphi' \sin \varphi' {\Phi^\ast}_v(r, \varphi')\left(1+ O(\frac{\rg}{r})\right)}{\Phi_v}$$
The light-bending corrections to $S_v$ are of the order of $O(\rg/{r_{\rm BLR}})\lesssim 10^{-3}$, and are likely weaker than the inaccuracy of the model due to the assumption that the BLR is a flat rotating disk, so it is not clear whether these corrections are observable[^2].
The observed $S_v$ will be diluted by the quasar continuum, which at optical and ultraviolet frequencies originates from the accretion disk. The continuum emission will have a negligible spectroastrometric signal, and therefore will dilute $S_v$ of the broad line by a factor of $\Phi_v / (\Phi_v+{\Phi_{v}^{\rm cont}})$, where ${\Phi_{v}^{\rm cont}}$ is the photon flux density of the continuum[^3]. Therefore, the final form of eqn. (\[eq: S\_v0\]) is $$\label{eq: S_v}
S_v = \cos j\cdot \frac{\int r\d r\int \d \varphi' \sin \varphi' {\Phi^\ast}_v(r, \varphi')\left(1+ O(\frac{\rg}{r})\right)}{\Phi_v + {\Phi_{v}^{\rm cont}}}$$
A Simplified BLR
----------------
In order to explore the dependence of $S_v$ on the disk properties, we begin by calculating $S_v$ for a simplified BLR which originates from a single thin ring $r={r_{\rm BLR}}$ and has negligible local line-broadening, i.e. $\sigma \ll {v_{\rm rot}}$, where $\sigma$ is the characteristic velocity of the local line-broadening mechanism, and ${v_{\rm rot}}$ is the rotational velocity. Under these conditions, ${v_{\rm rot}}=(G\mbh/{r_{\rm BLR}})^{1/2}$. We relax both assumptions below. In the figures, we use the fully relativistic expression for ${\Phi^\ast}_v(r,\varphi')$ from CH89[^4], and eqn. (\[eq: S\_v\]) to calculate $S_v$. However, since the relativistic corrections to ${\Phi^\ast}_v(r,\varphi')$ have a weak effect on the derived $S_v$, for increased readability we repeat here the Newtonian expressions for ${\Phi^\ast}_v$. In this simplified BLR, $\Phi_v^{\ast}$ is equal to $$\label{eq: Phi1}
\Phi_v^{\ast}(r, \varphi')\propto\delta(r-{r_{\rm BLR}})\delta(\sin\varphi'-\frac{v}{{v_{\rm rot}}\sin i})$$ The implied $\Phi_v$ is shown in the second panel of Fig. \[fig: setup\], assuming ${v_{\rm rot}}=10\,000\kms$, $i=30\degree$, and a total broad line photon flux of $4\times10^6~{\rm photons} \hr^{-1} \m^{-2}$. This broad line flux is appropriate for the broad line of a $L=10^{48}\ergs$ quasar at $z=2$, assuming an equivalent width (EW) of $570\AA$ compared to the AGN continuum (@SternLaor12a). We note that @SternLaor12a found that EW() is independent of $L$ at $L<10^{46}\ergs$, and therefore we assume that the same EW pertains at $L\gg10^{46}\ergs$. The implied EW in velocity units is $26,000\kms$, and the continuum flux density is derived from this assumed EW. The units are chosen for easy conversion into the number of photons expected in an observation. Using eqn. (\[eq: Phi1\]) in eqn. (\[eq: S\_v\]), the implied $S_v$ for the simplified BLR is $$\label{eq: Sv1}
\left. S_v\right|_{v\leq{{v_{\rm rot}}\sin i}} = \frac{\Phi_v}{\Phi_v+{\Phi_{v}^{\rm cont}}}\cdot{r_{\rm BLR}}\cos j \frac{v}{{{v_{\rm rot}}\sin i}} ~~~,$$ which is shown in the bottom panel of Fig. \[fig: setup\], assuming $j=30\degree$ and ${\theta({r_{\rm BLR}})}=100\muas$. The $v=5000\kms$ photons are offset from the $v =-5000\kms$ by $\approx2{r_{\rm BLR}}\cos j\cdot\Phi_v/(\Phi_v+{\Phi_{v}^{\rm cont}})=150\muas$.
Local Line Broadening {#sec: turbulence}
---------------------
![ The emission line profile (upper panel) and photocenter offsets (lower panel) of a BLR ring with local line broadening. At each location in the ring, the emitted photon wavelength distribution is assumed to be a Gaussian with width $\sigma$. The black lines assume negligible broadening as in Fig. \[fig: setup\], while other plotted lines assume the noted value of $\sigma / ({{v_{\rm rot}}\sin i})$. As $\sigma / ({{v_{\rm rot}}\sin i})$ increases, the two peaks in the line profile merge into a single peak, and the maximum value of $|S_v|$ decreases. []{data-label="fig: turbulence"}](turbulence.eps)
We now relax the assumption that the local line broadening at each location in the disk is negligible. Therefore, $$\label{eq: Phi2}
\Phi_v^{\ast}(r, \varphi')\propto
\delta(r-{r_{\rm BLR}}) {\rm e}^{-\frac{\left({{v_{\rm rot}}\sin i}\right)^2}{2\sigma^2}\left(\sin \varphi' - \frac{v}{{{v_{\rm rot}}\sin i}}\right)^2} ~~~.$$ In Figure \[fig: turbulence\] we plot $\Phi_v$ and $S_v$ for different $\sigma/({{v_{\rm rot}}\sin i})$. In order to match the emission line-profile of a putative observation, ${v_{\rm rot}}$ is chosen so that the line width remains constant in the different models. For convenience, we characterize the line width by the Half Width at Half Maximum ($\hwhm$).
Figure \[fig: turbulence\] illustrates that as $\sigma/({{v_{\rm rot}}\sin i})$ increases, $\Phi_v$ loses its double-peaked shape and the maximum spectroastrometric signal $|S_v|$ decreases. This behavior occurs because as $\sigma/({{v_{\rm rot}}\sin i})$ increases, photons from a wider range of $\varphi'$ contribute to a given value of $v$, thus smoothing the spectral profile and $S_v$.
What is the value of $\sigma$ expected in the BLR? There are some indirect observations which suggest that $\sigma \sim {{v_{\rm rot}}\sin i}\sim 0.5\,{v_{\rm rot}}$[^5]. In most quasars the BLR profiles do not exhibit a double-peaked profile (@Strateva+03), which is expected when $\sigma < 0.5\,{{v_{\rm rot}}\sin i}$ (see top panel of Fig. \[fig: turbulence\]). This lack of a double-peak is sometimes attributed to the BLR being emitted from a wide range of $r$, however this explanation is disfavored by the RM results which suggest that the BLR originates from a narrow range in $r$ (see §\[sec: intro\]). Therefore, the single-peaked profile typically observed is more likely due to $\sigma/{{v_{\rm rot}}\sin i}$ of order unity. Furthermore, [@Osterbrock78] argued that the statistics of BLR line widths suggest $\sigma/{v_{\rm rot}}\sim0.4$, which is also roughly consistent with the inferred BLR covering factor (CF) of $\sim0.3$ (@Korista+97), assuming that $\sigma/{v_{\rm rot}}$ is a measure of the height-to-diameter ratio of the BLR disk. Due to the above arguments, for the purpose of estimating the strength of the spectroastrometric signal we will henceforth assume $\sigma/({{v_{\rm rot}}\sin i}) \approx 1$. The bottom panel of Fig. \[fig: turbulence\] shows that $\sigma/({{v_{\rm rot}}\sin i}) \approx 1$ implies $S_{\pm\hwhm}\approx\pm0.25\,{\theta({r_{\rm BLR}})}$.
Distribution in $r$ {#sec: r-distribution}
-------------------
We now relax the idealized assumption that the BLR originates from a single $r$, and consider a general spatial distribution of the broad line emitting gas, parameterized by $f(r)$, the line emission per unit $\log r$: $$\label{eq: Phi3}
\Phi_v^{\ast}(r, \varphi') = \frac{f(r)}{r}{\rm e}^{-\frac{\left({{v_{\rm rot}}\sin i}\right)^2}{2\sigma^2}\left(\sin \varphi' - \frac{v}{{{v_{\rm rot}}\sin i}}\right)^2}$$ What is the expected shape of $f(r)$? As noted in §\[sec: intro\], RM studies suggest that the BLR emission originates from a small dynamical range in $r$, i.e. $f(r)$ is strongly peaked around $r={r_{\rm RM}}$. We therefore, parameterize $f(r)$ as a double power law $f(r)\propto r^{\pm\alpha}$, where the index is positive at $r<{r_{\rm RM}}$ and negative at $r>{r_{\rm RM}}$, and calculate $S_v$ for $\alpha = 1$ and $2$. These $f(r)$ are shown in the top panel of Figure \[fig: r-distribution\], together with a $\delta$-function (thin ring) for comparison with Figs. \[fig: setup\] and \[fig: turbulence\]. We assume that $f(r)$ covers the range $0.03\,{r_{\rm RM}}- 30\,{r_{\rm RM}}$. Additionally, we also consider a physically motivated $f(r)$, where $f(r)$ drops at $r>2{r_{\rm RM}}=\rsub$ due to absorption of the ionizing photons by dust grains, and also drops at $r<{r_{\rm RM}}$ due to collisional suppression of the line emission. This $f(r)$ was calculated for different lines by [@Baskin+14a], assuming a constant BLR CF per unit $\log r$. Since below we estimate $S_v$ for , we use the @Baskin+14a calculation of $f(r)$ of , which is similar in shape to $f(r)$ of seen in their fig. 5[^6]. We refer the reader to @Baskin+14a for the details of the calculation.
The middle and lower panels of Fig. \[fig: r-distribution\] show the implied $\Phi_v$ and $S_v$, for the different $f(r)$ shown in the top panel. As above, we assume ${\theta({r_{\rm RM}})}=100\muas$, $i=j=30\degree$, an EW of a typical ($=26\,000\kms$), and ${v_{\rm rot}}$ at $r={r_{\rm RM}}$ such that $\hwhm=5000\kms$. We assume ${v_{\rm rot}}(r)\propto r^{-1/2}$, and based on the arguments in the previous section, $\sigma(r)/{v_{\rm rot}}(r)\sin\,i = 1$. The middle panel shows that $\Phi_v$ of the different $f(r)$ are similar, up to somewhat more extended line wings in the $f(r)\propto r^{\pm1}$ and @Baskin+14a distributions compared to the $f(r)\propto r^{\pm2}$ and $\delta$-function distributions. This similarity is expected since all assumed $f(r)$ drop sharply when $r$ is significantly larger than or significantly smaller than ${r_{\rm RM}}$. In contrast, the shape of $S_v$ seen in the lower panel differ significantly for the different $f(r)$. At $v=2000\kms$, for example, $S_v=25\muas$ for $f(r)\propto r^{\pm2}$, compared to $S_v=60\muas$ for $f(r)\propto r^{\pm1}$ and $S_v=90\muas$ for the $f(r)$ calculated by @Baskin+14a However, at $|v|\approx\hwhm=5000\kms$, the dependence of $S_v$ on $f(r)$ is significantly weaker than at $v=2000\kms$, where $S_{v=\hwhm}$ is in the range $25-37\muas$ for all considered $f(r)$.
![ [**(Top)**]{} Possible radial distributions of the line emission. The black arrow is a $\delta$-function (thin ring), as used in Figs. \[fig: setup\] and \[fig: turbulence\]. The red solid line and the green dashed line are parameterizations of $f(r)$ which are consistent with RM studies, where $f(r)$ peaks at $r={r_{\rm RM}}$ and drops at higher and lower $r$. The top axis notes the physical scale for a quasar with $L_{1450\AA}=10^{47}\ergs$. The blue dash-dotted line is a physically-motivated -emission distribution from Baskin et al. (2014, fig. 5 there), where the line emission drops at $r>\rsub=2{r_{\rm RM}}$ due to absorption of the incident ionizing photons by dust grains, and also drops at $r\ll{r_{\rm RM}}$ due to collisional suppression of . [**(Middle)**]{} The line spectral profiles for the different $f(r)$ shown in the top panel, assuming $\sigma/({{v_{\rm rot}}\sin i})=1$ and $\hwhm=5000\kms$. The spectral profiles are similar to each other. [**(Bottom)**]{} The expected photocenter offsets for the different $f(r)$. At $|v|<\hwhm$, $S_v$ is highly sensitive to the relatively weak emission from $r\gg{r_{\rm RM}}$. At $|v|\approx\hwhm$, $S_v$ only weakly depends on the assumed $f(r)$. []{data-label="fig: r-distribution"}](rDistribution.eps)
Why is $S_v$ at $v<\hwhm$ so sensitive to emission from large $r$? As seen in eqn. (\[eq: S\_v\]), $S_v$ is a photon-weighted average of the contribution to $S_v$ from each $r$. Eqn. (\[eq: Sv1\]) demonstrates that the contribution to $S_v$ from a specific $r$ scales as $\propto r{v_{\rm rot}}^{-1}\propto r^{3/2}$, for $v\leq{v_{\rm rot}}(r)\sin i$. While eqn. (\[eq: Sv1\]) is derived assuming $\sigma\ll{{v_{\rm rot}}\sin i}$, the bottom panel of Fig. \[fig: turbulence\] shows that it is also roughly a correct description of $S_v$ for general $\sigma$, up to multiplication by a constant which depends on $\sigma$. Therefore, in the case where $f(r)\propto r^{-1}$ at $r>{r_{\rm RM}}$, this $r^{3/2}$ term implies that $S_v$ is dominated by photons from the largest $r$ with significant emission at $v$, despite that these large-$r$ photons have a negligible contribution to the line profile. This effect is even more pronounced for the physical $f(r)$ calculated by [@Baskin+14a], in which more photons originate from $r>10~{r_{\rm RM}}$ than in the $f(r)\propto r^{\pm1}$ case.
To conclude, the spectroastrometry signal at $v<\hwhm$ is very sensitive to $f(r)$ at $r>{r_{\rm RM}}$, and therefore can be used to probe the line emission from this regime, as further discussed in §\[sec: discussion\]. At $v$ comparable to the , $S_v$ probes ${r_{\rm RM}}$, i.e. $S_v$ probes the scale at which the bulk of the BLR emission is produced.
Contaminants
------------
In a real observation, some of the photons with wavelengths spanned by the observed broad line originate from non-BLR sources, and therefore we need to account for their affect on $S_v$. We have already referred to the effect of quasar continuum photons, which dilute $S_v$, and can be accounted for by dividing the observed $S_v$ by the fraction of line photons at each $v$ (see §\[sec: spectroastrometric signal\]). In this section, we discuss additional non-BLR contaminants and how to account for their effect on $S_v$.
### Narrow emission lines {#sec: NLR}
Photons from the NLR originate from $r\gg{r_{\rm BLR}}$, and hence any asymmetries on NLR scales (as seen in fig. 5 of @Bailey+98) may dominate $S_v$ at $v$ with narrow line emission. The largest relevant $r$ is ${r_{\rm PSF}}$, the physical size spanned by the [PSF]{}, since emission from $r>{r_{\rm PSF}}$ can likely be filtered out during the data reduction phase of the spectroastrometric observation. For an assumed ${\rm PSF}=70\mas$ we get ${r_{\rm PSF}}\approx 10^3\,{r_{\rm BLR}}$ (${r_{\rm PSF}}=600\pc$ at $z=2$). Now, since in the calculation of $S_v$ (eqn. \[eq: S\_v\]) there is a factor of $r$ in the integrand, NLR emission which originates from $10^3\,{r_{\rm BLR}}$ may dominate $S_v$, even if the NLR flux density is only $\sim0.1\%$ of the broad line flux density. This is the same argument made in §\[sec: r-distribution\] for why weak BLR emission from large scales can dominate $S_v$. Hence, photons from wavelengths near narrow emission lines will need to be disregarded when measuring the photocenter of the BLR.
For example, at a broad luminosity of $10^{44}\ergs$, which is equivalent to $L=10^{46}\ergs$ (@SternLaor12a), the mean fluxes of the narrow , , and lines are $0.3-3\%$ of the broad flux (fig. 3 in @SternLaor12b, fig. 1 in @SternLaor13). The flux [*density*]{} ratio of and the narrow is $\sim10$ times this value for a typical NLR-to-BLR ratio of $0.1$, while and , which reside on the broad wings, have even higher flux density ratios. Therefore, since the NLR to BLR flux density ratios is $\gg0.1\%$ at $v$ with these strong narrow lines, the relevant $S_v$ will likely be dominated by the NLR.
We search also for weaker lines in the (@Ferland+13) model of the NLR described in [@Stern+14][^7]. We find three additional lines which may be strong enough to dominate $S_v$. The lines are , , and , at $v_\Ha=-11\,500,~5840,$ and $5270\kms$, and have fluxes which are $0.08$, $0.03$, and $0.01$ times the flux of the narrow , respectively. The latter two lines can be seen as small bumps in the high resolution spectra of NGC 4151 (fig. 1 in @Arav+98). Photons near these lines may also have to be disregarded. In §\[sec: simulation\] below we demonstrate that the emission from even weaker lines will not significantly affect $S_v$.
We note that the above estimates for the NLR strength is likely an upper limit, since spectroastrometry candidates have $L\gg10^{46}\ergs$, while the narrow to broad flux ratio decreases with $L$ as $\sim L^{-0.3}$, probably due to the decrease in NLR covering factor with increasing $L$ (@SternLaor12b). Additionally, some of the narrow line emission may originate from $r>{r_{\rm PSF}}$ and therefore can be filtered out.
### Continuum emission from the host galaxy, distant scattering media, and dust grains {#sec: host}
At optical wavelengths, the polarized flux density in unobscured AGN is typically $0.5-5\%$ of the total flux density (@Smith+02), suggesting that some of the observed emission is scattered emission. The scattering medium may reside on scales $\gg{r_{\rm BLR}}$, and therefore, based on the same reasoning used above for the NLR, asymmetries in this scattering medium can in principle dominate $S_v$. However, because a distant scatter views the BLR as a point source, the opening angle covered by the scatterer will not be a function of location in the BLR, and hence not a function of velocity. Since the scattering efficiency is a weak function of $\lambda$, and a $v=10\,000\kms$ emission line spans merely $\d\lambda/\lambda=0.03$, the scattering efficiency is also a weak function of velocity. Therefore, we expect the contribution of the scattered flux to $S_v$ to be roughly constant across the emission line and the flanking continuum. We can account for the effect of a distant scatterer by interpolating $S_v$ from the continuum flanking the broad line profile, and subtracting out the result. This approach will also account for the effect of the host galaxy emission on $S_v$, since any asymmetry in the host galaxy is also likely to be a weak function $\lambda$.
We note that the rotation of the polarization angle across the profile found by [@Smith+02; @Smith+05] suggests the existence of scattering media also at $r\gtrsim {r_{\rm BLR}}$. Such nearby scatterers will not dominate $S_v$.
If one measures the spectroastrometric signal of a rest-frame IR emission line such as , then the dust IR continuum can affect $S_v$. The $S_v$ profile of the dust continuum is also expected to be a weak function of $\lambda$, and therefore can be accounted for using the same interpolation method noted above for accounting for scattering media. However, since the dust continuum is a significant fraction of the flux at IR wavelengths, in contrast with the host and scattered flux, any second-order dependence of $S_v$ on $\lambda$ which is not accounted for may significantly affect $S_v$, and therefore may pose a problem for performing spectroastrometric on rest-frame IR lines.
Observational Considerations
============================
Systematics {#sec: noise}
-----------
We note that most of the atmospheric and instrumental systematic errors relevant to imaging astrometry are irrelevant to BLR spectroastrometry, since in BLR spectroastrometry one measures the astrometric position of the broad emission line relative to the adjacent continuum, rather than in any absolute sense. Specifically, the red and blue sides of the broad emission line are expected to be offset from the continuum in different directions (Figs. \[fig: setup\] – \[fig: r-distribution\]). Most systematic effects on the photon centroid are expected to be a weak function of $\lambda$, since the line and continuum photons travel in the exact same way through the atmosphere and instrument, and are only split in the wavelength domain. Therefore, to mimic a spectroastrometric signal, any instrumental systematic needs to vary so strongly with $\lambda$ that within $\d\lambda/\lambda \approx 0.03$ it produces a relative offset between the red and blue wings of the emission line. Any systematic which is a weak function of $\lambda$ can be accounted for by interpolating the centroid of the continuum flanking the emission line, a method which accounts also for the effect of astrophysical continuum sources such as the host galaxy or scattering media (§\[sec: host\]). The error of this interpolation is limited by the photon count of the continuum, and is estimated below.
On the other hand, instrumental systematics which are a strong function of $\lambda$ may be a limiting factor, rather than the number of photons collected. [@Pontoppidan+11] showed that rotating the slit by $180\deg$ can mitigate such systematics which are related to flat-fielding and removal of telluric lines. This method is based on the fact that the spectroastrometric offset is a real offset between line photons on the sky, and therefore a $180\deg$ rotation of the slit flips the direction of the spectroastrometry offset in detector coordinates, while most instrumental systematics would be fixed in detector coordinates. By applying this $180\deg$ flip, @Pontoppidan+11 reached [*photon-limited*]{} accuracies of $100-500\muas$, i.e. at an accuracy of $100\muas$ their observations were still free of systematic effects. We therefore find it reasonable that systematics are, or can be reduced to $\sim25\muas$, the accuracy level required for BLR spectroastrometry.
An additional method to differentiate between a true signal and an instrumental systematic would be to test the predicted $j$-dependence of the spectroastrometric signal (eqn. \[eq: S\_v\]), using multiple slit orientations. In order for an instrumental systematic to mimic the spectroastrometric signal it would to have to behave exactly the same way with rotation of the slit.
Choice of Target and Emission Line
----------------------------------
The choice of an optimal target and emission line for spectroastrometry is motivated by four main factors. The angular size of the emitting region, the photon flux of the line, the strength of the non-BLR contaminants listed in the previous section, and the observed wavelength of the line. The $\Ha$ line scores highly given all of these considerations, since it is the strongest optical emission line, it is emitted from the outer part of the BLR (@Bentz+10), and its wavelength is near $1\mic$ where the accretion disk and dust continuum emission are at a minimum. Furthermore, at $z=2-2.5$ where ${\theta({r_{\rm RM}})}$ is near its peak value (Fig. \[fig: angular size\]), H$\alpha$ is redshifted into the near-IR K-band where adaptive-optics works best. Also, we note that the idea of an ordered rotational velocity field in the BLR is best established for low-ionization lines, such as the Balmer lines. Higher ionization lines appear also to have an outflowing wind component (@Richards+11, see further discussion in §\[sec: discussion\]). Hence, interpreting the spectroastrometic signal of high-ionization lines will likely be more complicated than interpreting the signal of low ionization lines.
Given the above arguments, below we simulate the expected spectroastrometry signal for . We consider also the prospect of measuring $S_v$ on , which falls in the NIR bands at $z=3-7.5$, and also $S_v$ of in the local quasar 3C 273, the brightest quasar on the sky.
Choosing the Slit Orientation {#sec: j}
-----------------------------
Since $S_v\propto\cos j$ (eqn. \[eq: S\_v\]), the required number of photons to achieve a given $\snr$ scales as $\cos^{-2}j$, where $j$ is not known a-priori. In the context of measuring the spectroastrometric signal of stellar disks, [@Pontoppidan+08] suggested an observing strategy where one observes the target at three slit orientations rotated by $60\degree$ from one another. Since $\left(\cos^2 (j) + \cos^2 (j-60) + \cos^2 (j-120)\right)/3 = 0.5$ for all $j$, than the required observing time with this technique is twice the observing time required for the $j=0$ case. Therefore, below we assume $j=0$, and multiply the required observing time by a factor of two to account for the unknown $j$.
We note that even if one performs a spectroastrometric observation with an Integral Field Unit (IFU), which maps the sky into 2D pixels and preforms simulatenous spectroscopy in each pixel, rotation of the unit is still likely to be required. An IFU maps one spatial direction continuously onto the detector, while the other spatial direction is mapped discontinuously. Constructing the spectroastrometic signal along the spatial dimension which is mapped continuously onto the detector is tantamount to constructing the signal in a long-slit observation. However, constructing the spectroastrometic signal along the dimension which is mapped discontinously requires modelling the projection of this spatial dimension onto the detector to the level of $10^{-3}$ of a pixel, as required for BLR spectroastrometry, and is therefore unlikely to be feasible. Therefore, using an IFU has no clear advantage over using a long-slit.
Simulated Spectroastrometry Signal {#sec: simulation}
==================================
In this section we simulate the spectroastrometric signal for a $z=2$ target on 8m and 39m telescopes, and then consider measurements of targets at other redshifts.
![The simulated spectroastrometric signal of the broad in J1521+5202, a ${L_{1450\AA}}=10^{47.6}\ergs$ quasar at $z=2.2$. [**(Top panel)**]{} The yellow line plots the normalized spectrum of J1540-0205, used as a template for J1521+5202. The black line plots the fit to the continuum, broad , and the noted narrow lines (marked by vertical dotted lines). [**(2$^{\rm nd}$ panel)**]{} The expected $S_v$, assuming $\sigma/({{v_{\rm rot}}\sin i})=1$ and the $f(r)\propto r^{\pm2}$ shown in Fig. \[fig: r-distribution\]. The narrow lines likely have asymmetry on large scales, therefore at $v$ with strong NLR emission $S_v$ is on the scale of the PSF, which is beyond the plotted axis. The weak and lines have a comparable contribution to $S_v$ as the broad . At $v$ which are not dominated by narrow lines (thick line), the broad and quasar continuum dominate $S_v$. The photocenter at $v=4000\kms$ is offset from the photocenter at $v=-4000\kms$ by $85\muas$. [**(3$^{\rm rd}$ panel)**]{} The solid line marks the expected $S_v$ at $v$ where the NLR does not dominate $S_v$. The error bars are the simulated centroid statistical uncertainties, for spectral bin sizes of $1000$ and $4000\kms$ in the emission line and continuum, respectively. We assume a $10\hr$ integration on an AO-assisted 8m telescope, with ${{\it Strehl}}=0.4$ and ${f_{\rm inst}}=0.2$. [**(Bottom panel)**]{} The expected $S_v$ and centroid uncertainties assuming the physically-motivated $f(r)$ shown in Fig. \[fig: r-distribution\]. []{data-label="fig: expected signal"}](expectedprofile.eps)
The H$\alpha$ Line of SDSS J1521+5202
-------------------------------------
SDSS J152156.48+520238.5 (hereafter J1521+5202) is the most luminous SDSS quasar where falls in the K-band (Fig. \[fig: angular size\]). J1521+5202 has ${L_{1450\AA}}=10^{47.6}\ergs$, ${\theta({r_{\rm RM}})}=111\muas$, and $z=2.208$. The $\fwhm$ of its line is $9300\kms$ (@Shen+11). As the NIR spectrum of J1521+5202 is not available, we use as a template the spectrum of SDSS J154019.57-020505.4 (hereafter J1540-0205), a ${L_{1450\AA}}=10^{45.4}\ergs$ quasar at $z=0.3$ with $\fwhm(\Ha)=9730\kms$. The two spectra should be similar since the $\fwhm$ of the Balmer lines is correlated with the $\fwhm$ of (@Shen+08). We multiply the observed photon flux of J1540-0205 by the photon flux ratio of the two quasars, $10^{47.6}/10^{45.4}\cdot(\dL(z=2.2)/\dL(z=0.3))^2\cdot(1+2.2)/(1+0.3)=3.1$. The normalized spectrum is shown in the top panel of Figure \[fig: expected signal\], together with the broad , narrow lines and continuum fits described in [@SternLaor12a], which we utilize below. We note that since J1521+5202 is more luminous than J1540-0205, the host galaxy is expected to be relatively weaker, as are the narrow emission lines (@SternLaor12b). We determine the flux of the weak , , and lines from the flux of the narrow and the model flux ratios mentioned in §\[sec: NLR\]. These lines are too weak to be observable in the SDSS spectrum.
In order to estimate $S_v$, the broad emission line is modeled as a rotating disk with $\sigma/({{v_{\rm rot}}\sin i})=1$ and either $f(r)\propto
r^{\pm2}$, or the $f(r)$ based on [@Baskin+14a]. We assume for simplicity that the angular size of the -emitting region is ${\theta({r_{\rm RM}})}$, though the true $\theta$ may be higher since ${r_{\rm RM}}$ was determined on , and is sometimes observed to have longer RM lags than (@Bentz+10). We set ${v_{\rm rot}}({r_{\rm RM}})\sin i$ of the model such that the $\fwhm$ in the model profile equals the observed $\fwhm$ of $9730\kms$, i.e. $2{v_{\rm rot}}({r_{\rm RM}})\sin i =
7400\kms$ for the $f(r)\propto r^{\pm2}$ case and $2{v_{\rm rot}}({r_{\rm RM}})\sin i
= 11\,000\kms$ for the $f(r)$ from @Baskin+14a We normalize the model $\Phi~(\equiv \int\Phi_v\d v)$ to give the same value of the observed $\Phi$. We note that at $|v|<5000\kms$, the observed $\Phi_v$ profile of the broad differs by up to 20% from the $\Phi_v$ calculated by the BLR model. However, since the goal of this study is to estimate $S_v$ rather than provide a model which accurately fits the observed emission line, we disregard this difference.
In order to estimate $S_v$ of the narrow lines, we assume the worst-case scenario where the narrow lines originate from $r={r_{\rm PSF}}=600\pc$ (assuming an 8m-telescope diffraction limited [PSF]{} of $70\mas$), where the contribution to $S_v$ is maximal (§\[sec: NLR\]). For simplicity, we assume the NLR emission also originates from a turbulent rotating disk with ${{v_{\rm rot}}\sin i}$ chosen to fit the observed narrow lines of $450\kms$. The total photon flux of the NLR model is chosen to fit the observed photon fluxes of the strong emission lines (, , and ), and the model photon fluxes of the weak emission lines (, , ). Finally, quasar continuum photons are assumed to originate from $r=0$.
The second panel in Fig. \[fig: r-distribution\] shows $S_v$ versus $v$ for $f(r)\propto r^{\pm2}$. Thin lines mark $v$ values which are contaminated by emission from narrow lines. The NLR spectroastrometric signal is on the scale of the PSF, which is beyond the plotted axis. Note that the weak signal of justifies our choice in §\[sec: NLR\] to disregard narrow lines with a flux which is lower than the flux of . The thick line marks the range of $v$ where the broad and quasar continuum dominate $S_v$, and only these $v$ are used further in the analysis. The photocenters at the blue and red wings are offset by up to $85\muas$ from each other.
To estimate the expected error on the measured photocenter, we assume a $t=10\hr$ observation on an $8{\rm m}$ telescope. Using the specifications of the Nasmyth Adaptive Optics System (NAOS) and COud[' e]{} Near Infrared CAmera (CONICA) as guidance[^8], we assume a collecting area of $A=38{\rm m}^2$, a diffraction-limited $\fwhm_{\rm {PSF}}=70\mas$, a [[*Strehl*]{}]{} ratio of $0.4$, and an instrument photon collecting efficiency of ${f_{\rm inst}}=0.2$. For the observed $\Phi_v\simeq10^6~[\m^{-2}\hr^{-1}\left(10^3 \kms\right)^{-1}]$ at half maximum of the broad (top panel of Fig. \[fig: r-distribution\]), the implied number of collected photons per $1000\kms$ is $$\label{eq: Nph}
\frac{\d N_{\rm ph}}{\d v} = \Phi_v \cdot A \cdot t \cdot {{\it Strehl}}\cdot {f_{\rm inst}}\cdot 0.5
= 15\times10^6 ~\left(10^3 \kms\right)^{-1} ,$$ where the factor of $0.5$ reduction arises from the slit orientation as discussed in §\[sec: j\].
The value of $\d N_{\rm ph}/\d v$ computed in eqn. (\[eq: Nph\]) implies a statistical $1\sigma$ photocenter positioning error $\epsilon_v$ of $$\label{eq: epsilon}
\epsilon_v = \frac{\fwhm_{\rm {PSF}}}{2.35\, \left(\frac{\d N_{\rm ph}}{\d v}\right)^{1/2}} = 7.7 \muas ~~~.$$ The third panel in Fig. \[fig: expected signal\] shows the expected errors in the photocenter measurements, for bins of size $\Delta v=1000\kms$ in the red and blue wings, and bins of size $\Delta v=4000\kms$ for the red and blue continuum regions. To simulate an observation, the center of the simulated errorbars is offset from the expected signal (solid line) by a random offset chosen from a Gaussian distribution with dispersion $\epsilon_v$. Fig. \[fig: expected signal\] demonstrates that if systematics can be reduced to the level of a few times the value of $\epsilon_v$ calculated in eq. \[eq: epsilon\], then the photocenter offset of the red wing from the blue wing is detectable with existing telescopes. For comparison, [@Pontoppidan+11] reached photon-limited uncertainties of $100\muas$ in the context of stellar disks, but this precision level was set by the number of photons collected rather than by systematic errors.
![Simulated spectroastrometric signal with a 39m telescope. As in the bottom panels of Fig. \[fig: expected signal\], the solid line denotes the expected $S_v$ at velocities in which the BLR and quasar continuum dominate $S_v$. Error bars denote the expected statistical uncertainty in the photocentroid measurement. In both panels, $f(r)$ from [@Baskin+14a] is assumed. [**(Top panel)**]{} The signal expected on the broad of J1521+5202, the object which appears also in Fig. \[fig: expected signal\]. The higher photon flux and smaller PSF in 39m telescopes compared to 8m telescopes enables constraining $S_v$ to an order of magnitude higher velocity resolution, with a fraction of the observing time. [**(Bottom panel)**]{} The signal expected on the broad line of SDSS J114816.64+525150.3, a $z=6.4$ quasar. []{data-label="fig: 39m"}](t39.eps)
In the bottom panel of Fig. \[fig: expected signal\], we show the expected spectroastrometric signal assuming the $f(r)$ based on [@Baskin+14a]. As expected from Fig. \[fig: r-distribution\], at $v<\hwhm$ the maximum expected $S_v$ is a factor of four larger than when assuming $f(r)\propto r^{\pm2}$, and the relative errorbar sizes are correspondingly smaller.
In the top panel of Figure \[fig: 39m\], we show the expected spectroastrometric signal of J1521+5202 for a $10\min$ observation on a next-generation $39\m$ telescope, zoomed-in on the emission line for clarity. We assume that $A\propto d^2$, $\fwhm_{\rm {PSF}}\propto d^{-1}$, a [[*Strehl*]{}]{} of $0.4$ (e.g. @Clenet+10), and ${f_{\rm inst}}=0.4$ (e.g. @Davies+10). Hence, eqs. \[eq: Nph\] and \[eq: epsilon\] imply that $\epsilon\propto d^{-2}$, and that the required observing time to reach a certain value of $\epsilon$ scales as $d^{-4}$. We therefore decrease the spectral bin sizes to $200\kms$. With a next generation telescope, one can derive $S_v$ to a high velocity resolution, and therefore yield strong constraints on the BLR kinematics.
Spectroastrometry Estimates for Additional Quasars
--------------------------------------------------
In this section, we estimate the statistical $\snr$ of the spectroastrometric signal of the most luminous quasars at different $z$. To this end, we define $\snr(S{_{\rm red}}-S{_{\rm blue}})$, to be the photocenter offset between the red and blue wings of a broad emission line, which is an aggregate $\snr$ that effectively combines the data points shown in Figs. \[fig: expected signal\] and \[fig: 39m\]. For each wing, we include all photons with velocities $0 < |v| < \fwhm$, excluding velocities which might be contaminated by narrow line emission. We address quasars in which , , or fall in one of the NIR transmission windows, and require $\lrest$ is at least $10\,000\kms$ away from the edge of the atmospheric window, in order to adequately sample the quasar continuum and avoid significant telluric absorption. These quasars are listed in Table \[tab: snr\], which appears in the Appendix. The of less luminous quasars can then be estimated by noting that $$\label{eq: snr vs. L}
\snr(z) \propto \frac{{\theta({r_{\rm RM}})}}{\epsilon_v} \propto \frac{L^{1/2}}{N_{\rm ph}^{-1/2}} \propto L$$ where we used eqs. \[eq: Kaspi05\] and \[eq: epsilon\] for the dependence of ${\theta({r_{\rm RM}})}$ and $\epsilon_v$ on $L$, respectively, and emphasize that eqn. (\[eq: snr vs. L\]) is only valid at a fixed redshift. We assume a [[*Strehl*]{}]{} ratio of $0.2$, $0.4$, and $0.4$, for the J, H and K bands, respectively, and ${f_{\rm inst}}=0.2~(0.4)$ for all bands for an 8m (39m) telescope[^9].
Potential narrow line contamination near and is derived in a similar fashion as for narrow lines near (§\[sec: NLR\]). Narrow lines with a large enough flux to significantly effect $S_v$ near include the narrow doublet, a line at $v_\mgiip=-7000\kms$, an line at $v_\mgiip=6000\kms$, and a line at $v_\mgiip=6100\kms$. Similarly, near there are the and lines, three He lines at $v_\Pa=-255,\,-1055$ and $-1855\kms$, and an line at $v_\Pa=2140\kms$. All photons with $v$ within $750\kms$ of these lines are excluded from the analysis. We note that some of the mentioned lines are permitted lines, and therefore should also exhibit broad line emission, however these broad lines are likely too weak to affect $S_v$ significantly.
The rotating disk parameters of the model are the same as in the previous section, with ${v_{\rm rot}}({r_{\rm RM}})$ and the normalization chosen so the model reproduces the estimated values of $\fwhm$ and $\Phi$ listed in Table \[tab: snr\] in the Appendix. Since not all targets have available NIR spectra, we estimate $\fwhm$ and $\Phi$ from observed emission lines, as detailed in the Appendix. For , we assume that the total flux is contributed from the two lines in the doublet (separated by $770\kms$) according to a $2:1$ ratio. Therefore, at a given $v$ the expected centroid is the photon-flux-weighted centroid of the individual lines in the doublet.
The value of $\snr(S{_{\rm red}}- S{_{\rm blue}})$ is the number of standard deviations from which $S{_{\rm red}}- S{_{\rm blue}}$ deviates from a straight line interpolation of the photon centroids of the continuum bins. It is equal to $$\label{eq: snr red - blue}
\snr(S{_{\rm red}}- S{_{\rm blue}}) = \frac{S{_{\rm red}}- S{_{\rm blue}}}{\left(\epsilon{_{\rm red}}^2 + \epsilon{_{\rm blue}}^2 + \epsilon{_{\rm cont}}^2\right)^{1/2}}$$ where $\epsilon{_{\rm blue}}$ and $\epsilon{_{\rm red}}$ are the statistical errors on the photocenters of blue and red wing photons, respectively, calculated from the expressions in eqs. \[eq: Nph\] and \[eq: epsilon\]. The error on the continuum level, $\epsilon{_{\rm cont}}$, is equal to the error on the evaluation of a straight line interpolation from the data points in the blue and red continuum regions: $$\label{eq: eps_cont}
\epsilon{_{\rm cont}}= \frac{v{_{\rm red}}- v{_{\rm blue}}}{v{_{\rm red\,cont}}- v{_{\rm blue\,cont}}} \left( \epsilon{_{\rm red\,cont}}^2 + \epsilon{_{\rm blue\,cont}}^2\right)^{1/2} ~~~,$$ where $\epsilon{_{\rm blue\,cont}}$ and $\epsilon{_{\rm red\,cont}}$ are the statistical errors on the photocenters of each continuum bin ($\fwhm<|v|<2\,\fwhm$), and the continuum flux is calculated from the emission line EW noted in Table \[tab: snr\]. The values of $v{_{\rm red}}$ and $v{_{\rm blue}}$ are the photon-flux average $v$ of the red and blue wing bins, respectively, while $v{_{\rm red\,cont}}=-v{_{\rm blue\,cont}}=1.5\,\fwhm$ are the average velocities of the continuum bins. We note that in the $\snr$ calculations shown in Fig. \[fig: SN vs. z\], $\epsilon{_{\rm cont}}^2 / (\epsilon{_{\rm red}}^2 + \epsilon{_{\rm blue}}^2) = 0.07 - 0.3$, and therefore including $\epsilon{_{\rm cont}}$ in the calculation does not significantly affect $\snr(S{_{\rm red}}- S{_{\rm blue}})$. However, in a real observation the amount of continuum photons may be limited if the emission line falls near the edge of an atmospheric window, which will increase $\epsilon{_{\rm cont}}$, and hence decrease $\snr(S{_{\rm red}}- S{_{\rm blue}})$.
The expected $\snr(S_{\rm red} - S_{\rm blue})$, for $8{\rm m}$ and $39{\rm m}$ telescopes, are shown in Figure \[fig: SN vs. z\]. At $1<z<2.5$, a night on an $8{\rm m}$ is sufficient to achieve $\snr(S_{\rm red} - S_{\rm blue}) \approx 10$, even when assuming $f(r)\propto r^{-2}$. Assuming the @Baskin+14a-based $f(r)$ increases the expected $\snr$ by a factor of $\approx2$, while an hour on a $39{\rm m}$ telescope increases the expected $\snr$ by a factor of $\approx10$ compared to $10$ hours on an $8{\rm m}$. Fig. \[fig: SN vs. z\] suggests that with next-generation telescopes, the spectroastrometic signal can be detected for quasars as distant as $z=6.5$. As an example, in the bottom panel of Fig. \[fig: 39m\] we plot a simulated observation of the line in SDSS J114816.64+525150.3, a $z=6.43$ quasar discovered by [@Fan+03].
Discussion {#sec: discussion}
==========
The Assumption that the BLR is a Rotating Disk
----------------------------------------------
For the purpose of considering ordered motion, we assumed that the BLR has a flattened geometry and that its kinematics are dominated by rotation. These assumptions are supported by several observations of the low-ionization BLR gas, which we briefly summarize below.
![ Expected S/N for spectroastrometry candidates at different $z$. Candidates are chosen from the most luminous quasars at each $z$ (Fig. \[fig: angular size\]), where a strong low-ionization emission line falls in one of the NIR bands. The y-axis is the S/N of the centroid offset of the blue and red wings of the emission line (eqn. \[eq: snr red - blue\]), for $10\hr$ on an 8m telescope (small symbols), and $1\hr$ on a 39m telescope (large symbols). The marker type denotes the assumed radial distribution of the line emission (Fig. \[fig: r-distribution\]), and bold markers denote the examples shown in Figs. \[fig: expected signal\] – \[fig: 39m\]. The spectroastrometric signal of the BLR in $1<z<2.5$ quasars is detectable at $\snr>10$ with existing telescopes, while next-generation telescopes will be able to observe the spectroastrometric signal up to $z=6.5$. []{data-label="fig: SN vs. z"}](SNR.eps)
Evidence for a flattened geometry comes from the AGN unification picture. According to the the standard AGN paradigm, in unobscured AGN (known as type 1 AGN, including quasars) the accretion disk is viewed at angles near face-on. When an AGN disk is viewed edge-on, the accretion disk and broad line emission are obscured by a geometrically thick dusty ‘torus’ (@Antonucci93). [@Maiolino+01] and [@Gaskell09] argued that if the BLR is spherically symmetric, then a fraction of the quasars, equal to the BLR covering factor of $\sim30\%$, should be viewed through the BLR clouds. Such BLR gas along the sightline would produce Lyman-edge absorption and low ionization broad absorption lines (LoBALs). However, only 1% of quasars are LoBALs (@Trump+06), in contrast with the prediction of a spherically symmetric BLR. Hence, a more likely picture is that the low-ionization BLR resides predominantly near the disk-plane in a flattened geometry, along line of sights where the quasar would be viewed as obscured.
As for our assumption of a rotating BLR, rotation is preferred over an outflow-dominated BLR, since in a scenario where the BLR is an outflow off the face of the disk, the redshifted and blueshifted photons will originate from opposite locations along the minor-axis of the projected BLR ellipse, in contrast to along the major-axis in the rotating BLR scenario (see top panel of Fig. \[fig: setup\]). This geometry implies a shorter response time of the blue wing compared to the response time of the red wing, which is ruled out in velocity-mapped RM studies of low ionization lines (@Maoz+91 [@Grier+13; @Pancoast+13]), albeit for low-$z$ low-$L$ AGN. Further support for the rotating BLR assumption comes from the mean redshift of the broad relative to systemic in SDSS quasars, which is consistent with the relativistic redshift expected in a rotating BLR (@Tremaine+14), and from the analysis of the microlensing signal of the quadruply lensed quasar HE 0435-1223 (@Braibant+14). However, the high-ionization broad lines likely have a significant outflowing component. Evidence for an outflow can be seen in the line profile of $\lambda1450$ (@BaskinLaor05 [@Richards+11]), and in the high-ionization states typically observed in broad absorption lines (@Hamann97), which are predominantly blueshifted, suggesting an outflow. Since in an outflow scenario the spatial separation of the redshifted and blueshifted gas is along the minor axis, which is smaller than the major axis by a factor of $\cos i$, the expected $S_v$ should also be correspondingly smaller (assuming the same ${r_{\rm BLR}}$). Since this signal reduction is relatively small, the spectroastrometric signal should be detectable also in an outflowing BLR.
Note that as the rotational motion in the BLR decreases relative to the random motion, $S_v$ is reduced as shown in the bottom panel of Fig. \[fig: turbulence\]. In J1521+5202, for example, the $\snr(S{_{\rm red}}-S{_{\rm blue}})=17$ expected for $10\hr$ on an 8m telescope (Fig. \[fig: SN vs. z\]) is reduced to $\snr(S{_{\rm red}}-S{_{\rm blue}})<5$ if $\sigma/({{v_{\rm rot}}\sin i})>3$. Therefore, even a null detection of the spectroastrometric signal would place an interesting upper limit on the fraction of the BLR motion that is ordered.
$\mbh$ Estimates
----------------
Above we have demonstrated that by measuring the BLR photocenter offset $S_v$, one can derive ${r_{\rm BLR}}$. Given a measurement of ${r_{\rm BLR}}$ and an estimate of the Keplerian velocity ${v_{\rm K}}$ based on the line width, one can derive $\mbh = {r_{\rm BLR}}{v_{\rm K}}^2 / G$, analogous to RM-studies. Therefore, since any uncertainty in the ${r_{\rm BLR}}$ measurement propagates to the $\mbh$ estimate, it is interesting to understand the implied uncertainty in ${r_{\rm BLR}}$ given a measurement of $S_v$.
For a given ${r_{\rm BLR}}$, different values of $\sigma$ imply changes of a factor of two in $S_v$ (Fig. \[fig: turbulence\]). Also, Fig. \[fig: r-distribution\] shows that while different radial distributions of the line emission can change the $S_v$ profile significantly, the range of $S_v$ at $|v|=\hwhm$ for different $f(r)$ spans merely a factor of $1.5$. So, in the worst case scenario where nothing is known about $\sigma$ and $f(r)$, for a given measurement of $S_v$ at $|v|=\hwhm$ one can estimate ${r_{\rm BLR}}$ to within a factor of $\sim3$. However, this uncertainty can be reduced by constraining $\sigma$ from the line profile, and by constraining $f(r)$ using the measurements of $S_v$ at $|v|<\hwhm$. For the purpose of estimating $\mbh$, these uncertainties on the ${r_{\rm BLR}}$ estimate should be added to the uncertainty on ${v_{\rm K}}$ due to the unknown inclination $i$ of the disk plane to the line of sight. The total implied accuracy to which $\mbh$ can be derived is an important topic for future study.
Furthermore, $\sigma$ and $f(r)$ are likely to be similar or the same for all BLRs. In this case, one can tie the spectroastrometry measurements of the BLR to lower-luminosity sources which have been reverberation mapped, potentially removing this source of ‘noise’. With next-generation telescopes, one can potentially perform spectroastrometry on objects which are low luminosity enough to be also reverberation mapped.
8m-class Telescopes, 30m-class Telescopes and Space Telescopes
--------------------------------------------------------------
Fig. \[fig: SN vs. z\] shows that existing telescopes can detect the BLR spectroastrometric signal in in the most luminous quasars at $1<z<2.5$. Therefore, with existing telescopes one can use spectroastrometry to estimate the $\mbh$ of the most luminous quasars at the peak of the quasar epoch. Also, the measured ${r_{\rm BLR}}$ can be compared to the extrapolation of the ${r_{\rm BLR}}-L$ relation deduced by RM studies on lower luminosity AGN (eqn. \[eq: Kaspi05\]).
In §\[sec: simulation\] we show that the required integration time to reach a given $\snr$ scales as $d^{-4}$, where $d$ is the telescope diameter. Therefore, with next generation 30m-class telescopes, the required integration times are lower by a factor of $(30/8)^4=200$ than with 8m telescopes. Also, the smaller [PSF]{} of larger telescopes implies that reducing the systematics to a given angular precision is likely less challenging.
The relatively short integration times required in 30m telescopes can be utilized to observe a large sample of quasars (from eqn. \[eq: snr red - blue\] and Fig. \[fig: SN vs. z\], reaching $\snr=10$ on quasars with $L_{1450}=10^{47}\ergs$ at $z\approx2$ requires $7-15$ minutes on a 39m). Thus, one can explore the dependence of $S_v$ on additional BLR parameters: e.g. the BLR , the EW of the lines, $\mbh$ and $\lledd$, whether the high-ionization BLR shows signs of an outflow (@Richards+11), and whether the quasar spectrum shows Broad Absorption Lines, which might be some proxy for orientation. The high photon flux can also be utilized to achieve a higher resolution in $v$, as shown in the top panel of Fig. \[fig: 39m\], and thus further constrain the kinematic model of the BLR. Additionally, with next generation telescopes one can explore fainter quasars, such as quasars as distant as $z=6.5$ (Fig. \[fig: SN vs. z\]), and fainter emission lines. This opens the possibility to perform spectroastrometry on multiple lines simultaneously, at specific $z$ where multiple lines fall in the NIR atmospheric windows (say $\ciii~\lambda1909$ in J and $\mgiip$ in H at $z\sim5$). Thus, one can constrain the BLR structure as a function of the ionization level of the line-emitting gas.
Another interesting possibility is to perform BLR spectroastrometry using the James Webb Space Telescope (JWST, @Gardner+06), which has a diameter of 6.6m and allows full coverage at $0.6-5\mic$ without the issue of telluric absorption. Since ${{\it Strehl}}=1$ in a space telescope, compared to ${{\it Strehl}}=0.4$ assumed above for an 8m telescope, the observing time requirements are comparable $( (6.5/8)^{-4}\times 0.4=0.9 )$. The continuous wavelength coverage can allow observing multiple emission lines for the same quasar. Utilizing the Hubble Space Telescope (HST) for BLR spectroastrometry may also be possible, since the increased observing time due to the relatively small telescope diameter of 2.4m is partially offset by the lower diffraction limit in the UV. We differ a more thorough analysis of the possibility to perform BLR spectroastrometry with space telescopes to future work.
Comparison with the Interferometry Results of [@Petrov+12]
----------------------------------------------------------
One can also use differential interferometry to measure the photocenter offset of the BLR relative to the continuum. [@Petrov+12] applied this approach to the P$\alpha$ line of 3C 273, and measured an angular radius of $\gtrsim400\muas$. As noted there, this angular radius is larger by a factor of $\gtrsim3$ than the radius implied by RM of . We suspect that this apparent discrepancy between ${r_{\rm BLR}}$ measured by interferometry and ${r_{\rm BLR}}$ measured by RM is due to the high sensitivity of the photocenter to emission from large $r$ (Fig. \[fig: r-distribution\] and §\[sec: r-distribution\]). Fig. \[fig: r-distribution\] demonstrates that when a small fraction of the line photons come from $r\gg{r_{\rm RM}}$ (i.e. when $f(r)\propto r^{\pm1}$ or when using the physically-motivated $f(r)$ calculated by @Baskin+14a), then the average $S_v$ at $0<v<\hwhm$ increases by a factor of $3-4$ compared to when assuming a $\delta$-function radial distribution. This increase is consistent with the discrepancy found by @Petrov+12
We note that a key advantage of spectroastrometry over interferometry is sky coverage. Currently interferometers require co-phasing of the telescopes with a bright star or AGN ($\lesssim 12\mag$ in the NIR), which needs to be at a small angular separation on the order of the isoplanatic angle at the science wavelength (typically $<20\arcsec$, @Esposito+00). Only a handful of nearby quasars (Fig. \[fig: angular size\]) are bright enough such that a nearby star is unnecessary. In contrast, spectroastrometry can benefit from the $\sim50\%$ sky coverage provided by Laser-Guide-Star Adaptive Optics (e.g. @Diolaiti10).
Constraints on Torus Models
---------------------------
As first noted by [@NetzerLaor93], the line emission from gas at $r>\rsub$ which is exposed to the quasar ionizing radiation is suppressed, due to absorption of the ionizing photons by dust grains, rather than absorption of the ionizing photons by gas particles. This suppression, seen as a drop at $r>2{r_{\rm RM}}\approx\rsub$ in the @Baskin+14a $f(r)$ profile shown in the top panel of Fig. \[fig: r-distribution\], explains the apparent gap in line emission between the BLR and the NLR. This dusty gas at $r>\rsub$, usually referred to as the ‘torus’, also creates the dichotomy of obscured / unobscured AGN invoked in the standard AGN unification model discussed above. Due to the weak line emission of the dusty gas, the physical properties of the torus have hitherto mainly been constrained via the observed IR emission and via the demographics of obscured and unobscured AGN. However, in Figs. \[fig: r-distribution\] – \[fig: expected signal\] we show that this weak line emission from $r>\rsub$ can have a dramatic effect on $S_v$ at $|v|<\hwhm$, and thus a measurement of $S_v$ can provide a new constraint on torus models.
Summary
=======
In this paper, we have shown that the extrapolation of the ${r_{\rm BLR}}-L$ relation found by RM studies of local AGN to the most luminous quasars on the sky imply ${\theta({r_{\rm BLR}})}\approx 100\muas$, with a weak dependence on $z$. Comparable angular sizes have previously been resolved in young stellar objects using spectroastrometry (@Pontoppidan+08 [@Pontoppidan+11]), suggesting that spectroastrometry is applicable also to the BLR of quasars.
Assuming that the BLR emission has significant contributions from ordered motions, we calculated the expected spectroastrometric signal, using a simple model of a rotation-dominated BLR. Our calculation suggests that the offset between the photocenter of the red-wing and blue-wing photons of the broad in luminous quasars is detectable ($\snr\approx10$) with modest time allocations on existing telescopes, and therefore can be used to constrain ${r_{\rm BLR}}$ in high-$L$ and high-$z$ quasars.
This estimate of ${r_{\rm BLR}}$ implies a new method to spatially resolve the kinematics of the BLR and directly test whether BLR motions are ordered. It also implies a new method to contrain $\mbh$. Since previous methods to estimate $\mbh$ are based on either RM or on host properties, which work best in low-$L$ low-$z$ quasars, spectroastrometry is novel in its ability to estimate $\mbh$ in luminous quasars during the peak of the quasar epoch.
With next-generation telescopes, BLR spectroastrometry should be routinely detectable for much fainter quasars out to $z\sim 6$, and can be expanded to include multiple broad emission lines, including weak lines, at an order of magnitude higher velocity resolution than achievable with existing telescopes. This will enable demographic and statistical studies of $\mbh$ and BLR properties as a function of luminosity and redshift, providing a compelling science case for next generation telescopes.
Acknowledgements {#acknowledgements .unnumbered}
================
We thank Aaron Barth for comments which significantly improved the manuscript. We thank Eduardo Ba[ñ]{}ados for providing the list of high-$z$ quasars, and Alexei Baskin for providing the $f(r)$ values of shown in Fig. \[fig: r-distribution\]. We also thank Michael Strauss and the members of the ENIGMA group[^10] at the Max Planck Institute for Astronomy (MPIA) for helpful discussions. J.F.H. acknowledges generous support from the Alexander von Humboldt foundation in the context of the Sofja Kovalevskaja Award. The Humboldt foundation is funded by the German Federal Ministry for Education and Research.
[100]{} Angione, R. J., & Smith, H. J. 1972, External Galaxies and Quasi-Stellar Objects, 44, 171 Antonucci, R. 1993, , 31, 473 Arav, N., Barlow, T. A., Laor, A., Sargent, W. L. W., & Blandford, R. D. 1998, , 297, 990 Bailey, J. A. 1998, , 3355, 932 Ba[ñ]{}ados, E., Venemans, B. P., Morganson, E., et al. 2014, , 148, 14 Baskin, A., & Laor, A. 2005, , 356, 1029 Baskin, A., Laor, A., & Stern, J. 2014, , 438, 604 Bentz, M. C., Peterson, B. M., Netzer, H., Pogge, R. W., & Vestergaard, M. 2009, , 697, 160 Bentz, M. C., Walsh, J. L., Barth, A. J., et al. 2009, , 705, 199 Bentz, M. C., Horne, K., Barth, A. J., et al. 2010, , 720, L46 Bentz, M. C., Denney, K. D., Grier, C. J., et al. 2013, , 767, 149 Blanco C[á]{}rdenas, M. W., K[ä]{}ufl, H. U., Guerrero, M. A., Miranda, L. F., & Seifahrt, A. 2014, , 566, AA133 Blandford, R. D., & McKee, C. F. 1982, , 255, 419 Braibant, L., Hutsem[é]{}kers, D., Sluse, D., Anguita, T., & Garc[í]{}a-Vergara, C. J. 2014, , 565, L11 Chelouche, D., Daniel, E., & Kaspi, S. 2012, , 750, L43 Chen, K., Halpern, J. P., & Filippenko, A. V. 1989, , 339, 742 Chen, K., & Halpern, J. P. 1989, , 344, 115 Cl[é]{}net, Y., Bernardi, P., Chapron, F., et al. 2010, , 7736, 77363Q Davidson K., Netzer H., 1979, , 51, 715 Davies, R., Ageorges, N., Barl, L., et al. 2010, , 7735, 77352A De Rosa, G., Venemans, B. P., Decarli, R., et al. 2014, , 790, 145 Dietrich, M., Wagner, S. J., Courvoisier, T. J.-L., Bock, H., & North, P. 1999, , 351, 31 Diolaiti, E. 2010, The Messenger, 140, 28 Esposito, S., Riccardi, A., & Femen[í]{}a, B. 2000, , 353, L29 Fan, X., Strauss, M. A., Schneider, D. P., et al. 2003, , 125, 1649 Ferland, G. J., Porter, R. L., van Hoof, P. A. M., et al. 2013, , 49, 137 Fine, S., Croom, S. M., Hopkins, P. F., et al. 2008, MNRAS, 390, 1413 Gardner, J. P., Mather, J. C., Clampin, M., et al. 2006, , 123, 485 Gaskell, C. M. 2009, , 53, 140 Gnerucci, A., Marconi, A., Capetti, A., Axon, D. J., & Robinson, A. 2013, , 549, A139 Grier, C. J., Peterson, B. M., Horne, K., et al. 2013, , 764, 47 Hamann, F. 1997, , 109, 279 Joergens, V., Bonnefoy, M., Liu, Y., et al. 2013, , 558, L7 Kaspi, S., Maoz, D., Netzer, H., et al. 2005, , 629, 61 Kaspi, S., Brandt, W. N., Maoz, D., et al. 2007, , 659, 997 Korista, K., Baldwin, J., Ferland, G., & Verner, D. 1997, , 108, 401 Laor, A. 2003, , 590, 86 Laor, A., Barth, A. J., Ho, L. C., & Filippenko, A. V. 2006, , 636, 83 Lynden-Bell, D. 1969, , 223, 690 Maiolino, R., Salvati, M., Marconi, A., & Antonucci, R. R. J. 2001, , 375, 25 Maoz, D., Netzer, H., Mazeh, T., et al. 1991, , 367, 493 Mathews, W. G., & Wampler, E. J. 1985, , 97, 966 Netzer, H., & Laor, A. 1993, , 404, L51 Osterbrock, D. E. 1978, Proceedings of the National Academy of Science, 75, 540 Pancoast, A., Brewer, B. J., Treu, T., et al. 2013, arXiv:1311.6475 P[â]{}ris, I., Petitjean, P., Aubourg, [É]{}., et al. 2014, , 563, A54 Peterson, B. M. 1993, , 105, 247 Peterson, B. M., Ferrarese, L., Gilbert, K. M., et al. 2004, , 613, 682 Petrov, R. G., Millour, F., Lagarde, S., et al. 2012, , 8445, Pontoppidan, K. M., Blake, G. A., van Dishoeck, E. F., et al. 2008, , 684, 1323 Pontoppidan, K. M., Blake, G. A., & Smette, A. 2011, , 733, 84 Richards, G. T., et al. 2006, ApJS, 166, 470 (R06) Richards, G. T., Kruczek, N. E., Gallagher, S. C., et al. 2011, , 141, 167 Schneider, D. P., et al. 2010, AJ, 139, 2360 Schmidt, M., & Green, R. F. 1983, , 269, 352 Shen, Y., Greene, J. E., Strauss, M. A., Richards, G. T., & Schneider, D. P. 2008, , 680, 169 Shen, Y., et al. 2011, ApJS, 194, 45 Shen, Y. 2012, , 757, 152 Smith, J. E., Young, S., Robinson, A., et al. 2002, , 335, 773 Smith, J. E., Robinson, A., Young, S., Axon, D. J., & Corbett, E. A. 2005, , 359, 846 Stern, J., & Laor, A. 2012a, , 423, 600 Stern, J., & Laor, A. 2012b, , 426, 2703 Stern, J., & Laor, A. 2013, , 431, 836 Stern, J., Laor, A., & Baskin, A. 2014, , 438, 901 Strateva, I. V., Strauss, M. A., Hao, L., et al. 2003, , 126, 1720 Tremaine, S., Shen, Y., Liu, X., & Loeb, A. 2014, arXiv:1406.2468 Trippe, S., Davies, R., Eisenhauer, F., et al. 2010, , 402, 1126 Trump, J. R., Hall, P. B., Reichard, T. A., et al. 2006, , 165, 1 Vanden Berk, D. E., Richards, G. T., Bauer, A., et al. 2001, , 122, 549 Vanden Berk, D. E., Wilhite, B. C., Kron, R. G., et al. 2004, , 601, 692 V[é]{}ron-Cetty, M.-P., & V[é]{}ron, P. 2010, , 518, AA10 Veilleux, S., & Zheng, W. 1991, , 377, 89
-------------------------- ------ ------------------ ------ ------ ------------ ---------------------- ---------------
Object $z$ $L_{1450}^{(1)}$ Line Band EW$^{(2)}$ $\Phi^{(3)}$ $\fwhm^{(4)}$
$(\ergs)$ ($\AA$) $(\m^{-2} \hr^{-1})$ $(\kms)$
3C 273 0.16 46.3 K 120 $2.8 \cdot 10^{7}$ 3400$^{(5)}$
SDSS J163302.66+234928.5 0.82 46.9 J 560 $1.8 \cdot 10^{7}$ 5050$^{(6)}$
PG 1634+706 1.33 47.5 H 560 $2.7 \cdot 10^{7}$ 4100$^{(7)}$
SDSS J152156.48+520238.5 2.21 47.7 K 560 $1.6 \cdot 10^{7}$ 9350$^{(6)}$
SDSS J155152.46+191104.0 2.85 47.8 J 33 $1.7 \cdot 10^{6}$ 7730$^{(8)}$
HS 0857+4227 3.29 47.6 J 33 $7.9 \cdot 10^{5}$ 4430$^{(9)}$
PSS J1347+4956 4.51 47.6 H 33 $5.2 \cdot 10^{5}$ 8940$^{(9)}$
SDSS J001115.23+144601.8 4.97 47.5 H 33 $3.5 \cdot 10^{5}$ 5930$^{(10)}$
SDSS J114816.64+525150.3 6.42 47.1 K 33 $8.6 \cdot 10^{4}$ 5930$^{(10)}$
ULAS J112001.48+064124.3 7.08 46.6 K 33 $2.5 \cdot 10^{4}$ 4410$^{(11)}$
-------------------------- ------ ------------------ ------ ------ ------------ ---------------------- ---------------
[^1]: None of these quasars appears in the list of lensed quasars compiled by [@VeronCettyVeron10].
[^2]: Note that relativistic corrections of orders $O((\rg/{r_{\rm BLR}})^{1/2})$ and $O(\rg/{r_{\rm BLR}})$ to the line [*profile*]{} have been observed (CH89; @Tremaine+14).
[^3]: Equivalent to eqn. 4 in [@Pontoppidan+08].
[^4]: The value of ${\Phi^\ast}_v(r,\varphi')$ is equal to the integrand in eqn. 7 in CH89, divided by the observed photon energy.
[^5]: The typical value of $\sin i$ is expected to be $\approx0.5$ in the standard AGN unification model, where the broad line region is visibile only if the quasar is viewed at pole-on inclinations.
[^6]: @Baskin+14a plot $f(r)$ in terms of the line EW, assuming a CF per unit $\log\,r$ of $0.3$. We use their BLR model with solar metallicity and an ionizing spectral slope of $-1.6$.
[^7]: We use the [@Stern+14] NLR model with an ionizing slope of $-1.6$, solar metallicity, and $r=600\,L_{47}^{1/2}\pc$, which corresponds to $r=1000\,{r_{\rm BLR}}\approx{r_{\rm PSF}}$, where the NLR contribution to $S_v$ peaks.
[^8]: See user manual at http://www.eso.org .
[^9]: Instrument throughput based on NACO and MICADO (Multi-AO Imaging Camera for Deep Observations, @Davies+10) specifications.
[^10]: http://www.mpia-hd.mpg.de/ENIGMA/
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The measured branching ratio of the $D$ meson semileptonic decay $D \to \rho e^+ \nu_e$, which is based on the $0.82~{\rm fb^{-1}}$ CLEO data taken at the peak of $\psi(3770)$ resonance, disagrees with the traditional SVZ sum rules analysis by about three times. In the paper, we show that this discrepancy can be eliminated by applying the QCD light-cone sum rules (LCSR) approach to calculate the $D\to \rho$ transition form factors $A_{1,2}(q^2)$ and $V(q^2)$. After extrapolating the LCSR predictions of these TFFs to whole $q^2$-region, we obtain $1/|V_{\rm cd}|^2 \times \Gamma(D \to \rho e \nu_e) =(55.45^{+13.34}_{-9.41})\times 10^{-15}~{\rm GeV}$. Using the CKM matrix element and the $D^0(D^+)$ lifetime from the Particle Data Group, we obtain ${\cal B} (D^0\to \rho^- e^+ \nu_e) = (1.749^{+0.421}_{-0.297}\pm 0.006)\times 10^{-3}$ and ${\cal B} (D^+ \to \rho^0 e^+ \nu_e) = (2.217^{+0.534}_{-0.376}\pm 0.015)\times 10^{-3}$, which agree with the CLEO measurements within errors. We also calculate the branching ratios of the two $D$ meson radiative processes and obtain ${\cal B}(D^0\to \rho^0 \gamma)= (1.744^{+0.598}_{-0.704})\times 10^{-5}$ and ${\cal B}(D^+ \to \rho^+ \gamma) = (5.034^{+0.939}_{-0.958})\times 10^{-5}$, which also agree with the Belle measurements within errors. Thus we think the LCSR approach is applicable for dealing with the $D$ meson decays.'
address:
- 'Institute of Particle Physics $\&$ Department of Physics, Guizhou Minzu University, Guiyang 550025, P.R. China'
- 'State Key Laboratory of Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing, 100190, P.R. China'
- 'Department of Physics, Chongqing University, Chongqing 401331, P.R. China'
author:
- 'Hai-Bing Fu'
- Long Zeng
- Rong Lü
- Wei Cheng
- 'Xing-Gang Wu'
title: 'The $D\to \rho$ semileptonic and radiative decays within the light-cone sum rules'
---
Introduction
============
The semileptonic decays of the heavy meson, which contains heavy $c$ or $b$ quark, are important for studying the weak and strong interactions and for studying the heavy-flavor physics. In the charm factories nowadays, such as Belle, LHCb, BES and PANDA, the $D$ meson semileptonic decays provide a good platform for the precision test of standard model (SM) and for searching of new physics (NP) beyond the SM. For examples, the CLEO Collaboration present the first measurement of the branching fraction of $D^+\to \omega e^+\nu_e$ [@Coan:2005iu; @Huang:2005iv]. Lately, the CLEO Collaboration finished a more precise measurement on the $D^0 \to \rho^- e^+ \nu_e$ and $D^+ \to \rho^0 e^+ \nu_e$ decays based on $0.82~{\rm fb^{-1}}$ data taken at the peak of $\psi(3770)$ resonance. The CLEO Collaboration gave the branching fractions ${\cal B}(D^0 \to \rho^- e^+ \nu_e) = (1.77\pm0.12\pm0.10)\times 10^{-3}$ and ${\cal B}(D^+ \to \rho^0 e^+ \nu_e) = (2.17\pm0.12^{+0.12}_{-0.22})\times 10^{-3}$ [@CLEO:2011ab]. Recently, the BES-III Collaboration published their improved results, $\mathcal B (D^0 \to \rho^- e^+ \nu_e) = ( 1.445 \pm 0.058 \pm 0.039) \times 10^{-3}$ and ${\cal B}(D^+ \to \rho^0 e^+ \nu_e) = (1.860 \pm 0.070 \pm 0.061) \times 10^{-3}$, by using a data sample corresponding to an integrated luminosity of 2.93 fb$^{-1}$ at a center-of-mass energy of 3.773 GeV [@Ablikim:2018qzz].
Meanwhile, the nonleptonic decays of the charm meson such as $D^0\to\rho^0\gamma$ and $D^+\to\rho^+\gamma$ have also been a subject of high priority for many years and has resulted in a wealth of experimental data. In particular, two measured observables by the Belle Collaboration are, ${\cal B}(D^0\to\rho^0\gamma) = (1.77\pm0.30\pm0.07) \times 10^{-5}$ and $A_{CP}(D^0\to\rho^0\gamma) = 0.056\pm0.152\pm0.006$ [@Abdesselam:2016yvr], where the charge-parity (CP) asymmetry $A_{CP} $ is defined as $$A_{CP}(D\to\rho\gamma) = \frac{\Gamma(D\to \rho\gamma) - \Gamma(\bar D\to \bar \rho\gamma)}{\Gamma(D\to \rho\gamma) + \Gamma(\bar D\to \bar \rho\gamma)} .$$
The $D\to\rho$ transition form factors (TFFs) are key components of the $D$ meson semileptonic decay $D\to \rho e \nu_e $ and the $D$ meson radiative decay $D\to \rho \gamma$. In order to achieve more precise predictions on the $D$ meson semileptonic and radiative decays, it is important to have more accurate $D\to\rho$ TFFs. Theoretically, the $D\to \rho$ TFFs have been calculated under various approaches, such as the 3-point QCD sum rules (3PSR) [@Ball:1993tp], the heavy quark effective field theory (HQEFT) [@Wang:2002zba; @Wu:2006rd], the relativistic harmonic oscillator potential model (RHOPM) [@Wirbel:1985ji], the quark model (QM) [@Isgur:1988gb; @Melikhov:2000yu], the light-front quark model (LFQM) [@Verma:2011yw], the heavy meson and chiral symmetries (HM$\chi$T) [@Fajfer:2005ug], and the Lattice QCD [@Lubicz:1991bi; @Bernard:1991bz]. Most results are consistent with the CLEO measurements within errors; while the QCD sum rules (SR) leads to much smaller branching ratios, i.e. ${\cal B}(D^0\to \rho^- e^+ \nu_e) = (0.5\pm0.1)\times 10^{-3}$ [@Ball:1993tp]. One may question the applicability of the QCD SR approach for those TFFs. There are large uncertainties for the 3PSR prediction, which is however based on the conventional Shifman-Vainshtein-Zakharov SR approach [@Shifman:1978bx] and the approach itself has many defaults in dealing with such kind of TFFs [@Braun:1997kw].
The QCD SR approach is applicable in both small and intermediate $q^2$-region, which provides a bridge between the pQCD and lattice QCD predictions, and then a more accurate QCD SR prediction is helpful for a better understanding of the TFFs. In the paper, as a new QCD SR analysis, we take the improved version of QCD SR approach, i.e. the light-cone sum rules (LCSR) [@Balitsky:1989ry; @Chernyak:1990ag; @Ball:1991bs], to recalculate the $D\to\rho$ TFFs. The LCSR avoids the problems of the conventional SVZ SR by making a partial resummation of the operator product expansion (OPE) to all orders and reorganize the OPE expansion in terms of the twists of relevant operators rather than their dimensions. The vacuum condensates of the SVZ SR are then substituted by the light-meson¡¯s light-cone distribution amplitudes (LCDAs) of increasing twists. The LCDA, which relates the matrix elements of the nonlocal light-ray operators sandwiched between the hadronic state and the vacuum, has a direct physical significance and provides the underlying links between the hadronic phenomena at small and large distances. Since its invention, the LCSR approach has been widely adopted for studying the heavy $\to$ light meson decays.
The remaining parts of the paper are organized as follows. In Sec. \[section:2\], we present the calculation technology for the $D\to \rho$ semileptonic and radiative decays, and their key components, e.g. the $D\to \rho$ TFFs, within the LCSR approach. In Sec. \[section:3\], we present our numerical results and discussions on the $D\to \rho$ TFFs, the $D\to \rho$ semileptonic decay widthes and branching ratios for two different channels, and the branching ratio and the direct CP asymmetry of the $D\to\rho\gamma$ decay. Section \[section:4\] is reserved for a summary.
Calculation technology {#section:2}
======================
The $D\to\rho e^+ \nu_e$ semileptonic decay
-------------------------------------------
![Typical diagram for the $D\to\rho e^+ \nu_e$ semileptonic decay.[]{data-label="semidecay"}](Fig3 "fig:"){width="35.00000%"}\
We present the typical diagram for the $D\to\rho e^+ \nu_e$ semileptonic decay in Fig.\[semidecay\]. The longitudinal and transverse helicity decay widths of $D\to \rho e^+ \nu_e$ can be expressed in terms of three helicity amplitudes $H_{0,\pm}(q^2)$: $$\begin{aligned}
\Gamma^a &=&\frac{G_F^2 |K_{\rm HV}|^2 }{192\pi^3 m_D^3}\int_{m_e^2}^{q^2_{\rm max}}q^2\sqrt{\lambda(q^2)} |H_a(q^2)|^2, \label{Eq:GammaLT0}\end{aligned}$$ where $a=+$, $-$, $0$, $q^2_{\rm max} = (m_D-m_\rho)^2$, and $\lambda(q^2)$ is the phase-space factor, which equals to $(m_D^2 + m_\rho^2 - q^2)^2-4 m_D^2 m_\rho^2$. The Fermi constant $G_F=1.166\times10^{-5}\;{\rm GeV}^{-2}$. The constant $K_{\rm HV}$ parameterizes the quark flavor mixing relevant to a particular transition, which equals to $|V_{\rm cd}|$ for $D^0 \to \rho^- e^+ \nu_e$ and $|V_{\rm cd}|/{\sqrt 2}$ for $D^+ \to \rho^0 e^+ \nu_e$. Then the total decay width of $D^{+/0}\to \rho^{0/-} e^+ \nu_e$ can be written as $$\Gamma=\Gamma^{\rm L} + \Gamma^{\rm T}, \label{Eq:Gamma}$$ where $\Gamma^{\rm T} = \Gamma^+ + \Gamma^-$ and $\Gamma^{\rm L}=\Gamma^0$.
Here we have adopted the helicity basis to express the decay width. In the helicity basis, the TFF $H_a(q^2)$ corresponds to a transition amplitude with definite spin-parity quantum number in the lepton pair center-of-mass frame. The transverse and longitudinal helicity TFF $H_{a}(q^2)$ can be calculated by relating them to the TFFs $A_{1,2}(q^2)$ and $V(q^2)$ via the following way $$\begin{aligned}
H_\pm(q^2) &=& (m_D+m_\rho) A_1(q^2)\mp\frac{\sqrt{\lambda(q^2)}} {m_D + m_\rho}V(q^2) \label{Hpm}\end{aligned}$$ and $$\begin{aligned}
H_0(q^2) &=& \frac{1}{2m_\rho\sqrt{q^2}}\bigg\{(m_D^2-m_\rho^2-q^2) (m_D+m_\rho) A_1(q^2)
\nonumber\\
&& -\frac{\lambda(q^2)}{m_D+m_\rho}A_2(q^2)\bigg\}. \label{H0}\end{aligned}$$ The TFFs $A_{1,2}(q^2)$ and $V(q^2)$ can be defined by relating them to the $D\to\rho$ matrix elements, e.g. $$\begin{aligned}
&&\langle \rho (\tilde p,\tilde \epsilon )|\bar q \gamma_\mu c|D(p)\rangle = \epsilon_{\mu\nu\alpha\beta}\tilde \epsilon^{*\nu} p^\alpha \tilde p^\beta \frac{2V(q^2)}{m_D + m_\rho }, \label{Drho:matrix1} \\
&&\langle \rho (\tilde p,\tilde \epsilon )|\bar q{\gamma _\mu } {\gamma _5}c|D(p)\rangle = i \tilde \epsilon^*_\mu (m_D + m_\rho )A_1(q^2)
\nonumber\\
&&\qquad - i(\tilde \epsilon^* \cdot p )(p + \tilde p)_\mu\frac{1}{m_D + m_\rho}A_2(q^2)
\nonumber\\
&&\qquad - i(\tilde \epsilon^* \cdot p )(p - \tilde p)_\mu \frac{2 m_\rho } {q^2}[A_3(q^2) - A_0(q^2)]. \label{Drho:matrix2}\end{aligned}$$ where $\tilde \epsilon$ is $\rho$-meson polarization vector and $q$ is the four-momentum transfer between the two mesons.
The $D\to \rho\gamma$ decay
---------------------------
We give a mini-review of the Cabibbo-suppressed $D\to\rho\gamma$ decay in this subsection, which is an important channel for testing the SM and for searching of new physics beyond the SM.
The amplitude of $D\to \rho\gamma$ decay can be decomposed into the following gauge invariant form, $$\begin{aligned}
&&{\cal A}[D(p)\to\rho(\tilde p, \tilde \epsilon)\gamma(q, \epsilon)]= \epsilon_{\mu\nu\alpha\beta} q^{\mu} \epsilon^{*\nu} p^{\alpha}\tilde\epsilon^{*\beta} {\cal A}_{\rm PC}^\rho
\nonumber\\
&& \qquad + i [(\tilde\epsilon^*\cdot q)(\epsilon\cdot\tilde p)-(\tilde p\cdot q)(\tilde \epsilon^*\cdot \epsilon^*)] {\cal A}_{\rm PV}^\rho ,\end{aligned}$$ where $\tilde\epsilon$ and $\epsilon$ are polarization vectors of $\rho$ and $\gamma$, respectively. ${\cal A}_{\rm PC}^\rho $ and ${\cal A}_{\rm PV}^\rho $ are parity conserving and parity violating amplitudes which can be calculated by using the effective $c\to u\gamma$ weak Lagrangian [@Burdman:1995te; @Greub:1996wn; @Fajfer:1997bh; @Isidori:2012yx]. Then the decay rate of $D\to \rho\gamma$ can be written as [@deBoer:2015boa; @deBoer:2017que] $$\begin{aligned}
\Gamma(D\to \rho\gamma) = \frac{m_D^3}{32 \pi} \bigg(1-\frac{m_\rho^2}{m_D^2}\bigg)^3 \big[|{\cal A}_{\rm PV}^\rho |^2 + |{\cal A}_{\rm PC}^\rho |^2 \big].\end{aligned}$$
Our remaining task is to calculate the amplitudes ${\cal A}_{\rm PV, PC}^\rho$, which contain both the long-distance contribution and the short-distance contribution, e.g. $${\cal A}_{\rm PV, PC}^\rho=({\cal A}_{\rm PV, PC}^\rho)^{\rm l.d.}+ ({\cal A}_{\rm PV, PC}^\rho)^{\rm s.d.}. \label{Eq:APCV_ld}$$
![The typical VME diagram via the penguin-like $c\to u\gamma$ transition.[]{data-label="Fig:Fig1"}](Fig1 "fig:"){width="35.00000%"}\
{width="90.00000%"}\
Firstly, the long-distance contributions for radiative decays $D\to\rho\gamma$ can be treated as originating from the nonleptonic transition $D\to \rho V_0$, followed by the conversion $V_0\to \gamma$ via the vector meson dominance (VME) mechanism [@Fajfer:1998dv], whose typical Feynman diagram via the penguin-like $c\to u\gamma$ transition is shown in Fig.\[Fig:Fig1\]. More precisely, the long-distance amplitudes $({\cal A}_{\rm PV, PC}^\rho)^{\rm l.d.}$ can be divided into three parts, e.g. $$({\cal A}_{\rm PV, PC}^\rho)^{\rm l.d.} = {\cal A}_{\rm PV, PC}^{\rm I} +{\cal A}_{\rm PV, PC}^{\rm II} +{\cal A}_{\rm PV, PC}^{\rm III}$$ with $8$ schematic diagrams to be calculated, are the hybrid model based on the heavy quark effective theory and chiral perturbation theory. The rectangle in each diagram of Fig.\[Fig:Fig2\] respects the weak transition due to the effective Lagrangian which can be found in Ref.[@Fajfer:1998dv]. The Lagrangian contains a product of two left-handed quark currents, which will be expressed in terms of the relevant hadronic degrees of freedom, $D$, $D^*$, $P$ and $V$, as described in Fig.\[Fig:Fig2\]:
- The first part ${\cal A}_{\rm PC}^{\rm I}$, which is represented by Fig.\[Fig:Fig2\](a) and Fig.\[Fig:Fig2\](b), denotes the photon emitted from the $D$ meson becomes $D^*$ meson and then weakly decays into $\rho$ meson.
- The second part ${\cal A}_{\rm PC}^{\rm II}$, which is shown by Fig.\[Fig:Fig2\](c), comes from the weak decay of $D$ meson, which firstly decays into an off-shell light pseudoscalar and then decays into $\rho\gamma$.
- As shown by Fig.\[Fig:Fig2\](d) and Fig.\[Fig:Fig2\](e), there are also long-distance penguin-like contributions, which contribute to both the parity-conserving amplitude ${\cal A}_{\rm PC}^{\rm III}$ and the parity-violating amplitude ${\cal A}_{\rm PV}^{\rm III}$. They will be vanished in the exact SU(3) flavour limit.
- Fig.\[Fig:Fig2\](f) and Fig.\[Fig:Fig2\](g) are bremsstrahlung-like diagrams, where the photon emission is due to the direct coupling to charged initial $D$ or final-state $\rho$-meson, both of which contribute to the parity violating amplitude ${\cal A}_{\rm PV}^{\rm I}$.
- Fig.\[Fig:Fig2\](h) denotes the remaining parity violating contribution from other quark-level picture [@Fajfer:1998dv], whose amplitude is represented as ${\cal A}_{\rm PV}^{\rm II}$.
It has been found that Fig.\[Fig:Fig2\](c) has sizable contribution, while Figs.\[Fig:Fig2\](d) and (e) have negligible contributions which are less than $1\%$. To derive the amplitudes ${\cal A}_{\rm PV, PC}^{\rm I, II, III}$, we need to know the coupling strengthes for the interaction of the light vector meson with $D$ or $D^*$ and the heavy quark-photon interaction, which can be characterized by the parameters $\lambda$ and $\lambda'$, respectively. To do our numerical calculation, we shall adopt $\lambda= 0.479 ~{\rm GeV}^{-1}$ [@Fajfer:1998dv], and $\lambda' = 1/(6m_c)$ which is predicted by approximately relating it to the charm quark magnetic moment [@Isgur:1989vq; @Yan:1992gz; @Cho:1992nt; @Amundson:1992yp].
Following the above discussions, the parity-conserving amplitudes for $D^+ \to \rho^+ \gamma$ are $$\begin{aligned}
{\cal A}_{\rm PC}^{\rm I} (D^+ \to \rho^+ \gamma) &=& 4a_1 {\cal G} \bigg| \lambda' - \lambda \frac{{\tilde g}_V}{2{\sqrt 2}} \left(\frac{f_\rho}{m_\rho}
- \frac{f_\omega}{3 m_\omega}\right) \bigg|
\nonumber\\
&&\times \frac{m_\rho m_{D^*} f_D f_\rho}{m_{D^*}^2 - m_\rho^2}
{\sqrt \frac{m_{D^*}}{m_D}},
\nonumber\\
\nonumber\\
{\cal A}_{\rm PC}^{\rm II} (D^+ \to \rho^+ \gamma) &=& 4a_1 {\cal G}
\frac{m_{D}^2 f_D f_\omega }{3 m_\omega (m_{D}^2 - m_{\pi}^2)} |C_{VV\Pi}|,
\nonumber\\
\nonumber\\
{\cal A}_{\rm PC}^{\rm III} (D^+ \to \rho^+ \gamma) &=& a_2 {\cal G}
\left(-f_\rho^2 + \frac1{3}f_\omega^2 - \frac{V_{cs}^*V_{us}}{V_{cd}^*V_{ud}}\frac{2}{3}f_\phi^2\right)
\nonumber\\
&&\times \frac{|V(0)|}{m_D + m_\rho}
\label{apcdpcs}\end{aligned}$$ and the parity-violating amplitudes for $D^+ \to \rho^+ \gamma$ are $$\begin{aligned}
&& {\cal A}_{\rm PV}^{\rm I} (D^+ \to \rho^+ \gamma) = 2a_1{\cal G}\frac{m_\rho f_{\rho} f_D }{m_D^2 - m_{\rho}^2},
\nonumber\\
\nonumber\\
&& {\cal A}_{\rm PV}^{\rm II} (D^+ \to \rho^+ \gamma) = - a_1{\cal G} \frac{f_\rho}{m_D^2 - m_\rho^2} \bigg[f_\rho (m_D + m_\rho)
\nonumber\\
&&\qquad\qquad \times A_1(m_\rho^2) - f_\omega \frac{m_\rho(m_D + m_\omega)}{3m_\omega} A_1^{D\omega}(m_\rho^2)\bigg],
\nonumber\\
\nonumber\\
&& {\cal A}_{\rm PV}^{\rm III} (D^+ \to \rho^+ \gamma) = a_2 {\cal G} \left(-f_\rho^2 + \frac1{3}f_\omega^2 - \frac{V_{cs}^*V_{us}}{V_{cd}^*V_{ud}}\frac{2}{3}f_\phi^2\right)
\nonumber\\
&&\qquad\qquad \times \frac{A_1(0)}{m_D - m_\rho},
\label{apvdpcs}\end{aligned}$$ where ${\cal G} = e G_F V_{ud} V_{cd}^*/\sqrt 2$ with $e = \sqrt{4\pi \alpha_{\rm em}}$, $\tilde{g}_{V} = 5.9$ which can be fixed by the flavor symmetry, $|C_{VV\Pi}| = 3 \tilde{g}_{V}^2/(32\pi^2) = 0.33$ [@Fajfer:1997bh], $a_1 = 1.26\pm0.1$ and $a_2 = -0.55\pm0.1$ [@Bauer:1986bm], $f_{\rho,\omega,\phi}$ are decay constants which can be fixed by the leptonic decays of these mesons. $A_1^{D\omega}(m_\rho^2)$ is the TFF for $D\to\omega$ [@Fajfer:1997bh].
The Cabibbo suppressed decay $D^0 \to \rho^0 \gamma$ involves the contribution from the $\eta - \eta^{\prime}$ mixing, whose parity-conserving amplitudes are $$\begin{aligned}
{\cal A}_{\rm PC}^{\rm I} (D^0 \to \rho^0 \gamma) &=& 4 a_2 b_\rho^0 {\cal G}
\bigg| \lambda^{\prime} +\lambda \frac{{\tilde g}_V}{2{\sqrt 2}} \left(\frac{f_\rho}{m_\rho} + \frac{f_{\omega}}{3 m_{\omega}}\right) \bigg|
\nonumber\\
&&\times \frac{m_\rho m_{D^*} f_D f_\rho }{m_{D^*}^2 - m_\rho^2} {\sqrt \frac{m_{D^*}}{m_D}},
\nonumber\\
\nonumber\\
{\cal A}_{\rm PC}^{\rm II} (D^0 \to \rho^0 \gamma) &=& 4a_2{\cal G} {\cal B}^\rho m_{D}^2 f_D | C_{VV\Pi} |,
\nonumber\\
\nonumber\\
{\cal A}_{\rm PC}^{\rm III} (D^0 \to \rho^0 \gamma) &=& a_2{\cal G} \left( -f_\rho^2 + \frac1{3}f_\omega^2 - \frac{V_{cs}^*V_{us}}{V_{cd}^*V_{ud}}\frac{2}{3}f_\phi^2 \right)
\nonumber\\
&&\times\frac{V(0)}{m_D + m_\rho},
\label{apcd0cs}\end{aligned}$$ where $b_{\rho}^0 = - 1/{\sqrt 2}$ and the coefficients $${\cal B}^\rho= \sum_{i=1}^{3} \frac{B^{\rho}_{P_i}}{m_D^2 - m_{P_i}^2}, \label{bv}$$ where $P_i$ stands for the light mesons $\pi$, $\eta$, and $\eta'$, respectively, and $B^\rho_\pi = f_\omega/(3 \sqrt 2 m_\omega)$, $B^{\rho}_\eta = -c(c - \sqrt 2 s)f_\rho /(\sqrt 2 m_\rho)$, $B^{\rho}_{\eta'} = - s ({\sqrt 2} c + s) f_{\rho} /(\sqrt 2 m_{\rho})$ with $s=\sin \theta$ and $c=\cos\theta$. Here $\theta$ is the $\eta-\eta'$ mixing angle and we set its value as $\theta = -20^{\rm o}$, which is consistent with the values derived in Refs.[@Tanabashi:2018oca; @Huang:2006as; @Klempt:2007cp; @Aaij:2014jna]. The parity-violating amplitudes for $D^0 \to \rho^0 \gamma$ are $$\begin{aligned}
&& {\cal A}_{\rm PV}^{\rm I} (D^0 \to \rho^0 \gamma) = 0,
\nonumber\\
\nonumber\\
&& {\cal A}_{\rm PV}^{\rm II} (D^0 \to \rho^0 \gamma)= - a_2{\cal G} \frac{f_\rho}{\sqrt 2( m_D^2 - m_\rho^2)} \bigg[f_\rho (m_D + m_\rho)
\nonumber\\
&&\qquad\qquad \times A_1(m_\rho^2) + f_\omega \frac{m_\rho(m_D + m_\omega)}{3m_\omega} A_1^{D\omega}(m_\rho^2)\bigg],
\nonumber\\
\nonumber\\
&& {\cal A}_{\rm PV}^{\rm III} (D^0 \to \rho^0 \gamma)= a_2 {\cal G} \left(-f_\rho^2 + \frac1{3}f_\omega^2 - \frac{V_{cs}^*V_{us}}{V_{cd}^*V_{ud}}\frac{2}{3}f_\phi^2\right)
\nonumber\\
&&\qquad \qquad \times \frac{A_1(0)}{m_D - m_\rho}.
\label{apvdp0s}\end{aligned}$$
Secondly, we need to deal with the short-distance contribution $({\cal A}_{\rm PC(PV)}^\rho)^{\rm s.d.}$. There are ten operators which enter into the weak interactions of the $D\to\rho$ decay [@Burdman:1995te; @Greub:1996wn; @Fajfer:1997bh; @Isidori:2012yx]. It has been observed that due to the GIM suppression and also the small magnitudes of the Wilson coefficients, only the operators ${\cal Q}_7$ and ${\cal Q}'_7$ have sizable contributions to the $D\to\rho$ decay; Thus the short-distance amplitudes $({\cal A}_{\rm PC(PV)}^\rho)^{\rm s.d.}$ have the following form [@Isidori:2012yx; @deBoer:2017que] $$\begin{aligned}
({\cal A}_{\rm PC(PV)}^\rho)^{\rm s.d.} = \frac{\sqrt{4\pi\alpha_e} Q_u G_F m_c}{2\sqrt 2\pi^2} (A_7 + A'_7)T(0) ,\end{aligned}$$ in which $T(0) = T_{1}(0)= T_{2}(0)$ and the tensor TFFs $T_{1,2}(q^2)$ are defined by $$\begin{aligned}
&&\langle \rho(\tilde p,\tilde \epsilon)|\bar u q^\nu \sigma_{\mu\nu}(1+\gamma_5)c|D(p)\rangle
\nonumber\\
\nonumber\\
&&\quad = -2i\epsilon_{\mu\nu\alpha\beta} \tilde \epsilon^{*\nu} p^\alpha \tilde p^\beta T_1(q^2)
\nonumber\\
\nonumber\\
&&\quad + \left[(m_D^2-m_\rho^2)\tilde\epsilon^*_\mu - (\tilde\epsilon^* \cdot p) (p + \tilde p)_\mu\right] T_2(q^2)
\nonumber\\
&&\quad + (\tilde \epsilon^* \cdot p )\left[(p - \tilde p)_\mu - (p + \tilde p)_\mu \frac{q^2}{m_D^2 - m_\rho^2}\right] T_3(q^2). \nonumber\\\end{aligned}$$ Those two tensor TFFs $T_{1,2}(q^2)$ can be calculated under the LCSR approach. For the needed $T_1$, by using the heavy quark effective field theory, it can be related to the TFFs $V$ and $A_1$ via the following way [@Wu:2006rd] $$T_1 (q^2) = \frac{m_D^2 - m_\rho^2 + q^2}{2m_D(m_D + m_\rho)}V(q^2) + \frac{m_D + m_\rho}{2m_D}A_1(q^2).$$ Furthermore, the coefficient $A_7^{(\prime)} = C_7^{(\prime){\rm eff}}+\cdots$, where the ellipse stands for the additional contribution from within and outside the SM.
As a final remark, for the Cabibbo suppressed decay $D^0 \to \rho^0 \gamma$, in the limit where the strong phases of the amplitudes have a mild energy dependence, and assuming that we can neglect the weak phase of the long-distance amplitudes, and the $CP$ asymmetry $|A_{CP}|$ is primarily sensitive to direct $CP$ asymmetry $|a_{\rho\gamma}^{\rm dir}|$, i.e. $$\begin{aligned}
|A_{CP}| \approx |a_{\rho\gamma}^{\rm dir}| = 2\frac{|{\rm Im}({\cal A}_{\rm PC, PV}^\rho)^{\rm s.d.}|}{|({\cal A}_{\rm PC, PV}^\rho)^{\rm l.d.}|}\times |\sin(\Delta\phi_{\rm strong})|. \nonumber \\
\label{Eq:CP_asymmetry}\end{aligned}$$
The $D\to \rho$ TFFs
--------------------
twist-2 twist-3 twist-4
---------------- ---------------------- -------------------------------------------------------------------------------------- ------------------------------------------------------------------------------------------------
$\delta^0$ $\phi_{2;\rho}^\bot$
$\delta^1$ $\phi_{2;\rho}^\|$ $\phi_{3;\rho}^\bot, \psi_{3;\rho}^\bot, \Phi_{3;\rho}^\|,\tilde\Phi_{3;\rho }^\bot$
$\delta^2$ $\phi_{3;\rho}^\|, \psi_{3;\rho}^\|, \Phi_{3;\rho}^\bot$ $ \phi_{4;\rho}^\bot ,\psi_{4;\rho}^\bot,\Psi_{4;\rho}^\bot, \widetilde{\Psi} _{4;\rho}^\bot$
$\delta^3$ $\phi_{4;\rho}^\|,\psi_{4;\rho}^\|$
: The $\rho$-meson LCDAs with different twist-structures up to $\delta^3$ [@Ball:2004rg], where $\delta \simeq m_\rho/m_c$.[]{data-label="DA_delta"}
As mentioned in the Introduction, we shall adopt the LCSR approach to calculate the $D\to\rho$ TFFs. The LCSR approach is based on the operator production expansion near the light cone, and in different to the traditional QCD SR approach which parameterizes all the non-perturbative dynamics into vacuum condensates, the LCSR approach parameterizes those non-perturbative dynamics into LCDAs with increasing twists. Due to the complex structures of the $\rho$-meson LCDAs, it is convenient to arrange them by the parameter $\delta \simeq m_\rho/m_c\sim 52\%$ [@Ball:2004rg; @Ball:1998sk]. A collection of the $\rho$-meson twist-2, twist-3 and twist-4 LCDAs up to $\delta^3$-order are shown in Table \[DA\_delta\].
Up to twist-4 level, there are totally fifteen $\rho$-meson LCDAs, all of which, especially the high-twist DAs, are far from affirmation. As a tricky point of the LCSR approach, one may choose proper current for the correlation function (correlator) so as to suppress less certain high-twist terms and improve the accuracy of the LCSR prediction [@Huang:1998gp; @Huang:2001xb; @Wan:2002hz; @Zuo:2006dk; @Wu:2007vi; @Wu:2009kq]. For example, one can adopt a right-handed current to do the LCSR calculation, i.e., starting from the following chiral correlator $$\begin{aligned}
\Pi_\mu(\tilde p,q) &=& i\int d^4x e^{iq\cdot x} \nonumber\\
&\times& \langle\rho (\tilde p,\tilde\epsilon)|{\rm T} \big\{\bar q_1(x)\gamma_\mu(1-\gamma_5)c(x), j_D^\dag (0)\big\} |0\rangle, \nonumber\\ \label{correlator}\end{aligned}$$ where the current $j_D^\dag (x)=i \bar c(x)(1 + \gamma_5)q_2(x)$. This chiral correlator highlights the contributions from the chiral-odd LCDAs $\phi_{2;\rho}^\bot$ at the $\delta^0$-order, $\phi_{3;\rho}^\|$, $\psi_{3;\rho}^\|$, $\Phi_{3;\rho}^\bot$, $\phi_{4;\rho}^\bot$, $\psi_{4;\rho}^\bot$, $\Psi_{4;\rho}^\bot$, $\widetilde{\Psi} _{4;\rho}^\bot$ at the $\delta^2$-order; while all the contributions from the chiral-even $\rho$-meson LCDAs are negligible.
Following the standard LCSR procedures, we can obtain the LCSRs for the $D \to \rho$ TFFs $A_{1,2}(q^2)$ and $V(q^2)$ by using the above chiral correlator, which are similar to the $B\to \rho$ TFFs. The $B\to \rho$ TFFs have calculated by various groups under the LCSR approach, and a recent work is done by applying the vacuum-to-$B$-meson correlation function with an interpolating current for the light vector meson [@Gao:2019lta]. The lengthy analytic expressions for the $D \to \rho$ TFFs with the help of the present choice of a chiral correlator can be obtained from Ref.[@Fu:2014pba] by replacing the $B$-meson inputs there as the present $D$ meson ones. To short the length of the paper, we shall not list those formulas here and the interesting reader may turn to Ref.[@Fu:2014pba] for detail.
Numerically, we observe that the leading-twist terms are dominant for the chiral LCSRs of the TFFs, agreeing well with the usual $\delta$-power counting rule. Thus, those TFFs shall provide us a useful platform for testing the properties of the leading-twist $\phi_{2;\rho}^\bot$ via comparisons with the data or predictions from other theoretical approaches, such as those of Refs.[@Ball:2007zt; @Choi:2007yu; @Forshaw:2012im; @Xu:2018mpf].
More over, it is convenient to define two ratios over the three TFFs $A_{1,2}(q^2)$ and $V(q^2)$, $$\begin{aligned}
r_V = \frac{V(0)}{A_1(0)} \qquad {\rm and}\qquad r_2=\frac{A_2(0)}{A_1(0)}.\end{aligned}$$ Those two ratios could suppress the theoretical errors for each TFF within the LCSR approach, and also suppress the differences of the predictions on the TFFs from various approaches. Thus a better comparison with the data can be achieved.
Numerical results and discussion {#section:3}
================================
Distribiton amplitudes and TFFs
-------------------------------
To do the numerical calculation, we take the decay constant $f_\rho^\bot = 0.165(9) {\rm GeV}$ [@Ball:2007zt]. The $\rho$ and $D$ meson masses are taken as $m_\rho=0.775~{\rm GeV}$ and $m_D = 1.864$ GeV [@Tanabashi:2018oca]. The Cabibbo-Kobayashi-Maskawa matrix element $|V_{\rm cd}| = 0.2252 \pm 0.0007$ [@Tanabashi:2018oca], and the $D$ meson decay constant $f_D$ shall be determined by using the QCD sum rules approach [@Fu:2013wqa]. We adopt the WH model [@Wu:2010zc] as the $\rho$-meson transverse leading twist wavefunction, whose radial part is from the BHL-prescription [@BHL] and the spin-space wavefunction $\chi_\rho^{h_1 h_2} (x,{\bf k}_\bot)$ is from Wigner-Melosh rotation. And then, after integrating out the transverse moment dependence, we obtain the $\rho$-meson LCDA $$\begin{aligned}
&& \phi _{2;\rho }^\bot (x,\mu) = \frac{{A_{2;\rho}^\bot \sqrt {3x\bar x} {m_q}}}{{8{\pi ^{\frac{3}{2}}}\widetilde f_\rho^\bot b_{2;\rho}^\bot }}[1 + {B_{2;\rho}^\bot }C_2^{\frac{3}{2}}(\xi )]\nonumber\\
&& \quad \times \left[ {{\rm{Erf}}\left( {b_{2;\rho}^\bot \sqrt {\frac{{{\mu^2} + m_q^2}}{{x\bar x}}} } \right) - {\rm{Erf}}\left( {b_{2;\rho}^\bot \sqrt {\frac{{m_q^2}}{{x\bar x}}} } \right)} \right], \nonumber\\ \label{Eq:DAWH}\end{aligned}$$ where $\widetilde{f}_\rho^\bot = f_\rho^{\bot}/\sqrt{3}$ and the error function, $\textrm{Erf}(x) = 2 \int^x_0 e^{-t^2} dt/ \sqrt \pi$. The constitute light-quark mass is taken as $m_q \simeq 300~{\rm MeV}$. The two parameters $A_{2;\rho}^\bot$ and $b_{2;\rho}^\bot$ can be fixed by the normalization condition and the average value of the squared transverse momentum $\langle {\bf k}_\bot^2 \rangle_{2;\rho}^{1/2} = 0.37 \pm 0.02 ~{\rm GeV}$ [@Fu:2014pba]. The parameter $B_{2;\rho}^\bot$ can be fixed by using the second Gegenbauer moment, i.e. $a_{2;\rho}^\bot(\mu_0 = 1~{\rm GeV}) = 0.14(6)$ [@Ball:2007zt].
Using those input values, we obtain $$\begin{aligned}
&& A_{2;\rho}^{\bot,{\rm C}} = 23.808,~b_{2;\rho}^{\bot,{\rm C}}=0.572,~B_{2;\rho}^{\bot,{\rm C}} = 0.100; \\
&& A_{2;\rho}^{\bot,{\rm U}} = 22.679,~b_{2;\rho}^{\bot,{\rm U}}=0.555,~B_{2;\rho}^{\bot,{\rm U}} = 0.151; \\
&& A_{2;\rho}^{\bot,{\rm D}} = 25.212,~b_{2;\rho}^{\bot,{\rm D}}=0.595,~B_{2;\rho}^{\bot,{\rm D}} = 0.050,\end{aligned}$$ where $C$, $U$ and $D$ stand for center, upper and lower values, respectively.
![The $\rho$-meson leading-twist LCDA $\phi_{2;\rho}^\bot(x,\mu_0 = 1~{\rm GeV})$ predicted from the WH model. As a comparison, the BB prediction [@Ball:2007zt], the Linear PM and HOPM [@Choi:2007yu], the AdS/QCD prediction [@Forshaw:2012im], and the asymptotic form have also been presented. []{data-label="Fig:DA"}](DA){width="45.00000%"}
We present the $\rho$ meson transverse twist-2 LCDA $\phi_{2;\rho}^\bot(x,\mu_0 = 1~{\rm GeV})$ in Fig. \[Fig:DA\]. As a comparison, the Ball and Braun (BB) model [@Ball:2007zt], the Linear potential model (PM) and the harmonic oscillator potential model (HOPM) [@Choi:2007yu], the AdS/QCD model [@Forshaw:2012im], and the asymptotic model have also been presented. Fig. \[Fig:DA\] shows that the shape of $\phi_{2;\rho}^\bot$ is still unfixed, which varies from a single peaked behavior to a double peaked behavior. The $\phi_{2;\rho}^\bot$ shape is primarily controlled by the magnitude of $B_{2;\rho}^\bot$ or equivalently the second Gegenbauer moment $a_{2;\rho}^\bot$. Thus the WH model provides a convenient way to fix the behavior of $\phi_{2;\rho}^\bot$ via comparing with the data.
To set the Borel windows for the LCSRs of the $D \to \rho$ TFFs, we adopt the following criteria
- We require the continuum contribution to be less than $30\%$ of the total LCSR.
- We require all the high-twist LCDAs’ contributions to be less than $15\%$ of the total LCSR.
- The derivatives of LCSRs for TFFs with respect to $(-1/M^2)$ give three LCSRs for the $D$ meson mass $m_{D}$. We require the predicted $D$ meson mass to be fulfilled in comparing with the experiment one, e.g. $|m^{\rm th}_{D}-m^{\rm exp}_{D}|/m^{\rm exp}_{D}$ less than $0.1\%$.
We take the continuum thresholds for $D\to\rho$ TFFs $A_{1,2}(q^2)$ and $V(q^2)$ as $s_0(A_1) = 6.1(3)~{\rm GeV}^2$, $s_0(A_2) = 7.1(3)~{\rm GeV}^2$ and $ s_0(V) = 6.6(3)~{\rm GeV}^2$, which are close to the squared mass of the $D$ meson’s first excited state $D_1(2420)$. Numerically, we observe that the TFFs change slightly with $s_0$, thus the uncertainties caused by different choices $s_0$ are small [^1].
![The determined Borel windows for the TFFs at the large recoil point, $A_{1,2}(0)$ and $V(0)$. []{data-label="Fig:FM2"}](FM2){width="45.00000%"}
Following those criteria, the determined Borel windows are $M^2= 4.5(3) {\rm GeV}^2$ for $A_1$, $M^2 = 6.2(3) {\rm GeV}^2$ for $A_2$, and $M^2 = 5.0(3) {\rm GeV}^2$ for $V$, respectively. More explicitly, we present the Borel windows for those TFFs at the large recoil point $q^2=0$ in Fig. \[Fig:FM2\], which shows that the TFFs change slightly within the Borel windows, being consistent with conventionally adopted qualitative criterion that the TFF should be flat within the Borel window (since the physical observable should not depend on this artificial parameter). Contributions for different LCDAs for those TFFs are presented in Table \[Tab:TFF01\]. It shows that the $\delta^0$-order twist-2 LCDA $\phi_{2;\rho}^\bot$ provides dominate contribution, while the contributions from the $\delta^2$-order high-twist LCDAs are small.
$A_1(0)$ $A_2(0)$ $V(0)$
---------------------- ------------------- ----------------- --------------
$\phi_{2;\rho}^\bot$ 0.539 1.067 0.989
$\psi_{3;\rho}^\| $ 0.089 $-0.187$ /
$\phi_{4;\rho}^\bot$ $-0.011$ $-0.157$ $-0.174$
$I_L$ $-0.016$ $-0.091$ /
$H_3$ $-0.021$ $-0.170$ /
$\Phi_{3;\rho}^\bot$ / $0.0003$ /
Total 0.580 0.468 0.815
: Contributions from the LCDAs with various twist structures for the $D\to\rho$ TFFs $A_{1,2}(0)$ and $V(0)$, in which only the main non-zero contributions are listed.[]{data-label="Tab:TFF01"}
$A_1(0)$ $A_2(0)$ $V(0)$
---------------------------- ------------------------------ ------------------------------ ---------------------------
This work $0.580^{+0.065}_{-0.050}$ $0.468^{+0.052}_{-0.053}$ $0.815^{+0.070}_{-0.051}$
CLEO2013 [@CLEO:2011ab] $0.56(1)^{+0.02}_{-0.03}$ $0.47(6)(4)$ $0.84(9)^{+0.05}_{-0.06}$
3PSR [@Ball:1993tp] $0.5(2)$ $0.4(1)$ $1.0(2)$
HQETF-I [@Wang:2002zba] $0.57(8)$ $0.52(7)$ $0.72(10)$
HQETF-II [@Wu:2006rd] $0.599^{+0.035}_{-0.030}$ $0.372^{+0.026}_{-0.031}$ $0.801^{+0.044}_{-0.036}$
RHOPM [@Wirbel:1985ji] 0.78 0.92 1.23
QM-I [@Isgur:1988gb] 0.59 0.23 1.34
QM-II [@Melikhov:2000yu] 0.59 0.49 0.90
LFQM [@Verma:2011yw] $0.60(1)$ $0.47(0)$ $0.88(3)$
HM$\chi$T [@Fajfer:2005ug] 0.61 0.31 1.05
Lattice [@Lubicz:1991bi] $0.45(4)$ $0.02(26)$ $0.78(12)$
Lattice [@Bernard:1991bz] $0.65(15){^{+0.24}_{-0.23}}$ $0.59(31){^{+0.28}_{-0.25}}$ $1.07(49)(35)$
: The $D\to \rho$ TFFs $A_{1,2}(q^2)$ and $V(q^2)$ at the large recoil region $q^2\simeq 0$. The errors are squared averages of all the mentioned error sources. As a comparison, we also present the prediction from various methods.[]{data-label="Tab:Tform factor0"}
Table \[Tab:Tform factor0\] shows the $D\to\rho$ TFFs at the large recoil region $q^2 \rightsquigarrow 0~{\rm GeV^2}$, where the uncertainties are squared averages of all the mentioned error sources for the LCSR approach. As a comparison, we also present the predictions from various approaches in Table \[Tab:Tform factor0\], i.e. from the CLEO collaboration [@CLEO:2011ab], the 3PSR [@Ball:1993tp], the HQEFT [@Wang:2002zba; @Wu:2006rd], RHOPM [@Wirbel:1985ji], the QM [@Isgur:1988gb; @Melikhov:2000yu], the LFQM [@Verma:2011yw], the HM$\chi$T [@Fajfer:2005ug], and the Lattice QCD predictions [@Lubicz:1991bi; @Bernard:1991bz], respectively.
$F_i$ $J^P$ $m_{R,i}/{\rm GeV}$
-------------- -------------- ----------------------------
$V $ $1^-$ 2.007
$A_1,A_2$ $1^+$ 2.427
: Resonance masses of quantum number $J^P$ as indicated necessary for the parameterisation of $D\to\rho$ TFFs $A_{1,2}$ and $V$.[]{data-label="Tab:mRi"}
{width="45.00000%"} {width="45.00000%"}
Table \[Tab:Tform factor0\] shows that the TFFs under many approaches are consistent with each other within reasonable errors. To show the relative importance of various TFFs within different approaches more clearly, we present a comparison of the ratios $r_2$ and $r_V$ in Fig. \[Fig\_Tform factorratio\]. The LCSR uncertainties for the TFFs are $\left(^{+13\%}_{-12\%}\right)$ for $r_V$ and $\left(^{+15\%}_{-15\%}\right)$ for $r_2$, which are much smaller than the previous 3PSR predictions (which are $\pm45\%$ and $\pm 48\%$ [@Ball:1993tp], respectively). Thus, by using the LCSR approach, more accurate QCD SR predictions can be obtained.
{width="33.00000%"} {width="33.00000%"} {width="33.00000%"}
The physically allowable range for the TFFs is $0\leq q^2 \leq (m_D-m_\rho)^2=1.18 {\rm GeV}^2$. Theoretically, the LCSRs for the $D\to \rho$ TFFs are applicable in low and intermediate $q^2$-regions, e.g. $q^2\in[0,0.8]\;{\rm GeV}^2$. We can extrapolate them to whole $q^2$-regions via a rapidly converging series over the $z(t)$-expansion [@Khodjamirian:2010vf; @Straub:2015ica] $$\begin{aligned}
F_i(q^2) = P_i (q^2) \sum_{k=0,1,2}a_k^i [z(q^2)-z(0)]^k,\end{aligned}$$ where $$\begin{aligned}
z(t)=\frac{\sqrt{t_+ - t}-\sqrt{t_+ - t_0}}{\sqrt{t_+ - t}+\sqrt{t_+ - t_0}}\end{aligned}$$ with $t_\pm=(m_D\pm m_\rho)^2$, $t_0=t_+(1-\sqrt{1-t_-/t_+})$, and the $F_i$ are three TFFs $A_{1,2}$ and $V$, respectively. The $P_i (q^2) = (1-q^2/m_{R,i}^2)^{-1}$ is a simple pole corresponding to the first resonance in the spectrum. The appropriate resonance masses are given in Table \[Tab:mRi\]. The parameters $a_k^i$ can be fixed by requiring $\Delta < 0.1\%$, and the results are put in Table \[analytic\]. Here the parameter $\Delta$ is introduced to measure the quality of extrapolation, $$\Delta=\frac{\sum_t\left|F_i(t)-F_i^{\rm fit}(t)\right|} {\sum_t\left|F_i(t)\right|}\times 100, \label{delta}$$ where $t\in[0,\frac{1}{40},\cdots,\frac{40}{40}]\times 0.8 {\rm GeV}^2$.
$A_1$ $A_2$ $V$
---------- --------------- --------------- -------------
$a_1^i$ 0.711 $-1.149$ $-0.797$
$a_2^i$ 29.23 $15.108$ $10.370$
$\Delta$ 0.05% 0.04% 0.01%
: The fitted parameters $a^i_{1,2}$ for the $D\to \rho$ TFFs $F_i$, in which all the LCSR parameters are set to be their central values. $\Delta$ is the measure of the quality of extrapolation.[]{data-label="analytic"}
The extrapolated TFFs in whole $q^2$-region are presented in Fig. \[Fig:Tform factor\], where the shaded bands are uncertainties from various input parameters. As a comparison, we also give the results from various approaches, which are from the CLEO Collaboration [@CLEO:2011ab], the QM [@Melikhov:2000yu], the LFQM [@Verma:2011yw], the HL$\chi$PT [@Fajfer:2005ug], and the Lattice predictions [@Flynn:1997ca], respectively. The CLEO Collaboration only issues the TFFs at large recoil region, and the present CLEO curves are fitted ones from the large energy chiral-quark model [@Palmer:2013yia].
The semi-leptonic decay $D\to\rho e\nu_e$
-----------------------------------------
$1/|V_{\rm cd}|^2 \times \Gamma$ $\Gamma^{\rm L}/ \Gamma^{\rm T}$ $\Gamma^+/ \Gamma^-$
-------------------------- ---------------------------------- ---------------------------------- ------------------------
This paper $55.45^{+13.34}_{-9.41}$ $1.18^{+0.14}_{-0.13}$ $0.22^{+0.04}_{-0.03}$
3PSR [@Ball:1993tp] $15.80\pm4.61$ $1.31\pm0.11$ $0.24\pm0.03$
HQEFT [@Wang:2002zba] $71\pm14$ $1.17\pm0.09$ $0.29\pm0.13$
RHOPM[@Wirbel:1985ji] $90.83$ 0.91 0.19
QM [@Isgur:1988gb] $88.86$ 1.33 0.11
Lattice[@Lubicz:1991bi] $54.63\pm12.51$ $1.86\pm0.56$ 0.16
Lattice[@Bernard:1991bz] 71.75 1.10 0.18
: Total decay width $1/|V_{\rm cd}|^2 \times \Gamma$, the ratio of longitudinal and transverse decay width $\Gamma^L/ \Gamma^T$, and the ratio of positive and negative decay width $\Gamma^+/ \Gamma^-$.[]{data-label="Tab:GammaRatio"}
By using the extrapolated $D\to\rho$ TFFs, we calculate the total decay width $1/|V_{\rm cd}|^2 \times \Gamma$, the ratio of longitudinal and transverse decay width $\Gamma^{\rm L}/ \Gamma^{\rm T}$ for the $D\to\rho$ semileptonic decays, and the ratio of positive and negative decay width $\Gamma^+/ \Gamma^-$. The results are presented in Table \[Tab:GammaRatio\], where the results under various approaches have also been presented. Table \[Tab:GammaRatio\] shows that our LCSR predictions for $1/|V_{\rm cd}|^2 \times \Gamma$, $\Gamma^{\rm L}/ \Gamma^{\rm T}$ and $\Gamma^+/ \Gamma^-$ are consistent with other approaches within errors, only the value of $1/|V_{\rm cd}|^2 \times \Gamma$ is quite larger than the 3PSR prediction [@Ball:1993tp].
Decay Mode $D^0\to \rho^- e^+ \nu_e$ $D^+ \to \rho^0 e^+ \nu_e$
----------------------------------------- ------------------------------------ ------------------------------------
This paper $1.749^{+0.421}_{-0.297}\pm 0.006$ $2.217^{+0.534}_{-0.376}\pm 0.015$
CLEO2005 [@Huang:2005iv] $1.94\pm0.39\pm0.13$ $2.1\pm0.4\pm0.1$
CLEO2013 [@CLEO:2011ab] $1.77\pm0.12\pm0.10$ $2.17\pm0.12^{+0.12}_{-0.22}$
3PSR [@Ball:1993tp] $0.5\pm0.1$ –
HQEFT [@Wang:2002zba] $1.4\pm0.3$ –
NWA [@Shi:2017pgh]+HQEFT [@Wu:2006rd] $1.67\pm0.27$ $2.16\pm0.36$
NWA [@Shi:2017pgh]+LFQM [@Verma:2011yw] $1.73\pm0.07$ $2.24\pm0.09$
FK [@Fajfer:2005ug] 2.0 2.5
ISGW2 [@Scora:1995ty] 1.0 1.3
: The branching ratios of the semileptonic decays $D^0\to \rho^- e^+ \nu_e$ and $D^+ \to \rho^0 e^+ \nu_e$ (in unit: $10^{-3}$). As a comparison, we also present the results of CLEO Collaboration [@CLEO:2011ab; @Huang:2005iv], 3PSR [@Ball:1993tp], HQEFT [@Wang:2002zba], NWA [@Shi:2017pgh] with HQEFT [@Wu:2006rd] and LFQM [@Verma:2011yw], FK [@Fajfer:2005ug] and ISGW2 [@Scora:1995ty]. []{data-label="Table:Branching_se"}
As a further step, we calculate the branching ratios for the two $D\to\rho$ semileptonic decays. One is the $D^0$-type decay via the process $D^0 \to \rho^- e^+\nu_e$ with the lifetime $\tau(D^0)= 0.410\pm 0.002$ ps, and the other one is the $D^+$-type decay via the process $D^+ \to \rho^0 e^+\nu_e$ with the lifetime $\tau(D^+)= 1.040\pm 0.007$ ps [@Tanabashi:2018oca]. The results are given in Table \[Table:Branching\_se\], where the first uncertainty is squared average of the mentioned error sources, and the second uncertainty is from the experimental errors for the lifetime. As a comparison, we also list the branching ratios derived from various approaches in Table \[Table:Branching\_se\]. It indicates that a smaller $1/|V_{\rm cd}|^2 \times \Gamma$ predicted by 3PSR leads to a smaller branching ratio. This explains why the previous SR prediction is inconsistent with other approaches. However, by using the LCSR approach, we observe that a more reasonable and accurate SR prediction can be achieved. The LCSR predictions for the branching ratios of the two $D\to\rho$ semileptonic decays also show better agreement with the CLEO measurements.
The radiative decay $D\to\rho \gamma$
-------------------------------------
${\mathcal B}(D^0\to \rho^0 \gamma)$ ${\mathcal B}(D^+ \to \rho^+ \gamma)$
------------------------------------------- --------------------------------------- ------------------------------------------
This paper (SM) $1.744^{+0.598}_{-0.704}$ $5.034^{+0.939}_{-0.958}$
QCD SM [@Khodjamirian:1995uc] $0.38$ $0.46$
Hybrid [@deBoer:2017que] $0.606\pm0.565$ $1.174\pm1.157$
FS [@Fajfer:1997bh] $0.55\pm0.45$ $3.35\pm2.95$
Pole Diagrams and VMD [@Burdman:1995te] $0.3\pm0.2$ $4\pm2$
Belle Collaboration [@Abdesselam:2016yvr] $1.77\pm0.31$ $-$
: The branching ratios for $D\to \rho\gamma$ decays (in unit: $10^{-5}$). As a comparison, we also present other theoretical predictions.[]{data-label="Table:Br_rediative deccay"}
After inputting the $D\to \rho$ TFFs into the parity conserving and parity violating effective couplings ${\cal A}_{\rm PV, PC}^\rho$, we get the branching ratio for the two $D$ meson radiative processes $D^0\to \rho^0\gamma$ and $D^+\to\rho^+\gamma$. The results are given in Table \[Table:Br\_rediative deccay\], where the uncertainties are squared average of the theoretical and experimental error sources. As a comparison, we also listed the branching ratios derived from various approaches. It indicates that the LCSR predictions for the branching ratios for radiative decay $D^0\to \rho^0 \gamma$ shows a better agreement with the Belle Collaboration result, which is larger than other theoretical predictions.
For the direct CP asymmetry of the $D\to \rho\gamma$ decay, we recall that the maximal value of Eq. can be reached in the limit of maximal constructive interference (namely of $\pm \pi/2$ strong phase difference) of the amplitudes with different weak phases. This way we get the upper limit of our predictions for the direct CP asymmetry.
The Wilson coefficient $C_7^{(\prime)\rm eff}$ mainly comes from the hard spectator interaction and weak annihilation contributions within the framework of SM, which results in $C_7^{\rm SM}(m_c) \in [-0.00949+0.0014 i, -0.00019+0.002 i]$ [@deBoer:2017que]. By taking input parameters into Eq., we obtain the SM prediction for the CP asymmetry of $D^0\to \rho^0\gamma$, e.g. $$\begin{aligned}
A_{CP}^{\rm SM} &=& \big[1.329 (\pm0.234)_{C_7} (\pm0.089)_{m_c} \left(^{+0.094}_{-0.073}\right)_{\rm F.F.}
\nonumber\\
&& \left(^{+0.295}_{-0.204}\right)_{a_2} \left(^{+0.052}_{-0.049}\right)_{\rm D.C.} \big]\times 10^{-2}. \\
&=& \left(1.329^{+0.402}_{-0.335}\right)\times 10^{-2}. \label{Eq:Value_ACP}\end{aligned}$$ In the first line, the separate uncertainties are caused by the errors of the quantities $C_7$, $m_c$, TFFs (F.F.), $a_2$ and decay constant (D.C.), respectively, and in the second line, all the errors are added up in quadrature. The central value of the SM prediction is smaller than the Belle result, $A^{\rm exp}_{CP} = 0.056\pm 0.152\pm0.006$ [@Abdesselam:2016yvr] by almost $3$ times. Because of the large statistical error for the present Belle measurements, the SM prediction roughly agrees with the Belle result within errors.
One may hope that the possible discrepancy can be accommodated by a well-motivated extension of the SM. To quantify the size of the Wilson coefficients, one can normalize the effective Hamiltonian within new-physics contributions as $$\begin{aligned}
{\cal H}^{\rm eff-NP} = \frac{G_F}{\sqrt 2} \sum_i C_i {\cal Q}_i + h.c.,~\label{EQ:H}\end{aligned}$$ where the complete list of potentially relevant operators can be found in Ref. [@deBoer:2015boa]. The Wilson coefficients $\delta C_{7,8}^{(\prime)}(M)$ are generically related beyond SM models, with $M$ denotes the matching scale. The initial conditions of the four operators are assuming $M > m_t$, taking into account the renormalization group evolution of the operators at the leading log level, leads to $$\begin{aligned}
&&C^{(\prime)}_7 (m_c) = \tilde\eta[\eta C^{(\prime)}_7 (M) + 8(\eta -1)C^{(\prime)}_8(M)],\label{Eq:C7} \\
&&C^{(\prime)}_8 (m_c) = \tilde\eta C^{(\prime)}_8 (M), \label{Eq:C8}\end{aligned}$$ where the coefficients $\eta$ and $\tilde\eta$ can be found in Ref.[@Buras:1999da]. A non-vanishing value for $\Delta A_{\rm CP} = A_{\rm CP}(K^+ K^-) - A_{\rm CP}(\pi^+ \pi^-)$ has been observed by LHCb and CDF collaborations [@Aaij:2011in; @Chatrchyan:2011fq], which could be used to restrict the new physics contribution from the operator ${\cal Q}_8$ by using the relationship $\Delta A_{\rm CP} \approx -1.8 |{\rm Im} [C_8^{\rm NP}(m_c)]$ [@Giudice:2012qq]. Considering the world average value $\Delta A_{\rm CP}^{\rm exp} = -(0.67\pm0.16)\%$ [@Aaltonen:2011se], one can get $|{\rm Im} [C_8^{\rm NP}(m_c)]| \approx 0.4\times 10^{-2}$. Furthermore the renormalization group evolution implies $|{\rm Im}[C_7^{\rm NP}(m_c)]| \approx |{\rm Im}[C_8^{\rm NP}(m_c)]|$ if the initial scale $M$ is set to be around $1$ TeV, which, for instance, could be happened for super-symmetry theory. Taking into account the uncertainties in determining $|{\rm Im}[C_8^{\rm NP}(m_c)]|$ and the uncertainties from the initial conditions of $|C_7^{\rm NP}(M)|$, one can obtain a conservative range $(0.2-0.8) \times 10^{-2}$ for $|{\rm Im}[C_7^{\rm NP}(m_c)]|$ [@Isidori:2012yx]. Thus we obtain a prediction for the CP asymmetry of $D^0\to \rho^0\gamma$, $$\begin{aligned}
A_{CP}^{\rm NP} &=& \big[3.907 (\pm 2.344)_{C_7} (\pm 0.260)_{m_c} \left(^{+0.276}_{-0.215}\right)_{\rm F.F.}
\nonumber\\
&& \left(^{+0.868}_{-0.601}\right)_{a_2} \left(^{+0.153}_{-0.145}\right)_{\rm D.C.} \big]\times 10^{-2} \\
&=& \left(3.907^{+2.533}_{-2.448}\right) \times 10^{-2}.\end{aligned}$$ In the second line, all the errors are added up in quadrature. This value agrees with the Belle Collaboration result within errors.
Summary {#section:4}
=======
In the paper, we have investigated the $D \to \rho$ TFFs within the LCSR approach. As shown by Table \[Tab:Tform factor0\] and Fig.\[Fig\_Tform factorratio\], more accurate QCD SR predictions for the TFFs $A_{1,2}(q^2)$ and $V(q^2)$ can be achieved by applying the LCSR approach other than the 3PSR approach. To compare with the CLEO measurements, the LCSR approach can give reasonable explanations for the $D \to \rho$ TFFs. The pQCD factorization approach is applicable in large recoil region $q^2 \rightsquigarrow 0$ and the lattice QCD approach is applicable in very large $q^2$-region, thus the extrapolation of the results under those two approaches shall be strongly model dependent. The LCSR prediction provides a bridge between the pQCD and lattice QCD predictions, since it is applicable for a wider range, i.e. in both small and intermediate $q^2$-region.
After extrapolating the $D\to\rho$ TFFs to whole $q^2$-region, the LCSR predictions for the branching ratios of the two $D$ meson semileptonic decays are ${\cal B} (D^0\to \rho^- e^+ \nu_e) = (1.749^{+0.421}_{-0.297}\pm 0.006)\times 10^{-3}$ and ${\cal B} (D^+ \to \rho^0 e^+ \nu_e) = (2.217^{+0.534}_{-0.376}\pm 0.015)\times 10^{-3}$, respectively, which agree with the CLEO measurements within errors. And as shown by Table \[Table:Br\_rediative deccay\], the branching ratios of the $D$ meson radiative decay $D^0 \to \rho^0 \gamma$ and $D^+ \to \rho^+ \gamma$ also show good agreement with the experimental results in comparing with other theoretical predictions.
Based on the short and long distance parity conserving and parity violating amplitudes, i.e. $( \mathcal{A}_{\mathrm{PV/PC}}^{\rho})^{\rm s.d./l.d.}$, and in the limit of maximal constructive interference of the amplitudes, we get the SM prediction for the CP asymmetry of $D^0\to \rho^0\gamma$, $ A_{CP}^{\rm SM}=(1.329^{+0.402}_{-0.335})\times 10^{-2}$, which has large discrepancy with the Belle Collaboration data. By taking $|{\rm Im}[C_7^{\rm NP}(m_c)]| = (0.2-0.8) \times 10^{-2}$ for the NP Wilson coefficient, we obtain the NP prediction, $A_{CP}^{\rm NP} = (3.907^{+2.533}_{-2.448})\times 10^{-2}$, which agrees with the Belle Collaboration result within errors. Since the short-distance amplitude is relatively much smaller than the long-distance amplitude, the NP effect induced by the operator ${\cal Q}_8$ shall give negligible effect (less than 0.1%) to the SM predictions of the branching ratios. Thus we think that the LCSR approach can provide a self-consistent way to deal with the $D$ meson decays, and the $D$ meson involved processes could be adopted for testing NP beyond the SM.
[**Acknowledgments**]{}: This work was supported in part by the National Science Foundation of China under Grant No.11881240255, No.11765007 and No.11625520, No.11847301, the Project of Guizhou Provincial Department of Science and Technology under Grant No.KY\[2017\]1089 and No.KY\[2019\]1171, the China Postdoctoral Science Foundation under Grant No.2019TQ0329, the Project of Guizhou Minzu University under Grant No.GZMU\[2019\]YB19.
[999]{}
T. E. Coan *et al.* \[CLEO Collaboration\], Absolute branching fraction measurements of exclusive $D^0$ semileptonic decays, Phys. Rev. Lett. [**95**]{}, 181802 (2005).
G. S. Huang *et al.* \[CLEO Collaboration\], Absolute branching fraction measurements of exclusive $D^+$ semileptonic decays, Phys. Rev. Lett. [**95**]{}, 181801 (2005).
S. Dobbs *et al.* \[CLEO Collaboration\], First Measurement of the Form Factors in the Decays $D^0 \to \rho^- e^+ \nu_e$ and $D^+ \to \rho^0 e^+ \nu_e$, Phys. Rev. Lett. [**110**]{}, 131802 (2013).
M. Ablikim [*et al.*]{} \[BESIII Collaboration\], “Observation of $D^+ \to f_0(500) e^+\nu_e$ and Improved Measurements of $D \to\rho e^+\nu_e$,” Phys. Rev. Lett. [**122**]{}, 062001 (2019).
A. Abdesselam [*et al.*]{} \[Belle Collaboration\], “Observation of $D^0\to \rho^0\gamma$ and search for $CP$ violation in radiative charm decays,” Phys. Rev. Lett. [**118**]{},051801 (2017).
P. Ball, The Semileptonic decays $D \to \pi (\rho) e \nu$ and $B \to \pi (\rho) e \nu$ from QCD sum rules, Phys. Rev. D [**48**]{}, 3190 (1993).
W. Y. Wang, Y. L. Wu and M. Zhong, Heavy to light meson exclusive semileptonic decays in effective field theory of heavy quarks, Phys. Rev. D [**67**]{}, 014024 (2003).
Y. L. Wu, M. Zhong and Y. B. Zuo, $B_{(s)}, D_{(s)} \to \pi, K, \eta, \rho, K^*, \omega, \phi$ transition form factors and decay rates with extraction of the CKM parameters $|V_{ub}|$, $|V_{cs}|$, $|V_{cd}|$, Int. J. Mod. Phys. A [**21**]{}, 6125 (2006).
M. Wirbel, B. Stech and M. Bauer, Exclusive semileptonic decays of heavy mesons, Z. Phys. C [**29**]{}, 637 (1985).
N. Isgur, D. Scora, B. Grinstein and M. B. Wise, Semileptonic $B$ and $D$ decays in the quark model, Phys. Rev. D [**39**]{}, 799 (1989).
D. Melikhov and B. Stech, Weak form-factors for heavy meson decays: An Update, Phys. Rev. D [**62**]{}, 014006 (2000).
R. C. Verma, Decay constants and form factors of $s$-wave and $p$-wave mesons in the covariant light-front quark model, J. Phys. G [**39**]{}, 025005 (2012).
S. Fajfer and J. F. Kamenik, Charm meson resonances and $D \to V$ semileptonic form-factors, Phys. Rev. D [**72**]{}, 034029 (2005).
V. Lubicz, G. Martinelli, M. S. McCarthy and C. T. Sachrajda, Semileptonic decays of $D$ mesons in a lattice QCD, Phys. Lett. B [**274**]{}, 415 (1992).
C. W. Bernard, A. X. El-Khadra and A. Soni, Lattice study of semileptonic decays of charm mesons into vector mesons, Phys. Rev. D [**45**]{}, 869 (1992).
M. A. Shifman, A. I. Vainshtein and V. I. Zakharov, QCD and Resonance Physics. Theoretical Foundations, Nucl. Phys. B [**147**]{}, 385 (1979).
V. M. Braun, Light cone sum rules, hep-ph/9801222.
I. I. Balitsky, V. M. Braun and A. V. Kolesnichenko, Radiative Decay $\Sigma^+ \to p \gamma$ in Quantum Chromodynamics, Nucl. Phys. B [**312**]{}, 509 (1989).
V. L. Chernyak and I. R. Zhitnitsky, $B$ meson exclusive decays into baryons, Nucl. Phys. B [**345**]{}, 137 (1990).
P. Ball, V. M. Braun and H. G. Dosch, “Form-factors of semileptonic D decays from QCD sum rules,” Phys. Rev. D [**44**]{}, 3567 (1991).
G. Burdman, E. Golowich, J. L. Hewett and S. Pakvasa, Radiative weak decays of charm mesons, Phys. Rev. D [**52**]{}, 6383 (1995).
C. Greub, T. Hurth, M. Misiak and D. Wyler, “The $c \to u \gamma$ contribution to weak radiative charm decay,” Phys. Lett. B [**382**]{}, 415 (1996).
S. Fajfer and P. Singer, Long distance $c \to u \gamma$ effects in weak radiative decays of $D$ mesons, Phys. Rev. D [**56**]{}, 4302 (1997).
G. Isidori and J. F. Kamenik, Shedding light on $CP$ violation in the charm system via $D \to V \gamma$ decays, Phys. Rev. Lett. [**109**]{}, 171801 (2012)
S. de Boer and G. Hiller, “Flavor and new physics opportunities with rare charm decays into leptons,” Phys. Rev. D [**93**]{}, 074001 (2016).
S. de Boer and G. Hiller, Rare radiative charm decays within the standard model and beyond, JHEP [**1708**]{}, 091 (2017).
S. Fajfer, S. Prelovsek and P. Singer, Long distance contributions in $D \to V \gamma$ decays, Eur. Phys. J. C [**6**]{}, 471 (1999).
N. Isgur and M. B. Wise, Weak Decays of Heavy Mesons in the Static Quark Approximation, Phys. Lett. B [**232**]{}, 113 (1989).
T. M. Yan, H. Y. Cheng, C. Y. Cheung, G. L. Lin, Y. C. Lin and H. L. Yu, Heavy quark symmetry and chiral dynamics, Phys. Rev. D [**46**]{}, 1148 (1992).
P. L. Cho and H. Georgi, Electromagnetic interactions in heavy hadron chiral theory, Phys. Lett. B [**296**]{}, 408 (1992).
J. F. Amundson, C. G. Boyd, E. E. Jenkins, M. E. Luke, A. V. Manohar, J. L. Rosner, M. J. Savage and M. B. Wise, Radiative D\* decay using heavy quark and chiral symmetry, Phys. Lett. B [**296**]{}, 415 (1992).
M. Bauer, B. Stech and M. Wirbel, Exclusive Nonleptonic Decays of $D$, $D_s$, and $B$ Mesons, Z. Phys. C [**34**]{}, 103 (1987).
M. Tanabashi [*et al.*]{} \[Particle Data Group\], Review of Particle Physics, Phys. Rev. D [**98**]{}, 030001 (2018).
T. Huang and X. G. Wu, Determination of the $\eta$ and $\eta'$ Mixing Angle from the Pseudoscalar Transition Form Factors, Eur. Phys. J. C [**50**]{}, 771 (2007).
E. Klempt and A. Zaitsev, Glueballs, Hybrids, Multiquarks. Experimental facts versus QCD inspired concepts, Phys. Rept. [**454**]{}, 1 (2007).
R. Aaij [*et al.*]{} \[LHCb Collaboration\], Study of $\eta-\eta^{\prime}$ mixing from measurement of $B^0_{(s)} \rightarrow J/\psi \eta^{(\prime)}$ decay rates, JHEP [**1501**]{}, 024 (2015).
P. Ball, V. M. Braun, Y. Koike and K. Tanaka, Higher twist distribution amplitudes of vector mesons in QCD: Formalism and twist-three distributions, Nucl. Phys. B [**529**]{}, 323 (1998).
P. Ball and R. Zwicky, $B_{d,s} \to \rho, \omega, K^*, \phi$ decay form-factors from light-cone sum rules revisited, Phys. Rev. D [**71**]{}, 014029 (2005).
H. B. Fu, X. G. Wu, H. Y. Han and Y. Ma, $B \to \rho$ transition form factors and the $\rho$-meson transverse leading-twist distribution amplitude, J. Phys. G [**42**]{}, 055002 (2015).
T. Huang and Z. H. Li, $B \to K^*$ gamma in the light cone QCD sum rule, Phys. Rev. D [**57**]{}, 1993 (1998).
T. Huang, Z. H. Li and X. Y. Wu, Improved approach to the heavy to light form-factors in the light cone QCD sum rules, Phys. Rev. D [**63**]{}, 094001 (2001).
Z. G. Wang, M. Z. Zhou and T. Huang, $B \pi$ weak form-factor with chiral current in the light cone sum rules, Phys. Rev. D [**67**]{}, 094006 (2003).
F. Zuo, Z. H. Li and T. Huang, Form factor for $B \to D l \nu$ in light-cone sum rules with chiral current correlator,” Phys. Lett. B [**641**]{}, 177 (2006).
X. G. Wu, T. Huang and Z. Y. Fang, ${\rm SU}_f$(3)-symmetry breaking effects of the $B \to K$ transition form-factor in the QCD light-cone sum rules, Phys. Rev. D [**77**]{}, 074001 (2008).
X. G. Wu and T. Huang, Radiative corrections on the $B \to P$ form factors with chiral current in the light-cone sum rules, Phys. Rev. D [**79**]{}, 034013 (2009).
J. Gao, C. D. L¨¹, Y. L. Shen, Y. M. Wang and Y. B. Wei, Precision calculations of $B \to V$ form factors in QCD, arXiv:1907.11092.
P. Ball, V. M. Braun and A. Lenz, Twist-4 distribution amplitudes of the $K^*$ and $\phi$ mesons in QCD, J.High Energy Phys. 08 (2007) 090.
H. M. Choi and C. R. Ji, Distribution amplitudes and decay constants for ($\pi, K, \rho, K^*$) mesons in light-front quark model, Phys. Rev. D [**75**]{}, 034019 (2007).
J. R. Forshaw and R. Sandapen, An AdS/QCD holographic wavefunction for the $\rho$-meson and diffractive $\rho$ meson electroproduction, Phys. Rev. Lett. [**109**]{}, 081601 (2012).
J. Xu, Q. A. Zhang and S. Zhao, Light-cone distribution amplitudes of vector meson in large momentum effective theory, arXiv:1804.01042 \[hep-ph\].
H. B. Fu, X. G. Wu, H. Y. Han, Y. Ma and T. Zhong, $|V_{cb}|$ from the semileptonic decay $B\to D \ell \bar{\nu}_{\ell}$ and the properties of the $D$ meson distribution amplitude, Nucl. Phys. [**B884**]{}, 172 (2014). X. G. Wu and T. Huang, An implication on the pion distribution amplitude from the pion-photon transition form factor with the new BABAR data, Phys. Rev. D [**82**]{}, 034024 (2010).
G. P. Lepage, S. J. Brodsky, T. Huang, and P. B. Mackezie, in Particles and Fields, Proceedings of the Banff Summer Institute on Particle Physics, Banff, Alberta, Canada, 1981 2, edited by A. Z. Capri and A. N. Kamal (Plenum, New York, 1983), p.83; T. Huang, in Proceedings of XXth International Conference on High Energy Physics, Madison, Wisconsin, 1980, edited by L. Durand and L.G. Pondrom, AIP Conf. Proc. No. 69 (AIP, New York, 1981), p. 1000. A. Bharucha, D. M. Straub and R. Zwicky, $B\to V \ell^+\ell^-$ in the standard model from light-cone sum rules, arXiv:1503.05534.
A. Khodjamirian, T. Mannel, A. A. Pivovarov and Y. M. Wang, Charm-loop effect in $B \to K^{(*)} \ell^{+} \ell^{-}$ and $B\to K^*\gamma$, J. High Energy Phys. 09 (2010) 089.
J. M. Flynn and C. T. Sachrajda, Heavy quark physics from lattice QCD, Adv. Ser. Direct. High Energy Phys. [**15**]{}, 402 (1998).
T. Palmer and J. O. Eeg, Form factors for semileptonic $D$ decays, Phys. Rev. D [**89**]{}, 034013 (2014).
Y. J. Shi, W. Wang and S. Zhao, Chiral dynamics, $S$-wave contributions and angular analysis in $D\to \pi \pi \ell \bar{\nu}$, Eur. Phys. J. C [**77**]{}, 452 (2017).
D. Scora and N. Isgur, Semileptonic meson decays in the quark model: An update, Phys. Rev. D [**52**]{}, 2783 (1995).
A. Khodjamirian, G. Stoll and D. Wyler, Calculation of long distance effects in exclusive weak radiative decays of $B$ meson, Phys. Lett. B [**358**]{}, 129 (1995).
A. J. Buras, G. Colangelo, G. Isidori, A. Romanino and L. Silvestrini, Connections between $\epsilon^\prime / \epsilon$ and rare kaon decays in supersymmetry, Nucl. Phys. B [**566**]{}, 3 (2000).
R. Aaij [*et al.*]{} \[LHCb Collaboration\], “Evidence for CP violation in time-integrated $D^0 \to h^-h^+$ decay rates,” Phys. Rev. Lett. [**108**]{}, 111602 (2012).
S. Chatrchyan [*et al.*]{} \[CMS Collaboration\], “Search for signatures of extra dimensions in the diphoton mass spectrum at the Large Hadron Collider,” Phys. Rev. Lett. [**108**]{}, 111801 (2012).
G. F. Giudice, G. Isidori and P. Paradisi, Direct CP violation in charm and flavor mixing beyond the SM, JHEP [**1204**]{}, 060 (2012).
T. Aaltonen [*et al.*]{} \[CDF Collaboration\], Measurement of CP violating asymmetries in $D^0\to\pi^+\pi^-$ and $D^0\to K^+K^-$ decays at CDF, Phys. Rev. D [**85**]{}, 012009 (2012).
[^1]: Such a small $s_0$ dependence also plays a role to suppress the unwanted scalar contribution due to the choice of chiral correlator.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We formulate a theory of the AC spin Hall magnetoresistance (SMR) in a bilayer system consisting of a magnetic insulator such as yttrium iron garnet (YIG) and a heavy metal such as platinum (Pt). We derive expressions for the DC voltage generation based on the drift-diffusion spin model and quantum mechanical boundary condition at the interface that reveal a spin torque ferromagnetic resonance (ST-FMR). We predict that ST-FMR experiments will reveal valuable information on the current-induced magnetization dynamics of magnetic insulators and AC spin Hall effect.'
author:
- 'Takahiro Chiba$^{1}$, Gerrit E. W. Bauer$^{1,2,3}$, and Saburo Takahashi$^{1}$'
title: 'Current-induced spin torque resonance of magnetic insulators'
---
Ferrimagnetic insulators such as yttrium iron garnet (YIG) with high critical temperatures and very low magnetization damping have been known for decades to be choice materials for in optical, microwave or data storage technologies [@Wu]. Near-dissipationless propagation of spin waves make YIG wires and circuits interesting for low power data transmission and logic devices. A crucial breakthrough was the discovery that the magnetization in YIG can be excited electrically by Pt contacts [@Kajiwara], thereby creating an interface between electronic/spintronic and magnonic circuits. However, the generation of coherent spin waves by the current-induced spin orbit torques in Pt is a strongly non-linear process and the low critical threshold currents found by the experiments [@Kajiwara] cannot yet be explained by theory [@Xiao]. Here we suggest and model a simpler method to get to grips with the important current-magnetization interaction in the YIG$|$Pt system without a problematic threshold, *viz*. by employing the recently discovered magnetoresistance of YIG$|$Pt bilayers or Spin Hall Magnetoresistance (SMR) [@Nakayama; @Chen] to detect current-induced spin torque ferromagnetic resonance (ST-FMR).
The SMR is the dependence of the electrical resistance of the normal metal on the magnetization angle of a proximity insulator and is caused by a concerted action of the Spin Hall Effect (SHE) [@review12] and its inverse (ISHE). An alternative mechanism of the SMR phenomenology in terms of an equilibrium proximity magnetization close to the YIG interface has been proposed [@Chen]. However, this interpretation has been challenged by experiments [@Nakayama; @Althammer]. Moreover, while experiments of many groups are described quantitatively well by the SMR model with one set of parameters [@Vlietstra; @Hahn1; @VlietstraX; @Isasa; @Marmion], we are not aware of a transport theory that explains the observed magnetoresistance in terms of a monolayer-order magnetic Pt.
Current-induced spin torque ferromagnetic resonance (ST-FMR) has been demonstrated [@Liu; @Kondou; @Ganguly] in bilayer thin films made from metallic ferromagnets (FM) and nonmagnetic metals (N). In these experiments the SHE transforms an in-plane alternating current (AC) into an oscillating transverse spin current. The resultant spin transfer resonates with the magnetization at the FMR frequency. The effects induced simultaneously by the Oersted field can be distinguished by a different symmetry of the resonance on detuning. The magnetization dynamics leads to a time dependence of the bilayer resistance by the anisotropic magnetoresistance (AMR). Mixing the applied current and the oscillating resistance generates a DC voltage that is referred to as spin torque diode effect [@Tulapurkar; @Sankey].
The longitudinal spin Seebeck effect was found to be frequency independent up to 30 MHz [@Roschewsky]. The DC ISHE induced by spin pumping has been observed by many groups, but detection of the AC spin Hall effect [@Jiao13] has only recently been reported in metallic structures [@Weiler0; @Wei; @Hyde] as well as in Pt$|$YIG under parametric microwave excitation [@Hahn2]. A DC voltage can be generated in Pt$|$YIG under FMR conditions by rectification of the AC spin Hall effect by means of the SMR, but this signal was found to be swamped by the DC spin Hall effect [@Iguchi]. A study of the spin Hall impedance concludes that the material constants of Pt$|$YIG bilayers do not depend on frequency up to 4 GHz [@Lotze].
In this paper we suggest to combine the principles sketched above to realize ST-FMR for bilayers of a ferro- or ferrimagnetic insulator (FI) such as YIG and a normal metal with spin orbit interaction (N) such as platinum [@Mosendz10] (see Fig. \[ACSMR\]). We derive the magnetization dynamics and DC voltages generated by the SMR-induced spin torque diode effect as a function of the external magnetic field. Our theory should help to better understand the elusive current-induced magnetization dynamics of ferromagnetic insulators which should pave the way for low-power devices based on magnetic insulators [@Wu].
![Schematic set-up to observe the SMR mediated spin torque diode effect. The light blue rectangle is a normal metal (N) film with a finite spin Hall angle, while F is a ferromagnetic insulator. The F$|$N bilayer film is patterned into a strip with width $w$ and length $h$. A Bias-Tee allows detection of a DC voltage under an AC bias.[]{data-label="ACSMR"}](chiba_fig1.eps){width="45.00000%"}
The spin current through an F$|$N interface is governed by the complex spin-mixing conductance $G^{\uparrow\downarrow}$ [@Brataas00]. The prediction of a large $\operatorname{Re}G^{\uparrow\downarrow}$ for interfaces between YIG and simple metals by first-principle calculations [@Jia11] has been confirmed by recent experiments [@Burrowes12; @Weiler]. The spin transport in N (spin Hall system) can be treated by spin-diffusion theory with quantum mechanical boundary conditions at the interface to the insulating ferromagnet [@Chen; @Chiba]. The AC current with frequency $\omega_{a}=2\pi
f_{a}$ induces a spin accumulation distribution $\boldsymbol{\mu}_{s}(z,t)$ in N that obeys the spin-diffusion equation $$\partial_{t}\boldsymbol{\mu}_{s}=D\partial_{z}^{2}\boldsymbol{\mu}_{s}%
-\frac{\boldsymbol{\mu}_{s}}{\tau_{\mathrm{sf}}}, \label{tdsd}%$$ where $D$ is the charge diffusion constant and $\tau_{\mathrm{sf}}$ spin-flip relaxation time in N. In position-frequency space the solution for the spatiotemporal dependence of the spin accumulation reads $\boldsymbol{\mu}%
_{s}(z,\omega)=\boldsymbol{A}e^{-\kappa(\omega)z}+\boldsymbol{B}%
e^{\kappa(\omega)z}$, where $\kappa^{2}(\omega)=(1+i\omega\tau_{\mathrm{sf}%
})/\lambda^{2}$, $\lambda=\sqrt{D\tau_{\mathrm{sf}}}$ is the spin-diffusion length, and the constant column vectors $\boldsymbol{A}$ and $\boldsymbol{B}$ are determined by the boundary conditions for the spin current density in the $z$-direction $\boldsymbol{J}_{s,z}(z)$, where $\boldsymbol{J}_{s,z}%
/\left\vert \boldsymbol{J}_{s,z}\right\vert $ is the spin polarization vector, which is continuous at the interface to the ferromagnet at $z=0$ and vanishes at the vacuum interface at $z=d_{N}$. For planar interfaces $$\boldsymbol{J}_{s,z}(z,\omega)=-\theta_{\mathrm{SH}}J_{c}(\omega
)\mathbf{\hat{y}}-\sigma\partial_{z}\frac{\boldsymbol{\mu}_{s}(z,\omega)}{2e},
\label{isx}%$$ where $\theta_{\mathrm{SH}}$ is the spin Hall angle, $\sigma$ the electrical conductivity, and $J_{c}(\omega)=2\pi J_{c}^{0}\delta(\omega_{a}-\omega)$ the currents not accounting for spin-orbit interaction. $\boldsymbol{J}%
_{s,z}\left( d_{N},\omega\right) =0\ $and$\boldsymbol{J}_{s,z}%
(0,\omega)=\int_{-\infty}^{\infty}\boldsymbol{J}_{s,z}(0,t)e^{-i\omega t}dt$, where $\boldsymbol{J}_{s,z}\left( 0,t\right) =\boldsymbol{J}_{s}%
^{\mathrm{T}}+\boldsymbol{J}_{s}^{\mathrm{P}}=\boldsymbol{J}_{s}%
^{\mathrm{(F)}}$ with $$\begin{gathered}
\boldsymbol{J}_{s}^{\mathrm{T}}=\frac{G_{r}}{e}\hat{\mathbf{M}}\times\left(
\hat{\mathbf{M}}\times\boldsymbol{\mu}_{s}(0)\right) +\frac{G_{i}}{e}%
\hat{\mathbf{M}}\times\boldsymbol{\mu}_{s}(0),\\
\boldsymbol{J}_{s}^{\mathrm{P}}=\frac{\hbar}{e}\left( G_{r}\hat{\mathbf{M}%
}\times\partial_{t}\hat{\mathbf{M}}+G_{i}\partial_{t}\hat{\mathbf{M}}\right)
,\end{gathered}$$ where $\hat{\mathbf{M}}$ is the unit vector along the FI magnetization and $G^{\uparrow\downarrow}=G_{r}+iG_{i}$ the complex spin-mixing interface conductance per unit area of the FI$|$N interface. The imaginary part $G_{i}$ can be interpreted as an effective exchange field acting on the spin accumulation, which is usually much smaller than the real part. A positive $\boldsymbol{J}_{s}^{\mathrm{(F)}}$ [@Jia11; @VlietstraX] corresponds here to up-spins flowing from FI into N. For Pt(Ta) $\omega_{a}\tau_{\mathrm{sf}%
}^{\mathrm{Pt(Ta)}}=1(15)\times10^{-3}$ at the FMR frequency $f_{a}=15.5\,$GHz with $\tau_{\mathrm{sf}}^{\mathrm{Pt(Ta)}}=0.01(0.15)\,$ps, indicating that the condition $\omega_{a}\tau_{\mathrm{sf}}^{\mathrm{Pt(Ta)}}\ll1$ is fulfilled for these metals [@Jiao13]. In this limit the frequency dependence of the spin diffusion length may be disregarded such that $$\begin{aligned}
& \boldsymbol{\mu}_{s}(z,t)\nonumber\\
& \rightarrow-\mathbf{\hat{y}}\mu_{s0}(t)\frac{\sinh\frac{2z-d_{N}}{2\lambda
}}{\sinh\frac{d_{N}}{2\lambda}}+\boldsymbol{J}_{s}^{\mathrm{(F)}}%
\frac{2e\lambda}{\sigma}\frac{\cosh\frac{z-d_{N}}{\lambda}}{\sinh\frac{d_{N}%
}{\lambda}},\end{aligned}$$$$\begin{aligned}
\boldsymbol{J}_{s}^{\mathrm{(F)}} & =\frac{\mu_{s0}(t)}{e}\left[
\hat{\mathbf{M}}\times\left( \hat{\mathbf{M}}\times\mathbf{\hat{y}}\right)
\operatorname{Re}+\hat{\mathbf{M}}\times\mathbf{\hat{y}}\operatorname{Im}%
\right] T\nonumber\\
& +\frac{\hbar}{e}\left[ \left( \hat{\mathbf{M}}\times\partial_{t}%
\hat{\mathbf{M}}\right) \operatorname{Re}+\partial_{t}\hat{\mathbf{M}%
}\operatorname{Im}\right] T, \label{jsf}%\end{aligned}$$ where $\mu_{s0}(t)=(2e\lambda/\sigma)\theta_{\mathrm{SH}}J_{c}^{0}\left(
t\right) \tanh\left[ d_{N}/\left( 2\lambda\right) \right] $ with $J_{c}^{0}(t)=J_{c}^{0}\operatorname{Re}(e^{i\omega_{a}t})$ and $T=\sigma
G^{\uparrow\downarrow}/\left[ \sigma+2\lambda G^{\uparrow\downarrow}%
\coth(d_{N}/\lambda)\right] $. The ISHE drives a charge current in the $x$-$y$ plane by the diffusion spin current along $z$. The total charge current density reads $$\boldsymbol{J}_{c}(z,t)=J_{c}^{0}(t)\mathbf{\hat{x}}+\sigma\theta
_{\mathrm{SH}}\left( \nabla\times\frac{\boldsymbol{\mu}_{s}(z,t)}{2e}\right)
.$$ The averaged current density over the film thickness is $\overline{J_{c,x}%
}(t)=d_{N}^{-1}\int_{0}^{d_{N}}J_{c,x}(z,t)dz=J_{\mathrm{SMR}}%
(t)+J_{\mathrm{SP}}(t)$ with $$\begin{aligned}
J_{\mathrm{SMR}}(t) & =J_{c}^{0}(t)\left[ 1-\frac{\Delta\rho_{0}}{\rho
}-\frac{\Delta\rho_{1}}{\rho}\left( 1-\hat{M}_{y}^{2}\right) \right]
,\label{SMRcurrent}\\
J_{\mathrm{SP}}(t) & =J_{r}^{P}\omega_{a}^{-1}\left( \hat{\mathbf{M}}%
\times\partial_{t}\hat{\mathbf{M}}\right) _{y}+J_{i}^{P}\omega_{a}%
^{-1}\partial_{t}\hat{M}_{y},\label{ISHEcurrent}\\
& J_{r(i)}^{P}=\frac{\hbar\omega_{a}}{2ed_{N}\rho}\theta_{\mathrm{SH}%
}\operatorname{Re}\left( \operatorname{Im}\right) \eta,\nonumber\end{aligned}$$ where $J_{\mathrm{SMR}}(t)$ and $J_{\mathrm{SP}}(t)$ are SMR rectification and spin pumping-induced charge currents, $\rho=\sigma^{-1}$ is the resistivity of the bulk normal metal layer and we recognize the conventional DC SMR with $\Delta\rho_{0}=-\rho\theta_{\mathrm{SH}}^{2}(2\lambda/d_{N})\tanh
(d_{N}/2\lambda)$ and $\Delta\rho_{1}=-\Delta\rho_{0}\operatorname{Re}\eta/2$, where $$\eta=\frac{2\lambda\rho G^{\uparrow\downarrow}\tanh\frac{d_{N}}{2\lambda}%
}{1+2\lambda\rho G^{\uparrow\downarrow}\coth\frac{d_{N}}{\lambda}},$$ are effective resistivities that do not depend on frequency [@Chen].
ST-FMR experiments employ the AC impedance of the oscillating transverse spin Hall current caused by the induced magnetization dynamics that is described by the Landau-Lifshitz-Gilbert (LLG) equation, including the transverse spin current Eq. (\[jsf\]), $$\partial_{t}\hat{\mathbf{M}}=-\gamma\hat{\mathbf{M}}\times
\mathbf{H_{\mathrm{eff}}}+\alpha_{0}\hat{\mathbf{M}}\times\partial_{t}%
\hat{\mathbf{M}}+\frac{\gamma\hbar\boldsymbol{J}_{s}^{(F)}}{2eM_{s}d_{F}},$$ where $\mathbf{H_{\mathrm{eff}}}=\mathbf{H_{\mathrm{ex}}}%
+\mathbf{H_{\mathrm{dy}}}$ with an external magnetic field $\mathbf{H_{\mathrm{ex}}}$ and the sum of the AC current-induced Oersted field and the (thin film limit of) the dynamic demagnetization $\mathbf{H}%
_{\mathrm{dy}}=\mathbf{H}_{ac}(t)+\mathbf{H}_{d}(t)=\left( 0,H_{ac}%
e^{i\omega_{a}t},-4\pi M_{z}(t)\right) $. $\gamma$, $\alpha_{0}$, $M_{s}$ and $d_{F}$ are the gyromagnetic ratio, the Gilbert damping constant of the isolated film, the saturation magnetization, and the thickness of the FI film, respectively.
We henceforth disregard the very low in-plane magnetocrystalline anisotropy field of $H_{k}\sim3$Oe reported [@Vlietstra]. The external magnetic field $\mathbf{H_{\mathrm{ex}}}$ is applied at a polar angle $\theta$ in the $x$-$y$ plane. It is convenient to consider the magnetization dynamics in the $XYZ$-coordinate system (Fig. \[ACSMR\]) in which the magnetization is stabilized along the $X$-axis by a sufficiently strong external magnetic field. Denoting the transformation matrix as $R(\theta)$, the magnetization $\mathbf{M}_{R}(t)=R(\theta)\mathbf{M}(t)$ precesses around the $X$-axis, where $\mathbf{M}_{R}(t)=\mathbf{M}_{R}^{0}+\mathbf{m}_{R}(t)\approx\left(
M_{s},m_{Y}(t),m_{Z}(t)\right) $ as shown in Fig. \[ACSMR\]. $\mathbf{M}%
_{R}^{0}$ and $\mathbf{m}_{\mathrm{R}}(t)$ are the static and the dynamic components of the magnetization, respectively. The LLG equation in the $XYZ$-system then becomes $\beta\partial_{t}\mathbf{M}_{R}=-\gamma
\mathbf{M}_{R}\times\mathbf{H}_{\mathrm{eff},R}+\alpha\hat{\mathbf{M}}%
_{R}\times\partial_{t}\mathbf{M}_{R}$ where the effective magnetic field in the $XYZ$-system is $\mathbf{H}_{\mathrm{eff},R}=H_{X}\hat{\mathbf{X}}%
+H_{Y}e^{i\omega_{a}t}\mathbf{\hat{Y}}+\left( H_{Z}e^{i\omega_{a}t}-4\pi
m_{Z}(t)\right) \mathbf{\hat{Z}}$ with $H_{X}=H_{ex}$, $H_{Y}=(H_{ac}%
+H_{i})\cos\theta$ and $H_{Z}=H_{r}\cos\theta$ with $$H_{r(i)}=\frac{\hbar}{2eM_{s}d_{F}}\theta_{\mathrm{SH}}J_{c}^{0}%
\operatorname{Re}\left( \operatorname{Im}\right) \eta,$$ a modulated damping $\alpha=\alpha_{0}+\Delta\alpha$ and g-factor $\beta=1-\Delta\beta$ with $\Delta\alpha\left( \Delta\beta\right)
=\gamma\hbar^{2}/(2e^{2}M_{s}d_{F})\operatorname{Re}T\left( \operatorname{Im}%
T\right) $. For a small-angle precession around the equilibrium direction $\mathbf{M}_{R}^{0}$, $\mathbf{m}_{R}(t)=(0,\delta m_{Y}e^{i\omega_{a}%
t},\delta m_{Z}e^{i\omega_{a}t})$ $\left( \operatorname{Re}[\delta
m_{Y}]\,\operatorname{Re}[\delta m_{Z}]\ll M_{s}\right) $. Disregarding higher orders in $\delta m_{Y(Z)}$ in the $R$-transformed LLG equation we arrive at the (Kittel) relation between AC current frequency and resonant magnetic field $H_{F}=-2\pi M_{s}+\sqrt{(2\pi M_{s})^{2}+(\omega_{a}%
/\gamma)^{2}}$.
A DC voltage is generated by two different mechanisms, viz. the time-dependent oscillations of the SMR in N (spin torque diode effect) and the ISHE generated by spin pumping. This is quite analogous to electrically detected FMR in which the magnetization is driven by microwaves in cavities or coplanar wave guides. In metallic bilayers, the spin pumping signal due to the ISHE can be separated from effects of the magnetoresistance of the metallic ferromagnet by sample design and angular dependences [@Bai; @Obstbaum]. Here we focus on the current-induced magnetization dynamics that induces down-converted DC and second harmonic components in the normal metal. Indicating time-average by $\left\langle \cdots\right\rangle _{t}$ the open-circuit DC voltage is $V_{DC}=h\rho\langle\overline{J_{c,x}}(t)\rangle_{t}=V_{\mathrm{SMR}%
}+V_{\mathrm{SP}},$ where $V_{\mathrm{X}}=h\rho\langle J_{\mathrm{X}%
}(t)\rangle_{t}$. The SMR rectification and spin pumping induced DC voltage are
$$\begin{aligned}
V_{\mathrm{SMR}}&=-\frac{h\Delta\rho_{1}J_{c}^{0}}{4}\frac{F_{S}(H_{\mathrm{ex}}%
)}{\Delta}\left[ C\left( H_{r}+\alpha H_{ac}\right) +C_{+}H_{ac} \frac{H_{\mathrm{ex}}-H_{F}}{\Delta
}\right] \cos\theta\sin2\theta, \label{VSMR}\\
V_{\mathrm{SP}}&=\frac{h\rho J_{r}^{P}}{4}\frac{F_{S}%
(H_{\mathrm{ex}})}{\Delta}%
C\left[ C_{-} \frac{H_{r}^{2}+\alpha H_{r}H_{ac}}{\Delta}+C_{+} \frac{H_{ac}^{2}-\alpha
H_{r}H_{ac}}{\Delta}\right] \cos\theta\sin2\theta,\label{VSP}\end{aligned}$$
where $C=\tilde{\omega}_{a}/\sqrt{1+\tilde{\omega}_{a}^{2}}$ and $C_{\pm}=1\pm1/\sqrt{1+\tilde{\omega}_{a}^{2}}$ with $\tilde{\omega}%
_{a}=\omega_{a}/(2\pi M_{s}\gamma)$, $F_{S}(H_{\mathrm{ex}})=\Delta
^{2}/[\left( H_{\mathrm{ex}}-H_{F}\right) ^{2}+\Delta^{2}]$, $\Delta
=\alpha\omega_{a}/\gamma$ the line width, $H_{ac}=2\pi J_{c}^{0}d_{N}/c$ the Oersted field from the AC current determined by Ampère’s Law (in the limit of an extended film), and $c$ speed of light. Using the material parameters for YIG [@Kajiwara] and Pt [@Weiler; @Obstbaum] shown in Tables \[tab.YIG\] and \[tab.Pt\] we compute the DC voltages in Eqs. (\[VSMR\]) and (\[VSP\]). The calculated $V_{\mathrm{SMR}}$ is plotted in Fig. \[fig.VSMR\] as a function of an external magnetic field and for different $d_{F}$, resolved in terms of the contributions to the FMR caused by the spin transfer torque (symmetric) and the Oersted magnetic field (asymmetric). In Fig. \[fig.VDC\] we show the total DC voltage with both spin torque diode and spin pumping contributions. The DC voltage in F$|$Pt bilayers depends more sensitively on $d_{F}$ for $\mathrm{F=YIG}$ than $\mathrm{F=Py/CoFeB}$ because spin pumping is more important when the Gilbert damping is small. ST-FMR measurements are carried out at relatively high current density, so Joule heating in Pt may cause observable effects, the most notable being the spin Seebeck effect (SSE), which adds a constant background DC voltage to the SMR rectification signal [@Michael14].
$\gamma\ [\mathrm{T^{-1}s^{-1}}]$ $M_{s}\ [\mathrm{Am^{-1}}]$ $\alpha_{0}$
----- ----------------------------------- ----------------------------- --------------------- --
YIG 1.76$\times 10^{11}$ 1.56$\times 10^{5}$ 6.7$\times 10^{-5}$
: \[tab.YIG\]Material parameters for the FI layer.
$G_{r}\ [\Omega^{-1}\mathrm{m}^{-2}]$ $\rho\ [\mu\Omega \mathrm{cm}]$ $\lambda\ [\mathrm{nm}]$ $\theta_{\mathrm{SH}}$
---- --------------------------------------- --------------------------------- -------------------------- ------------------------
Pt 3.8$\times 10^{14}$ 41 1.4 0.12
: \[tab.Pt\]Material parameters for the N layer.
[Reference . ]{}
![(a) The ferromagnet thickness dependence of calculated SMR rectificied voltage for YIG$|$Pt at $f_{a}=9\,$GHz with current density $J_{c}^{0}=10^{10}%
\ \mathrm{A/m^{2}}$ and F(N) layer length and width $h=w=30\,\mathrm{\mu}$m and $\theta=45^{\circ}$. (b) $d_{F(N)}=4(6)$ nm.[]{data-label="fig.VSMR"}](chiba_fig2.eps){width="45.00000%"}
![Dependence of the ST-FMR spectra on $d_{F}$ at $f_{a}=9\,$GHz and $\theta=45^{\circ}$. Inset: Contributions by SMR rectification and spin pumping for $d_{F}=4\,$nm.[]{data-label="fig.VDC"}](chiba_fig3.eps){width="45.00000%"}
The ST-FMR spectra in Fig. \[fig.VDC\] are enhanced for thicker F layers, but these are dominated by the Oersted field actuation. These less-interesting contributions can be eliminated in a tri-layer structure as in Fig. \[fig.Vtri\] in which the magnetic insulator is sandwiched by two normal metal films with the same electric impedance. The second film N2 should be Cu or another metal with negligible spin-orbit interaction and thereby contributions to the ST-FMR, the quality of the YIG$\vert$N2 interface is therefore less of an issue. In Fig. \[fig.Vtri\] we plot pure ST-FMR signals obtained by setting $H_{ac}=0$ in Eqs. (\[VSMR\]) and (\[VSP\]), which may now be observed also for thick magnetic layers.
![ST-FMR spectra dependence on $d_{F}$ in a trilayer set-up to observe the spin torque induced DC voltages without artifacts of the Oersted field. ($f_{a}=9\,$GHz and $\theta=45^{\circ}$)[]{data-label="fig.Vtri"}](chiba_fig4.eps){width="45.00000%"}
In summary, we predict observable AC current-driven ST-FMR in bilayer systems consisting of a ferromagnetic insulator such as YIG and a normal metal with spin-orbit interaction such as Pt. Our main results are the DC voltages caused by an AC current as a function of in-plane external magnetic field and film thickness of a magnetic insulator. The DC voltages generated in YIG$|$Pt bilayers depend sensitively on the ferromagnet layer thickness because of the small bulk Gilbert damping. The predictions can be tested experimentally by ST-FMR-like experiments with a magnetic insulator that would yield important insights into the nature of the conduction electron spin-magnon exchange interaction and current-induced spin wave excitations at the interface of metals and magnetic insulators.
This work was supported by KAKENHI (Grants-in-Aid for Scientific Research) Nos. 22540346, 25247056, 25220910, and 268063, FOM (Stichting voor Fundamenteel Onderzoek der Materie), the ICC-IMR, the EU-RTN Spinicur, EU-FET grant InSpin 612759, and DFG Priority Programme 1538 (Grant No. BA 2954/1).
[99]{}
Recent Advances in Magnetic Insulators - From Spintronics to Microwave Applications, edited by M. Wu and A. Hoffmann, Solid State Physics **64** (Academic Press, 2013).
Y. Kajiwara, K. Harii, S. Takahashi, J. Ohe, K. Uchida, M. Mizuguchi, H. Umezawa, H. Kawai, K. Ando, K. Takanashi, S. Maekawa, and E. Saitoh, Transmission of electrical signals by spin-wave interconversion in a magnetic insulator, Nature **464**, 262 (2010).
Y. Zhou, H. J. Jiao, Y.-T. Chen, G. E. W. Bauer, and J. Xiao, Current-induced spin-wave excitation in Pt/YIG bilayer, Phys. Rev. B **88**, 184403 (2013).
H. Nakayama, M. Althammer, Y.-T. Chen, K. Uchida, Y. Kajiwara, D. Kikuchi, T. Ohtani, S. Geprägs, M. Opel, S. Takahashi, R. Gross, G. E. W. Bauer, S. T. B. Goennenwein, and E. Saitoh, Spin Hall Magnetoresistance Induced by a Nonequilibrium Proximity Effect, Phys. Rev. Lett. **110**, 206601 (2013).
Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Theory of spin Hall magnetoresistance, Phys. Rev. B **87**, 144411 (2013).
S. R. Marmion, M. Ali, M. McLaren, D. A. Williams, and B. J. Hickey, Temperature dependence of spin Hall magnetoresistance in thin YIG/Pt films, Phys. Rev. B **89**, 220404(R) (2014).
M. Althammer, S. Meyer, H. Nakayama, M. Schreier, S. Altmannshofer, M. Weiler, H. Huebl, S. Geprgs, M. Opel, R. Gross, D. Meier, C. Klew, T. Kuschel, J.-M. Schmalhors, G. Reiss, L. Shen, A. Gupta, Y.-T. Chen, G. E. W. Bauer, E. Saitoh, and S. T. B. Goennenwein, Quantitative study of the spin Hall magnetoresistance in ferromagnetic insulator/normal metal hybrids, Phys. Rev. B **87**, 224401 (2013).
N. Vlietstra, J. Shan, V. Castel, B. J. van Wees, and J. Ben Youssef, Spin-Hall magnetoresistance in platinum on yttrium iron garnet: Dependence on platinum thickness and in-plane/out-of-plane magnetization, Phys. Rev. B **87**, 184421 (2013).
C. Hahn, G. de Loubens, O. Klein, M. Viret, V. V. Naletov, and J. B. Youssef, Comparative measurements of inverse spin Hall effects and magnetoresistance in YIG/Pt and YIG/Ta, Phys. Rev. B **87**, 174417 (2013).
N. Vlietstra, J. Shan, V. Castel, J. Ben Youssef, G. E. W. Bauer and B. J. van Wees, Exchange magnetic field torques in YIG/Pt bilayers observed by the spin-Hall magnetoresistance, Appl. Phys. Lett. **103** , 032401 (2013).
M. Isasa, A. B.-Pinto, F. Golmar, F. Sánchez, L. E. Hueso, J. Fontcuberta, and F. Casanova, Spin Hall magnetoresistance as a probe for surface magnetization, arXiv:1307.1267.
S. R. Marmion, M. Ali, M. McLaren, D. A. Williams, and B. J. Hickey, unpublished.
L. Liu, T. Moriyama, D. C. Ralph, and R. A. Buhrman, Spin-Torque Ferromagnetic Resonance Induced by the Spin Hall Effect, Phys. Rev. Lett. **106**, 036601 (2011).
K. Kondou, H. Sukegawa, S. Mitani, K. Tsukagoshi, and S. Kasai, Evaluation of Spin Hall Angle and Spin Diffusion Length by Using Spin Current-Induced Ferromagnetic Resonance, Appl. Phys. Express **5**, 073002 (2012).
A. Ganguly, K. Kondou, H. Sukegawa, S. Mitani, S. Kasai, Y. Niimi, Y. Otani, and A. Barman, Thickness dependence of spin torque ferromagnetic resonance in $\mathrm{Co_{75}Fe_{25}}$/Pt bilayer films, Appl. Phys. Lett. **104**, 072405 (2014).
A. A. Tulapurkar, Y. Suzuki, A. Fukushima, H. Kubota, H. Maehara, K. Tsunekawa, D. D. Djayaprawira, N. Watanabe, and S. Yuasa, Spin-torque diode effect in magnetic tunnel junctions, Nature **438**, 339 (2005).
J. C. Sankey, P. M. Braganca, A. G. F. Garcia, I. N. Krivorotov, R. A. Buhrman, and D. C. Ralph, Spin-Transfer-Driven Ferromagnetic Resonance of Individual Nanomagnets, Phys. Rev. Lett. **96**, 227601 (2006).
N. Roschewsky, M. Schreier, A. Kamra, F. Schade, K. Ganzhorn, S. Meyer, H. Huebl, S. Geprägs, R. Gross, S. T. B. Goennenwein, Time resolved spin Seebeck effect experiments as a probe of magnon-phonon thermalization time, arXiv:1309.3986.
M. Weiler, J. M. Shaw, H. T. Nembach, and T. J. Silva, Phase-sensitive detection of spin pumping via the ac inverse spin Hall effect, arXiv:1401.6469.
D. Wei, M. Obstbaum, M. Ribow, C. H. Back, and G. Woltersdorf, Spin Hall voltages from a.c. and d.c. spin currents, Nature Commun. **5**, 3768 (2013).
P. Hyde, Lihui Bai, D. M. J. Kumar, B. W. Southern, C.-M. Hu, S. Y. Huang, B. F. Miao, and C. L. Chien, Electrical detection of direct and alternating spin current injected from a ferromagnetic insulator into a ferromagnetic metal, Phys. Rev. B **89**, 180404 (2014).
C. Hahn, G. de Loubens, M. Viret, O. Klein, V. V. Naletov, and J. B. Youssef, Detection of Microwave Spin Pumping Using the Inverse Spin Hall Effect, Phys. Rev. Lett. **111**, 217204 (2013); (E) **112**, 179901 (2014).
H. J. Jiao and G. E.W. Bauer, Spin Backflow and ac Voltage Generation by Spin Pumping and the Inverse Spin Hall Effect, Phys. Rev. Lett. **110**, 217602 (2013).
R. Iguchi, K. Sato, D. Hirobe, S. Daimon, and E. Saitoh, Effect of spin Hall magnetoresistance on spin pumping measurements in insulating magnet/metal systems, Appl. Phys. Express **7**, 013003 (2014).
J. Lotze, H. Hübl, R. Gross, and S. T. B. Gönnenwein, Spin Hall Magnetoimpedance, arXiv:1404.7432.
O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer, S. D. Bader, and A. Hoffmann, Quantifying Spin Hall Angles from Spin Pumping: Experiments and Theory, Phys. Rev. Lett. **104**, 046601 (2010).
A. Brataas, Yu. V. Nazarov, and G. E. W. Bauer, Spin-transport in multi-terminal normal metal-ferromagnet systems with non-collinear magnetizations, Eur. Phys. J. B **22**, 99 (2001).
X. Jia, K. Liu, K. Xia, and G. E. W. Bauer, Spin transfer torque on magnetic insulators, Europhys. Lett. **96**, 17005 (2011).
C. Burrowes, B. Heinrich, B. Kardasz, E. A. Montoya, E. Girt, Y. Sun, Y. Y. Song, and M. Wu, Enhanced spin pumping at yttrium iron garnet/Au interfaces, Appl. Phys. Lett. **100**, 092403 (2012).
M. Weiler, M. Althammer, M. Schreier, J. Lotze, M. Pernpeintner, S. Meyer, H. Huebl, R. Gross, A. Kamra, J. Xiao, Y.-T. Chen, H. Jiao, G.E.W. Bauer and S.T.B. Gönnenwein, Experimental Test of the Spin Mixing Interface Conductivity Concept, Phys. Rev. Lett. **111**, 176601 (2013).
T. Chiba, G. E. W. Bauer, and S. Takahashi, Spin torque transistor revisited, Appl. Phys. Lett. **102**, 192412 (2013).
L. Bai, P. Hyde, Y. S. Gui, and C.-M. Hu, V. Vlaminck, J. E. Pearson, S. D. Bader, and A. Hoffmann, Universal Method for Separating Spin Pumping from Spin Rectification Voltage of Ferromagnetic Resonance, Phys. Rev. Lett. **111**, 217602 (2013).
M. Obstbaum, M. Härtinger, H. G. Bauer, T. Meier, F. Swientek, C. H. Back, and G. Woltersdorf, Inverse spin Hall effect in $\mathrm{Ni_{81}Fe_{19}}$/normal-metal bilayers, Phys. Rev. B **89**, 060407 (2014).
M. Schreier, G. E. W. Bauer, V. Vasyuchka, J. Flipse, K. Uchida, J. Lotze, V. Lauer, A. Chumak, A. Serga, S. Daimon, T. Kikkawa, E. Saitoh, B. J. van Wees, B. Hillebrands, R. Gross, S. T. B. Goennenwein, Sign of inverse spin Hall voltages generated by ferromagnetic resonance and temperature gradients in yttrium iron garnet$|$platinum bilayers, arXiv:1404.3490.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Collapsing Bose-Einstein condensates are rich and complex quantum systems for which quantitative explanation by simple models has proved elusive. We present new experimental data on the collapse of high density [$^{85}$]{} condensates with attractive interactions and find quantitative agreement with the predictions of the Gross-Pitaevskii equation. The collapse data and measurements of the decay of atoms from our condensates allow us to put new limits on the value of the [$^{85}$]{} three-body loss coefficient $K_3$ at small positive and negative scattering lengths.'
author:
- 'P. A. Altin, G. R. Dennis, G. D. McDonald, D. Döring, J. E. Debs,J. D. Close, C. M. Savage and N. P. Robins'
bibliography:
- 'bsnv.bib'
title: 'Bosenova and three-body loss in a [$^{85}$]{} Bose-Einstein condensate'
---
While most experiments with dilute gas Bose-Einstein condensates have employed atomic species with repulsive interactions, it has long been known that interesting and exotic physics manifests in attracting condensates. These include macroscopic quantum tunnelling [@Ueda:1998], the formation of soliton trains and vortex rings [@Cornish:2006; @Lahaye:2008], and the violent collapse and explosion known as the ‘bosenova’ [@Saito:2001; @Saito:2001a; @Donley:2001]. The first evidence for the collapse of attracting Bose-Einstein condensates was found by Sackett and coworkers, who analysed the thermal equilibration of a sample of $^7$Li atoms with negative scattering length that was cooled below the critical temperature [@Sackett:1999]. Soon after this work, condensate collapse was directly observed in pioneering experiments at JILA [@Donley:2001], which revealed a host of interesting dynamics and prompted a surge of theoretical interest [@Saito:2002; @Adhikari:2002; @Santos:2002; @Savage:2003; @Milstein:2003; @Bao:2004; @Adhikari:2004; @Wuster:2005]. More recently, the collapse of dipolar chromium BECs has been observed, displaying the striking $d$-wave symmetry of long-range dipole-dipole interactions in excellent agreement with theory [@Lahaye:2008].
However, while initial mean-field analysis of the JILA experiment using the Gross-Pitaevskii (GP) equation was able to qualitatively account for most of the experimental observations, including the formation of atomic ‘bursts’ and ‘jets’ [@Saito:2002; @Milstein:2003; @Bao:2004; @Adhikari:2004; @Wuster:2005], further investigation exposed a quantitative discrepancy between theory and experiment in the time taken for the condensates to collapse [@Savage:2003]. This was especially puzzling as the short time, low density phase of the experiment is exactly where the GP equation should be an excellent approximation. This disagreement, of about 100%, could not be eliminated by more complex quantum field calculations [@Wuster:2005; @Wuster:2007], and has led to the development of competing models for the collapse mechanism [@Calzetta:2008]. Yet amid the extensive theoretical work on this phenomenon that has continued in recent years, there has been a notable absence of further experimental data, and the discrepancy between theory and the [$^{85}$]{} experiment remains unresolved.
Here we present the first results on this phenomenon from a new [$^{85}$]{} BEC machine [@Altin:2010a], finding good agreement between the measured collapse times and those predicted by a GP model. Although we use the same atom, our experiment has several important differences from the original JILA work. Most notably, our condensates are confined in a purely optical potential, with a homogeneous magnetic bias field applied to manipulate the interatomic interactions. In addition, we have measured condensate collapses with $4\times10^4$ atoms in a tighter trap, which together result in an initial density over an order of magnitude larger than in Ref. [@Donley:2001]. This leads to shorter collapse times and lower values of the critical scattering length, but should not affect the ability of mean-field theory to describe the evolution of the system. It also allows us to investigate three-body recombination rates in a high density regime where they are the dominant source of atom loss.
Our apparatus for producing Bose-Einstein condensates of [$^{85}$]{} with tunable interactions has been described in detail elsewhere [@Altin:2010a]. In brief, we employ sympathetic cooling using [$^{87}$]{} as a refrigerant, initially in a quadrupole-Ioffe configuration magnetic trap and subsequently in a weak, large-volume crossed optical dipole trap. During the final evaporation, a magnetic bias field of 167G is applied to reduce losses due to two-body inelastic collisions [@Roberts:2000; @Altin:2010]. We can create condensates of up to $10^5$ [$^{85}$]{} [[$\left|{F=2, \; m_F=-2}\right\rangle$]{}]{} atoms with a thermal fraction below 10% in a trap with harmonic oscillation frequencies $\omega_{x,y,z} = 2\pi\times\{53,22,27\}$Hz. Condensates form at scattering lengths between $a=+50a_0$ and $a=+200 a_0$, where $a_0$ is the Bohr radius. We determine the scattering length from the applied magnetic bias field using the known parameters of the 155G [$^{85}$]{} Feshbach resonance [@Claussen:2003]. The field is calibrated by addressing radiofrequency transitions between the $m_F$ sublevels of the $F=2$ manifold; the transition frequency is related to the magnetic field strength by the Breit-Rabi equation [@Breit:1931]. The magnetic field can be determined in this way to within 5mG, which near the zero crossing of the scattering length corresponds to an uncertainty in $a$ of $\pm0.2a_0$.
To observe condensate collapse, we follow the procedure of Donley [*et al.*]{}, tuning the atomic interactions using the Feshbach bias magnetic field as shown in Figure \[fig:collapse\](a). First, the scattering length is ramped smoothly from $a=+89.1a_0$ to $a_\text{init}=+0.2a_0$ over 100ms to produce a near-ideal, noninteracting gas. The magnetic field is then increased suddenly ($<100\,\mu$s) to a value at which the interactions are attractive $a_\text{collapse}<0$, and held there for a time $\tau$ before the trap is switched off and the scattering length simultaneously increased to $a=+50a_0$. The condensate is allowed to expand ballistically at this value for 15ms, after which the magnetic field is switched off, changing the scattering length to $a_\text{bg} = -443a_0$. Following a futher 5ms of free evolution, the number of atoms present is determined by absorption imaging.
The number of atoms remaining as a function of evolution time at $a_\text{collapse}=-20a_0$ is shown in Figure \[fig:collapse\](b). In agreement with the original work of Donley [*et al.*]{}, we observe a sudden and delayed onset of atom loss. This is explained by density-dependent three-body recombination; when the interactions are made attractive, the condensate begins to contract slowly and its peak density $n_0$ increases, although not enough to cause significant three-body loss. As the condensate shrinks, however, the contraction accelerates, resulting eventually in a sudden implosion which increases the density by several orders of magnitude. This induces significant recombination losses (the loss rate scales with $n^3$) which ultimately halt the growth in density. The subsequent dynamics include further sporadic local implosions, which effect decay of the atom number in an approximately exponential form. We have observed remnant clouds surviving long after the collapse which contain several times the critical number of atoms $N_\text{cr} \simeq 0.6\,a_\text{ho}/|a|$, where $a_\text{ho} = \sqrt{\hbar/m\omega_\text{ho}}$ is the harmonic oscillator length [@Ruprecht:1995] (the critical number for a condensate with $a=-20a_0$ in our trap is $N_\text{cr} \simeq 1200$; c.f. Figure \[fig:collapse\]). Such configurations have been shown to achieve stability through the formation of mutually repelling bright solitons [@Cornish:2006].
![(a) Manipulation of the scattering length to induce and observe condensate collapse. After a variable evolution time $\tau$, the atoms are released from the optical trap simultaneously with an increase in $a$ from $a_\text{collapse}$ to $+50a_0$. The cloud is allowed to expand at this value for 15ms before the magnetic bias field is switched off. (b) Measured atom number as a function of $\tau$ for $a_\text{collapse}=-20a_0$. The solid line is a fit of the experimental data to equation (\[eqn:collapsefit\]). The atom number remains approximately constant for a time $t_\text{collapse}$, before a sudden onset of loss due to three-body recombination.[]{data-label="fig:collapse"}](figures/collapse.pdf)
The discrepancy between the JILA experiment and theoretical models concerns one of the most elemental characteristics of the bosenova: the time for which the atom number remains constant before the first density implosion – the so-called ‘collapse time’ $t_\text{collapse}$. Although the dynamics after the collapse are predicted to be complex and may exhibit behaviour beyond mean-field effects, the evolution prior to the first implosion should be captured in the mean-field approximation, and is determined almost exclusively by the initial density which is experimentally constrained. In particular, it has been noted that $t_\text{collapse}$ does not depend strongly on the the three-body recombination rate $K_3$ [@Saito:2002; @Savage:2003], which is not well-determined in the vicinity of the Feshbach resonance. Despite this, Gross-Pitaevskii (GP) simulations were found to systematically overestimate the collapse time measured in the JILA system by almost 100%, a discrepancy at the $2\sigma$ level given the experimental uncertainties [@Savage:2003].
As in previous work, we determine the collapse time by fitting plots of the measured atom number versus time to the function $$N(t) = (N_0 - N_f) \exp\left[-\frac{(t-t_\text{collapse})}{\tau_\text{decay}}\right] + N_f \,,
\label{eqn:collapsefit}$$ for $t>t_\text{collapse}$, where $N_0$ and $N_f$ denote the atom number at $t<t_\text{collapse}$ and $t\gg t_\text{collapse}$ respectively. Figure \[fig:collapsetimes\] shows the collapse time determined in this way as a function of $a_\text{collapse}$ for samples of $N_0=4\times10^4$ atoms. As expected, the collapse time is shorter for larger $|a_\text{collapse}|$, as stronger attraction between the condensate atoms results in more rapid contraction. The data are in qualitative agreement with the original experiment of Ref. [@Donley:2001] and later theoretical work.
, with error bars denoting the statistical uncertainty in the fit of equation (\[eqn:collapsefit\]). The solid line is the result of GP simulations for our experimental parameters, and shows good quantitative agreement with the experimental data. The dotted lines represents the variation in the simulated collapse time due to experimental uncertainties in $a_\text{init}$, $a_\text{collapse}$, $\bar\omega$ and $N_0$. The dashed vertical line shows the critical scattering length for collapse at this atom number, below which the condensate’s kinetic energy stabilizes it against implosion.[]{data-label="fig:collapsetimes"}](figures/collapsetimes.pdf)
To ascertain the ability of mean-field theory to *quantitatively* reproduce our experimental data, we have performed numerical simulations for the parameters of our system using the Gross-Pitaevskii equation for the condensate wave function $\Psi$: $$i\hbar \frac{\partial\Psi}{\partial t} = \left[ -\frac{\hbar^2}{2m} \nabla^2 + V_\text{trap} + \frac{4\pi\hbar^2a}{m} \left|\Psi\right|^2 - i\frac{\hbar}{2} K_3 \left|\Psi\right|^4 \right] \Psi \,,
\label{eqn:gpe}$$ where $V_\text{trap}$ is the confining potential. Three-body recombination is modelled by the phenomenological inclusion of an imaginary loss term proportional to the three-body loss rate coefficient $K_3$ (which differs by a Bose statistical factor of $3!$ from the coefficient for noncondensed atoms). This term leads to loss proportional to the cube of the atomic density: $$\frac{\partial}{\partial t} \int |\Psi|^2 d{\mathbf{r}} = -\int K_3 |\Psi|^6 d{\mathbf{r}} \,,$$ and entails the assumption that the products of recombination collisions leave the trap without interacting with the remaining atoms. As three-body processes dominate at the high densities relevant to this experiment [@Saito:2002], we do not include the effect of two-body inelastic collisions. To make the computation tractable, in integrating equation (\[eqn:gpe\]) we assume a cylindrically symmetric trap with oscillation frequencies $\omega_{z,\rho} = 2\pi\times\{53,24\}$Hz, such that the mean trap frequency $\bar\omega$ matches that of our crossed dipole trap (the collapse time has been found to be relatively robust to asymmetry in the trapping potential [@Savage:2003]). The simulation includes the 100ms magnetic field ramp from $a=+89.1a_0$ to $a_\text{init}=+0.2a_0$, but we neglect the expansion of the condensate in our simulation, as the density spikes which trigger the recombination losses cease once the interactions are made repulsive.
The results of this simulation for $N_0 = 4\times10^4$ and $a_\text{init} = +0.2a_0$ are overlaid with the experimental data in Figure \[fig:collapsetimes\] (solid line). For these simulations the three-body loss coefficient was scaled with $a_\text{collapse}$ as $K_3 = 8\times10^{-14}\,a^2$ cm$^4$/s [@Bao:2004]. The dotted lines show the variation in the simulated collapse time due to the combined experimental uncertainties in the initial scattering length and the trap frequencies, as well as run-to-run number fluctuations of 20%. The simulations show good quantitative agreement with the experimental data.
It should be noted that the 100ms ramp of the scattering length from $a=+89.1a_0$ to $a=+0.2a_0$ is not truly adiabatic. Our GP simulations show that the ramp excites breathing mode oscillations, despite the duration of the ramp exceeding the mean trap oscillation period by a factor of 3. The oscillation is predominantly along the weak trapping axes, and has an amplitude of approximately 10% of the condensate radius. It occurs because, although $a$ is varied smoothly, the condensate size does not depend linearly on the scattering length – in the Thomas-Fermi limit, the radius scales as $a^{1/5}$. This excitation accelerates the contraction of the condensate, decreasing $t_\text{collapse}$ by approximately 15%. The effect is included in the simulations shown in Figure \[fig:collapsetimes\]. It could be reduced by tailoring the magnetic field ramp to ensure that the condensate radius decreases smoothly. Our simulations show that ramping $a^{1/4}$ smoothly over 100ms reduces the breathing mode oscillations to below 1%, and causes the collapse times to be indistinguishable from those for a condensate that is in the ground state immediately prior to the collapse.
We now turn our attention to the possible systematics which may affect the agreement between theory and experiment. The source of the largest experimental uncertainty in our system is the oscillation frequencies of the crossed dipole trap, which vary during the evaporation to BEC as the intensity of the trapping laser is reduced. For technical reasons, we cannot directly measure the trap frequencies at the end of the evaporation. Instead, we make several measurements at higher intensities, which we fit to an analytic model of the dipole potential including the effect of gravity in the vertical direction. The model is further constrained by knowledge of the intensity $I_0$ at which gravity overcomes the dipole potential and the trap vanishes. Due to the strong dependence of the vertical trap frequency on the laser intensity near $I_0$, and the variation in the intensity itself, we estimate an uncertainty in $\bar\omega$ of 10%, predominantly in the vertical direction. As the peak density of a noninteracting condensate scales as $\bar\omega^{3/4}$, this corresponds to an estimated uncertainty in $n_0$ of 7%, which in turn produces an uncertainty in the simulated collapse time of approximately 15%. This is not sufficient to explain the inconsistency between our results and the JILA experiment.
We must also consider the possibility of a systematic error in our determination of atom number. $N_0$ is calculated using the theoretical optical cross-section by integrating the optical depth of an absorption image. We image the atoms on resonance with circularly polarized light and apply a small bias magnetic field along the imaging direction to provide a quantization axis. The calculation therefore makes use of the resonant cross-section and saturation intensity of the cycling transition ${{\ensuremath{\left|{F=3, \; m_F=\pm3}\right\rangle}}}\rightarrow{{\ensuremath{\left|{F^\prime=4, \; m_{F^\prime}=\pm4}\right\rangle}}}$. As a result, our measured value of $N_0$ is a lower bound: any errors in the polarization or detuning of the imaging light, or in the alignment of the quantization field, will reduce the measured atom number. We estimate the uncertainty in $N_0$ due to these effects to be less than 5%. Furthermore, if the atom number were undercounted then correcting for this would decrease the simulated collapse times, as a higher initial density speeds up the contraction. This effect therefore also cannot explain the disparity between our results and the original experiment, for which GP simulations *over*estimated the collapse times.
Although $t_\text{collapse}$ is only weakly dependent on the three-body loss coefficient $K_3$, the shape of the loss curves is affected by varying this parameter. The values of $K_3$ used in simulations of the original JILA experiment ranged from $K_3 = 2\times10^{-28}$ cm$^6$/s [@Saito:2002] to $K_3 = 2\times10^{-26}$ cm$^6$/s [@Savage:2003]. Several authors also considered a relationship between the loss coefficient and the scattering length of the form $K_3 \sim |a|^2$ for $a<0$ [@Moerdijk:1996; @Adhikari:2004], with Bao [*et al.*]{} deducing $K_3 = 2.68\times10^{-13}\,a^2$ cm$^4$/s [@Bao:2004]. We find that these values cannot reproduce the shape of our measured loss curves.
Figure \[fig:k3collapse\] shows the results of GP simulations using values of $K_3$ between $5\times10^{-27}$ cm$^6$/s and $5\times10^{-29}$ cm$^6$/s overlaid with experimental data for $a_\text{collapse}=-8.4a_0$. At higher loss rates, the high initial density of our sample causes significant loss during the contraction of the condensate in the simulation during the moments leading up to the collapse. In fact, for $K_3 > 10^{-27}$ cm$^6$/s this initial loss is so great that there is no sudden implosion of the condensate and no discernible elbow in the loss curve [^1]. In order to obtain the abrupt onset of loss that we observe in the experiment, a three-body loss rate of $K_3 \leq 5\times10^{-29}$ cm$^6$/s at $a=-8.4a_0$ is required. From a similar analysis of the $a_\text{collapse}=-20a_0$ data shown in Figure \[fig:collapse\](b) we find $K_3 \leq 1\times10^{-28}$ cm$^6$/s at that scattering length. These limits are more than an order of magnitude below most of the values used to simulate the original experiment. Assuming a scaling with $|a|^2$, they imply $K_3 \lesssim 1\times10^{-14} \, a^2$ cm$^4$/s. In this regime, loss after the initial implosion is caused by intermittent local density spikes between which three-body loss is negligible. This was first predicted by Saito and Ueda [@Saito:2001] even before the JILA experiment. These discrete implosions result in the numerous plateaus apparent in the simulated loss curve, although the scatter in our experimental data – caused primarily by run-to-run fluctuations in atom number – is too large to observe these directly.
![Comparison of experimental and simulated collapse data for $a_\text{collapse}=-8.4a_0$. The data points show the measured atom number $N$ (normalized to $N_0$) as a function of evolution time $\tau$ at $a<0$, and the lines represent GP simulation results with different values of the three-body loss coefficient $K_3$. A value of $K_3 \leq 5\times10^{-29}$ cm$^6$/s is necessary to replicate the sudden onset of loss detected in the experiment.[]{data-label="fig:k3collapse"}](figures/k3collapse.pdf)
We have investigated the inelastic loss rates further by measuring the depletion of our condensates over time with positive scattering lengths, at which the condensates are stable. The rate at which atoms are lost due to two- and three-body inelastic collisions depends on the density profile of the condensate. In the limit that $a\rightarrow0$, the density is given by the modulus squared of the ground state harmonic oscillator wavefunction, and the loss rate equation $\dot{N}/N =-\sum_i K_i \langle n^{i-1} \rangle$ becomes: $$\begin{aligned}
\dot{N} = -N/\tau - \eta_2 K_2 N^2 - \eta_3 K_3 N^3 \,,
\label{eqn:losseqnNI}\end{aligned}$$ where $\tau$ represents the one-body loss rate, $\eta_2 = (2\pi a_\text{ho}^2)^{-3/2}$ and $\eta_3 = (\sqrt{3}\pi a_\text{ho}^2)^{-3}$. In the Thomas-Fermi limit $Na/a_\text{ho} \gg 1$, the condensate density takes on the shape of the confining potential and the loss rate equation evaluates to: $$\begin{aligned}
\dot{N} = -N/\tau - \gamma_2 K_2 N^{7/5} - \gamma_3 K_3 N^{9/5} \,,
\label{eqn:losseqnTF}\end{aligned}$$ with $\gamma_2 = 15^{2/5}/[14\pi a^{3/5}a_\text{ho}^{12/5}]$ and $\gamma_3 = 5^{4/5}/[56\pi^2 3^{1/5} a^{6/5}a_\text{ho}^{24/5}]$. It should be noted that these expressions are valid only when the loss rate is small compared with the trap frequencies $K_i \langle n^{i-1} \rangle \ll \bar\omega$, so that the atomic density profile does not change significantly.
Figure \[fig:losscurves\] shows the number of atoms remaining as a function of time in condensates with $a=0$ and $a=+37.6a_0$. The lines plot the best-fit solutions to (\[eqn:losseqnNI\]) and (\[eqn:losseqnTF\]) respectively, assuming that the loss is entirely due to two-body (solid) or three-body (dashed) inelastic collisions. It is difficult to separate the contributions of two- and three-body loss purely from the shape of the decay curve, as has been noted in previous work [@Roberts:2000]. Nonetheless, attributing all of the measured loss to two- or three-body processes allows us to place an upper limit on the value of $K_2$ and $K_3$ at these scattering lengths. Figure \[fig:lossrates\](a) shows these upper bounds for scattering lengths between $0$ and $+100a_0$. In our system, $Na/a_\text{ho} \simeq a/a_0$ and we use the Thomas-Fermi approximation (\[eqn:losseqnTF\]) except at $a=0$. The error bars represent the statistical uncertainties in the fits; we assign an additional systematic error of 10% to incorporate the uncertainty in $\bar\omega$.
![Measurements of inelastic losses in [$^{85}$]{} condensates. The data points show the atom number as a function of hold time in the optical trap for condensates with $a=0$ and $a=+37.6a_0$. The solid lines are fits of the solutions of (\[eqn:losseqnNI\]) and (\[eqn:losseqnTF\]) to the experimental data, assuming $K_3=0$ (solid) and $K_2=0$ (dashed). Although the contributions of two- and three-body processes cannot be distinguished in this manner, these fits may be used to place upper bounds on the values of $K_2$ and $K_3$.[]{data-label="fig:losscurves"}](figures/losscurves.pdf)
![(a) Upper bounds on $K_2$ (open circles) and $K_3$ (filled circles) as a function of scattering length, calculated from fits to the solutions of (\[eqn:losseqnNI\]) and (\[eqn:losseqnTF\]). The error bars represent statistical uncertainties. (b) Locus of two- and three-body loss coefficients for which the solution to (\[eqn:losseqnNI\]) fits the experimental data for $a=0$ shown in Figure \[fig:losscurves\]. The $x$ and $y$ intercepts correspond to the upper bounds shown in (a). Assuming $K_2 \simeq 1.2\times10^{-14}$ cm$^3$/s [@Roberts:2000], the data suggest a three-body loss coefficient $K_3 \leq 10^{-30}$ cm$^6$/s.[]{data-label="fig:lossrates"}](figures/lossrates.pdf)
Theoretical calculations suggest that the recombination rate should vary strongly with the two-body elastic scattering cross-section, with several authors predicting a universal $K_3 \sim a^4$ scaling in the zero-temperature limit [@Fedichev:1996; @Nielsen:1999; @Esry:1999]. Our observations are consistent with a strong suppression of the recombination rate at the zero crossing of the $s$-wave scattering length, with the measured upper bound $K_3 \leq (3.9\pm0.7)\times10^{-29}$ cm$^6$/s at $a=0$ an order of magnitude below that for $a>+50a_0$, and more than three orders of magnitude below the loss rate far from the Feshbach resonance, $K_3 = 7\times10^{-26}$ cm$^6$/s [@Roberts:2000].
We can combine this latest data with previous measurements of the two-body loss rate to further constrain the three-body recombination coefficient. Figure \[fig:lossrates\](b) shows the locus of possible $K_2,K_3$ values for which the solution to (\[eqn:losseqnNI\]) best fits our experimental loss curve at $a=0$. In Ref. [@Roberts:2000], Roberts [*et al.*]{} measured $K^\text{nc}_2 \simeq 2.4\times10^{-14}$ cm$^3$/s for thermal clouds in the vicinity of $a=0$, corresponding to a value of $K_2 \simeq 1.2\times10^{-14}$ cm$^3$/s for condensed atoms. This matches our measured upper bound of $K_2 \leq (1.2\pm0.2)\times10^{-14}$ cm$^3$/s. Coupled with this result, our data is consistent with a value of the three-body loss coefficient $K_3 \leq 10^{-30}$ cm$^6$/s. Ref. [@Esry:1999] predicts $K_3 \simeq 5\times10^{-32}$ cm$^6$/s at $a=0$. In comparison, the three-body loss coefficient for [$^{87}$]{} is $K_3 = 6\times10^{-30}$ cm$^6$/s [@Burt:1997].
In conclusion, we have presented new experimental data on the collapse of [$^{85}$]{} Bose-Einstein condensates with attractive interactions in an optical dipole trap. Our results qualitatively match those of the original JILA bosenova experiment, but in addition agree quantitatively with GP simulations. We find that a lower value of the three-body loss coefficient $K_3$ than was used in simulating the original experiment is needed to reproduce the sudden onset of loss that we observe. We have also analysed the decay of atoms from our condensates and thereby placed further constraints on the three-body loss coefficient $K_3$ at small positive scattering lengths. We expect that this work will inform future experimental and theoretical investigations of this rich quantum system.
We thank J. Hope, M. Johnsson, S. Szigeti and M. Hush for helpful discussions. This work was supported by the Australian Research Council Centre of Excellence for Quantum-Atom Optics.
[^1]: To aid comparison of the elbow in the loss curve, the trap frequencies in the simulations have been adjusted to give the observed collapse time in Figure \[fig:k3collapse\]. Nonetheless, as can be seen from Figure \[fig:collapsetimes\], the experimental collapse time at $a_\text{collapse}=-8.4a_0$ is slightly below – although within the error bars of – the simulated collapse time.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'We report the first high spatial resolution near-infrared imaging of the Vela pulsar in the $K_s$ band obtained with the new adaptive optics system recently mounted on the Gemini-South telescope. For the first time, we have firmly detected the pulsar in this band with $K_s$ $\approx$ 218, and have resolved in detail an extended feature barely detected previously in the immediate vicinity of the pulsar in the $J_sH$ bands. The pulsar $K_s$ flux is fully consistent with the extension of the flat optical spectrum of the pulsar towards the infrared and does not confirm the strong infrared flux excess in the pulsar emission suggested earlier by the low spatial resolution data. The extended feature is about two times brighter than the pulsar and is likely associated with its X-ray counter-jet. It extends $\sim$ 2 southwards of the pulsar along the X-ray counter-jet and shows knot-like structures and a red spectrum.'
author:
- 'D. Zyuzin, Yu. Shibanov and A. Danilenko'
- 'R. E. Mennickent'
- 'S. Zharikov'
bibliography:
- 'vela002.bib'
title: 'The Vela pulsar and its likely counter-jet in the $K_s$ band '
---
Introduction
============
After the young Crab and B0540$-$69 pulsars with the visual magnitudes of 165 and 226 the 11 kyr old Vela pulsar with the magnitude of 236 is the third brightest in the optical among all isolated neutron stars (NSs) known. The relative brightness has allowed to perform detailed optical studies including the successful timing, spectral, and polarization observations. Similar to the Crab pulsar, Vela has an almost flat spectrum from the near infrared (IR) to the ultra violet (UV) [@mignani2007AsAp]. The similarity has been unexpectedly broken by recent *Spitzer* observations in the mid-IR, where the Vela pulsar has shown a strong flux excess over its optical-UV spectrum extension towards IR [@danilenko2011MNRAS]. This is completely different from Crab whose mid-IR and optical-UV spectra are described by a single power law [@sandberg2009AsAp]. Similar excesses have been detected only for two magnetars 4U 0142+61 and 1E 2259+586 [@wang2006Natur; @kaplan2009ApJ], where they were interpreted as an emission from hypothetical fall-back X-ray irradiated discs around those NSs. However, Vela is not a magnetar. It is an ordinary rotation powered pulsar emitting from radio to gamma-rays and powering, like Crab, a bright (in X-rays) torus-like pulsar wind nebula (PWN) with polar jets [@helfand2001ApJ]. The fall-back disk survival around such an active pulsar with a strong relativistic particle wind appears to be problematic [see, e.g., @jones2007MNRAS]. Two other possibilities have been suggested to explain the excess [@danilenko2011MNRAS]: a complicated distribution function of emitting particles in the NS magnetosphere; a possible contamination of the pulsar flux by an unresolved PWN structure. The first one looks very unusual, while the second is supported by the presence of a faint nebulosity 2 away from the pulsar tentatively detected in the near-IR VLT/ISAAC $J_s$ and $H$ images [@shibanov2003AsAp]. The images had a higher spatial resolution than the *Spitzer* ones. It is projected onto the origin of the pulsar X-ray counter-jet and has a red colour, which could, in principle, explain the mid-IR excess [@danilenko2011MNRAS]. To check that, one needs a higher spatial resolution imaging in the near-IR.
The Vela pulsar has never been observed in the $K$ band. Motivated by this and by the excess problem, we have carried out a high spatial resolution imaging of the Vela field in the $K_s$ band with the new generation of the adaptive optics (AO) system recently mounted on the Gemini-South telescope [@carrasco2012SPIE]. The observations and data reduction are described in Sect. 2, the results are presented in Sect. 3 and discussed in Sect. 4.
Gemini-South data
=================
Observations, data reduction, and calibration
---------------------------------------------
The Vela pulsar was observed on January 30, 2013 in the $K_s$ band with the Gemini Multi-Conjugate Adaptive Optics System (GeMS) and its near-infrared imager, the Gemini South Adaptive Optics Imager (GSAOI), mounted on the Gemini-South telescope[^1]. The observations were carried out in the service mode during the GeMS + GSAOI System Verification science program. The GSAOI science array is a 2$\times$2 mosaic of four Rockwell HAWAII-2RG detectors forming a 4080 $\times$ 4080 pixel focal plane with a field of view of $85\asec \times 85\asec$ and a pixel scale of 002. Each detector also contains a programmable On-Detector Guide Window (ODGW) which can provide a tip-tilt information for a combination of up to four natural guide stars in GeMS. Three natural optical guide stars, NOMAD 0448-0138807 ($R$ $\sim$ 156), 0448-0138766 (142), and 0448-0138794 (143) were used for the CANOPUS tip-tilt wave front sensor (CWFS), which is the AO bench of GeMS. The latter star was also the ODGW infrared guide star. The pulsar was exposed on chip 1 of the GSAOI array and we focus below on the data obtained from this chip. The observing conditions were photometric with seeing $\la$ 055.
We have obtained 19 dithered 100 s science exposures with $\sim$10 offsets at an airmass of $\sim$ 1.04. Gemini GeMS + GSAOI twilight flat-fields were taken and used to create a master flat-field frame. Data reduction, linearity correction, sky subtraction, flat-fielding and bad pixel correction were performed using the `IRAF` `gemini.gsaoi` tool. The effective mean seeing, defined as the full width at half maximum (FWHM) of a stellar profile, on the chip 1 AO corrected frames was in the range of 007–009, depending on the individual exposure and a source position on the frame. The best seeing concentrated within $\sim$03 region around the pulsar, in accordance with expectations for our AO observational setup. Differences of instrumental magnitudes of several field stars around the pulsar in different exposures were less than 1%. The reduced frames were aligned to a single reference frame of the best image quality and summed[^2]. The resulting effective mean seeing and integration time were 0085 and 1900 s, respectively. The magnitude zero-point of $C_{K_s}$ = 2553(4) (instrumental fluxes in ADUs) for the chip 1 was derived based on the photometric standards (9132, 9136, 9144, 9146) from @persson1998AJ obtained on the same night. The atmospheric extinction of 0065 in $K_s$ was taken from the GSAOI web-site.
(145,52)(0,0) (-10,0) []{} (74,0) []{}
Astrometry
----------
We have performed the relative astrometry of the $K_s$-band resulting image using the VLT/ISAAC $J_s$-band image obtained twelve years ago by @shibanov2003AsAp as a reference frame. Nine unsaturated isolated stars located in the pulsar vicinity were selected as the reference points, and used in the `IRAF` `geomap/ccmap` tools to obtain the plate solutions. Formal [*rms*]{} uncertainties of the astrometric fit were $\la$ 0014 with maximum residuals of $\la$ 0032 for both coordinates. Taking into account the uncertainties of the reference star positions, which were $\la$ 8 mas and $\la$ 0.2 mas for the VLT and Gemini images, respectively, a conservative 1$\sigma$ referencing uncertainty of the Gemini image with respect to the VLT ones is $\la$ 002. Adopting the VLT absolute astrometric image referencing from @shibanov2003AsAp, the absolute astrometric uncertainties of the Gemini $K_s$-band image is $\approx$ 021.
Results
=======
Identification of the pulsar and a likely counterpart of its counter-jet
------------------------------------------------------------------------
A fragment of the resulting $K_s$ image demonstrating the Vela pulsar vicinity is shown in the [*right panel*]{} of Fig. \[fig:gsao-im\]. It is compared with a similar fragment obtained in the $H$ band with the VLT/ISAAC by @shibanov2003AsAp ([*left panel*]{}). We consider a point-like object marked by “+” and detected in $K_s$ at $\sim$ 20$\sigma$ significance as the Vela pulsar counterpart candidate. In the $K_s$ image, there is also an extended feature adjacent to the counterpart candidate from the south. The feature is also barely resolved in the $H$ band, which means that it is not an artifact.
The position of the suggested pulsar counterpart in the $K_s$ image is shifted by 068 $\pm$ 005 with respect to the pulsar position at the epoch of the VLT observations (2001). The shift implies a proper motion $\mu$ = 56 $\pm$ 4 mas yr$^{-1}$ with positional angle PA = 2960 $\pm$ 43, which is consistent with the pulsar proper motion $\mu$ = 58 $\pm$ 0.1 mas yr$^{-1}$ and PA = 3010 $\pm$ 18 based on radio observations [@dodson2003ApJ]. This is a strong evidence that the point-like object detected in the $K_s$ image is the pulsar. None other point-like object in the pulsar vicinity demonstrates any significant proper motion.
The extended feature was named “counter-jet?” by @shibanov2003AsAp, assuming its possible association with the counter-jet of the Vela torus-like X-ray PWN. The Gemini AO observations confirm this structure with a higher significance and reveal important details of its morphology. It extends southwards from the pulsar by about 2 along the X-ray counter-jet axis with PA $\approx$ 130 [@helfand2001ApJ] marked by the dashed line in the [*right*]{} panel of Fig. \[fig:gsao-im\]. This supports the association of the extended feature with the Vela X-ray counter-jet. In X-rays the counter-jet extends to a much larger distance up to 1. In the near-IR it is likely that we see only its origin near the pulsar. The $K_s$-band counter-jet demonstrates a non-uniform morphology with several knots. The brightest knot is in the center of the feature. It is marked in Fig. \[fig:gsao-im\] and is reminiscent of the knot structure 06 away from the Crab pulsar, which is also projected onto the counter-jet origin of the Crab PWN [@hester1995ApJ]. The Vela IR counter-jet is not detected in the optical range with the *HST* [@shibanov2003AsAp] whose spatial resolution is comparable to that of Gemini/GSAOI. This implies that it has a very red spectrum. Also it becomes evident, that the putative counter-jet and the pulsar cannot be resolved from each other in the significantly lower spatial resolution *Spitzer* mid-IR images analyzed by @danilenko2011MNRAS.
In the Gemini image we do not find any other extended structure that could be identified with other parts of the Vela PWN. In particular, we do not see the inner arc whose marginal detection in $J_s$ was discussed by @shibanov2003AsAp. In addition to the o1–o3 point-like objects from the pulsar neighborhood considered earlier [@shibanov2003AsAp; @danilenko2011MNRAS] we find a red point-like source “o4” detected east of the pulsar at $\sim$5$\sigma$ confidence only in the $K_s$ image. It is significantly fainter than the pulsar and was partially overlapped with it at the VLT observation epoch.
[lllll]{} $\lambda_{eff}$ & &\
(band), & Mag.$^a$ & $\log$ F$_\nu$, & Mag. & $\log$ F$_\nu$,\
$\mu$m & & $\mu$Jy & & $\mu$Jy\
\
1.23($J_s$)$^b$ & 22.71(10) & 0.14(4) & 22.66(10) & 0.16(4)\
1.65($H$)$^b$ & 22.04(18) & 0.21(7) & 22.00(18) & 0.23(7)\
& [**21.76(6)**]{} & [**0.12(2)**]{} & [**21.74(6)**]{} & [**0.13(2)**]{}\
\
1.23($J_s$) & 24.14(37) & $-$0.43(15) & 24.09(37) & $-$0.41(15)\
1.65($H$) & 22.35(10) & 0.09(8) & 22.31(10) & 0.11(8)\
& [**21.16(8)**]{} & [**0.36(3)**]{} & [**21.14(8)**]{} & [**0.37(3)**]{}\
\
\
3.6$^c$ & 18.48(22) & 1.06(9) & 18.48(22) & 1.06(9)\
4.5$^c$ & $\ga$ 16.84 & $\la$ 1.52 & $\ga$ 16.84 & $\la$ 1.52\
5.8$^c$ & 16.38(27) & 1.51(11) & 16.38(27) & 1.51(11)\
8.0$^c$ & $\ga$ 15.58 & $\la$ 1.70 & $\ga$ 15.58 & $\la$ 1.70\
\[t:flux\]
Photometry
----------
For stellar-like object photometry we used a circular aperture with four pixels radius, comparable to the effective seeing of the $K_s$ image. The correction for finite aperture, derived from the point spread function (PSF) of bright stars located close to the pulsar, is $\delta m$ = 0.64(2). In this case, uncertainties include the PSF variation over the selected image section. For the counter-jet we used a polygon aperture shown in Fig. \[fig:gsao-im\], although the final results depend weakly on the specific aperture shape and background region parameters.
We have also remeasured the spatially integrated brightnesses of the counter-jet in the $J_s$ and $H$ bands. The photometry was performed on the VLT/ISAAC pulsar subtracted images [@shibanov2003AsAp Fig. 1]. Our results differ significantly from those presented by @danilenko2011MNRAS. The reason is that the authors used mean surface brightnesses from @shibanov2003AsAp which were accidentally overestimated by 1.94 $mag~arcsec^{-2}$ in both bands. After this correction the fluxes from @shibanov2003AsAp are consistent with our ones. All measured fluxes were de-reddened using $E_{B-V}$ = 0.055(5) and $R_V$ = 3.1 [@shibanov2003AsAp]. The results are summarized in Table \[t:flux\].
For completeness, $K_s$ magnitudes of o1, o2, o3, o4 field objects are 2248(8), 2053(4), 1877(4), and 232(2), respectively. The derived 3$\sigma$ detection limit for a point-like object for a 03 aperture (corresponding to $\ga$ 90% of a point-like source flux) centered at some position in the pulsar vicinity free from any sources is 236.
Spectra of the Vela pulsar and its “counter-jet”
------------------------------------------------
Using our photometry results, in Fig. \[fig:multi\_spec\] we show an upgrade of the IR-UV spectra of the pulsar and its likely counter-jet compiled recently by @danilenko2011MNRAS. One sees that the $K_s$ flux of the pulsar agrees well with the extrapolation of its flat power law UV-optical spectrum toward the longer wavelengths. This makes the Vela pulsar spectral energy distribution (SED) similar to that of the younger Crab [@sandberg2009AsAp].
At the same time, the counter-jet spatially integrated $K_s$ flux (shown by red symbol in Fig. \[fig:multi\_spec\]) is about twice as high as the pulsar flux. The counter-jet $J_sHK_s$ fluxes demonstrate a very steep power law SED with a spectral index of about three. The Spitzer fluxes are compatible with the long wavelength extrapolation of this SED, suggesting a common nature of the near-IR and mid-IR emission. We have checked that the essentially non-thermal, likely synchrotron, spectrum of the counter-jet is different from blackbody-like SEDs of the nearby stellar objects o1, o2, o3. This means that the extended feature can hardly be just a combination of faint blended stars.
Discussion
==========
The Gemini-South ground-based observations with the new generation of the AO system have provided us with the superb image quality comparable to that of the *HST*. This allowed us to firmly detect, for the first time, the Vela pulsar in the $K_s$ band. The measured pulsar flux is consistent with the extrapolation of the pulsar optical-UV spectrum towards the IR, which shows that the spectrum remains flat in this range and does not demonstrate any excess suggested early by the lower spatial resolution IR data. The AO observations also enabled us to confirm and resolve the feature extended immediately behind the pulsar. The elongation of the extended feature in the direction of the X-ray counter-jet and in the opposite direction of the pulsar proper motion suggests that it can be associated with the X-ray counter-jet. This is supported by the compact and relatively bright knot-like structure within the feature (Fig. \[fig:gsao-im\]), which is reminiscent of the well known optical-near-IR knot within the south-east jet of the Crab pulsar.
The likely non-thermal spectrum of the Vela IR counter-jet is also similar to the spectrum of the Crab knot, which follows the power law with a spectral index of 1 and is responsible for an apparent excess of the Crab-pulsar mid-IR fluxes observed with *Spitzer* [@sandberg2009AsAp]. The spectrum of the Vela counter-jet is even steeper resulting in a much stronger apparent excess of the Vela flux in the mid-IR. Our data practically rule out alternative interpretations of the mid-IR excess discussed by @danilenko2011MNRAS, such as the fall-back disk or the complicated emitting particle distribution function in the pulsar magnetosphere. The Vela counter-jet spectrum is also consistent with typically red spatially integrated optical-IR spectra of PWNe [@zharikov2013AsAp].
The Crab knot [@sandberg2009AsAp], and Vela jets [@pavlov2001ApJb] are known to demonstrate a high temporal variability even on a week scale. The current data do not allow one to infer whether the Vela IR counter-jet is variable and move together with the pulsar. Detection of such variability would thus serve as a strong evidence for its PWN nature. This can also help to discard possible alternative interpretations, such as a Vela supernova remnant filament or a background galaxy at a cosmological distance. Some marginal evidence of the feature variability has been reported by @shibanov2003AsAp based on the VLT near-IR observations. Forthcoming VLT/NACO and *Spitzer* data will hopefully find out whether the detected feature is variable, clarify its spectrum and real nature. Observations at longer wavelengths are important to confirm power law spectrum of the feature. To detect other parts of the Vela PWN, deeper near-IR observations are necessary.
We are grateful to the Gemini AO system team for performing excellent observations and to the anonymous referee for useful comments. The work was partially supported by the Russian Foundation for Basic Research (grants 11-02-00253 and 13-02-12017-ofi-m), RF Presidential Program (Grant NSh 4035.2012.2), and by CONACYT 151858 project. REM acknowledges support by the BASAL Centro de Astrofisica y Tecnologias Afines (CATA) PFB–06/2007.
[*Facilities:*]{} .
[^1]: http://www.gemini.edu/sciops/instruments/gsaoi/documents\
http://www.gemini.edu/sciops/instruments/gems/documents
[^2]: final image located on\
http://www.ioffe.ru/astro/NSG/obs/index.html
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Synthesising 3D facial motion from speech is a crucial problem manifesting in a multitude of applications such as computer games and movies. Recently proposed methods tackle this problem in controlled conditions of speech. In this paper, we introduce the first methodology for 3D facial motion synthesis from speech captured in arbitrary recording conditions (“in-the-wild”) and independent of the speaker. For our purposes, we captured 4D sequences of people uttering $500$ words, contained in the Lip Reading Words (LRW) a publicly available large-scale in-the-wild dataset, and built a set of 3D blendshapes appropriate for speech. We correlate the 3D shape parameters of the speech blendshapes to the LRW audio samples by means of a novel time-warping technique, named Deep Canonical Attentional Warping (DCAW), that can simultaneously learn hierarchical non-linear representations and a warping path in an end-to-end manner. We thoroughly evaluate our proposed methods, and show the ability of a deep learning model to synthesise 3D facial motion in handling different speakers and continuous speech signals in uncontrolled conditions.'
author:
- |
Panagiotis Tzirakis$^1$, Athanasios Papaioannou$^{1,2}$, Alexander Lattas$^1$, Michail Tarasiou$^1$,\
Bj[ö]{}rn Schuller$^{1,3}$, Stefanos Zafeiriou$^{1,4}$\
$^1$ Department of Computing, Imperial College London, UK\
$^2$ Great Ormond Street Institute of Child Health, University College London, UK\
$^3$ ZD.B. Chair of Embedded Intelligence for Health Care and Wellbeing, University of Augsburg, Germany\
$^4$ Center for Machine Vision and Signal Analysis, University of Oulu, Finland
title: 'Synthesising 3D Facial Motion from “In-the-Wild" Speech'
---
{width="100.00000%"} \[fig:Blendshapes2\]
Introduction
============
Synthesis of 3D talking faces is very crucial to many applications including but not limited to computer games, movies post-production (e.g., dubbing), talking faces in virtual reality applications, etc. Currently, the highest quality 3D face synthesis is performed by using facial capture rigs which make use of markers or sensors. Recently, machine learning methods [@edwards2016jali; @Karras3d; @phamend], and in particular deep neural networks, have been used in order to train systems that reconstruct the 3D facial geometry of talking faces directly from audio sources. Nevertheless, till now the proposed methods are person or rig-specific [@visemenet] [^1], hence do not generalise to arbitrary audio sequences. In this paper, we present the first methodology for estimation of the 3D facial motion related to speech directly from raw audio stream.
Estimating the 3D facial motion directly from raw audio samples captured in arbitrary recording conditions is an ill-pose problem since a great number of people uttering a large amount of diverse words need to be captured in 4D (i.e. 3D geometry in time). Even though many efforts have been performed towards collecting 4D expressive faces [@cheng20184dfab; @yin20063d; @yin2008high] there is a lack of datasets with talking 4D faces (i.e., 3D speech in time). One such dataset has been proposed by Marshall et al. [@marshall20154d], which captured four people in dyadic interaction (17mins in total). A limited number of words have been captured in [@cheng20184dfab] for biometric application, nevertheless the data are not publicly available. Hence, it is very difficult to train a generic method for 3D facial motion reconstruction from audio streams. The most closely related work to ours is by Phamend et al. [@phamend], which used a statistical blendshape model, trained on facial expressions. As we show, these blendshapes cannot model adequately 3D facial motion related to speech.
In this paper, we make the first, to the best of our knowledge, comprehensive effort to estimate 3D facial motion from arbitrary audio streams. To this end, we first capture 4D sequences of people uttering $500$ words, contained in an in-the-wild dataset named Lip Reading Words (LRW). After registering the 3D meshes with an adaptive template approach, we learn 3D blendshapes for the speech, which we make publicly available. Each 3D mesh can be parameterised as a set of 3D shape parameters through these 3D blendshapes. Employing a novel time-warping algorithm, named Deep Canonical Attentional Warping (DCAW), we align the speech, that we have 3D ground-truth on, with the corresponding “in-the-wild” speech signals in LRW, and propagate the 3D shape parameters, creating the “in-the-wild” LRW-3D dataset. We train a deep learning model on this dataset and show the ability of the model to predict 3D face motion for speech captured under uncontrolled conditions. Fig. \[fig:process\] shows the pipeline of our approach. In summary, the contributions of this work are the following:
- We collect a 4D dataset of people uttering $500$ words and learn the first statistical blendshape model for speech which we provide publicly available.
- In order to train accurate blendshapes, we propose an adaptive shape template method to accelerate the convergence of registration algorithms and achieve a better final shape correspondence.
- We propose Deep Canonical Attentional Warping (DCAW), a method which learns hierarchical non-linear representations and temporal alignment of two audio signals in an end-to-end manner. Using DCAW we create LRW-3D, and we make publicly available the aligned 3D shape parameters.
- Finally, we train a speech to 3D deep facial motion model that can operate in nearly real-time, and independently of the speaker in uncontrolled conditions of speech.
{width="8.4cm" height="3.3cm"} \[fig:process\]
Related Work {#related_work}
============
There are several traditional approaches [@fan2015photo; @katsamanis2009face; @mattheyses2015audiovisual; @sako2000hmm; @ohman1999using; @salvi2009synface; @wang2007assembling; @xie2007realistic] that exploit audio signal for 2D or 3D facial animation. More recent approaches utilise deep neural networks (DNN) for this task. These can be categorised in two main groups: linguistic-driven and audio-driven approaches.
**Linguistic-driven approaches**. Language-based methods take advantage of the mapping between phonemes and their visual counterpart visemes. For example, Edwards et al. [@edwards2016jali] proposed the JAw and LIp (JALI) model, a two-dimensional space that represents the jaw and lip movements of a facial animation based on psycholinguistic considerations. The main disadvantage of their study is the need for the speech signal, its text transcript and their alignment to create the facial animation. In another study, Taylor et al. [@taylor2012dynamic] first proposed generating dynamic units for visual speech for realistic visual speech animation. In a more recent study, the authors [@taylor2017deep] initially transcribe the speech signal to phoneme labels, which are then fed to a deep fully connected network to predict person-specific shape and appearance parameters obtained by Active Appearance Model (AAM). A main limitation of this work is the need for speech to phoneme labels conversion.
**Audio-driven approaches**. Audio-based methods drive facial animations using only audio cues. Zhou et al. [@visemenet] utilises audio features for an automatic and near real-time animation by driving a JALI face-rig. The authors propose VisemeNet, a three stage deep learning model based on Long Short-Term Memory (LSTM) networks that are fed with audio features. The first two stages comprises of extracting phonemes and landmarks, while the third extracts visemes. In a different study, Karras et al. [@Karras3d] proposed a deep convolutional neural network for 3D facial animation using auto-correlation audio features. The network models unknown variation of the audio cues by learning a small dimensional vector. Its output is the entire face and is trained using a three-way loss function that ensures temporal stability and correct responses under animation. The main limitation of their method is that the trained network is person-specific and as such it cannot generalise well to different speakers. Pham et al. [@phamend] used the same network architecture as Karras et al. but with melspectrograms as input. Their model was trained to predict the rotation and expression blending parameters extracted by a 3D face tracker. Suwajanakorn et al. [@suwajanakorn2017synthesizing] used audio features to synthesize videos of Obama focusing on the mouth region and taking the rest of the head and torso from stock footage. However, their method is person-specific with tens of hours of audio signal, and requires heavy hand-engineered work.
In comparison with the aforementioned methods, our approach is able to provide 3D face motion estimation not only in uncontrolled speech conditions but also is speaker-independent, i.e., independent of the speaker. On top of that, we are the first to propose a statistical blendshape model appropriate for speech.
Building Speech Blendshapes Model {#sec_speech_blendshapes}
=================================
In this section, we describe the pipeline for building the speech blendshapes model. The process consists of: (a) collecting a dataset of 3D meshes with various people uttering a set of words (Sec. \[data\_acquisition\]), (b) registering all the meshes to a common template (Sec. \[registration\]), and (c) building a statistical model on the registered meshes (Sec. \[speech\_blendshapes\]).
Data Acquisition {#data_acquisition}
----------------
To construct a set of blendshapes appropriate for speech we needed to capture 4D sequences (i.e. 3D geometry in time) of people talking. The choice of the utterance spoken by our participants is driven by our goal to train a model that can operate under unconstrained audio conditions (in-the-wild), and is speaker independent. To this purpose, we utilise the $500$ words contained in the publicly available Lip Reading Words (LRW) in-the-wild dataset [@chung2016lip] which contains almost $1000$ videos per word, summing to approximately $450,000$ videos. These videos were captured from TV broadcasts (e.g. news or interviews) in uncontrolled environments and contain $1000$ speakers, making it an excellent fit for our purposes.
We used the DI4D dynamic system[^2] to capture and build 4D faces. This system consists of six cameras (two pairs of stereo cameras and one pair of texture cameras, $30$FPS, $1200\times1600$). Before every recording, a calibration was necessary and was performed by utilising a $10\times10$, 20mm checkerboard. Two 4-lamp fluorescent lights were placed on each side to provide consistent and uniform lights. One microphone is used to capture the audio signal in $44,1$kHz sampling rate.
We captured 2 native and 2 non-native English speakers reading out the $500$ words from the aforementioned dataset. We should note that the choice of the non-native speakers is to add extra variability to the mouth region when we develop our blendshapes. In total, the whole process took 20min for each participant, and, approximately, $20,000$ 3D meshes were acquired for each individual (equivalent of $660$s of recording).
Registration
------------
**Automatic Annotation.** Before processing our 3D captured meshes, we need to re-parameterise them such that all the meshes have the same number of vertices joined into a common triangulation. Two approaches exist that perform such task and they differ on the space the registration is performed. The first method performs the dense registration in the 2D space, namely the UV-space [@patel20093d; @cosker2011facs]. The second method registers directly (i.e. in the 3D space) the mesh and the template [@amberg2007optimal; @myronenko2010point].
In this paper, we follow the latter approach as UV-based correspondence approaches introduce non-linearities into the process and require an extra step for rasterizing the UV image [@booth2018large]. To this end, we perform the registration in the 3D space between a neutral 3D mesh and the mean shape of Large Scale Facial Model (LSFM) [@booth2018large]. The registration is performed by utilising Non-rigid Iterative Closest Point (NICP), which has as a prerequisite the template and the mesh to be close in terms of euclidean distance.
The high number of 3D meshes do not allow us to manually annotate 3D facial landmarks. To automate the process we utilise a sparse alignment method proposed by Booth et al. [@booth20163d] that can automatically compute sparse annotations on the 3D meshes. More specifically, for each mesh we apply the face detection and alignment framework proposed by Zhang et al. [@zhang2016joint] on its corresponding 2D texture image, where the correspondence to the 3D coorditanes are known, to robustly locate a set of $68$ sparse annotations (landmarks) [@zafeiriou20183d] in the 2D space. Exploiting the correspondence between the texture image and the 3D mesh, we get the corresponding 3D positions of the landmarks. **Adaptive Template and Dense Registration.** The majority of NICP shape registration methods use the same 3D template shape to deform all 3D meshes of a dataset [@booth2018large]. Even though this approach can be sufficient for large datasets with neutral shapes, it is not the case when there is a large variation in terms of expressions and lip movements in each mesh. In our captured 3D meshes the position and shape of the mouth change frequently so if a single template shape is used important parts of it would not be close to the corresponding parts of 3D meshes. Hence, the registered results would have visible errors and inaccurate correspondences.
{width="\textwidth"} \[fig:registration\_pipeline\]
To alleviate from this problem, we propose an adaptive template approach where for each 3D mesh in our dataset we adapt the original template using sparse shape information (i.e. $68$ landmarks). In particular, by leveraging the expression blendshapes created by Cheng et al. [@cheng20184dfab], we compute the 3D shape parameters for each mesh through linear regression between the landmarks of the neutral shape and the landmarks of the reconstructed instance as
$$\mathbf{c}_o = \arg\min_{\mathbf{c}} || \boldsymbol l_{n} - \mathbf{A}(\mathbf{x}_{n}+\mathbf{U}_s\mathbf{c})||^2_F
\label{Eq:Min}$$
where $\boldsymbol l_{n} \in \mathbb{R}^{3m}$ is a vector with $m$ landmarks of the neutral shape, $\mathbf{A} \in \mathbb{R}^{3m \times 3n}$ is an indicator matrix, indicating the position of the 3D landmarks on the reconstructed mesh, $\mathbf{x}_{n} \in \mathbb{R}^{3n}$ is the neutral registered shape, $\mathbf{U}_s \in \mathbb{R}^{3n \times q}$ is a matrix with the $q$ blend shapes and $\mathbf{c} \in \mathbb{R}^{q}$ is the 3D shape parameters vector.
After the calculation of the 3D shape parameters $\mathbf{c}_o$ that minimise the expression in Eq. \[Eq:Min\], we can adapt the neutral template to the current mesh: $\mathbf{x}_{adapt} = \mathbf{x}_n + \mathbf{U}_s\mathbf{c}_o$. Finally, NICP can be performed between the adaptive template and the mesh. We should note that even though the blendshapes ($\mathbf{U}_s$) describe various expressions and not speech, they give a good prior for our template. The registration process is illustrated Fig. \[fig:registration\_pipeline\].
Creating Speech Blendshapes {#speech_blendshapes}
---------------------------
Our final step is to build a set blendshapes appropriate for speech. We start by subtracting the neutral mesh from each registered mesh in the sequence, creating vectors of differences, i.e., $\mathbf{d} = \mathbf{x}_{n} - \mathbf{x}_{adapt} \in \mathbb{R}^{3n}$. After we stack these vectors into a matrix $\mathbf{D} = [\mathbf{d}_1, \dots, \mathbf{d}_T] \in \mathbb{R}^{3n \times T}$, we apply Principal Component Analysis (PCA) to identify the deformation components $\mathbf{U}_b$. We keep $28$ blendshapes corresponding to $99.9\%$ of the total variance in the sequence. Hence, a new shape instance can be generated: $\mathbf{x}_{new} = \mathbf{x}_n + \mathbf{U}_b \boldsymbol \lambda$, where $\boldsymbol \lambda$ are the 3D shape parameters of our model. Finally, for each mesh in our sequence, we compute the 3D shape parameters that constitute our ground truth.
Lip Reading Words in 3D (LRW-3D) {#dcaw_3d_propagation}
================================
With the construction of our speech-driven statistical model, we can now use the speech signal of our participant to propagate the 3D shape parameters to the speech signals of the LRW dataset. Our process starts by segmenting our participants’ speech signal to the $500$ words, which is accomplished in a semi-automatic manner. More particularly, we first utilise the approach proposed by Elsner et al. [@elsner2017speech] to segment our speech signal. This step is crucial to our pipeline as a single faulty segmentation can result in almost a thousand false samples in our dataset, which can lead to ill-generalisable models. To this purpose, we listened each segment, and, when required, we manually adjusted them.
Deep Canonical Attentional Warping (DCAW)
-----------------------------------------
To accurately propagate the 3D shape parameters of our participant to the LRW dataset, we need to eliminate any temporal variations arising in the data. Hence, we compute a temporal alignment between the signal of each word uttered by our participant and the corresponding signals of the LRW. To compute this alignment, we propose *Deep Canonical Attentional Warping (DCAW)*, a novel method that can maximally correlate two data sequences (or views) and find a temporal alignment in an *end-to-end* manner. We leverage deep recurrent convolutional neural networks to spatially transform the raw speech signals, and utilise attention mechanism for the alignment.
The *attentional warping* is performed by computing attention weights between each feature frame of the one view (source) with all the features from the other view (target). We should point out that our data are monotonic and as such we utilise a monotonic attention mechanism [@raffel2017online]. Mathematically, given two views $\mathbf{X}_k \in \mathbb{R}^{d_k \times T_k}$, and their features $\textbf{h}_i^k$, $i = \{1, 2, ..., T_k\}$, where $d_k$ and $T_k$ are the dimensionality and length of view $k \in \{1, 2\}$, respectively, the attention between each target feature and the source ones is computed as follows:
$$\begin{aligned}
\label{eq:attention1}
\alpha_{\textbf{h}_i^1, \textbf{h}_j^2} = \frac{\exp(\operatorname{score}(\textbf{h}_i^1, \textbf{h}_j^2))}{\sum_{k=1}^{T_2} \exp(\operatorname{score}(\textbf{h}_i^1, \textbf{h}_k^2))},\end{aligned}$$
where $\exp$ is the exponential function, and $\operatorname{score}$ can be any attention function, such as Bahdanau [@bahdanau2014neural] or Luong [@luong2015effective]. For our purposes, we use the Bahdanau score:
$$\begin{aligned}
\label{eq:attention2}
\operatorname{score}(\mathbf{h}_i^1, \mathbf{h}_j^2) = \mathbf{v}_i^T \tanh(\mathbf{W}_t\mathbf{h}_i^1 + \mathbf{W}_s\mathbf{h}_j^2),\end{aligned}$$
where $\mathbf{v}_i^T$, $\mathbf{W}_t$ and $\mathbf{W}_s$ are learnable parameters of the model.
The outcome of Eq. \[eq:attention1\], is the attentional matrix $\mathbf{A} \in \mathbb{R}^{T_1 \times T_2}$ that includes the weights between the two views. We use this matrix to warp the target features with the source ones by multiplying it with the source features, i.e., $\mathbf{H}^1\mathbf{A}$, where $\mathbf{H}^1 \in \mathbb{R}^{d \times T_1}$ is the feature matrix containing all the features of the first view.
The alignment between the two views is found such that the features between the views are maximally correlated. This optimisation is formulated as a least-square problem: $$\begin{array}{rl}
\arg\min_{\theta_1, \theta_2} &|| f_1(\mathbf{X}_1;\theta_1) - f_2(\mathbf{X}_2;\theta_2) ||_F^2
\\ \\
\mbox{subject to}\, &f_1(\mathbf{X}_1;\theta_1)f_1(\mathbf{X}_1;\theta_1)^T = \mathbf{I} \\
&f_2(\mathbf{X}_2;\theta_2)f_2(\mathbf{X}_2;\theta_2)^T = \mathbf{I}\\
&f_1(\mathbf{X}_1;\theta_1)f_2(\mathbf{X}_2;\theta_2)^T = \mathbf{D} \\
&f_1(\mathbf{X}_1;\theta_1)\mathbf{1} = f_2(\mathbf{X}_2;\theta_2)\mathbf{1} = \mathbf{0},
\end{array}
\label{Projection_matrix}$$\
where $f_k(\mathbf{X}_k;\theta_k)$ with $k \in \{1, 2\}$ represents the output of the two neural networks with parameters $\theta_k$ and input $\mathbf{X}_k$, respectively, $\mathbf{D}$ is a diagonal matrix, and $\mathbf{1}$ is an appropriate dimensionality vector of all ones.
We find the optimal parameters for each network with the use of backpropagation. Our problem is a variant of Deep Canonical Correlation Analysis (DCCA) [@andrew2013deep], and as such the optimal objective value can be computed as the sum of the $k$ largest singular values of $\mathbf{K}_{DCAW} = \mathbf{\Sigma}_{11}^{-1/2} \mathbf{\Sigma}_{12} \mathbf{\Sigma}_{22}^{-1/2}$, where $\mathbf{\Sigma}_{ij} = \frac{1}{T_2 - 1} f_i(\mathbf{X}_i;\theta_i) \mathbf{C}_T f_j(\mathbf{X}_j;\theta_j)^T$ and $\mathbf{C}_T = \mathbf{I} - \frac{1}{T_2} \mathbf{1} \mathbf{1}^T$ is the centering matrix. The optimal objective is found by maximising the nuclear norm $||\mathbf{K}_{DCAW} ||_{*} = tr(\sqrt{\mathbf{K}\mathbf{K}^T})$, i.e.,
$$\begin{aligned}
\label{eq:on2}
\arg\max_{\theta_1, \theta_2} ||\mathbf{K}_{DCAW} ||_{*}\end{aligned}$$
We use gradient ascent to optimise Eq. \[eq:on2\]. Since the gradient cannot be computed analytically we use the subgradient [@bach2008consistency] by computing the singular value decomposition of $\mathbf{K}_{DCAW} = \mathbf{U} \mathbf{S} \mathbf{V}^T$, then the subgradient for the last layer of the network can be defined as follows
$$\begin{aligned}
\label{eq:attenti}
\mathbf{L}_+ &= \mathbf{\Sigma}_{11}^{-1/2} \mathbf{U} \mathbf{V}^T \mathbf{\Sigma}_{22}^{-1/2} f_2(\mathbf{X}_2;\theta_2)\mathbf{C}_T \\
\mathbf{L}_- &= \mathbf{\Sigma}_{11}^{-1/2} \mathbf{U} \mathbf{S} \mathbf{U}^T \mathbf{\Sigma}_{11}^{-1/2} f_1(\mathbf{X}_1;\theta_1)\mathbf{C}_T \\
&\frac{\vartheta ||\mathbf{K}_{DCAW} ||_{*}}{\vartheta f_1(\mathbf{X}_1;\theta_1)} = \frac{1}{T_2 - 1} ( \mathbf{L}_+ - \mathbf{L}_-).\end{aligned}$$
Finally, we should point out that DCAW can be extended to handle multiple data sequences by utilising an objective similar to Multi-set CCA, i.e., $$\begin{aligned}
\sum_{i,j} ||K_{DCAW}^{i,j}||_*.\end{aligned}$$
Word Alignment
--------------
We can now use DCAW to compute an alignment path between the speech signal of words uttered by our participant and the corresponding signals of the LRW dataset. To this end, we train a recurrent convolutional neural network for each word (i.e. $500$ networks - see Sec. \[training\] for topology) by fixing one of the views to our participant’s speech signal and the other to the corresponding speech signals of LRW, and get the alignment path between them.
Utilising the alignment path of the speech signals, we propagate the 3D shape parameters computed for our participant to the LRW dataset. In the case where an audio frame of the LRW is aligned to multiple frames from our dataset, the mean of the 3D shape parameters of these frames is computed and assigned as the ground truth of that frame. If the opposite holds, namely, an audio frame from our speech signal is aligned to multiple audio frames of the LRW, then the same ground truth is assigned to all LRW audio frames.
By transferring our 3D shape parameters to the LRW dataset, we create a large word-level dataset in-the-wild for 3D dense shape estimation from speech. This allows us to train models that can generalise to every speaker, and at the same time to in-the-wild speech signals.
Training
========
Three aspects are relevant to our training: (i) the input representation, (ii) the network topology, and (iii) the objective function utilised for training and evaluating the model. We describe in detail in the rest of the section.
**Input Representation.** All audio signals have sampling rate at $44.1$kHz, and after we remove the DC offset, we normalise its volume to $0$dB, namely, using the full $[-1, 1]$ range. No other pre-processing step takes place.
We use mel-spectograms as our input represenation of the audio signal. This representation is appropriate for our task because it is derived by approximating the frequencies perceived by the human cochlea [@rabiner2011theory]. Thus, the information is similar to the perceived human hearing.
For each visual frame, we derive mel-spectrograms in an audio window of length $400$ms so that we can take into account co-articulation effects in the signal. For each audio window, we compute $128$ mel-frequency parameters utilising a window of length $20$ms ($882$ samples) with $10$ms ($441$ samples) overlap. Hence, a 2D representation is formed of size $41 \times 128$. By calculating its first and second temporal derivatives, and place all three 2D representations in a different channel, we form a $41 \times 128 \times 3$ representation, which is the input to our convolutional recurrent neural network.
**Network Topology.** Our deep neural network topology is inspired by Karras et al. [@Karras3d], and is comprised of four parts: (a) frequency extractor, where features are extracted vertically, namely, exploiting the frequency domain of the input representation, with kernel and stride size of 3 and 2, respectively, (b) short term temporal extractor, where features are extracted horizontally, namely, from the temporal domain of the extracted representation of the previous step, with kernel and stride size of 3 and 2, respectively, (c) a non-linear transformer, that non-linearly transform of the convolutional extracted features 128 dimensionality, and (d) a long term temporal extractor represented with a recurrent neural network of 1-layer LSTM cell of 128 dimensions, which captures the long term temporal dynamics in the data. The last layer is a fully connected that produces the 3D shape parameters of our model. The filter size for all convolution layers is set to 64.
**Objective Function.** Most of the studies in the literature use as objective function the Mean Squared Error (MSE). However, we propose to use an objective function that is based on the Concordance Correlation Coefficient ($\rho_c$), which is also used as our evaluation metric. The correlation coefficient evaluates the agreement level between the predictions and the ground truth by scaling their correlation coefficient with their mean square difference. More particularly, for each shape parameter $i$ we define the concordance loss $\mathcal{L}_{c}^i$ between the ground truth $\mathbf{x}$ and the prediction $\mathbf{y}$ as follows:
$$\begin{aligned}
\nonumber
\mathcal{L}_{c}^i
&= 1 - \rho_c = 1 - \dfrac{2 \sigma_{xy}^2}{\sigma_x^2 + \sigma_y^2 + (\mu_x - \mu_y)^2}, \\
\label{eq:ccc_loss}\end{aligned}$$
where $\mu_{x} = \mathbb{E}(\mathbf x)$, ${\mu_{y} = \mathbb{E}(\mathbf y)}, {\sigma_x^2 = \mbox{var}(\mathbf x)}, \sigma_y^2 = \mbox{var}(\mathbf y)$ and $\sigma^2_{xy} = \mbox{cov}(\mathbf x, \mathbf y)$.
For our purposes we train our networks to simultaneously predict all $28$ 3D shape parameters. We should note that each shape parameter explains different variability percentage in our data, and hence contributes differently to the final 3D reconstructed mesh. We take this fact into account and we add as weight $w_i$ to each shape parameter concordance loss $\mathcal{L}_c^i$, the variability percentage the shape parameter represents. Our overall loss function is defined as $\mathcal{L} = \sum_{i=1}^{28} w_i \mathcal{L}_c^i$.
Experiments
===========
We perform extensive evaluation of our proposed methods by: (i) comparing the DCAW with the current state-of-the-art method for maximally correlating two views and finding an alignment (Sec. \[dcaw\]), (ii) comparing the proposed speech blendshapes with the blendshapes of the FaceWarehouse (Sec. \[blendshapes\_comparison\]), and (iii) validating our approach for 3D face motion from “in-the-wild” speech (Sec. \[audio\_experiments\]).
Experimental Setup {#hyperpar}
------------------
The hyperparameters of the models are kept the same throughout all of the experiments. As our optimisation function, we use the Adam optimiser [@adam] with $\beta_1 = 0.9$, $\beta_2 = 0.99$, and a initial learning rate of $5\times10^{-4}$, with batch size set to $50$. Finally, the initialisation of the feature extraction network was performed following He et al. [@he2015delving], whereas our recurrent network weights are initialised following Glorot based initialisation [@glorot2010understanding]. We should note that zero padding was used to samples that do not match the maximum sequence length of the batch. We discard the zero-padded frames that do not belong to the sample by applying a mask.
Deep Canonical Attentional Warping {#dcaw}
----------------------------------
We compare our proposed DCAW with the state-of-the-art method for representation learning and temporal warping, the Deep Canonical Temporal Warping (DCTW) [@trigeorgis2018deep]. The performance of the two methods is evaluated on two datasets: (a) the MMI Facial Expression Dataset [@pantic2005web], which contains more than $2900$ videos of $75$ different subjects, each performing a particular combination of Action Units (i.e., facial muscle activations). We predict the AU12 and utilise the same approach and network architecture (with a recurrent network on top) as in [@trigeorgis2018deep]. (b) We use the LRW dataset to perform template matching, where one of the views is a speech signal of a word and the other is a speech signal of a sentence containing, in a random location, the word. The methods need to accurately find the word in the sentence. For our purposes, we use $10$ randomly chosen words with $100$ samples.
The performance measure is the alignment error introduced in [@zhou2012generalized]. More particularly, given $m$ sequences, each algorithm infers a warping path, i.e., $P_{alg} = [p_{alg}^1, ..., p_{alg}^m] \in \mathbb{R}^{l_{alg} \times m}$, and the alignment error is computed with the ground truth path $P_{grd} = [p_{grd}^1, ..., p_{grd}^m] \in \mathbb{R}^{l_{grd} \times m}$ as follows:
$$\text{Error} = \frac{dist(P_{grd}, P_{alg}) + dist(P_{alg}, P_{grd})}{l_{grd} + l_{alg}},$$
where $dist(P_{grd}, P_{alg}) = \sum_{i=1}^{l_1} min_{j=1}^{l_2} ||p^{(i)}_1 - p^{(j)}_2||_2$.
Table \[dcaw\_result\] depicts the results for both experiments. Our method outperforms DCTW in both of them, and for the LRW one we find the result to be statistically significant ($a < 0.05$). These results validate our choice of using DCAW for aligning our participant’s speech signal with the ones from the LRW dataset.
Dataset DCTW DCAW
--------- ------ ----------
MMI 0.59 **0.61**
LRW 0.64 **0.72**
: Results (wrt the mean alignment error) of the DCAW and DCTW methods on the MMI and LRW datasets.[]{data-label="dcaw_result"}
Blendshapes Comparison {#blendshapes_comparison}
----------------------
We compare quantitatively and qualitatively the proposed blendshapes with the FaceWarehouse ones, that were used by Pham et al. [@phamend], by testing the generalisation capacity of the blendshapes to represent unseen 3D facial meshes.
More particularly, we compute the error as the per-vertex Euclidean distance between every mesh of the test subject (i.e. not included in the training process) and its corresponding projection to the subspace defined by the blendshapes. An average value is computed over all vertices. For the whole sequence the mean error of the FareWarehouse is 1.29, and our proposed blendshapes is 0.87. Fig. \[fig:Blendshapes\_comparison\] depicts the error in a sequence of $2,550$ frames. It is clear that the proposed blendshapes outperform the FaceWarehouse ones by a high margin. Finally, for qualitative purposes, we show three samples of the original 3D facial shapes and how they are reconstructed by the FaceWarehouse and the proposed blendshapes.
![Average Euclidean distance for a sequence of frames between the FaceWarehouse and the proposed blendshapes. - Best viewed in colour.[]{data-label="fig:Blendshapes_comparison"}](drawing_1c-crop.png){width="8cm"}
To further demonstrate the generalisation capability of our speech blendshapes, we show our model’s prediction and the ground truth 3D shape parameters on three individuals. Fig. \[fig:qualitative\_results\], shows the results.
![Depicting the ground truth 3D shape parameters (left column) and the predictions (right column) for three individuals. - Best viewed in colour.[]{data-label="fig:qualitative_results"}](qualitative_results.png){height="8cm"}
Dataset Methodology $\mu_{28}$ $c_1$ $c_2$ $c_3$ $c_4$ $c_5$
--------- ------------------- ----------------- ----------------- ----------------- ----------------- ----------------- -----------------
Speaker Ind. $.621$ ($.679$) $.643$ ($.712$) $.607$ ($.684$) $.563$ ($.596$) $.652$ ($.683$) $.582$ ($.641$)
Word/Speaker Ind. $.502$ ($.554$) $.536$ ($.556$) $.582$ ($.618$) $.557$ ($.624$) $.395$ ($.405$) $.411$ ($.426$)
LRS Continuous Speech $.463$ $.482$ $.414$ $.443$ $.475$ $.504$
Audio Experiments {#audio_experiments}
-----------------
We start the validity of our method by performing two kind of experiments using the LRW-3D dataset: (i) *speaker-independent*, where we measure the performance of our model on different speakers that pronounce the same words, and (ii) *word- and speaker-independent*, where we evaluate our model on different speakers that pronounce different words. For both experiments our test set is comprised of one of the participant’s audio signal with the corresponding 3D shape parameters, which are excluded from the training process. Finally, we measure the performance of our model on continuous speech signal.
**Speaker-Independent.** In our first experiment we test the performance of the model when the same words are uttered by different individuals. To this end, we split the $1000$ speakers of the dataset to $900$ for training, and the rest $100$ for validation. Hence, our training set is comprised of $370,000$ samples, and our validation set of the rest $80,000$. Table \[word\_results\] depicts the results of the mean, and the first five shape parameters in terms of the $\rho_c$ metric. The results indicate the validity of our method to generalise to every speaker and in uncontrolled conditions.
**Word- and Speaker-Independent.** Considering the high performance of the model for different speakers, in this experiment we test its performance for different words and speakers. Hence, we split our dataset to a training set that contains $450$ words and $900$ speakers, and the validation set contains the rest of the $50$ words and $100$ speakers. In total, the training set contains approximately $355,000$ samples, and the validation set approximately $95,000$ samples. We should point out that the words that comprise the validation set were chosen such that they contain the same phonemes as the ones in the training set.
To improve the generalisation capacity and reduce overfitting of our model, we perform a random time-segmentation of our training samples. More particularly, each sample in the training batch is randomly segmented from its both ends but always keeping at least $50$% of the frames of the original sample. This is particularly beneficial to our recurrent network architectures as now the temporal dynamics in the training set vary.
Table \[word\_results\] depicts the results of the mean, and the first five shape parameters in terms of the $\rho_c$ metric. The performance drops compared to the previous experiment as now the temporal information in the validation set is different from the training one. However, this does not limit the capacity of the model to be able to accurately reconstruct the 3D meshes.
**Sentence-level Experiments.** We also test the performance of our model in continuous speech signals. More specifically, we captured a native speaker (different than the previous experiments) pronouncing $50$ sentences (3 to 8sec long), taken from the Lip Reading Sentences (LRS) in-the-wild dataset [@chung2017lip], and extracted his 3D parameters. After the extraction of mel-spectrograms from the raw waveform, we feed them to the model and estimate its test performance in continuous speech signals. Table \[word\_results\] depicts the results of the mean, and the first five shape parameters in terms of the $\rho_c$ metric. We observe that the performance of our model remains also high in this experiment. We should point out that our model is trained with short temporal dynamics ($10$ to $30$ frames long), and as such it cannot accurately predict 3D shape parameters for longer sequences such as sentences. We tackle this difficulty by splitting the speech signal of the sentences to sequences of $15$ frames long and feed them separately our model. We apply a temporal filter on the predictions of our model, to remove temporal discontinuities added by the LSTM. In the supplementary material videos are provided that show the effectiveness of our method.
Conclusions
===========
We presented a methodology for constructing 3D facial meshes from speech cues captured in uncontrolled conditions. More particularly, we learned a statistical blendshape model by capturing 4D sequences of people uttering $500$ words selected from the Lip Reading Words (LRW) in-the-wild dataset. To align the words uttered from our participant with the words of the LRW, we proposed Deep Canonical Attentional Warping (DCAW), a novel method that simultaneously learns deep representations and an alignment path between two sequences. We thoroughly experimented with our proposed methods and showed the ability of a trained deep learning model on the create LRW-3D to generalise to different speakers and in uncontrolled conditions of speech.
For future work we intend to incorporate expression in our blendshapes such that accurate emotional speech can be obtained. In addition, we will test our model on languages different than English. On top of that, we will capture individuals talking on different languages to expand the generalisation capacity of our blendshapes.
[^1]: In [@Karras3d] even though the method was trained on a single person, the authors claim that they used the method on other people and the results were meaningful. The above contradicts the current practices in machine learning which require large and diverse sets for good generalisation. Because we could not reproduce their experiments, we contacted the authors but received no reply.
[^2]: http://www.di3d.com
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'The development of robot control programs is a complex task. Many robots are different in their electrical and mechanical structure which is also reflected in the software. Specific robot software environments support the program development, but are mainly text-based and usually applied by experts in the field with profound knowledge of the target robot. This paper presents a graphical programming environment which aims to ease the development of robot control programs. In contrast to existing graphical robot programming environments, our approach focuses on the composition of parallel action sequences. The developed environment allows to schedule independent robot actions on parallel execution lines and provides mechanism to avoid side-effects of parallel actions. The developed environment is platform-independent and based on the model-driven paradigm. The feasibility of our approach is shown by the application of the sequencer to a simulated service robot and a robot for educational purpose.'
author:
-
title: '**A Platform-independent Programming Environment for Robot Control** '
---
INTRODUCTION {#sec:intro}
============
The development of robot control programs is a difficult, error-prone and complex exercise. At first, the developer needs a good understanding about the task which the robot should fulfill. This includes knowledge about the environment in which the robot is embedded and an idea of the desired behavior of the robot. Second, platform-specific knowledge about the robot platform itself is needed. For instance, a vacuum cleaning robot is small and usually equipped with a cleaning device, whereas a service robot is tall and equipped with a manipulator to reach high cupboards and to grasp objects. This diversity in the mechanical and electrical structure reflects not only in the appearance and hardware design, but also in the software system. A robot application developer must be aware of this diversity in order to develop control programs in an optimal way. However, such a robot platform-oriented development of robot control programs leads (very often) to non-reusable control programs. Though, reuse of robot control programs is desirable, because the control program describes the logic of a specific task (e.g. an office delivery task) which could be applied on different robot platforms and reused (or partly reused) in similar tasks. To overcome the dependency of a specific platform (e.g. operating systems or even programming languages) the Software Engineering community introduced the concept of Platform-independent Models (PIM) and Platform-specific Models (PSM) within the model-driven software development paradigm. Here, models play a vital role to capture and abstract generic aspects from domain-specific aspects. Recently, the model-driven paradigm has been adopted by the robotics community. In [@c3] Baer et al. introduced a platform independent modeling language providing modeling elements for communication and collaboration infrastructures. The approach has been applied successfully for cooperative soccer playing robots. In Schlegel et al. [@c4] a component-based robot software development approach has been applied to the model-driven paradigm in order to generate a wide variety of target-specific source code (e.g. for embedded controllers) without modifying the platform-independent model. Another model-driven approach is described in [@c7]. Here, Alonso et al. introduced a platform-independent component-model which captures certain aspects of a robotic software system in three distinct views, namely behavioral, structural, and component. The approach has been applied to generate Ada source-code for a Cartesian robot.\
However, to the best of our knowledge, the model-driven paradigm has been so far not applied to the development of platform-independent robot control programs, even so this domain could benefit from it. In particular, the development of robot control programs where concurrency is crucial (e.g. in behavior-based and hybrid control approaches as [@c8]) is an error-prone task, because actions that may technically run in parallel can have side-effects. For instance, a manipulator can grasp an object while a camera processes images to detect objects. In such a situation the manipulator may block the field of view and the camera may not find the appropriate object. These pitfalls make the entry-level in robot programming very high and complicates the fast creation of robot control programs. Therefore, supporting tools and frameworks to ease the development are needed.\
One promising graphical programming environment with parallelism support is the Urbi Studio by the company Gostai [@d14]. Urbi Studio provides a graphical state-machine based programming environment which is bound to their custom programming language urbiscript. In addition to state-machine based programming Urbi Studio provides a time-based graphical sequencer to initialize and set variables over time – a specialty provided by the framework. However, the environment is tightly bound to the Urbi execution engine which allows to monitor the program execution in the graphical environment. The company aims to make all robots compatible, which leads into a concrete implementation and C++ wrapping of the robot framework to the Urbi framework. Another graphical programming environment which also provides the possibility to describe parallel sequences is the Nao Choregraphe from Aldebaran Robotics [@d15]. It is delivered with their Nao robot. This environment allows flow diagram based programming of Nao robots and provides a time-based sequencer to create motor control commands. The environment is able to execute code on the Nao robot either in Urbi or the Python language.\
In this paper we present a programming environment for graphical sequencing of robot actions. The approach is based on the model-driven paradigm. Exchangeable domain-specific languages are used to generate platform-specific source code for a wide variety of robots without changing the control logic. The paper is structured as follows. In Section \[sec:problemstatement\] the robot control approach, namely the sequencing of concurrent actions is introduced. Our model-driven approach to achieve platform-independence is described in Section \[sec:approach\]. The application of the graphical sequencer is shown in Section \[sec:usecases\] by the application to a simulated service robot and a a robot for educational purpose. In Section \[sec:conclusion\] our approach will be discussed and we conclude with some lessons learned.
ACTION SEQUENCING FOR ROBOT CONTROL {#sec:problemstatement}
===================================
We are interested in the programming of robot tasks. For instance, the autonomous delivery of documents in an office environment. A task $\mathbf{T} = \{\mathbf{a_{1}}, \mathbf{...}, \mathbf{a_{n}}\}$ is composed of a set of finite actions (e.g. grasping an object, following a person, or moving to a location) from the set of actions $\mathbf{A}$ and $\mathbf{T} \subseteq \mathbf{A}$. Each action $\mathbf{a} \in \mathbf{A}$ is a tuple of the following form $$\mathbf{a} = \langle \mathbf{n}, \mathbf{C} \rangle$$ where $\mathbf{n}$ is an unique name of the action and $\mathbf{C}$ is a set of execution constraints $\mathbf{e}$. Furthermore, a global data space $\mathbf{D}$ is used to store global variables which are needed, modified and shared by the actions. An execution constraint is formally defined as $$\mathbf{e} = \langle \mathbf{\Lambda}, \mathbf{a} \rangle \quad \mathbf{a} \in \mathbf{A}$$
where $\mathbf{\Lambda}$ is an execution operator which constraints the execution of the action $\mathbf{a}$ in a temporal manner. In our case $\mathbf{\Lambda}$ describes the order of actions. Hence, our robot control approach may be described as a dependency graph [@d16] with nodes as actions and edges as execution constraints describing the predecessor and successor relationship between actions. Figure 1 shows an example of such an execution order. Here, the action **A**, **B**, and **C** are executed initially, since no constraints are defined. Action **D** is constrained. **D** must only start as soon as the predecessors **A** and **B** finished. Furthermore, the action **E** is constrained. **E** is the last action and must only be executed, if all predecessors (namely **A**, **C** and **D**) finished.
=\[fill=blue,draw=none,text=white\]
\(C) [$\mathbf{C}$]{}; (B) [$\mathbf{B}$]{}; (C) \[below of=B\] [$\mathbf{C}$]{}; (B) \[above of=C\] [$\mathbf{B}$]{}; (A) \[above of=B\] [$\mathbf{A}$]{}; (D) \[right of=B\] [$\mathbf{D}$]{}; (E) \[right of=D\] [$\mathbf{E}$]{};
\(A) edge node (E) (A) edge node (D) (B) edge node (D) (C) edge node (E) (D) edge node (E);
\[fig:graph\]
APPROACH {#sec:approach}
========
In this section we will describe our approach that is based on a model driven paradigm. We show how the concepts of meta-modeling and domain-specific languages are applied to develop a platform and robot system independent programming environment which focuses on parallel action sequences.
Requirements
------------
The programming environment shall assist robot programmers to develop robot control sequences as those described in Section \[sec:problemstatement\]. Thereby it shall fulfill two main requirements. First, a simplification of action sequences, so that no experts in the field are required. In order to achieve this, an intuitive graphical program editor must be provided. Besides the simplification of action sequence creation the environment must assist the developer to reduce failures that occur due to parallelism in such sequences, which could be e.g. a manipulator that grasps an object and blocks the field of view of a camera recording images.
Secondly, it is required that the environment is independent of the target robot. It must be possible to describe robot action sequences for any robot platform. Thereby, different robot classes can have different programming elements e.g. a vacuum cleaning robot may not have a manipulator and a programmer should not have the possibility to use a grasp action on these robot types. The programming elements must be adaptable and even the same control programs shall be applicable to different robots.
Meta-model
----------
The programming model is structured by a meta-model which describes the structure of the introduced action sequences, namely the abstract syntax and semantics. The concept follows the scheme of Model-Driven Software Development (MDSD) defined by Stahl and Völte [@c9]. The meta-model describes the basic programming elements ‘Action’, ‘Resource Component’, and ‘Variable’. See Figure \[fig:PaperMetaModel\].
- **Action:** An action is a certain thing that a robot can do, like ‘move to a position’, ‘set motor speeds’ or ‘capture an image’. Actions can work on a set of predefined global variables which are used as parameter or return values, comparable to function calls. Actions can be put into an execution sequence as defined in Section \[sec:problemstatement\]. Further, to assist concurrent programming, every action has a list of concrete actions, which are not allowed to be scheduled simultaneously. Every action belongs to a resource component.
- **Resource Component:** A resource component enables the usage of a certain action set. For instance, the definition of a concrete resource component ‘manipulator’ enables the programmer to use the actions ‘move arm to pose’ and ‘grasp’. If a robot does not have a manipulator, it also should not enable the programmer to use those actions. A resource component can schedule a single action at a time in a serial manner. Only the use of multiple resource components allows the creation of parallel action sequences. So, if a robot has only a single resource component ‘manipulator’ it shall not be able to execute multiple ‘grasp’ actions simultaneously. An additional resource component like human machine interface would allow simultaneous speech output to grasping actions. The term resource component refers not necessarily to a specific hardware device. It may also represent a computational unit like a functional library for planning, numerical computation, mapping or navigation.
- **Variable:** The meta-model further contains the structure for variables that are globally defined. Variables are either simple or arbitrarily cascaded like structs in the C language.
Domain-specific language
------------------------
The meta-model defines abstract constructs and does not define concrete functionality. Only the abstract elements action, resource component and variable are described. The concrete instances which are available to the programmer are formalized by the Domain Specific Language (DSL). The DSL can be described for different types of robot classes. Every DSL has to define which concrete resource components are provided by a specific robot class e.g. vacuum cleaners. For every resource component it must be defined which specific types of actions are provided by the resource. Those actions must each explicitly declare which other actions are not allowed to run simultaneously. Further, the DSL specifies the possible variables that a programmer can use. As mentioned, those can be basic variables like integer or float, or more complex structures build from other defined variables like the structure of an address. An example of a concrete DSL is sketched in Figure \[fig:PaperExampleDSL\]. It shows an excerpt of a DSL for a vacuum cleaning robot and defines the resource components *DriveBase* with the actions *MoveFwd* and *Stop* and the resource component *CleaningDevice* with the action *Discharge*. The shown DSL has adds a parallelism constraint. It describes that the action *moveFwd* may never run simultaneously to the action *discharge*. This removes the semantical error that a vacuum cleaning robot could move while distributing all collected dirt on the floor.
Finally, to get running robot code, a template-based code generator is used. The code generator is the first place in the process where a concrete robot systems is targeted since the program sequence and DSL can be defined for a complete robot class. Every resource component and every action can have an arbitrary number of code templates. Those templates must be written based on the robot framework and programming language. They must implement the action sequences for the robot. This makes it possible to generate code for different robot systems from the same sequence. Further, an arbitrary number of main templates can be specified. This may be necessary for robot code in e.g. C++ where multiple code files are required (source files and headers). Further, it enables the definition of code templates in different languages or frameworks for the same robot types but different concrete instances.
An overview of the system in the MDSD terminology is shown in Figure \[fig:SystemConceptMDSD\]. The figure shows the relation between the action sequence programmer, the programming environment and configuration as well as the domain expert that describes the domain specific elements.
- A **programmer** can create a concrete program model respective formal model with the programming environment which is an action sequence for a robot class. It is based on a DSL that contains the description of the concrete programming elements for the robot class. This DSL is based on a meta-model which provides structure and is the same for any formal model and robot.
- The **domain expert** specifies the DSL via a configuration file. Profound knowledge in programming a certain robot respective robot class is needed. Moreover, the domain expert has to provide the code templates for a specific robot to enable the translation from formal model to specific robot code.
Implementation
--------------
The system is developed with Java and the Eclipse Rich Client Platform RCP [@c13] and is operating system independent. The graphical editor is implemented with the Eclipse Graphical Editing Framework [@c11] and allows the creation of the described action sequences. The meta-model is formalized as a set of Java classes while the DSL configuration can be separately loaded as XML files. Moreover, the existing sequences can be robot independent serialized to XML. The editor validates the generated action sequences on completeness, e.g. are all names unique, variables instantiated and parameters set and verifies if the defined parallelism constraints are not violated. For code generation the Apache Velocity Engine [@c12] is integrated. This code genaration library was chosen because of its very clear and powerful template language. The amount, location and names of the templates are also configurable with XML.
USE CASES {#sec:usecases}
=========
Service robotics use case
-------------------------
In the first use case a simulated service robot system with a holonomic drive system and two manipulators has been integrated in the environment. An example sequence which makes the robot move to a certain position and grasp an object is shown in Figure \[fig:sampleMRDS\]. For this, a minimalistic DSL containing the resource components for each gripper and the drive base has been designed with actions for closing a gripper, moving the arm to certain positions and moving the robots base. An excerpt of the DSL is given in Listing \[list:dslconfigscheduler\].
{}
<ResourceComponent type="Manipulator">
<Action returnType="String"
actionIdentifier="MoveManipulator">
<ParameterList>
<Parameter type="Vector3" name="targetPose">
</Parameter>
<Parameter type="Vector3"name="orientation">
</Parameter>
</ParameterList>
<NotAllowedSimultaneousActionTypes>
<NotAllowedSimultaneousAction type="MoveTo">
</NotAllowedSimultaneousAction>
</NotAllowedSimultaneousActionTypes>
</Action>
</ResourceComponent>
To generate code, every action and resource component has its custom code template. An excerpt of the moveManipulator action is given in Listing \[list:MRDSTemplateMoveManipulator\]. The template uses keywords of the velocity engine which are indicated by a ’\#’ for control structures and ’\$’ for data access from the formal model. Listing \[list:MRDSGeneratedMoveManipulator\] shows the generated code that results for this template.
{}
//Create list of parameters
parameters = new List<ParameterVariable>();
//fill list of parameters
#foreach($Parameter in $Action.getParameters())
//Add previous initialized variables
parameters.Add(getVariable(
"$Parameter.getVariable().getName()"));
#end
//Create robot specific action
ExecutionElement $Action.getName() =
new ExecElement(MOVE_MANIPULATOR, parameters));
{}
//Create list of parameters
parameters = new List<ParameterVariable>();
//Add previous initialized variables
parameters.Add(getVariable("targetPose"));
parameters.Add(getVariable("orientation"));
//Create robot system specific action
ExecutionElement MoveMani =
new ExecElement(MOVE_MANIPULATOR, parameters));
Educational use case
--------------------
A further use case based on a different configured DSL has been implemented with the same programming environment. This DSL configures the environment to program a Lego Mindstorm NXT educational robot. The robot was constructed with a left and right ultra sonic sensor, that are oriented at the robot front. The robot is actuated by a differential drive base. This use case shows the implementation of a Braitenbergs vehicles [@c10] so that the robot avoids obstacles. The program sequence is shown in Figure \[fig:sampleLegoObstacle\]. In Braitenberg vehicles the sensor readings are directly mapped to actuators which can result in complex reactive behaviors. For the use case the robots left sonar is mapped to the left motor and the right sonar is mapped to the right motor. The closer the robot senses an object on its right sensor, the higher the sensor value respective right motor speed which results in a turning behavior away from the obstacle. In the sequence shown in the figure, the sensors are read simultaneously and their results are passed to the left and right motors. This is done by two global variables storing the left and right ultra sonic sensor readings. After reading, these variables are used by the motors to set the speed. The left sensor shares a variable with the right motor and the right sensor shares a variable with the left motor. If an obstacle is detected in the left sensor, the motor on the right gets slower since it acts directly on the sensor reading that gets smaller by closer obstacle distance. This results in a turn away from the obstacle. The sequence is translated to NXC code that must be compiled and transfered to the robot.
CONCLUSION {#sec:conclusion}
==========
This paper presented a programming environment that allows the graphical programming of robot action sequences comparable to dependency graphs. Main focus of the environment is the simplification of robot programming especially in the sense of parallel program development. The designed system is based on the model-driven software development aspects which makes it possible to re-use it with any robot system or even transform the same described program into code for different robot platforms. In order to achieve this, a meta-model has been designed while robot specific DSLs and code generation settings can be loaded via XML configuration files. Two exemplary use-cases were described in which a a service robot and a robot for educational purpose were programmed with the presented environment. The environment is successfully used in lectures to teach students the programming of Lego NXT Robots. There it has been proven as a valuable tool to teach parallel programming on real robots.
In the future we will develop domain-specific languages and code generation templates for different robot platforms so for example our service robot Johnny Jackanapes [@c14] which is participating in the service robot competition RoboCup@Home. Further, the programming model will be enhanced by decisional elements to allow a state machine oriented programming.
ACKNOWLEDGMENTS
===============
The research leading to these results has received funding from the European Community’s Seventh Framework Programme (FP7/2007-2013) under grant agreement no. FP7-ICT-231940-BRICS (Best Practice in Robotics).
[99]{}
P. A. Baer, R. Reichle, and K. Geihs. The Spica Development Framework – Model-Driven Software Development for Autonomous Mobile Robots. In [*IAS-10, pp. 211–220*]{}. 2008.
Christian Schlegel, Thomas Ha"sler, Alex Lotz and Andreas Steck. Robotic Software Systems: From Code-Driven to Model-Driven Designs. In [*Proc. 14th Int. Conf. on Advanced Robotics (ICAR), Munich*]{}. 2009.
Diego Alonso, Cristina Vincente-Chicote, Francisco Ortiz, Juan Pastor, Barbara Alvarez. V3CMM: A 3-View Component Meta-Model for Model-Driven Robotic Software Development. In [*Journal of Software Engineering for Robotics*]{}. 2010
Joachim Hertzberg and Frank Schönherr. Concurrency in the DD&P Robot Control Architecture. In [*Proc. of The Int. NAISO Congress on Information Science Innovations (ISIÕ2001)*]{}. 2001.
Gostai. Urbi Studio, [*\[Online\] http://www.gostai.com/products/studio/*]{}, May 2010.
Aldebaran Robotics. Nao Choregraphe, [*\[Online\] http://www.aldebaran-robotics.com/en*]{}, July 2010.
Susan Horwitz and Thomas Reps. The use of program dependence graphs in software engineering, [*Proceedings of the 14th international conference on Software engineering*]{}, p.392-411, May 1992.
Thomas Stahl and Markus Völter. Model-Driven Software Development. [*Wiley & Sons*]{}, 2006,
Jeff McAffer, Jean-Michel Lemieux and Chris Aniszczyk. Eclipse Rich Client Platform - Designing, Coding and Packaging Java Applications, [*Addison-Wesley, the eclipse series*]{}, 2006.
The Eclipse Foundation. The Graphical Editing Framework, [*\[Online\] http://www.eclipse.org/gef*]{}, July 2010.
Apache Software Foundation. The Apache Velocity Framework, [*\[Online\] http://velocity.apache.org/*]{}, July 2010.
Valentin Braitenberg. Vehicles: Experiments in synthetic psychology, [*MIT Press*]{}, 1984.
Thomas Breuer, Geovanny Giorgana, Frederik Hegger, Christian Müller, Zha Jin, Michael Reckhaus, Jan Paulus, Nico Hochgeschwender, Ronny Hartanto, Paul Ploeger and Gerhard Kraetzschmar. The b-it-bots RoboCup@Home 2010 Team Description Paper, RoboCup-2010, Singapore, 2010.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
Let $\{B_k\}_{k=1}^\infty, \{X_k\}_{k=1}^\infty$ all be independent random variables. Assume that $\{B_k\}_{k=1}^\infty$ are $\{0,1\}$-valued Bernoulli random variables satisfying $B_k\stackrel{\text{dist}}{=}\text{Ber}(p_k)$, with $\sum_{k=1}^\infty p_k=\infty$, and assume that $\{X_k\}_{k=1}^\infty$ satisfy $X_k>0$ and $\mu_k\equiv EX_k<\infty$. Let $M_n=\sum_{k=1}^np_k\mu_k$, assume that $M_n\to\infty$ and define the normalized sum of independent random variables $W_n=\frac1{M_n}\sum_{k=1}^nB_kX_k$. We give a general condition under which $W_n\stackrel{\text{dist}}{\to}c$, for some $c\in[0,1]$, and a general condition under which $W_n$ converges weakly to a distribution from a family of distributions that includes the generalized Dickman distributions GD$(\theta),\theta>0$. In particular, we obtain the following result, which reveals a strange domain of attraction to generalized Dickman distributions. Assume that $\lim_{k\to\infty}\frac{X_k}{\mu_k}\stackrel{\text{dist}}{=}1$. Let $J_\mu,J_p$ be nonnegative integers, let $c_\mu,c_p>0$ and let
$
\mu_n\sim c_\mu n^{a_0}\prod_{j=1}^{J_\mu}(\log^{(j)}n)^{a_j}, \
p_n\sim c_p\big({n^{b_0}\prod_{j=1}^{J_p}(\log^{(j)}n)^{b_j}}\big)^{-1}, \ b_{J_p}\neq0.
$ If $$\begin{aligned}
&i.\ J_p\le J_\mu;\\
&ii.\ b_j=1, \ 0\le j\le J_p;\\
&iii.\ a_j=0, \ 0\le j\le J_p-1,\ \text{and}\ \ a_{J_p}>0,
\end{aligned}$$ then $ \lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}\frac1{\theta}\text{GD}(\theta),\ \text{where}\ \theta=\frac{c_p}{a_{J_p}}.
$ Otherwise, $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}c$, for some $c\in[0,1]$. We also give an application to the statistics of the number of inversions in certain random shuffling schemes.
address: |
Department of Mathematics\
Technion—Israel Institute of Technology\
Haifa, 32000\
Israel
author:
- 'Ross G. Pinsky'
title: On the strange domain of attraction to generalized Dickman distributions for sums of independent random variables
---
Introduction and Statement of Results
=====================================
The Dickman function $\rho_1$ is the unique function, continuous on $(0,\infty)$, and satisfying the differential-delay equation $$\begin{aligned}
&\rho_1(x)=0,\ x\le0;\\
&\rho_1(x)=1,\ x\in(0,1];\\
&x\rho_1'(x)+\rho_1(x-1)=0, \ x>1.
\end{aligned}$$ This function has an interesting role in number theory and probability, which we describe briefly at the end of this section. With a little work, one can show that the Laplace transform of $\rho_1$ is given by $\int_0^\infty \rho_1(x)e^{-\lambda x}dx=\exp(\gamma+\int_0^1\frac{e^{-\lambda x}-1}xdx)$, where $\gamma$ is Euler’s constant. From this it follows that $\int_0^\infty \rho_1(x)dx=e^\gamma$, and consequently, that $e^{-\gamma}\rho_1$ is a probability density on $[0,\infty)$. We will call this probability distribution the *Dickman distribution. We denote its density by $p_1(x)=e^{-\gamma}\rho_1(x)$, and we denote by $D_1$ a random variable distributed according to the Dickman distribution. Differentiating the Laplace transform $E\exp(-\lambda D_1)=\exp(\int_0^1\frac{e^{-\lambda x}-1}xdx)$ of $D_1$ at $\lambda=0$ shows that $ED_1=1$. The distribution decays very rapidly; indeed, it is not hard to show that $p_1(x)\le \frac{e^{-\gamma}}{\Gamma(x+1)},\ x\ge0$ [@MV].*
In fact, for all $\theta>0$, $\exp(\theta\int_0^1\frac{e^{-\lambda x}-1}xdx)$ is the Laplace transform of a probability distribution. (We will prove this directly; however, this fact follows from the theory of infinitely divisible distributions, and shows that the distribution in question is infinitely divisible.) This distribution has density $p_\theta=\frac{e^{-\theta\gamma}}{\Gamma(\theta)}\rho_\theta$, where $\rho_\theta$ satisfies the differential-delay equation $$\label{rhotheta}
\begin{aligned}
&\rho_\theta(x)=0,\ x\le0;\\
&\rho_\theta(x)=x^{\theta-1},\ 0< x\le 1;\\
&x\rho_\theta'(x)+(1-\theta)\rho_\theta(x)+\theta\rho_\theta(x-1)=0,\ x>1.
\end{aligned}$$ We will call such distributions *generalized Dickman distributions and denote them by GD$(\theta)$. We denote by $D_\theta$ a random variable with the GD$(\theta)$ distribution. Differentiating its Laplace transform at $\lambda=0$ shows that $ED_\theta=\theta$. These distribution decays very rapidly; indeed, it is not hard to show that $p_\theta(x)\le \frac{C_\theta}{\Gamma(x+1)}, x\ge1$, for an appropriate constant $C_\theta$. A fundamental fact about these distributions is that $$\label{fund}
D_\theta\stackrel{\text{dist}}{=}U^{\frac1\theta}(D_\theta+1),$$ where $U$ is distributed according to the uniform distribution on $[0,1]$, and $U$ and $D_\theta$ on the right hand side above are independent. From it is immediate that $$D_\theta\stackrel{\text{dist}}{=}U_1^\frac1\theta+(U_1U_2)^\frac1\theta+(U_1U_2U_3)^\frac1\theta+\cdots,$$ where $\{U_n\}_{n=1}^\infty$ are IID random variables distributed according to the uniform distribution on $[0,1]$. It will follow from the proof of Theorem \[1\] below that $\exp(\theta\int_0^1\frac{e^{-\lambda x}-1}xdx)$ is the Laplace transform of a probability distribution. In section \[background\] we will prove that a random variable with such a distribution satisfies , and that if a random variable satisfies , then it has a density of the form $c_\theta\rho_\theta$, where $\rho_\theta$ satisfies . Thus, this paper is self-contained with regard to all the above noted facts, with the exception of the rate of decay and the value $\frac{e^{-\theta\gamma}}{\Gamma(\theta)}$ of the normalizing constant $c_\theta$ in $p_\theta$. For more on these distributions, including a derivation of the normalizing constant, see, for example, [@ABT] and [@PW].*
In fact, the scope of this paper leads us to consider a more general family of distributions than the generalized Dickman distributions. Let $\mathcal{X}\ge0$ be a random variable satisfying $E\mathcal{X}\le 1$. Then, as we shall see, for $\theta>0$, there exists a distribution whose Laplace transform is $\exp\big(\theta\int_0^1\frac{E\exp(-\lambda x\mathcal{X})-1}xdx\big)$. We will denote this distribution by $GD^{(\mathcal{X})}(\theta)$ and we denote a random variable with this distribution by $D_\theta^{(\mathcal{X})}$. (When $\mathcal{X}\equiv1$, we revert to the previous notation for generalized Dickman distributions.) Differentiating the Laplace transform at $\lambda=0$ shows that $ED_\theta^{(\mathcal{X})}=\theta E\mathcal{X}$.
Mimicking the proof of that we give in section \[background\] shows that $$\label{fundagain}
D_\theta^{(\mathcal{X})}=U^{\frac1\theta}(D_\theta^{(\mathcal{X})}+\mathcal{X}),$$ where $U$ is distributed according to the uniform distribution on $[0,1]$, and $U$, $D_\theta^{(\mathcal{X})}$ and $\mathcal{X}$ on the right hand side above are independent. From it is immediate that $$D_\theta\stackrel{\text{dist}}{=}\mathcal{X}_1U_1^{\frac1\theta}+\mathcal{X}_2(U_1U_2)^{\frac1\theta}+\mathcal{X}_3(U_1U_2U_3)^{\frac1\theta}+\cdots,$$ where $\{U_n\}_{n=1}^\infty$ and $\{\mathcal{X}_n\}_{n=1}^\infty$ are mutually independent sequences of IID random variables, with $U_1$ distributed according to the uniform distribution on $[0,1]$ and $\mathcal{X}_1$ distributed according to the distribution of $\mathcal{X}$.
It is known that the generalized Dickman distribution $GD(\theta)$ arises as the limiting distribution of $\frac1n\sum_{k=1}^nkY_k$, where the $\{Y_k\}_{k=1}^\infty$ are independent random variables with $Y_k$ distributed according to the Poisson distribution with parameter $\frac \theta k$ [@ABT]. It is also known that the Dickman distribution $GD(1)$ arises as the limiting distribution of $\frac1n\sum_{k=1}^n kY_k$ as $n\to\infty$, where the $\{Y_k\}_{k=1}^\infty$ are independent Bernoulli random variables satisfying $P(Y_k=1)=1-P(Y_k=0)=\frac1k$. Such behavior is in distinct contrast to the law of large numbers behavior of a “well-behaved” sequence of independent random variables $\{Z_k\}_{k=1}^\infty$ with finite first moments; namely, that $\frac1{M_n}\sum_{k=1}^nZ_k$ converges in distribution to 1 as $n\to\infty$, where $M_n=\sum_{k=1}^nEZ_k$.
The purpose of this paper is to understand when the law of large numbers fails and a distribution from the family $\text{GD}^{(\mathcal{X})}(\theta)$ arises in its stead. From the above examples, we see that generalized Dickman distributions sometimes arise as limits of normalized sums from a sequence $\{V_k\}_{k=1}^\infty$ of independent random variables which are are non-negative and satisfy the following three conditions: (i) $\lim_{k\to\infty}P(V_k=0)=1$, (ii) $\lim_{k\to\infty}\frac{V_k|V_k>0}{E(V_k|V_k>0)}\stackrel{\text{dist}}{=}1$ and (iii) $\sum_{k=1}^\infty EV_k=\infty$. (In the above examples, $kY_k$ plays the role of $V_k$.) It turns out that these three conditions are very far from sufficient for a generalized Dickman distribution to arise. In fact, as we shall see in Theorem \[2\] below, such distributions arise only in a strange sequence of very narrow windows of opportunity.
In light of the above discussion, we will consider the following setting. Let $\{B_k\}_{k=1}^\infty, \{X_k\}_{k=1}^\infty$ be mutually independent sequences of independent random variables. Assume that $\{B_k\}_{k=1}^\infty$ are Bernoulli random variables satisfying: $$\label{Ber}
P(B_k=1)=1-P(B_k=0)=p_k\in[0,1),$$ and assume that $\{X_k\}_{k=1}^\infty$ satisfy: $$\label{Xn}
X_k>0,\ \ \ \mu_k\equiv EX_k<\infty.$$ Let $$\label{Mn}
M_n=\sum_{k=1}^np_k\mu_k,$$ and define $$\label{W}
W_n=\frac1{M_n}\sum_{k=1}^nB_kX_k.$$ We will be interested in the limiting behavior of $W_n$. In order to avoid trivialities, we will assume that $$\label{nontriv}
\lim_{n\to\infty}M_n=\infty\ \ \text{and} \ \ \ \sum_{k=1}^\infty p_k=\infty,$$ since otherwise $\sum_{n=1}^\infty B_kX_k$ is almost surely finite.
Note that for the example brought with the $\text{Pois}(\frac\theta k)$-distribution, we have $p_k=1-e^{-\frac\theta k}$, $X_k$ is distributed according to $kY_k|\{Y_k>0\}$, where $Y_k$ has the Pois$(\frac\theta k)$ distribution, $\mu_k=\frac\theta{1-e^{-\frac\theta k}}$ and $M_n=n\theta$. And for the example with the $\text{Ber}(\frac1k)$-distribution, we have $p_k=\frac1k$, $X_k=k$ deterministically, $\mu_k=k$ and $M_n=n$. In the first of these two examples, $\frac{X_k}{\mu_k}\stackrel{\text{dist}}{\to}1$, and in the second one, $\frac{X_k}{\mu_k}\stackrel{\text{dist}}{=}1$ for all $k$.
Our first theorem gives a general condition for $W_n\stackrel{\text{dist}}{\to}c$ (which is the law of large numbers if $c=1$), and a general condition for convergence to a limiting distribution from the family of distributions $\text{GD}^{(\mathcal{X})}(\theta)$. Using this theorem, we can prove our second theorem, which reveals the strange domain of attraction to generalized Dickman distributions. (Of course, we are using the term “domain of attraction” not in its classical sense, since our sequence of random variables, although independent, are not identically distributed.) Let $\delta_c$ denote the degenerate distribution at $c$.
\[1\] Let $W_n$ be as in , where $\{B_k\}_{k=1}^\infty$, $\{X_k\}_{k=1}^\infty$ and $M_n$ are as in - and .
i\. Assume that $\{\frac{X_k}{\mu_k}\}_{k=1}^\infty$ is uniformly integrable (which occurs automatically if $\lim_{k\to\infty}\frac{X_k}{\mu_k}\stackrel{\text{dist}}{=}1$).
*a. Assume also that $$\label{maxmueasy}
\lim_{n\to\infty}\frac{\max_{1\le k\le n}\thinspace\mu_k}{M_n}=0.$$ Then $$\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}1.$$*
*b. Assume also that there exists a sequence $\{K_n\}_{n=1}^\infty$ such that $$\label{Kn}
\lim_{n\to\infty}\sum_{k=K_n+1}^n p_k=0,$$ and $$\label{maxmu}
\lim_{n\to\infty}\frac{\max_{1\le k\le K_n}\thinspace\mu_k}{M_n}=0.$$ If $$\label{cexists}
c\equiv\lim_{n\to\infty}\frac{M_{K_n}}{M_n}\ \text{exists},$$ then $$\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}c.$$ If does not hold, then the distributions of $\{W_n\}_{n=1}^\infty$ form a tight sequence whose set of accumulation points is $\{\delta_c: c\in A\}$, where $A$ denotes the set of accumulation points of the sequence $\{\frac{M_{K_n}}{M_n}\}_{n=1}^\infty$.*
ii\. Assume that there exists a random variable $\mathcal{X}$ such that $$\label{weakconvtoX}
\lim_{k\to\infty}\frac{X_k}{\mu_k}\stackrel{\text{dist}}{=}\mathcal{X}.$$ Assume also that $\{\mu_k\}_{k=1}^\infty$ is increasing, that $\lim_{k\to\infty}p_k=0$ and that there exist $\theta,L\in(0,\infty)$ such that $$\label{twocond}
\lim_{k\to\infty}\frac{p_k\mu_k}{\mu_{k+1}-\mu_k}=\theta,\ \ \ \lim_{k\to\infty}\frac{\mu_k}{M_k}=L.$$ Then $$\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}LD^{(\mathcal{X})}(\theta),$$ where $D^{(\mathcal{X})}(\theta)$ is a random variable with the $\text{GD}^{(\mathcal{X})}(\theta)$ distribution.
**Remark 1. In , necessarily $L\le\frac1\theta$. Indeed, if $\{p_k\}_{k=1}^{\infty}$ and $\{\theta_k\}_{k=1}^\infty$ satisfy the conditions of part (ii), and we choose $X_k=\mu_k$, then $W_n\stackrel{\text{dist}}{\to}LD_\theta$. Since $EW_n=1$ and $ED_\theta=\theta$, it follows from Fatou’s lemma that $L\le \frac1\theta$. In most cases of interest, one has $L=\frac1\theta$.**
**Remark 2. By Fatou’s lemma, the random variable $\mathcal{X}$ in part (ii) must satisfy $E\mathcal{X}\le 1$.**
**Remark 3. The uniform integrability of $\{\frac{X_k}{\mu_k}\}_{k=1}^\infty$ in part (i) occurs automatically if $\lim_{k\to\infty}\frac{X_k}{\mu_k}\stackrel{\text{dist}}{=}1$, because if a sequence $\{Y_k\}_{k=1}^\infty$ of random variables satisfies $Y_k\stackrel{\text{dist}}{\to}Y$, and $E|Y_k|<\infty$, then $E|Y_k|\to E|Y|$ is equivalent to uniform integrability.**
**Remark 4. In the case that $X_k=\mu_k$, or more generally, if $EX_k\le C\mu_k^2$, for all $k$ and some $C>0$, then $$Var(W_n)\le\frac{C\sum_{k=1}^Np_k\mu_k^2}{M_n^2}=C\frac{\sum_{k=1}^np_k\mu_k^2}{(\sum_{k=1}^np_k\mu_k)^2}\le
C\frac{\sup_{1\le k\le n}\mu_k}{M_n}.$$ Thus, in this case part (i-a) follows directly from the second moment method.**
Using Theorem 1, we can prove the following theorem that exhibits the strange domain of attraction to generalized Dickman distributions. Let $\log^{(j)}$ denote the $j$th iterate of the logarithm, and make the convention $\prod_{j=1}^0=1$.
\[2\] Let $W_n$ be as in , where $\{B_k\}_{k=1}^\infty$, $\{X_k\}_{k=1}^\infty$ and $M_n$ are as in -. Assume also that $\lim_{k\to\infty}\frac{X_k}{\mu_k}\stackrel{\text{dist}}{=}1$. Let $J_\mu,J_p$ be nonnegative integers, let $c_\mu,c_p>0$ and define $$\begin{aligned}
&\mu(x)= c_\mu x^{a_0}\prod_{j=1}^{J_\mu}(\log^{(j)}x)^{a_j}, \\
&p(x)= c_p\big({x^{b_0}\prod_{j=1}^{J_p}(\log^{(j)}x)^{b_j}}\big)^{-1},
\end{aligned}$$ with $b_{J_p}\neq0$. Assume that $$\begin{aligned}
&\mu_k\sim\mu(k),\ \ p_k\sim p(k);\\
&\mu_{k+1}-\mu_k\sim\mu'(k).
\end{aligned}$$ Assume that the exponents $\{a_j\}_{j=0}^{J_\mu}, \{b_j\}_{j=0}^{J_p}$ have been chosen so that holds. If $$\label{3cond}
\begin{aligned}
&i.\ J_p\le J_\mu;\\
&ii.\ b_j=1, \ 0\le j\le J_p;\\
&iii.\ a_j=0, \ 0\le j\le J_p-1,\ \text{and}\ \ a_{J_p}>0,
\end{aligned}$$ then $$\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}\frac1\theta D_\theta,\ \text{with}\ \theta=\frac{c_p}{a_{J_p}},$$ where $D_\theta$ is a random variable with the GD$(\theta)$ distribution.
Otherwise, $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}c$, where $c\in\{0,1\}$. To determine $c$, let $$\label{kappa}
\kappa_\mu=\min\{0\le j\le J_\mu:a_j\neq0\}\ \ \text{and}\ \ \kappa_p=\min\{0\le j\le J_p:b_j\neq1\}.$$ If $\{0\le j\le J_\mu:a_j\neq0\}$ is not empty, $a_{\kappa_\mu}>0$ and either $\{0\le j\le J_p:b_j\neq1\}$ is empty and $\kappa_\mu<J_p$, or $\{0\le j\le J_p:b_j\neq1\}$ is not empty and $\kappa_\mu<\kappa_p$, then $c=0$; otherwise, $c=1$.
**Remark 1. Note that if one chooses $\mu_k=\mu(k)$ and $p_k=p(k)$, then the condition $\mu_{k+1}-\mu_k\sim \mu'(k)$ is always satisfied.**
**Remark 2. Theorem \[2\] shows that to obtain a generalized Dickman distribution, $\{p_k\}_{k=1}^\infty$ in particular must be set in a very restricted fashion. For some intuition regarding this phenomenon, take the situation where $X_k=\mu_k$, and consider the sequence $\{\sigma^2(W_n)\}_{n=1}^\infty$ of variances. This sequence converges to 0 in the cases where $W_n$ converges to 1, converges to $\infty$ in the cases where $W_n$ converges to 0, and converges to a positive number in the cases where $W_n$ converges to a generalized Dickman distribution.**
We now state explicitly what Theorem \[2\] yields in the cases $J_p=0,1$.
$\bf J_p=0$. We have $$p_n\sim \frac{c_p}{n^{b_0}},\ b_0>0, \ \ \mu_n\sim c_\mu n^{a_0}\prod_{j=1}^{J_\mu}(\log^{(j)}n)^{a_j}.$$ In order that hold, we require $b_0\le1$. We also require either: $a_0-b_0>-1$; or $a_0-b_0=-1$ and $a_1>-1$; or $a_0-b_0=a_1=-1$ and $a_2>-1$; etc.
*If $b_0=1$ and $a_0>0$, then $$\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=} \frac1\theta D_\theta,\
\text{where}\ \theta=\frac{c_p}{a_0}.$$ Otherwise, $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}1.$*
$\bf J_p=1$. We have $$p_n\sim\frac{c_p}{n^{b_0}(\log n)^{b_1}}, \ b_1\neq0,\ \ \ \mu_n\sim c_\mu n^{a_0}\prod_{j=1}^{J_\mu}(\log^{(j)}n)^{a_j}.$$ In order that hold, we require either $b_0=0$ and $b_1>0$, or $0<b_0<1$, or $b_0=1$ and $b_1\le1$. We also require either: $a_0-b_0>-1$; or $a_0-b_0=-1$ and $a_1-b_1>-1$; or $a_0-b_0=a_1-b_1=-1$ and $a_2>-1$; etc.
*If $J_\mu\ge1$, $b_0=b_1=1$, $a_0=0$ and $a_1>0$, then $$\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=} \frac1\theta D_\theta,
\ \text{where}\ \theta=\frac{c_p}{a_1}.$$*
If $b_0=1$ and $a_0>0$, then $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}0$.
Otherwise, $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}1$.
**Remark. In [@CS] and [@Pin], where the GD$(1)$ distribution arises, one has $b_0=b_1=1, a_0=0, a_1=1, c_p=c_\mu=1$.**
The organization of the rest of the paper is as follows. In section \[application\] we use Theorems \[1\] and \[2\] to investigate a question raised in [@HT] concerning the statistics of the number of inversions in certain random shuffling schemes. In sections \[pfthm1\] and \[pfthm2\] respectively we prove Theorems \[1\] and \[2\]. Finally, in section \[background\] we prove the basic facts about the Dickman distribution and its density, as was promised earlier in this section.
As mentioned above, we end this section with a little background concerning the Dickman function $\rho\equiv \rho_1$. The Dickman function arises in probabilistic number theory in the context of so-called *smooth numbers; that is, numbers all of whose prime divisors are “small.” Let $\Psi(x,y)$ denote the number of positive integers less than or equal to $x$ with no prime divisors greater than $y$. Numbers with no prime divisors greater than $y$ are called $y$-*smooth numbers. Then for $s\ge1$, $\Psi(N,N^\frac1s)\sim N\rho(s)$, as $N\to\infty$. This result was first proved by Dickman in 1930 [@Dick], whence the name of the function, with later refinements by de Bruijn [@deB1]. See also [@MV] or [@Tene]. Let $[n]=\{1,\ldots, n\}$ and let $p^{+}(n)$ denote the largest prime divisor of $n$. Then Dickman’s result states that the random variable $\frac{\log p^{+}(j)}{\log n}, j\in[n]$, on the probability space $[n]=\{1,\ldots, n\}$ with the uniform distribution converges in distribution as $n\to\infty$ to the distribution whose distribution function is $\rho(\frac1x)$, $x\in[0,1]$, and whose density is $-\frac{\rho'(\frac1x)}{x^2}=\frac{\rho(\frac1x-1)}x,\ x\in[0,1]$. It is easy to see that an equivalent statement of Dickman’s result is that the random variable $\frac{\log p^{+}(j)}{\log j}$, $j\in[n]$, on the probability space $[n]$ with the uniform distribution converges in distribution as $n\to\infty$ to the distribution whose distribution function is $\rho(\frac1x)$, $x\in[0,1]$, We note that the length of the longest cycle of a uniformly random permutation of $[n]$, normalized by dividing by $n$, also converges to a limiting distribution whose distribution function is $\rho(\frac1x)$. If instead of using the uniform measure on $S_n$, the set of permutations of $[n]$, one uses the Ewens sampling distribution on $S_n$, obtained by giving each permutation $\sigma\in S_n$ the probability proportional to $\theta^{C(\sigma)}$, where $C(\sigma)$ denotes the number of cycles in $\sigma$, then the length of the longest cycle of such a random permutation of $[n]$, normalized by dividing by $n$, converges to a limiting distribution whose distribution function is $\rho_\theta(\frac1x)$, $x\in[0,1]$. This distribution is also the distribution of the first coordinate of the Poisson-Dirichlet distribution PD$(\theta)$ (see [@ABT]).**
The examples in the above paragraph lead to limiting distributions where the Dickman function arises as a distribution function, not as a density as is the case with the GD$(\theta)$ distributions discussed in this paper. The GD$(\theta)$ distribution arises as a normalized limit in the context of certain natural probability measures that one can place on $\mathbb{N}$; see [@CS], [@Pin].
An application to random permutations {#application}
=====================================
We consider a setup that appeared in [@HT], and which in the terminology of this paper can be described as follows. For each $k\in\mathbb{N}$, let $E_k\subset\{1,\ldots, k-1\}$. Let $X_k$ be uniformly distributed on $E_k$, and let $B_k\stackrel{\text{dist}}{=}\text{Ber}(\frac{|E_k|}{k})$. So $$\mu_k=\frac1{|E_k|}\sum_{l\in E_k}l,\ \ \ p_k=\frac{|E_k|}k.$$ Define $$I_n=\sum_{k=1}^n B_kX_k.$$ We allow $E_k=\emptyset$, in which case $B_k=0$ and $X_k$ is not defined. In such a case, we define $B_kX_k=0$ and $\mu_k=0$. We always have $E_1=\emptyset$.
Consider first the case that $E_k=\{1,\ldots, k-1\}$. Then $B_1X_1=0$ and for $2\le k\le n$, $B_kX_k$ is uniformly distributed over $\{0,1\ldots, k-1\}$. In this case, $I_n$ has the distribution of the number of inversions in a uniformly random permutation from $S_n$. (The authors in [@HT] have a typo and wrote $E_k=\{1,\ldots, k\}$ instead.) To see this, consider the following shuffling procedure for $n$ cards, numbered from 1 to $n$. The cards are to be inserted in a row, one by one, in order of their numbers. At step one, card number 1 is set down. The number of inversions created by this step is zero, which is given by $B_1X_1$. At step $k$, for $k\in\{2,\ldots, n\}$, card number $k$ is randomly inserted in the current row of cards, numbered 1 to $k-1$. Thus, for any $j\in \{0,1,\ldots, k-1\}$, card number $k$ has probability $\frac1k$ of being placed in the position with $j$ cards to its right (and $k-1-j$ cards to its left), in which case this step will have created $j$ new inversions, and this is represented by $B_kX_k$. It is clear from the construction that the random variables $\{B_kX_k\}_{k=1}^n$ are independent. Thus, $I_n$ indeed gives the number of inversions in a uniformly random permutation from $S_n$. It is well-known that the law of large numbers and the central limit theorem hold for $I_n$ in this case.
Consider now the general case that $E_k\subset\{1,\ldots, k-1\}$. Then $I_n$ gives the number of inversions in a random permutation created by a shuffling procedure in the same spirit as the above one. At step $k$, with probability $1-\frac{|E_k|}k$, card number $k$ is inserted at the right end of the row, thereby creating no new inversions, and for each $j\in E_k$, with probability $\frac1k$ it is inserted in the position with $j$ cards to its right, thereby creating $j$ new inversions.
In particular, as a warmup consider the cases $E_k=\{1\}$ and $E_k=\{k-1\}$, $2\le k\le n$. In each of these two cases, at step $k$, $2\le k\le n$, card number $k$ is inserted at the right end of the row with probability $1-\frac1k$. In the first case, with probability $\frac1k$ card number $k$ is inserted immediately to the left of the right most card, thereby creating one new inversion, while in the second case, with probability $\frac1k$ card number $k$ is inserted at the left end of the row, thereby creating $k-1$ new inversions. In both cases $\frac{X_n}{\mu_n}\stackrel{\text{dist}}{=}1$ for all $n$, and in both cases, $p_k=\frac1k$. In the first case, $\mu_k=1$ while in the second case, $\mu_k=k-1$. Thus, in the first case, $M_n=\sum_{k=1}^n p_k\mu_k\sim\log n$, and in the second case, $M_n\sim n$. Therefore, it follows from Theorem \[1\] or \[2\] that in the first case $\frac{I_n}{\log n}$ converge in distribution to 1, while in the second case, $\frac{I_n}n$ converges in distribution to GD$(1)$.
The authors of [@HT] ask which choices of $\{E_k\}_{k=1}^\infty$ lead to the Dickman distribution and which choices lead to the central limit theorem. Of course, the law of large numbers is a prerequisite for the central limit theorem. The following theorem gives sufficient conditions for the law of large numbers to hold and sufficient conditions for convergence to a distribution from the family $\text{GD}^{(\mathcal{X})}(\theta)$. In order to avoid trivialities, we need to assume that holds. Recalling that $\mu_k=0$ when $|E_k|=0$, and that $\mu_k\ge1$ otherwise, note that $$M_n=EI_n=\sum_{k=1}^\infty\frac{|E_k|}k\mu_k\ge\sum_{k=1}^\infty\frac{|E_k|}k =\sum_{k=1}^\infty p_k.$$ Thus, in the present context the requirement is $$\label{nontrivagain}
\sum_{k=1}^\infty\frac{|E_k|}k=\infty,$$ which holds in particular if $E_k\neq\emptyset$ for all sufficiently large $k$.
\[shufflethm\] Assume that holds.
i\. Assume that at least one of the following conditions holds:
a\. $\lim_{k\to\infty}|E_k|=\infty$ and $\{\frac{\mu_n}{\sum_{k=1}^n\frac{\mu_k}k}\}_{n=1}^\infty$ is bounded;
b\. $\lim_{n\to\infty}\frac{\mu_n}{\sum_{k=1}^n\frac{\mu_k}k}=0$.
Then $\frac{I_n}{EI_n}\stackrel{\text{dist}}{\to}1$.
ii\. Assume that $|E_k|=N\ge1$, for all large $k$, and that $\frac{X_k}{\mu_k}\stackrel{\text{dist}}{\to} \mathcal{X}$. Also assume that $\mu_k\sim\mu(k)$ and $\mu_{k+1}-\mu_k\sim\mu'(k)$, where $\mu(x)= c_\mu x^{a_0}\prod_{j=1}^{J_\mu}(\log^{(j)}x)^{a_j}$, with $a_0>0$.
Then $\frac{I_n}{EI_n}\stackrel{\text{dist}}{\to}\frac1\theta D_\theta^{(\mathcal{X})}$, with $\theta=\frac{N}{a_0}$, where $D_\theta^{(\mathcal{X})}$ is a random variable with the $\text{GD}^{(\mathcal{X})}(\theta)$ distribution.
**Remark 1. The condition on $\{\mu_k\}$ in part (i-a) is just a very weak regularity requirement on its growth rate (recall that $1\le \mu_k<k-1$). The condition in part (i-b) is fulfilled if $\mu_k\sim\mu(k)$ and $j_{k+1}-j_k\sim\mu'(k)$, where $\mu(x)= c_\mu \prod_{j=1}^{J_\mu}(\log^{(j)}x)^{a_j}$ with $J_\mu\ge0$.**
**Remark 2. Note that the random variable $\mathcal{X}$ in part (ii) takes on no more than $N$ distinct values.**
Assume first that the condition in part (i-a) holds. We claim that since $\{\frac{\mu_n}{\sum_{k=1}^n\frac{\mu_k}k}\}_{n=1}^\infty$ is bounded, there exists a sequence of positive integers $\{\gamma_n\}_{n=1}^\infty$ satisfying $\lim_{n\to\infty}\gamma_n=\infty$ and such that $\{\frac{\mu_n}{\sum_{k=\gamma_n+1}^n\frac{\mu_k}k}\}_{n=1}^\infty$ is also bounded. Indeed, assume to the contrary. Then, in particular, $\{\mu_n\}_{n=1}^\infty$ is unbounded. Also, since $\mu_k<k$, we have $\sum_{k=1}^n\frac{\mu_k}k<\gamma_n+\sum_{k=\gamma_n+1}^n\frac{\mu_k}k$, and it would follow that $\{\frac{\mu_n}{\gamma_n}\}_{n=1}^\infty$ is bounded for all sequences $\{\gamma_n\}_{n=1}^\infty$ satisfying $\lim_{n\to\infty}\gamma_n=\infty$, which is a contradiction.
Let $\{\gamma_n\}_{n=1}^\infty$ be such a sequence. Then $$M_n=\sum_{k=1}^n\frac{|E_k|}k\mu_k\ge(\min_{k>\gamma_n}|E_k|)\sum_{k=\gamma_n}^n\frac{\mu_k}k.$$ Thus, the condition in (i-a) guarantees that holds.
Now assume that the condition in part (i-b) holds. Since $M_n\ge\sum_{k=1}^n\frac{\mu_k}k$, it follows again that holds.
Thus, assuming either (i-a) or (i-b), it follows from part (i-a) of Theorem \[1\] that $\frac{I_n}{EI_n}\stackrel{\text{dist}}{\to}1$.
Now assume that the condition in part (ii) holds. Then $p_k=\frac Nk$, for large $k$, and $\mu_k\sim c_\mu k^{a_0}\prod_{j=1}^{J_\mu}(\log^{(j)}k)^{a_j}$, with $a_0>0$. Thus, $$M_n=\sum_{k=1}^n\frac{|E_k|}k\mu_k\sim\frac {Nc_\mu}{a_0}n^{a_0}\prod_{j=1}^{J_\mu}(\log^{(j)}n)^{a_j},$$ and $\lim_{k\to\infty}\frac{\mu_k}{M_k}=\frac{a_0}N$. Also, if the condition in part (ii) holds, then $\mu_{k+1}-\mu_k\sim a_0c_\mu k^{a_0-1} \prod_{j=1}^{J_\mu}(\log^{(j)}k)^{a_j}$. Thus, $\lim_{k\to\infty}\frac{p_k\mu_k}{\mu_{k+1}-\mu_k}=\frac N{a_0}$. We conclude from part (ii) of Theorem \[1\] that $\frac{I_n}{EI_n}\stackrel{\text{dist}}{\to}\frac1\theta\text{GD}^{(\mathcal{X})}(\theta)$, where $\theta=\frac N{a_0}$.
Proof of Theorem \[1\] {#pfthm1}
======================
Since $EW_n=1$, for all $n$, the distributions of $\{W_n\}_{n=1}^\infty$ are tight. Thus, since the random variables are nonnegative, it suffices to show that their Laplace transforms $E\exp(-\lambda W_n)$ converge under the conditions of part (i) to $\exp(-\lambda c)$, for the specified value of $c$, and under the conditions of part (ii) to $\exp(\theta\int_0^1\frac{Ee^{-L\lambda x\mathcal{X}}-1}xdx)$, which is the Laplace transform of $LD^{(\mathcal{X})}(\theta)$.
*Proof of part (i).Note that part (i-a) is the particular case of part (i-b) in which one can choose $K_n=n$, and then holds with $c=1$. Thus, it suffices to consider part (i-b). We have for $\lambda>0$, $$\label{LT}
\begin{aligned}
&E\exp(-\lambda W_n)=\\
&=\prod_{k=1}^nE\exp(-\frac\lambda{M_n} B_kX_k)=\prod_{k=1}^n\Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big)=\\
&\prod_{k=1}^{K_n}\Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big)\prod_{k=K_n+1}^n\Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big).
\end{aligned}$$ Since $$\prod_{k=K_n+1}^n(1-p_k)\le\prod_{k=K_n+1}^n\Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big)\le1,$$ it follows from assumption that $$\label{Knassump}
\lim_{n\to\infty}\prod_{k=K_n+1}^n\Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big)=1.$$*
Applying the mean value theorem to $E\exp(-\frac\lambda{M_n}X_k)$ as a function of $\lambda$, and recalling that $\mu_k=EX_k$, we have $$\label{Taylor}
\frac{\lambda}{M_n}EX_k\exp(-\frac{\lambda}{M_n}X_k) \le1-E\exp(-\frac\lambda{M_n}X_k)\le\lambda\frac{\mu_k}{M_n}.$$ In light of and the assumption that $\{\frac{X_k}{\mu_k}\}_{k=1}^\infty$ is uniformly integrable, it follows that for all $\epsilon>0$, there exists an $n_\epsilon$ such that $$\label{fromwc}
\begin{aligned}
&\frac\lambda{M_n}EX_k\exp(-\frac{\lambda}{M_n}X_k)=\lambda\frac{\mu_k}{M_n} E\frac{X_k}{\mu_k}\exp(-\lambda\frac{\mu_k}{M_n}\frac{X_k}{\mu_k})\ge
(1-\epsilon)\lambda\frac{\mu_k}{M_n},\\
& 1\le k\le K_n,\ n\ge n_\epsilon.
\end{aligned}$$ Thus, and yield $$\label{key}
(1-\epsilon)\lambda\frac{\mu_k}{M_n} \le1-E\exp(-\frac\lambda{M_n}X_k)\le\lambda\frac{\mu_k}{M_n},\ 1\le k\le K_n,\ n\ge n_\epsilon.$$ Since for any $\epsilon>0$, there exists an $x_\epsilon>0$ such that $-(1+\epsilon)x\le \log (1-x)\le -x$, for $0<x<x_\epsilon$, it follows from and that there exists an $n'_\epsilon$ such that $$\label{finallog}
\begin{aligned}
&-(1+\epsilon)\lambda p_k\frac{\mu_k}{M_n}\le\log\Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big) \Big)\le-(1-\epsilon)\lambda p_k\frac{\mu_k}{M_n},\\
& 1\le k\le K_n, \ n\ge n'_\epsilon.
\end{aligned}$$
From we have $$\label{applylog}
\begin{aligned}
&-(1+\epsilon)\lambda\frac{\sum_{k=k_\epsilon}^{K_n}p_k\mu_k}{M_n}\le \log\prod_{k=1}^{K_n}\Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big)\le\\
&-(1-\epsilon)\lambda\frac{\sum_{k=k_\epsilon}^{K_n}p_k\mu_k}{M_n},\ n\ge n'_\epsilon.
\end{aligned}$$ If $$\label{expectationcalc}
c\equiv\lim_{n\to\infty}\frac{M_{K_n}}{M_n}= \lim_{n\to\infty}\frac{\sum_{k=1}^{K_n}p_k\mu_k}{M_n}$$ exists, then from , , and , along with the fact that $\epsilon>0$ is arbitrary, we conclude that $$\lim_{n\to\infty}E\exp(-\lambda W_n)=\exp(-\lambda c),$$ which proves that $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}c$. The rest of the results in part (i-b), concerning accumulation points, follow in the same manner.
*Proof of part (ii).From , we have $$\label{logLT}
\log E\exp(-\lambda W_n)=\sum_{k=1}^n\log \Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big).$$ Since by assumption $\lim_{k\to\infty}p_k=0$, for any $\epsilon>0$ there exists a $k_\epsilon$ such that $$\label{logapprox}
\begin{aligned}
&-(1+\epsilon)p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\le \log \Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big)\le\\
&-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big), \ k\ge k_\epsilon.
\end{aligned}$$*
We now show that for any $\epsilon>0$ there exists a $k'_\epsilon$ such that $$\label{Xkmuk}
(1-\epsilon)E\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X})\le E\exp(-\frac\lambda{M_n}X_k)\le(1+\epsilon)E\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X}),\ k\ge k'_\epsilon.$$ By assumption and the assumption that $\{\mu_n\}_{n=1}^\infty$ is increasing, there exists a $C$ such that $\frac{\mu_k}{M_n}\le C$, for $1\le k\le n$ and $n\ge1$. By assumption, $\frac{X_k}{\mu_k}\stackrel{\text{dist}}{\to}\mathcal{X}$. Without loss of generality, we assume that all of these random variables are defined on the same space and that $\frac{X_k}{\mu_k}\to \mathcal{X}$ a.s. For $\delta>0$, let $$A_{k;\delta}=\{\sup_{l\ge k}|\frac{X_l}{\mu_l}-\mathcal{X}|\le\delta\}.$$ Then $A_{k;\delta}$ is increasing in $k$ and $\lim_{k\to\infty}P(A_{k;\delta})=1$. We have $$\label{Adeltcompl}
\int_{A^c_{k;\delta}}\exp(-\frac\lambda{M_n}X_k)dP\le P(A^c_{k;\delta}),$$ and $$\label{Adelt}
\begin{aligned}
&\exp(-\lambda C\delta)\int_{A_{k;\delta}}\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X})dP
\le\int_{A_{k;\delta}}\exp(-\lambda\frac{\mu_k}{M_n}\frac{X_k}{\mu_k})dP\le\\
& \exp(\lambda C\delta)\int_{A_{k;\delta}}\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X})dP.
\end{aligned}$$ Now follows from and .
Letting $k^{''}_\epsilon=\max(k_\epsilon,k'_\epsilon)$, it follows from and that $$\label{goodlogapprox}
\begin{aligned}
&-(1+\epsilon)p_k\Big(1-(1-\epsilon)E\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X})\Big)\le \log \Big(1-p_k\big(1-E\exp(-\frac\lambda{M_n}X_k)\big)\Big)\le\\
&-p_k\Big(1-(1+\epsilon)E\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X})\Big),\ k\ge k^{''}_\epsilon.
\end{aligned}$$ From and we have $$\label{forRiemsum}
\begin{aligned}
&-\sum_{k=k^{''}_\epsilon}^np_k(1+\epsilon)\Big(1-(1-\epsilon)E\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X})\Big)+o(1)\le\log E\exp(-\lambda W_n)\le\\
&-\sum_{k=k^{''}_\epsilon}^np_k\Big(1-(1+\epsilon)E\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X})\Big),\ \text{as}\ n\to\infty.
\end{aligned}$$
Define $x_k^{(n)}=\frac{\mu_k}{M_n}$, $k^{''}_\epsilon\le k\le n$, and $\Delta_k^{(n)}=x_{k+1}^{(n)}-x_k^{(n)}=\frac{\mu_{k+1}-\mu_k}{M_n}$, $k^{''}_\epsilon\le k\le n-1$. Then we have $$\label{Riemannsum}
\begin{aligned}
&\sum_{k=k^{''}_\epsilon}^np_k\Big(1-(1\pm\epsilon)E\exp(-\lambda\frac{\mu_k}{M_n}\mathcal{X})\Big)=\\
&\sum_{k=k^{''}_\epsilon}^n\frac{1-(1\pm\epsilon)E\exp(-\lambda x^{(n)}_k\mathcal{X})}{x_k^{(n)}}\Delta_k^{(n)}\big(p_k\frac{\mu_k}{\mu_{k+1}-\mu_k}\big).
\end{aligned}$$ By assumption, $\{\mu_k\}_{k=1}^\infty$ is increasing; thus $\{x^{(n)}_k\}_{k=k^{''}_\epsilon}^n$ is a partition of $[\frac{\mu_{k^{''}_\epsilon}}{M_n},\frac{\mu_n}{M_n}]$. By assumption, $\lim_{n\to\infty}\frac{\mu_{k^{''}_\epsilon}}{M_n}=0$ and $\lim_{n\to\infty}\frac{\mu_n}{M_n}=L$. We now show that the mesh, $\max_{k^{''}_\epsilon\le k\le n-1}\Delta^{(n)}_k$, of the partition converges to 0 as $n\to\infty$. Let $\Delta^{(n)}_{j_n}= \max_{k^{''}_\epsilon\le k\le n-1}\Delta^{(n)}_k$, where $k^{''}_\epsilon\le j_n\le n$. Without loss of generality, assume either that $\{j_n\}$ is bounded or that $\lim_{n\to\infty}j_n=\infty$. In the former case it is clear that $\max_{k^{''}_\epsilon\le k\le n-1}\Delta^{(n)}_k=\Delta^{(n)}_{j_n}=
\frac{\mu_{j_n+1}-\mu_{j_n}}{M_n}\stackrel{n\to\infty}{\to}0$. Now consider the latter case. From assumption and the assumption that $\lim_{k\to\infty}p_k=0$, it follows that $\lim_{n\to\infty}\frac{\mu_{n+1}-\mu_n}{M_n}=0$. Then we have $$\max_{k^{''}_\epsilon\le k\le n-1}\Delta^{(n)}_k=\Delta^{(n)}_{j_n}=
\frac{\mu_{j_n+1}-\mu_{j_n}}{M_n}=\frac{\mu_{j_n+1}-\mu_{j_n}}{M_{j_n}}\frac{M_{j_n}}{M_n}\le
\frac{\mu_{j_n+1}-\mu_{j_n}}{M_{j_n}}\stackrel{n\to\infty}{\to}0.$$ Finally, we note that from we have $\lim_{k\to\infty}p_k\frac{\mu_k}{\mu_{k+1}-\mu_k}=\theta$. In light of these facts, along with , and the fact that $\epsilon>0$ is arbitrary, it follows that $$\lim_{n\to\infty}\log E\exp(-\lambda W_n)=\theta\int_0^L\frac{E\exp(-\lambda x\mathcal{X})-1}xdx=
\theta\int_0^1\frac{E\exp(-\lambda Lx\mathcal{X})-1}xdx,$$
$\square$
Proof of Theorem \[2\] {#pfthm2}
======================
We will assume that $J_p,J_\mu\ge1$ so that we can use a uniform notation, leaving it to the reader to verify that the proof also goes through if $J_p$ or $J_\mu$ is equal to zero.
First assume that holds. Then by the assumptions in the theorem, $$\begin{aligned}
&1\le J_p\le J_\mu;\\
&\mu_k\sim c_\mu\prod_{j=J_p}^{J_\mu}(\log^{(j)}k)^{a_j},\ a_{J_p}>0;\\
&p_k\sim c_p\big(x\prod_{j=1}^{J_p}\log^{(j)}k\big)^{-1};\\
&\mu_{k+1}-\mu_k\sim c_\mu a_{J_P}\thinspace\frac{(\log^{(J_P)}k)^{a_{J_p}-1}}{x\prod_{j=1}^{J_p-1}\log^{(j)}k}\thinspace
\prod_{j=J_p+1}^{J_\mu}(\log^{(j)}k)^{a_j}.
\end{aligned}$$ Thus, $$M_n=\sum_{k=1}^np_k\mu_k\sim c_\mu c_p\frac{(\log^{(J_p)}n)^{a_{J_p}}}{a_{J_p}}\prod_{j=J_p+1}^{J_\mu}(\log^{( j)}n)^{a_j}.$$ Consequently, $$\label{secondcond}
\lim_{k\to\infty}\frac{\mu_k}{M_k}=\frac{a_{J_p}}{c_p}\ \ \text{and}\ \ \lim_{k\to\infty}\frac{p_k\mu_k}{\mu_{k+1}-\mu_k}=\frac{c_p}{a_{J_p}}.$$ Thus, from part (ii) of Theorem \[1\] it follows that $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}\frac1{\theta}D_\theta$, where $\theta=\frac{c_p}{a_{J_p}}$.
Now assume that does not hold. We need to show that $\{K_n\}_{n=1}^\infty$ can be defined so that and hold, and so that holds with $c\in\{0,1\}$. We also have to show when $c=0$ and when $c=1$. Recall the definitions in . If $\{0\le j\le J_\mu:a_j\neq0\}$ is empty, or if it is not empty and $a_{\kappa_\mu}<0$, then $\{\mu_k\}_{k=1}^\infty$ is bounded. Therefore, and hold with $K_n=n$ and it follows from part (i-a) of Theorem \[1\] that $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}1$. Thus, from now on we assume that $\{0\le j\le J_\mu:a_j\neq0\}$ is not empty and that $a_{\kappa_\mu}>0$. In order to use uniform notation, we will assume that $\kappa_\mu>0$, leaving the reader to verify that the proof goes through if $\kappa_\mu=0$. Thus, we have $$\label{muformprelim}
\mu_k\sim\prod_{j=\kappa_\mu}^{J_\mu}(\log^{(j)}k)^{a_j},\ \ \kappa_\mu\ge1,\ a_{\kappa_\mu}>0.$$ In order to simplify notation, for the rest of this proof, we will let $\mathcal{L}_l(k)$ denote a positive constant multiplied by a product of powers (possibly of varying sign) of iterated logarithms $\log^{(j)}k$, where the smallest $j$ is strictly larger than $l$. The exact from of this expression may vary from line to line. Sometimes we will need to distinguish between two such expressions in the same formula, in which case we will use the notation $\mathcal{L}^{(1)}_l(k), \mathcal{L}^{(2)}_l(k)$. Thus, we rewrite as $$\label{muform}
\mu_k\sim(\log^{(\kappa_\mu)}k)^{a_{\kappa_\mu} }\mathcal{L}_{\kappa_\mu}(k),\ \ \kappa_\mu\ge1,\ a_{\kappa_\mu}>0.$$
If $\{0\le j\le J_p:b_j\neq1\}$ is empty, then the second condition in is fulfilled and we have $$\label{pkempty}
p_k\sim c_p\big(x\prod_{j=1}^{J_p}\log^{(j)}k\big)^{-1}.$$ Since we are assuming that does not hold, at least one of the other two conditions in must fail. This forces $\kappa_\mu\neq J_p$. (Recall that we are assuming that $\{0\le j\le J_\mu:a_j\neq0\}$ is not empty and that $a_{\kappa_\mu}>0$.)
Consider first the case that $\kappa_\mu>J_p$. Then from and we have $$\label{Mnfirst}
\begin{aligned}
&M_n=\sum_{k=1}^np_k\mu_k\sim
&(\log^{(J_p+1)}n)(\log^{(\kappa_\mu)}n)^{a_{\kappa_\mu}}\mathcal{L}_{\kappa_\mu}(n),
\ \text{where}\ \kappa_\mu\ge J_p+1.
\end{aligned}$$ From and it follows that and hold by choosing $K_n=n$. Thus, from part (i-a) of Theorem \[1\], $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}1$.
Now consider the case $\kappa_\mu<J_p$. Then from and we have $$\label{Mncase1}
\begin{aligned}
&M_n=\sum_{k=1}^np_k\mu_k\sim
&(\log^{(\kappa_\mu)}n)^{a_{\kappa_\mu}}\mathcal{L}_{\kappa_\mu}(n),
\ \text{where}\ \kappa_\mu\le J_p-1,
\end{aligned}$$ and for any $K_n$ satisfying $K_n\to\infty$ and $K_n\le n$, we have $$\label{nKn}
\sum_{k=K_n}^np_k\sim c_p\big(\log^{(J_p+1)}n-\log^{(J_p+1)}K_n\big)=c_p\log\frac{\log^{(J_p)}n}{\log^{(J_p)}K_n}.$$ From and we have $$\label{muMn}
\begin{aligned}
&\frac{\mu_{K_n}}{M_n}
\sim\Big(\frac{\log^{(\kappa_\mu)}K_n}{\log^{(\kappa_\mu)}n}\Big)^{a_{\kappa_\mu}}\frac{\mathcal{L}^{(1)}_{\kappa_\mu}(K_n)}{\mathcal{L}^{(2)}_{\kappa_\mu}(n)},\ \kappa_\mu\le J_p-1,\ a_{\kappa_\mu}>0;\\
&\frac{M_{K_n}}{M_n}
\sim\Big(\frac{\log^{(\kappa_\mu)}K_n}{\log^{(\kappa_\mu)}n}\Big)^{a_{\kappa_\mu}}\frac{\mathcal{L}^{(1)}_{\kappa_\mu}(K_n)}{\mathcal{L}^{(2)}_{\kappa_\mu}(n)},\ \kappa_\mu\le J_p-1,\ a_{\kappa_\mu}>0;\\
\end{aligned}$$ As we explain in some detail below, since $\kappa_\mu<J_p$, we can choose $\{K_n\}_{n=1}^\infty$ so that $$\label{Knbutn}
\lim_{n\to\infty}\frac{\log^{(J_p)}K_n}{\log^{(J_p)}n}=1\ \ \text{and}\ \
\lim_{n\to\infty}\Big(\frac{\log^{(\kappa_\mu)}K_n}{\log^{(\kappa_\mu)}n}\Big)^{a_{\kappa_\mu}}\frac{\mathcal{L}^{(1)}_{\kappa_\mu}(K_n)}{\mathcal{L}^{(2)}_{\kappa_\mu}(n)}=0.$$ From and -, we conclude that $\{K_n\}$ can be defined so that and hold, and so that holds with $c=0$. This proves that $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}0$.
To explain , note that $\frac{\mathcal{L}^{(1)}_{\kappa_\mu}(K_n)}{\mathcal{L}^{(2)}_{\kappa_\mu}(n)}\le(\log^{(\kappa_\mu+1)}n)^A$, for some $A>0$ and all large $n$. (Recall that the powers of the iterated logarithms in $\mathcal{L}_{k_\mu}^{(2)}$ can be negative.) Thus, in place of the second limit in , it suffices to show that $\delta_n\equiv \Big(\frac{\log^{(\kappa_\mu)}K_n}{\log^{(\kappa_\mu)}n}\Big)^{a_{\kappa_\mu}}(\log^{(\kappa_\mu+1)}n)^A\stackrel{n\to\infty}{\to}0$. We have $$\log^{(\kappa_\mu)}K_n=(\delta_n)^\frac1{a_{\kappa_\mu}}(\log^{(\kappa_\mu+1)}n)^{-\frac A{a_{\kappa_\mu}}}\log^{(\kappa_\mu)}n;$$ thus, $$\label{+1+2}
\frac{\log^{(\kappa_\mu+1)}K_n}{\log^{(\kappa_\mu+1)}n}=\frac{\log\delta_n}{a_{\kappa_\mu}\log^{(\kappa_\mu+1)}n}-\frac{A\log^{(\kappa_\mu+2)}n}{a_{\kappa_\mu}\log^{(\kappa_\mu+1)}n}+1.$$ Defining $K_n$ by choosing $\delta_n=(\log^{(\kappa_\mu+1)}n)^{-1}$, it follows from and the fact that $J_p\ge \kappa_\mu+1$ that the two equalities in hold.
We now consider the case that $\{0\le j\le J_p:b_j\neq1\}$ is not empty. Then in order to fulfill the second condition in , we have $b_{\kappa_p}<1$. We write $$\label{pknotempty}
p_k\sim c_p\big(x\prod_{j=1}^{\kappa_p-1}\log^{(j)}k\big)^{-1}\big(\log^{(\kappa_p)}k\big)^{-b_{\kappa_p}}\big(\prod_{j=\kappa_p+1}^{J_p}\log^{(j)}k\big)^{-b_j}.$$ From and it follows that $M_n=\sum_{k=1}^np_k\mu_k$ satisfies $$\label{Mncases}
M_n\sim\begin{cases}(\log^{(\kappa_\mu)}n)^{a_{\kappa_\mu}}\thinspace\mathcal{L}_{\kappa\mu}(n),\ \kappa_\mu<\kappa_p;\\
(\log^{(\kappa_p)}n)^{a_{\kappa_p}-b_{\kappa_p}+1}\thinspace\mathcal{L}_{\kappa_p}(n),\ \kappa_\mu=\kappa_p;\\
(\log^{(\kappa_p)}n)^{1-b_{\kappa_p}}\thinspace\mathcal{L}_{\kappa_p}(n),\ \kappa_\mu>\kappa_p,\end{cases}$$ and from it follows that for any $K_n$ satisfying $K_n\to\infty$ and $K_n\le n$, $$\label{nKnagain}
\begin{aligned}
&\sum_{k=K_n}^np_k\sim\\
&\frac{c_p}{1-b_{\kappa_p}}\Big[\big(\log^{(\kappa_p)}n\big)^{1-b_{\kappa_p}}\big(\prod_{j=\kappa_p+1}^{J_p}\log^{(j)}n\big)^{-b_j}
-\big(\log^{(\kappa_p)}K_n\big)^{1-b_{\kappa_p}}\big(\prod_{j=\kappa_p+1}^{J_p}\log^{(j)}K_n\big)^{-b_j}\Big].
\end{aligned}$$ From and we have $$\label{muMnagain}
\frac{\mu_{K_n}}{M_n}\sim\begin{cases}\big(\frac{\log^{(\kappa_\mu)}K_n}{\log^{(\kappa_\mu)}n}\big)^{a_{\kappa_\mu}}\thinspace\frac{\mathcal{L}^{(1)}_{\kappa_\mu}(K_n)}{\mathcal{L}^{(2)}_{\kappa_\mu}(n)},\ \kappa_\mu<\kappa_p;\\
\frac{\big(\log^{(\kappa_p)}K_n)^{a_{\kappa_p}}}{(\log^{(\kappa_p)}n\big)^{a_{\kappa_p}-b_{\kappa_p}+1}}
\thinspace \frac{\mathcal{L}^{(1)}_{\kappa_p}(K_n)}{\mathcal{L}^{(2)}_{\kappa_p}(n)},\ \kappa_\mu=\kappa_p;\\
\frac{\big(\log^{(\kappa_\mu)}K_n\big)^{a_{\kappa_\mu}}}{\big(\log^{(\kappa_p)}n\big)^{1-b_{\kappa_p}}}\thinspace\frac{\mathcal{L}^{(1)}_{\kappa_\mu}(K_n)}{\mathcal{L}^{(2)}_{\kappa_p}(n)},\ \kappa_\mu>\kappa_p.
\end{cases}$$
It is immediate and that if $\kappa_\mu\ge \kappa_p$, then and hold by choosing $K_n=n$. (For the case $\kappa_\mu=\kappa_p$, recall that $b_{\kappa_p}\in(0,1)$.) Thus, from part (i-a) of Theorem \[1\], $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}1$.
Now consider the case $\kappa_\mu< \kappa_p$. For simplicity, we will assume that the higher order iterated logarithmic terms do not appear; that is, we will assume from - that $$\label{easier}
\begin{aligned}
&\sum_{k=K_n}^np_k\sim\frac{c_p}{1-b_{\kappa_p}}\Big[\big(\log^{(\kappa_p)}n\big)^{1-b_{\kappa_p}}
-\big(\log^{(\kappa_p)}K_n\big)^{1-b_{\kappa_p}}\Big];\\
&\frac{\mu_{K_n}}{M_n}\sim\big(\frac{\log^{(\kappa_\mu)}K_n}{\log^{(\kappa_\mu)}n}\big)^{a_{\kappa_\mu}};\\
&\frac{M_{K_n}}{M_n}\sim\big(\frac{\log^{(\kappa_\mu)}K_n}{\log^{(\kappa_\mu)}n}\big)^{a_{\kappa_\mu}}.
\end{aligned}$$ The additional logarithmic terms can be dealt with similarly to the way they were dealt with for , as explained in the paragraph following . Applying the mean value theorem to the function $x^{1-b_{\kappa_p}}$, we obtain $$\label{MVT}
\big(\log^{(\kappa_p)}n\big)^{1-b_{\kappa_p}}
-\big(\log^{(\kappa_p)}K_n\big)^{1-b_{\kappa_p}}
=\frac{(1-b_{\kappa_p})\log^{(\kappa_p)}\frac n{K_n}}{(\log^{(\kappa_p)}n^*)^{b_{\kappa_p}}},$$ where $n^*\in(K_n,n)$. Since $\kappa_\mu<\kappa_p$, we can choose $K_n\to\infty$ such that $\lim_{n\to\infty}\frac{\log^{(\kappa_\mu)}K_n}{\log^{(\kappa_\mu)}n}=0$, but $\lim_{n\to\infty}\log^{(\kappa_p)}\frac {K_n}n=1$. For such a choice of $\{K_n\}$, it follows from , and that and hold, and that holds with $c=0$; thus, $\lim_{n\to\infty}W_n\stackrel{\text{dist}}{=}0$. $\square$
Basic Facts Concerning Generalized Dickman Distributions {#background}
========================================================
We proved in Theorem \[1\] that $\exp(\theta\int_0^1\frac{e^{-\lambda x}-1}xdx)$ is in fact the Laplace transform of a probability distribution, which we have denoted by GD$(\theta)$. In particular, if we let $X_k=\mu_k=k$ and $p_k=\frac\theta k$, in which case $M_n=\sum_{k=1}^np_k\mu_k=\theta n$, then it follows from Theorem \[1\] that $$\label{dickconv}
\hat W_n\equiv\theta W_n=\frac1n\sum_{k=1}^n kB_k\stackrel{\text{dist}}{\to}D_\theta,$$ where $D_\theta\stackrel{\text{dist}}{\sim} GD(\theta)$.
We now prove . Let $$J^+_n=\max\{k\le n: B_k\neq0\},$$ with $\max\emptyset\equiv0$. We write $$\label{Whatn}
\hat W_n\equiv\frac1 n\sum_{n=1}^n kB_k=\frac{J^+_n-1}n\Big(\frac1{J^+_n-1}\sum_{k=1}^{J^+_n-1}kB_k\Big)+
\frac{J^+_n}n,$$ where the first of the two summands on the right hand side above is interpreted as equal to 0 if $J^+_n\le1$. We have $$\label{JNasym}
\begin{aligned}
P(\frac{J^+_n} n\le x)=\prod_{k=[xn+1]}^n(1-\frac\theta k)\sim x^\theta,\ x\in(0,1).
\end{aligned}$$ Also, by the independence of $\{B_k\}_{k=1}^\infty$, we have $$\label{selfsim}
\frac1{J^+_n-1}\sum_{k=1}^{J^+_n-1}kB_k\mid \{J^+_n=k_0\}\stackrel{\text{dist}}{=}\frac1{k_0-1}\sum_{k=1}^{k_0-1}kB_k=\hat W_{k_0-1},\ k_0\ge2.$$ Letting $n\to\infty$ in and using , and , we conclude that holds, where $U$ is a distributed according to the uniform distribution on $[0,1]$, $D_\theta\stackrel{\text{dist}}{\sim}\text{GD}(\theta)$ and $U$ and $D_\theta$ on the right hand side are independent.
We now use to show that the GD$(\theta)$ distribution has a density function $p_\theta$ satisfying $p_\theta=c_\theta\rho_\theta$, for some $c_\theta>0$, where $\rho_\theta$ satisfies . Let $F_\theta(x)=P(D_\theta\le x)$ denote the distribution function for the GD$(\theta)$ distribution. Then from we have $$\label{Ftheta}
\begin{aligned}
&F_\theta(x)=P(D_\theta\le x)=P(U^\frac1\theta(D_\theta+1)\le x)=\int_0^1P(D_\theta+1\le xy^{-\frac1\theta})dy=\\
&\int_0^1F_\theta(xy^{-\frac1\theta}-1)dy.
\end{aligned}$$ For $x>0$, making the change of variables, $v=xy^{-\frac1\theta}-1$, we can rewrite as $$\label{Fthetaagain}
F_\theta(x)=\theta x^\theta\int_{x-1}^\infty F_{D_\theta}(v)(1+v)^{-1-\theta}dv,\ x>0.$$ From and the fact that $F_\theta(x)=0$, for $x\le0$, it follows that $F_\theta$ is continuous on $\mathbb{R}$. Also, since $F_\theta(x)=0$, for $x\le0$, we have $$\int_{x-1}^\infty F_{D_\theta}(v)(1+v)^{-1-\theta}dv=\int_0^\infty F_{D_\theta}(v)(1+v)^{-1-\theta}dv,\ x\le 1.$$ Consequently, it follows from that $F_\theta(x)=C_\theta x^\theta$, for $x\in[0,1]$, where $C_\theta=\theta\int_0^\infty F_{D_\theta}(v)(1+v)^{-1-\theta}dv$. From this and it follows that $F$ is differentiable on $(0,1)$ and on $(1,\infty)$, and that, letting $p_\theta=F_\theta'$, $$\label{density01}
p_\theta=c_\theta x^{\theta-1},\ 0<x<1,\ c_\theta=\theta^2\int_0^\infty F_{D_\theta}(v)(1+v)^{-1-\theta}dv,$$ and $$\label{density}
\begin{aligned}
&p_\theta(x)=\theta^2x^{\theta-1}\int_{x-1}^\infty F_{D_\theta}(v)(1+v)^{-1-\theta}dv-\theta x^{-1} F_\theta(x-1)=\\
&\frac\theta x (F_\theta(x)-F_\theta(x-1)), \ x>1.
\end{aligned}$$ From , it follows that $p_\theta$ is differentiable on $x>1$, and that $(xp_\theta(x))'=\theta\big(p_\theta(x)-p_\theta(x-1)\big)$, for $x>1$, or equivalently, $$\label{densityderiv}
xp'_\theta(x)+(1-\theta)p_\theta(x)+\theta p_\theta(x-1)=0,\ x>1.$$ From and we conclude that $p_\theta(x)=c_\theta\rho_\theta$, where $\rho_\theta$ satisfies . Integrating by parts in the formula for $c_\theta$ in shows that $$c_\theta=\theta\int_0^\infty(1+v)^{-\theta}p_\theta(v)dv=\theta E(1+D_\theta)^{-\theta}.$$
[99]{}
Arratia, R., Barbour, A. and Tavaré, S., *Logarithmic Combinatorial Structures: A Probabilistic Approach*, EMS Monographs in Mathematics, European Mathematical Society (EMS), Zurich, (2003).
de Bruijn, N., *On the number of positive integers $\leq x$ and free of prime factors $>y$*, Nederl. Acad. Wetensch. Proc. Ser. A. 54, (1951), 50–60.
Cellarosi, F. and Sinai, Y., *Non-standard limit theorems in number theory*, Prokhorov and Contemporary Probability Theory, 197–213, Springer Proc. Math. Stat., 33, Springer, Heidelberg, (2013).
Dickman, K., *On the frequency of numbers containing prime factors of a certain relative magnitude*, Ark. Math. Astr. Fys. 22, 1-14 (1930).
Hwang, H.-K. and Tsai, T.-H. *Quickselect and the Dickman function*, Combin. Probab. Comput. 11 (2002), 353-371.
Montgomery, H. and Vaughan, R., *Multiplicative Number Theory. I. Classical Theory*, Cambridge Studies in Advanced Mathematics, 97, Cambridge University Press, Cambridge, (2007).
Pinsky R., *A Natural Probabilistic Model on the Integers and its Relation to Dickman-Type Distributions and Buchstab’s Function*, preprint, (2016).
Penrose, M. and Wade, A., *Random minimal directed spanning trees and Dickman-type distributions*, Adv. in Appl. Probab. 36 (2004), 691–714.
Tenenbaum, G., *Introduction to Analytic and Probabilistic Number Theory*, Cambridge Studies in Advanced Mathematics, 46, Cambridge University Press, Cambridge, (1995).
| {
"pile_set_name": "ArXiv"
} |
---
author:
-
bibliography:
- 'main.bib'
title: 'AI-Mediated Exchange Theory'
---
<ccs2012> <concept> <concept\_id>10003120.10003130.10003131</concept\_id> <concept\_desc>Human-centered computing Collaborative and social computing theory, concepts and paradigms</concept\_desc> <concept\_significance>500</concept\_significance> </concept> </ccs2012>
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Learning abstract and systematic relations has been an open issue in neural network learning for over 30 years. It has been shown recently that neural networks do not learn relations based on identity and are unable to generalize well to unseen data. The Relation Based Pattern (RBP) approach has been proposed as a solution for this problem. In this work, we extend RBP by realizing it as a Bayesian prior on network weights to model the identity relations. This weight prior leads to a modified regularization term in otherwise standard network learning. In our experiments, we show that the Bayesian weight priors lead to perfect generalization when learning identity based relations and do not impede general neural network learning. We believe that the approach of creating an inductive bias with weight priors can be extended easily to other forms of relations and will be beneficial for many other learning tasks.'
author:
- |
Radha Kopparti\
Department of Computer Science\
City, University of London\
London, United Kingdom\
`[email protected]`\
Tillman Weyde\
Department of Computer Science\
City, University of London\
London, United Kigndom\
`[email protected]`\
bibliography:
- 'neurips\_2019.bib'
title:
- Bayesian Weight Priors for Learning Identity Relations
- Weight Priors for Learning Identity Relations
---
Introduction {#sec:intro}
============
Humans, including infants. are very effective at learning patterns of relations based on identity from sensory input and systematically applying them to new stimuli, even after a very brief exposure, while current neural networks in their standard form fail to detect these relations in unseen data [@GaryMarcus1999]. It has become evident that there are still relevant limitations to systematic generalization in current neural network architectures [@lake2018generalization; @GaryMarcus2018]. Neural networks are seen to be good at memorizing the numerical patterns seen in the training set but often not to extrapolate this representation outside the training set [@liska-et-al-2018-memorize]. It was found that standard neural networks do not seem to learn identity relations, i.e. the equality of two vectors [@weyde_kopparti_2018], which are fundamental for many higher level tasks.
A well known study in this direction was conducted by [@GaryMarcus1999], where a recurrent neural network failed to distinguish abstract patterns, based on equality or identity relations between the input stimuli, although seven-month-old infants showed the ability to distinguish them after a few minutes of exposure. This was followed by an lively exchange on rule learning by neural networks and in human language acquisition in general, where results by [@Elman1999; @Altmann2] and [@Shultz1] could not be reproduced by [@vilcu2001generalization; @vilcu2005two], and [@shultz2006neural] disputed claims by [@vilcu2005two]. Other approaches, such as those by [@Shastri-1999-spatiotemporal] and [@Dominey; @Alhama], use specialized network architectures or different problem formulations or evaluation methods.
Recently, the Relation Based Patterns (RBP) approach has been introduced as a way to create a suitable inductive bias for the problem of learning identity relations on binary vectors [@weyde_kopparti_2018; @weyde_kopparti_2019]. The task is to classify whether two halves of the input vectors are equal. i.e. $u_i = u_{n+i} \forall i \in \{1, \dots , n\}$ for an input vector with $2n$ dimensions. The RBP model introduced in [@weyde_kopparti_2018; @weyde_kopparti_2019] is based on the comparison of input neurons that correspond to each other in a relation, e.g. the corresponding dimensions in a pair of binary vectors. For the comparison they introduce Differentiator-Rectifier (DR) units, which calculate the absolute difference of two inputs: $f(x,y) = |x-y|$. For each dimension of the input vectors, a DR unit is introduced. This simplifies the learning problem because the summation of activations of the DR units is sufficient for generalisable identity detection. The results show no restriction on the learning of other tasks in practice.
There are different ways of integrating DR units into neural networks in RBP: *Early* and *Mid* Fusion. In *Early Fusion*, DR units are concatenated to input units, and in *Mid Fusion* they are concatenated to the hidden layer. In both cases, the existing input and hidden units are unchanged.
Adding units with hard-wired connections and unusual activation function limits the flexibility of the RBP approach. In this work, we introduce a modified RBP structure that can be formulated as a Bayesian prior to model the weight structure of Mid Fusion RBP in a standard feed forward network setting.
A Bayesian Approach to Relation Based Patterns
==============================================
In the weight prior approach, we replace each DR units with two standard neurons and model the fixed weights with a default weight matrix $D$. This matrix contains the weights that enable a dimension-wise comparison of the inputs.
A diagram of the structure is given in Figure \[fig:rbp4\](a) where $\alpha$ and $\beta$ are the two input values being compared. A small example of this default matrix is shown in Figure \[fig:rbp4\]b). The incoming connections from two corresponding inputs to a neuron in the hidden layer to be compared have values of 1 and $-1$. For the same pair we use another hidden neuron with inverted signs of the weights, as in rows 1 and 2. With a ReLU activation function, this means that the value of one of the hidden neurons will be positive, if one input neurons have different activations. We therefore need at least $n$ hidden units in the first hidden layer for a comparison of n two $n/2$-dimensional vectors. All other incoming connection weights are set to 0, including the bias. This ensures that in all cases where corresponding inputs are not equal, there will be a positive activation in one of the neurons.
------------------------------------------------------------------------ --------------------------------------------------
a\) $\vcenter{\hbox{{\includegraphics[width=5cm]{rbp_init1_1.png} }}}$ b\) [$\displaystyle
D = {\text{\boldmath$
\begin{pmatrix}
+1 & 0 & 0 & -1 & 0 & 0 \\
-1 & 0 & 0 & +1 & 0 & 0 \\
0 & +1 & 0 & 0 & -1 & 0 \\
0 & -1 & 0 & 0 & +1 & 0 \\
0 & 0 & +1 & 0 & 0 & -1 \\
0 & 0 & -1 & 0 & 0 & +1 \\
0 & 0 & 0 & 0 & 0 & 0 \\
&\vdots & & &\vdots & \\
\end{pmatrix}$}}
$]{}
------------------------------------------------------------------------ --------------------------------------------------
The matrix $D$ is then used to define default values of the network weights, i.e. we impose a loss based on the difference between $D$ and the actual weight matrix $W$, that connects the input neurons to the first hidden layer. This loss term $l_{RBP_{1/2}}$ for L1 or L2 loss are defined as: $$l_{RBP_1} = \sum_{i=1}^k |W_k - D_k|, \hspace{1cm} l_{RBP_2} = \sum_{i=1}^k (W_k - D_k)^2,$$ where $k$ is the number of elements in $W$. These loss functions correspond to Bayesian priors on the weights with the mean defined by the values of $D$. The $L2$ loss corresponds to a Gaussian and $L1$ loss to a Laplacian prior, such that backpropagation maximizes the posterior likelihood of the weights given the data [@williams1995bayesian]. The overall training loss $l_t$ is defined as $$l_t = l_c + \lambda \times l_{RBP}$$ where $l_c$ is the cross entropy and $\lambda$ is the regularization parameter, corresponding to the inverse of the variance of the prior, effectively regulating the strength of the RBP regularization. We call these methods of embedding the RBP into a standard network ERBP L1 and ERBP L2 respectively.
Generalizing and Learning of Identity Relations
===============================================
For the task of learning identity relations, we generate synthetic data and use a standard feed-forward neural network. The input vector is binary and the target values are $[0,1]$ for unequal and $[1,0]$ for equal vector halves as described in section \[sec:intro\].
We use a grid search over hyper-parameters: the number of epochs \[10,20,30\], the number of neurons per hidden layer was varied as \[10,20,30\]. For the Bayesian weight prior, we varied the regularization parameter $\lambda$ with values \[0.01,0.03,0.1,0.3,1,3,10,30\]. We ran a total of 10 simulations using the SGD and Adam optimizers, training for 20 epochs. We used a single hidden layer and a batch size of 1 unless indicated otherwise. The networks have been implemented using the PyTorch library[^1].
The train/test split was set to 75% / 25% for all tasks. We tested with vector dimensionalities $n=3,10,30$. We generate all vectors with equal halves and take a random sample of all those with unequal halves to balance the classes. We downsample the size of the dataset to 1000 when it is greater.
Identity Relations
------------------
Identity is an abstract relation in the sense that it is independent of the actual values of the individual arguments, it just depends on their combined configuration. In this task, pairs of vectors are presented to a feed-forward network and the task is to distinguish whether the two vectors are equal or not. We evaluated the performance of the network in different configurations on a held-out test set and the results are tabulated in Table \[tab:exprs\] for different vector dimensions. As observed by [@weyde_kopparti_2018], standard networks do not improve much over random guessing, while ERBP L1 and ERBP L2, as well as Mid Fusion, almost always achieve perfect generalizsation with sufficiently strong $\lambda$ (see next section for details).
Type Standard Early Fusion Mid Fusion ERBP L1 ERBP L2
-------- ----------- -------------- ------------ ------------ ------------
$n=3$ 55 (1.91) 65 (1.34) 100 (0.04) 100 (0.00) 100 (0.00)
$n=10$ 51 (1.67) 65 (1.32) 100 (0.08) 100 (0.04) 100 (0.02)
$n=30$ 50 (1.52) 65 (1.27) 100 (0.07) 100 (0.05) 100 (0.04)
: Test set classification accuracy (in %) and standard deviation over 10 simulations (in brackets) using different models for Identity Learning (vector dimensions $n=3,10,30$). The networks were trained with the Adam optimizer for 20 epochs. []{data-label="tab:exprs"}
Parameter Variations
--------------------
We study the effect of several parameters: number of hidden layers, choice of optimizer, regularization factor and weight initialization on identity relation learning tasks.
**Network Depth**: We tested identity learning with deeper neural network models, using $h = {2,3,4,5}$ hidden layers. The results are tabulated in Table \[tab:hid\], showing only minor improvements in the network performance for deeper networks. However, ERBP L1 and L2 generalization is consistent and independent of the network depth.
-------- ----------- ----------- ------------ ------------ ------------
Hidden No Early Mid ERBP L1 ERBP L2
layers RBP Fusion Fusion
h = 2 55 (1.65) 65 (1.26) 100 (0.02) 100 (0.00) 100 (0.00)
h = 3 55 (1.67) 67 (1.14) 100 (0.03) 100 (0.00) 100 (0.00)
h = 4 58 (1.63) 68 (1.25) 100 (0.02) 100 (0.00) 100 (0.00)
h = 5 59 (1.68) 72 (1.23) 100 (0.02) 100 (0.00) 100 (0.00)
-------- ----------- ----------- ------------ ------------ ------------
: Test set classification accuracy (in %) with standard deviation (in brackets) for identity learning ($n=3$) using deeper networks. The networks were trained with the Adam optimizer for 20 epochs. []{data-label="tab:hid"}
Type ERBP L1 ERBP L2
------ ------------ ------------
Adam 100 (0.00) 100 (0.00)
SGD 98 (0.06) 96 (0.05)
: Accuracy (in %) and standard deviation of identity learning ($n=3$) using Adam and SGD optimizer for ERBP L1 and L2. SGD can also lead to 100% accuracy on the test set, but needs higher $\lambda$ values.[]{data-label="tab:opt1"}
**Optimiser**: We used both Stochastic Gradient Descent (SGD) and the Adam optimizer [@adam_ref] for training the ERBP L1 and L2. We observed faster convergence and greater improvement in the overall accuracy with the Adam compared to SGD. We observed similar results for both ERBP L1 and L2. Table \[tab:opt1\] below summarizes the results of identity learning for both the optimizers with the regularization parameter $\lambda$ set to 1. We observe that the SGD does not reach full generalization in this setting, however it does so at higher values of $\lambda$.
**Regularization Factor**: We varied the regularization factor $\lambda$ in the loss function of the ERBP models. We observed that a factor of 3 or above reliably leads to perfect generalization in identity learning task. Figure \[fig:example3\], shows how the effect depends on the size the regularization factor $\lambda$ using L1 and L2 loss functions for learning identity relations.
![Test set classification accuracy (in %) of the network with ERBP L1 and L2 when varying the regularization parameter $\lambda$ (shown in logarithmic scale) for identity learning with ($n=3$) and using the Adam optimizer. []{data-label="fig:example3"}](equality_plot.png){width="7cm"}
Jointly learning identity relations and bit patterns
----------------------------------------------------
In order to test, whether the RBP and ERBP impedes other learning tasks, we also tested learning with some non-relational patterns that are based on values of specific neurons. The simplest form is that we have one bit that determines the target class. This case is learned with 100% generalization performance for a network with 2 additional outputs for the pattern classification.
Furthermore, we tested the learning of classification based on all even or odd elements in the vector being $0$. In this case, we also get perfect generalization for $\lambda$ up to 10. For $\lambda = 30$ we see first deterioration of the accuracy, but there is a wide range of values where both the identity relations and the even/odd classification generalize perfectly.
Conclusions
===========
Identity based relations are a fundamental form of relational learning. In this work, we re-visit the problem of learning identity relations using standard neural networks and show that creating a weight prior on the network weights leads to generalisable solutions of learning identity based relations. This also did not affect learning of non-relational patterns in our preliminary experiments, although more thorough testing remains to be done here. We believe that addressing these issues and coming up with effective solutions is necessary for more higher level relational learning tasks and also is relevant to address problems in general neural network learning. In future, we would like to extend this work towards learning other complex relational learning tasks and come with effective ways of creating an inductive bias using weight priors within the standard neural network architectures.
[^1]: <http://pytorch.org>
| {
"pile_set_name": "ArXiv"
} |
---
abstract: |
We investigate the properties of the $B$-band Tully-Fisher (T-F) relation for 25 compact group galaxies, using $V_{\rm max}$ derived from 2-D velocity maps. Our main result is that the majority of the Hickson Compact Group galaxies lie on the T-F relation. However, about 20% of the galaxies, including the lowest-mass systems, have higher $B$ luminosities for a given mass, or alternatively, a mass which is too low for their luminosities. We favour a scenario in which outliers have been brightened due to either enhanced star formation or merging. Alternatively, the T-F outliers may have undergone truncation of their dark halo due to interactions. It is possible that in some cases, both effects contribute. The fact that the $B$-band T-F relation is similar for compact group and field galaxies tells us that these galaxies show common mass-to-size relations and that the halos of compact group galaxies have not been significantly stripped inside $R_{25}$.
We find that 75% of the compact group galaxies studied (22 out of 29) have highly peculiar velocity fields. Nevertheless, a careful choice of inclination, position angle and center, obtained from the velocity field, and an average of the velocities over a large sector of the galaxy enabled the determination of fairly well-behaved rotation curves for the galaxies. However, two of the compact group galaxies which are the most massive members in M51–like pairs, HCG 91a and HCG 96a, have very asymmetric rotation curves, with one arm rising and the other one falling, indicating, most probably, a recent perturbation by the small close companions.
author:
- 'C. Mendes de Oliveira , P. Amram , H. Plana'
- 'C. Balkowski'
title: |
Dynamical effects of interactions and the\
Tully-Fisher relation for Hickson compact groups
---
Introduction
============
The influence of environmental effects on the internal dynamics and matter distribution of compact group galaxies has not yet been clearly established, mostly due to lack of reliable kinematic data. A recent study of the internal kinematics of 30 galaxies by Nishiura et al. (2000, hereafter N2000) found that asymmetric and peculiar rotation curves are more frequently seen in the Hickson Compact Groups (HCG) spiral galaxies than in field or cluster spirals and the dynamical properties of the galaxies do not seem to correlate with any group or galaxy parameter. An older but very influential study is that by Rubin et al. (1991, hereafter R1991), who analyzed rotation curves for 32 Hickson compact group galaxies and found that 2/3 of them had peculiar rotation curves. For the subsample of galaxies for which rotation curves were possible to be derived, they found a large offset of the T-F relation with respect to the field relation in the sense that galaxies in compact groups have “too low velocities for their luminosities or, alternatively, luminosities which are overbright for their rotation velocities.” This suggested to the authors that spiral galaxies in compact groups have low mass-to-light ratios compared to field galaxies by about 30% which, in turn, could be explained if compact group galaxies have smaller dark halos than their field counterparts. Given that compact groups are environments where tidal encounters are common, it may be expected that interactions have stripped or disrupted the galaxy dark halos at some level. These conclusions have important consequences for the determination of group lifetimes and understanding how compact groups evolve and eventually merge.
We revisit this important problem using a new dataset for 25 galaxies, of which 13 are in common with either the sample of R1991 or N2000 or both. In this paper we re-examine the T-F relation for compact group galaxies. Our study differs from the previous work in that it uses rotation curves obtained from 2-dimensional velocity fields. In a number of cases this allows a more accurate determination of the rotation curves than was possible with slit spectroscopy. Additionally, a fuller characterization of the kinematics of each galaxy is possible. We show that, with these new data, the T-F relation for compact group galaxies is similar to that for galaxies in less dense environments. The exceptions to this conclusion are seen for some of the least massive galaxies in our sample.
This paper is organized as follows. In Section 2 we present the set of rotation curves that are used. In Section 3 we illustrate and discuss comparisons of rotation curves for several galaxies in common with R1991 and N2000 and we show that 2-D spectroscopy is needed if one wants to accurately describe the kinematics of interacting galaxies. Sections 4 and 5 show the results on the T-F relation and a discussion respectively.
Data
====
The data used in this paper are gathered from three publications which studied the kinematics of galaxies in nine compact groups (HCG 10, 16, 19, 87, 88, 89, 90, 91, 96, 100). A detailed description of the observations and data reduction can be found in Mendes de Oliveira et al. (1998), Plana et al. (2003) and Amram et al. (2003). In summary, our dataset consists of H$\alpha$ emission-line velocity maps obtained with a Fabry-Perot system, from which rotation curves were obtained. For the T-F study we excluded from the sample all the elliptical galaxies. We also excluded galaxies HCG 10a, 10c, 16d and 100a because their rotation curves have a very short extension, well short of R$_{25}$ and galaxies HCG 16b and HCG 100b due to their extremely peculiar rotation curves. We included HCG 19a in our sample, although Hickson (1993) classifies it as an E2 galaxy, because we have reclassified it as an S0 based on its kinematic properties. In addition, we included in our sample unpublished data for two other galaxies (HCG 07c and HCG 79d).
We show in Figs. 1a and 1b the rotation curves for the galaxies in the sample studied in this paper. The x-axis plots r/R$_{25}$, the galaxy radius along its major axis normalized by R$_{25}$, the length of the major axis at the 25 mag arcsec$^{-2}$ isophote (as given by Hickson 1993; note, however that in Plana et al. 2003 and Amram et al. 2003, the value for R$_{25}$ was taken from de Vaucouleurs et al. 1991, the RC3, resulting in slightly different numbers from those used here). In order to obtain a V$_{max}$ for use in the T-F relation (represented by a black dot in each subpannel of Figs. 1a and 1b) we have computed the maximum velocity of the average velocity curve (single average of the values for the receding and approaching sides). V$_{max}$ is not the best kinematic parameter to be used in the T-F relation, since V$_{flat}$, the velocity of the flat portion of the curve or V$_{2d}$, the velocity at two times the galaxy scale length are known to yield less scatter in the T-F relation. However, the use of V$_{max}$ allowed us to use the largest possible number of compact group galaxies.
The control sample used for the T-F comparison was the sample of Sb and Sc galaxies from Courteau (1997), where we used their “Vmax” as the equivalent to the maximum velocity determined by us. We also used the sample of galaxies in the Ursa Major cluster (Verheijen 2001), in order to fill in the lower end of the mass sequence in the T-F diagram. We note that our values of V$_{max}$ are adjusted by the cosmological correction (1+z) as are also the values given for the galaxies in the control samples. All velocities were corrected for the same Virgocentric infall (Paturel et al. 1997) when distances were obtained using the Hubble law (for our sample and Courteau’s sample).
We note that in Fig. 1 the rotation of HCG 87a is one-sided because half of this almost edge-on galaxy is not observed due to a strong dust lane.
Comparisons with previous long-slit spectroscopy
================================================
The 2-D velocity maps obtained with a Fabry-Perot instrument allowed us to mimic a slit through our data in order to recover the rotation curves obtained from long-slit spectroscopy. In this way, we could make a direct comparison with previous rotation curves obtained by other authors in order to investigate the reason for existing differences in the shapes and amplitudes of the rotation curves between different studies, for a given galaxy.
We give an example in Fig. 2, where it is clear why with long-slit spectroscopy the derived rotation curves do not always correspond to the overall kinematics of the galaxies. The rotation curve derived by R1991 for HCG 88a (reproduced in Fig. 2) shows asymmetries with disagreement between the two sides of the galaxy (a bifurcation of the curve at a radius of 10 arcsec). We overplot on Fig. 2 the Fabry-Perot data for HCG 88a, with values restricted to a narrow cone around the galaxy major axis (to mimic a slit), using the particular values of center, inclination and position angle derived by R1991 photometrically. Inspecting the figure we see that for radii larger than 10 arcsec, there is good agreement between the slit-spectroscopy and Fabry-Perot data, if the photometrically derived center, inclination and position angle are used. However, the overall shape of the curve changes when such parameters are derived from the velocity map and a large sector of the galaxy is averaged (the resulting rotation curve is shown in Plana et al. 2003 and in Fig. 1). In particular, if the kinematic parameters are used and velocities over a large sector of the galaxies are averaged, the bifurcation present in Fig. 2, at r $\sim$ 10 arcsec, disappears, and the galaxy has normal kinematics, with both sides of the curve matching.
In order to understand the magnitude of the variations of the parameters measured from the kinematic and photometric data we list in Table 2 values for the inclination, PA and differences between centers measured from the continuum images and velocity fields. Detailed discussion on the method used to obtain the center, PA of the major axis, and inclination of the galaxies are given in Amram et al. (1996). We included in the table not only the galaxies used in the Tully-Fisher relation but also those that were considered too peculiar or for which the gas extension was too short. We list in the last four columns the values of inclination and PA given by R1991 and N2000, when available.
A comparison of the inclinations derived from the Fabry-Perot map and from the continuum images was done by subtracting columns 2 and 3 of Table 2 (and normalized to one of the two). The differences spread approximately around zero, with an rms of 20% (for the galaxies used in the Tully-Fisher relation). This indicates that there is no systematic error introduced in the determination of the maximum velocities due to the measurement of the inclination from either photometric or kinematic data. However, a similar comparison with the smaller subsample in common with R1991 and N2000 show a slight overestimate of the inclinations obtained by these authors, as compared to our estimates. This would indicate that the V$_{max}$ derived by them would be expected to be lower on average than those obtained by us (and in fact, the values of V$_{max}$ derived by R1991 in particular tend to be lower than those derived in our study).
A comparison between columns 4 and 5 of Table 2, of the kinematic and photometric PA’s for the galaxies used in the T-F analysis, shows that about 1/3 of the galaxies have misalignments (greater than 10 degrees) between the kinematic and photometric axes. This may be an indication that the gas is not in equilibrium in these galaxies (HCG 10d, 16c, 19a, 88b, 89d, 91a, 91c, 96a). If that is the case, and for an example the gas is collapsing and/or there is additional dispersion, then the V$_{max}$ we compute using the rotation curve of the gas is a lower limit to the real V$_{max}$ (indicated by the total mass). In other words, the real mass of the galaxy would then be higher than what we infer from our measurements. We note that four of the six galaxies that were too peculiar or had too short gas extensions to enter the T-F analysis also present misalignments between the kinematic and photometric axes (HCG 16b, 16d, 10c, 100b).
Similar results are obtained when comparing the kinematic PA’s to the corresponding photometric values derived from either R1991 and/or N2000 for the smaller subsample of galaxies in common (comparing columns 4 and 8 and columns 4 and 10 of Table 2).
We are not able to make a direct comparison between the centers we measure with those measured by R1991 and/or N2000 since we do not know the exact value they used (this problem is related to the fact that we do not have information from our Fabry-Perot data about the systemic velocities of the galaxies, given the scanning nature of the instrument, Amram et al. 1992, and also related to the fact that we do not know exactly where they positioned the slits). We then list in column 6 of Table 2 only the difference in arcsec between the measured kinematic and photometric center. The center of the continuum image was taken as the point of maximum intensity. The center of the velocity field was chosen to be a point along the major axis which made the rotation curve symmetric or as symmetric as possible (with similar amplitudes for the receding and approaching sides). This freedom with the Fabry-Perot data was, in fact, one reason why it was possible to construct fairly well behaved rotation curves even when the velocity maps were highly peculiar due to localized non-circular motions. The values for $\vert$$\Delta_{kin-cont}$$\vert$ varied from zero to six arcseconds.
Our conclusion is that the velocity fields of compact group galaxies are, in several cases, significantly affected by non-circular motions, local asymmetries and misalignments between the kinematic and stellar axes. A fine tuning of map parameters (center, inclination and position angle of the velocity fields) and an average of the velocities over a large sector of the galaxy are needed to derive reliable rotation curves and representative values of V$_{max}$.
A few other comparisons with data from R1991 and N2000 are exemplified in Fig. 3. In all cases we plot rotation curves resulting from a cut through the Fabry-Perot data, using the parameters given by the author’s (photometric parameters). An important point becomes clear when inspecting Fig. 3, especially Figs. 3a, 3b and 3f. In these cases the rotation curves derived from the 2-D maps extend beyond the long-slit ones, which is also another reason for the discrepant values of derived V$_{max}$, using these two different techniques, Fabry-Perot or long-slit. We should have in mind, however, that an over/underestimation of the inclination leads to an over/underestimation of the extension of the rotation curves.
Another point we should note is that the rotation curves derived by Fabry-Perot spectroscopy are much more well behaved in the sense that more often the two sides of the curves approximately agree and they are either flat or rising in the last measured point. In fact, we have no example of galaxy with a rotation curve that drops in both sides as one would expect in the Keplerian case.
Results
=======
Table 1 summarizes the main parameters for each galaxy. In columns (1) and (2) we show the identification and length of semi-major axis of the galaxy as given by Hickson (1993). The morphological classification for the galaxies taken from Hickson (1993) and from NED are given in columns (3) and (4) respectively . The values for the total B magnitudes of the galaxies, listed in column (5), B$_{Tcor}$, were obtained from the B$_T$ listed in Hickson et al. (1989), corrected by galactic extinction (Schlegel et al. 1998) and by the extinction due to inclination (Tully et al. 1998). The same corrections were also applied to the magnitudes of the galaxies in the control samples. The velocities listed in (6) are the maximum rotational velocities obtained from the average velocity curves as shown in Fig. 1. In (7) we list vvir, the heliocentric radial velocity of the galaxy, corrected for the local group infall onto Virgo, from the Hyperleda database. The absolute magnitudes in column (8) were obtained from $M=B_{Tcor}-5*LG(vvir/75)$. Finally, in column (9) we marked with a $P$ those galaxies whose velocity fields were considered highly disturbed either by Amram et al. (2003), Plana et al. (2003) or Mendes de Oliveira et al. (1998).
The morphological types are represented in the sample in the following proportion $S0:Sa:Sb:Sc:Sd:Sm:I$=1:2:3:12:3:2:2, with preponderance of the Sc and Scd types (both counted in the Sc bin). We note that two galaxies, HCG 16c and HCG 96d, were classified as “irregular” by Hickson (1993), indicating that they had peculiar morphologies and not that they are truly irregular galaxies. The magnitudes of the galaxies ranged from $M_B$ $\sim$ –18.0 to –22.0 $+$ 5 log $h_{75}$.
We show in Fig. 4 the T-F relation in the $B$ band for the Hickson compact group galaxies, for the Sb and Sc galaxies in the lower density environments studied by Courteau (1997), and for Sb-Sd galaxies in the Ursa Major cluster studied by Verheijen (2001). The $R$ magnitudes given in Courteau (1996) were transformed to the $B$ magnitudes using the observational relation between morphological type index and the $B-R$ colors (de Jong 1996). The solid line in Fig. 4 represents a linear-squares fit to the Courteau’s data (M$_B$ = –7.05 log (V$_{max}$) – 4.57). The broken lines indicate the $\pm$ one sigma dispersion (rms=0.63) for their data. The dispersion for the HCG galaxies around the solid line is rms=0.82 for the whole sample of 25 galaxies and rms=0.65 (similar to that for the control sample of Courteau 1997) if the outliers HCG 89d, 91c, 96c, 96d are not considered (see section 5.1 for the discussion why these galaxies could have a peculiar location in the T-F relation). The sample of Verheijen (2001) for the Ursa Major cluster spirals presents a [*smaller*]{} dispersion around the solid line of rms=0.43, as expected for [*cluster*]{} galaxies.
Our main result is that galaxies in compact groups follow the T-F relation, with a few exceptions. This result is in contrast with an earlier result by R1991, who found an offset of the T-F relation for most galaxies in compact groups in the sense that galaxies in compact groups have lower velocities for a given luminosity. The disagreement is explained by the differences in the amplitudes of the derived rotation curves, V$_{max}$, as discussed in Section 3, mainly due to a different choice of galactic parameters (center, inclination and position angle) and a less extended rotation curve in the case of curves obtained from long-slit spectroscopy. Generally, a choice of parameters guided by photometry alone (when no kinematic maps are available) will result in a lower V$_{max}$ of a galaxy, affecting the T-F relation in the direction of the R1991 result.
Discussion and future prospects
===============================
The fact that the $B$-band T-F relation is similar for compact group and field galaxies tells us that these galaxies show common mass-to-size relations. However, since the parameters for the T-F relation are mostly derived in the inner parts of the galaxy, this agreement does not tell us much about the dark matter, a dominant component in the [*outer*]{} parts (although for the latest-type galaxies we do expect a significant contribution of the dark component in the inner regions as well, Blais-Ouellette et al. 1999). Since the internal velocity dispersions of the compact group galaxy members (250 km s$^{-1}$) are of the same order of their orbital velocities, interactions in the compact-group environment are likely to lead to mergers. In fact, N-body simulations of compact group formation and evolution (Barnes 1989) show that the fate of a compact group is merging in a few crossing times. In order to avoid fast merging, the theoretical expectation is that the compact group environment should change the shapes of the dark matter halos of galaxies that traverse the group, specifically by transforming the single-galaxy halos in a common halo around the group (e.g. Athanassoula, Makino & Bosma, 1997). Our observations cannot test (directly) this hypothesis but do show that the galaxy halos have not been significantly stripped inside R$_{25}$. Nevertheless, mass models using a common halo for the groups added to the individual star distributions of each galaxy should be performed to test this scenario.
The outliers of the Tully-Fisher relation
-----------------------------------------
It is clear from Fig. 4 that a few of the lowest-mass members of compact groups lie well “above” the relationship. If they once had the T-F parameters similar to galaxies in less dense environment, they could have either brightened by 1 to 2 magnitudes in the $B$ band, to get to their present position, they could have lost a substantial amount of mass, or they could have undergone a combination of these processes.
A natural way that the smaller members could have brightened is by forming stars, induced by the interactions the galaxy may have suffered within the group. For a given rate of star formation, the brightening of a galaxy will be much more noticeable for a low-mass than for a high-mass galaxy. As an example, for a large spiral galaxy with mass of 10$^{11}$ solar masses, an episode of star formation that generates, say, 10 solar masses per year, would increase the luminosity of the galaxy in the $B$ band only by about 20%, while for a galaxy with a mass ten times lower, the same episode of star formation would triple the luminosity of the galaxy, enhancing its magnitude by 1.2 mag. Moreover, the two galaxies in the lowest-mass end of the diagram (HCG 89d and 96d) may have large quantities of gas, given their late morphological types. In fact, for HCG 96d, for which a study of HI has been made by Verdes Montenegro et al. (1997), the HI gas contributes with half of the total mass. HCG 89d and HCG 96d have masses (obtained from their maximum velocities inside R$_{25}$ and the virial theorem) well below 10$^{10}$ solar masses and their colours are quite blue (B–R=0.8 and 0.97 respectively, Hickson 1993 and Verdes-Montenegro et al. 1997). It is therefore quite probable that HCG 89d and HCG 96d are “above” the T-F relation mainly due to brightening caused by star formation.
HCG 96c, which may deviate from the T-F relation for similar reasons, is the least massive galaxy in an M51-like pair (with the Seyfert galaxy HCG 96a). HCG 96c is noted by Laurikainen and Moles (1988) as the galaxy with the highest star-formation rate per unit area among interacting galaxies. Its colours (B-V)=0.95 and (U-B)=0.38 are consistent with this scenario (Laurikainen and Moles 1988).
Another galaxy that lies “above” the T-F relation is HCG 91c, which has a colour of B–R=1.08 (Hickson 1993) and a mass inside R$_{25}$ of 2.5 $\times$ 10$^{10}$ M$\odot$. This galaxy clearly shows two dynamic components with quite similar rotation curves (plotted in Fig. 1 as C1 and C2, also see Amram et al. 2003). There are two points we would like to make about this galaxy. First, given the double kinematic component, we strongly suspect that this galaxy is probably the result of a merger of two similar-mass galaxies although we have no signs of this in the photometric profile of the object. In the T-F relation plotted in Fig. 4 we chose to represent this galaxy with the V$_{max}$ of the C1 component and the total B luminosity of the system as a whole. We, therefore, expect, in a naive scenario, that the galaxy will be higher in the T-F relation by at least 0.75 magnitudes, in addition to the amount of brightening due to star formation, if there is any, which, in turn, would explain why this galaxy lies above the T-F relation. The second point, which is related to the first, is that given our arbitrary choice to use only one component (C1) to represent a galaxy which clearly has two equally important kinematic structures, V$_{max}$ for HCG 91c is most probably an underestimate in Fig. 4 and a correction for this would then move the galaxy towards the T-F mean line.
A second mechanism to leave the outliers in the position where they are in the T-F relation would be by mass loss, within R$_{25}$. This could be caused, for example by disruption or ablation of a part of the dark halo due to strong interactions in the group. If the galaxy loses mass, in first approximation it would move to the left of the diagram. We then calculated how much mass the outliers had to lose in order to move from the position they should have (to lie on the T-F relation) to their actual position in the diagram, if mass-loss alone were acting. The answer is an impossible number – 300-500% of the total mass – making this mechanism very unlikely. We then conclude that the reason why the low-mass galaxies are off the T-F relation is most probably due to enhanced star formation and not mass loss but could be a combination of both. Two galaxies which may have suffered truncation are HCG 89d and HCG 96d, the most deviant galaxies from the T-F relation, which have morphological types Sm and Im respectively. Given their late morphological types we suspect that they contain a significant contribution from a dark component even in their inner regions, although detailed modelling has to be done to check this possibility. Their position in the T-F diagram could, then, be the result of a stripping of this inner dark halo combined with brightening of the galaxy due to star formation (given their blue colors).
A trend of overbright galaxies for their mass was seen, at the low-mass end of the relation, when the T-F diagram of binary galaxies was plotted by Márquez et al. (2002). Three of the least massive galaxies in interacting pairs showed a deviation of the T-F relation in the same sense as observed by us for galaxies in compact groups. Moreover, Kannappan et al. (2002), studying a sample of bright nearby galaxies brighter than M$_R=-18$, noted that interacting galaxies systematically lied above while Sa galaxies fell below the T-F relation. In fact, they proposed that there is a correlation of the residuals of the T-F relation with the star formation history of the galaxy and when a second parameter such as colour or equivalent width of H$\alpha$ is taken into account, the scatter on the T-F relation is significantly diminished.
Note that the lowest-mass HCG members deviate from the T-F relation in the opposite sense of those gas-rich galaxies studied by McGaugh et al. (2000) for which a “baryonic correction” (to take into account the gas mass into the “luminosity” of the galaxy) was necessary.
Fraction of HCG galaxies with peculiar velocity fields
------------------------------------------------------
In the last column of Table 1 we marked with a $P$ the galaxies whose velocity fields were considered highly disturbed either by Amram et al. (2003), Plana et al. (2003) or Mendes de Oliveira et al. (1998). A total of 17 from 24 galaxies (all in Table 1 but HCG 87a, which could not be classified) are marked peculiar. If we add to this sample those that were not included in the Tully-Fisher analysis because of their strong peculiarities or short gas extension, we find that 22 out of 29 galaxies, or 75% of the sample of studied galaxies have highly disturbed velocity fields. With this high rate of peculiar kinematics, it is striking that the corresponding rotation curves were possible to be obtained and that, on average, galaxies in compact groups follow the T-F relation.
Asymmetric rotation curves
--------------------------
We can see in Fig. 1 that the rotation curves of HCG 91a (NGC 7214) and HCG96a (NGC 7674) present strong asymmetries, with one arm rising and the other one falling. Both HCG 91a and HCG 96a are the most massive members in M51-like pairs and they are also known to be Seyfert galaxies. We have checked the rotation curves of other most-massive galaxies in M51-like pairs in the sample of Márquez et al. (2002) and we have found several examples of similar rotation curves: NGC 3395/3396, NGC 5395/5394 and NGC 5774/5775. In fact, most galaxies in pairs of galaxies of unequal mass seem to have rotation curves with either one side rising and the other one falling or both falling. Examples of the latter class are NGC 4496a/4496b and NGC 2535/2536 (Amram et al. 1989). Other binary galaxies from Márquez’ sample show a rotation curve with high scatter and a trend for asymmetry (e.g. NGC 3769/3769A). Such rotation curves are rare among isolated galaxies. On the other hand, some rotation curves of galaxies of nearly-equal mass (e.g. NGC 5257/5258, Fuentes-Carrera 2003) curiously do not seem to show such peculiarities, presenting more symmetric rotation curves.
If there was mass stripping from a galaxy and if the system had time to relax, then, the potential should be axisymmetric and both sides of the rotation curve should fall. We observe no such rotation curves among compact group galaxies. However, if the interaction is very recent and the system did not have time to relax, the potential could be non-axisymmetric (triaxial) and the resulting rotation curve could have the observed shape. Another possibility is that the halo could be just shaken (not stripped), in which case we would also expect to observe non-axisymmetric rotation curves such as those we observe among compact group galaxies. More N-body simulations are necessary to further clarify these issues.
The prospects for the future are (1) to use near-infrared imaging to check this result, (2) combine rotation curves with surface brightness profiles to model what role the dark matter plays in the mass distribution, specially to detect galaxies that are dominated by dark matter in their central parts, if there are any and (3) to use other more external (to the group) test-particles to test the halo properties at larger radii (e.g. planetary nebula, globular clusters and/or dwarf galaxies). Finally, a detailed comparison of the results presented in this paper with tidal stripping of dark matter halos in cosmological N-body simulations of compact groups would be of great interest.
CMdO deeply acknowledges the funding and hospitality of the Max Planck für Extraterrestrische Physik and the Universitaets-Sternwarte der Ludwig-Maximilians-Universität, where this work was finalized. CMdO would also like to thank the Brazilian FAPESP (projeto temático 01/07342-7), the PICS program for financial support of two visits to the Observatoire de Marseille and Paris/Meudon and the Brazilian PRONEX program. H.P. acknowledges financial support of Brazilian Cnpq under contract 150089/98-8. The authors thank J. Boulesteix and J.L. Gach for helping during the observations and Mike Bolte for very useful comments on the manuscript. We made use of the Hyperleda database and the NASA/IPAC Extragalactic Database (NED). The latter is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with NASA.
Athanassoula, E., Makino, J. & Bosma, A., 1997, MNRAS, 286, 825 Amram, P., Le Coarer, E., Marcelin, M., Balkowski, C., Sullivan, W.T., III, Cayatte, V., 1992, A&AS, 94, 175. Amram, P., Balkowski, C., Boulesteix, J., Cayatte, V., Marcelin, M., & Sullivan, W. T., III. 1996, A&A, 310, 737 Amram P., Plana H., Mendes de Oliveira C., Balkowski C., Boulesteix J., 2003, , 402, 865 Amram P., Marcelin M., Boulesteix J., Le Coarer E., 1989, A&A Supplement series, 81, 59 Barnes J., 1989, Nature, 338, 123 Blais-Ouellette S., Carignan C., Amram P., Côté S., 1999, AJ, 118, 2123 Courteau S., 1996, , 103, 363 Courteau S., 1997, , 114, 2402 de Jong R.S, 1996, A&A Supplement series, 118, 557 de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H. G., Buta, J., Paturel, G. and Fouqué, P., 1991, [*Third Reference Catalogue of Bright Galaxies*]{}, Springer Verlag (RC3) Fuentes-Carrera, I., Ph.D. Thesis, 2003. Hickson P., Kindl E., Auman J. 1989, , 70, 687 Hickson P. 1993, Astroph.Letter & Comm. 29, 1 Kannappan, S.J., Fabricant, D.G. and Franx, M., 2002, , 123, 2358. Márquez, I., Masegosa, J., Moles, M., Varela, J., Bettoni, D., Galletta, G., 2002, A&A, 393, 389. McGaugh S.S., Schombert J.M., Bothun G.D., de Blok W.J.G., 2000, , 533, L99 Mendes de Oliveira C., Plana H., Amram P., Balkowski C., Boulesteix J. 1998, , 507, 691 Nishiura S., Shimada M., Ohyama Y., Murayama T., Taniguchi Y. 2000, , 120, 169 (N2000) Paturel G., Gouguenheim L., Lanoix P., Marthinet M., Petit C., Rousseau J., Theureau G., Vauglin I. 1997, A&ASS, 124, 109 Plana H., Mendes de Oliveira C., Amram P., Boulesteix J., 1998, , 116, 2123 Plana H., Amram P., Mendes de Oliveira C., Balkowski C., Boulesteix J., 2003, , 125, 1736 Rubin V.C., Hunter D.A., Ford W.K.Jr. 1991, , 76, 153 (R1991) Schlegel, D., Finkbeiner, D. P., Davis, M. 1998, ApJ, 500, 525 Tully, R. B., Pierce, M. J., Huang, J. S., Saunders, W., Verheijen, M. A. W., & Witchalls, P. L. 1998, AJ, 115, 2264 Tully, R.B. & Pierce, M.J. 2000, ApJ, 533, 744 Verdes-Montenegro L., del Olmo A., Perea J., Athanassoula E., Márquez I., Augarde R., 1997, , 321, 409 Verheijen M.A.W, 2001, , 563, 694
[lcccccccc]{}
07c & 52.3 & SBc & SAB(r)c & 12.67 & 168 & 4369 & -21.08 &\
10d & 29.2 & Sc & Scd & 14.41 & 149 & 4781 & -19.69 &\
16a & 34.9 & SBab & SAB(r)ab& 12.82 & 247 & 3748 & -20.68 &\
16c & 31.5 & Im/S0a & SA(rs)0 & 13.09 & 229 & 3766 & -20.40 & P\
19a & 26.2 & S0 & S0 & 13.93 & 144 & 4157 & -19.77 & P\
19b & 24.1 & Scd & SB(r)a & 15.28 & 90 & 4109 & -18.42 & P\
19c & 32.4 & Sdm & SBm & 14.41 & 115 & 4032 & -19.28 &\
79d & 28.1 & Sdm & SB(s)c & 15.91 & 100 & 4678 & -17.98 & P\
87a & 39.8 & Sbc & S0 & 12.88 & 410 & 8645 & -22.46 & –\
87c & 21.4 & Sd & Sb & 14.78 & 195 & 8858 & -20.57 & P\
88a & 45.1 & Sb & Sb & 12.84 & 315 & 5972 & -21.69 &\
88b & 34.0 & SBb & SB(r)a& 13.28 & 303 & 6026 & -21.24 & P\
88c & 27.6 & Scd & SAB(r)bc & 14.04 & 123 & 6074 & -20.49 &\
88d & 32.7 & Sm & Sc & 14.3 & 153 & 6052 & -20.22 &\
89a & 29.1 & Sc & Sc & 14.04 & 210 & 8858 & -21.33 & P\
89b & 20.9 & SBc & SBc & 14.90 & 138 & 8950 & -20.47 & P\
89c & 14.4 & Scd & Scd & 15.51 & 180 & 8893 & -19.86 & P\
89d & 8.1 & Sm & Sm & 16.56 & 65 & 8878 & -18.81 & P\
91a & 43.2 & SBc & SB(s)bc & 12.71 & 271 & 6539 & -22.14 & P\
91c & 23.9 & Sc & Sc & 14.64 & 112 & 7185 & -20.21 & P\
96a & 33.3 & Sc & SA(r)bc & 13.52 & 201 & 8799 & -21.83 & P\
96c & 12.2 & Sa & Sa & 15.72 & 88 & 8709 & -19.63 & P\
96d & 6.6 & Im/Sd & Im & 16.72 & 64 & 9011 & -18.63 & P\
100c & 22.1 & SBc & SBc & 15.12 & 87 & 5507 & -19.19 & P\
100d & 16.0 & Scd & Scd & 15.74 & 114 & 5635 & -18.57 & P\
[rccccccccc]{}
HCG 07 c & 48 & 36 & 133 & 133 & 0.9 & & & &\
& & & & & & & & &\
HCG 10 c & 48 & 64 &153 & 33 & 4.2 & & &\
d & 68 & 75 & 147 & 135 &1.2 & & & &\
& & & & & & & & &\
HCG 16 a & 43 &57 & 3 & 0& 1.3 & 42 & 8 & &\
b & 72 & 58 & 70 & 86 & 2.6 & 64 & 86 & &\
c & 60 & 48 & 120 & 115& 6 & 38 & 78 & &\
d & 47 & 27 & 70 & 80 & 0.4 & 60 & 86 & &\
& & & & & & & & &\
HCG 19 a & 53 & 57 & 67 & 44 & 3.7 & & & &\
b & 60 & 35 & 90 & 87 & 5.9 & & & &\
c & 55 & 50 & 103 & 105 & 0.4 & & & &\
& & & & & & & & &\
HCG 79 d & 58 & 78 & 170 & 180 & 0.4 & 80 & 191 & 90 & 0\
& & & & & & & & &\
HCG 87 a & 85 & 70 & 52 & 50 & 1.6 & & & 90 & 56\
c & 50 & 60 & 82 & 90 & 1.2 & & & 76 & 90\
& & & & & & & & &\
HCG 88 a & 65 & 55 & 128 & 126 & 1.2 & 68 & 127 & 64 & 132\
b & 55 & 45 & 160 & 34 & 3.8 & & & 43 & 31\
c & 42 & ... & 150 & ... & 1.2 & 34 & 160 & 32 & 31\
d & 70 & 73 & 70 & 70 & 2.9 & & & 90 & 71\
& & & & & & & & &\
HCG 89 a & 45 & 51 & 54 & 44 & 0.6 & 60 & 52 & 51 & 57\
b & 49 & 65 & 165 & 155 & 1.2 & 63 & 141 & &\
c & 46 & 65 & 0 & 0 & 0 & & & &\
d & 50 & 45 & 70 & 87 & 2.5 & & & &\
& & & & & & & & &\
HCG 91 a & 50 & 50 & 0 & 50 & 2.5 & & & &\
c1 & 40 & 50 & 141 & 125 & 1.2 & & & &\
c2 & 45 & 50 & 125 & 125 & 1.2 & & & &\
& & & & & & & & &\
HCG 96 a & 50 & 60 & 132 & 150 & 1.8 & & & 25 & 90\
c & 57 & 53 & 33 & 33 & 0 & & & &\
d & 48 & 56 & 10 & 0 & 1.8 & & & &\
& & & & & & & & &\
HCG 100 a & 50 & 47 & 78 & 78 & 1.3& 60 & 85 & &\
b & 52 & 57 & 145 & 165 & 2.8 & & & &\
c & 66 & 65 & 72 & 70 & 2.5 & 68 & 73 & &\
d & 70 & 64 & 53 & 47 & 1.9 & & & &\
$^1$ Galaxy identification. Note that HCG 10c, 16b, 16d, 100a and 100b were not included in the T-F analysis\
$^2$ Inclination in degrees from our velocity field\
$^3$ Inclination in degrees from our continuum map\
$^4$ Position angle in degrees from in velocity field\
$^5$ Position angle in degrees from in continuum map\
$^6$ Difference between the kinematical center and the continuum center, in arsec\
$^7$ Inclination in degrees measured by R1991\
$^8$ PA in degrees measured by R1991\
$^9$ Inclination in degrees measured by N2000\
$^{10}$ PA in degrees measured by N2000\
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Continuous time Bayesian network classifiers are designed for temporal classification of multivariate streaming data when time duration of events matters and the class does not change over time. This paper introduces the CTBNCToolkit: an open source toolkit which provides a stand-alone application for temporal classification and a library for continuous time Bayesian network classifiers. CTBNCToolkit implements the inference algorithm, the parameter learning algorithm, and the structural learning algorithm for continuous time Bayesian network classifiers. The structural learning algorithm is based on scoring functions: the marginal log-likelihood score and the conditional log-likelihood score are provided. CTBNCToolkit provides also an implementation of the expectation maximization algorithm for clustering purpose. The paper introduces continuous time Bayesian network classifiers. How to use the CTBNToolkit from the command line is described in a specific section. Tutorial examples are included to facilitate users to understand how the toolkit must be used. A section dedicate to the library is proposed to help further code extensions.'
author:
- |
Daniele Codecasa $\ddag$\
[email protected]\
DISCo, Università degli Studi\
di Milano-Bicocca
- |
Fabio Stella $\ddag$\
[email protected]\
DISCo, Università degli Studi\
di Milano-Bicocca
bibliography:
- 'references.bib'
title: 'CTBNCToolkit: Continuous Time Bayesian Network Classifier Toolkit'
---
$\ddag$ Authors’ contributions {#ddag-authors-contributions .unnumbered}
------------------------------
The toolkit was developed by Daniele Codecasa, who also wrote the paper. Fabio Stella read the paper and made valuable suggestions. A former MATLAB implementation of the CTBNC inference algorithm was made available by Fabio Stella in order to test the correctness of the inference using the CTBNCToolkit[^1].
[^1]: CTBNCToolkit website and the repository will be updated soon.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'For systems of unstable particles that mix with each other, an approximation of the fully momentum-dependent propagator matrix is presented in terms of a sum of simple Breit–Wigner propagators that are multiplied with finite on-shell wave function normalisation factors. The latter are evaluated at the complex poles of the propagators. The pole structure of general propagator matrices is carefully analysed, and it is demonstrated that in the proposed approximation imaginary parts arising from absorptive parts of loop integrals are properly taken into account. Applying the formalism to the neutral MSSM Higgs sector with complex parameters, very good numerical agreement is found between cross sections based on the full propagators and the corresponding cross sections based on the described approximation. The proposed approach does not only technically simplify the treatment of propagators with non-vanishing off-diagonal contributions, it is shown that it can also facilitate an improved theoretical prediction of the considered observables via a more precise implementation of the total widths of the involved particles. It is also well-suited for the incorporation of interference effects arising from overlapping resonances.'
author:
- |
<span style="font-variant:small-caps;">Elina Fuchs</span>$^{a,}$^1^, <span style="font-variant:small-caps;">Georg Weiglein</span>$^{b,}$^2^\
*(a) Department of Particle Physics and Astrophysics,*\
*Weizmann Institute of Science, Rehovot 76100, Israel*\
*(b) DESY, Deutsches Elektronen-Synchrotron, Notkestr. 85,*\
*D-22607 Hamburg, Germany*
bibliography:
- 'mthesis\_literature.bib'
- 'gNWALoop.bib'
- 'PhDthesis.bib'
- 'bbH.bib'
- 'propagator.bib'
title: '\'
---
Introduction {#chap:intro}
============
As a consequence of electroweak symmetry breaking, particles with the same quantum numbers of the electric charge and colour can mix with each other. This is reflected in their propagators contributing to production or decay processes in particle physics. Beyond lowest order those propagators receive contributions from all possible mixings with other particles, such that the propagators of the system of particles that can mix with each other are of matrix-type. Since the involved particles are in general unstable, a proper treatment of imaginary parts is necessary. The mass of an unstable particle is determined from the real part of the (gauge-invariant) complex pole [@Willenbrock:1991hu; @Sirlin:1991fd; @Stuart:1991xk; @Veltman:1992tm], while the imaginary part yields the total decay width of the particle. The eigenstates associated with the masses obtained from the complex poles of the propagator matrix in general differ from the interaction eigenstates that contribute at lowest order.
As an example, in the Standard Model (SM) the neutral interaction eigenstates arising from the ${\rm SU}(2)_{\rm I}$ and ${\rm U}(1)_{\rm Y}$ gauge groups mix with each other to form the mass eigenstates photon, $\gamma$, and $Z$. Higher-order contributions lead to a mixing between these states, giving rise to a $\gamma-Z$ propagator matrix. In a renormalisable gauge, the (would-be) Goldstone boson as an unphysical degree of freedom does not vanish, and mixing contributions with this unphysical scalar particle need to be taken into account as well. In the quark sector of the SM the relation between the interaction and the mass eigenstates is encoded in terms of the CKM matrix [@Cabibbo:1963yz; @Kobayashi:1973fv]. The additional states in models of physics beyond the SM (BSM) usually follow the same pattern, i.e. interaction eigenstates mix with each other to form mass eigenstates, and the propagators of the latter are of matrix-type as a consequence of mixing contributions at higher orders (in some cases it can also be convenient to derive loop-corrected mass eigenstates directly from lowest-order interaction eigenstates, avoiding the introduction of lowest-order mass eigenstates). In particular, extended Higgs sectors usually contain several neutral Higgs states that can mix with each other. If ${\mathcal{CP}}$-conservation is assumed, ${\mathcal{CP}}$-even and ${\mathcal{CP}}$-odd states can only mix among themselves. In the general case where the possibility of ${\mathcal{CP}}$-violating interactions affecting the Higgs sector is taken into account, non-vanishing mixing contributions can occur between all neutral Higgs bosons of the model. Examples are general two-Higgs-doublet models, the Higgs sector of the minimal supersymmetric extension of the SM (MSSM), as well as the Higgs sectors of non-minimal supersymmetric models. Similarly, propagator mixing between vector resonances can occur for instance for the Kaluza–Klein excitations of the $Z$-boson and the photon ($Z'$ and $A'$, respectively), which can have similar masses and have a large mixing with each other[@Cacciapaglia:2009ic]. New, possibly nearly degenerate vector bosons can also arise in extended gauge symmetries and theories with a strongly coupled sector, e.g. composite Higgs models. For the case of the mixing of two $Z'$-like vectors, see e.g. Ref.[@deBlas:2012qp]. An example involving mixing with a new fermion is the mixing between the top quark $t$ and a top partner $t'$[@Pilaftsis:1997dr].
In order to properly treat external particles in physical processes, the correct on-shell properties of in- and outgoing particles have to be ensured. While the S-matrix theory is formulated only for stable in- and outgoing states [@Eden:AnaSmatrix], in practical applications one often needs to deal with the case of unstable external particles. The proper normalisation of external particles requires in particular that all external particles are on their mass shell, that their residues are equal to one, and that the mixing contributions with all other particles vanish on-shell at the considered order. This requirement holds independently from the specific renormalisation scheme that has been adopted for a certain calculation. In case the renormalisation conditions of the chosen scheme do not impose the proper normalisation of external particles, the correct on-shell properties need to be established via UV-finite wave function normalisation factors (“Z-factors”). In the case of unstable particles, finite-width effects have to be taken into account, and more generally imaginary parts arising from the absorptive parts of loop integrals need to be properly treated. A further complication arises if ${\mathcal{CP}}$-violating effects are taken into account, since in this case the presence of complex parameters yields an additional source of imaginary parts. In the SM the different sources of imaginary parts affect for instance the renormalisation of the CKM matrix from the two-loop level onwards, see e.g. the discussion in Ref. [@Espriu:2002xv] and references therein. In the MSSM with complex parameters the corresponding effects arising from absorptive parts of loop integrals and complex model parameters enter predictions for physical observables in the chargino and neutralino sector already at the one-loop level, see Refs. [@Fowler:2009ay; @Fowler:2010eba; @Bharucha:2012nx].
In the present paper we demonstrate that on-shell wave function normalisation factors, evaluated at the complex poles, are well suited for approximating the full mixing propagators also off-shell. In order to achieve this, the elements of the propagator matrix are expressed in terms of simple Breit–Wigner propagators that are multiplied with wave function normalisation factors. We show for various examples that this approximation works very well in practice. In this context we discuss in detail the uncertainties that are associated with the neglected terms in the approximation and provide estimates of their impact, which we find to be in good agreement with the observed numerical deviations between the full result and the approximation. We furthermore demonstrate that besides very significant technical simplifications our approach also has conceptual advantages.
Concerning technical simplifications, the evaluation of the propagator matrix for a given value of the external momentum squared requires the full momentum dependence of all contributing diagonal and off-diagonal self-energies. In our approximation based on wave function normalisation factors that are evaluated at the complex poles, on the other hand, the only momentum dependence that needs to be taken into account is the one of the simple Breit–Wigner propagators. We emphasize that our approach goes significantly beyond a simple “effective coupling”-type approximation where all external momenta at the loop level are neglected. While the latter approximation would disregard the important absorptive parts of the loop functions, those contributions are taken account in our approach through the evaluation of the Z-factors at the complex poles. We will demonstrate in particular that expanding the full propagators around all of their complex poles indeed results in the sum of Breit–Wigner propagators in combination with on-shell Z-factors. Approximating the full propagator matrix in this simple and convenient way has several conceptual advantages. In particular, the formulation in terms of Breit–Wigner propagators and Z-factors facilitates the implementation of a more precise total width than the width obtained from the imaginary part of the complex pole with self-energies evaluated at the same order. Furthermore, for states with masses that are nearly degenerate (i.e. the difference between the masses of the two states is of the same order as the sum of the total widths of the two states) the mixing becomes resonantly enhanced and the resonances overlap, see e.g. Refs. [@Pilaftsis:1997dr; @Frank:2006yh; @Cacciapaglia:2009ic; @deBlas:2012qp; @Fuchs:2014ola; @Gomez-Bock:2015sia]. In such a case a single-pole approximation is not applicable [@Cacciapaglia:2009ic; @deBlas:2012qp; @Fuchs:2014ola]. Instead, the contributions from the relevant complex poles of the propagators have to be taken into account. The approximation of the propagator matrix in terms of on-shell wave function normalisation factors evaluated at the complex poles and Breit–Wigner propagators allows the implementation of interference contributions using the prescription of Ref. [@Fuchs:2014ola].
We perform in this paper a careful treatment of imaginary parts and analyse the pole structure of matrix-type propagators in detail. While in the no-mixing case every element of an $(n \times n)$ propagator matrix has a single pole, we will demonstrate that in general every single element of an $(n \times n)$ propagator matrix (diagonal and off-diagonal) by itself has $n$ poles. This fact makes the association of the (loop-corrected) mass eigenstates with the propagator poles non-trivial. We will show that in principle all permutations of the $n$ propagator poles with the $n$ mass eigenstates are physically equivalent. For practical purposes, however, it is important to make a choice that leads to numerically stable and well-behaved results.
To be specific, in the following we use the language of the Higgs sector of the MSSM with complex parameters. In this case ${\mathcal{CP}}$-violating effects entering at the loop level give rise to a mixing between the lowest-order mass eigenstates $h, H, A$, which are ${\mathcal{CP}}$-eigenstates ($h, H$ are ${\mathcal{CP}}$-even states, while $A$ is a ${\mathcal{CP}}$-odd state). The $(3 \times 3)$ mixing between the neutral Higgs states is the smallest type of propagator mixing that exhibits the qualitative features of general $(n \times n)$ propagator matrices. Furthermore, the MSSM example is well-suited to emphasise the distinction between interaction eigenstates (in this case the component field of the two Higgs doublets, e.g. $\phi_1$ and $\phi_2$), lowest-order mass eigenstates ($h, H, A$) and loop-corrected mass eigenstates, which in the MSSM case are called $h_1, h_2, h_3$. The MSSM case also illustrates the possibility of mixing with other spin states ($\gamma, Z$) and with unphysical degrees of freedom ($G$). An important feature of models like the MSSM is the fact that the masses are [*predicted*]{} from the input parameters of the model, which makes the need for an appropriate assignment of mass eigenstates with propagator poles particularly apparent. The generalisation of our results to other models, cases with $(n \times n)$ mixing for $n > 3$ and other spin states should be straightforward.
The paper is organised as follows. After reviewing aspects of higher-order contributions and mixing effects in a system of three scalar states in Section\[sect:LoopMix\], we derive the pole structure of the propagator matrix and resulting relations between the wave function normalisation factors in Section\[sect:poles\]. An analytical derivation of the multi-Breit–Wigner approximation of the full propagators including an uncertainty estimate is given in Section\[sect:fullBWZ\]. We perform a detailed numerical comparison of the two approaches in Section\[sect:NumBWfull\] before we conclude in Section\[sect:conclusions\].
Loop-level mixing of scalar propagators {#sect:LoopMix}
=======================================
In a system of $n$ particles which have the same conserved quantum numbers in general a non-trivial mixing between the different states will occur. While at lowest order the mass matrix given in terms of the interaction eigenstates can be diagonalised in order to obtain the mass eigenstates, at the loop level momentum-dependent self-energy corrections enter. These contributions give rise to a momentum-dependent $(n \times n)$ propagator matrix which is in general non-diagonal. The physical masses that are associated with the loop-corrected propagator matrix need to be determined from the (complex) poles of the propagator matrix.
While the case of $(2\times 2)$ mixing provides the minimal setup for studying such a propagator structure, the issue of how to associate the mass eigenstates with the $n$ poles of the propagator matrix becomes fully apparent only from the $(3\times 3)$ case onwards. The case of $(3\times 3)$ mixing captures all relevant aspects of the pole structure and permutation of the states so that the conclusions can be generalised for larger mixing systems.
For illustration and definiteness we use in this paper the language and the formalism of the neutral Higgs sector of the MSSM with ${\mathcal{CP}}$-violating mixing, i.e. for the three lowest-order mass eigenstates $i,j,k = h,H,A$ and their mixing into the loop-corrected mass eigenstates $h_a,~a=1,2,3$. Since the analytic discussions in Chapters \[sect:LoopMix\]-\[sect:fullBWZ\] do not rely on any model-dependent relations of the MSSM, the results can be transferred also to scalar sectors of other models by replacing $h,H,A$ by a different set of $i,j,k$, and analogously $h_a$ by a general $X_a$, where the generalisation to the case of mixing among more than three particles should be straightforward. Upon incorporation of the appropriate spin structure, the results can also be generalised to the case of mixing propagators of vector bosons or fermions.
The Higgs propagators in the MSSM in fact receive contributions from the mixing with the longitudinal components of the neutral gauge bosons $Z$ and $\gamma$. While the $Z$-boson propagator contributes to the diagonal elements of the Higgs propagator matrix at the two-loop level, the photon enters in those elements only at the four-loop level. Besides the mixing with the longitudinal component of the $Z$ boson, also the mixing with the unphysical (would-be) Goldstone boson $G$ needs to be taken into account. Instead of considering a $(6\times6)$ propagator matrix of the states $\left\lbrace h,H,A,G,Z,\gamma\right\rbrace$, we focus on the full mixing contributions of the physical Higgs fields, i.e. we treat the case of a $(3
\times 3)$ propagator matrix (which would generalise to $(n \times n)$ for the case of $n$ physical scalar fields). For the $(3 \times 3)$ propagator matrix of physical Higgs fields, higher-order effects from the inversion of the matrix of the irreducible 2-point vertex functions are taken into account. In contrast, we treat the mixing contributions with the gauge bosons and the (would-be) Goldstone boson in a strict perturbative expansion up to the desired order. It should be noted that Higgs–$G/Z$ mixing contributions already appear at the one-loop level in processes with external Higgs bosons, see e.g. Refs. [@Williams:2007dc; @Fowler:2009ay; @Williams:2011bu; @Bharucha:2012nx], and therefore need to be included in order to obtain a complete one-loop result for processes of this kind. For contributions to the propagator matrix with an incoming and outgoing Higgs boson the mixing contributions with $G$ and $Z$ enter from the order of $({\hat{\Sigma}}_{i,G/Z})^{2}$ onwards, which is of subleading two-loop order. For our numerical analysis below, see Sect.\[sect:NumBWfull\], we will use use the program `FeynHiggs`[@Heinemeyer:1998np; @Heinemeyer:1998yj; @Degrassi:2002fi; @Heinemeyer:2007aq], where dominant two-loop corrections and leading higher-order contributions are incorporated in the irreducible self-energies. Since sub-leading two-loop contributions that are of the same order as the mixing contributions with $G$ and $Z$[@Frank:2006yh] are neglected in this analysis, we include no further corrections beyond the $(3 \times 3)$ propagator matrix of the physical Higgs fields in this case.
In this section we introduce the relevant quantities and fix our notation for the propagator matrix and the wave function normalisation factors, see Refs.[@Dabelstein:1994hb; @Dabelstein:1995js; @Heinemeyer:2000fa; @Heinemeyer:2001iy; @Hahn:2002gm; @Frank:2006yh; @Hahn:2006np; @Williams:2007dc; @Fowler:2010eba; @Williams:2011bu]. In Sect.\[sect:poles\] we will analyse the pole structure of propagator matrices of unstable particles.
Propagator matrix and the effective self-energy
-----------------------------------------------
As explained above, we use the case of the MSSM with complex parameters in order to illustrate propagator mixing between three physical scalar fields. If ${\mathcal{CP}}$ were assumed to be conserved in the MSSM, the ${\mathcal{CP}}$-violating self-energies would vanish, ${\hat{\Sigma}}_{hA}={\hat{\Sigma}}_{HA}=0$, so that only the two ${\mathcal{CP}}$-even states $h$ and $H$ would mix with each other. Our treatment corresponds to the case where non-zero phases from complex parameters are taken into account. Hence all renormalised self-energies ${\hat{\Sigma}_{ij}}{(p^{2})}$ of the Higgs bosons $i,j = h,H,A$ are in general non-vanishing, so that the matrix ${\textbf{M}}$ of mass squares consists of the tree-level masses $m_i^{2}$ on the diagonal and renormalised self-energies on the diagonal and off-diagonal entries. Expressed in terms of the lowest-order mass eigenstates $h, H, A$, which are also ${\mathcal{CP}}$ eigenstates, the matrix takes the form $${\textbf{M}}({p^{2}}) = {\begin{pmatrix}}m_h^{2}-\hat{\Sigma}_{hh}({p^{2}}) & -\hat{\Sigma}_{hH}({p^{2}}) &-\hat{\Sigma}_{hA}({p^{2}})\\
-\hat{\Sigma}_{Hh}({p^{2}}) & m_H^{2}-\hat{\Sigma}_{HH}({p^{2}}) &-\hat{\Sigma}_{HA}({p^{2}})\\
-\hat{\Sigma}_{Ah}({p^{2}}) & -\hat{\Sigma}_{AH}({p^{2}}) &m_A^{2}-\hat{\Sigma}_{AA}({p^{2}})
{\end{pmatrix}}. \label{eq:HiggsMassMatrix}$$ The renormalised irreducible 2-point vertex functions $$\hat{\Gamma}_{ij}{(p^{2})}= i\left[({p^{2}}-m_i^{2})\delta_{ij}+{\hat{\Sigma}}_{ij}{(p^{2})}\right] \label{eq:IrepVert}$$ can be collected in the $3\times3$ matrix ${\hat{\boldsymbol{\Gamma}}}_{hHA}$ in terms of ${\textbf{M}}$ as $${\hat{\boldsymbol{\Gamma}}}_{hHA}({p^{2}}) = i\left[{p^{2}}\textbf{1} -{\textbf{M}}({p^{2}}) \right] \label{eq:Gamma}.$$ Finally, the propagator matrix ${\boldsymbol{\Delta}}_{hHA}$ equals, up to the sign, the inverse of ${\hat{\boldsymbol{\Gamma}}}_{hHA}$, $${\boldsymbol{\Delta}}_{hHA}({p^{2}}) = - \left[{\hat{\boldsymbol{\Gamma}}}_{hHA}({p^{2}}) \right]^{-1}. \label{eq:DeltaGamInv}$$ Accordingly, the matrix inversion yields the individual propagators $\Delta_{ij}{(p^{2})}$ as the the $(ij)$ elements of the $3\times3$ matrix ${\boldsymbol{\Delta}}_{hHA}({p^{2}})$, $${\boldsymbol{\Delta}}_{hHA} = {\begin{pmatrix}}\Delta_{hh} &\Delta_{hH} &\Delta_{hA}\\
\Delta_{Hh} &\Delta_{HH} &\Delta_{HA}\\
\Delta_{Ah} &\Delta_{AH} &\Delta_{AA}
{\end{pmatrix}}\label{eq:DeltaMat}.$$ The off-diagonal entries (for $i\neq j$) result in: $$\Delta_{ij}({p^{2}}) = \frac{{\hat{\Gamma}_}{ij}{\hat{\Gamma}_}{kk}- {\hat{\Gamma}_}{jk}{\hat{\Gamma}_}{ki}}{{\hat{\Gamma}_}{ii}{\hat{\Gamma}_}{jj}{\hat{\Gamma}_}{kk} +2{\hat{\Gamma}_}{ij}{\hat{\Gamma}_}{jk}{\hat{\Gamma}_}{ki}-{\hat{\Gamma}_}{ii}{\hat{\Gamma}_}{jk}^{2} -{\hat{\Gamma}_}{jj}{\hat{\Gamma}_}{ki}^{2}-{\hat{\Gamma}_}{kk}{\hat{\Gamma}_}{ij}^{2}} \label{eq:Dij}.$$ All 2-point vertex functions $\hat{\Gamma}{(p^{2})}$ depend on ${p^{2}}$ via Eq.(\[eq:IrepVert\]). Here we do not write the ${p^{2}}$-dependence explicitly for the purpose of a simpler notation, but the full ${p^{2}}$-dependence is implied also below. The solutions of the diagonal propagators, $\Delta_{ii}$, can be expressed in the following compact way: $$\begin{aligned}
\Delta_{ii}{(p^{2})}&= \frac{{\hat{\Gamma}_}{jj}{\hat{\Gamma}_}{kk}-{\hat{\Gamma}_}{jk}^{2}}{-{\hat{\Gamma}_}{ii}{\hat{\Gamma}_}{jj}{\hat{\Gamma}_}{kk} + {\hat{\Gamma}_}{ii}{\hat{\Gamma}_}{jk}^{2} - 2{\hat{\Gamma}_}{ij}{\hat{\Gamma}_}{jk}{\hat{\Gamma}_}{ki} + {\hat{\Gamma}_}{jj}{\hat{\Gamma}_}{ki}^{2} +{\hat{\Gamma}_}{kk}{\hat{\Gamma}_}{ij}^{2}}\label{eq:Dii}\\
&= \frac{i}{{p^{2}}-m_i^{2} + {\hat{\Sigma}^{\rm{eff}}_{ii}}{(p^{2})}},\label{eq:Diieff}\end{aligned}$$ where the effective self-energy is introduced, $$\begin{aligned}
{\hat{\Sigma}^{\rm{eff}}_{ii}}(p^{2}) &= \hat{\Sigma}_{ii}(p^{2}) - i \frac{2{\hat{\Gamma}_}{ij}(p^{2}){\hat{\Gamma}_}{jk}(p^{2}){\hat{\Gamma}_}{ki}(p^{2}) - {\hat{\Gamma}_}{ki}^{2}({p^{2}}){\hat{\Gamma}_}{jj}({p^{2}})-{\hat{\Gamma}_}{ij}^{2}({p^{2}}){\hat{\Gamma}_}{kk}({p^{2}})}{{\hat{\Gamma}_}{jj}({p^{2}}){\hat{\Gamma}_}{kk}({p^{2}})-{\hat{\Gamma}_}{jk}^{2}({p^{2}})} \label{eq:seff}.\end{aligned}$$ It contains the diagonal self-energy, ${\hat{\Sigma}_{ii}}$ (which exists already at 1-loop order), and the mixing 2-point functions (whose products only contribute to ${\hat{\Sigma}^{\rm{eff}}_{ii}}$ from 2-loop order onwards). Hence, replacing the pure self-energy $\hat{\Sigma}_{ii}$ by the effective one, ${\hat{\Sigma}^{\rm{eff}}_{ii}}$, includes also the $3\times 3$ mixing contributions to the diagonal propagator in Eq.(\[eq:Diieff\]) while preserving formally the structure of the propagator as in the unmixed case. In the limit of no mixing, the second term in Eq.(\[eq:seff\]) vanishes.
The composition of ${\hat{\Sigma}^{\rm{eff}}_{ii}}$ in terms of the unmixed and the mixing contributions in Eq.(\[eq:seff\]) can be written in the alternative way [@KarinaPhD; @Fowler:2010eba; @Fuchs:2015jwa] $$\begin{aligned}
{\hat{\Sigma}^{\rm{eff}}_{ii}}{(p^{2})}&= {\hat{\Sigma}_{ii}}{(p^{2})}+ \frac{\Delta_{ij}{(p^{2})}}{\Delta_{ii}{(p^{2})}}{\hat{\Sigma}_{ij}}{(p^{2})}+ \frac{\Delta_{ik}{(p^{2})}}{\Delta_{ii}{(p^{2})}}{\hat{\Sigma}_{ik}}{(p^{2})}\label{eq:seffratio}\end{aligned}$$ for $i,j,k$ all different. Eq.(\[eq:seffratio\]) represents the sum of the diagonal and off-diagonal self-energies involving a Higgs boson $i$ where the off-diagonal contributions are weighted by the ratios of the respective off-diagonal and the diagonal propagators. This expression follows from Eq.(\[eq:seff\]) with the help of the equality $$\begin{aligned}
\frac{\Delta_{ij}}{\Delta_{ii}} &= -\frac{{\hat{\Gamma}_}{ij}{\hat{\Gamma}_}{kk}-{\hat{\Gamma}_}{jk}{\hat{\Gamma}_}{ki}}{{\hat{\Gamma}_}{jj}{\hat{\Gamma}_}{kk}-{\hat{\Gamma}_}{jk}^{2}}
= -\frac{{\hat{\Sigma}_{ij}}\,(D_k+{\hat{\Sigma}_{kk}})-{\hat{\Sigma}_{jk}}{\hat{\Sigma}_{ki}}}{(D_j+{\hat{\Sigma}_{jj}})\,(D_k+{\hat{\Sigma}_{kk}})-{\hat{\Sigma}_{jk}}^{2}}
,\label{eq:ratioijii}\end{aligned}$$ where the shorthand $$\begin{aligned}
D_i{(p^{2})}&= {p^{2}}- m_i^{2} \label{eq:Di}\end{aligned}$$ has been used, analogously for $j\leftrightarrow k$, and ${\hat{\Sigma}}_{ij}=-i{\hat{\Gamma}_}{ij}$ from Eq.(\[eq:IrepVert\]) with $i\neq j$ (for the special case of $(2\times 2)$ mixing, this relation can be directly read off from Eqs.(\[eq:Dii2x2\]-\[eq:seff2x2\]) given below).
On-shell wave function normalisation factors {#sect:onshellZ}
--------------------------------------------
### Complex poles
The complex poles ${\mathcal{M}^{2}}_{a}$, where $a = 1, 2, 3$, are determined as solutions of the equations $$\begin{aligned}
{p^{2}}-m_i^{2} + {\hat{\Sigma}^{\rm{eff}}_{ii}}{(p^{2})}= 0 ,\label{eq:complexpole}\end{aligned}$$ with $i = h, H, A$, see Sect. \[sect:poles\] below. While the propagators have poles for ${p^{2}}={\mathcal{M}^{2}}_a$, it is interesting to note that the ratios of off-diagonal and diagonal propagators $$R_{ij}^{(a)} := \frac{\Delta_{ij}{(p^{2})}}{\Delta_{ii}{(p^{2})}}\bigg\rvert_{{p^{2}}={\mathcal{M}^{2}}_a},\label{eq:Rija}$$ stay finite at the complex poles ${\mathcal{M}^{2}}_a,\,~a=1,2,3$.
### Z-matrix for on-shell properties of external particles {#sect:Zmatrix}
Higgs bosons appearing as *external* particles in a process need the appropriate on-shell properties for a correct normalisation of the S-matrix. In the hybrid on-shell/$\overline{\rm{DR}}$ renormalisation scheme [@Frank:2006yh], the masses are renormalised on-shell, but the $\overline{\textrm{DR}}$ renormalisation conditions for the fields and $\tan\beta$ [@Frank:2006yh], $$\begin{aligned}
\delta Z_{\mathcal{H}_1}^{\overline{\rm{DR}}} &=& -\text{Re}\left[\Sigma^{'(\rm{div})}_{HH}(m_{H}^{2})\right]_{\alpha=0}\label{eq:dZH1},\\
\delta Z_{\mathcal{H}_2}^{\overline{\rm{DR}}} &=& -\text{Re}\left[\Sigma^{'(\rm{div})}_{hh}(m_{h}^{2})\right]_{\alpha=0}\label{eq:dZH2},\\
\delta \tan\beta^{\overline{\rm{DR}}} &=& \frac{1}{2}(\delta Z_{\mathcal{H}_2}^{\overline{\rm{DR}}} - \delta Z_{\mathcal{H}_1}^{\overline{\rm{DR}}})\label{eq:dtanb},\end{aligned}$$ where $\alpha$ is the mixing angle between $h$ and $H$, see Eq.(\[eq:ncHmix\]) below, do not ensure proper on-shell properties of the Higgs bosons. In fact, the loop-corrected mass eigenstates $h_1,h_2,h_3$, which occur as external, on-shell, particles e.g. in decay processes, are a mixture of the lowest order states $h,H,A$. Thus, finite wave function normalisation factors need to be introduced in order to guarantee that the mixing vanishes on-shell and that the propagators of the external particles have unit residue.
These so-called $Z$-factors for a neutral Higgs boson $i=h,H,A$ on an external line are obtained from the residue of the propagators at the complex pole ${\mathcal{M}^{2}}_a,\,a=1,2,3$,[@Dabelstein:1995js; @Heinemeyer:2001iy] $$\begin{aligned}
{\hat{Z}}_i^{a} &:= \textrm{Res}_{{\mathcal{M}^{2}}_a}\left\lbrace\Delta_{ii}{(p^{2})}\right\rbrace \label{eq:ZhaRes}.
\end{aligned}$$ Expanding ${\hat{\Sigma}^{\rm{eff}}_{ii}}({p^{2}})$ around the complex pole ${p^{2}}={\mathcal{M}^{2}}_a$, one obtains the diagonal propagator $$\begin{aligned}
\Delta_{ii}(p^{2}) &=\frac{i}{p^{2}-m_i^{2}+{\hat{\Sigma}^{\rm{eff}}_{ii}}(p^{2})}
=\frac{i}{p^{2}-{\mathcal{M}^{2}}_a}\cdot
\frac{1}{1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_a)+ \mathcal{O}({p^{2}}- {\mathcal{M}^{2}}_a)}
\label{eq:DiiApprdSeff}\,.\end{aligned}$$ Performing the limit ${p^{2}}\rightarrow {\mathcal{M}^{2}}_a$ yields the residue of Eq.(\[eq:ZhaRes\]), $$\begin{aligned}
{\hat{Z}}_i^{a}
&= \frac{1}{\frac{\partial}{\partial {p^{2}}}\frac{i}{\Delta_{ii}{(p^{2})}}}\bigg\rvert_{{p^{2}}={\mathcal{M}^{2}}_a}
=\frac{1}{1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_a)} \label{eq:Zha}\,.\end{aligned}$$ Considering a diagram with the Higgs boson $i$ on an external line, whose propagator has three poles (see Sect. \[sect:poles\]), there are three possibilities which residue to compute. If the amputated Green’s function is evaluated at ${\mathcal{M}^{2}}_a$, the external $i$-line has to be multiplied by $\sqrt{{\hat{Z}}_i^{a}}$ for the correct S-matrix normalisation[^1]. So the resulting mass eigenstate as an outgoing particle is $h_a$. Alternatively, if the Green’s function is evaluated at ${\mathcal{M}^{2}}_b$, it has to be normalised by $$\begin{aligned}
\sqrt{{\hat{Z}}_i^{b}}=\frac{1}{\sqrt{1+{\hat{\Sigma}^{\rm{eff'}}}_{ii}({\mathcal{M}^{2}}_b)}}\end{aligned}$$ to achieve the correct S-matrix element. For this choice, the external mass eigenstate is $h_b$. Moreover, the wave function normalisation factor for $i-j$ mixing on an external on-shell line at ${\mathcal{M}^{2}}_a$ is composed of the overall normalisation factor $\sqrt{{\hat{Z}}_i^{a}}$ times the on-shell transition ratio $${\hat{Z}}_{ij}^{a} \equiv R_{ij}^{(a)} = \frac{\Delta_{ij}{(p^{2})}}{\Delta_{ii}{(p^{2})}}\bigg\rvert_{{p^{2}}={\mathcal{M}^{2}}_a} \label{eq:transij}\,.$$ Since $\Delta_{ii}$ and $\Delta_{ij}$ have – in the case of $3\times 3$ mixing – 3 complex poles (see Sect. \[sect:poles\]), any of them can be chosen for the evaluation of ${\hat{Z}}_i$, ${\hat{Z}}_{ij}$ and ${\hat{Z}}_{ik}$; in the considered example this is ${\mathcal{M}^{2}}_a$. Correspondingly, ${\hat{Z}}_j, {\hat{Z}}_{ji}$ and ${\hat{Z}}_{jk}$ will be evaluated at ${\mathcal{M}^{2}}_b$ and ${\hat{Z}}_k, {\hat{Z}}_{ki}$ and ${\hat{Z}}_{kj}$ at ${\mathcal{M}^{2}}_c$ where $a,b,c$ are a permutation of 1,2,3, and $i,j,k$ a permutation of $h,H,A$[@Williams:2007dc; @Fuchs:2015jwa]. For $2\times 2$ mixing, of course only the two indices involved in the mixing can be permuted.
All choices allowed by the mixing structure are generally possible because of the pole structure of each propagator. However, they might not be equally numerically stable. If the loop-corrected mass eigenstate $h_a$ contains only a small admixture of the lowest-order state $i$, the propagator still has a pole at ${\mathcal{M}^{2}}_a$, but the contribution of $i$ to $h_a$ is suppressed at ${p^{2}}\neq {\mathcal{M}^{2}}_a$.
In order to be definite, it is in any case necessary to define at which pole to evaluate which normalisation and mixing ${\hat{Z}}$-factor. This choice corresponds to fixing an assignment between the indices $i,j,k$ of the lowest-order states and the indices $a,b,c$ of the higher-order mixed states and then using it consistently. The assignment $(i,a), (j,b), (k,c)$, which we label as scheme $I$, prescribes to evaluate ${\hat{Z}}_i$, ${\hat{Z}}_{ij}$ and ${\hat{Z}}_{ik}$ at ${\mathcal{M}^{2}}_a$. Once the indices have been assigned we can clear up the notation by writing $${\hat{Z}}_a\big\rvert_I := {\hat{Z}}_{i}^{a},~~ {\hat{Z}}_{aj}\big\rvert_I := {\hat{Z}}_{ij}^{a},~~ {\hat{Z}}_{bi}\big\rvert_I := {\hat{Z}}_{ji}^{b}\,,\label{eq:Zidentification}$$ and accordingly for the other indices such that the first index always refers to a loop-corrected mass eigenstate ($a,b,c\in\{1,2,3\}$)[^2] and the second index to a lowest-order state ($i,j,k\in\{h,H,A\}$). Note that ${\hat{Z}}_{ai}={\hat{Z}}_{bj}={\hat{Z}}_{ck}\equiv1$ in the index scheme $I$ defined above. Once the index scheme has been specified, one can leave out the subscript $I$. For the scheme-independence of physical results, see Sect.\[sect:schemeIndep\].
Furthermore, it is convenient[@Frank:2006yh] to arrange the products of the normalisation factors $\sqrt{{\hat{Z}}_a}$ and transition ratios ${\hat{Z}}_{aj}$ as $$\begin{aligned}
{\hat{\textbf{Z}}}_{aj} &= \sqrt{{\hat{Z}}_a}{\hat{Z}}_{aj}
\label{eq:ZiZij}\end{aligned}$$ (note the difference between ${\hat{Z}}_{aj}$ and ${\hat{\textbf{Z}}}_{aj}$) into a non-unitary matrix: $$\begin{aligned}
\hat{\textbf{Z}} &= {\begin{pmatrix}}\sqrt{\hat{Z}_1}{\hat{Z}}_{1h} & \sqrt{\hat{Z}_1}\hat{Z}_{1H} &\sqrt{\hat{Z}_1}\hat{Z}_{1A}\\
\sqrt{\hat{Z}_2}\hat{Z}_{2h} & \sqrt{\hat{Z}_2}{\hat{Z}}_{2H} &\sqrt{\hat{Z}_2}\hat{Z}_{2A}\\
\sqrt{\hat{Z}_3}\hat{Z}_{3h} & \sqrt{\hat{Z}_3}\hat{Z}_{3H} &\sqrt{\hat{Z}_3}{\hat{Z}}_{3A}
{\end{pmatrix}}\label{eq:Zmatrix}.\end{aligned}$$ The ${\hat{\textbf{Z}}}$-matrix defined in Eq.(\[eq:Zmatrix\]) fulfils the on-shell conditions (unit residue and vanishing mixing), which can be written in the following compact form[@Williams:2007dc; @KarinaPhD; @Fowler:2010eba; @Williams:2011bu]: $$\begin{aligned}
\lim_{{p^{2}}\rightarrow {\mathcal{M}^{2}}_a} -\frac{i}{{p^{2}}-{\mathcal{M}^{2}}_a}\left({\hat{\textbf{Z}}}\cdot
{\hat{\boldsymbol{\Gamma}}}_{hHA}\cdot {\hat{\textbf{Z}}}^{T} \right)_{hh} ~&= 1
\label{eq:Reshha},\\
\lim_{{p^{2}}\rightarrow {\mathcal{M}^{2}}_b} -\frac{i}{{p^{2}}-{\mathcal{M}^{2}}_b}\left({\hat{\textbf{Z}}}\cdot
{\hat{\boldsymbol{\Gamma}}}_{hHA}\cdot
{\hat{\textbf{Z}}}^{T} \right)_{HH} &= 1
\label{eq:ResHHb},\\
\lim_{{p^{2}}\rightarrow {\mathcal{M}^{2}}_c} -\frac{i}{{p^{2}}-{\mathcal{M}^{2}}_c}\left({\hat{\textbf{Z}}}\cdot
{\hat{\boldsymbol{\Gamma}}}_{hHA}\cdot {\hat{\textbf{Z}}}^{T} \right)_{AA} &= 1
\label{eq:ResAAc},\end{aligned}$$ with vanishing off-diagonal entries. It is equally possible to begin with these equations (\[eq:Reshha\]-\[eq:ResAAc\]), i.e. to require unit residues and to demand vanishing mixing on-shell, to derive the elements of the ${\hat{\textbf{Z}}}$-matrix whose solutions are given in Eqs.(\[eq:Zha\]) and (\[eq:transij\]).
In Ref.[@Williams:2007dc] the evaluation at the full complex poles and the inclusion of imaginary parts were introduced (see Eq.(\[eq:SigijReIm\]) below for the treatment of loop functions of complex arguments). The evaluation at the complex poles leads to numerically more stable results than the evaluation at real ${p^{2}}=M_{h_a}^{2}$, as well as to the physically equivalent choices of index assignments. The calculation of the ${\hat{\textbf{Z}}}$-factors of MSSM Higgs bosons can be performed with the program `FeynHiggs`.
Use of Z-factors for external states {#sect:ExtH}
------------------------------------
The ${\hat{\textbf{Z}}}$-factors have been introduced for the correct normalisation of matrix elements with external (on-shell) Higgs bosons $h_a,~{p^{2}}={\mathcal{M}^{2}}_a$. It should be noted that ${\hat{\textbf{Z}}}$ does not provide a unitary transformation between the basis of lowest-order states and the basis of loop-corrected mass eigenstates. The fact that ${\hat{\textbf{Z}}}$ is a non-unitary matrix is related to the imaginary parts appearing in the propagators of unstable particles. Using the ${\hat{\textbf{Z}}}$-matrix, one can express the one-particle irreducible (1PI) vertex functions ${\hat{\Gamma}_}{h_a}$ involving a loop-corrected mass eigenstate $h_1,h_2,h_3$ as an external particle as a linear combination of the 1PI vertex functions of the lowest-order states, ${\hat{\Gamma}_}i$: $$\begin{aligned}
{\hat{\Gamma}_}{h_a}&={\hat{\textbf{Z}}}_{ah}{\hat{\Gamma}_}h+{\hat{\textbf{Z}}}_{aH}{\hat{\Gamma}_}H+{\hat{\textbf{Z}}}_{aA}{\hat{\Gamma}_}A + ...\label{eq:Zextsum}\\
&=\sqrt{{\hat{Z}}_a}\,\left({\hat{Z}}_{ah}{\hat{\Gamma}_}h+{\hat{Z}}_{aH}{\hat{\Gamma}_}H+{\hat{Z}}_{aA}{\hat{\Gamma}_}A\right) + \dots \label{eq:Vertexha},\end{aligned}$$ where the ellipsis refers to additional terms arising from the mixing with Goldstone and vector bosons, which are not described by the ${\hat{\textbf{Z}}}$-matrix. Thus, the overall normalisation factor $\sqrt{{\hat{Z}}_a}$ accounts for the particle $h_a$ appearing at an external line. In addition, the factors ${\hat{Z}}_{ai}$ given in Eqs.(\[eq:transij\]) and (\[eq:Zidentification\]) as ratios of propagators at ${p^{2}}={\mathcal{M}^{2}}_a$ describe the transition between the states $h_a$ and $i$. The transition factor ${\hat{Z}}_{ai}$ occurs in a diagram where $h_a$ is the external particle, but $i$ directly couples to the vertex. All possibilities for $i=h,H,A$ need to be included for each $h_a$, hence the sum arises. This is depicted in Fig.\[fig:Zext\] (cf. also Refs.[@Hahn:2007fq; @KarinaPhD]).
at (2.5,-0.6) [${p^{2}}={\mathcal{M}^{2}}_a$]{}; (1,0)–(4,0); at (2.5,0.6) [$\boldsymbol{h_a}$]{}; (4.3,0) circle (0.3); (5.6,1)–(4.6,0.1); (5.6,-1)–(4.6,-0.1); at (5.5,0) [$\hat\Gamma_{{\color{blue} h_a}}$]{}; at (7.2,0) [= $\sqrt{{\hat{Z}}_{{\color{blue} a}}}\Bigg($]{};
at (2.5,-0.8) [${\hat{Z}}_{{\color{blue}a}{\color{OliveGreen}h } }$]{}; (1,0)–(2.2,0); (2.8,0)–(4,0); at (1.6,0.6) [$\boldsymbol{h_a}$]{}; at (3.4,0.6) [$\boldsymbol{h}$]{}; (2.5,0) circle (0.3); (5,1)–(4,0)–(5,-1); at (5,0) [$\hat\Gamma_{{\color{OliveGreen} h}}$]{}; at (6,0) [+]{};
at (2.5,-0.8) [${\hat{Z}}_{{\color{blue}a}{\color{OliveGreen}H } }$]{}; (1,0)–(2.2,0); (2.8,0)–(4,0); at (1.6,0.6) [$\boldsymbol{h_a}$]{}; at (3.4,0.6) [$\boldsymbol{H}$]{}; (2.5,0) circle (0.3); (5,1)–(4,0)–(5,-1); at (5,0) [$\hat\Gamma_{{\color{OliveGreen} H}}$]{}; at (6,0) [+]{};
at (2.5,-0.8) [${\hat{Z}}_{{\color{blue}a}{\color{OliveGreen}A } }$]{}; (1,0)–(2.2,0); (2.8,0)–(4,0); at (1.6,0.6) [$\boldsymbol{h_a}$]{}; at (3.4,0.6) [$\boldsymbol{A}$]{}; (2.5,0) circle (0.3); (5,1)–(4,0)–(5,-1); at (5,0) [$\hat\Gamma_{{\color{OliveGreen} A}}$]{};
at (25,0) [$\Bigg)_{{p^{2}}={\mathcal{M}^{2}}_a}$]{}; at (26.4,0) [$+\dots$]{};
Conveniently, Eq.(\[eq:Vertexha\]) can be written in matrix form for all $h_1,h_2,h_3$ as $${\begin{pmatrix}}{\hat{\Gamma}_}{h_1}\\{\hat{\Gamma}_}{h_2}\\{\hat{\Gamma}_}{h_3}{\end{pmatrix}}= {\hat{\textbf{Z}}}\cdot {\begin{pmatrix}}{\hat{\Gamma}_}{h}\\{\hat{\Gamma}_}{H}\\{\hat{\Gamma}_}{A}{\end{pmatrix}}+\dots\label{eq:vertZ}~.$$ In this way, propagator corrections at external legs are effectively absorbed into the vertices of neutral Higgs bosons. If ${\hat{\textbf{Z}}}$-factors are applied to supplement the Born result, only non-Higgs propagator type corrections (such as mixing with the Goldstone and $Z$-bosons) as well as vertex, box and real corrections need to be calculated individually.
As it is the case for the usual (UV-divergent) on-shell field renormalisation constants, the finite wave function normalisation factors are in general gauge-dependent quantities. While the gauge-parameter dependence drops out at a fixed order in perturbation theory, the inclusion of higher-order contributions into the ${\hat{\textbf{Z}}}$-factors described above can give rise to a residual gauge-parameter dependence, which is formally of sub-leading higher order. This is caused by the fact that the ${\hat{\textbf{Z}}}$-factors contain not only gauge-parameter independent leading higher-order contributions but also sub-leading higher-order contributions that are in general gauge-parameter dependent. Those sub-leading higher-order contributions would cancel with corresponding sub-leading higher-order contributions from other Green functions such as vertex and box contributions.
Effective couplings {#sect:Ueff}
-------------------
Since the ${\hat{\textbf{Z}}}$-matrix is not unitary, it does not represent a unitary transformation between the $\{h,H,A\}$ and the $\{h_1,h_2,h_3\}$ basis. However, it is not necessary to diagonalise the mass matrix for the determination of the poles of the propagators. Hence there is a priori no need to introduce a unitary transformation. Though, if a unitary matrix ${\textbf{U}}$ is desired for the definition of effective couplings, an approximation of the momentum dependence of ${\hat{\textbf{Z}}}$ is required. There is no unique prescription of how to achieve a unitary mixing matrix as an approximation of the ${\hat{\textbf{Z}}}$-matrix, but a possible choice is the ${p^{2}}=0$ approximation[@Frank:2006yh; @Hahn:2007fq]. As in the effective potential approach, the external momentum ${p^{2}}$ is set to zero in the renormalised self-energies ${\hat{\Sigma}_{ij}}({p^{2}})\rightarrow {\hat{\Sigma}_{ij}}(0)$ so that they become real. Then ${\textbf{U}}$ diagonalises the real matrix ${\textbf{M}}(0)$. ${\textbf{U}}$ can be chosen real and it transforms the ${\mathcal{CP}}$-eigenstates into the mass eigenstates, $$\begin{aligned}
{\begin{pmatrix}}h_1\\ h_2\\ h_3 {\end{pmatrix}}= {\textbf{U}}{\begin{pmatrix}}h\\ H\\ A {\end{pmatrix}},~~~~
{\textbf{U}}= {\begin{pmatrix}}{\textbf{U}}_{1h}& {\textbf{U}}_{1H}& {\textbf{U}}_{1A}\\
{\textbf{U}}_{2h}& {\textbf{U}}_{2H}& {\textbf{U}}_{2A}\\
{\textbf{U}}_{3h}& {\textbf{U}}_{3H}& {\textbf{U}}_{3A}
{\end{pmatrix}}\label{eq:Urotate},\end{aligned}$$ so that $U_{aA}^{2}$ quantifies the admixture of a ${\mathcal{CP}}$-odd component inside $h_a$[@Frank:2006yh]. The elements of ${\textbf{U}}$ can then be used to introduce effective couplings of the loop-corrected states $h_a$ to any other particles $X$ in terms of the couplings of the unmixed states $i$ by the relation $$C_{h_a X}^{U} = \sum_{i=h,H,A} {\textbf{U}}_{ai} C_{i X} \label{eq:EffCoup}.$$ absorbing some higher-order corrections, but neglecting imaginary parts and the full momentum dependence of the self-energies. Hence, the application of ${\textbf{U}}$ resembles the use of ${\hat{\textbf{Z}}}$-factors in Eq.(\[eq:Zextsum\]) for the purpose of implementing partial higher-order effects into an improved Born result. Yet, the rotation matrix ${\textbf{U}}$ introduced for effective couplings as a unitary approximation is conceptually different from the ${\hat{\textbf{Z}}}$-matrix arising from propagator corrections and introduced for the correct normalisation of the $S$-matrix.
Technical treatment of imaginary parts
--------------------------------------
In order to account for complex momenta and imaginary parts of self-energies, we employ an expansion of the self-energies around the real part of the complex momentum, $$\begin{aligned}
{p^{2}}&\equiv {p^{2}}_r + i {p^{2}}_i \label{eq:psqReIm},\\
{\hat{\Sigma}_{ij}}({p^{2}}) &\simeq {\hat{\Sigma}_{ij}}({p^{2}}_r) + i {p^{2}}_i {\hat{\Sigma}^{'}}_{ij}({p^{2}}_r)\label{eq:SigijReIm},\end{aligned}$$ where ${\hat{\Sigma}^{'}}_{ij}{(p^{2})}\equiv \frac{d{\hat{\Sigma}}_{ij}{(p^{2})}}{d{p^{2}}}$. This expansion enables the calculation of self-energies at complex momentum in terms of self-energies evaluated at real momentum (e.g. by `FeynHiggs`). For the inclusion of all products of real and imaginary parts, we do not expand the effective self-energy from Eq.(\[eq:seff\]) directly according to Eq.(\[eq:psqReIm\]). Instead, in the same way as in Refs.[@KarinaPhD; @Fowler:2010eba; @Fuchs:2015jwa], we expand all $\hat{\Gamma}_{ij}{(p^{2})}$ in Eq.(\[eq:seff\]) individually before combining them into ${\hat{\Sigma}^{\rm{eff}}_{ii}}$.
Pole structure of propagator matrices {#sect:poles}
=====================================
In this section, we will first discuss in detail the solutions of the poles of the propagator matrix depending on the mixing. Subsequently we will prove the equivalence of different assignments between the lowest-order states and the loop-corrected mass eigenstates for the physical ${\hat{\textbf{Z}}}$-matrix.
Determination of the poles
--------------------------
Imaginary parts of the self-energies lead to propagator poles in the complex momentum plane. The Higgs masses are determined from the real part of the complex poles ${\mathcal{M}^{2}}$ of the diagonal propagators. Equivalently, the complex poles are obtained as the zeros of the inverse diagonal propagators. Due to $\boldsymbol{A}^{-1}\propto \det\left[\boldsymbol{A}\right]^{-1}$, $\det\left[\boldsymbol{A}^{-1}\right]=\det\left[\boldsymbol{A}\right]^{-1}$ and $\det\left[-\boldsymbol{A}\right]=(-1)^n\,\det\left[\boldsymbol{A}\right]$ for any $n\times n$-matrix $\boldsymbol{A}$, finding the roots of ${\boldsymbol{\Delta}}^{-1}{(p^{2})}$ is equivalent to solving $$\begin{aligned}
-\frac{1}{\det\left[{\boldsymbol{\Delta}}_{hHA}{(p^{2})}\right]}
&=\det\left[{\hat{\boldsymbol{\Gamma}}}_{hHA}{(p^{2})}\right] \stackrel{!}{=} 0,\label{eq:det0}
\end{aligned}$$ because of the relation between ${\boldsymbol{\Delta}}_{hHA}$ and ${\hat{\boldsymbol{\Gamma}}}_{hHA}$ from Eq.(\[eq:DeltaGamInv\])[^3]. Then the loop-corrected masses $M$ are obtained from the real parts of the complex poles and the total widths $\Gamma$ from the imaginary parts via $${\mathcal{M}^{2}}= M^{2} - i M\Gamma. \label{eq:MGamma}$$ In the following, the impact of higher-order and mixing contributions on the pole structure of the propagators will be discussed.
#### Lowest order
At lowest order, the self-energy contributions in Eq.(\[eq:IrepVert\]) are absent and the matrix ${\hat{\boldsymbol{\Gamma}}}$ simply reads $$\begin{aligned}
{\hat{\boldsymbol{\Gamma}}}^{(0)}_{hHA}{(p^{2})}&= i\,\textrm{diag}\left\lbrace D_h{(p^{2})}, D_H{(p^{2})},
D_A{(p^{2})}\right\rbrace \label{eq:GamLowestOrder},\end{aligned}$$ where the shorthand of Eq.(\[eq:Di\]) has been used. The solutions of Eq.(\[eq:det0\]) are the three tree-level masses $m_i^{2}$.
#### Higher order without mixing
Beyond tree-level, the self-energies are added at the available order. Restricting them to the unmixed case, ${\hat{\Sigma}}_{ij}=0$ for $i\neq j$, leads to $$\begin{aligned}
{\hat{\boldsymbol{\Gamma}}}^{(\textrm{no mix})}_{hHA}{(p^{2})}&= i\,\textrm{diag}\left\lbrace D_h{(p^{2})}+ {\hat{\Sigma}}_{hh}{(p^{2})}, D_H{(p^{2})}+{\hat{\Sigma}}_{HH}{(p^{2})}, D_A{(p^{2})}+{\hat{\Sigma}}_{AA}{(p^{2})}\right\rbrace \label{eq:GamNoMix},\end{aligned}$$ so that $$\det\left[{\hat{\boldsymbol{\Gamma}}}^{(\textrm{no mix})}_{hHA}{(p^{2})}\right]=\prod_{i=h,H,A}\left(D_i{(p^{2})}+{\hat{\Sigma}}_{ii}{(p^{2})}\right)=0 \label{eq:DetUnmix}$$ is achieved if ${p^{2}}$ fulfils the following on-shell relation $${p^{2}}- m_i^{2} + {\hat{\Sigma}}_{ii}{(p^{2})}= 0 \label{eq:0unmixed}$$ for any $i=h,H,A$. Thus, the full propagator matrix ${\boldsymbol{\Delta}}$ has three poles and each propagator $\Delta_{ii}{(p^{2})}$ has exactly one pole ${p^{2}}={\mathcal{M}^{2}}_i$ that solves Eq.(\[eq:0unmixed\]) in this case.
#### Higher order with 2x2 mixing
If now the mixing between $h$ and $H$ is taken into account, corresponding to the ${\mathcal{CP}}$-conserving case, the matrix ${\hat{\boldsymbol{\Gamma}}}$ becomes block-diagonal with the $2\times 2$ matrix ${\hat{\boldsymbol{\Gamma}}}_{hH}$ and the 2-point vertex function of $A$, which does not mix with the other states: $$\begin{aligned}
{\hat{\boldsymbol{\Gamma}}}_{hHA}{(p^{2})}&= {\begin{pmatrix}}{\hat{\boldsymbol{\Gamma}}}_{hH}{(p^{2})}&0\\
0 &\hat{\Gamma}_A{(p^{2})}{\end{pmatrix}}, \label{eq:blockdiag}\\
\det\left[{\hat{\boldsymbol{\Gamma}}}_{hHA}{(p^{2})}\right] &= \det\left[{\hat{\boldsymbol{\Gamma}}}_{hH}{(p^{2})}\right]\cdot\hat\Gamma_{A}{(p^{2})}\,.\end{aligned}$$ For a closer look at the relation between the roots of the determinant and the roots of the inverse propagator, we write down the propagators and the effective self-energy of the $\left\lbrace h,H\right\rbrace$ system explicitly. They follow from Eqs.(\[eq:Dij\]), (\[eq:Dii\]) and (\[eq:seff\]) by setting ${\hat{\Sigma}}_{hA}={\hat{\Sigma}}_{HA}=0$ or equivalently from the inversion of the $2\times 2$ submatrix ${\hat{\boldsymbol{\Gamma}}}_{hH}$: $$\begin{aligned}
\Delta_{ii}({p^{2}}) &= \frac{i\left[D_j{(p^{2})}+ {\hat{\Sigma}_{jj}}({p^{2}})\right]}{\left[D_i{(p^{2})}+{\hat{\Sigma}_{ii}}({p^{2}})\right] \left[D_j{(p^{2})}+{\hat{\Sigma}_{jj}}({p^{2}})\right]-{\hat{\Sigma}_{ij}}^{2}({p^{2}})}
=\frac{i}{{p^{2}}-m_i^{2}+{\hat{\Sigma}^{\rm{eff}}_{ii}}({p^{2}})},\label{eq:Dii2x2}\\
\Delta_{ij}({p^{2}})
&= \frac{-i{\hat{\Sigma}_{ij}}({p^{2}})}{\left[D_i{(p^{2})}+{\hat{\Sigma}_{ii}}({p^{2}})\right]\,\left[D_j{(p^{2})}+{\hat{\Sigma}_{jj}}({p^{2}})\right]-{\hat{\Sigma}_{ij}}^{2}({p^{2}})}\label{2x2eq:Dij},\\
{\hat{\Sigma}^{\rm{eff}}_{ii}}({p^{2}}) &= {\hat{\Sigma}_{ii}}({p^{2}}) -\frac{{\hat{\Sigma}_{ij}}^{2}({p^{2}})}{D_j({p^{2}})+{\hat{\Sigma}_{jj}}({p^{2}})}\label{eq:seff2x2}.\end{aligned}$$ Comparing the inverse diagonal propagators with the determinant of the submatrix ${\hat{\boldsymbol{\Gamma}}}_{hH}$, we find for $i,j\in\left\lbrace h,H \right\rbrace,~i\neq j,$ $$\begin{aligned}
\frac{1}{\Delta_{ii}{(p^{2})}} = \frac{i}{D_j{(p^{2})}+{\hat{\Sigma}}_{jj}{(p^{2})}}\,\det\left[{\hat{\boldsymbol{\Gamma}}}_{hH}{(p^{2})}\right]. \label{eq:InvPropDet2x2}\end{aligned}$$ Eq.(\[eq:InvPropDet2x2\]) reveals that both inverse diagonal propagators, $1/\Delta_{hh}$ and $1/\Delta_{HH}$, are proportional to the determinant of ${\hat{\boldsymbol{\Gamma}}}_{hH}$, which has two zeros. As opposed to the unmixed case, both zeros of $\det\left[{\hat{\boldsymbol{\Gamma}}}_{hH}{(p^{2})}\right]$ are poles of *each* of the diagonal propagators $\Delta_{hh},\Delta_{HH}$. The lowest-order states $h$ and $H$ are mixed into the loop-corrected mass eigenstates $h_1$ and $h_2$ with the loop-corrected masses $M_{h_1}, M_{h_2}$. The corresponding poles ${p^{2}}={\mathcal{M}^{2}}_{h_1}, {\mathcal{M}^{2}}_{h_2}$ solve $${p^{2}}- m_i^{2} + \hat{\Sigma}^{\textrm{eff}}_{ii}({p^{2}}) = 0 \label{eq:0mixed}$$ for ${p^{2}}={\mathcal{M}^{2}}_{h_a}$ in any combination of $i=h,H$ and $a=1,2$, where the effective self-energy is given in Eq.(\[eq:seff2x2\]). In the $2\times 2$ mixing system, it is convenient to choose $M_{h_1}\leq M_{h_2}$. As for the nomenclature in the $2\times 2$ case, the lighter mass eigenstate $h_1$ is often denoted as $h$ and the heavier one as $H$ because both are ${\mathcal{CP}}$-even states. It should be noted that both roots of ${\hat{\boldsymbol{\Gamma}}}_{hH}$, ${\mathcal{M}^{2}}_{h_1}$ and ${\mathcal{M}^{2}}_{h_2}$, are also complex poles of the off-diagonal propagators $\Delta_{hH}{(p^{2})}\equiv\Delta_{Hh}{(p^{2})}$ due to $$\begin{aligned}
\frac{1}{\Delta_{ij}{(p^{2})}} = \frac{-i}{{\hat{\Sigma}}_{ij}{(p^{2})}}\,\det\left[{\hat{\boldsymbol{\Gamma}}}_{hH}{(p^{2})}\right]. \label{eq:InvPropijDet2x2}\end{aligned}$$ Since in this case $A$ does not mix with $h$ and $H$, the third pole ${\mathcal{M}^{2}}_A$ solely solves $${\mathcal{M}^{2}}_{A} - m_A^{2} + {\hat{\Sigma}}_{AA}({\mathcal{M}^{2}}_{A}) = 0 \label{eq:solveA},$$ but no other combination of $A$ and $h_a$ satisfies the on-shell condition. $M_A$ is the loop-corrected mass of the (mass and interaction) eigenstate $A$.
#### Higher order with 3x3 mixing
Now we turn to the case where complex MSSM parameters lead to ${\mathcal{CP}}$-violating self-energies ${\hat{\Sigma}}_{hA}, {\hat{\Sigma}}_{HA}$. Thus, all three neutral Higgs lowest-order and ${\mathcal{CP}}$-eigenstates $h,H,A$ mix into the loop-corrected mass eigenstates $h_1, h_2, h_3$, which have no longer well-defined ${\mathcal{CP}}$ quantum numbers, but are admixtures of ${\mathcal{CP}}$-even and ${\mathcal{CP}}$-odd components. In this framework, ${\hat{\boldsymbol{\Gamma}}}_{hHA}$ is a full $3\times 3$ matrix with the determinant $$\begin{aligned}
\det[{\hat{\boldsymbol{\Gamma}}}_{hHA}] &= -i\left[(D_{h}+{\hat{\Sigma}}_{hh})(D_{H}+{\hat{\Sigma}}_{HH})(D_{A}+{\hat{\Sigma}}_{AA})+2{\hat{\Sigma}}_{hH}{\hat{\Sigma}}_{HA}{\hat{\Sigma}}{hA}\right.\nonumber\\
&\left.~~~~~~~~-(D_{h}+{\hat{\Sigma}}_{hh}){\hat{\Sigma}}_{HA}^2-(D_{H}+{\hat{\Sigma}}_{HH}){\hat{\Sigma}}_{hA}^2-(D_{A}+{\hat{\Sigma}}_{AA}){\hat{\Sigma}}_{hH}^2 \right] \label{eq:Det3mix},\end{aligned}$$ where we dropped the explicit ${p^{2}}$-dependence of each term for an ease of notation. Comparing Eq.(\[eq:Det3mix\]) with the diagonal and off-diagonal propagators from Eqs.(\[eq:Dii\]) and (\[eq:Dij\]), respectively, we see that their inverse is proportional to the determinant of ${\hat{\boldsymbol{\Gamma}}}_{hHA}$: $$\begin{aligned}
\frac{1}{\Delta_{ii}} &= \frac{\det\left[{\hat{\boldsymbol{\Gamma}}}_{hHA}\right]}{(D_{j}+{\hat{\Sigma}}_{jj})(D_{k}+{\hat{\Sigma}}_{kk})-{\hat{\Sigma}}_{jk}^{2}} \label{eq:InvPropDet},\\
\frac{1}{\Delta_{ij}} &= \frac{\det\left[{\hat{\boldsymbol{\Gamma}}}_{hHA}\right]}{{\hat{\Sigma}}_{jk}{\hat{\Sigma}}_{ki}-{\hat{\Sigma}}_{ij}(D_k+{\hat{\Sigma}}_{kk}) } \label{eq:InvPropijDet}\,,\end{aligned}$$ where $i\neq j\neq k\neq i$ and without summation over the indices. From Eq.(\[eq:InvPropDet\]) we conclude that all three roots ${p^{2}}={\mathcal{M}^{2}}_{h_a},\,a=1,2,3,$ of $\det\left[{\hat{\boldsymbol{\Gamma}}}_{hHA}{(p^{2})}\right]$ are complex poles of *each* of the three diagonal propagators $\Delta_{ii},\,i=h,H,A$. This means that $${\mathcal{M}^{2}}_{h_a} - m_i^{2}+\hat{\Sigma}^{\rm{eff}}_{ii}({\mathcal{M}^{2}}_{h_a}) =0\label{eq:SolvePole}$$ holds for any combination of $i$ and $a$ in the presence of $3\times 3$ mixing. Moreover, Eq.(\[eq:InvPropijDet\]) implies that also the off-diagonal propagators have as many poles as the determinant has zeros, namely three in the case of ${\mathcal{CP}}$-violating mixing. In the unmixed case, ${\hat{\Sigma}^{\rm{eff}}_{ii}}= {\hat{\Sigma}_{ii}}$ and each propagator has exactly one pole so that there is a unique mapping between $i$ and $a$, see Eq.(\[eq:0unmixed\]). On the other hand, for the general mixing case it is not unique how to relate the mass eigenstates to the interaction eigenstates. An assignment will be needed for the definition of on-shell wave function normalisation factors in Sect.\[sect:onshellZ\].
Index scheme independence of the Z-matrix {#sect:schemeIndep}
-----------------------------------------
As discussed above, the pole structure of the full propagators provides the freedom at which pole to evaluate which ${\hat{Z}}$-factor, or equivalently which loop-corrected mass eigenstate index ($a,b,c$) to assign to a lowest-order mass eigenstate index ($i,j,k$). We denote two choices for such index schemes by $$\begin{aligned}
I &\leftrightarrow (i,a), (j,b), (k,c)\label{eq:schemeI},\\
II &\leftrightarrow (j,a), (i,b), (k,c)\label{eq:schemeII}.\end{aligned}$$ This initial ambiguity, however, results in physically equivalent results. Using properties of ratios of diagonal and off-diagonal propagators and exploiting relations at a complex pole (for details, see Ref.[@Fuchs:2015jwa]), we are able to show that $$\begin{aligned}
\frac{1+{\hat{\Sigma}^{\rm{eff'}}}_{jj}({\mathcal{M}^{2}}_a)}{1+{\hat{\Sigma}^{\rm{eff'}}}_{ii}({\mathcal{M}^{2}}_a)}
=\left(\frac{\Delta_{ji}{(p^{2})}}{\Delta_{jj}{(p^{2})}}\right)^{2}_{{p^{2}}={\mathcal{M}^{2}}_a}.\end{aligned}$$ This equality provides a transformation between scheme $I$ (where $i$ and $a$ are associated, hence ${\hat{Z}}_{ai}\equiv1$) and scheme $II$ (where $j$ and $a$ are matched): $$\begin{aligned}
{\hat{\textbf{Z}}}_{ai}|_{I} = \left(\sqrt{{\hat{Z}}_a}{\hat{Z}}_{ai}\right)_{I} &= \frac{1}{\sqrt{1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_a)}}\nonumber\\
&= \frac{1}{\sqrt{1+{\hat{\Sigma}^{\rm{eff'}}}_{jj}({\mathcal{M}^{2}}_a)}}\frac{\Delta_{ji}}{\Delta_{jj}}\bigg\rvert_{{p^{2}}={\mathcal{M}^{2}}_a}
= \left(\sqrt{{\hat{Z}}_a} {\hat{Z}}_{ai} \right)_{II} = {\hat{\textbf{Z}}}_{ai}\bigg\rvert_{II}\label{eq:TransformScheme}.
\end{aligned}$$ While the values of ${\hat{Z}}_a$ and ${\hat{Z}}_{ai}$ depend on the choice of the index mapping, Eq.(\[eq:TransformScheme\]) ensures that the elements of the ${\hat{\textbf{Z}}}$-matrix, which appear in physical processes, are invariant under the choice of $a,b,c$ as a permutation of ${1,2,3}$. We also tested this relation numerically for various parameter points and always found agreement within the numerical precision.
Breit–Wigner approximation of the full propagators {#sect:fullBWZ}
==================================================
The propagator matrix depends on the momentum ${p^{2}}$ in a twofold way. On the one hand, the tree-level propagator factors $D_i{(p^{2})}={p^{2}}-m_i^{2}$ give rise to an explicit ${p^{2}}$-dependence. On the other hand, the self-energies ${\hat{\Sigma}_{ij}}{(p^{2})}$ depend on the momentum as well, but away from thresholds their ${p^{2}}$ dependence is not particularly pronounced. In this chapter, we will develop an approximation of the full mixing propagators with the aim to maintain the leading momentum dependence, but to greatly simplify the mixing contributions by making use of the on-shell ${\hat{\textbf{Z}}}$-factors.
Unstable particles and the total decay width
--------------------------------------------
In the context of determining complex poles of propagators, we now briefly discuss resonances and unstable particles, see e.g.Refs. [@Wackeroth:1996hz; @Freitas:2002ja; @Grunewald:2000ju; @Argyres:1995ym; @Beenakker:1996kn]. Stable particles are associated with a real pole of the S-matrix, whereas the self-energies of unstable particles develop an imaginary part, so that the pole of the propagator is located within the complex momentum plane off the real axis. For a single pole, the scattering amplitude $\mathcal{A}$ can be schematically written near the complex pole ${\mathcal{M}^{2}}_a$ in a gauge-invariant way as $$\mathcal{A}(s) = \frac{R}{s-{\mathcal{M}^{2}}_a}+F(s) \label{eq:gaugeinv},$$ where $s$ is the squared centre-of-mass energy, $R$ denotes the residue, while $F$ represents non-resonant contributions. The mass $M_{h_a}$ of the unstable particle $h_a$ is obtained from the real part of the complex pole ${\mathcal{M}^{2}}_a=M_{h_a}^{2}-iM_{h_a}\Gamma_{h_a}$, while the imaginary part gives rise to the total width. Accordingly, the expansion around the complex pole ${\mathcal{M}^{2}}_a$ leads to a Breit–Wigner propagator with a constant width, $${\Delta^{\textrm{BW}}}_a{(p^{2})}:=\frac{i}{{p^{2}}- {\mathcal{M}^{2}}_{a}} = \frac{i}{{p^{2}}- M_{h_a}^2 + i M_{h_a}\Gamma_{h_a}} . \label{eq:BWdef}$$ In the following, we will use a Breit–Wigner propagator of this form to describe the contribution of the unstable scalar $h_a$ with mass $M_{h_a}$ and total width $\Gamma_{h_a}$ in the resonance region.
Expansion of the full propagators around the complex poles
----------------------------------------------------------
Eqs.(\[eq:InvPropDet\]) and (\[eq:InvPropijDet\]) imply for $3\times
3$ mixing that each propagator $\Delta_{ii}, \Delta_{ij}$ has a pole at ${\mathcal{M}^{2}}_1,{\mathcal{M}^{2}}_2$ and ${\mathcal{M}^{2}}_3$. Because of this structure, an expansion of the full propagators near one single pole is not expected to yield a sufficient approximation. Instead, we will perform an expansion of the full propagators around all of their complex poles. The final expression obtained from combining the contributions from the different poles will constitute a main result of the present paper.
### Expansion of the diagonal propagators
We begin with an expansion of $\Delta_{ii}{(p^{2})}$ in the vicinity of ${\mathcal{M}^{2}}_a$ as in Eq.(\[eq:DiiApprdSeff\]), where the first factor equals the definition of the Breit–Wigner propagator of the state $h_a$, and the second factor corresponds to ${\hat{Z}}_a$ in scheme $I$ where $i$ and $a$ are associated indices. On top of that, ${\hat{Z}}_a\big\rvert_I={\hat{\textbf{Z}}}_{ai}^{2}$ as defined in Eq.(\[eq:ZiZij\]), and the elements of the ${\hat{\textbf{Z}}}$-matrix are independent of the index scheme (see Eq.(\[eq:TransformScheme\])). Thus, the following scheme-independent approximation holds for ${p^{2}}\simeq {\mathcal{M}^{2}}_a$: $$\begin{aligned}
\Delta_{ii}{(p^{2})}&\simeq {\Delta^{\textrm{BW}}}_a{(p^{2})}\,{\hat{Z}}_a\big\rvert_I = {\Delta^{\textrm{BW}}}_a{(p^{2})}\,{\hat{\textbf{Z}}}_{ai}^{2}
\label{eq:DeltaiiBWa}.\end{aligned}$$ In this approach, the mixing contributions are summarised in the on-shell $Z$-factor evaluated at ${\mathcal{M}^{2}}_a$. In contrast, the leading momentum dependence is contained in the Breit–Wigner propagator parametrised by the loop-corrected mass $M_{h_a}$ and the total width $\Gamma_{h_a}$ from the complex pole. In addition, $\Delta_{ii}({p^{2}})$ has a second pole at ${\mathcal{M}^{2}}_b$ because ${p^{2}}-m_i^{2}+{\hat{\Sigma}^{\rm{eff}}_{ii}}(p^{2})=0$ holds also at ${p^{2}}={\mathcal{M}^{2}}_b$. Analogously, we can expand ${\hat{\Sigma}^{\rm{eff}}_{ii}}$ around ${\mathcal{M}^{2}}_b$ and obtain for the diagonal propagator $$\begin{aligned}
\Delta_{ii}(p^{2}) &=\frac{i}{p^{2}-m_i^{2}+{\hat{\Sigma}^{\rm{eff}}_{ii}}(p^{2})}\\
&\simeq \frac{i}{p^{2}-m_i^{2} + {\hat{\Sigma}^{\rm{eff}}_{ii}}({\mathcal{M}^{2}}_b) +(p^{2}-{\mathcal{M}^{2}}_b)\cdot {\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_b)}\\
&=\frac{i}{(p^{2}-{\mathcal{M}^{2}}_b)\cdot\left[1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_b)\right]}\label{eq:DeltaiiMb}.\end{aligned}$$ Formally, $\frac{1}{1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_b)}$ has the structure of the definition of a ${\hat{Z}}$-factor from Eq.(\[eq:Zha\]), but in the index scheme $II$ where $b$ is assigned to $i$, whereas Eq.(\[eq:DeltaiiBWa\]) has been obtained in scheme $I$ with the $(i,a)$ assignment. Using the relation (\[eq:TransformScheme\]), we can rewrite Eq.(\[eq:DeltaiiMb\]) as $$\begin{aligned}
\Delta_{ii}({p^{2}}) &\simeq {\Delta^{\textrm{BW}}}_b({p^{2}}) \cdot \frac{1}{1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_b)}\\
&= {\Delta^{\textrm{BW}}}_b({p^{2}})\cdot \frac{1}{1+{\hat{\Sigma}^{\rm{eff'}}}_{jj}({\mathcal{M}^{2}}_b)} \left(\frac{\Delta_{ji}}{\Delta_{jj}}\right)^{2}_{{p^{2}}={\mathcal{M}^{2}}_b} \label{eq:BWbinDeltaii}\\
&= {\Delta^{\textrm{BW}}}_b({p^{2}})\cdot \left({\hat{Z}}_b\, {\hat{Z}}_{bi}^{2}\right)\bigg\rvert_{I}\label{eq:BWbSchemeia}\\
&= {\Delta^{\textrm{BW}}}_b({p^{2}})\cdot {\hat{\textbf{Z}}}_{bi}^{2} \label{eq:BWbZbi},\end{aligned}$$ where the ${\hat{Z}}$-factors in Eq.(\[eq:BWbSchemeia\]) are expressed in the same scheme as in Eq.(\[eq:DeltaiiBWa\]). Hence, in the vicinity of ${p^{2}}\simeq {\mathcal{M}^{2}}_b$, the diagonal propagator $\Delta_{ii}$ can be approximated by the Breit–Wigner propagator of $h_b$ weighted by the square of ${\hat{\textbf{Z}}}_{bi}$ that ensures the coupling to the incoming fields as Higgs boson $i$, propagation as the mass eigenstate $h_b$ and the coupling to the outgoing fields again as Higgs boson $i$. In the same manner, $\Delta_{ii}$ can be expanded around the third complex pole, ${\mathcal{M}^{2}}_c$, yielding $$\begin{aligned}
\Delta_{ii}({p^{2}}) &\simeq \frac{i}{(p^{2}-{\mathcal{M}^{2}}_c)\cdot\left[1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_c)\right]}\label{eq:DeltaiiMc}\\
&\simeq {\Delta^{\textrm{BW}}}_c({p^{2}})\cdot \frac{1}{1+{\hat{\Sigma}^{\rm{eff'}}}_{kk}({\mathcal{M}^{2}}_c)} \left(\frac{\Delta_{ki}}{\Delta_{kk}}\right)^{2}\bigg\rvert_{{p^{2}}={\mathcal{M}^{2}}_c} \label{eq:BWcinDeltaii}\\
&= {\Delta^{\textrm{BW}}}_c({p^{2}})\cdot {\hat{\textbf{Z}}}_{ci}^{2}\label{eq:BWcZci}.\end{aligned}$$ Thus, close to one of the complex poles (e.g. ${\mathcal{M}^{2}}_a$), the dominant contribution to the full propagator $\Delta_{ii}$ can be approximated by the corresponding Breit–Wigner propagator (${\Delta^{\textrm{BW}}}_a$) multiplied by the square of the respective ${\hat{\textbf{Z}}}$-factor, (${\hat{\textbf{Z}}}_{ai}^{2}$). However, close-by poles may cause overlapping resonances. In order to include this possibility and to extend the range of validity of the Breit–Wigner approximation to a more general case, we take the sum of all three Breit–Wigner contributions into account: $$\begin{aligned}
\Delta_{ii}({p^{2}})\simeq {\Delta^{\textrm{BW}}}_a({p^{2}}) \,{\hat{\textbf{Z}}}_{ai}^{2} + {\Delta^{\textrm{BW}}}_b({p^{2}}) \,{\hat{\textbf{Z}}}_{bi}^{2} + {\Delta^{\textrm{BW}}}_c({p^{2}}) \,{\hat{\textbf{Z}}}_{ci}^{2}
= \sum_{a=1}^{3}{\Delta^{\textrm{BW}}}_a({p^{2}})\,{\hat{\textbf{Z}}}_{ai}^{2}
\label{eq:iiBWsum}.\end{aligned}$$
### Expansion of the off-diagonal propagators {#sec:ExpansionOffD}
We proceed similarly for the off-diagonal propagators, which also have three complex poles so that we can expand the propagators around them. Note that ${\hat{\textbf{Z}}}_{ai}=\sqrt{{\hat{Z}}_a}$ and ${\hat{\textbf{Z}}}_{aj}=\sqrt{{\hat{Z}}_a}{\hat{Z}}_{aj}$ as defined in Eq.(\[eq:ZiZij\]). Starting at ${p^{2}}\simeq{\mathcal{M}^{2}}_a$, we express the ${\hat{Z}}$-factors in scheme $I$, $$\begin{aligned}
\Delta_{ij}{(p^{2})}&= \frac{\Delta_{ij}{(p^{2})}}{\Delta_{ii}{(p^{2})}} \Delta_{ii}{(p^{2})}\simeq {\hat{Z}}_{aj} {\hat{\textbf{Z}}}_{ai}^{2}\,{\Delta^{\textrm{BW}}}_a{(p^{2})}= {\hat{\textbf{Z}}}_{aj}{\hat{\textbf{Z}}}_{ai}\,{\Delta^{\textrm{BW}}}_a{(p^{2})}\label{eq:DeltaijBWa},\end{aligned}$$ Next, we approximate $\Delta_{ij}$ near ${p^{2}}={\mathcal{M}^{2}}_b$: $$\begin{aligned}
\Delta_{ij}{(p^{2})}&= \frac{\Delta_{ji}{(p^{2})}}{\Delta_{jj}{(p^{2})}} \Delta_{jj}{(p^{2})}\simeq {\hat{Z}}_{bi} {\hat{\textbf{Z}}}_{bj}^{2}\,{\Delta^{\textrm{BW}}}_b{(p^{2})}= {\hat{\textbf{Z}}}_{bi}{\hat{\textbf{Z}}}_{bj}\,{\Delta^{\textrm{BW}}}_b{(p^{2})}\label{eq:DeltaijBWb}.\end{aligned}$$ For ${p^{2}}\simeq{\mathcal{M}^{2}}_c$, we switch to a scheme where the indices $i$ and $c$ belong together. Thereby we can write $$\begin{aligned}
\Delta_{ij}{(p^{2})}&= \frac{\Delta_{ij}{(p^{2})}}{\Delta_{ii}{(p^{2})}} \Delta_{ii}{(p^{2})}\simeq {\hat{Z}}_{cj} {\hat{\textbf{Z}}}_{ci}^{2}\,{\Delta^{\textrm{BW}}}_c{(p^{2})}= {\hat{\textbf{Z}}}_{cj}{\hat{\textbf{Z}}}_{ci}\,{\Delta^{\textrm{BW}}}_c{(p^{2})}\label{eq:DeltaijBWc},\end{aligned}$$ which is expressed in terms of scheme-invariant ${\hat{\textbf{Z}}}$-factors. Finally, we take the sum of Eqs.(\[eq:DeltaijBWa\])-(\[eq:DeltaijBWc\]) to obtain $$\begin{aligned}
\Delta_{ij}({p^{2}}) &\simeq \sum_{a=1}^{3}{\hat{\textbf{Z}}}_{ai}\,{\Delta^{\textrm{BW}}}_a({p^{2}})\,{\hat{\textbf{Z}}}_{aj} \label{eq:ijBWsum}.\end{aligned}$$ This sum is illustrated diagrammatically in Fig.\[fig:fullvsBWZdiagram\].
(0,0)–(2,0); at (1,0.6) [$\boldsymbol{i}$]{}; (2.5,0) circle (0.5); (3,0)–(5,0); at (4,0.6) [$\boldsymbol{j}$]{}; at (5.6,0) [$\simeq$]{};
(0,0)–(1,0); at (0.5,0.6) [$\boldsymbol{i}$]{}; (1.3,0) circle (0.3); (1.6,0)–(3,0); at (2.3,0.6) [$\boldsymbol{h_1}$]{}; (3.3,0) circle (0.3); (3.6,0)–(4.6,0); at (4.1,0.6) [$\boldsymbol{j}$]{}; at (5.1,0) [$+$]{}; at (1.3,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}1}{\color{OliveGreen}i } }$]{}; at (3.3,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}1}{\color{OliveGreen}j } }$]{};
(0,0)–(1,0); at (0.5,0.6) [$\boldsymbol{i}$]{}; (1.3,0) circle (0.3); (1.6,0)–(3,0); at (2.3,0.6) [$\boldsymbol{h_2}$]{}; (3.3,0) circle (0.3); (3.6,0)–(4.6,0); at (4.1,0.6) [$\boldsymbol{j}$]{}; at (5.1,0) [$+$]{}; at (1.3,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}2}{\color{OliveGreen}i } }$]{}; at (3.3,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}2}{\color{OliveGreen}j } }$]{};
(0,0)–(1,0); at (0.5,0.6) [$\boldsymbol{i}$]{}; (1.3,0) circle (0.3); (1.6,0)–(3,0); at (2.3,0.6) [$\boldsymbol{h_3}$]{}; (3.3,0) circle (0.3); (3.6,0)–(4.6,0); at (4.1,0.6) [$\boldsymbol{j}$]{}; at (1.3,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}3}{\color{OliveGreen}i } }$]{}; at (3.3,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}3}{\color{OliveGreen}j } }$]{};
Eq.(\[eq:ijBWsum\]) represents the central result of this section, covering also the diagonal propagators in the special case of $i=j$. It shows how the full propagator can be approximated by the contributions of the three resonance regions, expressed by the Breit–Wigner propagators $\Delta_a{(p^{2})},\,a=1,2,3$, reflecting the main momentum dependence. The mixing among the Higgs bosons is comprised in the ${\hat{\textbf{Z}}}$-factors which are evaluated on-shell. Nonetheless, even a part of the momentum dependence of the self-energies is accounted for because the derivation of Eq.(\[eq:iiBWsum\]) is based on a first-order expansion of the momentum-dependent effective self-energies. Furthermore, the ${\hat{\textbf{Z}}}$-factors serve as transition factors between the loop-corrected mass eigenstates $h_a$ and the lowest-order states $i$ (although ${\hat{\textbf{Z}}}$ is not a unitary matrix transforming the states into each other). Pictorially, $\Delta_{ij}$ is a propagator that begins on the state $i$ and ends on $j$ while mixing occurs in between. Also on the RHS of Fig.\[fig:fullvsBWZdiagram\] and Eq.(\[eq:ijBWsum\]) the propagator in the $h_a$-basis begins with $i$ and ends on $j$. Thus, the coupling to the rest of any diagram connected to the propagator is well-defined. In between, each of the $h_a$ can propagate, and the correct transition is ensured by ${\hat{\textbf{Z}}}_{ai}$ and ${\hat{\textbf{Z}}}_{aj}$. All three combinations are visualised in Fig.\[fig:fullvsBWZdiagram\].
If ${\mathcal{CP}}$ is conserved and only $h$ and $H$ mix, or if two states are nearly degenerate and their resonances widely separated from the remaining complex pole, the full $3\times 3$ mixing is (exactly or approximately) reduced to the $2\times 2$ mixing case. Then the off-diagonal ${\hat{\textbf{Z}}}$-factors involving the unmixed state vanish or become negligible so that some terms in Eq.(\[eq:ijBWsum\]) approach zero.
Beyond that, if no mixing occurs among the neutral Higgs bosons, all off-diagonal full propagators as well as the off-diagonal ${\hat{\textbf{Z}}}$-factors vanish and each diagonal full propagator consists of only a single Breit–Wigner term where the ${\hat{\textbf{Z}}}$-factor is based on the diagonal self-energy instead of the effective self-energy. Thus, Eq.(\[eq:ijBWsum\]) covers all special cases of the a priori $3\times 3$ mixing among the neutral Higgs bosons.
### Uncertainty estimate {#sec:uncertainty}
As for the usual Breit–Wigner approximation in the case of a single pole, the approximation of Eq.(\[eq:ijBWsum\]) contains all resonant contributions for the case with mixing, while the non-resonant contributions that are neglected in this approximation are suppressed by powers of $\Gamma_{h_a}/M_{h_a}$. While a narrow width enhances the resonance contribution near its pole, it also causes it to fall rapidly for momenta away from the pole and thereby suppresses the non-resonant, far off-shell propagators, as well as their possible interference with a resonant term. The approximation in the off-shell regime is therefore of a comparable quality, determined by the ratios $\Gamma_{h_a}/M_{h_a}$ for the different resonances, as the standard narrow-width approximation without mixing. Interference effects between different resonances are accounted for by employing the approximation of Eq.(\[eq:ijBWsum\]). As we will discuss in more detail below, interference effects can be large for resonances that are close in mass, while on the other hand interferences between widely separated resonances are suppressed. The effects of the interference of resonances with a non-resonant background continuum can be inferred from the case without mixing.
While the on-shell approximation of Eq.(\[eq:ijBWsum\]) is based on the expansion around the complex pole, the description of the momentum dependence can be improved by incorporating higher powers of $({p^{2}}- {\mathcal{M}^{2}}_a)$. It should be noted in this context that apart from thresholds the self-energies generally depend only rather mildly on the incoming momentum.
In the following, we will assess the main sources of uncertainties that apply in the vicinity of the complex poles. The approximation given in Eq.(\[eq:ijBWsum\]) of the fully momentum-dependent propagators in terms of Breit–Wigner propagators and on-shell ${\hat{\textbf{Z}}}$-factors introduces an uncertainty by neglecting higher orders in the expansion of the effective self-energies and in the evaluation of the self-energies depending on complex momenta in Eq.(\[eq:SigijReIm\]). Besides, the pole condition is numerically not exactly fulfilled, which may represent an additional source of uncertainty.
#### Expansion of self-energies around the real part of a complex momentum
As defined in Eq.(\[eq:SigijReIm\]), we expand all self-energies as functions of a complex squared momentum ${p^{2}}$ around the real part of ${p^{2}}$ up to the first order in the imaginary part $p_i^2\equiv{\rm Im}\,{p^{2}}$. Furthermore, higher powers of $p_i^2$ can enter, $$\begin{aligned}
{\hat{\Sigma}_{ij}}({p^{2}}) &\simeq {\hat{\Sigma}_{ij}}({p^{2}}_r) + i {p^{2}}_i\, {\hat{\Sigma}^{'}}_{ij}({p^{2}}_r)
+ \frac{1}{2}\left(i {p^{2}}_i\right)^2\, {\hat{\Sigma}^{''}}_{ij}({p^{2}}_r)\,.\end{aligned}$$ In order to expand the propagators in terms of the second-order contribution we define $$\begin{aligned}
\epsilon^{{\rm Im}}_{ij} := \frac{\frac{1}{2}\left(i {p^{2}}_i\right)^2\, {\hat{\Sigma}^{''}}_{ij}({p^{2}}_r)}{{\hat{\Sigma}_{ij}}({p^{2}}_r) + i {p^{2}}_i\, {\hat{\Sigma}^{'}}_{ij}({p^{2}}_r)}\,,\end{aligned}$$ and calculate $\Delta_{ii}$ to first order in $ \epsilon^{{\rm Im}}_{ij}$. One should keep in mind, though, that we also evaluate the full propagators using the above-mentioned expansion of self-energies around the real part of the momentum. Thus, the comparison between the full and the approximated propagators is not directly affected by omitting terms of $\mathcal{O}(\epsilon^{{\rm Im}}_{ij})$, but it represents a source of uncertainty in both methods.
#### Expansion of effective self-energies around a complex pole
Regarding the effective self-energies, the expansion of ${\hat{\Sigma}^{\rm{eff}}_{ii}}{(p^{2})}$ around ${\mathcal{M}^{2}}_a$ in Eq.(\[eq:DiiApprdSeff\]) includes the term of $\mathcal{O}\left(p^2-{\mathcal{M}^{2}}_a\right)$. Additional terms in the expansion up to $\mathcal{O}\left((p^2-{\mathcal{M}^{2}}_a)^2\right)$, $$\begin{aligned}
p^{2}-m_i^{2}+{\hat{\Sigma}^{\rm{eff}}_{ii}}(p^{2})
\simeq\left(p^{2}-{\mathcal{M}^{2}}_a\right)\cdot
\left(1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_a)+\frac{1}{2}({p^{2}}- {\mathcal{M}^{2}}_a){\hat{\Sigma}^{\rm{eff''}}_{ii}}({\mathcal{M}^{2}}_a)\right)
\label{eq:d2Seff}\,,\end{aligned}$$ contribute if the term involving the second derivative of the effective self-energy at the complex pole, ${\hat{\Sigma}^{\rm{eff''}}_{ii}}({\mathcal{M}^{2}}_a)$, is non-negligible. We define $$\begin{aligned}
\epsilon^p_{a,i} &= \frac{\frac{1}{2}({p^{2}}- {\mathcal{M}^{2}}_a){\hat{\Sigma}^{\rm{eff''}}_{ii}}({\mathcal{M}^{2}}_a)}{1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_a)}
\simeq \frac{1}{2} (x-1) {\mathcal{M}^{2}}_a\, {\hat{\Sigma}^{\rm{eff''}}_{ii}}({\mathcal{M}^{2}}_a)\,,
\label{eq:epsp}\end{aligned}$$ where the approximation holds for ${\rm Re}\,{\hat{\Sigma}^{\rm{eff''}}_{ii}}({\mathcal{M}^{2}}_a),\,{\rm Im}\,{\hat{\Sigma}^{\rm{eff''}}_{ii}}({\mathcal{M}^{2}}_a) \ll 1$ and the variation of ${p^{2}}= x{\mathcal{M}^{2}}_a$ around ${\mathcal{M}^{2}}_a$. The impact of the neglected terms on the propagator $\Delta_{ii}$ reads at first order in $\epsilon^p_{a,i}$: $$\begin{aligned}
\Delta_{ii}{(p^{2})}\simeq \frac{i}{({p^{2}}-{\mathcal{M}^{2}}_a)\, (1+{\hat{\Sigma}^{\rm{eff'}}_{ii}}({\mathcal{M}^{2}}_a))}\, \cdot (1-\epsilon^p_{a,i})\,.\end{aligned}$$
The higher powers of $({p^{2}}- {\mathcal{M}^{2}}_a)$ directly translate to an expansion in terms of $\Gamma_{h_a}/M_{h_a}$ via ${\rm Im}\,{\mathcal{M}^{2}}_a = -iM_{h_a} \Gamma_{h_a}$, analogously to the unmixed case. Expanding the diagonal propagator $\Delta_{ii}{(p^{2})}$ around the complex pole ${p^{2}}\sim {\mathcal{M}^{2}}_a$ yields $$\begin{aligned}
\Delta_{ii}{(p^{2})}\simeq &\frac{i}{{p^{2}}- M_{h_a}^2 + iM_{h_a}\Gamma_{h_a} }
\cdot \left[ 1-\sum_{n=1}^{\infty} \frac{1}{n!} \left(\frac{{p^{2}}}{M_{h_a}^2 }-1 +
i\frac{\Gamma_{h_a} }{M_{h_a} }\right)^{n-1}\,M_{h_a}^{2(n-1)}\,
\hat\Sigma_{ii}^{\textrm{eff},(n)}\left({\mathcal{M}^{2}}_a\right)
\right]\,, \label{eq:expandGammaOverM}\end{aligned}$$ where $\hat\Sigma_{ii}^{\textrm{eff},(n)}\left({\mathcal{M}^{2}}_a\right) \equiv \left.\frac{\partial^n {\hat{\Sigma}^{\rm{eff}}_{ii}}{(p^{2})}}{\partial ({p^{2}})^n}\right|_{{\mathcal{M}^{2}}_a}$ denotes the $n$th derivative of the effective self-energy evaluated at the complex pole ${\mathcal{M}^{2}}_a$. The expression $\frac{{p^{2}}}{M_{h_a}^2 }-1$ approaches zero at the resonance for ${p^{2}}\rightarrow M_{h_a}$, both in the Breit–Wigner propagator in front of the square bracket and in the correction term. Beyond that, the correction scales with powers of $\Gamma_{h_a}/M_{h_a}$ multiplied by derivatives of ${\hat{\Sigma}^{\rm{eff}}_{ii}}$. While the term of the first derivative giving rise to the diagonal ${\hat{\textbf{Z}}}$-factor $Z_{ai}^2$ is included in the approximation in Eq.(\[eq:iiBWsum\]), the term linear in $\Gamma_{h_a}/M_{h_a}$ times the second derivative, and higher-order terms, are omitted and therefore represent a source of uncertainty. These higher terms exist at the real part of the resonance, ${p^{2}}=M_{h_a}^2$, but vanish at the complex pole, ${p^{2}}= {\mathcal{M}^{2}}_a$. In any case, they are not only suppressed by a narrow width compared to the mass, but also by the smallness of the higher derivatives at the pole. We confirmed numerically that already the second derivatives are suppressed by several orders compared to the first derivatives.
The expansion in Eq.(\[eq:expandGammaOverM\]) indeed corresponds exactly to the well-known unmixed case for the simplification of $M_{h_a}\rightarrow
M, \Gamma_{h_a} \rightarrow \Gamma$ and ${\hat{\Sigma}^{\rm{eff}}_{ii}}\rightarrow \Sigma$. In contrast, the presence of mixing necessitates the expansion around all complex poles. The mixing is accounted for by the effective self-energies, but the structure of the dependence on $\Gamma_{h_a}/M_{h_a}$ remains in analogy to the unmixed case.
Even in the case of no mixing, the inclusion of the diagonal ${\hat{\textbf{Z}}}$-factor does not only represent a conceptual, but also a numerical improvement of the agreement between the full and the approximate propagator compared to the pure Breit–Wigner propagator.
#### Pole condition
As an additional source of uncertainties, Eq.(\[eq:0mixed\]) is numerically not exactly fulfilled. Instead, $${\mathcal{M}^{2}}_a - m_i^{2} + \hat{\Sigma}^{\textrm{eff}}_{ii}({\mathcal{M}^{2}}_a) = o_{a,i} \label{eq:PoleNumericalZero}$$ sums up to a small number $o_{a,i}$. However, evaluating the dimensionless ratio $\epsilon^o_{a,i}:= \frac{o_{a,i}}{m_i^2}$ for all combinations of $a$ and $i$ and its impact on the propagators, this source of uncertainty is found to be negligible for the scenarios analysed in Sections \[sect:prop3light\] and \[sect:propMixextreme\].
### Amplitude with mixing based on full or Breit–Wigner propagators
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{h}$]{}; at (4,0.6) [$\boldsymbol{h}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{hh}}$]{};
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{h}$]{}; at (4,0.6) [$\boldsymbol{H}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{hH}}$]{};
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{h}$]{}; at (4,0.6) [$\boldsymbol{A}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{hA}}$]{};
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{H}$]{}; at (4,0.6) [$\boldsymbol{h}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{Hh}}$]{};
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{H}$]{}; at (4,0.6) [$\boldsymbol{H}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{HH}}$]{};
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{H}$]{}; at (4,0.6) [$\boldsymbol{A}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{HA}}$]{};
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{A}$]{}; at (4,0.6) [$\boldsymbol{h}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{Ah}}$]{};
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{A}$]{}; at (4,0.6) [$\boldsymbol{H}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{AH}}$]{};
(-1,1)–(0,0)–(-1,-1); (0,0)–(2,0); (2.5,0) circle (0.5); (3,0)–(5,0); (6,1)–(5,0)–(6,-1); at (1,0.6) [$\boldsymbol{A}$]{}; at (4,0.6) [$\boldsymbol{A}$]{}; at (2.5,-1) [$\boldsymbol{\Delta_{AA}}$]{};
In a physical process where neutral Higgs bosons can appear as intermediate particles, all of them need to be included in the prediction, see Fig.\[fig:FullPropAmpl\] and Ref.[@Fowler:2010eba]. The Higgs part of the amplitude then contains a sum over the irreducible vertex functions ${\hat{\Gamma}_}i^{X}$ (for a coupling of Higgs $i$ at the first vertex $X$) and ${\hat{\Gamma}_}j^{Y}$ (for a coupling of Higgs $j$ at the second vertex $Y$) times the fully momentum-dependent mixing propagators, $$\begin{aligned}
\mathcal{A} = \sum_{i,j=h,H,A} {\hat{\Gamma}_}i^{X} \,\Delta_{ij}{(p^{2})}\, {\hat{\Gamma}_}j^{Y} \label{eq:FullSum}.\end{aligned}$$ Applying Eq.(\[eq:ijBWsum\]), the amplitude in Eq.(\[eq:FullSum\]) can be approximated by the sum over Breit–Wigner propagators multiplied by on-shell $Z$-factors, in agreement with Ref.[@Fowler:2010eba], $$\begin{aligned}
\mathcal{A} &\simeq \sum_{i,j=h,H,A} {\hat{\Gamma}_}i^{X} \,\left[\sum_{a=1}^{3}{\hat{\textbf{Z}}}_{ai}\,{\Delta^{\textrm{BW}}}_a{(p^{2})}\,{\hat{\textbf{Z}}}_{aj}\right]\, {\hat{\Gamma}_}j^{Y} \label{eq:BWsum}\\
&= \sum_{a=1}^{3} \left(\sum_{i=h,H,A}{\hat{\textbf{Z}}}_{ai}{\hat{\Gamma}_}i^{X}\right)\,{\Delta^{\textrm{BW}}}_a{(p^{2})}\,\left(\sum_{j=h,H,A}{\hat{\textbf{Z}}}_{aj}{\hat{\Gamma}_}j^{Y}\right)\label{eq:BWsums}\\
&=\sum_{a=1}^{3} {\hat{\Gamma}_}{h_a}^{X}\,{\Delta^{\textrm{BW}}}_a{(p^{2})}\,{\hat{\Gamma}_}{h_a}^{Y}\label{eq:VerthaBW}.\end{aligned}$$ The first bracket in Eq.(\[eq:BWsums\]) represents ${\hat{\Gamma}_}{h_a}^{X}$, i.e., the vertex $X$ connected to the mass eigenstate $h_a$ as for an external Higgs boson in Eq.(\[eq:Vertexha\]). Subsequently, the second bracket is equal to the coupling of $h_a$ at vertex $Y$, ${\hat{\Gamma}_}{h_a}^{Y}$. As opposed to Sect.\[sect:ExtH\], the $h_a$ is not on-shell here, but a propagator with momentum ${p^{2}}$ between the vertices $X$ and $Y$, represented by the Breit–Wigner propagator ${\Delta^{\textrm{BW}}}_a{(p^{2})}$. So the ${\hat{\textbf{Z}}}$-factors are not only useful for the on-shell properties of external Higgs bosons, but they can also be used as an on-shell approximation of the mixing between Higgs propagators. This will be investigated numerically in Sect.\[sect:NumBWfull\].
Concerning the propagators of unstable particles, the long-standing problems related to the treatment of unstable particles in quantum field theory should be kept in mind. While the introduction of a non-zero width into the propagator of an unstable particle is indispensable for treating the resonance region where the unstable particle is close to its mass shell, such a modification of the propagator in general mixes orders of perturbation theory. This can give rise to a violation of unitarity and / or to gauge-dependent results, see e.g. Ref.[@Denner:2014zga] for a recent discussion of this issue and references therein. The approximation of the full off-shell propagators, where the effective self-energy appears in the denominator of the diagonal propagator according to Eq.(\[eq:Diieff\]), in terms of on-shell ${\hat{\textbf{Z}}}$-factors and Breit–Wigner propagators is not meant to provide a solution to this long-standing problem. It should be noted, however, that the expression of the off-shell propagator in terms of on-shell ${\hat{\textbf{Z}}}$-factors and simple Breit–Wigner factors significantly remedies the gauge dependence of the off-shell propagator. This is due to the fact that the approximation of Eq. (\[eq:DeltaiiBWa\]) and Eq. (\[eq:ijBWsum\]) is on the one hand numerically close to the full propagator (in the considered kinematic region and applying the Feynman gauge) but on the other hand does not contain gauge-dependent contributions that depend on the squared momentum $p^2$ of the propagator. While those $p^2$-dependent contributions of the gauge part of a Dyson-resummed self-energy in the denominator of a propagator often lead to violations of unitarity for high values of $p^2$ (see e.g. Ref.[@Denner:1996gb]), the Breit–Wigner factors multiplied with ${\hat{\textbf{Z}}}$-factors that are evaluated on-shell are much better behaved in the limit where $p^2$ gets large. We defer a more detailed discussion of this issue to future work.
Calculation of interference terms in the Breit–Wigner formulation {#sect:IntBW}
-----------------------------------------------------------------
In Eq.(\[eq:ijBWsum\]), the Breit–Wigner propagators are combined such that they approximate a given full propagator. Conversely, we will now separate the $h_a$ part from the contribution of the other mass eigenstates in the amplitude with Higgs exchange between the vertices $X$ and $Y$: $$\begin{aligned}
\mathcal{A}_{h_a} \equiv {\hat{\Gamma}_}{h_a}^{X}\, {\Delta^{\textrm{BW}}}_a{(p^{2})}\, {\hat{\Gamma}_}{h_a}^{Y}
= \sum_{i,j=h,H,A} {\hat{\Gamma}_}i^{X}\, {\hat{\textbf{Z}}}_{ai}\, {\Delta^{\textrm{BW}}}_a{(p^{2})}\, {\hat{\textbf{Z}}}_{aj}\, {\hat{\Gamma}_}j^{Y} \label{eq:amplha},\end{aligned}$$ i.e. the exchange of the state $h_a$ coupling with the mixed vertices ${\hat{\Gamma}_}{h_a}$ from Eq.(\[eq:Vertexha\]) as for an external Higgs.
at (0.3,0) [$\hat\Gamma_{{\color{blue} h_a}}^{X}$]{}; (0.4,1)–(1.4,0.1); (0.4,-1)–(1.4,-0.1); (1.7,0) circle (0.3); (2,0)–(4,0); at (3,0.6) [$\boldsymbol{h_a}$]{}; (4.3,0) circle (0.3); (5.6,1)–(4.6,0.1); (5.6,-1)–(4.6,-0.1); at (5.5,0) [$\hat\Gamma_{{\color{blue} h_a}}^{Y}$]{}; at (-0.5,-3) [=]{};
at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} h}}^{X}$]{}; (0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}h } }$]{}; (3.2,0)–(4.7,0); (5,0) circle (0.3); at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}h } }$]{}; (5.3,0)–(6.5,0); at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2,0.6) [$\boldsymbol{h}$]{}; at (5.9,0.6) [$\boldsymbol{h}$]{}; (7.5,1)–(6.5,0)–(7.5,-1); at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} h}}^{Y}$]{}; at (8.4,0) [+]{};
(0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); (3.2,0)–(4.7,0); (5,0) circle (0.3); (5.3,0)–(6.5,0); (7.5,1)–(6.5,0)–(7.5,-1); at (8.4,0) [+]{}; at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}h } }$]{}; at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}H } }$]{}; at (2,0.6) [$\boldsymbol{h}$]{}; at (5.9,0.6) [$\boldsymbol{H}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} h}}^{X}$]{}; at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} H}}^{Y}$]{};
(0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); (3.2,0)–(4.7,0); (5,0) circle (0.3); (5.3,0)–(6.5,0); (7.5,1)–(6.5,0)–(7.5,-1); at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}h } }$]{}; at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}A } }$]{}; at (2,0.6) [$\boldsymbol{h}$]{}; at (5.9,0.6) [$\boldsymbol{A}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} h}}^{X}$]{}; at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} A}}^{Y}$]{};
at (-0.6,0) [+]{}; (0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); (3.2,0)–(4.7,0); (5,0) circle (0.3); (5.3,0)–(6.5,0); (7.5,1)–(6.5,0)–(7.5,-1); at (8.4,0) [+]{}; at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}H } }$]{}; at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}h } }$]{}; at (2,0.6) [$\boldsymbol{H}$]{}; at (5.9,0.6) [$\boldsymbol{h}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} H}}^{X}$]{}; at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} h}}^{Y}$]{};
(0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); (3.2,0)–(4.7,0); (5,0) circle (0.3); (5.3,0)–(6.5,0); (7.5,1)–(6.5,0)–(7.5,-1); at (8.4,0) [+]{}; at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}H } }$]{}; at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}H } }$]{}; at (2,0.6) [$\boldsymbol{H}$]{}; at (5.9,0.6) [$\boldsymbol{H}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} H}}^{X}$]{}; at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} H}}^{Y}$]{};
(0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); (3.2,0)–(4.7,0); (5,0) circle (0.3); (5.3,0)–(6.5,0); (7.5,1)–(6.5,0)–(7.5,-1); at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}H } }$]{}; at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}A } }$]{}; at (2,0.6) [$\boldsymbol{H}$]{}; at (5.9,0.6) [$\boldsymbol{A}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} H}}^{X}$]{}; at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} A}}^{Y}$]{};
at (-0.6,0) [+]{}; (0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); (3.2,0)–(4.7,0); (5,0) circle (0.3); (5.3,0)–(6.5,0); (7.5,1)–(6.5,0)–(7.5,-1); at (8.4,0) [+]{}; at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}A } }$]{}; at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}h } }$]{}; at (2,0.6) [$\boldsymbol{A}$]{}; at (5.9,0.6) [$\boldsymbol{h}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} A}}^{X}$]{}; at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} h}}^{Y}$]{};
(0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); (3.2,0)–(4.7,0); (5,0) circle (0.3); (5.3,0)–(6.5,0); (7.5,1)–(6.5,0)–(7.5,-1); at (8.4,0) [+]{}; at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}A } }$]{}; at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}H } }$]{}; at (2,0.6) [$\boldsymbol{A}$]{}; at (5.9,0.6) [$\boldsymbol{H}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} A}}^{X}$]{}; at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} H}}^{Y}$]{};
(0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); (3.2,0)–(4.7,0); (5,0) circle (0.3); (5.3,0)–(6.5,0); (7.5,1)–(6.5,0)–(7.5,-1); at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}A } }$]{}; at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}A } }$]{}; at (2,0.6) [$\boldsymbol{A}$]{}; at (5.9,0.6) [$\boldsymbol{A}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} A}}^{X}$]{}; at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} A}}^{Y}$]{};
at (-2,0) [$=\displaystyle{\sum\limits_{{\color{OliveGreen}\boldsymbol{i,j=h,H,A}}}}$]{}; at (0.3,0) [$\hat\Gamma_{{\color{OliveGreen} \boldsymbol{i} }}^{X}$]{}; (0.4,1)–(1.4,0)–(0.4,-1); (1.4,0)–(2.6,0); (2.9,0) circle (0.3); at (2.9,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}i } }$]{}; (3.2,0)–(4.7,0); (5,0) circle (0.3); at (5,-0.8) [${\hat{\textbf{Z}}}_{{\color{blue}a}{\color{OliveGreen}j } }$]{}; (5.3,0)–(6.5,0); at (3.95,0.6) [$\boldsymbol{h_a}$]{}; at (2,0.6) [$\boldsymbol{i}$]{}; at (5.9,0.6) [$\boldsymbol{j}$]{}; (7.5,1)–(6.5,0)–(7.5,-1); at (7.4,0) [$\hat\Gamma_{{\color{OliveGreen} \boldsymbol{j}}}^{Y}$]{};
In order to calculate the squared amplitude as a *coherent* sum, all contributions of $h_1, h_2, h_3$ are summed up first before taking the absolute square, $$\begin{aligned}
|\mathcal{A}|^{2}_{\textrm{coh}} = \bigg\lvert \sum_{a=1}^{3} \mathcal{A}_{h_a} \bigg\rvert^{2} \label{eq:amplcoh}\,.\end{aligned}$$ On the contrary, the *incoherent* sum is the sum of the squared individual amplitudes, which misses the interference contribution, $$\begin{aligned}
|\mathcal{A}|^{2}_{\textrm{incoh}} = \sum_{a=1}^{3} \bigg\lvert\mathcal{A}_{h_a} \bigg\rvert^{2} \label{eq:amplincoh}\,.\end{aligned}$$ Thus, an advantage of the Breit–Wigner propagators is also the possibility to conveniently discern the interference of several resonances from their individual contributions in a squared amplitude $$\begin{aligned}
|\mathcal{A}|^{2}_{\textrm{int}} = |\mathcal{A}|^{2}_{\textrm{coh}} -|\mathcal{A}|^{2}_{\textrm{inccoh}}
= \sum_{a<b}2\,\textrm{Re}\left[\mathcal{A}_{h_a} \mathcal{A}_{h_b}^{*}\right].\end{aligned}$$ In contrast, the squared amplitude based on the full propagators $$\begin{aligned}
\lvert \mathcal{A}_{\textrm{full}}\rvert^{2} &= \Big\lvert\sum_{i,j=h,H,A} {\hat{\Gamma}_}i^{X} \,\Delta_{ij}{(p^{2})}\, {\hat{\Gamma}_}j^{Y} \Big\rvert^{2} \label{eq:fullA}\end{aligned}$$ does not allow for a straightforward determination of the pure interference term.
Numerical comparison in the MSSM with complex parameters {#sect:NumBWfull}
========================================================
After the analytical considerations so far, we will now compare the full propagators with their approximation as a combination of Breit–Wigner propagators and ${\hat{\textbf{Z}}}$-factors numerically. For definiteness, we will evaluate the propagators in the MSSM with complex parameters. We use `FeynHiggs-2.10.2` [^4][@Heinemeyer:1998np; @Heinemeyer:1998yj; @Degrassi:2002fi; @Heinemeyer:2007aq] for the numerical evaluation of Higgs masses, widths and mixing properties. In order to investigate the applicability of the expansion of the full propagators in one or all three resonance regions, we will first use as input a complex squared momentum around the three complex poles. For the later application to physical processes where the squared momentum equals the centre-of-mass energy $s$, we will also evaluate the propagators at ${p^{2}}=s$ near the real parts of the complex poles. In Sect.\[sect:prop3light\] we choose a scenario where all three Higgs bosons are relatively light so that we can study their mutual overlap. As a test of the ${\hat{\textbf{Z}}}$-factor approximation, we work in a scenario with large mixing between $H$ and $A$ in Sect.\[sect:propMixextreme\].
The MSSM with complex parameters at tree level {#chap:MSSM}
----------------------------------------------
In this section we fix the notation for the different particle sectors of the MSSM with complex parameters, following Ref.[@Frank:2006yh].
#### Sfermion sector {#sect:sfermion}
The mixing of sfermions $\tilde{f}_L, \tilde{f}_R$ within one generation into mass eigenstates $\tilde{f}_1, \tilde{f}_2$ is parametrised by the matrix $$\begin{aligned}
M_{\tilde{f}}^{2}&=\left(
\begin{matrix}
M_{\tilde{f}_L}^{2}+m_f^{2}+M_Z^{2}\cos 2 \beta (I_f^{3}-Q_f{s_{W}}^{2}) & m_f X_f^{*}\\
m_f X_f & M_{\tilde{f}_R}^{2}+m_{f}^{2}+M_Z^{2}\cos2\beta Q_f{s_{W}}^{2}
\end{matrix} \right),\label{eq:Msf}\\
X_f &:= A_f - \mu^{*}\cdot \left\{ \begin{matrix}
\cot \beta,~f = \text{up-type}~~~~\\\tan \beta,~f = \text{down-type}.
\end{matrix} \right.\end{aligned}$$ The trilinear couplings $A_f=|A_f|e^{i\phi_{A_f}}$, as well as $\mu=|\mu|e^{i\phi_{\mu}}$, can be complex. These phases enter the Higgs sector via sfermion loops starting at one-loop order. Diagonalising $M_{\tilde{f}}^{2}$ for all $\tilde{f}$ separately, one obtains the sfermion masses $m_{\tilde{f}_1}\leq m_{\tilde{f}_2}$.
#### Gluino sector
The gluino $\tilde{g}^{a},~a=1,2,3,$ has a mass of $ m_{\tilde{g}}=|M_3|$, where $M_3=|M_3|\,e^{i\phi_{M_3}}$ is the possibly complex gluino mass parameter. Since the gluino does not directly couple to Higgs bosons, the phase $\phi_{M_3}$ enters the Higgs sector only at the two-loop level, but has an impact for example on the bottom Yukawa coupling already at one-loop order.
#### Neutralino and chargino sector {#sect:neucha}
At tree-level, mixing in the chargino sector is governed by the higgsino and wino mass parameters $\mu$ and $M_2$, respectively. In the neutralino sector it additionally depends on the bino mass parameter $M_1$. The charginos $\widetilde{\chi}_i^{\pm},~i=1,2$, as mass eigenstates are superpositions of the charged winos $\widetilde{W}^{\pm}$ and higgsinos $\widetilde{H}^{\pm}$, with the mass matrix, $$X = {\begin{pmatrix}}M_2 & \sqrt{2}M_W{s_{\beta}}\\ \sqrt{2}M_W {c_{\beta}}& \mu {\end{pmatrix}}.
\label{eq:X}$$ Likewise in the neutralino sector, the neutral electroweak gauginos $\widetilde{B},\,\widetilde{W}^{3}$ and the neutral Higgsinos $\tilde{h}_d^0,\, \tilde{h}_u^0$ mix into the mass eigenstates ${\widetilde{\chi}_i^{0}},\,i=1,...,4$. The mixing is encoded in the gaugino mass matrix $Y$, $$Y = \left(\begin{matrix}
M_1 & 0 & -M_Z {c_{\beta}}s_W & M_Z {s_{\beta}}s_W\\
0 & M_2 & M_Z {c_{\beta}}c_W & -M_Z {s_{\beta}}c_W \\
-M_Z {c_{\beta}}s_W & M_Z {c_{\beta}}c_W & 0 &-\mu\\
M_Z {s_{\beta}}s_W & -M_Z {s_{\beta}}c_W & -\mu &0\\
\end{matrix} \right) .
\label{eq:neutmix}$$ The gaugino mass parameters $M_1$ and $M_2$ as well as the higgsino mass parameter can in principle be complex. However, only two of the phases are independent, and a frequently used convention is to set $\phi_{M_2}=0$.
#### Higgs sector {#sect:Higgstree}
The MSSM requires two complex scalar Higgs doublets with opposite hypercharge $Y_{\mathcal{H}_{1,2}} = \pm 1$, $$\begin{aligned}
\mathcal{H}_1 &= \begin{pmatrix}h_d^{0}\\h_d^{-}\end{pmatrix}
={\begin{pmatrix}}v_d+\frac{1}{\sqrt{2}}(\phi_1^{0}-i\chi_1^{0})\\ -\phi_1^{-}{\end{pmatrix}}\label{eq:H1doublet}\\
\mathcal{H}_2 &= \begin{pmatrix}h_u^{+}\\h_u^{0}\end{pmatrix}
= {\begin{pmatrix}}\phi_2^{+}\\v_u+\frac{1}{\sqrt{2}}(\phi_2^{0}+i\chi_2^{0}){\end{pmatrix}}. \label{eq:H2doublet}\end{aligned}$$ A relative phase between both Higgs doublets vanishes at the minimum of the Higgs potential, and the possible phase of the coefficient of the bilinear term in the Higgs potential can be rotated away. Hence, the Higgs sector conserves ${\mathcal{CP}}$ at lowest order, and the $4\times 4$ mass matrix of the neutral states $\mathbf{M}_{\phi\phi \chi\chi}$ becomes block-diagonal. The neutral tree level mass eigenstates are obtained by a diagonalisation of $\mathbf{M}_{\phi\phi \chi\chi}$: $$\begin{aligned}
{\begin{pmatrix}}h\\H\\A\\G {\end{pmatrix}}= {\begin{pmatrix}}-{s_{\alpha}}& {c_{\alpha}}& 0 & 0\\ {c_{\alpha}}&{s_{\alpha}}&0&0\\ 0&0& -{s_{\beta_n}}&{c_{\beta_n}}\\ 0&0 & {c_{\beta_n}}& {s_{\beta_n}}{\end{pmatrix}}{\begin{pmatrix}}\phi_1^{0}\\ \phi_2^{0}\\ \chi_1^{0}\\ \chi_2^{0}{\end{pmatrix}},\hspace*{1cm}
{\begin{pmatrix}}H^{\pm}\\G^{\pm}{\end{pmatrix}}= {\begin{pmatrix}}-{s_{\beta_c}}& {c_{\beta_c}}\\ {c_{\beta_c}}&{s_{\beta_c}}{\end{pmatrix}}{\begin{pmatrix}}\phi_1^{\pm}\\ \phi_2^{\pm}{\end{pmatrix}}\label{eq:ncHmix},\end{aligned}$$ where we introduced the short-hand notation $s_x\equiv\sin x,~c_x\equiv\cos x$. For later use we define ${t_{\beta}}\equiv \tan\beta$. The mixing angle $\alpha$ is applied for the $\mathcal{CP}$-even Higgs bosons $h,H$; $\beta_n$ for the neutral $\mathcal{CP}$-odd Higgs $A$ and Goldstone boson $G$, and $\beta_c$ for the charged Higgs $H^{\pm}$ and the charged Goldstone boson $G^{\pm}$. The minimum conditions for $V_H$ lead to $\beta = \beta_n = \beta_c$ at tree level. At higher orders, however, $\tan\beta$ is renormalised whereas the mixing angles $\alpha, \beta_n$ and $\beta_c$ are not renormalised. The masses of the $\mathcal{CP}$-odd and the charged Higgs bosons are at tree level related by $$\begin{aligned}
m_{H^{\pm}}^{2} &=& m_A^{2} + M_W^{2}\label{eq:mH+}.\end{aligned}$$ At lowest order, the Higgs sector is fully determined by the two SUSY input parameters (in addition to SM masses and gauge couplings) $\tan\beta$ and $m_{H^{\pm}}$ (or, for conserved $\mathcal{CP}$, equivalently $m_A$).
#### Higher-order corrections
Higher-order corrections in the MSSM Higgs sector are very relevant. Particles from other sectors contribute via loop diagrams to Higgs observables, in particular the trilinear couplings $A_f$, the stop and sbottom masses, and in the sub-leading terms the higgsino mass parameter $\mu$. Thus, beyond the lowest order, the Higgs sector is influenced by more parameters than only $M_{H^{\pm}}$ (or $M_A$) and $\tan\beta$. We adopt the hybrid on-shell and $\overline{\rm{DR}}$-renormalisation scheme defined in Ref.[@Frank:2006yh].
Scenario with three light Higgs bosons {#sect:prop3light}
--------------------------------------
For the numerical evaluation of the propagators, we work in the ${M_h^{\rm{mod+}}}$ scenario[@Carena:2013qia]. In this example, we fix the variable parameters $$\begin{aligned}
\mu &= 200{\,\textrm{GeV}},\nonumber\\
{M_{H^{\pm}}}&= 160{\,\textrm{GeV}},\nonumber\\
\tan\beta &= 50,\label{eq:choiceTbMHp}\end{aligned}$$ and choose for the complex phase ${\phi_{A_t}}=\pi/4$ to allow for ${\mathcal{CP}}$-violating mixing. This parameter choice is not meant to be the experimentally viable; in fact it has been experimentally ruled out by searches for $H,A\rightarrow \tau\tau$. The purpose is rather to provide a setting for illustration with nearby, but resolvable resonances so that the test of the Breit–Wigner approximation is not limited to well separated poles. These parameter values result in the following complex poles: $$\begin{aligned}
{\mathcal{M}^{2}}_1 &=(15791-70i){\,\textrm{GeV}}^{2},~~~{\mathcal{M}^{2}}_2=(16202-525i){\,\textrm{GeV}}^{2},~~~{\mathcal{M}^{2}}_3=(17388-385i){\,\textrm{GeV}}^{2}\label{eq:polevalues}.
\end{aligned}$$ All of the loop-corrected masses obtained from the real parts of the complex poles listed above are relatively light: $$\begin{aligned}
M_{h_1}&=125.7{\,\textrm{GeV}},~~~M_{h_2}=127.3{\,\textrm{GeV}},~~~M_{h_3}=131.9{\,\textrm{GeV}}\label{eq:massvalues}
\end{aligned}$$ so that the mass differences are of the order of – but not smaller than – the total widths from the imaginary parts of the complex poles, $\Gamma_{h_1}=0.6{\,\textrm{GeV}},~\Gamma_{h_2}=4.1{\,\textrm{GeV}},~\Gamma_{h_3}=2.9{\,\textrm{GeV}}$. The phase of $A_t$ induces ${\mathcal{CP}}$-violating mixing, and the on-shell mixing properties are reflected by the ${\hat{\textbf{Z}}}$-matrix obtained with `FeynHiggs`, $$\begin{aligned}
{\hat{\textbf{Z}}}&= {\begin{pmatrix}}0.95-0.04i & 0.34+0.09i &-0.05-0.05i\\
0.05-0.05i & 0.02+0.03i & 0.99-0.02i\\
-0.35-0.09i & 0.94-0.05i &-0.006-0.003i
{\end{pmatrix}}\label{eq:Zvalues},\end{aligned}$$ which indicates that $h_1$ couples mostly $h$-like, $h_2$ mostly $A$-like and $h_3$ mostly $H$-like. Thus, the contribution of $h_a$ to $\Delta_{ij}$, i.e. $$\Delta_{ij}\Big\rvert_{h_a}{(p^{2})}={\hat{\textbf{Z}}}_{ai}\,{\Delta^{\textrm{BW}}}_a{(p^{2})}\,{\hat{\textbf{Z}}}_{aj}\label{eq:hacontrib},$$ is only significant if the product ${\hat{\textbf{Z}}}_{ai}{\hat{\textbf{Z}}}_{aj}$ is not suppressed. In this case, we can already estimate that, for example, $h_3$ hardly contributes to $\Delta_{AA}$.
We have analysed all propagators $\Delta_{ij}$ around ${\mathcal{M}^{2}}_1,\,{\mathcal{M}^{2}}_2$ and ${\mathcal{M}^{2}}_3$ as well as for real momenta. In the following, we show and discuss a selection of these cases.
### Propagators depending on complex momenta
The analytical derivation of Eq.(\[eq:ijBWsum\]) builds on the expansion of the full propagators around the complex poles, and the on-shell condition in Eq.(\[eq:SolvePole\]) holds exactly only at complex momentum. Therefore, we evaluate the self-energies and propagators around the complex poles. Fig.\[fig:diag1\] displays $\Delta_{hh}{(p^{2})}$ for ${p^{2}}=
0.5\,{\mathcal{M}^{2}}_1~...~1.5\,{\mathcal{M}^{2}}_1$. In particular, Fig.\[fig:diag1Re\] shows $\textrm{Re}\left[\Delta_{hh}\right]$, and Fig.\[fig:diag1Im\] shows $\textrm{Im}\left[\Delta_{hh}\right]$ versus the ratio $x_1={p^{2}}/{\mathcal{M}^{2}}_1$ such that $x_1=1$ corresponds to the complex pole ${p^{2}}={\mathcal{M}^{2}}_1$. The black line (labelled by $\Delta$ full) represents the fully momentum-dependent mixing propagator from Eq.(\[eq:Diieff\]). Since the three poles do not have the same ratio between the real and imaginary parts, scaling $x_1$ does not run into ${\mathcal{M}^{2}}_2$ and ${\mathcal{M}^{2}}_3$. $\Delta_{hh}$ has a pole at $x=1$ and a second peak at $x\simeq1.1$ which is close to the real part of ${\mathcal{M}^{2}}_3$. This structure is very precisely reproduced (for a discussion of the uncertainties, see below) by the sum $\sum_{a=1}^{3}{\hat{\textbf{Z}}}_{ah}^{2}\,{\Delta^{\textrm{BW}}}_a{(p^{2})}$ according to Eq.(\[eq:iiBWsum\]) – as can be seen from the red dotted line (labelled as $\sum\textrm{BW}\cdot Z$), which lies directly on top of the black solid line.
In order to understand which of the Breit–Wigner propagators and ${\hat{\textbf{Z}}}$-factors dominate at which momentum, the individual contributions are shown by the dashed curves. The blue line (labelled by $h_1$) represents the contribution of $h_1$ to $\Delta_{hh}$, i.e., ${\hat{\textbf{Z}}}_{1h}^{2}\,{\Delta^{\textrm{BW}}}_1{(p^{2})}$. It clearly displays the pole at $x=1$, but strongly deviates from the full propagator at momenta away from ${\mathcal{M}^{2}}_1$. The orange line (labelled by $h_2$) represents ${\hat{\textbf{Z}}}_{2h}^{2}\,{\Delta^{\textrm{BW}}}_2{(p^{2})}$. Since ${\hat{\textbf{Z}}}_{2h}$ is small in this scenario, the contribution of $h_2$ to $\Delta_{hh}$ is numerically suppressed, but a tiny share is visible. The green line (labelled by $h_3$) stands for ${\hat{\textbf{Z}}}_{3h}^{2}\,{\Delta^{\textrm{BW}}}_3{(p^{2})}$ and it contributes significantly to $\Delta_{hh}$ near ${\mathcal{M}^{2}}_3$ because ${\hat{\textbf{Z}}}_{3h}=-0.35-0.09i$ is sizeable in this scenario. So we notice that none of the individual Breit–Wigner propagators multiplied by the appropriate ${\hat{\textbf{Z}}}$-factors suffices to approximate the full propagator, which has three complex poles. On the other hand, the sum of all three Breit–Wigner propagators times ${\hat{\textbf{Z}}}$-factors yields an accurate approach to the full mixing. This holds for the real and the imaginary part.
Having discussed the example of a diagonal propagator, we will now assess whether the ${\hat{\textbf{Z}}}$-factor approximation succeeds also for off-diagonal propagators. For instance, Fig.\[fig:offdiag2\] depicts $\Delta_{HA}$ versus $x_2={p^{2}}/{\mathcal{M}^{2}}_2$ such that $x=1$ matches ${p^{2}}={\mathcal{M}^{2}}_2$ where the propagator diverges. As above, owing to the different ratio between the real and imaginary part of each complex pole, scaling $x_2$ does not run into ${\mathcal{M}^{2}}_1$ and ${\mathcal{M}^{2}}_3$, but $\Delta_{HA}$ peaks close to their real parts. As in Fig.\[fig:diag1\], the black line representing the full propagator and the red, dotted line representing the sum of Breit–Wigner propagators according to Eq.(\[eq:ijBWsum\]) agree very well. Additionally, one can see the individual Breit–Wigner shapes. Because the products of the relevant ${\hat{\textbf{Z}}}$-factors, here ${\hat{\textbf{Z}}}_{aH}{\hat{\textbf{Z}}}_{aA}$, are non-negligible for all $a=1,2,3$, each Breit–Wigner propagator is important in the approximation of both the real part (Fig.\[fig:HARe\]) and the imaginary part (Fig.\[fig:HAIm\]) of $\Delta_{HA}$. The other diagonal and off-diagonal propagators in this scenario which are not displayed here have an equally good agreement between the full calculation and the approximation.
In order to quantify this agreement, we compute the relative deviation of the approximated propagators from the full ones, for each diagonal and off-diagonal element of the propagator matrix, depending on the momentum around each of the complex poles. For ${p^{2}}=x\cdot {\mathcal{M}^{2}}_1$, $0.5\leq x\leq 1.5$ as in Fig.\[fig:diag1\], the relative deviation of the approximated propagators from the full ones is below $4\cdot 10^{-3}$ for $\Delta_{hh}$, $\Delta_{HH}$ and below $9\cdot 10^{-3}$ for $\Delta_{AA}$ (only $\Delta_{hh}$ is displayed here), apart from the regions where the propagator itself goes through zero and changes sign. A similar agreement holds both for the diagonal and off-diagonal propagators such as in Fig.\[fig:offdiag2\], and for their real and imaginary parts. As expected, the relative deviation is minimal at the complex pole ($x=1$). The largest deviation between the full and the approximated propagators occurs (apart from the sign flips) at momenta far away from the pole where the full propagators approach zero and in some cases the individual propagators of the states $h_a$ contribute with opposite signs to sum.
The relative deviation for the displayed momenta should be compared to the expected uncertainty based on the estimates in Sect.\[sec:uncertainty\]. The higher powers of the imaginary part of the momentum that are neglected in the approximate evaluation of the self-energies at complex momenta give rise to only a small effect ($\epsilon^{\rm Im}_{ij}\sim \mathcal{O}(10^{-7})$. This induces an effect on the propagator of $\mathcal{O}(10^{-9}-10^{-6})$), which does not account for the actual difference (as explained above, both the full and the approximated propagators are subject to this kind of uncertainties arising from the evaluation at complex momenta). In contrast, the relative effect of the expansion of the momentum-dependent effective self-energy around the complex pole has a more significant effect of up to $\epsilon^p_{a,i}\lesssim 6\cdot 10^{-3}$ for the analysed momentum range, where the largest deviations occur furthest away from the poles. As a conclusion, the uncertainty arising from the approximate evaluation of the propagators at complex momenta is found to be negligible, while the expected uncertainty from the expansion around a complex pole is very close to the observed deviation.
### Propagators depending on the real momentum ${p^{2}}=s$
The calculation of the propagators at and around the complex poles together with the evaluation of the self-energies at complex momenta according to Eq.(\[eq:SigijReIm\]) was needed to fulfil the assumptions of the approximation. However, in collider processes, the Higgs propagator might appear for example in the s-channel of a $2\rightarrow 2$ scattering process where the squared momentum equals the square of the centre-of-mass energy $s$. Therefore here we will check the Breit–Wigner approximation around the real parts of the complex poles.
Fig.\[fig:hhsRe\] shows $\textrm{Re}\left[\Delta_{hh}\right]$ in the range $\sqrt{{p^{2}}}\simeq M_{h_1}, M_{h_2}, M_{h_3}$ of the three loop-corrected masses given in Eq.(\[eq:massvalues\]). The propagator has a pronounced peak around $M_{h_1}$ and a smaller and broader one at $M_{h_3}$. Again, the approximation (red, dotted) defined in Eq.(\[eq:iiBWsum\]) meets the full propagator (black) very precisely. The contribution of $h_1$ multiplied by ${\hat{\textbf{Z}}}_{1h}^{2}$ of $\mathcal{O}(1)$ dominates near $M_{h_1}$. At $M_{h_3}$ the Breit–Wigner shape of $h_3$ is dominant although multiplied only by ${\hat{\textbf{Z}}}_{3h}^{2}=(-0.35-0.09i)^{2}$, but also the tail of ${\Delta^{\textrm{BW}}}_1$ is relevant. The resonance of $h_2$ is strongly suppressed by the small ${\hat{\textbf{Z}}}_{2h}$. The described behaviour is analogous in $\textrm{Re}\left[\Delta_{HH}\right]$ (not shown here), where the strongest peak around $M_{h_3}$ is dominated by the $h_3$-contribution and the second peak near $M_{h_1}$ by the $h_1$-contribution, whereas the $h_2$ component is negligible. Also in this case, the tails of the propagators of the relevant mass eigenstates extend also to the other resonance regions. Fig.\[fig:AAsRe\] visualises $\textrm{Re}\left[\Delta_{AA}\right]$ with a broad peak at $M_{h_2}$. The black curve of the full propagator is again directly underneath the red, dotted curve of the Breit–Wigner approximation, which in this case stems nearly entirely from $h_2$ because of ${\hat{\textbf{Z}}}_{2A}\simeq 1$. Within $\Delta_{AA}$, the contribution of $h_1$ only has a minor impact, which can be seen as a small kink in Fig.\[fig:AAsRe\]. Although ${\Delta^{\textrm{BW}}}_1{(p^{2})}$ gets close to its pole, the resonance of $h_1$ is strongly suppressed by the small ${\hat{\textbf{Z}}}$-factor ${\hat{\textbf{Z}}}_{1A}=0.05 (1+i)$. As we anticipated above from the structure of ${\hat{\textbf{Z}}}$ in Eq.(\[eq:Zvalues\]), ${\Delta^{\textrm{BW}}}_{3}$ is a negligible component of $\Delta_{AA}$ for this parameter point.\
So far we have seen that the Breit–Wigner formulation combined with on-shell ${\hat{\textbf{Z}}}$-factors accurately reproduces (up to a relative deviation of ${\mathcal{O}}(10^{-3}- 10^{-2})$) the momentum dependence of the full diagonal and off-diagonal propagators by adding the contributions from all resonance regions. If all resonances are sufficiently separated (not shown in this example) or if all but one contributing products of ${\hat{\textbf{Z}}}$-factors are negligible, a single Breit–Wigner term is enough to approximate the full propagator in one of the resonance regions. In the general case, however, all Breit–Wigner terms need to be included. Even if the peaks are not located very close to each other compared to their widths, the tail of one resonance supported by a substantial product of ${\hat{\textbf{Z}}}$-factors can leak into another resonance region.
Scenario with large mixing {#sect:propMixextreme}
--------------------------
While the scenario in the previous section was characterised by three relatively similar masses, we now choose a setting with quasi degenerate heavy states $h_2$ and $h_3$. In Sect.\[sect:prop3light\] we considered the $M_h^{\rm{mod}+}$-scenario with the standard value of $\mu= 200{\,\textrm{GeV}}$ in combination with the phase $\phi_{A_t}=\pi/4$, leading to a moderate mixing predominantly between $h$ and $A$. In Ref.[@Carena:2013qia] it was suggested to choose also different values, $\mu=\pm200,\pm500, \pm1000{\,\textrm{GeV}}$. So in addition to the choice above, we now apply the following modification of the parameters in Eq.(\[eq:choiceTbMHp\]): $$\begin{aligned}
\mu &= 1000{\,\textrm{GeV}},\nonumber\\
{M_{H^{\pm}}}&= 650{\,\textrm{GeV}},\nonumber\\
\tan\beta &= 20.\label{eq:choicemue1000}\end{aligned}$$ This results in the complex poles $${\mathcal{M}^{2}}_1 = (15797-0.2i){\,\textrm{GeV}}^{2},~~{\mathcal{M}^{2}}_2 = (415336-1673i){\,\textrm{GeV}}^{2},~~{\mathcal{M}^{2}}_3 = (415554-1857i){\,\textrm{GeV}}^{2},$$ and therefore in similar masses of the heavy Higgs bosons, $$\begin{aligned}
M_{h_1} &= 125.33\,{\,\textrm{GeV}},~~~M_{h_2}= 644.47{\,\textrm{GeV}},~~~M_{h_3}= 644.63{\,\textrm{GeV}}\,.\end{aligned}$$ There is a large mixing between $H$ and $A$, visible in the ${\hat{\textbf{Z}}}$-factors evaluated with `FeynHiggs`, $$\begin{aligned}
{\hat{\textbf{Z}}}&\simeq {\begin{pmatrix}}1.01 &0 &0\\
0 &1.15-0.27i &-0.47-0.66i\\
0 &0.49+0.65i &1.13-0.28i
{\end{pmatrix}}\,.\end{aligned}$$ In the scenario in Sect.[\[sect:prop3light\]]{}, ${\hat{\textbf{Z}}}$ is approximately unitary, $$\begin{aligned}
{\hat{\textbf{Z}}}\cdot{\hat{\textbf{Z}}}^{\dagger}\Big\rvert_{\textrm{Eq.}~(\ref{eq:choiceTbMHp})} &\simeq {\begin{pmatrix}}1 &-0.1i &0.2i\\
0.1i &1 &0\\
-0.2i &0 &1
{\end{pmatrix}}\simeq \mathbb{1}.\end{aligned}$$ On the contrary, in the present scenario with $\mu=1000{\,\textrm{GeV}}$, the product ${\hat{\textbf{Z}}}\cdot{\hat{\textbf{Z}}}^{\dagger}$ deviates strongly from $\mathbb{1}$: $$\begin{aligned}
{\hat{\textbf{Z}}}\cdot{\hat{\textbf{Z}}}^{\dagger}\Big\rvert_{\textrm{Eq.}~(\ref{eq:choicemue1000})}&\simeq {\begin{pmatrix}}1 &0 &0\\
0 &2 &-1.7i\\
0 &1.7i &2 \label{eq:Znonunitary}
{\end{pmatrix}}.\end{aligned}$$ Hence it is useful to examine whether the Breit–Wigner propagators with ${\hat{\textbf{Z}}}$-factors still yield a viable approximation of the fully momentum-dependent propagators in this scenario with large mixing.
### Propagators depending on complex momenta
Due to the large difference between $M_{h_1}$ and $M_{h_2}\simeq M_{h_3}$, the propagators $\Delta_{hh},\,\Delta_{hH}$ and $\Delta_{hA}$ are strongly suppressed around ${\mathcal{M}^{2}}_2$. Fig.\[fig:mue1000ReHH\] shows $\textrm{Re}\left[\Delta_{HH}\right]$ for complex momentum around ${\mathcal{M}^{2}}_2$. The full calculation (black) and the Breit–Wigner approximation (red, dotted) are in relatively good agreement with each other although the curves do not lie directly on top of one another as in the scenario of Sect.\[sect:prop3light\]. The two heavy states $h_2$ and $h_3$ with a mass difference of less than $0.2{\,\textrm{GeV}}$ and total widths of $\Gamma_{h_2}=2.6{\,\textrm{GeV}}$ and $\Gamma_{h_3}=2.9{\,\textrm{GeV}}$ are too close to be resolved. ${\hat{\textbf{Z}}}_{2H}^{2}\Delta_{h_2}$ (orange) and ${\hat{\textbf{Z}}}_{3H}^{2}\Delta_{h_3}$ (green) contribute with similar magnitude, but opposite signs so that the result differs strongly from the single terms. Indeed, only their sum provides a reliable approximation. A comparable situation is shown in Fig.\[fig:mue1000ReHA\] for $\textrm{Re}\left[\Delta_{HA}\right]$.
In this scenario of highly admixed, almost mass-degenerate states $h_2,\,h_3$, the relative deviation of the approximated propagators from the full ones is larger, in particular in the region where the propagator itself is close to zero. The uncertainty of the evaluation of complex momenta is negligible here, as in the scenario of Sect.\[sect:prop3light\]. Instead, the reasons for the resulting deviation are the following. The neglected terms from the expansion around the complex pole (see Eq.(\[eq:epsp\])) amount to $\epsilon^p_{a,i}\sim{\mathcal{O}}(10^{-2}-10^{-1})$ for the states $i=H,\,A$ involved in the mixing around the nearby complex poles ${\mathcal{M}^{2}}_a,\,a=2,3$. The relative effect of the deviation is enhanced by the fact that the full propagator is close to zero. The approximation consists of the sum of two resonant contributions ($h_2,\,h_3$), which are in this case several orders of magnitude larger than the full propagator and occur with opposite signs. Hence, despite a cancellation of several orders of magnitude the approximation yields the correct sign of the propagator and a good description of its behaviour and size. In contrast, a single-resonance approach would result in a prediction that would be off by several orders of magnitude. As a result, although slight deviations between the the approximated propagator and the full one are visible in this scenario that is characterised by large mixing effects, the approximation based on Eq.(\[eq:ijBWsum\]) still represents a very significant improvement compared to a single-resonance treatment.
### Propagators depending on the real momentum ${p^{2}}=s$
Fig.\[fig:mue1000S\] shows the same selection of propagators as in Fig.\[fig:mue1000Mm2\], but in this case evaluated at real momentum. The approximation from Eq.(\[eq:ijBWsum\]) leads to a good agreement between the propagators in the full mixing calculation (black) and the Breit–Wigner formulation (red, dotted), displayed for $\textrm{Re}[\Delta_{HH}]$ in Fig.\[fig:mue1000sReHH\] and for $\textrm{Re}[\Delta_{HA}]$ in Fig.\[fig:mue1000sReHA\]. While the $h_1$-part (blue) is negligible due to the much lower mass $M_{h_1}$, $h_2$ (orange) and $h_3$ (green) both contribute substantially because their complex poles are very close to each other. For the evaluation at real momentum, the diagonal propagators $\Delta_{HH}, \Delta_{AA}$ are reproduced with an accuracy of ${\mathcal{O}}(10^{-3})$ by the approximation while the relative deviation is ${\mathcal{O}}(10^{-1})$ for $\Delta_{HA}$ where a cancellation of larger terms occurs. These comparisons show that the Breit–Wigner approximation is also applicable in scenarios of quasi-degenerate states and a strong resonance-enhanced mixing. We note, however, that the agreement between the full propagators and those with on-shell mixing factors is less accurate here than in the scenario with moderate, nearly unitary mixing.
### Comparison of ${\hat{\textbf{Z}}}$-factors with effective couplings {#sect:UZ}
The effective coupling approach mentioned in Sect.\[sect:Ueff\] makes use of the unitary, real ${\textbf{U}}$-matrix instead of the ${\hat{\textbf{Z}}}$-matrix. However, ${\textbf{U}}$ is not evaluated at the complex pole, but at ${p^{2}}=0$ (see Sect.\[sect:Ueff\]) and it does not comprise the imaginary parts of the self-energies. We compare the effective coupling approach based on the matrix ${\textbf{U}}$ with the ${\hat{\textbf{Z}}}$-factor approach which does take the imaginary parts into account, but cannot be directly interpreted as a unitary transformation between the states of the different bases.
Based on ${\hat{\textbf{Z}}}$- and ${\textbf{U}}$-factors from `FeynHiggs`, Fig.\[fig:mue1000UZ\] displays in the scenario of Eq.(\[eq:choicemue1000\]) real parts of $\Delta_{HH}$ and $\Delta_{HA}$ at real momentum around $M_{h_2}\simeq M_{h_3}$, calculated as a fully momentum-dependent mixing propagator (black), using the ${\hat{\textbf{Z}}}$-matrix approach (red, dotted) defined in Eq.(\[eq:ijBWsum\]), i.e.$\Delta_{ij}^{Z}({p^{2}})\simeq \sum_{a=1}^{3}{\hat{\textbf{Z}}}_{ai}\,{\Delta^{\textrm{BW}}}_a({p^{2}})\,{\hat{\textbf{Z}}}_{aj}$, and the ${\textbf{U}}$-matrix variant (grey, dashed) in the ${p^{2}}=0$ approximation as $$\Delta_{ij}^{U}({p^{2}})\simeq \sum_{a=1}^{3}{\textbf{U}}_{ai}({p^{2}}=0)\,{\Delta^{\textrm{BW}}}_a({p^{2}})\,{\textbf{U}}_{aj}({p^{2}}=0) \label{eq:UijBWsum}.$$
\
The black curves in Fig.\[fig:mue1000UZ\] representing the full propagators are identical to those shown in Fig.\[fig:mue1000S\]. While in Fig.\[fig:UZReHH\] the curve representing the approximation in terms of ${\hat{\textbf{Z}}}$-factors is almost identical to the curve of the full $\Delta_{HH}$ (with a relative deviation at the peak of $0.8\%$), the approximation in terms of ${\textbf{U}}$-factors differs from the full result by up to $14\%$. In Fig.\[fig:UZReHA\], not even the shape of $\Delta_{HA}$ is correctly approximated by the approach using the ${\textbf{U}}$-matrix whereas the approach using the ${\hat{\textbf{Z}}}$-matrix comes close to the full $\Delta_{HA}$ (up to the small deviations that can be seen in the plots).
This analysis indicates that the ${\hat{\textbf{Z}}}$-factors combined with Breit–Wigner propagators are well-suited to describe the Higgs propagators including their mixing also in scenarios with close-by resonances and strong mixing. This approach captures the leading momentum dependence and adequately accounts for the imaginary parts. In contrast, the combination of ${\textbf{U}}$-factors and Breit–Wigner propagators is – despite its unitary nature – incomplete with respect to the mixing effects in the resonance region and regarding the significance of imaginary parts.
Breit–Wigner and full propagators in cross sections {#sect:fullBWxsec}
---------------------------------------------------
As an application of the derivations above, we calculate a cross section with Higgs exchange. We study the example process $b\overline{b} \rightarrow h, H, A \rightarrow \tau^{+}\tau^{-}$, where the intermediate Higgs bosons are once represented by the full mixing propagators $\Delta_{ij}$ and once by Breit–Wigner propagators multiplied by ${\hat{\textbf{Z}}}$-factors. In order to disentangle this investigation from other higher-order effects, we restrict the Higgs-fermion-fermion vertices to the tree-level and do not include the emission of real particles in the initial or final state, but focus on the propagator corrections.
For the implementation of the full propagator method, we extended a `FeynArts` model file. New scalars $ij$ are introduced that correspond to the full propagator $\Delta_{ij}{(p^{2})}$ and couple to the first vertex as the interaction eigenstate $i$ and to the second vertex as $j$. Those propagators are used in the `FormCalc` calculation supplemented by self-energies from `FeynHiggs` incorporating corrections up to the two-loop level and the full momentum dependence at the one-loop level.
Considering only mixing between $h$ and $H$ at this point, we choose as a ${\mathcal{CP}}$-conserving scenario the $M_{h}^{\textrm{max}}$-scenario[@Carena:1999xa; @Carena:2002qg] with $\tan\beta = 50,\, {M_{H^{\pm}}}= 153\,$GeV, but we modify it by setting $A_{f_3}=2504\,$GeV. As before, this scenario has been selected for illustration purposes and it is not meant to be phenomenologically viable. An outcome of this parameter choice are large off-diagonal $Z$-factors $\hat{\textbf{Z}}_{12}\simeq 0.65+0.29i,\, \hat{\textbf{Z}}_{21}\simeq
-0.64-0.29i$, and $\hat{\textbf{Z}}_{11}\simeq 0.85-0.22i,\,
\hat{\textbf{Z}}_{22}\simeq 0.84-0.23i$. The masses of the $\mathcal{CP}$-even Higgs bosons are very close to each other, $M_{h_1}=126.20\,$GeV and $M_{h_2}=127.55$GeV, while the widths obtained from the imaginary part of the complex poles are $\Gamma_{h_1}=0.94\,$GeV and $\Gamma_{h_2}=1.21\,$GeV. Despite its large width of $\Gamma_A=3.58\,$GeV, the third neutral Higgs boson does not overlap significantly with the other two resonances due to the mass of $M_{h_3}=119.91\,$GeV, and no mixing with the other two states occurs because we are considering here the ${\mathcal{CP}}$-conserving case of real parameters.
at (0,0) [![The partonic cross section $\hat\sigma(b\overline{b}\rightarrow
\tau^{+}\tau^{-})$ in a modified $M_{h}^{\textrm{max}}$-scenario with $\tan\beta
= 50$ and ${M_{H^{\pm}}}= 153\,$GeV. The cross section is calculated with the full mixing propagators (blue, solid), approximated by the coherent sum of Breit–Wigner propagators times ${\hat{\textbf{Z}}}$-factors with the interference term (red, dashed) and the incoherent sum without the interference term (grey, dot-dashed). The individual contributions mediated by $h_1$ (light blue), $h_2$ (green) and $h_3$ (purple) are shown as dotted lines.[]{data-label="fig:DeltaBWMhmax"}](figures/bbtautau_Mhmax_DeltaBW_hHA_p2 "fig:"){width="80.00000%"}]{}; at (-67pt,112pt) [**\^**]{};
Fig.\[fig:DeltaBWMhmax\] shows the partonic cross section $\hat\sigma(b\overline{b} \rightarrow h, H, A \rightarrow \tau^{+}\tau^{-})$ as a function of the centre-of-mass energy $\sqrt{\hat s}$, where $\hat s =
(p_b+p_{\overline{b}})^{2}$ is the squared sum of the momenta of the $b$- and $\overline{b}$-quarks in the initial state. The calculation based on the full propagators (represented by the blue, solid line) is in very good agreement with the cross section based on the coherent sum of the $h_1, h_2,
h_3$ contributions of the Breit–Wigner propagators multiplied with ${\hat{\textbf{Z}}}$-factors (red, dashed) according to Eq.(\[eq:amplcoh\]). Both curves lie on top of each other and contain two peaks originating from $h_1$ (light blue, dotted) and $h_2$ (green, dotted). The resonances of $h_1$ and $h_2$ partly overlap as the mass difference is of the order of the total widths, but the two peaks can still be distinguished. The $h_3$ contribution peaks at a lower mass in this scenario, but for completeness it is also shown (purple, dotted). The incoherent sum $|h_1|^{2}+|h_2|^{2}+|h_3|^{2}$ (grey, dash-dotted) from Eq.(\[eq:amplincoh\]) clearly overestimates the full cross section as a consequence of the missing interference term that turns out to be destructive in this case. It is taken into account in the full calculation and in the coherent sum of Breit–Wigner propagators with ${\hat{\textbf{Z}}}$-factors. For the efficient calculation of interference terms of quasi degenerate resonances, see e.g. Refs.[@Fuchs:2014ola; @Fuchs:2014zra].
While the comparison in Fig.\[fig:DeltaBWMhmax\] is restricted to the case of $2\times 2$-mixing due to a scenario with real parameters, the agreement of the cross section $\hat\sigma(b\overline{b} \rightarrow h, H, A \rightarrow \tau^{+}\tau^{-})$ calculated with the full or Breit–Wigner propagators can be seen also in a scenario with complex parameters in Fig.\[fig:epsilon\]. For the numerical evaluation, we choose a scenario with large mixing defined in Eq.(\[eq:choicemue1000\]), i.e. the ${M_h^{\rm{mod+}}}$ scenario with $\mu=1000{\,\textrm{GeV}}$ and the phase ${\phi_{A_t}}=\pi/4$ where we vary ${M_{H^{\pm}}}$ and ${t_{\beta}}$.
We investigate the relative deviation $\epsilon$ between the cross section $\sigma_{\rm{full}}$ based on the full propagators and the cross section $\sigma_{\rm{coh}}^{\rm{BW}{\hat{\textbf{Z}}}}$ based on the coherent sum of Breit–Wigner propagators with ${\hat{\textbf{Z}}}$-factors, where the total widths are obtained from the imaginary parts of the complex poles, $\Gamma^{\textrm{Im}}$, defined in Eq.(\[eq:GammaIm\]), $$\begin{aligned}
\epsilon &= \frac{\sigma_{\rm{coh}}^{\rm{BW}{\hat{\textbf{Z}}}}({\phi_{A_t}})
}{\sigma_{\rm{full}}({\phi_{A_t}}) }-1\label{eq:epsilon}.\end{aligned}$$ Fig.\[fig:epsilon\] reveals that both methods agree very well, with a maximum deviation of $\pm2\%$ around ${M_{H^{\pm}}}=500{\,\textrm{GeV}},~\tan\beta=28$ and of about $0.8\%$ along the green band from this parameter point to larger values of ${M_{H^{\pm}}}$ and lower values of ${t_{\beta}}$. Otherwise the two calculations lead to the same results within $0.1\%$. Hence the use of Breit–Wigner propagators is suitable also for phenomenological applications such as the calculation of cross sections even in a scenario with complex parameters and large, ${\mathcal{CP}}$-violating mixing. A phenomenological investigation of such a scenario will be addressed in a forthcoming publication[@Fuchs:2017poi; @Fuchs:2017wkq; @IntCalc:InProgress].
Impact of the total width {#sect:Gtot}
-------------------------
This section addresses the impact of the precise value of the total width. So far, we have obtained the Higgs widths from the imaginary part of the complex poles as in Eq.(\[eq:MGamma\]) in order to consistently compare with the full propagator mixing. If the self-energies in ${\hat{\Sigma}^{\rm{eff}}_{ii}}$ are calculated at the one-loop level, the total width extracted from a complex pole of $\Delta_{ii}$ is then a tree-level width. Correspondingly, partial two-loop contributions to the imaginary parts of the self-energies give rise to partial one-loop corrections of the decay width. However, two-loop self-energies evaluated at ${p^{2}}=0$, as they are approximated in `FeynHiggs`, do not contribute to the imaginary part of the pole so that the width determined from the imaginary part of the complex pole remains at its tree-level value. Corrections to Higgs boson decays in the MSSM at and beyond the one-loop level are known and have been found to be important, see e.g. Refs.[@Dabelstein:1995js; @Djouadi:1995gt; @Williams:2011bu; @Heinemeyer:2015pfa]. Thus, the sum of the partial decay widths into any final state $X$ of a Higgs boson $h_a$, $$\Gamma_{h_a}^{\textrm{tot}} = \sum\limits_X \Gamma(h_a \rightarrow X),\label{eq:GammaTot}$$ leads to a more accurate result for its total width than from the imaginary part of the corresponding complex pole, $$\Gamma_{h_a}^{\textrm{Im}}=-\textrm{Im}[{\mathcal{M}^{2}}_a]/M_{h_a}, \label{eq:GammaIm}$$ even if ${\mathcal{M}^{2}}_a$ is based on self-energies at the same order as used for the calculation of $\Gamma_{h_a}^{\textrm{tot}} $ as in Eq.(\[eq:GammaTot\]). `FeynHiggs` contains the partial Higgs decay widths and their sum at the leading two-loop order. Having checked the compelling agreement between the full propagators and the Breit–Wigner propagators with the width from the imaginary part of the complex pole in the previous sections, now we implement the total width from `FeynHiggs` into the Breit–Wigner propagators in order to obtain the most precise phenomenological prediction.
at (0,0) [![Effect of the total width as an input for Breit–Wigner propagators: The partonic cross section $\hat\sigma(b\overline{b}\rightarrow \tau^{+}\tau^{-})$ in the same modified $M_{h}^{\textrm{max}}$-scenario as in Fig.\[fig:DeltaBWMhmax\] with $\tan\beta = 50$ and $M_{H^{+}} = 153\,$GeV. The Breit–Wigner propagators with the total widths from the imaginary part of the complex pole including (red, dashed) and excluding (grey, dash-dotted) the interference term (as in Fig.\[fig:DeltaBWMhmax\]) are compared with the Breit–Wigner propagators where the total widths are obtained from `FeynHiggs`. The corresponding results are shown including (black, solid) and excluding (black, dotted) the interference term. []{data-label="fig:GammaImTotMhmax"}](figures/bbtautau_Mhmax_BW_Im_Gtot "fig:"){width="80.00000%"}]{}; at (-60.8pt,110.5pt) [**\^**]{}; at (-42pt,114.5pt) [-]{}; at (-176pt,-1pt) [**\^**]{}; at (13pt,-109pt) [**\^**]{};
In the modified ${M_h^{\rm{max}}}$ scenario, the higher-order corrections have a significant impact on the Higgs decay widths so that $\Gamma_{h_1}^{\textrm{tot}}=2.55\,$GeV and $\Gamma_{h_2}^{\textrm{tot}}=3.24\,$GeV are much larger than the widths obtained from the imaginary part of the complex pole. This affects the order of magnitude of the cross-section $\hat\sigma\left(b\overline{b}\rightarrow \tau^{+}\tau^{-}\right)$ and the structure of the resonances, as can be seen in Fig.\[fig:GammaImTotMhmax\]. The coherent sum of Breit–Wigner propagators including the interference term (red, dashed) and the incoherent sum without the interference term (grey, dash-dotted) using $\Gamma^{\textrm{Im}}$ from the imaginary parts of the complex poles are the same as in Fig.\[fig:DeltaBWMhmax\]. In contrast, the total widths $\Gamma^{\textrm{tot}}_{\textrm{FH}}$ obtained from `FeynHiggs` as the sum of higher-order partial widths are implemented into the Breit–Wigner propagators in the cross section based on the coherent sum of all $h_a$-contributions (black, solid) and the incoherent sum (black, dotted).
The large increase in the widths from $\Gamma_{h_a}^{\textrm{Im}}$ to $\Gamma_{h_a}^{\textrm{tot}}$ has a very significant effect in Fig.\[fig:GammaImTotMhmax\], since the resonant behaviour is smeared out by the larger widths. As a consequence, the separate resonances are less pronounced, and the cross section is suppressed. Here, the incoherent sum without the interference term again overestimates the cross-sections. In addition, it lacks the two-peak structure. This observation emphasizes the importance of including the total width at the highest available precision and to take the interference term into account. One can also see that two resonances might be too close to be resolved if they are smeared by large widths.
Conclusions {#sect:conclusions}
===========
We have shown that the momentum-dependent propagator matrices of systems of unstable particles that mix with each other can be accurately approximated by a combination of Breit–Wigner propagators and wave function normalisation factors, where the latter are evaluated at the complex poles of the propagator matrix. For illustration, we have applied this approach to the example of the neutral Higgs bosons of the MSSM with complex parameters, where the ${\mathcal{CP}}$-violating interactions give rise to a $3\times 3$ mixing between the lowest-order mass eigenstates $h,H,A$ and the loop-corrected mass eigenstates $h_1,h_2,h_3$ (further mixing contributions with unphysical Goldstone bosons and vector bosons can be incorporated separately up to the considered order in perturbation theory). In the special case of ${\mathcal{CP}}$-conservation the mixing between the physical Higgs bosons is reduced to a $2\times 2$ mixing between the states $h$ and $H$.
Analysing the pole structure of propagator matrices, we have shown that for the case of $3\times 3$ mixing each entry of the propagator matrix has three complex poles, while for the $2\times 2$ mixing case each entry of the propagator matrix has two complex poles. Consequently, a single-pole approximation is not sufficient to approximate the momentum dependence of the full propagators. In particular for close-by states with sizeable widths, the different resonance regions may overlap.
We have derived in a process-independent way how the full propagators can be expanded around all of their complex poles. This approximation results in the sum of Breit–Wigner propagators of the corresponding resonances, weighted by wave function normalisation factors which encompass the relation between the lowest-order mass eigenstates and the loop-corrected mass eigenstates. As a key result, we have demonstrated that wave function normalisation factors that have been derived to ensure the correct on-shell properties of external particles in physical processes at their (in general complex) poles are a useful tool also for describing off-shell propagators, i.e. including momentum dependence. It has been shown that the momentum dependence of the full propagators can be accurately approximated by simple Breit–Wigner propagators. The complex wave function normalisation factors properly incorporate the imaginary parts of the self-energies arising from absorptive parts of loop integrals, which in contrast are neglected in effective coupling approaches where the contributing self-energies are evaluated at vanishing external momentum.
In our numerical analysis we have found very good agreement between the approximation presented in this paper and the full propagators. Besides the numerical comparison of the full and approximated propagators around the complex poles or depending on real momenta, we have provided a detailed discussion of the uncertainties involved in the approximation. In particular, the omitted higher-order terms from the expansion of self-energies around the real part of complex momenta, the expansion of the momentum-dependent effective self-energies around the complex poles and numerical effects in the on-shell condition have been quantified. The uncertainty estimates of these sources are in general well reflected in our numerical results. We have also discussed the validity of the approximation far away from the poles.
The formalism of Breit–Wigner propagators and on-shell wave function normalisation factors has several appealing advantages in describing the propagators of unstable particles that mix with each other. It avoids the momentum-dependent evaluation of self-energies and thereby significantly simplifies and accelerates the calculation of higher-order contributions. Besides, it enables the separation of the individual resonant contributions of the mass eigenstates $h_a$ within a given process and the straightforward calculation of their interference term. Moreover, the Breit–Wigner propagator turns into a $\delta$-distribution in the limit of a vanishing width, thus facilitating the separate calculation of the production and decay of an intermediate unstable particle by means of the narrow-width approximation and its generalisation to the case of overlapping and interfering resonances. In addition, the Breit–Wigner formulation allows the implementation of a more precise total width by incorporating important higher-order effects from the partial widths that are not included in the imaginary part of the complex pole with self-energies evaluated at the same perturbative order as for the partial widths. This feature is very useful for phenomenological predictions of processes involving the exchange of unstable particles, benefitting from the use of quantities computed at the highest available order in perturbation theory.
The explicit calculations presented in this paper have been performed in the MSSM, but the introduced method itself can be easily extended to the cases of particles with non-zero spin and mixing among more than three particles, such as, for example, in different models with a non-minimal scalar sector or new vector resonances.
Acknowledgements {#acknowledgements .unnumbered}
================
We thank Alison Fowler for her contributions at an early stage of this work. E.F. thanks the DESY theory group where most of this work was carried out. The work of E.F. was partially supported by the German National Academic Foundation. The work of G.W. is supported in part by the Collaborative Research Centre SFB 676 of the DFG, “Particles, Strings and the Early Universe”, and by the European Commission through the “HiggsTools” Initial Training Network PITN-GA-2012-316704.
[^1]: In order to avoid sign ambiguities, taking the square root of a $Z$-factor, which in general has two solutions in the complex plane, refers here and in the following always to the principal square root, i.e.the solution with a non-negative real part.
[^2]: This index notation differs from the conventions in Refs.[@Frank:2006yh; @Hahn:2006np; @Williams:2007dc; @Fowler:2010eba; @Williams:2011bu].
[^3]: Explicitly, all roots of ${\boldsymbol{\Delta}}_{hHA}{(p^{2})}^{-1}$ are roots of $\det\left[{\boldsymbol{\Delta}}_{hHA}{(p^{2})}\right]^{-1}
\equiv \det\left[-{\hat{\boldsymbol{\Gamma}}}_{hHA}{(p^{2})}^{-1}\right]^{-1} =
-\det\left[{\hat{\boldsymbol{\Gamma}}}_{hHA}{(p^{2})}\right]$.
[^4]: The additional contributions contained in more recent versions are not essential for the numerical comparison carried out here.
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'An algorithm is designed which decomposes a tropical univariate rational function into a composition of tropical binomials and trinomials. When a function is monotone, the composition consists just of binomials. Similar algorithms are designed for decomposing tropical algebraic rational functions being (in the classical language) piece-wise linear functions with rational slopes of their linear pieces. In addition, we provide a criterion when the composition of two tropical polynomials commutes (for classical polynomials a similar question was answered by J. Ritt).'
author:
-
---
**Decomposing tropical rational functions**
Introduction {#introduction .unnumbered}
============
We study decomposing tropical univariate rational functions (compositions of tropical rational functions find applications in deep learning of neural networks, see e. g. [@neural]). A tropical rational function is the tropical quotient (which corresponds to the subtraction in the classical sense) of two tropical polynomials. Thus, a tropical rational function is (classically) a piece-wise linear function with integer slopes of its linear pieces. A tropical root of a tropical rational function is defined as a point at which the function is not differentiable.
Relaxing the requirement that the slopes are integers allowing them to be rationals, we arrive to the concept of [*tropical algebraic rational functions*]{} or tropical Newton-Puiseux rational functions [@G] playing the role of algebraic functions in tropical algebra.
In classical algebra the problem of decomposing polynomials, rational and algebraic functions was elaborated in [@Gathen], [@Kozen], [@Rubio]. In tropical algebra the answer to the decomposing problem differs essentially from its classical counterpart. We show that a tropical rational function is a composition of binomials and trinomials. The similar holds for tropical algebraic rational functions.
In section \[one\] we introduce tropical monotone rational and algebraic rational functions and bound the number of tropical roots of compositions of tropical polynomials, monotone rational functions and rational functions.
In section \[two\] we design an algorithm which decomposes a tropical algebraic function and also a tropical monotone algebraic rational function into a composition of tropical binomials. In addition, we design an algorithm which decomposes a tropical algebraic rational function into a composition of tropical binomials and trinomials. Moreover, we provide a bound on the number of composants.
In section \[three\] decompositions of tropical rational functions (so, with integer slopes of their linear pieces) are studied. We design an algorithm which decomposes a tropical monotone rational function into a composition of tropical monotone binomials and monotone trinomials. Also we design an algorithm which decomposes a tropical rational function into a composition of tropical binomials and trinomials. In addition, a criterion is provided, when a tropical monotone trinomial is decomposable. Finally, bounds on the number of composants are given.
In section \[four\] we prove that the composition of two tropical polynomials $f,\, g$ without free terms commutes: $f\circ g=g\circ f$ iff there is a common fixed point $x_0$ (perhaps, $x_0=\infty$) for both $f,\, g$, i. e. $f(x_0)=g(x_0)=x_0$ and there exist a tropical increasing algebraic rational function $h$ and integers $a,\, b\ge 1,\, k,\, m\ge 0$ such that either $f=h^k,\, g=h^m$ or $f=ax+x_0(1-a),\, g=bx+x_0(1-b)$ on the interval $(-\infty, x_0]$ (similar conditions hold on the interval $[x_0,\, \infty)$), unless $f=x+c_1,\, g=x+c_2,\, x\in \RR$ for some $c_1,\, c_2 \in \RR$. Also we provide an example of a pair of increasing tropical rational functions with commuting composition which do not satisfy the latter conditions. For classical polynomials the answer to commutativity was given in [@R22], [@R23] (more recent generalizations and further references one can find in [@P]), in which commuting Chebyshev polynomials play a crucial role.
In section \[five\] we introduce tropical polynomial (respectively, Laurent polynomial and rational) parametrizations of polygonal lines. We show that any polygonal line admits a tropical rational parametrization and provide criteria when it does admit a tropical polynomial (respectively, Laurent polynomial) parametrization.
Tropical monotone rational functions {#one}
====================================
Recall (see e. g. [@MS]) that a (univariate) tropical polynomial has a form $f=\min_{0\le i\le d}\{ix+a_i\},\, a_i\in \RR \cup \{\infty\}$. Linear functions $ix+a_i,\, 0\le i\le d$ are called tropical monomials. So, the minimum plays the role of the addition in tropical algebra, while the addition plays the role of the multiplication. Thus, $f$ is a convex piece-wise linear function with integer slopes of the edges of its graph (sometimes, slightly abusing the terminology we call them the edges of $f$). We consider the natural ordering of the edges from the left to the right. A point $x\in \RR$ is a tropical root of $f$ if the minimum in $f$ is attained at least for two linear functions $ix+a_i,\, 0\le i\le d$. In other words, tropical roots of $f$ are the points at which $f$ is not differentiable.
A tropical rational function is a difference (which plays the role of the division in tropical algebra) of two tropical polynomials. It is a piece-wise linear function. So, its graph consists of several edges. Conversely, any continuous piece-wise linear function with integer slopes of its linear pieces (edges of its graph) is a tropical rational function (cf. [@G] where one can find further references). As the roots of a tropical rational function we again mean the points at which the function is not differentiable.
If $g,\, h$ are tropical rational functions with $p,\, q$ tropical roots, respectively, then (see [@GP]) the number of the roots of
$\bullet$ $\min\{g,\, h\}$ is at most $p+q+1$;
$\bullet$ $g+h$ or $g-h$ is at most $p+q$.
In this paper we study compositions $g\circ h$ (being tropical rational functions as well). Note that if $g,\, h$ are tropical polynomials then $g\circ h$ is also a tropical polynomial. If $s_1,\dots,s_k$ are consecutive (integer) slopes of (the linear pieces of) a tropical rational function $g$ then $g$ is a tropical polynomial iff $s_1>\cdots >s_k\ge 0$.
In a tropical [*monotone*]{} increasing (or decreasing, respectively) rational function $g$ its slopes are positive (respectively, negative). Note that any tropical polynomial $\min_{1\le i\le d} \{ix+a_i\}$ without free term is monotone increasing.
One can directly verify the following proposition.
\[roots\] If $g,\, h$ are tropical monotone rational functions with $p,\, q$ tropical roots, respectively, then the tropical monotone rational function $g\circ h$ has at most $p+q$ tropical roots. Moreover, if on an interval $[a,b]\subset \RR$ function $h$ is linear with a slope $s$, and $g$ is linear with a slope $l$ on the interval $[h(a),\, h(b)]$ (respectively, $[h(b),\, h(a)]$) when $h$ increases (respectively, decreases) then $g\circ h$ is linear on the interval $[a,\, b]$ with the slope $sl$.
\[roots-rational\] In general, the number of tropical roots of the composition $g\circ h$ of tropical rational functions does not exceed $pq+p+q$. Moreover, if $s_0,\dots,s_p$ (respectively, $t_0,\dots,t_q$) are the slopes of (the graph of) $g$ (respectively, $h$) listed with possible repetitions (multiplicities), then the slopes of $g\circ h$ are among $s_it_j,\, 0\le i\le p,\, 0\le j\le q$.
For a tropical rational function $g=\min\{x+1,\, -x+1\}$ the number of the tropical roots of $k$ iterations of $g^k:=g\circ\cdots\circ g$ is $2^k-1$ [@neural] (see also [@GP]).
Admitting rational coefficients in $\min_i \{b_ix+a_i\},\, 0\le b_i\in \QQ$, we arrive to the concept of [*tropical algebraic functions*]{} (or [*tropical Newton-Puiseux polynomials*]{}) [@G]. Respectively, we consider [*tropical algebraic rational functions*]{} being differences of tropical algebraic functions [@G].
The above statement in Proposition \[roots\] on the slopes of tropical rational functions holds for tropical algebraic rational functions as well with the difference that now we admit rational slopes rather than just integers. The above bounds on the number of tropical roots also hold literally for tropical algebraic rational functions.
Decomposing tropical algebraic rational functions {#two}
=================================================
In this section we consider tropical algebraic rational functions. As a [*tropical algebraic rational binomial*]{} we mean a function of the form either $\min\{b_1x+a_1,\, b_2x+a_2\},\, 0\neq b_1,\, b_2 \in \QQ$ or $\max\{b_1x+a_1,\, b_2x+a_2\}$. In the geometric language the former function is a convex piece-wise linear function with two (unbounded) edges (and we call it a [*tropical algebraic binomial*]{}), while the latter one is concave. If $b_1,\, b_2 >0$ then in both cases the functions are monotone increasing.
\[monotone-algebraic\] (i) There is an algorithm which for a tropical algebraic function $f$ with $k$ tropical roots yields a decomposition of $f$ into $k$ tropical algebraic binomials;
\(ii) let $f$ be a tropical monotone algebraic rational function with $k$ tropical roots. Then the algorithm yields a decomposition of $f$ into $k$ tropical monotone algebraic rational binomials.
\[monotone-algebraic-number\] Due to Proposition \[roots\] and taking into the account that each tropical algebraic rational binomial has a single tropical root, we conclude that in Proposition \[monotone-algebraic\] one can’t take less than $k$ composants.
[**Proof**]{}. The proofs for both items (i), (ii) proceed similarly. Let $f$ have consecutive slopes $s_0,\dots,s_k$ of its linear pieces. Recall that $s_0>s_1>\cdots >s_k\ge 0$ in case (i) and $s_1,\dots,s_k>0$ in case (ii). Denote by $x_l$ the $l$-th tropical root of $f,\, 1\le l\le k$. Take a (piece-wise linear) function $h$ with $k$ slopes $$s_0.s_1,\dots,s_{l-1},s_{l+1}\cdot s_{l-1}/s_l,\dots,s_k\cdot s_{l-1}/s_l$$ coinciding with $f$ for $x\le x_l$ and replacing $f$ by the composition with the linear function $((s_{l-1}/s_l)x+f(x_l)(1-s_{l-1}/s_l))\circ f$ for $x\ge x_l$. Thus, $h$ has the tropical roots $x_1,\dots,x_{l-1},x_{l+1},\dots,x_k$. The described procedure replacing $f$ by $h$ we call [*straightening*]{}: one tropical root (at $x_l$) disappears.
Take a tropical algebraic rational binomial $g$ coinciding with the identity function $x\to x$ for $x\le f(x_l)$ and with the linear function $(s_l/s_{l-1})x+f(x_l)(1-s_l/s_{l-1})$ for $x\ge f(x_l)$. Then $f=g\circ h$. Note that in case (i) $g$ is a tropical algebraic binomial since $s_{l-1}/s_l<1$.
Proceeding by induction on $k$ we complete the proof of the Proposition. $\Box$
Observe that each tropical root of $f$ corresponds to a suitable composant in a decomposition of $f$. Thus, by choosing (in the proof of Proposition \[monotone-algebraic\] above) the tropical roots in different orders, we obtain $k!$ “combinatorially different types” of decompositions of $f$.
Now let $f$ be a tropical algebraic rational function. Let for definiteness the first edge of $f$ with a non-zero slope have a positive slope. Consider tropical roots $x$ of $f$ such that $f$ has an edge with a negative slope to the right of $x$. If there does not exist such $x$ then $f$ is (non-strictly) monotone increasing, and we proceed to study the monotone case later. Among such $x$ pick $x_0$ (perhaps, if not unique then pick any of them) with the maximal value $f(x_0)$. Take a tropical root $x_1>x_0$ of $f$ with the minimal value $f(x_1)$. We have $f(x_1)<f(x_0),\, f([x_0,\, x_1])=[f(x_1),\, f(x_0)]$ and $\max\{f(x):\, x\in (-\infty,\, x_0]\}=f(x_0)$.
First we consider the case when $f(x)\le f(x_0)$ for all $x\ge x_1$. Note that in this case $f(x)\le f(x_0)$ for all $x\ge x_0$ due to the choice of $x_0$.
If both adjacent to $x_0$ edges of $f$ have non-zero slopes then the edge to the left from $x_0$ has a positive slope $s_0>0$, while the edge to the right from $x_0$ has a negative slope $s_1<0$ (due to the choice of $x_0$). Take as $g$ a tropical algebraic rational binomial which coincides with the identity function $x\to x$ for $x\le f(x_0)$ and with a linear function $(s_1/s_0)x+f(x_0)(1-s_1/s_0)$ for $x\ge f(x_0)$. So, $g$ is a tropical non-monotone algebraic rational binomial.
As $h$ take a tropical algebraic rational function which coincides with $f$ for $x\le x_0$ and coincides with the composition with the linear function $((s_0/s_1)x+f(x_0)(1-s_0/s_1))\circ f$ for $x\ge x_0$. Then $f=g\circ h$. By a [*block of edges*]{} of $f$ we mean a sequence of consecutive edges of the equal signs of their slopes (ignoring edges with zero slopes). Observe that $h$ has one less block of edges than $f$ does. Thus, by passing from $f$ to $h$ we straighten $f$ at point $x_0$.
Otherwise, if one of adjacent to $x_0$ edges has zero slope then as $g$ take a tropical binomial coinciding with the identity function $x\to x$ for $x\le f(x_0)$ and with the linear function $-x+2f(x_0)$ for $x\ge f(x_0)$. As $h$ take a tropical algebraic rational function which coincidies with $f$ for $x\le x_0$ and with the composition with the linear function $(-x+2f(x_0))\circ f$ for $x\ge x_0$. Then again $f=g\circ h$, and $h$ has one less block of edges than $f$ does. On the other hand, $h$ has the same number of tropical roots as $f$ does, so one does not straighten a piece-wise linear function at a point if one of two adjacent edges to this point has zero slope.
Now we proceed to the case when $f(x)$ takes a value greater than $f(x_0)$ for some $x>x_1$. Then $\min\{f(x): \, x\ge x_0\} = f(x_1)$.
A [*tropical regular algebraic rational trinomial*]{} is a piece-wise linear function with 3 edges having rational non-zero slopes. If the slopes are decreasing or increasing positive integers we talk about a [*tropical trinomial*]{}.
Construct the following tropical algebraic rational functions $h,\, g$. If both the edge of $f$ with the right (and respectively, the left) end-point $(x_0,f(x_0))$ has a non-zero slope $s_+$ (respectively, a non-zero slope $s_-$) then $h$ on the interval $(-\infty,\, x_0]$ coincides with the composition $(-(s_-/s_+)x+f(x_0)(1+s_-/s_+))\circ f$. Note that $s_+>0,\, s_-<0$. As $g$ take a function which on the interval $(-\infty,\, f(x_0)]$ coincides with the linear function $-(s_+/s_-)x+f(x_0)(1+s_+/s_-)$. We have $\max\{h(x): \, x\le x_0\}=f(x_0)$ and $g((-\infty,\, f(x_0)])=(-\infty,\, f(x_0)]$. In case if $s_+\cdot s_- =0$ then $h$ on the interval $(-\infty,\, x_0]$ coincides with $f$, and $g$ on the interval $(-\infty,\, f(x_0)]$ coincides with the identity function $x\to x$.
On the interval $[x_0,\, x_1]$ the function $h$ in both cases coincides with the composition $(-x+2f(x_0))\circ f$, and $g$ on the interval $[f(x_0),\, 2f(x_0)-f(x_1)]$ coincides with the linear function $-x+2f(x_0)$. Then $h([x_0,\, x_1])=[f(x_0),\, 2f(x_0)-f(x_1)]$ and $g([f(x_0,\, 2f(x_0)-f(x_1)])=[f(x_1),\, f(x_0)]$.
Finally, define $h$ on the interval $[x_1,\infty)$ and $g$ on the interval $[2f(x_0)-f(x_1),\, \infty)$. Similar to the consideration above of the interval $(-\infty, x_0]$ denote by $t_-$ (respectively, $t_+$) the slope of the edge of $f$ with the right (respectively, the left) end-point $(x_1,\, f(x_1))$. If $t_-\cdot t_+ \neq 0$ (in this case $t_-<0,\, t_+>0$ due to the choice of $x_1$) then $h$ on the interval $[x_1,\infty)$ coincides with the composition with the linear function $(-(t_-/t_+)x+2f(x_0)+f(x_1)(t_-/t_+-1))\circ f$. In this case $g$ on the interval $[2f(x_0)-f(x_1),\, \infty)$ coincides with the linear function $-t_+/t_-x + f(x_1)+t_+/t_-(2f(x_0)-f(x_1))$. Then $\min\{h(x)\, : \, x_1\le x<\infty\}=
2f(x_0)-f(x_1)$ and $g([2f(x_0)-f(x_1),\, \infty))=[f(x_1),\, \infty)$.
Otherwise, if $t_-\cdot t_+ =0$ then $h$ on the interval $[x_1,\, \infty)$ coincides with the composition $(x+2f(x_0)-2f(x_1))\circ f$, and $g$ on the interval $[2f(x_0)-f(x_1),\, \infty)$ coincides with the linear function $x-2f(x_0)+2f(x_1)$. In this case again $\min\{h(x)\, :\, x_1\le x<\infty\}=2f(x_0)-f(x_1)$ and $g([2f(x_0)-f(x_1),\, \infty))=[f(x_1),\, \infty)$. Thus, $f=g\circ h$.
Observe that $h$ has two less blocks of edges than $f$ does. Also note that $x$ is a tropical root of $h$ with the adjacent to $x$ edges of $h$ with the equal signs of their slopes iff $x$ is a tropical root of $f$ satisfying the same property (we call such $x$ a non-extremal tropical root of $h$ because $x$ is not a local extremal of $h$). In addition, the numbers of edges with zero slope are the same for $f$ and for $h$.
Thus, applying two described decomposition procedures to $f$ and obtaining $g$ to be either a tropical non-monotone algebraic rational binomial or a tropical regular algebraic rational trinomial, while it is possible, we arrive to a tropical algebraic rational function $f_0$ which is non-decreasing, so the slopes of its edges are non-negative. Thus, $$\begin{aligned}
\label{1}
f=g_1\circ \cdots \circ g_k \circ f_0\end{aligned}$$ where each of $g_1,\dots,g_k$ is either a tropical non-monotone algebraic rational binomial (their number among $g_1,\dots,g_k$ denote by $k_2$) or a tropical regular algebraic rational trinomial (their number denote by $k_3:=k-k_2$). Therefore, $k_2+2k_3$ equals the number of blocks of edges of $f$.
Now take a non-extremal tropical root $x_4$ of $f_0$. Let $s_->0$ (respectively, $s_+>0$) be the slope of the adjacent to $x_4$ left edge (respectively, right edge) of $f_0$. Denote by $g^{(1)}$ a tropical monotone increasing algebraic rational binomial which coincides with the identity function on the interval $(-\infty,\, f_0(x_4)]$ and which coincides with the linear function $(s_+/s_-)x+f(x_0)(1-s_+/s_-)$ on the interval $[f_0(x_4),\, \infty)$. Denote by $h_0$ a tropical non-decreasing algebraic rational function which coincides with $f_0$ on the interval $(-\infty,\, x_4]$ and which coincides with the composition $((s_-/s_+)x+f_0(x_4)(1-s_-/s_+))\circ f_0$ on the interval $[x_4,\, \infty)$.
Then $f_0=g^{(1)}\circ h_0$ and $h_0$ has no tropical root at $x_4$, while having all other tropical roots of $f_0$, so $h_0$ is a straightening of $f_0$. Applying the just described procedure to all non-extremal tropical roots of $f_0$, we obtain a decomposition $$\begin{aligned}
\label{2}
f_0=g^{(1)}_1\circ \cdots \circ g^{(1)}_{k_1}\circ f^{(1)}\end{aligned}$$ where each of $g^{(1)}_1,\dots,g^{(1)}_{k_1}$ is a tropical increasing algebraic rational binomial. Every second edge of $f^{(1)}$ has zero slope, and the number $k_0$ of edges with zero slope of $f^{(1)}$ equals the same number of $f$. Observe that $k_1$ equals the number of non-extremal tropical roots of $f$ (and also equals the number of non-extremal tropical roots of $f_0$). Hence $k_1+k_2+2k_3$ does not exceed the number of edges with non-zero slopes of $f$.
As a [*tropical singular algebraic rational trinomial*]{} we mean a trinomial whose middle edge has zero slope. Slightly abusing the terminology, we admit singular trinomials without one or two edges with non-zero slopes.
We are looking for a decomposition $$\begin{aligned}
\label{3}
f^{(1)}=g^{(0)}_{k_0}\circ \cdots \circ g^{(0)}_1\end{aligned}$$ where $g^{(0)}_i,\, 1\le i\le k_0$ is a tropical singular algebraic trinomial. To decompose take the left-most interval $[x_0,\, x_1]$ on which $f^{(1)}$ is constant, in other words, the edge of $f^{(1)}$ on $[x_0,\, x_1]$ has zero slope. It can happen that $x_0=-\infty$, in this case some of the following considerations become void. Define $g^{(0)}_1$ on the interval $(-\infty,\, x_0]$ as the identity function and on the interval $[x_0,\, x_1]$ as the constant function with the value $x_0$.
Let $f^{(1)}$ on the interval $(-\infty,\, x_0]$ equal a linear function $sx+r$ (so, $s$ is the slope of the left-most edge of $f^{(1)}$), in particular $sx_0+r=f^{(1)}(x_0)$. Define $g^{(0)}$ (later we’ll get that $g^{(0)}=g^{(0)}_{k_0}\circ \cdots \circ g^{(0)}_2$) on the interval $(-\infty,\, x_0]$ as the linear function $sx+r$. Therefore, $g^{(0)}\circ g^{(0)}_1$ on the interval $(-\infty,\, x_0]$ coincides with $f^{(1)}$. The same coincidence holds on the interval $[x_0,\, x_1]$ as well. Let the edge of $f^{(1)}$ with the left end-point $(x_1,\, f^{(1)}(x_1)=sx_0+r)$ have a slope $p$. Then define $g^{(0)}_1$ on the interval $[x_1,\, \infty)$ as the linear function $(p/s)x+x_0-(p/s)x_1$. Also define $g^{(0)}$ on the interval $[x_0,\, \infty)$ as the composition $f^{(1)}\circ ((s/p)x-(s/p)x_0+x_1)$. Then $f^{(1)}=g^{(0)}\circ g^{(0)}_1$.
Now we observe that $g^{(0)}$ is a continuous non-decreasing piece-wise linear function: we have constructed it by gluing at $x_0$ two non-decreasing piece-wise linear functions both having the value $sx_0+r=f^{(1)}(x_0)=f^{(1)}(x_1)$ at $x_0$. Moreover, the slope of the edge of $g^{(0)}$ with the right end-point $(x_0,\, f^{(1)}(x_0))$ equals $s$ which coincides with the slope of the edge of $g^{(0)}$ with the left end-point $(x_0,\, f^{(1)}(x_0))$. Therefore, $g^{(0)}$ has no tropical root at $x_0$ (so, $g^{(0)}$ is a straightening of $f^{(1)}$), and $g^{(0)}$ is of a similar shape as $f^{(1)}$, i. e. $g^{(0)}$ is a non-decreasing piece-wise linear function whose every second edge has zero slope. On the other hand, $g^{(0)}$ has one less edge with zero slope than $f^{(1)}$ does. Continuing in this way, we construct a required decomposition (\[3\]).
Combining (\[1\]), (\[2\]) and (\[3\]) we complete the proof of the following theorem.
\[algebraic-rational\] There is an algorithm which decomposes a tropical algebraic rational function $$\begin{aligned}
\label{4}
f=g_1\circ \cdots \circ g_k\circ g^{(1)}_1\circ \cdots \circ g^{(1)}_{k_1} \circ g^{(0)}_{k_0} \circ \cdots \circ g^{(0)}_1\end{aligned}$$ where each $g_i,\, 1\le i\le k$ is either a tropical regular algebraic rational trinomial or a tropical non-monotone algebraic rational binomial (cf. (\[1\])), each $g^{(1)}_j,\, 1\le j\le k_1$ is a tropical monotone algebraic rational binomial (cf. (\[2\])), and each $g^{(0)}_l,\, 1\le l\le k_0$ is a tropical singular algebraic trinomial (cf. (\[3\])).
Moreover, if $k_3$ is the number of tropical regular algebraic rational trinomials, and $k_2$ is the number of tropical non-monotone algebraic rational binomials in (\[4\]), so $k_3+k_2=k$ then $2k_3+k_2$ is the number of blocks of edges of $f$ of the equal (non-zero) signs of their slopes. The number $2k_3+k_2+k_1$ does not exceed the number of edges of $f$ with non-zero slopes, finally $k_0$ equals the number of edges with zero slopes.
The number of tropical roots of $f$ is greater or equal to $k_1+2k_0$, and on the other hand, is less or equal to $2k_3+k_2+k_1+k_0$, the latter number also equals the total number of tropical roots in the composants of $f$ from (\[4\]) (cf. Remark \[roots-rational\]).
Tropical rational functions {#three}
===========================
In this section we study decompositions of tropical rational functions, we recall that the slopes of edges of a piece-wise rational function $f$ are integers (unlike the section \[two\] in which the slopes could be rationals).
\[tropical-rational\] (i) There is an algorithm which for a tropical monotone rational function $f$ yields its decomposition into tropical monotone binomials and tropical monotone trinomials. The number of composants does not exceed the number of tropical roots of $f$ (cf. Proposition \[monotone-algebraic\] and Remark \[monotone-algebraic-number\]);
\(ii) there is an algorithm which decomposes a tropical rational function $$f=g_1\circ \cdots \circ g_k\circ h_1\circ \cdots \circ h_m\circ g^{(0)}_{k_0}\circ \cdots \circ g^{(0)}_1$$ (cf. (\[4\])) where each $g_i,\, 1\le i\le k$ is either a tropical non-monotone rational binomial with $\pm 1$ slopes or a tropical non-monotone rational trinomial with $\pm 1$ slopes (cf. (\[1\])), each $h_j,\, 1\le j\le m$ is either a tropical regular monotone binomial or a tropical regular monotone trinomial, and each $g^{(0)}_l,\, 1\le l\le k_0$ is a tropical singular monotone trinomial. The number of binomials among $g_1,\dots,g_k$ plus the double number of trinomials among $g_1,\dots,g_k$ does not exceed the number of blocks of edges of $f$ (cf. Theorem \[algebraic-rational\]). The number $m$ does not exceed the number of edges of $f$, and the number $k_0$ equals the number of edges of $f$ with zero slopes (again cf. Theorem \[algebraic-rational\] and (\[3\]));
\(iii) let $f$ be a tropical monotone rational function (respectively, a tropical polynomial) with the slopes of its edges $a_0,\dots,a_n\ge 1$ (respectively, $a_0>\dots >a_n\ge 1$) and denote by $q_i,\, 1\le i\le n$ the denominator of the irreducible fraction $a_i/a_{i-1}$. Then $f$ is a composition of tropical rational binomials (respectively, tropical binomials) iff $(q_1\cdots q_n)|a_0$.
If $f$ satisfies the latter condition in (iii) we call $f$ [*completely decomposable*]{}.
This condition in (iii) is equivalent to a more symmetric one: for any $m\ge 1$ and $j\ge 0$ such that $m+2j<n$ it holds $$\prod_{0\le i\le j} a_{m+2i} \quad | \prod_{0\le i\le j+1} a_{m+2i-1}.$$ In particular, for $n=2$ (trinomials), the condition in (iii) for $a_0,\, a_1,\, a_2$ is equivalent to $a_1|(a_0a_2)$.
[**Proof**]{}. (i). If an increasing $f$ has at least 4 edges then take any its tropical root $x_0$ being neither the left-most nor the right-most. Define $h$ to coincide with $f$ on the interval $(-\infty,\, x_0]$ and $g$ to coincide with the identity function on the interval $(-\infty,\, f(x_0)]$. Then define $h$ on the interval $[x_0,\, \infty)$ to coincide with the linear function $x+f(x_0)-x_0$, and define $g$ on the interval $[f(x_0),\, \infty)$ to coincide with the composition $f\circ (x-f(x_0)+x_0)$. Then $f=g\circ h$.
Continuing in this way, applying further the described construction to $g,\, h$ we complete the proof of (i).
(ii). First, similar to the proof of Theorem \[algebraic-rational\] one represents (by means of straightening) $f=g\circ h$ (assume w.l.o.g. that the first edge of $f$ with non-zero slope has a positive slope), where $g$ is either a tropical non-monotone binomial with the slopes of its edges $1$ and $-1$ or a tropical trinomial with the slopes $1,\, -1,\, 1$, while $h$ being a tropical rational function with less number of blocks of edges than $f$.
Continuing in this way, while it is possible, we arrive to a tropical non-decreasing rational function $f^{(0)}$ such that $f=g_1\circ \cdots \circ g_k \circ f^{(0)}$ (cf. (\[1\])). Applying to $f^{(0)}$ the constructions from the proof of Theorem \[algebraic-rational\] (cf. (\[3\])) and from the proof of Theorem \[tropical-rational\] (i), we complete the proof of (ii).
(iii). The proofs for both cases $f$ being a tropical increasing rational function or a tropical polynomial go similarly.
Let $f=g_1\circ \cdots \circ g_k$ where each $g_i,\, 1\le i \le k$ is a tropical increasing rational binomial (respectively, a tropical binomial) with slopes $b_i,\, c_i,\, 1\le i\le k$ (respectively, $b_i>c_i$). Denote by $r_i,\, 1\le i\le k$ the unique tropical root of $g_i$. Partition $\RR$ into intervals with the end-points $(g_{i+1}\circ \cdots \circ g_k)^{-1}(r_i),\, 1\le i\le k$. In case of tropical polynomials $f$ all these end-points are the tropical roots of $f$. In case of tropical increasing rational functions $f$ all $g_i$ for which $(g_{i+1}\circ \cdots \circ g_k)^{-1}(r_i)$ being not a tropical root of $f$, give a contribution into $g_1\circ \cdots \circ g_k$ by multiplying all the slopes of its edges on the intervals by the same integer, so w.l.o.g. one can assume that each $(g_{i+1}\circ \cdots \circ g_k)^{-1}(r_i),\, 1\le i\le k$ is a tropical root of $f$.
For $1\le j\le n$ take the set $I_j$ of $1\le i\le k$ such that $(g_{i+1}\circ \cdots \circ g_k)^{-1}(r_i)$ is $j$-th root $t_j$ of $f$. Then $a_j/a_{j-1}=\prod _{i\in I_j} (c_i/b_i)$. Therefore, $q_j| \prod _{i\in I_j} b_i$. Since $\prod _i b_i =a_0$, we conclude that $(q_1\cdots q_n)|a_0$.
Conversely, let $(q_1\cdots q_n)|a_0$. Put integers $b_j:=q_j,\, 1\le j\le n-1,\, b_n:=a_0/(b_1\cdots b_{n-1})$ and $c_j:=b_ja_j/a_{j-1},\, 1\le j\le n$.
Construct $g_n,\dots,g_1$ recursively. As a base of recursion take $g_n$ such that its unique tropical root coincides with $t_n$ (observe that $g_n$ is defined uniquely up to an additive shift, in other words, one can replace $g_n$ by $g_n+e,\, e\in \RR$). Assume that $g_n,\dots,g_{m+1}$ are already constructed by recursion. Then take $g_m$ such that its unique tropical root equals $(g_{m+1}\circ \cdots \circ g_n)(t_m)$. At the very last step of recursion we adjust $g_1$ by a suitable additive shift to make $g_1\circ \cdots \circ g_n$ coincide with $f$ at one (arbitrary) point. Hence $f= g_1\circ \cdots \circ g_n$. $\Box$
The algorithms designed in sections \[two\], \[three\] have polynomial complexity since after each procedure yielding a composant (cf. (\[1\]), (\[2\]), (\[3\])) either the number of blocks of edges or the number of edges drops at least by one.
Tropical polynomials with commuting composition {#four}
===============================================
Let $f,\, g$ be tropical polynomials without free terms (some statements below hold also for more general tropical increasing algebraic rational functions). In this section we give a criterion when $f\circ g=g\circ f$. Note that the inverse $f^{-1}$ (i. e. $f\circ f^{-1}=Id$ equals the identity function) is a tropical increasing algebraic rational function. Denote by $f^k:= f\circ \cdots \circ f$ the $k$ times iteration of $f$. We agree that $f^0:=Id$. Note that tropical increasing algebraic rational functions constitute a group with respect to the composition.
\[fixed-point\] Let $f,\, g$ be tropical increasing algebraic rational functions and $f\circ g=g\circ f$ hold. We call $x$ a [*fixed point*]{} of $f$ if $f(x)=x$. The set $F_f\subset \RR$ of fixed points is a finite union of disjoint closed intervals $\{[x_i,\, y_i]\}_i$ (including isolated points, i. e. $x_i=y_i$). Since $f\circ g(x)=g\circ f(x)=g(x)$ we conclude that $g(F_f)=F_f$, therefore $g(x_i)=x_i,\, g(y_i)=y_i$ for all $i$ since $g$ is increasing.
Observe that either $g(x)=x$ for any point $y_i<x<x_{i+1}$, either $g(x)<x$ for any point $y_i<x<x_{i+1}$ or $g(x)>x$ for any point $y_i<x<x_{i+1}$. Indeed, otherwise consider the set of fixed points $F_g\cap [y_i,\, x_{i+1}]$, and arguing as above in the previous paragraph we get $f(F_g\cap [y_i,\, x_{i+1}])=F_g\cap [y_i,\, x_{i+1}]$, again $f(x)=x$ for any end-point of an interval of $F_g\cap [y_i,\, x_{i+1}]$, which contradicts the choice of $y_i,\, x_{i+1}$, unless $F_g\supset (y_i,\, x_{i+1})$, in other words $g(x)=x$ for any point $y_i<x<x_{i+1}$. We allow intervals with $\pm \infty$ end-points.
In the case of tropical polynomials $f,\, g$ without free terms there is at most one end-point of the intervals of fixed points of $f,\, g$, which we denote by $x_0$, due to the convexity of $f,\, g$ and taking into the account that the slopes of edges of $f,\, g$ are greater or equal than 1. Thus, there are at most two intervals $(-\infty,\, x_0],\, [x_0,\, \infty)$ or just one interval $(-\infty,\, \infty)$ when $x_0=\infty$.
\[commuting\] Tropical polynomials $f,\, g$ commute: $f\circ g=g\circ f$ iff either $x_0=\infty$ and $f=x+c_1,\, g=x+c_2, x\in \RR$ for some $c_1,\, c_2\in \RR$ or $-\infty<x_0<\infty$ and the following is valid.
There exists a tropical increasing algebraic rational function $h$ such that $h(x_0)=x_0$ (see Remark \[fixed-point\]) and
$\bullet$ either $f=h^{p},\, g=h^{q}$ for suitable non-negative integers $p,\, q$
$\bullet$ or $f=ax+x_0(1-a),\, g=bx+x_0(1-b)$ for suitable integers $a,\, b\ge 1$
holds on the interval $[x_0,\, \infty)$. Similarly,
$\bullet$ either $f=h^{k},\, g=h^{m}$ for suitable non-negative integers $k,\, m$
$\bullet$ or $f=dx+x_0(1-d),\, g=ex+x_0(1-e)$ for suitable integers $d,\, e\ge 1$
holds on the interval $(-\infty,\, x_{0}]$,
Note that $h$ is not necessary a tropical polynomial.
[**Proof**]{}. In one direction, namely when either such appropriate $h$ does exist or $c_1,\, c_2$ do exist, obviously $f\circ g=g\circ f$ holds.
From now on let $f\circ g=g\circ f$. If $f(x)=x$ (respectively, $g(x)=x$) for any $x\ge x_0$ one can put $h:=g,\, f=h^0$ (respectively, $h:= f,\, g=h^0$) on the interval $(x_0,\, \infty)$. Thus, from now on we suppose that $f(x)>x,\, g(x)>x$ for any $x>x_0$ (cf. Remark \[fixed-point\]). We construct (increasing) $h$ on the interval $(-\infty,\, x_0)$ and separately on the interval $(x_0,\, \infty)$ such that $h(x_0)=x_0$ and after that glue them together and obtain a tropical increasing algebraic rational function $h$ required in Theorem \[commuting\].
\[value\] Let $f,\, g$ be tropical increasing algebraic rational functions, $f\circ g=g\circ f$ and for some point $y_i<x<x_{i+1}$ it holds $f(x)=g(x)$. Then $f$ coincides with $g$ on the interval $(y_i,\, x_{i+1})$.
[**Proof of Lemma \[value\]**]{}. Since neither $f$ nor $g$ has a fixed point in the interval $(y_i,\, x_{i+1})$ one can assume for definiteness that $f(y)>y,\, g(y)>y$ for any $y_i<y<x_{i+1}$ (see Remark \[fixed-point\]). For each integer $k$ we have $f(f^k(x))=g(f^k(x))$. The increasing sequence $x<f(x)<f^2(x)<\cdots$ tends to $x_{i+1}$ taking into the account that $F_f\cap (y_i\, x_{i+1})=\emptyset$. Therefore, the right-most edges of $f$ and $g$ on the interval $(y_i,\, x_{i+1})$ coincide. Suppose that $f$ and $g$ do not coincide on the interval $(y_i,\, x_{i+1})$.
Take the left-most point $z_0\in (y_i,\, x_{i+1})$ such that $f(y)=g(y)$ for any $z_0\le y<x_{i+1}$. Hence on a sufficiently small interval $[z,\, z_0]$ function $f$ (respectively, $g$) is linear $ax-az_0+f(z_0)$ (respectively, $bx-bz_0+f(z_0)$) and $a\neq b$. Since $x_{i+1}>f(z_0)=g(z_0)>z_0$ and due to the choice of $z_0$ there exists a linear function $cx+d$ such that on a sufficiently small interval $[z,\, z_0]$ the composition $f\circ g$ coincides with the linear function $cbx-cbz_0+cf(z_0)+d$, while the composition $g\circ f$ on $[z,z_0]$ coincides with the linear function $cax-caz_0+cf(z_0)+d$, which contradicts to the commutativity $f\circ g=g\circ f$ and proves Lemma \[value\]. $\Box$
Fix an interval $(y_i,\, x_{i+1})$ for the time being and denote by $T_f$ the set of tropical roots of $f$. We considered the case when $f(x)=x$ for any $x\in (y_i,\, x_{i+1})$ or when $g(x)=x$ for any $x\in (y_i,\, x_{i+1})$ above, so we assume that either $f(x)>x$ for any $x\in (y_i,\, x_{i+1})$ or $f(x)<x$ for any $x\in (y_i,\, x_{i+1})$, and either $g(x)>x$ for any $x\in (y_i,\, x_{i+1})$ or $g(x)<x$ for any $x\in (y_i,\, x_{i+1})$ (cf. Remark \[fixed-point\]). First, we study the case $(T_f\cup T_g)\cap (y_i,\, x_{i+1}) =\emptyset$. Since $g(y_i)=f(y_i)=y_i,\, g(x_{i+1})=f(x_{i+1})=x_{i+1}$ we conclude that one or both end-points of the interval $(y_i,\, x_{x+1})$ equal $\pm \infty$.
When both $y_i=-\infty,\, x_{i+1}=\infty$, we have $f=x+c_1,\, g=x+c_2$ for some $c_1,\, c_2 \in \RR$.
If $y_i\in\RR,\, x_{i+1}=\infty$ (the case $y_i=-\infty,\, x_{i+1}\in \RR$ is analyzed in a similar way) then $f$ (respectively, $g$) coincides on the interval $(y_i,\, \infty)$ with a linear function $ax-ay_i+y_i$ (respectively, $bx-by_i+y_i$) for suitable rationals $a,\, b>0$ which establishes Theorem \[commuting\] in the case $(T_f\cup T_g)\cap (y_i,\, x_{i+1}) =\emptyset$.
From now on we again assume $f,\, g$ to be tropical polynomials and let $(T_f\cup T_g) \cap (-\infty,\, x_0) \neq \emptyset$ (cf. Remark \[fixed-point\]).
\[group\] (i) If $h_1,\, h_2$ are tropical algebraic rational functions and $x\in T_{h_1\circ h_2}$ then either $x\in T_{h_2}$ or $h_2(x)\in T_{h_1}$. For tropical polynomials $f,\, g$ the converse is true: if either $x\in T_g$ or $g(x)\in T_f$ then $x\in T_{f\circ g}$;
\(ii) let $f\circ g=g\circ f$. If $x\in T_g$ then either $f^{-1}(x)\in T_g$ or $g\circ f^{-1} (x)\in T_f$;
\(iii) let $f\circ g=g\circ f$. If $x\in T_f\setminus T_g$ then $g(x)\in T_f$.
[**Proof of Lemma \[group\]**]{}. (i). For the converse statement the convexity of $f,\, g$ is used.
(ii). Denote $y:= f^{-1}(x)$. Due to (i) $y\in T_{g\circ f}$, hence either $y\in T_g$ or $g(y)\in T_f$ again due to (i) and taking into the account that $f\circ g=g\circ f$;
\(iii) Since due to (i) $x\in T_{g\circ f} =T_{f\circ g}$ we conclude that $g(x)\in T_f$ again by means of (i). $\Box$
Consider a directed graph $G$ with the nodes being the points from $(T_f\cup T_g) \cap (-\infty,\, x_0)$ and the arrows according to Lemma \[group\] as follows (recall that $G$ is not empty, the case of empty $G$ was studied above). From every node $x\in T_g \cap (-\infty,\, x_0)$ there is an arrow labeled by $f^{-1}$ to the node $f^{-1}(x)$, provided that $f^{-1}(x) \in T_g$, and there is an arrow labeled by $g\circ f^{-1}$ to the node $g\circ f^{-1}(x)$, provided that $g\circ f^{-1}(x) \in T_f$ (observe that $f^{-1}(x),\, g\circ f^{-1}(x) \in (-\infty,\, x_0)$). In addition, there is an arrow labeled by $g$ from every node $x\in T_f\setminus T_g$ to the node $g(x)\in T_f$ (again $g(x) \in (-\infty,\, x_0)$).
There is a cycle in $G$ (due to Lemma \[group\]), let it contain a node $x$. Denote by $t$ the composition of the labels of the arrows (starting with $x$) in this cycle. Then $t(x)=x$ and one can represent $t=g^s\circ f^{-r}$ (taking into the account that $f\circ g=g\circ f$) for some non-negative integers $s,\, r$ at least one of which being positive. Observe that in fact, $s,\, r>0$ since $f(x_1)<x_1,\, g(x_1)<x_1$ for any $x_1<x_0$ (cf. Remark \[fixed-point\]).
Hence $g^s(x)=f^r(x)$. Lemma \[value\] implies that $g^s$ coincides with $f^r$ on the interval $(-\infty,\, x_0)$. Denote $n:=GCD(s,\, r)$, then $(g^{s/n}\circ f^{-r/n})^n=Id$ on the interval $(-\infty,\, x_0)$. The function $u:= g^{s/n}\circ f^{-r/n}$ is increasing piece-wise linear. Therefore, one can partition $\RR$ into a finite number of intervals (including unbounded ones) such that on each of these intervals $[y_0,\, y_1]$ it holds $u(y_0)=y_0,\, u(y_1)=y_1$, and either $u(y)>y$ for any $y_0<y<y_1$, either $u(y)<y$ for any $y_0<y<y_1$ or $u(y)=y$ for any $y_0<y<y_1$ (cf. Remark \[fixed-point\]). Hence $u=Id$, i. e. $g^{s/n}=f^{r/n}$ on the interval $(-\infty,\, x_0)$.
For appropriate positive integers $i,\, j$ it holds $1=-i(s/n)+j(r/n)$. Consider a tropical increasing algebraic rational function $h:=g^j\circ f^{-i}$. Then
$h^{r/n}=g^{jr/n}\circ f^{-ir/n}=g^{jr/n}\circ g^{-is/n}=g$;
$h^{s/n}=g^{js/n}\circ f^{-is/n}=f^{jr/n}\circ f^{-is/n}=f$
on the interval $(-\infty,\, x_0)$.
In a similar way one produces $h$ on the interval $(x_0,\, \infty)$, provided that $x_0<\infty$. This completes the proof of Theorem \[commuting\]. $\Box$
It would be interesting to give a criterion for commuting tropical increasing algebraic rational functions, and more generally, for tropical non-monotone algebraic rational functions.
We exhibit an example of a commuting pair of increasing tropical rational functions $f,\, g$ (defined on the interval $[x_0,\, \infty),\, f(0)=g(0)=0$, thereby $x_0=0$, see Theorem \[commuting\]) not satisfying the conclusion of Theorem \[commuting\].
Pick a real $0<t$, integers $a>\alpha>1,\, b\ge 1$ such that $a\neq b,\, a|(b\alpha)$. denote $u:=\alpha t,\, v:=at,\, w:=\alpha a t$. Define $g$ to be piece-wise linear whose graph on $[x_0,\, \infty)$ consists of three edges having the slopes $a,\, b,\, a$, respectively, and with the tropical roots at the points $u,\, w$ (this defines $g$ uniquely). Similarly, define $f$ whose graph also has three edges with the slopes $\alpha,\, (b\alpha)/a,\, \alpha$, respectively, and with the tropical roots $v,\, w$. Then the graph on $[x_0,\, \infty)$ of the increasing tropical rational function $f\circ g=g\circ f$ has three edges as well with the slopes $\alpha a,\, \alpha b,\, \alpha a$, respectively, and with tropical roots $t,\, w$.
In case when $f^k=g^m$ (cf. Theorem \[commuting\]) we get that $\alpha^k=a^m$. Thus, if there are no such integers $k,\, m$ the conclusion of Theorem \[commuting\] for $f,\, g$ is not fulfilled.
Tropical polynomial and rational parametrizations {#five}
=================================================
We call a [*polygonal line*]{} $L\subset \RR^n$ with $k+1$ intervals a sequence of intervals with endpoints $v_1,\dots,v_k \in \QQ^n$ such that $i$-th interval has endpoints $v_i,\, v_{i+1}$ for $1\le i\le k-1$, while unbounded $0$-th interval (a ray) has $v_1$ as its right endpoint, and unbounded $k$-th interval (a ray) has $v_k$ as its left endpoint. The [*vector of slopes*]{} of $i$-th interval, $1\le i\le k-1$ is defined as $(a_{i,1},\dots,a_{i,n}):=v_{i+1}-v_i$, similarly one can define a vector of slopes $(a_{0,1},\dots,a_{0,n})$ of $0$-th and $(a_{k,1},\dots,a_{k,n})$ of $k$-th intervals, respectively (we assume that the latter two vectors of slopes are also rational). Note that the vector of slopes is determined up to a positive factor.
We say that tropical rational functions $f_1,\dots,f_n$ in one variable $t$ provide a [*tropical rational parametrisation*]{} of $L$ if the map $(f_1,\dots,f_n):\RR \to L$ is a bijection and (for definiteness) $(f_1,\dots,f_n)^{-1}(v_i)< (f_1,\dots,f_n)^{-1}(v_{i+1}),\, 1\le i\le k-1$. In particular, $\{v_1,\dots,v_k\}$ coincides with the set of all the tropical roots of $f_1,\dots,f_n$, i. e. the points where one of the functions $f_1,\dots,f_n$ is not smooth. We suppose w.l.o.g. that one can’t discard any $v_i,\, 1\le i\le k$ while keeping the propery of $L$ to be a polygonal line. When $f_1,\dots,f_n$ are tropical polynomials (respectively, Laurent polynomials), we talk about [*tropical polynomial*]{} (respectively, [*Laurent polynomial) parametrisation*]{} of $L$.
In case if $L$ is a subset of a tropical curve one can treat a parametrisation of $L$ as a parametrisation of the tropical curve (cf. [@G]) since a parametrisation provides a parametric family of solutions of a system of tropical equations.
Let $T\subset \RR^2$ be a tropical curve (a tropical line) defined by a tropical polynomial $\min\{x,\, y,\, 0\}$. Then $L\subset T$ consisting of two rays $\{x=0\le y\} \cup \{y=0\le x\}$ admits a tropical rational parametrisation with $f_1:=-\min\{t,\, 0\},\, f_2:=-\min\{-t,\, 0\}$.
\[parametrisation\] A polygonal line $L$ has
\(i) always a tropical rational parametrisation;
\(ii) a tropical polynomial parametrisation iff $a_{i,j}\ge 0,\, 0\le i\le k,\, 1\le j\le n$, and $a_{i,j}=0$ implies $a_{l,j}=0$ for all $l\ge i$;
\(iii) a tropical Laurent polynomial parametrisation iff
$\bullet$ $a_{i,j}<0$ implies $a_{i+1,j}<0$;
$\bullet$ $a_{i+1,j}>0$ implies $a_{i,j}>0$;
$\bullet$ $a_{i,j_0}>0,\, a_{i+1,j_0}>0,\, a_{i,j}<0,\, a_{i+1,j}<0$ imply $a_{i,j_0}/a_{i,j}\le a_{i+1,j_0}/a_{i+1,j}$
for all $0\le i\le k-1,\, 1\le j\neq j_0 \le n$.
[**Proof**]{}. (i) We have to construct tropical univariate rational functions $f_1,\dots,f_n$. First we construct piece-wise linear functions $g_1,\dots,g_n$ with rational slopes (in [@G] such functions are called tropical Newton-Puiseux rational functions). As a set of tropical roots of $g_1,\dots,g_n$ we take points $1,\dots,k$. The vector of the values of $g_1,\dots,g_n$ at point $i$ we put $v_i,\, 1\le i\le k$. Thereby, $g_1,\dots,g_n$ are defined on interval $[1,\, k]\subset \RR$. To extend $g_1,\dots,g_n$ to interval $(-\infty,\, 1]$ (respectively, $[k,\, \infty)$) use the vector of the slopes of $0$-th (respectively, $k$-th) interval of $L$.
To proceed to tropical rational functions $f_1,\dots,f_n$ (so, piece-wise linear functions with integer slopes), denote by $M$ the least common multiple of all the denominators of the slopes of $g_1,\dots,g_n$ (i. e. the slopes of $L$). As the set of tropical roots of $f_1,\dots,f_n$ take points $1/M,\dots,k/M$. The vector of the values at point $i/M$ we put $v_i,\, 1\le i\le k$. In other words, the corresponding slopes of $f_1,\dots,f_n$ are obtained from the corresponding slopes of $g_1,\dots,g_n$ multiplying by $M$. Satisfying also the latter condition, one extends $f_1,\dots,f_n$ to intervals $(-\infty,\, 1/M]$ and $[k/M,\, \infty)$.
\(ii) If $f_1,\dots,f_n$ constitute a tropical polynomial parametrization of $L$ then since the slopes of each $f_j,\, 1\le j\le n$ (being a convex function) are non-increasing non-negative integers we get the conditions stated in (ii).
Conversely, if the latter conditions are fulfilled one can recursively on $i$ choose positive rationals $c_0=1,\, c_1,\dots,c_k$ in such a way that $c_{i+1}\cdot a_{i+1,j}\le c_i \cdot a_{i,j},\, 0\le i\le k-1,\, 1\le j\le n$ taking each $c_{i+1}$ to be the maximal possible among satisfying the latter inequalities. Therefore, one can take $c_i\cdot a_{i,j},\, 1\le j\le n$ as the slopes of (Newton-Puiseux polynomials [@G], i. e. convex piece-wise linear functions with rational non-negative slopes) $g_j,\, 1\le j\le n$ with the tropical roots at points $1,\dots,k$. Then as at the end of the proof of (i) one can obtain tropical polynomials $f_j,\, 1\le j\le n$ with the non-negative integer slopes $M\cdot c_i\cdot a_{i,j},\, 0\le i\le k,\, 1\le j\le n$ and with the tropical roots at points $1/M,\dots,k/M$. Then $f_1,\dots,f_n$ provide a required parametrization of $L$.
\(iii) If there exists a tropical Laurent polynomial parametrization $f_1,\dots,f_n$ of $L$ then the slopes $b_{i,j},\, 0\le i\le k,\, 1\le j\le n$ of $f_j,\, 1\le j\le n$, respectively, being integers fulfil the conditions $b_{i,j}\ge b_{i+1,j},\, 0\le i\le k-1,\, 1\le j\le n$. On the other hand, there exist positive rationals $c_0,\dots,c_k$ such that $b_{i,j}=c_i\cdot a_{i,j},\, 0\le i\le k,\, 1\le j\le n$. This entails the conditions from (iii).
Conversely, let the conditions from (iii) be fulfilled. Construct positive rationals $c_0=1,\, c_1,\dots, c_k$ such that $c_i\cdot a_{i,j}\ge c_{i+1}\cdot a_{i+1.j},\, 0\le i\le k-1,\, 1\le j\le n$ by recursion on $i$. Assume that $c_0=1,\, c_1,\dots,c_i$ are already constructed. Take the maximal possible $c_{i+1}>0$ such that $c_i\cdot a_{i,j}\ge c_{i+1}\cdot a_{i+1.j}$ for all $1\le j\le n$ such that $a_{i,j}>0,\, a_{i+1,j}>0$. Then for suitable $j_0$ for which $a_{i,j_0}>0,\, a_{i+1,j_0}>0$ it holds $c_i\cdot a_{i,j_0}=c_{i+1}\cdot a_{i+1,j_0}$. For every $1\le j\le n$ for which $a_{i,j}<0,\, a_{i+1,j}<0$ the condition from (iii) $a_{i,j_0}/a_{i,j}\le a_{i+1,j_0}/a_{i+1,j}$ implies $c_{i+1}\cdot a_{i+1,j} \le c_i\cdot a_{i,j}$.
Thus, as in (i), (ii) one first constructs piece-wise linear functions $g_j,\, 1\le j\le n$ with rational non-increasing slopes $c_i\cdot a_{i,j},\, 1\le j\le n$ and with the tropical roots at points $1,\dots,k$. Denote by $M$ the common denominator of these slopes and construct tropical Laurent polynomials $f_1,\dots,f_n$ with the slopes obtained from the slopes of $g_j,\, 1\le j\le n$ multiplying them by $M$ and with the tropical roots $1/M,\dots,k/M$. Then $f_1,\dots,f_n$ provide a required parametrization of $L$. $\Box$
One can construct the required parametrizations in Proposition \[parametrisation\] within polynomial complexity following the proofs of (i), (ii), (iii).
It would be interesting to extend parametrizations from 1-dimensional polygonal lines to multidimensional polyhedral complexes.
[**Acknowledgements**]{}. The author is grateful to the grant RSF 16-11-10075 and to MCCME for inspiring atmosphere.
[99]{} J. von zur Gathen. Functional decomposition of polynomials: the wild case. , 10:437–452, 1990.
J. von zur Gathen, J. Gutierrez and R. Rubio. Multivariate polynomial decomposition. , 14:11–31, 2003.
D. Grigoriev. Tropical Newton-Puiseux polynomials. , 11077:177–186, 2018.
D. Grigoriev and V. Podolskii. Tropical combinatorial Nullstellensatz and fewnomials testing. , 10472:284–297, 2017.
D. Kozen, S. Landau and R. Zippel. Decomposition of algebraic functions. , 22:235–246, 1996.
D. Maclagan and B. Sturmfels. , volume 161 of [*Graduate Studies in Mathematics*]{}. American Mathematical Society, 2015.
G. F. Montúfar, R. Pascanu, K. Cho and Y. Bengio. On the number of linear regions of deep neural networks. , Montreal, 2924–2932, 2014.
F. Pakovich. Semiconjugate Rational Functions: A Dynamical Approach. , 4:59–68, 2018.
J. Ritt. Prime and composite polynomials. , 23:51–66, 1922.
J. Ritt. Permutable rational functions. , 25:399–448, 1923.
| {
"pile_set_name": "ArXiv"
} |
---
---
[A dissertation submitted to the Jagiellonian University in Kraków for the degree of Doctor of Philosophy]{}
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'Ten very low luminosity objects were observed multiple times in the 8.5 GHz continuum in search of protostellar magnetic activities. A radio outburst of IRAM 04191+1522 IRS was detected, and the variability timescale was about 20 days or shorter. The results of this survey and archival observations suggest that IRAM 04191+1522 IRS is in active states about half the time. Archival data show that L1014 IRS and L1148 IRS were detectable previously and suggest that at least 20%–30% of very low luminosity protostars are radio variables. Considering the variability timescale and flux level of IRAM 04191+1522 IRS and the previous detection of the circular polarization of L1014 IRS, the radio outbursts of these protostars are probably caused by magnetic flares. However, IRAM 04191+1522 IRS is too young and small to develop an internal convective dynamo. If the detected radio emission is indeed coming from magnetic flares, the discovery implies that the flares may be caused by the fossil magnetic fields of interstellar origin.'
author:
- 'Minho Choi$^{1}$, Jeong-Eun Lee$^2$, and Miju Kang$^{1,3}$'
title: Radio Variability Survey of Very Low Luminosity Protostars
---
INTRODUCTION
============
Magnetosphere is an integral part of a star and plays important roles in the dynamics of ionized gas in and around the star (Donati & Landstreet 2009; Mestel 2012). The magnetosphere manifests itself through powerful outbursts of energy such as radio and X-ray flares, and such magnetic activities start well before the onset of stable hydrogen burning (Feigelson & Montmerle 1999). (In this paper, we use the term “outburst” to mean a sudden increase of brightness and the term “flare” to mean an outburst event caused by magnetic reconnection. This flare is not necessarily a solar-type flare.) Many protostars drive well-collimated outflows (Arce et al. 2007), and the outflows are generated through magnetocentrifugal ejection mechanism (Pudritz et al. 2007), which suggests that protostars may possess well-organized magnetic fields. The magnetic fields of young stellar objects (YSOs) may be seeded by the interstellar magnetic fields in the natal cloud (Mestel 1965; Tayler 1987). However, it is difficult to study the origin of stellar magnetic fields and to obtain observational evidence for the magnetic fields of young protostars, because Class 0 protostars, objects in the earliest stage of stellar evolution (Andr[é]{} et al. 1993), are surrounded by opaque layers of gas that block the radio and X-ray emission from the view of outside observers (Feigelson & Montmerle 1999). Especially, the partially ionized winds of typical protostars are opaque to radio waves (Andr[é]{} 1996).
The free-free optical depth of a protostellar wind can be roughly estimated from the mass accretion rate. For a typical protostar, the mass accretion rate is $\sim$2 $\times$ 10$^{-6}$ $M_\odot$ yr$^{-1}$ (Shu et al. 1987). Assuming an accretion-to-outflow conversion factor of 10% and an ionization fraction of 1%, the mass-loss rate of ionized wind is $\sim$2 $\times$ 10$^{-9}$ $M_\odot$ yr$^{-1}$. Using Equation (5) of Andr[é]{} et al. (1988), and assuming a wind speed of 200 km s$^{-1}$, an emission-region radius of 3–9 $R_\odot$ (1–3 times the protostellar radius), and a wind temperature of 10$^4$ K, the optical depth would be 400–10,000 at 8.5 GHz. Therefore, magnetic flares of typical protostars are unobservable in the radio wavelength band.
Recently discovered low-luminosity protostars present a new opportunity to directly observe the protostellar atmospheres. These very low luminosity objects (VeLLOs) were revealed by observations in infrared wavelengths and have internal luminosities less than 0.1 $L_\odot$ (Young et al. 2004; Dunham et al. 2006). VeLLOs provide interesting insights into the physics and chemistry of low-mass star formation (Lee 2007; Dunham et al. 2008). They may have smaller mass accretion rates, hence smaller mass-loss rates, than typical protostars.
The internal luminosities of VeLLOs are in the range of 0.02–0.1 $L_\odot$ (Dunham et al. 2008). For a protostar of $\sim$0.075 $M_\odot$ (the stellar–substellar boundary) with a radius of 3 $R_\odot$, the mass accretion rate would be 0.3–1.3 $\times$ 10$^{-7}$ $M_\odot$ yr$^{-1}$. Using the assumptions above, the free–free optical depth of a VeLLO wind would be 0.08–40. Therefore, the magnetic flares of VeLLOs are detectable, especially when observed during powerful and large-size flare events.
Much of the knowledge about the magnetic fields of YSOs comes from the radio and X-ray observations of more evolved objects such as T Tauri stars, Class I protostars, and the late-stage protostar R CrA IRS 5 (Andr[é]{} 1996; G[ü]{}del 2002; Forbrich et al. 2006; Choi et al. 2008, 2009; Hamaguchi et al. 2008). These objects showed that one of the key characteristics is the flux variability of emission from nonthermal electrons. The variability timescale of magnetic flares of YSOs is usually hours to days.
In this paper, we present the results of our monitoring survey of VeLLOs using the Very Large Array (VLA) of the National Radio Astronomy Observatory (NRAO). We describe our observations and archival data in Section 2. In Section 3, we report the results. In Section 4, we discuss the magnetic activities of protostars.
OBSERVATIONS AND DATA
=====================
Source Selection
----------------
We observed the VeLLOs in the list compiled by Dunham et al. (2008). Out of the 15 VeLLOs in the list, 10 objects of high declination were selected. These sources stay over the observable elevation limit for relatively long durations and are suitable for a survey with VLA. The selected sources are listed in Table 1. The target sources are protostellar objects in nearby star-forming molecular clouds. The sources are divided into two groups depending on the right ascension so that all the sources in a group can be observed in a single observing run.
VLA Observations
----------------
The survey targets were observed using VLA in 2009 in the standard $X$-band continuum mode (8.5 GHz or $\lambda$ = 3.5 cm). The total bandwidth for the continuum observations was 172 MHz (43 MHz for each of the four intermediate-frequency channels). The observations were carried out in the D-array configuration toward the infrared source positions determined using the [*Spitzer Space Telescope*]{} (Dunham et al. 2008). Each target source was observed three times with an interval of $\sim$20 days (Group A sources on October 18, November 8, and November 30, and Group B sources on October 31, November 8, and November 30) so that the flux variability can be determined if the source went through a radio outburst. In each observing run, the on-source integration time for each target source was $\sim$12 minutes.
The phase was determined by observing nearby quasars. The flux was calibrated by setting the flux densities of the quasars 0542+498 (3C 147) and 1331+305 (3C 286) to 4.7 and 5.2 Jy, respectively. Maps were made using a CLEAN algorithm. With a natural weighting, the 8.5 GHz continuum data produced synthesized beams of $\sim$9$''$ in FWHM.
VLA Archival Data
-----------------
In addition to the survey data, several VLA data sets retrieved from the NRAO Data Archive System were also analyzed. We searched for 8.5 GHz continuum observations of the target regions and found 10 data sets. All of them have a continuum bandwidth of 172 MHz. These data sets were analyzed following the standard VLA data reduction procedure. The archival data by themselves are not adequate for the confirmation of magnetic flares because their time resolution is worse than a year.
RESULTS
=======
In our survey, only DCE 1 was detected. For each of the undetected sources, data from all the three observing runs were combined to reduce the noise, but no additional detection was made. In the archival data sets, DCE 1, 32, and 38 were detected, indicating that they are variable radio sources. The results are summarized in Tables 2 and 3.
IRAM 04191+1522 IRS (DCE 1)
---------------------------
DCE 1 corresponds to IRAM 04191+1522 IRS, a Class 0 protostar driving a highly-collimated bipolar outflow (Andr[é]{} et al. 1999; Dunham et al. 2006). IRAM 04191+1522 IRS was detected clearly in the first observing run (Figure 1). The position of the 8.5 GHz continuum source was $\alpha_{2000}$ = 04$^{\rm h}$21$^{\rm m}$56$\fs$79 and $\delta_{2000}$ = 15$^\circ$29$'$45$\farcs$5, which agrees with that of the infrared source within 1$\farcs$4. Considering the 9$''$ beam size, the radio source is positionally coincident with the infrared source. Circular polarization was not detected. From the noise level in the Stokes-$V$ map, the 3$\sigma$ upper limit on the polarization fraction is 21%.
IRAM 04191+1522 IRS weakened in the subsequent observing runs (Figure 1). Figure 2(a) shows the light curve. While IRAM 04191+1522 IRS is clearly a variable radio source, the timescale of the variability in 2009 October–November can be constrained only roughly because the number of data points is small. In the simplest case, the light curve may be showing a single outburst event. If the peak intensity occurred before the first run, the light curve in Figure 2(a) covers the waning phase only, and the decay half-life is $\sim$20 days. If it occurred between the first and second runs, the FWHM timescale may be in the range of 10 to 40 days. (The 40 day limit is for a box-shaped light curve, and a more likely value for Gaussian-like profiles is $\sim$20 days.) More realistically, the light curves of YSO radio outbursts show month-scale periods of enhanced activities that consist of several day-scale outbursts (Bower et al. 2003; Forbrich et al. 2006). If this is the case for IRAM 04191+1522 IRS, and if the detections of the first and second runs were from separate outburst events, the timescale of each event may be shorter than $\sim$10 days. In summary, the timescale of the IRAM 04191+1522 IRS flux variability may be $\sim$20 days or shorter.
The data over 17 yr from 1992 to 2009 show that IRAM 04191+1522 IRS was detectable ($>$ 0.05 mJy) about 50% of the time (Table 2, Figure 2(b)), which suggests that the radio outburst is not a rare phenomenon on this protostar. The 1996/1997 observations and results were reported previously by Andr[é]{} et al. (1999). In all the four runs when IRAM 04191+1522 IRS was detected, the source was unresolved.
The radio emission of IRAM 04191+1522 IRS in the quiescent state can be constrained from the observations of 2003 January. The 3$\sigma$ upper limit of the quiescent flux density is 0.02 mJy. Then the 2009 October–November outburst was stronger than the quiescent level by a factor of seven or larger. If we limit our attention only to the 2009 observations covering 43 days, the peak flux (first run) is $\sim$3.4 times the upper limit of the third run.
L1148 IRS (DCE 32)
------------------
DCE 32 corresponds to L1148 IRS, a Class I protostar driving a compact outflow and residing in a low-mass cloud core (Kauffmann et al. 2011). L1148 IRS was detected in 2005 (Table 3). The radio source position agrees with that of the infrared source within 0$\farcs$9. L1148 IRS was undetected in our survey. L1148 IRS seems to be a time-variable source (Figure 3), but it is difficult to understand the nature of this radio source because it was weak and detected only once.
L1014 IRS (DCE 38)
------------------
DCE 38 corresponds to L1014 IRS, a Class 0 protostar driving a compact outflow (Young et al. 2004; Bourke et al. 2005). L1014 IRS was detected in 2004 and 2005 (Table 3). The observations and results were reported by Shirley et al. (2007) along with the information on the detections at 4.9 GHz. The radio source position agrees with that of the infrared source within 0$\farcs$7.
As L1014 IRS was not detected in 2009, it is clearly a time-variable source (Figure 3). The detections in 2004 and 2005 may represent separate radio outbursts, and the outburst timescale of each event cannot be constrained. Alternatively, it is in principle possible that the 2004/2005 detections were from a single event with a timescale longer than 1.4 yr. In 2004, circular polarization of the 8.5 GHz continuum was marginally detectable with a Stokes-$V$ flux density of 0.030 $\pm$ 0.008 mJy. The polarization fraction is $\sim$30%. Shirley et al. (2007) reported that the 4.9 GHz continuum flux is also variable and that circular polarization was detected in 2004 August with a polarization fraction of $\sim$50%.
DISCUSSION
==========
The Nature of Variable Radio Emission
-------------------------------------
The results of this survey, together with the archival data, suggest that at least 20%–30% of VeLLOs are radio variables. Out of the 10 target sources, our survey and the archival data revealed that three sources showed both detections and nondetections (for a detection threshold of $\sim$0.05 mJy), and the signal-to-noise ratios at the time of detections were five or higher. The 20 day (or shorter) variability timescale of the IRAM 04191+1522 IRS outburst is comparable to that of the magnetic flares of YSOs. For example, the flares of GMR-A showed a timescale of several days (Bower et al. 2003) to $\sim$15 days (Furuya et al. 2003), and those of R CrA IRS 5b showed a timescale of $\sim$10 days (Forbrich et al. 2006; Choi et al. 2008).
The flux variability caused by other kinds of mechanism, such as the propagation of thermal radio jets, are much slower, typically taking several years or longer (e.g., Rodr[í]{}guez et al. 2000), and can be ruled out. If the variable radio emission were thermal radiation from molecular gas, it should have been easily detectable at shorter wavelengths. For example, the expected flux densities of the thermal emission are at least 20 mJy at 3 mm and at least 200 mJy at 1 mm. The observed values, however, were always much weaker: undetectable ($<$ 2 mJy) at 3.3 mm (Lee et al. 2005) and 6–9 mJy at 1.3 mm (Andr[é]{} et al. 1999; Belloche et al. 2002; Chen et al. 2012). Therefore, the thermal emission can be ruled out.
The modest amount of circular polarization of L1014 IRS (Section 3.3; Shirley et al. 2007) suggests that the detected emission was gyrosynchrotron radiation from mildly relativistic electrons, most likely accelerated in magnetic activities. The nondetection of circular polarization in the IRAM 04191+1522 IRS outburst does not necessarily contradict the nonthermal nature of the emission because the amount of polarization depends on many factors such as geometrical effects and because the detection upper limit ($\sim$20%) is rather high. For example, the polarization fraction of R CrA IRS 5b is $\sim$17% (Choi et al. 2009).
Considering the variability timescale and flux level of IRAM 04191+1522 IRS and the circular polarization of L1014 IRS together, the variable radio emission of VeLLOs may be coming from magnetic flares. Therefore, about 20%–30% of VeLLOs may be magnetically active, which may reflect the magnetic properties of the natal clouds (Nakano et al. 2002). These observations of VeLLO radio outbursts may be considered as the detection of light directly coming from the atmospheres of the youngest stellar objects. However, given the limited information available, the nature of the variable radio emission of VeLLOs is not firmly settled and needs more extensive investigations. For example, more well-sampled light curves are necessary to constrain the timescale better.
It should be noted the observations presented in this paper were designed to detect variabilities on timescales of weeks to months. It is not clear whether the radio variability of VeLLOs discussed above is analogous to solar-type flares that show timescales shorter than an hour and flux enhancements over orders of magnitude. The detected variability may be caused by either several solar-type flares or a different kind of phenomenon unique to protostars. To address this issue, detections of circular polarization and measurements of spectral index at the time of outbursts are needed to confirm the nonthermal radiation mechanism, and nearly continuous observations over several days are necessary.
The undetected VeLLOs were probably magnetically inactive during the time of observations. It is also possible that their protostellar winds may be opaque and blocking the radio emission from flares, considering that some VeLLOs can have optically thick winds (see the range of optical depth in Section 1). The existence of magnetic activities may apply to more typical protostars (with luminosities higher than those of VeLLOs), because magnetic properties do not explicitly depend on the luminosity. However, their optically thick winds make it difficult to observationally confirm any magnetic activity.
Implications of Magnetic Flares
-------------------------------
In this section we discuss some implications of the radio-variable VeLLOs assuming that the radio outbursts were caused by magnetic flares. However, the evidences are not absolutely conclusive, and the issues discussed below should be revisited when more and better data are available in the future.
### The Origin of Magnetic Fields
The detection of VeLLO radio flare suggests that at least some Class 0 protostars are magnetically active. The origin of the magnetic fields is an intriguing issue. The magnetic fields of more evolved objects, such as T Tauri stars and main-sequence stars, may be generated by either solar type or convective dynamos that require gas circulations such as convection in the stellar interior (Donati & Landstreet 2009; Mestel 2012). Protostars may start to develop a convection zone when the mass reaches $\sim$0.36 $M_\odot$, which corresponds to a time soon after the onset of deuterium burning (Stahler et al. 1980a, 1980b). However, IRAM 04191+1522 IRS has a collapse age of $\sim$10,000 yr and a stellar mass of $\sim$0.05 $M_\odot$ (Andr[é]{} et al. 1999), assuming a mass accretion rate of typical protostars. Considering the low luminosity, the accretion rate and the stellar mass can be even smaller. The small age and mass suggest that this protostar has not developed any convection zone yet. That is, IRAM 04191+1522 IRS probably has an inert internal region and may be unable to operate a magnetic dynamo. The cases of L1014 IRS and L1148 IRS are less certain because their age and mass are poorly known. If their masses are smaller than 0.1 $M_\odot$ (Young et al. 2004; Kauffmann et al. 2011), they also have no convection zone and thus no convective dynamo.
The magnetic fields of pre-main-sequence stars are often considered to be a combination of those from the convective dynamo and those from the interstellar space (Mestel 1965; Tayler 1987). Since the VeLLOs above cannot operate a convective dynamo, the source of their magnetic fields may be the fossil fields that are interstellar magnetic fields of the parent molecular cloud, dragged into the protostar by the accreting matter. Therefore, the flares of VeLLOs may be caused by the fossil fields.
However, there is no direct evidence for the existence of fossil fields in and around protostars so far. It is in principle possible that protostars may have an as yet unknown type of internal motion that can drive a magnetic dynamo. For example, the fast spin of protostars may induce internal circulations, and protostars in a close binary system may have tidally-induced internal motions. Theoretical studies of such possibilities should be helpful but have not been pursued yet.
### High-energy Photons
While the magnetic flares are most easily detected at radio wavelengths, most of the energy is released at shorter wavelengths such as X-ray. Though this issue can be important in the physics and chemistry of circumstellar material, there is not much known because VeLLOs have not been detected in X-ray. Below, we try extrapolating the empirical relations of more evolved objects to investigate this issue. However, their applicability to Class 0 protostars has not been verified observationally, and the arguments below are speculative.
There is an empirical relation between radio and X-ray luminosities of large flares: $L_{\rm XF} / L_{\rm RF} = 10^{15\pm1}$ Hz, where $L_{\rm XF}$ and $L_{\rm RF}$ are X-ray and radio luminosities during flare events, respectively (G[ü]{}del & Benz 1993; G[ü]{}del 2002). Many magnetically active objects, including YSOs, follow this relation (e.g., Bower et al. 2003). If the 2009 October flare of IRAM 04191+1522 IRS followed this relation, the X-ray luminosity might have been $\sim$8 $\times$ 10$^{-4}$ $L_\odot$. By contrast, the X-ray luminosity of a quiescent protostar can be estimated using another empirical relation: $L_{\rm XP} / L^*$ = 10$^{-4}$ – 10$^{-3}$, where $L_{\rm XP}$ and $L^*$ are X-ray and bolometric luminosities of an YSO, respectively (G[ü]{}del et al. 2007). Most of the X-ray luminosity of a quiescent YSO may come from the stellar corona or numerous small flares (G[ü]{}del et al. 2003, 2007). From the internal luminosity of IRAM 04191+1522 IRS (Dunham et al. 2008), the X-ray luminosity of the protostar may be $\sim$10$^{-5}$ $L_\odot$. Comparison of the two X-ray luminosities suggests that, during large flare events, the magnetic flare may outshine the protostar by a large factor in the high-energy regime.
Considering that IRAM 04191+1522 IRS is in active states about half the time, most of the ionizing photons around this protostar may be produced by the flares rather than the stellar corona. The copious amount of high-energy photons from the magnetic flares may enhance the ionization fraction of the circumstellar material (Glassgold et al. 2000) and significantly affect the gas dynamics such as magnetic breaking and ambipolar diffusion. Therefore, the magnetic activity can be an important control factor that may regulate the growth of protostars (Pudritz & Silk 1987; Feigelson & Montmerle 1999; Mestel 2012).
We thank Jungyeon Cho for helpful discussions. M.C. was supported by the Core Research Program of the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT and Future Planning of the Korean government (grant No. NRF-2011-0015816). J.-E.L. was supported by the Basic Science Research Program through NRF funded by the Ministry of Education of the Korean government (grant No. NRF-2012R1A1A2044689). NRAO is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.
Andr[é]{}, P. 1996, in ASP Conf. Ser. 93, Radio Emission from the Stars and the Sun, ed. A. R. Taylor & J. M. Paredes (San Francisco, CA: ASP), 273 Andr[é]{}, P., Montmerle, T., Feigelson, E. D., Stine, P. C., & Klein, K.-L. 1988, ApJ, 335, 940 Andr[é]{}, P., Motte, F., & Bacmann, A. 1999, ApJL, 513, L57 Andr[é]{}, P., Ward-Thompson, D., & Barsony, M. 1993, ApJ, 406, 122 Arce, H. G., Shepherd, D., & Gueth, F., et al. 2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil (Tucson, AZ: Univ. Arizona Press), 245 Belloche, A., Andr[é]{}, P., Despois, D., & Blinder, S. 2002, A&A, 393, 927 Bourke, T. L., Crapsi, A., Myers, P. C., et al. 2005, ApJL, 633, L129 Bower, G. C., Plambeck, R. L., Bolatto, A., et al. 2003, ApJ, 598, 1140 Chen, H., Myers, P. C., Ladd, E. F., & Wood, D. O. S. 1995, ApJ, 445, 377 Chen, X., Arce, H. G., Dunham, M. M., & Zhang, Q. 2012, ApJL, 747, L43 Choi, M., Hamaguchi, K., Lee, J.-E., & Tatematsu, K. 2008, ApJ, 687, 406 Choi, M., Tatematsu, K., Hamaguchi, K., & Lee, J.-E. 2009, ApJ, 690, 1901 Donati, J.-F., & Landstreet, J. D. 2009, ARA&A, 47, 333 Dunham, M. M., Crapsi, A., Evans, N. J., II, et al. 2008, ApJS, 179, 249 Dunham, M. M., Evans, N. J., II, Bourke, T. L., et al. 2006, ApJ, 651, 945 Enoch, M. L., Young, K. E., Glenn, J., et al. 2006, ApJ, 638, 293 Feigelson, E. D., & Montmerle, T. 1999, ARA&A, 37, 363 Forbrich, J., Preibisch, T., & Menten, K. M. 2006, A&A, 446, 155 Furuya, R. S., Shinnaga, H., Nakanishi, K., Momose, M., & Saito, M. 2003, PASJ, 55, L83 Glassgold, A. E., Feigelson, E. D., & Montmerle, T. 2000, in Protostars and Planets IV, ed. V. Mannings, A. P. Boss, & S. S. Russell (Tucson, AZ: Univ. Arizona Press), 429 G[ü]{}del, M. 2002, ARA&A, 40, 217 G[ü]{}del, M., Audard, M., Kashyap, V. L., Drake, J. J., & Guinan, E. F. 2003, ApJ, 582, 423 G[ü]{}del, M., & Benz, A. O. 1993, ApJL, 405, L63 G[ü]{}del, M., Briggs, K. R., Arzner, K., et al. 2007, A&A, 468, 353 Hamaguchi, K., Choi, M., Corcoran, M. F., et al. 2008, ApJ, 687, 425 Kauffmann, J., Bertoldi, F., Bourke, T. L., et al. 2011, MNRAS, 416, 2341 Kenyon, S. J., Dobrzycka, D., & Hartmann, L. 1994, AJ, 108, 1872 Lee, C.-F., Ho, P. T. P., & White, S. M. 2005, ApJ, 619, 948 Lee, J.-E. 2007, JKAS, 40, 83 Maheswar, G., Lee, C. W., & Dib, S. 2011, A&A, 536, A99 Mestel, L. 1965, QJRAS, 6, 265 Mestel, L. 2012, Stellar Magnetism (New York: Oxford Univ. Press) Nakano, T., Nishi, R., & Umebayashi, T. 2002, ApJ, 573, 199 Pudritz, R. E., Ouyed, R., Fendt, C., & Brandenburg, A. 2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil (Tucson, AZ: Univ. Arizona Press), 277 Pudritz, R. E., & Silk, J. 1987, ApJ, 316, 213 Rodr[í]{}guez, L. F., Delgado-Arellano, V. G., G[ó]{}mez, Y., et al. 2000, AJ, 119, 882 Shirley, Y. L., Claussen, M. J., Bourke, T. L., Young, C. H., & Blake, G. A. 2007, ApJ, 667, 329 Shu, F. H., Adams, F. C., & Lizano, S. 1987, ARA&A, 25, 23 Stahler, S. W., Shu, F. H., & Taam, R. E. 1980a, ApJ, 241, 637 Stahler, S. W., Shu, F. H., & Taam, R. E. 1980b, ApJ, 242, 226 Strai[ž]{}ys, V., [Č]{}ernis, K., & Barta[š]{}i[ū]{}t[ė]{}, S. 2003, A&A, 405, 585 Tayler, R. J. 1987, MNRAS, 227, 553 Young, C. H., J[ø]{}rgensen, J. K., Shirley, Y. L., et al. 2004, ApJS, 154, 396
| {
"pile_set_name": "ArXiv"
} |
---
abstract: 'To extend the applicability of density functional theory for superconductors (SCDFT) to systems with significant particle-hole asymmetry, we construct a new exchange-correlation kernel entering the gap equation. We show that the kernel is numerically stable and does not diverge even in the low temperature limit. Solving the gap equation for model systems with the present kernel analytically and numerically, we find that the asymmetric component of electronic density of states, which has not been considered with the previous kernel, systematically decreases transition temperature ($T_{\rm c}$). We present a case where the decrease of $T_{\rm c}$ amounts to several tens of percent.'
author:
- 'Ryosuke Akashi$^{1}$'
- 'Ryotaro Arita$^{1,2}$'
title: 'Density Functional Theory for Superconductors with Particle-hole Asymmetric Electronic Structure '
---
Introduction
============
Non-empirical calculation of superconducting transition temperature ($T_{\rm c}$) has been a great challenge in theoretical physics. Conventionally, $T_{\rm c}$s of phonon-mediated superconductors are estimated by the McMillan-Allen-Dynes formula,[@McMillan; @AllenDynes] $$T_{\rm c}=\frac{\omega_{\rm ln}}{1.2}\exp \left( -\frac{1.04(1+\lambda)}{\lambda-\mu^\ast(1+0.62\lambda)}\right),$$ where $\omega_{\rm ln}$, $\lambda$ and $\mu^\ast$ (Ref. ) are the logarithmically averaged phonon frequency, electron-phonon coupling, and Coulomb pseudo-potential, respectively. This formula clearly indicates the extreme sensitivity of $T_{\rm c}$ to the phonon and electron properties; to achieve a reliable calculation of $T_{\rm c}$, we need to evaluate these factors with high accuracy. For the phonon properties, developments in methods based on density functional theory[@Baroni-review; @Kunc-Martin-frozen] enables one to obtain reliable values very efficiently. On the other hand, for $\mu^{\ast}$, while some calculation based on the static random-phase approximation has been made,[@Lee-Chang-Cohen] quantitative description of the retardation effect[@Morel-Anderson] still remains a serious obstacle for the calculation of $T_{\rm c}$.
Recently, the situation has changed by the progress in the density-functional theory for superconductors (SCDFT).[@Oliveira; @Kreibich; @GrossI] There, a non-empirical scheme treating both the electron-phonon and electron-electron interactions was formulated referring the Migdal-Eliashberg (ME) theory of strong-coupling superconductivity.[@Migdal; @Eliashberg; @Scalapino; @Schrieffer] Subsequent applications have shown that this scheme can reproduce experimental $T_{\rm c}$s of various conventional phonon-mediated superconductors with error less than a few K.[@GrossII; @Floris-MgB2; @Sanna-CaC6; @Bersier-CaBeSi]
An important advantage of the current SCDFT-based scheme is that the particle-hole symmetric component of electronic structures around the Fermi level is accurately treated on the basis of the density functional calculations. In the context of the ME theory, its effect on $T_{\rm c}$ has been intensively studied for A15 compounds.[@Testardi; @Ho-Cohen-Pickett; @Pickett-general; @Lie-Carbotte; @Mitrovic-Carbotte; @Schachinger-Carbotte; @Yokoya-Nakamura; @Yokoya] However, the particle-hole antisymmetric component ignored in the current SCDFT has also been shown to have a significant impact on $T_{\rm c}$.[@Yokoya] Moreover, there has also been a great experimental progress of electric-field-induced carrier-doping technique, and it has stimulated current interest to superconductivity in doped insulators having significant particle-hole asymmetry.[@Caviglia-LAOSTO; @Ueno-Kawasaki-SrTiO3; @Iwasa-ZrNCl-EDLT; @Ueno-Kawasaki-KTaO3; @Taniguchi-MoS2; @Iwasa-MoS2] With this background, it is important to extend the applicability of SCDFT to systems with particle-hole asymmetric electronic structure.
In Ref. , the assumption of particle-hole symmetry with respect to the Fermi energy ($E_{\rm F}$) was originally introduced to avoid numerical divergence in the exchange-correlation kernel. In this paper, we show that we can construct an exchange correlation functional for systems with asymmetric electronic structure which does not have any divergences. In Sec. \[sec:theory\], we show how the kernel is constructed, and discuss how it is numerically stable. In Sec. \[sec:calc\], we perform calculations of $T_{\rm c}$ for the cases where the present improvement becomes indeed important. Section \[sec:summary\] is devoted to summary and conclusion.
gap-equation kernel for particle-hole asymmetry {#sec:theory}
===============================================
In this section, we describe the improvement of the gap equation based on the current SCDFT to include the particle-hole asymmetry effect. In the current SCDFT,[@GrossI] we solve the following gap equation: $$\begin{aligned}
\Delta_{n{\bf k}}\!=\!-\mathcal{Z}_{n\!{\bf k}}\!\Delta_{n\!{\bf k}}
\!-\!\frac{1}{2}\!\sum_{n'\!{\bf k'}}\!\mathcal{K}_{n\!{\bf k}\!n'{\bf k}'}
\!\frac{\mathrm{tanh}[(\!\beta/2\!)\!E_{n'{\bf k'}}\!]}{E_{n'{\bf k'}}}\!\Delta_{n'\!{\bf k'}}.
\label{eq:gap-eq}\end{aligned}$$ Here, $n$ and ${\bf k}$ denote the band index and crystal momentum, respectively, $\Delta$ is the gap function, and $\beta$ is the inverse temperature. The energy $E_{n {\bf k}}$ is defined as $E_{n {\bf k}}$=$\sqrt{\xi_{n {\bf k}}^{2}+\Delta_{n {\bf k}}^{2}}$ and $\xi_{n {\bf k}}=\epsilon_{n {\bf k}}-\mu$ is the one-electron energy measured from the chemical potential $\mu$, where $\epsilon_{n {\bf k}}$ is obtained by solving the normal Kohn-Sham equation in density functional theory (DFT).
Functions $\mathcal{Z}$ and $\mathcal{K}$ are the exchange-correlation kernels, which represent the effects of mass renormalization and effective pairing interactions, respectively. In the SCDFT, $\mathcal{Z}$ is treated by only the phonon contribution as $$\begin{aligned}
\mathcal{Z}_{n\!{\bf k}}=\mathcal{Z}^{\rm ph}_{n\!{\bf k}} \end{aligned}$$ with $\mathcal{Z}^{\rm ph}_{n{\bf k}}$ representing the renormalization of electronic states by the phonon exchange, and $\mathcal{K}$ consists of both the electron-phonon $\mathcal{K}^{\rm ph}_{n\!{\bf k}\!n'{\bf k}'}$ and electron-electron $\mathcal{K}^{\rm el}_{n\!{\bf k}\!n'{\bf k}'}$ exchange contributions $$\begin{aligned}
\mathcal{K}_{n\!{\bf k}\!n'{\bf k}'}=
\mathcal{K}^{\rm ph}_{n\!{\bf k}\!n'{\bf k}'}+
\mathcal{K}^{\rm el}_{n\!{\bf k}\!n'{\bf k}'}.\end{aligned}$$
Formally, $\mathcal{Z}^{\rm ph}_{n{\bf k}}$ is derived from the functional derivative of the free energy given by the Kohn-Sham perturbation theory with respect to the anomalous density [@GrossI] as $$\begin{aligned}
\mathcal{Z}^{\rm ph,PT}_{n{\bf k}}
\equiv
\mathcal{Z}^{\rm ph}_{1,n{\bf k}}
+
\mathcal{Z}^{\rm ph}_{2,n{\bf k}}
\label{eq:Z-original}\end{aligned}$$ with $$\begin{aligned}
\mathcal{Z}^{\rm ph}_{1,n{\bf k}}
&=&
\frac{1}{{\rm tanh}\bigl[(\beta/2)\xi_{n{\bf k}}\bigr]}
\sum_{n'{\bf k}'}
\sum_{\nu}
|g^{n{\bf k},n'{\bf k}'}_{\nu {\bf k}-{\bf k}'}|^{2}
\nonumber \\
&&
\times
\biggl\{\!
\frac{1}{\xi_{n{\bf k}}}
[
I(\xi_{n{\bf k}},\xi_{n'{\bf k}'},\omega_{\nu {\bf k}-{\bf k}'})\!\!
-\!\!
I(\xi_{n{\bf k}},-\xi_{n'{\bf k}'},\omega_{\nu {\bf k}-{\bf k}'})
]
\nonumber
\\
&&\hspace{40pt}
-2
I'(\xi_{n{\bf k}},\xi_{n'{\bf k}'},\omega_{\nu {\bf k}-{\bf k}'})
\biggr\}
\label{eq:Z1-original}\end{aligned}$$ and $$\begin{aligned}
\mathcal{Z}^{\rm ph}_{2,n{\bf k}}
&=&
-
\frac
{2}
{\sum_{n'{\bf k}'}(\beta/2)/{\rm cosh}^{2}
\bigl[(\beta/2)\xi_{n'{\bf k}'} \bigr]}
\nonumber \\
&&
\times
\biggl[
\frac{1}{\xi_{n{\bf k}}}
-
\frac
{\beta/2}
{{\rm sinh\bigl[(\beta/2)\xi_{n{\bf k}}\bigr]}
{\rm cosh\bigl[(\beta/2)\xi_{n{\bf k}}\bigr]}}
\biggr]
\nonumber \label{eq:Z1-original}
\\
&&
\times \!\!
\sum_{\substack{m_{1}{\bf l}_{1}\\m_{2}{\bf l}_{2}}}\!\!
\sum_{\nu}
|g^{m_{1}{\bf l}_{1},m_{2}{\bf l}_{2}}_{\nu {\bf l}_{1}-{\bf l}_{2}}|^{2}
I'\!(\xi_{m_{1}{\bf l}_{1}},\xi_{m_{2}{\bf l}_{2}},\omega_{\nu {\bf l}_{1}-{\bf l}_{2}})
.\label{eq:Z2-original}\end{aligned}$$ Here, $g^{n{\bf k},n'{\bf k}'}_{\nu {\bf k}-{\bf k}'}$ denotes the electron-phonon coupling, and $I$ and $I'$ are defined as $$\begin{aligned}
I(\xi,\xi',\omega)
&=&
f_{\beta}(\xi)f_{\beta}(\xi')n_{\beta}(\omega)
\nonumber \\
&&
\biggl[
\frac{e^{\beta \xi}-e^{\beta(\xi'+\omega)}}{\xi-\xi'-\omega}
-
\frac{e^{\beta \xi'}-e^{\beta(\xi+\omega)}}{\xi-\xi'+\omega}
\biggr]
,\label{eq:I-func}
\\
I'(\xi,\xi',\omega)
&=&
\frac{\partial}{\partial \xi}
I(\xi,\xi',\omega), \label{eq:Idiff-func}\end{aligned}$$ with $f_{\beta}$ and $n_{\beta}$ denoting the Fermi and Bose distribution functions, respectively. We note that $\mathcal{Z}^{\rm ph}_{1}$ originates from explicit dependence of the free energy on the anomalous density, whereas $\mathcal{Z}^{\rm ph}_{2}$ comes from the implicit dependence via the chemical potential.
Lüders *et al.*[@GrossI] reported that this form suffers from numerical divergence and that the divergent contribution is antisymmetric with respect to $E_{\rm F}$. Accordingly, in their paper they symmetrized $\mathcal{Z}^{\rm ph}_{1}$ in $\xi_{n{\bf k}}$, and dropped $\mathcal{Z}^{\rm ph}_{2}$ because of its antisymmetric dependence on $\xi_{n{\bf k}}$. The resulting form is $$\begin{aligned}
\!\!\!\!\mathcal{Z}^{\rm ph,sym}_{n{\bf k}}
&=&
-\frac{1}{{\rm tanh}\bigl[(\beta/2) \xi_{n{\bf k}}\bigr]}
\sum_{n'{\bf k}'}
\sum_{\nu}
|g^{n{\bf k},n'{\bf k}'}_{\nu {\bf k}-{\bf k}'}|^{2}
\nonumber \\
&& \!
\times \!
\bigl[\!
I'\!(\xi_{n{\bf k}},\xi_{n'{\bf k}'}\!,\omega_{\nu {\bf k}\!-\!{\bf k}'})\!
\!+\!\!
I'\!(\xi_{n{\bf k}},\!-\xi_{n'{\bf k}'}\!,\omega_{\nu {\bf k}\!-\!{\bf k}'}\!)
\!\bigr]
.
\label{eq:Z-ph-sym}\end{aligned}$$ Further, they modified an unphysically large component of $\mathcal{Z}^{\rm ph,sym}_{n{\bf k}}$ around $E_{\rm F}$, and obtained the following form $$\begin{aligned}
\!\!\!
\mathcal{Z}^{\rm ph}_{n{\bf k}}
&=&
-\frac{1}{{\rm tanh}\bigl[(\beta/2) \xi_{n{\bf k}}\bigr]}
\sum_{n'{\bf k}'}
\sum_{\nu}
|g^{n{\bf k},n'{\bf k}'}_{\nu {\bf k}-{\bf k}'}|^{2}
\nonumber \\
&& \!
\times \! \!
\bigl[
J\!(\xi_{n{\bf k}},\xi_{n'{\bf k}'}\!,\omega_{\nu {\bf k}\!-\!{\bf k}'})\!
+\!
J\!(\xi_{n{\bf k}},\!-\xi_{n'{\bf k}'}\!,\omega_{\nu {\bf k}\!-\!{\bf k}'})
\bigr]
,
\label{eq:Z-ph-prev}\end{aligned}$$ where $J$ is defined by Eq. (80) in Ref. . So far, only this form has been used in practice.
In Eqs. (\[eq:Z-ph-sym\]) and (\[eq:Z-ph-prev\]), the antisymmetric parts of the electronic density of states (DOS) and $|g^{n{\bf k},n'{\bf k}'}_{\nu {\bf k}-{\bf k}'}|^{2}$ with respect to $E_{\rm F}$ within the phonon energy scale are ignored. Within the standard ME theory, the effect of their antisymmetric parts is assumed to be minor[@Pickett-general; @Mitrovic] and often ignored for simplicity. [@Schrieffer] On the other hand, in systems with rapidly varying DOS, this treatment is also known to lead to a considerable error in the estimate of $T_{\rm c}$ and the isotope-effect coefficient.[@Yokoya] Below, we will show a numerically stable form of $\mathcal{Z}$ without the symmetrization, with which the asymmetry effect is properly treated.
Cancellation of divergent terms {#subsec:cancel}
-------------------------------
First, we analytically examine the divergence of $\mathcal{Z}^{\rm ph,PT}$. For simplicity, we deal with the $n{\bf k}$-averaged form of $\mathcal{Z}^{\rm ph,PT}$, defined by $$\begin{aligned}
\mathcal{Z}^{\rm ph,PT}(\xi)
&=&
\mathcal{Z}^{\rm ph}_{1}(\xi)
+
\mathcal{Z}^{\rm ph}_{2}(\xi)
,\label{eq:Z-ave-original}
\\
\mathcal{Z}^{\rm ph}_{i}(\xi)
&\equiv&
\frac{1}{N(\xi)}
\sum_{n{\bf k}} \delta(\xi-\xi_{n{\bf k}}) \mathcal{Z}^{\rm ph}_{i,n{\bf k}}
\hspace{5pt} (i=1,2)
,
\label{eq:Z-ave-Z1Z2}\end{aligned}$$ where $N(\xi)$ is the electronic DOS. We also introduce an approximation for the electron-phonon coupling $$\begin{aligned}
&&
\sum_{n{\bf k} n'{\bf k}' \atop \nu} \!\!\!\!
\delta(\xi \!-\! \xi_{n{\bf k}})
\delta(\xi' \!-\! \xi_{n'{\bf k}'})
\delta(\omega \!-\! \omega_{\nu{\bf k}-{\bf k}'})
|g^{n{\bf k},n'{\bf k}'}_{\nu {\bf k}-{\bf k}'}|^{2}
\nonumber \\
&&\hspace{50pt}
\simeq
\frac{N(\xi)N(\xi')}{N(0)} \alpha^{2}F(\omega)
,
\label{eq:G2-approx2}\end{aligned}$$ where the antisymmetric component of DOS is retained. The function $\alpha^{2}F(\omega)$ denotes the Eliashberg function, with which the electron-phonon coupling coefficient $\lambda$ and the characteristic frequency $\omega_{\rm ln}$ are defined as $$\begin{aligned}
\lambda
&=&
2\int d\omega
\frac{
\alpha^{2}F(\omega)
}{\omega}
,
\label{eq:lambda-def}
\\
\omega_{\rm ln}
&=&
{\rm exp}\left[
\frac{2}{\lambda}
\int d\omega
\frac{
\alpha^{2}F(\omega)
}{\omega}
{\rm ln}\omega
\right]
.
\label{eq:omega-def}\end{aligned}$$
Starting from the above form for $\mathcal{Z}$, in brief, we show that the numerically divergent terms analytically cancel with each other so that we can construct a numerically stable form. Readers who are not interested in the detail of this analysis can skip the remaining part of this subsection.
Let us first focus on $\mathcal{Z}^{\rm ph}_{1}$. As we show below, the divergence in this part basically comes from the slow decay of order $O(\xi'^{-1})$ of the integrands. In order to extract this, we transform the following factor appearing in $I$ and $I'$ \[Eqs. (\[eq:I-func\]) and (\[eq:Idiff-func\])\] as $$\begin{aligned}
\frac{1}{\xi-\xi'\pm\omega}
=
{\rm P}
\frac{\xi}{(\xi-\xi'\pm\omega)(\xi'\mp\omega)}
-
{\rm P}
\frac{1}{\xi'\mp\omega}
,\end{aligned}$$ The principal value integral ${\rm P}$ has been introduced to deal with the poles at $\xi'$$=$$\pm \omega$ in the right hand side; note that the expression in the left hand side does not have these poles. Substituting this expression into Eq. (\[eq:Z-ave-Z1Z2\]), inserting identities $1=\int d\xi' \delta(\xi'-\xi_{n'{\bf k}'})$ and $1=\int d\omega \delta(\omega-\omega_{\nu{\bf k}-{\bf k}'})$, and using Eq. (\[eq:G2-approx2\]), we obtain
$$\begin{aligned}
\mathcal{Z}^{\rm ph}_{1}(\xi)
&=&
\frac{1}{{\rm tanh}[(\beta/2)\xi]}
\int d\omega
\int d\xi'
\frac{N(\xi')}{N(0)} \alpha^{2}F(\omega)
\biggl\{
I_{0}(\xi,\xi',\omega)
+
I_{1}(\xi,\xi',\omega)
\nonumber \\
&& \hspace{70pt}
-2
[J_{0}(\xi,\xi',\omega)
+J_{1}(\xi,\xi',\omega)
+J_{2}(\xi,\xi',\omega)
]
\biggr\}
, \label{eq:Z-IJ-def}
\\
I_{0}(\xi,\xi',\omega)
&=&
-
\frac{1}{\xi}
f_{\beta}(\xi)f_{\beta}(\xi')n_{\beta}(\omega)
\bigl[
{\rm P}\frac{e^{\beta \xi} \!-\! e^{\beta(\xi'+\omega)}}{\xi'+\omega}
\!-\!
{\rm P}\frac{e^{\beta \xi'} \!-\! e^{\beta(\xi+\omega)}}{\xi'-\omega}
\!-\!
{\rm P}\frac{1 \!-\! e^{\beta(\xi+\xi'+\omega)}}{\xi'+\omega}
\!+\!
{\rm P}\frac{e^{\beta (\xi+\xi')} \!-\! e^{\beta(\omega)}}{\xi'-\omega}
\bigr]
,\label{eq:I0-def}
\\
I_{1}(\xi,\xi',\omega)
&=&
f_{\beta}(\xi)\!f_{\beta}(\xi'\!)\!n_{\beta}(\omega\!)\!
\biggl[\!
{\rm P}\frac{e^{\beta \xi}-e^{\beta(\xi'+\omega)}}{(\xi \!\!-\!\! \xi' \!\!\!-\!\! \omega)(\xi' \!\!\!+\!\! \omega)}
\!-\!
{\rm P}\frac{e^{\beta \xi'}-e^{\beta(\xi+\omega)}}{(\xi \!\!-\!\! \xi' \!\!\!+\!\! \omega)(\xi' \!\!\!-\!\! \omega)}
\!-\!
{\rm P}\frac{1-e^{\beta(\xi+\xi'+\omega)}}{(\xi \!\!+\!\! \xi' \!\!\!+\!\! \omega)(\xi' \!\!\!+\!\! \omega)}
\!+\!
{\rm P}\frac{e^{\beta (\xi+\xi')}-e^{\beta \omega}}{(\xi \!\!+\!\! \xi' \!\!\!-\!\! \omega)(\xi' \!\!\!-\!\! \omega)}
\!
\biggr],
\label{eq:I1-def}
\\
J_{0}(\xi,\xi',\omega)
&=&
f'_{\beta}(\xi)f_{\beta}(\xi')n_{\beta}(\omega)
\bigl[
{\rm P}\frac{1+e^{\beta(\xi'+\omega)}}{\xi'+\omega}
+
{\rm P}\frac{e^{\beta \xi'}+e^{\beta \omega}}{\xi'-\omega}
\bigr]
,\label{eq:J0-def}
\\
J_{1}(\xi,\xi',\omega)
&=&
-f'_{\beta}(\xi)f_{\beta}(\xi')n_{\beta}(\omega)\xi
\biggl[
{\rm P}\frac{1+e^{\beta(\xi'+\omega)}}{(\xi-\xi'-\omega)(\xi'+\omega)}
+
{\rm P}\frac{e^{\beta\xi'}+e^{\beta \omega}}{(\xi-\xi'+\omega)(\xi'-\omega)}
\biggr]
,
\label{eq:J1-def}
\\
J_{2}(\xi,\xi',\omega)
&=&
-f_{\beta}(\xi)f_{\beta}(\xi')n_{\beta}(\omega)
\biggl[
\frac{e^{\beta \xi}-e^{\beta(\xi'+\omega)}}{(\xi-\xi'-\omega)^{2}}
-
\frac{e^{\beta \xi'}-e^{\beta(\xi+\omega)}}{(\xi-\xi'+\omega)^{2}}
\biggr]
\label{eq:J2-def}
.\end{aligned}$$
Here $I_{i}$ ($i$$=$$0,1$) and $J_{i}$ ($i$$=$$0,1,2$) denote the terms originating from $I$ and $I'$, respectively. From this expression, the divergent contribution $\mathcal{Z}^{\rm ph,div}_{1}$ is extracted as $$\begin{aligned}
\mathcal{Z}^{\rm ph,div}_{1}(\xi)
\!&=&\!
\frac{1}{{\rm tanh}[(\beta/2)\xi]}
\int d\omega
\int d\xi'
\frac{N(\xi')}{N(0)} \alpha^{2}F(\omega)
\nonumber \\
&&\hspace{20pt}
\times
\biggl\{
I_{0}(\xi,\xi'\!,\omega)
\!-\!
2J_{0}(\xi,\xi'\!,\omega)
\biggr\}
.\label{eq:Z1-div}\end{aligned}$$ We do not explicitly write the variables $(\xi,\xi',\omega)$ unless necessary in the following.
Let us examine $\mathcal{Z}^{\rm ph,div}_{1}$. For simplicity, we start from the case of constant electronic DOS $N(\xi)$$=$$N(0)$ with asymmetric energy cutoffs $[-L_{2},L_{1}]$ as depicted in Fig. \[fig:DOS-const\], which is the simplest case of particle-hole asymmetry. We then get to
$$\begin{aligned}
\mathcal{Z}^{\rm ph,div}_{1}(\xi)
\!=\!
\frac{1}{{\rm tanh}[(\beta/2)\xi]}
\!\int \!\!d\omega \alpha^{2}F(\omega)
\int^{L_{1}}_{-L_{2}} \!\!d\xi'
\nonumber \\
\hspace{20pt}\times
\biggl\{
I_{0}
\!-\!
2J_{0}
\biggr\}
.\end{aligned}$$
![Constant DOS with asymmetric cutoff energies, $L_{1}$ and $-L_{2}$.[]{data-label="fig:DOS-const"}](DOS_const_fig_130425.eps){width="6cm"}
Next we carry out the $\xi'$ integral. Since the integrand is of order $O(\xi'^{-1})$, integration results in cutoff-dependent terms as $$\begin{aligned}
\int^{L_{1}}_{-L_{2}} d\xi' I_{0}
&\simeq&
\frac{1}{\xi}
\bigl[
f_{\beta}(\xi)
-
f_{\beta}(-\xi)
\bigr]
{\rm ln}\left|\frac{L_{1}+\omega}{L_{2}+\omega}\right|
,
\\
\int^{L_{1}}_{-L_{2}} d\xi' J_{0}
&\simeq&
f'_{\beta}(\xi)
{\rm ln}\left|\frac{L_{1}+\omega}{L_{2}+\omega}\right|
.\end{aligned}$$ Here we have neglected terms of order $O(e^{-\beta \omega})$, $O(e^{-\beta L_{1}})$ and $O(e^{-\beta L_{2}})$, and this approximation is implicitly used in the following equations with “$\simeq$" in this section. We consequently get to $$\begin{aligned}
\mathcal{Z}^{\rm ph,div}_{1}(\xi)
&\simeq&
-
\left\{
\frac{1}{\xi}
-
\frac{\beta/2}{{\rm sinh[(\beta/2)\xi]}{\rm cosh[(\beta/2)\xi]}}
\right\}
\nonumber \\
&& \hspace{35pt}
\times
\int d\omega
\alpha^{2}F(\omega)
{\rm ln}\left|\frac{L_{1}+\omega}{L_{2}+\omega}\right|
.
\label{eq:I0J0-final}\end{aligned}$$ The $\xi$-dependent factor here is reminiscent of $\mathcal{Z}^{\rm ph}_{2}$ \[Eq. (\[eq:Z2-original\])\]. In fact, we prove that this term exactly cancels $\mathcal{Z}^{\rm ph}_{2}(\xi)$ in the following.
Next, we consider $\mathcal{Z}^{\rm ph}_{2}(\xi)$ using the approximations of the same level. The $n{\bf k}$-averaged form is explicitly written as $$\begin{aligned}
\mathcal{Z}^{\rm ph}_{2}(\xi)
&=&
-
\biggl[
\frac{1}{\xi}
-
\frac
{\beta/2}
{{\rm sinh\bigl[(\beta/2)\xi\bigr]}
{\rm cosh\bigl[(\beta/2)\xi\bigr]}}
\biggr]
\nonumber
\\
&& \!\!\!\!\!\!\!\!
\times \!\!
\int \! d\omega \alpha^{2}F(\omega) \! \!
\int^{L_{1}}_{-L_{2}}\!\!\!d\xi' \!
\int^{L_{1}}_{-L_{2}}\!\!\!d\xi \!
I'(\xi,\xi',\omega)
,\end{aligned}$$ where we have used $$\begin{aligned}
\sum_{n'{\bf k}'}(\beta/2)/{\rm cosh}^{2}
\bigl[(\beta/2)\xi_{n'{\bf k}'} \bigr]
\simeq
2N(0)
.\end{aligned}$$ By carrying out the integration $\int d\xi \int d\xi'$ and ignoring terms of order $O[(T/L_{1})^{2},(T/L_{2})^{2}]$, we get $$\begin{aligned}
\int^{L_{1}}_{-L_{2}} d \xi'
\int^{L_{1}}_{-L_{2}} d \xi
I'(\xi,\xi',\omega)
\simeq
{\rm ln}\left|\frac{L_{2}+\omega}{L_{1}+\omega} \right|
,\end{aligned}$$ and immediately obtain $$\begin{aligned}
\mathcal{Z}^{\rm ph}_{2}(\xi)
&\simeq&
\biggl[
\frac{1}{\xi}
-
\frac
{\beta/2}
{{\rm sinh\bigl[(\beta/2)\xi\bigr]}
{\rm cosh\bigl[(\beta/2)\xi\bigr]}}
\biggr]
\nonumber
\\
&&
\times
\int d\omega \alpha^{2}F(\omega)
{\rm ln}\left|\frac{L_{1}+\omega}{L_{2}+\omega} \right|
.
\label{eq:Z2-final}\end{aligned}$$ This expression exactly cancels out $\mathcal{Z}^{\rm ph,div}_{1}(\xi)$ \[Eq. (\[eq:I0J0-final\])\].
One can observe the same cancellation for the nonconstant DOS case. The cutoff-dependent parts $\mathcal{Z}^{\rm ph,div}_{1}$ and $\mathcal{Z}^{\rm ph}_{2}$ then read $$\begin{aligned}
\mathcal{Z}^{\rm ph,div}_{1}(\xi)
&\simeq&
\frac{1}{{\rm tanh}\frac{\beta}{2}\xi}\!
\int \!\! d\omega a^{2}F(\omega) \!\!
\int \!\! d\xi' \frac{N(\xi')}{N(0)}
\bigl[\!
I_{0}
\!-\!2J_{0}
\!\bigr]
,\label{eq:Z1-nonconst}
\\
\mathcal{Z}^{\rm ph}_{2}(\xi)
&\simeq&
-
\biggl\{\!
\frac{1}{\xi}
\!-\!
\frac{\beta/2}{{\rm cosh}[(\beta/2)\xi]{\rm sinh}[(\beta/2)\xi]}
\biggr\}\!\!
\int \!\!d\omega
\alpha^{2}\!F(\omega)
\nonumber \\
&&
\times
\int^{L_{1}}_{-L_{2}}\!\!\!d\xi
\frac{N(\xi)}{N(0)}
\int^{L_{1}}_{-L_{2}}\!\!\!d\xi'
\frac{N(\xi')}{N(0)}
I'(\xi,\xi',\omega)
.
\label{eq:Z2-nonconst}\end{aligned}$$
![Decomposition procedure of DOS. Typical DOS is represented as red-shaded area.[]{data-label="fig:DOS-general"}](DOS_general_fig_130425.eps){width="8cm"}
Here, we approximate the DOS as a step-like function with a certain set of energy tics $\{\pm\epsilon_{i}\}$ $(i=1, 2, . . ., N_{\epsilon}; \epsilon_{i} < {\rm min}[L_{1},L_{2}])$. The approximate function is then decomposed into functions $\theta^{+}_{i}(\xi) \equiv \theta(\xi-\epsilon_{i}) + \theta(-\xi-\epsilon_{i})$ and $\theta^{-}_{i}(\xi) \equiv \theta(\xi-\epsilon_{i}) - \theta(-\xi-\epsilon_{i})$ as $$\begin{aligned}
N(\xi)
=
N(0)
\biggl\{
1+
\sum_{i=1}^{N_{\epsilon}}
[
N^{-}_{i}\theta^{-}_{i}(\xi)
+
N^{+}_{i}\theta^{+}_{i}(\xi)
]
\biggr\}
.
\label{eq:DOS-step}\end{aligned}$$ This procedure is schematically depicted in Fig. \[fig:DOS-general\]. For $\mathcal{Z}^{\rm ph,div}_{1}$, straightforward calculations yield $$\begin{aligned}
&&
\mathcal{Z}^{\rm ph,div}_{1}\!(\xi)
\!\simeq\!
-
\biggl\{\!
\frac{1}{\xi}
\!-\!
\frac{\beta/2}{{\rm cosh}[(\beta/2)\xi]{\rm sinh}[(\beta/2)\xi]}
\biggr\}\!\!\!
\int \!\! d\omega \alpha^{2}\!F(\omega)
\nonumber \\
&& \hspace{40pt}
\times
\biggl\{
{\rm ln}
\left|
\frac{L_{1} \!+\! \omega}{L_{2}\!+\!\omega}
\right|
\!+\!
\sum_{i}
\biggl[
N^{-}_{i}
{\rm ln}
\left|
\frac{(L_{1} \!+\! \omega)(L_{2} \!+\! \omega)}{(\epsilon_{i} \!+\! \omega)^{2}}
\right|
\nonumber \\
&& \hspace{100pt}
+
N^{+}_{i}
{\rm ln}
\left|
\frac{L_{1} \!+\! \omega}{L_{2} \!+\! \omega}
\right|
\biggr]
\biggr\}
.
\label{eq:Z1-generalDOS}\end{aligned}$$ For $\mathcal{Z}^{\rm ph}_{2}$, terms proportional to $N^{\pm}_{i}N^{\pm}_{j}$ (any double sign) seemingly become nonzero, but careful calculations show that these terms cancel out each other (See Appendix \[app:eval-Z2\]). Finally, we see that the cutoff-dependent part in $\mathcal{Z}^{\rm ph}_{2}$ exactly cancels Eq. (\[eq:Z1-generalDOS\]). By taking limit for the number of tics $N_{\epsilon} \rightarrow \infty$, the present cancellation is proved even for general DOS.[@note-proof]
Now let us point out the important aspect of Eq. (\[eq:Z-ave-original\]): With the present approximation, the cutoff dependent divergence cancels if analytically calculated, and the apparently divergent contribution is in fact nonsingular with respect to $\xi$, since it is proportional to a nonsingular function $
\frac{1}{\xi}
-
\frac{\beta/2}{{\rm cosh}[(\beta/2)\xi]{\rm sinh}[(\beta/2)\xi]}
$. In order to verify this, we have calculated $\mathcal{Z}^{\rm ph,PT}(\xi)$ \[Eq. (\[eq:Z-ave-original\])\] with a certain asymmetric model DOS \[depicted later in Sec. \[sec:calc\]\]. We have generated the energy grid so that it becomes uniform in logarithmic scale. The calculated result is shown in Fig. \[fig:Zconv\]. By lowering the minimum energy cutoff for the grid, convergence within order of $\lambda$ has been achieved, where the apparent divergence gradually vanishes. However, we have also found that the convergence requires formidably accurate integration within the energy scale smaller than temperature, and so it is extremely difficult to achieve in practical calculations.
![Calculated values of $\mathcal{Z}^{\rm ph,PT}(\xi)$ \[Eq. (\[eq:Z-ave-original\])\] with the approximation given in Eq. \[eq:G2-approx2\]. Model DOS depicted in Fig. \[fig:Tc-ratio\] with $E_{\rm F}=0$ and model Eliashberg function given by Eq. (\[eq:a2F-model\]) with $\lambda=1$, $\omega_{E}=400$cm$^{-1}$, and $\sigma=\omega_{E}/10$ were used. The calculations were done with $T=0.1$K. Logarithmically uniform energy grids with mininum energy cutoff of $10^{-7}$ (red solid line), $10^{-8}$ (green dashed line), $10^{-9}$ (blue dotted line), and $10^{-10}$ (violet, dot-dashed line) were used. The DOS values for $|\xi|$$<$$10T$ were fixed to $N(0)$ so that it is consistent with the condition ($\epsilon_{i}$$\gg$$T$) assumed in Sec. \[subsec:cancel\].[]{data-label="fig:Zconv"}](Model_step_Z_prev_conv_combine_130502.eps){width="9cm"}
$\mathcal{Z}$ kernel for particle-hole asymmetic systems {#subsec:derive-Z-new}
--------------------------------------------------------
The above analysis specifies the numerically diverging (or *unstable*) but analytically irrelevant terms. In order to circumvent the numerical difficulty, we propose to simply subtract $\mathcal{Z}^{\rm ph,div}_{1}$ and $\mathcal{Z}^{\rm ph}_{2}$, obtaining $$\begin{aligned}
\mathcal{Z}^{\rm ph,aux}(\xi)
&=&
\frac{1}{{\rm tanh}[(\beta/2)\xi]}
\int d\omega
\alpha^{2}F(\omega)
\int d\xi'
\frac{N(\xi')}{N(0)}
\nonumber \\
&&
\times
\bigl[
I_{1}
-2J_{1}-2J_{2}
\bigr]
.
\label{eq:Z-aux}\end{aligned}$$
This auxiliary expression shows stable convergence, includes the effect of the antisymmetric part of electronic states, and agrees with Eq. (\[eq:Z-ave-original\]) with error of order $O[(T/L_{1})^{2}, (T/L_{2})^{2}]$ in the limit where Eq. (\[eq:G2-approx2\]) becomes exact. The form given in Eq. (\[eq:Z-aux\]) is a generalization of the $n{\bf k}$-averaged form of Eq. (\[eq:Z-ph-sym\]) $$\begin{aligned}
\mathcal{Z}^{\rm ph,sym}(\xi)
&=&
-\frac{1}{{\rm tanh}\bigl[(\beta/2) \xi\bigr]}
\int d\omega \alpha^{2}F(\omega)
\int d\xi'
\frac{N(\xi')}{N(0)}
\nonumber \\
&& \!
\times \!
\bigl[
I'\!(\xi,\xi',\omega)\!
+\!
I'\!(\xi,\!-\xi',\omega)
\bigr]
:
\label{eq:Z-ph-ave-sym}\end{aligned}$$ Symmetrization of Eq. (\[eq:Z-aux\]) in $\xi$ yields Eq. (\[eq:Z-ph-ave-sym\]), the limiting value $\lim_{\xi \rightarrow 0}[\mathcal{Z}^{\rm ph,aux}(\xi)-\mathcal{Z}^{\rm ph,sym}(\xi)]=0$ because of the fact that the integrand for $\xi$$=$$0$ is symmetric in $\xi'$, and the two forms show almost the same temperature dependence.
Following the procedure of Ref. , we next modify $\mathcal{Z}^{\rm ph,aux}(\xi)$ around $E_{\rm F}$ and obtain $$\begin{aligned}
&&
\mathcal{Z}^{\rm ph,aux2}(\xi)
\nonumber \\
&& \hspace{10pt}
=
\frac{1}{{\rm tanh}[(\beta/2)\xi]}
\int d\omega
\alpha^{2}F(\omega)
\int d\xi'
\frac{N(\xi')}{N(0)}
\nonumber \\
&& \hspace{30pt}
\times
\bigl[
I_{1}
\!-\! 2J'_{1}\!-\! 2J_{2}
\bigr]
,
\label{eq:Z-l2-def}
\\
&&
J'_{1}
=
f'_{\beta}(\xi)\xi
\biggl[ \!
\bigl\{
\! f_{\beta}(\xi \!-\! \omega) \!+\! n_{\beta}(-\omega) \!
\bigr\}
{\rm P}\frac{1}{(\xi \!-\! \xi' \!-\! \omega)(\xi' \!+\! \omega)}
\nonumber \\
&& \hspace{30pt}
-
\bigl\{
f_{\beta}(\xi \!+\! \omega) \!+\! n_{\beta}(\omega)
\bigr\}
{\rm P}\frac{1}{(\xi \!-\! \xi' \!+\! \omega)(\xi' \!-\! \omega)}
\biggr]
.
\label{eq:J1p-def}\end{aligned}$$
In practice, we also found that the principal-value integral around the singularity at $\xi'=\pm\omega$ is numerically unstable and its convergence is difficult to achieve with a practical computational cost. We therefore introduce an even smoothing function $p(\xi'\pm\omega)$ whose value is unity for $|\xi'\pm\omega| \gtrsim T$ and of order $O[(\xi'\pm\omega)^{2}]$ for $|\xi'\pm\omega|\ll T$. Substituting $p(\xi'\pm\omega)/(\xi'\pm\omega)$ for ${\rm P}[1/(\xi'\pm\omega)]$ and rearranging the terms yield[@comment-subtract] $$\begin{aligned}
\mathcal{Z}^{\rm ph,new}(\xi)
&=&
\frac{1}{{\rm tanh}[(\beta/2)\xi]}
\!\!\int \!\! d\omega
\alpha^{2}\!F(\omega) \!\!
\int \!\! d\xi'
\frac{N(\xi')}{N(0)}
\bigl[
\mathcal{I}
\!-\! 2\mathcal{J}
\bigr]
, \nonumber \\
\label{eq:Z-ave-new}
\\
\mathcal{I}(\xi\!,\xi'\!\!,\omega)
&=&
\tilde{\mathcal{I}}(\xi\!,\xi'\!\!,\omega)
\!-\!
\tilde{\mathcal{I}}(\xi\!,\xi'\!\!,\!-\omega)
\nonumber \\
&& \hspace{40pt}
\!-\!
\tilde{\mathcal{I}}(\xi\!,\!-\xi'\!\!,\omega)
\!+\!
\tilde{\mathcal{I}}(\xi\!,-\xi'\!\!,\!-\omega)
,\label{eq:mathcalI-def}
\\
\tilde{\mathcal{I}}(\xi\!,\xi'\!\!,\omega)
&=&
[f_{\beta}(\xi)\!+\!n_{\beta}(\omega)]\frac{f_{\beta}(\xi')\!\!-\!\!f_{\beta}(\xi\!-\!\omega)}{\xi\!-\!\xi'\!-\!\omega}
\frac{p(\xi'\!\!+\!\omega)}{\xi'\!\!+\!\omega}
,\label{eq:tildeI-def}
\\
\mathcal{J}(\xi\!,\xi'\!\!,\omega)
&=&
\tilde{\mathcal{J}}(\xi\!,\xi'\!\!,\omega)
\!-\!
\tilde{\mathcal{J}}(\xi\!,\xi'\!\!,-\omega)
,\label{eq:mathcalJ-def}
\\
\tilde{\mathcal{J}}(\xi\!,\xi'\!\!,\omega)
&=&
-
\frac{f_{\beta}(\xi)\!+\!n_{\beta}(\omega)}{\xi\!-\!\xi'\!-\!\omega}
p(\xi'\!\!+\!\omega)
\biggl[
\frac{f_{\beta}(\xi')\!-\!f_{\beta}(\xi\!-\!\omega)}{\xi\!-\!\xi'\!-\!\omega}
\nonumber \\
&&\hspace{40pt}
-
\beta f_{\beta}(\xi\!-\!\omega)f_{\beta}(\!-\xi\!+\!\omega)
\frac{\xi}{\xi'\!+\!\omega}
\biggr]
.\label{eq:tildeJ-def}\end{aligned}$$ We have used $p(x)=[{\rm tanh}(500\beta x)]^{4}$ in this paper, though the calculated value is not sensitive to a specfic choice of $p(x)$.
With the above form, the asymmetry effect is properly treated in a numerically stable manner; the smoothness of the integrands at $\xi'=\pm\xi\pm\omega, \pm \omega$, and $\xi=0$ is retained. In addition, for approximately symmetric DOS, the calculated results are guaranteed to be quite close to those with the $n{\bf k}$-averaged form of Eq. (\[eq:Z-ph-prev\]) $$\begin{aligned}
\mathcal{Z}^{\rm ph}(\xi)
&=&
-\frac{1}{{\rm tanh}[(\beta/2) \xi]}
\int d\omega \alpha^{2}F(\omega)
\int d\xi'
\frac{N(\xi')}{N(0)}
\nonumber \\
&& \!
\times \!
\bigl[
J\!(\xi,\xi',\omega)\!
+\!
J\!(\xi,\!-\xi',\omega)
\bigr]
\label{eq:Z-ph-ave-prev}\end{aligned}$$ because of the following properties: (i) $\lim_{\xi \rightarrow 0}[\mathcal{Z}^{\rm ph, new}(\xi)-\mathcal{Z}^{\rm ph}(\xi)]\simeq 0$, and particularly for constant DOS, the desirable behavior[@GrossI] $\lim_{\xi \rightarrow 0}\mathcal{Z}^{\rm ph,new}(\xi)\simeq\lim_{\xi \rightarrow 0}\mathcal{Z}^{\rm ph}(\xi) \simeq \lambda$ is satisfied, (ii) temperature dependence of the calculated value is similar to that of $\mathcal{Z}^{\rm ph}(\xi)$, and (iii) the symmetric part of the integrand, that is, $$\begin{aligned}
\frac{
\mathcal{I}(\xi,\xi' \!\!,\omega)
\!+\!
\mathcal{I}(\xi,\!-\xi' \!\!,\omega)
\!-\!
2\mathcal{J}(\xi,\xi' \!\!,\omega)
\!-\!
2\mathcal{J}(\xi,\!-\xi' \!\!,\omega)
}{2}
\label{eq:sym-part}\end{aligned}$$ agrees with $J\!(\xi,\xi',\omega)\!+\!J\!(\xi,\!-\xi',\omega)$ in Eq. (\[eq:Z-ph-ave-prev\]) when we impose a condition $\omega$$\gtrsim$$|\xi|$$\gg$$T$. These properties are demonstrated in the following calculations. Further, we have found that the $n{\bf k}$-resolved counterpart of Eq. (\[eq:Z-ave-new\]) given by $$\begin{aligned}
\!\!\!\!\!\!\!\!
\mathcal{Z}^{\rm ph, new}_{n{\bf k}}\!\!
&=&
\frac{1}{{\rm tanh}[(\beta/2)\xi]}
\sum_{n'{\bf k}' \nu}
|g^{{\bf k}-{\bf k}'\nu}_{n{\bf k},n'{\bf k}'}|^{2}
\nonumber \\
&& \hspace{-10pt}
\times
\bigl[
\mathcal{I}\!(\xi_{n{\bf k}},\xi_{n'{\bf k}'}\!,\omega_{\nu{\bf k}-{\bf k}'}\!)
\!-\!
2\mathcal{J}\!(\xi_{n{\bf k}},\xi_{n'{\bf k}'}\!,\omega_{\nu{\bf k}-{\bf k}'}\!)
\bigr]
\label{eq:Z-new}\end{aligned}$$ is also numerically stable, which is a reasonable generalization of Eq. (\[eq:Z-ph-prev\]). Equations (\[eq:Z-ave-new\]) and (\[eq:Z-new\]) are those we propose as the new forms of $\mathcal{Z}$ for particle-hole asymmetry.
Asymmetry effect on $T_{\rm c}$ {#sec:calc}
===============================
In the rest of the present paper we get insights into the asymmetry effect included in the present improvement. In Sec. \[subsec:analytic\], we analytically construct a formula of $T_{\rm c}$ to discuss in what situation the asymmetry significantly affects $T_{\rm c}$. In Sec. \[subsec:modelcalc\], a typical model case is presented where our new kernel and the previous one give quite different $T_{\rm c}$, and some remarks on the application to elemental metals are given. We here note that we focus on the $n{\bf k}$-averaged form \[Eq. (\[eq:Z-ave-new\])\], and thereby the present scope includes only the asymmetry of DOS, not that of the electron-phonon matrix elements.
Analytic calculation {#subsec:analytic}
--------------------
We first analytically solve the energy-averaged gap equation[@GrossI] $$\begin{aligned}
\Delta(\xi)
&=&
-\mathcal{Z}(\xi)\Delta(\xi)
\nonumber \\
&& \hspace{5pt}
-\frac{1}{2}
\int \!\! d\xi' \!
N(\xi') \mathcal{K}(\xi\!,\xi')\frac{{\rm tanh}[(\beta/2)\xi']}{\xi'}\Delta(\xi')
,
\label{eq:gap-eq-ave}\end{aligned}$$ with the electron-phonon kernels only. We introduce a simple model for which DOS and kernels are given by the constant part and the antisymmetric part, given as $\mathcal{Z}(\xi)=\mathcal{Z}_{\rm c}(\xi) +\mathcal{Z}_{\rm a}(\xi)$, $N(\xi)=N_{\rm c}(\xi) +N_{\rm a}(\xi)$, $\mathcal{K}(\xi, \xi')=\mathcal{K}_{\rm c}(\xi,\xi') +\mathcal{K}_{\rm a}(\xi,\xi')$, with each part defined within the Debye frequency $\omega_{\rm D}$ by $$\begin{aligned}
\mathcal{Z}_{\rm c}(\xi)
&=&
\!Z_{\rm c}
, \ \
\mathcal{Z}_{\rm a}(\xi)
\!=\!
\left\{
\begin{array}{ll}
\!-\!Z_{\rm a}, & \ ( \xi \!\leq\!-\omega' ) \\
0, & \ (|\xi| \!<\! \omega') \\
Z_{\rm a}, & \ ( \xi \!\geq\! \omega') \\
\end{array}
\right.
\label{eq:Z-McM}
\\
N_{\rm c}(\xi)
&=&
\!N_{\rm c}
,
N_{\rm a}(\xi)
\!=\!
\left\{
\begin{array}{ll}
\!-\!N_{\rm a}, & \ ( \xi \!\leq\!-\omega' ) \\
0, & \ (|\xi| \!<\! \omega') \\
N_{\rm a}, & \ (\xi \!\geq\! \omega') \\
\end{array}
\right.
\label{eq:N-McM}
\\
\mathcal{K}_{\rm c}(\xi,\xi')
&=&
K_{\rm c}
, \nonumber \\
\mathcal{K}_{\rm a}(\xi,\xi')
\!&=&\!
\left\{
\begin{array}{ll}
\!-\! K_{\rm a}, & \ \bigl[( \xi \leq -\omega' \ \ {\rm and} \ \ \xi'\leq -\omega' ) \\
& \ \ {\rm or} \ \ ( \xi \geq \omega' \ \ {\rm and} \ \ \xi' \geq \omega') \bigr] \\
K_{\rm a}, & \ \bigl[(\xi\leq -\omega' \ \ {\rm and} \ \ \xi' \geq \omega') \\
& \ \ {\rm or} \ \ ( \xi \geq \omega' \ \ {\rm and} \ \ \xi'\leq -\omega' ) \bigr] \\
0, & \ \bigl(|\xi| < \omega' \ \ {\rm or}\ \ |\xi'| < \omega' \bigr) \\
\end{array}
\right.
.
\label{eq:K-McM}\end{aligned}$$ Here, in addition to $\omega_{\rm D}$, we have also introduced $\omega'$ which specifies the energy scale where DOS and kernels substantially deviate from the values at $E_{\rm F}$. These forms are depicted in Fig. \[fig:terms\].
![(Color online) Schematic description of terms defined by Eqs. (\[eq:Z-McM\])–(\[eq:K-McM\]).[]{data-label="fig:terms"}](terms_130501.eps){width="8cm"}
Similarly to McMillan’s procedure,[@McMillan] we assume the following rectangular form for the solution $$\begin{aligned}
\Delta(\xi)
=
\left\{
\begin{array}{ll}
\Delta_{-}, & \ \ (-\omega' \geq \xi \geq -\omega_{\rm D}) \\
\Delta_{0}, & \ \ (|\xi| < \omega') \\
\Delta_{+}, & \ \ (\omega_{\rm D} \geq \xi \geq \omega') \\
\end{array}
\right.
.
\label{eq:gap-McM}\end{aligned}$$ The condition that a nonzero solution for $\Delta$ exists gives $T_{\rm c}$. By retaining the lowest order term with respect to the values with subscript “a", we obtain $$\begin{aligned}
\!\!
T_{\rm c}
&\propto&
{\rm exp}\!
\left\{\!
\frac{1\!\!+\!\!Z_{\rm c}}{K_{\rm c}\!N_{\rm c}}
\!-\!
\frac{N_{\rm a}\!Z_{\rm a}{\rm ln}\!\left[\omega_{\rm D}/\omega'\right]}{N_{\rm c}(1+Z_{\rm c})}
\!+\!
\frac{Z_{\rm a}^{2}{\rm ln}\!\left[\omega_{\rm D}/\omega'\right]}{(1+Z_{\rm c})^{2}}\!
\right\}
.
\label{eq:Tc-McMillan}\end{aligned}$$ Since $Z_{\rm a}$ is dependent on $N_{\rm a}$ through Eq. (\[eq:Z-ave-new\]), we can substitute an approximate form for $Z_{\rm a}$ in terms of $N_{\rm a}/N_{\rm c}$, and then we obtain $$\begin{aligned}
T_{\rm c}
&\propto&
{\rm exp}
\left\{\!
-
\frac{1+\lambda}{\lambda}
-
\left(
\frac{N_{\rm a}}{N_{\rm c}}
\right)^{2}
\Lambda(r, \lambda)
{\rm ln}r
\right\}
.
\label{eq:Tc-McMillan2}\end{aligned}$$ Here, $r\equiv\omega_{\rm D}/\omega'$, and we have used the following relation $K_{\rm c}N_{\rm c}=-\lambda$ and $Z_{\rm c}=\lambda$ (Ref. ). The first term in the bracket corresponds to the zeroth order, which appears in literatures.[@McMillan] The function $\Lambda(r,\lambda)$ is a positive function which monotonically increases by increasing $r$ or $\lambda$ and converges to a finite value $<1$ in the limit $r \rightarrow \infty$ and $\lambda \rightarrow \infty$. The detail of the derivation of Eqs. (\[eq:Tc-McMillan\]) and (\[eq:Tc-McMillan2\]) is given in Appendix \[app:McM\].
From Eq. (\[eq:Tc-McMillan2\]), we see how the antisymmetric part affects $T_{\rm c}$: It reduces $T_{\rm c}$ regardless of the sign of $N_{\rm a}$ and its amount becomes substantial when the ratios $N_{\rm a}/N_{\rm c}$ or $r\equiv\omega_{\rm D}/\omega'$ are large. In addition, we find the antisymmetric part of the nondiagonal kernel ($K_{\rm a}$) does not contribute to $T_{\rm c}$ within the lowest order.
Numerical calculation {#subsec:modelcalc}
---------------------
The above analysis tells us that the asymmetry effect becomes pronounced for the cases where $N_{\rm a}/N_{\rm c}$ and $r$ are large. To quantify the effect more explicitly, we calculated $T_{\rm c}$ by numerically solving Eq. (\[eq:gap-eq-ave\]) using a model DOS. We included both the phonon and electron contribution to the kernels, $\mathcal{K}
=
\mathcal{K}^{\rm ph}
+
\mathcal{K}^{\rm el}
$ and $\mathcal{Z}=\mathcal{Z}^{\rm ph}$. We employed the $n{\bf k}$-averaged form for $\mathcal{K}^{\rm ph}$ \[Eq. (23) in Ref. \], and we treated $\mathcal{K}^{\rm el}$ by an approximate form $$\begin{aligned}
\mathcal{K}^{\rm el}(\xi,\xi')
=
\frac{\mu}{N(0)}\end{aligned}$$ with a certain energy cutoff. For $\mathcal{Z}^{\rm ph}$, we employed the previous form \[Eq. (\[eq:Z-ph-ave-prev\])\] and the new one \[Eq. (\[eq:Z-ave-new\])\].
We introduced a step-like DOS as depicted in the upper panel of Fig. \[fig:Tc-ratio\] so that $N_{\rm a}/N_{\rm c}$ and $r$ can be large. This form is characterized by two parameters; the ratio of the high-energy and low-energy values $N_{+}/N_{-}$, and the energy range where DOS changes $d$. For the Eliashberg function, we used the following model function $$\begin{aligned}
\alpha^{2}F(\omega)
=
\frac{\lambda}{2\sigma \sqrt{\pi}}
\omega
{\rm exp}
\left[
-\biggl(\frac{\omega-\omega_{\rm D}}{\sigma}\biggr)^{2}
\right]
\label{eq:a2F-model}\end{aligned}$$ with width $\sigma=\omega_{\rm D}/10$. We set $\omega_{\rm D}=400$\[cm$^{-1}$\], $d=\omega_{\rm D}/2$, and $N_{+}/N_{-}=6$. This setting is realistic in that one often see a similar dependence in doped semiconductors with quasi-2D Fermi surfaces.[@Felser-MNCl; @Akashi-MNCl; @Iwasa-MoS2]
With these settings, we calculated $T_{\rm c}$ for various settings of $E_{\rm F}$. The calculation was performed with a logarithmically uniform energy grid, where the minimum energy cutoff was set to $|\xi|=10^{-6}$ eV and the number of points per digit was 20. The energy cutoff for $\mathcal{K}^{\rm el}$ was fixed to 20eV. The parameters $\lambda$ and $\mu$ were fixed to $1.0$ and $0.5$ for all the setting of $E_{\rm F}$ so that the resulting $T_{\rm c}$ becomes $\gtrsim$ $10$K.
In the lower panel of Fig. \[fig:Tc-ratio\], we plot the ratio of the calculated $T_{\rm c}$ using $\mathcal{Z}^{\rm ph, new}(\xi)$ \[Eq. (\[eq:Z-ave-new\])\] to that using $\mathcal{Z}^{\rm ph}(\xi)$ \[Eq. (\[eq:Z-ph-ave-prev\])\] as a function of $E_{\rm F}$, where the absolute values of $T_{\rm c}$ are given in the inset.[@note-Tc] The ratio is systematically less than unity; the effect of asymmetry lowers $T_{\rm c}$ for any $E_{\rm F}$, which is consistent with Eq. (\[eq:Tc-McMillan2\]). As the Fermi level approaches to the lower edge of the slope, the decrease of the ratio becomes significant and finally amounts to more than 20%. When the Fermi level is located on the upper edge of the slope, the decrease is not appreciable. This $E_{\rm F}$ dependence can also be consistently understood with Eq. (\[eq:Tc-McMillan2\]); the characteristic energy scale ($\omega'$) decreases as the Fermi level approaches to the slope, and the ratio $N_{\rm a}/N_{\rm c}$ is relatively small when the Fermi level is on the upper edge. Thus, the change of $T_{\rm c}$ by the present improvement becomes crucial when $E_{\rm F}$ is close to a point where DOS increases rapidly.
![(Color online) (Bottom) Ratio of $T_{\rm c}$ calculated using $\mathcal{Z}^{\rm ph, new}$ \[Eq. (\[eq:Z-ave-new\])\] ($T_{\rm c}^{\rm asym}$) to that calculated using $\mathcal{Z}^{\rm ph}$ \[Eq. (\[eq:Z-ph-ave-prev\])\] ($T_{\rm c}^{\rm sym}$) plotted as a function of $E_{\rm F}$. In the inset, absolute values of $T_{\rm c}^{\rm asym}$ and $T_{\rm c}^{\rm sym}$ are shown. (Top) DOS used for the calculations. We set $N(\xi)$$=$$N(E_{\rm F})$ for $|\xi|>5$eV.[]{data-label="fig:Tc-ratio"}](Model_step_Tc_rate_dos_combine_variousEF_130501.eps){width="8.5cm"}
![(Color online) Calculated values of $\mathcal{Z}^{\rm ph, new}(\xi)$ (red solid line) and $\mathcal{Z}^{\rm ph}(\xi)$ (blue dotted line) for $E_{\rm F}$$=$$-0.3\omega_{\rm D}$. The green shaded area depicts DOS (arb. unit.) in the same energy scale. Note that the values for $T=T^{\rm asym}_{\rm c}$ (thin lines) are multiplied by 0.5.[]{data-label="fig:compare-Z"}](model_step_Z_ed_130502.eps)
We show the calculated value of $\mathcal{Z}^{\rm ph}$ for $E_{\rm F}$$=$$-0.3\omega_{\rm D}$ in Fig. \[fig:compare-Z\] to examine the properties of the present form. The model DOS is also depicted in the same energy scale (green shaded area). For both $T$$=$$0.01$K (thick lines) and $T$$\simeq$$T^{\rm asym}_{\rm c}$ (thin lines), $\mathcal{Z}^{\rm ph, new}(\xi)$ becomes asymmetric in $\xi$, but they are similar to $\mathcal{Z}^{\rm ph}(\xi)$. In particular, the limiting value for $\xi \rightarrow 0$ agrees with $\lim_{\xi \rightarrow 0}\mathcal{Z}^{\rm ph}(\xi)$. These features well represent the properties (i)–(iii) described in the last paragraph of Sec. \[subsec:derive-Z-new\]. We observe a difference between $\mathcal{Z}^{\rm ph,new}(\xi)$ and $\mathcal{Z}^{\rm ph}(\xi)$ around $|\xi|\gtrsim \omega_{\rm D}$, which is responsible for the difference of $T^{\rm asym}_{\rm c}$ and $T^{\rm sym}_{\rm c}$ in Fig. \[fig:Tc-ratio\]. Clearly, $\mathcal{Z}^{\rm ph, new}(\xi)$ becomes larger for the energy region with larger density of states, whereas $\mathcal{Z}^{\rm ph}(\xi)$ does not. The dependence of the former is reasonable because larger DOS should result in stronger mass renormalization due to stronger total electron-phonon coupling. Thus we were able to see that the particle-hole asymmetric electronic structure is properly treated with $\mathcal{Z}^{\rm ph,new}$ in a numerically stable manner.
Finally, let us comment on the application to actual superconductors. For typical superconductors, aluminum and niobium, we have carried out stable calculations of $T_{\rm c}$ (see appendix \[app:abinitio\]), confirming again that the listed properties (i)–(iii) of $\mathcal{Z}^{\rm ph,new}$ are valid. The asymmetry effect on $T_{\rm c}$ was, however, estimated to be less than 0.1 % for the both systems. This is mainly due to the small $N_{\rm a}/N_{\rm c}$ in these systems. The present weak asymmetry effect gives support to the validity of the particle-hole symmetrizing treatment[@Schrieffer; @Mitrovic; @Pickett-general; @GrossI] for many cases of conventional superconductors.
Summary and Conclusions {#sec:summary}
=======================
We constructed a new exchange-correlation kernel $\mathcal{Z}^{\rm ph,new}$ in SCDFT to treat the particle-hole asymmetry of electronic structure. The obtained $n{\bf k}$-averaged forms \[Eq. (\[eq:Z-ave-new\])\] and $n{\bf k}$-resolved form \[Eq. (\[eq:Z-new\])\] do not show any divergences or instabilities, and, for the systems with good symmetry, well agree with the previous symmetrized forms. By analytically deriving $T_{\rm c}$ formula, we found that the asymmetry systematically decreases $T_{\rm c}$. The amount of the decrease becomes substantial when the following two ratios are large: The ratio of the antisymmetic component of DOS to the constant component, and that of the Debye frequency to the characteristic energy scale of the DOS variation. We also calculated $T_{\rm c}$ with a model step-like DOS, showing that the amount of the reduction of $T_{\rm c}$ can be more than 20%. With the present work, we successfully reinforced the theoretical foundation of SCDFT to extend its applicability to systems with significant particle-hole asymmetric electronic structure.
The authors thank Kazuma Nakamura, Shiro Sakai, Hideyuki Miyahara, and Jianting Ye for enlightening comments. This work was supported by Grants-in-Aid for Scientic Research from JSPS (No. 23340095), Funding Program for World-Leading Innovative R & D on Science and Technology (FIRST program) on Quantum Science on Strong Correlation", JST-PRESTO and the Next Generation Super Computing Project and Nanoscience Program from MEXT, Japan.
Evaluation of $\mathcal{Z}^{\rm ph}_{2}(\xi)$ for nonconstant DOS {#app:eval-Z2}
=================================================================
We here describe the detail of evaluation of $\mathcal{Z}^{\rm ph}_{2}(\xi)$ \[Eq. (\[eq:Z2-nonconst\])\] for the nonconstant DOS. Under the decomposition approximation of DOS introduced in Sec. \[subsec:cancel\] \[Fig. \[fig:DOS-general\], Eq. (\[eq:DOS-step\])\], $\mathcal{Z}^{\rm ph,2}(\xi)$ is given by $$\begin{aligned}
&&
\mathcal{Z}^{\rm ph,2}(\xi)
\nonumber \\
&&=
-
\biggl\{
\frac{1}{\xi}
-
\frac{\beta/2}{{\rm cosh}[(\beta/2)\xi]{\rm sinh}[(\beta/2)\xi]}
\biggr\}
\nonumber \\
&& \hspace{5pt} \times \!\!
\int \!\! d\omega
\alpha^{2}F(\omega)
\int^{L_{1}}_{-L_{2}}d\xi
\biggl\{
1 \!+\!
\sum_{i}[
N^{-}_{i}\theta^{-}_{i}(\xi)
\!+\!
N^{+}_{i}\theta^{+}_{i}(\xi)
]
\biggr\}
\nonumber \\
&& \hspace{10pt}
\times \!\!
\int^{L_{1}}_{-L_{2}}\!\! d\xi'
\biggl\{ \!
1 \!+\!
\sum_{j}[
N^{-}_{j}\theta^{-}_{j}(\xi')
\!+\!
N^{+}_{j}\theta^{+}_{j}(\xi')
]
\biggr\}
I'(\xi,\xi',\omega)
.\nonumber \\\end{aligned}$$ With omitting the temperature-dependent terms of order $O[(T/\epsilon_{i})^{2}]$, $O(e^{-\beta \epsilon_{i}})$, $O(e^{-\beta L_{1}})$ and $O(e^{-\beta L_{2}})$, the integral for each term is carried out as $$\begin{aligned}
&&
\int^{L_{1}}_{-L_{2}} d\xi
\int^{L_{1}}_{-L_{2}} d\xi'
\theta^{\pm}_{i}(\xi)
\theta^{\pm}_{j}(\xi')
I'(\xi,\xi',\omega)
\nonumber \\
&&\simeq
\pm
{\rm ln}
\left|
\frac{(L_{1} \!+\! \epsilon_{i} \!+\! \omega)(L_{2} \!+\! \epsilon_{j} \!+\! \omega)}{(L_{1} \!+\! \epsilon_{j} \!+\! \omega)(L_{2} \!+\! \epsilon_{i} \!+\! \omega)}
\right|
,
\\
&&
\int^{L_{1}}_{-L_{2}} d\xi
\int^{L_{1}}_{-L_{2}} d\xi'
\theta^{\pm}_{i}(\xi)
\theta^{\mp}_{j}(\xi')
I'(\xi,\xi',\omega),
\nonumber \\
&&\simeq
\pm
{\rm ln}
\left|
\frac
{(L_{1} \!+\! \epsilon_{i}\!+\! \omega)(L_{1}\!+\! \epsilon_{j} \!+\! \omega)
(L_{2} \!+\! \epsilon_{i} \!+\! \omega)(L_{2} \!+\! \epsilon_{j} \!+\! \omega)}
{(L_{1} \!+\! L_{2} \!+\! \omega)^{2}(\epsilon_{i} \!+\! \epsilon_{j} \!+\! \omega)^{2}}
\right|
,
\nonumber \\
\\
&&
\int^{L_{1}}_{-L_{2}} d\xi
\int^{L_{1}}_{-L_{2}} d\xi'
\theta^{-}_{i}(\xi)
I'(\xi,\xi',\omega),
\nonumber \\
&&\simeq
{\rm ln}
\left|
\frac
{(L_{1} \!+\! L_{2} \!+\! \omega)^{2}(\epsilon_{i} \!+\! \omega)^{2}}
{(L_{1} \!+\! \omega)(L_{1} \!+\! \epsilon_{i} \!+\! \omega)
(L_{2} \!+\! \omega)(L_{2} \!+\! \epsilon_{i}\!+\! \omega)}
\right|
,
\\
&&
\int^{L_{1}}_{-L_{2}} d\xi
\int^{L_{1}}_{-L_{2}} d\xi'
\theta^{-}_{i}(\xi')
I'(\xi,\xi',\omega),
\nonumber \\
&&\simeq
{\rm ln}
\left|
\frac
{(L_{1} \!+\!\epsilon_{i} \!+\! \omega)
(L_{2} \!+\! \epsilon_{i} \!+\! \omega)}
{(L_{1} \!+\! L_{2} \!+\! \omega)^{2}}
\right|
,
\\
&&
\int^{L_{1}}_{-L_{2}} d\xi
\int^{L_{1}}_{-L_{2}} d\xi'
\theta^{+}_{i}(\xi)
I'(\xi,\xi',\omega),
\nonumber \\
&&\simeq
{\rm ln}
\left|
\frac
{(L_{1} \!+\! \epsilon_{i} \!+\! \omega)
(L_{2} \!+\! \omega)}
{(L_{2} \!+\! \epsilon_{i} \!+\! \omega)
(L_{1} \!+\! \omega)}
\right|
,
\\
&&
\int^{L_{1}}_{-L_{2}} d\xi
\int^{L_{1}}_{-L_{2}} d\xi'
\theta^{+}_{i}(\xi')
I'(\xi,\xi',\omega),
\nonumber \\
&&\simeq
{\rm ln}
\left|
\frac
{L_{2} \!+\! \epsilon_{i} \!+\! \omega}
{L_{1} \!+\! \epsilon_{i} \!+\! \omega}
\right|
.\end{aligned}$$ Using these relations, $\mathcal{Z}^{\rm ph}_{2}(\xi)$ is transformed to the same form as Eq. (\[eq:Z1-generalDOS\]) with the opposite sign.
Derivation of Eq. (\[eq:Tc-McMillan\]) and Eq. (\[eq:Tc-McMillan2\]) {#app:McM}
====================================================================
Let us first plug the kernels and the gap function given in Eqs. (\[eq:Z-McM\])–(\[eq:gap-McM\]) into the energy-averaged gap equation \[Eq. (\[eq:gap-eq-ave\])\]. The condition to have a nonzero solution is
$$\begin{aligned}
{\rm det} M&=&0, \nonumber \\
M
&\equiv&
\left[
\begin{array}{ccc}
1 \!+\! Z_{\rm c} \!-\! Z_{\rm a}+\frac{1}{2}(N_{\rm c} \!-\! N_{\rm a})(K_{\rm c} \!-\! K_{\rm a})q &\ \ N_{\rm c}K_{\rm c}p &\ \ \frac{1}{2}(N_{\rm c} \!+\! N_{\rm a})(K_{\rm c} \!+\! K_{\rm a})q \\
\frac{1}{2}(N_{\rm c} \!-\! N_{\rm a})K_{\rm c}q &\ \ 1 \!+\! Z_{\rm c} + N_{\rm c}K_{\rm c}p &\ \ \frac{1}{2}(N_{\rm c} \!+\! N_{\rm a})K_{\rm c}q \\
\frac{1}{2}(N_{\rm c} \!-\! N_{\rm a})(K_{\rm c} \!+\! K_{\rm a})q &\ \ N_{\rm c}K_{\rm c}p &\ \ 1 \!+\! Z_{\rm c} \!+\! Z_{\rm a} + \frac{1}{2}(N_{\rm c} \!+\! N_{\rm a})(K_{\rm c} \!-\! K_{\rm a})q
\end{array}
\right]
,\end{aligned}$$
with $p$ and $q$ defined by $$\begin{aligned}
p\!=\!\!\!
\int_{0}^{\omega'} \!\!\!\! d\xi'
\frac{{\rm tanh}[(\beta_{\rm c}/2)\xi']}{\xi'}
,
q\!=\!\!\!
\int_{\omega'}^{\omega_{\rm D}} \!\!\!\! d\xi'
\frac{{\rm tanh}[(\beta_{\rm c}/2)\xi']}{\xi'}
.\end{aligned}$$ Here $\beta_{\rm c}$ denotes the inverse of $T_{\rm c}$. In order to treat the lowest-order contribution, we keep the terms up to the second order with respect to the values with subscript “a" as $$\begin{aligned}
{\rm det}M
&=& 0 \nonumber \\
&\simeq&
(1 \!+\! Z_{\rm c})^{3}
+
[K_{\rm c}N_{\rm c}C \!-\! K_{\rm a}N_{\rm c}q](1 \!+\!Z_{\rm c})^{2}
\nonumber \\
&& \hspace{10pt}
+
[
-\! K_{\rm c}K_{\rm a}N_{\rm c}^{2}qC
\!-\!
K_{\rm c}N_{\rm a}Z_{\rm a}q
\!-\!
Z_{\rm a}^{2}
]
(1+Z_{\rm c})
\nonumber \\
&& \hspace{20pt}
-
K_{\rm c}N_{\rm c}Z_{\rm a}^{2}p
,\end{aligned}$$ where $C\equiv p+q$ has been introduced. Retaining the lowest-order terms, we obtain $$\begin{aligned}
C
&\simeq&
-
\frac{1+Z_{\rm c}}{K_{\rm c}N_{\rm c}}
+
\frac{N_{\rm a}Z_{\rm a}q}{N_{\rm c}(1+Z_{\rm c})}
-
\frac{Z_{\rm a}^{2}q}{(1+Z_{\rm c})^{2}}
.\end{aligned}$$ Using $
C={\rm ln}
\left[
2e^{\gamma}\omega_{\rm D} /(\pi T_{c})
\right]
$ with $\gamma$ being the Euler constant, we get Eq. (\[eq:Tc-McMillan\]).
Next we consider the relation between $Z_{\rm a}$ and $N_{\rm a}$. Let us start from our newly developed form \[Eq. (\[eq:Z-ave-new\])\]. The antisymmetric part of the integrand in $\xi'$ can be written as $$\begin{aligned}
&&
\tilde{\mathcal{I}}(\xi,\xi',\omega)
-
\tilde{\mathcal{I}}(\xi,\xi',-\omega)
-
\tilde{\mathcal{I}}(\xi,-\xi',\omega)
+
\tilde{\mathcal{I}}(\xi,-\xi',-\omega)
\nonumber \\
&&
\hspace{10pt}
-\!
\tilde{\mathcal{J}}(\xi,\xi',\omega)
\!+\!
\tilde{\mathcal{J}}(\xi,\xi',-\omega)
\!+\!
\tilde{\mathcal{J}}(\xi,-\xi',\omega)
\!-\!
\tilde{\mathcal{J}}(\xi,-\xi',-\omega)
.\label{eq:antisym-part}
\nonumber \\\end{aligned}$$ Using the facts $\tilde{\mathcal{I}}(-\xi,-\xi',-\omega)=\tilde{\mathcal{I}}(\xi,\xi',\omega)$ and $\tilde{\mathcal{J}}(-\xi,-\xi',-\omega)=\tilde{\mathcal{J}}(\xi,\xi',\omega)$, one can easily find the above part divided by ${\rm tanh}[(\beta/2) \xi]$ is antisymmetric in $\xi$. The symmetric part \[Eq. (\[eq:sym-part\])\] divided by ${\rm tanh}[(\beta/2) \xi]$ is, on the other hand, symmetric in $\xi$. Consequently, the antisymmetric part of $\mathcal{Z}^{\rm ph, new}(\xi)$ is yielded by only the antisymmetric part of $N(\xi')$.
Substituting $N(\xi')=N_{\rm c}(\xi')+N_{\rm a}(\xi')$, for the antisymmetric contribution, we obtain $$\begin{aligned}
&&\int d\xi'
\frac{N_{\rm a}(\xi')}{N_{0}}
[\mathcal{I}(\xi,\xi',\omega) \!-\! 2\mathcal{J}(\xi,\xi',\omega)]
\nonumber \\
&& \hspace{10pt}
=
2\frac{N_{\rm a}}{N_{\rm c}}
\int_{\omega'}^{\omega_{\rm D}}d\xi'
[(\textrm{antisym.})]
,\end{aligned}$$ where (antisym.) represents the terms in Eq. (\[eq:antisym-part\]). This integration is easily performed since we can assume $\xi \gg T$, $\xi' \gg T$, and $\omega \gg T$, so that we get $$\begin{aligned}
&&
2\frac{N_{\rm a}}{N_{\rm c}}
\int_{\omega'}^{\omega_{\rm D}}d\xi'
[(\textrm{antisym.})]
\nonumber \\
&& \simeq
2\frac{N_{\rm a}}{N_{\rm c}}
\int_{\omega'}^{\omega_{\rm D}}d\xi'
\frac{1}{(\xi+\xi'+\omega)^{2}}
\nonumber \\
&& =
2\frac{N_{\rm a}}{N_{\rm c}}
\left[
\frac{1}{\xi+\omega'+\omega}
-
\frac{1}{\xi+\omega_{\rm D}+\omega}
\right]
.\end{aligned}$$ Finally, by using the Einstein spectrum $\alpha^{2}F(\omega)=\frac{\lambda}{2}\omega_{\rm D}\delta(\omega-\omega_{\rm D})$, we obtain $$\begin{aligned}
&&
\frac{1}{2}
\left[
\mathcal{Z}^{\rm ph,new}(\xi) \!-\! \mathcal{Z}^{\rm ph,new}(-\xi)
\right]
\nonumber \\
&& \hspace{10pt}=
{\rm sgn}(\xi)
\lambda
\frac{N_{\rm a}}{N_{\rm c}}
\frac{\omega_{\rm D}(\omega_{\rm D}-\omega')}{(\xi \!+\! \omega' \!+\! \omega_{\rm D})(\xi \!+\! 2\omega_{\rm D})}
,
\label{eq:z-asym-def}\end{aligned}$$ where the sign function comes from $\lim_{\beta \rightarrow \infty}$ $1/{\rm tanh}[(\beta/2)\xi]$.
Subsequently, we consider to approximate this form as the rectangular form \[Eq. (\[eq:Z-McM\])\]. Using the value at $\xi=\omega_{\rm D}$ yields $$\begin{aligned}
Z_{\rm a}
&=&
\lambda \frac{N_{\rm a}}{N_{\rm c}}
\frac{r-1}{3(2r+1)}
\nonumber \\
&\equiv&
\lambda \frac{N_{\rm a}}{N_{\rm c}}
h(r)
,\end{aligned}$$ where $r$$=$$\omega_{\rm D}/\omega'$$>$$1$. By using this form in Eq. (\[eq:Tc-McMillan\]) and substituting $K_{\rm a}N_{\rm c}=-\lambda$, $Z_{\rm c}=\lambda$, we get to $$\begin{aligned}
T_{\rm c}
&\propto&
{\rm exp}
\Biggl\{
-\frac{1\!+\!\lambda}{\lambda}
\nonumber \\
&&
-
\left(
\frac{N_{\rm a}}{N_{\rm c}}
\right)^{2} \!\!
\left[
\frac{\lambda h(r)}{1 \!+\! \lambda}
\!-\!
\left(
\frac{\lambda h(r)}{1\!+\! \lambda}
\right)^{2}
\right]
{\rm ln}r
\Biggr\}
.
\label{eq:Tc-appendix}\end{aligned}$$ This expression is equivalent to Eq. (\[eq:Tc-McMillan2\]), where the function $\Lambda(r, \lambda)$ is defined as the part in the square bracket in Eq. (\[eq:Tc-appendix\]). Its positiveness, monotonic dependences and the convergence in the limit $r \rightarrow \infty$ and $\lambda \rightarrow \infty$ referred in Sec. \[subsec:analytic\] are easily confirmed with this expression. We also note that $\Lambda(r, \lambda)$ has these properties for any values of $\xi$ ($\omega_{\rm D}$$>$$\xi$$>$$\omega'$) in Eq. (\[eq:z-asym-def\]) when we get an approximate form of $Z_{\rm a}$.
[*Ab initio*]{} calculation {#app:abinitio}
===========================
![(Color online) Calculated DOS for Al \[(a)\] and Nb \[(b)\]. $\mathcal{Z}$ kernels for Al \[(c)\] and Nb \[(d)\] calculated from Eq. (\[eq:Z-ave-new\]) (solid line) and Eq. (\[eq:Z-ph-ave-prev\]) (dashed line): The input Eliashberg function (dotted line) and DOS (green shaded area) are given in the same energy scale, with which $\lambda$ and $\omega_{\rm ln}$ were estimated for Al (Nb) as 0.417 (1.267) and 314 K (176 K).[]{data-label="fig:Al-Nb-Z"}](Al_Nb_Z_sym_asym_a2F_DOS_ed_130502.eps)
We applied our formalism to typical weak- and strong-coupling superconductors, Al and Nb. We solved the gap equation \[Eq. (\[eq:gap-eq\])\] with the energy-averaged approximation for the phonon kernels $$\begin{aligned}
\mathcal{K}_{n{\bf k},n'{\bf k}'}
&=&
\mathcal{K}^{\rm ph}(\xi_{n{\bf k}},\xi_{n'{\bf k}'})
+
\mathcal{K}^{\rm el}_{n{\bf k},n'{\bf k}'}
,
\\
\mathcal{Z}_{n{\bf k}}
&=&
\mathcal{Z}(\xi_{n{\bf k}})
,\end{aligned}$$ where $\mathcal{K}^{\rm ph}(\xi_{n{\bf k}},\xi_{n'{\bf k}'})$ and $\mathcal{K}^{\rm el}_{n{\bf k},n'{\bf k}'}$ are defined by Eq. (23) in Ref. and Eq. (13) in Ref. , respectively. For the diagonal part $\mathcal{Z}(\xi_{n{\bf k}})$, we employed $\mathcal{Z}^{\rm ph, new}(\xi_{n{\bf k}})$ \[Eq. (\[eq:Z-ave-new\])\] and $\mathcal{Z}^{\rm ph}(\xi_{n{\bf k}})$ \[Eq. (\[eq:Z-ph-ave-prev\])\]. We used an accurate random-sampling scheme given in Ref. , and the resulting sampling error was not more than a few percent. The detailed condition of the calculations is given in Ref. .
In Fig. \[fig:Al-Nb-Z\], the calculated DOS for Al \[(a)\] and Nb \[(b)\] are shown. The calculated values of $\mathcal{Z}_{n{\bf k}}$ with the two forms are also shown in panel (c) and (d) together with the input Eliashberg function $\alpha^{2}F$. For the both systems, the DOS (green shaded area) has small but nonzero antisymmetric component in the corresponding energy scale. However, the two forms of $\mathcal{Z}$ yield approximately the same value. For the calculated $T_{\rm c}$, regardless of the form of $\mathcal{Z}_{n{\bf k}}$, we obtained 0.7 and 8.7 K for Al and Nb, respectively. Although the gap values calculated with $\mathcal{Z}^{\rm ph, new}(\xi_{n{\bf k}})$ were approximately 0.1% smaller than those calculated with $\mathcal{Z}^{\rm ph}(\xi_{n{\bf k}})$ for each set of the sampling points, this difference was within the sampling error. The present result indicates that the asymmetry effect is not significant for $T_{\rm c}$ in Al and Nb.
[999]{} W. L. McMillan, Phys. Rev. [**167**]{}, 331 (1968). P. B. Allen and R. C. Dynes, Phys. Rev. B [**12**]{}, 905 (1975). P. Morel and P. W. Anderson, Phys. Rev. **125**, 1263 (1962); N. N. Bogoliubov, V. V. Tolmachev, and D. V. Shirkov, [*A New Method in the Theory of Superconductivity*]{} (1958) (translated from Russian: Consultants Bureau, Inc., New York, 1959). S. Baroni, S. deGironcoli, A. Dal Corso, and P. Giannozzi, Rev. Mod. Phys. **73**, 515 (2001). K. Kunc and R. M. Martin, in *Ab initio Calculation of Phonon Spectra*, edited by J. T. Devreese, V. E. van Doren, and P. E. van Camp (Plenum, New York, 1983), p. 65. K. H. Lee, K. J. Chang, and M. L. Cohen, Phys. Rev. B **52**, 1425 (1995). L. N. Oliveira, E. K. U. Gross, and W. Kohn, Phys. Rev. Lett. **60**, 2430 (1988). T. Kreibich and E. K. U. Gross, Phys. Rev. Lett. **86**, 2984 (2001). M. Lüders, M. A. L. Marques, N. N. Lathiotakis, A. Floris, G. Profeta, L. Fast, A. Continenza, S. Massidda, and E. K. U. Gross, Phys. Rev. B **72**, 024545 (2005). A. B. Migdal, Sov. Phys. JETP **7**, 996 (1958). G. M. Eliashberg, Sov. Phys. JETP **11**, 696 (1960). D. J. Scalapino, in [*Superconductivity*]{} edited by R. D. Parks, (Marcel Dekker, New York, 1969) VOLUME 1. J. R. Schrieffer,[*Theory of superconductivity; Revised Printing*]{}, (Westview Press, Colorado, 1971) M. A. L. Marques, M. Lüders, N. N. Lathiotakis, G. Profeta, A. Floris, L. Fast, A. Continenza, E. K. U. Gross, and S. Massidda, Phys. Rev. B **72**, 024546 (2005). A. Floris, G. Profeta, N. N. Lathiotakis, M. Lüders, M. A. L. Marques, C. Franchini, E. K. U. Gross, A. Continenza, and S. Massidda, Phys. Rev. Lett. **94**, 037004 (2005). A. Sanna, G. Profeta, A. Floris, A. Marini, E. K. U. Gross, and S. Massidda, Phys. Rev. B **75**, 020511(R) (2007). C. Bersier, A. Floris, A. Sanna, G. Profeta, A. Continenza, E. K. U. Gross, and S. Massidda, Phys. Rev. B **79**, 104503 (2009). L. R. Testardi, Rev. Mod. Phys. **47**, 637 (1975) K. M. Ho, M. L. Cohen, and W. E. Pickett, Phys. Rev. Lett. **41**, 815 (1978). W. E. Pickett, Phys. Rev. B **21**, 3897 (1980); **26**, 1186 (1982); W. E. Pickett and B. M. Klein, Solid State Commun. **38**, 95 (1981). S. G. Lie and J. P. Carbotte, Solid State Commun. **26**, 511 (1978); **34**, 599 (1980); **35**, 127 (1980). B. Mitrovic and J. P. Carbotte, Can. J. Phys. **61**, 784 (1983). E. Schachinger, M. G. Greeson, and J. P. Carbotte, Phys. Rev. B **42**, 406 (1990). Y. Yokoya and Y. O. Nakamura, Solid State Commun. **98**, 133 (1996); Physica C **262**, 187 (1996). Y. Yokoya, Phys. Rev. B **56**, 6107 (1997). A. D. Caviglia, S. Gariglio, N. Reyren, D. Jaccard, T. Schneider, M. Gabay, S. Thiel, G. Hammerl, J. Mannhart, and J. M. Triscone, Nature **456**, 624 (2008). K. Ueno, S. Nakamura, H. Shimotani, A. Ohtomo, N. Kimura, T. Nojima, H. Aoki, Y. Iwasa, and M. Kawasaki, Nat. Mater. **7**, 855 (2008) J. T. Ye, S. Inoue, K. Kobayashi, Y. Kasahara, H. T. Yuan, H. Shimotani, and Y. Iwasa, Nat. Mater. **9**, 125 (2010). K. Ueno, S. Nakamura, H. Shimotani, H. T. Yuan, N. Kimura, T. Nojima, H. Aoki, Y. Iwasa, and M. Kawasaki, Nat. Nanotechnol. **6**, 408 (2011). K. Taniguchi, A. Matsumoto, H. Shimotani, and H. Takagi, Appl. Phys. Lett. **101**, 042603 (2012). J. T. Ye, Y. J. Zhang, R. Akashi, M. S. Bahramy, R. Arita, and Y. Iwasa, Science **338**, 1193 (2012). P. B. Allen and B. Mitrović, in *Solid State Physics*, edited by H. Ehrenreich, F. Seitz, and D. Turnbull (Academic, New York, 1982), Vol. 37, p. 1. We note that the condition $\epsilon_{i}$$\gg$$T$ must be retained when we take limit $N_{\epsilon}\rightarrow\infty$. We found that it is necessary to multiply $p(\xi'\pm\omega)$ also for the terms originating from $J_{2}$; otherwise, numerically $I_{1}$$-$$2J_{2}$ does not become zero at $\xi$$=$$0$ and consequently the kernel diverges for $\xi\rightarrow 0$. R. Akashi, K. Nakamura, R. Arita, and M. Imada, Phys. Rev. B **86**, 054513 (2012). C. Felser and R. Seshadri, J. Mater. Chem. **9**, 459 (1999). The non-monotonic dependence of $T_{\rm c}^{\rm asym}$ and $T_{\rm c}^{\rm sym}$ on $E_{\rm F}$ is due to our approximation that $\lambda$ and $\mu$ are constant regardless of $E_{\rm F}$, and so we do not focus on that. S. Massidda, F. Bernardini, C. Bersier, A. Continenza, P. Cudazzo, A. Floris, H. Glawe, M. Monni, S. Pittalis, G. Profeta, A. Sanna, S. Sharma, and E. K. U. Gross, Supercond. Sci. Technol. **22**, 034006 (2009). To calculate exchange-correlation kernels and electronic energy eigenvalues, we performed electronic structure and phonon calculations within the local-density approximation [@Ceperley-Alder; @PZ81] using [*ab initio*]{} plane-wave pseudopotential calculation codes [Quantum Espresso]{}.[@Espresso; @Troullier-Martins] The pseudopotential for Al was generated in the configuration of (3$s$)$^{2.0}$(3$p$)$^{0.5}$(3$d$)$^{0.0}$, whereas that for Nb was generated in the configuration of (4$s$)$^{2.0}$(4$p$)$^{6.0}$(4$d$)$^{4.0}$. The plane-wave energy cutoff was set to 20 and 80 Ry for Al and Nb, respectively. The charge density was calculated with the 16$\times$16$\times$16 $k$ points in the Monkhorst-Pack grid for Al and Nb. The optimized lattice parameters used for the calculations were 7.483 and 6.099 bohr for Al (fcc) and Nb (bcc). The dynamical matrices for Al and Nb were calculated on the 8$\times$8$\times$8 $q$ points, where the gaussian of width 0.200 Ry was used for the Fermi-surface integration. The phonon linewidths for Al (Nb) were integrated using supplementary 32$\times$32$\times$32 (48$\times$48$\times$48) $k$ points and the gaussian of width 0.020 Ry (the first-order Hermite-gaussian function[@Methfessel-Paxton] of width 0.020 Ry). The Eliashberg functions were calculated using the tetrahedron interpolation method [@tetrahedra]. Using the output of [pw]{} code, the electron dielectric functions $\varepsilon$ were calculated within the random-phase approximation[@Hybertsen-Louie] on the 7$\times$7$\times$7 (7$\times$7$\times$7) $q$ points using the tetrahedron interpolation method with the Rath-Freeman treatment [@Rath-Freeman] considering 30 (40) bands for Al (Nb). The SCDFT gap equation was solved for Al (Nb) with 15000 (18000) $k$ points for bands crossing the Fermi level and 500 (600) points for the other bands, where we considered 12 (26) bands. D. M. Ceperley and B. J. Alder, Phys. Rev. Lett. **45**, 566 (1980). J. P. Perdew and A. Zunger, Phys. Rev. B **23**, 5048 (1981). P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. Dal Corso, S. Fabris, G. Fratesi, S. de Gironcoli, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcovitch, J. Phys.: Condens. Matter **21**, 395502 (2009); http://www.quantum-espresso.org/ N. Troullier and J. L. Martins, Phys. Rev. B **43**, 1993 (1991). M. Methfessel and A. T. Paxton, Phys. Rev. B **40**, 3616 (1989). G. Lehmann and M. Taut, Phys. Stat. Sol. **54**, 469 (1972). M. S. Hybertsen and S. G. Louie, Phys. Rev. B **35**, 5585 (1987); **35**, 5602 (1987). J. Rath and A. J. Freeman, Phys. Rev. B **11**, 2109 (1975).
| {
"pile_set_name": "ArXiv"
} |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.